Next Article in Journal
FKBP8 Enhances Protein Stability of the CLC-1 Chloride Channel at the Plasma Membrane
Next Article in Special Issue
Melatonin May Increase Anticancer Potential of Pleiotropic Drugs
Previous Article in Journal
Melatonin and Its Metabolites Ameliorate UVR-Induced Mitochondrial Oxidative Stress in Human MNT-1 Melanoma Cells
Previous Article in Special Issue
Targeting Cancer with Phytochemicals via Their Fine Tuning of the Cell Survival Signaling Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chemoprevention of Colorectal Cancer by Dietary Compounds

by
Teodora Costea
1,†,
Ariana Hudiță
2,†,
Oana-Alina Ciolac
3,
Bianca Gălățeanu
2,
Octav Ginghină
4,5,
Marieta Costache
2,
Constanța Ganea
3 and
Maria-Magdalena Mocanu
3,*
1
Department of Pharmacognosy, Phytochemistry and Phytotherapy, ”Carol Davila” University of Medicine and Pharmacy, 020956 Bucharest, Romania
2
Department of Biochemistry and Molecular Biology, University of Bucharest, 050095 Bucharest, Romania
3
Department of Biophysics, ”Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
4
Department of Surgery, ”Sf. Ioan” Emergency Clinical Hospital, 042122 Bucharest, Romania
5
Department II, Faculty of Dental Medicine, “Carol Davila” University of Medicine and Pharmacy, 030167 Bucharest, Romania
*
Author to whom correspondence should be addressed.
These two authors equally contributed to the work.
Int. J. Mol. Sci. 2018, 19(12), 3787; https://doi.org/10.3390/ijms19123787
Submission received: 30 October 2018 / Revised: 18 November 2018 / Accepted: 23 November 2018 / Published: 28 November 2018
(This article belongs to the Special Issue The Effect of Dietary Factors on Cancer)

Abstract

:
Colorectal cancer is one of the leading causes of death, and the third most diagnosed type of cancer, worldwide. It is most common amongst men and women over 50 years old. Risk factors include smoking, alcohol, diet, physical inactivity, genetics, alterations in gut microbiota, and associated pathologies (diabetes, obesity, chronic inflammatory bowel diseases). This review will discuss, in detail, the chemopreventive properties of some dietary compounds (phenolic compounds, carotenoids, iridoids, nitrogen compounds, organosulfur compounds, phytosterols, essential oil compounds, polyunsaturated fatty acids and dietary fiber) against colorectal cancer. We present recent data, focusing on in vitro, laboratory animals and clinical trials with the previously mentioned compounds. The chemopreventive properties of the dietary compounds involve multiple molecular and biochemical mechanisms of action, such as inhibition of cell growth, inhibition of tumor initiation, inhibition of adhesion, migration and angiogenesis, apoptosis, interaction with gut microbiota, regulation of cellular signal transduction pathways and xenobiotic metabolizing enzymes, etc. Moreover, this review will also focus on the natural dietary compounds’ bioavailability, their synergistic protective effect, as well as the association with conventional therapy. Dietary natural compounds play a major role in colorectal chemoprevention and continuous research in this field is needed.

Graphical Abstract

1. Introduction

The alarming incidence of colorectal cancer (CRC) has led to a pressing demand in developing novel therapeutic strategies that could overcome the limitations of conventional therapies. Depending on the cancer stage, its treatment usually involves surgery, radiation and chemotherapy. Because CRC is generally diagnosed in the late stages, when patients frequently present with distant metastases, chemotherapy represents the backbone of CRC treatment [1,2]. Despite the extensive use of chemotherapeutical agents, such as fluorouracil, oxaliplatin or irinotecan, chemotherapy presents several disadvantages like severe toxicity associated with serious adverse reactions, development of drug resistance, and a lack of specificity in targeting solely tumor cells [3,4].
Reports that sustain the beneficial effects of natural compounds administration for a broad spectrum of diseases associated with the numerous sources available for compound isolation have opened an emerging interest in using natural compounds as novel therapeutic candidates for CRC treatment. Natural compound based CRC therapy can be used through the entire process of CRC management. A favorable diet plan can prevent the disease based on a high consumption of vegetables, fruits and fibers. It had been shown that natural compounds can target tumor cells after disease occurrence and prevent tumor recurrence or metastasis [5]. Moreover, one of the most important features of natural compounds remains their capacity to induce tumor cells sensitivity for chemotherapeutic agents after the development of drug resistance [6,7]. It has been reported that the dietary compounds, such as quercetin and curcumin, succeeded to overcome the multidrug resistance in several malignant cell lines [8,9,10]. This chemosensitivity potential of natural compounds favor their use as adjuvant therapy in conventional treatment protocols, but exclusively natural compound treatment can be also based on a different mechanism of action against tumor cells [8,11]. Therefore, natural compounds can exert anti-tumor effects due to their antioxidant capacity, ability to inhibit cellular growth and trigger tumor cell apoptosis or by modulating the metastatic cascade.

2. General Aspects of CRC

CRC is one of the leading causes of death and the third most diagnosed cancer in the world along with breast, prostate, lung, ovary, pancreas and bladder cancers [12,13]. The premalignant lesions associated with CRC are multiple aberrant crypt foci (AFC) which precedes the evolution of the adenomatous polyps [14,15]. Regarding the genomic instability, there are known three main pathways involved in the initiation and progression of CRC: (i) The chromosomal instability (CIN) pathway, (ii) global genome hypermethylation correlated with the shutdown of tumor suppressor genes, known as CpG island methylator phenotype (CIMP), and (iii) DNA microsatellite instability (MSI) phenotype [16,17,18].
Chronic inflammation, which is usually present in inflammatory bowel diseases (IBD), such as Chron’s disease or ulcerative colitis, is also a trigger for CRC development, especially for patients under 30 years. The molecular and cellular changes in chronic inflammation of the bowel may include alteration in cellular immunity, over-expression of cyclooxygenase (COX-2), activation of toll-like receptor 4 (TLR4), nucleotide-binding oligomerization domain-like receptors (NLRs), overexpression of vascular endothelial growth factor (VEGF) and genetic factors [19,20,21]. Inflammation in patients with CRC is associated with increased cellularity of Peyer’s patches, a lymphoid tissue associated with the small intestine [22].
It is well known that genetics, race, gut microbiota and environmental factors play a key role in the tremendous increase of CRC worldwide. Among genetic factors, mutations in the several genes (the homolog of Kirsten rat sarcoma proto-oncogene, KRAS; the homolog of rapidly accelerated fibrosarcoma proto-oncogene, BRAF) and chromosomal instability (manifested by frequent mutations) are involved in CRC etiology. Moreover, the polymorphism of several proteins is responsible for CRC incidence. In this category might be included nucleic acid binding protein 1 (NABP1, a protein responsible for DNA repair), laminin-1, cyclin D2, and transcription factors which target wingless-related integration site (Wnt)/β-catenin pathway [12,23,24,25].
In addition, there are patients that have a truly inherited predisposition for CRC development, those with familial adenomatous polyposis, Peutz-Jeghers syndrome, serrated polyposis syndrome and Lynch syndrome [26].
Gut microbiota is involved in tumor pathogenesis through the production of detrimental metabolites (polyamines, hydrogen sulfide and secondary bile acids), bacterially derived genotoxins, host defense modulation, inflammation, and oxidative stress [27]. Detrimental metabolites can cause direct DNA damage through DNA alkylation or induce chronic inflammation via interleukin 6 (IL-6) and tumor necrosis factor alpha (TNF-α) production [28,29]. Intestinal dysbiosis is common for CRC and it involves high levels of gut bacteria, such as Fusobacterium nucleatum, enterotoxigenic Bacteroides fragilis, adherent-invasive Escherichia coli, Peptostreptococcus anaerobius, Enterococcus faecalis, Bilophila wadsworthia, Solobacterium moorei and low levels of Bifidobacterium sp., Roseburia sp., Blautia sp., Lactobacillus sp. [29,30,31,32].
Intestinal dysbiosis, it is also characterized by defective production of short chain fatty acids, such as butyrate and acetic/propionic acids. Butyrate, produced by Eubacterium rectale and Roseburia sp., represents the energy substrate for epithelial cells of the colon, increases mucus production, raise the activity of antioxidant enzymes, and has anti-inflammatory properties. Acetic and propionic acids have been shown to inhibit the activation of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and to stimulate the apoptosis of tumor cells [19]. Pathogenic bacteria are responsible for a chronic intestinal inflammatory state via the recognition of microorganism-associated molecular pattern by Toll-like receptor 4 (TLR4), that induce activation of T helper 17 cells and production of pro-inflammatory interleukin 23 (IL-23) [28]. Moreover, these harmful bacteria are responsible for the increased activity of bacterial enzymes, such as β-glucuronidase, β-glucosidase, azoreductase, nitrate-reductase, alcohol dehydrogenase involved in converting heterocyclic aromatic amines, polycyclic aromatic hydrocarbons and primary bile acids into active carcinogens. In addition, harmful bacteria decrease the production of mucins by goblet cells and synthetize phenols, cresols and nitroso-compounds [19].
Environmental factors which favor the increased incidence of CRC include: Alcohol, smoking [33], consumption of red meat [34], associated pathologies, such as diabetes, overweight, and obesity [35], exposure to different chemicals (aromatic hydrocarbons), radiation [29,36], physical inactivity [37,38] and diet [38,39]. Red meat is a source of iron porphyrin pigment, which is responsible for the induction of carcinogenesis through the formation of nitroso-compounds [34]. Furthermore, the meat intake was related to CRC due to several factors: High fat-diet that might stimulate carcinogenesis through insulin resistance or fecal bile acids, heterocyclic amines produced by cooking the meat at high temperatures, N-nitroso compounds from processed meat, such as sausages, hot dogs, bacon [40,41].
Under these circumstances, it was reported that an increased intake of fruits and vegetables along with the limitation of a western diet (rich in fast food products and sweets), folic acid and vitamin D supplementation have a positive impact towards the reduction of CRC incidence [39,42].
Hormonal status also plays an important role in colon carcinogenesis mainly in women. Post-menopausal women have a high incidence of CRC, probably due to loss of cell cycle regulatory properties mediated by estrogen ERβ receptors or an increase in ERα receptors [43].
Screening remains the golden standard for preventing CRC and it is usually applied for people above 50 years. Screening methods include stool-based tests, visualization (colonoscopy) and serology/tissue markers. To reduce mortality and to improve the prognosis of CRC, earlier detection of serum markers is critical. The common biomarkers used in serological screening methods may include carcinoembryonic antigen (CEA), seprase, soluble cytokeratin-19 fragment (CYFRA 21-1), osteopontin (OPN), ferritin, anti-p53 [44,45].

3. General Aspects Regarding the Dietary Compounds

Healthy eating strategies, with increased consumption of fruits, vegetables, cereal grains, edible macrofungi and a balanced intake of meat and high fat products, have a beneficial role in the prevention of chronic diseases, such as heart disease, diabetes, cancer, neurodegenerative diseases or stroke [22,46]. Dietary phytochemicals can be divided into several categories, such as phenolic compounds, carotenoids, iridoids, nitrogen containing compounds, organosulfur compounds, phytosterols, dietary fiber, essential oils, polyunsaturated fatty acids [22,47].

3.1. Phenolic Compounds

Phenolic compounds are one of the widely distributed secondary metabolites in plants kingdom that share a common characteristic, the presence of at least one aromatic ring hydroxyl substituted. They usually bound to other molecules, such as glycosides and proteins and are well known for their powerful antioxidant properties [48]. Phenolic compounds can be divided according to their chemical structure into flavonoid and non-flavonoid constituents. Flavonoids are water-soluble pigments with a C6-C3-C6 skeleton, which have been subdivided in flavones, flavonols, flavanones, flavanols, isoflavones, and anthocyanidins [49]. Non-flavonoids compounds include tannins, phenolcarboxylic acids (derivatives of hydroxybenzoic and hydroxycinnamic acids), lignans, stilbens and other compounds. Phenolic compounds are widely distributed in fruits, vegetables, coffee, black tea, green tea and wine [50].

3.2. Carotenoids

Carotenoids are terpenic fat-soluble pigments containing 40 carbon atoms that share a polyisoprenoid structure, a long conjugated chain of double bonds and with a symmetry around the central bond [51]. According to their chemical structures, carotenoids are classified as carotens (α-carotene, β-carotene, lycopen—which are linear hydrocarbons that are cyclized at one end or both ends of the molecule) and oxygenated carotenoids (xantophylls), such as lutein, zeaxanthin [52]. Other important carotenoids are crocetin and its ester crocin. Dietary carotenoids are associated with low cardiovascular risk [53], eye health [54], prevention of infectious disease, anti-cancer and anti-inflammatory effects [51].

3.3. Iridoids

Iridoids are a type of monoterpenoids that have a cyclopentanopyran backbone, and are typically found in plants, such as glycosides. One of the most studied iridoids is oleuropein extracted from olives, which was known for its protective activity against cardiovascular and metabolic diseases [55].

3.4. Nitrogen Compounds

Nitrogen compounds are characterized by the presence of nitrogen atoms and are classified as alkaloids (a group of secondary plant metabolites, biosynthesized from amino acids) and non-alkaloid derivatives, such as protoalkaloids, pseudoalkaloids, alkamides, lectines, cyanogenic glycosides etc. [56]. Nitrogen compounds are characterized by great structural diversity and diverse pharmacological effects, such as analgesic, anti-inflammatory, antitussive, anticancer, antimalarial etc. [57].

3.5. Organosulfur Compounds

Organosulfur compounds contain at least one atom of sulfur in their molecule and are found mainly in garlic and cruciferous vegetables [58,59]. These compounds are well-known for their antifungal, antibacterian, anti-parasitic, antiviral, antithrombotic, immunomodulatory and anticancer effects [59].

3.6. Phytosterols

Phytosterols are plant derived lipid compounds, which resemble cholesterol, but vary in carbon side chains or presence/absence of a double bound. They are classified as sterols, (unsaturated compounds) and stanols (saturated molecules). Phytosterols are usually found in plants or macro fungi and share a wide range of therapeutic effects, such as hypolipidaemic, anti-inflammatory or anticancer effects [60].

3.7. Essential Oil Compounds

Essential oils are complex mixtures of chemical compounds with great structural diversity. Monoterpenes, sesquiterpenes and aromatic compounds are the main classes of bioactive substances found in essential oils. They have a wide spectrum of pharmacological effects, such as antibacterial, antifungal, antiviral, anti-mutagenic, cancer preventive and anti-inflammatory properties. Aromatic plants and spices (rosemary, fennel, coriander, thyme, cinnamon etc.) are excellent sources of essential oils [61,62].

3.8. Polyunsaturated Fatty Acids (PUFA)

Fatty acids are lipophilic compounds, for which the carbon chain can vary from two to 40 carbon atoms. They are essential components of membrane phospholipids. Fatty acids are usually classified as short chain (up to six carbon atoms), medium chain (eight-twelve carbons) and long chain (above 12 carbons). In addition, fatty acids are classified according to the presence or the absence of the double bonds in their molecules [63]. Long chain PUFA are categorized in omega-3 (n-3) and omega-6 (n-6) depending on the position of the first double bond from the methyl group of the fatty acid [64]. Humans are not able to synthesize the essential omega-3 fatty acids (α-linolenic acid) and omega-6 fatty acids (linoleic acid), so they depend exclusively on dietary sources [65]. Other omega-6, omega-3 fatty acids, such as γ-linolenic acid (GLA), di-homo-γ-linolenic acid (DHGLA), docosahexaenoic acid (DHA), eicosapentaenoic acid (EPA), arachidonic acid (AA) are generated during PUFA production, under the influence of different enzymes: ∆6 desaturase, elongase etc. [66]. The ratio between n-6/n-3 is very important for a balanced synthesis of eicosanoids [66,67] and the ideal ratio is 1/5–10 [66]. Recent findings support the role of PUFA in cardiovascular, neurodegenerative diseases or depression [64,68].

3.9. Dietary Fiber

Dietary fibers are found in fruits, vegetables, grains and macro fungi. They include non-starch polysaccharides, resistant starch fructo-oligosaccharides, galacto-oligosaccharides, glucomannans, lignin, pectin and β-glucans [63,69,70]. This large category of natural compounds has been classified according to their solubility, their quality of being fermentable and their physiological effects. Soluble fiber dissolves in water and form gels. They consist of inulin, pectin, gums and β-glucans. Insoluble fibers are represented by lignin, cellulose and some hemicelluloses [63]. Dietary fiber has various health benefits, such as lowering cardiovascular risk, glycemic and body weight control, laxation, prebiotic effect [63,69,71].
Some of the most well-known dietary compounds, belonging to different categories are presented in Table 1 (phenolic compounds) and Table 2 (non-phenolic compounds), including members, chemical structures and dietary sources.

4. In Vitro Studies

4.1. Polyphenols

The potential use of polyphenols for CRC prevention and treatment had been widely investigated (Figure 1). Research results highlight the potential of the dietary bioactive compounds to interfere with tumorigenesis at all steps tumorigenesis, including initiation, promotion and progression. In vitro mechanisms of dietary polyphenols demonstrated their ability to modulate cellular processes, such as, gene expression, apoptosis or differentiation [80]. Several recent studies have demonstrated that cancer treatment through a combinatorial approach might show increased efficiency than the use of drugs only [81].

4.1.1. Flavones

Quercetin is a flavonol present in vegetables and fruits like as onion or apples. Quercetin reduces proliferation in CRC cell lines by downregulating RAS p21 protein activator 1 gene [82]. The combination of quercetin with tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) proved a TRAIL-induced apoptosis in a synergistic manner in colorectal adenocarcinoma cells (SW-620 and HT-29 cell lines), while this combination resulted in an additive effect in the case of Caco-2 human epithelial colorectal adenocarcinoma cells [83]. Quercetin has been also tested in vitro in combination with fluorouracil (5-FU) in CO115 human colon carcinoma cells (p53 positive) and HCT15 colorectal adenocarcinoma cells (p53 negative). This combination of drugs showed higher apoptosis levels in CO115 cell line, in a synergistic manner, but as an additive effect in HCT15 cells [84].
Kaempferol can be found in black tea, broccoli, propolis or grapefruit acts on different types of cancer cells [85] inducing apoptosis through cytochrome c mitochondrial release, caspase-3 cleavage activation and p53-dependent growth inhibition [86,87]. Cho et al. showed in 2012 that kaempferol may induce G1 and G2/M cell cycle arrest by inhibiting the activity of cyclin-dependent kinase 2, 4 (CDK 2, 4) and cell division cycle protein 2 [88]. In HT-29 cells, kaempferol induces apoptosis and inhibits insulin-like growth factor 1 receptor (IGF-1R) and receptor tyrosine kinase ErbB3 signaling pathways [89]. Kaempherol was tested in combination with TRAIL on SW480 and DLD-1 CRC cell lines and the results showed an increase in apoptotic induction in a kaempferol-dose dependent manner, probably by up-regulating death receptor-5 [90].
Apigenin is one of the well-known flavones which is present in several fruits and vegetables, such as parsley, garlic, Chinese cabbage, celery, bell pepper, and guava [72]. It is a chemopreventive agent with strong cytostatic and anti-angiogenic effects in vitro [72]. In CRC cells, the activity of apigenin was correlated with a blockage in cell cycle progression, induction of apoptosis and inhibition of the cell growth [91,92,93]. However, to increase the anti-neoplastic activity of apigenin additional combination therapy with drugs, such as 5-FU, oxaliplatin or irinotecan, might be required.

4.1.2. Isoflavones

Genistein is an isoflavone, which can be found in high concentrations in soybeans, lentils, beans, and chickpeas. A negative correlation was reported between the soybean diet and the occurrence of the CRC [94,95]. This isoflavone aroused a growing interest as a pro-apoptotic agent because of succeeded to be more effective in CRC cells compared to their normal counterparts [96]. Genistein acts by increasing the expression of B cell lymphoma 2-associated X protein (Bax) or cyclin-dependent kinase inhibitor, p21 [97], by inhibiting Nuclear factor kappa-light-chain-enhancer of activated B cells/NF-κB [98] and topoisomerase II [99]. In addition, genistein displayed anti-cancer activities by regulating ErbB proteins expression [100] and by suppressing the carcinogen induction of Wnt/β-catenin signaling pathway [101]. In addition, genistein was tested together with 5-FU on HT-29 cells and the results showed a synergistic effect on cell growth blocking, probably due to the over-expression of pro-apoptotic p53 and p21 and down-regulation of COX-2 [102].

4.1.3. Phenocarboxilic Acids

Caffeic acid showed a pro-apoptotic effect on HT-29 cells [103,104] starting from 80 μM concentration, while chlorogenic acid [105] did not show any significant activity against different human colorectal carcinomas [106]. Ferulic acid displayed inhibition of CRC progression acting on cells adhesion and migration mechanisms [103]. Gallic acid is the only one among benzoic acid derivatives that acts on CRC cells by upregulating Bax and downregulating Bcl-2 [107,108].

4.1.4. Stilbens

Resveratrol is a phytoalexin found in many plant species, including edible plants like grapes or berries. Resveratrol is synthesized by plants in response to environmental stress and pathogenic invasion as a natural inhibitor of cell proliferation [81]. The first study revealing the resveratrol’s chemopreventive activity was published by Jang et al. in 1977, who described the effect of the topical administration of resveratrol [109]. Since then, resveratrol was used in many studies for various malignancies, including CRC. In vitro studies demonstrated that resveratrol might modulate and inhibit tumorigenesis by reduction of the COX-2 and cytochrome P450 activity. Moreover, resveratrol can exert synergistic activities in combination with other active compounds, such as quercetin or chemotherapeutic agents, such as 5-FU, oxaliplatin or mitomycin C [110].
Consequently, the combination of resveratrol and quercetin stimulated apoptosis in HT-29 cell line, by reducing the RNA and proteins levels for several transcription factors [111]. This combination of the natural compounds suppressed human colon cancer stem cells proliferation and down-regulated c-Myc [112]. The toxicity of 5-FU was increased in CRC cell lines by synergistic activity of resveratrol. The mechanism of action of the combined treatment included an increase in the oxidative stress associated with inhibition of Akt and STAT3 pathways [113]. In addition, resveratrol together with 5-FU significantly induced apoptosis and reduced migration in CRC cells [114].
Other recent studies show that resveratrol chemosensitizes HT-29 and HCT-116 cells to oxaliplatin by upregulating miR-34c [115]. In addition, resveratrol and oxaliplatin synergistically inhibited cell growth of Caco-2 cells via apoptosis and necrosis induction [116]. At the same time, resveratrol can induce p21WAF1/CIP1 overexpression regardless of p53 status, and a combined treatment of resveratrol and mitomycin C repressed the proliferation of mitomycin C-resistant CRC cells [117].

4.1.5. Other Compounds

Curcumin, a diarylheptanoid found in turmeric, was identified as a chemopreventive dietary compound in CRC [118,119,120]. Curcumin showed a synergistic effect with dasatanib, a potent inhibitor of Src and Abl kinases, in CRC cell lines with resistant phenotype to FOLFOX (folinic acid, fluorouracil, oxaliplatin) chemotherapy [120]. The anti-neoplastic effect of 5-FU was reported to be increased by curcumin [121,122]. Curcumin can potentiate, as well the pro-apoptotic and anti-metastatic effects of capecitabine [123].

4.2. Non-polyphenolic Compounds—In Vitro Mechanism of Action

Non-polyphenolic compounds display chemopreventive effects on CRC cells in vitro modulating signaling pathways with impact on cell cycle arrest, apoptosis, invasion, inflammation and much more (Figure 2).

4.2.1. Carotenoids

Even if the benefits of a lycopene diet in CRC prevention have not been clearly demonstrated [124], lycopene remains a potential chemopreventive agent for CRC management due to its excellent antioxidant capacity [125]. Lycopene inhibits the cellular growth of HT-29 cells in a dose-dependent manner by effectively inhibiting the phosphorylation of Protein kinase B (Akt) and therefore silencing Phosphatidylinositol-4,5-bisphosphate 3-kinase (PIK3)/Akt, a signaling pathway that is associated with colorectal tumor development [126]. Moreover, lycopene suppresses tumor cell invasion mediated by leptin as proteomic assays revealed that lycopene exposure significantly decreased expression levels of MMP7 and thereby reduce the tumor cells invasion capacity [127]. Tanga et al. highlighted a synergistic mechanism between lycopene and eicosapentaenoic acid that significantly increased expression of proapoptotic molecules Bax and Fas. Eicosapentaenoid acid and lycopene reduce tumor cell growth also by suppression of PIK3/Akt pathway and by further blocking the activation of downstream molecule mTOR [128]. Another promising effect of lycopene use as a chemopreventive agent was based on its anti-inflammatory potential against different inflammation-related proteins associated with CRC development [129]. In SW480 CRC cell culture, lycopene modulates the inflammatory cascade by inhibiting the protein expression of NF-κB and JNK. In human CRC cells SW480, after the stimulation with lipopolysaccharides, the level of inflammatory cytokines (TNF-α, IL-1 and IL-6) as well the inflammation-inducing enzymes (COX-2 and iNOS) was reduced when treated with lycopene. Consequently, this compound shows promising effects for modulating the NF-κB signaling pathway in CRC associated inflammation [130].
β-carotene significantly reduce in vitro cell proliferation of CRC lines LS180, SW620 and HCT-15 and its cytotoxic potential can be easily tailored by adjusting different experimental parameters, such as dose, exposure time and cell seeding density [131]. However, the cytotoxic potential of β-carotene was strongly affected by the cell capacity to incorporate the carotenoid as different cell lines present various patterns of growth inhibition in presence of the same concentration of compound [132]. Briviba et al. associated β-carotene exposure of HT-29 cells with induction of apoptosis through a mechanism independent expression of the MEK/ERK signaling pathway as the expression of extracellular signal regulated kinases ERK1 and ERK2 was not affected [133]. The anti-cancer potential of β-carotene was attributed to cell cycle arrest and apoptosis induction by modulating the expression of different key regulator proteins. In the presence of β-carotene, COLO320HSR cells exhibit cell cycle arrest in G2/M phase mediated by the down-regulation of cyclin A expression and apoptotic death correlated with decreased levels of apoptotic inhibitors Bcl-2 and Bcl-xL [134]. β-carotene modulates also the apoptotic pathway mediated by Bcl-2 through its pro-oxidant activity, as it is capable to stimulate ROS production in CRC lines [135].
Crocetin is a promising chemopreventive agent for CRC mainly due to its antioxidant and anti-inflammatory potential [136]. Crocetin extracts have proven to be effective in inhibiting the cellular growth of HT-29, HCT-116 and SW480 CRC cell lines in a dose-dependent manner [137]. Even if several mechanisms of action have been described for crocetin use in different types of cancer, the role of crocetin in CRC development inhibition has not been fully elucidated [138,139]. Crocetin promotes apoptosis via p53 dependent or independent mechanisms, increasing the attraction of crocetin use in defective p53 tumors [140,141]. Moreover, HTC-116 colon cancer cells treated with crocetin presented a significant reduction in the expression of inflammation related genes like IL-6, IL-8 and High mobility group box 1 protein (HMGB1), associated with an increased expression of NAG-1 gene that encodes a protein with high anti-tumorigenic activity [142].

