Next Article in Journal
Differences in Accuracy and Radiation Dose in Placement of Iliosacral Screws: Comparison between 3D and 2D Fluoroscopy
Previous Article in Journal
Postoperative Complications and Swallowing Function after Jejunal and Skin Flap Reconstruction for Hypopharyngeal Carcinoma—A Multicenter Retrospective Study
Previous Article in Special Issue
Detection of Human Papillomavirus DNA in Paired Peripheral Blood and Cervix Samples in Patients with Cervical Lesions and Healthy Individuals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Bioactives: Back to the Future in the Fight against Human Papillomavirus? A Narrative Review

1
Biotechnology Laboratory, Casaccia Research Center, Biotechnology and Agro-Industry Division, Italian National Agency for New Technologies, Energy and Sustainable Economic Development (ENEA), 00123 Rome, Italy
2
Department of Agriculture and Forest Sciences (DAFNE), University of Tuscia, 01100 Viterbo, Italy
3
HPV-Unit, Unità Operativa Semplice Dipartimentale (UOSD) Tumor Immunology and Immunotherapy, IRCCS Regina Elena National Cancer Institute, 00144 Rome, Italy
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2022, 11(5), 1465; https://doi.org/10.3390/jcm11051465
Submission received: 19 January 2022 / Revised: 25 February 2022 / Accepted: 4 March 2022 / Published: 7 March 2022
(This article belongs to the Special Issue Oncogenic Human Papillomaviruses)

Abstract

:
Human papillomavirus (HPV) still represents an important threat to health worldwide. Better therapy in terms of further improvement of outcomes and attenuation of related side-effects is desirable. The pharmaceutical industry has always targeted natural substances—phytochemicals in particular—to identify lead compounds to be clinically validated and industrially produced as antiviral and anticancer drugs. In the field of HPV, numerous naturally occurring bioactives and dietary phytochemicals have been investigated as potentially valuable in vitro and in vivo. Interference with several pathways and improvement of the efficacy of chemotherapeutic agents have been demonstrated. Notably, some clinical trials have been conducted. Despite being endowed with general safety, these natural substances are in urgent need of further assessment to foresee their clinical exploitation. This review summarizes the basic research efforts conducted so far in the study of anti-HPV properties of bio-actives with insights into their mechanisms of action and highlights the variety of their natural origin in order to provide comprehensive mapping throughout the different sources. The clinical studies available are reported, as well, to highlight the need of uniformity and consistency of studies in the future to select those natural compounds that may be suited to clinical application.

1. Introduction

Cancer is the main cause of death in wealthy countries with a constant increase in the number of cases around the globe. In 2020, 19.3 million new cancer cases were estimated (18.1 million, omitting nonmelanoma skin cancer) and almost 10 million cancer deaths (9.9 million, omitting nonmelanoma skin cancer) occurred [1]. About 15–20% of all human cancers are related to viral infections [2,3] and human papillomavirus (HPV)-related cancers account for 5% of all human cancers, with HPV being the most common sexually transmitted virus [4]. HPV tumors mainly originate from persistent infection with high-risk types HPV-16 and HPV-18 [5,6]. HPV infection is linked to anogenital cancers (cervical, vulvar, vaginal, anal, and penial) and head and neck cancers in both men and women [4,7]. Among HPV-associated tumors, cervical cancer still ranks fifth in terms of both incidence (6.5% of total female cases) and mortality (7.7% of total female deaths for cancer) of all cancers in the female population [7]. Information about transmission, cytological screening, and preventive vaccination are essential factors in limiting HPV infection and tumor spread in the global population [8], and thus in reducing the related incidence and mortality [9].
Despite preventive vaccination having been a big step in the direction of HPV cancer prevention [10], the incidence of HPV-related cancers continues to rise. Thus, HPV disease is still considered a serious public health concern [2]. Suggested for women up to 45 years of age, but mainly administered in the 9–15 years cohort, the impact of vaccination is visible mainly in the long run. The options available for therapy of HPV cancers (i.e., surgery, and chemo- and radio-therapy) are insufficient. Novel promising targeted agents with potential anti-HPV cancer effects comprise repurposed drugs (i.e., rimantadine) [11], in silico defined small molecules [12], and immunotherapeutics such as CAR-T cells, immune checkpoint inhibitors (ICIs), single-chain intrabodies, and therapeutic vaccines [13,14,15,16].
Taking advantage of the presence of a reliable tumor-associated antigen, represented by HPV E6 and E7 oncoproteins, the development of therapeutic vaccines able to provide better or comparable results to those treatments already in use which would be highly feasible and desirable. Many experimental efforts to develop therapeutic HPV vaccines have shown promising results, but no therapeutic vaccines are currently licensed for use [17]. Among therapeutic experimental HPV vaccines, ‘plant-inspired’ formulations have been developed, demonstrating the promise of plant-based natural compounds in making immunity more oriented to mount a response to the vaccine. As an example, a Nicotiana benthamiana crude extract containing HPV-16 E7 protein induced a cell-mediated immune response, protecting mice from experimental tumor challenge, in absence of concomitant adjuvant administration [18,19]. The same extract induced the maturation of human dendritic cells that were then able to induce human blood-derived T lymphocytes from healthy individuals to exert HPV-16 E7-specific cytotoxic activity [20]. A similar increased immune “visibility” was proven for an E7 protein-based vaccine produced through the use of the Chlamydomonas reinhardtii microalga [21]. The common feature of these plant-based vaccines is administration in the form of extracts and their intrinsic adjuvant activity [22].
Increasing attention is being given to the possible impact of purified natural substances and extracts on the development of HPV tumors [23,24,25]. The prophylactic inhibitory or reverting effect exerted by synthetic, natural, or biological chemicals against carcinogenesis commonly defines the chemoprevention ability of a substance. Chemoprevention may have importance in reducing, preventing, and inhibiting the development of different types of cancer, as nearly 33% of them can be avoided by lifestyle changes, including diet [26]. The ideal chemopreventive compound should be safe, highly efficient, and available at low production cost. Oral administration with diet would be a useful ‘plus’. Very often, indeed, the heavier budget items in the production of plant-based compounds include purification and downstream processing. This is the reason why population studies on plasma levels of phytochemicals and natural diet-related substances have been carried out in this field as well.
In addition to decreasing the risk of cancer development, natural compounds can influence one given routine chemo- and radio-therapeutic treatment by improving the sensitivity of tumors to chemo- and/or radio-therapeutics and, as a consequence, enhance related beneficial effects [27,28]. These effects are known as chemo- and radio-sensitization.
It is worth noting that many phytochemicals have antioxidant activities that, by eliminating free oxygen radicals, prevent cell DNA damage and, therefore, counteract one of the mechanisms of carcinogenesis [29]. Hence, antioxidant phytochemicals, beside reducing the risk of cancer development per se, can improve the cytotoxic effects of chemotherapeutics [22,23].
Despite research on plant-derived natural substances having been documented for years in the field of HPV, the real clinical applications of these substances still need further research and organic comprehension (for a list and details of published clinical trials grouped by formulation, see Supplementary Table S1). The current lack of longitudinal results, especially from in vivo and clinical points of view, does not allow us to foresee immediate implications for clinical practice. As a consequence, this review aims to highlight the efforts done in the study of anti-HPV properties of bioactives from different sources, and to provide a more comprehensive mapping of these properties. In fact, besides the well-known compounds of plant origin, the potential application of bacterial-derived and marine-bioactive compounds has been shown against HPV as well (Figure 1). In addition, the anti-HPV potential of some natural compounds has also been shown in silico by molecular modelling.
This review also offers a spotlight on plant (but also microbial) cells as tools to set up biotechnological factories of value-added specialized metabolites and proteins when naturally occurring sources are rare, endangered, only seasonally available, or when these molecules need qualitative or quantitative improvement.

2. Methodology

2.1. Criteria for Reference Selection

The search for reference material was carried out by consulting online databases for existing work relevant to the topic. PubMed, MEDLINE, ScienceDirect, Scopus, and Google Scholar (the latter in order to identify both published and unpublished information) were searched, applying publication date and other filters. In order to obtain a broader range search, also older publications were considered, when relevant. To minimize the risk of bias that may be associated with narrative reviews, our search was comprehensive in order to identify both research studies as well as reviews. This search was performed through an analysis of keywords throughout the specific index terms related to the articles and the words contained in the whole text of the retrieved papers.

2.2. Research Question

Search query was performed using the following keywords: HPV disease, natural compounds, bioactives, phytochemicals, combination treatments, therapy, infection, name of compounds, and class of compounds.

2.3. Study Selection

The results of the database searches were transferred to a Mendeley library and then were de-duplicated. Irrelevant records were removed before identifying records that met the inclusion criteria. Inclusion criteria involved reviews and/or research papers on in vitro/in vivo/clinical randomized and non-randomized studies published in indexed and peer-reviewed scientific journals. Exclusion criteria involved abstracts or unpublished works with insufficient or invalid information established on the basis of the expertise of co-authors.

2.4. Charting the Data

Data were synthesized and interpreted using a narrative/descriptive approach. All published papers matching the above-mentioned criteria have been quoted in the manuscript and in the bibliography section, after a critical revision of the exposure experimental methods.

3. Results

Results obtained from literature scan following the Methods described in Section 2 are provided in Table 1.
Results were summarized using an analytic approach to guide the narrative account of existing literature. Due to the huge availability of entities tested against HPV, for ease of consultation, compounds and formulations tested in the different studies have been reported in alphabetical order by chemical family and type of formulation (i.e., purified entities alone or in combination, extracts, mixtures), describing natural source, type of study through which they have been investigated (i.e., in vitro, in vivo, case study, clinical trial) and related outcomes. For completeness and ease of consultation, a summary of all the phytocompounds (i.e., the vast majority of natural chemical entities tested against HPV) reported in the present review, together with further phytocompounds cited by additional literature, is reported in Supplementary Table S2.

4. Purified Phytochemicals

Since HPV-18-positive HeLa cells were discovered, and before the HPV etiology was established, phytochemicals and traditional ethnobotanical compounds have been tested against cervical cancer [30,31,32]
Plant cells have the highest diversity of metabolite production among live cells [33,34]. Indeed, plants necessitate specialized metabolites for their development and survival. Phytochemicals, as purified and characterized entities (i.e., the so-called specialized metabolites) (Figure 2), will be reviewed in this section (for a complete list, see Supplementary Table S2).

4.1. Polyphenols

Numerous in vivo and/or in vitro studies have mainly focused on the anticancer effects of purified plant-derived metabolites and, specially, on polyphenols, most of which are flavonoids [23,24,25]. This group of phenolic compounds, which is easily encountered in daily nutrition, is diverse and is composed of sub-groups of compounds. Polyphenols will be presented in their outcomes through listing by their four major classes, starting from the initial biosynthetic steps of the pathway: phenolic acids, flavonoids, stilbenes, lignans, and curcuminoids. The effects of these compounds have been demonstrated to rely on the activation or inhibition of several pathways and, in some cases, their action has been demonstrated to be exerted directly on E6 and E7 oncoproteins or their targets.

4.1.1. Phenolic Acids

Caffeic acid (hydroxycinnamic acid) from Ocimum gratissimum (1 mM and 10 mM), has been demonstrated to significantly reduce the proliferation of HeLa cells by apoptosis through decreased levels of caspase-3 and Bcl-2, and upregulation of caspase-3 and p53 in a concentration-dependent manner, triggered 24–48 h after exposure. This indicates that caffeic acid induces apoptosis via the mitochondrial apoptotic pathway [35].
Caffeic acid has shown induction of cervical cancer cell death in vitro by generating reactive oxygen species (ROS), arresting cycle in S and G2/M phases, and by binding to histone deacetylase (HDAC) followed by apoptosis induction, as well [36].
Ellagic acid is found in strawberries, raspberries, walnuts, and pomegranate. In particular, ellagic acid from pomegranate peel has been shown to induce inhibition of HeLa cells by blocking the AKT/mTOR signaling pathway through enhanced expression of insulin-like growth factor-binding protein 7 (IGFBP7). The invasion ability and apoptosis rate of treated HeLa cells was demonstrated, indeed, to be significantly lower compared to controls [37]. However, the authors reported that the ellagic acid use has some limitations due to intrinsic characteristics such as poor absorption and bioavailability and lack of information regarding its safety in the long term.
Ferulic acid (hydroxycinnamic acid) has been shown to significantly inhibit proliferation and invasion of HeLa and CaSki cells through the reduction in MMP-9 mRNA expression and by an induced arrest in G0/G1 phase, also inhibiting autophagy in both cancer cell types in a dose-dependent mechanism (p < 0.05) [38].
In addition, ferulic acid has been demonstrated to increase the radiosensitivity of HeLa and Me180 cells by potentiation of apoptosis through a ROS-dependent mechanism [39].
Gallic acid is a hydroxybenzoic acid found in grapes, blackberries, raspberries, pomegranate, tea, vanilla, and in the leaves of many plants. This compound has demonstrated induction of apoptosis and necrosis in HeLa cells accompanied by ROS increase and glutathione (GSH) depletion [40].
Gallic acid-related downregulation of epidermal growth factor (EGFR), phosphorylated serine/threonine kinase 1 (Akt/p-Akt), and phosphorylated extracellular signal-regulated kinase (Erk/p-Erk) signaling pathways has been proposed as a possible tool to induce suppression of cancer progression in vivo as well [41].

4.1.2. Flavonoids

Flavanones

Hesperetin is mainly found in Citrus species and has been demonstrated to induce apoptosis in SiHa cells by upregulating caspases, p53, Bax, and Fas death receptor and its adaptor protein [42,43].
Naringenin, mainly accumulated in grapefruit and citrus fruits, has been shown to inhibit the proliferation of SiHa cells by cell cycle arrest at the G2/M phase and apoptosis triggering [44]. Due to its intrinsic instability, studies have been conducted on naringenin-loaded nanoparticles. The use of these nanoparticles has shown better outcomes over free naringenin in HeLa cells [45] where they were able to determine dose-dependent cytotoxicity, apoptosis, decreased intracellular glutathione levels, alterations in mitochondrial membrane potential, increased intracellular levels of ROS, and lipid peroxidation. Naringenin was found to induce apoptosis of HeLa cells as well by reducing expression of NF-κB p65 subunit, COX-2, and caspase-1. In addition, naringenin treatment induced p21WAF1 expression and downregulated expression of cyclins and cyclin-dependent kinases (CDKs). Moreover, treatment with naringenin induced phosphorylation of extracellular signal-regulated kinase (ERK), p38 mitogen-activated protein kinase, and c-Jun N-terminal kinase. Despite these results, naringenin application in cancer is limited due to its suboptimal pharmacokinetics and poor bioavailability at the tumor sites [46,47,48,49].

Flavones

Apigenin, the most active flavone in plants, is found in fruits and vegetables and it is of particular interest due to its lower intrinsic toxicity [50,51,52,53,54]. It has been extensively tested in vitro.
CaSki, HeLa, and C-33A growth was inhibited by G1 phase arrest, p53-dependent apoptosis, and upregulation of p21WAF1 expression, Fas/APO-1 death receptor, and caspase-3. Apigenin has been demonstrated to downregulate the antiapoptotic factor Bcl-2 [55].
Interference with HeLa cell gap junctions has been considered responsible for the apigenin-induced attenuation of their motility and translocation in vitro, revealing a possible tool to reduce invasive potential in vivo through this phytocompound [56]. Apigenin has also been demonstrated to sensitize HeLa and SiHa cells to paclitaxel through apoptosis induction triggered by intracellular ROS accumulation [57]. Upon apigenin administration, TC-1 tumor cells (primary mice lung cells transformed with HPV-16 E6, HPV-16 E7, and ras oncogenes) have been demonstrated to become more susceptible to lysis by E7-specific cytotoxic CD8+ T cells and to apoptosis in vitro in a dose-dependent fashion.
In the context of a combinatorial in vivo pre-clinical study, administration of apigenin in association with an E7-HSP70 DNA vaccine has been shown to increase the levels of primary and memory E7-specific CD8+ T cells, leading to therapeutic anti-tumor effects against E7-expressing tumors in TC-1 tumor-bearing mice [58].
Jaceosidin, common in Compositae, has been shown to block E6 and E7 oncoproteins activity by impeding binding to p53 and pRb in SiHa and CaSki cells [59]. In addition, jaceosidin increased cleavage of caspase-9 and caspase-3 in oral squamous carcinoma cells and downregulated Akt phosphorylation. Jaceosidin demonstrated no effects on HaCaT normal epithelial cell viability, indicating a possible selective chemotherapeutic potential [60].
Luteolin is most often found in perilla, broccoli, celery, parsley, thyme, olive oil, carrots, oregano, oranges, and peppermint. It has been demonstrated to induce apoptosis in HeLa cells and to be an E6 inhibitor through binding to a hydrophobic pocket at the interface between E6 and E6AP [61]. In particular, luteolin enhanced the expression of death receptor downstream factors—such as Fas/FasL, DR5/TRAIL, and FADD in HeLa cells—and activated caspase cascades, especially caspase-3 and caspase-8 in a dose-dependent manner.
Wogonin from Scutellaria baicalensis has been found to trigger apoptosis through suppression of E6 and E7 and increased p53 and pRb in SiHa and CaSki cells [62]. Wogonin demonstrated A549 and HeLa cell sensitization to cisplatin by ROS generation as well [63].

