Next Article in Journal
In Vitro Activities of LCB 01-0648, a Novel Oxazolidinone, against Gram-Positive Bacteria
Next Article in Special Issue
Induction of G2M Arrest by Flavokawain A, a Kava Chalcone, Increases the Responsiveness of HER2-Overexpressing Breast Cancer Cells to Herceptin
Previous Article in Journal
Glycerol as Precursor of Organoselanyl and Organotellanyl Alkynes
Previous Article in Special Issue
Impact of Age and Insulin-Like Growth Factor-1 on DNA Damage Responses in UV-Irradiated Human Skin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cancer Chemoprevention by Phytochemicals: Nature’s Healing Touch

1
Department of Oncologic Sciences, Mitchell Cancer Institute, University of South Alabama, Mobile, AL 36604, USA
2
Department of Molecular Biology and Biochemistry, College of Medicine, University of South Alabama, Mobile, AL 36688, USA
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(3), 395; https://doi.org/10.3390/molecules22030395
Submission received: 24 January 2017 / Revised: 27 February 2017 / Accepted: 28 February 2017 / Published: 3 March 2017
(This article belongs to the Special Issue Cancer Chemoprevention)

Abstract

:
Phytochemicals are an important part of traditional medicine and have been investigated in detail for possible inclusion in modern medicine as well. These compounds often serve as the backbone for the synthesis of novel therapeutic agents. For many years, phytochemicals have demonstrated encouraging activity against various human cancer models in pre-clinical assays. Here, we discuss select phytochemicals—curcumin, epigallocatechin-3-gallate (EGCG), resveratrol, plumbagin and honokiol—in the context of their reported effects on the processes of inflammation and oxidative stress, which play a key role in tumorigenesis. We also discuss the emerging evidence on modulation of tumor microenvironment by these phytochemicals which can possibly define their cancer-specific action. Finally, we provide recent updates on how low bioavailability, a major concern with phytochemicals, is being circumvented and the general efficacy being improved, by synthesis of novel chemical analogs and nanoformulations.

1. Introduction

Cancer is currently the second leading cause of death worldwide and a major health problem throughout the world. It is estimated that in 2017, the United States alone will have 1,688,780 new cancer diagnoses and 600,920 cancer-related deaths [1]. While significant progress has been made to improve diagnosis and surveillance, this has not helped much to improve the overall cancer survival rates. This had led to a surge in molecular targeted therapies to develop better clinical outcomes for cancer patients [2]. Unfortunately, these strategies have also not provided substantial improvement due to the development of resistance against therapies [3], reviving an interest in the prospects of phytochemicals, natural anticancer agents from plants, due to the multitude of effects of these agents on diverse molecular signaling pathways, with no or minimal toxicity in normal cells [4].
It is believed that the number of new cancer cases can be reduced and many cancer-related deaths can be prevented. The studies focused on ‘cancer prevention’ are a step in this direction. The overall aims of cancer prevention are broad. The primary aim is to completely prevent or at least delay the onset of cancer through maintenance of healthy lifestyle, avoidance of exposure to toxicants/carcinogens, and dietary consumption of chemopreventive agents and drugs. The secondary aim depends on the early detection of cancer in the precancerous or early stage tumors, thereby helping in the better management and treatment of these tumors, and the tertiary aim of cancer prevention involves reducing the risk of metastases, development of secondary tumors and recurrence, using preventive agents. Natural products (from botanicals, herbs, etc.) as well as minerals and vitamins have been demonstrated to affect all three areas of cancer prevention [5,6]. Among natural products, phytochemicals represent the class of compounds that have been extensively studied for their biological effects.
Consumption of fruits and herbal medicines in the diet is a convenient and effective method of administering phytochemicals in a cost-effective manner [7,8]. Overall, at least 20% of all cancers can be prevented by consumption of diets rich in vegetables and fruits (>400 g/day) [9]. According to an analysis, of all the 175 small molecules approved for cancer therapy from 1940s to the year 2014, 85 (49%) were natural products or directly derived therefrom [4]. Phytochemicals continue to enter clinical trials or provide leads for the synthesis of semi-synthetic medicinal agents. However, the laboratory success of phytochemicals has not been reproduced in the clinic. This is partly due to the inherent differences between in vitro laboratory experiments and the human physiological conditions. As with other drugs, improvements in formulation of compounds with potential anticancer properties are being made for better therapeutic outcomes. In this article, we briefly discuss the mechanism of action of phytochemicals and, instead of touching upon the classical activities such as induction of apoptosis/cell cycle arrest, we discuss the effect of phytochemicals on reactive oxygen species (ROS) production and the tumor microenvironment (TME), along with updates on current studies to improve the efficacy of these compounds.

2. Preventive and Therapeutic Mechanisms of Phytochemicals

The proposed mechanisms by which vegetables and fruits affect human cancers are multiple and complex. Various stages of carcinogenesis may be inhibited, and various in vitro or in vivo systems are used to model these inhibitory effects in preclinical studies. Therefore, characterizing the active chemical components of these plant products and accumulation of compelling in vitro and animal study data prior to clinical studies is necessary. Phytochemicals, due to their dietary origin, are presumed safer and are better tolerated with relatively low toxicity. For the ease of discussion and to keep the discussion focused, we have chosen five representative phytochemicals, viz. curcumin, epigallocatechin-3-gallate (EGCG), resveratrol, plumbagin and honokiol. Chemical structures of these phytochemicals are provided in Figure 1.

2.1. Modulation of Oxidative Stress

Healthy cells maintain an intricate balance of redox homeostasis wherein the levels of ROS and reactive nitrogen species (RNS) are fine-tuned by the antioxidant defense system [10]. ROS/RNS are by-products of normal cell metabolism with physiological roles at moderate and low concentrations [10]. Oxidative stress is induced as a result of imbalance in the redox status of the cells, and has been suggested to be responsible for the pathophysiology of several diseases, including cancer. In cancer, oxidative stress by physical and chemical agents, inflammation and infection can give rise to direct DNA damage inducing tumorigenesis [11]. Oxidative stress also alters protein conformation and function, thus affecting regular functions of affected proteins. Phytochemicals are generally known for their antioxidant activity and have been demonstrated to counteract the damaging effects of oxidation in vitro by a direct quenching of ROS. In addition, phytochemicals also upregulate the expression of genes that detoxify reactive species, metabolize toxic compounds, and maintain cellular homeostasis.
The ability of phytochemicals to inhibit chemically-induced carcinogenesis in mice models has been a cornerstone for their chemopreventive property. A delay in tumor promotion has been shown, upon concurrent topical application of 12-O-tetradecanoylphorbol-13-acetate (TPA) and phenolic compounds such as caffeic acid, chlorogenic acid, curcumin and ferulic acid [12], or resveratrol and ursolic acid [13]. TPA is a commonly used phorbol diester, a known tumor promoter, employed to over-activate the PKC signaling, and is a potent generator of superoxide anions [14]. Thus, the ability of phytochemicals to inhibit/delay the progression of TPA-induced carcinogenesis is indicative of their antioxidant effects. Plumbagin [15] and honokiol [16] have also been reported to inhibit TPA-induced effects. In addition, the promotion and progression of 4-nitroquinoline-1-oxide (4-NQO)-induced tongue carcinogenesis in rats was inhibited by these polyphenols when administered in the diet [17]. 4-NQO is a carcinogenic agent that introduces DNA damage through the production of ROS [18]. Curcumin, the most bioactive constituent of turmeric and an integral part of the Indian diet, inhibits diethylnitrosamine (DEN)-induced hepatocarcinogenesis in mice at a concentration of 0.2% in the diet [19]. DEN specifically induces carcinogenesis of the gastrointestinal tract, especially liver, through the downregulation of ROS-detoxifying enzymes [20]. Green tea and black tea constituents have also demonstrated potential inhibition of 7,12-dimethylbenz[a]anthracene (DMBA)-treated UVB-induced skin carcinogenesis [21]. Administration of decaffeinated green or black tea to mice has been reported to significantly reduce 4-(methylnitrosamine)-1-(3-pyridyl)-1 butanone-induced tumor formation in mice [22].

2.2. Inhibition of Inflammation

The persistent inflammation and inflammatory mechanisms have been implicated as the basis of several diseases, such as chronic conditions of old age [23]. Studies have also provided significant evidence to demonstrate a positive relation between inflammation and cancer [24,25]. Chronic inflammatory states are triggered by multiple factors such as microbial infections, obesity, autoimmune diseases, etc. [26]. Underlying infections or inflammatory responses have been linked to nearly 15%–20% of cancer-associated deaths [27]. A number of phytochemicals have been suggested to interfere with inflammation-related pathways, partially explaining their anti-cancer potential [28] (Table 1). Curcumin is a well-known anti-inflammatory agent [29]. Its effect on inflammation is mediated by multiple mechanisms, such as transforming growth factor beta1 (TGF-β1) up-regulation and down-regulation of inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX2) [30], as well as by modulation of toll-like receptors (TLR)/interleukin-1 receptor (IL-1R) pathway [31]. Resveratrol, another known anti-cancer agent [32,33], also works as an anti-inflammatory agent [34]. It inhibits the production of pro-inflammatory cytokines [interleukin (IL)-6/8 and tumor necrosis factor-alpha (TNF-α)] and pro-inflammatory miR-155, while also inducing the anti-inflammatory cytokines and miR-663 [35]. Resveratrol has also been shown to interfere with pro-inflammatory events triggered by IL1-β [36].
A number of signaling pathways play a critical role in inflammation, and often connect inflammation with cancer onset and progression. Signaling through nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) is one well studied pathway that links inflammation with cancer [49]. It also happens to be the pathway that has been investigated in detail, with respect to the anticancer effects of phytochemicals [45,50,51,52]. Almost all phytochemicals have been shown to affect NF-κB signaling, including curcumin, resveratrol, EGCG, plumbagin and honokiol [42,44,45,46,47,48,53]. The Kelch-like erythroid cell-derived protein with Cap’n’collar homology-associated protein (Keap1)/NF-E2 p45-related factor 2 (Nrf2) pathway is another pathway that regulates inflammation-related gene expression [54]. Nrf2 coordinates antioxidant response to several stimuli, thereby preventing oxidative damage and onset of inflammation-responsive diseases, including cancer. Keap1, under normal conditions, keeps Nrf2 sequestered in the cytoplasm thus preventing Nrf2 from discharging its antioxidant functions. It is therefore proposed that targeting of Keap1 by phytochemicals can favorably induce the antioxidant activity of Nrf2. Indeed, several phytochemicals are well known to target Keap1/Nrf2 pathway [30,37,39,40,41,55].

2.3. Prooxidant Activity

While antioxidant activity of phytochemicals has traditionally been at the forefront of investigations into their putative anticancer action, interestingly, many of these agents also exhibit prooxidant action, particularly in the presence of transition metal ions, especially copper [33,56,57]. There seems to be enough evidence to support phytochemicals-mediated production of ROS, a prooxidant action that is responsible for their ability to induce apoptosis in cancer cells [58,59]. Several phytochemicals that are antioxidants at some concentrations become prooxidants at other concentrations. The relevance of copper in the prooxidant action of phytochemicals stems from the observation that preneoplastic and neoplastic cells have elevated copper levels, compared to normal cells [60]. Such cancer cells with endogenously elevated copper levels are much more sensitive to electron transfer, and generation of ROS, presumably through redox recycling of copper ions [32,56]. Therefore, the prooxidant action of phytochemicals, in the presence of redox active transition ions, represents an important pathway through which transformed cells are selectively targeted by phytochemicals while normal cells survive. Such prooxidant action has been demonstrated for many phytochemicals, including curcumin [61,62], EGCG [59], plumbagin [63] and resveratrol [32]. Copper is not the only metal ion with associated prooxidant activity. Iron [64] and zinc ions [65] have also been reported to generate ROS leading to prooxidant activity. Interestingly, the anticancer activity of vitamin C [66,67] has also been reported to involve prooxidant production of ROS. Further, a role of vitamin C in Fenton reaction type production of ROS, involving redox recycling on iron ions, is well known [68]. An investigation into prooxidant activity of known antioxidants and vitamins reported prooxidant potential of vitamins A and C in synergy with iron and copper ions, with combination of vitamin C and copper being the most effective [69]. Based on these evidences, it is conceivable that the prooxidant mechanism of induction of apoptosis better explains the anticancer effects of phytochemicals, through generation of ROS close to the DNA. It also helps explain how anticancer agents with diverse chemical structures function similarly, and exhibit preferential cytotoxicity against cancer cells.

2.4. Modulation of Tumor Metabolism

Metabolism in tumor cells is intricately connected with ROS production [70]. Cancer cells extensively alter their metabolic activity to meet the growing needs associated with survival and growth [71]. The two most important metabolites supporting tumor growth are glucose and glutamine [72]. Breakdown of glucose supports energy demand and the generation of biosynthetic metabolites for the growth of cancer cells. In addition, glutamine, the most abundant amino acid in human plasma, not only contributes to the energy pool, but also provides nitrogen for the biosynthesis of nitrogen-containing compounds. Thus, targeting the altered tumor cell metabolism can yield therapeutic outcomes.
A number of studies have now identified the regulation of glucose and glutamine metabolism by phytochemicals through different mechanisms [73,74]. Phytochemicals have been observed to directly inhibit the basal transport of glucose in cancer cells to improve their response to chemotherapy [75]. Curcumin, in addition to glucose uptake, has been shown to alter glutathione as well as lipid metabolism in association with docetaxel [76]. The ability of curcumin to interfere with glucose transport can be explained by its direct binding [77] and inhibition of glucose transporter 1 (GLUT1) [78]. Similar to curcumin, resveratrol [79] and plumbagin [80] can also down-regulate GLUT1. As a proof that glucose metabolism can be effectively targeted in cancer cells, particularly in the cancer cells with metastatic mesenchymal phenotype, glucose-coated magnetic nanoparticles were observed to be preferentially taken up by mesenchymal cells, as opposed to epithelial cells [81]. Blocking of GLUT1 affected the uptake of nanoparticles, suggesting an important role of glucose shell on the nanoparticles. In addition to curcumin [82,83] and resveratrol [83,84], EGCG [85,86], plumbagin [87] and honokiol [88] have also been reported to alter glucose metabolism leading to anti-cancer effects. Given the relevance of glutamine metabolism in cancer cells, phytochemicals have also been tested for their ability to modulate glutamine metabolism. It was reported that resveratrol-induced cell death in castration-resistant C4-2 prostate cancer cells depends on glutamine metabolism [89]. Further, curcumin’s cytotoxicity against colorectal cancer stem cells was suggested to involve blocking of glutamine’s entry into the cells though coupling of curcumin to the CD44 receptors [90]. Thus, there is evidence in literature to support an effect of phytochemicals on glucose and glutamine metabolism as a mechanism for their anti-tumor properties.
The predominance of aerobic glycolysis in cancer cells, the ‘Warburg effect [91]’ has generated a lot of attention in last several decades. Interestingly, phytochemicals have been shown to reverse this phenomenon. Curcumin could reverse inflammatory TNF-α mediated Warburg effect in breast cancer cells [92]. Resveratrol was recently shown to partially reverse Warburg effect resulting in increased cell death through increased oxygen consumption, hyperpolarization of mitochondrial membrane and ROS generation [93]. While the contribution of ‘Warburg effect’ to tumorigenesis has been challenged in recent years [94], particularly with the realization that a number of tumor cells have functional mitochondria, the overall role of altered metabolism in tumor cells still remains an attractive target for therapy [95]. With multi-targeted effects against different metabolic pathways, phytochemicals are prime candidates for further testing in clinical settings.

