Skip to main content

Conserved genes regulating human sex differentiation, gametogenesis and fertilization

Abstract

The study of the functional genome in mice and humans has been instrumental for describing the conserved molecular mechanisms regulating human reproductive biology, and for defining the etiologies of monogenic fertility disorders. Infertility is a reproductive disorder that includes various conditions affecting a couple’s ability to achieve a healthy pregnancy. Recent advances in next-generation sequencing and CRISPR/Cas-mediated genome editing technologies have facilitated the identification and characterization of genes and mechanisms that, if affected, lead to infertility. We report established genes that regulate conserved functions in fundamental reproductive processes (e.g., sex determination, gametogenesis, and fertilization). We only cover genes the deletion of which yields comparable fertility phenotypes in both rodents and humans. In the case of newly-discovered genes, we report the studies demonstrating shared cellular and fertility phenotypes resulting from loss-of-function mutations in both species. Finally, we introduce new model systems for the study of human reproductive biology and highlight the importance of studying human consanguineous populations to discover novel monogenic causes of infertility. The rapid and continuous screening and identification of putative genetic defects coupled with an efficient functional characterization in animal models can reveal novel mechanisms of gene function in human reproductive tissues.

Background

Successful sexual reproduction requires the recognition and fusion of sperm and oocytes, which leads to the conception of new embryos. To produce healthy and functional gametes, the embryo must develop healthy gonads, which are populated by germ cells that differentiate at the onset of puberty. Molecular genetic studies have shown that thousands of genes play a role in regulating and preserving human reproduction. When one of the reproductive processes fails, the individual becomes unable to conceive. Infertility is defined as the inability to conceive after 6–12 months of unprotected sexual intercourse and affects 16.5–17.8% of couples globally [1].

Since the inception of human in vitro fertilization (IVF) [2], intracytoplasmic sperm injection (ICSI) [3], and the achievement of the first IVF pregnancies [4], extensive research has focused on optimizing in vitro insemination and embryo culture conditions in order to improve IVF outcomes. In addition, next-generation sequencing (NGS) technologies have also been adopted to decipher the causes of fertility phenotypes for faster and more precise diagnostics and treatments. Typical monogenic fertility disorders affect sperm number, motility, or morphology [5] and may lead to primary ovarian insufficiency (POI), abnormal zygote cleavage, and embryo development arrest [6]. These phenotypes are caused by loss-of-function variants affecting genes regulating mammalian reproduction [5, 6].

We searched PubMed and the Online Mendelian Inheritance in Man (OMIM) for established and newly identified genes that regulate conserved functions in fundamental mammalian reproductive processes (e.g., sex determination, gametogenesis, fertilization, and early embryo development). Of note, we included only genes for which studies on genome-edited rodents demonstrated conserved functions and comparable fertility phenotypes upon homozygous (or compound heterozygous) loss of function mutations (i.e., mutations leading to lack of protein expression) in humans [7,8,9]. For the established genes (genes with a well-documented conserved role in mouse and human), here we briefly describe their conserved role in mouse and human reproductive biology and reference only studies reporting the human fertility phenotypes. For the most recently identified genes, we describe shared cellular and fertility phenotypes resulting from loss-of-function mutations in both species.

We do not cover associative studies on genes for which no animal models have been generated. We do not include genes for which a null-mutant animal model has been generated, but no clear loss of function mutations has been found in humans (e.g., a homozygous putative deleterious missense variant without functional validation, or a heterozygous frameshift variant would not be considered as clear loss of function mutations). In addition, we do not include fertility phenotypes due to chromosomal structural aberrations, abnormal sex chromosome numbers (e.g., Y chromosome microdeletions, Klinefelter, XXY- XXXXY males), or genetic systemic disorders associated with infertility such as the Kartagener’s, fragile X, Noonan syndromes, myotonic dystrophy, sickle cell anaemia, and β-thalassemia.

Genetic control of sex determination

Human sex determination occurs early in embryogenesis, and the embryo develops bipotential gonadal primordia, which through genetic regulation, can differentiate as either testes or ovaries [10]. Six weeks post-conception, the sex-determining region Y (SRY) gene activates Sry-related HMG box gene-9 (SOX9), which induces the expression of the Anti-Müllerian duct hormone (AMH), thus actively controlling testis development, Sertoli cell differentiation, and the general maleness of XY individuals [11]. Secreted by the Sertoli cells, AMH induces the degeneration of the Müllerian duct [12]. Sox9 also activates fibroblast growth factor-9 (FGF9), which represses the WNT4 expression and the ovarian development [10]. In addition, SRY works in concert with the steroidogenic factor-1 (encoded by the Nuclear Receptor NR5A1) to maintain Sox9 expression [13].

Conversely, the Nuclear Receptor Subfamily 0 Group B Member-1 (NR0B1 or DAX1), antagonizes the function of SRY [14] while downregulating NR5A1 expression [15] (Fig. 1). DNA variants affecting SRY and SOX9 [16] leads to sex reversal in humans, and, on the other end, XX individuals carrying extra copies of either SRY or SOX9 develop as males [16]. Similarly, mutations in human NR5A1 frequently lead to 46 XY disorders of sex development [17]. Consistent with its function, duplication of the X region containing NR0B1 is associated with male-to-female sex reversal in XY individuals, and loss-of-function mutations in NR0B1 are responsible for X-linked adrenal hypoplasia congenita, a disorder characterized by hypogonadotropic hypogonadism (Fig. 1) [18]. When gene expression favors the pro-ovarian Rspo1/Wnt4–β-catenin signaling pathway over Fgf9, a different set of genes takes over to regulate female sex determination (Fig. 1).

Fig. 1
figure 1

Genetics of human sex determination. Genes that mediate the differentiation of bipotent gonad or reproductive ducts and that are associated with comparable sex developmental disorders in mice and humans (OMIM gene ID).

The Wingless-Type MMTV Integration Site Family, Member-4 (WNT4), is expressed in the genital ridge while still in its bipotential stage [11] and becomes undetectable in XY gonads. In contrast, it is maintained in XX as their gonads differentiate into ovaries. WNT4 expression is regulated by RSpondin-1 (RSPO1), a secreted activator protein that upregulates the canonical WNT/β-catenin signaling pathway to promote ovarian development while antagonizing testis formation. In humans, mutations in WNT4 lead to 46 XX virilization, primary amenorrhea, uterine hypoplasia, and follicle depletion [19]. Similarly, XX individuals lacking RSPO1 show female-to-male sex reversal, and XY individuals with a duplicated Chromosome 1 region encompassing RSPO1 and WNT4 present male-to-female sex reversal [20]. These mechanisms are conserved across evolution, and gene deletion in transgenic mice leads to comparable sex reversal phenotypes and infertility. Of note, recent studies have shown that the elimination of the Wolffian ducts is an active process regulated by the Nuclear Receptor Subfamily-2-Group-F-Member-2 (NR2F2). This ligand-inducible transcription factor suppresses the mesenchyme-epithelium crosstalk, which is necessary to conserve the Wolffian ducts [21]. Gene deletion in mice leads to intersex individuals presenting female and male reproductive tracts [21]. In humans, loss-of-function mutations affecting NR2F2 lead to testis [22] or ovotestis development [23] in XX individuals (Fig. 1).

Sex reversal can also manifest upon hormonal imbalance during embryo development. Mutations in AMH or its receptor AMHR2 induce Persistent Müllerian Duct Syndrome and internal hermaphroditism [24,25,26,27,28]. In addition, deleterious variants in genes regulating hormonal biosynthesis are typically associated with sex-determination phenotypes. For example, deficiency in the enzymes regulating cortisol biosynthesis, such as Cytochrome P450 Family-11-Subfamily-B-Member-1 (CYP11B1) lead to congenital adrenal hyperplasia [29, 30]. Also, the Hydroxysteroid 17-beta Dehydrogenase-3 (HSD17B3) code for enzymes responsible for the biogenesis of testosterone and dihydrotestosterone. Lack of HSD17B3 leads to pseudohermaphroditism in 46 XY individuals, impaired maturation of Leydig cells, and under-masculinization in men and mice [31]. Similar fertility phenotypes are observed when Amh, Ahmr2, Cyp11b1, are deleted in transgenic mice.

While bipotent gonads commit to becoming ovaries or testes, gonadal somatic cells support the development of sex-specific germ cell lineages, precursors of eggs and sperm.

Genes regulating germ cell physiology and meiosis in men

Over the past decade, advanced next-generation sequencing has unveiled the causes of numerous cases of male-factor infertility. Simultaneously, the generation of genome-edited mouse lines has shed light on the conserved functions of key mammalian genes, preserving male fertility in both mice and humans. Normal gametogenesis in mammalian males originates during early embryonic development from isomorphic primordial germ cells (PGCs) [32]. At this stage, the FA Complementation Group M (FANCM) gene is necessary to preserve genomic stability by regulating mammalian DNA replication and repair [33]. Deleterious variants in human FANCM lead to oligoasthenozoospermia [34] or Sertoli Cell-Only Syndrome, where only Sertoli cells outline the seminiferous tubules, with no sperm detectable [35], and Fancm-null male mice show reduced proliferation and loss of PGCs [36].

Male PGCs proliferate and migrate into the developing testis, where they will differentiate postnatally, into spermatogonia stem cells (SSCs) [37]. SSCs are unipotent cells that complement self-renewing with differentiating divisions to preserve the stem cell pool while maintaining adequate sperm production throughout the male reproductive lifespan [38]. Factors maintaining a balance between proliferating and differentiating SSCs help prevent the premature depletion of the SSC pool, while regulating controlled differentiation [38]. Human Nanos C2hc-Type Zinc Finger-2 (NANOS2) is necessary to prevent XY germ cells from prematurely entering meiosis, and one homozygous deleterious mutation has been found to segregate with Sertoli cell-only syndrome in humans [8]. Nanos2 role is conserved in mammals: Nanos2 deletion by Cas9 genome-editing in mice, pigs, goats, and cattle leads to germline-depleted testes and male infertility [39]. While Nanos2 prevents differentiation, the TATA-box Binding Protein Associated Factor-4b (TAF4B) controls the expression of genes promoting differentiation and self-renewal of SSCs, and TAF4B deletion results in nonobstructive azoospermia (NOA) or oligozoospermia in mice and men (Fig. 2) [40].

Spermatogonia undergo premeiotic DNA replication, differentiate into primary spermatocytes, and eventually enter prophase I of meiosis. The alignment and synapsis of the homologous chromosomes and genetic recombination occur during prophase I of meiosis. Meanwhile, the duplication of centrioles, the formation of DNA double-strand breaks (DSBs)(leptotene stage), the assembly and maintenance of the synaptonemal complex (zygotene stage), and the formation of crossing overs (pachytene stage) ensure the formation of genetically intact and fertile sperm (Fig. 2). The Polo-Like Kinase-4 (PLK4) controls centriole duplication, which is necessary for primary spermatocytes to undergo meiosis. One patient, carrier of a heterozygous deletion in the Ser/Thr kinase domain of PLK4, presented with infertility due to Sertoli cell-only syndrome, similar to mice heterozygous for a Plk4 -null mutation [41].

The synaptonemal complex ensures the association between homologous chromosomes, while the cohesin complex mediates sister chromatid cohesion, and the telomeres adhere to and move on the nuclear envelope to regulate chromosome mobility and homologous pairing. Meiotic double-stranded break formation protein-1 (MEI1), Meiosis Specific with Ob-Fold (MEIOB), Testis Expressed-11 (TEX11), 14 (TEX14) and 15 (TEX15), help to induce the formation of DSBs, while contributing to the assembly of the synaptonemal complex and generation of crossovers between homologous chromosomes. Spermatocyte arrest is observed upon gene deletion of human MEI1 [42] or TEX11 [43,44,45], and in men carrying homozygous loss-of-function mutations in MEIOB [46,47,48,49]. In addition, depletion of TEX14 leads to Sertoli cell-only syndrome [8, 50], whereas deleterious variants in TEX15 lead to NOA and crypto/oligozoospermia [45]. In addition, Minichromosome Maintenance Domain-Containing Protein-2 (MCMDC2) and Ring-Finger Protein-212 (RNF212) are necessary for the formation and maintenance of the synaptonemal complex and resolution of DSBs [8, 51, 52]. Moreover, three factors assemble in a complex that promotes telomere adhesion to the nuclear envelope, namely Telomere Repeat-Binding Bouquet Formation Protein-1 and − 2 (TERB1, TERB2), Membrane-Anchored Junction Protein (MAJIN) [52], and SAD1-and-UNC84-Domain-Containing-1 (SUN1) [53]. The TERB1-TERB2-MAJIN complex is necessary for mouse and human meiosis [54], and its disruption leads to NOA in men [52].

Screening more NOA patient genomes has confirmed the role of several other factors (that were already established as necessary for male meiosis in mice) in regulating human meiosis. Genes coding for meiosis-specific recombinases (DNA Meiotic Recombinase-1, DMC1) [55], proteins repairing DNA inter-strand crosslink and DSBs (FANCA) [56], transcription or post-transcriptional regulation such as Tudor-Domain Containing Protein-7 (TDRD7) [57] and Zinc Finger Mynd-Containing Protein-15 (ZMYND15) [40], protein kinases such as Serine Protease Inhibitor-Kazal-Type-2 (SPINK2) [58], or genes controlling meiotic progression (Meiosis-1-Associated-Protein, M1AP) [59] are necessary regulators of male meiosis in men and mice (Fig. 2).

Moreover, during male meiosis, the P-element–induced wimpy testis (PIWI)-interacting RNAs (piRNAs) (which are expressed in the pachytene spermatocytes) interact with PIWI proteins [60,61,62] (PIWIL1, or MIWI, PIWIL2, or MILI, and PIWIL4, or MIWI2) to preserve the germline genome against the mobilization of transposable elements [63]. The Hen Methyltransferase-1 (HENMT1) controls the 2’ O-methylation of piRNAs. HENMT1 deficiency results in piRNA instability and NOA [64]. In addition, adult meiotic and haploid germ cells undergo TE de-repression, resulting in the premature expression of haploid transcripts, increased DNA damage, and spermiogenesis arrest [65]. The Poly(A)-Specific RNAse-Like Domain Containing-1 (PNLDC1) regulates the processing of piRNAs by trimming the 3′ ends [66, 67]. Pnldc1 deletion in men and mice affected the expression of piRNA-processing proteins (e.g., PIWIL1, PIWIL4) and pachytene piRNAs in the testes, leading to NOA [66,67,68]. TDRD9 silences Line-1 (L1) retrotransposons in the male germ line [69], and loss of TDRD9 leads to cryptozoospermia or azoospermia [69].

Fig. 2
figure 2

Genes regulating meiosis and spermiogenesis during spermatogenesis. Top, genes regulating different phases of Prophase I at meiosis or spermiogenesis stages (OMIM gene ID). Deletions of any of these genes present a comparable fertility phenotype in mice and humans (reported in OMIM or by case reports). Bottom, cellular differentiation stages of spermatogenesis

Genetic control of female gametogenesis

DNA replication and recombination of oocyte meiosis occur in the fetal ovary, and the maturing oocytes arrest at the dictyate (diplotene) stage. At this stage, homologous chromosomes are held together in a bivalent configuration through crossover recombination between homologous chromosomes and cohesion between sister chromatids. In humans, such a configuration is maintained for decades until ovulation. At the resumption of oocyte meiosis, the completion of meiosis I coincides with the spindle formation, and the segregation of homologous chromosomes depends on the correct assembly of the spindle [70]. Tubulin-Beta-8 (TUBB8) is a necessary component of the mouse and human oocyte spindles. TUBB8 DNA variants acting as dominant-negative led to infertility due to defective oocyte maturation, and an abnormal or completely-absent spindle [70, 71].

Spindle pole instability is another major cause of human fertility disorders, as it may lead to aneuploidy. Humans, bovine and porcine oocytes are depleted of acentriolar microtubule-organizing centers (aMTOCs), and the Nuclear Mitotic Apparatus Protein (NUMA)-mediated clustering of microtubule minus ends focuses the spindle poles [72]. In mice, the aMTOC-free oocytes present stable spindles owing to a spindle-stabilizing protein, the KINESIN Superfamily Protein-C1 (KIFC1), constitutively absent in human oocytes [72]. Kifc1 deletion in mice leads to spindle instability, while exogenous KIFC1 injected in human oocytes rescues spindle instability [72].

At meiosis I, homologous recombination mediates accurate segregation of homologous chromosomes. DMC1 regulates meiotic recombination by promoting homologous chromosome pairing and DNA strand transfer from nicked dsDNA to homologous ssDNA. Dmc1-null mice show aberrant oogenesis during fetal development and adult ovaries devoid of germ cells [55]. Similarly, in humans, homozygous deleterious mutation in DMC1 presented with primary infertility due to POI [55]. Psmc3-Interacting Protein (PSMC3IP) is another factor that facilitates meiotic recombination to promote DNA strand exchange at meiosis. Homozygous deleterious mutations in PSMC3IP have been shown to segregated with POI [73]. Similarly to humans, Psmc3ip null female mice present ovaries deprived of follicles and are infertile [73].

DNA repair by homologous recombination is a necessary step for the development of fertile oocytes. The repair of DNA double-strand breaks and stalled DNA replication forks during meiosis is mediated by two components of the mini-chromosome maintenance protein group, the Minichromosome Maintenance 8 Homologous Recombination Repair and Minichromosome Maintenance 9 Homologous Recombination Repair Factors (MCM8, MCM9). Several studies reported pathogenic variants in either MCM8 [74,75,76,77] or MCM9 [78, 79] segregating with POI. Comparably to humans, the deletion of mouse Mcm8 or Mcm9 leads to infertility due to ovaries lacking mature follicles [80]. During homologous recombination, the REC114 Meiotic Recombination Protein (REC114) mediates the formation of DNA DSBs in unsynapsed regions, a necessary step for the completion of synapsis. Lack of Rec114 leads to NOA and POI in mice, due to defective DSB formation and aberrant homologous synapsis [81]; in women, REC114 gene deletion leads to infertility due to supernumerary pronuclei formation at fertilization and early embryonic arrest [82].

Genes that cause defective gametogenesis in men and women

Several genes regulate shared molecular pathways during male and female gametogenesis. These pathways include the pairing and recombination of homologous chromosomes, DNA repair, formation of crossovers, synaptonemal or cohesin complexes, or regulate gene expression during gametogenesis, independently from meiosis. Loss-of-function mutations affecting these genes may often result in NOA in men and POI in women [55, 73,74,75,76,77,78,79,80, 83,84,85]. Examples include MutS Homolog 4 and 5 (MSH4, and MSH5), X-Ray Repair Cross-Complementing-2 (XRCC2), Kash Domain-Containing Protein 5 (KASH5), DNA-Binding Protein-Synaptonemal Complex Protein-3 (SYCP3), Synaptonemal Complex Central Element Protein-1 (SYCE1), Thyroid Hormone Receptor Interactor-13 (TRIP13), Stromal Antigen 3 (STAG3), Spermatogenesis And Oogenesis Specific Basic Helix-Loop-Helix 1 and 2 (SOHLH1 and SOHLH2).

MSH4 mediates recombination and segregation of homologous chromosomes at meiosis in testes and ovaries, and gene deletion leads to POI [86] and NOA [87]. MSH5 regulates DNA mismatch repair and meiotic recombination, and deleterious variants also result in POI or NOA [88]. XRCC2 controls the homologous recombination DNA repair pathway during chromosomal fragmentation, translocations or deletions. While constitutional lack of Xrcc2 leads to almost complete fetal or perinatal lethality, mice carrying a deleterious variant in Xxrc2 present male infertility due to NOA and female infertility or severe female subfertility due to atrophic ovaries deprived of follicles [89]. Similarly, a deleterious variant in XRCC2 has been shown to cause POI and NOA in humans [90]. During recombination, Helicase For Meiosis 1 (HFM1) regulates formation of crossover and complete synapsis of homologous chromosomes and its deletion leads to POI and NOA in mice [91] and POI in humans [92]. KASH5 regulates pairing of homologous chromosomes [93], and gene deletion leads to POI and NOA [8, 94].

SYCP3 also regulates homologous chromosome pairing and meiotic recombination. SYCP3 DNA variants are associated with recurrent pregnancy loss, and Sycp3 deletion in mice leads to oocyte aneuploidy [95]. Synaptonemal Complex Central Element Protein 1 (SYCE1) also connects homologous chromosomes during meiotic prophase I, and it is necessary for crossover formation. SYCE1 interfaces with Chromosome-14-Orf-39 (C14ORF39), a meiotic protein expressed in the central element of the synaptonemal complex. Deletion of SYCE1 in two infertile sisters from a consanguineous family [96] or deleterious mutations in C14orf39 result in POI or NOA [51, 85, 96]. TRIP13 is a negative regulator of crucial elements of the synaptonemal complex, namely the HORMA proteins, HORMAD1 and HORMAD2 [97]. Deleterious missense mutations in TRIP13 result in lower TRIP13 protein expression and to POI due to an aberrant intracellular accumulation of HORMAD2 mRNA and protein, which had a dominant effect, leading to oocyte meiotic arrest [98]; similarly, deletion of Trip13 in mice leads to POI or NOA [99].

Cohesin is a chromosome-associated multi-subunit protein complex that preserves cohesion between replicated sister chromatids, and it is necessary for chromosome segregation and DNA repair. STAG3 is a subunit of the cohesin complex. Homozygous STAG3 missense pathogenic variants associated with POI and NOA and Stag3-deficient mice show comparable phenotypes [83]. During gonad development, oocyte and spermatogonia differentiation are regulated by two transcription factors, SOHLH1 and 2. In males, Sohlh1 and 2 suppress genes that control SSCs maintenance and promote the expression of genes inducing spermatogonial differentiation. In females, they control oogenesis and folliculogenesis in the embryonic gonad. Both Sohlh1-null and Sohlh2-null mutant female mice are infertile due to severe lack of follicles [100], and Sohlh1-null and Sohlh2-null mutant males present spermatogonia that precociously enter meiosis [101]. In humans, lack of SOHLH1 leads to ovarian dysgenesis [102].

Spermiogenesis regulates the formation of fully differentiated sperm

Haploid spermatocytes undergo a series of key morphological changes, including acrosome biogenesis, DNA repackaging, head reshaping, and flagellum formation. These changes are orchestrated by several genes, whose mutations lead to fertility conditions related to sperm morphology and motility (e.g., globozoospermia, asthenozoospermia, and asthenoteratozoospermia) [103].

During acrosome biogenesis, small Golgi-derived proacrosomic vesicles amass and merge into a single spherical acrosomic vesicle that connects to the nucleus [103]. Deleterious mutations in genes regulating acrosome biogenesis, such as Spermatogenesis Associated-16 (SPATA16), Dpy-19 Like-2 (DPY19L2), Zona Pellucida Binding Protein-1 (ZPBP1) and the Coiled-Coil Domain-Containing-62 (CCDC62), typically lead to globozoospermia, defined by sperm with a round-shaped head, and an atrophied, misplaced, or virtually absent acrosome, as shown by studies in consanguineous and non-consanguineous human populations [104,105,106]. Meanwhile, the sperm nucleus is efficiently compacted to facilitate the delivery of the paternal genome to the egg. Nucleoporin − 210-Like (NUP210L) controls nuclear trafficking at spermiogenesis, and loss-of-function variants in human NUP210L in one infertile consanguineous man resulted in low sperm count, poor motility, and large-headed sperm presenting uncondensed nuclear DNA [107].

The structural reshaping of the spermatocyte head is regulated by a transient microtubular structure defined as the manchette, which mediates the condensation and elongation of the sperm head and the development of the flagellum [108]. During the manchette assembly, the IQ Motif-Containing-N (IQCN) regulates microtubule nucleation through calmodulin and calmodulin-related binding proteins [109]. The Hook Microtubule-Tethering Protein-1 (HOOK1) mediates the formation of the manchette and the intracellular transport of proteins [110]. Sperm Associated Antigen-17 (SPAG17) is a component of the sperm manchette and axoneme [111]. The ATP/GTP Binding Protein 1 (AGTPBP1) regulates the polyglutamation of tubulin during spermiogenesis. AGTPBP1 is expressed in the mouse and human manchette, and its absence leads to teratozoospermia and infertility in mice and humans [112]. The Cilia And Flagella Associated Protein-52 (CFAP52) codes for an inner microtubule protein necessary for the ciliary or flagellar beating. CFAP52 works with CFAP45 and axonemal dynein subunit DNAH11 and localizes to the spermatid manchette and the sperm tail. In addition, CFAP69 regulates head and flagellum development [113], whereas the Centrosomal Protein 70 (CEP70) mediates flagellar formation and acrosome biogenesis [114, 115]. Deleterious variants affecting the manchette development lead to aberrant acrosomal morphology [109], decaudated heads or headless tails [110], severely reduced sperm motility [116], or asthenozoospermia [117].

The flagellum is another essential structural component of mature sperm as it confers progressive and hyperactive motility, both necessary for fertilization. Numerous factors define the flagellum formation and elongation. The development and function of each of these factors are regulated by individual genes that, if deleted, lead to asthenozoospermia and multiple morphological abnormalities of the flagella (MMAF; defined by short, coiled, irregular, or absent sperm tails). The centrosomes are organelles that play a dual role, before and after fertilization. Before fertilization, centrosomes link the head and tail and regulate sperm flagellar beating; after fertilization, centrosomes mediate the formation of the zygote cytoskeleton [118]. The Centrosomal Protein-128 and 135 (CEP128; CEP135) mediate centriole biogenesis [119, 120]. The Intraflagellar transport (IFT) complex controls protein transport along the developing flagellum, and the Tetratricopeptide Repeat Domain-21a (TTC21A) [121] and − 29 (TTC29) are key regulators of the IFTs [122, 123]. The fibrous sheath (FS) provides the sperm with proper structure, flexibility, and regulation of motility through the activity of A-Kinase Anchoring Protein-3 and − 4 (AKAP3, AKAP4), and the Fibrous Sheath Interacting Protein-2 (FSIP2) [124, 125]. The axoneme of the flagellum is defined by a “9 + 2” structure consisting of a central pair of two singlet microtubules surrounded by nine doublet microtubules. It confers motility to the sperm through the Inner and Outer Dynein Arms (IDAs, ODAs) motor activity [126]. Sperm Flagellar Protein-2 (SPEF2) is necessary to develop the axoneme [126, 127].

In addition, several axonemal dynein proteins, including DNAH1, DNAH2, DNAH6, DNAH10, DNAH17, and DNALI1 [128,129,130,131,132,133,134,135,136,137,138], Cilia And Flagella Associated Proteins (CFAP43, CFAP44, CFAP47, CFAP54, CFAP57, CFAP65, CFAP70) are main constituents of the IDAs and ODAs and loss of function variants lead to defective spermiogenesis and male infertility [139,140,141,142,143,144,145]. Finally, the radial spoke is a multiprotein complex (CFAP61, CFAP91, CFAP206, CFAP251) [146,147,148,149] serving as a mechanochemical transducer between the central and peripheral pair microtubule doublets while controlling flagellar beating. Also, during flagellar formation, the ubiquitin-proteasome pathway eliminates abnormal proteins, organelles, and sperm cells. Glutamine Rich-2 (QRICH2) stabilizes the expression of proteins necessary for flagellar development by suppressing the ubiquitination-dependent degradation of these proteins [150, 151]. Indeed, the exact splicing of pre-mRNAs from genes regulating spermatid head structuring and acrosome and tail biogenesis is imperative for proper spermiogenesis. The RNA-Binding Motif Protein-5 (RBM5) is a key component of the spliceosome A complex, and deleterious mutations lead to defective spermatid differentiation and NOA [152].

Spermiogenesis finally results in the development of mature spermatozoa that are released from the seminiferous tubules into the epididymis to undergo post-testicular maturation before becoming competent for natural fertilization. Different molecular changes in the growing ovarian follicle and developing oocyte are necessary to develop a fertile egg.

Follicular maturation and formation of the zona pellucida

The control of ovarian function, somatic granulosa cells’ fate, and oocyte’s developmental competence are under the concerted action of the Growth/Differentiation Factor-9 (GDF9) and the Bone Morphogenetic Protein-15 (BMP15). GDF9 and BMP15 belong to the TGF-beta of ligands that activate SMAD family transcription factors. Homozygous frameshift deletion in GDF9 leads to POI and infertility in women and mice [153]. BMP15 can heterodimerize with GDF9 to promote oocyte maturation and follicular development by regulating gene expression in granulosa cells.

Genome integrity and appropriate control of mRNA metabolism are also necessary during oocyte maturation. The transcription factor p63 (TP63) maintains the female germline genome intact during oogenesis. Heterozygous TP63 gene deletion leads to the formation of an aberrantly activated mutant TP63 tetramer, which acts in a dominant negative fashion by increasing the expression of apoptosis-inducing factors, leading to cell apoptosis in the ovary and POI in mice and humans [154]. Pat1-Homolog-2 (PATL2) regulates transcription and translation during oogenesis, and loss of PATL2 leads to infertility in women due to oocyte maturation arrest [155]. Deletion of Patl2 in mice affects the expression of key transcripts during oocyte maturation, leading to a decreased number of ovulated MII oocytes and defective early embryo development [156].

Fig. 3
figure 3

Genetics of oogenesis and follicular development. Top, genes regulating different stages of meiosis I and II, or phases at meiosis I (gene OMIM ID). Deletions of these genes present a comparable fertility phenotype in mice and humans. Bottom, stages of folliculogenesis at pre- or post-birth and genes regulating oocyte maturation (OMIM gene ID). Deletions of these genes present a comparable female (or female and male, for genes regulating meiosis) fertility phenotype in mice and humans (reported in OMIM or by case reports)

In addition, aberrant ovarian development leads to ovarian dysgenesis likely resulting in atrophic ovaries and absence of folliculogenesis. Indeed, lack of Follicle Stimulating Hormone Receptor (FSHR) leads to atrophic ovaries, loss of folliculogenesis and defective ovulation in mice [157] and ovarian dysgenesis in humans [158].

Human oocytes are surrounded by the extracellular zona pellucida (ZP), composed of 4 glycoproteins, ZP1-4 [7]. Mutations of the ZP genes affect the zona structure; homozygous frameshift or compound heterozygous variants affecting ZP1 result in mutant ZP1 that sequesters ZP3 during zona biogenesis [159, 160] or prevent the establishment of filament crosslinking in the matrix, which typically preserves the structural stability of the zona, leading to zona absence and infertility [161,162,163]. In addition, women or female mice with heterozygous nonsense mutations in ZP2 and frameshift mutations in ZP3 had no zona formation and primary infertility [164]. Moreover, deleterious missense mutations in ZP3 prevent proper interaction with ZP2, leading to empty follicle syndrome due to the absence of zona formation [165]. A structurally intact zona surrounding a genetically-intact MII oocyte is the prerequisite for successful fertilization (Fig. 3).

Fertilization and early embryo development

Following asymmetric cytokinesis, oocytes enter metaphase II (MII) and complete meiosis II only after fusion with the fertilizing sperm. Inhibition of Cyclin-Dependent Kinase-1 (CDK1) induces meiosis II completion after gamete fusion. The inhibition of CDK1 is regulated by the WEE2-Oocyte Meiosis-Inhibiting Kinase (WEE2), which is a regulator of cell cycle progression (Fig. 3). Women with deleterious variants in WEE2 show oocyte maturation defects [166] or recurrent fertilization failure (due to aberrant oocyte maturation) [167]. For successful fertilization, sperm must swim toward the unfertilized egg, undergo acrosome exocytosis, and bind and cross the extracellular zona. After gamete adhesion and fusion, egg activation allows the resumption of meiosis and the beginning of preimplantation embryo development [7] (Fig. 4). The Solute Carrier Family 9 Member-C1 (SLC9C1) is a sodium/proton exchanger that controls sperm motility through soluble adenylyl cyclase in men and mice [168]. In addition, the Potassium Channel, Subfamily U, Member-1 (KCNU1), mediates sperm membrane hyperpolarization and acrosome exocytosis, and men from consanguineous families and mice lacking KCNU1 are infertile due to impaired acrosome exocytosis and zona penetration [169].

Also, Acrosin, a trypsin-like serine protease in the sperm acrosome, mediates sperm passage through the zona, and gene deletion in consanguineous men and hamsters leads to infertility [170, 171]. Before crossing the zona, sperm bind to ZP2 [172]. Two homozygous ZP2 variants in consanguineous women led to zonae that could not support sperm binding, and women were infertile [173]. To cross the zona, sperm must also acquire hyperactive motility, a vigorous non-linear swimming pattern, which, in mice, is mediated by CatSper, a sperm-flagellar specific and Ca2+-selective channel [174]. CatSper is composed of nine known different proteins coded by individual genes. Loss of CATSPER1 or CATSPER2 leads to infertility in men due to reduced sperm count [175] or asthenoteratozoospermia [176].

After crossing the zona, acrosome-reacted sperm adhere to the oolemma through the direct interaction between the sperm membrane ligand IZUMO1 and its oocyte receptor, JUNO [177]. After adhesion, gametes fuse, and the sperm protein Phospholipase C-Zeta-1 (PLCζ1) mediates egg activation through Ca2+ signaling. Consanguineous men lacking PLCζ1 present normal sperm parameters, yet they cannot fertilize eggs due to a defective oocyte activation [178, 179]. In addition, deleterious mutations in PLCζ1 induce a mislocalized and decreased expression of PLCζ1 in the sperm head, which results in infertility [180]. Also, men with homozygous or compound heterozygous pathogenic variants in Actin-Like7a (ACTL7A) and 9 (ACTL9) present sperm lacking PLCζ expression in the head, which leads to failed oocyte activation [181, 182]. Contrarily to humans, transgenic male mice lacking PLCζ1 are subfertile, as their sperm often (though not always) fail to activate the MII oocytes [183].

Fig. 4
figure 4

Genetics of fertilization and preimplantation development. Genes coding for proteins mediating sperm physiology, gamete interaction, and early embryo development (gene OMIM ID). Deletions of these genes present comparable or dissimilar fertility phenotypes in mice vs. humans

After fertilization, preimplantation embryo development is regulated by preserving a precise balance between cell pluripotency and cell differentiation, which is necessary for successful implantation (Fig. 4). In the blastocyst, the Caudal Type Homeobox-2 (CDX2)-expressing trophectoderm and the pluripotent inner cell mass (ICM) define two distinct lineage specifications. In humans, Pou Domain-Class-5-Transcription Factor-1 (POU5F1) transcripts are detected at the four-to-eight-cell stage, yet the OCT4 protein is detectable at the eight-cell stage [184]. Genome editing studies have reported that deletion of POU5F1 by CRISPR/Cas9 in human zygotes resulted in developmental defects before blastocyst formation [185]. Loss of POU5F1 in human embryos leads to failure to complete blastocyst formation and embryonic lethality [185].

Conclusions

Almost half a century after the first baby was conceived through IVF [2], our understanding of the genetics and molecular biology underlying fertility disorders has remarkably increased. Genetic reports on infertile individuals from populations with high consanguinity rates have successfully identified more infertility-causing mutations. Besides, CRISPR/Cas-genome editing tools have facilitated the functional characterization of genes regulating mammalian reproduction. Thus, the increased available numbers of known genes and variants causing fertility phenotypes allow fertility specialists to treat patients with personalized therapies based on their genetics. However, despite the increased understanding of the functions of individual reproductive proteins, a few key questions still need to be answered.

Can deleterious heterozygous variants in multiple genes affect the fertility of one individual? Recent studies in mice show how even deleterious heterozygous variants affecting distinct but functionally related genes may lead to reproductive phenotypes such as male infertility due to MMAF [186]. These discoveries raise the hypothesis that some idiopathic male infertility cases could be explained by heterozygous deleterious variants affecting multiple loci within the same intracellular pathways.

Is the mouse the best organism to model human reproductive disorders? Combining genome sequencing of infertile consanguineous individuals with the deletion of candidate genes in transgenic mice can be an effective strategy for studying the genetics of infertility. However, deleting conserved genes in different mammalian species may lead to differences in the severity of fertility phenotypes or the sex affected. For example, the genetic ablation of Acrosin results in no fertility phenotype in mice, subfertility in rats, or complete infertility in hamsters and humans [170, 171]. Moreover, because Piwi genes regulate spermatogenesis in mice, their role in preserving women’s fertility has been neglected. However, in hamsters, Piwi genes control oogenesis and early embryo development, and deleting genes regulating the Piwi-interacting RNAs pathway in hamsters leads to male and female infertility [187, 188]. Interestingly, the expression pattern of Piwi genes in humans is highly similar to hamsters. Therefore, the hamster represents another promising model for studying human infertility.

Does inbreeding increase the incidence of monogenic forms of infertility? From a public health perspective, it is still unclear whether consanguineous populations experience a higher incidence of monogenic forms of infertility [189]. Several studies report that inbreeding increases the relative risk for monogenic recessive disorders; thus, it is reasonable to hypothesize that inbreeding augments the risk of monogenic infertility.

To address these questions, future research should continue combining NGS of infertile patients with mammalian models of reproductive disorders to expand the knowledge on the genetics of infertility. In addition, genomic data from inbred populations will help determine whether high consanguinity rates lead to increased primary infertility cases.

Data availability

The data underlying this article are available in the article.

Abbreviations

NOA:

nonobstructive azoospermia

POI:

primary ovarian insufficiency

MMAF:

multiple morphological abnormalities of the flagella

NGS:

Next generation sequencing

ICSI:

Intracytoplasmic sperm injection

IVF:

In vitro fertilization

References

  1. World Health Organization (WHO). Infertility Prevalence Estimates, 1990–2021 [Internet]. Geneva: World Health Organization (WHO). 2023 Apr pp. 1–79. https://www.who.int/publications/i/item/978920068315.

  2. Steptoe PC, Edwards RG. Birth after the reimplantation of a human embryo. Lancet Lond Engl. 1978;2:366.

    Article  CAS  Google Scholar 

  3. Palermo G, Joris H, Devroey P, Van Steirteghem AC. Pregnancies after intracytoplasmic injection of single spermatozoon into an oocyte. Lancet Lond Engl. 1992;340:17–8.

    Article  CAS  Google Scholar 

  4. Steptoe PC, Edwards RG, Purdy JM. Clinical aspects of pregnancies established with cleaving embryos grown in vitro. Br J Obstet Gynaecol. 1980;87:757–68.

    Article  CAS  PubMed  Google Scholar 

  5. Krausz C, Riera-Escamilla A. Genetics of male infertility. Nat Rev Urol. 2018;15:369–84.

    Article  CAS  PubMed  Google Scholar 

  6. Yatsenko SA, Rajkovic A. Genetics of human female infertility. Biol Reprod. 2019;101:549–66.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Bhakta HH, Refai FH, Avella MA. The molecular mechanisms mediating mammalian fertilization. Development. 2019;146:dev176966.

    Article  CAS  PubMed  Google Scholar 

  8. Fakhro KA, Elbardisi H, Arafa M, Robay A, Rodriguez-Flores JL, Al-Shakaki A, et al. Point-of-care whole-exome sequencing of idiopathic male infertility. Genet Med off J Am Coll Med Genet. 2018;20:1365–73.

    CAS  Google Scholar 

  9. Jiao S-Y, Yang Y-H, Chen S-R. Molecular genetics of infertility: loss-of-function mutations in humans and corresponding knockout/mutated mice. Hum Reprod Update. 2021;27:154–89.

    Article  CAS  PubMed  Google Scholar 

  10. Capel B. Vertebrate sex determination: evolutionary plasticity of a fundamental switch. Nat Rev Genet. 2017;18:675–89.

    Article  CAS  PubMed  Google Scholar 

  11. Garcia-Alonso L, Lorenzi V, Mazzeo CI, Alves-Lopes JP, Roberts K, Sancho-Serra C, et al. Single-cell roadmap of human gonadal development. Nature. 2022;607:540–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Tran D, Muesy-Dessole N, Josso N. Anti-Müllerian hormone is a functional marker of foetal sertoli cells. Nature. 1977;269:411–2.

    Article  CAS  PubMed  Google Scholar 

  13. Sekido R, Lovell-Badge R. Sex determination involves synergistic action of SRY and SF1 on a specific Sox9 enhancer. Nature. 2008;453:930–4.

    Article  CAS  PubMed  Google Scholar 

  14. Swain A, Narvaez V, Burgoyne P, Camerino G, Lovell-Badge R. Dax1 antagonizes sry action in mammalian sex determination. Nature. 1998;391:761–7.

    Article  CAS  PubMed  Google Scholar 

  15. Nachtigal MW, Hirokawa Y, Enyeart-VanHouten DL, Flanagan JN, Hammer GD, Ingraham HA. Wilms’ tumor 1 and Dax-1 modulate the orphan nuclear receptor SF-1 in sex-specific gene expression. Cell. 1998;93:445–54.

    Article  CAS  PubMed  Google Scholar 

  16. Subrini J, Turner J. Y chromosome functions in mammalian spermatogenesis. Capel B, Cheah KSE, editors. eLife. 2021;10:e67345.

  17. Köhler B, Lin L, Ferraz-de-Souza B, Wieacker P, Heidemann P, Schröder V, et al. Five novel mutations in steroidogenic factor 1 (SF1, NR5A1) in 46,XY patients with severe underandrogenization but without adrenal insufficiency. Hum Mutat. 2008;29:59–64.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Bardoni B, Zanaria E, Guioli S, Floridia G, Worley KC, Tonini G, et al. A dosage sensitive locus at chromosome Xp21 is involved in male to female sex reversal. Nat Genet. 1994;7:497–501.

    Article  CAS  PubMed  Google Scholar 

  19. Biason-Lauber A, Konrad D, Navratil F, Schoenle EJ. A WNT4 mutation associated with Müllerian-duct regression and virilization in a 46,XX woman. N Engl J Med. 2004;351:792–8.

    Article  CAS  PubMed  Google Scholar 

  20. Parma P, Radi O, Vidal V, Chaboissier MC, Dellambra E, Valentini S, et al. R-spondin1 is essential in sex determination, skin differentiation and malignancy. Nat Genet. 2006;38:1304–9.

    Article  CAS  PubMed  Google Scholar 

  21. Zhao F, Franco HL, Rodriguez KF, Brown PR, Tsai M-J, Tsai SY, et al. Elimination of the male reproductive tract in the female embryo is promoted by COUP-TFII in mice. Science. 2017;357:717–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Bashamboo A, Eozenou C, Jorgensen A, Bignon-Topalovic J, Siffroi J-P, Hyon C, et al. Loss of function of the Nuclear receptor NR2F2, Encoding COUP-TF2, causes Testis Development and Cardiac defects in 46,XX children. Am J Hum Genet. 2018;102:487–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Carvalheira G, Malinverni AM, Moysés-Oliveira M, Ueta R, Cardili L, Monteagudo P, et al. The natural history of a Man with Ovotesticular 46,XX DSD caused by a Novel 3-Mb 15q26.2 deletion containing NR2F2 gene. J Endocr Soc. 2019;3:2107–13.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Armendares S, Buentello L, Frenk S. Two male sibs with uterus and fallopian tubes. A rare, probably inherited disorder. Clin Genet. 1973;4:291–6.

    Article  CAS  PubMed  Google Scholar 

  25. Beheshti M, Churchill BM, Hardy BE, Bailey JD, Weksberg R, Rogan GF. Familial persistent müllerian duct syndrome. J Urol. 1984;131:968–9.

    Article  CAS  PubMed  Google Scholar 

  26. Behringer RR, Finegold MJ, Cate RL. Müllerian-inhibiting substance function during mammalian sexual development. Cell. 1994;79:415–25.

    Article  CAS  PubMed  Google Scholar 

  27. Brook CG, Wagner H, Zachmann M, Prader A, Armendares S, Frenk S, et al. Familial occurrence of persistent mullerian structures in otherwise normal males. Br Med J. 1973;1:771–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Imbeaud S, Belville C, Messika-Zeitoun L, Rey R, di Clemente N, Josso N, et al. A 27 base-pair deletion of the anti-müllerian type II receptor gene is the most common cause of the persistent müllerian duct syndrome. Hum Mol Genet. 1996;5:1269–77.

    Article  CAS  PubMed  Google Scholar 

  29. Bhangoo A, Wilson R, New MI, Ten S. Donor splice mutation in the 11beta-hydroxylase (CypllB1) gene resulting in sex reversal: a case report and review of the literature. J Pediatr Endocrinol Metab JPEM. 2006;19:1267–82.

    Article  CAS  PubMed  Google Scholar 

  30. Curnow KM, Slutsker L, Vitek J, Cole T, Speiser PW, New MI, et al. Mutations in the CYP11B1 gene causing congenital adrenal hyperplasia and hypertension cluster in exons 6, 7, and 8. Proc Natl Acad Sci U S A. 1993;90:4552–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sipilä P, Junnila A, Hakkarainen J, Huhtaniemi R, Mairinoja L, Zhang FP, et al. The lack of HSD17B3 in male mice results in disturbed Leydig cell maturation and endocrine imbalance akin to humans with HSD17B3 deficiency. FASEB J off Publ Fed Am Soc Exp Biol. 2020;34:6111–28.

    Google Scholar 

  32. Hancock GV, Wamaitha SE, Peretz L, Clark AT. Mammalian primordial germ cell specification. Dev Camb Engl. 2021;148:dev189217.

    CAS  Google Scholar 

  33. Yan Z, Delannoy M, Ling C, Daee D, Osman F, Muniandy PA, et al. A histone-fold complex and FANCM form a conserved DNA-remodeling complex to maintain genome stability. Mol Cell. 2010;37:865–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Yin H, Ma H, Hussain S, Zhang H, Xie X, Jiang L, et al. A homozygous FANCM frameshift pathogenic variant causes male infertility. Genet Med off J Am Coll Med Genet. 2019;21:62–70.

    Google Scholar 

  35. Kasak L, Punab M, Nagirnaja L, Grigorova M, Minajeva A, Lopes AM, et al. Bi-allelic recessive loss-of-function variants in FANCM cause non-obstructive azoospermia. Am J Hum Genet. 2018;103:200–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Luo Y, Hartford SA, Zeng R, Southard TL, Shima N, Schimenti JC. Hypersensitivity of primordial germ cells to compromised replication-associated DNA repair involves ATM-p53-p21 signaling. PLoS Genet. 2014;10:e1004471.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Law NC, Oatley JM. Developmental underpinnings of spermatogonial stem cell establishment. Andrology. 2020;8:852–61.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Fayomi AP, Orwig KE. Spermatogonial stem cells and spermatogenesis in mice, monkeys and men. Stem Cell Res. 2018;29:207–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Ciccarelli M, Giassetti MI, Miao D, Oatley MJ, Robbins C, Lopez-Biladeau B, et al. Donor-derived spermatogenesis following stem cell transplantation in sterile NANOS2 knockout males. Proc Natl Acad Sci U S A. 2020;117:24195–204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ayhan Ö, Balkan M, Guven A, Hazan R, Atar M, Tok A, et al. Truncating mutations in TAF4B and ZMYND15 causing recessive azoospermia. J Med Genet. 2014;51:239–44.

    Article  CAS  PubMed  Google Scholar 

  41. Miyamoto T, Bando Y, Koh E, Tsujimura A, Miyagawa Y, Iijima M, et al. A PLK4 mutation causing azoospermia in a man with sertoli cell-only syndrome. Andrology. 2016;4:75–81.

    Article  CAS  PubMed  Google Scholar 

  42. Ben Khelifa M, Ghieh F, Boudjenah R, Hue C, Fauvert D, Dard R, et al. A MEI1 homozygous missense mutation associated with meiotic arrest in a consanguineous family. Hum Reprod Oxf Engl. 2018;33:1034–7.

    Article  CAS  Google Scholar 

  43. Yang F, Silber S, Leu NA, Oates RD, Marszalek JD, Skaletsky H, et al. TEX11 is mutated in infertile men with azoospermia and regulates genome-wide recombination rates in mouse. EMBO Mol Med. 2015;7:1198–210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Yatsenko AN, Georgiadis AP, Röpke A, Berman AJ, Jaffe T, Olszewska M, et al. X-linked TEX11 mutations, meiotic arrest, and azoospermia in infertile men. N Engl J Med. 2015;372:2097–107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Sha Y, Zheng L, Ji Z, Mei L, Ding L, Lin S, et al. A novel TEX11 mutation induces azoospermia: a case report of infertile brothers and literature review. BMC Med Genet. 2018;19:63.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Gershoni M, Hauser R, Barda S, Lehavi O, Arama E, Pietrokovski S, et al. A new MEIOB mutation is a recurrent cause for azoospermia and testicular meiotic arrest. Hum Reprod Oxf Engl. 2019;34:666–71.

    Article  CAS  Google Scholar 

  47. Krausz C, Riera-Escamilla A, Moreno-Mendoza D, Holleman K, Cioppi F, Algaba F, et al. Genetic dissection of spermatogenic arrest through exome analysis: clinical implications for the management of azoospermic men. Genet Med off J Am Coll Med Genet. 2020;22:1956–66.

    CAS  Google Scholar 

  48. Wang Y, Liu L, Tan C, Meng G, Meng L, Nie H, et al. Novel MEIOB variants cause primary ovarian insufficiency and non-obstructive azoospermia. Front Genet. 2022;13:936264.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zhu X, Hu K, Cheng H, Wu H, Li K, Gao Y et al. Novel MEIOB pathogenic variants including a homozygous non-canonical splicing variant, cause meiotic arrest and human non-obstructive azoospermia. Clin Genet. 2023.

  50. Gershoni M, Hauser R, Yogev L, Lehavi O, Azem F, Yavetz H, et al. A familial study of azoospermic men identifies three novel causative mutations in three new human azoospermia genes. Genet Med off J Am Coll Med Genet. 2017;19:998–1006.

    CAS  Google Scholar 

  51. Fan S, Jiao Y, Khan R, Jiang X, Javed AR, Ali A, et al. Homozygous mutations in C14orf39/SIX6OS1 cause non-obstructive azoospermia and premature ovarian insufficiency in humans. Am J Hum Genet. 2021;108:324–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Salas-Huetos A, Tüttelmann F, Wyrwoll MJ, Kliesch S, Lopes AM, Goncalves J, et al. Disruption of human meiotic telomere complex genes TERB1, TERB2 and MAJIN in men with non-obstructive azoospermia. Hum Genet. 2021;140:217–27.

    Article  CAS  PubMed  Google Scholar 

  53. Meng Q, Shao B, Zhao D, Fu X, Wang J, Li H, et al. Loss of SUN1 function in spermatocytes disrupts the attachment of telomeres to the nuclear envelope and contributes to non-obstructive azoospermia in humans. Hum Genet. 2023;142:531–41.

    Article  CAS  PubMed  Google Scholar 

  54. K I HS. Y W. The TRF1-binding protein TERB1 promotes chromosome movement and telomere rigidity in meiosis. Nat Cell Biol [Internet]. 2014 [cited 2022 Sep 7];16. https://pubmed.ncbi.nlm.nih.gov/24413433/.

  55. He W-B, Tu C-F, Liu Q, Meng L-L, Yuan S-M, Luo A-X, et al. DMC1 mutation that causes human non-obstructive azoospermia and premature ovarian insufficiency identified by whole-exome sequencing. J Med Genet. 2018;55:198–204.

    Article  CAS  PubMed  Google Scholar 

  56. Krausz C, Riera-Escamilla A, Chianese C, Moreno-Mendoza D, Ars E, Rajmil O, et al. From exome analysis in idiopathic azoospermia to the identification of a high-risk subgroup for occult fanconi anemia. Genet Med off J Am Coll Med Genet. 2019;21:189–94.

    CAS  Google Scholar 

  57. Tan Y-Q, Tu C, Meng L, Yuan S, Sjaarda C, Luo A, et al. Loss-of-function mutations in TDRD7 lead to a rare novel syndrome combining congenital cataract and nonobstructive azoospermia in humans. Genet Med off J Am Coll Med Genet. 2019;21:1209–17.

    CAS  Google Scholar 

  58. Kherraf Z-E, Christou-Kent M, Karaouzene T, Amiri-Yekta A, Martinez G, Vargas AS, et al. SPINK2 deficiency causes infertility by inducing sperm defects in heterozygotes and azoospermia in homozygotes. EMBO Mol Med. 2017;9:1132–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Wyrwoll MJ, Temel ŞG, Nagirnaja L, Oud MS, Lopes AM, van der Heijden GW, et al. Bi-allelic mutations in M1AP are a frequent cause of meiotic arrest and severely impaired spermatogenesis leading to male infertility. Am J Hum Genet. 2020;107:342–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Aravin A, Gaidatzis D, Pfeffer S, Lagos-Quintana M, Landgraf P, Iovino N, et al. A novel class of small RNAs bind to MILI protein in mouse testes. Nature. 2006;442:203–7.

    Article  CAS  PubMed  Google Scholar 

  61. Girard A, Sachidanandam R, Hannon GJ, Carmell MA. A germline-specific class of small RNAs binds mammalian piwi proteins. Nature. 2006;442:199–202.

    Article  PubMed  Google Scholar 

  62. Grivna ST, Beyret E, Wang Z, Lin H. A novel class of small RNAs in mouse spermatogenic cells. Genes Dev. 2006;20:1709–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Ozata DM, Gainetdinov I, Zoch A, O’Carroll D, Zamore PD. PIWI-interacting RNAs: small RNAs with big functions. Nat Rev Genet. 2019;20:89–108.

    Article  CAS  PubMed  Google Scholar 

  64. Kherraf Z-E, Cazin C, Bouker A, Fourati Ben Mustapha S, Hennebicq S, Septier A, et al. Whole-exome sequencing improves the diagnosis and care of men with non-obstructive azoospermia. Am J Hum Genet. 2022;109:508–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Lim SL, Qu ZP, Kortschak RD, Lawrence DM, Geoghegan J, Hempfling A-L, et al. HENMT1 and piRNA Stability are required for adult male germ cell transposon repression and to define the Spermatogenic Program in the mouse. PLOS Genet. 2015;11:e1005620.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Ding D, Liu J, Dong K, Midic U, Hess RA, Xie H, et al. PNLDC1 is essential for piRNA 3’ end trimming and transposon silencing during spermatogenesis in mice. Nat Commun. 2017;8:819.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Nishimura T, Nagamori I, Nakatani T, Izumi N, Tomari Y, Kuramochi-Miyagawa S, et al. PNLDC1, mouse pre-piRNA Trimmer, is required for meiotic and post-meiotic male germ cell development. EMBO Rep. 2018;19:e44957.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Nagirnaja L, Mørup N, Nielsen JE, Stakaitis R, Golubickaite I, Oud MS, et al. Variant PNLDC1, defective piRNA Processing, and Azoospermia. N Engl J Med. 2021;385:707–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Arafat M, Har-Vardi I, Harlev A, Levitas E, Zeadna A, Abofoul-Azab M, et al. Mutation in TDRD9 causes non-obstructive azoospermia in infertile men. J Med Genet. 2017;54:633–9.

    Article  CAS  PubMed  Google Scholar 

  70. Feng R, Sang Q, Kuang Y, Sun X, Yan Z, Zhang S, et al. Mutations in TUBB8 and human oocyte meiotic arrest. N Engl J Med. 2016;374:223–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Cao T, Guo J, Xu Y, Lin X, Deng W, Cheng L, et al. Two mutations in TUBB8 cause developmental arrest in human oocytes and early embryos. Reprod Biomed Online. 2021;43:891–8.

    Article  CAS  PubMed  Google Scholar 

  72. So C, Menelaou K, Uraji J, Harasimov K, Steyer AM, Seres KB, et al. Mechanism of spindle Pole organization and instability in human oocytes. Science. 2022;375:eabj3944.

    Article  CAS  PubMed  Google Scholar 

  73. Al-Agha AE, Ahmed IA, Nuebel E, Moriwaki M, Moore B, Peacock KA, et al. Primary ovarian insufficiency and azoospermia in carriers of a homozygous PSMC3IP stop Gain Mutation. J Clin Endocrinol Metab. 2018;103:555–63.

    Article  PubMed  Google Scholar 

  74. AlAsiri S, Basit S, Wood-Trageser MA, Yatsenko SA, Jeffries EP, Surti U, et al. Exome sequencing reveals MCM8 mutation underlies ovarian failure and chromosomal instability. J Clin Invest. 2015;125:258–62.

    Article  PubMed  Google Scholar 

  75. Bouali N, Francou B, Bouligand J, Imanci D, Dimassi S, Tosca L, et al. New MCM8 mutation associated with premature ovarian insufficiency and chromosomal instability in a highly consanguineous Tunisian family. Fertil Steril. 2017;108:694–702.

    Article  CAS  PubMed  Google Scholar 

  76. Tenenbaum-Rakover Y, Weinberg-Shukron A, Renbaum P, Lobel O, Eideh H, Gulsuner S, et al. Minichromosome maintenance complex component 8 (MCM8) gene mutations result in primary gonadal failure. J Med Genet. 2015;52:391–9.

    Article  PubMed  Google Scholar 

  77. Zhang Y-X, He W-B, Xiao W-J, Meng L-L, Tan C, Du J, et al. Novel loss-of-function mutation in MCM8 causes premature ovarian insufficiency. Mol Genet Genomic Med. 2020;8:e1165.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Fauchereau F, Shalev S, Chervinsky E, Beck-Fruchter R, Legois B, Fellous M, et al. A non-sense MCM9 mutation in a familial case of primary ovarian insufficiency. Clin Genet. 2016;89:603–7.

    Article  CAS  PubMed  Google Scholar 

  79. Wood-Trageser MA, Gurbuz F, Yatsenko SA, Jeffries EP, Kotan LD, Surti U, et al. MCM9 mutations are associated with ovarian failure, short stature, and chromosomal instability. Am J Hum Genet. 2014;95:754–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Lutzmann M, Grey C, Traver S, Ganier O, Maya-Mendoza A, Ranisavljevic N, et al. MCM8- and MCM9-deficient mice reveal gametogenesis defects and genome instability due to impaired homologous recombination. Mol Cell. 2012;47:523–34.

    Article  CAS  PubMed  Google Scholar 

  81. Kumar R, Oliver C, Brun C, Juarez-Martinez AB, Tarabay Y, Kadlec J, et al. Mouse REC114 is essential for meiotic DNA double-strand break formation and forms a complex with MEI4. Life Sci Alliance. 2018;1:e201800259.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Wang W, Dong J, Chen B, Du J, Kuang Y, Sun X, et al. Homozygous mutations in REC114 cause female infertility characterised by multiple pronuclei formation and early embryonic arrest. J Med Genet. 2020;57:187–94.

    Article  CAS  PubMed  Google Scholar 

  83. Jaillard S, McElreavy K, Robevska G, Akloul L, Ghieh F, Sreenivasan R, et al. STAG3 homozygous missense variant causes primary ovarian insufficiency and male non-obstructive azoospermia. Mol Hum Reprod. 2020;26:665–77.

    Article  CAS  PubMed  Google Scholar 

  84. Llano E, Gomez- HL, García-Tuñón I, Sánchez-Martín M, Caburet S, Barbero JL, et al. STAG3 is a strong candidate gene for male infertility. Hum Mol Genet. 2014;23:3421–31.

    Article  CAS  PubMed  Google Scholar 

  85. Sánchez-Sáez F, Gómez-H L, Dunne OM, Gallego-Páramo C, Felipe-Medina N, Sánchez-Martín M, et al. Meiotic chromosome synapsis depends on multivalent SYCE1-SIX6OS1 interactions that are disrupted in cases of human infertility. Sci Adv. 2020;6:eabb1660.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Carlosama C, Elzaiat M, Patiño LC, Mateus HE, Veitia RA, Laissue P. A homozygous donor splice-site mutation in the meiotic gene MSH4 causes primary ovarian insufficiency. Hum Mol Genet. 2017;26:3161–6.

    CAS  PubMed  Google Scholar 

  87. Li P, Ji Z, Zhi E, Zhang Y, Han S, Zhao L, et al. Novel bi-allelic MSH4 variants causes meiotic arrest and non-obstructive azoospermia. Reprod Biol Endocrinol RBE. 2022;20:21.

    Article  CAS  Google Scholar 

  88. Wyrwoll MJ, van Walree ES, Hamer G, Rotte N, Motazacker MM, Meijers-Heijboer H, et al. Bi-allelic variants in DNA mismatch repair proteins MutS Homolog MSH4 and MSH5 cause infertility in both sexes. Hum Reprod Oxf Engl. 2021;37:178–89.

    Article  CAS  Google Scholar 

  89. Yang Y, Guo J, Dai L, Zhu Y, Hu H, Tan L, et al. XRCC2 mutation causes meiotic arrest, azoospermia and infertility. J Med Genet. 2018;55:628–36.

    Article  CAS  PubMed  Google Scholar 

  90. Zhang Y-X, Li H-Y, He W-B, Tu C, Du J, Li W, et al. XRCC2 mutation causes premature ovarian insufficiency as well as non-obstructive azoospermia in humans. Clin Genet. 2019;95:442–3.

    Article  CAS  PubMed  Google Scholar 

  91. Guiraldelli MF, Eyster C, Wilkerson JL, Dresser ME, Pezza RJ. Mouse HFM1/Mer3 is required for crossover formation and complete synapsis of homologous chromosomes during meiosis. PLoS Genet. 2013;9:e1003383.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Wang J, Zhang W, Jiang H, Wu B-L. Primary ovarian insufficiency collaboration. Mutations in HFM1 in recessive primary ovarian insufficiency. N Engl J Med. 2014;370:972–4.

    Article  CAS  PubMed  Google Scholar 

  93. Horn HF, Kim DI, Wright GD, Wong ESM, Stewart CL, Burke B, et al. A mammalian KASH domain protein coupling meiotic chromosomes to the cytoskeleton. J Cell Biol. 2013;202:1023–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Hou X, Zeb A, Dil S, Zhou J, Zhang H, Shi B, et al. A homozygous KASH5 frameshift mutation causes diminished ovarian reserve, recurrent miscarriage, and non-obstructive azoospermia in humans. Front Endocrinol. 2023;14:1128362.

    Article  Google Scholar 

  95. Sazegari A, Kalantar SM, Pashaiefar H, Mohtaram S, Honarvar N, Feizollahi Z, et al. The T657C polymorphism on the SYCP3 gene is associated with recurrent pregnancy loss. J Assist Reprod Genet. 2014;31:1377–81.

    Article  PubMed  PubMed Central  Google Scholar 

  96. de Vries L, Behar DM, Smirin-Yosef P, Lagovsky I, Tzur S, Basel-Vanagaite L. Exome sequencing reveals SYCE1 mutation associated with autosomal recessive primary ovarian insufficiency. J Clin Endocrinol Metab. 2014;99:E2129–2132.

    Article  PubMed  Google Scholar 

  97. Clairmont CS, Sarangi P, Ponnienselvan K, Galli LD, Csete I, Moreau L, et al. TRIP13 regulates DNA repair pathway choice through REV7 conformational change. Nat Cell Biol. 2020;22:87–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Zhang Z, Li B, Fu J, Li R, Diao F, Li C, et al. Bi-allelic missense pathogenic variants in TRIP13 cause female infertility characterized by oocyte maturation arrest. Am J Hum Genet. 2020;107:15–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Pacheco S, Marcet-Ortega M, Lange J, Jasin M, Keeney S, Roig I. The ATM signaling cascade promotes recombination-dependent pachytene arrest in mouse spermatocytes. PLoS Genet. 2015;11:e1005017.

    Article  PubMed  PubMed Central  Google Scholar 

  100. Shin Y-H, Ren Y, Suzuki H, Golnoski KJ, Ahn HW, Mico V, et al. Transcription factors SOHLH1 and SOHLH2 coordinate oocyte differentiation without affecting meiosis I. J Clin Invest. 2017;127:2106–17.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Suzuki H, Ahn HW, Chu T, Bowden W, Gassei K, Orwig K, et al. SOHLH1 and SOHLH2 coordinate spermatogonial differentiation. Dev Biol. 2012;361:301–12.

    Article  CAS  PubMed  Google Scholar 

  102. Bayram Y, Gulsuner S, Guran T, Abaci A, Yesil G, Gulsuner HU, et al. Homozygous loss-of-function mutations in SOHLH1 in patients with nonsyndromic hypergonadotropic hypogonadism. J Clin Endocrinol Metab. 2015;100:E808–814.

    Article  PubMed  PubMed Central  Google Scholar 

  103. Khawar MB, Gao H, Li W. Mechanism of Acrosome Biogenesis in mammals. Front Cell Dev Biol. 2019;7:195.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Celse T, Cazin C, Mietton F, Martinez G, Martinez D, Thierry-Mieg N, et al. Genetic analyses of a large cohort of infertile patients with globozoospermia, DPY19L2 still the main actor, GGN confirmed as a guest player. Hum Genet. 2021;140:43–57.

    Article  CAS  PubMed  Google Scholar 

  105. ElInati E, Fossard C, Okutman O, Ghédir H, Ibala-Romdhane S, Ray PF, et al. A new mutation identified in SPATA16 in two globozoospermic patients. J Assist Reprod Genet. 2016;33:815–20.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Oud MS, Okutman Ö, Hendricks LaJ, de Vries PF, Houston BJ, Vissers LELM, et al. Exome sequencing reveals novel causes as well as new candidate genes for human globozoospermia. Hum Reprod Oxf Engl. 2020;35:240–52.

    Article  CAS  Google Scholar 

  107. Arafah K, Lopez F, Cazin C, Kherraf Z-E, Tassistro V, Loundou A, et al. Defect in the nuclear pore membrane glycoprotein 210-like gene is associated with extreme uncondensed sperm nuclear chromatin and male infertility: a case report. Hum Reprod Oxf Engl. 2021;36:693–701.

    Article  Google Scholar 

  108. Lehti MS, Sironen A. Formation and function of the manchette and flagellum during spermatogenesis. Reprod Camb Engl. 2016;151:R43–54.

    Article  CAS  Google Scholar 

  109. Dai J, Li Q, Zhou Q, Zhang S, Chen J, Wang Y, et al. IQCN disruption causes fertilization failure and male infertility due to manchette assembly defect. EMBO Mol Med. 2022;14:e16501.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Chen H, Zhu Y, Zhu Z, Zhi E, Lu K, Wang X, et al. Detection of heterozygous mutation in hook microtubule-tethering protein 1 in three patients with decapitated and decaudated spermatozoa syndrome. J Med Genet. 2018;55:150–7.

    Article  CAS  PubMed  Google Scholar 

  111. Kazarian E, Son H, Sapao P, Li W, Zhang Z, Strauss JF, et al. SPAG17 is required for male germ cell differentiation and fertility. Int J Mol Sci. 2018;19:E1252.

    Article  Google Scholar 

  112. Lin Y-H, Wang Y-Y, Lai T-H, Teng J-L, Lin C-W, Ke C-C et al. Deleterious genetic changes in AGTPBP1 result in teratozoospermia with sperm head and flagella defects. J Cell Mol Med. 2023.

  113. Dong FN, Amiri-Yekta A, Martinez G, Saut A, Tek J, Stouvenel L, et al. Absence of CFAP69 causes male infertility due to multiple morphological abnormalities of the Flagella in Human and Mouse. Am J Hum Genet. 2018;102:636–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Liu Q, Guo Q, Guo W, Song S, Wang N, Chen X, et al. Loss of CEP70 function affects acrosome biogenesis and flagella formation during spermiogenesis. Cell Death Dis. 2021;12:478.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Ruan T, Yang Y, Jiang C, Shen G, Li D, Shen Y. Identification of biallelic variations of CEP70 in patients with male infertility. Front Endocrinol. 2023;14:1133222.

    Article  Google Scholar 

  116. Xu X, Sha Y-W, Mei L-B, Ji Z-Y, Qiu P-P, Ji H, et al. A familial study of twins with severe asthenozoospermia identified a homozygous SPAG17 mutation by whole-exome sequencing. Clin Genet. 2018;93:345–9.

    Article  CAS  PubMed  Google Scholar 

  117. Dougherty GW, Mizuno K, Nöthe-Menchen T, Ikawa Y, Boldt K, Ta-Shma A, et al. CFAP45 deficiency causes situs abnormalities and asthenospermia by disrupting an axonemal adenine nucleotide homeostasis module. Nat Commun. 2020;11:5520.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Avidor-Reiss T, Mazur M, Fishman EL, Sindhwani P. The role of sperm Centrioles in Human Reproduction - the known and the unknown. Front Cell Dev Biol. 2019;7:188.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Sha Y-W, Xu X, Mei L-B, Li P, Su Z-Y, He X-Q, et al. A homozygous CEP135 mutation is associated with multiple morphological abnormalities of the sperm flagella (MMAF). Gene. 2017;633:48–53.

    Article  CAS  PubMed  Google Scholar 

  120. Zhang X, Wang L, Ma Y, Wang Y, Liu H, Liu M, et al. CEP128 is involved in spermatogenesis in humans and mice. Nat Commun. 2022;13:1395.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Liu W, He X, Yang S, Zouari R, Wang J, Wu H, et al. Bi-allelic mutations in TTC21A induce asthenoteratospermia in humans and mice. Am J Hum Genet. 2019;104:738–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Liu C, He X, Liu W, Yang S, Wang L, Li W, et al. Bi-allelic mutations in TTC29 cause male subfertility with asthenoteratospermia in humans and mice. Am J Hum Genet. 2019;105:1168–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Lorès P, Dacheux D, Kherraf Z-E, Nsota Mbango J-F, Coutton C, Stouvenel L, et al. Mutations in TTC29, encoding an evolutionarily conserved axonemal protein, result in Asthenozoospermia and male infertility. Am J Hum Genet. 2019;105:1148–67.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Baccetti B, Collodel G, Estenoz M, Manca D, Moretti E, Piomboni P. Gene deletions in an infertile man with sperm fibrous sheath dysplasia. Hum Reprod Oxf Engl. 2005;20:2790–4.

    Article  CAS  Google Scholar 

  125. Liu W, Wu H, Wang L, Yang X, Liu C, He X, et al. Homozygous loss-of-function mutations in FSIP2 cause male infertility with asthenoteratospermia. J Genet Genomics Yi Chuan Xue Bao. 2019;46:53–6.

    Article  CAS  PubMed  Google Scholar 

  126. Lehti MS, Zhang F-P, Kotaja N, Sironen A. SPEF2 functions in microtubule-mediated transport in elongating spermatids to ensure proper male germ cell differentiation. Dev Camb Engl. 2017;144:2683–93.

    CAS  Google Scholar 

  127. Liu W, Sha Y, Li Y, Mei L, Lin S, Huang X, et al. Loss-of-function mutations in SPEF2 cause multiple morphological abnormalities of the sperm flagella (MMAF). J Med Genet. 2019;56:678–84.

    Article  CAS  PubMed  Google Scholar 

  128. Amiri-Yekta A, Coutton C, Kherraf Z-E, Karaouzène T, Le Tanno P, Sanati MH, et al. Whole-exome sequencing of familial cases of multiple morphological abnormalities of the sperm flagella (MMAF) reveals new DNAH1 mutations. Hum Reprod Oxf Engl. 2016;31:2872–80.

    Article  CAS  Google Scholar 

  129. Ben Khelifa M, Coutton C, Zouari R, Karaouzène T, Rendu J, Bidart M, et al. Mutations in DNAH1, which encodes an inner arm heavy chain dynein, lead to male infertility from multiple morphological abnormalities of the sperm flagella. Am J Hum Genet. 2014;94:95–104.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Sha Y, Yang X, Mei L, Ji Z, Wang X, Ding L, et al. DNAH1 gene mutations and their potential association with dysplasia of the sperm fibrous sheath and infertility in the Han Chinese population. Fertil Steril. 2017;107:1312–e13182.

    Article  CAS  PubMed  Google Scholar 

  131. Wang X, Jin H, Han F, Cui Y, Chen J, Yang C, et al. Homozygous DNAH1 frameshift mutation causes multiple morphological anomalies of the sperm flagella in Chinese. Clin Genet. 2017;91:313–21.

    Article  CAS  PubMed  Google Scholar 

  132. Hwang JY, Nawaz S, Choi J, Wang H, Hussain S, Nawaz M, et al. Genetic defects in DNAH2 Underlie Male Infertility with multiple morphological abnormalities of the sperm flagella in humans and mice. Front Cell Dev Biol. 2021;9:662903.

    Article  PubMed  PubMed Central  Google Scholar 

  133. Li Y, Sha Y, Wang X, Ding L, Liu W, Ji Z, et al. DNAH2 is a novel candidate gene associated with multiple morphological abnormalities of the sperm flagella. Clin Genet. 2019;95:590–600.

    Article  CAS  PubMed  Google Scholar 

  134. Zhang B, Ma H, Khan T, Ma A, Li T, Zhang H, et al. A DNAH17 missense variant causes flagella destabilization and asthenozoospermia. J Exp Med. 2020;217:e20182365.

    Article  PubMed  Google Scholar 

  135. Tu C, Nie H, Meng L, Yuan S, He W, Luo A, et al. Identification of DNAH6 mutations in infertile men with multiple morphological abnormalities of the sperm flagella. Sci Rep. 2019;9:15864.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Li L, Sha Y-W, Xu X, Mei L-B, Qiu P-P, Ji Z-Y et al. DNAH6 is a novel candidate gene associated with sperm head anomaly. Andrologia. 2018.

  137. Wang R, Yang D, Tu C, Lei C, Ding S, Guo T et al. Dynein axonemal heavy chain 10 deficiency causes primary ciliary dyskinesia in humans and mice. Front Med. 2023.

  138. Wu H, Liu Y, Li Y, Li K, Xu C, Gao Y, et al. DNALI1 deficiency causes male infertility with severe asthenozoospermia in humans and mice by disrupting the assembly of the flagellar inner dynein arms and fibrous sheath. Cell Death Dis. 2023;14:127.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Zhao X, Ge H, Xu W, Cheng C, Zhou W, Xu Y et al. Lack of CFAP54 causes primary ciliary dyskinesia in a mouse model and human patients. Front Med. 2023.

  140. Sha Y-W, Wang X, Xu X, Su Z-Y, Cui Y, Mei L-B, et al. Novel mutations in CFAP44 and CFAP43 cause multiple morphological abnormalities of the sperm flagella (MMAF). Reprod Sci. 2019;26:26–34.

    Article  CAS  PubMed  Google Scholar 

  141. Tang S, Wang X, Li W, Yang X, Li Z, Liu W, et al. Biallelic mutations in CFAP43 and CFAP44 cause male infertility with multiple morphological abnormalities of the sperm flagella. Am J Hum Genet. 2017;100:854–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Wu H, Li W, He X, Liu C, Fang Y, Zhu F, et al. NovelCFAP43 andCFAP44 mutations cause male infertility with multiple morphological abnormalities of the sperm flagella (MMAF). Reprod Biomed Online. 2019;38:769–78.

    Article  CAS  PubMed  Google Scholar 

  143. Ma A, Zhou J, Ali H, Abbas T, Ali I, Muhammad Z, et al. Loss-of-function mutations in CFAP57 cause multiple morphological abnormalities of the flagella in humans and mice. JCI Insight. 2023;8:e166869.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Liu C, Tu C, Wang L, Wu H, Houston BJ, Mastrorosa FK, et al. Deleterious variants in X-linked CFAP47 induce asthenoteratozoospermia and primary male infertility. Am J Hum Genet. 2021;108:309–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Beurois J, Martinez G, Cazin C, Kherraf Z-E, Amiri-Yekta A, Thierry-Mieg N, et al. CFAP70 mutations lead to male infertility due to severe astheno-teratozoospermia. A case report. Hum Reprod. 2019;34:2071–9.

    Article  PubMed  Google Scholar 

  146. Kherraf Z-E, Amiri-Yekta A, Dacheux D, Karaouzène T, Coutton C, Christou-Kent M, et al. A homozygous ancestral SVA-Insertion-mediated deletion in WDR66 induces multiple morphological abnormalities of the sperm flagellum and male infertility. Am J Hum Genet. 2018;103:400–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Liu S, Zhang J, Kherraf ZE, Sun S, Zhang X, Cazin C, et al. CFAP61 is required for sperm flagellum formation and male fertility in human and mouse. Dev Camb Engl. 2021;148:dev199805.

    CAS  Google Scholar 

  148. Shen Q, Martinez G, Liu H, Beurois J, Wu H, Amiri-Yekta A, et al. Bi-allelic truncating variants in CFAP206 cause male infertility in human and mouse. Hum Genet. 2021;140:1367–77.

    Article  CAS  PubMed  Google Scholar 

  149. Martinez G, Beurois J, Dacheux D, Cazin C, Bidart M, Kherraf Z-E, et al. Biallelic variants in MAATS1 encoding CFAP91, a calmodulin-associated and spoke-associated complex protein, cause severe astheno-teratozoospermia and male infertility. J Med Genet. 2020;57:708–16.

    Article  CAS  PubMed  Google Scholar 

  150. Kherraf Z-E, Cazin C, Coutton C, Amiri-Yekta A, Martinez G, Boguenet M, et al. Whole exome sequencing of men with multiple morphological abnormalities of the sperm flagella reveals novel homozygous QRICH2 mutations. Clin Genet. 2019;96:394–401.

    Article  CAS  PubMed  Google Scholar 

  151. Shen Y, Zhang F, Li F, Jiang X, Yang Y, Li X, et al. Loss-of-function mutations in QRICH2 cause male infertility with multiple morphological abnormalities of the sperm flagella. Nat Commun. 2019;10:433.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Oud MS, Smits RM, Smith HE, Mastrorosa FK, Holt GS, Houston BJ, et al. A de novo paradigm for male infertility. Nat Commun. 2022;13:154.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. França MM, Funari MFA, Nishi MY, Narcizo AM, Domenice S, Costa EMF, et al. Identification of the first homozygous 1-bp deletion in GDF9 gene leading to primary ovarian insufficiency by using targeted massively parallel sequencing. Clin Genet. 2018;93:408–11.

    Article  PubMed  Google Scholar 

  154. Huang C, Zhao S, Yang Y, Guo T, Ke H, Mi X, et al. TP63 gain-of-function mutations cause premature ovarian insufficiency by inducing oocyte apoptosis. J Clin Invest. 2023;133:e162315.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Chen B, Zhang Z, Sun X, Kuang Y, Mao X, Wang X, et al. Biallelic mutations in PATL2 cause female infertility characterized by oocyte maturation arrest. Am J Hum Genet. 2017;101:609–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Christou-Kent M, Kherraf Z, Amiri‐Yekta A, Le Blévec E, Karaouzène T, Conne B, et al. PATL2 is a key actor of oocyte maturation whose invalidation causes infertility in women and mice. EMBO Mol Med. 2018;10:e8515.

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  158. Meduri G, Touraine P, Beau I, Lahuna O, Desroches A, Vacher-Lavenu MC, et al. Delayed puberty and primary amenorrhea associated with a novel mutation of the human follicle-stimulating hormone receptor: clinical, histological, and molecular studies. J Clin Endocrinol Metab. 2003;88:3491–8.

    Article  CAS  PubMed  Google Scholar 

  159. Huang H-L, Lv C, Zhao Y-C, Li W, He X-M, Li P, et al. Mutant ZP1 in familial infertility. N Engl J Med. 2014;370:1220–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Yuan P, Li R, Li D, Zheng L, Ou S, Zhao H, et al. Novel mutation in the ZP1 gene and clinical implications. J Assist Reprod Genet. 2019;36:741–7.

    Article  PubMed  PubMed Central  Google Scholar 

  161. Nishimura K, Dioguardi E, Nishio S, Villa A, Han L, Matsuda T, et al. Molecular basis of egg coat cross-linking sheds light on ZP1-associated female infertility. Nat Commun. 2019;10:1–15.

    Article  Google Scholar 

  162. Cao Q, Zhao C, Zhang X, Zhang H, Lu Q, Wang C, et al. Heterozygous mutations in ZP1 and ZP3 cause formation disorder of ZP and female infertility in human. J Cell Mol Med. 2020;24:8557–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Luo G, Zhu L, Liu Z, Yang X, Xi Q, Li Z, et al. Novel mutations in ZP1 and ZP2 cause primary infertility due to empty follicle syndrome and abnormal zona pellucida. J Assist Reprod Genet. 2020;37:2853–60.

    Article  PubMed  PubMed Central  Google Scholar 

  164. Liu W, Li K, Bai D, Yin J, Tang Y, Chi F, et al. Dosage effects of ZP2 and ZP3 heterozygous mutations cause human infertility. Hum Genet. 2017;136:975–85.

    Article  CAS  PubMed  Google Scholar 

  165. Chen T, Bian Y, Liu X, Zhao S, Wu K, Yan L, et al. A recurrent missense mutation in ZP3 causes empty follicle syndrome and female infertility. Am J Hum Genet. 2017;101:459–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Sang Q, Li B, Kuang Y, Wang X, Zhang Z, Chen B, et al. Homozygous mutations in WEE2 cause fertilization failure and female infertility. Am J Hum Genet. 2018;102:649–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Zhang Z, Mu J, Zhao J, Zhou Z, Chen B, Wu L, et al. Novel mutations in WEE2: expanding the spectrum of mutations responsible for human fertilization failure. Clin Genet. 2019;95:520–4.

    Article  CAS  PubMed  Google Scholar 

  168. Cavarocchi E, Whitfield M, Chargui A, Stouvenel L, Lorès P, Coutton C, et al. The sodium/proton exchanger SLC9C1 (sNHE) is essential for human sperm motility and fertility. Clin Genet. 2021;99:684–93.

    Article  CAS  PubMed  Google Scholar 

  169. Liu R, Yan Z, Fan Y, Qu R, Chen B, Li B, et al. Bi-allelic variants in KCNU1 cause impaired acrosome reactions and male infertility. Hum Reprod Oxf Engl. 2022;37:1394–405.

    Article  CAS  Google Scholar 

  170. Hua R, Xue R, Liu Y, Li Y, Sha X, Li K et al. ACROSIN deficiency causes total fertilization failure in humans by preventing the sperm from penetrating the zona pellucida. Hum Reprod Oxf Engl. 2023;dead059.

  171. Hirose M, Honda A, Fulka H, Tamura-Nakano M, Matoba S, Tomishima T, et al. Acrosin is essential for sperm penetration through the zona pellucida in hamsters. Proc Natl Acad Sci. 2020;117:2513–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Avella MA, Baibakov B, Dean J. A single domain of the ZP2 zona pellucida protein mediates gamete recognition in mice and humans. J Cell Biol. 2014;205:801–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Dai C, Hu L, Gong F, Tan Y, Cai S, Zhang S et al. ZP2 pathogenic variants cause in vitro fertilization failure and female infertility. Genet Med off J Am Coll Med Genet. 2018.

  174. Chung J-J, Shim S-H, Everley RA, Gygi SP, Zhuang X, Clapham DE. Structurally distinct ca(2+) signaling domains of sperm flagella orchestrate tyrosine phosphorylation and motility. Cell. 2014;157:808–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Avenarius MR, Hildebrand MS, Zhang Y, Meyer NC, Smith LLH, Kahrizi K, et al. Human male infertility caused by mutations in the CATSPER1 channel protein. Am J Hum Genet. 2009;84:505–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Dgany O, Avidan N, Delaunay J, Krasnov T, Shalmon L, Shalev H, et al. Congenital dyserythropoietic Anemia type I is caused by mutations in Codanin-1. Am J Hum Genet. 2002;71:1467–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Aydin H, Sultana A, Li S, Thavalingam A, Lee JE. Molecular architecture of the human sperm IZUMO1 and egg JUNO fertilization complex. Nature. 2016;534:562–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Dai J, Dai C, Guo J, Zheng W, Zhang T, Li Y, et al. Novel homozygous variations in PLCZ1 lead to poor or failed fertilization characterized by abnormal localization patterns of PLCζ in sperm. Clin Genet. 2020;97:347–51.

    Article  CAS  PubMed  Google Scholar 

  179. Escoffier J, Lee HC, Yassine S, Zouari R, Martinez G, Karaouzène T, et al. Homozygous mutation of PLCZ1 leads to defective human oocyte activation and infertility that is not rescued by the WW-binding protein PAWP. Hum Mol Genet. 2016;25:878–91.

    Article  CAS  PubMed  Google Scholar 

  180. Peng Y, Lin Y, Deng K, Shen J, Cui Y, Liu J, et al. Mutations in PLCZ1 induce male infertility associated with polyspermy and fertilization failure. J Assist Reprod Genet. 2023;40:53–64.

    Article  PubMed  Google Scholar 

  181. Dai J, Zhang T, Guo J, Zhou Q, Gu Y, Zhang J, et al. Homozygous pathogenic variants in ACTL9 cause fertilization failure and male infertility in humans and mice. Am J Hum Genet. 2021;108:469–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Wang J, Zhang J, Sun X, Lin Y, Cai L, Cui Y, et al. Novel bi-allelic variants in ACTL7A are associated with male infertility and total fertilization failure. Hum Reprod Oxf Engl. 2021;36:3161–9.

    Article  CAS  Google Scholar 

  183. Hachem A, Godwin J, Ruas M, Lee HC, Ferrer Buitrago M, Ardestani G, et al. PLCζ is the physiological trigger of the Ca2 + oscillations that induce embryogenesis in mammals but conception can occur in its absence. Dev Camb Engl. 2017;144:2914–24.

    CAS  Google Scholar 

  184. Blakeley P, Fogarty NME, del Valle I, Wamaitha SE, Hu TX, Elder K, et al. Defining the three cell lineages of the human blastocyst by single-cell RNA-seq. Dev Camb Engl. 2015;142:3151–65.

    CAS  Google Scholar 

  185. Fogarty NME, McCarthy A, Snijders KE, Powell BE, Kubikova N, Blakeley P, et al. Genome editing reveals a role for OCT4 in human embryogenesis. Nature. 2017;550:67–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Martinez G, Coutton C, Loeuillet C, Cazin C, Muroňová J, Boguenet M, et al. Oligogenic heterozygous inheritance of sperm abnormalities in mouse. eLife. 2022;11:e75373.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Loubalova Z, Fulka H, Horvat F, Pasulka J, Malik R, Hirose M, et al. Formation of spermatogonia and fertile oocytes in golden hamsters requires piRNAs. Nat Cell Biol. 2021;23:992–1001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Hasuwa H, Iwasaki YW, Au Yeung WK, Ishino K, Masuda H, Sasaki H, et al. Production of functional oocytes requires maternally expressed PIWI genes and piRNAs in golden hamsters. Nat Cell Biol. 2021;23:1002–12.

    Article  CAS  PubMed  Google Scholar 

  189. Romdhane L, Mezzi N, Hamdi Y, El-Kamah G, Barakat A, Abdelhak S. Consanguinity and inbreeding in Health and Disease in north African populations. Annu Rev Genomics Hum Genet. 2019;20:155–79.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We are grateful to our colleagues in the Sidra Medicine Research and Reproductive Medicine Unit for their valuable feedback and discussion. We apologize to colleagues whose work could not be cited owing to space limitations.

Funding

This study was supported by the by the Sidra Medicine Grant (SDR400185) to MA, the FERTILITY-IQ Study, NPRP12S-0318-190394 from the Qatar National Research Fund (QNRF) to KF and MA.

Author information

Authors and Affiliations

Authors

Contributions

K.F. and J.A. contributed equally to this work. Literature research: M.A., K.F., J.A. Figures: M.A., S.G. Writing and editing manuscript: M.A., S.G., L.S.

Corresponding author

Correspondence to Matteo Avella.

Ethics declarations

Conflict of interest

The authors declare no conflict of interest.

Additional information

Publisher’s Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fakhro, K.A., Awwad, J., Garibova, S. et al. Conserved genes regulating human sex differentiation, gametogenesis and fertilization. J Transl Med 22, 473 (2024). https://doi.org/10.1186/s12967-024-05162-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-024-05162-2

Keywords