GET THE APP

Key Features of Genomic Imprinting during Mammalian Spermatogenes
Anatomy & Physiology: Current Research

Anatomy & Physiology: Current Research
Open Access

ISSN: 2161-0940

+44 1300 500008

Review Article - (2016) Volume 6, Issue 5

Key Features of Genomic Imprinting during Mammalian Spermatogenesis: Perspectives for Human assisted Reproductive Therapy: A Review

Miriama Štiavnická1,2*, Olga García Álvarez1, Jan Nevoral1,2, Milena Králícková1,2 and Peter Sutovsky3,4
1Laboratory of Reproductive Medicine, Biomedical Centre, Department of Histology and Embryology, Charles University in Prague, Czech Republic
2Department of Histology and Embryology, Charles University in Prague, Czech Republic
3Division of Animal Science, and Departments of Obstetrics, Gynecology and Women’s Health, University of Missouri, Columbia, MO, USA
4omi_departments of Obstetrics, Gynaecology and Women’s Health, University of Missouri, Columbia, MO, USA
*Corresponding Author: Miriama Štiavnická, Laboratory of Reproductive Medicine, Biomedical Centre, Department of Histology and Embryology, Charles University in Prague, Czech Republic, Tel: +420377593808 Email:

Abstract

Increasing influence of epigenetics is obvious in all medical fields including reproductive medicine. Epigenetic alterations of the genome and associated post-translational modifications of DNA binding histones equally impact gamete development and maturation, as well as embryogenesis. Relationships between methylation and acetylation of histones and involvement of DNA methylation are important not only for chromatin remodelling but also significant to gene imprinting and thus affecting the gene expression in the embryo. Therefore, gene silencing accompanied with methylation of histones and DNA is a result of heterochromatin establishment in haploid male germ cell, the spermatid. Complex epigenetic changes leading to the establishment of histone code and heterochromatin are regulated by a broad range of factors, such as histone deacetylases, histone methyltransferases, non-coding RNAs, and small protein modifiers ubiquitin and SUMO. These factors are candidate diagnostic targets for reproductive medicine when anomalous gene imprinting or histone modification of the gametes may disrupt embryo development or cause developmental disorders in the offspring. Using advanced non-invasive techniques for sperm selection based on testing of epigenetic markers is a possible approach to more successful assisted reproductive therapy (ART) as well as prevention of epigenetic-origin disorders in ART babies.

Keywords: Sperm; Epigenetics; Histone methylation; Gene imprinting

Introduction

Spermatogenesis is the process of proliferation and differentiation of spermatogonial stem cells into adult spermatozoa within seminiferous tubules. In the first, diploid phase, spermatogonial stem cells multiply by mitosis and subsequently undergo meiotic division to reach the haploid phase, thus becoming spermatids. During this second phase, called spermiogenesis, spermatids differentiate to spermatozoa capable of acquiring potential for motility and fertilization [1].

Spermatogenesis encompasses distinct epigenetic events that are characterised as heritable changes in gene expression without changes of nucleotide sequence, but with an influence on cell phenotype [2]. Epigenetic events of spermatogenesis include DNA imprinting and chromatin remodelling leading to its condensation. These changes are defined by post-translational modifications of histone proteins, such as the acetylation, methylation, phosphorylation and ubiquitination, and gradual replacement of histones in the nucleosome by protamines [3-7].

Protamines are basic proteins that wrap around DNA more stringently than histones and create compact toroidal structure, which protects DNA. Protamines are also subject to post-translational modification such as phosphorylation of serine, threonine and tyrosine residues, the purpose of which is poorly understood. Another posttranslational modification of protamines is the formation of disulphide bonds that prevent dislocation from DNA [8-10]. Post-translational modifications either stimulate or repress developmentally regulated gene expression and are crucial for male fertility and paternally influenced aspects of embryo development. An incorrect replacement of histones, mutations in genes that encode for enzymes necessary for DNA methylation, and protein methylation and/or acetylation disrupt spermatogenesis and spermiogenesis [11,12]. Apart from these processes, noncoding RNAs also have an indispensable role in spermatogenesis [6,7].

Histone-protamine replacement is a gradual process confined to distinct areas of the spermatid genome. However, about 15% of original spermatocyte histones remain associated with the sperm DNA [13]. Originally considered a carryover from incomplete histoneprotamine exchange, these retained histones now appear to have an essential regulatory role, conveyed by their post-translational modifications, such as acetylation and methylation. In the next step of histone-protamine exchange, hyperacetylated testicular histones allow for DNA relaxation in the nucleosome, and are gradually replaced. First, testis-specific histone variants (H2B and TH2B) are incorporated into spermatid chromatin by transition protein 1 and 2 (TNP1, TNP2) which are then supplanted by protamines 1 and 2 (PRM1, PRM2) [5,11]. Initially, protamines are phosphorylated by serine/arginine protein specific kinase 1 (SRPK1) and calcium/calmodulin-dependent protein kinase 4 (CAMK4), targeting PRM1 and PRM2, respectively. Rapid dephosphorylation follows, allowing for the formation of disulphide bonds between the unmasked cysteine residues of dephosphorylated protamines [14,15]. These events take place during spermatid elongation [5,11].

The PRM1:PRM1 ratio of the mammalian sperm genome is close to 1:1 [16,17]. The shift of this ratio has been associated with male reproductive disorders [11,18]. After protamine-rich chromatin establishment, spermatid nucleus becomes more compact. Such a hypercondensed state of sperm chromatin conveys hydrodynamic sperm phenotype and protects sperm DNA from damage during sperm transport in both the male and female reproductive tracts [12].

Understanding sperm epigenetics will be beneficial for human assisted reproductive therapy (ART), wherein fertilization protocols do not approximate in vivo conditions. Currently, we do not know the crucial sperm epigenetic factors, and broad experiments elucidating their role in fertilization process and pre-implantation embryonic development are needed.

Histone Acetylation: Residual Sperm Histones Carry Developmentally Relevant Information

Acetylation of protein is characterised by transfer of an acetyl moiety from acetyl CoA to a free amino-group of the target protein. Histone acetylation is catalyzed by specific enzymes, namely histone acetyltransferases (HATs) and histone deacetylases (HDACs). Based on the binding site for acetyl group, there are two types of acetylation: lysine acetylation and N-terminal acetylation [19-21].

Sperm histone H3 is acetylated on lysine (K) residues K9, K18 and K23. The level of H3 acetylation changes throughout spermatogenesis, influencing both the proliferation and differentiation of male germ cells. Hyperacetylation of H3K9, H3K18 and H3K23 peaks in spermatogonia, and later tapers off to recur in elongating spermatids and dissipate in fully differentiated spermatozoa [22,23]. Contrary to H4 acetylation, the role of H3 acetylation is not known, and it awaits further examination [24].

