Next Article in Journal
Exosomal microRNAs miR-30d-5p and miR-126a-5p Are Associated with Heart Failure with Preserved Ejection Fraction in STZ-Induced Type 1 Diabetic Rats
Previous Article in Journal
Changes of Subjective Symptoms and Tear Film Biomarkers following Femto-LASIK
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Early Gonadal Development and Sex Determination in Mammal

1
National Engineering Research Center for Breeding Swine Industry, College of Animal Science, South China Agricultural University, Guangzhou 510630, China
2
Guangdong Provincial Key Laboratory of Agro-Animal Genomics and Molecular Breeding, South China Agricultural University, Guangzhou 510630, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(14), 7500; https://doi.org/10.3390/ijms23147500
Submission received: 6 June 2022 / Revised: 29 June 2022 / Accepted: 5 July 2022 / Published: 6 July 2022
(This article belongs to the Section Molecular Biology)

Abstract

:
Sex determination is crucial for the transmission of genetic information through generations. In mammal, this process is primarily regulated by an antagonistic network of sex-related genes beginning in embryonic development and continuing throughout life. Nonetheless, abnormal expression of these sex-related genes will lead to reproductive organ and germline abnormalities, resulting in disorders of sex development (DSD) and infertility. On the other hand, it is possible to predetermine the sex of animal offspring by artificially regulating sex-related gene expression, a recent research hotspot. In this paper, we reviewed recent research that has improved our understanding of the mechanisms underlying the development of the gonad and primordial germ cells (PGCs), progenitors of the germline, to provide new directions for the treatment of DSD and infertility, both of which involve manipulating the sex ratio of livestock offspring.

1. Introduction

The most important issues in human reproductive medicine are disorders of sex development (DSD) and infertility. According to data, the proportion of DSD patients ranges between 1:2000–1:4500 [1]; however, the underlying mechanisms of this congenital disease remain unclear, and genetic diagnostic cannot be performed on up to 75% of patients [2]. As for infertility, 8–12% of couples reproductive-aged are affected by this condition globally [3], which can also be caused by genetic background. In order to diagnose and treat DSD and infertility, a substantial amount of research has been devoted to sex determination, whose abnormality can result in these. In mammal, this process is governed by sex chromosomes [4] and involves the sexual differentiation of bipotential gonads and primordial germ cells (PGCs) [5]. In livestock production, the molecular mechanism of sex determination has been a research hotspot for many years because it has the potential to significantly increase production efficiency; however, a lack of theoretical knowledge prevents commercial application impossible [6]. In this review, we focused primarily on molecular, cellular, and genetic studies on early gonadal development and sexual differentiation of bipotential gonads, both with PGC formation, migration, and gender-specific differentiation. Although some studies in this field have been reviewed elsewhere [7,8,9,10,11,12], we summarized these review articles and combined single-cell RNA sequencing results on the basis of the original research to discuss some potential mechanistic links between bipotential gonads and PGC and find genes that play key roles in multiple stages of early gonadal development and sex determination. On the other hand, we described novel gene editing experiments that lead to sex reversal, potential gender differences before sex-specific differentiation, and the function of epigenetic regulation in these processes, with the hope of informing future studies on abnormal development of reproductive organs and infertility in humans, as well as sex manipulation technologies in livestock.

2. Genital Ridge Formation

In mammal, both the testis and ovary develop from the genital ridge (GR), which first appears at approximately four-five weeks of pregnancy in humans [13] and around embryonic day (E) nine and a half in mice [8], when coelomic epithelial cells begin to proliferate on the ventromedial surface of the mesonephros [14,15,16]. Each mesonephros contains a Wolffian duct and a Müllerian duct, which give rise to the epididymis, vas deferens, and seminal vesicles [17,18] or the fallopian tubes, uterus, and a portion of the vagina [19]. At around 32 days of pregnancy in humans (E10.5 in mice), coelomic epithelial cells differentiate into two distinct somatic precursor lineages (supporting cell precursors and steroidogenic cell precursors) [20,21]. Recent single-cell sequencing research has confirmed that mammalian gonadal cells originate from the same progenitor [22,23]. However, little is known about the development of the gonad prior to sex determination, as the bipotential gonad contains multiple uncharacterized subpopulations lacking specific markers [22]. In the last few decades, knockout mice models and mutation data from DSD patients have been used extensively to investigate transcription factors essential for genital ridge formation. These important genes are detailed in Table 1 and Figure 1. However, these genes, such as GATA4, POD1, PBX1, and ODD1, were also involved in the formation of various organs [24], which hinders the effectiveness of research in the field.

3. Differentiation of the Bipotential Gonads

At approximately 6–7 weeks of pregnancy in humans (E12.5 in mice), testis cords are observed in XY gonads, indicating the start of sex differentiation in bipotential gonads [51,52,53,54]. Controlled by gene expression dynamics and antagonistic genetic programs, sexual differentiation begins when the antagonistic network’s balance is tilted toward males or females. Furthermore, these antagonistic genetic programs will be maintained into adulthood to preserve gonad stability and reproductive capacity [55]. The SRY on the Y chromosome is the “master switch” for testis determination in mammal. When SRY is expressed in bipotential gonads during a critical window of fetal development, SOX9 expression and a male-promoting regulatory network are activated, resulting in testis differentiation. In contrast, ovary differentiation will be triggered when the balance is tilted towards a female-promoting regulatory network [56]. However, recent research has not identified genes with a similar function role to SRY in female sexual differentiation. By inhibiting SOX9 expression, WNT4, RSPO1, and FOXL2 were previously involved in the ovarian-determination pathway.

3.1. Sex-Determining Region Y, SRY, and Sox9

In 1989, a 35-kb region from the Y chromosome was identified in 46,XX DSD patients, and it was believed to be the possible carrier of the testis-determining factor gene [57]. A year later, an open reading frame (ORF) encoding a new gene, later designated SRY (sex-determining region Y), was discovered in this Y fragment [58]. The SRY encodes a transcription factor with an N-terminal domain (NTD), a high conserved mobility group (HMG) domain, and a C-terminal domain (CTD) [59,60]. The SRY mutation analysis revealed that the NTD in SRY is associated with nuclear importation [61,62], whereas the CTD may contribute to the conformation and function of SRY [63] and be required for SOX9 activation [60]. Most human male-to-female sex reversal syndrome cases are believed to be caused by a mutation located in the HMG domain of SRY [64,65]. Recent gene-editing research demonstrated that knockout of the HMG domain of the porcine SRY gene could result in male-to-female sex reversal [66]; however, additional tests are necessary to confirm the fertility of these transgenic animals. Interestingly, a study found that XX mice carrying a chimeric SRY/SOX construct (replacing the HMG domain of SRY with the HMG domain of SOX3 or SOX9) also exhibited sexual reversal [67], suggesting that SOX3 and SOX9 HMG domains can functionally replace SRY HMG domain. In addition, a two-exon SRY transcript was required for male testis determination, disproving the theory that SRY is a single-exon gene [68].
During gonadal development in mice, the SRY is initially expressed in Sertoli cell progenitors at E10.5, reaches its maximum expression at E11.5, and then disappears at E12.5. In contrast, SRY expression in humans begins around 41 days of pregnancy and peaks at 44 days [69]. In contrast to mice, human SRY expression gradually decreases to a base level around 60 days of pregnancy and is maintained until adulthood [70]. In addition, SRY expression in the post-testis determination stage has been detected in goats, sheep, pigs, rabbits, and cattle [71], prompting additional research into the mechanism of the male-promoting regulatory network in mice following sex determination. SRY expression is non-synchronous in the gonad; the wave of SRY expression moves from the center to the poles between E10.5 and E12.5, limiting transcriptomics research on sex determination to some extent [72,73].
The SOX9 expression reaches a plateau around 48 days of pregnancy in human testis [70] due to the synergistic action of SRY and NR5A1 [74] (E11.5–12.5 in mice [75]), thereby activating the male-promoting regulatory network and testis determination [76,77], followed by suppressing SRY expression and binding to NR5A1 to maintain expression [78]. In addition, FGF9 and PGD2 signaling pathways are activated after activation of testis-related genes, repression of anti-testis genes, inducing Supporting-to-Sertoli cell differentiation [79]. The SOX9 was sufficient to induce testis differentiation in the absence of SRY in transgenic XX gonads [80,81], which was later confirmed in a 46,XX mosaic male patient [82]. Therefore, SRY may only regulate SOX9 expression during testis development, which requires further investigation.

3.2. WNT Family Member 4 (WNT4)

The WNT4 is a member of the WNT family, essential for early embryonic development, the transition between naive and primed embryonic stem cells (ESC), and tissue homeostasis in adults [83]. The WNT4 is initially expressed in the undifferentiated early gonad at E11.25 [84], and WNT4 knockout translated into a significant increase in steroidogenic cells in both sexes [85]. In addition, the proliferation of coelomic epithelial cells was reported to be inhibited in the early gonads of WNT4-RSPO1-double-knockout mice [84], indicating that WNT4 may play the same role in both sexes during the early stages of gonadal development. Moreover, Müllerian duct formation failed in male and female WNT4−/− mice [86]. Similarly, WNT4 mutant 46,XX was found to have Müllerian duct abnormalities [87] and regression [88]. This suggests that WNT4 is necessary for regulating the histogeny of the Müllerian duct in both sexes.
The WNT4 is a component of the WNT/β-catenin signaling pathway and is essential in female sexual differentiation. The activation of the WNT/β-catenin signaling pathway is first detected in both sexes’ bipotential gonads at E11.5, acting as an anti-testicular agent by limiting the expression of SOX9 [84], but is downregulated by SRY in males [89]. An increase in WNT4 copies in humans was shown to result in a male-to-female sex reversal in 46,XY patients [90], while WNT4 inactivation or mutation resulted in sex reversion–kidneys, adrenal, and lung dysgenesis (SERKAL) syndrome [91] or virilization [88]. The WNT4 participates in accessory gland development by regulating hormone secretion, e.g., in WNT4−/− XX mice, genes involved in testosterone (a hormone involved in the formation of the epididymis, vas deferens, and seminal vesicles) synthesis were found elevated [92]. In contrast, steroidogenic enzymes 3β-hydroxysteroid dehydrogenase and 17α-hydroxylase, which are required for testosterone synthesis, were expressed in ovaries [86]. However, these transgenic models do not affect steroidogenic cell differentiation [93], revealing that WNT4 inhibits testosterone secretion by antagonizing steroidogenic cell migration rather than steroidogenic cell differentiation [94]. In humans, WNT4 mutant 46,XX patients were reported to suffer hyperandrogenism [87].
The WNT4 is required after sexual differentiation to prevent the formation of testis-specific vasculature, one of the earliest morphological changes during testicular differentiation [93,95]. In addition, it plays an essential role in the survival of oocytes and the maintenance of ovarian function [93,96]. Moreover, it is required for secreting steroid hormones in granulosa cells, which regulate normal ovarian follicle development and female fertility [97].

3.3. R-Spondin 1 (RSPO1)

The RSPO1 was discovered in the dorsal neural tube of mice in 2004 [98]. Since then, the RSPO1 family has been extensively studied, and the other three family members (RSPO2, RSPO3, and RSPO4) were discovered later. In mammal, these four RSPO1 family members have similar domain organization and are essential for embryogenesis, development, and tumorigenesis [98,99,100].
The RSPO1 is mainly expressed in mice XX gonadal somatic cells during ovary determination and suppressed in mice XY gonad, with only interstitial cells having low expression [101,102]. Loss-of-function experiment showed that RSPO1 knockout led to sex reversal and formation of ovotestis in XX mice [103]. In humans, the RSPO1 mutation caused hermaphroditism [104], palmoplantar hyperkeratosis, and squamous cell carcinoma [101]. Furthermore, RSPO1 functions conservatively in various vertebrates during ovarian development. In a recent study, goat BAC clones containing the RSPO1 gene (gRSPO1) were injected into mouse oocytes, which resulted in the restoration of sex-reversal in RSPO1 knockout XX mice [105]. Although the function of RSPO1 in inhibiting testicular differentiation still requires further research, the result described above provides essential insights into DSD treatment.
With the deepening of research, RSPO1 protein has been identified as an agonist of the WNT/β-catenin signaling pathway [106]; RSPO1−/− mice demonstrated an absence of activation of WNT4 [102]. In addition, the ovarian phenotype of RSPO1 knockout mice recapitulated with those of WNT4 knockout female mice [107]. The RSPO1 mutation 46,XX ovotestis, reduced expression of β-catenin protein and WNT4 mRNA, restricted ovarian differentiation. Transfection of RSPO1 resulted in activation of the β-catenin responsive TOPFLASH reporter (1.8-fold maximum), whereas RSPO1 and CTNNB1 (encoding β-catenin) synergy resulted in a 10-folds increased activation [108]. Above all, RSPO1 functions as an enhancer of β-catenin signaling during early ovary development. Interestingly, a novel role of RSPO1 in steroid hormone secretion independent of WNT/β-catenin signaling was discovered. After luminal cells-specific RSPO1 knockout, ESR1 (estrogen receptor alpha) expression was decreased, and mammary side branches were reduced. However, ESR1 expression was increased after luminal cell-specific knockout of WNT4, both with the attenuation of WNT/β-catenin signaling activities [109], revealing RSPO1 may involve in other signaling pathways that regulate female sexual differentiation.
The RSPO1 is also reported to participate in oocyte differentiation and meiosis after sex determination, as germ cell proliferation, STRA8 (early meiotic marker) expression, and the number of germ cells entering meiosis were all reported impaired in the RSPO1−/− fetal ovary [110]. However, in human disease, RSPO1 was found to promote progression in ovarian cancer by increasing the proliferation and migration of ovarian cancer cells and reducing ovarian cancer cells’ apoptosis [111].

3.4. Forkhead Box L2 (FOXL2)

The FOXL2 is one of the earliest markers of ovary differentiation in a mammal, which is sexual-specific and expressed in female gonads after E12.5 [112]. In FOXL2-knockout XX mice, granulosa cells and steroidogenic theca cells were reprogrammed into Sertoli-like cells and Leydig-like cells under the repression of SOX9 [113]. In the in vitro system, up-expression of NR5A1 was antagonized by FOXL2, and a 2-fold increase in NR5A1 expression was detected in FOXL2−/− mice relative to wild-type mutant [114]. According to findings, FOXL2 may regulate early ovarian development by directly suppressing the expression of testis-specific genes. Although FOXL2 plays a vital role in ovarian development in goats, it is more involved in fetal development than postnatal maintenance when compared to mice [115]. In XY transgenic mice, over-expression of FOXL2 led to the impairment of testis tubule differentiation [116], and RSPO1-FOXL2-double-knockout mice showed a similar phenotype earlier stage of sex reversal than RSPO1 knockout mice, revealing a potential interaction between these two female sex determination genes [117].
The FOXL2 becomes involved in follicle development by inducing Follicle-stimulating hormone (FSH) synthesis following sex determination [118,119], whose expression is regulated by ovarian hormones [120,121]. Furthermore, FOXL2 plays a role in the development and maintenance of the ovary via interacting with STAR [122], ESR2 [123]), and p27 [124]. Moreover, FOXL2 is expressed in the other components of the female reproductive tract, including the uterus, cervix, and oviduct, and plays a crucial role in postnatal uterine maturation [125]. The FOXL2 mutations are linked to Blepharophimosis-Ptosis-Epicanthus Inversus syndrome (BPES) [126,127,128], adult ovarian granulosa-cell tumor [129], testicular adult-type granulosa cell tumors [130,131], ovarian Sertoli-Leydig cell tumors [132], incompletely differentiated sex cord-stromal tumors [131] and ovarian sex cord-stromal tumors [133] in human.
When these sexual differentiation-related genes are taken together, they regulate the testis- and ovarian-determination network (listed in Table 2 and Figure 1) during embryonic development and throughout adulthood. They are also associated with developing other organs, reproductive capacity, and health. As a result, research aimed at developing animal models and modifying offspring sex ratios using gene-editing technology has been hampered for a long time due to organ failure. Although some research has investigated the underlying mechanisms of these sexual differentiation-related genes and has produced sexual reversal offspring, the development of offspring reproductive organs was significantly retarded. The number of available knockout offspring is lacking, limiting the study on growth performance. In addition, recent research has emphasized the importance of epigenetics in regulating sexual differentiation [134], inspiring future research in exploring the function of DNA methylation, histone modifications, non-coding RNA, and RNA methylation during sex determination gonads.

