Skip to main content
Journal of Assisted Reproduction and Genetics logoLink to Journal of Assisted Reproduction and Genetics
. 2007 Sep 27;24(11):513–520. doi: 10.1007/s10815-007-9167-0

Expression profile of genes coding for DNA repair in human oocytes using pangenomic microarrays, with a special focus on ROS linked decays

Yves Jr Menezo 1,2,6,, GianLuigi Russo 3, Elisabetta Tosti 4, Said El Mouatassim 5, Moncef Benkhalifa 5
PMCID: PMC3455023  PMID: 17899356

Abstract

Purpose

To determine the level of expression for mRNAs that regulate DNA repair activity in oocytes at the germinal vesicle (GV) stage. Reactive oxygen species (ROS) have been shown to play a major role in the appearance of deleterious DNA decays, and this study focuses on the repair of damage linked to decay caused by the action of ROS. The oocyte needs a mechanism for repairing DNA decays in the early preimplantation embryo before the onset of genomic activation, since in the absence of repair, residual DNA damage would lead to either apoptosis or tolerance. Tolerance of DNA damage is a source of potential mutations.

Method

GV oocytes were selected for this study, both for the ethical reason that they are unsuitable for patient treatment, and because no transcription takes place during the period from GV to MII and then prior to genomic activation. The GV oocyte is therefore a good model for looking at DNA during the first cleavages of early preimplantation development. Six cohorts of GV oocytes were pooled for extraction of mRNA; the DNA was analysed using Affimetrix HG-UG133 Plus 2, containing 54,675 probe sets; spike and housekeeping genes were also added as internal controls.

Results

In GV oocytes, DNA repair pathways for oxidized bases are redundant. One step repair procedure (OSR), BER (base excision repair), MMR (mismatch repair) and NER (Nucleotide excision repair) are present. All the recognition proteins are also present. The chromatin assembly factors necessary for the maintenance of genomic stability are highly expressed.

Conclusion

Gene expression analysis shows that the oocyte does not allow a high level of tolerance for DNA decays. This regulatory mechanism should avoid transmitting mutations into the next generation.

Keywords: Human oocyte, ROS, mRNA, DNA repair

Introduction

The impact of male infertility on human reproduction is still a matter of debate. For many years, male fertility has been defined as the ability of sperm to fertilize oocytes that develop into early cleavage stage embryos. In human IVF, the development of two to four cell embryos was considered to be a test for sperm fertilizing ability, with the assumption that all of the resulting embryos had the same developmental potential, irrespective of sperm and oocyte quality. However, it is now admitted that differences in male fertility are not related simply to sperm penetration failure, but also to the ability of the oocyte to support zygote development post fertilization. Attention has therefore been more carefully focused towards the quality of sperm DNA: a range of methods have been developed to evaluate paternal DNA integrity that are independent of all semen parameters including sperm morphology, concentration and motility [1]. The same conclusion is reached irrespective of the methodology used to examine DNA integrity: sperm DNA fragmentation and decays related to reactive oxygen species (ROS) lead to reduced fertility [2, 3], and ROS have a heavily deleterious impact on sperm DNA [4, 5]. A major end product of ROS damage to DNA is the formation of apurinic/apyrimidic (AP) sites [6, 7], but 8 oxo deoxyguanosine (8 OXO dG or 8 OH dG) can also be produced. We recently described the formation of ethenoadenosine and ethenoguanosine (1,N2-etheno-2′-deoxyguanosine (ɛdGuo), and 1,N6-etheno-2′-deoxyadenosine (ɛdAdo) in human sperm [8]. Shortened telomeres, in relation to age and sperm DNA fragmentation [9] may lead to inappropriate end to end fusion and the formation of chromosome bridges. On the other hand, paternal DNA always carries a varying level of DNA strand breaks [110]. These decays are the result of ROS and/or anomalies of topoisomerase activity during spermatogenesis. Other types of decays may arise, especially in sperm with high teratospermia. Demethylation mediated by DNA repair is also a potent promoter of epigenetic activation [11].

