Skip to main content

Microenvironment of spermatogonial stem cells: a key factor in the regulation of spermatogenesis

Abstract

Spermatogonial stem cells (SSCs) play a crucial role in the male reproductive system, responsible for maintaining continuous spermatogenesis. The microenvironment or niche of SSCs is a key factor in regulating their self-renewal, differentiation and spermatogenesis. This microenvironment consists of multiple cell types, extracellular matrix, growth factors, hormones and other molecular signals that interact to form a complex regulatory network. This review aims to provide an overview of the main components of the SSCs microenvironment, explore how they regulate the fate decisions of SSCs, and discuss the potential impact of microenvironmental abnormalities on male reproductive health.

Graphical Abstract

Introduction

Definition and significance of SSCs

Normal spermatogenesis in mammals is fundamental to the maintenance of male fertility, and spermatogenesis is a complex and highly efficient process involving a series of mitotic, meiotic, and morphological changes in the cells [1]. Spermatogenesis is the process of formation of mature spermatozoa in the testis, which is divided into three main stages: mitosis of spermatogonia, two meiotic divisions of spermatocytes, and metamorphosis and maturation of spermatocytes [1]. The classification of spermatogonia in mammals is different in primates and rodents [2] (as shown in Fig. 1). In primates, there are two main types of spermatogonia, type A and type B. Type A spermatogonia are stem cells with the ability to self-renewal and proliferate, and type A spermatogonia are subdivided into Adark and Apale cells. Type B spermatogonia are the precursor cells that differentiate into spermatocytes, and they undergo a series of cell divisions and differentiation processes that lead to the formation of mature spermatids [3, 4]. Adark spermatogonia are usually considered true SSCs, while Apale spermatogonia proliferate and differentiate as male germline progenitors into further germ cells [2, 5]. In contrast, spermatogonia in rodents are mainly classified into Asingle, Apaired, and Aaligned types. Aaligned spermatogonia differentiate into type A1 spermatogonia, which subsequently differentiate into type A1–A4, intermediate, and type B spermatogonia. Asingle spermatogonia are considered the sole spermatogonial type with a self-renewal capacity [6]. The balance and transformation between type A and type B spermatogonia are essential for the maintenance of normal spermatogenesis. Through the self-renewal and differentiation of SSCs, the male can continuously produce new spermatozoa and maintain the normal function of the reproductive system. Primary spermatocytes are cells formed by mitosis of spermatogonia. Primary spermatocytes undergo meiosis I to form two secondary spermatocytes, the two secondary spermatocytes undergo meiosis II to form four round spermatids, and the round spermatids undergo a series of morphological and functional changes to form mature spermatozoa with complete structures and functions [1]. This process is regulated by various factors, including the influence of hormones, growth factors and cytokines, among which neurotrophic factors such as Glial cell line-derived neurotrophic factor (GDNF) play an important role in regulating spermatogenesis [7].

Fig. 1
figure 1

Schematic diagram of testicular spermatogenesis. The formation of spermatogonia differs between mouse and human spermatogonia, with type A spermatogonia in mice mainly divided into Asingle, Aparied, Aaligned, and A1–A4 spermatogonia; whereas human type A spermatogonia are divided into Adark and Apale spermatogonia

SSCs are a key cell type in the male reproductive system and they play a crucial role in reproduction and fertility maintenance. Firstly, studying the molecular regulatory network of SSCs helps us understand the molecular mechanisms of male infertility and may identify new targets for its treatment [8]. Secondly, SSCs are the only adult stem cells that can transmit genetic information to offspring. Therefore, gene editing of SSCs may offer new possibilities for treating genetic disorders [9]. By gene editing SSCs in vitro (e.g. using CRISPR/Cas9 technology), defective genes can be repaired or replaced. Then these repaired cells can be transplanted back into the body, making it possible to treat genetic disorders in germline cells [10]. Thirdly, cryopreservation and transplantation of SSCs can be used to preserve fertility in childhood cancer patients [11], where SSCs are extracted from the testes of cancer patients for cryopreservation, and thawed SSCs can be used for in vitro culture to become functional spermatozoa to be used for in vitro fertilization (IVF). Fourthly, the in vitro differentiation potential of SSCs offers potential applications in regenerative medicine. In particular, SSCs could potentially be used to repair or replace damaged tissues and organs in treating neurodegenerative diseases or heart disease [12]. Multipotent SSCs (mSSCs), like embryonic stem cells (ESCs), have self-renewal and differentiation properties and are capable of differentiating into vascular endothelial cells (VECs) and smooth muscle cells in vitro [13]. SSCs, as germ cell initiating cells, are the cornerstone of male reproductive health, and they maintain sperm production through self-renewal and differentiation. In turn, the proper performance of these functions of SSCs relies on their stable microenvironment, a complex network of multiple biological signals that ensures the correct behavior of SSCs. Understanding the composition and function of this microenvironment is important for unravelling the underlying mechanisms of male infertility and developing new therapeutic approaches.

SSCs microenvironment

The concept of the "niche" was first proposed in the context of hematopoietic stem cells (HSCs) [14]. Niche refers to the specific microenvironment or cell population in which stem cells are found in a tissue or organ, as well as the niche-specific ECM [15]. The niche of SSCs, also known as the reproductive ecosystem or the microenvironment within the seminiferous tubules, is a complex system comprising a variety of cell types and molecular signals that work together to maintain the self-renewal and differentiation of SSCs. The microenvironment of SSCs consists mainly of cellular fractions of Sertoli cells, Leydig cells, peritubular myoid cells (PMCs), macrophages, and VECs, as well as the ECM [16, 17].

Sertoli Cells are one of the most important cell types in the SSCs niche [18]. They bind tightly to SSCs, provide physical support and secrete a variety of growth factors and cytokines, such as GDNF, fibroblast growth factor 2 (FGF2), and colony stimulating factor 1 (CSF1), which are essential for the self-renewal and differentiation of SSCs [19, 20]. Leydig cells are mainly responsible for the production of testosterone, an important sex hormone. Testosterone has a major influence on the differentiation of SSCs and the entire spermatogenesis process [21]. PMCs surround the seminiferous tubules and not only provide structural support but may also regulate the function of SSCs by secreting growth factors or cytokines [22]. Macrophages play an immunosurveillance role in the niche of SSCs and may influence the functions of SSCs by secreting cytokines [23]. VECs constitute a vascular network within the testis that provides oxygen and nutrients to SSCs, and VECs can act as feeder layer cells to maintain the proliferative and self-renewal capacity of SSCs while retaining their stemness properties [24]. Together, these cell types form a complex network that interacts through direct signaling exchanges and works together to maintain the homeostasis and proper function of SSCs [25]. Clinical studies suggest that idiopathic male infertility may be associated with disturbed signaling communication in the testicular microenvironment [26]. Imbalances in the microenvironment may lead to abnormalities in spermatogenesis, thereby affecting male fertility. Therefore, understanding how these cells interact is important for revealing the regulatory mechanisms of SSCs and developing new strategies for treating male infertility.

Impact of the microenvironment on SSCs

Influence of microenvironmental cellular fractions on SSC

The role of Sertoli cells

During spermatogenesis, Sertoli cells provide structural support, form the blood-testis barrier (BTB), supply nutrients, and regulate metabolism. They also participate in complex intercellular communication through secreted signaling molecules, which are essential for maintaining the stem cell state of SSCs and regulating their differentiation into mature spermatozoa [18]. At the same time, Sertoli cells can also participate in maintaining the number of Leydig cells with the normal functioning of PMCs [27]. Here, we focus on the function of signals secreted by Sertoli cells in regulating SSC self-renewal as well as differentiation through paracrine manners (Table 1).

Table 1 Sertoli cells regulate SSC self-renewal and differentiation

Sertoli cells regulate the self-renewal of SSCs


GDNF


Signaling molecules such as GDNF, FGF, Ets variant 5 (ETV5/ERM), WNT, NOTCH, etc., which are expressed in Sertoli cells, play important roles in the regulation of SSC self-renewal (e.g., Fig. 2). GDNF is an important neurotrophic factor with protective and survival-promoting effects on many types of neurons [28]. In 2000, Meng et al. demonstrated that GDNF is involved in the self-renewal of the SSC [29]. In mammalian testis, GDNF is mainly secreted by Sertoli cells but is also produced by PMCs stimulated by testosterone [22]. GDNF binds to its high-affinity receptor, GDNF family receptor alpha (GFRα). This binding is typically followed by an interaction with the tyrosine kinase receptor RET, which initiates downstream signaling pathways. The GDNF-RET-GFRα signaling axis is essential for the self-renewal of SSCs [19, 30, 31]. Additionally, GDNF can bind to GFRα independently of RET to mediate signaling [32]. In SSCs, GDNF signaling secreted by Sertoli cells is particularly important for maintaining the self-renewal of SSCs and preventing their premature differentiation. The testes of Gdnf-deficient mice are morphologically normal until adulthood, whereas the testes of older Gdnf ± mice have only Sertoli cells and no spermatogonia, and show the Sertoli cell-only syndrome (SCOS) phenotype [29]. The high expression of Gdnf inhibits spermatogonia differentiation, leading to the overproliferation of undifferentiated spermatogonia, leading to seminoma [33]. This result was also validated in human testicular tissues, where mRNA and protein expression of GDNF was significantly lower in SCOS tissues than in normal [34]. Restoration of GDNF expression in mice deficient in GDNF signaling and a small amount of SSCs in the tubules can reconstitute the SSC pool [35], we hope that this research will enable SCOS patients to bear their own children.

Fig. 2
figure 2

Regulatory Network of SSC Self-Renewal by Sertoli Cells. Sertoli cells regulate SSC self-renewal through key signaling pathways such as GDNF, which is essential for SSC maintenance, and FGF2, which synergizes with GDNF to enhance SSC proliferation. ETV5 aids in the transcriptional regulation of SSC self-renewal genes. Other signaling molecules and pathways also contribute to maintaining the balance between SSC self-renewal and differentiation

GDNF is a key factor in the maintenance of self-renewal of SSCs, and multiple signaling molecules in Sertoli cells mediate SSC self-renewal by regulating GDNF expression [7]. Thus, GDNF and its receptor system play a central role in regulating the fate of SSCs and spermatogenesis. Both Sertoli cells-derived GDNF and recombinant GDNF induce spermatogonia migration in vitro, and GDNF-induced migration is dependent on the expression of the GDNF co-receptor GFRα1 and upregulates the actin regulatory protein vasodilator-stimulated phosphoprotein (VASP) in a dose-dependent manner, which in turn affects cell migration [36]. Similarly, macrophage migration inhibitory factor (MIF) is expressed in Sertoli cells, whereas its receptor MIFR is expressed on spermatogonia, Leydig cells and Sertoli cells. MIF is involved in spermatogonia migration and MIF was found to synergize with GDNF to increase spermatogonia migration [37]. Many signaling molecules in Sertoli cells are involved in the maintenance of SSC self-renewal through the regulation of GDNF and thus the maintenance of SSCs. Cell division control protein 42 (Cdc42) is a major regulator of the actin cytoskeleton [38], which maintains the SSC pool by mitogen-activated protein kinase 1/3 (MAPK1/3)-dependent secretion of GDNF, and Sertoli cells-specific knockout of Cdc42 downregulates GDNF, leading to loss of SSCs [39]. AT-rich interaction domain 4B (ARID4B) is involved in maintaining the balance between self-renewal capacity and differentiation of HSCs [40]. SSCs from Sertoli cell-specific Arid4b knockout mice cannot maintain self-renewal, and the Arid4b knockout leads to a reduced expression of GDNF, which in turn affects the self-renewal of SSCs [41]. Follicle-stimulating hormone (FSH) has been shown to regulate GDNF expression, and FSH stimulates the expression of nuclear orphan receptor Nur77 in Sertoli cells during prenatal and early postnatal, whereas Nur77 binds to the promoter of GDNF to promote GDNF expression. The FSH/Nur77/GDNF pathway affects SSC proliferation [42].


FGF2


In addition to GDNF, which primarily regulates SSC self-renewal, researchers have identified the FGF2 signaling pathway as perhaps a novel mode of self-renewal regulation independent of GDNF [43]. FGF2 is expressed in Sertoli cells in mouse testis [44], and its receptor, fibroblast growth factor receptor 3 (FGFR3), is expressed in A-type spermatogonia [45]. Without GDNF signaling, the expansion of SSC cultured in vitro can be maintained by adding FGF2 [43]. In the presence of testosterone, FGF2 secreted by Sertoli cells causes activation of the Extracellular signal-regulated kinase (ERK) signaling pathway by interacting with receptors in germ cells, where the transcription factor cAMP-response element binding protein (CREB) binds to the 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 4 (PFKFB4) promoter and initiates transcription [46]. Previous studies indicate that FGF2 regulates the self-renewal of SSCs. However, recent research suggests that FGF2 might also promote spermatogonia differentiation. Additionally, FGF2 has been found to inhibit GDNF production and retinoic acid (RA) metabolism, potentially favoring differentiation. Therefore, the precise spatiotemporal role of FGF2 warrants further investigation [47]. FGF2 also stimulates the expression of ETV5 along with GDNF in Sertoli cells through the MAPK and phosphatidylinositol-4,5-bisphosphate 3-kinase (PI3K) signaling cascade response, but unlike GDNF, the expression of ETV5 is increased in response to epidermal growth factor (EGF) [48]. The transcription factor Etv5 is expressed in Sertoli cells, and loss of Etv5 prevents SSCs from maintaining self-renewal but does not affect the differentiation process. This ultimately results in progressive germ cell loss, leading to SCOS [49]. Furthermore, Etv5 does not affect the maintenance of SSCs through the modulation of GDNF and consequently SSCs [50]. In addition, microtubule-associated serine/threonine kinase 4 (MAST4) in Sertoli cells has been found to regulate SSC self-renewal via the FGF2-ETV5 pathway, and Mast4 knockout testes have a similar phenotype to SCOS, with germ cell loss [51]. Further exploration revealed that MAST4 was able to regulate the cell cycle by inducing cyclin-dependent kinase 2 (CDK2) to activate promyelocytic leukemia zinc-finger (PLZF), and the activated PLZF inhibited the transcription of p21, p53, and cyclin-A2 (Ccna2) to keep SSCs in the stem cell state [52]. Unlike Fgf2, Fgf9 is predominantly expressed in mouse Leydig cells, and Fgf9 promotes the proliferation of SSCs [53]. By activating its signaling cascade within the SSCs, it induces the phosphorylation of p38 MAPK, upregulates Etv5 and increases B-cell CLL/lymphoma 6 member B protein (Bcl6b) expression, ultimately promoting SSC proliferation [54]. Berberine (BBR), an antioxidant and anti-inflammatory agent, enhances the self-renewal capacity of SSCs by strengthening the reduced Sertoli cells and Leydig cells in the testis, which in turn upregulates the expression of Etv5, Gdnf, and Bcl6b [55], perhaps BBR could be applied in the future long-term in vitro culture of SSCs. Phosphoglycerate mutase 1 (Pgam1) in Sertoli cells increases the expression of Fgf2 and Gdnf through the glycolytic pathway, which in turn promotes self-renewal of SSCs [56]. Thus, FGF2 plays an important role in the balance of SSC self-renewal and differentiation.


GATA4


GATA Binding Protein 4 (GATA4) is also a key regulator in Sertoli cells involved in stem cell pool maintenance. Rubicon, a negative regulator of autophagy in Sertoli cells, maintains the normal function of Sertoli cells by preventing autophagic degradation of the transcription factor Gata4, thereby maintaining the normal function of the SSC pool [57]. Silencing of Gata4 in Sertoli cells affects germ cell development, leading to loss of the SSC pool and also allowing increased permeability of the BTB [58, 59]. GATA4 is also expressed in Leydig cells, and defects in Gata4 in Leydig cells lead to reduced expression of genes related to androgen synthesis and cause an increase in apoptotic cells, suggesting that Gata4 also acts as a critical transcription factor in Leydig cells [60].


WNT, NOTCH


The WNT and NOTCH signaling pathways in Sertoli cells are involved in regulating SSC self-renewal. The specific activation of beta-catenin 1 (CTNNB1) in Sertoli cells reduces the activity of endogenous SSCs through Wnt4 signaling, which results in germ cell apoptosis and failure of spermatogenesis [61]. Sertoli cells-secreted Wnt5a maintains SSC self-renewal by promoting SSC survival [62], whereas luteinizing hormone (LH) negatively regulates SSC self-renewal by inhibiting Wnt5a expression via testosterone [63]. The cell-specific secretion of Wnt6 supports the activation of Wnt/β-catenin signaling and the expression of the Wnt target gene Axin2, which promotes SSC proliferation but does not affect SSC self-renewal [64]. Deletion of Collaborator of ARF (Carf) in Sertoli cells downregulates Wnt signaling and Gdnf expression, which in turn impairs self-renewal of SSCs and proliferation of undifferentiated spermatogonia, leading to SCOS [65]. Wilms' tumor 1 (WT1) also regulates spermatogenesis through the WNT signaling pathway. WT1 is specifically expressed in Sertoli cells. When Wt1 is defective, undifferentiated spermatogonia accumulates in the testis, and germ cells fail to undergo normal meiosis [66]. This defect leads to a progressive loss of developing germ cells and Sertoli cells, ultimately destroying the spermatogenic tubule structure [67]. Defects in Wt1 result in decreased expression of Wnt4 and Wnt11. This downregulation of Wnt4 subsequently reduces the expression of cell polarity-related genes, such as Par6b and E-cadherin, which in turn disrupts spermatogenesis [68]. The heterogeneous nuclear ribonucleoprotein U (HNRNPU) in Sertoli cells interacts with WT1 and SOX9 and binds directly to the promoter regions of Sox8 and Sox9, which in turn regulates the development of both Sertoli cells and germ cells. Deficiency in Hnrnpu impairs spermatogonia proliferation and differentiation [69]. Leydig cells can be classified as adult Leydig cells (ALCs) and fetal Leydig cells (FLCs) [70, 71]. Deletion of Wt1 downregulates NOTCH signaling via a paracrine pathway, disrupting the differentiation of FLCs from PMCs. Additionally, Wt1 may play a role in maintaining the balance between FLCs and ALCs differentiation [72]. Deletion of Wilms tumor 1-associated protein (Wtap) in Sertoli cells can lead to defects in SSC self-renewal and proliferation, leading to a depletion of the SSC pool. However, the absence of Wtap does not affect meiosis [73]. In contrast to the WNT signaling pathway, NOTCH signaling in Sertoli cells acts as a negative regulator, modulating GDNF and controlling SSC self-renewal [74]. Overexpression of NOTCH signaling can lead to premature differentiation and progressive apoptosis of germ cells [75]. Transcription factor Rbpj is a key factor of NOTCH signaling, and the absence of Rbpj in Sertoli cells does not affect the development of the Sertoli cells themselves, but leads to an increase in the number of SSCs and enhances the expression of key factors that maintain undifferentiated spermatogonia [76]. This indicates that NOTCH signaling plays a crucial role in maintaining the balance between SSC self-renewal and differentiation.


CXCL12/CXCR4


The CXC motif chemokine ligand-12/CXC motif chemokine receptor-4 (CXCL12/CXCR4) signaling regulates SSC migration and ensures they settle in specific locations within the seminiferous tubules to maintain their stem cell properties. Both CXCL12 and CXCR4 are crucial for maintaining the SSC pool. CXCL12 is expressed by postnatal mouse testicular Sertoli cells, while CXCR4 is expressed by undifferentiated spermatogonia [77]. Overexpression of Cxcl12 in Sertoli cells increased SSC colonization in vivo [78]. Sin3a is required for the establishment and maintenance of germline stem cells. The absence of Sin3a in Sertoli cells affects the expression of stem cell markers and results in the absence of Cxcl12/Cxcr4 for the maintenance of undifferentiated spermatogonia [77]. Consequently, this inhibition of Cxcl12/Cxcr4 leads to a loss of SSCs [79]. The expression of CXCL12 in Sertoli cells and CXCR4 in germ cells has also been demonstrated in the adult testis [80]. However, recent single-cell RNA sequencing revealed that in human testis, CXCL12 is mainly expressed in Leydig cells, whereas CXCR4 is expressed in both macrophages and spermatogonia [17]. The interaction network of CXCR12 and CXCR4 needs to be further investigated. ASK-1 interacting protein (Aip1) is also involved in regulating germ cell migration, and Aip1 deficiency in Sertoli cells reduces SSC self-renewal but increases differentiated spermatogonia [81].


Other factors


Maintaining the SSC pool also depends on several essential paracrine factors secreted by Sertoli cells, including Insulin-like growth factor 1 (IGF1), Insulin Growth Factor Binding Protein 7 (IGFBP7), the Na+-K+-Cl-transporter isoform 1 (NKCC1), and protein-tyrosine phosphatase, non-receptor type 11 (PTPN11, also known as SHP2), all of which promote SSC stemness and proliferation [82,83,84]. IGF1 secreted by Sertoli cells, Leydig cells and PMCs stimulates mSSCs proliferation by promoting the G2/M phase of the cell cycle [85]. Previous studies have shown that combining IGF1 with GDNF enhances SSC activity in in vitro cultures [7]. And IGFBP7 secreted by Sertoli cells maintains SSC stemness by inhibiting the downstream PI3K signaling pathway through competition with IGF1 for binding to IGF1R [82]. NKCC1 is expressed in both germ cells and Sertoli cells, and deletion of NKCC1 leads to meiosis and spermatogenesis abnormalities [83]. Further investigation of its role in Sertoli cells revealed that NKCC1 is a downstream target gene of oxidative stress-responsive kinase 1 (OSR1) and SPS1-related proline/alanine-rich kinase (SPAK). Double knockout of Osr1 and Spak resulted in a decrease in Nkcc1 activity in Sertoli cells, resulting in loss of germ cells and finally a SCOS phenotype [86]. PTPN11 in Sertoli cells is crucial for balancing SSC self-renewal and differentiation. Although the absence of Ptpn11 does not affect Sertoli cell development, it leads to premature SSC differentiation, depleting the SSC pool, and ultimately resulting in the absence of germ cells in the seminiferous tubules [84]. PTPN11 has been reported to be expressed in Leydig cells and to participate in testosterone production. However, its potential role in regulating spermatogenesis warrants further investigation [87, 88].


Sertoli cells regulate the differentiation of SSCs


Sertoli cells are involved not only in regulating SSC self-renewal but also in the differentiation of SSCs by secreting a variety of growth factors and cytokines (as shown in Fig. 3).

Fig. 3
figure 3

Regulatory network of SSC differentiation by Sertoli cells. VEGFA165B regulates SSC differentiation by binding to VEGFR, and both SCF and KITL are involved in SSC differentiation by binding to c-kit. Also LRH1, WT1, Cx43, Erbb4/Nrg1, and Rdh10 regulate SSC differentiation

Retinoic acid (RA) has an important effect on SSC differentiation in the reproductive system. RA is considered one of the important signaling molecules to promote the differentiation of spermatogonia into spermatozoa [89]. It has been shown that RA induces SSCs into a differentiated state and promotes the progression of spermatogonia to the spermatocyte lineage [90]. RA affects the differentiation process of spermatogonia by binding to the retinoic acid receptor (RAR), regulating the expression of downstream genes and promoting spermatogonia differentiation and maturation [91]. Retinol dehydrogenase 10 (RDH10) is essential for RA biosynthesis [92]. The absence of Rdh10 in the Sertoli cells causes a defect in spermatogonia differentiation, resulting in only Sertoli cells with undifferentiated spermatogonia in the seminiferous tubules [93].

The KIT ligand (KITL)-KIT signaling pathway is essential for spermatogenesis, and KITL is expressed in Sertoli cells, and its receptor c-kit is expressed in spermatogonia and Leydig cells [94]. Kitl mutant mice exhibit impaired sperm production. When testicular tissues from these mutant mice were cultured with recombinant KITL and CSF1, it was observed that CSF1 enhanced the effects of KITL and promoted spermatogenesis [95]. Further exploration revealed that low concentrations of KITL induced the differentiation of stem Leydig cells (SLCs), while high concentrations of KITL inhibited the differentiation of SLCs. Thus, KITL is an essential growth factor that regulates the development of SLCs [96]. Stem cell factor (SCF) also binds to c-kit to regulate spermatogonia differentiation and apoptosis. While SCF is broadly expressed in various testicular cell types, the absence of Scf specifically in Sertoli cells impairs spermatogonia differentiation, leading to male infertility. The lack of SCF in Sertoli cells results in the loss of most c-kit+ spermatogonia and subsequent germ cells, underscoring the critical role of Sertoli cell-specific SCF production in spermatogenesis [97]. In vitro experiments demonstrated that β-estradiol induced an increase in the expression level of SCF in human fetal Sertoli cells, which in turn promoted the proliferation and inhibited the apoptosis of SSCs [98]. Neuregulin-1 (Nrg1) is expressed in Sertoli cells, and Nrg1Ser−/− leads to spermatogonia cell death and inhibits spermatogonia proliferation and differentiation into spermatocytes [99]. NRG1 and SCF may interact to promote the proliferation of type A spermatogonia, while the differentiation of spermatogonia into spermatocytes may depend on the NRG1/ERBB4 signaling pathway. Deletion of Erbb4 in Sertoli cells leads to apoptosis and impaired spermatogenesis. Additionally, dysregulation of ERBB4 affects the formation of the BTB, disrupts Sertoli and Leydig cell function, and ultimately compromises male fertility [100].

Connexin 43 (Cx43, also known as GJA1) is regulated in Sertoli cells by androgens through WT1 [101]. Knockout of Cx43 in Sertoli cells prevents spermatogenesis [102], leading to male sterility by impeding spermatogonia differentiation [103, 104]. Cx43 is a major gap junction protein in the BTB, and its knockout also disrupts the integrity of the BTB [105], and the overexpression of Cx43 rescues the function of the BTB and re-initiates meiosis [106]. Knockout of Cx43 in Sertoli cells adversely impacts both germ cell development and the integrity of the BTB, while also affecting the Sertoli cells’ maturation [107].

Vascular endothelial growth factor A (VEGFA) is a key molecule that is highly expressed on Sertoli cells to regulate the balance between SSC self-renewal and differentiation [108], and multiple isoforms of VEGFA act by binding to vascular endothelial growth factor receptor (VEGFR) expressed on SSCs. VEGFA164 promotes SSC self-renewal, while VEGFA165b promotes SSC differentiation [109, 110]. Liver receptor homologue-1 (Lrh1) is also expressed in Leydig cells, Sertoli cells and germ cells. Lrh1 deficiency in Leydig cells has no significant effect on spermatogenesis. However, Lrh1 deficiency in Sertoli cells results in the dysfunction of Sertoli cells, and unable to maintain spermatogenesis, and Lrh1 in germ cells is mainly involved in the maintenance and differentiation of SSCs [111]. The lack of RING finger protein 20 (Rnf20) in Sertoli cells leads to the loss of H2BK120ub, which further reduces the expression of the Claudin11 (Cldn11) in Sertoli cells, disrupting cell adhesion and leading to apoptosis of spermatogonia and spermatocytes, resulting in a SCOS phenotype [112].

The normal formation of the BTB is essential for spermatogenesis, and deletion of F-Box and WD Repeat Domain Containing 7 (Fbxw7) in Sertoli cells leads to excessive germ cell loss and spermatogenesis blockage, age-dependent tubular atrophy, and subsequent infertility in mice. Knockout of Fbxw7 in Sertoli cells disrupts the integrity of the BTB and causes reduced testosterone secretion by Leydig cells [113]. In contrast, in a previous study, Fbxw7 in SSCs was found to negatively regulate its self-renewal by degrading MYC [114]. Thus, Fbxw7 may play different roles in different cells, and the exact molecular mechanisms remain to be explored. Alpha-ketoglutarate-dependent dioxygenase alkB homolog 5 (ALKBH5) is expressed in Leydig cells, Sertoli cells and germ cells [115], and depletion of Alkbh5 in Sertoli cells affects the stability of N- cadherin (Cdh2) mRNA stability which disrupts BTB integrity [116].

Contribution of Leydig cells

Leydig cells are endocrine cells located in the testicular interstitium that synthesize and secrete testosterone, which is essential for the development and function of the male reproductive system (as shown in Fig. 4). Leydig cells and Sertoli cells may originate from the same progenitor cells. Knockout of the Wt1 gene reprogrammed testicular Sertoli cells into Leydig cells [117]. Nes-GFP+ cells were identified as SLCs, and Nes-GFP+ cells transplanted into Leydig cells functionally damaged or senescent testes were able to increase testosterone and restore germ cell meiosis [118]. In our previous study, human testicular SLCs were isolated and cultured in vitro, and we found that WNT5A promotes proliferation, DNA synthesis and stemness maintenance of human testicular SLCs through activation of the β-catenin signaling pathway [119]. Future studies may investigate whether WNT5A in SLCs affects human SSC function through co-culture. Leydig cells indirectly influence the function of SSCs and the spermatogenesis process mainly through secreted testosterone, which is secreted by ALCs originating from the proliferation and differentiation of SLCs [120]. Testosterone is essential for promoting the differentiation of SSCs to mature spermatocytes during spermatogenesis. It affects the microenvironment of SSCs through endocrine pathways, thereby regulating the proliferation and differentiation of SSCs. Testosterone can affect the function of Sertoli cells, including the levels of growth factors and cytokines they secrete, which in turn influence the behavior of SSCs. Specifically, Testosterone regulates the levels of GDNF and FGF2 secreted by Sertoli cells, which are key factors in maintaining the self-renewal of SSCs [121, 122]. Testosterone regulates the pituitary–gonadal axis through an endocrine feedback mechanism, affecting the release of gonadotropins, such as FSH and LH, which further influence the proliferation and differentiation of SSCs [123].

Fig. 4
figure 4

Regulation of SSCs by Leydig cells, PMCs, VECs, and macrophage. Leydig cells regulate SSCs mainly by secreting CSF1, ESR2, and IGF1. PMCs affect SSCs mainly by secreting AR, GDNF, CSF1, and pnliprp2. VECs affect SSCs by secreting GDNF, and macrophages regulate SSCs by secreting CSF1 and NR2C2

Testosterone activates FGF2 expression in spermatogonia through an internal ribosome entry site (IRES)-dependent mechanism in the adult testis, while FSH rather than testosterone is used to support spermatogenesis in the fetal testis [122]. Casein kinase 1α (Ck1α), tumor necrosis factor induced protein 3 (Tnfaip3), and Nrg1 regulate testosterone secretion in Leydig cells. CK1α regulates testosterone synthesis through the LH/PKA/EGFR/ERK1/2 signaling pathway, and Ck1α deficiency in Leydig cells results in fewer germ cells, aberrant spermatogenesis and reduced male fertility [124]. Tnfaip3 inhibits apoptosis in Leydig cells and promotes testosterone production by upregulating CCAAT/enhancer-binding protein β (Cebpb) expression [125]. Nrg1 is an LH target gene, and knockout of Nrg1 in Leydig cells significantly reduced the proliferation of infant testicular Leydig cells, leading to reduced testosterone production and causing abnormal spermatogenesis in mutant mice, but deletion of Nrg1 did not affect SSCs development [126]. D-aspartate (D-asp) may further promote the proliferation of GC-1 cells by stimulating the secretion of estradiol/estrogen receptor β (Esr2) from Leydig cells [127]. Csf1 is mainly expressed in Leydig cells and PMCs in the testis [128]. The absence of Leydig cells leads to insufficient secretion of CSF1, which affects the self-renewal ability of SSCs, whereas melatonin reduces apoptosis of Leydig cells, increases CSF1 secretion, and ameliorates the reduction of spermatogenesis induced by diabetes [129]. Unlike the knockout of Nuclear Receptor Subfamily 5 Group A Member 1 (Nr5a1) in Sertoli cells, which does not impact male fertility, the deletion of Nr5a1 in Leydig cells affects Leydig cell development but also does not impair fertility. However, a double knockout of Nr5a1 in both Sertoli and Leydig cells leads to abnormal Sertoli cell localization, morphological abnormalities in Leydig cells, and degeneration of spermatogenic tubules [130]. Similar to CSF1, IGF1 secreted by Leydig cells activates the PI3K/Akt signaling pathway by binding to IGF1 receptors on SSCs, thereby promoting SSC self-renewal [131].

Functions of macrophages, PMCs, and TECs in the microenvironment of SSCs

Macrophages, PMCs and TECs play complex and important roles in the microenvironment of SSCs. They maintain the stability of the microenvironment, support the function of SSCs and participate in the regulation of the spermatogenic process through cytokine secretion and intercellular signaling.

Macrophages in the testis are divided into a peritubular macrophage population and an interstitial macrophage population, which play different roles due to different localization [132]. The peritubular macrophage population can stimulate the differentiation of spermatogonia, whereas the interstitial macrophage is mainly involved in intercellular communication with Leydig cells or other surrounding somatic cells [133, 134]. It has been shown that a short-term absence of peritubular macrophages leads to a block in spermatogonia differentiation without affecting the maintenance of the SSC pool [23]. Moreover, macrophages secrete CSF1 and RA, which suggests they may promote SSC self-renewal in a manner similar to Leydig cells through CSF1 secretion, and regulate SSC differentiation through RA secretion [23]. However, experimental validation is required to confirm these proposed functions. Nuclear receptor subfamily 2 group C member 2 (NR2C2) is expressed in macrophages in the testis, and NR2C2 promotes the expression of Interleukin 1β (IL-1β) and IL-6 in macrophages through the NF-κB signaling pathway, exerting a pro-inflammatory effect, which in turn inhibits the proliferation of mouse spermatogonia [135].

GDNF production by PMCs is essential for developing undifferentiated spermatogonia in vivo [22, 136], and GDNF production by PMCs is regulated by testosterone secreted by Leydig cells [137]. PMCs also express the androgen receptor (AR), and progressive loss of spermatogonia occurs after AR knockout in PMCs [138]. However, knockout of AR in Sertoli cells does not affect the number of spermatogonia and mainly affects meiosis in germ cells [139]. Overexpression of AR in Sertoli cells also affects Sertoli cell/germ cell homeostasis and male fertility [140]. Pancreatic lipase-related protein 2 (Pnliprp2) is specifically expressed in testicular PMCs, and Pnliprp2−/− leads to a disruption of homeostasis in undifferentiated spermatogonia that does not affect meiotic progression but results in a decrease in the number of SSCs [141]. Leucine-rich repeat-containing G protein-coupled receptor 4 (Lgr4) is specifically expressed in PMCs, and the absence of Lgr4 may affect the balance between self-renewal and differentiation of SSCs, leading to an increase in undifferentiated spermatogonia and a decrease in differentiated germ cells [142].

VECs are also vital components of the niche. Co-culturing VECs with mouse SSC promoted SSC self-renewal, and the addition of GDNF, FGF2, and VECs-specific secretion of IGFBP2, CXCL12, and macrophage inflammatory protein 2 (MIP2) was found to be sufficient for long-term culture of human and mouse SSCs in the absence of feeder layer cells [143]. Additionally, FGF2 binds to FGFR1 on VECs, thereby promoting the secretion of GDNF from these cells [143]. Then, researchers also demonstrated that VECs enhanced the self-renewal capacity and preserved the stemness of rat SSCs [24].

Role of the ECM

The process of self-renewal and differentiation of SSCs requires the regulation of different signals in the microenvironment. The signals provided by the microenvironment change to facilitate the transition of SSCs into mature sperm. For example, changes in the composition of the ECM may affect cell migration and differentiation. The ECM modulates gene expression patterns by influencing the interaction between chromatin and the nuclear matrix and affecting the association between the cytoskeleton and mRNAs [144]. ECM is continuously degraded, remodeled and synthesized by cells in each tissue, rendering it tissue-specific [145]. The ECM of SSCs is a complex network of laminin, fibronectin, collagen and proteoglycans [146,147,148,149], which provide structural support and biochemical signals to SSCs and are essential for the processes of SSC self-renewal, differentiation and spermatogenesis (Table 2).

Table 2 The influence of extracellular matrix on SSCs

Collagen is the most abundant protein in the ECM and provides the structural framework for spermatogenic tubules [150]. In the ECM of SSCs, collagens contribute to a three-dimensional support network that maintains tubule morphology and mechanical stability [151]. It has been suggested that collagen may not only provide structural support but may be involved in the formation of tight junctions [152]. Collagen has been found to promote SSCs localization in in vitro culture [153]. Collagen type I alpha 1 chain (Col1a1) is also involved in SSC self-renewal, and Col1a1 deficiency inhibits SSC self-renewal and promoted spermatogonia differentiation [154]. In an in vitro culture of porcine SSCs, collagen type IV (Col IV) promotes SSCs growth and differentiation [155]. Laminins are crucial structural components in the testis, serving as key basement membrane elements that bind to integrin receptors and other cell surface receptors and are involved in cell adhesion, differentiation, and migration [152, 156,157,158]. Laminin alpha2 (Lama2) deficiency leads to abnormalities of the testicular basement membrane, which causes male infertility and can be rescued by overexpression of laminin alpha1 (Lama1) [159]. In vitro culture of testicular tissue reveals that Lama1 is a key component in maintaining the niche of the SSCs, and the loss of germ cells may be associated with a reduction of Lama1 in the culture medium [160]. Laminin deficiency and altered expression patterns of all ECM proteins in the testes of SCOS patients [161] and dysregulation of ECM genes in infertile patients were also identified by single-cell RNA sequencing [162]. Proteoglycans are molecules consisting of a protein core and one or more glycosylated polysaccharide chains. They act as fillers in the ECM and are involved in water retention and regulation of cell signaling. In the ECM of SSCs, proteoglycans such as hyaluronic acid and chondroitin sulphate are essential for maintaining the physical properties of the microenvironment and regulating cell behavior. It was found that the use of cell scaffolds containing hyaluronic acid (HA) promoted the proliferation and differentiation of SSCs more than the 2D culture system, and that the combination of silicone nanoparticles (SNs) with HA promoted spermatogenesis [163]. Integrins are cell surface receptors crucial for the adhesion and localization of SSCs, with their activity and expression patterns being vital for these processes. The absence of β1-integrin in Sertoli cells impairs SSC colonization in vivo and decreases SSC homing. Additionally, integrin binding to adhesion molecules has been shown to co-regulate SSC homing [21, 164]. α9β1-integrin is expressed in porcine SSCs and is essential for sustaining SSC self-renewal in pigs. [165]. Exposure of rat testes to petrol exhaust leads to abnormal spermatogenesis. Petrol exhaust impairs SSC self-renewal by down-regulating α6-integrin and β1-integrin [166].

Extracellular vesicles (EVs) in the testis are also part of the ECM [167]. EVs include exosomes and microvesicles, small membrane structures released by cells into the extracellular environment, which have a significant effect on the microenvironment of SSCs [168]. These vesicles can contain a variety of biomolecules such as proteins, lipids, RNA and DNA, and testicular EVs can cross the BTB and play a crucial role in intercellular communication. Extracellular vesicles from different sources serve distinct functions. miR-486-5p in extracellular vesicles secreted by Sertoli cells can promote SSCs differentiation by interacting with PTEN [169]. Similarly, miR-493-5p in Sertoli cells-derived exosomes can inhibit GDNF expression, thereby inhibiting self-renewal and promoting differentiation [170]. However, inhibition of exosomes released from Sertoli cells reduces spermatogonia proliferation [171], and undifferentiated A spermatogonia-derived EVs are also able to inhibit the proliferation of SSCs in vitro [172]. Recent studies have found that palmitoylation of VMP1 affects the secretion of EVs from Sertoli cells and thus the maintenance of SSCs-like cells (SSCLCs). Disruption of VMP1 palmitoylation leads to a decrease in EV release, inhibiting the self-renewal and proliferation of SSCLCs and increasing the number of apoptotic cells [173]. Exosomes can cross the BTB, and exosomes secreted by rat Sertoli cells deliver Ccl20 mRNA to Leydig cells, which promotes Leydig cell survival via the Akt pathway [174]. Consequently, exosomes are pivotal mediators of intercellular signaling. Future research should explore the molecular mechanisms by which EVs facilitate cellular communication within the testicular microenvironment.

Together, these ECM components form a dynamic network that not only provides physical support but also participates in cell fate regulation. The ECM, with its diverse and complex composition, ensures that SSCs can effectively self-renew and differentiate within the appropriate microenvironment. The removal of testicular cells from decellularized testes, leaving only the ECM components, produces what is known as the decellularized testicular matrix (DTM) [175]. DTM promotes SSC proliferation and differentiation in the in vitro culture of SSCs [176]. Studies have shown that a serum-free, feeder layer-free system using ECM extracted from homologous tissue can support the culture of hSSCs for a period. However, further research is needed to optimize this in vitro culture system for hSSCs [177].

Correlation between microenvironmental abnormalities and male infertility

Microenvironmental abnormalities are strongly correlated with male infertility. In non-obstructive azoospermia (NOA) patients, abnormalities in Sertoli cells, Leydig cells, and peritubular myoid cells (PMCs) have been observed. Disruptions in intercellular communication are likely to be the molecular mechanisms underlying the development of NOA [178]. The homozygous + 48845 G > T (TT allele) variant of ETV5 has been identified as a genetic risk factor for developing SCOS [179]. Abnormal expression of various genes in Sertoli cells may lead to male infertility, and the expression of GDNF, FGF8, BMP4, FGF9 and CX43 were significantly reduced in SCOS patients, suggesting that the abnormal expression of these genes may be associated with the development of SCOS [180,181,182]. In KS and SCOS patients, ECM proteins are abnormally expressed and laminin expression is notably absent compared to normal controls [183]. In cryptorchidism, the testicular ECM is also disrupted, with decreased expression of Col IV and laminin [184]. Disruption of homeostasis in Sertoli cells may accelerate testicular senescence, as revealed by single-cell transcriptome analyses [185]. Whole exome sequencing of 529 patients with non-obstructive azoospermia (NOA) identified six missense mutations in WT1 gene. Knockout of Wt1 revealed that most seminiferous tubules contained only Sertoli cells, with extensive germ cell apoptosis. This suggests that WT1 mutations may be a potential cause of NOA in humans [68]. Gene chip and bioinformatics analyses revealed abnormal expression of numerous genes, including members of the G-protein-coupled receptors (GPCR) superfamily, in Sertoli cells from NOA patients [186].

Researchers have attempted to treat male infertility caused by microenvironmental abnormality. Since abnormal Leydig cell function leads to insufficient testosterone production, which impairs spermatogenesis and results in infertility, Leydig cell transplantation has been investigated as a means to restore testosterone levels and address the issue. In a recent study, it was found that transplantation of CXCR4-SF1 bifunctional adipose-derived stem cells (CXCR4-SF1-ADSCS) into BPA-induced Leydig cells injury model mice using tail vein injection was able to differentiate into Leydig-like cells, increasing the expression of testosterone and restoring the function of Leydig cells, rescuing the defects in reproductive function in the future. CXCR4-SF1-ADSCs are expected to treat male infertility caused by Leydig cells dysfunction in the microenvironment [187, 188]. Lhcgr is essential for the maturation of Leydig cells. Deletion of Lhcgr in Leydig cells impairs their function, and mutations in Lhcgr lead to hypoplasia of Leydig cells [189]. AAV8-Lhcgr treatment restores testosterone in the testis and significantly promotes the formation of round and elongated spermatozoa, increases sperm count and viability, and rescues spermatogenesis [190]. Currently, these findings are based on animal studies, but it is hoped that they will eventually translate into clinical applications, offering viable solutions for male infertility patients.

Directions for future research

Although several genes involved in the self-renewal and differentiation of SSCs have been identified, understanding the complex regulatory mechanisms within the testicular microenvironment remains incomplete. Key challenges include the difficulty in obtaining human testicular tissue, which is both scarce and requires ethical approval. Additionally, the testicular microenvironment is dynamically influenced by hormonal changes, affecting cellular states and interactions. While in vitro models can partially simulate this environment, they may not fully replicate its complexity. Consequently, current studies fall short of fully elucidating the regulatory networks of the testicular microenvironment.

With the development of sequencing technology, the signaling communication between somatic cells and germ cells can be revealed using single-cell RNA sequencing [191]. Single-cell RNA sequencing analysis reveals that classical signaling networks, such as GDNF-GFRa1/Ret and FGFs-FGFRs, are involved in the interaction between Sertoli cells and spermatogonia. Additionally, new signaling networks were identified, such as INHA produced by Sertoli cells, with its receptors,  ACVR2B and TGFBR3, being expressed in spermatogonia and Leydig cells, respectively [192]. The specific molecular mechanisms of the interactions need to be demonstrated by in vivo and in vitro experiments in future work. Existing studies on the molecular mechanisms of microenvironmental regulation of SSCs have relied on the following three approaches (e.g., Fig. 5). The first is in vivo experiments, in which somatic cells are infected in vivo by injecting specific lentiviruses infecting somatic cells or constructing knockout mice [54, 193, 194]. The second is in vitro transplantation experiments, in which Sertoli cells, Leydig cells, etc., are modified in vitro and then transplanted back into the mouse testis [195]. The third is somatic cells and SSCs co-culture system [78]. A new option in the co-culture system is to construct a three-dimensional scaffold in vitro to study the functions played by the niche components of SSCs and their molecular mechanisms. A study used a gel scaffold composed of hyaluronic acid (HA), chitosan, and decellularized testicular matrix (DTM) for in vitro SSCs culture, and this scaffold was found to support SSC differentiation and proliferation in vitro [196]. With the development of technology, many researchers are culturing testicular organoids, which may be used in subsequent studies to investigate the specific regulatory mechanisms of the microenvironment on SSC self-renewal and differentiation [197]. The construction and application of testicular organoids can not only study the signal communication between the testicular microenvironment and germ cells in infertile patients, but perhaps can also be applied to the recovery of spermatogenesis in infertile patients [198].

Fig. 5
figure 5

Four approaches to explore microenvironmental regulation of SSCs. The four research methods mainly include in vivo experiments (A), in vitro transplantation experiments (B), establishment of in vitro co-culture systems (C), and construction of testicular organoids (D)

The combination of testicular organoid culture technology and spatial proteomics to study the testicular microenvironment is a novel field of research that allows a deeper understanding of the complexity of the testicular microenvironment. Spatial proteomics is an emerging direction that can provide spatial information on proteins in tissue sections, revealing their localization and function in the testicular microenvironment [199]. Based on spatial proteomics, the composition and conditions of testicular organoid culture models can be optimized to more closely mimic the natural testicular microenvironment and improve the biological relevance of experimental results. Potential drugs or therapeutic regimens can be tested in organoid models to evaluate their effects on the testicular microenvironment and SSC function [200].

Conclusion

In this article, we detail the mechanisms by which Sertoli cells, Leydig cells, PMCs, macrophages, and VECs in the microenvironment regulate SSCs, summarize the role of ECM in the microenvironment such as collagen, laminin, and EVs on SSCs, and finally also explore the correlation between microenvironmental abnormalities and male infertility. The microenvironment of SSCs is a complex network, and normal signaling communication in the microenvironment is essential for the self-renewal and differentiation of SSCs. Future studies could further reveal the specific roles of the various components of this microenvironment and how they interact under different conditions through advanced sequencing techniques such as spatial proteomics, which will help develop new therapeutic strategies for male infertility.

Availability of data and materials

Not applicable.

Abbreviations

SSCs:

Spermatogonial stem cells

ECM:

Extracellular matrix

GDNF:

Glial cell line-derived neurotrophic factor

IVF:

In vitro fertilization

mSSCs:

Multipotent SSCs

ESCs:

Embryonic stem cells

VECs:

Vascular endothelial cells

HSCs:

Hematopoietic stem cells

CSF1:

Colony stimulating factor 1

BTB:

Blood-testis barrier

PMCs:

Peritubular myoid cells

GFRα:

GDNF family receptor α

SCOS:

Sertoli cell-only syndrome

VASP:

Vasodilator-stimulated phosphoprotein

MIF:

Macrophage migration inhibitory factor

Cdc42:

Cell division control protein 42

ARID4B:

AT-rich interaction domain 4B

FSH:

Follicle-stimulating hormone

FGF2:

Fibroblast growth factor 2

FGFR3:

Fibroblast growth factor receptor 3

ERK:

Extracellular signal-regulated kinase

CREB:

CAMP-response element binding protein

PFKFB4:

6-Phosphofructo-2-kinase/fructose-2,6-bisphosphatase 4

RA:

Retinoic acid

ETV5/ERM:

Ets variant 5

MAPK:

Mitogen-activated protein kinase

PI3K:

Phosphatidylinositol-4,5-bisphosphate 3-kinase

EGF:

Epidermal growth factor

MAST4:

Microtubule-associated serine/threonine kinase 4

CDK2:

Cyclin-dependent kinase 2

PLZF:

Promyelocytic leukemia zinc-finger

Ccna2:

Cyclin-A2

Bcl6b:

B-cell CLL/lymphoma 6 member B protein

BBR:

Berberine

Pgam1:

Phosphoglycerate mutase 1

GATA4:

GATA binding protein 4

CTNNB1:

Beta-catenin 1

LH:

Luteinizing hormone

Carf:

Collaborator of ARF

WT1:

Wilms' tumor 1

HNRNPU:

Heterogeneous nuclear ribonucleoprotein U

ALCs:

Adult Leydig cells

FLCs:

Fetal Leydig cells

Wtap:

Wilms' tumor 1-associated protein

CXCL12:

CXC motif chemokine ligand-12

CXCR4:

CXC motif chemokine receptor-4

Aip1:

ASK-1 interacting protein

IGFBP7:

Insulin growth factor binding protein 7

NKCC1:

Na+-K+-Cl-transporter isoform 1

PTPN11:

Protein-tyrosine phosphatase, nonreceptor type 11

IGF1:

Insulin-like growth factor 1

OSR1:

Oxidative stress-responsive kinase 1

SPAK:

SPS1-related proline/alanine-rich kinase

RAR:

Retinoic acid receptor

RDH10:

Retinol dehydrogenase 10

KITL:

KIT ligand

SLCs:

Stem Leydig cells

SCF:

Stem cell factor

Nrg1:

Neuregulin-1

ERBB4:

Erb-b2 receptor tyrosine kinase 4

Cx43:

Connexin 43

VEGFA:

Vascular endothelial growth factor A

VEGFR:

Vascular endothelial growth factor receptor

Lrh1:

Liver receptor homologue-1

Rnf20:

RING finger protein 20

Cldn11:

Claudin11

Fbxw7:

F-box and WD repeat domain containing 7

ALKBH5:

Alpha-ketoglutarate-dependent dioxygenase alkB homolog 5

Cdh2:

Cadherin 2

IRES:

Internal ribosome entry site

CK1α:

Casein kinase 1α

Tnfaip3:

Tumor necrosis factor induced protein 3

Cebpb:

CCAAT/enhancer-binding protein β

D-asp:

D-Aspartate

Esr2:

Estradiol/estrogen receptor β

Nr5a1:

Nuclear receptor subfamily 5 group A member 1

NR2C2:

Nuclear receptor subfamily 2 group C member 2

IL-1β:

Interleukin 1β

AR:

Androgen receptor

Pnliprp2:

Pancreatic lipase-related protein 2

Lgr4:

Leucine-rich repeat-containing G protein-coupled receptor 4

MIP2:

Macrophage inflammatory protein 2

Col1a1:

Collagen type I α1 chain

Lama2:

Laminin α2

Lama1:

Laminin α1

HA:

Hyaluronic acid

SNs:

Silicone nanoparticles

EVs:

Extracellular vesicles

SSCLCs:

SSCs-like cells

VMP1:

Vacuole membrane protein 1

DTM:

Decellularized testicular matrix

NOA:

Non-obstructive azoospermia

Col IV:

Collagen type IV

GPCR:

G-protein-coupled receptors

CXCR4-SF1-ADSCS:

CXCR4-SF1 bifunctional adipose-derived stem cells

INHA:

Inhibin subunit α

DTM:

Decellularized testicular matrix

References

  1. de Kretser DM, Loveland KL, Meinhardt A, Simorangkir D, Wreford N. Spermatogenesis. Hum Reprod. 1998;13(Suppl 1):1–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/13.suppl_1.1.

    Article  PubMed  Google Scholar 

  2. Hermann BP, Sukhwani M, Hansel MC, Orwig KE. Spermatogonial stem cells in higher primates: are there differences from those in rodents? Reproduction. 2010;139:479–93. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-09-0255.

    Article  CAS  PubMed  Google Scholar 

  3. Hermann BP, Sukhwani M, Simorangkir DR, Chu T, Plant TM, Orwig KE. Molecular dissection of the male germ cell lineage identifies putative spermatogonial stem cells in rhesus macaques. Hum Reprod. 2009;24:1704–16. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/dep073.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  4. Simorangkir DR, Marshall GR, Ehmcke J, Schlatt S, Plant TM. Prepubertal expansion of dark and pale type A spermatogonia in the rhesus monkey (Macaca mulatta) results from proliferation during infantile and juvenile development in a relatively gonadotropin independent manner. Biol Reprod. 2005;73:1109–15. https://doiorg.publicaciones.saludcastillayleon.es/10.1095/biolreprod.105.044404.

    Article  CAS  PubMed  Google Scholar 

  5. van Alphen MM, van de Kant HJ, de Rooij DG. Repopulation of the seminiferous epithelium of the rhesus monkey after X irradiation. Radiat Res. 1988;113:487–500. https://doiorg.publicaciones.saludcastillayleon.es/10.2307/3577245.

    Article  PubMed  Google Scholar 

  6. Ehmcke J, Wistuba J, Schlatt S. Spermatogonial stem cells: questions, models and perspectives. Hum Reprod Update. 2006;12:275–82. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humupd/dmk001.

    Article  CAS  PubMed  Google Scholar 

  7. Kubota H, Avarbock MR, Brinster RL. Growth factors essential for self-renewal and expansion of mouse spermatogonial stem cells. Proc Natl Acad Sci USA. 2004;101:16489–94. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.0407063101.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  8. Diao L, Turek PJ, John CM, Fang F, Reijo Pera RA. Roles of spermatogonial stem cells in spermatogenesis and fertility restoration. Front Endocrinol (Lausanne). 2022;13:895528. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2022.895528.

    Article  PubMed  Google Scholar 

  9. Mulder CL, Zheng Y, Jan SZ, Struijk RB, Repping S, Hamer G, et al. Spermatogonial stem cell autotransplantation and germline genomic editing: a future cure for spermatogenic failure and prevention of transmission of genomic diseases. Hum Reprod Update. 2016;22:561–73. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humupd/dmw017.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  10. Wu Y, Zhou H, Fan X, Zhang Y, Zhang M, Wang Y, et al. Correction of a genetic disease by CRISPR-Cas9-mediated gene editing in mouse spermatogonial stem cells. Cell Res. 2015;25:67–79. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/cr.2014.160.

    Article  CAS  PubMed  Google Scholar 

  11. Geens M, Goossens E, De Block G, Ning L, Van Saen D, Tournaye H. Autologous spermatogonial stem cell transplantation in man: current obstacles for a future clinical application. Hum Reprod Update. 2008;14:121–30. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humupd/dmm047.

    Article  PubMed  Google Scholar 

  12. Guan K, Wagner S, Unsold B, Maier LS, Kaiser D, Hemmerlein B, et al. Generation of functional cardiomyocytes from adult mouse spermatogonial stem cells. Circ Res. 2007;100:1615–25. https://doiorg.publicaciones.saludcastillayleon.es/10.1161/01.RES.0000269182.22798.d9.

    Article  CAS  PubMed  Google Scholar 

  13. Im JE, Song SH, Kim JY, Kim KL, Baek SH, Lee DR, et al. Vascular differentiation of multipotent spermatogonial stem cells derived from neonatal mouse testis. Exp Mol Med. 2012;44:303–9. https://doiorg.publicaciones.saludcastillayleon.es/10.3858/emm.2012.44.4.034.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  14. Schofield R. The relationship between the spleen colony-forming cell and the haemopoietic stem cell. Blood Cells. 1978;4:7–2.

    CAS  PubMed  Google Scholar 

  15. Chacon-Martinez CA, Koester J, Wickstrom SA. Signaling in the stem cell niche: regulating cell fate, function and plasticity. Development. 2018. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.165399.

    Article  PubMed  Google Scholar 

  16. Oatley JM, Brinster RL. The germline stem cell niche unit in mammalian testes. Physiol Rev. 2012;92:577–95. https://doiorg.publicaciones.saludcastillayleon.es/10.1152/physrev.00025.2011.

    Article  CAS  PubMed  Google Scholar 

  17. Guo J, Grow EJ, Mlcochova H, Maher GJ, Lindskog C, Nie X, et al. The adult human testis transcriptional cell atlas. Cell Res. 2018;28:1141–57. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41422-018-0099-2.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  18. O’Donnell L, Smith LB, Rebourcet D. Sertoli cells as key drivers of testis function. Semin Cell Dev Biol. 2022;121:2–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.semcdb.2021.06.016.

    Article  CAS  PubMed  Google Scholar 

  19. Parekh PA, Garcia TX, Hofmann MC. Regulation of GDNF expression in Sertoli cells. Reproduction. 2019;157:R95–107. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-18-0239.

    Article  CAS  PubMed  Google Scholar 

  20. de Rooij DG. The spermatogonial stem cell niche. Microsc Res Tech. 2009;72:580–5. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/jemt.20699.

    Article  CAS  PubMed  Google Scholar 

  21. Zhou R, Wu J, Liu B, Jiang Y, Chen W, Li J, et al. The roles and mechanisms of Leydig cells and myoid cells in regulating spermatogenesis. Cell Mol Life Sci. 2019;76:2681–95. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00018-019-03101-9.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  22. Chen LY, Willis WD, Eddy EM. Targeting the Gdnf Gene in peritubular myoid cells disrupts undifferentiated spermatogonial cell development. Proc Natl Acad Sci USA. 2016;113:1829–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1517994113.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  23. DeFalco T, Potter SJ, Williams AV, Waller B, Kan MJ, Capel B. Macrophages contribute to the spermatogonial niche in the adult testis. Cell Rep. 2015;12:1107–19. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.celrep.2015.07.015.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  24. Kim YH, Oh MG, Bhang DH, Kim BJ, Jung SE, Kim SM, et al. Testicular endothelial cells promote self-renewal of spermatogonial stem cells in ratsdagger. Biol Reprod. 2019;101:360–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/biolre/ioz105.

    Article  PubMed  Google Scholar 

  25. Shami AN, Zheng X, Munyoki SK, Ma Q, Manske GL, Green CD, et al. Single-cell RNA sequencing of human, macaque, and mouse testes uncovers conserved and divergent features of mammalian spermatogenesis. Dev Cell. 2020;54(529–547): e512. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.devcel.2020.05.010.

    Article  CAS  Google Scholar 

  26. Andersson AM, Jorgensen N, Frydelund-Larsen L, Rajpert-De Meyts E, Skakkebaek NE. Impaired Leydig cell function in infertile men: a study of 357 idiopathic infertile men and 318 proven fertile controls. J Clin Endocrinol Metab. 2004;89:3161–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/jc.2003-031786.

    Article  CAS  PubMed  Google Scholar 

  27. Rebourcet D, O’Shaughnessy PJ, Monteiro A, Milne L, Cruickshanks L, Jeffrey N, et al. Sertoli cells maintain Leydig cell number and peritubular myoid cell activity in the adult mouse testis. PLoS ONE. 2014;9: e105687. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0105687.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  28. Airaksinen MS, Saarma M. The GDNF family: signalling, biological functions and therapeutic value. Nat Rev Neurosci. 2002;3:383–94. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nrn812.

    Article  CAS  PubMed  Google Scholar 

  29. Meng X, Lindahl M, Hyvonen ME, Parvinen M, de Rooij DG, Hess MW, et al. Regulation of cell fate decision of undifferentiated spermatogonia by GDNF. Science. 2000;287:1489–93. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.287.5457.1489.

    Article  CAS  PubMed  Google Scholar 

  30. Naughton CK, Jain S, Strickland AM, Gupta A, Milbrandt J. Glial cell-line derived neurotrophic factor-mediated RET signaling regulates spermatogonial stem cell fate. Biol Reprod. 2006;74:314–21. https://doiorg.publicaciones.saludcastillayleon.es/10.1095/biolreprod.105.047365.

    Article  CAS  PubMed  Google Scholar 

  31. Jing S, Wen D, Yu Y, Holst PL, Luo Y, Fang M, et al. GDNF-induced activation of the ret protein tyrosine kinase is mediated by GDNFR-alpha, a novel receptor for GDNF. Cell. 1996;85:1113–24. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/s0092-8674(00)81311-2.

    Article  CAS  PubMed  Google Scholar 

  32. Saarma M, Sariola H. Other neurotrophic factors: glial cell line-derived neurotrophic factor (GDNF). Microsc Res Tech. 1999;45:292–302. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/(SICI)1097-0029(19990515/01)45:4/5%3c292::AID-JEMT13%3e3.0.CO;2-8.

    Article  CAS  PubMed  Google Scholar 

  33. Meng X, de Rooij DG, Westerdahl K, Saarma M, Sariola H. Promotion of seminomatous tumors by targeted overexpression of glial cell line-derived neurotrophic factor in mouse testis. Cancer Res. 2001;61:3267–71.

    CAS  PubMed  Google Scholar 

  34. Singh D, Paduch DA, Schlegel PN, Orwig KE, Mielnik A, Bolyakov A, et al. The production of glial cell line-derived neurotrophic factor by human sertoli cells is substantially reduced in sertoli cell-only testes. Hum Reprod. 2017;32:1108–17. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/dex061.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  35. Wright WW. The regulation of spermatogonial stem cells in an adult testis by glial cell line-derived neurotrophic factor. Front Endocrinol (Lausanne). 2022;13: 896390. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2022.896390.

    Article  PubMed  Google Scholar 

  36. Dovere L, Fera S, Grasso M, Lamberti D, Gargioli C, Muciaccia B, et al. The niche-derived glial cell line-derived neurotrophic factor (GDNF) induces migration of mouse spermatogonial stem/progenitor cells. PLoS ONE. 2013;8: e59431. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0059431.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  37. Huleihel M, Abofoul-Azab M, Abarbanel Y, Einav I, Levitas E, Lunenfeld E. Production of macrophage inhibitory factor (MIF) by primary sertoli cells; its possible involvement in migration of spermatogonial cells. J Cell Physiol. 2017;232:2869–77. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/jcp.25718.

    Article  CAS  PubMed  Google Scholar 

  38. Munemitsu S, Innis MA, Clark R, McCormick F, Ullrich A, Polakis P. Molecular cloning and expression of a G25K cDNA, the human homolog of the yeast cell cycle gene CDC42. Mol Cell Biol. 1990;10:5977–82. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.10.11.5977-5982.1990.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  39. Mori Y, Takashima S, Kanatsu-Shinohara M, Yi Z, Shinohara T. Cdc42 is required for male germline niche development in mice. Cell Rep. 2021;36: 109550. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.celrep.2021.109550.

    Article  CAS  PubMed  Google Scholar 

  40. Young IC, Wu B, Andricovich J, Chuang ST, Li R, Tzatsos A, et al. Differentiation of fetal hematopoietic stem cells requires ARID4B to restrict autocrine KITLG/KIT-Src signaling. Cell Rep. 2021;37: 110036. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.celrep.2021.110036.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  41. Wu RC, Zeng Y, Chen YF, Lanz RB, Wu MY. Temporal-spatial establishment of initial niche for the primary spermatogonial stem cell formation is determined by an ARID4B regulatory network. Stem Cells. 2017;35:1554–65. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/stem.2597.

    Article  CAS  PubMed  Google Scholar 

  42. Ding LJ, Yan GJ, Ge QY, Yu F, Zhao X, Diao ZY, et al. FSH acts on the proliferation of type A spermatogonia via Nur77 that increases GDNF expression in the Sertoli cells. FEBS Lett. 2011;585:2437–44. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.febslet.2011.06.013.

    Article  CAS  PubMed  Google Scholar 

  43. Takashima S, Kanatsu-Shinohara M, Tanaka T, Morimoto H, Inoue K, Ogonuki N, et al. Functional differences between GDNF-dependent and FGF2-dependent mouse spermatogonial stem cell self-renewal. Stem Cell Rep. 2015;4:489–502. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stemcr.2015.01.010.

    Article  CAS  Google Scholar 

  44. Mullaney BP, Skinner MK. Basic fibroblast growth factor (bFGF) gene expression and protein production during pubertal development of the seminiferous tubule: follicle-stimulating hormone-induced Sertoli cell bFGF expression. Endocrinology. 1992;131:2928–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/endo.131.6.1446630.

    Article  CAS  PubMed  Google Scholar 

  45. von Kopylow K, Staege H, Schulze W, Will H, Kirchhoff C. Fibroblast growth factor receptor 3 is highly expressed in rarely dividing human type A spermatogonia. Histochem Cell Biol. 2012;138:759–72. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00418-012-0991-7.

    Article  CAS  Google Scholar 

  46. Gomez M, Manzano A, Figueras A, Vinals F, Ventura F, Rosa JL, et al. Sertoli-secreted FGF-2 induces PFKFB4 isozyme expression in mouse spermatogenic cells by activation of the MEK/ERK/CREB pathway. Am J Physiol Endocrinol Metab. 2012;303:E695-707. https://doiorg.publicaciones.saludcastillayleon.es/10.1152/ajpendo.00381.2011.

    Article  CAS  PubMed  Google Scholar 

  47. Masaki K, Sakai M, Kuroki S, Jo JI, Hoshina K, Fujimori Y, et al. FGF2 has distinct molecular functions from GDNF in the mouse germline niche. Stem Cell Rep. 2018;10:1782–92. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stemcr.2018.03.016.

    Article  CAS  Google Scholar 

  48. 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. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.yexcr.2007.05.002.

    Article  CAS  PubMed  Google Scholar 

  49. Hess RA, Cooke PS, Hofmann MC, Murphy KM. Mechanistic insights into the regulation of the spermatogonial stem cell niche. Cell Cycle. 2006;5:1164–70. https://doiorg.publicaciones.saludcastillayleon.es/10.4161/cc.5.11.2775.

    Article  CAS  PubMed  Google Scholar 

  50. Chen C, Ouyang W, Grigura V, Zhou Q, Carnes K, Lim H, et al. ERM is required for transcriptional control of the spermatogonial stem cell niche. Nature. 2005;436:1030–4. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nature03894.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  51. Lee SJ, Park J, Lee DJ, Otsu K, Kim P, Mizuno S, et al. Mast4 knockout shows the regulation of spermatogonial stem cell self-renewal via the FGF2/ERM pathway. Cell Death Differ. 2021;28:1441–54. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41418-020-00670-2.

    Article  CAS  PubMed  Google Scholar 

  52. Lee SJ, Kim KH, Lee DJ, Kim P, Park J, Kim SJ, et al. MAST4 controls cell cycle in spermatogonial stem cells. Cell Prolif. 2023;56: e13390. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/cpr.13390.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  53. Jiang X, Skibba M, Zhang C, Tan Y, Xin Y, Qu Y. The roles of fibroblast growth factors in the testicular development and tumor. J Diabetes Res. 2013;2013: 489095. https://doiorg.publicaciones.saludcastillayleon.es/10.1155/2013/489095.

    Article  PubMed Central  PubMed  Google Scholar 

  54. Yang F, Whelan EC, Guan X, Deng B, Wang S, Sun J, et al. FGF9 promotes mouse spermatogonial stem cell proliferation mediated by p38 MAPK signalling. Cell Prolif. 2021;54: e12933. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/cpr.12933.

    Article  CAS  PubMed  Google Scholar 

  55. Rashtbari H, Razi M, Hassani-Bafrani H, Najaran H. Berberine reinforces Sertoli cells niche and accelerates spermatogonial stem cells renewal in experimentally-induced varicocele condition in rats. Phytomedicine. 2018;40:68–78. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.phymed.2017.12.036.

    Article  CAS  PubMed  Google Scholar 

  56. An XJ, Li Q, Chen N, Li TT, Wang HH, Su MC, et al. Effects of Pgam1-mediated glycolysis pathway in Sertoli cells on Spermatogonial stem cells based on transcriptomics and energy metabolomics. Front Vet Sci. 2022;9:992877. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fvets.2022.992877.

    Article  PubMed Central  PubMed  Google Scholar 

  57. Yamamuro T, Nakamura S, Yamano Y, Endo T, Yanagawa K, Tokumura A, et al. Rubicon prevents autophagic degradation of GATA4 to promote Sertoli cell function. PLoS Genet. 2021;17: e1009688. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pgen.1009688.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  58. Kyronlahti A, Euler R, Bielinska M, Schoeller EL, Moley KH, Toppari J, et al. GATA4 regulates Sertoli cell function and fertility in adult male mice. Mol Cell Endocrinol. 2011;333:85–95. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mce.2010.12.019.

    Article  CAS  PubMed  Google Scholar 

  59. Schrade A, Kyronlahti A, Akinrinade O, Pihlajoki M, Fischer S, Rodriguez VM, et al. GATA4 regulates blood-testis barrier function and lactate metabolism in mouse Sertoli cells. Endocrinology. 2016;157:2416–31. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2015-1927.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  60. Schrade A, Kyronlahti A, Akinrinade O, Pihlajoki M, Hakkinen M, Fischer S, et al. GATA4 is a key regulator of steroidogenesis and glycolysis in mouse Leydig cells. Endocrinology. 2015;156:1860–72. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2014-1931.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  61. Boyer A, Yeh JR, Zhang X, Paquet M, Gaudin A, Nagano MC, et al. CTNNB1 signaling in sertoli cells downregulates spermatogonial stem cell activity via WNT4. PLoS ONE. 2012;7: e29764. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0029764.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  62. Yeh JR, Zhang X, Nagano MC. Wnt5a is a cell-extrinsic factor that supports self-renewal of mouse spermatogonial stem cells. J Cell Sci. 2011;124:2357–66. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/jcs.080903.

    Article  CAS  PubMed  Google Scholar 

  63. Tanaka T, Kanatsu-Shinohara M, Lei Z, Rao CV, Shinohara T. The luteinizing hormone-testosterone pathway regulates mouse spermatogonial stem cell self-renewal by suppressing WNT5A expression in Sertoli cells. Stem Cell Rep. 2016;7:279–91. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stemcr.2016.07.005.

    Article  CAS  Google Scholar 

  64. Takase HM, Nusse R. Paracrine Wnt/beta-catenin signaling mediates proliferation of undifferentiated spermatogonia in the adult mouse testis. Proc Natl Acad Sci USA. 2016;113:E1489-1497. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1601461113.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  65. Cui W, He X, Zhai X, Zhang H, Zhang Y, Jin F, et al. CARF promotes spermatogonial self-renewal and proliferation through Wnt signaling pathway. Cell Discov. 2020;6:85. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41421-020-00212-7.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  66. Zheng QS, Wang XN, Wen Q, Zhang Y, Chen SR, Zhang J, et al. Wt1 deficiency causes undifferentiated spermatogonia accumulation and meiotic progression disruption in neonatal mice. Reproduction. 2014;147:45–52. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-13-0299.

    Article  CAS  PubMed  Google Scholar 

  67. Gao F, Maiti S, Alam N, Zhang Z, Deng JM, Behringer RR, et al. The Wilms tumor gene, Wt1, is required for Sox9 expression and maintenance of tubular architecture in the developing testis. Proc Natl Acad Sci USA. 2006;103:11987–92. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.0600994103.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  68. Wang XN, Li ZS, Ren Y, Jiang T, Wang YQ, Chen M, et al. The Wilms tumor gene, Wt1, is critical for mouse spermatogenesis via regulation of sertoli cell polarity and is associated with non-obstructive azoospermia in humans. PLoS Genet. 2013;9: e1003645. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pgen.1003645.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  69. Wen Y, Ma X, Wang X, Wang F, Dong J, Wu Y, et al. hnRNPU in Sertoli cells cooperates with WT1 and is essential for testicular development by modulating transcriptional factors Sox8/9. Theranostics. 2021;11:10030–46. https://doiorg.publicaciones.saludcastillayleon.es/10.7150/thno.66819.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  70. Griswold SL, Behringer RR. Fetal Leydig cell origin and development. Sex Dev. 2009;3:1–15. https://doiorg.publicaciones.saludcastillayleon.es/10.1159/000200077.

    Article  CAS  PubMed  Google Scholar 

  71. O’Shaughnessy PJ, Fowler PA. Endocrinology of the mammalian fetal testis. Reproduction. 2011;141:37–46. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-10-0365.

    Article  CAS  PubMed  Google Scholar 

  72. Wen Q, Wang Y, Tang J, Cheng CY, Liu YX. Sertoli cell Wt1 regulates peritubular Myoid cell and fetal Leydig cell differentiation during fetal testis development. PLoS ONE. 2016;11: e0167920. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0167920.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  73. Jia GX, Lin Z, Yan RG, Wang GW, Zhang XN, Li C, et al. WTAP function in Sertoli cells is essential for sustaining the spermatogonial stem cell niche. Stem Cell Rep. 2020;15:968–82. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stemcr.2020.09.001.

    Article  CAS  Google Scholar 

  74. Garcia TX, Parekh P, Gandhi P, Sinha K, Hofmann MC. The NOTCH Ligand JAG1 regulates GDNF expression in Sertoli cells. Stem Cells Dev. 2017;26:585–98. https://doiorg.publicaciones.saludcastillayleon.es/10.1089/scd.2016.0318.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  75. Garcia TX, DeFalco T, Capel B, Hofmann MC. Constitutive activation of NOTCH1 signaling in Sertoli cells causes gonocyte exit from quiescence. Dev Biol. 2013;377:188–201. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ydbio.2013.01.031.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  76. Garcia TX, Farmaha JK, Kow S, Hofmann MC. RBPJ in mouse Sertoli cells is required for proper regulation of the testis stem cell niche. Development. 2014;141:4468–78. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.113969.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  77. Payne CJ, Gallagher SJ, Foreman O, Dannenberg JH, Depinho RA, Braun RE. Sin3a is required by sertoli cells to establish a niche for undifferentiated spermatogonia, germ cell tumors, and spermatid elongation. Stem Cells. 2010;28:1424–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/stem.464.

    Article  CAS  PubMed  Google Scholar 

  78. Kanatsu-Shinohara M, Inoue K, Takashima S, Takehashi M, Ogonuki N, Morimoto H, et al. Reconstitution of mouse spermatogonial stem cell niches in culture. Cell Stem Cell. 2012;11:567–78. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stem.2012.06.011.

    Article  CAS  PubMed  Google Scholar 

  79. Yang QE, Kim D, Kaucher A, Oatley MJ, Oatley JM. CXCL12-CXCR4 signaling is required for the maintenance of mouse spermatogonial stem cells. J Cell Sci. 2013;126:1009–20. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/jcs.119826.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  80. Gilbert DC, Chandler I, McIntyre A, Goddard NC, Gabe R, Huddart RA, et al. Clinical and biological significance of CXCL12 and CXCR4 expression in adult testes and germ cell tumours of adults and adolescents. J Pathol. 2009;217:94–102. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/path.2436.

    Article  CAS  PubMed  Google Scholar 

  81. Xu J, Wan P, Wang M, Zhang J, Gao X, Hu B, et al. AIP1-mediated actin disassembly is required for postnatal germ cell migration and spermatogonial stem cell niche establishment. Cell Death Dis. 2015;6:e1818. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/cddis.2015.182.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  82. Jin C, Wang Z, Li P, Tang J, Jiao T, Li Y, et al. Decoding the spermatogonial stem cell niche under physiological and recovery conditions in adult mice and humans. Sci Adv. 2023;9:eabq3173. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/sciadv.abq3173.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  83. Pace AJ, Lee E, Athirakul K, Coffman TM, O’Brien DA, Koller BH. Failure of spermatogenesis in mouse lines deficient in the Na(+)-K(+)-2Cl(-) cotransporter. J Clin Invest. 2000;105:441–50. https://doiorg.publicaciones.saludcastillayleon.es/10.1172/JCI8553.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  84. Hu X, Tang Z, Li Y, Liu W, Zhang S, Wang B, et al. Deletion of the tyrosine phosphatase Shp2 in Sertoli cells causes infertility in mice. Sci Rep. 2015;5:12982. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/srep12982.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  85. Wang S, Wang X, Wu Y, Han C. IGF-1R signaling is essential for the proliferation of cultured mouse spermatogonial stem cells by promoting the G2/M progression of the cell cycle. Stem Cells Dev. 2015;24:471–83. https://doiorg.publicaciones.saludcastillayleon.es/10.1089/scd.2014.0376.

    Article  CAS  PubMed  Google Scholar 

  86. Liu YL, Yang SS, Chen SJ, Lin YC, Chu CC, Huang HH, et al. OSR1 and SPAK cooperatively modulate Sertoli cell support of mouse spermatogenesis. Sci Rep. 2016;6:37205. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/srep37205.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  87. Duarte A, Poderoso C, Cooke M, Soria G, Maciel FC, Gottifredi V, et al. Mitochondrial fusion is essential for steroid biosynthesis. PLoS ONE. 2012;7:e45829. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0045829.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  88. Puri P, Walker WH. The regulation of male fertility by the PTPN11 tyrosine phosphatase. Semin Cell Dev Biol. 2016;59:27–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.semcdb.2016.01.020.

    Article  CAS  PubMed  Google Scholar 

  89. 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. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12958-020-00686-w.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  90. Griswold MD. Spermatogenesis: the commitment to meiosis. Physiol Rev. 2016;96:1–17. https://doiorg.publicaciones.saludcastillayleon.es/10.1152/physrev.00013.2015.

    Article  CAS  PubMed  Google Scholar 

  91. 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. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/sj.emboj.7601447.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  92. Sandell LL, Lynn ML, Inman KE, McDowell W, Trainor PA. RDH10 oxidation of Vitamin A is a critical control step in synthesis of retinoic acid during mouse embryogenesis. PLoS ONE. 2012;7: e30698. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0030698.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  93. Tong MH, Yang QE, Davis JC, Griswold MD. Retinol dehydrogenase 10 is indispensible for spermatogenesis in juvenile males. Proc Natl Acad Sci USA. 2013;110:543–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1214883110.

    Article  PubMed  Google Scholar 

  94. Vincent S, Segretain D, Nishikawa S, Nishikawa SI, Sage J, Cuzin F, et al. Stage-specific expression of the Kit receptor and its ligand (KL) during male gametogenesis in the mouse: a Kit-KL interaction critical for meiosis. Development. 1998;125:4585–93. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.125.22.4585.

    Article  CAS  PubMed  Google Scholar 

  95. Sato T, Yokonishi T, Komeya M, Katagiri K, Kubota Y, Matoba S, et al. Testis tissue explantation cures spermatogenic failure in c-Kit ligand mutant mice. Proc Natl Acad Sci USA. 2012;109:16934–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1211845109.

    Article  PubMed Central  PubMed  Google Scholar 

  96. Liu S, Chen X, Wang Y, Li L, Wang G, Li X, et al. A role of KIT receptor signaling for proliferation and differentiation of rat stem Leydig cells in vitro. Mol Cell Endocrinol. 2017;444:1–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mce.2017.01.023.

    Article  CAS  PubMed  Google Scholar 

  97. Peng YJ, Tang XT, Shu HS, Dong W, Shao H, Zhou BO. Sertoli cells are the source of stem cell factor for spermatogenesis. Development. 2023. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.200706.

    Article  PubMed Central  PubMed  Google Scholar 

  98. Tao K, Sun Y, Chao Y, Xing L, Leng L, Zhou D, et al. beta-estradiol promotes the growth of primary human fetal spermatogonial stem cells via the induction of stem cell factor in Sertoli cells. J Assist Reprod Genet. 2021;38:2481–90. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10815-021-02240-y.

    Article  PubMed Central  PubMed  Google Scholar 

  99. Zhang J, Eto K, Honmyou A, Nakao K, Kiyonari H, Abe S. Neuregulins are essential for spermatogonial proliferation and meiotic initiation in neonatal mouse testis. Development. 2011;138:3159–68. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.062380.

    Article  CAS  PubMed  Google Scholar 

  100. Naillat F, Veikkolainen V, Miinalainen I, Sipila P, Poutanen M, Elenius K, et al. ErbB4, a receptor tyrosine kinase, coordinates organization of the seminiferous tubules in the developing testis. Mol Endocrinol. 2014;28:1534–46. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/me.2013-1244.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  101. Xia Q, Zhang D, Wang J, Zhang X, Song W, Chen R, et al. Androgen indirectly regulates gap junction component connexin 43 through Wilms Tumor-1 in Sertoli cells. Stem Cells Dev. 2020;29:169–76. https://doiorg.publicaciones.saludcastillayleon.es/10.1089/scd.2019.0166.

    Article  CAS  PubMed  Google Scholar 

  102. Sridharan S, Brehm R, Bergmann M, Cooke PS. Role of connexin 43 in Sertoli cells of testis. Ann N Y Acad Sci. 2007;1120:131–43. https://doiorg.publicaciones.saludcastillayleon.es/10.1196/annals.1411.004.

    Article  CAS  PubMed  Google Scholar 

  103. Rode K, Weider K, Damm OS, Wistuba J, Langeheine M, Brehm R. Loss of connexin 43 in Sertoli cells provokes postnatal spermatogonial arrest, reduced germ cell numbers and impaired spermatogenesis. Reprod Biol. 2018;18:456–66. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.repbio.2018.08.001.

    Article  PubMed  Google Scholar 

  104. Weider K, Bergmann M, Giese S, Guillou F, Failing K, Brehm R. Altered differentiation and clustering of Sertoli cells in transgenic mice showing a Sertoli cell specific knockout of the connexin 43 gene. Differentiation. 2011;82:38–49. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.diff.2011.03.001.

    Article  CAS  PubMed  Google Scholar 

  105. Zhao Y, Chen MS, Wang JX, Cui JG, Zhang H, Li XN, et al. Connexin-43 is a promising target for lycopene preventing phthalate-induced spermatogenic disorders. J Adv Res. 2023;49:115–26. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.jare.2022.09.001.

    Article  CAS  PubMed  Google Scholar 

  106. Li N, Mruk DD, Mok KW, Li MW, Wong CK, Lee WM, et al. Connexin 43 reboots meiosis and reseals blood-testis barrier following toxicant-mediated aspermatogenesis and barrier disruption. FASEB J. 2016;30:1436–52. https://doiorg.publicaciones.saludcastillayleon.es/10.1096/fj.15-276527.

    Article  CAS  PubMed  Google Scholar 

  107. Liang J, Wang N, He J, Du J, Guo Y, Li L, et al. Induction of Sertoli-like cells from human fibroblasts by NR5A1 and GATA4. Elife. 2019. https://doiorg.publicaciones.saludcastillayleon.es/10.7554/eLife.48767.

    Article  PubMed Central  PubMed  Google Scholar 

  108. Bott RC, McFee RM, Clopton DT, Toombs C, Cupp AS. Vascular endothelial growth factor and kinase domain region receptor are involved in both seminiferous cord formation and vascular development during testis morphogenesis in the rat. Biol Reprod. 2006;75:56–67. https://doiorg.publicaciones.saludcastillayleon.es/10.1095/biolreprod.105.047225.

    Article  CAS  PubMed  Google Scholar 

  109. Caires KC, de Avila JM, Cupp AS, McLean DJ. VEGFA family isoforms regulate spermatogonial stem cell homeostasis in vivo. Endocrinology. 2012;153:887–900. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2011-1323.

    Article  CAS  PubMed  Google Scholar 

  110. Sargent KM, Clopton DT, Lu N, Pohlmeier WE, Cupp AS. VEGFA splicing: divergent isoforms regulate spermatogonial stem cell maintenance. Cell Tissue Res. 2016;363:31–45. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00441-015-2297-2.

    Article  CAS  PubMed  Google Scholar 

  111. Agrimson KS, Minkina A, Sadowski D, Wheeler A, Murphy MW, Gearhart MD, et al. Lrh1 can help reprogram sexual cell fate and is required for Sertoli cell development and spermatogenesis in the mouse testis. PLoS Genet. 2022;18: e1010088. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pgen.1010088.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  112. Lai F, Wang H, Zhao X, Yang K, Cai L, Hu M, et al. RNF20 is required for male fertility through regulation of H2B ubiquitination in the Sertoli cells. Cell Biosci. 2023;13:71. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13578-023-01018-2.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  113. Zhang H, Chen F, Dong H, Xie M, Zhang H, Chen Y, et al. Loss of Fbxw7 in Sertoli cells impairs testis development and causes infertility in micedagger. Biol Reprod. 2020;102:963–74. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/biolre/ioz230.

    Article  PubMed  Google Scholar 

  114. Kanatsu-Shinohara M, Onoyama I, Nakayama KI, Shinohara T. Skp1-Cullin-F-box (SCF)-type ubiquitin ligase FBXW7 negatively regulates spermatogonial stem cell self-renewal. Proc Natl Acad Sci USA. 2014;111:8826–31. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1401837111.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  115. Cai Z, Niu Y, Li H. RNA N6-methyladenosine modification, spermatogenesis, and human male infertility. Mol Hum Reprod. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/molehr/gaab020.

    Article  PubMed  Google Scholar 

  116. Cai Z, Zhang Y, Yang L, Ma C, Fei Y, Ding J, et al. ALKBH5 in mouse testicular Sertoli cells regulates Cdh2 mRNA translation to maintain blood-testis barrier integrity. Cell Mol Biol Lett. 2022;27:101. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s11658-022-00404-x.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  117. Zhang L, Chen M, Wen Q, Li Y, Wang Y, Wang Y, et al. Reprogramming of Sertoli cells to fetal-like Leydig cells by Wt1 ablation. Proc Natl Acad Sci USA. 2015;112:4003–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.1422371112.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  118. Jiang MH, Cai B, Tuo Y, Wang J, Zang ZJ, Tu X, et al. Characterization of Nestin-positive stem Leydig cells as a potential source for the treatment of testicular Leydig cell dysfunction. Cell Res. 2014;24:1466–85. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/cr.2014.149.

    Article  PubMed Central  PubMed  Google Scholar 

  119. Liu W, Du L, Cui Y, He C, He Z. WNT5A regulates the proliferation, apoptosis and stemness of human stem Leydig cells via the beta-catenin signaling pathway. Cell Mol Life Sci. 2024;81:93. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00018-023-05077-z.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  120. Chen H, Wang Y, Ge R, Zirkin BR. Leydig cell stem cells: Identification, proliferation and differentiation. Mol Cell Endocrinol. 2017;445:65–73. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mce.2016.10.010.

    Article  CAS  PubMed  Google Scholar 

  121. Zaker H, Razi M, Mahmoudian A, Soltanalinejad F. Boosting effect of testosterone on GDNF expression in Sertoli cell line (TM4); comparison between TM3 cells-produced and exogenous testosterone. Gene. 2022;812: 146112. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.gene.2021.146112.

    Article  CAS  PubMed  Google Scholar 

  122. Gonzalez-Herrera IG, Prado-Lourenco L, Pileur F, Conte C, Morin A, Cabon F, et al. Testosterone regulates FGF-2 expression during testis maturation by an IRES-dependent translational mechanism. FASEB J. 2006;20:476–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1096/fj.04-3314fje.

    Article  PubMed  Google Scholar 

  123. 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. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mce.2018.12.012.

    Article  CAS  PubMed  Google Scholar 

  124. Guo H, Zhang D, Zhou Y, Sun L, Li C, Luo X, et al. Casein kinase 1alpha regulates testosterone synthesis and testis development in adult mice. Endocrinology. 2023. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/endocr/bqad042.

    Article  PubMed Central  PubMed  Google Scholar 

  125. Xing D, Jin Y, Sun D, Liu Y, Cai B, Gao C, et al. Protective effect of TNFAIP3 on testosterone production in Leydig cells under an aging inflammatory microenvironment. Arch Gerontol Geriatr. 2024;117:105274. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.archger.2023.105274.

    Article  CAS  PubMed  Google Scholar 

  126. Umehara T, Kawashima I, Kawai T, Hoshino Y, Morohashi KI, Shima Y, et al. Neuregulin 1 regulates proliferation of Leydig cells to support spermatogenesis and sexual behavior in adult mice. Endocrinology. 2016;157:4899–913. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2016-1478.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  127. Di Fiore MM, Santillo A, Falvo S, Longobardi S, Chieffi Baccari G. Molecular mechanisms elicited by d-Aspartate in Leydig cells and spermatogonia. Int J Mol Sci. 2016;17:1127. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ijms17071127.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  128. Oatley JM, Oatley MJ, Avarbock MR, Tobias JW, Brinster RL. Colony stimulating factor 1 is an extrinsic stimulator of mouse spermatogonial stem cell self-renewal. Development. 2009;136:1191–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.032243.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  129. Du Z, Xu S, Hu S, Yang H, Zhou Z, Sidhu K, et al. Melatonin attenuates detrimental effects of diabetes on the niche of mouse spermatogonial stem cells by maintaining Leydig cells. Cell Death Dis. 2018;9:968. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41419-018-0956-4.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  130. Smith OE, Morin F, Roussel V, Bertucci MC, Boyer A, Murphy BD. The role of steroidogenic factor 1 (SF-1) in steroidogenic cell function of the testes and ovaries of mature mice. Reproduction. 2023;165:1–17. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-22-0049.

    Article  CAS  PubMed  Google Scholar 

  131. Huang YH, Chin CC, Ho HN, Chou CK, Shen CN, Kuo HC, et al. Pluripotency of mouse spermatogonial stem cells maintained by IGF-1-dependent pathway. FASEB J. 2009;23:2076–87. https://doiorg.publicaciones.saludcastillayleon.es/10.1096/fj.08-121939.

    Article  CAS  PubMed  Google Scholar 

  132. Mossadegh-Keller N, Gentek R, Gimenez G, Bigot S, Mailfert S, Sieweke MH. Developmental origin and maintenance of distinct testicular macrophage populations. J Exp Med. 2017;214:2829–41. https://doiorg.publicaciones.saludcastillayleon.es/10.1084/jem.20170829.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  133. Gillette R, Tiwary R, Voss J, Hewage SN, Richburg JH. Peritubular macrophages are recruited to the testis of peripubertal rats after mono-(2-ethylhexyl) phthalate exposure and is associated with increases in the numbers of spermatogonia. Toxicol Sci. 2021;182:288–96. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/toxsci/kfab059.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  134. Mossadegh-Keller N, Sieweke MH. Testicular macrophages: guardians of fertility. Cell Immunol. 2018;330:120–5. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.cellimm.2018.03.009.

    Article  CAS  PubMed  Google Scholar 

  135. Min Z, Wan J, Xin H, Liu X, Rao X, Fan Z, et al. NR2C2 of macrophages promotes inflammation via NF-kappaB in LPS-induced orchitis in mice. Reproduction. 2023;166:209–20. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-23-0041.

    Article  CAS  PubMed  Google Scholar 

  136. Spinnler K, Kohn FM, Schwarzer U, Mayerhofer A. Glial cell line-derived neurotrophic factor is constitutively produced by human testicular peritubular cells and may contribute to the spermatogonial stem cell niche in man. Hum Reprod. 2010;25:2181–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/deq170.

    Article  CAS  PubMed  Google Scholar 

  137. Chen LY, Brown PR, Willis WB, Eddy EM. Peritubular myoid cells participate in male mouse spermatogonial stem cell maintenance. Endocrinology. 2014;155:4964–74. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2014-1406.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  138. Welsh M, Saunders PT, Atanassova N, Sharpe RM, Smith LB. Androgen action via testicular peritubular myoid cells is essential for male fertility. FASEB J. 2009;23:4218–30. https://doiorg.publicaciones.saludcastillayleon.es/10.1096/fj.09-138347.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  139. De Gendt K, Swinnen JV, Saunders PT, Schoonjans L, Dewerchin M, Devos A, et al. A Sertoli cell-selective knockout of the androgen receptor causes spermatogenic arrest in meiosis. Proc Natl Acad Sci USA. 2004;101:1327–32. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.0308114100.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  140. Hazra R, Upton D, Desai R, Noori O, Jimenez M, Handelsman DJ, et al. Elevated expression of the Sertoli cell androgen receptor disrupts male fertility. Am J Physiol Endocrinol Metab. 2016;311:E396-404. https://doiorg.publicaciones.saludcastillayleon.es/10.1152/ajpendo.00159.2016.

    Article  PubMed  Google Scholar 

  141. Tao HP, Lu TF, Li S, Jia GX, Zhang XN, Yang QE, et al. Pancreatic lipase-related protein 2 is selectively expressed by peritubular myoid cells in the murine testis and sustains long-term spermatogenesis. Cell Mol Life Sci. 2023;80:217. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00018-023-04872-y.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  142. Qian Y, Liu S, Guan Y, Pan H, Guan X, Qiu Z, et al. Lgr4-mediated Wnt/beta-catenin signaling in peritubular myoid cells is essential for spermatogenesis. Development. 2013;140:1751–61. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/dev.093641.

    Article  CAS  PubMed  Google Scholar 

  143. Bhang DH, Kim BJ, Kim BG, Schadler K, Baek KH, Kim YH, et al. Testicular endothelial cells are a critical population in the germline stem cell niche. Nat Commun. 2018;9:4379. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41467-018-06881-z.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  144. Bissell MJ, Hall HG, Parry G. How does the extracellular matrix direct gene expression? J Theor Biol. 1982;99:31–68. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/0022-5193(82)90388-5.

    Article  CAS  PubMed  Google Scholar 

  145. Zhang Y, He Y, Bharadwaj S, Hammam N, Carnagey K, Myers R, et al. Tissue-specific extracellular matrix coatings for the promotion of cell proliferation and maintenance of cell phenotype. Biomaterials. 2009;30:4021–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.biomaterials.2009.04.005.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  146. Tung PS, Skinner MK, Fritz IB. Cooperativity between Sertoli cells and peritubular myoid cells in the formation of the basal lamina in the seminiferous tubule. Ann N Y Acad Sci. 1984;438:435–46. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/j.1749-6632.1984.tb38304.x.

    Article  CAS  PubMed  Google Scholar 

  147. Bashiri Z, Amiri I, Gholipourmalekabadi M, Falak R, Asgari H, Maki CB, et al. Artificial testis: a testicular tissue extracellular matrix as a potential bio-ink for 3D printing. Biomater Sci. 2021;9:3465–84. https://doiorg.publicaciones.saludcastillayleon.es/10.1039/d0bm02209h.

    Article  CAS  PubMed  Google Scholar 

  148. Dym M. Basement membrane regulation of Sertoli cells. Endocr Rev. 1994;15:102–15. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/edrv-15-1-102.

    Article  CAS  PubMed  Google Scholar 

  149. Baert Y, Stukenborg JB, Landreh M, De Kock J, Jornvall H, Soder O, et al. Derivation and characterization of a cytocompatible scaffold from human testis. Hum Reprod. 2015;30:256–67. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/deu330.

    Article  CAS  PubMed  Google Scholar 

  150. Hadley MA, Dym M. Immunocytochemistry of extracellular matrix in the lamina propria of the rat testis: electron microscopic localization. Biol Reprod. 1987;37:1283–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1095/biolreprod37.5.1283.

    Article  CAS  PubMed  Google Scholar 

  151. Kahsai TZ, Enders GC, Gunwar S, Brunmark C, Wieslander J, Kalluri R, et al. Seminiferous tubule basement membrane. Composition and organization of type IV collagen chains, and the linkage of alpha3(IV) and alpha5(IV) chains. J Biol Chem. 1997;272:17023–32. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.272.27.17023.

    Article  CAS  PubMed  Google Scholar 

  152. Siu MK, Lee WM, Cheng CY. The interplay of collagen IV, tumor necrosis factor-alpha, gelatinase B (matrix metalloprotease-9), and tissue inhibitor of metalloproteases-1 in the basal lamina regulates Sertoli cell-tight junction dynamics in the rat testis. Endocrinology. 2003;144:371–87. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2002-220786.

    Article  CAS  PubMed  Google Scholar 

  153. Shiva R, Ghasem S, Masoud H, Sadat KL, Ali K, Reza DM. Comparison of colony formation of human spermatogonial stem cells (SSCs) with and without collagen. J Pak Med Assoc. 2016;66:285–91.

    PubMed  Google Scholar 

  154. Chen SH, Li D, Xu C. Downregulation of Col1a1 induces differentiation in mouse spermatogonia. Asian J Androl. 2012;14:842–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/aja.2012.66.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  155. Zhao H, Nie J, Zhu X, Lu Y, Liang X, Xu H, et al. In vitro differentiation of spermatogonial stem cells using testicular cells from Guangxi Bama mini-pig. J Vet Sci. 2018;19:592–9. https://doiorg.publicaciones.saludcastillayleon.es/10.4142/jvs.2018.19.5.592.

    Article  PubMed Central  PubMed  Google Scholar 

  156. Siu MK, Cheng CY. Interactions of proteases, protease inhibitors, and the beta1 integrin/laminin gamma3 protein complex in the regulation of ectoplasmic specialization dynamics in the rat testis. Biol Reprod. 2004;70:945–64. https://doiorg.publicaciones.saludcastillayleon.es/10.1095/biolreprod.103.023606.

    Article  CAS  PubMed  Google Scholar 

  157. Colognato H, Yurchenco PD. Form and function: the laminin family of heterotrimers. Dev Dyn. 2000;218:213–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/(SICI)1097-0177(200006)218:2%3c213::AID-DVDY1%3e3.0.CO;2-R.

    Article  CAS  PubMed  Google Scholar 

  158. Yan HHN, Cheng CY. Laminin alpha 3 forms a complex with beta3 and gamma3 chains that serves as the ligand for alpha 6beta1-integrin at the apical ectoplasmic specialization in adult rat testes. J Biol Chem. 2006;281:17286–303. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M513218200.

    Article  CAS  PubMed  Google Scholar 

  159. Hager M, Gawlik K, Nystrom A, Sasaki T, Durbeej M. Laminin alpha1 chain corrects male infertility caused by absence of laminin alpha2 chain. Am J Pathol. 2005;167:823–33. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/s0002-9440(10)62054-8.

    Article  PubMed Central  PubMed  Google Scholar 

  160. Kurek M, Akesson E, Yoshihara M, Oliver E, Cui Y, Becker M, et al. Spermatogonia loss correlates with LAMA 1 expression in human prepubertal testes stored for fertility preservation. Cells. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/cells10020241.

    Article  PubMed Central  PubMed  Google Scholar 

  161. Van Saen D, Vloeberghs V, Gies I, De Schepper J, Tournaye H, Goossens E. Characterization of the stem cell niche components within the seminiferous tubules in testicular biopsies of Klinefelter patients. Fertil Steril. 2020;113(1183–1195):e1183. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.fertnstert.2020.01.018.

    Article  CAS  Google Scholar 

  162. Di Persio S, Tekath T, Siebert-Kuss LM, Cremers JF, Wistuba J, Li X, et al. Single-cell RNA-seq unravels alterations of the human spermatogonial stem cell compartment in patients with impaired spermatogenesis. Cell Rep Med. 2021;2:100395. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.xcrm.2021.100395.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  163. Saharkhiz S, Abdolmaleki Z, Eslampour MA. Hyaluronic acid/silicon nanoparticle scaffold induces proliferation and differentiation of mouse spermatogonial stem cells transplanted to epididymal adipose tissue. Cell Tissue Bank. 2024;25:231–43. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10561-023-10093-1.

    Article  CAS  PubMed  Google Scholar 

  164. Kanatsu-Shinohara M, Takehashi M, Takashima S, Lee J, Morimoto H, Chuma S, et al. Homing of mouse spermatogonial stem cells to germline niche depends on beta1-integrin. Cell Stem Cell. 2008;3:533–42. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stem.2008.08.002.

    Article  CAS  PubMed  Google Scholar 

  165. Park MH, Yun JI, Lee E, Lee ST. Integrin heterodimer alpha(9) beta(1) is localized on the surface of porcine spermatogonial stem cells in the undifferentiated state. Reprod Domest Anim. 2019;54:1497–500. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/rda.13555.

    Article  CAS  PubMed  Google Scholar 

  166. Luo L, Li E, Zhao S, Wang J, Zhu Z, Liu Y, et al. Gasoline exhaust damages spermatogenesis through downregulating alpha6-integrin and beta1-integrin in the rat model. Andrologia. 2018;50:e13045. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/and.13045.

    Article  CAS  PubMed  Google Scholar 

  167. Rilla K, Mustonen AM, Arasu UT, Harkonen K, Matilainen J, Nieminen P. Extracellular vesicles are integral and functional components of the extracellular matrix. Matrix Biol. 2019;75–76:201–19. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.matbio.2017.10.003.

    Article  CAS  PubMed  Google Scholar 

  168. Ma Y, Ma QW, Sun Y, Chen XF. The emerging role of extracellular vesicles in the testis. Hum Reprod. 2023;38:334–51. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/dead015.

    Article  CAS  PubMed  Google Scholar 

  169. Li Q, Li H, Liang J, Mei J, Cao Z, Zhang L, et al. Sertoli cell-derived exosomal MicroRNA-486-5p regulates differentiation of spermatogonial stem cell through PTEN in mice. J Cell Mol Med. 2021;25:3950–62. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/jcmm.16347.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  170. Tian H, Wang X, Li X, Song W, Mi J, Zou K. Regulation of spermatogonial stem cell differentiation by Sertoli cells-derived exosomes through paracrine and autocrine signaling. J Cell Physiol. 2024. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/jcp.31202.

    Article  PubMed  Google Scholar 

  171. Thiageswaran S, Steele H, Voigt AL, Dobrinski I. A role for exchange of extracellular vesicles in porcine spermatogonial co-culture. Int J Mol Sci. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ijms23094535.

    Article  PubMed Central  PubMed  Google Scholar 

  172. Lin Y, Fang Q, He Y, Gong X, Wang Y, Liang A, et al. Thy1-positive spermatogonia suppress the proliferation of spermatogonial stem cells by extracellular vesicles in vitro. Endocrinology. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/endocr/bqab052.

    Article  PubMed  Google Scholar 

  173. Qu M, Liu X, Wang X, Li Z, Zhou L, Li H. Palmitoylation of vacuole membrane protein 1 promotes small extracellular vesicle secretion via interaction with ALIX and influences intercellular communication. Cell Commun Signal. 2024;22:150. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12964-024-01529-6.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  174. Ma Y, Zhou Y, Zou SS, Sun Y, Chen XF. Exosomes released from Sertoli cells contribute to the survival of Leydig cells through CCL20 in rats. Mol Hum Reprod. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/molehr/gaac002.

    Article  PubMed  Google Scholar 

  175. Garcia-Gareta E, Abduldaiem Y, Sawadkar P, Kyriakidis C, Lali F, Greco KV. Decellularised scaffolds: just a framework? Current knowledge and future directions. J Tissue Eng. 2020;11:2041731420942903. https://doiorg.publicaciones.saludcastillayleon.es/10.1177/2041731420942903.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  176. Yang Y, Lin Q, Zhou C, Li Q, Li Z, Cao Z, et al. A testis-derived hydrogel as an efficient feeder-free culture platform to promote mouse spermatogonial stem cell proliferation and differentiation. Front Cell Dev Biol. 2020;8:250. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fcell.2020.00250.

    Article  PubMed Central  PubMed  Google Scholar 

  177. Murdock MH, David S, Swinehart IT, Reing JE, Tran K, Gassei K, et al. Human testis extracellular matrix enhances human spermatogonial stem cell survival in vitro. Tissue Eng Part A. 2019;25:663–76. https://doiorg.publicaciones.saludcastillayleon.es/10.1089/ten.TEA.2018.0147.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  178. Piechka A, Sparanese S, Witherspoon L, Hach F, Flannigan R. Molecular mechanisms of cellular dysfunction in testes from men with non-obstructive azoospermia. Nat Rev Urol. 2024;21:67–90. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41585-023-00837-9.

    Article  CAS  PubMed  Google Scholar 

  179. O’Bryan MK, Grealy A, Stahl PJ, Schlegel PN, McLachlan RI, Jamsai D. Genetic variants in the ETV5 gene in fertile and infertile men with nonobstructive azoospermia associated with Sertoli cell-only syndrome. Fertil Steril. 2012;98(827–835):e821-823. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.fertnstert.2012.06.013.

    Article  CAS  Google Scholar 

  180. Chung CL, Lu CW, Cheng YS, Lin CY, Sun HS, Lin YM. Association of aberrant expression of sex-determining gene fibroblast growth factor 9 with Sertoli cell-only syndrome. Fertil Steril. 2013;100(1547–1554):e1541-1544. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.fertnstert.2013.08.004.

    Article  CAS  Google Scholar 

  181. Defamie N, Berthaut I, Mograbi B, Chevallier D, Dadoune JP, Fenichel P, et al. Impaired gap junction connexin43 in Sertoli cells of patients with secretory azoospermia: a marker of undifferentiated Sertoli cells. Lab Invest. 2003;83:449–56. https://doiorg.publicaciones.saludcastillayleon.es/10.1097/01.lab.0000059928.82702.6d.

    Article  CAS  PubMed  Google Scholar 

  182. Paduch DA, Hilz S, Grimson A, Schlegel PN, Jedlicka AE, Wright WW. Aberrant gene expression by Sertoli cells in infertile men with Sertoli cell-only syndrome. PLoS ONE. 2019;14:e0216586. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0216586.

    Article  PubMed Central  PubMed  Google Scholar 

  183. Cina DP, Flannigan R. Failing to mature: somatic cell dysfunction in the spermatogenic niche among patients with Sertoli cell-only and Klinefelter syndromes. Fertil Steril. 2020;113:1155–6. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.fertnstert.2020.03.006.

    Article  PubMed  Google Scholar 

  184. Wang H, Yuan L, Song J, Wang Q, Zhang Y. Distribution of extracellular matrix related proteins in normal and cryptorchid ziwuling black goat testes. Anim Reprod. 2022;19:e20220005. https://doiorg.publicaciones.saludcastillayleon.es/10.1590/1984-3143-AR2022-0005.

    Article  PubMed Central  PubMed  Google Scholar 

  185. Huang D, Zuo Y, Zhang C, Sun G, Jing Y, Lei J, et al. A single-nucleus transcriptomic atlas of primate testicular aging reveals exhaustion of the spermatogonial stem cell reservoir and loss of Sertoli cell homeostasis. Protein Cell. 2023;14:888–907. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/procel/pwac057.

    Article  CAS  PubMed  Google Scholar 

  186. Hashemi Karoii D, Azizi H, Skutella T. Altered G-protein transduction protein gene expression in the testis of infertile patients with nonobstructive azoospermia. DNA Cell Biol. 2023;42:617–37. https://doiorg.publicaciones.saludcastillayleon.es/10.1089/dna.2023.0189.

    Article  CAS  PubMed  Google Scholar 

  187. Li X, Xu A, Li K, Zhang J, Li Q, Zhao G, et al. CXCR4-SF1 bifunctional adipose-derived stem cells benefit for the treatment of Leydig cell dysfunction-related diseases. J Cell Mol Med. 2020;24:4633–45. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/jcmm.15128.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  188. Chen Y, Li C, Ji W, Wang L, Chen X, Zhao S, et al. Differentiation of human adipose derived stem cells into Leydig-like cells with molecular compounds. J Cell Mol Med. 2019;23:5956–69. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/jcmm.14427.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  189. Hassan HA, Essawi ML, Mekkawy MK, Mazen I. Novel mutations of the LHCGR gene in two families with 46, XY DSD causing Leydig cell hypoplasia I. Hormones (Athens). 2020;19:573–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s42000-020-00226-6.

    Article  PubMed  Google Scholar 

  190. Xia K, Wang F, Lai X, Dong L, Luo P, Zhang S, et al. AAV-mediated gene therapy produces fertile offspring in the Lhcgr-deficient mouse model of Leydig cell failure. Cell Rep Med. 2022;3:100792. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.xcrm.2022.100792.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  191. Salehi N, Totonchi M. The construction of a testis transcriptional cell atlas from embryo to adult reveals various somatic cells and their molecular roles. J Transl Med. 2023;21:859. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12967-023-04722-2.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  192. Di Persio S, Neuhaus N. Human spermatogonial stem cells and their niche in male (in)fertility: novel concepts from single-cell RNA-sequencing. Hum Reprod. 2023;38:1–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humrep/deac245.

    Article  CAS  PubMed  Google Scholar 

  193. Ikawa M, Tergaonkar V, Ogura A, Ogonuki N, Inoue K, Verma IM. Restoration of spermatogenesis by lentiviral gene transfer: offspring from infertile mice. Proc Natl Acad Sci USA. 2002;99:7524–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.072207299.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  194. Abel MH, Baker PJ, Charlton HM, Monteiro A, Verhoeven G, De Gendt K, et al. Spermatogenesis and sertoli cell activity in mice lacking sertoli cell receptors for follicle-stimulating hormone and androgen. Endocrinology. 2008;149:3279–85. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/en.2008-0086.

    Article  CAS  PubMed  Google Scholar 

  195. Ruthig VA, Lamb DJ. Updates in Sertoli cell-mediated signaling during spermatogenesis and advances in restoring Sertoli cell function. Front Endocrinol (Lausanne). 2022;13:897196. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2022.897196.

    Article  PubMed  Google Scholar 

  196. Naeemi S, Eidi A, Khanbabaee R, Sadri-Ardekani H, Kajbafzadeh AM. Differentiation and proliferation of spermatogonial stem cells using a three-dimensional decellularized testicular scaffold: a new method to study the testicular microenvironment in vitro. Int Urol Nephrol. 2021;53:1543–50. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s11255-021-02877-9.

    Article  CAS  PubMed  Google Scholar 

  197. Alves-Lopes JP, Stukenborg JB. Testicular organoids: a new model to study the testicular microenvironment in vitro? Hum Reprod Update. 2018;24:176–91. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/humupd/dmx036.

    Article  CAS  PubMed  Google Scholar 

  198. Kanbar M, Vermeulen M, Wyns C. Organoids as tools to investigate the molecular mechanisms of male infertility and its treatments. Reproduction. 2021;161:R103–12. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-20-0499.

    Article  CAS  PubMed  Google Scholar 

  199. Lundberg E, Borner GHH. Spatial proteomics: a powerful discovery tool for cell biology. Nat Rev Mol Cell Biol. 2019;20:285–302. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41580-018-0094-y.

    Article  CAS  PubMed  Google Scholar 

  200. Wu S, Li X, Li P, Li T, Huang G, Sun Q, et al. Developing rat testicular organoid models for assessing the reproductive toxicity of antidepression drugs in vitro. Acta Biochim Biophys Sin (Shanghai). 2022;54:1748–52. https://doiorg.publicaciones.saludcastillayleon.es/10.3724/abbs.2022164.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This study was supported by Hunan Province Natural Science Foundation, Grant/Award Number: 2024JJ5518 and 2022JJ40253; Science and Technology Planning Project of the Hunan Provincial Department of Science and Technology, Grant/Award Number: 2023ZK4175; Scientific research project of Hunan Health Commission, Grant/Award Number: 202102041763; Changsha Municipal Natural Science Foundation, Grant/Award Number: kq2014267; Hunan Cancer Hospital Climb Plan, Grant/Award Number: 2023NSFC-B004; and the Hunan Provincial Innovation Foundation for Postgraduate, Grant/Award Number: CX20230465.

Author information

Authors and Affiliations

Authors

Contributions

Wei Liu: Investigation, Writing-original draft. Li Du: Investigation, Writing-original draft. Junjun Li: Visualization. Yan He: Project administration, Funding acquisition. Mengjie Tang: Project administration, Funding acquisition. All authors actively participated in the revision of the manuscript, carefully reviewed it, and approved the final version for submission.

Corresponding authors

Correspondence to Yan He or Mengjie Tang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

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-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, 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 you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. 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-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, W., Du, L., Li, J. et al. Microenvironment of spermatogonial stem cells: a key factor in the regulation of spermatogenesis. Stem Cell Res Ther 15, 294 (2024). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13287-024-03893-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13287-024-03893-z

Keywords