Next Article in Journal
The Role of GPR15 Function in Blood and Vasculature
Next Article in Special Issue
Autoimmunity to the Follicle-Stimulating Hormone Receptor (FSHR) and Luteinizing Hormone Receptor (LHR) in Polycystic Ovarian Syndrome
Previous Article in Journal
Amyloid-β Processing in Aged S100B Transgenic Mice Is Sex Dependent
Previous Article in Special Issue
Influence of Vitamin D on the Vasoactive Effect of Estradiol in a Rat Model of Polycystic Ovary Syndrome
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transcriptome Analyses Identify Potential Key microRNAs and Their Target Genes Contributing to Ovarian Reserve

1
Department of Obstetrics and Gynecology, Seoul National University Hospital, Seoul 03080, Korea
2
Institute of Reproductive Medicine and Population, Medical Research Center, Seoul National University, Seoul 03080, Korea
3
Transdisciplinary Department of Medicine & Advanced Technology, Seoul National University Hospital, Seoul 03080, Korea
4
Department of Obstetrics and Gynecology, Korea University College of Medicine, Goryeodae-ro 73, Seongbuk-gu, Seoul 02841, Korea
*
Author to whom correspondence should be addressed.
Both authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(19), 10819; https://doi.org/10.3390/ijms221910819
Submission received: 5 August 2021 / Revised: 13 September 2021 / Accepted: 20 September 2021 / Published: 6 October 2021

Abstract

:
Female endocrinological symptoms, such as premature ovarian inefficiency (POI) are caused by diminished ovarian reserve and chemotherapy. The etiology of POI remains unknown, but this can lead to infertility. This has accelerated the search for master regulator genes or other molecules that contribute as enhancers or silencers. The impact of regulatory microRNAs (miRNAs) on POI has gained attention; however, their regulatory function in this condition is not well known. RNA sequencing was performed at four stages, 2-(2 W), 6-(6 W), 15-(15 W), and 20-(20 W) weeks, on ovarian tissue samples and 5058 differentially expressed genes (DEGs) were identified. Gene expression and enrichment were analyzed based on the gene ontology and KEGG databases, and their association with other proteins was assessed using the STRING database. Gene set enrichment analysis was performed to identify the key target genes. The DEGs were most highly enriched in 6 W and 15 W groups. Figla, GDF9, Nobox, and Pou51 were significantly in-creased at 2 W compared with levels at 6 W and 20 W, whereas the expression of Foxo1, Inha, and Taf4b was significantly de-creased at 20 W. Ccnd2 and Igf1 expression was maintained at similar levels in each stage. In total, 27 genes were upregulated and 26 genes interacted with miRNAs; moreover, stage-specific upregulated and downregulated interactions were demonstrated. Increased and decreased miRNAs were identified at each stage in the ovaries. The constitutively expressed genes, Ccnd2 and Igf1, were identified as the major targets of many miRNAs (p < 0.05), and Fshr and Foxo3 interacted with miRNAs, namely mmu-miR-670-3p and mmu-miR-153-3p. miR-26a-5p interacted with Piwil2, and its target genes were downregulated in the 20 W mouse ovary. In this study, we aimed to identify key miRNAs and their target genes encompassing the reproductive span of mouse ovaries using mRNA and miRNA sequencing. These results indicated that gene sets are regulated in the reproductive stage-specific manner via interaction with miRNAs. Furthermore, consistent expression of Ccnd2 and Igf1 is considered crucial for the ovarian reserve and is regulated by many interactive miRNAs.

1. Introduction

Diminished ovarian reserve, such as premature ovarian inefficiency (POI) is an increasing problem and leads to infertility in women under the age of 40 years. In previous studies, genes involved in POI have been identified (Table 1), but the critical regulators of this condition have not been fully elucidated. In addition to regulatory genes, non-coding RNAs are considered candidate regulators. miRNAs are well-known conserved regulators of gene expression [1], and their contribution to germ cell developmental control has been studied. Their involvement in ovarian follicles and oocyte development has also been demonstrated [2,3,4]. The correlation between miRNAs and POI has been studied, and several miRNAs have been suggested to be associated with POI [5]. Among these, miR-23a and miR-27a are involved in granulosa cell proliferation and apoptosis by controlling XIAP [6,7] and IGFBP2 [8].
The fundamental reasons for POI are not fully elucidated due to its complex regulation; however, advancements in array technology have enhanced the identification of major genes and other regulators, such as non-coding RNAs, including microRNAs (miRNAs) [9,10] and long non-coding RNAs [11,12]. miRNAs are small non-coding RNAs that contribute to the regulation of gene expression at transcriptional and post-transcriptional levels [1,13]. Their regulatory functions in the ovary are also known [14,15], and their regulatory role [3,16] and expression in patients with POI have been studied [17,18].
The reproductive span of oocytes changes dramatically during the life cycle of women [19]. The number of oocytes in a female is pre-determined, and these are not reproducible and are diminished until menopause via menstrual cycles [20]. However, the reproductive cycle becomes irregular with endocrinological symptoms such as POI and other exogenous factors such as chemotherapy. With advancements in array technologies, some key genes regulating follicle development, ovarian reserve, and POI have been reported (Table 1).
Table 1. List of genes related to premature ovarian inefficiency (POI) from previous studies.
Table 1. List of genes related to premature ovarian inefficiency (POI) from previous studies.
GeneBiological CategoryFunctionsReference
AMH(19q13.3)Anti-Müllerian hormoneControl of the formation of primary follicles by inhibiting excessive follicular recruitment by FSH[21]
AMHR2(12q13)AMH receptorAMH signal transduction[22]
BMPR2BMP receptorSignal transduction between oocytes and somatic cells[23]
BMP15(Xq11.2)Growth factorGrowth and differentiation of granulosa cells (GCs)[24]
ESR1Estrogen receptorRegulation of follicle growth and maturation and oocyte release[25]
FIGLA(2q13.3)bHLH transcription factorRegulation of multiple oocyte-specific genes, including genes involved in folliculogenesis and those that encode the zona pellucida[26]
FMR1(Xq27)Highly polymorphic CGG repeat in the 5′ untranslated region (UTR) of the exon 1Transcriptional regulation[27]
FSHR(2q21-p16)ReceptorFollicular development and ovarian steroidogenesis[28]
FOXL2(3q23)Transcription factorDifferentiation and growth of granulosa cells[29]
FOXO3ATranscription factorRegulating primordial follicle growth activation[30]
GDF9(5q31.1)Growth factorGrowth and differentiation of granulosa cells proliferation[31]
INHA variantsGrowth factorMaturation of ovarian follicles by FSH inhibition[32]
KHDRBS1Signal transduction activatorAlter mRNA expression level and alternative splicing[33]
LHRLutropin-choriogonadotropic hormone receptorRegulation of ovarian follicle maturation, steroidogenesis, and ovulation[34]
LHX8Transcription factorGerm-cell-specific critical regulator of early oogenesis[35]
NOBOX(7q35)Transcription factorFollicle development[36]
PGRMC1(Xq22-q24)Heme-binding proteinRegulation of apoptosis[37]
POLR3HRNA polymerase III subunit HRegulation of cell cycle, cell growth, and differentiation[38]
SOHLH1*Transcription factorEarly folliculogenesis[39]
WT1(11q13)Transcription factorGranulosa cell differentiation and oocyte-granulosa cell interaction[40]
Especially, POI-related gene families and pathways, such as the Fas ligand (FasL)-Fas pathway, nuclear factor-kappa B (NF-κB), inhibitory kappa B α (IκBα), interleukin-1 receptor-associated kinase (IRAK1), and tumor necrosis factor receptor-associated factor 6 (TRAF6), have been reported in humans [41,42,43].
Other studies have demonstrated the relationship between mRNA-miRNA interaction during ovarian follicle development and POI regulation [44,45]. Correlations between miRNAs and POI have been studied, and miR-23a, miR-27a, miR-22-3p, miR-146a, miR-196a, miR-290-295, miR-423, and miR-608 have been suggested to be associated with this condition [5]. Among these miRNAs, miR-23a and miR-27a are involved in granulosa cell proliferation and apoptosis by controlling XIAP [6,7] and IGFBP2 [8]. The apoptosis of granulosa cells is critical in POI, therefore, understanding the regulatory roles of miRNAs in granulosa cells enhances our understanding of the pathogenesis of POI [46].
The Piwil (Piwi) protein family is a small-RNA-bound effector complex [47,48], and the mouse genome encoded three Piwi proteins, PIWIL1/MIWI, PIWIL2/MILI, and PIWIL4/MIWI2 [49]. Mammalian Piwil proteins associate with Piwi-interacting RNAs (piRNAs) and piRNA expression is largely restricted to the germline [50]. The deficiency of genes required for piRNA biogenesis leads to infertility in males [51]; however, females with this genotype are not infertile [52]. Whereas the role of regulatory miRNAs in POI has been studied, their regulatory function in this condition is not well known.
In this study, we aimed to identify the miRNAs that regulate the ovarian age-specific genes, which might contribute to the regulation of POI. We also analyzed the correlation between Piwil and specific miRNAs.

