Skip to main content

Genetic insights into the complexity of premature ovarian insufficiency

Abstract

Premature Ovarian Insufficiency (POI) is a highly heterogeneous condition characterized by ovarian dysfunction in women occurring before the age of 40, representing a significant cause of female infertility. It manifests through primary or secondary amenorrhea. While more than half of POI cases are idiopathic, genetic factors play a pivotal role in all instances with known causes, contributing to approximately 20–25% of cases. This article comprehensively reviews the genetic factors associated with POI, delineating the primary candidate genes. The discussion delves into the intricate relationship between these genes and ovarian development, elucidating the functional consequences of diverse mutations to underscore the fundamental impact of genetic effects on POI. The identified genetic factors, encompassing gene mutations and chromosomal abnormalities, are systematically classified based on whether the resulting POI is syndromic or non-syndromic. Furthermore, this paper explores the genetic interplay between mitochondrial genes, such as Required for Meiotic Nuclear Division 1 homolog Gene (RMND1), Mitochondrial Ribosomal Protein S22 Gene (MRPS22), Leucine-rich Pentapeptide Repeat Gene (LRPPRC), and non-coding RNAs, including both microRNAs and Long non-coding RNAs, with POI. The insights provided serve to consolidate and enhance our understanding of the etiology of POI, contributing to establishing a theoretical foundation for diagnosing and treating POI patients, as well as for exploring the mechanisms underlying the disease.

Background

In April, 2023, the World Health Organization reported that approximately 17.5% of couples of childbearing age worldwide experience infertility. Female factors account for 30–50% of these cases [1,2,3,4]. Premature Ovarian Insufficiency (POI) emerges as a significant cause of female infertility. POI is characterized by the cessation of ovarian function in women before the age of 40, leading to amenorrhea and infertility. Diagnostic criteria include a minimum of 4 months of amenorrhea, elevated Follicle-stimulating hormone (FSH) levels exceeding 25 IU/L, and/or reduced estrogen levels persisting for at least 4 weeks, along with the absence of antral follicles in ovarian ultrasound examination [5, 6]. POI symptoms encompass hot flashes, sweating, tension, loss of libido, weakness, dry skin, mucous membrane dryness, reduced bone mineral density, and metabolic disorders, with infertility being the most profound manifestation [7].

As a heterogeneous condition, POI has various causes, including genetic factors (20–25%), autoimmune factors, iatrogenic influences (both contributing around 10%), and environmental factors (Fig. 1). Nevertheless, the etiology for more than half of patients remains elusive [8].

In a recent study involving 1030 POI patients, 242 cases (23.5%) were identified as associated with the pathogenicity and potential pathogenic mutations of POI-related genes, both known and novel [9]. More than 50 gene mutations associated with POI have been identified, impacting diverse processes, including gonadal development, DNA replication/meiosis, DNA repair, transcription processes, signal transduction, RNA metabolism and translation, and mitochondrial function [9]. Recent studies have unveiled a potential connection between non-coding RNAs (ncRNAs) and POI, such as miRNAs, lnc-RNAs [10,11,12,13]. In addition to non-syndrome-associated POI, there are gene variants with POI as a clinical phenotype, and some chromosomal abnormalities contribute to 10–13% of POI cases [14]. Genetic research on these symptomatic POIs is progressively advancing.

This literature review aims to synthesize genetic factors and potential mechanisms related to POI across four dimensions: syndrome-associated POI, non-syndrome POI, mitochondrial dysfunction-associated POI, and non-coding RNA-associated POI. For non-syndrome-associated POI, we categorize the genetic factors based on the roles played by different genes at various stages of follicular development and maturation. Specifically, two novel areas— related to mitochondrial dysfunction-associated POI and non-coding RNA-associated POI— are highlighted to enhance our understanding of POI. This review seek to present current knowledge and hypotheses, identify areas requiring further investigation, and outline future research directions.

Fig. 1
figure 1

Pathogenic model of premature ovarian insufficiency (POI). Under the regulation of intraovarian factors and gonadotropins, primary follicles develop into preantral and early antral follicles, which are particularly susceptible to atresia, or follicle death. Subsequently, they progress to preovulatory follicles, leading to oocyte release and corpora lutea formation. Defects in folliculogenesis—such as a decrease in primordial follicles, an increase in atresia, and altered follicular maturation—cause POI. Diagnostic criteria for POI include a minimum of four months of amenorrhea, elevated follicle-stimulating hormone (FSH) levels exceeding 25 IU/L, and/or reduced estrogen levels persisting for at least four weeks, alongside the absence of antral follicles on ovarian ultrasound examination. The impact of POI on health is multifaceted, with female infertility primarily resulting from oocyte depletion and dysfunctions in follicular development. The heterogeneity of POI is attributed to genetic causes, including karyotype abnormalities, nuclear genes, mitochondrial genes, and non-coding RNAs, as well as non-genetic causes, such as autoimmunity and environmental toxins

The genetic etiology of syndrome-related POI

Chromosome abnormality

X chromosome aneuploidies

Turner Syndrome (TS) is a genetic disorder arising from the complete or partial absence of an X chromosome, typical ly discerned through a 45, X karyotype. It is characterized by POI, hypogonadism, and short stature [15]. Prevalent in approximately 1 in 2500 live births [16], TS is a significant contributor to POI cases, constituting 4–5% of the total [17]. Female individuals with TS undergo follicle loss during early ovarian development, leading to elevated FSH levels in infancy. This elevation can result in primary amenorrhea and ovarian dysplasia. The underlying mechanism involves the deletion of an X chromosome during meiosis or the loss of crucial X-linked genes, such as the Short-stature homeobox(SHOX) gene [18]. Normal ovarian function hinges on the presence of two copies of X-linked genes; the loss of one copy manifests as TS-related phenotypes. Recent research has additionally implicated telomere function, length, and epigenetic modifications in TS-related POI [19].

Trisomy X Syndrome (TXS), identified by a 47, XXX karyotype, is a sex chromosome aneuploidy disorder impacting approximately 1 in 1000 women, with a diagnostic rate of merely 10% [20]. Despite limited research, a 2020 study revealed diminished Anti-Mullerian Hormone (AMH) levels in TXS patients, suggesting an augmented risk of POI [21]. TXS individuals may also exhibit elevated FSH and Luteinizing Hormone (LH), correlating with menstrual cycle disorders and POI. Earlier investigations allude to a potential correlation between TXS and POI [22].

Structural chromosomal abnormality

In the realm of structural rearrangements involving the X chromosome, the isochromosome (46, Xi(X)(q10)) emerges as a prevalent occurrence, often associated with the TS phenotype, indistinguishable from 45, X patients. Deletions and translocations within the long arm of the X chromosome are also reported in cases of POI. Deletions typically exhibit breakpoints in the Xq24–Xq27 region, while translocation breakpoints predominantly occur from Xq13 to Xq21. These observations designate the Xq24–q27 and Xq13.1–q21.33 regions as POI critical regions 1 and 2, respectively [23].

In addition, the correlation between X-autosomal translocation and POI cannot be ignored. Various genes on the X chromosome have been implicated through X-autosomal translocations, encompassing Diaphanous Related Formin 2, Premature Ovarian Failure, 1B, Progesterone receptor membrane component 1, among others. Despite the rarity of such rearrangements (1:30,000), distinct patterns in the localization of X-chromosome breakpoints have been discerned – 80% of the breakpoints fall within the Xq21 cytoband of the POI2 region [24]. The correlation between X-autosomal translocation and POI remains inconclusive, with three prevailing hypotheses: Gene disruption, Meiosis error, and Position effect hypothesis [25].

While evidence substantiates the role of X chromosome abnormalities in POI, recent research in this domain has been limited. Analysis of data from four studies reveals a prevalence of X-chromosomal structural abnormalities and X-autosomal translocations in POI ranging from 4.2 to 12.0% [26]. This underscores the imperative for further exploration into the molecular and genetic mechanisms underpinning these abnormalities and their contribution to POI.

Apart from structural abnormalities in the X chromosome, Autosomal translocations, microdeletions, gene mutations, epistasis, and epigenetic changes linked to autosomal genes also constitute factors in POI. Previous studies have documented 28 cases of autosomal abnormalities associated with POI, encompassing 10 Robertsonian translocations, 10 reverse translocations, 5 chromosome inversions, and 3 autosomal chromosome microdeletions. These abnormalities manifest across diverse races, including Chinese, Thai, and American populations [27, 28].

Gene mutation

Autoimmune disease-related POI

Autoimmune polyendocrine syndrome type 1 (APS-1) arises from mutations in the Autoimmune regulator (AIRE) gene, situated on 21q22.3 [29]. The AIRE gene, responsible for encoding a protein with two zinc fingers, functions as a transcription factor crucial for autoimmune tolerance. Its primary role lies in facilitating the expression of specific autoantigens in thymic stromal cells. Approximately 50 mutations in the AIRE gene have been identified, inherited through autosomal recessive transmission [30]. Notably, around 41% of APS-1 patients experience complications related to POI. The primary association between APS-1 and POI is autoimmune lymphocytic ovarian inflammation [31].

Ataxia-telangiectasia (AT) is a rare autosomal recessive genetic disorder that predominantly affects the nerves, blood vessels, skin, mononuclear macrophage system, and endocrine system, resulting in primary immunodeficiency [32]. The causative gene for AT is the Ataxia Telangiectasia Mutated (ATM) gene, which plays a crucial role in DNA damage repair, cell cycle regulation, and immune response, potentially influencing sexual maturity. The pathology of AT arises from the loss of protein expression due to ATM gene mutation. Clinical manifestations encompass cerebellar ataxia, ocular motility disorders, telangiectasia, immunodeficiency, tumor susceptibility, chromosome instability, and gonadal dysplasia. Female AT patients frequently present with ovarian hypoplasia and disorders in the development of primordial germ cells [33, 34].

Metabolic disease-related POI

Galactosemia is an autosomal recessive hereditary disorder arising from the deficiency of Galactose-1-Phosphate Uridylyltransferase (GALT) gene. This condition can result in severe complications in organs expressing high levels of GALT, including the liver, kidneys, ovaries, and heart [35]. POI manifests in 80-90% of female patients with homologous mutations in the GALT gene. GALT gene mutations can lead to the accumulation of galactose in the ovary, hindering its metabolism and inducing toxic effects, thereby fostering premature follicular atresia. The elevation of FSH in galactosemia patients initiates from birth to early adolescence, with varying onset times for ovarian function impairment [36]. However, primary amenorrhea predominates in most patients. Despite a minimal number of patients achieving natural conception, key biochemical indicators such as AMH, estradiol, and FSH consistently reflect ovarian failure [37].

Carbohydrate-Deficient Glycoprotein Syndrome is a congenital metabolic disorder stemming from impaired glycan binding to other complexes, such as proteins or lipids. The etiology of this syndrome is linked to a mutation in the Phosphomannomutase 2 gene. The Phosphomannomutase gene encodes a phosphomannose mutase responsible for catalyzing the conversion of mannose-6-phosphate to mannose-1-phosphate. A functional defect in this enzyme can disrupt ovarian glycoprotein glycosylation, potentially contributing to POI by affecting ovarian glucose metabolism [38].

Endocrine disease-related POI

Blepharophimosis-ptosis-epicanthus in-versus syndrome (BPES) is an autosomal dominant genetic disorder which manifests in two distinct types, with Type I commonly associated with POI, and its etiology correlated with mutations in the Forkhead transcription factor L2 (FOXL2) [39]. Over 260 FOXL2 gene mutations have been found. These include changes in the polyalanine chain, leading to Type I or Type II BPES. FOXL2 mutations can affect protein-protein or protein-DNA interactions [40]. However, some mutations don’t fully match the ovarian phenotype, especially missense mutations in the Forkhead domain [41]. Therefore, individuals with Type II BPES mutations may still need ovarian function monitoring.

Pseudohypoparathyroidism is a rare disorder, patients may encounter gonadal issues, such as delayed or impaired sexual development, absence of menstrual periods, irregular menstruation, or infertility. Approximately 70-80% of individuals with pseudohypoparathyroidism type 1a have a clearly identifiable genetic cause, predominantly attributed to GNAS Complex Locus (GNAS) gene mutations or methylation changes [42]. The GNAS gene encodes the alpha subunit of the stimulatory heterotrimeric G protein, a critical component of the gonadotropin receptor signaling pathway. Individuals with gonadotropin resistance and POI may harbor mutant maternal GNAS alleles in the gonads, disrupting the gonadotropin receptor signaling pathway [43].

Nervous system disease-related POI

Ovarioleukodystrophy is a severe hereditary neurodegenerative disorder characterized by white matter loss and POI [44]. The condition is associated with mutations in the Eukaryotic initiation factor 2B (EIF2B) gene, a pivotal protein in the initiation of eukaryotic translation. EIF2B comprises five subunits: α, β, γ, δ, and ε. Mutations in any of these subunits can lead to the development of the disease. Individuals with mutations in the EIF2B gene often display progressive degenerative changes in the nervous system, and the severity of ovarian insufficiency is positively correlated with age [45]. However, in some cases, POI may manifest before the onset of neurological symptoms or during subclinical neurological lesions.

Progressive external ophthalmoplegia is characterized by a gradual onset of facial ptosis and restricted eye movement. Its pathogenesis is associated with mutations in the DNA Polymerase Gamma, Catalytic Subunit (POLG) gene, which can manifest as either an autosomal dominant or recessive genetic disorder. This disorder falls within the category of mitochondrial diseases. Mutations in the POLG gene not only result in ocular symptoms but are also frequently accompanied by POI and Parkinson’s disease [46].

Perrault syndrome is a pleiotropic autosomal recessive disorder characterized by ovarian failure in females, sensorineural hearing loss in both genders, and neurological manifestations in select patients. In a European mixed-race family comprising two sisters, researchers detected compound heterozygous variants of HSD17B4 c.1704 A > G and c.17T > A [47]. Subsequent investigations revealed mutations in Histidyl-TRNA Synthetase 2, Mitochondrial (HARS2), Leucyl-TRNA Synthetase 2, Mitochondrial (LARS2), Caseinolytic Mitochondrial Matrix Peptidase Proteolytic Subunit (CLPP), and Twinkle mtDNA Helicase (TWNK) within the Perrault syndrome framework, utilizing similar genomic methodologies such as linkage analysis or NGS, all crucial for maintaining normal mitochondrial function [48].

Other diseases-related POI

Premature aging syndromes, such as Bloom, Werner, and GAPO, are frequently associated with both female POI and male asthenospermia. These syndromes arise from specific gene mutations leading to autosomal recessive inheritance. Bloom syndrome, stemming from BLM RecQ Like Helicase gene mutations affecting DNA helicase, is characterized by short stature, a photosensitive rash, immunodeficiency, an elevated risk of tumors, and hypogonadism [49]; Werner syndrome, a typical human premature aging syndrome, results from WRN RecQ Like Helicase gene mutations. Symptoms encompass skin aging, scleroderma, cataracts, arteriosclerosis, a heightened risk of tumors, and gonadal dysplasia [50]; GAPO syndrome, caused by mutations in the ANTXR Cell Adhesion Molecule 1 gene involved in cell adhesion and migration, exhibits features such as growth retardation, alopecia, impacted teeth, optic nerve atrophy, and ovarian abnormalities in women [51]; Epidemiological research indicates a robust association between POI and premature aging syndromes. This suggests that POI can be considered a form of accelerated ovarian aging, even without obvious syndromic features.

Fragile X syndrome arises from a mutation in the Fragile X Messenger Ribonucleoprotein 1 (FMR1) gene on the X chromosome. The FMR1 gene possesses a variable CGG repeat sequence, with a normal range of 6 to 44. When the repeats increase to 55 to 200, it is referred to as a premutation. Approximately 1 in 250 women carry the FMR1 gene premutation, and 15–24% of them experience POI, known as Fragile X-related ovarian insufficiency (FXPOI). Studies indicate that about 11.5% of familial POI and 3.2% of sporadic POI patients harbor FMR1 gene premutations. Therefore, FMR1 gene premutations represent a primary genetic cause of POI [52]. The mechanism of FXPOI remains unclear, but it is suggested that an increased number of premutant CGG repeats hinder the normal binding of the FMR1 gene product to the ribosome 40 S subunit. This affects the translation of the protein FMRP, leading to reduced FMRP expression and a subsequent rise in transcription factors, resulting in elevated FMR1 gene mRNA levels [52]. FMRP protein is prominently expressed in germ cells and fetal ovaries, regulating the growth of oogonia and the initial ovarian size. Interestingly, FMRP expression is observed in granulosa cells (GCs) of mature follicles but not in primordial and primary follicles. This shift in expression highlights the role of FMRP in regulating follicular maturation. Abnormal FMRP expression directly impacts follicle and egg development. Moreover, the increase in CGG repeats can have a toxic effect on GCs, leading to increased follicular death [53].

Genetic etiology of non-syndromic POI

Numerous genes are intimately associated with non-syndromic POI, implicating diverse biological processes. The involved processes and related genes are as followed. Homologous recombination and DNA damage repair: helicase for meiosis (HFM1), synaptonemal complex central element protein 1 (SYCE1), stromal antigen 3 (STAG3), mutS homolog 4 (MSH4), mutS homolog 5 (MSH5), etc. Follicle development and maturation: growth differentiation factor 9 (GDF9), FOXL2, FSHR, etc. Mitosis: nanos C2HC-type zinc finger 3 (NANOS3). Follicle assembly: folliculogenesis specific basic Helix-Loop-Helix gene A (FIGLA), SF-1. Follicle activation: LIM homeobox 8 (LHX8), neonatal ovarian homologous box (NOBOX), spermatogenesis and oogenesis specific basic Helix-Loop-Helix 1(SOHLH1), and oogenesis specific basic Helix-Loop-Helix 1 (SOHLH2). As well as cell death pathways: FMR1, and eukaryotic translation initiation factor 4E nuclear import factor 1 (EIF4ENIF1) (Table 1). These POI-associated genes exert their influence across all stages of oocyte and follicular development (Fig. 2).

Table 1 List of candidate genes implicated in non-syndromic POI

Meiosis and DNA damage repair factors

An adequate oocyte reserve is imperative for women of reproductive age in ensuring the birth of healthy offspring. Nonetheless, gene mutations involved in meiosis, DNA replication, or DNA repair can result oogenesis abnormality and consequent infertility. Hence, a comprehensive investigation into the genes governing these key processes, is essential for an enhanced comprehension of POI.

Meiosis and DNA replication related-genes

HFM1 plays a pivotal role in homologous chromosome crossover and synapsis completion during meiosis. A specific HFM1 mutation (c.3470G > A) was identified in a Chinese family with POI, with bioinformatics tools suggesting its potential association with mRNA splicing [54]. Hfm1 knockout mice exhibit decreased follicular reserve and fertility. Subsequent investigations suggest that HFM1 may contribute to the pathogenesis of POI by influencing spindle assembly and division [55].

Another key gene is SYCE1 which encodes a constituent of the synaptonemal complex that mediates homologous chromosomes synapsis and promotes crossover formation [56]. In 2014, the first POI-associated SYCE1 variant (c.613 C > T) was reported [57]. Mouse models carrying this variant exhibit infertility and significantly reduce Syce1 translation levels [58]. In POI patients, SYCE1 mutations have been revealed can disrupt the interfaces formed between SYCE1 and synaptonemal complex component six6 opposite strand transcript 1 [59].

STAG3 as another member of the Synaptonemal Complex, facilitates the nuclear localization of meiotic recombination protein REC8 through direct interaction. Meiotic recombination protein REC8-STAG3-containing cohesin regulates chromosomal topology and maintains sister chromatid cohesion [60]. In 2019, whole-exome sequencing of a Caucasian family with POI identified a truncation mutation (c.3052delC) and a missense mutation (c.659T > G), both predicted to significantly affect domain structural integrity [61].

Co-expressed with STAG3, minichromosome maintenance 8 (MCM8) and minichromosome maintenance 9 (MCM9) likely share similar functions. MCM8/9 facilitates normal replication fork progression and protects stalled forks from excessive degradation [62]. In 2016 a patient with heterozygous mutations in both MCM8 and MCM9 was identified [63]. Knockout experiments in mice revealed Mcm8 and Mcm9 are crucial for gametogenesis and genomic stability [64]. Cells with homozygous MCM8 mutations exhibited significantly higher chromosomal breakage compared to controls when exposed to mitomycin C [65]. Recent research indicates that MCM9 mutations reduce double-strand break (DSB) repair functionality [66].

Putative homologs of MCM8 found in other organisms include meiosis-specific with OB domain (MEIOB), a single-strand DNA-binding protein necessary for maintaining the appropriate number of RAD51 recombinase (RAD51) and DNA meiotic recombinase 1 (DMC1) foci after the zygotene stage. MEIOB forms a heterodimer with spermatogenesis associated 22 (SPATA22), interacting with the RPA heterotrimer, the primary complex in DNA metabolism [67]. In 2019, Caburet et al. reported the first POI-associated MEIOB mutation: c.1218G > A [68]. Recently, Wang et al. discovered two novel homozygous frameshift mutations (c.258_259del, c.1072_1073del) and a new homozygous nonsense mutation (c.814 C > T) in MEIOB [69]. These mutations disrupt theMEIOB-SPATA22 interaction [68, 69]. Moreover, Meiob-deficient mouse exhibit meiotic failure and infertility in both sexes [67].

Taken together, genes involved in meiosis and DNA replication related POI include HFM1, SYCE1, STAG3, MCM8, MCM9, MEIOB. HFM1mutations may cause POI by affecting spindle assembly and division [55]. As members of the synaptonemal complex, SYCE1 and STAG3 mutations might lead to POI [57, 61, 70]. Potential links between mutations in other complex components and POI require further exploration. MCM8 and MCM9, as members of the minichromosome maintenance family, regulate cell division, and mutations have been related with POI [63]. MEIOB forms a heterodimer with SPATA22 for proper homologous recombination, and experimental evidence demonstrates that POI-associated MEIOB mutations disrupt this interaction [68, 69]. Another gene such as PSMC3IP, its mutations have found in POI patients, but the pathogenic mechanisms remain unclear.

DNA repair related-genes

Mismatch repair systems, ubiquitous across organisms, are crucial for genomic stability. By rectifying base pair mismatches and small nucleotide indels during DNA replication and recombination, mismatch repair system ensures replication and recombination fidelity. MSH4 and MSH5, as members of the DNA mismatch repair mutS family, encode meiosis-specific proteins. MSH4 and MSH5 heterodimers uniquely bind Holliday Junctions, stabilizing and preserving meiotic DSB repair intermediates [71]. Carlosama et al. identified a homozygous donor splicing site mutation (c.2355 + 1G > A) in MSH4 in Colombian families with POI, predicted to inactivate the highly conserved Walker B motif in the ATP binding domain [72]. A recent study has suggested a correlation between MSH4 variants and reduced oocyte quality by identifying biallelic MSH4 variants in a diminished ovarian reserve patient [73]. For MSH5, Guo et al. reported a homozygous missense mutation (c.1459G > T) in Chinese sisters with POI, confirming its deleterious effects on DNA homologous recombination repair [74]. In 2020, Wang et al. proposed that through YB1, MSH5 and DNA damage repair transcriptionally regulated by the long non-coding RNA HCP5, implicating MSH5 in the etiology of POI [75].

RAD51 and DMC1 play a pivotal role in DSB repair by facilitating the nucleoprotein filament to search for and invade its homologous chromosome [76]. Heat shock transcription factor 2 binding protein (HSF2BP) regulates RAD51/DMC1 localization at meiotic DSB sites through interaction with breast cancer type 2 susceptibility protein [77]. Felipe-Medina et al. reported a candidate missense variant of HSF2BP (c.500 C > T) in three families with POI. Further investigation revealed that 19 open reading frame 57 and MEIOB associated 1 (BRME1) serves as a potent interaction agent and stabilizer of HSF2BP. In meiotic cells carrying this mutation, staining of both HSF2BP and BRME1 reduced, along with fewer RAD51/DMC1 foci, suggesting a potential mechanism by which HSF2BP contributes to POI development [78].

During replication the Shu complex can regulate RAD51 recruitment to DNA repair foci. As a member of Shu complex, SPIDR plays a pivotal role in diverse processes, including homologous recombination regulation and RAD51 foci formation [79]. In 2017, a study reported a biallelic mutation in SPIDR associated with ovarian dysgenesis, altering SPIDR activity in homologous recombination [80]. Female mice lacking Spidr exhibit decreased fertility, while some Spidr−/− oocytes still complete meiosis [81].

As mentioned above, MSH4, MSH5, HSF2BP and SPIDR are involved in DNA damage repair, participating in POI. MSH4 and MSH5 are key genes in DNA damage repair. In 2020, the pathogenic mechanism of POI-associated MSH5 mutations was proposed [75], while the mechanism of POI-related MSH4 mutations requires further study. HSF2BP regulates the localization of RAD51/DMC1 [77], and recent studies suggest that its gene mutations, through interactor chromosome 19 open reading frame 57/BRME1, may lead to POI [78]. SPIDR and another member of Shu complex, zinc finger SWIM type containing 7, both of which have been found associated with POI [80, 82, 83]. Meiosis and DNA damage repair are critical processes during oogenesis, and genetic variations affecting these processes could lead to female reproductive abnormalities. In addition to the aforementioned genes, many other genes that potentially associated with POI, warrant further investigation.