The worldwide use of ICSI raises questions regarding oocyte DNA repair activity. All DNA lesions have profound effects on cell viability, leading to apoptosis and then cell death, which is a recurrent feature in IVF. The most highly detrimental situation is probably when the cell tolerates these unrepaired decays—this leads sooner or later to mutations and oncogenic transformation. Fortunately, mammalian cells, including oocytes, have systems that allow DNA repair [1215]. DNA repair is probably one of the most important processes taking place during and post-fertilization in the oocyte and zygote, in order to allow complete embryonic development and avoid developmental arrests. The quality of maternal DNA during oocyte maturation at the genetic and epigenetic level is a matter for conjecture, but there is indirect evidence showing a positive effect of antioxidants on oocyte quality [16]. GV oocytes collected during a stimulated IVF cycle are unsuitable for treatment and are therefore surplus; therefore, primarily for ethical reasons as the French law forbids any type of work destroying human embryos, these were selected to examine the expression profile of genes involved in DNA repair activity. However, the mRNA content in GV oocytes is the same as that of MII oocytes, since there is no transcription during the final stages of oocyte maturation, only a remodeling of mRNA that is basically related to polyadenylation [17]. The first steps of preimplantation embryo development, at least until maternal to zygotic transition (i.e. four to eight-cell stage), are carried out with the use of maternal stores of protein and mRNA [1820].

Materials and methods

The oocytes were collected from two different IVF units, at the time of preparation for ICSI. The current protocol is semi long analog treatment followed by stimulation with recombinant FSH. The mean age of the patient was 33.1 years. After careful denudation of all oocytes retrieved, any GV oocytes present were used for this study, a total of 160 oocytes. The pooled oocytes were divided into six samples: the first four samples had ten oocytes per batch, and the remaining two samples were allocated sixty oocytes each, in order to increase the starting quantity of mRNA.

The analyses were performed with microarrays, using the Affymetrix HG U-133 plus 2.0 chips. Amplification was performed via double IVT (in vitro transcription, two-cycle cDNA synthesis kit, Affimetrix). Data were normalised using the algorithm MAS5, in order to generate a signal for each of the 54,613 transcripts of the chips (Probe set ID) that was proportional to the level of expression level. A call signal points out and specifies if the gene is expressed at a level above the background. The expression level is given as high, medium or low according to the signal. Housekeeping and spike controls were used for each determination. Expression was interpreted as low when the signal was above the background (Signal = 15) and up to a signal of 200 in at least three of the samples. The expression was classified as medium for a signal between 200 and 1,000; it was considered as high for a signal repeatedly over 1,000. Our references for the human DNA repair genes were based on Wood et al. [12, 21] and Cline and Hanawalt [22].

Results

Out of 54,613 probes tested, the overall number detected was 15,615 (27.7%, range 23.7–33.7). The number of probes at the limit of detection was between 1.2 and 1.5%, with a maximum of 879 with no clear cut absence or presence. No one of the DNA repair probe was in this borderline zone of expression. Five housekeeping gene probes and 17 spikes were used as internal controls. Absence of Amphiregulin (AREG), epiregulin (EREG) and betacellulin (BTC) that are normally present in cumulus cells, was checked in order to confirm that the GV oocytes were completely denuded.

The list of the genes involved in DNA repair activity [12, 21, 22], with their level of expression is listed in Table 1.

Table 1.

Expression levels of mRNAs coding for DNA repair in human oocytes

Gene Expression level Role in DNA repair
Base excision repair (BER)
UNG High Releases U
SMUG1 Low Releases altered U
MBD4 Medium Releases U or T opposite G at CpG
TDG Low Releases U, T or EthenoC opposite G
OGG1 Medium Releases 8 oxo or FAPY G opposite C
MUTYH (MYH) Medium Releases A opposite 8 oxo
NTHL1 (NTH1) Medium Releases fragmented or ring saturated pyrimidines (Fapy), thymine glycol, 5 OHU, 6-OHC
MPG High Removes 3me-A, ethenoA, hypoxanthine
NEIL1 Low Remove thymidine glycol, 8-oxoG, FapyG, FapyA, 5OH–U, 5OH–C
NEIL2 Low Removes oxidative products of pyrimidines, 5OH–U, 5OH–C
APEX1 High AP endonuclease
APEX2 Medium AP endonuclease
LIG3 Low Ligase III, DNA, ATP-dependent, main ligation function
XRCC1 High+ Ligase accessory factor, main ligation function
PNKP Low Convert DNA breaks to ligatable ends
PARP1 (ADPRT) High ADP-ribosyltransferase, protects against strand interruption
PARP2 (ADPRTL2) High Id
Direct reversal of damage
MGMT Medium O-6-methylguanine–DNA methyltransferase
MGC90512 (ABH2) High Alkylation repair: 1-Me A dioxygenase
DEPC-1 (ABH3) Medium Alkylation repair: 1-Me A dioxygenase
Repair of DNA–protein cross-links
TDP1 High Tyrosyl–DNA phosphodiesterase: removes TOP1–DNA complex
Mismatch excision repair (MMR)
MSH2 High DNA mismatch repair protein, mismatch and loop recognition
MSH3 High Id
MSH6 High Id
MSH4 Not expressed
MSH5 Medium Id, specialized for meiosis
PMS1 Medium DNA mismatch repair protein
MLH1 Medium Mut forming heterodimers
PMS2 Medium DNA mismatch repair protein
MLH3 Not expressed Mut of unknown function
PMS2L3 Low Id
PMS2L4 (PMS6) Not expressed
Nucleotide excision repair (NER)
XPC High DNA repair protein, binds DNA damaged
RAD23B (HR23B) High Excision repair protein
CETN2 High Binds damaged DNA as complex
RAD23A (HR23A) Medium Substitute for RAD 23B
XPA High Binds DNA damaged in preincision complex
RPA1 Very High Replication protein, binds DNA in preincision complex
RPA2 Low Replication protein, ID
RPA3 High Replication protein, ID
TFIIH Catalyses unwinding in preincision complex
ERCC3 (XPB) Not expressed « ‘to 5’ DNA helicase
ERCC2 (XPD) Not expressed 5′ to 3′ DNA helicase
GTF2H1 High General transcription factor
GTF2H2 High ID
GTF2H3 Low ID
GTF2H4 Low ID
GTF2H5 (TTDA) High ID
CDK7 High Kinase subunit of TFIIH
CCNH High Cyclin H, kinase subunit of TFIIH
MNAT1 Not expressed Kinase subunit of TFIIH
ERCC5 (XPG) High 3′ Incision
ERCC1 Not expressed 5′ Incision
ERCC4 (XPF) Not expressed 5′ Incision
LIG1 Medium DNA joining
NER-related
CKN1 (CSA) Not expressed
ERCC6 (CSB) High Putative transcription-repair coupling factor
XAB2 (HCNP) Medium PA binding protein 2
DDB1 High Damage-specific DNA binding protein
DDB2 Medium Damage-specific DNA binding protein
MMS19L (MMS19) Medium Transcription and NER
Homologous recombination
RAD51 Low Homologous pairing, DNA repair protein
RAD51L1(RAD51B) Medium ID
RAD51C High ID
RAD51L3(RAD51D) Not expressed
DMC1 Low Meiotic recombination protein
XRCC2 Not expressed DNA break and cross-link repair
XRCC3 Not expressed ID
RAD52 Not expressed Accessory factor for recombination
RAD54L Medium DNA repair and recombination protein
RAD54B Low DNA repair and recombination protein
BRCA1 High Accessory factor for transcription and recombination
BRCA2 Medium Cooperation with RAD1 for homologous pairing
SHFM1 (DSS1) Low Associated with BRCA2
RAD50 Medium ATPase
MRE11A High 3′Exonuclease
EME1 (MMS4L) High Essential meiotic endonuclease
Non-homologous end-joining
G22P1 (Ku70) Not expressed DNA end binding
XRCC5 (Ku80) Medium ATP dependant DNA helicase
PRKDC Not expressed Protein kinase, DNA-activated, catalytic polypeptide
LIG4 Medium Ligase IV, DNA, ATP-dependent
XRCC4 High Ligase accessory factor
DCLRE1C Medium DNA cross-link repair, nuclease
XLF NHEJ1) Not expressed XRCC4–LIG4 interacting factor
Modulation (sanitization) of nucleotide pools
NUDT1 (MTH1) Medium 8-OxoGTPase
DUT High DUTP pyrophosphatase
RRM2B (p53R2) High Ribonucleotide reductase (TP53 inducible)
DNA polymerases (catalytic subunits)
POLB High Polymerase beta nuclear DNA directed
POLG Medium Polymerase delta, mitochondrial DNA directed
POLD1 Low Polymerase DNA directed, NER and MMR
POLE Low Polymerase DNA directed, epsilon
PCNA High+ Proliferating cell nuclear antigen, sliding clamp for POLD and POLE
REV3L (POLZ) Medium Catalytic subunit of DNA polymerase zeta
(REV7)MAD2L2 Medium Catalytic subunit of DNA polymerase zeta
REV1L (REV1) High DCMP transferase
POLH Medium Polymerase DNA directed, eta
POLI (RAD30B) Medium Polymerase DNA directed iota, lesion bypass
POLQ High Polymerase, DNA directed, theta, DNA cross-link repair
POLK (DINB1) Not expressed Polymerase, DNA directed, kappa
POLL Not expressed Polymerase DNA directed, lambda
POLM Not expressed Polymerase (DNA directed), mu
POLN (POL4P) Not expressed Polymerase (DNA directed) nu
Editing and processing nucleases
FEN1(DNaseIV) High Flap structure-specific endonuclease, 5′ nuclease
TREX2 Not expressed 3′ Repair exonuclease, 3exo′ nuclease
EXO1(HEX1) High 5′ Exonuclease
SPO11 Not expressed Endonuclease
FLJ35220 (ENDOV) Low Incision 3′ of hypoxanthine and uracil
Rad6 pathway
UBE2A (RAD6A) Not expressed Ubiquitin-conjugating enzyme
UBE2B (RAD6B) Not expressed Ubiquitin-conjugating enzyme
RAD18 Medium E 3 ubiquitin ligase, assists repair or replication of damaged DNA
UBE2N (UBC13) High Ubiquitin-conjugating enzyme
Chromatin Structure
H2AFX (H2AX) High H2A histone family, phosphorylated after DNA damage, DNA damage signalling
CHAF1A (CAF1) High Chromatin assembly in DNA replication and DNA repair
Other identified genes with a suspected DNA repair function
DCLRE1A (SNM1) High DNA cross-link repair 1A
DCLRE1B(SNM1B) High DNA cross-link repair
RPA4 Medium Replication protein, binds DNA in preincision complex
APTX (aprataxin) Medium Aprataxin: processing of DNA single strand interruption
NEIL3 Low Endonuclease
RECQL (RECQ1) Low RecQ protein-like (DNA helicase Q1-like)
RECQL5 High DNA helicase
HEL308 Medium DNA helicase
RAD52B (RDM1) Not Expressed
Other conserved DNA damage response genes
ATR High DNA damage sensor
RAD1 Low Cell cycle checkpoint protein
RAD9A High Cell cycle checkpoint control protein
HUS1 Low Checkpoint protein
RAD17 (RAD24) Medium Cell cycle checkpoint protein
CHEK2 High CHK2 checkpoint homolog, effector kinase
TP53 Medium Tumor suppressor 53, regulation of the cell cycle, apoptosis

Discussion and conclusion

In all eukaryotic cells, DNA damage must be highlighted by DNA sensors that initiate a cellular response. Three options are then open in response to this signal: DNA repair, cell cycle arrest or apoptosis—events that occur frequently in human IVF. In order to maintain genomic stability, the early embryo must have the capacity to fully repair DNA, in particular damage linked to ROS-related decays in sperm DNA. DNA oxidation leads globally to three different oxidized molecules for each of the bases, the most important being 8 oxo guanine, 8 oxo adenine, FapyA and G (aA and G) and thymine glycol. Most of the genes with a proven or a suspected DNA repair activity for these decays are expressed in the oocyte.

Using BER, a multistep pathway involving several protein complexes, the removal of a single base is performed by a DNA glycosylase in the oocyte. The remaining AP site (apurinic/apyrimidic) site is hydrolyzed by an AP endonuclease. The single nucleotide gap is filled by polymerase beta (highly expressed in the oocyte—see Table 1—DNA polymerase, catalytic subunits) and ligated. BER pathways seem able to remove etheno-adducts and other oxidative products. APEX, capable of initiating the repair of apurinic/apyrimidic (AP) sites, the most common decay in damaged DNA, is highly expressed. APEX-1 also known to play an important role in response to oxidative stress, providing the reducing potential necessary for ref-1 activity in relation to thoredoxin:. [6, 23, 24]. The DNA-binding activity of APEX is modulated by a post translational mechanism involving reduction oxidation, and so is at least partly mediated by ROS: the RNAs coding for thiorexdoxin and Thioredoxin reductase and the transcripts for 6 members of the peroxiredoxin family are also present. This suggests that there is a very active complete pathway for the repair of ROS induced DNA decays. Surprisingly, OGG1, which removes 8 oxo DG, does not seem to have a high level of expression. The mRNA for 8-oxo-GTPase is highly expressed (see modulation of nucleotide pools): this enzyme hydrolyzes the oxidized nucleotide 8-oxo-dGTP to 8-oxo-dGMP, preventing incorporation of 8-oxo-dGTP into DNA.

After DNA replication, mismatch (MMR) repair is one of the pathways that is also associated with the repair of 8-oxoG lesions; MSH2–MSH6, highly expressed here, binds the 8-oxo containing DNA substrates, enhances ATPase and increases ADT–→ ADP exchange activity [25]. The oocyte seems to be well equipped for removal of DNA lesions linked to ROS, but not through the most common pathway, OGG1. We recently described the presence of etheno adducts in human sperm [8]. N-methyl purine DNA glycosylase (MPG) mRNA is highly expressed in the oocyte, and this is also true, to a lesser extent, for TDG (thymine DNA glycosylase): both enzymes remove the etheno adducts from DNA. TDP1 (tyrosyl–DNA phosphodiesterase 1), also highly expressed in the oocyte, removes glycolate from single-stranded DNA containing a 3′-phosphoglycolate, suggesting a role in the repair of free-radical mediated DNA double-strand breaks.

DNA methylation is a major epigenetic process, with a particular role in imprinting. SAM synthetase activity is required for DNA methylation activity through DMT (DNA methyl transferase), and has a high level of expression in both mouse and human oocytes and early preimplantation embryos [26]. The mechanism for direct reversal of damage, repairing alkylation damage in one step, is also expressed. This system is of major importance in the control of hypermethylation, i.e. the regulation of imprinting processes.

The oocyte has a full panel of mRNAs coding for Nucleotide excision repair (NER or global genome repair), a multistep pathway that involves more than 30 proteins. TCR, transcription-coupled-repair, is a more discrete pathway, responsible for removal of DNA insults that block transcription. Since there is no transcription before the 8-cell stage, this pathway is redundant at the earlier stages. However, depending on their mechanism of storage, these mRNAs can be translated at later stages post-fertilization. Other genes involved in this NER/TCR pathway are highly expressed: ERCC6(CSB), GTF2H1,2 and 5, and MMS19. Shortening of telomeres has a strong association with sperm DNA fragmentation [9], and this leads to aberrant DNA repair responses. The tumor suppressor gene TP53 plays a crucial role in controlling this inappropriate response to damage, by determining whether the DNA will be repaired, or whether the cell will undergo programmed cell death. Any decrease in TP53 activity will allow a permissive environment, leading to chromosome fusion and the formation of chromosome bridges. We found a surprisingly moderate level of TP53 transcription. TDP1 is a ubiquitous factor involved in repairing covalent topoisomerase I-DNA complexes (DNA-protein cross-links). DCLE1A and B, with DNA cross-link repair activity, are highly expressed (see Table 1: other identified genes with suspected DNA repair activity).

Maintaining a correct three-dimensional chromatin structure is mandatory, as the 3D structure is critical during chromosome segregation. This is especially true in human IVF, where a high incidence of mosaicism is a common feature. Histones are a central feature of transcription regulation, DNA repair, DNA replication and chromosomal stability; BRCA1 and 2, highly expressed in the oocyte, play a central role in maintaining genomic stability (See also Wells et al. 2005 [20]). We also found a very high level of expression for H2AFX (H2A histone family member X) and CHAF1 (chromatin assembly factor 1), both involved in chromatin structure regulation during and after DNA repair. It must be pointed out that several types of regulation, in particular polyadenylation [1727], will allow these mRNAs to be translated: portions of the mRNA can be lost both pre and post translation. Poly(A) binding proteins and the cytoplasmic polyadenylation elements are highly expressed in the oocyte (up to 500 times the background level). However, mRNA expression usually indicates that the corresponding proteins are being produced and further utilized, especially in the case of the early embryo prior to genomic activation [20].

In conclusion, high permissivity and tolerance for DNA decays is a highly detrimental situation for oocytes and early preimplantation embryos. Although not all of the mRNAs detected here are fully translated, and we have not considered the level of functional protein repair, the pathways identified are highly expressed, and are somewhat redundant. In human IVF, developmental arrest and implantation failure may be linked to one of the classical responses to DNA decays, i.e. apoptosis [28]. However, when DNA decays reach a very high level in gametes, as is the case in heavy smokers [29], the limits of DNA repair activity might be overreached, with an increase in childhood cancer due to DNA mutations as a potential result. This aspect should not be overlooked in Assisted reproductive technology, especially for IVF/ICSI using poor quality sperm with high levels of ROS-related decays and a high degree of sperm DNA fragmentation [30].

On a final note, growth hormone (GH) up regulates DNA repair capacity in liver cells [31], human naked oocytes and cumulus cells have receptors for GH [32], and co-stimulation with GH has been shown to be efficient in human IVF, especially in patients over 40 years of age where mRNA expression is decreasing. An attempt to increase the DNA repair capacity of the oocyte, perhaps with the use of GH would be an interesting approach to examine.

Acknowledgment

We would like to thank Dr Martine Dumont and Dr Anne Marie Junca for the generous gift of GV oocytes. Dr Soularue (Partnerchips) is acknowledged for his help in the realization of the project.

This work was sponsored by Organon France.

References

  • 1.Evenson DP, Larson KL, Jost LK. Sperm chromatin structure assay: its clinical use for detecting sperm DNA fragmentation in male infertility and comparisons with other techniques. J Androl. 2002;23:25–43. doi: 10.1002/j.1939-4640.2002.tb02599.x. [DOI] [PubMed] [Google Scholar]
  • 2.Henkel R, Hajimohammad M, Stalf TH, Hoogendijk C, Mehnert C, Menkveld R, et al. Influence of deoxyribonucleic acid damage on fertilization and pregnancy. Fertil Steril. 2004;81:965–972. doi: 10.1016/j.fertnstert.2003.09.044. [DOI] [PubMed] [Google Scholar]
  • 3.Kodama H, Yamaguchi R, Fukuda J, Kasai H, Tanaka T. Increased oxidative deoxyribonucleic acid damage in the spermatozoa of infertile male patients. Fertil Steril. 1997;68:519–524. doi: 10.1016/S0015-0282(97)00236-7. [DOI] [PubMed] [Google Scholar]
  • 4.Lopes S, Sun J, Jurisicova A, Meriano J, Casper RF. Semen deoxyribonucleic acid fragmentation is increased in poor quality semen samples and correlates with failed fertilisation in intracytoplasmic sperm injection. Fertil Steril. 1998;69:528–532. doi: 10.1016/S0015-0282(97)00536-0. [DOI] [PubMed] [Google Scholar]
  • 5.Lopes S, Jurisicova A, Sun JG, Meriano J, Casper RF. Reactive oxygen species: potential cause for DNA fragmentation in human spermatozoa. Hum Reprod. 1998;13:896–900. doi: 10.1093/humrep/13.4.896. [DOI] [PubMed] [Google Scholar]
  • 6.Hsieh MM, Hegde V, Kelley MR, Deutsch WA. Activation of APE/Ref-1 redox activity is mediated by reactive oxygen species and PKC phosphorylation. Nucleic Acids Res. 2001;29:3116–3122. doi: 10.1093/nar/29.14.3116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Xu YJ, Kim Y. Demple B excision of c-4′-oxidised deoxyribose lesions from double-stranded DNA by human apurinic/apyrimidic endonuclease (Ape1 protein) and DNA polymerase beta. J Biol Chem. 1998;273:28837–28844. doi: 10.1074/jbc.273.44.28837. [DOI] [PubMed] [Google Scholar]
  • 8.Badouard C, Ménézo Y, Panteix G, Ravanat JL, Douki T, Cadet J, et al. Determination of new types of DNA lesions in human sperm. Zygote 2007;15 in press [DOI] [PubMed]
  • 9.Rodriquez S, Goyanes V, Sequrelles E, Blasco M, Gosálvez J, Fernández JL. Critically short telomeres are associated with sperm DNA fragmentation. Fertil Steril. 2005;84:843–845. doi: 10.1016/j.fertnstert.2005.05.014. [DOI] [PubMed] [Google Scholar]
  • 10.Guerin P, Matillon C, Bleau G, Levy R, Menezo Y. Impact of sperm DNA fragmentation on ART outcome. Gynecol Obstet Fertil. 2005;33:665–668. doi: 10.1016/j.gyobfe.2005.07.015. [DOI] [PubMed] [Google Scholar]
  • 11.Baretto G, Schaffer A, Marholg G, Stach D, Swaminathan SK, Handa V, et al. Gadd45a promotes epigenetic gene activation by repair-mediated DNA demethylation. Nature 2007;445:671–67. [DOI] [PubMed]
  • 12.Wood RD, Mitchell M, Sgouros JG, Lindahl T. Human DNA repair genes. Science. 2001;291:1284–1289. doi: 10.1126/science.1056154. [DOI] [PubMed] [Google Scholar]
  • 13.Bessho T, Tano K, Kasai H, Ohtsuka E, Nishimura S. Evidence for two DNA repair enzymes for 8-hydroxyguanosine (7,8 dihydro-8-oxoguanine) in human cells. J Biol Chem. 1996;268:19416–19421. [PubMed] [Google Scholar]
  • 14.Zeng F, Baldwin DA, Shultz RA. Transcript profiling during preimplantation mouse development. Dev Biol. 2004;272:483–496. doi: 10.1016/j.ydbio.2004.05.018. [DOI] [PubMed] [Google Scholar]
  • 15.Kocabas MA, Crosby J, Ross PJ, Otu HH, Beyhan Z, Can H, et al. The transcriptome of human oocytes. Proc Natl Acad Sci. 2006;103:14027–14032. doi: 10.1073/pnas.0603227103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Tarin JJ, Perez-Albala S, Cano A. Oral antioxidants counteract the negative effects of female aging on oocyte quantity and quality in the mouse. Mol Reprod Dev. 2002;61:385–397. doi: 10.1002/mrd.10041. [DOI] [PubMed] [Google Scholar]