Flavonols

Epigallocatechin-3-gallate (EGCG) is considered the most bioactive agent of green tea from Camellia sinensis, and it has shown antiproliferative, antiangiogenic, anti-metastatic, and proapoptotic effects by affecting a wide array of signal transduction pathways including JAK/STAT, MAPK, PI3K/AKT, Wnt, and Notch [64]. The inhibition of HPV E6/E7 expression and estrogen receptor α and aromatase is supposed to be the main mechanism [65] of EGCG-mediated apoptosis in HPV cancer cells [66]. It has been demonstrated that the cell type involved in cervical cancer development can affect the result of the treatment with this compound. Indeed, when treated with 50 µg/mL of EGCG, cell growth was inhibited by 100% in squamous cell carcinoma cell line Me180, while growth inhibition was less evident in adenocarcinoma cell line HeLa (85.6%) [67]. A genome-wide study has demonstrated that EGCG also affected DNA methylation in the oral squamous cell carcinoma CAL 27 cell line [68]. EGCG was able to suppress HeLa, CaSki, and C-33A cells by regulating the expression of miRNAs [69].
A combination of tea polyphenols with EGCG sensitized HeLa cells to bleomycin by enhanced apoptosis induction triggered through caspase-3, -8, -9, and by upregulation of p53 and Bcl-2 [70].
In addition, a study associating EGCG with E7-based DNA vaccination demonstrated enhanced tumor-specific T-cell response and antitumor effects relative to either DNA vaccine or EGCG alone in tumor-bearing mice. Combined DNA vaccination and oral EGCG treatment provided long-term antitumor protection in healed mice that were afterward able to specifically reject experimental E7-expressing tumors, showing antigen-specific immune responses [71].
Fisetin, found mainly in strawberries, apples, onions, cucumbers, and khaki, has shown HeLa anti-proliferation properties through apoptosis mediated by inhibition of ERK1/2 and activation of caspase-8 and-3 [72]. Moreover, treatment of HeLa cells with fisetin induced a sustained activation of the phosphorylation of ERK1/2, and inhibition of ERK1/2 by PD98059 (MEK1/2 inhibitor).
Kaempferol, found in many fruits and vegetables, has been shown to block HeLa cell proliferation by G2/M phase cell cycle block, decreasing cyclin B1 and CDK1, inhibition of NF-kB nuclear translocation, upregulation of Bax, and downregulation of Bcl-2 [73]. In addition, kaempferol inhibited the proliferation of SiHa cells in a time- and dose-dependent manner—inducing cell apoptosis, mitochondrial membrane potential disruption, and intracellular free calcium elevation [74].
Quercetin, which is found in many fruits, vegetables, and grains, has been demonstrated to induce apoptosis through G2/M phase block, inhibition of anti-apoptotic AKT, and Bcl-2 expression in HeLa cells [75]. Quercetin sensitized HeLa cells to apoptosis upon paclitaxel treatment by an enhancement of intracellular ROS level [76]. Quercetin has also been shown to increase the radiosensitivity of HeLa cells in vitro and of DLD1 human colorectal cancer xenografts in vivo by enhancing apoptosis through a time- and dose-dependent mechanism determining ROS modulation and downregulation of E6 and E7 expression [77].

Isoflavonoids

Daidzein is represented in legumes (mainly in soybeans) and is the most frequent isoflavone in nature. At 6.25–100 µmol/L concentrations, it blocked the growth of HeLa cells affecting the cell cycle and downregulation of telomerase catalytic subunit mRNA, especially at the lowest concentrations [78].
Formononetin, accumulated by several plants and herbs such as red clover, has been demonstrated to sensitize cervical cancer (HeLa) cells in vitro to epirubicin via ROS production [79].
Genistein, mostly found in legumes, has been shown to block the growth of CaSki and HeLa cells in vitro through cell cycle arrest and activation of the AKT gene [80,81,82]. However, the safety of genistein demonstrated in this work was judged controversial as the concentration tested was higher than that reported to have anti-cancer effects (i.e., as tested against MCF-7 cells) in literature. The effect of genistein on cancer cells is still controversial, since a study reported a stimulatory effect of genistein and quercetin on MCF-7 cells [83,84].

Anthocyanins

These natural pigments are mainly found in black currant, eggplant, cherry, cranberries, blueberries, and black rice and are known for their ROS scavenging ability. In the case of HPV, cyanidin 3-glucoside from black rice has been demonstrated to inhibit proliferation of HeLa cells in a dose- and time-dependent manner by apoptosis induced by Bax/Bcl-2 [85].

Flavolignans

Silibinin has shown cell cycle modulation (G2 arrest) and dose- and time-dependent apoptosis induction abilities via the activation of the dynamin-related protein 1 in HeLa cells [86,87].

Lignans

Methylenedioxy lignan showed inhibition of HeLa cell proliferation through apoptosis and inhibition of telomerase [88]. In addition, methylenedioxy lignan inhibits Bcl2 and activates caspase-3 and caspase-8.
Nor-dihydro-guaiaretic acid from Larrea tridentata downregulated HPV E6 and E7 expression and promoted apoptosis in SiHa cells through suppression of HPV transcription. In addition, nor-dihydro-guaiaretic acid induced cell cycle arrest at the G1 phase and promoted acetylation of histone H3 and p21 gene-associated chromatin [89].
Sesamin, a major lignan from sesame oil, has been demonstrated to inhibit the proliferation and migration of HeLa cells and to induce ER stress-mediated apoptosis through the IRE1alpha/JNK pathway. Furthermore, it was able to activate autophagy [90].

4.1.3. Stilbenes

Resveratrol accumulates in grapes, blueberries, raspberries, and peanuts. In HeLa cells, it reduced the generation of ROS and suppressed invasion and migration [91]. In CaSki cervical cancer cells it induced a decreased expression of metalloproteinases. Inhibition of AKT and ERK1/2, decrease in the angiogenic activity, destabilization of lysosomes, induction of autophagy, and increased cytosol translocation are other activities that resveratrol was found to exert in the above-mentioned cervical cancer cells [92]. Resveratrol has shown increased radiosensitivity and potentiation of apoptosis of HeLa and SiHa cells through dose-dependent increased cytotoxicity and cell cycle alteration [93]. However, doubts have arisen regarding the characteristics of the patients enrolled, the dose used and duration of supplementation in the above-mentioned study, as well as resveratrol toxicity at high doses [94].

4.1.4. Curcuminoids

Curcumin is a hydrophobic polyphenol derived from the rhizome of Curcuma longa. It shows several pharmacological properties.
The effect of curcumin on tumor HeLa, SiHa, and C-33A cells has been found to involve the induction of apoptosis and the downregulation of COX-2. HPV-specific effects have been demonstrated for HeLa and SiHa cells that were able to downregulate the viral oncogenes E6 and E7, to inhibit TNF-α-induced NFκB and EGF-induced AP-1 activation, iNOS, and cyclin D1 [95].
Curcumin also exerts downregulation of HPV-18 transcription, inhibition of AP-1 binding activity and expression of c-fos and fra-1 in HeLa cells [96].
In addition, curcumin has been found to upregulate Bax, promote the release of cytochrome c, and downregulate Bcl-2 and Bcl-XL, inducing cell growth inhibition and apoptosis [97]. The cytotoxic potential of curcumin analogs such as diarylpentanoids has been studied in HeLa and CaSki cells as well. Among them, 1,5-Bis(2-hydroxyphenyl)-1,4-pentadien-3-one has been demonstrated to have cytotoxic, anti-proliferative, and apoptotic effects in CaSki cells due to activation of caspase-3. Downregulation of HPV-18 and HPV-16 E6 and E7 oncogene expression has been found as well [98].
Curcumin has been demonstrated to be a sensitizing agent to taxol- and to cisplatin for cervical cancer (HeLa) cells by augmenting apoptosis through downregulation of NF-kB [99,100,101]. Tetrahydrocurcuminoids increased the sensitivity to vinblastine, mitoxantrone, and etoposide of a drug-resistant human cervical carcinoma cell line [102]. Curcumin has also been demonstrated to increase radiosensitivity and apoptosis in HeLa and SiHa cells through an ROS-dependent mechanism [103].
Highly purified curcumin has been also evaluated in clinical setting in a phase I clinical trial of oral administration of 0.5–12 mg for 3 months in four cervical intraepithelial neoplasia (CIN) cases, as well. This dosage has been demonstrated to be clinically safe and has resulted in histological improvements in one out of four cases. Bioavailability and toxicology were also investigated, concluding that the safe dose of curcumin may be up to 8 mg by day (Supplementary Table S1, curcumin-based formulation no.1) [104].

4.2. Other Specialized Metabolites

Although polyphenols may hold the crown of the most tested compounds against HPV in vitro, many other phytochemical species have been investigated, amplifying the panorama of tools provided by plant compounds in this field.

4.2.1. Terpenoids

Several case-control studies have shown significant negative correlation between consumption of terpenoids such as carotenoids (naturally found in sweet potatoes, carrots, dark leafy greens, lettuce, red peppers, tomato, etc.), retinol (accumulating in amaranth, spinach and chard, orange-fleshed sweet potatoes, carrots, pumpkins, yellow maize, mango, and papaya), and tocopherols (retrievable in many plant-derived oils, nuts, seeds, fruits, and greens) antioxidant micronutrients and risk of cervical cancer [105,106,107,108].
Beta-carotene, tocopherols, and retinol have shown no association with risk of cervical cancer from a prospective dietary study [109].
Since studies implying dietary administration of compounds may suffer from inaccuracy in measuring their effective intake and bioavailability, researchers have proposed to put in relationship effective, constitutive plasma levels of recruited patients with the risk of cervical cancer.
Alpha-carotene, β-carotene, lycopene, or tocopherols have shown an inverse association between serum levels and cervical cancer risk in two studies [108,110]. The same findings had not been obtained by Lehtinen and colleagues [111], probably due to the not very consistent construction of the study.
Alpha-carotene, beta-carotene, lutein, zeaxanthin, and alpha-tocopherol plasma levels, individually, have been evaluated in a study enrolling Chinese women. Constitutive plasma levels of these compounds in patients with histologically confirmed cervical cancer before the initiation of radiotherapy or chemotherapy were compared with plasma levels of these anti-oxidants in healthy controls with no history of cervical cancer. A statistically significant inverse correlation was found between plasma levels of these compounds and diagnosis of cervical cancer in a case-control study. In particular, alpha-carotene (odds ratios = 0.50, 95%; confidence interval = 0.38–0.67, p < 0.001), beta-carotene (OR = 0.70, 95% CI = 0.56–0.88, p = 0.002), lutein and zeaxanthin (OR = 0.75, 95% CI = 0.59–0.95, p = 0.015) have showed the most statistically reduction in risk [112].
Lycopene is mostly found in tomato, watermelon, pink grapefruit, guava, and papaya [113]. Lycopene has been demonstrated to be a synergistic agent of cisplatin, able to increase HeLa cells’ sensitivity to cisplatin in vitro. Lycopene determined a significant enhancement of Bax expression and decreased Bcl-2 expression and has markedly activated NRF2 expression and suppressed the NF-κB signaling pathway [114,115].
Saikasaponins from Bupleurum falcatum sensitized HeLa and SiHa cells to cisplatin through intracellular ROS accumulation [116,117].
Saponins are mainly found in chickpeas and soybeans but are accumulated in several herbs, as well. In HeLa cells, saponins induced apoptosis by increased levels of cleaved caspase-3 and caspase-9, Bax, and Bcl-2 [118].
Tanshinone IIA from Salvia spp. has shown to be able to downregulate E6 and E7, induced apoptosis, cell cycle arrest, and inhibition of growth in HeLa, SiHa, CaSki, and C-33A cells [119]. In HPV positive cells, it has been shown to downregulate expression of HPV E6 and E7 genes and modulate associated proteins E6AP and E2F1, to cause S phase cell cycle arrest, induce accumulation of p53 and alter expression of p53-dependent targets, and modulate pRb and related proteins. In particular, authors have demonstrated that the tashinone IIA can cause p53-mediated apoptosis by moderating Bcl2, Bax, and caspase-3.
Ursolic acid can be found in apples, cranberries, peppermint, prunes, oregano, and thyme. Ursolic acid nanoparticles of poly(DL-lactide-co-glycolide) have been able to suppress cervical cancer cell proliferation, invasion, and migration through caspases and p53 in vitro in CaSki, HeLa, C4-1, and SiHa cells. In vivo, this compound has been able to reduce the size of CaSki, HeLa, and SiHa xenografts [120,121].

4.2.2. Thiols

Allicin is mainly found in garlic and onion. In vitro studies conducted in SiHa cells have shown that allicin can induce apoptosis mainly by inhibiting the expression of NRF2 (nuclear factor erythroid 2-related factor 2) and increasing the level of glutathione and superoxide dismutase [122].
Indole-3-carbinol (I-3-C) can be found mainly in broccoli and brussels sprouts. Initial in vitro studies to test anti-cancer properties of I-3-C have been performed on C-33A cells and 17-beta-estradiol (E2)-treated mice have shown selective apoptotic potential on transformed cells and abolishment of the anti-apoptotic effects of estradiol, respectively [123].
These results have been upgraded by several clinical trials performed on I-3-C and its derivative 3,30-diindolylmethane (DIM, a stable form of I-3-C) (Supplementary Table S1, indole-3-carbinol-based formulations no.1–no.6).
A phase I trial, conducted in 1998 on 18 patients with recurrent respiratory papillomatosis, revealed that the use of I-3-C administered orally reduced the papilloma growth rate in 33% of patients (6 out of 18), resulting in cessation of papilloma growth and no surgery in 33% of them (6 out of 18), while no clinical response was found in the resulting 33% (6 out of 18) [124]. This phase I trial also established that I-3-C is safe and well-tolerated by children (Supplementary Table S1, indole-3-carbinol-based formulation no.1).
In addition, a placebo-controlled trial involving 27 patients diagnosed with CIN II–CIN III, resulted in a statistically significant complete regression in four out of eight patients orally supplied with I-3-C 200 mg/day and in four out of nine patients on I-3-C 400 mg/day. A change of the 2/16a-hydroxyestrone ratio was found (Supplementary Table S1, indole-3-carbinol-based formulation no.2) [125].
In 2004, an open-label clinical trial involving 33 patients with recurrent papillomatosis resulted in remission of papillomatous growth and no surgery in 11 (33%) patients treated with I-3-C and reduction in papillomatous growth and less frequent surgery in 10 (30%) patients, while there was no clinical response in 12 (36%) patients (Supplementary Table S1, indole-3-carbinol-based formulation no.3) [126].
Furthermore, a randomized double-blind placebo-controlled phase III trial (64 patients with CIN II or CIN III) demonstrated that CIN decreased by 1–2 grades or to normal histology in 47% (21) of the subjects on DIM treatment and that stratification by the level of dysplasia, age, race, HPV status, tobacco use, contraceptive use did not alter the results. This trial found no systemic toxicities as well (Supplementary Table S1, indole-3-carbinol-based formulation no.4) [127]. Conversely, a double-blind, randomized, controlled, primary prevention trial with 551 patients with low-grade cytological abnormalities demonstrated that DIM oral supplementation had no statistically relevant effect on cytology or infection compared to a placebo, with the development of CIN II or worse in 9% on DIM and 12% on placebo (risk ratio (RR) 0.7). Development of CIN III or worse (RR 0.9) occurred in 4.6% on DIM and 5.1% on placebo and no sign of disease in 27.3% on DIM and 34.3% on placebo at 6 months (RR 0.8) were found (Supplementary Table S1, indole-3-carbinol-based formulation no.5) [128].
Finally, a randomized, double-blind, multi-center clinical trial with a placebo was carried out to evaluate the effects and safety of 3,30-diindolylmethane (DIM, a stable form of I-3-C) administered as a vaginal suppository applied for 180 days for the treatment of CIN [129]. In total, 78 patients 18–39 years of age (56 out of 78 diagnosed with CIN I or CIN II) were enrolled. 200 mg/day and 100 mg/day determined a 100% (CI 95% = 82.35–100.00%) and 90.5% (CI 95% = 69.62–98.83%) regression of CIN lesions, respectively with significantly lower regression in the placebo group (61.1%, CI 95% = 35.75–82.70%) (Supplementary Table S1, indole-3-carbinol-based formulation no.6).
Sulforaphane, a isothiocyanate of cruciferous vegetables [130] in HeLa, Cx, and CxWJ cervical cancer cell models has shown dose-dependent cytotoxicity associated with arrest in G2/M phase through Cyclin B1 downregulation [131].

4.3. Miscellaneous Specialized Metabolites

Berberine (benzylisoquinoline alkaloid) mainly accumulated by plants of the Berberis genus and possesses general anti-inflammatory and anti-cancer properties through inhibition of AP-1 transcription factor with no known toxicity. In SiHa and HeLa cells, berberine specifically downregulates the expression of oncogenic c-Fos [132]. Moreover, berberine has been able to reduce expression of E6 and E7 with a concomitant increase in p53 and pRb expression, loss of telomerase protein, hTERT, resulting in growth inhibition of cervical cancer cells [132]. Furthermore, in HeLa cells, berberine has been shown to alter epigenetic modifications and to disrupt the microtubule network by targeting p53 [133].
Betaines are common in beetroot and have demonstrated proliferation inhibition of HeLa cells through G1/S or S/G2 cell cycle arrest and dose-dependent apoptosis induction through Bax, p53, caspase-3, and Bcl-2 modulation [134].
Decursin and decursinol (pyranocoumarin compounds) from the roots of Angelica gigas, were able to make HeLa cells more sensitive to tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) by activating caspase-8 and caspase-9 [135].
Piperine (1-pipeoylpeperdine) can be found in black and long peppers (Piper nigrum, Piper longum). Endowed with anti-cancer activities, it has been tested against HPV models as well. In HeLa cells, it was demonstrated to induce dose-dependent apoptosis through caspase-3, DNA fragmentation, ROS generation, and nuclear condensation [136].
Withaferin A (steroid lactones) from Withania somnifera induced apoptosis of CaSki cells through E6 and E7 downregulation and consequent inhibition of tumor growth [137]. In particular, withaferin A alters expression levels of p53-mediated apoptotic markers Bcl2, Bax, caspase-3, and cleaved PARP.