Transcription factor related-factors

Although transcription factors cannot directly participate in many important biological processes, they play a crucial role in regulating body activity by controlling various target genes. In female reproduction, they regulate the expression time, location, and level of reproductive genes to ensure the smooth progress of each reproductive process.

SF-1 and Neonatal Ovarian Homologous Box (NOBOX) gene have been identified as the earliest transcription factor specifically involved in gonadotropin cells [84,85,86]. They play an important role in early ovarian development, including follicle assembly and growth maintaining the primordial follicle (PF) pool and oocyte-specific gene expression during early follicular development [84, 85]. In ovaries with SF-1 deficiency, multiple key follicular related mechanisms are impaired. These damaged pathways lead to increased oocyte death and a significant decrease in ovarian reserve [84]. The absence of SF-1 in the pituitary and GCs before and after ovulation indicates that this nuclear receptor plays an important role in the pituitary gonadal axis [85, 87].

In human gonads, NOBOX is preferentially expressed in primordial follicles and metaphase II oocytes [88]. A study of 96 white women with premature ovarian failure found two mutations c.1064G > A and c.1079G > A in the NOBOX gene that disrupted its DNA binding activity, leading to ovarian dysfunction [88]. Additionally, sumoylation has been shown to have an inhibitory effect on NOBOX transcriptional activation using Gdf9 as the promoter [89]. The function of Spermatogenesis and SOHLH1 and SOHLH2 is to coordinate oocyte differentiation without affecting meiosis and regulate the co expression of oocyte specific transcription regulatory factors. The function of SOHLH1 and SOHLH2 is similar to NOBOX [90]. A study suggests that ubiquitination can directly regulate SOHLH1 or serve as an indirect regulatory factor in SOHLH2 localization [89]. A 2020 study explored FIGLA, LHX8 and SOHLH1, showing that they are jointly involved in a multi-functional network, which plays an important role in regulating the maintenance and differentiation of germ cells during early mouse oogenesis [90].

LHX8 is expressed in germ cells and is crucial for early oogenesis. The global knockout and conditional knockout of LHX8 leads to the primary follicle death and reduced reserve of secondary follicle pools; Another study found that LHX8 ablation leads to DNA damage in oocytes, leading to a significant increase in autophagy and premature depletion of ovarian reserves. A study based on gene load testing identified LOF mutations in LHX8, characterized by delayed oocyte maturation, suggesting haploid dysfunction effects [91].

A study on Chinese patients with premature ovarian failure found three different FIGLA mutations c.11 C > A, c.625G > A and c.84 C > A in four patients. These mutations in FIGLA may affect the transcription of zona pellucida genes and cause ovarian dysfunction [92]. FIGLA is a key gene that influences the balance of other transcription factors. Without FIGLA, the expression of meiotic related genes (such as SYCP3, RAD51, YBX2) and oocyte growth/differentiation genes (like NOBOX, LHX8, TAF4B, SOHLH1, SOHLH2, GDF9) becomes imbalanced. The imbalance significantly affects meiosis, leading to DNA damage and oocyte apoptosis [90].

Transcription factors are pivotal in orchestrating various biological processes by controlling the expression of target genes. In female reproduction, they govern the timing, location, and levels of reproductive genes crucial for the smooth progression of reproductive processes. These findings of SF-1, NOBOX, SOHLH1, SOHLH2 and FIGLA highlight the intricate regulatory network orchestrated by transcription factors in female reproductive health.

Signal pathway

Signal molecules, such as hormones, have the ability to transmit information between cells and regulate cell activity by binding to receptors both inside and outside the cell. Multiple hormones and receptors are involved in many important processes related to female reproduction and fertility, and their expression levels in the human body can serve as indicators of female reproductive status.

FSHR, Anti-Müllerian Hormone Receptor Type II (AMHR2), Bone Morphogenetic Protein Receptor Type 1 A (BMPR1A) and Bone Morphogenetic Protein Receptor Type 1B (BMPR1B) are expressed in ovarian granulosa [93,94,95]. They play a crucial role in normal gonadal function, regulating follicle growth, estrogen production, and oocyte maturation. In patients with primary ovarian insufficiency and partial loss of FSHR function, folliculogenesis can proceed only up to the small antral stages. In addition, the complete inactivation of the FSHR in women may be caused by the absence of follicular recruitment during the reproductive life of these patients due to the inactive FSHR [94]. A 2020 study reported a novel missense mutation c.1268T > C in the FSHR gene in a female with normal puberty but primary amenorrhea, which severely affected cAMP/PKA signaling [96].

AMHR2 is mainly expressed in GCs that grow follicles, and in the cumulus of large antral follicles. AMHR2 mutations can trigger persistent Müllerian disorder syndrome in humans, characterized by the persistence of Müllerian tube derivatives and early failure of ovarian follicle banks. Other residues of AMHR2 are also crucial for AMH signaling [97, 98]. Two new missense AMHR2 variants (c.627T > A and c.1060 C > T) were found in a Chinese POI cohort. Mutations located in the intracellular domain, for instance c.627T > A, which is thought to disrupt the substrate-binding site of the kinase domain, migrate normally to the cell surface but are unable to transduce the AMH signal [99, 100].

BMPR1A or ALK3 is serine threonine kinase type I receptors expressed in many tissues. BMPR1B plays an important role in cumulus cell expansion, ovulation cycle, and skeletal system development. Ovarian specific Bmpr1a−/− mice become infertile within 6 months after reproduction due to the cessation of follicular development. For Bmpr1b−/− mice, the mutant exhibits cumulus expansion damage and endometrial development defects [95, 101].

GDF9 and Bone Morphogenetic Protein 15 (BMP15) play a crucial role in follicle formation and development before the ovarian cycle. They form heterodimers or homodimers to activate primary follicles and participate in follicular development, ultimately mediating germ cell formation and ovulation processes [102, 103]. In 2014, high-resolution array comparative genomic hybridization (CGH) analysis revealed the first POI-related mutation that may affect the GDF9 regulatory region [103]. A study found BMP15 mutations c.791G > A and c.1076 C > T in two sisters with POI that affect the water molecules around the protein and the thermal stability [104]. Scientists also conducted in vitro cell line experiments and found that BMP15 function was impaired. Two other studies also found harmful heterozygous BMP15 variants c.309T > G, c.551T > C, and c.406G > C in POI patients [105, 106].

Linker for Activation of T Cells (LAT) and Vascular Endothelial Growth Factor A (VEGFA) gene mutation may be potential mechanisms for POI development. It is reported that a novel single compound heterozygous pathogenic variant of the LAT gene (c.245 C > T and c.181 C > G) has been discovered in a POI family, and for the first time, it has been reported that LAT, which was initially expressed in T cells, exists in ovarian GCs. These two variants reduce the expression of LAT, inhibit the proliferation of GCs, and promote GCs apoptosis, which may be related to the etiology of POI [107].

A study showed that VEGFA − 1154G > A and 936 C > T variants are linked to the susceptibility of Chinese Han women to primary ovarian insufficiency [108]. VEGFA essentially affects the ovarian folliculogenesis as well as the normal reproductive function. It has been found to lead to the extravasation of protein-rich fluid in ovarian follicles [109]. Hence, VEGFA genetic variants may be capable of modulating its expression and affecting POI.

Signal molecules, including hormones, orchestrate cell activity by binding to receptors, both inside and outside the cell. In female reproduction and fertility, numerous hormones and receptors are pivotal, serving as indicators of reproductive status. These findings of FSHR, AMHR2, BMPR1A, BMPR1B ,GDF9, BMP15, LAT and VEGFA underscore the intricate interplay between signaling molecules and receptors in regulating female reproductive processes and highlight their significance in POI.

RNA metabolism and translation related-factors

RNA is closely related to DNA and proteins. Therefore, damaged RNA activity, including metabolism and translation, has adverse effects on proteins. These proteins may play important roles in various biological processes, such as follicle formation, steroid production, cell proliferation, and apoptosis.

In 2013, a study demonstrated that the c.457 C > T substitution decreases the stability of NANOS3, potentially resulting in a hypomorph. Furthermore, an investigation of the relationship between the number of primordial germ cells (PGCs) and the dosage of NANOS3 in mouse models showed that the population of PGCs is controlled by the level of NANOS3 protein [110]. There are three homologues: NANOS1, NANOS2, and NANOS3. Destroying NANOS1 in mice does not affect germ cell development, but knocking out NANOS2 or NANOS3 can lead to infertility, resulting in a decrease in gonadal size due to the loss of PGCs [26].

In a large French Canadian family, 7 women with POI had heterozygous premature termination codons c.1286 C > A in EIF4ENIF1, which were not present in the unaffected members, indicating a dominant genetic pattern of POI leading to EIF4ENIF1 mutations [111]. In addition, a study found a rare mutation in the EIF4ENIF1 gene in a family with POI. This mutation may lead to mutations around the site α- Damage to the spiral structure or α- Reduction in spiral length. This mutation is the second new EIF4ENIF1 mutation found in POI patients [112]. EIF4ENIF1 is a factor that transports the translation initiation factor eIF4E. It competitively blocks the effective binding of eIF4E and eIF4G, thereby regulating ribosomal delay and reducing protein synthesis by interfering with their interaction [112].

RNA is intricately linked to DNA and proteins, and disruptions in RNA activity, including metabolism and translation, can have detrimental effects on protein function. These proteins are integral to various biological processes such as follicle formation, steroid production, cell proliferation, and apoptosis. These findings NANOS3 and EIF4ENIF1 highlight the critical role of RNA in regulating protein funtion and its implications for POI.

Fig. 2
figure 2

POI causative genes associated with oogenesis and folliculogenesis. The figure illustrates the correlation between pathogenic genes linked to POI and the developmental processes of oocytes and follicles. Initially, primordial germ cells undergo mitosis to differentiate into oogonia. These oogonia then enter meiosis, forming primary oocytes that pause at the diploid stage during early meiosis. Pregranulosa cells surround the oocytes during this phase, creating primordial follicles, which represent the dormant follicular reserve within the ovary. Oocytes lacking the encapsulation of primordial follicles may undergo apoptosis, resulting in their loss. The majority of primordial follicles remain quiescent, with only a minority undergoing activation. Once activated, these follicles undergo continuous growth, eventually either persisting in an atretic state or being released through ovulation in the later stages of follicular development. It is noteworthy that pathogenic genes associated with POI exert their influence across all stages of oocyte and follicular development, extending beyond the aforementioned developmental stages

Mitochondrial dysfunction-associated POI

Several genes associated with POI, whether in isolated instances or as part of syndromes, function within the mitochondria. These genes include Required for Meiotic Nuclear Division 1 homolog Gene (RMND1), Mitochondrial Ribosomal Protein S22 Gene (MRPS22), POLG, DNA Polymerase Gamma 2, Accessory Subunit (POLG2), Transcription Factor A, Mitochondrial (TFAM), Leucine-rich Pentapeptide Repeat Gene (LRPPRC), ClpB Family Mitochondrial Disaggregase (CLPB), CLPP, TWNK, Tu Translation Elongation Factor, Mitochondrial (TUFM), LARS2, Mitochondrial Alanyl-tRNA synthetase 2 (AARS2), and HARS2 (Table 2). Broadly, their functions encompass mtDNA replication, transcription, translation as well as protein degradation (Fig. 3).

mtDNA replication-related factors

The p140 catalytic subunit, PolγA (encoded by POLG), is a crucial component of DNA polymerase gamma (Polγ), responsible for replicating the mitochondrial genome. Mutations in POLG can lead to various mitochondrial diseases, including mitochondrial epilepsy, polyneuropathy, ataxia, and progressive external ophthalmoplegia [113]. Additionally, POI is an important clinical symptom resulting from POLG mutation. Previous studies typically describe POI caused by POLG mutations as part of a syndrome rather than isolated [114, 115]. However, in 2018, Chen et al. identified the first homozygous pathogenic POLG mutation, c.2890 C > T, causing non-syndromic POI through Whole exome sequencing (WES) [116]. Therefore, further studies are needed to establish a clear POLG genotype-phenotype correlation.

The genes encoding the four core components of mtDNA replicators are POLG, POLG2, TWNK, and SSBP1. Studies have confirmed that mutations in POLG, POLG2, and TWNK are associated with POI (Table 2). However, no mutations related to POI have been reported in the SSBP1 gene, highlighting a need for further research in this area.

mtDNA transcription-related factors

Mitochondrial RNA polymerase is primarily responsible for mtDNA transcription, requiring TFAM and mitochondrial transcription factor B2 (TFB2M) to initiate mtDNA transcription. TFAM plays a pivotal role in mtDNA transcription, maintenance, and replication. During transcription initiation, TFAM binds to mtDNA, bending it into a U-turn to activate transcription [117]. Mutations in TFAM are primarily associated with neurodegenerative diseases such as Alzheimer’s disease and Parkinson’s disease [118]. In 2020, Tucker et al. identified the most promising candidate mutation as a homozygous missense mutation, c.694 C > T, in TFAM while studying a patient with POI in Pakistan [119]. Subsequently, in 2021, Farid Ullah et al. confirmed the mutation’s pathogenicity through in vivo complementation studies in zebrafish [120].

Despite TFAM being identified as a strong candidate pathogenic gene for POI, its pathogenic mechanism remains unclear. Additionally, other genes involved in mtDNA transcription, particularly TFB2M, which shares functional similarities with TFAM, have not been reported to be associated with POI. Hence, further research in this area is warranted.

Mitochondrial translation-related factors

Mammalian mitochondrial protein translation is a highly complex process divided into initiation, extension, termination, and circulation stages. It encompasses mitochondrial ribosome assembly, mRNA stability, and aminoacylation of amino acids with corresponding tRNA. Mutations in genes governing these processes impact translation, potentially causing various diseases [121]. Below, we discuss some gene mutations related to POI within this translation process.

The spectrum of diseases associated with MRPS22 mutations ranges from isolated POI to severe mitochondrial disorders in infants [122]. MRPS22, a crucial component of the mitochondrial ribosomal 28 S small subunit, plays a key role in translating mtDNA-encoded peptides and maintaining mitochondrial ribosome complex stability [123]. A study by Chen et al. identified homozygous missense variants (c.404G > A and c.605G > A) in independent families and confirmed MRPS22 mutations as an autosomal recessive cause of POI [124]. Isolated POI cases with MRPS22 mutations do not typically exhibit other neurological or syndromic features, and the loss of MRPS22 does not affect oxidative phosphorylation (OXPHOS), suggesting a non-energy-related mitochondrial function in POI [125].

LRPPRC, an RNA-binding protein, collaborating with SLIRP, it enhances mRNA structure, accuracy of translation, and stability, with distinct roles in mRNA polyadenylation and ribosomal guidance [126]. LRPPRC also interacts with eIF4e in the nucleus, impacting nuclear export and translation [127]. Mutations in LRPPRC lead to a rare disease known as Leigh syndrome, prevalent in the Saguenay-Lac-Saint-Jean region of Quebec and colloquially referred to as the French Canadian syndrome. Individuals with Leigh syndrome face a heightened risk of death due to neurological or acidosis crises triggered by factors such as infection, stress, diet, disease, or exercise. The average lifespan is less than 5 years, though some individuals may live longer. Female survivors may experience POI after puberty, characterized by elevated FSH levels, small ovaries with few follicles, and absence of menstrual periods [122]. Given LRPPRC’s impact on both mitochondrial and nuclear gene expression, further investigation is essential when considering its role in mitochondrial dysfunction-related POI.

In addition to MRPS22 and LRPPRC, mutations in other genes that play a role in mitochondrial translation, including RMND1, HARS2, LARS2, AARS2, and TUFM, have also been shown to be associated with POI (Table 2)., emphasizing the importance of mitochondrial translation process for ovarian function. However, several unresolved issues persist: (i) the specific molecular mechanism underlying POI due to these gene mutations remains unclear; (ii) MRPS22 mutations have been reported in both isolated and symptomatic POI cases, establishing the genotype-phenotype relationship of MRPS22 is a key challenge; (iii) LRPPRC impacts both mitochondrial and nuclear gene expression. Whether its mutation causes POI by causing mitochondrial dysfunction remains to be confirmed.

Protein degradation-related factors

Mammalian mitochondrial protein translation is a highly intricate process, categorized into four stages. One crucial protein is ClpB, also known as Suppressor of potassium transport defect 3 (SKD3), encoded by the CLPB gene and belonging to the AAA + family of ATPases. Located within mitochondria, ClpB plays a pivotal role in various activities, particularly in protein folding and the degradation of misfolded proteins. Despite its significance, its role in ovarian function remains unclear [128]. A study in 2022 revealed that female carriers of dominant ClpB pathogenic variants often experienced POI after puberty [129]. Another essential protein, ClpP, shares a similar function in degrading damaged or misfolded proteins. Encoded by the CLPP gene in humans, ClpP is a crucial component in forming endopeptidase Clp protein complexes [130]. In a study in 2013, recessive mutations of CLPP were identified in patients with Perrault syndrome, correlated with POI and sensorineural hearing loss, along with growth retardation. The disease’s pathological causes may involve insufficient clearance of mitochondrial components and inflammatory tissue damage [131]. Additionally, a novel variant of CLPP c.628G > A was discovered in a Chinese POI patient in 2023. Researchers also confirmed that CLPP deficiency impaired oxidative respiratory chain IV function by affecting COX5A, leading to reactive oxygen species (ROS) accumulation. This, in turn, is speculated to be the potential mechanism behind CLPP abnormalities accelerating ovarian follicular atresia, ultimately resulting in POI [132]. However, this hypothesis requires further investigation for validation.

In conclusion, the ClpB protein encoded by the CLPB gene and the ClpP protein encoded by the CLPP gene both contribute to protein degradation. Mutations in both the CLPB and CLPP genes have been linked to POI, although the exact pathogenic mechanism of CLPB mutations remains unclear. CLPP mutation may lead to POI by damaging the function of oxidative respiratory chain IV, leading to ROS accumulation, activating the intrinsic apoptotic pathway, and ultimately affecting the apoptosis of GCs. However, the hypothesis needs further verification.

Fig. 3
figure 3

Schematic diagram of mitochondrial genes associated with POI. This figure illustrates the diverse molecular functions of proteins within mitochondria, which have been thoroughly characterized, establishing that mutations in these genes can lead to POI. The DNA polymerase γ (Polγ) consists of PolγA, encoded by the POLG gene, and PolγB, encoded by the POLG2 gene. Additionally, TWNK encodes a mitochondrial helicase crucial for mtDNA replication processes. Together, they form the core machinery for mitochondrial DNA (mtDNA) replication. During transcription initiation, TFAM binds to mtDNA, activating transcription. MRPS22 is essential for the translation of mtDNA-encoded peptides and maintaining mitochondrial ribosomal complex stability. LRPPRC, an RNA-binding protein, enhances mRNA structure, translation accuracy, and stability, impacting the expression of both mitochondrial and nuclear genes. RMND1 contributes to mitochondrial translation, particularly in the synthesis of mtDNA-encoded peptides. The genes HARS2, LARS2, and AARS2 encode aminoacyl-tRNA synthetases (aaRS), pivotal for translating mitochondrial-encoded genes. TUFM, a nuclear-encoded mitochondrial translation elongation factor, is crucial for mitochondrial protein synthesis. Moreover, CLpP and CLpB are involved in degrading damaged or misfolded mitochondrial proteins, maintaining mitochondrial health and function

ncRNAs abnormalities in POI

Up to now, the pathogenic genes of POI have been extensively and deeply studied in the coding region. However, it is estimated that human protein coding sequences account for only 1.5% of the entire genome. ncRNA serves as an epigenetic marker regulating gene expression through interactions with DNA or RNA sequences and proteins. The primary categories of ncRNA include miRNAs and lncRNAs, both playing crucial roles in a diverse array of physiological and pathological processes. In recent years, more and more studies have shown that they are an important part of the genetic etiology of POI.

Role of miRNAs in POI

MicroRNAs (miRNAs) have been implicated in an array of reproductive disorders, including POI, through the regulation of fertility-related genes. Despite this, the exact pathogenic mechanisms underlying their involvement remain poorly understood. Recent studies have shed light on the role of aberrantly expressed miRNAs in influencing GCs by modulating apoptosis and proliferation via specific signaling pathways.

Zhang and colleagues performed a comparative miRNA profiling analysis across multiple mammalian species, which highlighted the critical roles played by certain miRNA families and clusters in GCs apoptosis and follicular atresia. These include the let-7 family, miR-23-27-24 cluster, miR-183-96-182 cluster, and miR-17-92 cluster along with their corresponding pathways [133]. In findings from both Zhang et al. and Dang et al., it was reported that miR-127-5p and miR-379-5p exhibit significantly elevated expression levels in women with biochemically diagnosed POI. Overexpression of these miRNAs is correlated with suppressed proliferation of mouse GCs and a decline in DNA repair mechanisms, achieved by targeting key genes such as high mobility group box 2 (Hmgb2), poly ADP-ribose polymerase 1 (Parp1), and X-ray repair cross-complementing 6 (Xrcc6) [134, 135]. Separately, Chen et al. conducted a study where they identified a set of differentially expressed miRNAs in the plasma of POI patients. This set comprised ten upregulated miRNAs (miR-202, miR-146a, miR-125b-2*, miR-139-3p, miR-654-5p, miR-27a, miR-765, miR-23a, miR-342-3p, and miR-126) and two downregulated miRNAs (let-7c and miR-144). Notably, miR-146a was found to facilitate GCs apoptosis by directly targeting interleukin-1 receptor-associated kinase 1 (IRAK1) and tumor necrosis factor receptor-associated factor 6 (TRAF6), thus activating the caspase cascade pathway [136]. In another investigation by Yang et al., miR-23a was shown to induce apoptosis in human GCs by suppressing the expression of the anti-apoptotic protein XIAP and upregulating the cleaved, pro-apoptotic form of caspase-3 [137]. Furthermore, miR-23a also reduced the expression of SIRT1 and augmented apoptosis through the extracellular signal-regulated kinase 1/2 (ERK1/2) pathway [138].

These findings collectively suggest that miRNAs may serve as potential biomarkers and therapeutic targets for understanding and treating reproductive disorders like POI, given their regulatory influence on GCs function via modulation of apoptosis and proliferation pathways.

Role of lncRNAs in POI

Two distinct mechanisms have been proposed to elucidate the role of long non-coding RNAs (lncRNAs) in POI. Firstly, the interaction between lncRNAs and miRNAs may contribute to the induction of apoptosis in ovarian GCs. For instance, Duan et al.‘s study revealed that the lncRNA DDGC stabilizes RAD51 mRNA through competitive binding to miR-589-5p, thus preventing ubiquitin-mediated degradation of the WT1 protein via interaction with heat shock protein 90 (HSP90),this mechanism impacts the DNA damage repair capability and differentiation of GCs [139]. Luo et al.‘s study demonstrated that the lncRNA HOTAIR acts as a molecular sponge for miR-148b-3p, relieving the inhibition of ATG14 and promoting autophagy activity and cell proliferation in GCs. Under normal physiological conditions, HOTAIR maintains appropriate levels of autophagy and GCs proliferation by regulating the miR-148b-3p/ATG14 axis. However, in the case of POI, down-regulation of HOTAIR may lead to autophagy disorders and impaired GCs function [140]. Another study by Zhang et al. observed a decreased expression of translation regulatory long non-coding RNA 1 (TRERNA1) in POI patients, suggesting its potential to sponge miR-23a and suppress GCs apoptosis [141].

Secondly, certain lncRNAs are implicated in regulating critical proteins and signaling pathways. For example, Wang et al.‘s study revealed that the lncRNA HCP5 plays a crucial role in stabilizing the binding of YB1 to the promoter region of the MSH5 gene, consequently enhancing the transcription and expression of MSH5. Additionally, HCP5 acts as an RNA scaffold, facilitating the interaction between YB1 and ILF2, which promotes the transport of YB1 into the nucleus. This timely regulation effectively modulates the function of MSH5 at DNA damage sites. Notably, the down-regulation of HCP5 expression in follicular GCs of POI patients impairs DNA damage repair mechanisms and increases GCs apoptosis [75]. Duan et al.‘s study found that lncRNA ZNF674-AS1 was down-regulated in follicular GCs of patients with biochemical ovarian insufficiency and was associated with serum levels of clinical ovarian reserve indicators. Further studies revealed that decreased expression of ZNF674-AS1 reduces aldolase A (ALDOA) enzyme activity and promotes the interaction between ALDOA and the proton pump (v-ATPase), thereby activating lysosomal-located adenylate-activated protein kinase (AMPK) and regulating glycolysis and GCs proliferation [142]. Another study by Duan et al. showed that lncRNA GCAT1 regulates CDKN1B (p27) translation by competitively binding to polypyrimidine tract-binding protein 1 (PTBP1) in GCs. In POI patients, down-regulation of GCAT1 inhibits G1/S cell cycle progression in GCs, thereby inhibiting GCs proliferation [143].

In conclusion, the above studies highlight the intricate roles of various lncRNAs in regulating key processes such as apoptosis, autophagy, DNA damage repair, and cell cycle progression in GCs, shedding light on underlying molecular mechanisms of POI.

Conclusions, discussions, and perspectives

Currently, research on POI has identified numerous relevant genetic factors. This literature review synthesizes genetic factors and potential mechanisms related to POI across four dimensions: syndrome-associated POI, non-syndrome-associated POI (Table 1), mitochondrial dysfunction-associated POI (Table 2), and non-coding RNA-associated POI. The discussion delves into the intricate relationship between these genes and ovarian development, elucidating the functional consequences of diverse mutations to underscore the fundamental impact of genetic effects on POI. These insights serve to consolidate and enhance our comprehension of POI’s etiology, providing a theoretical basis for the diagnosis and treatment of POI patients and exploration of the disease mechanism.

Table 2 List of candidate genes implicated in mitochondrial dysfunction-associated POI

However, several challenges persist in genetic research on POI patients. Firstly, the detailed mechanisms of action of many pathogenic genes and the specific interactions between different genes in the context of POI are not yet fully elucidated. Moreover, the specific mechanisms of pathways related to meiosis, such as DNA damage repair pathways, and their role in the development of POI remain unclear. It is also uncertain whether other genes involved in meiotic pathways can contribute to the development of POI. Transcription factor genes are known to play critical roles in early ovarian development, with their balance regulated by FIGLA. However, the precise interactions among these genes are not completely understood. The binding of signaling molecules to receptors is crucial for guiding follicle development. For instance, genes like BMP15 and GDF9 exhibit moderate expression levels in primary and early follicular populations, but their expression dramatically increases in mid-follicles. Nevertheless, the exact mechanism of interaction between BMP15 and GDF9 remains elusive [102].

Secondly, findings from animal experiments may not directly translate to humans. The differences in the gonadotropin-dependent stage between humans and rodents suggest that results obtained from animal models may not be directly applicable to humans due to genetic and physiological differences [94]. Furthermore, mouse models may not fully replicate the intricate interactions between molecules, cells, organs, organisms, and the environment.

There are several unresolved issues in current research on mitochondria and POI. For example, while certain gene mutations can lead to POI as mentioned in the text, mutations in other mitochondria genes with similar functions do not seem to have the same effect. The underlying reasons for this discrepancy remain unclear. Furthermore, the mechanism by which mutations in mitochondria genes can lead to a spectrum of diseases, ranging from isolated POI to multi-organ syndrome diseases, is not well understood. Additionally, there is a question of whether we can predict genotype-phenotype correlations in these cases.

Lastly, due to the heterogeneity of POI and the various modes of mutation transmission, the prevalence and types of genetic mutations vary significantly across different regions and races. Therefore, future genetic research should consider population stratification to accurately analyze genetic changes. It is crucial to involve different ethnic groups and larger sample sizes in future studies. Clinically, caution should be exercised when applying data from different racial groups in consultations.

In conclusion, genetic factors primarily affect ovarian development and follicle formation in POI, but the specific mechanisms of many gene variations and their effects are still unclear. Future research could focus on investigating ovarian development, follicle formation processes, and the functions of various genes at different stages of follicle development. Additionally, the association between POI and mitochondria, as well as non-coding RNAs, is a promising research area that will require further attention. These efforts aim to enhance clinical understanding and diagnostic approaches for POI.

Data availability

No datasets were generated or analysed during the current study.

Abbreviations

POI:

Premature Ovarian Insufficiency

FSH:

Follicle-stimulating hormone

GCs:

Germ cells

PGCs:

Primordial germ cells

ncRNAs:

Non-coding RNAs

HFM1:

Helicase for meiosis

SYCE1:

Synaptonemal complex central element protein 1

STAG3:

Stromal antigen3

MCM8:

Minichromosome maintenance 8

MCM9:

Minichromosome maintenance 9

DSB:

Double-strand break

MEIOB:

Meiosis-specific with OB domain

RAD51:

RAD51 recombinase

DMC1:

DNA meiotic recombinase 1

SPATA22:

Spermatogenesis associated 22

PSMC3IP:

PSMC3 interaction protein

MSH4:

MutS homolog 4

MSH5:

MutS homolog 5

HSF2BP:

Heat shock transcription factor 2 binding protein

BRME1:

MEIOB associated 1

ZSWIM7:

Zinc finger SWIM-type containing 7

SPIDR:

Scaffold protein involved in DNA repair

NOBOX:

Neonatal Ovarian Homologous Box

SOHLH1:

Spermatogenesis and Oogenesis Specific Basic Helix-Loop-Helix 1

SOHLH2:

Spermatogenesis and Oogenesis Specific Basic Helix-Loop-Helix 2

FIGLA:

Folliculogenesis Specific Basic Helix-Loop-Helix Gene A

LHX8:

LIM homeobox 8

AMHR2:

Anti-Müllerian Hormone Receptor Type II

BMPR1A:

Bone Morphogenetic Protein Receptor Type 1 A

BMPR1B:

Bone Morphogenetic Protein Receptor Type 1B

GDF9:

Growth Differentiation Factor 9

BMP15:

Bone Morphogenetic Protein 15

LAT:

Linker for Activation of T Cells

VEGFA:

Vascular Endothelial Growth Factor A

TCR:

T cell antigen receptor

NANOS3:

Nanos C2HC-type zinc finger 3

EIF4ENIF1:

Eukaryotic translation initiation factor 4E nuclear import factor 1

TS:

Turner Syndrome

TXS:

Trisomy X Syndrome

SHOX:

Short stature homeobox

AMH:

Anti-Mullerian Hormone

LH:

Luteinizing Hormone

APS-1:

Autoimmune polyendocrine syndrome type 1

AIRE:

Autoimmune regulator

AT:

Ataxia-telangiectasia

ATM:

Ataxia Telangiectasia Mutated

GALT:

Galactose-1-Phosphate Uridylyltransferase

BPES:

Blepharophimosis-ptosis-epicanthus in-versus syndrome

FOXL2:

Forkhead transcription factor L2

SF-1:

Steroidogenic factor 1

GNAS:

GNAS Complex Locus

EIF2B:

Eukaryotic initiation factor 2B

POLG:

DNA Polymerase Gamma, Catalytic Subunit

POLG2:

DNA Polymerase Gamma 2, Accessory Subunit

HARS2:

Histidyl-TRNA Synthetase 2, Mitochondrial

LARS2:

Leucyl-TRNA Synthetase 2, Mitochondrial

CLPP:

Caseinolytic Mitochondrial Matrix Peptidase Proteolytic Subunit

TWNK:

Twinkle mtDNA Helicase

FMR1:

Fragile X Messenger Ribonucleoprotein 1

FXPOI:

Fragile X-related ovarian insufficiency

RMND1:

Required for Meiotic Nuclear Division 1 homolog

MRPS22:

Mitochondrial Ribosomal Protein S22

TFAM:

Transcription Factor A, Mitochondrial

CLPB:

ClpB Family Mitochondrial Disaggregase

TUFM:

Tu Translation Elongation Factor, Mitochondrial

AARS2:

Mitochondrial Alanyl-tRNA synthetase 2

WES:

Whole exome sequencing

aaRSs:

Aminoacyl-tRNA synthetases

References

  1. Vander Borght M, Wyns C. Fertility and infertility: definition and epidemiology. Clin Biochem. 2018;62:2–10.

    Article  PubMed  Google Scholar 

  2. Fertility evaluation of. Infertile women: a committee opinion. Fertil Steril. 2021;116(5):1255–65.

    Article  Google Scholar 

  3. Carson SA, Kallen AN. Diagnosis and management of infertility: a review. JAMA. 2021;326(1):65–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Farquhar CM, Bhattacharya S, Repping S, Mastenbroek S, Kamath MS, Marjoribanks J, et al. Female subfertility. Nat Reviews Disease Primers. 2019;5(1):7.

    Article  PubMed  Google Scholar 

  5. Ishizuka B. Current understanding of the etiology, Symptomatology, and Treatment options in premature ovarian insufficiency (POI). Front Endocrinol. 2021;12:626924.

    Article  Google Scholar 

  6. Stuenkel CA, Gompel A, Davis SR, Pinkerton JV, Lumsden MA, Santen RJ. Approach to the patient with New-Onset secondary amenorrhea: is this primary ovarian insufficiency? J Clin Endocrinol Metab. 2022;107(3):825–35.

    Article  PubMed  Google Scholar 

  7. Jankowska K. Premature ovarian failure. Przeglad Menopauzalny = Menopause Rev. 2017;16(2):51–6.

    Article  Google Scholar 

  8. Huhtaniemi I, Hovatta O, La Marca A, Livera G, Monniaux D, Persani L, et al. Advances in the Molecular Pathophysiology, Genetics, and treatment of primary ovarian insufficiency. Trends Endocrinol Metab. 2018;29(6):400–19.

    Article  CAS  PubMed  Google Scholar 

  9. Ke H, Tang S, Guo T, Hou D, Jiao X, Li S, et al. Landscape of pathogenic mutations in premature ovarian insufficiency. Nat Med. 2023;29(2):483–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Akhade VS, Pal D, Kanduri C, Long Noncoding RNA. Genome Organization and mechanism of action. Adv Exp Med Biol. 2017;1008:47–74.

    Article  CAS  PubMed  Google Scholar 

  11. Zhang J, Chen S, Liu K. Structural insights into piRNA biogenesis. Biochim et Biophys acta Gene Regul Mech. 2022;1865(2):194799.

    Article  CAS  Google Scholar 

  12. Alshaer W, Zureigat H, Al Karaki A, Al-Kadash A, Gharaibeh L, Hatmal MM, et al. siRNA: mechanism of action, challenges, and therapeutic approaches. Eur J Pharmacol. 2021;905:174178.

    Article  CAS  PubMed  Google Scholar 

  13. St Laurent G, Wahlestedt C, Kapranov P. The Landscape of long noncoding RNA classification. Trends Genet. 2015;31(5):239–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Katari S, Aarabi M, Kintigh A, Mann S, Yatsenko SA, Sanfilippo JS, et al. Chromosomal instability in women with primary ovarian insufficiency. Hum Reprod (Oxford England). 2018;33(3):531–8.

    Article  CAS  Google Scholar 

  15. Dai HL, Zhou X, Guang XF. Turner syndrome. QJM: Monthly J Association Physicians. 2023;116(2):136–7.

    Article  Google Scholar 

  16. Huang AC, Olson SB, Maslen CL. A review of recent developments in Turner Syndrome Research. J Cardiovasc Dev Disease. 2021;8(11).

  17. Mastellari E, La Marca A. Genetic conditions impairing female fertility. Panminerva Med. 2020;62(4):260–7.

    PubMed  Google Scholar 

  18. Viuff M, Gravholt CH. Turner Syndrome and Fertility. Ann Endocrinol. 2022;83(4):244–9.

    Article  Google Scholar 

  19. Mikheev RK, Grigoryan OR, Pankratova MS, Andreeva EN, Sheremetyeva EV, Absatarova YS, et al. [Telomere pathology in ontogenesis in patients with Turner syndrome]. Problemy Endokrinologii. 2022;68(2):128–32.

    Article  CAS  PubMed  Google Scholar 

  20. Berglund A, Viuff MH, Skakkebæk A, Chang S, Stochholm K, Gravholt CH. Changes in the cohort composition of turner syndrome and severe non-diagnosis of Klinefelter, 47,XXX and 47,XYY syndrome: a nationwide cohort study. Orphanet J Rare Dis. 2019;14(1):16.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Davis SM, Soares K, Howell S, Cree-Green M, Buyers E, Johnson J et al. Diminished Ovarian Reserve in Girls and Adolescents with Trisomy X Syndrome. Reproductive sciences (Thousand Oaks, Calif). 2020;27(11):1985-91.

  22. Rafique M, AlObaid S, Al-Jaroudi D. 47, XXX syndrome with infertility, premature ovarian insufficiency, and streak ovaries. Clin case Rep. 2019;7(6):1238–41.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Di-Battista A, Moysés-Oliveira M, Melaragno MI. Genetics of premature ovarian insufficiency and the association with X-autosome translocations. Reprod (Cambridge England). 2020;160(4):R55–64.

    Article  CAS  Google Scholar 

  24. Di-Battista A, Favilla BP, Zamariolli M, Nunes N, Defelicibus A, Armelin-Correa L, et al. Premature ovarian insufficiency is associated with global alterations in the regulatory landscape and gene expression in balanced X-autosome translocations. Epigenetics Chromatin. 2023;16(1):19.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Di-Battista A, Moysés-Oliveira M, Melaragno MI. Genetics of premature ovarian insufficiency and the association with X-autosome translocations %J Reproduction. 2020;160(4):R55–64.

  26. Qin Y, Jiao X, Simpson JL, Chen ZJ. Genetics of primary ovarian insufficiency: new developments and opportunities. Hum Reprod Update. 2015;21(6):787–808.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Yuan K, He M, Fang Y, Zhu J, Liang L, Wang C. Autosomal chromosome microdeletions in three adolescent girls with premature ovarian insufficiency: a case report. Turk J Pediatr. 2022;64(4):729–35.

    Article  PubMed  Google Scholar 

  28. Chen M, Jiang H, Zhang C. Selected genetic factors Associated with primary ovarian insufficiency. Int J Mol Sci. 2023;24(5).

  29. Besnard M, Padonou F, Provin N, Giraud M, Guillonneau C. AIRE deficiency, from preclinical models to human APECED disease. Dis Models Mech. 2021;14(2).

  30. Ferré EMN, Schmitt MM, Lionakis MS. Autoimmune polyendocrinopathy-candidiasis-ectodermal dystrophy. Front Pead. 2021;9:723532.

    Article  Google Scholar 

  31. Szeliga A, Calik-Ksepka A, Maciejewska-Jeske M, Grymowicz M, Smolarczyk K, Kostrzak A et al. Autoimmune diseases in patients with premature ovarian insufficiency-our current state of knowledge. Int J Mol Sci. 2021;22(5).

  32. Aguado J, Gómez-Inclán C, Leeson HC, Lavin MF, Shiloh Y, Wolvetang EJ. The hallmarks of aging in Ataxia-Telangiectasia. Ageing Res Rev. 2022;79:101653.

    Article  PubMed  Google Scholar 

  33. Subramanian GN, Yeo AJ, Gatei MH, Coman DJ, Lavin MF. Metabolic stress and mitochondrial dysfunction in Ataxia-Telangiectasia. Antioxid (Basel Switzerland). 2022;11(4).

  34. Kuhn K, Lederman HM, McGrath-Morrow SA. Ataxia-telangiectasia clinical trial landscape and the obstacles to overcome. Expert Opin Investig Drugs. 2023;32(8):693–704.

    Article  CAS  PubMed  Google Scholar 

  35. Teixeira LF, Prauchner GRK, Gusso D, Wyse ATS. Classical Hereditary galactosemia: findings in patients and animal models. Metab Brain Dis. 2024;39(1):239–48.

    Article  CAS  PubMed  Google Scholar 

  36. Succoio M, Sacchettini R, Rossi A, Parenti G, Ruoppolo M. Galactosemia: Biochemistry, Molecular Genetics, Newborn Screening, and treatment. Biomolecules. 2022;12(7).

  37. Kruszewska J, Laudy-Wiaderny H, Krzywdzinska S, Grymowicz M, Smolarczyk R, Meczekalski B. Two consecutive pregnancies in a patient with premature ovarian insufficiency in the course of classic galactosemia and a review of the literature. Gynecol Endocrinology: Official J Int Soc Gynecol Endocrinol. 2022;38(2):186–9.

    Article  Google Scholar 

  38. Masunaga Y, Mochizuki M, Kadoya M, Wada Y, Okamoto N, Fukami M, et al. Primary ovarian insufficiency in a female with phosphomannomutase-2 gene (PMM2) mutations for congenital disorder of glycosylation. Endocr J. 2021;68(5):605–11.

    Article  CAS  PubMed  Google Scholar 

  39. Li F, Chen H, Wang Y, Yang J, Zhou Y, Song X, et al. Functional studies of Novel FOXL2 variants in Chinese families with blepharophimosis-ptosis-epicanthus Inversus Syndrome. Front Genet. 2021;12:616112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Méjécase C, Nigam C, Moosajee M, Bladen JC. The genetic and clinical features of FOXL2-Related blepharophimosis, ptosis and Epicanthus Inversus Syndrome. Genes. 2021;12(3).

  41. Xue M, Zheng J, Zhou Q, Hejtmancik JF, Wang Y, Li S. Novel FOXL2 mutations in two Chinese families with blepharophimosis-ptosis-epicanthus inversus syndrome. BMC Med Genet. 2015;16:73.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Singla M, Garg G, Gupta A. Pseudohypoparathyroidism type 1a. QJM: Monthly J Association Physicians. 2018;111(5):331–3.

    Article  CAS  Google Scholar 

  43. Jüppner H. Molecular definition of pseudohypoparathyroidism variants. J Clin Endocrinol Metab. 2021;106(6):1541–52.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Ibitoye RT, Renowden SA, Faulkner HJ, Scolding NJ, Rice CM. Ovarioleukodystrophy due to EIF2B5 mutations. Pract Neurol. 2016;16(6):496–9.

    Article  CAS  PubMed  Google Scholar 

  45. Rodríguez-Palmero A, Schlüter A, Verdura E, Ruiz M, Martínez JJ, Gourlaouen I, et al. A novel hypomorphic splice variant in EIF2B5 gene is associated with mild ovarioleukodystrophy. Ann Clin Transl Neurol. 2020;7(9):1574–9.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Hirano M, Pitceathly RDS. Progressive external ophthalmoplegia. Handb Clin Neurol. 2023;194:9–21.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Morino H, Pierce SB, Matsuda Y, Walsh T, Ohsawa R, Newby M, et al. Mutations in Twinkle primase-helicase cause Perrault syndrome with neurologic features. Neurology. 2014;83(22):2054–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lerat J, Jonard L, Loundon N, Christin-Maitre S, Lacombe D, Goizet C, et al. An application of NGS for Molecular investigations in Perrault Syndrome: study of 14 families and review of the literature. Hum Mutat. 2016;37(12):1354–62.

    Article  CAS  PubMed  Google Scholar 

  49. Gönenc II, Elcioglu NH, Martinez Grijalva C, Aras S, Großmann N, Praulich I, et al. Phenotypic spectrum of BLM- and RMI1-related Bloom syndrome. Clin Genet. 2022;101(5–6):559–64.

    Article  PubMed  Google Scholar 

  50. Tsuge K, Shimamoto A. Research on Werner Syndrome: Trends from Past to Present and Future prospects. Genes. 2022;13(10).

  51. Troxell TN, Piccinin MA, Smith CM, Parsons ME, Drew GS. GAPO syndrome: a rare genodermatosis presenting with unique features. Int J Dermatol. 2018;57(6):727–8.

    Article  PubMed  Google Scholar 

  52. Fink DA, Nelson LM, Pyeritz R, Johnson J, Sherman SL, Cohen Y, et al. Fragile X Associated Primary ovarian insufficiency (FXPOI): Case Report and Literature Review. Front Genet. 2018;9:529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Tang R, Yu Q. The significance of FMR1 CGG repeats in Chinese women with premature ovarian insufficiency and diminished ovarian reserve. Reproductive biology and endocrinology: RB&E; 2020;18(1):82.

    Google Scholar 

  54. Zhe J, Chen S, Chen X, Liu Y, Li Y, Zhou X, et al. A novel heterozygous splice-altering mutation in HFM1 may be a cause of premature ovarian insufficiency. J Ovarian Res. 2019;12(1):61.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Wang H, Zhong C, Yang R, Yin Y, Tan R, Gao L, et al. Hfm1 participates in golgi-associated spindle assembly and division in mouse oocyte meiosis. Cell Death Dis. 2020;11(6):490.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Bolcun-Filas E, Hall E, Speed R, Taggart M, Grey C, de Massy B, et al. Mutation of the mouse Syce1 gene disrupts synapsis and suggests a link between synaptonemal complex structural components and DNA repair. PLoS Genet. 2009;5(2):e1000393.

    Article  PubMed  PubMed Central  Google Scholar 

  57. 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(10):E2129–32.

    Article  PubMed  Google Scholar 

  58. Hernandez-Lopez D, Geisinger A, Trovero MF, Santinaque FF, Brauer M, Folle GA, et al. Familial primary ovarian insufficiency associated with an SYCE1 point mutation: defective meiosis elucidated in humanized mice. Mol Hum Reprod. 2020;26(7):485–97.

    Article  PubMed  Google Scholar 

  59. Sánchez-Sáez F, Gómez HL, 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(36).

  60. Choi EH, Yoon S, Koh YE, Hong TK, Do JT, Lee BK, et al. Meiosis-specific cohesin complexes display essential and distinct roles in mitotic embryonic stem cell chromosomes. Genome Biol. 2022;23(1):70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Heddar A, Dessen P, Flatters D, Misrahi M. Novel STAG3 mutations in a caucasian family with primary ovarian insufficiency. Mol Genet Genomics: MGG. 2019;294(6):1527–34.

    Article  CAS  PubMed  Google Scholar 

  62. Griffin WC, McKinzey DR, Klinzing KN, Baratam R, Eliyapura A, Trakselis MA. A multi-functional role for the MCM8/9 helicase complex in maintaining fork integrity during replication stress. Nat Commun. 2022;13(1):5090.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Desai S, Wood-Trageser M, Matic J, Chipkin J, Jiang H, Bachelot A, et al. MCM8 and MCM9 nucleotide variants in women with primary ovarian insufficiency. J Clin Endocrinol Metab. 2017;102(2):576–82.

    PubMed  Google Scholar 

  64. 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(4):523–34.

    Article  CAS  PubMed  Google Scholar 

  65. 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(4):694–702.

    Article  CAS  PubMed  Google Scholar 

  66. Guo T, Zheng Y, Li G, Zhao S, Ma J, Qin Y. Novel pathogenic mutations in minichromosome maintenance complex component 9 (MCM9) responsible for premature ovarian insufficiency. Fertil Steril. 2020;113(4):845–52.

    Article  CAS  PubMed  Google Scholar 

  67. Luo M, Yang F, Leu NA, Landaiche J, Handel MA, Benavente R, et al. MEIOB exhibits single-stranded DNA-binding and exonuclease activities and is essential for meiotic recombination. Nat Commun. 2013;4:2788.

    Article  PubMed  Google Scholar 

  68. Caburet S, Todeschini AL, Petrillo C, Martini E, Farran ND, Legois B, et al. A truncating MEIOB mutation responsible for familial primary ovarian insufficiency abolishes its interaction with its partner SPATA22 and their recruitment to DNA double-strand breaks. EBioMedicine. 2019;42:524–31.

    Article  PubMed  PubMed Central  Google Scholar 

  69. 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 

  70. Mellone S, Zavattaro M, Vurchio D, Ronzani S, Caputo M, Leone I et al. A Long Contiguous Stretch of homozygosity disclosed a Novel STAG3 biallelic pathogenic variant causing primary ovarian insufficiency: a Case Report and Review of the literature. Genes (Basel). 2021;12(11).

  71. Snowden T, Acharya S, Butz C, Berardini M, Fishel R. hMSH4-hMSH5 recognizes Holliday junctions and forms a meiosis-specific sliding clamp that embraces homologous chromosomes. Mol Cell. 2004;15(3):437–51.

    Article  CAS  PubMed  Google Scholar 

  72. 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(16):3161–6.

    CAS  PubMed  Google Scholar 

  73. Wan Y, Hong Z, Ma B, He X, Ma L, Wang M, et al. Identification of compound heterozygous variants in MSH4 as a novel genetic cause of diminished ovarian reserve. Reproductive biology and endocrinology: RB&E; 2023;21(1):76.

    Google Scholar 

  74. Guo T, Zhao S, Zhao S, Chen M, Li G, Jiao X, et al. Mutations in MSH5 in primary ovarian insufficiency. Hum Mol Genet. 2017;26(8):1452–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Wang X, Zhang X, Dang Y, Li D, Lu G, Chan WY, et al. Long noncoding RNA HCP5 participates in premature ovarian insufficiency by transcriptionally regulating MSH5 and DNA damage repair via YB1. Nucleic Acids Res. 2020;48(8):4480–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hinch AG, Becker PW, Li T, Moralli D, Zhang G, Bycroft C, et al. The configuration of RPA, RAD51, and DMC1 binding in meiosis reveals the nature of critical recombination intermediates. Mol Cell. 2020;79(4):689–e70110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Brandsma I, Sato K, van Rossum-Fikkert SE, van Vliet N, Sleddens E, Reuter M, et al. HSF2BP interacts with a conserved domain of BRCA2 and is required for mouse spermatogenesis. Cell Rep. 2019;27(13):3790–e87.

    Article  CAS  PubMed  Google Scholar 

  78. Felipe-Medina N, Caburet S, Sanchez-Saez F, Condezo YB, de Rooij DG, Gomez HL et al. A missense in HSF2BP causing primary ovarian insufficiency affects meiotic recombination by its novel interactor C19ORF57/BRME1. Elife. 2020;9.

  79. Wan L, Han J, Liu T, Dong S, Xie F, Chen H, et al. Scaffolding protein SPIDR/KIAA0146 connects the Bloom syndrome helicase with homologous recombination repair. Proc Natl Acad Sci USA. 2013;110(26):10646–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Smirin-Yosef P, Zuckerman-Levin N, Tzur S, Granot Y, Cohen L, Sachsenweger J, et al. A biallelic mutation in the homologous recombination repair gene SPIDR is Associated with human gonadal dysgenesis. J Clin Endocrinol Metab. 2017;102(2):681–8.

    Article  PubMed  Google Scholar 

  81. Huang T, Wu X, Wang S, Bao Z, Wan Y, Wang Z, et al. SPIDR is required for homologous recombination during mammalian meiosis. Nucleic Acids Res. 2023;51(8):3855–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. McGlacken-Byrne SM, Le Quesne Stabej P, Del Valle I, Ocaka L, Gagunashvili A, Crespo B, et al. ZSWIM7 is Associated with Human Female meiosis and familial primary ovarian insufficiency. J Clin Endocrinol Metab. 2022;107(1):e254–63.

    Article  PubMed  Google Scholar 

  83. Yatsenko SA, Gurbuz F, Topaloglu AK, Berman AJ, Martin PM, Rodrigue-Escriba M, et al. Pathogenic variants in ZSWIM7 cause primary ovarian insufficiency. J Clin Endocrinol Metab. 2022;107(6):e2359–64.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Hughes CHK, Smith OE, Meinsohn MC, Brunelle M, Gévry N, Murphy BD. Steroidogenic factor 1 (SF-1; Nr5a1) regulates the formation of the ovarian reserve. Proc Natl Acad Sci USA. 2023;120(32):e2220849120.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Smith OE, Roussel V, Morin F, Ongaro L, Zhou X, Bertucci MC et al. Steroidogenic factor 1 regulation of the Hypothalamic-Pituitary-Ovarian Axis of Adult Female mice. Endocrinology. 2022;163(4).

  86. Yan H, Wen J, Zhang T, Zheng W, He M, Huang K, et al. Oocyte-derived E-cadherin acts as a multiple functional factor maintaining the primordial follicle pool in mice. Cell Death Dis. 2019;10(3):160.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Meinsohn MC, Smith OE, Bertolin K, Murphy BD. The Orphan Nuclear receptors Steroidogenic Factor-1 and liver receptor Homolog-1: structure, regulation, and essential roles in mammalian Reproduction. Physiol Rev. 2019;99(2):1249–79.

    Article  CAS  PubMed  Google Scholar 

  88. Qin Y, Choi Y, Zhao H, Simpson JL, Chen ZJ, Rajkovic A. NOBOX homeobox mutation causes premature ovarian failure. Am J Hum Genet. 2007;81(3):576–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Patton BK, Madadi S, Briley SM, Ahmed AA, Pangas SA. Sumoylation regulates functional properties of the oocyte transcription factors SOHLH1 and NOBOX. FASEB Journal: Official Publication Federation Am Soc Experimental Biology. 2023;37(2):e22747.

    Article  CAS  Google Scholar 

  90. Wang Z, Liu CY, Zhao Y, Dean J, FIGLA. LHX8 and SOHLH1 transcription factor networks regulate mouse oocyte growth and differentiation. Nucleic Acids Res. 2020;48(7):3525–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Zhao L, Li Q, Kuang Y, Xu P, Sun X, Meng Q, et al. Heterozygous loss-of-function variants in LHX8 cause female infertility characterized by oocyte maturation arrest. Genet Medicine: Official J Am Coll Med Genet. 2022;24(11):2274–84.

    Article  CAS  Google Scholar 

  92. Mei L, Huang Y, Wu X, He H, Ye R, Ma J, et al. Mutations in FIGLA Associated with premature ovarian insufficiency in a Chinese Population. Front Med. 2021;8:714306.

    Article  Google Scholar 

  93. di Clemente N, Racine C, Pierre A, Taieb J. Anti-Müllerian hormone in Female Reproduction. Endocr Rev. 2021;42(6):753–82.

    Article  PubMed  Google Scholar 

  94. Oduwole OO, Huhtaniemi IT, Misrahi M. The roles of luteinizing hormone, follicle-stimulating hormone and testosterone in spermatogenesis and folliculogenesis revisited. Int J Mol Sci. 2021;22(23).

  95. Renault L, Patiño LC, Magnin F, Delemer B, Young J, Laissue P et al. BMPR1A and BMPR1B missense mutations cause primary ovarian insufficiency. J Clin Endocrinol Metab. 2020;105(4).

  96. Zariñán T, Mayorga J, Jardón-Valadez E, Gutiérrez-Sagal R, Maravillas-Montero JL, Mejía-Domínguez NR, et al. A novel mutation in the FSH receptor (I423T) affecting receptor activation and leading to primary ovarian failure. J Clin Endocrinol Metab. 2021;106(2):e534–50.

    Article  PubMed  Google Scholar 

  97. Malone SA, Papadakis GE, Messina A, Mimouni NEH, Trova S, Imbernon M et al. Defective AMH signaling disrupts GnRH neuron development and function and contributes to hypogonadotropic hypogonadism. eLife. 2019;8.

  98. Hart KN, Pépin D, Czepnik M, Donahoe PK, Thompson TB. Mutational analysis of the putative Anti-Müllerian hormone (AMH) binding interface on its type II receptor, AMHR2. Endocrinology. 2020;161(7).

  99. Qin C, Yuan Z, Yao J, Zhu W, Wu W, Xie J. AMH and AMHR2 genetic variants in Chinese women with primary ovarian insufficiency and normal age at natural menopause. Reprod Biomed Online. 2014;29(3):311–8.

    Article  CAS  PubMed  Google Scholar 

  100. Li L, Zhou X, Wang X, Wang J, Zhang W, Wang B, et al. A dominant negative mutation at the ATP binding domain of AMHR2 is associated with a defective anti-Müllerian hormone signaling pathway. Mol Hum Reprod. 2016;22(9):669–78.

    Article  CAS  PubMed  Google Scholar 

  101. Zhao Y, Pan S, Li Y, Wu X. Exosomal mir-143-3p derived from follicular fluid promotes granulosa cell apoptosis by targeting BMPR1A in polycystic ovary syndrome. Sci Rep. 2022;12(1):4359.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Liu MN, Zhang K, Xu TM. The role of BMP15 and GDF9 in the pathogenesis of primary ovarian insufficiency. Hum Fertil (Cambridge England). 2021;24(5):325–32.

    Article  CAS  Google Scholar 

  103. Norling A, Hirschberg AL, Rodriguez-Wallberg KA, Iwarsson E, Wedell A, Barbaro M. Identification of a duplication within the GDF9 gene and novel candidate genes for primary ovarian insufficiency (POI) by a customized high-resolution array comparative genomic hybridization platform. Hum Reprod (Oxford England). 2014;29(8):1818–27.

    Article  CAS  Google Scholar 

  104. Zhang T, Ma Q, Shen Q, Jiang C, Zou F, Shen Y, et al. Identification of novel biallelic variants in BMP15 in two siblings with premature ovarian insufficiency. J Assist Reprod Genet. 2022;39(9):2125–34.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Ferrarini E, De Marco G, Orsolini F, Gianetti E, Benelli E, Fruzzetti F, et al. Characterization of a novel mutation V136L in bone morphogenetic protein 15 identified in a woman affected by POI. J Ovarian Res. 2021;14(1):85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Afkhami F, Shahbazi S, Farzadi L, Danaei S. Novel bone morphogenetic protein 15 (BMP15) gene variants implicated in premature ovarian insufficiency. Reproductive biology and endocrinology: RB&E; 2022;20(1):42.

    Google Scholar 

  107. Chu K, He Y, Li Z, Jiang Z, Wang L, Ji Y, et al. Novel LAT pathogenic variants in a POI family and its role in the Ovary. Front Genet. 2021;12:764160.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Li Y, Liu F, Li S, Tan S. Association of Variants in vascular endothelial growth factor a gene and VEGFA serum levels with the risk of primary ovarian insufficiency: a case-control study. Gynecol Obstet Invest. 2021;86(5):469–76.

    Article  CAS  PubMed  Google Scholar 

  109. Li J, Zeng Z, Chang Y, Li M, Wu Q, Chen P, et al. Suppressive effects of Ursolic Acid on Human endometriotic stromal cells survival. Gynecol Obstet Invest. 2020;85(1):72–81.

    Article  CAS  PubMed  Google Scholar 

  110. Wu X, Wang B, Dong Z, Zhou S, Liu Z, Shi G, et al. A NANOS3 mutation linked to protein degradation causes premature ovarian insufficiency. Cell Death Dis. 2013;4(10):e825.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Kasippillai T, MacArthur DG, Kirby A, Thomas B, Lambalk CB, Daly MJ, et al. Mutations in eIF4ENIF1 are associated with primary ovarian insufficiency. J Clin Endocrinol Metab. 2013;98(9):E1534–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Zhao M, Feng F, Chu C, Yue W, Li L. A novel EIF4ENIF1 mutation associated with a diminished ovarian reserve and premature ovarian insufficiency identified by whole-exome sequencing. J Ovarian Res. 2019;12(1):119.

    Article  PubMed  PubMed Central  Google Scholar 

  113. Rahman S, Copeland WC. POLG-related disorders and their neurological manifestations. Nat Reviews Neurol. 2019;15(1):40–52.

    Article  CAS  Google Scholar 

  114. Pagnamenta AT, Taanman JW, Wilson CJ, Anderson NE, Marotta R, Duncan AJ, et al. Dominant inheritance of premature ovarian failure associated with mutant mitochondrial DNA polymerase gamma. Hum Reprod (Oxford England). 2006;21(10):2467–73.

    Article  CAS  Google Scholar 

  115. Luoma P, Melberg A, Rinne JO, Kaukonen JA, Nupponen NN, Chalmers RM, et al. Parkinsonism, premature menopause, and mitochondrial DNA polymerase gamma mutations: clinical and molecular genetic study. Lancet (London England). 2004;364(9437):875–82.

    Article  CAS  PubMed  Google Scholar 

  116. Chen B, Li L, Wang J, Zhou Y, Zhu J, Li T, et al. Identification of the first homozygous POLG mutation causing non-syndromic ovarian dysfunction. Climacteric: J Int Menopause Soc. 2018;21(5):467–71.

    Article  CAS  Google Scholar 

  117. Gustafsson CM, Falkenberg M, Larsson NG. Maintenance and expression of mammalian mitochondrial DNA. Annu Rev Biochem. 2016;85:133–60.

    Article  CAS  PubMed  Google Scholar 

  118. Song Y, Wang W, Wang B, Shi Q. The protective mechanism of TFAM on mitochondrial DNA and its role in neurodegenerative diseases. Mol Neurobiol. 2023.

  119. Tucker EJ, Rius R, Jaillard S, Bell K, Lamont PJ, Travessa A, et al. Genomic sequencing highlights the diverse molecular causes of Perrault syndrome: a peroxisomal disorder (PEX6), metabolic disorders (CLPP, GGPS1), and mtDNA maintenance/translation disorders (LARS2, TFAM). Hum Genet. 2020;139(10):1325–43.

    Article  CAS  PubMed  Google Scholar 

  120. Ullah F, Rauf W, Khan K, Khan S, Bell KM, de Oliveira VC, et al. A recessive variant in TFAM causes mtDNA depletion associated with primary ovarian insufficiency, seizures, intellectual disability and hearing loss. Hum Genet. 2021;140(12):1733–51.

    Article  CAS  PubMed  Google Scholar 

  121. Boczonadi V, Ricci G, Horvath R. Mitochondrial DNA transcription and translation: clinical syndromes. Essays Biochem. 2018;62(3):321–40.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Tiosano D, Mears JA, Buchner DA. Mitochondrial dysfunction in primary ovarian insufficiency. Endocrinology. 2019;160(10):2353–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Amunts A, Brown A, Toots J, Scheres SHW, Ramakrishnan V. Ribosome. The structure of the human mitochondrial ribosome. Sci (New York NY). 2015;348(6230):95–8.

    Article  CAS  Google Scholar 

  124. Chen A, Tiosano D, Guran T, Baris HN, Bayram Y, Mory A, et al. Mutations in the mitochondrial ribosomal protein MRPS22 lead to primary ovarian insufficiency. Hum Mol Genet. 2018;27(11):1913–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Jolly A, Bayram Y, Turan S, Aycan Z, Tos T, Abali ZY, et al. Exome sequencing of a primary ovarian insufficiency cohort reveals common Molecular etiologies for a spectrum of Disease. J Clin Endocrinol Metab. 2019;104(8):3049–67.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Siira SJ, Spåhr H, Shearwood AJ, Ruzzenente B, Larsson NG, Rackham O, et al. LRPPRC-mediated folding of the mitochondrial transcriptome. Nat Commun. 2017;8(1):1532.

    Article  PubMed  PubMed Central  Google Scholar 

  127. Volpon L, Culjkovic-Kraljacic B, Sohn HS, Blanchet-Cohen A, Osborne MJ, Borden KLB. A biochemical framework for eIF4E-dependent mRNA export and nuclear recycling of the export machinery. New York, NY: RNA; 2017;23(6)927–37.

    Google Scholar 

  128. Spaulding Z, Thevarajan I, Schrag LG, Zubcevic L, Zolkiewska A, Zolkiewski M. Human mitochondrial AAA + ATPase SKD3/CLPB assembles into nucleotide-stabilized dodecamers. Biochem Biophys Res Commun. 2022;602:21–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Tucker EJ, Baker MJ, Hock DH, Warren JT, Jaillard S, Bell KM, et al. Premature ovarian insufficiency in CLPB Deficiency: transcriptomic, proteomic and phenotypic insights. J Clin Endocrinol Metab. 2022;107(12):3328–40.

    Article  PubMed  PubMed Central  Google Scholar 

  130. Luo B, Ma Y, Zhou Y, Zhang N, Luo Y. Human ClpP protease, a promising therapy target for diseases of mitochondrial dysfunction. Drug Discovery Today. 2021;26(4):968–81.

    Article  CAS  PubMed  Google Scholar 

  131. Gispert S, Parganlija D, Klinkenberg M, Dröse S, Wittig I, Mittelbronn M, et al. Loss of mitochondrial peptidase clpp leads to infertility, hearing loss plus growth retardation via accumulation of CLPX, mtDNA and inflammatory factors. Hum Mol Genet. 2013;22(24):4871–87.

    Article  CAS  PubMed  Google Scholar 

  132. Yuan X, Ma W, Chen S, Wang H, Zhong C, Gao L et al. CLPP inhibition triggers apoptosis in human ovarian granulosa cells via COX5A abnormality–mediated mitochondrial dysfunction. 2023;14:1141167.

  133. Zhang J, Xu Y, Liu H, Pan Z. MicroRNAs in ovarian follicular atresia and granulosa cell apoptosis. Reproductive Biology Endocrinology: RB&E. 2019;17(1):9.

    Article  Google Scholar 

  134. Dang Y, Wang X, Hao Y, Zhang X, Zhao S, Ma J, et al. MicroRNA-379-5p is associate with biochemical premature ovarian insufficiency through PARP1 and XRCC6. Cell Death Dis. 2018;9(2):106.

    Article  PubMed  PubMed Central  Google Scholar 

  135. Zhang X, Dang Y, Liu R, Zhao S, Ma J, Qin Y. MicroRNA-127-5p impairs function of granulosa cells via HMGB2 gene in premature ovarian insufficiency. J Cell Physiol. 2020;235(11):8826–38.

    Article  CAS  PubMed  Google Scholar 

  136. Chen X, Xie M, Liu D, Shi K. Downregulation of microRNA–146a inhibits ovarian granulosa cell apoptosis by simultaneously targeting interleukin–1 receptor–associated kinase and tumor necrosis factor receptor–associated factor 6. Mol Med Rep. 2015;12(4):5155–62.

    Article  CAS  PubMed  Google Scholar 

  137. Yang X, Zhou Y, Peng S, Wu L, Lin HY, Wang S, et al. Differentially expressed plasma microRNAs in premature ovarian failure patients and the potential regulatory function of mir-23a in granulosa cell apoptosis. Reprod (Cambridge England). 2012;144(2):235–44.

    Article  CAS  Google Scholar 

  138. Luo H, Han Y, Liu J, Zhang Y. Identification of microRNAs in granulosa cells from patients with different levels of ovarian reserve function and the potential regulatory function of miR-23a in granulosa cell apoptosis. Gene. 2019;686:250–60.

    Article  CAS  PubMed  Google Scholar 

  139. Li D, Xu W, Wang X, Dang Y, Xu L, Lu G, et al. lncRNA DDGC participates in premature ovarian insufficiency through regulating RAD51 and WT1. Mol Therapy Nucleic Acids. 2021;26:1092–106.

    Article  CAS  Google Scholar 

  140. Luo C, Wei L, Qian F, Bo L, Gao S, Yang G, et al. LncRNA HOTAIR regulates autophagy and proliferation mechanisms in premature ovarian insufficiency through the miR-148b-3p/ATG14 axis. Cell Death Discovery. 2024;10(1):44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Zhang L, Mao B, Zhao X, Yuan Y, Wang W, Lin S. Translation regulatory long non-coding RNA 1 (TRERNA1) sponges microRNA-23a to suppress granulosa cell apoptosis in premature ovarian failure. Bioengineered. 2022;13(2):2173–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Li D, Wang X, Li G, Dang Y, Zhao S, Qin Y. LncRNA ZNF674-AS1 regulates granulosa cell glycolysis and proliferation by interacting with ALDOA. Cell Death Discovery. 2021;7(1):107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Li D, Wang X, Dang Y, Zhang X, Zhao S, Lu G, et al. lncRNA GCAT1 is involved in premature ovarian insufficiency by regulating p27 translation in GCs via competitive binding to PTBP1. Mol Therapy Nucleic Acids. 2021;23:132–41.

    Article  CAS  Google Scholar 

  144. 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(3).

  145. Caburet S, Arboleda VA, Llano E, Overbeek PA, Barbero JL, Oka K, et al. Mutant cohesin in premature ovarian failure. New Engl J Med. 2014;370(10):943–9.

  146. 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(4):523–34.

  147. Hartford SA, Luo YH, Southard TL, Min IM, Lis JT, Schimenti JC. Minichromosome maintenance helicase paralog MCM9 is dispensible for DNA replication but functions in germ-line stem cells and tumor suppression. P Natl Acad Sci USA. 2011;108(43):17702–7.

  148. Wu Y, Li Y, Murtaza G, Zhou J, Jiao Y, Gong C, et al. Whole-exome sequencing of consanguineous families with infertile men and women identifies homologous mutations in SPATA22 and MEIOB. Hum Reprod. 2021;36(10):2793–804.

  149. Souquet B, Abby E, Hervé R, Finsterbusch F, Tourpin S, Le Bouffant R, et al. MEIOB targets single-strand DNA and is necessary for meiotic recombination. Plos Genet. 2013;9(9).

  150. Luo MC, Yang F, Leu NA, Landaiche J, Handel MA, Benavente R, et al. MEIOB exhibits single-stranded DNA-binding and exonuclease activities and is essential for meiotic recombination. Nat Commun. 2013;4.

  151. Sirchia F, Giorgio E, Cucinella L, Valente EM, Nappi RE. Biallelic mutations in PSMC3IP are associated with secondary amenorrhea: expanding the spectrum of premature ovarian insufficiency. Journal of assisted reproduction and genetics. 2022;39(5):1177–81.

  152. Petukhova GV, Romanienko PJ, Camerini-Otero RD. The Hop2 protein has a direct role in promoting interhomolog interactions during mouse meiosis. Dev Cell. 2003;5(6):927–36.

  153. Felipe-Medina N, Caburet S, Sánchez-Sáez F, Condezo YB, de Rooij DG, Gómez HL, et al. A missense in HSF2BP causing primary ovarian insufficiency affects meiotic recombination by its novel interactor C19ORF57/BRME1. eLife. 2020;9.

  154. Zhang JJ, Fujiwara Y, Yamamoto S, Shibuya H. A meiosis-specific BRCA2 binding protein recruits recombinases to DNA double-strand breaks to ensure homologous recombination. Nat Commun. 2019;10:722.

  155. Yatsenko SA, Gurbuz F, Topaloglu AK, Berman AJ, Martin PM, Rodrigue-Escribà M, et al. Pathogenic variants in ZSWIM7 cause primary ovarian insufficiency. J Clin Endocrinol Metab. 2022;107(6):e2359–e64.

  156. Smirin-Yosef P, Zuckerman-Levin N, Tzur S, Granot Y, Cohen L, Sachsenweger J, et al. A biallelic mutation in the homologous recombination repair gene SPIDR is associated with human gonadal dysgenesis. J Clin Endocrinol Metab. 2017;102(2):681–8.

  157. Rajkovic A, Pangas SA, Ballow D, Suzumori N, Matzuk MM. NOBOX deficiency disrupts early folliculogenesis and oocyte-specific gene expression. Sci. 2004;305(5687):1157–9.

  158. Soyal SM, Amleh A, Dean J. FIGα, a germ cell-specific transcription factor required for ovarian follicle formation. Development. 2000;127(21):4645–54.

  159. Alvaro Mercadal B, Imbert R, Demeestere I, Gervy C, De Leener A, Englert Y, et al. AMH mutations with reduced in vitro bioactivity are related to premature ovarian insufficiency. Hum Reprod. 2015;30(5):1196–202.

  160. Demain LAM, Antunes D, O'Sullivan J, Bhaskhar SS, O'Keefe RT, Newman WG. A known pathogenic variant in the essential mitochondrial translation gene RMND1 causes a Perrault-like syndrome with renal defects. Clin Genet. 2018;94(2):276–7.

  161. Oziębło D, Pazik J, Stępniak I, Skarżyński H, Ołdak M. Two novel pathogenic variants confirm RMND1 causative role in perrault syndrome with renal involvement. Genes. 2020;11(9).

  162. Trifunovic A, Wredenberg A, Falkenberg M, Spelbrink JN, Rovio AT, Bruder CE, et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature. 2004;429(6990):417–23.

  163. Humble MM, Young MJ, Foley JF, Pandiri AR, Travlos GS, Copeland WC. Polg2 is essential for mammalian embryogenesis and is required for mtDNA maintenance. Hum Mol Genet. 2013;22(5):1017–25.

  164. Dosekova P, Dubiel A, Karlowicz A, Zietkiewicz S, Rydzanicz M, Habalova V, et al. Whole exome sequencing identifies a homozygous POLG2 missense variant in an adult patient presenting with optic atrophy, movement disorders, premature ovarian failure and mitochondrial DNA depletion. Eur J Med Genet. 2020;63(4):103821.

  165. Larsson NG, Wang J, Wilhelmsson H, Oldfors A, Rustin P, Lewandoski M, et al. Mitochondrial transcription factor A is necessary for mtDNA maintenance and embryogenesis in mice. Nat Genet. 1998;18(3):231–6.

  166. Cuillerier A, Honarmand S, Cadete VJJ, Ruiz M, Forest A, Deschênes S, et al. Loss of hepatic LRPPRC alters mitochondrial bioenergetics, regulation of permeability transition and trans-membrane ROS diffusion. Hum Mol Genet. 2017;26(16):3186–201.

  167. Mourier A, Ruzzenente B, Brandt T, Kühlbrandt W, Larsson NG. Loss of LRPPRC causes ATP synthase deficiency. Hum Mol Genet. 2014;23(10):2580–92.

  168. Ghaddhab C, Morin C, Brunel-Guitton C, Mitchell GA, Van Vliet G, Huot C. Premature ovarian failure in french canadian leigh syndrome. J Pediatr. 2017;184:227–9.e1.

  169. Gispert S, Parganlija D, Klinkenberg M, Dröse S, Wittig I, Mittelbronn M, et al. Loss of mitochondrial peptidase Clpp leads to infertility, hearing loss plus growth retardation via accumulation of CLPX, mtDNA and inflammatory factors. Hum Mol Genet. 2013;22(24):4871–87.

  170. Xu P, Wang L, Peng H, Liu H, Liu H, Yuan Q, et al. Disruption of Hars2 in cochlear hair cells causes progressive mitochondrial dysfunction and hearing loss in mice. Front Cell Neurosci. 2021;15:804345.

  171. Zhang J, Zhou XY, Wang A, Lai YH, Zhang XF, Liu XT, et al. Novel Tu translation elongation factor, mitochondrial (TUFM) homozygous variant in a consanguineous family with premature ovarian insufficiency. Clin Genet. 2023;104(5):516–27.

  172. Dallabona C, Diodato D, Kevelam SH, Haack TB, Wong LJ, Salomons GS, et al. Novel (ovario) leukodystrophy related to AARS2 mutations. Neurol. 2014;82(23):2063–71.

Download references

Acknowledgements

No/Not applicable.

Funding

This project has received funding from National Natural Science Foundation of China (NSFC) (No. 82301817 to MW); Innovation and Entrepreneurship Training Program for College Students (S202410486359, to MW, ZH, LN, XW, and SW).

Author information

Authors and Affiliations

Authors

Contributions

MW and ZH conceived and drafted the manuscript. LN, XW, and SW drafted the manuscript. XW and SW drew the figures. LN, XW, and SW prepared the tables. LN, XW, and SW collected the references. MW and ZH proofread the manuscript and made critical revisions. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Zhidan Hong or Mei Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

An informed written consent was signed by the participant.

Competing interests

The authors declare no competing interests.

Author information

Linhang Nie, Xiaojie Wang, and Songyuan Wang contributed equally to this work.

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

Nie, L., Wang, X., Wang, S. et al. Genetic insights into the complexity of premature ovarian insufficiency. Reprod Biol Endocrinol 22, 94 (2024). https://doi.org/10.1186/s12958-024-01254-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12958-024-01254-2