4.2.2. Nitrogen Compounds

Piperine, the most common dietary alkaloid was intensively used in clinics based on its strong antioxidant and anti-inflammatory capacity, but various studies revealed the great potential of using this natural compound as an active anti-neoplastic agent. Piperine specifically inhibits tumor cells proliferation and arrest cell cycle in G1 phase as revealed after HT-29 cells exposure to piperine. The cell cycle lock is commonly associated with a decreased expression of cyclins D1 and D3, up-regulation of p21WAF1 and p27KIP1 expression and a diminished phosphorylation of Rb protein [143]. Piperine mediated apoptosis can be activated as a result of ROS production increase [144] or by modulating mTORC1 signaling cascade as Caco-2 and HT-29 cells exposure to piperine is associated with inhibition of mTORC1, a key regulator of cellular autophagy [145].
Capsaicin is a promising natural agent in CRC therapy as various favorable effects on different CRC cell lines have been reported regarding the capacity of capsaicin to suppress tumor cell expansion [146,147,148,149]. However, in vitro capsaicin administrated doses need to be tightly adjusted since low concentrations of capsaicin promote metastasis by interfering with key molecules involved in the metastatic cascade. At low concentrations, capsaicin favors EMT transition by inhibiting E-cadherin expression, sustains tumor cell migration by inducing over-expression of MMP2 and MMP9 and activates Akt/mTOR signaling pathway, molecular events that together enhance the migration and invasive potential of SW480 cells. In contrast, at proper concentrations capsaicin is effective in reducing the metastatic burden by inhibiting overproduction of MMPs and the epithelial mesenchymal transition [150]. One of the main molecular mechanisms of action that underlies the role of capsaicin in cell proliferation inhibition is cell cycle arrest and subsequent induction of apoptosis mainly by modulation of p53 activity and other apoptosis linked molecules [151]. Colorectal tumor apoptotic cells are characterized by AMP-activated protein kinase (AMPK) signaling cascade activation [152], activation of the pro-apoptotic caspase-3 [149] and an altered expression pattern of β-catenin and transcription factor 4 (TCF-4) that blocks their interaction [153]. Apoptosis can be orchestrated by NO levels that trigger apoptotic pathways dependent or independent of p53 activity or by ROS levels that generate an impairment of the mitochondrial membrane potential [149,154]. Additionally, capsaicin also exhibits an efficient immunomodulatory effect as reveled by a dose-dependent decrease of numerous inflammatory cytokines after capsaicin treatment of HT-29 and RKO tumor cells [155]. Capsaicin presents a superior anti-tumor effect in combination with 3,3′-diindolylmethane by modulating the transcriptional activity of p53 and NF-κB together with other apoptosis related genes [156].
In conclusion, recent in vitro studies reveal that polyphenols act on colon cancer cells by inducing apoptosis and by inhibiting cells growth, migration, adhesion and tumor initiation through Wnt/β-catenin signaling pathway. Non-polyphenolic compounds induce apoptosis and inhibit cell growth, inflammation, angiogenesis and tumor initiation through PIK3/Akt signaling pathway. Considering the evidence provided by the in vitro studies regarding the mechanisms of action of the dietary polyphenols and non-polyphenolic compounds in CRC cells, further investigations should focus on underlying their efficacy and safety use in combination with chemotherapy and/or radiotherapy.

5. In Vivo Studies in CRC Animal Models

5.1. The Effect of Phenolic Compounds

The summary of the effects of phenolic compounds is presented in Table 3.

5.1.1. Isoflavones

In vivo studies revealed that genistein treated high-fat mice with CRC induced by administration of azoxymethane/dextran sulfat sodium decreased the expression of inflammatory factors [157]. Genistein has also shown tumor suppressive activity in mice colon cancer through apoptotic effects, reduction of tumor weight and reduction of wingless-related integration site (Wnt) signaling [101,158]. Genistein was tested on BALB/c mice bearing and CT26 xenografts under radiotherapy and the results showed less non-tumorigenic apoptotic cells and improved morphological changes in healthy intestinal tissue [158]. More than 70% of patients with CRC undergoing radiotherapy displayed side effects at the gastrointestinal level. Normal colonocytes might be frequently damaged mainly as a result of an increased oxidative stress induced by radiation therapy, resulting in inflammatory or ulcerative lesions [110]. Administration of natural compounds as radioprotective mediators might represent a possible solution to avoid the occurrence of the mucosities, as side effects of the ionizing radiation therapy [172]. Xenografts’ mice with colon cancer tumors are treated with 5–10 Gy, and genistein was administrated one day prior to irradiation. The authors reported that genistein protected against intestinal injury induced by radiation therapy in mice [158].

5.1.2. Anthocyanidins

Regarding administration of anthocyanidins, cyanidin from tart cherries significantly reduced adenomas of the cecum [159]. Shi N. and coworkers (2015) investigated the effect of dietary lyophilized strawberries, which are known to be rich in cyanidin and pelargonidin glycosides, on colon carcinogenesis induced by administration of azoxymethane. The results are promising with inhibition of tumor development associated with reduced phosphorylation of PI3K/Akt, declining levels of NF-κB and decreased expression of pro-inflammatory markers [160]. Recent researches have shown that administration of grape juice concentrate rich in phenolic compounds (peonidin glucoside, malvidin-glucoside) reduced inflammation, through decreased expression of COX-2, if taken before or after induction of colon cancer [161]. Other authors reported that administration of anthocyanin-rich diet in an animal model of CRC showed significant reduction of colon tumors and a positive effect on gut microbiota [162].

5.1.3. Phenocarboxilic Acids

Phenolcarboxylic acids, chlorogenic acid showed chemopreventive effects in a mouse model of CRC [163]. Banerjee N. and coworkers demonstrated that administration of a plum beverage rich in phenolic compounds (chlorogenic and neochlorogenic acid) decreased the number of aberrant crypt foci, modulated the activity of several intracellular payhways and reduced inflammation [164]. Cinnamic acid upregulated the expression of protein Bax, a pro-apoptotic protein, in colorectal xenografts grown in athymic mice, probably by inhibition of histone deacetylases [165]. Gallic acid, rosmarinic acid and p-coumaric acid have a chemopreventive effect colon carcinogenesis induced by 1,2-dimethyl hydrazine induced. The phenolcarboxylic acids decreased tumor incidence, induced the inhibition of precancerous lesions, reduced gut bacterial enzymes (mucinase, nitroreductase, sulphatase, β- glucosidase) and displayed antioxidant properties [166,167,168]. It is well known that phase I metabolizing enzymes are required for activation of different carcinogens into forms capable of binding to proteins or DNA and leading to mutations. An increase in phase II metabolizing activity has been related to the elimination of carcinogens, through the formation of hydrophilic compounds, such as glucoronides, glutathione conjugates or glutathione sulphates [167]. Moreover, the phenolcarboxylic acids significant increased phase II xenobiotic metabolizing enzymes (glutathione S-transferase, gamma-glutamyl-transpeptidase) and decreased phase I xenobiotic metabolizing enzymes, such as cytochrome P450 and cytochrome b5 [166,167,168].

5.1.4. Lignans

Lignans have strong anticancer effects against CRC; administration of a rich secoisolariciresinol extract from flaxseeds, reduced cancer biomarker levels and decreased the number of the proliferating cells [169]. Moreover, deflated flaxseeds reduced the incidence of precancerous lesions in the proximal, distal and middle colon. The exact mechanism of action is not known, since no changes are observed in tumor protein p53 (p53), cyclin-dependent kinase inhibitor 1 (p21CIP1/WAF1) and multiple tumor suppressor 1 (p16) expression and the authors did not notice an increase in fecal short chain fatty acids [170]. Sesamol, a lignin from sesame seeds, significantly reduced the number of intestinal polyps in mice with colorectal cancer, through suppression of COX-2 and prostaglandin E2 receptor expression levels [171].

5.2. The Effect of Non-Phenolic Compounds

The summary of in vivo experiments regarding CRC and non-phenolic compounds is presented in Table 4.

5.2.1. Carotenoids

The effect of carotenoids (lycopene, crocin) on CRC was observed using animal models. Results are encouraging, since both dietary products showed chemopreventive effects and reduced the incidence of pre-neoplasic polyps, through various mechanisms, such as up-regulation of p21CIP1/WAF1 proteins, suppression of proliferating cell nuclear antigen (PCNA) expression and nuclear levels of β-catenin in tumor tissues. Moreover, an anti-inflammatory effect has been seen through suppression of COX-2 and prostaglandin E2 (PGE2) gene expression. Lycopene also inhibits matrix metallopeptidase protein (MMP-9) in correlation with reduction of angiogenesis, tumor invasion and metastasis. Crocin reduced the chronic inflammation by inhibition of the activated nuclear factor NF-κB and displayed a protective role regarding the toxicity of xenobiotics by increasing the level of Nrf2, nuclear factor (erythroid-derived 2)-like 2 [173,174,175].

5.2.2. Iridoids

Oleuropein, an iridoid found mainly in olive leaves and fruits, can prevent CRC by regulation of Wnt/β-catenin, NF-κB, PI3K/Akt pathways, anti-inflammatory activity associated with significant decreased the intestinal concentrations of several interleukins (IL-6, IL-17), TNF-α and inhibited COX-2 activity. Besides oleuropein up-regulated Bax protein expression and induced apoptosis in intestinal tumor cells [176].

5.2.3. Nitrogen Compounds

Pre-clinical studies have demonstrated that capsaicin may be a chemopreventive agent in CRC through anti-proliferative, anti-genotoxic, induced expression of apoptotic genes, and up-regulation of genes involved in cell differentiation [177]. Since previous reports indicated that capsaicin might promote metastasis, acting as co-carcinogen in cancer skin [196] and promoting breast cells metastasis [197]. Yang et al., investigated the correlation between capsaicin and metastasis in a tumor xenografts mouse model for CRC [150]. To investigate the ability of capsaicin to induce metastasis CT-26 murine CRC cell previous treated or not with 100 μM are intravenously injected in mice. The investigation of the pulmonary metastatic nodules in 15 days after injection validated the previous reports. Namely, the capsacin-treated cells increased the number of pulmonary metastatic nodules in mouse models [150]. Under these circumstances, caution should be taken when various doses of capsaicin are used for designing in vivo experiments or clinical trials, if recommended.

5.2.4. Phytosterols

Among phytosterols, β-sitosterol has proved antioxidant activity in 1,2 dimethylhydrazine (DMH) colon carcinogenesis and restored endogenous antioxidant enzyme levels, such as superoxide dismutase, catalase, glutathione reductase, glutathione peroxidase and vitamins, such as vitamin E and C [179].

5.2.5. Organosulfur Compounds

Down-regulation of histone deacetylase (HDAC) enzymes play a key role in cell cycle arrest and apoptosis, whilst up-regulation of Nrf2 expression is involved in the increased activity of phase II metabolizing enzymes [198,199]. Sulforaphane, from broccoli, inhibits CRC carcinogenesis by modulating Nrf2 activity and inhibition of HDAC enzymes [178].

5.2.6. Essential Oils

Former data about the administration of essential oil compounds, mainly thymoquinone to male albino WISTAR rats with chemical induction of CRC reported a reduction in tumor incidence, volume and multiplicity and decrease in vascular endothelial growth factor (VEFG) concentration [180]. In addition, the association between thymoquinone and mesalazine, an anti-inflammatory drug, significantly reduced tumor incidence and multiplicity in transgenic mice [181]. Recent data reported that cinnamaldehyde, the main compound of cinnamon essential oil, induces Nrf2 activation in CRC tissues, without any toxicity [182]. Carvacrol is a monoterpenic phenol that occurs in essential oils of the Labiatae family (Thymus sp., Origanum sp., Satureja sp.). Administration of carvacrol to adult male albino rats with chemical induced CRC, revealed that carvacrol reduced the number of colon tumors, dysplastic polyps, aberrant crypts foci, increased glutathione peroxidase and glutathione reductase activities and restored levels of liver peroxidation products [183]. Moreover, combined treatment between carvacrol and X-radiation significantly decreased tumor incidence [200].

5.2.7. Polyunsaturated Fatty Acids

Dietary feeding of polyunsaturated omega-3 fatty acids in C57 BL6 mice with CRC induced by inoculation of MC38 murine colon adenocarcinoma cells significantly inhibited the tumor growth. The possible molecular mechanisms included the augmentation in the levels of omega-3 metabolites (epoxydocosapentaenoic acids), suppression of inflammation through the reduction of the level of the pro-inflammatory agents (arachidonic acid eicosanoids), and decrement in the expression of human retroviral correspondent identified in myelocytomatosis, MYC protooncogene [184]. Nevertheless, more recent data reported that a ketogenic diet (low carbohydrates and high fat diet) with or without omega-3 fatty acids supplementation delayed tumor growth and suppressed tumor neovascularization [185]. Supplementation of diet with eicosapentaenoic acid in C57BL/6J mice with CRC induced by administration of azoxymethane suppressed tumor growth, increased apoptosis, and decreased systemic inflammation. In addition, supplementation with eicosapentaenoic acid increased Notch-1 signaling pathway activity, which is involved in apoptosis, in particular in the initiation phase [186]. In transgenic C57BL/6J mice, carrying a mutation for the adenomatous polyposis coli (APC) gene, supplementation of diet with olive oil and salmon oil (rich in omega-3 fatty acids) inhibited the cancer development. The effects of oil diet are evaluated according to the level of signal transduction and activator of transcription 3 (p-STAT3), a transcription factor involved in up-regulation of anti-apoptotic genes and the level of fatty acid synthase, a key enzyme in neoplastic lipogenesis activities. The diet rich in omega-3 fatty acids inhibited the development of the malignancy by the reduction in the levels of STAT-3 and fatty acid synthase [187].

5.2.8. Dietary Fiber

Regarding dietary fiber, a polysaccharide from Lentinus edodes mushroom suppressed tumor growth in nude mice, upregulated caspase-3, -9 activity, increased Bax/Bcl2 ratio, increased the generation of free radicals in tumor tissues and TNF-α production [188]. Masuda Y. and co-workers demonstrated that a soluble β-glucan from Grifola frondosa inhibits tumor growth in murine cancer models through induced systemic tumor-antigen specific T cell response, increased activity of T-cells in tumor and decreased number of tumors caused immunosuppressive cells [189]. Supplementation of diet with prebiotics, such as inulin, in Sprague Dawley rats with azoxymethane induced CRC showed a significant increase in Lactobacillus sp. and Bifidobacteria sp. biomass along with a reduction in E. coli activity. In addition, the authors reported a reduction in microbial enzyme (β-glucuronidase, nitroreductase, azoreducatse) activity along with a reduction in aberrant crypt foci formation [190]. Besides, a combination of inulin with galacto-oligosaccharides significantly reduced β-glucosidase, nitroreducatse, azoreductase activity and enhanced short-chain fatty acid production [191]. Recent data reported that inulin decreased COX-2 and NF-κB expression in the colon with a significant reduction of inflammation markers [192]. Reduced inflammatory response in the jejuna and colon mucosa was also observed for co-administration of inulin and Lactobacillus plantarum in rats exposed to dimethylhydrazine [194].
Experiments performed with Sprague Dawley rats with 1,2 dimethylhydrazine induced CRC, showed that inulin significantly decreased the activity of colon enzymes (β-glucosidase, β-glucuronidase) and the effect was more pronounced compared to lactulose, a non-absorbable sugar [193]. Association between water-soluble polysacharide (glucomannan) and prebiotics (inulin), before initiation of CRC with azoxymethane (AOM) in a mouse model up-regulated gene expression of antioxidant enzymes, such as glutathione peroxidase 2 (GPX2), glutathione-S-transferase (GST) and catalase (CAT) and increased short fatty acid chain fatty acids production [195].
In conclusion, administration of dietary compounds to laboratory animals with induced CRC has indicated beneficial effects on different stages of carcinogenesis (Table 4). Several mechanisms are involved: Inhibition of tumor growth, modulation of several pathways (Wnt/β catenin, PI3K/Akt, Notch-1), anti-inflammatory activity, up-regulation of antioxidant enzymes and NRf2 expression, down-regulation of microbial enzymes activity, activation of caspases or increased Bax/Bcl-2 ratio.

6. Chemoprevention of CRC by Dietary Compounds in Humans

6.1. Phenolic Compounds

6.1.1. Isoflavones

Keeping in mind the important role of healthy diet (the consumption of fruits, vegetables, cereals) for CRC prevention [201], we would like to consider the role of phenolic and non-phenolic compounds in chemoprevention of CRC in humans. An epidemiological case-control study performed on Japanese patients (721 cases and 697 control subjects) between 2004–2008 revealed an inverse association between dietary isoflavone intake (as tofu or miso soup) and the risk of colorectal carcinoma in both men and women [202]. According to another case control study performed in Korea, which investigated the effects of isoflavone intake (for 901 cancer cases and 2669 controls) for CRC prevention, the highest intake of soy products and isoflavones was associated with a significantly reduced CRC risk for both men and women. The observed effects are more pronounced for rectal (in women) and distal (in men) colon cancers. A significantly reduced CRC risk was also observed for post-menopausal women [203]. Reports from a Spanish case control study (The Bellvitge CRC study) showed that exposure to isoflavones was inversely related to CRC risk. Isoflavones protect against colorectal carcinogenesis through their estrogenic properties and cause up-regulation of estrogen ERβ receptors. Moreover, isoflavones inhibit the development of CRC via up-regulation of protein p21 expression, decrease in the expression of the proliferating cell antigen (PCNA), decrease in the extracellular signal regulated kinase (ERK), Akt and nuclear factor NF-κB expression [204]. Plasma phytoestrogens were strongly linked with decreased incidence of the CRC. Analysis of plasma phytoestrogens levels in Korean and Vietnamese patients, revealed an inverse correlation between plasma genistein concentration and CRC [205]. Furthermore, genistein has shown anti-oncogenic action by increasing the expression of antioxidant enzymes [206], and by preventing CRC metastases [207].

6.1.2. Lignans

According to earlier reports association between insoluble fibers, flaxseed dry extract (with 20% secoisolariciresinol) and milk thistle extract (with 30% silibinin, the major active constituent of sylimarin) in patients with sporadic colonic adenomas, significantly increased ERβ proteins and ERβ/ERα ratio. Moreover, increased ERβ/ERα ratio was associated with pro-aptotic effects, such as activation of caspase-3 and caspase-8 activity [208]. Another report that evaluated oral supplementation with a mixture of sylimarin and secoisolariciresinol diglucoside in patients with familial adenomatous polyposis revealed a significant reduction in the number and the size of the polyps. The observed effect was the consequence of lignans on ERβ selective agonist activity [209]. In a case control study in Korean and Vietnamese population (2003–2007), the authors did not find a correlation regarding enterolactone (metabolite of lignans) levels [205].
The summary of the association between CRC risk and phenolic/non-phenolic compounds is presented in Table 5.

6.1.3. Anthocyanidins

Administration of a bilberry dry extract (mirtocyan), with a 36% anthocyanin concentration (represented by cyanidin-3-galactoside, delphinidin-3-galactoside, delphinidin-3-arabinoside, cyanidin-3-glucoside) in patients with CRC conducted to decreased proliferation of cancer cells and reduction of the proliferation marker Ki-67 level [212]. Recent research has shown that administration (orally as a powder or as rectal suppositories) of a black raspberry freeze-dried extract rich in anthocyanins (cyanidin-3-glucoside, cyanidin-3-sambubioside, cyaniding-3-rutinoside, cyanidin-3-xylorutinoside) to patients with familial adenomatous polyposis (FAP), significantly decreased the burden of polyps. Several mechanisms are involved: Decrease of the DNA (cytosine-5)-methyltransferase 1 (DNMT1) which was correlated with reduced methylation of p16 protein and regulatory genes in the Wnt signaling pathway [213].

6.2. Non-phenolic Compounds

6.2.1. Organosulfur Compounds

In a double-blind placebo control study, using 51 patients, which are diagnosed with colorectal adenomas using colonoscopy, administration of aged garlic extract, especially at high dose significantly reduced the number and size of colon adenoma after 12 months [214]. Aged garlic extract is a processed garlic product, for which allicin is transformed into more stable compounds, such as S-allylmercaptocysteine [214]. However, other authors did not find a strong support regarding garlic chemopreventive effects, even after seven years of follow up [215]. Although, preclinical data support garlic use for CRC, a clinical study performed on both women and men for up to 24 years did not support garlic intake (even > 1 serving/week) or garlic supplementation for CRC chemoprevention [216]. Contradictory results are also reported by other authors, which even found a significant increase of colorectal incidence with 35% at a five-year follow up [217]. The contradictory effects regarding garlic chemopreventive role, are probably due to heterogeneity of studies regarding the assessment of garlic intake, type of garlic and pathology (CRC, adenoma etc). A meta-analysis of observational studies regarding the consumption of cruciferous vegetables (cabbage, broccoli) revealed an inverse correlation between dietary intake and CRC risk [236].

6.2.2. Carotenoids

A large European case control study showed an inverse correlation between dietary concentration of β-carotene, vitamin C and CRC incidence, mainly in the distal colon [219]. Kabat C. G. and co-workers have also found an inverse correlation between β-carotene plasma levels and CRC incidence in post-menopausal women [220].

6.2.3. Polyunsaturated Fatty Acids

In a phase II double-blind randomized clinical trial for patients with CRC and liver metastasis, pre-operative administration of EPA for a median of 30 days conducted to increased levels of EPA and decreased level of PGE2 in tumor tissues compared to control tissue. Moreover, EPA also showed anti-angiogenic activity, whilst preoperative treatment with omega-3 fatty acids had postoperative benefit, in the first 18 months (regarding cancer recurrence). This effect may be the consequence of EPA prolonged plasma biological half-life [221]. According to a prospective study is US men and women, intake of marine omega-3 fatty acids and fish are inversely associated with rectal cancer incidence [222]. In a Japanese based prospective study, administration of omega-3 fatty acids reduced the incidence of proximal colon cancer [223].
According to earlier studies, administration of encapsulated fish oil to patients with CRC, undergoing chemotherapy, significantly reduced C reactive protein (CRP) and CRC/albumin ratio compared to controls. Increased values of CRP and CRP/albumin ratio are associated with a chronic inflammatory state and poor clinical prognosis for CRC patients. Administration of fish oil also increased EPA, DHA and decreased arachidonic acid plasma levels [224]. It is well known that polyunsaturated fatty acids have anti-inflammatory effects; a meta-analysis of clinical trials with CRC patients (that included administration of fish oil or supplementation of diet with omega-3 fatty acids) showed a significant decrease of IL-6 plasma levels and an increase for albumin concentration. For patients undergoing chemotherapy supplementation with EPA and DHA significantly reduced CRP/albumin ratio [225]. Other authors reported that administration of an enriched omega-3 nutritional supplement significantly increased EPA, DHA and decreased arachidonic acid concentration in granulocytes. These findings suggest that omega 3 fatty acids might exert an immuno-stimulatory effect [226].
However, other studies failed to demonstrate a beneficial role of omega-3 fatty acids for CRC patients. According to Ma C.J. and coworkers, administration of a lipid emulsion composed of soybean oil, medium chain triglycerides and polyunsaturated fatty acids after surgery of CRC patients did not improve the level of the inflammatory markers, instead had a positive effect on lipid profile [227]. In addition, subjects with a high intake of fish and marine food appeared to have an increased risk of distal colon cancer. These differences might be the consequence of high pH values in the distal colon (due to low production of short chain fatty acids), that could attenuate omega-3 fatty acids effects or the presence of contaminants in fish (lead, mercury etc.) [222].

6.2.4. Dietary Fiber

Recent research has shown that supplementation with agave inulin in healthy patients has a beneficial role upon gut microbiota. Administration of the agave inulin resulted in significant increase of Bifidobacterium genus. A significant decrease was observed for Rumincoccus sp. and Desulfovibrio sp., which might have a benefit towards CRC incidence [228]. Still other authors did not find a correlation between inulin intake and CRC risk. In a randomized phase II chemoprevention trial, with patients with aberrant crypt foci (ACF) the administration of inulin did not show a significant reduction in ACF number compared to control [229].
A high fibers intake is associated with low risk of CRC. In a prospective cohort study, intake of cereal-derived fibers was associated with a low risk of Fusobacterium nucleatum positive CRC. Recent findings have shown that a high amount of F. nucleatum in tumor tissues was linked to cancer severity and high mortality [230]. A meta-analysis of studies regarding the benefit of dietary fiber consumption, have also shown an inverse correlation between fiber intake and risk of colorectal adenoma [231,232]. According to Kunzmann A. and co-workers a diet rich in fibers significantly reduced CRC risk. Nonetheless, the association was stronger for males than females [233]. In another multicenter prospective cohort study, a high intake of dietary fibers was significantly associated with low risk of distal/proximal colon and rectum cancers [234]. Navarro S. and co-workers investigated the effect of dietary fiber and omega-3, -6 fatty acids in a women’s health initiative prospective cohort during their 11.7-year follow up. The results pointed out a reduced incidence of CRC for the association between a low dose of soluble fiber, a high dose of insoluble fiber and a high dose of EPA and DHA [237]. However, results from other studies regarding the benefit role of dietary fiber for CRC prevention showed contradictory results. In a randomized control trial, administration of resistant starch, which exerts similar effects as dietary fiber in patients with Lynch syndrome, a hereditary non-polyposis colon cancer, over 29 months, had no detectable effect on cancer development [235].
In conclusion, most of the presented clinical trials have shown an inverse correlation between dietary intake of carotenoids, phytoestrogens (lignans, isoflavones), polyunsaturated fatty acids, fibers and CRC incidence (Table 5). However, results should be carefully interpreted keeping in mind the individual variability, the number of dietary compounds consumed by patients and the complexity of CRC.

