Next Article in Journal
Topical Plant Polyphenols Prevent Type I Interferon Signaling in the Skin and Suppress Contact Hypersensitivity
Previous Article in Journal
Identification of Putative Novel Class-I Lanthipeptides in Firmicutes: A Combinatorial In Silico Analysis Approach Performed on Genome Sequenced Bacteria and a Close Inspection of Z-Geobacillin Lanthipeptide Biosynthesis Gene Cluster of the Thermophilic Geobacillus sp. Strain ZGt-1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structure-Based Classification and Anti-Cancer Effects of Plant Metabolites

1
Collage of Pharmacy and Research Institute of Pharmaceutical Sciences, Gyeongsang National University, Jinju 52828, Korea
2
Division of Applied Life Science (BK21Plus), Plant Molecular Biology and Biotechnology Research Center (PMBBRC), Research Institute of Life Sciences (RILS), Gyeongsang National University, Jinju 52828, Korea
3
College of Pharmacy, Chung-Ang University, Seoul 06974, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2018, 19(9), 2651; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms19092651
Submission received: 24 July 2018 / Revised: 4 September 2018 / Accepted: 5 September 2018 / Published: 6 September 2018
(This article belongs to the Section Bioactives and Nutraceuticals)

Abstract

:
A variety of malignant cancers affect the global human population. Although a wide variety of approaches to cancer treatment have been studied and used clinically (surgery, radiotherapy, chemotherapy, and immunotherapy), the toxic side effects of cancer therapies have a negative impact on patients and impede progress in conquering cancer. Plant metabolites are emerging as new leads for anti-cancer drug development. This review summarizes these plant metabolites with regard to their structures and the types of cancer against which they show activity, organized by the organ or tissues in which each cancer forms. This information will be helpful for understanding the current state of knowledge of the anti-cancer effects of various plant metabolites against major types of cancer for the further development of novel anti-cancer drugs.

1. Introduction

Cancer is characterized by uncontrolled/unlimited cell growth, which can result in death [1]. Although a variety of methods to overcome and treat cancers have been researched, the number of cancer patients continues to increase each year. Furthermore, an estimated 15.5 million people in the world will become cancer patients by 2030, and 11.5 million of these cases are expected to be fatal [2]. Therefore, cancer is the leading cause of mortality and morbidity worldwide [3]. Cancers have been reported to be caused by the dysregulation of key cellular processes, such as growth signaling, anti-apoptotic signaling, immune response, gene stability, and regulation of the stromal microenvironment [1,4]. The treatment of cancer has been focused on re-regulating these cellular functions. Up to the present date, numerous clinical trials have investigated potential cures for cancer via radiation, chemotherapy, antibody treatment, and immunotherapy [5]. Radiation and chemotherapy have severe side effects due to their cytotoxicity to normal cells [3]. Antibody treatment and immunotherapy show highly specific cancer targeting ability, but have a limited target range and can be very expensive [5]. Additionally, many types of cancer tend to relapse and acquire resistance after treatment [3,5]. Currently, combination therapies involving several drugs or therapies are being used to attempt to overcome the limitations and the drawbacks of individual therapies [3,5]. Furthermore, to reduce the side effects of anti-cancer drugs and to discover more effective drugs, new approaches have been developed to identify novel molecules with anti-cancer activity from new sources [3].
Plant species have been used in medical treatment for millennia [3,4]. Additionally, plant-derived metabolites have been reported to be useful for a variety of therapeutic purposes and biotechnological applications [6]. Plant metabolites exhibit a wide range of biological functions, including anti-cancer, analgesic, anti-inflammation, and anti-microbial activities [3]. Plants have generated about 25% of clinically used drugs [7]. More than 60% of drugs with anti-cancer activity originated from plants [8]. As discussed above, the development of new molecules for cancer treatment with fewer side effects and greater efficacy is essential. Plant-derived metabolites are good sources of new anti-cancer drugs with reduced cytotoxicity and increased activity [9]. In this review, we categorize such plant metabolites according to their structure and summarize their activity according to type of cancer.

2. Phytochemicals as Bioactive Metabolites

Phytochemicals are constitutive metabolites that are produced by various parts of plants through their primary or secondary metabolism, and have essential functions in the plant for general growth and defense against animals, insects, microorganisms, and abiotic stress [10,11]. Primary metabolites such as carbohydrates, lipids, and proteins have a direct relationship to the growth and metabolism of the plant. Secondary metabolites, which are biosynthetically derived from primary metabolites, are not necessary for survival, but are involved in important functions in the plant, such as protection, competition, and species interactions [12,13]. These can be classified into three major groups based on their biosynthetic origins: phenolic compounds, terpenoids, and nitrogen/sulfur-containing compounds [14]. These compounds have been investigated for use in carcinomatous-related diseases, and have been reported to have diverse anti-cancer properties, such as anti-proliferation and apoptotic cell death activity. In this review, we categorize these plant metabolites according to their structure and discuss their structure and anti-cancer activity.

2.1. Phenolic Compounds

Phenolic compounds (Figure 1), a type of plant secondary metabolites, are polyhydroxylated phytochemicals found in plant, fruits, vegetables, spices, nuts, and grains [15]. They are one of the most abundant and widely distributed groups of natural compounds available to human beings [16]. Secondary metabolites with phenolic structures play key roles in various ecological relationships between plants and other living things and their physical environment [15,16]. The structures of polyphenol compounds are characterized by at least one aromatic ring with one or more hydroxyl groups [17]. They are categorized by the structural components binding these rings to one another, and by the number of phenol rings that they contain. Polyphenolic compounds are believed to have anti-cancer activity, and include flavonoids, stilbenes, and phenolic acids [18].

2.1.1. Flavonoids

Flavonoids (Figure 1a) are the largest and most diverse sub-group of polyphenolic compounds that are produced as plant secondary metabolites [19]. These compounds are found in various fruits and vegetables, including several medicinal plants, and they also have critical roles in the growth, development, and defense of plants [19]. The basic structure of flavonoids consists of two benzene rings (A and B) linked by a heterocyclic ring (C) with a carbon bridge [20]. Most of the more than 6,000 flavonoids that have been identified from a variety of plants can be categorized into the flavonol, flavone, flavanol, isoflavone, flavanone, or anthocyanidin subclasses according to their structure [21].
Flavonols (Figure 1b) are the most ubiquitous subclass of flavonoids, and are found in plants and fruits such as olives, onions, kale, apples, beans, and green leaves [22]. The main representatives of this subclass are quercetin, kaempferol, myricetin, isorhamnetin, and rutin. Flavonols have a hydroxyl group (-OH) on the 3-position of the C-ring. These hydroxyl groups are present in a glycosylated form in plants in combination with a sugar (commonly glucose or rhamnose) [23,24]. The biological activities of flavonols have been reported to play an important role in preventing carcinogenesis through anti-proliferation, anti-oxidation, and apoptosis activity in various cancer cell lines [25].
Flavones are mainly found in fruits, spices, and vegetables such as celery, olives, onion, garlic, citrus fruits, pepper, and parsley [22,23]. Although the flavone 2-phenyl-4H-1-benzopyran-4-one is the core structure of flavonoids, flavones are much less common than flavonols among plant metabolites [26]. Flavones (Figure 1c) are present chiefly as 7-O-glycosides. They are mainly present in forms such as luteolin and apigenin, while less abundant flavones include tangeretin, nobiletin, baicalein, wogonin, and chrysin [23]. The chemical structure of these flavones consists of a 3-hydroxyflavone backbone, which is the simplest flavone structure, and may contain a broad range of functional groups, including hydroxyl groups, carbonyl groups, and conjugated double bonds [18]. Flavones have been reported to have a variety of biological activities, including antioxidant, anti-proliferative, anti-tumor, anti-microbial, estrogenic, acetyl cholinesterase, and anti-inflammatory activities, and are used for controlling various types of disease, such as cancer, cardiovascular disease, and neurodegenerative disorders [26].
Flavanols, which are sometimes referred to as flavan-3-ols, are derivatives of flavans (Figure 1d). Flavanols have a hydroxyl group at the C3 position [27]. They are the most varied and complex subgroup of flavonoids, and exist in states ranging from single molecules to oligomers, polymers, and other derivatives [28]. Flavanol compounds include catechin, epicatechin, epicatechin-3-O-gallate, theaflavins, epigallocatechin-3-O-gallate, proanthocyanidins, and thearubigins [27,29]. Moreover, they are present in fruits and vegetables such as pears, green leaves, berries, cherries, red grapes, currants, and apples [30]. The flavanols have been reported to exhibit several biological activities such as anti-oxidation, anti-carcinogenesis, cardioprotective, and anti-viral effects [31]. However, most flavanol-related data has been derived from medium/small-scale and short-term (from weeks to several months) dietary intervention studies [32].
Isoflavones (Figure 1e) are secondary metabolites of flavonoids that occur naturally in members of the Leguminosae/Fabaceae family, such as kudzu, lupine, soybeans, red clover, peanuts, chickpeas, broccoli, cauliflower, barley, fava beans, and alfalfa [33,34]. The benzene ring (B) of isoflavones is linked to C3 of the heterocyclic ring by a carbon bridge. The isoflavone compounds include genistein, daidzein, biochanin A, glycitein, and formononetin [34]. Isoflavones are also classified as phytoestrogens because of their structural similarities with estrogen, particularly 17-β-estradiol (a human female hormone), and can bind to both alpha and beta estrogen receptors [24,33,35]. Therefore, they can exert various bioactivities in some hormone-dependent diseases by modulating the expression of genes that control cell survival [35,36].
Flavanones (Figure 1f) are non-planar flavonoids that are derived chiefly in mono- and di-glycoside forms, but are less frequently present in aglycone form [23]. Although flavanones are found in tomatoes and selected aromatic plants such as mints, they are almost exclusively present in high concentrations in citrus fruit [24]. The most common flavanone glycosides, which are generally glycosylated by a disaccharide, are neohesperidin, naringenin, and hesperetin [18]. These glycosides are abundant in the fruit of oranges, grapefruit, and tomatoes, and also found in the peels of citrus, bitter oranges, and grapefruit [37,38].
Anthocyanins (glycosylated forms of anthocyanidin (Figure 1g)) are polyphenolic pigments that belong to the water-soluble flavonoid group, and impart red, blue, and purple colors to plants in a pH-dependent manner [39,40]. They are found in plant organs such as fruits, flowers, and leaves, including those of grapes, berries, pomegranate, red cabbage, purple corn, apples, radishes, tulips, roses, and orchids [39]. More than 700 anthocyanin derivatives have been verified in nature [41]. Anthocyanins vary in their number of hydroxyl groups and the degree of methylation of the aglycone molecule. Additionally, the number and the location of sugars connected to the aglycone molecule, and the number and the character of aliphatic or aromatic acids connected to these sugars, can also vary [23,42]. The most abundant anthocyanins are cyanidin, peonidin, pelargonidin, delphinidin, petunidin, and malvidin [43]. Although anthocyanins are non-essential nutrients, they may promote the maintenance of health and can confer protection against chronic diseases [41]. Recently, research into anthocyanins has been highlighted due to their potential preventative and/or therapeutic effects for a variety of diseases [40].

2.1.2. Stilbenes

Stilbenes (Figure 1h) are a class of nonflavonoid polyphenol phytochemicals [18]. Their molecular backbone consists of 1,2-diphenylethylene units. Stilbenes can be categorized as monomeric and oligomeric stilbenes [44]. These compounds are somewhat limited in plants, since the core enzyme in stilbene biosynthesis, stilbene synthase, is not universally expressed [45]. However, due to their bioactive properties and low toxicity, stilbenes have a remarkable potential for the prevention and treatment of a variety of diseases, including cancer [46,47]. The most representative stilbene derivatives are the stilbenoids, which are hydroxylated derivatives of stilbene that can act as phytoalexins. Such compounds include resveratrol, pterostilbene, gnetol, and piceatannol, and are derived from grapes, berries, peanuts, and other plant sources [45,46]. Among these, resveratrol is the most widely studied stilbenoid. Resveratrol is found as cis- and trans-isomers, as well as conjugated derivatives (trans-resveratrol-3-O-glucoside) [18]. In addition, resveratrols have been reported to show cancer chemopreventive properties by blocking carcinogenesis [48,49,50].

2.1.3. Phenolic Acids

Phenolic acids are secondary metabolites that are present in almost all plant-derived foods including mushrooms, berries, black currants, kiwis, plums, apples, pears, chicory, and potatoes [30,51]. These compounds can be classified into two major groups, hydroxybenzoic and hydroxycinnamic acids, which are derived from the non-phenolic benzoic and cinnamic acids [51]. The most common hydroxybenzoic acids (Figure 1i) are gallic, p-hydroxybenzoic, syringic, vanillic, and protocatechuic acid, while the corresponding hydroxycinnamic acids (Figure 1j) are caffeic, chlorogenic, coumaric, ferulic, and sinapic acid [24]. These compounds are present in both free and bound forms in all plant-derived foods. The bound forms are most frequently esters, glycosides, and bound complexes [52]. Phenolic acids have been reported to have powerful antioxidant properties and biological activities including cardioprotective, anti-carcinogenic, antimicrobial, and hepatoprotective properties [53].

2.2. Terpenoids

Terpenoids (Figure 2), which are also known as isoprenoids, are one of the most numerous and structurally diverse classes of metabolites [54]. They are flammable non-saturated hydrocarbons that exist in the liquid state, and are typically found in essential oils, resins, or oleoresins [55]. Terpenoids are based on linear arrangements of isoprene, and their carbon skeletons consist of two or more carbon units [56,57]. In particular, terpenoids can be classified as mono-, di-, or tetraterpenoids based on isoprenoid biosynthesis in the plastid [18].

2.2.1. Monoterpenoids

Monoterpenoid structures comprise two isoprene units (C10) and can be divided into three sub-groups: acyclic, monocycles, and bicycles (Figure 2a) [56]. The monoterpenoids within each group are simple unsaturated hydrocarbons and can have functional groups such as alcohols, aldehydes, and ketones [56]. The most important representatives are myrcene, citral, linalool, α-terpineol, limonene, thymol, menthol, carvone, eucalyptol, α/β-pinene, borneol, and camphor [58]. Monoterpenoids can be isolated from the fragrant oils of many plants, and are also found in many marine organisms, where they are generally halogenated. In addition, they are well known as components of the essential oils of flowers and herbs, pollinator attractants, and defense compounds [18]. Moreover, monoterpenoids have been reported to potentially act as antioxidants and are widely used as medicines with antimicrobial, antiseptic, disinfectant, and wound-healing properties [59].

2.2.2. Diterpenoids

Diterpenoids constitute a large group of compounds derived from geranylgeranyl pyrophosphate (Figure 2b) [18]. Their structure comprises a C20 carbon skeleton based on four isoprene units [56], and they can be classified into linear, bicyclic, tricyclic, tetracyclic, pentacyclic, or macrocyclic subgroups based on their skeletal core [58]. Diterpenoids are present in higher plants, fungi, insects, and marine organisms [57]. They are typically found in polyoxygenated form with ketone and hydroxyl groups [56]. Typical compounds of this group include phytol, sclareol, marrubiin, salvinorin A, abietic acid, 9-geranyl-α-terpineol, gibberellin A1, ginkgolide A, and taxol [18,60]. Diterpenoids have been reported to have cytotoxic and anti-proliferative properties [61].

2.2.3. Tetraterpenoids (Carotenoids)

Tetraterpenoids consist of eight isoprene units and have a 40-carbon backbone [56]. Carotenoids (Figure 2c), the most common class of tetraterpenoids, are a group of natural pigments produced in plants, algae, bacteria, and fungi [62]. They are the key source of the yellow, orange, and red colors in many plants, including the orange-red colors of oranges, tomatoes, and carrots and the yellow colors of many flowers [62,63]. Carotenoids are essential both in plants and animals. However, they cannot be synthesized in animals, and therefore must be obtained from dietary sources. In addition, carotenoids are known to have protective activity against some forms of cancer, particularly lung cancer [64]. Their beneficial effects are thought to be due to their role as antioxidants [65]. Based on their chemical structure, carotenoids can be generally classified into two classes, carotenes and xanthophylls [58]. Carotenes are non-oxygenated carotenoids that may be linear or possess cyclic hydrocarbons, and include β-carotene, α-carotene, and lycopene [56]. Xanthophylls are the oxygenated derivatives of carotenes, and include β-cryptoxanthin, lutein, zeaxanthin, meso-zeaxanthin, astaxanthin, and canthaxanthin [64]. Carotenoids play a critical role in various biological processes such as the immune response, prevention of cell propagation, induction of apoptosis, and suppression of several cancers [66,67]. Therefore, carotenoid deficiency can cause health problems in human beings.

2.3. Nitrogen-Containing Alkaloids and Sulfur-Containing Compounds

2.3.1. Alkaloids

Alkaloids are secondary metabolites containing a basic nitrogen, and are found primarily in plants [68]. The most common forms are derived from amino acids, whereas others originate from the modification of various classes of molecules such as polyphenols, terpenes, or steroids [14]. Alkaloids are produced by a large variety of organisms including bacteria, fungi, and animals [69]. Alkaloids have diverse biological functions, including anti-cancer, anti-microbial, anti-inflammatory, and antinociceptive properties [70]. Therefore, they play roles as protective agents against various diseases [70,71]. Individual plant species produce only a few kinds of alkaloids [68]. Certain plant species, such as Papaveraceae, Ranunculaceae, Solanaceae, and Amaryllidaceae, are particularly rich in alkaloids [68,72]. Although there is no uniform classification scheme for alkaloids, they can be generally divided into the following major groups: true alkaloids, protoalkaloids, and pseudoalkaloids [69,73]. True alkaloids (Figure 3a) are derived from amino acids, and have a nitrogen-atom-containing heterocyclic ring [74]. This group is further divided into 14 sub-groups according to the ring structure: pyrrolidine, pyrrolizidine, piperidine, tropone, quinoline, isoquinoline, acridine, quinolizidine, benzopyrrole, indolizidine, imidazole, purine, quinolizidine, and oxazole. The second group, protoalkaloids (Figure 3b), are derived from amino acids but do not contain a nitrogen-atom-bearing heterocyclic ring. These are less commonly found in nature in comparison with true alkaloids. The protoalkaloids include hordenine, mescaline, ephedrine, colchicine, erythromycin, jurubin, pachysandrine A, and taxol. Finally, although pseudoalkaloids (Figure 3c) are not derived from amino acids, they contain a nitrogen atom in a heterocyclic ring, and include subclasses such as terpene- and steroid-like alkaloids: delphinine, aconitine and solanidine [69,73].

2.3.2. Organosulfur Compounds

Organosulfur compounds (OSC) are sulfur-containing organic compounds (Figure 3d) [75]. Some essential amino acids and enzymes, sulfides, disulfides, and other OSCs are generated in the bodies of all living creatures and the natural environment [75,76]. OSCs can both maintain normal health in the human body and contribute to the development of disease by determining the thiol/disulfide redox states in body [75,77]. There are two major groups of vegetables that contain OSCs with special properties [75,76]. One is the Allium genus (family Amaryllidaceae), which produces S-alk(en)yl-l-cysteine sulfoxides, and includes plants such as garlic, onions, shallots, leeks, and chives. The second group includes members of the Brassica genus, including cabbage, cauliflower, Brussels sprouts, and kale and the members of the Eruca genus of the mustard or cruciferous family, which includes plants such as rucola; this group contains S-methyl cysteine-l-sulfoxide. The OSCs of vegetables from the Allium, Eruca, and Brassica genera include cycloalliin, thiosulfonates, cysteine alkyl disulfides, glucosinolates, goitrin, and epithionitrile [75,76]. There is an abundance of epidemiological and experimental evidence that indicates that OSCs have protective effects against several cancers, including breast cancer [75,76].

3. Anti-Tumor Activity of Plant Metabolites in Various Malignant Cancers

3.1. Colorectal Cancer

Colorectal cancer is the major cause of cancer-mediated death worldwide. Nutrients and food play an important role in the development of colorectal cancer, and eating mostly food of plant origin rather than red and processed meat is recommended for cancer prevention [6]. Secondary metabolites from potatoes have been found to inhibit the growth of colon cancer cells [78]. The maximum cancer cell growth inhibition was achieved when HT-29 colon cancer cells were exposed to extracts of potatoes with red-and purple-fleshed tubers. This indicates that some metabolites of potatoes with red and purple tubers could be valuable as a dietary intervention against developing the colon cancer [79]. Dichamanetin, a secondary metabolite from Piper sarmentosum, which is an edible herb used as a spice in Southeast Asia, was reported to reduce cell viability in HT-29 colon cancer cells [80]. This metabolite showed dose-dependent cytotoxic effects on this cancer cell type via the induction of ROS, and also arrested their cell cycle, suggesting that it could be used to block cancer cell proliferation [80]. Active oxyprenylated natural products from citrus fruits belonging to the Rutaceae family have been considered as interesting phytochemicals for several decades [81]. For example, 4′-geranyloxyferulic acid (GOFA) has been reported to have chemopreventive activity against cancer since it was first extracted in 1966 from Acronychia baueri Schott (Fam. Rutaceae) [82]. 3-(4′-Geranyloxy-3′;-methoxyphenyl)-l-alanyl-l-proline (GAP), a peptide prodrug of GOFA, was discovered to suppress colitis-related carcinogenesis in the colon in the azoxymethane (AOM)/dextran sodium sulfate (DSS)-induced cancer model in animals [83]. Similarly, GOFA/β-CD (the β-cyclodextrin inclusion compound of GOFA) inhibited the development of colonic carcinoma in the AOM/DSS model [84]. Auraptene (7-geranyloxycoumarin, AUR), one of the lead anti-cancer compounds from the Rutaceae family, was also found to inhibit the growth of both wild-type and chemo-resistant colon cancer cells and suppress the formation of colonospheres, suggesting that it could prevent the recurrence of cancer stem cells [85]. Curcumin is the main secondary metabolite derived from Curcuma longa and other Curcuma spp, and has been widely studied as a therapeutic agent having antiangiogenic, anti-inflammatory, and antioxidant activity [86]. Recently, curcumin was found to suppress the oncogenicity of human colon cancer cells by reducing the stability of SIRT1 (a NAD+ dependent histone/protein deacetylase) and to suppress the growth of HCT-116 tumor xenografts [87]. Genistein, a phenolic compound found in soybeans, is known to act as a chemopreventive agent against various tumors [88].
This had inhibitory effects on colorectal cancer cells HCT 116 and LoVo; it inhibited cell proliferation and induced apoptosis [89,90]. It also inhibited the invasion and migration of colorectal cancer cells and inhibited the metastasis of human colorectal cancer cells implanted in nude mice [91]. Combinatorial treatment of genistein and indole-3-carbinol synergistically induced apoptosis of HCT 116 cells [92]. Benzyl isothiocyanate (BITC), an organosulfur compound, suppressed the viability of HCT 116 cells and activated the PI3K/Akt/forkhead box O pathway, which influences drug resistance in various human cancer cells [93]. A combination treatment with an inhibitor of the PI3K/Akt/forkhead box O pathway potentiated cell death of colorectal cancer cells induced by BITC [93]. Sulforaphane, another isothiocyanate, has an anticancer effect on the human colon cancer cell line HT-29. It downregulates the expression of microsomal prostaglandin E synthase-1, which is involved in the synthesis of prostaglandin E2 known to be highly expressed in colorectal cancer [94]. Dietary phenethyl isothiocyanate (PEITC) improved adenocarcinoma in azoxymethane (AOM) and dextran induced colitis associated cancer mouse models [95]. Recently, 6-(methylsulfinyl)hexyl isothiocyanate (6-MSITC), obtained from Wasabia japonica, was found to induce apoptosis in human colorectal cancer cells ( HCT 116 p53+/+ and HCT 116 p53/ ) via p53-independent mitochondrial dysfunction [96].

3.2. Gastric Cancer

Gastric cancer, also known as stomach cancer, is one of the most common cancers, and has a poor prognosis [97]. Although many other factors contribute to gastric tumorigenesis, there is strong evidence that H. pylori infection is the predominant etiological factor in the induction of gastric cancer [98]. Many plant phytochemicals used as anti-gastric-cancer agents have been found to not only affect cancer cells directly but also to inhibit H. pylori. Resveratrol (3,4,5′-hydroxystilbene), a polyphenol flavonoid, is known to be produced by a limited number of plants (about 31 genera), and has the ability to inhibit H. pylori growth and the proliferation of gastric cancer cells [99]. Isothiocyanates (ITCs) are phytochemicals derived from cruciferous plants, including allyl isothiocyanate, sulforaphane (SFN), benzyl isothiocyanate (BITC), and phenethyl isothiocyanate (PEITC). ITCs have been reported not only to have bactericidal activity toward H. pylori and to reduce the colonization of H. pylori in the stomach, but also to have chemopreventive effects on gastric cancer in vitro and in vivo [100]. SFN was found to eradicate extracellular and intracellular H. pylori and block benzo[a]pyrene-induced stomach tumors in mice [101]. PEITC induced cell cycle arrest and apoptosis by disrupting microtubule filaments in MKN74 and Kato-III human gastric cancer cells [102]. Another group demonstrated that PEITC reduced the invasion and the migration of AGS human gastric cancer cells through blocking the mitogen-activated protein kinase (MAPK) signaling pathways that regulate the expression of matrix metalloproteinases (MMPs)-2 and -9 [103]. BITC was also found to inhibit the migration and invasion of AGS human gastric cancer cells in a dose-dependent manner [104]. In addition to colorectal cancer, curcumin has also been extensively investigated for its chemopreventive effects on gastric cancer. In an in vitro study, curcumin was shown to inhibit the proliferation of SGC-7901 human gastric cancer cells by facilitating the collapse of the mitochondrial membrane potential, and in an in vivo study, the growth of xenograft tumors was reduced by curcumin [105]. In addition, another in vivo study showed that curcumin reduced lymphatic vessel density (LVD) in gastric-tumor bearing nude mice [106]. Quercetin, a natural flavonoid present in various fruits, was reported to induce apoptosis in BGC-823 human gastric cancer cells [107]. Recently, a combined treatment with curcumin and quercetin was found to significantly inhibit proliferation and induce apoptosis in BGC-823 cells [108]. Allicin, an active compound derived from garlic, was found to have chemopreventive effects on gastric cancer by inhibiting cell growth, arresting the cell cycle, and inducing apoptosis [109].

3.3. Lung Cancer

Lung cancer is the most common cancer, and has the highest cancer-related mortality worldwide [110]. Several secondary metabolites have been discovered to have inhibitory activity against lung cancer. Epigallocatechin gallate (EGCG), a major component of green tea from Camellia sinensis, has been reported to have preventive effects on carcinogenesis [111]. There are several reports that EGCG can inhibit lung cancer in vitro. Recently, EGCG was shown to inhibit the growth of several types of human lung cancer cells via upregulating p53 expression, increasing p53 stability, and inhibiting p53 ubiquitination [112]. Another study indicated that EGCG was involved in increasing miR210, a major miRNA (micro RNA) regulated by HIF-1α, resulting in a significant reduction of the proliferation and growth of mouse and human lung cancer cells [113]. Liu et al. reported that EGCG inhibited not only TGF-β-induced cell migration and invasion but also TGF-β-induced epithelial-to-mesenchymal transition (EMT) via inhibition of the Smad2 and ERK1/2 signaling pathways in nonsmall cell lung cancer (NSCLC) cells [113]. EGCG has also been found to inhibit telomerase and induce apoptosis in both drug-sensitive and drug-resistant small cell lung cancer (SCLC) cells [114].
In addition to their activity against gastric cancer discussed above, ITCs have also been reported to have anti-lung cancer activity via various molecular mechanisms [111]. There are three different types of ITCs [115]: BITC, PEITC, and SFN. All three ITCs arrested the growth of human lung cancer A549 cells by binding to tubulin, with their relative activities following the order BITC > PEITC > SFN [115]. BITC inhibited the growth of NSCLC cells that are resistant to gefitinib, which is widely used in treatment of NSCLC, via cell cycle arrest and reactive oxygen species generation [116]. BITC was also reported to inhibit tumorigenesis of A/J mice induced by the polycyclic aromatic hydrocarbons (PAHs) found in cigarette smoke [117]. In addition, PEITC induced the apoptosis of NSCLC cells by inducing the disassembly of actin stress fibers and degradation of tubulin, resulting in the inhibition of NSCLC cell growth [118]. In another study, both BITC and PEITC were shown to induce the apoptosis of highly metastatic lung cancer L9981 cells by activating three mitogen-activated protein kinases (MAPKs): JNK, ERK1/2, and p38 [113]. Oral SFN treatment of mice with lung cancer induced by benzo(a)pyrene (B(a)P) was proved to rehabilitate carcinogenic lungs via decreasing H2O2 production and inducing apoptosis [119]. Combination treatment with SFN and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) induced apoptosis in A549 lung adenocarcinoma cells, which are resistant to the apoptotic effect of TRAIL, through downregulation of ERK and Akt [120].
Indole-3-carbinol (I3C) is a hydrolysis product of glucosinolate, which is a natural component in members of the Brassica family including broccoli, cabbage, cauliflower, and Brussels sprouts, and is known to have various anti-tumor activities [111]. I3C has lung cancer-preventive activity during the progression of tobacco carcinogen induced lung adenocarcinoma in mice and is involved in the modulation of apoptosis-related proteins in lung cancer A549 cells [121]. Choi et al. showed that I3C induced cell cycle arrest at the G0/G1 phase through increasing the expression of phosphorylated p53 and cyclin D1 and activated caspase-8 mediated apoptosis via increasing Fas mRNA in lung cancer A549 cells [122]. The anti-lung cancer activity of I3C in combination with silibinin, the major active constituent of Silybum marianum, is stronger than that of single treatment and avoids undesirable side effects in A549 and H460 lung cancer cells and in vivo 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK)-induced lung tumors [123]. The overexpression or underexpression of microRNAs (miRNAs), which function as tumor suppressors, during tumorigenesis has been studied. It has been reported that I3C can inhibit carcinogenesis by modulating the expression of several miRNAs in the vinyl carbamate (VC)-induced lung cancer model [124].
Genistein inhibits SCLC cell proliferation and migration and induces apoptosis in the SCLC cells H446 through downregulation of FoxM1, whose target genes regulate the cell cycle and apoptosis [125]. Several reports have also indicated that genistein has synergistic effects with other well-known anti-cancer drugs. The combination of genistein with gefitinib, a drug widely used in the treatment of various cancers, can inhibit cell proliferation and induce apoptosis in drug resistant H1975 NSCLC cells, which harbor an epidermal growth factor receptor (EGFR) mutation [126]. Another report showed that treating H460 lung cancer cells with a combination of genistein and the chemotherapeutic agents cisplatin, docetaxel, or doxorubicin inhibited cell growth and induced apoptosis with greater anti-cancer activity than single treatment alone. Furthermore, genistein can inhibit the induction of nuclear factor kappaB (NF-κB) activity by chemotherapeutic agents, which enables cancer cells to become drug resistant [127].
Fisetin (3,3′,4′,7-tetrahydroxyflavone) is a polyphenolic flavonoid found in many fruits and vegetables, and has been reported to possess anti-inflammatory, antiangiogenic, and anti-tumor activities [128]. It has dual inhibitory effects on phosphatidylinositol-3 kinase (PI3K)/Akt and the mammalian target rapamycin (mTOR) signaling in A549 human NSCLC cells and inhibits the cell viability and colony-forming ability of A549 cells [129]. Fisetin is also involved in inhibiting the invasion and migration of A549 NSCLC cells through the inactivation of the extracellular signal-regulated kinase (ERK) signaling pathway and reducing the expression of MMP-2 and urokinase-type plasminogen activator (u-PA) [130]. Orally administered fisetin inhibits lung carcinogenesis by alleviating mitochondrial dysfunction and inducing apoptosis in the B(a)P-induced lung cancer mouse model [131]. In another in vivo study, fisetin inhibited angiogenesis and tumor growth in Lewis lung carcinoma bearing mice, and the combination of fisetin with cyclophosphamide (CPA), a medication used as chemotherapy, showed markedly improved anti-tumor activity over fisetin or CPA alone without toxic side effects [132].
Punicalagin (PC) is an ellagitannin, a type of phenolic compound found in Punica granatum (pomegranate), which has been shown to exert antioxidant, anti-mutagenic, and anti-cancer activity [133]. PC has anti-mutagenic potential and shows dose-dependent anti-proliferative effects in A549 and H1299 human lung cancer cells [134]. Pomegranate fruit extracts (PFE) inhibit not only the growth and viability of A549 lung cancer cells in vitro but also the growth of A549 lung cancer cells in nude mice in vivo [135]. Additionally, PFE has been reported to inhibit tumorigenesis in the B(a)P-induced lung cancer mice model [136].
Curcumin has also been reported to have anti-cancer activity in both NSCLC and SCLC cell lines [111]. In NSCLC cells, curcumin inhibits cell growth and invasion by suppressing the expression of Metastasis-associated protein 1 (MTA1) and subsequently inactivating the Wnt/β-catenin pathway, which has a cooperative role in promoting lung tumorigenesis [137]. Curcumin downregulates the expression of Cdc42, which is known to be involved in the proliferation, metastasis, and invasion of cancer cells, resulting in inhibition of the invasion of lung cancer cells [126]. One of the underlying mechanisms for the inhibition of lung cancer cell growth by curcumin was the induction of autophagy via activating the AMP-activated protein kinase (AMPK) signaling pathway [138]. In addition, curcumin is involved in lowering the resistance of NSCLC cells against erlotinib, a drug used for NSCLC [139]. In SCLC cells, curcumin suppressed cell proliferation, migration, invasion, and angiogenesis through inhibiting the signal transducer and activator of transcription 3 (STAT3) and downregulating the expression of STAT3-regulated gene products (Cyclin B1, Bcl-XL, survivin, vascular endothelial growth factor, MMP-2, -7, and intercellular adhesion molecule-1) [140]. Curcumin-induced apoptosis was accompanied by mechanisms that increased the intracellular reactive oxygen species (ROS) level [141].