4.4. Polysaccharides

Crude polysaccharides from Glycyrrhiza uralensis have been recently investigated as adjuvants of HPV DNA vaccine tested in a preclinical model to detect the efficacy of combinatorial approaches in improving antigen-specific cellular and humoral immune responses. C57BL/6 mice were co-immunized with HPV-DNA vaccine and polysaccharides. Crude polysaccharides enhanced the DNA vaccine-induced antigen-specific CD4+ and CD8+ T cell responses and the levels of IgG, IgG1, and IgG2a. Researchers also demonstrated that administration of Glycyrrhiza crude polysaccharides had no adverse effects in mice but significantly increased B cells and macrophages in the spleen. Intragastric administration of polysaccharides also significantly increased the levels of CD4+ and CD8+ T splenocytes [138].

4.5. Proteins and Peptides

Lectins from different species among which Clematis montana and Astragalus mongholicus have been tested against HeLa cells in vitro showing antiproliferative properties through induction of apoptosis via caspase-dependent pathways and S-phase arrest, upregulation of p21 and p27, and reduction in active complex cyclin E/CDK2 kinase [139].
Peptides from Triticum aestivum, Abrus precatorius, and Trichosanthes kirilowii have been shown to have antiproliferative activities in HeLa cells in vitro by different mechanisms such as induction of DNA damage and G2 arrest and induction of apoptosis with the generation of ROS [140,141,142].

5. Combinations of Purified Phytocompounds

Specialized metabolites, as purified and characterized entities used in combination, are reviewed in this section.
Alpha–, beta–, and gamma–carotene and zeaxanthin carotenoids, previously studied as individual compounds [108,110], have been evaluated in combination case studies, as well, demonstrating that the risk of developing cervical cancer is inversely correlated with blood levels of this combination of compounds in patients [143].
Curcumin and ellagic acid combined at different concentrations have shown improved inhibitory effects on HeLa cells compared to the single compounds administered individually by MTT assay. The combination of curcumin and ellagic acid restored p53, increased ROS levels, and induced DNA damage as well. It has been further demonstrated that curcumin and ellagic acid determined a downregulation of the HPV E6 oncoprotein as well [144].
Purified green tea compounds have been tested in different clinical settings.
Different ECGC-containing (200 mg) formulations, such as polyphenon E (poly E, decaffeinated and enriched green tea catechin extract, containing 85–95% total catechins, with 56–72% as EGCG) and/or ECGC capsules, have been evaluated in 90 patients with HPV cervical lesions from chronic cervicitis to severe dysplasia. Clinical effects have been proven both through intravaginal and oral delivery of the two formulations (on a daily basis, for 8–12 weeks). Indeed, 74% of patients treated topically with poly E ointment, 75% of patients treated both topically and orally with poly E ointment + oral EGCG, 50% of patients treated orally with poly E, and 60% of patients treated orally with EGCG, showed a significant ameliorating response (Supplementary Table S1, catechin-based formulation no.1) [145].
A phase II randomized, double-blind, placebo-controlled trial (98 patients with persistent high-risk HPV infection and CIN I, daily treated for 4 months with oral poly E capsules), demonstrated no statistically relevant promotion of neither clearance of persistent high-risk HPV or normal histopathology but demonstrated that poly E is acceptable, safe, and well-tolerated (Supplementary Table S1, catechin-based formulation no.2) [146].
Rutin (a flavonoid common in capers, olives, buckwheat, asparagus) purified from Citrus sinensis peel was combined with the fucoidan polysaccharide from Fucus vesiculosus algae in a rutin-fucoidan (Ru-Fu) to form a complex with low bioavailability of flavonoids. The Ru-Fu complex has been shown to effectively induce anti-proliferation, S-phase cell cycle i and apoptosis in HeLa cells via nuclear fragmentation, ROS generation, and loss of mitochondrial membrane potential [147].

6. Plant Extracts

The use of fractions or extracts from plant parts may represent an easier and cheaper approach compared to highly purified phytocompounds. For this reason, extracts of different origins and types have been tested for anti-HPV properties.
Extracts from the stem of Cudrania tricuspidata have been shown to induce dose-dependent cytotoxicity in SiHa tumor cells, but not in HaCaT keratinocytes at 0.125–0.5 mg/mL concentrations. The mechanism underneath this selective effect has been related to downregulation of E6 and E7 and apoptosis triggering related to upregulation of Fas, death receptor 5, TRAIL, activation of caspases-3 and -8, and cleavage of PARP (poly-ADP-ribose polymerase) [148].
Ficus carica fruit latex has been demonstrated to inhibit the rapid growth and invasiveness of CaSki and HeLa cells through downregulation of p16 and E6 and E7 gene expression [149].
Aqueous extract from Crocus sativus and its two main purified compounds (i.e., picrocrocin, a monoterpene aldehyde; and the carotenoid crocin) reduced TC-1 cell proliferation in vitro in a concentration- and time-dependent fashion. When orally administered in association with a DNA vaccine encoding E7-NT (gp96), both extracts and purified compounds acted synergistically in the inhibition of experimental tumor growth in mice, acting either as preventative (aqueous extract, picrocrocin) or therapeutic (crocin) agents [150].

7. Plant Extracts Mixtures

Extracts of different origins have also been tested in combination against HPV to possibly integrate the benefits of multiple components.
Inhibition of AP-1 and STAT3—which are known to promote cervical carcinogenesis, and specifically downregulate the expression of viral oncogenes E6 and E7—has been found to represent the major mechanism of action of a mixture of extracts from Pinellia pedatisecta (rhizome extract), Bryophyllum pinnata (fractionated leaf extract, enriched in bryophyllin A), and Phyllanthus emblica (fruit, crude extract), Brucea javanica (oil emulsion) in different human cervical cancer in vitro models [151,152,153].
Basant, a mixture of extracts from Emblica officinalis and Aloe vera—added with curcumin, saponins, and Mentha citrata oil and Praneem—blocked transduction of HPV-16 pseudovirus in HeLa cells, when diluted at non-cytotoxic concentrations (IC50 1:20,000). Basant was found to be safe according to pre-clinical toxicology carried out on rabbit vagina after application for 7 consecutive days or twice daily for 3 weeks [154].
Therefore, a placebo-controlled phase II randomized study in which either intra-vaginal application of curcumin-containing capsules or basant ointment was performed in 280 HPV-infected women in absence of high grade CIN. This study has demonstrated a statistically relevant resolution of cervical HPV infection in the case of basant ointment application for 30 days and no severe adverse events have been reported (Supplementary Table S1, curcumin-based formulation no.2) [155].
All the 11 HPV-infected women with low-grade cervical lesions enrolled in a placebo-controlled phase I trial with basant, cleared HPV-16 infection upon basant application (Supplementary Table S1, curcumin-based formulation no.3) [156].
Praneem polyherbal, a mixture of extracts from Azadirachta indica and Emblica officinalis—combined with saponins, Mentha citrata oil, and Aloe vera ge—have been examined in 20 women diagnosed with low-grade squamous intraepithelial lesions in a placebo-controlled clinical study [157]. Topical administration (500 mg, intra-vaginally, 30 day-treatment) cleared HPV-16 in 60% of patients. A subsequent treatment was administered to women who had not cleared the infection, determining the end of infection in 80% of patients overall (Supplementary Table S1, other polyherbal-based formulation no. 1). Microbicidal action against infectious agents that are known to cooperate in the promotion of HPV carcinogenesis has been demonstrated for Praneem and may disclose further effects on cervical lesions [158].

8. Phytocompounds with Anti-HPV Potential Inferred by In Silico Approaches

With the evolution of sophisticated molecular modeling, in silico approaches have been utilized to target HPV oncoproteins to make them non-functional.
As a consequence of these studies, it was inferred that daphnoretin (a protein kinase C activator isolated from the plant Wilkostroemia indica), withaferin A, and artemisinin (a sesquiterpene lactone derived from Artemisia annua) have docking capacity on E6 from HPV-16 and HPV-18 and may exert an inhibitory effect nearly 20 times stronger than curcumin, ECGC, or jaceosidin, which are known to functionally interfere with E6 [159,160].
Furthermore, 20 plant-derived compounds from different sources have been collected by Nabati and colleagues and tested against E6 [161]. In this study, the molecular interaction of HPV-16 E6 protein with plant-derived inhibitors in the proximity of E6AP, p53, and Myc binding sites were investigated using docking analysis. Ginkgetin, hypericin, and apigetrin have shown significant ability to block three binding sites on HPV-16 E6 oncoprotein with minimum binding energy [161].
With the aim to get a stronger immune response against HPV infection, an in silico model to predict the adjuvant activity of plant-derived compounds in the context of preventive immunization was built as well. This study predicted that neohesperidin, a flavanone glycoside mainly found in citrus fruits, may be able to elicit an immune response and subsequently validated this hypothesis in vivo in a mouse model. This integrated methodology (virtual screening combined with in vivo testing) allowed us to set up a protocol to confer humoral immune response against HPV infection [162].
Despite the power, speed, and potential of this approach, further studies and translations into experimental settings are needed to validate findings and to develop rationally designed drugs.

9. The (Plant) Cell Factory

As highlighted, plant cells represent a natural factory of useful metabolites. For pharmaceutical production ends, when naturally occurring sources of a compound of interest are rare, endangered, available only on a seasonal basis, or when these molecules need qualitative or quantitative improvement, ‘artificial’ production in a bioreactor can be accomplished. Microbes, such as bacteria and yeasts—but especially plant-based production platforms—can be used to this end. The use of plants as biofactories for the production of either small molecules (i.e., specialized metabolites, sugars, and others) or complex polymers like proteins (such as lectins) that may be either natural or exogenous is known as ‘plant molecular farming’ (PMF; Figure 3) [163,164,165]. Different bioreactor formats have been derived from plants to this end: whole plants, cell suspension, and organ cultures. Plant cell totipotency and plasticity, low-cost upstream production, intrinsic safety based on the inability of human pathogens to replicate in plant cells, and potential for industrial large-scale production with manufacturing capacity rapidly scalable to meet market demands, are benefits generally applicable to plant-based production platforms making PMF attractive for future applications [165,166,167].
In the case of small molecules, once the biosynthetic pathways responsible for the biosynthesis of a desired product and its regulation by the involved enzymes and transcription factors are clarified, the expression of one or more sets of the necessary enzymes and/or transcription factors can be accomplished. Stable expression in transgenic whole plants, plant cells, and plant organs has been the main expression strategy applied to PMF of specialized metabolites and it has been accomplished in natural hosts or in exogenous species if technologically feasible [168] (Figure 3).
The reconstitution of the flavonoids pathway was demonstrated in both E. coli and S. cerevisiae [169,170]. Nevertheless, modifications—such as glycosilation, methylation, and acylation—that are needed to obtain more ‘authentic’, stable, and biologically active flavonoids have to be assured to occur in the requested pattern in these systems. Engineered microbial systems have also been demonstrated to be useful to produce derivatives of flavonoids that do not occur in nature, and these compounds are being investigated for their pharmaceutical potential [171]. Carotenoids, akaloids, betalains, and glucosinolates are among the further specialized metabolites that have been mostly approached by metabolic engineering [172].
Especially regarding the production of proteins with anti-cancer activity, many plant species and the different culture formats thereof can produce them naturally or have been used as expression platforms for recombinant, exogenous production [165]. Plant cells can synthesize proteins with complex structure, preserving their authentic post-translational modifications (e.g., glycosylation, disulfide bond formation) and eventually improving function (‘biobetter’ pharmaceuticals). An application of these advantages occurs in the plant-based production of eterodimeric lectins such as viscumin. Lectins are anti-tumoral complex glycoproteins that also have proven effectiveness against HPV in vitro. These are ‘familiar’ compounds for plants and heterologous lectins are commonly less toxic for plant cells than for other cell types. In plants, viscumin can be produced from a single transcript and the monomeric precursor can be processed as in the native host, which further lowers process complexity. Otherwise, lectins should be produced by separate bacterial fermentations and should require out-of-bioreactor refolding and glycosilation (necessary for stability and efficacy), leading to poor recoveries. On the other hand, production by yeasts would decorate lectins with high-mannose rather than complex-type glycans [173,174,175]. Mammalian cells are unlikely to express lectins at high levels because they would be toxic [176].

10. Marine Substances

A few marine-derived bioactive compounds have been shown to have anti-HPV and related cancer activities, demonstrating the anti-HPV potential of marine resources ‘competing’ with the well-known plant-derived compounds. Polysaccharides, among other bioactive compounds, present in different marine organisms—from microbes to algae to invertebrates—have been examined for potential application against HPV.

10.1. Algae

Carrageenan, a sulfated polysaccharide of D-galactose extracted from red algae, is widely used as a thickener in cosmetics and food products [177]. Carrageenan has been demonstrated to act by impairing the binding of HPV to cells and, subsequently, by blocking infection through heparin sulfate-independent mechanism [177]. This ability was maintained by milk-based products containing carrageenan, even though this was at a lower efficiency than pure carrageenan [177]. Carrageenan has been shown to block HPV genital transmission in a mouse cervicovaginal model as well. The sulfated galactose structure and content have been defined as important factors for the anti-HPV effect of carrageenans [178].
Fucan (i.e., L-fucose-containing sulfated polysaccharides) contained in a polysaccharide-rich extract from the brown seaweed Sargassum filipendula showed significant reduction in HeLa cell proliferation by release of apoptosis-inducing factor (AIF) into cytosol from mitochondria [179]. In another work, an isolated antioxidant heterofucan (SF-1.5v) inhibited HeLa cell proliferation by a molecular mechanism that remains unclear [180].
A fucose-enriched fraction, isolated from the brown seaweed Sargassum stenophyllum has been shown to significantly alter cellular morphology of HeLa cells and to reduce cell growth in a dose-dependent manner (40–80 µg/mL) [181].
Heparin and heparinoids are sulfate polysaccharides that have been demonstrated to interfere with the initial attachment of HPV viral particles to the host cell [182]. Through the same mechanism, the marine heparinoid polysaccharides, alginic acid and fucoidan, from brown algae have been demonstrated to effectively block HPV pseudovirions infection [177]. In addition to blocking infection, alginic acid from the brown seaweed Laminaria brasiliensis demonstrated an inhibitory effect on HeLa cells by promoting atypical mitoses and nuclear fragmentation [181].

10.2. Marine Invertebrates

Chitosan, which can derived from deacetylation of chitin from the shells of crustaceans, has been demonstrated to have inhibitory effects on HPV infection-induced chronic cervicitis with a superior negative conversion rate compared to interferon α-2b gel [182]. Diethylaminoethyl chitosan has been found to trigger apoptosis in HeLa cells through upregulation of caspases, p53, and Bax, and downregulation of Bcl-2 [183].
Hemocyanins are a large family of respiratory glycoproteins found in some mollusks. They possess several beneficial effects linked to potent Th1 adjuvant activity. Keyhole limpet hemocyanin (KLH) and two new hemocyanins from Concholepas concholepas (CCH) and Fissurella latimarginata (FLH) have been utilized in a pre-clinical mouse model of orthotopic oral HPV-associated cancer. In order to potentiate the antitumor activity, the two hemocyanins were applied in combination with QS-21, a synthetic saponin. The KLH and FLH in combination with the QS-21 adjuvant showed reduced tumor development and greater overall survival. This study suggests potential for the effective use of hemocyanins plus adjuvants for the development of immunotherapies against head and neck cancers [184].
Manzamine A, an alkaloid from an Indo-Pacific sponge, has been demonstrated to have antiproliferative activity against C33A, HeLa, SiHa, and CaSki cells [185]. In SiHa and CaSki cells, the antiproliferative effect of manzamine has been demonstrated to occur at non-cytotoxic concentrations (up to 4 μM) through the block of cell cycle at G1/S phase and by regulation of cell cycle-related genes and restoration of p21 and p53. Along these anti-tumor properties, interestingly, manzamine A administration led to a significant decrease in SIX1 and CK2α proteins intracellular level along with the regulation of cell-cycle related checkpoint proteins in the mentioned cell lines, suggesting possible activity in vivo against cervical cancer, since it expresses high levels of SIX1. In a direct comparison to apigenin, a known inhibitor of cyclin-dependent kinases, manzamin demonstrated a 10-fold inhibiting activity on CK2α and SIX1 proteins than apigenin.
Mer2, a low molecular weight polypeptide from the clam Meretrix meretrix has been demonstrated to significantly block the growth of HeLa cells through a dose- and time-dependent mechanism and has induced morphology changes and apoptosis [186].
SJAMP, a polysaccharide isolated from the marine echinoderm Apostichopus japonicus, have been shown to significantly inhibit the proliferation of HeLa cells [187]. In this model, SJAMP downregulated the proliferating cell nuclear antigen (PCNA) and the cell cycle inhibitor protein Mdm2, and induced the arrest of cell cycle in G1 phase [187].

