1932

Abstract

The presence of human-made chemical contaminants in the environment has increased rapidly during the past 70 years. Harmful effects of such contaminants were first reported in the late 1950s in wildlife and later in humans. These effects are predominantly induced by endocrine disrupting chemicals (EDCs), chemicals that mimic the actions of endogenous hormones and leave marks at several levels of organization in organisms, from physiological outcomes (phenotypes) to molecular alterations, including epigenetic modifications. Epigenetic mechanisms play pivotal roles in the developmental processes that contribute to determining adult phenotypes, through so-called epigenetic programming. While there is increasing evidence that EDC exposure during sensitive periods of development can perturb epigenetic programming, it is unclear whether these changes are truly predictive of adverse outcomes. Understanding the mechanistic links between EDC-induced epigenetic changes and phenotypic endpoints will be critical for providing improved regulatory tools to better protect the environment and human health from exposure to EDCs.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-environ-102016-061111
2017-10-17
2024-04-16
Loading full text...

Full text loading...

/deliver/fulltext/energy/42/1/annurev-environ-102016-061111.html?itemId=/content/journals/10.1146/annurev-environ-102016-061111&mimeType=html&fmt=ahah

Literature Cited

  1. Carson R. 1.  1962. Silent Spring. Boston: Houghton Mifflin
  2. Tyler CR, Jobling S, Sumpter JP. 2.  1998. Endocrine disruption in wildlife: a critical review of the evidence. Crit. Rev. Toxicol. 28:4319–61 [Google Scholar]
  3. Guillette LJ Jr, Gross TS, Masson GR, Matter JM, Percival HF, Woodward AR. 3.  1994. Developmental abnormalities of the gonad and abnormal sex hormone concentrations in juvenile alligators from contaminated and control lakes in Florida. Environ. Health Perspect. 102:8680–88 [Google Scholar]
  4. Sonne C, Letcher RJ, Bechshøft , Rigét FF, Muir DCG. 4.  et al. 2012. Two decades of biomonitoring polar bear health in Greenland: a review. Acta Vet. Scand. 54:1S15 [Google Scholar]
  5. Herbst AL, Ulfelder H, Poskanzer DC. 5.  1971. Adenocarcinoma of the vagina—association of maternal stilbestrol therapy with tumor appearance in young women. N. Engl. J. Med. 284:15878–81 [Google Scholar]
  6. Colborn TC, Clement C. 6.  1992. Chemically-Induced Alterations in Sexual and Functional Development: The Wildlife/Human Connection, Vol. XXI Advances in Modern Environmental Toxicology Princeton, NJ: Princeton Sci. Publ.
  7. Colborn T, Dumanoski D, Peterson Myers J. 7.  1996. Our Stolen Future. New York: Dutton Penguin
  8. Jenkins S, Betancourt AM, Wang J, Lamartiniere CA. 8.  2012. Endocrine-active chemicals in mammary cancer causation and prevention. J. Steroid Biochem. Mol. Biol. 129:3–5191–200 [Google Scholar]
  9. Giwercman A, Carlsen E, Keiding N, Skakkebaek NE. 9.  1993. Evidence for increasing incidence of abnormalities of the human testis: a review. Environ. Health Perspect. 101:Suppl. 265–71 [Google Scholar]
  10. 10. World Health Org., UN Environ. Program. 2013. State of the Science of Endocrine Disrupting Chemicals—2012 Washington, DC: WHO
  11. 11. Organ. Econ. Coop. Dev. 2012. Detailed Review Paper on the State of the Science on Novel In Vitro and In Vivo Screening and Testing Methods and Endpoints for Evaluating Endocrine Disruptors Ser. Test. Assess. 178 Washington, DC: OECD
  12. Greally JM, Jacobs MN. 12.  2013. In vitro and in vivo testing methods of epigenomic endpoints for evaluating endocrine disruptors. ALTEX 30:4445–71 [Google Scholar]
  13. Marczylo EL, Jacobs MN, Gant TW. 13.  2016. Environmentally induced epigenetic toxicity: potential public health concerns. Crit. Rev. Toxicol. 46:8676–700 [Google Scholar]
  14. Gard P. 14.  1998. Endocrinology London, UK: Taylor & Francis
  15. Weigel NL. 15.  1996. Steroid hormone receptors and their regulation by phosphorylation. Biochem. J. 319:657–67 [Google Scholar]
  16. Alberts B, Bray D, Johnson A, Lewis J, Raff M. 16.  et al. 1997. Essential Cell Biology. New York: Garland:
  17. Jacobs M. 17.  2005. Nuclear receptors and dietary ligands. Nutrients and Cell Signaling J Zempleni, K Dakshinamurti 35–90 New York: CRC Press [Google Scholar]
  18. Jacobs MN, Dickins M, Lewis DF. 18.  2003. Homology modelling of the nuclear receptors: human oestrogen receptor β (hERβ), the human pregnane-X-receptor (PXR), the Ah receptor (AhR) and the constitutive androstane receptor (CAR) ligand binding domains from the human oestrogen receptor α (hERα) crystal structure, and the human peroxisome proliferator activated receptor α (PPARα) ligand binding domain from the human PPARγ crystal structure. J. Steroid Biochem. Mol. Biol. 84:2–3117–32 [Google Scholar]
  19. Thornton JW. 19.  2001. Evolution of vertebrate steroid receptors from an ancestral estrogen receptor by ligand exploitation and serial genome expansions. PNAS 98:105671–76 [Google Scholar]
  20. Brzozowski AM, Pike ACW, Dauter Z, Hubbard RE, Bonn T. 20.  et al. 1997. Molecular basis of agonism and antagonism in the oestrogen receptor. Nature 389:6652753–58 [Google Scholar]
  21. Kuiper GG, Kuiper GJM, Enmark E, Pelto-Huikko M, Nilsson S, Gustafsson J-A. 21.  1996. Cloning of a novel receptor expressed in rat prostate and ovary. PNAS 93:125925–30 [Google Scholar]
  22. 22. Eur. Food Saf. Auth. 2010. Scientific report of the Endocrine Active Substances Task Force. EFSA J 8:111932 [Google Scholar]
  23. Damstra T, Barlow S, Bergman A, Kavlock R, Van Der Kraak G. 23. , eds. 2002. Global Assessment of the State-of-the-Science of Endocrine Disruptors Washington, DC: WHO