7. Bioavailability of the Natural Dietary Compounds

Regularly, when discussing the health value of the dietary compounds, these are evaluated according to their ability to be released from the ingested food, be absorbed in the gastrointestinal tract, undergo the metabolic steps and reach the target tissue. In the case of CRC, the last steps are overcome, since the interaction between dietary compounds and cancer cells is more direct. At the intestinal lumen, the lipophilic compounds can pass into enterocytes by facilitated diffusion, while the hydrophilic compounds (such as polyphenols) can enter the cells after the aglycone is liberated through the enzymatic hydrolysis [238]. However, the bioavailability can be reduced by the presence of ABC transporters (ATP-Binding Cassette) or other multidrug mechanisms, which increase the efflux of xenobiotic [3]. In CRC cells this effect might be reversed by the administration of the flavonoids from Citrus sp. plants [239]. With the aim of understanding the beneficial properties of the dietary compounds, their bioavailability is required to be studied (Table 6).
The plasma concentration of the dietary compounds is another key topic to discuss. Although experiments performed in cancer cells are important for developing new therapeutic agents, differences occur when the same substance is administrated in animal models or used in clinical trials due to its bioavailability, metabolisms or interaction with gut microbiota. A series of extensive reviews presented by Manach et al., regarding the bioavailability of the polyphenols, introduced the average values of the polyphenols in plasma at a short time after ingestion [240]. The concentration values of the plasma polyphenols vary in range of low micromolar levels. For instance, after the ingestion of ~100 mg quercetin from apples the plasma level of quercetin reached 0.3 μM at 2.5 h after consumption, while ingestion of ~125 mg hesperitine from orange juice reached 2.2 μM at 5.4 h after consumption. Similar plasma concentration levels are identified for epigallocatechin-3-O-gallate (EGCG) from green tee, namely 0.16–0.96 μM after ingestion of 200–800 mg EGCG or genistein from soymilk, i.e., 1.14 μM subsequently to the consumption of 0.59 mg/kg [49,72,240,241]. Several strategies, such as modulation of the administration, inhibition of the metabolic pathways, administration in combination with other drugs or encapsulation methods attempts to improve the bioavailability of the dietary compounds [120,238]. The bioavailability of the phenolic compounds was extensive studied [120,238,242,243,244], while little is known about the bioavailability of dietary non-phenolic products. The maximum concentration of ellagitannis in 1 h after the intake of 318 mg ellagitannins from pomegranate juice was about 0.06 μM [245]. The non-alkaloid capsacin administrated to in rats (30 mg/kg body weight) was detected in serum after 1 h in ranges of 1.9 ± 1.2 μM and the concentration levels decreased within time [246]. An estimation of 50 g of olive oil/day uptake leads to the detection of ~13 μM hydroxytyrosol (a metabolite of oleuropein obtained after the hydrolysis of oleuropein-aglycone) in plasma, a concentration much lower than 50–100 μM required for the antioxidant effect. The authors concluded that the olive oil compounds are well absorbed, but the plasma level of their metabolites are too low to induce substantial biological effects [247]. Further research in this field is needed to analyze and improve the bioavailability of non-phenolic compounds in order to obtain significant effects on human health.

7.1. Bioavailability of Phenolic Compounds

Studies demonstrated the small bioavailability of curcumin was due to poor absorption, rapid metabolism and rapid elimination. In order to increase curcumin bioavailability it was associated with other compounds, for instance piperine, the main component of black pepper, which increased its bioavailability by 2000% [248]. Another method, by which it has been demonstrated to obtain a better bioavailability of curcumin, was the method of using nanocurcumin, a polymeric nanoparticle encapsulated formulation of curcumin. Nanocurcumin has shown a better activity in cancer cell lines by inhibition of NF-κB [259]. There are studies in vivo made in both humans and animals, to determine the bioavailability of curcumin. Experiments on rats showed a maximum serum concentration (Cmax) of 1.35 ± 0.23 μg/mL in 0.83 h compared to humans where Cmax in the same period was much lower consisting in 0.006 ± 0.005 μg/mL [120].
Studies made with EGCG both in vivo and in vitro are conducted to show the systemic absorption of the most abundant catechin found in the green tea. Derliz Mereles and Werner Hunstein proved that for optimizing EGCG bioavailability there must be taken into consideration the pharmacokinetic parameters that can diminish or enhance the bioavailability. Thereby among the factors that improve the absorption of EGCG are alkaloids like piperine, vitamin C, proteins like albumin or even fish oil. On the other hand, the absorption of EGCG can be reduced by air contact oxidation, metal ions like Ca2+ and Mg2+, temperature or beverages like milk. In vivo studies conducted on human subjects, after administration of one oral dose of EGCG with a fasting period overnight, Cmax was reached in 1–2 h and the elimination half-life of EGCG was at 3.4 ± 0.3 h [249]. Thereby the matter of bioavailability is a major concern for all the scientists who are trying to find a better way for natural compounds to be assimilated and to keep their effects powerful.
Beside curcumin and EGCG, the natural compound resveratrol is also considered to have a poor solubility and bioavailability [260]. Thus, studies on resveratrol showed that a better bioavailability can be achieved by combining resveratrol with other foods, beverages or even with other polyphenols. Calvo-Castro, L. A. et al. demonstrated that the impact of trans-resveratrol, on a group of six men and women, was increased by a liquid micellar formulation of a grapevine-shoot extract, which contains a high amount of resveratrol monomers and oligomers [250]. The experiment demonstrated that a single dose of 500 mg of grapevine-shoot extract (30 mg trans-resveratrol, 75 mg trans-ε-viniferin) influenced the trans-resveratrol concentrations with a Cmax by 10.6-fold higher with no detection of trans-ε-viniferin nor in plasma or urine [250].
Quercetin a flavonoid recognized for the powerful effect in cancer cell lines was also studied for its bioavailability. According to Khan, F. et al. a study conducted on rodents demonstrated that after an administration of quercetin intravenously, after plasma assays are performed, no quercetin was found [261]. These results might be explained by the fact that is difficult to assess the intracellular effect of the compound at the organ site and nowadays is generally accepted the correlation between the plasma concentration of the compound and its therapeutic action [262]. A high quantity of quercetin is found in onion, which is considered to be more effective than quercetin supplement regarding its bioavailability, presumably for the increased intestinal absorption of the food matrix [263]. It has been demonstrated that the absorption of quercetin might depend on the type of sugar residue attached to quercetin. For instance, quercetin glicosides found in onion are better absorbed unlike the major quercetin glycoside in tea. The assessment of quercetin glucoside absorption was between 3%−17% in healthy subjects, after a dose of 100 mg which is relatively low due to poor absorption, extensive metabolism or rapid elimination [251].
Another flavonoid, worth to be taken into consideration is genistein found in soy and is well known for its beneficial properties, including multiple molecular effects, such as the promotion of apoptosis, anti-inflammatory properties, modulation of metabolic pathways and steroidal hormone receptors [264]. For a better understanding of how genistein can work as a chemopreventive agent, it is required to know its bioavailability. Pharmacokinetic studies demonstrated the low oral bioavailability of genistein, whilst the plasma or tissue concentrations are reduced compared to in vitro values of half maximal inhibitory concentration, IC50. Yang. Z. et al. demonstrated in their study that after intravenous and oral administration of 20 mg/kg genistein in FVB mice, genistein was transformed mostly to glucuronosides and sulfates and the genistein aglycone bioavailability was 23.4% [252]. Regarding the oral bioavailability of genistein there are studies that showed a better bioavailability of total genistein in mice than genistein aglycone. After feeding female Balb/c mice with soy protein, Yang Z. et al. demonstrated a difference between bioavailability of total genistein, which was almost 90%, and genistein aglycone, less than 15% [252]. In addition, it should be taken into account supplementary factors which can contribute to poor absorption of genistein, such as age, gender or dose frequency, but the most important factor is related to absorption, distribution, metabolism and excretion (ADME) properties [252].
Anthocyanins are a group of molecules that belong to flavonoid family and they are found in a large group of fruits, flowers, roots and leaves responsible for the blue, purple and red color [265]. In contradiction to other flavonoids, the bioavailability of anthocyanins (delphinidin, malvidin, cyaniding and pelargonidin) is rather different. These molecules can be absorbed in the stomach or intestines. In the gastroinstestinal wall, it may be found the pure form of cyanidin-3-glucoside anthocyanin and pelargonidin-3-glucoside, which can be decomposed by microbiota right after reaching to the large intestine [265]. In a pilot study, Muller D. et al. demonstrated that anthocyanins could reach the small intestine within one hour and the level of the compound at the intestinal area was about 30–50% of the ingested substance, while the plasma level was very low [253]. Anthocyanins can be methylated due to the metabolic transformation carried out by the enzymes, such as cathechol-O-methyltransferase (COMT), thereby the anticancer effect of anthocyanins may be limited. However, there are inhibitors that are able to decrease the methylation of polyphenols [266].
Proanthocyanidins are phytochemicals represented by a group of flavonoids which are found in a variety of plants and aliments, such as apples, cinnamon, grapes, cranberry, green tea etc. [267]. However, the main downside of proanthocyanidins is regarding its bioavailability. Multiple studies are conducted to elucidate the bioavailability issue of this class of oligomeric flavonoids. While the monomeric flavonoids are rather absorbed in the small intestine, the metabolism of oligomeric and polymeric proanthocyanidins in the colon, is not so much understood. In that matter Choy Y. et al. conducted a study to investigate the presence of proanthocyanidins in the colon after ingestion of grape seed extract [254]. The evidence from this study suggests the presence of proanthocyanidins in the colon as intact compounds which might be beneficial for maintaining a healthy digestive system [254].

7.2. Bioavailability of Non-Phenolic Compounds

Capsaicin is one of the most pungent ingredients consumed worldwide and is naturally found in chili peppers. Capsaicin is known to possess many beneficial effects on the human body, such as anti-inflammatory, antimicrobial, anticancer and it was described to work as a topical analgesic. Beside capsaicin, capsaicinoids are also represented by dihydrocapsaicin (DHC), nordihydrocapsaicin (n-DHC), homocapsaicin (h-C) and homodihydrocapsaicin (h-DHC). However, capsaicin low bioavailability is a concern for restricting its application [268]. Rollysons, W.D. et al. conducted a study in vivo to explore the intestinal absorption of capsaicin, using lab rats to isolate intestinal sacs. The capsaicin was absorbed into intestinal tissues, jejunum and serosa fluid [255]. For instance, the absorption of the compound was differently regarding the intestinal region of interest. Accordingly, 1 mM of capsaicin was absorbed in a proportion of 50% in the stomach, 80% in the jejunum and 70% in the ileum [255]. Kawada et al. conducted a study on WISTAR rats are the effect of capsaicin and DHC are absorbed in the stomach and small intestine in a percentage of 85% [246].
Piperine belongs to Piperaceae family and is one of the most important alkaloids found in black pepper (Piper nigrum) and in long pepper (Piperum longum). Because of its anticancer effect, it is mandatory to have knowledge about the potential of absorption and the bioavailability of piperine. It is well known that piperine is insoluble in water and presents a low bioavailability, which may limit its use in clinical experiments. Nevertheless, piperine can be used in clinical assays single or in addition to other dietary agents. Several studies demonstrated the ability of piperine to act as a bioavailability enhancer for many chemopreventive agents, such as resveratrol, leading to increased levels of revesterol in plasma [269]. Johnson J. J et al. demonstrated that the administration of resveratrol alone or in combination with piperine in vivo, enhanced the serum bioavailability of resveratrol by almost 229% [256]. In addition, piperine was known to enhance the effect of curcumin and lycopene because of its ability to inhibit intestinal and hepatic glucoronidation. For instance, piperine enhanced the bioavailability of curcumin by 2000%, in both rats and human experiments [257].
Allicin is derived from isothiocyanate phytochemical and is a compound found in garlic. Among its properties are included antifungal, anti-neoplastic and antibacterian effects [270]. However, these properties are questionable because of the poor bioavailability of allicin. According to Lawson et al. after oral administration of garlic and pure allicin, there was no detection of it neither in urine nor in blood [258].
In conclusion, the oral bioavailability of dietary agents was variable depending on each compound of interest, being able to be increased or decreased by other agents, vitamins, proteins according to each experiment. Further studies should be carried out to have an objective vision regarding the ability of the natural compounds to help in chemoprevention.

7.3. Encapsulation Strategies for Increased Bioavailability

Since the absorption of the dietary compounds, after their oral administration, might by restricted by the insufficient gastric residence time, low permeability, low solubility, instability during food processing (pH, enzymes, presence of other nutrients) the health benefits of these compounds are limited. During the last decades the nanoparticles are investigated due to their ability to transport and deliver drugs. Several advantages might recommend the nanoparticles as potent delivery agents: Increased bioavailability, reduced toxicity and improved solubility in aqueous medium [271]. Furthermore, the nanoparticle can accumulate in the solid tumors as a result of the reduced lymphatic drainage [272]. In this view, many anticancer drug delivery systems are developed based on nanoscale strategy, by using nanoparticles of different compositions [273]. “Yallapu et al. demonstrated that curcumin loaded PLGA nanoparticles display an increased effect on metastatic cancer cells than curcumin alone [274], while Radu et al. developed a poly alkanoate nanocarrier for silymarin delivery with good drug release and biocompatibility properties [275]. Despite its extensive in vitro study, the drugs encapsulation approach is still under development and only a few clinical trials are currently running [276].
The most common carrier agents used in micro- and/or nano-encapsulation of dietary bioactive compounds are (i) polysaccharides, such as: Starch [277], dextrins [278], maltodextrins [279], cyclodextrins [280], (ii) celulloses, such as: Carboxymethyl cellulose, methylcellulose [281] and cellulose ethers [282], (iii) pectins [283], (iv) chitosan [284] etc. To obtain these micro- and/or nano-encapsulations several techniques are developed, of which spray drying, freeze drying, complex coacervation, emulsification, anti-solvent precipitation, extrusion, electro-spinning, layer-by-layer deposition and solid dispersion are the most frequently used.

8. Conclusions

Natural compounds can exert anti-tumor effects due to their antioxidant capacity and their ability to inhibit cellular growth, capacity to trigger tumor cells apoptosis or to modulate the metastatic cascade. The use of natural bioactive compounds could minimize chemotherapy and/or radiotherapy side effects, such as neutropenia, diarrhea, cardiotoxicity, nephrotoxicity, hepatotoxicity, etc.
Both in vitro and in vivo studies demonstrated that the administration of dietary active compounds induces growth inhibition, apoptosis and inhibition of adhesion and migration. Furthermore, they exert anti-inflammatory effects and modulate Wnt/β catenin, PI3K/Akt and Notch-1 key pathways for tumor initiation.
Surprisingly, some clinical trials presented in this review have shown an inverse correlation between dietary intake of carotenoids, phytoestrogens (lignans, isoflavones), polyunsaturated fatty acids, fibers and CRC incidence, but these results should be carefully interpreted considering the individual variability, co-morbidities and cancer development status. Consequently, further investigations should focus on underlying dietary compounds efficacy and the safety use in combination with chemotherapy and/or radiotherapy, as well as their bioavailability (intestinal absorption and metabolism).
With this respect, the oral bioavailability of the dietary agents was shown to be variable depending on each compound and to be influenced by other agents. However, most of the studies conclude that generally, the dietary bioactive compounds show low bioavailability. Therefore, the development of new strategies to increase their bioavailability and adjust their administration doses would be of great interest. A promising approach in this view is the development of nano-/micro-shuttles able to carry active bio-compounds and to release these molecules in a controlled manner.

Author Contributions

T.C., O.-A.C., C.G.: Writing the manuscript, preparing the tables, critical revising of the manuscript; A.H., B.G., O.G., M.C.: Writing the manuscript, preparation of the figures, critical revising of the manuscript; M.-M.M.: Concept, drafting, editing and critical revising of the manuscript, supervising the manuscript preparation.

Funding

This research was funded by the Romanian Ministry of Education and by the UEFISCDI, grant number PN-III-P2-2.1-PTE-2016-0149/19PTE – TUMFLOW.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van der Stok, E.P.; Spaander, M.C.W.; Grunhagen, D.J.; Verhoef, C.; Kuipers, E.J. Surveillance after curative treatment for colorectal cancer. Nat. Rev. Clin. Oncol. 2017, 14, 297–315. [Google Scholar] [CrossRef] [PubMed]
  2. Van Cutsem, E.; Cervantes, A.; Adam, R.; Sobrero, A.; Van Krieken, J.H.; Aderka, D.; Aranda Aguilar, E.; Bardelli, A.; Benson, A.; Bodoky, G.; et al. ESMO consensus guidelines for the management of patients with metastatic colorectal cancer. Ann. Oncol. 2016, 27, 1386–1422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hammond, W.A.; Swaika, A.; Mody, K. Pharmacologic resistance in colorectal cancer: A review. Ther. Adv. Med. Oncol. 2016, 8, 57–84. [Google Scholar] [CrossRef] [PubMed]
  4. Braun, M.S.; Seymour, M.T. Balancing the efficacy and toxicity of chemotherapy in colorectal cancer. Ther. Adv. Med. Oncol. 2011, 3, 43–52. [Google Scholar] [CrossRef] [PubMed]
  5. Fernandez, J.; Redondo-Blanco, S.; Gutierrez-del-Rio, I.; Miguelez, E.M.; Villar, C.J.; Lombo, F. Colon microbiota fermentation of dietary prebiotics towards short—Chain fatty acids and their roles as anti-inflammatory and antitumor agents: A review. J. Funct. Foods 2016, 25, 511–522. [Google Scholar] [CrossRef]
  6. Aung, T.N.; Qu, Z.; Kortschak, R.D.; Adelson, D.L. Understanding the Effectiveness of Natural Compound Mixtures in Cancer through Their Molecular Mode of Action. Int. J. Mol. Sci. 2017, 18, 656. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, P.; Yang, H.L.; Yang, Y.J.; Wang, L.; Lee, S.C. Overcome Cancer Cell Drug Resistance Using Natural Products. Evid.-Based Complement. Altern. Med. ECAM 2015, 2015, 1–14. [Google Scholar] [CrossRef] [PubMed]
  8. Chai, S.; To, K.K.; Lin, G. Circumvention of multi-drug resistance of cancer cells by Chinese herbal medicines. Chin. Med. 2010, 5, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Tang, X.Q.; Bi, H.; Feng, J.Q.; Cao, J.G. Effect of curcumin on multidrug resistance in resistant human gastric carcinoma cell line SGC7901/VCR. Acta Pharmacol. Sin. 2005, 26, 1009–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Ganta, S.; Amiji, M. Coadministration of Paclitaxel and curcumin in nanoemulsion formulations to overcome multidrug resistance in tumor cells. Mol. Pharm. 2009, 6, 928–939. [Google Scholar] [CrossRef] [PubMed]
  11. Housman, G.; Byler, S.; Heerboth, S.; Lapinska, K.; Longacre, M.; Snyder, N.; Sarkar, S. Drug resistance in cancer: An overview. Cancers 2014, 6, 1769–1792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Witold, K.; Anna, K.; Maciej, T.; Jakub, J. Adenomas-Genetic factors in colorectal cancer prevention. Rep. Pract. Oncol. Radiother. 2018, 23, 75–83. [Google Scholar] [CrossRef] [PubMed]
  13. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Dyba, T.; Randi, G.; Bettio, M.; Gavin, A.; Visser, O.; Bray, F. Cancer incidence and mortality patterns in Europe: Estimates for 40 countries and 25 major cancers in 2018. Eur. J. Cancer 2018, 103, 356–387. [Google Scholar] [CrossRef] [PubMed]
  14. Sakai, E.; Nakajima, A.; Kaneda, A. Accumulation of aberrant DNA methylation during colorectal cancer development. World J. Gastroenterol. 2014, 20, 978–987. [Google Scholar] [CrossRef] [PubMed]
  15. Oines, M.; Helsingen, L.M.; Bretthauer, M.; Emilsson, L. Epidemiology and risk factors of colorectal polyps. Best Pract. Res. Clin. Gastroenterol. 2017, 31, 419–424. [Google Scholar] [CrossRef] [PubMed]
  16. Colussi, D.; Brandi, G.; Bazzoli, F.; Ricciardiello, L. Molecular pathways involved in colorectal cancer: Implications for disease behavior and prevention. Int. J. Mol. Sci. 2013, 14, 16365–16385. [Google Scholar] [CrossRef] [PubMed]
  17. Nazemalhosseini Mojarad, E.; Kuppen, P.J.; Aghdaei, H.A.; Zali, M.R. The CpG island methylator phenotype (CIMP) in colorectal cancer. Gastroenterol. Hepatol. Bed Bench 2013, 6, 120–128. [Google Scholar] [PubMed]
  18. Worthley, D.L.; Leggett, B.A. Colorectal cancer: Molecular features and clinical opportunities. Clin. Biochem. Rev. 2010, 31, 31–38. [Google Scholar] [PubMed]
  19. Dos Reis, S.A.; da Conceicao, L.L.; Siqueira, N.P.; Rosa, D.D.; da Silva, L.L.; Peluzio, M.D. Review of the mechanisms of probiotic actions in the prevention of colorectal cancer. Nutr. Res. 2017, 37, 1–19. [Google Scholar] [CrossRef] [PubMed]
  20. Hu, T.; Li, L.; Shen, J.; Zhang, L.; Cho, C. Chronic Inflammation and Colorectal Cancer: The Role of Vascular Endothelial Growth Factor. Curr. Pharm. Des. 2015, 21, 2960–2967. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, E.R.; Chang, D.K. Colorectal cancer in inflammatory bowel disease: The risk, pathogenesis, prevention and diagnosis. World J. Gastroenterol. 2014, 20, 9872–9881. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, Z.; Cao, A.T.; Cong, Y. Microbiota regulation of inflammatory bowel disease and colorectal cancer. Semin. Cancer Biol. 2013, 23, 543–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Peters, U.; Jiao, S.; Schumacher, F.R.; Hutter, C.M.; Aragaki, A.K.; Baron, J.A.; Berndt, S.I.; Bezieau, S.; Brenner, H.; Butterbach, K.; et al. Identification of Genetic Susceptibility Loci for Colorectal Tumors in a Genome-Wide Meta-analysis. Gastroenterology 2013, 144, 799–807.e724. [Google Scholar] [CrossRef] [PubMed]
  24. Coppede, F.; Lopomo, A.; Spisni, R.; Migliore, L. Genetic and epigenetic biomarkers for diagnosis, prognosis and treatment of colorectal cancer. World J. Gastroenterol. 2014, 20, 943–956. [Google Scholar] [CrossRef] [PubMed]
  25. Shen, Y.; Wang, J.; Han, X.; Yang, H.; Wang, S.; Lin, D.; Shi, Y. Effectors of epidermal growth factor receptor pathway: The genetic profiling ofKRAS, BRAF, PIK3CA, NRAS mutations in colorectal cancer characteristics and personalized medicine. PLoS ONE 2013, 8, e81628. [Google Scholar] [CrossRef] [PubMed]
  26. Bogaert, J.; Prenen, H. Molecular genetics of colorectal cancer. Ann. Gastroenterol. 2014, 27, 9–14. [Google Scholar] [PubMed]
  27. Gagniere, J.; Raisch, J.; Veziant, J.; Barnich, N.; Bonnet, R.; Buc, E.; Bringer, M.A.; Pezet, D.; Bonnet, M. Gut microbiota imbalance and colorectal cancer. World J. Gastroenterol. 2016, 22, 501–518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Louis, P.; Hold, G.L.; Flint, H.J. The gut microbiota, bacterial metabolites and colorectal cancer. Nat. Rev. Microbiol. 2014, 12, 661–672. [Google Scholar] [CrossRef] [PubMed]
  29. Nistal, E.; Fernandez-Fernandez, N.; Vivas, S.; Olcoz, J.L. Factors Determining Colorectal Cancer: The Role of the Intestinal Microbiota. Front. Oncol. 2015, 5, 220. [Google Scholar] [CrossRef] [PubMed]
  30. Rezasoltani, S.; Asadzadeh Aghdaei, H.; Dabiri, H.; Akhavan Sepahi, A.; Modarressi, M.H.; Nazemalhosseini Mojarad, E. The association between fecal microbiota and different types of colorectal polyp as precursors of colorectal cancer. Microb. Pathog. 2018, 124, 244–249. [Google Scholar] [CrossRef] [PubMed]
  31. Niederreiter, L.; Adolph, T.E.; Tilg, H. Food, microbiome and colorectal cancer. Dig. Liver Dis. 2018, 50, 647–652. [Google Scholar] [CrossRef] [PubMed]
  32. Zackular, J.P.; Rogers, M.A.; Ruffin, M.T.T.; Schloss, P.D. The human gut microbiome as a screening tool for colorectal cancer. Cancer Prev. Res. 2014, 7, 1112–1121. [Google Scholar] [CrossRef] [PubMed]
  33. Fagunwa, I.O.; Loughrey, M.B.; Coleman, H.G. Alcohol, smoking and the risk of premalignant and malignant colorectal neoplasms. Best Pract. Res. Clin. Gastroenterol. 2017, 31, 561–568. [Google Scholar] [CrossRef] [PubMed]
  34. Kruger, C.; Zhou, Y. Red meat and colon cancer: A review of mechanistic evidence for heme in the context of risk assessment methodology. Food Chem. Toxicol. 2018, 118, 131–153. [Google Scholar] [CrossRef] [PubMed]
  35. Voutsadakis, I.A. Obesity and diabetes as prognostic factors in patients with colorectal cancer. Diabetes Metab. Syndr. 2017, 11 (Suppl. 1), S109–S114. [Google Scholar] [CrossRef] [PubMed]
  36. Talibov, M.; Sormunen, J.; Hansen, J.; Kjaerheim, K.; Martinsen, J.-I.; Sparen, P.; Tryggvadottir, L.; Weiderpass, E.; Pukkala, E. Benzene exposure at workplace and risk of colorectal cancer in four Nordic countries. Cancer Epidemiol. 2018, 55, 156–161. [Google Scholar] [CrossRef] [PubMed]
  37. Patel, P.; De, P. Trends in colorectal cancer incidence and related lifestyle risk factors in 15–49-year-olds in Canada, 1969–2010. Cancer Epidemiol. 2016, 42, 90–100. [Google Scholar] [CrossRef] [PubMed]
  38. Grimmett, C.; Simon, A.; Lawson, V.; Wardle, J. Diet and physical activity intervention in colorectal cancer survivors: A feasibility study. Eur. J. Oncol. Nurs. 2015, 19, 1–6. [Google Scholar] [CrossRef] [PubMed]
  39. Tayyem, R.F.; Bawadi, H.A.; Shehadah, I.; Agraib, L.M.; AbuMweis, S.S.; Al-Jaberi, T.; Al-Nusairr, M.; Bani-Hani, K.E.; Heath, D.D. Dietary patterns and colorectal cancer. Clin. Nutr. 2017, 36, 848–852. [Google Scholar] [CrossRef] [PubMed]
  40. Santarelli, R.L.; Pierre, F.; Corpet, D.E. Processed meat and colorectal cancer: A review of epidemiologic and experimental evidence. Nutr. Cancer 2008, 60, 131–144. [Google Scholar] [CrossRef] [PubMed]
  41. Bouvard, V.; Loomis, D.; Guyton, K.Z.; Grosse, Y.; Ghissassi, F.E.; Benbrahim-Tallaa, L.; Guha, N.; Mattock, H.; Straif, K. Carcinogenicity of consumption of red and processed meat. Lancet Oncol. 2015, 16, 1599–1600. [Google Scholar] [CrossRef] [Green Version]
  42. Pericleous, M.; Mandair, D.; Caplin, M.E. Diet and supplements and their impact on colorectal cancer. J. Gastrointest. Oncol. 2013, 4, 409–423. [Google Scholar] [PubMed]
  43. Barzi, A.; Lenz, A.M.; Labonte, M.J.; Lenz, H.J. Molecular pathways: Estrogen pathway in colorectal cancer. Clin. Cancer Res. 2013, 19, 5842–5848. [Google Scholar] [CrossRef] [PubMed]
  44. Wild, N.; Andres, H.; Rollinger, W.; Krause, F.; Dilba, P.; Tacke, M.; Karl, J. A combination of serum markers for the early detection of colorectal cancer. Clin. Cancer Res. 2010, 16, 6111–6121. [Google Scholar] [CrossRef] [PubMed]
  45. Berretta, M.; Alessandrini, L.; De Divitiis, C.; Nasti, G.; Lleshi, A.; Di Francia, R.; Facchini, G.; Cavaliere, C.; Buonerba, C.; Canzonieri, V. Serum and tissue markers in colorectal cancer: State of art. Crit. Rev. Oncol. Hematol. 2017, 111, 103–116. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Y.J.; Gan, R.Y.; Li, S.; Zhou, Y.; Li, A.N.; Xu, D.P.; Li, H.B. Antioxidant Phytochemicals for the Prevention and Treatment of Chronic Diseases. Molecules 2015, 20, 21138–21156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Dzialo, M.; Mierziak, J.; Korzun, U.; Preisner, M.; Szopa, J.; Kulma, A. The Potential of Plant Phenolics in Prevention and Therapy of Skin Disorders. Int. J. Mol. Sci. 2016, 17, 160. [Google Scholar] [CrossRef] [PubMed]
  48. Martins, S.; Mussatto, S.I.; Martinez-Avila, G.; Montanez-Saenz, J.; Aguilar, C.N.; Teixeira, J.A. Bioactive phenolic compounds: Production and extraction by solid-state fermentation. A review. Biotechnol. Adv. 2011, 29, 365–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Manach, C.; Donovan, J.L. Pharmacokinetics and metabolism of dietary flavonoids in humans. Free Radic. Res. 2004, 38, 771–785. [Google Scholar] [CrossRef] [PubMed]
  50. Reis Giada, M.D.L. Food Phenolic Compounds: Main Classes, Sources and Their Antioxidant Power. In Oxidative Stress and Chronic Degenerative Diseases—A Role for Antioxidants; Croatia InTech: Rijeka, Croatia, 2013. [Google Scholar] [CrossRef] [Green Version]
  51. Milani, A.; Basirnejad, M.; Shahbazi, S.; Bolhassani, A. Carotenoids: Biochemistry, pharmacology and treatment. Br. J. Pharmacol. 2017, 174, 1290–1324. [Google Scholar] [CrossRef] [PubMed]
  52. Mezzomo, N.; Ferreira, S.R.S. Carotenoids Functionality, Sources, and Processing by Supercritical Technology: A Review. J. Chem. 2016, 2016, 1–16. [Google Scholar] [CrossRef]
  53. Wang, Y.; Chung, S.J.; McCullough, M.L.; Song, W.O.; Fernandez, M.L.; Koo, S.I.; Chun, O.K. Dietary carotenoids are associated with cardiovascular disease risk biomarkers mediated by serum carotenoid concentrations. J. Nutr. 2014, 144, 1067–1074. [Google Scholar] [CrossRef] [PubMed]
  54. Abdel-Aal el, S.M.; Akhtar, H.; Zaheer, K.; Ali, R. Dietary sources of lutein and zeaxanthin carotenoids and their role in eye health. Nutrients 2013, 5, 1169–1185. [Google Scholar] [CrossRef] [PubMed]
  55. Bulotta, S.; Celano, M.; Lepore, S.M.; Montalcini, T.; Pujia, A.; Russo, D. Beneficial effects of the olive oil phenolic components oleuropein and hydroxytyrosol: Focus on protection against cardiovascular and metabolic diseases. J. Transl. Med. 2014, 12, 219. [Google Scholar] [CrossRef] [PubMed]
  56. Irchhaiya, R.; Kumar, A.; 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. 2014, 4, 287–305. [Google Scholar]
  57. Kabera, J.N.; Semana, E.; Mussa, A.R.; Xin, H. Plant Secondary Metabolites: Biosynthesis, Classification, Function and Pharmacological Properties. J. Pharm. Pharmacol. 2014, 2, 377–392. [Google Scholar]
  58. Gonzalez-Vallinas, M.; Gonzalez-Castejon, M.; Rodriguez-Casado, A.; Ramirez de Molina, A. Dietary phytochemicals in cancer prevention and therapy: A complementary approach with promising perspectives. Nutr. Rev. 2013, 71, 585–599. [Google Scholar] [CrossRef] [PubMed]
  59. Mikaili, P.; Maadirad, S.; Moloudizargari, M.; Aghajanshakeri, S.; Sarahroodi, S. Therapeutic uses and pharmacological properties of garlic, shallot, and their biologically active compounds. Iran. J. Basic Med. Sci. 2013, 16, 1031–1048. [Google Scholar] [PubMed]
  60. Ogbe, R.J.; Ochalefu, D.O.; Mafulul, S.G.; Olaniru, O.B. A review on dietary phytosterols: Their occurrence, metabolism and health benefits. Asian J. Plant Sci. Res. 2015, 5, 10–21. [Google Scholar]
  61. Bayala, B.; Bassole, I.H.; Gnoula, C.; Nebie, R.; Yonli, A.; Morel, L.; Figueredo, G.; Nikiema, J.B.; Lobaccaro, J.M.; Simpore, J. Chemical composition, antioxidant, anti-inflammatory and anti-proliferative activities of essential oils of plants from Burkina Faso. PLoS ONE 2014, 9, e92122. [Google Scholar] [CrossRef] [PubMed]
  62. Raut, J.S.; Karuppayil, S.M. A status review on the medicinal properties of essential oils. Ind. Crops Prod. 2014, 62, 250–264. [Google Scholar] [CrossRef]
  63. Slavin, J. Fiber and prebiotics: Mechanisms and health benefits. Nutrients 2013, 5, 1417–1435. [Google Scholar] [CrossRef] [PubMed]
  64. Zarate, R.; El Jaber-Vazdekis, N.; Tejera, N.; Perez, J.A.; Rodriguez, C. Significance of long chain polyunsaturated fatty acids in human health. Clin. Transl. Med. 2017, 6, 25. [Google Scholar] [CrossRef] [PubMed]
  65. Ibarguren, M.; Lopez, D.J.; Escriba, P.V. The effect of natural and synthetic fatty acids on membrane structure, microdomain organization, cellular functions and human health. Biochim. Biophys. Acta 2014, 1838, 1518–1528. [Google Scholar] [CrossRef] [PubMed]
  66. Abedi, E.; Sahari, M.A. Long-chain polyunsaturated fatty acid sources and evaluation of their nutritional and functional properties. Food Sci. Nutr. 2014, 2, 443–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Li, F.; Duan, Y.; Li, Y.; Tang, Y.; Geng, M.; Oladele, O.A.; Kim, S.W.; Yin, Y. Effects of dietary n-6:n-3 PUFA ratio on fatty acid composition, free amino acid profile and gene expression of transporters in finishing pigs. Br. J. Nutr. 2015, 113, 739–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Song, C.; Shieh, C.H.; Wu, Y.S.; Kalueff, A.; Gaikwad, S.; Su, K.P. The role of omega-3 polyunsaturated fatty acids eicosapentaenoic and docosahexaenoic acids in the treatment of major depression and Alzheimer’s disease: Acting separately or synergistically? Prog. Lipid Res. 2016, 62, 41–54. [Google Scholar] [CrossRef] [PubMed]
  69. Mudgil, D.; Barak, S. Composition, properties and health benefits of indigestible carbohydrate polymers as dietary fiber: A review. Int. J. Biol. Macromol. 2013, 61, 1–6. [Google Scholar] [CrossRef] [PubMed]
  70. Dahl, W.J.; Stewart, M.L. Position of the Academy of Nutrition and Dietetics: Health Implications of Dietary Fiber. J. Acad. Nutr. Diet. 2015, 115, 1861–1870. [Google Scholar] [CrossRef] [PubMed]
  71. Otles, S.; Ozgoz, S. Health effects of dietary fiber. Acta Sci. Pol. Technol. Aliment. 2014, 13, 191–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. 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]
  73. Oroian, M.; Escriche, I. Antioxidants: Characterization, natural sources, extraction and analysis. Food Res. Int. 2015, 74, 10–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Smeriglio, A.; Barreca, D.; Bellocco, E.; Trombetta, D. Proanthocyanidins and hydrolysable tannins: Occurrence, dietary intake and pharmacological effects. Br. J. Pharmacol. 2017, 174, 1244–1262. [Google Scholar] [CrossRef] [PubMed]
  75. Uddin, M.S.; Sarker, M.Z.; Ferdosh, S.; Akanda, M.J.; Easmin, M.S.; Bt Shamsudin, S.H.; Bin Yunus, K. Phytosterols and their extraction from various plant matrices using supercritical carbon dioxide: A review. J. Sci. Food Agric. 2015, 95, 1385–1394. [Google Scholar] [CrossRef] [PubMed]
  76. Johnson, M.; Bradford, C. Omega-3, Omega-6 and Omega-9 Fatty Acids: Implications for Cardiovascular and Other Diseases. J. Glycomics Lipidomics 2014, 4. [Google Scholar] [CrossRef]
  77. Villares, A.; Mateo-Vivaracho, L.; Garcia-Lafuente, A.; Guillamon, E. Storage temperature and UV-irradiation influence on the ergosterol content in edible mushrooms. Food Chem. 2014, 147, 252–256. [Google Scholar] [CrossRef] [PubMed]
  78. Naumoska, K.; Vovk, I. Analysis of triterpenoids and phytosterols in vegetables by thin-layer chromatography coupled to tandem mass spectrometry. J. Chromatogr. A 2015, 1381, 229–238. [Google Scholar] [CrossRef] [PubMed]
  79. Mudannayake, D.C.; Wimalasiri, K.M.S.; Silva, K.; Ajlouni, S. Selected Sri Lankan food plants and other herbs as potential sources of inulin-type fructans. J. Natl. Sci. Found. Sri Lanka 2015, 43. [Google Scholar] [CrossRef]
  80. Pan, M.H.; Lai, C.S.; Wu, J.C.; Ho, C.T. Molecular mechanisms for chemoprevention of colorectal cancer by natural dietary compounds. Mol. Nutr. Food Res. 2011, 55, 32–45. [Google Scholar] [CrossRef] [PubMed]
  81. Singh, C.K.; George, J.; Ahmad, N. Resveratrol-based combinatorial strategies for cancer management. Ann. N. Y. Acad. Sci. 2013, 1290, 113–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ranelletti, F.O.; Maggiano, N.; Serra, F.G.; Ricci, R.; Larocca, L.M.; Lanza, P.; Scambia, G.; Fattorossi, A.; Capelli, A.; Piantelli, M. Quercetin inhibits p21-RAS expression in human colon cancer cell lines and in primary colorectal tumors. Int. J. Cancer 2000, 85, 438–445. [Google Scholar] [CrossRef] [Green Version]
  83. Psahoulia, F.H.; Drosopoulos, K.G.; Doubravska, L.; Andera, L.; Pintzas, A. Quercetin enhances TRAIL-mediated apoptosis in colon cancer cells by inducing the accumulation of death receptors in lipid rafts. Mol. Cancer Ther. 2007, 6, 2591–2599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Xavier, C.P.; Lima, C.F.; Rohde, M.; Pereira-Wilson, C. Quercetin enhances 5-fluorouracil-induced apoptosis in MSI colorectal cancer cells through p53 modulation. Cancer Chemother. Pharmacol. 2011, 68, 1449–1457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gutiérrez-del-Río, I.; Villar, C.J.; Lombó, F. Therapeutic Uses of Kaempferol: Anticancer and Anti-Inflammatory Activity; Nova Science Publishers, Inc: New York, NY, USA, 2016. [Google Scholar]
  86. Li, W.; Du, B.; Wang, T.; Wang, S.; Zhang, J. Kaempferol induces apoptosis in human HCT116 colon cancer cells via the Ataxia-Telangiectasia Mutated-p53 pathway with the involvement of p53 Upregulated Modulator of Apoptosis. Chem. Biol. Interact. 2009, 177, 121–127. [Google Scholar] [CrossRef] [PubMed]
  87. Lee, H.S.; Cho, H.J.; Yu, R.; Lee, K.W.; Chun, H.S.; Park, J.H. Mechanisms underlying apoptosis-inducing effects of Kaempferol in HT-29 human colon cancer cells. Int. J. Mol. Sci. 2014, 15, 2722–2737. [Google Scholar] [CrossRef] [PubMed]
  88. Cho, H.J.; Park, J.H. Kaempferol Induces Cell Cycle Arrest in HT-29 Human Colon Cancer Cells. J. Cancer Prev. 2013, 18, 257–263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Lee, H.S.; Cho, H.J.; Kwon, G.T.; Park, J.H. Kaempferol Downregulates Insulin-like Growth Factor-I Receptor and ErbB3 Signaling in HT-29 Human Colon Cancer Cells. J. Cancer Prev. 2014, 19, 161–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Yoshida, T.; Konishi, M.; Horinaka, M.; Yasuda, T.; Goda, A.E.; Taniguchi, H.; Yano, K.; Wakada, M.; Sakai, T. Kaempferol sensitizes colon cancer cells to TRAIL-induced apoptosis. Biochem. Biophys. Res. Commun. 2008, 375, 129–133. [Google Scholar] [CrossRef] [PubMed]
  91. Zhong, Y.; Krisanapun, C.; Lee, S.H.; Nualsanit, T.; Sams, C.; Peungvicha, P.; Baek, S.J. Molecular targets of apigenin in colorectal cancer cells: Involvement of p21, NAG-1 and p53. Eur. J. Cancer 2010, 46, 3365–3374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Lee, Y.; Sung, B.; Kang, Y.J.; Kim, D.H.; Jang, J.Y.; Hwang, S.Y.; Kim, M.; Lim, H.S.; Yoon, J.H.; Chung, H.Y.; et al. Apigenin-induced apoptosis is enhanced by inhibition of autophagy formation in HCT116 human colon cancer cells. Int. J. Oncol. 2014, 44, 1599–1606. [Google Scholar] [CrossRef] [PubMed]
  93. Yang, L.; Allred, K.F.; Dykes, L.; Allred, C.D.; Awika, J.M. Enhanced action of apigenin and naringenin combination on estrogen receptor activation in non-malignant colonocytes: Implications on sorghum-derived phytoestrogens. Food Funct. 2015, 6, 749–755. [Google Scholar] [CrossRef] [PubMed]
  94. Spector, D.; Anthony, M.; Alexander, D.; Arab, L. Soy consumption and colorectal cancer. Nutr. Cancer 2003, 47, 1–12. [Google Scholar] [CrossRef] [PubMed]
  95. Rossi, M.; Negri, E.; Talamini, R.; Bosetti, C.; Parpinel, M.; Gnagnarella, P.; Franceschi, S.; Dal Maso, L.; Montella, M.; Giacosa, A.; et al. Flavonoids and colorectal cancer in Italy. Cancer Epidemiol. Biomark. Prev. 2006, 15, 1555–1558. [Google Scholar] [CrossRef] [PubMed]
  96. Marin, L.; Miguelez, E.M.; Villar, C.J.; Lombo, F. Bioavailability of dietary polyphenols and gut microbiota metabolism: Antimicrobial properties. Biomed. Res. Int. 2015, 2015, 1–18. [Google Scholar] [CrossRef] [PubMed]
  97. Yu, Z.; Li, W.; Liu, F. Inhibition of proliferation and induction of apoptosis by genistein in colon cancer HT-29 cells. Cancer Lett. 2004, 215, 159–166. [Google Scholar] [CrossRef] [PubMed]
  98. Luo, Y.; Wang, S.X.; Zhou, Z.Q.; Wang, Z.; Zhang, Y.G.; Zhang, Y.; Zhao, P. Apoptotic effect of genistein on human colon cancer cells via inhibiting the nuclear factor-kappa B (NF-kappaB) pathway. Tumour Biol. 2014, 35, 11483–11488. [Google Scholar] [CrossRef] [PubMed]
