Next Article in Journal
Fusion of Normoxic- and Hypoxic-Preconditioned Myoblasts Leads to Increased Hypertrophy
Next Article in Special Issue
Mechanisms of Sperm–Egg Interactions: What Ascidian Fertilization Research Has Taught Us
Previous Article in Journal
Cumulative Metabolic and Epigenetic Effects of Paternal and/or Maternal Supplementation with Arachidonic Acid across Three Consecutive Generations in Mice
Previous Article in Special Issue
SCF Ligases and Their Functions in Oogenesis and Embryogenesis—Summary of the Most Important Findings throughout the Animal Kingdom
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ubiquitin-Proteasome System–Regulated Protein Degradation in Spermatogenesis

1
Zhejiang University-University of Edinburgh Institute (ZJU-UoE Institute), Zhejiang University School of Medicine, International Campus, Zhejiang University, 718 East Haizhou Rd, Haining 314400, China
2
Key Laboratory of Reproductive Dysfunction Management of Zhejiang Province, Assisted Reproduction Unit, Department of Obstetrics and Gynecology, School of Medicine, Zhejiang University, Sir Run Run Shaw Hospital, 3 East Qing Chun Rd, Hangzhou 310020, China
3
College of Life Sciences, Zhejiang University, 866 Yuhangtang Rd, Hangzhou 310058, China
4
Department of Dermatology, The Second Affiliated Hospital, Zhejiang University School of Medicine, Zhejiang University, Hangzhou 310029, China
*
Author to whom correspondence should be addressed.
Cells 2022, 11(6), 1058; https://doi.org/10.3390/cells11061058
Submission received: 28 January 2022 / Revised: 14 March 2022 / Accepted: 18 March 2022 / Published: 21 March 2022

Abstract

:
Spermatogenesis is a prolonged and highly ordered physiological process that produces haploid male germ cells through more than 40 steps and experiences dramatic morphological and cellular transformations. The ubiquitin proteasome system (UPS) plays central roles in the precise control of protein homeostasis to ensure the effectiveness of certain protein groups at a given stage and the inactivation of them after this stage. Many UPS components have been demonstrated to regulate the progression of spermatogenesis at different levels. Especially in recent years, novel testis-specific proteasome isoforms have been identified to be essential and unique for spermatogenesis. In this review, we set out to discuss our current knowledge in functions of diverse USP components in mammalian spermatogenesis through: (1) the composition of proteasome isoforms at each stage of spermatogenesis; (2) the specificity of each proteasome isoform and the associated degradation events; (3) the E3 ubiquitin ligases mediating protein ubiquitination in male germ cells; and (4) the deubiquitinases involved in spermatogenesis and male fertility. Exploring the functions of UPS machineries in spermatogenesis provides a global picture of the proteome dynamics during male germ cell production and shed light on the etiology and pathogenesis of human male infertility.

1. Introduction

Spermatogenesis, which occurs within seminiferous tubules of male testes, ensures sustainable production of male gametes, the spermatozoa [1]. Generally, this sophisticated process could be divided into three main phases: (i) the mitotic phase that A single (As) spermatogonia undergo several rounds of mitosis to give rise to undifferentiated spermatogonia (A paired and A aligned), differentiating spermatogonia (A1, A2, A3, A4 and intermediate) and ultimately, type B spermatogonia, which are capable of entering meiosis and forming preleptotene spermatocytes; (ii) the meiotic phase that starts with complicated chromosomal behaviors in meiotic prophase I, such as DNA double strand break (DSB) formation and repair, homologous recombination, and synapsis, followed by two rounds of cell divisions to allow the separation of homologous chromosomes and sister chromatids and to produce haploid spermatids; (iii) the postmeiotic phase, or spermiogenesis, that haploid spermatids mature to motile spermatozoa via 8 steps of round spermatids and 8 steps of elongating spermatids [2,3] (Figure 1).
In mammals, this continuous spermatogenic program, from As spermatogonia to elongated spermatozoa, involves more than 40 steps and extends for months (42 days in mice and 85 days in humans, respectively) [4,5]. Through spermatogenesis, male germ cells experience dramatic changes in morphology, cellular components, genetics as well as epigenetics and therefore require coordinated, sequential functioning of different protein groups, both functional and structural, to ensure stage-specific molecular interactions and cellular events [6].

2. The Ubiquitination-Proteasome System

In all cell types, precise control of the effective window of a given protein relies largely on multiple machineries that regulate its abundance via tightly and precisely controlled protein synthesis and degradation. In recent years, genome-wide RNA-sequencing (RNA-seq) of purified male germ cells or isolated single cells has provided fruitful knowledge on the transcriptional regulations during mammalian spermatogenesis [7,8]. However, RNA-seq offers barely no information on protein degradation.
The ubiquitination-proteasome system (UPS) plays crucial roles in the regulation of protein degradation, and in some cases, protein activity [9,10] (Figure 2). Ubiquitin is a 76-amino-acid polypeptide with diversities only in 2 amino acids from yeast to human. Action of the UPS is initiated by a three-step enzyme cascade, consisted of ubiquitin-activating enzyme (E1), ubiquitin-conjugating enzyme (E2), and ubiquitin ligase (E3). This cascade transfers ubiquitin to the substrates via formation of covalent bonds between C-terminal glycine residues (Gly76) on the ubiquitin as well as lysine residues on the substrates or ubiquitin. The recognition between E3 ubiquitin ligases and their substrates is critical. In most cases, poly-ubiquitinated substrates are then recognized by the proteasomes for proteolysis and degradation, which terminates the function of the substrates [11]. However, in some cases, the ubiquitin chain could also be removed by deubiquitinating enzymes (DUBs), which recycles both functional proteins and ubiquitin [12].
Proteasomes, comprised of a 20S core particle (CP) and one or two regulatory particles (RPs), are central executors to the UPS and protein degradation [13]. The 20S CP is a 28-subunit complex with the molecular weight of around 750 kDa. These subunits are arranged as four stacked rings, two identical outer α rings and two identical inner β rings, forming a barrel-like structure. Each ring, either α or β, contains seven evolutionally-conserved subunits, i.e., α1–α7 and β1–β7, and among which, β1, β2 and β5 subunits possess the caspase-like, trypsin-like and chymotrypsin-like proteolysis activities, respectively [14,15]. This is the most common 20S proteasome species and is termed constitutive 20S proteasome, or c20S. In immunocytes or cells under inflammatory conditions, the three catalytic β subunits are replaced by β1i, β2i and β5i subunits, forming the so-called immune-20S proteasome, or i20S, which facilitates antigen presenting [16]. Recently, several studies uncovered the physiological functions of a male germ cell-specific spermato-20S proteasome, or s20S CP, in the regulation of homologous recombination and meiotic progression [17,18,19]. In the s20S CP, the α4 subunit is replaced by its analogue, α4s [20]. Accumulative evidences have shown that the 20S CP alone has proteolysis activity to degrade unfolded or disordered proteins under stress conditions [21,22].
However, the proteolytic activity and specificity of proteasomes are significantly increased by adding RPs to the 20S proteasome [13]. The RPs are attached to the α rings of the 20S CP, acting as the gates of the proteasome to recognize certain proteins for the subsequent proteolysis and meanwhile to prevent the degradation of other proteins. 19S RP is the most common RP and is made up of 17 subunits with a molecular weight of 700 kDa. The 19S RP binds to one or two ends of the 20S CP to form the 26S proteasome, which is the most abundant proteasome isoform responsible for the degradation of majority of poly-ubiquitinated proteins [23]. Besides the 19S RP, additional RPs are found to be associated with the 20S proteasomes, such as the PA28 RPs in immunocytes and the PA200 RP in male germ cells [24,25].
Events of proteasome degradation are common and crucial to spermatogenesis [26]. First of all, male germ cells have diverse proteasome isoforms, which possess specialized activities and exert their specific roles in protein degradation at distinct stages of spermatogenesis. Secondly, massive proteomic alterations take place during spermatogenic processes. For example, H2B dissociation from the chromosomes upon ubiquitination is important for both homologous recombination and meiotic double strand break (DSB) repair [27]. The dissociation of ubiquitinated H2A and H2B also ensures successful histone-to-protamine replacement during spermatid chromatin condensation [28]. In addition to histones, the tremendous loss of mitochondria in elongating spermatids is mostly carried out by the 26S proteasome as the form of enfolded cytoplasmic lobe, which contains both mitochondria and other organelles [29,30]. Single mitochondrial degradation also occurs when its outer membrane proteins (e.g., Mitofusins) are ubiquitinated and presented to the 26S proteasome [31]. Dissociation of mitochondrial outer membrane exposes its inner contents (e.g., prohibitin) for further degradation [30]. Finally, mutations of many UPS-related genes, including genes encoding E3 ubiquitin ligases, DUBs and proteasome components, results in dysregulated spermatogenesis and male infertility in humans and mice. In this review, we summarize the functions of the UPS components in spermatogenesis and intend to draw a comprehensive schematic of protein degradation during this process.

3. Proteasomes in Male Germ Cells

Testis specificity of the proteasomes relies largely on its unique isoforms of RP and CP subunits. In Drosophila, nearly one third of the 20S CP subunits possess alternative isoforms, the majority of which are solely expressed in testis [32]. For example, the α3 isoform, Prosα3T; the α4 isoform, Prosα4T1 (Pros28.1A) and Prosα4T2 (Pros28.1B); and the α6 isoform, Prosα6T [32,33,34]. These 20S isoforms are exclusively expressed in Drosophila testes from meiotic to post-meiotic germ cells, and gradually replace the constitutive isoforms to exerts their biological functions. Prosα6T, for instance, plays a pivotal role in sperm individualization and nuclear maturation, and therefore, Drosophila carrying Prosα6T mutation show disrupted individualization complexes and abnormal nuclear bundles [33].
Mammalian testis also has its distinctive features in proteasome composition, mainly due to the expression of male germ cell-associated proteasomal components, α4s and PA200, which are encoded by Psma8 and Psme4, respectively (Figure 2). A very recent preprint has revealed the complexity of proteasome composition in developing male germ cells by mass spectrometry (MS) [35]. Dramatic dynamics have been observed throughout the whole process of spermatogenesis, such as an increase in the amount of 20S proteasomes in spermatids, replacement of α4 (PSMA7) by α4s (PSMA8) in spermatocytes, as well as assembly of PA200-20S proteasomes in spermatocytes and spermatids. Therefore, it is of great significance to understand the stage-specific proteasome composition and relevant degradation events during spermatogenesis.

3.1. PSMA8

In mitotic spermatogonia and spermatocytes before the pachytene stage, around 98% 20S CP are made up of PSMA7, the constitutive α4 subunit [17,35]. However, from the pachytene stage onwards, the composition of 20S proteasomes are gradually shifted to PSMA8-associated s20S proteasomes [17,20]. This transition, from c20S to s20S, is mainly due to the expression of PSMA8 in pachynema. However, in PSMA8-null spermatocytes beyond the pachytene stage, the amount of c20S proteasomes is significantly decreased, suggesting that this c20S to s20S transition is an active process, through disassembly of the c20S and subsequent reassembly of the s20S (Figure 1) [17,19,35]. This type of male germ cell-specific s20S CP persists from meiotic spermatocytes to haploid spermatids and is required for proper protein degradation in these cells. Several recent studies have investigated the physiological and biochemical functions of PSMA8, or s20S CP, in spermatogenesis [17,18,19]. Although the exact function of PSMA8 in spermatogenesis remains controversial, one consensus that has been reached is its indispensability for Meiosis I (MI) progression. Spermatocytes of PSMA8-deficient mice show delayed entry into metaphase I and are finally arrested at this stage, resulting in few, if any, spermatids and mature sperm in the seminiferous tubules or epididymis, and ultimately, male infertility [17,18,19].
Progression from Prophase I to Anaphase I can be affected by a series of events that take place during Prophase I, such as assembly and disassembly of the synaptonemal complex (SC), generation and repair of meiotic DSBs, homologous recombination, and cell cycle control [3]. Most studies suggest that meiotic DSB generation and repair, synapsis, homologous recombination as well as crossing-over formation are less affected upon PSMA8-deletion [17]. This conclusion was supported by immunofluorescent (IF) staining of many protein markers (RPA proteins, ZMM proteins, MLH1, etc.), which are comparable in WT and PSMA8-deficient spermatocytes at similar developmental stages.
Efforts have been made to identify the direct targets of s20S CP, through methodologies of Western blotting (WB), IF staining and MS. In consistent with its function in protein proteolysis, the level of protein ubiquitination is accumulated in PSMA8-deleted testes [17,18]. Moreover, PSMA8-deficient male germ cells are incompetent to degrade certain Prophase I proteins (Figure 2). For example, RAD51 and RPA1 are highly expressed from zygotene to early-pachytene stages, but supposed to be degraded at late-pachytene and diplotene stages in normal spermatogenesis. However, these two proteins are found accumulated in PSMA8-deficient spermatocytes at late-pachytene and diplotene stages [17]. According to another study, PSMA8 is localized to the central region of SC, which is in consistent with a previous study demonstrating the chromosomal localization of proteasomes [18,36]. Correspondingly, abnormal accumulation of the transverse filament protein SYCP1 and the axial element protein SYCP3 is found in PSMA8-deficient spermatocytes. In a third paper, the authors proposed the functions of s20S CP, in complex with PA200, in the degradation of core histones upon DNA damage and during spermiogenesis [19]. Additionally, maturation-promoting factor (MPF) complex, consisting of cyclin-dependent kinase 1 (CDK1) and cyclin B, is central to cell cycle regulation. MPF activity is required for metaphase entry, however its inactivation is crucial for metaphase-anaphase transition. PSMA8 deficiency leads to elevated MPF activity, which is consistent with the phenotype of MI arrest [17,18].
Further experiments suggest that PA200 is co-immunoprecipitated with PSMA8. PSMA8 deficiency causes significant decrease in PA200 level, as well as the intensity of 20S catalytic subunits β1, β2 and β5, which ultimately results in abortive proteasome assembly [17,18]. Since the majority of 20S CPs in meiotic and post-meiotic male germ cells possess PSMA8 instead of PSMA7 and s20S associates with PA200, it was proposed that PA200-s20S is the main proteasome isoform functioning in protein degradation in spermatocytes and spermatids. However, another study found that other than PSMA8, more PSMA7 protein fragments are recovered in anti-PA200 immunoprecipitants, while on the other hand, s20S CP preferentially binds to the 19S RP, suggesting a controversial composition of proteasome isoforms in male germ cells [35]. This selectivity of combinations between RPs and CPs might be explained by the domain flexibility between PSMA7 and PSMA8. Nevertheless, the underlying mechanisms determining the differences between c20S and s20S CPs and the specificity of s20S CPs to RPs and protein substrates need further investigation.

3.2. PA200

PA200 is encoded by Blm10 in budding yeast, PSME4 in human and Psme4 in mouse, and is first identified as a 200-kDa monomeric novel proteasome activator in the nucleus [37]. PA200 associates with 20S CP in both bovine testes and HeLa cells and forms foci following irradiation, suggesting its role in DNA repair and chromatin remodeling [37,38]. Further studies suggest that the PA200 RP exhibits high testis preference and is associated with spermatogenesis and male sub-fertility by enforcing the degradation of core histones via an acetylation-dependent manner, during DSB repair and histone-protamine replacement [25,39].
In bovine testes, nearly 90% of the proteasomes contain PA200, while usually lower than 10% proteasomes are capped with PA200 in other tissues, for example, muscle [39]. PA200 deficiency causes significant abnormality in mouse spermatogenesis. PA200-deficient testis exhibits polynucleated giant germ cells, insufficient spermatogenesis, spermatic phagocytosis and accumulation of residual bodies [25,39]. Histone-protamine replacement is a hallmark of spermiogenesis and enables the compaction of sperm DNA into a higher architecture [40]. Despite the constitutive UPS pathway, core histone degradation during spermatogenesis can also be achieved by the testis-specific PA200-proteasome in an acetylation-dependent manner (Figure 2). Unlike the 19S RP, PA200 cannot recognize ubiquitin-linked protein substrates. Instead, PA200 contains an acetyl-lysine binding bromodomain (BRD) capable of recognizing acetylated histones [39]. The C-terminal residues of PA200 anchor to the binding pocket at the α1–α2 interface of 20S CP in the pattern of PA200-20S or PA200-20S-PA200 [39,41,42]. Binding of PA200 to acetylated substrates causes conformational rearrangements of the α rings, creating an open channel for substrate entry [42]. Changes of the α rings further extend to β subunits, resulting in increased trypsin-like, caspase-like and chymotryptic activities [35,42]. Consequently, large amount of H2A, H2B and H4 are found retaining in both round and elongating spermatids, suggesting PA200-deficient germ cells are incapable of histone degradation. The excessive histones can decrease genome stability and DNA damage resistance, so that the PA200-deficient testis reflects strong apoptotic and necrotic signals in both meiotic spermatocytes (mainly at and after the pachytene stage) and post-meiotic spermatids [25].
There are two pivotal stages of spermatogenesis which largely depend on acetylated histone dissociation from the nucleosomes, and are therefore vulnerable to PA200 deficiency. Besides the histone-protamine replacement during spermiogenesis, in meiotic Prophase I DSBs trigger the acetylation of core histones and their subsequent recognition and degradation by PA200-proteasomes [39]. Degradation of core histones from the DSB loci leaves space for the recruitment of DNA damage repair-related proteins. However, this point of view is controversial since some studies have reported that PA200 knockout mouse embryonic stem cells exhibit no significant impairment of their DSB repair ability upon DNA damage [38,41]. Besides, both lymphocyte development and immunoglobulin class switching, which depend on the generation and repair of DSBs, are not disrupted in mice with conventional PA200 knockout [25,37,38]. Moreover, PA200 also incorporates into the constitutive 26S proteasome, generating PA200-20S-19S hybrids [39,43]. This type of proteasome is supposed to be involved in the degradation of both ubiquitin- and acetyl-labelled substrates.
However, according to another study, at any stage of spermatogenesis (spermatogonia, spermatocytes and spermatids), the 19S RP is the main RP [35]. Even in the most PA200-abundant spermatocytes and spermatids samples, PA200-proteasomes only represents around 5% of the total proteasomes. Moreover, as stated in the PSMA8 section, 19S RP mostly associates with PSMA8 while PA200 has higher or similar affinity to PSMA7 in testis [35]. These biochemical evidences demonstrate that diverged proteasome species (including at least PA200-s20S, PA200-c20S, 19S-s20S) function simultaneously in a single cell at the same or different stages to regulate protein homeostasis. The specificity and activity of these proteasome species will be studied to illustrate the protein degradation events in spermatogenesis. Because most of the current studies were performed with mixed germ cell pools, it might be informative to reinvestigate or further investigate these results with purified synchronized male germ cells from each distinct stage.

4. E3 Ubiquitin Ligases in Spermatogenesis

Spermatogenesis requires efficient actions of E3 ubiquitin ligases which are master regulators of UPS to mediate protein turnover. As the last step of the enzymatic cascade, E3 confers substrate specificity and is of high abundance and diversity during spermatogenesis. Till now, at least 73 different E3 have been found highly or uniquely expressed in testis, with some of them having identified substrates and functions in spermatogenesis, such as CUL2, CUL4, UBR2 and RNF8 (Table 1) [44]. However, substrates of most testis-specific E3 ligases are still unclear so that the current understanding of their potential functions remains largely speculative and more comprehensive research are required in the future.

4.1. Cullin-Ring Ligase Family

Cullin-RING Ligases (CRLs) family of multi-subunit E3 ligase complexes is the largest E3 ligase family [97]. In CRL complex, Cullin protein is the scaffold protein, recruiting substrate-targeting modules on its N-terminal for substrate-targeting specificity and RING finger proteins, usually ROC1/2, on its C-terminal for E2 combination. The diversity of CRLs depends mainly on the substrate-targeting module, i.e., the adaptor protein and the substrate receptor. The mammalian Cullin protein family comprises CUL1 to 7, as well as PARC. These CRLs are involved in multiple biological processes like cell-cycle control, DNA replication and developmental regulation [98]. Among them, CRL1, CRL2, CRL3, CRL4A and CRL4B have been identified to be spermatogenesis-related.
The CRL1 complex, or SCFβ-TrCP, is the most known E3 ubiquitin ligase and is expressed ubiquitously in all cell types. This E3 complex comprises the scaffold protein CUL1, the adaptor protein SKP1 (S phase kinase-associated protein), the substrate receptor β-TrCP and the E2-interacting protein ROC1/2. There are two paralogs of β-TrCP in mammals, β-TrCP1 (BTRC) and β-TrCP2 (FBXW11) [69]. β-TrCP1 is expressed in spermatogonia at medium level but highly expressed in spermatocytes. In contrast, β-TrCP2 shows medium expression level in spermatogonia and low expression level in meiotic cells. β-TrCP1 deficiency in mice leads to prolonged and abnormal meiosis as indicated by accumulating MI spermatocytes and multinucleated spermatids [70]. In these β-TrCP1-deleted spermatocytes, stabilization of EMI1 is identified. EMI1 is a bona fide substrate of β-TrCP1, regulating metaphase to anaphase progression in both mitosis and meiosis via its inhibitive effects on the anaphase promoting complex/cyclosome, or APC/C [99]. While to progress through and exit from metaphase, APC/C needs to be reactivated, depending on EMI1 degradation. However, in spermatocytes lacking β-TrCP1, APC/C complex is inhibited by the accumulated level of EMI1. On the other hand, β-TrCP2 deficiency alone doesn’t cause any defects in spermatogenesis since β-TrCP1 is preserved at sufficient functional levels in all types of cells for compensation [69]. Double deletion of β-TrCP1 and β-TrCP2 leads to severe testis abnormality as indicated by absence of spermatocytes, spermatids and mature sperm. Surprisingly, these mice show dislocated spermatogonia in their seminiferous tubules, which is not observed in mice carrying either single deletion [69]. Snail1 is another substrate of β-TrCP, transcriptionally controlling the level of E-cadherin [100]. Deficiency of β-TrCP1 and β-TrCP2 results in a significantly increased level of Snail1 in these dislocated spermatogonia. As a consequence, accumulated Snail1 reduces E-cadherin and thereby disrupts the integrity of adherent junctions and the trafficking of differentiating cells from the basement of seminiferous tubules to the lumen [69].
Similar to the SCF complex, CRL4 complex is made up of the scaffold protein CUL4A/B, the adaptor protein DDB1 (DNA damage binding protein 1), the substrate receptor (DDB1- and CUL4-associated factors, DCAFs) and ROC1/2 [101]. CUL4A is predominantly expressed in primary spermatocytes at the pachytene and diplotene stages [45]. γH2AX indicates the sites of DSBs in meiotic prophase I and can be only detected on XY chromosome pairs in pachytene and diplotene spermatocytes. However, CUL4A deficiency leads to persistent γH2AX signals in mouse pachytene spermatocytes [46]. Similarly, foci of the late homologous recombination protein, MLH1, which are supposed to disappear from diplotene, remain detectable on chromosome pairs of diplotene spermatocytes [45]. As a consequence, significant decrease of post-meiotic spermatids and accumulation of malformed meiotic spermatocytes are evident in Cul4a knockout testes [45,46]. In contrast to CUL4A, CUL4B is expressed in spermatogonia and spermatids, predominantly regulating post-meiotic differentiation to guarantee sperm motility [45]. Although CUL4B-deficient mice is still proficient in producing spermatids, most of their spermatozoa possess low motility or become completely immotile, indicating normal meiosis but impaired post-meiotic specification [48]. Diminished mitochondrial activity with low level of ATP production and missing, superabundant or misshapen microtubule doublets on the flagella axonemes are found in spermatozoa derived from CUL4B-deficient mice. DDB1 is the adaptor protein and deletion of which abolishes the activity of all CRL4 complexes [101,102]. Strikingly, conditional deletion of DDB1 in male germ cells with a Ddx4-Cre results in complete loss of male germ cells [103]. DCAF12 is one of the substrate receptors of the CRL4 complexes and directly interacts with the Moloney Leukemia Virus 10 (MOV10) [47]. DCAF12 deficiency leads to decreased sperm counts accompanied by elevated level of MOV10. MOV10 is a RNA helicase, included in RNA-induced Silencing Complex (RISC), and plays a role in RNA interference and silencing of transposons, virus, and recently duplicated genes [104]. Increased amount of MOV10 results in the abnormal lower expression of two key meiotic proteins, SYCP3 and γH2AX, suggesting that CRL4A-DCAF12 mediates the degradation of MOV10 during spermatogenesis in a UPS dependent manner to maintain proper levels of meiotic-associated proteins. However, the entire substrates of CRL4 complexes in spermatogenesis remain largely elusive.
Besides SCF and CRL4 complexes, CRL2 and CRL3 also exhibit certain roles in spermatogenesis in C. elegans or Drosophila. In C. elegans, CUL2 is highly expressed in germline and participates in multiple developmental processes in complex with distinct substrate receptors [52]. The Leucine Rich Repeat protein LRR-1, a substrate receptor protein of CRL2, is expressed throughout the germline [53]. In testis, CUL2 associates with LRR-1 to assembly the CRL2 complex and acts with the 26S proteasome to degrade HTP-3, which is a meiotic-related protein recruiter in the mitotic zone of the germline, preparing chromosomes for meiosis [52,105]. Normally, HTP-3 expression shows a low-to-high gradient from the most distal germ cells to the most proximal ones. However, dysfunction of the CRL2 complex leads to uniform distribution of HTP-3 along C. elegans gonads, resulting in premature meiotic spermatocytes [105].
Drosophila has four CUL3 isoforms and one of them is expressed exclusively in male germ cells (CUL3testis), playing an essential role during post-meiotic differentiation. CUL3testis combines with its adaptor protein Klhl10 and targets dBruce, a caspase inhibitor for degradation [49,50]. This CRL3 E3 ubiquitin ligase complex thereby ensures active function of apoptotic proteins to eliminate superfluous cytoplasm as well as unwanted organelles during post-meiotic differentiation. In mice, CUL3 emerges from post-meiotic differentiation step 9 and keeps detectable through the remaining stages of this process [50]. However, the specific roles of mammalian CUL3 during spermatogenesis remain unsolved since CUL3 deficiency in mice unexpectedly leads to early embryonic lethality by interfering cell cycle progression [51].

4.2. Other E3 Ubiquitin Ligases

Besides the CRLs family, many other E3 ubiquitin ligases also participate in the regulation of spermatogenesis. In addition to be regulated by the SCF-EMI1 ubiquitination pathway, APC/C itself is an E3 ubiquitin ligase complex and mediates poly-ubiquitination of substrates for proteasomal degradation [106]. During mitosis, APC/C associates with the spindle apparatus to promote proper separation of sister chromatids by mediating cyclin B1 and securin degradation [107]. In addition, APC/C functions to degrade proteins that are necessary for S phase and inactivation of APC/C has been proposed as the commitment point for cell-cycle entry [108]. It is likely that these functions also apply to meiosis, albeit not yet fully characterized. However, a unique role of APC/C in meiosis has been proposed as mediating piRNA-triggered PIWI destruction in late spermatids (both elongating and elongated spermatids) [60]. PIWI-interacting RNA (piRNA) is a novel germline-specific small noncoding RNA and cooperates with PIWI to maintain genome integrity by silencing transposons [109]. Spermatocytes and round spermatids show high level of piRNA/PIWI, while this level begins to decrease in late spermatids, and is ultimately eliminated in mature sperm. Knockdown of endogenous APC/C abolishes PIWI ubiquitination and degradation in late spermatids, resulting in maturation stagnation of these cells [60]. Besides, piRNA is synchronously degraded with PIWI in late spermatids, suggesting a feedforward mechanism for their elimination. That is, piRNA binds to PIWI during late spermatogenesis, causing a conformational change of PIWI, which facilitates its binding to APC/C. Meanwhile, PIWI degradation in turn leaves the piRNA unprotected and therefore primed for elimination [60].
SIAH1A confers cell growth arrest in somatic cells, while in male germ cells, it plays an important role to promote spermatogenesis progression beyond MI [110]. Siah1a deficiency in mice induces severe impact on spermatogenesis characterized by large amounts of spermatocytes accumulated at metaphase to telophase of meiosis I. Besides, anaphase cells are largely bi- or multi-nucleated because of incomplete chromosome segregation [66]. Since ubiquitin-dependent degradation of anaphase inhibitors such as securin is crucial for the transition from metaphase to anaphase [111], it is likely that E3 ubiquitin ligase SIAH1A also targets these proteins as substrates to facilitate anaphase entry.
HEI10 is an E3 ubiquitin ligase important for the regulation of homologous recombination during pachytene [68]. Following DSB repair mediated by RAD51, HEI10 mediates the degradation of CCNB3 to free CDK2. In HEI10 mutant mouse spermatocytes, homologous recombination initiates normally with proper DSB formation and repair. However, these spermatocytes fail to form chiasmata which leads to premature homolog separation, and ultimately undergo meiotic arrest at MI or apoptosis [112].
RNF8, or RING finger protein 8, is essentially a DNA damage response (DDR) protein that ubiquitinates histone H2A and H2AX to facilitate downstream recruitment of DDR factors to DSB-flanking chromosomes [55]. In pachytene spermatocytes, ubiquitinated H2AX (ubH2AX) is significantly enriched in the XY body, a process known as meiotic sex chromosome inactivation, or MSCI [113]. RNF8 deficiency leads to severe loss of ubH2AX in the XY body, indicating a role of RNF8 in ubiquitinating H2AX during this process. However, neither XY body formation nor the subsequent meiosis is affected upon RNF8 deficiency, indicating that RNF8 may facilitate but is not necessary for MSCI or meiosis. In contrast, RNF8 is indispensable during post-meiotic differentiation, when nucleosome removal occurs accompanied by H2A and H2B ubiquitination [55]. RNF8 deficiency in mice leads to reduced ubH2A and ubH2B in elongating spermatids as well as impaired histone-protamine transition. Sperm production in RNF8-deficient mice is significantly impaired with accumulation of defective sperm possessing malformed sperm heads [55]. Notably, although RNF8 itself has no acetyltransferase activity, it is supposed that RNF8 indirectly regulates mammalian acetyltransferase MOF through ubH2A and ubH2B. Therefore, RNF8 deficiency leads to decreased MOF level as well as H4 acetylation [114]. RNF8 also participates in epigenetic modification in post-meiotic spermatids. One study has shown that through H2A ubiquitination, RNF8 is able to achieve active epigenetic modifications such as H3K4 dimethylation (H3K4me2), which further activate escaped genes from silenced sex chromosomes in post-meiotic spermatids [58].
UBR2 also targets histone H2A for ubiquitination conjugation. UBR2 primarily localizes to the unsynapsed chromatin regions where it transfers ubiquitin from its E2 enzyme HR6B to histone H2A, mediating meiotic silencing of unsynapsed chromatin (MSUC) [64]. Furthermore, UBR2-mediated histone ubiquitination is linked to a pachytene checkpoint. UBR2 deficiency impairs multiple meiotic processes such as homologous recombination, MSUC and synapsis, which trigger meiotic arrest through the pachytene checkpoint system [61,62,65].

5. Protein Deubiquitination in Spermatogenesis

Deubiquitinases (DUBs) are cysteine proteases, which remove and recycle ubiquitin. Till now, more than 100 DUB enzymes have been identified in human [115]. DUBs participate in various biological processes such as apoptosis, cell cycle progression, and protection of proteins from degradation. DUBs are classified into six catalogues: USP (ubiquitin-specific processing proteases), UCH (ubiquitin C-terminal hydrolases), JAMM (Jab1/Pab1/MPN domain-containing metalloenzymes), OUT (Otu-domain ubiquitin aldehyde-binding proteins), MCPIPs (Monocyte chemotactic protein-induced proteases) and Ataxin-3/Josephin [116]. Some DUBs have been demonstrated to play a requisite role during spermatogenesis, which largely belong to the USP and UCH family (Table 2) [115].
USP7 localizes preferentially to the XY body in early pachytene spermatocytes but gradually decreases its level as meiosis progresses, which is highly concurrent with the expression manner of SCML2, a testis-specific polycomb protein [119]. Biochemically, USP7 forms a complex with SCML2 to counteract histone H2A ubiquitination in the XY chromatin during meiosis. In SCML2-deficient mice, USP7 is absent from H2A ubiquitination sites, leading to augmented H2A monoubiquitination, which in turn causes spermatogenic impairments characterized by massive pachytene spermatocyte apoptosis [119]. Though the direct effects of USP7 have not been studied yet, it can be drawn from this study that localization of USP7 to the XY body relies on SCML2 and these two proteins cooperate to regulate male meiosis, most likely trough mediating transcriptional silencing.
USP8 is expressed in both brain and testis and interacts with MSJ1, STAM2, EEA1 and VSP54 [121]. In mouse testis, USP8 shows remarkable increase during post-meiotic differentiation and forms small plaques. USP8 localizes specifically to the nuclear envelope of round spermatids, as well as the centrosome and acrosome vesicle of elongating spermatids [120]. In mouse spermatids, USP8 associates with MSJ1 and the 20S CP to move towards the developing acrosome and centrosome [121,145]. STAM2 gives rise to the endosomal-sorting complex ESCRT-0 and EEA1 is an early endosome antigen. Colocalization of USP8 with STAM2 and EEA1 is found at acrosomal vacuole and acrosome surface, respectively [146]. VSP54 is responsible for retrograde transport from early endosomes, and during post-meiotic differentiation, it follows the same migration trajectory as USP8 until complete acrosome formation is achieved [147]. Taken together, these results suggest that USP8 is a key participant of the endosome pathway during acrosome generation.
USP2 is restrictedly expressed in elongating spermatids, possessing speculative function during post-meiotic differentiation. USP2 deficiency in mice leads to defective sperm whose motility is highly vulnerable to environmental changes [118]. Besides, these sperm exhibit poor fertilization capacity due to failure in binding or penetrating to the zona pellucida. USP14 is important for post-meiotic differentiation. In Drosophila, USP14 deficiency impairs spermatid individualization, during which syncytial spermatids are separated into individual cells [128]. USP14 deficiency causes loss of synchronization and abnormal distribution of the actin cones. Besides, USP14-deficient mice exhibit significantly reduced sperm counts as well as severe sperm malformation such as multiple nucleus, missing of sperm head or dual sperm tails [127].
Mammalian USP9X is the functional orthologue of drosophila deubiquitinating enzyme fat facets (Faf), which interacts with and impedes the degradation of a DEAD-box RNA helicase Vasa, a highly-conserved marker for germ cells [148]. USP9X is predominantly expressed in spermatogonia and weakly expressed in early spermatocytes before pachytene stage. USP9X-deficiency in mouse germ cells leads to male infertility with various abnormalities along the progression of spermatogenesis: reduced number of spermatocytes, degenerated spermatids with residual body-like structures, as well as aberrantly retained mature spermatozoa in seminiferous tubules [126]. However, the maintenance and proliferation of spermatogonia were less affected in USP9X-deficient testes, indicating a critical role of USP9X since mitosis-to-meiosis transition in male germ cells.
USP26 has been considered as a potential infertility gene due to its restricted expression in mammalian testis and X-chromosome-localization (single copy in males) [149,150,151]. Several studies have reported the polymorphisms in USP26 associated with non-obstructive azoospermia or asthenozoospermia, suggesting a causal relationship with human male infertility. However, most of the identified USP26 polymorphisms can’t disrupt its enzymatic function [152]. In addition, USP26 is indispensable for mouse fertility in both sexes [153]. It has recently been found the effects of USP26 mutation on male fertility dependent on the genetic background of mice [154]. USP26 mutants in DBA/2, rather than C57BL/6 background exhibit impaired spermatogenesis with obvious deficiency and malformation of spermatozoa, resulting in infertile or sub-fertile males. These results implicate that strain-specific genetic components interact with USP26 mutations to interfere spermatogenesis; such components may also exist in human but further investigations are needed.
UCHL3 expression exhibits differentiation-dependent pattern during spermatogenesis with minimal amount in spermatogonia but to a sequentially increasing extent in meiotic pachytene spermatocytes and post-meiotic spermatids [141]. Besides, a later study has detected intensive UCHL3 expression in sperm acrosomes and flagella, suggesting the role of UCHL3 in regulating both meiosis and post-meiotic differentiation [155]. This study also examines the UCHL3 level in patients with asthenozoospermia (A) and oligoasthenozoospermia (OA) who possess malformed spermatozoa as characterized by small or absent acrosome, as well as atrophic or distorted tail. According to the results, OA patients show significantly lower UCHL3 content and activity as compared to normozoospermia controls; UCHL3 condition is slightly better in A patients but still far from normal level [155]. Meanwhile, it is proposed that the amount and activity of UCHL3 are positively related to a series of fertility indicators including sperm counts, sperm concentration and sperm motility, highlighting the importance of this DUB during spermatogenesis.
UCHL1 shares high sequence similarity with UCHL3, however, its distribution pattern in male germ cells is totally different, indicating distinct function during spermatogenesis [156]. Predominant UCHL1 expression is found in spermatogonia as well as in Sertoli cells. The level of UCHL1 precisely determines spermatogonia fate of either self-renewal or meiotic differentiation as indicated by two markers Plzf and c-Kit, respectively [156]. In mouse testis, UCHL1 overexpression leads to drastic loss of post-meiotic germ cells and a large amount of arrested pachytene spermatocytes retards in seminiferous tubules which further undergo apoptosis [138]. On the contrary, UCHL1 deficiency significantly increases the number of spermatogonia and preleptotene spermatocytes while immensely inhibits germ cell apoptosis during the first round of spermatogenesis. Additionally, these mice have reduced sperm motility and more abnormal spermatozoa even though complete sterility is not observed [139]. Taken together, UCHL1 and UCHL3 express strongly but reciprocally during spermatogenesis, where the former mediates fate determination of spermatogonia as well as elimination of defective spermatozoa to maintain testicular homeostasis, while the latter facilitates post meiotic maturation to produce fertilization-competent spermatozoa.
Besides the DUBs in USP and UCH families, CYLD (cylindromatosis) is also found to be required for spermatogenesis by regulating early wave of germ cell apoptosis mainly in spermatogonia, a requisite process to eliminate excessive germ cells and thereby maintain the balance between germ cells and Sertoli cells [157]. CYLD directly deubiquitinates the receptor-interacting protein 1 (RIP1), to prevent the activation of IKK and NF-𝜅B signaling and the expression of downstream anti-apoptotic genes, ultimately promote cell apoptosis [144]. Loss of CYLD leads to reduced spermatozoa, failure in radial organization of round spermatids, as well as malformation of acrosomes in elongating spermatids. On the contrary, spermatogonia and early spermatocytes are aberrantly accumulated in CYLD-deficient seminiferous tubules.

6. Conclusions and Perspective

For a prolonged and highly ordered physiological process such as spermatogenesis, the precise control of protein homeostasis to ensure the functioning of certain protein groups at a given stage and the inactivation of them after this stage (actively or passively) is critical for the step-by-step progression. In this review, we summarized the biochemical and physiological functions of the UPS components in male spermatogenesis, including the diverse proteasomes that target ubiquitinated proteins for proteolysis, the E3 ubiquitin ligases that tag their substrates for ubiquitination, as well as the DUBs that remove ubiquitin chains from the substrates. We propose further investigation on the underlying mechanisms of UPS-mediated protein homeostasis during spermatogenesis, in order to reveal the unique protein degradation events in male germ cells, interpret the specificity of testis-associated proteasome isoforms in targeting different groups of substrates, and clarify the pathogenesis of human male infertility related to UPS dysfunction. However, most related studies do not take the big picture of the UPS into consideration. For example, some studies only investigate an E3 and sometimes including its substrates but do not analyze the following degradation events, or some other studies focus on a proteasome isoform but omit its substrate specificity and the biochemical uniqueness. The main reason should be the complexity of both the UPS and the spermatogenic procedure, and one way to improve it might be focusing on a certain protein in spermatogenesis, to systematically investigate its expression pattern transcriptionally and translationally, its E3 and DUB, the ubiquitination cascade and the proteasome isoform that proteolyzes it. The other way is to investigate the proteomic alterations during each step of spermatogenesis with advanced techniques, such as quantitative MS and ubiquitination-related MS.
Since the regulation of protein ubiquitination and degradation is always specific and precise to a narrow time window, the significance of a single degradation event might be underestimated, if the result is drawn with male germ cells of mixed stages. Due to the development of a WIN18,446/RA-based protocol to synchronize spermatogenesis into a “tsunami” [158], it is possible to study proteomic changes with synchronized, pure germ cell samples. And by investigating the specialized physiological process of spermatogenesis, we may reveal the unique regulatory mechanisms of UPS in physiological conditions and broaden our understanding of both UPS and spermatogenesis.
Around 8–12% of couples worldwide suffer from infertility and male factors contribute to up to 20–30% of the cases independently [159]. Causes of male infertility, such as azoospermia, oligozoospermia, teratozoospermia, etc., have been comprehensively studied in both medical and basic research, which largely influence the therapeutic methods and success rates in IVF-ET (In vitro fertilization and embryo transfer). Regulatory functions of UPS components found in mouse spermatogenesis have been confirmed in humans. For example, PA200-proteasomes are also responsible for histone-protamine transition in human spermiogenesis; deletions of USP9Y have been found in infertile male patients; declined level and activity of UCHL3 cause defects in human sperm count and motility [124,155,160]. Therefore, identification of novel genes in UPS correlating with spermatogenetic failure could unravel the genetic cause of male infertility and promote the prevalence of PGD (Preimplantation genetic diagnosis) before ICSI (Intracytoplasmic sperm injection), ultimately improve the birth quality of human beings.

Author Contributions

Y.X. and Q.Z. reviewed the literature and wrote the manuscript, while C.Y. supervised it and did revision. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by National Natural Science Foundation of China (82171598 to C.Y.), and Zhejiang Provincial Natural Science Fund (LR21C120001 to C.Y. and LQ21C050002 to Q.Z.).

Acknowledgments

We thank all the members in the laboratory of Qianting Zhang for discussions. Due to limited space, we apologize to those whose research work is not cited in this manuscript.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Hess, R.A.; de Franca, L.R. Spermatogenesis and cycle of the seminiferous epithelium. Adv. Exp. Med. Biol. 2008, 636, 1–15. [Google Scholar] [CrossRef] [PubMed]
  2. Oatley, J.M.; Brinster, R.L. The germline stem cell niche unit in mammalian testes. Physiol. Rev. 2012, 92, 577–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hunter, N. Meiotic recombination: The essence of heredity. Cold Spring Harb. Perspect. Biol. 2015, 7, a016618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Adler, I.D. Comparison of the duration of spermatogenesis between male rodents and humans. Mutat. Res. 1996, 352, 169–172. [Google Scholar] [CrossRef]
  5. Marchetti, F.; Wyrobek, A.J. Mechanisms and consequences of paternally-transmitted chromosomal abnormalities. Birth Defects Res. Part. C Embryo Today Rev. 2005, 75, 112–129. [Google Scholar] [CrossRef]
  6. Griswold, M.D. Spermatogenesis: The commitment to meiosis. Physiol. Rev. 2016, 96, 1–17. [Google Scholar] [CrossRef] [Green Version]
  7. Chen, Y.; Zheng, Y.; Gao, Y.; Lin, Z.; Yang, S.; Wang, T.; Wang, Q.; Xie, N.; Hua, R.; Liu, M.; et al. Single-cell RNA-seq uncovers dynamic processes and critical regulators in mouse spermatogenesis. Cell Res. 2018, 28, 879–896. [Google Scholar] [CrossRef] [Green Version]
  8. Soumillon, M.; Necsulea, A.; Weier, M.; Brawand, D.; Zhang, X.; Gu, H.; Barthès, P.; Kokkinaki, M.; Nef, S.; Gnirke, A.; et al. Cellular source and mechanisms of high transcriptome complexity in the mammalian testis. Cell Rep. 2013, 3, 2179–2190. [Google Scholar] [CrossRef]
  9. Glickman, M.H.; Ciechanover, A. The ubiquitin-proteasome proteolytic pathway: Destruction for the sake of construction. Physiol. Rev. 2002, 82, 373–428. [Google Scholar] [CrossRef]
  10. Sadowski, M.; Suryadinata, R.; Tan, A.R.; Roesley, S.N.A.; Sarcevic, B. Protein monoubiquitination and polyubiquitination generate structural diversity to control distinct biological processes. IUBMB Life 2012, 64, 136–142. [Google Scholar] [CrossRef]
  11. Hochstrasser, M. Ubiquitin-dependent protein degradation. Annu. Rev. Genet. 1996, 30, 405–439. [Google Scholar] [CrossRef]
  12. Nijman, S.M.B.; Luna-Vargas, M.P.A.; Velds, A.; Brummelkamp, T.R.; Dirac, A.M.G.; Sixma, T.K.; Bernards, R. A genomic and functional inventory of deubiquitinating enzymes. Cell 2005, 123, 773–786. [Google Scholar] [CrossRef] [Green Version]
  13. Tanaka, K. The proteasome: Overview of structure and functions. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2009, 85, 12–36. [Google Scholar] [CrossRef] [Green Version]
  14. Huang, X.; Luan, B.; Wu, J.; Shi, Y. An atomic structure of the human 26S proteasome. Nat. Struct. Mol. Biol. 2016, 23, 778–785. [Google Scholar] [CrossRef] [PubMed]
  15. Groll, M.; Ditzel, L.; Löwe, J.; Stock, D.; Bochtler, M.; Bartunik, H.D.; Huber, R. Structure of 20S proteasome from yeast at 2.4 Å resolution. Nature 1997, 386, 463–471. [Google Scholar] [CrossRef]
  16. Tanaka, K.; Kasahara, M. The MHC class I ligand-generating system: Roles of immunoproteasomes and the interferon-gamma-inducible proteasome activator PA28. Immunol. Rev. 1998, 163, 161–176. [Google Scholar] [CrossRef]
  17. Zhang, Q.; Ji, S.-Y.; Busayavalasa, K.; Shao, J.; Yu, C. Meiosis I progression in spermatogenesis requires a type of testis-specific 20S core proteasome. Nat. Commun. 2019, 10, 3387. [Google Scholar] [CrossRef] [Green Version]
  18. Gómez-H, L.; Felipe-Medina, N.; Condezo, Y.B.; Garcia-Valiente, R.; Ramos, I.; Suja, J.A.; Barbero, J.L.; Roig, I.; Sánchez-Martín, M.; de Rooij, D.G.; et al. The PSMA8 subunit of the spermatoproteasome is essential for proper meiotic exit and mouse fertility. PLoS Genet. 2019, 15, e1008316. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, Z.-H.; Jiang, T.-X.; Chen, L.-B.; Zhou, W.; Liu, Y.; Gao, F.; Qiu, X.-B. Proteasome subunit alpha4s is essential for formation of spermatoproteasomes and histone degradation during meiotic DNA repair in spermatocytes. J. Biol. Chem. 2021, 296, 100130. [Google Scholar] [CrossRef]
  20. Uechi, H.; Hamazaki, J.; Murata, S. Characterization of the testis-specific proteasome subunit alpha4s in mammals. J. Biol. Chem. 2014, 289, 12365–12374. [Google Scholar] [CrossRef] [Green Version]
  21. Asher, G.; Reuven, N.; Shaul, Y. 20S proteasomes and protein degradation “by default”. Bioessays 2006, 28, 844–849. [Google Scholar] [CrossRef]
  22. Tenzer, S.; Stoltze, L.; Schönfisch, B.; Dengjel, J.; Müller, M.; Stevanović, S.; Rammensee, H.-G.; Schild, H. Quantitative analysis of prion-protein degradation by constitutive and immuno-20S proteasomes indicates differences correlated with disease susceptibility. J. Immunol. 2004, 172, 1083–1091. [Google Scholar] [CrossRef] [PubMed]
  23. Coux, O.; Zieba, B.A.; Meiners, S. The proteasome system in health and disease. Adv. Exp. Med. Biol. 2020, 1233, 55–100. [Google Scholar] [CrossRef] [PubMed]
  24. Preckel, T.; Fung-Leung, W.P.; Cai, Z.; Vitiello, A.; Salter-Cid, L.; Winqvist, O.; Wolfe, T.G.; Von Herrath, M.; Angulo, A.; Ghazal, P.; et al. Impaired immunoproteasome assembly and immune responses in PA28-/- mice. Science 1999, 286, 2162–2165. [Google Scholar] [CrossRef] [PubMed]
  25. Khor, B.; Bredemeyer, A.L.; Huang, C.-Y.; Turnbull, I.R.; Evans, R.; Maggi, L.B., Jr.; White, J.M.; Walker, L.M.; Carnes, K.; Hess, R.A.; et al. Proteasome activator PA200 is required for normal spermatogenesis. Mol. Cell. Biol. 2006, 26, 2999–3007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Sutovsky, P. Ubiquitin-dependent proteolysis in mammalian spermatogenesis, fertilization, and sperm quality control: Killing three birds with one stone. Microsc. Res. Tech. 2003, 61, 88–102. [Google Scholar] [CrossRef] [PubMed]
  27. Robzyk, K.; Recht, L.; Osley, M.A. Rad6-dependent ubiquitination of histone H2B in yeast. Science 2000, 287, 501–504. [Google Scholar] [CrossRef]
  28. Lewis, J.D.; Abbott, D.W.; Ausió, J. A haploid affair: Core histone transitions during spermatogenesis. Biochem. Cell Biol. 2003, 81, 131–140. [Google Scholar] [CrossRef]
  29. Hecht, N.B. The making of a spermatozoon: A molecular perspective. Dev. Genet. 1995, 16, 95–103. [Google Scholar] [CrossRef]
  30. Karbowski, M.; Youle, R.J. Regulating mitochondrial outer membrane proteins by ubiquitination and proteasomal degradation. Curr. Opin. Cell Biol. 2011, 23, 476–482. [Google Scholar] [CrossRef] [Green Version]
  31. Ziviani, E.; Tao, R.N.; Whitworth, A.J. Drosophila parkin requires PINK1 for mitochondrial translocation and ubiquitinates mitofusin. Proc. Natl. Acad. Sci. USA 2010, 107, 5018–5023. [Google Scholar] [CrossRef] [Green Version]
  32. Ma, J.; Katz, E.; Belote, J.M. Expression of proteasome subunit isoforms during spermatogenesis in Drosophila melanogaster. Insect Mol. Biol. 2002, 11, 627–639. [Google Scholar] [CrossRef]
  33. Zhong, L.; Belote, J.M. The testis-specific proteasome subunit Prosalpha6T of D. melanogaster is required for individualization and nuclear maturation during spermatogenesis. Development 2007, 134, 3517–3525. [Google Scholar] [CrossRef] [Green Version]
  34. Yuan, X.; Miller, M.; Belote, J.M. Duplicated proteasome subunit genes in Drosophila melanogaster encoding testes-specific isoforms. Genetics 1996, 144, 147–157. [Google Scholar] [CrossRef]
  35. Zivkovic, D.; Dafun, A.S.; Menneteau, T.; Schahl, A.; Lise, S.; Kervarrec, C.; Rêgo, A.T.; da Fonseca, P.C.A.; Chavent, M.; Pineau, C.; et al. Proteasome complexes experience profound structural and functional rearrangements throughout mammalian spermatogenesis. Biorxiv Prepr. Serv. Biol. 2021. preprint. [Google Scholar] [CrossRef]
  36. Rao, H.B.D.P.; Qiao, H.; Bhatt, S.K.; Bailey, L.R.; Tran, H.D.; Bourne, S.L.; Qiu, W.; Deshpande, A.; Sharma, A.N.; Beebout, C.J.; et al. A sumo-ubiquitin relay recruits proteasomes to chromosome axes to regulate meiotic recombination. Science 2017, 355, 403–407. [Google Scholar] [CrossRef] [Green Version]
  37. Ustrell, V.; Hoffman, L.; Pratt, G.; Rechsteiner, M. PA200, a nuclear proteasome activator involved in DNA repair. EMBO J. 2002, 21, 3516–3525. [Google Scholar] [CrossRef] [Green Version]
  38. Blickwedehl, J.; McEvoy, S.; Wong, I.; Kousis, P.; Clements, J.; Elliott, R.; Cresswell, P.; Liang, P.; Bangia, N. Proteasomes and proteasome activator 200 kda (PA200) accumulate on chromatin in response to ionizing radiation. Radiat. Res. 2007, 167, 663–674. [Google Scholar] [CrossRef]
  39. Qian, M.-X.; Pang, Y.; Liu, C.H.; Haratake, K.; Du, B.-Y.; Ji, D.-Y.; Wang, G.-F.; Zhu, Q.-Q.; Song, W.; Yu, Y.; et al. Acetylation-mediated proteasomal degradation of core histones during DNA repair and spermatogenesis. Cell 2013, 153, 1012–1024. [Google Scholar] [CrossRef] [Green Version]
  40. Braun, R.E. Packaging paternal chromosomes with protamine. Nat. Genet. 2001, 28, 10–12. [Google Scholar] [CrossRef]
  41. Savulescu, A.F.; Glickman, M.H. Proteasome activator 200: The heat is on. Mol. Cell Proteom. 2011, 10, R110.006890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Guan, H.; Wang, Y.; Yu, T.; Huang, Y.; Li, M.; Saeed, A.F.U.H.; Perčulija, V.; Li, D.; Xiao, J.; Wang, D.; et al. Cryo-EM structures of the human PA200 and PA200-20S complex reveal regulation of proteasome gate opening and two PA200 apertures. PLoS Biol. 2020, 18, e3000654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hou, C.-C.; Yang, W.-X. New insights to the ubiquitin-proteasome pathway (UPP) mechanism during spermatogenesis. Mol. Biol. Rep. 2013, 40, 3213–3230. [Google Scholar] [CrossRef]
  44. Richburg, J.H.; Myers, J.L.; Bratton, S.B. The role of E3 ligases in the ubiquitin-dependent regulation of spermatogenesis. Semin. Cell Dev. Biol. 2014, 30, 27–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yin, Y.; Lin, C.; Kim, S.T.; Roig, I.; Chen, H.; Liu, L.; Veith, G.M.; Jin, R.U.; Keeney, S.; Jasin, M.; et al. The E3 ubiquitin ligase cullin 4A regulates meiotic progression in mouse spermatogenesis. Dev. Biol. 2011, 356, 51–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Kopanja, D.; Roy, N.; Stoyanova, T.; Hess, R.A.; Bagchi, S.; Raychaudhuri, P. Cul4A is essential for spermatogenesis and male fertility. Dev. Biol. 2011, 352, 278–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Lidak, T.; Baloghova, N.; Korinek, V.; Sedlacek, R.; Balounova, J.; Kasparek, P.; Cermak, L. CRL4-DCAF12 ubiquitin ligase controls MOV10 RNA helicase during spermatogenesis and T cell activation. Int. J. Mol. Sci. 2021, 22, 5394. [Google Scholar] [CrossRef]
  48. Yin, Y.; Liu, L.; Yang, C.; Lin, C.; Veith, G.M.; Wang, C.; Sutovsky, P.; Zhou, P.; Ma, L. Cell autonomous and nonautonomous function of CUL4B in mouse spermatogenesis. J. Biol. Chem. 2016, 291, 6923–6935. [Google Scholar] [CrossRef] [Green Version]
  49. Arama, E.; Bader, M.; Rieckhof, G.E.; Steller, H. A ubiquitin ligase complex regulates caspase activation during sperm differentiation in Drosophila. PLoS Biol. 2007, 5, e251. [Google Scholar] [CrossRef] [Green Version]
  50. Wang, S.H.; Zheng, H.L.; Esaki, Y.; Kelly, F.; Yan, W. Cullin3 is a KLHL10-interacting protein preferentially expressed during late spermiogenesis. Biol. Reprod. 2006, 74, 102–108. [Google Scholar] [CrossRef]
  51. Singer, J.D.; Gurian-West, M.; Clurman, B.; Roberts, J.M. Cullin-3 targets cyclin E for ubiquitination and controls S phase in mammalian cells. Genes Dev. 1999, 13, 2375–2387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Burger, J.; Merlet, J.; Tavernier, N.; Richaudeau, B.; Arnold, A.; Ciosk, R.; Bowerman, B.; Pintard, L. CRL2(LRR-1) E3-ligase regulates proliferation and progression through meiosis in the Caenorhabditis elegans germline. PLoS Genet. 2013, 9, e1003375. [Google Scholar] [CrossRef] [Green Version]
  53. Merlet, J.; Burger, J.; Tavernier, N.; Richaudeau, B.; Gomes, J.-E.; Pintard, L. The CRL2LRR-1ubiquitin ligase regulates cell cycle progression during C. elegans development. Development 2010, 137, 3857–3866. [Google Scholar] [CrossRef] [Green Version]
  54. Huen, M.S.Y.; Grant, R.; Manke, I.; Minn, K.; Yu, X.; Yaffe, M.B.; Chen, J. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 2007, 131, 901–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lu, L.-Y.; Wu, J.; Ye, L.; Gavrilina, G.B.; Saunders, T.L.; Yu, X. RNF8-dependent histone modifications regulate nucleosome removal during spermatogenesis. Dev. Cell 2010, 18, 371–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li, L.; Halaby, M.-J.; Hakem, A.; Cardoso, R.; El Ghamrasni, S.; Harding, S.; Chan, N.; Bristow, R.; Sanchez, O.; Durocher, D.; et al. Rnf8 deficiency impairs class switch recombination, spermatogenesis, and genomic integrity and predisposes for cancer. J. Exp. Med. 2010, 207, 983–997. [Google Scholar] [CrossRef] [Green Version]
  57. Ma, T.; Keller, J.A.; Yu, X.C. RNF8-dependent histone ubiquitination during DNA damage response and spermatogenesis. Acta Bioch. Bioph. Sin. 2011, 43, 339–345. [Google Scholar] [CrossRef] [Green Version]
  58. Sin, H.S.; Barski, A.; Zhang, F.; Kartashov, A.V.; Nussenzweig, A.; Chen, J.; Andreassen, P.R.; Namekawa, S.H. RNF8 regulates active epigenetic modifications and escape gene activation from inactive sex chromosomes in post-meiotic spermatids. Genes Dev. 2012, 26, 2737–2748. [Google Scholar] [CrossRef] [Green Version]
  59. Ozata, D.M.; Gainetdinov, I.; Zoch, A.; O’Carroll, D.; Zamore, P.D. PIWI-interacting RNAs: Small RNAs with big functions. Nat. Rev. Genet. 2019, 20, 89–108. [Google Scholar] [CrossRef] [Green Version]
  60. Zhao, S.; Gou, L.-T.; Zhang, M.; Zu, L.-D.; Hua, M.-M.; Hua, Y.; Shi, H.-J.; Li, Y.; Li, J.; Li, D.; et al. piRNA-triggered MIWI ubiquitination and removal by APC/C in late spermatogenesis. Dev. Cell 2013, 24, 13–25. [Google Scholar] [CrossRef] [Green Version]
  61. Kwon, Y.T.; Xia, Z.; An, J.Y.; Tasaki, T.; Davydov, I.V.; Seo, J.W.; Sheng, J.; Xie, Y.; Varshavsky, A. Female lethality and apoptosis of spermatocytes in mice lacking the UBR2 ubiquitin ligase of the N-end rule pathway. Mol. Cell Biol. 2003, 23, 8255–8271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Ouyang, Y.; Kwon, Y.T.; An, J.Y.; Eller, D.; Tsai, S.-C.; Diaz-Perez, S.; Troke, J.J.; Teitell, M.A.; Marahrens, Y. Loss of Ubr2, an E3 ubiquitin ligase, leads to chromosome fragility and impaired homologous recombinational repair. Mutat. Res. 2006, 596, 64–75. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, F.; Cheng, Y.; An, J.Y.; Kwon, Y.T.; Eckardt, S.; Leu, N.A.; McLaughlin, K.J.; Wang, P.J. The ubiquitin ligase Ubr2, a recognition E3 component of the N-end rule pathway, stabilizes Tex19.1 during spermatogenesis. PLoS ONE 2010, 5, e14017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. An, J.Y.; Kim, E.-A.; Jiang, Y.; Zakrzewska, A.; Kim, D.E.; Lee, M.J.; Mook-Jung, I.; Zhang, Y.; Kwon, Y.T. UBR2 mediates transcriptional silencing during spermatogenesis via histone ubiquitination. Proc. Natl. Acad. Sci. USA 2010, 107, 1912–1917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. An, J.Y.; Kim, E.; Zakrzewska, A.; Yoo, Y.D.; Jang, J.M.; Han, D.H.; Lee, M.J.; Seo, J.W.; Lee, Y.J.; Kim, T.Y.; et al. UBR2 of the N-end rule pathway is required for chromosome stability via histone ubiquitylation in spermatocytes and somatic cells. PLoS ONE 2012, 7, e37414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Dickins, R.A.; Frew, I.J.; House, C.M.; O’Bryan, M.K.; Holloway, A.J.; Haviv, I.; Traficante, N.; de Kretser, D.M.; Bowtell, D.D.L. The ubiquitin ligase component Siah1a is required for completion of meiosis I in male mice. Mol. Cell Biol. 2002, 22, 2294–2303. [Google Scholar] [CrossRef] [Green Version]
  67. Frew, I.J.; Hammond, V.E.; Dickins, R.A.; Quinn, J.M.W.; Walkley, C.R.; Sims, N.A.; Schnall, R.; Della, N.G.; Holloway, A.J.; Digby, M.R.; et al. Generation and analysis of Siah2 mutant mice. Mol. Cell Biol. 2003, 23, 9150–9161. [Google Scholar] [CrossRef] [Green Version]
  68. Ward, J.O.; Reinholdt, L.G.; Motley, W.W.; Niswander, L.M.; Deacon, D.C.; Griffin, L.B.; Langlais, K.K.; Backus, V.L.; Schimenti, K.J.; O’Brien, M.J.; et al. Mutation in mouse hei10, an E3 ubiquitin ligase, disrupts meiotic crossing over. PLoS Genet. 2007, 3, e139. [Google Scholar] [CrossRef] [Green Version]
  69. Kanarek, N.; Horwitz, E.; Mayan, I.; Leshets, M.; Cojocaru, G.; Davis, M.; Tsuberi, B.Z.; Pikarsky, E.; Pagano, M.; Ben-Neriah, Y. Spermatogenesis rescue in a mouse deficient for the ubiquitin ligase SCFβ-TrCP by single substrate depletion. Genes Dev. 2010, 24, 470–477. [Google Scholar] [CrossRef] [Green Version]
  70. Guardavaccaro, D.; Kudo, Y.; Boulaire, J.; Barchi, M.; Busino, L.; Donzelli, M.; Margottin-Goguet, F.; Jackson, P.K.; Yamasaki, L.; Pagano, M. Control of meiotic and mitotic progression by the F box protein β-Trcp1 in vivo. Dev. Cell 2003, 4, 799–812. [Google Scholar] [CrossRef] [Green Version]
  71. van der Laan, R.; Uringa, E.-J.; Wassenaar, E.; Hoogerbrugge, J.W.; Sleddens, E.; Odijk, H.; Roest, H.P.; de Boer, P.; Hoeijmakers, J.H.J.; Grootegoed, J.A.; et al. Ubiquitin ligase Rad18Sc localizes to the XY body and to other chromosomal regions that are unpaired and transcriptionally silenced during male meiotic prophase. J. Cell Sci. 2004, 117, 5023–5033. [Google Scholar] [CrossRef] [Green Version]
  72. van der Laan, R.; Roest, H.P.; Hoogerbrugge, J.W.; Smit, E.M.; Slater, R.; Baarends, W.M.; Hoeijmakers, J.H.; Grootegoed, J.A. Characterization of mRAD18Sc, a mouse homolog of the yeast postreplication repair gene RAD18. Genomics 2000, 69, 86–94. [Google Scholar] [CrossRef]
  73. Gillot, I.; Matthews, C.; Puel, D.; Vidal, F.; Lopez, P. Ret finger protein: An E3 ubiquitin ligase juxtaposed to the XY body in meiosis. Int. J. Cell Biol. 2009, 2009, 524858. [Google Scholar] [CrossRef] [Green Version]
  74. Tezel, G.; Nagasaka, T.; Iwahashi, N.; Asai, N.; Iwashita, T.; Sakata, K.; Takahashi, M. Different nuclear/cytoplasmic distributions of RET finger protein in different cell types. Pathol. Int. 1999, 49, 881–886. [Google Scholar] [CrossRef]
  75. Riparbelli, M.G.; Callaini, G. The Drosophila parkin homologue is required for normal mitochondrial dynamics during spermlogenesis. Dev. Biol. 2007, 303, 108–120. [Google Scholar] [CrossRef] [Green Version]
  76. Greene, J.C.; Whitworth, A.J.; Kuo, I.; Andrews, L.A.; Feany, M.B.; Pallanck, L.J. Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc. Natl. Acad. Sci. USA 2003, 100, 4078–4083. [Google Scholar] [CrossRef] [Green Version]
  77. Smith, C.L.; DeVera, D.G.; Lamb, D.J.; Nawaz, Z.; Jiang, Y.-H.; Beaudet, A.L.; O’Malley, B.W. Genetic ablation of the steroid receptor coactivator-ubiquitin ligase, E6-AP, results in tissue-selective steroid hormone resistance and defects in reproduction. Mol. Cell Biol. 2002, 22, 525–535. [Google Scholar] [CrossRef] [Green Version]
  78. Liu, Y.-Q.; Bai, G.; Zhang, H.; Su, D.; Tao, D.-C.; Yang, Y.; Ma, Y.-X.; Zhang, S.-Z. Human RING finger protein ZNF645 is a novel testis-specific E3 ubiquitin ligase. Asian J. Androl. 2010, 12, 658–666. [Google Scholar] [CrossRef]
  79. Dwyer, J.L.; Richburg, J.H. Age-dependent alterations in spermatogenesis in itchy mice. Spermatogenesis 2012, 2, 104–116. [Google Scholar] [CrossRef] [Green Version]
  80. Lui, W.Y.; Lee, W.M. cAMP perturbs inter-Sertoli tight junction permeability barrier in vitro via its effect on proteasome-sensitive ubiquitination of occludin. J. Cell Physiol. 2005, 203, 564–572. [Google Scholar] [CrossRef] [PubMed]
  81. Liu, Z.; Oughtred, R.; Wing, S.S. Characterization of E3Histone, a novel testis ubiquitin protein ligase which ubiquitinates histones. Mol. Cell. Biol. 2005, 25, 2819–2831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Liu, Z.; Miao, D.; Xia, Q.; Hermo, L.; Wing, S.S. Regulated expression of the ubiquitin protein ligase, E3(Histone)/LASU1/Mule/ARF-BP1/HUWE1, during spermatogenesis. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2007, 236, 2889–2898. [Google Scholar] [CrossRef]
  83. Rodriguez, C.I.; Stewart, C.L. Disruption of the ubiquitin ligase HERC4 causes defects in spermatozoon maturation and impaired fertility. Dev. Biol. 2007, 312, 501–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Manku, G.; Wing, S.S.; Culty, M. Expression of the ubiquitin proteasome system in neonatal rat gonocytes and spermatogonia: Role in gonocyte differentiation. Biol. Reprod. 2012, 87, 44. [Google Scholar] [CrossRef]
  85. Oughtred, R.; Bedard, N.; Adegoke, O.A.J.; Morales, C.R.; Trasler, J.; Rajapurohitam, V.; Wing, S.S. Characterization of rat100, a 300-kilodalton ubiquitin-protein ligase induced in germ cells of the rat testis and similar to the Drosophila hyperplastic discs gene. Endocrinology 2002, 143, 3740–3747. [Google Scholar] [CrossRef] [Green Version]
  86. Mansfield, E.; Hersperger, E.; Biggs, J.; Shearn, A. Genetic and molecular analysis of hyperplastic disks, a gene whose product is required for regulation of cell-proliferation in Drosophila melanogaster imaginal disks and germ cells. Dev. Biol. 1994, 165, 507–526. [Google Scholar] [CrossRef]
  87. Pertceva, J.A.; Dorogova, N.V.; Bolobolova, E.U.; Nerusheva, O.O.; Fedorova, S.A.; Omelyanchuk, L.V. The role of Drosophila hyperplastic discs gene in spermatogenesis. Cell Biol. Int. 2010, 34, 991–996. [Google Scholar] [CrossRef]
  88. Lerer-Goldshtein, T.; Bel, S.; Shpungin, S.; Pery, E.; Motro, B.; Goldstein, R.S.; Bar-Sheshet, S.I.; Breitbart, H.; Nir, U. TMF/ARA160: A key regulator of sperm development. Dev. Biol. 2010, 348, 12–21. [Google Scholar] [CrossRef] [Green Version]
  89. Vyas, R.; Kumar, R.; Clermont, F.; Helfricht, A.; Kalev, P.; Sotiropoulou, P.; Hendriks, I.A.; Radaelli, E.; Hochepied, T.; Blanpain, C.; et al. RNF4 is required for DNA double-strand break repair in vivo. Cell Death Differ. 2013, 20, 490–502. [Google Scholar] [CrossRef] [Green Version]
  90. Rivkin, E.; Kierszenbaum, A.L.; Gil, M.; Tres, L.L. Rnf19a, a ubiquitin protein ligase, and Psmc3, a component of the 26S proteasome, tether to the acrosome membranes and the head-tail coupling apparatus during rat spermatid development. Dev. Dyn. 2009, 238, 1851–1861. [Google Scholar] [CrossRef]
  91. Nian, H.; Zhang, W.; Shi, H.X.; Zhao, Q.Z.; Xie, Q.; Liao, S.Y.; Zhang, Y.; Zhang, Z.Q.; Wang, C.; Han, C.S. Mouse RING finger protein RNF133 is a testis-specific endoplasmic reticulum-associated E3 ubiquitin ligase. Cell Res. 2008, 18, 800–802. [Google Scholar] [CrossRef]
  92. Nian, H.; Fan, C.; Liao, S.; Shi, Y.; Zhang, K.; Liu, Y.; Han, C. RNF151, a testis-specific RING finger protein, interacts with dysbindin. Arch. Biochem. Biophys. 2007, 465, 157–163. [Google Scholar] [CrossRef]
  93. Bohgaki, T.; Bohgaki, M.; Cardoso, R.; Panier, S.; Zeegers, D.; Li, L.; Stewart, G.S.; Sanchez, O.; Hande, M.P.; Durocher, D.; et al. Genomic instability, defective spermatogenesis, immunodeficiency, and cancer in a mouse model of the RIDDLE syndrome. PLoS Genet. 2011, 7, e1001381. [Google Scholar] [CrossRef]
  94. Zhao, B.; Ito, K.; Iyengar, P.V.; Hirose, S.; Nakamura, N. MARCH7 E3 ubiquitin ligase is highly expressed in developing spermatids of rats and its possible involvement in head and tail formation. Histochem. Cell Biol. 2013, 139, 447–460. [Google Scholar] [CrossRef]
  95. Iyengar, P.V.; Hirota, T.; Hirose, S.; Nakamura, N. Membrane-associated RING-CH 10 (MARCH10 protein) is a microtubule-associated E3 ubiquitin ligase of the spermatid flagella. J. Biol. Chem. 2011, 286, 39082–39090. [Google Scholar] [CrossRef] [Green Version]
  96. Morokuma, Y.; Nakamura, N.; Kato, A.; Notoya, M.; Yamamoto, Y.; Sakai, Y.; Fukuda, H.; Yamashina, S.; Hirata, Y.; Hirose, S. MARCH-XI, a novel transmembrane ubiquitin ligase implicated in ubiquitin-dependent protein sorting in developing spermatids. J. Biol. Chem. 2007, 282, 24806–24815. [Google Scholar] [CrossRef] [Green Version]
  97. Zheng, N.; Shabek, N. Ubiquitin ligases: Structure, function, and regulation. Annu. Rev. Biochem. 2017, 86, 129–157. [Google Scholar] [CrossRef]
  98. Sarikas, A.; Hartmann, T.; Pan, Z.-Q. The cullin protein family. Genome Biol. 2011, 12, 220. [Google Scholar] [CrossRef] [Green Version]
  99. Hsu, J.Y.; Reimann, J.D.R.; Sørensen, C.S.; Lukas, J.; Jackson, P.K. E2F-dependent accumulation of hEmi1 regulates S phase entry by inhibiting APC(Cdh1). Nat. Cell Biol. 2002, 4, 358–366. [Google Scholar] [CrossRef]
  100. Zhou, B.P.; Deng, J.; Xia, W.; Xu, J.; Li, Y.M.; Gunduz, M.; Hung, M.-C. Dual regulation of Snail by GSK-3β-mediated phosphorylation in control of epithelial-mesenchymal transition. Nat. Cell Biol. 2004, 6, 931–940. [Google Scholar] [CrossRef]
  101. Yu, C.; Zhang, Y.-L.; Pan, W.-W.; Li, X.-M.; Wang, Z.-W.; Ge, Z.-J.; Zhou, J.-J.; Cang, Y.; Tong, C.; Sun, Q.-Y.; et al. CRL4 complex regulates mammalian oocyte survival and reprogramming by activation of TET proteins. Science 2013, 342, 1518–1521. [Google Scholar] [CrossRef] [PubMed]
  102. Yu, C.; Ji, S.-Y.; Sha, Q.-Q.; Sun, Q.-Y.; Fan, H.-Y. CRL4-DCAF1 ubiquitin E3 ligase directs protein phosphatase 2A degradation to control oocyte meiotic maturation. Nat. Commun. 2015, 6, 8017. [Google Scholar] [CrossRef]
  103. Yu, J.; Lan, X.; Chen, X.; Yu, C.; Xu, Y.; Liu, Y.; Xu, L.; Fan, H.-Y.; Tong, C. Protein synthesis and degradation are essential to regulate germline stem cell homeostasis in Drosophila testes. Development 2016, 143, 2930–2945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Fischer, S.E.J.; Ruvkun, G. Caenorhabditis elegans ADAR editing and the ERI-6/7/MOV10 RNAi pathway silence endogenous viral elements and LTR retrotransposons. Proc. Natl. Acad. Sci. USA 2020, 117, 5987–5996. [Google Scholar] [CrossRef] [Green Version]
  105. Goodyer, W.; Kaitna, S.; Couteau, F.; Ward, J.D.; Boulton, S.J.; Zetka, M. HTP-3 links DSB formation with homolog pairing and crossing over during C. elegans meiosis. Dev. Cell 2008, 14, 263–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Peters, J.-M. The anaphase promoting complex/cyclosome: A machine designed to destroy. Nat. Rev. Mol. Cell Biol. 2006, 7, 644–656. [Google Scholar] [CrossRef]
  107. Lara-Gonzalez, P.; Moyle, M.W.; Budrewicz, J.; Mendoza-Lopez, J.; Oegema, K.; Desai, A. The G2-to-M transition is ensured by a dual mechanism that protects cyclin B from degradation by Cdc20-activated APC/C. Dev. Cell 2019, 51, 313–325.e10. [Google Scholar] [CrossRef]
  108. Cappell, S.D.; Chung, M.; Jaimovich, A.; Spencer, S.L.; Meyer, T. Irreversible APCCdh1 inactivation underlies the point of no return for cell cycle entry. Cell 2016, 166, 167–180. [Google Scholar] [CrossRef] [Green Version]
  109. Siomi, M.C.; Sato, K.; Pezic, D.; Aravin, A.A. PIWI-interacting small RNAs: The vanguard of genome defence. Nat. Rev. Mol. Cell Biol. 2011, 12, 246–258. [Google Scholar] [CrossRef]
  110. Liu, J.; Stevens, J.; Rote, C.A.; Yost, H.J.; Hu, Y.X.; Neufeld, K.L.; White, R.L.; Matsunami, N. Siah-1 mediates a novel beta-catenin degradation pathway linking p53 to the adenomatous polyposis coli protein. Mol. Cell 2001, 7, 927–936. [Google Scholar] [CrossRef]
  111. Singleton, M.R.; Uhlmann, F. Separase-securin complex: A cunning way to control chromosome segregation. Nat. Struct. Mol. Biol. 2017, 24, 337–339. [Google Scholar] [CrossRef]
  112. Qiao, H.; Prasada Rao, H.B.D.; Yang, Y.; Fong, J.H.; Cloutier, J.M.; Deacon, D.C.; Nagel, K.E.; Swartz, R.K.; Strong, E.; Holloway, J.K.; et al. Antagonistic roles of ubiquitin ligase HEI10 and SUMO ligase RNF212 regulate meiotic recombination. Nat. Genet. 2014, 46, 194–199. [Google Scholar] [CrossRef]
  113. Turner, J.M.A. Meiotic sex chromosome inactivation. Development 2007, 134, 1823–1831. [Google Scholar] [CrossRef] [Green Version]
  114. Sonnack, V.; Failing, K.; Bergmann, M.; Steger, K. Expression of hyperacetylated histone H4 during normal and impaired human spermatogenesis. Andrologia 2002, 34, 384–390. [Google Scholar] [CrossRef]
  115. Suresh, B.; Lee, J.; Hong, S.-H.; Kim, K.-S.; Ramakrishna, S. The role of deubiquitinating enzymes in spermatogenesis. Cell Mol. Life Sci. 2015, 72, 4711–4720. [Google Scholar] [CrossRef]
  116. Mevissen, T.E.T.; Komander, D. Mechanisms of deubiquitinase specificity and regulation. Annu. Rev. Biochem. 2017, 86, 159–192. [Google Scholar] [CrossRef] [Green Version]
  117. Lin, H.; Keriel, A.; Morales, C.R.; Bedard, N.; Zhao, Q.; Hingamp, P.; Lefrançois, S.; Combaret, L.; Wing, S.S. Divergent N-terminal sequences target an inducible testis deubiquitinating enzyme to distinct subcellular structures. Mol. Cell Biol. 2000, 20, 6568–6578. [Google Scholar] [CrossRef]
  118. Bedard, N.; Yang, Y.; Gregory, M.; Cyr, D.G.; Suzuki, J.; Yu, X.; Chian, R.-C.; Hermo, L.; O’Flaherty, C.; Smith, C.E.; et al. Mice lacking the USP2 deubiquitinating enzyme have severe male subfertility associated with defects in fertilization and sperm motility. Biol. Reprod. 2011, 85, 594–604. [Google Scholar] [CrossRef] [Green Version]
  119. Luo, M.; Zhou, J.; Leu, N.A.; Abreu, C.M.; Wang, J.; Anguera, M.C.; de Rooij, D.G.; Jasin, M.; Wang, P.J. Polycomb protein SCML2 associates with USP7 and counteracts Histone H2A ubiquitination in the XY chromatin during male meiosis. PLoS Genet. 2015, 11, e1004954. [Google Scholar] [CrossRef] [Green Version]
  120. Berruti, G.; Martegani, E. mUBPy and MSJ-1, a deubiquitinating enzyme and a molecular chaperone specifically expressed in testis, associate with the acrosome and centrosome in mouse germ cells. Ann. N. Y. Acad. Sci. 2002, 973, 5–7. [Google Scholar] [CrossRef]
  121. Berruti, G.; Martegani, E. The deubiquitinating enzyme mUBPy interacts with the sperm-specific molecular chaperone MSJ-1: The relation with the proteasome, acrosome, and centrosome in mouse male germ cells. Biol. Reprod. 2005, 72, 14–21. [Google Scholar] [CrossRef] [Green Version]
  122. Lee, K.H.; Song, G.J.; Kang, I.S.; Kim, S.W.; Paick, J.-S.; Chung, C.H.; Rhee, K. Ubiquitin-specific protease activity of USP9Y, a male infertility gene on the Y chromosome. Reprod. Fertil. Dev. 2003, 15, 129–133. [Google Scholar] [CrossRef] [PubMed]
  123. Sun, C.; Skaletsky, H.; Birren, B.; Devon, K.; Tang, Z.; Silber, S.; Oates, R.; Page, D.C. An azoospermic man with a de novo point mutation in the Y-chromosomal gene USP9Y. Nat. Genet. 1999, 23, 429–432. [Google Scholar] [CrossRef] [PubMed]
  124. Luddi, A.; Margollicci, M.; Gambera, L.; Serafini, F.; Cioni, M.; De Leo, V.; Balestri, P.; Piomboni, P. Spermatogenesis in a man with complete deletion of USP9Y. N. Engl. J. Med. 2009, 360, 881–885. [Google Scholar] [CrossRef] [PubMed]
  125. Krausz, C.; Degl’Innocenti, S.; Nuti, F.; Morelli, A.; Felici, F.; Sansone, M.; Varriale, G.; Forti, G. Natural transmission of USP9Y gene mutations: A new perspective on the role of AZFa genes in male fertility. Hum. Mol. Genet. 2006, 15, 2673–2681. [Google Scholar] [CrossRef] [Green Version]
  126. Kishi, K.; Uchida, A.; Takase, H.M.; Suzuki, H.; Kurohmaru, M.; Tsunekawa, N.; Kanai-Azuma, M.; Wood, S.A.; Kanai, Y. Spermatogonial deubiquitinase USP9X is essential for proper spermatogenesis in mice. Reproduction 2017, 154, 135–143. [Google Scholar] [CrossRef]
  127. Crimmins, S.; Sutovsky, M.; Chen, P.-C.; Huffman, A.; Wheeler, C.; Swing, D.A.; Roth, K.; Wilson, J.; Sutovsky, P.; Wilson, S. Transgenic rescue of ataxia mice reveals a male-specific sterility defect. Dev. Biol. 2009, 325, 33–42. [Google Scholar] [CrossRef] [Green Version]
  128. Kovács, L.; Nagy, A.; Pál, M.; Deák, P. Usp14 is required for spermatogenesis and ubiquitin stress responses in Drosophila melanogaster. J. Cell Sci. 2020, 133, jcs237511. [Google Scholar] [CrossRef]
  129. Shi, Y.-C.; Wei, L.; Cui, Y.-X.; Shang, X.-J.; Wang, H.-Y.; Xia, X.-Y.; Zhou, Y.-C.; Li, H.; Jiang, H.-T.; Zhu, W.-M.; et al. Association between ubiquitin-specific protease USP26 polymorphism and male infertility in Chinese men. Clin. Chim. Acta 2011, 412, 545–549. [Google Scholar] [CrossRef]
  130. Luddi, A.; Crifasi, L.; Quagliarello, A.; Governini, L.; De Leo, V.; Piomboni, P. Single nucleotide polymorphisms of USP26 in azoospermic men. Syst. Biol. Reprod. Med. 2016, 62, 372–378. [Google Scholar] [CrossRef] [Green Version]
  131. Zhang, J.; Qiu, S.-D.; Li, S.-B.; Zhou, D.-X.; Tian, H.; Huo, Y.-W.; Ge, L.; Zhang, Q.-Y. Novel mutations in ubiquitin-specific protease 26 gene might cause spermatogenesis impairment and male infertility. Asian J. Androl. 2007, 9, 809–814. [Google Scholar] [CrossRef] [PubMed]
  132. Stouffs, K.; Lissens, W.; Tournaye, H.; Van Steirteghem, A.; Liebaers, I. Possible role of USP26 in patients with severely impaired spermatogenesis. Eur. J. Hum. Genet. 2005, 13, 336–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Zhang, J.; Shao, X.-G.; Shi, Y.-B.; Yan, L.; Wang, L.; Tian, H.; Qiu, S.-D. Polymorphism of Usp26 correlates with idiopathic male infertility. Zhonghua Nan Ke Xue 2012, 18, 105–108. [Google Scholar]
  134. Stouffs, K.; Lissens, W.; Tournaye, H.; Van Steirteghem, A.; Liebaers, I. Alterations of the USP26 gene in Caucasian men. Int. J. Androl. 2006, 29, 614–617. [Google Scholar] [CrossRef] [PubMed]
  135. Kim, Y.-K.; Kim, Y.-S.; Yoo, K.-J.; Lee, H.-J.; Lee, D.-R.; Yeo, C.Y.; Baek, K.-H. The expression of Usp42 during embryogenesis and spermatogenesis in mouse. Gene Expr. Patterns 2007, 7, 143–148. [Google Scholar] [CrossRef]
  136. Devi, L.; Pawar, R.M.; Makala, H.; Goel, S. Conserved expression of ubiquitin carboxyl-terminal esterase L1 (UCHL1) in mammalian testes. Indian. J. Exp. Biol. 2015, 53, 305–312. [Google Scholar] [PubMed]
  137. Toliczenko-Bernatowicz, D.; Matuszczak, E.; Tylicka, M.; Szymańska, B.; Komarowska, M.; Gorodkiewicz, E.; Debek, W.; Hermanowicz, A. Overexpression of ubiquitin carboxyl-terminal hydrolase 1 (UCHL1) in boys with cryptorchidism. PLoS ONE 2018, 13, e0191806. [Google Scholar] [CrossRef]
  138. Wang, Y.-L.; Liu, W.Z.; Sun, Y.-J.; Kwon, J.; Setsuie, R.; Osaka, H.; Noda, M.; Aoki, S.; Yoshikawa, Y.; Wada, K. Overexpression of ubiquitin carboxyl-terminal hydrolase L1 arrests spermatogenesis in transgenic mice. Mol. Reprod. Dev. 2006, 73, 40–49. [Google Scholar] [CrossRef]
  139. Kwon, J.; Mochida, K.; Wang, Y.-L.; Sekiguchi, S.; Sankai, T.; Aoki, S.; Ogura, A.; Yoshikawa, Y.; Wada, K. Ubiquitin C-terminal hydrolase L-1 is essential for the early apoptotic wave of germinal cells and for sperm quality control during spermatogenesis. Biol. Reprod. 2005, 73, 29–35. [Google Scholar] [CrossRef] [Green Version]
  140. Kwon, J. The new function of two ubiquitin C-terminal hydrolase isozymes as reciprocal modulators of germ cell apoptosis. Exp. Anim. 2007, 56, 71–77. [Google Scholar] [CrossRef] [Green Version]
  141. Kwon, J.; Wang, Y.-L.; Setsuie, R.; Sekiguchi, S.; Sakurai, M.; Sato, Y.; Lee, W.-W.; Ishii, Y.; Kyuwa, S.; Noda, M.; et al. Developmental regulation of ubiquitin C-terminal hydrolase isozyme expression during spermatogenesis in mice. Biol. Reprod. 2004, 71, 515–521. [Google Scholar] [CrossRef] [PubMed]
  142. Kurihara, L.J.; Semenova, E.; Levorse, J.M.; Tilghman, S.M. Expression and functional analysis of Uch-L3 during mouse development. Mol. Cell Biol. 2000, 20, 2498–2504. [Google Scholar] [CrossRef] [Green Version]
  143. Osawa, Y.; Wang, Y.L.; Osaka, H.; Aoki, S.; Wada, K. Cloning, expression, and mapping of a mouse gene, Uchl4, highly homologous to human and mouse Uchl3. Biochem. Biophys. Res. Commun. 2001, 283, 627–633. [Google Scholar] [CrossRef]
  144. Wright, A.; Reiley, W.W.; Chang, M.; Jin, W.; Lee, A.J.; Zhang, M.; Sun, S.-C. Regulation of early wave of germ cell apoptosis and spermatogenesis by deubiquitinating enzyme CYLD. Dev. Cell. 2007, 13, 705–716. [Google Scholar] [CrossRef] [Green Version]
  145. Berruti, G.; Paiardi, C. USP8/UBPy-regulated sorting and the development of sperm acrosome: The recruitment of MET. Reproduction 2015, 149, 633–644. [Google Scholar] [CrossRef] [Green Version]
  146. Berruti, G.; Ripolone, M.; Ceriani, M. USP8, a regulator of endosomal sorting, is involved in mouse acrosome biogenesis through interaction with the spermatid ESCRT-0 complex and microtubules. Biol. Reprod. 2010, 82, 930–939. [Google Scholar] [CrossRef] [Green Version]
  147. Quenneville, N.R.; Chao, T.-Y.; McCaffery, J.M.; Conibear, E. Domains within the GARP subunit Vps54 confer separate functions in complex assembly and early endosome recognition. Mol. Biol. Cell 2006, 17, 1859–1870. [Google Scholar] [CrossRef] [Green Version]
  148. Lasko, P. The DEAD-box helicase Vasa: Evidence for a multiplicity of functions in RNA processes and developmental biology. Biochim. Biophys. Acta. 2013, 1829, 810–816. [Google Scholar] [CrossRef]
  149. Wosnitzer, M.S.; Mielnik, A.; Dabaja, A.; Robinson, B.; Schlegel, P.N.; Paduch, D.A. Ubiquitin specific protease 26 (USP26) expression analysis in human testicular and extragonadal tissues indicates diverse action of USP26 in cell differentiation and tumorigenesis. PLoS ONE 2014, 9, e98638. [Google Scholar] [CrossRef] [Green Version]
  150. Wang, P.J.; Page, D.C.; McCarrey, J.R. Differential expression of sex-linked and autosomal germ-cell-specific genes during spermatogenesis in the mouse. Hum. Mol. Genet. 2005, 14, 2911–2918. [Google Scholar] [CrossRef] [Green Version]
  151. Nishimune, Y.; Tanaka, H. Infertility caused by polymorphisms or mutations in spermatogenesis-specific genes. J. Androl. 2006, 27, 326–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Liu, Y.-L.; Zheng, J.; Mi, Y.-J.; Zhao, J.; Tian, Q.-B. The impacts of nineteen mutations on the enzymatic activity of USP26. Gene 2018, 641, 292–296. [Google Scholar] [CrossRef] [PubMed]
  153. Felipe-Medina, N.; Gómez-H, L.; Condezo, Y.B.; Sanchez-Martín, M.; Barbero, J.L.; Ramos, I.; Llano, E.; Pendás, A.M. Ubiquitin-specific protease 26 (USP26) is not essential for mouse gametogenesis and fertility. Chromosoma 2019, 128, 237–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Sakai, K.; Ito, C.; Wakabayashi, M.; Kanzaki, S.; Ito, T.; Takada, S.; Toshimori, K.; Sekita, Y.; Kimura, T. Usp26 mutation in mice leads to defective spermatogenesis depending on genetic background. Sci. Rep. 2019, 9, 13757. [Google Scholar] [CrossRef]
  155. Wang, M.; Yu, T.; Hu, L.; Cheng, Z.; Li, M. Ubiquitin carboxy-terminal hydrolaseL3 correlates with human sperm count, motility and fertilization. PLoS ONE 2016, 11, e0165198. [Google Scholar] [CrossRef]
  156. Luo, J.; Megee, S.; Dobrinski, I. Asymmetric distribution of UCH-L1 in spermatogonia is associated with maintenance and differentiation of spermatogonial stem cells. J. Cell Physiol. 2009, 220, 460–468. [Google Scholar] [CrossRef] [Green Version]
  157. Rodriguez, I.; Ody, C.; Araki, K.; Garcia, I.; Vassalli, P. An early and massive wave of germinal cell apoptosis is required for the development of functional spermatogenesis. EMBO J. 1997, 16, 2262–2270. [Google Scholar] [CrossRef] [Green Version]
  158. Hogarth, C.A.; Evanoff, R.; Mitchell, D.; Kent, T.; Small, C.; Amory, J.K.; Griswold, M.D. Turning a spermatogenic wave into a tsunami: Synchronizing murine spermatogenesis using WIN 18,446. Biol. Reprod. 2013, 88, 1–9. [Google Scholar] [CrossRef]
  159. Krausz, C.; Riera-Escamilla, A. Genetics of male infertility. Nat. Rev. Urol. 2018, 15, 369–384. [Google Scholar] [CrossRef]
  160. Toste Rêgo, A.; da Fonseca, P.C.A. Characterization of fully recombinant human 20S and 20S-PA200 proteasome complexes. Mol. Cell 2019, 76, 138–147.e135. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The processes of mammalian spermatogenesis. In males, primordial germ cells (PGCs) are specialized at early stages of fetal development and are arrested at G1 phase. After birth, PGCs differentiate to give rise to spermatogonia in male gonads, or testis. Spermatogonia undergo several rounds of mitosis to produce Asingle (As), Apaired (Apr), Aaligned (Aal), A1–4, intermediate and B types of spermatogonia, progressively. Type B spermatogonia are able to enter meiosis in response to the retinoic acid (RA) signalling. Through homologous recombination and synapsis at Prophase I, homologous chromosome segregation at Metaphase I and sister chromatids separation at Metaphase II, one spermatocyte divides to form 4 haploid round spermatids with ploidy reduction. Following the process of spermiogenesis, round spermatids mature to become spermatozoa with the capacity to fertilize MII eggs. The timelines of spermatogenesis in mouse and human are indicated (not to scale). The degradation timing of regulatory proteins involved in meiotic prophase I (Prophase I proteins) and histones as substrates of UPS, as well as the expression pattern of 20S proteasome subunits PSMA7 and PSMA8 are labeled with Gradient bar.
Figure 1. The processes of mammalian spermatogenesis. In males, primordial germ cells (PGCs) are specialized at early stages of fetal development and are arrested at G1 phase. After birth, PGCs differentiate to give rise to spermatogonia in male gonads, or testis. Spermatogonia undergo several rounds of mitosis to produce Asingle (As), Apaired (Apr), Aaligned (Aal), A1–4, intermediate and B types of spermatogonia, progressively. Type B spermatogonia are able to enter meiosis in response to the retinoic acid (RA) signalling. Through homologous recombination and synapsis at Prophase I, homologous chromosome segregation at Metaphase I and sister chromatids separation at Metaphase II, one spermatocyte divides to form 4 haploid round spermatids with ploidy reduction. Following the process of spermiogenesis, round spermatids mature to become spermatozoa with the capacity to fertilize MII eggs. The timelines of spermatogenesis in mouse and human are indicated (not to scale). The degradation timing of regulatory proteins involved in meiotic prophase I (Prophase I proteins) and histones as substrates of UPS, as well as the expression pattern of 20S proteasome subunits PSMA7 and PSMA8 are labeled with Gradient bar.
Cells 11 01058 g001
Figure 2. The schematic showing the functions of the UPS to regulate protein homeostasis during spermatogenesis. In male germ cells, the stability of proteins (or substrates here) are tightly regulated by the UPS system, including the E3 ubiquitin ligase that tags substrates for ubiquitination and the DUB that removes ubiquitin (ub) chain from the substrates to recycle both ubiquitin and substrates. Proteasomes are capable of degrading ubiquitin-linked substrates for proteolysis or degradation. So far, there are at least four types of proteasomes in testes (19S-c20S, 19S-s20S, PA200-c20S, PA200-s20S), probably expressed in different stages with distinct functions during spermatogenesis (A,B). RAD51, RPA1 and SYCP3, which are involved in synapsis and homologous recombination, are largely degraded by 19S-s20S proteasome to mediate meiotic progression; correspondingly, the unique substrates and functions for 19S-c20S proteasome remain to be explored (A). PA200 subunit is responsible for the degradation of acetylated histones to promote histone-to-protamine replacement during spermiogenesis; and it remains to be determined whether c20S or s20S is involved in this process (B).
Figure 2. The schematic showing the functions of the UPS to regulate protein homeostasis during spermatogenesis. In male germ cells, the stability of proteins (or substrates here) are tightly regulated by the UPS system, including the E3 ubiquitin ligase that tags substrates for ubiquitination and the DUB that removes ubiquitin (ub) chain from the substrates to recycle both ubiquitin and substrates. Proteasomes are capable of degrading ubiquitin-linked substrates for proteolysis or degradation. So far, there are at least four types of proteasomes in testes (19S-c20S, 19S-s20S, PA200-c20S, PA200-s20S), probably expressed in different stages with distinct functions during spermatogenesis (A,B). RAD51, RPA1 and SYCP3, which are involved in synapsis and homologous recombination, are largely degraded by 19S-s20S proteasome to mediate meiotic progression; correspondingly, the unique substrates and functions for 19S-c20S proteasome remain to be explored (A). PA200 subunit is responsible for the degradation of acetylated histones to promote histone-to-protamine replacement during spermiogenesis; and it remains to be determined whether c20S or s20S is involved in this process (B).
Cells 11 01058 g002
Table 1. Overview of E3 ubiquitin ligases associated with spermatogenesis.
Table 1. Overview of E3 ubiquitin ligases associated with spermatogenesis.
NamesExpression Pattern
(In Testis)
SubstratesMutant PhenotypesSpeciesReferences
CUL4ASg, Sc, St (low)Histone H3, H4;
MOV10
Infertility; apoptosis of Sc; decreased St and Sz; abnormal Sz with low motility and defective acrosome formationM. musculus[45,46,47]
CUL4BSg, StUnknownSz with low or no motilityM. musculus[48]
CUL3StdBruceReduced or eliminated effector caspase activation in StD. melanogaster[49]
CUL3StUnknownEarly embryonic deathM. musculus[50,51]
CUL2Male germ cells (high in Sc)HTP-3Infertility; cell cycle arrest; premature meiotic entryC. elegans[52,53]
RNF8UndefinedHistones H2A, H2AX, H2BInfertility; defective nucleosome removal in StM. musculus[54,55,56,57,58]
APC/CSc, StPIWI/MIWI Defective maturation of StM. musculus[59,60]
UBR2TestisHistones H2A, H2BInfertility; meiotic arrest; defective homologous recombinational repair, MSUC, and synapsis M. musculus[61,62,63,64,65]
SIAH1ATestisUnknownInfertility; metaphase to telophase arrest during meiosis I; bi- or multi-nucleated anaphase cellsM. musculus[66,67]
Mei4/HEI10UndefinedCCNB3 (?)Infertility; failure of crossover formation; meiotic arrest; apoptosis of ScM. musculus[68]
SCFβ-TrCPSg, ScSnail1, Emi1 (?)β-TrCP1 deficiency: accumulation of metaphase I Sc and multinucleated St;
β-TrCP1/2 double deficiency: absence of Sc, St and Sz
M. musculus[69,70]
RAD18/18scTestis (high in Sc)UnknownSubfertility; defective meiotic DSB repair and MSCIM. musculus[71,72]
RFPMale germ cellsUnknownUndefinedM. musculus[73,74]
ParkinUndefinedUnknownInfertility; defective mitochondrial organization and St individualizationD. melanogaster[75,76]
E6-APUndefinedUnknownSubfertility; reduced SzM. musculus[77]
ZNF645Sc, St, Leydig cellsUnknownUndefinedH. sapiens[78]
ITCHTestisOccludinSubfertility; age-dependent impairment; delayed St development and organizationM. musculus[79,80]
HUWE1/LASU1Sg, ScHistones (?)UndefinedR. norvegicus[81,82]
HERC4Sg, Sc, StUnknownSubfertility; reduced and abnormal SzM. musculus[83]
UBR5/HYD/EDDSc, St (low)UnknownRat: Undefined
Drosophila: Infertility; defects in St elongation
R. norvegicus
D. melanogaster
[84,85,86,87]
TMF/ARA160Sc, StUnknownInfertility; reduced motility and malformation of Sz M. musculus[88]
RNF4UndefinedMDC1, BRCA1Infertility; age dependent testicular atrophy; germ cell depletionM. musculus[89]
RNF19APachytene Sc, St, SzUnknownUndefinedR. norvegicus[90]
RNF133StUnknownUndefinedM. musculus[91]
RNF149Male germ cellsUnknownUndefinedR. norvegicus[84]
RNF151StDysbindin (?)UndefinedM. musculus[92]
RNF168UndefinedUnknownSubfertility or infertility; age-dependent impairment; testis atrophyM. musculus[93]
MARCH7Testis (high in St)UnknownUndefinedR. norvegicus[94]
MARCH10StUnknownUndefinedR. norvegicus[95]
MARCH11StUnknownUndefinedR. norvegicus[96]
Sg: spermatogonia; Sc: spermatocyte; St: Spermatids; Sz: spermatozoa. “?” indicates further studies are required to verify the substrates.
Table 2. Overview of DUBs associated with spermatogenesis.
Table 2. Overview of DUBs associated with spermatogenesis.
NamesExpression Pattern
(In Testis)
SubstratesMutant PhenotypesSpeciesReferences
USP2Ubiquitous
(high in elongating St)
UnknownSubfertility; abnormal aggregation of StM. musculus[117,118]
USP7Sc (Early-pachytene)RNF2UndefinedM. musculus[119]
USP8Testis (high in St, Sz)MSJ1 (?)UndefinedM. musculus[120,121]
USP9YUndefinedUnknownSNP associated with azoospermia/oligospermiaH. sapiens[122,123,124,125]
USP9XSg (high), Sc (low)UnknownInfertility; reduced reduced Sc; malformation of St and SzM. musculus[126]
USP14St and SzUnknownDefective St individualization; reduced and abnormal SzD. melanogaster, M. musculus[127,128]
USP26TestisUnknownPolymorphisms associated with azoospermia/oligozoospermia/asthenozoospermiaH. sapiens[129,130,131,132,133,134]
UPS42St UnknownUndefinedM. musculus[135]
UCHL1Sg, Sertoli cellsUnknownIncreased Sg; reduced motility and malformation of Sz; resistant to early wave of germ cell apoptosis M. musculus[136,137,138,139]
UCHL3Pachytene Sc, StUnknownIncreased germ cell loss after cryptorchid injuryM. musculus[140,141,142]
UCHL4SgUnknownUndefinedM. musculus[141,143]
UCHL5Sc, StUnknownUndefinedM. musculus[141]
CYLDUbiquitousRIP1Infertility; deficiency and malformation of St; attenuates the early wave of germ cell apoptosisM. musculus[144]
Sg: spermatogonia; Sc: spermatocyte; St: Spermatids; Sz: spermatozoa. “?” indicates further studies are required to verify the substrates.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiong, Y.; Yu, C.; Zhang, Q. Ubiquitin-Proteasome System–Regulated Protein Degradation in Spermatogenesis. Cells 2022, 11, 1058. https://doi.org/10.3390/cells11061058

AMA Style

Xiong Y, Yu C, Zhang Q. Ubiquitin-Proteasome System–Regulated Protein Degradation in Spermatogenesis. Cells. 2022; 11(6):1058. https://doi.org/10.3390/cells11061058

Chicago/Turabian Style

Xiong, Yi, Chao Yu, and Qianting Zhang. 2022. "Ubiquitin-Proteasome System–Regulated Protein Degradation in Spermatogenesis" Cells 11, no. 6: 1058. https://doi.org/10.3390/cells11061058

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

Article Metrics

Back to TopTop