  24. McGowan PO, Szyf M. 24.  2010. Environmental epigenomics: understanding the effects of parental care on the epigenome. Essays Biochem 48:1275–87 [Google Scholar]
  25. Kornberg RD, Lorch Y. 25.  1999. Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosome. Cell 98:3285–94 [Google Scholar]
  26. Suganuma T, Workman JL. 26.  2011. Signals and combinatorial functions of histone modifications. Annu. Rev. Biochem. 80:473–99 [Google Scholar]
  27. Bannister AJ, Kouzarides T. 27.  2011. Regulation of chromatin by histone modifications. Cell Res 21:3381–95 [Google Scholar]
  28. Bird A, Taggart M, Frommer M, Miller OJ, Macleod D. 28.  1985. A fraction of the mouse genome that is derived from islands of nonmethylated, CpG-rich DNA. Cell 40:191–99 [Google Scholar]
  29. Deaton AM, Bird A. 29.  2011. CpG islands and the regulation of transcription. Genes Dev 25:101010–22 [Google Scholar]
  30. Wu H, Zhang Y. 30.  2014. Reversing DNA methylation: mechanisms, genomics, and biological functions. Cell 156:1–245–68 [Google Scholar]
  31. Chen J, Xue Y. 31.  2016. Emerging roles of non-coding RNAs in epigenetic regulation. Sci. China Life Sci. 59:3227–35 [Google Scholar]
  32. Yang JX, Rastetter RH, Wilhelm D. 32.  2016. Non-coding RNAs: an introduction. Adv. Exp. Med. Biol 88613–32 [Google Scholar]
  33. Hamm CA, Costa FF. 33.  2015. Epigenomes as therapeutic targets. Pharmacol. Ther. 151:72–86 [Google Scholar]
  34. Amarasekera M, Martino D, Tulic MK, Saffery R, Prescott S. 34.  2012. Epigenetic aberrations in human allergic diseases. Epigenetics in Human Disease T Tollefsbol 372–80 San Diego, CA: Academic [Google Scholar]
  35. Schmitz SU, Grote P, Herrmann BG. 35.  2016. Mechanisms of long noncoding RNA function in development and disease. Cell Mol. Life Sci. 73:132491–509 [Google Scholar]
  36. Ho SM, Cheong A, Adgent MA, Veevers J, Suen AA. 36.  et al. 2016. Environmental factors, epigenetics, and developmental origin of reproductive disorders. Reprod. Toxicol. 68:85–104 [Google Scholar]
  37. Bayoumi AS, Sayed A, Broskova Z, Teoh J-P, Wilson J. 37.  et al. 2016. Crosstalk between long noncoding RNAs and microRNAs in health and disease. Int. J. Mol. Sci. 17:3356 [Google Scholar]
  38. Rzeczkowska PA, Hou H, Wilson MD, Palmert MR. 38.  2014. Epigenetics: a new player in the regulation of mammalian puberty. Neuroendocrinology 99:3–4139–55 [Google Scholar]
  39. Pierron F, Bureau du Colombier S, Moffett A, Caron A, Peluhet L. 39.  et al. 2014. Abnormal ovarian DNA methylation programming during gonad maturation in wild contaminated fish. Environ. Sci. Technol. 48:1911688–95 [Google Scholar]
  40. Guillette LJ Jr., Parrott BB, Nilsson E, Haque MM, Skinner MK. 40.  2016. Epigenetic programming alterations in alligators from environmentally contaminated lakes. Gen. Comp. Endocrinol. 238:4–12 [Google Scholar]
  41. Mirbahai L, Yin G, Bignell JP, Li N, Williams TD, Chipman JK. 41.  2011. DNA methylation in liver tumorigenesis in fish from the environment. Epigenetics 6:111319–33 [Google Scholar]
  42. Huang D, Zhang Y, Wang Y, Xie Z, Ji W. 42.  2007. Assessment of the genotoxicity in toad Bufo raddei exposed to petrochemical contaminants in Lanzhou Region, China. Mutat Res 629:281–88 [Google Scholar]
  43. Mirbahai L, Southam AD, Sommer U, Williams TD, Bignell JP. 43.  et al. 2013. Disruption of DNA methylation via S-adenosylhomocysteine is a key process in high incidence liver carcinogenesis in fish. J. Proteome Res. 12:62895–904 [Google Scholar]
  44. Vandegehuchte MB, Janssen CR. 44.  2014. Epigenetics in an ecotoxicological context. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 764–765:36–45 [Google Scholar]
  45. Metzger DC, Schulte PM. 45.  2016. Epigenomics in marine fishes. Mar. Genom. 30:43–54 [Google Scholar]
  46. Cameron BE, Craig PM, Trudeau VL. 46.  2016. Implication of microRNA deregulation in the response of vertebrates to endocrine disrupting chemicals. Environ. Toxicol. Chem. 35:4788–93 [Google Scholar]
  47. Suarez-Ulloa V, Gonzalez-Romero R, Eirin-Lopez JM. 47.  2015. Environmental epigenetics: a promising venue for developing next-generation pollution biomonitoring tools in marine invertebrates. Mar. Pollut. Bull. 98:1–25–13 [Google Scholar]
  48. Wilks A, Seldran M, Jost JP. 48.  1984. An estrogen-dependent demethylation at the 5′ end of the chicken vitellogenin gene is independent of DNA synthesis. Nucleic Acids Res 12:21163–77 [Google Scholar]
  49. Thomassin H, Flavin M, Espinás ML, Grange T. 49.  2001. Glucocorticoid-induced DNA demethylation and gene memory during development. EMBO J 20:81974–83 [Google Scholar]
  50. Bernal AJ, Jirtle RL. 50.  2010. Epigenomic disruption: the effects of early developmental exposures. Birth Defects Res. A 88:10938–44 [Google Scholar]
  51. Bollati V, Baccarelli A. 51.  2010. Environmental epigenetics. Heredity 105:1105–12 [Google Scholar]
  52. Rosenfeld CS, Sieli PT, Warzak DA, Ellersieck MR, Pennington KA, Michael Roberts R. 52.  2013. Maternal exposure to bisphenol A and genistein has minimal effect on Avy/α offspring coat color but favors birth of agouti over nonagouti mice. PNAS 110:2537–42 [Google Scholar]
  53. Wong RL, Wang Q, Treviño LS, Bosland MC, Chen J. 53.  et al. 2015. Identification of secretaglobin Scgb2a1 as a target for developmental reprogramming by BPA in the rat prostate. Epigenetics 10:2127–34 [Google Scholar]
  54. Ho SM, Tang WY, Belmonte de Frausto J, Prins GS. 54.  2006. Developmental exposure to estradiol and bisphenol A increases susceptibility to prostate carcinogenesis and epigenetically regulates phosphodiesterase type 4 variant 4. Cancer Res 66:115624–32 [Google Scholar]
