Hostname: page-component-586b7cd67f-t8hqh Total loading time: 0 Render date: 2024-11-24T16:39:21.616Z Has data issue: false hasContentIssue false

Alterations in the expression pattern of some epigenetic-related genes and microRNAs subsequent to oocyte cryopreservation

Published online by Cambridge University Press:  20 June 2023

Ali Shadmanesh*
Affiliation:
Reproductive Biotechnology in Veterinary, Islamic Azad University, Eqlid Branch, Iran
Hassan Nazari
Affiliation:
Research Institute of Animal Embryo Technology, Shahrekord University, Shahrekord, Iran
*
Corresponding author: Ali Shadmanesh; Emails: [email protected]; [email protected]
Rights & Permissions [Opens in a new window]

Summary

MicroRNAs (miRNAs) are small non-encoding RNAs that actively regulate biological and physiological processes, and play an important role in regulating gene expression in all cells, especially in most animal cells, including oocytes and embryos. The expression of miRNAs at the right time and place is crucial for the oocyte’s maturation and the embryo’s subsequent development. Although assisted reproductive techniques (ART) have helped to solve many infertility problems, they cause changes in the expression of miRNA and genes in oocytes and preimplantation embryos, and the effect of these changes on the future of offspring is unknown, and has caused concerns. The relevant genomic alterations commonly imposed on embryos during cryopreservation may have potential epigenetic risks. Understanding the biological functions of miRNAs in frozen maturated oocytes may provide a better understanding of embryonic development and a comparison of fertility conservation in female mammals. With the development of new techniques for genomic evaluation of preimplantation embryos, it has been possible to better understand the effects of ART. The results of various articles have shown that freezing of oocytes and the cryopreservation method are effective for the expression of miRNAs and, in some cases, cause changes in the expression of miRNAs and epigenetic changes in the resulting embryo. This literature review study aimed to investigate the effects of oocyte cryopreservation in both pre-maturation and post-maturation stages, the cryopreservation method and the type of cryoprotectants (CPA) used on the expression of some epigenetic-related genes and miRNAs.

Type
Review Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Introduction

Cryopreservation is a very important technique that exposes the oocyte, embryo and various cells to different physical and chemical conditions, and today is an essential assisted reproductive technology (ART). According to studies, vitrification affects oocyte physiology to a lesser extent than slow freezing (Kuwayama et al., Reference Kuwayama, Vajta, Ieda and Kato2005; Cao et al., Reference Cao, Xing, Zhang, Wei, Zhou and Cong2009; Borini and Bianchi, Reference Borini and Bianchi2010). In general, survival and fertilization, as well as the rate of cleavage and implantation of embryos, were higher in vitrified human mature oocytes than in slowly frozen counterparts (Dessolle et al., Reference Dessolle, de Larouzière, Ravel, Berthaut, Antoine and Mandelbaum2009). However, high concentrations of cryoprotectants and rapid freezing rates may affect ultrastructure such as spindles and chromosomes (Trapphoff et al., Reference Trapphoff, Heiligentag, Simon, Staubach, Seidel, Otte, Fröhlich, Arnold and Eichenlaub-Ritter2016), mitochondrial function (Nohales-Córcoles et al., Reference Nohales-Córcoles, Sevillano-Almerich, Di Emidio, Tatone, Cobo, Dumollard and De los Santos Molina2016), gene expression patterns, microRNAs (miRNAs), and epigenetic stabilities of the oocytes and the resulting embryos (Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012). A study of the Danish population showed an increased risk of childhood cancer after frozen/thawed embryo transfer (FET) in a population of 174,881 children conceived through ART compared with 910,291 children naturally conceived (Hargreave et al., Reference Hargreave, Jensen, Hansen, Dehlendorff, Winther, Schmiegelow and Kjær2019). Data from four counties (Denmark, Finland, Norway, and Sweden) showed that children born after FET had a higher risk of childhood cancer than children born after fresh embryo transfer and spontaneous conception (Sargisian et al., Reference Sargisian, Lannering, Petzold, Opdahl, Gissler, Pinborg, Henningsen, Tiitinen, Romundstad, Spangmose, Bergh and Wennerholm2022). Therefore, it is important to determine the effect of cryopreservation protocols on miRNA and especially gene expression, as one of the negative modulators on epigenetics.

miRNAs, a subset of the non-coding RNA class, have been identified as central regulators of gene expression (Kim et al., Reference Kim, Kang, Kim, Lee and Lee2005). Studies have reported significant changes in miRNA expression patterns in frozen blastocysts (Zhao et al., Reference Zhao, Hao, Du and Zhu2015) and frozen sperm (Capra et al., Reference Capra, Turri, Lazzari, Cremonesi, Gliozzi, Fojadelli, Stella and Pizzi2017). Conversely, suppression of miR-762 has been reported to be beneficial for the survival and developmental capacity of frozen oocytes (Wang et al., Reference Wang, Zhang, Zhao, Kuang and Xue2018). One study found that the miR-21 level in pig MII oocytes was enhanced compared with germinal vesicle (GV) oocytes (Song et al., Reference Song, Peng, Chen, Jin, Yao, Shi, Yang, Zhang and Sun2016). and also reported a significant increase in miR-21 levels in in vitro maturation of vitrified bovine oocytes, although the pre-miR-21 transcript was not statistically changed between immature and in vitro-matured oocytes (Tscherner et al., Reference Tscherner, Brown, Stalker, Kao, Dufort, Sirard and LaMarre2018). Several pieces of molecular evidence on the effect of cryopreservation on oocyte gene expression profile have been presented (Liu et al., Reference Liu, He and Rosenwaks2003; Lee et al., Reference Lee, Li, Lu, Ho, Chen and Yeh2008; Anchamparuthy et al., Reference Anchamparuthy, Pearson and Gwazdauskas2010; Di Pietro et al., Reference Di Pietro, Vento, Guglielmino, Borzì, Santonocito, Ragusa, Barbagallo, Duro, Majorana, De Palma, Garofalo, Minutolo, Scollo and Purrello2010; Chamayou et al., Reference Chamayou, Bonaventura, Alecci, Tibullo, Di Raimondo, Guglielmino and Barcellona2011). Most of this evidence is connected to mammalian species, and has shown that, in metaphase II (MII) oocytes, cryopreservation mainly affected the expression of genes connected with oxidative stress (members of the heat shock protein family, superoxide dismutase 1), apoptosis (members of the BCL2 family, death receptors) and cell cycle (cyclin B, members of the histone family, polymerases; Liu et al., Reference Liu, He and Rosenwaks2003; Succu et al., Reference Succu, Leoni, Berlinguer, Madeddu, Bebbere, Mossa, Bogliolo, Ledda and Naitana2007; Lee et al., Reference Lee, Li, Lu, Ho, Chen and Yeh2008; Anchamparuthy et al., Reference Anchamparuthy, Pearson and Gwazdauskas2010). Some studies did not report the negative effects of cryopreservation at the molecular level (such as DNA damage, altered gene expression and miRNAs) in pigs, humans (Di Pietro et al., Reference Di Pietro, Vento, Guglielmino, Borzì, Santonocito, Ragusa, Barbagallo, Duro, Majorana, De Palma, Garofalo, Minutolo, Scollo and Purrello2010), mice (Liu et al., Reference Liu, He and Rosenwaks2003; Gao et al., Reference Gao, Jia, Li, Ma, Huang, Zhu, Hou and Fu2017), and canine (Turathum et al., Reference Turathum, Saikhun, Sangsuwan and Kitiyanant2010), However other studies have shown signs of adverse effects in ovine (Succu et al., Reference Succu, Leoni, Berlinguer, Madeddu, Bebbere, Mossa, Bogliolo, Ledda and Naitana2007) murine (Lee et al., Reference Lee, Li, Lu, Ho, Chen and Yeh2008) bovine (Wang et al., Reference Wang, Li, Zhu, Hao, Wang, Yan, Zhao, Du, Wang, Liu, Pang and Zhao2017a; Anchamparuthy et al., Reference Anchamparuthy, Pearson and Gwazdauskas2010) and human (Chamayou et al., Reference Chamayou, Bonaventura, Alecci, Tibullo, Di Raimondo, Guglielmino and Barcellona2011; Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012; Shahedi et al., Reference Shahedi, Hosseini, Ali Khalili and Yeganeh2017) species. It is apparent that these adverse effects might subsequently be corrected or repaired during further growth and development (Succu et al., Reference Succu, Leoni, Berlinguer, Madeddu, Bebbere, Mossa, Bogliolo, Ledda and Naitana2007). Any changes in gene expression because of cryopreservation may explain some of the differences in viability between fresh and post-thaw cells. Moreover, it has been shown that changes in gene expression may also lead to significant defects in the brain, ear, eye and kidneys. In a study by Azizi et al. (Reference Azizi, Ghaffari Novin, Naji, Amidi, Hosseinirad and Shams Mofarahe2021) on mouse embryos it was shown that let-7a expression significantly decreased in IVF-driven blastocysts versus in vivo ones (Azizi et al., Reference Azizi, Ghaffari Novin, Naji, Amidi, Hosseinirad and Shams Mofarahe2021). Zhao et al. (Reference Zhao, Hao, Du and Zhu2015) showed that vitrification of mouse blastocysts resulted in significant changes in the miRNA transcriptome, which may affect the implantation potential of vitrified blastocysts. In one study, a significant decrease in the expression of miR-16, as well as a significant increase in the expression level of the BCL2 gene (as one of the targets of miR-16), was reported in the vitrified and re-vitrified human blastocysts (groups) compared with the fresh group (Daneshvar et al., Reference Daneshvar, Movahedin, Salehi and Noruzinia2021). The aim of this review was to discuss the effects of cryopreservation on miRNA and the expression of some genes in oocytes and embryos of vertebrate species.

Oocyte cryopreservation

Effect of method cryopreservation

In recent years, many studies have revealed the superiority of vitrification over slow freeze protocols in reducing chilling damage to oocytes during cryopreservation (Saragusty and Arav, Reference Saragusty and Arav2011). Fadini et al. (Reference Fadini, Brambillasca, Renzini, Merola, Comi, De Ponti and Dal Canto2009) reported higher pregnancy rates for vitrification (18.2 vs. 7.6%). In addition, vitrification has been proposed as a cheaper and more reasonable strategy for the cryopreservation of human and animal oocytes instead of the slow freezing technique using a higher concentration of cryoprotectants and a rapid cooling rate to prevent the formation of intracellular ice crystals (Antinori et al., Reference Antinori, Licata, Dani, Cerusico, Versaci and Antinori2007). Furthermore, the vitrification method has also been applied to cryopreserve oocytes or embryos produced from IVM programmes (Mogas, Reference Mogas2019). Unlike the slow freezing method, vitrification presents the capacity to control solute penetration, the dehydration rate, and the maintenance of physiological temperature during the equilibration process. Breaking the zona pellucida and formation of growing ice crystals are possible during the slow freezing process, but not with vitrification (Alcolak et al., Reference Alcolak, Abu Marar, Mytas, Chalvatzas, Palapelas, Schöpper, Diedrich and Al-Hasani2011).

Results showed that the addition of melatonin to the freezing medium can effectively improve the development ability of frozen human oocytes by increasing mitochondrial function, reducing oxidative damage (by expressing proteins related to the phosphorylation pathway), maintaining reactive oxygen species (ROS)/GSH homeostasis and, as a result, normal ATP production and the ROS/GSH ratio were improved (Zhu et al., Reference Zhu, Ding, Yang, Zou, Yang, Wang, Zhang, Chen, Ji, Hao, Xue, Xu, Wang, Wang, Yan, Cao, Zou and Zhang2022). Limited numbers of studies on oocytes have provided controversial results regarding the effect on DNA fragmentation, sister chromatid exchange (SCE), and aneuploidy. A study of human embryos has shown that vitrification affects DNA integrity much less than slow freezing. Animal studies have shown increased mitochondrial DNA mutations in embryos after cryopreservation (Kopeika et al., Reference Kopeika, Thornhill and Khalaf2015). Studies have shown that slow freezing and vitrification alter the distribution of mitochondria in mouse embryos. Also, in the vitrification method, ROS are higher in the morula stages. Therefore, in addition to the freezing method, the embryonic stage also determines the changes in the cells caused by freezing (Somoskoi et al., Reference Somoskoi, Martino, Cardone, Lacalandra, Dell’Aquila and Cseh2015; Hayashi et al., Reference Hayashi, Kansaku, Abe, Ueda and Iwata2019). A study on rabbit embryos showed that there was no significant difference in the expression of genes (SCGB1A1, EMP1, C1QTNF1, ANXA3, EGFLAM and TNFAIP6) between embryos frozen with the two techniques (slow freezing and vitrification; Saenz-de-Juano et al., Reference Saenz-de-Juano, Marco-Jimenez, Viudes-de-Castro, Lavara and Vicente2014). A comparison between the transcriptional profiles of two vitrified oocyte techniques (semi-automated and manual vitrification) using single-cell RNA sequencing showed that the transcriptional changes between the two techniques were very low and not significant, but specific adverse effects of oocyte freezing on genomic expression were prominent compared with fresh oocytes. Given that fertility preservation through egg freezing will be developed worldwide in the coming years, More studies in this field are needed for the health of children born from frozen oocytes (Barberet et al., Reference Barberet, Ducreux, Bruno, Guilleman, Simonot, Lieury, Guilloteau, Bourc’his and Fauque2022).

Type of CPA and miRNAs

Results of the assessment of IVM bovine oocytes submitted to vitrification showed that DMSO had a significant effect on the incidence of abnormal microtubule distribution when compared with fresh oocytes (Verheijen et al., Reference Verheijen, Lienhard, Schrooders, Clayton, Nudischer, Boerno, Timmermann, Selevsek, Schlapbach, Gmuender, Gotta, Geraedts, Herwig, Kleinjans and Caiment2019). Frequencies found when comparing the DMSO group to glycerol were similar (P > 0.05). Additionally, no differences were found when comparing fresh oocytes to those submitted to vitrification using the combination of ethylene glycol (EG) and glycerol (P > 0.05). No differences between oocytes vitrified using EG and glycerol and fresh oocytes were found. Conversely, oocytes vitrified using the combination of EG and glycerol presented a higher incidence of normal microtubule configuration. It has been widely suggested that the incubation process allows the reorganization of microtubules. The chromosomal arrangement was not affected by vitrification and the incubation process did not affect it (Verheijen et al., Reference Verheijen, Lienhard, Schrooders, Clayton, Nudischer, Boerno, Timmermann, Selevsek, Schlapbach, Gmuender, Gotta, Geraedts, Herwig, Kleinjans and Caiment2019). In a recent study, DMSO effects were assessed with complete transcriptome analysis (using ribodepleted total RNA sequencing and microRNA sequencing), whole-genome methylation profiling (using MeDIP-seq) and proteomics analysis (using mass spectrometry). The results demonstrated that DMSO could not be considered biologically inert, but induced significant changes in miRNAs and the epigenetic landscape in cells in culture even at very low concentrations, especially in the maturing cardiac model. Zhang et al. (Reference Zhang, Mu, Ding, Zou, Li, Chen, Leung, Chang, Zhu, Wang, Xue, Xu, Zou, Zhou, Wei and Cao2021) evaluated the effects of EG as a cryoprotectant in open vitrification medium in human oocytes. Global gene expression studies based on fragments per kilobase million (FPKM) cluster analyses showed that 3740 genes were upregulated and 956 genes were downregulated in vitrified donor oocytes vs. fresh donor oocytes.

Stage of oocyte cryopreservation

Despite the proven superiority of the vitrification method over slow freezing, there is still debate about the vitrification stage of the oocyte (before and after in vitro maturation). Some studies in species such as pigs (Martin and Zhang, Reference Martin and Zhang2007) and humans (Ma et al., Reference Ma, Pan, Montgomery, Olson and Schultz2012; Gao et al., Reference Gao, Jia, Li, Ma, Huang, Zhu, Hou and Fu2017) have reported that vitrified oocytes at the GV stage have a lower survival rate and efficiency than at the metaphase II stage (MII). Conversely, one study reported that vitrification of mouse oocytes in the MII stage was more suitable for transcription (Gao et al., Reference Gao, Jia, Li, Ma, Huang, Zhu, Hou and Fu2017). Also, a study using the same cohort of oocytes (oocytes from the same donors) reported that vitrified oocytes did not show any statistically significant differences in fertilization, embryo quality, or clinical outcome compared with fresh oocytes (Solé et al., Reference Solé, Santaló, Boada, Clua, Rodríguez, Martínez, Coroleu, Barri and Veiga2013). A study using RNA-seq showed that vitrification of in vitro-matured bovine oocytes caused changes in the expression of some genes (Wang et al., Reference Wang, Li, Zhu, Hao, Wang, Yan, Zhao, Du, Wang, Liu, Pang and Zhao2017a, Reference Wang, Zhang, Zhao, Kuang and Xue2018). Song et al. (Reference Song, Peng, Chen, Jin, Yao, Shi, Yang, Zhang and Sun2016) also used GV-stage and GVBD (MI)-stage oocytes generated from IVF cycles. A comparison between matured oocytes in vivo without vitrification with groups of vitrified oocytes before and after maturation in vitro showed that, after thawing, the cleavage rate in vitrified oocytes at the MII stage after IVM was higher than in vitrified immature oocytes before IVM.

One study evaluated the expression of DNMT1, apoptosis-related (Bcl-2 and Bax) and stress-related (Sod1 and HSP70) genes in fresh IVM oocytes, vitrified GV (vGV) and vitrification of in vitro-matured oocytes (vIVM). The results showed that the expression of Bcl-2, Bax, Sod1, and HSP70 genes in the vIVM group was significantly increased, and also the DNMT1 gene in both groups [vitrification before (vGV) and after (vIVM) maturation]. But the rate of expression was reduced in vitrified IVM (vIVM) oocytes (Shahedi et al., Reference Shahedi, Hosseini, Ali Khalili and Yeganeh2017). Therefore, vitrification after maturation caused more stress on the oocytes, and higher expression of stress-related genes in the oocyte. The major problem associated with freezing mature oocytes is meiotic spindle sensitivity to low temperatures and cryoprotectants. To solve this problem, GV oocytes were used instead of mature oocytes. However, poor maturation, fertilization, and embryo development were the main disadvantages associated with the cryostorage of immature oocytes (Walls et al., Reference Walls, Junk, Ryan and Hart2012).

One study found that vitrification acted as a stressor and altered the expression of these genes, thereby reducing the developmental competence of mouse zygotes (Borini and Bianchi, Reference Borini and Bianchi2010). In addition, it should be noted that various factors in the freezing process, such as high concentrations of cryoprotectants, freezing, and osmotic stress, changed the pattern of gene expression as well as the onset of apoptosis, which threatened the viability and development of oocytes (Men et al., Reference Men, Monson, Parrish and Rutledge2003). Men et al. (Reference Men, Monson, Parrish and Rutledge2003) showed that the degeneration mechanism of cryopreserved oocytes was apoptosis. The results showed that the percentage of oocyte maturation in the oocytes matured in vitro before vitrification was significantly higher than in immature oocytes, which were vitrified, and then matured in vitro. Cao et al. (Reference Cao, Xing, Zhang, Wei, Zhou and Cong2009) similarly announced that the oocyte maturation rate in oocytes that were vitrified at the GV stage and afterwards matured in vitro, compared with oocytes that were first matured in vitro and followed by vitrification, was significantly reduced.

Epigenetic effects of cryopreservation on oocytes (such as non-coding RNAs; miRNAs)

Vitrification of porcine oocytes prompted changes in histone H4 acetylation and histone H3 methylation of lysine 9 (Spinaci et al., Reference Spinaci, Vallorani, Bucci, Tamanini, Porcu and Galeati2012). Also, vitrification significantly reduced the expression of DNMT1 mRNA in mouse metaphase II oocytes (Zhao et al., Reference Zhao, Ren, Du, Hao, Wang, Qin, Liu and Zhu2013; Cheng et al., Reference Cheng, Fu, Zhang, Jia, Hou and Zhu2014). The results of experimental studies showed that the expression level of DNMT1 decreased in the vitrified IVM group. Decreased DNMT1 gene expression usually alters DNA methylation patterns, leading to impaired gene expression, genomic printing, and genome stabilization, ultimately leading to cell death. Various studies have shown that, following the freezing and thawing of oocytes, H3K9 methylation and H4K5 acetylation increased sharply. Usually, no H3K14 acetylation occurs in fresh or frozen MII oocytes. Vitrified oocytes are highly sensitive to changes in H3K9 and H4K5 (Yan et al., Reference Yan, Yan, Qiao, Zhao and Liu2010; Rasti and Vaquero, Reference Rasti and Vaquero2018). Histone methylation involves arginine and lysine residues and plays an important role in regulating transcription, genome integrity and epigenetic inheritance. Overall, the results showed that cryopreservation may be associated with changes in physiological epigenetic marks with putative long-term effects on the cells and/or their derivatives.

According to some reports, vitrification of oocytes did not affect zygote cleavage rates, but reduced the blastocyst rate. Chen et al. (Reference Chen, Zhang, Deng, Zou, Wang, Quan and Zhang2016) reported that the degrees of DNA methylation and H3K9me3 in vitrified oocytes and early cleavage embryos were lower than in the control group (non-vitrified group), albeit in the vitrified group the level of acH3K9 increased in the early cleavage stages. These changes in methylation may be due to changes in the expression of micoRNAs in vitrified oocytes. The expression levels of putative-imprinted genes (PEG10, XIST, and KCNQ1O1T) were upregulated in blastocysts. These epigenetic abnormalities may be partially explained by the altered expression of genes associated with epigenetic regulations. DNA methylation and H3K9 modification suggested that oocyte vitrification may excessively relax the chromosomes of oocytes and early cleavage embryos. Consequently, these epigenetic indexes can be used as damage markers of oocyte vitrification during early embryonic development (Chen et al., Reference Chen, Zhang, Deng, Zou, Wang, Quan and Zhang2016). Overall, methylation in bovine MII oocytes has been reported to decrease significantly after vitrification with EG and DMSO as cryoprotectants and 0.25-ml straws as a container (Hu et al., Reference Hu, Marchesi, Qiao and Feng2012). In mice, DNA methylation was also significantly decreased in vitrified MII oocytes and in 2-cell- to 8-cell-stage embryos compared with the control. The data indicated that vitrification of the oocyte seemed to always results in the reduction or loss of DNA methylation in the oocyte or embryo, which may lead to impaired embryonic development and qualities in later stages (Hu et al., Reference Hu, Marchesi, Qiao and Feng2012).

miRNAs

Specific characteristics of miRNAs in oocytes (oocyte maturation)

miRNAs are an important class of small non-coding RNAs of ∼22 nucleoids that act as effectors of post-transcriptional gene silencing, and also act as a potent regulator of gene expression in most biological systems and cellular contexts. In eukaryotic organisms, miRNA genes are highly conserved, affect specific biological processes such as amplification, differentiation, and signal transduction, and are widely involved in development, physiology, and pathology (Flynt and Lai, Reference Flynt and Lai2008; Sun and Lai, Reference Sun and Lai2013). miRNAs typically suppress the expression of ‘target’ genes by anti-sense base-pairing interactions with the 3′ untranslated region (3′UTR). An active miRNA molecule binds to the ‘target’ mRNA through Watson–Crick base pairs using an essential seven-nucleotide seed sequence (Sun and Lai, Reference Sun and Lai2013). Due to this short pairing, a miRNA has the potential to bind to hundreds of 3′UTR targets. Hundreds of miRNAs have been examined based on in silico analysis; miRNAs were initially reported to regulate ∼30% of the expressed mammalian genes (Sevignani et al., Reference Sevignani, Calin, Siracusa and Croce2006), which increased to more than 60% (Friedman et al., Reference Friedman, Farh, Burge and Bartel2009). Given the potential for these small molecules to dramatically alter cellular behaviour and function, it continues to be crucial to understand how miRNAs are themselves regulated. miRNA biosynthesis is a sequential process that provides ample opportunity for complex regulation.

General and specific studies of miRNA profiles have shown that changes in the expression pattern of biologically relevant miRNAs occur during the growth and development of mammalian oocytes (Tesfaye et al., Reference Tesfaye, Worku, Rings, Phatsara, Tholen, Schellander and Hoelker2009; Xu et al., Reference Xu, Wang, Ding, Li, Gu and Zhou2011). For example, miR-21 is a miRNA that potentially targets genes involved in fatty acid metabolism and fatty acid biosynthesis, and is expressed in different patterns during oocyte maturation (Song et al., Reference Song, Peng, Chen, Jin, Yao, Shi, Yang, Zhang and Sun2016). Inhibition of miR-21 by an anti-miR21 peptide nucleic acid leads to a reduced ratio of porcine oocytes to the MII stage and further embryo development (Wright et al., Reference Wright, Hale, Yang, Njoka and Ross2016). Similar results by inhibiting the activity of let-7c, miR-27a and miR-322 in mouse oocytes (Kim et al., Reference Kim, Ku, Kim, Liu, Chi, Kim, Choi, Kim and Moon2013), miR-15/16 in Xenopus oocytes (Wilczynska et al., Reference Wilczynska, Git, Argasinska, Belloc and Standart2016) and miR-378 in the porcine oocyte (Pan et al., Reference Pan, Toms, Shen and Li2015) have been reported. Conversely, increasing the expression of miR-224 in cumulus cells reduced the rate of oocyte maturation and blastocyst growth by regulating PTX3 expression (Li et al., Reference Li, Wang, Sheng and Wang2017).

The results showed that specific changes in miRNA expression were associated with significant changes in the function and morphology of aged oocytes. Experimental studies on the function of key proteins predicted by KEGG analysis and injection of miR-98 mimics or inhibitors have confirmed that miRNAs have stimulatory/inhibitory roles in post-ovulatory oocyte ageing (Wang et al., Reference Wang, Zhang, Zhu, Lian, Yuan, Gao, Luo and Tan2017b). Whereas miR-21 acts as an anti-apoptotic, miR-29b, miR-15a and miR-16, miR-128, and miR-98 act as pro-apoptotics (Xia et al., Reference Xia, Lu, Wang, Yang, Zhou and Huang2018). However, all these six apoptosis-related miRNAs significantly increased their expression from 18 to 24 h after hCG injection. The study showed that miR-98 upregulated the expression of caspase-3, which increased the release of Ca2+ from the endoplasmic reticulum into the oocyte, leading to increased spontaneous ovulation activity (SA). Therefore, the results suggested that increased oocytes in SA were associated with increased cytoplasmic Ca2+ (Zhang et al., Reference Zhang, Cui, Zhang, Zhang, Wang, Zhu, Jiao and Tan2014). Experimental results showed that suppressed expression of miR-224 in cumulus cells (CC) was essential for oocyte maturation. Overexpression of miR-224 stopped or delayed the progression of oocyte maturation, whereas inhibition of miR-224 induced oocyte development. As a result, overexpression of miRNA-224 led to the decreased expression of genes and proteins associated with oocyte maturation such as GDF9, BMP15, CX37, and ZP3 in CC and oocytes (Li et al., Reference Li, Wang, Sheng and Wang2017).

Oocyte cryopreservation and miRNAs

In general, growth and development in oocytes occurred as a result of various cellular and molecular changes during ovarian follicular development. These changes are controlled by endocrine and paracrine factors coordinated by the oocyte and surrounding somatic cells. These processes are regulated by the precise expression and interaction of many genes in different parts of the ovary (oocytes, granulosa, and theca cells) to achieve oocyte development (Bonnet et al., Reference Bonnet, Tran and Sirard2007) and are possibly regulated by miRNAs. Studies on miRNAs have shown that miRNAs are involved in the regulation of oocyte cross-communication and CC (Assou et al., Reference Assou, Al-Edani, Haouzi, Philippe, Lecellier, Piquemal, Commes, Aït-Ahmed, Dechaud and Hamamah2013). Therefore, altering the expression of miRNAs at different stages of oocyte development can have a negative effect on the maturation and competence of the oocyte and the resulting embryo.

In most mammals, the process of meiotic division in oocytes arrests at the stage of the dictate of prophase I until the resumption of meiosis. Despite the extensive transcript turnover in the oocyte, it is almost accepted that the meiotic maturity of the oocyte is transcriptionally quiescent. Therefore, the fertilization and early development of the embryo depend on the maternal transcripts that have accumulated during the development of the oocyte. This mother-to-zygotic transition occurs in different species at different times in the embryonic stage. Recent studies have detailed conflicting reports on the occurrence of transcription following the resumption of oocyte meiotic maturation. Increased expression of some transcripts during oocyte maturation has been reported between immature (GV) and in vivo matured oocytes (MII) in mice, cattle, and humans (Assou et al., Reference Assou, Anahory, Pantesco, Le Carrour, Pellestor, Klein, Reyftmann, Dechaud, De Vos and Hamamah2006). The comparison of results between GV and MII oocytes showed an increased expression of four miRNAs (hsa-miR-193a-5p, hsa-miR-297, hsa-miR-625, and hsa-miR-602), and decreased expression of 11 miRNAs (hsa-miR-888, hsa-miR-212, hsa-miR-662, hsa-miR-299–5p, hsa-miR-339–5p, hsa-miR-20a, hsa-miR-486–5p, hsa-miR-141, hsa-miR-768–5p, hsa-miR-376a, and hsa-miR-15a; Xu et al., Reference Xu, Wang, Ding, Li, Gu and Zhou2011). Expression of specific miRNAs in human oocytes showed dynamic changes during meiosis. High concentrations of FSH in IVM medium had the opposite effect on the expression of hsa-miR-15a and hsa-miR-20a. A study to confirm miRNA dynamic changes in bovine GV, MII oocytes, and presumptive zygotes (PZ) showed that bta-miR-155, bta-miR-222, bta-miR-21, bta-let-7d, bta-let-7i, and bta-miR-190a had progressive changes between stages from GV to PZ. pri-miR-155 and pri-miR-222 are not normally detected in GV oocytes, but pri-miR-155 is present in MII oocytes, indicating transcription during maturation. miR-148a was found in large amounts with stable expression at all stages (Gilchrist et al., Reference Gilchrist, Tscherner, Nalpathamkalam, Merico and LaMarre2016).

However, in the maturing oocyte, several significant pathways have the strong potential for repression by miRNAs that are expressed, including phosphatidyl inositol phosphate binding (miR-222), transcriptional regulatory region DNA binding (miR-21), and regulation of the protein kinase cascade activity (let-7d). Studies of miRNAs in other cells, such as cancer tissue, have shown that p53 is targeted by the let-7 family of miRNAs. p53 is known to be widely involved in cell cycle checkpoints and is present in early bovine embryos (Tomek and Smiljakovic, Reference Tomek and Smiljakovic2005). In addition, miR-155 targets the mRNA of inositol 5-phosphatase 1 (INPP5D) and reduces inositol 5-phosphate 1 (O’Connell et al., Reference O’Connell, Chaudhuri, Rao and Baltimore2009). A decrease in INPP5D has been shown to increase AKT activity, a pathway involved in bovine oocyte maturation (Yamanaka et al., Reference Yamanaka, Tagawa, Takahashi, Watanabe, Guo, Iwamoto, Yamashita, Saitoh, Kameoka, Shimizu, Ichinohasama and Sawada2009). The data showed that let-7 (Takahashi et al., Reference Takahashi, Nakaoka and Yamashita2012) and miR-27a (Kim et al., Reference Kim, Ku, Kim, Liu, Chi, Kim, Choi, Kim and Moon2013) regulate cell cycle signalling pathways. Studies have shown that miR-322 modulates ovarian development, folliculogenesis, and cell differentiation (Leem et al., Reference Leem, Han, Lee, Ha, Kwon, Ho, Kim, Tran, Bae and Kang2011; Kim et al., Reference Kim, Ku, Kim, Liu, Chi, Kim, Choi, Kim and Moon2013).

The results showed that several miRNAs were expressed in human oocytes and blastocysts, targeting key genes involved in DNA repair and cell cycle checkpoints. Most miRNAs are generally expressed at lower levels in the blastocyst than in the oocyte. The results of the correlation analysis showed that there was an inverse and direct relationship between miRNAs and their target mRNAs (Tulay et al., Reference Tulay, Naja, Cascales-Roman, Doshi, Serhal and SenGupta2015). This study investigated the expression of 10 mRNAs and 20 miRNAs in human oocytes and blastocysts. The results showed that all mRNAs studied and 11 miRNAs were expressed in both oocytes and blastocysts. However, mRNA expression levels in oocytes were higher relative to blastocysts. This result appeared normal because the oocyte must transfer enough mRNAs to support the resulting zygote’s early development and maintain survival and development until embryonic genome activation. The results of the molecular analysis showed that miRNA can be involved in various processes of oocyte maturation. The importance of miRNA in oocyte maturation has initially been demonstrated in knockdown or knockout studies, and it has become clear that Dicer is required for meiotic spindle integrity and completion of meiosis I (Murchison et al., Reference Murchison, Stein, Xuan, Pan, Zhang, Schultz and Hannon2007). Many factors, such as oocyte maturation (in vitro or in vivo), cryopreservation methods and cryopreservation procedures (slow freezing or vitrification), oocyte vitrification stages (GV or MII oocyte), maternal age etc., can adversely cause changes in the expression pattern of miRNAs in the oocyte (Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012; Wang et al., Reference Wang, Li, Zhu, Hao, Wang, Yan, Zhao, Du, Wang, Liu, Pang and Zhao2017a).

According to some studies, ∼520 miRNAs have been detected in oocytes, of which only 22 miRNAs were significantly expressed in vitrified oocytes compared with fresh oocytes (Li et al., Reference Li, Yang, Liu, Song and Liu2019). Changes in miRNA expression are dynamic during oocyte vitrification. The data showed that miR-21–3p expression increased after cryopreservation. It has been identified that miR-21–3p target genes are involved in physiological and pathological processes (Yan et al., Reference Yan, Chen, Gong, Yin, Zhou, Chaugai and Wang2015; Xia et al., Reference Xia, Lu, Wang, Yang, Zhou and Huang2018). This suggested that miR-21–3p suppressed PTEN gene expression in umbilical cord blood plasma exosomes (Hu et al., Reference Hu, Rao, Wang, Cao, Tan, Luo, Li, Zhang, Chen and Xie2018). It also proved that miR-21–3p reduced PTEN gene expression by targeting its 3′UTR. PTEN always acts towards the PI3K/AKT pathway, important signalling in regulating cell apoptosis through the regulation of oxidative stress (Matsuda et al., Reference Matsuda, Nakagawa, Kitagishi, Nakanishi and Murai2018). Studies have shown that decreased PTEN expression is associated with increased levels of ROS (Li et al., Reference Li, Mao, Zheng, Dong and Ling2013; Noh et al., Reference Noh, Park, Song, Kim, Lee, Song, Hong, Whang, Han, Kwon, Kim and Lee2016), which acts as a factor in reducing the developmental potential of vitrified oocytes (Pan et al., Reference Pan, Toms, Shen and Li2015; Wang et al., Reference Wang, Zhang, Zhao, Kuang and Xue2018). In the study, ultrastructural malformations such as increased vacuolation, aberrant dynamic variations in mitochondria–smooth endoplasmic reticulum and mitochondria–vesicle complexes, and scarce cortical granules were observed in vitrified oocytes (Nottola et al., Reference Nottola, Albani, Coticchio, Palmerini, Lorenzo, Scaravelli, Borini, Levi-Setti and Macchiarelli2016). In addition, disturbed mitochondrial localization has been reported in vitrified mouse oocytes (Yan et al., Reference Yan, Yan, Qiao, Zhao and Liu2010, Reference Yan, Chen, Gong, Yin, Zhou, Chaugai and Wang2015). Also, vitrification causes damage to microtubules, actin filaments, and chromosome integrity in the oocytes (Wen et al., Reference Wen, Zhao, Chao, Yu, Song, Shen, Chen and Deng2014).

Studies have shown that storage duration did not affect the gene expression pattern in frozen oocytes and did not alter their expression (Goldman et al., Reference Goldman, Kramer, Hodes-Wertz, Noyes, McCaffrey and Grifo2015; Stigliani et al., Reference Stigliani, Moretti, Anserini, Casciano, Venturini and Scaruffi2015). Oocyte developmental competence has been shown to depend on the accumulation of maternal RNAs and proteins during oogenesis. Conversely, reduced developmental competence in the oocyte is one of the main reasons for IVF failure. Therefore, this proves that the developmental potential of frozen/thawed oocytes is not related to their retention time in liquid nitrogen. Therefore, the possible damage to frozen/thawed oocytes is only due to the method and procedure of cryopreservation, which seems to be more important than the storage time. In support of this, clinical studies have been previously performed on controlled-rate cryopreserved and vitrified human oocytes (Goldman et al., Reference Goldman, Kramer, Hodes-Wertz, Noyes, McCaffrey and Grifo2015).

Under in vitro conditions, the dynamic miRNA profile changes are partly attributed to the in vitro maturation environment or ingredients used under in vivo conditions. The miRNA profiles could be affected by physiological conditions such as the animal’s age. For example, in humans, treatment of metaphase I (MI) human oocytes with insulin-like growth factor 1 activated the expression of miR-133a, miR-205–5p and 145 miRNAs and suppressed 200 others, including miR-152 and miR-142–5p (Xiao et al., Reference Xiao, Xia, Yang, Liu, Du, Kang, Lin, Guan, Yan and Tang2014). Conversely, altered expression of 12 miRNAs (including let-7b-5p and let-7e-5p) in oocytes derived from older women compared with young women suggested that the in vivo miRNA profile in the oocytes was affected by oocyte ageing (Battaglia et al., Reference Battaglia, Vento, Ragusa, Barbagallo, La Ferlita, Di Emidio, Borzí, Artini, Scollo, Tatone, Purrello and Di Pietro2016).

A study in mice showed that, although the expression pattern of the miRNA biogenesis pathway varied between 8-cell embryos and blastocysts, vitrification did not affect the expression level of these genes in preimplantation embryos. Expression levels of miR-21 and let-7a were significantly reduced in vitrified 8-cell embryos and fresh blastocysts compared with fresh 8-cell embryos. STAT-3 expression in blastocysts decreased significantly after vitrification (Azizi et al., Reference Azizi, Ghaffari Novin, Naji, Amidi, Hosseinirad and Shams Mofarahe2021). However, changes in the expression pattern of miRNAs in the oocyte and multicellular embryos were compensated in the blastocyst stage. These changes were very small in vitrified and fresh embryos in the blastocyst stage. The results indicated that re-vitrification of embryos changed the expression of miR-16, miR-let-7a and their target genes. These changes caused increased expression of BCl-2 and ITGβ3 genes which play important roles in embryo survival and implantation, respectively (Daneshvar et al., Reference Daneshvar, Movahedin, Salehi and Noruzinia2021).

Method of preservation (slow freezing or vitrification) and miRNAs

Vitrification is relatively simple, does not require expensive programmable freezing equipment, and uses a small amount of liquid nitrogen for freezing. In addition, cryopreservation of oocytes using vitrification has been suggested to maintain female fertility by servicing and freezing their oocytes at optimal times. Experimental data showed that both cryopreservation methods (slow freezing and glass vitrification) had a negative effect on miRNAs and the pattern and gene expression of human MII oocytes compared with non-frozen MII oocytes, but the effects of vitrification on oocyte physiology (developmental competence of oocyte) was less in slow freezing (Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012; Quan et al., Reference Quan, Wu and Hong2017). However, slow-frozen and vitrified MI oocytes showed signatures for specific gene expression. Slow freezing with downregulation of genes that maintain chromosomal structure (KIF2C and KIF3A) and cell cycle regulation (CHEK2 and CDKN1B) may reduce oocyte developmental competence. In vitrified oocytes, many genes of the ubiquitination pathway are downregulated (including members of the ubiquitin-specific peptidase family and subunits of the 26S proteasome; Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012). The controlled-rate cryopreservation method (slow freezing) seriously alters specific transcript levels, leading to loss of transcript content (Chamayou et al., Reference Chamayou, Bonaventura, Alecci, Tibullo, Di Raimondo, Guglielmino and Barcellona2011; Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012; Stigliani et al., Reference Stigliani, Moretti, Anserini, Casciano, Venturini and Scaruffi2015).

Experimental data showed that the expression profiles of cryopreserved MII oocytes significantly differed from those of non-cryopreserved oocytes in 107 probe sets corresponding to 73 downregulated and 29 upregulated unique transcripts. Gene Ontology analysis using the DAVID bioinformatics resource disclosed that cryopreservation deregulates genes involved in oocyte function and early embryo development (such as chromosome organization, RNA splicing and processing, cell cycle, cellular response to DNA damage and to stress, DNA repair, calcium ion binding, malate dehydrogenase activity, and mitochondrial activity; Stigliani et al., Reference Stigliani, Moretti, Anserini, Casciano, Venturini and Scaruffi2015). A study on mice indicated that the influence of vitrification on the transcriptome of oocytes was negligible as no differentially expressed genes were found between vitrified and fresh oocytes. The MII stage is more suitable for oocyte vitrification with respect to the transcriptome (Gao et al., Reference Gao, Jia, Li, Ma, Huang, Zhu, Hou and Fu2017).

Comparison of IVF results from slow-frozen and vitrified oocytes showed that vitrification improved survival, fertilization, and pregnancy rates (Cao et al., Reference Cao, Xing, Zhang, Wei, Zhou and Cong2009; Fadini et al., Reference Fadini, Brambillasca, Renzini, Merola, Comi, De Ponti and Dal Canto2009), although only Fadini et al. (Reference Fadini, Brambillasca, Renzini, Merola, Comi, De Ponti and Dal Canto2009) reported significantly higher pregnancy rates (18.2 vs. 7.6%). Increasing evidence on the effectiveness of IVF with vitrified oocytes showed that it could achieve similar results to IVF using fresh oocytes, with an oocyte survival rate of more than 84% (Rienzi et al., Reference Rienzi, Cobo, Paffoni, Scarduelli, Capalbo, Vajta, Remohí, Ragni and Ubaldi2012). A meta-analysis of five studies showed that the fertilization rate, embryo cleavage, high-quality embryos and ongoing pregnancy progress did not differ between the vitrification and fresh oocyte groups (Cobo and Diaz, Reference Cobo and Diaz2011). Some IVF programmes now use vitrification to cryopreserve oocytes (Brison et al., Reference Brison, Cutting, Clarke and Wood2012; Glujovsky et al., Reference Glujovsky, Riestra, Sueldo, Fiszbajn, Repping, Nodar, Papier and Ciapponi2014).

Applications of miRNAs in assisted reproduction technology

Due to the important role of miRNAs in the suppression or expression of post-translational mRNAs, in addition to treating some diseases, they can be used to improve the maturity of oocytes, in vitro-produced embryos and determine the non-invasive quality of oocytes. Data showed that increased expression of HDAC6 by miR-762 suppression may improve the current protocol for oocyte vitrification. Overexpression of HDAC6 or reduction of miR-762 improved the survival rate, cleavage rate, and blastocyst rate of oocytes after oocyte vitrification (Wang et al., Reference Wang, Zhang, Zhao, Kuang and Xue2018). The results showed that fresh ovine embryos expressed the HDAC1 gene at a higher level than their vitrified counterpart embryos (Bahr et al., Reference Bahr, Robey, Luchenko, Basseville, Chakraborty, Kozlowski, Pauly, Patel, Schneider, Gottesman and Bates2016). Histone acetylation is critical for proper cellular functions, including chromosome condensation, DNA double-stranded breakage repair, and mRNA transcription, but there is no evidence for the effect of HDAC6 on in vivo development of the blastocysts from cryopreserved oocytes. The interpretation of this issue will depend on future works.

Conclusion

Oocyte vitrification has been proven to be a successful approach for the long-term storage of oocytes and maintaining women’s fertility. Various molecular studies have shown that cryopreservation can alter the expression of miRNAs in vitrified oocytes compared with fresh oocytes. These changes can affect the expression of some genes and lead to physiological and metabolic changes in the vitrified oocytes and early embryo development, which can reduce the final competence of the frozen oocytes and the resulting embryos (Liu et al., Reference Liu, He and Rosenwaks2003; Di Pietro et al., Reference Di Pietro, Vento, Guglielmino, Borzì, Santonocito, Ragusa, Barbagallo, Duro, Majorana, De Palma, Garofalo, Minutolo, Scollo and Purrello2010; Gao et al., Reference Gao, Jia, Li, Ma, Huang, Zhu, Hou and Fu2017). However, the rate of change resulting from cryopreservation is much lower than in vitro maturation of oocytes. Conversely, several studies have reported that cryopreservation did not alter gene and miRNA expression (Anchamparuthy et al., Reference Anchamparuthy, Pearson and Gwazdauskas2010; Monzo et al., Reference Monzo, Haouzi, Roman, Assou, Dechaud and Hamamah2012; Stigliani et al., Reference Stigliani, Moretti, Anserini, Casciano, Venturini and Scaruffi2015; Shahedi et al., Reference Shahedi, Hosseini, Ali Khalili and Yeganeh2017; Wang et al., Reference Wang, Li, Zhu, Hao, Wang, Yan, Zhao, Du, Wang, Liu, Pang and Zhao2017a). Therefore, further research is needed to investigate the expression of genes and miRNAs in vitrified oocytes and their embryos. It seems that these changes in the expression of miRNAs and genes can be compensated for in other stages of development, because these minor changes may be resolved due to ‘embryo adaptability’ and may not significantly affect embryo competence.

Availability of data and material

All data searched in this study are included in this publication.

Acknowledgements

We are thankful to the anonymous reviewers for their critical reading of the manuscript and suggestions for improvement.

Funding

This research received no specific grant from any funding agency, commercial or not-for-profit sectors.

Code availability

Not applicable.

Consent to participate

Not applicable.

Competing interests

The authors declare no conflict of interest. All authors approved the final manuscript for publication.

Ethics approval

This is a literature review article that examines past studies and the references used in the text.

References

Alcolak, E., Abu Marar, E., Mytas, S. C., Chalvatzas, N., Palapelas, V., Schöpper, B., Diedrich, K. and Al-Hasani, S. (2011) Comparison of two different media for vitrification and rewarming of human zygotes: Prospective randomized study. Middle East Fertility Society Journal, 16(3), 189193. doi: 10.1016/j.mefs.2011.02.003 CrossRefGoogle Scholar
Anchamparuthy, V. M., Pearson, R. E. and Gwazdauskas, F. C. (2010). Expression pattern of apoptotic genes in vitrified-thawed bovine oocytes. Reproduction in Domestic Animals, 45(5), e83e90. doi: 10.1111/j.1439-0531.2009.01527.x Google ScholarPubMed
Antinori, M., Licata, E., Dani, G., Cerusico, F., Versaci, C. and Antinori, S. (2007). Cryotop vitrification of human oocytes results in high survival rate and healthy deliveries. Reproductive Biomedicine Online, 14(1), 7279. doi: 10.1016/s1472-6483(10)60766-3 CrossRefGoogle ScholarPubMed
Assou, S., Anahory, T., Pantesco, V., Le Carrour, T., Pellestor, F., Klein, B., Reyftmann, L., Dechaud, H., De Vos, J. and Hamamah, S. (2006). The human cumulus–oocyte complex gene-expression profile. Human Reproduction, 21(7), 17051719. doi: 10.1093/humrep/del065 CrossRefGoogle ScholarPubMed
Assou, S., Al-Edani, T., Haouzi, D., Philippe, N., Lecellier, C. H., Piquemal, D., Commes, T., Aït-Ahmed, O., Dechaud, H. and Hamamah, S. (2013). MicroRNAs: New candidates for the regulation of the human cumulus–oocyte complex. Human Reproduction, 28(11), 30383049. doi: 10.1093/humrep/det321 CrossRefGoogle ScholarPubMed
Azizi, E., Ghaffari Novin, M., Naji, M., Amidi, F., Hosseinirad, H. and Shams Mofarahe, Z. (2021). Effect of vitrification on biogenesis pathway and expression of development-related microRNAs in preimplantation mouse embryos. Cell and Tissue Banking, 22(1), 103114. doi: 10.1007/s10561-020-09870-z CrossRefGoogle ScholarPubMed
Bahr, J. C., Robey, R. W., Luchenko, V., Basseville, A., Chakraborty, A. R., Kozlowski, H., Pauly, G. T., Patel, P., Schneider, J. P., Gottesman, M. M. and Bates, S. E. (2016). Blocking downstream signaling pathways in the context of HDAC inhibition promotes apoptosis preferentially in cells harboring mutant Ras. Oncotarget, 7(43), 6980469815. doi: 10.18632/oncotarget.12001 CrossRefGoogle ScholarPubMed
Barberet, J., Ducreux, B., Bruno, C., Guilleman, M., Simonot, R., Lieury, N., Guilloteau, A., Bourc’his, D. and Fauque, P. (2022). Comparison of oocyte vitrification using a semi-automated or a manual closed system in human siblings: Survival and transcriptomic analyses. Journal of Ovarian Research, 15(1), 128. doi: 10.1186/s13048-022-01064-3 CrossRefGoogle ScholarPubMed
Battaglia, R., Vento, M. E., Ragusa, M., Barbagallo, D., La Ferlita, A., Di Emidio, G., Borzí, P., Artini, P. G., Scollo, P., Tatone, C., Purrello, M. and Di Pietro, C. (2016). MicroRNAs are stored in human MII oocyte and their expression profile changes in reproductive ageing. Biology of Reproduction, 95(6), 131. doi: 10.1095/biolreprod.116.142711 CrossRefGoogle Scholar
Bonnet, A., Tran, R. D. and Sirard, M. (2007). Opportunities and challenges in applying genomics to the study of oogenesis and follicullogenesis in farm animals. 2. International Meeting on Mammalian Embryogenomics. doi:10.1530/rep-07–0331 CrossRefGoogle Scholar
Borini, A. and Bianchi, V. (2010). Cryopreservation of mature and immature oocytes. Clinical Obstetrics and Gynecology, 53(4), 763774. doi: 10.1097/GRF.0b013e3181f96f01 CrossRefGoogle ScholarPubMed
Brison, D., Cutting, R., Clarke, H. and Wood, M. (2012). ACE consensus meeting report. ACE consensus meeting report: Oocyte and embryo cryopreservation Sheffield 17.05.11. Human Fertility, 15(2), 6974. doi: 10.3109/14647273.2012.687124 CrossRefGoogle ScholarPubMed
Cao, Y., Xing, Q., Zhang, Z. G., Wei, Z. L., Zhou, P. and Cong, L. (2009). Cryopreservation of immature and in-vitro matured human oocytes by vitrification. Reproductive Biomedicine Online, 19(3), 369373. doi: 10.1016/s1472-6483(10)60170-8 CrossRefGoogle ScholarPubMed
Capra, E., Turri, F., Lazzari, B., Cremonesi, P., Gliozzi, T. M., Fojadelli, I., Stella, A. and Pizzi, F. (2017). Small RNA sequencing of cryopreserved semen from single bull revealed altered miRNAs and piRNAs expression between high- and low-motile sperm populations. BMC Genomics, 18(1), 14. doi: 10.1186/s12864-016-3394-7 CrossRefGoogle ScholarPubMed
Chamayou, S., Bonaventura, G., Alecci, C., Tibullo, D., Di Raimondo, F., Guglielmino, A. and Barcellona, M. L. (2011). Consequences of metaphase II oocyte cryopreservation on mRNA content. Cryobiology, 62(2), 130134. doi: 10.1016/j.cryobiol.2011.01.014 CrossRefGoogle ScholarPubMed
Chen, H., Zhang, L., Deng, T., Zou, P., Wang, Y., Quan, F. and Zhang, Y. (2016). Effects of oocyte vitrification on epigenetic status in early bovine embryos. Theriogenology, 86(3), 868878. doi: 10.1016/j.theriogenology.2016.03.008 CrossRefGoogle ScholarPubMed
Cheng, K. R., Fu, X. W., Zhang, R. N., Jia, G. X., Hou, Y. P. and Zhu, S. E. (2014). Effect of oocyte vitrification on deoxyribonucleic acid methylation of H19, Peg3, and Snrpn differentially methylated regions in mouse blastocysts. Fertility and Sterility, 102(4), 11831190.e3. e1183. doi: 10.1016/j.fertnstert.2014.06.037 CrossRefGoogle ScholarPubMed
Cobo, A. and Diaz, C. (2011). Clinical application of oocyte vitrification: A systematic review and meta-analysis of randomized controlled trials. Fertility and Sterility, 96(2), 277285. doi: 10.1016/j.fertnstert.2011.06.030 CrossRefGoogle ScholarPubMed
Daneshvar, M., Movahedin, M., Salehi, M. and Noruzinia, M. (2021). Alterations of miR-16, miR-let-7a and their target genes expression in human blastocysts following vitrification and re-vitrification. Reproductive Biology and Endocrinology: RB&E, 19(1), 155. doi: 10.1186/s12958-021-00842-w CrossRefGoogle ScholarPubMed
Dessolle, L., de Larouzière, V., Ravel, C., Berthaut, I., Antoine, J. M. and Mandelbaum, J. (2009). Slow freezing and vitrification of human mature and immature oocytes. Gynecologie, Obstetrique et Fertilite, 37(9), 712719. doi: 10.1016/j.gyobfe.2009.04.026 CrossRefGoogle ScholarPubMed
Di Pietro, C., Vento, M., Guglielmino, M. R., Borzì, P., Santonocito, M., Ragusa, M., Barbagallo, D., Duro, L. R., Majorana, A., De Palma, A., Garofalo, M. R., Minutolo, E., Scollo, P. and Purrello, M. (2010). Molecular profiling of human oocytes after vitrification strongly suggests that they are biologically comparable with freshly isolated gametes. Fertility and Sterility, 94(7), 28042807. doi: 10.1016/j.fertnstert.2010.04.060 CrossRefGoogle ScholarPubMed
Fadini, R., Brambillasca, F., Renzini, M. M., Merola, M., Comi, R., De Ponti, E. and Dal Canto, M. B. (2009). Human oocyte cryopreservation: Comparison between slow and ultrarapid methods. Reproductive Biomedicine Online, 19(2), 171180. doi: 10.1016/s1472-6483(10)60069-7 CrossRefGoogle ScholarPubMed
Flynt, A. S. and Lai, E. C. (2008). Biological principles of microRNA-mediated regulation: Shared themes amid diversity. Nature Reviews. Genetics, 9(11), 831842. doi: 10.1038/nrg2455 CrossRefGoogle ScholarPubMed
Friedman, R. C., Farh, K. K.-H., Burge, C. B. and Bartel, D. P. (2009). Most mammalian mRNAs are conserved targets of microRNAs. Genome Research, 19(1), 92105. doi: 10.1101/gr.082701.108 CrossRefGoogle ScholarPubMed
Gao, L., Jia, G., Li, A., Ma, H., Huang, Z., Zhu, S., Hou, Y. and Fu, X. (2017). RNA-Seq transcriptome profiling of mouse oocytes after in vitro maturation and/or vitrification. Scientific Reports, 7(1), 13245. doi: 10.1038/s41598-017-13381-5 CrossRefGoogle ScholarPubMed
Gilchrist, G. C., Tscherner, A., Nalpathamkalam, T., Merico, D. and LaMarre, J. (2016). MicroRNA expression during bovine oocyte maturation and fertilization. International Journal of Molecular Sciences, 17(3), 396. doi: 10.3390/ijms17030396 CrossRefGoogle ScholarPubMed
Glujovsky, D., Riestra, B., Sueldo, C., Fiszbajn, G., Repping, S., Nodar, F., Papier, S. and Ciapponi, A. (2014). Vitrification versus slow freezing for women undergoing oocyte cryopreservation. Cochrane Database of Systematic Reviews, 9(9), CD010047. doi: 10.1002/14651858.CD010047.pub2 Google Scholar
Goldman, K. N., Kramer, Y., Hodes-Wertz, B., Noyes, N., McCaffrey, C. and Grifo, J. A. (2015). Long-term cryopreservation of human oocytes does not increase embryonic aneuploidy. Fertility and Sterility, 103(3), 662668. doi: 10.1016/j.fertnstert.2014.11.025 CrossRefGoogle Scholar
Hargreave, M., Jensen, A., Hansen, M. K., Dehlendorff, C., Winther, J. F., Schmiegelow, K. and Kjær, S. K. (2019). Association between fertility treatment and cancer risk in children. JAMA, 322(22), 22032210. doi: 10.1001/jama.2019.18037 CrossRefGoogle ScholarPubMed
Hayashi, T., Kansaku, K., Abe, T., Ueda, S. and Iwata, H. (2019). Effects of resveratrol treatment on mitochondria and subsequent embryonic development of bovine blastocysts cryopreserved by slow freezing. Animal Science Journal, 90(7), 849856. doi: 10.1111/asj.13219 CrossRefGoogle ScholarPubMed
Hu, W., Marchesi, D., Qiao, J. and Feng, H. L. (2012). Effect of slow freeze versus vitrification on the oocyte: An animal model. Fertility and Sterility, 98(3), 752760.e3. e753. doi: 10.1016/j.fertnstert.2012.05.037 CrossRefGoogle ScholarPubMed
Hu, Y., Rao, S. S., Wang, Z. X., Cao, J., Tan, Y. J., Luo, J., Li, H. M., Zhang, W. S., Chen, C. Y. and Xie, H. (2018). Exosomes from human umbilical cord blood accelerate cutaneous wound healing through miR-21–3p-mediated promotion of angiogenesis and fibroblast function. Theranostics, 8(1), 169184. doi: 10.7150/thno.21234 CrossRefGoogle ScholarPubMed
Kim, S. S., Kang, H. G., Kim, N. H., Lee, H. C. and Lee, H. H. (2005). Assessment of the integrity of human oocytes retrieved from cryopreserved ovarian tissue after xenotransplantation. Human Reproduction, 20(9), 25022508. doi: 10.1093/humrep/dei099 Google ScholarPubMed
Kim, Y. J., Ku, S. Y., Kim, Y. Y., Liu, H. C., Chi, S. W., Kim, S. H., Choi, Y. M., Kim, J. G. and Moon, S. Y. (2013). MicroRNAs transfected into granulosa cells may regulate oocyte meiotic competence during in vitro maturation of mouse follicles. Human Reproduction, 28(11), 30503061. doi: 10.1093/humrep/det338 CrossRefGoogle ScholarPubMed
Kopeika, J., Thornhill, A. and Khalaf, Y. (2015). The effect of cryopreservation on the genome of gametes and embryos: Principles of cryobiology and critical appraisal of the evidence. Human Reproduction Update, 21(2), 209227. doi: 10.1093/humupd/dmu063 CrossRefGoogle ScholarPubMed
Kuwayama, M., Vajta, G., Ieda, S. and Kato, O. (2005). Comparison of open and closed methods for vitrification of human embryos and the elimination of potential contamination. Reproductive Biomedicine Online, 11(5), 608614. doi: 10.1016/s1472-6483(10)61169-8 CrossRefGoogle ScholarPubMed
Lee, R. K.-K., Li, S. H., Lu, C. H., Ho, H. Y., Chen, Y. J. and Yeh, H. I. (2008). Abnormally low expression of connexin 37 and connexin 43 in subcutaneously transplanted cryopreserved mouse ovarian tissue. Journal of Assisted Reproduction and Genetics, 25(9–10), 489497. doi: 10.1007/s10815-008-9264-8 CrossRefGoogle ScholarPubMed
Leem, Y. E., Han, J. W., Lee, H. J., Ha, H. L., Kwon, Y. L., Ho, S. M., Kim, B. G., Tran, P., Bae, G. U. and Kang, J. S. (2011). Gas1 cooperates with CDO and promotes myogenic differentiation via activation of p38MAPK. Cellular Signalling, 23(12), 20212029. doi: 10.1016/j.cellsig.2011.07.016 CrossRefGoogle ScholarPubMed
Li, P., Mao, W. M., Zheng, Z. G., Dong, Z. M. and Ling, Z. Q. (2013). Down-regulation of PTEN expression modulated by dysregulated miR-21 contributes to the progression of esophageal cancer. Digestive Diseases and Sciences, 58(12), 34833493. doi: 10.1007/s10620-013-2854-z CrossRefGoogle Scholar
Li, X., Wang, H., Sheng, Y. and Wang, Z. (2017). MicroRNA-224 delays oocyte maturation through targeting Ptx3 in cumulus cells. Mechanisms of Development, 143, 2025. doi: 10.1016/j.mod.2016.12.004 CrossRefGoogle ScholarPubMed
Li, J., Yang, X., Liu, F., Song, Y. and Liu, Y. (2019). Evaluation of differentially expressed microRNAs in vitrified oocytes by next generation sequencing. International Journal of Biochemistry and Cell Biology, 112, 134140. doi: 10.1016/j.biocel.2019.05.006 CrossRefGoogle ScholarPubMed
Liu, H. C., He, Z. and Rosenwaks, Z. (2003). Mouse ovarian tissue cryopreservation has only a minor effect on in vitro follicular maturation and gene expression. Journal of Assisted Reproduction and Genetics, 20(10), 421431. doi: 10.1023/a:1026284609730 CrossRefGoogle Scholar
Ma, P., Pan, H., Montgomery, R. L., Olson, E. N. and Schultz, R. M. (2012). Compensatory functions of histone deacetylase 1 (HDAC1) and HDAC2 regulate transcription and apoptosis during mouse oocyte development. Proceedings of the National Academy of Sciences of the United States of America, 109(8), E481E489. doi: 10.1073/pnas.1118403109 Google ScholarPubMed
Martin, C. and Zhang, Y. (2007). Mechanisms of epigenetic inheritance. Current Opinion in Cell Biology, 19(3), 266272. doi: 10.1016/j.ceb.2007.04.002 CrossRefGoogle ScholarPubMed
Matsuda, S., Nakagawa, Y., Kitagishi, Y., Nakanishi, A. and Murai, T. (2018). Reactive oxygen species, superoxide dimutases, and PTEN-p53-AKT-MDM2 signaling loop network in mesenchymal stem/stromal cells regulation. Cells, 7(5), 36. doi: 10.3390/cells7050036 CrossRefGoogle Scholar
Men, H., Monson, R. L., Parrish, J. J. and Rutledge, J. J. (2003). Degeneration of cryopreserved bovine oocytes via apoptosis during subsequent culture. Cryobiology, 47(1), 7381. doi: 10.1016/s0011-2240(03)00070-1 CrossRefGoogle ScholarPubMed
Mogas, T. (2019). Update on the vitrification of bovine oocytes and in vitro-produced embryos. Reproduction, Fertility, and Development, 31(1), 105117. doi: 10.1071/RD18345 CrossRefGoogle Scholar
Monzo, C., Haouzi, D., Roman, K., Assou, S., Dechaud, H. and Hamamah, S. (2012). Slow freezing and vitrification differentially modify the gene expression profile of human metaphase II oocytes. Human Reproduction, 27(7), 21602168. doi: 10.1093/humrep/des153 CrossRefGoogle ScholarPubMed
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 and Development, 21(6), 682693. doi: 10.1101/gad.1521307 CrossRefGoogle ScholarPubMed
Noh, E. M., Park, J., Song, H. R., Kim, J. M., Lee, M., Song, H. K., Hong, O. Y., Whang, P. H., Han, M. K., Kwon, K. B., Kim, J. S. and Lee, Y. R. (2016). Skin aging-dependent activation of the PI3K signaling pathway via downregulation of PTEN increases intracellular ROS in human dermal fibroblasts. Oxidative Medicine and Cellular Longevity, 2016, 6354261. doi: 10.1155/2016/6354261 CrossRefGoogle ScholarPubMed
Nohales-Córcoles, M., Sevillano-Almerich, G., Di Emidio, G., Tatone, C., Cobo, A. C., Dumollard, R. and De los Santos Molina, M. J. (2016). Impact of vitrification on the mitochondrial activity and redox homeostasis of human oocyte. Human Reproduction, 31(8), 18501858. doi: 10.1093/humrep/dew130 CrossRefGoogle ScholarPubMed
Nottola, S. A., Albani, E., Coticchio, G., Palmerini, M. G., Lorenzo, C., Scaravelli, G., Borini, A., Levi-Setti, P. E. and Macchiarelli, G. (2016). Freeze/thaw stress induces organelle remodeling and membrane recycling in cryopreserved human mature oocytes. Journal of Assisted Reproduction and Genetics, 33(12), 15591570. doi: 10.1007/s10815-016-0798-x CrossRefGoogle ScholarPubMed
O’Connell, R. M., Chaudhuri, A. A., Rao, D. S. and Baltimore, D. (2009). Inositol phosphatase SHIP1 is a primary target of miR-155. Proceedings of the National Academy of Sciences of the United States of America, 106(17), 71137118. doi: 10.1073/pnas.0902636106 CrossRefGoogle ScholarPubMed
Pan, B., Toms, D., Shen, W. and Li, J. (2015). MicroRNA-378 regulates oocyte maturation via the suppression of aromatase in porcine cumulus cells. American Journal of Physiology. Endocrinology and Metabolism, 308(6), E525E534. doi: 10.1152/ajpendo.00480.2014 CrossRefGoogle ScholarPubMed
Quan, G., Wu, G. and Hong, Q. (2017). Oocyte cryopreservation based in sheep: The current status and future perspective. Biopreservation and Biobanking, 15(6), 535547. doi: 10.1089/bio.2017.0074 CrossRefGoogle ScholarPubMed
Rasti, G. and Vaquero, A. (2018). Epigenetic modifications of histones. In Epigenetics and assisted reproduction (pp. 1728). CRC Press. E-book ISBN 9781315208701.CrossRefGoogle Scholar
Rienzi, L., Cobo, A., Paffoni, A., Scarduelli, C., Capalbo, A., Vajta, G., Remohí, J., Ragni, G. and Ubaldi, F. M. (2012). Consistent and predictable delivery rates after oocyte vitrification: An observational longitudinal cohort multicentric study. Human Reproduction, 27(6), 16061612. doi: 10.1093/humrep/des088 CrossRefGoogle ScholarPubMed
Saenz-de-Juano, M. D., Marco-Jimenez, F., Viudes-de-Castro, M. P., Lavara, R. and Vicente, J. S. (2014). Direct comparison of the effects of slow freezing and vitrification on late blastocyst gene expression, development, implantation and offspring of rabbit morulae. Reproduction in Domestic Animals, 49(3), 505511. doi: 10.1111/rda.12320 CrossRefGoogle ScholarPubMed
Saragusty, J. and Arav, A. (2011). Current progress in oocyte and embryo cryopreservation by slow freezing and vitrification. Reproduction, 141(1), 119. doi: 10.1530/REP-10-0236 CrossRefGoogle ScholarPubMed
Sargisian, N., Lannering, B., Petzold, M., Opdahl, S., Gissler, M., Pinborg, A., Henningsen, A.-K. A., Tiitinen, A., Romundstad, L. B., Spangmose, A. L., Bergh, C. and Wennerholm, U. B. (2022). Cancer in children born after frozen–thawed embryo transfer: A cohort study. PLOS Medicine, 19(9), e1004078. doi: 10.1371/journal.pmed.1004078 CrossRefGoogle ScholarPubMed
Sevignani, C., Calin, G. A., Siracusa, L. D. and Croce, C. M. (2006). Mammalian microRNAs: A small world for fine-tuning gene expression. Mammalian Genome, 17(3), 189202. doi: 10.1007/s00335-005-0066-3 CrossRefGoogle ScholarPubMed
Shahedi, A., Hosseini, A., Ali Khalili, M. and Yeganeh, F. (2017). Vitrification affects nuclear maturation and gene expression of immature human oocytes. Research in Molecular Medicine, 5(1), 2733. doi: 10.29252/rmm.5.1.27 Google Scholar
Solé, M., Santaló, J., Boada, M., Clua, E., Rodríguez, I., Martínez, F., Coroleu, B., Barri, P. N. and Veiga, A. (2013). How does vitrification affect oocyte viability in oocyte donation cycles? A prospective study to compare outcomes achieved with fresh versus vitrified sibling oocytes. Human Reproduction, 28(8), 20872092. doi: 10.1093/humrep/det242 CrossRefGoogle ScholarPubMed
Somoskoi, B., Martino, N. A., Cardone, R. A., Lacalandra, G. M., Dell’Aquila, M. E. and Cseh, S. (2015). Different chromatin and energy/redox responses of mouse morulae and blastocysts to slow freezing and vitrification. Reproductive Biology and Endocrinology: RB&E, 13, 22. doi: 10.1186/s12958-015-0018-z CrossRefGoogle ScholarPubMed
Song, W. Y., Peng, Z. F., Chen, X. M., Jin, H. X., Yao, G. D., Shi, S. L., Yang, H. Y., Zhang, X. Y. and Sun, Y. P. (2016). Effects of vitrification on outcomes of in vivo-mature, in vitro-mature and immature human oocytes. Cellular Physiology and Biochemistry, 38(5), 20532062. doi: 10.1159/000445564 CrossRefGoogle ScholarPubMed
Spinaci, M., Vallorani, C., Bucci, D., Tamanini, C., Porcu, E. and Galeati, G. (2012). Vitrification of pig oocytes induces changes in histone H4 acetylation and histone H3 lysine 9 methylation (H3K9). Veterinary Research Communications, 36(3), 165171. doi: 10.1007/s11259-012-9527-9 CrossRefGoogle ScholarPubMed
Stigliani, S., Moretti, S., Anserini, P., Casciano, I., Venturini, P. L. and Scaruffi, P. (2015). Storage time does not modify the gene expression profile of cryopreserved human metaphase II oocytes. Human Reproduction, 30(11), 25192526. doi: 10.1093/humrep/dev232 CrossRefGoogle Scholar
Succu, S., Leoni, G. G., Berlinguer, F., Madeddu, M., Bebbere, D., Mossa, F., Bogliolo, L., Ledda, S. and Naitana, S. (2007). Effect of vitrification solutions and cooling upon in vitro matured prepubertal ovine oocytes. Theriogenology, 68(1), 107114. doi: 10.1016/j.theriogenology.2007.04.035 CrossRefGoogle ScholarPubMed
Sun, K. and Lai, E. C. (2013). Adult-specific functions of animal microRNAs. Nature Reviews. Genetics, 14(8), 535548. doi: 10.1038/nrg3471 CrossRefGoogle ScholarPubMed
Takahashi, N., Nakaoka, T. and Yamashita, N. (2012). Profiling of immune-related microRNA expression in human cord blood and adult peripheral blood cells upon proinflammatory stimulation. European Journal of Haematology, 88(1), 3138. doi: 10.1111/j.1600-0609.2011.01707.x CrossRefGoogle ScholarPubMed
Tesfaye, D., Worku, D., Rings, F., Phatsara, C., Tholen, E., Schellander, K. and Hoelker, M. (2009). Identification and expression profiling of microRNAs during bovine oocyte maturation using heterologous approach. Molecular Reproduction and Development, 76(7), 665677. doi: 10.1002/mrd.21005 CrossRefGoogle ScholarPubMed
Tomek, W. and Smiljakovic, T. (2005). Activation of Akt (protein kinase B) stimulates metaphase I to metaphase II transition in bovine oocytes. Reproduction, 130(4), 423430. doi: 10.1530/rep.1.00754 CrossRefGoogle ScholarPubMed
Trapphoff, T., Heiligentag, M., Simon, J., Staubach, N., Seidel, T., Otte, K., Fröhlich, T., Arnold, G. J. and Eichenlaub-Ritter, U. (2016). Improved cryotolerance and developmental potential of in vitro and in vivo matured mouse oocytes by supplementing with a glutathione donor prior to vitrification. Molecular Human Reproduction, 22(12), 867881. doi: 10.1093/molehr/gaw059 Google ScholarPubMed
Tscherner, A., Brown, A. C., Stalker, L., Kao, J., Dufort, I., Sirard, M. A. and LaMarre, J. (2018). STAT3 signaling stimulates miR-21 expression in bovine cumulus cells during in vitro oocyte maturation. Scientific Reports, 8(1), 11527. doi: 10.1038/s41598-018-29874-w CrossRefGoogle ScholarPubMed
Tulay, P., Naja, R. P., Cascales-Roman, O., Doshi, A., Serhal, P. and SenGupta, S. B. (2015). Investigation of microRNA expression and DNA repair gene transcripts in human oocytes and blastocysts. Journal of Assisted Reproduction and Genetics, 32(12), 17571764. doi: 10.1007/s10815-015-0585-0 CrossRefGoogle ScholarPubMed
Turathum, B., Saikhun, K., Sangsuwan, P. and Kitiyanant, Y. (2010). Effects of vitrification on nuclear maturation, ultrastructural changes and gene expression of canine oocytes. Reproductive Biology and Endocrinology: RB&E, 8, 70. doi: 10.1186/1477-7827-8-70 CrossRefGoogle ScholarPubMed
Verheijen, M., Lienhard, M., Schrooders, Y., Clayton, O., Nudischer, R., Boerno, S., Timmermann, B., Selevsek, N., Schlapbach, R., Gmuender, H., Gotta, S., Geraedts, J., Herwig, R., Kleinjans, J. and Caiment, F. (2019). DMSO induces drastic changes in human cellular processes and epigenetic landscape in vitro . Scientific Reports, 9(1), 4641. doi: 10.1038/s41598-019-40660-0 CrossRefGoogle ScholarPubMed
Walls, M., Junk, S., Ryan, J. P. and Hart, R. (2012). IVF versus ICSI for the fertilization of in-vitro matured human oocytes. Reproductive Biomedicine Online, 25(6), 603607. doi: 10.1016/j.rbmo.2012.08.001 CrossRefGoogle ScholarPubMed
Wang, N., Li, C. Y., Zhu, H. B., Hao, H. S., Wang, H. Y., Yan, C. L., Zhao, S. J., Du, W. H., Wang, D., Liu, Y., Pang, Y. W. and Zhao, X. M. (2017a). Effect of vitrification on the mRNA transcriptome of bovine oocytes. Reproduction in Domestic Animals, 52(4), 531541. doi: 10.1111/rda.12942 CrossRefGoogle ScholarPubMed
Wang, T. Y., Zhang, J., Zhu, J., Lian, H. Y., Yuan, H. J., Gao, M., Luo, M. J. and Tan, J. H. (2017b). Expression profiles and function analysis of microRNAs in postovulatory aging mouse oocytes. Aging, 9(4), 11861201. doi: 10.18632/aging.101219 CrossRefGoogle ScholarPubMed
Wang, Y., Zhang, M. L., Zhao, L. W., Kuang, Y. P. and Xue, S. G. (2018). Enhancement of the efficiency of oocyte vitrification through regulation of histone deacetylase 6 expression. In Journal of Assisted Reproduction and Genetics (2018/07/04. Springer, 35(7), 11791185). doi: 10.1007/s10815-018-1221-6 CrossRefGoogle ScholarPubMed
Wen, Y., Zhao, S., Chao, L., Yu, H., Song, C., Shen, Y., Chen, H. and Deng, X. (2014). The protective role of antifreeze protein 3 on the structure and function of mature mouse oocytes in vitrification. Cryobiology, 69(3), 394401. doi: 10.1016/j.cryobiol.2014.09.006 CrossRefGoogle ScholarPubMed
Wilczynska, A., Git, A., Argasinska, J., Belloc, E. and Standart, N. (2016). CPEB and miR-15/16 co-regulate translation of cyclin E1 mRNA during Xenopus oocyte maturation. PLOS ONE, 11(2), e0146792. doi: 10.1371/journal.pone.0146792 CrossRefGoogle ScholarPubMed
Wright, E. C., Hale, B. J., Yang, C. X., Njoka, J. G. and Ross, J. W. (2016). MicroRNA-21 and PDCD4 expression during in vitro oocyte maturation in pigs. Reproductive Biology and Endocrinology: RB&E, 14(1), 21. doi: 10.1186/s12958-016-0152-2 CrossRefGoogle ScholarPubMed
Xia, B., Lu, J., Wang, R., Yang, Z., Zhou, X. and Huang, P. (2018). miR-21–3p regulates influenza A virus replication by targeting histone deacetylase-8. Frontiers in Cellular and Infection Microbiology, 8, 175. doi: 10.3389/fcimb.2018.00175 CrossRefGoogle ScholarPubMed
Xiao, G., Xia, C., Yang, J., Liu, J., Du, H., Kang, X., Lin, Y., Guan, R., Yan, P. and Tang, S. (2014). MiR-133b regulates the expression of the actin protein TAGLN2 during oocyte growth and maturation: A potential target for infertility therapy. PLOS ONE, 9(6), e100751. doi: 10.1371/journal.pone.0100751 CrossRefGoogle ScholarPubMed
Xu, Y. W., Wang, B., Ding, C. H., Li, T., Gu, F. and Zhou, C. (2011). Differentially expressed micoRNAs in human oocytes. Journal of Assisted Reproduction and Genetics, 28(6), 559566. doi: 10.1007/s10815-011-9590-0 CrossRefGoogle ScholarPubMed
Yamanaka, Y., Tagawa, H., Takahashi, N., Watanabe, A., Guo, Y. M., Iwamoto, K., Yamashita, J., Saitoh, H., Kameoka, Y., Shimizu, N., Ichinohasama, R. and Sawada, K. (2009). Aberrant overexpression of microRNAs activate AKT signaling via down-regulation of tumor suppressors in natural killer-cell lymphoma/leukemia. Blood, 114(15), 32653275. doi: 10.1182/blood-2009-06-222794 CrossRefGoogle ScholarPubMed
Yan, L. Y., Yan, J., Qiao, J., Zhao, P. L. and Liu, P. (2010). Effects of oocyte vitrification on histone modifications. Reproduction, Fertility, and Development, 22(6), 920925. doi: 10.1071/RD09312 CrossRefGoogle ScholarPubMed
Yan, M., Chen, C., Gong, W., Yin, Z., Zhou, L., Chaugai, S. and Wang, D. W. (2015). miR-21–3p regulates cardiac hypertrophic response by targeting histone deacetylase-8. Cardiovascular Research, 105(3), 340352. doi: 10.1093/cvr/cvu254 CrossRefGoogle ScholarPubMed
Zhang, C. X., Cui, W., Zhang, M., Zhang, J., Wang, T. Y., Zhu, J., Jiao, G. Z. and Tan, J. H. (2014). Role of Na+/Ca2+ exchanger (NCX) in modulating postovulatory aging of mouse and rat oocytes. PLOS ONE, 9(4), e93446. doi: 10.1371/journal.pone.0093446 CrossRefGoogle ScholarPubMed
Zhang, Z., Mu, Y., Ding, D., Zou, W., Li, X., Chen, B., Leung, P. C., Chang, H. M., Zhu, Q., Wang, K., Xue, R., Xu, Y., Zou, H., Zhou, P., Wei, Z. and Cao, Y. (2021). Melatonin improves the effect of cryopreservation on human oocytes by suppressing oxidative stress and maintaining the permeability of the oolemma. Journal of Pineal Research, 70(2), e12707. doi: 10.1111/jpi.12707 CrossRefGoogle ScholarPubMed
Zhao, X. M., Ren, J. J., Du, W. H., Hao, H. S., Wang, D., Qin, T., Liu, Y. and Zhu, H. B. (2013). Effect of vitrification on promoter CpG island methylation patterns and expression levels of DNA methyltransferase 1o, histone acetyltransferase 1, and deacetylase 1 in metaphase II mouse oocytes. Fertility and Sterility, 100(1), 256261. doi: 10.1016/j.fertnstert.2013.03.009 CrossRefGoogle ScholarPubMed
Zhao, X., Hao, H., Du, W. and Zhu, H. (2015). Effect of vitrification on the microRNA transcriptome in mouse blastocysts. PLOS ONE, 10(4), e0123451. doi: 10.1371/journal.pone.0123451 CrossRefGoogle ScholarPubMed
Zhu, Q., Ding, D., Yang, H., Zou, W., Yang, D., Wang, K., Zhang, C., Chen, B., Ji, D., Hao, Y., Xue, R., Xu, Y., Wang, Q., Wang, J., Yan, B., Cao, Y., Zou, H. and Zhang, Z. (2022). Melatonin protects mitochondrial function and inhibits oxidative damage against the decline of human oocytes development caused by prolonged cryopreservation. Cells, 11(24), 4018. doi: 10.3390/cells11244018 CrossRefGoogle ScholarPubMed