Skip to main content

Follicle-stimulating hormone signaling in Sertoli cells: a licence to the early stages of spermatogenesis

Abstract

Follicle-stimulating hormone signaling is essential for the initiation and early stages of spermatogenesis. Follicle-stimulating hormone receptor is exclusively expressed in Sertoli cells. As the only type of somatic cell in the seminiferous tubule, Sertoli cells regulate spermatogenesis not only by controlling their own number and function but also through paracrine actions to nourish germ cells surrounded by Sertoli cells. After follicle-stimulating hormone binds to its receptor and activates the follicle-stimulating hormone signaling pathway, follicle-stimulating hormone signaling will establish a normal Sertoli cell number and promote their differentiation. Spermatogonia pool maintenance, spermatogonia differentiation and their entry into meiosis are also positively regulated by follicle-stimulating hormone signaling. In addition, follicle-stimulating hormone signaling regulates germ cell survival and limits their apoptosis. Our review summarizes the aforementioned functions of follicle-stimulating hormone signaling in Sertoli cells. We also describe the clinical potential of follicle-stimulating hormone treatment in male patients with infertility. Furthermore, our review may be helpful for developing better therapies for treating patients with dysfunctional follicle-stimulating hormone signaling in Sertoli cells. 

Background

Spermatogenesis is a process that is under complex regulation to achieve successive germ cell proliferation and differentiation [1]. Starting from spermatogonia stem cells (SSCs) producing differentiated spermatogonia, differentiated spermatogonia transform into spermatocytes. Spermatocytes undergo meiosis to produce round spermatids in which the chromosome number is reduced from diploid to haploid [2]. Round spermatids then undergo transformation to form the final spermatozoa which are released into the lumen. In mammals, only Sertoli cells and undifferentiated spermatogonia are detected during prepubertal and juvenile periods while spermatogenesis is initiated at puberty when undifferentiated spermatogonia begin to differentiate and enter meiosis [3,4,5].

As the only type of somatic cell in seminiferous tubule, Sertoli cell (SC) functions as a ‘nurse’ to care for spermatogenesis via paracrine actions to provide necessary nutrition and factors as well as forming necessary structures such as the blood-testis barrier (BTB) and Sertoli cell-Germ cell adhesion complex [6,7,8,9]. In higher vertebrates, spermatogenesis requires hormonal regulation by the hypothalamic-pituitary gonadal axis [10, 11]. Gonadotropin-releasing hormone (GnRH) is synthesized in the hypothalamus and is released into the pituitary gland where it stimulates the secretion of two gonadotropins, follicle-stimulating hormone (FSH) and luteinizing hormone (LH) [12]. FSH and LH then enter the circulation system to act on testis. Only undifferentiated spermatogonia and Sertoli cells are present in seminiferous tubule in the absence of FSH and LH [13]. Hormonal regulation of spermatogenesis is important and is mediated indirectly by SCs. Here, we will focus on the function of FSH signaling in spermatogenesis.

FSH is a glycoprotein that plays an essential role in prepubertal preparation for spermatogenesis and pubertal spermatogenesis regulation [14]. Its receptor FSH receptor (FSHR) is exclusively expressed on the cellular membrane of Sertoli cells [15]. In early life of both primates and rodents, physiological role of FSH signaling in spermatogenesis is to stimulate the transcription of genes related to DNA replication and cell cycle progression [16, 17]. Decades of study using the hypogonadal (hpg) model [18], FSHβ subunit knockout model [19], GnRH-immunized model [20] and FSHR knockout model [21,22,23] have revealed pivotal roles for FSH in regulating Sertoli cell function, increasing spermatogonia number, promoting entry into meiosis and limiting overall germ cell apoptosis. Since adult FSHR knockout mice are fertile but exhibit a reduced sperm output and completion of meiosis mainly depends on testosterone action, FSH is suggested to play a dominant role in establishing the most important parameter for testicular development and spermatogenesis prior to puberty in rodents [21, 24,25,26,27]. In men, FSH is essential to maintain fertility. Subfertility with quantitatively reduced spermatogenesis will occur in the absence of FSHR function [28] while a mutation in FSHβ subunit leads to azoospermia and infertility [29]. Although studies have provided a better understanding of the spermatogenesis processes that are regulated by FSH signaling, few molecules participating in these regulating activities have been precisely identified. We are also unable to elucidate the exact role of FSH in human spermatogenesis [30]. As a result, a review of the present work about FSH signaling in SCs is necessary and suggestions for future studies should be proposed. Moreover, FSH treatment has the potential to improve sperm number and motility in patients with hypogonadotropic hypogonadism or normogonadotropic patients with idiopathic impairment of spermatogenesis, highlighting the importance of obtaining a better understanding of FSH signaling in humans [31,32,33,34]. All of our hard work aims to achieve the ‘bench to bedside’ translation and cure more patients with FSH signaling dysfunction.

We surveyed articles in PubMed using the following keywords: ‘Sertoli cell’, ‘FSH’, ‘spermatogenesis’, ‘Sertoli cell proliferation’, ‘Sertoli cell differentiation’, ‘Spermatogonia stem cell self-renewal’, ‘meiosis’, ‘Spermatogonia proliferation’, ‘apoptosis’, ‘hypogonadotropic hypogonadism’, ‘normogonadotropic’, ‘FSH treatment’. We will present this review at the cellular and molecular levels, covering four parts: 1) Sertoli cell proliferation, differentiation and apoptosis; 2) Spermatogonia pool maintenance, differentiation and spermatogonia survival; 3) Entry into meiosis and spermatocyte survival; 4) Potential use of follicle-stimulating hormone in treating male infertility. Experimental species include rats, mice, zebrafish, sheep, bovines, goats, newts, trout and men. Our review is focused on the function of FSH signaling in SCs during the early stages of spermatogenesis.

Overview of follicle-stimulating hormone signaling in Sertoli cell

FSH is a glycoprotein composed of ɑ and β subunits. FSHɑ is a subunit shared with other glycoproteins, while FSHβ subunit is unique to FSH [35]. FSH exerts its function through the interaction between FSHβ and FSHR [14, 36]. According to a recent analysis of crystal structure, FSHR, which is a heterotrimeric guanine nucleotide–binding proteins (G proteins)-coupled receptor, is composed of a hormone binding domain, hinge region, hairpin loop, seven-transmembrane α helical domains and an intracellular domain [37,38,39]. The binding of FSH to the FSHR hormone-binding domain leads to a conformational change in FSHR, which facilitates the interaction between residues of FSHɑ and FSHβ subunit with the residues of the hinge region of FSHR. This interaction will further alter the conformation of seven-transmembrane α helical domains, resulting in the transmission of the signal to the intracellular domain, where coupling to effectors, recruiting adaptor proteins and transmitting FSH signaling downstream happen [39,40,41]. For more crystal structures, please see a review [39].

FSHR is present in the testis before a significant concentration of hormone appears in the foetal circulation [42]. In both rodents and primates, FSHR expression begins in the second half of gestation [43]. The interaction between FSH and FSHR is important for the function of FSH signaling. The variation of FSH/FSHR interaction at different age depends on the amount of SCs expressing FSHR with respect to those not expressing FSHR. In mice, FSH binding peaks between Days 7 and 21 but decreases significantly between Days 20 and 37 [44]. In rats, the Fshr mRNA level increases until Day 7, remains constant for 10 days and then decreases sharply on Day 40 [14]. The initial increase correlates with the proliferation of SCs and the increase of FSHR density per SC. While the sharp decrease correlates with the wide appearance of spermatocytes and spermatids so that the ratio of SCs to germ cells drops per seminiferous tubule. During one cycle of spermatogenesis in rats, FSH binding and Fshr mRNA level peak in stages XIII, XIV and I during the early development of germ cells but reach their lowest levels in stages VII and VIII when germ cells have developed to mature state [45,46,47]. Based on these observations, it can be concluded that FSH signaling mainly participates in the initiation and early development phases of spermatogenesis.

To date, at least five FSH signaling pathways have been identified in SCs: cyclic adenosine monophosphate (cAMP)/protein kinase A (PKA) pathway, extracellular-regulated kinase (ERK)/mitogen-activated protein kinase (MAPK) pathway, phosphoinositide 3-kinase (PI3K) pathway, calcium pathway and phospholipase A2 pathway. Here we mainly review the first three types of these five types pathways as well as the newly found retinoid acid pathway. Only these four types of pathways are included in Fig. 1. The cAMP/ PKA signaling pathway was the first to be identified. Upon FSH binding to FSHR on the plasma membrane of SCs, FSHR couples to the Gɑs subunit to activate adenylate cyclase (AC). Activated AC recruits ATP and transforms it into cAMP [48]. cAMP then binds to the regulatory subunits of PKA to release catalytic subunits of PKA [49, 50]. Catalytic subunits translocate into the nucleus and phosphorylate cyclic AMP response-element binding protein (CREB) at Ser133 or some cAMP-responsive elemental modulators [14]. These factors bind to the cAMP-response element of target genes to regulate their transcriptional activity during spermatogenesis [51]. Additionally, FSH activates ERK/MAPK pathway by coupling to both the Gɑi and Gɑs subunits in vitro [52]. The interaction of FSHR with the Gɑs subunit leads to ERK activation via a cAMP/PKA dependent pathway while the exact pathway mediated by FSHR coupling to Gɑi remains to be determined. Moreover, coupling of FSHR to Gβγ also activates PI3K [53, 54]. Activated PI3K triggers the transition from phosphatidylinositol bisphosphate (PIP2) to phosphatidylinositol 3,4,5-trisphosphate (PIP3) [55]. Accumulation of PIP3 leads to the phosphorylation of protein kinase B (Akt) and mammalian/mechanistic target of rapamycin (mTOR) [53, 56,57,58,59]. Activated mTOR phosphorylates 70-kDa ribosomal S6 kinase (p70S6K) to promote protein translation and gene expression [53, 60, 61]. Beta-arrestin promotes the internalization of FSHR to sustain the prolonged activation of signaling. This internalization is mediated by clathrin proteins [31, 62]. Interestingly, the levels and biochemical characteristics of signaling messengers are stage-specific upon FSH stimulation. The FSH-induced cAMP production level increases from birth to puberty while the FSH-induced PIP3 production level decreases from birth to puberty. Additionally, the p70S6K phosphorylation sites differ between different developmental periods. p70S6K is phosphorylated at T389, T421 and S424 via the cAMP/PKA pathway and PI3K/Akt pathway during proliferating stage but is only phosphorylated at T389 by the cAMP/PKA pathway during differentiating state [53]. The signaling pathway adopted is also stage-specific. For example, FSH mediated ERK activation in vitro was only detected in 5 and 11 day old rats, not in 19 day-old rats. Recently, FSH signaling was linked to retinoic acid (RA) signaling in SCs. FSH stimulates RA synthesis during the postnatal period via cAMP-dependent upregulation of retinol dehydrogenase 10 (RDH10) and aldehyde dehydrogenase 1A1 (ALDH1A1) [63,64,65]. During the pubertal period, FSH facilitates the translocation of retinoic acid receptor ɑ (RARɑ) into the nucleus. With the help of cytoplasmic RA-binding protein 2 (CRABP2), RA interacts with the RAR/retinoid X receptor (RXR) heterodimer and binds to RA response elements (RAREs) to regulate gene transcription [66, 67]. Other signaling pathways, including the calcium pathway and phospholipase A2 pathway have been reviewed by other researchers (For reviews, please see [68,69,70]).

Fig. 1
figure 1

FSH signaling pathway in SCs. After FSH binds to FSHR on the membrane of SCs, FSH signaling is activated. FSHR recruits different types of G proteins to mediate different signaling pathways. Recruitment of Gβγ subunits activates the PI3K/Akt/mTOR signaling pathway, promoting the translation of mRNAs. Recruitment of the Gɑ subunit activates the cAMP/PKA signaling pathway. Activated PKA directly phosphorylates the CREB protein in the nucleus. Phosphorylated CREB binds to CREs of target genes to regulate transcriptional activity. In addition, PKA activates ERK during Sertoli cell proliferation. ERK activation is also mediated by recruiting the Gɑi subunit. Recently, FSH-induced RA signaling was reported. FSH promotes RA biosynthesis through a cAMP-dependent pathway. RA translocates into the nucleus and binds to RAR/RXR to regulate target gene transcription. The cAMP/PKA signaling pathway participates in Sertoli cell differentiation, SSC self-renewal and differentiation, spermatogonia proliferation and their entry into meiosis, as well as BTB dynamics. The cAMP/PKA/ERK signaling pathway and PI3K/Akt/mTOR signaling pathway induce Sertoli cell proliferation. The cAMP/RA signaling pathway has been shown to participate in SSC differentiation, spermatogonia differentiation and their entry into meiosis. TF: transcription factor 

Follicle-stimulating hormone signaling in Sertoli cell regulates early stages of spermatogenesis

Roles of FSH mediated signaling in Sertoli cell proliferation, differentiation and apoptosis

SCs create a microenvironment and provide necessary nutrition for germ cells to complete spermatogenesis. The final number of SCs in adulthood is determined by the proliferation activity during the prepubertal period. SC differentiation during puberty endows SCs with their functions in spermatogenesis. Additionally, SC apoptosis maintains a healthy SC pool. All three processes are regulated by FSH signaling (Fig. 2).

Fig. 2
figure 2

Functions of FSH signaling in SC proliferation, differentiation and apoptosis. Binding of FSH to FSHR recruits Gɑs to FSHR, activating AC. AC transforms ATP into cAMP, and cAMP binds to regulatory subunits of PKA to release catalytic subunits of PKA. Activated PKA induces the expression of c-Myc and Cyclin D1 to promote SC proliferation (red). c-Myc expression is also upregulated by the PI3K/Akt/mTOR pathway. The cAMP/PKA/ERK pathway leads to the expression of Cyclin D1 only during SC proliferation. Through an unknown pathway, FSH induces the expression of HIF2 to promote c-Myc and Cyclin D1 expression. Moreover, inter-SC junction dynamics are also mediated by FSH signaling (green). FSH induces tPA expression via a cAMP/PKA-dependent pathway and three components of the BTB, N-cadherin, ɑ6β1-integrin and claudin 11, through an unknown pathway. Combined with testosterone, FSH stimulates Gja expression via the Wnt3 pathway. Additionally, FSH participates in SC differentiation (blue). Klf4, NF-κB and ASNS expression are mediated by the cAMP/PKA signaling pathway. HIF1 and Aqp8 expression are mediated by an unknown pathway. HIF1 mainly participates in the glycolytic pathway, and Aqp8 is important for water balance. Finally, the cAMP/PKA-mediated balance of FAAH and DAX1 maintains normal SC apoptosis (orange). The dotted line represents an unknown mechanism. For more information, please see the main text 

Sertoli cell proliferation

The final number of SCs determines the quality and quantity of spermatogenesis. SC proliferation occurs in the foetal or neonatal period and in the peripubertal period in all species [16]. Decades of studies using decapitation model, FSH antagonist model, transgenic model and FSHR knockdown model have revealed that FSH signaling in SCs is essential for SC proliferation.