  55. Tang WY, Morey LM, Cheung YY, Birch L, Prins GS, Ho S-M. 55.  2012. Neonatal exposure to estradiol/bisphenol A alters promoter methylation and expression of Nsbp1 and Hpcal1 genes and transcriptional programs of Dnmt3a/b and Mbd2/4 in the rat prostate gland throughout life. Endocrinology 153:142–55 [Google Scholar]
  56. Cheong A, Zhang X, Cheung YY, Chen J. 56.  et al. 2016. DNA methylome changes by estradiol benzoate and bisphenol A links early-life environmental exposures to prostate cancer risk. Epigenetics 11:674–89 [Google Scholar]
  57. Prados J, Stenz L, Somm E, Stouder C, Dayer A, Paoloni-Giacobino A. 57.  2015. Prenatal exposure to DEHP affects spermatogenesis and sperm DNA methylation in a strain-dependent manner. PLOS ONE 10:7e0132136 [Google Scholar]
  58. Anway MD, Rekow SS, Skinner MK. 58.  2008. Transgenerational epigenetic programming of the embryonic testis transcriptome. Genomics 91:130–40 [Google Scholar]
  59. Zama AM, Uzumcu M. 59.  2009. Fetal and neonatal exposure to the endocrine disruptor methoxychlor causes epigenetic alterations in adult ovarian genes. Endocrinology 150:104681–91 [Google Scholar]
  60. Zhou R, Chen F, Chang F, Bai Y, Chen L. 60.  2013. Persistent overexpression of DNA methyltransferase 1 attenuating GABAergic inhibition in basolateral amygdala accounts for anxiety in rat offspring exposed perinatally to low-dose bisphenol A. J. Psychiatr. Res. 47:101535–44 [Google Scholar]
  61. Liu J, Zhang L, Winterroth LC, Garcia M, Weiman S. 61.  et al. 2013. Epigenetically mediated pathogenic effects of phenanthrene on regulatory T cells. J. Toxicol. 2013:967029 [Google Scholar]
  62. Meunier L, Siddeek B, Vega A, Lakhdari N, Inoubli L. 62.  et al. 2012. Perinatal programming of adult rat germ cell death after exposure to xenoestrogens: role of microRNA miR-29 family in the down-regulation of DNA methyltransferases and Mcl-1. Endocrinology 153:41936–47 [Google Scholar]
  63. Jefferson WN, Chevalier DM, Phelps JY, Cantor AM, Padilla-Banks E. 63.  et al. 2013. Persistently altered epigenetic marks in the mouse uterus after neonatal estrogen exposure. Mol. Endocrinol. 27:101666–77 [Google Scholar]
  64. Kurian JR, Louis S, Keen KL, Wolfe A, Terasawa E, Levine JE. 64.  2016. The methylcytosine dioxygenase ten-eleven translocase-2 (tet2) enables elevated GnRH gene expression and maintenance of male reproductive function. Endocrinology 157:93588–603 [Google Scholar]
  65. Takeda T, Fujii M, Taura J, Ishii Y, Yamada H. 65.  2012. Dioxin silences gonadotropin expression in perinatal pups by inducing histone deacetylases: a new insight into the mechanism for the imprinting of sexual immaturity by dioxin. J. Biol. Chem. 287:2218440–50 [Google Scholar]
  66. Zhang J, Ali HI, Bedi YS, Choudhury M. 66.  2015. The plasticizer BBP selectively inhibits epigenetic regulator sirtuins. Toxicology 338:130–41 [Google Scholar]
  67. Greathouse KL, Bredfeldt T, Everitt JI, Lin K, Berry T. 67.  et al. 2012. Environmental estrogens differentially engage the histone methyltransferase EZH2 to increase risk of uterine tumorigenesis. Mol. Cancer Res. 10:4546–57 [Google Scholar]
  68. Wang Q, Trevino LS, Wong RLY, Medvedovic M, Chen J. 68.  et al. 2016. Reprogramming of the epigenome by MLL1 links early-life environmental exposures to prostate cancer risk. Mol. Endocrinol. 30:8856–71 [Google Scholar]
  69. Carretero MV, Latasa MU, Garcia-Trevijano ER, Corrales FJ, Wagner C. 69.  et al. 2001. Inhibition of liver methionine adenosyltransferase gene expression by 3-methylcolanthrene: protective effect of S-adenosylmethionine. Biochem. Pharmacol. 61:91119–28 [Google Scholar]
  70. Dolinoy DC, Huang D, Jirtle RL. 70.  2007. Maternal nutrient supplementation counteracts bisphenol A-induced DNA hypomethylation in early development. PNAS 104:3213056–61 [Google Scholar]
  71. Yuan C, Zhang Y, Liu Y, Zhang T, Wang Z. 71.  2016. Enhanced GSH synthesis by bisphenol A exposure promoted DNA methylation process in the testes of adult rare minnow Gobiocypris rarus. . Aquat. Toxicol. 178:99–105 [Google Scholar]
  72. Martens JH, Rao NA, Stunnenberg HG. 72.  2011. Genome-wide interplay of nuclear receptors with the epigenome. Biochim. Biophys. Acta 18128818–23 [Google Scholar]
  73. Kouzmenko A, Ohtake F, Fujiki R, Kato S. 73.  2010. Hormonal gene regulation through DNA methylation and demethylation. Epigenomics 2:6765–74 [Google Scholar]
  74. Liu Y, Duong W, Krawczyk C, Bretschneider N, Borbély N. 74.  et al. 2016. Oestrogen receptor β regulates epigenetic patterns at specific genomic loci through interaction with thymine DNA glycosylase. Epigenet. Chromatin 9:7 [Google Scholar]
  75. Kitraki E, Nalvarte I, Alavian-Ghavanini A, Rüegg J. 75.  2015. Developmental exposure to bisphenol A alters expression and DNA methylation of Fkbp5, an important regulator of the stress response. Mol. Cell Endocrinol. 417:191–99 [Google Scholar]
  76. Jorgensen EM, 3rd Alderman MH, Taylor HS. 76.  2016. Preferential epigenetic programming of estrogen response after in utero xenoestrogen (bisphenol-A) exposure. FASEB J 30:93194–201 [Google Scholar]
  77. Papoutsis AJ, Selmin OI, Borg JL, Romagnolo DF. 77.  2015. Gestational exposure to the AhR agonist 2,3,7,8-tetrachlorodibenzo-p-dioxin induces BRCA-1 promoter hypermethylation and reduces BRCA-1 expression in mammary tissue of rat offspring: preventive effects of resveratrol. Mol. Carcinog. 54:4261–69 [Google Scholar]
  78. Klinge CM. 78.  2015. miRNAs regulated by estrogens, tamoxifen, and endocrine disruptors and their downstream gene targets. Mol. Cell Endocrinol. 418:273–97 [Google Scholar]
  79. Bhan A, Hussain I, Ansari KI, Bobzean SAM, Perrotti LI, Mandal SS. 79.  2014. Bisphenol-A and diethylstilbestrol exposure induces the expression of breast cancer associated long noncoding RNA HOTAIR in vitro and in vivo. J. Steroid Biochem. Mol. Biol. 141:160–70 [Google Scholar]
  80. Gordon MW, Yan F, Zhong X, Mazumder PB, Xu-Monette ZY. 80.  et al. 2015. Regulation of p53-targeting microRNAs by polycyclic aromatic hydrocarbons: implications in the etiology of multiple myeloma. Mol. Carcinog. 54:101060–69 [Google Scholar]
  81. Guerrero-Bosagna C, Skinner MK. 81.  2012. Environmentally induced epigenetic transgenerational inheritance of phenotype and disease. Mol. Cell Endocrinol. 354:1–23–8 [Google Scholar]
  82. Alonso-Magdalena P, Rivera FJ, Guerrero-Bosagna C. 82.  2016. Bisphenol-A and metabolic diseases: epigenetic, developmental and transgenerational basis. Environ. Epigen. 2:3dvw022 [Google Scholar]
  83. Newbold RR, Hanson RB, Jefferson WN, Bullock BC, Haseman J, McLachlan JA. 83.  2000. Proliferative lesions and reproductive tract tumors in male descendants of mice exposed developmentally to diethylstilbestrol. Carcinogenesis 21:71355–63 [Google Scholar]
  84. Anway MD, Cupp AS, Uzumcu M, Skinner MK. 84.  2005. Epigenetic transgenerational actions of endocrine disruptors and male fertility. Science 308:57271466–69 [Google Scholar]
  85. Uzumcu M, Suzuki H, Skinner MK. 85.  2004. Effect of the anti-androgenic endocrine disruptor vinclozolin on embryonic testis cord formation and postnatal testis development and function. Reprod. Toxicol. 18:6765–74 [Google Scholar]
  86. Anway MD, Leathers C, Skinner MK. 86.  2006. Endocrine disruptor vinclozolin induced epigenetic transgenerational adult-onset disease. Endocrinology 147:125515–23 [Google Scholar]
  87. Anway MD, Memon MA, Uzumcu M, Skinner MK. 87.  2006. Transgenerational effect of the endocrine disruptor vinclozolin on male spermatogenesis. J. Androl. 27:6868–79 [Google Scholar]
  88. Guerrero-Bosagna C, Covert TR, Haque MM, Settles M, Nilsson EE. 88.  et al. 2012. Epigenetic transgenerational inheritance of vinclozolin induced mouse adult onset disease and associated sperm epigenome biomarkers. Reprod. Toxicol. 34:4694–707 [Google Scholar]
  89. Guerrero-Bosagna C, Settles M, Lucker B, Skinner MK. 89.  2010. Epigenetic transgenerational actions of vinclozolin on promoter regions of the sperm epigenome. PLOS ONE 5:9e13100 [Google Scholar]
  90. Stouder C, Paoloni-Giacobino A. 90.  2010. Transgenerational effects of the endocrine disruptor vinclozolin on the methylation pattern of imprinted genes in the mouse sperm. Reproduction 139:2373–79 [Google Scholar]
  91. Song Y, Wu N, Wang S, Gao M, Song P. 91.  et al. 2014. Transgenerational impaired male fertility with an Igf2 epigenetic defect in the rat are induced by the endocrine disruptor p,p′-DDE. Hum. Reprod. 29:112512–21 [Google Scholar]
  92. Ma J, Chen X, Liu Y, Xie Q, Sun Y. 92.  et al. 2015. Ancestral TCDD exposure promotes epigenetic transgenerational inheritance of imprinted gene Igf2: methylation status and DNMTs. Toxicol. Appl. Pharmacol. 289:2193–202 [Google Scholar]
  93. Skinner MK, Guerrero-Bosagna C, Haque M, Nilsson E, Bhandari R, McCarrey JR. 93.  2013. Environmentally induced transgenerational epigenetic reprogramming of primordial germ cells and the subsequent germ line. PLOS ONE 8:7e66318 [Google Scholar]
  94. Guerrero-Bosagna C, Savenkova M, Haque MM, Nilsson E, Skinner MK. 94.  2013. Environmentally induced epigenetic transgenerational inheritance of altered Sertoli cell transcriptome and epigenome: molecular etiology of male infertility. PLOS ONE 8:3e59922 [Google Scholar]
  95. Nilsson E, Larsen G, Manikkam M, Guerrero-Bosagna C, Savenkova MI, Skinner MK. 95.  2012. Environmentally induced epigenetic transgenerational inheritance of ovarian disease. PLOS ONE 7:5e36129 [Google Scholar]
  96. Guerrero-Bosagna C, Weeks S, Skinner MK. 96.  2014. Identification of genomic features in environmentally induced epigenetic transgenerational inherited sperm epimutations. PLOS ONE 9:6e100194 [Google Scholar]
  97. Manikkam M, Guerrero-Bosagna C, Tracey R, Haque MM, Skinner MK. 97.  2012. Transgenerational actions of environmental compounds on reproductive disease and epigenetic biomarkers of ancestral exposures. PLOS ONE 7:2e31901 [Google Scholar]
  98. Manikkam M, Haque MM, Gerrero-Bosagna C, Nilsson EE, Skinner MK. 98.  2014. Pesticide methoxychlor promotes the epigenetic transgenerational inheritance of adult-onset disease through the female germline. PLOS ONE 9:7e102091 [Google Scholar]
  99. Skinner MK, Manikkam M, Tracey R, Gerrero-Bosagna C, Haque MM, Nilsson EE. 99.  2013. Ancestral dichlorodiphenyltrichloroethane (DDT) exposure promotes epigenetic transgenerational inheritance of obesity. BMC Med 11:228 [Google Scholar]
  100. Guerrero-Bosagna C, Jensen P. 100.  2015. Globalization, climate change and transgenerational epigenetic inheritance: Will our descendants be at risk?. Clin. Epigenet. 7:8 [Google Scholar]
  101. Doyle TJ, Bowman JL, Windell VL, McLean DJ, Kim KH. 101.  2013. Transgenerational effects of di-(2-ethylhexyl) phthalate on testicular germ cell associations and spermatogonial stem cells in mice. Biol. Reprod. 88:5112 [Google Scholar]
  102. Bruner-Tran KL, Osteen KG. 102.  2010. Developmental exposure to TCDD reduces fertility and negatively affects pregnancy outcomes across multiple generations. Reprod. Toxicol. 31:3344–50 [Google Scholar]
  103. Nilsson EE, Anway MD, Stanfield J, Skinner MK. 103.  2008. Transgenerational epigenetic effects of the endocrine disruptor vinclozolin on pregnancies and female adult onset disease. Reproduction 135:5713–21 [Google Scholar]
