EMAN RESEARCH PUBLISHING | <p>Phytochemicals in Cancer with Special Emphasis on Ovarian Cancer</p>
Inflammation Cancer Angiogenesis Biology and Therapeutics | Impact 0.1 (CiteScore) | Online ISSN  2207-872X
REVIEWS   (Open Access)

Phytochemicals in Cancer with Special Emphasis on Ovarian Cancer

Meher U Nessa1,2, Md S Anwar1, Fazlul Huq3*

+ Author Affiliations

Journal of Angiotherapy 4 (1) 156-166 https://doi.org/10.25163/angiotherapy.41209822608080720

Submitted: 26 May 2020 Revised: 30 June 2020  Published: 08 July 2020 


Abstract

Ovarian cancer is one of the most prevalent malignancies and is the deadliest among all gynecologic cancers. Although tremendous progress has been made on cancer biology and tumour treatment, knowledge on the mechanisms of cancer development including that for ovarian cancer remains incomplete. Thus, in addition to cancer prevention it is urgent to develop effective therapeutic modalities even for the advanced stage and drug-resistant forms of the disease with ovarian cancer being no exception. Multistep tumourigenesis is activated by various environmental carcinogens, inflammatory agents and tumour promoters. Carcinogens modulate the transcription factors, anti-apoptotic and pro-apoptotic proteins, protein kinases, cell cycle proteins, cell adhesion molecules and growth factor signaling pathways and thus produce malignancies. Phytochemicals exert their antitumour effects by numerous signaling pathways which in turn affect multiple steps in the various cellular pathways leading to tumourigenesis. Antitumour active phytochemicals have cytotoxic action in all ovarian cancer cell lines with comparable or greater activity in the resistant cell lines than in the parent cell line. As phytochemicals have been a part of human diet without toxic effects, they are likely to protect normal cells and tissues caused by the direct and bystander effects of platinum drugs. Therefore, it is quite logical to assume that phytochemicals possessing both cancer preventive and therapeutic attributes, can be ideal candidates in cancer prevention and synergistic outcomes from their combination with targeted therapy.

Key Words: Ovarian cancer, Phytochemicals, Cell signaling pathways, Chemoprevention, Targeted therapy

Introduction

GO

Cancer is one of the most dreaded diseases of our time that invokes death sentence in many minds. Ovarian cancer is one of the most prevalent malignancies and is the fifth leading cause of cancer-related death among women, being the deadliest of all gynecologic cancers (Anon, 2019). The treatment of ovarian cancer remains a major challenge. Despite much research effort, improvement in surgical techniques, and development of an ever expanding number of chemotherapeutic agents, it is still the most lethal gynaecological tumour and the cause of considerable morbidity for the suffers (Marsden et al., 2000). Like other cancer, various treatments available for ovarian cancer often face two main problems namely development of resistance associated with prolonged use and presence of severe side effects. Although tremendous progress has been made on cancer biology and tumour treatment, knowledge on the mechanisms of cancer development remains incomplete. This is the reason why the means of treatment against cancer are far from being ideal and are often met with failure. Thus, in addition to cancer prevention there is an urgent need for the development of effective therapeutic modalities even for advanced stage and the drug-resistant forms. Recently, there has been a growing interest in the use of dietary chemopreventive agents (such as phytochemicals) which can prevent cancer initiation, suppress proliferation and induce apoptosis in tumour cells.  In this review the mechanism of cancer development and actions of antitumour active phytochemicals will be discussed briefly.

Materials and Methods

GO

For this review relevant literatures were collected from the major scientific databases including Pubmed, Science direct, Medline and Google scholar to find out mechanisms and development of cancer, molecular mechanisms of anti tumour activities of phytochemicals. Some articles were cited through cross citations from other publications or by directly accessing the journals’ web-site. More emphasis was given to the literatures published in recent years. The keyword combinations for the search included: causes, mechanisms and development of ovarian cancer as well as other cancers, tumour active phytochemicals and their mechanisms of actions, drug resistance and phytochemicals. Additional information was assimilated by using some other keywords such as cancer development and cell cycle control, cell proliferation, apoptosis, oxidative stress, cancer stem cell. A total of 182 research articles reporting on the development of cancer, drug resistance, use of phytochemicals in prevention and treatment of cancer are recovered and presented in this review. Cytotoxic action of phytochemicals on various ovarian cancer cells are collected from the laboratory results of the Cancer Research Team, Discipline of Biomedical Science, The university of Sydney, Australia.

Ovarian Cancer: A short overview

GO

Among ovarian cancer, the most common (90% of cases) is epithelial type which arises from the cells on the outside of the ovary; about 4% of cases are the germ cell type that arises from the cells which produce eggs; and the rests are stromal type arising from supporting tissues within the ovary (Council, 2020). The World Health Organization has categorized epithelial ovarian carcinoma according to the predominant epithelial cell type, which are serous carcinoma, endometrioid carcinoma, mucinous carcinoma and clear-cell carcinoma of the ovary. Number of genetic and epigenetic changes lead to ovarian carcinoma cell transformation (Lengyel, 2010). Various studies confirmed that genes expressed in different ovarian carcinomas are concordantly expressed in the normal tissues they resemble histologically (Marquez et al., 2005). It is also evident that ovarian tumorigenesis can progress either along a stepwise mutation process from a slow growing borderline tumor to a well-differentiated carcinoma known as type I or encompasses as a genetically unstable high-grade serous carcinoma that metastasizes rapidly known as type II (Lengyel, 2010).

Mechanisms of Cancer Development

GO

Causes of Cancer: Cancer development is believed to be a multiple step process including initiation, promotion and progression (Surh, 2003). Very recent and convincing hypothesis is that inflammation contributes to every step of carcinogenesis, including tumour initiation, promotion and progression (Grivennikov et al., 2010). Components of the inflammatory pathway, including free radicals, cytokines, NF-?B, STAT-3, iNOS, COX-2, prostaglandins and VEGF have been shown to contribute to the development of various malignancies including ovarian cancer (Seo et al., 2004). Aberrant expression of microRNAs (miRNAs/miRs) are involved in the development and progression of ovarian cancer supported by the observation that downregulation of miR-183 markedly inhibited cell proliferation, migration and invasion, and promoted apoptosis in ovarian cancer cells (Zhou et al., 2019).

One strong risk factor found in epidemiologic studies is a positive family history (Amos and Struewing, 1993, Schuijer et al., 2003). This might be the results of homeostatic imbalances that can happen due to factors such as genetic mutation and if not corrected by genes such as tumour suppressor genes, will lead to the development of cancerous cells. For example, a woman who carries the mutated tumour suppressor genes BRCA1 and BRCA2 is at a higher risk of developing ovarian and breast cancers especially having the history of either breast or ovarian cancer (Ford et al., 1998, Narod, 2002).

Endometriosis is another risk factor for ovarian cancer.  Some epidemiological studies and some historical pathological observations show that both clear cell ovarian and endometrioid ovarian carcinomas may arise from endometriosis. Other neoplasms such as seromucinous borderline, low-grade serous ovarian carcinomas, adenosarcomas and endometrial stromal sarcomas may also arise from endometriosis (Dawson et al., 2018).

Ageing is also a major risk factor for the development of cancer. This is most likely due to the inevitable time-dependent decline in physiological organ function and the decline of cellular repair mechanisms (Aunan et al., 2017).

One very prominent hypothesis for ovarian carcinogenesis is ovulation hypothesis, which relates ovarian cancer risk to continual ovulation (Fathalla, 1971). In support of the hypothesis it can be cited that extensive epidemiologic evidence indicates that oral contraceptive use, multiple pregnancies, and prolonged breast-feeding can decrease the cancer risk (Fathalla, 1971, Kim et al., 2011).

Cancer development: Cancer is a disease in which cells divide without control and invade other tissues. Cancer cells can spread to other parts of the body through the blood and lymph systems. Autonomy in growth signals, insensitivity to growth inhibitory signals, avoidance of apoptosis, unlimited replicative potential, persistent angiogenesis, tissue invasion and metastasis make cancer cells distinct from normal cells (Hanahan and Weinberg, 2000). For understanding the underlying mechanisms of cancer development and to plan for proper treatment method, a brief discussion on cell proliferation and cell death and other important characteristics for the management and progression of cancer are considered further. 

Cell proliferation and cell death: Cell proliferation and cell death are two diametrically opposed cellular fates that are coupled at various levels through individual molecular players (Evan and Vousden, 2001, Lowe et al., 2004). In normal cells growth stimulating and growth limiting signals are finely controlled and very well balanced, such that cell proliferation occurs only when it is required. Whereas in tumour cells this balance gets disrupted and continued cell propagation takes place (Hahn and Weinberg, 2002). As apoptosis is the major cause of cell death, agents that trigger apoptosis/cell death, could be the most promising candidates as therapeutics for cancer (Ravindran et al., 2009). In addition to angiogenesis, metastasis and suppression of apoptosis, uncontrolled proliferation of cancer cells also lies at the heart of the disease. Therefore, clear knowledge on cell proliferation and its control is a must to find out the therapeutic targets of successful tumour regression.

Cell cycle control points: Advances in understanding of the cell cycle machinery during the last few years have demonstrated that disruption of normal cell cycle control is a hallmark of human cancer (Sa and Das, 2008). Cyclin dependent kinases (CDKs) are highly regulated family of enzymes that remain at the heart of the regulatory apparatus during the cell cycle progression (Norbury and Nurse, 1992, Hartwell and Kastan, 1994, Nurse et al., 1998, Park and Lee, 2003). CDKs (CDK4, CDK6, CDK2, and CDC2) and cyclins associates (cyclin D, E, A, and B) are positive regulators that induce cell cycle progression from G1 to mitosis whereas important negative regulators such as cyclin dependent kinase inhibitors (CKIs) act as brakes to stop the cell cycle progression in response to regulatory signals. By direct association with CDK, CKIs can negatively regulate CDK activity (Park and Lee, 2003). Two families of CKIs have been characterized according to their structures and CDK targets are INK and CIP/KIP family. The four members of the INK family, INK4A (p16), INK4B (p15), INK4C (p18), and INK4D (p19), exclusively bind to CDK4 and CDK6, preventing their association with D-type cyclins. The three members of the CIP/KIP family, CIP1 (p21), KIP1 (p27), and KIP2 (p57), form heterotrimeric complexes with the G1/S CDKs i.e. bind to both cyclin and CDK subunits, inhibit cyclin E- and A-dependent kinases but act as positive regulators of cyclin D-dependent kinases. CKIs come into play in response to different cellular processes (Sherr and Roberts, 1999, Sherr, 2000). For instance, the KIP1 levels are generally high in quiescent cells. CIP1 is one of the effectors of tumour suppressor p53 that is important in checkpoint associated with DNA damage (Park and Lee, 2003).

As applied to cell cycle control, the critical event remains the restriction point. After passing this point, the cell is irreversibly committed to the next phase of the cell cycle. The primary substrates of CDK4/6 and CDK2 in G1 progression are members of the retinoblastoma protein family pRB, p107, and p130, and are the as negative regulators at the restriction point (Morgan, 1997, Adams, 2001). In mammalian cell, E2F family of transcription factors plays key roles in regulation of cell cycle and DNA synthesis. E2F1, 2 and E2F3a act as the activators while and E2F3b, E2F4-8 act as suppressors (Harbour and Dean, 2000, Sozzani et al., 2006).

Possible targets for therapeutic intervention have been revealed from loss of cell cycle regulation in cancer. It is thought that restoration of proper restriction point control would allow cancer cells to return to the state of quiescence. As phytochemicals have specific cellular targets even in cell cycle component, the use of phytochemicals could be a promising strategy for the treatment of cancer.

Apoptosis: The ability of cancer cells to evade apoptosis (programmed cell death) is a major characteristic that enables them to undergo uncontrolled growth. The efficiency of treatment depends on the successful stimulation of apoptosis in cancer cells (Melet et al., 2008). For example, activation of the JNK signaling pathway is frequently observed in apoptosis. A number of apoptotic molecules including p53, c-Myc, Bcl-2 and Bcl-xL are prime targets for JNK-mediated phosphorylation which are involved in regulation of cytochrome c release, the key event in caspases activation (Fuchs et al., 1998, Fuchs et al., 1998, Noguchi et al., 1999, Yamamoto et al., 1999, Tournier et al., 2000). Another very important pathway is Akt pathway. Akt-mediated phosphorylation can impact on the transcriptional regulation of apoptosis. For example, Akt-mediated phosphorylation and inactivation of forkhead (FKHRL1) serve to limit transcription of its target genes including FasL, IGF-BP1 and Bim, all of which function to promote apoptosis (Kasibhatla and Tseng, 2003). In contrast, Akt is able to accelerate the degradation of I?B, and thus potentiate the activity of NF-?B, which in turn accelerates the expression of its target genes the anti-apoptotic Bcl-2 protein A1, TNF receptor-associated factors and the caspase inhibitors  (Kane et al., 1999, Zong et al., 1999, Kasibhatla and Tseng, 2003). Thus, successful treatment also depends on the suppression of this pro-survival Akt pathway.

Abnormalities in apoptotic pathways are believed to be associated with reduced activity of pro-apoptotic or tumour suppressor genes or proteins or increased expressions of oncogenes or anti-apoptotic proteins. As the development of tumours arises as a consequence of dysregulated proliferation and suppression of apoptosis, each primary defects provides an obvious opportunity for clinical intervention (Kasibhatla and Tseng, 2003). It is hoped that greater insights into the field of apoptosis and cancer will uncover new and effective strategies to tackle the complexity of tumour chemoresistance.

Oxidative Stress: Although cancer cells have many adaptive mechanisms to minimize the effect of oxidative damage and baseline or even a controlled reactive oxygen species (ROS) levels can produce a pro-survival effect, excessive levels of ROS can disrupt redox homeostasis and hence can affect cell death or survival which ultimately cause cellular harm. It does so either by irreversibly damaging cellular macromolecules including DNA, carbohydrates, protein, and lipids (Dalle-Donne et al., 2003) or by modulating redox-sensitive signaling proteins at the levels of transduction or transcriptional regulation (or both), thus trigger apoptotic cell death (Trachootham et al., 2008). Moreover, modifications induced by ROS would target several other proteins such as NF-?B, AP-1, HRas, MAPK, IP3 kinase, PKC-e, Ras, p53, HIF-1, ASK-1, Bcl-2, caspases, JNK, and p38 MAPK that play key roles in cell death and survival (England and Cotter, 2005).

The overall redox status is the net result of ROS generation and elimination. This means compounds that heighten ROS generation or suppress its elimination can favour ROS accumulation and hence induce damage to cancer cells or cause its death (Trachootham et al., 2009). Compounds that promote ROS generation, i.e., mitochondrial electron transport chain modulators, redox-cycling compounds, or that disrupt antioxidant defenses, i.e., GSH depleting agents, and inhibitors of super oxide dismutase (SOD), and catalase could selectively sensitize cancer cells to overbalanced ROS so as to cause cell death (Barbieri et al., 1994, Trachootham et al., 2009, Gibellini et al., 2010).

Drug resistance is a major problem in cancer chemotherapy affecting the clinical outcome. Among the various mechanisms of resistance, phase II detoxification system involving glutathione is a major factor. Besides high concentration of GSH and increased expression of GSTs, GSH-transporters are a common feature of transformed cells that, in turn, are associated with high resistance to chemotherapeutic agents (Yang et al., 2006, Singh et al., 2012).  Manipulation of intracellular GSH level using compounds that stimulate ROS synthesis or compounds that inhibit synthesis of GSH can be used to increase the sensitivity of different tumour cell lines to therapy and thus selective differential chemotherapy responses of normal versus tumour cells is possible (Williamson et al., 1982). Although different cells respond differently to oxidative stress inducing therapies (Mattson David et al., 2009), it is found that manipulation of intracellular oxidant status of tumour cells can be clinically useful. Increasing ROS or decreasing free radical scavengers such as GSH thus can be a therapeutic strategy for overcoming resistance. 

Drug resistance: Considering that the cytotoxic outcome of chemotherapy is a multifaceted process, relating to drug entry into cells to the final stages of apoptosis, it follows that intracellular events interfering with any of these processes will inhibit apoptosis and lead to drug resistance. Resistance mechanisms, therefore, arise as a consequence of ample intracellular modifications including changes in cellular uptake, drug efflux, increased detoxification, inhibition of apoptosis and increased DNA repair (Siddik, 2003).

Cisplatin is a platinum based anticancer drug commonly used to treat various cancers including ovarian, testicular and head and neck cancers (Crul et al., 2002, Boulikas and Vougiouka, 2004). However, acquired resistance remains one of the problems associated with the use of cisplatin (Binju et al., 2019). Of the various mechanisms of platinum resistance, one is associated with reduced cellular accumulation (Garmann et al., 2008). Since reduced accumulation can be shown over a wide range of cisplatin concentrations, it is logical to think that resistance occurs as a result of changes in passive drug diffusion (Kelland, 2000), reduction of energy-dependent active transport involving Na+K+-ATPase or a gated ion channel (Kishimoto et al., 2006), overexpression of Enhancer of zeste homolog 2 (EZH2) (Binju et al., 2019) and repression of high-affinity copper transporter CTR1 located at the plasma membrane. In addition to reduced cellular accumulation of cisplatin, an active efflux pump for cisplatin in some cases also associates with cisplatin resistance (Komatsu et al., 2000). Overexpression of adenosine triphosphate (ATP7A and ATP7B)(Holzer and Howell, 2006, Holzer et al., 2006) and  ATP-binding cassette, sub-family C2 (ABCC2) (Surowiak et al., 2006) is found to cause reduced accumulation of cisplatin due to increased efflux of the drug out of the cell. Among the various mechanisms, phase II detoxification system involving glutathione is believed to a major factor of resistance which is supported by increased level of glutathione (GSH) and glutathione-S-transferases (GSTs)(Singh et al., 2012). Over-expression of the MRP/GS-X pump for excretion of GSH conjugates is also observed in cisplatin-resistant human leukemia HL-60 cells (Ishikawa et al., 1996). Other efflux proteins such as MDR1 and MRP1 have also been found to couple with sequestration of cisplatin and other platinum drugs (Samimi et al., 2004).

The bulky DNA adducts generated by platinum drugs are mainly repaired by the nucleotide excision repair (NER) pathway composed of at least 17 different proteins. Upregulation of only a few rate-limiting proteins is necessary to increase the excision repair capacity in resistant tumour cells (Siddik, 2003, Rocha et al., 2018). Increases in the excision repair cross-complementing ERCC1 or ERCC1/XPF complexes and over-expression of the NER-related XPA gene contributing to enhanced repair, are also observed in resistant cells (Wang et al., 2011, McNeil and Melton, 2012). In addition, mismatch repair (MMR) complex maintains the integrity of the genome through repair of DNA mismatch lesions, but when it fails to do so would activate the apoptotic signal (Galluzzi et al., 2011). Downregulation or mutations in MMR genes hMLH1, hMSH2 and hMSH6 are observed consistently in resistant cells thus fail to activate the apoptotic signal (Rosell et al., 2003, Siddik, 2003, McNeil and Melton, 2012).

Cancer stem cells:  In recent times, it has been anticipated that cancer stem cells (CSCs) are responsible for the cellular heterogeneity of the tumour, resistance to therapy, self-restoration and unrestricted spread of the disease (Solomon et al., 2008, Chan et al., 2018). CSCs possess characteristics associated with normal stem cells, specifically the ability to give rise to all cell types found in a particular cancer sample. They may generate tumours through the stem cell processes of self-renewal and differentiation into multiple cell types (Zhan et al., 2013).

Although chemotherapy can reduce tumour mass, an aggressive population of CSCs within the tumour may be capable of resisting chemotherapeutic drugs and are believed to be an underlying cause for tumour recurrence and metastasis (Brian T. Kawasaki and Farrar, 2008). Development of specific therapies targeted at CSCs holds hope for improvement of survival and quality of life of cancer patients, especially for sufferers of metastatic disease. Elimination of ovarian CSCs has been challenging in part due to heterogeneity of tumour. CSCs are resistant to conventional anti-proliferative drugs. It has been suggested that conventional chemotherapies may only kill differentiated or differentiating cells (which form the bulk of the tumour) but not the CSCs (Chan et al., 2018). In order to be cured, it is imperative that CSCs must be eliminated by the cancer therapy. Combination treatments or phytochemicals alone due to their multiple specific actions can target both differentiated and un-differentiated cells thus providing a new direction to cancer treatment in the future (Valent et al., 2012, Zhan et al., 2013).

Mechanisms of anti-tumour action of phytochemicals

GO

From the above discussion, it is clear that tumourigenesis is a multistep process that can be activated by any of various environmental carcinogens, inflammatory agents and tumour promoters. These carcinogens bring about changes in transcription factors, anti-apoptotic proteins, pro-apoptotic proteins, protein kinases, cell cycle proteins, cell adhesion molecules and growth factor signaling pathways and thus lead to malignancies. 

Despite significant innovations in cancer therapy over the past several decades, the global burden of cancer continues to increase so that cancer has become one of the most devastating diseases worldwide. Therefore, cancer prevention has become an important avenue through which the fight against cancer could be feasible (Sarkar and Li, 2007). Since tumour active plant compounds often exert their anti-tumour activity through the regulation of cell signaling pathways different from those of platinum drugs, it is logical to think that phytochemicals in combination with the platinum drugs or alone may exert enhanced anti-tumour activity through synergistic action and/or compensation of the adverse effects (Chan and Fong, 2007). The combination therapy or treatment with phytochemicals alone may also decrease the systemic toxicity caused by chemotherapies or radiotherapies because of lower doses required. 

Thus, based on mode of action number of phytochemicals were chosen to test their activity against ovarian cancer cell lines (both parent and resistant) to justify the use of phytochemicals towards improvement in cancer treatment and to overcome resistance. Specific molecular mechanisms and cytotoxic action of the selected phytochemicals are discussed below to find out the specific logical way to fight against cancer.

 Anethole: Anethole (1-methoxy-4-(prop-1-enyl) benzene), the chief component of anise oil, fennel oil and camphor, and its derivative anethole ditholethione (ADT) have been shown to block carcinogenesis. They act as antioxidants (Park et al., 2003),  increase intracellular levels of GSH and GST (Bouthillier et al., 1996, Drukarch et al., 1997, Chen and deGraffenried, 2012), suppress TNF-induced lipid peroxidation and ROI generation and reduce oxidative levels by scavenging hydroxyl radicals (Chainy et al., 2000, Chen and deGraffenried, 2012). Anethole suppresses JNK and MAPK- kinase and NF-?B activation process (Aggarwal and Shishodia, 2006). Anethole also suppresses TNF induced activation of AP-1, which is involved in carcinogenesis (Karin, 1996, Chainy et al., 2000). AP-1 activation requires the activation of c-Jun N-terminal kinase (JNK) and mitogen-activated protein kinase (MAPK) kinase (MAPKK or MEK) (Karin, 1996).

Betulinic acid: Betulinic acid (3Ăź, hydroxy-lup-20(29)-en-28-oic acid) is a naturally occurring pentacyclic triterpenoid available in the outer bark of a variety of tree species, e.g. white-barked birch trees (Tan et al., 2003, Fulda, 2008). It has potent antitumour properties and can be used as an effective alternative when certain chemotherapy drugs fail. Non-malignant cells and normal tissues are not affected by BA as it exerts its effects directly on the mitochondrion to trigger death of cancerous cells (Ali-Seyed et al., 2016). It also has been shown to exert their effect through mechanisms that involve modulation of NF-?B. Without activating ERK, BA activates p38 and SAP/JNK early in the programmed cell death process and thus induction of apoptosis (Fulda et al., 1999, Tan et al., 2003). Further, betulinic acid exhibits anti-angiogenic activity due to activation of selective proteasome-dependent degradation of the transcription factors specificity protein 1 (Sp1), Sp3, and Sp4, which regulate vascular endothelial growth (VEGF) expression (Chintharlapalli et al., 2007).

Capsaicin: Capsaicin (trans-8-methyl-N-Vanilyl-6-nonenamide) is the pungent ingredient of red pepper and hot chili pepper is a cancer-suppressing agent. It is capable of inhibiting, retarding or reversing the multi-stage carcinogenesis and also angiogenesis and it does so by blocking translocation of NF-?B, and by inhibiting AP-1 and STAT3 signaling pathway (Oyagbemi et al., 2010). By upregulating tribbles-

related protein 3 (TRIB3) expression it also promotes apoptotic cell death in cancer cells (Lin et al., 2018). Capsaicin inhibits constitutive activation of STAT3 and thus down-regulates the expression of the STAT3-regulated gene products, such as cyclin D1, Bcl-2, Bcl-xL, survivin and VEGF (Bhutani et al., 2007). Capsaicin also induces the accumulation of cells in G1 phase, inhibits proliferation, and induces apoptosis by ROS generation, alteration of the mitochondrial inner transmembrane potential, JNK activation, activation of caspase 3, ceramide accumulation and extracellular signal-regulated protein kinase (ERK) activation (Sánchez et al., 2007, Sánchez et al., 2008, Oyagbemi et al., 2010).  In addition, generation of ROS in cells with the resultant induction of apoptosis and cell cycle arrest are beneficial for cancer chemoprevention.

Cholecalciferol: Mammalian skin produces cholecalciferol (vitamin D3) in the presence of sunlight (Norman, 2008). Though the major function of cholecalciferol is sustaining plasma calcium concentration and bone health (Chang et al., 2015), it has the ability to decrease interferon-gamma and NF-?B expression towards inducing autophagy (Wu and Sun, 2011). It is associated with reducing anti-apoptotic proteins expression such as cyclin D1, Bcl-2  and it also increases pro-apoptotic BAX protein expression (Yang et al., 2008). In addition, it inhibits cancer multiplication, progression and triggers apoptosis (Tokar and Webber, 2005, Fleet et al., 2012, Giammanco et al., 2015).

Curcumin: Curcumin (diferuloylmethane) is a key component of turmeric governs a number of intracellular targets, including proteins involved in antioxidant response, immune response, apoptosis, cell cycle regulation and tumor progression (Kumar et al., 2016). Curcumin suppresses the expression of TNFa and NF-?B, thus decreases the expression of inflammatory enzymes cyclooxygenase (COX2) and inducible nitric oxide synthase (iNOS) and ultimately sensitizes resistant cancer cells towards apoptosis by cisplatin and taxol (HemaIswarya and Doble, 2006, Sarkar et al., 2006). Curcumin also down-regulates the expression of IL1, IL6, TNFa (II), and angiogenic factors such as vascular endothelial growth factor (VEGF) (Lin et al., 2007, Tan et al., 2010). Curcumin modulates CKis, CDK-cyclin and Rb-E2F complexes to render G1-arrest and alters CDK/cyclin B complex formation to block G2/M transition (Sablina Anna et al., 2005). Curcumin promotes caspase-3-mediated cleavage of Ăź-catenin, decreases Ăź-catenin/Tcf-Lef transactivation capacity for c-Myc and cyclin D1 (Jaiswal et al., 2002).

Genistein: Genistein is a major component of soybean isoflavone and has multiple functions associated with anti-tumour effects (Suzuki et al., 2002). Genistein upregulates pro-apoptotic proteins (Bad and Bax) and miRNA-218 expression and also induces activation of cleaved caspase-3. It also reduces the activated NF-?B signalling and overproduction of pro-inflammatory cytokines (TNF-alpha, IL-1beta and IL-6), nuclear translocation of p65 and subsequent gene expression in cancer cells (Zheng et al., 2017). Genistein also inhibits AKT pathway and thus can enhance necrotic-like cancer cell death (Satoh et al., 2003). Moreover, genistein antagonizes estrogen- and androgen-mediated signaling pathways in the processes of carcinogenesis (Mathieson and Kitts, 1980). It also inhibits topoisomerase II (Okura et al., 1988), 17Ăź-hydroxysteroid dehydrogenase and 5a-reductase (Evans et al., 1995). Genistein has also been found to have antioxidant properties, and thus acts as a potent inhibitor of angiogenesis and metastasis (Banerjee et al., 2008). It is believed to have the ability to increase the expression of gene for glutathione peroxidase, without any significant change in the expression of gene for other antioxidant enzymes such as superoxide dismutase and catalase (Suzuki et al., 2002). Finally, inhibition of proteasome activity by genistein might contribute to its cancer-preventive properties (Kazi et al., 2003).

Honokiol: Honokiol is derived from seed/leaf/bark of Magnolia virginia magnolia tree (Fried and Arbiser, 2009). Honokiol is reported to suppress NF-?B, EFGR, STAT3 and mTOR transduction pathways (Arora et al., 2012) and induce caspase-dependent apoptosis and blocks VEGF expression to prevent new blood vessel formation (Fried and Arbiser, 2009). Honokiol inhibits tumour cell proliferation and decreases human epidermal growth receptor 2 expressions (Liu et al., 2008). Honokiol also inhibits G1 cell cycle phase progression and increases pro-apoptotic BAX expression (Arora et al., 2011). Honokiol has the ability to induce apoptosis via both p53 dependent and  p53 independent pathways (Wang et al., 2004, Guo et al., 2015).

Kaempferol:  Kaempferol is isolated from different plants and fruits, i.e. green tea, broccoli, apple, grapes and tomato (Kim and Choi, 2013). Kaempferol significantly decreases VEGF biomarker expression towards preventing formation of the new blood vessel. It inhibits cell cycle G2/M phase transition by preventing cyclin-dependent kinase 1 and cyclin B. It blocks AP-1 and ERKp38- JNK signaling pathways to prevent the tumour cell invasion. It dampens the NF-?B expression as well as Bcl-2 and Bcl-xL anti-apoptotic proteins. Kaempferol greatly induces pro-apoptotic p53, BAX and BAD proteins expression (Kim and Choi, 2013). 

Mycophenolic acid: Mycophenolic acid is a secondary metabolite produced by marine fungi Penicillium brevicompactum (Rovirosa et al., 2006). Mycophenolic acid is a potent inhibitor of inosine monophosphate dehydrogenase enzyme in the de-novo guanosine nucleotides synthesis (Allison and Eugui, 2000), which prevents T cells synthesis namely TNF-a, IL-17 and Interferon-? (He et al., 2011). Therefore, it could inhibit replication of DNA. It increases pro-apoptotic p53 protein activity towards facilitating apoptosis (Sun et al., 2008) and blocks vascular VEGF-a secretion towards preventing new blood vessel formation (Monguilhott Dalmarco et al., 2011). It also inhibits AKT/mTOR and NF-?B pathways (Morales et al., 2008, He et al., 2011, Zheng et al., 2011) towards promoting anticancer activity.

Paclitaxel: Paclitaxel (Taxol), was first discovered in the bark of the western yew tree, Taxus brevifolia (Kaye, 1996).  The combined treatment of paclitaxel with cisplatin in first-line therapy marked the next major advance in the treatment efficacy for advanced ovarian cancer. It is a mitotic inhibitor that arrests cells at G2/M phase. It interferes with the normal breakdown of microtubules during cell division by stabilizing microtubules, and thus it destroys cell's ability to use its cytoskeleton in a flexible manner. In addition paclitaxel also acts as a molecular mop by sequestering free tubulin effectively depleting the cells supply of tubulin monomers and/or dimers and thus trigger apoptosis (Foss et al., 2008). Paclitaxel also induces apoptosis in cancer cells by binding to anti-apoptotic protein Bcl-2 (Haldar et al., 1995). Recently, numerous preclinical studies have suggested that the combination of paclitaxel and curcumin may be an ideal strategy to reverse multi drug resistance and synergistically improve therapeutic efficacy in cancer therapy (Wei et al., 2017).

Quercetin: Quercetin (3,3’,4’,5,7-pentahydroxyflavone) is an important dietary flavonoid, present in different vegetables, fruits, seeds, nuts and tea It strongly increases intracellular ROS levels, by producing quercetin radicals (Quercetin-O•) (Jeong et al., 2008), and thus causes free radical-induced apoptosis through the ROS/AMPKa1/ASK1/p38 and the AMPKa1/COX2 signaling pathways (Lee et al., 2010). In addition, quercetin radicals also lowers the intracellular pool of GSH and thus triggers apoptosis through mitochondrial depolarization (Lugli et al., 2005, Gibellini et al., 2010). Quercetin reduces drug-induced up-regulation of p53, p21 and Bax and reduces the levels of cyclin B1 and survivin proteins (Samuel et al., 2012). These events go along with the cleavage of procaspase 9 and poly (ADP-ribose) polymerase (PARP) (Granado-Serrano et al., 2006, Lee et al., 2006, Yang et al., 2006). Quercetin also inhibits the activation of NF-?B and PI3K/AKT pathway (Aggarwal and Shishodia, 2006, Granado-Serrano et al., 2006, Gibellini et al., 2010) and upregulates the death receptor (DR)-5, which is activated by TNF-related apoptosis-inducing ligand (TRAIL) (Kim et al., 2008, Jung et al., 2010).

Resveratrol: Resveratrol (3,5,4’-trihydroxystilben) present in red grapes, mulberries and some nuts induces apoptosis and overcomes chemo-resistance through the down-regulation of NF-?B activation and signal transducer and activator transcription factor 3 (STAT3), anti-apoptotic and cell survival gene products such as Bcl-2, and also by regulating the cyclooxygenase (COX) expression  (Sexton et al., 2006, Hu et al., 2007, Singh et al., 2009). Resveratrol also up-regulates the tumour suppressor p53 and cytokine (MIC-1) that possesses anti-tumourigenic activity (Golkar et al., 2007). It also increases the expression of glutathione S-transferase (GST) and glutathione peroxidase (GPx) or their activity resulting in lowered glutathione level (Hu et al., 2007). In the absence of oxygen, resveratrol also acts as a protector against radiation (Bader and Getoff, 2006). Resveratrol downregulates expression of related Wnt/beta-catenin signaling pathway target genes, such as Ăź-catenin, c-myc, cyclin D1, MMP-2 and MMP-9, and upregulates E-cadherin level as well. Resveratrol also suppresses the activity of Wnt/ Ăź-catenin signaling pathway (Xie et al., 2017).

Thymoquinone: Thymoquinone (2-Isopropyl-5-methylbenzo-1,4-quinone) is a plant compound present in the Nigella sativa has potent anti-proliferative, pro-apoptotic, anti-oxidant, cytotoxic, anti-metastatic, and NK-dependent cytotoxic effects. The most significant pathways through which thymoquinone mediates its anti-cancer activity are p53, NF- ?B, PPARgamma, STAT3, MAPK, and PI3K/AKT signaling pathways (Majdalawieh et al., 2017). Apoptosis induction by thymoquinone is associated with an increase in mRNA expression of p53 and the downstream p53 target gene p21WAF1 and a significant inhibition of anti-apoptotic Bcl-2 protein (Gali-Muhtasib et al., 2004). Thymoquinone suppresses activation of AKT and ERK, TNF, NF-?B and the NF-?B-dependent reporter gene expression (Aggarwal et al., 2008). It also has been found to down-regulate the expression of antiapoptotic (IAP1, IAP2, XIAP Bcl-2, Bcl-xL, and survivin), proliferative (cyclin D1, COX-2, and c-myc), and angiogenic (MMP-9 and VEGF) gene products (El-Dakhakhny et al., 2002, El-Mahmoudy et al., 2002, Gali-Muhtasib et al., 2004, El-Mahmoudy et al., 2005, El Mezayen et al., 2006, Roepke et al., 2007).

6-shogaol: 6-shogaol is mainly isolated from Ginger Zingiber officinale Roscoe (Ok and Jeong, 2012). 6-shogaol inhibits cell proliferation  and triggers autophagy by blocking mTOR/AKT pathway (Hung et al., 2009, Ray et al., 2015, Li and Chiang, 2017). It induces mitotic arrest and reduces the cancer cells viability (Ishiguro et al., 2007) and inhibits cell cycle G2/M phase transition (Li and Chiang, 2017). 6-shogaol exceedingly attenuates NF-?B genes such as MMP-9, survivin, c-MYC, and cyclin D1 (Ling et al., 2010, Saha et al., 2014). It increases PPAR-? dependent apoptosis (Tan et al., 2013) and also causes programmed cell death through caspase-dependent oxidative stress pathway (Chen et al., 2007, Liu et al., 2013) and initiates up-regulation of p27, interleukin-7 and BAX pro-apoptotic factors to assist apoptosis (Saha et al., 2014).

Phytochemicals in cancer treatment

GO

From the above discussions it is clear that phytochemicals exert their anti-tumour effects by bringing into play numerous cellular proteins, signaling pathways (summarized in Table-1) which in turn affect multiple steps in the pathways leading to tumourigenesis. In addition, based on the specific mode of action in cancer treatment and prevention, the selected phytochemicals are classified broadly in six major groups are shown in the Figure-1. This is also supported by the cytotoxic action of the selected phytochemicals by in vitro Cytotoxicity tests on ovarian cancer cell lines (summarized in Table-2).