3. Modulation of Tumor Microenvironment

The tumor microenvironment (TME) refers to the immediate vicinity of tumor cells that is populated by many different types of cells/factors: immune cells, fibroblasts, cytokines, blood vessels, etc. The dynamic interactions between several components within the TME are now considered drivers of cancer progression [96]. The bi-directional talk between tumor cells and the surroundings in the TME facilitates their proliferation, invasion and metastasis, in addition to bestowing upon them the ability to evade therapeutic insults. In view of the important role that the TME plays in tumor progression, it is critical for any putative therapy to be able to modulate the TME favorably, countering the many advantages that the TME confers to the tumor cells. There are reports documenting the effects of phytochemicals on the TME as basis of their anti-cancer activity [97,98].
In pancreatic cancer the TME is particularly marked by severe hypoxia, which, in turn, triggers the activation of hedgehog (Hh) signaling [99,100]. The cross-talk between tumor cells and the surrounding stroma in the hypoxic TME has a profound dependence on Hh signaling [100] making Hh signaling an attractive target for therapy. A recent report has documented the inhibitory action of curcumin against Hh signaling in hypoxic pancreatic cancer cells Panc-1 TME [101]. In addition to its action against Hh signaling, curcumin was also demonstrated to reverse hypoxia-induced epithelial to mesenchymal transition (EMT). Resveratrol is also a potent inhibitor of Hh signaling, as revealed by its action in hypoxic TME of pancreatic cancer cells BxPC-3 and Panc-1, through a mechanism that involved suppression of ROS [102]. Similar to these effects of curcumin and resveratrol, our own investigations have revealed an effect of honokiol against Hh signaling in pancreatic cancer TME [103]. In addition to Hh signaling, we found a potent effect of honokiol against C-X-C chemokine receptor type 4 (CXCR4), another important factor that mediates tumor-stromal cross-talk. Further, EGCG can decrease hypoxia in non-small cell lung cancer (NSCLC) A549 cells through a rebalance of angiopoietins, resulting in sensitization of these cells to cisplatin [104].
In addition to induction of EMT, there is also evidence for enrichment of cancer stem cells (CSCs) in the TME, which can be effectively targeted by curcumin [105]. Not just in a pancreatic cancer model, CSCs are also enriched in the TME of colorectal cancer as well, where curcumin is able to interfere with the cross-talk between CSCs and stromal fibroblasts, resulting in reversal of EMT and the associated metastasis [106]. In the colorectal TME, the therapeutic potential of curcumin is underlined by its ability to potentiate 5-FU activity [106,107]. The same research group has reported a very similar action of resveratrol as well [108] which also involves potentiation of 5-FU activity in a 3D-alginate microenvironment, through reversal of EMT and inhibition of NF-κB signaling.
There is also evidence for immunomodulatory potential of phytochemicals within the TME. For example, curcumin, when delivered as polyethylene glycol (PEG) conjugate along with Trp2 peptide vaccine, resulted in reduced IL-6 levels and down-regulation of immunosuppressive factors (regulatory T cells, myeloid-derived suppressor cells) in a melanoma-bearing mouse model [109]. Resveratrol, however, in a renal cell carcinoma model, could only reduce regulatory T cells, but had no effect on myeloid-derived suppressor cells [110]. Further, COP9 signalosome 5 (CSN5) stabilized programmed cell death-ligand-1 (PD-L1) plays important role in TNF-α-induced cancer immunosuppression in TME, and curcumin can inhibit CSN5, leading to sensitization of cancer cells to immunotherapy [111].
Plumbagin seems to be a promising phytochemical to target bone metastasis of breast cancer [112] because of its ability to target tumor-bone microenvironment [113]. It specifically disrupts association of receptor activator of nuclear factor of κB (RANK) with TNF receptor-associated factor 6 (TRAF6), thus abrogating mitogen-activated protein kinases (MAPK) and NF-κB signaling [113]. IL-6 levels can be reduced by EGCG in breast cancer TME, as part of regulation of tumor-associated macrophages (TAMs) [114], and IL-18 can be inhibited by resveratrol in melanoma TME, leading to reduced metastasis [115]. Also, plumbagin has been demonstrated to inhibit the activity of c-MYB [116], an important modulator of tumor-stromal cross-talk and pancreatic cancer signaling [117]. In addition, EGCG has been shown to inhibit prostate cancer-associated myofibroblast differentiation [118]. Thus, there seems to be ample evidence in support of an action of phytochemicals against various components of the TME (Table 2).

4. Challenges for Phytochemicals in Cancer Therapy and Emerging Alternatives

Despite many promises and the demonstrated success in in vitro and pre-clinical studies, there has been little to no progress in the transition of phytochemicals to the clinic as the first line therapy. The limitations of in vitro testing models are well known. A direct exposure of cancer cell lines during in vitro testing causes an acute presentation of phytochemicals, inducing significant anticancer and antiproliferative action at concentrations usually not achieved under normal physiological conditions even upon consumption of the pure compound extract. While these in vitro studies provide significant insights into the cellular signaling mechanisms, they do not provide information on the effect of test agent on the organism as a whole. However, the obvious practical and ethical limitations of involving human studies without substantial laboratory evidence still makes it mandatory to rely on such in vitro models as the first step. Thus, there is a pressing need for in vitro and/or preclinical models that can mimic systemic exposure to phytochemicals, with resulting metabolomic and pharmacokinetic changes.
The first and foremost challenge is the problem of bioavailability [119,120,121]. Since a majority of these phytochemicals are part of normal human diet, they are efficiently metabolized and cleared by body. They do not persist in physiological systems and the therapeutic effects are usually short-lived [119]. The other aspect of using phytochemicals in cancer therapy is the lack of target specificity. While this is viewed by some as a limitation, it is increasingly being realized that such lack of specificity, and the multi-targeted effects of phytochemicals, the ‘pleiotropic’ effects, underline the very essence of these anticancer agents [51,122], particularly in view of the knowledge that when challenged with targeted therapies, tumor cells often activate alternate pathways which results in failure of targeted therapy. Under such circumstances, a multi-targeted therapeutic agent is likely to be a comparatively more effective because of its ability to check the activation of alternate survival pathways.
While the issue of bioavailability cannot be resolved just by increasing the administered dose or the frequency of administration, there are certain alternate ways which are being pursued to circumvent this problem. These include: (a) chemical syntheses of novel analogs of phytochemicals to increase the efficacy and bioavailability; (b) novel formulations to selectively and more effectively deliver the phytochemicals to their intended target organs; and (c) formulation of novel delivery systems that modulate the pharmacokinetics of the anticancer agent. In this section, we will highlight some recent advancements in these fields, particularly related to the phytochemicals discussed so far, to provide an overview of the broader research field.

4.1. Synthesis of Chemical Analogs

Curcumin is an exemplary phytochemical that has shown lots of potential in pre-clinical studies, only to fail in clinical settings. It also remains one of the most extensively modified phytochemicals, with so many reported analogs that it will be beyond the scope of this article to comment on every single curcumin analog that has been tested and reported for its enhanced anticancer activity. Just to put this into perspective, some curcumin analogs (Table 3) reported within the past two years include C-150 (inhibits NF-κB in glioblastoma cells [123]), Da0324 (inhibits NF-κB in gastric cancer cells [124]), 2,2’-fluoromonocarbonyl analog (modulates ROS in lung cancer cells [125]), A17 (induces ER stress in lung cancer cells [126]), MC37 (induces cell cycle arrest in colorectal cancer cells [127]), HO-3867 (STAT3 inhibitor in pancreatic cancer cells [128]), BDMC-A (inhibits NF-κB in breast cancer cells [129]), GO-Y078 (inhibits invasion of endothelial cells [130]), DM-1 (induces apoptosis in melanoma cells [131]), FLLL12 (induces apoptosis in lung cancer cells [132]), BHBA (activates Nrf2 in lung cancer models [133]) and L49H37 (induces apoptosis in pancreatic stellate cells [134]). Other than these, there are a few other curcumin analogs that have been investigated in comparatively more detail. These include difluorinated curcumin (CDF, inhibits MMP2 in lung cancer cells [135], restores PTEN in colorectal cancer cells [136] and inhibits cancer stem cells in pancreatic and colorectal cancer models [137,138]), WZ35 (modulates ROS in gastric cancer [139,140] and prostate cancer cells [141]) and EF24 (inhibits src phosphorylation in hepatocellular carcinoma cells [142], suppresses EMT through miR-33b in melanoma cells [143], induces ROS-dependent apoptosis in colorectal cancer cells [144], suppresses NF-κB in cholangiocarcinoma [145] and induces apoptosis in pancreatic cancer cells [146]).
Resveratrol has also generated considerable interest because of its anti-cancer potential, and consequently, a number of groups have synthesized analogs of resveratrol with an aim to enhance its efficacy [147,148,149,150,151,152,153] (Table 3). The DMU-212 analog of resveratrol has not just been reported to exhibit enhanced anti-tumor effects, but, interestingly, its mechanism of action has been reported to be distinct from the parent compound, resveratrol [154]. DMU-212 has also been suggested to suppress pro-inflammatory factors, especially NF-κB [155], a mechanism which happens to be similar to resveratrol. The interest in DMU-212 has led to synthesis of its own analogs that have been tested against a panel of 60 human cancer cell lines with mixed results [156]. HS-1793 is another synthesized analog of resveratrol with multiple reported properties ([157,158,159,160,161,162]).
In contrast to curcumin and resveratrol, only a handful of studies have reported synthesis of EGCG analogs. The d-ring analog of EGCG was one of the first to be made, and it was observed to target VEGF in breast cancer cells [163]. Such VEGF-targeting activity of another EGCG analog, a methylated one, has also been reported [164]. Other analogs of EGCG include the fluoro-substituted analogs that have shown promise in inhibiting proteasomal activity of leukemia [165] and breast cancer cells [166,167]. There is evidence for synthesis of novel plumbagin analog [168] but it has not yet been tested in pre-clinical cancer models. For the phytochemical honokiol, a dichloroacetate ester has been synthesized which inhibits androgen receptor signaling in prostate cancer cells [169] and re-sensitizes vemurafenib-resistant melanomas [170].

4.2. Novel Formulations

Nanotechnology has expanded the horizon of anti-cancer therapy in general [171], and the activity of several phytochemicals in particular [172]. Due to the overwhelming interest in curcumin in pre-clinical studies, a number of reports are available on the nanoformulations of curcumin as well as its analogs, all with the premises of recapitulating the pre-clinical success of curcumin in clinical settings [173,174,175]. Again, similar to the sub-section above, we will list here reports from only last two years, to keep the discussion meaningful.
Poly(d,l-lactic acid)-glycerol (PDLLA)-G-based curcumin nanoparticles have been reported to be as effective as free curcumin in in vitro assays, and their in vivo clearance was comparatively lower [176]. Such delayed clearance of curcumin nanoparticles can result in enhanced in vivo activity, as confirmed in a cervical cancer model [177]. Curcumin-loaded monomethoxy polyethylene glycol (mPEG)- poly(ε-caprolactone) (PCL) micelles showed increased plasma retention [178] while curcumin’s oligosaccharide of hyaluronan conjugate nanoparticle has been shown to be more stable and less toxic, relative to curcumin [179]. Similarly, curcumin-cyclodextrin/cellulose nanocrystal complexes kill colorectal and prostate cancer cells at IC50 values lower than curcumin [180]. Poly(lactic-co-glycolic acid) (PLGA)-curcumin particles have been demonstrated to deliver active curcumin directly to the cells’ cytosolic compartment, resulting in enhanced therapeutic activity [181]. This is similar to amphiphilic polyaspartamide polyelectrolytes that increase cellular uptake of curcumin [182]. Targeted delivery to prostate cancer tissue has been reported via lipid-polymer hybrid nanoparticles that encapsulate curcumin as well as docetaxel [183]. Conjugating curcumin on the hydrophilic terminals of pluronic F68 chains through cis-aconitic anhydride linkers is proposed to improve delivery of curcumin [184] while curcumin micelles have been shown to overcome multidrug resistance [185].
Not only the parent compound curcumin, there has been interest in nanoformulation of curcumin analogs as well. Nanomicelles of curcumin analog curcumin difluorinated (CDF) with amphiphilic styrene-maleic acid copolymer (SMA) increased the solubility by 5%–15% and exhibited enhanced antitumor effect [186] while liposome encapsulation of CDF sensitized the cisplatin-resistant head and neck cancer stem cells [187]. The hyaluronic acid-SMA-CDF micelles were reported to specifically target pancreatic cancer stem cells [188], similar to hyaluronic acid-PAMAM dendrimer formulation of CDF [189]. Another curcumin analog, EF24, has also been nanoencapsulated in pegylated liposomes resulting in enhanced anticancer effects against pancreatic cancer cells in vitro and in a pancreatic xenograft model [190]. Pegylated curcumin nanoparticles have also shown promise in real-time monitoring of drug release [191]. Another concept of embedding phytochemicals in the biopolymer PCL to continuously deliver the small molecule for extending periods of time has also been demonstrated for the delivery of curcuminoids [192]. The PCL-implant leads to the release of the agent in a two-step process. An initial burst releases the drug present on the surface into circulation. This initial burst-release can be controlled by coating an empty polymer; followed by the slow-steady release of the agent present in the matrix. Moreover, the site of implantation is hypothesized to increase drug accumulation at the site and a controlled release [192].
There are reports on nanoformulations of resveratrol as well. Resveratrol’s PCL and PLGA-PEG-COOH nanoparticles have been shown to improve release of resveratrol in hypoxia-relevant acidic TME, with efficient take up by the prostate cancer cells [193]. One of the objectives of nanoformulations is to reduce effective toxicity and this was reported in case of lipid nanoparticles of resveratrol when tested in lung cancer cells exposed to cigarette smoke condensate [194]. Another objective is to increase solubility and this has been demonstrated by nanocomplexation of resveratrol with soy protein isolate [195]. The gold nanoparticles of resveratrol have been shown to be biologically active with activity against multiple signaling cascades [196].
Efforts have also been made to formulate EGCG nanoparticles. In an early report on the topic [197], EGCG was conjugated with FITC, and the complex could enter the cytoplasm as well as the nucleus. In another report, gelatinized EGCG was observed to retain its biological activity against breast cancer cells, and was as potent as EGCG [198]. Similarly, PLA-PEG-EGCG nanoparticles also retained their activity, and even reported a 10-fold dose advantage in in vitro as well as in in vivo assays [199,200]. While free EGCG was completely degraded in 4 hours, the EGCG nanoparticles had a significantly longer half-life [201,202]. Co-encapsulation of EGCG with paclitaxel within a targeted PLGA-casein nanoparticle has been reported to re-sensitize paclitaxel-resistant breast cancer cells [203].
For the phytochemical plumbagin, silver nanoparticles were synthesized to overcome the lack of sensitivity and selectivity towards cancer cells, and such formulation reduced the effective concentration required to induce apoptosis to half of the free compound [204] in addition to enhancing the cellular uptake [205]. The reduced toxicity in normal cells, and better bioavailability of nanoformulated plumbagin has been reported by another independent group as well [206]. There has been some interest in nanoformulations of honokiol as well. The PCL-PEG-PCL-honokiol nanoparticles were shown to release honokiol over extended time [207]. The MPEG-PLA-encapsulation also had similar effect on honokiol release [208] and, moreover, this formulation also made honokiol injectable [209]. The stability of honokiol was remarkably improved by MPEG, to the extent that whereas it took 2 h to release 20% honokiol from the formulation in plasma, it took more than 200 h to release the same amount of honokiol in PBS [210].
Recently nanoformulations of bioactive lipids have been described. These lipids, such as the fatty acid docosahexaenoic acid, have been observed to demonstrate potent anticancer properties against a variety of cancer types such as breast cancer, ovarian cancer, glioblastomas, etc. [211,212,213,214]. The nanocapsules, comprising shells of bioactive lipids, are able to deliver drugs and other biologically sensitive molecules to specific cells or organs, thus enhancing the potency and cancer therapeutic potential [211,215]. This represents a promising area of novel nanoformulation where some progress has been made in recent years [216,217]. As a whole, the developments in the field of nanotechnology have raised the hopes of using phytochemicals as chemotherapeutic anticancer agents in near future.

5. Conclusions and Perspectives

The action of phytochemicals against cancer cells, via inhibition of proliferation, invasion, angiogenesis and metastasis, is well documented. A number of derived analogs have also been tested in different pre-clinical models. New investigations into phytochemicals’ mechanism of action have suggested that many of the observed pre-clinical effects of phytochemicals can possibly be explained by their ability to regulate TME (Figure 2). The importance of TME in cancer progression is well recognized. However, the studies involving regulation of TME and its various components can get quite challenging, with unavailability of assay systems that can accurately mimic the complexities of TME. This is specially challenging for studies with phytochemicals, where often there is a need for quick and efficient screening of several novel phytochemicals or their synthetic analogs. As a step in this direction, a 3-dimensional microfluidic device has been reported, with a capability of assaying multiple compounds simultaneously for their anti-metastatic potential [218]. The co-culture of endothelial and cancer cells in this device, and their 3-dimensional morphology, better represents TME. Such advancements in experimental capabilities provide hope for possible use of phytochemicals in clinics in near future.
Nanotechnology has been used to enhance chemoprevention outcome and the resulting ‘nanochemoprevention’ [219] seems to be relevant to enhancing the efficacy of phytochemicals [201]. However, nanoparticles themselves are often toxic, and often not suitable for oral consumption. This has led to novel ways to prepare nanoparticles, like for example chitosan-based nanoparticles [220]. Further, the individual chemicals used in nanoformulations are not alike, and differ in their ability to reinforce the intended biological effect. For example, silver –based nanoparticles are superior than zinc and titanium-based nanoparticles in protection against UV-induced DNA damage [221]. Thus, the promises of nanotechnology rest on fine tuning and careful characterization of the underlying materials and methods. Interestingly, the concept of ‘green’ nanotechnology, using biodegradable and eco-friendly materials, is gaining ground [222,223]. The combination of green nanotechnology and the natural phytochemicals sounds promising. By producing phytochemicals, nature has provided a healing touch to the very problems it frequently presents, including diseases such as cancer. Persistent and innovative methods to improve the anti-cancer efficacy of phytochemicals should not halt, until the pre-clinical success of these agents is realized in clinics.

Acknowledgments

This work was supported in part by NIH grants to A.P.S. (R01CA175772 and U01CA185490) and S.S. (R01CA204801 and R03CA186223). A.P.S. and S.S. also received funding and resource support from the University of South Alabama Mitchell Cancer Institute.

Author Contributions

H.Z., S.A., A.A. and A.P.S. conceived this work; H.Z., A.A., M.A.K. and G.K.P. helped in the writing and proof reading of article; S.S. and A.P.S. provided resources, infrastructure and oversight.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study or the writing of manuscript.

Abbreviations

The following abbreviations are used in this manuscript:
4-NQO4-nitroquinoline-1-oxide
5-FU5-FluoroUracil
CDFCurcumin Difluorinated
COX2CycloOxygenase2
CSCCancer Stem Cells
CSN5COP9 Signalosome 5
CXCR4C-X-C chemokine receptor type 4
DENDiEthylNitrosamine
DMBA7,12-Dimethylbenz[a]anthracene
EGCGEpiGalloCatechin Gallate
EMTEpithelial-Mesenchymal Transition
FITCFluorescein IsoThioCyanate
HhHedgehog
ILInterLeukin
iNOSinducible Nitric Oxide Synthase
Keap1Kelch-like erythroid cell-derived protein with Cap’n’collar homology-associated protein
MAPKMitogen-Activated Protein Kinases
MMP2Matrix MetalloProteinase2
MPEGMono methoxy Poly Ethylene Glycol
NF-κBNuclear Factor Kappa-light-chain-enhancer of activated B cells
Nrf2NF-E2 p45-related factor 2
NSCLCNon-Small Cell Lung Cancer
PCLpoly (ε-CaproLactone)
PD-L1Programmed cell Death-Ligand-1
PDLLA-GPoly(d,l-lactic acid)-glycerol
PEGPolyEthylene Glycol
PKCProtein Kinase C
PLAPoly Lactic Acif
PLGAPoly(Lactic-co-Glycolic Acid)
PTENPhosphatase and Tensin Homolog
RANKReceptor Activator of Nuclear Factor of κB
RNSReactive Nitrogen Species
ROSReactive Oxygen Species
SMAStyrene-Maleic Acid copolymer
TGF-βTransforming Growth Factor beta1
TLRToll-Like Receptors
TMETumor MicroEnvironment
TNF-αTumor Necrosis Factor-Alpha
TPA12-O-tetradecanoylphorbol-13-acetate
TRAF6TNF Receptor Associated Factor 6
UVUltraViolet
VEGFVascular Endothelial Growth Factor

References

  1. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2017. CA Cancer J. Clin. 2017, 67, 7–30. [Google Scholar] [CrossRef] [PubMed]
  2. Okimoto, R.A.; Bivona, T.G. Recent advances in personalized lung cancer medicine. Per Med 2014, 11, 309–321. [Google Scholar] [CrossRef] [PubMed]
  3. Krepler, C.; Xiao, M.; Sproesser, K.; Brafford, P.A.; Shannan, B.; Beqiri, M.; Liu, Q.; Xu, W.; Garman, B.; Nathanson, K.L.; et al. Personalized preclinical trials in braf inhibitor-resistant patient-derived xenograft models identify second-line combination therapies. Clin. Cancer Res. 2016, 22, 1592–1602. [Google Scholar] [CrossRef] [PubMed]
  4. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661. [Google Scholar] [CrossRef] [PubMed]
  5. Janakiram, N.B.; Mohammed, A.; Madka, V.; Kumar, G.; Rao, C.V. Prevention and treatment of cancers by immune modulating nutrients. Mol. Nutr. Food Res. 2016, 60, 1275–1294. [Google Scholar] [CrossRef] [PubMed]
  6. Chih, H.J.; Lee, A.H.; Colville, L.; Binns, C.W.; Xu, D. A review of dietary prevention of human papillomavirus-related infection of the cervix and cervical intraepithelial neoplasia. Nutr. Cancer 2013, 65, 317–328. [Google Scholar] [CrossRef] [PubMed]
  7. Terry, P.; Wolk, A. Tea consumption and the risk of colorectal cancer in sweden. Nutr. Cancer 2001, 39, 176–179. [Google Scholar] [CrossRef] [PubMed]
  8. Van Duyn, M.A.; Pivonka, E. Overview of the health benefits of fruit and vegetable consumption for the dietetics professional: Selected literature. J. Am. Diet. Assoc. 2000, 100, 1511–1521. [Google Scholar] [CrossRef]
  9. Tseng, M. Diet, cancer and public health nutrition. Public Health Nutr. 2009, 12, 737–738. [Google Scholar] [CrossRef]
  10. Blokhina, O.; Fagerstedt, K.V. Oxidative metabolism, ros and no under oxygen deprivation. Plant Physiol. Biochem. 2010, 48, 359–373. [Google Scholar] [CrossRef] [PubMed]
  11. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative stress, inflammation, and cancer: How are they linked? Free Radic. Biol. Med. 2010, 49, 1603–1616. [Google Scholar] [CrossRef] [PubMed]
  12. Huang, M.T.; Smart, R.C.; Wong, C.Q.; Conney, A.H. Inhibitory effect of curcumin, chlorogenic acid, caffeic acid, and ferulic acid on tumor promotion in mouse skin by 12-o-tetradecanoylphorbol-13-acetate. Cancer Res. 1988, 48, 5941–5946. [Google Scholar] [PubMed]
  13. Cho, J.; Rho, O.; Junco, J.; Carbajal, S.; Siegel, D.; Slaga, T.J.; DiGiovanni, J. Effect of combined treatment with ursolic acid and resveratrol on skin tumor promotion by 12-o-tetradecanoylphorbol-13-acetate. Cancer Prev. Res. (Phila) 2015, 8, 817–825. [Google Scholar] [CrossRef] [PubMed]
  14. Blumberg, P.M. Protein kinase c as the receptor for the phorbol ester tumor promoters: Sixth rhoads memorial award lecture. Cancer Res. 1988, 48, 1–8. [Google Scholar] [PubMed]
  15. Shieh, J.M.; Chiang, T.A.; Chang, W.T.; Chao, C.H.; Lee, Y.C.; Huang, G.Y.; Shih, Y.X.; Shih, Y.W. Plumbagin inhibits tpa-induced mmp-2 and u-pa expressions by reducing binding activities of nf-kappab and ap-1 via erk signaling pathway in a549 human lung cancer cells. Mol. Cell Biochem. 2010, 335, 181–193. [Google Scholar] [CrossRef] [PubMed]
  16. Konoshima, T.; Kozuka, M.; Tokuda, H.; Nishino, H.; Iwashima, A.; Haruna, M.; Ito, K.; Tanabe, M. Studies on inhibitors of skin tumor promotion, ix. Neolignans from magnolia officinalis. J. Nat. Prod. 1991, 54, 816–822. [Google Scholar] [CrossRef] [PubMed]
  17. Yanaida, Y.; Kohno, H.; Yoshida, K.; Hirose, Y.; Yamada, Y.; Mori, H.; Tanaka, T. Dietary silymarin suppresses 4-nitroquinoline 1-oxide-induced tongue carcinogenesis in male f344 rats. Carcinogenesis 2002, 23, 787–794. [Google Scholar] [CrossRef] [PubMed]
  18. Nunoshiba, T.; Demple, B. Potent intracellular oxidative stress exerted by the carcinogen 4-nitroquinoline-n-oxide. Cancer Res. 1993, 53, 3250–3252. [Google Scholar] [PubMed]
  19. Chuang, S.E.; Kuo, M.L.; Hsu, C.H.; Chen, C.R.; Lin, J.K.; Lai, G.M.; Hsieh, C.Y.; Cheng, A.L. Curcumin-containing diet inhibits diethylnitrosamine-induced murine hepatocarcinogenesis. Carcinogenesis 2000, 21, 331–335. [Google Scholar] [CrossRef] [PubMed]
  20. Ajiboye, T.O.; Komolafe, Y.O.; Bukoye Oloyede, H.O.; Yakubu, M.T.; Adeoye, M.D.; Abdulsalami, I.O.; Oladiji, A.T.; Akanji, M.A. Diethylnitrosamine-induced redox imbalance in rat microsomes: Protective role of polyphenolic-rich extract from sorghum bicolor grains. J. Basic Clin. Physiol. Pharmacol. 2013, 24, 41–49. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, Z.Y.; Huang, M.T.; Lou, Y.R.; Xie, J.G.; Reuhl, K.R.; Newmark, H.L.; Ho, C.T.; Yang, C.S.; Conney, A.H. Inhibitory effects of black tea, green tea, decaffeinated black tea, and decaffeinated green tea on ultraviolet b light-induced skin carcinogenesis in 7,12-dimethylbenz[a]anthracene-initiated skh-1 mice. Cancer Res. 1994, 54, 3428–3435. [Google Scholar]
  22. Wang, Z.Y.; Hong, J.Y.; Huang, M.T.; Reuhl, K.R.; Conney, A.H.; Yang, C.S. Inhibition of n-nitrosodiethylamine- and 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone-induced tumorigenesis in a/j mice by green tea and black tea. Cancer Res. 1992, 52, 1943–1947. [Google Scholar] [PubMed]
  23. Finch, C.E.; Crimmins, E.M. Inflammatory exposure and historical changes in human life-spans. Science 2004, 305, 1736–1739. [Google Scholar] [CrossRef] [PubMed]
  24. Ahmad, A.; Banerjee, S.; Wang, Z.; Kong, D.; Majumdar, A.P.; Sarkar, F.H. Aging and inflammation: Etiological culprits of cancer. Curr. Aging Sci. 2009, 2, 174–186. [Google Scholar] [CrossRef] [PubMed]
  25. He, Y.; Yue, Y.; Zheng, X.; Zhang, K.; Chen, S.; Du, Z. Curcumin, inflammation, and chronic diseases: How are they linked? Molecules 2015, 20, 9183–9213. [Google Scholar] [CrossRef] [PubMed]
  26. Ercolini, A.M.; Miller, S.D. The role of infections in autoimmune disease. Clin. Exp. Immunol. 2009, 155, 1–15. [Google Scholar] [CrossRef] [PubMed]
  27. Anand, P.; Kunnumakkara, A.B.; Sundaram, C.; Harikumar, K.B.; Tharakan, S.T.; Lai, O.S.; Sung, B.; Aggarwal, B.B. Cancer is a preventable disease that requires major lifestyle changes. Pharm. Res. 2008, 25, 2097–2116. [Google Scholar] [CrossRef] [PubMed]
  28. Tili, E.; Michaille, J.J. Promiscuous effects of some phenolic natural products on inflammation at least in part arise from their ability to modulate the expression of global regulators, namely micrornas. Molecules 2016, 21, 1263. [Google Scholar] [CrossRef] [PubMed]
  29. Schaffer, M.; Schaffer, P.M.; Bar-Sela, G. An update on curcuma as a functional food in the control of cancer and inflammation. Curr. Opin. Clin. Nutr. Metab. Care 2015, 18, 605–611. [Google Scholar] [CrossRef] [PubMed]
  30. Das, L.; Vinayak, M. Long term effect of curcumin in restoration of tumour suppressor p53 and phase-II antioxidant enzymes via activation of Nrf2 signalling and modulation of inflammation in prevention of cancer. PLoS ONE 2015, 10, e0124000. [Google Scholar] [CrossRef] [PubMed]
  31. Rana, M.; Maurya, P.; Reddy, S.S.; Singh, V.; Ahmad, H.; Dwivedi, A.K.; Dikshit, M.; Barthwal, M.K. A standardized chemically modified curcuma longa extract modulates irak-mapk signaling in inflammation and potentiates cytotoxicity. Front Pharmacol. 2016, 7, 223. [Google Scholar] [CrossRef] [PubMed]
  32. Ahmad, A.; Farhan Asad, S.; Singh, S.; Hadi, S.M. DNA breakage by resveratrol and cu(II): Reaction mechanism and bacteriophage inactivation. Cancer Lett. 2000, 154, 29–37. [Google Scholar] [CrossRef]
  33. Ahmad, A.; Syed, F.A.; Singh, S.; Hadi, S.M. Prooxidant activity of resveratrol in the presence of copper ions: Mutagenicity in plasmid DNA. Toxicol. Lett. 2005, 159, 1–12. [Google Scholar] [CrossRef] [PubMed]
  34. Inoue, H.; Nakata, R. Resveratrol targets in inflammation. Endocr. Metab. Immune Disord. Drug Targets 2015, 15, 186–195. [Google Scholar] [CrossRef] [PubMed]
  35. Latruffe, N.; Lancon, A.; Frazzi, R.; Aires, V.; Delmas, D.; Michaille, J.J.; Djouadi, F.; Bastin, J.; Cherkaoui-Malki, M. Exploring new ways of regulation by resveratrol involving mirnas, with emphasis on inflammation. Ann. N. Y. Acad. Sci. 2015, 1348, 97–106. [Google Scholar] [CrossRef] [PubMed]
  36. Limagne, E.; Lancon, A.; Delmas, D.; Cherkaoui-Malki, M.; Latruffe, N. Resveratrol interferes with il1-beta-induced pro-inflammatory paracrine interaction between primary chondrocytes and macrophages. Nutrients 2016, 8, 280. [Google Scholar] [CrossRef] [PubMed]
  37. Shanmugam, T.; Selvaraj, M.; Poomalai, S. Epigallocatechin gallate potentially abrogates fluoride induced lung oxidative stress, inflammation via Nrf2/keap1 signaling pathway in rats: An in vivo and in-silico study. Int. Immunopharmacol. 2016, 39, 128–139. [Google Scholar] [CrossRef] [PubMed]
  38. Kweon, M.H.; Adhami, V.M.; Lee, J.S.; Mukhtar, H. Constitutive overexpression of nrf2-dependent heme oxygenase-1 in a549 cells contributes to resistance to apoptosis induced by epigallocatechin 3-gallate. J. Biol. Chem. 2006, 281, 33761–33772. [Google Scholar] [CrossRef] [PubMed]
  39. Gao, D.Q.; Qian, S.; Ju, T. Anticancer activity of honokiol against lymphoid malignant cells via activation of ros-jnk and attenuation of nrf2 and nf-kappab. J. BUON 2016, 21, 673–679. [Google Scholar]
  40. Pan, S.T.; Qin, Y.; Zhou, Z.W.; He, Z.X.; Zhang, X.; Yang, T.; Yang, Y.X.; Wang, D.; Zhou, S.F.; Qiu, J.X. Plumbagin suppresses epithelial to mesenchymal transition and stemness via inhibiting Nrf2-mediated signaling pathway in human tongue squamous cell carcinoma cells. Drug Des. Dev. Ther. 2015, 9, 5511–5551. [Google Scholar]
  41. Singh, B.; Shoulson, R.; Chatterjee, A.; Ronghe, A.; Bhat, N.K.; Dim, D.C.; Bhat, H.K. Resveratrol inhibits estrogen-induced breast carcinogenesis through induction of nrf2-mediated protective pathways. Carcinogenesis 2014, 35, 1872–1880. [Google Scholar] [CrossRef] [PubMed]
  42. Kim, J.H.; Gupta, S.C.; Park, B.; Yadav, V.R.; Aggarwal, B.B. Turmeric (Curcuma longa) inhibits inflammatory nuclear factor (NF)-kappab and nf-kappab-regulated gene products and induces death receptors leading to suppressed proliferation, induced chemosensitization, and suppressed osteoclastogenesis. Mol. Nutr. Food Res. 2012, 56, 454–465. [Google Scholar] [CrossRef] [PubMed]
  43. Pan, H.; Chen, J.; Shen, K.; Wang, X.; Wang, P.; Fu, G.; Meng, H.; Wang, Y.; Jin, B. Mitochondrial modulation by epigallocatechin 3-gallate ameliorates cisplatin induced renal injury through decreasing oxidative/nitrative stress, inflammation and NF-kB in mice. PLoS ONE 2015, 10, e0124775. [Google Scholar] [CrossRef] [PubMed]
  44. Syed, D.N.; Afaq, F.; Kweon, M.H.; Hadi, N.; Bhatia, N.; Spiegelman, V.S.; Mukhtar, H. Green tea polyphenol egcg suppresses cigarette smoke condensate-induced nf-kappab activation in normal human bronchial epithelial cells. Oncogene 2007, 26, 673–682. [Google Scholar] [CrossRef] [PubMed]
  45. Arora, S.; Singh, S.; Piazza, G.A.; Contreras, C.M.; Panyam, J.; Singh, A.P. Honokiol: A novel natural agent for cancer prevention and therapy. Curr. Mol. Med. 2012, 12, 1244–1252. [Google Scholar] [CrossRef] [PubMed]
  46. Singh, T.; Katiyar, S.K. Honokiol inhibits non-small cell lung cancer cell migration by targeting pge(2)-mediated activation of beta-catenin signaling. PLoS ONE 2013, 8, e60749. [Google Scholar]
  47. Ahmad, A.; Banerjee, S.; Wang, Z.; Kong, D.; Sarkar, F.H. Plumbagin-induced apoptosis of human breast cancer cells is mediated by inactivation of nf-kappab and bcl-2. J. Cell Biochem. 2008, 105, 1461–1471. [Google Scholar] [CrossRef] [PubMed]
  48. Adhami, V.M.; Afaq, F.; Ahmad, N. Suppression of ultraviolet b exposure-mediated activation of nf-kappab in normal human keratinocytes by resveratrol. Neoplasia 2003, 5, 74–82. [Google Scholar] [CrossRef]
  49. Ben-Neriah, Y.; Karin, M. Inflammation meets cancer, with nf-kappab as the matchmaker. Nat. Immunol. 2011, 12, 715–723. [Google Scholar] [CrossRef] [PubMed]
  50. Sung, B.; Prasad, S.; Yadav, V.R.; Aggarwal, B.B. Cancer cell signaling pathways targeted by spice-derived nutraceuticals. Nutr. Cancer 2012, 64, 173–197. [Google Scholar] [CrossRef] [PubMed]
  51. Ahmad, A.; Ginnebaugh, K.R.; Li, Y.; Padhye, S.B.; Sarkar, F.H. Molecular targets of naturopathy in cancer research: Bridge to modern medicine. Nutrients 2015, 7, 321–334. [Google Scholar] [CrossRef] [PubMed]
  52. Ahmad, A.; Biersack, B.; Li, Y.; Kong, D.; Bao, B.; Schobert, R.; Padhye, S.B.; Sarkar, F.H. Targeted regulation of pi3k/akt/mtor/nf-kappab signaling by indole compounds and their derivatives: Mechanistic details and biological implications for cancer therapy. Anticancer Agents Med. Chem. 2013, 13, 1002–1013. [Google Scholar] [CrossRef] [PubMed]
  53. Murtaza, I.; Adhami, V.M.; Hafeez, B.B.; Saleem, M.; Mukhtar, H. Fisetin, a natural flavonoid, targets chemoresistant human pancreatic cancer aspc-1 cells through dr3-mediated inhibition of NF-kappaB. Int. J. Cancer 2009, 125, 2465–2473. [Google Scholar] [CrossRef] [PubMed]
  54. Ahmed, S.M.; Luo, L.; Namani, A.; Wang, X.J.; Tang, X. Nrf2 signaling pathway: Pivotal roles in inflammation. Biochim. Biophys. Acta 2017, 1863, 585–597. [Google Scholar] [CrossRef] [PubMed]
  55. Qin, S.; Hou, D.X. Multiple regulations of keap1/Nrf2 system by dietary phytochemicals. Mol. Nutr. Food Res. 2016, 60, 1731–1755. [Google Scholar] [CrossRef] [PubMed]
  56. Hadi, S.M.; Ullah, M.F.; Azmi, A.S.; Ahmad, A.; Shamim, U.; Zubair, H.; Khan, H.Y. Resveratrol mobilizes endogenous copper in human peripheral lymphocytes leading to oxidative DNA breakage: A putative mechanism for chemoprevention of cancer. Pharm. Res. 2010, 27, 979–988. [Google Scholar] [CrossRef] [PubMed]
  57. Khan, H.Y.; Zubair, H.; Ullah, M.F.; Ahmad, A.; Hadi, S.M. Oral administration of copper to rats leads to increased lymphocyte cellular DNA degradation by dietary polyphenols: Implications for a cancer preventive mechanism. Biometals 2011, 24, 1169–1178. [Google Scholar] [CrossRef] [PubMed]
  58. Zubair, H.; Azim, S.; Khan, H.Y.; Ullah, M.F.; Wu, D.; Singh, A.P.; Hadi, S.M.; Ahmad, A. Mobilization of intracellular copper by gossypol and apogossypolone leads to reactive oxygen species-mediated cell death: Putative anticancer mechanism. Int. J. Mol. Sci. 2016, 17, 973. [Google Scholar] [CrossRef] [PubMed]
  59. Farhan, M.; Khan, H.Y.; Oves, M.; Al-Harrasi, A.; Rehmani, N.; Arif, H.; Hadi, S.M.; Ahmad, A. Cancer therapy by catechins involves redox cycling of copper ions and generation of reactive oxygen species. Toxins (Basel) 2016, 8, 37. [Google Scholar] [CrossRef] [PubMed]
  60. Gupte, A.; Mumper, R.J. Elevated copper and oxidative stress in cancer cells as a target for cancer treatment. Cancer Treat. Rev. 2009, 35, 32–46. [Google Scholar] [CrossRef] [PubMed]
  61. Yoshino, M.; Haneda, M.; Naruse, M.; Htay, H.H.; Tsubouchi, R.; Qiao, S.L.; Li, W.H.; Murakami, K.; Yokochi, T. Prooxidant activity of curcumin: Copper-dependent formation of 8-hydroxy-2′-deoxyguanosine in DNA and induction of apoptotic cell death. Toxicol. In Vitro 2004, 18, 783–789. [Google Scholar] [CrossRef] [PubMed]
  62. Khan, M.A.; Gahlot, S.; Majumdar, S. Oxidative stress induced by curcumin promotes the death of cutaneous t-cell lymphoma (hut-78) by disrupting the function of several molecular targets. Mol. Cancer Ther. 2012, 11, 1873–1883. [Google Scholar] [CrossRef] [PubMed]
  63. Nazeem, S.; Azmi, A.S.; Hanif, S.; Ahmad, A.; Mohammad, R.M.; Hadi, S.M.; Kumar, K.S. Plumbagin induces cell death through a copper-redox cycle mechanism in human cancer cells. Mutagenesis 2009, 24, 413–418. [Google Scholar] [CrossRef] [PubMed]
  64. Foundation, T.B.N. Iron as a pro-oxidant. In Iron: Nutritional and Physiological Significance the Report of the British Nutrition Foundation’s Task Force; Springer: Dordrecht, The Netherlands, 1995; pp. 33–41. [Google Scholar]
  65. Ninsontia, C.; Phiboonchaiyanan, P.P.; Chanvorachote, P. Zinc induces epithelial to mesenchymal transition in human lung cancer h460 cells via superoxide anion-dependent mechanism. Cancer Cell Int. 2016, 16, 48. [Google Scholar] [CrossRef] [PubMed]
  66. Lim, J.Y.; Kim, D.; Kim, B.R.; Jun, J.S.; Yeom, J.S.; Park, J.S.; Seo, J.H.; Park, C.H.; Woo, H.O.; Youn, H.S.; et al. Vitamin C induces apoptosis in ags cells via production of ros of mitochondria. Oncol. Lett. 2016, 12, 4270–4276. [Google Scholar] [CrossRef] [PubMed]
  67. Amatore, C.; Arbault, S.; Ferreira, D.C.M.; Tapsoba, I.; Verchier, Y. Vitamin c stimulates or attenuates reactive oxygen and nitrogen species (ros, rns) production depending on cell state: Quantitative amperometric measurements of oxidative bursts at plb-985 and raw 264.7 cells at the single cell level. J. Electroanal. Chem. 2008, 615, 34–44. [Google Scholar] [CrossRef]
  68. Du, J.; Cullen, J.J.; Buettner, G.R. Ascorbic acid: Chemistry, biology and the treatment of cancer. Biochim. Biophys. Acta 2012, 1826, 443–457. [Google Scholar] [CrossRef] [PubMed]
  69. Bergstrom, T.; Ersson, C.; Bergman, J.; Moller, L. Vitamins at physiological levels cause oxidation to the DNA nucleoside deoxyguanosine and to DNA—Alone or in synergism with metals. Mutagenesis 2012, 27, 511–517. [Google Scholar] [CrossRef] [PubMed]
  70. Weinberg, S.E.; Chandel, N.S. Targeting mitochondria metabolism for cancer therapy. Nat. Chem. Biol. 2015, 11, 9–15. [Google Scholar] [CrossRef] [PubMed]
  71. Weber, G.F. Time and circumstances: Cancer cell metabolism at various stages of disease progression. Front. Oncol. 2016, 6, 257. [Google Scholar] [CrossRef] [PubMed]
  72. DeBerardinis, R.J. Is cancer a disease of abnormal cellular metabolism? New angles on an old idea. Genet. Med. 2008, 10, 767–777. [Google Scholar] [CrossRef] [PubMed]
  73. Hanhineva, K.; Torronen, R.; Bondia-Pons, I.; Pekkinen, J.; Kolehmainen, M.; Mykkanen, H.; Poutanen, K. Impact of dietary polyphenols on carbohydrate metabolism. Int. J. Mol. Sci. 2010, 11, 1365–1402. [Google Scholar] [CrossRef] [PubMed]
  74. Qiu, P.; Sun, J.; Man, S.; Yang, H.; Ma, L.; Yu, P.; Gao, W. Curcumin attenuates n-nitrosodiethylamine-induced liver injury in mice by utilizing the method of metabonomics. J. Agric. Food Chem. 2017. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, W.; Liu, Y.; Chen, X.; Bergmeier, S.C. Novel inhibitors of basal glucose transport as potential anticancer agents. Bioorg. Med. Chem. Lett. 2010, 20, 2191–2194. [Google Scholar] [CrossRef] [PubMed]
  76. Bayet-Robert, M.; Morvan, D. Metabolomics reveals metabolic targets and biphasic responses in breast cancer cells treated by curcumin alone and in association with docetaxel. PLoS ONE 2013, 8, e57971. [Google Scholar] [CrossRef] [PubMed]
  77. Gunnink, L.K.; Alabi, O.D.; Kuiper, B.D.; Gunnink, S.M.; Schuiteman, S.J.; Strohbehn, L.E.; Hamilton, K.E.; Wrobel, K.E.; Louters, L.L. Curcumin directly inhibits the transport activity of glut1. Biochimie 2016, 125, 179–185. [Google Scholar] [CrossRef] [PubMed]
  78. Liao, H.; Wang, Z.; Deng, Z.; Ren, H.; Li, X. Curcumin inhibits lung cancer invasion and metastasis by attenuating glut1/mt1-mmp/mmp2 pathway. Int. J. Clin. Exp. Med. 2015, 8, 8948–8957. [Google Scholar] [PubMed]
  79. Gwak, H.; Haegeman, G.; Tsang, B.K.; Song, Y.S. Cancer-specific interruption of glucose metabolism by resveratrol is mediated through inhibition of akt/glut1 axis in ovarian cancer cells. Mol. Carcinog. 2015, 54, 1529–1540. [Google Scholar] [CrossRef] [PubMed]
  80. Sinha, S.; Pal, K.; Elkhanany, A.; Dutta, S.; Cao, Y.; Mondal, G.; Iyer, S.; Somasundaram, V.; Couch, F.J.; Shridhar, V.; et al. Plumbagin inhibits tumorigenesis and angiogenesis of ovarian cancer cells in vivo. Int. J. Cancer 2013, 132, 1201–1212. [Google Scholar] [CrossRef]
  81. Venturelli, L.; Nappini, S.; Bulfoni, M.; Gianfranceschi, G.; Dal Zilio, S.; Coceano, G.; Del Ben, F.; Turetta, M.; Scoles, G.; Vaccari, L.; et al. Glucose is a key driver for glut1-mediated nanoparticles internalization in breast cancer cells. Sci. Rep. 2016, 6, 21629. [Google Scholar] [CrossRef] [PubMed]
  82. Jung, K.H.; Lee, J.H.; Park, J.W.; Moon, S.H.; Cho, Y.S.; Choe, Y.S.; Lee, K.H. Effects of curcumin on cancer cell mitochondrial function and potential monitoring with (1)(8)f-fdg uptake. Oncol. Rep. 2016, 35, 861–868. [Google Scholar] [CrossRef]
  83. Malhotra, A.; Nair, P.; Dhawan, D.K. Study to evaluate molecular mechanics behind synergistic chemo-preventive effects of curcumin and resveratrol during lung carcinogenesis. PLoS ONE 2014, 9, e93820. [Google Scholar] [CrossRef] [PubMed]
  84. Li, W.; Ma, X.; Li, N.; Liu, H.; Dong, Q.; Zhang, J.; Yang, C.; Liu, Y.; Liang, Q.; Zhang, S.; et al. Resveratrol inhibits hexokinases ii mediated glycolysis in non-small cell lung cancer via targeting akt signaling pathway. Exp. Cell Res. 2016, 349, 320–327. [Google Scholar] [CrossRef] [PubMed]
  85. Moreira, L.; Araujo, I.; Costa, T.; Correia-Branco, A.; Faria, A.; Martel, F.; Keating, E. Quercetin and epigallocatechin gallate inhibit glucose uptake and metabolism by breast cancer cells by an estrogen receptor-independent mechanism. Exp. Cell Res. 2013, 319, 1784–1795. [Google Scholar] [CrossRef] [PubMed]
  86. Wu, A.H.; Spicer, D.; Stanczyk, F.Z.; Tseng, C.C.; Yang, C.S.; Pike, M.C. Effect of 2-month controlled green tea intervention on lipoprotein cholesterol, glucose, and hormone levels in healthy postmenopausal women. Cancer Prev. Res. (Phila) 2012, 5, 393–402. [Google Scholar] [CrossRef] [PubMed]
  87. Parimala, R.; Sachdanandam, P. Effect of plumbagin on some glucose metabolising enzymes studied in rats in experimental hepatoma. Mol. Cell Biochem. 1993, 125, 59–63. [Google Scholar] [CrossRef] [PubMed]
  88. Wu, J.P.; Zhang, W.; Wu, F.; Zhao, Y.; Cheng, L.F.; Xie, J.J.; Yao, H.P. Honokiol: An effective inhibitor of high-glucose-induced upregulation of inflammatory cytokine production in human renal mesangial cells. Inflamm. Res. 2010, 59, 1073–1079. [Google Scholar] [CrossRef] [PubMed]
  89. Freeman, M.R.; Kim, J.; Lisanti, M.P.; Di Vizio, D. A metabolic perturbation by u0126 identifies a role for glutamine in resveratrol-induced cell death. Cancer Biol. Ther. 2011, 12, 966–977. [Google Scholar] [CrossRef] [PubMed]
  90. Huang, Y.T.; Lin, Y.W.; Chiu, H.M.; Chiang, B.H. Curcumin induces apoptosis of colorectal cancer stem cells by coupling with cd44 marker. J. Agric. Food Chem. 2016, 64, 2247–2253. [Google Scholar] [CrossRef] [PubMed]
  91. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed]
  92. Vaughan, R.A.; Garcia-Smith, R.; Dorsey, J.; Griffith, J.K.; Bisoffi, M.; Trujillo, K.A. Tumor necrosis factor alpha induces warburg-like metabolism and is reversed by anti-inflammatory curcumin in breast epithelial cells. Int. J. Cancer 2013, 133, 2504–2510. [Google Scholar] [CrossRef] [PubMed]
  93. Blanquer-Rossello, M.D.; Hernandez-Lopez, R.; Roca, P.; Oliver, J.; Valle, A. Resveratrol induces mitochondrial respiration and apoptosis in sw620 colon cancer cells. Biochim. Biophys. Acta 2017, 1861, 431–440. [Google Scholar] [CrossRef] [PubMed]
  94. Potter, M.; Newport, E.; Morten, K.J. The warburg effect: 80 years on. Biochem. Soc. Trans. 2016, 44, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, N.; Kim, D. Cancer metabolism: Fueling more than just growth. Mol. Cells 2016, 39, 847–854. [Google Scholar] [PubMed]
  96. Chang, C.H.; Qiu, J.; O'Sullivan, D.; Buck, M.D.; Noguchi, T.; Curtis, J.D.; Chen, Q.; Gindin, M.; Gubin, M.M.; van der Windt, G.J.; et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell 2015, 162, 1229–1241. [Google Scholar] [CrossRef] [PubMed]
  97. Pistollato, F.; Giampieri, F.; Battino, M. The use of plant-derived bioactive compounds to target cancer stem cells and modulate tumor microenvironment. Food Chem. Toxicol. 2015, 75, 58–70. [Google Scholar] [CrossRef] [PubMed]
  98. Sharma, S.H.; Thulasingam, S.; Nagarajan, S. Chemopreventive agents targeting tumor microenvironment. Life Sci. 2016, 145, 74–84. [Google Scholar] [CrossRef] [PubMed]
  99. Onishi, H.; Kai, M.; Odate, S.; Iwasaki, H.; Morifuji, Y.; Ogino, T.; Morisaki, T.; Nakashima, Y.; Katano, M. Hypoxia activates the hedgehog signaling pathway in a ligand-independent manner by upregulation of smo transcription in pancreatic cancer. Cancer Sci. 2011, 102, 1144–1150. [Google Scholar] [CrossRef] [PubMed]
  100. Spivak-Kroizman, T.R.; Hostetter, G.; Posner, R.; Aziz, M.; Hu, C.; Demeure, M.J.; Von Hoff, D.; Hingorani, S.R.; Palculict, T.B.; Izzo, J.; et al. Hypoxia triggers hedgehog-mediated tumor-stromal interactions in pancreatic cancer. Cancer Res. 2013, 73, 3235–3247. [Google Scholar] [CrossRef] [PubMed]
  101. Cao, L.; Xiao, X.; Lei, J.; Duan, W.; Ma, Q.; Li, W. Curcumin inhibits hypoxia-induced epithelialmesenchymal transition in pancreatic cancer cells via suppression of the hedgehog signaling pathway. Oncol. Rep. 2016, 35, 3728–3734. [Google Scholar] [PubMed]
  102. Li, W.; Cao, L.; Chen, X.; Lei, J.; Ma, Q. Resveratrol inhibits hypoxia-driven ros-induced invasive and migratory ability of pancreatic cancer cells via suppression of the hedgehog signaling pathway. Oncol. Rep. 2016, 35, 1718–1726. [Google Scholar] [CrossRef] [PubMed]
  103. Averett, C.; Bhardwaj, A.; Arora, S.; Srivastava, S.K.; Khan, M.A.; Ahmad, A.; Singh, S.; Carter, J.E.; Khushman, M.; Singh, A.P. Honokiol suppresses pancreatic tumor growth, metastasis and desmoplasia by interfering with tumor-stromal cross-talk. Carcinogenesis 2016, 37, 1052–1061. [Google Scholar] [CrossRef] [PubMed]
  104. Deng, P.B.; Hu, C.P.; Xiong, Z.; Yang, H.P.; Li, Y.Y. Treatment with egcg in nsclc leads to decreasing interstitial fluid pressure and hypoxia to improve chemotherapy efficacy through rebalance of ang-1 and ang-2. Chin. J. Nat. Med. 2013, 11, 245–253. [Google Scholar] [CrossRef] [PubMed]
  105. Bao, B.; Ahmad, A.; Li, Y.; Azmi, A.S.; Ali, S.; Banerjee, S.; Kong, D.; Sarkar, F.H. Targeting cscs within the tumor microenvironment for cancer therapy: A potential role of mesenchymal stem cells. Exp. Opin. Ther. Targets 2012, 16, 1041–1054. [Google Scholar] [CrossRef] [PubMed]
  106. Buhrmann, C.; Kraehe, P.; Lueders, C.; Shayan, P.; Goel, A.; Shakibaei, M. Curcumin suppresses crosstalk between colon cancer stem cells and stromal fibroblasts in the tumor microenvironment: Potential role of emt. PLoS ONE 2014, 9, e107514. [Google Scholar] [CrossRef] [PubMed]
  107. Shakibaei, M.; Kraehe, P.; Popper, B.; Shayan, P.; Goel, A.; Buhrmann, C. Curcumin potentiates antitumor activity of 5-fluorouracil in a 3d alginate tumor microenvironment of colorectal cancer. BMC Cancer 2015, 15, 250. [Google Scholar] [CrossRef] [PubMed]
  108. Buhrmann, C.; Shayan, P.; Kraehe, P.; Popper, B.; Goel, A.; Shakibaei, M. Resveratrol induces chemosensitization to 5-fluorouracil through up-regulation of intercellular junctions, epithelial-to-mesenchymal transition and apoptosis in colorectal cancer. Biochem. Pharmacol. 2015, 98, 51–68. [Google Scholar] [CrossRef] [PubMed]
  109. Lu, Y.; Miao, L.; Wang, Y.; Xu, Z.; Zhao, Y.; Shen, Y.; Xiang, G.; Huang, L. Curcumin micelles remodel tumor microenvironment and enhance vaccine activity in an advanced melanoma model. Mol. Ther. 2016, 24, 364–374. [Google Scholar] [CrossRef] [PubMed]
  110. Chen, L.; Yang, S.; Liao, W.; Xiong, Y. Modification of antitumor immunity and tumor microenvironment by resveratrol in mouse renal tumor model. Cell Biochem. Biophys. 2015, 72, 617–625. [Google Scholar] [CrossRef] [PubMed]
  111. Lim, S.O.; Li, C.W.; Xia, W.; Cha, J.H.; Chan, L.C.; Wu, Y.; Chang, S.S.; Lin, W.C.; Hsu, J.M.; Hsu, Y.H.; et al. Deubiquitination and stabilization of pd-l1 by csn5. Cancer Cell 2016, 30, 925–939. [Google Scholar] [CrossRef] [PubMed]
  112. Yan, W.; Wang, T.Y.; Fan, Q.M.; Du, L.; Xu, J.K.; Zhai, Z.J.; Li, H.W.; Tang, T.T. Plumbagin attenuates cancer cell growth and osteoclast formation in the bone microenvironment of mice. Acta Pharmacol. Sin. 2014, 35, 124–134. [Google Scholar] [CrossRef] [PubMed]
  113. Li, Z.; Xiao, J.; Wu, X.; Li, W.; Yang, Z.; Xie, J.; Xu, L.; Cai, X.; Lin, Z.; Guo, W.; et al. Plumbagin inhibits breast tumor bone metastasis and osteolysis by modulating the tumor-bone microenvironment. Curr. Mol. Med. 2012, 12, 967–981. [Google Scholar] [CrossRef] [PubMed]
  114. Jang, J.Y.; Lee, J.K.; Jeon, Y.K.; Kim, C.W. Exosome derived from epigallocatechin gallate treated breast cancer cells suppresses tumor growth by inhibiting tumor-associated macrophage infiltration and m2 polarization. BMC Cancer 2013, 13, 421. [Google Scholar] [CrossRef] [PubMed]
  115. Salado, C.; Olaso, E.; Gallot, N.; Valcarcel, M.; Egilegor, E.; Mendoza, L.; Vidal-Vanaclocha, F. Resveratrol prevents inflammation-dependent hepatic melanoma metastasis by inhibiting the secretion and effects of interleukin-18. J. Transl. Med. 2011, 9, 59. [Google Scholar] [CrossRef] [PubMed]
  116. Uttarkar, S.; Piontek, T.; Dukare, S.; Schomburg, C.; Schlenke, P.; Berdel, W.E.; Muller-Tidow, C.; Schmidt, T.J.; Klempnauer, K.H. Small-molecule disruption of the myb/p300 cooperation targets acute myeloid leukemia cells. Mol. Cancer Ther. 2016, 15, 2905–2915. [Google Scholar] [CrossRef] [PubMed]
  117. Azim, S.; Zubair, H.; Srivastava, S.K.; Bhardwaj, A.; Zubair, A.; Ahmad, A.; Singh, S.; Khushman, M.; Singh, A.P. Deep sequencing and in silico analyses identify myb-regulated gene networks and signaling pathways in pancreatic cancer. Sci. Rep. 2016, 6, 28446. [Google Scholar] [CrossRef] [PubMed]
  118. Gray, A.L.; Stephens, C.A.; Bigelow, R.L.; Coleman, D.T.; Cardelli, J.A. The polyphenols (−)-epigallocatechin-3-gallate and luteolin synergistically inhibit tgf-beta-induced myofibroblast phenotypes through rhoa and erk inhibition. PLoS ONE 2014, 9, e109208. [Google Scholar] [CrossRef] [PubMed]
  119. Stylos, E.; Chatziathanasiadou, M.V.; Syriopoulou, A.; Tzakos, A.G. Liquid chromatography coupled with tandem mass spectrometry (LC-MS/MS) based bioavailability determination of the major classes of phytochemicals. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2016. [Google Scholar] [CrossRef] [PubMed]
  120. Maru, G.B.; Hudlikar, R.R.; Kumar, G.; Gandhi, K.; Mahimkar, M.B. Understanding the molecular mechanisms of cancer prevention by dietary phytochemicals: From experimental models to clinical trials. World J. Biol. Chem. 2016, 7, 88–99. [Google Scholar] [CrossRef] [PubMed]
  121. Faggian, M.; Sut, S.; Perissutti, B.; Baldan, V.; Grabnar, I.; Dall'Acqua, S. Natural deep eutectic solvents (nades) as a tool for bioavailability improvement: Pharmacokinetics of rutin dissolved in proline/glycine after oral administration in rats: Possible application in nutraceuticals. Molecules 2016, 21, 1531. [Google Scholar] [CrossRef] [PubMed]
  122. Ahmad, A.; Li, Y.; Sarkar, F.H. The bounty of nature for changing the cancer landscape. Mol. Nutr. Food Res. 2016, 60, 1251–1263. [Google Scholar] [CrossRef] [PubMed]
  123. Hackler, L., Jr.; Ozsvari, B.; Gyuris, M.; Sipos, P.; Fabian, G.; Molnar, E.; Marton, A.; Farago, N.; Mihaly, J.; Nagy, L.I.; et al. The curcumin analog c-150, influencing nf-kappab, upr and akt/notch pathways has potent anticancer activity in vitro and in vivo. PLoS ONE 2016, 11, e0149832. [Google Scholar] [CrossRef] [PubMed]
  124. Jin, R.; Xia, Y.; Chen, Q.; Li, W.; Chen, D.; Ye, H.; Zhao, C.; Du, X.; Shi, D.; Wu, J.; et al. Da0324, an inhibitor of nuclear factor-kappab activation, demonstrates selective antitumor activity on human gastric cancer cells. Drug Des. Dev. Ther. 2016, 10, 979–995. [Google Scholar]
  125. Liu, G.Y.; Zhai, Q.; Chen, J.Z.; Zhang, Z.Q.; Yang, J. 2,2'-fluorine mono-carbonyl curcumin induce reactive oxygen species-mediated apoptosis in human lung cancer nci-h460 cells. Eur. J. Pharmacol. 2016, 786, 161–168. [Google Scholar] [CrossRef] [PubMed]
  126. Ye, H.; Wei, X.; Wang, Z.; Zhang, S.; Ren, J.; Yao, S.; Shi, L.; Yang, L.; Qiu, P.; Wu, J.; et al. A novel double carbonyl analog of curcumin induces the apoptosis of human lung cancer h460 cells via the activation of the endoplasmic reticulum stress signaling pathway. Oncol. Rep. 2016, 36, 1640–1648. [Google Scholar] [CrossRef] [PubMed]
  127. Liang, B.; Liu, Z.; Cao, Y.; Zhu, C.; Zuo, Y.; Huang, L.; Wen, G.; Shang, N.; Chen, Y.; Yue, X.; et al. Mc37, a new mono-carbonyl curcumin analog, induces g2/m cell cycle arrest and mitochondria-mediated apoptosis in human colorectal cancer cells. Eur. J. Pharmacol. 2016, 796, 139–148. [Google Scholar] [CrossRef] [PubMed]
  128. Hu, Y.; Zhao, C.; Zheng, H.; Lu, K.; Shi, D.; Liu, Z.; Dai, X.; Zhang, Y.; Zhang, X.; Hu, W.; et al. A novel stat3 inhibitor ho-3867 induces cell apoptosis by reactive oxygen species-dependent endoplasmic reticulum stress in human pancreatic cancer cells. Anticancer Drugs 2017. [Google Scholar] [CrossRef] [PubMed]
  129. Mohankumar, K.; Sridharan, S.; Pajaniradje, S.; Singh, V.K.; Ronsard, L.; Banerjea, A.C.; Somasundaram, D.B.; Coumar, M.S.; Periyasamy, L.; Rajagopalan, R. Bdmc-a, an analog of curcumin, inhibits markers of invasion, angiogenesis, and metastasis in breast cancer cells via nf-kappab pathway--a comparative study with curcumin. Biomed. Pharmacother. 2015, 74, 178–186. [Google Scholar] [CrossRef] [PubMed]
  130. Sugiyama, S.; Yoshino, Y.; Kuriyama, S.; Inoue, M.; Komine, K.; Otsuka, K.; Kohyama, A.; Yamakoshi, H.; Ishioka, C.; Tanaka, M.; et al. A curcumin analog, go-y078, effectively inhibits angiogenesis through actin disorganization. Anticancer Agents Med. Chem. 2016, 16, 633–647. [Google Scholar] [CrossRef] [PubMed]
  131. Faiao-Flores, F.; Quincoces Suarez, J.A.; Fruet, A.C.; Maria-Engler, S.S.; Pardi, P.C.; Maria, D.A. Curcumin analog dm-1 in monotherapy or combinatory treatment with dacarbazine as a strategy to inhibit in vivo melanoma progression. PLoS ONE 2015, 10, e0118702. [Google Scholar] [CrossRef] [PubMed]
  132. Haque, A.; Rahman, M.A.; Fuchs, J.R.; Chen, Z.G.; Khuri, F.R.; Shin, D.M.; Amin, A.R. Flll12 induces apoptosis in lung cancer cells through a p53/p73-independent but death receptor 5-dependent pathway. Cancer Lett. 2015, 363, 166–175. [Google Scholar] [CrossRef] [PubMed]
  133. Shen, T.; Jiang, T.; Long, M.; Chen, J.; Ren, D.M.; Wong, P.K.; Chapman, E.; Zhou, B.; Zhang, D.D. A curcumin derivative that inhibits vinyl carbamate-induced lung carcinogenesis via activation of the nrf2 protective response. Antioxid. Redox. Signal 2015, 23, 651–664. [Google Scholar] [CrossRef] [PubMed]
  134. Gundewar, C.; Ansari, D.; Tang, L.; Wang, Y.; Liang, G.; Rosendahl, A.H.; Saleem, M.A.; Andersson, R. Antiproliferative effects of curcumin analog l49h37 in pancreatic stellate cells: A comparative study. Ann. Gastroenterol. 2015, 28, 391–398. [Google Scholar] [PubMed]
  135. Ahmad, A.; Sayed, A.; Ginnebaugh, K.R.; Sharma, V.; Suri, A.; Saraph, A.; Padhye, S.; Sarkar, F.H. Molecular docking and inhibition of matrix metalloproteinase-2 by novel difluorinatedbenzylidene curcumin analog. Am. J. Transl. Res. 2015, 7, 298–308. [Google Scholar] [PubMed]
  136. Roy, S.; Yu, Y.; Padhye, S.B.; Sarkar, F.H.; Majumdar, A.P. Difluorinated-curcumin (cdf) restores pten expression in colon cancer cells by down-regulating mir-21. PLoS ONE 2013, 8, e68543. [Google Scholar] [CrossRef] [PubMed]
  137. Bao, B.; Ali, S.; Banerjee, S.; Wang, Z.; Logna, F.; Azmi, A.S.; Kong, D.; Ahmad, A.; Li, Y.; Padhye, S.; et al. Curcumin analogue cdf inhibits pancreatic tumor growth by switching on suppressor micrornas and attenuating ezh2 expression. Cancer Res. 2012, 72, 335–345. [Google Scholar] [CrossRef] [PubMed]
  138. Kanwar, S.S.; Yu, Y.; Nautiyal, J.; Patel, B.B.; Padhye, S.; Sarkar, F.H.; Majumdar, A.P. Difluorinated-curcumin (cdf): A novel curcumin analog is a potent inhibitor of colon cancer stem-like cells. Pharm. Res. 2011, 28, 827–838. [Google Scholar] [CrossRef] [PubMed]
  139. Chen, M.; Zhou, B.; Zhong, P.; Rajamanickam, V.; Dai, X.; Karvannan, K.; Zhou, H.; Zhang, X.; Liang, G. Increased intracellular reactive oxygen species mediates the anti-cancer effects of wz35 via activating mitochondrial apoptosis pathway in prostate cancer cells. Prostate 2016. [Google Scholar] [CrossRef] [PubMed]
  140. Zou, P.; Zhang, J.; Xia, Y.; Kanchana, K.; Guo, G.; Chen, W.; Huang, Y.; Wang, Z.; Yang, S.; Liang, G. Ros generation mediates the anti-cancer effects of wz35 via activating jnk and er stress apoptotic pathways in gastric cancer. Oncotarget 2015, 6, 5860–5876. [Google Scholar] [CrossRef] [PubMed]
  141. Zhang, X.; Chen, M.; Zou, P.; Kanchana, K.; Weng, Q.; Chen, W.; Zhong, P.; Ji, J.; Zhou, H.; He, L.; et al. Curcumin analog wz35 induced cell death via ros-dependent er stress and g2/m cell cycle arrest in human prostate cancer cells. BMC Cancer 2015, 15, 866. [Google Scholar] [CrossRef] [PubMed]
  142. Zhao, R.; Tin, L.; Zhang, Y.; Wu, Y.; Jin, Y.; Jin, X.; Zhang, F.; Li, X. Ef24 suppresses invasion and migration of hepatocellular carcinoma cells in vitro via inhibiting the phosphorylation of src. Biomed. Res. Int. 2016, 2016, 8569684. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, P.; Bai, H.; Liu, G.; Wang, H.; Chen, F.; Zhang, B.; Zeng, P.; Wu, C.; Peng, C.; Huang, C.; et al. Microrna-33b, upregulated by ef24, a curcumin analog, suppresses the epithelial-to-mesenchymal transition (emt) and migratory potential of melanoma cells by targeting hmga2. Toxicol. Lett. 2015, 234, 151–161. [Google Scholar] [CrossRef] [PubMed]
  144. He, G.; Feng, C.; Vinothkumar, R.; Chen, W.; Dai, X.; Chen, X.; Ye, Q.; Qiu, C.; Zhou, H.; Wang, Y.; et al. Curcumin analog ef24 induces apoptosis via ros-dependent mitochondrial dysfunction in human colorectal cancer cells. Cancer Chemother. Pharmacol. 2016, 78, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  145. Yin, D.L.; Liang, Y.J.; Zheng, T.S.; Song, R.P.; Wang, J.B.; Sun, B.S.; Pan, S.H.; Qu, L.D.; Liu, J.R.; Jiang, H.C.; et al. Ef24 inhibits tumor growth and metastasis via suppressing nf-kappab dependent pathways in human cholangiocarcinoma. Sci. Rep. 2016, 6, 32167. [Google Scholar] [CrossRef] [PubMed]
  146. Yar Saglam, A.S.; Yilmaz, A.; Onen, H.I.; Alp, E.; Kayhan, H.; Ekmekci, A. Hdac inhibitors, ms-275 and salermide, potentiates the anticancer effect of ef24 in human pancreatic cancer cells. Excli J. 2016, 15, 246–255. [Google Scholar] [PubMed]
  147. Lu, J.; Ho, C.H.; Ghai, G.; Chen, K.Y. Resveratrol analog, 3,4,5,4′-tetrahydroxystilbene, differentially induces pro-apoptotic p53/bax gene expression and inhibits the growth of transformed cells but not their normal counterparts. Carcinogenesis 2001, 22, 321–328. [Google Scholar] [CrossRef] [PubMed]
  148. Nam, K.A.; Kim, S.; Heo, Y.H.; Lee, S.K. Resveratrol analog, 3,5,2′,4′-tetramethoxy-trans-stilbene, potentiates the inhibition of cell growth and induces apoptosis in human cancer cells. Arch. Pharm. Res. 2001, 24, 441–445. [Google Scholar] [CrossRef] [PubMed]
  149. Hong, Y.B.; Kang, H.J.; Kim, H.J.; Rosen, E.M.; Dakshanamurthy, S.; Rondanin, R.; Baruchello, R.; Grisolia, G.; Daniele, S.; Bae, I. Inhibition of cell proliferation by a resveratrol analog in human pancreatic and breast cancer cells. Exp. Mol. Med. 2009, 41, 151–160. [Google Scholar] [CrossRef] [PubMed]
  150. Saiko, P.; Graser, G.; Giessrigl, B.; Steinmann, M.T.; Schuster, H.; Lackner, A.; Grusch, M.; Krupitza, G.; Jaeger, W.; Somepalli, V.; et al. Digalloylresveratrol, a novel resveratrol analog inhibits the growth of human pancreatic cancer cells. Invest. New Drugs 2013, 31, 1115–1124. [Google Scholar] [CrossRef] [PubMed]
  151. Aldawsari, F.S.; Aguayo-Ortiz, R.; Kapilashrami, K.; Yoo, J.; Luo, M.; Medina-Franco, J.L.; Velazquez-Martinez, C.A. Resveratrol-salicylate derivatives as selective dnmt3 inhibitors and anticancer agents. J. Enzyme Inhib. Med. Chem. 2016, 31, 695–703. [Google Scholar] [CrossRef] [PubMed]
  152. Aldawsari, F.S.; Aguiar, R.P.; Wiirzler, L.A.; Aguayo-Ortiz, R.; Aljuhani, N.; Cuman, R.K.; Medina-Franco, J.L.; Siraki, A.G.; Velazquez-Martinez, C.A. Anti-inflammatory and antioxidant properties of a novel resveratrol-salicylate hybrid analog. Bioorg. Med. Chem. Lett. 2016, 26, 1411–1415. [Google Scholar] [CrossRef] [PubMed]
  153. Thomas, E.; Gopalakrishnan, V.; Hegde, M.; Kumar, S.; Karki, S.S.; Raghavan, S.C.; Choudhary, B. A novel resveratrol based tubulin inhibitor induces mitotic arrest and activates apoptosis in cancer cells. Sci. Rep. 2016, 6, 34653. [Google Scholar] [CrossRef] [PubMed]
  154. Ma, Z.; Molavi, O.; Haddadi, A.; Lai, R.; Gossage, R.A.; Lavasanifar, A. Resveratrol analog trans 3,4,5,4′-tetramethoxystilbene (dmu-212) mediates anti-tumor effects via mechanism different from that of resveratrol. Cancer Chemother. Pharmacol. 2008, 63, 27–35. [Google Scholar] [CrossRef] [PubMed]
  155. Cichocki, M.; Baer-Dubowska, W.; Wierzchowski, M.; Murias, M.; Jodynis-Liebert, J. 3,4,5,4′-trans-tetramethoxystilbene (dmu-212) modulates the activation of nf-kappab, ap-1, and stat3 transcription factors in rat liver carcinogenesis induced by initiation-promotion regimen. Mol. Cell Biochem. 2014, 391, 27–35. [Google Scholar] [CrossRef] [PubMed]
  156. Penthala, N.R.; Thakkar, S.; Crooks, P.A. Heteroaromatic analogs of the resveratrol analog dmu-212 as potent anti-cancer agents. Bioorg. Med. Chem. Lett. 2015, 25, 2763–2767. [Google Scholar] [CrossRef] [PubMed]
  157. Jeong, N.Y.; Yoon, Y.G.; Rho, J.H.; Lee, J.S.; Lee, S.Y.; Yoo, K.S.; Song, S.; Suh, H.; Choi, Y.H.; Yoo, Y.H. The novel resveratrol analog hs-1793-induced polyploid lncap prostate cancer cells are vulnerable to downregulation of bcl-xl. Int. J. Oncol. 2011, 38, 1597–1604. [Google Scholar] [PubMed]
  158. Jeong, M.H.; Yang, K.M.; Choi, Y.J.; Kim, S.D.; Yoo, Y.H.; Seo, S.Y.; Lee, S.H.; Ryu, S.R.; Lee, C.M.; Suh, H.; et al. Resveratrol analog, HS-1793 enhance anti-tumor immunity by reducing the CD4+CD25+ regulatory T cells in FM3A tumor bearing mice. Int. Immunopharmacol. 2012, 14, 328–333. [Google Scholar] [CrossRef] [PubMed]
  159. Jeong, S.H.; Song, I.S.; Kim, H.K.; Lee, S.R.; Song, S.; Suh, H.; Yoon, Y.G.; Yoo, Y.H.; Kim, N.; Rhee, B.D.; et al. An analogue of resveratrol hs-1793 exhibits anticancer activity against mcf-7 cells via inhibition of mitochondrial biogenesis gene expression. Mol. Cells 2012, 34, 357–365. [Google Scholar] [CrossRef] [PubMed]
  160. Jeong, S.K.; Yang, K.; Park, Y.S.; Choi, Y.J.; Oh, S.J.; Lee, C.W.; Lee, K.Y.; Jeong, M.H.; Jo, W.S. Interferon gamma induced by resveratrol analog, hs-1793, reverses the properties of tumor associated macrophages. Int. Immunopharmacol. 2014, 22, 303–310. [Google Scholar] [CrossRef] [PubMed]
  161. Kim, H.J.; Yang, K.M.; Park, Y.S.; Choi, Y.J.; Yun, J.H.; Son, C.H.; Suh, H.S.; Jeong, M.H.; Jo, W.S. The novel resveratrol analogue hs-1793 induces apoptosis via the mitochondrial pathway in murine breast cancer cells. Int. J. Oncol. 2012, 41, 1628–1634. [Google Scholar] [PubMed]
  162. Choi, Y.J.; Heo, K.; Park, H.S.; Yang, K.M.; Jeong, M.H. The resveratrol analog hs-1793 enhances radiosensitivity of mouse-derived breast cancer cells under hypoxic conditions. Int. J. Oncol. 2016, 49, 1479–1488. [Google Scholar] [CrossRef] [PubMed]
  163. Waleh, N.S.; Chao, W.R.; Bensari, A.; Zaveri, N.T. Novel d-ring analog of epigallocatechin-3-gallate inhibits tumor growth and vegf expression in breast carcinoma cells. Anticancer Res. 2005, 25, 397–402. [Google Scholar] [PubMed]
  164. Hashimoto, O.; Nakamura, A.; Nakamura, T.; Iwamoto, H.; Hiroshi, M.; Inoue, K.; Torimura, T.; Ueno, T.; Sata, M. Methylated-(3″)-epigallocatechin gallate analog suppresses tumor growth in huh7 hepatoma cells via inhibition of angiogenesis. Nutr. Cancer 2014, 66, 728–735. [Google Scholar] [CrossRef] [PubMed]
  165. Yu, Z.; Qin, X.L.; Gu, Y.Y.; Chen, D.; Cui, Q.C.; Jiang, T.; Wan, S.B.; Dou, Q.P. Prodrugs of fluoro-substituted benzoates of egc as tumor cellular proteasome inhibitors and apoptosis inducers. Int. J. Mol. Sci. 2008, 9, 951–961. [Google Scholar] [CrossRef] [PubMed]
  166. Yang, H.; Sun, D.K.; Chen, D.; Cui, Q.C.; Gu, Y.Y.; Jiang, T.; Chen, W.; Wan, S.B.; Dou, Q.P. Antitumor activity of novel fluoro-substituted (−)-epigallocatechin-3-gallate analogs. Cancer Lett. 2010, 292, 48–53. [Google Scholar] [CrossRef] [PubMed]
  167. Huo, C.; Yang, H.; Cui, Q.C.; Dou, Q.P.; Chan, T.H. Proteasome inhibition in human breast cancer cells with high catechol-o-methyltransferase activity by green tea polyphenol egcg analogs. Bioorg. Med. Chem. 2010, 18, 1252–1258. [Google Scholar] [CrossRef] [PubMed]
  168. Dandawate, P.; Vemuri, K.; Venkateswara Swamy, K.; Khan, E.M.; Sritharan, M.; Padhye, S. Synthesis, characterization, molecular docking and anti-tubercular activity of plumbagin-isoniazid analog and its beta-cyclodextrin conjugate. Bioorg. Med. Chem. Lett. 2014, 24, 5070–5075. [Google Scholar] [CrossRef] [PubMed]
  169. Hahm, E.R.; Karlsson, A.I.; Bonner, M.Y.; Arbiser, J.L.; Singh, S.V. Honokiol inhibits androgen receptor activity in prostate cancer cells. Prostate 2014, 74, 408–420. [Google Scholar] [CrossRef] [PubMed]
  170. Bonner, M.Y.; Karlsson, I.; Rodolfo, M.; Arnold, R.S.; Vergani, E.; Arbiser, J.L. Honokiol bis-dichloroacetate (honokiol dca) demonstrates activity in vemurafenib-resistant melanoma in vivo. Oncotarget 2016, 7, 12857–12868. [Google Scholar] [PubMed]
  171. Arora, S.; Tyagi, N.; Bhardwaj, A.; Rusu, L.; Palanki, R.; Vig, K.; Singh, S.R.; Singh, A.P.; Palanki, S.; Miller, M.E.; et al. Silver nanoparticles protect human keratinocytes against uvb radiation-induced DNA damage and apoptosis: Potential for prevention of skin carcinogenesis. Nanomed.: Nanotechnol. Biol. Med. 2015, 11, 1265–1275. [Google Scholar] [CrossRef] [PubMed]
  172. Aras, A.; Khokhar, A.R.; Qureshi, M.Z.; Silva, M.F.; Sobczak-Kupiec, A.; Pineda, E.A.; Hechenleitner, A.A.; Farooqi, A.A. Targeting cancer with nano-bullets: Curcumin, egcg, resveratrol and quercetin on flying carpets. Asian Pac. J. Cancer Prev. 2014, 15, 3865–3871. [Google Scholar] [CrossRef] [PubMed]
  173. Nayak, A.P.; Mills, T.; Norton, I. Lipid based nanosystems for curcumin: Past, present and future. Curr. Pharm. Des. 2016, 22, 4247–4256. [Google Scholar] [PubMed]
  174. Rahimi, H.R.; Nedaeinia, R.; Sepehri Shamloo, A.; Nikdoust, S.; Kazemi Oskuee, R. Novel delivery system for natural products: Nano-curcumin formulations. Avicenna J. Phytomed. 2016, 6, 383–398. [Google Scholar] [PubMed]
  175. Strojny, B.; Grodzik, M.; Sawosz, E.; Winnicka, A.; Kurantowicz, N.; Jaworski, S.; Kutwin, M.; Urbanska, K.; Hotowy, A.; Wierzbicki, M.; et al. Diamond nanoparticles modify curcumin activity: In vitro studies on cancer and normal cells and in ovo studies on chicken embryo model. PLoS ONE 2016, 11, e0164637. [Google Scholar] [CrossRef] [PubMed]
  176. Yoon, I.S.; Park, J.H.; Kang, H.J.; Choe, J.H.; Goh, M.S.; Kim, D.D.; Cho, H.J. Poly(d,l-lactic acid)-glycerol-based nanoparticles for curcumin delivery. Int. J. Pharm. 2015, 488, 70–77. [Google Scholar] [CrossRef] [PubMed]
  177. Zaman, M.S.; Chauhan, N.; Yallapu, M.M.; Gara, R.K.; Maher, D.M.; Kumari, S.; Sikander, M.; Khan, S.; Zafar, N.; Jaggi, M.; et al. Curcumin nanoformulation for cervical cancer treatment. Sci. Rep. 2016, 6, 20051. [Google Scholar] [CrossRef] [PubMed]
  178. Kheiri Manjili, H.; Ghasemi, P.; Malvandi, H.; Mousavi, M.S.; Attari, E.; Danafar, H. Pharmacokinetics and in vivo delivery of curcumin by copolymeric mpeg-pcl micelles. Eur. J. Pharm. Biopharm. 2016. [Google Scholar] [CrossRef] [PubMed]
  179. Chen, D.; Lian, S.; Sun, J.; Liu, Z.; Zhao, F.; Jiang, Y.; Gao, M.; Sun, K.; Liu, W.; Fu, F. Design of novel multifunctional targeting nano-carrier drug delivery system based on cd44 receptor and tumor microenvironment ph condition. Drug Deliv. 2016, 23, 808–813. [Google Scholar] [CrossRef] [PubMed]
  180. Ndong Ntoutoume, G.M.; Granet, R.; Mbakidi, J.P.; Bregier, F.; Leger, D.Y.; Fidanzi-Dugas, C.; Lequart, V.; Joly, N.; Liagre, B.; Chaleix, V.; et al. Development of curcumin-cyclodextrin/cellulose nanocrystals complexes: New anticancer drug delivery systems. Bioorg. Med. Chem. Lett. 2016, 26, 941–945. [Google Scholar] [CrossRef] [PubMed]
  181. Yallapu, M.M.; Khan, S.; Maher, D.M.; Ebeling, M.C.; Sundram, V.; Chauhan, N.; Ganju, A.; Balakrishna, S.; Gupta, B.K.; Zafar, N.; et al. Anti-cancer activity of curcumin loaded nanoparticles in prostate cancer. Biomaterials 2014, 35, 8635–8648. [Google Scholar] [CrossRef] [PubMed]
  182. Fatima, M.T.; Chanchal, A.; Yavvari, P.S.; Bhagat, S.D.; Gujrati, M.; Mishra, R.K.; Srivastava, A. Cell permeating nano-complexes of amphiphilic polyelectrolytes enhance solubility, stability, and anti-cancer efficacy of curcumin. Biomacromolecules 2016, 17, 2375–2383. [Google Scholar] [CrossRef] [PubMed]
  183. Yan, J.; Wang, Y.; Zhang, X.; Liu, S.; Tian, C.; Wang, H. Targeted nanomedicine for prostate cancer therapy: Docetaxel and curcumin co-encapsulated lipid-polymer hybrid nanoparticles for the enhanced anti-tumor activity in vitro and in vivo. Drug Deliv. 2016, 23, 1757–1762. [Google Scholar] [CrossRef] [PubMed]
  184. Fang, X.B.; Zhang, J.M.; Xie, X.; Liu, D.; He, C.W.; Wan, J.B.; Chen, M.W. Ph-sensitive micelles based on acid-labile pluronic f68-curcumin conjugates for improved tumor intracellular drug delivery. Int. J. Pharm. 2016, 502, 28–37. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, J.; Ma, W.; Tu, P. Synergistically improved anti-tumor efficacy by co-delivery doxorubicin and curcumin polymeric micelles. Macromol. Biosci. 2015, 15, 1252–1261. [Google Scholar] [CrossRef] [PubMed]
  186. Kesharwani, P.; Banerjee, S.; Padhye, S.; Sarkar, F.H.; Iyer, A.K. Parenterally administrable nano-micelles of 3,4-difluorobenzylidene curcumin for treating pancreatic cancer. Colloids Surf. B Biointerfaces 2015, 132, 138–145. [Google Scholar] [CrossRef] [PubMed]
  187. Basak, S.K.; Zinabadi, A.; Wu, A.W.; Venkatesan, N.; Duarte, V.M.; Kang, J.J.; Dalgard, C.L.; Srivastava, M.; Sarkar, F.H.; Wang, M.B.; et al. Liposome encapsulated curcumin-difluorinated (cdf) inhibits the growth of cisplatin resistant head and neck cancer stem cells. Oncotarget 2015, 6, 18504–18517. [Google Scholar] [CrossRef] [PubMed]
  188. Kesharwani, P.; Banerjee, S.; Padhye, S.; Sarkar, F.H.; Iyer, A.K. Hyaluronic acid engineered nanomicelles loaded with 3,4-difluorobenzylidene curcumin for targeted killing of cd44+ stem-like pancreatic cancer cells. Biomacromolecules 2015, 16, 3042–3053. [Google Scholar] [CrossRef] [PubMed]
  189. Kesharwani, P.; Xie, L.; Banerjee, S.; Mao, G.; Padhye, S.; Sarkar, F.H.; Iyer, A.K. Hyaluronic acid-conjugated polyamidoamine dendrimers for targeted delivery of 3,4-difluorobenzylidene curcumin to cd44 overexpressing pancreatic cancer cells. Colloids Surf. B Biointerfaces 2015, 136, 413–423. [Google Scholar] [CrossRef] [PubMed]
  190. Bisht, S.; Schlesinger, M.; Rupp, A.; Schubert, R.; Nolting, J.; Wenzel, J.; Holdenrieder, S.; Brossart, P.; Bendas, G.; Feldmann, G. A liposomal formulation of the synthetic curcumin analog ef24 (lipo-ef24) inhibits pancreatic cancer progression: Towards future combination therapies. J. Nanobiotechnol. 2016, 14, 57. [Google Scholar] [CrossRef] [PubMed]
  191. Zhang, J.; Li, S.; An, F.F.; Liu, J.; Jin, S.; Zhang, J.C.; Wang, P.C.; Zhang, X.; Lee, C.S.; Liang, X.J. Self-carried curcumin nanoparticles for in vitro and in vivo cancer therapy with real-time monitoring of drug release. Nanoscale 2015, 7, 13503–13510. [Google Scholar] [CrossRef] [PubMed]
  192. Gupta, R.C.; Bansal, S.S.; Aqil, F.; Jeyabalan, J.; Cao, P.; Kausar, H.; Russell, G.K.; Munagala, R.; Ravoori, S.; Vadhanam, M.V. Controlled-release systemic delivery—A new concept in cancer chemoprevention. Carcinogenesis 2012, 33, 1608–1615. [Google Scholar] [CrossRef] [PubMed]
  193. Sanna, V.; Siddiqui, I.A.; Sechi, M.; Mukhtar, H. Resveratrol-loaded nanoparticles based on poly(epsilon-caprolactone) and poly(d,l-lactic-co-glycolic acid)-poly(ethylene glycol) blend for prostate cancer treatment. Mol. Pharm. 2013, 10, 3871–3881. [Google Scholar] [CrossRef] [PubMed]
  194. Pulliero, A.; Wu, Y.; Fenoglio, D.; Parodi, A.; Romani, M.; Soares, C.P.; Filaci, G.; Lee, J.L.; Sinkam, P.N.; Izzotti, A. Nanoparticles increase the efficacy of cancer chemopreventive agents in cells exposed to cigarette smoke condensate. Carcinogenesis 2015, 36, 368–377. [Google Scholar] [CrossRef] [PubMed]
  195. Pujara, N.; Jambhrunkar, S.; Wong, K.Y.; McGuckin, M.; Popat, A. Enhanced colloidal stability, solubility and rapid dissolution of resveratrol by nanocomplexation with soy protein isolate. J. Colloid Interface Sci. 2017, 488, 303–308. [Google Scholar] [CrossRef] [PubMed]
  196. Park, S.Y.; Chae, S.Y.; Park, J.O.; Lee, K.J.; Park, G. Gold-conjugated resveratrol nanoparticles attenuate the invasion and mmp-9 and cox-2 expression in breast cancer cells. Oncol. Rep. 2016, 35, 3248–3256. [Google Scholar] [CrossRef] [PubMed]
  197. Han, D.W.; Matsumura, K.; Kim, B.; Hyon, S.H. Time-dependent intracellular trafficking of fitc-conjugated epigallocatechin-3-o-gallate in l-929 cells. Bioorg. Med. Chem. 2008, 16, 9652–9659. [Google Scholar] [CrossRef] [PubMed]
  198. Shutava, T.G.; Balkundi, S.S.; Vangala, P.; Steffan, J.J.; Bigelow, R.L.; Cardelli, J.A.; O'Neal, D.P.; Lvov, Y.M. Layer-by-layer-coated gelatin nanoparticles as a vehicle for delivery of natural polyphenols. ACS Nano 2009, 3, 1877–1885. [Google Scholar] [CrossRef] [PubMed]
  199. Siddiqui, I.A.; Mukhtar, H. Nanochemoprevention by bioactive food components: A perspective. Pharm. Res. 2010, 27, 1054–1060. [Google Scholar] [CrossRef] [PubMed]
  200. Wang, D.; Taylor, E.W.; Wang, Y.; Wan, X.; Zhang, J. Encapsulated nanoepigallocatechin-3-gallate and elemental selenium nanoparticles as paradigms for nanochemoprevention. Int. J. Nanomed. 2012, 7, 1711–1721. [Google Scholar]
  201. Siddiqui, I.A.; Adhami, V.M.; Ahmad, N.; Mukhtar, H. Nanochemoprevention: Sustained release of bioactive food components for cancer prevention. Nutr. Cancer 2010, 62, 883–890. [Google Scholar] [CrossRef] [PubMed]
  202. Shafiei, S.S.; Solati-Hashjin, M.; Samadikuchaksaraei, A.; Kalantarinejad, R.; Asadi-Eydivand, M.; Abu Osman, N.A. Epigallocatechin gallate/layered double hydroxide nanohybrids: Preparation, characterization, and in vitro anti-tumor study. PLoS ONE 2015, 10, e0136530. [Google Scholar] [CrossRef] [PubMed]
  203. Narayanan, S.; Mony, U.; Vijaykumar, D.K.; Koyakutty, M.; Paul-Prasanth, B.; Menon, D. Sequential release of epigallocatechin gallate and paclitaxel from plga-casein core/shell nanoparticles sensitizes drug-resistant breast cancer cells. Nanomed.: Nanotechnol. Biol. Med. 2015, 11, 1399–1406. [Google Scholar] [CrossRef] [PubMed]
  204. Duraipandy, N.; Lakra, R.; Kunnavakkam Vinjimur, S.; Samanta, D.; K, P.S.; Kiran, M.S. Caging of plumbagin on silver nanoparticles imparts selectivity and sensitivity to plumbagin for targeted cancer cell apoptosis. Metallomics 2014, 6, 2025–2033. [Google Scholar] [CrossRef] [PubMed]
  205. Appadurai, P.; Rathinasamy, K. Plumbagin-silver nanoparticle formulations enhance the cellular uptake of plumbagin and its antiproliferative activities. IET Nanobiotechnol. 2015, 9, 264–272. [Google Scholar] [CrossRef] [PubMed]
  206. Nair, H.A.; Snima, K.S.; Kamath, R.C.; Nair, S.V.; Lakshmanan, V.K. Plumbagin nanoparticles induce dose and ph dependent toxicity on prostate cancer cells. Curr. Drug Deliv. 2015, 12, 709–716. [Google Scholar] [CrossRef] [PubMed]
  207. Gou, M.; Zheng, L.; Peng, X.; Men, K.; Zheng, X.; Zeng, S.; Guo, G.; Luo, F.; Zhao, X.; Chen, L.; et al. Poly(epsilon-caprolactone)-poly(ethylene glycol)-poly(epsilon-caprolactone) (pcl-peg-pcl) nanoparticles for honokiol delivery in vitro. Int. J. Pharm. 2009, 375, 170–176. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, B.; Gou, M.; Zheng, X.; Wei, X.; Gong, C.; Wang, X.; Zhao, Y.; Luo, F.; Chen, L.; Qian, Z.; et al. Co-delivery honokiol and doxorubicin in mpeg-pla nanoparticles. J. Nanosci. Nanotechnol. 2010, 10, 4166–4172. [Google Scholar] [CrossRef] [PubMed]
  209. Zheng, X.; Kan, B.; Gou, M.; Fu, S.; Zhang, J.; Men, K.; Chen, L.; Luo, F.; Zhao, Y.; Zhao, X.; et al. Preparation of mpeg-pla nanoparticle for honokiol delivery in vitro. Int. J. Pharm. 2010, 386, 262–267. [Google Scholar] [CrossRef] [PubMed]
  210. Qiu, N.; Cai, L.L.; Xie, D.; Wang, G.; Wu, W.; Zhang, Y.; Song, H.; Yin, H.; Chen, L. Synthesis, structural and in vitro studies of well-dispersed monomethoxy-poly(ethylene glycol)-honokiol conjugate micelles. Biomed. Mater. 2010, 5, 065006. [Google Scholar] [CrossRef] [PubMed]
  211. Alaarg, A.; Jordan, N.Y.; Verhoef, J.J.; Metselaar, J.M.; Storm, G.; Kok, R.J. Docosahexaenoic acid liposomes for targeting chronic inflammatory diseases and cancer: An in vitro assessment. Int. J. Nanomed. 2016, 11, 5027–5040. [Google Scholar] [CrossRef] [PubMed]
  212. Molfino, A.; Amabile, M.I.; Monti, M.; Arcieri, S.; Rossi Fanelli, F.; Muscaritoli, M. The role of docosahexaenoic acid (dha) in the control of obesity and metabolic derangements in breast cancer. Int. J. Mol. Sci. 2016, 17, 505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Wan, X.H.; Fu, X.; Ababaikeli, G. Docosahexaenoic acid induces growth suppression on epithelial ovarian cancer cells more effectively than eicosapentaenoic acid. Nutr. Cancer 2016, 68, 320–327. [Google Scholar] [CrossRef] [PubMed]
  214. Harvey, K.A.; Xu, Z.; Saaddatzadeh, M.R.; Wang, H.; Pollok, K.; Cohen-Gadol, A.A.; Siddiqui, R.A. Enhanced anticancer properties of lomustine in conjunction with docosahexaenoic acid in glioblastoma cell lines. J. Neurosurg. 2015, 122, 547–556. [Google Scholar] [CrossRef] [PubMed]
  215. Roy, J.; Oliveira, L.T.; Oger, C.; Galano, J.M.; Bultel-Ponce, V.; Richard, S.; Guimaraes, A.G.; Vilela, J.M.; Andrade, M.S.; Durand, T.; et al. Polymeric nanocapsules prevent oxidation of core-loaded molecules: Evidence based on the effects of docosahexaenoic acid and neuroprostane on breast cancer cells proliferation. J. Exp. Clin. Cancer Res. 2015, 34, 155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Zou, L.; Zheng, B.; Zhang, R.; Zhang, Z.; Liu, W.; Liu, C.; Xiao, H.; McClements, D.J. Enhancing the bioaccessibility of hydrophobic bioactive agents using mixed colloidal dispersions: Curcumin-loaded zein nanoparticles plus digestible lipid nanoparticles. Food Res. Int. 2016, 81, 74–82. [Google Scholar] [CrossRef]
  217. Neves, A.R.; Lucio, M.; Martins, S.; Lima, J.L.; Reis, S. Novel resveratrol nanodelivery systems based on lipid nanoparticles to enhance its oral bioavailability. Int. J. Nanomed. 2013, 8, 177–187. [Google Scholar]
  218. Niu, Y.; Bai, J.; Kamm, R.D.; Wang, Y.; Wang, C. Validating antimetastatic effects of natural products in an engineered microfluidic platform mimicking tumor microenvironment. Mol. Pharm. 2014, 11, 2022–2029. [Google Scholar] [CrossRef] [PubMed]
  219. Siddiqui, I.A.; Adhami, V.M.; Bharali, D.J.; Hafeez, B.B.; Asim, M.; Khwaja, S.I.; Ahmad, N.; Cui, H.; Mousa, S.A.; Mukhtar, H. Introducing nanochemoprevention as a novel approach for cancer control: Proof of principle with green tea polyphenol epigallocatechin-3-gallate. Cancer Res. 2009, 69, 1712–1716. [Google Scholar] [CrossRef] [PubMed]
  220. Siddiqui, I.A.; Bharali, D.J.; Nihal, M.; Adhami, V.M.; Khan, N.; Chamcheu, J.C.; Khan, M.I.; Shabana, S.; Mousa, S.A.; Mukhtar, H. Excellent anti-proliferative and pro-apoptotic effects of (−)-epigallocatechin-3-gallate encapsulated in chitosan nanoparticles on human melanoma cell growth both in vitro and in vivo. Nanomed.: Nanotechnol. Biol. Med. 2014, 10, 1619–1626. [Google Scholar] [CrossRef] [PubMed]
  221. Tyagi, N.; Srivastava, S.K.; Arora, S.; Omar, Y.; Ijaz, Z.M.; Al-Ghadhban, A.; Deshmukh, S.K.; Carter, J.E.; Singh, A.P.; Singh, S. Comparative analysis of the relative potential of silver, zinc-oxide and titanium-dioxide nanoparticles against uvb-induced DNA damage for the prevention of skin carcinogenesis. Cancer Lett. 2016, 383, 53–61. [Google Scholar] [CrossRef] [PubMed]
  222. Ahmed, S.; Annu; Chaudhry, S.A.; Ikram, S. A review on biogenic synthesis of zno nanoparticles using plant extracts and microbes: A prospect towards green chemistry. J. Photochem. Photobiol. B 2017, 166, 272–284. [Google Scholar] [CrossRef] [PubMed]
  223. Lam, P.L.; Wong, W.Y.; Bian, Z.; Chui, C.H.; Gambari, R. Recent advances in green nanoparticulate systems for drug delivery: Efficient delivery and safety concern. Nanomedicine (Lond.) 2017. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phytochemicals and their sources. EGCG: Epigallocatechin gallate.
Figure 1. Phytochemicals and their sources. EGCG: Epigallocatechin gallate.
Molecules 22 00395 g001
Figure 2. Role of phytochemicals in human malignancies. Phytochemicals potentially scavenge reactive oxygen species (ROS) or upregulate anti-oxidant signalling to combat ROS generation in cancer cells to inhibit growth. There are also context dependent evidences which advocate the prooxidant cell death inducing behaviour of phytochemicals. Phytochemicals have also been shown to inhibit inflammation via targeting NF-κB pathway. Tumor micro environment (TME) plays a vital role in many solid tumors pathogenesis, and phytochemicals have been shown to target both tumor and stromal compartments.
Figure 2. Role of phytochemicals in human malignancies. Phytochemicals potentially scavenge reactive oxygen species (ROS) or upregulate anti-oxidant signalling to combat ROS generation in cancer cells to inhibit growth. There are also context dependent evidences which advocate the prooxidant cell death inducing behaviour of phytochemicals. Phytochemicals have also been shown to inhibit inflammation via targeting NF-κB pathway. Tumor micro environment (TME) plays a vital role in many solid tumors pathogenesis, and phytochemicals have been shown to target both tumor and stromal compartments.
Molecules 22 00395 g002
Table 1. Inflammation-influencing signaling factors/pathways modulated by phytochemicals.
Table 1. Inflammation-influencing signaling factors/pathways modulated by phytochemicals.
Signaling Factor/PathwayPhytochemicalReference
COX2Curcumin[30]
IL-1βResveratrol[36]
IL-6Resveratrol[35]
IL-8Resveratrol[35]
iNOSCurcumin[30]
TLR/IL-1RCurcumin[31]
Keap1/Nrf2Curcumin[30]
EGCG[37,38]
Honokiol[39]
Plumbagin[40]
Resveratrol[41]
NF-κBCurcumin[42]
EGCG[43,44]
Honokiol[45,46]
Plumbagin[47]
Resveratrol[48]
Table 2. Tumor microenvironment components affected by phytochemicals.
Table 2. Tumor microenvironment components affected by phytochemicals.
TME ComponentPhytochemicalCancer ModelReference
CSCsCurcuminPancreatic[105]
Colorectal[106]
Hh signalingCurcuminPancreatic[101]
HonokiolPancreatic[103]
ResveratrolPancreatic[102]
CXCR4HonokiolPancreatic[103]
IL-6CurcuminMelanoma[109]
ResveratrolRenal[110]
EGCGBreast[114]
IL-18ResveratrolMelanoma[115]
MicrovasculatureEGCGNSCLC[104]
Myofibroblast DifferentiationEGCGProstate[118]
RANKPlumbaginBreast[113]
Regulatory T-cellsCurcuminMelanoma[109]
ResveratrolRenal[110]
Table 3. Chemical analogues of phytochemicals and their reported effects.
Table 3. Chemical analogues of phytochemicals and their reported effects.
PhytochemicalAnalogueReported ActivityReference
CurcuminC-150Inhibits NF-κB[123]
Da0324Inhibits NF-κB[124]
2-2′-fluorine mono-carbonyl analogModulates ROS[125]
A17Induces ER stress[126]
MC37Induces cell cycle arrest[127]
HO-3867Inhibits STAT3[128]
BDMC-AInhibits NF-κB[129]
GO-Y078Inhibits invasion of endothelial cells[130]
DM-1Induces apoptosis[131]
FLLL12Induces apoptosis[132]
BHBAActivates nrf2[133]
L49H37Induces apoptosis in pancreatic stellate cells[134]
CDFMultiple pathways affected[135,136,137,138]
WZ35Modulates ROS[139,140,141]
EF24Multiple pathways affected[142,143,144,145,146]
EGCGD-ring analogTargets VEGF[163]
Methylated analogTargets VEGF[164]
Fluoro-substitutedInhibits proteasomal activity[165,166]
HonokiolDichloroacetate esterInhibits AR, chemosensitizes[169,170]
PlumbaginIsoniazid analogMultiple pathways affected[168]
ResveratrolDMU-212Inhibits NF-κB[155]
HS-1793Multiple pathways affected[157,158,159,160,161,162]

Share and Cite

MDPI and ACS Style

Zubair, H.; Azim, S.; Ahmad, A.; Khan, M.A.; Patel, G.K.; Singh, S.; Singh, A.P. Cancer Chemoprevention by Phytochemicals: Nature’s Healing Touch. Molecules 2017, 22, 395. https://doi.org/10.3390/molecules22030395

AMA Style

Zubair H, Azim S, Ahmad A, Khan MA, Patel GK, Singh S, Singh AP. Cancer Chemoprevention by Phytochemicals: Nature’s Healing Touch. Molecules. 2017; 22(3):395. https://doi.org/10.3390/molecules22030395

Chicago/Turabian Style

Zubair, Haseeb, Shafquat Azim, Aamir Ahmad, Mohammad Aslam Khan, Girijesh Kumar Patel, Seema Singh, and Ajay Pratap Singh. 2017. "Cancer Chemoprevention by Phytochemicals: Nature’s Healing Touch" Molecules 22, no. 3: 395. https://doi.org/10.3390/molecules22030395

Article Metrics

Back to TopTop