10.3. Marine Microbes

Marine microbes can be sources of different bioactive compounds with anti-HPV cancer activity as well.
Antimycin alkaloids and related analogs isolated from the lab-grown marine Streptomyces sp. THS-55 has exhibited cytotoxic activity in vitro against HeLa cells. One of them in particular, denoted as NADA, showed the highest potency by inhibition of proliferation, cell cycle arrest, and apoptosis induction. NADA has been shown to degrade HPV E6 and E7 viral oncoproteins and to inhibit their function through ROS-mediated ubiquitin-dependent proteasome system activation [188].
Gliotoxin, which can be produced by marine Aspergillus species, has reduced the proliferation of HeLa cells by apoptosis associated with a drop of mitochondrial membrane potential and activation of Bax, caspase-3, caspase-8, and caspase-9, together with suppression of Bcl-2 [189].
Neoechinulin A (prenylated indole alkaloid) produced by cultured marine-derived Microsporum sp. [190] has shown cytotoxic effect on HeLa cells through downregulation of Bcl-2 and upregulation of Bax and activating the caspase-3 pathway.
Despite the potential to block HPV infection, as well as to have in vitro effects, these marine-derived compounds need to be explored and characterized further. Current research suggests that marine natural products of proven efficacy show low cytotoxicity, low production costs, and acceptability as novel drug tools [191,192].

11. Bacterial Substances

There is evidence that, in one case, bacterial compounds have effects on HPV. HPV-16, -18, and -6 pseudovirions have been inhibited by E. coli K5 capsular polysaccharides due to their similarity with the backbone of heparin/heparan biosynthetic precursor [193].

12. Discussion

HPV-related lesions still represent a major concern for public health. The number of recorded cancer cases is likely to increase due to demographic changes and to improved and more widely available diagnostics. Combined with increasing access to healthcare systems in developing countries, this will enhance the demand for anti-HPV agents. More importantly for developing countries, approaches satisfying criteria such as to be possibly self-applied, non-destructive, with minimal side effects—while being effective and affordable—should be fulfilled to improve pharmacological treatment of HPV. Reducing side effects of currently available radio- and chemo-therapy is central as well. Therefore, exploring additional approaches to HPV prevention and therapy is useful to possibly corroborate current intervention methods.
Plants are the main source of chemicals that demonstrated effects against HPV. Despite the ubiquitous polyphenols receiving the most attention in in vitro/pre-clinical studies, several different mechanisms are exerted by natural compounds against HPV in various models: apoptosis induction, growth arrest, effects on DNA, alteration of signal transduction, alteration of redox balance, as well as capacity to specifically modulate the expression and activity of HPV oncogenes. These mechanisms play roles in different steps of the cascade of HPV-induced cell transformation and indirectly show the possible contribution of these compounds in the fight against HPV.
One encouraging feature of plant-derived dietary compounds is their behavior as safe and effective chemo- and radio-sensitizer agents. A possible drawback is the eventual poor bioavailability of some of these compounds. To improve this aspect, encapsulation or complexation of plant compounds with other (bio)materials to form nanoparticles or complexes, respectively, has been considered. Indeed, it has been demonstrated that complexes may improve stability and, as a consequence, targeted release. On the other hand, nanosized phytochemicals have demonstrated high stability, improved absorption rates, and reduced compound degradation as well. Nevertheless, nanotechnology remains an issue for clinical success, mainly due to safety.
Considering the different and peculiar targets of action of each natural compound that has been tested against HPV, the synergistic use of molecules may be considered as a tool to amplify their effects on HPV. The few, recent studies available on the improved effects of association of purified phytochemicals compared to separate entities seem to point in this direction. In this sense, purified phytochemicals, alone or in combination—due to wider availability, precise identity, quantity and batch-to-batch reproducibility—may probably represent the safest and the most feasible entities to be translated to clinics. Other formulations—such as extracts, suffering from batch-to-batch changes—may be unfeasible for real clinical applications. Phytocompounds derived from food and traditional medicine may have an advantage over others, due to use-proven non-toxicity within certain ranges. Compounds that directly affect HPV E6 and/or E7 are probably the most promising candidates for HPV therapy, due to their impact on the major and most specific players of HPV oncogenesis.
Despite plants remaining the main source of compounds with effects on HPV, notably, a few studies on marine compounds of different origins (algal, microbial, animal) are available, and they offer examples that further enlarge the number of compounds with pharmacological values against HPV.
Undoubtedly, additional basic research to further evaluate the properties and effects of natural (and, especially, plant-derived) compounds is needed—in vitro and in vivo—to better clarify mechanisms of action and to evaluate the entity of the interference with cell pathways. Studies on the effects on the immune system are lacking. These studies may be of significance to investigate if natural compounds may have a role in modifying the immunosuppressive tumor microenvironment as well. It would be also desirable to study and delve into the role of natural bioactive molecules in persistent HPV infection, although diagnostic methods for detecting HPV viral latency are still lacking. Natural products, and especially phytochemicals, may help in this area. Immune-enhancing natural compounds that can potentiate innate or antigen-specific cell-mediated immune responses should be taken into consideration. Results from the few studies available suggest that combinatorial treatment strategies, such as combining immunotherapy with the administration of a natural compound with tumor-killing properties, may be a more effective anticancer strategy than single-modality of treatments.
Several clinical trials (mostly phase I–II) have been published demonstrating the possible clinical use of phytocompounds (listed also in Supplementary Table S1), mostly with no concerns regarding safety issues, and some of them with encouraging results. Apart from the outcome, ablative procedures on advanced HPV lesions require specialized infrastructure or expensive approaches. Natural compounds would represent a simpler, safe, and cost-effective approach with extensive pre-clinical validation. They would be easily applied both orally and topically if clinical elimination of HPV infection or reversion of lesions could be effectively and finally proven. These features would make them well suited to application in low-income developing countries. To this end, more structured, uniform, large-scale, and multi-centered clinical trials than those that have been already reported are strongly needed. These are crucial to provide the necessary information for a clear assessment of their human use.
Among the most clinically studied (at least in the western countries) substances of natural origin, DIM appears promising. As already introduced, DIM is a naturally occurring indole found in cruciferous vegetables. It is a well-known dietary supplement. DIM is derived from the transformation of indole-3-carbinol (I-3-C) in the acidic environment of the stomach. The chemopreventive, anti-mutagenic, and anti-carcinogenic properties of DIM and I-3-C also have potential against HPV. As a derivative of I-3-C, DIM reduces the levels of 16-hydroxy estrogen metabolites and increases the formation of 2-hydroxy metabolites, thereby increasing antioxidant activities. Clinical reports of success have been published in the treatment of HPV-related lesions of the cervix and the larynx with this family of compounds (Supplementary Table S1). Despite mostly favorable data, some of these studies failed to demonstrate a statistically significant advantage in the use of these compounds. Sometimes, this may be explained with the wrong choice of the placebo (i.e., Del Priore et al. [131] used rice bran that may have potential activity against HPV lesions). Nevertheless, it is difficult to arrive at a final conclusion due to lack of uniformity in dosage and administration route (DIM intravaginal route seem to be the most promising in the case of cervix lesions). In order of importance (i.e., as per the number of published clinical studies), curcumin-derived compounds follow those based on DIM and I-3-C, especially in Eastern countries. Nevertheless, these studies suffer from either a very low number of enrolled patients or, when enrolling wider patient numbers, lack full statistical relevance. In addition, despite not directly concerning HPV, a study carried out on breast cancer patients revealed that the use of curcumin could reduce tamoxifen and endoxifen concentrations below the levels of therapeutic efficacy, suggesting that tamoxifen treatment efficacy should be adequately monitored in these patients [194]. This study highlighted the importance of monitoring the activity of bioactive compounds on concomitant treatments with other drugs.
Available HPV clinical data, regardless of the compound and formulation, seem to point out to a general safety of phyto-compounds. However, clinical research has still to fill the current gaps of knowledge about optimal doses and pharmaco-dynamics. Clinical research should also focus on a more detailed evaluation of the interactions of these compounds with other drugs and possible adverse effects for comparative studies with canonical treatments and within combinational schedules. Further investigation on the synergistic effects among compounds with approved HPV therapy schedules should be accomplished. The latter approach may be the most interesting application of phytochemicals if improvement of action and mitigation of adverse effects of canonical radio- and chemo-therapy-based approaches could be proven.
In the absence of highly finalized clinical research, the wealth of currently available studies runs the risk of providing only information limited to the huge variety of compounds, formulations, and targets that have been tested so far for potential against HPV. All the work done will remain necessarily associated with strong limitations until the most promising compounds are chosen for conclusive clinical investigation and possibly authorized for safe human use.
Once these goals will be achieved, thanks to current knowledge in the regulation and control of metabolic networks in plants and bacterial biofactories, both synthetic biology and scalable industrial production of natural compounds could be guaranteed and qualitatively/quantitatively improved for wide clinical application. The use of plant biotechnology and its advances will especially be of help in delivering such agents, since established manufacturing platforms based on mammalian cells and microbes are endowed with either high production costs, anti-cancer-agent inherent toxicity, or synthesis quality limitations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jcm11051465/s1, Table S1: Summary of plant-derived formulations that have been utilized in clinical trials against HPV, listed in chronological order, by chemical group association; Table S2: Summary of phytochemicals that have been screened against HPV listed in alphabetical order within the different chemical groups (↑: upregulation/induction/increase; ↓: downregulation/reduction/decrease). References [195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, S.M., F.P. and A.V.; Writing—original draft preparation, S.M. and R.P.; Writing—review and editing, S.M., R.P., F.P. and A.V.; Supervision, S.M. and A.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by “Fondazione Federico Calabresi Onlus” (grant 20/09/R/46).

Acknowledgments

Silvia Massa acknowledges support from the ‘InnCoCells’ project funded by the European Union’s Horizon 2020 research and innovation programme under grant agreement 101000373.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Mathers, C.; Parkin, D.M.; Piñeros, M.; Znaor, A.; Bray, F. Estimating the global cancer incidence and mortality in 2018: GLOBOCAN sources and methods. Int. J. Cancer 2018, 144, 1941–1953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Dagenais, G.R.; Leong, D.P.; Rangarajan, S.; Lanas, F.; Lopez-Jaramillo, P.; Gupta, R.; Diaz, R.; Avezum, A.; Oliveira, G.B.F.; Wielgosz, A.; et al. Variations in common diseases, hospital admissions, and deaths in middle-aged adults in 21 countries from five continents (PURE): A prospective cohort study. Lancet 2020, 395, 785–794. [Google Scholar] [CrossRef]
  4. De Martel, C.; Plummer, M.; Vignat, J.; Franceschi, S. Worldwide burden of cancer attributable to HPV by site, country and HPV type. Int. Agency Res. Cancer (IARC/WHO) 2017, 141, 664–670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Badaracco, G.; Savarese, A.; Micheli, A.; Rizzo, C.; Paolini, F.; Carosi, M.; Cutillo, G.; Vizza, E.; Arcangeli, G.; Venuti, A. Persistence of HPV after radio-chemotherapy in locally advanced cervical cancer. Oncol. Rep. 2010, 23, 1093–1099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Burd, E.M. Human Papillomavirus and Cervical Cancer. Clin. Microbiol. Rev. 2003, 16, 1–17. [Google Scholar] [CrossRef] [Green Version]
  7. World Health Organisation. International Agency for Research on Cancer IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. In Human Papillomaviruses; World Health Organisation: Geneva, Switzerland, 2007; Volume 90, ISBN 9789283212904. [Google Scholar]
  8. Kessler, T.A. Cervical Cancer: Prevention and Early Detection. Semin. Oncol. Nurs. 2017, 33, 172–183. [Google Scholar] [CrossRef]
  9. Ledford, L.R.C.; Lockwood, S. Scope and Epidemiology of Gynecologic Cancers: An Overview. Semin. Oncol. Nurs. 2019, 35, 147–150. [Google Scholar] [CrossRef]
  10. Cordeiro, M.N.; De Lima, R.d.C.P.; Paolini, F.; Melo, A.R.d.S.; Campos, A.P.F.; Venuti, A.; De Freitas, A.C. Current research into novel therapeutic vaccines against cervical cancer. Expert Rev. Anticancer Ther. 2018, 18, 365–376. [Google Scholar] [CrossRef]
  11. Miyauchi, S.; Sanders, P.D.; Guram, K.; Kim, S.S.; Paolini, F.; Venuti, A.; Cohen, E.E.W.; Gutkind, J.S.; Califano, J.A.; Sharabi, A.B. HPV16 E5 Mediates Resistance to PD-L1 Blockade and Can Be Targeted with Rimantadine in Head and Neck Cancer. Cancer Res. 2020, 80, 732–746. [Google Scholar] [CrossRef] [Green Version]
  12. Gomes, D.; Silvestre, S.; Duarte, A.P.; Venuti, A.; Soares, C.P.; Passarinha, L.; Sousa, Â. In Silico Approaches: A Way to Unveil Novel Therapeutic Drugs for Cervical Cancer Management. Pharmaceuticals 2021, 14, 741. [Google Scholar] [CrossRef] [PubMed]
  13. Schepisi, G.; Casadei, C.; Toma, I.; Poti, G.; Iaia, M.L.; Farolfi, A.; Conteduca, V.; Lolli, C.; Ravaglia, G.; Brighi, N.; et al. Immunotherapy and Its Development for Gynecological (Ovarian, Endometrial and Cervical) Tumors: From Immune Checkpoint Inhibitors to Chimeric Antigen Receptor (CAR)-T Cell Therapy. Cancers 2021, 13, 840. [Google Scholar] [CrossRef] [PubMed]
  14. Attademo, L.; Tuninetti, V.; Pisano, C.; Cecere, S.C.; Di Napoli, M.; Tambaro, R.; Valabrega, G.; Musacchio, L.; Setola, S.V.; Piccirillo, P.; et al. Immunotherapy in cervix cancer. Cancer Treat. Rev. 2020, 90, 102088. [Google Scholar] [CrossRef] [PubMed]
  15. Vici, P.; Pizzuti, L.; Mariani, L.; Zampa, G.; Santini, D.; Di Lauro, L.; Gamucci, T.; Natoli, C.; Marchetti, P.; Barba, M.; et al. Targeting immune response with therapeutic vaccines in premalignant lesions and cervical cancer: Hope or reality from clinical studies. Expert Rev. Vaccines 2016, 15, 1327–1336. [Google Scholar] [CrossRef] [PubMed]
  16. Vici, P.; Mariani, L.; Pizzuti, L.; Sergi, D.; Di Lauro, L.; Vizza, E.; Tomao, F.; Tomao, S.; Cavallotti, C.; Paolini, F.; et al. Immunologic treatments for precancerous lesions and uterine cervical cancer. J. Exp. Clin. Cancer Res. 2014, 33, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Cheng, L.; Wang, Y.; Du, J. Human Papillomavirus Vaccines: An Updated Review. Vaccines 2020, 8, 39. [Google Scholar] [CrossRef] [PubMed]
  18. Massa, S.; Franconi, R.; Brandi, R.; Muller, A.; Mett, V.; Yusibov, V.; Venuti, A. Anti-cancer activity of plant-produced HPV16 E7 vaccine. Vaccine 2007, 25, 3018–3021. [Google Scholar] [CrossRef]
  19. Franconi, R.; Di Bonito, P.; Dibello, F.; Accardi, L.; Muller, A.; Cirilli, A.; Simeone, P.; Donà, M.G.; Venuti, A.; Giorgi, C. Plant-derived human papillomavirus 16 E7 oncoprotein induces immune response and specific tumor protection. Cancer Res. 2002, 62, 3654–3658. [Google Scholar]
  20. Di Bonito, P.; Grasso, F.; Mangino, G.; Massa, S.; Illiano, E.; Franconi, R.; Fanales-Belasio, E.; Falchi, M.; Affabris, E.; Giorgi, C.; et al. Immunomodulatory Activity of a Plant Extract Containing Human Papillomavirus 16-E7 Protein in Human Monocyte-Derived Dendritic Cells. Int. J. Immunopathol. Pharmacol. 2009, 22, 967–978. [Google Scholar] [CrossRef] [Green Version]
  21. Demurtas, O.C.; Massa, S.; Ferrante, P.; Venuti, A.; Franconi, R.; Giuliano, G. A Chlamydomonas-Derived Human Papillomavirus 16 E7 Vaccine Induces Specific Tumor Protection. PLoS ONE 2013, 8, e61473. [Google Scholar] [CrossRef] [Green Version]
  22. Franconi, R.; Massa, S.; Paolini, F.; Vici, P.; Venuti, A. Plant-Derived Natural Compounds in Genetic Vaccination and Therapy for HPV-Associated Cancers. Cancers 2020, 12, 3101. [Google Scholar] [CrossRef] [PubMed]
  23. Moga, M.A.; Dimienescu, O.G.; Arvatescu, C.A.; Mironescu, A.; Dracea, L.; Ples, L. The role of natural polyphenols in the prevention and treatment of cervical cancer—An overview. Molecules 2016, 21, 1055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Drețcanu, G.; Iuhas, C.I.; Diaconeasa, Z. The involvement of natural polyphenols in the chemoprevention of cervical cancer. Int. J. Mol. Sci. 2021, 22, 8812. [Google Scholar] [CrossRef] [PubMed]
  25. Di Domenico, F.; Foppoli, C.; Coccia, R.; Perluigi, M. Antioxidants in cervical cancer: Chemopreventive and chemotherapeutic effects of polyphenols. Biochim. Biophys. Acta Mol. Basis Dis. 2012, 1822, 737–747. [Google Scholar] [CrossRef] [Green Version]
  26. Chen, H.; Liu, R.H. Potential Mechanisms of Action of Dietary Phytochemicals for Cancer Prevention by Targeting Cellular Signaling Transduction Pathways. J. Agric. Food Chem. 2018, 66, 3260–3276. [Google Scholar] [CrossRef]
  27. Arora, I.; Sharma, M.; Tollefsbol, T.O. Combinatorial Epigenetics Impact of Polyphenols and Phytochemicals in Cancer Prevention and Therapy. Int. J. Mol. Sci. 2019, 20, 4567. [Google Scholar] [CrossRef] [Green Version]
  28. Liu, C.-J.; Singh, S.; Sharma, B.; Kanwar, S.S.; Kumar, A. Lead Phytochemicals for Anticancer Drug Development. Front. Plant Sci. 2016, 7, 1667. [Google Scholar] [CrossRef] [Green Version]
  29. Kotecha, R.; Takami, A.; Espinoza, J.L. Dietary Phytochemicals and Cancer Chemoprevention: A Review of the Clinical Evidence. Oncotarget 2016, 7, 52517. [Google Scholar] [CrossRef] [Green Version]
  30. Dubensky, T.W.; Kanne, D.B.; Leong, M.L. Rationale, progress and development of vaccines utilizing STING-activating cyclic dinucleotide adjuvants. Ther. Adv. Vaccines 2013, 1, 131–143. [Google Scholar] [CrossRef]
  31. Barber, G.N. Cytoplasmic DNA innate immune pathways. Immunol. Rev. 2011, 243, 99–108. [Google Scholar] [CrossRef]
  32. Pilla, L.; Rivoltini, L.; Patuzzo, R.; Marrari, A.; Valdagni, R.; Parmiani, G. Multipeptide vaccination in cancer patients. Expert Opin. Biol. Ther. 2009, 9, 1043–1055. [Google Scholar] [CrossRef] [PubMed]
  33. Mett, V.; Farrance, C.E.; Green, B.J.; Yusibov, V. Plants as biofactories. Biologicals 2008, 36, 354–358. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, T.; Kerbler, S.M.; Fernie, A.R.; Zhang, Y. Plant cell cultures as heterologous bio-factories for secondary metabolite production. Plant Commun. 2021, 2, 100235. [Google Scholar] [CrossRef] [PubMed]
  35. Chang, W.-C.; Hsieh, C.-H.; Hsiao, M.-W.; Lin, W.-C.; Hung, Y.-C.; Ye, J.-C. Caffeic acid induces apoptosis in human cervical cancer cells through the mitochondrial pathway. Taiwan. J. Obstet. Gynecol. 2010, 49, 419–424. [Google Scholar] [CrossRef] [Green Version]
  36. Anantharaju, P.G.; Reddy, D.B.; Padukudru, M.A.; Kumari Chitturi, C.M.; Vimalambike, M.G.; Madhunapantula, S.V. Induction of colon and cervical cancer cell death by cinnamic acid derivatives is mediated through the inhibition of Histone Deacetylases (HDAC). PLoS ONE 2017, 12, e0186208. [Google Scholar] [CrossRef]
  37. 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] [Green Version]
  38. Gao, J.; Yu, H.; Guo, W.; Kong, Y.; Li, Q.; Yang, S.; Zhang, Y.; Wang, Y. The anticancer effects of ferulic acid is associated with induction of cell cycle arrest and autophagy in cervical cancer cells. Cancer Cell Int. 2018, 18, 102. [Google Scholar] [CrossRef] [Green Version]
  39. Karthikeyan, S.; Kanimozhi, G.; Prasad, N.R.; Mahalakshmi, R. Radiosensitizing effect of ferulic acid on human cervical carcinoma cells in vitro. Toxicol. In Vitro 2011, 25, 1366–1375. [Google Scholar] [CrossRef]
  40. You, B.R.; Moon, H.J.; Han, Y.H.; Park, W.H. Gallic acid inhibits the growth of HeLa cervical cancer cells via apoptosis and/or necrosis. Food Chem. Toxicol. 2010, 48, 1334–1340. [Google Scholar] [CrossRef]
  41. Zhao, B.; Hu, M. Gallic acid reduces cell viability, proliferation, invasion and angiogenesis in human cervical cancer cells. Oncol. Lett. 2013, 6, 1749–1755. [Google Scholar] [CrossRef] [Green Version]
  42. Alshatwi, A.A.; Ramesh, E.; Periasamy, V.S.; 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] [PubMed]
  43. Roohbakhsh, A.; Parhiz, H.; Soltani, F.; Rezaee, R.; Iranshahi, M. Molecular mechanisms behind the biological effects of hesperidin and hesperetin for the prevention of cancer and cardiovascular diseases. Life Sci. 2015, 124, 64–74. [Google Scholar] [CrossRef] [PubMed]
  44. 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]
  45. Krishnakumar, N.; Sulfikkarali, N.; RajendraPrasad, N.; Karthikeyan, S. Enhanced anticancer activity of naringenin-loaded nanoparticles in human cervical (HeLa) cancer cells. Biomed. Prev. Nutr. 2011, 1, 223–231. [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] [Green Version]
  47. Vanamala, J.; Leonardi, T.; Patil, B.S.; Taddeo, S.S.; Murphy, M.E.; Pike, L.M.; Chapkin, R.S.; Lupton, J.R.; Turner, N.D. Suppression of colon carcinogenesis by bioactive compounds in grapefruit. Carcinogenesis 2006, 27, 1257–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kim, D.I.; Lee, S.J.; Lee, S.B.; Park, K.; Kim, W.J.; Moon, S.K. Requirement for Ras/Raf/ERK pathway in naringin-induced G1 -cell-cycle arrest via p21WAF1 expression. Carcinogenesis 2008, 29, 1701–1709. [Google Scholar] [CrossRef] [Green Version]
  49. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Yang, C.S.; Landau, J.M.; Huang, M.T.; Newmark, H.L. Inhibition of carcinogenesis by dietary polyphenolic compounds. Annu. Rev. Nutr. 2001, 21, 381–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Czeczot, H.; Tudek, B.; Kusztelak, J.; Szymczyk, T.; Dobrowolska, B.; Glinkowska, G.; Malinowski, J.; Strzelecka, H. Isolation and studies of the mutagenic activity in the Ames test of flavonoids naturally occurring in medical herbs. Mutat. Res. Toxicol. 1990, 240, 209–216. [Google Scholar] [CrossRef]
  52. Del Rio, D.; Rodriguez-Mateos, A.; Spencer, J.P.E.; Tognolini, M.; Borges, G.; Crozier, A. Dietary (poly)phenolics in human health: Structures, bioavailability, and evidence of protective effects against chronic diseases. Antioxid. Redox Signal. 2013, 18, 1818–1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Rodriguez-Mateos, A.; Vauzour, D.; Krueger, C.G.; Shanmuganayagam, D.; Reed, J.; Calani, L.; Mena, P.; Del Rio, D.; Crozier, A. Bioavailability, bioactivity and impact on health of dietary flavonoids and related compounds: An update. Arch. Toxicol. 2014, 88, 1803–1853. [Google Scholar] [CrossRef] [PubMed]
  54. Brglez Mojzer, E.; Knez Hrnčič, M.; Škerget, M.; Knez, Ž.; Bren, U. Polyphenols: Extraction Methods, Antioxidative Action, Bioavailability and Anticarcinogenic Effects. Molecules 2016, 21, 901. [Google Scholar] [CrossRef] [PubMed]
  55. Zheng, P.W.; Chiang, L.C.; Lin, C.C. Apigenin induced apoptosis through p53-dependent pathway in human cervical carcinoma cells. Life Sci. 2005, 76, 1367–1379. [Google Scholar] [CrossRef]
  56. Czyz, J.; Madeja, Z.; Korohoda, W.; Hü Lser, D.F. Flavonoid apigenin inhibits motility and invasiveness of carcinoma cells in vitro. Int. J. Cancer 2004, 114, 12–18. [Google Scholar] [CrossRef]
  57. Xu, Y.; Xin, Y.; Diao, Y.; Lu, C.; Fu, J.; Luo, L.; Yin, Z. Synergistic Effects of Apigenin and Paclitaxel on Apoptosis of Cancer Cells. PLoS ONE 2011, 6, e29169. [Google Scholar] [CrossRef] [Green Version]
  58. Chuang, C.M.; Monie, A.; Wu, A.; Hung, C.F. Combination of apigenin treatment with therapeutic HPV DNA vaccination generates enhanced therapeutic antitumor effects. J. Biomed. Sci. 2009, 16, 1–11. [Google Scholar] [CrossRef] [Green Version]
  59. Lee, H.G.; Yu, K.A.; Oh, W.K.; Baeg, T.W.; Oh, H.C.; Ahn, J.S.; Jang, W.C.; Kim, J.W.; Lim, J.S.; Choe, Y.K.; et al. Inhibitory effect of jaceosidin isolated from Artemisia argyi on the function of E6 and E7 oncoproteins of HPV 16. J. Ethnopharmacol. 2005, 98, 339–343. [Google Scholar] [CrossRef]
  60. Han, H.Y.; Kim, H.J.; Jeong, S.H.; Kim, J.; Jeong, S.H.; Kim, G.C.; Hwang, D.S.; Kim, U.K.; Ryu, M.H. The Flavonoid Jaceosidin from Artemisia princeps Induces Apoptotic Cell Death and Inhibits the Akt Pathway in Oral Cancer Cells. Evid.-Based Complement. Altern. Med. 2018, 2018, 5765047. [Google Scholar] [CrossRef] [Green Version]
  61. Cherry, J.J.; Rietz, A.; Malinkevich, A.; Liu, Y.; Xie, M.; Bartolowits, M.; Davisson, V.J.; Baleja, J.D.; Androphy, E.J. Structure Based Identification and Characterization of Flavonoids That Disrupt Human Papillomavirus-16 E6 Function. PLoS ONE 2013, 8, 84506. [Google Scholar] [CrossRef]
  62. 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] [PubMed]
  63. 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] [PubMed]
  64. Hussain, S.; Ashafaq, M. Epigallocatechin-3-Gallate (EGCG): Mechanisms, perspectives and clinical applications in cervical cancer. J. Cancer Prev. Curr. Res. 2018, 9, 178–182. [Google Scholar] [CrossRef]
  65. Qiao, Y.; Cao, J.; Xie, L.; Shi, X. Cell Growth Inhibition and Gene Expression Regulation by (-)-Epigal-locatechin-3-Gallate in Human Cervical Cancer Cells. Arch. Pharm Res. 2009, 32, 1309–1315. [Google Scholar] [CrossRef] [PubMed]
  66. Sharma, C.; Nusri, Q.E.A.; Begum, S.; Javed, E.; Rizvi, T.A.; Hussain, A. (-)-Epigallocatechin-3-gallate induces apoptosis and inhibits invasion and migration of human cervical cancer cells. Asian Pac. J. Cancer Prev. 2012, 13, 4815–4822. [Google Scholar] [CrossRef] [Green Version]
  67. Zou, C.; Liu, H.; Feugang, J.M.; Hao, Z.; Chow, H.H.S.; Garcia, F. Green tea compound in chemoprevention of cervical cancer. Int. J. Gynecol. Cancer 2010, 20, 617–624. [Google Scholar] [CrossRef] [Green Version]
  68. Chen, L.; Han, W.; Geng, Y.; Su, J. A genome-wide study of DNA methylation modified by epigallocatechin-3-gallate in the CAL-27 cell line. Mol. Med. Rep. 2015, 12, 5886–5890. [Google Scholar] [CrossRef] [Green Version]
  69. Zhu, Y.; Huang, Y.; Liu, M.; Yan, Q.; Zhao, W.; Yang, P.; Gao, Q.; Wei, J.; Zhao, W.; Ma, L. Epigallocatechin gallate inhibits cell growth and regulates miRNA expression in cervical carcinoma cell lines infected with different high-risk human papillomavirus subtypes. Exp. Ther. Med. 2019, 17, 1742–1748. [Google Scholar] [CrossRef] [Green Version]
  70. Alshatwi, A.A.; Periasamy, V.S.; Athinarayanan, J.; Elango, R. Synergistic anticancer activity of dietary tea polyphenols and bleomycin hydrochloride in human cervical cancer cell: Caspase-dependent and independent apoptotic pathways. Chem. Biol. Interact. 2016, 247, 1–10. [Google Scholar] [CrossRef]
  71. Kang, T.H.; Lee, J.H.; Song, C.K.; Han, H.D.; Shin, B.C.; Pai, S.I.; Hung, C.-F.; Trimble, C.; Lim, J.-S.; Kim, T.W.; et al. Epigallocatechin-3-Gallate Enhances CD8+ T Cell–Mediated Antitumor Immunity Induced by DNA Vaccination. Cancer Res. 2007, 67, 802–811. [Google Scholar] [CrossRef] [Green Version]
  72. 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] [PubMed]
  73. Xu, W.; Liu, J.; Li, C.; Wu, H.Z.; Liu, Y.W. Kaempferol-7-O-β-d-glucoside (KG) isolated from Smilax china L. rhizome induces G2/M phase arrest and apoptosis on HeLa cells in a p53-independent manner. Cancer Lett. 2008, 264, 229–240. [Google Scholar] [CrossRef] [PubMed]
  74. 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] [PubMed] [Green Version]
  75. 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]
  76. 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]
  77. Lin, C.; Yu, Y.; Zhao, H.G.; Yang, A.; Yan, H.; Cui, Y. Combination of quercetin with radiotherapy enhances tumor radiosensitivity in vitro and in vivo. Radiother. Oncol. 2012, 104, 395–400. [Google Scholar] [CrossRef]
  78. Guo, J.M.; Kang, G.Z.; Xiao, B.X.; Liu, D.H.; Zhang, S. Effect of daidzein on cell growth, cell cycle, and telomerase activity of human cervical cancer in vitro. Int. J. Gynecol. Cancer 2004, 14, 882–888. [Google Scholar] [CrossRef]
  79. Lo, Y.L.; Wang, W. Formononetin potentiates epirubicin-induced apoptosis via ROS production in HeLa cells in vitro. Chem. Biol. Interact. 2013, 205, 188–197. [Google Scholar] [CrossRef]
  80. Yashar, C.M.; Spanos, W.J.; Taylor, D.D.; Gercel-Taylor, C. Potentiation of the radiation effect with genistein in cervical cancer cells. Gynecol. Oncol. 2005, 99, 199–205. [Google Scholar] [CrossRef]
  81. Zhang, B.; Liu, J.-Y.; Pan, J.-S.; Han, S.-P.; Yin, X.-X.; Wang, B.; Hu, G. Combined Treatment of Ionizing Radiation With Genistein on Cervical Cancer HeLa Cells. J. Pharmacol. Sci. 2006, 102, 129–135. [Google Scholar] [CrossRef] [Green Version]
  82. Shin, J.I.; Shim, J.H.; Kim, K.H.; Choi, H.S.; Kim, J.W.; Lee, H.G.; Kim, B.Y.; Park, S.N.; Park, O.J.; Yoon, D.Y. Sensitization of the apoptotic effect of γ-irradiation in genistein-pretreated CaSki cervical cancer cells. J. Microbiol. Biotechnol. 2008, 18, 523–531. [Google Scholar] [PubMed]
  83. Zheng, W.-Y.; Nie, S.-P.; Li, W.-J.; Hu, X.-J.; Xie, M.-Y. Stimulatory effects of genistein and quercetin on the proliferation of MCF-7 cells. Food Biosci. 2013, 2, 15–23. [Google Scholar] [CrossRef]
  84. Liu, H.; Du, J.; Hu, C.; Qi, H.; Wang, X.; Wang, S.; Liu, Q.; Li, Z. Delayed activation of extracellular-signal-regulated kinase 1/2 is involved in genistein- and equol-induced cell proliferation and estrogen-receptor-α-mediated transcription in MCF-7 breast cancer cells. J. Nutr. Biochem. 2010, 21, 390–396. [Google Scholar] [CrossRef] [PubMed]
  85. Qian, S.; Jin, L.-Q. The Studies of Cyanidin 3-Glucoside-Induced Apoptosis in Human Cervical Cancer Hela Cells and its Mechanism. Chin. J. Biochem. Pharm. 2008, 6, 369–373. [Google Scholar]
  86. 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. 2012, 30, 243–248. [Google Scholar] [CrossRef]
  87. 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] [Green Version]
  88. Giridharan, P.; Somasundaram, S.T.; Perumal, K.; Vishwakarma, R.A.; Karthikeyan, N.P.; Velmurugan, R.; Balakrishnan, A. Novel substituted methylenedioxy lignan suppresses proliferation of cancer cells by inhibiting telomerase and activation of c-myc and caspases leading to apoptosis. Br. J. Cancer 2002, 87, 98–105. [Google Scholar] [CrossRef]
  89. Gao, P.; Zhai, F.; Guan, L.; Zheng, J. Nordihydroguaiaretic acid inhibits growth of cervical cancer SiHa cells by upregulating p21. Oncol. Lett. 2011, 2, 123–128. [Google Scholar] [CrossRef] [Green Version]
  90. Dou, H.; Yang, S.; Hu, Y.; Xu, D.; Liu, L.; Li, X. Sesamin induces ER stress-mediated apoptosis and activates autophagy in cervical cancer cells. Life Sci. 2018, 200, 87–93. [Google Scholar] [CrossRef]
  91. Kim, Y.S.; Sull, J.W.; Sung, H.J. Suppressing effect of resveratrol on the migration and invasion of human metastatic lung and cervical cancer cells. Mol. Biol. Rep. 2012, 39, 8709–8716. [Google Scholar] [CrossRef]
  92. García-Zepeda, S.P.; García-Villa, E.; Díaz-Chávez, J.; Hernández-Pando, R.; Gariglio, P. Resveratrol induces cell death in cervical cancer cells through apoptosis and autophagy. Eur. J. Cancer Prev. 2013, 22, 577–584. [Google Scholar] [CrossRef]
  93. Zoberi, I.; Bradbury, C.M.; Curry, H.A.; Bisht, K.S.; Goswami, P.C.; Roti, J.L.; Gius, D. Radiosensitizing and anti-proliferative effects of resveratrol in two human cervical tumor cell lines. Cancer Lett. 2002, 175, 165–173. [Google Scholar] [CrossRef]
  94. Shaito, A.; Posadino, A.M.; Younes, N.; Hasan, H.; Halabi, S.; Alhababi, D.; Al-Mohannadi, A.; Abdel-Rahman, W.M.; Eid, A.H.; Nasrallah, G.K.; et al. Potential adverse effects of resveratrol: A literature review. Int. J. Mol. Sci. 2020, 21, 2084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Divya, C.S.; Pillai, M.R. Antitumor action of curcumin in human papillomavirus associated cells involves downregulation of viral oncogenes, prevention of NFkB and AP-1 translocation, and modulation of apoptosis. Mol. Carcinog. 2006, 45, 320–332. [Google Scholar] [CrossRef] [PubMed]
  96. Maher, D.M.; Bell, M.C.; O’Donnell, E.A.; Gupta, B.K.; Jaggi, M.; Chauhan, S.C. Curcumin suppresses human papillomavirus oncoproteins, restores p53, rb, and ptpn13 proteins and inhibits benzo[a]pyrene-induced upregulation of HPV E7. Mol. Carcinog. 2011, 50, 47–57. [Google Scholar] [CrossRef] [PubMed]
  97. Anto, R.J.; Mukhopadhyay, A.; Denning, K.; Aggarwal, B.B. Curcumin (diferuloylmethane) induces apoptosis through activation of caspase-8, BID cleavage and cytochrome c release: Its suppression by ectopic expression of Bcl-2 and Bcl-xl. Carcinogenesis 2002, 23, 143–150. [Google Scholar] [CrossRef] [PubMed]
  98. Paulraj, F.; Abas, F.; Lajis, N.H.; Othman, I.; Hassan, S.S.; Naidu, R. The curcumin analogue 1,5-bis(2-hydroxyphenyl)-1,4-pentadiene-3-one induces apoptosis and downregulates E6 and E7 oncogene expression in HPV16 and HPV18-infected cervical cancer cells. Molecules 2015, 20, 11830–11860. [Google Scholar] [CrossRef]
  99. Bava, S.V.; Puliappadamba, V.T.; Deepti, A.; Nair, A.; Karunagaran, D.; Anto, R.J. Sensitization of taxol-induced apoptosis by curcumin involves downregulation of nuclear factor-κB and the serine/threonine kinase Akt and is independent of tubulin polymerization. J. Biol. Chem. 2005, 280, 6301–6308. [Google Scholar] [CrossRef] [Green Version]
  100. Bava, S.V.; Sreekanth, C.N.; Thulasidasan, A.K.T.; Anto, N.P.; Cheriyan, V.T.; Puliyappadamba, V.T.; Menon, S.G.; Ravichandran, S.D.; Anto, R.J. Akt is upstream and MAPKs are downstream of NF-κB in paclitaxel-induced survival signaling events, which are downregulated by curcumin contributing to their synergism. Int. J. Biochem. Cell Biol. 2011, 43, 331–341. [Google Scholar] [CrossRef]
  101. Prusty, B.K.; Das, B.C. Constitutive activation of transcription factor AP-1 in cervical cancer and suppression of human papillomavirus (HPV) transcription and AP-1 activity in HeLa cells by curcumin. Int. J. Cancer 2005, 113, 951–960. [Google Scholar] [CrossRef]
  102. Limtrakul, P.; Chearwae, W.; Shukla, S.; Phisalphong, C.; Ambudkar, S.V. Modulation of function of three ABC drug transporters, P-glycoprotein (ABCB1), mitoxantrone resistance protein (ABCG2) and multidrug resistance protein 1 (ABCC1) by tetrahydrocurcumin, a major metabolite of curcumin. Mol. Cell. Biochem. 2007, 296, 85–95. [Google Scholar] [CrossRef] [PubMed]
  103. Javvadi, P.; Hertan, L.; Kosoff, R.; Datta, T.; Kolev, J.; Mick, R.; Tuttle, S.W.; Koumenis, C. Thioredoxin reductase-1 mediates curcumin-induced radiosensitization of squamous carcinoma cells. Cancer Res. 2010, 70, 1941–1950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Cheng, A.L.; Hsu, C.H.; Lin, J.K.; Hsu, M.M.; Ho, Y.F.; Shen, T.S.; Ko, J.Y.; Lin, J.T.; Lin, B.R.; Ming-Shiang, W.; et al. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer Res. 2001, 21, 2895–2900. [Google Scholar]
  105. Verreault, R.; Chu, J.; Mandelson, M.; Shy, K. A case-control study of diet and invasive cervical cancer. Int. J. Cancer 1989, 43, 1050–1054. [Google Scholar] [CrossRef] [PubMed]
  106. Herrero, R.; Potischman, N.; Brinton, L.A.; Reeves, W.C.; Brenes, M.M.; Tenorio, F.; de Britton, R.C.; Gaitan, E. A Case-Control Study of Nutrient Status and Invasive Cervical Cancer: I. Dietary Indicators. Am. J. Epidemiol. 1991, 134, 1335–1346. [Google Scholar] [CrossRef] [PubMed]
  107. Ghosh, C.; Baker, J.A.; Moysich, K.B.; Rivera, R.; Brasure, J.R.; McCann, S.E. Dietary intakes of selected nutrients and food groups and risk of cervical cancer. Nutr. Cancer 2008, 60, 331–341. [Google Scholar] [CrossRef]
  108. Tomita, L.Y.; Longatto Filho, A.; Costa, M.C.; Andreoli, M.A.A.; Villa, L.L.; Franco, E.L.; Cardoso, M.A. Diet and serum micronutrients in relation to cervical neoplasia and cancer among low-income Brazilian women. Int. J. Cancer 2010, 126, 703–714. [Google Scholar] [CrossRef]
  109. González, C.A.; Travier, N.; Luján-Barroso, L.; Castellsagué, X.; Bosch, F.X.; Roura, E.; Bueno-de-Mesquita, H.B.; Palli, D.; Boeing, H.; Pala, V.; et al. Dietary factors and in situ and invasive cervical cancer risk in the European prospective investigation into cancer and nutrition study. Int. J. Cancer 2011, 129, 449–459. [Google Scholar] [CrossRef]
  110. Batieha, A.M.; Armenian, H.K.; Norkus, E.P.; Morris, J.S.; Spate, V.E.; Comstock, G.W. Serum micronutrients and the subsequent risk of cervical cancer in a population-based nested case-control study. Cancer Epidemiol. Biomark. Prev. 1993, 2, 335–339. [Google Scholar]
  111. Lehtinen, M.; Luostarinen, T.; Youngman, L.D.; Anttila, T.; Dillner, J.; Hakulinen, T.; Koskela, P.; Lenner, P.; Hallmans, G. Low levels of serum vitamins A and E in blood and subsequent risk for cervical cancer: Interaction with HPV seropositivity. Nutr. Cancer 1999, 34, 229–234. [Google Scholar] [CrossRef]
  112. Zhang, Y.Y.; Lu, L.; Abliz, G.; Mijit, F. Serum carotenoid, retinol and tocopherol concentrations and risk of cervical cancer among chinese women. Asian Pac. J. Cancer Prev. 2015, 16, 2981–2986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Ratnam, D.V.; Ankola, D.D.; Bhardwaj, V.; Sahana, D.K.; Kumar, M.N.V.R. Role of antioxidants in prophylaxis and therapy: A pharmaceutical perspective. J. Control. Release 2006, 113, 189–207. [Google Scholar] [CrossRef] [PubMed]
  114. Aktepe, O.H.; Şahin, T.K.; Güner, G.; Arik, Z.; Yalçin, Ş. Lycopene sensitizes the cervical cancer cells to cisplatin via targeting nuclear factor- kappa B (NF-κB) pathway. Turk. J. Med. Sci. 2021, 51, 368–374. [Google Scholar] [CrossRef] [PubMed]
  115. Overwijk, W.W.; Wang, E.; Marincola, F.M.; Rammensee, H.G.; Restifo, N.P. Mining the mutanome: Developing highly personalized Immunotherapies based on mutational analysis of tumors. J. Immunother. Cancer 2013, 1, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Wang, Q.; Zheng, X.L.; Yang, L.; Shi, F.; Gao, L.B.; Zhong, Y.J.; Sun, H.; He, F.; Lin, Y.; Wang, X. Reactive oxygen species-mediated apoptosis contributes to chemosensitization effect of saikosaponins on cisplatin-induced cytotoxicity in cancer cells. J. Exp. Clin. Cancer Res. 2010, 29, 159. [Google Scholar] [CrossRef] [Green Version]
  117. Yuan, B.; Yang, R.; Ma, Y.; Zhou, S.; Zhang, X.; Liu, Y. A systematic review of the active saikosaponins and extracts isolated from Radix Bupleuri and their applications. Pharm. Biol. 2017, 55, 620–635. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, W.; Zhang, D.; Ma, X.; Liu, Z.; Li, F.; Wu, D. Paris saponin VII suppressed the growth of human cervical cancer Hela cells. Eur. J. Med. Res. 2014, 19, 41. [Google Scholar] [CrossRef] [Green Version]
  119. 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]
  120. Guo, J.-L.; Han, T.; Bao, L.; Li, X.-M.; Ma, J.-Q.; Tang, L.-P. Ursolic acid promotes the apoptosis of cervical cancer cells by regulating endoplasmic reticulum stress. J. Obstet. Gynaecol. Res. 2019, 45, 877–881. [Google Scholar] [CrossRef]
  121. Wang, S.; Meng, X.; Dong, Y. Ursolic acid nanoparticles inhibit cervical cancer growth in vitro and in vivo via apoptosis induction. Int. J. Oncol. 2017, 50, 1330–1340. [Google Scholar] [CrossRef] [Green Version]
  122. Zhang, Q.; Yang, D. Allicin suppresses the migration and invasion in cervical cancer cells mainly by inhibiting NRF2. Exp. Ther. Med. 2018, 17, 1523–1528. [Google Scholar] [CrossRef]
  123. Chen, D.Z.; Qi, M.; Auborn, K.J.; Carter, T.H. Indole-3-carbinol and diindolylmethane induce apoptosis of human cervical cancer cells and in murine HPV16-transgenic preneoplastic cervical epithelium. J. Nutr. 2001, 131, 3294–3302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rosen, C.A.; Woodson, G.E.; Thompson, J.W.; Hengesteg, A.P.; Bradlow, H.L. Preliminary results of the use of indole-3-carbinol for recurrent respiratory papillomatosis. Otolaryngol. Neck Surg. Off. J. Am. Acad. Otolaryngol. Neck Surg. 1998, 118, 810–815. [Google Scholar] [CrossRef]
  125. Bell, M.C.; Crowley-Nowick, P.; Bradlow, H.L.; Sepkovic, D.W.; Schmidt-Grimminger, D.; Howell, P.; Mayeaux, E.J.; Tucker, A.; Turbat-Herrera, E.A.; Mathis, J.M. Placebo-controlled trial of indole-3-carbinol in the treatment of CIN. Gynecol. Oncol. 2000, 78, 123–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Rosen, C.A.; Bryson, P.C. Indole-3-carbinol for recurrent respiratory papillomatosis: Long-term results. J. Voice 2004, 18, 248–253. [Google Scholar] [CrossRef]
  127. Del Priore, G.; Gudipudi, D.K.; Montemarano, N.; Restivo, A.M.; Malanowska-Stega, J.; Arslan, A.A. Oral diindolylmethane (DIM): Pilot evaluation of a nonsurgical treatment for cervical dysplasia. Gynecol. Oncol. 2010, 116, 464–467. [Google Scholar] [CrossRef]
  128. Castañon, A.; Tristram, A.; Mesher, D.; Powell, N.; Beer, H.; Ashman, S.; Rieck, G.; Fielder, H.; Fiander, A.; Sasieni, P. Effect of diindolylmethane supplementation on low-grade cervical cytological abnormalities: Double-blind, randomised, controlled trial. Br. J. Cancer 2012, 106, 45–52. [Google Scholar] [CrossRef]
  129. Ashrafian, L.; Sukhikh, G.; Kiselev, V.; Paltsev, M.; Drukh, V.; Kuznetsov, I.; Muyzhnek, E.; Apolikhina, I.; Andrianova, E. Double-blind randomized placebo-controlled multicenter clinical trial (phase IIa) on diindolylmethane’s efficacy and safety in the treatment of CIN: Implications for cervical cancer prevention. EPMA J. 2015, 6, 25. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, S.J.; Zheng, C.J.; Peng, C.; Zhang, H.; Jiang, Y.P.; Han, T.; Qin, L.P. Plants and cervical cancer: An overview. Expert Opin. Investig. Drugs 2013, 22, 1133–1156. [Google Scholar] [CrossRef]
  131. Cheng, Y.-M.; Tsai, C.-C.; Hsu, Y.-C. Sulforaphane, a Dietary Isothiocyanate, Induces G 2/M Arrest in Cervical Cancer Cells through CyclinB1 Downregulation and GADD45β/CDC2 Association. Int. J. Mol. Sci. Artic. 2016, 17, 1530. [Google Scholar] [CrossRef] [Green Version]
  132. Mahata, S.; Bharti, A.C.; Shukla, S.; Tyagi, A.; Husain, S.A.; 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]
  133. Saha, S.K.; Khuda-Bukhsh, A.R. Berberine alters epigenetic modifications, disrupts microtubule network, and modulates HPV-18 E6-E7 oncoproteins by targeting p53 in cervical cancer cell HeLa: A mechanistic study including molecular docking. Eur. J. Pharmacol. 2015, 744, 132–146. [Google Scholar] [CrossRef] [PubMed]
  134. Guo, Y.; Xu, L.S.; Zhang, D.; Liao, Y.P.; Wang, H.P.; Lan, Z.H.; Guan, W.J.; Liu, C.Q. Betaine effects on morphology, proliferation, and p53-induced apoptosis of hela cervical carcinoma cells in vitro. Asian Pac. J. Cancer Prev. 2015, 16, 3195–3201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Yim, N.H.; Lee, J.H.; Cho, W.K.; Yang, M.C.; Kwak, D.H.; Ma, J.Y. Decursin and decursinol angelate from Angelica gigas Nakai induce apoptosis via induction of TRAIL expression on cervical cancer cells. Eur. J. Integr. Med. 2011, 3, e299–e307. [Google Scholar] [CrossRef]
  136. Jafri, A.; Siddiqui, S.; Rais, J.; Sultan Ahmad, M.; 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]
  137. Munagala, R.; Kausar, H.; Munjal, C.; Gupta, R.C. Withaferin a induces p53-dependent apoptosis by repression of HPV oncogenes and upregulation of tumor suppressor proteins in human cervical cancer cells. Carcinogenesis 2011, 32, 1697–1705. [Google Scholar] [CrossRef] [Green Version]
  138. Cai, S.; Adila, A.; Alimu, A.; Chen, Q.; Li, Y. Glycyrrhiza uralensis Fisch L. crude polysaccharides enhance mouse immunity and immune responses induced by HPV-DNA vaccine. Chin. J. Microbiol. Immunol. 2018, 38, 774–781. [Google Scholar]
  139. Yan, Q.; Li, Y.; Jiang, Z.; Sun, Y.; Zhu, L.; Ding, Z. Antiproliferation and apoptosis of human tumor cell lines by a lectin (AMML) of Astragalus mongholicus. Phytomedicine 2009, 16, 586–593. [Google Scholar] [CrossRef]
  140. Wang, P.; Li, J.C. Trichosanthin-induced specific changes of cytoskeleton configuration were associated with the decreased expression level of actin and tubulin genes in apoptotic Hela cells. Life Sci. 2007, 81, 1130–1140. [Google Scholar] [CrossRef]
  141. Mancinelli, L.; De Angelis, P.M.; Annulli, L.; Padovini, V.; Elgjo, K.; Gianfranceschi, G.L. A class of DNA-binding peptides from wheat bud causes growth inhibition, G2 cell cycle arrest and apoptosis induction in HeLa cells. Mol. Cancer 2009, 8, 55. [Google Scholar] [CrossRef] [Green Version]
  142. Bhutia, S.K.; Mallick, S.K.; Maiti, T.K. In vitro immunostimulatory properties of Abrus lectins derived peptides in tumor bearing mice. Phytomedicine 2009, 16, 776–782. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, Q.; Tang, X.; Lu, Q.Y.; Zhang, Z.F.; 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] [PubMed] [Green Version]
  144. 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]
  145. Ahn, W.S.; Yoo, J.; Huh, S.W.; Kim, C.K.; Lee, J.M.; Namkoong, S.E.; Bae, S.M.; Lee, I.P. Protective effects of green tea extracts (polyphenon E and EGCG) on human cervical lesions. Eur. J. Cancer Prev. 2003, 12, 383–390. [Google Scholar] [CrossRef] [PubMed]
  146. Garcia, F.A.R.; Cornelison, T.; Nuño, T.; Greenspan, D.L.; Byron, J.W.; Hsu, C.-H.; Alberts, D.S.; Chow, H.-H.S. Results of a phase II randomized, double-blind, placebo-controlled trial of Polyphenon E in women with persistent high-risk HPV infection and low-grade cervical intraepithelial neoplasia. Gynecol. Oncol. 2014, 132, 377–382. [Google Scholar] [CrossRef] [Green Version]
  147. Deepika, M.S.; Thangam, R.; Sheena, T.S.; Sasirekha, R.; Sivasubramanian, S.; Babu, M.D.; Jeganathan, K.; Thirumurugan, R. A novel rutin-fucoidan complex based phytotherapy for cervical cancer through achieving enhanced bioavailability and cancer cell apoptosis. Biomed. Pharmacother. 2019, 109, 1181–1195. [Google Scholar] [CrossRef]
  148. Kwon, S.B.; Kim, M.J.; Yang, J.M.; Lee, H.P.; Hong, J.T.; Jeong, H.S.; Kim, E.S.; Yoon, D.Y. Cudrania tricuspidata stem extract induces apoptosis via the extrinsic pathway in SiHa cervical cancer cells. PLoS ONE 2016, 11, e0150235. [Google Scholar] [CrossRef]
  149. Ghanbari, A.; Le Gresley, A.; Naughton, D.; Kuhnert, N.; Sirbu, D.; Ashrafi, G.H. Biological activities of Ficus carica latex for potential therapeutics in Human Papillomavirus (HPV) related cervical cancers. Sci. Rep. 2019, 9, 1013. [Google Scholar] [CrossRef] [Green Version]
  150. Khavari, A.; Bolhassani, A.; Alizadeh, F.; Bathaie, S.Z.; Balaram, P.; Agi, E.; Vahabpour, R. Chemo-immunotherapy using saffron and its ingredients followed by E7-NT (gp96) DNA vaccine generates different anti-tumor effects against tumors expressing the E7 protein of human papillomavirus. Arch. Virol. 2015, 160, 499–508. [Google Scholar] [CrossRef]
  151. Mahata, S.; Maru, S.; Shukla, S.; Pandey, A.; Mugesh, G.; Das, B.C.; Bharti, A.C. Anticancer property of Bryophyllum pinnata (Lam.) Oken. leaf on human cervical cancer cells. BMC Complement. Altern. Med. 2012, 12, 15. [Google Scholar] [CrossRef] [Green Version]
  152. Li, G.L.; Jiang, W.; Xia, Q.; Chen, S.H.; Ge, X.R.; Gui, S.Q.; Xu, C.J. HPV E6 downregulation and apoptosis induction of human cervical cancer cells by a novel lipid-soluble extract (PE) from Pinellia pedatisecta Schott in vitro. J. Ethnopharmacol. 2010, 132, 56–64. [Google Scholar] [CrossRef]
  153. Mahata, S.; Pandey, A.; Shukla, S.; Tyagi, A.; Husain, S.A.; Das, B.C.; Bharti, A.C. Anticancer activity of Phyllanthus emblica Linn. (Indian Gooseberry): Inhibition of transcription factor ap-1 and HPV gene expression in cervical cancer cells. Nutr. Cancer 2013, 65, 88–97. [Google Scholar] [CrossRef] [PubMed]
  154. Talwar, G.P.; Dar, S.A.; Rai, M.K.; Reddy, K.V.R.; Mitra, D.; Kulkarni, S.V.; Doncel, G.F.; Buck, C.B.; Schiller, J.T.; Muralidhar, S.; et al. A novel polyherbal microbicide with inhibitory effect on bacterial, fungal and viral genital pathogens. Int. J. Antimicrob. Agents 2008, 32, 180–185. [Google Scholar] [CrossRef] [PubMed]
  155. Basu, P.; Dutta, S.; Begum, R.; Mittal, S.; Dutta, P.D.; Bharti, A.C.; Panda, C.K.; Biswas, J.; Dey, B.; Talwar, G.P.; et al. Clearance of cervical human papillomavirus infection by topical application of curcumin and curcumin containing polyherbal cream: A phase II randomized controlled study. Asian Pac. J. Cancer Prev. 2013, 14, 5753–5759. [Google Scholar] [CrossRef] [Green Version]
  156. Talwar, G.P.; Sharma, R.; Singh, S.; Das, B.C.; Bharti, A.C.; Sharma, K.; Singh, P.; Atrey, N.; Gupta, J.C. BASANT, a Polyherbal Safe Microbicide Eliminates HPV-16 in Women with Early Cervical Intraepithelial Lesions. J. Cancer Ther. 2015, 6, 1163. [Google Scholar] [CrossRef] [Green Version]
  157. Shukla, S.; Bharti, A.C.; Hussain, S.; Mahata, S.; Hedau, S.; Kailash, U.; Kashyap, V.; Bhambhani, S.; Roy, M.; Batra, S.; et al. Elimination of high-risk human papillomavirus type HPV16 infection by ‘Praneem’ polyherbal tablet in women with early cervical intraepithelial lesions. J. Cancer Res. Clin. Oncol. 2009, 135, 1701–1709. [Google Scholar] [CrossRef] [PubMed]
  158. Talwar, G.P.; Raghuvanshi, P.; Mishra, R.; Banerjee, U.; Rattan, A.; Whaley, K.J.; Zeitlin, L.; Achilles, S.L.; Barré-Sinoussi, F.; David, A.; et al. Polyherbal formulations with wide spectrum antimicrobial activity against reproductive tract infections and sexually transmitted pathogens. Am. J. Reprod. Immunol. 2000, 43, 144–151. [Google Scholar] [CrossRef]
  159. Mamgain, S.; Sharma, P.; Pathak, R.K.; Baunthiyal, M. Computer aided screening of natural compounds targeting the E6 protein of HPV using molecular docking. Bioinformation 2015, 11, 236. [Google Scholar] [CrossRef] [Green Version]
  160. Kumar, S.; Jena, L.; Mohod, K.; Daf, S.; Varma, A.K. Virtual Screening for Potential Inhibitors of High-Risk Human Papillomavirus 16 E6 Protein. Interdiscip. Sci. Comput. Life Sci. 2015, 7, 136–142. [Google Scholar] [CrossRef]
  161. Nabati, F.; Moradi, M.; Mohabatkar, H. In silico analyzing the molecular interactions of plant-derived inhibitors against E6AP, p53, and c-Myc binding sites of HPV type 16 E6 oncoprotein. Mol. Biol. Res. Commun. 2020, 9, 71–82. [Google Scholar] [CrossRef]
  162. Pennisi, M.; Russo, G.; Ravalli, S.; Pappalardo, F. Combining agent based-models and virtual screening techniques to predict the best citrus-derived vaccine adjuvants against human papilloma virus. BMC Bioinform. 2017, 18, 87–97. [Google Scholar] [CrossRef]
  163. Buyel, J.F. Plants as sources of natural and recombinant anti-cancer agents. Biotechnol. Adv. 2018, 36, 506–520. [Google Scholar] [CrossRef] [PubMed]
  164. Obembe, O.O.; Popoola, J.O.; Leelavathi, S.; Reddy, S.V. Advances in plant molecular farming. Biotechnol. Adv. 2011, 29, 210–222. [Google Scholar] [CrossRef] [PubMed]
  165. Spiegel, H.; Boes, A.; Voepel, N.; Beiss, V.; Edgue, G.; Rademacher, T.; Sack, M.; Schillberg, S.; Reimann, A.; Fischer, R. Application of a Scalable Plant Transient Gene Expression Platform for Malaria Vaccine Development. Front. Plant Sci. 2015, 6, 1169. [Google Scholar] [CrossRef] [PubMed]
  166. Commandeur, U.; Twyman, R.M.; Fischer, R. The Biosafety of Molecular Farming in Plants; CABI: Wallingford, UK, 2003. [Google Scholar] [CrossRef]
  167. Buyel, J.F.; Stöger, E.; Bortesi, L. Targeted genome editing of plants and plant cells for biomanufacturing. Transgenic Res. 2021, 30, 401–426. [Google Scholar] [CrossRef]
  168. Zhu, H.; Li, C.; Gao, C. Applications of CRISPR–Cas in agriculture and plant biotechnology. Nat. Rev. Mol. Cell Biol. 2020, 21, 661–677. [Google Scholar] [CrossRef]
  169. Kim, B.G.; Yang, S.M.; Kim, S.Y.; Cha, M.N.; Ahn, J.H. Biosynthesis and production of glycosylated flavonoids in Escherichia coli: Current state and perspectives. Appl. Microbiol. Biotechnol. 2015, 99, 2979–2988. [Google Scholar] [CrossRef]
  170. Wang, H.; Yang, Y.; Lin, L.; Zhou, W.; Liu, M.; Cheng, K.; Wang, W. Engineering Saccharomyces cerevisiae with the deletion of endogenous glucosidases for the production of flavonoid glucosides. Microb. Cell Fact. 2016, 15, 134. [Google Scholar] [CrossRef] [Green Version]
  171. Menezes, J.C.; Orlikova, B.; Morceau, F.; Diederich, M. Natural and Synthetic Flavonoids: Structure–Activity Relationship and Chemotherapeutic Potential for the Treatment of Leukemia. Crit. Rev. Food Sci. Nutr. 2016, 56, S4–S28. [Google Scholar] [CrossRef]
  172. Birchfield, A.S.; McIntosh, C.A. Metabolic engineering and synthetic biology of plant natural products—A minireview. Curr. Plant Biol. 2020, 24, 100163. [Google Scholar] [CrossRef]
  173. Lannoo, N.; Vervecken, W.; Proost, P.; Rougé, P.; Van Damme, E.J.M. Expression of the nucleocytoplasmic tobacco lectin in the yeast Pichia pastoris. Protein Expr. Purif. 2007, 53, 275–282. [Google Scholar] [CrossRef] [PubMed]
  174. Oliveira, C.; Felix, W.; Moreira, R.A.; Teixeira, J.A.; Domingues, L. Expression of frutalin, an α-d-galactose-binding jacalin-related lectin, in the yeast Pichia pastoris. Protein Expr. Purif. 2008, 60, 188–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Strasser, R. Plant protein glycosylation. Glycobiology 2016, 26, 926–939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Endo, Y.; Tsurugi, K.; Franz, H. The Site of Action of the A-Chain of Mistletoe Lectin I on Eukaryotic Ribosomes the RNA IV-Glycosidase Activity of the Protein. FEBS Lett. 1988, 231, 378–380. [Google Scholar] [CrossRef] [Green Version]
  177. Buck, C.B.; Thompson, C.D.; Roberts, J.N.; Müller, M.; Lowy, D.R.; Schiller, J.T. Carrageenan is a potent inhibitor of papillomavirus infection. PLoS Pathog. 2006, 2, e69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Roberts, J.N.; Buck, C.B.; Thompson, C.D.; Kines, R.; Bernardo, M.; Choyke, P.L.; Lowy, D.R.; Schiller, J.T. Genital transmission of HPV in a mouse model is potentiated by nonoxynol-9 and inhibited by carrageenan. Nat. Med. 2007, 13, 857–861. [Google Scholar] [CrossRef] [Green Version]
  179. Costa, L.S.; Fidelis, G.P.; Cordeiro, S.L.; Oliveira, R.M.; Sabry, D.A.; Câmara, R.B.G.; Nobre, L.T.D.B.; Costa, M.S.S.P.; Almeida-Lima, J.; Farias, E.H.C.; et al. Biological activities of sulfated polysaccharides from tropical seaweeds. Biomed. Pharmacother. 2010, 64, 21–28. [Google Scholar] [CrossRef]
  180. Wang, W.; Wang, S.X.; Guan, H.S. The antiviral activities and mechanisms of marine polysaccharides: An overview. Mar. Drugs 2012, 10, 2795–2816. [Google Scholar] [CrossRef]
  181. Stevan, F.R.; Oliveira, M.B.; Bucchi, D.F.; Noseda; Iacomini, M.; Duarte, M.E. Cytotoxic effects against HeLa cells of polysaccharides from seaweeds. J. Submicrosc. Cytol. Pathol. 2001, 33, 477–484. [Google Scholar]
  182. Wang, S.-X.; Zhang, X.-S.; Guan, H.-S.; Wang, W. Potential Anti-HPV and Related Cancer Agents from Marine Resources: An Overview. Mar. Drugs 2014, 12, 2019–2035. [Google Scholar] [CrossRef] [Green Version]
  183. Lee, S.H.; Ryu, B.; Je, J.Y.; Kim, S.K. Diethylaminoethyl chitosan induces apoptosis in HeLa cells via activation of caspase-3 and p53 expression. Carbohydr. Polym. 2011, 84, 571–578. [Google Scholar] [CrossRef]
  184. Román, J.J.M.; Del Campo, M.; Villar, J.; Paolini, F.; Curzio, G.; Venuti, A.; Jara, L.; Ferreira, J.; Murgas, P.; Lladser, A.; et al. Immunotherapeutic potential of mollusk hemocyanins in combination with human vaccine adjuvants in murine models of oral cancer. J. Immunol. Res. 2019, 2019, 7076942. [Google Scholar] [CrossRef]
  185. Karan, D.; Dubey, S.; Pirisi, L.; Nagel, A.; Pina, I.; Choo, Y.-M.; Hamann, M.T. The Marine Natural Product Manzamine A Inhibits Cervical Cancer by Targeting the SIX1 Protein. J. Nat. Prod. 2020, 83, 286–295. [Google Scholar] [CrossRef]
  186. Zhang, J.; Kang, J.; Liu, F.; Fan, C.; Li, H.; Chen, Q. Effect of the polypeptides from Meretrix meretrix Linnaeus on proliferation of cervical cancer Hela cells. J. Xiamen Univ. Nat. Sci. 2009, 5, 2019–2035. [Google Scholar]
  187. Lu, Y.; Zhang, B.Y.; Dong, Q.; Wang, B.L.; Sun, X.B. The Effects of Stichopus japonicus Acid Mucopolysaccharide on the Apoptosis of the Human Hepatocellular Carcinoma Cell Line HepG2. Am. J. Med. Sci. 2010, 339, 141–144. [Google Scholar] [CrossRef]
  188. Zhang, W.; Che, Q.; Tan, H.; Qi, X.; Li, J.; Li, D.; Gu, Q.; Zhu, T.; Liu, M. Marine Streptomyces sp. derived antimycin analogues suppress HeLa cells via depletion HPV E6/E7 mediated by ROS-dependent ubiquitin-proteasome system. Sci. Rep. 2017, 7, 42180. [Google Scholar] [CrossRef] [Green Version]
  189. Nguyen, V.-T.; Suck Lee, J.; Qian, Z.-J.; Li, Y.-X.; Kim, K.-N.; Heo, S.-J.; Jeon, Y.-J.; Park, W.S.; Choi, I.-W.; Je, J.-Y.; et al. Gliotoxin Isolated from Marine Fungus Aspergillus sp. Induces Apoptosis of Human Cervical Cancer and Chondrosarcoma Cells. Mar. Drugs 2014, 12, 69–87. [Google Scholar] [CrossRef]
  190. Wijesekara, I.; Li, Y.-X.; Vo, T.-S.; Van Ta, Q.; Ngo, D.-H.; Kim, S.-K. Induction of apoptosis in human cervical carcinoma HeLa cells by neoechinulin A from marine-derived fungus Microsporum sp. Process Biochem. 2013, 48, 68–72. [Google Scholar] [CrossRef]
  191. Ghosh, T.; Chattopadhyay, K.; Marschall, M.; Karmakar, P.; Mandal, P.; Ray, B. REVIEW Focus on antivirally active sulfated polysaccharides: From structure-activity analysis to clinical evaluation. Glycobiology 2009, 19, 2–15. [Google Scholar] [CrossRef]
  192. Vo, T.-S.; Kim, S.-K. Potential Anti-HIV Agents from Marine Resources: An Overview. Mar. Drugs 2010, 8, 2871–2892. [Google Scholar] [CrossRef] [Green Version]
  193. Lembo, D.; Donalisio, M.; Rusnati, M.; Bugatti, A.; Cornaglia, M.; Cappello, P.; Giovarelli, M.; Oreste, P.; Landolfo, S. Sulfated K5 Escherichia coli Polysaccharide Derivatives as Wide-Range Inhibitors of Genital Types of Human Papillomavirus. Antimicrob. Agents Chemother. 2008, 52, 1374–1381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Hussaarts, K.G.A.M.; Hurkmans, D.P.; Oomen-de Hoop, E.; van Harten, L.J.; Berghuis, S.; van Alphen, R.J.; Spierings, L.E.A.; van Rossum-Schornagel, Q.C.; Vastbinder, M.B.; van Schaik, R.H.N.; et al. Impact of Curcumin (with or without Piperine) on the Pharmacokinetics of Tamoxifen. Cancers 2019, 11, 403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Lian, H.; Hui, Y.; Xiaoping, T.; Wei, T.; Jiyi, X.; Xiaolan, Y. Baicalein suppresses the proliferation of human cervical cancer cells via Notch 1/Hes signaling pathway. J. Cancer Res. Ther. 2019, 15, 1216. [Google Scholar] [CrossRef] [PubMed]
  196. 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. Onco-Targets Ther. 2021, 14, 905. [Google Scholar] [CrossRef] [PubMed]
  197. Peng, Y.; Li, Q.; Li, K.; Zhao, H.; Han, Z.; Li, F.; Sun, M.; Zhang, Y. Antitumor activity of baicalein on the mice bearing U14 cervical cancer. Afr. J. Biotechnol. 2011, 10, 14169–14176. [Google Scholar] [CrossRef] [Green Version]
  198. Yang, H.; Xia, J.; Li, Y.; Cao, Y.; Tang, L.; Yu, X. Baicalein inhibits the invasion of human cervical cancer cells by inhibiting the hedgehog/Gli signaling pathway. Trop. J. Pharm. Res. 2020, 19, 115–120. [Google Scholar] [CrossRef]
  199. Bai, X.; Ma, Y.; Zhang, G. Butein suppresses cervical cancer growth through the PI3K/AKT/mTOR pathway. Oncol. Rep. 2015, 33, 3085–3092. [Google Scholar] [CrossRef] [Green Version]
  200. Tong, X.-Q.; Lv, W.; Luo, W.-Q.; Lu, W.-X.; Li, N. Treatment of human cervical cancer cells with butein leads to apoptosis and DNA damage. Oncol. Rep. 2016, 9, 11084–11089. [Google Scholar]
  201. Yang, P.-Y.; Hu, D.-N.; Kao, Y.-H.; Lin, I.-C.; Liu, F.-S. Butein induces apoptotic cell death of human cervical cancer cells. Oncol. Lett. 2018, 16, 6615–6623. [Google Scholar] [CrossRef]
  202. Zhang, D.; Sun, G.; Peng, L.; Tian, J.; Zhang, H. Calycosin inhibits viability, induces apoptosis, and suppresses invasion of cervical cancer cells by upregulating tumor suppressor miR-375. Arch. Biochem. Biophys. 2020, 691, 108478. [Google Scholar] [CrossRef]
  203. Khoo, B.Y.; Chua, S.L.; Balaram, P. Apoptotic effects of chrysin in human cancer cell lines. Int. J. Mol. Sci. 2010, 11, 2188–2199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Jung, J. Emerging Utilization of Chrysin Using Nanoscale Modification. J. Nanomater. 2016, 2016, 2894089. [Google Scholar] [CrossRef] [Green Version]
  205. Pratiwi, R.; Tunjung, W.A.S.; Rumiyati, R.; Amalia, A.R. Black Rice Bran Extracts and Fractions Containing Cyanidin 3-glucoside and Peonidin 3-glucoside Induce Apoptosis in Human Cervical Cancer Cells. Indones. J. Biotechnol. 2016, 20, 69–76. [Google Scholar] [CrossRef] [Green Version]
  206. Lee, K.; Hyun Lee, D.; Jung, Y.J.; Shin, S.Y.; Lee, Y.H. The natural flavone eupatorin induces cell cycle arrest at the G2/M phase and apoptosis in HeLa cells. Appl. Biol. Chem. 2016, 59, 193–199. [Google Scholar] [CrossRef]
  207. Abd Razak, N.; Yeap, S.K.; Alitheen, N.B.; Ho, W.Y.; Yong, C.Y.; Tan, S.W.; Tan, W.S.; Long, K. Eupatorin Suppressed Tumor Progression and Enhanced Immunity in a 4T1 Murine Breast Cancer Model. Integr. Cancer Ther. 2020, 19, 1534735420935625. [Google Scholar] [CrossRef]
  208. Wu, Z.; Zou, B.; Zhang, X.; Peng, X. Eupatilin regulates proliferation and cell cycle of cervical cancer by regulating hedgehog signalling pathway. Cell Biochem. Funct. 2020, 38, 428–435. [Google Scholar] [CrossRef]
  209. Tousi, M.S.; Sepehri, H.; Khoee, S.; Farimani, M.M.; Delphi, L.; Mansourizadeh, F. Evaluation of apoptotic effects of mPEG-b-PLGA coated iron oxide nanoparticles as a eupatorin carrier on DU-145 and LNCaP human prostate cancer cell lines. J. Pharm. Anal. 2021, 11, 108–121. [Google Scholar] [CrossRef]
  210. Chien, S.T.; Der Shi, M.; Lee, Y.C.; Te, C.C.; Shih, Y.W. Galangin, a novel dietary flavonoid, attenuates metastatic feature via PKC/ERK signaling pathway in TPA-treated liver cancer HepG2 cells. Cancer Cell Int. 2015, 15, 15. [Google Scholar] [CrossRef] [Green Version]
  211. Kumar, R.; Tiku, A.B. Galangin induces cell death by modulating the expression of glyoxalase-1 and Nrf-2 in HeLa cells. Chem. Biol. Interact. 2018, 279, 1–9. [Google Scholar] [CrossRef]
  212. Wang, K.L.; Yu, Y.C.; Hsia, S.M. Perspectives on the role of isoliquiritigenin in cancer. Cancers 2021, 13, 115. [Google Scholar] [CrossRef]
  213. Hirchaud, F.; Hermetet, F.; Ablise, M.; Fauconnet, S.; Vuitton, D.A.; Prétet, J.L.; Mougin, C. Isoliquiritigenin Induces Caspase-Dependent Apoptosis via Downregulation of HPV16 E6 Expression in Cervical Cancer Ca Ski Cells. Planta Med. 2013, 79, 1628–1635. [Google Scholar] [CrossRef] [PubMed]
  214. Wei, J.; Su, H.; Bi, Y.; Li, J.; Feng, L.; Sheng, W. Anti-proliferative effect of isorhamnetin on HeLa cells through inducing G2/M cell cycle arrest. Exp. Ther. Med. 2018, 15, 3917–3923. [Google Scholar] [CrossRef] [PubMed]
  215. Gong, G.; Guan, Y.Y.; Zhang, Z.L.; Rahman, K.; Wang, S.J.; Zhou, S.; Luan, X.; Zhang, H. Isorhamnetin: A review of pharmacological effects. Biomed. Pharmacother. 2020, 128, 110301. [Google Scholar] [CrossRef] [PubMed]
  216. Park, C.; Cha, H.J.; Choi, E.O.; Lee, H.; Hwang-Bo, H.; Ji, S.Y.; Kim, M.Y.; Kim, S.Y.; Hong, S.H.; Cheong, J.H.; et al. Isorhamnetin induces cell cycle arrest and apoptosis via reactive oxygen species-mediated AMP-activated protein kinase signaling pathway activation in human bladder cancer cells. Cancers 2019, 11, 1494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Cai, F.; Zhang, Y.; Li, J.; Huang, S.; Gao, R. Isorhamnetin inhibited the proliferation and metastasis of androgen-independent prostate cancer cells by targeting the mitochondrion-dependent intrinsic apoptotic and PI3K/Akt/mTOR pathway. Biosci. Rep. 2020, 40, BSR20192826. [Google Scholar] [CrossRef] [Green Version]
  218. Zhao, X.; Zhu, W.; Zhang, R.; Zhang, M.; Zhao, J.; Hou, J.; Zhang, W. Targeted juglone blocks the invasion and metastasis of HPV-positive cervical cancer cells. J. Pharmacol. Sci. 2019, 140, 211–217. [Google Scholar] [CrossRef]
  219. Lu, Z.; Chen, H.; Zheng, X.M.; Chen, M.L. Experimental study on the apoptosis of cervical cancer Hela cells induced by juglone through c-Jun N-terminal kinase/c-Jun pathway. Asian Pac. J. Trop. Med. 2017, 10, 572–575. [Google Scholar] [CrossRef]
  220. Zhang, Q.; Zhang, F.; Thakur, K.; Wang, J.; Wang, H.; Hu, F.; Zhang, J.G.; Wei, Z.J. Molecular mechanism of anti-cancerous potential of Morin extracted from mulberry in Hela cells. Food Chem. Toxicol. 2018, 112, 466–475. [Google Scholar] [CrossRef]
  221. Solairaja, S.; Andrabi, M.Q.; Dunna, N.R.; Venkatabalasubramanian, S. Overview of Morin and Its Complementary Role as an Adjuvant for Anticancer Agents. Nutr. Cancer 2021, 73, 927–942. [Google Scholar] [CrossRef]
  222. Yi, J.L.; Shi, S.; Shen, Y.L.; Wang, L.; Chen, H.Y.; Zhu, J.; Ding, Y. Myricetin and methyl eugenol combination enhances the anticancer activity, cell cycle arrest and apoptosis induction of cis-platin against HeLa cervical cancer cell lines. Int. J. Clin. Exp. Pathol. 2015, 8, 1116–1127. [Google Scholar]
  223. Devi, K.P.; Rajavel, T.; Habtemariam, S.; Nabavi, S.F.; Nabavi, S.M. Molecular mechanisms underlying anticancer effects of myricetin. Life Sci. 2015, 142, 19–25. [Google Scholar] [CrossRef] [PubMed]
  224. Taheri, Y.; Suleria, H.A.R.; Martins, N.; Sytar, O.; Beyatli, A.; Yeskaliyeva, B.; Seitimova, G.; Salehi, B.; Semwal, P.; Painuli, S.; et al. Myricetin bioactive effects: Moving from preclinical evidence to potential clinical applications. BMC Complement. Med. Ther. 2020, 20, 241. [Google Scholar] [CrossRef] [PubMed]
  225. Sheng, S.Q.; Yu, L.Y.; Zhou, X.W.; Pan, H.Y.; Hu, F.Y.; Liu, J.L. Paeonol prevents migration and invasion, and promotes apoptosis of cervical cancer cells by inhibiting 5-lipoxygenase. Mol. Med. Rep. 2021, 23, 401. [Google Scholar] [CrossRef] [PubMed]
  226. Du, J.; Song, D.; Li, J.; Li, Y.; Li, B.; Li, L. Paeonol triggers apoptosis in HeLa cervical cancer cells: The role of mitochondria-related caspase pathway. Psychopharmacol. 2021. [Google Scholar] [CrossRef]
  227. Hsiao, Y.H.; Hsieh, M.J.; Yang, S.F.; Chen, S.P.; Tsai, W.C.; Chen, P.N. Phloretin suppresses metastasis by targeting protease and inhibits cancer stemness and angiogenesis in human cervical cancer cells. Phytomedicine 2019, 62, 152964. [Google Scholar] [CrossRef]
  228. Lagoa, R.; Silva, J.; Rodrigues, J.R.; Bishayee, A. Advances in phytochemical delivery systems for improved anticancer activity. Biotechnol. Adv. 2020, 38, 107381. [Google Scholar] [CrossRef]
  229. Chatterjee, K.; Mukherjee, S.; Vanmanen, J.; Banerjee, P.; Fata, J.E. Dietary polyphenols, resveratrol and pterostilbene exhibit antitumor activity on an HPV E6-positive cervical cancer model: An in vitro and in vivo analysis. Front. Oncol. 2019, 9, 352. [Google Scholar] [CrossRef] [Green Version]
  230. 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]
  231. Zhang, L.; Chinnathambi, A.; Alharbi, S.A.; Veeraraghavan, V.P.; Mohan, S.K.; Zhang, G. Punicalagin promotes the apoptosis in human cervical cancer (ME-180) cells through mitochondrial pathway and by inhibiting the NF-kB signaling pathway. Saudi J. Biol. Sci. 2020, 27, 1100–1106. [Google Scholar] [CrossRef]
  232. Jing, Z.; Fei, W.; Zhou, J.; Zhang, L.; Chen, L.; Zhang, X.; Liang, X.; Xie, J.; Fang, Y.; Sui, X.; et al. Salvianolic acid B, a novel autophagy inducer, exerts antitumor activity as a single agent in colorectal cancer cells. Oncotarget 2016, 7, 61509–61519. [Google Scholar] [CrossRef]
  233. Mears, B.; Shiao, F.; Toney, N.; Chen, C.-M.; Choi, K.; Chow, C.; Navin, M.; Shiau, H. Salvianolic Acid B Inhibits Growth of Cervical Cancer Cells In Vitro via Induction of Apoptosis through the Extrinsic Pathway; University of Maryland: College Park, MD, USA, 2014. [Google Scholar]
  234. Tian, Q.; Wang, L.; Sun, X.; Zeng, F.; Pan, Q.; Xue, M. Scopoletin exerts anticancer effects on human cervical cancer cell lines by triggering apoptosis, cell cycle arrest, inhibition of cell invasion and PI3K/AKT signalling pathway. J. BUON 2019, 24, 997–1002. [Google Scholar]
  235. Yong, W.K.; Abd Malek, S.N. Xanthohumol induces growth inhibition and apoptosis in Ca Ski human cervical cancer cells. Evid.-Based Complement. Altern. Med. 2015, 2015, 921306. [Google Scholar] [CrossRef] [PubMed]
  236. Jiang, C.H.; Sun, T.L.; Xiang, D.X.; Wei, S.S.; Li, W.Q. Anticancer activity and mechanism of xanthohumol: A prenylated flavonoid from hops (Humulus lupulus L.). Front. Pharmacol. 2018, 9, 530. [Google Scholar] [CrossRef] [PubMed]
  237. Su, K.; Wang, C.F.; Zhang, Y.; Cai, Y.J.; Zhang, Y.Y.; Zhao, Q. The inhibitory effects of carnosic acid on cervical cancer cells growth by promoting apoptosis via ROS-regulated signaling pathway. Biomed. Pharmacother. 2016, 82, 180–191. [Google Scholar] [CrossRef] [PubMed]
  238. Li, J.; Li, Q.W.; Gao, D.W.; Han, Z.S.; Lu, W.Z. Antitumor and immunomodulating effects of polysaccharides isolated from Solanum nigrum Linne. Phyther. Res. 2009, 23, 1524–1530. [Google Scholar] [CrossRef]
Figure 1. Sources (plants, marine environment, bacteria) and nature (purified compounds and their combinations, fractions, extracts, mixtures) of formulations that have been investigated against HPV.
Figure 1. Sources (plants, marine environment, bacteria) and nature (purified compounds and their combinations, fractions, extracts, mixtures) of formulations that have been investigated against HPV.
Jcm 11 01465 g001
Figure 2. Phytochemicals investigated against HPV.
Figure 2. Phytochemicals investigated against HPV.
Jcm 11 01465 g002
Figure 3. The plant molecular farming approach. Through stable or transient expression of recombinant proteins, either exogenous proteins and metabolites accumulation or improvement of biosynthesis pathways can be achieved (NPBT: new plant breeding techniques; Crispr-Cas: clustered regularly interspaced short palindromic repeats–Crispr-associated protein).
Figure 3. The plant molecular farming approach. Through stable or transient expression of recombinant proteins, either exogenous proteins and metabolites accumulation or improvement of biosynthesis pathways can be achieved (NPBT: new plant breeding techniques; Crispr-Cas: clustered regularly interspaced short palindromic repeats–Crispr-associated protein).
Jcm 11 01465 g003
Table 1. Literature search details.
Table 1. Literature search details.
Database/Platform:MEDLINE (PubMed)
Date coverage:From 1995 to present
Library:National Library of Medicine, National Center for Biotechnology Information
Date of last search:20 October 2021
Limits:In: “Article Title, Abstract, Keywords”
Published: “All years” to “Present”
Document type: “All”
Subject Areas: All checked (default)
Search query:HPV disease + natural compounds + bioactives + phytochemicals + combination treatments + therapy + infection + name of compounds/class of compounds
Number of hits:22
Database/Platform:Science Direct (Elsevier)
Date coverage:From 1998 to present
Library:Free access
Date of last search:20 October 2021
Limits:In: “Article Title, Abstract, Keywords”
Published: “All years” to “Present”
Document type: “All”
Subject Areas: All checked (default)
Search query:HPV disease + natural compounds + bioactives + phytochemicals + combination treatments + therapy + infection + name of compounds/class of compounds
Number of hits:2870
Database/Platform:Scopus (Elsevier)
Date coverage:From 1995 to present
Library:Free access
Date of last search:20 October 2021
Limits:In: “Article Title, Abstract, Keywords”
Published: “All years” to “Present”
Document type: “All”
Subject Areas: All checked (default)
Search query:HPV disease + natural compounds + bioactives + phytochemicals + combination treatments + therapy + infection + name of compounds/class of compounds
Number of hits:20
Database/Platform:Google Scholar
Date coverage:From 1995 to present
Library:Free access
Date of last search:20 October 2021
Limits:In: “Article Title, Abstract, Keywords”
Published: “All years” to “Present”
Document type: “All”
Subject Areas: All checked (default)
Search query:HPV disease + natural compounds + bioactives + phytochemicals + combination treatments + therapy + infection + name of compounds/class of compounds
Number of hits:1330
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Massa, S.; Pagliarello, R.; Paolini, F.; Venuti, A. Natural Bioactives: Back to the Future in the Fight against Human Papillomavirus? A Narrative Review. J. Clin. Med. 2022, 11, 1465. https://doi.org/10.3390/jcm11051465

AMA Style

Massa S, Pagliarello R, Paolini F, Venuti A. Natural Bioactives: Back to the Future in the Fight against Human Papillomavirus? A Narrative Review. Journal of Clinical Medicine. 2022; 11(5):1465. https://doi.org/10.3390/jcm11051465

Chicago/Turabian Style

Massa, Silvia, Riccardo Pagliarello, Francesca Paolini, and Aldo Venuti. 2022. "Natural Bioactives: Back to the Future in the Fight against Human Papillomavirus? A Narrative Review" Journal of Clinical Medicine 11, no. 5: 1465. https://doi.org/10.3390/jcm11051465

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