2. Results

2.1. Identification of Transcripts in Different Reproductive Stages of Mouse Ovaries

We identified a total of 5058 DEGs (Figure 1 and Figure S1) including 2658 upregulated genes and 2400 downregulated genes in the reproductively aged mouse ovary based on comparisons among stages (Table 2). Correlations of total gene count (Figure 2A) from each stage and filtered count numbers were determined by respective FPKM and TPM analyses and displayed by heatmaps (Figure S2A,B). The count data displayed by the PCA plot indicate the distance of expressed transcripts (Figure 2B). The distribution of transcripts is displayed as a cluster in Figure S3. The DEGs among the reproductively aged mouse ovaries were compared and different comparisons were divided into six groups as follows: (1) 6 weeks (W) vs. 2 W, (2) 15 W vs. 2 W, (3) 20 W vs. 2 W, (4) 15 W vs. 6 W, (5) 20 W vs. 6 W, (6) 20 W vs. 15 W. The DEGs are displayed as a volcano plot based on logFC values (Figure 3A–F).

2.2. Gene Ontology (GO) Enrichment of DEGs

The overall biological processes associated with DEGs were analyzed, and the p value cutoff was set at <0.05. Gene ontology biological process, gene ontology molecular function, and gene ontology cellular component (GOCC) analyses were performed.
In the 6 W vs. 2 W group, the DEGs were mainly enriched in the steroid biosynthetic process, regulation of cholesterol metabolic process, organonitrogen compound biosynthetic process, and cholesterol biosynthetic process (Figure S4A). In the 15 W vs. 2 W set, DEG sets comprising the regulation of cell migration, negative regulation of the cellular process, and extracellular matrix organization were enriched. Upregulated DEGs were mainly enriched in the sterol biosynthetic process, regulation of cholesterol metabolic process, and cholesterol biosynthetic process. Downregulated DEGs were enriched in the cellular protein metabolic process (Figure S4B). Analysis revealed that lysosome, Golgi subcompartments, and Golgi membranes were enriched among cellular components (Figure S6A).
In the 20 W vs. 2 W set, DEGs were highly enriched in cytokine-medicated signaling pathways, and downregulated DEGs were mostly enriched in cellular protein metabolic process (Figure S4C), neuropilin binding (Figure S5A), and the endoplasmic reticulum lumen (Figure S6B). In the 15 W and 6 W sets, DEGs were enriched in the cellular response to cytokine stimulus (Figure S4D) and cytosolic part (Figure S6C). In the 20 W vs. 6 W set, DEGs were enriched in the regulation of transcription from RNA polymerase II promoter, regulation of cell proliferation, and positive regulation of transcription (Figure S4E). Furthermore, they were enriched in cytokine activity (Figure S5B) and platelet alpha granules (Figure S6D). For 20 W vs. 15 W, DEGs were enriched in the regulation of transcription from RNA polymerase II promoter and positive regulation of transcription (Figure S4F). They were also enriched in RNA polymerase II regulatory region sequence-specific DNA binding (Figure S5C). These data indicated that the transcriptomes in the ovary are altered and stage-specific DEGs are increased in the reproductively aged 20 W ovary.

2.3. KEGG Analysis of DEGs

The upregulated DEGs in each group were enriched in cytokine-cytokine receptors for 6 W vs. 2 W (Figure S7A) and cell adhesion molecules for 15 W vs. 2 W (Figure S7B). In the 20 W vs. 2 W set, DEGs were enriched in cytokine-cytokine receptors (Figure S7 C). Hormones, steroids, ovarian signaling, and cortisol synthesis were upregulated in the 15 W vs. 6 W group (Figure S7D). DEGs of 20 W vs. 6 W were enriched in TNF signaling, IL-17 signaling, MAPK signaling, and cytokine-cytokine receptor (Figure S7E). TNF signaling and IL-17 signaling were highly enriched for DEGs that were upregulated (Figure S7F).

2.4. Protein-Protein Interaction (PPI) Network Construction and Clusters Analyses

The PPI networks were built using STRING based on DEGs logFC values and are shown in Table 3 and Figure S8. The PPI networks of upregulated and downregulated DEGs were generated, and their biological processes, molecular functions, and KEGG pathways were analyzed. In the 6 W vs. 2 W set, Klk1, which has a role in the positive regulation of steroid hormone biosynthetic process, and the well-known Figla were downregulated (Table 3A and Figure S8A,B). Steroid biosynthesis-related Hsd17b7 and cholesterol metabolism-related Star were upregulated for 15 W vs. 2 W (Table 3B and Figure S8C). The RNA polymerase regulator, Atf3 was downregulated (Table 3B and Figure S8D), and this phenomenon was also observed in the 15 W vs. 6 W set. Steroid hormone biosynthesis-related Akr1c18 and Hsd17b7 were also enriched in 20 W vs. 2 W (Table 3C and Figure S8E). Eif3j1, which has translation initiation factor activity was upregulated for 15 W vs. 6 W (Table 3D and Figure S8F). In the 20 W vs. 6 W set, Oog1 was downregulated, and it is involved female gamete generation (Table 3E and Figure S8G). For 20 W vs. 15 W, the IL-17 signaling pathway-related Cxcl1 and Atf3, and the regulation of transcription from RNA polymerase II promoter were upregulated in response to endoplasmic reticulum stress (Table 3F and Figure S8H).

2.5. Identification of Key Target Gene and Their Validation

Among the DEGs identified in each reproductive stage, we selected 90 highly expressed genes after filtering and selected 13, Bmp15, Ccnd2, Figla, Foxo1, Foxo3, Fshr, Gdf9, Igf1, Inha, Nobox, Smad3, Taf4b, and Pou5f1, as key target genes based on their correlation with the ovaries (Table 4). Their expression levels and ranked correlations were calculated and filtered using the FPKM and TPM values (p < 0.05).
The expression levels of Ccnd2 and Igf1 were similar throughout the reproductively-aged mouse ovaries. The levels of Figla, GDF9, Nobox, and Pou5f1 were decreased as reproductive aging progressed. Foxo1, Inha, and Taf4b demonstrated a more than two-fold increase in expression. Finally, in reproductively aged individuals, Fshr increased at 20 W.

2.6. Correlation with miRNAs

Interacting miRNAs with the key target genes were identified based on the ranked correlation (p < 0.05) and are listed in Table 5. Their interactions with miRNAs were confirmed using databases (Table S3). Continuously expressed genes, such as Ccnd2 and Igf1, are target genes of various miRNAs (Table 5). Genes regulating oocyte and follicle development, such as Figla, Foxo3, and Fshr, were determined to interact with only one miRNA, mmu-miR-669d-2-3p, mmu-miR-153-3p, and mmu-miR-670-3p, respectively.
In total, 27 genes were increased based on the regulatory effects of miRNAs, and 26 genes were decreased between the 2 W and 20 W sets (p < 0.05, Table S4A). During the reproductive period, some genes were downregulated following transient up-regulation. The genes expressed at high levels at 20 W and decreased at 2 W (Table S4B, left panel), as well as the genes expressed at high levels at 2 W and decreased at 20 W, set are listed in Table S4B, right panel. Genes upregulated following downregulation in the 20 W set (Table S4C, left panel) and 2 W set (Table S4C, right panel) are also listed with their interacting miRNAs.
Moreover, the up- and downregulated DEGs between 2 W and 20 W, and their correlations with miRNAs occurring in a reproductive-age-specific manner in the ovary were further analyzed using the database. mmu-miR-9-5p, mmu-miR-26a-5p, mmu-miR-29b-3p, mmu-miR-26a-5p, mmu-miR-706 were the miRNAs predominantly correlated with downregulated DEGs (Table S5B).

2.7. Interaction with Piwil Gene Family

The Piwil gene family is essential for developmental regulation. Therefore, we additionally analyzed the correlation between DEGs from reproductively aged mouse ovaries and this family (Table S6A). Target genes and each Piwil gene were analyzed and are listed in Table S6B. Target genes of Piwil included well-known early oocyte development and follicle assembly genes (Table S6B). Piwil4 was determined to interact with mmu-miR-210-5p and mmu-miR-3470b, whereas Piwil1 and Piwil2 were found to interact with more miRNAs (Table 6).