  104. de Assis S, Warri A, Cruz MI, Laja O, Tian Y. 104.  et al. 2012. High-fat or ethinyl-oestradiol intake during pregnancy increases mammary cancer risk in several generations of offspring. Nat. Commun. 3:1053 [Google Scholar]
  105. Crews D, Gore AC, Hsu TS, Dangleben NL, Spinetta M. 105.  et al. 2007. Transgenerational epigenetic imprints on mate preference. PNAS 104:145942–46 [Google Scholar]
  106. Wolstenholme JT, Edwards M, Shetty SRJ, Gatewood JD, Taylor JA. 106.  et al. 2012. Gestational exposure to bisphenol a produces transgenerational changes in behaviors and gene expression. Endocrinology 153:83828–38 [Google Scholar]
  107. Wolstenholme JT, Goldsby JA, Rissman EF. 107.  2013. Transgenerational effects of prenatal bisphenol A on social recognition. Horm. Behav. 64:5833–39 [Google Scholar]
  108. Derouiche L, Keller M, Duittoz AH, Pillon D. 108.  2015. Developmental exposure to ethinylestradiol affects transgenerationally sexual behavior and neuroendocrine networks in male mice. Sci. Rep. 5:17457 [Google Scholar]
  109. Choi CS, Gonzales EL, Kim KC, Yang SM, Kim JW. 109.  et al. 2016. The transgenerational inheritance of autism-like phenotypes in mice exposed to valproic acid during pregnancy. Sci. Rep. 6:36250 [Google Scholar]
  110. Bhandari RK, vom Saal FS, Tillitt DE. 110.  2015. Transgenerational effects from early developmental exposures to bisphenol A or 17α-ethinylestradiol in medaka. Oryzias latipes. Sci. Rep. 5:9303 [Google Scholar]
  111. Baker TR, Peterson RE, Heideman W. 111.  2014. Using zebrafish as a model system for studying the transgenerational effects of dioxin. Toxicol. Sci. 138:2403–11 [Google Scholar]
  112. Salian S, Doshi T, Vanage G. 112.  2011. Perinatal exposure of rats to bisphenol A affects fertility of male offspring: an overview. Reprod. Toxicol. 31:3359–62 [Google Scholar]
  113. Salian S, Doshi T, Vanage G. 113.  2009. Impairment in protein expression profile of testicular steroid receptor coregulators in male rat offspring perinatally exposed to bisphenol A. Life Sci 85:1–211–18 [Google Scholar]
  114. Mohamed El-SA, Song W-H, Oh S-A, Park Y-J, You Y-A. 114.  et al. 2010. The transgenerational impact of benzo(a)pyrene on murine male fertility. Hum. Reprod. 25:102427–33 [Google Scholar]
  115. Sharma A. 115.  2013. Transgenerational epigenetic inheritance: focus on soma to germline information transfer. Prog. Biophys. Mol. Biol. 113:3439–46 [Google Scholar]
  116. Rajender S, Avery K, Agarwal A. 116.  2011. Epigenetics, spermatogenesis and male infertility. Mutat. Res. 727:362–71 [Google Scholar]
  117. Vlachogiannis G, Niederhuth CE, Tuna S, Stathopoulou A, Viiri K. 117.  et al. 2015. The Dnmt3L ADD domain controls cytosine methylation establishment during spermatogenesis. Cell Rep 10:944–56 [Google Scholar]
  118. Mukherjee A, Koli S, Reddy KV. 118.  2014. Regulatory non-coding transcripts in spermatogenesis: shedding light on ‘dark matter.’. Andrology 2:3360–69 [Google Scholar]
  119. Zhang X, Ho SM. 119.  2011. Epigenetics meets endocrinology. J. Mol. Endocrinol. 46:1R11–32 [Google Scholar]
  120. Chung I, Karpf AR, Muindi JR, Conroy JM, Nowak NJ. 120.  et al. 2007. Epigenetic silencing of CYP24 in tumor-derived endothelial cells contributes to selective growth inhibition by calcitriol. J. Biol. Chem. 282:128704–14 [Google Scholar]
  121. Kim H, Lapointe J, Kaygusuz G, Ong DE, Li C. 121.  et al. 2005. The retinoic acid synthesis gene ALDH1a2 is a candidate tumor suppressor in prostate cancer. Cancer Res 65:188118–24 [Google Scholar]
  122. Tian M, Zhao B, Zhang J, Martin FL, Huang Q. 122.  et al. 2016. Association of environmental benzo[a]pyrene exposure and DNA methylation alterations in hepatocellular carcinoma: a Chinese case-control study. Sci. Total Environ. 541:1243–52 [Google Scholar]
  123. Huo W, Cai P, Chen M, Li H, Tang J. 123.  et al. 2016. The relationship between prenatal exposure to BP-3 and Hirschsprung's disease. Chemosphere 144:1091–97 [Google Scholar]
  124. Kundakovic M, Gudsnuk K, Herbstman JB, Tang D, Perera FP, Champagne FA. 124.  2015. DNA methylation of BDNF as a biomarker of early-life adversity. PNAS 112:226807–13 [Google Scholar]
  125. Zhang C, Liang Y, Lei L, Zhu G, Chen X. 125.  et al. 2013. Hypermethylations of RASAL1 and KLOTHO is associated with renal dysfunction in a Chinese population environmentally exposed to cadmium. Toxicol. Appl. Pharmacol. 271:178–85 [Google Scholar]
  126. Morales E, Bustamente M, Vilahur N, Escaramis G, Montfort M. 126.  et al. 2012. DNA hypomethylation at ALOX12 is associated with persistent wheezing in childhood. Am. J. Respir. Crit. Care Med. 185:9937–43 [Google Scholar]
  127. Wang J, Zhang Y, Zhang W, Jin Y, Dai J. 127.  2012. Association of perfluorooctanoic acid with HDL cholesterol and circulating miR-26b and miR-199–3p in workers of a fluorochemical plant and nearby residents. Environ. Sci. Technol. 46:179274–81 [Google Scholar]
  128. Wang IJ, Karmaus WJJ, Chen S-L, Holloway JW, Ewart S. 128.  2015. Effects of phthalate exposure on asthma may be mediated through alterations in DNA methylation. Clin. Epigenetics 7:127 [Google Scholar]
  129. Perera F, Tang W-y, Herbstman J, Tang D, Levin L. 129.  et al. 2009. Relation of DNA methylation of 5′-CpG island of ACSL3 to transplacental exposure to airborne polycyclic aromatic hydrocarbons and childhood asthma. PLOS ONE 4:2e4488 [Google Scholar]
  130. Yang P, Ma J, Zhang B, Duan H, He Z. 130.  et al. 2012. CpG site-specific hypermethylation of p16INK4α in peripheral blood lymphocytes of PAH-exposed workers. Cancer Epidemiol. Biomark. Prev. 21:1182–90 [Google Scholar]
  131. White AJ, Chen J, Teitelbaum SL, McCullough LE, Xu X. 131.  et al. 2016. Sources of polycyclic aromatic hydrocarbons are associated with gene-specific promoter methylation in women with breast cancer. Environ. Res. 145:93–100 [Google Scholar]