4. Current Knowledge of PGCs

The PGCs are distinct stem cells that can give rise to other stem cell types and pass on their genome to the next generation. PGC research offers new hope for treating infertility patients by in vitro mediating PGC differentiation, even though germ cell yields remain low. As a result, research on the formation of PGCs will be an important future research direction envisaged to promote in vitro derivation of human PGCs. Furthermore, there may be gender differences during the migration and differentiation of PGCs, which could provide a theoretical foundation for manipulating offspring sex ratios in livestock production by changing the ratio of Y- and X-chromosome-bearing sperm through gene editing.

4.1. Formation of PGCs

PGCs originate from a subpopulation of cells in the proximal epiblast (PE) at around two weeks of pregnancy in humans (around E6.5 in mice) [157]. Subsequently, these cells cluster and are located in the base of incipient allantois [158]. Current research has identified that bone morphogenic proteins (BMPs) mainly induce PGCs specification signals secreted from surrounding extraembryonic ectoderm (BMP4, BMP8b) [159,160] and visceral endoderm (BMP2) [161]. However, BMPs signals alone could not determine PGCs fate because only a subset of PE cells can induce differentiation into germ cells. Therefore, several in vivo and in vitro studies identified positive and negative signals directing PGC fate (listed in Table 3 and Figure 1). Furthermore, significant differences in PGC formation regulatory actions have been observed between humans and mice, such as SOX2 is required for PGC development in mice, while SOX17 is required in humans [162,163,164]. Moreover, KLF4 is only involved in maintaining pluripotency in human PGCs [165]. Interestingly, LncPGCAT-1 was found to positively regulate the formation of PGCs by elevating the expression of Cvh and C-kit and repressing the NANOG in vitro and in vivo [166], providing a new direction for research into the underlying biology of PGCs formation. In addition, recent single-cell sequencing research showed that the germline development between bovines and humans [167] and between mice and humans [165] were similar, which may provide new model organisms for the research on the development of PGCs.

4.2. In Vitro Derivation of PGCs

To examine the mammalian germline’s developmental mechanism, mice PGCs were first isolated in 1982 [184]. Since then, PGCs of other species have been successfully isolated, including goats [185], rabbits [186], sheep [187], and humans [188]. However, due to the low PGCs generation rate, the current research was devoted to deriving PGCs from pluripotent cells. So far, primordial-germ-cell-like cells (PGCLCs) [189] and long-term expanded PGCLCs [190] have been developed to generate fertile mice oocytes [191] and produce offspring [192] in vitro. On the other hand, the same gene expression patterns were observed for human PGCLCs and PGCs [165], cementing the feasibility of researching PGCs formation in vitro. Several genes important in PGCs formation and maintenance, such as TFAP2C [193,194], SSEA1 [195], DND1 [196], and SOX15 [197], were identified by employing sequencing technology, cell biology techniques, and genome editing technology on PGCLCs. However, there were still many challenges ahead; for instance, human PGCLCs derived in vitro could not meiosis completely during the embryonic stage [198]. Therefore, recent research is dedicated to the optimization of PGCs derivation routes.
According to recent studies, niche environments are important for differentiating human PGCs from pluripotent cells. Franklin D. West et al. discovered that co-culturing with mouse embryonic fibroblasts increased the expression of germ-cell-specific genes [199]. One year later, human fetal gonadal stromal cells were used for co-culturing with human embryonic stem cells (ESCs), significantly improving PGCs generation efficiency [200].
On the other hand, research on optimizing cell culture medium was carried out since Niels Geijsen et al. derived PGCs from ESCs by culturing with leukemia inhibitory factor in 2004 [201]. Until now, there have been several biochemical agents used in inducing PGCLCs differentiation in vitro, such as retinoic acid promoting the differentiation of PGCLCs from skin-derived stem cells [202]; retinoic acid combined with CHIR99021 promoting the differentiation of PGCLCs from human ESCs [203]; luteinizing hormone regulating the proliferation of porcine PGCLCs through ceRNA network [204]. Furthermore, recent research has identified the role of epigenetic modification in the differentiation of PGCLCs in vitro. MIR-10B has been discovered to play a role in differentiating PGCLCs from human mesenchymal stem cells [205]. In addition, α-ketoglutarate can promote PGCLCs specialization by regulating epigenetic reprogramming [206]. Similarly, the cell adhesion microenvironment was found to contribute to the differentiation of ESCs, which provide new ideas for PGCs derivation in vitro, where mesh substrates were found to induce self-organize and differentiation of ESCs, transiting to a PGCs-like state without the addition of biochemical inducers [207]. Interestingly, sex differences were found in the associations between Bisphenol A and PGCLC proliferation, with downregulated X-linked gene expression and PGCLC proliferation inhibited in XX cells but not in XY cells [208], providing a theoretical basis for intervening in the fate of different gender PGCs.
The differentiation of PGCs to embryonic germ cells (EGCs) has a lot of promise in studying the mechanisms of PGC survival, proliferation, and regulation. During the conversion process from PGCs to EGCs, the whole-transcriptome analysis revealed that BLIMP1 and Akt were involved in the specification and reprogramming of PGCs, respectively [209]. Further research showed that Akt activation promoted G1-S transition and enhanced PGCs reprogramming by downregulating H3K27me3 [210]. In addition, methylation changes at imprinting control centers (ICCs) during this conversion process were also discovered, stating that methylated ICCs are critical for PGCs derivation from ESCs [211]. Moreover, many new cell models have been developed to study factors regulating PGCs biologies, such as induced pluripotent stem cells [212] and PGCs derivation from nuclear transfer ESCs [213].

4.3. Migration of PGCs

Successful migration of PGCs to gonads is essential for gametogenesis in mammal, while anomalous migration of PGCs is required for the origin of endometriosis [214]. Although, with molecular biology development, the stages in PGC migration, with the underlying transcriptional regulatory network and signal pathways, have gradually been discovered and reviewed before [10,215,216,217], how PGCs migrates remains an important question.
Following PGC specification, PGCs first move from the primitive streak to allantois, where members of the interferon-inducing transmembrane protein (IFITM) family play a role in PGC incorporation into the hindgut [218]. Subsequently, the hindgut elongated, and PGCs moved into the dorsal mesentery through a fragmented basement membrane and finally colonized the GRs. Jingjing Sun et al. [219] found that, in the absence of MSX1 and MSX2, PGCs migration defected. The number of PGCs was reduced due to the reduction in the expression level WNT5A, which promoted directional migration of PGCs [220]. With improved molecular technology, several other regulatory RNAs have been discovered in the recent years such as NUP50 [221], SMAD4 [222], XVLG1 [223], HSP70 [224], PRDM1 [225], Ptch2/Gas1 and Ptch1/Boc [226].
During migration, the epigenome of PGCs undergoes comprehensive remodeling, including global DNA-demethylation, erasure of genomic imprints, and removal of H3K9me2; however, how they occur in PGCs is yet unknown. Anna Mallol et al. identified that PRDM14 was involved in global and X-chromosomal reprogramming, which upregulated the repressive H3K9me2 dose dependently and removed H3K27me3 from the inactive X-chromosome [227]. In addition, the DNA methylome between human PGCs and mice PGCs was found to be roughly comparable before PGCs differentiation [165], providing a basis for the future establishment of animal models in epigenetic research. However, a recent study indicated that PGCs migration mechanisms vary among mammals. PE Høyer et al. found an association between human PGCs and autonomic nerve fibers, which suggested that PGCs might be guided by nerve fibers [228], which was confirmed by Mollgard K et al. [229]. However, in mice and a non-human primate (marmoset monkey), most PGCs maintained a minimum distance of 50 µm from the closest neuron during different stages of embryonic development. More importantly, PGCs were discovered to reach the gonads before the emergence of neurons around the gonads [230]. Above all, whether PGCs migration mechanisms in different species are diversified remains controversial.
Another factor that affects PGCs migration is DNA damage response (DDR) which is present at all embryonic development stages and results in apoptosis or delayed proliferation of PGCs. However, the underlying mechanisms remain partially known. Recent genetic studies showed that FANCM or MCM9 deficiency reduced the number of PGCs before and after arriving in gonads. Interestingly, FANCM- MCM9-double-knockout mice showed an additive reduction of PGCs number [231], indicating that different DDR pathways can cause impaired PGCs migration. In another recent study, conditional knockout of PRMT5 activated DDR inducing sterility through PIWI-interacting RNA (piRNA) pathway indicated that PRMT5 was an important DNA protector [232]. The DDR was further studied with Ionizing radiation (IR), where, following germ cell differentiation and uncoupling of meiotic initiation in IR-treated female PGCs, gender differences were observed. In contrast, piRNA metabolism repression and transposon de-repression were detected in IR-treated male PGCs [233]. Importantly, this work provided new ideas for the research on sex manipulation by identifying genes that fit the established XX or XY germline.

4.4. Proliferation of PGCs and Gametogenesis

The PGCs begin to increase during migration and continue until a global change in gene expression occurs; PGCs are ready for gametogenesis. However, the mechanisms regulating the balance between proliferation and differentiation of PGCs remained unclear. Andrea V Cantú et al. discovered that WNT5A involves the proliferation of PGCs in different niches by repressing β-catenin-dependent and ROR2-mediated pathways [234], revealing that the tissue microenvironment regulated PGCs proliferation during migration rather than embryonic age. Another research using conditional knockout models showed that MASTL is vital for anaphase entry in female PGCs. Simultaneous deletion of PPP2R1A in MASTL-knockout PGCs can rescue the failure of PGCs to proceed beyond the metaphase-like stage, demonstrating that MASTL with PPP2A was essential for establishing female germline by regulating PGCs proliferation through phosphatase activity [235].
Furthermore, proteomic techniques were used to investigate PGC proliferation mechanisms, and it was reported that fatty acid degradation might play an important role in PGC proliferation. Furthermore, in vitro experiments demonstrated that when fatty acid degradation was suppressed, the number of PGCs decreased. Moreover, the expression levels of AMPK (p53 activator to induce cell cycle arrest), phosphorylated AMPK, phosphorylated p53, and cyclin-dependent kinase inhibitor 1 were increased, indicating that fatty acid degradation is involved in the proliferation of female PGCs via the p53 pathway [236]. Interestingly, some genes functioned at both proliferation and differentiation. For instance, ERK1-2 was expressed in PGCs at E8.5–E10.5 and gradually increased from E12.5–E14.5. After culturing PGCs with U0126 (MEK-specific inhibitor), ERK-12 expression was repressed, reducing PGCs at E8.5. Moreover, there were sex differences in controlling meiosis that only progression through meiotic prophase I of female PGCs treated with U0126 were slowed down [237]. In addition to participating in sexual differentiation of bipotential gonad, FGF9 was dose-dependent in regulating mice XY PGCs fate. Low doses of FGF9 (0.2 ng/mL) increased male-specific genes expression (DNMT3L and NANOS2) in XY PGCs, while a high dose of FGF9 (25 ng/mL) repressed the expression of male-specific genes and stimulated XY PGCs proliferation, revealing that FGF9 regulates the balance between proliferation and differentiation of XY PGCs in a dose-dependent manner [238]. These could be used as a selective mechanism to favor male or female migrators by repressing the proliferation or differentiation of one through conditional knockouts or conditional overexpression. Interestingly, EMX2 regulated the FGF9 pathway in somatic cells [239], which was important for GR formation, demonstrating that sex determination occurs throughout mammals’ lives.
Before gametogenesis, PGCs required permission to start meiosis and sexual differentiation; however, it remained unknown whether this permission was cell-autonomous or gonad-independent. Yueh-Chiang Hu et al. built a GATA4 (gene only expressed in somatic cells) conditional knockout model. They found that PGCs in GATA4-knockout embryos can migrate to the genital ridge but fail to start meiotic [240], indicating that gonad signaling is essential for gametogenesis. To fully explore the function of gonads, single-cell transcriptomics analysis was used in human fetal gonads. Four major signaling pathways (WNT, NOTCH, TGFβ/BMP, and receptor tyrosine kinases) were found to be involved in ligand-receptor interactions between PGCs and gonadal somatic cells using the CellPhoneDB algorithm [241]. WNT signaling pathway has been studied in depth because it is believed to be involved in sex determination throughout the life cycle. Anne-Amandine Chassot et al. found that spermatogonial proliferation was repressed and spermatocyte apoptosis increased following activation of the WNT/β-catenin pathway [242], which is consistent with the theory mentioned above that the WNT signaling pathway inhibits male-related biological processes. Another study identified WNT signaling as a “central gatekeeper” in female gametogenesis. PGCs maintained pluripotency or entered prematurely in the β-catenin gain- and loss-of-function models.
Additionally, by interacting with POU5F1, β-catenin was involved in pluripotency maintenance, and germ cell differentiation occurred when the WNT/β-catenin pathway was repressed after ZNRF3 upregulation [243]. The FGF signaling has been shown to regulate PGCs differentiation in two ways, i.e., by repressing female-related gene expression and activating downstream nodal/activin signaling to promote male gamete differentiation through degrading retinoic acid [244] and by activating the expression of NANOS2 (male germ cell marker) [245], which can prevent XX PGCs meiosis and induce male-like differentiation [246]. Interestingly, Quan Wu et al. found that SMAD2, a putative gene downstream from nodal/activin signaling, promoted male differentiation through a retinoic acid-independent routine because retinoic acid signaling suppression did not rescue male-specific gene expression in SMAD2 conditional knockout testes [247]. In mice PGCs proteomic research, there was no close correlation between proteomic data with published transcriptomic data using comparative analysis [248], revealing that the molecular mechanisms of gametogenesis may extend beyond the scope of the transcriptome, providing us essential inspiration for human gametogenesis research.
PGCs undergo a wide range of epigenetic reprogramming before sex-specific differentiation. DNA methylation has been extensively studied in PGCs, associated with chromatin reorganization, genomic imprinting erasure, and X-chromosome reactivation [165,249,250]. This process is mainly achieved by repressing DNA methylation-related genes (such as DNMT3A/B/L) and activating TET proteins, though there are still many unknowns in this field. Several upstream regulatory genes have been identified using gene-editing technology such as PRDM14 [251]. Moreover, Peter W S Hill et al. found that TET1 was involved in the maintenance of DNA demethylation rather than activation, providing a complete understanding of the TET family [252]. The SMARCB1 was discovered to have gender differences in regulating PGC epigenetic reprogramming. In SMARCB1-null female mice, meiosis-related genes were repressed, resulting in defects in synapse formation and DNA double-strand break repair. In contrast, in mutant male mice, the expression of genes related to growth and de novo DNA methylation was abnormal, resulting in mitotic arrest delay and hypomethylation of retrotransposons and imprinted genes [253].
Furthermore, DND1 was identified as a negative regulator of pluripotency and a positive regulator of epigenetic modifiers in male germ cell differentiation. In DND1Ter/Ter mutant mice, genes associated with pluripotency, cell cycle, male differentiation, and chromatin regulation were repressed, translating into entering G1/G0 impairment and teratomas formation [254]. These findings supported manipulating sex-dependent differentiation of PGCs; however, the function of these genes in humans remains unknown. As a result, recent research has examined the transcriptome and DNA methylome landscapes of human PGCs, laying the groundwork for understanding the complex relationship between gene regulatory networks and DNA methylation during the global epigenetic reprogramming process of human PGCs [165]. In addition, DNA methylomes of human PGCs during epigenetic reprogramming were roughly similar to mice [165,249]. However, human PGCs also show a unique gene regulatory network in epigenetic modification different from mice PGCs [249,250]. On the other hand, recent studies have identified additional epigenetic reprogramming of PGCs before sexual differentiation, such as histone acetylation [255] and noncoding RNAs [256].
Under the influence of a male or female regulatory network, PGCs give rise to spermatogonial stem cells or oogonia. It is worth noting that the previously mentioned antagonistic network still determines the fate of these germ cells. WNT4, activated by CTNNB1 signaling, can suppress spermatogonial stem cell activity in Sertoli cells [257], while female germ cell survival in the ovary is maintained by the WNT4/β-catenin pathway [258].

