Next Article in Journal
Tight Spaces, Tighter Signals: Spatial Constraints as Drivers of Peripheral Myelination
Next Article in Special Issue
Biological Aging and Uterine Fibrosis in Cattle: Reproductive Trade-Offs from Enhanced Productivity
Previous Article in Journal
Lower Zinc but Higher Calcium Content in Rodent Spinal Cord Compared to Brain
Previous Article in Special Issue
Molecular Mechanisms of Cellular Senescence in Age-Related Endometrial Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Spotlight on Proteases: Roles in Ovarian Health and Disease

Department of Biochemistry and Molecular Biology, Indiana University, Indianapolis, IN 46202, USA
*
Author to whom correspondence should be addressed.
Cells 2025, 14(12), 921; https://doi.org/10.3390/cells14120921
Submission received: 30 April 2025 / Revised: 2 June 2025 / Accepted: 4 June 2025 / Published: 18 June 2025
(This article belongs to the Special Issue Cellular and Molecular Mechanisms in Gynecological Disorders)

Abstract

Proteases play crucial roles in ovarian folliculogenesis, regulating several processes from primordial follicle activation to ovulation and corpus luteum formation. This review synthesizes the current knowledge on the diverse functions of proteases in ovarian physiology and pathology. We discuss the classification and regulation of proteases, highlighting their importance in extracellular matrix remodeling, cell signaling, and apoptosis during ovarian follicular development. We explore the roles of several proteases including matrix metalloproteinases, tissue inhibitors of metalloproteinases, the plasminogen activator system, and cathepsins, and their roles in the critical functions of ovarian biology including follicle dynamics and senescence. Furthermore, we address the involvement of proteases in ovarian pathologies, including cancer, polycystic ovary syndrome, and primary ovarian insufficiency. By integrating recent findings from clinical genomics and animal models, this review provides a comprehensive overview of protease functions in the ovary, emphasizing their potential use for therapeutic interventions in reproductive medicine.

1. Introduction

Several ovarian processes from follicle activation to the maturation and release of oocytes require the catalytic action of protease enzymes [1,2,3,4]. Since first reported in 1905 by P. A. Levene, proteases remain at the forefront of basic and clinical research due to their regulatory activity of fundamental and ubiquitous cellular functions [5]. Recent innovations in large-scale genome sequencing and high-throughput differential screening methods have established that the mouse and human genomes contain 628 and 566 protease-encoding genes, respectively [6,7,8,9,10]. In humans, there are 273 extracellular proteases, 277 intracellular proteases, and 16 integral membrane proteases. The mouse degradome includes at least 628 members with 514 true orthologs of human proteases [11,12,13]. The specific function of several proteases remains unclear, and roughly 100 of these proteins are believed to be inactive due to the lack of critical catalytic domains [14]. Proteases are classified into two main groups: exopeptidases, which cleave peptide bonds near the amino or carboxy terminus (aminopeptidases and carboxypeptidases), and endopeptidases, which cut inside proteins, away from the ends. Based on the type of catalytic domain, proteases are grouped into six classes [15,16,17]. Serine, cysteine, and threonine proteases use nucleophilic amino acids for cleavage, whereas aspartic, glutamic, and metalloproteases activate water molecules as nucleophiles for the hydrolysis of peptide bonds (Figure 1) [18,19,20]. Most proteases are synthesized as inactive precursors called zymogens or proenzymes. These contain inhibitory pro-domains that prevent premature proteolytic activity [17,21]. Proteases lacking pro-domains typically require cofactor binding or post-translational modifications to achieve an active state (Figure 1) [22,23,24].
Proteases modulate growth factor/cytokine bioavailability, control the levels of cell surface/receptor proteins, and are involved in the reorganization of the extracellular matrix (ECM) [25,26,27]. Therefore, their activity requires fine regulation through several mechanisms including inhibition by specific proteins, degradation by other proteases, and subcellular relocalization [28].
In the ovary, several proteases, including matrix metalloproteinases (MMPs), a disintegrin and metalloproteinase with thrombospondin-like motifs (ADAMTS), cathepsins (CTS), and plasminogen activator (PA)/plasmin system enzymes, have been reported to play critical roles in folliculogenesis, tissue remodeling, and ovulation [1,3,29,30,31,32,33]. These critical ovarian processes require the reorganization of the collagen-rich ECM surrounding the follicles, as well as the basement membrane separating granulosa and theca cells [34,35,36,37,38]. In addition, proteases regulate other fundamental processes including apoptosis and autophagy [39,40,41,42,43,44]. Ovarian folliculogenesis is tightly regulated by endocrine, paracrine, and autocrine factors, which affect the growth and maturation of oocytes and surrounding granulosa cells [25,45]. MMPs, including MMP-2, MMP-9, MMP-13, and membrane-type MMP-14 and MMP-16, serve as zinc- and calcium-dependent enzymes that collectively degrade proteinaceous ECM components (proteoglycans, laminin, and collagen) [6,46,47]. Their activities are tightly controlled by tissue inhibitors of metalloproteinases (TIMPs) to maintain the critical MMP/TIMP ratio necessary to regulate cell proliferation, differentiation, migration, and survival throughout follicular development [19,35]. The luteinizing hormone surge triggers the transcriptional activation of key proteases, particularly ADAMTS-1 (a disintegrin and metalloproteinase with thrombospondin motifs) and cathepsin L, which play critical roles in follicle rupture [48,49] (Figure 2). ADAMTS-1 facilitates structural remodeling during folliculogenesis by degrading proteoglycans and maintaining basement membrane integrity, whereas its disruption leads to follicle dysgenesis and impaired ovulation [48,49,50,51,52]. Cathepsin B and L, lysosomal cysteine proteases, regulate granulosa cell apoptosis, proliferation, and steroidogenesis while also modulating autophagy. Cathepsin L activity correlates directly with meiotic progression and embryonic development [53,54]. This network of proteases exhibits specific spatio-temporal expression patterns, and their coordinated action enables the cyclical tissue remodeling essential for follicular growth, ovulation, and corpus luteum formation [11]. Despite their ubiquitous function in both germ and somatic cells, specific roles of proteases in ovarian biology and disease are still far from being understood. This review summarizes the current knowledge on this class of proteins and their functions in follicle dynamics, ovulation, and corpus luteum development, as well as their involvement in ovarian disorders.

2. Materials and Methods

We searched databases including PubMed, Web of Science, and Scopus for the following keywords: “Protease” OR “Proteinase” OR “Peptidase, Ovary” OR “Ovarian, Folliculogenesis” OR “Follicle development, Ovarian pathology” OR “Ovarian disease, Metalloprotease” OR “MMP, Serine protease, Cysteine protease, Aspartic protease, ADAM” OR “ADAMTS, Cathepsin, Ovulation, Atresia, Polycystic ovary syndrome” OR “PCOS, Premature ovarian failure” OR “POF, Ovarian cancer”. Inclusion Criteria: Studies focusing on proteases in the ovary, including research on proteases involved in folliculogenesis and ovarian pathologies, and in vivo and in vitro studies in both humans and mammalian models. To ensure relevance while capturing important foundational work, we considered only research articles and reviews published in peer-reviewed journals in the last 30 years. Exclusion Criteria: Studies focusing solely on non-ovarian tissues or without direct relevance to ovarian function or pathology, articles not peer-reviewed, publications older than 30 years, case reports (unless they provide unique insights not available elsewhere), articles not available in English, and conference abstracts.

3. Roles of Proteases in Ovarian Follicle Development

3.1. Metalloproteinases, Plasminogen Activators, and Their Inhibitors

MMPs, tissue inhibitors of metalloproteinases (TIMPs), and PAs play crucial roles in regulating the remodeling of the ovarian ECM (Table 1) [3,4,6]. MMPs are zinc-dependent endopeptidases that are involved in physiological processes, such as embryonic development, tissue morphogenesis, and angiogenesis, and play critical roles in pathological conditions, including inflammation, cancer, and cardiovascular disease [7,8,9]. MMPs are classified into several groups based on their substrate specificity and structural features: collagenases (MMP1, 8, 13, and 18), gelatinases (MMP2 and 9), stromelysins (MMP3, 10, and 11), matrilysins (MMP7 and 26), membrane-type MMPs (MMP14, 15, 16, 17, 24, and 25), metalloelastase (MMP12), epilysin (MMP28), and enamelysin (MMP20) [8,10,11,14,15,16]. MMPs are secreted as zymogens (pro-MMPs) and their function is regulated at multiple levels, including transcription, zymogen activation, and inhibition by TIMPs [6,15,17]. TIMPs are a family of four proteins (TIMP-1, -2, -3, and -4) that bind to the active site of MMPs in a 1:1 stoichiometry, thereby inhibiting their proteolytic activity [7,16,18,19]. In addition, TIMPs have MMP-independent functions and are involved in cell growth and differentiation, angiogenesis, and apoptosis [20,21,22,23]. The balance between MMPs and TIMPs is critical for maintaining ECM homeostasis and regulating follicle development in the ovary [2,24]. The activation of primordial follicles involves the breakdown of their basement membrane, allowing the proliferation of granulosa cells [25,27,28]. The inhibition of MMP activity results in a significant reduction in primary and preantral follicle numbers [1,2,34,35]. MMP2 and MMP9 are expressed throughout folliculogenesis in both granulosa cells and oocytes [12,29,30,31,45]. TIMP-1 and TIMP-2 are also expressed in the granulosa cells and oocytes of primordial and primary follicles in humans [30]. Their expression decreases during the transition from primordial to primary follicles, suggesting that a reduction in TIMP activity may be necessary for follicular activation [30]. TIMP-1-deficient mice show increased MMP activity, particularly MMP2, and MMP9, and display an increased number of primary and preantral follicular compared to wild-type mice [36,37,38,55] (Table 2).
In addition to MMPs and TIMPs, PAs play an important role in the ovary [56]. PAs convert inactive plasminogen into plasmin, which is involved in the degradation of the ECM proteins, including fibrin, fibronectin, and laminin [57,58,59]. The PA system consists of a tissue-type plasminogen activator (tPA) and a urokinase-type plasminogen activator (uPA). PAs are regulated by protease inhibitors α2-antiplasmin and α2-macroglobulin, and plasminogen activator inhibitors (PAIs), which belong to the serine proteinase inhibitor (serpin) gene superfamily and include PAI-1, PAI-2, PAI-3, and the protease nexin I [39,40,41,42,43,57,60]. In the ovary, PAs are produced in granulosa cells and oocytes, whereas the majority of PAI-1 is produced within the theca (interstitium) [4,61,62,63]. uPA and tPA are upregulated during follicle activation and maturation, and their expression is regulated by gonadotropins, growth factors, and cytokines (Table 1) [57,59,64,65,66]. Li et al. showed that TGFα treatment increases uPA levels and activity in hen granulosa cells across all the follicle stages [66]. Similarly, FSH induces both uPA and tPA, whereas IL-1β suppresses their activity in vitro [67]. In rat ovaries, Hurwitz et al. also reported that IL-1β functions as an inhibitor of the PA system [68].
Other serine proteases that have been reported in the regulation of folliculogenesis include LONP1, FURIN, PAPPA, and matriptase (Table 2). The mitochondrial LONP1 is a multifunctional protease involved in mitochondrial quality control including oxidized protein degradation, protein folding, and mitochondrial DNA copy number homeostasis. The oocyte-specific ablation of Lonp using Gdf9-cre or Zp3-cre results in female infertility due to impaired follicle development, loss of the ovarian reserve, and progressive oocyte death ([69]. FURIN encodes a transmembrane serine protease localized in the Golgi apparatus, endosomes, and plasma membrane. The conditional ablation of Furin using Gdf9-cre or Zp3-cre leads to female infertility due to arrested folliculogenesis at the secondary follicle stage [70]. PAPPA encodes an extracellular metalloprotease, and Pappa knockout females exhibit decreased litter size and ovulatory capacity, which was attributed to the decreased bioavailability of insulin-like growth factor [71,72,73]. Matriptase, encoded by Tmprss6, is a type II transmembrane serine protease that regulates iron homeostasis by cleaving cell surface proteins associated with iron absorption. Tmprss6-null females show a severe delay in follicle maturation, likely due to a significant decrease in plasma iron levels [74]. Notably, defective follicle development and female infertility can be reproduced by a low-iron diet [75].
Table 1. Protease functions during ovarian physiology and follicular development.
Table 1. Protease functions during ovarian physiology and follicular development.
ProteaseMechanism of RegulationSpecific StageFunctionSpeciesReference
MMP1Increased expression following hCG administrationPreovulatoryDegradation of collagenous ECMRhesus monkey[18,76]
MMP2 and MMP9Increased in the granulosa and thecal cells of atretic follicles during proestrus and in corpus luteum during metestrusPreovulatory folliclesECM remodelingGuinea pigs[45]
MMP2, and MMP9Localized to the oogonium/oocyte cytoplasm and surface epitheliumFolliculogenesisECM remodeling during gonadal development and cell–matrix interactionsHuman[77]
MMP1 and MMP13Expression increased in response to LH surgePreovulatoryDegradation of collagenous ECMBovine[78]
MMP1, MMP2, MMP3, MMP9, MMP13 Increased in mRNA expression by gonadotropins Prehierarchical white (WFs), yellowish (YFs), and preovulatory folliclesInvolved in the atresia of the early stage of follicle while not participating in the regulation of advanced stage atresiaChicken[24,79]
MMP1, MMP3, and MMP9Increased MMP1 and MMP3 expression levels in granulosa FolliculogenesisMMP9 induced by TGFB1; MMP1and MMP3 stimulated by FSH, LH, P4, and E2Chicken[80]
MMP10 and MMP11Expression patterns changes following hCG administration Ovulation and luteogenesisMmp10 mRNA was increased and MMP11 decreased in granulosa and theca cells during OvulationHuman and Rats[81]
MMP13, MMP14, MMP16, ADAMT1Increased expression in cumulus cells following hCG administration Ovulation and luteogenesisMigratory phenotype of the cumulus–oocyte complex at the time of ovulationRat[82]
MMP19Localized to granulosa and theca-interstitial cells with temporal increases following hCG administrationPreovulatory folliclesECM remodeling and tissue degradationMouse, Rat, Bovineand Human[76,82,83]
MMP2, MMP9, TIMP-1, and TIMP-2The ratio of MMP-2/TIMP-2 decreased in small antral follicles; the MMP-9/TIMP-1 ratio increased in large-preovulatory folliclesPreovulatory folliclesTissue reorganization during ovulationEquine[29]
TIMP-2 and TIMP-3Increased transcript abundance of TIMP-2 in yellow atretic follicles; decreased mRNA expression of TIMP-3Prehierarchical white (WFs), yellowish (YFs), and preovulatory folliclesInvolved in the atresiaChicken[24,79]
TIMP4Increased significantly during the luteinization process of granulosa cellsLocalized to the theca of antral and preovulatory follicles and adjacent ovarian stromaMaintenance of luteal functionMice, Rat[84,85]
tPA and uPAActivity increased during the periovulatory periodGranulosa and theca cellsConversion of plasminogen to plasmin during ovulationRat[66,86,87]
tPA and uPATNFα suppressed FSH-stimulated tPA activity but potentiated FSH-induced uPA activity in undifferentiated granulosa cellsUndifferentiated granulosa cells of preantral and antral folliclesFollicular wall remodeling during ovarian follicular developmentRat[67,88]
PAI-1 and PAI-2mRNAs upregulated after the gonadotrophin surgePAI-1 localized to the thecal layer of preovulatory follicles. PAI-2 localized to the granulosa cell Control plasminogen activator activity associated with ovulation and early corpus luteum formation.Bovine[89]
CTSLExpression increased following hCG administrationOocyte meiosis, Preovulatory to ovulationDegradation of the follicular wallRat, Rhesus monkey, Bovine [54,76]
CTSBExpression increased following hCG administration, Autophagy inductionPreovulatory to ovulationRegulation of follicular developmentMice, Bovine[53,90]
CTSB, K, L, and HExpressed in germinal epithelium throughout the estrous cycleOocytes and granulosa cells of primordial, primary follicles and corpus luteumDegradation of extracellular matrixMice[91,92]
KallikreinsResponse to steroid hormones (androgens and estrogens); various expression patterns with eCG/hCG stimulationPrimordial to ovulationProteolytic processing of growth factors and hormones; angiogenesisRat[93,94]