  99. Mizushina, Y.; Shiomi, K.; Kuriyama, I.; Takahashi, Y.; Yoshida, H. Inhibitory effects of a major soy isoflavone, genistein, on human DNA topoisomerase II activity and cancer cell proliferation. Int. J. Oncol. 2013, 43, 1117–1124. [Google Scholar] [CrossRef] [PubMed]
  100. Pampaloni, B.; Palmini, G.; Mavilia, C.; Zonefrati, R.; Tanini, A.; Brandi, M.L. In vitro effects of polyphenols on colorectal cancer cells. World J. Gastrointest. Oncol. 2014, 6, 289–300. [Google Scholar] [CrossRef] [PubMed]
  101. Zhang, Y.; Li, Q.; Zhou, D.; Chen, H. Genistein, a soya isoflavone, prevents azoxymethane-induced up-regulation of WNT/beta-catenin signalling and reduces colon pre-neoplasia in rats. Br. J. Nutr. 2013, 109, 33–42. [Google Scholar] [CrossRef] [PubMed]
  102. Hwang, J.T.; Ha, J.; Park, O.J. Combination of 5-fluorouracil and genistein induces apoptosis synergistically in chemo-resistant cancer cells through the modulation of AMPK and COX-2 signaling pathways. Biochem. Biophys. Res. Commun. 2005, 332, 433–440. [Google Scholar] [CrossRef] [PubMed]
  103. Nasr Bouzaiene, N.; Kilani Jaziri, S.; Kovacic, H.; Chekir-Ghedira, L.; Ghedira, K.; Luis, J. The effects of caffeic, coumaric and ferulic acids on proliferation, superoxide production, adhesion and migration of human tumor cells in vitro. Eur. J. Pharmacol. 2015, 766, 99–105. [Google Scholar] [CrossRef] [PubMed]
  104. Murad, L.D.; Soares Nda, C.; Brand, C.; Monteiro, M.C.; Teodoro, A.J. Effects of caffeic and 5-caffeoylquinic acids on cell viability and cellular uptake in human colon adenocarcinoma cells. Nutr. Cancer 2015, 67, 532–542. [Google Scholar] [CrossRef] [PubMed]
  105. Choi, D.W.; Lim, M.S.; Lee, J.W.; Chun, W.; Lee, S.H.; Nam, Y.H.; Park, J.M.; Choi, D.H.; Kang, C.D.; Lee, S.J.; et al. The Cytotoxicity of Kahweol in HT-29 Human Colorectal Cancer Cells Is Mediated by Apoptosis and Suppression of Heat Shock Protein 70 Expression. Biomol. Ther. 2015, 23, 128–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Zheng, Q.; Hirose, Y.; Yoshimi, N.; Murakami, A.; Koshimizu, K.; Ohigashi, H.; Sakata, K.; Matsumoto, Y.; Sayama, Y.; Mori, H. Further investigation of the modifying effect of various chemopreventive agents on apoptosis and cell proliferation in human colon cancer cells. J. Cancer Res. Clin. Oncol. 2002, 128, 539–546. [Google Scholar] [CrossRef] [PubMed]
  107. Faried, A.; Kurnia, D.; Faried, L.S.; Usman, N.; Miyazaki, T.; Kato, H.; Kuwano, H. Anticancer effects of gallic acid isolated from Indonesian herbal medicine, Phaleria macrocarpa (Scheff.) Boerl, on human cancer cell lines. Int. J. Oncol. 2007, 30, 605–613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Giftson, J.S.; Jayanthi, S.; Nalini, N. Chemopreventive efficacy of gallic acid, an antioxidant and anticarcinogenic polyphenol, against 1,2-dimethyl hydrazine induced rat colon carcinogenesis. Investig. New Drugs 2010, 28, 251–259. [Google Scholar] [CrossRef] [PubMed]
  109. 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]
  110. Redondo-Blanco, S.; Fernandez, J.; Gutierrez-Del-Rio, I.; Villar, C.J.; Lombo, F. New Insights toward Colorectal Cancer Chemotherapy Using Natural Bioactive Compounds. Front. Pharmacol. 2017, 8, 109. [Google Scholar] [CrossRef] [PubMed]
  111. Del Follo-Martinez, A.; Banerjee, N.; Li, X.; Safe, S.; Mertens-Talcott, S. Resveratrol and quercetin in combination have anticancer activity in colon cancer cells and repress oncogenic microRNA-27a. Nutr. Cancer 2013, 65, 494–504. [Google Scholar] [CrossRef] [PubMed]
  112. Reddivari, L.; Charepalli, V.; Radhakrishnan, S.; Vadde, R.; Elias, R.J.; Lambert, J.D.; Vanamala, J.K. Grape compounds suppress colon cancer stem cells in vitro and in a rodent model of colon carcinogenesis. BMC Complement. Altern. Med. 2016, 16, 278. [Google Scholar] [CrossRef] [PubMed]
  113. Santandreu, F.M.; Valle, A.; Oliver, J.; Roca, P. Resveratrol potentiates the cytotoxic oxidative stress induced by chemotherapy in human colon cancer cells. Cell. Physiol. Biochem. 2011, 28, 219–228. [Google Scholar] [CrossRef] [PubMed]
  114. Mohapatra, P.; Preet, R.; Choudhuri, M.; Choudhuri, T.; Kundu, C.N. 5-fluorouracil increases the chemopreventive potentials of resveratrol through DNA damage and MAPK signaling pathway in human colorectal cancer cells. Oncol. Res. 2011, 19, 311–321. [Google Scholar] [CrossRef] [PubMed]
  115. Yang, S.; Li, W.; Sun, H.; Wu, B.; Ji, F.; Sun, T.; Chang, H.; Shen, P.; Wang, Y.; Zhou, D. Resveratrol elicits anti-colorectal cancer effect by activating miR-34c-KITLG in vitro and in vivo. BMC Cancer 2015, 15, 969. [Google Scholar] [CrossRef] [PubMed]
  116. Kaminski, B.M.; Weigert, A.; Scherzberg, M.C.; Ley, S.; Gilbert, B.; Brecht, K.; Brune, B.; Steinhilber, D.; Stein, J.; Ulrich-Ruckert, S. Resveratrol-induced potentiation of the antitumor effects of oxaliplatin is accompanied by an altered cytokine profile of human monocyte-derived macrophages. Apoptosis 2014, 19, 1136–1147. [Google Scholar] [CrossRef] [PubMed]
  117. Ali, I.; Braun, D.P. Resveratrol enhances mitomycin C-mediated suppression of human colorectal cancer cell proliferation by up-regulation of p21WAF1/CIP1. Anticancer Res. 2014, 34, 5439–5446. [Google Scholar] [PubMed]
  118. Chauhan, D.P. Chemotherapeutic potential of curcumin for colorectal cancer. Curr. Pharm. Des. 2002, 8, 1695–1706. [Google Scholar] [CrossRef] [PubMed]
  119. Goel, A.; Kunnumakkara, A.B.; Aggarwal, B.B. Curcumin as “Curecumin”: From kitchen to clinic. Biochem. Pharmacol. 2008, 75, 787–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Prasad, S.; Tyagi, A.K.; Aggarwal, B.B. Recent developments in delivery, bioavailability, absorption and metabolism of curcumin: The golden pigment from golden spice. Cancer Res. Treat. 2014, 46, 2–18. [Google Scholar] [CrossRef] [PubMed]
  121. Du, B.; Jiang, L.; Xia, Q.; Zhong, L. Synergistic inhibitory effects of curcumin and 5-fluorouracil on the growth of the human colon cancer cell line HT-29. Chemotherapy 2006, 52, 23–28. [Google Scholar] [CrossRef] [PubMed]
  122. Srimuangwong, K.; Tocharus, C.; Yoysungnoen Chintana, P.; Suksamrarn, A.; Tocharus, J. Hexahydrocurcumin enhances inhibitory effect of 5-fluorouracil on HT-29 human colon cancer cells. World J. Gastroenterol. 2012, 18, 2383–2389. [Google Scholar] [CrossRef] [PubMed]
  123. Kunnumakkara, A.B.; Diagaradjane, P.; Anand, P.; Harikumar, K.B.; Deorukhkar, A.; Gelovani, J.; Guha, S.; Krishnan, S.; Aggarwal, B.B. Curcumin sensitizes human colorectal cancer to capecitabine by modulation of cyclin D1, COX-2, MMP-9, VEGF and CXCR4 expression in an orthotopic mouse model. Int. J. Cancer 2009, 125, 2187–2197. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, X.; Yang, H.H.; Liu, Y.; Zhou, Q.; Chen, Z.H. Lycopene Consumption and Risk of Colorectal Cancer: A Meta-Analysis of Observational Studies. Nutr. Cancer 2016, 68, 1083–1096. [Google Scholar] [CrossRef] [PubMed]
  125. Trejo-Solis, C.; Pedraza-Chaverri, J.; Torres-Ramos, M.; Jimenez-Farfan, D.; Cruz Salgado, A.; Serrano-Garcia, N.; Osorio-Rico, L.; Sotelo, J. Multiple molecular and cellular mechanisms of action of lycopene in cancer inhibition. Evid.-Based Complement. Altern. Med. ECAM 2013, 2013, 1–17. [Google Scholar] [CrossRef] [PubMed]
  126. Tang, F.Y.; Shih, C.J.; Cheng, L.H.; Ho, H.J.; Chen, H.J. Lycopene inhibits growth of human colon cancer cells via suppression of the Akt signaling pathway. Mol. Nutr. Food Res. 2008, 52, 646–654. [Google Scholar] [CrossRef] [PubMed]
  127. Lin, M.C.; Wang, F.Y.; Kuo, Y.H.; Tang, F.Y. Cancer chemopreventive effects of lycopene: Suppression of MMP-7 expression and cell invasion in human colon cancer cells. J. Agric. Food Chem. 2011, 59, 11304–11318. [Google Scholar] [CrossRef] [PubMed]
  128. Tang, F.Y.; Cho, H.J.; Pai, M.H.; Chen, Y.H. Concomitant supplementation of lycopene and eicosapentaenoic acid inhibits the proliferation of human colon cancer cells. J. Nutr. Biochem. 2009, 20, 426–434. [Google Scholar] [CrossRef] [PubMed]
  129. Danese, S.; Mantovani, A. Inflammatory bowel disease and intestinal cancer: A paradigm of the Yin-Yang interplay between inflammation and cancer. Oncogene 2010, 29, 3313–3323. [Google Scholar] [CrossRef] [PubMed]
  130. Cha, J.H.; Kim, W.K.; Ha, A.W.; Kim, M.H.; Chang, M.J. Anti-inflammatory effect of lycopene in SW480 human colorectal cancer cells. Nutr. Res. Pract. 2017, 11, 90–96. [Google Scholar] [CrossRef] [PubMed]
  131. Iftikhar, S.; Lietz, H.; Mobarhan, S.; Frommel, T.O. In vitro beta-carotene toxicity for human colon cancer cells. Nutr. Cancer 1996, 25, 221–230. [Google Scholar] [CrossRef] [PubMed]
  132. Palozza, P.; Serini, S.; Torsello, A.; Di Nicuolo, F.; Piccioni, E.; Ubaldi, V.; Pioli, C.; Wolf, F.I.; Calviello, G. Beta-carotene regulates NF-kappaB DNA-binding activity by a redox mechanism in human leukemia and colon adenocarcinoma cells. J. Nutr. 2003, 133, 381–388. [Google Scholar] [CrossRef] [PubMed]
  133. Briviba, K.; Schnabele, K.; Schwertle, E.; Blockhaus, M.; Rechkemmer, G. Beta-carotene inhibits growth of human colon carcinoma cells in vitro by induction of apoptosis. Biol. Chem. 2001, 382, 1663–1668. [Google Scholar] [CrossRef] [PubMed]
  134. Palozza, P.; Serini, S.; Maggiano, N.; Angelini, M.; Boninsegna, A.; Di Nicuolo, F.; Ranelletti, F.O.; Calviello, G. Induction of cell cycle arrest and apoptosis in human colon adenocarcinoma cell lines by beta-carotene through down-regulation of cyclin A and Bcl-2 family proteins. Carcinogenesis 2002, 23, 11–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Palozza, P.; Calviello, G.; Serini, S.; Maggiano, N.; Lanza, P.; Ranelletti, F.O.; Bartoli, G.M. Beta-Carotene at high concentrations induces apoptosis by enhancing oxy-radical production in human adenocarcinoma cells. Free Radic. Biol. Med. 2001, 30, 1000–1007. [Google Scholar] [CrossRef]
  136. Nam, K.N.; Park, Y.M.; Jung, H.J.; Lee, J.Y.; Min, B.D.; Park, S.U.; Jung, W.S.; Cho, K.H.; Park, J.H.; Kang, I.; et al. Anti-inflammatory effects of crocin and crocetin in rat brain microglial cells. Eur. J. Pharmacol. 2010, 648, 110–116. [Google Scholar] [CrossRef] [PubMed]
  137. Aung, H.H.; Wang, C.Z.; Ni, M.; Fishbein, A.; Mehendale, S.R.; Xie, J.T.; Shoyama, C.Y.; Yuan, C.S. Crocin from Crocus sativus possesses significant anti-proliferation effects on human colorectal cancer cells. Exp. Oncol. 2007, 29, 175–180. [Google Scholar] [PubMed]
  138. Dhar, A.; Mehta, S.; Dhar, G.; Dhar, K.; Banerjee, S.; Van Veldhuizen, P.; Campbell, D.R.; Banerjee, S.K. Crocetin inhibits pancreatic cancer cell proliferation and tumor progression in a xenograft mouse model. Mol. Cancer Ther. 2009, 8, 315–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Mousavi, S.H.; Tavakkol-Afshari, J.; Brook, A.; Jafari-Anarkooli, I. Role of caspases and Bax protein in saffron-induced apoptosis in MCF-7 cells. Food Chem. Toxicol. 2009, 47, 1909–1913. [Google Scholar] [CrossRef] [PubMed]
  140. Ray, P.; Guha, D.; Chakraborty, J.; Banerjee, S.; Adhikary, A.; Chakraborty, S.; Das, T.; Sa, G. Crocetin exploits p53-induced death domain (PIDD) and FAS-associated death domain (FADD) proteins to induce apoptosis in colorectal cancer. Sci. Rep. 2016, 6, 32979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Bajbouj, K.; Schulze-Luehrmann, J.; Diermeier, S.; Amin, A.; Schneider-Stock, R. The anticancer effect of saffron in two p53 isogenic colorectal cancer cell lines. BMC Complement. Altern. Med. 2012, 12, 69. [Google Scholar] [CrossRef] [PubMed]
  142. Zhuang, X.; Dong, A.; Wang, R.; Shi, A. Crocetin treatment inhibits proliferation of colon cancer cells through down-regulation of genes involved in the inflammation. Saudi J. Biol. Sci. 2017. [Google Scholar] [CrossRef]
  143. Yaffe, P.B.; Power Coombs, M.R.; Doucette, C.D.; Walsh, M.; Hoskin, D.W. Piperine, an alkaloid from black pepper, inhibits growth of human colon cancer cells via G1 arrest and apoptosis triggered by endoplasmic reticulum stress. Mol. Carcinog. 2015, 54, 1070–1085. [Google Scholar] [CrossRef] [PubMed]
  144. Yaffe, P.B.; Doucette, C.D.; Walsh, M.; Hoskin, D.W. Piperine impairs cell cycle progression and causes reactive oxygen species-dependent apoptosis in rectal cancer cells. Exp. Mol. Pathol. 2013, 94, 109–114. [Google Scholar] [CrossRef] [PubMed]
  145. Kaur, H.; He, B.; Zhang, C.; Rodriguez, E.; Hage, D.S.; Moreau, R. Piperine potentiates curcumin-mediated repression of mTORC1 signaling in human intestinal epithelial cells: Implications for the inhibition of protein synthesis and TNFalpha signaling. J. Nutr. Biochem. 2018, 57, 276–286. [Google Scholar] [CrossRef] [PubMed]
  146. Ginghina, O.; Negrei, C.; Hudita, A.; Ioana-Lavric, V.; Galateanu, B.; Dragomir, S. In vitro impact of some natural compounds on HT-29 colorectal adenocarcinoma cells. Farmacia 2017, 65, 947–953. [Google Scholar]
  147. Lee, S.H.; Clark, R. Anti-Tumorigenic Effects of Capsaicin in Colon Cancer. J. Food Chem. Nanotechnol. 2016, 2, 162–167. [Google Scholar] [CrossRef]
  148. Lu, H.F.; Chen, Y.L.; Yang, J.S.; Yang, Y.Y.; Liu, J.Y.; Hsu, S.C.; Lai, K.C.; Chung, J.G. Antitumor activity of capsaicin on human colon cancer cells in vitro and colo 205 tumor xenografts in vivo. J. Agric. Food Chem. 2010, 58, 12999–13005. [Google Scholar] [CrossRef] [PubMed]
  149. Yang, K.M.; Pyo, J.O.; Kim, G.Y.; Yu, R.; Han, I.S.; Ju, S.A.; Kim, W.H.; Kim, B.S. Capsaicin induces apoptosis by generating reactive oxygen species and disrupting mitochondrial transmembrane potential in human colon cancer cell lines. Cell. Mol. Biol. Lett. 2009, 14, 497–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Yang, J.; Li, T.Z.; Xu, G.H.; Luo, B.B.; Chen, Y.X.; Zhang, T. Low-concentration capsaicin promotes colorectal cancer metastasis by triggering ROS production and modulating Akt/mTOR and STAT-3 pathways. Neoplasma 2013, 60, 364–372. [Google Scholar] [CrossRef] [PubMed]
  151. Jin, J.; Lin, G.; Huang, H.; Xu, D.; Yu, H.; Ma, X.; Zhu, L.; Ma, D.; Jiang, H. Capsaicin mediates cell cycle arrest and apoptosis in human colon cancer cells via stabilizing and activating p53. Int. J. Biol. Sci. 2014, 10, 285–295. [Google Scholar] [CrossRef] [PubMed]
  152. Kim, Y.M.; Hwang, J.T.; Kwak, D.W.; Lee, Y.K.; Park, O.J. Involvement of AMPK signaling cascade in capsaicin-induced apoptosis of HT-29 colon cancer cells. Ann. N. Y. Acad. Sci. 2007, 1095, 496–503. [Google Scholar] [CrossRef] [PubMed]
  153. Lee, S.H.; Richardson, R.L.; Dashwood, R.H.; Baek, S.J. Capsaicin represses transcriptional activity of beta-catenin in human colorectal cancer cells. J. Nutr. Biochem. 2012, 23, 646–655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Kim, M.Y.; Trudel, L.J.; Wogan, G.N. Apoptosis induced by capsaicin and resveratrol in colon carcinoma cells requires nitric oxide production and caspase activation. Anticancer Res. 2009, 29, 3733–3740. [Google Scholar] [PubMed]
  155. Bessler, H.; Djaldetti, M. Capsaicin Modulates the Immune Cross Talk Between Human Mononuclears and Cells from Two Colon Carcinoma Lines. Nutr. Cancer 2017, 69, 14–20. [Google Scholar] [CrossRef] [PubMed]
  156. Clark, R.; Lee, J.; Lee, S.H. Synergistic anticancer activity of capsaicin and 3,3′-diindolylmethane in human colorectal cancer. J. Agric. Food Chem. 2015, 63, 4297–4304. [Google Scholar] [CrossRef] [PubMed]
  157. Song, S.; Cheng, D.; Wei, S.; Wang, X.; Niu, Y.; Qi, W.; Wang, C. Preventive effect of genistein on AOM/DSS-induced colonic neoplasm by modulating the PI3K/AKT/FOXO3 signaling pathway in mice fed a high-fat diet. J. Funct. Foods 2018, 46, 237–242. [Google Scholar] [CrossRef]
  158. Son, T.G.; Gong, E.J.; Bae, M.J.; Kim, S.D.; Heo, K.; Moon, C.; Yang, K.; Kim, J.S. Protective effect of genistein on radiation-induced intestinal injury in tumor bearing mice. BMC Complement. Altern. Med. 2013, 13, 103. [Google Scholar] [CrossRef] [PubMed]
  159. Kang, S.-Y.; Seeram, N.P.; Nair, M.G.; Bourquin, L.D. Tart cherry anthocyanins inhibit tumor development in ApcMin mice and reduce proliferation of human colon cancer cells. Cancer Lett. 2003, 194, 13–19. [Google Scholar] [CrossRef]
  160. Shi, N.; Clinton, S.K.; Liu, Z.; Wang, Y.; Riedl, K.M.; Schwartz, S.J.; Zhang, X.; Pan, Z.; Chen, T. Strawberry phytochemicals inhibit azoxymethane/dextran sodium sulfate-induced colorectal carcinogenesis in Crj: CD-1 mice. Nutrients 2015, 7, 1696–1715. [Google Scholar] [CrossRef] [PubMed]
  161. Silva, R.M.; Campanholo, V.M.; Paiotti, A.P.; Artigiani Neto, R.; Oshima, C.T.; Ribeiro, D.A.; Forones, N.M. Chemopreventive activity of grape juice concentrate (G8000TM) on rat colon carcinogenesis induced by azoxymethane. Environ. Toxicol. Pharmacol. 2015, 40, 870–875. [Google Scholar] [CrossRef] [PubMed]
  162. Fernandez, J.; Garcia, L.; Monte, J.; Villar, C.J.; Lombo, F. Functional Anthocyanin-Rich Sausages Diminish Colorectal Cancer in an Animal Model and Reduce Pro-Inflammatory Bacteria in the Intestinal Microbiota. Genes 2018, 9, 133. [Google Scholar] [CrossRef] [PubMed]
  163. Matsunaga, K.; Katayama, M.; Sakata, K.; Kuno, T.; Yoshida, K.; Yamada, Y.; Hirose, Y.; Yoshimi, N.; Mori, H. Inhibitory Effects of Chlorogenic Acid on Azoxymethane-induced Colon Carcinogenesis in Male F344 Rats. Asian Pac. J. Cancer Prev. 2002, 3, 163–166. [Google Scholar] [PubMed]
  164. Banerjee, N.; Kim, H.; Talcott, S.T.; Turner, N.D.; Byrne, D.H.; Mertens-Talcott, S.U. Plum polyphenols inhibit colorectal aberrant crypt foci formation in rats: Potential role of the miR-143/protein kinase B/mammalian target of rapamycin axis. Nutr. Res. 2016, 36, 1105–1113. [Google Scholar] [CrossRef] [PubMed]
  165. Zhu, B.; Shang, B.; Li, Y.; Zhen, Y. Inhibition of histone deacetylases by trans-cinnamic acid and its antitumor effect against colon cancer xenografts in athymic mice. Mol. Med. Rep. 2016, 13, 4159–4166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Sharma, S.H.; Chellappan, D.R.; Chinnaswamy, P.; Nagarajan, S. Protective effect of p-coumaric acid against 1,2 dimethylhydrazine induced colonic preneoplastic lesions in experimental rats. Biomed. Pharmacother. 2017, 94, 577–588. [Google Scholar] [CrossRef] [PubMed]
  167. Venkatachalam, K.; Gunasekaran, S.; Jesudoss, V.A.; Namasivayam, N. The effect of rosmarinic acid on 1,2-dimethylhydrazine induced colon carcinogenesis. Exp. Toxicol. Pathol. 2013, 65, 409–418. [Google Scholar] [CrossRef] [PubMed]
  168. Giftson Senapathy, J.; Jayanthi, S.; Viswanathan, P.; Umadevi, P.; Nalini, N. Effect of gallic acid on xenobiotic metabolizing enzymes in 1,2-dimethyl hydrazine induced colon carcinogenesis in Wistar rats—A chemopreventive approach. Food Chem. Toxicol. 2011, 49, 887–892. [Google Scholar] [CrossRef] [PubMed]
  169. Shah, N.R.; Patel, B.M. Secoisolariciresinol diglucoside rich extract of L. usitatissimum prevents diabetic colon cancer through inhibition of CDK4. Biomed. Pharmacother. 2016, 83, 733–739. [Google Scholar] [CrossRef] [PubMed]
  170. Gomides, A.F.; Paula, S.O.; Rosa, D.D.; Oliveira, L.L.; Comastri, D.S.; Peluzio Mdo, C. Use of defatted flaxseed meal reduces precancerous colon lesions in C57BL/6 mice. Acta Cir. Bras. 2013, 28, 607–613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Shimizu, S.; Fujii, G.; Takahashi, M.; Nakanishi, R.; Komiya, M.; Shimura, M.; Noma, N.; Onuma, W.; Terasaki, M.; Yano, T.; et al. Sesamol suppresses cyclooxygenase-2 transcriptional activity in colon cancer cells and modifies intestinal polyp development in Apc (Min/+) mice. J. Clin. Biochem. Nutr. 2014, 54, 95–101. [Google Scholar] [CrossRef] [PubMed]
  172. Jagetia, G.C. Radioprotection and radiosensitization by curcumin. Adv. Exp. Med. Biol. 2007, 595, 301–320. [Google Scholar] [PubMed]
  173. Tang, F.Y.; Pai, M.H.; Wang, X.D. Consumption of lycopene inhibits the growth and progression of colon cancer in a mouse xenograft model. J. Agric. Food Chem. 2011, 59, 9011–9021. [Google Scholar] [CrossRef] [PubMed]
  174. Dias, M.C.; Vieiralves, N.F.; Gomes, M.I.; Salvadori, D.M.; Rodrigues, M.A.; Barbisan, L.F. Effects of lycopene, synbiotic and their association on early biomarkers of rat colon carcinogenesis. Food Chem. Toxicol. 2010, 48, 772–780. [Google Scholar] [CrossRef] [PubMed]
  175. Kawabata, K.; Tung, N.H.; Shoyama, Y.; Sugie, S.; Mori, T.; Tanaka, T. Dietary Crocin Inhibits Colitis and Colitis-Associated Colorectal Carcinogenesis in Male ICR Mice. Evid.-Based Complement. Altern. Med. ECAM 2012, 2012, 1–13. [Google Scholar] [CrossRef] [PubMed]
  176. Giner, E.; Recio, M.C.; Rios, J.L.; Cerda-Nicolas, J.M.; Giner, R.M. Chemopreventive effect of oleuropein in colitis-associated colorectal cancer in c57bl/6 mice. Mol. Nutr. Food Res. 2016, 60, 242–255. [Google Scholar] [CrossRef] [PubMed]
  177. Caetano, B.F.R.; Tablas, M.B.; Pereira, N.E.F.; de Moura, N.A.; Carvalho, R.F.; Rodrigues, M.A.M.; Barbisan, L.F. Capsaicin reduces genotoxicity, colonic cell proliferation and preneoplastic lesions induced by 1,2-dimethylhydrazine in rats. Toxicol. Appl. Pharmacol. 2018, 338, 93–102. [Google Scholar] [CrossRef] [PubMed]
  178. Rajendran, P.; Dashwood, W.M.; Li, L.; Kang, Y.; Kim, E.; Johnson, G.; Fischer, K.A.; Lohr, C.V.; Williams, D.E.; Ho, E.; et al. Nrf2 status affects tumor growth, HDAC3 gene promoter associations, and the response to sulforaphane in the colon. Clin. Epigenetics 2015, 7, 102. [Google Scholar] [CrossRef] [PubMed]
  179. Baskar, A.A.; Al Numair, K.S.; Gabriel Paulraj, M.; Alsaif, M.A.; Muamar, M.A.; Ignacimuthu, S. beta-sitosterol prevents lipid peroxidation and improves antioxidant status and histoarchitecture in rats with 1,2-dimethylhydrazine-induced colon cancer. J. Med. Food 2012, 15, 335–343. [Google Scholar] [CrossRef] [PubMed]
  180. Asfour, W.; Almadi, S.; Haffar, L. Thymoquinone Suppresses Cellular Proliferation, Inhibits VEGF Production and Obstructs Tumor Progression and Invasion in the Rat Model of DMH-Induced Colon Carcinogenesis. Pharmacol. Pharm. 2013, 4, 7–17. [Google Scholar] [CrossRef]
  181. Kortum, B.; Campregher, C.; Lang, M.; Khare, V.; Pinter, M.; Evstatiev, R.; Schmid, G.; Mittlbock, M.; Scharl, T.; Kucherlapati, M.H.; et al. Mesalazine and thymoquinone attenuate intestinal tumour development in Msh2(loxP/loxP) Villin-Cre mice. Gut 2015, 64, 1905–1912. [Google Scholar] [CrossRef] [PubMed]
  182. Long, M.; Tao, S.; Rojo de la Vega, M.; Jiang, T.; Wen, Q.; Park, S.L.; Zhang, D.D.; Wondrak, G.T. Nrf2-dependent suppression of azoxymethane/dextran sulfate sodium-induced colon carcinogenesis by the cinnamon-derived dietary factor cinnamaldehyde. Cancer Prev. Res. 2015, 8, 444–454. [Google Scholar] [CrossRef] [PubMed]
  183. Sivaranjani, A.; Sivagami, G.; Nalini, N. Chemopreventive effect of carvacrol on 1,2-dimethylhydrazine induced experimental colon carcinogenesis. J. Cancer Res. Ther. 2016, 12, 755–762. [Google Scholar] [PubMed]
