Next Article in Journal
Biocatalytic Production of Aldehydes: Exploring the Potential of Lathyrus cicera Amine Oxidase
Next Article in Special Issue
Combined Anti-Adipogenic Effects of Hispidulin and p-Synephrine on 3T3-L1 Adipocytes
Previous Article in Journal
Natural Product Isoliquiritigenin Activates GABAB Receptors to Decrease Voltage-Gate Ca2+ Channels and Glutamate Release in Rat Cerebrocortical Nerve Terminals
Previous Article in Special Issue
Postprandial Bioactivity of a Spread Cheese Enriched with Mountain Tea and Orange Peel Extract in Plasma Oxidative Stress Status, Serum Lipids and Glucose Levels: An Interventional Study in Healthy Adults
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Potential Mechanisms of Plant-Derived Natural Products in the Treatment of Cervical Cancer

Xinjiang Key Laboratory of Biological Resources and Genetic Engineering, College of Life Science and Technology, Xinjiang University, Urumqi 830017, China
*
Authors to whom correspondence should be addressed.
Biomolecules 2021, 11(10), 1539; https://doi.org/10.3390/biom11101539
Submission received: 27 August 2021 / Revised: 1 October 2021 / Accepted: 13 October 2021 / Published: 18 October 2021

Abstract

:
Cervical cancer is the second most common gynecological malignancy globally; it seriously endangers women’s health because of its high morbidity and mortality. Conventional treatments are prone to drug resistance, recurrence and metastasis. Therefore, there is an urgent need to develop new drugs with high efficacy and low side effects to prevent and treat cervical cancer. In recent years, plant-derived natural products have been evaluated as potential anticancer drugs that preferentially kill tumor cells without severe adverse effects. A growing number of studies have shown that natural products can achieve practical anti-cervical-cancer effects through multiple mechanisms, including inhibition of tumor-cell proliferation, induction of apoptosis, suppression of angiogenesis and telomerase activity, enhancement of immunity and reversal of multidrug resistance. This paper reviews the therapeutic effects and mechanisms of plant-derived natural products on cervical cancer and provides references for developing anti-cervical-cancer drugs with high efficacy and low side effects.

1. Introduction

Cervical cancer is one of the second most common gynecologic malignancies in the world, with more than 570,000 women diagnosed with cervical cancer and 311,000 deaths annually worldwide [1]. However, persistent high-risk human papillomavirus (HPV) infection is the primary causative agent of cervical carcinogenesis, especially HPV16 and HPV18, which cause 70% to 72% of invasive cervical cancers. HPV contains two oncoproteins (E6 and E7) that play an essential role in the cervical cancer development and progression [2,3]. Currently, the main treatments for cervical cancer include surgical resection, radiotherapy, local targeted therapy and immunotherapy. Although traditional therapies are effective for early stage cervical cancer, they have limited efficacy for locally staged and metastatic cervical cancer due to severe side effect, drug resistance, multiple recurrences and metastases [4,5]. Therefore, there is an urgent need to develop drugs with a better safety profile and higher efficacy for cervical cancer treatment.
In recent years, plant-derived natural products have been considered as the most promising candidates for oncology therapies. They can preferentially kill tumor cells with low toxic effect on normal cells. This is attributed to their chemical diversity, structural complexity, inherent biological activity and low side effects [6,7]. A variety of plant-derived natural products with antitumor activities have been identified, such as flavonoids, terpenoids, alkaloids and phenols, which can inhibit tumor-cell proliferation, induce apoptosis, reduce telomerase activity, suppress angiogenesis, improve immune function, reverse multidrug resistance, etc. [8,9,10]. In this paper, we review the recent studies on the role and mechanisms of plant-derived natural products in the treatment of cervical cancer.

2. Methods

We collected the relevant experimental works in the literature published in the last five years that elucidate the anticancer effects of natural products on cervical cancer, using PUBMED (including Medline) and the Google Scholar database. Upon searching for the appropriate studies, we used “natural product, cervical cancer” as keywords. After completing the initial search, we removed duplicate works from the literature. Chemical structures of compounds derived from natural products were cited from the NCBI PubChem website, available online: https://pubmed.ncbi.nlm.nih.gov/ (accessed on 10 June 2021).

3. Flavonoids

Flavonoids are phenolic phytochemicals that are commonly found in fruits, vegetables and plant-based beverages (e.g., green tea and wine) [11]. More than 8000 species have been identified and isolated from plants. Their structures consist of C6-C3-C6 skeletons labeled with A, B and C rings, which can be classified into flavones, flavanones, flavonols, flavanols, isoflavones and anthocyanidins based on their structural diversity [12,13]. Many studies have reported that flavonoids have a significant role in tumor prevention and treatment attributed to their wide range of biological activities, such as anti-inflammatory, antioxidant, anti-hyperlipidemic, anti-fatigue, anti-aging, etc. [14,15,16]. In addition, they can induce apoptosis of tumor cells by inhibiting various pro-cancer pathways and genes in tumor cells.

3.1. Flavones

Scutellaria baicalensis is one of the versatile herbs traditionally and has been used in China to treat inflammatory diseases, hypertension, cardiovascular diseases, bacterial and viral infections. A large amount of evidence suggests that S. baicalensis also has potent anticancer activity, and its main bioactive components are baicalein, Wogonin, and baicalin [17,18]. Baicalein is a flavonoid derived from the roots of S. baicalensis and has a variety of pharmacological activities, which can inhibit cell proliferation and migration; induce apoptosis and cell-cycle arrest [19]. Cyclin D1 is a potential therapeutic target for cervical cancer. Baicalein inhibits cyclin D1 overexpression and arrests the cell cycle at G0/G1 phase through Wnt/β-catenin and protein kinase B/glycogen synthase kinase-3β (AKT/GSK-3β) signaling pathways to suppress proliferation and induce apoptosis of human cervical cancer HeLa and SiHa cells [20,21]. According to reports, baicalein inhibited tumor necrosis factor alpha (TNF-α)-induced the activation of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and the expression of apoptosis protein 1 (cIAP-1), cIAP-2, FLIP, B-cell lymphoma 2 (Bcl-2), matrix metalloproteinase 2 (MMP2), MMP9, first apoptosis signal receptor (Fas), FasL, caspase 8 and vascular endothelial factor (VEGF) in a dose-dependent manner to suppress HeLa cell invasion and migration. On the other hand, baicalein blocked TNF-α-induced nuclear translocation of p65 through inhibiting the phosphorylation and degradation of the inhibitory subunit of NF-κB (IκBα), activated caspase 8 and promoted the cleavage of poly (ADP-ribose) polymerase (PARP) expression to induce apoptosis in HeLa cells. In addition, baicalein had strong anti-inflammatory activity by inhibiting the phosphorylation of extracellular signal-regulated kinases 1/2 (ERK1/2) and p38, which reduced the expression of the inflammatory cytokines, including interleukin 8 (IL-8) and monocyte chemoattractant protein 1 (MCP1) [22,23,24]. Moreover, baicalein inactivated the AKT/mammalian target of rapamycin (mTOR) pathway by targeting the circHIAT1/miR-19a-3p axis to inhibit the proliferation of cervical cancer cells [25]. Another study illustrated that baicalin significantly suppressed cervical cancer xenograft tumor growth and metastasis in vivo through intraperitoneally injecting 10 mg/kg/d baicalin for four weeks. It was also reported that baicalein downregulated the expression of long non-coding RNA (lncRNA) in a dose- and time-dependent manner and named baicalein down-regulated long non-coding RNA (BDLNR). As it physically bound itself to Y-box binding protein 1 (YBX1), recruited YBX1 to the PIK3CA promoter, and it mediated the anti-cancer effects of baicalein in cervical cancer via activating PI3K/Akt pathway [26]. It is suggested that BDLNR may be a potential therapeutic target to enhance the anticancer effect of baicalein.
Wogonin is a natural monoflavonoid with the potential for selective tumor therapy in vitro and in vivo [27]. Wogonin could decrease the expression of HPV oncoproteins E6 and E7 and induce apoptosis in SiHa and CaSki cells through the mitochondria-mediated pathway, thus reducing mitochondrial membrane potential (MMP); elevating the Bcl-2-associated X protein (Bax)/Bcl-2 expression ratio; leading to cytochrome c (Cyt c) release; and triggering the cleavage of caspase 3, caspase 9 and PARP [28]. Wogonin also inhibited HeLa cell proliferation by inducing G1-phase cell-cycle arrest and apoptosis via decreasing the expression of cyclin D1, cyclin-dependent kinase 4 (CDK 4), pRb and nuclear transcription factor E2F-1, as well as increasing the expression of cyclin-dependent kinase inhibitor 1A/CDK-interacting protein 1(p21cip1) at the mRNA and protein levels through a p53-dependent mechanism [29]. Moreover, wogonin can enhance the effect of cisplatin on the induction of cancer cell apoptosis through reactive oxygen species (ROS)-dependent mechanism [30]. The results indicate that wogonin is a potential anticancer agent.
Apigenin is an edible natural flavonoid found in various dietary plant foods, such as vegetables and fruits [31]. It has a strong inhibitory effect on tumor-cell viability in vivo and in vitro. Its derivative apigenin 7-glucoside has anti-inflammatory, antioxidant and anticancer activities [32]. Apigenin 7-glucoside significantly suppressed the proliferation of HeLa cells, and the IC50 value at 48 h was 47.26 μmol/L. It induced apoptosis in HeLa cells through the death receptor pathway and the mitochondrial pathway, which effectively increased the expression of ROS, Fas, FasL, TNF-α, TNF-r1, Fas-associated death domain (FADD), TNF receptor-associated death domain (TRADD), caspase 3 and caspase 9 and decreased the expression of pro-caspase 8, caspase 10, Bcl-2 and Bcl-2 extra-large protein (Bcl-xl). Additionally, apigenin 7-glucoside treatment increased p16 INK4A expression and reduced Cyclin (A, D, E) and CDK2/6 expression. Meanwhile, it inhibited HeLa cell migration by targeting the phosphatase and tensin homolog (PTEN)/PI3K/AKT pathway [33]. Some studies reported that apigenin showed selective sensitivity for human cervical cancer HeLa, SiHa, CaSki and C33A cells with IC50 values of 10, 68, 76 and 40 μmol/L, respectively. Apigenin also induced mitochondrial redox damage, decreased MMP and lipid peroxidation and inhibited the migration and invasion of cervical cancer cells [34]. Zhang et al. found that Apigenin suppressed cervical tumor growth in vivo by attenuating histamine-induced abnormal estrogen receptor signaling by increasing the estrogen receptor β/estrogen receptor α (ERβ/ERα) ratio when tumor model mice were administered intraperitoneally with 100 mg/kg apigenin +1 mg/kg histamine every 3 d. Moreover, apigenin induced autophagy and apoptosis in HeLa cells through the PI3K/AKT/mTOR signaling pathway [35]. Consequently, apigenin and its derivatives may prevent the development and progression of cervical cancer and represent promising drugs for cervical cancer therapy.
Luteolin is a common plant flavonoid found in various plants, including fruits, vegetables and medicinal herbs. Luteolin-rich plants have been used to treat multiple diseases, such as hypertension, cancer and inflammation [36,37]. Luteolin could dose-dependently reduce the proliferation of HeLa cells with IC50 value of 21.8 μmol/L. Luteolin induced apoptosis and G2/M phase cell-cycle arrest in HeLa cells by downregulating UHRF1 and DNA methyltransferase 1 (DNMT1) with a reduction of overall DNA methylation and upregulating p16 INK4A expression [38]. On the other hand, luteolin promoted apoptosis by inhibiting TNF-α-induced NF-κB activation, downregulating A20 and c-IAP1 gene expression and enhancing c-Jun N-terminal kinase (JNK) activity [39]. It was also reported that luteoloside inhibited HeLa cell proliferation by endogenous and exogenous pathways, which increased the ratio of Bax/Bcl-2 to reduce MMP and release Cyt c, upregulated Fas expression and activated caspase 8 and caspase 3. At the same time, luteoloside inhibited mTOR and activated p38 mitogen-activated protein kinase (MAPK) signaling pathways to exert anti-cervical-cancer effects [40]. In addition, luteolin combined with TNF-related apoptosis-inducing ligand (TRAIL) synergistically induced apoptosis in HeLa cells through upregulation of death receptor 5 (DR5) and Bid cleavage, and activation of caspase-8 [41]. These data suggest that luteolin may serve as a new therapeutic strategy for cervical cancer.

3.2. Flavanones

Citrus fruits are rich in flavonoids and are known for their health-promoting and chemopreventive properties. Naringin, a flavonoid glycoside, can be isolated from citrus fruits, such as orange, tangerine, lemon and lime, and has several pharmacological activities [42]. Naringin can significantly reduce cancer cell viability and proliferation without toxicity to normal tissue cells [43]. The IC50 values of naringin for HeLa, SiHa and C33A cells after 24-h treatment were 793, 764 and 745 μmol/L, respectively, and showed a dose-dependent relationship. Naringin increased the level of endoplasmic reticulum (ER) stress sensors, phosphorylated eIF2α and activated the apoptosis-related protein CHOP and other associated pro-apoptotic proteins (PARP1). Importantly, naringin also blocked the β-catenin signaling pathway by decreasing β-catenin (Ser576) and GSK-3β (Ser9) protein expression and phosphorylation and induced cell-cycle arrest at G0/G1 phase by increasing the expression of cell-cycle checkpoint proteins p21/cip and p27/kip to trigger apoptosis in cervical cancer cells [44]. Another study reported that naringin promoted the expressions of caspases (3, 8 and 9), p53, Bax, Fas death receptor and its adaptor protein FADD to induce apoptosis in HeLa cells through the death receptor pathway and the mitochondrial pathway. The activation of the exogenous pathway was associated with increased caspase 8, which activated the death receptor by cleaving Bid into tBid and the crosstalk between the death receptor and mitochondrial pathways [45]. Moreover, naringin also induced growth inhibition and apoptosis in HeLa cells by decreasing the expression of NF-κBp65, COX-2 and caspase 1 [46]. The results suggest that naringin is a potentially effective drug for treating human cervical cancer.
Hesperidin is a citrus flavonoid and has various functions, including antioxidation, antitumor and anti-angiogenesis functions, which can protect mitochondrial membranes from free radical attack and inhibit tumor-cell migration and invasion [47,48]. Hesperetin inhibited the cell viability of SiHa cells in a dose- and time-dependent manner with IC50 values of 650 μmol/L. Hesperetin induced cell-cycle arrest at G2/M phase and apoptosis in SiHa cells by death receptor and mitochondrial pathways, characterized by depolarizing MMP and increasing the expression of caspase 3, caspase 8, caspase 9, p53, Bax, Fas and FADD [49]. In addition, hesperidin inhibited HeLa cell proliferation and induced apoptosis through ER stress and mitochondria-mediated pathways via decreasing the expression of cyclin D1, cyclin E1 and CDK2, promoting Cyt c release, and activating the expression of apoptosis-inducing factor (AIF), GADD153/CHOP and glucose-regulated protein of 78 kDa (GRP78) [50]. Hesperidin has potential in preventing and treating cervical cancer and may open new avenues for cancer treatment.
Silibinin is a bioactive polyphenolic flavonoid isolated from the fruits and seeds of silybum marianum. It has been used to treat various diseases, especially the liver, gallbladder and kidney [51,52]. Silibinin inhibited the growth of Hela and SiHa cells in a dose- and time-dependent manner, with IC50 values of 332 and 275 μmol/L for HeLa cells, and 250 and 195 μmol/L for SiHa cells at 48 h and 72 h, respectively. It reduced ATP content, mtDNA copy number and MMP, activated the dynamin-related protein 1(Drp1)-mediated mitochondrial fission pathway, induced G2/M cell-cycle arrest through reducing the expression of CDK1, cyclin B1 and Cdc25C [53]. Silibinin also suppressed angiogenesis and promoted apoptosis in HeLa cells through inhibiting the mTOR/p70S6K/4E-BP1 signaling pathway and decreasing the accumulation and transcriptional activity of hypoxia-inducible factor-1α (HIF-1α) and the release of VEGF. Moreover, the anti-angiogenic ability of silibinin was enhanced by blocking AKT activation through PI3K/AKT inhibitor LY294002 [54]. At the same time, silibinin induced apoptosis in HeLa cells through the mitochondrial pathway and the death-inducing pathway, resulting in decreased expression of CDK1 and CDK2 proteins and increased ratio of Bax/Bcl-2, followed by Cyt c release and activation of caspase 9, as well as increased expression of Fas and FasL and activation of caspase 8 [55]. The evidence has shown that hydroxyl radical (-OH) is the main form of silibinin-induced ROS [56]. Silibinin inhibits HeLa cell growth by activating apoptotic vesicles and caspase 3, and promoting the phosphorylation of p53 and JNK in a dose-dependent manner. p53 subsequently interferes with mitochondrial function through the p53-upregulated modulator of apoptosis (PUMA) pathway, which upregulates the Bax/Bcl-2 ratio, reduces MMP and increases ROS production. Then ROS induces autophagy and apoptosis in HeLa cells. Furthermore, p53-mediated glutathione (GSH) depletion significantly enhanced the cytotoxicity of NO in HeLa cells [57,58]. It may be a classical candidate for the design of anticancer drugs.

3.3. Flavonols

Kaempferol is a natural flavonol and widely distributes in many plant families. The growing evidence has shown that kaempferol is a potential cancer therapeutic agent with potent antitumor, anti-inflammatory and antioxidant properties [59]. It inhibited SiHa cell growth and proliferation in a dose- and time-dependent manner, and promoted SiHa cell apoptosis by disrupting MMP and elevating intracellular free Ca2+ concentration, which caused shrinkage of spindle-shaped SiHa cells and damage of their microtubule networks [60]. Another study showed that Kaempferol inhibited HeLa cell growth in a time- and concentration-dependent manner and the IC50 value was 10.48 μmol/L at 72 h, while it had weak toxicity for normal cells with IC50 value of 707 μmol/L at 72 h. It induced apoptosis and senescence in HeLa cells by inhibiting PI3K/AKT pathway and human telomerase reverse transcriptase (hTERT) expression and promoting the p53 pathway [61]. Telomerase is considered a new and potentially selective target for tumor therapy. Telomerase inhibition by kaempferol may provide a safe and effective approach for the treatment of cervical cancer. Additionally, overexpression of P-glycoprotein (P-gp) causes efflux of chemotherapeutic drugs from cells and is considered to be one of the crucial mechanisms of multidrug resistance (MDR) in cancer [62]. It was reported that kaempferol significantly decreased the activity and function of P-gp in MDR human cervical cancer KB-V1 cells in a dose-dependent manner, reduced drug efflux, improved sensitivity to chemotherapeutic drugs vinblastine and paclitaxel as well as cytotoxicity, which in turn induced apoptosis and reversed MDR in KB-V1 cells [63]. Kaempferol might be a potential candidate for the prevention and treatment of cancer due to its safety and low-cost advantages.
Quercetin is a dietary polyphenolic compound with wide distribution in vegetables and fruits and has various activities, including anti-allergic, anti-inflammatory and antitumor activities [64,65]. Quercetin can suppress tumor-cell proliferation and induce apoptosis, cell-cycle arrest and DNA damage through intrinsic apoptosis pathway, which involved in PI3K, MAPK and Wnt [66]. Priyadarsini et al. showed that quercetin could target both opposing signaling pathways, p53 and NF-κB to inhibit the proliferation of HeLa cells [67]. Quercetin induced apoptosis of HeLa and SiHa cells by blocking the interaction of E6/E6AP complex, reactivating p53 and upregulating the expression of transcriptional target p21 [68]. Further, quercetin promoted apoptosis and reduced migration of HeLa cells within 18 h by downregulating the expression of AKT and Bcl-2, and blocked the cell cycle at the G2/M phase. The accumulation of ROS increased Cyt c release and MMP depolarization and activated caspase 3, which in turn exhibited significant anti-proliferative and pro-apoptotic effects on HeLa cells [69]. Interestingly, the combination of quercetin with other chemical agents effectively enhanced the antitumor effect. Quercetin inhibited the viability of HeLa and SiHa cells in a dose- and time-dependent manner, with an IC50 of 30 μmol/L for HeLa cells at 24 h and 50 μmol/L for SiHa cells at 48 h. Treatment with cisplatin or quercetin alone did not reduce the expression of MMP2 protein in cervical cancer cells but their combination significantly decreased MMP2 expression, which inhibited the migration and invasion of cervical cancer cells. Meanwhile, quercetin enhanced the chemosensitivity of cervical cancer cells by downregulating the expression of P-gp and methyltransferase-like 3 (METTL3), which mediated HeLa cell proliferation and apoptosis [70]. Quercetin also increased the sensitivity of HeLa cells to cisplatin by inhibiting the expression of the multidrug resistance-associated protein (MRP) and heat shock protein Hsp72, which induced apoptosis and reversed cellular resistance [71]. These studies provided the experimental basis for the treatment of cisplatin-resistant patients.
Fisetin is a natural flavonols found in vegetables, fruits and nuts and has antitumor, anti-invasive, anti-angiogenic, antidiabetic, cardioprotective and neuroprotective activities [72,73]. Fisetin suppressed urokinase-type plasminogen activator (u-PA) expression by blocking the phosphorylation of p38 MAPK and the nuclear translocation of NF-κB, which inhibited the migration and invasion of SiHa cells. Moreover, the addition of the p38 MAPK inhibitor SB203580 further enhanced the inhibitory effect of fisetin on u-PA activity and expression [74]. It was also reported that fisetin not only promoted apoptosis protease activating factor-1 (Apaf-1) expression and Cyt c release to activate caspase 3 and caspase 9 but also inhibited ERK1/2 phosphorylation, COX-2 expression and prostaglandin E2 (PGE2) production by blocking the NF-κB/p300 signaling pathway, which in turn promoted HeLa cell apoptosis [75]. Furthermore, fisetin significantly inhibited the growth rate of tumors with inhibition rates of 82.65% and 92.62% in a mouse model of HeLa cell line injection without significant side effects [76]. It was also experimentally proven that fisetin combined with sorafenib could activate the DR5-mediated death receptor pathway and mitochondria-dependent pathway, which upregulated Bax/Bcl-2 ratio and promoted MMP depolarization, caspase 3/caspase 8 activation and PARP cleavage to induce apoptosis in HeLa cells [77]. Fisetin is expected to be an anticancer drug for the clinical treatment of cervical cancer.

3.4. Flavanols

Green tea (Camellia sinensis) is one of the most commonly used herbs globally and is widely known for its effectiveness in preventing chronic diseases and tumors. This is mainly attributed to the biologically active catechin compounds in green tea, such as (−)-epigallocatechin gallate (EGCG), (−)epigallocatechin 3-gallate (ECG), (−)epigallocatechin (EGC) and (+) catechin. Among them, EGCG is the main component of catechins and has substantial antioxidative, antitumor and anti-angiogenic effects [78,79]. After treatment with 150 and 300 μg/mL of catechin for 72 h, apoptosis rates of SiHa cells reached 31.62% and 34.8%, with an IC50 value of 196.07 μg/mL. In addition, catechin inhibited the proliferation and induced apoptosis of SiHa cells partly by regulating TP53 and caspase 3, caspase 8 and caspase 9 [80]. EGCG concentration- and time-dependently inhibited HeLa cell proliferation with an IC50 value of 20 µg/mL. EGCG inhibited the activation of AKT and NF-κB by blocking the phosphorylation and degradation of the inhibitory κBα and κBβ subunits to result in the downregulation of COX-2 expression. Furthermore, EGCG treatment led to mitochondrial dysfunction through increasing ROS production, p53 and Bax/Bcl-2 ratios to promote Cyt c release and activation of caspase cascade, which induced apoptosis in HeLa cells [81]. Recently, one study indicated that EGCG inhibited transforming growth factor-β (TGF-β)-induced epithelial-mesenchymal transition (EMT) in Hela and SiHa cells via the ROS/Smad signaling pathway to inhibit cell migration and invasion [82]. Besides, EGCG can be used as an anti-angiogenic agent in the treatment of cervical cancer. EGCG significantly suppressed hypoxia and serum-induced accumulation of HIF-1a protein, as well as the expression of VEGF by blocking PI3K/Akt/mTOR and ERK1/2 signaling pathways, thereby inhibiting HeLa cell angiogenesis [83]. Meanwhile, EGCG enhances the sensitivity of cisplatin to cervical cancer cells by inhibiting the mTOR signaling pathway and the levels of p-p70S6K1 and p-4E-BP1, which in turn inhibits cell viability and induces apoptosis in HeLa cells [84]. It has also been shown that EGCG can induce apoptosis in HeLa cells by inhibiting telomerase activity [85]. It is suggested that telomerase inhibition may be one of the critical mechanisms of EGCG treatment.

3.5. Isoflavones

Genistein is considered the primary isoflavone in soy foods and is one of the most widely studied phytoestrogens in the Asian diet [86]. It was demonstrated that oral administration of 20 mg/kg of genistein exerted anti-proliferative and immunomodulatory effects in cervical cancer mouse model, which promoted lymphocyte proliferation and lactate dehydrogenase (LDH) release, enhanced cytolytic activity and IFN-γ production, and thus induced protective antitumor immunity [87]. AKT signaling pathway plays a crucial role in controlling cell survival and apoptosis. Genistein enhanced the activity of cisplatin by inhibiting the expression of the AKT/mTOR pathway, p-p70S6K1 and p-4E-BP1, which led to the growth inhibition of cervical cancer cells [88]. Several studies indicated that inhibition of ERK or PI3K signaling pathways, along with activation of the p38-JNK pathway, shifted the balance toward apoptosis to result in higher levels of growth inhibition in a variety of tumor-cell lines [89,90]. Kim et al. found that genistein effectively inhibited the growth of HeLa and CaSki cells with IC50 values of 20 and 60 μmol/L after 48 h. It inhibited cervical cancer cell proliferation by attenuating ERK1/2 activity and activating p38 MAPK, where AKT and JNK were partially involved in genistein-induced cancer cell growth inhibition [91]. In previous study, phytoestrogens have also been found to have an anti-tumor effect [92]. Genistein blocked the estrogen receptor-mediated PI3K/AKT-NF-κB signaling pathway and exerted antitumor effects by downregulating Bcl-2, VEGF, and tumor expression necrosis factor-associated receptor factor 1 (TRAF1) and promoting apoptosis in HeLa cells [93]. Genistein also induced apoptosis in HeLa cells via a p53-dependent pathway that disrupted the interaction between p53 and APE1 to increase the intracellular stability of p53 [94]. Therefore, genistein may have potential clinical application in the treatment of cervical cancer.
Puerarin is the main active ingredient extracted from Pueraria tuberosa, which has beneficial effects on cardiovascular diseases, neurological dysfunction, osteoporosis, liver injury and inflammation. Puerarin could effectively inhibit the production of pro-inflammatory cytokines in a variety of disease models [95]. Puerarin also inhibited tumor-cell proliferation in a dose- and time-dependent manner through inducing apoptosis and blocking cellular cyclin D1 expression by PI3K/AKT/NF-κB signaling pathway [96]. Moreover, puerarin gavaged with 500 mg/kg for 15 d could increase the level of IL-2 and superoxide dismutase (SOD) activity in the plasma of U14 cervical cancer mice to enhance the immune function and antioxidant enzyme activity to scavenge excessive free radicals and reduce the damage of ROS in the body, thus achieving the antitumor effect [97]. Another study proved that puerarin (12.5–50 μmol/L) could inhibit HeLa cell proliferation and induce apoptosis through suppressing the β-catenin/Wnt/p53 signaling pathway [98]. Puerarin also significantly inhibited the expression of p-mTOR protein and activated caspase 3/9 in HeLa cells to reduce their proliferation and migration [99]. Thus, puerarin is expected to be a new drug for the prevention and treatment of cervical cancer.
Formononetin is a 7-hydroxy-40-methoxy herbal isoflavone isolated from Pycnanthus angolensis, red clovers, chickpea and other plants, and it possesses a variety of pharmacological activities, including antioxidant, chemopreventive, anti-inflammatory, anti-allergic, antibacterial and cardio-preventive effects [100,101]. According to the research, formononetin might mediate MRP (MRP1 and MRP2) inhibition through ROS while activate the mitochondrial apoptosis pathway, which resulted in MMP loss, increased Bax/Bcl-2 ratio and activation of caspase 3 and caspase 9. On the other hand, formononetin could also enhance the cytotoxicity of epirubicin in HeLa cells through the death receptor/caspase 8 apoptotic pathway [102]. Apart from this, formononetin inhibited HeLa cell proliferation in a concentration-dependent manner with an IC50 value of 72.112 ± 5.671 µmol/L. Multi-walled carbon nanotubes are the best carrier of formononetin, and their combination can further enhance ROS-mediated mitochondrial dysfunction [103]. Moreover, oral gavage of 10 mg/kg of formononetin for 31 d significantly inhibited tumor growth in BALB/c mouse model inoculated with HeLa cells and reduced the expression of HIF-1α and VEGF in tumor tissues without severe side effects [104]. Moreover, formononetin (oral gavage of 20 and 40 mg/kg for 35 d) activated caspase 3 and upregulated the expression of Bax by inhibiting the transduction of PI3K/AKT signal pathway, accompanied by the release of Cyt c, and then induced apoptosis of HeLa cells [105]. It might be a potential drug for cervical cancer treatment.

3.6. Anthocyanins

Anthocyanins are natural pigments in the plant kingdom and endow plants their blue color and antioxidant potential. Anthocyanins are mainly found in small berries, such as strawberries, blueberries, cranberries, cherries and black raspberries. More than 600 structurally diverse anthocyanins have been identified in nature [106]. Cyanidin 3-O-glucoside is the main active component of anthocyanins, which has preventive and therapeutic effects on various diseases such as cerebral ischemia and Alzheimer’s disease [107]. The study showed that the inhibitory effects of cisplatin alone, cyanidin 3-O-glucoside alone and their combination on the proliferation of HeLa cells were 17.43%, 34.98% and 63.38%, respectively. Cyanidin 3-O-glucoside treatment downregulated the PI3K/AKT/mTOR signaling pathway; activated Bax, p53 and tissue inhibitor of metalloproteinase-1 (TIMP-1); decreased cyclin D1 and Bcl-2 expression; and blocked the cell cycle at the G1 phase, thus promoting apoptosis in HeLa cells [108]. Importantly, cyanidin 3-O-glucoside enhanced the sensitivity of HeLa cells to cisplatin. Their combination significantly inhibited the activities of SOD, catalase (CAT) and glutathione peroxidase (GSH-PX) by reducing the expression of NAD(P)H quinone dehydrogenase 1 (NQO1) and heme oxygenase-1 (HO-1), as well as modulating the nuclear factor E2-related factor 2 (Nrf2) signaling pathway to induce oxidative stress and apoptosis in HeLa cells [109]. It is suggested that cyanidin 3-O-glucoside may increase the antitumor activity of cisplatin to reduce the adverse effects associated with chemotherapy in cervical cancer and is a potential strategy for the treatment of cervical cancer.
The chemical structures of flavonoids are summarized in Figure 1. Table 1 summarizes their effects and mechanisms on cervical cancer.

4. Terpenoids

4.1. Monoterpenoids

Paeoniflorin, a natural monoterpene glycoside compound, is the most important active component of the Chinese medicinal herb Paeonia lactiflora and has various protective effects on the cardiovascular system as mediating anti-inflammatory, antioxidant, apoptotic and autophagic regulation [110,111]. Recently, it has been found that paeoniflorin dose- and time-dependently inhibited the proliferation of human endometrial cancer RL95-2 cells by activating p38 MAPK and NF-κB signaling pathways [112]. It also inhibited the proliferation of human gastric cancer MGC-803 cells by inducing apoptosis through upregulation of microRNA-124 (miR-124) and inhibition of PI3K/AKT and transcription activator 3 (STAT3) signaling pathway [113]. In addition, the growth inhibition rate of paeoniflorin on human cervical cancer HeLa cells showed a concentration- and time- dependent relationship, with IC50 values of 5054, 2965 and 2459 μg/mL for 24, 48 and 72 h of treatment. It increased Bax/Bcl-2 ratio and mitochondrial outer membrane permeability, caused the release of Cyt c into the cytoplasm, and activated Apaf-1 and caspase 3/9, which induced apoptosis in HeLa cells [114], and it is considered as a potential agent for the treatment of cervical cancer.
Carvacrol is a monoterpene phenol produced by many aromatic plants, including thyme, oregano and other plants. It has various biological and pharmacological properties, such as antibacterial, anticancer, anti-inflammatory, hepatoprotective, antispasmodic and vasodilator properties [115,116]. Carvacrol dose-dependently enhanced cytotoxicity in HeLa cells with IC50 value of 556 ± 39 μmol/L after 24 h treatment. In addition, the cytotoxicity of carvacrol was further increased by MEK inhibitor PD325901, which inhibited ERK and increased the levels of LC3β-I/II. Carvacrol induced apoptosis through promoting caspase 9 expression and PARP cleavage [117]. Moreover, carvacrol induced cell-cycle arrest by increasing p53 expression and decreasing cyclin D1 expression in HeLa cells [118]. It may be a promising adjuvant therapeutic agent.

4.2. Sesquiterpenoids

In recent years, artemisinin and its derivatives (e.g., dihydroartemisinin and artesunate) have been recognized as effective antimalarial agents [119]. They also have anticancer and anti-angiogenic properties with cytotoxic effects against several cancer types in vitro and in vivo [120,121,122]. Dihydroartemisinin inhibited the proliferation of HeLa and Caski cells in a dose- and time-dependent manner with IC50 values of 22.08 and 18.20 μmol/L, respectively. It induced apoptosis through upregulation of Raf kinase inhibitor protein (RKIP) and downregulation of Bcl-2. In a BALB/c mouse model inoculated with HeLa cells or Caski cells, intraperitoneal injection of dihydroartemisinin for 3 weeks significantly inhibited tumor growth with inhibition rates among 70–80% [123]. Moreover, dihydroartemisinin induced apoptosis in HeLa cells and human endometrial cancer cells by decreasing the expression of TfR mRNA and increasing the expression of caspase 3 mRNA. It also led to an increase in LC3-I/II ratio and a decrease in p62 protein levels, which induced autophagic pathway-mediated cell death. The inhibition of the autophagic pathway by using 3-MA further enhanced the cytotoxicity of dihydroartemisinin to HeLa cells [124]. In addition, dihydroartemisinin promoted the production of ROS by upregulating γH2AX protein and foci formation, led to DNA double-strand breaks, while upregulated the phosphorylation of Bcl-2 (Ser70) and mTOR (Ser2448), increased expression of the pro-autophagic protein Beclin-1 and induced autophagic death of Hela cells [125]. Zhang et al. found that dihydroartemisinin significantly inhibited the cell viability of HeLa cells by upregulating the expression of caveolin 1 (Cav1) and mitochondrial carrier homolog 2 (MTCH2) to activate p53 and decrease NQO1 expression, which contributed to the activation of the cell-death pathway and promoted apoptosis in HeLa cells [126].
Artesunate effectively enhanced TRAIL-mediated cytotoxicity by reducing the expression of pro-survival proteins (survivin, XIAP and Bcl-xl) via inhibiting the activation of NF-κB and AKT, and increased apoptosis induced by TRAIL [127]. Besides, artesunate dose-dependently inhibited HeLa and SiHa cell proliferation with IC50 values of 5.47 and 6.34 μmol/L, respectively. Interestingly, artesunate in vitro and in vivo increased the radiosensitivity of HeLa cells by promoting apoptosis and G2/M phase transition induced by X-ray irradiation, which was related with the increased expression of cytosolic cyclin B1. It was suggested that it could be applied as an effective radiosensitizer in cancer therapy [128]. Meanwhile, artesunate inhibited cell proliferation by suppressing COX-2 expression in HeLa and CaSki cells, and it decreased PGE2 production and the percentage of CD4+CD25+Foxp3+ T cells [129]. It was reported that subcutaneous injection of 100 mg/kg of artesunate for 15 d effectively reduced the growth and angiogenesis of xenograft tumors by suppressing the secretion of VEGF and the expression of vascular endothelial growth factor receptor KDR/flk-1 on tumors [130]. The results indicate that dihydroartemisinin and artesunate are promising candidates for cervical cancer treatment.

4.3. Diterpenoids and Terperpenoids

Tanshinone IIA is a diterpenoid naphthoquinone found in traditional Chinese medicine Salvia miltiorrhiza. It has anti-inflammatory and antioxidant activities and has been commonly used in prevention and treatment of cardiovascular disease [131,132]. Radha et al. found that tanshinone IIA caused a significant increase in the expression of p53, p21cip1/waf1, pRb and p130, and activated p53-dependent anticancer activity by inhibiting the expression of HPV E6 and E7 proteins, leading to growth inhibition of cervical cancer cells. In preclinical studies, intraperitoneal injection of 30 mg/kg of tanshinone IIA for 8 weeks significantly reduced the expression of the proliferation marker PCNA in tumor tissues and the volume of cervical cancer transplanted tumors in nude mice by more than 66% [133]. In addition, tanshinone IIA also reduced glycolysis by inhibiting intracellular AKT/mTOR and HIF-1α activities. In vivo, intraperitoneal injection of 40 mg/kg of tanshinone IIA, administered every 2 d for 20 d, effectively inhibited tumor growth and metastasis in U14 cervical cancer mice with an inhibition rate of 72.7% [134]. Another study found that tanshinone IIA inhibited the proliferation of HeLa, SiHa, CaSki and C33A cells in a dose-dependent manner with IC50 values of 6.97, 14.47, 5.51 and 9.89 μmol/L respectively, and showed a higher anti-proliferative effect than paclitaxel (18.45 μmol/L). Furthermore, tanshinone IIA treatment activated caspase 9 and 3 to cleave PARP, increased the ratio of Bax/Bcl-2 and JNK, released Cyt c that interacted with Apaf-1, and phosphorylated p38 to trigger ER stress-mediated apoptosis through IRE1 and PERK-related pathways. Notably, tanshinone IIA could improve paclitaxel chemosensitivity, suggesting that it may be a potential strategy to overcome paclitaxel resistance [135]. Importantly, tanshinone IIA inhibited the migration and invasion of cervical cancer stem cells (CSCs) by inhibiting the transfer of HuR from the nucleus to the cytoplasm to reduce the stability and transcriptional activity YAP gene. It was also found that tanshinone IIA not only directly killed the activity of cervical CSCs, but also restored the sensitivity of cervical CSCs to adriamycin [136]. It can be used as an effective therapeutic agent for treating patients with cervical cancer and chemotherapy resistance.
Oridonin is an active diterpene isolated from Rabdosia rubescens with various pharmacological and physiological effects, including antibacterial, anti-inflammatory and antitumor [137]. Oridonin as an AKT inhibitor suppressed the growth of cancer cells by attenuating AKT signaling [138]. Hu et al. reported that oridonin inhibited the cell viability of HeLa cells in a dose- and time-dependent manner with an IC50 value of 4.13 μmol/L (48 h) [139]. Several studies have shown that oridonin induced apoptosis involving multiple molecular pathways. It significantly inhibited the activity of constitutively activated targets of PI3K (Akt, forkhead box O protein (FOXO) and GSK3) in HeLa cell lines. Treatment of HeLa cells with oridonin activated the cell-death pathway by downregulating Akt kinase signaling, causing loss of MMP to trigger Cyt c release from the mitochondria, leading to downstream activation of caspase 9, caspase 3 and downregulating the expression of the survival proteins (cIAP1, XIAP and survivin), which led to cervical cancer cell death [140,141]. Additionally, ROS is the initiation signal for mitochondrial and caspase-dependent apoptosis as well as autophagy. Oridonin induced ROS production in HeLa cells in a dose-dependent manner, significantly increased Bax and caspase 8 and decreased pro-caspase 3, pro-caspase 9 and Bcl-2, which induced apoptosis and autophagy in HeLa cells [142]. Oridonin also induced apoptosis in HeLa cells through downregulating the level of p-AKT protein and inhibiting the expression and activity of cellular telomerase FKHRL and GSK3β [143]. It may be a key drug in the treatment of cervical cancer.
Ginsenoside Rh2 is a biologically active compound derived from ginseng and has various health effects, including stimulation of immune function, enhancement of cardiovascular health and anti-stress capacity [144]. Ginsenoside Rh2 activated mitochondria-dependent apoptosis pathway and inhibited mitochondrial oxidative phosphorylation and glycolysis in HeLa cells. In addition, ginsenoside Rh2 inhibited energy metabolism and induced apoptosis in HeLa cells by upregulating voltage-dependent anion channel 1 (VDAC1), which caused MMP depolarization and ROS production [145]. It is suggested that VDAC1 is a new target of ginsenoside Rh2. Ginsenoside Rh2 inhibited Hela cell proliferation in a dose- and time-dependent manner by targeting the Akt pathway, and prevented HeLa cell migration and invasion by inhibiting Akt/GSK3β, Snail expression and EMT occurrence [146]. Besides, the combination of ginsenoside Rh2 and betulinic acid induced Bax translocation to mitochondria and released Cyt c, activated caspase 8 and Bid cleavage, and sensitized tumor cells through Bax-dependent mechanism, which decreased cell viability and induced apoptosis [147]. Therefore, ginsenoside Rh2 has the potential to be a novel anticancer drug for cervical cancer.
Betulinic acid, a natural pentacyclic triterpene found in white birch bark, is one of the most promising cancer therapeutic compound with protective effects against tumor progression, inflammation, metabolic diseases and cardiovascular diseases [148,149]. Betulinic acid induced apoptosis in HeLa cells through modulation of PI3K/AKT and mitochondrial pathways. It dose-dependently inhibited the viability of HeLa cells with an IC50 value of 30.42 ± 2.39 μmol/L at 48 h of treatment, the phosphorylation of AKT at Thr308 and Ser473, activated Bad, caspase 9 and cell-cycle regulatory factors p27Kip and p21Waf1/Cip1, and induced cell-cycle arrest at G0/G1 phase [150]. Betulinic acid also mediated the accumulation of HIF-1α and inhibited expression of HIF target genes VEGF, GLUT1 and PDK1 in HeLa cells by directly activating proteasome β1, β2 and β5 [151]. Further, betulinic acid decreased Bcl-2 and cyclin D1 and increased the expression of Bax genes, which induced apoptosis and inhibited cell proliferation and migration [152]. It has been shown that betulinic acid is not toxic in vivo at a maximum dose of 500 mg/kg [153]. Betulinic acid is expected to be an adjuvant candidate for human cancer therapy shortly.
The chemical structures of terpenoids are summarized in Figure 2. Table 2 summarizes their effects and mechanisms on cervical cancer.

5. Alkaloids

5.1. Piperine

Piperine is a nitrogenous stimulant found in the fruits of black pepper and long pepper. It has various pharmacological properties in vitro and in vivo, such as anticancer, antibacterial, antiulcer hepatoprotective and immunomodulatory properties [154,155]. Asif et al. showed that piperine treatment at 50, 100 and 200 μmol/L dose-dependently reduced the cell viability of HeLa cells to 69.90%, 49.27% and 33.54%, respectively. Piperine inhibited HeLa cell proliferation by increasing ROS production and nuclear cohesion, delaying wound healing and blocking the G2/M phase cell cycle. In addition, piperine induced apoptosis in HeLa cells by disrupting MMP and activating caspase-3 [156]. Recently, it has been shown that piperine has the potential to reverse drug resistance in human cervical cancer cells. Piperine (50 μmol/L) downregulated Mcl-1, phosphorylated AKT and cooperated with paclitaxel to increase paclitaxel-induced apoptosis of drug-resistant cervical cancer cells [157]. Apart from this, piperine inhibited Bcl-2 signaling pathway and increased Bid, caspase and PARP activities by blocking STAT3/NF-κB, thereby enhancing therapeutic and antiproliferative effects of mitomycin C (MMC) on drug-resistant HeLa/MMC cells [158]. Consequently, the combination of these two drugs can effectively reduce tumor growth in vivo and is a potential anticancer agent for treating human cervical cancer.

5.2. Matrine

Sophora flavescens is a kind of Chinese medicine and contains active ingredient matrine. Because of its wide range of pharmacological effects, it has been considered as anticancer drug in China. It has been used clinically to treat viral hepatitis, neuropathic pain, heart disease, skin diseases and other disorders for a long time [159,160]. S. flavescens inhibits cervical-cancer-cell proliferation and metastasis and induces apoptosis by inhibiting the AKT/mTOR signaling pathway; reducing the expression of Bcl-2, cyclin A and MMP2; blocking the G2/M phase cell cycle; and stimulating Bax and E-calmodulin. Matrine reduces vasodilator-stimulated phosphoprotein (VASP) phosphorylation by inhibiting protein kinase A (PKA) activity, thereby inhibiting HeLa cell adhesion and migration [161]. Besides, in mice injected with HeLa cells, intraperitoneal injection of matrine with 50 mg/kg/d for 21 d effectively inhibited the growth of cervical cancer xenografts with an inhibition rate of 58.33%. It also reduces the expression activity of extracellular matrix factors MMP2 and MMP9 by downregulating the p38 signal pathway, thus inhibiting the proliferation, metastasis and invasion of cervical cancer cells [162]. Matrine has the potential for development as therapeutic or adjuvant agent for human cervical cancer.

5.3. Berberine

Berberine, an isoquinoline alkaloid extracted from the herb, has various biological activities, such as antibacterial, anti-inflammatory, antidiabetic, anti-angiogenic and anticancer effects. It is widely used as an antibacterial agent in clinical practice [163,164]. Berberine inhibited the proliferation of HeLa and SiHa cells in a dose-dependent manner. It reduced cell viability and hTERT expression through a mitochondria-mediated pathway and activated caspase3, thereby inducing apoptosis. Moreover, berberine can selectively inhibit the transcription factor Activator Protein-1 (AP-1) activity and the transcription of HPV oncoproteins E6 and E7 and increase the expression of p53 and Rb to play an anti-cervical-cancer effect [165]. A study showed that berberine inhibited the invasion of SiHa cells by reducing the transcriptional activity of MMP2 and u-PA, mediated the downregulation of TGF-β1 expression and inhibited the angiogenic potential of SiHa cells. Notably, berberine (20 mg/kg/d for 21 d by oral gavage) was able to downregulate the NF-κB signaling pathway, increase E-calmodulin expression and selectively inhibit snail-1 gene expression, thereby reversing EMT and reducing cancer cell growth and lung metastasis [166]. It was also found that the IC50 value of berberine was 300 μmol/L after the 48-h treatment of HeLa cells. Berberine activated the death receptor pathway in HeLa cells by upregulating Fas, FasL, TNF-α and TRAF-1. Simultaneously, berberine induced apoptosis in HeLa cells by activating the mitochondrial pathway, releasing Cyt c and significantly increasing the ratio of Bax/Bcl-2, which blocked the cell cycle at the S phase. Moreover, berberine activated the MAPK pathway and increased p53 expression, suggesting that p53 may be a drug target in the treatment of cervical cancer [167]. In addition, N-Mannich bases of berberine (methoxy functional group, acid functionality and cyano group), berberine analogs, had significant cytotoxicity in cervical cancer HeLa and CaSki cells [168]. It suggested the potential antitumor effect of berberine and its analogs on cervical cancer cells.
The chemical structures of alkaloids are summarized in Figure 3. Table 3 summarizes their effects and mechanisms on cervical cancer.

6. Phenols

6.1. Curcumin

Curcumin is a phenol compound extracted from the rhizome of the perennial herb turmeric in Zingiberaceae, and has been commonly used as a spice and food coloring. Its safety and efficacy as a traditional herbal medicine have been shown in previous studies. It also exerts anticancer effects through various mechanisms, including inhibition of cell proliferation, invasion and metastasis, and epigenetic alterations [169,170]. Curcumin directly inhibited telomerase activity in HeLa, SiHa and CaSki cells in a dose-dependent manner, induced Cyt c release, activated AIF and inhibited cyclin D1 and Ras proteins, thereby downregulating downstream targets of the Ras/Raf pathway and ultimately triggering apoptosis via the mitochondrial pathway. Curcumin also mediates the anti-proliferative and anti-inflammatory activities of cervical cancer cells by inhibiting COX-2 and inducible nitric oxide synthase (iNOS) expression and activating caspase 3 and caspase 9 [171]. It has been reported that oral gavage of curcumin (500, 1000 and 1500 mg/kg/d) for 30 d in a mouse model injected with CaSki cells significantly inhibited tumor-cell growth (21.03%) and the expression of VEGF and EGFR, which inhibited angiogenesis [172]. In addition, curcumin inhibited the proliferation of HeLa, CaSki, C33A and ME180 cells in a time- and concentration-dependent manner; and it activated ER-resident unfolded protein response (UPR) sensors (PERK, IRE-1a and ATF6) and their downstream signaling proteins in cervical cancer cells, especially CHOP. Activation of CHOP elevated the Bax/Bcl-2 ratio and increased ROS production, which induced apoptosis in cervical cancer cells [173]. On the other hand, curcumin induced G2/M-phase cell-cycle arrest and triggered cell autophagy and apoptosis through downregulation of cyclin B1 and Cdc25. At the same time, it mediated cellular senescence by elevating the p53/p21 pathway [174]. Furthermore, curcumin treatment effectively reduced P-gp expression and increased the sensitivity of drug-resistant KB-V1 cells to vinblastine [175]. To enhance the bioavailability of curcumin, Warayuth et al. developed a delivery vehicle for oral curcumin. The cytotoxicity test showed that the IC50 values of drug-loaded (curcumin) micelles on HeLa, SiHa and C33a cell lines were 4.7, 3.6 and 12.2 times lower than those of free curcumin (21.17, 16.28 and 54.29 μmol/L). Compared with free curcumin, the number of micelles carrying curcumin is significantly increased by 6-fold. It is worth noting that the percentage of early apoptosis of HeLa, SiHa and C33a cells is increased by curcumin micelles to 30–55% [176]. These data provide a possible theoretical basis for clinical application of cervical cancer.

6.2. Ellagic Acid

Ellagic acid is a natural phenol compound primarily found in fruits and nuts, including raspberries, pomegranates and walnuts, with strong anti-proliferative, anti-metastatic and anti-angiogenic activities [177,178]. Ellagic acid inhibited SiHa cell proliferation in a dose-dependent manner with an IC50 value of 48.7 ± 2.5 μmol/L. Ellagic acid promoted SiHa cell apoptosis by upregulating p53, Bcl-2 and caspase 3/9; downregulating Bax; and inducing the G2-phase cell-cycle arrest [179]. Li et al. reported that ellagic acid induced G1-phase cell-cycle arrest by inhibiting the activation of janus kinase 2 (JAK2)/STAT3 pathway and downregulating the expression of Cyclin D1, Bcl-xl and Mcl-1, which in turn promoted apoptosis [180]. Moreover, ellagic acid inhibits the AKT/mTOR signaling pathway by increasing insulin-like growth factor-binding protein 7 (IGFBP7), thereby suppressing HeLa cell proliferation and invasion [181]. A recent study showed that treatment of HeLa cell with ellagic acid, curcumin and their combination for 72 h resulted in IC50 values of 19.47, 16.52 and 10.9 μmol/L, which revealed the synergistic anticancer effect of ellagic acid and curcumin. Their combination synergistically restored p53 protein expression, induced ROS formation and increased DNA damage, accompanied by increased expression of the downstream effectors p21 and Bax, which induced apoptosis in HeLa cells [182]. Ellagic acid has the potential to be an ideal drug for the treatment of cervical cancer.

6.3. Resveratrol

Resveratrol is a dietary polyphenol derived from grapes, berries, peanuts and other plant sources. It has several potential biological effects, including cancer prevention and treatment [183]. Resveratrol inhibited the proliferation of HeLa and CaSki cells in a dose- and time-dependent manner with IC50 values of 40.06 and 59.07 μmol/L at 72 h. Resveratrol inhibited HeLa and CaSki cell growth and development by inhibiting the expression of HPV oncoproteins E6 and E7; promoting p53, Bax and p16; and inducing cell-cycle arrest at the G1/S phase [184]. A study proved that resveratrol inhibited cervical cancer cell proliferation by upregulating suppressors of cytokine signaling 3 (SOCS3) and activated STAT3 (PIAS3) expression, blocking the DNA-binding activity of STAT3 and causing STAT3 inactivation [185]. Furthermore, resveratrol suppressed the degradation of EMT and extracellular matrix (ECM) by inhibiting STAT3 (Tyr705) phosphorylation [186]. He et al. found that resveratrol induced apoptosis in HeLa cells by downregulating hTERT mRNA and protein levels, which in turn inhibited telomerase activity [187]. Importantly, resveratrol also inhibited the expression of HIF-1α and VEGF in cervical cancer cells by blocking ERK1/2 and PI3K/Akt signaling pathways, thereby effectively reversing the angiogenic activity induced by HPV-16 E6 and E7 oncoproteins [188]. Another experiment proved that intraperitoneal injection of 10 mg/kg of resveratrol for 28 d significantly inhibited the growth of xenograft tumors in nude mice. Resveratrol significantly decreased the mRNA and protein expression of phospholipid scramblase 1 (PLSCR1), thereby inhibiting the proliferation and development of HeLa cells. In addition, resveratrol also inhibited migration and invasion of HeLa cells by suppressing NF-κB and AP-1-mediated MMP9 expression [189]. Resveratrol offers potential therapeutic value for the treatment of cervical cancer.
The chemical structures of phenols are summarized in Figure 4. Table 4 summarizes their mechanisms in cervical cancer. Figure 5 outlines the mechanism of natural products against cervical cancer

7. Conclusions

The high mortality and morbidity rates of cervical cancer remain a primary challenge for scientific research. Although conventional treatments, such as radiotherapy and chemotherapy, are effective, the five-year survival rate of cervical cancer patients is meager and often ends in failure. Moreover, chemotherapy is prone to drug resistance and toxic side effects, causing great suffering to patients. In recent years, plant-derived natural products have been considered as the most promising candidates for oncology therapies. In 2010, Bar-Sela et al. [190] reported a review article on curcumin as an anticancer agent, reviewing and analyzing data from clinical trials focusing on colon and pancreatic cancer and summarizing the main anticancer mechanisms of curcumin, but the article was published relatively early, ten years ago. Therefore, there is a need for further exploration and systematic review and sorting of the latest effective antitumor studies. In 2019, Hsiao et al. reviewed the effects and potential mechanisms of Chinese herbal medicines on cervical cancer and classified them according to crude extracts and compounds, outlining their effects on induction of apoptosis and inhibition of migration in vitro and in vivo, but the number of compounds was small [191]. In 2021, Park et al. reviewed the therapeutic effects of a variety of natural products on cervical cancer and their mechanisms, including apoptosis, anti-angiogenesis, anti-metastasis, drug resistance and microRNA modulation. However the inhibition of telomerase activity and enhancement of immune function were not included [192]. Our study reviewed 30 natural products that have antitumor effects on cervical cancer, which showed possible benefits in treating patients with cervical cancer through mechanisms, such as induction of apoptosis, inhibition of cell proliferation and angiogenesis (Figure 6). Therefore, the search for more readily available natural antitumor active ingredients and precursor drugs could provide effectively alternative or adjuvant treatment strategies for cancer patients.
Plant-derived natural products, a rich treasure trove of drug resources, have driven the development of novel inhibitors with their unique mode of action and molecular structural diversity. They play an important role in human cancer therapy with high safety and low side effects. However, some unresolved issues limit the translation of plant-derived natural products to clinical applications. First, some have poor water solubility, which seriously affects bio-availability. Second, potential toxicity and adverse effects of plant-derived natural products are lack of sufficient evaluation in vivo and clinical trial data. Again, further research on the mechanism of natural antitumor drugs is still the direction of future efforts. We believe that, with additional research, plant-derived natural products will be indeed applied in clinical medicine.

Funding

This work was supported by the National Natural Science Foundation of China (32060229, 31860258, U1803381).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, L.; Hua, K. Cervical Cancer: Emerging Immune Landscape and Treatment. OncoTargets Ther. 2020, 13, 8037–8047. [Google Scholar] [CrossRef] [PubMed]
  2. Arbyn, M.; Xu, L.; Simoens, C.; Martin-Hirsch, P.P. Prophylactic vaccination against human papillomaviruses to prevent cervical cancer and its precursors. Cochrane Database Syst. Rev. 2018, 5, CD009069. [Google Scholar] [CrossRef] [PubMed]
  3. José, F.X.B.; Quint, W.G.; Alemany, L.; Geraets, D.T.; Klaustermeier, J.E.; Lloveras, B.; Tous, S.; Felix, A.; Bravo, L.E.; Shin, H.-R.; et al. Human papillomavirus genotype attribution in invasive cervical cancer: A retrospective cross-sectional worldwide study. Lancet Oncol. 2010, 11, 1048–1056. [Google Scholar] [CrossRef]
  4. Rogers, L.; Siu, S.S.N.; Luesley, D.; Bryant, A.; O Dickinson, H. Radiotherapy and chemoradiation after surgery for early cervical cancer. Cochrane Database Syst. Rev. 2012, 5, CD007583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tambaro, R.; Scambia, G.; Di Maio, M.; Pisano, C.; Barletta, E.; Iaffaioli, V.R.; Pignata, S. The role of chemotherapy in locally advanced, metastatic and recurrent cervical cancer. Crit. Rev. Oncol. 2004, 52, 33–44. [Google Scholar] [CrossRef] [PubMed]
  6. Mann, J. Natural products in cancer chemotherapy: Past, present and future. Nat. Rev. Cancer 2002, 2, 143–148. [Google Scholar] [CrossRef] [PubMed]
  7. Aung, T.N.; Qu, Z.; Kortschak, R.D.; Adelson, D.L. Understanding the Effectiveness of Natural Compound Mixtures in Cancer through Their Molecular Mode of Action. Int. J. Mol. Sci. 2017, 18, 656. [Google Scholar] [CrossRef]
  8. Harvey, A.L. Natural products in drug discovery. Drug Discov. Today 2008, 13, 894–901. [Google Scholar] [CrossRef]
  9. Ouyang, L.; Luo, Y.; Tian, M.; Zhang, S.-Y.; Lu, R.; Wang, J.-H.; Kasimu, R.; Li, X. Plant natural products: From traditional compounds to new emerging drugs in cancer therapy. Cell Prolif. 2014, 47, 506–515. [Google Scholar] [CrossRef]
  10. Kikuchi, H.; Yuan, B.; Hu, X.; Okazaki, M. Chemopreventive and anticancer activity of flavonoids and its possibility for clinical use by combining with conventional chemotherapeutic agents. Am. J. Cancer Res. 2019, 9, 1517–1535. [Google Scholar]
  11. Nijveldt, R.J.; van Nood, E.; van Hoorn, D.E.; Boelens, P.G.; van Norren, K.; van Leeuwen, P.A. Flavonoids: A review of probable mechanisms of action and potential applications. Am. J. Clin. Nutr. 2001, 74, 418–425. [Google Scholar] [CrossRef] [PubMed]
  12. Alseekh, S.; de Souza, L.P.; Benina, M.; Fernie, A.R. The style and substance of plant flavonoid decoration; towards defining both structure and function. Phytochemistry 2020, 174, 112347. [Google Scholar] [CrossRef]
  13. Zhang, Q.; Yang, W.; Liu, J.; Liu, H.; Lv, Z.; Zhang, C.; Chen, D.; Jiao, Z. Identification of Six Flavonoids as Novel Cellular Antioxidants and Their Structure-Activity Relationship. Oxid. Med. Cell. Longev. 2020, 2020, 4150897. [Google Scholar] [CrossRef]
  14. Serafini, M.; Peluso, I.; Raguzzini, A. Flavonoids as anti-inflammatory agents. Proc. Nutr. Soc. 2010, 69, 273–278. [Google Scholar] [CrossRef] [Green Version]
  15. Khater, M.; Ravishankar, D.; Greco, F.; Osborn, H.M. Metal complexes of flavonoids: Their synthesis, characterization and enhanced antioxidant and anticancer activities. Future Med. Chem. 2019, 11, 2845–2867. [Google Scholar] [CrossRef]
  16. Ma, Y.; Zeng, M.; Sun, R.; Hu, M. Disposition of Flavonoids Impacts their Efficacy and Safety. Curr. Drug Metab. 2015, 15, 841–864. [Google Scholar] [CrossRef]
  17. Dinda, B.; Dinda, S.; Das Sharma, S.; Banik, R.; Chakraborty, A.; Dinda, M. Therapeutic potentials of baicalin and its aglycone, baicalein against inflammatory disorders. Eur. J. Med. Chem. 2017, 131, 68–80. [Google Scholar] [CrossRef]
  18. Liao, H.; Ye, J.; Gao, L.; Liu, Y. The main bioactive compounds of Scutellaria baicalensis Georgi. for alleviation of inflammatory cytokines: A comprehensive review. Biomed. Pharmacother. 2021, 133, 110917. [Google Scholar] [CrossRef]
  19. Cheng, C.-S.; Chen, J.; Tan, H.-Y.; Wang, N.; Chen, Z.; Feng, Y. Scutellaria baicalensis and Cancer Treatment: Recent Progress and Perspectives in Biomedical and Clinical Studies. Am. J. Chin. Med. 2018, 46, 25–54. [Google Scholar] [CrossRef]
  20. Xia, X.; Xia, J.; Yang, H.; Li, Y.; Liu, S.; Cao, Y.; Tang, L.; Yu, X. Baicalein blocked cervical carcinoma cell proliferation by targeting CCND1 via Wnt/β-catenin signaling pathway. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2729–2736. [Google Scholar] [CrossRef]
  21. Wu, X.; Yang, Z.; Dang, H.; Peng, H.; Dai, Z. Baicalein Inhibits the Proliferation of Cervical Cancer Cells Through the GSK3β-Dependent Pathway. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2018, 26, 645–653. [Google Scholar] [CrossRef]
  22. Li, J.; Ma, J.; Wang, K.S.; Mi, C.; Wang, Z.; Piao, L.X.; Xu, G.H.; Li, X.; Lee, J.J.; Jin, X. Baicalein inhibits TNF-α-induced NF-κB activation and expression of NF-κB-regulated target gene products. Oncol. Rep. 2016, 36, 2771–2776. [Google Scholar] [CrossRef] [Green Version]
  23. Peng, Y.; Guo, C.; Yang, Y.; Li, F.; Zhang, Y.; Jiang, B.; Li, Q. Baicalein induces apoptosis of human cervical cancer HeLa cells in vitro. Mol. Med. Rep. 2014, 11, 2129–2134. [Google Scholar] [CrossRef] [Green Version]
  24. Yu, X.; Lu, K.; Xia, J.; Mao, X. Baicalein induces HeLa cell growth inhibition by down-regulation of matrix metalloproteinases and activating extracellular signal-regulated kinase. Chin. J. Cell. Mol. Immunol. 2014, 30, 798–801. [Google Scholar]
  25. Hu, J.; Wang, R.; Liu, Y.; Zhou, J.; Shen, K.; Dai, Y. Baicalein Represses Cervical Cancer Cell Growth, Cell Cycle Progression and Promotes Apoptosis via Blocking AKT/mTOR Pathway by the Regulation of circHIAT1/miR-19a-3p Axis. OncoTargets Ther. 2021, 14, 905–916. [Google Scholar] [CrossRef]
  26. Yu, X.; Yang, Y.; Li, Y.; Cao, Y.; Tang, L.; Chen, F.; Xia, J. Baicalein inhibits cervical cancer progression via downregulating long noncoding RNA BDLNR and its downstream PI3 K/Akt pathway. Int. J. Biochem. Cell Biol. 2018, 94, 107–118. [Google Scholar] [CrossRef]
  27. He, L.; Lu, N.; Dai, Q.; Zhao, Y.; Zhao, L.; Wang, H.; Li, Z.; You, Q.; Guo, Q. Wogonin induced G1 cell cycle arrest by regulating Wnt/β-catenin signaling pathway and inactivating CDK8 in human colorectal cancer carcinoma cells. Toxicology 2013, 312, 36–47. [Google Scholar] [CrossRef]
  28. Kim, M.S.; Bak, Y.; Park, Y.S.; Lee, D.H.; Kim, J.H.; Kang, J.W.; Song, H.-H.; Oh, S.-R.; Yoon, D.Y. Wogonin induces apoptosis by suppressing E6 and E7 expressions and activating intrinsic signaling pathways in HPV-16 cervical cancer cells. Cell Biol. Toxicol. 2013, 29, 259–272. [Google Scholar] [CrossRef]
  29. Yang, L.; Zhang, H.-W.; Hu, R.; Yang, Y.; Qi, Q.; Lu, N.; Liu, W.; Chu, Y.-Y.; You, Q.-D.; Guo, Q.-L. Wogonin induces G1 phase arrest through inhibiting Cdk4 and cyclin D1 concomitant with an elevation in p21Cip1 in human cervical carcinoma HeLa cells. Biochem. Cell Biol. 2009, 87, 933–942. [Google Scholar] [CrossRef] [PubMed]
  30. He, F.; Wang, Q.; Zheng, X.-L.; Yan, J.-Q.; Yang, L.; Sun, H.; Hu, L.-N.; Lin, Y.; Wang, X. Wogonin potentiates cisplatin-induced cancer cell apoptosis through accumulation of intracellular reactive oxygen species. Oncol. Rep. 2012, 28, 601–605. [Google Scholar] [CrossRef]
  31. Madunić, J.; Madunic, I.V.; Gajski, G.; Popić, J.; Garaj-Vrhovac, V. Apigenin: A dietary flavonoid with diverse anticancer properties. Cancer Lett. 2018, 413, 11–22. [Google Scholar] [CrossRef]
  32. Wang, W.; Yue, R.-F.; Jin, Z.; He, L.-M.; Shen, R.; Du, D.; Tang, Y.-Z. Efficiency comparison of apigenin-7-O-glucoside and trolox in antioxidative stress and anti-inflammatory properties. J. Pharm. Pharmacol. 2020, 72, 1645–1656. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, M.-M.; Ma, R.-H.; Ni, Z.-J.; Thakur, K.; Cespedes-Acuña, C.L.; Jiang, L.; Wei, Z.-J. Apigenin 7-O-glucoside promotes cell apoptosis through the PTEN/PI3K/AKT pathway and inhibits cell migration in cervical cancer HeLa cells. Food Chem. Toxicol. 2020, 146, 111843. [Google Scholar] [CrossRef] [PubMed]
  34. Souza, R.P.; Bonfim-Mendonça, P.D.S.; Gimenes, F.; Ratti, B.A.; Kaplum, V.; Bruschi, M.L.; Nakamura, C.V.; Silva, S.O.; Maria-Engler, S.; Consolaro, M.E.L. Oxidative Stress Triggered by Apigenin Induces Apoptosis in a Comprehensive Panel of Human Cervical Cancer-Derived Cell Lines. Oxidative Med. Cell. Longev. 2017, 2017, 1512745. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, E.; Zhang, Y.; Fan, Z.; Cheng, L.; Han, S.; Che, H. Apigenin Inhibits Histamine-Induced Cervical Cancer Tumor Growth by Regulating Estrogen Receptor Expression. Molecules 2020, 25, 1960. [Google Scholar] [CrossRef] [Green Version]
  36. Seelinger, G.; Merfort, I.; Schempp, C.M. Anti-Oxidant, Anti-Inflammatory and Anti-Allergic Activities of Luteolin. Planta Med. 2008, 74, 1667–1677. [Google Scholar] [CrossRef]
  37. Lin, Y.; Shi, R.; Wang, X.; Shen, H.-M. Luteolin, a Flavonoid with Potential for Cancer Prevention and Therapy. Curr. Cancer Drug Targets 2008, 8, 634–646. [Google Scholar] [CrossRef]
  38. Krifa, M.; Alhosin, M.; Muller, C.D.; Gies, J.-P.; Chekir-Ghedira, L.; Ghedira, K.; Mély, Y.; Bronner, C.; Mousli, M. Limoniastrum guyonianum aqueous gall extract induces apoptosis in human cervical cancer cells involving p16INK4A re-expression related to UHRF1 and DNMT1 down-regulation. J. Exp. Clin. Cancer Res. 2013, 32, 30. [Google Scholar] [CrossRef] [Green Version]
  39. Shi, R.-X.; Ong, C.-N.; Shen, H.-M. Luteolin sensitizes tumor necrosis factor-α-induced apoptosis in human tumor cells. Oncogene 2004, 23, 7712–7721. [Google Scholar] [CrossRef] [Green Version]
  40. Shao, J.; Wang, C.; Li, L.; Liang, H.; Dai, J.; Ling, X.; Tang, H. Luteoloside Inhibits Proliferation and Promotes Intrinsic and Extrinsic Pathway-Mediated Apoptosis Involving MAPK and mTOR Signaling Pathways in Human Cervical Cancer Cells. Int. J. Mol. Sci. 2018, 19, 1664. [Google Scholar] [CrossRef] [Green Version]
  41. Horinaka, M.; Yoshida, T.; Shiraishi, T.; Nakata, S.; Wakada, M.; Nakanishi, R.; Nishino, H.; Sakai, T. The combination of TRAIL and luteolin enhances apoptosis in human cervical cancer HeLa cells. Biochem. Biophys. Res. Commun. 2005, 333, 833–838. [Google Scholar] [CrossRef]
  42. Bharti, S.; Rani, N.; Krishnamurthy, B.; Arya, D.S. Preclinical Evidence for the Pharmacological Actions of Naringin: A Review. Planta Med. 2014, 80, 437–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ghanbari-Movahed, M.; Jackson, G.; Farzaei, M.H.; Bishayee, A. A Systematic Review of the Preventive and Therapeutic Effects of Naringin Against Human Malignancies. Front. Pharmacol. 2021, 12, 639840. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, S.; Lin, R.; Hu, X.; Shi, Q.; Chen, H. Naringin induces endoplasmic reticulum stress-mediated apoptosis, inhibits β-catenin pathway and arrests cell cycle in cervical cancer cells. Acta Biochim. Pol. 2020, 67, 181–188. [Google Scholar] [CrossRef]
  45. Ramesh, E.; Alshatwi, A.A. Naringin induces death receptor and mitochondria-mediated apoptosis in human cervical cancer (SiHa) cells. Food Chem. Toxicol. 2013, 51, 97–105. [Google Scholar] [CrossRef]
  46. Zeng, L.; Zhen, Y.; Chen, Y.; Zou, L.; Zhang, Y.; Hu, F.; Feng, J.; Shen, J.; Wei, B. Naringin inhibits growth and induces apoptosis by a mechanism dependent on reduced activation of NF-κB/COX-2-caspase-1 pathway in HeLa cervical cancer cells. Int. J. Oncol. 2014, 45, 1929–1936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pandey, P.; Khan, F. A mechanistic review of the anticancer potential of hesperidin, a natural flavonoid from citrus fruits. Nutr. Res. 2021, 92, 21–31. [Google Scholar] [CrossRef] [PubMed]
  48. de Oliveira, J.M.P.F.; Santos, C.; Fernandes, E. Therapeutic potential of hesperidin and its aglycone hesperetin: Cell cycle regulation and apoptosis induction in cancer models. Phytomedicine 2020, 73, 152887. [Google Scholar] [CrossRef] [PubMed]
  49. Alshatwi, A.A.; Ramesh, E.; Periasamy, V.; Subash-Babu, P. The apoptotic effect of hesperetin on human cervical cancer cells is mediated through cell cycle arrest, death receptor, and mitochondrial pathways. Fundam. Clin. Pharmacol. 2013, 27, 581–592. [Google Scholar] [CrossRef]
  50. Wang, Y.; Yu, H.; Zhang, J.; Gao, J.; Ge, X.; Lou, G. Hesperidin inhibits HeLa cell proliferation through apoptosis mediated by endoplasmic reticulum stress pathways and cell cycle arrest. BMC Cancer 2015, 15, 682. [Google Scholar] [CrossRef] [Green Version]
  51. Abenavoli, L.; Capasso, R.; Milic, N.; Capasso, F. Milk thistle in liver diseases: Past, present, future. Phytother. Res. 2010, 24, 1423–1432. [Google Scholar] [CrossRef]
  52. Chu, C.; Li, D.; Zhang, S.; Ikejima, T.; Jia, Y.; Wang, D.; Xu, F. Role of silibinin in the management of diabetes mellitus and its complications. Arch. Pharmacal. Res. 2018, 41, 785–796. [Google Scholar] [CrossRef] [PubMed]
  53. You, Y.; He, Q.; Lu, H.; Zhou, X.; Chen, L.; Liu, H.; Lu, Z.; Liu, D.; Liu, Y.; Zuo, D.; et al. Silibinin Induces G2/M Cell Cycle Arrest by Activating Drp1-Dependent Mitochondrial Fission in Cervical Cancer. Front. Pharmacol. 2020, 11, 271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. García-Maceira, P.; Mateo, J.; Garc, P. Silibinin inhibits hypoxia-inducible factor-1α and mTOR/p70S6K/4E-BP1 signalling pathway in human cervical and hepatoma cancer cells: Implications for anticancer therapy. Oncogene 2008, 28, 313–324. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, Y.; Ge, Y.; Chen, Y.; Li, Q.; Chen, J.; Dong, Y.; Shi, W. Cellular and molecular mechanisms of silibinin induces cell-cycle arrest and apoptosis on HeLa cells. Cell Biochem. Funct. 2011, 30, 243–248. [Google Scholar] [CrossRef] [PubMed]
  56. Fan, S.; Li, L.; Chen, S.; Yü, Y.; Qi, M.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. Silibinin induced-autophagic and apoptotic death is associated with an increase in reactive oxygen and nitrogen species in HeLa cells. Free Radic. Res. 2011, 45, 1307–1324. [Google Scholar] [CrossRef] [PubMed]
  57. Fan, S.; Qi, M.; Yü, Y.; Li, L.; Yao, G.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. P53 activation plays a crucial role in silibinin induced ROS generation via PUMA and JNK. Free Radic. Res. 2012, 46, 310–319. [Google Scholar] [CrossRef] [PubMed]
  58. Fan, S.; Yu, Y.; Qi, M.; Sun, Z.; Li, L.; Yao, G.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. P53-mediated GSH depletion enhanced the cytotoxicity of NO in silibinin-treated human cervical carcinoma HeLa cells. Free Radic. Res. 2012, 46, 1082–1092. [Google Scholar] [CrossRef] [PubMed]
  59. Imran, M.; Rauf, A.; Shah, Z.A.; Saeed, F.; Imran, A.; Arshad, M.U.; Ahmad, B.; Bawazeer, S.; Atif, M.; Peters, D.G.; et al. Chemo-preventive and therapeutic effect of the dietary flavonoid kaempferol: A comprehensive review. Phytother. Res. 2019, 33, 263–275. [Google Scholar] [CrossRef]
  60. Tu, L.-Y.; Bai, H.-H.; Cai, J.-Y.; Deng, S.-P. The mechanism of kaempferol induced apoptosis and inhibited proliferation in human cervical cancer SiHa cell: From macro to nano. Scanning 2016, 38, 644–653. [Google Scholar] [CrossRef] [Green Version]
  61. Kashafi, E.; Moradzadeh, M.; Mohamadkhani, A.; Erfanian, S. Kaempferol increases apoptosis in human cervical cancer HeLa cells via PI3K/AKT and telomerase pathways. Biomed. Pharmacother. 2017, 89, 573–577. [Google Scholar] [CrossRef]
  62. Kumar, A.; Jaitak, V. Natural products as multidrug resistance modulators in cancer. Eur. J. Med. Chem. 2019, 176, 268–291. [Google Scholar] [CrossRef]
  63. Limtrakul, P.; Khantamat, O.; Pintha, K. Inhibition of P-Glycoprotein Function and Expression by Kaempferol and Quercetin. J. Chemother. 2005, 17, 86–95. [Google Scholar] [CrossRef]
  64. Andres, S.; Pevny, S.; Ziegenhagen, R.; Bakhiya, N.; Schäfer, B.; Hirsch-Ernst, K.I.; Lampen, A. Safety Aspects of the Use of Quercetin as a Dietary Supplement. Mol. Nutr. Food Res. 2018, 62, 1700447. [Google Scholar] [CrossRef]
  65. Mrkus, L.; Batinić, J.; Bjeliš, N.; Jakas, A. Synthesis and biological evaluation of quercetin and resveratrol peptidyl derivatives as potential anticancer and antioxidant agents. Amino Acids 2018, 51, 319–329. [Google Scholar] [CrossRef]
  66. Sundaram, M.K.; Raina, R.; Afroze, N.; Bajbouj, K.; Hamad, M.; Haque, S.; Hussain, A. Quercetin modulates signaling pathways and induces apoptosis in cervical cancer cells. Biosci. Rep. 2019, 39, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Vidya Priyadarsini, R.; Senthil Murugan, R.; Maitreyi, S.; Ramalingam, K.; Karunagaran, D.; Nagini, S. The flavonoid quercetin induces cell cycle arrest and mitochondria-mediated apoptosis in human cervical cancer (HeLa) cells through p53 induction and NF-κB inhibition. Eur. J. Pharmacol. 2010, 649, 84–91. [Google Scholar] [CrossRef] [PubMed]
  68. Clemente-Soto, A.F.; Salas-Vidal, E.; Milan-Pacheco, C.; Sánchez-Carranza, J.N.; Peralta-Zaragoza, O.; González-Maya, L. Quercetin induces G2 phase arrest and apoptosis with the activation of p53 in an E6 expression-independent manner in HPV-positive human cervical cancer-derived cells. Mol. Med. Rep. 2019, 19, 2097–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Bishayee, K.; Ghosh, S.; Mukherjee, A.; Sadhukhan, R.; Mondal, J.; Khuda-Bukhsh, A.R. Quercetin induces cytochrome-c release and ROS accumulation to promote apoptosis and arrest the cell cycle in G2/M, in cervical carcinoma: Signal cascade and drug-DNA interaction. Cell Prolif. 2013, 46, 153–163. [Google Scholar] [CrossRef] [PubMed]
  70. Xu, W.; Xie, S.; Chen, X.; Pan, S.; Qian, H.; Zhu, X. Effects of Quercetin on the Efficacy of Various Chemotherapeutic Drugs in Cervical Cancer Cells. Drug Des. Dev. Ther. 2021, 15, 577–588. [Google Scholar] [CrossRef] [PubMed]
  71. Jakubowicz-Gil, J.; Paduch, R.; Piersiak, T.; Głowniak, K.; Gawron, A.; Kandefer-Szerszeń, M. The effect of quercetin on pro-apoptotic activity of cisplatin in HeLa cells. Biochem. Pharmacol. 2005, 69, 1343–1350. [Google Scholar] [CrossRef]
  72. Adhami, V.M.; Syed, D.N.; Khan, N.; Mukhtar, H. Dietary flavonoid fisetin: A novel dual inhibitor of PI3K/Akt and mTOR for prostate cancer management. Biochem. Pharmacol. 2012, 84, 1277–1281. [Google Scholar] [CrossRef] [Green Version]
  73. Kashyap, D.; Sharma, A.; Sak, K.; Tuli, H.S.; Buttar, H.S.; Bishayee, A. Fisetin: A bioactive phytochemical with potential for cancer prevention and pharmacotherapy. Life Sci. 2018, 194, 75–87. [Google Scholar] [CrossRef]
  74. Chou, R.-H.; Hsieh, S.-C.; Yu, Y.-L.; Huang, M.-H.; Huang, Y.-C.; Hsieh, Y.-H. Fisetin inhibits migration and invasion of human cervical cancer cells by down-regulating urokinase plasminogen activator expression through suppressing the p38 MAPK-dependent NF-κB signaling pathway. PLoS ONE 2013, 8, e71983. [Google Scholar] [CrossRef] [Green Version]
  75. Liu, L.-Q.; Guo, W.; Yu, W.-D.; You, Z.-S. Fisetin Simultaneously Targets Apaf-1, ERK, and COX-2 Signaling Leading to Growth Inhibition and Apoptosis in Human Cervical Carcinoma Cell In Vitro. J. Sun Yat-Sen Univ. Med. Sci. 2013, 34, 75. [Google Scholar]
  76. Ying, T.-H.; Yang, S.-F.; Tsai, S.-J.; Hsieh, S.-C.; Huang, Y.-C.; Bau, D.-T.; Hsieh, Y.-H. Fisetin induces apoptosis in human cervical cancer HeLa cells through ERK1/2-mediated activation of caspase-8-/caspase-3-dependent pathway. Arch. Toxicol. 2012, 86, 263–273. [Google Scholar] [CrossRef]
  77. Lin, M.-T.; Lin, C.-L.; Lin, T.-Y.; Cheng, C.-W.; Yang, S.-F.; Lin, C.-L.; Wu, C.-C.; Hsieh, Y.-H.; Tsai, J.-P. Synergistic effect of fisetin combined with sorafenib in human cervical cancer HeLa cells through activation of death receptor-5 mediated caspase-8/caspase-3 and the mitochondria-dependent apoptotic pathway. Tumor Biol. 2015, 37, 6987–6996. [Google Scholar] [CrossRef] [PubMed]
  78. Musial, C.; Kuban-Jankowska, A.; Gorska-Ponikowska, M. Beneficial Properties of Green Tea Catechins. Int. J. Mol. Sci. 2020, 21, 1744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Singh, B.N.; Shankar, S.; Srivastava, R.K. Green tea catechin, epigallocatechin-3-gallate (EGCG): Mechanisms, perspectives and clinical applications. Biochem. Pharmacol. 2011, 82, 1807–1821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Al-Hazzani, A.A.; Alshatwi, A.A. Catechin hydrate inhibits proliferation and mediates apoptosis of SiHa human cervical cancer cells. Food Chem. Toxicol. 2011, 49, 3281–3286. [Google Scholar] [CrossRef]
  81. Singh, M.; Singh, R.; Bhui, K.; Tyagi, S.; Mahmood, Z.; Shukla, Y. Tea polyphenols induce apoptosis through mitochondrial pathway and by inhibiting nuclear factor-kappaB and Akt activation in human cervical cancer cells. Oncol. Res. 2011, 19, 245–257. [Google Scholar] [CrossRef]
  82. Panji, M.; Behmard, V.; Zare, Z.; Malekpour, M.; Nejadbiglari, H.; Yavari, S.; Dizaj, T.N.; Safaeian, A.; Maleki, N.; Abbasi, M.; et al. Suppressing effects of green tea extract and Epigallocatechin-3-gallate (EGCG) on TGF-β- induced Epithelial-to-mesenchymal transition via ROS/Smad signaling in human cervical cancer cells. Gene 2021, 794, 145774. [Google Scholar] [CrossRef]
  83. Zhang, Q.; Tang, X.; Lu, Q.; Zhang, Z.; Rao, J.; Le, A.D. Green tea extract and (−)-epigallocatechin-3-gallate inhibit hypoxia- and serum-induced HIF-1α protein accumulation and VEGF expression in human cervical carcinoma and hepatoma cells. Mol. Cancer Ther. 2006, 5, 1227–1238. [Google Scholar] [CrossRef] [Green Version]
  84. Kilic, U.; Sahin, K.; Tuzcu, M.; Basak, N.; Orhan, C.; Elibol-Can, B.; Kilic, E.; Sahin, F.; Kucuk, O. Enhancement of Cisplatin sensitivity in human cervical cancer: Epigallocatechin-3-gallate. Front. Nutr. 2014, 1, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Yokoyama, M.; Noguchi, M.; Nakao, Y.; Pater, A.; Iwasaka, T. The tea polyphenol, (-)-epigallocatechin gallate effects on growth, apoptosis, and telomerase activity in cervical cell lines. Gynecol. Oncol. 2004, 92, 197–204. [Google Scholar] [CrossRef] [PubMed]
  86. Dixon, R.A.; Ferreira, D. Genistein. Phytochemistry 2002, 60, 205–211. [Google Scholar] [CrossRef]
  87. Ghaemi, A.; Soleimanjahi, H.; Razeghi, S.; Gorji, A.; Tabarraei, A.; Moradi, A.; Alizadeh, A.; Vakili, M.A. Genistein induces a protective immunomodulatory effect in a mouse model of cervical cancer. Iran. J. Immunol. 2012, 9, 119–127. [Google Scholar] [PubMed]
  88. Sahin, K.; Tuzcu, M.; Basak, N.; Caglayan, B.; Kilic, U.; Sahin, F.; Kucuk, O. Sensitization of Cervical Cancer Cells to Cisplatin by Genistein: The Role of NFκB and Akt/mTOR Signaling Pathways. J. Oncol. 2012, 2012, 461562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Molton, S.A.; Todd, D.E.; Cook, S.J. Selective activation of the c-Jun N-terminal kinase (JNK) pathway fails to elicit Bax activation or apoptosis unless the phosphoinositide 3′-kinase (PI3K) pathway is inhibited. Oncogene 2003, 22, 4690–4701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Aguirre-Ghiso, J.A.; Estrada, Y.; Liu, D.; Ossowski, L. ERK(MAPK) activity as a determinant of tumor growth and dormancy; regulation by p38(SAPK). Cancer Res. 2003, 63, 1684–1695. [Google Scholar] [CrossRef]
  91. Kim, S.-H.; Kim, Y.-B.; Jeon, Y.-T.; Lee, S.-C.; Song, Y.-S. Genistein Inhibits Cell Growth by Modulating Various Mitogen-Activated Protein Kinases and AKT in Cervical Cancer Cells. Ann. N. Y. Acad. Sci. 2009, 1171, 495–500. [Google Scholar] [CrossRef]
  92. Hsieh, C.J.; Hsu, Y.L.; Huang, Y.F.; Tsai, E.M. Molecular Mechanisms of Anticancer Effects of Phytoestrogens in Breast Cancer. Curr. Protein Pept. Sci. 2018, 19, 323–332. [Google Scholar] [CrossRef]
  93. Chen, H.-H.; Chen, S.-P.; Zheng, Q.-L.; Nie, S.-P.; Li, W.-J.; Hu, X.-J.; Xie, M.-Y. Genistein Promotes Proliferation of Human Cervical Cancer Cells Through Estrogen Receptor-Mediated PI3K/Akt-NF-κB Pathway. J. Cancer 2018, 9, 288–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhu, J.; Zhang, C.; Qing, Y.; Cheng, Y.; Jiang, X.; Li, M.; Yang, Z.; Wang, N. Genistein induces apoptosis by stabilizing intracellular p53 protein through an APE1-mediated pathway. Free Radic. Biol. Med. 2015, 86, 209–218. [Google Scholar] [CrossRef] [PubMed]
  95. Bharti, R.; Chopra, B.S.; Raut, S.; Khatri, N. A Review on Traditional Uses, Pharmacology, and Phytochemistry. Front. Pharmacol. 2020, 11, 582506. [Google Scholar] [CrossRef] [PubMed]
  96. Gan, M.; Yin, X. Puerarin Induced in Mantle Cell Lymphoma Apoptosis and its Possible Mechanisms Involving Multi-signaling Pathway. Cell Biophys. 2015, 71, 367–373. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, L.-J.; Sun, L.; Wang, Z.-D. Effects of Pueraria Mirifica on immune and oxidative functions in U14 cervical cancer mice. Heilongjiang Med. Pharm. 2011, 34, 42–43. [Google Scholar]
  98. Hu, Y.-L.; Li, G.-L.; Lin, Z.-X. Effect and mechanism of Puerarin on the proliferation of human cervical cancer HeLa cells in vitro. Chin. J. Control Endem. Dis. 2017, 32, 1427–1428. [Google Scholar]
  99. Jia, L.; Hu, Y.; Yang, G.; Li, P. Puerarin suppresses cell growth and migration in HPV-positive cervical cancer cells by inhibiting the PI3K/mTOR signaling pathway. Exp. Ther. Med. 2019, 18, 543–549. [Google Scholar] [CrossRef] [Green Version]
  100. Jiang, D.; Rasul, A.; Batool, R.; Sarfraz, I.; Hussain, G.; Tahir, M.M.; Qin, T.; Selamoglu, Z.; Ali, M.; Li, J.; et al. Potential Anticancer Properties and Mechanisms of Action of Formononetin. BioMed Res. Int. 2019, 2019, 5854315. [Google Scholar] [CrossRef]
  101. Tay, K.-C.; Tan, L.T.-H.; Chan, C.K.; Hong, S.L.; Chan, K.G.; Yap, W.H.; Pusparajah, P.; Lee, L.-H.; Goh, B.-H. Formononetin: A Review of Its Anticancer Potentials and Mechanisms. Front. Pharmacol. 2019, 10, 820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Lo, Y.-L.; Wang, W. Formononetin potentiates epirubicin-induced apoptosis via ROS production in HeLa cells in vitro. Chem. Interact. 2013, 205, 188–197. [Google Scholar] [CrossRef] [PubMed]
  103. Guo, B.; Liao, C.; Liu, X.; Yi, J. Preliminary study on conjugation of formononetin with multiwalled carbon nanotubes for inducing apoptosis via ROS production in HeLa cells. Drug Des. Dev. Ther. 2018, 12, 2815–2826. [Google Scholar] [CrossRef] [Green Version]
  104. Zhang, Y.; Chen, C.; Zhang, J. Effects and significance of formononetin on expression levels of HIF-1α and VEGF in mouse cervical cancer tissue. Oncol. Lett. 2019, 18, 2248–2253. [Google Scholar] [CrossRef]
  105. Jin, Y.-M.; Xu, T.-M.; Zhao, Y.-H.; Wang, Y.-C.; Cui, M.-H. In vitro and in vivo anti-cancer activity of formononetin on human cervical cancer cell line HeLa. Tumor Biol. 2014, 35, 2279–2284. [Google Scholar] [CrossRef] [PubMed]
  106. Yousuf, B.; Gul, K.; Wani, A.A.; Singh, P. Health Benefits of Anthocyanins and Their Encapsulation for Potential Use in Food Systems: A Review. Crit. Rev. Food Sci. Nutr. 2016, 56, 2223–2230. [Google Scholar] [CrossRef]
  107. Zhang, J.; Wu, J.; Liu, F.; Tong, L.; Chen, Z.; Chen, J.; He, H.; Xu, R.; Ma, Y.; Huang, C. Neuroprotective effects of anthocyanins and its major component cyanidin-3-O-glucoside (C3G) in the central nervous system: An outlined review. Eur. J. Pharmacol. 2019, 858, 172500. [Google Scholar] [CrossRef]
  108. Li, X.; Zhao, J.; Yan, T.; Mu, J.; Lin, Y.; Chen, J.; Deng, H.; Meng, X. Cyanidin-3-O-glucoside and cisplatin inhibit proliferation and downregulate the PI3K/AKT/mTOR pathway in cervical cancer cells. J. Food Sci. 2021, 86, 2700–2712. [Google Scholar] [CrossRef]
  109. Li, X.; Mu, J.; Lin, Y.; Zhao, J.; Meng, X. Combination of cyanidin-3-O-glucoside and cisplatin induces oxidative stress and apoptosis in HeLa cells by reducing activity of endogenous antioxidants, increasing bax/bcl-2 mRNA expression ratio, and downregulating Nrf2 expression. J. Food Biochem. 2021, 45, e13806. [Google Scholar] [CrossRef]
  110. Zhou, Y.-X.; Gong, X.-H.; Zhang, H.; Peng, C. A review on the pharmacokinetics of paeoniflorin and its anti-inflammatory and immunomodulatory effects. Biomed. Pharmacother. 2020, 130, 110505. [Google Scholar] [CrossRef]
  111. Wang, J.-S.; Huang, Y.; Zhang, S.; Yin, H.-J.; Zhang, L.; Zhang, Y.-H.; Song, Y.-W.; Li, D.-D. A Protective Role of Paeoniflorin in Fluctuant Hyperglycemia-Induced Vascular Endothelial Injuries through Antioxidative and Anti-Inflammatory Effects and Reduction of PKC1. Oxid. Med. Cell Longev. 2019, 2019, 5647219. [Google Scholar] [CrossRef] [Green Version]
  112. Zhang, J.; Wang, F.; Wang, H.; Wang, Y.; Wu, Y.; Xu, H.; Su, C. Paeoniflorin inhibits proliferation of endometrial cancer cells via activating MAPK and NF-κB signaling pathways. Exp. Ther. Med. 2017, 14, 5445–5451. [Google Scholar] [CrossRef] [Green Version]
  113. Zheng, Y.-B.; Xiao, G.-C.; Tong, S.-L.; Ding, Y.; Wang, Q.-S.; Li, S.-B.; Hao, Z.-N. Paeoniflorin inhibits human gastric carcinoma cell proliferation through up-regulation of microRNA-124 and suppression of PI3K/Akt and STAT3 signaling. World J. Gastroenterol. 2015, 21, 7197–7207. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, L.; Zhang, S. Modulating Bcl-2 Family Proteins and Caspase-3 in Induction of Apoptosis by Paeoniflorin in Human Cervical Cancer Cells. Phytother. Res. 2011, 25, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  115. Suntres, Z.E.; Coccimiglio, J.; Alipour, M. The Bioactivity and Toxicological Actions of Carvacrol. Crit. Rev. Food Sci. Nutr. 2014, 55, 304–318. [Google Scholar] [CrossRef] [PubMed]
  116. Silva, E.R.; De Carvalho, F.O.; Teixeira, L.G.B.; Santos, N.G.L.; Felipe, F.A.; Santana, H.S.R.; Shanmugam, S.; Júnior, L.J.Q.; Araujo, A.A.d.S.; Nunes, P.S. Pharmacological Effects of Carvacrol in In vitro Studies: A Review. Curr. Pharm. Des. 2018, 24, 3454–3465. [Google Scholar] [CrossRef]
  117. Potočnjak, I.; Gobin, I.; Domitrović, R. Carvacrol induces cytotoxicity in human cervical cancer cells but causes cisplatin resistance: Involvement of MEK-ERK activation. Phytother. Res. 2018, 32, 1090–1097. [Google Scholar] [CrossRef]
  118. Gökalp, F. The effective natural compounds for inhibiting Cervical cancer. Med. Oncol. 2021, 38, 1–4. [Google Scholar] [CrossRef]
  119. Cui, L.; Su, X.-Z. Discovery, mechanisms of action and combination therapy of artemisinin. Expert Rev. Anti-Infect. Ther. 2009, 7, 999–1013. [Google Scholar] [CrossRef]
  120. Ji, P.; Huang, H.; Yuan, S.; Wang, L.; Wang, S.; Chen, Y.; Feng, N.; Veroniaina, H.; Wu, Z.; Wu, Z.; et al. ROS-Mediated Apoptosis and Anticancer Effect Achieved by Artesunate and Auxiliary Fe(II) Released from Ferriferous Oxide-Containing Recombinant Apoferritin. Adv. Health Mater. 2019, 8, e1900911. [Google Scholar] [CrossRef]
  121. Wong, Y.K.; Xu, C.; Kalesh, K.A.; Chengchao, X.; Lin, Q.; Wong, W.S.F.; Shen, H.-M.; Wang, J. Artemisinin as an anticancer drug: Recent advances in target profiling and mechanisms of action. Med. Res. Rev. 2017, 37, 1492–1517. [Google Scholar] [CrossRef]
  122. Chen, H.-H.; Zhou, H.-J.; Fang, X. Inhibition of human cancer cell line growth and human umbilical vein endothelial cell angiogenesis by artemisinin derivatives in vitro. Pharmacol. Res. 2003, 48, 231–236. [Google Scholar] [CrossRef]
  123. Hu, C.-J.; Zhou, L.; Cai, Y. Dihydroartemisinin induces apoptosis of cervical cancer cells via upregulation of RKIP and downregulation of bcl-2. Cancer Biol. Ther. 2013, 15, 279–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Tang, T.; Xia, Q.; Xi, M. Dihydroartemisinin and its anticancer activity against endometrial carcinoma and cervical cancer: Involvement of apoptosis, autophagy and transferrin receptor. Singap. Med. J. 2021, 62, 96–103. [Google Scholar] [CrossRef]
  125. Wang, L.; Li, J.; Shi, X.; Li, S.; Tang, P.M.K.; Li, Z.; Li, H.; Wei, C. Antimalarial Dihydroartemisinin triggers autophagy within HeLa cells of human cervical cancer through Bcl-2 phosphorylation at Ser70. Phytomedicine 2019, 52, 147–156. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, T.; Hu, Y.; Wang, T.; Cai, P. Dihydroartemisinin inhibits the viability of cervical cancer cells by upregulating caveolin 1 and mitochondrial carrier homolog 2: Involvement of p53 activation and NAD(P)H:quinone oxidoreductase 1 downregulation. Int. J. Mol. Med. 2017, 40, 21–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Thanaketpaisarn, O.; Waiwut, P.; Sakurai, H.; Saiki, I. Artesunate enhances TRAIL-induced apoptosis in human cervical carcinoma cells through inhibition of the NF-κB and PI3K/Akt signaling pathways. Int. J. Oncol. 2011, 39, 279–285. [Google Scholar]
  128. Luo, J.; Zhu, W.; Tang, Y.; Cao, H.; Zhou, Y.; Ji, R.; Zhou, X.; Lu, Z.; Yang, H.; Zhang, S.; et al. Artemisinin derivative artesunate induces radiosensitivity in cervical cancer cells in vitro and in vivo. Radiat. Oncol. 2014, 9, 84. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, L.-X.; Liu, Z.-N.; Ye, J.; Sha, M.; Qian, H.; Bu, X.-H.; Luan, Z.-Y.; Xu, X.-L.; Huang, A.-H.; Yuan, D.-L.; et al. Artesunate exerts an anti-immunosuppressive effect on cervical cancer by inhibiting PGE2production and Foxp3 expression. Cell Biol. Int. 2014, 38, 639–646. [Google Scholar] [CrossRef]
  130. Chen, H.-H.; Zhou, H.-J.; Wu, G.-D.; Lou, X.-E. Inhibitory Effects of Artesunate on Angiogenesis and on Expressions of Vascular Endothelial Growth Factor and VEGF Receptor KDR/flk-1. Pharmacology 2004, 71, 1–9. [Google Scholar] [CrossRef]
  131. Yen, J.-H.; Huang, S.-T.; Huang, H.-S.; Fong, Y.-C.; Wu, Y.-Y.; Chiang, J.-H.; Su, Y.-C. HGK-sestrin 2 signaling-mediated autophagy contributes to antitumor efficacy of Tanshinone IIA in human osteosarcoma cells. Cell Death Dis. 2018, 9, 1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Li, Z.M.; Xu, S.W.; Liu, P.Q. Salvia miltiorrhizaBurge (Danshen): A golden herbal medicine in cardiovascular therapeutics. Acta Pharmacol. Sin. 2018, 39, 802–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Munagala, R.; Aqil, F.; Jeyabalan, J.; Gupta, R.C. Tanshinone IIA inhibits viral oncogene expression leading to apoptosis and inhibition of cervical cancer. Cancer Lett. 2015, 356, 536–546. [Google Scholar] [CrossRef]
  134. Liu, Z.; Zhu, W.; Kong, X.; Chen, X.; Sun, X.; Zhang, W.; Zhang, R. Tanshinone IIA inhibits glucose metabolism leading to apoptosis in cervical cancer. Oncol. Rep. 2019, 42, 1893–1903. [Google Scholar] [CrossRef] [PubMed]
  135. Pan, T.-L.; Wang, P.-W.; Hung, Y.-C.; Huang, C.-H.; Rau, K.-M. Proteomic analysis reveals tanshinone IIA enhances apoptosis of advanced cervix carcinoma CaSki cells through mitochondria intrinsic and endoplasmic reticulum stress pathways. Proteomics 2013, 13, 3411–3423. [Google Scholar] [CrossRef]
  136. Qin, J.; Shi, H.; Xu, Y.; Zhao, F.; Wang, Q. Tanshinone IIA inhibits cervix carcinoma stem cells migration and invasion via inhibiting YAP transcriptional activity. Biomed. Pharmacother. 2018, 105, 758–765. [Google Scholar] [CrossRef]
  137. Cheng, W.; Huang, C.; Ma, W.; Tian, X.; Zhang, X. Recent Development of Oridonin Derivatives with Diverse Pharmacological Activities. Mini-Rev. Med. Chem. 2018, 19, 114–124. [Google Scholar] [CrossRef]
  138. Song, M.; Liu, X.; Liu, K.; Zhao, R.; Huang, H.; Shi, Y.; Zhang, M.; Zhou, S.; Xie, H.; Chen, H.; et al. Targeting AKT with Oridonin Inhibits Growth of Esophageal Squamous Cell Carcinoma In Vitro and Patient-Derived Xenografts In Vivo. Mol. Cancer Ther. 2018, 17, 1540–1553. [Google Scholar] [CrossRef] [Green Version]
  139. Zhang, Y.-N.; Zhang, Z.-Y.; Liu, H.-Y.; Wang, S.-H. Effects of oridonin on mitochondrial apoptotic pathway of human cervical carcinoma Hela cells. China J. Mod. Med. 2021, 31, 1–7. [Google Scholar]
  140. Hu, H.-Z.; Yang, Y.-B.; Xu, X.-D.; Shen, H.-W.; Shu, Y.-M.; Ren, Z.; Li, X.-M.; Shen, H.-M.; Zeng, H.-T. Oridonin induces apoptosis via PI3K/Akt pathway in cervical carcinoma HeLa cell line. Acta Pharmacol. Sin. 2007, 28, 1819–1826. [Google Scholar] [CrossRef] [PubMed]
  141. Zhang, C.-L.; Wu, L.-J.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. Oridonin induced A375-S2 cell apoptosis VIA BAX-regulated caspase pathway activation, dependent on the cytochromeC/CASPASE-9 apoptosome. J. Asian Nat. Prod. Res. 2004, 6, 127–138. [Google Scholar] [CrossRef]
  142. Zhang, Y.-H.; Wu, Y.-L.; Tashiro, S.-I.; Onodera, S.; Ikejima, T. Reactive oxygen species contribute to oridonin-induced apoptosis and autophagy in human cervical carcinoma HeLa cells. Acta Pharmacol. Sin. 2011, 32, 1266–1275. [Google Scholar] [CrossRef] [Green Version]
  143. Shen, H.; Hu, H.; Zeng, H.; Ke, P.; Yang, Y.; Li, X.; Ren, Z. Oridonin Induces Apoptosis Via PI3K/AKT Pathway in Cervical Car cinoma HeLa Cell Line. J. Sun Yat-sen Univ. Med. Sci. 2007, 28, 1819–1826. [Google Scholar]
  144. Li, X.; Chu, S.; Lin, M.; Gao, Y.; Liu, Y.; Yang, S.; Zhou, X.; Zhang, Y.; Hu, Y.; Wang, H.; et al. Anticancer property of ginsenoside Rh2 from ginseng. Eur. J. Med. Chem. 2020, 203, 112627. [Google Scholar] [CrossRef]
  145. Liu, Y.; Wang, J.; Qiao, J.; Liu, S.; Wang, S.; Zhao, D.; Bai, X.; Liu, M. Ginsenoside Rh2 inhibits HeLa cell energy metabolism and induces apoptosis by upregulating voltage-dependent anion channel 1. Int. J. Mol. Med. 2020, 46, 1695–1706. [Google Scholar] [CrossRef]
  146. Shi, X.; Yang, J.; Wei, G. Ginsenoside 20(S)-Rh2 exerts anti-cancer activity through the Akt/GSK3β signaling pathway in human cervical cancer cells. Mol. Med. Rep. 2018, 17, 4811–4816. [Google Scholar] [CrossRef] [PubMed]
  147. Li, Q.; Li, Y.; Wang, X.; Fang, X.; He, K.; Guo, X.; Zhan, Z.; Sun, C.; Jin, Y.-H. Co-treatment with ginsenoside Rh2 and betulinic acid synergistically induces apoptosis in human cancer cells in association with enhanced capsase-8 activation, bax translocation, and cytochrome c release. Mol. Carcinog. 2011, 50, 760–769. [Google Scholar] [CrossRef]
  148. Fulda, S. Betulinic acid: A natural product with anticancer activity. Mol. Nutr. Food Res. 2009, 53, 140–146. [Google Scholar] [CrossRef] [PubMed]
  149. Ríos, J.L.; Máñez, S. New Pharmacological Opportunities for Betulinic Acid. Planta Med. 2018, 84, 8–19. [Google Scholar] [CrossRef] [Green Version]
  150. Xu, T.; Pang, Q.; Wang, Y.; Yan, X. Betulinic acid induces apoptosis by regulating PI3K/Akt signaling and mitochondrial pathways in human cervical cancer cells. Int. J. Mol. Med. 2017, 40, 1669–1678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Kim, H.-J.; Cho, H.-S.; Ban, H.S.; Nakamura, H. Suppression of HIF-1α accumulation by betulinic acid through proteasome activation in hypoxic cervical cancer. Biochem. Biophys. Res. Commun. 2020, 523, 726–732. [Google Scholar] [CrossRef] [PubMed]
  152. Rzeski, W.; Stepulak, A.; Szymański, M.; Sifringer, M.; Kaczor, J.; Wejksza, K.; Zdzisińska, B.; Kandefer-Szerszeń, M. Betulinic acid decreases expression of bcl-2 and cyclin D1, inhibits proliferation, migration and induces apoptosis in cancer cells. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2006, 374, 11–20. [Google Scholar] [CrossRef] [PubMed]
  153. Sarek, J.; Klinot, J.; Dzubák, P.; Klinotová, E.; Nosková, V.; Krecek, V.; Korínková, G.; Thomson, J.O.; Janost’áková, A.; Wang, S.; et al. New lupane derived compounds with pro-apoptotic activity in cancer cells: Synthesis and structure-activity relationships. J. Med. Chem. 2003, 46, 5402–5415. [Google Scholar] [CrossRef] [PubMed]
  154. Aumeeruddy, M.Z.; Mahomoodally, M.F. Combating breast cancer using combination therapy with 3 phytochemicals: Piperine, sulforaphane, and thymoquinone. Cancer 2019, 125, 1600–1611. [Google Scholar] [CrossRef] [PubMed]
  155. Haq, I.U.; Imran, M.; Nadeem, M.; Tufail, T.; Gondal, T.A.; Mubarak, M.S. Piperine: A review of its biological effects. Phytother. Res. 2020, 35, 680–700. [Google Scholar] [CrossRef] [PubMed]
  156. Jafri, A.; Siddiqui, S.; Rais, J.; Ahmad, S.; Kumar, S.; Jafar, T.; Afzal, M.; Arshad, M. Induction of apoptosis by piperine in human cervical adenocarcinoma via ROS mediated mitochondrial pathway and caspase-3 activation. EXCLI J. 2019, 18, 154–164. [Google Scholar] [CrossRef] [PubMed]
  157. Xie, Z.; Wei, Y.; Xu, J.; Lei, J.; Yu, J. Alkaloids from Piper nigrum Synergistically Enhanced the Effect of Paclitaxel against Paclitaxel-Resistant Cervical Cancer Cells through the Downregulation of Mcl-1. J. Agric. Food Chem. 2019, 67, 5159–5168. [Google Scholar] [CrossRef] [PubMed]
  158. Han, S.-Z.; Liu, H.-X.; Yang, L.-Q.; Cui, L.-d.; Xu, Y. Piperine (PP) enhanced mitomycin-C (MMC) therapy of human cervical cancer through suppressing Bcl-2 signaling pathway via inactivating STAT3/NF-κB. Biomed. Pharmacother. 2017, 96, 1403–1410. [Google Scholar] [CrossRef]
  159. Zhang, H.; Chen, L.; Sun, X.; Yang, Q.; Wan, L.; Guo, C. Matrine: A Promising Natural Product with Various Pharmacological Activities. Front. Pharmacol. 2020, 11, 588. [Google Scholar] [CrossRef]
  160. You, L.; Yang, C.; Du, Y.; Liu, Y.; Chen, G.; Sai, N.; Dong, X.; Yin, X.; Ni, J. Matrine Exerts Hepatotoxic Effects via the ROS-Dependent Mitochondrial Apoptosis Pathway and Inhibition of Nrf2-Mediated Antioxidant Response. Oxid. Med. Cell. Longev. 2019, 2019, 1045345. [Google Scholar] [CrossRef] [Green Version]
  161. Zhang, L.; Wang, T.; Wen, X.; Wei, Y.; Peng, X.; Li, H.; Wei, L. Effect of matrine on HeLa cell adhesion and migration. Eur. J. Pharmacol. 2007, 563, 69–76. [Google Scholar] [CrossRef] [PubMed]
  162. Wu, X.; Zhou, J.; Cai, D.; Xiaoling, W. Matrine inhibits the metastatic properties of human cervical cancer cells via downregulating the p38 signaling pathway. Oncol. Rep. 2017, 38, 1312–1320. [Google Scholar] [CrossRef]
  163. Imenshahidi, M.; Hosseinzadeh, H. Berberis Vulgaris and Berberine: An Update Review. Phytother. Res. 2016, 30, 1745–1764. [Google Scholar] [CrossRef] [PubMed]
  164. Wang, K.; Feng, X.; Chai, L.; Cao, S.; Qiu, F. The metabolism of berberine and its contribution to the pharmacological effects. Drug Metab. Rev. 2017, 49, 139–157. [Google Scholar] [CrossRef] [PubMed]
  165. Mahata, S.; Bharti, A.C.; Shukla, S.; Tyagi, A.; A Husain, S.; Das, B.C. Berberine modulates AP-1 activity to suppress HPV transcription and downstream signaling to induce growth arrest and apoptosis in cervical cancer cells. Mol. Cancer 2011, 10, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Chu, S.-C.; Yu, C.-C.; Hsu, L.-S.; Chen, K.-S.; Su, M.-Y.; Chen, P.-N. Berberine Reverses Epithelial-to-Mesenchymal Transition and Inhibits Metastasis and Tumor-Induced Angiogenesis in Human Cervical Cancer Cells. Mol. Pharmacol. 2014, 86, 609–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Lu, B.; Hu, M.; Liu, K.; Peng, J. Cytotoxicity of berberine on human cervical carcinoma HeLa cells through mitochondria, death receptor and MAPK pathways, and in-silico drug-target prediction. Toxicol. Vitr. 2010, 24, 1482–1490. [Google Scholar] [CrossRef]
  168. Mistry, B.M.; Shin, H.-S.; Keum, Y.-S.; Kim, D.H.; Moon, S.H.; Kadam, A.A.; Shinde, S.K.; Patel, R.V. Synthesis and Evaluation of Antioxidant and Cytotoxicity of the N-Mannich Base of Berberine Bearing Benzothiazole Moieties. Anti-Cancer Agents Med. Chem. 2018, 17, 1652–1660. [Google Scholar] [CrossRef]
  169. Li, Y.; Zhang, T. Targeting cancer stem cells by curcumin and clinical applications. Cancer Lett. 2014, 346, 197–205. [Google Scholar] [CrossRef]
  170. Tomeh, M.A.; Hadianamrei, R.; Zhao, X. A Review of Curcumin and Its Derivatives as Anticancer Agents. Int. J. Mol. Sci. 2019, 20, 1033. [Google Scholar] [CrossRef] [Green Version]
  171. Singh, M.; Singh, N. Molecular mechanism of curcumin induced cytotoxicity in human cervical carcinoma cells. Mol. Cell. Biochem. 2009, 325, 107–119. [Google Scholar] [CrossRef]
  172. Yoysungnoen-Chintana, P.; Bhattarakosol, P.; Patumraj, S. Antitumor and Antiangiogenic Activities of Curcumin in Cervical Cancer Xenografts in Nude Mice. BioMed Res. Int. 2014, 2014, 817972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Kim, B.; Kim, H.S.; Jung, E.-J.; Lee, J.Y.; K Tsang, B.; Lim, J.M.; Song, Y.S. Curcumin induces ER stress-mediated apoptosis through selective generation of reactive oxygen species in cervical cancer cells. Mol. Carcinog. 2016, 55, 918–928. [Google Scholar] [CrossRef]
  174. Wang, T.; Wu, X.; Al Rudaisat, M.; Song, Y.; Cheng, H. Curcumin induces G2/M arrest and triggers autophagy, ROS generation and cell senescence in cervical cancer cells. J. Cancer 2020, 11, 6704–6715. [Google Scholar] [CrossRef] [PubMed]
  175. Anuchapreeda, S.; Leechanachai, P.; Smith, M.M.; Ambudkar, S.V.; Limtrakul, P.-N. Modulation of P-glycoprotein expression and function by curcumin in multidrug-resistant human KB cells. Biochem. Pharmacol. 2002, 64, 573–582. [Google Scholar] [CrossRef]
  176. Sajomsang, W.; Gonil, P.; Saesoo, S.; Ruktanonchai, U.R.; Srinuanchai, W.; Puttipipatkhachorn, S. Synthesis and anticervical cancer activity of novel pH responsive micelles for oral curcumin delivery. Int. J. Pharm. 2014, 477, 261–272. [Google Scholar] [CrossRef] [PubMed]
  177. Ceci, C.; Lacal, P.M.; Tentori, L.; De Martino, M.G.; Miano, R.; Graziani, G. Experimental Evidence of the Antitumor, Antimetastatic and Antiangiogenic Activity of Ellagic Acid. Nutrients 2018, 10, 1756. [Google Scholar] [CrossRef] [Green Version]
  178. Zhang, H.-M.; Zhao, L.; Li, H.; Xu, H.; Chen, W.-W.; Tao, L. Research progress on the anticarcinogenic actions and mechanisms of ellagic acid. Cancer Biol. Med. 2014, 11, 92–100. [Google Scholar] [CrossRef]
  179. Chen, W.; Shen, X.; Ma, L.; Chen, R.; Yuan, Q.; Zheng, Y.; Li, C.; Peng, G. Phenolic Compounds from Polygonum chinense Induce Growth Inhibition and Apoptosis of Cervical Cancer SiHa Cells. BioMed Res. Int. 2020, 2020, 8868508. [Google Scholar] [CrossRef]
  180. Li, L.; Na, C.; Tian, S.; Chen, J.; Ma, R.; Gao, Y.; Lou, G. Ellagic acid induces HeLa cell apoptosis via regulating signal transducer and activator of transcription 3 signaling. Exp. Ther. Med. 2018, 16, 29–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Guo, H.; Zhang, D.; Fu, Q. Inhibition of Cervical Cancer by Promoting IGFBP7 Expression Using Ellagic Acid from Pomegranate Peel. Med. Sci. Monit. 2016, 22, 4881–4886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Kumar, D.; Basu, S.; Parija, L.; Rout, D.; Manna, S.; Dandapat, J.; Debata, P.R. Curcumin and Ellagic acid synergistically induce ROS generation, DNA damage, p53 accumulation and apoptosis in HeLa cervical carcinoma cells. Biomed. Pharmacother. 2016, 81, 31–37. [Google Scholar] [CrossRef] [PubMed]
  183. Bishayee, A. Cancer Prevention and Treatment with Resveratrol: From Rodent Studies to Clinical Trials. Cancer Prev. Res. 2009, 2, 409–418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Sun, X.; Fu, P.; Xie, L.; Chai, S.; Xu, Q.; Zeng, L.; Wang, X.; Jiang, N.; Sang, M. Resveratrol inhibits the progression of cervical cancer by suppressing the transcription and expression of HPV E6 and E7 genes. Int. J. Mol. Med. 2020, 47, 335–345. [Google Scholar] [CrossRef]
  185. Zhang, P.; Yang, B.; Yao, Y.-Y.; Zhong, L.-X.; Chen, X.-Y.; Kong, Q.-Y.; Wu, M.-L.; Li, C.; Li, H.; Liu, J. PIAS3, SHP2 and SOCS3 Expression patterns in Cervical Cancers: Relevance with activation and resveratrol-caused inactivation of STAT3 signaling. Gynecol. Oncol. 2015, 139, 529–535. [Google Scholar] [CrossRef]
  186. Sun, X.; Xu, Q.; Zeng, L.; Xie, L.; Zhao, Q.; Xu, H.; Wang, X.; Jiang, N.; Fu, P.; Sang, M. Resveratrol suppresses the growth and metastatic potential of cervical cancer by inhibiting STAT3 Tyr705 phosphorylation. Cancer Med. 2020, 9, 8685–8700. [Google Scholar] [CrossRef] [PubMed]
  187. He, M.-J.; Yang, S.-Y.; Lu, J.; Han, S.-S.; Liu, R.; Lian, F.-Z. Effects of resveratrol on telomerase activity and gene expression in human cervical cancer HeLa cells. J. Hyg. Res. 2019, 48, 1001–1003. [Google Scholar]
  188. Tang, X.; Zhang, Q.; Nishitani, J.; Brown, J.; Shi, S.; Le, A.D. Overexpression of Human Papillomavirus Type 16 Oncoproteins Enhances Hypoxia-Inducible Factor 1α Protein Accumulation and Vascular Endothelial Growth Factor Expression in Human Cervical Carcinoma Cells. Clin. Cancer Res. 2007, 13, 2568–2576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Zhao, Y.; Yuan, X.; Li, X.; Zhang, Y. Resveratrol significantly inhibits the occurrence and development of cervical cancer by regulating phospholipid scramblase 1. J. Cell. Biochem. 2019, 120, 1527–1531. [Google Scholar] [CrossRef]
  190. Epelbaum, R.; Schaffer, M. Curcumin as an Anti-Cancer Agent: Review of the Gap Between Basic and Clinical Applications. Curr. Med. Chem. 2010, 17, 190–197. [Google Scholar] [CrossRef] [Green Version]
  191. Hsiao, Y.-H.; Lin, C.-W.; Wang, P.-H.; Hsin, M.-C.; Yang, S.-F. The Potential of Chinese Herbal Medicines in the Treatment of Cervical Cancer. Integr. Cancer Ther. 2019, 18, 1534735419861693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Park, S.-H.; Kim, M.; Lee, S.; Jung, W.; Kim, B. Therapeutic Potential of Natural Products in Treatment of Cervical Cancer: A Review. Nutrients 2021, 13, 154. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of several flavonoids against cervical cancer.
Figure 1. Chemical structure of several flavonoids against cervical cancer.
Biomolecules 11 01539 g001
Figure 2. Chemical structure of several terpenoids in anti-cervical-cancer.
Figure 2. Chemical structure of several terpenoids in anti-cervical-cancer.
Biomolecules 11 01539 g002
Figure 3. Chemical structure of several alkaloids in anti-cervical-cancer.
Figure 3. Chemical structure of several alkaloids in anti-cervical-cancer.
Biomolecules 11 01539 g003
Figure 4. Chemical structure of several phenols against cervical cancer.
Figure 4. Chemical structure of several phenols against cervical cancer.
Biomolecules 11 01539 g004
Figure 5. Mechanism of natural products against cervical cancer. Baicalein inhibits Cyclin D1 expression through downregulation of the Wnt/β-catenin pathway, which in turn inhibits cell proliferation. Genistein inhibits cell proliferation by activating the p38 MAPK pathway. It also induces apoptosis by inhibiting the PI3K/AKT-NF-κB pathway, which in turn induces apoptosis. Naringin promotes apoptosis by up-regulating the expression of Bax and caspase 3/8/9, which in turn promotes apoptosis. Kaempferol is pro-apoptotic by inhibiting PI3K/AKT and activating the p53 pathway. EGCG induces apoptosis by increasing ROS production and Bax/Bcl-2 expression. Berberine induces apoptosis by activating JNK and p53, while Curcumin promotes apoptosis by inducing Cyt c release, thereby downregulating the Ras/Raf pathway. Tanshinone IIA induces apoptosis through inhibition of AKT/mTOR and activation of the JNK pathway. Betulinic acid induces cell cycle arrest and apoptosis by activating caspase 9 and inhibiting the PI3K/AKT pathway. ↑: Upregulation, ↓: Downregulation.
Figure 5. Mechanism of natural products against cervical cancer. Baicalein inhibits Cyclin D1 expression through downregulation of the Wnt/β-catenin pathway, which in turn inhibits cell proliferation. Genistein inhibits cell proliferation by activating the p38 MAPK pathway. It also induces apoptosis by inhibiting the PI3K/AKT-NF-κB pathway, which in turn induces apoptosis. Naringin promotes apoptosis by up-regulating the expression of Bax and caspase 3/8/9, which in turn promotes apoptosis. Kaempferol is pro-apoptotic by inhibiting PI3K/AKT and activating the p53 pathway. EGCG induces apoptosis by increasing ROS production and Bax/Bcl-2 expression. Berberine induces apoptosis by activating JNK and p53, while Curcumin promotes apoptosis by inducing Cyt c release, thereby downregulating the Ras/Raf pathway. Tanshinone IIA induces apoptosis through inhibition of AKT/mTOR and activation of the JNK pathway. Betulinic acid induces cell cycle arrest and apoptosis by activating caspase 9 and inhibiting the PI3K/AKT pathway. ↑: Upregulation, ↓: Downregulation.
Biomolecules 11 01539 g005
Figure 6. Overview of signaling pathways against cervical cancer.
Figure 6. Overview of signaling pathways against cervical cancer.
Biomolecules 11 01539 g006
Table 1. Effect and mechanism of flavonoids on cervical cancer.
Table 1. Effect and mechanism of flavonoids on cervical cancer.
Chemical FamilyMolecule NameConcentrationCell LineMechanism
FlavonesBaicalein100 μmol/LHeLaIncreased: activation of caspase 3; PARP cleavage
Decreased: cIAP-1, cIAP-2, FLIP, Bcl-2, MMP2, MMP9, caspase 8, Fas, FasL, VEGF, COX-2, cyclin D1, IL-8, MCP1; Inhibit NF-κB, ERK1/2; G1 phase cell block
10 mg/kg for 28 d Xenografts of cervical cancer HeLa cells in female BALB/c miceDecreased: tumor growth
Wogonin0–100 μmol/LSiHa and CaSkiIncreased: Bax; activation of caspase 3 and 9; Cyt c release; PARP cleavage
Decreased: MMP; Bcl-2
40–160 μmol/LSiHa and CaSkiIncreased: Bax, p53, p21, p27, pRb
Decreased: E6, E7
Apigenin 7-glucosideIC50 = 47.26 μmol/LHeLaIncreased: ROS, Fas, FasL, TNF-α, TNF-r1, FADD, RADD; activation of caspase 3 and 9
Decreased: Bcl-2, Bcl-xl, Cyclin (A, D, E), CDK2/6, MMP; Inhibit PTEN/PI3K/AKT;
ApigeninIC50 = 10, 68, 76 and 40 μmol/LHeLa, SiHa, CaSki and C33AIncreased: ROS, H2O2
Decreased: MMP
100 mg/kg for 30 dXenografts of cervical cancer HeLa cells in female BALB/c miceDecreased: ERβ/ERα, tumor growth
LuteolinIC50 = 21.8 μmol/LHeLaIncreased: p16 INK4A, JNK
Decreased: UHRF1, DNMT1, A20, c-IAP1; G2/M phase cell block
FlavanonesNaringinIC50 = 750 μmol/LSiHaIncreased: p53, Bax, Fas, FADD; activation of caspase 3, 8 and 9
Hesperidin0–100 μmol/LHeLaIncreased: AIF, Cyt c release; activation of caspase 3
Decreased: MMP, cyclin D1, cyclin E1, CDK2; G0/G1 phase cell block
SilibininIC50 = 250,195 μmol/LHeLa and SiHaIncreased: activation of Drp1
Decreased: ATP, MMP, ROS, CDK1, Cdc25C, cyclinB1; G2/M phase cell block
FlavonolsKaempferolIC50 = 10.48 μmol/LHeLaIncreased: Bax, p53
Decreased: Bcl-2, hTERT; Inhibit PI3K/AKT
QuercetinIC50 = 110.38 ± 0.66 μmol/LHeLaIncreased: Bax, p53, ROS; activation of caspase 3; Cyt c release
Decreased: Bcl-2, AKT, MMP; G2/M cell block
0–200 μmol/LHeLa and SiHaIncreased: Bax, p53, p21
Decreased: E6/E6AP, G2 phase cell block
FisetinIC50 = 36.0 ± 0.5 μmol/LHeLaIncreased: ERK1/2, activation of caspase 3 and 8
2–4 mg/kg for 35 dXenografts of cervical cancer HeLa cells in male BALB/c miceDecreased: tumor growth, with inhibition rates of 82.65% and 92.62%
FlavanolsEpigallocatechin gallateIC50 = 20 μmol/LHeLaIncreased: Bax, p53, ROS; Cyt c release
Decreased: Bcl-2, COX-2; inhibition AKT and NF-κB
IsoflavonesGenisteinIC50 = 20 and 60 μmol/LHeLa and CaSkiIncreased: p38 MAPK, p38-JNK
Decreased: ERK1/2, AKT
20 mg/kg C57BL/6 cervical cancer cell mice modelDecreased: tumor growth
Puerarin12.5–50 μmol/LHeLaIncreased: Bax
Decreased: Wnt/β-catenin, p21, p53, Bcl-2
0–2000 μmol/LHeLaIncreased: Bax
Decreased: Inhibit PI3K/AKT/mTOR
500 mg/kg for 15 dcervical cancer cell line U14 mice modelsIncreased: IL-2, SOD
Decreased: tumor growth
Formononetin0–100 μmol/LHeLaIncreased: Bax, ROS; activation of caspase 3 and 9
Decreased: Bcl-2, MRP1 and MRP2, MMP
20 and 40 mg/kg for 35 dXenografts of cervical cancer HeLa cells in BALB/c nude miceDecreased: tumor growth
AnthocyaninsCyanidin 3-O-glucoside400 μmol/LHeLaIncreased: Bax, p53, TIMP-1
Decreased: Bcl-2, cyclin D1; G2/M phase cell block; PI3K/AKT/mTOR
Table 2. Effect of terpenoids on cervical cancer.
Table 2. Effect of terpenoids on cervical cancer.
Chemical FamilyMolecule NameConcentrationCell LineMechanism
MonoterpenoidsPaeoniflorinIC50 = 2459 μg/mLHeLaIncreased: Bax, Apaf-1; activation of caspase 3; Cyt c release
Decreased: Bcl-2
CarvacrolIC50 = 556 ± 39 μmol/LHeLaIncreased: LC3β-I/II; activation of caspase 9; PARP cleavage
Decreased: ERK
SesquiterpenoidsDihydroartemisininIC50 = 22.08 and 18.20 μmol/LHeLa and CaskiIncreased: RKIP
Decreased: Bcl-2
20 μmol/L for 15 dXenografts of cervical cancer HeLa or Caski cells in BALB/c miceDecreased: tumor growth, with inhibition rates of 70–80%
Artesunate60 μg/mLHeLaIncreased: AKT; activation of caspase 3; PARP cleavage
Decreased: survivin, XIAP, Bcl-xl; G2/M phase cell block; Inhibit NF-κB
100 mg/kg for 15 dXenografts of cervical cancer HeLa cells in BALB/c nude miceDecreased: tumor growth and inhibition of angiogenesis
DiterpenoidsTanshinone IIAIC50 = 6.97, 14.47, 5.51, and 9.89 μmol/LHeLa, SiHa, CaSki and C33AIncreased: Bax, PERK, IRE1, p38, JNK; activation of caspase 3 and 9; PARP cleavage; Cyt c and Ca2+ release
Decreased: Bcl-2
0–10 μmol/LHeLa, SiHa, CaSkiIncreased: p53, p21, p130, pRb
Decreased: E6, E7
40 mg/kg for 20 dCervical cancer cell line U14 mice modelsDecreased: metastasis and tumor growth with inhibition rates of 72.7%
OridoninC50 = 4.13 μmol/LHeLaIncreased: Bax; activation of caspase 3 and 9; Cyt c release
Decreased: Bcl-2, MMP
TriterpenoidsGinsenoside Rh2C50 = 35 μmol/LHeLaIncreased: Bax, ROS, VDAC1; Cyt c release
Decreased: MMP
Betulinic acidC50 = 30.42 ± 2.39 μmol/LHeLaIncreased: Bad, ROS, p27Kip and p21Waf1/Cip1; activation of caspase 9
Decreased: G0/G1 phase cell block; Inhibit PI3K/AKT
Table 3. Effect of alkaloids on cervical cancer.
Table 3. Effect of alkaloids on cervical cancer.
Chemical FamilyMolecule NameConcentrationCell LineMechanism
AlkaloidsPiperine0–200 μmol/LHeLaIncreased: ROS; activation of caspase 3
Decreased: MMP; G2/M phase cell block
Matrine50 mg/kg for 21 dXenografts of cervical cancer HeLa cells in BALB/c nude miceDecreased: p38, MMP2 and 9; tumor growth with inhibition rates of 58.33%
BerberineIC50 = 300 μmol/LHeLaIncreased: Bax, Fas, FasL, TNF-α, TRAF-1, p53, MAPK; DNA fragmentation; activation of caspase 3
Decreased: Bcl-2; S phase cell block
0–250 μg/mLHeLa and SiHaIncreased: p53, pRb
Decreased: E6, E7, AP-1, c-Jun, c-Fos
Table 4. Effect of phenols on cervical cancer.
Table 4. Effect of phenols on cervical cancer.
Chemical FamilyMolecule NameConcentrationCell LineMechanism
PhenolsCurcumin50 and 100 μmol/LHeLa, SiHa and CaSkiIncreased: AIF; activation of caspase 3 and 9; Cyt c release
Decreased: COX-2, iNOS, cyclin D1; Inhibit Ras/Raf
500, 1000 and 1500 mg/kgXenografts of cervical cancer CaSki cells in BALB/c miceDecreased: tumor growth, VEGF and EGFR
Ellagic acidC50 = 48.7 ± 2.5 μmol/LSiHaIncreased: Bcl-2, p53; activation of caspase 3 and 9
Decreased: Bax; G2 phase cell block
ResveratrolC50 = 40.06, 59.07 μmol/LHeLaIncreased: Bax, p53, p16
Decreased: G1/S phase cell block
5–250 μmol/LHeLa and CaSkiDecreased: Inhibit PI3K/AKT, ERK1/2, VEGF, HIF-1α accumulation
10 kg/mg for 28 dXenografts of cervical cancer HeLa cells in nude miceDecreased: tumor growth
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

He, M.; Xia, L.; Li, J. Potential Mechanisms of Plant-Derived Natural Products in the Treatment of Cervical Cancer. Biomolecules 2021, 11, 1539. https://doi.org/10.3390/biom11101539

AMA Style

He M, Xia L, Li J. Potential Mechanisms of Plant-Derived Natural Products in the Treatment of Cervical Cancer. Biomolecules. 2021; 11(10):1539. https://doi.org/10.3390/biom11101539

Chicago/Turabian Style

He, Meizhu, Lijie Xia, and Jinyao Li. 2021. "Potential Mechanisms of Plant-Derived Natural Products in the Treatment of Cervical Cancer" Biomolecules 11, no. 10: 1539. https://doi.org/10.3390/biom11101539

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