3.2. Cathepsins

Cathepsins are classified into three main groups based on their catalytic site residues: cysteine cathepsins (Types B, C, F, H, K, L, O, S, V, W, and X), aspartic cathepsins (Types D and E), and serine cathepsins (Type G) [95,96]. In the ovary, cathepsins are expressed in oocytes, granulosa, and theca cells, and their expression is regulated by gonadotropins and growth factors (Table 1) [92].
  • Cathepsin B (CTSB): CTSB can function both as endo- and exo-(carboxy) peptidase [97,98]. CTSB has been identified as a critical regulator of ovarian reserve maintenance in mice [99]. The inhibition of Ctsb by myricetin significantly increased the number of primordial and primary follicles, suggesting a role in follicle activation. This effect seems mediated by the inhibition of autophagy and upregulation of the IGF1R and AKT-mTOR pathways [99]. Similarly, Liang et al. reported that the inhibition of CTSB activity preserved oocyte quality and enhanced developmental competence by mitigating age-related mitochondrial dysfunction and oxidative stress [100]. Chen et al. reported that the silencing of Ctsb in mouse granulosa cells decreased apoptosis by downregulating TNF-α, Casp8, and Casp3 while upregulating Bcl2 expression [53]. Ctsb knockdown also increased granulosa cell proliferation by activating the p-Akt and p-ERK pathways [53]. Komatsu et al. reported that Stefin A, an inhibitor of CTSB, blocked the activation of primordial follicles in mouse newborn ovaries in vitro [101]. In the follicle fluid of pregnant women undergoing ICSI, Bastu et al. found higher levels of CTSB compared to non-pregnant patients [102].
  • Cathepsin L (CTSL): Ctsl is involved in the activation of primordial follicles adjacent to ovulatory follicles, and its inhibition results in a significant reduction in growing follicle numbers [92]. Ctsl expression was detected in large cuboidal cells of small, developing corpora lutea, suggesting possible roles in corpus luteum function [91,103]. Ezz et al. showed that CTSL regulates oocyte meiosis, and its supplementation improves oocyte quality and early embryo development in the bovine [54] (Table 2).
  • Cathepsin S (CTSS): Song et al. reported that Ctss overexpression significantly increased progesterone (P4) and estrogen (E2) production by upregulating Star and Cyp19a1 in rabbit granulosa cells [104,105]. The overexpression of Ctss also increased granulosa cell proliferation while decreasing apoptosis by enhancing the expression of Pcna and Bcl2. Conversely, Ctss knockdown significantly decreased the secretion of P4 and E2 while increasing apoptosis [104].
Table 2. Effects of protease gene knockout or protease inhibition on follicle development in model organisms.
Table 2. Effects of protease gene knockout or protease inhibition on follicle development in model organisms.
ProteaseKO/InhibitorEffect on Follicular DevelopmentSpecific StageMolecular MechanismLocalizationSpeciesReference
MMP1, MMP9, MMP10, and MMP19Inhibitor (GM6001)Reduced ovulation ratePreovulatory to ovulationDegradation of the follicular wallGranulosa and theca cellsRhesus monkey[76]
MMP2Inhibitor (ZK158252)Inhibited hCG-induced ovulation and MMP-2 activationPreovulatory to ovulationLeukotriene B4-receptor antagonismOvarian folliclesRat[106]
MMP1, MMP2, and MMP3Inhibitor (GM6001)Reduction in CL and E2 with GM6001Preovulatory to ovulationInhibits MMP activity in photostimulated ovaries-Hamster[107,108]
MMP10Inhibitor AG1478 Up-regulation of Mmp10 by LH.Ovulation and luteinization.Suppressed the induction of Mmp10 mRNAGranulosa cellsRat [81]
TIMP-1KOIncreased number of primary and preantral folliclesPrimordial to primary/preantralRegulation of MMP activity-Rodent [38,109]
CTSBInhibitor (Myricetin)Increased oocyte reserve Primordial to primaryInhibition of autophagy and upregulation of the IGF1R and AKT-mTOR pathwaysOocytesMouse[99]
CTSLsiRNAEnhanced fertilization capability and blastocyst formationOocytesIncreasing mitochondrial function, reducing accumulated ROS, lowering apoptosis, and recovering lysosome capacityOocytesMouse[91]
CTSLKOReduced ovulation ratePreovulatory to ovulationDegradation of the follicular wallGranulosa and theca cellsMice[48]
CTSLrCTSL supplementationRegulated oocyte meiosis during maturation and early embryo developmentOocyte maturationMeiotic regulationOocytesBovine[54,76]
CTSBStefin ABlocked activation of primordial folliclesPrimordial follicles17β-estradiol increased Stefin A mRNA expression and inhibited follicle development-Mouse[101]
ADAMTS1KOLower numbers of mature follicles and impaired ovulationAntral to ovulationMaintenance of follicular basement membrane integrityGranulosa cellsMouse[110,111]
ADAMTS1KOFailure of ovulation and fertilizationPreovulatory to ovulationExpansion of cumulus–oocyte complexes (COCs)COCsMouse[49,52,112]
ADAMTS9KOOvarian malformation and inability to ovulatePrimordial to ovulation--Zebrafish[113]
LONP1Oocyte-specific KOImpaired follicular development and progressive oocyte deathPrimordial to antralRegulation of mitochondrial functionOocytesMouse[69]
FURINOocyte-specific KOArrested oogenesis at early secondary folliclesPrimary to secondary-OocytesMouse[70]
PAPPAKODecreased litter size and ovulatory capacityAntral to ovulationRegulation of IGF bioavailability-Mouse[71,72,73]
TMPRSS6KORetardation in ovarian maturationPrimordial to antralRegulation of iron homeostasis-Mouse[74]
tPA and uPAInhibitor (PAI-1)Significantly reduced ovulation ratePreovulatory to ovulationECM degradation-Rat, Human[87,114,115,116,117]
PAProtease nexin-1 (SerpinE2)tPA activity higher in cells from small follicles; SerpinE2 levels higher in large folliclesAntral and basal granulosa cellsSerpinE2 secretion regulated at the transcriptional levelGranulosa cellsBovine[41]

4. Role of Proteases in Antral Follicle Development and Ovulation

4.1. Metalloproteinases, Plasminogen Activators, and Their Inhibitors

Ovulation is a complex process involving the rupture of the preovulatory follicle and the release of the oocyte [118,119]. A critical step in this process is the degradation of the basal membrane and ECM surrounding the mature follicle, which are rich in collagen, laminin, and fibronectin [2,57]. MMP2 and MMP9 are expressed in the granulosa and theca cells of preovulatory follicles of both rodents and humans, and their levels increase as the follicles approach ovulation [1,31,34,46,81,118]. Conversely, the inhibition of their activity significantly reduces ovulation rates [31,120,121,122].
Several studies have analyzed changes in Mmp expression following the administration of human chorionic gonadotropin (hCG), which mimics the luteinizing hormone (LH) surge necessary for ovulation. In rhesus monkeys, MMP1 was found to increase following hCG treatment [18]. Mmp11 decreased in both humans and rats [81], whereas Mmp13 increased in bovines [78], but not in humans [83], and Mmp19 increased in mice [63], rats [123], and humans [83]. ADAMTS1 is a secreted metalloproteinase expressed in the granulosa cell layer of mature follicles in the ovary [51,111,112]. LH stimulates the expression of Adamts1 in the granulosa cells of the preovulatory follicles and is sustained in a progesterone-dependent manner [48]. Adamts1-null female mice display lower numbers of growing follicles and impaired ovulation due to mature oocytes remaining trapped within the antral follicles [110,111]. In Adamts1 knockout mice, the development of the ovarian medullary vascular network and lymphatic system were severely delayed, and the expansion of cumulus–oocyte complex (COCs) was reduced, causing lower ovulation and fertilization rates [49,52,112]. These findings highlight a role for Adamts1 in maintaining the structural integrity of follicle basement membranes and supporting lymphangiogenesis, therefore providing new mechanistic insight into ovarian development and disease.
PAs and their inhibitors PAIs are key regulators of the proteolytic cascade involved in ovulation [57,64,124]. During follicle growth, granulosa and theca cells produce tPA and uPA, which facilitate the breakdown of the surrounding basement membrane and the subsequent release of the oocyte [57,59,125,126,127]. Consistent with these observations, PAI-1 expression was reported to decrease during follicle maturation in the human ovary, allowing higher PA activity [128]. Similarly, in rats and porcine ovaries, tPA and uPA activity increased during the periovulatory period, peaking just before ovulation [86,87,114,124,129]. In addition to tPA and uPA, kallikreins are serine proteases that participate in plasminogen activation and promote vascularization [130,131,132,133]. Accumulating evidence suggests that kallikrein expression is regulated by steroid hormones (Table 1). Estrogens stimulate the secretion of KLK1, KLK10, KLK11, and KLK14, whereas androgens trigger the secretion of KLK3 [134,135]. Interestingly, combined stimulation with both androgens and estrogens downregulates KLK3 expression, consistently with the antagonistic effect of estrogens on androgen receptor activity [135,136,137,138]. KLK5 and KLK6 were found to be involved in the proteolytic processing of growth factors and hormones essential for follicular development and ovulation [94,139,140,141].

4.2. Cathepsins

Several studies have reported the involvement of both CTSB and CTSL in regulating ovulation [100,142,143]. In bovine ovaries, Balboula et al. found an increased expression of CTSB in follicle granulosa and theca cells following hCG administration, whereas inhibition of CTSB activity resulted in a significant reduction in ovulation rates [144,145]. The expression of Ctsl increases in the granulosa cells of mouse preovulatory follicles following PMSG and hCG administration [48]. In addition, Sriraman et al. found an increased expression of CTSL in the granulosa cells of human preovulatory follicles, with levels peaking just before ovulation [142]. García et al. proposed that the transient expression of progesterone receptor (PR) in human granulosa cells of the preovulatory follicle may play a role in the activation of CTSL [146]. Interestingly, Zhang et al. demonstrated a significant increase in Ctsl levels in oocytes of aged mice (8–9 months and 11–12 months), and found that the overexpression of Ctsl in the oocytes of young mice (6–8 weeks) substantially diminished their quality, which was restored upon Ctsl inhibition [91].

5. Role of Proteases in Corpus Luteum Formation and Function

Following ovulation, the remainder of the ruptured follicle transforms into the corpus luteum (CL). The CL is a transient endocrine gland that is necessary to sustain pregnancy through the production of progesterone, which prepares the uterus for implantation and pregnancy [147,148,149]. If fertilization does not occur, the CL degenerates leading to the initiation of a new cycle. Dramatic ECM remodeling and angiogenesis are involved in the development, activity, and regression of the CL. MMP2, MMP13, and MMP14 seem to be involved in the regulation of CL dynamics [46,92]. The degradation of the ECM by plasmin allows for the rapid remodeling and vascularization of the CL, which is essential for establishing its function [2,115,135,150]. Wahlberg et al. investigated the formation and function of the CL in plasminogen (Plg)-deficient mice, with or without the administration of galardin, a broad-spectrum synthetic MMP inhibitor [151]. The study revealed that CL formation occurred in Plg-deficient mice, galardin-treated wild-type mice, and galardin-treated Plg-deficient mice, demonstrating that neither the plasminogen activator nor the MMP system is necessary for CL formation. However, serum progesterone levels in Plg-deficient mice were reduced by approximately 50%, and galardin treatment did not further decrease progesterone concentrations. These findings suggest that plasmin, but not MMPs, seems to play a significant role in maintaining luteal function, possibly through the proteolytic activation of growth factors and other paracrine factors [151].

6. Role of Proteases in Ovarian Disease

Due to the critical role proteases play in numerous biological processes, including cell proliferation, migration, and programmed cell death, the dysregulation of their function can lead to several ovarian pathologies (Figure 3).

6.1. Ovarian Cancer

Ovarian cancer is the fifth leading cause of cancer death in females with long-term survival rates of 20% or less in advanced stages III-IV [152,153,154]. The dysregulation of protease activity has been linked to ovarian cancer pathogenesis (Table 3) [33,155,156].
  • MMPs and TIMPs: MMPs participate in several processes that are involved in ovarian cancer progression, including the degradation of the ECM, the promotion of angiogenesis, and the induction of epithelial–mesenchymal transition (EMT) [6,35,46,157,158,159]. Several studies have shown the upregulation of MMPs, such as MMP2 and MMP9, in ovarian cancer tissues compared to normal or benign ovarian tissues, and their expression levels correlate with clinical stage, tumor invasiveness, and metastatic potential [155,157,160,161]. Tumor-derived MMP2 and MMP9 expression has been identified as a negative prognostic indicator in ovarian cancer patients, predicting lower overall survival rates [162,163,164,165,166]. Ovarian cancer cells (Ovcar3) treated with an activator of the PKC pathway, phorbol-12-myristate 13-acetate (PMA), increased MMP7 and MMP10 mRNA [167,168]. MMP14 was shown to activate pro-MMP2 to MMP2, playing a role in the development of vasculogenic-like networks and matrix remodeling by aggressive ovarian cancer cells [168,169,170]. MMP1 activates PAR1, inducing the secretion of angiogenic factors in ovarian carcinoma cells [171]. MMP3 is involved in the estradiol-induced migration and invasion of SKOV3 ovarian cancer cells via the PI3K/Akt/FOXO3 pathway [172]. MMP7 promotes the invasion and metastasis of ovarian cancer cells by activating gelatinases and through the MAPK/ERK and JNK pathways [173,174]. MMP8 upregulates IL-1β, whose expression levels correlate with tumor grade and poor prognosis [175]. The MMP12 82A/G polymorphism has been associated with increased susceptibility to ovarian cancer [176,177], and MMP13 in ascitic fluids of ovarian cancer patients has been identified as a potential marker for disease risk and survival outcomes [178]. Taken together, these studies underline the association between the dysregulation of MMP expression and activity and ovarian cancer.
    In addition to MMPs, several TIMPs, including TIMP1 and TIMP3, have been found upregulated in ovarian cancer [179,180]. However, Davidson et al. found decreased TIMP levels alongside increased MMP2 in ovarian cancer [181]. These seemingly contradictory results highlight the complex mechanisms involved in ovarian cancer and show how dysregulation of the MMP/TIMP balance may have a more significant impact than the overexpression of a single class of proteins [157,182]. In addition, there are several possible explanations for the conflicting findings of increased levels of both MMPs and their inhibitors: (1) TIMPs regulate processes independent of their protease inhibitory activity, including cell growth, migration, and angiogenesis [183]; (2) the stoichiometric balance between MMPs and TIMPs may be more critical than absolute levels, and elevated TIMP levels may sometimes be insufficient to counteract excessive MMP activity in aggressive cancers [184]; (3) TIMPs have been found to activate MMPs in certain instances [19]; and (4) different tissue compartments may have varying MMP ratios, allowing MMPs to remain active in specific microenvironments despite elevated TIMP levels [185].
  • The PA and PAI system: In vitro analyses have shown that uPA is highly expressed in several types of cancer cells, including ovarian cancer [186,187,188,189,190,191,192,193]. The overexpression of uPA and PAI-1 was found in more than 75% of primary ovarian carcinomas, and in most metastatic epithelial ovarian cancer (EOC) [194]. Further, Kenny et al. reported that in vitro and in vivo treatments with a uPA receptor (uPAR) antibody inhibited ovarian cancer cell invasion, migration, and adhesion by inhibiting α5-integrin and decreasing the expression of urokinase, uPAR, β3-integrin, and fibroblast growth factor receptor-1 [195]. High levels of PAI-1 have been associated with poor clinical outcomes in ovarian serous carcinoma [187,196]. In ovarian cancer cells, PAI-1 inhibition resulted in cell cycle arrest and decreased proliferation, and, in xenograft models, significantly reduced peritoneal dissemination [196]. Similarly, PAI-1 silencing in SKOV3 cells disrupted the platelet-induced upregulation of the genes involved in proliferation and ECM remodeling [197]. At the molecular level, some reports suggest that PAI-1 inhibits cell adhesion and migration by blocking vitronectin (VN) binding to integrins or by displacing uPAR from VN in the extracellular matrix [198,199]. However, other studies have shown that PAI-1 can enhance cancer cell adhesion [200,201].
    As for MMPs/TIMPs, it is unclear why the upregulation of both uPA and PAIs correlates with cancer progression and poor clinical outcomes. Several mechanisms may explain this apparent contradiction: (1) PAI have additional functions beyond uPA inhibition, including activation of pathways that promote tumor growth, angiogenesis, and cell detachment [202,203,204]; (2) the PAI-1/uPA/uPAR complex can be internalized and recycled, potentially leading to increased uPAR on the cell surface and enhanced invasiveness [205]; (3) PAI-1 can elicit inflammatory responses and immune cell recruitment in the tumor microenvironment, potentially promoting a pro-tumorigenic milieu [206]. Overall, these findings suggest that PAI function is context-dependent and highlight the complex regulation of the PA/PAI system in ovarian cancer.
  • Cathepsins: Cathepsins and their inhibitors cystatins have also been associated with ovarian cancer. Liu et al. showed that CTSB and its binding proteins AMBP and TSRC1 modulated TNF-induced apoptosis in ovarian cancer cells [207]. Additionally, Ctsl knockdown inhibited proliferation, invasion, and tumor growth both in vitro and in vivo, while Ctsl overexpression had the opposite effects [208,209]. In malignant serous tumors, cystic fluid levels of CTSB, CTSL, and their inhibitor Cystatin C (Cst3) were significantly elevated compared to benign serous tumors [210]. Gashenko et al. found significantly increased levels of procathepsin B, cystatin B (CstB), and Cst3 in serum and ascite fluids of ovarian cancer patients compared to controls, suggesting their possible use as disease biomarkers [211]. Nishikawa et al. found significantly elevated levels of Cst3, but not CTSB, in ovarian cancer compared to benign samples and healthy controls [212]. Interestingly, invasion assays showed that the inhibition of Cst3 or CTSB suppressed cancer cell invasion in a dose-dependent manner [212]. Once again, some of these findings appear contradictory. Elevated CysC in cancer may represent a compensatory mechanism to control excessive cathepsin activity [213]. In addition, similar to other protease inhibitors, Cst3 may have additional functions including the regulation of immune response and cell signaling [214]. Furthermore, changes in the balance between cathepsins and cystatins may be more important than absolute expression levels of either protein [215]. Finally. Cst3 primarily regulates extracellular cathepsin activity, while pro-tumorigenic effects of CTSB may be, at least in part, intracellular [216].
Xu et al. reported elevated levels of serum CTSK in ovarian cancer patients compared to healthy controls [217]. CTSD was also found associated with ovarian cancer, with epithelial CTSD expression more common (65.1%) in ovarian tumors with low malignant potential (LMP) compared to invasive tumors (43.7%) [218,219]. CTSD was also found to significantly increase the proliferation and migration of human omental microvascular endothelial cells, and the knockdown of CTSD was exhibited to be able to suppress migration and invasion in an EOC cell line [220,221]. In addition, the inhibition of CTSS induced apoptosis in cancer cells by downregulating Bcl-2 and c-FLIP [222].
Another serine protease, Hepsin is overexpressed in several cancers including ovarian cancer, but its specific role remains to be elucidated [223,224,225,226]. Miao et al. reported that membrane-associated Hepsin localized at desmosomal junctions with its putative proteolytic substrate hepatocyte growth factor (HGF), and showed that Hepsin overexpression promoted ovarian tumor growth in a mouse model [227]. The dysregulation of Hepsin expression could disrupt the integrity and function of epithelial barriers. In addition, the Hepsin-mediated cleavage of substrates including pro-HGF and pro-uPA could promote metastasis [228,229]. Despite a role in mediating cancer formation and progression, sosme reports suggest that high levels of Hepsin could have antitumor effects by reducing oncogenic signaling and increasing autophagy [230]. It is possible that specific functions in cancer depend on the microenvironment or disease stage.
Table 3. Proteases in ovarian cancer.
Table 3. Proteases in ovarian cancer.
ProteaseFinding in Ovarian CancerLocalizationSpeciesPrognostic ValueRoleReference
MMP1Activates PAR1Ovarian carcinoma cellsHuman, Epithelial ovary cell linesNot reportedInduces the secretion of angiogenic factors[231,232,233,234]
MMP2Upregulated in ovarian cancer tissues compared to normal/benign ovarian tissuesOvarian cancer tissue, epithilial, stroma HumanNegative prognostic indicator with lower overall survival ratesDegrades ECM, promotes angiogenesis, and induces EMT[161,162,163,182,235,236,237]
MMP3Involved in estradiol-induced migration and invasionSKOV3 ovarian cancer cellsHuman cell lineNot reportedMediates estrogen-induced cancer progression via PI3K/Akt/FOXO3 pathway[172]
MMP7Promotes invasion and metastasisOvarian cancer cellsHuman cell lineNot reportedActs through MAPK/ERK and JNK pathways; activates gelatin enzymes[174]
MMP8Upregulates IL-1βOvarian cancer tissueHumanCorrelates with tumor grade and poor prognosisPromotes inflammatory microenvironment[175]
MMP9Upregulated in ovarian cancer tissuesOvarian cancer tissueHumanCorrelates with clinical stage, tumor invasiveness, and metastatic potentialDegrades ECM, promotes angiogenesis, and induces EMT; cleaves fibronectin and type IV collagen[23,165,165,238,239]
MMP10Increased expression with PKC pathway activationOvarian cancer cells (Ovcar3, EOC)Human cell lineNot reportedRegulated by PKC pathway, Wnt signaling [167,240]
MMP11Overexpression in stromal cellsOvarian carcinomas Human Overexpression not correlates with survival.Tumor progression[241]
MMP1282 A/G polymorphism associated with increased susceptibilityGenetic studyHumanNot reportedGenetic predisposition factor[176,177]
MMP13Elevated in ascitic fluidsAscitic fluidHumanPotential marker for disease risk and survival outcomesNot reported[178,179]
MMP14Activates pro-MMP2 to MMP2Ovarian cancer cellsHumanAssociated with vasculogenic-like networksMatrix remodeling; activates pro-MMP2[82,168,242]
TIMP1Upregulated in ovarian cancerOvarian cancer tissueHumanNot reportedComplex: May have MMP-independent roles in cell growth, migration, and angiogenesis[180,181]
TIMP3Upregulated in ovarian cancerOvarian cancer tissueHumanNot reportedComplex regulatory roles beyond MMP inhibition[243,244,245]
uPAHighly expressed in cancer cells; overexpressed in >75% of primary ovarian carcinomas and metastatic EOC samplesOvarian cancer cellsHuman and cell linesAssociated with invasion and metastasisPromotes invasion, migration, and adhesion[187,188,189,190,191]
PAI-1High levels in ovarian serous carcinomaOvarian cancer tissueHuman, cell lines, and xenograft modelsAssociated with poor clinical outcomesComplex: Inhibits cell adhesion by blocking vitronectin; disrupts platelet-induced gene upregulation[189,197,198]
CTSBModulates TNF-induced apoptosis; elevated in cystic fluid and serumCystic fluid, serum, and cancer cellsHumanSerum procathepsin B significantly elevated compared to healthy controlsBinding proteins AMBP and TSRC1 involved in TNF-induced apoptosis[210,211,212]
CTSLOverexpressed; knockdown inhibits proliferation, invasion, and tumor growthCancer cellsHuman, cell lines, and mouse modelsAssociated with paclitaxel resistancePromotes proliferation and migration; confers chemoresistance[208,209]
CTSSInhibition stimulates TRAIL-induced apoptosisCancer cellsHumanNot reportedDownregulation of Bcl-2 and Cbl-mediated c-FLIP by ROS-mediated p53 expression[222]
Cst3Elevated in malignant tissues, serum, and cystic fluidMalignant tissue, Serum, Cystic fluidHumanElevated in ovarian cancer compared to benign samplesComplex: May represent failed compensatory mechanism; has additional immune and signaling functions[210,212,213,214,215]
CTSKOverexpressed in peritoneal metastatic ovarian carcinomas; elevated serum levelsPeritoneal metastases, SerumHumanPotential biomarkerAssociated with peritoneal metastasis[217,246]
CTSDExpression more common (65.1%) in tumors with low malignant potential vs. invasive tumors (43.7%); promotes the proliferation and migration of endothelial cellsEpithelial cells, Stromal cellsHuman and cell linesIndependent prognostic factor for disease-free survival in invasive ovarian cancerPro-angiogenic and pro-metastatic role via ERK1/2 and AKT activation; correlates with microvessel density[218,219,220,221]
HepsinOverexpressed in ovarian cancerDesmosomal junctionsHuman and mouse modelNot reportedCleaves HGF and pro-uPA; localizes with substrate HGF; disrupts epithelial barriers[223,224,225,227,229]