The main reason for histone acetylation in spermatids is to facilitate chromatin remodelling, leading to genome-wide cessation of gene expression in the spermatid nuclei. Histone acetylation levels change during spermatogenesis. As a result, spermatozoa display acetylation of histone H4 on lysine residues K5, K8, K12 and K16. Hyperacetylation of histone H4 yields sites for binding of bromodomain testis associated proteins (BRDT) of the BET subfamily, including BRDT, BRD2, BRDT3 and BRDT4 [25-27]. During spermiogenesis, the BRDT proteins participate in chromatin remodelling required for successful spermatid differentiation [23,24]. Mice homozygous for BrdtΔBD1/ ΔBD1 mutation lacked BRDT from their spermatids and were sterile due to a defect in chromatin remodelling causing chromocenter fragmentation in spermatid nuclei during spermiogenesis [28,29]. With regard to other bromodomain proteins, there is little information about their function in spermatogenesis. However, there is a proven link between BRD2, BRD4 and mouse embryo lethality [30-32]. It is known that BRD2 has an impact on the development of the neural system, and its presence was also shown in oocytes and early embryos. The activity of BRD4 influences embryo development: the mutation of this gene has a negative effect on an embryo, and it is associated with epilepsy and neural developmental defects [31,33]. Physiological roles of BRDT proteins have mainly been studied in mice and rats [30-32]. Altogether, BRDT-driven changes on the genomic or epigenomic level impact fertilization, embryo implantation and development to term in rodents, supporting an assumption that similar effects are conveyed by them in human [34].

Berkovits with collective proposed that in normal healthy spermatids, the BRDT protein keeps SIRT1 (sirtuin-1, NAD+-dependent HDAC) out of the chromocenter, the pericentromeric heterochromatin-based structure. Ablation of the murine Sirt1 gene did not affect post-meiotic male germ cells, but pre-meiotic Sirt1 inactivation delayed pre-meiotic germ cell differentiation, conveyed abnormal shape of spermatids with higher levels of DNA fragmentation and an overall smaller testis size in the mutant males [29]. Also noticeable was abnormal histones-protamine replacement and defects of chromatin condensation in spermatids [1,29].

Compared to wild type, disruption of Sirt1 gene results in smaller (in diameter), abnormally shaped seminiferous tubules, a reduced number of fully differentiated spermatozoa, a higher level of sperm DNA damage and compromised genome integrity. Consequently, mutant spermatozoa have a reduced ability to produce viable zygotes and offspring by in vitro fertilization (IVF) and embryo transfer, most likely due to high incidence of implantation failure [35].

Histone Methylation as a Regulator of Spermatogenesis

Histone acetylation is counterbalanced by methylation events that in general have an opposite effect on chromatin structure and gene expression. The DNA methyltransferases (DNMTs) perform both de novo methylation and maintenance methylation [36,37]. Cellular content of DNMTS reflects the level of methylation in genome. Apart from histone methylation, genomic DNA methylation also takes place [37,38].

Methylation activates gene expression, which can then be repressed by demethylation; it is defined as the binding of the methyl unit from the S-adenosyl-L-methionine to the 5th position of cytosine residues in nucleotides using enzymes methyltransferases [39]. Methylation of DNA is mostly observed in cytosine-phosphate-guanine dinucleotides (CpGs) [40]. Key chromatin/histone methylation events occur on histones H3 and H4. Mono, di- and tri- methylated histones have been detected during spermatogenesis [41,42].

To date, methylation of lysine residues K4, K9, K27, K36 and K79 of histone H3 and lysine K20 of histone H4 have been described [43]. While methylations of H3K4, H3K36 and H3K79 are typical for euchromatin, methylations of H3K9, H3K27 and H4K20 are more common for transcriptionally silent heterochromatin associated regions of the genome [44]. Changing levels of histone methylation during spermatogenesis suggest its impact on germ cell differentiation. While H3K4me1, H3K4me3 and H3K27me2/3 methylations were increased in spermatogonia and round spermatids, but not in elongating spermatids, methylation of H4K20 was detected throughout spermatogenesis [22,45]. However, during differentiation of spermatogonia, the level of H4K20me3 was reduced and H4K20me1 was increased, underscoring the influence of histone methylation on differentiation of spermatogonia to spermatids [45,46]. Methylation of histone H4 is connected with the process of chromatin remodeling, mediated by the replacement of histones with protamines. At the corresponding steps of spermatid elongation, the level of H4K20 methylation is reduced and H4 acetylation is increased, collectively allowing for histone replacement [47].

The equilibrium of histone methylation during spermatogenesis is maintained by demethylase enzymes. Overexpression of H3K4 demethylase KDM1A reduced H3K4me2 in spermatozoa and caused transgenerational developmental defect of offspring [48]. Conversely, genetic ablation of H3K9 demethylase JMJD1A caused germ cells apoptosis and anomalous spermatid elongation [49,50]. From observations about effect of histone methylation, gene exposure to translation or gene silencing for methylation of H3K4 and K9 we can assume the necessity of chromatin stability.

Histone Ubiquitination and Sumoylation: Epigenetic Factors or Mediators of Histone Degradation?

Protein ubiquitination is a stable, yet reversible post-translational modification by the covalent binding of the small chaperone protein ubiquitin to lysine residues of substrate proteins. This process is ATPdependent and catalyzed by three different classes of enzymes, the ubiquitin-activating enzyme E1 (UBA1), ubiquitin-conjugating enzymes (UBE2) and ubiquitin ligases (UBE3) [51]. Ubiquitination during spermatogenesis serves the substrate specific, developmentally regulated degradation of various proteins by the ubiquitin proteasome system (UPS). Apart from protein turnover, ubiquitination participates in the regulation of transcription, protein transport and cell signaling, among others [52]. Within the male reproductive system, ubiquitination has been implicated both in spermatogenesis and in epididymal sperm maturation as quality control to eliminate defective spermatozoa and dead epididymal epithelial cells [53-55]. This proposed role is supported by the detection of high amounts of ubiquitin in epididymal epithelia and luminal fluid. Ubiquitin as well as the enzyme of ubiquitin-substrate conjugation machinery are secreted by epididymal epithelium to eliminate defective spermatozoa by subsequent phagocytosis. Nevertheless, some of the defective spermatozoa tagged by extracellular/cell surface ubiquitination are carried over into ejaculate [56,57].

Whereas polyubiquitination plays a central role in the protein turnover by ubiquitin proteasome system, monoubiquitination influences gene expression [58]. The other role of ubiquitin is probably in meiotic sex chromosome inactivation (MSCI), deduced from high level of enzyme ubiquitin ligase UBR2 in unpaired XY axes as well as ubiquitination of histone H2. Ubiquitination of histone H2 is one of the most common histone ubiquitinations in mammals and has been implicated in transcriptional silencing such as X inactivation. Deficiency in UBE2B is associated with impaired spermatogenesis and male sterility [58-61]. Testis specific E2 enzyme UBC4 (UBC4-testis isoform), that mediates the first step of histone degradation and replacement by protamines during spermatid elongation [62,63], has been assigned an important role in spermatid histone ubiquitination.