3.4. Breast Cancer

Breast cancer represents the most common and highest-mortality malignancy in females around the world [142]. Naturally occurring compounds have been studied for their chemopreventive effects on breast cancer. Tomatine is a glycoalkaloid secondary plant metabolite occurring in the Solanaceae family of plants that is known to have defensive activities against phytopathogens [143]. It can also induce cell cytotoxicity and apoptosis and decrease metastasis-related MMP-2, -9 activity in MCF-7 human breast cancer cells [144].
I3C shows effective anti-tumor properties in estrogen receptor α (ERα)-positive breast cancer cells through the ligand-activated aryl hydrocarbon receptor (AhR), which amplifies ERα signaling via ROS induction by the upregulation of cyclic-AMP-dependent transcription factor (ATF)-3 and downstream pro-apoptotic BH3-only proteins [145]. Also, I3C inhibits tumor sphere formation in breast cancer cells with stem/progenitor cell-like character by selectively stimulating the interaction of nucleostemin (a cancer stem/progenitor cell marker highly expressed in breast cancer stem cells) with MDM2 (an inhibitor of p53 tumor suppressor) [146].
Triterpenoids are secondary metabolites found in various plants, and are known to have antioxidant, anti-microbial, anti-allergic, and anti-angiogenic activity. Dozens of triterpenoids have been reported to have chemopreventive potential against breast cancer [147]. Curcubitane-type triterpenoids isolated from Cucurbitaceae family inhibit the growth of several types of human breast cancer cells [148,149,150], exhibit cytotoxicity against these cells [151,152], and induce apoptosis [153,154,155]. Dammarane triterpenoids isolated from the tropical plant Chisocheton penduliflorus exhibit weak cytotoxicity in breast cancer cells [156]. Two major friedelane triterpenoids, pristimerin and celastrol, have been found to be active against breast cancer cells. Pristimerin acts as a mitochondrial-targeting compound and induces caspase-mediated apoptosis and cytochrome c release in MDA-MB-231 breast cancer cells [157]. Celastrol has been shown to not only inhibit the growth and induce apoptosis of W256 rat breast cancer cells, but also suppress their migration by acting as an inhibitor of IκB kinase (IKK) [158]. Meliavolkenin, a limonoid triterpene isolated from Melia volkensii (Meliaceae), has cytotoxic effects on MCF-7 breast cancer cells [159]. Betulinic acid (BA), a pentacyclic triterpenoid, has anti-proliferative activity in MCF-7 and T47D breast cancer cells [160], in which a decrease in bcl2 and cyclin D1 gene expression and an increase in the bax gene were also observed [161]. In another study, most breast cancer cell lines (SKBR3, MDA231, MDL13E, BT483, BT474, T47D, and BT 549) except for MCF7 and ZR-75-1 cells were sensitive to BA treatment [162]. Lupeol, another natural pentacyclic triterpenoid, inhibits proliferation in estrogen receptor alpha (ERα)-negative MDA-MB-231 cells [163]. Ursolic acid, a pentacyclic triterpenoid widely found in the peels of fruits, has been studied as a potential inhibitor of breast tumors. Ursolic acid inhibits MCF-7 cell proliferation through arresting the cell cycle at G1 [164] and possesses cytotoxic activity against MCF-7 and MDA-MB-231 cells [165,166,167]. Additionally, ursolic acid is involved in inducing apoptosis through modulation of the glucocorticoid receptor (GR) and Activator protein 1 (AP1) in MCF-7 cells [168]. Yeh et al. observed that it has suppressing effects on migration and invasion through inactivation of c-Jun N-terminal kinase (JNK), Akt, and mTOR signaling in highly metastatic MDA-MB-231 breast cancer cells [169]. Another pentacyclic triterpenoid, asiatic acid, which is extracted from the tropical medicinal plant Centella asiatica, was found to inhibit cell growth by inducing S-G2/M phase cell cycle arrest and executing apoptosis through the activation of mitochondrial pathways in MCF-7 and MDA-MB-231 cells [170]. Remangilones A and C, which are oleanane triterpenoids isolated from Physena madagascariensis, exhibit cytotoxicity against two breast cancer cell lines, MDA-MB-231 and MDA-MB-435, and induce apoptosis [171]. Amooranin (AMR), a triterpene acid isolated from the tropical tree Amoora rohituca, was shown to have cytotoxicity against MCF-7 cells [172]. Also, in studies of the mechanism of AMR-related cell death, AMR was reported to induce apoptosis through elevating caspase activity in MCF-7 and multidrug resistant MCF-7/TH cells, to suppress cell growth by arresting the cell cycle, and to induce apoptosis by regulating Bcl-2 family proteins and caspases in MDA-468 and MCF-7 cells [173,174]. Tirucallane-type triterpenoids extracted from Amphipterygium adstringens had cytotoxic effects on MCF-7 cells [175]. A newly discovered triterpenoid, Ailanthus excelsa chloroform extract-1 (AECHL-1) from Ailanthus excelsa Roxb, was shown to regress tumor volume in nude mice injected with MCF-7 cells [176].
Recently, curcumin has been also studied as an inhibitor of breast cancer cell proliferation. It was found to prevent the proliferation of Bisphenol A (BPA) induced MCF-7 cells by suppressing BPA-upregulated expression of miRNA-19, a key oncogenic miRNA [177]. Resveratrol and resveratrol sulfates reduced the cell viability of breast cancer cells (MCF-7, ZR-75-1, and MDA-MB-231) [178]. Avicennia marina extracts, used in traditional medicine, were shown to induce apoptosis in breast cancer cells (AU565, MDA-MB-231, and BT483) and inhibit tumor growth in MDA-MB-231 transplanted nude mice [179]. Additionally, these extracts were found to be rich in polyphenols [179]. In a recent study, hydroxycinnamic acid and flavonol derivatives, present in Bursera copallifera, were shown to be involved in inhibiting the migration of MCF-7 and MDA-MB-231 cells [180]. As in the case of colorectal cancer, dichamanetin also reduced the cell viability of MDA-MB-231 cells [80].

3.5. Prostate Cancer

Prostate cancer is one of the most commonly diagnosed cancers in men worldwide. Diet and lifestyle are thought to be major contributors to prostate cancer development, and therefore, the ability of bioactive natural plant chemicals to inhibit prostate cancer has been widely studied [181]. Recently, decursinol, a metabolite of Angelica gigas, has been shown to decrease tumor growth in mice with xenografts of human DU145 and PC3 prostate cancer cells [182], and another group has reported that decursin and decursinol angelate (DA) from Angelica gigas Nakai (AGN) have inhibitory effects on the growth of prostate epithelium in the transgenic adenocarcinoma of mouse prostate (TRAMP) model [183]. Both resveratrol and γ-viniferin, a tetramer of resveratrol, inhibit the growth of LNCaP prostate cancer cells by arresting the cell cycle at the G1 phase; γ-viniferin has more potent growth-inhibiting activity than resveratrol [184]. Another plant polyphenol, fisetin, has been found to be involved in regulating microtubule stability through increasing the amount of acetylated α-tubulin and microtubule associated proteins (MAP)-2 and 4 in PU3 and DU145 cells and downregulating nuclear migration protein (NudC), which plays an essential role in mitosis and cytokinesis [185]. Prostate-cancer-associated mortality is mainly caused by metastasis. Therefore, it is important to develop anti-cancer compounds to inhibit its metastasis. Genistein was found to act as an anti-metastatic agent to inhibit cellular invasion in prostate cancer cells through decreasing MMP expression and decreasing the formation of metastases in mice implanted with the PC3-M human prostate cancer cell line [186]. Curcumin was discovered to inhibit cancer-associated fibroblast (CAF)-induced EMT and invasion in PC3 cells by suppressing the monoamine oxidase A (MAOA)/mTOR/HIF-1α signaling pathway [187]. Additionally, it has anti-cancer effects through the inhibition of prostate cancer cell growth and metastasis [188,189]. Both SFN and I3C attenuate Akt/NKκB signaling and induce growth arrest and apoptosis in prostate cancer [181].

3.6. Hematologic Cancer

Hematologic cancer, also called blood cancer, develops in blood-forming tissue or in immune-system-related cells and includes leukemia, myeloma, and lymphoma [190]. Its overall prognosis is poor despite extensive research into cytotoxic agents to combat it. Recently, hypericin, a secondary metabolite from Hypericum (Saint John’s wort), was discovered to potentiate the mitoxantrone (MTX)-induced death of the HL-60 subclone human leukemia cells, in which the ABC transporter is overexpressed [191]. The anti-cancer mechanisms of the natural polyphenol resveratrol have been widely studied. Azmi et al. showed that resveratrol induces DNA breakage in the presence of copper in human peripheral lymphocytes, suggesting a novel anti-cancer mechanism involving the mobilization of endogenous copper, which is known to be increased in various malignancies [192]. Another group discovered that resveratrol inhibited cell proliferation, arrested the cell cycle in the S-phase, and induced apoptosis in the acute myeloid leukemia cells OCI-ANK3 and OCIM2 [193]. Similarly to in lung cancer, EGCG also induced the apoptotic death of the human B lymphoblastoid cell line (Ramos cells) in a dose- and time-dependent manner [194]. In addition, I3C was found to have anticancer properties in B cell precursor acute lymphoblastic leukemia in NALM-6 cells. It caused the arrest of the G1 phase in cell cycle and triggered apoptosis [195].

3.7. Skin Cancer

Skin cancer is one of the tumors causing malignancies around the world, and its incidence is increasing alarmingly [196]. Skin cancer is believed to develop through co-carcinogenic effects, and many natural metabolites have been widely studied as anti-carcinogens. In particular, allyl sulfides including diallyl sulfide (DAS), diallyl disulfide (DADS), and diallyl trisulfide (DATS) have been reported to prevent the progression of skin cancer [197]. DAS was found to have anti-mutagenic properties against 7,12-dimethylbenz[a]anthracene (DMBA), a carcinogenic polycyclic aromatic hydrocarbon that induces DNA strand breaks in mouse skin [198]. Additionally, DAS induced apoptosis in DMBA-mediated mouse skin tumors through multiple mechanisms, including the up-regulation of tumor suppressor protein p53, its downstream proteins, and proapoptotic proteins such as Bax, and the reduction of Ras onco-protein expression [199,200]. Pomolic acid, a triterpenoid found in Polylepis racemosa, was reported to have cytotoxic effects on M-14 melanoma cells [167].

3.8. Head and Neck Cancer

Head and neck cancer is one of the leading causes of death worldwide [201]. Current medical and surgical treatments for these malignancies result in functional morbidity and side effects; thus, chemopreventive phytochemicals have been widely studied [201]. β-Carotene is one of the most abundant carotenoids, which are natural pigments found in plants and that are well known to be effective antioxidants [202]. Recently, β-carotene has been reported to enhance the inhibitory effect of 5-FU, a medication used against cancer, on tumor growth of xenografts of Eca109 esophageal squamous cell carcinoma (ESCC) cells in nude mice and to inhibit cell proliferation in the ESCC cells EC1 and Eca109 [203]. EGCG has been found to have cytotoxic effects via arrest of the cell cycle at G1 and the induction of apoptosis in the human head and neck squamous cell carcinoma (HNSCC) cell lines YCU-N861 and YCU-H891 [204]. It has also been reported to synergistically inhibit the growth of HNSCC cells via inhibition of the NF-κB signaling pathway when used in conjunction with erlotinib, a tyrosine kinase inhibitor of EGFR, which is frequently overexpressed in HNSCC cells [205]. In addition, EGCG was shown to inhibit the invasion and migration of the human oral cancer cell line OC2 through decreasing MMP-2, -9, and uPA in a dose dependent manner without cytotoxicity [206].

4. Conclusions and Perspectives

Up to the present date, several thousands of different metabolites have been identified in plants and studied for their effectiveness in a wide variety of applications [6]. We have categorized plant-derived metabolites into several major classes based on their structure, and the structural characteristics of each class were discussed. Also, natural compounds with anti-cancer activity were summarized according to type of cancer (Table 1 and Table 2). Medicinal plants have been used since ancient times, and are still used as a primary source of medical treatment in developing countries [3]. Plant-derived substances have advantages including their low cost and the rapid speed of discovery of new drugs; their main disadvantage is the absence of common international standards for evaluating their quality, efficacy, and safety [3]. Additionally, the incidence of various malignant cancers has been growing, and conventional cancer therapies have limitations, including the high toxicity and side effects of anti-cancer drugs [3]. For this reason, a broad multidisciplinary research approach involving ethnopharmacology, botany, pharmacognosy, and phytochemistry is required for the successful application of phytochemicals in the treatment or prevention of cancer [207]. Also, the biotechnological production of secondary metabolites of naturally occurring plant substances and the combination of phytochemicals with existing anti-cancer drugs or other chemical compounds represent alternative approaches to natural-product-based drug development. Furthermore, besides the cytotoxic effects of plant metabolites, additional therapies that treat cancers by different mechanisms are required for the development of new drugs from plant metabolites. One of the new cancer treatment method focuses on the immunomodulation of the tumor microenvironment. Therefore, the development of natural-product-based drugs that can regulate the functioning of the immune system in the tumor microenvironment will be a novel cancer treatment option in the future. This review provides comprehensive information on the various classes of plant-derived metabolites and bioactive plant compounds that have shown anti-cancer activity in vitro or in vivo models of different types of cancer. The data we have summarized clearly suggests that natural metabolites from plants play a major role as the most prominent source of anti-cancer treatments.