FSH regulates SC proliferation only during foetal and early postnatal life. Pioneering works using the [3H] -thymidine assay, which indicates mitosis activation, found that low levels of endogenous FSH decrease [3H]-thymidine incorporation and the final SC number both in vivo and in vitro [71,72,73]. Injection of human FSH into immature rats with FSH withdrawal restored the mitotic activation and the final SC number [74,75,76,77]. In addition, treating hpg mice with recombinant FSH or expressing an FSH transgene in hpg mice also counteracted the negative effect of FSH deprivation on SC proliferation. FSHR mutation also decreased the SC number in mice, supporting these results [21, 22, 78].

Molecular mechanism underlying this stimulatory effect has been elucidated (Table 1). The main pathway that is included in this period is the PI3K/Akt signaling pathway [69, 79]. PI3K/Akt pathway phosphorylates p70S6K at T389, T421 and S424 [53]. Furthermore, Riera et al. reported that FSH also regulates SC proliferation via the PI3K/Akt/mammalian target of rapamycin complex 1(mTORC1) pathway. Supporting these results, phosphorylated Akt, phosphorylated proline-rich Akt substrate of 40 kDa (PRAS40), phosphorylated mTOR and phosphorylated p70S6K were detected after FSH stimulation in vitro [80]. Moreover, Crépieux et al. showed that FSH supports cAMP/PKA dependent extracellular signal-regulated protein kinases 1 and 2 (ERK1/2) activation and subsequent activation of the MAPK cascade in vitro [52].

Table 1 Factors that are under FSH signaling regulation during early stages of spermatogenesis 

Furthermore, genes regulated by FSH signaling that promote SC proliferation have been identified. Most genes are related to DNA replication, the cell cycle, cytoskeletal rearrangement and stem cell factors. Among them, cell-derived Myc (c-Myc) and type D1 cyclin (Cyclin D1) have been linked to the FSH signaling pathway. The proto-oncogene c-myc encodes the transcription factor c-Myc, which is important for promoting cell growth and maintaining vitality [81]. In prepubertal rats, the expression of c-myc mRNA was elevated by FSH stimulation via a cAMP-dependent pathway [82]. Further study using rat SCs found that PI3K/Akt/mTOR signaling participate in FSH stimulation of c-myc expression [80]. Cyclin D1, a member of cyclin, binds to cyclin-dependent kinase 4 and 6 to form a complex that promotes cell cycle progression from G1 to S phase [83]. By activating the cAMP-dependent ERK pathway in rat SCs, FSH stimulates Cyclin D1 expression in neonatal rat testes to promote SC proliferation [52]. Hypoxia inducible factor (HIF) is a transcription factor that regulates cell metabolism [84,85,86]. HIF1 regulates the expression of genes in the glycolytic pathway, while HIF2 regulates the expression of genes related to cell cycle progression [129,130,131,132]. During rat SC proliferation, FSH only upregulates HIF2 expression to increase c-Myc and Cyclin D1 expression both in vivo and in vitro through an unknown pathway [130]. Other genes that are regulated by FSH include hairy/enhancer of split gene 1, max-interacting protein repressor and Nur-related protein 1 in murine SCs [133]. Moreover, FSH signaling also cross-talks with insulin growth factor signaling to promote mouse SC proliferation. It is reported that FSH amplifies insulin growth factor signaling mediated Akt phosphorylation [134]. Interestingly, in female mice, FSH can stimulate granulosa cells proliferation via inducing Octamer-binding transcription factor 4 (OCT4) expression [135]. OCT4 is also found to be expressed in human SCs [136]. Whether FSH signaling can promote human SCs proliferation via OCT4 is proposed to be investigated. The precise signaling pathway regulating target gene expression after FSH binds to FSHR remains to be determined.

Sertoli cell differentiation

SC differentiation begins after SC proliferation cessation during puberty in all species [16]. During SC differentiation, SCs form the BTB to separate the adluminal area and basalarea. Also, SCs undergo metabolism to provide nutrition for germ cells between them. FSH is maintained at a relatively high level during this stage and promotes SC differentiation via an absolutely different signaling pathway compared with SC proliferation [1].

The main pathway by which FSH regulates SC differentiation is the cAMP/PKA signaling pathway. Although debates exist regarding whether FSH promotes SC differentiation, some evidences support our hypothesis. Firstly, FSH deprivation or FSHR knockout in mature mouse SCs led to low sperm counts and the SC transition from differentiation to proliferation [137]. Secondly, FSH activates ERK in immature rat SCs but inhibits its activation in mature SCs via cAMP/PKA signaling [52]. Thirdly, p70S6K is only phosphorylated at T389 in mature rat SCs, while p70S6K is phosphorylated at T389, T421 and S424 in proliferating cells [53]. Fourthly, cAMP and stimulatory Ga production in pubertal rat SCs are greater than those in neonatal rat SCs and FSH mediated cAMP signaling increases stimulatory Ga production in pubertal rat SCs [138]. Further support is obtained from evidence that FSH inhibits Yes-associated protein (YAP) expression to inhibit the Ste20-like protein kinase Hippo (Hippo) signaling pathway in pubertal rat SCs [139]. Hippo signaling pathway is known to promote cell proliferation [140]. Additionally, an increase in FSH level during puberty promotes RARa to translocate into the nucleus, which is important for SC differentiation [67]. In summary, FSH mainly regulates the cAMP/PKA signaling pathway to promote SC differentiation.

FSH regulates SC differentiation directly and indirectly via targeting direct functional factors and transcription factors respectively (Table 1). First class is transcription factors. Krüppel-like factor 4 (Klf4) is a pleiotropic zinc finger transcription factor that induces the expression of genes involved in SC differentiation. Klf4 expression is induced via cAMP/PKA signaling pathway in the TM4 Sertoli cell line [87, 88]. A recent in vivo study using mice demonstrates that FSH is able to induce expression of Klf4 via suppressing microRNA-92a-3p [141]. Nuclear factor (NF)-κB, a transcription factor that induces expression of genes related to SC function such as androgen binding protein, androgen receptor, is activated during SC differentiation following FSH stimulation in rats [89]. In rat mature SCs, both HIF1 and HIF2 expression are induced under FSH regulation. HIF1 increases glucose transporter 1 (Glut1) mRNA level to augment glucose uptake while HIF2 promotes the expression of tight junction protein ZO-1, ZO-2 and Occludin levels to establish the BTB [90]. Also, in vitro study using rat SCs indicated that FSH inhibits the expression of cell-derived jun proto-oncogene (c-jun) and increases jun-B mRNA level to regulate the transcription factor activator protein-1 (AP-1). AP-1 participates in the transcription response to hormones and growth factors which are necessary for SC differentiation [91].

In addition to transcription factors, direct functional factors involved in structural establishment, biochemical reactions and cell morphology were identified. Asp synthetase (ASNS), which promotes Asp accumulation in SCs, is regulated by FSH to induce its transcription in rat SCs. FSH activates the cAMP/PKA signaling pathway to regulate ASNS expression and its function in SC metabolism [94]. Besides Asp accumulation, FSH also positively regulates lactate production via glycolysis process in SCs [142]. Through interaction with PI3K, FSH promotes translocation of Glut1 to plasma membrane to absorb more glucose in rat SCs [143]. Also, FSH induces the transcription level of bifunctional enzyme 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase (PFKFB) isoform 1 (PFKFB1) and 3 (PFKFB3) in rat SCs to regulate synthesis and degradation of fructose 2,6-biphosphate [101]. What’s more, FSH can inhibit the transition of pyruvate to acetyl-coA through increasing the expression of pyruvate dehydrogenasekinase 3 (PDK3) in rat SCs [101]. All the efforts are to produce more lactate to nourish germ cells. Aquaporin 8 (Aqp8), which is involved in the water balance of rat and mouse SCs, is also stimulated by FSH [95, 96]. FSH is shown to be permissive for the formation of BTB and Sertoli cell – Germ cell junction such as ectoplasmic specialization and adherent junctions [133, 144]. Neural-cadherin (N- cadherin) and alpha6beta1-integrin (a6β1 integrin), two molecules known to make up of ectoplasmic specialization, are upregulated during the rat Sertoli cell differentiation process. FSH participates in this promotion in vitro, along with claudin-11, which belongs to tight junction protein [98,99,100]. Combination of FSH and testosterone also regulates the expression of gap junction protein Gja via wingless-type MMTV integration site family, member 3 (Wnt3) pathway in mice [97]. Gap junctions are pivotal for germ cell development [145]. Tissue plasminogen activator (tPA) was found to be induced by cAMP/PKA signaling pathway in rat and bovine SCs [92]. As a protease, tPA degrades tight junction proteins to regulate BTB dynamics [93]. Further studies could focus on the linkage between the transcription factor and molecules directly related to SC differentiation, which is helpful to elucidate the complete signaling network downstream FSHR.

Sertoli cell apoptosis

Apoptosis is important for maintaining a healthy microenvironment and cell number. SC apoptosis is also regulated by FSH, through the cAMP/PKA signaling pathway. FSH activates the cAMP/PKA signaling pathway to stimulate the expression of N-arachidonoylethanolamine hydrolase (FAAH) and inhibits SC apoptosis that caused by N-arachidonoylethanolamine (AEA) in mice [102]. AEA initiates apoptosis by inducing DNA fragmentation [146]. Additionally, activation of the cAMP/PKA signaling pathway downregulates nuclear receptor subfamily 0 Group B member 1 (DAX1) in maturing rat SCs [147]. Downregulated DAX1 is associated with a higher number of apoptotic cells [103]. The mechanism by which FSH achieves a balance between apoptosis and survival requires further investigation, but is probably mediated by regulating different transcription factors.

In summary, FSH promotes SC proliferation via the cAMP/PKA/ERK and PI3K/Akt/mTORC1 pathways while regulating SC differentiation and apoptosis via the cAMP/PKA pathway. Through sequential and corporative regulation of these three processes, FSH provides a healthy and functional microenvironment for spermatogenesis. However, the pathways that are activated in different processes and molecule function vary between different experiments. One explanation is that no precise boundaries exist between different stages. At the time of SC differentiation, SC proliferation can also occur. Besides, different species have different developmental stages. The two stages may overlap in some species such as rats [16]. Different cell culture conditions might also explain the difference between the results. In addition, it is suggested to consider whether the autocrine action is involved in SC proliferation and differentiation, as well as whether the autocrine action is regulated by FSH.

Role of FSH mediated signaling in Spermatogonia pool maintenance, differentiation and survival

Residing in the basement area of the seminiferous tubule, spermatogonia stem cells function as the original sources for the final spermatozoa [148]. In rodents, undifferentiated type A spermatogonia are classified as SSCs and subsequent progenitors [149]. Differentiating spermatogonia are classified into A1, A2, A3, A4, intermediate and type B spermatogonia [150, 151]. In humans, undifferentiated spermatogonia are categorized into Apale and Adark spermatogonia. Type B spermatogonia are the differentiating cells [152]. Decades of studies have provided insights into the function of FSH signaling in the spermatogonia pool (Fig. 3).

Fig. 3
figure 3

Functions of FSH signaling in early stages of spermatogenesis. Activating cAMP/PKA signaling pathway induces expression of GDNF and FGF2 in SC which will be secreted to SSC to maintain spermatogonia pool. Notably, FSH can inhibit degradation of GDNF via suppressing lysosomal biogenesis. This process is validated by activating PI3K/Akt/mTOR pathway. Additionally, FSH negatively regulates PGE2 through inhibiting AMH. PGE2 is also important for spermatogonia pool maintenance. SLF, SCF, Igf3 and BMP4 that is activated via cAMP/PKA pathway, along with transferrin through unknown pathway, contribute to spermatogonia differentiation. FSH ensures entry into meiosis through regulating DNA synthesis and chromosome dynamics. FSH regulates DNA synthesis via temporally inhibiting IL-6, Activin A and inducing Inhibin B. Chromosome dynamics is positively regulated by nociception via cAMP/PKA signaling pathway. Also, FSH upregulates Nrg1 and Nrg3 to induce Stra8 expression in spermatocytes. Besides, limiting gem cell apoptosis by inhibiting Gal3 and Bok is another task of FSH signaling in SCs. Dotted line represents unknown mechanism

Spermatogonia pool maintenance

Maintaining the spermatogonia pool ensures normal spermatogonia stem cell self-renewal and proliferation of undifferentiated progenitors. FSH has been shown to positively regulate spermatogonia pool maintenance in vivo and in vitro [153,154,155]. Impaired FSH signaling in immature SCs or mature SCs decreases the colonization of SSCs [156].

Molecular mechanism behind FSH regulation has been elucidated (Table 1). Glial cell line-derived neurotrophic factor (GDNF) and fibroblast growth factor 2 (FGF2) are two factors that are secreted by SCs and positively regulate SSC self-renewal and undifferentiated spermatogonia proliferation [157]. Among them, GDNF activates Akt and MAP kinse-ERK kinase (MEK) signaling pathway, resulting in the production of reactive oxygen species in SSCs. Reactive oxygen species stimulate SSCs self-renewal via the 38 kDa protein (p38) pathway [104, 105]. FSH inhibits autophagy of GDNF in goat SCs via activating the PI3K/Akt/mTOR pathway, which inhibits the translocation of transcription factor EB (TFEB) into the nucleus. Otherwise TFEB induces the expression of lysosomal biogenesis-related genes to degrade GDNF in goat SCs [158]. Recent study reported that GDNF receptor, GDNF family receptor α1, which is expressed by undifferentiated spermatogonia, is also positively regulated by FSH signaling in prepubertal trout testis, though the mechanism is unknown [159]. FGF2 is considered to be a bifunctional factor. For one thing, FGF2 promotes SSC self-renewal along with GDNF [106]. For another thing, FGF2 creates a more suitable environment for SSC differentiation by suppressing GDNF and cytochrome P450 family 26 subfamily B member 1 (Cyp26b1) expression [107]. Possible explanation may be that the combination of GDNF and FGF2 prepares the environment for the formation of progenitors that are ready to differentiate under FSH stimulation. In vivo and in vitro studies using bovine testis and rat testis demonstrates that FSH activates a cAMP-dependent signaling pathway to increase the mRNA levels and protein levels of these two factors [160, 161]. In zebrafish, FSH negatively regulates prostaglandin E2 (PGE2) in SCs by inhibiting Anti Mullerian Hormone (AMH). Otherwise PGE2 promotes SSC self-renewal and inhibits SSC differentiation [108].

Spermatogonia differentiation

Undifferentiated type A spermatogonia are under regulation of signaling network to differentiate into differentiated type B spermatogonia and then preleptotene spermatocytes [162]. FSH seems to initiate type A spermatogonia differentiation and induce differentiating spermatogonia proliferation.

Molecular mechanism behind FSH regulation has been elucidated (Table 1). FSH activates RA signaling by increasing RDH10, ALDH1A1, CRABP2 levels in primate SCs and this will provide an environment for induction of spermatogonia differentiation [1]. Stem cell factor (SCF) and steel factor (SLF) are two factors secreted by Sertoli cells during postnatal stages that are essential for the expansion of differentiating spermatogonia [111, 160]. FSH signaling induces transcription of SCF and SLF in prepubertal mouse testis via the cAMP-dependent signaling pathway [112]. The same phenomenon was also observed in adult rat testes [111, 160]. Both SCF and SLF are v-Kit Hardy-Zuckerman 4 Feline Sarcoma Viral Oncogene Homolog (kit) ligands (pleiotropic growth factor) and bind to kit (kit ligands receptor, CD117 is the cluster number for KIT receptor tyrosine kinase) on the surface of differentiating spermatogonia [163]. Owning to alternative splicing, kit ligand has transmembrane form or soluble form at different developmental stages and soluble form is favored for SSC differentiation [112]. Supporting this, transmembrane form of kit ligand is detected immunohistochemically in stages VII-VIII of the mouse seminiferous epithelium, during which SCs are less-sensitive to FSH signaling in mice [111, 164, 165]. Bone morphogenetic protein-4 (BMP4), secreted by SC during early postnatal stage, is proposed to promote SSC differentiation after binding to its receptor on spermatogonia [109, 110]. Its expression is under FSH/cAMP regulation. cAMP analogues downregulate BMP4 expression in prepubertal and pubertal mouse SCs while RA upregulates BMP4 expression level in prepubertal and pubertal mouse SCs [166, 167]. In zebrafish, FSH activates insulin growth factor 3 (Igf3) production via cAMP/PKA pathway [113]. Igf3 promotes SSC differentiation via beta-catenin (β-catenin) pathway in SSCs [114]. Other factors that are regulated by FSH and promote SSC differentiation include transferrin and Doublesex (sex determination and differentiation gene) and mab-3 (sex determination and differentiation gene) related transcription factor [161]. Transferrin functions as an ion transport to provide necessary ions for differentiating spermatogonia [115, 116].

Spermatogonia survival

FSH has been shown to protect spermatogonia from apoptosis which is important for steadiness of spermatogonia pool. Bcl-2-related ovarian killer (Bok) is proapoptotic member of the Bcl-2 gene family. FSH downregulates Bok mRNA level in rat testes to inhibit apoptosis [117]. Supporting this result, FSH suppression in immature rat SCs activates the caspase 9 mediated intrinsic apoptotic pathway [118]. Activation of the intrinsic apoptotic pathway is partially attributed to the activation of Bcl-2 gene family [168].

To sum up, FSH signaling in SCs induces paracrine action to maximize the capacity of spermatogonia ecology by maintaining the undifferentiated spermatogonia pool, promoting spermatogonia differentiation and spermatogonia survival. How to achieve a balance between molecules promoting self-renewal and molecules promoting differentiation in response to FSH stimulation remains to be further investigated. The answer may reside in spermatogonia themselves because it was reported that germ cells can control the local balance of GDNF, BMP4 and kit ligand levels [161].

Entry into meiosis and spermatocyte survival

Transition of type B spermatogonia into spermatocytes facilitates meiosis while surviving spermatocytes are essential for quantitative spermatogenesis. FSH is shown to be indispensable for entry into meiosis and positively regulates spermatocyte survival (Fig. 3).

Entry into meiosis

In vitro study using coculture system containing SCs and spermatogonia showed that FSH initiates the differentiation of secondary spermatogonia into primary spermatocytes in newts [169]. Injection of hpg mice with exogeneous FSH or transgenic expression of FSH restores the number of spermatogonia and spermatocytes [170, 171]. This observation is supported by hypophysectomised or GnRH-immunized adult rat models which lack normal circulating FSH levels [26, 172]. Furthermore, knocking out of FSHR and FSHβ in mice resulted in decreased numbers of spermatogonia and spermatocytes, let alone spermatids [25, 173]. These results indicate that FSH is necessary for spermatogonia to differentiate into primary spermatocytes, promoting entry into meiosis.

Detailed mechanisms that FSH adopts to guarantee entry into meiosis are as follows (Table 1). Activin alpha (Activin A, growth and differentiation gene) and Inhibin beta (Inhibin B, growth and differentiation gene) are two structurally-related factors that belong to the transforming growth factor β family. Activin A promotes DNA synthesis in spermatocytes while Inhibin B inhibits this biological process [119]. Through the cAMP/PKA signaling pathway, FSH can activate the production of Inhibin B while inhibit the production of Activin A near the beginning of meiosis [120]. Thus, FSH functions as a monitor for the end of DNA synthesis. This result is further supported by a study of another factor interleukin 6 (IL-6). IL-6 is reported to negatively regulate DNA synthesis. IL-6 expression is downregulated by FSH via the cAMP dependent pathway during stages VII-VIII and upregulated by FSH via the PKC dependent pathway during stages IX-XI in rats [121]. Stages VII-VIII corelate with the initiation of meiotic DNA synthesis while stages IX-XI correlate with DNA synthesis termination. Recently, Eto et al. reported that FSH can promote nociception expression via cAMP/PKA signaling in murine Sertoli cells [122]. Nociceptin (17-residue neuropeptide) is secreted by Sertoli cells and binds to nociception receptor opioid related nociceptin receptor 1 (OPRL1) on the surface of spermatocytes [123]. Binding of nociception to its receptor leads to REC8 meiotic recombination protein (Rec8) phosphorylation in spermatocytes which promotes meiotic chromosome dynamics to prepare for the subsequent meiosis [174]. Similarly, FSH, combined with retinoic acid, stimulates Neuregulin 1 (Nrg1) and Neuregulin 3 (Nrg3) expression in mouse SCs [124, 125]. Nrg1 and Nrg3 are secreted from SCs and bind to their receptor EGFR – Mouse Genome Informatics 4 (ERBB4) on the surface of pre-spermatocytes which will trigger stimulated retinoic acid gene 8 (Stra8) expression [125]. Upregulated Stra8 expression promotes the early stage of meiotic prophase [175]. However, the exact signaling pathway that is adopted by FSH remains to be elucidated. In summary, these results demonstrates that FSH positively regulates entry into meiosis by temporally ensuring the initiation of DNA synthesis and termination of DNA synthesis, as well as monitoring meiotic chromosome dynamics.

Spermatocyte survival

FSH is also pivotal for spermatocyte survival. In FSH-suppressed adult rats and gonadotropin-suppressed adult men, the spermatocyte apoptosis rate showed a significant increase [176, 177].When the androgen level is normal, the suppression of FSH reduced pachytene spermatocytes numbers in rats [178]. In human SCs, Sá et al. found that the combination of FSH and testosterone maximally maintains spermatocytes because FSH alone was not enough to limit spermatocytes apoptosis during the second week of vero cell conditioned medium [179, 180]. Previous study demonstrated that spermatocyte apoptosis is related to both extrinsic (Caspase 8) and intrinsic (Caspase 9) apoptotic pathways [181, 182]. Supporting this finding, FSH signaling in rats has been shown to inhibit both the extrinsic and intrinsic apoptotic pathways during the first wave of spermatogenesis to promote spermatocyte survival [118].

Another factor, Galectin-3 (Gal-3) is reported to inhibit both the intrinsic and extrinsic apoptotic pathways by blocking cytochrome c release and Fas (death receptor)/Fas-ligand (member of the tumor necrosis factor family of death-inducing ligands)cross linking respectively [126]. FSH induces expression of Gal-3 in porcine and rat SCs at the initiation stage of meiosis and protects spermatocytes from apoptosis, probably via cAMP/PKA dependent pathway [127].

Additionally, FSH can inhibit early meiotic spermatocyte apoptosis via inhibition of transcription factor AP-1 in human SCs. Activation of AP-1 occurs before the activation of effector caspase such as caspase 3 [128]. Caspase 3 was shown to be expressed in human SCs and germ cells [183]. This indicates that FSH signaling in human SCs may control germ cell death via paracrine action.

In summary, FSH exerts its effect at the beginning of meiosis by promoting entry into meiosis and the survival of spermatocytes. This effect may be mainly due to the sufficient number of Sertoli cells and spermatogonia. Further studies are recommended to focus on the effect of FSH on the transition from spermatogonia to spermatocytes as well as whether FSH has effect on the transition from primary spermatocytes to the secondary spermatocytes.

Potential use of follicle-stimulating hormone in treating male infertility

In humans, FSH induces SC proliferation and spermatogonia proliferation and differentiation, while FSH alone is not essential to complete meiosis and spermiogenesis [184, 185]. Testosterone is more important from the beginning of spermatocyte development [186]. The differences in the functions of FSH signaling in spermatogenesis between humans and other experimental animals suggest that further studies should be conducted to understand the FSH regulation in human or we should develop more appropriate experimental animal models.

Currently, FSH treatment is mainly administrated to two types of patients: patients with hypogonadotropic hypogonadism (HH) and normogonadotropic patients with idiopathic impairment of spermatogenesis [187]. The findings described above indicate that FSH promotes the final sperm production by positively regulating Sertoli cell biology. As expected, questions about whether excess FSH is harmful for spermatogenesis arise before FSH treatment. In rodents, high serum FSH levels result in better testis development [78, 188]. Men with pituitary adenoma secreting excess FSH also show normal spermatogenesis and normal testicular development [189]. Supporting this result, enhanced receptor activity resulted from gain of function mutations in FSHR, such as FSHR-D567G and FSHR-N431I, also appears to have little effect on normal spermatogenesis [190, 191]. These results provide a theoretical support for FSH treatment.

In patients with hypogonadotropic hypogonadism, the lack of gonadotropin FSH stimulation or defects in gonadotropin-releasing hormone synthesis and secretion lead to azoospermia or severe oligozoospermia [187, 192]. One method to treat HH is pulsatile GnRH administration which may lead to the secretion of gonadotropin from the pituitary gland [29, 193]. The secreted FSH can stimulate Sertoli cell growth to support normal spermatogenesis. However, this method is costly and troublesome since external GnRH must be pumped subcutaneously [194]. Another method to treat HH is the administration of exogenous gonadotropins. This method, which involves treating patients with human chorionic gonadotropin alone or in combination with FSH, is more direct and may be more successful in most cases [195, 196]. Human chorionic gonadotropin (hCG) functions similarly to luteinizing hormone, but with different bioactivities [197]. hCG is observed to restore sperm production in men. This effect may be enhanced when hCG is administered in combination with FSH [198, 199]. However, the precise dosage and timing of FSH treatment in this method remains controversial. Though this method seems useful at present, limitations have also been noted. For example, exogenous gonadotropin administration is not the same as the gonadotropin secretion stimulated by GnRH. The intrinsic regulatory network cannot be simulated using this method.

FSH treatment appears to be beneficial for normogonadotropic patients with idiopathic impairment of spermatogenesis. Meta-analyses revealed that FSH therapy in these patients increases the rate of clinical pregnancies in female partners [200]. However, since the number of participants taking part in the experiment was relatively low, studies are still needed to determine whether FSH therapy truly affects normogonadotropic patients. As a result, the selection of appropriate normogonadotropic patients to receive FSH therapy is necessary. First, no identifiable and generally accepted cause for male infertility should be detected [201]. Second, FSH pharmacogenetics is promising in this evaluation. Testing for the single nucleic polymorphism (SNP) p.N680S in patients receiving FSH therapy is important. Male patients with the p.N680S homozygous N polymorphism exhibit a significantly decreased DNA fragmentation index of sperm in the ejaculate after FSH treatment [202,203,204]. Also, SNP of the FSHβ is another marker to select the normogonadotropic patients to receive FSH therapy [205, 206]. More well-organized and sufficient randomized studies are needed to determine whether FSH therapy is truly helpful for normogonadotropic patients as well as the dosage and timing needed to carry out therapy. In summary, Precision Medicine matters a lot!

Let’s return to the exogeneous gonadotropin administration method. How to expand the half-life of gonadotropin is important for treatment efficiency. One way is to conjugate the gonadotropin to polyethylene glycol (PEG). PEGylated FSH not only retains FSH activity but also results in improved bioavailability [207]. Another way is to develop a single-chain recombinant analogue of gonadotropin [208]. These molecules were engineered with the β-subunits oriented at the N-terminus of the α subunit and used the hCGβ carboxy-terminal peptide (CTP) sequence as a linker [209, 210]. These analogues have an increased serum half-life and increased biopotency. Using this method, we obtain dual FSH and LH analogues, such as FSHβ-CTP-LHβ-CTP-α [211]. Recently, fusion analogues of FSH consisting FSHα, FSHβ and immunoglobulin constant fragments were constructed [212]. This type of analogue can improve pharmacokinetics. All the aforementioned analogues have great beneficial to female infertility treatment and ovarian development of experimental animals such as sheep and monkeys. It remains to be determined if present analogues can treat male infertility. Analogs that can be used in clinical trials are being researched [31].

Recently, the relationship between diabetes mellitus and male infertility attracts attention [213,214,215]. As the metabolic modulator in seminiferous tubule, SC metabolism dysfunction is thought to be one link between diabetes mellitus and male infertility [216, 217]. In human, Glut1 and Glut3 transport glucose into SCs. With the help of lactate dehydrogenase (LDH), glucose can be converted into lactate which will be transported out of SCs and supplied to germ cells via monocarboxylate transporter (MCT) [218,219,220]. Diabetes mellitus patients displayed low level of FSH, low mRNA levels of Glut1, Glut3 and low protein level of LDH in SCs [221]. Insulin-deprived human SCs, which was similar to diabetes mellitus, presented decreased transcript level of LDH, MCT4, Glut3 and increased transcript level of Glut1 [222]. Moreover, decreased level of sirtuin 1 and increased level of ghrelin in diabetes mellitus patients impair the hypothalamus-pituitary–gonadal axis which leads to low level of FSH [86, 223, 224]. In Klinefelter syndrome male patients which is prone to suffer from diabetes mellitus, high level of FSH along with increased mRNA expression of Glut3 and decreased mRNA expression of Glut1 in SCs may be a try to rescue spermatogenesis [225]. Based on this, drugs can be developed to rescue FSH in diabetes mellitus patients so that normal SC metabolism can occur and sufficient energy can be provided to germ cells.

Hope still exists. Conversation in spermatogenic processes between human and mouse are revealed in previous studies [226]. Since the phenotypes of Fshr-knockout mice and men carrying Fshr mutations are similar, the Fshr-knockout mouse model still has great clinical potential [29]. The identification of additional genes that are regulated by FSH in mice and developing targeted medicines are feasible. Last, the combination of FSH and testosterone treatment is more efficient than a single hormone treatment, since testosterone can augment FSH signaling in SCs [227].

Conclusion

FSH signaling in SCs establishes the appropriate microenvironment for spermatogenesis. FSH signaling plays a dominant role in determining the number and function of SCs. FSH signaling also maintains the spermatogonia pool and induces spermatogonia differentiation through paracrine actions. In addition, FSH signaling promotes entry into meiosis and the survival of germ cells. However, few molecules involved in these paracrine actions have been found. We could detect the changes within protein and mRNA expression level of receptors on the surface of gem cells that are associated with different spermatogenic processes, and then determine whether the levels of their ligands changed after the administration of FSH signaling in SCs. Mass spectrometry and single-cell transcriptomics will be helpful. Moreover, combining the transgenic mouse model with human infertility is necessary to develop therapies for diseases related to dysfunctional FSH signaling. 

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Abbreviations

ɑ6β1 integrin:

Alpha6beta1-integrin

β-catenin:

Beta-catenin

AC:

Adenylate cyclase

Activin alpha:

Activin A

AEA:

N-arachidonoylethanolamine

Akt:

Protein kinase B

ALDH1A1:

Aldehyde dehydrogenase 1A1

AMH:

Anti Mullerian Hormone

AP-1:

Activator protein-1

Aqp8:

Aquaporin 8

ASNS:

Asp synthetase

BMP4:

Bone morphogenetic protein-4

Bok:

Bcl-2-related ovarian killer

BTB:

Blood-testis barrier

cAMP:

Cyclic adenosine monophosphate

c-Myc:

Cell-derived Myc

c-jun:

Cell-derived jun proto-oncogene

CRABP2:

Cytoplasmic RA-binding protein 2

CREB:

Cyclic AMP response-element binding protein

CTP:

Carboxy-terminal peptide

Cyclin D1:

Type D1 cyclin

Cyp26b1:

Cytochrome P450 family 26 subfamily B member 1

DAX1:

Nuclear receptor subfamily 0 group B member 1

ERBB4:

EGFR – Mouse Genome Informatics 4

ERK:

Extracellular-regulated kinase

ERK1/2:

Extracellular signal-regulated protein kinases 1 and 2

FAAH:

N-arachidonoylethanolamine hydrolase

FGF2:

Fibroblast growth factor 2

FSH:

Follicle-stimulating hormone

FSHR:

Follicle-stimulating hormone receptor

G proteins:

Heterotrimeric guanine nucleotide–binding proteins

Gal-3:

Galectin-3

GDNF:

Glial cell line-derived neurotrophic factor

GnRH:

Gonadotropin-releasing hormone

hCG:

Human chorionic gonadotropin

HH:

Hypogonadotropic hypogonadism

HIF:

Hypoxia inducible factor

hpg:

Hypogonadal

Hippo:

Ste20-like protein kinase Hippo

Igf3:

Insulin growth factor 3

IL-6:

Interleukin 6

Inhibin B:

Inhibin beta

Kit:

V-Kit Hardy-Zuckerman 4 Feline Sarcoma Viral Oncogene Homolog

Klf4:

Krüppel-like factor 4

LDH:

Lactate dehydrogenase

LH:

Luteinizing hormone

MAPK:

Mitogen-activated protein kinase

MCT:

Monocarboxylate transporter

MEK:

MAP kinase-ERK kinase

mTOR:

Mammalian/mechanistic target of rapamycin

mTORC1:

Mammalian target of rapamycin complex 1

N-cadherin:

Neural-cadherin

NF:

Nuclear factor

Nrg1:

Neuregulin 1

Nrg3:

Neuregulin 3

OCT4:

Octamer-binding transcription factor 4

p38:

38 KDa protein

OPRL1:

Opioid related nociceptin receptor 1

p70S6K:

Phosphorylated 70-kDa ribosomal S6 kinase

PDK3:

Pyruvate dehydrogenasekinase 3

PEG:

Polyethylene glycol

PFKFB:

6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase

PFKFB1:

6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase isoform 1

PFKFB3:

6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase isoform 3

PGE2:

Prostaglandin E2

PI3K:

Phosphoinositide 3-kinase

PIP2:

Phosphatidylinositol bisphosphate

PIP3:

Phosphatidylinositol 3,4,5-trisphosphate

PKA:

Protein kinase A

PRAS40:

Proline-rich Akt substrate of 40 kDa

RA:

Retinoic acid

RARE:

RA response element

RARÉ‘:

Retinoic acid receptor É‘

RDH10:

Retinol dehydrogenase 10

Rec8:

REC8 meiotic recombination protein

RXR:

Retinoid X receptor

SC:

Sertoli cell

SCF:

Stem cell factor

SLF:

Steel factor

SNP:

Single nucleic polymorphism

SSC:

Spermatogonia stem cell

Stra8:

Stimulated retinoic acid gene 8

TFEB:

Transcription factor EB

tPA:

Tissue plasminogen activator

Wnt3:

Wingless-type MMTV integration site family, member 3

YAP:

Yes-associated protein

References

  1. Khanehzad M, Abbaszadeh R, Holakuyee M, Modarressi MH, Nourashrafeddin SM. FSH regulates RA signaling to commit spermatogonia into differentiation pathway and meiosis. Reprod Biol Endocrinol. 2021;19:4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. de Kretser DM, Loveland KL, Meinhardt A, Simorangkir D, Wreford N. Spermatogenesis. Hum Reprod. 1998;13(Suppl 1):1–8.

    Article  PubMed  Google Scholar 

  3. Plant TM. Undifferentiated primate spermatogonia and their endocrine control. Trends Endocrinol Metab. 2010;21:488–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Feng CW, Bowles J, Koopman P. Control of mammalian germ cell entry into meiosis. Mol Cell Endocrinol. 2014;382:488–97.

    Article  CAS  PubMed  Google Scholar 

  5. Walker WH. Regulation of mammalian spermatogenesis by miRNAs. Semin Cell Dev Biol. 2022;121:24–31.

    Article  CAS  PubMed  Google Scholar 

  6. Ni FD, Hao SL, Yang WX. Molecular insights into hormone regulation via signaling pathways in Sertoli cells: With discussion on infertility and testicular tumor. Gene. 2020;753:144812.

    Article  CAS  PubMed  Google Scholar 

  7. Oliveira PF, Alves MG, Rato L, Silva J, Sá R, Barros A, Sousa M, Carvalho RA, Cavaco JE, Socorro S. Influence of 5α-dihydrotestosterone and 17β-estradiol on human Sertoli cells metabolism. Int J Androl. 2011;34:e612-620.

    Article  CAS  PubMed  Google Scholar 

  8. Walker WH. Non-classical actions of testosterone and spermatogenesis. Philos Trans R Soc Lond B Biol Sci. 2010;365:1557–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Crisóstomo L, Alves MG, Calamita G, Sousa M, Oliveira PF. Glycerol and testicular activity: the good, the bad and the ugly. Mol Hum Reprod. 2017;23:725–37.

    Article  PubMed  CAS  Google Scholar 

  10. Ilgin S. The adverse effects of psychotropic drugs as an endocrine disrupting chemicals on the hypothalamic-pituitary regulation in male. Life Sci. 2020;253:117704.

    Article  CAS  PubMed  Google Scholar 

  11. Kaprara A, Huhtaniemi IT. The hypothalamus-pituitary-gonad axis: Tales of mice and men. Metabolism. 2018;86:3–17.

    Article  CAS  PubMed  Google Scholar 

  12. Abreu AP, Kaiser UB. Pubertal development and regulation. Lancet Diabetes Endocrinol. 2016;4:254–64.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Plant TM, Marshall GR. The functional significance of FSH in spermatogenesis and the control of its secretion in male primates. Endocr Rev. 2001;22:764–86.

    Article  CAS  PubMed  Google Scholar 

  14. Simoni M, Gromoll J, Nieschlag E. The follicle-stimulating hormone receptor: biochemistry, molecular biology, physiology, and pathophysiology. Endocr Rev. 1997;18:739–73.

    CAS  PubMed  Google Scholar 

  15. Oduwole OO, Peltoketo H, Huhtaniemi IT. Role of Follicle-Stimulating Hormone in Spermatogenesis. Front Endocrinol (Lausanne). 2018;9:763.

    Article  Google Scholar 

  16. Sharpe RM, McKinnell C, Kivlin C, Fisher JS. Proliferation and functional maturation of Sertoli cells, and their relevance to disorders of testis function in adulthood. Reproduction. 2003;125:769–84.

    Article  CAS  PubMed  Google Scholar 

  17. Zimmermann C, Stévant I, Borel C, Conne B, Pitetti JL, Calvel P, Kaessmann H, Jégou B, Chalmel F, Nef S. Research resource: the dynamic transcriptional profile of sertoli cells during the progression of spermatogenesis. Mol Endocrinol. 2015;29:627–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. McLachlan RI, Wreford NG, Tsonis C, De Kretser DM, Robertson DM. Testosterone effects on spermatogenesis in the gonadotropin-releasing hormone-immunized rat. Biol Reprod. 1994;50:271–80.

    Article  CAS  PubMed  Google Scholar 

  19. Kumar TR, Wang Y, Lu N, Matzuk MM. Follicle stimulating hormone is required for ovarian follicle maturation but not male fertility. Nat Genet. 1997;15:201–4.

    Article  CAS  PubMed  Google Scholar 

  20. Lerchl A, Sotiriadou S, Behre HM, Pierce J, Weinbauer GF, Kliesch S, Nieschlag E. Restoration of spermatogenesis by follicle-stimulating hormone despite low intratesticular testosterone in photoinhibited hypogonadotropic Djungarian hamsters (Phodopus sungorus). Biol Reprod. 1993;49:1108–16.

    Article  CAS  PubMed  Google Scholar 

  21. Dierich A, Sairam MR, Monaco L, Fimia GM, Gansmuller A, LeMeur M, Sassone-Corsi P. Impairing follicle-stimulating hormone (FSH) signaling in vivo: targeted disruption of the FSH receptor leads to aberrant gametogenesis and hormonal imbalance. Proc Natl Acad Sci U S A. 1998;95:13612–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Abel MH, Wootton AN, Wilkins V, Huhtaniemi I, Knight PG, Charlton HM. The effect of a null mutation in the follicle-stimulating hormone receptor gene on mouse reproduction. Endocrinology. 2000;141:1795–803.

    Article  CAS  PubMed  Google Scholar 

  23. Chang C, Chen YT, Yeh SD, Xu Q, Wang RS, Guillou F, Lardy H, Yeh S. Infertility with defective spermatogenesis and hypotestosteronemia in male mice lacking the androgen receptor in Sertoli cells. Proc Natl Acad Sci U S A. 2004;101:6876–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. De Gendt K, Swinnen JV, Saunders PT, Schoonjans L, Dewerchin M, Devos A, Tan K, Atanassova N, Claessens F, Lécureuil C, et al. A Sertoli cell-selective knockout of the androgen receptor causes spermatogenic arrest in meiosis. Proc Natl Acad Sci U S A. 2004;101:1327–32.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Abel MH, Baker PJ, Charlton HM, Monteiro A, Verhoeven G, De Gendt K, Guillou F, O’Shaughnessy PJ. Spermatogenesis and sertoli cell activity in mice lacking sertoli cell receptors for follicle-stimulating hormone and androgen. Endocrinology. 2008;149:3279–85.

    Article  CAS  PubMed  Google Scholar 

  26. McLachlan RI, Wreford NG, de Kretser DM, Robertson DM. The effects of recombinant follicle-stimulating hormone on the restoration of spermatogenesis in the gonadotropin-releasing hormone-immunized adult rat. Endocrinology. 1995;136:4035–43.

    Article  CAS  PubMed  Google Scholar 

  27. Haywood M, Spaliviero J, Jimemez M, King NJ, Handelsman DJ, Allan CM. Sertoli and germ cell development in hypogonadal (hpg) mice expressing transgenic follicle-stimulating hormone alone or in combination with testosterone. Endocrinology. 2003;144:509–17.

    Article  CAS  PubMed  Google Scholar 

  28. Layman LC, Porto AL, Xie J, da Motta LA, da Motta LD, Weiser W, Sluss PM. FSH beta gene mutations in a female with partial breast development and a male sibling with normal puberty and azoospermia. J Clin Endocrinol Metab. 2002;87:3702–7.

    CAS  PubMed  Google Scholar 

  29. Zheng J, Mao J, Cui M, Liu Z, Wang X, Xiong S, Nie M, Wu X. Novel FSHβ mutation in a male patient with isolated FSH deficiency and infertility. Eur J Med Genet. 2017;60:335–9.

    Article  PubMed  Google Scholar 

  30. Murao K, Imachi H, Muraoka T, Fujiwara M, Kushida Y, Haba R, Ishida T. Isolated follicle-stimulating hormone (FSH) deficiency without mutation of the FSHbeta gene and successful treatment with human menopausal gonadotropin. Fertil Steril. 2012;2008(90):e2017-2019.

    Google Scholar 

  31. Santi D, Crépieux P, Reiter E, Spaggiari G, Brigante G, Casarini L, Rochira V, Simoni M. Follicle-stimulating Hormone (FSH) Action on Spermatogenesis: A Focus on Physiological and Therapeutic Roles. J Clin Med. 2020;9(4):1014.

    Article  CAS  PubMed Central  Google Scholar 

  32. Santi D, De Vincentis S, Alfano P, Balercia G, Calogero AE, Cargnelutti F, Coccia ME, Condorelli RA, Dal Lago A, de Angelis C, et al. Use of follicle-stimulating hormone for the male partner of idiopathic infertile couples in Italy: Results from a multicentre, observational, clinical practice survey. Andrology. 2020;8:637–44.

    Article  CAS  PubMed  Google Scholar 

  33. Luján S, Caroppo E, Niederberger C, Arce JC, Sadler-Riggleman I, Beck D, Nilsson E, Skinner MK. Sperm DNA Methylation Epimutation Biomarkers for Male Infertility and FSH Therapeutic Responsiveness. Sci Rep. 2019;9:16786.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Simoni M, Brigante G, Rochira V, Santi D, Casarini L. Prospects for FSH Treatment of Male Infertility. J Clin Endocrinol Metab. 2020;105(7):dgaa243.

    Article  PubMed  Google Scholar 

  35. Das N, Kumar TR. Molecular regulation of follicle-stimulating hormone synthesis, secretion and action. J Mol Endocrinol. 2018;60:R131-r155.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Szymańska K, Kałafut J, Przybyszewska A, Paziewska B, Adamczuk G, Kiełbus M, Rivero-Müller A. FSHR Trans-Activation and Oligomerization. Front Endocrinol (Lausanne). 2018;9:760.

    Article  Google Scholar 

  37. Thal DM, Vuckovic Z, Draper-Joyce CJ, Liang YL, Glukhova A, Christopoulos A, Sexton PM. Recent advances in the determination of G protein-coupled receptor structures. Curr Opin Struct Biol. 2018;51:28–34.

    Article  CAS  PubMed  Google Scholar 

  38. Navarro G, Cordomí A, Casadó-Anguera V, Moreno E, Cai NS, Cortés A, Canela EI, Dessauer CW, Casadó V, Pardo L, et al. Evidence for functional pre-coupled complexes of receptor heteromers and adenylyl cyclase. Nat Commun. 2018;9:1242.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Ulloa-Aguirre A, Zariñán T, Jardón-Valadez E, Gutiérrez-Sagal R, Dias JA. Structure-Function Relationships of the Follicle-Stimulating Hormone Receptor. Front Endocrinol (Lausanne). 2018;9:707.

    Article  Google Scholar 

  40. Jiang X, Liu H, Chen X, Chen PH, Fischer D, Sriraman V, Yu HN, Arkinstall S, He X. Structure of follicle-stimulating hormone in complex with the entire ectodomain of its receptor. Proc Natl Acad Sci U S A. 2012;109:12491–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Jiang X, Dias JA, He X. Structural biology of glycoprotein hormones and their receptors: insights to signaling. Mol Cell Endocrinol. 2014;382:424–51.

    Article  CAS  PubMed  Google Scholar 

  42. Tsutsui K, Shimizu A, Kawamoto K, Kawashima S. Developmental changes in the binding of follicle-stimulating hormone (FSH) to testicular preparations of mice and the effects of hypophysectomy and administration of FSH on the binding. Endocrinology. 1985;117:2534–43.

    Article  CAS  PubMed  Google Scholar 

  43. Warren DW, Huhtaniemi IT, Tapanainen J, Dufau ML, Catt KJ. Ontogeny of gonadotropin receptors in the fetal and neonatal rat testis. Endocrinology. 1984;114:470–6.

    Article  CAS  PubMed  Google Scholar 

  44. Ito I, Minegishi T, Hasegawa Y, Shinozaki H, Nakamura K, Igarashi S, Nakamura M, Miyamoto K, Ibuki Y. Developmental changes of testicular activin and FSH receptor mRNA and plasma FSH and inhibin levels in the rat. Life Sci. 1993;53:1299–307.

    Article  CAS  PubMed  Google Scholar 

  45. Heckert L, Griswold MD. Expression of the FSH receptor in the testis. Recent Prog Horm Res. 1993;48:61–77.

    Article  CAS  PubMed  Google Scholar 

  46. Heckert LL, Griswold MD. Expression of follicle-stimulating hormone receptor mRNA in rat testes and Sertoli cells. Mol Endocrinol. 1991;5:670–7.

    Article  CAS  PubMed  Google Scholar 

  47. Kangasniemi M, Kaipia A, Toppari J, Perheentupa A, Huhtaniemi I, Parvinen M. Cellular regulation of follicle-stimulating hormone (FSH) binding in rat seminiferous tubules. J Androl. 1990;11:336–43.

    CAS  PubMed  Google Scholar 

  48. Northup JK, Sternweis PC, Smigel MD, Schleifer LS, Ross EM, Gilman AG. Purification of the regulatory component of adenylate cyclase. Proc Natl Acad Sci U S A. 1980;77:6516–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Rangel-Aldao R, Rosen OM. Dissociation and reassociation of the phosphorylated and nonphosphorylated forms of adenosine 3’:5’ -monophosphate-dependent protein kinase from bovine cardiac muscle. J Biol Chem. 1976;251:3375–80.

    Article  CAS  PubMed  Google Scholar 

  50. Landmark BF, Fauske B, Eskild W, Skålhegg B, Lohmann SM, Hansson V, Jahnsen T, Beebe SJ. Identification, characterization, and hormonal regulation of 3’, 5’-cyclic adenosine monophosphate-dependent protein kinases in rat Sertoli cells. Endocrinology. 1991;129:2345–54.

    Article  CAS  PubMed  Google Scholar 

  51. de Kretser DM, Loveland KL, Meehan T, O’Bryan MK, Phillips DJ, Wreford NG. Inhibins, activins and follistatin: actions on the testis. Mol Cell Endocrinol. 2001;180:87–92.

    Article  PubMed  Google Scholar 

  52. Crépieux P, Marion S, Martinat N, Fafeur V, Vern YL, Kerboeuf D, Guillou F, Reiter E. The ERK-dependent signalling is stage-specifically modulated by FSH, during primary Sertoli cell maturation. Oncogene. 2001;20:4696–709.

    Article  PubMed  Google Scholar 

  53. Musnier A, Heitzler D, Boulo T, Tesseraud S, Durand G, Lécureuil C, Guillou H, Poupon A, Reiter E, Crépieux P. Developmental regulation of p70 S6 kinase by a G protein-coupled receptor dynamically modelized in primary cells. Cell Mol Life Sci. 2009;66:3487–503.

    Article  CAS  PubMed  Google Scholar 

  54. Dupont J, Musnier A, Decourtye J, Boulo T, Lécureuil C, Guillou H, Valet S, Fouchécourt S, Pitetti JL, Nef S, et al. FSH-stimulated PTEN activity accounts for the lack of FSH mitogenic effect in prepubertal rat Sertoli cells. Mol Cell Endocrinol. 2010;315:271–6.

    Article  CAS  PubMed  Google Scholar 

  55. Xia P, Xu XY. PI3K/Akt/mTOR signaling pathway in cancer stem cells: from basic research to clinical application. Am J Cancer Res. 2015;5:1602–9.

    PubMed  PubMed Central  Google Scholar 

  56. McDonald CA, Millena AC, Reddy S, Finlay S, Vizcarra J, Khan SA, Davis JS. Follicle-stimulating hormone-induced aromatase in immature rat Sertoli cells requires an active phosphatidylinositol 3-kinase pathway and is inhibited via the mitogen-activated protein kinase signaling pathway. Mol Endocrinol. 2006;20:608–18.

    Article  CAS  PubMed  Google Scholar 

  57. Nechamen CA, Thomas RM, Cohen BD, Acevedo G, Poulikakos PI, Testa JR, Dias JA. Human follicle-stimulating hormone (FSH) receptor interacts with the adaptor protein APPL1 in HEK 293 cells: potential involvement of the PI3K pathway in FSH signaling. Biol Reprod. 2004;71:629–36.

    Article  CAS  PubMed  Google Scholar 

  58. Chen YJ, Hsiao PW, Lee MT, Mason JI, Ke FC, Hwang JJ. Interplay of PI3K and cAMP/PKA signaling, and rapamycin-hypersensitivity in TGFbeta1 enhancement of FSH-stimulated steroidogenesis in rat ovarian granulosa cells. J Endocrinol. 2007;192:405–19.

    Article  CAS  PubMed  Google Scholar 

  59. Jesus TT, Oliveira PF, Sousa M, Cheng CY, Alves MG. Mammalian target of rapamycin (mTOR): a central regulator of male fertility? Crit Rev Biochem Mol Biol. 2017;52:235–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Lécureuil C, Tesseraud S, Kara E, Martinat N, Sow A, Fontaine I, Gauthier C, Reiter E, Guillou F, Crépieux P. Follicle-stimulating hormone activates p70 ribosomal protein S6 kinase by protein kinase A-mediated dephosphorylation of Thr 421/Ser 424 in primary Sertoli cells. Mol Endocrinol. 2005;19:1812–20.

    Article  PubMed  CAS  Google Scholar 

  61. Alam H, Maizels ET, Park Y, Ghaey S, Feiger ZJ, Chandel NS, Hunzicker-Dunn M. Follicle-stimulating hormone activation of hypoxia-inducible factor-1 by the phosphatidylinositol 3-kinase/AKT/Ras homolog enriched in brain (Rheb)/mammalian target of rapamycin (mTOR) pathway is necessary for induction of select protein markers of follicular differentiation. J Biol Chem. 2004;279:19431–40.

    Article  CAS  PubMed  Google Scholar 

  62. Geva E, Amit A, Lerner-Geva L, Lessing JB. Autoimmunity and reproduction. Fertil Steril. 1997;67:599–611.

    Article  CAS  PubMed  Google Scholar 

  63. Guo X, Morris P, Gudas L. Follicle-stimulating hormone and leukemia inhibitory factor regulate Sertoli cell retinol metabolism. Endocrinology. 2001;142:1024–32.

    Article  CAS  PubMed  Google Scholar 

  64. Theodosiou M, Laudet V, Schubert M. From carrot to clinic: an overview of the retinoic acid signaling pathway. Cell Mol Life Sci. 2010;67:1423–45.

    Article  CAS  PubMed  Google Scholar 

  65. Vernet N, Dennefeld C, Guillou F, Chambon P, Ghyselinck NB, Mark M. Prepubertal testis development relies on retinoic acid but not rexinoid receptors in Sertoli cells. Embo j. 2006;25:5816–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Dufour JM, Kim KH. Cellular and subcellular localization of six retinoid receptors in rat testis during postnatal development: identification of potential heterodimeric receptors. Biol Reprod. 1999;61:1300–8.

    Article  CAS  PubMed  Google Scholar 

  67. Santos NC, Kim KH. Activity of retinoic acid receptor-alpha is directly regulated at its protein kinase A sites in response to follicle-stimulating hormone signaling. Endocrinology. 2010;151:2361–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Walker WH, Cheng J. FSH and testosterone signaling in Sertoli cells. Reproduction. 2005;130:15–28.

    Article  CAS  PubMed  Google Scholar 

  69. Gloaguen P, Crépieux P, Heitzler D, Poupon A, Reiter E. Mapping the follicle-stimulating hormone-induced signaling networks. Front Endocrinol (Lausanne). 2011;2:45.

    Article  CAS  Google Scholar 

  70. Casarini L, Crépieux P. Molecular Mechanisms of Action of FSH. Front Endocrinol (Lausanne). 2019;10:305.

    Article  Google Scholar 

  71. Griswold MD, Solari A, Tung PS, Fritz IB. Stimulation by follicle-stimulating hormone of DNA synthesis and of mitosis in cultured Sertoli cells prepared from testes of immature rats. Mol Cell Endocrinol. 1977;7:151–65.

    Article  CAS  PubMed  Google Scholar 

  72. Orth JM. The role of follicle-stimulating hormone in controlling Sertoli cell proliferation in testes of fetal rats. Endocrinology. 1984;115:1248–55.

    Article  CAS  PubMed  Google Scholar 

  73. Griswold MD, Mably ER, Fritz IB. FSH stimulation of DNA synthesis in Sertoli cells in culture. Mol Cell Endocrinol. 1976;4:139–49.

    Article  CAS  PubMed  Google Scholar 

  74. Orth JM, Higginbotham CA, Salisbury RL. Hemicastration causes and testosterone prevents enhanced uptake of [3H] thymidine by Sertoli cells in testes of immature rats. Biol Reprod. 1984;30:263–70.

    Article  CAS  PubMed  Google Scholar 

  75. Almirón I, Chemes H. Spermatogenic onset. II. FSH modulates mitotic activity of germ and Sertoli cells in immature rats. Int J Androl. 1988;11:235–46.

    Article  PubMed  Google Scholar 

  76. Singh J, Handelsman DJ. The effects of recombinant FSH on testosterone-induced spermatogenesis in gonadotrophin-deficient (hpg) mice. J Androl. 1996;17:382–93.

    CAS  PubMed  Google Scholar 

  77. Meachem SJ, McLachlan RI, de Kretser DM, Robertson DM, Wreford NG. Neonatal exposure of rats to recombinant follicle stimulating hormone increases adult Sertoli and spermatogenic cell numbers. Biol Reprod. 1996;54:36–44.

    Article  CAS  PubMed  Google Scholar 

  78. Allan CM, Garcia A, Spaliviero J, Zhang FP, Jimenez M, Huhtaniemi I, Handelsman DJ. Complete Sertoli cell proliferation induced by follicle-stimulating hormone (FSH) independently of luteinizing hormone activity: evidence from genetic models of isolated FSH action. Endocrinology. 2004;145:1587–93.

    Article  CAS  PubMed  Google Scholar 

  79. Ulloa-Aguirre A, Reiter E, Crépieux P. FSH Receptor Signaling: Complexity of Interactions and Signal Diversity. Endocrinology. 2018;159:3020–35.

    Article  CAS  PubMed  Google Scholar 

  80. Riera MF, Regueira M, Galardo MN, Pellizzari EH, Meroni SB, Cigorraga SB. Signal transduction pathways in FSH regulation of rat Sertoli cell proliferation. Am J Physiol Endocrinol Metab. 2012;302:E914-923.

    Article  CAS  PubMed  Google Scholar 

  81. Rebello RJ, Oing C, Knudsen KE, Loeb S, Johnson DC, Reiter RE, Gillessen S, Van der Kwast T, Bristow RG. Prostate cancer Nat Rev Dis Primers. 2021;7:9.

    Article  PubMed  Google Scholar 

  82. Lim K, Hwang BD. Follicle-stimulating hormone transiently induces expression of protooncogene c-myc in primary Sertoli cell cultures of early pubertal and prepubertal rat. Mol Cell Endocrinol. 1995;111:51–6.

    Article  CAS  PubMed  Google Scholar 

  83. Parra-Herran C, McCluggage WG. Ovarian microcystic stromal tumour: From morphological observations to syndromic associations. Histopathology. 2022;80(6):898–904.

    Article  PubMed  Google Scholar 

  84. Ke Q, Costa M. Hypoxia-inducible factor-1 (HIF-1). Mol Pharmacol. 2006;70:1469–80.

    Article  CAS  PubMed  Google Scholar 

  85. Albadari N, Deng S, Li W. The transcriptional factors HIF-1 and HIF-2 and their novel inhibitors in cancer therapy. Expert Opin Drug Discov. 2019;14:667–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Rato L, Alves MG, Silva BM, Sousa M, Oliveira PF. Sirtuins: Novel Players in Male Reproductive Health. Curr Med Chem. 2016;23:1084–99.

    Article  CAS  PubMed  Google Scholar 

  87. Godmann M, Kosan C, Behr R. Krüppel-like factor 4 is widely expressed in the mouse male and female reproductive tract and responds as an immediate early gene to activation of the protein kinase A in TM4 Sertoli cells. Reproduction. 2010;139:771–82.

    Article  CAS  PubMed  Google Scholar 

  88. Godmann M, Katz JP, Guillou F, Simoni M, Kaestner KH, Behr R. Krüppel-like factor 4 is involved in functional differentiation of testicular Sertoli cells. Dev Biol. 2008;315:552–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Delfino F, Walker WH. Stage-specific nuclear expression of NF-kappaB in mammalian testis. Mol Endocrinol. 1998;12:1696–707.

    CAS  PubMed  Google Scholar 

  90. Gruber M, Mathew LK, Runge AC, Garcia JA, Simon MC. EPAS1 Is Required for Spermatogenesis in the Postnatal Mouse Testis. Biol Reprod. 2010;82:1227–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Hamil KG, Conti M, Shimasaki S, Hall SH. Follicle-stimulating hormone regulation of AP-1: inhibition of c-jun and stimulation of jun-B gene transcription in the rat Sertoli cell. Mol Cell Endocrinol. 1994;99:269–77.

    Article  CAS  PubMed  Google Scholar 

  92. Guo J, Shi YQ, Yang W, Li YC, Hu ZY, Liu YX. Testosterone upregulation of tissue type plasminogen activator expression in Sertoli cells : tPA expression in Sertoli cells. Endocrine. 2007;32:83–9.

    Article  CAS  PubMed  Google Scholar 

  93. White S, Lin L, Hu K. NF-κB and tPA Signaling in Kidney and Other Diseases. Cells. 2020;9(6):1348.

    Article  CAS  PubMed Central  Google Scholar 

  94. Hongo S, Chiyo T, Sato T. Induction of asparagine synthetase by follicle-stimulating hormone in primary cultures of rat Sertoli cells. Arch Biochem Biophys. 1994;313:222–8.

    Article  CAS  PubMed  Google Scholar 

  95. Tani T, Koyama Y, Nihei K, Hatakeyama S, Ohshiro K, Yoshida Y, Yaoita E, Sakai Y, Hatakeyama K, Yamamoto T. Immunolocalization of aquaporin-8 in rat digestive organs and testis. Arch Histol Cytol. 2001;64:159–68.

    Article  CAS  PubMed  Google Scholar 

  96. Soffientini U, Rebourcet D, Abel MH, Lee S, Hamilton G, Fowler PA, Smith LB, O’Shaughnessy PJ. Identification of Sertoli cell-specific transcripts in the mouse testis and the role of FSH and androgen in the control of Sertoli cell activity. BMC Genomics. 2017;18:972.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Basu S, Arya SP, Usmani A, Pradhan BS, Sarkar RK, Ganguli N, Shukla M, Mandal K, Singh S, Sarda K, Majumdar SS. Defective Wnt3 expression by testicular Sertoli cells compromise male fertility. Cell Tissue Res. 2018;371:351–63.

    Article  CAS  PubMed  Google Scholar 

  98. Anawalt BD, Bebb RA, Bremner WJ, Matsumoto AM. A lower dosage levonorgestrel and testosterone combination effectively suppresses spermatogenesis and circulating gonadotropin levels with fewer metabolic effects than higher dosage combinations. J Androl. 1999;20:407–14.

    CAS  PubMed  Google Scholar 

  99. Salanova M, Ricci G, Boitani C, Stefanini M, De Grossi S, Palombi F. Junctional contacts between Sertoli cells in normal and aspermatogenic rat seminiferous epithelium contain alpha6beta1 integrins, and their formation is controlled by follicle-stimulating hormone. Biol Reprod. 1998;58:371–8.

    Article  CAS  PubMed  Google Scholar 

  100. Shobana N, Kumar MK, Navin AK, Akbarsha MA, Aruldhas MM. Prenatal exposure to excess chromium attenuates transcription factors regulating expression of androgen and follicle stimulating hormone receptors in Sertoli cells of prepuberal rats. Chem Biol Interact. 2020;328:109188.

    Article  CAS  PubMed  Google Scholar 

  101. Regueira M, Artagaveytia SL, Galardo MN, Pellizzari EH, Cigorraga SB, Meroni SB, Riera MF. Novel molecular mechanisms involved in hormonal regulation of lactate production in Sertoli cells. Reproduction. 2015;150:311–21.

    Article  CAS  PubMed  Google Scholar 

  102. Maccarrone M, Cecconi S, Rossi G, Battista N, Pauselli R, Finazzi-Agrò A. Anandamide activity and degradation are regulated by early postnatal aging and follicle-stimulating hormone in mouse Sertoli cells. Endocrinology. 2003;144:20–8.

    Article  CAS  PubMed  Google Scholar 

  103. Park SY, Lee EJ, Emge D, Jahn CL, Jameson JL. A phenotypic spectrum of sexual development in Dax1 (Nr0b1)-deficient mice: consequence of the C57BL/6J strain on sex determination. Biol Reprod. 2008;79:1038–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Simon L, Ekman GC, Tyagi G, Hess RA, Murphy KM, Cooke PS. Common and distinct factors regulate expression of mRNA for ETV5 and GDNF, Sertoli cell proteins essential for spermatogonial stem cell maintenance. Exp Cell Res. 2007;313:3090–9.

    Article  CAS  PubMed  Google Scholar 

  105. Morimoto H, Iwata K, Ogonuki N, Inoue K, Atsuo O, Kanatsu-Shinohara M, Morimoto T, Yabe-Nishimura C, Shinohara T. ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell. 2013;12:774–86.

    Article  CAS  PubMed  Google Scholar 

  106. Kubota H, Avarbock MR, Brinster RL. Growth factors essential for self-renewal and expansion of mouse spermatogonial stem cells. Proc Natl Acad Sci U S A. 2004;101:16489–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Masaki K, Sakai M, Kuroki S, Jo JI, Hoshina K, Fujimori Y, Oka K, Amano T, Yamanaka T, Tachibana M, et al. FGF2 Has Distinct Molecular Functions from GDNF in the Mouse Germline Niche. Stem Cell Reports. 2018;10:1782–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Crespo D, Lemos MS, Zhang YT, Safian D, Norberg B, Bogerd J, Schulz RW. PGE2 inhibits spermatogonia differentiation in zebrafish: interaction with Fsh and an androgen. J Endocrinol. 2020;244:163–75.

    Article  CAS  PubMed  Google Scholar 

  109. Li Y, Zhang Y, Zhang X, Sun J, Hao J. BMP4/Smad signaling pathway induces the differentiation of mouse spermatogonial stem cells via upregulation of Sohlh2. Anat Rec (Hoboken). 2014;297:749–57.

    Article  CAS  Google Scholar 

  110. Mikedis MM, Fan Y, Nicholls PK, Endo T, Jackson EK, Cobb SA, de Rooij DG, Page DC. DAZL mediates a broad translational program regulating expansion and differentiation of spermatogonial progenitors. Elife. 2020;9:e56523.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Rossi P, Dolci S, Albanesi C, Grimaldi P, Ricca R, Geremia R. Follicle-stimulating hormone induction of steel factor (SLF) mRNA in mouse Sertoli cells and stimulation of DNA synthesis in spermatogonia by soluble SLF. Dev Biol. 1993;155:68–74.

    Article  CAS  PubMed  Google Scholar 

  112. Rossi P, Sette C, Dolci S, Geremia R. Role of c-kit in mammalian spermatogenesis. J Endocrinol Invest. 2000;23:609–15.

    Article  CAS  PubMed  Google Scholar 

  113. Nóbrega RH, Morais RD, Crespo D, de Waal PP, de França LR, Schulz RW, Bogerd J. Fsh Stimulates Spermatogonial Proliferation and Differentiation in Zebrafish via Igf3. Endocrinology. 2015;156:3804–17.

    Article  PubMed  CAS  Google Scholar 

  114. Safian D, Bogerd J, Schulz RW. Igf3 activates β-catenin signaling to stimulate spermatogonial differentiation in zebrafish. J Endocrinol. 2018;238:245–57.

    Article  CAS  PubMed  Google Scholar 

  115. Nayak G, Rao A, Mullick P, Mutalik S, Kalthur SG, Adiga SK, Kalthur G. Ethanolic extract of Moringa oleifera leaves alleviate cyclophosphamide-induced testicular toxicity by improving endocrine function and modulating cell specific gene expression in mouse testis. J Ethnopharmacol. 2020;259:112922.

    Article  CAS  PubMed  Google Scholar 

  116. Darebna P, Spicka J, Kucera R, Topolcan O, Navratilova E, Ruzicka V, Volny M, Novak P, Pompach P. Detection and Quantification of Carbohydrate-Deficient Transferrin by MALDI-Compatible Protein Chips Prepared by Ambient Ion Soft Landing. Clin Chem. 2018;64:1319–26.

    Article  CAS  PubMed  Google Scholar 

  117. Suominen JS, Yan W, Toppari J, Kaipia A. The expression and regulation of Bcl-2-related ovarian killer (Bok) mRNA in the developing and adult rat testis. Eur J Endocrinol. 2001;145:771–8.

    Article  CAS  PubMed  Google Scholar 

  118. Ruwanpura SM, McLachlan RI, Stanton PG, Loveland KL, Meachem SJ. Pathways involved in testicular germ cell apoptosis in immature rats after FSH suppression. J Endocrinol. 2008;197:35–43.

    Article  CAS  PubMed  Google Scholar 

  119. Appiah Adu-Gyamfi E, Tanam Djankpa F, Nelson W, Czika A, Kumar Sah S, Lamptey J, Ding YB, Wang YX. Activin and inhibin signaling: From regulation of physiology to involvement in the pathology of the female reproductive system. Cytokine. 2020;133:155105.

    Article  CAS  PubMed  Google Scholar 

  120. Hedger MP, Winnall WR. Regulation of activin and inhibin in the adult testis and the evidence for functional roles in spermatogenesis and immunoregulation. Mol Cell Endocrinol. 2012;359:30–42.

    Article  CAS  PubMed  Google Scholar 

  121. Hakovirta H, Syed V, Jégou B, Parvinen M. Function of interleukin-6 as an inhibitor of meiotic DNA synthesis in the rat seminiferous epithelium. Mol Cell Endocrinol. 1995;108:193–8.

    Article  CAS  PubMed  Google Scholar 

  122. Eto K. Nociceptin and meiosis during spermatogenesis in postnatal testes. Vitam Horm. 2015;97:167–86.

    Article  CAS  PubMed  Google Scholar 

  123. Eto K, Shiotsuki M, Sakai T, Abe S. Nociceptin is upregulated by FSH signaling in Sertoli cells in murine testes. Biochem Biophys Res Commun. 2012;421:678–83.

    Article  CAS  PubMed  Google Scholar 

  124. Falls DL. Neuregulins: functions, forms, and signaling strategies. Exp Cell Res. 2003;284:14–30.

    Article  CAS  PubMed  Google Scholar 

  125. Zhang J, Eto K, Honmyou A, Nakao K, Kiyonari H, Abé S. Neuregulins are essential for spermatogonial proliferation and meiotic initiation in neonatal mouse testis. Development. 2011;138:3159–68.

    Article  CAS  PubMed  Google Scholar 

  126. Yu F, Finley RL Jr, Raz A, Kim HR. Galectin-3 translocates to the perinuclear membranes and inhibits cytochrome c release from the mitochondria. A role for synexin in galectin-3 translocation. J Biol Chem. 2002;277:15819–27.

    Article  CAS  PubMed  Google Scholar 

  127. Deschildre C, Ji JW, Chater S, Dacheux F, Selva J, Albert M, Bailly M, Hatey F, Benahmed M. Expression of galectin-3 and its regulation in the testes. Int J Androl. 2007;30:28–40.

    Article  CAS  PubMed  Google Scholar 

  128. Suomalainen L, Dunkel L, Ketola I, Eriksson M, Erkkilä K, Oksjoki R, Taari K, Heikinheimo M, Pentikäinen V. Activator protein-1 in human male germ cell apoptosis. Mol Hum Reprod. 2004;10:743–53.

    Article  CAS  PubMed  Google Scholar 

  129. Galardo MN, Gorga A, Merlo JP, Regueira M, Pellizzari EH, Cigorraga SB, Riera MF, Meroni SB. Participation of HIFs in the regulation of Sertoli cell lactate production. Biochimie. 2017;132:9–18.

    Article  CAS  PubMed  Google Scholar 

  130. Gorga A, Rindone G, Regueira M, Riera MF, Pellizzari EH, Cigorraga SB, Meroni SB, Galardo MN. HIF involvement in the regulation of rat Sertoli cell proliferation by FSH. Biochem Biophys Res Commun. 2018;502:508–14.

    Article  CAS  PubMed  Google Scholar 

  131. Raval RR, Lau KW, Tran MG, Sowter HM, Mandriota SJ, Li JL, Pugh CW, Maxwell PH, Harris AL, Ratcliffe PJ. Contrasting properties of hypoxia-inducible factor 1 (HIF-1) and HIF-2 in von Hippel-Lindau-associated renal cell carcinoma. Mol Cell Biol. 2005;25:5675–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Lim JH, Lee YM, Chun YS, Chen J, Kim JE, Park JW. Sirtuin 1 modulates cellular responses to hypoxia by deacetylating hypoxia-inducible factor 1alpha. Mol Cell. 2010;38:864–78.

    Article  CAS  PubMed  Google Scholar 

  133. Sadate-Ngatchou PI, Pouchnik DJ, Griswold MD. Follicle-stimulating hormone induced changes in gene expression of murine testis. Mol Endocrinol. 2004;18:2805–16.