  132. Nishihara R, Wang M, Qian ZR, Baba Y, Yamauchi M. 132.  et al. 2014. Alcohol, one-carbon nutrient intake, and risk of colorectal cancer according to tumor methylation level of IGF2 differentially methylated region. Am. J. Clin. Nutr. 100:61479–88 [Google Scholar]
  133. Burris HH, Baccarelli AA, Motta V, Byun H-M, Just AC. 133.  et al. 2014. Association between length of gestation and cervical DNA methylation of PTGER2 and LINE 1-HS. Epigenetics 9:81083–91 [Google Scholar]
  134. Hinz D, Bauer M, Röder S, Olek S, Huehn J. 134.  et al. 2012. Cord blood Tregs with stable FOXP3 expression are influenced by prenatal environment and associated with atopic dermatitis at the age of one year. Allergy 67:3380–89 [Google Scholar]
  135. Haworth KE, Farrell WE, Emes RD, Ismail KMK, Carroll WD. 135.  et al. 2013. Combined influence of gene-specific cord blood methylation and maternal smoking habit on birth weight. Epigenomics 5:137–49 [Google Scholar]
  136. Wang IJ, Chen S-L, Lu T-P, Chuang EY, Chen P-C. 136.  2013. Prenatal smoke exposure, DNA methylation, and childhood atopic dermatitis. Clin. Exp. Allergy 43:5535–43 [Google Scholar]
  137. Stroud LR, Papadonatos GD, Rodriguez D, McCallum M, Salisbury AL. 137.  et al. 2014. Maternal smoking during pregnancy and infant stress response: test of a prenatal programming hypothesis. Psychoneuroendocrinology 48:29–40 [Google Scholar]
  138. Kupers LK, Xu X, Jankipersadsing SA, Vaez A, la Bastide-van Gemert S. 138.  et al. 2015. DNA methylation mediates the effect of maternal smoking during pregnancy on birthweight of the offspring. Int. J. Epidemiol. 44:41224–37 [Google Scholar]
  139. Lee KW, Abrahamowicz M, Leonard GT, Richer L, Perron M. 139.  et al. 2015. Prenatal exposure to cigarette smoke interacts with OPRM1 to modulate dietary preference for fat. J. Psychiatry Neurosci. 40:138–45 [Google Scholar]
  140. Morales E, Vilahur N, Salas LA, Motta V, Fernandez MF. 140.  et al. 2016. Genome-wide DNA methylation study in human placenta identifies novel loci associated with maternal smoking during pregnancy. Int. J. Epidemiol. 45:51644–55 [Google Scholar]
  141. Hayashi H, Yazawa T, Okudela K, Nagai J-i, Ito T. 141.  et al. 2002. Inactivation of O6-methylguanine-DNA methyltransferase in human lung adenocarcinoma relates to high-grade histology and worse prognosis among smokers. Jpn. J. Cancer Res. 93:2184–89 [Google Scholar]
  142. Yanagawa N, Tamura G, Oizumi H, Takahashi N, Shimazaki Y, Motoyama T. 142.  2002. Frequent epigenetic silencing of the p16 gene in non-small cell lung cancers of tobacco smokers. Jpn. J. Cancer Res. 93:101107–13 [Google Scholar]
  143. Marsit CJ, Kim D-H, Liu M, Hinds PW, Wiencke JK. 143.  et al. 2005. Hypermethylation of RASSF1A and BLU tumor suppressor genes in non-small cell lung cancer: implications for tobacco smoking during adolescence. Int. J. Cancer 114:2219–23 [Google Scholar]
  144. Marsit CJ, Houseman EA, Schned AR, Karagas MR, Kelsey KT. 144.  2007. Promoter hypermethylation is associated with current smoking, age, gender and survival in bladder cancer. Carcinogenesis 28:81745–51 [Google Scholar]
  145. Lin RK, Hsieh Y-S, Lin P, Hsu H-S, Chen C-Y. 145.  et al. 2010. The tobacco-specific carcinogen NNK induces DNA methyltransferase 1 accumulation and tumor suppressor gene hypermethylation in mice and lung cancer patients. J. Clin. Investig. 120:2521–32 [Google Scholar]
  146. Ostrow KL, Hoque MO, Loyo M, Brait M, Greenberg A. 146.  et al. 2010. Molecular analysis of plasma DNA for the early detection of lung cancer by quantitative methylation-specific PCR. Clin. Cancer Res. 16:133463–72 [Google Scholar]
  147. Van Pottelberge GR, Mestdagh P, Bracke KR, Thas O, van Durme YMTA. 147.  et al. 2011. MicroRNA expression in induced sputum of smokers and patients with chronic obstructive pulmonary disease. Am. J. Respir. Crit. Care Med. 183:7898–906 [Google Scholar]
  148. Liu R, Liao J, Yang M, Shi Y, Peng Y. 148.  et al. 2012. Circulating miR-155 expression in plasma: a potential biomarker for early diagnosis of esophageal cancer in humans. J. Toxicol. Environ. Health A 75:181154–62 [Google Scholar]
  149. Ostrow KL, Michalidi C, Guerrero-Preston R, Hoque MO, Greenberg A. 149.  et al. 2013. Cigarette smoke induces methylation of the tumor suppressor gene NISCH. Epigenetics 8:4383–88 [Google Scholar]
  150. Shenker NS, Polidoro S, van Veldhoven K, Sacerdote C, Ricceri F. 150.  et al. 2013. Epigenome-wide association study in the European Prospective Investigation into Cancer and Nutrition (EPIC-Turin) identifies novel genetic loci associated with smoking. Hum. Mol. Genet. 22:5843–51 [Google Scholar]
  151. Stánitz E, Juhász K, Tóth C, Gombos K, Natali PG, Ember I. 151.  2013. Evaluation of microRNA expression pattern of gastric adenocarcinoma associated with socioeconomic, environmental and lifestyle factors in northwestern Hungary. Anticancer Res 33:83195–200 [Google Scholar]
  152. Sato T, Arai E, Kohno T, Takahashi Y, Miyata S. 152.  et al. 2014. Epigenetic clustering of lung adenocarcinomas based on DNA methylation profiles in adjacent lung tissue: its correlation with smoking history and chronic obstructive pulmonary disease. Int. J. Cancer 135:2319–34 [Google Scholar]
  153. Xie L, Wu M, Lin H, Liu C, Yang H. 153.  et al. 2014. An increased ratio of serum miR-21 to miR-181a levels is associated with the early pathogenic process of chronic obstructive pulmonary disease in asymptomatic heavy smokers. Mol. Biosyst. 10:51072–81 [Google Scholar]
  154. Zhang Y, Yang R, Burwinkel B, Breitling LP, Holleczek B. 154.  et al. 2014. F2RL3 methylation in blood DNA is a strong predictor of mortality. Int. J. Epidemiol. 43:41215–25 [Google Scholar]
  155. Reynolds LM, Wan M, Ding J, Taylor JR, Lohman K. 155.  et al. 2015. DNA methylation of the aryl hydrocarbon receptor repressor associations with cigarette smoking and subclinical atherosclerosis. Circ. Cardiovasc. Genet. 8:5707–16 [Google Scholar]
  156. Shui IM, Wong C-J, Zhao S, Kolb S, Ebot EM. 156.  et al. 2016. Prostate tumor DNA methylation is associated with cigarette smoking and adverse prostate cancer outcomes. Cancer 122:142168–77 [Google Scholar]
  157. Runyon RS, Cachola LM, Rajeshuni N, Hunter T, Garcia M. 157.  et al. 2012. Asthma discordance in twins is linked to epigenetic modifications of T cells. PLOS ONE 7:11e48796 [Google Scholar]
  158. Sood A, Petersen H, Blanchette CM, Meek P, Picchi MA. 158.  et al. 2010. Wood smoke exposure and gene promoter methylation are associated with increased risk for COPD in smokers. Am. J. Respir. Crit. Care Med. 182:91098–104 [Google Scholar]
  159. Liang J, Zhu H, Li C, Ding Y, Zhou Z, Wu Q. 159.  2012. Neonatal exposure to benzo[a]pyrene decreases the levels of serum testosterone and histone H3K14 acetylation of the StAR promoter in the testes of SD rats. Toxicology 302:2–3285–91 [Google Scholar]
  160. Doshi T, Mehta SS, Dighe V, Balasinor N, Vanage G. 160.  2011. Hypermethylation of estrogen receptor promoter region in adult testis of rats exposed neonatally to bisphenol A. Toxicology 289:2–374–82 [Google Scholar]
  161. Doshi T, D'Souza C, Vanage G. 161.  2013. Aberrant DNA methylation at Igf2H19 imprinting control region in spermatozoa upon neonatal exposure to bisphenol A and its association with post implantation loss. Mol. Biol. Rep. 40:84747–57 [Google Scholar]
  162. Hong J, Chen F, Wang X, Bai Y, Zhou R. 162.  et al. 2016. Exposure of preimplantation embryos to low-dose bisphenol A impairs testes development and suppresses histone acetylation of StAR promoter to reduce production of testosterone in mice. Mol. Cell Endocrinol. 427:101–11 [Google Scholar]
  163. Zhang L, Ding S, Qiao P, Dong L, Yu M. 163.  et al. 2016. n-Butylparaben induces male reproductive disorders via regulation of estradiol and estrogen receptors. J. Appl. Toxicol. 36:91223–34 [Google Scholar]
  164. Sekaran S, Jagadeesan A. 164.  2015. In utero exposure to phthalate downregulates critical genes in Leydig cells of F1 male progeny. J. Cell Biochem. 116:71466–77 [Google Scholar]
  165. Choi JS, Oh J-H, Park H-J, Choi M-S, Park S-M. 165.  et al. 2011. miRNA regulation of cytotoxic effects in mouse Sertoli cells exposed to nonylphenol. Reprod. Biol. Endocrinol. 9:126 [Google Scholar]
  166. Trapphoff T, Heiligentag M, El Hajj N, Haaf T, Eichenlaub-Ritter U. 166.  2013. Chronic exposure to a low concentration of bisphenol A during follicle culture affects the epigenetic status of germinal vesicles and metaphase II oocytes. Fertil. Steril. 100:61758–67.e1 [Google Scholar]
  167. Chao HH, Zhang X-F, Chen B, Pan B, Zhang L-J. 167.  et al. 2012. Bisphenol A exposure modifies methylation of imprinted genes in mouse oocytes via the estrogen receptor signaling pathway. Histochem. Cell Biol. 137:2249–59 [Google Scholar]
  168. Savabieasfahani M, Kannan K, Astapova O, Evans NP, Padmanabhan V. 168.  2006. Developmental programming: differential effects of prenatal exposure to bisphenol-A or methoxychlor on reproductive function. Endocrinology 147:125956–66 [Google Scholar]
  169. Veiga-Lopez A, Luense LJ, Christenson LK, Padmanabhan V. 169.  2013. Developmental programming: gestational bisphenol-A treatment alters trajectory of fetal ovarian gene expression. Endocrinology 154:51873–84 [Google Scholar]
  170. Avissar-Whiting M, Veiga KR, Uhl KM, Maccani MA, Gagne LA. 170.  et al. 2010. Bisphenol A exposure leads to specific microRNA alterations in placental cells. Reprod. Toxicol. 29:4401–6 [Google Scholar]
  171. Altamirano GA, Ramos JG, Gomez AL, Luque EH, Muñoz-de-Toro M, Kass L. 171.  2016. Perinatal exposure to bisphenol A modifies the transcriptional regulation of the β-Casein gene during secretory activation of the rat mammary gland. Mol. Cell Endocrinol. 439:5407–18 [Google Scholar]
  172. Zhang XF, Zhang L-J, Li L, Feng Y-N, Chen B. 172.  et al. 2013. Diethylhexyl phthalate exposure impairs follicular development and affects oocyte maturation in the mouse. Environ. Mol. Mutagen. 54:5354–61 [Google Scholar]
  173. Meruvu S, Zhang J, Bedi YS, Choudhury M. 173.  2016. Mono-(2-ethylhexyl) phthalate induces apoptosis through miR-16 in human first trimester placental cell line HTR-8/SVneo. Toxicol. Vitro 31:35–42 [Google Scholar]
  174. Matsumoto Y, Hannigan B, Crews D. 174.  2014. Embryonic PCB exposure alters phenotypic, genetic, and epigenetic profiles in turtle sex determination, a biomarker of environmental contamination. Endocrinology 155:114168–77 [Google Scholar]
  175. Cai JL, Liu L-L, Hu Y, Jiang X-M, Qui H-L. 175.  et al. 2016. Polychlorinated biphenyls impair endometrial receptivity in vitro via regulating mir-30d expression and epithelial mesenchymal transition. Toxicology 365:25–34 [Google Scholar]
  176. Kamstra JH, Hruba E, Blumberg B, Janesick A, Mandrup S. 176.  et al. 2014. Transcriptional and epigenetic mechanisms underlying enhanced in vitro adipocyte differentiation by the brominated flame retardant BDE-47. Environ. Sci. Technol. 48:74110–19 [Google Scholar]
  177. Ke ZH, Pan JX, Jin LY, Xu HY, Yu TT. 177.  et al. 2016. Bisphenol A exposure may induce hepatic lipid accumulation via reprogramming the DNA methylation patterns of genes involved in lipid metabolism. Sci. Rep. 6:31331 [Google Scholar]
  178. Rajesh P, Balasubramanian K. 178.  2014. Phthalate exposure in utero causes epigenetic changes and impairs insulin signalling. J. Endocrinol. 223:147–66 [Google Scholar]
  179. Fernandez SV, Huang Y, Snider KE, Zhou Y, Pogash TJ, Russo J. 179.  2012. Expression and DNA methylation changes in human breast epithelial cells after bisphenol A exposure. Int. J. Oncol. 41:1369–77 [Google Scholar]
  180. Qin XY, Fukuda T, Yang L, Zaha H, Akanuma H. 180.  et al. 2012. Effects of bisphenol A exposure on the proliferation and senescence of normal human mammary epithelial cells. Cancer Biol. Ther. 13:5296–306 [Google Scholar]
  181. Weinhouse C, Sartor MA, Faulk C, Anderson OS, Sant KE. 181.  et al. 2016. Epigenome-wide DNA methylation analysis implicates neuronal and inflammatory signaling pathways in adult murine hepatic tumorigenesis following perinatal exposure to bisphenol A. Environ. Mol. Mutagen. 57:6435–46 [Google Scholar]
  182. Kim JH, Sartor MA, Rozek LS, Faulk C, Anderson OS. 182.  et al. 2014. Perinatal bisphenol A exposure promotes dose-dependent alterations of the mouse methylome. BMC Genom 15:30 [Google Scholar]
  183. Li Y, Hu Y, Dong C, Lu H, Zhang C. 183.  et al. 2016. Vimentin-mediated steroidogenesis induced by phthalate esters: involvement of DNA demethylation and nuclear factor κB. PLOS ONE 11:1e0146138 [Google Scholar]
  184. Lu H, Zhang C, Hu Y, Qin H, Gu A. 184.  et al. 2016. miRNA-200c mediates mono-butyl phthalate-disrupted steroidogenesis by targeting vimentin in Leydig tumor cells and murine adrenocortical tumor cells. Toxicol. Lett. 241:95–102 [Google Scholar]
  185. Huang Q, Chen Y, Chen Q, Zhang H, Lin Y. 185.  et al. 2016. Dioxin-like rather than non-dioxin-like PCBs promote the development of endometriosis through stimulation of endocrine-inflammation interactions. Arch. Toxicol. 91:1915 [Google Scholar]
  186. Yoshioka W, Higashiyama W, Tohyama C. 186.  2011. Involvement of microRNAs in dioxin-induced liver damage in the mouse. Toxicol. Sci. 122:2457–65 [Google Scholar]
  187. Kundakovic M, Gudsnuk K, Franks B, Madrid J, Miller RL. 187.  et al. 2013. Sex-specific epigenetic disruption and behavioral changes following low-dose in utero bisphenol A exposure. PNAS 110:249956–61 [Google Scholar]
  188. Lesiak A, Zhu M, Chen H, Appleyard SM, Impey S. 188.  et al. 2014. The environmental neurotoxicant PCB 95 promotes synaptogenesis via ryanodine receptor-dependent miR132 upregulation. J. Neurosci. 34:3717–25 [Google Scholar]
  189. Jiang Y, Xia W, Yang J, Zhu Y, Chang H. 189.  et al. 2015. BPA-induced DNA hypermethylation of the master mitochondrial gene PGC-1α contributes to cardiomyopathy in male rats. Toxicology 329:21–31 [Google Scholar]
  190. Patel BB, Raad M, Sebag IA, Chalifour LE. 190.  2013. Lifelong exposure to bisphenol A alters cardiac structure/function, protein expression, and DNA methylation in adult mice. Toxicol. Sci. 133:1174–85 [Google Scholar]
  191. Haddad R, Kasneci A, Mepham K, Sebag IA, Chalifour LE. 191.  2013. Gestational exposure to diethylstilbestrol alters cardiac structure/function, protein expression and DNA methylation in adult male mice progeny. Toxicol. Appl. Pharmacol. 266:127–37 [Google Scholar]
  192. Haddad R, Kasneci A, Sebag IA, Chalifour LE. 192.  2013. Cardiac structure/function, protein expression, and DNA methylation are changed in adult female mice exposed to diethylstilbestrol in utero. Can. J. Physiol. Pharmacol. 91:9741–49 [Google Scholar]
  193. Singh NP, Singh UP, Guan H, Nagarkatti P, Nagarkatti M. 193.  2012. Prenatal exposure to TCDD triggers significant modulation of microRNA expression profile in the thymus that affects consequent gene expression. PLOS ONE 7:9e45054 [Google Scholar]
  194. Camacho IA, Nagarkatti M, Nagarkatti PS. 194.  2004. Evidence for induction of apoptosis in T cells from murine fetal thymus following perinatal exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Toxicol. Sci. 78:196–106 [Google Scholar]
  195. Chen X, Xu B, Han X, Mao Z, Talbot P. 195.  et al. 2013. Effect of bisphenol A on pluripotency of mouse embryonic stem cells and differentiation capacity in mouse embryoid bodies. Toxicol. In Vitro 27:82249–55 [Google Scholar]
  196. Chen X, Xu B, Han X, Mao Z, Chen M. 196.  et al. 2015. The effects of triclosan on pluripotency factors and development of mouse embryonic stem cells and zebrafish. Arch. Toxicol. 89:4635–46 [Google Scholar]
  197. Gao L, Yin J, Wu W. 197.  2016. Long non-coding RNA H19-mediated mouse cleft palate induced by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Exp. Ther. Med. 11:62355–60 [Google Scholar]
  198. Manikkam M, Tracey R, Guerrero-Bosagna C, Skinner MK. 198.  2012. Plastics derived endocrine disruptors (BPA, DEHP and DBP) induce epigenetic transgenerational inheritance of obesity, reproductive disease and sperm epimutations. PLOS ONE 8:1e55387 [Google Scholar]
  199. Tracey R, Manikkam M, Guerrero-Bosagna C, Skinner MK. 199.  2012. Hydrocarbon (jet fuel JP-8) induces epigenetic transgenerational inheritance of adult-onset disease and sperm epimutations. Reprod. Toxicol. 36:104–16 [Google Scholar]
  200. Manikkam M, Tracey R, Guerrero-Bosagna C, Skinner MK. 200.  2012. Pesticide and insect repellent mixture (permethrin and DEET) induces epigenetic transgenerational inheritance of disease and sperm epimutations. Reprod. Toxicol. 34:4708–19 [Google Scholar]
  201. Manikkam M, Tracey R, Guerrero-Bosagna C, Skinner MK. 201.  2012. Dioxin (TCDD) induces epigenetic transgenerational inheritance of adult onset disease and sperm epimutations. PLOS ONE 7:9e46249 [Google Scholar]
/content/journals/10.1146/annurev-environ-102016-061111
Loading
/content/journals/10.1146/annurev-environ-102016-061111
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error