Furthermore, protein ubiquitination is related to sumoylation that involves the attachment of small ubiquitin-like modifier (SUMO) to lysine residues of substrate proteins [64,65]. There are four SUMO isoforms: SUMO1, SUMO2, SUMO3 and SUMO4. Since SUMO2 and SUMO3 have nearly identical amino acid sequences, a collective name SUMO2/3 applies [66-68]. The SUMO1 has been localized to the nucleus and midpiece of human spermatozoa [69].

Sumoylation has been implicated in chromatin inactivation and transcriptional repression, supported by the presence of SUMO1 in XY bodies of pachytene spermatocytes, and near the centromere [70,71]. It is suspected that sumoylation participates in MSCI. Vigodner [72] evaluated early presence of SUMO1 on sex chromosomes during MSCI which was followed by γH2AX accumulation [72]. SUMO1 is important in the regulation of meiotic division of spermatocytes, where it binds to synaptonemal proteins SCP1 and SCP2 that form the synaptonemal complex as a prerequisite for recombination [73]. Accumulation of SUMO1 in human spermatozoa coincides with reduced motility and abnormal sperm morphology. Specific targets of sumoylation in male germ cells include the dynamin-related protein 1 (DRP1) in the sperm tail mid-piece, and Ran GTPase-activating protein 1 (RANGAP1) and DNA Topoisomerase IIa (TOP2A) in the postacrosomal region of the sperm head. Expression of SUMO1 and its binding to the aforementioned substrate proteins was significantly higher in males with reduced sperm motility and abnormal sperm morphology. In addition, a positive correlation was found between sperm DNA fragmentation and SUMO1 [74].

Furthermore, SUMO1 is important for organization of constitutive chromatin because the regions rich in SUMO1 also carry the hallmark heterochromatin proteins HP1α/CBX5, trimethylated H3K9 and trimethylated H4K20. On the other hand, sumoylation may block histone H4 methylation, an important step for subsequent H4 acetylation and histone-protamine replacement [75].

Aforementioned post-translational modifications represent essential epigenetics factors with key roles during gene imprinting, fertilization and embryonic development. These modifications are considered to be in active cross-talk with another epigenetic phenomenon-non coding RNAs [76,77], important for sperm-driven signalling in the fertilized oocyte/zygote.

Non-coding RNAs - Big Device for Small Spermatozoon

Noncoding RNAs are not translated to protein but have important regulatory function in cellular gene expression. Their role in spermatogenesis [78] and fertilization [79] has already been established. The basic classification of non-coding RNAs defines long noncoding RNAs (lnRNAs) and small noncoding RNAs (snRNAs). The small noncoding RNA group includes small interfering RNAs (siRNAs), microRNAs (miRNAs/MiR) and PIWI-interacting RNAs (piRNAs) [80-82]. The piRNAs differ from the rest in their biogenesis but share similar regulatory functions [7].

The miRNAs are small RNAs of 20-24 nucleotides that are complementary to their target RNAs. Synthesis of miRNA starts in the nucleus and finishes in the cytoplasm; it is catalyzed by enzymes of the RNA interference machinery, including DROSHA and DICER. The function of miRNAs is to inhibit translation of mRNA to protein and thus silence mRNA manifestation [83,84]. During spermatogenesis, differential expression patterns and wide repertoires of miRNAs were observed. Liu with collective identified 559 miRNAs at distinct stages of spermatogenesis [85]. Spermatozoal miRNAs may also control expression of genes in early embryonic development after fertilization [86].

Likewise, siRNAs have an important role in gene silencing. The length of these molecules is also 20-24 nucleotides. Similar to miRNAs, biosynthesis of siRNA requires endoribonuclease DICER. Together, miRNAs with siRNAs participate in the RNA-induced silencing complex (RISC) that is the main cellular tool for silencing of gene expression in spermatogenesis [87]. The siRNAs are often used as an experimental tool for knocking down gene expression and to deduce gene function [88,89].

Piwi RNAs specifically interact with proteins from PIWI family, such as MIWI, MIWI2 and MILI [90]. In comparison to siRNAs and miRNAs, piRNAs have 26-31 nucleotides and are independent of DICER. The role of piRNAs is in silencing transposome elements in germ line during gametogenesis [91]. When mutated, the MIWI2 protein holds spermatogonia in the leptotene stage of meiosis [92]. MIWI knock out blocks spermatid differentiation, causing male infertility [93]. Insufficiency of MILI and MIWI proteins is connected todefects of methylation that influence expression of retrotransposons in fetal male germ cells [94].

Long noncoding RNAs (lncRNAs) are longer than 200 bp and have an indispensable role in the regulation of spermatogenesis [95]; they inhibit binding of transcription factors to specific DNA sites and enhance DNA methylation [96,97]. The long noncoding RNAs can also be a resource for DROSHA and DICER to produce small noncoding RNA [98].

In addition to spermatogenesis, ncRNAs are essential for oocytecumulus complexes as well. Cumulus expansion is a hallmark of oocyte maturation [99,100]. During oogenesis, miR124 targets pentraxin 3 gene (Ptx3) that is important for cumulus expansion, and its silencing disrupts oogenesis [101]. In addition to this interaction, miR-378- aromatase [102] and miR-207-BDNF [103] interactions have been described, both significantly related to cumulus expansion and oocyte maturation. Contrary to miRNAs, the interactions and complex effects of ncRNAs, either sperm or oocyte derived, during fertilization and early embryogenesis remain unclear. Possible cross-talk between ncRNAs and other epigenetic regulatory mechanisms, such as DNA methylation and histone modifications remains to be characterized [79,104].

The Role of Sperm Histone Code in Fertilization and Early Embryogenesis

Immediately after fertilization, maternal and paternal pronuclei develop in the zygote. Rapid demethylation of DNA and histones is essential for the formation of a dedifferentiated embryo. Although rapid active DNA demethylation of paternal chromatin followed by paternal pronucleus development is observed [105,106], specific loci still remain methylated [107]. These epigenetic marks are necessary for further embryonic development. Aforementioned, locus-specific epigenetic modifications are closely tied with gene imprinting (Figure 1). Accurate parent-of-origin gene dosage is crucial for successful embryo development. Therefore, monoallelic gene activity, affecting about 150 genes, is observed during mouse embryonic development, respective of maternal or paternal allele-specific gene expression [108]. Both DNA methylation and the histone code established by posttranslational modifications regulate gene promoters and are required during gametogenesis, resulting in gene imprinting in the embryo (Figure 1) [109]. Accordingly, histone methylation, such as H3K9me3 and H3K20me2, is often associated with methylated DNA in the embryo and placenta [110]. Collectively, such parent-of-originspecific epigenetic changes promote differential expression of select genes in the embryo, particularly those encoding for transcription and growth factors [111]. Although relatively few genes are paternally imprinted, the necessity of sperm-derived gene imprinting is obvious [107].

anatomy-physiology-histone-modifications

Figure 1: The involvement of histone modifications in gene imprinting during spermatogenesis.