5. Conclusions

DSD has been a problem for humans for many years, and identifying functional variants of sex-related genes in DSD patients remains challenging. Infertility is another significant medical issue for which no effective treatments exist. The mechanisms underlying reproductive organogenesis and gametogenesis remain unknown despite considerable progress in recent years. However, it is worth mentioning that we have identified the interaction between somatic and germ cells and that signaling from somatic cells was essential for the proliferation and differentiation of PGCs. In contrast, female germ cells contributed to ovary maintenance. In addition, it was found that genes from males- and female-promoting antagonistic network primarily regulated the mammalian sex determination, which begins during embryonic development and continues throughout the life cycle. Nevertheless, several genes in this antagonistic network are also involved in the biological processes of organ maintenance and development, limiting the application of transgenic technology. In addition, transgenic efficiency remains low, and the sex-reversal trait cannot be stably transmitted to the next generation.
Alternatively, epigenetic changes during reproductive organogenesis and gametogenesis may explain the inability to identify DSD through genetic diagnosis. In order to address the issues above, it is possible to divide further future research into three distinct areas: (i) Intercellular signaling mechanisms must be first investigated, (ii) the structure and regulatory regions of known sex-related genes, intergenic regulatory networks, and identification of novel sex-related genes should be focused on in the future, and (iii) using sequencing technology, changes in DNA methylation, histone modifications, non-coding RNA, and RNA methylation need to be identified during sex determination. With additional research, we will better understand the processes underlying the development of the gonad and germline in humans, mice, and other mammals, which will aid in diagnosing and treating DSD and human infertility. In addition, these studies can offer theoretical support for manipulating offspring sex ratios in livestock production.