Author Contributions

S.-A.S., S.Y.M., H.H.P., and C.S.L. wrote the manuscript with guidance from H.H.P. and C.S.L., W.-Y.K., and S.-M.P. provided intellectual contributions in this study. S.-A.S. and S.Y.M. contributed equally to the work.

Funding

This research was supported by a grant from Next-Generation BioGreen21 Program (PJ01327302), Rural Development Administration, Korea.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

OSCOrganosulfur compounds
GOFA 4′-geranyloxyferulic acid
GAP3-(4′-geranyloxy-3′;-methoxyphenyl)-l-alanyl-l-proline
AOMAzoxymethane
DSSDextran sodium sulfate
β-CDβ-cyclodextrin
AURAuraptene
SIRT1Sirtuin 1
ITCsIsothiocyanates
SFNSulforaphane
BITCBenzyl isothiocyanate
PEITCPhenethyl isothiocyanate
MMPsMatrix metalloproteinases
LVDLymphatic vessel density
EGCGEpigallocatechin gallate
miRNAmicro RNA
HIF-1αHypoxia-inducible factor 1alpha
TGF-βTransforming growth factor β
EMTEpithelial-to-mesenchymal transition
ERK1/2Extracellular signal–regulated kinases 1/2
NSCLCNonsmall cell lung cancer
SCLCSmall cell lung cancer
MAPKMitogen-activated protein kinase
B(a)PBenzo(a)pyrene
TRAILTumor necrosis factor-related apoptosis-inducing ligand
I3CIndole-3-carbinol
VCVinyl carbamate
EGFREpidermal growth factor receptor
NF-κBNuclear factor kappaB
mTORmammalian target of rapamycin
u-PAurokinase-type plasminogen activator
CPACyclophosphamide
PCPunicalagin
PFEPomegranate fruit extracts
MTA1Metastasis-associated protein1
AMPKAMP-activated protein kinase
STAT3Signal transducer and activator of transcription 3
ROSReactive oxygen species
ERαEstrogen receptor α
AhRAryl hydrocarbon receptor
ATF-3cyclic AMP dependent transcription factor
IKKInhibitor against IκB kinase
BABetulinic acid
GRGlucocorticoid receptor
AP1Activator protein 1
JNKc-Jun N-terminal Kinase
AMRAmooranin
AECHL-1Ailanthus excelsa chloroform extract-1
BPABisphenol A
miRNA-19microRNA-19
AGNAngelica gigas Nakai
TRAMPTransgenic adenocarcinoma of mouse prostate
MAPMicrotubule associated proteins
NudCNuclear migration protein
CAFCancer associated fibroblast
MAOAMonoamine oxidase A
MTXMitoxantrone
DASDiallyl sulfide
DADSDiallyl disulfide
DATSDiallyl trisulfide
DMBA7,12-dimethylbenz[a]anthracene
ESCCEsophageal squamous cell carcinoma
HNSCCHead and neck squamous cell carcinoma

References

  1. Lee, C.S.; Baek, J.; Han, S.Y. The role of kinase modulators in cellular senescence for use in cancer treatment. Molecules 2017, 22, 1411. [Google Scholar] [CrossRef] [PubMed]
  2. Amin, A.R.; Kucuk, O.; Khuri, F.R.; Shin, D.M. Perspectives for cancer prevention with natural compounds. J. Clin. Oncol. 2009, 27, 2712–2725. [Google Scholar] [CrossRef] [PubMed]
  3. Fridlender, M.; Kapulnik, Y.; Koltai, H. Plant derived substances with anti-cancer activity: From folklore to practice. Front. Plant Sci. 2015, 6, 799. [Google Scholar] [CrossRef] [PubMed]
  4. Gali-Muhtasib, H.; Hmadi, R.; Kareh, M.; Tohme, R.; Darwiche, N. Cell death mechanisms of plant-derived anticancer drugs: Beyond apoptosis. Apoptosis 2015, 20, 1531–1562. [Google Scholar] [CrossRef] [PubMed]
  5. Morrissey, K.M.; Yuraszeck, T.M.; Li, C.C.; Zhang, Y.; Kasichayanula, S. Immunotherapy and novel combinations in oncology: Current landscape, challenges, and opportunities. Clin. Transl. Sci. 2016, 9, 89–104. [Google Scholar] [CrossRef] [PubMed]
  6. Korkina, L.; Kostyuk, V. Biotechnologically produced secondary plant metabolites for cancer treatment and prevention. Curr. Pharm. Biotechnol. 2012, 13, 265–275. [Google Scholar] [CrossRef] [PubMed]
  7. Schmidt, B.M.; Ribnicky, D.M.; Lipsky, P.E.; Raskin, I. Revisiting the ancient concept of botanical therapeutics. Nat. Chem. Biol. 2007, 3, 360–366. [Google Scholar] [CrossRef] [PubMed]
  8. Gordaliza, M. Natural products as leads to anticancer drugs. Clin. Transl. Oncol. 2007, 9, 767–776. [Google Scholar] [CrossRef] [PubMed]
  9. Ijaz, S.; Akhtar, N.; Khan, M.S.; Hameed, A.; Irfan, M.; Arshad, M.A.; Ali, S.; Asrar, M. Plant derived anticancer agents: A green approach towards skin cancers. Biomed. Pharmacother. 2018, 103, 1643–1651. [Google Scholar] [CrossRef] [PubMed]
  10. Molyneux, R.J.; Lee, S.T.; Gardner, D.R.; Panter, K.E.; James, L.F. Phytochemicals: The good, the bad and the ugly? Phytochemistry 2007, 68, 2973–2985. [Google Scholar] [CrossRef] [PubMed]
  11. Santhi, K.S.; Sengottuvel, R. Qualitative and quantitative phytochemical analysis of moringa concanensis nimmo. Int. J. Curr. Microbiol. App. Sci. 2016, 5, 633–640. [Google Scholar] [CrossRef]
  12. Pichersky, E.; Gang, D.R. Genetics and biochemistry of secondary metabolites in plants: An evolutionary perspective. Trends. Plant. Sci. 2000, 5, 439–445. [Google Scholar] [CrossRef]
  13. Baxter, H.; Harborne, J.B.; Moss, G.P. Phytochemical Dictionary: A Handbook of Bioactive Compounds from Plants; CRC Press: New York, NY, USA, 1998. [Google Scholar]
  14. Kumar, A.; Irchhaiya, R.; Yadav, A.; Gupta, N.; Kumar, S.; Gupta, N.; Kumar, S.; Yadav, V.; Prakash, A.; Gurjar, H. Metabolites in plants and its classification. World J. Pharm. Pharm. Sci. 2015, 4, 287–305. [Google Scholar]
  15. Mocanu, M.M.; Nagy, P.; Szollosi, J. Chemoprevention of breast cancer by dietary polyphenols. Molecules 2015, 20, 22578–22620. [Google Scholar] [CrossRef] [PubMed]
  16. Varoni, E.M.; Lodi, G.; Sardella, A.; Carrassi, A.; Iriti, M. Plant polyphenols and oral health: Old phytochemicals for new fields. Curr. Med. Chem. 2012, 19, 1706–1720. [Google Scholar] [CrossRef] [PubMed]
  17. Crozier, A.; Jaganath, I.B.; Clifford, M.N. Dietary phenolics: Chemistry, bioavailability and effects on health. Nat. Prod. Rep. 2009, 26, 1001–1043. [Google Scholar] [CrossRef] [PubMed]
  18. Crozier, A.; Clifford, M.N.; Ashihara, H. Plant Secondary Metabolites: Occurrence, Structure and Role in the Human Diet; John Wiley & Sons: Oxford, UK, 2008. [Google Scholar]
  19. Ravishankar, D.; Rajora, A.K.; Greco, F.; Osborn, H.M. Flavonoids as prospective compounds for anti-cancer therapy. Int. J. Biochem. Cell Biol. 2013, 45, 2821–2831. [Google Scholar] [CrossRef] [PubMed]
  20. Beecher, G.R. Overview of dietary flavonoids: Nomenclature, occurrence and intake. J Nutr 2003, 133, 3248S–3254S. [Google Scholar] [CrossRef] [PubMed]
  21. Xiao, J.; Cao, H.; Wang, Y.; Zhao, J.; Wei, X. Glycosylation of dietary flavonoids decreases the affinities for plasma protein. J. Agric. Food Chem. 2009, 57, 6642–6648. [Google Scholar] [CrossRef] [PubMed]
  22. Leo, C.H.; Woodman, O.L. Flavonols in the prevention of diabetes-induced vascular dysfunction. J. Cardiovasc. Pharmacol. 2015, 65, 532–544. [Google Scholar] [CrossRef] [PubMed]
  23. Fantini, M.; Benvenuto, M.; Masuelli, L.; Frajese, G.V.; Tresoldi, I.; Modesti, A.; Bei, R. In vitro and in vivo antitumoral effects of combinations of polyphenols, or polyphenols and anticancer drugs: Perspectives on cancer treatment. Int. J. Mol. Sci. 2015, 16, 9236–9282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Manach, C.; Scalbert, A.; Morand, C.; Remesy, C.; Jimenez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef] [PubMed]
  25. Lea, M.A. Flavonol regulation in tumor cells. J. Cell Biochem. 2015, 116, 1190–1194. [Google Scholar] [CrossRef] [PubMed]
  26. Singh, M.; Kaur, M.; Silakari, O. Flavones: An important scaffold for medicinal chemistry. Eur. J. Med. Chem. 2014, 84, 206–239. [Google Scholar] [CrossRef] [PubMed]
  27. Sies, H.; Hollman, P.C.; Grune, T.; Stahl, W.; Biesalski, H.K.; Williamson, G. Protection by flavanol-rich foods against vascular dysfunction and oxidative damage: 27th hohenheim consensus conference. Adv. Nutr. 2012, 3, 217–221. [Google Scholar] [CrossRef] [PubMed]
  28. Mena, P.; Dominguez-Perles, R.; Girones-Vilaplana, A.; Baenas, N.; Garcia-Viguera, C.; Villano, D. Flavan-3-ols, anthocyanins, and inflammation. IUBMB Life 2014, 66, 745–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Quideau, S.; Deffieux, D.; Douat-Casassus, C.; Pouysegu, L. Plant polyphenols: Chemical properties, biological activities, and synthesis. Angew. Chem. Int. Ed. Engl. 2011, 50, 586–621. [Google Scholar] [CrossRef] [PubMed]
  30. King, A.; Young, G. Characteristics and occurrence of phenolic phytochemicals. J. Am. Diet. Assoc. 1999, 99, 213–218. [Google Scholar] [CrossRef]
  31. Aron, P.M.; Kennedy, J.A. Flavan-3-ols: Nature, occurrence and biological activity. Mol. Nutr. Food Res. 2008, 52, 79–104. [Google Scholar] [CrossRef] [PubMed]
  32. Kuhnle, G.G.C. Nutrition epidemiology of flavan-3-ols: The known unknowns. Mol. Aspects Med. 2018, 61, 2–11. [Google Scholar] [CrossRef] [PubMed]
  33. Ko, K.P. Isoflavones: Chemistry, analysis, functions and effects on health and cancer. Asian Pac. J. Cancer Prev. 2014, 15, 7001–7010. [Google Scholar] [CrossRef] [PubMed]
  34. Bircsak, K.M.; Aleksunes, L.M. Interaction of isoflavones with the bcrp/abcg2 drug transporter. Curr. Drug. Metab. 2015, 16, 124–140. [Google Scholar] [CrossRef] [PubMed]
  35. Vitale, D.C.; Piazza, C.; Melilli, B.; Drago, F.; Salomone, S. Isoflavones: Estrogenic activity, biological effect and bioavailability. Eur. J. Drug Metab. Pharmacokinet. 2013, 38, 15–25. [Google Scholar] [CrossRef] [PubMed]
  36. Mahmoud, A.M.; Yang, W.; Bosland, M.C. Soy isoflavones and prostate cancer: A review of molecular mechanisms. J. Steroid Biochem. Mol. Biol. 2014, 140, 116–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Tomás-Barberán, F.A.; Clifford, M.N. Flavanones, chalcones and dihydrochalcones–nature, occurrence and dietary burden. J. Sci. Food Agric. 2000, 80, 1073–1080. [Google Scholar] [CrossRef]
  38. Chanet, A.; Milenkovic, D.; Manach, C.; Mazur, A.; Morand, C. Citrus flavanones: What is their role in cardiovascular protection? J. Agric. Food. Chem. 2012, 60, 8809–8822. [Google Scholar] [CrossRef] [PubMed]
  39. Fang, J. Bioavailability of anthocyanins. Drug Metab. Rev. 2014, 46, 508–520. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, L.S.; Stoner, G.D. Anthocyanins and their role in cancer prevention. Cancer Lett. 2008, 269, 281–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Wallace, T.C.; Giusti, M.M. Anthocyanins. Adv. Nutr. 2015, 6, 620–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Marzocchella, L.; Fantini, M.; Benvenuto, M.; Masuelli, L.; Tresoldi, I.; Modesti, A.; Bei, R. Dietary flavonoids: Molecular mechanisms of action as anti- inflammatory agents. Recent Pat. Inflamm. Allergy Drug Discov. 2011, 5, 200–220. [Google Scholar] [CrossRef] [PubMed]
  43. Bellik, Y.; Boukraa, L.; Alzahrani, H.A.; Bakhotmah, B.A.; Abdellah, F.; Hammoudi, S.M.; Iguer-Ouada, M. Molecular mechanism underlying anti-inflammatory and anti-allergic activities of phytochemicals: An update. Molecules 2012, 18, 322–353. [Google Scholar] [CrossRef] [PubMed]
  44. Shen, T.; Wang, X.N.; Lou, H.X. Natural stilbenes: An overview. Nat. Prod. Rep. 2009, 26, 916–935. [Google Scholar] [CrossRef] [PubMed]
  45. Riviere, C.; Pawlus, A.D.; Merillon, J.M. Natural stilbenoids: Distribution in the plant kingdom and chemotaxonomic interest in vitaceae. Nat. Prod. Rep. 2012, 29, 1317–1333. [Google Scholar] [CrossRef] [PubMed]
  46. Sirerol, J.A.; Rodriguez, M.L.; Mena, S.; Asensi, M.A.; Estrela, J.M.; Ortega, A.L. Role of natural stilbenes in the prevention of cancer. Oxid. Med. Cell Longev. 2016, 2016, 3128951. [Google Scholar] [CrossRef] [PubMed]
  47. De Filippis, B.; Ammazzalorso, A.; Fantacuzzi, M.; Giampietro, L.; Maccallini, C.; Amoroso, R. Anticancer activity of stilbene-based derivatives. ChemMedChem. 2017, 12, 558–570. [Google Scholar] [CrossRef] [PubMed]
  48. Kundu, J.K.; Surh, Y.J. Cancer chemopreventive and therapeutic potential of resveratrol: Mechanistic perspectives. Cancer Lett. 2008, 269, 243–261. [Google Scholar] [CrossRef] [PubMed]
  49. Bishayee, A. Cancer prevention and treatment with resveratrol: From rodent studies to clinical trials. Cancer Prev. Res. (Phila) 2009, 2, 409–418. [Google Scholar] [CrossRef] [PubMed]
  50. Jang, M.; Cai, L.; Udeani, G.O.; Slowing, K.V.; Thomas, C.F.; Beecher, C.W.; Fong, H.H.; Farnsworth, N.R.; Kinghorn, A.D.; Mehta, R.G.; et al. Cancer chemopreventive activity of resveratrol, a natural product derived from grapes. Science 1997, 275, 218–220. [Google Scholar] [CrossRef] [PubMed]
  51. Heleno, S.A.; Martins, A.; Queiroz, M.J.; Ferreira, I.C. Bioactivity of phenolic acids: Metabolites versus parent compounds: A review. Food Chem. 2015, 173, 501–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Andjelković, M.; Van Camp, J.; De Meulenaer, B.; Depaemelaere, G.; Socaciu, C.; Verloo, M.; Verhe, R. Iron-chelation properties of phenolic acids bearing catechol and galloyl groups. Food Chem. 2006, 98, 23–31. [Google Scholar] [CrossRef]
  53. Kaushik, P.; Andujar, I.; Vilanova, S.; Plazas, M.; Gramazio, P.; Herraiz, F.J.; Brar, N.S.; Prohens, J. Breeding vegetables with increased content in bioactive phenolic acids. Molecules 2015, 20, 18464–18481. [Google Scholar] [CrossRef] [PubMed]
  54. Mahmoud, S.S.; Croteau, R.B. Strategies for transgenic manipulation of monoterpene biosynthesis in plants. Trends Plant Sci. 2002, 7, 366–373. [Google Scholar] [CrossRef]
  55. Pichersky, E.; Raguso, R.A. Why do plants produce so many terpenoid compounds? New Phytol. 2016. [Google Scholar] [CrossRef] [PubMed]
  56. McCreath, S.B.; Delgoda, R. Pharmacognosy: Fundamentals, Applications and Strategies; Academic Press: London, UK, 2017. [Google Scholar]
  57. Bruckingham, J. Dictionary of Natural Products on Cd-Rom; Champman and Hall: New York, NY, USA, 2000. [Google Scholar]
  58. Harborne, J.B. hytochemical Methods: A Guide to Modern Techniques of Plant Analysis; Chapman and Hall: London, UK, 1980. [Google Scholar]
  59. Bhatti, H.N.; Khan, S.S.; Khan, A.; Rani, M.; Ahmad, V.U.; Choudhary, M.I. Biotransformation of monoterpenoids and their antimicrobial activities. Phytomedicine 2014, 21, 1597–1626. [Google Scholar] [CrossRef] [PubMed]
  60. Rao, A.V.; Ray, M.R.; Rao, L.G. Lycopene. Adv. Food Nutr. Res. 2006, 51, 99–164. [Google Scholar] [PubMed]
  61. Akaberi, M.; Mehri, S.; Iranshahi, M. Multiple pro-apoptotic targets of abietane diterpenoids from salvia species. Fitoterapia 2015, 100, 118–132. [Google Scholar] [CrossRef] [PubMed]
  62. Soares Nda, C.; Teodoro, A.J.; Lotsch, P.F.; Granjeiro, J.M.; Borojevic, R. Anticancer properties of carotenoids in prostate cancer. A review. Histol. Histopathol. 2015, 30, 1143–1154. [Google Scholar] [PubMed]
  63. Stahl, W.; Sies, H. Separation of geometrical isomers of beta-carotene and lycopene. Methods Enzymol. 1994, 234, 388–400. [Google Scholar] [PubMed]
  64. Bendich, A.; Olson, J.A. Biological actions of carotenoids. FASEB. J. 1989, 3, 1927–1932. [Google Scholar] [CrossRef] [PubMed]
  65. Johnson, E.J. The role of carotenoids in human health. Nutr. Clin. Care 2002, 5, 56–65. [Google Scholar] [CrossRef] [PubMed]
  66. Chew, B.P.; Park, J.S. Carotenoid action on the immune response. J. Nutr. 2004, 134, 257S–261S. [Google Scholar] [CrossRef] [PubMed]
  67. Bolhassani, A. Cancer chemoprevention by natural carotenoids as an efficient strategy. Anticancer Agents Med. Chem. 2015, 15, 1026–1031. [Google Scholar] [CrossRef] [PubMed]
  68. Ng, Y.P.; Or, T.C.; Ip, N.Y. Plant alkaloids as drug leads for alzheimer’s disease. Neurochem. Int. 2015, 89, 260–270. [Google Scholar] [CrossRef] [PubMed]
  69. Evans, W.C. Trease and Evans’ Pharmacognosy E-Book; Elsevier Health Sciences: New York, NY, USA, 2009. [Google Scholar]
  70. Jiang, Q.W.; Chen, M.W.; Cheng, K.J.; Yu, P.Z.; Wei, X.; Shi, Z. Therapeutic potential of steroidal alkaloids in cancer and other diseases. Med. Res. Rev. 2016, 36, 119–143. [Google Scholar] [CrossRef] [PubMed]
  71. Cushnie, T.P.; Cushnie, B.; Lamb, A.J. Alkaloids: An overview of their antibacterial, antibiotic-enhancing and antivirulence activities. Int. J. Antimicrob. Agents 2014, 44, 377–386. [Google Scholar] [CrossRef] [PubMed]
  72. Wilkinson, A.; McNaught, A. Iupac Compendium of Chemical Terminology, (the “Gold Book”); International Union of Pure and Applied Chemistry: Zürich, Switzerland, 1997. [Google Scholar]
  73. Hesse, M. Alkaloids: Nature’s Curse or Blessing? John Wiley & Sons: Zurich, Switzerland, 2002. [Google Scholar]
  74. Cooper, R.; Nicola, G. Natural Products Chemistry: Sources, Separations and Structures; CRC Press: New York, NY, USA, 2014. [Google Scholar]
  75. Gupta, R.C. Nutraceuticals: Efficacy, Safety and Toxicity; Academic Press: London, UK, 2016. [Google Scholar]
  76. Moriarty, R.M.; Naithani, R.; Surve, B. Organosulfur compounds in cancer chemoprevention. Mini Rev. Med. Chem. 2007, 7, 827–838. [Google Scholar] [CrossRef] [PubMed]
  77. de Figueiredo, S.M.; Binda, N.S.; Nogueira-Machado, J.A.; Vieira-Filho, S.A.; Caligiorne, R.B. The antioxidant properties of organosulfur compounds (sulforaphane). Recent. Pat. Endocr. Metab. Immune. Drug Discov. 2015, 9, 24–39. [Google Scholar] [CrossRef] [PubMed]
  78. Reddivari, L.; Vanamala, J.; Chintharlapalli, S.; Safe, S.H.; Miller, J.C., Jr. Anthocyanin fraction from potato extracts is cytotoxic to prostate cancer cells through activation of caspase-dependent and caspase-independent pathways. Carcinogenesis 2007, 28, 2227–2235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Zuber, T.; Holm, D.; Byrne, P.; Ducreux, L.; Taylor, M.; Kaiser, M.; Stushnoff, C. Optimization of in vitro inhibition of ht-29 colon cancer cell cultures by solanum tuberosum l. Extracts. Food Funct. 2015, 6, 72–83. [Google Scholar] [CrossRef] [PubMed]
  80. Yong, Y.; Matthew, S.; Wittwer, J.; Pan, L.; Shen, Q.; Kinghorn, A.D.; Swanson, S.M.; De Blanco, E.J. Dichamanetin inhibits cancer cell growth by affecting ros-related signaling components through mitochondrial-mediated apoptosis. Anticancer Res. 2013, 33, 5349–5355. [Google Scholar] [PubMed]
  81. Genovese, S.; Fiorito, S.; Locatelli, M.; Carlucci, G.; Epifano, F. Analysis of biologically active oxyprenylated ferulic acid derivatives in citrus fruits. Plant Foods Hum. Nutr. 2014, 69, 255–260. [Google Scholar] [CrossRef] [PubMed]
  82. Genovese, S.; Epifano, F. Recent developments in the pharmacological properties of 4′-geranyloxyferulic acid, a colon cancer chemopreventive agent of natural origin. Curr. Drug Targets 2012, 13, 1083–1088. [Google Scholar] [CrossRef] [PubMed]
  83. Miyamoto, S.; Epifano, F.; Curini, M.; Genovese, S.; Kim, M.; Ishigamori-Suzuki, R.; Yasui, Y.; Sugie, S.; Tanaka, T. A novel prodrug of 4′-geranyloxy-ferulic acid suppresses colitis-related colon carcinogenesis in mice. Nutr. Cancer 2008, 60, 675–684. [Google Scholar] [CrossRef] [PubMed]
  84. Tanaka, T.; de Azevedo, M.B.; Duran, N.; Alderete, J.B.; Epifano, F.; Genovese, S.; Tanaka, M.; Tanaka, T.; Curini, M. Colorectal cancer chemoprevention by 2 beta-cyclodextrin inclusion compounds of auraptene and 4′-geranyloxyferulic acid. Int. J. Cancer 2010, 126, 830–840. [Google Scholar] [CrossRef] [PubMed]
  85. Epifano, F.; Genovese, S.; Miller, R.; Majumdar, A.P. Auraptene and its effects on the re-emergence of colon cancer stem cells. Phytother Res. 2013, 27, 784–786. [Google Scholar] [CrossRef] [PubMed]
  86. Lestari, M.L.; Indrayanto, G. Curcumin. Profiles Drug Subst. Excip. Relat. Methodol. 2014, 39, 113–204. [Google Scholar] [PubMed]
  87. Lee, Y.H.; Song, N.Y.; Suh, J.; Kim, D.H.; Kim, W.; Ann, J.; Lee, J.; Baek, J.H.; Na, H.K.; Surh, Y.J. Curcumin suppresses oncogenicity of human colon cancer cells by covalently modifying the cysteine 67 residue of sirt1. Cancer Lett. 2018, 431, 219–229. [Google Scholar] [CrossRef] [PubMed]
  88. Spagnuolo, C.; Russo, G.L.; Orhan, I.E.; Habtemariam, S.; Daglia, M.; Sureda, A.; Nabavi, S.F.; Devi, K.P.; Loizzo, M.R.; Tundis, R.; et al. Genistein and cancer: Current status, challenges, and future directions. Adv. Nutr. 2015, 6, 408–419. [Google Scholar] [CrossRef] [PubMed]
  89. Qin, J.; Teng, J.; Zhu, Z.; Chen, J.; Huang, W.J. Genistein induces activation of the mitochondrial apoptosis pathway by inhibiting phosphorylation of akt in colorectal cancer cells. Pharm. Biol. 2016, 54, 74–79. [Google Scholar] [CrossRef] [PubMed]
  90. Qin, J.; Chen, J.X.; Zhu, Z.; Teng, J.A. Genistein inhibits human colorectal cancer growth and suppresses mir-95, akt and sgk1. Cell Physiol. Biochem. 2015, 35, 2069–2077. [Google Scholar] [CrossRef] [PubMed]
  91. Xiao, X.; Liu, Z.; Wang, R.; Wang, J.; Zhang, S.; Cai, X.; Wu, K.; Bergan, R.C.; Xu, L.; Fan, D. Genistein suppresses flt4 and inhibits human colorectal cancer metastasis. Oncotarget 2015, 6, 3225–3239. [Google Scholar] [CrossRef] [PubMed]
  92. Nakamura, Y.; Yogosawa, S.; Izutani, Y.; Watanabe, H.; Otsuji, E.; Sakai, T. A combination of indol-3-carbinol and genistein synergistically induces apoptosis in human colon cancer ht-29 cells by inhibiting akt phosphorylation and progression of autophagy. Mol. Cancer 2009, 8, 100. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, X.; Takano, C.; Shimizu, T.; Yokobe, S.; Abe-Kanoh, N.; Zhu, B.; Nakamura, T.; Munemasa, S.; Murata, Y.; Nakamura, Y. Inhibition of phosphatidylinositide 3-kinase ameliorates antiproliferation by benzyl isothiocyanate in human colon cancer cells. Biochem. Biophys. Res. Commun. 2017, 491, 209–216. [Google Scholar] [CrossRef] [PubMed]
  94. Tafakh, M.S.; Saidijam, M.; Ranjbarnejad, T.; Malih, S.; Mirzamohammadi, S.; Najafi, R. Sulforaphane, a chemopreventive compound, inhibits cyclooxygenase-2 and microsomal prostaglandin e synthase-1 expression in human ht-29 colon cancer cells. Cells Tissues Organs 2018, 1–8. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, Y.; Dey, M. Dietary phenethyl isothiocyanate protects mice from colitis associated colon cancer. Int. J. Mol. Sci. 2017, 18, 1908. [Google Scholar] [CrossRef]
  96. Yano, S.; Wu, S.; Sakao, K.; Hou, D.X. Wasabi 6-(methylsulfinyl)hexyl isothiocyanate induces apoptosis in human colorectal cancer cells through p53-independent mitochondrial dysfunction pathway. BioFactors 2018. [Google Scholar] [CrossRef] [PubMed]
  97. Guo, Q.; Yuan, Y.; Jin, Z.; Xu, T.; Gao, Y.; Wei, H.; Li, C.; Hou, W.; Hua, B. Association between tumor vasculogenic mimicry and the poor prognosis of gastric cancer in china: An updated systematic review and meta-analysis. Biomed. Res. Int. 2016, 2016, 2408645. [Google Scholar] [CrossRef] [PubMed]
  98. Kuipers, E.J. Review article: Exploring the link between helicobacter pylori and gastric cancer. Aliment. Pharmacol. Ther. 1999, 13 (Suppl. 1), 3–11. [Google Scholar] [CrossRef]
  99. Zulueta, A.; Caretti, A.; Signorelli, P.; Ghidoni, R. Resveratrol: A potential challenger against gastric cancer. World J. Gastroenterol. 2015, 21, 10636–10643. [Google Scholar] [CrossRef] [PubMed]
  100. Overby, A.; Zhao, C.M.; Chen, D. Plant phytochemicals: Potential anticancer agents against gastric cancer. Curr. Opin. Pharmacol. 2014, 19, 6–10. [Google Scholar] [CrossRef] [PubMed]
  101. Fahey, J.W.; Haristoy, X.; Dolan, P.M.; Kensler, T.W.; Scholtus, I.; Stephenson, K.K.; Talalay, P.; Lozniewski, A. Sulforaphane inhibits extracellular, intracellular, and antibiotic-resistant strains of helicobacter pylori and prevents benzo[a]pyrene-induced stomach tumors. Proc. Natl. Acad. Sci. USA. 2002, 99, 7610–7615. [Google Scholar] [CrossRef] [PubMed]
  102. Overby, A.; Zhao, C.M.; Bones, A.M.; Chen, D. Naturally occurring phenethyl isothiocyanate-induced inhibition of gastric cancer cell growth by disruption of microtubules. J. Gastroenterol. Hepatol. 2014, 29 (Suppl. 4), 99–106. [Google Scholar] [CrossRef] [Green Version]
  103. Yang, M.D.; Lai, K.C.; Lai, T.Y.; Hsu, S.C.; Kuo, C.L.; Yu, C.S.; Lin, M.L.; Yang, J.S.; Kuo, H.M.; Wu, S.H.; et al. Phenethyl isothiocyanate inhibits migration and invasion of human gastric cancer ags cells through suppressing mapk and nf-kappab signal pathways. Anticancer Res. 2010, 30, 2135–2143. [Google Scholar] [PubMed]
  104. Ho, C.C.; Lai, K.C.; Hsu, S.C.; Kuo, C.L.; Ma, C.Y.; Lin, M.L.; Yang, J.S.; Chung, J.G. Benzyl isothiocyanate (bitc) inhibits migration and invasion of human gastric cancer ags cells via suppressing erk signal pathways. Hum. Exp. Toxicol. 2011, 30, 296–306. [Google Scholar] [CrossRef] [PubMed]
  105. Liu, X.; Sun, K.; Song, A.; Zhang, X.; Zhang, X.; He, X. Curcumin inhibits proliferation of gastric cancer cells by impairing atp-sensitive potassium channel opening. World J. Surg. Oncol. 2014, 12, 389. [Google Scholar] [CrossRef] [PubMed]
  106. Da, W.; Zhu, J.; Wang, L.; Sun, Q. Curcumin suppresses lymphatic vessel density in an in vivo human gastric cancer model. Tumour. Biol. 2015, 36, 5215–5223. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, P.; Zhang, K.; Zhang, Q.; Mei, J.; Chen, C.J.; Feng, Z.Z.; Yu, D.H. Effects of quercetin on the apoptosis of the human gastric carcinoma cells. Toxicol. In Vitro 2012, 26, 221–228. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, J.Y.; Lin, M.T.; Zhou, M.J.; Yi, T.; Tang, Y.N.; Tang, S.L.; Yang, Z.J.; Zhao, Z.Z.; Chen, H.B. Combinational treatment of curcumin and quercetin against gastric cancer mgc-803 cells in vitro. Molecules 2015, 20, 11524–11534. [Google Scholar] [CrossRef] [PubMed]
  109. Luo, R.; Fang, D.; Hang, H.; Tang, Z. The mechanism in gastric cancer chemoprevention by allicin. Anticancer Agents Med. Chem. 2016, 16, 802–809. [Google Scholar] [CrossRef] [PubMed]
  110. Oser, M.G.; Niederst, M.J.; Sequist, L.V.; Engelman, J.A. Transformation from non-small-cell lung cancer to small-cell lung cancer: Molecular drivers and cells of origin. Lancet Oncol. 2015, 16, e165–172. [Google Scholar] [CrossRef]
  111. Khan, N.; Mukhtar, H. Dietary agents for prevention and treatment of lung cancer. Cancer Lett. 2015, 359, 155–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Jin, L.; Li, C.; Xu, Y.; Wang, L.; Liu, J.; Wang, D.; Hong, C.; Jiang, Z.; Ma, Y.; Chen, Q.; et al. Epigallocatechin gallate promotes p53 accumulation and activity via the inhibition of mdm2-mediated p53 ubiquitination in human lung cancer cells. Oncol. Rep. 2013, 29, 1983–1990. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, H.; Bian, S.; Yang, C.S. Green tea polyphenol egcg suppresses lung cancer cell growth through upregulating mir-210 expression caused by stabilizing hif-1alpha. Carcinogenesis 2011, 32, 1881–1889. [Google Scholar] [CrossRef] [PubMed]
  114. Sadava, D.; Whitlock, E.; Kane, S.E. The green tea polyphenol, epigallocatechin-3-gallate inhibits telomerase and induces apoptosis in drug-resistant lung cancer cells. Biochem. Biophys. Res. Commun. 2007, 360, 233–237. [Google Scholar] [CrossRef] [PubMed]
  115. Mi, L.; Xiao, Z.; Hood, B.L.; Dakshanamurthy, S.; Wang, X.; Govind, S.; Conrads, T.P.; Veenstra, T.D.; Chung, F.L. Covalent binding to tubulin by isothiocyanates. A mechanism of cell growth arrest and apoptosis. J. Biol. Chem. 2008, 283, 22136–22146. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, B.N.; Yan, H.Q.; Wu, X.; Pan, Z.H.; Zhu, Y.; Meng, Z.W.; Zhou, Q.H.; Xu, K. Apoptosis induced by benzyl isothiocyanate in gefitinib-resistant lung cancer cells is associated with akt/mapk pathways and generation of reactive oxygen species. Cell Biochem. Biophys. 2013, 66, 81–92. [Google Scholar] [CrossRef] [PubMed]
  117. Hecht, S.S.; Kenney, P.M.; Wang, M.; Upadhyaya, P. Benzyl isothiocyanate: An effective inhibitor of polycyclic aromatic hydrocarbon tumorigenesis in a/j mouse lung. Cancer Lett. 2002, 187, 87–94. [Google Scholar] [CrossRef]
  118. Pawlik, A.; Szczepanski, M.A.; Klimaszewska, A.; Gackowska, L.; Zuryn, A.; Grzanka, A. Phenethyl isothiocyanate-induced cytoskeletal changes and cell death in lung cancer cells. Food Chem. Toxicol. 2012, 50, 3577–3594. [Google Scholar] [CrossRef] [PubMed]
  119. Kalpana Deepa Priya, D.; Gayathri, R.; Gunassekaran, G.R.; Murugan, S.; Sakthisekaran, D. Apoptotic role of natural isothiocyanate from broccoli (brassica oleracea italica) in experimental chemical lung carcinogenesis. Pharm. Biol. 2013, 51, 621–628. [Google Scholar] [CrossRef] [PubMed]
  120. Jin, C.Y.; Moon, D.O.; Lee, J.D.; Heo, M.S.; Choi, Y.H.; Lee, C.M.; Park, Y.M.; Kim, G.Y. Sulforaphane sensitizes tumor necrosis factor-related apoptosis-inducing ligand-mediated apoptosis through downregulation of erk and akt in lung adenocarcinoma a549 cells. Carcinogenesis 2007, 28, 1058–1066. [Google Scholar] [CrossRef] [PubMed]
  121. Qian, X.; Melkamu, T.; Upadhyaya, P.; Kassie, F. Indole-3-carbinol inhibited tobacco smoke carcinogen-induced lung adenocarcinoma in a/j mice when administered during the post-initiation or progression phase of lung tumorigenesis. Cancer Lett. 2011, 311, 57–65. [Google Scholar] [CrossRef] [PubMed]
  122. Choi, H.S.; Cho, M.C.; Lee, H.G.; Yoon, D.Y. Indole-3-carbinol induces apoptosis through p53 and activation of caspase-8 pathway in lung cancer a549 cells. Food Chem. Toxicol. 2010, 48, 883–890. [Google Scholar] [CrossRef] [PubMed]
  123. Dagne, A.; Melkamu, T.; Schutten, M.M.; Qian, X.; Upadhyaya, P.; Luo, X.; Kassie, F. Enhanced inhibition of lung adenocarcinoma by combinatorial treatment with indole-3-carbinol and silibinin in a/j mice. Carcinogenesis 2011, 32, 561–567. [Google Scholar] [CrossRef] [PubMed]
  124. Melkamu, T.; Zhang, X.; Tan, J.; Zeng, Y.; Kassie, F. Alteration of microrna expression in vinyl carbamate-induced mouse lung tumors and modulation by the chemopreventive agent indole-3-carbinol. Carcinogenesis 2010, 31, 252–258. [Google Scholar] [CrossRef] [PubMed]
  125. Tian, T.; Li, J.; Li, B.; Wang, Y.; Li, M.; Ma, D.; Wang, X. Genistein exhibits anti-cancer effects via down-regulating foxm1 in h446 small-cell lung cancer cells. Tumour Biol. 2014, 35, 4137–4145. [Google Scholar] [CrossRef] [PubMed]
  126. Zhu, H.; Cheng, H.; Ren, Y.; Liu, Z.G.; Zhang, Y.F.; De Luo, B. Synergistic inhibitory effects by the combination of gefitinib and genistein on nsclc with acquired drug-resistance in vitro and in vivo. Mol. Biol. Rep. 2012, 39, 4971–4979. [Google Scholar] [CrossRef] [PubMed]
  127. Li, Y.; Ahmed, F.; Ali, S.; Philip, P.A.; Kucuk, O.; Sarkar, F.H. Inactivation of nuclear factor kappab by soy isoflavone genistein contributes to increased apoptosis induced by chemotherapeutic agents in human cancer cells. Cancer Res. 2005, 65, 6934–6942. [Google Scholar] [CrossRef] [PubMed]
  128. Kashyap, D.; Sharma, A.; Sak, K.; Tuli, H.S.; Buttar, H.S.; Bishayee, A. Fisetin: A bioactive phytochemical with potential for cancer prevention and pharmacotherapy. Life Sci 2018, 194, 75–87. [Google Scholar] [CrossRef] [PubMed]
  129. Khan, N.; Afaq, F.; Khusro, F.H.; Mustafa Adhami, V.; Suh, Y.; Mukhtar, H. Dual inhibition of phosphatidylinositol 3-kinase/akt and mammalian target of rapamycin signaling in human nonsmall cell lung cancer cells by a dietary flavonoid fisetin. Int. J. Cancer 2012, 130, 1695–1705. [Google Scholar] [CrossRef] [PubMed]
  130. Liao, Y.C.; Shih, Y.W.; Chao, C.H.; Lee, X.Y.; Chiang, T.A. Involvement of the erk signaling pathway in fisetin reduces invasion and migration in the human lung cancer cell line a549. J. Agric. Food Chem. 2009, 57, 8933–8941. [Google Scholar] [CrossRef] [PubMed]
  131. Ravichandran, N.; Suresh, G.; Ramesh, B.; Manikandan, R.; Choi, Y.W.; Vijaiyan Siva, G. Fisetin modulates mitochondrial enzymes and apoptotic signals in benzo(a)pyrene-induced lung cancer. Mol. Cell Biochem. 2014, 390, 225–234. [Google Scholar] [CrossRef] [PubMed]
  132. Touil, Y.S.; Seguin, J.; Scherman, D.; Chabot, G.G. Improved antiangiogenic and antitumour activity of the combination of the natural flavonoid fisetin and cyclophosphamide in lewis lung carcinoma-bearing mice. Cancer Chemother. Pharmacol. 2011, 68, 445–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Turrini, E.; Ferruzzi, L.; Fimognari, C. Potential effects of pomegranate polyphenols in cancer prevention and therapy. Oxid. Med. Cell Longev. 2015, 2015, 938475. [Google Scholar] [CrossRef] [PubMed]
  134. Zahin, M.; Ahmad, I.; Gupta, R.C.; Aqil, F. Punicalagin and ellagic acid demonstrate antimutagenic activity and inhibition of benzo[a]pyrene induced DNA adducts. Biomed. Res. Int. 2014, 2014, 467465. [Google Scholar] [CrossRef] [PubMed]
  135. Khan, N.; Hadi, N.; Afaq, F.; Syed, D.N.; Kweon, M.H.; Mukhtar, H. Pomegranate fruit extract inhibits prosurvival pathways in human a549 lung carcinoma cells and tumor growth in athymic nude mice. Carcinogenesis 2007, 28, 163–173. [Google Scholar] [CrossRef] [PubMed]
  136. Khan, N.; Afaq, F.; Kweon, M.H.; Kim, K.; Mukhtar, H. Oral consumption of pomegranate fruit extract inhibits growth and progression of primary lung tumors in mice. Cancer Res. 2007, 67, 3475–3482. [Google Scholar] [CrossRef] [PubMed]
  137. Lu, Y.; Wei, C.; Xi, Z. Curcumin suppresses proliferation and invasion in non-small cell lung cancer by modulation of mta1-mediated wnt/beta-catenin pathway. In Vitro Cell. Dev. Biol. Anim. 2014, 50, 840–850. [Google Scholar] [CrossRef] [PubMed]
  138. Xiao, K.; Jiang, J.; Guan, C.; Dong, C.; Wang, G.; Bai, L.; Sun, J.; Hu, C.; Bai, C. Curcumin induces autophagy via activating the ampk signaling pathway in lung adenocarcinoma cells. J. Pharmacol. Sci. 2013, 123, 102–109. [Google Scholar] [CrossRef] [PubMed]
  139. Li, S.; Liu, Z.; Zhu, F.; Fan, X.; Wu, X.; Zhao, H.; Jiang, L. Curcumin lowers erlotinib resistance in non-small cell lung carcinoma cells with mutated egf receptor. Oncol. Res. 2013, 21, 137–144. [Google Scholar] [CrossRef] [PubMed]
  140. Yang, C.L.; Liu, Y.Y.; Ma, Y.G.; Xue, Y.X.; Liu, D.G.; Ren, Y.; Liu, X.B.; Li, Y.; Li, Z. Curcumin blocks small cell lung cancer cells migration, invasion, angiogenesis, cell cycle and neoplasia through janus kinase-stat3 signalling pathway. PLoS ONE 2012, 7, e37960. [Google Scholar] [CrossRef] [PubMed]
  141. Yang, C.L.; Ma, Y.G.; Xue, Y.X.; Liu, Y.Y.; Xie, H.; Qiu, G.R. Curcumin induces small cell lung cancer nci-h446 cell apoptosis via the reactive oxygen species-mediated mitochondrial pathway and not the cell death receptor pathway. DNA Cell Biol. 2012, 31, 139–150. [Google Scholar] [CrossRef] [PubMed]
  142. Sauter, E.R. Breast cancer prevention: Current approaches and future directions. Eur. J. Breast Health 2018, 14, 64–71. [Google Scholar] [CrossRef] [PubMed]
  143. Friedman, M. Chemistry and anticarcinogenic mechanisms of glycoalkaloids produced by eggplants, potatoes, and tomatoes. J. Agric. Food Chem. 2015, 63, 3323–3337. [Google Scholar] [CrossRef] [PubMed]
  144. Yelken, B.O.; Balci, T.; Susluer, S.Y.; Kayabasi, C.; Avci, C.B.; Kirmizibayrak, P.B.; Gunduz, C. The effect of tomatine on metastasis related matrix metalloproteinase (mmp) activities in breast cancer cell model. Gene 2017, 627, 408–411. [Google Scholar] [CrossRef] [PubMed]
  145. Caruso, J.A.; Campana, R.; Wei, C.; Su, C.H.; Hanks, A.M.; Bornmann, W.G.; Keyomarsi, K. Indole-3-carbinol and its n-alkoxy derivatives preferentially target eralpha-positive breast cancer cells. Cell Cycle 2014, 13, 2587–2599. [Google Scholar] [CrossRef] [PubMed]
  146. Tin, A.S.; Park, A.H.; Sundar, S.N.; Firestone, G.L. Essential role of the cancer stem/progenitor cell marker nucleostemin for indole-3-carbinol anti-proliferative responsiveness in human breast cancer cells. BMC Biol. 2014, 12, 72. [Google Scholar] [CrossRef] [PubMed]
  147. Bishayee, A.; Ahmed, S.; Brankov, N.; Perloff, M. Triterpenoids as potential agents for the chemoprevention and therapy of breast cancer. Front. Biosci. (Landmark Ed.) 2011, 16, 980–996. [Google Scholar] [CrossRef] [PubMed]
  148. Jayaprakasam, B.; Seeram, N.P.; Nair, M.G. Anticancer and antiinflammatory activities of cucurbitacins from cucurbita andreana. Cancer Lett. 2003, 189, 11–16. [Google Scholar] [CrossRef]
  149. Ramalhete, C.; Mansoor, T.A.; Mulhovo, S.; Molnar, J.; Ferreira, M.J. Cucurbitane-type triterpenoids from the african plant momordica balsamina. J. Nat. Prod. 2009, 72, 2009–2013. [Google Scholar] [CrossRef] [PubMed]
  150. Wakimoto, N.; Yin, D.; O’Kelly, J.; Haritunians, T.; Karlan, B.; Said, J.; Xing, H.; Koeffler, H.P. Cucurbitacin b has a potent antiproliferative effect on breast cancer cells in vitro and in vivo. Cancer Sci. 2008, 99, 1793–1797. [Google Scholar] [CrossRef] [PubMed]
  151. Kongtun, S.; Jiratchariyakul, W.; Kummalue, T.; Tan-ariya, P.; Kunnachak, S.; Frahm, A.W. Cytotoxic properties of root extract and fruit juice of trichosanthes cucumerina. Planta Med. 2009, 75, 839–842. [Google Scholar] [CrossRef] [PubMed]
  152. Rodriguez, N.; Vasquez, Y.; Hussein, A.A.; Coley, P.D.; Solis, P.N.; Gupta, M.P. Cytotoxic cucurbitacin constituents from sloanea zuliaensis. J. Nat. Prod. 2003, 66, 1515–1516. [Google Scholar] [CrossRef] [PubMed]
  153. Blaskovich, M.A.; Sun, J.; Cantor, A.; Turkson, J.; Jove, R.; Sebti, S.M. Discovery of jsi-124 (cucurbitacin i), a selective janus kinase/signal transducer and activator of transcription 3 signaling pathway inhibitor with potent antitumor activity against human and murine cancer cells in mice. Cancer Res. 2003, 63, 1270–1279. [Google Scholar] [PubMed]
  154. Sun, J.; Blaskovich, M.A.; Jove, R.; Livingston, S.K.; Coppola, D.; Sebti, S.M. Cucurbitacin q: A selective stat3 activation inhibitor with potent antitumor activity. Oncogene 2005, 24, 3236–3245. [Google Scholar] [CrossRef] [PubMed]
  155. Yang, L.; Wu, S.; Zhang, Q.; Liu, F.; Wu, P. 23,24-dihydrocucurbitacin b induces g2/m cell-cycle arrest and mitochondria-dependent apoptosis in human breast cancer cells (bcap37). Cancer Lett. 2007, 256, 267–278. [Google Scholar] [CrossRef] [PubMed]
  156. Phongmaykin, J.; Kumamoto, T.; Ishikawa, T.; Suttisri, R.; Saifah, E. A new sesquiterpene and other terpenoid constituents of chisocheton penduliflorus. Arch. Pharm. Res. 2008, 31, 21–27. [Google Scholar] [CrossRef] [PubMed]
  157. Wu, C.C.; Chan, M.L.; Chen, W.Y.; Tsai, C.Y.; Chang, F.R.; Wu, Y.C. Pristimerin induces caspase-dependent apoptosis in mda-mb-231 cells via direct effects on mitochondria. Mol. Cancer Ther. 2005, 4, 1277–1285. [Google Scholar] [CrossRef] [PubMed]
  158. Idris, A.I.; Libouban, H.; Nyangoga, H.; Landao-Bassonga, E.; Chappard, D.; Ralston, S.H. Pharmacologic inhibitors of ikappab kinase suppress growth and migration of mammary carcinosarcoma cells in vitro and prevent osteolytic bone metastasis in vivo. Mol. Cancer Ther. 2009, 8, 2339–2347. [Google Scholar] [CrossRef] [PubMed]
  159. Zeng, L.; Gu, Z.M.; Chang, C.J.; Wood, K.V.; McLaughlin, J.L. Meliavolkenin, a new bioactive triterpenoid from melia volkensii (meliaceae). Bioorg. Med. Chem. 1995, 3, 383–390. [Google Scholar] [CrossRef]
  160. Amico, V.; Barresi, V.; Condorelli, D.; Spatafora, C.; Tringali, C. Antiproliferative terpenoids from almond hulls (prunus dulcis): Identification and structure-activity relationships. J. Agric. Food Chem. 2006, 54, 810–814. [Google Scholar] [CrossRef] [PubMed]
  161. Rzeski, W.; Stepulak, A.; Szymanski, M.; Sifringer, M.; Kaczor, J.; Wejksza, K.; Zdzisinska, B.; Kandefer-Szerszen, M. Betulinic acid decreases expression of bcl-2 and cyclin d1, inhibits proliferation, migration and induces apoptosis in cancer cells. Naunyn Schmiedebergs Arch. Pharmacol. 2006, 374, 11–20. [Google Scholar] [CrossRef] [PubMed]
  162. Kessler, J.H.; Mullauer, F.B.; de Roo, G.M.; Medema, J.P. Broad in vitro efficacy of plant-derived betulinic acid against cell lines derived from the most prevalent human cancer types. Cancer Lett. 2007, 251, 132–145. [Google Scholar] [CrossRef] [PubMed]
  163. Lambertini, E.; Lampronti, I.; Penolazzi, L.; Khan, M.T.; Ather, A.; Giorgi, G.; Gambari, R.; Piva, R. Expression of estrogen receptor alpha gene in breast cancer cells treated with transcription factor decoy is modulated by bangladeshi natural plant extracts. Oncol. Res. 2005, 15, 69–79. [Google Scholar] [PubMed]
  164. Es-Saady, D.; Simon, A.; Jayat-Vignoles, C.; Chulia, A.J.; Delage, C. Mcf-7 cell cycle arrested at g1 through ursolic acid, and increased reduction of tetrazolium salts. Anticancer Res. 1996, 16, 481–486. [Google Scholar] [PubMed]
  165. Chen, Y.H.; Chang, F.R.; Wu, C.C.; Yen, M.H.; Liaw, C.C.; Huang, H.C.; Kuo, Y.H.; Wu, Y.C. New cytotoxic 6-oxygenated 8,9-dihydrofurocoumarins, hedyotiscone A-C, from hedyotis biflora. Planta Med. 2006, 72, 75–78. [Google Scholar] [CrossRef] [PubMed]
  166. Martin-Cordero, C.; Reyes, M.; Ayuso, M.J.; Toro, M.V. Cytotoxic triterpenoids from erica andevalensis. Z. Naturforsch. C 2001, 56, 45–48. [Google Scholar] [CrossRef] [PubMed]
  167. Neto, C.C.; Vaisberg, A.J.; Zhou, B.N.; Kingston, D.G.; Hammond, G.B. Cytotoxic triterpene acids from the peruvian medicinal plant polylepis racemosa. Planta Med. 2000, 66, 483–484. [Google Scholar] [CrossRef] [PubMed]
  168. Kassi, E.; Sourlingas, T.G.; Spiliotaki, M.; Papoutsi, Z.; Pratsinis, H.; Aligiannis, N.; Moutsatsou, P. Ursolic acid triggers apoptosis and bcl-2 downregulation in mcf-7 breast cancer cells. Cancer Investig. 2009, 27, 723–733. [Google Scholar] [CrossRef] [PubMed]
  169. Yeh, C.T.; Wu, C.H.; Yen, G.C. Ursolic acid, a naturally occurring triterpenoid, suppresses migration and invasion of human breast cancer cells by modulating c-jun n-terminal kinase, akt and mammalian target of rapamycin signaling. Mol. Nutr. Food Res. 2010, 54, 1285–1295. [Google Scholar] [CrossRef] [PubMed]
  170. Hsu, Y.L.; Kuo, P.L.; Lin, L.T.; Lin, C.C. Asiatic acid, a triterpene, induces apoptosis and cell cycle arrest through activation of extracellular signal-regulated kinase and p38 mitogen-activated protein kinase pathways in human breast cancer cells. J. Pharmacol. Exp. Ther. 2005, 313, 333–344. [Google Scholar] [CrossRef] [PubMed]
  171. Deng, Y.; Jiang, T.Y.; Sheng, S.; Tianasoa-Ramamonjy, M.; Snyder, J.K. Remangilones a-c, new cytotoxic triterpenes from physena madagascariensis. J. Nat. Prod 1999, 62, 471–476. [Google Scholar] [CrossRef] [PubMed]
  172. Rabi, T.; Karunagaran, D.; Krishnan Nair, M.; Bhattathiri, V.N. Cytotoxic activity of amooranin and its derivatives. Phytother. Res. 2002, 16 (Suppl. 1), S84–S86. [Google Scholar] [CrossRef]
  173. Rabi, T.; Ramachandran, C.; Fonseca, H.B.; Nair, R.P.; Alamo, A.; Melnick, S.J.; Escalon, E. Novel drug amooranin induces apoptosis through caspase activity in human breast carcinoma cell lines. Breast Cancer Res. Treat. 2003, 80, 321–330. [Google Scholar] [CrossRef] [PubMed]
  174. Rabi, T.; Wang, L.; Banerjee, S. Novel triterpenoid 25-hydroxy-3-oxoolean-12-en-28-oic acid induces growth arrest and apoptosis in breast cancer cells. Breast Cancer Res. Treat. 2007, 101, 27–36. [Google Scholar] [CrossRef] [PubMed]
  175. Chavez, I.O.; Apan, T.R.; Martinez-Vazquez, M. Cytotoxic activity and effect on nitric oxide production of tirucallane-type triterpenes. J. Pharm. Pharmacol. 2005, 57, 1087–1091. [Google Scholar] [CrossRef] [PubMed]
  176. Lavhale, M.S.; Kumar, S.; Mishra, S.H.; Sitasawad, S.L. A novel triterpenoid isolated from the root bark of ailanthus excelsa roxb (tree of heaven), aechl-1 as a potential anti-cancer agent. PLoS ONE 2009, 4, e5365. [Google Scholar] [CrossRef] [PubMed]
  177. Li, X.; Xie, W.; Xie, C.; Huang, C.; Zhu, J.; Liang, Z.; Deng, F.; Zhu, M.; Zhu, W.; Wu, R.; et al. Curcumin modulates mir-19/pten/akt/p53 axis to suppress bisphenol a-induced mcf-7 breast cancer cell proliferation. Phytother. Res. 2014, 28, 1553–1560. [Google Scholar] [CrossRef] [PubMed]
  178. Miksits, M.; Wlcek, K.; Svoboda, M.; Kunert, O.; Haslinger, E.; Thalhammer, T.; Szekeres, T.; Jager, W. Antitumor activity of resveratrol and its sulfated metabolites against human breast cancer cells. Planta Med. 2009, 75, 1227–1230. [Google Scholar] [CrossRef] [PubMed]
  179. Huang, C.; Lu, C.K.; Tu, M.C.; Chang, J.H.; Chen, Y.J.; Tu, Y.H.; Huang, H.C. Polyphenol-rich avicennia marina leaf extracts induce apoptosis in human breast and liver cancer cells and in a nude mouse xenograft model. Oncotarget 2016, 7, 35874–35893. [Google Scholar] [CrossRef] [PubMed]
  180. Dominguez, F.; Maycotte, P.; Acosta-Casique, A.; Rodriguez-Rodriguez, S.; Moreno, D.A.; Ferreres, F.; Flores-Alonso, J.C.; Delgado-Lopez, M.G.; Perez-Santos, M.; Anaya-Ruiz, M. Bursera copallifera extracts have cytotoxic and migration-inhibitory effects in breast cancer cell lines. Integr. Cancer Ther. 2018. [Google Scholar] [CrossRef] [PubMed]
  181. G, W.W.; L, M.B.; D, E.W.; R, H.D.; Ho, E. Phytochemicals from cruciferous vegetables, epigenetics, and prostate cancer prevention. AAPS J. 2013, 15, 951–961. [Google Scholar]
  182. Wu, W.; Tang, S.N.; Zhang, Y.; Puppala, M.; Cooper, T.K.; Xing, C.; Jiang, C.; Lu, J. Prostate cancer xenograft inhibitory activity and pharmacokinetics of decursinol, a metabolite of angelica gigas pyranocoumarins, in mouse models. Am. J. Chin. Med. 2017, 45, 1773–1792. [Google Scholar] [CrossRef] [PubMed]
  183. Tang, S.N.; Zhang, J.; Wu, W.; Jiang, P.; Puppala, M.; Zhang, Y.; Xing, C.; Kim, S.H.; Jiang, C.; Lu, J. Chemopreventive effects of korean angelica versus its major pyranocoumarins on two lineages of transgenic adenocarcinoma of mouse prostate carcinogenesis. Cancer Prev. Res. (Phila) 2015, 8, 835–844. [Google Scholar] [CrossRef] [PubMed]
  184. Empl, M.T.; Albers, M.; Wang, S.; Steinberg, P. The resveratrol tetramer r-viniferin induces a cell cycle arrest followed by apoptosis in the prostate cancer cell line lncap. Phytother. Res. 2015, 29, 1640–1645. [Google Scholar] [CrossRef] [PubMed]
  185. Mukhtar, E.; Adhami, V.M.; Sechi, M.; Mukhtar, H. Dietary flavonoid fisetin binds to beta-tubulin and disrupts microtubule dynamics in prostate cancer cells. Cancer Lett. 2015, 367, 173–183. [Google Scholar] [CrossRef] [PubMed]
  186. Pavese, J.M.; Krishna, S.N.; Bergan, R.C. Genistein inhibits human prostate cancer cell detachment, invasion, and metastasis. Am. J. Clin. Nutr. 2014, 100 (Suppl. 1), 431S–436S. [Google Scholar] [CrossRef]
  187. Du, Y.; Long, Q.; Zhang, L.; Shi, Y.; Liu, X.; Li, X.; Guan, B.; Tian, Y.; Wang, X.; Li, L.; et al. Curcumin inhibits cancer-associated fibroblast-driven prostate cancer invasion through maoa/mtor/hif-1alpha signaling. Int. J. Oncol. 2015, 47, 2064–2072. [Google Scholar] [CrossRef] [PubMed]
  188. Dorai, T.; Diouri, J.; O’Shea, O.; Doty, S.B. Curcumin inhibits prostate cancer bone metastasis by up-regulating bone morphogenic protein-7 in vivo. J. Cancer Ther. 2014, 5, 369–386. [Google Scholar] [CrossRef] [PubMed]
  189. Zhou, D.Y.; Ding, N.; Du, Z.Y.; Cui, X.X.; Wang, H.; Wei, X.C.; Conney, A.H.; Zhang, K.; Zheng, X. Curcumin analogues with high activity for inhibiting human prostate cancer cell growth and androgen receptor activation. Mol. Med. Rep. 2014, 10, 1315–1322. [Google Scholar] [CrossRef] [PubMed]
  190. Taylor, J.; Xiao, W.; Abdel-Wahab, O. Diagnosis and classification of hematologic malignancies on the basis of genetics. Blood 2017, 130, 410–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Jendzelovska, Z.; Jendzelovsky, R.; Hilovska, L.; Koval, J.; Mikes, J.; Fedorocko, P. Single pre-treatment with hypericin, a st. John’s wort secondary metabolite, attenuates cisplatin- and mitoxantrone-induced cell death in a2780, a2780cis and hl-60 cells. Toxicol. In Vitro 2014, 28, 1259–1273. [Google Scholar] [CrossRef] [PubMed]
  192. Azmi, A.S.; Bhat, S.H.; Hanif, S.; Hadi, S.M. Plant polyphenols mobilize endogenous copper in human peripheral lymphocytes leading to oxidative DNA breakage: A putative mechanism for anticancer properties. FEBS Lett. 2006, 580, 533–538. [Google Scholar] [CrossRef] [PubMed]
  193. Estrov, Z.; Shishodia, S.; Faderl, S.; Harris, D.; Van, Q.; Kantarjian, H.M.; Talpaz, M.; Aggarwal, B.B. Resveratrol blocks interleukin-1beta-induced activation of the nuclear transcription factor nf-kappab, inhibits proliferation, causes s-phase arrest, and induces apoptosis of acute myeloid leukemia cells. Blood 2003, 102, 987–995. [Google Scholar] [CrossRef] [PubMed]
  194. Noda, C.; He, J.; Takano, T.; Tanaka, C.; Kondo, T.; Tohyama, K.; Yamamura, H.; Tohyama, Y. Induction of apoptosis by epigallocatechin-3-gallate in human lymphoblastoid b cells. Biochem. Biophys. Res. Commun. 2007, 362, 951–957. [Google Scholar] [CrossRef] [PubMed]
  195. Safa, M.; Tavasoli, B.; Manafi, R.; Kiani, F.; Kashiri, M.; Ebrahimi, S.; Kazemi, A. Indole-3-carbinol suppresses nf-kappab activity and stimulates the p53 pathway in pre-b acute lymphoblastic leukemia cells. Tumour Biol. 2015, 36, 3919–3930. [Google Scholar] [CrossRef] [PubMed]
  196. Linares, M.A.; Zakaria, A.; Nizran, P. Skin cancer. Prim. Care 2015, 42, 645–659. [Google Scholar] [CrossRef] [PubMed]
  197. Wang, H.C.; Pao, J.; Lin, S.Y.; Sheen, L.Y. Molecular mechanisms of garlic-derived allyl sulfides in the inhibition of skin cancer progression. Ann. N. Y. Acad. Sci. 2012, 1271, 44–52. [Google Scholar] [CrossRef] [PubMed]
  198. Nigam, N.; Shukla, Y. Preventive effects of diallyl sulfide on 7,12-dimethylbenz[a]anthracene induced DNA alkylation damage in mouse skin. Mol. Nutr. Food Res. 2007, 51, 1324–1328. [Google Scholar] [CrossRef] [PubMed]
  199. Arora, A.; Shukla, Y. Induction of apoptosis by diallyl sulfide in dmba-induced mouse skin tumors. Nutr. Cancer 2002, 44, 89–94. [Google Scholar] [CrossRef] [PubMed]
  200. Kalra, N.; Arora, A.; Shukla, Y. Involvement of multiple signaling pathways in diallyl sulfide mediated apoptosis in mouse skin tumors. Asian Pac. J. Cancer Prev. 2006, 7, 556–562. [Google Scholar] [PubMed]
  201. Chang, H.P.; Sheen, L.Y.; Lei, Y.P. The protective role of carotenoids and polyphenols in patients with head and neck cancer. J. Chin. Med. Assoc. 2015, 78, 89–95. [Google Scholar] [CrossRef] [PubMed]
  202. Beta-carotene. In Drugs and Lactation Database (LactMed). Bethesda (MD). Available online: www.ncbi.nlm.nih.gov/books/NBK501922/ (accessed on 22 July 2018).
  203. Zhang, Y.; Zhu, X.; Huang, T.; Chen, L.; Liu, Y.; Li, Q.; Song, J.; Ma, S.; Zhang, K.; Yang, B.; et al. Beta-carotene synergistically enhances the anti-tumor effect of 5-fluorouracil on esophageal squamous cell carcinoma in vivo and in vitro. Toxicol. Lett. 2016, 261, 49–58. [Google Scholar] [CrossRef] [PubMed]
  204. Masuda, M.; Suzui, M.; Weinstein, I.B. Effects of epigallocatechin-3-gallate on growth, epidermal growth factor receptor signaling pathways, gene expression, and chemosensitivity in human head and neck squamous cell carcinoma cell lines. Clin. Cancer Res. 2001, 7, 4220–4229. [Google Scholar] [PubMed]
  205. Amin, A.R.; Khuri, F.R.; Chen, Z.G.; Shin, D.M. Synergistic growth inhibition of squamous cell carcinoma of the head and neck by erlotinib and epigallocatechin-3-gallate: The role of p53-dependent inhibition of nuclear factor-kappab. Cancer Prev. Res. (Phila) 2009, 2, 538–545. [Google Scholar] [CrossRef] [PubMed]
  206. Ho, Y.C.; Yang, S.F.; Peng, C.Y.; Chou, M.Y.; Chang, Y.C. Epigallocatechin-3-gallate inhibits the invasion of human oral cancer cells and decreases the productions of matrix metalloproteinases and urokinase-plasminogen activator. J. Oral. Pathol. Med. 2007, 36, 588–593. [Google Scholar] [CrossRef] [PubMed]
  207. Atanasov, A.G.; Waltenberger, B.; Pferschy-Wenzig, E.M.; Linder, T.; Wawrosch, C.; Uhrin, P.; Temml, V.; Wang, L.; Schwaiger, S.; Heiss, E.H.; et al. Discovery and resupply of pharmacologically active plant-derived natural products: A review. Biotechnol. Adv. 2015, 33, 1582–1614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Structure of phenolic compounds.
Figure 1. Structure of phenolic compounds.
Ijms 19 02651 g001
Figure 2. Structure of terpenoids.
Figure 2. Structure of terpenoids.
Ijms 19 02651 g002
Figure 3. Structure of nitrogen-containing alkaloids and sulfur-containing compounds.
Figure 3. Structure of nitrogen-containing alkaloids and sulfur-containing compounds.
Ijms 19 02651 g003
Table 1. Structural classification of active metabolites with anticancer activity.
Table 1. Structural classification of active metabolites with anticancer activity.
ClassActive MetaboliteStructure
Phenolic compoundsCurcumin Ijms 19 02651 i001
Decursin Ijms 19 02651 i002
Decursinol Ijms 19 02651 i003
Decursinol angelate Ijms 19 02651 i004
Dichamanetin Ijms 19 02651 i005
Epigallocatechin gallate (EGCG) Ijms 19 02651 i006
Fisetin Ijms 19 02651 i007
Genistein Ijms 19 02651 i008
Hydroxycinnamic acid Ijms 19 02651 i009
Hypericin Ijms 19 02651 i010
Quercetin Ijms 19 02651 i011
Phenolic compoundsResveratrol Ijms 19 02651 i012
Punicalagin (PC) Ijms 19 02651 i013
γ-viniferin Ijms 19 02651 i014
TerpenoidsAsiatic acid Ijms 19 02651 i015
Ailanthus excelsa chloroform extract-1 (AECHL-1) Ijms 19 02651 i016
Amooranin (AMR) Ijms 19 02651 i017
Auraptene (AUR) Ijms 19 02651 i018
Betulinic acid (BA) Ijms 19 02651 i019
Celastrol Ijms 19 02651 i020
Curcubitane-type triterpenoids
(Balsaminapentaol)
Ijms 19 02651 i021
Dammarane triterpenoid
(Cabraleadiol)
Ijms 19 02651 i022
Lupeol Ijms 19 02651 i023
Meliavolkenin Ijms 19 02651 i024
Pomolic acid Ijms 19 02651 i025
Pristimerin Ijms 19 02651 i026
Remangilones A Ijms 19 02651 i027
Remangilones C Ijms 19 02651 i028
Tirucallane-type triterpenoids Ijms 19 02651 i029
Ursolic acid Ijms 19 02651 i030
β-carotene Ijms 19 02651 i031
Nitrogen-containing alkaloids & sulfur-containing compoundsAllicin Ijms 19 02651 i032
Benzyl isothiocyanate (BITC) Ijms 19 02651 i033
Diallyl sulfide (DAS) Ijms 19 02651 i034
Indole-3-carbinol (I3C) Ijms 19 02651 i035
Phenethyl isothiocyanate (PEITC) Ijms 19 02651 i036
Sulforaphane (SFN) Ijms 19 02651 i037
Tomatine Ijms 19 02651 i038
6-MSITC Ijms 19 02651 i039
Table 2. Anti-cancer effects of active metabolites from plants in different types of cancer.
Table 2. Anti-cancer effects of active metabolites from plants in different types of cancer.
Type of CancerActive MetabolitesIn Vitro or In Vivo EffectsIC50 & Effective Concentration (EC) (μM).Ref.
Colorectal cancerDichamanetinInduction of ROS and cell cycle arrest in HT-29 colon cancer cellsIC50: 13.8[80]
GAPSuppression of colon carcinogenesis in DSS miceEC: 0.01 % or 0.05 % in diet[83]
AurapteneInhibition of the growth of colon cancer cells and suppression of colonosphere formationEC: 10[85]
CurcuminSuppression of the oncogenicity of human colon cancer cells and the growth of HCT-116 tumor xenograftsEC: 10[87]
GenisteinInhibition of cell proliferation and induction of apoptosis in HCT 116 and LoVo cellsEC: 135[89,90,91]
Inhibition of metastasis in colorectal cancer cell implanted nude mice
I3C+GenisteinInduction of apoptosis in HT 29 colon cancer cellsI3C EC: 300
Genistein EC: 40
[92]
BITCSuppression of viability in HCT 116 colon cancer cellsEC: 5–20[93]
SFNInduction of apoptosis and inhibition of proliferation in HT 29 colon cancer cellsEC: 5–20[94]
PEITCReduction of colon carcinogenesis in AOM/DSS induced miceEC: 0.12 % in diet[95]
6-MSITCInduction of apoptosis in HCT 116 colon cancer cellsIC50: 0.92–10.01[96]
Gastric cancerResveratrolInhibition of proliferation in gastric cancer cellsEC: 50–200[99]
SFNPrevention of benzo[a]pyrene-induced stomach tumors in miceEC: 1.33 mg per mouse[101]
PEITCInduction of cell cycle arrest and apoptosis in gastric cancer cells MKN74 and Kato-IIIEC: 17.8[102,103]
Inhibition of migration and invasion in AGS gastric cancer cells
BITCInhibition of migration and invasion in AGS gastric cancer cellsEC: 0.25–0.5[104]
CurcuminInhibition of proliferation in SGC-7901 gastric cancer cells EC: 15–60 [105,106]
Reduction of xenograft tumor growth in mice
Reduction of LVD in gastric cancer bearing nude mice
QuercetinInduction of apoptosis in BGC-823 gastric cancer cells EC: 30–120[107]
AllicinInhibition of gastric cancer cell growthEC: 184.88[109]
Lung cancerEGCGInduction of cell cycle arrest and apoptosis in lung cancer cells
Reduction of proliferation and growth in lung cancer cells
Inhibition of TGF-β-induced cell migration, invasion, and EMT in NSCLC cells
IC50: 70[112,113,114]
BITCInhibition of growth in A549 lung cancer cellsEC: 10[115,116,117]
Inhibition of tumorigenesis in PAH-induced A/J mice
PEITCInduction of apoptosis in NSCLC cells EC: 12.5–20 [113,118]
SFNInduction of apoptosis in NSCLC cellsEC: 10[119,120]
Alleviation of carcinogenic lung in B(a)P induced lung cancer bearing mice
I3CInduction of apoptosis in A549 lung adenocarcinoma cells in combination with TRAILEC: 100–500[121,122,123,124]
Inhibition of progression of tobacco carcinogen induced lung adenocarcinoma progression
Induction of cell cycle arrest and apoptosis in A549 lung cancer cells
Inhibition of NNK-induced lung tumors in combination with silibinin in mice
GenisteinInhibition of carcinogenesis in mice with VC-induced lung cancer
Inhibition of cell proliferation and induction of apoptosis in H446 SCLC cells
IC50: 81[125,126,127]
Inhibition of cell proliferation and induction of apoptosis in combination with gefitinib in H1975 NSCLC cells
FisetinInhibition of cell growth and induction of apoptosis in combination with chemotherapeutic agents in H460 NSCLC cellsIC50: 59[129,130,131,132]
Inhibition of cell viability and colony-forming activity in A549 NSCLC cells
Inhibition of the invasion and migration of A549 NSCLC cells
Inhibition of lung carcinogenesis in B(a)P-induced mice
Inhibition of angiogenesis and tumor growth in Lewis lung carcinoma bearing mice
PunicalaginAnti-proliferative effects on A549 and H1299 NSCLC cells
Inhibition of tumor growth in mice with xenografts of A549 NSCLC cells
Inhibition of B(a)P-induced tumorigenesis in A/J mice
EC: 11.52–184.3[134,135,136]
CurcuminInhibition of cell growth and invasion in NSCLC cells
Lowering the resistance of NSCLC cells against erlotinib
Suppression of cell proliferation, the cell cycle, migration, invasion, and angiogenesis in SCLC cells
Induction of apoptosis in SCLC cells
EC: 30[137,138,139,140,141]
Breast cancerTomatineInduction of cell cytotoxicity and apoptosis in MCF-7 breast cancer cellsIC50: 7.07[144]
I3CIncreasing apoptotic cell death and decreasing the proliferation of the ERα-positive breast cancer cells
Disruption of in vitro 10AT-Her2 cell tumorsphere formation and in vivo tumor xenograft growth
IC50: 204[145,146]
Curcubitane-type triterpenoidsInhibition of cell growth and induction of apoptosis in human breast cancer cellsEC: 0.5–35.7[148,149,150,151,152,153,154,155]
Dammarane triterpenoidsCytotoxicity against breast cancer cellsEC: 20.97[156]
PristimerinInduction of apoptosis in MDA-MB-231 breast cancer cellsEC: 1–3[157]
CelastrolInhibition of cell growth and invasion and induction of apoptosis in W256 breast cancer cells EC: 1[158]
MeliavolkeninCytotoxicity against MCF7 breast cancer cellsEC: 6.05[159]
Betulinic acidInduction of anti-proliferation in MCF7 and T47D breast cancer cellsIC50: 2.4[160,161,162]
LupeolInhibition of MDA-MB-231 ERα-negative cell proliferationEC: 1–30[163]
Ursolic acidInhibition of proliferation and induction of apoptosis in MCF7 cells
Suppression of migration and invasion in MDA-MB-231 cells
IC50: 3.26[164,165,166,167,168,169]
Asiatic acidInhibition of cell growth and induction of apoptosis in MCF7 and MDA-MB 231 cellsIC50: 5.95–8.12[170]
Remangilones A and CCytotoxicity against MDA-MB-231 and MDA-MB-435 cellsRemangilonesA IC50: 6.6–8.5
RemangilonesC IC50: 1.6–2.0
[171]
AmooraninInduction of apoptosis and suppression of cell growth in MDA-468 and MCF7 cellsIC50: 3.82-7.22[172,173,174]
Tirucallane-type triterpenoidsCytotoxicity against MCF7 cellsIC50: 41.33–86.14[175]
AECHL-1Regression of MCF7 xenograft tumors in nude miceEC: 5–100[176]
CurcuminAnti-proliferation of BPA-induced MCF7 cellsEC: 1[177]
ResveratrolReduction of cell viability in breast cancer cells (MCF-7, ZR-75-1, and MDA-MB-231)IC50: 67.6–82.2[178]
Hydroxycinnamic acidInhibition of migration in MCF-7 and MDA-MB-231 cellsIC50: 75.71[180]
DichamanetinInduction of ROS and cell cycle arrest in MDA-MB-231 cellsEC: 8.7[80]
Prostate cancerDecursinolSuppression of tumor growth in mice with xenografted DU145 and PC3 prostate cancersEC: 4.5 mg per mouse[182]
Decursin & Decursinol angelateInhibition of prostate epithelium growth in the TRAMP modelEC: 3 mg per mouse[183]
Resveratrol & γ-viniferinInhibition of the growth of LNCaP prostate cancer cellResveratrol IC50: 10.23-228.3
γ-viniferin IC50: 8.93–90.1
[184]
FisetinInhibition of cell growth and proliferation in PU3 and DU145 cellsEC: 20–80[185]
GenisteinInhibition of cellular invasion in in vitro prostate cancer and in vivo metastasis formation in mice with xenografts of PC3-M prostate cancerEC: 10[186]
CurcuminInhibition of CAF-induced EMT and invasion in PC3 cells
Induction of cell cycle arrest and apoptosis in in vitro prostate cancer cells and the in vivo TRAMP model
EC: 25[187,188,189]
SFN and I3CInduction of cell cycle arrest and apoptosis of PC3, LNCaP, and DU145 cells in vitroSFN EC: 40
I3C EC: 30–100
[181]
Hematologic cancerHypericinAttenuation of MTX cytotoxicity in HL-60 promyelocytic leukemia cellsEC: 0.1–0.5[191]
ResveratrolInduction of DNA breakage in human peripheral lymphocytes
Induction of apoptosis in OCI-ANK3 and OCIM2 acute myeloid leukemia cells
EC: 10–75 [192,193]
EGCGInduction of apoptotic death in Ramos B lymphoblastoid cellsEC: 60–100[194]
I3CInhibition of cell growth and induction of apoptosis in pre-B acute lymphoblastic leukemia cellsEC: 60[195]
Skin cancerDiallyl sulfideReduction of DNA strand breaks in DMBA induced mouse skin
Induction of apoptosis in DMBA-induced mouse skin tumors
EC: 25[197,198,199,200]
Pomolic acidCytotoxic effects against M-14 melanoma cellsEC: 14.6[167]
Head and neck cancerβ-caroteneInhibition of tumor growth in nude mice with xenografts of Eca109 ESCC cell xenograftsEC: 30[203]
EGCGInduction of cell cycle arrest and apoptosis in YCU-N861 and YCU-H891 HNSCC cells
Inhibition of cell growth in combination with erlotinib in HNSCC cells
Inhibition of the invasion and migration in oral cancer cell OC2
EC: 30–60[204,205,206]

Share and Cite

MDPI and ACS Style

Shin, S.-A.; Moon, S.Y.; Kim, W.-Y.; Paek, S.-M.; Park, H.H.; Lee, C.S. Structure-Based Classification and Anti-Cancer Effects of Plant Metabolites. Int. J. Mol. Sci. 2018, 19, 2651. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms19092651

AMA Style

Shin S-A, Moon SY, Kim W-Y, Paek S-M, Park HH, Lee CS. Structure-Based Classification and Anti-Cancer Effects of Plant Metabolites. International Journal of Molecular Sciences. 2018; 19(9):2651. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms19092651

Chicago/Turabian Style

Shin, Seong-Ah, Sun Young Moon, Woe-Yeon Kim, Seung-Mann Paek, Hyun Ho Park, and Chang Sup Lee. 2018. "Structure-Based Classification and Anti-Cancer Effects of Plant Metabolites" International Journal of Molecular Sciences 19, no. 9: 2651. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms19092651

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop