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

Cytoplasmic and nuclear determinants of the maternal-to-embryonic transition

Anilkumar Bettegowda A B , Kyung-Bon Lee A B and George W. Smith A B C D
+ Author Affiliations
- Author Affiliations

A Laboratory of Mammalian Reproductive Biology and Genomics, Michigan State University, East Lansing, MI 48824, USA.

B Department of Animal Science, Michigan State University, East Lansing, MI 48824, USA.

C Department of Physiology, Michigan State University, East Lansing, MI 48824, USA.

D Corresponding author. Email: smithge7@msu.edu

Reproduction, Fertility and Development 20(1) 45-53 https://doi.org/10.1071/RD07156
Published: 12 December 2007

Abstract

Although improvements in culture systems have greatly enhanced in vitro embryo production, success rates under the best conditions are still far from ideal. The reasons for developmental arrest of the majority of in vitro produced embryos are unclear, but likely attributable, in part, to intrinsic and extrinsic influences on the cytoplasmic and/or nuclear environment of an oocyte and/or early embryo that impede normal progression through the maternal-to-embryonic transition. The maternal-to-embryonic transition is the time period during embryonic development spanning from fertilisation until when control of early embryogenesis changes from regulation by oocyte-derived factors to regulation by products of the embryonic genome. The products of numerous maternal effect genes transcribed and stored during oogenesis mediate this transition. Marked epigenetic changes to chromatin during this window of development significantly modulate embryonic gene expression. Depletion of maternal mRNA pools is also an obligatory event during the maternal-to-embryonic transition critical to subsequent development. An increased knowledge of the fundamental mechanisms and mediators of the maternal-to-embryonic transition is foundational to understanding the regulation of oocyte quality and future breakthroughs relevant to embryo production.


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.
PubMed |

Alizadeh, Z. , Kageyama, S. , and Aoki, F. (2005). Degradation of maternal mRNA in mouse embryos: selective degradation of specific mRNAs after fertilization. Mol. Reprod. Dev. 72, 281–290.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Aoki, F. , Hara, K. T. , and Schultz, R. M. (2003). Acquisition of transcriptional competence in the 1-cell mouse embryo: requirement for recruitment of maternal mRNAs. Mol. Reprod. Dev. 64, 270–274.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bachvarova, R. , De Leon, V. , Johnson, A. , Kaplan, G. , and Paynton, B. V. (1985). Changes in total RNA, polyadenylated RNA, and actin mRNA during meiotic maturation of mouse oocytes. Dev. Biol. 108, 325–331.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bartel, D. P. (2004). MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116, 281–297.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Beaujean, N. , Hartshorne, G. , Cavilla, J. , Taylor, J. , Gardner, J. , Wilmut, I. , Meehan, R. , and Young, L. (2004). Non-conservation of mammalian preimplantation methylation dynamics. Curr. Biol. 14, R266–R267.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bensaude, O. , Babinet, C. , Morange, M. , and Jacob, F. (1983). Heat shock proteins, first major products of zygotic gene activity in mouse embryo. Nature 305, 331–333.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bettegowda, A. , and Smith, G. W. (2007). Mechanisms of maternal mRNA regulation: implications for mammalian early embryonic development. Front. Biosci. 12, 3713–3726.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bouniol, C. , Nguyen, E. , and Debey, P. (1995). Endogenous transcription occurs at the 1-cell stage in the mouse embryo. Exp. Cell Res. 218, 57–62.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Brevini, T. A. , Cillo, F. , Colleoni, S. , Lazzari, G. , Galli, C. , and Gandolfi, F. (2004). Expression pattern of the maternal factor zygote arrest 1 (Zar1) in bovine tissues, oocytes, and embryos. Mol. Reprod. Dev. 69, 375–380.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Bultman, S. J. , Gebuhr, T. C. , Pan, H. , Svoboda, P. , Schultz, R. M. , and Magnuson, T. (2006). Maternal BRG1 regulates zygotic genome activation in the mouse. Genes Dev. 20, 1744–1754.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Burns, K. H. , Viveiros, M. M. , Ren, Y. , Wang, P. , DeMayo, F. J. , Frail, D. E. , Eppig, J. J. , and Matzuk, M. M. (2003). Roles of NPM2 in chromatin and nucleolar organization in oocytes and embryos. Science 300, 633–636.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Camous, S. , Kopecny, V. , and Flechon, J. E. (1986). Autoradiographic detection of the earliest stage of [3H]-uridine incorporation into the cow embryo. Biol. Cell 58, 195–200.
PubMed |

Chandolia, R. K. , Peltier, M. R. , Tian, W. , and Hansen, P. J. (1999). Transcriptional control of development, protein synthesis, and heat-induced heat shock protein 70 synthesis in 2-cell bovine embryos. Biol. Reprod. 61, 1644–1648.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Christians, E. , Campion, E. , Thompson, E. M. , and Renard, J. P. (1995). Expression of the HSP 70.1 gene, a landmark of early zygotic activity in the mouse embryo, is restricted to the first burst of transcription. Development 121, 113–122.
PubMed |

Christians, E. , Michel, E. , Adenot, P. , Mezger, V. , Rallu, M. , Morange, M. , and Renard, J. P. (1997). Evidence for the involvement of mouse heat shock factor 1 in the atypical expression of the HSP70.1 heat shock gene during mouse zygotic genome activation. Mol. Cell. Biol. 17, 778–788.
PubMed |

Christians, E. , Davis, A. A. , Thomas, S. D. , and Benjamin, I. J. (2000). Maternal effect of Hsf1 on reproductive success. Nature 407, 693–694.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Conover, J. C. , Temeles, G. L. , Zimmermann, J. W. , Burke, B. , and Schultz, R. M. (1991). Stage-specific expression of a family of proteins that are major products of zygotic gene activation in the mouse embryo. Dev. Biol. 144, 392–404.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Dean, W. , Santos, F. , Stojkovic, M. , Zakhartchenko, V. , Walter, J. , Wolf, E. , and Reik, W. (2001). Conservation of methylation reprogramming in mammalian development: aberrant reprogramming in cloned embryos. Proc. Natl Acad. Sci. USA 98, 13 734–13 738.
Crossref | GoogleScholarGoogle Scholar | PubMed |

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.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Flach, G. , Johnson, M. H. , Braude, P. R. , Taylor, R. A. , and Bolton, V. N. (1982). The transition from maternal to embryonic control in the 2-cell mouse embryo. EMBO J. 1, 681–686.
PubMed |

Frei, R. E. , Schultz, G. A. , and Church, R. B. (1989). Qualitative and quantitative changes in protein synthesis occur at the 8–16-cell stage of embryogenesis in the cow. J. Reprod. Fertil. 86, 637–641.
PubMed |

Fulka, H. , Mrazek, M. , Tepla, O. , and Fulka, J. (2004). DNA methylation pattern in human zygotes and developing embryos. Reproduction 128, 703–708.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Giraldez, A. J. , Cinalli, R. M. , Glasner, M. E. , Enright, A. J. , Thomson, J. M. , Baskerville, S. , Hammond, S. M. , Bartel, D. P. , and Schier, A. F. (2005). MicroRNAs regulate brain morphogenesis in zebrafish. Science 308, 833–838.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Giraldez, A. J. , Mishima, Y. , Rihel, J. , Grocock, R. J. , Van Dongen, S. , Inoue, K. , Enright, A. J. , and Schier, A. F. (2006). Zebrafish MiR-430 promotes deadenylation and clearance of maternal mRNAs. Science 312, 75–79.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Hyttel, P. , Viuff, D. , Avery, B. , Laurincik, J. , and Greve, T. (1996). Transcription and cell cycle-dependent development of intranuclear bodies and granules in two-cell bovine embryos. J. Reprod. Fertil. 108, 263–270.
PubMed |

Kanka, J. (2003). Gene expression and chromatin structure in the pre-implantation embryo. Theriogenology 59, 3–19.
Crossref | GoogleScholarGoogle Scholar | PubMed |

King, W. A. , Niar, A. , Chartrain, I. , Betteridge, K. J. , and Guay, P. (1988). Nucleolus organizer regions and nucleoli in preattachment bovine embryos. J. Reprod. Fertil. 82, 87–95.
PubMed |

Kingston, R. E. , and Narlikar, G. J. (1999). ATP-dependent remodeling and acetylation as regulators of chromatin fluidity. Genes Dev. 13, 2339–2352.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Kopecny, V. , Flechon, J. E. , Camous, S. , and Fulka, J. (1989). Nucleologenesis and the onset of transcription in the eight-cell bovine embryo: fine-structural autoradiographic study. Mol. Reprod. Dev. 1, 79–90.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Lane, N. , Dean, W. , Erhardt, S. , Hajkova, P. , Surani, A. , Walter, J. , and Reik, W. (2003). Resistance of IAPs to methylation reprogramming may provide a mechanism for epigenetic inheritance in the mouse. Genesis 35, 88–93.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Latham, K. E. , and Schultz, R. M. (2001). Embryonic genome activation. Front. Biosci. 6, D748–D759.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Le Douarin, B. , Nielsen, A. L. , Garnier, J. M. , Ichinose, H. , Jeanmougin, F. , Losson, R. , and Chambon, P. (1996). A possible involvement of TIF1 alpha and TIF1 beta in the epigenetic control of transcription by nuclear receptors. EMBO J. 15, 6701–6715.
PubMed |

Lonergan, P. (2007). State-of-the-art embryo technologies in cattle. Soc. Reprod. Fertil. 64((Suppl.)), 315–325.


Ma, J. , Zeng, F. , Schultz, R. M. , and Tseng, H. (2006). Basonuclin: a novel mammalian maternal-effect gene. Development 133, 2053–2062.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Matsumoto, K. , Anzai, M. , Nakagata, N. , Takahashi, A. , Takahashi, Y. , and Miyata, K. (1994). Onset of paternal gene activation in early mouse embryos fertilized with transgenic mouse sperm. Mol. Reprod. Dev. 39, 136–140.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Matzuk, M. M. , Burns, K. H. , Viveiros, M. M. , and Eppig, J. J. (2002). Intercellular communication in the mammalian ovary: oocytes carry the conversation. Science 296, 2178–2180.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Mayer, W. , Niveleau, A. , Walter, J. , Fundele, R. , and Haaf, T. (2000). Demethylation of the zygotic paternal genome. Nature 403, 501–502.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Memili, E. , and First, N. L. (2000). Zygotic and embryonic gene expression in cow: a review of timing and mechanisms of early gene expression as compared with other species. Zygote 8, 87–96.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Misirlioglu, M. , Page, G. P. , Sagirkaya, H. , Kaya, A. , Parrish, J. J. , First, N. L. , and Memili, E. (2006). Dynamics of global transcriptome in bovine matured oocytes and preimplantation embryos. Proc. Natl Acad. Sci. USA 103, 18 905–18 910.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Morgan, H. D. , Santos, F. , Green, K. , Dean, W. , and Reik, W. (2005). Epigenetic reprogramming in mammals. Hum. Mol. Genet. 14, R47–R58.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Murchison, E. P. , Stein, P. , Xuan, Z. , Pan, H. , Zhang, M. Q. , Schultz, R. M. , and Hannon, G. J. (2007). Critical roles for Dicer in the female germline. Genes Dev. 21, 682–693.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Nielsen, A. L. , Ortiz, J. A. , You, J. , Oulad-Abdelghani, M. , Khechumian, R. , Gansmuller, A. , Chambon, P. , and Losson, R. (1999). Interaction with members of the heterochromatin protein 1 (HP1) family and histone deacetylation are differentially involved in transcriptional silencing by members of the TIF1 family. EMBO J. 18, 6385–6395.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Olek, A. , and Walter, J. (1997). The pre-implantation ontogeny of the H19 methylation imprint. Nat. Genet. 17, 275–276.
Crossref | GoogleScholarGoogle Scholar | PubMed |

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.
Crossref | GoogleScholarGoogle Scholar | PubMed |

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.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Paynton, B. V. , Rempel, R. , and Bachvarova, R. (1988). Changes in state of adenylation and time course of degradation of maternal mRNAs during oocyte maturation and early embryonic development in the mouse. Dev. Biol. 129, 304–314.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Pennetier, S. , Uzbekova, S. , Perreau, C. , Papillier, P. , Mermillod, P. , and Dalbies-Tran, R. (2004). Spatio-temporal expression of the germ cell marker genes MATER, ZAR1, GDF9, BMP15, and VASA in adult bovine tissues, oocytes, and preimplantation embryos. Biol. Reprod. 71, 1359–1366.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Pennetier, S. , Perreau, C. , Uzbekova, S. , Thelie, A. , Delaleu, B. , Mermillod, P. , and Dalbies-Tran, R. (2006). MATER protein expression and intracellular localization throughout folliculogenesis and preimplantation embryo development in the bovine. BMC Dev. Biol. 6, 26.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Peterson, C. L. , and Workman, J. L. (2000). Promoter targeting and chromatin remodeling by the SWI/SNF complex. Curr. Opin. Genet. Dev. 10, 187–192.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Philpott, A. , Leno, G. H. , and Laskey, R. A. (1991). Sperm decondensation in Xenopus egg cytoplasm is mediated by nucleoplasmin. Cell 65, 569–578.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Pillai, R. S. (2005). MicroRNA function: multiple mechanisms for a tiny RNA? RNA 11, 1753–1761.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Plante, L. , Plante, C. , Shepherd, D. L. , and King, W. A. (1994). Cleavage and 3H-uridine incorporation in bovine embryos of high in vitro developmental potential. Mol. Reprod. Dev. 39, 375–383.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Plasterk, R. H. (2006). Micro RNAs in animal development. Cell 124, 877–881.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Roest, H. P. , Baarends, W. M. , de Wit, J. , van Klaveren, J. W. , Wassenaar, E. , Hoogerbrugge, J. W. , van Cappellen, W. A. , Hoeijmakers, J. H. , and Grootegoed, J. A. (2004). The ubiquitin-conjugating DNA repair enzyme HR6A is a maternal factor essential for early embryonic development in mice. Mol. Cell. Biol. 24, 5485–5495.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Sagata, N. , Watanabe, N. , Vande Woude, G. F. , and Ikawa, Y. (1989). The c-mos proto-oncogene product is a cytostatic factor responsible for meiotic arrest in vertebrate eggs. Nature 342, 512–518.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Santos, F. , and Dean, W. (2004). Epigenetic reprogramming during early development in mammals. Reproduction 127, 643–651.
Crossref | GoogleScholarGoogle Scholar | PubMed |

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.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Schultz, R. M. (1993). Regulation of zygotic gene activation in the mouse. Bioessays 15, 531–538.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Schultz, R. M. , and Worrad, D. M. (1995). Role of chromatin structure in zygotic gene activation in the mammalian embryo. Semin. Cell Biol. 6, 201–208.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Su, Y. Q. , Sugiura, K. , Woo, Y. , Wigglesworth, K. , Kamdar, S. , Affourtit, J. , and Eppig, J. J. (2007). Selective degradation of transcripts during meiotic maturation of mouse oocytes. Dev. Biol. 302, 104–117.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Tang, F. , Kaneda, M. , O’Carroll, D. , Hajkova, P. , Barton, S. C. , Sun, Y. A. , Lee, C. , Tarakhovsky, A. , Lao, K. , and Surani, M. A. (2007). Maternal microRNAs are essential for mouse zygotic development. Genes Dev. 21, 644–648.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Telford, N. A. , Watson, A. J. , and Schultz, G. A. (1990). Transition from maternal to embryonic control in early mammalian development: a comparison of several species. Mol. Reprod. Dev. 26, 90–100.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Thompson, E. M. , Legouy, E. , and Renard, J. P. (1998). Mouse embryos do not wait for the MBT: chromatin and RNA polymerase remodeling in genome activation at the onset of development. Dev. Genet. 22, 31–42.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Tian, Q. , Kopf, G. S. , Brown, R. S. , and Tseng, H. (2001). Function of basonuclin in increasing transcription of the ribosomal RNA genes during mouse oogenesis. Development 128, 407–416.
PubMed |

Tong, Z. B. , and Nelson, L. M. (1999). A mouse gene encoding an oocyte antigen associated with autoimmune premature ovarian failure. Endocrinology 140, 3720–3726.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Tong, Z. B. , Gold, L. , Pfeifer, K. E. , Dorward, H. , Lee, E. , Bondy, C. A. , Dean, J. , and Nelson, L. M. (2000). Mater, a maternal effect gene required for early embryonic development in mice. Nat. Genet. 26, 267–268.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Tong, Z. B. , Bondy, C. A. , Zhou, J. , and Nelson, L. M. (2002). A human homologue of mouse Mater, a maternal effect gene essential for early embryonic development. Hum. Reprod. 17, 903–911.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Torres-Padilla, M. E. , and Zernicka-Goetz, M. (2006). Role of TIF1alpha as a modulator of embryonic transcription in the mouse zygote. J. Cell Biol. 174, 329–338.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Tseng, H. , Biegel, J. A. , and Brown, R. S. (1999). Basonuclin is associated with the ribosomal RNA genes on human keratinocyte mitotic chromosomes. J. Cell Sci. 112, 3039–3047.
PubMed |

Uzbekova, S. , Roy-Sabau, M. , Dalbies-Tran, R. , Perreau, C. , and Papillier, P. , et al. (2006). Zygote arrest 1 gene in pig, cattle and human: evidence of different transcript variants in male and female germ cells. Reprod. Biol. Endocrinol. 4, 12.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Wright, S. J. (1999). Sperm nuclear activation during fertilization. Curr. Top. Dev. Biol. 46, 133–178.
PubMed |

Wu, X. , Viveiros, M. M. , Eppig, J. J. , Bai, Y. , Fitzpatrick, S. L. , and Matzuk, M. M. (2003a). Zygote arrest 1 (Zar1) is a novel maternal-effect gene critical for the oocyte-to-embryo transition. Nat. Genet. 33, 187–191.
Crossref | GoogleScholarGoogle Scholar | PubMed |

Wu, X. , Wang, P. , Brown, C. A. , Zilinski, C. A. , and Matzuk, M. M. (2003b). Zygote arrest 1 (Zar1) is an evolutionarily conserved gene expressed in vertebrate ovaries. Biol. Reprod. 69, 861–867.
Crossref | GoogleScholarGoogle Scholar | PubMed |