  184. Wang, W.; Yang, J.; Nimiya, Y.; Lee, K.S.S.; Sanidad, K.; Qi, W.; Sukamtoh, E.; Park, Y.; Liu, Z.; Zhang, G. omega-3 Polyunsaturated fatty acids and their cytochrome P450-derived metabolites suppress colorectal tumor development in mice. J. Nutr. Biochem. 2017, 48, 29–35. [Google Scholar] [CrossRef] [PubMed]
  185. Hao, G.W.; Chen, Y.S.; He, D.M.; Wang, H.Y.; Wu, G.H.; Zhang, B. Growth of Human Colon Cancer Cells in Nude Mice is Delayed by Ketogenic Diet With or Without Omega-3 Fatty Acids and Medium-chain Triglycerides. Asian Pac. J. Cancer Prev. 2015, 16, 2061–2068. [Google Scholar] [CrossRef] [PubMed]
  186. Piazzi, G.; D’Argenio, G.; Prossomariti, A.; Lembo, V.; Mazzone, G.; Candela, M.; Biagi, E.; Brigidi, P.; Vitaglione, P.; Fogliano, V.; et al. Eicosapentaenoic acid free fatty acid prevents and suppresses colonic neoplasia in colitis-associated colorectal cancer acting on Notch signaling and gut microbiota. Int. J. Cancer 2014, 135, 2004–2013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Barone, M.; Notarnicola, M.; Caruso, M.G.; Scavo, M.P.; Viggiani, M.T.; Tutino, V.; Polimeno, L.; Pesetti, B.; Di Leo, A.; Francavilla, A. Olive oil and omega-3 polyunsaturated fatty acids suppress intestinal polyp growth by modulating the apoptotic process in ApcMin/+ mice. Carcinogenesis 2014, 35, 1613–1619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Wang, J.; Li, W.; Huang, X.; Liu, Y.; Li, Q.; Zheng, Z.; Wang, K. A polysaccharide from Lentinus edodes inhibits human colon cancer cell proliferation and suppresses tumor growth in athymic nude mice. Oncotarget 2017, 8, 610–623. [Google Scholar] [CrossRef] [PubMed]
  189. Masuda, Y.; Inoue, H.; Ohta, H.; Miyake, A.; Konishi, M.; Nanba, H. Oral administration of soluble beta-glucans extracted from Grifola frondosa induces systemic antitumor immune response and decreases immunosuppression in tumor-bearing mice. Int. J. Cancer 2013, 133, 108–119. [Google Scholar] [CrossRef] [PubMed]
  190. Pattananandecha, T.; Sirilun, S.; Duangjitcharoen, Y.; Sivamaruthi, B.S.; Suwannalert, P.; Peerajan, S.; Chaiyasut, C. Hydrolysed inulin alleviates the azoxymethane-induced preneoplastic aberrant crypt foci by altering selected intestinal microbiota in Sprague-Dawley rats. Pharm. Biol. 2016, 54, 1596–1605. [Google Scholar] [CrossRef] [PubMed]
  191. Qamar, T.R.; Syed, F.; Nasir, M.; Rehman, H.; Zahid, M.N.; Liu, R.H.; Iqbal, S. Novel Combination of Prebiotics Galacto-Oligosaccharides and Inulin-Inhibited Aberrant Crypt Foci Formation and Biomarkers of Colon Cancer in Wistar Rats. Nutrients 2016, 8, 465. [Google Scholar] [CrossRef] [PubMed]
  192. Hijová, E.; Szabadosova, V.; Štofilová, J.; Hrčková, G. Chemopreventive and metabolic effects of inulin on colon cancer development. J. Vet. Sci. 2013, 14, 387–393. [Google Scholar] [CrossRef] [PubMed]
  193. Verma, A.; Shukla, G. Administration of prebiotic inulin suppresses 1,2 dimethylhydrazine dihydrochloride induced procarcinogenic biomarkers fecal enzymes and preneoplastic lesions in early colon carcinogenesis in Sprague Dawley rats. J. Funct. Foods 2013, 5, 991–996. [Google Scholar] [CrossRef]
  194. Stofilova, J.; Szabadosova, V.; Hrckova, G.; Salaj, R.; Bertkova, I.; Hijova, E.; Strojny, L.; Bomba, A. Co-administration of a probiotic strain Lactobacillus plantarum LS/07 CCM7766 with prebiotic inulin alleviates the intestinal inflammation in rats exposed to N,N-dimethylhydrazine. Int. Immunopharmacol. 2015, 24, 361–368. [Google Scholar] [CrossRef] [PubMed]
  195. Wu, W.T.; Yang, L.C.; Chen, H.L. Effects of konjac glucomannan, inulin and cellulose on acute colonic responses to genotoxic azoxymethane. Food Chem. 2014, 155, 304–310. [Google Scholar] [CrossRef] [PubMed]
  196. Georgescu, S.R.; Sarbu, M.I.; Matei, C.; Ilie, M.A.; Caruntu, C.; Constantin, C.; Neagu, M.; Tampa, M. Capsaicin: Friend or Foe in Skin Cancer and Other Related Malignancies? Nutrients 2017, 9, 1365. [Google Scholar] [CrossRef] [PubMed]
  197. Erin, N.; Boyer, P.J.; Bonneau, R.H.; Clawson, G.A.; Welch, D.R. Capsaicin-mediated denervation of sensory neurons promotes mammary tumor metastasis to lung and heart. Anticancer Res. 2004, 24, 1003–1009. [Google Scholar] [PubMed]
  198. Ververis, K.; Hiong, A.; Karagiannis, T.C.; Licciardi, P.V. Histone deacetylase inhibitors (HDACIs): Multitargeted anticancer agents. Biologics 2013, 7, 47–60. [Google Scholar] [PubMed]
  199. Ma, Q. Role of nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef] [PubMed]
  200. Arivalagan, S.; Thomas, N.S.; Chandrasekaran, B.; Mani, V.; Siddique, A.I.; Kuppsamy, T.; Namasivayam, N. Combined therapeutic efficacy of carvacrol and X-radiation against 1,2-dimethyl hydrazine-induced experimental rat colon carcinogenesis. Mol. Cell. Biochem. 2015, 410, 37–54. [Google Scholar] [CrossRef] [PubMed]
  201. Vogtmann, E.; Xiang, Y.B.; Li, H.L.; Levitan, E.B.; Yang, G.; Waterbor, J.W.; Gao, J.; Cai, H.; Xie, L.; Wu, Q.J.; et al. Fruit and vegetable intake and the risk of colorectal cancer: Results from the Shanghai Men’s Health Study. Cancer Causes Control 2013, 24, 1935–1945. [Google Scholar] [CrossRef] [PubMed]
  202. Akhter, M.; Iwasaki, M.; Yamaji, T.; Sasazuki, S.; Tsugane, S. Dietary isoflavone and the risk of colorectal adenoma: A case-control study in Japan. Br. J. Cancer 2009, 100, 1812–1816. [Google Scholar] [CrossRef] [PubMed]
  203. Shin, A.; Joo, J.; Yang, H.R.; Bak, J.; Park, Y.; Kim, J.; Oh, J.H.; Nam, B.H. Risk prediction model for colorectal cancer: National Health Insurance Corporation study, Korea. PLoS ONE 2014, 9, e88079. [Google Scholar] [CrossRef] [PubMed]
  204. Zamora-Ros, R.; Not, C.; Guino, E.; Lujan-Barroso, L.; Garcia, R.M.; Biondo, S.; Salazar, R.; Moreno, V. Association between habitual dietary flavonoid and lignan intake and colorectal cancer in a Spanish case-control study (the Bellvitge Colorectal Cancer Study). Cancer Causes Control 2013, 24, 549–557. [Google Scholar] [CrossRef] [PubMed]
  205. Ko, K.P.; Yeo, Y.; Yoon, J.H.; Kim, C.S.; Tokudome, S.; Ngoan, L.T.; Koriyama, C.; Lim, Y.K.; Chang, S.H.; Shin, H.R.; et al. Plasma phytoestrogens concentration and risk of colorectal cancer in two different Asian populations. Clin. Nutr. 2018, 37, 1675–1682. [Google Scholar] [CrossRef] [PubMed]
  206. Ganai, A.A.; Farooqi, H. Bioactivity of genistein: A review of in vitro and in vivo studies. Biomed. Pharmacother. 2015, 76, 30–38. [Google Scholar] [CrossRef] [PubMed]
  207. 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] [Green Version]
  208. Principi, M.; Di Leo, A.; Pricci, M.; Scavo, M.P.; Guido, R.; Tanzi, S.; Piscitelli, D.; Pisani, A.; Ierardi, E.; Comelli, M.C.; et al. Phytoestrogens/insoluble fibers and colonic estrogen receptor beta: Randomized, double-blind, placebo-controlled study. World J. Gastroenterol. 2013, 19, 4325–4333. [Google Scholar] [CrossRef] [PubMed]
  209. Calabrese, C.; Pratico, C.; Calafiore, A.; Coscia, M.; Gentilini, L.; Poggioli, G.; Gionchetti, P.; Campieri, M.; Rizzello, F. Eviendep(R) reduces number and size of duodenal polyps in familial adenomatous polyposis patients with ileal pouch-anal anastomosis. World J. Gastroenterol. 2013, 19, 5671–5677. [Google Scholar] [CrossRef] [PubMed]
  210. Shin, A.; Lee, J.; Lee, J.; Park, M.S.; Park, J.W.; Park, S.C.; Oh, J.H.; Kim, J. Isoflavone and Soyfood Intake and Colorectal Cancer Risk: A Case-Control Study in Korea. PLoS ONE 2015, 10, e0143228. [Google Scholar] [CrossRef] [PubMed]
  211. Yang, G.; Shu, X.O.; Li, H.; Chow, W.H.; Cai, H.; Zhang, X.; Gao, Y.T.; Zheng, W. Prospective cohort study of soy food intake and colorectal cancer risk in women. Am. J. Clin. Nutr. 2009, 89, 577–583. [Google Scholar] [CrossRef] [PubMed]
  212. Thomasset, S.; Berry, D.P.; Cai, H.; West, K.; Marczylo, T.H.; Marsden, D.; Brown, K.; Dennison, A.; Garcea, G.; Miller, A.; et al. Pilot study of oral anthocyanins for colorectal cancer chemoprevention. Cancer Prev. Res. 2009, 2, 625–633. [Google Scholar] [CrossRef] [PubMed]
  213. Wang, L.S.; Burke, C.A.; Hasson, H.; Kuo, C.T.; Molmenti, C.L.; Seguin, C.; Liu, P.; Huang, T.H.; Frankel, W.L.; Stoner, G.D. A phase Ib study of the effects of black raspberries on rectal polyps in patients with familial adenomatous polyposis. Cancer Prev. Res. 2014, 7, 666–674. [Google Scholar] [CrossRef] [PubMed]
  214. Tanaka, S.; Haruma, K.; Yoshihara, M.; Kajiyama, G.; Kira, K.; Amagase, H.; Chayama, K. Aged garlic extract has potential suppressive effect on colorectal adenomas in humans. J. Nutr. 2006, 136, 821s–826s. [Google Scholar] [CrossRef] [PubMed]
  215. McCullough, M.L.; Jacobs, E.J.; Shah, R.; Campbell, P.T.; Gapstur, S.M. Garlic consumption and colorectal cancer risk in the CPS-II Nutrition Cohort. Cancer Causes Control 2012, 23, 1643–1651. [Google Scholar] [CrossRef] [PubMed]
  216. Meng, S.; Zhang, X.; Giovannucci, E.L.; Ma, J.; Fuchs, C.S.; Cho, E. No association between garlic intake and risk of colorectal cancer. Cancer Epidemiol. 2013, 37, 152–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Satia, J.A.; Littman, A.; Slatore, C.G.; Galanko, J.A.; White, E. Associations of herbal and specialty supplements with lung and colorectal cancer risk in the VITamins and Lifestyle study. Cancer Epidemiol. Biomark. Prev. 2009, 18, 1419–1428. [Google Scholar] [CrossRef] [PubMed]
  218. Lu, M.S.; Fang, Y.J.; Chen, Y.M.; Luo, W.P.; Pan, Z.Z.; Zhong, X.; Zhang, C.X. Higher intake of carotenoid is associated with a lower risk of colorectal cancer in Chinese adults: A case-control study. Eur. J. Nutr. 2015, 54, 619–628. [Google Scholar] [CrossRef] [PubMed]
  219. Leenders, M.; Leufkens, A.M.; Siersema, P.D.; van Duijnhoven, F.J.; Vrieling, A.; Hulshof, P.J.; van Gils, C.H.; Overvad, K.; Roswall, N.; Kyro, C.; et al. Plasma and dietary carotenoids and vitamins A, C and E and risk of colon and rectal cancer in the European Prospective Investigation into Cancer and Nutrition. Int. J. Cancer 2014, 135, 2930–2939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Kabat, G.C.; Kim, M.Y.; Sarto, G.E.; Shikany, J.M.; Rohan, T.E. Repeated measurements of serum carotenoid, retinol and tocopherol levels in relation to colorectal cancer risk in the Women’s Health Initiative. Eur. J. Clin. Nutr. 2012, 66, 549–554. [Google Scholar] [CrossRef] [PubMed]
  221. Cockbain, A.J.; Volpato, M.; Race, A.D.; Munarini, A.; Fazio, C.; Belluzzi, A.; Loadman, P.M.; Toogood, G.J.; Hull, M.A. Anticolorectal cancer activity of the omega-3 polyunsaturated fatty acid eicosapentaenoic acid. Gut 2014, 63, 1760–1768. [Google Scholar] [CrossRef] [PubMed]
  222. Song, M.; Chan, A.T.; Fuchs, C.S.; Ogino, S.; Hu, F.B.; Mozaffarian, D.; Ma, J.; Willett, W.C.; Giovannucci, E.L.; Wu, K. Dietary intake of fish, omega-3 and omega-6 fatty acids and risk of colorectal cancer: A prospective study in U.S. men and women. Int. J. Cancer 2014, 135, 2413–2423. [Google Scholar] [CrossRef] [PubMed]
  223. Sasazuki, S.; Inoue, M.; Iwasaki, M.; Sawada, N.; Shimazu, T.; Yamaji, T.; Takachi, R.; Tsugane, S.; Japan Public Health Center-Based Prospective Study Group. Intake of n-3 and n-6 polyunsaturated fatty acids and development of colorectal cancer by subsite: Japan Public Health Center-based prospective study. Int. J. Cancer 2011, 129, 1718–1729. [Google Scholar] [CrossRef] [PubMed]
  224. Mocellin, M.C.; Pastore e Silva Jde, A.; Camargo Cde, Q.; Fabre, M.E.; Gevaerd, S.; Naliwaiko, K.; Moreno, Y.M.; Nunes, E.A.; Trindade, E.B. Fish oil decreases C-reactive protein/albumin ratio improving nutritional prognosis and plasma fatty acid profile in colorectal cancer patients. Lipids 2013, 48, 879–888. [Google Scholar] [CrossRef] [PubMed]
  225. Mocellin, M.C.; Camargo, C.Q.; Nunes, E.A.; Fiates, G.M.R.; Trindade, E. A systematic review and meta-analysis of the n-3 polyunsaturated fatty acids effects on inflammatory markers in colorectal cancer. Clin. Nutr. 2016, 35, 359–369. [Google Scholar] [CrossRef] [PubMed]
  226. Sorensen, L.S.; Thorlacius-Ussing, O.; Schmidt, E.B.; Rasmussen, H.H.; Lundbye-Christensen, S.; Calder, P.C.; Lindorff-Larsen, K. Randomized clinical trial of perioperative omega-3 fatty acid supplements in elective colorectal cancer surgery. Br. J. Surg. 2014, 101, 33–42. [Google Scholar] [CrossRef] [PubMed]
  227. Ma, C.J.; Wu, J.M.; Tsai, H.L.; Huang, C.W.; Lu, C.Y.; Sun, L.C.; Shih, Y.L.; Chen, C.W.; Chuang, J.F.; Wu, M.H.; et al. Prospective double-blind randomized study on the efficacy and safety of an n-3 fatty acid enriched intravenous fat emulsion in postsurgical gastric and colorectal cancer patients. Nutr. J. 2015, 14, 9. [Google Scholar] [CrossRef] [PubMed]
  228. Holscher, H.D.; Bauer, L.L.; Gourineni, V.; Pelkman, C.L.; Fahey, G.C., Jr.; Swanson, K.S. Agave Inulin Supplementation Affects the Fecal Microbiota of Healthy Adults Participating in a Randomized, Double-Blind, Placebo-Controlled, Crossover Trial. J. Nutr. 2015, 145, 2025–2032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Limburg, P.J.; Mahoney, M.R.; Ziegler, K.L.; Sontag, S.J.; Schoen, R.E.; Benya, R.; Lawson, M.J.; Weinberg, D.S.; Stoffel, E.; Chiorean, M.; et al. Randomized phase II trial of sulindac, atorvastatin, and prebiotic dietary fiber for colorectal cancer chemoprevention. Cancer Prev. Res. 2011, 4, 259–269. [Google Scholar] [CrossRef] [PubMed]
  230. Mehta, R.S.; Nishihara, R.; Cao, Y.; Song, M.; Mima, K.; Qian, Z.R.; Nowak, J.A.; Kosumi, K.; Hamada, T.; Masugi, Y.; et al. Association of Dietary Patterns With Risk of Colorectal Cancer Subtypes Classified by Fusobacterium nucleatum in Tumor Tissue. JAMA Oncol. 2017, 3, 921–927. [Google Scholar] [CrossRef] [PubMed]
  231. Ben, Q.; Sun, Y.; Chai, R.; Qian, A.; Xu, B.; Yuan, Y. Dietary fiber intake reduces risk for colorectal adenoma: A meta-analysis. Gastroenterology 2014, 146, 689–699.e6. [Google Scholar] [CrossRef] [PubMed]
  232. Hansen, L.; Skeie, G.; Landberg, R.; Lund, E.; Palmqvist, R.; Johansson, I.; Dragsted, L.O.; Egeberg, R.; Johnsen, N.F.; Christensen, J.; et al. Intake of dietary fiber, especially from cereal foods, is associated with lower incidence of colon cancer in the HELGA cohort. Int. J. Cancer 2012, 131, 469–478. [Google Scholar] [CrossRef] [PubMed]
  233. Kunzmann, A.T.; Coleman, H.G.; Huang, W.Y.; Kitahara, C.M.; Cantwell, M.M.; Berndt, S.I. Dietary fiber intake and risk of colorectal cancer and incident and recurrent adenoma in the Prostate, Lung, Colorectal, and Ovarian Cancer Screening Trial. Am. J. Clin. Nutr. 2015, 102, 881–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Murphy, N.; Norat, T.; Ferrari, P.; Jenab, M.; Bueno-de-Mesquita, B.; Skeie, G.; Dahm, C.C.; Overvad, K.; Olsen, A.; Tjonneland, A.; et al. Dietary fibre intake and risks of cancers of the colon and rectum in the European prospective investigation into cancer and nutrition (EPIC). PLoS ONE 2012, 7, e39361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Mathers, J.C.; Movahedi, M.; Macrae, F.; Mecklin, J.-P.; Moeslein, G.; Olschwang, S.; Eccles, D.; Evans, G.; Maher, E.R.; Bertario, L.; et al. Long-term effect of resistant starch on cancer risk in carriers of hereditary colorectal cancer: An analysis from the CAPP2 randomised controlled trial. Lancet Oncol. 2012, 13, 1242–1249. [Google Scholar] [CrossRef]
  236. Wu, Q.J.; Yang, Y.; Vogtmann, E.; Wang, J.; Han, L.H.; Li, H.L.; Xiang, Y.B. Cruciferous vegetables intake and the risk of colorectal cancer: A meta-analysis of observational studies. Ann. Oncol. 2013, 24, 1079–1087. [Google Scholar] [CrossRef] [PubMed]
  237. Navarro, S.L.; Neuhouser, M.L.; Cheng, T.D.; Tinker, L.F.; Shikany, J.M.; Snetselaar, L.; Martinez, J.A.; Kato, I.; Beresford, S.A.; Chapkin, R.S.; et al. The Interaction between Dietary Fiber and Fat and Risk of Colorectal Cancer in the Women’s Health Initiative. Nutrients 2016, 8, 779. [Google Scholar] [CrossRef] [PubMed]
  238. Rein, M.J.; Renouf, M.; Cruz-Hernandez, C.; Actis-Goretta, L.; Thakkar, S.K.; da Silva Pinto, M. Bioavailability of bioactive food compounds: A challenging journey to bioefficacy. Br. J. Clin. Pharmacol. 2013, 75, 588–602. [Google Scholar] [CrossRef] [PubMed]
  239. Wesolowska, O.; Wisniewski, J.; Sroda-Pomianek, K.; Bielawska-Pohl, A.; Paprocka, M.; Dus, D.; Duarte, N.; Ferreira, M.J.; Michalak, K. Multidrug resistance reversal and apoptosis induction in human colon cancer cells by some flavonoids present in citrus plants. J. Nat. Prod. 2012, 75, 1896–1902. [Google Scholar] [CrossRef] [PubMed]
  240. Manach, C.; Williamson, G.; Morand, C.; Scalbert, A.; Remesy, C. Bioavailability and bioefficacy of polyphenols in humans. I. Review of 97 bioavailability studies. Am. J. Clin. Nutr. 2005, 81, 230s–242s. [Google Scholar] [CrossRef] [PubMed]
  241. Williamson, G.; Manach, C. Bioavailability and bioefficacy of polyphenols in humans. II. Review of 93 intervention studies. Am. J. Clin. Nutr. 2005, 81, 243s–255s. [Google Scholar] [CrossRef] [PubMed]
  242. Karakaya, S. Bioavailability of phenolic compounds. Crit. Rev. Food Sci. Nutr. 2004, 44, 453–464. [Google Scholar] [CrossRef] [PubMed]
  243. D’Archivio, M.; Filesi, C.; Vari, R.; Scazzocchio, B.; Masella, R. Bioavailability of the polyphenols: Status and controversies. Int. J. Mol. Sci. 2010, 11, 1321–1342. [Google Scholar] [CrossRef] [PubMed]
  244. Abourashed, E.A. Bioavailability of Plant-Derived Antioxidants. Antioxidants 2013, 2, 309–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Seeram, N.P.; Henning, S.M.; Zhang, Y.; Suchard, M.; Li, Z.; Heber, D. Pomegranate juice ellagitannin metabolites are present in human plasma and some persist in urine for up to 48 hours. J. Nutr. 2006, 136, 2481–2485. [Google Scholar] [CrossRef] [PubMed]
  246. Lu, M.; Ho, C.T.; Huang, Q. Extraction, bioavailability, and bioefficacy of capsaicinoids. J. Food Drug Anal. 2017, 25, 27–36. [Google Scholar] [CrossRef] [PubMed]
  247. Vissers, M.N.; Zock, P.L.; Katan, M.B. Bioavailability and antioxidant effects of olive oil phenols in humans: A review. Eur. J. Clin. Nutr. 2004, 58, 955–965. [Google Scholar] [CrossRef] [PubMed]
  248. Hewlings, S.J.; Kalman, D.S. Curcumin: A Review of Its’ Effects on Human Health. Foods 2017, 6, 92. [Google Scholar] [CrossRef] [PubMed]
  249. Mereles, D.; Hunstein, W. Epigallocatechin-3-gallate (EGCG) for clinical trials: More pitfalls than promises? Int. J. Mol. Sci. 2011, 12, 5592–5603. [Google Scholar] [CrossRef] [PubMed]
  250. Calvo-Castro, L.A.; Schiborr, C.; David, F.; Ehrt, H.; Voggel, J.; Sus, N.; Behnam, D.; Bosy-Westphal, A.; Frank, J. The Oral Bioavailability of Trans-Resveratrol from a Grapevine-Shoot Extract in Healthy Humans is Significantly Increased by Micellar Solubilization. Mol. Nutr. Food Res. 2018, 62, e1701057. [Google Scholar] [CrossRef] [PubMed]
  251. Li, Y.; Yao, J.; Han, C.; Yang, J.; Chaudhry, M.T.; Wang, S.; Liu, H.; Yin, Y. Quercetin, Inflammation and Immunity. Nutrients 2016, 8, 167. [Google Scholar] [CrossRef] [PubMed]
  252. Yang, Z.; Kulkarni, K.; Zhu, W.; Hu, M. Bioavailability and pharmacokinetics of genistein: Mechanistic studies on its ADME. Anticancer Agents Med. Chem. 2012, 12, 1264–1280. [Google Scholar] [CrossRef] [PubMed]
  253. Mueller, D.; Jung, K.; Winter, M.; Rogoll, D.; Melcher, R.; Richling, E. Human intervention study to investigate the intestinal accessibility and bioavailability of anthocyanins from bilberries. Food Chem. 2017, 231, 275–286. [Google Scholar] [CrossRef] [PubMed]
  254. Choy, Y.Y.; Jaggers, G.K.; Oteiza, P.I.; Waterhouse, A.L. Bioavailability of intact proanthocyanidins in the rat colon after ingestion of grape seed extract. J. Agric. Food Chem. 2013, 61, 121–127. [Google Scholar] [CrossRef] [PubMed]
  255. Rollyson, W.D.; Stover, C.A.; Brown, K.C.; Perry, H.E.; Stevenson, C.D.; McNees, C.A.; Ball, J.G.; Valentovic, M.A.; Dasgupta, P. Bioavailability of capsaicin and its implications for drug delivery. J. Control. Release 2014, 196, 96–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Johnson, J.J.; Nihal, M.; Siddiqui, I.A.; Scarlett, C.O.; Bailey, H.H.; Mukhtar, H.; Ahmad, N. Enhancing the bioavailability of resveratrol by combining it with piperine. Mol. Nutr. Food Res. 2011, 55, 1169–1176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Kaur, S.; Sharma, R.; Sarangal, V.; Kaur, N.; Prashar, P. Evaluation of anti-inflammatory effects of systemically administered curcumin, lycopene and piperine as an adjunct to scaling and root planing: A clinical study. Ayu 2017, 38, 117–121. [Google Scholar] [CrossRef] [PubMed]
  258. Amagase, H.; Petesch, B.L.; Matsuura, H.; Kasuga, S.; Itakura, Y. Intake of garlic and its bioactive components. J. Nutr. 2001, 131, 955S–962S. [Google Scholar] [CrossRef] [PubMed]
  259. Toden, S.; Goel, A. The Holy Grail of Curcumin and Its Efficacy in Various Diseases: Is Bioavailability Truly a Big Concern? J. Restor. Med. 2017, 6, 27–36. [Google Scholar] [CrossRef]
  260. Salehi, B.; Mishra, A.P.; Nigam, M.; Sener, B.; Kilic, M.; Sharifi-Rad, M.; Fokou, P.V.T.; Martins, N.; Sharifi-Rad, J. Resveratrol: A Double-Edged Sword in Health Benefits. Biomedicines 2018, 6, 91. [Google Scholar] [CrossRef] [PubMed]
  261. Khan, F.; Niaz, K.; Maqbool, F.; Ismail Hassan, F.; Abdollahi, M.; Nagulapalli Venkata, K.C.; Nabavi, S.M.; Bishayee, A. Molecular Targets Underlying the Anticancer Effects of Quercetin: An Update. Nutrients 2016, 8, 529. [Google Scholar] [CrossRef] [PubMed]
  262. Shanmukhan, N.K.; Radhakishnan, A.; Maniarasan, V.; Kuppusamy, G.; Gideon, S. Bioformulative concepts on intracellular organ specific bioavailability. Ther. Deliv. 2018, 9, 775–796. [Google Scholar] [CrossRef] [PubMed]
  263. Terao, J. Factors modulating bioavailability of quercetin-related flavonoids and the consequences of their vascular function. Biochem. Pharmacol. 2017, 139, 15–23. [Google Scholar] [CrossRef] [PubMed]
  264. Mukund, V.; Mukund, D.; Sharma, V.; Mannarapu, M.; Alam, A. Genistein: Its role in metabolic diseases and cancer. Crit. Rev. Oncol. Hematol. 2017, 119, 13–22. [Google Scholar] [CrossRef] [PubMed]
  265. Fang, J. Bioavailability of anthocyanins. Drug Metab. Rev. 2014, 46, 508–520. [Google Scholar] [CrossRef] [PubMed]
  266. Grimes, K.L.; Stuart, C.M.; McCarthy, J.J.; Kaur, B.; Cantu, E.J.; Forester, S.C. Enhancing the Cancer Cell Growth Inhibitory Effects of Table Grape Anthocyanins. J. Food Sci. 2018, 83, 2369–2374. [Google Scholar] [CrossRef] [PubMed]
  267. Shoji, T. Chapter 26—Chemical Properties, Bioavailability, and Metabolomics of Fruit Proanthocyanidins. In Polyphenols in Human Health and Disease; Watson, R.R., Preedy, V.R., Zibadi, S., Eds.; Academic Press: San Diego, CA, USA, 2014; pp. 339–351. [Google Scholar] [CrossRef]
  268. Zhu, Y.; Zhang, J.; Zheng, Q.; Wang, M.; Deng, W.; Li, Q.; Firempong, C.K.; Wang, S.; Tong, S.; Xu, X.; et al. In vitro and in vivo evaluation of capsaicin-loaded microemulsion for enhanced oral bioavailability. J. Sci. Food Agric. 2015, 95, 2678–2685. [Google Scholar] [CrossRef] [PubMed]
  269. Rather, R.A.; Bhagat, M. Cancer Chemoprevention and Piperine: Molecular Mechanisms and Therapeutic Opportunities. Front. Cell Dev. Biol. 2018, 6, 10. [Google Scholar] [CrossRef] [PubMed]
  270. Borlinghaus, J.; Albrecht, F.; Gruhlke, M.C.; Nwachukwu, I.D.; Slusarenko, A.J. Allicin: Chemistry and biological properties. Molecules 2014, 19, 12591–12618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Khayata, N.; Abdelwahed, W.; Chehna, M.F.; Charcosset, C.; Fessi, H. Preparation of vitamin E loaded nanocapsules by the nanoprecipitation method: From laboratory scale to large scale using a membrane contactor. Int. J. Pharm. 2012, 423, 419–427. [Google Scholar] [CrossRef] [PubMed]
  272. Maeda, H.; Nakamura, H.; Fang, J. The EPR effect for macromolecular drug delivery to solid tumors: Improvement of tumor uptake, lowering of systemic toxicity, and distinct tumor imaging in vivo. Adv. Drug Deliv. Rev. 2013, 65, 71–79. [Google Scholar] [CrossRef] [PubMed]
  273. Veiseh, O.; Tang, B.C.; Whitehead, K.A.; Anderson, D.G.; Langer, R. Managing diabetes with nanomedicine: Challenges and opportunities. Nat. Rev. Drug Discov. 2015, 14, 45–57. [Google Scholar] [CrossRef] [PubMed]
  274. Yallapu, M.M.; Gupta, B.K.; Jaggi, M.; Chauhan, S.C. Fabrication of curcumin encapsulated PLGA nanoparticles for improved therapeutic effects in metastatic cancer cells. J. Colloid Interface Sci. 2010, 351, 19–29. [Google Scholar] [CrossRef] [PubMed]
  275. Radu, I.C.; Hudita, A.; Zaharia, C.; Stanescu, P.O.; Vasile, E.; Iovu, H.; Stan, M.; Ginghina, O.; Galateanu, B.; Costache, M.; et al. Poly(HydroxyButyrate-co-HydroxyValerate) (PHBHV) Nanocarriers for Silymarin Release as Adjuvant Therapy in Colo-rectal Cancer. Front. Pharmacol. 2017, 8, 508. [Google Scholar] [CrossRef] [PubMed]
  276. Cisterna, B.A.; Kamaly, N.; Choi, W.I.; Tavakkoli, A.; Farokhzad, O.C.; Vilos, C. Targeted nanoparticles for colorectal cancer. Nanomedicine 2016, 11, 2443–2456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Fathi, M.; Martín, Á.; McClements, D.J. Nanoencapsulation of food ingredients using carbohydrate based delivery systems. Trends Food Sci. Technol. 2014, 39, 18–39. [Google Scholar] [CrossRef]
  278. Alvani, K.; Qi, X.; Tester, R. Used of carbohydrates, including dextrin for oral delivery. Starch-Stärke 2011, 63, 424–431. [Google Scholar] [CrossRef]
  279. Yamashita, C.; Chung, M.M.S.; dos Santos, C.; Mayer, C.R.M.; Moraes, I.C.F.; Branco, I.G. Microencapsulation of an anthocyanin-rich blackberry (Rubus spp.) by product extract by freeze-drying. LWT Food Sci. Technol. 2017, 84, 256–262. [Google Scholar] [CrossRef]
  280. Wang, S.; Chen, Y.; Liang, H.; Chen, Y.; Shi, M.; Wu, J.; Liu, X.; Li, Z.; Liu, B.; Yuan, Q.; et al. Intestine-Specific Delivery of Hydrophobic Bioactives from Oxidized Starch Microspheres with an Enhanced Stability. J. Agric. Food Chem. 2015, 63, 8669–8675. [Google Scholar] [CrossRef] [PubMed]
  281. Rokhade, A.P.; Agnihotri, S.A.; Patil, S.A.; Mallikarjuna, N.N.; Kulkarni, P.V.; Aminabhavi, T.M. Semi-interpenetrating polymer network microspheres of gelatin and sodium carboxymethyl cellulose for controlled release of ketorolac tromethamine. Carbohydr. Polym. 2006, 65, 243–252. [Google Scholar] [CrossRef]
  282. Nienaltowska, K.; Perfetti, G.; Meesters, G.M.H.; Ronsse, F.; Pieters, J.G.; Dewettinck, K. Attrition strength of water-soluble cellulose derivatives coatings. Adv. Powder Technol. 2010, 198, 298–309. [Google Scholar] [CrossRef]
  283. Chomto, P.; Nunthanid, J. Physicochemical and powder characteristics of various citrus pectins and their application for oral pharmaceutical tablets. Carbohydr. Polym. 2017, 174, 25–31. [Google Scholar] [CrossRef] [PubMed]
  284. Akolade, J.O.; Oloyede, H.O.B.; Onyenekwe, P.C. Encapsulation in chitosan-based polyelectrolyte complexes enhances antidiabetic activity of curcumin. J. Funct. Foods 2017, 35, 584–594. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the polyphenols’ major mechanisms of action on colorectal cancer (CRC) cells. Polyphenols: (i) Determine cell cycle arrest by downregulating cyclin-dependent kinase 2, 4 (CDK 2, 4), (ii) induce apoptosis by upregulating B cell lymphoma 2 associated protein X and B cell lymphoma 2 (Bax-Bcl-2), caspase-3 and caspase-8, (iii) inhibit cell adhesion and migration through the downregulation of matrix metalloproteinase 9 (MMP-9) and cyclooxygenase 2 (COX2) and (iv) stop tumor initiation by altering nuclear factor kappa-light-chain-enhancer of B cells (NF-kb)/β-catenin signaling pathway.
Figure 1. Schematic representation of the polyphenols’ major mechanisms of action on colorectal cancer (CRC) cells. Polyphenols: (i) Determine cell cycle arrest by downregulating cyclin-dependent kinase 2, 4 (CDK 2, 4), (ii) induce apoptosis by upregulating B cell lymphoma 2 associated protein X and B cell lymphoma 2 (Bax-Bcl-2), caspase-3 and caspase-8, (iii) inhibit cell adhesion and migration through the downregulation of matrix metalloproteinase 9 (MMP-9) and cyclooxygenase 2 (COX2) and (iv) stop tumor initiation by altering nuclear factor kappa-light-chain-enhancer of B cells (NF-kb)/β-catenin signaling pathway.
Ijms 19 03787 g001
Figure 2. Schematic representation of the non-polyphenolic compounds’ major mechanisms of action on CRC cells. Non-polyphenolic compounds: (i) Determine cell cycle arrest by downregulating cyclin-dependent kinase 1 (CDK 1), (ii) induce apoptosis by upregulating B cell lymphoma 2 associated protein X (Bax), B cell lymphoma 2 (Blc-2) and by increasing the production of reactive oxygen species (ROS), (iii) exert anti-inflammatory effects by downregulating interleukin 1 (IL-1), interleukin 2 (IL-2), interleukin 6 (IL-6) and cyclooxygenase 2 (COX2), (iv) inhibit angiogenesis and invasion by downregulating matrix metalloproteinase 7 (MMP-7) and phosphoinositide 3-kinase (PKI3)/Protein kinase B (Akt) and (v) stop tumor initiation by altering the PKI3/Akt pathway.
Figure 2. Schematic representation of the non-polyphenolic compounds’ major mechanisms of action on CRC cells. Non-polyphenolic compounds: (i) Determine cell cycle arrest by downregulating cyclin-dependent kinase 1 (CDK 1), (ii) induce apoptosis by upregulating B cell lymphoma 2 associated protein X (Bax), B cell lymphoma 2 (Blc-2) and by increasing the production of reactive oxygen species (ROS), (iii) exert anti-inflammatory effects by downregulating interleukin 1 (IL-1), interleukin 2 (IL-2), interleukin 6 (IL-6) and cyclooxygenase 2 (COX2), (iv) inhibit angiogenesis and invasion by downregulating matrix metalloproteinase 7 (MMP-7) and phosphoinositide 3-kinase (PKI3)/Protein kinase B (Akt) and (v) stop tumor initiation by altering the PKI3/Akt pathway.
Ijms 19 03787 g002
Table 1. Main classes of phenolic compounds with representative members, chemical structure and dietary sources [58,72,73,74].
Table 1. Main classes of phenolic compounds with representative members, chemical structure and dietary sources [58,72,73,74].
Dietary CompoundsChemical StructureRepresentative CompoundsSources
Flavonoids
Flavones Ijms 19 03787 i001Apigenin, luteolinCelery (Apium graveolens L.), onions (Allium cepa L.), broccoli (Brassica oleracea italic Plenck.), red orange (Citrus x sinensis L. Osbeck), green pepper (Capsicum sp.)
Flavonols Ijms 19 03787 i002Quercetin, kaempferol, myricetinBroccoli (Brassica oleracea italic Plenck.), lettuce (Lactiva sativa L.), onions (Allium cepa L.), kale (Brassica oleracea L.), leek (Allium ampeloprasum L.), apricot (Prunus armeniaca L.), apple (Malus domestica Borkh.), blueberry (Vaccinium corymbosum Rydb.), black currant (Ribes nigrum L.)
Flavonones Ijms 19 03787 i003Naringenin, hesperitin, diosmetinOrange (Citrus x aurantium sp. L.), grepfruit (Citrus paradise Macfad.), lemon (Citrus limon L.)
Isoflavones Ijms 19 03787 i004Genistein, daidzein, glyciteinSoybeans (Glycine max L.) boiled, miso, tofu, soy milk, soy flour
Flavan-3-Ol
Monomeric catechins
Ijms 19 03787 i005Catechin, epicatechin, epigallocatechin, epigallocatechin-3-O-gallateApple (Malus domestica Borkh.), peach (Prunus persica L.), green tea, black tea (Camelia sinensis L.), red wine, grapes (Vitis vinifera L.), beans (Phaseolus vulgaris L.)
Protoanthocyanidins or condensed tannins Ijms 19 03787 i006
Protoanthocyanidin B
Proanthocyanidin A, BAmerican cranberry (Vaccinium macrocarpon Aiton.), cranberry (Vaccinium oxycoccus L.), chokeberries (Aronia melanocarpa Elliott.), plum (Prunus domestica L.), blueberry (Vaccinium myrtillus L.), red currant (Ribes rubrum L.), lingonberries (Vaccinium vitis-idaea L.), pear (Pyrus communis L.), apple (Malus domestica Borkh.), almonds (Prunus dulcis Mill.), hazelnuts (Corylus spp.), pecans (Carya illinoinensis Wangenh.), pistachio nuts (Pistacia vera L.), walnuts (Juglans regia L.), grape seeds (Vitis vinifera L.), beans (Phaseolus vulgaris L.), cowpea (Vigna unguiculata Walp.), lentils (Lens culinaris Medik.), barley (Hordeum vulgare L.), rice (Oryza sativa L.), sorghum (Sorghum bicolor L. Moench.)
Anthocyanidins/Anthocianins Ijms 19 03787 i007Delphinidin (R1, OH; R2, OH; R3, OH);
cyanidin (R1, OH; R2, OH; R3, H)
malvidin (R1, OH; R2, OCH3; R3, OCH3); pelargonidin (R1, OH; R2, H; R3, H)
Blackberry (Rubus fruticosus L.), raspberry (Rubus idaeus L.), bilberry (Vaccinium myrtillus L.), plum (Prunus sp), sour cherry (Prunus cerasus L.), black grapes (Vitis vinifera L.), elderberry (Sambucus nigra L.), red cabbage (Brassica oleracea L.)
Non-Flavonoids
Tannins
A. Gallotannins derivatives
Ijms 19 03787 i008Ellagic acidRaspberry (Rubus idaeus L.), strawberry (Fragaria x ananassa Duchesne.), pomegranate (Punica granatum L.), black currants (Ribes nigrum L.), blackberry (Rubus fruticosus L.), guava (Psidium guajava L.)
B. Elagotannins derivatives Ijms 19 03787 i009Gallic acid
Phenol-Carboxilic Acids
A. Hydroxybenzoic acids
Ijms 19 03787 i010Gallic acid (R1, OH; R2, OH; R3, OH),
syringic acid (R1, OCH3, R2, OCH3, R3, OH)
Broccoli (Brassica oleracea italica Plenck.), cowpea (Vigna unguiculata Walp.), spinach (Spinacia oleracea L.), black currant (Ribes nigrum L.), acai palm (Euterpe oleracea Mart.), sugar apple (Ananona squamosa L.)
B. Hydroxy-cinnamic acids Ijms 19 03787 i011Cinnamic acid (R1, H; R2, H; R3, H; R4, OH),
Caffeic acid (R1, OH; R2, OH; R3, H; R4, OH),
chlorogenic acid,
ferulic acid acid (R1, H; R2, OH; R3, OCH3; R4, OH),
sinapic acid (R1, OCH3; R2, OH; R3, OCH3; R4, OH),
p-coumaric acid (R1, H; R2, OH; R3, H; R4, OH),
rosmarinic acid
Pear (Pyrus communis L.), kiwi (Actinidia deliciosa C.F.Liang and A.R.Ferguson), plum (Prunus domestica L.), apple (Malus domestica Borkh.), artichoke (Cynara scolymus L.), potato (Solanum tuberosum L.), coffee (Coffea arabica L.), rosemary (Rosmarinus officinalis L.), basil (Ocimum basilicum L.), oregano (Origanum vulgare L.), sage (Salvia officinalis L.)
Lignans Ijms 19 03787 i012Secoisolariciresinol, sesaminol, sesamol, sesaminFlaxseed/ linseed (Linum usitassimum L.), sesameseed (Sesamum indicum L.)
Stilbens Ijms 19 03787 i013Resveratrol
(R1, OH; R2, OH; R3, OH)
Red wine, red grapes (Vitis vinifera L.), plum (Prunus domestica L.)
Other Compounds Ijms 19 03787 i014CurcuminTurmeric (Curcuma longa L.)
Ijms 19 03787 i015GingerolGinger (Zingiberis officinale Roscoe.)
Table 2. Main classes of dietary compounds (others then phenolic compounds) with representative members, chemical structure and dietary sources [52,55,62,66,72,73,75,76,77,78,79].
Table 2. Main classes of dietary compounds (others then phenolic compounds) with representative members, chemical structure and dietary sources [52,55,62,66,72,73,75,76,77,78,79].
Dietary CompoundsChemical StructureRepresentative CompoundsSources
Carotenoids Ijms 19 03787 i016LycopeneTomatoes (Solanum lycopersicum L.), guava (Psidium guajava L.), pumpkin (Cucurbita pepo L.), carrot (Daucus carota subsp. sativus), rose hip (Rosa rugosa Thunb.), watermelon (Citrullus lanatus Thunb.), saffron (Crocus sativa L.)
Ijms 19 03787 i017α-carotene
Ijms 19 03787 i018β-Carotene
Ijms 19 03787 i019Crocetin
Iridoids Ijms 19 03787 i020OleuropeinUnripe olive fruits, table olives, virgin olive oil (Olea europaea L.)
Nitrogen Compounds
A. Alkaloids
Ijms 19 03787 i021PiperineBlack pepper (Piper nigrum L.); long pepper (Piper longum L.)
B. Non-alkaloidic compounds Ijms 19 03787 i022CapsaicinChilli pepper (Capsicum sp.)
Organosulfur Compounds Ijms 19 03787 i023AllicinGarlic (Allium sativum L.), onion (Allium cepa L.), broccoli (Brassica oleracea italica Plenck.)
Ijms 19 03787 i024Allin
Ijms 19 03787 i025Sulforaphane
Phytosterols Ijms 19 03787 i026β-sitosterolZucchini (Cucurbita pepo L.), lettuce (Lactuca sativa L.), mangold (Betta vulgaris L.), white cabbage (Brassica oleracea L.), pumpkin seed (Cucurbita pepo L.), oat (Avena sativa L.), peanut (Arachys hypogaea L.), mushrooms (Agaricus bisporus, Lentinula edodes, Grifolia frondosa, Boletus edulis, Pleurotus osteatrus, Armillaria mellea)
Ijms 19 03787 i027Stigmasterol
Ijms 19 03787 i028Ergosterol
Ijms 19 03787 i029Campesterol
Essential Oil Compounds Ijms 19 03787 i030ThymolEssential oils of thyme (Thymus vulgaris L.), wild thyme (Thymus serphyllum L.), oregano (Origanum vulgare L.), lemon balm (Melissa officinalis L.), cinnamon (Cinnamomum sp.), aniseed (Pimpinella anisum L.), star anise (Illicium verum Hook.), fennel (Foeniculum vulgare Mill.), black cumin (Nigella sativa L.)
Ijms 19 03787 i031Carvacrol
Ijms 19 03787 i032Citronellal
Ijms 19 03787 i033Cinnamaldehyde
Ijms 19 03787 i034Anethole
Ijms 19 03787 i035Thymoquinone
Polyunsaturated Fatty Acids
A. Omega-3 fatty acids
Ijms 19 03787 i036α-linolenic acid (ALA)Catfish (Silurus glanis), anchovy (Engraulis encrasicolus), bluefish (Pomatomus saltatrix), sardine (Sardina pilchardus), tuna (Thunnus thynnus), spiny lobster (Panulirus argus), mussels (Mytilus edulis), oyster (Ostrea lurida), herring (Clupea harengus membras, C. h. harengus), mullet (Mugil cephalus), shrimps (Palaemon serratus), seaweeds (Rhodophyta phylum, Phaeophyceae class, Chlorophyta phylum); seed oil from: Hemp (Cannabis sativa L.), chia (Salvia hispanica L.), echium (Echium plantagineum L.), flax (Linum usitassinum L.); walnut oil (Juglans regia L.)
Ijms 19 03787 i037Docosahexaenoic acid (DHA)
Ijms 19 03787 i038Eicosapentaenoic acid (EPA)
B. Omega-6 fatty acids Ijms 19 03787 i039Linoleic acid (LA)Black currant (Ribes nigrum L.) seed oil, gooseberries (Ribes grossularia L.) seed oil, hemp (Cannabis sativa L.) seed oil; meet from beef (Bos taurus), lamb (Ovis aries), pork (Sus scrofa domesticus)
Ijms 19 03787 i040γ–linolenic acid (GLA)
Dietary Fiber Ijms 19 03787 i041InulinSweet potato (Ipomoea batatas L.), leek (Allium ampeloprasum L.), garlic (Allium sativum L.), onion (Allium cepa L.)
Ijms 19 03787 i042GalactooligosaccharidesApple (Malus domestica Borkh.),
pear (Pyrus communis L.),
potato (Solanum tuberosum L.),
beans (Phaseolus vulgaris L.), brusselsprouts (Brassica oleracea var. gemmifera L.), whole wheat (Triticum L.), whole corn (Zea mays L.), chickpea (Cicer arientinum L.), apricot (Prunus armeaniaca L.)
Ijms 19 03787 i043Fructooligosaccharides
Ijms 19 03787 i044β-glucansMushrooms (Agaricus bisporus, Lentinula edodes, Grifolia frondosa, Boletus edulis, Pleurotus osteatrus, Armillaria mellea)
Ijms 19 03787 i045GlucomannansKonjac (Amorphophallus konjac Koch.)
Table 3. Summary of in vivo experiments regarding CRC and phenolic compounds.
Table 3. Summary of in vivo experiments regarding CRC and phenolic compounds.
Author, YearAnimal ModelsDoses and Duration of AdministrationResults
Genistein
Song S. et al., 2018 [157]Mice with AOM/DSS CRC225 mg/kg diet; 450 mg/kg diet and 900 mg/kg diet—six monthSignificant improvement of colon architectural repair, anti-inflammatory activity ↓ COX-2, TNF-α; ↓ expression of PI3K/AKT pathway; ↑expression of FOXO3, Bax proteins
Zhang Y. et al., 2013 [101]Sprague–Dawley rats with AOM induced CRCPre-treatment diet supplementation with 140 mg/kg—six weeks, before induction of cancerInhibition of aberrant crypt foci, prevention of nuclear β-catenin accumulation, suppression of cyclin D1, c-myc expression and Wnt signaling genes (Wnt1, Wnt5a, Sfrp1, Sfrp5)
Son T.G. et al., 2013 [158]BALB/mice subcutaneously injected with CT26 mouse colon cancer cellsTumor bearing mice are treated with genistein 200 mg/kg 1 day before radiation (5, 10 Gy); evaluation of tumors after 12 h, 3.5 daysgenistein increased progenitor cell survival and cell death after radiation, recovery of intestinal damage after radiation (↑ Ki-67), significant tumor regression for combined treatment
Cyanidin/Pelargonidin/Malvidin
Kang S.Y. et al., 2003 [159]Apcmin mice-mutant mouse lineage predisposed to multiple intestinal neoplasia due to mutations in adeno-matous polyposis coli (APC) gene800 mg/L anthocyanidins/200 mg/L (rich in cyanidin glucosides) in the drinking water and modified diet with 200 g/kg freeze dried cherriesFewer and smaller adenomas in the cecum compared to control
Colon tumor volume was not significantly reduced
Shi N. et al., 2015 [160]Male CRJ:CD-1 (ICR) mice with CRC induced with AOM/DSS2.5%; 5%; 10% freeze-dried strawberries—cyanidine glucoside (1.67%) and pelargonidin glucoside (41.1%)—20 weeksInhibition of tumor development from 100% (control) to 74–44%,
↓ of nitrotyrosine production,
↓ Nf-kb, PI3K/AKT phosphorylation, ↓ COX-2, iNOS expression
Silva R.M. et al., 2015 [161]Wistar rats with AOM induced CRCAdministration of 1% (222 mg/zi) or 2% red grape juice (444 mg/zi), two weeks before AOM or 4 weeks after the last administration of AOM↓ COX-2 mRNA with 1% grape juice before AOM and with 2% juice after the last AOM administration
Fernandez J. et al., 2018 [162]Male Fischer 344 rats with AOM/DSS induced CRC20 g/day/rat of functional sausage with 0.11 anthocyanins (mixture of 1:1 dehydrated strawberries and blackberries powder with 59% cyanidin-3-glucosides and 41% pelargonidin-3-glucosides)Significant reduction of Peyer patches, caecum weight, number of polyps;
Significant reduction of Bilophila wadsworthia—a bacteria that generated high level of H2S and has pro-inflammatory effects
Chlorogenic Acid
Matsunaga K. et al., 2002 [163]Male F344 rats with CRC induced with AOM250 ppm chlorogenic acid was administered one week before and a one week after tumor induction with AOM; study duration 36 weeksSignificant decrease of colon tumors for pre- treatment with chlorogenic acid
Banerjee N. et al., 2016 [164]Sprague Dawly rats with AOM induced CRCPlum (Prunus salicina L.) beverage rich in chlorogenic and neochlorogenic acids, 10 weeksSignificant decrease of dysplastic polyps, ↓ expression of COX-2, Nf-kB, AKT/mTOR signaling pathway, ↑ miR-143
Cinnamic Acid
Zhu B. et al., 2016 [165]Female BALB/c nude mice inoculated with HT29 colon carcinoma cells1 and 1.5 mmol/kg
x3/week for two weeks
Significant inhibition of tumor growth, ↑ expression of Bax and caspase 3
p-Coumaric Acid
Sharma S.H. et al., 2017 [166]Male albino rats with DMH induced CRC50 mg/kg, 100 mg/kg, 200 mg/kg; 15 weeksSignificant dose-dependent reduction of polyps incidence and formation of pre-neoplasic lesions, reduction of oxidative stress; significant decrease in gut microbial enzymes (mucinases and β-dehydrogenases)
Rosmarinic Acid (RA)
Venkatachalam K. et al., 2013 [167]Male wistar rats with DMH induced CRC5 mg/kg received during administration of DMH (15 weeks) or one week after the last DMH dose (until 30 weeks) or through the whole period
(30 weeks)
Supplementation with RA for the whole period showed the highest tumor reduction, ↓ stress oxidative markers, ↓mucosal bacterial enzymes activity, regulation of xenobiotic metabolizing enzymes, up-regulation of apoptotic factors
Gallic Acid
Giftson J.S. et al., 2011 [168]Male albino Wistar rats with DMH induced CRC50 mg/kg received one week before DMH and continued 30 weeks (group 1), after cessation of DMH until 30 weeks (group 2), along the whole period (group 3)Supplementation with gallic acid for the whole period showed the highest tumor reduction, regulation xenobiotic metabolizing enzymes, decreased tumor incidence
Secoisoscilaresinol
Shah N.R. and Patel B.M., 2016 [169]Diabetic male Sprague Dawley rats with DMH induced CRC500 mg/kg p.o secoisolariciresinol rich extract of L. ussitatissimum—18 weeks↓ pro-inflammatory markers, ↓ PCNA, ↓ CEA, ↓ mRNA level of CDK4, reduction in hyperplastic cells
Gomides A.F. et al., 2016 [170]C57 BL6 mice with DMH induced CRC10% defatted flaxseed meal—15 weeksReduction of precancerous lesions in the distal colon
Sesamol
Shimizu S. et al., 2014 [171]C57/BL6-Apc Min/+ mice500 pp/8 weeks↓ pro-inflammatory factors, suppression of intestinal polyps formation
Legend: AOM—azoxymethane, DSS—dextran sulfate sodium; COX2—cycloygenase 2, TNF-α—tumor necrosis factor, PI3K/AKT, Ki-67—proliferative marker, iNOS—nitric oxide synthases, Nf-kB—nuclear factor kappa-light-chain-enhancer of activated B cells, DMH—1,2 dimethylhydrazine, Bcl-2—B-cell lymphoma 2, Bax—Bcl-2associated X protein, PCNA—proliferating cell nuclear antigen, CDK4—cyclin dependent kinase, CEA—carcinogenic embryonic antigen, mir-143—microRNA.
Table 4. Summary of in vivo experiments regarding CRC and non-phenolic compounds.
Table 4. Summary of in vivo experiments regarding CRC and non-phenolic compounds.
Author, YearAnimal ModelsDoses and Duration of AdministrationResults
Lycopene
Tang F.Y. et al., 2011 [173]BALB/cAnN-Foxn1 nude mice with CRC induced by inoculation of HT-29 cells3/6 mg/kg—5 weeksSignificant inhibition of tumor growth
Dias M.C. et al., 2010 [174]Male Wistar rats with DMH induced CRC300 mg/kg lycopen + symbiotic (60 mg oligofructose + 50 mg inulin + 109 CFU Bifidobacteria lactis)—eight weeks (before/during or after initiation with DMH)↓ PCNA, ↓ p-53 colonic cells, ↓ AFC, ↓colonic Paneth cells
Crocin
Kawabata K. et al., 2012 [175]CD1 (ICR) mice with AOM/DSS induced CRC50, 100, 200 ppm for 15 weeks after initiation of colon cancerSignificant reduction of inflammation and mucosal ulcers, multiplicity of adenocarcinoma
Oleuropein
Giner E. et al., 2016 [176]C57BL6 mice with AOM/DSS colorectal induced cancer50 mg/kg or 100 mg/kg—63 weeksInhibition of tumor formation, decreased cell proliferation, anti-inflammatory activity
Capsaicin
Caetano B.F.R. et al., 2018 [177]Male WISTAR rats with DMH induced colorectal cancer5 mg/kg or 50 mg/kg—four weeks↓Ki-67, significant ↓ of tumor volume and number of AFC
Sulforaphane/Organosulfur Compounds
Rajendran P. et al., 2015 [178]Male WT or Nrf2−/+ mice with DMH induced CRCMice received alternating or daily 400 ppm sulforaphane included in the diet for 25 weeksSignificant reduction in tumor multiplicity only after continuous treatment
β-Sitosterol
Baskar A.A. et al., 2012 [179]Male albino Wistar rats with DMH induced CRC5 mg/kg; 10 mg/kg; 20 mg/kg —16 weeksSignificant increase of antioxidant defense system, ↑ GSH, ↓ hyperplasic lesions
Thymoquinone (TQ)
Asfour W. et al., 2013 [180]Male albino rats with DMH induced CRCAdministration of 10 mg/kg for 10 weeks (in the initiation phase + DMH) and 11 weeks (in the post initiation phase, after induction of cancer)Chemopreventive effect, significant inhibition of tumor growth (for simultaneously administration in the initiation phase), ↓ PCNA, inhibition of VEGF production
Kortum B. et al., 2015 [181]Male and female
Msh2 loxP/loxP
Villin-Cre mice—transgenic mice that simulate intestinal carcinogenesis
Mice were divided in 5 groups: Group 1—regular chow; group 2—500 mg mesalazine/kg chow, group 3—2500 mg mesalazine/kg chow; group 4—37.5 mg TQ /kg chow; group 5—375 mg TQ/kg chow; treatment for 43 weeks↓ incidence of tumors dose dependent for TQ; no significant differences between TQ and mesalazine
Cinnamaldehyde
Long M. et al., 2015 [182]Experimental Nrf2+/+ and Nrf2−/− C57BL/6 mice with AOM/DSS induced CRCSupplementation of diet with 0.5% cinnamaldehyde—11 weeksSupplementation significantly attenuated colon carcinogenesis only for Nrf2+/+ mice; anti-inflammatory and antioxidant effects
Carvacrol
Sivaranjani A. et al., 2016 [183]Male Albino WISTAR rats with DMH induced CRCAdministration of 20, 40, 80 mg/kg for 16 weeksReduced tumor incidence, inhibition of aberrant crypts formation, ↑ in antioxidant defense system, ↓ activity of colonic bacterial enzymes
Omega-3/Omega-6 Fatty Acids
Wang W. et al., 2017 [184]C57 BL6 mice with CRC induced by inoculation of MC38 CRC cellsPre-treatment (3 weeks) with DHA diet (omega-6/omega-3 ratio = 1.26:1) and DHA high diet (omega-6/omega-3 ratio = 0.56:1) before tumor initiation (3 weeks)
DHA diet—22 g/kg LA, 0.31 g/kg ALA, 17.2 g/kg DHA; DHA high diet—12.5 g/kg—LA, 0.17 g/kg—ALA, 21.9 g/kg DHA
Inhibition of colon growth, modulation of fatty acids profile in colon tumors (↓ARA, ↑EPA, DHA), ↓ EETS
Hao G.W. et al., 2015 [185]BALB/c nude mice with CRC induced by inoculation of HCT116 colon cancer cellsKetogenic diet with or without omega-3 fatty acids; supplementation received until tumor volume was 600–700 mm3
(45 days)
Delayed tumor growth
↓ tumor vascularity for ketogenic diet supplemented with omega-3 fatty acids
Piazzi G. et al., 2014 [186]C57BLJ/6J mice with AOM/DSS induced CRCEffect of 1% eicosapentaenoic free fatty acid on both initiation and progression of carcinogenesis, 105 daysSuppression of tumor development, increase of apoptosis, anti-inflammatory effects, modulation of gut microbiota
Barone M. et al., 2014 [187]C57BLJ/6J mice with mutation for the Apc gene (Apc Min/+)Supplementation of diet with olive oil and omega-3 fatty acids, 10 weeksDecrease in polyps number, pro-apoptotic effects
Dietary Fibers
Wang J. et al., 2017 [188]Athymic male nude mice BALB/c-nu with CRCobtained by inoculation of HT-29 cancer cellsAdministration of a polysaccharide from Lentinus edodes (0.2 mg/kg; 1 mg/kg; 5 mg/kg) or 20 mg/kg 5-fluorouracil for 21 days after cancer induction↓ tumor growth, pro-apoptotic effects
Masuda Y. et al., 2013 [189]Female BALB/c, BALB/c-nude, C3H/HeJ mice inoculated with colon-26 cancer cellsAdministration of D fraction (β-glucan) from the maitake mushroom (Grifola frondolosa) for 19 days after cancer inductionSignificant decrease of tumor growth through systemic immune responses
Pattananandecha T. et al., 2016 [190]Male Sprague-Dawley rats with AOM induced CRCSupplementation of diet with 10% inulin for 17 weeksReduction of colonic AFC, reduction in bacterial colon enzymes, increase in Lactobacillus sp, Bifidobacteria
Qamar T.R. et al., 2016 [191]Male Wistar rats with DMH induced CRCAdministration of galacto-oligosaccharides (76–151 mg), inulin (114 mg) separately or co-administration for 16 weeksFor co-administration significant ↓ in AFC formation and fecal enzyme activities
Hijova E. et al., 2013 [192]Male and female Sprague-Dawly rats with DMH induced CRCSupplementation of diet with 80 g inulin/kg food for 28 weeksSignificant ↓ coliform counts and ↑ lactobacilli counts, ↓ fecal enzyme activities, anti-inflammatory effects
Verma A. and Shukla G., 2013 [193]Male Sprague Dawly rats with DMH induced colorectal cancerAdministration of inulin 10 mg/0.1 mL for a week before initiation of CRC and 6 weeks after initiation↓ of AFC and nitroreductase/β-glucosidase activity
Stofilova J. et al., 2015 [194]Male and female Sprague Dawly rats with DMH induced colorectal cancerCo-administration of oligofructose-enriched inulin preparation (95% fructan chains and 5% monosaccharide and disaccharide) with 109 CFU/mL for 28 weeks↓ inflammatory process in the jejuna and colon mucosa
Wu W.T. et al., 2014 [195]Male C57/BL/6J with AOM induced colorectal cancerAdministration of high-fat low fibre diet (1% cellulose) or high 5% fibre diet with konjac glucomannan, inulin, cellulose for 3 weeks before cancer initiationKonjac glucomannan and inulin have anti-genotoxic effects, increase cecal short chain fatty-acids, up-regulate antioxidant enzymes genes
Legend: AOM—azoxymethane, DSS—dextran sulfate sodium, DMH—1,2 dimethylhydrazine, PCNA—proliferating cell nuclear antigen, AFC—aberrant crypt foci, Ki-67—proliferative marker in the jejum, GSH—glutathione, VEGF C vascular endothelial growth factor, LA—Linoleic acid, ALA—α-linolenic acid, DHA—docosahexaenoic acid, EPA—eicosapentaenoic acid, ARA—arachidonic acid, EETS—epoxyeicosatrienoic acids, CD1 (ICR)—outbred mouse of Swiss origin.
Table 5. Summary of the association between CRC risk and phenolic/non-phenolic compounds.
Table 5. Summary of the association between CRC risk and phenolic/non-phenolic compounds.
Author, YearDate/Type of StudyCasesControl CasesDoseOR/HR/RR/IRR/SRR
(95%CI)/Observation
Conclusions
Isoflavones (IF)
Akhter M. et al. 2009 [202]2004–2005/control-study721; men and women 40–79 years old69724.77–62.41 mg IF/dayOR
0.49 (0.27–0.90) to 0.53 (0.28–0.98)
p < 0.05
Significant inverse association between high intake of isoflavone consumption and CRC in women
Shin A. et al. 2015 [210]Case control study/2010–2013901 men and women2669Daidzein 3.20–9.89 mg/day
Genistein 3.3–9.7 mg/day
Glycitein 0.85–2.44 mg/day
OR
Daidzein 1.25 (0.96–1.61) to 0.71 (0.54–0.95)
Genistein 1.18 (0.91–1.53) to 0.75 (0.57–1)
Glycitein 1.32 (1.03–1.70) to 0.39 (0.25–0.61)
A high intake of isoflavones is significantly associated with decreased risk of CRC in both men and women
Yang G. et al. 2009 [211]Prospective cohort study /1996–200568,412 women 40–70 years oldNA12.8–21 g soy food intake/day equivalent to 15.1–48.9 mg IF/dayRR
Soy food 0.88 (0.67–1.15) to 0.71 (0.53–0.95)
Soy isoflavones 0.91 (0.69–1.19)
to 0.80 (0.60–1.07)
High intake of soy food products and isoflavones is correlated with reduced incidence in CRC, especially for menopausal women
Ko K. et al. 2018 [205]Case control study in Korean (1993–2004) and Vietnamese population (2003–2007)101 (Korean study)
222 cases (Vietnamese study)
391 (Korean study)
226 (Vietnamese study)
Evaluation of plasma IF levels for patients with CRCOR for genistein 0.67 (0.34–1.31)
to 0.50 (0.25–0.98)—Korean patients
OR for genistein 0.97 (0.54–1.74)
to 0.43 (0.25–0.73)—Vietnamese patients
OR for daidzein (Vietnamese patients) 0.84 (0.47–1.49) to 0.48 (0.28–0.82)
Significant inverse correlation between high isoflavones plasma concentrations and reduced colorectal incidence
Lignans
Principi M, et al. 2013 [208]Randomized double blind placebo-controlled study30 patients30 placeboSupplementation of diet with 750 mg insoluble oat fiber, 50 mg flaxseed dry extract with 20% secoisolariciresinol diglycoside +175 mg milk thistle extract (70% silymarin and 30% silibinin)—60 days prior to colonoscopySignificant increase in ERβ/ERα ratio and activation of caspasesModulation of ERβ receptor is important for a chemo-preventive effect
Calabrese C. et al. 2013 [209]Open study /2012–201311 patients with familial adenomatous polyposis with ileal pouch anal anastomosisNA5 mg Eviendep® (30% silibinin + 40% secoisolariciresinol diglucoside + indigestible fibers 5% lignin) × 2/day for 3 monthSignificant reduction of number and size of polyps with 32% and 51% respectivelyChemo-preventive effect
Zamora-Ros R. et al. 2013 [204]Case control study /1996–19984264010.27–0.50 mg lignans/1000 kcal dayRR for Lignans 0.72 (0.47–1.10) to 0.59 (0.34–0.99)Significant inverse correlation between high intake of lignans and colorectal incidence
Anthocyanidins
Thomasset S. et al. 2009 [212]Pilot study 2006–200815 patients with histological confirmed; 10 patients with colorectal liver metastasisNA1.4/2.8/5.6 g of Mirtocyan (a standardized extract rich in anthocyanidins) for 7 days before surgeryMild decrease of tumor tissue only for 1.4 gPossible chemo-preventive effects in humans
Wang L. S. et al. 2014 [213]Randomized double blind placebo control study14 patients with familial adenomatous polyposisNAGroup I—7 patients placebo powder (60 g/day) + 2 rectal suppositories (each 720 mg freeze-dried black raspberry extract)
Group 2—60 g/day black raspberry freeze dried extract + 2 rectal suppositories—9 month treatment
Reduction of polyps mainly for suppositories
Significantly de-methylated regions in adenomas
Regressing of rectal polyps in patients with familial adenomatous polyposis
Organosulfur Compounds
Tanaka S. et al. 2006 [214]Preliminary double blind randomized clinical trial37 patients with colorectal adenomas which are removed if the size was > 5 mmNAGroup I—6 capsules of aged garlic extract (AGE) equivalent to 2.4 mL AGE/day
Group II—control (low dose) 6 capsules of AGE equivalent to 0.16 mL AGE/day. Patients are evaluated after 6, 12 months
AGE suppressed colorectal adenomas after 6, 12 monthsChemo-preventive effect in humans
McCullough M. et al. 2012 [215]CPSII Nutrition cohort 1999–200742,824 men
56,876 women
NASupplementation of diet with garlic cloves < 1 clove/month; 1–3 cloves/month; 1 clove/week; 2–4 cloves/week; 5–6 cloves/week, 1 clove dailyProtective effect - women HR for 1–3 cloves/week 1.08 (0.86–1.35); 0.95 (0.72–1.26) for 1 clove/week; 0.77(0.58–1.02) for 2–4 cloves/week; 0.74 (0.48–1.13) to 5–6 cloves/week and 0.87 (0.58–1.32) for 1 clove/dayWeak chemo-preventive effect of garlic consumption for women; but not for men
Meng S. et al. 2013 [216]Cohort study 1984–200876,208 women
45,592 men
NAAdministration of garlic cloves < 1 clove/month; 1–3 cloves/month; 1 clove/week; 2–4 cloves/week; 5–6 cloves/week, 1 clove dailyWomen HR 1–3 cloves/month 1.11 (0.94–1.31) compared to HR 1.21 (0.94–1.57) for 1 clove/day (p = 0.14)
Men HR 1–3 cloves/month 0.99 (0.84–1.16) compared to HR 1.03 (0.73–1.45) for 1 clove/day (p = 0.99)
No association was found between garlic intake and CRC risk
Satia J. A. et al. 2009 [217]Cohort study 2000–200242876,084Administration of garlic pills at least once a week for > 1 year during previous 10 yearsHR 1.35 (0.59–1.17) compared to 1.00 (Reference)
p = 0.04
Significant increase of CRC incidence with garlic administration
Carotenoids
Lu M. S. et al. 2015 [218]On-going case control study 2010–2013845845Food frequency questionnaire regarding intake of fruits and vegetables rich in carotenoidsα-carotene
OR 0.54 (0.42–0.70) to 0.41 (0.31–0.54)
β-carotene
OR 0.79 (0.61–1.03) to 0.62 (0.48–0.82)
lycopene
OR 0.66 (0.51–0.85) to 0.45 (0.35–0.60)
p < 0.001
Significant inverse correlation between carotenoids intake and CRC incidence
Leenders M. et al. 2014 [219]Cohort study 1992–2000Colon cancer 898 Rectum cancer 501898/501Food frequency questionnaire regarding intake of fruits and vegetables rich in carotenoids and vitaminsFor colon cancer
β-carotene
OR 0.89 (0.67–1.18) to 0.69 (0.52–0.94)
Vitamin C
OR 0.98 (0.74–1.29) to 0.76 (0.57–1.01)
Vitamin E
OR 0.88 (0.67–1.16) to 0.99 (0.74–1.33)
Significant inverse correlation between CRC incidence and mainly dietary β-carotene, vitamin C intake
Kabat G. C. et al. 2012 [220]Large, prospective, multicenter study88 CRC in post-menopausal women5389Analysis of antioxidants from fasting blood samples at baseline, 1/3/6 years follow-upFor CRC β-carotene
HR 0.65 (0.39–1.09) to 0.54 (0.31–0.96)
For colon cancer β-carotene
HR 0.57 (0.32–1.00) to 0.47 (0.25–0.88)
Significant inverse correlation between β-carotene plasma levels and CRC incidence
Polyunsaturated Fatty Acids
Cockbain A. J. et al., 2014 [221]Phase II double blind randomized placebo control trial
2010–2011
203 patients with CRC liver metastasisNA1.Placebo—43 patients
2. Patients receiving 2 g/day of EPA for 30 days prior to surgery, follow-up 18 months after surgery
Significant higher content of EPA in tumor tissues 1.82% compared to 1.30%
(for placebo) p = 0.0008,
decreased PGE2 in tumor tissues, anti-angiogenic activity (p = 0.075)
Pre-operative treatment has shown provide post-operative benefit
Song M. et al. 2014 [222]Study cohort
1984–2008
76,386 women
47,143 men
NAAdministration of fish 15–40 g/day (women) and 16–46 g/day (men), marine fish (0.15–0.30 g/day)Significant risk of distal colon cancer for both fish intake HR 1.12 (0.85–1.48) to 1.36 (1–1.85) and marine fish HR 1.19 (0.89–1.58) to 1.36 (1.03–1.80)
p ≤ 0.05
Associated risk between marine fish intake and CRC risk
Sasazuki S. at al. 2011 [223]Prospective study
1995–2006
827,833 subjectsNAFood frequency questioners regarding fish intake; marine fish 0.49–2.18 g/day for men and womenSignificant
Decrease associated with marine fish intake only for men
RR 0.97 (0.51–1.83) to 0.35 (0.14–0.88)
p = 0.05
Chemo-preventive effect of marine fish rich in omega-3 fatty acids
Mocellin M. C. et al. 2013 [224]Prospective randomized controlled trial
2011–2012
57 patients with CRC undergoing,
chemotherapy, only 11 are randomized
NA1. Control group (n = 5).
2. Supplemented group (n = 6) with 2g fish oil/day—9 weeks
2 g fish oil = 360 mg/day EPA, 240 mg/day DHA
Significant decrease of C-reactive protein from 18.14 mg/L to 1.14 mg/L (p = 0.04)
Significant increase of EPA, DHA compared to control group (p = 0.014,
p = 0.019), significant decrease of AA between baseline and 9 weeks follow-up for the supplemented group (p = 0.028)
Significant anti-inflammatory effects for patients undergoing chemotheraphy and increase for plasma fatty acid profile
Mocellin M. C. et al. 2016 [225]Meta-analysis of Nine trials475 patients with CRCNASupplementation of diet with omega-3 fatty acids or administration of 0.2 g/kg fish oil parenterally at post-operative period
Patients undergoing chemotherapy supplementation with 0.6 g/day EPA+DHA -9 weeks
Significant decrease of IL-6 (p = 0.024) and increase of albumin (p = 0.014)
Supplementation of EPA+ DHA during chemotherapy significantly reduced CRP concentration (p = 0.017) and CRP/albumin ratio (p = 0.016)
Use of omega-3 fatty acids have benefits, especially for inflammatory markers in CRC patients
Sorensen L. S. et al. 2014 [226]Randomized double blind placebo controlled trial148 patients awaiting for CRC surgeryNA1. Control group
2. Supplemen-tation group with 2 g EPA + 2 g DHA for 7 days before and 7 days after surgery
Pre-operative treatment with omega-3 fatty acids determined a significant increase of EPA, DHA levels in granulocytes and a significant decrease of AA (p < 0.001) compared to control groupPotential immune-stimulatory effects and prevention of post-operative infections
Ma C. J. et al. 2015 [227]Prospective randomized double-blind study 2009–201099 patients with gastric and CRCNA1. Control group
2. Supplementation group with a lipid emulsion containing soybean oil
(80–100 g/L), medium chain triglycerides (100 g/L) and PUFA (linoleic acid—38–58 g/L, α-linolenic acid—4–11 g/L) for 7 days after surgery
There are no significant differences regarding inflammatory markers between the control and the supplementation group.
Significant positive effect on lipid markers
(p < 0.05)
Improvements only in lipid metabolism
Dietary Fiber
Holscher H. D. et al. 2015 [228]Prospective Randomized double-blind placebo 3-period controlled study30 heathy patientsNA1. Placebo group
2. Administration of 5 g or 7.5 g inulin for 21 days with 7 days wash-out between periods
Significant increase of Bifidobacterium sp.
(p < 0.001), significant decrease of Ruminococcus sp. and Desulfovibrio sp. (p < 0.01); dietary intake was positively associated with fecal butyrate
(p = 0.005)
Beneficial changes in gastro-intestinal microbiota are correlated with decreased CRC incidence
Limburg P et al, 2011 [229]Randomized phase II clinical trial 2006–200885 patients with aberrant crypt foci ≥ 5 at baselineNA1. Control group
2. Atorvastatin 20 mg/day
3. Sulindac 150 mg x 2/day
4. Oligo-fructose enriched inulin 6 g powder x 2/day for 6 months
All treatment didn’t provide a significant decrease in AFC number and sizeNo association was found between inulin intake and CRC
Mehta RS et al, 2018 [230]Prospective cohort study 1980–2012121,700 females
51,529 males
NAFood frequency questionnaires regarding dietary fiber intakeHigh intake of fiber was associated with a low risk of Fusobacterium nucleatum positive CRCs
HR 0.54 (0.33–0.89) to 0.40 (0.24–0.67)
Intestinal microbiota plays an important role in mediating the association between consumption of high amount of dietary fiber/whole grains and CRC incidence
Ben Q et al, 2014 [231]Meta-analysis of 20 studies (case-control, cohort)10,984 patients with colorectal adenomaNAAdministration of 10 g/day fibersSRR for dietary fiber are 0.72 (0.63–0.83) in a high vs low intake, inverse association between total fiber intake and CRC risk SSR 0.66
(0.56–0.77)
Chemopreventive effect of dietary fiber
Hansen L et al, 2012 [232]Cohort study, 1997–2008108,081 patientsNAAdministration of fiber 16–28 g/day for men and 15–24 g for womenSignificant inverse correlation between CRC incidence and dietary fiber intake for men
IRR 0.93 (0.68–1.26) to 0.55 (0.38–0.79)
Chemopreventive effect of dietary fiber
Kunzman A et al, 2015 [233]Cohort study 1993–20092036 patients15,976Administration of fiber 9.9–12.8 g/1000 kcal/day from fruits/vegetablesSignificant inverse correlation between dietary fiber intake and distal colon or rectal adenoma in men
OR 0.88 (0.75–1.04) to 0.76 (0.63–0.91)
p = 0.003
Chemopreventive effect of dietary fiber against CRC
Murphy N et al, 2012 [234]On-going multicentre prospective cohort study 1992–2000477,312 patientsNAAdministration of dietary fiber 16.4–28.5 g/daySignificant inverse correlation between dietary intake and colon-distal cancer
HR 0.90(0.75–1.07) to 0.70 (0.53–0.92)
p = 0.021
colon proximal cancer
HR 0.93 (0.78–1.1) to 0.86 (0.69–1.07)
p = 0.16
rectum cancer
HR 1 (0.87–1.17) to 0.79 (0.65–0.96)
p = 0.012
Chemopreventive effect of dietary fiber against CRC
Mathers JC et al, 2012 [235]Randomized control trial937 eligible patients with Lynch syndromeNA1. 463 patients received 30 g resistant starch/day
2. 455 patients—resistant starch placebo
3. 19 patients 600 mg aspirin/day—29 months
No significant effect of resistant-starch administration on cancer development
IRR resistant starch vs resistant starch placebo 1.15 (0.66–2.00)
p = 0.61
No detectable effect on cancer development
Legend: HR—hazard ratio; RR—relative risk; OR—odds ratio; IRR—incidence rate ratio; SRR—summary relative risk; NA, not applicable.
Table 6. Summary of the bioavailability of the dietary compounds.
Table 6. Summary of the bioavailability of the dietary compounds.
Compound of InterestSourceBioavailabilityIn Vivo StudiesRef.
CurcuminTurmeric- poor absorption, rapid metabolism, rapid elimination
- enhanced by piperine with 2000%
- better results on animal experiments
- 2g/kg of curcumin in rats → Cmax = 1.35 ± 0.23 μg/mL in 0.83 h
- 2 g/kg of curcumin in human subjects →
Cmax = 0.006 ± 0.005 μg/mL
[120,248]
EGCGGreen tea- poor absorption
- alkaloids, vitamins, proteins and fish oil improve absorption
- air contact oxidation, metal ions like Ca2+ and Mg2+ and milk reduce absorption
- one oral dose of EGCG
half-time = 3.4 ± 0.3 h
[249]
ResveratrolGrapes- poor absorption
- bioavailability increased by a liquid micellar formulation
- 500 mg of Vineatrol →
Cmax of trans-resveratrol
= 10.6 fold higher
- no detection of trans-ε-viniferin in plasma or urine
[250]
QuercetinOnion
Apples
Beans
Broccoli
- better bioavailability of quercetin glucoside- 100 mg quercetin → absorption of quercetin glucoside = 3–17%[251]
GenisteinSoy- lower bioavailability in vivo
- total genistein—a better bioavailability than genistein aglycone
- 20 mg/kg of genistein in FVB mice → genistein aglycone bioavailability = 23.4%
- soy protein feeding Balb/c mice females → bioavailability of total genistein = 90%
→ bioavailability of genistein aglycone = < 15%
[252]
AnthocyaninsBilberries- poor bioavailability
- 30% of amthocyanins are stable in the upper intestine for 8 h.
- bioavailability of anthocyanins from bilberries → ↑ amount of anthocyanins and degradants in the heathy compared to ileostomists group[253]
Proantho-cyanidinApples
Grapes
Green tea
- oligomeric flavonoids with limited bioavailability- ad libitum diet of grape seed extract in lab rats → the presence of PAC in the colonic contents
- 11% of PAC—present in the feces
[254]
CapsaicinChilli- low bioavailability
- capsaicin was absorbed into intestinal tissues, jejunum and serosal fluid
- 1 mM of capsaicin in rats → absorption = 50% in the stomach, 80% in the jejunum and 70% in the ileum.[255]
PiperineBlack pepper- insoluble in water with a low bioavailability
- used in clinical assays single or as an enhancer for other dietary agents
- improved the bioavailability of resveratrol, curcumin and lycopene
- resveratrol + piperine in mice → piperine enhanced the bioavailability of resveratrol with 229%
- curcumin + piperine in rats and human subjects → piperine enhanced bioavailability of curcumin with 2000%
[256,257]
AliicinGarlic- poor bioavailability- administration of garlic/ pure allicin → no detection in urine or blood[258]

Share and Cite

MDPI and ACS Style

Costea, T.; Hudiță, A.; Ciolac, O.-A.; Gălățeanu, B.; Ginghină, O.; Costache, M.; Ganea, C.; Mocanu, M.-M. Chemoprevention of Colorectal Cancer by Dietary Compounds. Int. J. Mol. Sci. 2018, 19, 3787. https://doi.org/10.3390/ijms19123787

AMA Style

Costea T, Hudiță A, Ciolac O-A, Gălățeanu B, Ginghină O, Costache M, Ganea C, Mocanu M-M. Chemoprevention of Colorectal Cancer by Dietary Compounds. International Journal of Molecular Sciences. 2018; 19(12):3787. https://doi.org/10.3390/ijms19123787

Chicago/Turabian Style

Costea, Teodora, Ariana Hudiță, Oana-Alina Ciolac, Bianca Gălățeanu, Octav Ginghină, Marieta Costache, Constanța Ganea, and Maria-Magdalena Mocanu. 2018. "Chemoprevention of Colorectal Cancer by Dietary Compounds" International Journal of Molecular Sciences 19, no. 12: 3787. https://doi.org/10.3390/ijms19123787

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