6.2. Polycystic Ovary Syndrome (PCOS)

PCOS is an endocrine disorder characterized by ovulatory dysfunction, excess of androgen, and polycystic ovaries, often leading to infertility, insulin resistance, and cardiovascular disease [247,248,249,250,251].
  • MMPs and TIMPs: MMPs and TIMPs have been associated with the pathogenesis of PCOS (Table 4) [252]. It has been reported that MMP2 and MMP9 concentrations are elevated in the follicular fluid of patients with PCOS compared to healthy controls [252,253]. The increased MMP activity was associated with higher levels of androgens, insulin resistance, disrupted follicular development, and ovulatory dysfunction [6,253,254]. Consistent with these observations, it was found that the granulosa cells of women with PCOS express fewer MMP inhibitors TIMP-1 and TIMP-2 compared to healthy controls [252,255]. Recently, Butler et al. reported that women with PCOS showed significantly elevated MMP9 [254]. Interestingly, the ratios of MMP9 to all TIMPs were significantly higher in the PCOS group, while MMP17/TIMP-1 and MMP17/TIMP-2 were lower. Higher expression of Mmp2/9 was also observed in antral follicles compared to the preantral follicle and primordial follicle of a Letrozole-induced PCOS rat model [256].
  • The PA and PAI system: Elevated PAI-1 levels in plasma have been reported in patients with PCOS compared to controls (Table 4) [257,258,259,260,261,262]. However, findings regarding PAI-1 distribution within ovarian tissue have been inconsistent. Devin et al. reported increased PAI-1 in granulosa cells of cystic and atretic follicles in mouse models of PCOS [263]. Atiomo et al. detected PAI-1 in granulosa and theca cells without significant differences between PCOS and control ovaries, whereas other authors reported increased PAI-1 expression in follicular fluid from patients with PCOS [257,264,265]. Genetic predisposition seems to contribute to PAI-1 dysregulation in PCOS, which has been reported associated with the 4G/4G and 4G/5G genotypic subtypes in the PAI-1 promoter region, leading to increased protein levels [266].
    Kelly et al. observed increased tPA antigen levels inversely correlating with insulin resistance, whereas Tarkun et al. found a direct correlation of PAI-1 levels, even in lean PCOS women [197,259]. Orio et al. reported elevated PAI-1 activity independent of obesity, while Sahay et al. found a correlation with both insulin resistance and obesity [260,267]. Ma et al. provided mechanistic insight through a mouse model demonstrating that PAI-1 deficiency prevented diet-induced obesity and insulin resistance [268,269]. Finally, Ibrahim et al. reported the presence of KLK2 in the serum of women with PCOS in association with hirsutism, but the nature of this relationship remains unclear [270].
  • Cathepsins: The downregulation of CTSD has been reported in the ovaries of patients with PCOS [271]. CTSD downregulation may contribute to the abnormal follicle development associated with PCOS, leading to ovulatory dysfunction and infertility. Dawood et al. found significantly increased levels of CTSS, among patients with PCOS compared to healthy females [272]. Additionally, genetics may also play a role as CTSB polymorphisms have recently been associated with PCOS risks [273].
Table 4. Protease dysregulation in PCOS.
Table 4. Protease dysregulation in PCOS.
ProteaseFinding in PCOSLocalizationSpeciesProposed Pathogenic RoleReference
MMP2
And MMP9
Elevated in follicular fluid of patients with PCOS compared to healthy controlsFollicular fluid and serumHumanAssociated with higher levels of androgens, insulin resistance, disrupted follicular development, and ovulatory dysfunction.
Associated with higher MMP9/TIMP ratios, ECM remodeling, and follicular development
[252,253,256]
MMP2/9Higher expression in antral follicles compared to preantral and primordial folliclesOvarian folliclesRat (Letrozole-induced PCOS model)ECM remodeling in PCOS ovaries[256]
MMP17Lower MMP17/TIMP-1 and MMP17/TIMP-2 ratios in PCOSSerumHumanECM remodeling[254]
TIMP-1 and TIMP-2Decreased expression in granulosa cells of women with PCOSGranulosa cellsHumanExcessive ECM degradation[255,274]
PAI-1Elevated in plasma, granulosa cells, and follicular fluid; homogeneous distribution throughout PCOS ovariesPlasma, granulosa cells, follicular fluid, and theca cellsHuman and mouse (PCOS model)Associated with insulin resistance; higher expression in 4G/4G and 4G/5G genotypes[260,262,263,264,266,267,268,269]
PlasminogenUniquely present in small follicles of PCOS ovariesSmall folliclesHumanAltered proteolytic activity in early follicular development[258,259,261]
tPAIncreased antigen levels inversely correlating with insulin resistancePlasmaHumanAssociated with insulin resistance[275]
KLK2/3Present in the serum of women with PCOS in association with hirsutismSerumHumanAssociated with androgsen excess and hirsutism[270]
CTSBCTSB polymorphisms contribute to PCOS pathogenesisBloodHumanrs12898, rs8898, and rs3779659 variants associated with PCOS risk[273]
CTSDDownregulated in ovaries of patients with PCOSCytoplasm and cell membrane of stromal and granulosa cellsHumanAbnormal follicle development[271]
CTSSSignificantly increased levels in patients with PCOSSerumHumanInflammation associated with PCOS[272]

6.3. Primary Ovarian Insufficiency (POI)

Also known as premature ovarian failure (POF), POI is a condition characterized by the loss of ovarian function before the age of 40 [276,277]. The ovaries of mice with chemotherapy-induced POI showed increased expression of MMP2 and MMP9, and downregulation of TIMP-1 and TIMP-2 compared to control mice [278]. The greater activity of MMPs was associated with the increased apoptosis of the granulosa cells and abnormal folliculogenesis in the POI mice [278]. An et al. reported that the TIMP2G  >  C (rs8179090) and TIMP2G  >  A (rs2277698) genotypes were strongly associated with POI [279]. Pathogenic variants of serine protease LONP1 have been found associated with POI [69]. Affected patients lack large antral follicles and are infertile, but the molecular mechanisms remain unknown, and functional analyses have not been performed [69]. PAs and PAIs are also associated with POI, particularly regarding the accelerated depletion of the ovarian reserve. Similarly to MMPs, tPA activity was higher in the ovaries of mice with chemotherapy-induced POI and associated with the increased apoptosis of granulosa cells and accelerated depletion of the ovarian reserve [280].

7. Conclusions

Proteases play crucial roles in several processes of ovarian function, including folliculogenesis, ovulation, and corpus luteum development. The balance between proteases and their inhibitors is essential for maintaining normal ovarian function. MMPs, PAs, and cathepsins contribute to the degradation of the follicular basement membrane and the extracellular matrix, enabling follicle activation, maturation, and oocyte release. The dysregulation of protease activity has been associated with several ovarian disorders, including cancer, PCOS, and POI. However, despite extensive evidence of the important role proteases play in ovarian functions, specific molecular mechanisms remain elusive. As proteases represent promising targets for therapeutic applications, future investigations should focus on identifying mechanisms of action and relationships with critical signaling pathways of the ovary.

Author Contributions

Conceptualization—E.P. and B.K.; data curation—B.K.; writing of MS—E.P. and B.K., original draft preparation—B.K.; review and editing—E.P. and B.K.; supervision—E.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable. All the figures were created using BioRender (Kushawaha, B. 2025 https://BioRender.com/3htq4d3 accessed on 25 March 2025).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AbbreviationFull Form
ADAMTSA Disintegrin and Metalloproteinase with Thrombospondin-like Motifs
AIFM1Apoptosis-Inducing Factor Mitochondrion-associated 1
AKTProtein Kinase B
AMBPAlpha-1-Microglobulin/Bikunin Precursor
ATG5Autophagy Related 5
BCL2B-Cell Lymphoma 2
CASP3/8Caspase 3/8
CLCorpus Luteum
COCCumulus–Oocyte Complex
CTSBCathepsin B
CTSDCathepsin D
CTSECathepsin E
CTSGCathepsin G
CTSKCathepsin K
CTSLCathepsin L
CTSSCathepsin S
CstBCystatin B
Cst3Cystatin C
CYP17A1Cytochrome P450 Family 17 Subfamily A Member 1
CYP19A1Cytochrome P450 Family 19 Subfamily A Member 1
DNADeoxyribonucleic Acid
DpcDays Post Coitus
DppDays Postpartum
E2Estradiol
ECMExtracellular Matrix
EOCEpithelial Ovarian Cancer
ERKExtracellular Signal-Regulated Kinase
FAKFocal Adhesion Kinase
FFFollicular Fluid
FGF1Fibroblast Growth Factor 1
FSHFollicle-Stimulating Hormone
FURINPaired Basic Amino Acid Cleaving Enzyme
GCNAGerm Cell Nuclear Antigen
GDF9Growth Differentiation Factor 9
hCGHuman Chorionic Gonadotropin
HGFHepatocyte Growth Factor
ICSIIntracytoplasmic Sperm Injection
IGFInsulin-like Growth Factor
IGF1RInsulin-like Growth Factor 1 Receptor
IL-1βInterleukin 1 Beta
JNKc-Jun N-terminal Kinase
KLKKallikrein
LC3-IMicrotubule-associated Protein 1A/1B-Light Chain 3
LHLuteinizing Hormone
LONP1Lon Peptidase 1
LMPLow Malignant Potential
MAPKMitogen-Activated Protein Kinase
MMPMatrix Metalloproteinase
mRNAMessenger Ribonucleic Acid
mTORMammalian Target of Rapamycin
MYCMyelocytomatosis Oncogene
NFkBNuclear Factor Kappa B
P4Progesterone
PAPlasminogen Activator
PAIPlasminogen Activator Inhibitor
PAPPAPregnancy-Associated Plasma Protein A
PAR1Protease-Activated Receptor 1
PCNAProliferating Cell Nuclear Antigen
PCOSPolycystic Ovary Syndrome
PI3KPhosphoinositide 3-Kinase
PKCProtein Kinase C
PMAPhorbol-12-myristate 13-acetate
PMSGPregnant Mare Serum Gonadotropin
POF/POIPremature Ovarian Failure/Primary Ovarian Insufficiency
PRProgesterone Receptor
RGDArg-Gly-Asp (Arginine–Glycine–Aspartic acid)
ROSReactive Oxygen Species
SECSecurities and Exchange Commission
siRNASmall Interfering RNA
SMADSmall Mothers Against Decapentaplegic
SPINK1Serine Protease Inhibitor Kazal Type 1
STARSteroidogenic Acute Regulatory Protein
TGFαTransforming Growth Factor Alpha
TGF-βTransforming Growth Factor Beta
TIMPTissue Inhibitor of Metalloproteinases
TMPRSS6Transmembrane Serine Protease 6 (Matriptase-2)
TNF-αTumor Necrosis Factor Alpha
tPATissue-type Plasminogen Activator
TRAILTNF-Related Apoptosis-Inducing Ligand
TSRC1Thrombospondin and Calcium-binding domains 1
uPAUrokinase-type Plasminogen Activator
uPARUrokinase-type Plasminogen Activator Receptor
VEGFVascular Endothelial Growth Factor
VNVitronectin
ZMP-2Zinc Metalloproteinase-2
ZP3Zona Pellucida Glycoprotein 3

References

  1. Curry, T.E., Jr.; Osteen, K.G. Cyclic Changes in the Matrix Metalloproteinase System in the Ovary and Uterus. Biol. Reprod. 2001, 64, 1285–1296. [Google Scholar] [CrossRef]
  2. Curry, T.E., Jr.; Osteen, K.G. The Matrix Metalloproteinase System: Changes, Regulation, and Impact throughout the Ovarian and Uterine Reproductive Cycle. Endocr. Rev. 2003, 24, 428–465. [Google Scholar] [CrossRef] [PubMed]
  3. Rodgers, R.J.; Irving-Rodgers, H.F.; Russell, D.L. Extracellular matrix of the developing ovarian follicle. Reproduction 2003, 126, 415–424. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Y.-X. Plasminogen activator/plasminogen activator inhibitors in ovarian physiology. Front. Biosci. J. Virtual Libr. 2004, 9, 3356–3373. [Google Scholar] [CrossRef]
  5. Levene, P.A. The Cleavage Products of Proteoses. J. Biol. Chem. 1905, 1, 45–58. [Google Scholar] [CrossRef]
  6. Goldman, S. MMPS and TIMPS in ovarian physiology and pathophysiology. Front. Biosci. 2004, 9, 2474. [Google Scholar] [CrossRef] [PubMed]
  7. Vu, T.H.; Werb, Z. Matrix metalloproteinases: Effectors of development and normal physiology. Genes Dev. 2000, 14, 2123–2133. [Google Scholar] [CrossRef]
  8. Nagase, H.; Visse, R.; Murphy, G. Structure and function of matrix metalloproteinases and TIMPs. Cardiovasc. Res. 2006, 69, 562–573. [Google Scholar] [CrossRef]
  9. Hey, S.; Linder, S. Matrix metalloproteinases at a glance. J. Cell Sci. 2024, 137, jcs261898. [Google Scholar] [CrossRef]
  10. Page, M.J.; Di Cera, E. Serine peptidases: Classification, structure and function. Cell. Mol. Life Sci. 2008, 65, 1220–1236. [Google Scholar] [CrossRef]
  11. Levin, G.; Coelho, T.M.; Nóbrega, N.G.; Trombetta-Lima, M.; Sogayar, M.C.; Carreira, A.C.O. Spatio-temporal expression profile of matrix metalloproteinase (Mmp) modulators Reck and Sparc during the rat ovarian dynamics. Reprod. Biol. Endocrinol. RBE 2018, 16, 116. [Google Scholar] [CrossRef] [PubMed]
  12. Robinson, L.L.L.; Sznajder, N.A.; Riley, S.C.; Anderson, R.A. Matrix metalloproteinases and tissue inhibitors of metalloproteinases in human fetal testis and ovary. Mol. Hum. Reprod. 2001, 7, 641–648. [Google Scholar] [CrossRef]
  13. Miyakoshi, K.; Murphy, M.J.; Yeoman, R.R.; Mitra, S.; Dubay, C.J.; Hennebold, J.D. The Identification of Novel Ovarian Proteases Through the Use of Genomic and Bioinformatic Methodologies. Biol. Reprod. 2006, 75, 823–835. [Google Scholar] [CrossRef]
  14. Murphy, G.; Nagase, H. Progress in matrix metalloproteinase research. Mol. Aspects Med. 2008, 29, 290–308. [Google Scholar] [CrossRef]
  15. Almutairi, S.; Kalloush, H.M.; Manoon, N.A.; Bardaweel, S.K. Matrix Metalloproteinases Inhibitors in Cancer Treatment: An Updated Review (2013–2023). Molecules 2023, 28, 5567. [Google Scholar] [CrossRef]
  16. Verma, R.P.; Hansch, C. Matrix metalloproteinases (MMPs): Chemical–biological functions and (Q)SARs. Bioorg. Med. Chem. 2007, 15, 2223–2268. [Google Scholar] [CrossRef] [PubMed]
  17. Donepudi, M.; Grütter, M.G. Structure and zymogen activation of caspases. Biophys. Chem. 2002, 101–102, 145–153. [Google Scholar] [CrossRef] [PubMed]
  18. Chaffin, C.L.; Stouffer, R.L. Expression of Matrix Metalloproteinases and Their Tissue Inhibitor Messenger Ribonucleic Acids in Macaque Periovulatory Granulosa Cells: Time Course and Steroid Regulation1. Biol. Reprod. 1999, 61, 14–21. [Google Scholar] [CrossRef]
  19. Lambert, E.; Dassé, E.; Haye, B.; Petitfrère, E. TIMPs as multifacial proteins. Crit. Rev. Oncol. Hematol. 2004, 49, 187–198. [Google Scholar] [CrossRef]
  20. Moracho, N.; Learte, A.I.R.; Muñoz-Sáez, E.; Marchena, M.A.; Cid, M.A.; Arroyo, A.G.; Sánchez-Camacho, C. Emerging roles of MT-MMPs in embryonic development. Dev. Dyn. 2022, 251, 240–275. [Google Scholar] [CrossRef]
  21. Moore, C.S.; Crocker, S.J. An Alternate Perspective on the Roles of TIMPs and MMPs in Pathology. Am. J. Pathol. 2012, 180, 12–16. [Google Scholar] [CrossRef]
  22. Rundhaug, J.E. Matrix metalloproteinases and angiogenesis. J. Cell. Mol. Med. 2005, 9, 267–285. [Google Scholar] [CrossRef] [PubMed]
  23. Belotti, D.; Paganoni, P.; Manenti, L.; Garofalo, A.; Marchini, S.; Taraboletti, G.; Giavazzi, R. Matrix Metalloproteinases (MMP9 and MMP2) Induce the Release of Vascular Endothelial Growth Factor (VEGF) by Ovarian Carcinoma Cells: Implications for Ascites Formation1. Cancer Res. 2003, 63, 5224–5229. [Google Scholar] [PubMed]
  24. Wolak, D.; Hrabia, A. Alternations in the expression of selected matrix metalloproteinases (MMP-2, -9, -10, and -13) and their tissue inhibitors (TIMP-2 and -3) and MMP-2 and -9 activity in the chicken ovary during pause in laying induced by fasting. Theriogenology 2021, 161, 176–186. [Google Scholar] [CrossRef] [PubMed]
  25. Nilsson, E.; Skinner, M.K. Cellular Interactions That Control Primordial Follicle Development and Folliculogenesis. J. Soc. Gynecol. Investig. JSGI 2001, 8, S17–S20. [Google Scholar] [CrossRef]
  26. Jabłońska-Trypuć, A.; Matejczyk, M.; Rosochacki, S. Matrix metalloproteinases (MMPs), the main extracellular matrix (ECM) enzymes in collagen degradation, as a target for anticancer drugs. J. Enzyme Inhib. Med. Chem. 2016, 31, 177–183. [Google Scholar] [CrossRef]
  27. Pepling, M.E. From primordial germ cell to primordial follicle: Mammalian female germ cell development. Genesis 2006, 44, 622–632. [Google Scholar] [CrossRef]
  28. Mazaud, S.; Guyot, R.; Guigon, C.J.; Coudouel, N.; Le Magueresse-Battistoni, B.; Magre, S. Basal membrane remodeling during follicle histogenesis in the rat ovary: Contribution of proteinases of the MMP and PA families. Dev. Biol. 2005, 277, 403–416. [Google Scholar] [CrossRef]
  29. Sessions, D.R.; Vick, M.M.; Fitzgerald, B.P. Characterization of matrix metalloproteinase-2 and matrix metalloproteinase-9 and their inhibitors in equine granulosa cells in vivo and in vitro. J. Anim. Sci. 2009, 87, 3955–3966. [Google Scholar] [CrossRef]
  30. Duncan, W.C.; McNeilly, A.S.; Illingworth, P.J. The Effect of Luteal “Rescue” on the Expression and Localization of Matrix Metalloproteinases and Their Tissue Inhibitors in the Human Corpus Luteum. J. Clin. Endocrinol. Metab. 1998, 83, 2470–2478. [Google Scholar] [CrossRef]
  31. Bagavandoss, P. Differential distribution of gelatinases and tissue inhibitor of metalloproteinase-1 in the rat ovary. J. Endocrinol. 1998, 158, 221–228. [Google Scholar] [CrossRef] [PubMed]
  32. Robinson, R.S.; Woad, K.J.; Hammond, A.J.; Laird, M.; Hunter, M.G.; Mann, G.E. Angiogenesis and vascular function in the ovary. Reproduction 2009, 138, 869–881. [Google Scholar] [CrossRef] [PubMed]
  33. Downs, L.S., Jr.; Lima, P.H.; Bliss, R.L.; Blomquist, C.H. Cathepsins B and D Activity and Activity Ratios in Normal Ovaries, Benign Ovarian Neoplasms, and Epithelial Ovarian Cancer. J. Soc. Gynecol. Investig. 2005, 12, 539–544. [Google Scholar] [CrossRef]
  34. Smith, M.F.; McIntush, E.W.; Ricke, W.A.; Kojima, F.N.; Smith, G.W. Regulation of ovarian extracellular matrix remodelling by metalloproteinases and their tissue inhibitors: Effects on follicular development, ovulation and luteal function. J. Reprod. Fertil. Suppl. 1999, 54, 367–381. [Google Scholar] [CrossRef] [PubMed]
  35. Smith, M.F.; Ricke, W.A.; Bakke, L.J.; Dow, M.P.D.; Smith, G.W. Ovarian tissue remodeling: Role of matrix metalloproteinases and their inhibitors. Mol. Cell. Endocrinol. 2002, 191, 45–56. [Google Scholar] [CrossRef]
  36. Nothnick, W.B. Reduction in reproductive lifespan of tissue inhibitor of metalloproteinase 1 (TIMP-1)-deficient female mice. Reproduction 2001, 122, 923–927. [Google Scholar] [CrossRef]
  37. Nothnick, W.B. Tissue Inhibitor of Metalloproteinase-1 (TIMP-1) Deficient Mice Display Reduced Serum Progesterone Levels during Corpus Luteum Development. Endocrinology 2003, 144, 5–8. [Google Scholar] [CrossRef]
  38. Stilley, J.A.W.; Birt, J.A.; Nagel, S.C.; Sutovsky, M.; Sutovsky, P.; Sharpe-Timms, K.L. Neutralizing TIMP1 Restores Fecundity in a Rat Model of Endometriosis and Treating Control Rats with TIMP1 Causes Anomalies in Ovarian Function and Embryo Development1. Biol. Reprod. 2010, 83, 185–194. [Google Scholar] [CrossRef]
  39. Curry, T.E.; Dean, D.D.; Sanders, S.L.; Pedigo, N.G.; Jones, P.B.C. The role of ovarian proteases and their inhibitors in ovulation. Steroids 1989, 54, 501–521. [Google Scholar] [CrossRef]
  40. Bédard, J.; Brûlé, S.; Price, C.A.; Silversides, D.W.; Lussier, J.G. Serine protease inhibitor-E2 (SERPINE2) is differentially expressed in granulosa cells of dominant follicle in cattle. Mol. Reprod. Dev. 2003, 64, 152–165. [Google Scholar] [CrossRef]
  41. Cao, M.; Sahmi, M.; Lussier, J.G.; Price, C.A. Plasminogen activator and serine protease inhibitor-E2 (protease nexin-1) expression by bovine granulosa cells in vitro. Biol. Reprod. 2004, 71, 887–893. [Google Scholar] [CrossRef] [PubMed]
  42. Oh, H.S.; Kim, T.; Gu, D.-H.; Lee, T.S.; Kim, T.H.; Shin, S.; Shin, B.S. Pharmacokinetics of Nafamostat, a Potent Serine Protease Inhibitor, by a Novel LC-MS/MS Analysis. Molecules 2022, 27, 1881. [Google Scholar] [CrossRef] [PubMed]
  43. Murdoch, W.J.; Gottsch, M.L. Proteolytic Mechanisms in the Ovulatory Folliculo-Luteal Transformation. Connect. Tissue Res. 2003, 44, 50–57. [Google Scholar] [CrossRef]
  44. Eykelbosh, A.J.; Van Der Kraak, G. A role for the lysosomal protease cathepsin B in zebrafish follicular apoptosis. Comp. Biochem. Physiol. A. Mol. Integr. Physiol. 2010, 156, 218–223. [Google Scholar] [CrossRef]
  45. Li, J.R.; Shen, T. Expression characteristics of MMP-2 and MMP-9 in guinea pig ovaries during the estrous cycle. Genet. Mol. Res. 2015, 14, 17329–17340. [Google Scholar] [CrossRef]
  46. Vos, M.C.; van der Wurff, A.A.; Last, J.T.; de Boed, E.A.; Smeenk, J.M.; van Kuppevelt, T.H.; Massuger, L.F. Immunohistochemical expression of MMP-14 and MMP-2, and MMP-2 activity during human ovarian follicular development. Reprod. Biol. Endocrinol. 2014, 12, 12. [Google Scholar] [CrossRef] [PubMed]
  47. Kato, N.; Motoyama, T. Relation Between Laminin-5 γ2 Chain and Cell Surface Metalloproteinase MT1-MMP in Clear Cell Carcinoma of the Ovary. Int. J. Gynecol. Pathol. 2009, 28, 49. [Google Scholar] [CrossRef]
  48. Robker, R.L.; Russell, D.L.; Espey, L.L.; Lydon, J.P.; O’Malley, B.W.; Richards, J.S. Progesterone-regulated genes in the ovulation process: ADAMTS-1 and cathepsin L proteases. Proc. Natl. Acad. Sci. USA 2000, 97, 4689–4694. [Google Scholar] [CrossRef]
  49. Brown, H.M.; Dunning, K.R.; Robker, R.L.; Pritchard, M.; Russell, D.L. Requirement for ADAMTS-1 in extracellular matrix remodeling during ovarian folliculogenesis and lymphangiogenesis. Dev. Biol. 2006, 300, 699–709. [Google Scholar] [CrossRef]
  50. Russell, D.L.; Doyle, K.M.H.; Ochsner, S.A.; Sandy, J.D.; Richards, J.S. Processing and localization of ADAMTS-1 and proteolytic cleavage of versican during cumulus matrix expansion and ovulation. J. Biol. Chem. 2003, 278, 42330–42339. [Google Scholar] [CrossRef]
  51. Russell, D.L.; Brown, H.M.; Dunning, K.R. ADAMTS proteases in fertility. Matrix Biol. 2015, 44–46, 54–63. [Google Scholar] [CrossRef] [PubMed]
  52. Brown, H.M.; Dunning, K.R.; Robker, R.L.; Boerboom, D.; Pritchard, M.; Lane, M.; Russell, D.L. ADAMTS1 Cleavage of Versican Mediates Essential Structural Remodeling of the Ovarian Follicle and Cumulus-Oocyte Matrix During Ovulation in Mice1. Biol. Reprod. 2010, 83, 549–557. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, C.; Ahmad, M.J.; Ye, T.; Du, C.; Zhang, X.; Liang, A.; Yang, L. Cathepsin B Regulates Mice Granulosa Cells’ Apoptosis and Proliferation In Vitro. Int. J. Mol. Sci. 2021, 22, 11827. [Google Scholar] [CrossRef]
  54. Ezz, M.A.; Takahashi, M.; Rivera, R.M.; Balboula, A.Z. Cathepsin L regulates oocyte meiosis and preimplantation embryo development. Cell Prolif. 2023, 57, e13526. [Google Scholar] [CrossRef]
  55. Nothnick, W.B.; Soloway, P.; Curry, T.E., Jr. Assessment of the Role of Tissue Inhibitor of Metalloproteinase-1 (TIMP-1) during the Periovulatory Period in Female Mice Lacking a Functional TIMP-1 Gene1. Biol. Reprod. 1997, 56, 1181–1188. [Google Scholar] [CrossRef] [PubMed]
  56. Fata, J.E.; Ho, A.T.-V.; Leco, K.J.; Moorehead, R.A.; Khokha*, R. Cellular turnover and extracellular matrix remodeling in female reproductive tissues: Functions of metalloproteinases and their inhibitors. Cell. Mol. Life Sci. CMLS 2000, 57, 77–95. [Google Scholar] [CrossRef]
  57. Ny, T.; Wahlberg, P.; Brändström, I.J.M. Matrix remodeling in the ovary: Regulation and functional role of the plasminogen activator and matrix metalloproteinase systems. Mol. Cell. Endocrinol. 2002, 187, 29–38. [Google Scholar] [CrossRef]
  58. Bodén, I. The Roles of the Plasminogen Activator and Matrix Metalloproteinase Systems in Ovulation and Corpus Luteum Formation. Ph.D. Thesis, Umeå University, Umeå, Sweden, 2004. [Google Scholar]
  59. Epifano, O.; Riminucci, M.; Manna, C.; Apa, R.; Greco, E.; Lanzone, A.; Stefanini, M.; Canipari, R. In vitro production of plasminogen activator by human granulosa cells. J. Clin. Endocrinol. Metab. 1994, 78, 174–179. [Google Scholar] [CrossRef]
  60. Zhu, C.; Frederick Woessner, J., Jr. A Tissue Inhibitor of Metalloproteinases and α-Macroglobulins in the Ovulating Rat Ovary: Possible Regulators of Collagen Matrix Breakdown1. Biol. Reprod. 1991, 45, 334–342. [Google Scholar] [CrossRef]
  61. Peng, X.-R.; Hsueh, A.J.W.; Ny, T. Transient and cell-specific expression of tissue-type plasminogen activator and plasminogen-activator-inhibitor type 1 results in controlled and directed proteolysis during gonadotropin-induced ovulation. Eur. J. Biochem. 1993, 214, 147–156. [Google Scholar] [CrossRef]
  62. Hägglund, A.-C.; Ny, A.; Leonardsson, G.; Ny, T. Regulation and Localization of Matrix Metalloproteinases and Tissue Inhibitors of Metalloproteinases in the Mouse Ovary during Gonadotropin-Induced Ovulation. Endocrinology 1999, 140, 4351–4358. [Google Scholar] [CrossRef] [PubMed]
  63. Hägglund, A.-C.; Basset, P.; Ny, T. Stromelysin-3 Is Induced in Mouse Ovarian Follicles Undergoing Hormonally Controlled Apoptosis, but This Metalloproteinase Is Not Required for Follicular Atresia. Biol. Reprod. 2001, 64, 457–463. [Google Scholar] [CrossRef]
  64. Ogiwara, K.; Hagiwara, A.; Rajapakse, S.; Takahashi, T. The Role of Urokinase Plasminogen Activator and Plasminogen Activator Inhibitor-1 in Follicle Rupture During Ovulation in the Teleost Medaka. Biol. Reprod. 2015, 92, 1–17. [Google Scholar] [CrossRef] [PubMed]
  65. O’Connell, M.L.; Canipari, R.; Strickland, S. Hormonal regulation of tissue plasminogen activator secretion and mRNA levels in rat granulosa cells. J. Biol. Chem. 1987, 262, 2339–2344. [Google Scholar] [CrossRef]
  66. Li, M.; Karakji, E.G.; Xing, R.; Fryer, J.N.; Carnegie, J.A.; Rabbani, S.A.; Tsang, B.K. Expression of Urokinase-Type Plasminogen Activator and Its Receptor during Ovarian Follicular Development. Endocrinology 1997, 138, 2790–2799. [Google Scholar] [CrossRef]
  67. Karakji, E.G.; Tsang, B.K. Regulation of Rat Granulosa Cell Plasminogen Activator System: Influence of Interleukin-1β and Ovarian Follicular Development. Biol. Reprod. 1995, 53, 1302–1310. [Google Scholar] [CrossRef] [PubMed]
  68. Hurwitz, A.; Finci-Yeheskel, Z.; Dushnik, M.; Milwidsky, A.; Ben-Chetrit, A.; Yagel, S.; Adashi, E.Y.; Mayer, M. Cytokine-mediated regulation of rat ovarian function: Interleukin-1 inhibits plasminogen activator activity through the induction of plasminogen activator inhibitor-1 (PAI-1). Mol. Cell. Endocrinol. 1994, 101, 307–314. [Google Scholar] [CrossRef]
  69. Sheng, X.; Liu, C.; Yan, G.; Li, G.; Liu, J.; Yang, Y.; Li, S.; Li, Z.; Zhou, J.; Zhen, X.; et al. The mitochondrial protease LONP1 maintains oocyte development and survival by suppressing nuclear translocation of AIFM1 in mammals. eBioMedicine 2022, 75, 103790. [Google Scholar] [CrossRef]
  70. Meng, T.-G.; Hu, M.-W.; Ma, X.-S.; Huang, L.; Liang, Q.-X.; Yuan, Y.; Hou, Y.; Wang, H.; Schatten, H.; Wang, Z.-B.; et al. Oocyte-specific deletion of furin leads to female infertility by causing early secondary follicle arrest in mice. Cell Death Dis. 2017, 8, e2846. [Google Scholar] [CrossRef]
  71. Nyegaard, M.; Overgaard, M.T.; Su, Y.-Q.; Hamilton, A.E.; Kwintkiewicz, J.; Hsieh, M.; Nayak, N.R.; Conti, M.; Conover, C.A.; Giudice, L.C. Lack of Functional Pregnancy-Associated Plasma Protein-A (PAPPA) Compromises Mouse Ovarian Steroidogenesis and Female Fertility. Biol. Reprod. 2010, 82, 1129–1138. [Google Scholar] [CrossRef]
  72. Oxvig, C. The role of PAPP-A in the IGF system: Location, location, location. J. Cell Commun. Signal. 2015, 9, 177–187. [Google Scholar] [CrossRef] [PubMed]
  73. Mason, E.J.; Grell, J.A.; Wan, J.; Cohen, P.; Conover, C.A. Insulin-like growth factor (IGF)-I and IGF-II contribute differentially to the phenotype of pregnancy associated plasma protein-A knock-out mice. Growth Horm. IGF Res. 2011, 21, 243–247. [Google Scholar] [CrossRef]
  74. Folgueras, A.R.; de Lara, F.M.; Pendás, A.M.; Garabaya, C.; Rodríguez, F.; Astudillo, A.; Bernal, T.; Cabanillas, R.; López-Otín, C.; Velasco, G. Membrane-bound serine protease matriptase-2 (Tmprss6) is an essential regulator of iron homeostasis. Blood 2008, 112, 2539–2545. [Google Scholar] [CrossRef]
  75. Tonai, S.; Kawabata, A.; Nakanishi, T.; Lee, J.Y.; Okamoto, A.; Shimada, M.; Yamashita, Y. Iron deficiency induces female infertile in order to failure of follicular development in mice. J. Reprod. Dev. 2020, 66, 475–483. [Google Scholar] [CrossRef]
  76. Turk, V.; Turk, B.; Turk, D. Lysosomal cysteine proteases: Facts and opportunities. EMBO J. 2001, 20, 4629–4633. [Google Scholar] [CrossRef] [PubMed]
  77. Turk, V.; Stoka, V.; Vasiljeva, O.; Renko, M.; Sun, T.; Turk, B.; Turk, D. Cysteine cathepsins: From structure, function and regulation to new frontiers. Biochim. Biophys. Acta Proteins Proteom. 2012, 1824, 68–88. [Google Scholar] [CrossRef]
  78. Oksjoki, S.; Söderström, M.; Vuorio, E.; Anttila, L. Differential expression patterns of cathepsins B, H, K, L and S in the mouse ovary. Mol. Hum. Reprod. 2001, 7, 27–34. [Google Scholar] [CrossRef] [PubMed]
  79. Peluffo, M.C.; Murphy, M.J.; Talcott Baughman, S.; Stouffer, R.L.; Hennebold, J.D. Systematic Analysis of Protease Gene Expression in the Rhesus Macaque Ovulatory Follicle: Metalloproteinase Involvement in Follicle Rupture. Endocrinology 2011, 152, 3963–3974. [Google Scholar] [CrossRef]
  80. Lind, A.-K.; Dahm-Kähler, P.; Weijdegård, B.; Sundfeldt, K.; Brännström, M. Gelatinases and their tissue inhibitors during human ovulation: Increased expression of tissue inhibitor of matrix metalloproteinase-1. Mol. Hum. Reprod. 2006, 12, 725–736. [Google Scholar] [CrossRef]
  81. Bakke, L.J.; Li, Q.; Cassar, C.A.; Dow, M.P.D.; Pursley, J.R.; Smith, G.W. Gonadotropin Surge-Induced Differential Upregulation of Collagenase-1 (MMP-1) and Collagenase-3 (MMP-13) mRNA and Protein in Bovine Preovulatory Follicles. Biol. Reprod. 2004, 71, 605–612. [Google Scholar] [CrossRef]
  82. Hrabia, A.; Wolak, D.; Kwaśniewska, M.; Kieronska, A.; Socha, J.K.; Sechman, A. Expression of gelatinases (MMP-2 and MMP-9) and tissue inhibitors of metalloproteinases (TIMP-2 and TIMP-3) in the chicken ovary in relation to follicle development and atresia. Theriogenology 2019, 125, 268–276. [Google Scholar] [CrossRef] [PubMed]
  83. Zhu, G.; Kang, L.; Wei, Q.; Cui, X.; Wang, S.; Chen, Y.; Jiang, Y. Expression and Regulation of MMP1, MMP3, and MMP9 in the Chicken Ovary in Response to Gonadotropins, Sex Hormones, and TGFB1. Biol. Reprod. 2014, 90, 1–11. [Google Scholar] [CrossRef]
  84. McCord, L.A.; Li, F.; Rosewell, K.L.; Brännström, M.; Curry, T.E. Ovarian Expression and Regulation of the Stromelysins During the Periovulatory Period in the Human and the Rat. Biol. Reprod. 2012, 86, 78. [Google Scholar] [CrossRef]
  85. Shrestha, K.; Puttabyatappa, M.; Wynn, M.A.; Hannon, P.R.; Al-Alem, L.F.; Rosewell, K.L.; Akin, J.; Curry, T.E., Jr. Protease expression in the human and rat cumulus–oocyte complex during the periovulatory period: A role in cumulus–oocyte complex migration. Biol. Reprod. 2024, 111, 845–855. [Google Scholar] [CrossRef] [PubMed]
  86. Rosewell, K.L.; Al-Alem, L.; Zakerkish, F.; McCord, L.; Akin, J.W.; Chaffin, C.L.; Brännström, M.; Curry, T.E. Induction of proteinases in the human preovulatory follicle of the menstrual cycle by human chorionic gonadotropin. Fertil. Steril. 2015, 103, 826–833. [Google Scholar] [CrossRef]
  87. Bu, S.; Cao, C.; Yang, Y.; Miao, C.; Hu, Z.; Cao, Y.; Sang, Q.A.; Duan, E. Localization and temporal regulation of tissue inhibitor of metalloproteinases-4 in mouse ovary. Reproduction 2006, 131, 1099–1107. [Google Scholar] [CrossRef]
  88. Simpson, K.S.; Komar, C.M.; Curry, T.E., Jr. Localization and Expression of Tissue Inhibitor of Metalloproteinase-4 in the Immature Gonadotropin-Stimulated and Adult Rat Ovary. Biol. Reprod. 2003, 68, 214–221. [Google Scholar] [CrossRef] [PubMed]
  89. Liu, Y.-X.; Cajander, S.B.; Ny, T.; Kristensen, P.; Hsueh, A.J.W. Gonadotropin regulation of tissue-type and urokinase-type plasminogen activators in rat granulosa and theca-interstitial cells during the periovulatory period. Mol. Cell. Endocrinol. 1987, 54, 221–229. [Google Scholar] [CrossRef]
  90. Macchione, E.; Epifano, O.; Stefanini, M.; Belin, D.; Canipari, R. Urokinase Redistribution from the Secreted to the Cell-Bound Fraction in Granulosa Cells of Rat Preovulatory Follicles1. Biol. Reprod. 2000, 62, 895–903. [Google Scholar] [CrossRef]
  91. Karakji, E.G.; Tsang, B.K. Tumor Necrosis Factor Alpha Inhibits Rat Granulosa Cell Plasminogen Activator Activity in Vitro during Follicular Development. Biol. Reprod. 1995, 52, 745–752. [Google Scholar] [CrossRef]
  92. Dow, M.P.D.; Bakke, L.J.; Cassar, C.A.; Peters, M.W.; Pursley, J.R.; Smith, G.W. Gonadotropin Surge-Induced Up-Regulation of the Plasminogen Activators (Tissue Plasminogen Activator and Urokinase Plasminogen Activator) and the Urokinase Plasminogen Activator Receptor Within Bovine Periovulatory Follicular and Luteal Tissue. Biol. Reprod. 2002, 66, 1413–1421. [Google Scholar] [CrossRef] [PubMed]
  93. Li, J.; Balboula, A.Z.; Aboelenain, M.; Fujii, T.; Moriyasu, S.; Bai, H.; Kawahara, M.; Takahashi, M. Effect of autophagy induction and cathepsin B inhibition on developmental competence of poor quality bovine oocytes. J. Reprod. Dev. 2020, 66, 83–91. [Google Scholar] [CrossRef]
  94. Zhang, K.; Xu, R.; Zheng, L.; Zhang, H.; Qian, Z.; Li, C.; Xue, M.; He, Z.; Ma, J.; Li, Z.; et al. Elevated N-glycosylated cathepsin L impairs oocyte function and contributes to oocyte senescence during reproductive aging. Aging Cell 2025, 24, e14397. [Google Scholar] [CrossRef]
  95. Holland, A.; Findlay, J.; Clements, J. Kallikrein gene expression in the gonadotrophin-stimulated rat ovary. J. Endocrinol. 2001, 170, 243–250. [Google Scholar] [CrossRef] [PubMed]
  96. Clements, J.A.; Mukhtar, A.; Holland, A.M.; Ehrlich, A.R.; Fuller, P.J. Kallikrein gene family expression in the rat ovary: Localization to the granulosa cell. Endocrinology 1995, 136, 1137–1144. [Google Scholar] [CrossRef]
  97. Musil, D.; Zucic, D.; Turk, D.; Engh, R.A.; Mayr, I.; Huber, R.; Popovic, T.; Turk, V.; Towatari, T.; Katunuma, N. The refined 2.15 A X-ray crystal structure of human liver cathepsin B: The structural basis for its specificity. EMBO J. 1991, 10, 2321–2330. [Google Scholar] [CrossRef] [PubMed]
  98. Cavallo-Medved, D.; Moin, K.; Sloane, B. Cathepsin B: Basis sequence: Mouse. AFCS-Nat. Mol. Pages 2011, 2011, A000508. [Google Scholar]
  99. Mohanty, A.; Kumari, A.; Kumar, S.L.; Kumar, A.; Birajdar, P.; Beniwal, R.; Athar, M.; Kumar P, K.; Rao, H.B.D.P. Cathepsin B regulates ovarian reserve quality and quantity via mitophagy by modulating IGF1R turnover. Aging Cell 2024, e70066. [Google Scholar] [CrossRef]
  100. Liang, S.; Jiang, H.; Shen, X.-H.; Zhang, J.-B.; Kim, N.-H. Inhibition of cathepsin B activity prevents deterioration in the quality of in vitro aged porcine oocytes. Theriogenology 2018, 116, 103–111. [Google Scholar] [CrossRef]
  101. Komatsu, K.; Wei, W.; Murase, T.; Masubuchi, S. 17β-Estradiol and cathepsins control primordial follicle growth in mouse ovaries. Reproduction 2021, 162, 277–287. [Google Scholar] [CrossRef]
  102. Bastu, E.; Gokulu, S.G.; Dural, O.; Yasa, C.; Bulgurcuoglu, S.; Karamustafaoglu Balci, B.; Celik, C.; Buyru, F. The association between follicular fluid levels of cathepsin B, relaxin or AMH with clinical pregnancy rates in infertile patients. Eur. J. Obstet. Gynecol. Reprod. Biol. 2015, 187, 30–34. [Google Scholar] [CrossRef]
  103. Kondo, Y.; Rajapakse, S.; Ogiwara, K. Involvement of cathepsin L in the degradation and degeneration of postovulatory follicle of the medaka ovary. Biol. Reprod. 2023, 109, 904–917. [Google Scholar] [CrossRef] [PubMed]
  104. Matousek, M.; Mitsube, K.; Mikuni, M.; Brännström, M. Inhibition of ovulation in the rat by a leukotriene B4 receptor antagonist. Mol. Hum. Reprod. 2001, 7, 35–42. [Google Scholar] [CrossRef]
  105. Shahed, A.; Simmons, J.J.; Featherstone, S.L.; Young, K.A. Matrix metalloproteinase inhibition influences aspects of photoperiod stimulated ovarian recrudescence in Siberian hamsters. Gen. Comp. Endocrinol. 2015, 216, 46–53. [Google Scholar] [CrossRef] [PubMed]
  106. Whited, J.; Shahed, A.; McMichael, C.F.; Young, K.A. Inhibition of Matrix Metalloproteinases (MMPs) in Siberian Hamsters Impedes Photostimulated Recrudescence of Ovaries. Reproduction 2010, 140, 875–883. [Google Scholar] [CrossRef]
  107. Oksjoki, S.; Sallinen, S.; Vuorio, E.; Anttila, L. Cyclic expression of mRNA transcripts for connective tissue components in the mouse ovary. Mol. Hum. Reprod. 1999, 5, 803–808. [Google Scholar] [CrossRef] [PubMed]
  108. Shindo, T.; Kurihara, H.; Kuno, K.; Yokoyama, H.; Wada, T.; Kurihara, Y.; Imai, T.; Wang, Y.; Ogata, M.; Nishimatsu, H.; et al. ADAMTS-1: A metalloproteinase-disintegrin essential for normal growth, fertility, and organ morphology and function. J. Clin. Investig. 2000, 105, 1345–1352. [Google Scholar] [CrossRef]
  109. Mittaz, L.; Russell, D.L.; Wilson, T.; Brasted, M.; Tkalcevic, J.; Salamonsen, L.A.; Hertzog, P.J.; Pritchard, M.A. Adamts-1 Is Essential for the Development and Function of the Urogenital System. Biol. Reprod. 2004, 70, 1096–1105. [Google Scholar] [CrossRef]
  110. Shozu, M.; Minami, N.; Yokoyama, H.; Inoue, M.; Kurihara, H.; Matsushima, K.; Kuno, K. ADAMTS-1 is involved in normal follicular development, ovulatory process and organization of the medullary vascular network in the ovary. J. Mol. Endocrinol. 2005, 35, 343–355. [Google Scholar] [CrossRef]
  111. Carter, N.J.; Roach, Z.A.; Byrnes, M.M.; Zhu, Y. Adamts9 is necessary for ovarian development in zebrafish. Gen. Comp. Endocrinol. 2019, 277, 130–140. [Google Scholar] [CrossRef]
  112. Liu, K.; Brändström, A.; Liu, Y.X.; Ny, T.; Selstam, G. Coordinated expression of tissue-type plasminogen activator and plasminogen activator inhibitor type 1 during corpus luteum formation and luteolysis in the adult pseudopregnant rat. Endocrinology 1996, 137, 2126–2132. [Google Scholar] [CrossRef] [PubMed]
  113. Bacharach, E.; Itin, A.; Keshet, E. In vivo patterns of expression of urokinase and its inhibitor PAI-1 suggest a concerted role in regulating physiological angiogenesis. Proc. Natl. Acad. Sci. USA 1992, 89, 10686–10690. [Google Scholar] [CrossRef]
  114. Jones, P.B.; Muse, K.N.; Wilson, E.A.; Curry, T.E. Expression of plasminogen activator (PA) and a PA inhibitor in human granulosa cells from preovulatory follicles. J. Clin. Endocrinol. Metab. 1988, 67, 857–860. [Google Scholar] [CrossRef]
  115. Chen, B.; Chang, H.-M.; Zhang, Z.; Cao, Y.; Leung, P.C.K. ALK4-SMAD3/4 mediates the effects of activin A on the upregulation of PAI-1 in human granulosa lutein cells. Mol. Cell. Endocrinol. 2020, 505, 110731. [Google Scholar] [CrossRef] [PubMed]
  116. Song, G.; Jiang, Y.; Wang, Y.; Song, M.; Niu, X.; Xu, H.; Li, M. Modulation of Cathepsin S (CTSS) Regulates the Secretion of Progesterone and Estradiol, Proliferation, and Apoptosis of Ovarian Granulosa Cells in Rabbits. Animals 2021, 11, 1770. [Google Scholar] [CrossRef] [PubMed]
  117. Song, H.; Qin, Q.; Yuan, C.; Li, H.; Zhang, F.; Fan, L. Metabolomic Profiling of Poor Ovarian Response Identifies Potential Predictive Biomarkers. Front. Endocrinol. 2021, 12, 774667. [Google Scholar] [CrossRef]
  118. Espey, L.L. Ovarian Proteolytic Enzymes and Ovulation. Biol. Reprod. 1974, 10, 216–235. [Google Scholar] [CrossRef]
  119. Espey, L.L.; Richards, J.; Neill, J. Physiology of Reproduction; Elsevier: Amsterdam, The Netherlands, 2006; Volume 2006, pp. 425–474. [Google Scholar]
  120. Basini, G.; Bussolati, S.; Baioni, L.; Grasselli, F. Gelatinases (MMP2 and MMP9) in swine antral follicle. BioFactors 2011, 37, 117–120. [Google Scholar] [CrossRef]
  121. Deady, L.D.; Shen, W.; Mosure, S.A.; Spradling, A.C.; Sun, J. Matrix Metalloproteinase 2 Is Required for Ovulation and Corpus Luteum Formation in Drosophila. PLoS Genet. 2015, 11, e1004989. [Google Scholar] [CrossRef]
  122. Fujihara, M.; Yamamizu, K.; Wildt, D.; Songsasen, N. Expression pattern of matrix metalloproteinases changes during folliculogenesis in the cat ovary. Reprod. Domest. Anim. 2016, 51, 717–725. [Google Scholar] [CrossRef]
  123. Jo, M.; Curry, T.E., Jr. Regulation of Matrix Metalloproteinase-19 Messenger RNA Expression in the Rat Ovary. Biol. Reprod. 2004, 71, 1796–1806. [Google Scholar] [CrossRef] [PubMed]
  124. Politis, I.; Srikandakumar, A.; Turner, J.D.; Tsang, B.K.; Ainsworth, L.; Downey, B.R. Changes in and Partial Identification of the Plasminogen Activator and Plasminogen Activator Inhibitor Systems During Ovarian Follicular Maturation in the Pig. Biol. Reprod. 1990, 43, 636–642. [Google Scholar] [CrossRef] [PubMed]
  125. Zupanič, N.; Počič, J.; Leonardi, A.; Šribar, J.; Kordiš, D.; Križaj, I. Serine pseudoproteases in physiology and disease. FEBS J. 2023, 290, 2263–2278. [Google Scholar] [CrossRef] [PubMed]
  126. Canipari, R.; Strickland, S. Plasminogen activator in the rat ovary. Production and gonadotropin regulation of the enzyme in granulosa and thecal cells. J. Biol. Chem. 1985, 260, 5121–5125. [Google Scholar] [CrossRef]
  127. Canipari, R.; O’Connell, M.L.; Meyer, G.; Strickland, S. Mouse ovarian granulosa cells produce urokinase-type plasminogen activator, whereas the corresponding rat cells produce tissue-type plasminogen activator. J. Cell Biol. 1987, 105, 977–981. [Google Scholar] [CrossRef]
  128. Shen, X.; Minoura, H.; Yoshida, T.; Toyoda, N. Changes in Ovarian Expression of Tissue-Type Plasminogen Activator and Plasminogen Activator Inhibitor Type-1 Messenger Ribonucleic Acids during Ovulation in Rat. Endocr. J. 1997, 44, 341–348. [Google Scholar] [CrossRef]
  129. Liu, Y.-X. Regulation of the Plasminogen Activator System in the Ovary. Biol. Signals Recept. 1999, 8, 160–177. [Google Scholar] [CrossRef]
  130. Plendl, J.; Snyman, C.; Naidoo, S.; Sawant, S.; Mahabeer, R.; Bhoola, K.D. Expression of Tissue Kallikrein and Kinin Receptors in Angiogenic Microvascular Endothelial Cells. Biol. Chem. 2000, 381, 1103–1115. [Google Scholar] [CrossRef]
  131. Hazzard, T.M.; Stouffer, R.L. Angiogenesis in ovarian follicular and luteal development. Best Pract. Res. Clin. Obstet. Gynaecol. 2000, 14, 883–900. Available online: https://www.sciencedirect.com/science/article/pii/S1521693400901330 (accessed on 21 April 2025). [CrossRef]
  132. Tamanini, C.; De Ambrogi, M. Angiogenesis in Developing Follicle and Corpus Luteum. Reprod. Domest. Anim. 2004, 39, 206–216. [Google Scholar] [CrossRef]
  133. Kobayashi, M.; Yoshino, O.; Nakashima, A.; Ito, M.; Nishio, K.; Ono, Y.; Kusabiraki, T.; Kunitomi, C.; Takahashi, N.; Harada, M.; et al. Inhibition of autophagy in theca cells induces CYP17A1 and PAI-1 expression via ROS/p38 and JNK signalling during the development of polycystic ovary syndrome. Mol. Cell. Endocrinol. 2020, 508, 110792. [Google Scholar] [CrossRef] [PubMed]
  134. Paliouras, M.; Diamandis, E.P. Intracellular signaling pathways regulate hormone-dependent kallikrein gene expression. Tumour Biol. 2008, 29, 63–75. [Google Scholar] [CrossRef]
  135. Iwasaki, T.; Tokumori, M.; Matsubara, M.; Ojima, F.; Kamigochi, K.; Aizawa, S.; Ogoshi, M.; Kimura, A.P.; Takeuchi, S.; Takahashi, S. A regulatory mechanism of mouse kallikrein 1 gene expression by estrogen. Mol. Cell. Endocrinol. 2023, 577, 112044. [Google Scholar] [CrossRef] [PubMed]
  136. Lorenzetti, S.; Marcoccia, D.; Narciso, L.; Mantovani, A. Cell viability and PSA secretion assays in LNCaP cells: A tiered in vitro approach to screen chemicals with a prostate-mediated effect on male reproduction within the ReProTect project. Reprod. Toxicol. 2010, 30, 25–35. [Google Scholar] [CrossRef]
  137. Raimondo, S.; Gentile, M.; Esposito, G.; Gentile, T.; Ferrara, I.; Crescenzo, C.; Palmieri, M.; Cuomo, F.; De Filippo, S.; Lettieri, G.; et al. Could Kallikrein-Related Serine Peptidase 3 Be an Early Biomarker of Environmental Exposure in Young Women? Int. J. Environ. Res. Public Health 2021, 18, 8833. [Google Scholar] [CrossRef] [PubMed]
  138. Lawrence, M.G.; Lai, J.; Clements, J.A. Kallikreins on Steroids: Structure, Function, and Hormonal Regulation of Prostate-Specific Antigen and the Extended Kallikrein Locus. Endocr. Rev. 2010, 31, 407–446. [Google Scholar] [CrossRef] [PubMed]
  139. Borgoño, C.A.; Diamandis, E.P. The emerging roles of human tissue kallikreins in cancer. Nat. Rev. Cancer 2004, 4, 876–890. [Google Scholar] [CrossRef]
  140. Kurlender, L.; Yousef, G.M.; Memari, N.; Robb, J.-D.; Michael, I.P.; Borgoño, C.; Katsaros, D.; Stephan, C.; Jung, K.; Diamandis, E.P. Differential Expression of a Human Kallikrein 5 (KLK5) Splice Variant in Ovarian and Prostate Cancer. Tumor Biol. 2004, 25, 149–156. [Google Scholar] [CrossRef]
  141. Ahmed, N.; Dorn, J.; Napieralski, R.; Drecoll, E.; Kotzsch, M.; Goettig, P.; Zein, E.; Avril, S.; Kiechle, M.; Diamandis, E.P.; et al. Clinical relevance of kallikrein-related peptidase 6 (KLK6) and 8 (KLK8) mRNA expression in advanced serous ovarian cancer. Biol. Chem. 2016, 397, 1265–1276. [Google Scholar] [CrossRef]
  142. Sriraman, V.; Richards, J.S. Cathepsin L Gene Expression and Promoter Activation in Rodent Granulosa Cells. Endocrinology 2004, 145, 582–591. [Google Scholar] [CrossRef] [PubMed]
  143. Carnevali, O.; Cionna, C.; Tosti, L.; Lubzens, E.; Maradonna, F. Role of cathepsins in ovarian follicle growth and maturation. Gen. Comp. Endocrinol. 2006, 146, 195–203. [Google Scholar] [CrossRef] [PubMed]
  144. Balboula, A.Z.; Yamanaka, K.; Sakatani, M.; Hegab, A.O.; Zaabel, S.M.; Takahashi, M. Cathepsin B activity is related to the quality of bovine cumulus oocyte complexes and its inhibition can improve their developmental competence. Mol. Reprod. Dev. 2010, 77, 439–448. [Google Scholar] [CrossRef] [PubMed]
  145. Balboula, A.Z.; Yamanaka, K.; Sakatani, M.; Kawahara, M.; Hegab, A.O.; Zaabel, S.M.; Takahashi, M. Cathepsin B activity has a crucial role in the developmental competence of bovine cumulus–oocyte complexes exposed to heat shock during in vitro maturation. Reproduction 2013, 146, 407–417. [Google Scholar] [CrossRef]
  146. García, V.; Kohen, P.; Maldonado, C.; Sierralta, W.; Muñoz, A.; Villarroel, C.; Strauss, J.F.; Devoto, L. Transient expression of progesterone receptor and cathepsin-l in human granulosa cells during the periovulatory period. Fertil. Steril. 2012, 97, 707–713.e1. [Google Scholar] [CrossRef]
  147. Arosh, J.A.; Banu, S.K.; McCracken, J.A. Novel concepts on the role of prostaglandins on luteal maintenance and maternal recognition and establishment of pregnancy in ruminants1. J. Dairy Sci. 2016, 99, 5926–5940. [Google Scholar] [CrossRef] [PubMed]
  148. Wiltbank, M.C.; Mezera, M.A.; Toledo, M.Z.; Drum, J.N.; Baez, G.M.; García-Guerra, A.; Sartori, R. Physiological mechanisms involved in maintaining the corpus luteum during the first two months of pregnancy. Anim. Reprod. 2018, 15 (Suppl. 1), 805–821. [Google Scholar] [CrossRef]
  149. Cha, J.; Sun, X.; Dey, S.K. Mechanisms of implantation: Strategies for successful pregnancy. Nat. Med. 2012, 18, 1754–1767. [Google Scholar] [CrossRef]
  150. Ferrara, N.; Chen, H.; Davis-Smyth, T.; Gerber, H.-P.; Nguyen, T.-N.; Peers, D.; Chisholm, V.; Hillan, K.J.; Schwall, R.H. Vascular endothelial growth factor is essential for corpus luteum angiogenesis. Nat. Med. 1998, 4, 336–340. [Google Scholar] [CrossRef]
  151. Wahlberg, P.; Nylander, Å.; Ahlskog, N.; Liu, K.; Ny, T. Expression and Localization of the Serine Proteases High-Temperature Requirement Factor A1, Serine Protease 23, and Serine Protease 35 in the Mouse Ovary. Endocrinology 2008, 149, 5070–5077. [Google Scholar] [CrossRef]
  152. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2021. CA Cancer J. Clin. 2021, 71, 7–33. [Google Scholar] [CrossRef]
  153. Elias, K.M.; Guo, J.; Bast, R.C. Early Detection of Ovarian Cancer. Hematol. Oncol. Clin. N. Am. 2018, 32, 903–914. [Google Scholar] [CrossRef] [PubMed]
  154. Tingulstad, S.; Skjeldestad, F.E.; Halvorsen, T.B.; Hagen, B. jørn Survival and prognostic factors in patients with ovarian cancer. Obstet. Gynecol. 2003, 101, 885–891. [Google Scholar] [CrossRef] [PubMed]
  155. Carey, P.; Low, E.; Harper, E.; Stack, M.S. Metalloproteinases in Ovarian Cancer. Int. J. Mol. Sci. 2021, 22, 3403. [Google Scholar] [CrossRef] [PubMed]
  156. Scorilas, A.; Borgoño, C.A.; Harbeck, N.; Dorn, J.; Schmalfeldt, B.; Schmitt, M.; Diamandis, E.P. Human Kallikrein 13 Protein in Ovarian Cancer Cytosols: A New Favorable Prognostic Marker. J. Clin. Oncol. 2004, 22, 678–685. [Google Scholar] [CrossRef]
  157. Zhang, Y.; Chen, Q. Relationship between matrix metalloproteinases and the occurrence and development of ovarian cancer. Braz. J. Med. Biol. Res. 2017, 50, e6104. [Google Scholar] [CrossRef]
  158. Goldman, S.; Shalev, E. The role of the matrix metalloproteinases in human endometrial and ovarian cycles. Eur. J. Obstet. Gynecol. Reprod. Biol. 2003, 111, 109–121. [Google Scholar] [CrossRef]
  159. Irving-Rodgers, H.F.; Rodgers, R.J. Extracellular matrix in ovarian follicular development and disease. Cell Tissue Res. 2005, 322, 89–98. [Google Scholar] [CrossRef]
  160. Kenny, H.A.; Kaur, S.; Coussens, L.M.; Lengyel, E. The initial steps of ovarian cancer cell metastasis are mediated by MMP-2 cleavage of vitronectin and fibronectin. J. Clin. Investig. 2008, 118, 1367–1379. [Google Scholar] [CrossRef]
  161. Sakata, K.; Shigemasa, K.; Nagai, N.; Ohama, K. Expression of matrix metalloproteinases (MMP-2, MMP-9, MT1-MMP) and their inhibitors (TIMP-1, TIMP-2) in common epithelial tumors of the ovary. Int. J. Oncol. 2000, 17, 673–754. [Google Scholar] [CrossRef]
  162. Davidson, B.; Goldberg, I.; Gotlieb, W.H.; Kopolovic, J.; Ben-Baruch, G.; Nesland, J.M.; Berner, A.; Bryne, M.; Reich, R. High levels of MMP-2, MMP-9, MT1-MMP and TIMP-2 mRNA correlate with poor survival in ovarian carcinoma. Clin. Exp. Metastasis 1999, 17, 799–808. [Google Scholar] [CrossRef]
  163. Fu, Z.; Xu, S.; Xu, Y.; Ma, J.; Li, J.; Xu, P. The Expression of Tumor-Derived and Stromal-Derived Matrix Metalloproteinase 2 Predicted Prognosis of Ovarian Cancer. Int. J. Gynecol. Cancer 2015, 25, 356–362. [Google Scholar] [CrossRef] [PubMed]
  164. Davidson, B.; Goldberg, I.; Gotlieb, W.H.; Kopolovic, J.; Ben-Baruch, G.; Nesland, J.M.; Reich, R. The prognostic value of metalloproteinases and angiogenic factors in ovarian carcinoma. Mol. Cell. Endocrinol. 2002, 187, 39–45. [Google Scholar] [CrossRef] [PubMed]
  165. Sillanpää, S.; Anttila, M.; Voutilainen, K.; Ropponen, K.; Turpeenniemi-Hujanen, T.; Puistola, U.; Tammi, R.; Tammi, M.; Sironen, R.; Saarikoski, S.; et al. Prognostic significance of matrix metalloproteinase-9 (MMP-9) in epithelial ovarian cancer. Gynecol. Oncol. 2007, 104, 296–303. [Google Scholar] [CrossRef] [PubMed]
  166. Gong, W.; Liu, Y.; Diamandis, E.P.; Kiechle, M.; Bronger, H.; Dorn, J.; Dreyer, T.; Magdolen, V. Prognostic value of kallikrein-related peptidase 7 (KLK7) mRNA expression in advanced high-grade serous ovarian cancer. J. Ovarian Res. 2020, 13, 125. [Google Scholar] [CrossRef]
  167. Al-Alem, L.F.; McCord, L.A.; Southard, R.C.; Kilgore, M.W.; Curry, T.E., Jr. Activation of the PKC Pathway Stimulates Ovarian Cancer Cell Proliferation, Migration, and Expression of MMP7 and MMP101. Biol. Reprod. 2013, 89, 1–7. [Google Scholar] [CrossRef]
  168. Al-Alem, L.; Curry, T.E. Ovarian cancer: Involvement of the matrix metalloproteinases. Reproduction 2015, 150, R55–R64. [Google Scholar] [CrossRef]
  169. Kamat, A.A.; Fletcher, M.; Gruman, L.M.; Mueller, P.; Lopez, A.; Landen, C.N., Jr.; Han, L.; Gershenson, D.M.; Sood, A.K. The Clinical Relevance of Stromal Matrix Metalloproteinase Expression in Ovarian Cancer. Clin. Cancer Res. 2006, 12, 1707–1714. [Google Scholar] [CrossRef]
  170. Sood, A.K.; Fletcher, M.S.; Coffin, J.E.; Yang, M.; Seftor, E.A.; Gruman, L.M.; Gershenson, D.M.; Hendrix, M.J.C. Functional role of matrix metalloproteinases in ovarian tumor cell plasticity. Am. J. Obstet. Gynecol. 2004, 190, 899–909. [Google Scholar] [CrossRef]
  171. Agrawal, A.; Ambad, R.; Lahoti, R.; Muley, P.; Pande, P.S. Role of Artificial Intelligence in PCOS Detection. J. Datta Meghe Inst. Med. Sci. Univ. 2022, 17, 491. [Google Scholar] [CrossRef]
  172. Gao, X.-W.; Su, X.-T.; Lu, Z.-H.; Ou, J. 17β-Estradiol Prevents Extracellular Matrix Degradation by Downregulating MMP3 Expression via PI3K/Akt/FOXO3 Pathway. Spine 2020, 45, 292. [Google Scholar] [CrossRef]
  173. Wang, F.-Q.; So, J.; Reierstad, S.; Fishman, D.A. Matrilysin (MMP-7) promotes invasion of ovarian cancer cells by activation of progelatinase. Int. J. Cancer 2005, 114, 19–31. [Google Scholar] [CrossRef] [PubMed]
  174. Chang, M.-C.; Chen, C.-A.; Chen, P.-J.; Chiang, Y.-C.; Chen, Y.-L.; Mao, T.-L.; Lin, H.-W.; Lin Chiang, W.-H.; Cheng, W.-F. Mesothelin enhances invasion of ovarian cancer by inducing MMP-7 through MAPK/ERK and JNK pathways. Biochem. J. 2012, 442, 293–302. [Google Scholar] [CrossRef] [PubMed]
  175. Stadlmann, S.; Pollheimer, J.; Moser, P.L.; Raggi, A.; Amberger, A.; Margreiter, R.; Offner, F.A.; Mikuz, G.; Dirnhofer, S.; Moch, H. Cytokine-regulated expression of collagenase-2 (MMP-8) is involved in the progression of ovarian cancer. Eur. J. Cancer 2003, 39, 2499–2505. [Google Scholar] [CrossRef]
  176. Chen, S.-S.; Song, J.; Tu, X.-Y.; Zhao, J.-H.; Ye, X.-Q. The association between MMP-12 82 A/G polymorphism and susceptibility to various malignant tumors: A meta-analysis. Int. J. Clin. Exp. Med. 2015, 8, 10845–10854. [Google Scholar]
  177. Liu, L.; Sun, J.; Li, G.; Gu, B.; Wang, X.; Chi, H.; Guo, F. Association between MMP-12-82A/G polymorphism and cancer risk: A meta-analysis. Int. J. Clin. Exp. Med. 2015, 8, 11896–11904. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4612787/ (accessed on 21 April 2025).
  178. Hantke, B.; Harbeck, N.; Schmalfeldt, B.; Claes, I.; Hiller, O.; Luther, M.-O.; Welk, A.; Kuhn, W.; Schmitt, M.; Tschesche, H.; et al. Clinical Relevance of Matrix Metalloproteinase-13 Determined with a New Highly Specific and Sensitive ELISA in Ascitic Fluid of Advanced Ovarian Carcinoma Patients. Biol. Chem. 2003, 384, 1247–1251. [Google Scholar] [CrossRef] [PubMed]
  179. Escalona, R.M.; Kannourakis, G.; Findlay, J.K.; Ahmed, N. Expression of TIMPs and MMPs in Ovarian Tumors, Ascites, Ascites-Derived Cells, and Cancer Cell Lines: Characteristic Modulatory Response Before and After Chemotherapy Treatment. Front. Oncol. 2022, 11, 796588. [Google Scholar] [CrossRef]
  180. Escalona, R.M.; Bilandzic, M.; Western, P.; Kadife, E.; Kannourakis, G.; Findlay, J.K.; Ahmed, N. TIMP-2 regulates proliferation, invasion and STAT3-mediated cancer stem cell-dependent chemoresistance in ovarian cancer cells. BMC Cancer 2020, 20, 960. [Google Scholar] [CrossRef]
  181. Davidson, B.; Reich, R.; Berner, A.; Givant-Horwitz, V.; Goldberg, I.; Risberg, B.; Kristensen, G.B.; Trope, C.G.; Bryne, M.; Kopolovic, J.; et al. Ovarian carcinoma cells in serous effusions show altered MMP-2 and TIMP-2 mRNA levels. Eur. J. Cancer 2001, 37, 2040–2049. [Google Scholar] [CrossRef]
  182. Będkowska, G.E.; Piskór, B.; Gacuta, E.; Zajkowska, M.; Osada, J.; Szmitkowski, M.; Dąbrowska, M.; Ławicki, S. Diagnostic Power of Selected Cytokines, MMPs and TIMPs in Ovarian Cancer Patients–ROC Analysis. Anticancer Res. 2019, 39, 2575–2582. [Google Scholar] [CrossRef]
  183. Stetler-Stevenson, W.G. Tissue Inhibitors of Metalloproteinases in Cell Signaling: Metalloproteinase-Independent Biological Activities. Sci. Signal. 2008, 1, re6. [Google Scholar] [CrossRef] [PubMed]
  184. Brew, K.; Nagase, H. The tissue inhibitors of metalloproteinases (TIMPs): An ancient family with structural and functional diversity. Biochim. Biophys. Acta BBA-Mol. Cell Res. 2010, 1803, 55–71. [Google Scholar] [CrossRef]
  185. Cruz-Munoz, W.; Khokha, R. The Role of Tissue Inhibitors of Metalloproteinases in Tumorigenesis and Metastasis. Crit. Rev. Clin. Lab. Sci. 2008, 45, 291–338. [Google Scholar] [CrossRef] [PubMed]
  186. ASTEDT, B. Immunological identity of urokinase and ovarian carcinoma plasminogen activator released in tissue culture. Eur. J Cancer 1981, 17, 239–241. [Google Scholar] [CrossRef]
  187. Kuhn, W.; Schmalfeldt, B.; Reuning, U.; Pache, L.; Berger, U.; Ulm, K.; Harbeck, N.; Späthe, K.; Dettmar, P.; Höfler, H.; et al. Prognostic significance of urokinase (uPA) and its inhibitor PAI-1 for survival in advanced ovarian carcinoma stage FIGO IIIc. Br. J. Cancer 1999, 79, 1746–1751. [Google Scholar] [CrossRef]
  188. Cai, S.; Zhang, P.; Dong, S.; Li, L.; Cai, J.; Xu, M. Downregulation of SPINK13 Promotes Metastasis by Regulating uPA in Ovarian Cancer Cells. Cell. Physiol. Biochem. 2018, 45, 1061–1071. [Google Scholar] [CrossRef]
  189. Chambers, S.K.; Ivins, C.M.; Carcangiu, M.L. Plasminogen activator inhibitor-1 is an independent poor prognostic factor for survival in advanced stage epithelial ovarian cancer patients. Int. J. Cancer 1998, 79, 449–454. [Google Scholar] [CrossRef]
  190. Alberti, C.; Pinciroli, P.; Valeri, B.; Ferri, R.; Ditto, A.; Umezawa, K.; Sensi, M.; Canevari, S.; Tomassetti, A. Ligand-dependent EGFR activation induces the co-expression of IL-6 and PAI-1 via the NFkB pathway in advanced-stage epithelial ovarian cancer. Oncogene 2012, 31, 4139–4149. [Google Scholar] [CrossRef] [PubMed]
  191. Dorn, J.; Harbeck, N.; Kates, R.; Gkazepis, A.; Scorilas, A.; Soosaipillai, A.; Diamandis, E.; Kiechle, M.; Schmalfeldt, B.; Schmitt, M. Impact of expression differences of kallikrein-related peptidases and of uPA and PAI-1 between primary tumor and omentum metastasis in advanced ovarian cancer. Ann. Oncol. 2011, 22, 877–883. [Google Scholar] [CrossRef]
  192. Wang, L.; Madigan, M.C.; Chen, H.; Liu, F.; Patterson, K.I.; Beretov, J.; O’Brien, P.M.; Li, Y. Expression of urokinase plasminogen activator and its receptor in advanced epithelial ovarian cancer patients. Gynecol. Oncol. 2009, 114, 265–272. [Google Scholar] [CrossRef]
  193. Harbeck, N.; Krüger, A.; Sinz, S.; Kates, R.E.; Thomssen, C.; Schmitt, M.; Jänicke, F. Clinical Relevance of the Plasminogen Activator Inhibitor Type1—A Multifaceted Proteolytic Factor. Onkologie 2001, 24, 238–244. [Google Scholar] [CrossRef]
  194. van Dam, P.A.; Coelho, A.; Rolfo, C. Is there a role for urokinase-type plasminogen activator inhibitors as maintenance therapy in patients with ovarian cancer? Eur. J. Surg. Oncol. EJSO 2017, 43, 252–257. [Google Scholar] [CrossRef] [PubMed]
  195. Kenny, H.A.; Leonhardt, P.; Ladanyi, A.; Yamada, S.D.; Montag, A.; Im, H.K.; Jagadeeswaran, S.; Shaw, D.E.; Mazar, A.P.; Lengyel, E. Targeting the Urokinase Plasminogen Activator Receptor Inhibits Ovarian Cancer Metastasis. Clin. Cancer Res. 2011, 17, 459–471. [Google Scholar] [CrossRef]
  196. Nakatsuka, E.; Sawada, K.; Nakamura, K.; Yoshimura, A.; Kinose, Y.; Kodama, M.; Hashimoto, K.; Mabuchi, S.; Makino, H.; Morii, E.; et al. Plasminogen activator inhibitor-1 is an independent prognostic factor of ovarian cancer and IMD-4482, a novel plasminogen activator inhibitor-1 inhibitor, inhibits ovarian cancer peritoneal dissemination. Oncotarget 2017, 8, 89887–89902. [Google Scholar] [CrossRef]
  197. Kelly, T.E.; Spillane, C.L.; Ward, M.P.; Hokamp, K.; Huang, Y.; Tewari, P.; Martin, C.M.; Norris, L.A.; Mohamed, B.M.; Bates, M.; et al. Plasminogen activator inhibitor 1 is associated with high-grade serous ovarian cancer metastasis and is reduced in patients who have received neoadjuvant chemotherapy. Front. Cell Dev. Biol. 2023, 11, 1150991. [Google Scholar] [CrossRef] [PubMed]
  198. Seiffert, D.; Loskutoff, D.J. Type 1 plasminogen activator inhibitor induces multimerization of plasma vitronectin: A suggested mechanism for the generation of the tissue form of vitronectin in vivo. J. Biol. Chem. 1996, 271, 29644–29651. [Google Scholar] [CrossRef]
  199. Seiffert, D.; Ciambrone, G.; Wagner, N.V.; Binder, B.R.; Loskutoff, D.J. The somatomedin B domain of vitronectin. Structural requirements for the binding and stabilization of active type 1 plasminogen activator inhibitor. J. Biol. Chem. 1994, 269, 2659–2666. [Google Scholar] [CrossRef]
  200. Planus, E.; Barlovatz-Meimon, G.; Rogers, R.A.; Bonavaud, S.; Ingber, D.E.; Wang, N. Binding of urokinase to plasminogen activator inhibitor type-1 mediates cell adhesion and spreading. J. Cell Sci. 1997, 110, 1091–1098. [Google Scholar] [CrossRef] [PubMed]
  201. De Lorenzi, V.; Sarra Ferraris, G.M.; Madsen, J.B.; Lupia, M.; Andreasen, P.A.; Sidenius, N. Urokinase links plasminogen activation and cell adhesion by cleavage of the RGD motif in vitronectin. EMBO Rep. 2016, 17, 982–998. [Google Scholar] [CrossRef]
  202. Czekay, R.-P.; Wilkins-Port, C.E.; Higgins, S.P.; Freytag, J.; Overstreet, J.M.; Klein, R.M.; Higgins, C.E.; Samarakoon, R.; Higgins, P.J. PAI-1: An Integrator of Cell Signaling and Migration. Int. J. Cell Biol. 2011, 2011, 562481. [Google Scholar] [CrossRef]
  203. McMahon, G.A.; Petitclerc, E.; Stefansson, S.; Smith, E.; Wong, M.K.K.; Westrick, R.J.; Ginsburg, D.; Brooks, P.C.; Lawrence, D.A. Plasminogen Activator Inhibitor-1 Regulates Tumor Growth and Angiogenesis. J. Biol. Chem. 2001, 276, 33964–33968. [Google Scholar] [CrossRef]
  204. Andreasen, P.A.; Egelund, R.; Petersen, H.H. The plasminogen activation system in tumor growth, invasion, and metastasis. Cell. Mol. Life Sci. CMLS 2000, 57, 25–40. [Google Scholar] [CrossRef]
  205. Cortese, K.; Sahores, M.; Madsen, C.D.; Tacchetti, C.; Blasi, F. Clathrin and LRP-1-Independent Constitutive Endocytosis and Recycling of uPAR. PLoS ONE 2008, 3, e3730. [Google Scholar] [CrossRef]
  206. Placencio, V.R.; DeClerck, Y.A. Plasminogen Activator Inhibitor-1 in Cancer: Rationale and Insight for Future Therapeutic Testing. Cancer Res. 2015, 75, 2969–2974. [Google Scholar] [CrossRef]
  207. Liu, J.; Guo, Q.; Chen, B.; Yu, Y.; Lu, H.; Li, Y.-Y. Cathepsin B and its interacting proteins, bikunin and TSRC1, correlate with TNF-induced apoptosis of ovarian cancer cells OV-90. FEBS Lett. 2006, 580, 245–250. [Google Scholar] [CrossRef]
  208. Zhang, L.; Wei, L.; Shen, G.; He, B.; Gong, W.; Min, N.; Zhang, L.; Duan, Y.; Xie, J.; Luo, H.; et al. Cathepsin L is involved in proliferation and invasion of ovarian cancer cells. Mol. Med. Rep. 2015, 11, 468–474. [Google Scholar] [CrossRef]
  209. Zhang, W.; Wang, S.; Wang, Q.; Yang, Z.; Pan, Z.; Li, L. Overexpression of cysteine cathepsin L is a marker of invasion and metastasis in ovarian cancer. Oncol. Rep. 2014, 31, 1334–1342. [Google Scholar] [CrossRef]
  210. Kolwijck, E.; Massuger, L.F.A.G.; Thomas, C.M.G.; Span, P.N.; Krasovec, M.; Kos, J.; Sweep, F.C.G.J. Cathepsins B, L and cystatin C in cyst fluid of ovarian tumors. J. Cancer Res. Clin. Oncol. 2010, 136, 771–778. [Google Scholar] [CrossRef]
  211. Gashenko, E.A.; Lebedeva, V.A.; Brak, I.V.; Tsykalenko, E.A.; Vinokurova, G.V.; Korolenko, T.A. Evaluation of serum procathepsin B, cystatin B and cystatin C as possible biomarkers of ovarian cancer. Int. J. Circumpolar Health 2013, 72, 21215. [Google Scholar] [CrossRef]
  212. Nishikawa, H.; Ozaki, Y.; Nakanishi, T.; Blomgren, K.; Tada, T.; Arakawa, A.; Suzumori, K. The role of cathepsin B and cystatin C in the mechanisms of invasion by ovarian cancer. Gynecol. Oncol. 2004, 92, 881–886. [Google Scholar] [CrossRef]
  213. Kos, J.; Werle, B.; Lah, T.; Brunner, N. Cysteine proteinases and their inhibitors in extracellular fluids: Markers for diagnosis and prognosis in cancer. Int. J. Biol. Markers 2000, 15, 84–89. [Google Scholar] [CrossRef]
  214. Magister, S.; Kos, J. Cystatins in immune system. J. Cancer 2013, 4, 45–56. [Google Scholar] [CrossRef]
  215. Jedeszko, C.; Sloane, B.F. Cysteine cathepsins in human cancer. Biol. Chem. 2004, 385, 1017–1027. [Google Scholar] [CrossRef]
  216. Wallin, H.; Bjarnadottir, M.; Vogel, L.K.; Wassélius, J.; Ekström, U.; Abrahamson, M. Cystatins–Extra- and intracellular cysteine protease inhibitors: High-level secretion and uptake of cystatin C in human neuroblastoma cells. Biochimie 2010, 92, 1625–1634. [Google Scholar] [CrossRef]
  217. Xu, H.; Ma, Y.; Zhang, Y.; Pan, Z.; Lu, Y.; Liu, P.; Lu, B. Identification of Cathepsin K in the Peritoneal Metastasis of Ovarian Carcinoma Using In-silico, Gene Expression Analysis. J. Cancer 2016, 7, 722–729. [Google Scholar] [CrossRef]
  218. Lösch, A.; Schindl, M.; Kohlberger, P.; Lahodny, J.; Breitenecker, G.; Horvat, R.; Birner, P. Cathepsin D in ovarian cancer: Prognostic value and correlation with p53 expression and microvessel density. Gynecol. Oncol. 2004, 92, 545–552. [Google Scholar] [CrossRef]
  219. Scambia, G.; Benedetti, P.; Ferrandina, G.; Battaglia, F.; Baiocchi, G.; Mancuso, S. Cathepsin D assay in ovarian cancer: Correlation with pathological features and receptors for oestrogen, progesterone and epidermal growth factor. Br. J. Cancer 1991, 64, 182–184. [Google Scholar] [CrossRef]
  220. Pranjol, M.Z.I.; Gutowski, N.; Hannemann, M.; Whatmore, J. The Potential Role of the Proteases Cathepsin D and Cathepsin L in the Progression and Metastasis of Epithelial Ovarian Cancer. Biomolecules 2015, 5, 3260–3279. [Google Scholar] [CrossRef]
  221. Pranjol, M.Z.I.; Gutowski, N.J.; Hannemann, M.; Whatmore, J.L. Cathepsin D non-proteolytically induces proliferation and migration in human omental microvascular endothelial cells via activation of the ERK1/2 and PI3K/AKT pathways. Biochim. Biophys. Acta BBA-Mol. Cell Res. 2018, 1865, 25–33. [Google Scholar] [CrossRef]
  222. Seo, B.R.; Min, K.; Woo, S.M.; Choe, M.; Choi, K.S.; Lee, Y.-K.; Yoon, G.; Kwon, T.K. Inhibition of Cathepsin S Induces Mitochondrial ROS That Sensitizes TRAIL-Mediated Apoptosis Through p53-Mediated Downregulation of Bcl-2 and c-FLIP. Antioxid. Redox Signal. 2017, 27, 215–233. [Google Scholar] [CrossRef]
  223. Tripathi, M.; Nandana, S.; Yamashita, H.; Ganesan, R.; Kirchhofer, D.; Quaranta, V. Laminin-332 is a substrate for hepsin, a protease associated with prostate cancer progression. J. Biol. Chem. 2008, 283, 30576–30584. [Google Scholar] [CrossRef] [PubMed]
  224. Tanimoto, H.; Yan, Y.; Clarke, J.; Korourian, S.; Shigemasa, K.; Parmley, T.H.; Parham, G.P.; O’Brien, T.J. Hepsin, a Cell Surface Serine Protease Identified in Hepatoma Cells, Is Overexpressed in Ovarian Cancer. Cancer Res. 1997, 57, 2884–2887. [Google Scholar] [PubMed]
  225. Rosewell, K.; Al-Alem, L.; Li, F.; Kelty, B.; Curry, T.E., Jr. Identification of Hepsin and Protein Disulfide Isomerase A3 as Targets of Gelatinolytic Action in Rat Ovarian Granulosa Cells During the Periovulatory Period1. Biol. Reprod. 2011, 85, 858–866. [Google Scholar] [CrossRef]
  226. Duru, N.; Pawar, N.R.; Martin, E.W.; Buzza, M.S.; Conway, G.D.; Lapidus, R.G.; Liu, S.; Reader, J.; Rao, G.G.; Roque, D.M.; et al. Selective targeting of metastatic ovarian cancer using an engineered anthrax prodrug activated by membrane-anchored serine proteases. Proc. Natl. Acad. Sci. USA 2022, 119, e2201423119. [Google Scholar] [CrossRef]
  227. Miao, J.; Mu, D.; Ergel, B.; Singavarapu, R.; Duan, Z.; Powers, S.; Oliva, E.; Orsulic, S. Hepsin colocalizes with desmosomes and induces progression of ovarian cancer in a mouse model. Int. J. Cancer J. Int. Cancer 2008, 123, 2041–2047. [Google Scholar] [CrossRef]
  228. Lu, L.; Cole, A.; Huang, D.; Wang, Q.; Guo, Z.; Yang, W.; Lu, J. Clinical Significance of Hepsin and Underlying Signaling Pathways in Prostate Cancer. Biomolecules 2022, 12, 203. [Google Scholar] [CrossRef] [PubMed]
  229. Herter, S.; Piper, D.E.; Aaron, W.; Gabriele, T.; Cutler, G.; Cao, P.; Bhatt, A.S.; Choe, Y.; Craik, C.S.; Walker, N.; et al. Hepatocyte growth factor is a preferred in vitro substrate for human hepsin, a membrane-anchored serine protease implicated in prostate and ovarian cancers. Biochem. J. 2005, 390, 125–136. [Google Scholar] [CrossRef]
  230. Willbold, R.; Wirth, K.; Martini, T.; Sültmann, H.; Bolenz, C.; Wittig, R. Correction: Excess hepsin proteolytic activity limits oncogenic signaling and induces ER stress and autophagy in prostate cancer cells. Cell Death Dis. 2020, 11, 120. [Google Scholar] [CrossRef]
  231. Kenny, H.A.; Lengyel, E. MMP-2 functions as an early response protein in ovarian cancer metastasis. Cell Cycle 2009, 8, 683–688. [Google Scholar] [CrossRef]
  232. Morales-Vásquez, F.; Castillo-Sánchez, R.; Gómora, M.J.; Almaraz, M.Á.; Pedernera, E.; Pérez-Montiel, D.; Rendón, E.; López-Basave, H.N.; Román-Basaure, E.; Cuevas-Covarrubias, S.; et al. Expression of metalloproteinases MMP-2 and MMP-9 is associated to the presence of androgen receptor in epithelial ovarian tumors. J. Ovarian Res. 2020, 13, 86. [Google Scholar] [CrossRef]
  233. Wu, X.; Li, H.; Kang, L.; Li, L.; Wang, W.; Shan, B. Activated Matrix Metalloproteinase-2—A Potential Marker of Prognosis for Epithelial Ovarian Cancer. Gynecol. Oncol. 2002, 84, 126–134. [Google Scholar] [CrossRef]
  234. Zeng, L.; Qian, J.; Zhu, F.; Wu, F.; Zhao, H.; Zhu, H. The prognostic values of matrix metalloproteinases in ovarian cancer. J. Int. Med. Res. 2020, 48, 0300060519825983. [Google Scholar] [CrossRef]
  235. Li, L.-N.; Zhou, X.; Gu, Y.; Yan, J. Prognostic Value of MMP-9 in Ovarian Cancer: A Meta-analysis. Asian Pac. J. Cancer Prev. 2013, 14, 4107–4113. [Google Scholar] [CrossRef]
  236. Vos, M.C.; van der Wurff, A.A.M.; van Kuppevelt, T.H.; Massuger, L.F.A.G. The role of MMP-14 in ovarian cancer: A systematic review. J. Ovarian Res. 2021, 14, 101. [Google Scholar] [CrossRef]
  237. Agarwal, A.; Tressel, S.L.; Kaimal, R.; Balla, M.; Lam, F.H.; Covic, L.; Kuliopulos, A. Identification of a Metalloprotease-Chemokine Signaling System in the Ovarian Cancer Microenvironment: Implications for Antiangiogenic Therapy. Cancer Res. 2010, 70, 5880–5890. [Google Scholar] [CrossRef]
  238. Wang, F.; Fisher, J.; Fishman, D.A. MMP-1-PAR1 axis mediates LPA-induced epithelial ovarian cancer (EOC) invasion. Gynecol. Oncol. 2011, 120, 247–255. [Google Scholar] [CrossRef]
  239. Wang, L.; Kong, B. Analysis of the Association of Matrix Metalloproteinase-1 Gene Promoter (rs1799750) Polymorphism and Risk of Ovarian Cancer. Int. J. Gynecol. Cancer 2015, 25, 961–967. [Google Scholar] [CrossRef]
  240. Hobbs, C.; Huang, Z.; Murphy, S.; Berchuck, A. The function of matrix metalloproteinase 1 in ovarian cancer. Gynecol. Oncol. 2019, 154, 80. [Google Scholar] [CrossRef]
  241. Mariya, T.; Hirohashi, Y.; Torigoe, T.; Tabuchi, Y.; Asano, T.; Saijo, H.; Kuroda, T.; Yasuda, K.; Mizuuchi, M.; Saito, T.; et al. Matrix metalloproteinase-10 regulates stemness of ovarian cancer stem-like cells by activation of canonical Wnt signaling and can be a target of chemotherapy-resistant ovarian cancer. Oncotarget 2016, 7, 26806–26822. [Google Scholar] [CrossRef]
  242. Périgny, M.; Bairati, I.; Harvey, I.; Beauchemin, M.; Harel, F.; Plante, M.; Têtu, B. Role of Immunohistochemical Overexpression of Matrix Metalloproteinases MMP-2 and MMP-11 in the Prognosis of Death by Ovarian Cancer. Am. J. Clin. Pathol. 2008, 129, 226–231. [Google Scholar] [CrossRef]
  243. Hakamy, S.; Assidi, M.; Jafri, M.A.; Nedjadi, T.; Alkhatabi, H.; Al-Qahtani, A.; Al-Maghrabi, J.; Sait, K.; Al-Qahtani, M.; Buhmeida, A.; et al. Assessment of prognostic value of tissue inhibitors of metalloproteinase 3 (TIMP3) protein in ovarian cancer. Libyan J. Med. 2021, 16, 1937866. [Google Scholar] [CrossRef]
  244. Liu, M.C.P.; Choong, D.Y.H.; Hooi, C.S.F.; Williams, L.H.; Campbell, I.G. Genetic and epigenetic analysis of the TIMP-3 gene in ovarian cancer. Cancer Lett. 2007, 247, 91–97. [Google Scholar] [CrossRef]
  245. Cymbaluk-Płoska, A.; Chudecka-Głaz, A.; Pius-Sadowska, E.; Machaliński, B.; Menkiszak, J.; Sompolska-Rzechuła, A. Suitability assessment of baseline concentration of MMP3, TIMP3, HE4 and CA125 in the serum of patients with ovarian cancer. J. Ovarian Res. 2018, 11, 1. [Google Scholar] [CrossRef]
  246. Fan, X.; Wang, C.; Song, X.; Liu, H.; Li, X.; Zhang, Y. Elevated Cathepsin K potentiates metastasis of epithelial ovarian cancer. Histol. Histopathol. 2018, 33, 673–680. [Google Scholar] [CrossRef]
  247. Norman, R.J.; Dewailly, D.; Legro, R.S.; Hickey, T.E. Polycystic ovary syndrome. Lancet 2007, 370, 685–697. [Google Scholar] [CrossRef]
  248. Pasquali, R.; Casimirri, F.; Vicennati, V. Weight control and its beneficial effect on fertility in women with obesity and polycystic ovary syndrome. Hum. Reprod. 1997, 12 (Suppl. 1), 82–87. [Google Scholar] [CrossRef]
  249. Carmassi, F.; De Negri, F.; Fioriti, R.; De Giorgi, A.; Giannarelli, C.; Fruzzetti, F.; Pedrinelli, R.; Dell’Omo, G.; Bersi, C. Insulin resistance causes impaired vasodilation and hypofibrinolysis in young women with polycystic ovary syndrome. Thromb. Res. 2005, 116, 207–214. [Google Scholar] [CrossRef]
  250. Mannerås-Holm, L.; Baghaei, F.; Holm, G.; Janson, P.O.; Ohlsson, C.; Lönn, M.; Stener-Victorin, E. Coagulation and fibrinolytic disturbances in women with polycystic ovary syndrome. J. Clin. Endocrinol. Metab. 2011, 96, 1068–1076. [Google Scholar] [CrossRef]
  251. Burchall, G.F.; Piva, T.J.; Linden, M.D.; Gibson-Helm, M.E.; Ranasinha, S.; Teede, H.J. Comprehensive Assessment of the Hemostatic System in Polycystic Ovarian Syndrome. Semin. Thromb. Hemost. 2015, 42, 55–62. [Google Scholar] [CrossRef]
  252. Ranjbaran, J.; Farimani, M.; Tavilani, H.; Ghorbani, M.; Karimi, J.; Poormonsefi, F.; Khodadadi, I. Matrix metalloproteinases 2 and 9 and MMP9/NGAL complex activity in women with PCOS. Reproduction 2016, 151, 305–311. [Google Scholar] [CrossRef]
  253. Lewandowski, K.C.; Komorowski, J.; O’Callaghan, C.J.; Tan, B.K.; Chen, J.; Prelevic, G.M.; Randeva, H.S. Increased Circulating Levels of Matrix Metalloproteinase-2 and -9 in Women with the Polycystic Ovary Syndrome. J. Clin. Endocrinol. Metab. 2006, 91, 1173–1177. [Google Scholar] [CrossRef] [PubMed]
  254. Butler, A.E.; Nandakumar, M.; Sathyapalan, T.; Brennan, E.; Atkin, S.L. Matrix Metalloproteinases, Tissue Inhibitors of Metalloproteinases, and Their Ratios in Women with Polycystic Ovary Syndrome and Healthy Controls. Int. J. Mol. Sci. 2025, 26, 321. [Google Scholar] [CrossRef] [PubMed]
  255. Lahav-Baratz, S.; Kraiem, Z.; Shiloh, H.; Koifman, M.; Ishai, D.; Dirnfeld, M. Decreased expression of tissue inhibitor of matrix metalloproteinases in follicular fluid from women with polycystic ovaries compared with normally ovulating patients undergoing in vitro fertilization. Fertil. Steril. 2003, 79, 567–571. [Google Scholar] [CrossRef] [PubMed]
  256. Li, W.; Zhang, Y.; Li, F.; Shi, Y.; Wang, Y. Expression of Vascular Endothelial Growth Factor, Matrix Metalloproteinase-2 and Matrix Metalloproteinase-9 in Polycystic Ovarian Syndrome Rats and Its Implication. J. Biomater. Tissue Eng. 2021, 11, 841–846. Available online: https://www.ingentaconnect.com/contentone/asp/jbte/2021/00000011/00000005/art00006 (accessed on 22 April 2025). [CrossRef]
  257. Atiomo, W.U.; Hilton, D.; Fox, R.; Lee, D.; Shaw, S.; Friend, J.; Wilkin, T.J.; Prentice, A.G. Immunohistochemical detection of plasminogen activator inhibitor-1 in polycystic ovaries. Gynecol. Endocrinol. 2000, 14, 162–168. [Google Scholar] [CrossRef]
  258. Atiomo, W.U.; Bates, S.A.; Condon, J.E.; Shaw, S.; West, J.H.; Prentice, A.G. The plasminogen activator system in women with polycystic ovary syndrome. Fertil. Steril. 1998, 69, 236–241. [Google Scholar] [CrossRef]
  259. Tarkun, I.; Cantürk, Z.; Arslan, B.C.; Türemen, E.; Tarkun, P. The plasminogen activator system in young and lean women with polycystic ovary syndrome. Endocr. J. 2004, 51, 467–472. [Google Scholar] [CrossRef]
  260. Sahay, S.; Jain, M.; Dash, D.; Choubey, L.; Jain, S.; Singh, T.B. Role of plasminogen activator inhibitor type 1 (PAI-1) in PCOS patient. Int. J. Reprod. Contracept. Obstet. Gynecol. 2017, 6, 4052–4058. [Google Scholar] [CrossRef]
  261. Burchall, G.F.; Pouniotis, D.S.; Teede, H.J.; Ranasinha, S.; Walters, K.A.; Piva, T.J. Expression of the plasminogen system in the physiological mouse ovary and in the pathological polycystic ovary syndrome (PCOS) state. Reprod. Biol. Endocrinol. RBE 2019, 17, 33. [Google Scholar] [CrossRef]
  262. Ibáñez, L.; Aulesa, C.; Potau, N.; Ong, K.; Dunger, D.B.; De Zegher, F. Plasminogen Activator Inhibitor-1 in Girls with Precocious Pubarche: A Premenarcheal Marker for Polycystic Ovary Syndrome? Pediatr. Res. 2002, 51, 244–248. [Google Scholar] [CrossRef]
  263. Devin, J.K.; Johnson, J.E.; Eren, M.; Gleaves, L.A.; Bradham, W.S.; Bloodworth, J.R.; Vaughan, D.E. Transgenic overexpression of plasminogen activator inhibitor-1 promotes the development of polycystic ovarian changes in female mice. J. Mol. Endocrinol. 2007, 39, 9–16. [Google Scholar] [CrossRef]
  264. Shah, A.K.; Yadav, B.K.; Shah, A.K.; Suri, A.; Deo, S.K. Cardiovascular Risk Predictors High Sensitivity C-Reactive Protein and Plasminogen Activator Inhibitor-1 in Women with Lean Phenotype of Polycystic Ovarian Syndrome: A Prospective Case-Control Study. J. Lab. Physicians 2022, 15, 31–37. [Google Scholar] [CrossRef] [PubMed]
  265. Ambekar, A.S.; Kelkar, D.S.; Pinto, S.M.; Sharma, R.; Hinduja, I.; Zaveri, K.; Pandey, A.; Prasad, T.S.K.; Gowda, H.; Mukherjee, S. Proteomics of Follicular Fluid From Women With Polycystic Ovary Syndrome Suggests Molecular Defects in Follicular Development. J. Clin. Endocrinol. Metab. 2015, 100, 744–753. [Google Scholar] [CrossRef] [PubMed]
  266. Diamanti-Kandarakis, E.; Palioniko, G.; Alexandraki, K.; Bergiele, A.; Koutsouba, T.; Bartzis, M. The prevalence of 4G5G polymorphism of plasminogen activator inhibitor-1 (PAI-1) gene in polycystic ovarian syndrome and its association with plasma PAI-1 levels. Eur. J. Endocrinol. 2004, 150, 793–798. [Google Scholar] [CrossRef]
  267. Orio, F.; Palomba, S.; Cascella, T.; Tauchmanová, L.; Nardo, L.G.; Di Biase, S.; Labella, D.; Russo, T.; Tolino, A.; Zullo, F.; et al. Is plasminogen activator inhibitor-1 a cardiovascular risk factor in young women with polycystic ovary syndrome? Reprod. Biomed. Online 2004, 9, 505–510. [Google Scholar] [CrossRef]
  268. Ma, L.-J.; Mao, S.-L.; Taylor, K.L.; Kanjanabuch, T.; Guan, Y.; Zhang, Y.; Brown, N.J.; Swift, L.L.; McGuinness, O.P.; Wasserman, D.H. Prevention of obesity and insulin resistance in mice lacking plasminogen activator inhibitor 1. Diabetes 2004, 53, 336–346. Available online: https://diabetesjournals.org/diabetes/article-abstract/53/2/336/11495 (accessed on 22 April 2025). [CrossRef] [PubMed]
  269. Wang, L.; Chen, L.; Liu, Z.; Liu, Y.; Luo, M.; Chen, N.; Deng, X.; Luo, Y.; He, J.; Zhang, L.; et al. PAI-1 Exacerbates White Adipose Tissue Dysfunction and Metabolic Dysregulation in High Fat Diet-Induced Obesity. Front. Pharmacol. 2018, 9, 1087. [Google Scholar] [CrossRef]
  270. Obiezu, C.V.; Scorilas, A.; Magklara, A.; Thornton, M.H.; Wang, C.Y.; Stanczyk, F.Z.; Diamandis, E.P. Prostate-Specific Antigen and Human Glandular Kallikrein 2 Are Markedly Elevated in Urine of Patients with Polycystic Ovary Syndrome. J. Clin. Endocrinol. Metab. 2001, 86, 1558–1561. [Google Scholar] [CrossRef]
  271. Jin, M.; Cai, J.; Hu, Y.; Lu, X.; Huang, H. Cathepsin D expression in ovaries from polycystic ovarian syndrome patients. Zhejiang Xue Xue Bao Yi Xue Ban J. Zhejiang Univ. Med. Sci. 2007, 36, 429–432. [Google Scholar] [CrossRef]
  272. Dawood Salman, Z.; Ismaeel Jassim, W. Role of Cathepsin S and Insulin Resistance in Iraqi Women with Polycystic Ovary Syndrome. J. Obstet. Gynecol. Cancer Res. 2024, 9, 379–384. [Google Scholar] [CrossRef]
  273. Sabeti Akbar-Abad, M.; Majidpour, M.; Sargazi, S.; Ghasemi, M.; Saravani, R. Unraveling the Role of Cathepsin B Variants in Polycystic Ovary Syndrome: Insights from a Case-Control Study and Computational Analyses. Reprod. Sci. Thousand Oaks Calif 2025, 32, 1166–1179. [Google Scholar] [CrossRef] [PubMed]
  274. Shalev, E.; Goldman, S.; Ben-Shlomo, I. The balance between MMP-9 and MMP-2 and their tissue inhibitor (TIMP)-1 in luteinized granulosa cells: Comparison between women with PCOS and normal ovulatory women. Mol. Hum. Reprod. 2001, 7, 325–331. [Google Scholar] [CrossRef] [PubMed]
  275. Kelly, C.J.G.; Lyall, H.; Petrie, J.R.; Gould, G.W.; Connell, J.M.C.; Rumley, A.; Lowe, G.D.O.; Sattar, N. A Specific Elevation in Tissue Plasminogen Activator Antigen in Women with Polycystic Ovarian Syndrome. J. Clin. Endocrinol. Metab. 2002, 87, 3287–3290. [Google Scholar] [CrossRef] [PubMed]
  276. Pelosi, E.; Simonsick, E.; Forabosco, A.; Garcia-Ortiz, J.E.; Schlessinger, D. Dynamics of the Ovarian Reserve and Impact of Genetic and Epidemiological Factors on Age of Menopause. Biol. Reprod. 2015, 92, 130. [Google Scholar] [CrossRef]
  277. Beck-Peccoz, P.; Persani, L. Premature ovarian failure. Orphanet J. Rare Dis. 2006, 1, 9. [Google Scholar] [CrossRef]
  278. Yin, R.; Zhang, C.; Gen, A.; Li, Y.; Yang, H.; Tian, X.; Chi, Y. Propofol affects the biological behavior of ovarian cancer SKOV3 cells via ERK1/2-MMP-2/9 signaling pathway. Trop. J. Pharm. Res. 2020, 19, 233–238. [Google Scholar] [CrossRef]
  279. An, H.J.; Ahn, E.H.; Kim, J.O.; Park, H.S.; Ryu, C.S.; Cho, S.H.; Kim, J.H.; Lee, W.S.; Kim, N.K. Association between tissue inhibitor of metalloproteinase (TIMP) genetic polymorphisms and primary ovarian insufficiency (POI). Maturitas 2019, 120, 77–82. [Google Scholar] [CrossRef]
  280. Hsueh, A.J.; Billig, H.; Tsafriri, A. Ovarian follicle atresia: A hormonally controlled apoptotic process. Endocr. Rev. 1994, 15, 707–724. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of protease activation: (A) zymogen activation through proteolytic cleavage triggered by factors including specific proteases, autocatalysis, or cofactor binding; (B) activation of serine proteases featuring the Asp–His–Ser catalytic triad, which facilitates nucleophilic attack on peptide bonds; (C) activation of cysteine proteases where the thiolate group serves as the nucleophile; (D) activation of aspartic proteases involving dual aspartate residues that use water for peptide bond hydrolysis; (E) mechanisms of activation of tissue–type (tPA) and urokinase–type (uPA) plasminogen activators; (F) activation of metalloproteases through the cysteine switch and zinc-mediated catalysis.
Figure 1. Mechanisms of protease activation: (A) zymogen activation through proteolytic cleavage triggered by factors including specific proteases, autocatalysis, or cofactor binding; (B) activation of serine proteases featuring the Asp–His–Ser catalytic triad, which facilitates nucleophilic attack on peptide bonds; (C) activation of cysteine proteases where the thiolate group serves as the nucleophile; (D) activation of aspartic proteases involving dual aspartate residues that use water for peptide bond hydrolysis; (E) mechanisms of activation of tissue–type (tPA) and urokinase–type (uPA) plasminogen activators; (F) activation of metalloproteases through the cysteine switch and zinc-mediated catalysis.
Cells 14 00921 g001
Figure 2. Schematic Overview: Ovarian Follicle Development Through Corpus Luteum Formation and Involvement of Various Protease. Proteases regulate critical processes including follicular activation, differentiation, migration, ECM remodeling, and gonadotropin-dependent responses. Several extracellular matrix components (proteoglycans, laminin, and collagen) and cell types (fibroblasts and mesenchymal cells) interact with follicle cells during the developmental process. Proteases are involved in ECM remodeling, which is critical for follicle growth, oocyte maturation, and luteal phase transition.
Figure 2. Schematic Overview: Ovarian Follicle Development Through Corpus Luteum Formation and Involvement of Various Protease. Proteases regulate critical processes including follicular activation, differentiation, migration, ECM remodeling, and gonadotropin-dependent responses. Several extracellular matrix components (proteoglycans, laminin, and collagen) and cell types (fibroblasts and mesenchymal cells) interact with follicle cells during the developmental process. Proteases are involved in ECM remodeling, which is critical for follicle growth, oocyte maturation, and luteal phase transition.
Cells 14 00921 g002
Figure 3. Roles of proteases in ovarian health and disease. Matrix metalloproteinases (MMPs), tissue inhibitors of metalloproteinases (TIMPs), plasminogen activators (PAs), plasminogen activator inhibitors (PAIs), ADAMTS, cathepsins, and kallikreins are involved in critical physiological processes of the ovary, including folliculogenesis, extracellular matrix (ECM) remodeling, apoptosis, ovulation, corpus luteum (CL) development. The dysregulation of proteases can lead to several pathological conditions including cancer, polycystic ovary syndrome (PCOS), primary ovarian insufficiency (POI), and ovulatory irregularities.
Figure 3. Roles of proteases in ovarian health and disease. Matrix metalloproteinases (MMPs), tissue inhibitors of metalloproteinases (TIMPs), plasminogen activators (PAs), plasminogen activator inhibitors (PAIs), ADAMTS, cathepsins, and kallikreins are involved in critical physiological processes of the ovary, including folliculogenesis, extracellular matrix (ECM) remodeling, apoptosis, ovulation, corpus luteum (CL) development. The dysregulation of proteases can lead to several pathological conditions including cancer, polycystic ovary syndrome (PCOS), primary ovarian insufficiency (POI), and ovulatory irregularities.
Cells 14 00921 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kushawaha, B.; Pelosi, E. Spotlight on Proteases: Roles in Ovarian Health and Disease. Cells 2025, 14, 921. https://doi.org/10.3390/cells14120921

AMA Style

Kushawaha B, Pelosi E. Spotlight on Proteases: Roles in Ovarian Health and Disease. Cells. 2025; 14(12):921. https://doi.org/10.3390/cells14120921

Chicago/Turabian Style

Kushawaha, Bhawna, and Emanuele Pelosi. 2025. "Spotlight on Proteases: Roles in Ovarian Health and Disease" Cells 14, no. 12: 921. https://doi.org/10.3390/cells14120921

APA Style

Kushawaha, B., & Pelosi, E. (2025). Spotlight on Proteases: Roles in Ovarian Health and Disease. Cells, 14(12), 921. https://doi.org/10.3390/cells14120921

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

Article Metrics

Back to TopTop