3. Discussion

In this study, we demonstrated the stage-specific upregulation and downregulation of DEGs and their correlation with miRNAs in reproductively aged mouse ovaries. We identified 24,958 transcripts from each reproductive stage of ovaries, and 5058 genes were identified. In total, 2658 genes were upregulated and 2400 genes were downregulated. Furthermore, 53 upregulated and downregulated miRNAs were identified. GO analyses demonstrated that upregulated DEGs were enriched in different cellular processes in a stage-specific manner. DEGs from the early phase set of 2 W, ovaries were mainly enriched in the steroid biosynthetic process, regulation of cholesterol metabolic process, organonitrogen compound biosynthetic process, and cholesterol biosynthetic process. In ovaries from middle-phase sets, 6 W and 15 W, DEG sets associated with the regulation of cell migration, negative regulation of the cellular process, and extracellular matrix organization were enriched. KEGG analyses suggested that the upregulated genes were mostly enriched in cell adhesion molecules.
We also focused on the regulatory roles of miRNAs in reproductively aged mouse ovaries. The roles of miRNAs in folliculogenesis [4] and oocyte development have emerged [3]. We focused on the altered gene expression in a specific manner, and interactive regulatory miRNAs of target genes were identified based on a score > 90 (p <0.05). mmu-miR-136-5p, mmu-miR-335-5p, mmu-miR-665-3p, and mmu-miR-18a-5p targeted several upregulated genes in 20 W mouse ovaries (Table S5A). miRNAs targeted genes were involved in spermatogenesis, follicular development process. ESR1, SMAD2, and Spata1 are the target of mmu-miR-18a-5p. ESR1 has been involved in the genetic variation of female infertility, and Sparta is an acrosomal protein in sperm and related to genetic mutation in sperm. Its role on female infertility is still unknown. mmu-miR-136-5p targets Prdm16 and GDF6, genes involved in follicular development. mmu-miR-335-5p targets Calu, regulator of chromosome condensation. The major target of mmu-miR-665-3p is Pou2f3, which belongs to same family of Oct4, master regulator of pluripotency.
In the down-regulated gene set, mmu-miR-9-5p, mmu-miR-26a-5p, mmu-miR-706, and mmu-miR-29b-3p correlated with several genes (Table S5B). mmu-miR-9-5p targets several fertility and germ cell developmental genes, such as Frip2, Prdm6, Spam1, Prdm1, Sirt1 and Map3k2. mmu-miR-26a-5p mainly targets cell growth regulation by estrogen and voltage-dependent channel genes, Kcnq4 and Cacna1c. mmu-miR-706 targets mesodermal genes and mmu-miR-29b-3p targets Bcl2 modifying factor (Bmf) and vascular endothelial growth factor A (Vegfa). In this finding, we hypothesize that sperm activity function regulating genes is a major target of up and downregulated miRNAs. Therefore, we conclude that several male fertility genes might be the involved onset of diminishing ovarian reserve.
Among the highly expressed DEGs, well-known ovary-associated genes were identified. The master regulator of oocyte development, Figla was downregulated in later-stage ovaries. Figla (factor in the germline alpha), encodes a germ cell-specific basic helix-loop-helix transcription factor first identified as an activator of oocyte genes [53,54]. Figla regulates primordial follicle formation in the fetal ovary [55] and antagonizes spermatogenesis during embryo development [56]. This gene is related to POI, and its expression was decreased according to reproductive age. Its interacting miRNA was mmu-miR-669d-2-3p, and the main function of this is to modify sperm-related gene expression [57]. This result supports the known function of Figla and supports its decreased expression throughout the reproductive span.
Foxo3 and Fshr are also mainly regulated by miRNAs. Foxo3 is involved in the apoptosis of granulosa cells [58] and involved in the early development of follicles on its own [59] or with other genes [60]. Foxo3 interacts with mmu-miR-153-3p, which is known to suppress tumor growth in ovarian carcinoma [61]. Fshr (follicle-stimulating hormone receptor) is one of the most important receptors in the ovary and other female reproductive organs. It is associated with the granulosa cells of the follicle and is located in the granulosa cells [62]. Interactive mmu-miR-670-3p is expressed in the newborn ovary [63,64], however, its roles are still not well known.
Klk1 has been involved in positive regulation of the steroid hormone biosynthetic process, is related to the renin-angiotensin system, and was enriched in the early and middle phase sets. The renin-angiotensin system, with angiotensin, is known for its relation to ovarian follicle development [65] and acquisition of dominancy [66]. Components are expressed in an estrogen-dependent manner in the uterus and are involved in cell proliferation [67]. Furthermore, they also involved the relationship with umbilical veins [68].
The Piwil family has been identified as comprising regulatory proteins responsible for stem cell and germ cell differentiation [69], and it is co-expressed with early developmental genes [70]. To date, the Piwil family is known to be more involved in male fertility than female fertility [50]. POI does not result in infertility from the beginning; therefore, we hypothesized that fine regulatory components might be involved in the onset of POI. Furthermore, they would be responsive to E2, which is another important molecule in the ovary [71]. The main targets of this family are enriched in genes such as FOXL2, NANOG, POU5F1, LIN28A, SOX2, PRDM1, NANOS3, DAZL, and DDX4 (Table 6). These results indicate that Piwil might contribute to the regulation of certain DEGs in a stage-specific manner.
Among this family, Piwil4 interacted predominantly with two miRNAs, mmu-miR-210-5p, and mmu-miR-3470b. Further, mmu-miR-26a-5p, miRNA that was determined to interact with Piwil2, mostly interacted with downregulated genes, such as APCDD1, CTDSP2, EZH2, MGA4A, PATZ1, and RHOQ, based on the 2 W vs. 20 W comparison (Table S5 B). These results indicate that Piwil2 and mmu-miR-26a-5p might be a candidate that regulates POI. Piwi-interacting RNAs (piRNAs) are involved in infertility in males, but not in females. We hypothesized that Piwil2 does not lead to terminal infertility; however, it affects the ovarian reserve via downregulated genes. APCDD1 is involved in adipogenic differentiation [72], CTDSP2 is a target gene of FOXO and regulates cell cycle progression [72], EZH2 is an important regulator in the female reproductive tract [73], MGA4A is involved in embryo lethality and female infertility [74], and RHOQ regulates mitochondrial function [75]. All genes regulated by miR-26a-5p were related to female infertility and cell cycle progression. Therefore, their correlation with POI needs to be further studied. These findings also correlated with the target genes regulated by mmu-miR-26a-5p were male fertility-related genes.
In conclusion, we identified DEGs and interactive miRNAs in reproductively aged mouse ovaries. Stage-specific upregulation and downregulation of DEGs indicated the regulatory roles of miRNAs in the ovary and their correlation with ovarian reserve. We used three reproductively aged stages, which correspond to the reproductive span of a human. However, differences between species limit the interpretation of experimental results. POI is an endocrinological symptom only in women; therefore, identified miRNAs might have distinct roles in humans. Further studies are needed to elucidate the functions of these miRNAs as regulators of the ovarian reserve. We will further study the roles of these miRNAs during in vitro follicular development based on non-human primate models, which are physiologically closer to humans [76,77,78].

4. Materials and Methods

4.1. Ethics

All the experiments were conducted under the control of AAALAC guidelines, and the animal experimental plan was reviewed and approved by IACUC of Seoul National University Hospital (No.18-0029-C1A1).

4.2. Preparation of Ovaries and Extraction of Total RNA

The experimental scheme of this study is presented in Figure 1. Ovaries of 2, 8, 15, and 20 weeks old C57BL/6 strain female mice were collected immediately after cervical dislocation. The lipid pad near the ovaries was removed and the collected ovaries were washed with HBSS (Invitrogen, Waltham, MA, USA). And then, ovaries were minced into pieces by a surgical blade (Feather Safety Razor, Osaka, Japan). The pieces were digested with Collagenase Type IV (1 mg/mL, Invitrogen) at 37 °C for 45 min and centrifugated. Pelleted samples washed with PBS (Invitrogen) and total RNAs were extracted using Trizol kit (Invitrogen) according to manufacturer’s instruction. Extracted RNAs were quantified and qualified for further Total Omics Transcriptome analyses.

4.3. Total Transcriptome (RNA-Seq) Analyses

4.3.1. mRNA Sequencing

The libraries were prepared for 150 bp paired-end sequencing using TruSeq Stranded mRNA Sample Preparation Kit (Illumina, CA, USA). Namely, mRNA molecules were purified and fragmented from 1μg of total RNA using oligo (dT) magnetic beads. The fragmented mRNAs were synthesized as single-stranded cDNAs through random hexamer priming. By applying this as a template for second-strand synthesis, double- stranded cDNA was prepared. After the sequential process of end repair, A-tailing, and adapter ligation, cDNA libraries were amplified with PCR (Polymerase Chain Reaction). The quality of these cDNA libraries was evaluated with the Agilent 2100 BioAnalyzer (Agilent, CA, USA). They were quantified with the KAPA library quantification kit (Kapa Biosystems, MA, USA) according to the manufacturer’s library quantification protocol. Following cluster amplification of denatured templates, sequencing was progressed as paired-end (2 × 150 bp) using Illumina NovaSeq 6000 sequencer (Illumina, CA, USA).

4.3.2. microRNA Sequencing

The libraries were prepared for 50 bp single-end sequencing using the NEXTflex Small RNA-Seq Kit (Bioo Scientific Corp). Namely, small RNA molecules were isolated from 1 μg of total RNA via the adapter ligation. The isolated small RNAs were synthesized as single-stranded cDNAs through the RT (Reverse transcription) priming. By applying this as a template for second-strand synthesis, double-stranded cDNA was prepared through PCR (Polymerase Chain Reaction). And, the fragments around 150 bp were extracted for sequencing through size selection by gel electrophoresis. The quality of these cDNA libraries was evaluated with the Agilent 2100 BioAnalyzer (Agilent, CA, USA) followed by quantification with the KAPA library quantification kit (Kapa Biosystems, MA, USA) according to the manufacturer’s protocol. Following cluster amplification of denatured templates, sequencing was progressed as single-end (50 bp) using Illumina sequencing platform (Illumina, CA, USA).

4.4. Read Quality of the Experiment

The identification was performed by the Sanger method, using Illumina 1.9. Total sequences were 32219029 before fast qc and 31058105 after fast qc. Sequence length ranged from 36 to 101 (Figure S1A). The percentage of the GC was 49 and the GC content has not emerged from theoretical distribution (Figure S1B). No rRNA contamination was observed. Sequences filtered as the poor quality was none (Figure S1C) and the duplicated sequence was 36.32% (Figure S1D).

4.5. Data Processing

We evaluated the data quality by sample clustering based on Pearson’s correlation matrices between different samples. And a heatmap was drawn corresponding to the different expressions of probes.

4.6. Differentially Expressed Genes (DEGs)

We employed the “limma” R language package to screen the DEGs between uterine leiomyoma and normal myometrium. The adjusted p-value < 0.05 and |log2fold change (FC)| > 1 were considered statistically significant.

4.7. KEGG and GO Enrichment Analyses of DEGs

Gene Ontology (GO, http://geneontology.org/, accessed on 1 February 2021) provides an ontology of defined terms to represent gene functions (molecular function, cellular component and biological process). Besides, Kyoto Encyclopedia of Genes and Genomes (KEGG, http://www.genome.jp/kegg/ accessed on 1 March 2021) is a database resource for understanding high-level functions and utilities of the biological system. A package in R language called “clusterprofiler” was used to determine the biological significance of DEGs. The package is capable of providing GO and KEGG enrichment analyses and visualization for users to obtain more valuable biological information [36]. p value < 0.05 was considered a significant enrichment.

4.8. Gene Set Enrichment Analysis (GSEA)

To explore the potential functions of selected key genes and microRNAs in mouse ovaries, and samples of datasets divided into different groups following the expression levels of the key genes, respectively. GSEA was utilized to explore whether priority determined biological processes datasets were enriched in these groups derived from DEGs between the two groups [30]. The criteria were set as p-value < 0.05 and FDR < 0.25.

4.9. miRNA Profiling

Profiling of expressed miRNAs and their target genes were annotated and analyzed by multiple databases: miRDB (http://mirdb.org, accessed on 1 January 2021), miRTarBase (https://bio.tools›mirtarbase, accessed on 1 December 2020), TarBase (https://carolina.imis.athena-innovation.gr, accessed on 10 December 2020), and TargetScan (http://www.targetscan.org, accessed on 20 December 2020).

4.10. Protein Interaction

Prediction of interaction with other functional genes was predicted by STRING database (https://string-db.org, accessed on 1 April 2021).

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ijms221910819/s1.

Author Contributions

Y.-Y.K. and K.-S.K. contributed to experiments design, data generating, data analysis, and manuscript writing. Y.-J.K. contributed to analysis and manuscript writing. S.-W.K. and H.K. contributed to the analysis. S.-Y.K. contributed to experiment design and manuscript writing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by grants from the Ministry of Science, ICT, and Future Planning (2016R1E1A1A01943455, 2019R1A2C1010163 and 2020R1A2C1010293).

Institutional Review Board Statement

The animal experimental plan was reviewed and approved by IACUC of Seoul National University Hospital (No.18-0029-C1A1).

Informed Consent Statement

Not applicable.

Acknowledgments

The authors would like to express sincere thanks to the technical assistance of Myung Jae Chon, In Soon Jang, Jungwoo Kim, Jiwon Lim, and Bo Bin Choi.

Conflicts of Interest

None of the authors have any conflicts of interest.

References

  1. Kim, Y.Y.; Min, H.; Kim, H.; Choi, Y.M.; Liu, H.C.; Ku, S.Y. Differential MicroRNA expression profile of human embryonic stem cell-derived cardiac lineage cells. Tissue Eng. Regen. Med. 2017, 14, 163–169. [Google Scholar] [CrossRef]
  2. Kim, Y.J.; Ku, S.Y.; Rosenwaks, Z.; Liu, H.C.; Chi, S.W.; Kang, J.S.; Lee, W.J.; Jung, K.C.; Kim, S.H.; Choi, Y.M.; et al. MicroRNA expression profiles are altered by gonadotropins and vitamin C status during in vitro follicular growth. Reprod. Sci. 2010, 17, 1081–1089. [Google Scholar] [CrossRef]
  3. Kim, Y.J.; Ku, S.Y.; Kim, Y.Y.; Liu, H.C.; Chi, S.W.; Kim, S.H.; Choi, Y.M.; Kim, J.G.; Moon, S.Y. MicroRNAs transfected into granulosa cells may regulate oocyte meiotic competence during in vitro maturation of mouse follicles. Hum. Reprod. 2013, 28, 3050–3061. [Google Scholar] [CrossRef] [Green Version]
  4. Kim, Y.Y.; Tamadon, A.; Ku, S.Y. Potential Use of Antiapoptotic Proteins and Noncoding RNAs for Efficient In Vitro Follicular Maturation and Ovarian Bioengineering. Tissue Eng. Part B Rev. 2017, 23, 142–158. [Google Scholar] [CrossRef] [PubMed]
  5. Guo, Y.; Sun, J.; Lai, D. Role of microRNAs in premature ovarian insufficiency. Reprod. Biol. Endocrinol. 2017, 15, 38. [Google Scholar] [CrossRef] [PubMed]
  6. Nie, M.; Yu, S.; Peng, S.; Fang, Y.; Wang, H.; Yang, X. miR-23a and miR-27a promote human granulosa cell apoptosis by targeting SMAD5. Biol. Reprod. 2015, 93, 98. [Google Scholar] [CrossRef]
  7. Kim, Y.Y.; Kim, W.O.; Liu, H.C.; Rosenwaks, Z.; Kim, J.W.; Ku, S.Y. Effects of paclitaxel and cisplatin on in vitro ovarian follicle development. Arch. Med. Sci. 2019, 15, 1510–1519. [Google Scholar] [CrossRef]
  8. Armstrong, D.G.; McEvoy, T.G.; Baxter, G.; Robinson, J.J.; Hogg, C.O.; Woad, K.J.; Webb, R.; Sinclair, K.D. Effect of dietary energy and protein on bovine follicular dynamics and embryo production in vitro: Associations with the ovarian insulin-like growth factor system. Biol. Reprod. 2001, 64, 1624–1632. [Google Scholar] [CrossRef] [PubMed]
  9. McCallie, B.; Schoolcraft, W.B.; Katz-Jaffe, M.G. Aberration of blastocyst microRNA expression is associated with human infertility. Fertil. Steril. 2010, 93, 2374–2382. [Google Scholar] [CrossRef] [PubMed]
  10. Miles, J.R.; McDaneld, T.G.; Wiedmann, R.T.; Cushman, R.A.; Echternkamp, S.E.; Vallet, J.L.; Smith, T.P. MicroRNA expression profile in bovine cumulus-oocyte complexes: Possible role of let-7 and miR-106a in the development of bovine oocytes. Anim. Reprod. Sci. 2012, 130, 16–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Khalil, A.M.; Coller, J. Molecular Biology of Long Non-Coding RNAs; Springer: New York, NY, USA, 2013; p. 227. [Google Scholar]
  12. Morris, K.V. Long Non-coding RNAs in Human Disease. In Current Topics in Microbiology and Immunology, 1st ed.; Springer International Publishing: Cham, Switzerland, 2016; Volume 8, 262p. [Google Scholar] [CrossRef]
  13. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  14. Hawkins, S.M.; Matzuk, M.M. Oocyte-somatic cell communication and microRNA function in the ovary. Ann. Endocrinol. 2010, 71, 144–148. [Google Scholar] [CrossRef] [Green Version]
  15. McGinnis, L.K.; Luense, L.J.; Christenson, L.K. MicroRNA in ovarian biology and disease. Cold Spring. Harb. Perspect. Med. 2015, 5, a022962. [Google Scholar] [CrossRef] [Green Version]
  16. Rah, H.; Jeon, Y.J.; Lee, B.E.; Kim, J.O.; Shim, S.H.; Lee, W.S.; Choi, D.H.; Kim, J.H.; Kim, N.K. Association of polymorphisms in microRNA machinery genes (DROSHA, DICER1, RAN, and XPO5) with risk of idiopathic primary ovarian insufficiency in Korean women. Menopause 2013, 20, 1067–1073. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, X.; Zhou, Y.; Peng, S.; Wu, L.; Lin, H.Y.; Wang, S.; Wang, H. Differentially expressed plasma microRNAs in premature ovarian failure patients and the potential regulatory function of mir-23a in granulosa cell apoptosis. Reproduction 2012, 144, 235–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Dang, Y.J.; Zhao, S.D.; Qin, Y.Y.; Han, T.; Li, W.P.; Chen, Z.J. MicroRNA-22-3p is down-regulated in the plasma of Han Chinese patients with premature ovarian failure. Fertil. Steril. 2015, 103, U802–U807. [Google Scholar] [CrossRef] [PubMed]
  19. Pelosi, E.; Forabosco, A.; Schlessinger, D. Genetics of the ovarian reserve. Front. Genet. 2015, 6, 308. [Google Scholar] [CrossRef]
  20. Grive, K.J.; Freiman, R.N. The developmental origins of the mammalian ovarian reserve. Development 2015, 142, 2554–2563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Alvaro Mercadal, B.; Imbert, R.; Demeestere, I.; Gervy, C.; De Leener, A.; Englert, Y.; Costagliola, S.; Delbaere, A. AMH mutations with reduced in vitro bioactivity are related to premature ovarian insufficiency. Hum. Reprod. 2015, 30, 1196–1202. [Google Scholar] [CrossRef] [Green Version]
  22. Yoon, S.H.; Choi, Y.M.; Hong, M.A.; Kim, J.J.; Lee, G.H.; Hwang, K.R.; Moon, S.Y. Association study of anti-Mullerian hormone and anti-Mullerian hormone type II receptor polymorphisms with idiopathic primary ovarian insufficiency. Hum. Reprod. 2013, 28, 3301–3305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Patino, L.C.; Silgado, D.; Laissue, P. A potential functional association between mutant BMPR2 and primary ovarian insufficiency. Syst. Biol. Reprod. Med. 2017, 63, 145–149. [Google Scholar] [CrossRef] [PubMed]
  24. Di Pasquale, E.; Beck-Peccoz, P.; Persani, L. Hypergonadotropic ovarian failure associated with an inherited mutation of human bone morphogenetic protein-15 (BMP15) gene. Am. J. Hum. Genet. 2004, 75, 106–111. [Google Scholar] [CrossRef] [Green Version]
  25. de Mattos, C.S.; Trevisan, C.M.; Peluso, C.; Adami, F.; Cordts, E.B.; Christofolini, D.M.; Barbosa, C.P.; Bianco, B. ESR1 and ESR2 gene polymorphisms are associated with human reproduction outcomes in Brazilian women. J. Ovarian Res. 2014, 7, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhao, H.; Chen, Z.J.; Qin, Y.Y.; Shi, Y.H.; Wang, S.; Choi, Y.; Simpson, J.L.; Rajkovic, A. Transcription factor FIGLA is mutated in patients with Premature Ovarian Failure. Am. J. Hum. Genet. 2008, 82, 1342–1348. [Google Scholar] [CrossRef] [Green Version]
  27. Oostra, B.A.; Willemsen, R. FMR1: A gene with three faces. Biochim. Biophys. Acta 2009, 1790, 467–477. [Google Scholar] [CrossRef]
  28. Welt, C.K. Primary ovarian insufficiency: A more accurate term for premature ovarian failure. Clin. Endocrinol. 2008, 68, 499–509. [Google Scholar] [CrossRef] [PubMed]
  29. Bouilly, J.; Beau, I.; Barraud, S.; Bernard, V.; Azibi, K.; Fagart, J.; Fevre, A.; Todeschini, A.L.; Veitia, R.A.; Beldjord, C.; et al. Identification of multiple gene mutations accounts for a new genetic architecture of primary ovarian insufficiency. J. Clin. Endocrinol. Metab. 2016, 101, 4541–4550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Watkins, W.J.; Umbers, A.J.; Woad, K.J.; Harris, S.E.; Winship, I.M.; Gersak, K.; Shelling, A.N. Mutational screening of FOXO3A and FOXO1A in women with premature ovarian failure. Fertil. Steril. 2006, 86, 1518–1521. [Google Scholar] [CrossRef] [PubMed]
  31. Patino, L.C.; Walton, K.L.; Mueller, T.D.; Johnson, K.E.; Stocker, W.; Richani, D.; Agapiou, D.; Gilchrist, R.B.; Laissue, P.; Harrison, C.A. BMP15 Mutations Associated With Primary Ovarian Insufficiency Reduce Expression, Activity, or Synergy With GDF9. J. Clin. Endocr. Metab. 2017, 102, 1009–1019. [Google Scholar] [CrossRef]
  32. Dixit, H.; Deendayal, M.; Singh, L. Mutational analysis of the mature peptide region of inhibin genes in Indian women with ovarian failure. Hum. Reprod. 2004, 19, 1760–1764. [Google Scholar] [CrossRef] [Green Version]
  33. Wang, B.B.; Li, L.; Zhu, Y.; Zhang, W.; Wang, X.; Chen, B.L.; Li, T.Y.; Pan, H.; Wang, J.; Kee, K.; et al. Sequence variants of KHDRBS1 as high penetrance susceptibility risks for primary ovarian insufficiency by mis-regulating mRNA alternative splicing. Hum. Reprod. 2017, 32, 2138–2146. [Google Scholar] [CrossRef] [PubMed]
  34. Simpson, J.L. Genetic and phenotypic heterogeneity in ovarian failure: Overview of selected candidate genes. Ann. N. Y. Acad. Sci. 2008, 1135, 146–154. [Google Scholar] [CrossRef] [PubMed]
  35. Rossetti, R.; Ferrari, I.; Bonomi, M.; Persani, L. Genetics of primary ovarian insufficiency. Clin. Genet. 2017, 91, 183–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Rajkovic, A.; Pangas, S.A.; Ballow, D.; Suzumori, N.; Matzuk, M.M. NOBOX deficiency disrupts early folliculogenesis and oocyte-specific gene expression. Science 2004, 305, 1157–1159. [Google Scholar] [CrossRef] [Green Version]
  37. Venturella, R.; De Vivo, V.; Carlea, A.; D’Alessandro, P.; Saccone, G.; Arduino, B.; Improda, F.P.; Lico, D.; Rania, E.; De Marco, C.; et al. The Genetics of Non-Syndromic Primary Ovarian Insufficiency: A Systematic Review. Int. J. Fertil. Steril. 2019, 13, 161–168. [Google Scholar] [CrossRef]
  38. Franca, M.M.; Han, X.F.; Funari, M.F.A.; Lerario, A.M.; Nishi, M.Y.; Fontenele, E.G.P.; Domenice, S.; Jorge, A.A.L.; Garcia-Galiano, D.; Elias, C.F.; et al. Exome sequencing reveals the POLR3H gene as a novel cause of primary ovarian insufficiency. J. Clin. Endocr. Metab. 2019, 104, 2827–2841. [Google Scholar] [CrossRef]
  39. Zhao, S.; Li, G.; Dalgleish, R.; Vujovic, S.; Jiao, X.; Li, J.; Simpson, J.L.; Qin, Y.; Ivanisevic, M.; Ivovic, M.; et al. Transcription factor SOHLH1 potentially associated with primary ovarian insufficiency. Fertil. Steril. 2015, 103, 548–553. [Google Scholar] [CrossRef]
  40. Gao, F.; Zhang, J.; Wang, X.; Yang, J.; Chen, D.; Huff, V.; Liu, Y.X. Wt1 functions in ovarian follicle development by regulating granulosa cell differentiation. Hum. Mol. Genet. 2014, 23, 333–341. [Google Scholar] [CrossRef] [Green Version]
  41. Subramanian, A.; Tamayo, P.; Mootha, V.K.; Mukherjee, S.; Ebert, B.L.; Gillette, M.A.; Paulovich, A.; Pomeroy, S.L.; Golub, T.R.; Lander, E.S.; et al. Gene set enrichment analysis: A knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 2005, 102, 15545–15550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Pashaiasl, M.; Ebrahimi, M.; Ebrahimie, E. Identification of the key regulating genes of diminished ovarian reserve (DOR) by network and gene ontology analysis. Mol. Biol. Rep. 2016, 43, 923–937. [Google Scholar] [CrossRef]
  43. Samartzis, E.P.; Labidi-Galy, S.I.; Moschetta, M.; Uccello, M.; Kalaitzopoulos, D.R.; Perez-Fidalgo, J.A.; Boussios, S. Endometriosis-associated ovarian carcinomas: Insights into pathogenesis, diagnostics, and therapeutic targets-a narrative review. Ann. Transl. Med. 2020, 8, 1712. [Google Scholar] [CrossRef] [PubMed]
  44. Cho, S.H.; Ahn, E.H.; An, H.J.; Kim, J.H.; Ko, J.J.; Kim, Y.R.; Lee, W.S.; Kim, N.K. Association of miR-938G > A Polymorphisms with Primary Ovarian Insufficiency (POI)-Related Gene Expression. Int. J. Mol. Sci. 2017, 18, 1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Norling, A.; Hirschberg, A.L.; Rodriguez-Wallberg, K.A.; Iwarsson, E.; Wedell, A.; Barbaro, M. Identification of a duplication within the GDF9 gene and novel candidate genes for primary ovarian insufficiency (POI) by a customized high-resolution array comparative genomic hybridization platform. Hum. Reprod. 2014, 29, 1818–1827. [Google Scholar] [CrossRef] [Green Version]
  46. Tripurani, S.K.; Lee, K.B.; Wee, G.; Smith, G.W.; Yao, J. MicroRNA-196a regulates bovine newborn ovary homeobox gene (NOBOX) expression during early embryogenesis. BMC Dev. Biol. 2011, 11, 25. [Google Scholar] [CrossRef] [Green Version]
  47. Cenik, E.S.; Zamore, P.D. Argonaute proteins. Curr. Biol. 2011, 21, R446–R449. [Google Scholar] [CrossRef] [Green Version]
  48. Kawamura, Y.; Saito, K.; Kin, T.; Ono, Y.; Asai, K.; Sunohara, T.; Okada, T.N.; Siomi, M.C.; Siomi, H. Drosophila endogenous small RNAs bind to Argonaute 2 in somatic cells. Nature 2008, 453, 793–797. [Google Scholar] [CrossRef]
  49. Kuramochi-Miyagawa, S.; Kimura, T.; Yomogida, K.; Kuroiwa, A.; Tadokoro, Y.; Fujita, Y.; Sato, M.; Matsuda, Y.; Nakano, T. Two mouse piwi-related genes: Miwi and mili. Mech. Dev. 2001, 108, 121–133. [Google Scholar] [CrossRef]
  50. Yang, Q.; Hua, J.; Wang, L.; Xu, B.; Zhang, H.; Ye, N.; Zhang, Z.; Yu, D.; Cooke, H.J.; Zhang, Y.; et al. MicroRNA and piRNA profiles in normal human testis detected by next generation sequencing. PLoS ONE 2013, 8, e66809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Zheng, K.; Wang, P.J. Blockade of pachytene piRNA biogenesis reveals a novel requirement for maintaining post-meiotic germline genome integrity. PLoS Genet. 2012, 8, e1003038. [Google Scholar] [CrossRef] [Green Version]
  52. Malki, S.; van der Heijden, G.W.; O’Donnell, K.A.; Martin, S.L.; Bortvin, A. A role for retrotransposon LINE-1 in fetal oocyte attrition in mice. Dev. Cell 2014, 29, 521–533. [Google Scholar] [CrossRef] [Green Version]
  53. Yu, G.C.; Wang, L.G.; Han, Y.Y.; He, Q.Y. clusterProfiler: An R package for comparing biological themes among gene clusters. Omics 2012, 16, 284–287. [Google Scholar] [CrossRef] [PubMed]
  54. Hu, W.; Gauthier, L.; Baibakov, B.; Jimenez-Movilla, M.; Dean, J. FIGLA, a basic helix-loop-helix transcription factor, balances sexually dimorphic gene expression in postnatal oocytes. Mol. Cell. Biol. 2010, 30, 3661–3671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Huntriss, J.; Gosden, R.; Hinkins, M.; Oliver, B.; Miller, D.; Rutherford, A.J.; Picton, H.M. Isolation, characterization and expression of the human Factor In the Germline alpha (FIGLA) gene in ovarian follicles and oocytes. Mol. Hum. Reprod. 2002, 8, 1087–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Qiu, Y.; Sun, S.; Charkraborty, T.; Wu, L.; Sun, L.; Wei, J.; Nagahama, Y.; Wang, D.; Zhou, L. Figla Favors Ovarian Differentiation by Antagonizing Spermatogenesis in a Teleosts, Nile Tilapia (Oreochromis niloticus). PLoS ONE 2015, 10, e0123900. [Google Scholar] [CrossRef] [Green Version]
  57. Reilly, J.N.; McLaughlin, E.A.; Stanger, S.J.; Anderson, A.L.; Hutcheon, K.; Church, K.; Mihalas, B.P.; Tyagi, S.; Holt, J.E.; Eamens, A.L.; et al. Characterisation of mouse epididymosomes reveals a complex profile of microRNAs and a potential mechanism for modification of the sperm epigenome. Sci. Rep. 2016, 6, 31794. [Google Scholar] [CrossRef] [PubMed]
  58. Mikaeili, S.; Rashidi, B.H.; Safa, M.; Najafi, A.; Sobhani, A.; Asadi, E.; Abbasi, M. Altered FoxO3 expression and apoptosis in granulosa cells of women with polycystic ovary syndrome. Arch. Gynecol. Obstet. 2016, 294, 185–192. [Google Scholar] [CrossRef]
  59. Ezzati, M.M.; Baker, M.D.; Saatcioglu, H.D.; Aloisio, G.M.; Pena, C.G.; Nakada, Y.; Cuevas, I.; Carr, B.R.; Castrillon, D.H. Regulation of FOXO3 subcellular localization by Kit ligand in the neonatal mouse ovary. J. Assist. Reprod. Genet. 2015, 32, 1741–1747. [Google Scholar] [CrossRef] [Green Version]
  60. Albamonte, M.I.; Calabro, L.Y.; Albamonte, M.S.; Zuccardi, L.; Stella, I.; Halperin, J.; Vitullo, A.D. PTEN and FOXO3 expression in the prenatal and postnatal human ovary. J. Assist. Reprod. Genet. 2020, 37, 1613–1622. [Google Scholar] [CrossRef]
  61. Li, C.; Zhang, Y.; Zhao, W.; Cui, S.; Song, Y. miR-153-3p regulates progression of ovarian carcinoma in vitro and in vivo by targeting MCL1 gene. J. Cell. Biochem. 2019, 120, 19147–19158. [Google Scholar] [CrossRef]
  62. Nandedkar, T.D.; Moodbidri, S.B.; Raghavan, V.P. FSH binding to granulosa cells of normal and atretic follicles of mouse ovary in vivo and in vitro. Indian J. Exp. Biol. 1985, 23, 685–688. [Google Scholar]
  63. Ahn, H.W.; Morin, R.D.; Zhao, H.; Harris, R.A.; Coarfa, C.; Chen, Z.J.; Milosavljevic, A.; Marra, M.A.; Rajkovic, A. MicroRNA transcriptome in the newborn mouse ovaries determined by massive parallel sequencing. Mol. Hum. Reprod. 2010, 16, 463–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Mineno, J.; Okamoto, S.; Ando, T.; Sato, M.; Chono, H.; Izu, H.; Takayama, M.; Asada, K.; Mirochnitchenko, O.; Inouye, M.; et al. The expression profile of microRNAs in mouse embryos. Nucl. Acids Res. 2006, 34, 1765–1771. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, Y.Y.; Kim, Y.J.; Cho, K.M.; Kim, S.H.; Park, K.E.; Kang, B.C.; Jung, K.C.; Kim, M.S.; Ku, S.Y. The expression profile of angiotensin system on thawed murine ovaries. Tissue Eng. Regen. Med. 2016, 13, 724–731. [Google Scholar] [CrossRef]
  66. Kim, Y.J.; Kim, Y.Y.; Kang, B.C.; Kim, M.S.; Ko, I.K.; Liu, H.C.; Rosenwaks, Z.; Ku, S.Y. Induction of multiple ovulation via modulation of angiotensin II receptors in in vitro ovarian follicle culture models. J. Tissue Eng. Regen. Med. 2017, 11, 3100–3110. [Google Scholar] [CrossRef] [PubMed]
  67. Rajapakse, S.; Yamano, N.; Ogiwara, K.; Hirata, K.; Takahashi, S.; Takahashi, T. Estrogen-dependent expression of the tissue kallikrein gene (Klk1) in the mouse uterus and its implications for endometrial tissue growth. Mol. Reprod. Dev. 2007, 74, 1053–1063. [Google Scholar] [CrossRef]
  68. Dedio, J.; Wiemer, G.; Rutten, H.; Dendorfer, A.; Scholkens, B.A.; Muller-Esterl, W.; Wohlfart, P. Tissue kallikrein KLK1 is expressed de novo in endothelial cells and mediates relaxation of human umbilical veins. Biol. Chem. 2001, 382, 1483–1490. [Google Scholar] [CrossRef]
  69. Gomes Fernandes, M.; He, N.; Wang, F.; Van Iperen, L.; Eguizabal, C.; Matorras, R.; Roelen, B.A.J.; Chuva De Sousa Lopes, S.M. Human-specific subcellular compartmentalization of P-element induced wimpy testis-like (PIWIL) granules during germ cell development and spermatogenesis. Hum. Reprod. 2018, 33, 258–269. [Google Scholar] [CrossRef] [Green Version]
  70. Jin, Y.H.; Davie, A.; Migaud, H. Expression pattern of nanos, piwil, dnd, vasa and pum genes during ontogenic development in Nile tilapia Oreochromis niloticus. Gene 2019, 688, 62–70. [Google Scholar] [CrossRef] [Green Version]
  71. Yang, X.G.; Yue, H.M.; Ye, H.; Shan, X.S.; Xie, X.; Li, C.J.; Wei, Q.W. Identification and characterization of two piwi genes and their expression in response to E2 (17 beta-estradiol) in Dabry’s sturgeon Acipenser dabryanus. Fish. Sci. 2020, 86, 307–317. [Google Scholar] [CrossRef]
  72. Yiew, N.K.H.; Chatterjee, T.K.; Tang, Y.L.; Pellenberg, R.; Stansfield, B.K.; Bagi, Z.; Fulton, D.J.; Stepp, D.W.; Chen, W.; Patel, V.; et al. A novel role for the Wnt inhibitor APCDD1 in adipocyte differentiation: Implications for diet-induced obesity. J. Biol. Chem. 2017, 292, 6312–6324. [Google Scholar] [CrossRef] [Green Version]
  73. Fang, X.; Ni, N.; Lydon, J.P.; Ivanov, I.; Bayless, K.J.; Rijnkels, M.; Li, Q. Enhancer of zeste 2 polycomb repressive complex 2 subunit is required for uterine epithelial integrity. Am. J. Pathol. 2019, 189, 1212–1225. [Google Scholar] [CrossRef] [Green Version]
  74. Cousminer, D.L.; Stergiakouli, E.; Berry, D.J.; Ang, W.; Groen-Blokhuis, M.M.; Korner, A.; Siitonen, N.; Ntalla, I.; Marinelli, M.; Perry, J.R.; et al. Genome-wide association study of sexual maturation in males and females highlights a role for body mass and menarche loci in male puberty. Hum. Mol. Genet. 2014, 23, 4452–4464. [Google Scholar] [CrossRef] [PubMed]
  75. Yan, L.R.; Wang, A.; Lv, Z.; Yuan, Y.; Xu, Q. Mitochondria-related core genes and TF-miRNA-hub mrDEGs network in breast cancer. Biosci. Rep. 2021, 41, 1–17. [Google Scholar] [CrossRef]
  76. Yun, J.W.; Kim, Y.Y.; Ahn, J.H.; Kang, B.C.; Ku, S.Y. Use of nonhuman primates for the development of bioengineered female reproductive organs. Tissue Eng. Regen. Med. 2016, 13, 323–334. [Google Scholar] [CrossRef] [PubMed]
  77. Kim, Y.Y.; Yun, J.W.; Kim, J.M.; Park, C.G.; Rosenwaks, Z.; Liu, H.C.; Kang, B.C.; Ku, S.Y. Gonadotropin ratio affects the in vitro growth of rhesus ovarian preantral follicles. J. Investig. Med. 2016, 64, 888–893. [Google Scholar] [CrossRef] [Green Version]
  78. Kim, Y.Y.; Kang, B.C.; Yun, J.W.; Ahn, J.H.; Kim, Y.J.; Kim, H.; Rosenwaks, Z.; Ku, S.Y. Expression of transcripts in marmoset oocytes retrieved during follicle isolation without gonadotropin induction. Int. J. Mol. Sci. 2019, 20, 1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Experimental scheme of this study.
Figure 1. Experimental scheme of this study.
Ijms 22 10819 g001
Figure 2. Correlation among analyzed genes in the reproductively aged mouse ovary. (A) Count data and correlation among each ovary stage, indicated by based on a heatmap, (B) Count data indicated by a PCA plot.
Figure 2. Correlation among analyzed genes in the reproductively aged mouse ovary. (A) Count data and correlation among each ovary stage, indicated by based on a heatmap, (B) Count data indicated by a PCA plot.
Ijms 22 10819 g002
Figure 3. Up- and downregulated transcripts in the reproductively aged mouse ovary. Differently expressed transcripts among the differently aged mouse ovaries, based on the log2 fold changes and 5 fold up-and downregulation, filtered based on FPKM values. (A) 6 weeks (W) vs. 2 W, (B) 15 W vs. 2 W, (C) 20 W vs. 2 W, (D) 15 W vs. 6 W, (E) 20 W vs. 6 W, (F) 20 W vs. 15 W.
Figure 3. Up- and downregulated transcripts in the reproductively aged mouse ovary. Differently expressed transcripts among the differently aged mouse ovaries, based on the log2 fold changes and 5 fold up-and downregulation, filtered based on FPKM values. (A) 6 weeks (W) vs. 2 W, (B) 15 W vs. 2 W, (C) 20 W vs. 2 W, (D) 15 W vs. 6 W, (E) 20 W vs. 6 W, (F) 20 W vs. 15 W.
Ijms 22 10819 g003
Table 2. Numbers of differentially expressed genes (DEGs) based on comparisons between the reproductive stages of the mouse ovary. (A) Total expression and upregulated and downregulated DEG numbers based on comparison for each aged mouse ovaries. (B) Quantitative graph of up- and downregulated DEG numbers.
Table 2. Numbers of differentially expressed genes (DEGs) based on comparisons between the reproductive stages of the mouse ovary. (A) Total expression and upregulated and downregulated DEG numbers based on comparison for each aged mouse ovaries. (B) Quantitative graph of up- and downregulated DEG numbers.
A6 W vs. 2 W15 W vs. 2 W20 W vs. 2 W15 W vs. 6 W20 W vs. 6 W20 W vs. 15 W
Up-regulated545925696208170114
Down-regulated44774572719921765
Total expression99216701423407387179
Ijms 22 10819 i001
Table 3. Predicted protein interaction for up- and downregulated transcripts in the reproductively aged mouse ovary. Table 2. Fold changes among the groups were selected and their predictive interaction was analyzed using the STRING database. Redundant genes were indicated when they first appeared in the table. (A) 6 weeks (W) vs. 2 W, (B) 15 W vs. 2 W, (C) 20 W vs. 2 W, (D) 15 W vs. 6 W, (E) 20 W vs. 6 W, (F) 20 W vs. 15 W.
Table 3. Predicted protein interaction for up- and downregulated transcripts in the reproductively aged mouse ovary. Table 2. Fold changes among the groups were selected and their predictive interaction was analyzed using the STRING database. Redundant genes were indicated when they first appeared in the table. (A) 6 weeks (W) vs. 2 W, (B) 15 W vs. 2 W, (C) 20 W vs. 2 W, (D) 15 W vs. 6 W, (E) 20 W vs. 6 W, (F) 20 W vs. 15 W.
(A) 6 W vs. 2 W
Up-Regulation
GeneCategoryBiological Functions
BhmtBiological processHomocysteine catabolic process
Molecular FunctionOxidoreductase activity, acting on the CH-NH group of donors, flavin as acceptor
KEGG pathwayCystein and methionine metabolism
Spp1Biological processCalcium-independent cell-matrix adhesion
Molecular FunctionExtracellular matrix binding
KEGG pathwayECM-receptor interaction
Klk1Biological processPositive regulation of steroid hormone biosynthetic process
Molecular FunctionBradykinin receptor activity
KEGG pathwayRenin-angiotensin system
Serpina5Biological processBlood coagulation, fibrin clot formation
Molecular FunctionSerine-type endopeptidase activity
KEGG pathwayComplement and coagulation cascades
Sfrp4Biological processPositive regulation of dermatome development
Molecular FunctionChemoattractant activity involved in axon guidance
KEGG pathwayBasal cell carcinoma
Plin4Biological processNegative regulation of sequestering of triglyceride
Molecular FunctionTryglyceride lipase activity
KEGG pathwayRegulation of lipolysis in adipocytes
Wisp2Biological processRegulation of dermatome development
Molecular FunctionFibronectin binding
KEGG pathwayBladder cancer
GM1987
(Ccl21a)
Biological processLymphocyte chemotaxis across high endothelial venule
Molecular FunctionCXCR chemokine receptor binding
KEGG pathwayChemokine signaling pathway
Down-Regulation
GeneCategoryBiological Functions
BC100530, Stfa1Biological processNegative regulation of endopeptidase activity
Molecular FunctionCysteine-type endopeptidase inhibitor activity
FiglaBiological processNegative regulation of fertilization
Molecular FunctionAcrosin binding
KEGG pathwayOvarian steroidogenesis
S100a9, S100a8Biological processNeutrophil aggregation
Molecular FunctionToll-like receptor 4 binding
KEGG pathwayIL-17 signaling pathway
(B) 15 W vs. 2 W
Up-Regulation
GeneCategoryBiological Functions
Wnt10bBiological processWnt signaling pathway involved in midbrain dopaminergic neuron differentiation
Molecular FunctionFrizzled binding
KEGG pathwayBasal cell carcinoma
H2-T9 (Gm7030)Biological processAntigen processing and presentation of exogenous protein antigen via MHC class
Molecular FunctionMHC class I protein binding
KEGG pathwayAntigen processing and presentation
Atp1a3Biological processPositive regulation of sodium ion export across plasma membrane
Molecular FunctionSodium:potassium-exchanging atpase activity
KEGG pathwayProximal tubule bicarbonate reclamation
PtgfrBiological processPhospholipase c-activating angiotensin-activated signaling pathway
Molecular FunctionAngiotensin type I receptor activity
KEGG pathwayRenin-angiotensin system
LipgBiological processTriglyceride catabolic process
Molecular FunctionLipoprotein lipase activity
KEGG pathwayGlycerolipid metabolism
Abcb1bBiological processDrug export
Molecular FunctionXenobiotic transmembrane transporting atpase activity
KEGG pathwayBile secretion
Hsd17b7Biological processEstrogen biosynthetic process
Molecular FunctionEstradiol 17-beta-dehydrogenase activity
KEGG pathwaySteroid biosynthesis
StarBiological processVesicle tethering to endoplasmic reticulum
Molecular FunctionPorin activity
KEGG pathwayCholesterol metabolism
Plin4, Bhmt, Sfrp4Functions are indicated in previous page
Down-Regulation
GeneCategoryBiological Functions
Atf3Biological processPositive regulation of transcription from RNA polymerase
Molecular FunctionCamp response element binding protein binding
KEGG pathwayCocaine addiction
Fos, FosbBiological processResponse to gravity
Molecular FunctionNad-dependent histone deacetylase activity
KEGG pathwayAmphetamine addiction
S100a9, S100a8Functions are indicated in previous page
(C) 20 W vs. 2 W
Up-Regulation
GeneCategoryBiological Functions
Sprr1aBiological processPeptide cross-linking
Wisp2Biological processRegulation of dermatome development
Molecular FunctionFibronectin binding
KEGG pathwayBladder cancer
InhbaBiological processRegulation of follicle-stimulating hormone secretion
Molecular FunctionInhibin binding
KEGG pathwayTGF-beta signaling pathway
Mgst2Biological processXenobiotic catabolic process
Molecular FunctionGlutathione disulfide oxidoreductase activity
KEGG pathwayMetabolism of xenobiotics by cytochrome P450
Akr1c18Biological processPolyprenol catabolic process
Molecular FunctionEnone reductase activity
KEGG pathwaySteroid hormone biosynthesis
Sfrp4Biological processPositive regulation of dermatome development
Molecular FunctionChemoattractant activity involved in axon guidance
KEGG pathwayBasal cell carcinoma
Wnt10b, Plin4, Lipg, Bhmt,
Abcb1b, Spp1, Ptgfr, Hsd17b7
Functions are indicated in previous page
Down-Regulation
GeneCategoryBiological Functions
BC100530Biological processNegative regulation of endopeptidase activity
Molecular FunctionCysteine-type endopeptidase inhibitor activity
S100a8Functions are indicated in previous page
(D) 15 W vs. 6 W
Up-Regulation
GeneCategoryBiological Functions
Tgoln1
(Tgoln2)
Biological processGolgi ribbon formation
Molecular FunctionMannose binding
KEGG pathwaySNARE interactions in vesicular transport
Eif3j1Biological processViral translational termination-reinitiation
Molecular FunctionTranslation initiation factor activity
KEGG pathwayRNA transport
Nme2Biological processITP metabolic process
Molecular FunctionCTP synthase activity
KEGG pathwayPyrimidine metabolism
Tmem254bUnknown
H2-T9 (Gm7030)Functions are indicated in previous page
Down-Regulation
GeneCategoryBiological Functions
Plac9aUnknown
Atf3, Fosb,
GM1987 (Ccl21a), Fos
Functions are indicated in previous page
(E) 20 W vs. 6 W
Up-Regulation
GeneCategoryBiological Functions
Tmem254bUnknown
Eif3j1, Sprr1aFunctions are indicated in previous page
Down-Regulation
GeneCategoryBiological Functions
Ooog1Biological processFemale gamete generation
GM1987 (Ccl21a)Functions are indicated in previous page
(F) 20 W vs. 15 W
Up-Regulation
GeneCategoryBiological Functions
Cxcl1Biological processInterleukin-8-mediated signaling pathway
Molecular FunctionInterleukin-8 receptor activity
KEGG pathwayIL-17 signaling pathway
Atf3Biological processPositive regulation of transcription from RNA polymerase II promoter in response to endoplasmic reticulum stress
Molecular FunctionCamp response element binding protein binding
KEGG pathwayCocaine addiction
Hspa1a, Hspa1bBiological processTelomerase holoenzyme complex assembly
Molecular FunctionCTP binding
KEGG pathwayPrion diseases
Zfp36Biological processEpithelial cell proliferation involved in salivary gland morphogenesis
Molecular FunctionMrna 3’-UTR au-rich region binding
KEGG pathwayAntifolate resistance
Sprr1a, Fosb, GM1987 (Ccl21a), Nme2Functions are indicated in previous page
Down-Regulation
GeneCategoryBiological Functions
GM1987 (Ccl21a), Nme2Functions are indicated in previous page
Table 4. Key regulatory genes expressed in the reproductively-aged mouse ovary, analyzed by FPKM and TPM.
Table 4. Key regulatory genes expressed in the reproductively-aged mouse ovary, analyzed by FPKM and TPM.
GenesBmp15Ccnd2FiglaFoxo1Foxo3FshrGdf9Igf1InhaNoboxSmad3Taf4bPou5f1
2 WFPKM14.6444.1451.8419.510.553.07130.5715.29533.0133.3328.043.8224.29
6 W20.0244.181.6866.927.968.7152.1916.041396.527.4216.714.55.17
20 W9.5547.871.882.8910.8921.8733.55161915.616.4726.579.523.62
2 WTPM20.3361.2971.9727.0814.654.27181.2821.23740.0446.2738.935.3133.73
6 W31.0368.482.61103.7312.3413.4980.8924.872164.7711.5125.916.978.02
20 W16.4682.453.09142.7818.7537.6857.7927.553299.6511.1445.7616.46.23
Table 5. Correlated miRNAs that interact with key target genes.
Table 5. Correlated miRNAs that interact with key target genes.
GeneMature miRNAs
FPKMCcnd2mmu-miR-1251-5p, mmu-miR-503-5p, mmu-miR-497a-5p, mmu-miR-195b, mmu-miR-322-5p, mmu-miR-15a-5p, mmu-miR-16-5p, mmu-miR-195a-5p, mmu-miR-195a-5p, mmu-miR-6955-3p, mmu-miR-29a-3p, mmu-miR-29b-3p, mmu-miR-29c-3p, mmu-miR-302c-3p, mmu-miR-302b-3p, mmu-miR-302d-3p, mmu-miR-302a-3p, mmu-miR-294-3p, mmu-miR-295-3p, mmu-miR-291a-3p, mmu-miR-106a-5p, mmu-miR-106b-5p, mmu-miR-93-5p, mmu-miR-20a-5p, mmu-miR-20b-5p, mmu-miR-17-5p, mmu-miR-19b-3p, mmu-miR-19a-3p, mmu-miR-503-5p, mmu-miR-195a-5p, mmu-miR-16-5p, mmu-miR-195b, mmu-miR-322-5p, mmu-miR-15a-5p, mmu-miR-497a-5p, mmu-miR-322-5p, mmu-miR-195a-5p, mmu-miR-195b, mmu-miR-15a-5p, mmu-miR-16-5p, mmu-miR-15b-5p, mmu-miR-503-5p, mmu-miR-182-5p, mmu-miR-9-3p, mmu-miR-124-3p.1, mmu-miR-1193-3p
Figlammu-miR-669d-2-3p
Foxo1mmu-miR-370-3p, mmu-miR-183-5p, mmu-miR-183-5p.2, mmu-miR-96-5p, mmu-miR-144-3p, mmu-miR-27a-3p, mmu-miR-27b-3p, mmu-miR-411-5p, mmu-miR-145b, mmu-miR-145a-5p, mmu-miR-6715-5p, mmu-miR-135a-5p, mmu-miR-135b-5p
Foxo3mmu-miR-153-3p
Fshrmmu-miR-670-3p
Igf1mmu-miR-365-3p, mmu-miR-1a-3p, mmu-miR-206-3p, mmu-miR-483-3p.2, mmu-miR-18a-5p, mmu-miR-18b-5p, mmu-miR-452-5p, mmu-miR-29c-3p, mmu-miR-29b-3p, mmu-miR-29a-3p, mmu-miR-495-3p, mmu-miR-378c, mmu-miR-378a-3p, mmu-miR-1839-5p, mmu-let-7a-5p, mmu-let-7c-5p, mmu-miR-98-5p, mmu-let-7i-5p, mmu-let-7g-5p, mmu-let-7d-5p, mmu-let-7b-5p, mmu-let-7k, mmu-let-7f-5p, mmu-let-7e-5p, mmu-miR-192-5p, mmu-miR-215-5p, mmu-miR-29b-3p, mmu-miR-29a-3p, mmu-miR-29c-3p, mmu-miR-142a-5p, mmu-miR-340-5p, mmu-miR-489-3p, mmu-miR-425-5p, mmu-miR-186-5p
Smad3mmu-miR-129-2-3p, mmu-miR-129-1-3p, mmu-miR-145a-5p, mmu-miR-145b, mmu-miR-3154
Taf4bmmu-miR-148a-3p, mmu-miR-152-3p, mmu-miR-148b-3p
Pou5f1mmu-miR-218-5p, mmu-miR-1955-5p, mmu-miR-881-3p, mmu-miR-186-5p
Table 6. Correlation of Piwil gene family expression and frequently expressed target genes and microRNAs in the reproductively aged mouse ovary. Target genes of the Piwil gene family and interactive miRNAs of each Piwil gene were analyzed based on the FPKM correlation (p < 0.05, score 90+).
Table 6. Correlation of Piwil gene family expression and frequently expressed target genes and microRNAs in the reproductively aged mouse ovary. Target genes of the Piwil gene family and interactive miRNAs of each Piwil gene were analyzed based on the FPKM correlation (p < 0.05, score 90+).
GeneTarget Genes and Mature miRNAs
FPKMTarget genes
FOXL2, FOXO3, NANOG, POU5F1, LIN28A, SOX2, SOX3, KLF4, PRDM1, NANOS3, UTF1, CD38, DAZL,
DDX4, SYCP3, RNF17, TDRD5, TDRD9, SOX17, GATA4, TEAD4, TFAP2C, LHX9, WT1
Interactive miRNAs
Piwil1mmu-miR-1249-5p, mmu-miR-147-5p, mmu-miR-1892, mmu-miR-3084-3p, mmu-miR-330-5p,
mmu-miR-6540-3p, mmu-miR-7116-3p, mmu-miR-7669-5p
Piwil2mmu-miR-103-2-5p, mmu-miR-1198-5p, mmu-miR-139-5p, mmu-miR-145a-5p, mmu-miR-149-5p,
mmu-miR-1b-5p, mmu-miR-216b-5p, mmu-miR-223-5p, mmu-miR-23a-3p, mmu-miR-26a-5p,
mmu-miR-26b-3p, mmu-miR-346-5p, mmu-miR-425-3p, mmu-miR-466d-3p, mmu-miR-466f-3p,
mmu-miR-466g, mmu-miR-7065-3p, mmu-miR-760-3p, mmu-miR-7647-3p, mmu-miR-7661-5p,
mmu-miR-935
Piwil4mmu-miR-210-5p, mmu-miR-3470b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, Y.-Y.; Kim, K.-S.; Kim, Y.-J.; Kim, S.-W.; Kim, H.; Ku, S.-Y. Transcriptome Analyses Identify Potential Key microRNAs and Their Target Genes Contributing to Ovarian Reserve. Int. J. Mol. Sci. 2021, 22, 10819. https://doi.org/10.3390/ijms221910819

AMA Style

Kim Y-Y, Kim K-S, Kim Y-J, Kim S-W, Kim H, Ku S-Y. Transcriptome Analyses Identify Potential Key microRNAs and Their Target Genes Contributing to Ovarian Reserve. International Journal of Molecular Sciences. 2021; 22(19):10819. https://doi.org/10.3390/ijms221910819

Chicago/Turabian Style

Kim, Yoon-Young, Kwang-Soo Kim, Yong-Jin Kim, Sung-Woo Kim, Hoon Kim, and Seung-Yup Ku. 2021. "Transcriptome Analyses Identify Potential Key microRNAs and Their Target Genes Contributing to Ovarian Reserve" International Journal of Molecular Sciences 22, no. 19: 10819. https://doi.org/10.3390/ijms221910819

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop