Next Article in Journal
Attenuation of Oxidative Damage via Upregulating Nrf2/HO-1 Signaling Pathway by Protease SH21 with Exerting Anti-Inflammatory and Anticancer Properties In Vitro
Previous Article in Journal
A Protocol for Organoids from the Urine of Bladder Cancer Patients
Previous Article in Special Issue
Chromatin Morphology in Human Germinal Vesicle Oocytes and Their Competence to Mature in Stimulated Cycles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Advances in Cellular Approaches for Pathophysiology and Treatment of Polycystic Ovary Syndrome

1
Graduate Institute of Clinical Medical Sciences, College of Medicine, Chang Gung University, Taoyuan City 333, Taiwan
2
An-Ten Obstetrics and Gynecology Clinic, Kaohsiung City 802, Taiwan
3
Department of Chinese Medicine, Kaohsiung Chang Gung Memorial Hospital, College of Medicine, Chang Gung University, Kaohsiung City 833, Taiwan
4
Department of Biological Science, National Sun Yat-sen University, Kaohsiung 804, Taiwan
5
Center for Hormone and Reproductive Medicine Research, Department of Obstetrics and Gynecology, Kaohsiung Chang Gung Memorial Hospital, College of Medicine, Chang Gung University, Kaohsiung City 833, Taiwan
6
Division of Endocrinology and Metabolism, Department of Internal Medicine, Kaohsiung Chang Gung Memorial Hospital, College of Medicine, Chang Gung University, Kaohsiung City 833, Taiwan
*
Author to whom correspondence should be addressed.
Cells 2023, 12(17), 2189; https://doi.org/10.3390/cells12172189
Submission received: 3 July 2023 / Revised: 28 August 2023 / Accepted: 28 August 2023 / Published: 31 August 2023

Abstract

:
Polycystic ovary syndrome (PCOS) is a prevalent gynecological and endocrine disorder that results in irregular menstruation, incomplete follicular development, disrupted ovulation, and reduced fertility rates among affected women of reproductive age. While these symptoms can be managed through appropriate medication and lifestyle interventions, both etiology and treatment options remain limited. Here we provide a comprehensive overview of the latest advancements in cellular approaches utilized for investigating the pathophysiology of PCOS through in vitro cell models, to avoid the confounding systemic effects such as in vitro fertilization (IVF) therapy. The primary objective is to enhance the understanding of abnormalities in PCOS-associated folliculogenesis, particularly focusing on the aberrant roles of granulosa cells and other relevant cell types. Furthermore, this article encompasses analyses of the mechanisms and signaling pathways, microRNA expression and target genes altered in PCOS, and explores the pharmacological approaches considered as potential treatments. By summarizing the aforementioned key findings, this article not only allows us to appreciate the value of using in vitro cell models, but also provides guidance for selecting suitable research models to facilitate the identification of potential treatments and understand the pathophysiology of PCOS at the cellular level.

Graphical Abstract

1. Introduction

Polycystic ovarian syndrome (PCOS) is a common heterogeneous endocrine disorder that affects 7–12% of reproductive-age women and is significantly associated with infertility. The primary symptoms of PCOS include hyperandrogenism, absent ovulation, and polycystic ovaries. Currently, there are no definitive diagnostic criteria for PCOS, as its pathology involves multiple factors such as endocrinology and gynecology, making it difficult to categorize [1,2,3,4]. PCOS is characterized by an increased density of small pre-antral follicles and a higher proportion of early growing follicles, accompanied by abnormal granulosa cell proliferation. Additionally, PCOS is associated with the apoptosis of granulosa cells in antral follicles [5,6,7]. Granulosa cells (GCs) form a cuboidal layer on the surface of the oocyte and secrete steroid hormones that play a crucial role in follicle development, making them a key focus in PCOS pathogenesis [5,8]. During oocyte development, the oocyte and its surrounding granulosa cells exhibit a mutual interdependence, crucial for providing the oocyte with essential nutrients and growth regulators. In return, the oocyte plays a role in promoting the growth and differentiation of the granulosa cells. There is evidence to indicate that dysfunction in these cellular interactions may play a role in the abnormal folliculogenesis observed in PCOS [9]. Because the molecular mechanism responsible for triggering PCOS is unknown, it is difficult to offer guidance to women and healthcare professionals regarding the condition. Hence, discovering the specific mechanism that causes PCOS has become an area of intensive research.
In recent years, several in vitro models, including cell lines and animal models, have been developed to study PCOS due to ethical constraints in human studies. However, despite conducting numerous experiments, the underlying pathophysiology of PCOS still remains unclear. Nonetheless, the development of these experimental models holds significant potential in advancing our understanding of the condition and identifying the most effective treatment options [4]. Herein, we focus on up-to-date in vitro cell models for PCOS and summarize the mechanisms proposed or discussed in PCOS-like phenotypes in these models. We also briefly summarize the related mechanisms in hormonal and genetic alterations in developed PCOS, as well as the common pathophysiology that accompanies metabolic syndrome. We also summarize the common cellular/molecular mechanisms of drugs, compounds, and traditional Chinese herbal medicines, along with miRNA expression and specific genes implicated in PCOS pathogenesis. Finally, we discuss the effects of cells other than ovarian tissue on PCOS.

2. In Vitro Cell Models for PCOS research

Established animal models with PCOS characteristic are important for studying the pathophysiology and etiology of the disorder. Moreover, animal-derived cells or cell lines with an indefinite lifespan offer cost-effective and time-efficient alternatives for screening a broad spectrum of drugs before progressing to in-vivo models. Consequently, they provide a reliable and consistent cell source for experimentation. Figure 1 summarized the common in vitro cell models for PCOS research, as we discussed below.

2.1. Human Ovarian Granulosa Cell Lines

In vitro human cell line models have been used in fundamental research, including studies on PCOS. As outlined in Table 1, numerous human GC lines have been established successfully, including long-term cultures from human tumor cells [10,11,12,13,14,15] and immortalization through oncoviruses [16,17,18,19], lentiviruses [20,21,22,23], or retroviral systems [24,25]. GC function is defined by three distinct properties: (1) the production of 17β-estradiol in response to follicle stimulating hormone (FSH) through the action of P450 aromatase [26]; (2) support for oocyte growth during the development of the surrounding follicle [27,28]; and (3) the presence of specific molecular markers of apoptosis involved in follicle atresia [29]. To confirm the functional characteristics of these immortalized GCs, various measurement parameters were investigated for validation. GCs are essential in the ovary, playing a significant role in folliculogenesis and oocyte maturation. One of their important functions is secretion of steroid hormones, such as estrogen, progesterone, and AMH. Moreover, FSH receptor (FSHR) and aromatase are crucial markers of GCs. FSHR acts as an upstream regulator of aromatase, which plays a pivotal role in the production of estrogen by GCs through the aromatization of androstenedione. The level of sex steroid hormones in the culture medium can be measured to determine aromatase activity [24].
Most of the human GC cell lines described in Table 1 exhibit aromatase activity, although they are not all FSH-dependent. Additionally, several GC lines, including SVOG, HGL5, and HO23, are typically isolated and established by transfection of luteinized GCs obtained from IVF programs [16,17,18,19,22,23]. Due to the administration of human chorionic gonadotropin (hCG), luteinized GCs do not express a functional FSHR [30]. Both GCN and GCP, which are established from normal and PCOS GCs, express specific granulosa markers such as FSHR and aromatase, and their expressions of these markers are similar to those of primary GCs. These two GC lines provide a good in vitro model for understanding the cellular mechanisms of the normal and PCOS human ovary [24,25]. KGN, COV434, and HGrC1 were useful models because they maintain most physiological activities, such as the expression of functional FSHR and an FSH-dependent increase in aromatase activity [10,11,12,13,14,20,21]. Among these cell lines, KGN is particularly advantageous as it retains the physiologic features of GCs and exhibits a stable, long-term proliferative capability. Moreover, as it does not produce endogenous steroids and responds well to gonadotropins, KGN serves as a suitable model for the two-cell, two-gonadotropin model of ovarian steroidogenesis. Consequently, it is extensively used in research on the cellular functions of GCs and the molecular regulatory mechanisms involved in PCOS [10,19].
Table 1. Characteristics of human ovarian granulosa cell lines for PCOS research.
Table 1. Characteristics of human ovarian granulosa cell lines for PCOS research.
Cell LineGCPGCNKGNCOV434HTOGHGrC1SVOGHGL5HO23
PopulationIranianIranian.JapaneseNDJapaneseNDNDNDIsraeli
Age323663277435NDAdultAdult
CategoryTelomerase immortalized cell lineTelomerase immortalized cell lineCancer cell lineCancer cell lineCancer cell lineTransformed cell lineTransformed cell lineTransformed cell lineTransformed cell line
DiseasePCOSNormal patient with IVFOvarian granulosa cell tumorOvarian granulosa cell tumorOvarian granulosa cell tumorNDNDNDNormal patient
with IVF
TransformedhTERT/
c-Myc
hTERT/
c-Myc
NDNDNDE6/E7SV40E6/E7SV40
Doubling time75 h86 h46.4 h24 h25 h40 hND96 hND
LuteinizedND
FSH responsiveND
Androgen
responsive
NDNDNDNDNDND
AromataseNDND
P4
E2
AMHNDNDNDND
FOXL2 geneNDNDC134WWildtypeNDC134WNDNDND
Ref.[24,25][24,25][10,11,12,13,31,32][11,14,31,32][15][20,21][16][22,23,32][17,18,19,32]
ND, not determined; ✓, the feature shows in this cell line; ✗, the feature does not show in this cell line.
Both KGN and HGrC1 cell lines harbor a missense mutation in the Forkhead transcription factor gene, FOXL2 gene, which play a crucial role in the differentiation and function of the ovaries [33]. FOXL2 gene is involved in the proliferation and differentiation of GCs. Hence, there may be concerns regarding the use of KGN and HGrC1 in investigating the proliferation and apoptosis of GCs during normal folliculogenesis. By contrast, COV434, another granulosa cell tumor-derived cell line, has a wildtype FOXL2 genotype and does not express FOXL2 from a juvenile GC tumor, making it a better model for investigating these aspects [11,14,21].
Establishing appropriate in vitro GC cell line models can help in investigating the role of steroidogenesis, oogenesis, folliculogenesis, atresia, and luteinization in PCOS [19]. Many studies have used primary culture systems of GCs obtained from follicular fluids during oocyte retrieval in IVF programs. However, some limitations still exist in GC culturing such as low cell yield, difficulties in maintaining cell viability for extended cell generations and preparing uniform cell populations in sufficient amounts. By contrast, human GC cell lines have no limitations in cell number and show long-term viability and decreased proliferation.

2.2. Primary Ovary and Other Relevant Cell Lines for PCOS In Vitro Studies

Follicular fluid from patients undergoing IVF procedures is a common source of human GCs. Therefore, successfully isolating and purifying abundant levels of high-quality GCs from the follicular fluid is crucial for PCOS research [34,35]. As shown in Table 2, several human GC isolation techniques have been described in the literature, enabling us to investigate the development mechanisms of PCOS, including proliferation, apoptosis, insulin resistance (IR), and oxidative stress (OS) [36,37,38,39,40,41,42,43,44,45,46,47,48,49]. Hyperandrogenism is also a prominent feature of PCOS, characterized by excessive production of Δ4 steroids such as androstenedione (A4) and testosterone, which can lead to anovulation, oligomenorrhea, and infertility [50]. The increased expression of steroidogenic enzymes P450c17 (17α-hydroxylase, 17,20-lyase) and 3β-HSDII (3beta-hydroxysteroid dehydrogenase), which are essential for androgen production, causes enhanced androgen biosynthesis in PCOS [51,52,53]. NCI-H295R is a human adrenocortical cell line widely recognized as a valuable model for studying steroidogenesis [54,55,56,57,58,59]. These cells express all the genes that encode the steroidogenic enzymes found in all three layers of the adult adrenal cortex, including 3β-HSDII and P450c17.

2.2.1. Primary GC Isolation from Medicine-Induced PCOS Mouse Models In Vivo

Using human models for PCOS research does have limitations due to difficulties in obtaining samples, as well as ethical or logistical reasons. Therefore, based on the high degree of evolutionary conservation of the mammalian reproductive system, the development of animal models is essential for exploring the pathophysiology of PCOS both in vivo and in vitro. As shown in Table 2, PCOS models in animals can be induced using compounds such as dehydroepiandrosterone (DHEA), dihydrotestosterone (DHT), letrozole (LET), testosterone, or IL-5. Currently, there are no animal models that are precise enough to completely mirror the features of PCOS [62,63,64,65,68]. While the DHEA-induced PCOS rodent model displayed key features similar to PCOS in women, such as cystic follicles, it only induced minor metabolic changes in the rats. Notably, however, the cystic follicles in the rodent model had a thinner theca cell layer, contrasting with the thickened theca cell layer seen in human PCOS cases [75]. These findings deviate from the typical characteristics of PCOS, as women with PCOS often experience IR and glucose intolerance. The DHT-induced PCOS model is a valuable tool for examining the mechanisms of altered hormonal regulation and ovarian changes and can be used to investigate various aspects of PCOS such as ovarian function, pathophysiology, metabolic disturbances, and treatment options. This model closely mirrors most of the reproductive and metabolic alterations observed in PCOS women, making it a suitable choice for PCOS research [76,77]. With immunoassay-based techniques, testosterone levels in premenopausal women have been found to be about 40 ng/dL (1.4 nmol/L) and DHT levels about 10 ng/dL (0.34 nmol/L) [60], there were also testosterone mouse model which represent most of the PCOS phenome. In LET-induced PCOS models, metabolic dysregulation such as weight gain increased abdominal adiposity, elevated fasting blood glucose and insulin levels, and IR has been observed [78]. However, the pubertal LET model differs from the organizational effects of androgens during prenatal development, which result in permanent changes in the brain, including alterations in gonadotropin releasing hormone (GnRH) neurons and anxiety-like behavior. Despite this, hyperandrogenism, LH hypersecretion, polycystic ovaries, and a lack of corpora lutea are features found in LET-induced PCOS models and even in ER-α (ESR-1) knockout mice [65,79,80,81]. Therefore, the key features observed in LET-induced models of PCOS may be caused by impaired estrogen action, rather than the hyperandrogenic condition.

2.2.2. Primary GCs Isolated from Normal Animals and Treated with Medicine In Vitro

One of the most common methods of primary GC culture isolation is to isolate cells from ovaries in PCOS rodent models. However, several studies collected GCs from healthy animals and treated them with DHT, IL-15, and H2O2 [46,60,61,66,67,69,70]. Numerous studies have also demonstrated that ovarian GCs in patients with PCOS experience imbalanced OS, suggesting that autophagy of GCs triggered by OS may play a crucial role in the development and onset of PCOS pathology [46,82]. H2O2 has been extensively employed as an exogenous inducer in many studies to investigate damage mechanisms. By inducing H2O2 in rat GCs, researchers have established a model for investigating damage mechanisms such as apoptosis and autophagy that result from OS [60]. It is important to note that the antioxidant system in tissue is more intricate than it is at the cellular level. Therefore, further research is necessary to better understand the complexity of the antioxidant system in tissue and its potential implications in the context of OS and for conditions such as PCOS. Low-grade chronic inflammation may play a role in the development of PCOS, with the proinflammatory cytokine IL-15 contributing to the development of chronic inflammation, which leads to obesity-associated metabolic syndrome. Higher levels of IL-15 in follicular fluid were found in PCOS patients and PCOS animal models, although IL-15 levels in the serum did not increase significantly. It appears, therefore, that IL-15 is produced locally or accumulated in ovarian tissue, such as GC cells. Consequently, GC cells cultured with IL-15 could potentially be used to further explore the pathogenesis of PCOS by affecting the inflammation state, steroidogenesis, and GC survival rates [67].

2.2.3. Other Relevant Cell Lines for PCOS In Vitro Studies

Theca cells (TCs) play a key role during follicular growth and atresia and are one of the most important cell types in the follicles. Their functions include synthesizing androgens, promoting the function of GCs, facilitating oocyte development, and providing structural support for the follicle [83]. Studies have shown that IL-18-dependent regulation of proliferation and steroidogenesis in TCs may influence follicle development and result in similar pathologic features to PCOS [71]. Another study investigated the effects of androgen and the mechanism of Col6a5 in excess lipid accumulation and cell hypertrophy in ovarian stromal cells under a DHT-induced hyperandrogenic mouse model and hyperandrogenic cell models. However, it did not investigate mouse liver cells [72].

2.3. Mouse Primary Follicle Cultures

Another method for PCOS studies in vitro is follicle isolation and culture. The ovarian follicle is the fundamental functional unit of the ovary, comprising an oocyte, and surrounding GCs and TCs. The process of folliculogenesis is crucial for the generation of competent oocytes that can be fertilized and developed properly. This process involves the regulation of multiple signals, including hormonal regulation, paracrine signals, and bidirectional communication between the oocyte and the GCs [84]. Follicular development progresses from primordial follicles, primary follicles (PMF), secondary follicles (SF), preantral follicles, antral follicles (AF), and eventually to mature follicles [85]. The abnormal morphology of polycystic ovaries is linked to abnormal ovarian follicle development, which includes increased activation of PMF, ostensible accumulation of follicles at the primary stage, greater proliferation of GCs in small preantral follicles, and follicle arrest at the antral stage [86,87,88,89]. This indicates that ovarian impairment in PCOS originates during the earliest phases of follicle development, when gonadotrophin action is not essential, and where local factors such as androgens play a more prominent role in regulating follicle development [90]. Hence, the in vitro follicle culture system provides an opportunity to study folliculogenesis, oocyte maturation, and pathophysiology.
Hyperandrogenism is detected in PCOS patients. Many studies have suggested that both endogenous and exogenous sources of androgen excess may directly contribute to the development of polycystic ovary morphology [91]. Table 3 has summarized the various rodent primary follicle cultures that have been used to investigate the stimulation of follicle growth by androgens [84,90,92,93,94]. Among the strains of rodents provided by major suppliers, C57BL/6 and ICR mice are commonly used in studies related to aging, toxicology, pathology, and physiology [95]. These two rodent strains are not only easy to handle for their size but also have high fertility. They are also commercially available and widely used for various studies including in vitro culturing of ovarian follicles. In order to investigate follicle development, isolated growing follicles need to be collected from the ovaries of three-week-old C57BL/6 mice or six-week-old ICR mice. DHEA is a precursor to potent androgens and estrogens. In vitro, DHEA supplementation inhibits follicle growth and steroid hormone synthesis. However, blocking AR signaling can reverse the inhibition of follicle growth and ovulation by DHEA [84]. By contrast, DHT significantly promoted secondary follicle growth by potentiating FSH action. Administering DHT to SF in mice with low FSH increased FSHR expression and promoted follicular growth by enhancing FSH action [92]. Testosterone also promoted follicle development in rodent models during the early stages (D7) but inhibited follicle growth at a later stage (D7-14) [94].
In general terms, improper androgen production or conversion to estrogen can lead to hyperandrogenic conditions in PCOS cases [96,97]. The dysfunctional synthesis of androgens and estrogens in PCOS mainly originates from ovarian follicle theca and GCs, resulting in thickened theca and thin granulosa cell layers under polycystic conditions [98]. At the preantral stage, various androgens promote the growth of isolated mouse preantral follicles [99,100,101]. Androgen excess altered the growth patterns of the mice and this hyperandrogenic environment may have contributed to the antral follicle arrest observed in polycystic ovaries [90]. Therefore, rodent primary follicle cultures serve as a good platform to investigate how a hyperandrogenic environment can directly and permanently affect follicular development and ovarian function.

3. Mechanisms and Signaling Pathways Identified from In Vitro Models of PCOS

PCOS is a complex endocrine disease that involves endocrine disorders, androgen excess, infertility, IR, obesity, and glycometabolism disorder [102]. The PI3K-AKT signaling pathway was previously thought to be related to proliferation, apoptosis, and transfer in many diseases, including cancer. However, recent PCOS studies have shown that changes occur in the PI3K-AKT signaling pathway in response to IR, androgen excess, obesity, and follicular development [103]. In addition, emerging evidence indicates that the transforming growth factor beta (TGFβ) signaling pathway and NF-κB signaling pathway also play key roles in regulating the physiological conditions of PCOS [40]. Table 4 summarizes the most important molecular mechanisms involved in PCOS pathogenesis, providing a new perspective on the pathogenesis of PCOS and GC dysfunction.

3.1. Proliferation and Apoptosis

GCs are necessary for follicular development and oocyte ovulation as they provide nutrients and growth regulators. Therefore, decreases in GC proliferation and increases in GC apoptosis appear to be significantly involved in the pathogenesis of PCOS [104,105]. The PI3K/AKT/mTOR pathway is an important signaling pathway for cell cycle regulation and cell proliferation [106]. Several studies have indicated that the PI3K/AKT/mTOR pathway plays a role in regulating the proliferation or apoptosis of GCs in follicles (see Table 4) [36,70,107,108,109,110]. In addition, regulators involved in the PI3K/AKT/mTOR pathway, such as ribosomal protein S6 kinase 1 (S6K1), IGF-1, TLR4, and PTEN, may also be associated with the apoptosis/proliferation ratio in GCs of PCOS [36,61,107,109]. The TGFβ signaling pathway is another common pathway involved in multiple cellular processes such as including proliferation, migration, adhesion, and apoptosis, and its dysregulation has been associated with pathological conditions such as PCOS [37,111,112]. Dysregulated miRNAs, including miR-33b, miR-142, miR-125b, miR-203a, and miR-423, have been identified in GCs from PCOS patients and may be involved in regulating the TGFβ-signaling pathway, promoting cell proliferation, and repressing apoptosis [37,40,41]. Furthermore, expression of components in the hedgehog (Hh) signaling pathway has been observed in the GCs of PCOS patients as well as in ovarian tissues in mice [49,113]. Inhibiting the Hh signaling pathway has been shown to decrease GC apoptosis in PCOS, indicating that aberrant activation of the Hh signaling pathway is related to abnormal follicular development in PCOS patients [49]. Recent evidence from fish models suggests that androgens can trigger apoptosis in granulosa/theca (G/T) cells [73,114]. Testosterone activates the membrane receptor ZIP9, which induces G/T cell apoptosis through a mechanism involving Gsα subunit signaling and extracellular signal-regulated kinase (ERK) pathway activation. This mechanism coordinates androgen-induced G protein signaling pathways with zinc signaling to facilitate apoptosis, a crucial cellular function, in vertebrate cells [73].
Table 4. Pathological mechanisms and signaling pathways associated with PCOS in various in vitro models.
Table 4. Pathological mechanisms and signaling pathways associated with PCOS in various in vitro models.
Functional OutcomeAffected PathwayCells/FolliclesSource/TreatmentRef.
Proliferation/ApoptosisAKT/mTOR Human primary GCsPCOS patients[107]
Proliferation/ApoptosisPI3K/AKT/mTORRat primary GCsND[108]
Proliferation/ApoptosisPI3K/AKT/mTORKGNND[109]
Proliferation/ApoptosisPI3K/AKT/mTORKGNDHEA[47,115]
Proliferation/Apoptosisp53/TGF βHuman primary GC/KGNPCOS patients[40,41]
ProliferationTGF βHuman primary GCsPCOS patients[37]
ProliferationPI3K/AKTHuman primary GC/KGNmiR-let-7d-3p OE[36]
ProliferationPPARγ/PTEN/p-AKTRat primary GCsDHEA/DHT/FSH[61]
ApoptosisHedgehog signalingHuman primary GCsPCOS patients[49]
Apoptosis/AutophagyIGF1/p-AKT/BimELPorcine primary GCsFSH/PDTC[70]
ApoptosisGsα/MAPKfish primary G/T cellsTestosterone[73]
AutophagyPI3K/AKT/mTORCOV434ND[116]
AutophagyPI3K/AKT/mTORRat primary GCsND[60,64]
AutophagyPI3K/AKT/mTORKGNDHEA [47,115]
Autophagyp53/AMPKRat primary GCsTP[60,64]
AutophagyNF-κB/JNK Porcine primary GCsFSH/PDTC[69]
Insulin ResistanceIRS1/PI3K/AKTRat primary GCsDHEA-treated[62]
Insulin ResistanceIGF-1/PI3K/AKT/Bax/Bcl-2Rat primary GCsHFD/Letrozole[65]
Insulin ResistancePTEN/AKT/TLR2/TLR4/NF-κBHuman primary GCsPCOS patients[42,43]
Mitochondrial damage
Glycolysis defection
Insulin receptor/PI3K/AKT/ERK KGNSIRT3 KD/Insulin[44]
MitophagyPINK1/ParkinHuman primary GC/KGNDHT[48]
Mitochondrial dysfunctionPDK1/AKTKGNDHT[110]
Oxidative stressPI3K/AKT/mTORRat primary GCsH2O2[60]
Oxidative stressAMPK/AKT/Nrf2Human primary GCsPCOS patients[117]
Oxidative stressKeap1/NrF2Human primary GC/
Rat primary GCs
Vitamin K3[46]
Estrogen synthesisWNT2/FZD3/β-cateninCOV434FSH[118]
Follicular developmentPKA pathwayMice folliclesFSH[119]
Abbreviations: mTOR, mammalian target of rapamycin; OE, over expression; PPARγ, peroxisome proliferator-activated receptor gamma; BimEL, bcl-2-interacting mediator of cell death-extra long; AMPK, AMP-activated protein kinase; JNK, Jun amino-terminal kinase; Bax, Bcl-2 associated X; Bcl-2, B-cell lymphoma 2; HFD, high-fat diet; TLR2, toll-like receptor 2; KD, knock down; PINK1, PTEN-induced kinase 1; PDK1, phosphoinositide dependent proteinkinase 1; WNT2, Wnt family member 2; FZD3, frizzled 3; PKA, protein kinase A.

3.2. Autophagy

Distinct from apoptosis, autophagy is a cellular process that induces cell death and can facilitate the pathological progression of various diseases [120]. Increasing numbers of studies have revealed that obvious autophagy occurs in GCs in both humans and rats with PCOS and may be one of its primary causes [121]. The PI3K/AKT/mTOR signaling pathway is a classic pathway involved in the regulation of autophagy, and many studies have shown that the activation of the PI3K/AKT/mTOR signaling can inhibit autophagy in both human and rat PCOS models [47,60,116]. The tumor suppressor p53 is another modulator of autophagy in regulating cell death and survival [122]. One study used testosterone propionate to induce a PCOS cell model in vitro by activating autophagy and also showed that autophagy of GCs is inhibited by mediating the p53/AMPK signaling pathway [64]. FSH has been shown to regulate autophagy in rat GCs by activating NF-κB via PI3K/AKT/mTOR signaling [123,124]. By contrast, in porcine GCs, FSH inhibits NF-κB and then promotes autophagy via the JNK pathway, which provides new insights into the crosstalk between different signaling pathways during follicle development [69].

3.3. Insulin Resistance

Another key characteristic of PCOS is IR. Several studies have shown that IR regulates multiple mediators and pathways involved in the pathogenesis and development of PCOS [125]. Table 4 shows that insulin primarily regulates the PI3K/AKT signaling to mediate its metabolic regulation effect in both human and rat primary GCs [43,62,65]. Insulin sensitizers, such as humanin analog (HNG) and metformin, suppress IR by regulating the PI3K/AKT signaling pathway [62]. Furthermore, Cangfudaotan decoction, a Chinese medicine used to treat gynecological diseases, particularly PCOS, regulates the IGF-1-PI3K/AKT-Bax/Bcl-2 pathway to alleviate IR, improve follicular development, and inhibit apoptosis [65]. Additionally, SAA1 overproduction may contribute to IR development in GCs in PCOS patients by stimulating the TLR2/TLR 4 and NF-κB pathways [42].

3.4. Oxidative Stress and Mitochondrial Dysfunction

Numerous studies have increasingly linked low-grade systemic and ovarian chronic inflammation, as well as OS, to PCOS and its associated endocrinological dysfunction [46,126,127]. This indicates that imbalanced OS found in GCs may be an important factor in the development of PCOS pathology. Several studies have reported that OS contributes to the pathophysiology of PCOS through various signaling pathways (see Table 4). In a rat PCOS model, increasing OS-induced GCs autophagy through the PI3K/AKT/mTOR pathway can be ameliorated by treatment with metformin [60], while the increased production of reactive oxygen species (ROS), which causes OS, induced apoptosis in granulosa-lutein cells via activation of the AMPK/AKT/Nrf2 signaling pathway [117]. The Keap1/Nrf2 pathway has recently been identified as the pivotal pathway regulating OS. Nrf2 is a key molecule that becomes activated in response to OS and is usually sequestered by cytoplasmic Keap1 and targeted for proteasomal degradation under basal conditions. Humanin downregulation has been observed in ovarian GCs of PCOS patients and is associated with oxidative imbalance through modulation of the Keap1/Nrf2 signaling pathway [46,128]. Recent studies have indicated that mitochondrial injury in GCs, which is associated with PCOS pathogenesis, is linked to excessive OS [115,129,130]. The PI3K/AKT signaling pathway is one of the major pathways that help stabilize mitochondrial membrane potential and prevent mitochondrial membrane defects, thereby maintaining their primary biological functions. Zheng et al. demonstrated that GCs in PCOS patients contained damaged mitochondrial membranes, and that melatonin might activate PDK1/AKT by promoting SIRT1 expression to repair the damage [110]. By contrast, Yi et al. found that decreased SIRT1 expression in the GCs of PCOS might have caused excessive mitophagy and mitochondrial injury. In addition, melatonin was found to protect against mitochondrial injury in GCs of PCOS by enhancing SIRT1 expression to inhibit excessive PINK1/Parkin-mediated mitophagy [48].

3.5. Other Physiological Characteristics and Signaling Pathway

FSH enhances the differentiation capabilities of GCs, including their ability to produce E2 and facilitate preantral follicle growth [131]. Recent research has suggested that Wingless-type MMTV integration site family members (WNTs), working in conjunction with their frizzled (FZD) receptors, contribute to regulating normal folliculogenesis, luteogenesis, and ovarian steroidogenesis. In the cumulus cells (CCs) of patients with PCOS, FZD3 expression was found to be significantly upregulated, which, when coupled with the activation of the WNT2/β-Catenin pathway, was strongly linked to IR and estrogen deficiency. Thus, excessive FZD3 expression in CCs may act as an impediment to steroidogenic activation, which is normally overcome by FSH stimulation [118]. Numerous studies have also provided evidence that FSH can impact the growth of preantral follicles, through in vivo experiments and in vitro culture of ovarian tissue explants or isolated, multilayered preantral follicles. Moreover, FSH-induced growth was found to be suppressed by a PKA inhibitor, indicating that the PKA pathway is involved in FSH-induced follicle growth [119].

4. MicroRNA Expression and TARGET Genes in In Vitro Models of PCOS

MicroRNAs (miRNAs) are small, non-coding RNA molecules consisting of approximately 20–22 nucleotides, which negatively regulate target gene expression at the post-transcriptional level by imperfectly base pairing with the 3’-untranslated region (UTR) of target mRNAs. Numerous studies have demonstrated the impact of miRNAs on various biological processes, including development, differentiation, cell proliferation, apoptosis, metabolism, inflammatory responses, and various diseases [132]. miRNA dysregulation is linked to various pathophysiological processes, including PCOS, with increasing evidence suggesting that abnormal expression of miRNAs in GCs plays a crucial role in the onset and progression of PCOS [133,134,135].
Moreover, PCOS patients exhibit distinct patterns of miRNA expression during ovarian steroidogenesis. In women with PCOS, numerous miRNAs are modified in their serum, granulosa-lutein cells (GLCs), and follicular fluids, regulating key processes such as follicular development and maturation, insulin signaling, glucose and lipid metabolism, and steroid hormone synthesis [136,137,138]. A better understanding of how genetic regulation and environmental factors interact to produce varying miRNA expressions may provide valuable insights into the development of PCOS. miRNAs also have the potential to serve as noninvasive biomarkers for PCOS diagnosis and classification. However, the current understanding of the relationship between miRNAs and PCOS development is limited, mainly because a single miRNA can target numerous mRNA molecules and vice versa. Further functional studies exploring the connection between miRNAs and PCOS will be necessary [139].
According to evidence, isolated CCs from PCOS patients exhibited differential expression of several miRNAs compared with controls [140]. Several miRNAs have been shown to increase significantly in GCs from PCOS patients or cell lines. For example, in rat models of PCOS, and in the ovarian GCs of PCOS patients, miR-194 expression was found to be significantly upregulated. Suppressing miR-194 promoted the growth and proliferation of KGN cells, while its over-expression induced cell apoptosis [141]. The upregulation of miR-3188 levels in PCOS patients may enhance cell viability and progression of the cell cycle while suppressing cell apoptosis, an effect that can be achieved through the downregulation of KCNA5 (potassium voltage-gated channel subfamily A member 5) [142]. The expression of miR-21 and toll-like receptor 8 (TLR8) was significantly elevated in granulosa cells of PCOS patients when compared with normal GCs. miR-21 promotes the translation of TLR8 mRNA, leading to increased secretion of IFN-γ, TNF-α, and IL-12, which suggests that miR-21 and TLR8 are involved in PCOS-related inflammation [143]. In PCOS patients, miR-186 and miR-135a were found to be overexpressed. Song et al. identified ESR2 as a direct target of both miR-186 and miR-135a in GCs, establishing a link between dysregulated miRNAs and GC dysfunction in PCOS patients [39] (Table 5).
Several miRNAs were oppositely expressed in GC cells. miR-451a is downregulated in KGN cells, and its upregulation may inhibit GC proliferation, which could be a contributing factor to the development of abnormal follicles in PCOS patients. Additionally, miR-451a regulates the proliferation and apoptosis of ovarian GCs by targeting ATF2 [144]. Both PCOS patients and KGN cells showed downregulation in miR-206 expression. This miRNA targets CCND2, which functions as a negative regulator of miR-206. The regulation of cell viability and apoptosis of ovarian GCs by miR-206 suggests that it may play a critical role in PCOS pathogenesis [145]. High insulin concentrations were found to decrease miR-19b expression levels, promote cell proliferation, and increase IGF-1 levels. Moreover, both PCOS ovary tissues and KGN cells exhibited significantly lower expression levels of miR-19b. Additionally, miR-19b directly targeted IGF-1 and functioned as a negative regulator of its expression. The over-expression of IGF-1 was found to promote cell proliferation [146]. miR-323-3p downregulation was observed in human GCs of women with PCOS, while the inhibition of miR-323-3p levels upregulated steroidogenesis and promoted apoptosis in KGN cells. This miRNA was found to inhibit steroidogenesis and GCs apoptosis by targeting IGF-1, suggesting it may play a role in the development of PCOS [147]. The expression levels of miR-21 were found to be downregulated in the ovarian tissue of PCOS patients and KGN cells. Over-expression of miR-21 inhibited the proliferation of KGN cells and induced apoptosis. The miR-21/SNHG7 axis was shown to play a role in regulating GCs proliferation and apoptosis [148]. The downregulation of miR-320a expression in primary GCs from PCOS patients is associated with estrogen deficiency. IGF1 plays a role in regulating miR-320a expression in GCs. Through direct targeting of the 3’UTR of the osteogenic transcription factor RUNX2, miR-320a potentiates steroidogenesis in GCs by modulating the expression of CYP11A1 and CYP19A1. This suggests that miR-320a may play a critical role in the pathogenesis of PCOS [150]. miR-29a downregulation was observed in PCOS patients and was found to be correlated with an increase in the antral follicle count. The over-expression of miR-29a in KGN and COV434 cells led to the inhibition of cell proliferation, arrested cell cycle progression, and reduced aromatase expression and estradiol production. These findings suggest that miR-29a plays a critical role in GC proliferation and steroidogenesis and could provide new insights into the pathogenesis of PCOS [149].
In light of these findings, miRNAs appear to have the potential to serve as clinical biomarkers for diagnosing PCOS and as therapeutic targets for treating the condition. It may also partly explain the heterogeneity observed in PCOS women. However, despite recent progress, research surrounding miRNAs as potential diagnostic tools is still in the early stages. Additionally, while there is potential for miRNA-based therapeutics, they are yet to be developed. As such, developing commercially available miRNA-based diagnostics and therapeutic tools remains a long-term goal [152,153].

5. Pharmacological Approaches in In Vitro Models of PCOS

PCOS has a complex pathophysiology, with an etiology involving multiple factors such as genetics, diet, environment, or even social psychology [47,154]. Current treatments for PCOS tend to focus on symptom relief instead of addressing or curing the root cause. The most commonly used medications and compounds for PCOS therapy are summarized in Table 6. Metformin is a drug usually used to treat patients with PCOS and can improve hyperandrogenism and induce ovulation [60]. In addition to the drugs, there are currently many studies investigating the apoptosis or cell viability of GCs using different nutritional supplements, which could be a new therapeutic modality for PCOS treatment. Another category, Traditional Chinese medicine (TCM) and Herbal Compounds, has its role in PCOS treatment. TCM, a notable component of complementary and alternative medicine, has been used to manage PCOS for centuries. TCM prescriptions comprise multiple active compounds that have been shown to alleviate clinical symptoms and abnormal laboratory data of PCOS without significant side effects [155]. Formulas such as Cangfudaotan, Gui Zhu Yi Kun, Guizhi Fuling Wan, Xiao Yao San, Bu Shen Tian Jing, Bushen Huatan, Kunling Wan, and Gengnianchun have been widely used to treat PCOS in human and animal models. These TCM formulas can inhibit GC autophagy, reduce cell apoptosis, promote normal follicular development, regulate hormone aberrations, ameliorate irregular estrous cycles, reverse ovarian aging, and improve pregnancy outcomes [64,65,156]. Some compounds from herbs, such as cryptotanshinone (CRY) and curcumin, can inhibit granulosa cell apoptosis, regulate OS, improve hormone imbalance, reduce inflammation, and correct disturbances in the estrous cycle in cell and animal models [38,157,158,159]. Additionally, plumbagin exhibits an inhibitory effect on the proliferation and viability of various cancer cell lines and induces a significant concentration-dependent inhibition of rat ovarian GC proliferation. Moreover, a study reported that plumbagin reduce the pyroptosis of GC in PCOS mouse by inhibiting the Wilms tumor 1-associated protein-mediated N6-methylation of C-terminal caspase recruitment domain mRNA. Those study supports the profound potential of plumbagin in PCOS treatment [108,160,161]. All these data show that herbal medicine is an acceptable alternative therapy for PCOS patients. The current therapeutic mechanisms of pharmacological agents for PCOS, as mentioned in Section 4, are highly diverse. Therefore, this section focuses on in vitro studies to better understand their effects.

5.1. Small Molecular Drugs

Metformin

Metformin is one of the most widely used insulin-sensitizing drugs in the treatment of PCOS. It has also been shown to reduce hyperandrogenism and induce ovulation [166]. However, the use of metformin as an adjunct is limited and only favorable in the treatment of PCOS women who are resistant to CC alone [167,168]. In a rat model, metformin was shown to improve PCOS by decreasing excessive autophagy in primary culture GCs. Metformin also reduces the levels of OS and autophagy in H2O2-induced GCs through regulating the PI3K/AKT/mTOR signaling pathway [60]. Organic cation transporters (OCTs) OCT1, OCT2, and OCT3 control metformin uptake into primary culture GCs of rats, where metformin can decrease vascular endothelial growth factor (VEGF) and increase pAMPK levels. When OCTs were inhibited, these effects were reversed. Metformin has also been shown to act directly on ovarian cells by regulating cell metabolism and VEGF expression [169].

5.2. Nutritional Supplement

5.2.1. Melatonin

Secreted by the pineal gland, melatonin is a neuroendocrine hormone that plays a crucial role in regulating the reproductive functions of mammals [170]. It is a therapeutic agent that regulates autophagy and also has the potential to suppress autophagy and apoptosis in PCOS [47,171]. Additionally, studies have shown that melatonin can significantly reduce androgen levels and increase FSH levels in PCOS patients. Moreover, it has a positive effect on oocyte quality in cases of hypoestrogenia and hyperandrogenia [45,172].
Melatonin is not only used in PCOS treatment, it is also a mitochondria-targeted antioxidant and might repair mitochondrial damage [43,48,110,173]. In both GCs from PCOS patients and DHT-treated KGN cells, melatonin was found to increase the expression of SIRT1 and decrease the expression of PINK1/Parkin at the protein level, which improves mitochondrial dysfunction [48]. Melatonin decreases mitochondrial permeability transition pore (mPTP) opening and increases the JC-1 aggregate/monomer ratio, indicating that regulating mPTP can enhance mitochondrial membrane potential. It reduces the levels of cytochrome C and Bax both in vivo and in vitro while increasing the phosphorylation of PDK1 and AKT. This activation of the PDK1/AKT pathway is crucial in improving mitochondrial membrane function. Additionally, melatonin treatment increases the expression of SIRT1, while knocking down SIRT1 mRNA inhibits the protective effect of melatonin on the activation of PDK1/AKT [110].
Melatonin has been shown to have multiple effects on luteinized GCs of PCOS patients. Firstly, it upregulates the expression of CYP19A1 via the ERK pathway, which accelerates the conversion of androgen to 17β-estradiol. Secondly, it reduces the levels of inducible nitric oxide (NO) synthetase and NO in GCs. Thirdly, it increases the level of transcripts encoding Nrf2 and its downstream target HO1, which results in anti-inflammatory and antioxidant effects [45].
Melatonin has been shown to improve glucose uptake and insulin signaling in both GCs from PCOS patients and SVOG cells that were treated with palmitic acid. In these cells, melatonin increases IRS-1 and GLUT4 expression while decreasing p-IRS-1 (Ser307) expression. Palmitic acid has been shown to inhibit PI3K and AKT phosphorylation, but melatonin increases the levels of p-PI3K and p-AKT while decreasing IR via the PI3K/AKT signaling pathway in GCs and palmitic acid-induced SVOG cells [43].

5.2.2. Humanin

Humanin, a peptide derived from mitochondria, is involved in metabolic processes such as diabetes and PCOS. Wang et al. showed that local ovarian expression of humanin was downregulated in PCOS patients with IR compared with its expression in PCOS patients without IR. They also showed that exogenous humanin supplementation could improve body weight gain, ovarian morphological abnormalities, endocrinological disorders, and ovarian and systemic OS in PCOS rat models induced by DHEA [46].
Activating the Keap1/Nrf2 signaling pathway in response to OS is one mechanism by which humanin exerts its protective effects on ovarian GCs in patients with PCOS. By modulating this pathway, humanin can alleviate OS in GCs [46]. A study conducted that supplementation with a humanin analog could reduce elevated levels of fasting plasma glucose and fasting insulin in PCOS rat models induced by DHEA [62]. Moreover, exogenous humanin supplementation has been shown to improve the metabolic profile of PCOS rats by targeting the IRS1/PI3K/AKT insulin signaling pathway, and it decreased the phosphorylation of IRS1, PI3K, AKT, and GLUT4 proteins in the primary culture of ovarian GCs [62]. These results suggest that humanin has great potential as a therapeutic drug for patients with PCOS.

5.2.3. Vitamin D3

In PCOS patients, Vitamin D deficiency can lead to excessive androgen secretion, IR, and disrupted follicular growth. However, Vitamin D3 can increase the number of preantral and antral follicles in a DHEA-induced rat PCOS model compared with a healthy group [68]. Vitamin D3 was shown to improve mitochondrial biogenesis, membrane integrity, and mtDNA copy number in primary culture GCs of PCOS mice induced with DHEA, potentially enhancing follicular development and oocyte quality [68].

5.2.4. Sulforaphane

Sulforaphane has anti-tumor, immunoregulatory, and antioxidative effects, and can protect against OS by lowering the level of intracellular ROS and apoptosis in the GLCs of PCOS patients [117]. Sulforaphane activates the AMPK/AKT/Nrf2 signaling pathway, providing protection against OS in the GLCs of PCOS patients [117].

5.3. Traditional Chinese Medicine

An in vitro Gui Zhu Yi Kun formula study showed that primary culture GCs treated with Gui Zhu Yi Kun formula can increase in mTOR, phosphorylated mTOR, and AMPKα expression levels, and reduce p53 and sestrin2 expression levels. This finding implies that a decrease in ovarian primary culture GCs in rats with PCOS may be related to autophagy [64]. The Cangfudaotan decoction has been shown to improve IR and diminish ovary morphological damage, normalize abnormal serum hormone levels, and inhibit inflammatory cytokines in a PCOS rat model [65]. In the primary culture GCs studies, Cangfudaotan can improve cell viability and inhibit cell apoptosis, which is associated with the regulation of IGF-1-PI3K/AKT-Bax/Bcl-2 pathway-mediated gene expression [65]. Guizhi Fuling Wan has fewer atretic and cystic follicles, and more mature follicles and corpus lutea, as well as lower serum T, LH, LH/FSH ratios, HOMA-IR, and FINS levels in the PCOS rat model [164].
Other studies have demonstrated that Guizhi Fuling Wan can inhibit autophagy in primary culture GCs and promote follicular development to attenuate ovulation disorder in PCOS-IR, which is associated with the activation of the PI3K/AKT/mTOR signaling pathway [156]. Xiao Yao San can reduce the level of autophagy in GCs caused by noradrenaline and can also mitigate the autophagy of primary culture GCs through the AKT/mTOR/S6K1 pathway [162]. Additionally, Xiao Yao San has been shown to inhibit the apoptosis of primary culture GCs in the AF and the autophagy of GCs in the antral and cystic follicles in a chronic, unpredictable mild stress-induced PCOS rat model [162]. Bu Shen Tian Jing can ameliorate glucose tolerance, the estrous cycle, and ovarian morphology in the PCOS rat model, while Bu Shen Tian Jing relieves OS and increases SIRT3 expression of ovarian GCs [163]. In the above study on the KGN cell model, Bu Shen Tian Jing was shown to reverse palmitate-induced impaired SIRT3 expression and glucose uptake and decrease palmitate-induced mitochondrial ROS production mediated by SIRT3 [163]. Both Bushen Huatan and Kunling Wan can improve pregnancy outcomes, improve the viability of primary culture GCs and decrease the apoptosis of primary culture GCs in a PCOS rat model [164]. The study also showed that Bushen Huatan granules inhibit apoptosis by attenuating the dysfunctional mitochondrial of primary culture GC, while Kunling Wan can relieve endoplasmic reticulum stress [164]. Gengnianchun can decrease IR, reduce damage to the ovarian reserve, and reduce aging-related mRNA and protein levels such as p53, p16, and p21 in a mice PCOS model [165]. In a long-term insulin-treated KGN cells model, which results in senescence, differently expressed genes are mainly enriched to the ERK1 and ERK2 cascade pathway, which is an integral component of the plasma membrane and signals a level of receptor binding [165].

5.4. Herbal Compounds

In vitro models of compounds from herb plants, such as CRY, have been shown to reduce OS, decrease high mobility group box 1 (HMGB1) and Bax expression of primary culture GCs induced by ischemia-reperfusion through inhibiting ferroptosis, reduce GPX4 expression, and activate NF-κB via the MAKP signaling pathway [157]. Another study of CRY demonstrates that CTBP1-AS is highly expressed in PCOS patients. CTBP1-AS interacts with EZH2 and EED in primary culture GCs and balances the proliferation and apoptosis of primary culture GCs in PCOS patients. CRY has been shown to reduce the level of CTBP1-AS in a KGN cell model [38] and has also been shown to attenuate increases in body weight, ovarian quotiety, Lee’s index, and body mass index in a PCOS rat model [158]. CRY also reduces the proliferation of primary culture GCs and modulates TNF-α, TLR4, NF-κB/p65, and HMGB1 expression in vitro [158].
Curcumin can reverse the phenotype of PCOS model rats such as irregular estrus cycle, the increase in body and ovarian weight, the elevation of serum T and LH, and the decrease in FSH [159]. Curcumin also attenuates ER stress in primary culture GCs by activating the PI3K/AKT pathway and protects against the apoptosis of primary culture GCs, perhaps by activating the PI3K/AKT pathway and inhibiting IRE1α-XBP1 levels [159]. Plumbagin can deactivate the PI3K/AKT/mTOR pathway in the primary GCs of the PCOS rat model, which results in the apoptosis and inhibition of proliferation of the primary GCs.
In summary, compounds from herb plants have multiple regulatory mechanisms in GCs such as the MAKP signaling pathway, the PI3K/AKT pathway, HMGB1, TNF-α, TLR4, and NF-κB/p65, resulting in a more regular estrous cycle and normal serum testosterone, LH, and FSH levels in animal models.

6. Conclusions

Experimental models have been extensively used to gain a better understanding of PCOS. However, merely treating the symptoms is not enough; suitable models must be utilized to develop a cure for PCOS. While clinical observations have supported the role of androgen actions in PCOS, there is currently a lack of conclusive evidence. Recent experimental studies have provided substantial evidence supporting the significance of direct AR-mediated androgen actions in the development and progression of PCOS. However, because of the ethical limitations surrounding human experimentation, animal models using rodents, sheep, and non-human primates have had to be developed. Nevertheless, in vitro models of PCOS offer a unique opportunity for mechanistic experiments to be conducted and be reflected pathological features of PCOS (see Table 7), and provide insights into the pathogenesis of PCOS (Figure 2). In summary, by combining clinical observations with targeted in vitro models, we can have a better understanding of the mechanistic pathogenesis of PCOS. This approach offers the possibility of identifying target sites and key pathways involved in PCOS, which may, in turn, lead to the development of novel, evidence-based therapeutic treatment for PCOS.

Author Contributions

All authors contributed to the concepts of the manuscript. Y.-R.T. wrote the first draft of manuscript and designed Tables; Y.-N.L. wrote the manuscript, H.-Y.K. reviewed the manuscript and supervised the work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Chang Gung Memorial Hospital-Kaohsiung Medical Center, Kaohsiung, Taiwan (CMRPG6H0281, CMRPD8H0282-3, CMRPD8K0011-3, CMRPD8L0011-3) and the Ministry of Science and Technology of the Republic of China, Taiwan (Contract No. MOST107-2320-B-182-035-MY3 and MOST110-2320-B-182-004-MY3) for financially supporting this research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Chen Ju Tsai for technical supporting this research work.

Conflicts of Interest

The authors declare no conflict interests.

References

  1. Ch, M.; Marshall, J.C. Polycystic Ovary Syndrome. N. Engl. J. Med. 2016, 375, 1398–1399. [Google Scholar]
  2. Chen, Z.J.; Zhao, H.; He, L.; Shi, Y.; Qin, Y.; Shi, Y.; Li, Z.; You, L.; Zhao, J.; Liu, J.; et al. Genome-wide association study identifies susceptibility loci for polycystic ovary syndrome on chromosome 2p16.3, 2p21 and 9q33.3. Nat. Genet. 2011, 43, 55–59. [Google Scholar] [CrossRef]
  3. Teede, H.; Deeks, A.; Moran, L. Polycystic ovary syndrome: A complex condition with psychological, reproductive and metabolic manifestations that impacts on health across the lifespan. BMC Med. 2010, 8, 41. [Google Scholar] [CrossRef] [PubMed]
  4. Divyashree, S.; Janhavi, P.; Ravindra, P.V.; Muthukumar, S.P. Experimental models of polycystic ovary syndrome: An update. Life Sci. 2019, 237, 116911. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Z.; Dong, H.; Yang, L.; Yi, P.; Wang, Q.; Huang, D. The role of FDX1 in granulosa cell of Polycystic ovary syndrome (PCOS). BMC Endocr. Disord. 2021, 21, 119. [Google Scholar] [CrossRef] [PubMed]
  6. Cataldo, N.A.; Dumesic, D.A.; Goldsmith, P.C.; Jaffe, R.B. Immunolocalization of Fas and Fas ligand in the ovaries of women with polycystic ovary syndrome: Relationship to apoptosis. Hum. Reprod. 2000, 15, 1889–1897. [Google Scholar] [CrossRef]
  7. Franks, S.; Stark, J.; Hardy, K. Follicle dynamics and anovulation in polycystic ovary syndrome. Hum. Reprod. Update 2008, 14, 367–378. [Google Scholar] [CrossRef]
  8. Cooney, L.G.; Dokras, A. Beyond fertility: Polycystic ovary syndrome and long-term health. Fertil. Steril. 2018, 110, 794–809. [Google Scholar] [CrossRef]
  9. Das, M.; Djahanbakhch, O.; Hacihanefioglu, B.; Saridogan, E.; Ikram, M.; Ghali, L.; Raveendran, M.; Storey, A. Granulosa cell survival and proliferation are altered in polycystic ovary syndrome. J. Clin. Endocr. Metab. 2008, 93, 881–887. [Google Scholar] [CrossRef]
  10. Nishi, Y.; Yanase, T.; Mu, Y.; Oba, K.; Ichino, I.; Saito, M.; Nomura, M.; Mukasa, C.; Okabe, T.; Goto, K.; et al. Establishment and characterization of a steroidogenic human granulosa-like tumor cell line, KGN, that expresses functional follicle-stimulating hormone receptor. Endocrinology 2001, 142, 437–445. [Google Scholar] [CrossRef]
  11. Jamieson, S.; Butzow, R.; Andersson, N.; Alexiadis, M.; Unkila-Kallio, L.; Heikinheimo, M.; Fuller, P.J.; Anttonen, M. The FOXL2 C134W mutation is characteristic of adult granulosa cell tumors of the ovary. Mod. Pathol. 2010, 23, 1477–1485. [Google Scholar] [CrossRef]
  12. Astapova, O.; Seger, C.; Hammes, S.R. Ligand Binding Prolongs Androgen Receptor Protein Half-Life by Reducing its Degradation. J. Endocr. Soc. 2021, 5, bvab035. [Google Scholar] [CrossRef]
  13. Zhou, R.; Li, S.; Liu, J.; Wu, H.; Yao, G.; Sun, Y.; Chen, Z.J.; Li, W.; Du, Y. Up-regulated FHL2 inhibits ovulation through interacting with androgen receptor and ERK1/2 in polycystic ovary syndrome. EBioMedicine 2020, 52, 102635. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, H.; Vollmer, M.; De Geyter, M.; Litzistorf, Y.; Ladewig, A.; Durrenberger, M.; Guggenheim, R.; Miny, P.; Holzgreve, W.; De Geyter, C. Characterization of an immortalized human granulosa cell line (COV434). Mol. Hum. Reprod. 2000, 6, 146–153. [Google Scholar] [CrossRef]
  15. Ishiwata, I.; Ishiwata, C.; Soma, M.; Kobayashi, N.; Ishikawa, H. Establishment and characterization of an estrogen-producing human ovarian granulosa tumor cell line. J. Natl. Cancer Inst. 1984, 72, 789–800. [Google Scholar] [PubMed]
  16. Lie, B.L.; Leung, E.; Leung, P.C.; Auersperg, N. Long-term growth and steroidogenic potential of human granulosa-lutein cells immortalized with SV40 large T antigen. Mol. Cell Endocrinol. 1996, 120, 169–176. [Google Scholar] [CrossRef] [PubMed]
  17. Lan, K.C.; Chang, S.Y.; Huang, F.J.; Lin, H.J.; Lin, C.Y.; Huang, K.E.; Kang, H.Y. Analysis of androgen receptor and anti-Müllerian hormone pathways in human granulosa cells under luteinizing hormone treatment. Reprod. Biol. Endocrinol. 2013, 11, 11. [Google Scholar] [CrossRef]
  18. Hosokawa, K.; Dantes, A.; Schere-Levy, C.; Barash, A.; Yoshida, Y.; Kotsuji, F.; Vlodavsky, I.; Amsterdam, A. Induction of Ad4BP/SF-1, steroidogenic acute regulatory protein, and cytochrome P450scc enzyme system expression in newly established human granulosa cell lines. Endocrinology 1998, 139, 4679–4687. [Google Scholar] [CrossRef]
  19. Havelock, J.C.; Rainey, W.E.; Carr, B.R. Ovarian granulosa cell lines. Mol. Cell. Endocrinol. 2004, 228, 67–78. [Google Scholar] [CrossRef]
  20. Bayasula Iwase, A.; Kiyono, T.; Takikawa, S.; Goto, M.; Nakamura, T.; Nagatomo, Y.; Nakahara, T.; Kotani, T.; Kobayashi, H.; Kondo, M. Establishment of a human nonluteinized granulosa cell line that transitions from the gonadotropin-independent to the gonadotropin-dependent status. Endocrinology 2012, 153, 2851–2860. [Google Scholar] [CrossRef]
  21. Belli, M.; Secchi, C.; Stupack, D.; Shimasaki, S. FOXO1 Negates the Cooperative Action of FOXL2(C134W) and SMAD3 in CYP19 Expression in HGrC1 Cells by Sequestering SMAD3. J. Endocr. Soc. 2019, 3, 2064–2081. [Google Scholar] [CrossRef] [PubMed]
  22. Rainey, W.H.; Sawetawan, C.; Shay, J.W.; Michael, M.D.; Mathis, J.M.; Kutteh, W.; Byrd, W.; Carr, B.R. Transformation of human granulosa cells with the E6 and E7 regions of human papillomavirus. J. Clin. Endocrinol. Metab. 1994, 78, 705–710. [Google Scholar] [CrossRef] [PubMed]
  23. Salih, S.M.; Jamaluddin, M.; Salama, S.A.; Fadl, A.A.; Nagamani, M.; Al-Hendy, A. Regulation of catechol O-methyltransferase expression in granulosa cells: A potential role for follicular arrest in polycystic ovary syndrome. Fertil. Steril. 2008, 89, 1414–1421. [Google Scholar] [CrossRef] [PubMed]
  24. Hashemian, Z.; Afsharian, P.; Farzaneh, P.; Eftekhari-Yazdi, P.; Vakhshiteh, F.; Daneshvar Amoli, A.; Nasimian, A. Establishment and characterization of a PCOS and a normal human granulosa cell line. Cytotechnology 2020, 72, 833–845. [Google Scholar] [CrossRef]
  25. Hashemain, Z.; Amiri-Yekta, A.; Khosravifar, M.; Alvandian, F.; Shahhosseini, M.; Hosseinkhani, S.; Afsharian, P. CYP19A1 Promoters Activity in Human Granulosa Cells: A Comparison between PCOS and Normal Subjects. Cell J. 2022, 24, 170–175. [Google Scholar] [CrossRef]
  26. Erickson, G.F.; Hsueh, A.J. Stimulation of aromatase activity by follicle stimulating hormone in rat granulosa cells in vivo and in vitro. Endocrinology 1978, 102, 1275–1282. [Google Scholar] [CrossRef]
  27. Eppig, J.J. A comparison between oocyte growth in coculture with granulosa cells and oocytes with granulosa cell-oocyte junctional contact maintained in vitro. J. Exp. Zool. 1979, 209, 345–353. [Google Scholar] [CrossRef]
  28. Schultz, R.M. Roles of cell-to-cell communication in development. Biol. Reprod. 1985, 32, 27–42. [Google Scholar] [CrossRef]
  29. Kaipia, A.; Hsueh, A.J. Regulation of ovarian follicle atresia. Ann. Rev. Physiol. 1997, 59, 349–363. [Google Scholar] [CrossRef]
  30. Stocco, C.; Telleria, C.; Gibori, G. The molecular control of corpus luteum formation, function, and regression. Endocr. Rev. 2007, 28, 117–149. [Google Scholar] [CrossRef]
  31. Bildik, G.; Esmaeilian, Y.; Vatansever, D.; Bilir, E.; Taskiran, C.; Oktem, O. A comparative molecular analysis of DNA damage response, cell cycle progression, viability and apoptosis of malignant granulosa cells exposed to gemcitabine and cisplatin. Mol. Biol. Rep. 2020, 47, 3789–3796. [Google Scholar] [CrossRef]
  32. Sudhakaran, G.; Babu, S.R.; Mahendra, H.; Arockiaraj, J. Updated experimental cellular models to study polycystic ovarian syndrome. Life Sci. 2023, 322, 121672. [Google Scholar] [CrossRef]
  33. Nicol, B.; Grimm, S.A.; Chalmel, F.; Lecluze, E.; Pannetier, M.; Pailhoux, E.; Dupin-De-Beyssat, E.; Guiguen, Y.; Capel, B.; Yao, H.H. RUNX1 maintains the identity of the fetal ovary through an interplay with FOXL2. Nat. Commun. 2019, 10, 5116. [Google Scholar] [CrossRef]
  34. Chilvers, R.A.; Bodenburg, Y.H.; Denner, L.A.; Urban, R.J. Development of a novel protocol for isolation and purification of human granulosa cells. J. Assist. Reprod. Gen. 2012, 29, 547–556. [Google Scholar] [CrossRef]
  35. Aghadavod, E.; Zarghami, N.; Farzadi, L.; Zare, M.; Barzegari, A.; Movassaghpour, A.A.; Nouri, M. Isolation of granulosa cells from follicular fluid; applications in biomedical and molecular biology experiments. Adv. Biomed. Res. 2015, 4, 250. [Google Scholar] [CrossRef]
  36. Wu, W.; Duan, C.; Lv, H.; Song, J.; Cai, W.; Fu, K.; Xu, J. MiR-let-7d-3p inhibits granulosa cell proliferation by targeting TLR4 in polycystic ovary syndrome. Reprod. Toxicol. 2021, 106, 61–68. [Google Scholar] [CrossRef]
  37. Li, Y.; Xiang, Y.; Song, Y.; Wan, L.; Yu, G.; Tan, L. Dysregulated miR-142, -33b and -423 in granulosa cells target TGFBR1 and SMAD7: A possible role in polycystic ovary syndrome. Mol. Hum. Reprod. 2019, 25, 638–646. [Google Scholar] [CrossRef]
  38. Wen, M.; Dou, X.; Zhang, S.; Wang, B.; Xu, J.; Zhang, W.; Wang, F. CTBP1AS upregulation is associated with polycystic ovary syndrome and can be effectively downregulated by cryptotanshinone. Mol. Med. Rep. 2022, 26, 245. [Google Scholar] [CrossRef]
  39. Song, Y.; Yu, G.; Xiang, Y.; Li, Y.; Wan, L.; Tan, L. Altered miR-186 and miR-135a contribute to granulosa cell dysfunction by targeting ESR2: A possible role in polycystic ovary syndrome. Mol. Cell. Endocrinol. 2019, 494, 110478. [Google Scholar] [CrossRef]
  40. Zhang, D.; Tang, H.Y.; Tan, L.; Zhao, D.M. MALAT1 is involved in the pathophysiological process of PCOS by modulating TGFbeta signaling in granulosa cells. Mol. Cell. Endocrinol. 2020, 499, 110589. [Google Scholar] [CrossRef]
  41. Li, Y.; Xiang, Y.; Song, Y.; Zhang, D.; Tan, L. MALAT1 downregulation is associated with polycystic ovary syndrome via binding with MDM2 and repressing P53 degradation. Mol. Cell. Endocrinol. 2022, 543, 111528. [Google Scholar] [CrossRef]
  42. Zhu, Q.; Yao, Y.; Xu, L.; Wu, H.; Wang, W.; He, Y.; Wang, Y.; Lu, Y.; Qi, J.; Ding, Y.; et al. Elevated SAA1 promotes the development of insulin resistance in ovarian granulosa cells in polycystic ovary syndrome. Reprod. Biol. Endocrinol. 2022, 20, 4. [Google Scholar] [CrossRef]
  43. Guo, R.; Zheng, H.; Li, Q.; Qiu, X.; Zhang, J.; Cheng, Z. Melatonin alleviates insulin resistance through the PI3K/AKT signaling pathway in ovary granulosa cells of polycystic ovary syndrome. Reprod. Biol. 2022, 22, 100594. [Google Scholar] [CrossRef]
  44. Zhang, Q.; Ren, J.; Wang, F.; Pan, M.; Cui, L.; Li, M.; Qu, F. Mitochondrial and glucose metabolic dysfunctions in granulosa cells induce impaired oocytes of polycystic ovary syndrome through Sirtuin 3. Free Radic. Biol. Med. 2022, 187, 1–16. [Google Scholar] [CrossRef]
  45. Yu, K.; Wang, R.X.; Li, M.H.; Sun, T.C.; Zhou, Y.W.; Li, Y.Y.; Sun, L.H.; Zhang, B.L.; Lian, Z.X.; Xue, S.G.; et al. Melatonin Reduces Androgen Production and Upregulates Heme Oxygenase-1 Expression in Granulosa Cells from PCOS Patients with Hypoestrogenia and Hyperandrogenia. Oxidative Med. Cell. Longev. 2019, 2019, 8218650. [Google Scholar] [CrossRef]
  46. Wang, Y.; Li, N.; Zeng, Z.; Tang, L.; Zhao, S.; Zhou, F.; Zhou, L.; Xia, W.; Zhu, C.; Rao, M. Humanin regulates oxidative stress in the ovaries of polycystic ovary syndrome patients via the Keap1/Nrf2 pathway. Mol. Hum. Reprod. 2021, 27, gaaa081. [Google Scholar] [CrossRef]
  47. Xie, F.; Zhang, J.; Zhai, M.; Liu, Y.; Hu, H.; Yu, Z.; Zhang, J.; Lin, S.; Liang, D.; Cao, Y. Melatonin ameliorates ovarian dysfunction by regulating autophagy in PCOS via the PI3K-Akt pathway. Reproduction 2021, 162, 73–82. [Google Scholar] [CrossRef]
  48. Yi, S.; Zheng, B.; Zhu, Y.; Cai, Y.; Sun, H.; Zhou, J. Melatonin ameliorates excessive PINK1/Parkin-mediated mitophagy by enhancing SIRT1 expression in granulosa cells of PCOS. Am. J. Physiol. Endocrinol. Metab. 2020, 319, E91–E101. [Google Scholar] [CrossRef]
  49. Li, Y.; Xiong, G.; Tan, J.; Wang, S.; Wu, Q.; Wan, L.; Zhang, Z.; Huang, O. Aberrant activation of the Hedgehog signaling pathway in granulosa cells from patients with polycystic ovary syndrome. Bioengineered 2021, 12, 12123–12134. [Google Scholar] [CrossRef]
  50. Azziz, R.; Carmina, E.; Dewailly, D.; Diamanti-Kandarakis, E.; Escobar-Morreale, H.F.; Futterweit, W.; Janssen, O.E.; Legro, R.S.; Norman, R.J.; Taylor, A.E.; et al. The Androgen Excess and PCOS Society criteria for the polycystic ovary syndrome: The complete task force report. Fertil. Steril. 2009, 91, 456–488. [Google Scholar] [CrossRef]
  51. Nelson, V.L.; Qin, K.N.; Rosenfield, R.L.; Wood, J.R.; Penning, T.M.; Legro, R.S.; Strauss, J.F., III; McAllister, J.M. The biochemical basis for increased testosterone production in theca cells propagated from patients with polycystic ovary syndrome. J. Clin. Endocrinol. Metab. 2001, 86, 5925–5933. [Google Scholar] [CrossRef]
  52. Nelson, V.L.; Legro, R.S.; Strauss, J.F., III; McAllister, J.M. Augmented androgen production is a stable steroidogenic phenotype of propagated theca cells from polycystic ovaries. Mol. Endocrinol. 1999, 13, 946–957. [Google Scholar] [CrossRef]
  53. Hirsch, A.; Hahn, D.; Kempna, P.; Hofer, G.; Nuoffer, J.M.; Mullis, P.E.; Fluck, C.E. Metformin inhibits human androgen production by regulating steroidogenic enzymes HSD3B2 and CYP17A1 and complex I activity of the respiratory chain. Endocrinology 2012, 153, 4354–4366. [Google Scholar] [CrossRef] [PubMed]
  54. Lee, B.H.; Indran, I.R.; Tan, H.M.; Li, Y.; Zhang, Z.; Li, J.; Yong, E.L. A Dietary Medium-Chain Fatty Acid, Decanoic Acid, Inhibits Recruitment of Nur77 to the HSD3B2 Promoter In Vitro and Reverses Endocrine and Metabolic Abnormalities in a Rat Model of Polycystic Ovary Syndrome. Endocrinology 2016, 157, 382–394. [Google Scholar] [CrossRef]
  55. Pyun, B.J.; Yang, H.; Sohn, E.; Yu, S.Y.; Lee, D.; Jung, D.H.; Ko, B.S.; Lee, H.W. Tetragonia tetragonioides (Pall.) Kuntze Regulates Androgen Production in a Letrozole-Induced Polycystic Ovary Syndrome Model. Molecules 2018, 23, 1173. [Google Scholar] [CrossRef]
  56. Yu, J.; Liu, Y.; Zhang, D.; Zhai, D.; Song, L.; Cai, Z.; Yu, C. Baicalin inhibits recruitment of GATA1 to the HSD3B2 promoter and reverses hyperandrogenism of PCOS. J. Endocrinol. 2019, 240, 497–507. [Google Scholar] [CrossRef]
  57. Chugh, R.M.; Park, H.S.; Esfandyari, S.; Elsharoud, A.; Ulin, M.; Al-Hendy, A. Mesenchymal Stem Cell-Conditioned Media Regulate Steroidogenesis and Inhibit Androgen Secretion in a PCOS Cell Model via BMP-2. Int. J. Mol. Sci. 2021, 22, 9184. [Google Scholar] [CrossRef]
  58. Marti, N.; Bouchoucha, N.; Sauter, K.S.; Fluck, C.E. Resveratrol inhibits androgen production of human adrenocortical H295R cells by lowering CYP17 and CYP21 expression and activities. PLoS ONE 2017, 12, e0174224. [Google Scholar] [CrossRef]
  59. Zhu, W.; Han, B.; Fan, M.; Wang, N.; Wang, H.; Zhu, H.; Cheng, T.; Zhao, S.; Song, H.; Qiao, J. Oxidative stress increases the 17,20-lyase-catalyzing activity of adrenal P450c17 through p38alpha in the development of hyperandrogenism. Mol. Cell Endocrinol. 2019, 484, 25–33. [Google Scholar] [CrossRef]
  60. Xu, B.; Dai, W.; Liu, L.; Han, H.; Zhang, J.; Du, X.; Pei, X.; Fu, X. Metformin ameliorates polycystic ovary syndrome in a rat model by decreasing excessive autophagy in ovarian granulosa cells via the PI3K/AKT/mTOR pathway. Endocr. J. 2022, 69, 863–875. [Google Scholar] [CrossRef]
  61. Chen, M.J.; Chou, C.H.; Chen, S.U.; Yang, W.S.; Yang, Y.S.; Ho, H.N. The effect of androgens on ovarian follicle maturation: Dihydrotestosterone suppress FSH-stimulated granulosa cell proliferation by upregulating PPARgamma-dependent PTEN expression. Sci. Rep. 2015, 5, 18319. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, Y.; Zeng, Z.; Zhao, S.; Tang, L.; Yan, J.; Li, N.; Zou, L.; Fan, X.; Xu, C.; Huang, J.; et al. Humanin Alleviates Insulin Resistance in Polycystic Ovary Syndrome: A Human and Rat Model-Based Study. Endocrinology 2021, 162, bqab056. [Google Scholar] [CrossRef] [PubMed]
  63. Song, L.; Yu, J.; Zhang, D.; Li, X.; Chen, L.; Cai, Z.; Yu, C. Androgen Excess Induced Mitochondrial Abnormality in Ovarian Granulosa Cells in a Rat Model of Polycystic Ovary Syndrome. Front. Endocrinol. 2022, 13, 789008. [Google Scholar] [CrossRef] [PubMed]
  64. Xing, Y.; Liu, Y.X.; Liu, X.; Wang, S.L.; Li, P.; Lin, X.H.; Sui, C.L.; Xu, C.; Qi, B.; Tong, Q. Effects of Gui Zhu Yi Kun formula on the P53/AMPK pathway of autophagy in granulosa cells of rats with polycystic ovary syndrome. Exp. Ther. Med. 2017, 13, 3567–3573. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, C.; Ding, C.; Hua, Z.; Chen, C.; Yu, J. Cangfudaotan Decoction Alleviates Insulin Resistance and Improves Follicular Development in Rats with Polycystic Ovary Syndrome via IGF-1-PI3K/Akt-Bax/Bcl-2 Pathway. Mediat. Inflamm. 2020, 2020, 8865647. [Google Scholar] [CrossRef]
  66. Li, T.; Dong, G.; Kang, Y.; Zhang, M.; Sheng, X.; Wang, Z.; Liu, Y.; Kong, N.; Sun, H. Increased homocysteine regulated by androgen activates autophagy by suppressing the mammalian target of rapamycin pathway in the granulosa cells of polycystic ovary syndrome mice. Bioengineered 2022, 13, 10875–10888. [Google Scholar] [CrossRef]
  67. Liu, Y.; Li, Z.; Wang, Y.; Cai, Q.; Liu, H.; Xu, C.; Zhang, F. IL-15 Participates in the Pathogenesis of Polycystic Ovary Syndrome by Affecting the Activity of Granulosa Cells. Front. Endocrinol. 2022, 13, 787876. [Google Scholar] [CrossRef]
  68. Safaei, Z.; Bakhshalizadeh, S.H.; Nasr Esfahani, M.H.; Akbari Sene, A.; Najafzadeh, V.; Soleimani, M.; Shirazi, R. Effect of Vitamin D3 on Mitochondrial Biogenesis in Granulosa Cells Derived from Polycystic Ovary Syndrome. Int. J. Fertil. Steril. 2020, 14, 143–149. [Google Scholar] [CrossRef]
  69. Gao, H.; Lin, L.; Haq, I.U.; Zeng, S.M. Inhibition of NF-kappaB promotes autophagy via JNK signaling pathway in porcine granulosa cells. Biochem. Biophys. Res. Commun. 2016, 473, 311–316. [Google Scholar] [CrossRef]
  70. Han, Y.; Wang, S.; Wang, Y.; Zeng, S. IGF-1 Inhibits Apoptosis of Porcine Primary Granulosa Cell by Targeting Degradation of BimEL. Int. J. Mol. Sci. 2019, 20, 5356. [Google Scholar] [CrossRef]
  71. Zhang, H.Y.; Zhu, F.F.; Zhu, Y.J.; Hu, Y.J.; Chen, X. Effects of IL-18 on the proliferation and steroidogenesis of bovine theca cells: Possible roles in the pathogenesis of polycystic ovary syndrome. J. Cell. Mol. Med. 2021, 25, 1128–1139. [Google Scholar] [CrossRef] [PubMed]
  72. Sun, L.F.; Yang, Y.L.; Wang, M.Y.; Zhao, H.S.; Xiao, T.X.; Li, M.X.; Wang, B.B.; Huang, C.; Ren, P.G.; Zhang, J.V. Inhibition of Col6a5 Improve Lipid Metabolism Disorder in Dihydrotestosterone-Induced Hyperandrogenic Mice. Front. Cell Dev. Biol. 2021, 9, 669189. [Google Scholar] [CrossRef] [PubMed]
  73. Converse, A.; Zhang, C.; Thomas, P. Membrane Androgen Receptor ZIP9 Induces Croaker Ovarian Cell Apoptosis via Stimulatory G Protein Alpha Subunit and MAP Kinase Signaling. Endocrinology 2017, 158, 3015–3029. [Google Scholar] [CrossRef]
  74. Su, C.; Chen, M.; Huang, H.; Lin, J. Testosterone enhances lipopolysaccharide-induced interleukin-6 and macrophage chemotactic protein-1 expression by activating the extracellular signal-regulated kinase 1/2/nuclear factor-kappaB signalling pathways in 3T3-L1 adipocytes. Mol. Med. Rep. 2015, 12, 696–704. [Google Scholar] [CrossRef] [PubMed]
  75. Luchetti, C.G.; Solano, M.E.; Sander, V.; Arcos, M.L.; Gonzalez, C.; Di Girolamo, G.; Chiocchio, S.; Cremaschi, G.; Motta, A.B. Effects of dehydroepiandrosterone on ovarian cystogenesis and immune function. J. Reprod. Immunol. 2004, 64, 59–74. [Google Scholar] [CrossRef] [PubMed]
  76. Manneras, L.; Cajander, S.; Holmang, A.; Seleskovic, Z.; Lystig, T.; Lohn, M.; Stener-Victorin, E. A new rat model exhibiting both ovarian and metabolic characteristics of polycystic ovary syndrome. Endocrinology 2007, 148, 3781–3791. [Google Scholar] [CrossRef]
  77. Johansson, J.; Feng, Y.; Shao, R.; Lonn, M.; Billig, H.; Stener-Victorin, E. Intense electroacupuncture normalizes insulin sensitivity, increases muscle GLUT4 content, and improves lipid profile in a rat model of polycystic ovary syndrome. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E551–E559. [Google Scholar] [CrossRef]
  78. Skarra, D.V.; Hernandez-Carretero, A.; Rivera, A.J.; Anvar, A.R.; Thackray, V.G. Hyperandrogenemia Induced by Letrozole Treatment of Pubertal Female Mice Results in Hyperinsulinemia Prior to Weight Gain and Insulin Resistance. Endocrinology 2017, 158, 2988–3003. [Google Scholar] [CrossRef]
  79. Lubahn, D.B.; Moyer, J.S.; Golding, T.S.; Couse, J.F.; Korach, K.S.; Smithies, O. Alteration of reproductive function but not prenatal sexual development after insertional disruption of the mouse estrogen receptor gene. Proc. Natl. Acad. Sci. USA 1993, 90, 11162–11166. [Google Scholar] [CrossRef]
  80. Lindzey, J.; Korach, K.S. Developmental and physiological effects of estrogen receptor gene disruption in mice. Trends Endocrinol. Metab. 1997, 8, 137–145. [Google Scholar] [CrossRef]
  81. Couse, J.F.; Bunch, D.O.; Lindzey, J.; Schomberg, D.W.; Korach, K.S. Prevention of the polycystic ovarian phenotype and characterization of ovulatory capacity in the estrogen receptor-alpha knockout mouse. Endocrinology 1999, 140, 5855–5865. [Google Scholar] [CrossRef] [PubMed]
  82. Gong, Y.; Luo, S.; Fan, P.; Zhu, H.; Li, Y.; Huang, W. Growth hormone activates PI3K/Akt signaling and inhibits ROS accumulation and apoptosis in granulosa cells of patients with polycystic ovary syndrome. Reprod. Biol. Endocrinol. 2020, 18, 121. [Google Scholar] [CrossRef]
  83. Young, J.M.; McNeilly, A.S. Theca: The forgotten cell of the ovarian follicle. Reproduction 2010, 140, 489–504. [Google Scholar] [CrossRef]
  84. Liao, B.; Qi, X.; Yun, C.; Qiao, J.; Pang, Y. Effects of Androgen Excess-Related Metabolic Disturbances on Granulosa Cell Function and Follicular Development. Front. Endocrinol. 2022, 13, 815968. [Google Scholar] [CrossRef] [PubMed]
  85. Edson, M.A.; Nagaraja, A.K.; Matzuk, M.M. The mammalian ovary from genesis to revelation. Endocr. Rev. 2009, 30, 624–712. [Google Scholar] [CrossRef]
  86. Webber, L.J.; Stubbs, S.; Stark, J.; Trew, G.H.; Margara, R.; Hardy, K.; Franks, S. Formation and early development of follicles in the polycystic ovary. Lancet 2003, 362, 1017–1021. [Google Scholar] [CrossRef]
  87. Maciel, G.A.; Baracat, E.C.; Benda, J.A.; Markham, S.M.; Hensinger, K.; Chang, R.J.; Erickson, G.F. Stockpiling of transitional and classic primary follicles in ovaries of women with polycystic ovary syndrome. J. Clin. Endocrinol. Metab. 2004, 89, 5321–5327. [Google Scholar] [CrossRef]
  88. Stubbs, S.A.; Stark, J.; Dilworth, S.M.; Franks, S.; Hardy, K. Abnormal preantral folliculogenesis in polycystic ovaries is associated with increased granulosa cell division. J. Clin. Endocrinol. Metab. 2007, 92, 4418–4426. [Google Scholar] [CrossRef]
  89. Jonard, S.; Dewailly, D. The follicular excess in polycystic ovaries, due to intra-ovarian hyperandrogenism, may be the main culprit for the follicular arrest. Hum. Reprod. Update 2004, 10, 107–117. [Google Scholar] [CrossRef]
  90. Bertoldo, M.J.; Caldwell, A.S.L.; Riepsamen, A.H.; Lin, D.; Gonzalez, M.B.; Robker, R.L.; Ledger, W.L.; Gilchrist, R.B.; Handelsman, D.J.; Walters, K.A. A Hyperandrogenic Environment Causes Intrinsic Defects That Are Detrimental to Follicular Dynamics in a PCOS Mouse Model. Endocrinology 2019, 160, 699–715. [Google Scholar] [CrossRef]
  91. Pignatelli, D. Non-classic adrenal hyperplasia due to the deficiency of 21-hydroxylase and its relation to polycystic ovarian syndrome. Front. Horm. Res. 2013, 40, 158–170. [Google Scholar] [CrossRef]
  92. Fujibe, Y.; Baba, T.; Nagao, S.; Adachi, S.; Ikeda, K.; Morishita, M.; Kuno, Y.; Suzuki, M.; Mizuuchi, M.; Honnma, H.; et al. Androgen potentiates the expression of FSH receptor and supports preantral follicle development in mice. J. Ovarian Res. 2019, 12, 31. [Google Scholar] [CrossRef]
  93. Nagao, S.; Baba, T.; Fujibe, Y.; Adachi, S.; Ikeda, K.; Morishita, M.; Kuno, Y.; Honnma, H.; Endo, T.; Kiya, T.; et al. Pioglitazone suppresses excessive follicular development in murine preantral follicles. J. Ovarian Res. 2019, 12, 82. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, T.; Zhao, H.; Wang, J.; Shu, X.; Gao, Y.; Mu, X.; Gao, F.; Liu, H. The role of fructose-1,6-bisphosphatase 1 in abnormal development of ovarian follicles caused by high testosterone concentration. Mol. Med. Rep. 2017, 16, 6489–6498. [Google Scholar] [CrossRef] [PubMed]
  95. Hwang, D.J.; Song, H.K.; Kim, K.S.; Jung, Y.S.; Hwang, D.Y.; Cho, J.Y. Comparative analysis of basal locomotor activity-related metabolic phenotypes between C57BL/6 mice and ICR mice substrains derived from three different sources. Lab. Anim. Res. 2017, 33, 140–149. [Google Scholar] [CrossRef] [PubMed]
  96. Rachoń, D. Differential diagnosis of hyperandrogenism in women with polycystic ovary syndrome. Exp. Clin. Endocrinol. Diabetes 2012, 120, 205–209. [Google Scholar] [CrossRef]
  97. Sander, V.A.; Hapon, M.B.; Sícaro, L.; Lombardi, E.P.; Jahn, G.A.; Motta, A.B. Alterations of folliculogenesis in women with polycystic ovary syndrome. J. Steroid Biochem. Mol. Biol. 2011, 124, 58–64. [Google Scholar] [CrossRef]
  98. Patel, R.; Shah, G. High-fat diet exposure from pre-pubertal age induces polycystic ovary syndrome (PCOS) in rats. Reproduction 2018, 155, 141–151. [Google Scholar] [CrossRef]
  99. Murray, A.A.; Gosden, R.G.; Allison, V.; Spears, N. Effect of androgens on the development of mouse follicles growing in vitro. J. Reprod. Fertil. 1998, 113, 27–33. [Google Scholar] [CrossRef]
  100. Wang, H.; Andoh, K.; Hagiwara, H.; Xiaowei, L.; Kikuchi, N.; Abe, Y.; Yamada, K.; Fatima, R.; Mizunuma, H. Effect of adrenal and ovarian androgens on type 4 follicles unresponsive to FSH in immature mice. Endocrinology 2001, 142, 4930–4936. [Google Scholar] [CrossRef]
  101. Laird, M.; Thomson, K.; Fenwick, M.; Mora, J.; Franks, S.; Hardy, K. Androgen Stimulates Growth of Mouse Preantral Follicles In Vitro: Interaction With Follicle-Stimulating Hormone and With Growth Factors of the TGF beta Superfamily. Endocrinology 2017, 158, 920–935. [Google Scholar] [CrossRef] [PubMed]
  102. Sadeghi, H.M.; Adeli, I.; Calina, D.; Docea, A.O.; Mousavi, T.; Daniali, M.; Nikfar, S.; Tsatsakis, A.; Abdollahi, M. Polycystic Ovary Syndrome: A Comprehensive Review of Pathogenesis, Management, and Drug Repurposing. Int. J. Mol. Sci. 2022, 23, 583. [Google Scholar] [CrossRef]
  103. Li, T.T.; Mo, H.; Chen, W.F.; Li, L.; Xiao, Y.; Zhang, J.; Li, X.F.; Lu, Y. Role of the PI3K-Akt Signaling Pathway in the Pathogenesis of Polycystic Ovary Syndrome. Reprod. Sci. 2017, 24, 646–655. [Google Scholar] [CrossRef] [PubMed]
  104. Zhao, J.; Xu, J.; Wang, W.; Zhao, H.; Liu, H.; Liu, X.; Liu, J.; Sun, Y.; Dunaif, A.; Du, Y.; et al. Long non-coding RNA LINC-01572:28 inhibits granulosa cell growth via a decrease in p27 (Kip1) degradation in patients with polycystic ovary syndrome. EBioMedicine 2018, 36, 526–538. [Google Scholar] [CrossRef] [PubMed]
  105. Geng, X.; Zhao, J.; Huang, J.; Li, S.; Chu, W.; Wang, W.S.; Chen, Z.J.; Du, Y. lnc-MAP3K13-7:1 Inhibits Ovarian GC Proliferation in PCOS via DNMT1 Downregulation-Mediated CDKN1A Promoter Hypomethylation. Mol. Ther. 2021, 29, 1279–1293. [Google Scholar] [CrossRef]
  106. Li, X.; Liu, H.; Wang, H.; Sun, L.; Ding, F.; Sun, W.; Han, C.; Wang, J. Follistatin could promote the proliferation of duck primary myoblasts by activating PI3K/Akt/mTOR signalling. Biosci. Rep. 2014, 34, e00143. [Google Scholar] [CrossRef]
  107. Song, W.J.; Shi, X.; Zhang, J.; Chen, L.; Fu, S.X.; Ding, Y.L. Akt-mTOR Signaling Mediates Abnormalities in the Proliferation and Apoptosis of Ovarian Granulosa Cells in Patients with Polycystic Ovary Syndrome. Gynecol. Obstet. Investig. 2018, 83, 124–132. [Google Scholar] [CrossRef]
  108. Cai, Z.; He, S.; Li, T.; Zhao, L.; Zhang, K. Plumbagin inhibits proliferation and promotes apoptosis of ovarian granulosa cells in polycystic ovary syndrome by inactivating PI3K/Akt/mTOR pathway. Anim. Cells Syst. 2020, 24, 197–204. [Google Scholar] [CrossRef]
  109. Xiong, H.; Hu, Q.; Jiang, Q. Protective effects of lidocaine on polycystic ovary syndrome through modulating ovarian granulosa cell physiology via PI3K/AKT/mTOR pathway. Cytotechnology 2022, 74, 283–292. [Google Scholar] [CrossRef]
  110. Zheng, B.; Meng, J.; Zhu, Y.; Ding, M.; Zhang, Y.; Zhou, J. Melatonin enhances SIRT1 to ameliorate mitochondrial membrane damage by activating PDK1/Akt in granulosa cells of PCOS. J. Ovarian Res. 2021, 14, 152. [Google Scholar] [CrossRef]
  111. Itoh, S.; Itoh, F.; Goumans, M.J.; Ten Dijke, P. Signaling of transforming growth factor-beta family members through Smad proteins. Eur. J. Biochem. 2000, 267, 6954–6967. [Google Scholar] [CrossRef] [PubMed]
  112. Aydos, A.; Gurel, A.; Oztemur Islakoglu, Y.; Noyan, S.; Gokce, B.; Ecemis, T.; Kaya, C.; Aksu, A.T.; Gur Dedeoglu, B. Identification of Polycystic Ovary Syndrome (PCOS) Specific Genes in Cumulus and Mural Granulosa Cells. PLoS ONE 2016, 11, e0168875. [Google Scholar] [CrossRef] [PubMed]
  113. Russell, M.C.; Cowan, R.G.; Harman, R.M.; Walker, A.L.; Quirk, S.M. The hedgehog signaling pathway in the mouse ovary. Biol. Reprod. 2007, 77, 226–236. [Google Scholar] [CrossRef] [PubMed]
  114. Thomas, P.; Converse, A.; Berg, H.A. ZIP9, a novel membrane androgen receptor and zinc transporter protein. Gen. Comp. Endocrinol. 2018, 257, 130–136. [Google Scholar] [CrossRef] [PubMed]
  115. Li, Y.; Zheng, Q.; Sun, D.; Cui, X.; Chen, S.; Bulbul, A.; Liu, S.; Yan, Q. Dehydroepiandrosterone stimulates inflammation and impairs ovarian functions of polycystic ovary syndrome. J. Cell Physiol. 2019, 234, 7435–7447. [Google Scholar] [CrossRef]
  116. Luo, X.; Gong, Y.; Cai, L.; Zhang, L.; Dong, X. Chemerin regulates autophagy to participate in polycystic ovary syndrome. J. Int. Med. Res. 2021, 49, 3000605211058376. [Google Scholar] [CrossRef]
  117. Taheri, M.; Hayati Roudbari, N.; Amidi, F.; Parivar, K. The protective effect of sulforaphane against oxidative stress in granulosa cells of patients with polycystic ovary syndrome (PCOS) through activation of AMPK/AKT/NRF2 signaling pathway. Reprod. Biol. 2021, 21, 100563. [Google Scholar] [CrossRef]
  118. Qiao, G.Y.; Dong, B.W.; Zhu, C.J.; Yan, C.Y.; Chen, B.L. Deregulation of WNT2/FZD3/beta-catenin pathway compromises the estrogen synthesis in cumulus cells from patients with polycystic ovary syndrome. Biochem. Biophys. Res. Commun. 2017, 493, 847–854. [Google Scholar] [CrossRef]
  119. Hardy, K.; Fenwick, M.; Mora, J.; Laird, M.; Thomson, K.; Franks, S. Onset and Heterogeneity of Responsiveness to FSH in Mouse Preantral Follicles in Culture. Endocrinology 2017, 158, 134–147. [Google Scholar] [CrossRef]
  120. Canu, N.; Tufi, R.; Serafino, A.L.; Amadoro, G.; Ciotti, M.T.; Calissano, P. Role of the autophagic-lysosomal system on low potassium-induced apoptosis in cultured cerebellar granule cells. J. Neurochem. 2005, 92, 1228–1242. [Google Scholar] [CrossRef]
  121. Kumariya, S.; Ubba, V.; Jha, R.K.; Gayen, J.R. Autophagy in ovary and polycystic ovary syndrome: Role, dispute and future perspective. Autophagy 2021, 17, 2706–2733. [Google Scholar] [CrossRef] [PubMed]
  122. Zhang, X.D.; Qin, Z.H.; Wang, J. The role of p53 in cell metabolism. Acta Pharmacol. Sin. 2010, 31, 1208–1212. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, Y.; Chan, S.; Tsang, B.K. Involvement of inhibitory nuclear factor-kappaB (NFkappaB)-independent NFkappaB activation in the gonadotropic regulation of X-linked inhibitor of apoptosis expression during ovarian follicular development in vitro. Endocrinology 2002, 143, 2732–2740. [Google Scholar] [CrossRef] [PubMed]
  124. Choi, J.; Jo, M.; Lee, E.; Choi, D. AKT is involved in granulosa cell autophagy regulation via mTOR signaling during rat follicular development and atresia. Reproduction 2014, 147, 73–80. [Google Scholar] [CrossRef] [PubMed]
  125. Moghetti, P.; Tosi, F. Insulin resistance and PCOS: Chicken or egg? J. Endocrinol. Investig. 2021, 44, 233–244. [Google Scholar] [CrossRef]
  126. Gonzalez, F.; Sia, C.L.; Shepard, M.K.; Rote, N.S.; Minium, J. Hyperglycemia-induced oxidative stress is independent of excess abdominal adiposity in normal-weight women with polycystic ovary syndrome. Hum. Reprod. 2012, 27, 3560–3568. [Google Scholar] [CrossRef]
  127. Murri, M.; Luque-Ramirez, M.; Insenser, M.; Ojeda-Ojeda, M.; Escobar-Morreale, H.F. Circulating markers of oxidative stress and polycystic ovary syndrome (PCOS): A systematic review and meta-analysis. Hum. Reprod. Update 2013, 19, 268–288. [Google Scholar] [CrossRef]
  128. Itoh, K.; Wakabayashi, N.; Katoh, Y.; Ishii, T.; Igarashi, K.; Engel, J.D.; Yamamoto, M. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 1999, 13, 76–86. [Google Scholar] [CrossRef]
  129. Lee, S.H.; Chung, D.J.; Lee, H.S.; Kim, T.J.; Kim, M.H.; Jeong, H.J.; Im, J.A.; Lee, D.C.; Lee, J.W. Mitochondrial DNA copy number in peripheral blood in polycystic ovary syndrome. Metabolism 2011, 60, 1677–1682. [Google Scholar] [CrossRef]
  130. Azziz, R.; Carmina, E.; Chen, Z.; Dunaif, A.; Laven, J.S.; Legro, R.S.; Lizneva, D.; Natterson-Horowtiz, B.; Teede, H.J.; Yildiz, B.O. Polycystic ovary syndrome. Nat. Rev. Dis. Primers 2016, 2, 16057. [Google Scholar] [CrossRef]
  131. Dewailly, D.; Robin, G.; Peigne, M.; Decanter, C.; Pigny, P.; Catteau-Jonard, S. Interactions between androgens, FSH, anti-Müllerian hormone and estradiol during folliculogenesis in the human normal and polycystic ovary. Hum. Reprod. Update 2016, 22, 709–724. [Google Scholar] [CrossRef] [PubMed]
  132. Krol, J.; Loedige, I.; Filipowicz, W. The widespread regulation of microRNA biogenesis, function and decay. Nat. Rev. Genet. 2010, 11, 597–610. [Google Scholar] [CrossRef] [PubMed]
  133. Chen, B.; Xu, P.; Wang, J.; Zhang, C. The role of MiRNA in polycystic ovary syndrome (PCOS). Gene 2019, 706, 91–96. [Google Scholar] [CrossRef] [PubMed]
  134. Li, Y.; Fang, Y.; Liu, Y.; Yang, X. MicroRNAs in ovarian function and disorders. J. Ovarian Res. 2015, 8, 51. [Google Scholar] [CrossRef]
  135. Salehi, R.; Wyse, B.A.; Asare-Werehene, M.; Esfandiarinezhad, F.; Abedini, A.; Pan, B.; Urata, Y.; Gutsol, A.; Vinas, J.L.; Jahangiri, S.; et al. Androgen-induced exosomal miR-379-5p release determines granulosa cell fate: Cellular mechanism involved in polycystic ovaries. J. Ovarian Res. 2023, 16, 74. [Google Scholar] [CrossRef]
  136. Huang, X.; Liu, C.; Hao, C.F.; Tang, Q.Q.; Liu, R.M.; Lin, S.X.; Zhang, L.P.; Yan, W. Identification of altered microRNAs and mRNAs in the cumulus cells of PCOS patients: miRNA-509-3p promotes oestradiol secretion by targeting MAP3K8. Reproduction 2016, 151, 643–655. [Google Scholar] [CrossRef]
  137. Naji, M.; Nekoonam, S.; Aleyasin, A.; Arefian, E.; Mahdian, R.; Azizi, E.; Nashtaei, M.S.; Amidi, F. Expression of miR-15a, miR-145, and miR-182 in granulosa-lutein cells, follicular fluid, and serum of women with polycystic ovary syndrome (PCOS). Arch. Gynecol. Obstet. 2018, 297, 221–231. [Google Scholar] [CrossRef]
  138. Xue, Y.; Lv, J.; Xu, P.; Gu, L.; Cao, J.; Xu, L.; Xue, K.; Li, Q. Identification of microRNAs and genes associated with hyperandrogenism in the follicular fluid of women with polycystic ovary syndrome. J. Cell. Biochem. 2018, 119, 3913–3921. [Google Scholar] [CrossRef]
  139. Chen, Z.; Ou, H.; Wu, H.; Wu, P.; Mo, Z. Role of microRNA in the Pathogenesis of Polycystic Ovary Syndrome. DNA Cell Biol. 2019, 38, 754–762. [Google Scholar] [CrossRef]
  140. Liu, S.; Zhang, X.; Shi, C.; Lin, J.; Chen, G.; Wu, B.; Wu, L.; Shi, H.; Yuan, Y.; Zhou, W.; et al. Altered microRNAs expression profiling in cumulus cells from patients with polycystic ovary syndrome. J. Transl. Med. 2015, 13, 238. [Google Scholar] [CrossRef]
  141. Wu, Y.X.; Lin, Y.S.; Li, S.C.; Yao, X.; Cheng, M.; Zhu, L.; Liu, H.Y. microRNA-194 is increased in polycystic ovary syndrome granulosa cell and induce KGN cells apoptosis by direct targeting heparin-binding EGF-like growth factor. Reprod. Biol. Endocrinol. 2021, 19, 170. [Google Scholar] [CrossRef] [PubMed]
  142. Zhou, S.; Xia, L.; Chen, Y.; Guo, W.; Hu, J. miR-3188 Regulates proliferation and apoptosis of granulosa cells by targeting KCNA5 in the polycystic ovary syndrome. Acta Biochim. Pol. 2021, 68, 83–89. [Google Scholar] [CrossRef] [PubMed]
  143. Yu, Y.; Li, G.; He, X.; Lin, Y.; Chen, Z.; Lin, X.; Xu, H. MicroRNA-21 regulate the cell apoptosis and cell proliferation of polycystic ovary syndrome (PCOS) granulosa cells through target toll like receptor TLR8. Bioengineered 2021, 12, 5789–5796. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, T.; Wang, L.; Zhang, Y.; Zheng, J.; Liu, L. MicroRNA-451a plays a role in polycystic ovary syndrome by regulating ovarian granulosa cell proliferation and apoptosis. Exp. Ther. Med. 2021, 21, 583. [Google Scholar] [CrossRef] [PubMed]
  145. Zhou, J.; Jin, X.; Sheng, Z.; Zhang, Z. miR-206 serves an important role in polycystic ovary syndrome through modulating ovarian granulosa cell proliferation and apoptosis. Exp. Ther. Med. 2021, 21, 179. [Google Scholar] [CrossRef]
  146. Zhong, Z.; Li, F.; Li, Y.; Qin, S.; Wen, C.; Fu, Y.; Xiao, Q. Inhibition of microRNA-19b promotes ovarian granulosa cell proliferation by targeting IGF-1 in polycystic ovary syndrome. Mol. Med. Rep. 2018, 17, 4889–4898. [Google Scholar] [CrossRef]
  147. Wang, T.; Liu, Y.; Lv, M.; Xing, Q.; Zhang, Z.; He, X.; Xu, Y.; Wei, Z.; Cao, Y. miR-323-3p regulates the steroidogenesis and cell apoptosis in polycystic ovary syndrome (PCOS) by targeting IGF-1. Gene 2019, 683, 87–100. [Google Scholar] [CrossRef]
  148. Aldakheel, F.M.; Abuderman, A.A.; Alduraywish, S.A.; Xiao, Y.; Guo, W.W. MicroRNA-21 inhibits ovarian granulosa cell proliferation by targeting SNHG7 in premature ovarian failure with polycystic ovary syndrome. J. Reprod. Immunol. 2021, 146, 103328. [Google Scholar] [CrossRef]
  149. Li, Y.; Liu, Y.D.; Zhou, X.Y.; Chen, S.L.; Chen, X.; Zhe, J.; Zhang, J.; Zhang, Q.Y.; Chen, Y.X. MiR-29a regulates the proliferation, aromatase expression, and estradiol biosynthesis of human granulosa cells in polycystic ovary syndrome. Mol. Cell Endocrinol. 2019, 498, 110540. [Google Scholar] [CrossRef]
  150. Zhang, C.L.; Wang, H.; Yan, C.Y.; Gao, X.F.; Ling, X.J. Deregulation of RUNX2 by miR-320a deficiency impairs steroidogenesis in cumulus granulosa cells from polycystic ovary syndrome (PCOS) patients. Biochem. Biophys. Res. Commun. 2017, 482, 1469–1476. [Google Scholar] [CrossRef]
  151. Jiang, Y.C.; Ma, J.X. The role of MiR-324-3p in polycystic ovary syndrome (PCOS) via targeting WNT2B. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 3286–3293. [Google Scholar] [CrossRef] [PubMed]
  152. Vitale, S.G.; Fulghesu, A.M.; Mikus, M.; Watrowski, R.; D’Alterio, M.N.; Lin, L.T.; Shah, M.; Reyes-Munoz, E.; Sathyapalan, T.; Angioni, S. The Translational Role of miRNA in Polycystic Ovary Syndrome: From Bench to Bedside-A Systematic Literature Review. Biomedicines 2022, 10, 1816. [Google Scholar] [CrossRef] [PubMed]
  153. Abdalla, M.; Deshmukh, H.; Atkin, S.L.; Sathyapalan, T. miRNAs as a novel clinical biomarker and therapeutic targets in polycystic ovary syndrome (PCOS): A review. Life Sci. 2020, 259, 118174. [Google Scholar] [CrossRef] [PubMed]
  154. Sacks, D.; Baxter, B.; Campbell, B.C.V.; Carpenter, J.S.; Cognard, C.; Dippel, D.; Eesa, M.; Fischer, U.; Hausegger, K.; Hirsch, J.A.; et al. Multisociety Consensus Quality Improvement Revised Consensus Statement for Endovascular Therapy of Acute Ischemic Stroke. Int. J. Stroke 2018, 13, 612–632. [Google Scholar] [CrossRef]
  155. Moini Jazani, A.; Nasimi Doost Azgomi, H.; Nasimi Doost Azgomi, A.; Nasimi Doost Azgomi, R. A comprehensive review of clinical studies with herbal medicine on polycystic ovary syndrome (PCOS). Daru 2019, 27, 863–877. [Google Scholar] [CrossRef] [PubMed]
  156. Liu, M.; Zhu, H.; Zhu, Y.; Hu, X. Guizhi Fuling Wan reduces autophagy of granulosa cell in rats with polycystic ovary syndrome via restoring the PI3K/AKT/mTOR signaling pathway. J. Ethnopharmacol. 2021, 270, 113821. [Google Scholar] [CrossRef]
  157. Liu, H.; Xie, J.; Fan, L.; Xia, Y.; Peng, X.; Zhou, J.; Ni, X. Cryptotanshinone Protects against PCOS-Induced Damage of Ovarian Tissue via Regulating Oxidative Stress, Mitochondrial Membrane Potential, Inflammation, and Apoptosis via Regulating Ferroptosis. Oxidative Med. Cell. Longev. 2022, 2022, 8011850. [Google Scholar] [CrossRef]
  158. Yang, Y.; Yang, L.; Qi, C.; Hu, G.; Wang, L.; Sun, Z.; Ni, X. Cryptotanshinone alleviates polycystic ovary syndrome in rats by regulating the HMGB1/TLR4/NFkappaB signaling pathway. Mol. Med. Rep. 2020, 22, 3851–3861. [Google Scholar] [CrossRef]
  159. Zhang, Y.; Wang, L.; Weng, Y.; Wang, D.; Wang, R.; Wang, H.; Wang, L.; Shen, S.; Wang, H.; Li, Y.; et al. Curcumin Inhibits Hyperandrogen-Induced IRE1alpha-XBP1 Pathway Activation by Activating the PI3K/AKT Signaling in Ovarian Granulosa Cells of PCOS Model Rats. Oxidative Med. Cell. Longev. 2022, 2022, 2113293. [Google Scholar] [CrossRef]
  160. Wu, H.; Dai, X.; Wang, E. Plumbagin inhibits cell proliferation and promotes apoptosis in multiple myeloma cells through inhibition of the PI3K/Akt-mTOR pathway. Oncol. Lett. 2016, 12, 3614–3618. [Google Scholar] [CrossRef]
  161. Cai, Z.; He, S.; Liu, R.; Zhou, L.; Zhao, L. Plumbagin rescues the granulosa cell’s pyroptosis by reducing WTAP-mediated N6-methylation in polycystic ovary syndrome. J. Ovarian Res. 2022, 15, 126. [Google Scholar] [CrossRef] [PubMed]
  162. Sun, H.Y.; Li, Q.; Liu, Y.Y.; Wei, X.H.; Pan, C.S.; Fan, J.Y.; Han, J.Y. Xiao-Yao-San, a Chinese Medicine Formula, Ameliorates Chronic Unpredictable Mild Stress Induced Polycystic Ovary in Rat. Front. Physiol. 2017, 8, 729. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, Q.; Ren, J.; Wang, F.; Li, M.; Pan, M.; Zhang, H.; Qu, F. Chinese herbal medicine alleviates the pathogenesis of polycystic ovary syndrome by improving oxidative stress and glucose metabolism via mitochondrial Sirtuin 3 signaling. Phytomedicine 2023, 109, 154556. [Google Scholar] [CrossRef] [PubMed]
  164. Xu, Y.; Pan, C.S.; Li, Q.; Zhang, H.L.; Yan, L.; Anwaier, G.; Wang, X.Y.; Yan, L.L.; Fan, J.Y.; Li, D.; et al. The Ameliorating Effects of Bushen Huatan Granules and Kunling Wan on Polycystic Ovary Syndrome Induced by Dehydroepiandrosterone in Rats. Front. Physiol. 2021, 12, 525145. [Google Scholar] [CrossRef]
  165. Gao, H.; Gao, L.; Rao, Y.; Qian, L.; Li, M.; Wang, W. The Gengnianchun recipe attenuates insulin resistance-induced diminished ovarian reserve through inhibiting the senescence of granulosa cells. Front. Endocrinol. 2023, 14, 1133280. [Google Scholar] [CrossRef]
  166. Kim, C.H.; Chon, S.J.; Lee, S.H. Effects of lifestyle modification in polycystic ovary syndrome compared to metformin only or metformin addition: A systematic review and meta-analysis. Sci. Rep. 2020, 10, 7802. [Google Scholar] [CrossRef]
  167. Franik, S.; Eltrop, S.M.; Kremer, J.A.; Kiesel, L.; Farquhar, C. Aromatase inhibitors (letrozole) for subfertile women with polycystic ovary syndrome. Cochrane Database Syst. Rev. 2018, 5, CD010287. [Google Scholar] [CrossRef]
  168. Brown, J.; Farquhar, C. Clomiphene and other antioestrogens for ovulation induction in polycystic ovarian syndrome. Cochrane Database Syst. Rev. 2016, 12, CD002249. [Google Scholar] [CrossRef]
  169. Di Pietro, M.; Velazquez, C.; Matzkin, M.E.; Frungieri, M.B.; Peña, M.G.; de Zúñiga, I.; Pascuali, N.; Irusta, G.; Bianchi, M.S.; Parborell, F.; et al. Metformin has a direct effect on ovarian cells that is dependent on organic cation transporters. Mol. Cell. Endocrinol. 2020, 499, 110591. [Google Scholar] [CrossRef]
  170. Yu, K.; Deng, S.L.; Sun, T.C.; Li, Y.Y.; Liu, Y.X. Melatonin Regulates the Synthesis of Steroid Hormones on Male Reproduction: A Review. Molecules 2018, 23, 447. [Google Scholar] [CrossRef]
  171. Boga, J.A.; Caballero, B.; Potes, Y.; Perez-Martinez, Z.; Reiter, R.J.; Vega-Naredo, I.; Coto-Montes, A. Therapeutic potential of melatonin related to its role as an autophagy regulator: A review. J. Pineal Res. 2019, 66. [Google Scholar] [CrossRef]
  172. Tagliaferri, V.; Romualdi, D.; Scarinci, E.; Cicco, S.; Florio, C.D.; Immediata, V.; Tropea, A.; Santarsiero, C.M.; Lanzone, A.; Apa, R. Melatonin Treatment May Be Able to Restore Menstrual Cyclicity in Women With PCOS: A Pilot Study. Reprod. Sci. 2018, 25, 269–275. [Google Scholar] [CrossRef]
  173. Chen, Z.; Lei, L.; Wen, D.; Yang, L. Melatonin attenuates palmitic acid-induced mouse granulosa cells apoptosis via endoplasmic reticulum stress. J. Ovarian Res. 2019, 12, 43. [Google Scholar] [CrossRef]
Figure 1. The common in-vitro models used in PCOS studies are outlined. The primary GCs from humans and rodents, cell lines, and others related to the reproductive system or endocrine system are utilized in PCOS research.
Figure 1. The common in-vitro models used in PCOS studies are outlined. The primary GCs from humans and rodents, cell lines, and others related to the reproductive system or endocrine system are utilized in PCOS research.
Cells 12 02189 g001
Figure 2. PCOS is a disorder characterized by polycystic ovaries, hyperandrogenism, and ovulatory dysfunction. The common concepts pertaining to PCOS in in-vitro studies are shown. In-vitro models can be used to study signal transduction, microRNA regulation, and subsequent biological processes.
Figure 2. PCOS is a disorder characterized by polycystic ovaries, hyperandrogenism, and ovulatory dysfunction. The common concepts pertaining to PCOS in in-vitro studies are shown. In-vitro models can be used to study signal transduction, microRNA regulation, and subsequent biological processes.
Cells 12 02189 g002
Table 2. Primary cells from the ovary and other relevant cells for in vitro PCOS studies.
Table 2. Primary cells from the ovary and other relevant cells for in vitro PCOS studies.
Human Manipulation/TreatmentMainly DescriptionTargetRef.
Primary granulosa cellsfrom women underwent IVFProliferationTLR4[36]
TGFBR1, SMAD7[37]
Proliferation, apoptosisCTBP1-AS[38]
ESR2[39]
TGFBR1, TGFBR2[40]
MALAT1, MDM2[41]
Insulin resistanceSAA1[42]
IRS-1, GLUT4[43]
Mitochondrial damage
Glycolysis defection
SIRT3[44]
Androgen production
Antioxidative damage
CYP19A1, HO1[45]
Oxidative stressKeap1, Nrf2[46]
AutophagyBeclin-1, light chain 3[47]
MitophagySIRT1[48]
OvulationFHL2, AR[13]
RNA sequencingTNF-α[49]
NCI-H295R
(Adrenocortical carcinoma)
Decanoic acidAndrogen production3β-HSDII [54]
ForskolinAndrogen productionCYP17A1, 3β-HSDII [55]
NDAndrogen production3β-HSDII[56]
BMP2Steroidogenesis, androgen productionBMP2[57]
NCI-H295R
(Adrenocortical carcinoma)
NDSteroidogenesis, androgen productionCYP17, CYP21[58]
NCI-H295R
(Adrenocortical carcinoma)
Palmitate, H2O2, HNESteroidogenesis, androgen productionP450c17[59]
Animal
Rat primary granulosa cellsH2O2 (in vitro)Oxidative stress, autophagyBeclin-1, light chain 3[60]
DHT and FSH (in vitro)ProliferationPTEN[61]
DHEA (in vivo)Mitochondrial abnormality,
insulin resistance
NDUFB8, ATP5j,
IRS-1, GLUT4
[62,63]
TP (in vivo)AutophagyBeclin-1, light chain 3[64]
Vitamin K3 (in vitro)Oxidative stressKeap1, Nrf2[46]
Letrozole (in vivo)Insulin resistanceIGF-1, PI3K, AKT[65]
Mouse primary granulosa cellsDHT (in vitro)AutophagyBeclin-1, light chain 3[66]
IL-15 (in vitro)Proliferation, apoptosis, steroidogenesis, inflammationCYP17A1, Ifng, IL-1b[67]
DHEA (in vivo)Mitochondrial biogenesismtDNA[68]
Porcine primary granulosa cellsFSH and PDTC (in vitro)Autophagy, apoptosisNF-κB, IGF-1[69,70]
Bovine primary theca cellsIL-18 (in vitro)Proliferation, steroidogenesisCYP11A1, CYP17A1[71]
Mouse primary stroma cellsDHT and LH (in vitro)Lipid metabolism disorder, steroidogenesis, hypertrophyCol6a5[72]
Fish primary granulosa/theca cellsTestosterone (in vitro)ApoptosisMAPK[73]
3T3-L1
(Mouse preadipocyte)
Testosterone, LPSInflammationIL-6
MCP-1
[74]
DHTLipid metabolism disorderCol6a5[72]
NCTC1469
(Mouse liver cells
DHTLipid metabolism disorderCol6a5[72]
Abbreviations: TLR4, toll-like receptor 4; TGFBR, transforming growth factor beta receptor; SMAD7, mothers against decapentaplegic homolog 7; CTBP1-AS, C-terminal binding protein 1-antisense RNA; ESR2, estrogen receptor 2; MALAT1, metastasis associated lung adenocarcinoma transcript 1; MDM2, mouse double minute 2; SAA1, serum amyloid A1; IRS-1, insulin receptor substrate 1; GLUT4, glucose transporter type 4; SIRT, sirtuin; CYP19A1, cytochrome p450 family 19 subfamily A member 1; HO1, heme oxygenase 1; Keap1, kelch-like ECH-associated protein 1; Nrf2, nuclear factor like 2; FHL2, four and a half LIM domains 2; AR, androgen receptor; TNF-α, tumor necrosis factor- alpha; CYP17A1, cytochrome p450 family 17 subfamily A member 1; BMP2, bone morphogenetic protein 2; CYP21, cytochrome p450 family 21; HNE, hexanitroethane; PTEN, phosphatase and tensin homolog; NDUFB8, NADH: ubiquinone oxidoreductase subunit B8; ATP5j, ATP synthase peripheral stalk subunit F6; TP, testosterone propionate; IGF-1, insulin like growth factor 1; PI3K, phosphatidylinositol 3-kinase; AKT, protein kinase B; Ifng, interferon gamma; IL, interleukin; mtDNA, mitochondrial DNA; PDTC, pyrrolidine dithiocarbamate; NF-κB, nuclear factor kappa B; CYP11A1, cytochrome p450 family 11 subfamily A member 1; LH, luteinizing hormone; Col6a5, collagen type VI alpha 5 chain; MAPK, mitogen-activated protein kinase; LPS, Lipopolysaccharide; MCP-1, The monocyte chemotactic protein-1.
Table 3. Characterization of mouse primary follicle cultures from androgen-induced PCOS models.
Table 3. Characterization of mouse primary follicle cultures from androgen-induced PCOS models.
Mice SpeciesC57BL/6JC57BL/6JICRICRKunming
Sacrificed ageDay 18 to 21Week 13 to 16Week 6Week 6Day 14
follicle stageSecondary follicleSmall preantral follicle
(100 to 150 μm)
Large preantral follicle
(151 to 200 μm)
Small antral follicle
(201 to 250 μm)
Large antral follicle
(251 to 350 μm)
Preovulatory follicle
(351 to 450 μm)
Secondary follicle (100 to 160 μm)Secondary follicle
(100 to 160 μm)
Primary follicle
Culture Period6 days5 days13 days13 days14 days
Culture mediumαMEM +
10 mIU/mL rFSH
αMEM +
100 mIU/mL rFSH
αMEM +
33 mIU/mL rFSH or 100 mIU/mL rFSH
αMEM +
33 mIU/mL rFSH
αMEM +
100mIU/mL rFSH
and LH
Culture dish96 well dish4 well dish48 well dish96 well dish10 μL droplet
Treatment10 μM DHEA,
10 μM flutamide
DHT50, 500, 1250 ng/mL DHT500 ng/mL DHT
5 ng/mL Pioglitazone
10−6, 10−5, 10−4 M Testosterone
Reported outcomeDHEA induces impairment of follicle growth and ovulationProlonged exposure to excess DHT leads to aberrant follicle developmentDHT supports follicle development during the FSH-dependent preantral stagePioglitazone negatively affects follicular growthTestosterone promotes the follicle development
Ref.[84][90][92][93][94]
Abbreviations: αMEM, alpha minimal essential media; rFSH, recombinant FSH.
Table 5. MicroRNA expression and target genes in in vitro PCOS studies.
Table 5. MicroRNA expression and target genes in in vitro PCOS studies.
MicroRNATargetCellsCell ResponseRef.
miR-451aATF2KGNProliferation ↓, apoptosis ↑[144]
miR-194HB-EGFKGNProliferation ↓, apoptosis ↑[141]
miR-3188KCNA5KGNProliferation ↑, apoptosis ↓[142]
miR-206CCND2KGNProliferation ↓, apoptosis ↑[145]
miR-let-7d-3pTLR4KGNProliferation ↓[36]
miR-19bIGF-1KGNProliferation ↓[146]
miR-323-3pIGF-1KGNSteroidogenesis, apoptosis[147]
miR-21SNHG7KGNProliferation ↓, apoptosis ↑[148]
miR-186, miR-135aESR2KGNProliferation ↑, apoptosis ↓[39]
miR-29aNDKGN, COV434Proliferation ↓,
Aromatase expression↓, Estradiol biosynthesis ↓
[149]
miR-320aRUNX2Human primary GCsSteroidogenesis[150]
miR-21TLR8Human primary GCsProliferation ↑, apoptosis ↓[143]
miR-33b, miR-142TGFBR1Human primary GCsProliferation ↑, apoptosis ↓[37]
miR-423SMAD7Human primary GCsProliferation ↑, apoptosis ↓[37]
miR-324-3pWNT2BRat primary GCsProliferation ↓, apoptosis ↑[151]
↓, decrease; ↑, increase. Abbreviations: ATF2, cyclic AMP-dependent transcription factor 2; HB-EGF, Heparin-binding epidermal growth factor-like growth factor; CCND2, cyclin D2; SNHG7, small nucleolar RNA host gene 7; RUNX2, Runt-related transcription factor 2.
Table 6. Pharmacological agents used to investigate the pathophysiology of PCOS in in vitro models.
Table 6. Pharmacological agents used to investigate the pathophysiology of PCOS in in vitro models.
Agents for TreatmentsCellsSource/TreatmentTreatment OutcomeRef.
MetforminRat primary GCsH2O2 Decreased excessive autophagy in GCs[60]
MelatoninHuman primary GCsPCOS patientsReduced insulin resistance in GCs[43]
MelatoninHuman primary GCsPCOS patientsReduced androgen levels through ERK in GCs[45]
MelatoninKGN cellDHTProtected against mitochondrial injury in GCs of PCOS[48]
MelatoninKGN cellDHTAmeliorated mitochondrial membrane damage in GCs of PCOS[110]
MelatoninKGN cellDHEASuppressed autophagy and apoptosis[47]
HNG supplementation Rat primary GCsDHEA-inducedImproved local ovarian insulin resistance[62]
HNG supplementationCOV434Vitamin K3Alleviated oxidative stress in GCs of PCOS[46]
Vitamin D3Mice Primary GCsDHEAImproved mitochondrial biogenesis and membrane integrity[68]
SulforaphaneHuman primary GCsPCOS patientsAgainst oxidative stress by recudcing intracellular ROS and apoptosis levels[117]
Gui Zhu Yi Kun formulaRat primary GCsTPInhibited GC autophagy[64]
Cangfudaotan DecoctionRat primary GCsLetrozoleSuppressed insulin resistance and improves follicular development [65]
Guizhi Fuling WanRat primary GCsLetrozoleInhibited GC autophagy and promoted follicular development[156]
Xiao Yao SanRat primary GCsChronic unpredictable mild stressReduced apoptosis and autophagy of GCs[162]
Bu Shen Tian Jing FormulaKGNPalmitateImproving oxidative stress and glucose metabolism[163]
Bushen Huatan Granules and Kunling WanRat primary GCsDHEAProtected endoplasmic reticulum stress.[164]
Gengnianchun recipeKGNInsulinInhibited the senescence of GCs[165]
CryptotanshinoneRat primary GCshCG and insulinInhibited oxidative stress and apoptosis[157]
CryptotanshinoneKGNNDInhibited the proliferation and promote the apoptosis[38]
CryptotanshinoneRat primary GCshCG and insulinAttenuated hormone and inflammatory factor level[158]
CurcuminRat primary GCsDHEAInhibited endoplasmic reticulum stress[159]
PlumbaginRat primary GCsNDInhibited proliferation and promotes apoptosis[108]
Plumbaginmice primary GCsDHEAReduced the pyroptosis of GCs[161]
Table 7. The connection between parameters in in vitro models and pathological features of PCOS.
Table 7. The connection between parameters in in vitro models and pathological features of PCOS.
ParametersExperimental MeasurementsPathological Features of PCOS
Morphology
Apoptosis/AutophagyFlow cytometry, caspase 3 activity assay, TUNEL staining, LC3 Ⅱ/LC3 Ⅰ ratio, follicle survival rateFollicular atresia, follicular arrest[38,47,92]
Mitochondria structureElectron microscopy, mtDNA copy number measurementMitochondrial dysfunction, glycolysis defection, impaired oocytes[68]
ProliferationCell proliferation assay, follicle diameter measurementDisruption of folliculogenesis, follicular arrest[38,84]
Biological function
Lipid metabolismTriglyceride and Total Cholesterol measurement,
Oil Red O Staining
Dyslipidemia[72]
OvulationhCG-induced ovulation test,
RT-PCR for ovulation related gene
Oligo/Anovulation[84]
SteroidogenesisELISA, RT-PCR for steroidogenesis related geneSteroidogenesis imbalance[57]
Chemical/Stress response
AndrogensRT-PCR for AR related gene, western blottingHyperandrogenism[13]
CytokinesRT-PCR for pro-inflammatory cytokine related gene,
immunofluorescence, ELISA
Chronic inflammation[67]
InsulinIRS-1/GLUT4 RT-PCR, Glucose Consumption Assay
western blotting, immunofluorescence
Insulin resistance[62]
Oxidative stressMDA, SOD, and GSH detection Chronic inflammation, glycolysis defection, impaired oocytes, insulin resistance[72]
Abbreviations: TUNEL, Terminal deoxynucleotidyl transferase dUTP nick end labeling; LC3: Microtubule-associated protein light chain 3; RT-PCR: Reverse transcriptase PCR; ELISA: Enzyme-linked immunosorbent assay; MDA: Malondialdehyde; SOD: Superoxide dismutase; GSH: Reduced glutathione.
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

Tsai, Y.-R.; Liao, Y.-N.; Kang, H.-Y. Current Advances in Cellular Approaches for Pathophysiology and Treatment of Polycystic Ovary Syndrome. Cells 2023, 12, 2189. https://doi.org/10.3390/cells12172189

AMA Style

Tsai Y-R, Liao Y-N, Kang H-Y. Current Advances in Cellular Approaches for Pathophysiology and Treatment of Polycystic Ovary Syndrome. Cells. 2023; 12(17):2189. https://doi.org/10.3390/cells12172189

Chicago/Turabian Style

Tsai, Yi-Ru, Yen-Nung Liao, and Hong-Yo Kang. 2023. "Current Advances in Cellular Approaches for Pathophysiology and Treatment of Polycystic Ovary Syndrome" Cells 12, no. 17: 2189. https://doi.org/10.3390/cells12172189

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