While genes are imprinted in a sex-specific manner, the regulatory mechanisms of gene imprinting seem to be shared by spermatogenesis and oogenesis [112]. Sperm gene imprinting depends on posttranslational modifications of residual sperm histones, such as H3K20me3 and H4K20me3 (Figure 1) [43,107]. Acquisition of DNA methylation by male germ cell stem cells, the pro-spermatogonia occurs during foetal development, e.g. on day 14.5-18.5 in the mouse; it is completed postnatally, in pachytene spermatocytes [113,114]. These epigenetic changes regulate a limited number of paternal loci, e.g. those affecting Gtl2 and Rasgrf1 genes [115,116]. The DNA methyltransferases (DNMT3A, DNMT3L) and histone methyltransferases are involved in male germline gene imprinting [117-120]. In spite of rapid paternal DNA demethylation after fertilization [106], paternally imprinted genes, such as H19, Gtl2 and Rasgrf1 are protected against demethylation of their promoters [110,121]. However, the molecular mechanisms of DNA methylationand histone code-driven gene imprinting remain poorly understood.

Imprinted genes are often grouped in clusters and these are controlled by imprinting control regions. Within those clusters lie the differentially methylated regions (DMRs) [40,110,122]. The best known DMR is H19-Igf2 in human locus 11p15, associated with imprinting disorders, that contains gene encoding Insulin-like growth factor 2 (IGF2) and H19 [123]. While the paternal H19 allele is methylated and the maternal one transcribed into a noncoding RNA, expression of Igf2 gene is exclusively from paternal allele [124-126]. Apart from DNA methylation and histone modifications, regulation by noncoding RNA is also important for proper gene imprinting. Alterations of ncRNA profiles are connected with epigenetic developmental disorders such as Goiter, Kabuki, and Claes-Jensen Xlinked mental retardation syndrome [127-129]. In general, abnormal imprinting has been associated with Angelman, Beckwith-Wiedemann, Prader-Willi and Silver-Russell syndromes [130-132].

Histone code is stable during proliferative (spermatogonium) and meiotic (primary and secondary spermatocyte) phases of spermatogenesis. Post-meiotically, histone acetylation promotes histone-protamine exchange, although some residual histones remain associated with spermatid DNA. At this time, post-translational modifications of core histones, particularly histone H3 on their lysine (K) residua, play an essential role in epigenetic imprinting of a small number of paternally-silenced loci at imprinting control regions (ICRs). Although DNA methylation within the imprinted genes (H19- Igf2, Gtl2, Rasgrf1) is well documented, the association of posttranslationally modified histones with imprinted loci remains unclear. In addition to methylation and acetylation, other modifications are potentially capable of regulating gene expression. The code of residual histones seems to be more precise and more specific than DNA methylation. The failure of gene imprinting leads to severe developmental disorders, birth defects such as the Angelman, Beckwith-Wiedemann, Prader-Willi, and Silver-Russell syndromes. Therefore, epigenetic changes during spermatogenesis represent a crucial process aimed at preventing epigenetic diseases in offspring.

Conclusion and Perspectives

By various estimates, 15-19% of reproductive age couples experience difficulties conceiving, with male factor infertility contributing to approximately one half of those cases. The most common option for infertile couples is using assisted reproductive therapy. Based on recent knowledge, the epigenetics seem to be impactful phenomenon for fertilization and early embryonic development. In addition to these processes, gametogenesis is affected by epigenetic changes as well. However, less is known about the consequences of gamete epigenetics for the success of embryonic development. Spermatogenesis is an attractive model system to study the epigenome because a large sperm output is continuously produced and various epigenetic changes occur throughout this process. Without a doubt, a number of epigenetic factors with crucial effect on sperm quality and male fertility are awaiting discovery.

Correct sperm chromatin structure gradually achieved during spermatogenesis is required for male fertility [133-135]. Accordingly, more detailed insights into spermatid DNA structure, histone modifications, sperm protamination and histone-protamine exchange have uncovered unique epigenetic mechanisms leading to the acquisition of sperm fertilization ability and sperm contribution to the regulation of embryonic development [11,18,48,110]. Indeed, spermatogenesis represents sensitive process of epigenetic regulation of sperm chromatin packaging as well as correct gene imprinting. Systematic study of epigenetics promises to solve some problems associated with clinical infertility and prevention of genetically determined syndromes.

Based on the body of knowledge reviewed above, identification of relevant biomarkers of sperm quality and male fertility will help clinicians differentiate between healthy and unhealthy spermatozoa and chart the treatment course for ART couples. Some biomarkers have already been validated in retrospective clinical studies, such as ubiquitin, post-acrosomal WW-domain binding protein (PAWP) and others, mainly by using flow cytometry [136].

Intense study and testing of markers of epigenetic defects using relevant biological models as well as clinical routine identify suitable approaches for epigenetics-based sperm analysis and selection. Some of recently utilized methods, such as SUTI assay [57,137] and ejaculate nanopurification [138] would be helpful for modification and epigenetic markers utilization. This approach can lead to the elimination of epigenome determined diseases, such as syndromes based on failure of gene imprinting. On the other hand, a more precise understanding of the relationship between sperm epigenome and the phenotype of both spermatozoon and embryo is needed. Although more research is necessary, the advanced sperm screening and selection methods will eventually be translated into assisted reproductive therapy.

Acknowledgements

We thank Ms. Kathy Craighead for editorial assistance. The research discussed in this review is funded by Charles University in Prague (PRVOUK P-36) and the National Sustainability Program I (NPU-040) No.LO1503 provided by the Ministry of Education, Youth and Sports of the Czech Republic. Work in P.S. laboratory relevant to this review is supported by Agriculture and Food Research Initiative Competitive Grant no. 2015-67015-23231 from the USDA National Institute of Food and Agriculture and by seed funding from the Food for the 21st Century program of the University of Missouri.

References

  1. Cheng CY (2008) Molecular Mechanisms in Spermatogenesis (Advances in Experimental Medicine and Biology). Springer-Verlag New York.
  2. Handel AE, Ebers GC, Ramagopalan SV (2010) Epigenetics: molecular mechanisms and implications for disease. Trends Mol Med 16: 7-16.
  3. Cho C, Willis A, Goulding E, Jung-Ha H, Choi Y, et al. (2001) Haploinsufficiency of protamine-1 or -2 causes infertility in mice. Nat. Genet 1: 82-86.
  4. Govin J, Caron C, Lestrat C, Rousseaux S, Khochbin S (2004) The role of histones in chromatin remodelling during mammalian spermiogenesis. European journal of biochemistry 271: 3459-69.
  5. Montellier E, Boussouar F, Rousseaux S, Zhang K, Buchou T, et al. (2013) Chromatin-to-nucleoprotamine transition is controlled by the histone H2B variant TH2B. Genes & development 27: 1680-92.
  6. Jenuwein T, Allis CD (2001) Translating the histone code. Science 293: 1074-80.
  7. Tolia NH, Joshua-Tor L (2007) Slicer and the argonautes. Nat ChemBiol 3: 36-43.
  8. Green GR, Balhorn R, Poccia DL, Hecht NB (1994) Synthesis and processing of mammalian protamines and transition proteins. MolReprodDev 37:255-263.
  9. Vilfan ID, Conwell CC, Hud NV (2004) Formation of native-like mammalian sperm cell chromatin with folded bull protamine. J BiolChem 279: 20088-20095.
  10. Balhorn R (2007) The protamine family of sperm nuclear proteins. Genome Biol 8: 227.
  11. Carrell DT, Emery BR, Hammoud S (2007) Altered protamine expression and diminished spermatogenesis: what is the link? Human Reproduction Update 13: 313-327.
  12. Oliva R (2006) Protamines and male infertility. Human Reproduction Update 12: 417-435.
  13. Gatewood JM, Cook GR, Balhorn R, Bradbury EM, Schmid CW (1987) Sequence-specific packaging of DNA in human sperm chromatin. Science 236: 962-4.
  14. Papoutsopoulou S, Nikolakaki E, Chalepakis G, Kruft V, Chevaillier P, etal. (1999) SR protein-specific kinase 1 is highly expressed in testis and phosphorylates protamine 1. Nucleic Acids Research 27: 2972-2980.
  15. Wu JY, Ribar TJ, Cummings DE, Burton KA, McKnight GS, et al. (2000) Spermiogenesis and exchange of basic nuclear proteins are impaired in male germ cells lacking Camk4. Nat Genet 25: 448-52.
  16. Govin J, Escoffier E, Rousseaux S, Kuhn L, Ferro M, et al. (2007) Pericentric heterochromatin reprogramming by new histone variants during mouse spermiogenesis. J Cell Biol 29: 283-94.
  17. Wykes SM, Krawetz SA (2003) The Structural Organization of Sperm Chromatin. J BiolChem 278: 29471-7.19.
  18. Francis S, Yelumalai S, Jones C, Coward K (2014) Aberrant protamine content in sperm and consequential implications for infertility treatment. Hum Fertil (Camb) 17: 80-9
  19. Allfrey VG, Faulkner R, Mirsky AE (1964) Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. ProcNatlAcadSci USA 51: 786-94.
  20. Polevoda B, Sherman F (2000) Nalpha-terminal acetylation of eukaryotic proteins. J BiolChem 275: 36479-36482.
  21. Pang A, Rennert O (2014) Protein acetylation and spermatogenesis. ReprodSyst Sex Disord: Suppl 1:5.
  22. Song N, Liu J, An S, Nishino T, Hishikawa Y, et al. (2011) Immunohistochemical Analysis of Histone H3 Modifications in Germ Cells during Mouse Spermatogenesis. ActaHistochemCytochem 44: 183-190.
  23. Kim JH, Jee BC, Lee JM, Suh CS, Kim SH (2014) Histone acetylation level and histone acetyltransferase/deacetylase activity in ejaculated sperm from normozoospermic men. Yonsei Med J 55: 1333-1340.
  24. Dai L, Endo D, Akiyama N, Yamamoto-Fukuda T, Koji T (2015) Aberrant levels of histone H3 acetylation induce spermatid anomaly in mouse testis. Histochem Cell Biol 143: 209-24.
  25. Govin J, Lestrat C, Caron C, Pivot-Pajot C, Rousseaux S, et al. (2006) Histone acetylation-mediated chromatin compaction during mouse spermatogenesis. Ernst Schering Res Found Workshop 57: 155-172.
  26. Shang E, Nickerson HD, Wen D, Wang X, Wolgemuth DJ (2007) The first bromodomain of Brdt, a testis-specific member of the BET sub-family of double-bromodomain-containing proteins, is essential for male germ cell differentiation. Development 134: 3507-15.
  27. Gaucher J, Boussouar F, Montellier E, Curtet S, Buchou T, et al. (2012) Bromodomain-dependent stage-specific male genome programming by Brdt. EMBO J 31: 3809-3820.
  28. Leser K, Awe S, Barckmann B, Renkawitz-Pohl R, Rathke C (2012) The bromodomain-containing protein tBRD-1 is specifically expressed in spermatocytes and is essential for male fertility. Biol Open 1: 597-606.
  29. Berkovits BD, Wolgemuth DJ (2011) The first bromodomain of the testis-specific double bromodomain protein Brdt is required for chromocenter organization that is modulated by genetic background. Developmental biology 360: 358-368.
  30. Shang E, Cui Q, Wang X, Beseler C, Greenberg DA, et al. (2011) The bromodomain-containing gene BRD2 is regulated at transcription, splicing, and translation levels. J Cell Biochem 112: 2784-2793.
  31. Shang E, Wang X, Wen D, Greenberg DA, Wolgemuth DJ (2009) Double bromodomain-containing gene Brd2 is essential for embryonic development in mouse. DevDyn 238: 908-917.
  32. Houzelstein D, Bullock SL, Lynch DE, Grigorieva EF, Wilson VA, et al. (2002) Growth and early postimplantation defects in mice deficient for the bromodomain-containing protein Brd4. Mol Cell Biol 22: 3794-3802.
  33. Pal DK, Evgrafov OV, Tabares P, Zhang F, Durner M, et al. (2003) BRD2 (RING3) is a probable major susceptibility gene for common juvenile myoclonic epilepsy. Am J Hum Genet 73:261–270.
  34. Kennedy C, Ahlering P, Rodriguez H, Levy S, Sutovsky P (2011) Sperm chromatin structure correlates with spontaneous abortion and multiple pregnancy rates in assisted reproduction. Reprod Biomed Online 22: 272-276.
  35. Coussens M, Maresh JG, Yanagimachi R, Maeda G, Allsopp R (2008) Sirt1 deficiency attenuates spermatogenesis and germ cell function. Public Library of Science one 3: e1571.
  36. Okano M, Bell DW, Haber DA, Li E (1999) DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99: 247-57.
  37. La Salle S, Mertineit C, Taketo T, Moens PB, Bestor TH, et al. (2004) Windows for sex-specific methylation marked by DNA methyltransferase expression profiles in mouse germ cells. DevBiol 268: 403-415.
  38. Trojer P, Reinberg D (2006) Histone lysine demethylases and their impact on epigenetics. Cell 125: 213-217.
  39. Chen T, Li E (2004) Structure and function of eukaryotic DNA methyltransferases. Curr Top DevBiol 60:55-89.
  40. Jones PA (2012) Functions of DNA methylation: islands, start sites, gene bodies and beyond. Nature Reviews. Genetics 13: 484-492.
  41. Godmann M, Auger V, Ferraroni-Aguiar V, Di Sauro A, Sette C, et al. (2007) Dynamic regulation of histone H3 methylation at lysine 4 in mammalian spermatogenesis. BiolReprod 77: 754-64.
  42. Khalil AM, Boyar FZ, Driscoll DJ (2004) Dynamic histone modifications mark sex chromosome inactivation and reactivation during mammalian spermatogenesis. ProcNatlAcadSci USA 101: 16583-16587.
  43. Hammoud SS, Nix DA, Zhang H, Purwar J, Carrell DT, et al. (2009) Distinctive chromatin in human sperm packages genes for embryo development. Nature 460: 473-478.
  44. Sims RJ 3rd, Nishioka K, Reinberg D (2003) Histone lysine methylation: a signature for chromatin function. Trends Genet 19: 629-39.
  45. An J, Qin J, Wan Y, Zhang Y, Hu Y, et al. (2015) Histone lysine methylation exhibits a distinct distribution during spermatogenesis in pigs. Theriogenology 84: 1455-1462.
  46. Shirakata Y, Hiradate Y, Inoue H, Sato E, Tanemura K (2014) Histone h4 modification during mouse spermatogenesis. J ReprodDev 60: 383-387.
  47. Sonnack V, Failing K, Bergmann M, Steger K (2002) Expression of hyperacetylated histone H4 during normal and impaired human spermatogenesis. Andrologia 34: 384-390.
  48. Siklenka K, Erkek S, Godmann M, Lambrot R, McGraw S, et al. (2015) Disruption of histone methylation in developing sperm impairs offspring health transgenerationally. Science p:350.
  49. Liu Z, Zhou S, Liao L, Chen X, Meistrich M, et al. (2010) Jmjd1a demethylase-regulated histone modification is essential for cAMP-response element modulator-regulated gene expression and spermatogenesis. J BiolChem 285: 2758-2770.
  50. Okada Y, Scott G, Ray MK, Mishina Y, Zhang Y (2007) Histone demethylase JHDM2A is critical for Tnp1 and Prm1 transcription and spermatogenesis. Nature 450: 119-23.
  51. Glickman MH, Ciechanover A (2002) The ubiquitin-proteasome proteolytic pathway: destruction for the sake of construction. Physiol Rev 82: 373-428.
  52. Conaway RC, Brower CS, Conaway JW (2002) Emerging roles of ubiquitin in transcription regulation. Science 296: 1254-8.
  53. Da Silva N, Barton CR (2016) Macrophages and dendritic cells in the post-testicular environment. Cell Tissue Res 363: 97-104.
  54. Baska KM, Manandhar G, Feng D, Agca Y, Tengowski MW, etal. (2008) Mechanism of extracellular ubiquitination in the mammalian epididymis. J Cell Physiol 215: 684-696.
  55. Richburg JH, Myers JL, Bratton SB (2014) The role of E3 ligases in the ubiquitin-dependent regulation of spermatogenesis. Semin Cell DevBiol 30: 27-35.
  56. Sutovsky P, Moreno R, Ramalho-Santos J, Dominko T, Thompson WE, Schatten G (2001) A putative, ubiquitin-dependent mechanism for the recognition and elimination of defective spermatozoa in the mammalian epididymis. J Cell Sci114: 1665-1675.
  57. Sutovsky P, Terada Y, Schatten G (2001) Ubiquitin-based sperm assay for the diagnosis of male factor infertility. Hum Reprod16:250-258.
  58. MulugetaAchame E, Wassenaar E, Hoogerbrugge JW,Sleddens-Linkels E, Ooms M, et al. (2010) The ubiquitin-conjugating enzyme HR6B is required for maintenance of X chromosome silencing in mouse spermatocytes and spermatids. BMC Genomics 11: 367.
  59. An JY, Kim EA, Jiang Y, Zakrzewska A, Kim DE (2010) UBR2 mediates transcriptional silencing during spermatogenesis via histone ubiquitination. ProcNatlAcadSci U S A 107: 1912-1917.
  60. Roest HP, van Klaveren J, de Wit J, van Gurp CG, Koken MH, etal. (1996) Inactivation of the HR6B ubiquitin-conjugating DNA repair enzyme in mice causes male sterility associated with chromatin modification. Cell 86: 799-810.
  61. Gaucher J, Reynoird N, MontellierE, Boussouar F, Rousseaux S, et al. (2010) From meiosis to postmeiotic events: the secrets of histone disappearance. FEBS J 277: 599-604.
  62. Chen HY, Sun JM, Zhang Y, Davie JR, Meistrich ML (1998) Ubiquitination of histone H3 in elongating spermatids of rat testes. J BiolChem 273: 13165-13169.
  63. Rajapurohitam V, Bedard N, Wing SS (2002) Control of ubiquitination of proteins in rat tissues by ubiquitin conjugating enzymes and isopeptidases. Am J PhysiolEndocrinolMetab 282: E739-45.
  64. Gill G (2004) SUMO and ubiquitin in the nucleus: different functions, similar mechanisms?Genes Dev 18: 2046-2059.
  65. Geiss-Friedlander R, Melchior F (2007) Concepts in sumoylation: a decade on.Nat Rev Mol Cell Biol 8: 947-956.
  66. Dohmen RJ (2004) SUMO protein modification.BiochimBiophysActa 1695: 113-131.
  67. Bohren KM, Nadkarni V, Song JH, Gabbay KH, Owerbach D (2004) A M55V polymorphism in a novel SUMO gene (SUMO-4) differentially activates heat shock transcription factors and is associated with susceptibility to type I diabetes mellitus. Journal of Biological Chemistry 279: 27233-27238.
  68. Saitoh H, Hinchey J (2000) Functional heterogeneity of small ubiquitin-related protein modifiers, SUMO-1 versus SUMO-2/3. J BiolChem 275: 6252-58.
  69. Marchiani S, Tamburrino L, Giuliano L, Nosi D, Sarli V, et al. (2011) Sumo1-ylation of human spermatozoa and its relationship with semen quality. International Journal of Andrology 34: 581-593.
  70. Vigodner M Morris PL (2005) Testicular expression of small ubiquitin-related modifier-1 (SUMO-1) supports multiple roles in spermatogenesis: silencing of sex chromosomes in spermatocytes, spermatid microtubule nucleation, and nuclear reshaping. DevBiol 282: 480-92.
  71. Vigodner M, Ishikawa T, Schlegel PN, Morris PL (2006) SUMO-1, human male germ cell development, and the androgen receptor in the testis of men with normal and abnormal spermatogenesis. Am J PhysiolEndocrinolMetab 290: E1022-33.
  72. Vigodner M (2009) Sumoylation precedes accumulation of phosphorylated H2AX on sex chromosomes during their meiotic inactivation. Chromosome Res 17: 37-45.
  73. Brown PW, Hwang K, Schlegel PN, Morris PL (2008) Small ubiquitin-related modifier (SUMO)-1, SUMO-2/3 and SUMOylation are involved with centromeric heterochromatin of chromosomes 9 and 1 and proteins of the synaptonemal complex during meiosis in men. Hum Reprod 23: 2850-7.
  74. Marchiani S, Tamburrino L, Ricci B, Nosi D, Cambi M, et al. (2014) SUMO1 in human sperm: new targets, role in motility and morphology and relationship with DNA damage. Reproduction 148:453-67.
  75. Metzler-Guillemain C, Depetris D, Luciani JJ, Mignon-Ravix C, Mitchell MJ, et al. (2008) In human pachytene spermatocytes, SUMO protein is restricted to the constitutive heterochromatin. Chromosome Res 16:761-782.
  76. Lin FM, Kumar S, Ren J, Karami S, Bahnassy S, et al. (2016) SUMOylation of HP1α supports association with ncRNA to define responsiveness of breast cancer cells to chemotherapy. Oncotarget 7:21.
  77. Zhu C, Chen C, Huang J, Zhang H, Zhao X, et al. (2015) SUMOylation at K707 of DGCR8 controls direct function of primary microRNA. J BiolChem 290:20893-20903.
  78. Rinn JL, Chang HY (2012) Genome regulation by long noncoding RNAs. Annu Rev Biochem 81: 145-66.
  79. Gilchrist GC, Tscherner A, Nalpathamkalam T, Merico D, LaMarre J (2016) MicroRNA Expression during Bovine Oocyte Maturation and Fertilization. Int J MolSci 17: E396.
  80. Bernstein E, Caudy AA, Hammond SM, Hannon GJ (2001) Role for a bidentateribonuclease in the initiation step of RNA interference. Nature 409: 363-366.
  81. Lee Y, Ahn C, Han J, Choi H, Kim J, et al. (2003) The nuclear RNase III Drosha initiates microRNA processing.Nature 425: 415-419.
  82. Girard A, Sachidanandam R, Hannon GJ, Carmell MA (2006) A germline-specific class of small RNAs binds mammalian Piwi proteins.Nature 442: 199-202.
  83. Boyd SD (2008) Everything you wanted to know about small RNA but were afraid to ask.Lab Invest 88: 569-578.
  84. Inui M, Martello G, Piccolo S (2010) MicroRNA control of signal transduction.Nat Rev Mol Cell Biol 11: 252-263.
  85. Liu Y, Niu M, Yao C, Hai Y, Yuan Q, et al. (2015) Fractionation of human spermatogenic cells using STA–PUT gravity sedimentation and their miRNA profiling. Scientific Reports 5: 8084.
  86. Jodar M, Kalko S, Castillo J, Ballescà JL, Oliva R (2012) Differential RNAs in the sperm cells of asthenozoospermic patients.Hum Reprod 27: 1431-1438.
  87. Plasterk RH (2002) RNA silencing: the genome's immune system.Science 296: 1263-1265.
  88. Zhou Y, Zheng M, Shi Q, Zhang L, Zhen W, et al. (2008) An epididymis-specific secretory protein HongrES1 critically regulates sperm capacitation and male fertility. PLoS One 3: e4106.
  89. Zhao Y1, Diao H, Ni Z, Hu S, Yu H, et al. (2011) The epididymis-specific antimicrobial peptide β-defensin 15 is required for sperm motility and male fertility in the rat (Rattusnorvegicus).Cell Mol Life Sci 68: 697-708.
  90. Klattenhoff C1, Theurkauf W (2008) Biogenesis and germline functions of piRNAs.Development 135: 3-9.
  91. Kazazian HH Jr (2004) Mobile elements: drivers of genome evolution. Science 303: 1626-32.
  92. Carmell MA, Girard A, van de Kant HJ, Bourc'his D, Bestor TH, et al. (2007) MIWI2 is essential for spermatogenesis and repression of transposons in the mouse male germline.Dev Cell 12: 503-514.
  93. Deng W, Lin H (2002) miwi, a murine homolog of piwi, encodes a cytoplasmic protein essential for spermatogenesis. Dev Cell 2: 819-30.
  94. Kuramochi-Miyagawa S, Kimura T, Yomogida K, Kuroiwa A, Tadokoro Y, et al. (2008) DNA methylation of retrotransposon genes is regulated by Piwi family members MILI and MIWI2 in murine fetal testes. Genes Dev 22: 908-917.
  95. Mercer TR, Dinger ME, Mattick JS (2009) Long non-coding RNAs: insights into functions.Nat Rev Genet 10: 155-159.
  96. Hung T, Wang Y, Lin MF, Koegel AK, Kotake Y, et al. (2011) Extensive and coordinated transcription of noncoding RNAs within cell-cycle promoters. Nature Genetics 43: 621-629.
  97. Berghoff EG, Clark MF, Chen S, Cajigas I, Leib DE, et al. (2013) Evf2 (Dlx6as) lncRNA regulates ultraconserved enhancer methylation and the differential transcriptional control of adjacent genes. Development 140: 4407-4416.
  98. Keniry A, Oxley D, Monnier P, Kyba M, Dandolo L, et al. (2012) The H19 lincRNA is a developmental reservoir of miR-675 that suppresses growth and Igf1r. Nature Cell Biology 14: 659-665.
  99. Han ZB, Lan GC, Wu YG, Han D, Feng WG, et al. (2006) Interactive effects of granulosa cell apoptosis, follicle size, cumulus-oocyte complex morphology, and cumulus expansion on the developmental competence of goat oocytes: a study using the well-in-drop culture system. Reproduction 132: 749-58.
  100. Nevoral J, Orsák M, Klein P, Petr J, Dvoráková M, etal. (2015) Cumulus cell expansion, its role in oocyte biology and perspectives of measurement: A review ScientiaAgriculturaeBohemica 45: 212-225.
  101. Yao G, Liang M, Liang N, Yin M, Lü M, et al. (2014) MicroRNA-224 is involved in the regulation of mouse cumulus expansion by targeting Ptx3. Mol Cell Endocrinol 382: 244-53.
  102. Pan B, Toms D, Shen W, Li J (2015) MicroRNA-378 regulates oocyte maturation via the suppression of aromatase in porcine cumulus cells. Am J PhysiolEndocrinolMetab308:E525-34.
  103. Li C, Chen C, Chen L, Chen S, Li H, et al. (2016) BDNF-induced expansion of cumulus-oocyte complexes in pigs was mediated by microRNA-205. Theriogenology 85: 1476-82.
  104. Rosenkranz D, Han CT, Roovers EF, Zischler H, Ketting RF (2015) Piwi proteins and piRNAs in mammalian oocytes and early embryos: From sample to sequence. Genom Data 5: 309-313.
  105. Mayer W, Niveleau A, Walter J, Fundele R, Haaf T (2000) Demethylation of the zygotic paternal genome.Nature 403: 501-502.
  106. Guo F, Li X, Liang D, Li T, Zhu P, et al. (2014) Active and passive demethylation of male and female pronuclear DNA in the mammalian zygote.Cell Stem Cell 15: 447-458.
  107. Delaval K, Govin J, Cerqueira F,Rousseaux S, Khochbin S, et al. (2007) Differential histone modifications mark mouse imprinting control regions during spermatogenesis. The EMBO Journal 26: 720-729.
  108. Williamson CM, Blake A, Thomas S, Beechey CV, Hancock J, et al. (2013), MRC Harwell, Oxfordshire. World Wide Web Site - Mouse Imprinting Data and References.
  109. Edwards CA1, Ferguson-Smith AC (2007) Mechanisms regulating imprinted genes in clusters.CurrOpin Cell Biol 19: 281-289.
  110. Monk D, Wagschal A, Arnaud P, Müller PS, Parker-Katiraee L, et al. (2008) Comparative analysis of human chromosome 7q21 and mouse proximal chromosome 6 reveals a placental-specific imprinted gene, TFPI2/Tfpi2, which requires EHMT2 and EED for allelic-silencing. Genome research 18: 1270-1281.
  111. Kelsey G, Feil R (2013) New insights into establishment and maintenance of DNA methylation imprints in mammals. Philosophical Transactions of the Royal Society of London B: Biological Sciences 368: 1609.
  112. Ferguson-Smith AC (2011) Genomic imprinting: the emergence of an epigenetic paradigm.Nat Rev Genet 12: 565-575.
  113. Davis TL1, Yang GJ, McCarrey JR, Bartolomei MS (2000) The H19 methylation imprint is erased and re-established differentially on the parental alleles during male germ cell development.Hum Mol Genet 9: 2885-2894.
  114. Li JY, Lees-Murdock DJ, Xu GL, Walsh CP (2004) Timing of establishment of paternal methylation imprints in the mouse.Genomics 84: 952-960.
  115. Yoon BJ, Herman H, Sikora A, Smith LT, Plass C, , et al. (2002) Regulation of DNA methylation of Rasgrf1. Nat Genet.32: 90-6.
  116. Lin SP, Youngson N, Takada S, Seitz H, Reik W, et al. (2003) Asymmetric regulation of imprinting on the maternal and paternal chromosomes at the Dlk1-Gtl2 imprinted cluster on mouse chromosome 12. Nat Genet 35: 97-102.
  117. Bourc'his D, Xu GL, Lin CS, Bollman B, Bestor TH (2001) Dnmt3L and the establishment of maternal genomic imprints.Science 294: 2536-2539.
  118. Kaneda M, Okano M, Hata K, Sado T, Tsujimoto N, et al. (2004) Essential role for de novo DNA methyltransferase Dnmt3a in paternal and maternal imprinting.Nature 429: 900-903.
  119. Henckel A, Chebli K, Kota SK, Arnaud P, Feil R (2012) Transcription and histone methylation changes correlate with imprint acquisition in male germ cells. The EMBO journal 31: 606-615.
  120. An J, Zhang X, Qin J, Wan Y, Hu Y, et al. (2014) The histone methyltransferase ESET is required for the survival of spermatogonial stem/progenitor cells in mice. Cell death & disease 5: e1196.
  121. Nakamura T, Arai Y, Umehara H, Masuhara M, Kimura T, et al. (2007) PGC7/Stella protects against DNA demethylation in early embryogenesis. Nat Cell Biol 9: 64-71.
  122. Thorvaldsen JL, Duran KL, Bartolomei MS (1998) Deletion of the H19 differentially methylated domain results in loss of imprinted expression of H19 and Igf2. Genes Dev 12: 3693–3702.
  123. Bartolomei MS, Zemel S, Tilghman SM (1991) Parental imprinting of the mouse H19 gene.Nature 351: 153-155.
  124. Leighton PA, Saam JR, Ingram RS, Stewart CL, Tilghman SM (1995) An enhancer deletion affects both H19 and Igf2 expression. Genes Dev 9: 2079-89.
  125. Ferguson-Smith AC1, Sasaki H, Cattanach BM, Surani MA (1993) Parental-origin-specific epigenetic modification of the mouse H19 gene.Nature 362: 751-755.
  126. Rio Frio T, Bahubeshi A, Kanellopoulou C, Hamel N, Niedziela M, etal. (2011) DICER1 mutations in familial multinodular goiter with and without ovarian Sertoli-Leydig cell tumors. JAMA 305:68-77.
  127. Stark KL1, Xu B, Bagchi A, Lai WS, Liu H, et al. (2008) Altered brain microRNA biogenesis contributes to phenotypic deficits in a 22q11-deletion mouse model.Nat Genet 40: 751-760.
  128. Berdasco M1, Esteller M (2013) Genetic syndromes caused by mutations in epigenetic genes.Hum Genet 132: 359-383.
  129. Hirasawa R, Feil R (2010) Genomic imprinting and human disease.Essays Biochem 48: 187-200.
  130. Tycko B, Morison IM (2002) Physiological functions of imprinted genes.J Cell Physiol 192: 245-258.
  131. Paulsen M, Takada S, Youngson NA, Benchaib M, Charlier C, etal. (2001) Comparative sequence analysis of the imprinted Dlk1-Gtl2 locus in three mammalian species reveals highly conserved genomic elements and refines comparison with the Igf2-H19 region. Genome Res 11: 2085-94.
  132. Sutovsky P, Schatten G (2000) Paternal contributions to the mammalian zygote: fertilization after sperm-egg fusion. Int Rev Cytol. 195:1-65.
  133. van der Heijden GW, Ramos L, Baart EB, van den Berg IM, Derijck AA, et al. (2008) Sperm-derived histones contribute to zygotic chromatin in humans. BMC DevBiol 8:34.
  134. Dogan S, Vargovic P, Oliveira R, Belser LE, Kaya A, et al. (2015) Sperm protamine-status correlates to the fertility of breeding bulls.BiolReprod 92: 92.
  135. Sills ES (2015) Screening for the Single Euploid Embryo - Molecular Genetics in Reproductive Medicine, Springer International Publishing.
  136. Eskandari-Shahraki M, Tavalaee M, Deemeh MR, Jelodar GA, Nasr-Esfahani MH (2013) Proper ubiquitination effect on the fertilisation outcome post-ICSI.Andrologia 45: 204-210.
  137. Odhiambo JF, DeJarnette JM, Geary TW, Kennedy CE, Suarez SS, et al. (2014) Increased conception rates in beef cattle inseminated with nanopurified bull semen.BiolReprod 91: 97.
Citation: Štiavnická M, Álvarez OG, Nevoral J, Králícková M, Sutovsky P (2016) Key Features of Genomic Imprinting during Mammalian Spermatogenesis: Perspectives for Human assisted Reproductive Therapy: A Review. Anat Physiol 6:236.

Copyright: © 2016 Štiavnická M, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top