From the IC50 values given in Table 2, it can be seen that all of the phytochemicals generally have cytotoxic action in all ovarian cancer cell lines and have either comparable or greater activity in the resistant cell lines than in the parent cell line. When the major factors involved in the development of cancer and platinum resistance in ovarian cancer are counter acted by phytochemicals, it is not unexpected to find that the phytochemicals have comparable or greater activity in the parent and the resistant cell lines.

As plant derived products are natural products, another aim of using the phytochemicals would be to prevent the damage to normal cells and tissues caused by the direct and bystander effects of platinum drugs. As the molecular mechanisms of action of platinum drugs and the phytochemicals are found to be different, it is logical to assume that phytochemicals possessing both cancer preventive and therapeutic attributes, can be ideal candidates for combination with targeted therapy towards synergistic outcome.

 

Conclusion

GO

Cancer is the most dreaded disease of our time because of its ability to metastasize and develop resistance to drugs resulting in failure in treatment, and the lack of complete understanding of mechanisms of its development. Knowledge on mechanism of cytotoxic activity of phytochemicals in various cancer cell lines shows that phytochemicals bring antitumour effects by modulating numerous cellular proteins, signaling pathways which in turn affect multiple steps in the pathways leading to prevent tumourigenesis. Thus, it is logical to think that phytochemicals can be wonderful agent against cancer development and therapy especially related problems associated with conventional cancer therapies. Because of heterogeneity of ovarian cancer, relationships among histological group, stage of disease, tumour markers, patient characteristics and survival of different ovarian cancer patients are different, meaning one patient may benefit from one type of treatment whereas many others might fail to respond. Therefore, there should be continuous search for the new tumour active phytochemicals that will provide better patient care, efficacy and safety.

Future Directions: In depth in vivo studies involving the use of phytochemicals alone and in combination with targeted therapy against cancer are needed towards the development of less toxic and more affordable anticancer therapy including that for ovarian cancer.

Author Contributions

GO

Nessa and Anwar reviewed the current literature and drafted the manuscript. Huq conceptualized the project, provided overall guidance and edited the manuscript.

References


Adams, P.D. (2001). Regulation of the retinoblastoma tumor suppressor protein by cyclin/cdks, Biochimica et Biophysica Acta, Reviews on Cancer 1471, 3, M123-M133.
https://doi.org/10.1016/S0304-419X(01)00019-1

Aggarwal, B.B., Kunnumakkara, A.B., Harikumar, K.B., Tharakan, S.T., Sung, B. and Anand, P. (2008). Potential of spice-derived phytochemicals for cancer prevention, Planta Med 74, 13, 1560-1569.
https://doi.org/10.1055/s-2008-1074578
PMid:18612945

Aggarwal, B.B. and Shishodia, S. (2006). Molecular targets of dietary agents for prevention and therapy of cancer, Biochemical Pharmacology 71, 10, 1397-1421.
https://doi.org/10.1016/j.bcp.2006.02.009
PMid:16563357

Alam, M. (2018). Studies on novel palladiums alone and in combination with phytochemicals in tumour models PhD Thesis, The University of Sydney, Australia.

Alamro, A.A.S. (2015). Studies on combination between tumour active compounds in ovarian tumour models, PhD Thesis, The University of Sydney, Australia.

Ali-Seyed, M., Jantan, I., Vijayaraghavan, K. and Bukhari, S.N. (2016). Betulinic Acid: Recent Advances in Chemical Modifications, Effective Delivery, and Molecular Mechanisms of a Promising Anticancer Therapy, Chem Biol Drug Des 87, 4, 517-536.
https://doi.org/10.1111/cbdd.12682
PMid:26535952

Allison, A.C. and Eugui, E.M. (2000). Mycophenolate mofetil and its mechanisms of action, Immunopharmacology 47, 2-3, 85-118.
https://doi.org/10.1016/S0162-3109(00)00188-0

Amos, C. and Struewing, J. (1993). Genetic epidemiology of epithelial ovarian cancer, Cancer 71, S2, 566-572.
https://doi.org/10.1002/cncr.2820710212
PMid:8420678

Anon (2019). Ovarian Cancer Research Alliance, The American Cancer Society.

Anwar, M. (2018). Natural compounds in combination with platinum drugs administered to ovarian cancer models towards synergistic outcomes PhD Thesis, The University of Sydney, Australia.

Anwar, M.S., Yu, J.Q., Beale, P. and Huq, F. (2016). 6-Shogaol and mycophenolic acid are seen to act synergistically in combination with platinum drug in killing ovarian cancer cells, European Journal of Cancer 69, S18.
https://doi.org/10.1016/S0959-8049(16)32634-X

Anwar, M.S., Yu, J.Q., Beale, P. and Huq, F. (2017). Abstract 4212: Natural compounds alone and in combination with platinum drugs found to show significant anti-tumour activity against ovarian cancer cell lines, Cancer Research 77, 13 Supplement, 4212-4212.
https://doi.org/10.1158/1538-7445.AM2017-4212

Arora, S., Bhardwaj, A., Srivastava, S.K., Singh, S., McClellan, S., Wang, B. and Singh, A.P. (2011). Honokiol arrests cell cycle, induces apoptosis, and potentiates the cytotoxic effect of gemcitabine in human pancreatic cancer cells, PloS one 6, 6, e21573.
https://doi.org/10.1371/journal.pone.0021573
PMid:21720559 PMCid:PMC3123370

Arora, S., Singh, S., Piazza, G.A., Contreras, C.M., Panyam, J. and Singh, A.P. (2012). Honokiol: a novel natural agent for cancer prevention and therapy, Current molecular medicine 12, 10, 1244-1252.
https://doi.org/10.2174/156652412803833508
PMid:22834827 PMCid:PMC3663139

Aunan, J.R., Cho, W.C. and Soreide, K. (2017). The Biology of Aging and Cancer: A Brief Overview of Shared and Divergent Molecular Hallmarks, Aging dis 8, 5, 628-642.
https://doi.org/10.14336/AD.2017.0103
PMid:28966806 PMCid:PMC5614326

Bader, Y. and Getoff, N. (2006). Effect of resveratrol and mixtures of resveratrol and mitomycin C on cancer cells under irradiation, Anticancer Res 26, 6B, 4403-4408.

Banerjee, S., Li, Y., Wang, Z. and Sarkar, F.H. (2008). Multi-targeted therapy of cancer by genistein, Cancer Letters (Shannon, Ireland) 269, 2, 226-242.
https://doi.org/10.1016/j.canlet.2008.03.052
PMid:18492603 PMCid:PMC2575691

Barbieri, D., Grassilli, E., Monti, D., Salvioli, S., Franceschini, M.G., Franchini, A., Belelsia, E., Salomoni, P., Negro, P. and et al. (1994). D-Ribose and deoxy-D-ribose induce apoptosis in human quiescent peripheral blood mononuclear cells, Biochemical and Biophysical Research Communications 201, 3, 1109-1116.
https://doi.org/10.1006/bbrc.1994.1820
PMid:8024552

Bhutani, M., Pathak, A.K., Nair, A.S., Kunnumakkara, A.B., Guha, S., Sethi, G. and Aggarwal, B.B. (2007). Capsaicin Is a Novel Blocker of Constitutive and Interleukin-6-Inducible STAT3 Activation, Clinical Cancer Research 13, 10, 3024-3032.
https://doi.org/10.1158/1078-0432.CCR-06-2575
PMid:17505005

Binju, M., Padilla, M.A., Singomat, T., Kaur, P., Suryo Rahmanto, Y., Cohen, P.A. and Yu, Y. (2019). Mechanisms underlying acquired platinum resistance in high grade serous ovarian cancer - a mini review, Biochim 1863, 2, 371-378.
https://doi.org/10.1016/j.bbagen.2018.11.005
PMid:30423357

Boulikas, T. and Vougiouka, M. (2004). Recent clinical trials using cisplatin, carboplatin and their combination chemotherapy drugs (review), Oncology Reports 11, 3, 559-595.
https://doi.org/10.3892/or.11.3.559

Bouthillier, L., Charbonneau, M. and Brodeur, J. (1996). Assessment of the role of glutathione conjugation in the protection afforded by anethol dithiolthione against hexachloro-1,3-butadiene-induced nephrotoxicity, Toxicology and Applied Pharmacology 139, 1, 177-185.
https://doi.org/10.1006/taap.1996.0156
PMid:8685901

Brian T. Kawasaki, E.M.H., Tashan Mistree, and Farrar, a.W.L. (2008). Targeting cancer stem cells with phytochemicals, Molecular Interventions 8, 4, 11.
https://doi.org/10.1124/mi.8.4.9
PMid:18829843

Chainy, G.B.N., Manna, S.K., Chaturvedi, M.M. and Aggarwal, B.B. (2000). Anethole blocks both early and late cellular responses transduced by tumor necrosis factor: effect on NF-κB, AP-1, JNK, MAPKK and apoptosis, Oncogene 19, 25, 2943-2950.
https://doi.org/10.1038/sj.onc.1203614
PMid:10871845

Chan, M.M., Chen, R. and Fong, D. (2018). Targeting cancer stem cells with dietary phytochemical - Repositioned drug combinations, Cancer letters 433, 53-64.
https://doi.org/10.1016/j.canlet.2018.06.034
PMid:29960048 PMCid:PMC7117025

Chan, M.M. and Fong, D. (2007). Overcoming ovarian cancer drug resistance with phytochemicals and other compounds. In. Parsons, R.A. (Ed.), Prog. Cancer Drug Resist. Res. Nova Science Publishers, Inc., pppp1-28.

Chang, B., Schlussel, Y., Sukumar, D., Schneider, S.H. and Shapses, S.A. (2015). Influence of vitamin D and estrogen receptor gene polymorphisms on calcium absorption: BsmI predicts a greater decrease during energy restriction, Bone 81, 138-144.
https://doi.org/10.1016/j.bone.2015.07.011
PMid:26165414 PMCid:PMC4641000

Chen, C.-Y., Liu, T.-Z., Liu, Y.-W., Tseng, W.-C., Liu, R.H., Lu, F.-J., Lin, Y.-S., Kuo, S.-H. and Chen, C.-H. (2007). 6-shogaol (alkanone from ginger) induces apoptotic cell death of human hepatoma p53 mutant Mahlavu subline via an oxidative stress-mediated caspase-dependent mechanism, Journal of agricultural and food chemistry 55, 3, 948-954.
https://doi.org/10.1021/jf0624594
PMid:17263498

Chen, C.H. and deGraffenried, L.A. (2012). Anethole suppressed cell survival and induced apoptosis in human breast cancer cells independent of estrogen receptor status, Phytomedicine 19, 8-9, 763-767.
https://doi.org/10.1016/j.phymed.2012.02.017
PMid:22464689

Chintharlapalli, S., Papineni, S., Ramaiah, S.K. and Safe, S. (2007). Betulinic acid inhibits prostate cancer growth through inhibition of specificity protein transcription factors, Cancer Research 67, 6, 2816-2823.
https://doi.org/10.1158/0008-5472.CAN-06-3735
PMid:17363604

Council, C. (2020), April 2020. Understanding Ovarian Cancer. from https://www.cancer.org.au/content/about_cancer/ebooks/cancertypes/Understanding_Ovarian_Cancer_booklet_April_2020.pdf#_ga=2.102422786.2071573207.1593235825-1363732324.1554694293 (accessed 20.05.20).

Crul, M., van Waardenburg, R.C., Beijnen, J.H. and Schellens, J.H. (2002). DNA-based drug interactions of cisplatin, Cancer Treatment Reviews 28, 6, 291-303.
https://doi.org/10.1016/S0305-7372(02)00093-2

Dalle-Donne, I., Giustarini, D., Colombo, R., Rossi, R. and Milzani, A. (2003). Protein carbonylation in human diseases, Trends in Molecular Medicine 9, 4, 169-176.
https://doi.org/10.1016/S1471-4914(03)00031-5

Dawson, A., Fernandez, M.L., Anglesio, M., Yong, P.J. and Carey, M.S. (2018). Endometriosis and endometriosis-associated cancers: new insights into the molecular mechanisms of ovarian cancer development, Ecancermedicalscience 12, 803.
https://doi.org/10.3332/ecancer.2018.803
PMid:29456620 PMCid:PMC5813919

Drukarch, B., Schepens, E., Stoof, J.C. and Langeveld, C.H. (1997). Anethole dithiolethione prevents oxidative damage in glutathione-depleted astrocytes, European Journal of Pharmacology 329, 2/3, 259-262.
https://doi.org/10.1016/S0014-2999(97)89187-X

El-Dakhakhny, M., Madi, N., Lembert, N. and Ammon, H. (2002). < i> Nigella sativa</i> oil, nigellone and derived thymoquinone inhibit synthesis of 5-lipoxygenase products in polymorphonuclear leukocytes from rats, Journal of Ethnopharmacology 81, 2, 161-164.
https://doi.org/10.1016/S0378-8741(02)00051-X

El-Mahmoudy, A., Matsuyama, H., Borgan, M., Shimizu, Y., El-Sayed, M., Minamoto, N. and Takewaki, T. (2002). Thymoquinone suppresses expression of inducible nitric oxide synthase in rat macrophages, International immunopharmacology 2, 11, 1603-1611.
https://doi.org/10.1016/S1567-5769(02)00139-X

El-Mahmoudy, A., Shimizu, Y., Shiina, T., Matsuyama, H., Nikami, H. and Takewaki, T. (2005). Macrophage-derived cytokine and nitric oxide profiles in type I and type II diabetes mellitus: effect of thymoquinone, Acta diabetologica 42, 1, 23-30.
https://doi.org/10.1007/s00592-005-0170-6
PMid:15868110

El Mezayen, R., El Gazzar, M., Nicolls, M.R., Marecki, J.C., Dreskin, S.C. and Nomiyama, H. (2006). Effect of thymoquinone on cyclooxygenase expression and prostaglandin production in a mouse model of allergic airway inflammation, Immunology letters 106, 1, 72-81.
https://doi.org/10.1016/j.imlet.2006.04.012
PMid:16762422

England, K. and Cotter, T.G. (2005). Direct oxidative modifications of signalling proteins in mammalian cells and their effects on apoptosis, Redox Report 10, 5, 237-245.
https://doi.org/10.1179/135100005X70224
PMid:16354412

Evan, G.I. and Vousden, K.H. (2001). Proliferation, cell cycle and apoptosis in cancer, Nature (London, United Kingdom) 411, 6835, 342-348.
https://doi.org/10.1038/35077213
PMid:11357141

Evans, B.A.J., Griffiths, K. and Morton, M.S. (1995). Inhibition of 5α-reductase in genital skin fibroblasts and prostate tissue by dietary lignans and isoflavonoids, Journal of Endocrinology 147, 2, 295-302.
https://doi.org/10.1677/joe.0.1470295
PMid:7490559

Fathalla, M. (1971). Incessant ovulation--a factor in ovarian neoplasia?, Lancet 2, 7716, 163.
https://doi.org/10.1016/S0140-6736(71)92335-X

Fleet, J.C., DeSmet, M., Johnson, R. and Li, Y. (2012). Vitamin D and Cancer: A review of molecular mechanisms, The Biochemical journal 441, 1, 61-76.
https://doi.org/10.1042/BJ20110744
PMid:22168439 PMCid:PMC4572477

Ford, D., Easton, D., Stratton, M., Narod, S., Goldgar, D., Devilee, P., Bishop, D., Weber, B., Lenoir, G. and Chang-Claude, J. (1998). Genetic heterogeneity and penetrance analysis of the BRCA1 and BRCA2 genes in breast cancer families. The Breast Cancer Linkage Consortium, American journal of human genetics 62, 3, 676.
https://doi.org/10.1086/301749
PMid:9497246 PMCid:PMC1376944

Foss, M., Wilcox, B.W., Alsop, G.B., Zhang, D., Foss, M., Wilcox, B.W.L., Alsop, G.B. and Zhang, D. (2008). Taxol crystals can masquerade as stabilized microtubules, PLoS ONE [Electronic Resource] 3, 1, e1476.
https://doi.org/10.1371/journal.pone.0001476
PMid:18213384 PMCid:PMC2194920

Fried, L.E. and Arbiser, J.L. (2009). Honokiol, a Multifunctional Antiangiogenic and Antitumor Agent, Antioxidants & Redox Signaling 11, 5, 1139-1148.
https://doi.org/10.1089/ars.2009.2440
PMid:19203212 PMCid:PMC2842137

Fuchs, S.Y., Adler, V., Buschmann, T., Yin, Z., Wu, X., Jones, S.N. and Ronai, Z.e. (1998). JNK targets p53 ubiquitination and degradation in nonstressed cells, Genes & Development 12, 17, 2658-2663.
https://doi.org/10.1101/gad.12.17.2658
PMid:9732264 PMCid:PMC317120

Fuchs, S.Y., Adler, V., Pincus, M.R. and Ronai, Z.e. (1998). MEKK1/JNK signaling stabilizes and activates p53, Proceedings of the National Academy of Sciences 95, 18, 10541-10546.
https://doi.org/10.1073/pnas.95.18.10541
PMid:9724739 PMCid:PMC27930

Fulda, S. (2008). Betulinic acid for cancer treatment and prevention, Int. J. Mol. Sci. 9, 6, 1096-1107.
https://doi.org/10.3390/ijms9061096
PMid:19325847 PMCid:PMC2658785

Fulda, S., Jeremias, I., Steiner, H.H., Pietsch, T. and Debatin, K.M. (1999). Betulinic acid: A new cytotoxic agent against malignant brain-tumor cells, International Journal of Cancer 82, 3, 435-441.
https://doi.org/10.1002/(SICI)1097-0215(19990730)82:3<435::AID-IJC18>3.0.CO;2-1

Gali-Muhtasib, H., Diab-Assaf, M., Boltze, C., Al-Hmaira, J., Hartig, R., Roessner, A., Schneider-Stock, R., Gali-Muhtasib, H., Diab-Assaf, M., Boltze, C., Al-Hmaira, J., Hartig, R., Roessner, A. and Schneider-Stock, R. (2004). Thymoquinone extracted from black seed triggers apoptotic cell death in human colorectal cancer cells via a p53-dependent mechanism, International Journal of Oncology 25, 4, 857-866.

Gali-Muhtasib, H.U., Kheir, W.G.A., Kheir, L.A., Darwiche, N. and Crooks, P.A. (2004). Molecular pathway for thymoquinone-induced cell-cycle arrest and apoptosis in neoplastic keratinocytes, Anti-cancer drugs 15, 4, 389-399.
https://doi.org/10.1097/00001813-200404000-00012
PMid:15057144

Galluzzi, L., Senovilla, L., Vitale, I., Michels, J., Martins, I., Kepp, O., Castedo, M. and Kroemer, G. (2011). Molecular mechanisms of cisplatin resistance, Oncogene 31, 15, 1869-1883.
https://doi.org/10.1038/onc.2011.384
PMid:21892204

Garmann, D., Warnecke, A., Kalayda, G.V., Kratz, F. and Jaehde, U. (2008). Cellular accumulation and cytotoxicity of macromolecular platinum complexes in cisplatin-resistant tumor cells, J Controlled Release 131, 2, 100-106.
https://doi.org/10.1016/j.jconrel.2008.07.017
PMid:18691617

Giammanco, M., Di Majo, D., La Guardia, M., Aiello, S., Crescimannno, M., Flandina, C., Tumminello, F.M. and Leto, G. (2015). Vitamin D in cancer chemoprevention, Pharmaceutical Biology 53, 10, 1399-1434.
https://doi.org/10.3109/13880209.2014.988274
PMid:25856702

Gibellini, L., Pinti, M., Nasi, M., De Biasi, S., Roat, E., Bertoncelli, L. and Cossarizza, A. (2010). Interfering with ROS metabolism in cancer cells: the potential role of quercetin, Cancers 2, 1288-1311.
https://doi.org/10.3390/cancers2021288
PMid:24281116 PMCid:PMC3835130

Golkar, L., Ding, X.Z., Ujiki, M.B., Salabat, M.R., Kelly, D.L., Scholtens, D., Fought, A.J., Bentrem, D.J., Talamonti, M.S., Bell, R.H., Adrian, T.E., Golkar, L., Ding, X.-Z., Ujiki, M.B., Salabat, M.R., Kelly, D.L., Scholtens, D., Fought, A.J., Bentrem, D.J., Talamonti, M.S., Bell, R.H. and Adrian, T.E. (2007). Resveratrol inhibits pancreatic cancer cell proliferation through transcriptional induction of macrophage inhibitory cytokine-1, J Surg Res 138, 2, 163-169.
https://doi.org/10.1016/j.jss.2006.05.037
PMid:17257620

Granado-Serrano, A.B., Martín, M.A., Bravo, L., Goya, L. and Ramos, S. (2006). Quercetin induces apoptosis via caspase activation, regulation of Bcl-2, and inhibition of PI-3-kinase/Akt and ERK pathways in a human hepatoma cell line (HepG2), The Journal of nutrition 136, 11, 2715-2721.
https://doi.org/10.1093/jn/136.11.2715
PMid:17056790

Grivennikov, S.I., Greten, F.R. and Karin, M. (2010). Immunity, inflammation, and cancer, Cell 140, 6, 883-899.
https://doi.org/10.1016/j.cell.2010.01.025
PMid:20303878 PMCid:PMC2866629

Guo, Y.-B., Bao, X.-J., Xu, S.-B., Zhang, X.-D. and Liu, H.-Y. (2015). Honokiol induces cell cycle arrest and apoptosis via p53 activation in H4 human neuroglioma cells, International Journal of Clinical and Experimental Medicine 8, 5, 7168-7175.

Hahn, W.C. and Weinberg, R.A. (2002). Rules for making human tumor cells, New England Journal of Medicine 347, 20, 1593-1603.
https://doi.org/10.1056/NEJMra021902
PMid:12432047

Haldar, S., Jena, N. and Croce, C.M. (1995). Inactivation of Bcl-2 by phosphorylation, Proceedings of the National Academy of Sciences of the United States of America 92, 10, 4507-4511.
https://doi.org/10.1073/pnas.92.10.4507
PMid:7753834 PMCid:PMC41973

Hanahan, D. and Weinberg, R.A. (2000). The hallmarks of cancer, Cell (Cambridge, Mass.) 100, 1, 57-70.
https://doi.org/10.1016/S0092-8674(00)81683-9

Harbour, J.W. and Dean, D.C. (2000). The Rb/E2F pathway: expanding roles and emerging paradigms.
https://doi.org/10.1101/gad.813200
PMid:11018009

Hartwell, L.H. and Kastan, M.B. (1994). Cell cycle control and cancer, Science (Washington, D. C.) 266, 5192, 1821-1828.
https://doi.org/10.1126/science.7997877
PMid:7997877

He, X., Smeets, R.L., Koenen, H.J.P.M., Vink, P.M., Wagenaars, J., Boots, A.M.H. and Joosten, I. (2011). Mycophenolic Acid-Mediated Suppression of Human CD4+ T Cells: More Than Mere Guanine Nucleotide Deprivation, American Journal of Transplantation 11, 3, 439-449.
https://doi.org/10.1111/j.1600-6143.2010.03413.x
PMid:21342445

HemaIswarya, S. and Doble, M. (2006). Potential synergism of natural products in the treatment of cancer, Phytother Res 20, 4, 239-249.
https://doi.org/10.1002/ptr.1841
PMid:16557604

Holzer, A.K. and Howell, S.B. (2006). The internalization and degradation of human copper transporter 1 following cisplatin exposure, Cancer Research 66, 22, 10944-10952.
https://doi.org/10.1158/0008-5472.CAN-06-1710
PMid:17108132

Holzer, A.K., Manorek, G.H. and Howell, S.B. (2006). Contribution of the major copper influx transporter CTR1 to the cellular accumulation of cisplatin, carboplatin, and oxaliplatin, Molecular Pharmacology 70, 4, 1390-1394.
https://doi.org/10.1124/mol.106.022624
PMid:16847145

Hu, Y., Rahlfs, S., Mersch-Sundermann, V., Becker, K., Hu, Y., Rahlfs, S., Mersch-Sundermann, V. and Becker, K. (2007). Resveratrol modulates mRNA transcripts of genes related to redox metabolism and cell proliferation in non-small-cell lung carcinoma cells, Biol Chem 388, 2, 207-219.
https://doi.org/10.1515/BC.2007.023

Hung, J.-Y., Hsu, Y.-L., Li, C.-T., Ko, Y.-C., Ni, W.-C., Huang, M.-S. and Kuo, P.-L. (2009). 6-Shogaol, an active constituent of dietary ginger, induces autophagy by inhibiting the AKT/mTOR pathway in human non-small cell lung cancer A549 cells, Journal of agricultural and food chemistry 57, 20, 9809-9816.
https://doi.org/10.1021/jf902315e
PMid:19799425

Ishiguro, K., Ando, T., Maeda, O., Ohmiya, N., Niwa, Y., Kadomatsu, K. and Goto, H. (2007). Ginger ingredients reduce viability of gastric cancer cells via distinct mechanisms, Biochemical and biophysical research communications 362, 1, 218-223.
https://doi.org/10.1016/j.bbrc.2007.08.012
PMid:17706603

Ishikawa, T., Bao, J.-J., Yamane, Y., Akimaru, K., Frindrich, K., Wright, C.D. and Kuo, M.T. (1996). Coordinated induction of MRP/GS-X pump and γ-glutamylcysteine synthetase by heavy metals in human leukemia cells, Journal of Biological Chemistry 271, 25, 14981-14988.
https://doi.org/10.1074/jbc.271.25.14981
PMid:8663001

Ittiudomrak, T., Puthong, S., Roytrakul, S. and Chanchao, C. (2019). alpha-Mangostin and Apigenin Induced Cell Cycle Arrest and Programmed Cell Death in SKOV-3 Ovarian Cancer Cells, Toxicol 35, 2, 167-179.
https://doi.org/10.5487/TR.2019.35.2.167
PMid:31015899 PMCid:PMC6467359

Jaiswal, A., Marlow, B., Gupta, N. and Narayan, S. (2002). Beta-catenin-mediated transactivation and cell-cell adhesion pathways are important in curcumin (diferuylmethane)-induced growth arrest and apoptosis in colon cancer cells, Oncogene 21, 55, 8414-8427.
https://doi.org/10.1038/sj.onc.1205947
PMid:12466962

Jeong, J.-H., An, J.Y., Kwon, Y.T., Rhee, J.G. and Lee, Y.J. (2008). Effects of low dose quercetin: cancer cell-specific inhibition of cell cycle progression, Journal of Cellular Biochemistry 106, 1, 73-82.
https://doi.org/10.1002/jcb.21977
PMid:19009557 PMCid:PMC2736626

Jung, Y.H., Heo, J., Lee, Y.J., Kwon, T.K. and Kim, Y.H. (2010). Quercetin enhances TRAIL-induced apoptosis in prostate cancer cells via increased protein stability of death receptor 5, Life sciences 86, 9, 351-357.
https://doi.org/10.1016/j.lfs.2010.01.008
PMid:20096292 PMCid:PMC3003259

Kane, L.P., Shapiro, V.S., Stokoe, D. and Weiss, A. (1999). Induction of NF-κB by the Akt/PKB kinase, Current Biology 9, 11, 601-S601.
https://doi.org/10.1016/S0960-9822(99)80265-6

Karin, M. (1996). The regulation of AP-1 activity by mitogen-activated protein kinases, Philosophical Transactions of the Royal Society of London, Series B: Biological Sciences 351, 1336, 127-134.
https://doi.org/10.1098/rstb.1996.0008
PMid:8650258

Kasibhatla, S. and Tseng, B. (2003). Why target apoptosis in cancer treatment?, Molecular Cancer Therapeutics 2, 6, 573-580.

Kaye, S. (1996). Paclitaxel in Cancer Treatment, British journal of cancer 73, 3, 414.
https://doi.org/10.1038/bjc.1996.72
PMCid:PMC2074423

Kazi, A., Daniel, K.G., Smith, D.M., Kumar, N.B. and Dou, Q.P. (2003). Inhibition of the proteasome activity, a novel mechanism associated with the tumor cell apoptosis-inducing ability of genistein, Biochemical Pharmacology 66, 6, 965-976.
https://doi.org/10.1016/S0006-2952(03)00414-3

Kelland, L.R. (2000). Preclinical perspectives on platinum resistance, Drugs 59 Suppl 4, 1-8; discussion 37-38.
https://doi.org/10.2165/00003495-200059004-00001
PMid:10864225

Kim, J.Y., Kim, E.H., Park, S.S., Lim, J.H., Kwon, T.K. and Choi, K.S. (2008). Quercetin sensitizes human hepatoma cells to TRAIL-induced apoptosis via Sp1-mediated DR5 up-regulation and proteasome-mediated c-FLIPS down-regulation, Journal of Cellular Biochemistry 105, 6, 1386-1398.
https://doi.org/10.1002/jcb.21958
PMid:18980244

Kim, M.-K., Kim, K., Han, J.Y., Lim, J.M. and Song, Y.S. (2011). Modulation of inflammatory signaling pathways by phytochemicals in ovarian cancer, Genes and Nutrition 6, 2, 109-115.
https://doi.org/10.1007/s12263-011-0209-y
PMid:21484164 PMCid:PMC3092902

Kim, S.-H. and Choi, K.-C. (2013). Anti-cancer Effect and Underlying Mechanism(s) of Kaempferol, a Phytoestrogen, on the Regulation of Apoptosis in Diverse Cancer Cell Models, Toxicol 29, 4, 229-234.
https://doi.org/10.5487/TR.2013.29.4.229
PMid:24578792 PMCid:PMC3936174

Kishimoto, S., Kawazoe, Y., Ikeno, M., Saitoh, M., Nakano, Y., Nishi, Y., Fukushima, S. and Takeuchi, Y. (2006). Role of Na+, K+-ATPase alpha1 subunit in the intracellular accumulation of cisplatin, Cancer Chemother Pharmacol 57, 1, 84-90.
https://doi.org/10.1007/s00280-005-0003-x
PMid:16044341

Komatsu, M., Sumizawa, T., Mutoh, M., Chen, Z.S., Terada, K., Furukawa, T., Yang, X.L., Gao, H., Miura, N., Sugiyama, T. and Akiyama, S. (2000). Copper-transporting P-type adenosine triphosphatase (ATP7B) is associated with cisplatin resistance, Cancer Research 60, 5, 1312-1316.

Kumar, G., Mittal, S., Sak, K. and Tuli, H.S. (2016). Molecular mechanisms underlying chemopreventive potential of curcumin: Current challenges and future perspectives, Life Sci 148, 313-328.
https://doi.org/10.1016/j.lfs.2016.02.022
PMid:26876915

Lee, T.-J., Kim, O.H., Kim, Y.H., Lim, J.H., Kim, S., Park, J.-W. and Kwon, T.K. (2006). Quercetin arrests G2/M phase and induces caspase-dependent cell death in U937 cells, Cancer letters 240, 2, 234-242.
https://doi.org/10.1016/j.canlet.2005.09.013
PMid:16274926

Lee, Y.-K., Hwang, J.-T., Kwon Dae, Y., Surh, Y.-J. and Park Ock, J. (2010). Induction of apoptosis by quercetin is mediated through AMPKalpha1/ASK1/p38 pathway, Cancer letters 292, 2, 228-236.
https://doi.org/10.1016/j.canlet.2009.12.005
PMid:20083342

Lengyel, E. (2010). Ovarian cancer development and metastasis,  1, 3, 1053-1064.
https://doi.org/10.2353/ajpath.2010.100105
PMid:20651229 PMCid:PMC2928939

Li, T.-Y. and Chiang, B.-H. (2017). 6-shogaol induces autophagic cell death then triggered apoptosis in colorectal adenocarcinoma HT-29 cells, Biomedicine & Pharmacotherapy 93, 208-217.
https://doi.org/10.1016/j.biopha.2017.06.038
PMid:28641163

Lin, R.J., Wu, I.J., Hong, J.Y., Liu, B.H., Liang, R.Y., Yuan, T.M. and Chuang, S.M. (2018). Capsaicin-induced TRIB3 upregulation promotes apoptosis in cancer cells, Cancer Manag Res 10, 4237-4248.
https://doi.org/10.2147/CMAR.S162383
PMid:30323679 PMCid:PMC6177521

Lin, Y.G., Kunnumakkara, A.B., Nair, A., Merritt, W.M., Han, L.Y., Armaiz-Pena, G.N., Kamat, A.A., Spannuth, W.A., Gershenson, D.M., Lutgendorf, S.K., Aggarwal, B.B. and Sood, A.K. (2007). Curcumin inhibits tumor growth and angiogenesis in ovarian carcinoma by targeting the nuclear factor-kappaB pathway, Clinical cancer research : an official journal of the American Association for Cancer Research 13, 11, 3423-3430.
https://doi.org/10.1158/1078-0432.CCR-06-3072
PMid:17545551

Ling, H., Yang, H., Tan, S.H., Chui, W.K. and Chew, E.H. (2010). 6-Shogaol, an active constituent of ginger, inhibits breast cancer cell invasion by reducing matrix metalloproteinase-9 expression via blockade of nuclear factor-κB activation, British journal of pharmacology 161, 8, 1763-1777.
https://doi.org/10.1111/j.1476-5381.2010.00991.x
PMid:20718733 PMCid:PMC3010581

Liu, H., Zang, C., Emde, A., Planas-Silva, M.D., Rosche, M., Kühnl, A., Schulz, C.-O., Elstner, E., Possinger, K. and Eucker, J. (2008). Anti-tumor effect of honokiol alone and in combination with other anti-cancer agents in breast cancer, European Journal of Pharmacology 591, 1, 43-51.
https://doi.org/10.1016/j.ejphar.2008.06.026
PMid:18588872

Liu, Q., Peng, Y.-B., Zhou, P., Qi, L.-W., Zhang, M., Gao, N., Liu, E.H. and Li, P. (2013). 6-Shogaol induces apoptosis in human leukemia cells through a process involving caspase-mediated cleavage of eIF2α, Molecular Cancer 12, 135-135.
https://doi.org/10.1186/1476-4598-12-135
PMid:24215632 PMCid:PMC4176122

Lowe, S.W., Cepero, E. and Evan, G. (2004). Intrinsic tumor suppression, Nature (London, United Kingdom) 432, 7015, 307-315.
https://doi.org/10.1038/nature03098
PMid:15549092

Lugli, E., Troiano, L., Ferraresi, R., Roat, E., Prada, N., Nasi, M., Pinti, M., Cooper, E.L. and Cossarizza, A. (2005). Characterization of cells with different mitochondrial membrane potential during apoptosis, Cytometry Part A 68, 1, 28-35.
https://doi.org/10.1002/cyto.a.20188
PMid:16184612

Majdalawieh, A.F., Fayyad, M.W. and Nasrallah, G.K. (2017). Anti-cancer properties and mechanisms of action of thymoquinone, the major active ingredient of Nigella sativa, Crit Rev Food Sci Nutr 57, 18, 3911-3928.
https://doi.org/10.1080/10408398.2016.1277971
PMid:28140613

Marquez, R.T., Baggerly, K.A., Patterson, A.P., Liu, J., Broaddus, R., Frumovitz, M., Atkinson, E.N., Smith, D.I., Hartmann, L., Fishman, D., Berchuck, A., Whitaker, R., Gershenson, D.M., Mills, G.B., Bast, R.C., Jr. and Lu, K.H. (2005). Patterns of gene expression in different histotypes of epithelial ovarian cancer correlate with those in normal fallopian tube, endometrium, and colon, Clinical Cancer Research 11, 17, 6116-6126.
https://doi.org/10.1158/1078-0432.CCR-04-2509
PMid:16144910

Marsden, D.E., Friedlander, M. and Hacker, N.F. (2000). Current management of epithelial ovarian carcinoma: a review, Semin Surg Oncol 19, 1, 11-19.
https://doi.org/10.1002/1098-2388(200007/08)19:1<11::AID-SSU3>3.0.CO;2-3

Mathieson, R.A. and Kitts, W.D. (1980). Binding of phytoestrogen and estradiol-17β by cytoplasmic receptors in the pituitary gland and hypothalamus of the ewe, Journal of Endocrinology 85, 2, 317-325.
https://doi.org/10.1677/joe.0.0850317
PMid:7400718

Mattson David, M., Ahmad Iman, M., Dayal, D., Parsons Arlene, D., Aykin-Burns, N., Li, L., Orcutt Kevin, P., Spitz Douglas, R., Dornfeld Kenneth, J. and Simons Andrean, L. (2009). Cisplatin combined with zidovudine enhances cytotoxicity and oxidative stress in human head and neck cancer cells via a thiol-dependent mechanism, Free radical biology & medicine 46, 2, 232-237.
https://doi.org/10.1016/j.freeradbiomed.2008.10.023
PMid:18983911 PMCid:PMC2659778

Mazumder, M.E.H. (2013). Studies on New Tumour Active Palladium Complexes Targeted to Overcome Resistance in Ovarian Cancer PhD Thesis, University of Sydney, Australia.

Mazumder, M.E.H., Beale, P., Chan, C., Yu, J.Q. and Huq, F. (2012). Epigallocatechin gallate acts synergistically in combination with cisplatin and designed trans-palladiums in ovarian cancer cells, Anticancer Res 32, 11, 4851-4860.

McNeil, E.M. and Melton, D.W. (2012). DNA repair endonuclease ERCC1-XPF as a novel therapeutic target to overcome chemoresistance in cancer therapy, Nucleic acids research.
https://doi.org/10.1093/nar/gks818
PMid:22941649 PMCid:PMC3488251

Melet, A., Song, K., Bucur, O., Jagani, Z., Grassian, A.R. and Khosravi-Far, R. (2008). Apoptotic pathways in tumor progression and therapy, Advances in Experimental Medicine and Biology 615, Programmed Cell Death in Cancer Progression and Therapy, 47-79.
https://doi.org/10.1007/978-1-4020-6554-5_4
PMid:18437891

Monguilhott Dalmarco, E., Mendes de Córdova, C.M. and Fröde, T.S. (2011). Evidence of an anti-inflammatory effect of mycophenolate mofetil in a murine model of pleurisy, Experimental Lung Research 37, 7, 399-407.
https://doi.org/10.3109/01902148.2011.570416
PMid:21777147

Morales, R.R., Agrapart, V., Mencacci, C., Moretti, C., Frajese, G. and Frajese, G.V. (2008). Functional re-differentiation of prostate cancer derived cell lines by the anti-tumoral drug Mycophenolic Acid (MPA), European Journal of Cancer Supplements 6, 9, 146.
https://doi.org/10.1016/S1359-6349(08)71737-3

Morgan, D.O. (1997). Cyclin-dependent kinases: engines, clocks, and microprocessors, Annual Review of Cell and Developmental Biology 13, 261-291.
https://doi.org/10.1146/annurev.cellbio.13.1.261
PMid:9442875

Narod, S.A. (2002). Modifiers of risk of hereditary breast and ovarian cancer, Nature Reviews Cancer 2, 2, 113-123.
https://doi.org/10.1038/nrc726
PMid:12635174

Nessa, M.U. (2013). Studies on Combinations Between Platinum Drugs and Phytochemicals in Ovarian Tumour Models, University of Sydney, Australia.

Nessa, M.U., Beale, P., Chan, C., Yu, J.Q. and Huq, F. (2011). Synergism from combinations of cisplatin and oxaliplatin with quercetin and thymoquinone in human ovarian tumour models, Anticancer Res 31, 11, 3789-3797.

Nessa, M.U., Beale, P., Chan, C., Yu, J.Q. and Huq, F. (2012). Combinations of resveratrol, cisplatin and oxaliplatin applied to human ovarian cancer cells, Anticancer Res 32, 1, 53-59.

Nessa, M.U., Beale, P., Chan, C., Yu, J.Q. and Huq, F. (2012). Studies on combination of platinum drugs cisplatin and oxaliplatin with phytochemicals anethole and curcumin in ovarian tumour models, Anticancer Res 32, 11, 4843-4850.

Noguchi, K., Kitanaka, C., Yamana, H., Kokubu, A., Mochizuki, T. and Kuchino, Y. (1999). Regulation of c-Myc through phosphorylation at Ser-62 and Ser-71 by c-Jun N-terminal kinase, Journal of Biological Chemistry 274, 46, 32580-32587.
https://doi.org/10.1074/jbc.274.46.32580
PMid:10551811

Norbury, C. and Nurse, P. (1992). Animal cell cycles and their control, Annual Review of Biochemistry 61, 441-470.
https://doi.org/10.1146/annurev.bi.61.070192.002301
PMid:1497317

Norman, A.W. (2008). From vitamin D to hormone D: fundamentals of the vitamin D endocrine system essential for good health, The American journal of clinical nutrition 88, 2, 491S-499S.
https://doi.org/10.1093/ajcn/88.2.491S
PMid:18689389

Nurse, P., Masui, Y. and Hartwell, L. (1998). Understanding the cell cycle: past, present and future, Nature Medicine (New York) 4, 10, 1103-1106.
https://doi.org/10.1038/2594
PMid:9771732

Ok, S. and Jeong, W.-S. (2012). Optimization of Extraction Conditions for the 6-Shogaol-rich Extract from Ginger (Zingiber officinale Roscoe), Preventive Nutrition and Food Science 17, 2, 166-171.
https://doi.org/10.3746/pnf.2012.17.2.166
PMid:24471079 PMCid:PMC3866747

Okura, A., Arakawa, H., Oka, H., Yoshinari, T. and Monden, Y. (1988). Effect of genistein on topoisomerase activity and on the growth of [Val 12]Ha-ras-transformed NIH 3T3 cells, Biochemical and Biophysical Research Communications 157, 1, 183-189.
https://doi.org/10.1016/S0006-291X(88)80030-5

Oyagbemi, A., Saba, A. and Azeez, O. (2010). Capsaicin: a novel chemopreventive molecule and its underlying molecular mechanisms of action, Indian journal of cancer 47, 1, 53.
https://doi.org/10.4103/0019-509X.58860
PMid:20071791

Oyagbemi, A.A., Saba, A.B. and Azeez, O.I. (2010). Capsaicin: a novel chemopreventive molecule and its underlying molecular mechanisms of action, Indian J Cancer 47, 1, 53-58.
https://doi.org/10.4103/0019-509X.58860
PMid:20071791

Park, B.S., Lee, K.G., Shibamoto, T., Lee, S.E. and Takeoka, G.R. (2003). Antioxidant activity and characterization of volatile constituents of Taheebo (Tabebuia impetiginosa Martius ex DC), J Agric Food Chem 51, 1, 295-300.
https://doi.org/10.1021/jf020811h
PMid:12502424

Park, M.-T. and Lee, S.-J. (2003). Cell cycle and cancer, Journal of Biochemistry and Molecular Biology 36, 1, 60-65.
https://doi.org/10.5483/BMBRep.2003.36.1.060
PMid:12542976

Ravindran, J., Prasad, S. and Aggarwal Bharat, B. (2009). Curcumin and cancer cells: how many ways can curry kill tumor cells selectively?, Aaps J 11, 3, 495-510.
https://doi.org/10.1208/s12248-009-9128-x
PMid:19590964 PMCid:PMC2758121

Ray, A., Vasudevan, S. and Sengupta, S. (2015). 6-Shogaol inhibits breast cancer cells and stem cell-like spheroids by modulation of Notch signaling pathway and induction of autophagic cell death, PloS one 10, 9, e0137614.
https://doi.org/10.1371/journal.pone.0137614
PMid:26355461 PMCid:PMC4565635

Rocha, C.R.R., Silva, M.M., Quinet, A., Cabral-Neto, J.B. and Menck, C.F.M. (2018). DNA repair pathways and cisplatin resistance: an intimate relationship, Clinics 73, suppl 1, e478s.
https://doi.org/10.6061/clinics/2018/e478s

Roepke, M., Diestel, A., Bajbouj, K., Walluscheck, D., Schonfeld, P., Roessner, A., Schneider-Stock, R. and Gali-Muhtasib, H. (2007). Lack of p53 augments thymoquinone-induced apoptosis and caspase activation in human osteosarcoma cells, Cancer biology & therapy 6, 2, 160-169.
https://doi.org/10.4161/cbt.6.2.3575
PMid:17218778

Rosell, R., Taron, M., Barnadas, A., Scagliotti, G., Sarries, C. and Roig, B. (2003). Nucleotide excision repair pathways involved in Cisplatin resistance in non-small-cell lung cancer, Cancer Control 10, 4, 297-305.
https://doi.org/10.1177/107327480301000404
PMid:12915808

Rovirosa, J., Diaz-Marrero, A., Darías, J., Painemal, K. and San Martín, A. (2006). Secondary metabolites from marine Penicillium brevicompactum, Journal of the Chilean Chemical Society 51, 1, 775-778.
https://doi.org/10.4067/S0717-97072006000100004

Sa, G. and Das, T. (2008). Anti cancer effects of curcumin: cycle of life and death, Cell Division 3, No pp given.
https://doi.org/10.1186/1747-1028-3-14
PMid:18834508 PMCid:PMC2572158

Sablina Anna, A., Budanov Andrei, V., Ilyinskaya Galina, V., Agapova Larissa, S., Kravchenko Julia, E. and Chumakov Peter, M. (2005). The antioxidant function of the p53 tumor suppressor, Nature medicine 11, 12, 1306-1313.
https://doi.org/10.1038/nm1320
PMid:16286925 PMCid:PMC2637821

Saha, A., Blando, J., Silver, E., Beltran, L., Sessler, J. and DiGiovanni, J. (2014). 6-Shogaol from Dried Ginger Inhibits Growth of Prostate Cancer Cells Both and through Inhibition of STAT3 and NF-κB Signaling, Cancer Prevention Research 7, 6, 627-638.
https://doi.org/10.1158/1940-6207.CAPR-13-0420
PMid:24691500

Samimi, G., Safaei, R., Katano, K., Holzer, A.K., Rochdi, M., Tomioka, M., Goodman, M. and Howell, S.B. (2004). Increased expression of the copper efflux transporter ATP7A mediates resistance to cisplatin, carboplatin, and oxaliplatin in ovarian cancer cells, Clinical Cancer Research 10, 14, 4661-4669.
https://doi.org/10.1158/1078-0432.CCR-04-0137
PMid:15269138

Samuel, T., Fadlalla, K., Mosley, L., Katkoori, V., Turner, T. and Manne, U. (2012). Dual-mode interaction between quercetin and DNA-damaging drugs in cancer cells, Anticancer Res 32, 1, 61-71.

Sánchez, A.M., Malagarie-Cazenave, S., Olea, N., Vara, D., Chiloeches, A. and Díaz-Laviada, I. (2007). Apoptosis induced by capsaicin in prostate PC-3 cells involves ceramide accumulation, neutral sphingomyelinase, and JNK activation, Apoptosis 12, 11, 2013-2024.
https://doi.org/10.1007/s10495-007-0119-z
PMid:17828457

Sánchez, A.M., Martínez-Botas, J., Malagarie-Cazenave, S., Olea, N., Vara, D., Lasunción, M.A. and Díaz-Laviada, I. (2008). Induction of the endoplasmic reticulum stress protein GADD153/CHOP by capsaicin in prostate PC-3 cells: a microarray study, Biochemical and Biophysical Research Communications 372, 4, 785-791.
https://doi.org/10.1016/j.bbrc.2008.05.138
PMid:18533110

Sarkar, F.H. and Li, Y.-w. (2007). Targeting multiple signal pathways by chemopreventive agents for cancer prevention and therapy, Acta Pharmacologica Sinica 28, 9, 1305-1315.
https://doi.org/10.1111/j.1745-7254.2007.00689.x
PMid:17723164

Sarkar, F.H., Li, Y., Sarkar, F.H. and Li, Y. (2006). Using chemopreventive agents to enhance the efficacy of cancer therapy, Cancer Research 66, 7, 3347-3350.
https://doi.org/10.1158/0008-5472.CAN-05-4526
PMid:16585150

Satoh, H., Nishikawa, K., Suzuki, K., Asano, R., Virgona, N., Ichikawa, T., Hagiwara, K. and Yano, T. (2003). Genistein, a soy isoflavone, enhances necrotic-like cell death in a breast cancer cell treated with a chemotherapeutic agent, Research Communications in Molecular Pathology and Pharmacology 113-114, 149-158.

Schuijer, M., Berns, E.M., Schuijer, M. and Berns, E.M.J.J. (2003). TP53 and ovarian cancer, Hum Mutat 21, 3, 285-291.
https://doi.org/10.1002/humu.10181
PMid:12619114

Seo, S.S., Song, Y.S., Kang, D.-h., Park, I., Bang, Y.J., Kang, S.B. and Lee, H.P. (2004). Expression of cyclooxygenase-2 in association with clinicopathological prognostic factors and molecular markers in epithelial ovarian cancer, Gynecologic oncology 92, 3, 927-935.
https://doi.org/10.1016/j.ygyno.2003.11.055
PMid:14984962

Sexton, E., Van Themsche, C., LeBlanc, K., Parent, S., Lemoine, P., Asselin, E., Sexton, E., Van Themsche, C., LeBlanc, K., Parent, S., Lemoine, P. and Asselin, E. (2006). Resveratrol interferes with AKT activity and triggers apoptosis in human uterine cancer cells, Molecular Cancer 5, 45.
https://doi.org/10.1186/1476-4598-5-45
PMid:17044934 PMCid:PMC1626081

Sherr, C.J. (2000). The Pezcoller lecture: cancer cell cycles revisited, Cancer Research 60, 14, 3689-3695.

Sherr, C.J. and Roberts, J.M. (1999). CDK inhibitors: positive and negative regulators of G1-phase progression, Genes & Development 13, 12, 1501-1512.
https://doi.org/10.1101/gad.13.12.1501
PMid:10385618

Siddik, Z.H. (2003). Cisplatin: mode of cytotoxic action and molecular basis of resistance, Oncogene 22, 47, 7265-7279.
https://doi.org/10.1038/sj.onc.1206933
PMid:14576837

Singh, N., Nigam, M., Ranjan, V., Sharma, R., Balapure, A.K. and Rath, S.K. (2009). Caspase mediated enhanced apoptotic action of cyclophosphamide- and resveratrol-treated MCF-7 cells, Journal of Pharmacological Sciences 109, 4, 473-485.
https://doi.org/10.1254/jphs.08173FP
PMid:19372630

Singh, S., Khan, A.R. and Gupta, A.K. (2012). Role of glutathione in cancer pathophysiology and therapeutic interventions, Journal of Experimental Therapeutics and Oncology 9, 4, 303-316.

Solomon, L.A., Ali, S., Banerjee, S., Munkarah, A.R., Morris, R.T. and Sarkar, F.H. (2008). Sensitization of ovarian cancer cells to cisplatin by genistein: the role of NF-kappaB, J Ovarian Res 1, No pp given.
https://doi.org/10.1186/1757-2215-1-9
PMid:19025644 PMCid:PMC2611983

Sozzani, R., Maggio, C., Varotto, S., Canova, S., Bergounioux, C., Albani, D. and Cella, R. (2006). Interplay between Arabidopsis activating factors E2Fb and E2Fa in cell cycle progression and development, Plant Physiol 140, 4, 1355-1366.
https://doi.org/10.1104/pp.106.077990
PMid:16514015 PMCid:PMC1435807

Sun, X.-X., Dai, M.-S. and Lu, H. (2008). Mycophenolic acid activation of p53 requires ribosomal proteins L5 and L11, Journal of Biological Chemistry 283, 18, 12387-12392.
https://doi.org/10.1074/jbc.M801387200
PMid:18305114 PMCid:PMC2430998

Surh, Y.-J. (2003). Cancer chemoprevention with dietary phytochemicals, Nature Reviews Cancer 3, Copyright
https://doi.org/10.1038/nrc1189
PMid:14570043

(C) 2012 American Chemical Society (ACS). All Rights Reserved., 768-780.

Surowiak, P., Materna, V., Kaplenko, I., Spaczynski, M., Dolinska-Krajewska, B., Gebarowska, E., Dietel, M., Zabel, M. and Lage, H. (2006). ABCC2 (MRP2, cMOAT) can be localized in the nuclear membrane of ovarian carcinomas and correlates with resistance to cisplatin and clinical outcome, Clinical Cancer Research 12, 23, 7149-7158.
https://doi.org/10.1158/1078-0432.CCR-06-0564
PMid:17145840

Suzuki, K., Koike, H., Matsui, H., Ono, Y., Hasumi, M., Nakazato, H., Okugi, H., Sekine, Y., Oki, K. and Ito, K. (2002). Genistein, a soy isoflavone, induces glutathione peroxidase in the human prostate cancer cell lines LNCaP and PC-3, International Journal of Cancer 99, 6, 846-852.
https://doi.org/10.1002/ijc.10428
PMid:12115487

Tan, B.S., Kang, O., Mai, C.W., Tiong, K.H., Khoo, A.S.-B., Pichika, M.R., Bradshaw, T.D. and Leong, C.-O. (2013). 6-Shogaol inhibits breast and colon cancer cell proliferation through activation of peroxisomal proliferator activated receptor γ (PPARγ), Cancer letters 336, 1, 127-139.
https://doi.org/10.1016/j.canlet.2013.04.014
PMid:23612072

Tan, X., Sidell, N., Mancini, A., Huang, R.-P., Wang, S., Horowitz, I.R., Liotta, D.C., Taylor, R.N. and Wieser, F. (2010). Multiple anticancer activities of EF24, a novel curcumin analog, on human ovarian carcinoma cells, Reproductive Sciences 17, 10, 931-940.
https://doi.org/10.1177/1933719110374239
PMid:20693500

Tan, Y.M., Yu, R. and Pezzuto, J.M. (2003). Betulinic acid-induced programmed cell death in human melanoma cells involves mitogen-activated protein kinase activation, Clinical Cancer Research 9, 7, 2866-2875.

Tokar, E.J. and Webber, M.M. (2005). Chemoprevention of Prostate Cancer by Cholecalciferol (Vitamin D3): 25-hydroxylase (CYP27A1) in Human Prostate Epithelial Cells, Clinical & Experimental Metastasis 22, 3, 265.
https://doi.org/10.1007/s10585-005-8394-y
PMid:16158254

Tournier, C., Hess, P., Yang, D.D., Xu, J., Turner, T.K., Nimnual, A., Bar-Sagi, D., Jones, S.N., Flavell, R.A. and Davis, R.J. (2000). Requirement of JNK for stress-induced activation of the cytochrome c-mediated death pathway, Science 288, 5467, 870-874.
https://doi.org/10.1126/science.288.5467.870
PMid:10797012

Trachootham, D., Alexandre, J. and Huang, P. (2009). Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach?, Nature reviews. Drug discovery 8, 7, 579-591.
https://doi.org/10.1038/nrd2803
PMid:19478820

Trachootham, D., Lu, W., Ogasawara, M.A., Valle, N.R.-D. and Huang, P. (2008). Redox Regulation of Cell Survival, Antioxid Redox Signal 10, 8, 1343-1374.
https://doi.org/10.1089/ars.2007.1957
PMid:18522489 PMCid:PMC2932530

Valent, P., Bonnet, D., De Maria, R., Lapidot, T., Copland, M., Melo, J.V., Chomienne, C., Ishikawa, F., Schuringa, J.J., Stassi, G., Huntly, B., Herrmann, H., Soulier, J., Roesch, A., Schuurhuis, G.J., Wohrer, S., Arock, M., Zuber, J., Cerny-Reiterer, S., Johnsen, H.E., Andreeff, M. and Eaves, C. (2012). Cancer stem cell definitions and terminology: the devil is in the details, Nat Rev Cancer 12, 11, 767-775.
https://doi.org/10.1038/nrc3368
PMid:23051844

Wang, Q.-E., Milum, K., Han, C., Huang, Y.-W., Wani, G., Thomale, J. and Wani, A.A. (2011). Differential contributory roles of nucleotide excision and homologous recombination repair for enhancing cisplatin sensitivity in human ovarian cancer cells, Molecular Cancer 10, 24.
https://doi.org/10.1186/1476-4598-10-24
PMid:21385444 PMCid:PMC3064653

Wang, T., Chen, F., Chen, Z., Wu, Y.-F., Xu, X.-L., Zheng, S. and Hu, X. (2004). Honokiol induces apoptosis through p53-independent pathway in human colorectal cell line RKO, World Journal of Gastroenterology 10, 15, 2205-2208.
https://doi.org/10.3748/wjg.v10.i15.2205
PMid:15259066 PMCid:PMC4724979

Wei, Y., Pu, X. and Zhao, L. (2017). Preclinical studies for the combination of paclitaxel and curcumin in cancer therapy (Review), Oncol Rep 37, 6, 3159-3166.
https://doi.org/10.3892/or.2017.5593
PMid:28440434

Williamson, J.M., Boettcher, B. and Meister, A. (1982). Intracellular cysteine delivery system that protects against toxicity by promoting glutathione synthesis, Proceedings of the National Academy of Sciences of the United States of America 79, 20, 6246-6249.
https://doi.org/10.1073/pnas.79.20.6246
PMid:6959113 PMCid:PMC347097

Wu, S. and Sun, J. (2011). Vitamin D, Vitamin D Receptor, and Macroautophagy in Inflammation and Infection, Discovery Medicine 11, 59, 325-335.

Xie, D., Zheng, G.Z., Xie, P., Zhang, Q.H., Lin, F.X., Chang, B., Hu, Q.X., Du, S.X. and Li, X.D. (2017). Antitumor activity of resveratrol against human osteosarcoma cells: a key role of Cx43 and Wnt/beta-catenin signaling pathway, Oncotarget 8, 67, 111419-111432.
https://doi.org/10.18632/oncotarget.22810
PMid:29340064 PMCid:PMC5762332

Yamamoto, K., Ichijo, H. and Korsmeyer, S.J. (1999). BCL-2 is phosphorylated and inactivated by an ASK1/Jun N-terminal protein kinase pathway normally activated at G2/M, Molecular and cellular biology 19, 12, 8469-8478.
https://doi.org/10.1128/MCB.19.12.8469
PMid:10567572 PMCid:PMC84954

Yang, J.H., Hsia, T.C., Kuo, H.M., Chao, P.D.L., Chou, C.C., Wei, Y.H. and Chung, J.G. (2006). Inhibition of lung cancer cell growth by quercetin glucuronides via G2/M arrest and induction of apoptosis, Drug metabolism and disposition 34, 2, 296-304.
https://doi.org/10.1124/dmd.105.005280
PMid:16280456

Yang, K., Lamprecht, S.A., Shinozaki, H., Fan, K., Yang, W., Newmark, H.L., Kopelovich, L., Edelmann, W., Jin, B. and Gravaghi, C. (2008). Dietary calcium and cholecalciferol modulate cyclin D1 expression, apoptosis, and tumorigenesis in intestine of adenomatous polyposis coli1638N/+ mice, The Journal of nutrition 138, 9, 1658-1663.
https://doi.org/10.1093/jn/138.9.1658
PMid:18716166

Yang, P., Ebbert Jon, O., Sun, Z. and Weinshilboum Richard, M. (2006). Role of the glutathione metabolic pathway in lung cancer treatment and prognosis: a review, Journal of clinical oncology : official journal of the American Society of Clinical Oncology 24, 11, 1761-1769.
https://doi.org/10.1200/JCO.2005.02.7110
PMid:16603718

Zhan, Q., Wang, C. and Ngai, S. (2013). Ovarian cancer stem cells: a new target for cancer therapy, Biomed Res Int 2013, 916819.
https://doi.org/10.1155/2013/916819
PMid:23509802 PMCid:PMC3581273

Zheng, A.W., Chen, Y.Q., Zhao, L.Q. and Feng, J.G. (2017). Myricetin induces apoptosis and enhances chemosensitivity in ovarian cancer cells, Oncol 13, 6, 4974-4978.
https://doi.org/10.3892/ol.2017.6031
PMid:28588737 PMCid:PMC5452908

Zheng, Y., Liu, H. and Liang, Y. (2017). Genistein exerts potent antitumour effects alongside anaesthetic, propofol, by suppressing cell proliferation and nuclear factor-kappaB-mediated signalling and through upregulating microRNA-218 expression in an intracranial rat brain tumour model, J Pharm Pharmacol 69, 11, 1565-1577.
https://doi.org/10.1111/jphp.12781
PMid:28776680

Zheng, Z.-H., Yang, Y., Lu, X.-H., Zhang, H., Shui, X.-X., Liu, C., He, X.-B., Jiang, Q., Zhao, B.-H. and Si, S.-Y. (2011). Mycophenolic acid induces adipocyte-like differentiation and reversal of malignancy of breast cancer cells partly through PPARγ, European Journal of Pharmacology 658, 1, 1-8.
https://doi.org/10.1016/j.ejphar.2011.01.068
PMid:21349264

Zhou, J., Zhang, C., Zhou, B. and Jiang, D. (2019). miR-183 modulated cell proliferation and apoptosis in ovarian cancer through the TGF-beta/Smad4 signaling pathway, Int J Mol Med 43, 4, 1734-1746.
https://doi.org/10.3892/ijmm.2019.4082

Zong, W.-X., Edelstein, L.C., Chen, C., Bash, J. and Gélinas, C. (1999). The prosurvival Bcl-2 homolog Bfl-1/A1 is a direct transcriptional target of NF-κB that blocks TNFα-induced apoptosis, Genes Dev. 13, 4, 382-387.
https://doi.org/10.1101/gad.13.4.382
PMid:10049353 PMCid:PMC316475

Committee on Publication Ethics

PDF
Supplementary Material
Abstract
Export Citation

View Dimensions


View Plumx


View Altmetric




Save
0
Citation
752
View

Share