Author Contributions

Conceptualization: Y.X. and L.H.; writing—original draft preparation: Y.X.; writing—review and editing: Y.X., C.W., Z.L., Z.W. and L.H.; funding acquisition: Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Department of Science and Technology of Guangdong Province (2021B0707010007), the Guangdong Provincial Promotion Project on Preservation and UtiIization of Local Breed of Livestock and Poultry (2018143).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Witchel, S.F. Disorders of sex development. Best Pract. Res. Clin. Obstet. Gynaecol. 2018, 48, 90–102. [Google Scholar] [CrossRef] [PubMed]
  2. Arboleda, V.A.; Sandberg, D.E.; Vilain, E. DSDs: Genetics, underlying pathologies and psychosexual differentiation. Nat. Rev. Endocrinol. 2014, 10, 603–615. [Google Scholar] [CrossRef] [PubMed]
  3. Vander, B.M.; Wyns, C. Fertility and infertility: Definition and epidemiology. Clin. Biochem. 2018, 62, 2–10. [Google Scholar] [CrossRef] [PubMed]
  4. Rice, W.R. Sex chromosomes and the evolution of sexual dimorphism. Evolution 1984, 38, 735–742. [Google Scholar] [CrossRef] [PubMed]
  5. Lucas-Herald, A.K.; Bashamboo, A. Gonadal development. Underst. Differ. Disord. Sex Dev. 2014, 27, 1–16. [Google Scholar] [CrossRef]
  6. Xie, Y.; Xu, Z.; Wu, Z.; Hong, L. Sex Manipulation Technologies Progress in Livestock: A Review. Front. Vet. Sci. 2020, 7, 481. [Google Scholar] [CrossRef]
  7. She, Z.Y.; Yang, W.X. Sry and SoxE genes: How they participate in mammalian sex determination and gonadal development? Semin. Cell Dev. Biol. 2017, 63, 13–22. [Google Scholar] [CrossRef]
  8. Tanaka, S.S.; Nishinakamura, R. Regulation of male sex determination: Genital ridge formation and Sry activation in mice. Cell. Mol. Life Sci. 2014, 71, 4781–4802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Biason-Lauber, A. WNT4, RSPO1, and FOXL2 in sex development. Semin. Reprod. Med. 2012, 30, 387–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Nikolic, A.; Volarevic, V.; Armstrong, L.; Lako, M.; Stojkovic, M. Primordial Germ Cells: Current Knowledge and Perspectives. Stem Cells Int. 2016, 2016, 1741072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Marlow, F. Primordial Germ Cell Specification and Migration. F1000Research 2015, 4, Rev-1462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kumar, D.L.; DeFalco, T. Of Mice and Men: In Vivo and In Vitro Studies of Primordial Germ Cell Specification. Semin. Reprod. Med. 2017, 35, 139–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Jost, A. A new look at the mechanisms controlling sex differentiation in mammals. Johns Hopkins Med. J. 1972, 1, 38–53. [Google Scholar]
  14. Wartenberg, H.; Kinsky, I.; Viebahn, C.; Schmolke, C. Fine structural characteristics of testicular cord formation in the developing rabbit gonad. J. Electron Microsc. Tech. 1991, 19, 133–157. [Google Scholar] [CrossRef] [PubMed]
  15. Pelliniemi, L.J. Ultrastructure of gonadal ridge in male and female pig embryos. Anat. Embryol. 1975, 147, 20–34. [Google Scholar] [CrossRef]
  16. Gropp, A.; Ohno, S. The presence of a common embryonic blastema for ovarian and testicular parenchymal (follicular, interstitial and tubular) cells in cattle Bos taurus. Z. Zellforsch. Mikrosk. Anat. 1966, 74, 505–528. [Google Scholar] [CrossRef] [PubMed]
  17. Shaw, G.; Renfree, M.B. Wolffian duct development. Sex. Dev. 2014, 8, 273–280. [Google Scholar] [CrossRef] [PubMed]
  18. Hannema, S.E.; Hughes, I.A. Regulation of Wolffian duct development. Horm. Res. 2007, 67, 142–151. [Google Scholar] [CrossRef]
  19. Acien, P. Embryological observations on the female genital tract. Hum. Reprod. 1992, 7, 437–445. [Google Scholar] [CrossRef]
  20. Ross, A.J.; Capel, B. Signaling at the crossroads of gonad development. Trends Endocrinol. Metab. 2005, 16, 19–25. [Google Scholar] [CrossRef]
  21. Hirst, C.E.; Major, A.T.; Smith, C.A. Sex determination and gonadal sex differentiation in the chicken model. Int. J. Dev. Biol. 2018, 62, 153–166. [Google Scholar] [CrossRef] [PubMed]
  22. Stevant, I.; Kuhne, F.; Greenfield, A.; Chaboissier, M.C.; Dermitzakis, E.T.; Nef, S. Dissecting Cell Lineage Specification and Sex Fate Determination in Gonadal Somatic Cells Using Single-Cell Transcriptomics. Cell Rep. 2019, 26, 3272–3283. [Google Scholar] [CrossRef] [Green Version]
  23. Guo, J.; Sosa, E.; Chitiashvili, T.; Nie, X.; Rojas, E.J.; Oliver, E.; Plath, K.; Hotaling, J.M.; Stukenborg, J.B.; Clark, A.T.; et al. Single-cell analysis of the developing human testis reveals somatic niche cell specification and fetal germline stem cell establishment. Cell Stem Cell 2021, 28, 764–778. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, Y.; Workman, S.; Wilson, M. The molecular pathways underlying early gonadal development. J. Mol. Endocrinol. 2018, 62, R47–R64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Luo, X.; Ikeda, Y.; Parker, K.L. A cell-specific nuclear receptor is essential for adrenal and gonadal development and sexual differentiation. Cell 1994, 77, 481–490. [Google Scholar] [CrossRef]
  26. Sadovsky, Y.; Crawford, P.A.; Woodson, K.G.; Polish, J.A.; Clements, M.A.; Tourtellotte, L.M.; Simburger, K.; Milbrandt, J. Mice deficient in the orphan receptor steroidogenic factor 1 lack adrenal glands and gonads but express P450 side-chain-cleavage enzyme in the placenta and have normal embryonic serum levels of corticosteroids. Proc. Natl. Acad. Sci. USA 1995, 92, 10939–10943. [Google Scholar] [CrossRef] [Green Version]
  27. Shinoda, K.; Lei, H.; Yoshii, H.; Nomura, M.; Nagano, M.; Shiba, H.; Sasaki, H.; Osawa, Y.; Ninomiya, Y.; Niwa, O.; et al. Developmental defects of the ventromedial hypothalamic nucleus and pituitary gonadotroph in the Ftz-F1 disrupted mice. Dev. Dyn. 1995, 204, 22–29. [Google Scholar] [CrossRef]
  28. Bashamboo, A.; Ferraz-de-Souza, B.; Lourenco, D.; Lin, L.; Sebire, N.J.; Montjean, D.; Bignon-Topalovic, J.; Mandelbaum, J.; Siffroi, J.P.; Christin-Maitre, S.; et al. Human male infertility associated with mutations in NR5A1 encoding steroidogenic factor 1. Am. J. Hum. Genet. 2010, 87, 505–512. [Google Scholar] [CrossRef] [Green Version]
  29. Laan, M.; Kasak, L.; Timinskas, K.; Grigorova, M.; Venclovas, C.; Renaux, A.; Lenaerts, T.; Punab, M. NR5A1 c.991-1G > C splice-site variant causes familial 46,XY partial gonadal dysgenesis with incomplete penetrance. Clin. Endocrinol. 2021, 94, 656–666. [Google Scholar] [CrossRef] [PubMed]
  30. Na, X.; Mao, Y.; Tang, Y.; Jiang, W.; Yu, J.; Cao, L.; Yang, J. Identification and functional analysis of fourteen NR5A1 variants in patients with the 46 XY disorders of sex development. Gene 2020, 760, 145004. [Google Scholar] [CrossRef]
  31. Yu, B.; Liu, Z.; Gao, Y.; Mao, J.; Wang, X.; Hao, M.; Ma, W.; Huang, Q.; Zhang, R.; Nie, M.; et al. Novel NR5A1 mutations found in Chinese patients with 46, XY disorders of sex development. Clin. Endocrinol. 2018, 89, 613–620. [Google Scholar] [CrossRef] [PubMed]
  32. Achermann, J.C.; Ito, M.; Ito, M.; Hindmarsh, P.C.; Jameson, J.L. A mutation in the gene encoding steroidogenic factor-1 causes XY sex reversal and adrenal failure in humans. Nat. Genet. 1999, 22, 125–126. [Google Scholar] [CrossRef] [PubMed]
  33. Knarston, I.M.; Robevska, G.; van den Bergen, J.A.; Eggers, S.; Croft, B.; Yates, J.; Hersmus, R.; Looijenga, L.; Cameron, F.J.; Monhike, K.; et al. NR5A1 gene variants repress the ovarian-specific WNT signaling pathway in 46,XX disorders of sex development patients. Hum. Mutat. 2019, 40, 207–216. [Google Scholar] [CrossRef] [Green Version]
  34. Jaillard, S.; Sreenivasan, R.; Beaumont, M.; Robevska, G.; Dubourg, C.; Knarston, I.M.; Akloul, L.; van den Bergen, J.; Odent, S.; Croft, B.; et al. Analysis of NR5A1 in 142 patients with premature ovarian insufficiency, diminished ovarian reserve, or unexplained infertility. Maturitas 2020, 131, 78–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hu, Y.C.; Okumura, L.M.; Page, D.C. Gata4 is required for formation of the genital ridge in mice. PLoS Genet. 2013, 9, e1003629. [Google Scholar] [CrossRef] [Green Version]
  36. Tevosian, S.G.; Albrecht, K.H.; Crispino, J.D.; Fujiwara, Y.; Eicher, E.M.; Orkin, S.H. Gonadal differentiation, sex determination and normal Sry expression in mice require direct interaction between transcription partners GATA4 and FOG2. Development 2002, 129, 4627–4634. [Google Scholar] [CrossRef]
  37. Lourenco, D.; Brauner, R.; Rybczynska, M.; Nihoul-Fekete, C.; McElreavey, K.; Bashamboo, A. Loss-of-function mutation in GATA4 causes anomalies of human testicular development. Proc. Natl. Acad. Sci. USA 2011, 108, 1597–1602. [Google Scholar] [CrossRef] [Green Version]
  38. Hammes, A.; Guo, J.K.; Lutsch, G.; Leheste, J.R.; Landrock, D.; Ziegler, U.; Gubler, M.C.; Schedl, A. Two splice variants of the Wilms’ tumor 1 gene have distinct functions during sex determination and nephron formation. Cell 2001, 106, 319–329. [Google Scholar] [CrossRef] [Green Version]
  39. Barbaux, S.; Niaudet, P.; Gubler, M.C.; Grunfeld, J.P.; Jaubert, F.; Kuttenn, F.; Fekete, C.N.; Souleyreau-Therville, N.; Thibaud, E.; Fellous, M.; et al. Donor splice-site mutations in WT1 are responsible for Frasier syndrome. Nat. Genet. 1997, 17, 467–470. [Google Scholar] [CrossRef]
  40. Klamt, B.; Koziell, A.; Poulat, F.; Wieacker, P.; Scambler, P.; Berta, P.; Gessler, M. Frasier syndrome is caused by defective alternative splicing of WT1 leading to an altered ratio of WT1 +/-KTS splice isoforms. Hum. Mol. Genet. 1998, 7, 709–714. [Google Scholar] [CrossRef]
  41. Birk, O.S.; Casiano, D.E.; Wassif, C.A.; Cogliati, T.; Zhao, L.; Zhao, Y.; Grinberg, A.; Huang, S.; Kreidberg, J.A.; Parker, K.L.; et al. The LIM homeobox gene Lhx9 is essential for mouse gonad formation. Nature 2000, 403, 909–913. [Google Scholar] [CrossRef] [PubMed]
  42. Kusaka, M.; Katoh-Fukui, Y.; Ogawa, H.; Miyabayashi, K.; Baba, T.; Shima, Y.; Sugiyama, N.; Sugimoto, Y.; Okuno, Y.; Kodama, R.; et al. Abnormal epithelial cell polarity and ectopic epidermal growth factor receptor (EGFR) expression induced in Emx2 KO embryonic gonads. Endocrinology 2010, 151, 5893–5904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Fujimoto, Y.; Tanaka, S.S.; Yamaguchi, Y.L.; Kobayashi, H.; Kuroki, S.; Tachibana, M.; Shinomura, M.; Kanai, Y.; Morohashi, K.; Kawakami, K.; et al. Homeoproteins Six1 and Six4 regulate male sex determination and mouse gonadal development. Dev. Cell 2013, 26, 416–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Cui, S.; Ross, A.; Stallings, N.; Parker, K.L.; Capel, B.; Quaggin, S.E. Disrupted gonadogenesis and male-to-female sex reversal in Pod1 knockout mice. Development 2004, 131, 4095–4105. [Google Scholar] [CrossRef] [Green Version]
  45. Katoh-Fukui, Y.; Owaki, A.; Toyama, Y.; Kusaka, M.; Shinohara, Y.; Maekawa, M.; Toshimori, K.; Morohashi, K. Mouse Polycomb M33 is required for splenic vascular and adrenal gland formation through regulating Ad4BP/SF1 expression. Blood 2005, 106, 1612–1620. [Google Scholar] [CrossRef]
  46. Katoh-Fukui, Y.; Tsuchiya, R.; Shiroishi, T.; Nakahara, Y.; Hashimoto, N.; Noguchi, K.; Higashinakagawa, T. Male-to-female sex reversal in M33 mutant mice. Nature 1998, 393, 688–692. [Google Scholar] [CrossRef]
  47. Efstratiadis, A. Genetics of mouse growth. Int. J. Dev. Biol. 1998, 42, 955–976. [Google Scholar]
  48. Schnabel, C.A.; Selleri, L.; Cleary, M.L. Pbx1 is essential for adrenal development and urogenital differentiation. Genesis 2003, 37, 123–130. [Google Scholar] [CrossRef]
  49. Eozenou, C.; Bashamboo, A.; Bignon-Topalovic, J.; Merel, T.; Zwermann, O.; Lourenco, D.; Lottmann, H.; Lichtenauer, U.; Rojo, S.; Beuschlein, F.; et al. The TALE homeodomain of PBX1 is involved in human primary testis-determination. Hum. Mutat. 2019, 40, 1071–1076. [Google Scholar] [CrossRef]
  50. Wang, Q.; Lan, Y.; Cho, E.S.; Maltby, K.M.; Jiang, R. Odd-skipped related 1 (Odd 1) is an essential regulator of heart and urogenital development. Dev. Biol. 2005, 288, 582–594. [Google Scholar] [CrossRef] [Green Version]
  51. McLaren, A. Somatic and germ-cell sex in mammals. Philos. Trans. R. Soc. London B Biol. Sci. 1988, 322, 3–9. [Google Scholar] [CrossRef] [PubMed]
  52. Wyndham, N.R. A morphological study of testicular descent. J. Anat. 1943, 77, 179–188. [Google Scholar] [PubMed]
  53. Hacker, A.; Capel, B.; Goodfellow, P.; Lovell-Badge, R. Expression of Sry, the mouse sex determining gene. Development 1995, 121, 1603–1614. [Google Scholar] [CrossRef] [PubMed]
  54. Schmahl, J.; Eicher, E.M.; Washburn, L.L.; Capel, B. Sry induces cell proliferation in the mouse gonad. Development 2000, 127, 65–73. [Google Scholar] [CrossRef] [PubMed]
  55. Veitia, R.A. FOXL2 versus SOX9: A lifelong “battle of the sexes”. Bioessays 2010, 32, 375–380. [Google Scholar] [CrossRef]
  56. Minkina, A.; Matson, C.K.; Lindeman, R.E.; Ghyselinck, N.B.; Bardwell, V.J.; Zarkower, D. DMRT1 protects male gonadal cells from retinoid-dependent sexual transdifferentiation. Dev. Cell 2014, 29, 511–520. [Google Scholar] [CrossRef] [Green Version]
  57. Palmer, M.S.; Sinclair, A.H.; Berta, P.; Ellis, N.A.; Goodfellow, P.N.; Abbas, N.E.; Fellous, M. Genetic evidence that ZFY is not the testis-determining factor. Nature 1989, 342, 937–939. [Google Scholar] [CrossRef] [PubMed]
  58. Sinclair, A.H.; Berta, P.; Palmer, M.S.; Hawkins, J.R.; Griffiths, B.L.; Smith, M.J.; Foster, J.W.; Frischauf, A.M.; Lovell-Badge, R.; Goodfellow, P.N. A gene from the human sex-determining region encodes a protein with homology to a conserved DNA-binding motif. Nature 1990, 346, 240–244. [Google Scholar] [CrossRef] [Green Version]
  59. Bowles, J.; Schepers, G.; Koopman, P. Phylogeny of the SOX family of developmental transcription factors based on sequence and structural indicators. Dev. Biol. 2000, 227, 239–255. [Google Scholar] [CrossRef] [Green Version]
  60. Zhao, L.; Ng, E.T.; Davidson, T.L.; Longmuss, E.; Urschitz, J.; Elston, M.; Moisyadi, S.; Bowles, J.; Koopman, P. Structure-function analysis of mouse Sry reveals dual essential roles of the C-terminal polyglutamine tract in sex determination. Proc. Natl. Acad. Sci. USA 2014, 111, 11768–11773. [Google Scholar] [CrossRef] [Green Version]
  61. Harley, V.R.; Layfield, S.; Mitchell, C.L.; Forwood, J.K.; John, A.P.; Briggs, L.J.; McDowall, S.G.; Jans, D.A. Defective importin beta recognition and nuclear import of the sex-determining factor SRY are associated with XY sex-reversing mutations. Proc. Natl. Acad. Sci. USA 2003, 100, 7045–7050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Li, B.; Zhang, W.; Chan, G.; Jancso-Radek, A.; Liu, S.; Weiss, M.A. Human sex reversal due to impaired nuclear localization of SRY. A clinical correlation. J. Biol. Chem. 2001, 276, 46480–46484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Li, B.; Phillips, N.B.; Jancso-Radek, A.; Ittah, V.; Singh, R.; Jones, D.N.; Haas, E.; Weiss, M.A. SRY-directed DNA bending and human sex reversal: Reassessment of a clinical mutation uncovers a global coupling between the HMG box and its tail. J. Mol. Biol. 2006, 360, 310–328. [Google Scholar] [CrossRef] [PubMed]
  64. Battiloro, E.; Angeletti, B.; Tozzi, M.C.; Bruni, L.; Tondini, S.; Vignetti, P.; Verna, R.; D’Ambrosio, E. A novel double nucleotide substitution in the HMG box of the SRY gene associated with Swyer syndrome. Hum. Genet. 1997, 100, 585–587. [Google Scholar] [CrossRef] [PubMed]
  65. Hawkins, J.R. Mutational analysis of SRY in XY females. Hum. Mutat. 1993, 2, 347–350. [Google Scholar] [CrossRef]
  66. Kurtz, S.; Lucas-Hahn, A.; Schlegelberger, B.; Gohring, G.; Niemann, H.; Mettenleiter, T.C.; Petersen, B. Knockout of the HMG domain of the porcine SRY gene causes sex reversal in gene-edited pigs. Proc. Natl. Acad. Sci. USA 2021, 118, e2008743118. [Google Scholar] [CrossRef]
  67. Bergstrom, D.E.; Young, M.; Albrecht, K.H.; Eicher, E.M. Related function of mouse SOX3, SOX9, and SRY HMG domains assayed by male sex determination. Genesis 2000, 28, 111–124. [Google Scholar] [CrossRef]
  68. Miyawaki, S.; Kuroki, S.; Maeda, R.; Okashita, N.; Koopman, P.; Tachibana, M. The mouse Sry locus harbors a cryptic exon that is essential for male sex determination. Science 2020, 370, 121–124. [Google Scholar] [CrossRef]
  69. Hanley, N.A.; Hagan, D.M.; Clement-Jones, M.; Ball, S.G.; Strachan, T.; Salas-Cortes, L.; McElreavey, K.; Lindsay, S.; Robson, S.; Bullen, P.; et al. SRY, SOX9, and DAX1 expression patterns during human sex determination and gonadal development. Mech. Dev. 2000, 91, 403–407. [Google Scholar] [CrossRef]
  70. Mamsen, L.S.; Ernst, E.H.; Borup, R.; Larsen, A.; Olesen, R.H.; Ernst, E.; Anderson, R.A.; Kristensen, S.G.; Andersen, C.Y. Temporal expression pattern of genes during the period of sex differentiation in human embryonic gonads. Sci. Rep. 2017, 7, 15961. [Google Scholar] [CrossRef] [Green Version]
  71. Parma, P.; Radi, O. Molecular mechanisms of sexual development. Sex. Dev. 2012, 6, 7–17. [Google Scholar] [CrossRef] [PubMed]
  72. Warr, N.; Carre, G.A.; Siggers, P.; Faleato, J.V.; Brixey, R.; Pope, M.; Bogani, D.; Childers, M.; Wells, S.; Scudamore, C.L.; et al. Gadd45gamma and Map3k4 interactions regulate mouse testis determination via p38 MAPK-mediated control of Sry expression. Dev. Cell 2012, 23, 1020–1031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Bullejos, M.; Koopman, P. Spatially dynamic expression of Sry in mouse genital ridges. Dev. Dyn. 2001, 221, 201–205. [Google Scholar] [CrossRef] [PubMed]
  74. Sekido, R.; Lovell-Badge, R. Sex determination involves synergistic action of SRY and SF1 on a specific Sox9 enhancer. Nature 2008, 453, 930–934. [Google Scholar] [CrossRef] [PubMed]
  75. Sekido, R.; Bar, I.; Narvaez, V.; Penny, G.; Lovell-Badge, R. SOX9 is up-regulated by the transient expression of SRY specifically in Sertoli cell precursors. Dev. Biol. 2004, 274, 271–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Gubbay, J.; Collignon, J.; Koopman, P.; Capel, B.; Economou, A.; Munsterberg, A.; Vivian, N.; Goodfellow, P.; Lovell-Badge, R. A gene mapping to the sex-determining region of the mouse Y chromosome is a member of a novel family of embryonically expressed genes. Nature 1990, 346, 245–250. [Google Scholar] [CrossRef]
  77. Larney, C.; Bailey, T.L.; Koopman, P. Switching on sex: Transcriptional regulation of the testis-determining gene Sry. Development 2014, 141, 2195–2205. [Google Scholar] [CrossRef] [Green Version]
  78. Zangen, D.; Kaufman, Y.; Banne, E.; Weinberg-Shukron, A.; Abulibdeh, A.; Garfinkel, B.P.; Dweik, D.; Kanaan, M.; Camats, N.; Fluck, C.; et al. Testicular differentiation factor SF-1 is required for human spleen development. J. Clin. Investig. 2014, 124, 2071–2075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Chaboissier, M.C.; Kobayashi, A.; Vidal, V.I.; Lutzkendorf, S.; van de Kant, H.J.; Wegner, M.; de Rooij, D.G.; Behringer, R.R.; Schedl, A. Functional analysis of Sox8 and Sox9 during sex determination in the mouse. Development 2004, 131, 1891–1901. [Google Scholar] [CrossRef] [Green Version]
  80. Vidal, V.P.; Chaboissier, M.C.; de Rooij, D.G.; Schedl, A. Sox9 induces testis development in XX transgenic mice. Nat. Genet. 2001, 28, 216–217. [Google Scholar] [CrossRef]
  81. Qin, Y.; Bishop, C.E. Sox9 is sufficient for functional testis development producing fertile male mice in the absence of Sry. Hum. Mol. Genet. 2005, 14, 1221–1229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Huang, B.; Wang, S.; Ning, Y.; Lamb, A.N.; Bartley, J. Autosomal XX sex reversal caused by duplication of SOX9. Am. J. Med. Genet. 1999, 87, 349–353. [Google Scholar] [CrossRef]
  83. Steinhart, Z.; Angers, S. Wnt signaling in development and tissue homeostasis. Development 2018, 145, dev146589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Chassot, A.A.; Bradford, S.T.; Auguste, A.; Gregoire, E.P.; Pailhoux, E.; de Rooij, D.G.; Schedl, A.; Chaboissier, M.C. WNT4 and RSPO1 together are required for cell proliferation in the early mouse gonad. Development 2012, 139, 4461–4472. [Google Scholar] [CrossRef] [Green Version]
  85. Jeays-Ward, K.; Dandonneau, M.; Swain, A. Wnt4 is required for proper male as well as female sexual development. Dev. Biol. 2004, 276, 431–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Vainio, S.; Heikkila, M.; Kispert, A.; Chin, N.; McMahon, A.P. Female development in mammals is regulated by Wnt-4 signalling. Nature 1999, 397, 405–409. [Google Scholar] [CrossRef] [PubMed]
  87. Philibert, P.; Biason-Lauber, A.; Gueorguieva, I.; Stuckens, C.; Pienkowski, C.; Lebon-Labich, B.; Paris, F.; Sultan, C. Molecular analysis of WNT4 gene in four adolescent girls with mullerian duct abnormality and hyperandrogenism (atypical Mayer-Rokitansky-Kuster-Hauser syndrome). Fertil. Steril. 2011, 95, 2683–2686. [Google Scholar] [CrossRef] [PubMed]
  88. Biason-Lauber, A.; Konrad, D.; Navratil, F.; Schoenle, E.J. A WNT4 mutation associated with Mullerian-duct regression and virilization in a 46,XX woman. N. Engl. J. Med. 2004, 351, 792–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Bernard, P.; Sim, H.; Knower, K.; Vilain, E.; Harley, V. Human SRY inhibits beta-catenin-mediated transcription. Int. J. Biochem. Cell Biol. 2008, 40, 2889–2900. [Google Scholar] [CrossRef] [Green Version]
  90. Jaaskelainen, M.; Prunskaite-Hyyrylainen, R.; Naillat, F.; Parviainen, H.; Anttonen, M.; Heikinheimo, M.; Liakka, A.; Ola, R.; Vainio, S.; Vaskivuo, T.E.; et al. WNT4 is expressed in human fetal and adult ovaries and its signaling contributes to ovarian cell survival. Mol. Cell. Endocrinol. 2010, 317, 106–111. [Google Scholar] [CrossRef] [Green Version]
  91. Mandel, H.; Shemer, R.; Borochowitz, Z.U.; Okopnik, M.; Knopf, C.; Indelman, M.; Drugan, A.; Tiosano, D.; Gershoni-Baruch, R.; Choder, M.; et al. SERKAL syndrome: An autosomal-recessive disorder caused by a loss-of-function mutation in WNT4. Am. J. Hum. Genet. 2008, 82, 39–47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Heikkila, M.; Prunskaite, R.; Naillat, F.; Itaranta, P.; Vuoristo, J.; Leppaluoto, J.; Peltoketo, H.; Vainio, S. The partial female to male sex reversal in Wnt-4-deficient females involves induced expression of testosterone biosynthetic genes and testosterone production, and depends on androgen action. Endocrinology 2005, 146, 4016–4023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Jeays-Ward, K.; Hoyle, C.; Brennan, J.; Dandonneau, M.; Alldus, G.; Capel, B.; Swain, A. Endothelial and steroidogenic cell migration are regulated by WNT4 in the developing mammalian gonad. Development 2003, 130, 3663–3670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Heikkila, M.; Peltoketo, H.; Leppaluoto, J.; Ilves, M.; Vuolteenaho, O.; Vainio, S. Wnt-4 deficiency alters mouse adrenal cortex function, reducing aldosterone production. Endocrinology 2002, 143, 4358–4365. [Google Scholar] [CrossRef] [Green Version]
  95. Coveney, D.; Ross, A.J.; Slone, J.D.; Capel, B. A microarray analysis of the XX Wnt4 mutant gonad targeted at the identification of genes involved in testis vascular differentiation. Gene Expr. Patterns 2007, 7, 82–92. [Google Scholar] [CrossRef] [PubMed]
  96. Naillat, F.; Prunskaite-Hyyrylainen, R.; Pietila, I.; Sormunen, R.; Jokela, T.; Shan, J.; Vainio, S.J. Wnt4/5a signalling coordinates cell adhesion and entry into meiosis during presumptive ovarian follicle development. Hum. Mol. Genet. 2010, 19, 1539–1550. [Google Scholar] [CrossRef] [Green Version]
  97. Boyer, A.; Lapointe, E.; Zheng, X.; Cowan, R.G.; Li, H.; Quirk, S.M.; DeMayo, F.J.; Richards, J.S.; Boerboom, D. WNT4 is required for normal ovarian follicle development and female fertility. FASEB J. 2010, 24, 3010–3025. [Google Scholar] [CrossRef] [Green Version]
  98. Kamata, T.; Katsube, K.; Michikawa, M.; Yamada, M.; Takada, S.; Mizusawa, H. R-spondin, a novel gene with thrombospondin type 1 domain, was expressed in the dorsal neural tube and affected in Wnts mutants. Biochim. Biophys. Acta 2004, 1676, 51–62. [Google Scholar] [CrossRef]
  99. Kazanskaya, O.; Glinka, A.; Del, B.B.I.; Stannek, P.; Niehrs, C.; Wu, W. R-Spondin2 is a secreted activator of Wnt/beta-catenin signaling and is required for Xenopus myogenesis. Dev. Cell 2004, 7, 525–534. [Google Scholar] [CrossRef] [Green Version]
  100. Yoon, J.K.; Lee, J.S. Cellular signaling and biological functions of R-spondins. Cell. Signal. 2012, 24, 369–377. [Google Scholar] [CrossRef] [Green Version]
  101. Parma, P.; Radi, O.; Vidal, V.; Chaboissier, M.C.; Dellambra, E.; Valentini, S.; Guerra, L.; Schedl, A.; Camerino, G. R-spondin1 is essential in sex determination, skin differentiation and malignancy. Nat. Genet. 2006, 38, 1304–1309. [Google Scholar] [CrossRef] [PubMed]
  102. Chassot, A.A.; Ranc, F.; Gregoire, E.P.; Roepers-Gajadien, H.L.; Taketo, M.M.; Camerino, G.; de Rooij, D.G.; Schedl, A.; Chaboissier, M.C. Activation of beta-catenin signaling by Rspo1 controls differentiation of the mammalian ovary. Hum. Mol. Genet. 2008, 17, 1264–1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Chassot, A.A.; Gregoire, E.P.; Magliano, M.; Lavery, R.; Chaboissier, M.C. Genetics of ovarian differentiation: Rspo1, a major player. Sex. Dev. 2008, 2, 219–227. [Google Scholar] [CrossRef] [PubMed]
  104. Tomaselli, S.; Megiorni, F.; De Bernardo, C.; Felici, A.; Marrocco, G.; Maggiulli, G.; Grammatico, B.; Remotti, D.; Saccucci, P.; Valentini, F.; et al. Syndromic true hermaphroditism due to an R-spondin1 (RSPO1) homozygous mutation. Hum. Mutat. 2008, 29, 220–226. [Google Scholar] [CrossRef] [PubMed]
  105. Buscara, L.; Montazer-Torbati, F.; Chadi, S.; Auguste, A.; Laubier, J.; Chassot, A.A.; Renault, L.; Passet, B.; Costa, J.; Pannetier, M.; et al. Goat RSPO1 over-expression rescues sex-reversal in Rspo1-knockout XX mice but does not perturb testis differentiation in XY or sex-reversed XX mice. Transgenic Res. 2009, 18, 649–654. [Google Scholar] [CrossRef]
  106. Naasse, Y.; Bakhchane, A.; Charoute, H.; Jennane, F.; Bignon-Topalovic, J.; Malki, A.; Bashamboo, A.; Barakat, A.; Rouba, H.; McElreavey, K. A Novel Homozygous Missense Mutation in the FU-CRD2 Domain of the R-spondin1 Gene Associated with Familial 46,XX DSD. Sex. Dev. 2017, 11, 269–274. [Google Scholar] [CrossRef]
  107. Tomizuka, K.; Horikoshi, K.; Kitada, R.; Sugawara, Y.; Iba, Y.; Kojima, A.; Yoshitome, A.; Yamawaki, K.; Amagai, M.; Inoue, A.; et al. R-spondin1 plays an essential role in ovarian development through positively regulating Wnt-4 signaling. Hum. Mol. Genet. 2008, 17, 1278–1291. [Google Scholar] [CrossRef] [Green Version]
  108. Tomaselli, S.; Megiorni, F.; Lin, L.; Mazzilli, M.C.; Gerrelli, D.; Majore, S.; Grammatico, P.; Achermann, J.C. Human RSPO1/R-spondin1 is expressed during early ovary development and augments beta-catenin signaling. PLoS ONE 2011, 6, e16366. [Google Scholar] [CrossRef] [Green Version]
  109. Geng, A.; Wu, T.; Cai, C.; Song, W.; Wang, J.; Yu, Q.C.; Zeng, Y.A. A novel function of R-spondin1 in regulating estrogen receptor expression independent of Wnt/beta-catenin signaling. eLife 2020, 9, e56434. [Google Scholar] [CrossRef]
  110. Chassot, A.A.; Gregoire, E.P.; Lavery, R.; Taketo, M.M.; de Rooij, D.G.; Adams, I.R.; Chaboissier, M.C. RSPO1/beta-catenin signaling pathway regulates oogonia differentiation and entry into meiosis in the mouse fetal ovary. PLoS ONE 2011, 6, e25641. [Google Scholar] [CrossRef] [Green Version]
  111. Liu, Q.; Zhao, Y.; Xing, H.; Li, L.; Li, R.; Dai, J.; Li, Q.; Fang, S. The role of R-spondin 1 through activating Wnt/beta-catenin in the growth, survival and migration of ovarian cancer cells. Gene 2019, 689, 124–130. [Google Scholar] [CrossRef] [PubMed]
  112. Schmidt, D.; Ovitt, C.E.; Anlag, K.; Fehsenfeld, S.; Gredsted, L.; Treier, A.C.; Treier, M. The murine winged-helix transcription factor Foxl2 is required for granulosa cell differentiation and ovary maintenance. Development 2004, 131, 933–942. [Google Scholar] [CrossRef] [Green Version]
  113. Uhlenhaut, N.H.; Jakob, S.; Anlag, K.; Eisenberger, T.; Sekido, R.; Kress, J.; Treier, A.C.; Klugmann, C.; Klasen, C.; Holter, N.I.; et al. Somatic sex reprogramming of adult ovaries to testes by FOXL2 ablation. Cell 2009, 139, 1130–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Takasawa, K.; Kashimada, K.; Pelosi, E.; Takagi, M.; Morio, T.; Asahara, H.; Schlessinger, D.; Mizutani, S.; Koopman, P. FOXL2 transcriptionally represses Sf1 expression by antagonizing WT1 during ovarian development in mice. FASEB J. 2014, 28, 2020–2028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Boulanger, L.; Pannetier, M.; Gall, L.; Allais-Bonnet, A.; Elzaiat, M.; Le Bourhis, D.; Daniel, N.; Richard, C.; Cotinot, C.; Ghyselinck, N.B.; et al. FOXL2 is a female sex-determining gene in the goat. Curr. Biol. 2014, 24, 404–408. [Google Scholar] [CrossRef] [Green Version]
  116. Ottolenghi, C.; Pelosi, E.; Tran, J.; Colombino, M.; Douglass, E.; Nedorezov, T.; Cao, A.; Forabosco, A.; Schlessinger, D. Loss of Wnt4 and Foxl2 leads to female-to-male sex reversal extending to germ cells. Hum. Mol. Genet. 2007, 16, 2795–2804. [Google Scholar] [CrossRef] [Green Version]
  117. Auguste, A.; Chassot, A.A.; Gregoire, E.P.; Renault, L.; Pannetier, M.; Treier, M.; Pailhoux, E.; Chaboissier, M.C. Loss of R-spondin1 and Foxl2 amplifies female-to-male sex reversal in XX mice. Sex. Dev. 2011, 5, 304–317. [Google Scholar] [CrossRef]
  118. Ghochani, Y.; Saini, J.K.; Mellon, P.L.; Thackray, V.G. FOXL2 is involved in the synergy between activin and progestins on the follicle-stimulating hormone beta-subunit promoter. Endocrinology 2012, 153, 2023–2033. [Google Scholar] [CrossRef] [Green Version]
  119. Corpuz, P.S.; Lindaman, L.L.; Mellon, P.L.; Coss, D. FoxL2 Is required for activin induction of the mouse and human follicle-stimulating hormone beta-subunit genes. Mol. Endocrinol. 2010, 24, 1037–1051. [Google Scholar] [CrossRef] [Green Version]
  120. Herndon, M.K.; Nilson, J.H. Maximal expression of Foxl2 in pituitary gonadotropes requires ovarian hormones. PLoS ONE 2015, 10, e126527. [Google Scholar] [CrossRef] [Green Version]
  121. Eozenou, C.; Lesage-Padilla, A.; Mauffre, V.; Healey, G.D.; Camous, S.; Bolifraud, P.; Giraud-Delville, C.; Vaiman, D.; Shimizu, T.; Miyamoto, A.; et al. FOXL2 is a Progesterone Target Gene in the Endometrium of Ruminants. Int. J. Mol. Sci. 2020, 21, 1478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Pisarska, M.D.; Bae, J.; Klein, C.; Hsueh, A.J. Forkhead l2 is expressed in the ovary and represses the promoter activity of the steroidogenic acute regulatory gene. Endocrinology 2004, 145, 3424–3433. [Google Scholar] [CrossRef] [PubMed]
  123. Georges, A.; L’Hote, D.; Todeschini, A.L.; Auguste, A.; Legois, B.; Zider, A.; Veitia, R.A. The transcription factor FOXL2 mobilizes estrogen signaling to maintain the identity of ovarian granulosa cells. eLife 2014, 3, e04207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gustin, S.E.; Hogg, K.; Stringer, J.M.; Rastetter, R.H.; Pelosi, E.; Miles, D.C.; Sinclair, A.H.; Wilhelm, D.; Western, P.S. WNT/beta-catenin and p27/FOXL2 differentially regulate supporting cell proliferation in the developing ovary. Dev. Biol. 2016, 412, 250–260. [Google Scholar] [CrossRef]
  125. Bellessort, B.; Bachelot, A.; Heude, E.; Alfama, G.; Fontaine, A.; Le Cardinal, M.; Treier, M.; Levi, G. Role of Foxl2 in uterine maturation and function. Hum. Mol. Genet. 2015, 24, 3092–3103. [Google Scholar] [CrossRef] [Green Version]
  126. Crisponi, L.; Deiana, M.; Loi, A.; Chiappe, F.; Uda, M.; Amati, P.; Bisceglia, L.; Zelante, L.; Nagaraja, R.; Porcu, S.; et al. The putative forkhead transcription factor FOXL2 is mutated in blepharophimosis/ptosis/epicanthus inversus syndrome. Nat. Genet. 2001, 27, 159–166. [Google Scholar] [CrossRef] [PubMed]
  127. Moumne, L.; Dipietromaria, A.; Batista, F.; Kocer, A.; Fellous, M.; Pailhoux, E.; Veitia, R.A. Differential aggregation and functional impairment induced by polyalanine expansions in FOXL2, a transcription factor involved in cranio-facial and ovarian development. Hum. Mol. Genet. 2008, 17, 1010–1019. [Google Scholar] [CrossRef] [Green Version]
  128. Dipietromaria, A.; Benayoun, B.A.; Todeschini, A.L.; Rivals, I.; Bazin, C.; Veitia, R.A. Towards a functional classification of pathogenic FOXL2 mutations using transactivation reporter systems. Hum. Mol. Genet. 2009, 18, 3324–3333. [Google Scholar] [CrossRef] [Green Version]
  129. Shah, S.P.; Kobel, M.; Senz, J.; Morin, R.D.; Clarke, B.A.; Wiegand, K.C.; Leung, G.; Zayed, A.; Mehl, E.; Kalloger, S.E.; et al. Mutation of FOXL2 in granulosa-cell tumors of the ovary. N. Engl. J. Med. 2009, 360, 2719–2729. [Google Scholar] [CrossRef]
  130. Lima, J.F.; Jin, L.; de Araujo, A.R.; Erikson-Johnson, M.R.; Oliveira, A.M.; Sebo, T.J.; Keeney, G.L.; Medeiros, F. FOXL2 mutations in granulosa cell tumors occurring in males. Arch. Pathol. Lab. Med. 2012, 136, 825–828. [Google Scholar] [CrossRef] [Green Version]
  131. Hes, O.; Vanecek, T.; Petersson, F.; Grossmann, P.; Hora, M.; Perez, M.D.; Steiner, P.; Dvorak, M.; Michal, M. Mutational analysis (c.402C>G) of the FOXL2 gene and immunohistochemical expression of the FOXL2 protein in testicular adult type granulosa cell tumors and incompletely differentiated sex cord stromal tumors. Appl. Immunohistochem. Mol. Morphol. 2011, 19, 347–351. [Google Scholar] [CrossRef] [PubMed]
  132. Karnezis, A.N.; Wang, Y.; Keul, J.; Tessier-Cloutier, B.; Magrill, J.; Kommoss, S.; Senz, J.; Yang, W.; Proctor, L.; Schmidt, D.; et al. DICER1 and FOXL2 Mutation Status Correlates With Clinicopathologic Features in Ovarian Sertoli-Leydig Cell Tumors. Am. J. Surg. Pathol. 2019, 43, 628–638. [Google Scholar] [CrossRef]
  133. Al-Agha, O.M.; Huwait, H.F.; Chow, C.; Yang, W.; Senz, J.; Kalloger, S.E.; Huntsman, D.G.; Young, R.H.; Gilks, C.B. FOXL2 is a sensitive and specific marker for sex cord-stromal tumors of the ovary. Am. J. Surg. Pathol. 2011, 35, 484–494. [Google Scholar] [CrossRef] [PubMed]
  134. Gunes, S.O.; Metin, M.A.; Agarwal, A. Genetic and epigenetic effects in sex determination. Birth Defects Res. Part C Embryo Today Rev. 2016, 108, 321–336. [Google Scholar] [CrossRef]
  135. Kuroki, S.; Matoba, S.; Akiyoshi, M.; Matsumura, Y.; Miyachi, H.; Mise, N.; Abe, K.; Ogura, A.; Wilhelm, D.; Koopman, P.; et al. Epigenetic regulation of mouse sex determination by the histone demethylase Jmjd1a. Science 2013, 341, 1106–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Shendy, N.; Broadhurst, A.L.; Shoemaker, K.; Read, R.; Abell, A.N. MAP3K4 kinase activity dependent control of mouse gonadal sex determinationdagger. Biol. Reprod. 2021, 105, 491–502. [Google Scholar] [CrossRef]
  137. Johnen, H.; Gonzalez-Silva, L.; Carramolino, L.; Flores, J.M.; Torres, M.; Salvador, J.M. Gadd45g is essential for primary sex determination, male fertility and testis development. PLoS ONE 2013, 8, e58751. [Google Scholar] [CrossRef] [Green Version]
  138. Wen, Y.; Ma, X.; Wang, X.; Wang, F.; Dong, J.; Wu, Y.; Lv, C.; Liu, K.; Zhang, Y.; Zhang, Z.; et al. hnRNPU in Sertoli cells cooperates with WT1 and is essential for testicular development by modulating transcriptional factors Sox8/9. Theranostics 2021, 11, 10030–10046. [Google Scholar] [CrossRef] [PubMed]
  139. Josso, N.; Belville, C.; di Clemente, N.; Picard, J.Y. AMH and AMH receptor defects in persistent Mullerian duct syndrome. Hum. Reprod. Update 2005, 11, 351–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Ozisik, G.; Achermann, J.C.; Meeks, J.J.; Jameson, J.L. SF1 in the development of the adrenal gland and gonads. Horm. Res. 2003, 59 (Suppl. S1), 94–98. [Google Scholar] [CrossRef] [PubMed]
  141. Manuylov, N.L.; Fujiwara, Y.; Adameyko, I.I.; Poulat, F.; Tevosian, S.G. The regulation of Sox9 gene expression by the GATA4/FOG2 transcriptional complex in dominant XX sex reversal mouse models. Dev. Biol. 2007, 307, 356–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Wilhelm, D.; Hiramatsu, R.; Mizusaki, H.; Widjaja, L.; Combes, A.N.; Kanai, Y.; Koopman, P. SOX9 regulates prostaglandin D synthase gene transcription in vivo to ensure testis development. J. Biol. Chem. 2007, 282, 10553–10560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Rossitto, M.; Ujjan, S.; Poulat, F.; Boizet-Bonhoure, B. Multiple roles of the prostaglandin D2 signaling pathway in reproduction. Reproduction 2015, 149, R49–R58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Chen, M.; Zhang, L.; Cui, X.; Lin, X.; Li, Y.; Wang, Y.; Wang, Y.; Qin, Y.; Chen, D.; Han, C.; et al. Wt1 directs the lineage specification of sertoli and granulosa cells by repressing Sf1 expression. Development 2017, 144, 44–53. [Google Scholar] [CrossRef] [Green Version]
  145. Gomes, N.L.; de Paula, L.; Silva, J.M.; Silva, T.E.; Lerario, A.M.; Nishi, M.Y.; Batista, R.L.; Faria, J.J.; Moraes, D.; Costa, E.; et al. A 46,XX testicular disorder of sex development caused by a Wilms’ tumour Factor-1 (WT1) pathogenic variant. Clin. Genet. 2019, 95, 172–176. [Google Scholar] [CrossRef] [Green Version]
  146. Eozenou, C.; Gonen, N.; Touzon, M.S.; Jorgensen, A.; Yatsenko, S.A.; Fusee, L.; Kamel, A.K.; Gellen, B.; Guercio, G.; Singh, P.; et al. Testis formation in XX individuals resulting from novel pathogenic variants in Wilms’ tumor 1 (WT1) gene. Proc. Natl. Acad. Sci. USA 2020, 117, 13680–13688. [Google Scholar] [CrossRef]
  147. Bhandari, R.K.; Sadler-Riggleman, I.; Clement, T.M.; Skinner, M.K. Basic helix-loop-helix transcription factor TCF21 is a downstream target of the male sex determining gene SRY. PLoS ONE 2011, 6, e19935. [Google Scholar] [CrossRef] [Green Version]
  148. Katoh-Fukui, Y.; Miyabayashi, K.; Komatsu, T.; Owaki, A.; Baba, T.; Shima, Y.; Kidokoro, T.; Kanai, Y.; Schedl, A.; Wilhelm, D.; et al. Cbx2, a polycomb group gene, is required for Sry gene expression in mice. Endocrinology 2012, 153, 913–924. [Google Scholar] [CrossRef] [Green Version]
  149. Pitetti, J.L.; Calvel, P.; Romero, Y.; Conne, B.; Truong, V.; Papaioannou, M.D.; Schaad, O.; Docquier, M.; Herrera, P.L.; Wilhelm, D.; et al. Insulin and IGF1 receptors are essential for XX and XY gonadal differentiation and adrenal development in mice. PLoS Genet. 2013, 9, e1003160. [Google Scholar] [CrossRef] [Green Version]
  150. Jorgez, C.J.; Seth, A.; Wilken, N.; Bournat, J.C.; Chen, C.H.; Lamb, D.J. E2F1 regulates testicular descent and controls spermatogenesis by influencing WNT4 signaling. Development 2021, 148, dev191189. [Google Scholar] [CrossRef]
  151. Jameson, S.A.; Lin, Y.T.; Capel, B. Testis development requires the repression of Wnt4 by Fgf signaling. Dev. Biol. 2012, 370, 24–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Bagheri-Fam, S.; Bird, A.D.; Zhao, L.; Ryan, J.M.; Yong, M.; Wilhelm, D.; Koopman, P.; Eswarakumar, V.P.; Harley, V.R. Testis Determination Requires a Specific FGFR2 Isoform to Repress FOXL2. Endocrinology 2017, 158, 3832–3843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Yao, H.H.; Matzuk, M.M.; Jorgez, C.J.; Menke, D.B.; Page, D.C.; Swain, A.; Capel, B. Follistatin operates downstream of Wnt4 in mammalian ovary organogenesis. Dev. Dyn. 2004, 230, 210–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Jordan, B.K.; Mohammed, M.; Ching, S.T.; Delot, E.; Chen, X.N.; Dewing, P.; Swain, A.; Rao, P.N.; Elejalde, B.R.; Vilain, E. Up-regulation of WNT-4 signaling and dosage-sensitive sex reversal in humans. Am. J. Hum. Genet. 2001, 68, 1102–1109. [Google Scholar] [CrossRef] [Green Version]
  155. Kashimada, K.; Pelosi, E.; Chen, H.; Schlessinger, D.; Wilhelm, D.; Koopman, P. FOXL2 and BMP2 act cooperatively to regulate follistatin gene expression during ovarian development. Endocrinology 2011, 152, 272–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lindeman, R.E.; Gearhart, M.D.; Minkina, A.; Krentz, A.D.; Bardwell, V.J.; Zarkower, D. Sexual cell-fate reprogramming in the ovary by DMRT1. Curr. Biol. 2015, 25, 764–771. [Google Scholar] [CrossRef] [Green Version]
  157. Lawson, K.A.; Hage, W.J. Clonal analysis of the origin of primordial germ cells in the mouse. Germline Dev. 1994, 165, 68–84. [Google Scholar] [CrossRef]
  158. Ginsburg, M.; Snow, M.H.; McLaren, A. Primordial germ cells in the mouse embryo during gastrulation. Development 1990, 110, 521–528. [Google Scholar] [CrossRef]
  159. Lawson, K.A.; Dunn, N.R.; Roelen, B.A.; Zeinstra, L.M.; Davis, A.M.; Wright, C.V.; Korving, J.P.; Hogan, B.L. Bmp4 is required for the generation of primordial germ cells in the mouse embryo. Genes Dev. 1999, 13, 424–436. [Google Scholar] [CrossRef]
  160. Ying, Y.; Qi, X.; Zhao, G.Q. Induction of primordial germ cells from murine epiblasts by synergistic action of BMP4 and BMP8B signaling pathways. Proc. Natl. Acad. Sci. USA 2001, 98, 7858–7862. [Google Scholar] [CrossRef] [Green Version]
  161. Ying, Y.; Zhao, G.Q. Cooperation of endoderm-derived BMP2 and extraembryonic ectoderm-derived BMP4 in primordial germ cell generation in the mouse. Dev. Biol. 2001, 232, 484–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Avilion, A.A.; Nicolis, S.K.; Pevny, L.H.; Perez, L.; Vivian, N.; Lovell-Badge, R. Multipotent cell lineages in early mouse development depend on SOX2 function. Genes Dev. 2003, 17, 126–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Campolo, F.; Gori, M.; Favaro, R.; Nicolis, S.; Pellegrini, M.; Botti, F.; Rossi, P.; Jannini, E.A.; Dolci, S. Essential role of Sox2 for the establishment and maintenance of the germ cell line. Stem Cells 2013, 31, 1408–1421. [Google Scholar] [CrossRef] [Green Version]
  164. Irie, N.; Weinberger, L.; Tang, W.W.; Kobayashi, T.; Viukov, S.; Manor, Y.S.; Dietmann, S.; Hanna, J.H.; Surani, M.A. SOX17 is a critical specifier of human primordial germ cell fate. Cell 2015, 160, 253–268. [Google Scholar] [CrossRef] [Green Version]
  165. Guo, F.; Yan, L.; Guo, H.; Li, L.; Hu, B.; Zhao, Y.; Yong, J.; Hu, Y.; Wang, X.; Wei, Y.; et al. The Transcriptome and DNA Methylome Landscapes of Human Primordial Germ Cells. Cell 2015, 161, 1437–1452. [Google Scholar] [CrossRef] [Green Version]
  166. Zuo, Q.; Jin, J.; Jin, K.; Zhou, J.; Sun, C.; Song, J.; Chen, G.; Zhang, Y.; Li, B. P53 and H3K4me2 activate N6-methylated LncPGCAT-1 to regulate primordial germ cell formation via MAPK signaling. J. Cell. Physiol. 2020, 235, 9895–9909. [Google Scholar] [CrossRef]
  167. Soto, D.A.; Ross, P.J. Similarities between bovine and human germline development revealed by single-cell RNA sequencing. Reproduction 2021, 161, 239–253. [Google Scholar] [CrossRef]
  168. Ohinata, Y.; Ohta, H.; Shigeta, M.; Yamanaka, K.; Wakayama, T.; Saitou, M. A signaling principle for the specification of the germ cell lineage in mice. Cell 2009, 137, 571–584. [Google Scholar] [CrossRef] [Green Version]
  169. Ohinata, Y.; Payer, B.; O’Carroll, D.; Ancelin, K.; Ono, Y.; Sano, M.; Barton, S.C.; Obukhanych, T.; Nussenzweig, M.; Tarakhovsky, A.; et al. Blimp1 is a critical determinant of the germ cell lineage in mice. Nature 2005, 436, 207–213. [Google Scholar] [CrossRef]
  170. Mochizuki, K.; Hayashi, Y.; Sekinaka, T.; Otsuka, K.; Ito-Matsuoka, Y.; Kobayashi, H.; Oki, S.; Takehara, A.; Kono, T.; Osumi, N.; et al. Repression of Somatic Genes by Selective Recruitment of HDAC3 by BLIMP1 Is Essential for Mouse Primordial Germ Cell Fate Determination. Cell Rep. 2018, 24, 2682–2693. [Google Scholar] [CrossRef] [Green Version]
  171. Yamaguchi, Y.L.; Tanaka, S.S.; Kumagai, M.; Fujimoto, Y.; Terabayashi, T.; Matsui, Y.; Nishinakamura, R. Sall4 is essential for mouse primordial germ cell specification by suppressing somatic cell program genes. Stem Cells 2015, 33, 289–300. [Google Scholar] [CrossRef] [PubMed]
  172. West, J.A.; Viswanathan, S.R.; Yabuuchi, A.; Cunniff, K.; Takeuchi, A.; Park, I.H.; Sero, J.E.; Zhu, H.; Perez-Atayde, A.; Frazier, A.L.; et al. A role for Lin28 in primordial germ-cell development and germ-cell malignancy. Nature 2009, 460, 909–913. [Google Scholar] [CrossRef] [Green Version]
  173. Yamaji, M.; Seki, Y.; Kurimoto, K.; Yabuta, Y.; Yuasa, M.; Shigeta, M.; Yamanaka, K.; Ohinata, Y.; Saitou, M. Critical function of Prdm14 for the establishment of the germ cell lineage in mice. Nat. Genet. 2008, 40, 1016–1022. [Google Scholar] [CrossRef] [PubMed]
  174. Sybirna, A.; Tang, W.; Pierson, S.M.; Dietmann, S.; Gruhn, W.H.; Brosh, R.; Surani, M.A. A critical role of PRDM14 in human primordial germ cell fate revealed by inducible degrons. Nat. Commun. 2020, 11, 1282. [Google Scholar] [CrossRef]
  175. Aramaki, S.; Hayashi, K.; Kurimoto, K.; Ohta, H.; Yabuta, Y.; Iwanari, H.; Mochizuki, Y.; Hamakubo, T.; Kato, Y.; Shirahige, K.; et al. A mesodermal factor, T, specifies mouse germ cell fate by directly activating germline determinants. Dev. Cell 2013, 27, 516–529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Weber, S.; Eckert, D.; Nettersheim, D.; Gillis, A.J.; Schafer, S.; Kuckenberg, P.; Ehlermann, J.; Werling, U.; Biermann, K.; Looijenga, L.H.; et al. Critical function of AP-2 gamma/TCFAP2C in mouse embryonic germ cell maintenance. Biol. Reprod. 2010, 82, 214–223. [Google Scholar] [CrossRef] [Green Version]
  177. Tremblay, K.D.; Dunn, N.R.; Robertson, E.J. Mouse embryos lacking Smad1 signals display defects in extra-embryonic tissues and germ cell formation. Development 2001, 128, 3609–3621. [Google Scholar] [CrossRef]
  178. Chang, H.; Matzuk, M.M. Smad5 is required for mouse primordial germ cell development. Mech. Dev. 2001, 104, 61–67. [Google Scholar] [CrossRef]
  179. Mochizuki, K.; Tando, Y.; Sekinaka, T.; Otsuka, K.; Hayashi, Y.; Kobayashi, H.; Kamio, A.; Ito-Matsuoka, Y.; Takehara, A.; Kono, T.; et al. SETDB1 is essential for mouse primordial germ cell fate determination by ensuring BMP signaling. Development 2018, 145, dev164160. [Google Scholar] [CrossRef] [Green Version]
  180. Nady, N.; Gupta, A.; Ma, Z.; Swigut, T.; Koide, A.; Koide, S.; Wysocka, J. ETO family protein Mtgr1 mediates Prdm14 functions in stem cell maintenance and primordial germ cell formation. eLife 2015, 4, e10150. [Google Scholar] [CrossRef] [Green Version]
  181. Okamura, E.; Tam, O.H.; Posfai, E.; Li, L.; Cockburn, K.; Lee, C.; Garner, J.; Rossant, J. Esrrb function is required for proper primordial germ cell development in presomite stage mouse embryos. Dev. Biol. 2019, 455, 382–392. [Google Scholar] [CrossRef] [PubMed]
  182. Fang, F.; Angulo, B.; Xia, N.; Sukhwani, M.; Wang, Z.; Carey, C.C.; Mazurie, A.; Cui, J.; Wilkinson, R.; Wiedenheft, B.; et al. A PAX5-OCT4-PRDM1 developmental switch specifies human primordial germ cells. Nat. Cell Biol. 2018, 20, 655–665. [Google Scholar] [CrossRef] [PubMed]
  183. Zuo, Q.; Jin, K.; Song, J.; Zhang, Y.; Chen, G.; Li, B. Interaction of the primordial germ cell-specific protein C2EIP with PTCH2 directs differentiation of embryonic stem cells via HH signaling activation. Cell Death Dis. 2018, 9, 497. [Google Scholar] [CrossRef] [PubMed]
  184. De Felici, M.; McLaren, A. Isolation of mouse primordial germ cells. Exp. Cell Res. 1982, 142, 476–482. [Google Scholar] [CrossRef]
  185. Kuhholzer, B.; Baguisi, A.; Overstrom, E.W. Long-term culture and characterization of goat primordial germ cells. Theriogenology 2000, 53, 1071–1079. [Google Scholar] [CrossRef]
  186. Kakegawa, R.; Teramura, T.; Takehara, T.; Anzai, M.; Mitani, T.; Matsumoto, K.; Saeki, K.; Sagawa, N.; Fukuda, K.; Hosoi, Y. Isolation and culture of rabbit primordial germ cells. J. Reprod. Dev. 2008, 54, 352–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Liu, C.X.; Wang, W.L.; Zhao, R.Y.; Wang, H.T.; Liu, Y.Y.; Wang, S.Y.; Zhou, H.M. Isolation, culture, and characterization of primordial germ cells in Mongolian sheep. Vitr. Cell. Dev. Biol.-Anim. 2014, 50, 207–213. [Google Scholar] [CrossRef]
  188. Tilgner, K.; Atkinson, S.P.; Golebiewska, A.; Stojkovic, M.; Lako, M.; Armstrong, L. Isolation of primordial germ cells from differentiating human embryonic stem cells. Stem Cells 2008, 26, 3075–3085. [Google Scholar] [CrossRef]
  189. Costa, J.; Souza, G.B.; Soares, M.; Ribeiro, R.P.; van den Hurk, R.; Silva, J. In vitro differentiation of primordial germ cells and oocyte-like cells from stem cells. Histol. Histopathol. 2018, 33, 121–132. [Google Scholar] [CrossRef]
  190. Murase, Y.; Yabuta, Y.; Ohta, H.; Yamashiro, C.; Nakamura, T.; Yamamoto, T.; Saitou, M. Long-term expansion with germline potential of human primordial germ cell-like cells in vitro. EMBO J. 2020, 39, e104929. [Google Scholar] [CrossRef]
  191. Morohaku, K.; Tanimoto, R.; Sasaki, K.; Kawahara-Miki, R.; Kono, T.; Hayashi, K.; Hirao, Y.; Obata, Y. Complete in vitro generation of fertile oocytes from mouse primordial germ cells. Proc. Natl. Acad. Sci. USA 2016, 113, 9021–9026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Hayashi, K.; Ogushi, S.; Kurimoto, K.; Shimamoto, S.; Ohta, H.; Saitou, M. Offspring from oocytes derived from in vitro primordial germ cell-like cells in mice. Science 2012, 338, 971–975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Chen, D.; Liu, W.; Zimmerman, J.; Pastor, W.A.; Kim, R.; Hosohama, L.; Ho, J.; Aslanyan, M.; Gell, J.J.; Jacobsen, S.E.; et al. The TFAP2C-Regulated OCT4 Naive Enhancer Is Involved in Human Germline Formation. Cell Rep. 2018, 25, 3591–3602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Chen, D.; Sun, N.; Hou, L.; Kim, R.; Faith, J.; Aslanyan, M.; Tao, Y.; Zheng, Y.; Fu, J.; Liu, W.; et al. Human Primordial Germ Cells Are Specified from Lineage-Primed Progenitors. Cell Rep. 2019, 29, 4568–4582. [Google Scholar] [CrossRef] [Green Version]
  195. Ghasemi, H.H.; Sobhani, A.; Nazm, B.M. Multipotent SSEA1 Positive Cells Population Differentiation into Primordial Germ Cells and Subsequently Progress into Oocyte-like Cells. Arch. Iran. Med. 2015, 18, 404–410. [Google Scholar]
  196. Mall, E.M.; Lecanda, A.; Drexler, H.; Raz, E.; Scholer, H.R.; Schlatt, S. Heading towards a dead end: The role of DND1 in germ line differentiation of human iPSCs. PLoS ONE 2021, 16, e258427. [Google Scholar] [CrossRef]
  197. Pierson, S.M.; Sybirna, A.; Wong, F.; Surani, M.A. Testing the role of SOX15 in human primordial germ cell fate. Wellcome Open Res. 2019, 4, 122. [Google Scholar] [CrossRef]
  198. Abdyyev, V.K.; Sant, D.W.; Kiseleva, E.V.; Spangenberg, V.E.; Kolomiets, O.L.; Andrade, N.S.; Dashinimaev, E.B.; Vorotelyak, E.A.; Vasiliev, A.V. In vitro derived female hPGCLCs are unable to complete meiosis in embryoid bodies. Exp. Cell Res. 2020, 397, 112358. [Google Scholar] [CrossRef]
  199. West, F.D.; Machacek, D.W.; Boyd, N.L.; Pandiyan, K.; Robbins, K.R.; Stice, S.L. Enrichment and differentiation of human germ-like cells mediated by feeder cells and basic fibroblast growth factor signaling. Stem Cells 2008, 26, 2768–2776. [Google Scholar] [CrossRef]
  200. Park, T.S.; Galic, Z.; Conway, A.E.; Lindgren, A.; van Handel, B.J.; Magnusson, M.; Richter, L.; Teitell, M.A.; Mikkola, H.K.; Lowry, W.E.; et al. Derivation of primordial germ cells from human embryonic and induced pluripotent stem cells is significantly improved by coculture with human fetal gonadal cells. Stem Cells 2009, 27, 783–795. [Google Scholar] [CrossRef] [Green Version]
  201. Geijsen, N.; Horoschak, M.; Kim, K.; Gribnau, J.; Eggan, K.; Daley, G.Q. Derivation of embryonic germ cells and male gametes from embryonic stem cells. Nature 2004, 427, 148–154. [Google Scholar] [CrossRef] [PubMed]
  202. Tan, H.; Wang, J.J.; Cheng, S.F.; Ge, W.; Sun, Y.C.; Sun, X.F.; Sun, R.; Li, L.; Li, B.; Shen, W. Retinoic acid promotes the proliferation of primordial germ cell-like cells differentiated from mouse skin-derived stem cells in vitro. Theriogenology 2016, 85, 408–418. [Google Scholar] [CrossRef] [PubMed]
  203. Cheng, T.; Zhai, K.; Chang, Y.; Yao, G.; He, J.; Wang, F.; Kong, H.; Xin, H.; Wang, H.; Jin, M.; et al. CHIR99021 combined with retinoic acid promotes the differentiation of primordial germ cells from human embryonic stem cells. Oncotarget 2017, 8, 7814–7826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhang, M.Y.; Tian, Y.; Zhang, S.E.; Yan, H.C.; Ge, W.; Han, B.Q.; Yan, Z.H.; Cheng, S.F.; Shen, W. The proliferation role of LH on porcine primordial germ cell-like cells (pPGCLCs) through ceRNA network construction. Clin. Transl. Med. 2021, 11, e560. [Google Scholar] [CrossRef] [PubMed]
  205. Mahboudi, S.; Parivar, K.; Mazaheri, Z.; Irani, S.H. Mir-106b Cluster Regulates Primordial Germ Cells Differentiation from Human Mesenchymal Stem Cells. Cell J. 2021, 23, 294–302. [Google Scholar] [CrossRef]
  206. Xing, M.; Wang, N.; Zeng, H.; Zhang, J. alpha-ketoglutarate promotes the specialization of primordial germ cell-like cells through regulating epigenetic reprogramming. J. Biomed. Res. 2020, 35, 36–46. [Google Scholar] [CrossRef]
  207. Ando, Y.; Okeyo, K.O.; Adachi, T. Modulation of adhesion microenvironment using mesh substrates triggers self-organization and primordial germ cell-like differentiation in mouse ES cells. APL Bioeng. 2019, 3, 16102. [Google Scholar] [CrossRef]
  208. Ooi, S.; Jiang, H.; Kang, Y.; Allard, P. Examining the Developmental Trajectory of an in Vitro Model of Mouse Primordial Germ Cells following Exposure to Environmentally Relevant Bisphenol A Levels. Environ. Health Perspect. 2021, 129, 97013. [Google Scholar] [CrossRef]
  209. Nagamatsu, G.; Saito, S.; Takubo, K.; Suda, T. Integrative Analysis of the Acquisition of Pluripotency in PGCs Reveals the Mutually Exclusive Roles of Blimp-1 and AKT Signaling. Stem Cell Rep. 2015, 5, 111–124. [Google Scholar] [CrossRef] [Green Version]
  210. Takehara, A.; Matsui, Y. Shortened G1 phase of cell cycle and decreased histone H3K27 methylation are associated with AKT-induced enhancement of primordial germ cell reprogramming. Dev. Growth Differ. 2019, 61, 357–364. [Google Scholar] [CrossRef]
  211. Oliveros-Etter, M.; Li, Z.; Nee, K.; Hosohama, L.; Hargan-Calvopina, J.; Lee, S.A.; Joti, P.; Yu, J.; Clark, A.T. PGC Reversion to Pluripotency Involves Erasure of DNA Methylation from Imprinting Control Centers followed by Locus-Specific Re-methylation. Stem Cell Rep. 2015, 5, 337–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Bazley, F.A.; Liu, C.F.; Yuan, X.; Hao, H.; All, A.H.; De Los, A.A.; Zambidis, E.T.; Gearhart, J.D.; Kerr, C.L. Direct Reprogramming of Human Primordial Germ Cells into Induced Pluripotent Stem Cells: Efficient Generation of Genetically Engineered Germ Cells. Stem Cells Dev. 2015, 24, 2634–2648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Mansouri, V.; Salehi, M.; Nourozian, M.; Fadaei, F.; Farahani, R.M.; Piryaei, A.; Delbari, A. The ability of mouse nuclear transfer embryonic stem cells to differentiate into primordial germ cells. Genet. Mol. Biol. 2015, 38, 220–226. [Google Scholar] [CrossRef] [Green Version]
  214. Makiyan, Z. Endometriosis origin from primordial germ cells. Organogenesis 2017, 13, 95–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Molyneaux, K.; Wylie, C. Primordial germ cell migration. Int. J. Dev. Biol. 2004, 48, 537–544. [Google Scholar] [CrossRef] [PubMed]
  216. Richardson, B.E.; Lehmann, R. Mechanisms guiding primordial germ cell migration: Strategies from different organisms. Nat. Rev. Mol. Cell Biol. 2010, 11, 37–49. [Google Scholar] [CrossRef] [Green Version]
  217. De Felici, M. The Formation and Migration of Primordial Germ Cells in Mouse and Man. In Molecular Mechanisms of Cell Differentiation in Gonad Development; Springer: Cham, Switzerland, 2016; Volume 58, pp. 23–46. [Google Scholar] [CrossRef]
  218. Lange, U.C.; Adams, D.J.; Lee, C.; Barton, S.; Schneider, R.; Bradley, A.; Surani, M.A. Normal germ line establishment in mice carrying a deletion of the Ifitm/Fragilis gene family cluster. Mol. Cell. Biol. 2008, 28, 4688–4696. [Google Scholar] [CrossRef] [Green Version]
  219. Sun, J.; Ting, M.C.; Ishii, M.; Maxson, R. Msx1 and Msx2 function together in the regulation of primordial germ cell migration in the mouse. Dev. Biol. 2016, 417, 11–24. [Google Scholar] [CrossRef] [Green Version]
  220. Laird, D.J.; Altshuler-Keylin, S.; Kissner, M.D.; Zhou, X.; Anderson, K.V. Ror2 enhances polarity and directional migration of primordial germ cells. PLoS Genet. 2011, 7, e1002428. [Google Scholar] [CrossRef] [Green Version]
  221. Park, E.; Lee, B.; Clurman, B.E.; Lee, K. NUP50 is necessary for the survival of primordial germ cells in mouse embryos. Reproduction 2016, 151, 51–58. [Google Scholar] [CrossRef] [Green Version]
  222. Senft, A.D.; Bikoff, E.K.; Robertson, E.J.; Costello, I. Genetic dissection of Nodal and Bmp signalling requirements during primordial germ cell development in mouse. Nat. Commun. 2019, 10, 1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Shimaoka, K.; Mukumoto, Y.; Tanigawa, Y.; Komiya, T. Xenopus Vasa Homolog XVLG1 is Essential for Migration and Survival of Primordial Germ Cells. Zool. Sci. 2017, 34, 93–104. [Google Scholar] [CrossRef] [PubMed]
  224. Lejong, M.; Choa-Duterre, M.; Vanmuylder, N.; Louryan, S. Geldanamycin administration reduces the amount of primordial germ cells in the mouse embryo. Morphologie 2018, 102, 219–224. [Google Scholar] [CrossRef] [PubMed]
  225. Yamashiro, C.; Hirota, T.; Kurimoto, K.; Nakamura, T.; Yabuta, Y.; Nagaoka, S.I.; Ohta, H.; Yamamoto, T.; Saitou, M. Persistent Requirement and Alteration of the Key Targets of PRDM1 During Primordial Germ Cell Development in Mice. Biol. Reprod. 2016, 94, 7. [Google Scholar] [CrossRef]
  226. Kim, Y.; Lee, J.; Seppala, M.; Cobourne, M.T.; Kim, S.H. Ptch2/Gas1 and Ptch1/Boc differentially regulate Hedgehog signalling in murine primordial germ cell migration. Nat. Commun. 2020, 11, 1994. [Google Scholar] [CrossRef] [Green Version]
  227. Mallol, A.; Guirola, M.; Payer, B. PRDM14 controls X-chromosomal and global epigenetic reprogramming of H3K27me3 in migrating mouse primordial germ cells. Epigenetics Chromatin 2019, 12, 38. [Google Scholar] [CrossRef] [Green Version]
  228. Hoyer, P.E.; Byskov, A.G.; Mollgard, K. Stem cell factor and c-Kit in human primordial germ cells and fetal ovaries. Mol. Cell. Endocrinol. 2005, 234, 1–10. [Google Scholar] [CrossRef]
  229. Mollgard, K.; Jespersen, A.; Lutterodt, M.C.; Yding, A.C.; Hoyer, P.E.; Byskov, A.G. Human primordial germ cells migrate along nerve fibers and Schwann cells from the dorsal hind gut mesentery to the gonadal ridge. Mol. Hum. Reprod. 2010, 16, 621–631. [Google Scholar] [CrossRef] [Green Version]
  230. Wolff, E.; Suplicki, M.M.; Behr, R. Primordial germ cells do not migrate along nerve fibres in marmoset monkey and mouse embryos. Reproduction 2019, 157, 101–109. [Google Scholar] [CrossRef] [Green Version]
  231. Luo, Y.; Schimenti, J.C. MCM9 deficiency delays primordial germ cell proliferation independent of the ATM pathway. Genesis 2015, 53, 678–684. [Google Scholar] [CrossRef]
  232. Kim, S.; Gunesdogan, U.; Zylicz, J.J.; Hackett, J.A.; Cougot, D.; Bao, S.; Lee, C.; Dietmann, S.; Allen, G.E.; Sengupta, R.; et al. PRMT5 protects genomic integrity during global DNA demethylation in primordial germ cells and preimplantation embryos. Mol. Cell 2014, 56, 564–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Bloom, J.C.; Schimenti, J.C. Sexually dimorphic DNA damage responses and mutation avoidance in the mouse germline. Genes Dev. 2020, 34, 1637–1649. [Google Scholar] [CrossRef] [PubMed]
  234. Cantu, A.V.; Altshuler-Keylin, S.; Laird, D.J. Discrete somatic niches coordinate proliferation and migration of primordial germ cells via Wnt signaling. J. Cell Biol. 2016, 214, 215–229. [Google Scholar] [CrossRef] [Green Version]
  235. Risal, S.; Zhang, J.; Adhikari, D.; Liu, X.; Shao, J.; Hu, M.; Busayavalasa, K.; Tu, Z.; Chen, Z.; Kaldis, P.; et al. MASTL is essential for anaphase entry of proliferating primordial germ cells and establishment of female germ cells in mice. Cell Discov. 2017, 3, 16052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Teng, H.; Sui, X.; Zhou, C.; Shen, C.; Yang, Y.; Zhang, P.; Guo, X.; Huo, R. Fatty acid degradation plays an essential role in proliferation of mouse female primordial germ cells via the p53-dependent cell cycle regulation. Cell Cycle 2016, 15, 425–431. [Google Scholar] [CrossRef] [Green Version]
  237. Sorrenti, M.; Klinger, F.G.; Iona, S.; Rossi, V.; Marcozzi, S.; De Felici, M. Expression and possible roles of extracellular signal-related kinases 1–2 (ERK1–2) in mouse primordial germ cell development. J. Reprod. Dev. 2020, 66, 399–409. [Google Scholar] [CrossRef]
  238. Ulu, F.; Kim, S.M.; Yokoyama, T.; Yamazaki, Y. Dose-dependent functions of fibroblast growth factor 9 regulate the fate of murine XY primordial germ cells. Biol. Reprod. 2017, 96, 122–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Tian-Zhong, M.; Bi, C.; Ying, Z.; Xia, J.; Cai-Ling, P.; Yun-Shan, Z.; Mei-Wen, H.; Yan-Ru, N. Critical role of Emx2 in the pluripotency-differentiation transition in male gonocytes via regulation of FGF9/NODAL pathway. Reproduction 2016, 151, 673–681. [Google Scholar] [CrossRef] [Green Version]
  240. Hu, Y.C.; Nicholls, P.K.; Soh, Y.Q.; Daniele, J.R.; Junker, J.P.; van Oudenaarden, A.; Page, D.C. Licensing of primordial germ cells for gametogenesis depends on genital ridge signaling. PLoS Genet. 2015, 11, e1005019. [Google Scholar] [CrossRef] [Green Version]
  241. Overeem, A.W.; Chang, Y.W.; Spruit, J.; Roelse, C.M.; Chuva, D.S.L.S. Ligand-Receptor Interactions Elucidate Sex-Specific Pathways in the Trajectory From Primordial Germ Cells to Gonia During Human Development. Front. Cell Dev. Biol. 2021, 9, 661243. [Google Scholar] [CrossRef]
  242. Chassot, A.A.; Le Rolle, M.; Jourden, M.; Taketo, M.M.; Ghyselinck, N.B.; Chaboissier, M.C. Constitutive WNT/CTNNB1 activation triggers spermatogonial stem cell proliferation and germ cell depletion. Dev. Biol. 2017, 426, 17–27. [Google Scholar] [CrossRef] [PubMed]
  243. Le Rolle, M.; Massa, F.; Siggers, P.; Turchi, L.; Loubat, A.; Koo, B.K.; Clevers, H.; Greenfield, A.; Schedl, A.; Chaboissier, M.C.; et al. Arrest of WNT/beta-catenin signaling enables the transition from pluripotent to differentiated germ cells in mouse ovaries. Proc. Natl. Acad. Sci. USA 2021, 118, e2023376118. [Google Scholar] [CrossRef] [PubMed]
  244. Yadu, N.; Kumar, P.G. Retinoic acid signaling in regulation of meiosis during embryonic development in mice. Genesis 2019, 57, e23327. [Google Scholar] [CrossRef]
  245. Barrios, F.; Filipponi, D.; Pellegrini, M.; Paronetto, M.P.; Di Siena, S.; Geremia, R.; Rossi, P.; De Felici, M.; Jannini, E.A.; Dolci, S. Opposing effects of retinoic acid and FGF9 on Nanos2 expression and meiotic entry of mouse germ cells. J. Cell Sci. 2010, 123, 871–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Suzuki, A.; Saga, Y. Nanos2 suppresses meiosis and promotes male germ cell differentiation. Genes Dev. 2008, 22, 430–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Wu, Q.; Fukuda, K.; Weinstein, M.; Graff, J.M.; Saga, Y. SMAD2 and p38 signaling pathways act in concert to determine XY primordial germ cell fate in mice. Development 2015, 142, 575–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Wang, P.; Miao, Y.; Li, X.H.; Zhang, N.; Wang, Q.; Yue, W.; Sun, S.C.; Xiong, B.; Qiao, J.; Li, M. Proteome landscape and spatial map of mouse primordial germ cells. Sci. China Life Sci. 2021, 64, 966–981. [Google Scholar] [CrossRef]
  249. Eguizabal, C.; Herrera, L.; De Onate, L.; Montserrat, N.; Hajkova, P.; Izpisua, B.J. Characterization of the Epigenetic Changes During Human Gonadal Primordial Germ Cells Reprogramming. Stem Cells 2016, 34, 2418–2428. [Google Scholar] [CrossRef] [Green Version]
  250. Tang, W.W.; Dietmann, S.; Irie, N.; Leitch, H.G.; Floros, V.I.; Bradshaw, C.R.; Hackett, J.A.; Chinnery, P.F.; Surani, M.A. A Unique Gene Regulatory Network Resets the Human Germline Epigenome for Development. Cell 2015, 161, 1453–1467. [Google Scholar] [CrossRef] [Green Version]
  251. Seki, Y. PRDM14 Is a Unique Epigenetic Regulator Stabilizing Transcriptional Networks for Pluripotency. Front. Cell Dev. Biol. 2018, 6, 12. [Google Scholar] [CrossRef] [Green Version]
  252. Hill, P.; Leitch, H.G.; Requena, C.E.; Sun, Z.; Amouroux, R.; Roman-Trufero, M.; Borkowska, M.; Terragni, J.; Vaisvila, R.; Linnett, S.; et al. Epigenetic reprogramming enables the transition from primordial germ cell to gonocyte. Nature 2018, 555, 392–396. [Google Scholar] [CrossRef] [PubMed]
  253. Ito, T.; Osada, A.; Ohta, M.; Yokota, K.; Nishiyama, A.; Niikura, Y.; Tamura, T.; Sekita, Y.; Kimura, T. SWI/SNF chromatin remodeling complex is required for initiation of sex-dependent differentiation in mouse germline. Sci. Rep. 2021, 11, 24074. [Google Scholar] [CrossRef] [PubMed]
  254. Ruthig, V.A.; Friedersdorf, M.B.; Garness, J.A.; Munger, S.C.; Bunce, C.; Keene, J.D.; Capel, B. The RNA-binding protein DND1 acts sequentially as a negative regulator of pluripotency and a positive regulator of epigenetic modifiers required for germ cell reprogramming. Development 2019, 146, dev175950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Moratilla, A.; Sainz, D.L.M.D.; Cadenas, M.M.; Lopez-Iglesias, P.; Gonzalez-Peramato, P.; De Miguel, M.P. Inhibition of PKCepsilon induces primordial germ cell reprogramming into pluripotency by HIF1&2 upregulation and histone acetylation. Am. J. Stem Cells 2021, 10, 1–17. [Google Scholar]
  256. Fernandez-Perez, D.; Brieno-Enriquez, M.A.; Isoler-Alcaraz, J.; Larriba, E.; Del, M.J. MicroRNA dynamics at the onset of primordial germ and somatic cell sex differentiation during mouse embryonic gonad development. RNA 2018, 24, 287–303. [Google Scholar] [CrossRef] [Green Version]
  257. Boyer, A.; Yeh, J.R.; Zhang, X.; Paquet, M.; Gaudin, A.; Nagano, M.C.; Boerboom, D. CTNNB1 signaling in sertoli cells downregulates spermatogonial stem cell activity via WNT4. PLoS ONE 2012, 7, e29764. [Google Scholar] [CrossRef]
  258. Liu, C.F.; Parker, K.; Yao, H.H. WNT4/beta-catenin pathway maintains female germ cell survival by inhibiting activin betaB in the mouse fetal ovary. PLoS ONE 2010, 5, e10382. [Google Scholar] [CrossRef]
Figure 1. Schematic illustrations of the development of gonad and primordial germ cells (PGCs). At 4–5 weeks of pregnancy in humans (E9.5 in mice), coelomic epithelial cells (green) start to proliferate on the ventromedial surface of the mesonephros, and the genital ridge (brown) appears. At five weeks of pregnancy in humans (E10–E11 in mice), coelomic epithelial cells continue to proliferate and form bipotential gonads under the regulation of NR5A1, GATA4, etc. Subsequently, bipotential gonad differentiates into testis and ovary, respectively, through a sex-related genes antagonistic network. Among them, SRY, WNT4, RSPO1, and FOXL2 factors. In addition, the Wolffian duct (blue) and Müllerian duct (pink) form epididymis, vas deferens, and seminal vesicles or fallopian tubes, uterus, and part of the vagina, respectively. On the other hand, PGCs originate from a subpopulation of cells in the proximal epiblast (yellow). At two weeks of pregnancy in humans (E6.5 in mice), extraembryonic ectoderm and visceral endoderm secrete signals BMP signals, in turn, activate the expression of WNT3, PRDM1, etc. and induce the specification of PGCs. At five weeks of pregnancy in humans (E10–E11 in mice), PGCs have migrated into the bipotential gonad. After several rounds of cell division and a global change in gene expression, PGCs differentiate into sperm or oocyte after receiving signals from the forming testis or ovary and the nearby mesonephric tissue.
Figure 1. Schematic illustrations of the development of gonad and primordial germ cells (PGCs). At 4–5 weeks of pregnancy in humans (E9.5 in mice), coelomic epithelial cells (green) start to proliferate on the ventromedial surface of the mesonephros, and the genital ridge (brown) appears. At five weeks of pregnancy in humans (E10–E11 in mice), coelomic epithelial cells continue to proliferate and form bipotential gonads under the regulation of NR5A1, GATA4, etc. Subsequently, bipotential gonad differentiates into testis and ovary, respectively, through a sex-related genes antagonistic network. Among them, SRY, WNT4, RSPO1, and FOXL2 factors. In addition, the Wolffian duct (blue) and Müllerian duct (pink) form epididymis, vas deferens, and seminal vesicles or fallopian tubes, uterus, and part of the vagina, respectively. On the other hand, PGCs originate from a subpopulation of cells in the proximal epiblast (yellow). At two weeks of pregnancy in humans (E6.5 in mice), extraembryonic ectoderm and visceral endoderm secrete signals BMP signals, in turn, activate the expression of WNT3, PRDM1, etc. and induce the specification of PGCs. At five weeks of pregnancy in humans (E10–E11 in mice), PGCs have migrated into the bipotential gonad. After several rounds of cell division and a global change in gene expression, PGCs differentiate into sperm or oocyte after receiving signals from the forming testis or ovary and the nearby mesonephric tissue.
Ijms 23 07500 g001
Table 1. Genes involved in genital ridge formation.
Table 1. Genes involved in genital ridge formation.
GenesFunction in Organogenesis and DSDReferences
NR5A1NR5A1 is involved in the development of the gonad, adrenal gland, and pituitary.[25,26,27]
NR5A1 variants are associated with male infertility and DSD in humans males and females.[28,29,30,31,32,33,34]
GATA4GATA4 is required for the proliferation of coelomic epithelial cells and is involved in gonadal development by interacting with FOG2. [35,36]
The GATA4 mutant protein failed to bind with FOG2, resulting in DSD.[37]
WT1WT1 plays a distinct role in gonadal formation and development by maintaining somatic cell survival.[38]
WT1 mutations are responsible for Frasier syndrome with streak gonads.[39,40]
LHX9LHX9 participates in gonad formation by regulating cell proliferation.[41]
EMX2EMX2 is required for the migration and survival of cells in the mesenchymal compartment and involves GR formation by regulating NR5A1 expression.[42]
SIX1 and SIX4SIX1 and SIX4 have a functional redundancy and mainly function in the proliferation of supporting cell precursors and steroidogenic cell precursors.[43]
POD1POD1 is essential for gonadal development by regulating Nr5a1 expression.[44]
CBX2CBX2 is required for splenic vascular, adrenal gland, and gonad formation. [45,46]
INSR and IGF1RINSR and IGFIR regulate somatic progenitor cell proliferation by mediating Insulin and its growth factors (IGF1 and IGF2) during GR formation.[47]
PBX1PBX1 is involved in progenitor cell proliferation in GR by regulating the expression of NR5A1. [48]
PBX1 mutation abolishes its interaction with CBX2 and EMX2, causing gonadal dysgenesis and radiocubital synostosis in humans.[49]
ODD1ODD1 regulates the expression of LHX1, PAX2, and WT1, inhibiting cell apoptosis in nephrogenic mesenchyme and participating in gonadal development.[50]
Table 2. Genes involved in gonadal sexual differentiation.
Table 2. Genes involved in gonadal sexual differentiation.
GenesFunctionsReferences
JMJD1AJMJD1A is involved in the H3K9 demethylation in SRY, and JMJD1A deficiency presents a decrease in SRY expression. [135]
MAP3K4MAP3K4 is involved in SRY expression, and loss of MAP3K4 will lead to a male sex reversal.[136]
GADD45γGADD45γ is upstream of MAP3K4, without which will cause male sexual reversal.[72]
p38α and p38βp38α and p38β are members of p38 MAPK family, and loss of p38α and p38β will causes disruption to SRY expression and XY embryonic gonadal sex reversal.
GADD45GGADD45G is necessary for SRY expression.[137]
HNRNPUHNRNPU enhances the expression of SOX8 and SOX9 by interacting with WT1 and SOX9.[138]
AMH and AMHR2AMH and its receptor AMHR2 are involved in the normal development of the accessory gland. [139]
NR5A1NR5A1 is involved in testis formation by cooperating with other regulators such as WT1, DAX1, SRY, and SOX9.[140]
GATA4The interaction of GATA4 and FOG2 is important in sex differentiation because it regulates the expression of SRY and AMH.[36,141]
FOG2
PTGDSPTGDS is one of the downstream targets of SOX9, which involves the production of prostaglandin D2, maintaining the sustained expression of SOX9 in testis. [142,143]
WT1WT1 is a potential upstream of SRY and controls somatic cells’ fate through regulating NR5A1 expression. [38,144]
WT1 variants lead to 46,XX testicular DSD in humans.[145,146]
SIX1 and SIX4SIX1 and SIX4 regulate SRY expression by activating FOG2 expression, regulating male sex determination. [43]
POD1The downstream target of SRY is POD1, which is involved in the formation of testicular cords and testis-specific coelomic vessels.[44,147]
CBX2CBX2 regulates SRY expression by interfering with upstream steps. [46,148]
INSR and IGF1RINSR and IGF1R have potential feedback interactions between WNT4 and RSPO1 signaling pathways. [84]
INSR and IGF1R are involved in the adrenal specification, testicular differentiation, and ovarian development by regulating the expression of sex-related genes, including WT1, LHX9, and NR5A1.[149]
FGF9 and PGD2FGF9 and PGD2 are downstream of SOX9 and are involved in supporting-to-Sertoli cell differentiation by activating testis-related genes and repressing anti-testis genes. [79]
E2F1E2F1 regulates testicular descent and controls spermatogenesis by repressing WNT4 expression[150]
FGF9 and FGFR2FGF9 and FGFR2 are required in testis development by repressing the expression of WNT4 and FOXL2.[151,152]
FSTFST prevents testis-specific vasculature formation by antagonizing Activin B action through WNT4 activation.[153]
DAX1DAX1 is downstream of WNT4 and is involved in gonadal sexual differentiation by antagonizing SRY expression.[154]
BMP2BMP2 acts cooperatively with FOXL2 to regulate FST gene expression during ovarian development.[155]
DMRT1DMRT1 is required for sexual differentiation of somatic and germ cells by silencing FOXL2 expression.[156]
Table 3. Positive and negative signals directed PGC fate.
Table 3. Positive and negative signals directed PGC fate.
GenesFunctionsReferences
WNT3WNT3 enables PE cells to receive a BMP4 signal.[168]
PRDM1PRDM1 is Likely downstream of BMP4.[169]
PRDM1 is involved in PGCs formation by repressing somatic cell program genes through selective recruitment of HDAC3.[170]
SALL4SALL4 participates in the specification of PGCs by suppressing the expression of somatic cell program genes.[171]
LIN28LIN28 is involved in PGCs development by regulating PRDM1 transcript translation.[172]
PRDM14PRDM14 establishes germ cell lineage by inducing SOX2 expression and cooperating with TFAP2C and PRDM1, which upregulates pluripotency genes and represses somatic markers. [173,174]
TT specifies germ cell fate by activating the expression of PRDM1 and PRDM14.[175]
AP2CAP2C is most likely downstream of PRDM1 and is involved in maintaining PGCs.[176]
SMAD1SMAD1 is a downstream signal mediator for BMPs and is essential for PGCs formation.[177]
SMAD5SMAD5 is a downstream signal mediator for BMPs and is required for PGCs development.[178]
SETDB1SETDB1 is involved in PGCs fate determination by ensuring BMP signaling through repressing the expression of Dppa2, Otx2, and Utf1.[179]
MTGR1MTGR1 is involved in stem cell maintenance and PGCs formation by mediating PRDM14 functions.[180]
ESRRBESRRB functions as an upstream factor of BMP4 and regulates PGCs development.[181]
PAX5PAX5 participates in PGCs specification by activating germline and repressing somatic program genes through a PAX5-OCT4-PRDM1 core transcriptional network.[182]
C2EIPC2EIP promotes the generation of PGCs by activating the Hedgehog (Hh) signaling pathway via PTCH2 ubiquitination.[183]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xie, Y.; Wu, C.; Li, Z.; Wu, Z.; Hong, L. Early Gonadal Development and Sex Determination in Mammal. Int. J. Mol. Sci. 2022, 23, 7500. https://doi.org/10.3390/ijms23147500

AMA Style

Xie Y, Wu C, Li Z, Wu Z, Hong L. Early Gonadal Development and Sex Determination in Mammal. International Journal of Molecular Sciences. 2022; 23(14):7500. https://doi.org/10.3390/ijms23147500

Chicago/Turabian Style

Xie, Yanshe, Changhua Wu, Zicong Li, Zhenfang Wu, and Linjun Hong. 2022. "Early Gonadal Development and Sex Determination in Mammal" International Journal of Molecular Sciences 23, no. 14: 7500. https://doi.org/10.3390/ijms23147500

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