  • 17.Mouatassim S, Guerin P, Ménézo Y. Expression of genes encoding antioxidant enzymes in human and mouse oocytes during final stages of maturation. Mol Hum Reprod. 1999;5:720–725. doi: 10.1093/molehr/5.8.720. [DOI] [PubMed] [Google Scholar]
  • 18.Braude P, Bolton V, Moore S. Human gene expression first occurs between the four-and eight-cell stage of preimplantation development. Nature. 1988;332:459–461. doi: 10.1038/332459a0. [DOI] [PubMed] [Google Scholar]
  • 19.Telford NA, Watson AJ, Schultz GA. Transition from maternal to embryonic control in early mammalian development: a comparison of several species. Mol Reprod Dev. 1990;26:90–100. doi: 10.1002/mrd.1080260113. [DOI] [PubMed] [Google Scholar]
  • 20.Wells D, Bermudez MG, Steuerwald N, Thornhill AR, Walker DL, Malter H, et al. Expression of genes regulating chromosome segregation, the cell cycle and apoptosis during human preimplantation development. Human Reprod. 2005;20:1339–1348. doi: 10.1093/humrep/deh778. [DOI] [PubMed] [Google Scholar]
  • 21.Wood RD, Mitchell M, Lindahl T. Human DNA repair genes. Mutat Res. 2005;577:275–283. doi: 10.1016/j.mrfmmm.2005.03.007. [DOI] [PubMed] [Google Scholar]
  • 22.Cline SD, Hanawalt PC. Who’s on first in the cellular response to DNA damage? Nat Rev Mol Cell Biol. 2003;4:361–372. doi: 10.1038/nrm1101. [DOI] [PubMed] [Google Scholar]
  • 23.Kelley MR, Parsons SH. Redox regulation of the DNA repair functions of the human AP endonuclease Ape1/ref-1. Antioxid Redox Signal. 2001;3:671–683. doi: 10.1089/15230860152543014. [DOI] [PubMed] [Google Scholar]
  • 24.Hedley D, Pintilie M, Woo J, Nicklee T, Morrison A, Birle D, et al. Up-regulation of the redox mediators thioredoxin and apurinic/apyrimidinic excision (APE)/Ref-1 in hypoxic microregions of invasive cervical carcinomas, mapped using multispectral, wide-field fluorescence image analysis. Am J Pathol. 2004;164:557–565. doi: 10.1016/S0002-9440(10)63145-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Mazurek A, Berardini M, Fishel R. Activation of human MutS homologs by 8-oxo-guanine DNA damage. J Biol Chem. 2002;227:8260–8266. doi: 10.1074/jbc.M111269200. [DOI] [PubMed] [Google Scholar]
  • 26.Menezo Y, Khatchadourian Ch, Gharrib A, Hamidi J, Greenland T, Sarda N. Regulation of S-adenosyl methionine synthesis in the mouse embryo. Life Sci. 1989;44:1601–1609. doi: 10.1016/0024-3205(89)90455-4. [DOI] [PubMed] [Google Scholar]
  • 27.Seli E, Lalioti MD, Flaherty SM, Sakkas D, Terzi N, Steitz JA. An embryonic poly(A)-binding protein (ePAB) is expressed in mouse oocytes and early preimplantation embryos. Proc Natl Acad Sci USA. 2005;102:367–372. doi: 10.1073/pnas.0408378102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Jurisicova A, Antenos M, Varmusa S, Tilly JL, Casper RF. Expression of apoptosis-related genes during human preimplantation embryo development: potential roles for the Harakiri gene product and Caspase-3 in blastomere fragmentation. Mol Hum Reprod. 2003;9:133–141. doi: 10.1093/molehr/gag016. [DOI] [PubMed] [Google Scholar]
  • 29.Zenzes MT. Smoking and reproduction: gene damage to human gametes and embryos. Hum Reprod Updat. 2000;6:122–131. doi: 10.1093/humupd/6.2.122. [DOI] [PubMed] [Google Scholar]
  • 30.Aitken RJ, Krausz C. Oxidative stress, DNA damage and the Y chromosome. Reproduction. 2001;122:497–506. doi: 10.1530/rep.0.1220497. [DOI] [PubMed] [Google Scholar]
  • 31.Thompson BJ, Shang CA, Waters MJ. Identification of genes induced by growth hormone in rat liver using cDNA. Endocrinology. 2000;141:4321–4324. doi: 10.1210/en.141.11.4321. [DOI] [PubMed] [Google Scholar]
  • 32.Menezo Y, Mouatassim S, Chavrier M, Servy EJ, Nicolet B. Human oocytes and preimplantation embryos express mRNA for Growth Hormone Receptor (GhR*) Zygote. 2003;11:293–297. doi: 10.1017/S096719940300234X. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Assisted Reproduction and Genetics are provided here courtesy of Springer Science+Business Media, LLC

RESOURCES