    Article  CAS  PubMed  Google Scholar 

  134. Griffeth RJ, Bianda V, Nef S. The emerging role of insulin-like growth factors in testis development and function. Basic Clin Androl. 2014;24:12.

    Article  PubMed  PubMed Central  Google Scholar 

  135. Heng D, Wang Q, Ma X, Tian Y, Xu K, Weng X, Hu X, Liu W, Zhang C. Role of OCT4 in the Regulation of FSH-Induced Granulosa Cells Growth in Female Mice. Front Endocrinol (Lausanne). 2019;10:915.

    Article  Google Scholar 

  136. Sá R, Miranda C, Carvalho F, Barros A, Sousa M. Expression of stem cell markers: OCT4, KIT, ITGA6, and ITGB1 in the male germinal epithelium. Syst Biol Reprod Med. 2013;59:233–43.

    Article  PubMed  CAS  Google Scholar 

  137. O’Shaughnessy PJ, Monteiro A, Abel M. Testicular development in mice lacking receptors for follicle stimulating hormone and androgen. PLoS ONE. 2012;7:e35136.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Bhattacharya I, Sharma SS, Sarkar H, Gupta A, Pradhan BS, Majumdar SS. FSH mediated cAMP signalling upregulates the expression of Gα subunits in pubertal rat Sertoli cells. Biochem Biophys Res Commun. 2021;569:100–5.

    Article  CAS  PubMed  Google Scholar 

  139. Sen Sharma S, Vats A, Majumdar S. Regulation of Hippo pathway components by FSH in testis. Reprod Biol. 2019;19:61–6.

    Article  PubMed  Google Scholar 

  140. Ma S, Meng Z, Chen R, Guan KL. The Hippo Pathway: Biology and Pathophysiology. Annu Rev Biochem. 2019;88:577–604.

    Article  CAS  PubMed  Google Scholar 

  141. Gupta A, Vats A, Ghosal A, Mandal K, Sarkar R, Bhattacharya I, Das S, Pal R, Majumdar SS. Follicle-stimulating hormone-mediated decline in miR-92a-3p expression in pubertal mice Sertoli cells is crucial for germ cell differentiation and fertility. Cell Mol Life Sci. 2022;79:136.

    Article  CAS  PubMed  Google Scholar 

  142. Rato L, Meneses MJ, Silva BM, Sousa M, Alves MG, Oliveira PF. New insights on hormones and factors that modulate Sertoli cell metabolism. Histol Histopathol. 2016;31:499–513.

    CAS  PubMed  Google Scholar 

  143. Meroni SB, Riera MF, Pellizzari EH, Cigorraga SB. Regulation of rat Sertoli cell function by FSH: possible role of phosphatidylinositol 3-kinase/protein kinase B pathway. J Endocrinol. 2002;174:195–204.

    Article  CAS  PubMed  Google Scholar 

  144. Sluka P, O’Donnell L, Bartles JR, Stanton PG. FSH regulates the formation of adherens junctions and ectoplasmic specialisations between rat Sertoli cells in vitro and in vivo. J Endocrinol. 2006;189:381–95.

    Article  CAS  PubMed  Google Scholar 

  145. Stanton PG. Regulation of the blood-testis barrier. Semin Cell Dev Biol. 2016;59:166–73.

    Article  CAS  PubMed  Google Scholar 

  146. Rossi G, Gasperi V, Paro R, Barsacchi D, Cecconi S, Maccarrone M. Follicle-stimulating hormone activates fatty acid amide hydrolase by protein kinase A and aromatase-dependent pathways in mouse primary Sertoli cells. Endocrinology. 2007;148:1431–9.

    Article  CAS  PubMed  Google Scholar 

  147. Tamai KT, Monaco L, Alastalo TP, Lalli E, Parvinen M, Sassone-Corsi P. Hormonal and developmental regulation of DAX-1 expression in Sertoli cells. Mol Endocrinol. 1996;10:1561–9.

    CAS  PubMed  Google Scholar 

  148. Xie Y, Wei BH, Ni FD, Yang WX. Conversion from spermatogonia to spermatocytes: Extracellular cues and downstream transcription network. Gene. 2021;764:145080.

    Article  CAS  PubMed  Google Scholar 

  149. Huckins C. The spermatogonial stem cell population in adult rats. I. Their morphology, proliferation and maturation. Anat Rec. 1971;169:533–57.

    Article  CAS  PubMed  Google Scholar 

  150. Clermont Y. The cycle of the seminiferous epithelium in man. Am J Anat. 1963;112:35–51.

    Article  CAS  PubMed  Google Scholar 

  151. Leblond CP, Clermont Y. Definition of the stages of the cycle of the seminiferous epithelium in the rat. Ann N Y Acad Sci. 1952;55:548–73.

    Article  CAS  PubMed  Google Scholar 

  152. Di Persio S, Saracino R, Fera S, Muciaccia B, Esposito V, Boitani C, Berloco BP, Nudo F, Spadetta G, Stefanini M, et al. Spermatogonial kinetics in humans. Dev. 2017;144:3430–9.

    Article  CAS  Google Scholar 

  153. Patel H, Bhartiya D. Direct action of FSH on testicular stem cells. Stem Cell Res Ther. 2019;10:261.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. McLachlan RI, O’Donnell L, Meachem SJ, Stanton PG. de K, Pratis K, Robertson DM: Hormonal regulation of spermatogenesis in primates and man: insights for development of the male hormonal contraceptive. J Androl. 2002;23:149–62.

    CAS  PubMed  Google Scholar 

  155. Ruwanpura SM, McLachlan RI, Stanton PG, Meachem SJ. Follicle-stimulating hormone affects spermatogonial survival by regulating the intrinsic apoptotic pathway in adult rats. Biol Reprod. 2008;78:705–13.

    Article  CAS  PubMed  Google Scholar 

  156. Sofikitis N, Giotitsas N, Tsounapi P, Baltogiannis D, Giannakis D, Pardalidis N. Hormonal regulation of spermatogenesis and spermiogenesis. J Steroid Biochem Mol Biol. 2008;109:323–30.

    Article  CAS  PubMed  Google Scholar 

  157. Jabarpour M, Tajik P. Evaluation of the effect of follicular stimulating hormone on the in vitro bovine spermatogonial stem cells self-renewal: An experimental study. Int J Reprod Biomed. 2017;15:795–802.

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Xi H, Ren F, Li Y, Xian M, Wang L, Hu J. FSH inhibits autophagy and lysosomal biogenesis to regulate protein degradation in cultured goat Sertoli cells. Mol Cell Endocrinol. 2022;540:111505.

    Article  CAS  PubMed  Google Scholar 

  159. Bellaïche J, Goupil AS, Sambroni E, Lareyre JJ, Le Gac F. Gdnf-Gfra1 pathway is expressed in a spermatogenetic-dependent manner and is regulated by Fsh in a fish testis. Biol Reprod. 2014;91:94.

    Article  PubMed  CAS  Google Scholar 

  160. Bhattacharya I, Pradhan BS, Sarda K, Gautam M, Basu S, Majumdar SS. A switch in Sertoli cell responsiveness to FSH may be responsible for robust onset of germ cell differentiation during prepubartal testicular maturation in rats. Am J Physiol Endocrinol Metab. 2012;303:E886-898.

    Article  CAS  PubMed  Google Scholar 

  161. Rossi P, Dolci S. Paracrine mechanisms involved in the control of early stages of Mammalian spermatogenesis. Front Endocrinol (Lausanne). 2013;4:181.

    Article  Google Scholar 

  162. Pellegrini M, Filipponi D, Gori M, Barrios F, Lolicato F, Grimaldi P, Rossi P, Jannini EA, Geremia R, Dolci S. ATRA and KL promote differentiation toward the meiotic program of male germ cells. Cell Cycle. 2008;7:3878–88.

    Article  CAS  PubMed  Google Scholar 

  163. Tassinari V, Campolo F, Cesarini V, Todaro F, Dolci S, Rossi P. Fgf9 inhibition of meiotic differentiation in spermatogonia is mediated by Erk-dependent activation of Nodal-Smad2/3 signaling and is antagonized by Kit Ligand. Cell Death Dis. 2015;6:e1688.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Rossi P, Albanesi C, Grimaldi P, Geremia R. Expression of the mRNA for the ligand of c-kit in mouse Sertoli cells. Biochem Biophys Res Commun. 1991;176:910–4.

    Article  CAS  PubMed  Google Scholar 

  165. Manova K, Huang EJ, Angeles M, De Leon V, Sanchez S, Pronovost SM, Besmer P, Bachvarova RF. The expression pattern of the c-kit ligand in gonads of mice supports a role for the c-kit receptor in oocyte growth and in proliferation of spermatogonia. Dev Biol. 1993;157:85–99.

    Article  CAS  PubMed  Google Scholar 

  166. Carlomagno G, van Bragt MP, Korver CM, Repping S, de Rooij DG, van Pelt AM. BMP4-induced differentiation of a rat spermatogonial stem cell line causes changes in its cell adhesion properties. Biol Reprod. 2010;83:742–9.

    Article  CAS  PubMed  Google Scholar 

  167. Pellegrini M, Grimaldi P, Rossi P, Geremia R, Dolci S. Developmental expression of BMP4/ALK3/SMAD5 signaling pathway in the mouse testis: a potential role of BMP4 in spermatogonia differentiation. J Cell Sci. 2003;116:3363–72.

    Article  CAS  PubMed  Google Scholar 

  168. Ashkenazi A, Fairbrother WJ, Leverson JD, Souers AJ. From basic apoptosis discoveries to advanced selective BCL-2 family inhibitors. Nat Rev Drug Discov. 2017;16:273–84.

    Article  CAS  PubMed  Google Scholar 

  169. Ito R, Abé SI. FSH-initiated differentiation of newt spermatogonia to primary spermatocytes in germ-somatic cell reaggregates cultured within a collagen matrix. Int J Dev Biol. 1999;43:111–6.

    CAS  PubMed  Google Scholar 

  170. O’Shaughnessy PJ, Monteiro A, Verhoeven G, De Gendt K, Abel MH. Effect of FSH on testicular morphology and spermatogenesis in gonadotrophin-deficient hypogonadal mice lacking androgen receptors. Reproduction. 2010;139:177–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Baines H, Nwagwu MO, Hastie GR, Wiles RA, Mayhew TM, Ebling FJ. Effects of estradiol and FSH on maturation of the testis in the hypogonadal (hpg) mouse. Reprod Biol Endocrinol. 2008;6:4.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  172. Russell LD, Corbin TJ, Borg KE, De França LR, Grasso P, Bartke A. Recombinant human follicle-stimulating hormone is capable of exerting a biological effect in the adult hypophysectomized rat by reducing the numbers of degenerating germ cells. Endocrinology. 1993;133:2062–70.

    Article  CAS  PubMed  Google Scholar 

  173. Wreford NG, Rajendra Kumar T, Matzuk MM, de Kretser DM. Analysis of the testicular phenotype of the follicle-stimulating hormone beta-subunit knockout and the activin type II receptor knockout mice by stereological analysis. Endocrinology. 2001;142:2916–20.

    Article  CAS  PubMed  Google Scholar 

  174. Eto K, Shiotsuki M, Abe S. Nociceptin induces Rec8 phosphorylation and meiosis in postnatal murine testes. Endocrinology. 2013;154:2891–9.

    Article  CAS  PubMed  Google Scholar 

  175. Gewiss RL, Shelden EA, Griswold MD. STRA8 induces transcriptional changes in germ cells during spermatogonial development. Mol Reprod Dev. 2021;88:128–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Meachem SJ, Ruwanpura SM, Ziolkowski J, Ague JM, Skinner MK, Loveland KL. Developmentally distinct in vivo effects of FSH on proliferation and apoptosis during testis maturation. J Endocrinol. 2005;186:429–46.

    Article  CAS  PubMed  Google Scholar 

  177. Hildorf S, Cortes D, Thorup J, Clasen-Linde E, Hutson J, Li R. During infancy low levels of follicle-stimulating hormone may result in high rate of germ cell apoptosis. J Pediatr Surg. 2021;56:2399–406.

    Article  PubMed  Google Scholar 

  178. Meachem SJ, McLachlan RI, Stanton PG, Robertson DM, Wreford NG. FSH immunoneutralization acutely impairs spermatogonial development in normal adult rats. J Androl. 1999;20:756–62 (discussion 755).

    CAS  PubMed  Google Scholar 

  179. Sá R, Neves R, Fernandes S, Alves C, Carvalho F, Silva J, Cremades N, Malheiro I, Barros A, Sousa M. Cytological and expression studies and quantitative analysis of the temporal and stage-specific effects of follicle-stimulating hormone and testosterone during cocultures of the normal human seminiferous epithelium. Biol Reprod. 2008;79:962–75.

    Article  PubMed  CAS  Google Scholar 

  180. Sá R, Graça I, Silva J, Malheiro I, Carvalho F, Barros A, Sousa M. Quantitative analysis of cellular proliferation and differentiation of the human seminiferous epithelium in vitro. Reprod Sci. 2012;19:1063–74.

    Article  PubMed  CAS  Google Scholar 

  181. Moreno RD, Lizama C, Urzúa N, Vergara SP, Reyes JG. Caspase activation throughout the first wave of spermatogenesis in the rat. Cell Tissue Res. 2006;325:533–40.

    Article  CAS  PubMed  Google Scholar 

  182. Codelia VA, Cisternas P, Moreno RD. Relevance of caspase activity during apoptosis in pubertal rat spermatogenesis. Mol Reprod Dev. 2008;75:881–9.

    Article  CAS  PubMed  Google Scholar 

  183. Almeida C, Correia S, Rocha E, Alves A, Ferraz L, Silva J, Sousa M, Barros A. Caspase signalling pathways in human spermatogenesis. J Assist Reprod Genet. 2013;30:487–95.

    Article  PubMed  PubMed Central  Google Scholar 

  184. Sousa M, Cremades N, Alves C, Silva J, Barros A. Developmental potential of human spermatogenic cells co-cultured with Sertoli cells. Hum Reprod. 2002;17:161–72.

    Article  PubMed  Google Scholar 

  185. Recchia K, Jorge AS, Pessôa LVF, Botigelli RC, Zugaib VC, de Souza AF, Martins DDS, Ambrósio CE, Bressan FF, Pieri NCG. Actions and Roles of FSH in Germinative Cells. Int J Mol Sci. 2021;22(18):10110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Walker WH. Androgen Actions in the Testis and the Regulation of Spermatogenesis. Adv Exp Med Biol. 2021;1288:175–203.

    Article  PubMed  Google Scholar 

  187. Behre HM. Clinical Use of FSH in Male Infertility. Front Endocrinol (Lausanne). 2019;10:322.

    Article  Google Scholar 

  188. Allan CM, Haywood M, Swaraj S, Spaliviero J, Koch A, Jimenez M, Poutanen M, Levallet J, Huhtaniemi I, Illingworth P, Handelsman DJ. A novel transgenic model to characterize the specific effects of follicle-stimulating hormone on gonadal physiology in the absence of luteinizing hormone actions. Endocrinology. 2001;142:2213–20.

    Article  CAS  PubMed  Google Scholar 

  189. Snyder PJ. Gonadotroph cell pituitary adenomas. Endocrinol Metab Clin North Am. 1987;16:755–64.

    Article  CAS  PubMed  Google Scholar 

  190. Gromoll J, Simoni M, Nieschlag E. An activating mutation of the follicle-stimulating hormone receptor autonomously sustains spermatogenesis in a hypophysectomized man. J Clin Endocrinol Metab. 1996;81:1367–70.

    CAS  PubMed  Google Scholar 

  191. Casas-González P, Scaglia HE, Pérez-Solís MA, Durand G, Scaglia J, Zariñán T, Dias JA, Reiter E, Ulloa-Aguirre A. Normal testicular function without detectable follicle-stimulating hormone. A novel mutation in the follicle-stimulating hormone receptor gene leading to apparent constitutive activity and impaired agonist-induced desensitization and internalization. Mol Cell Endocrinol. 2012;364:71–82.

    Article  PubMed  CAS  Google Scholar 

  192. Mao JF, Liu ZX, Nie M, Wang X, Xu HL, Huang BK, Zheng JJ, Min L, Kaiser UB, Wu XY. Pulsatile gonadotropin-releasing hormone therapy is associated with earlier spermatogenesis compared to combined gonadotropin therapy in patients with congenital hypogonadotropic hypogonadism. Asian J Androl. 2017;19:680–5.

    Article  CAS  PubMed  Google Scholar 

  193. Sykiotis GP, Hoang XH, Avbelj M, Hayes FJ, Thambundit A, Dwyer A, Au M, Plummer L, Crowley WF Jr, Pitteloud N. Congenital idiopathic hypogonadotropic hypogonadism: evidence of defects in the hypothalamus, pituitary, and testes. J Clin Endocrinol Metab. 2010;95:3019–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Casarini L, Crépieux P, Reiter E, Lazzaretti C, Paradiso E, Rochira V, Brigante G, Santi D, Simoni M. FSH for the Treatment of Male Infertility. Int J Mol Sci. 2020;21(7):2270.

    Article  CAS  PubMed Central  Google Scholar 

  195. Rastrelli G, Corona G, Mannucci E, Maggi M. Factors affecting spermatogenesis upon gonadotropin-replacement therapy: a meta-analytic study. Andrology. 2014;2:794–808.

    Article  CAS  PubMed  Google Scholar 

  196. MacLeod J, Pazianos A, Ray B. The restoration of human spermatogenesis and of the reproductive tract with urinary gonadotropins following hypophysectomy. Fertil Steril. 1966;17:7–23.

    Article  CAS  PubMed  Google Scholar 

  197. Casarini L, Lispi M, Longobardi S, Milosa F, La Marca A, Tagliasacchi D, Pignatti E, Simoni M. LH and hCG action on the same receptor results in quantitatively and qualitatively different intracellular signalling. PLoS ONE. 2012;7:e46682.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Dwyer AA, Raivio T, Pitteloud N. Gonadotrophin replacement for induction of fertility in hypogonadal men. Best Pract Res Clin Endocrinol Metab. 2015;29:91–103.

    Article  CAS  PubMed  Google Scholar 

  199. Liu Z, Mao J, Wu X, Xu H, Wang X, Huang B, Zheng J, Nie M, Zhang H. Efficacy and Outcome Predictors of Gonadotropin Treatment for Male Congenital Hypogonadotropic Hypogonadism: A Retrospective Study of 223 Patients. Medicine (Baltimore). 2016;95:e2867.

    Article  CAS  Google Scholar 

  200. Santi D, Granata AR, Simoni M. FSH treatment of male idiopathic infertility improves pregnancy rate: a meta-analysis. Endocr Connect. 2015;4:R46-58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Barbonetti A, Calogero AE, Balercia G, Garolla A, Krausz C, La Vignera S, Lombardo F, Jannini EA, Maggi M, Lenzi A, et al. The use of follicle stimulating hormone (FSH) for the treatment of the infertile man: position statement from the Italian Society of Andrology and Sexual Medicine (SIAMS). J Endocrinol Invest. 2018;41:1107–22.

    Article  CAS  PubMed  Google Scholar 

  202. Behre HM, Greb RR, Mempel A, Sonntag B, Kiesel L, Kaltwasser P, Seliger E, Röpke F, Gromoll J, Nieschlag E, Simoni M. Significance of a common single nucleotide polymorphism in exon 10 of the follicle-stimulating hormone (FSH) receptor gene for the ovarian response to FSH: a pharmacogenetic approach to controlled ovarian hyperstimulation. Pharmacogenet Genomics. 2005;15:451–6.

    Article  CAS  PubMed  Google Scholar 

  203. Simoni M, Santi D, Negri L, Hoffmann I, Muratori M, Baldi E, Cambi M, Marcou M, Greither T, Baraldi E, et al. Treatment with human, recombinant FSH improves sperm DNA fragmentation in idiopathic infertile men depending on the FSH receptor polymorphism p.N680S: a pharmacogenetic study. Hum Reprod. 2016;31:1960–9.

    Article  PubMed  Google Scholar 

  204. Muratori M, Baldi E. Effects of FSH on Sperm DNA Fragmentation: Review of Clinical Studies and Possible Mechanisms of Action. Front Endocrinol (Lausanne). 2018;9:734.

    Article  Google Scholar 

  205. Ferlin A, Vinanzi C, Selice R, Garolla A, Frigo AC, Foresta C. Toward a pharmacogenetic approach to male infertility: polymorphism of follicle-stimulating hormone beta-subunit promoter. Fertil Steril. 2011;96:1344-1349.e1342.

    Article  CAS  PubMed  Google Scholar 

  206. Selice R, Ferlin A, Garolla A, Caretta N, Foresta C. Effects of endogenous FSH on normal human spermatogenesis in adults. Int J Androl. 2011;34:e511-517.

    Article  CAS  PubMed  Google Scholar 

  207. Uchiyama Y, Hosoe M, Sato T, Takahashi T. Biological and immunological characteristics of porcine follicle-stimulating hormone chemically modified with a polyethylene glycol derivative. Vet J. 2010;184:208–11.

    Article  CAS  PubMed  Google Scholar 

  208. Sugahara T, Pixley MR, Minami S, Perlas E, Ben-Menahem D, Hsueh AJ, Boime I. Biosynthesis of a biologically active single peptide chain containing the human common alpha and chorionic gonadotropin beta subunits in tandem. Proc Natl Acad Sci U S A. 1995;92:2041–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Garcia-Campayo V, Sato A, Hirsch B, Sugahara T, Muyan M, Hsueh AJ, Boime I. Design of stable biologically active recombinant lutropin analogs. Nat Biotechnol. 1997;15:663–7.

    Article  CAS  PubMed  Google Scholar 

  210. Lemke EP, Adams BM, Jablonka-Shariff A, Boime I, Adams TE. Single-chain human gonadotropin analogs induce follicle development in sheep. J Endocrinol. 2008;196:593–600.

    Article  CAS  PubMed  Google Scholar 

  211. Rutigliano HM, Adams BM, Jablonka-Shariff A, Boime I, Adams TE. Effect of single-chain ovine gonadotropins with dual activity on ovarian function in sheep. Reproduction. 2014;148:129–36.

    Article  CAS  PubMed  Google Scholar 

  212. Zhang YL, Guo KP, Ji SY, Liu XM, Wang P, Wu J, Gao L, Jiang TQ, Xu T, Fan HY. Development and characterization of a novel long-acting recombinant follicle stimulating hormone agonist by fusing Fc to an FSH-β subunit. Hum Reprod. 2016;31:169–82.

    Article  PubMed  CAS  Google Scholar 

  213. Martins AD, Monteiro MP, Silva BM, Barros A, Sousa M, Carvalho RA, Oliveira PF, Alves MG. Metabolic dynamics of human Sertoli cells are differentially modulated by physiological and pharmacological concentrations of GLP-1. Toxicol Appl Pharmacol. 2019;362:1–8.

    Article  CAS  PubMed  Google Scholar 

  214. Oliveira PF, Sousa M, Silva BM, Monteiro MP, Alves MG. Obesity, energy balance and spermatogenesis. Reproduction. 2017;153:R173-r185.

    Article  CAS  PubMed  Google Scholar 

  215. Meneses MJ, Bernardino RL, Sá R, Silva J, Barros A, Sousa M, Silva BM, Oliveira PF, Alves MG. Pioglitazone increases the glycolytic efficiency of human Sertoli cells with possible implications for spermatogenesis. Int J Biochem Cell Biol. 2016;79:52–60.

    Article  CAS  PubMed  Google Scholar 

  216. Bernardino RL, Dias TR, Moreira BP, Cunha M, Barros A, Oliveira E, Sousa M, Alves MG, Oliveira PF. Carbonic anhydrases are involved in mitochondrial biogenesis and control the production of lactate by human Sertoli cells. Febs j. 2019;286:1393–406.

    Article  CAS  PubMed  Google Scholar 

  217. Dias TR, Alves MG, Silva J, Barros A, Sousa M, Casal S, Silva BM, Oliveira PF. Implications of epigallocatechin-3-gallate in cultured human Sertoli cells glycolytic and oxidative profile. Toxicol In Vitro. 2017;41:214–22.

    Article  CAS  PubMed  Google Scholar 

  218. Bernardino RL, Carrageta DF, Sousa M, Alves MG, Oliveira PF. pH and male fertility: making sense on pH homeodynamics throughout the male reproductive tract. Cell Mol Life Sci. 2019;76:3783–800.

    Article  CAS  PubMed  Google Scholar 

  219. Dias TR, Bernardino RL, Alves MG, Silva J, Barros A, Sousa M, Casal S, Silva BM, Oliveira PF. L-Theanine promotes cultured human Sertoli cells proliferation and modulates glucose metabolism. Eur J Nutr. 2019;58:2961–70.

    Article  CAS  PubMed  Google Scholar 

  220. Dias TR, Alves MG, Almeida SP, Silva J, Barros A, Sousa M, Silva BM, Silvestre SM, Oliveira PF. Dehydroepiandrosterone and 7-oxo-dehydroepiandrosterone in male reproductive health: Implications of differential regulation of human Sertoli cells metabolic profile. J Steroid Biochem Mol Biol. 2015;154:1–11.

    Article  CAS  PubMed  Google Scholar 

  221. Alves MG, Martins AD, Moreira PI, Carvalho RA, Sousa M, Barros A, Silva J, Pinto S, Simões T, Oliveira PF. Metabolic fingerprints in testicular biopsies from type 1 diabetic patients. Cell Tissue Res. 2015;362:431–40.

    Article  CAS  PubMed  Google Scholar 

  222. Oliveira PF, Alves MG, Rato L, Laurentino S, Silva J, Sá R, Barros A, Sousa M, Carvalho RA, Cavaco JE, Socorro S. Effect of insulin deprivation on metabolism and metabolism-associated gene transcript levels of in vitro cultured human Sertoli cells. Biochim Biophys Acta. 2012;1820:84–9.

    Article  CAS  PubMed  Google Scholar 

  223. Martins AD, Sá R, Monteiro MP, Barros A, Sousa M, Carvalho RA, Silva BM, Oliveira PF, Alves MG. Ghrelin acts as energy status sensor of male reproduction by modulating Sertoli cells glycolytic metabolism and mitochondrial bioenergetics. Mol Cell Endocrinol. 2016;434:199–209.

    Article  CAS  PubMed  Google Scholar 

  224. Alves MG, Jesus TT, Sousa M, Goldberg E, Silva BM, Oliveira PF. Male fertility and obesity: are ghrelin, leptin and glucagon-like peptide-1 pharmacologically relevant? Curr Pharm Des. 2016;22:783–91.

    Article  CAS  PubMed  Google Scholar 

  225. Alves MG, Martins AD, Jarak I, Barros A, Silva J, Sousa M, Oliveira PF. Testicular lactate content is compromised in men with Klinefelter Syndrome. Mol Reprod Dev. 2016;83:208–16.

    Article  CAS  PubMed  Google Scholar 

  226. Hermann BP, Cheng K, Singh A, Roa-De La Cruz L, Mutoji KN, Chen IC, Gildersleeve H, Lehle JD, Mayo M, Westernströer B, et al. The Mammalian Spermatogenesis Single-Cell Transcriptome, from Spermatogonial Stem Cells to Spermatids. Cell Rep. 2018;25:1650-1667.e1658.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Bhattacharya I, Basu S, Pradhan BS, Sarkar H, Nagarajan P, Majumdar SS. Testosterone augments FSH signaling by upregulating the expression and activity of FSH-Receptor in Pubertal Primate Sertoli cells. Mol Cell Endocrinol. 2019;482:70–80.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors want to show their appreciation to all members of the Sperm Laboratory in Zhejiang University for their support, encouragement, and assistance.

Funding

This research was funded by Natural Science Foundation of Zhejiang Province, China (No. LY20H040010) and National Natural Science Foundation of China (No. 32072954).

Author information

Authors and Affiliations

Authors

Contributions

J.-M. Wang conceived, wrote, revised the manuscript, and prepared all the figures; Z.-F. Li wrote part of the manuscript and participated in discussion; W.-X. Yang was involved in constructing the frame of the manuscript, participated in the revision and supplied the fund for this research; F.-Q. Tan participated in discussion and revision, and provides the funding of the publication of this work. The author(s) read and approved the final manuscript. 

Corresponding authors

Correspondence to Wan-Xi Yang or Fu-Qing Tan.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, JM., Li, ZF., Yang, WX. et al. Follicle-stimulating hormone signaling in Sertoli cells: a licence to the early stages of spermatogenesis. Reprod Biol Endocrinol 20, 97 (2022). https://doi.org/10.1186/s12958-022-00971-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12958-022-00971-w

Keywords