Register      Login
Reproduction, Fertility and Development Reproduction, Fertility and Development Society
Vertebrate reproductive science and technology
RESEARCH ARTICLE

Mechanisms of epigenetic remodelling during preimplantation development

Pablo Juan Ross A C and Sebastian Canovas B
+ Author Affiliations
- Author Affiliations

A Department of Animal Science, University of California, Davis, CA 95616 USA.

B LARCEL (Laboratorio Andaluz de Reprogramación Celular), BIONAND, Centro Andaluz de Nanomedicina y Biotecnología Campanillas, Malaga 29590, Spain.

C Corresponding author. Email: pross@ucdavis.edu

Reproduction, Fertility and Development 28(2) 25-40 https://doi.org/10.1071/RD15365
Published: 3 December 2015

Abstract

Epigenetics involves mechanisms independent of modifications in the DNA sequence that result in changes in gene expression and are maintained through cell divisions. Because all cells in the organism contain the same genetic blueprint, epigenetics allows for cells to assume different phenotypes and maintain them upon cell replication. As such, during the life cycle, there are moments in which the epigenetic information needs to be reset for the initiation of a new organism. In mammals, the resetting of epigenetic marks occurs at two different moments, which both happen to be during gestation, and include primordial germ cells (PGCs) and early preimplantation embryos. Because epigenetic information is reversible and sensitive to environmental changes, it is probably no coincidence that both these extensive periods of epigenetic remodelling happen in the female reproductive tract, under a finely controlled maternal environment. It is becoming evident that perturbations during the extensive epigenetic remodelling in PGCs and embryos can lead to permanent and inheritable changes to the epigenome that can result in long-term changes to the offspring derived from them, as indicated by the Developmental Origins of Health and Disease (DOHaD) hypothesis and recent demonstration of inter- and trans-generational epigenetic alterations. In this context, an understanding of the mechanisms of epigenetic remodelling during early embryo development is important to assess the potential for gametic epigenetic mutations to contribute to the offspring and for new epimutations to be established during embryo manipulations that could affect a large number of cells in the offspring. It is of particular interest to understand whether and how epigenetic information can be passed on from the gametes to the embryo or offspring, and whether abnormalities in this process could lead to transgenerationally inheritable phenotypes. The aim of this review is to highlight recent progress made in understanding the nature and mechanisms of epigenetic remodelling that ensue after fertilisation.

Additional keywords: DNA methylation, embryo, epigenetic inheritance, histone mofidication, imprinting, nuclear reprogramming.


References

Adenot, P. G., Mercier, Y., Renard, J. P., and Thompson, E. M. (1997). Differential H4 acetylation of paternal and maternal chromatin precedes DNA replication and differential transcriptional activity in pronuclei of 1-cell mouse embryos. Development 124, 4615–4625.
| 1:CAS:528:DyaK1cXht12jsA%3D%3D&md5=5202c4d0f3d3117efa41c52e72e02dc3CAS | 9409678PubMed |

Ahmad, K., and Henikoff, S. (2002). The histone variant H3.3 marks active chromatin by replication-independent nucleosome assembly. Mol. Cell 9, 1191–1200.
The histone variant H3.3 marks active chromatin by replication-independent nucleosome assembly.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38Xlt1Sksbs%3D&md5=cf80b5a8e16c468a4dbd0374357b5a3eCAS | 12086617PubMed |

Akiyama, T., Suzuki, O., Matsuda, J., and Aoki, F. (2011). Dynamic replacement of histone H3 variants reprograms epigenetic marks in early mouse embryos. PLoS Genet. 7, e1002279.
Dynamic replacement of histone H3 variants reprograms epigenetic marks in early mouse embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXhtl2mtbjN&md5=18042c357d6c32c2ed21026caf907bcfCAS | 21998593PubMed |

Alder, O., Lavial, F., Helness, A., Brookes, E., Pinho, S., Chandrashekran, A., Arnaud, P., Pombo, A., O’Neill, L., and Azuara, V. (2010). Ring1B and Suv39h1 delineate distinct chromatin states at bivalent genes during early mouse lineage commitment. Development 137, 2483–2492.
Ring1B and Suv39h1 delineate distinct chromatin states at bivalent genes during early mouse lineage commitment.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXhtF2mt7bN&md5=20c3e57beafe17b5129e7f7308ad4b77CAS | 20573702PubMed |

Alukal, J. P., and Lipshultz, L. I. (2008). Safety of assisted reproduction, assessed by risk of abnormalities in children born after use of in vitro fertilization techniques. Nat. Clin. Pract. Urol. 5, 140–150.
Safety of assisted reproduction, assessed by risk of abnormalities in children born after use of in vitro fertilization techniques.Crossref | GoogleScholarGoogle Scholar | 18253110PubMed |

Aoshima, K., Inoue, E., Sawa, H., and Okada, Y. (2015). Paternal H3K4 methylation is required for minor zygotic gene activation and early mouse embryonic development. EMBO Rep. 16, 803–812.
Paternal H3K4 methylation is required for minor zygotic gene activation and early mouse embryonic development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtFWgtbjL&md5=7b3cad2bc29f00dbe9ad218a512a241fCAS | 25925669PubMed |

Arand, J., Spieler, D., Karius, T., Branco, M. R., Meilinger, D., Meissner, A., Jenuwein, T., Xu, G., Leonhardt, H., Wolf, V., and Walter, J. (2012). In vivo control of CpG and non-CpG DNA methylation by DNA methyltransferases. PLoS Genet. 8, e1002750.
In vivo control of CpG and non-CpG DNA methylation by DNA methyltransferases.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XpvVWru74%3D&md5=7a7026d94b4c1909b72608ae794daac3CAS | 22761581PubMed |

Arand, J., Wossidlo, M., Lepikhov, K., Peat, J. R., Reik, W., and Walter, J. (2015). Selective impairment of methylation maintenance is the major cause of DNA methylation reprogramming in the early embryo. Epigenetics Chromatin 8, 1.
Selective impairment of methylation maintenance is the major cause of DNA methylation reprogramming in the early embryo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXjtlKnurc%3D&md5=4905014481e05dc6810d5b109ad8ec16CAS | 25621012PubMed |

Aravind, L., and Koonin, E. V. (2000). SAP: a putative DNA-binding motif involved in chromosomal organization. Trends Biochem. Sci. 25, 112–114.
SAP: a putative DNA-binding motif involved in chromosomal organization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXhsFOisrs%3D&md5=b9a0a8994aaac94ced83dabaaec84310CAS | 10694879PubMed |

Athanasiadou, R., de Sousa, D., Myant, K., Merusi, C., Stancheva, I., and Bird, A. (2010). Targeting of de novo DNA methylation throughout the Oct-4 gene regulatory region in differentiating embryonic stem cells. PLoS One 5, e9937.
Targeting of de novo DNA methylation throughout the Oct-4 gene regulatory region in differentiating embryonic stem cells.Crossref | GoogleScholarGoogle Scholar | 20376339PubMed |

Azuara, V., Perry, P., Sauer, S., Spivakov, M., Jorgensen, H. F., John, R. M., Gouti, M., Casanova, M., Warnes, G., Merkenschlager, M., and Fisher, A. G. (2006). Chromatin signatures of pluripotent cell lines. Nat. Cell Biol. 8, 532–538.
Chromatin signatures of pluripotent cell lines.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XksVGnurk%3D&md5=2f14b1fa2b9170ed868a11a77a7353c7CAS | 16570078PubMed |

Bakhtari, A., and Ross, P. J. (2014). DPPA3 prevents cytosine hydroxymethylation of the maternal pronucleus and is required for normal development in bovine embryos. Epigenetics 9, 1271–1279.
DPPA3 prevents cytosine hydroxymethylation of the maternal pronucleus and is required for normal development in bovine embryos.Crossref | GoogleScholarGoogle Scholar | 25147917PubMed |

Bakhtari, A., Rahmani, H. R., Bonakdar, E., Jafarpour, F., Asgari, V., Hosseini, S. M., Hajian, M., Edriss, M. A., and Nasr-Esfahani, M. H. (2014). The interfering effects of superovulation and vitrification upon some important epigenetic biomarkers in mouse blastocyst. Cryobiology 69, 419–427.
The interfering effects of superovulation and vitrification upon some important epigenetic biomarkers in mouse blastocyst.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhsl2ks73K&md5=a9f331b1af34dca65785c829074790c0CAS | 25307438PubMed |

Batcheller, A., Cardozo, E., Maguire, M., DeCherney, A. H., and Segars, J. H. (2011). Are there subtle genome-wide epigenetic alterations in normal offspring conceived by assisted reproductive technologies? Fertil. Steril. 96, 1306–1311.
Are there subtle genome-wide epigenetic alterations in normal offspring conceived by assisted reproductive technologies?Crossref | GoogleScholarGoogle Scholar | 22035969PubMed |

Beaujean, N. (2014). Histone post-translational modifications in preimplantation mouse embryos and their role in nuclear architecture. Mol. Reprod. Dev. 81, 100–112.
Histone post-translational modifications in preimplantation mouse embryos and their role in nuclear architecture.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhvV2lur3K&md5=59b1ea67fefc14dcec02e90211c344f7CAS | 24150914PubMed |

Beaujean, N. (2015). Epigenetics, embryo quality and developmental potential. Reprod. Fertil. Dev. 27, 53–62.
Epigenetics, embryo quality and developmental potential.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXitVylt7jP&md5=1ed6470449cf51083ce61130297207f0CAS |

Behboodi, E., Anderson, G. B., BonDurant, R. H., Cargill, S. L., Kreuscher, B. R., Medrano, J. F., and Murray, J. D. (1995). Birth of large calves that developed from in vitro-derived bovine embryos. Theriogenology 44, 227–232.
Birth of large calves that developed from in vitro-derived bovine embryos.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BD28zgtVCkuw%3D%3D&md5=3496b4d96bf9d6b230b4d01166373f08CAS | 16727722PubMed |

Bernstein, B. E., Mikkelsen, T. S., Xie, X., Kamal, M., Huebert, D. J., Cuff, J., Fry, B., Meissner, A., Wernig, M., Plath, K., Jaenisch, R., Wagschal, A., Feil, R., Schreiber, S. L., and Lander, E. S. (2006). A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 125, 315–326.
A bivalent chromatin structure marks key developmental genes in embryonic stem cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xkt1Oqur4%3D&md5=0df58a80e6e881253cee232a3b86966dCAS | 16630819PubMed |

Bian, C., and Yu, X. (2014). PGC7 suppresses TET3 for protecting DNA methylation. Nucleic Acids Res. 42, 2893–2905.
PGC7 suppresses TET3 for protecting DNA methylation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXktleku78%3D&md5=76527456716e2c740b330b8f8165753aCAS | 24322296PubMed |

Black, J. C., Van Rechem, C., and Whetstine, J. R. (2012). Histone lysine methylation dynamics: establishment, regulation, and biological impact. Mol. Cell 48, 491–507.
Histone lysine methylation dynamics: establishment, regulation, and biological impact.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhslKqtLvM&md5=d846ad4c28b9efbcf80bc368e2896de4CAS | 23200123PubMed |

Bogliotti, Y. S., and Ross, P. J. (2012). Mechanisms of histone H3 lysine 27 trimethylation remodeling during early mammalian development. Epigenetics 7, 976–981.
Mechanisms of histone H3 lysine 27 trimethylation remodeling during early mammalian development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhs1aru7k%3D&md5=2859f7e98cdfeab9f42bbcbe09eebd0aCAS | 22895114PubMed |

Bogliotti, Y. S., and Ross, P. J. (2015). Molecular mechanisms of transcriptional and chromatin remodeling around embryonic genome activation. Anim. Reprod. 12, 52–61.

Bostick, M., Kim, J. K., Esteve, P. O., Clark, A., Pradhan, S., and Jacobsen, S. E. (2007). UHRF1 plays a role in maintaining DNA methylation in mammalian cells. Science 317, 1760–1764.
UHRF1 plays a role in maintaining DNA methylation in mammalian cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXhtVGmsLbN&md5=f904b6dd94c55f39139014d1cb6ad414CAS | 17673620PubMed |

Bourc’his, D., Xu, G. L., Lin, C. S., Bollman, B., and Bestor, T. H. (2001). Dnmt3L and the establishment of maternal genomic imprints. Science 294, 2536–2539.
Dnmt3L and the establishment of maternal genomic imprints.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38Xns1Or&md5=a0ab69688cbbcc9b714bb7e74de40896CAS | 11719692PubMed |

Branco, M. R., Ficz, G., and Reik, W. (2012). Uncovering the role of 5-hydroxymethylcytosine in the epigenome. Nat. Rev. Genet. 13, 7–13.
| 1:CAS:528:DC%2BC3MXhsVKjtr7L&md5=ed08b06b3742dce28b2af93467c7419eCAS |

Calarco, J. P., Borges, F., Donoghue, M. T., Van Ex, F., Jullien, P. E., Lopes, T., Gardner, R., Berger, F., Feijo, J. A., Becker, J. D., and Martienssen, R. A. (2012). Reprogramming of DNA methylation in pollen guides epigenetic inheritance via small RNA. Cell 151, 194–205.
Reprogramming of DNA methylation in pollen guides epigenetic inheritance via small RNA.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xhtl2rsr3P&md5=7870a0699e821912b10a464261adad84CAS | 23000270PubMed |

Canovas, S., Cibelli, J. B., and Ross, P. J. (2012). Jumonji domain-containing protein 3 regulates histone 3 lysine 27 methylation during bovine preimplantation development. Proc. Natl Acad. Sci. USA 109, 2400–2405.
Jumonji domain-containing protein 3 regulates histone 3 lysine 27 methylation during bovine preimplantation development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XivFSjsLs%3D&md5=c5a9191c63fe3fa02cb5eeac3ab421d0CAS | 22308433PubMed |

Cantone, I., and Fisher, A. G. (2013). Epigenetic programming and reprogramming during development. Nat. Struct. Mol. Biol. 20, 282–289.
Epigenetic programming and reprogramming during development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXjsFOnsbk%3D&md5=f3e6ee6e41e4f405e8e58e0cf1f96cc2CAS | 23463313PubMed |

Cao, Z., Zhou, N., Zhang, Y., Zhang, Y., Wu, R., Li, Y., Zhang, Y., and Li, N. (2014). Dynamic reprogramming of 5-hydroxymethylcytosine during early porcine embryogenesis. Theriogenology 81, 496–508.
Dynamic reprogramming of 5-hydroxymethylcytosine during early porcine embryogenesis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhvFyru73I&md5=6608fd1963a9b6e3e0afb58946b2dfdbCAS | 24315686PubMed |

Chen, S. S., Raval, A., Johnson, A. J., Hertlein, E., Liu, T. H., Jin, V. X., Sherman, M. H., Liu, S. J., Dawson, D. W., Williams, K. E., Lanasa, M., Liyanarachchi, S., Lin, T. S., Marcucci, G., Pekarsky, Y., Davuluri, R., Croce, C. M., Guttridge, D. C., Teitell, M. A., Byrd, J. C., and Plass, C. (2009). Epigenetic changes during disease progression in a murine model of human chronic lymphocytic leukemia. Proc. Natl Acad. Sci. USA 106, 13 433–13 438.
Epigenetic changes during disease progression in a murine model of human chronic lymphocytic leukemia.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhtVKqsL%2FL&md5=7e21bc53f71b209759ae28e28cb73574CAS |

Cirio, M. C., Ratnam, S., Ding, F., Reinhart, B., Navara, C., and Chaillet, J. R. (2008). Preimplantation expression of the somatic form of Dnmt1 suggests a role in the inheritance of genomic imprints. BMC Dev. Biol. 8, 9.
Preimplantation expression of the somatic form of Dnmt1 suggests a role in the inheritance of genomic imprints.Crossref | GoogleScholarGoogle Scholar | 18221528PubMed |

Cloos, P. A., Christensen, J., Agger, K., and Helin, K. (2008). Erasing the methyl mark: histone demethylases at the center of cellular differentiation and disease. Genes Dev. 22, 1115–1140.
Erasing the methyl mark: histone demethylases at the center of cellular differentiation and disease.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXlslCqsr0%3D&md5=0c34133918123406831b92998f8b7a40CAS | 18451103PubMed |

Cortellino, S., Xu, J., Sannai, M., Moore, R., Caretti, E., Cigliano, A., Le Coz, M., Devarajan, K., Wessels, A., Soprano, D., Abramowitz, L. K., Bartolomei, M. S., Rambow, F., Bassi, M. R., Bruno, T., Fanciulli, M., Renner, C., Klein-Szanto, A. J., Matsumoto, Y., Kobi, D., Davidson, I., Alberti, C., Larue, L., and Bellacosa, A. (2011). Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell 146, 67–79.
Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXoslCks74%3D&md5=3adcfac08a98f8bc7e30bdd8f26dc3feCAS | 21722948PubMed |

Cotton, A. M., Avila, L., Penaherrera, M. S., Affleck, J. G., Robinson, W. P., and Brown, C. J. (2009). Inactive X chromosome-specific reduction in placental DNA methylation. Hum. Mol. Genet. 18, 3544–3552.
Inactive X chromosome-specific reduction in placental DNA methylation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhtFagsbbP&md5=6c76cb58e074dbe46ade3928f290f629CAS | 19586922PubMed |

Dahl, J. A., Reiner, A. H., Klungland, A., Wakayama, T., and Collas, P. (2010). Histone H3 lysine 27 methylation asymmetry on developmentally-regulated promoters distinguish the first two lineages in mouse preimplantation embryos. PLoS One 5, e9150.
Histone H3 lysine 27 methylation asymmetry on developmentally-regulated promoters distinguish the first two lineages in mouse preimplantation embryos.Crossref | GoogleScholarGoogle Scholar | 20161773PubMed |

Davies, M. J., Moore, V. M., Willson, K. J., Van Essen, P., Priest, K., Scott, H., Haan, E. A., and Chan, A. (2012). Reproductive technologies and the risk of birth defects. N. Engl. J. Med. 366, 1803–1813.
Reproductive technologies and the risk of birth defects.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xnt1Gjsrk%3D&md5=d51f492861570a8cc6ec93ae7129ae6eCAS | 22559061PubMed |

Dawlaty, M. M., Ganz, K., Powell, B. E., Hu, Y. C., Markoulaki, S., Cheng, A. W., Gao, Q., Kim, J., Choi, S. W., Page, D. C., and Jaenisch, R. (2011). Tet1 is dispensable for maintaining pluripotency and its loss is compatible with embryonic and postnatal development. Cell Stem Cell 9, 166–175.
Tet1 is dispensable for maintaining pluripotency and its loss is compatible with embryonic and postnatal development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXpvFGit74%3D&md5=d945d53e99dcdd15edd82e14e0b6c4f4CAS | 21816367PubMed |

Dawlaty, M. M., Breiling, A., Le, T., Raddatz, G., Barrasa, M. I., Cheng, A. W., Gao, Q., Powell, B. E., Li, Z., Xu, M., Faull, K. F., Lyko, F., and Jaenisch, R. (2013). Combined deficiency of Tet1 and Tet2 causes epigenetic abnormalities but is compatible with postnatal development. Dev. Cell 24, 310–323.
Combined deficiency of Tet1 and Tet2 causes epigenetic abnormalities but is compatible with postnatal development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhsFCqsLc%3D&md5=0a2d78e16ef8da96290e79763a7d57f1CAS | 23352810PubMed |

Dawlaty, M. M., Breiling, A., Le, T., Barrasa, M. I., Raddatz, G., Gao, Q., Powell, B. E., Cheng, A. W., Faull, K. F., Lyko, F., and Jaenisch, R. (2014). Loss of Tet enzymes compromises proper differentiation of embryonic stem cells. Dev. Cell 29, 102–111.
Loss of Tet enzymes compromises proper differentiation of embryonic stem cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXmt1agu78%3D&md5=2bf9c168ac9f8b4ecee939f2441e1beeCAS | 24735881PubMed |

Dean, W., Santos, F., and Reik, W. (2003). Epigenetic reprogramming in early mammalian development and following somatic nuclear transfer. Semin. Cell Dev. Biol. 14, 93–100.
Epigenetic reprogramming in early mammalian development and following somatic nuclear transfer.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXnvFar&md5=7a3806dd4c3fb2f82f04bafb43e55e2eCAS | 12524012PubMed |

DeBaun, M. R., Niemitz, E. L., and Feinberg, A. P. (2003). Association of in vitro fertilization with Beckwith–Wiedemann syndrome and epigenetic alterations of LIT1 and H19. Am. J. Hum. Genet. 72, 156–160.
Association of in vitro fertilization with Beckwith–Wiedemann syndrome and epigenetic alterations of LIT1 and H19.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXis1ygug%3D%3D&md5=7ca01c0685aa0ebf02f9b20ab89fbbbbCAS | 12439823PubMed |

Du, M., Tong, J., Zhao, J., Underwood, K. R., Zhu, M., Ford, S. P., and Nathanielsz, P. W. (2010). Fetal programming of skeletal muscle development in ruminant animals. J. Anim. Sci. 88, E51–E60.
Fetal programming of skeletal muscle development in ruminant animals.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BC3c3pslOisw%3D%3D&md5=618e292d9c7c07415e4ae1b7a6c744c5CAS | 19717774PubMed |

Dwivedi, R. S., Herman, J. G., McCaffrey, T. A., and Raj, D. S. (2011). Beyond genetics: epigenetic code in chronic kidney disease. Kidney Int. 79, 23–32.
Beyond genetics: epigenetic code in chronic kidney disease.Crossref | GoogleScholarGoogle Scholar | 20881938PubMed |

Ecker, D. J., Stein, P., Xu, Z., Williams, C. J., Kopf, G. S., Bilker, W. B., Abel, T., and Schultz, R. M. (2004). Long-term effects of culture of preimplantation mouse embryos on behavior. Proc. Natl Acad. Sci. USA 101, 1595–1600.
Long-term effects of culture of preimplantation mouse embryos on behavior.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXhsFaqtrg%3D&md5=d9a26e29fe43fda4a58fb7db45e23fdbCAS | 14747652PubMed |

Farin, P. W., and Farin, C. E. (1995). Transfer of bovine embryos produced in vivo or in vitro: survival and fetal development. Biol. Reprod. 52, 676–682.
Transfer of bovine embryos produced in vivo or in vitro: survival and fetal development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2MXjslChtr8%3D&md5=2cf4a42c71de5c5b7e217bd30abed7b5CAS | 7756461PubMed |

Farin, C. E., Farin, P.W., and Piedrahita, J. A. (2004). Development of fetuses from in vitro-produced and cloned bovine embryos. J. Anim. Sci. 82, E53–E62.
| 15471815PubMed |

Farin, P. W., Piedrahita, J. A., and Farin, C. E. (2006). Errors in development of fetuses and placentas from in vitro-produced bovine embryos. Theriogenology 65, 178–191.
Errors in development of fetuses and placentas from in vitro-produced bovine embryos.Crossref | GoogleScholarGoogle Scholar | 16266745PubMed |

Feil, R., and Fraga, M. F. (2011). Epigenetics and the environment: emerging patterns and implications. Nat. Rev. Genet. 13, 97–109.

Fernández-Gonzalez, R., Moreira, P., Bilbao, A., Jiménez, A., Pérez-Crespo, M., Ramírez, M. A., Rodríguez De Fonseca, F., Pintado, B., and Gutiérrez-Adán, A. (2004). Long-term effect of in vitro culture of mouse embryos with serum on mRNA expression of imprinting genes, development, and behavior. Proc. Natl Acad. Sci. USA 101, 5880–5885.
Long-term effect of in vitro culture of mouse embryos with serum on mRNA expression of imprinting genes, development, and behavior.Crossref | GoogleScholarGoogle Scholar | 15079084PubMed |

Figueroa, M. E., Lugthart, S., Li, Y., Erpelinck-Verschueren, C., Deng, X., Christos, P. J., Schifano, E., Booth, J., van Putten, W., Skrabanek, L., Campagne, F., Mazumdar, M., Greally, J. M., Valk, P. J., Lowenberg, B., Delwel, R., and Melnick, A. (2010). DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia. Cancer Cell 17, 13–27.
DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXkvFKgs70%3D&md5=7ae468ebe7908ce8290d44b0a9460fbcCAS | 20060365PubMed |

Gao, Y., Hyttel, P., and Hall, V. J. (2010). Regulation of H3K27me3 and H3K4me3 during early porcine embryonic development. Mol. Reprod. Dev. 77, 540–549.
Regulation of H3K27me3 and H3K4me3 during early porcine embryonic development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXlt1agt7k%3D&md5=749987cda4f35f9c31fea29cf59bdfb3CAS | 20422712PubMed |

Gardner, D. K. (1994). Mammalian embryo culture in the absence of serum or somatic cell support. Cell Biol. Int. 18, 1163–1180.
Mammalian embryo culture in the absence of serum or somatic cell support.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DyaK2M3hvVOntQ%3D%3D&md5=533d34515277b8c437629b3b198bf05fCAS | 7703956PubMed |

Gaspar, R. C., Arnold, D. R., Correa, C. A., da Rocha, C. V., Penteado, J. C., Del Collado, M., Vantini, R., Garcia, J. M., and Lopes, F. L. (2015). Oxygen tension affects histone remodeling of in vitro-produced embryos in a bovine model. Theriogenology 83, 1408–1415.
Oxygen tension affects histone remodeling of in vitro-produced embryos in a bovine model.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtFKru7k%3D&md5=d604ed7109c0eacf8e01418c21d70e4fCAS | 25777077PubMed |

Gkountela, S., and Clark, A. T. (2014). A big surprise in the little zygote: the curious business of losing methylated cytosines. Cell Stem Cell 15, 393–394.
A big surprise in the little zygote: the curious business of losing methylated cytosines.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhsF2murfF&md5=2389357deafab9b12ebc94a68170ceb5CAS | 25280211PubMed |

Gkountela, S., Zhang, K. X., Shafiq, T. A., Liao, W. W., Hargan-Calvopina, J., Chen, P. Y., and Clark, A. T. (2015). DNA demethylation dynamics in the human prenatal germline. Cell 161, 1425–1436.
DNA demethylation dynamics in the human prenatal germline.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXovVCnurc%3D&md5=f65b8d7ac53927adecc951284bb5bbe0CAS | 26004067PubMed |

Gluckman, P. D., and Hanson, M. A. (2004). Developmental origins of disease paradigm: a mechanistic and evolutionary perspective. Pediatr. Res. 56, 311–317.
Developmental origins of disease paradigm: a mechanistic and evolutionary perspective.Crossref | GoogleScholarGoogle Scholar | 15240866PubMed |

Goldberg, A. D., Banaszynski, L. A., Noh, K. M., Lewis, P. W., Elsaesser, S. J., Stadler, S., Dewell, S., Law, M., Guo, X., Li, X., Wen, D., Chapgier, A., DeKelver, R. C., Miller, J. C., Lee, Y. L., Boydston, E. A., Holmes, M. C., Gregory, P. D., Greally, J. M., Rafii, S., Yang, C., Scambler, P. J., Garrick, D., Gibbons, R. J., Higgs, D. R., Cristea, I. M., Urnov, F. D., Zheng, D., and Allis, C. D. (2010). Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691.
Distinct factors control histone variant H3.3 localization at specific genomic regions.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXlsVSgtrs%3D&md5=124ec6827f0fc9336bf61000b2225992CAS | 20211137PubMed |

Gopalakrishnan, S., Sullivan, B. A., Trazzi, S., Della Valle, G., and Robertson, K. D. (2009). DNMT3B interacts with constitutive centromere protein CENP-C to modulate DNA methylation and the histone code at centromeric regions. Hum. Mol. Genet. 18, 3178–3193.
DNMT3B interacts with constitutive centromere protein CENP-C to modulate DNA methylation and the histone code at centromeric regions.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXps1ajtro%3D&md5=315f540cebb1db1b48ff919e6567fbfdCAS | 19482874PubMed |

Gu, T. P., Guo, F., Yang, H., Wu, H. P., Xu, G. F., Liu, W., Xie, Z. G., Shi, L., He, X., Jin, S. G., Iqbal, K., Shi, Y. G., Deng, Z., Szabo, P. E., Pfeifer, G. P., Li, J., and Xu, G. L. (2011). The role of Tet3 DNA dioxygenase in epigenetic reprogramming by oocytes. Nature 477, 606–610.
The role of Tet3 DNA dioxygenase in epigenetic reprogramming by oocytes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXhtFersL7M&md5=700c4d2113e0c792f7ca54cf4944535bCAS | 21892189PubMed |

Guo, F., Li, X., Liang, D., Li, T., Zhu, P., Guo, H., Wu, X., Wen, L., Gu, T. P., Hu, B., Walsh, C. P., Li, J., Tang, F., and Xu, G. L. (2014a). Active and passive demethylation of male and female pronuclear DNA in the mammalian zygote. Cell Stem Cell 15, 447–458.
Active and passive demethylation of male and female pronuclear DNA in the mammalian zygote.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhs1ymurvJ&md5=ce4f67b08a12dee9440811e4edf50ff6CAS | 25220291PubMed |

Guo, H., Zhu, P., Yan, L., Li, R., Hu, B., Lian, Y., Yan, J., Ren, X., Lin, S., Li, J., Jin, X., Shi, X., Liu, P., Wang, X., Wang, W., Wei, Y., Li, X., Guo, F., Wu, X., Fan, X., Yong, J., Wen, L., Xie, S. X., Tang, F., and Qiao, J. (2014b). The DNA methylation landscape of human early embryos. Nature 511, 606–610.
The DNA methylation landscape of human early embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXht1ChurbK&md5=1257f3afa4dd277251651fd463d11e06CAS | 25079557PubMed |

Guo, F., Yan, L., Guo, H., Li, L., Hu, B., Zhao, Y., Yong, J., Hu, Y., Wang, X., Wei, Y., Wang, W., Li, R., Yan, J., Zhi, X., Zhang, Y., Jin, H., Zhang, W., Hou, Y., Zhu, P., Li, J., Zhang, L., Liu, S., Ren, Y., Zhu, X., Wen, L., Gao, Y. Q., Tang, F., and Qiao, J. (2015). The transcriptome and DNA methylome landscapes of human primordial germ cells. Cell 161, 1437–1452.
The transcriptome and DNA methylome landscapes of human primordial germ cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtVejurfO&md5=0185549962a70bffbe83bde475d9db2aCAS | 26046443PubMed |

Hammoud, S. S., Nix, D. A., Zhang, H., Purwar, J., Carrell, D. T., and Cairns, B. R. (2009). Distinctive chromatin in human sperm packages genes for embryo development. Nature 460, 473–478.
| 1:CAS:528:DC%2BD1MXnt1Sht7w%3D&md5=c2ccc7d918fa5f510adb39b7504a311bCAS | 19525931PubMed |

Hasler, J. F. (2000). In-vitro production of cattle embryos: problems with pregnancies and parturition. Hum. Reprod. 15, 47–58.
In-vitro production of cattle embryos: problems with pregnancies and parturition.Crossref | GoogleScholarGoogle Scholar | 11263537PubMed |

Hasler, J. F., Henderson, W. B., Hurtgen, P. J., Jin, Z. Q., McCauley, A. D., Mower, S. A., Neely, B., Shuey, L. S., Stokes, J. E., and Trimmer, S. A. (1995). Production, freezing and transfer of bovine IVF embryos and subsequent calving results. Theriogenology 43, 141–152.
Production, freezing and transfer of bovine IVF embryos and subsequent calving results.Crossref | GoogleScholarGoogle Scholar |

He, Y. F., Li, B. Z., Li, Z., Liu, P., Wang, Y., Tang, Q., Ding, J., Jia, Y., Chen, Z., Li, L., Sun, Y., Li, X., Dai, Q., Song, C. X., Zhang, K., He, C., and Xu, G. L. (2011). Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 333, 1303–1307.
Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXhtV2jt7nO&md5=35ea2582d21dbb5911ac187d8012083fCAS | 21817016PubMed |

Heard, E., and Martienssen, R. A. (2014). Transgenerational epigenetic inheritance: myths and mechanisms. Cell 157, 95–109.
Transgenerational epigenetic inheritance: myths and mechanisms.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXmtVCnurs%3D&md5=bfb2e7c7775ebae15ba5316cc600f602CAS | 24679529PubMed |

Heijmans, B. T., Tobi, E. W., Stein, A. D., Putter, H., Blauw, G. J., Susser, E. S., Slagboom, P. E., and Lumey, L. H. (2008). Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proc. Natl Acad. Sci. USA 105, 17 046–17 049.
Persistent epigenetic differences associated with prenatal exposure to famine in humans.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXhtlKqt7%2FP&md5=8421cd8de311a098eb9ce0ee365340d6CAS |

Hellman, A., and Chess, A. (2007). Gene body-specific methylation on the active X chromosome. Science 315, 1141–1143.
Gene body-specific methylation on the active X chromosome.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXhvFWrs7o%3D&md5=f8591ab5bc011edfeb6b3850415468fdCAS | 17322062PubMed |

Himes, K. P., Koppes, E., and Chaillet, J. R. (2013). Generalized disruption of inherited genomic imprints leads to wide-ranging placental defects and dysregulated fetal growth. Dev. Biol. 373, 72–82.
Generalized disruption of inherited genomic imprints leads to wide-ranging placental defects and dysregulated fetal growth.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xhs1agsbvL&md5=be02d43fe39dc06c6360b18e60e2a231CAS | 23085235PubMed |

Hong, S. H., Rampalli, S., Lee, J. B., McNicol, J., Collins, T., Draper, J. S., and Bhatia, M. (2011). Cell fate potential of human pluripotent stem cells is encoded by histone modifications. Cell Stem Cell 9, 24–36.
Cell fate potential of human pluripotent stem cells is encoded by histone modifications.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXosFagu7g%3D&md5=09e231cc5e2940dea5bc488384bd85a1CAS | 21726831PubMed |

Hoshi, H. (2003). In vitro production of bovine embryos and their application for embryo transfer. Theriogenology 59, 675–685.
In vitro production of bovine embryos and their application for embryo transfer.Crossref | GoogleScholarGoogle Scholar | 12499011PubMed |

Howell, C. Y., Bestor, T. H., Ding, F., Latham, K. E., Mertineit, C., Trasler, J. M., and Chaillet, J. R. (2001). Genomic imprinting disrupted by a maternal effect mutation in the Dnmt1 gene. Cell 104, 829–838.
Genomic imprinting disrupted by a maternal effect mutation in the Dnmt1 gene.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXisVyisrc%3D&md5=26a97e66f4a491f82bd994a38cbbdcd8CAS | 11290321PubMed |

Huang, J., Zhang, H., Wang, X., Dobbs, K. B., Yao, J., Qin, G., Whitworth, K., Walters, E. M., Prather, R. S., and Zhao, J. (2015). Impairment of preimplantation porcine embryo development by histone demethylase KDM5B knockdown through disturbance of bivalent H3K4me3–H3K27me3 modifications. Biol. Reprod. 92, 72.
Impairment of preimplantation porcine embryo development by histone demethylase KDM5B knockdown through disturbance of bivalent H3K4me3–H3K27me3 modifications.Crossref | GoogleScholarGoogle Scholar | 25609834PubMed |

Inoue, A., and Zhang, Y. (2011). Replication-dependent loss of 5-hydroxymethylcytosine in mouse preimplantation embryos. Science 334, 194.
Replication-dependent loss of 5-hydroxymethylcytosine in mouse preimplantation embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXht1yksbbN&md5=764ac31646cbccf6467b7612b1d7e74aCAS | 21940858PubMed |

Inoue, A., and Zhang, Y. (2014). Nucleosome assembly is required for nuclear pore complex assembly in mouse zygotes. Nat. Struct. Mol. Biol. 21, 609–616.
Nucleosome assembly is required for nuclear pore complex assembly in mouse zygotes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXpsVamsbw%3D&md5=aa872bf602c43af32332577cbade90afCAS | 24908396PubMed |

Iqbal, K., Jin, S. G., Pfeifer, G. P., and Szabo, P. E. (2011). Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc. Natl Acad. Sci. USA 108, 3642–3647.
Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXivFKqs7Y%3D&md5=ce4a13e5f1603d78078dac2a0b36fc4fCAS | 21321204PubMed |

Jacobsen, H., Schmidt, M., Holm, P., Sangild, P. T., Greve, T., and Callesen, H. (2000). Ease of calving, blood chemistry, insulin and bovine growth hormone of newborn calves derived from embryos produced in vitro in culture systems with serum and co-culture or with PVA. Theriogenology 54, 147–158.
Ease of calving, blood chemistry, insulin and bovine growth hormone of newborn calves derived from embryos produced in vitro in culture systems with serum and co-culture or with PVA.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXmvVaisrk%3D&md5=5bd90e086a84dc54e4dde1358b8032b6CAS | 10990356PubMed |

Jaenisch, R., and Bird, A. (2003). Epigenetic regulation of gene expression: how the genome integrates intrinsic and environmental signals. Nat. Genet. 33, 245–254.
Epigenetic regulation of gene expression: how the genome integrates intrinsic and environmental signals.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXhsV2kt7s%3D&md5=851249e65bc9a919653927cfc5718a87CAS | 12610534PubMed |

Jahangiri, M., Shahhoseini, M., and Movaghar, B. (2014). H19 and MEST gene expression and histone modification in blastocysts cultured from vitrified and fresh two-cell mouse embryos. Reprod. Biomed. Online 29, 559–566.
H19 and MEST gene expression and histone modification in blastocysts cultured from vitrified and fresh two-cell mouse embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhs1Squ7bF&md5=9795abdd37267dcff9be1b7cdb6f561bCAS | 25257147PubMed |

Jeffries, M. A., Dozmorov, M., Tang, Y., Merrill, J. T., Wren, J. D., and Sawalha, A. H. (2011). Genome-wide DNA methylation patterns in CD4+ T cells from patients with systemic lupus erythematosus. Epigenetics 6, 593–601.
Genome-wide DNA methylation patterns in CD4+ T cells from patients with systemic lupus erythematosus.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XitF2ku7o%3D&md5=30645aca0c7fc1bfa3850e0222909159CAS | 21436623PubMed |

Jenuwein, T., Laible, G., Dorn, R., and Reuter, G. (1998). SET domain proteins modulate chromatin domains in eu- and heterochromatin. Cell. Mol. Life Sci. 54, 80–93.
SET domain proteins modulate chromatin domains in eu- and heterochromatin.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1cXns1CltA%3D%3D&md5=a12250dac488c0a66eed5134a98fe80bCAS | 9487389PubMed |

Jjingo, D., Conley, A. B., Yi, S. V., Lunyak, V. V., and Jordan, I. K. (2012). On the presence and role of human gene-body DNA methylation. Oncotarget 3, 462–474.
| 22577155PubMed |

Kang, J., Lienhard, M., Pastor, W. A., Chawla, A., Novotny, M., Tsagaratou, A., Lasken, R. S., Thompson, E. C., Surani, M. A., Koralov, S. B., Kalantry, S., Chavez, L., and Rao, A. (2015). Simultaneous deletion of the methylcytosine oxidases Tet1 and Tet3 increases transcriptome variability in early embryogenesis. Proc. Natl Acad. Sci. USA 112, E4236–E4245.
Simultaneous deletion of the methylcytosine oxidases Tet1 and Tet3 increases transcriptome variability in early embryogenesis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtF2ku7rP&md5=63d0f3bf9f9dd84f1136467fec6d1803CAS | 26199412PubMed |

Katari, S., Turan, N., Bibikova, M., Erinle, O., Chalian, R., Foster, M., Gaughan, J. P., Coutifaris, C., and Sapienza, C. (2009). DNA methylation and gene expression differences in children conceived in vitro or in vivo. Hum. Mol. Genet. 18, 3769–3778.
DNA methylation and gene expression differences in children conceived in vitro or in vivo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhtFyhurfE&md5=96429da6897fb9c7399f94bd19fb75cfCAS | 19605411PubMed |

Khosla, S., Dean, W., Brown, D., Reik, W., and Feil, R. (2001). Culture of preimplantation mouse embryos affects fetal development and the expression of imprinted genes. Biol. Reprod. 64, 918–926.
Culture of preimplantation mouse embryos affects fetal development and the expression of imprinted genes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXhsVKjtrc%3D&md5=7ddf0f1cd89d4b8356ee25312c6d10c5CAS | 11207209PubMed |

Kleijkers, S. H., van Montfoort, A. P., Smits, L. J., Viechtbauer, W., Roseboom, T. J., Nelissen, E. C., Coonen, E., Derhaag, J. G., Bastings, L., Schreurs, I. E., Evers, J. L., and Dumoulin, J. C. (2014). IVF culture medium affects post-natal weight in humans during the first 2 years of life. Hum. Reprod. 29, 661–669.
IVF culture medium affects post-natal weight in humans during the first 2 years of life.Crossref | GoogleScholarGoogle Scholar | 24549211PubMed |

Klose, R. J., Kallin, E. M., and Zhang, Y. (2006a). JmjC-domain-containing proteins and histone demethylation. Nat. Rev. Genet. 7, 715–727.
JmjC-domain-containing proteins and histone demethylation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xot1Ggu74%3D&md5=a021ed800014f5b834ebeeaa00ea330cCAS | 16983801PubMed |

Klose, R. J., Yamane, K., Bae, Y., Zhang, D., Erdjument-Bromage, H., Tempst, P., Wong, J., and Zhang, Y. (2006b). The transcriptional repressor JHDM3A demethylates trimethyl histone H3 lysine 9 and lysine 36. Nature 442, 312–316.
The transcriptional repressor JHDM3A demethylates trimethyl histone H3 lysine 9 and lysine 36.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XmvF2jsr0%3D&md5=c54d00dc744c2da06fd9c8081bc81cdcCAS | 16732292PubMed |

Kobayashi, H., Sakurai, T., Imai, M., Takahashi, N., Fukuda, A., Yayoi, O., Sato, S., Nakabayashi, K., Hata, K., Sotomaru, Y., Suzuki, Y., and Kono, T. (2012). Contribution of intragenic DNA methylation in mouse gametic DNA methylomes to establish oocyte-specific heritable marks. PLoS Genet. 8, e1002440.
Contribution of intragenic DNA methylation in mouse gametic DNA methylomes to establish oocyte-specific heritable marks.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhtVyitrg%3D&md5=45ad88cfba4aa97bcc2c6f8f4b391360CAS | 22242016PubMed |

Kohli, R. M., and Zhang, Y. (2013). TET enzymes, TDG and the dynamics of DNA demethylation. Nature 502, 472–479.
TET enzymes, TDG and the dynamics of DNA demethylation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhs1KrurzF&md5=44b39ee626db7246d17afae5d8f1d4baCAS | 24153300PubMed |

Kono, T., Sotomaru, Y., Sato, Y., and Nakahara, T. (1993). Development of androgenetic mouse embryos produced by in vitro fertilization of enucleated oocytes. Mol. Reprod. Dev. 34, 43–46.
Development of androgenetic mouse embryos produced by in vitro fertilization of enucleated oocytes.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DyaK3s7hvFKlsA%3D%3D&md5=3e7b701ec039bc4ba87fb8ab5af62b53CAS | 8418815PubMed |

Kornberg, R. D. (1974). Chromatin structure: a repeating unit of histones and DNA. Science 184, 868–871.
Chromatin structure: a repeating unit of histones and DNA.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaE2cXksVKntLY%3D&md5=27cd51efb2ad68786ab743a4db7a165eCAS | 4825889PubMed |

Kruip, T. A. M., and den Daas, J. H. G. (1997). In vitro produced and cloned embryos: effects on pregnancy, parturition and offspring. Theriogenology 47, 43–52.
In vitro produced and cloned embryos: effects on pregnancy, parturition and offspring.Crossref | GoogleScholarGoogle Scholar |

Kruip, T. A., Bevers, M. M., and Kemp, B. (2000). Environment of oocyte and embryo determines health of IVP offspring. Theriogenology 53, 611–618.
Environment of oocyte and embryo determines health of IVP offspring.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BD3c7pvFahug%3D%3D&md5=cb80f0d44acf3dfb91299073cd72225bCAS | 10735053PubMed |

Lee, K. K., and Workman, J. L. (2007). Histone acetyltransferase complexes: one size doesn’t fit all. Nat. Rev. Mol. Cell Biol. 8, 284–295.
Histone acetyltransferase complexes: one size doesn’t fit all.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXjtlygsrw%3D&md5=a1b3bd144c47940d521e4eac1003a18bCAS | 17380162PubMed |

Lei, H., Oh, S. P., Okano, M., Juttermann, R., Goss, K. A., Jaenisch, R., and Li, E. (1996). De novo DNA cytosine methyltransferase activities in mouse embryonic stem cells. Development 122, 3195–3205.
| 1:CAS:528:DyaK28Xms1Ogsb0%3D&md5=fde85f59a7a0a2dc304618ba1df0d42eCAS | 8898232PubMed |

Lewis, P. W., Muller, M. M., Koletsky, M. S., Cordero, F., Lin, S., Banaszynski, L. A., Garcia, B. A., Muir, T. W., Becher, O. J., and Allis, C. D. (2013). Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 340, 857–861.
Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXnsFSms7k%3D&md5=081f0ab50e411d10843a782ec0001ac0CAS | 23539183PubMed |

Li, E. (2002). Chromatin modification and epigenetic reprogramming in mammalian development. Nat. Rev. Genet. 3, 662–673.
Chromatin modification and epigenetic reprogramming in mammalian development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38Xmslagt7w%3D&md5=5b08812d334e242d683b623eb72ef246CAS | 12209141PubMed |

Li, E., Bestor, T. H., and Jaenisch, R. (1992). Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 69, 915–926.
Targeted mutation of the DNA methyltransferase gene results in embryonic lethality.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK38XksVGgsr0%3D&md5=2b718f8d499364d040a2b98cc8afc61eCAS | 1606615PubMed |

Li, Z., Cai, X., Cai, C. L., Wang, J., Zhang, W., Petersen, B. E., Yang, F. C., and Xu, M. (2011). Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood 118, 4509–4518.
Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXhsVGqu73F&md5=c4bad8403a24848c7d178976aac3f712CAS | 21803851PubMed |

Lin, C. J., Conti, M., and Ramalho-Santos, M. (2013). Histone variant H3.3 maintains a decondensed chromatin state essential for mouse preimplantation development. Development 140, 3624–3634.
Histone variant H3.3 maintains a decondensed chromatin state essential for mouse preimplantation development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhsFOktL%2FL&md5=05f684855eba7bd9bb9c9eb6b0a8fbecCAS | 23903189PubMed |

Lin, C. J., Koh, F. M., Wong, P., Conti, M., and Ramalho-Santos, M. (2014). Hira-mediated H3.3 incorporation is required for DNA replication and ribosomal RNA transcription in the mouse zygote. Dev. Cell 30, 268–279.
Hira-mediated H3.3 incorporation is required for DNA replication and ribosomal RNA transcription in the mouse zygote.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXht1Oqs7fO&md5=c5267a51f2100db776624f8be2873408CAS | 25087892PubMed |

Liu, H., Galka, M., Iberg, A., Wang, Z., Li, L., Voss, C., Jiang, X., Lajoie, G., Huang, Z., Bedford, M. T., and Li, S. S. (2010). Systematic identification of methyllysine-driven interactions for histone and nonhistone targets. J. Proteome Res. 9, 5827–5836.
Systematic identification of methyllysine-driven interactions for histone and nonhistone targets.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXht1enurjE&md5=b5c36dd4390127db1aed8e0f281c247dCAS | 20836566PubMed |

Luger, K., and Richmond, T. J. (1998). The histone tails of the nucleosome. Curr. Opin. Genet. Dev. 8, 140–146.
The histone tails of the nucleosome.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1cXjtlWju74%3D&md5=4fde29e1a8b640efed1eb0b9f17156c5CAS | 9610403PubMed |

Lyall, K., Pauls, D. L., Spiegelman, D., Santangelo, S. L., and Ascherio, A. (2012). Fertility therapies, infertility and autism spectrum disorders in the Nurses’ Health Study II. Paediatr. Perinat. Epidemiol. 26, 361–372.
Fertility therapies, infertility and autism spectrum disorders in the Nurses’ Health Study II.Crossref | GoogleScholarGoogle Scholar | 22686388PubMed |

Ma, P., and Schultz, R. M. (2008). Histone deacetylase 1 (HDAC1) regulates histone acetylation, development, and gene expression in preimplantation mouse embryos. Dev. Biol. 319, 110–120.
| 1:CAS:528:DC%2BD1cXnt1yjtrs%3D&md5=06dd589b247ecdcc9ec1e5e3b70f3befCAS | 18501342PubMed |

Ma, P., Pan, H., Montgomery, R. L., Olson, E. N., and Schultz, R. M. (2012). Compensatory functions of histone deacetylase 1 (HDAC1) and HDAC2 regulate transcription and apoptosis during mouse oocyte development. Proc. Natl Acad. Sci. USA 109, E481–E489.
Compensatory functions of histone deacetylase 1 (HDAC1) and HDAC2 regulate transcription and apoptosis during mouse oocyte development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XjsFyrtbw%3D&md5=1638c5a4162d511c7e6697267c712172CAS | 22223663PubMed |

Maalouf, W. E., Alberio, R., and Campbell, K. H. (2008). Differential acetylation of histone H4 lysine during development of in vitro fertilized, cloned and parthenogenetically activated bovine embryos. Epigenetics 3, 199–209.
Differential acetylation of histone H4 lysine during development of in vitro fertilized, cloned and parthenogenetically activated bovine embryos.Crossref | GoogleScholarGoogle Scholar | 18698155PubMed |

Mansouri-Attia, N., Sandra, O., Aubert, J., Degrelle, S., Everts, R. E., Giraud-Delville, C., Heyman, Y., Galio, L., Hue, I., Yang, X., Tian, X. C., Lewin, H. A., and Renard, J. P. (2009). Endometrium as an early sensor of in vitro embryo manipulation technologies. Proc. Natl Acad. Sci. USA 106, 5687–5692.
Endometrium as an early sensor of in vitro embryo manipulation technologies.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXkvFGhurY%3D&md5=049ad45de0d722a3086ea994b9d4a7b5CAS | 19297625PubMed |

Marks, H., Kalkan, T., Menafra, R., Denissov, S., Jones, K., Hofemeister, H., Nichols, J., Kranz, A., Stewart, A. F., Smith, A., and Stunnenberg, H. G. (2012). The transcriptional and epigenomic foundations of ground state pluripotency. Cell 149, 590–604.
The transcriptional and epigenomic foundations of ground state pluripotency.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xmt1Gqtbw%3D&md5=340c6bd6d0e5a08df9f137a4d7b5a801CAS | 22541430PubMed |

Mayer, W., Niveleau, A., Walter, J., Fundele, R., and Haaf, T. (2000). Demethylation of the zygotic paternal genome. Nature 403, 501–502.
Demethylation of the zygotic paternal genome.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXht1ShsbY%3D&md5=af4af7648baabcf3522f57bb9a45515eCAS | 10676950PubMed |

McEvoy, T. G., Sinclair, K. D., Young, L. E., Wilmut, I., and Robinson, J. J. (2000). Large offspring syndrome and other consequences of ruminant embryo culture in vitro: relevance to blastocyst culture in human ART. Hum. Fertil. (Camb.) 3, 238–246.
Large offspring syndrome and other consequences of ruminant embryo culture in vitro: relevance to blastocyst culture in human ART.Crossref | GoogleScholarGoogle Scholar | 11844385PubMed |

McGraw, S., Oakes, C. C., Martel, J., Cirio, M. C., de Zeeuw, P., Mak, W., Plass, C., Bartolomei, M. S., Chaillet, J. R., and Trasler, J. M. (2013). Loss of DNMT1o disrupts imprinted X chromosome inactivation and accentuates placental defects in females. PLoS Genet. 9, e1003873.
Loss of DNMT1o disrupts imprinted X chromosome inactivation and accentuates placental defects in females.Crossref | GoogleScholarGoogle Scholar | 24278026PubMed |

Messerschmidt, D. M. (2012). Should I stay or should I go: protection and maintenance of DNA methylation at imprinted genes. Epigenetics 7, 969–975.
Should I stay or should I go: protection and maintenance of DNA methylation at imprinted genes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhs1aru7s%3D&md5=f9b2b818c0af61b6ebc0f877c64dfb5aCAS | 22869105PubMed |

Messerschmidt, D. M., de Vries, W., Ito, M., Solter, D., Ferguson-Smith, A., and Knowles, B. B. (2012). Trim28 is required for epigenetic stability during mouse oocyte to embryo transition. Science 335, 1499–1502.
Trim28 is required for epigenetic stability during mouse oocyte to embryo transition.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XktFWksbk%3D&md5=bcef671bb37fcb3e7300d60f42b5e3b9CAS | 22442485PubMed |

Mikkelsen, T. S., Ku, M., Jaffe, D. B., Issac, B., Lieberman, E., Giannoukos, G., Alvarez, P., Brockman, W., Kim, T. K., Koche, R. P., Lee, W., Mendenhall, E., O’Donovan, A., Presser, A., Russ, C., Xie, X., Meissner, A., Wernig, M., Jaenisch, R., Nusbaum, C., Lander, E. S., and Bernstein, B. E. (2007). Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448, 553–560.
Genome-wide maps of chromatin state in pluripotent and lineage-committed cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXosVSrtrc%3D&md5=a4d778047390c4de0d495da648d020a0CAS | 17603471PubMed |

Moran-Crusio, K., Reavie, L., Shih, A., Abdel-Wahab, O., Ndiaye-Lobry, D., Lobry, C., Figueroa, M. E., Vasanthakumar, A., Patel, J., Zhao, X., Perna, F., Pandey, S., Madzo, J., Song, C., Dai, Q., He, C., Ibrahim, S., Beran, M., Zavadil, J., Nimer, S. D., Melnick, A., Godley, L. A., Aifantis, I., and Levine, R. L. (2011). Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation. Cancer Cell 20, 11–24.
Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXoslCms7w%3D&md5=4d9121d5696162b855d6f2d85d634d9aCAS | 21723200PubMed |

Nakamura, T., Liu, Y. J., Nakashima, H., Umehara, H., Inoue, K., Matoba, S., Tachibana, M., Ogura, A., Shinkai, Y., and Nakano, T. (2012). PGC7 binds histone H3K9me2 to protect against conversion of 5mC to 5hmC in early embryos. Nature 486, 415–419.
| 1:CAS:528:DC%2BC38XovFyrtro%3D&md5=0f82bff88005aa5c56ef71878f2e0c61CAS | 22722204PubMed |

Nashun, B., Yukawa, M., Liu, H., Akiyama, T., and Aoki, F. (2010). Changes in the nuclear deposition of histone H2A variants during pre-implantation development in mice. Development 137, 3785–3794.
Changes in the nuclear deposition of histone H2A variants during pre-implantation development in mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXhsF2js7bJ&md5=b41b2cdf8750e79930a0d805ecd31405CAS | 20943707PubMed |

Oda, H., Okamoto, I., Murphy, N., Chu, J., Price, S. M., Shen, M. M., Torres-Padilla, M. E., Heard, E., and Reinberg, D. (2009). Monomethylation of histone H4-lysine 20 is involved in chromosome structure and stability and is essential for mouse development. Mol. Cell. Biol. 29, 2278–2295.
Monomethylation of histone H4-lysine 20 is involved in chromosome structure and stability and is essential for mouse development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXks1SrsL8%3D&md5=6e34f2964ab607b90c76fb7160a8eff5CAS | 19223465PubMed |

Okae, H., Chiba, H., Hiura, H., Hamada, H., Sato, A., Utsunomiya, T., Kikuchi, H., Yoshida, H., Tanaka, A., Suyama, M., and Arima, T. (2014). Genome-wide analysis of DNA methylation dynamics during early human development. PLoS Genet. 10, e1004868.
Genome-wide analysis of DNA methylation dynamics during early human development.Crossref | GoogleScholarGoogle Scholar | 25501653PubMed |

Okano, M., Bell, D. W., Haber, D. A., and Li, E. (1999). DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99, 247–257.
DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXnt1Gqsrc%3D&md5=44602078c3a4166ddda06bc1c70feb98CAS | 10555141PubMed |

Ooga, M., Suzuki, M. G., and Aoki, F. (2013). Involvement of DOT1L in the remodeling of heterochromatin configuration during early preimplantation development in mice. Biol. Reprod. 89, 145.
Involvement of DOT1L in the remodeling of heterochromatin configuration during early preimplantation development in mice.Crossref | GoogleScholarGoogle Scholar | 24132959PubMed |

Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, A., Fundele, R., Dean, W., Reik, W., and Walter, J. (2000). Active demethylation of the paternal genome in the mouse zygote. Curr. Biol. 10, 475–478.
Active demethylation of the paternal genome in the mouse zygote.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXislyltLo%3D&md5=60f5d776667b8ee8deb1e065bfabb5e2CAS | 10801417PubMed |

Pandey, S., Shetty, A., Hamilton, M., Bhattacharya, S., and Maheshwari, A. (2012). Obstetric and perinatal outcomes in singleton pregnancies resulting from IVF/ICSI: a systematic review and meta-analysis. Hum. Reprod. Update 18, 485–503.
Obstetric and perinatal outcomes in singleton pregnancies resulting from IVF/ICSI: a systematic review and meta-analysis.Crossref | GoogleScholarGoogle Scholar | 22611174PubMed |

Pascual, M., Suzuki, M., Isidoro-Garcia, M., Padron, J., Turner, T., Lorente, F., Davila, I., and Greally, J. M. (2011). Epigenetic changes in B lymphocytes associated with house dust mite allergic asthma. Epigenetics 6, 1131–1137.
Epigenetic changes in B lymphocytes associated with house dust mite allergic asthma.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XitlGksrk%3D&md5=5023d55091312a5494864367d96edec9CAS | 21975512PubMed |

Payer, B., Saitou, M., Barton, S. C., Thresher, R., Dixon, J. P., Zahn, D., Colledge, W. H., Carlton, M. B., Nakano, T., and Surani, M. A. (2003). Stella is a maternal effect gene required for normal early development in mice. Curr. Biol. 13, 2110–2117.
Stella is a maternal effect gene required for normal early development in mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXps1Oqsb4%3D&md5=db94dd426566fde1d7e0e8b17f34a8adCAS | 14654002PubMed |

Peat, J. R., Dean, W., Clark, S. J., Krueger, F., Smallwood, S. A., Ficz, G., Kim, J. K., Marioni, J. C., Hore, T. A., and Reik, W. (2014). Genome-wide bisulfite sequencing in zygotes identifies demethylation targets and maps the contribution of TET3 oxidation. Cell Reports 9, 1990–2000.
Genome-wide bisulfite sequencing in zygotes identifies demethylation targets and maps the contribution of TET3 oxidation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXitFKnurbL&md5=86ebb2c94516f5aef619ccef5b7dada8CAS | 25497087PubMed |

Petridou, E. T., Sergentanis, T. N., Panagopoulou, P., Moschovi, M., Polychronopoulou, S., Baka, M., Pourtsidis, A., Athanassiadou, F., Kalmanti, M., Sidi, V., Dessypris, N., Frangakis, C., Matsoukis, I. L., Stefanadis, C., Skalkidou, A., Stephansson, O., Adami, H. O., and Kieler, H. (2012). In vitro fertilization and risk of childhood leukemia in Greece and Sweden. Pediatr. Blood Cancer 58, 930–936.
In vitro fertilization and risk of childhood leukemia in Greece and Sweden.Crossref | GoogleScholarGoogle Scholar | 21618418PubMed |

Qi, H. H., Sarkissian, M., Hu, G. Q., Wang, Z., Bhattacharjee, A., Gordon, D. B., Gonzales, M., Lan, F., Ongusaha, P. P., Huarte, M., Yaghi, N. K., Lim, H., Garcia, B. A., Brizuela, L., Zhao, K., Roberts, T. M., and Shi, Y. (2010). Histone H4K20/H3K9 demethylase PHF8 regulates zebrafish brain and craniofacial development. Nature 466, 503–507.
Histone H4K20/H3K9 demethylase PHF8 regulates zebrafish brain and craniofacial development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXos1ahsr8%3D&md5=22b1cb7cef1ff23ea2fe4f103033081eCAS | 20622853PubMed |

Reik, W., Dean, W., and Walter, J. (2001). Epigenetic reprogramming in mammalian development. Science 293, 1089–1093.
Epigenetic reprogramming in mammalian development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXmtVWltL8%3D&md5=a4d32ea9cf4c51f15d70243c38da206cCAS | 11498579PubMed |

Rimm, A. A., Katayama, A. C., Diaz, M., and Katayama, K. P. (2004). A meta-analysis of controlled studies comparing major malformation rates in IVF and ICSI infants with naturally conceived children. J. Assist. Reprod. Genet. 21, 437–443.
A meta-analysis of controlled studies comparing major malformation rates in IVF and ICSI infants with naturally conceived children.Crossref | GoogleScholarGoogle Scholar | 15704519PubMed |

Ross, P. J., Ragina, N. P., Rodriguez, R. M., Iager, A. E., Siripattarapravat, K., Lopez-Corrales, N., and Cibelli, J. B. (2008). Polycomb gene expression and histone H3 lysine 27 trimethylation changes during bovine preimplantation development. Reproduction 136, 777–785.
Polycomb gene expression and histone H3 lysine 27 trimethylation changes during bovine preimplantation development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXns1CrsQ%3D%3D&md5=df147260882b97e53cea27eb8a874325CAS | 18784248PubMed |

Sachs, M., Onodera, C., Blaschke, K., Ebata, K. T., Song, J. S., and Ramalho-Santos, M. (2013). Bivalent chromatin marks developmental regulatory genes in the mouse embryonic germline in vivo. Cell Reports 3, 1777–1784.
Bivalent chromatin marks developmental regulatory genes in the mouse embryonic germline in vivo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXos12qs78%3D&md5=465b93302771aca3e5e25ff8ddf1a335CAS | 23727241PubMed |

Sadic, D., Schmidt, K., Groh, S., Kondofersky, I., Ellwart, J., Fuchs, C., Theis, F. J., and Schotta, G. (2015). Atrx promotes heterochromatin formation at retrotransposons. EMBO Rep. 16, 836–850.
Atrx promotes heterochromatin formation at retrotransposons.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXoslCnsrY%3D&md5=b3b541013cccb45cf01e7ae15470676aCAS | 26012739PubMed |

Santenard, A., Ziegler-Birling, C., Koch, M., Tora, L., Bannister, A. J., and Torres-Padilla, M. E. (2010). Heterochromatin formation in the mouse embryo requires critical residues of the histone variant H3.3. Nat. Cell Biol. 12, 853–862.
Heterochromatin formation in the mouse embryo requires critical residues of the histone variant H3.3.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXhtFSju7%2FE&md5=f93dc1c2a5845d4b43389f2cb8dc2a78CAS | 20676102PubMed |

Santos, F., Hendrich, B., Reik, W., and Dean, W. (2002). Dynamic reprogramming of DNA methylation in the early mouse embryo. Dev. Biol. 241, 172–182.
Dynamic reprogramming of DNA methylation in the early mouse embryo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXptVWhsrg%3D&md5=ce9c51ef214d402134d8000ffbe72399CAS | 11784103PubMed |

Santos, F., Zakhartchenko, V., Stojkovic, M., Peters, A., Jenuwein, T., Wolf, E., Reik, W., and Dean, W. (2003). Epigenetic marking correlates with developmental potential in cloned bovine preimplantation embryos. Curr. Biol. 13, 1116–1121.
| 1:CAS:528:DC%2BD3sXlt1entrY%3D&md5=58133119c69fbabf4a58880d31016539CAS | 12842010PubMed |

Sarmento, O. F., Digilio, L. C., Wang, Y., Perlin, J., Herr, J. C., Allis, C. D., and Coonrod, S. A. (2004). Dynamic alterations of specific histone modifications during early murine development. J. Cell Sci. 117, 4449–4459.
Dynamic alterations of specific histone modifications during early murine development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXovVehtr0%3D&md5=cafcb079a34844fb805cee35af15ab72CAS | 15316069PubMed |

Sasaki, H., and Matsui, Y. (2008). Epigenetic events in mammalian germ-cell development: reprogramming and beyond. Nat. Rev. Genet. 9, 129–140.
Epigenetic events in mammalian germ-cell development: reprogramming and beyond.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXnt1eguw%3D%3D&md5=2c1b572b40403b806aa625773f6be89dCAS | 18197165PubMed |

Schmidt, M., Greve, T., Avery, B., Beckers, J. F., Sulon, J., and Hansen, H. B. (1996). Pregnancies, calves and calf viability after transfer of in vitro produced bovine embryos. Theriogenology 46, 527–539.
Pregnancies, calves and calf viability after transfer of in vitro produced bovine embryos.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BD2svhtlOguw%3D%3D&md5=e7e159de60034a7629a689d1acc1655eCAS | 16727920PubMed |

Schroeder, D. I., Jayashankar, K., Douglas, K. C., Thirkill, T. L., York, D., Dikinson, P. J., Williams, L. E., Samollow, P. B., Ross, P. J., Bannasch, D. L., Douglas, G. C., and LaSalle, J. M. (2015). Early developmental and evolutionary origins of gene body DNA methylation patterns in mammalian placentas. PLoS Genet. 11, e1005442.
Early developmental and evolutionary origins of gene body DNA methylation patterns in mammalian placentas.Crossref | GoogleScholarGoogle Scholar | 26241857PubMed |

Schultz, M. D., He, Y., Whitaker, J. W., Hariharan, M., Mukamel, E. A., Leung, D., Rajagopal, N., Nery, J. R., Urich, M. A., Chen, H., Lin, S., Lin, Y., Jung, I., Schmitt, A. D., Selvaraj, S., Ren, B., Sejnowski, T. J., Wang, W., and Ecker, J. R. (2015). Human body epigenome maps reveal noncanonical DNA methylation variation. Nature 523, 212–216.
Human body epigenome maps reveal noncanonical DNA methylation variation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtFaitrvL&md5=4f3c40b141d72f938e4164d6d92a35e4CAS | 26030523PubMed |

Sen, P., Dang, W., Donahue, G., Dai, J., Dorsey, J., Cao, X., Liu, W., Cao, K., Perry, R., Lee, J. Y., Wasko, B. M., Carr, D. T., He, C., Robison, B., Wagner, J., Gregory, B. D., Kaeberlein, M., Kennedy, B. K., Boeke, J. D., and Berger, S. L. (2015). H3K36 methylation promotes longevity by enhancing transcriptional fidelity. Genes Dev. 29, 1362–1376.
H3K36 methylation promotes longevity by enhancing transcriptional fidelity.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXht12qurfK&md5=e0a5f255167a5d47963ae78c450cde3cCAS | 26159996PubMed |

Sharif, J., Muto, M., Takebayashi, S., Suetake, I., Iwamatsu, A., Endo, T. A., Shinga, J., Mizutani-Koseki, Y., Toyoda, T., Okamura, K., Tajima, S., Mitsuya, K., Okano, M., and Koseki, H. (2007). The SRA protein Np95 mediates epigenetic inheritance by recruiting Dnmt1 to methylated DNA. Nature 450, 908–912.
The SRA protein Np95 mediates epigenetic inheritance by recruiting Dnmt1 to methylated DNA.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXhtl2gsb3L&md5=e6b132b672fe3f62879f17c2c7dc3842CAS | 17994007PubMed |

Shen, L., Inoue, A., He, J., Liu, Y., Lu, F., and Zhang, Y. (2014). Tet3 and DNA replication mediate demethylation of both the maternal and paternal genomes in mouse zygotes. Cell Stem Cell 15, 459–470.
Tet3 and DNA replication mediate demethylation of both the maternal and paternal genomes in mouse zygotes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhs1Gnu7jE&md5=db7e8d7a236cce888c11968d3345a96aCAS | 25280220PubMed |

Shen, J., Jiang, D., Fu, Y., Wu, X., Guo, H., Feng, B., Pang, Y., Streets, A. M., Tang, F., and Huang, Y. (2015). H3K4me3 epigenomic landscape derived from ChIP-Seq of 1000 mouse early embryonic cells. Cell Res. 25, 143–147.
H3K4me3 epigenomic landscape derived from ChIP-Seq of 1000 mouse early embryonic cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhsV2ntr7J&md5=6476cd0cbe0370868190f5ee6ad1e892CAS | 25178839PubMed |

Shinagawa, T., Takagi, T., Tsukamoto, D., Tomaru, C., Huynh, L. M., Sivaraman, P., Kumarevel, T., Inoue, K., Nakato, R., Katou, Y., Sado, T., Takahashi, S., Ogura, A., Shirahige, K., and Ishii, S. (2014). Histone variants enriched in oocytes enhance reprogramming to induced pluripotent stem cells. Cell Stem Cell 14, 217–227.
Histone variants enriched in oocytes enhance reprogramming to induced pluripotent stem cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXitV2ks78%3D&md5=f28ab038071816fc95e45aa5e0fd896fCAS | 24506885PubMed |

Shirane, K., Toh, H., Kobayashi, H., Miura, F., Chiba, H., Ito, T., Kono, T., and Sasaki, H. (2013). Mouse oocyte methylomes at base resolution reveal genome-wide accumulation of non-CpG methylation and role of DNA methyltransferases. PLoS Genet. 9, e1003439.
Mouse oocyte methylomes at base resolution reveal genome-wide accumulation of non-CpG methylation and role of DNA methyltransferases.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXnt1ejurw%3D&md5=6223c3a95289d4d6ae0882d1803df594CAS | 23637617PubMed |

Sinclair, K. D., Young, L. E., Wilmut, I., and McEvoy, T. G. (2000). In-utero overgrowth in ruminants following embryo culture: lessons from mice and a warning to men. Hum. Reprod. 15, 68–86.
In-utero overgrowth in ruminants following embryo culture: lessons from mice and a warning to men.Crossref | GoogleScholarGoogle Scholar | 11263539PubMed |

Smallwood, S. A., Lee, H. J., Angermueller, C., Krueger, F., Saadeh, H., Peat, J., Andrews, S. R., Stegle, O., Reik, W., and Kelsey, G. (2014). Single-cell genome-wide bisulfite sequencing for assessing epigenetic heterogeneity. Nat. Methods 11, 817–820.
Single-cell genome-wide bisulfite sequencing for assessing epigenetic heterogeneity.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhslelsLvN&md5=ecdb781d67806af52d7f92371b2a5c07CAS | 25042786PubMed |

Smith, Z. D., Chan, M. M., Humm, K. C., Karnik, R., Mekhoubad, S., Regev, A., Eggan, K., and Meissner, A. (2014). DNA methylation dynamics of the human preimplantation embryo. Nature 511, 611–615.
DNA methylation dynamics of the human preimplantation embryo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXht1ChurfJ&md5=2eeaa2d7b375c64a332d376e37889f84CAS | 25079558PubMed |

Stitzel, M. L., Sethupathy, P., Pearson, D. S., Chines, P. S., Song, L., Erdos, M. R., Welch, R., Parker, S. C., Boyle, A. P., Scott, L. J., Margulies, E. H., Boehnke, M., Furey, T. S., Crawford, G. E., and Collins, F. S. (2010). Global epigenomic analysis of primary human pancreatic islets provides insights into type 2 diabetes susceptibility loci. Cell Metab. 12, 443–455.
Global epigenomic analysis of primary human pancreatic islets provides insights into type 2 diabetes susceptibility loci.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXhtlKhtbzL&md5=b23e654092c6e315618fdbd5b420206eCAS | 21035756PubMed |

Suo, L., Meng, Q., Pei, Y., Fu, X., Wang, Y., Bunch, T. D., and Zhu, S. (2010). Effect of cryopreservation on acetylation patterns of lysine 12 of histone H4 (acH4K12) in mouse oocytes and zygotes. J. Assist. Reprod. Genet. 27, 735–741.
Effect of cryopreservation on acetylation patterns of lysine 12 of histone H4 (acH4K12) in mouse oocytes and zygotes.Crossref | GoogleScholarGoogle Scholar | 20838874PubMed |

Szenker, E., Lacoste, N., and Almouzni, G. (2012). A developmental requirement for HIRA-dependent H3.3 deposition revealed at gastrulation in Xenopus. Cell Reports 1, 730–740.
A developmental requirement for HIRA-dependent H3.3 deposition revealed at gastrulation in Xenopus.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhtVOrsrnP&md5=18c13cf8e3ed3099f61d961cc0094636CAS | 22813747PubMed |

Tang, M. C., Jacobs, S. A., Mattiske, D. M., Soh, Y. M., Graham, A. N., Tran, A., Lim, S. L., Hudson, D. F., Kalitsis, P., O’Bryan, M. K., Wong, L. H., and Mann, J. R. (2015a). Contribution of the two genes encoding histone variant h3.3 to viability and fertility in mice. PLoS Genet. 11, e1004964.
Contribution of the two genes encoding histone variant h3.3 to viability and fertility in mice.Crossref | GoogleScholarGoogle Scholar | 25675407PubMed |

Tang, W. W., Dietmann, S., Irie, N., Leitch, H. G., Floros, V. I., Bradshaw, C. R., Hackett, J. A., Chinnery, P. F., and Surani, M. A. (2015b). A unique gene regulatory network resets the human germline epigenome for development. Cell 161, 1453–1467.
A unique gene regulatory network resets the human germline epigenome for development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtVejurfM&md5=b07178362e00ee11226f8d287cf34bb4CAS | 26046444PubMed |

Thélie, A., Papillier, P., Pennetier, S., Perreau, C., Traverso, J. M., Uzbekova, S., Mermillod, P., Joly, C., Humblot, P., and Dalbiès-Tran, R. (2007). Differential regulation of abundance and deadenylation of maternal transcripts during bovine oocyte maturation in vitro and in vivo. BMC Dev. Biol. 7, 125.
Differential regulation of abundance and deadenylation of maternal transcripts during bovine oocyte maturation in vitro and in vivo.Crossref | GoogleScholarGoogle Scholar | 17988387PubMed |

Thompson, J. G., and Peterson, A. J. (2000). Bovine embryo culture in vitro: new developments and post-transfer consequences. Hum. Reprod. 15, 59–67.
Bovine embryo culture in vitro: new developments and post-transfer consequences.Crossref | GoogleScholarGoogle Scholar | 11263538PubMed |

Toyota, M., Ahuja, N., Ohe-Toyota, M., Herman, J. G., Baylin, S. B., and Issa, J. P. (1999). CpG island methylator phenotype in colorectal cancer. Proc. Natl Acad. Sci. USA 96, 8681–8686.
CpG island methylator phenotype in colorectal cancer.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXkslOktLg%3D&md5=ef0d9eba6adb103ed251eabe7c90ab21CAS | 10411935PubMed |

van de Werken, C., van der Heijden, G. W., Eleveld, C., Teeuwssen, M., Albert, M., Baarends, W. M., Laven, J. S., Peters, A. H., and Baart, E. B. (2014). Paternal heterochromatin formation in human embryos is H3K9/HP1 directed and primed by sperm-derived histone modifications. Nat. Commun. 5, 5868.
Paternal heterochromatin formation in human embryos is H3K9/HP1 directed and primed by sperm-derived histone modifications.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXjvFWqu7s%3D&md5=2f7daad1701bfc9ef8189a8225719557CAS | 25519718PubMed |

van der Heijden, G. W., Dieker, J. W., Derijck, A. A., Muller, S., Berden, J. H., Braat, D. D., van der Vlag, J., and de Boer, P. (2005). Asymmetry in histone H3 variants and lysine methylation between paternal and maternal chromatin of the early mouse zygote. Mech. Dev. 122, 1008–1022.
Asymmetry in histone H3 variants and lysine methylation between paternal and maternal chromatin of the early mouse zygote.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXntVGjtr4%3D&md5=bcb2b07769dfc4d558f699d83a5eba40CAS | 15922569PubMed |

van Wagtendonk-de Leeuw, A. M., Aerts, B. J., and den Daas, J. H. (1998). Abnormal offspring following in vitro production of bovine preimplantation embryos: a field study. Theriogenology 49, 883–894.
Abnormal offspring following in vitro production of bovine preimplantation embryos: a field study.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BD3c7ps1KntA%3D%3D&md5=b4884ab9f544d95a828f76f47e1928f7CAS | 10732097PubMed |

Vastenhouw, N. L., Zhang, Y., Woods, I. G., Imam, F., Regev, A., Liu, X. S., Rinn, J., and Schier, A. F. (2010). Chromatin signature of embryonic pluripotency is established during genome activation. Nature 464, 922–926.
Chromatin signature of embryonic pluripotency is established during genome activation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXjvVSgsrc%3D&md5=ec71fe9a8a644eff89183b8495245491CAS | 20336069PubMed |

Velker, B. A., Denomme, M. M., and Mann, M. R. (2012). Embryo culture and epigenetics. Methods Mol. Biol. 912, 399–421.
| 1:CAS:528:DC%2BC3sXmslajtg%3D%3D&md5=31a7d765eeb384e248cbcc98c26d02bfCAS | 22829387PubMed |

Venkatesh, S., Smolle, M., Li, H., Gogol, M. M., Saint, M., Kumar, S., Natarajan, K., and Workman, J. L. (2012). Set2 methylation of histone H3 lysine 36 suppresses histone exchange on transcribed genes. Nature 489, 452–455.
Set2 methylation of histone H3 lysine 36 suppresses histone exchange on transcribed genes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xht1Ghsb%2FI&md5=98298008a8c14335d13210d4940725bfCAS | 22914091PubMed |

Waddington, C. H. (2012). The epigenotype. 1942. Int. J. Epidemiol. 41, 10–13.
The epigenotype. 1942.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DC%2BC38zhs1WltA%3D%3D&md5=25dad2cd5df0b29bb45565283c52eb1cCAS | 22186258PubMed |

Walker, S. K., Hartwich, K. M., and Seamark, R. F. (1996). The production of unusually large offspring following embryo manipulation: concepts and challenges. Theriogenology 45, 111–120.
The production of unusually large offspring following embryo manipulation: concepts and challenges.Crossref | GoogleScholarGoogle Scholar |

Wang, F., Kou, Z., Zhang, Y., and Gao, S. (2007). Dynamic reprogramming of histone acetylation and methylation in the first cell cycle of cloned mouse embryos. Biol. Reprod. 77, 1007–1016.
Dynamic reprogramming of histone acetylation and methylation in the first cell cycle of cloned mouse embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXhsVSgtLbE&md5=db6e84356323c0981a1397add9d6f7d2CAS | 17823087PubMed |

Wang, L., Zhang, J., Duan, J., Gao, X., Zhu, W., Lu, X., Yang, L., Li, G., Ci, W., Li, W., Zhou, Q., Aluru, N., Tang, F., He, C., Huang, X., and Liu, J. (2014). Programming and inheritance of parental DNA methylomes in mammals. Cell 157, 979–991.
Programming and inheritance of parental DNA methylomes in mammals.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXotFGntb0%3D&md5=757ea1fc9b13b4c3705ef41c19c2e6a7CAS | 24813617PubMed |

Watkins, A. J., and Fleming, T. P. (2009). Blastocyst environment and its influence on offspring cardiovascular health: the heart of the matter. J. Anat. 215, 52–59.
Blastocyst environment and its influence on offspring cardiovascular health: the heart of the matter.Crossref | GoogleScholarGoogle Scholar | 19215321PubMed |

Wen, D., Banaszynski, L. A., Rosenwaks, Z., Allis, C. D., and Rafii, S. (2014). H3.3 replacement facilitates epigenetic reprogramming of donor nuclei in somatic cell nuclear transfer embryos. Nucleus 5, 369–375.
H3.3 replacement facilitates epigenetic reprogramming of donor nuclei in somatic cell nuclear transfer embryos.Crossref | GoogleScholarGoogle Scholar | 25482190PubMed |

Wen, L., Fu, L., Guo, X., Chen, Y., and Shi, Y. B. (2015). Histone methyltransferase Dot1L plays a role in postembryonic development in Xenopus tropicalis. FASEB J. 29, 385–393.
Histone methyltransferase Dot1L plays a role in postembryonic development in Xenopus tropicalis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXislGjtLY%3D&md5=4c7c00da7b501478b504efa2f86f6b4bCAS | 25366346PubMed |

Wossidlo, M., Nakamura, T., Lepikhov, K., Marques, C. J., Zakhartchenko, V., Boiani, M., Arand, J., Nakano, T., Reik, W., and Walter, J. (2011). 5-Hydroxymethylcytosine in the mammalian zygote is linked with epigenetic reprogramming. Nat. Commun. 2, 241.
5-Hydroxymethylcytosine in the mammalian zygote is linked with epigenetic reprogramming.Crossref | GoogleScholarGoogle Scholar | 21407207PubMed |

Wrenzycki, C., Herrmann, D., Lucas-Hahn, A., Gebert, C., Korsawe, K., Lemme, E., Carnwath, J. W., and Niemann, H. (2005). Epigenetic reprogramming throughout preimplantation development and consequences for assisted reproductive technologies. Birth Defects Res. C Embryo Today 75, 1–9.
Epigenetic reprogramming throughout preimplantation development and consequences for assisted reproductive technologies.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXktFSltLc%3D&md5=4e6169b3b245df5ae70adb0f62f11a31CAS | 15838918PubMed |

Wu, X., Li, Y., Xue, L., Wang, L., Yue, Y., Li, K., Bou, S., Li, G. P., and Yu, H. (2011). Multiple histone site epigenetic modifications in nuclear transfer and in vitro fertilized bovine embryos. Zygote 19, 31–45.
Multiple histone site epigenetic modifications in nuclear transfer and in vitro fertilized bovine embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXjtVSgtw%3D%3D&md5=1a10057b9cb39d6c3d3823addaa34375CAS | 20609268PubMed |

Wu, F. R., Liu, Y., Shang, M. B., Yang, X. X., Ding, B., Gao, J. G., Wang, R., and Li, W. Y. (2012). Differences in H3K4 trimethylation in in vivo and in vitro fertilization mouse preimplantation embryos. Genet. Mol. Res. 11, 1099–1108.
Differences in H3K4 trimethylation in in vivo and in vitro fertilization mouse preimplantation embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XnsVGiur0%3D&md5=9f6373bbfd09afb43806d450f9408e6bCAS | 22614279PubMed |

Young, J. B., and Morrison, S. F. (1998). Effects of fetal and neonatal environment on sympathetic nervous system development. Diabetes Care 21, B156–B160.
| 9704244PubMed |

Young, L. E., Sinclair, K. D., and Wilmut, I. (1998). Large offspring syndrome in cattle and sheep. Rev. Reprod. 3, 155–163.
Large offspring syndrome in cattle and sheep.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1cXntlaltL8%3D&md5=f8b310dcfcc67e4119d2de803ac03ef6CAS | 9829550PubMed |

Young, L. E., Fernandes, K., McEvoy, T. G., Butterwith, S. C., Gutierrez, C. G., Carolan, C., Broadbent, P. J., Robinson, J. J., Wilmut, I., and Sinclair, K. D. (2001). Epigenetic change in IGF2R is associated with fetal overgrowth after sheep embryo culture. Nat. Genet. 27, 153–154.
Epigenetic change in IGF2R is associated with fetal overgrowth after sheep embryo culture.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXhtFGktL8%3D&md5=c65829986653ffe791fcf55227fbe680CAS | 11175780PubMed |

Zhang, M., Wang, F., Kou, Z., Zhang, Y., and Gao, S. (2009). Defective chromatin structure in somatic cell cloned mouse embryos. J. Biol. Chem. 284, 24 981–24 987.
Defective chromatin structure in somatic cell cloned mouse embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhtV2nsLbI&md5=ba6e716bb864b6a9ee2e7c2d8b12f610CAS |

Zhang, A., Xu, B., Sun, Y., Lu, X., Gu, R., Wu, L., Feng, Y., and Xu, C. (2012). Dynamic changes of histone H3 trimethylated at positions K4 and K27 in human oocytes and preimplantation embryos. Fertil. Steril. 98, 1009–1016.
Dynamic changes of histone H3 trimethylated at positions K4 and K27 in human oocytes and preimplantation embryos.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhtV2nu7bI&md5=c22bc0db1124b69088c284118383eb18CAS | 22818287PubMed |

Zhang, K., Haversat, J. M., and Mager, J. (2013). CTR9/PAF1c regulates molecular lineage identity, histone H3K36 trimethylation and genomic imprinting during preimplantation development. Dev. Biol. 383, 15–27.
CTR9/PAF1c regulates molecular lineage identity, histone H3K36 trimethylation and genomic imprinting during preimplantation development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXhsFems7fL&md5=ebd96048bd17466d1fd9acc396068639CAS | 24036311PubMed |

Zhang, Y., Wei, C., Zhang, P., Li, X., Liu, T., Pu, Y., Li, Y., Cao, Z., Cao, H., Liu, Y., Zhang, X., and Zhang, Y. (2014). Efficient reprogramming of naïve-like induced pluripotent stem cells from porcine adipose-derived stem cells with a feeder-independent and serum-free system. PLoS One 9, e85089.
Efficient reprogramming of naïve-like induced pluripotent stem cells from porcine adipose-derived stem cells with a feeder-independent and serum-free system.Crossref | GoogleScholarGoogle Scholar | 24465482PubMed |

Zhou, L., Opalinska, J., Sohal, D., Yu, Y., Mo, Y., Bhagat, T., Abdel-Wahab, O., Fazzari, M., Figueroa, M., Alencar, C., Zhang, J., Kambhampati, S., Parmar, S., Nischal, S., Hueck, C., Suzuki, M., Freidman, E., Pellagatti, A., Boultwood, J., Steidl, U., Sauthararajah, Y., Yajnik, V., McMahon, C., Gore, S. D., Platanias, L. C., Levine, R., Melnick, A., Wickrema, A., Greally, J. M., and Verma, A. (2011). Aberrant epigenetic and genetic marks are seen in myelodysplastic leukocytes and reveal Dock4 as a candidate pathogenic gene on chromosome 7q. J. Biol. Chem. 286, 25 211–25 223.
Aberrant epigenetic and genetic marks are seen in myelodysplastic leukocytes and reveal Dock4 as a candidate pathogenic gene on chromosome 7q.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXosFGmtL0%3D&md5=58c33dea41eb3dfc2192f6a449d491a0CAS |

Zhou, L., Wang, P., Zhang, J., Heng, B. C., and Tong, G. Q. (2015). ING2 (inhibitor of growth protein-2) plays a crucial role in preimplantation development. Zygote , .
ING2 (inhibitor of growth protein-2) plays a crucial role in preimplantation development.Crossref | GoogleScholarGoogle Scholar | 25672483PubMed |