Category Archives: DU145

Cryptotanshinone (See also Tanshinone)

Cancer:
Prostate, breast, cervical., leukemia, hepatocellular carcinoma

Action: Anti-inflammatory, cell-cycle arrest, inhibits dihydrotestosterone (DHT), anti-proliferative, hepato-protective

Cryptotanshinone is a major constituent of tanshinones from Salvia miltiorrhiza (Bunge).

Tanshinone IIA and cryptotanshinone could induce CYP3A activity (Qiu et al., 2103).

Anti-proliferative Agent

Cryptotanshinone (CPT), a natural compound, is a potential anti-cancer agent. Chen et al., (2010) have shown that CPT inhibited cancer cell proliferation by arresting cells in G(1)-G(0) phase of the cell-cycle. This is associated with the inhibition of cyclin D1 expression and retinoblastoma (Rb) protein phosphorylation.

Furthermore, they found that CPT inhibited the signaling pathway of the mammalian target of rapamycin (mTOR), a central regulator of cell proliferation. This is evidenced by the findings that CPT inhibited type I insulin-like growth factor I- or 10% fetal bovine serum-stimulated phosphorylation of mTOR, p70 S6 kinase 1, and eukaryotic initiation factor 4E binding protein 1 in a concentration- and time-dependent manner. Expression of constitutively active mTOR conferred resistance to CPT inhibition of cyclin D1 expression and Rb phosphorylation, as well as cell growth. The results suggest that CPT is a novel anti-proliferative agent.

Anti-inflammatory; COX-2, PGE2

Cyclooxygenase-2 (COX-2) is a key enzyme that catalyzes the biosynthesis of prostaglandins from arachidonic acid and plays a critical role in some pathologies including inflammation, neurodegenerative diseases and cancer. Cryptotanshinone is a major constituent of tanshinones and has well-documented anti-oxidative and anti-inflammatory effects.

This study confirmed the remarkable anti-inflammatory effect of cryptotanshinone in the carrageenan-induced rat paw edema model. Since the action of cryptotanshinone on COX-2 has not been previously described, in this study, Jin et al. (2006) examined the effect of cryptotanshinone on cyclooxygenase activity in the exogenous arachidonic acid-stimulated insect sf-9 cells, which highly express human COX-2 or human COX-1, and on cyclooxygenases expression in human U937 promonocytes stimulated by lipopolysaccharide (LPS) plus phorbolmyristate acetate (PMA).

Cryptotanshinone reduced prostaglandin E2 synthesis and reactive oxygen species generation catalyzed by COX-2, without influencing COX-1 activity in cloned sf-9 cells. In PMA plus LPS-stimulated U937 cells, cryptotanshinone had negligible effects on the expression of COX-1 and COX-2, at either a mRNA or protein level. These results demonstrate that the anti-inflammatory effect of cryptotanshinone is directed against enzymatic activity of COX-2, not against the transcription or translation of the enzyme.

Prostate Cancer

Cryptotanshinone was identified as a potent STAT3 inhibitor. Cryptotanshinone rapidly inhibited STAT3 Tyr705 phosphorylation in DU145 prostate cancer cells and the growth of the cells through 96 hours of the treatment. Inhibition of STAT3 Tyr705 phosphorylation in DU145 cells decreased the expression of STAT3 downstream target proteins such as cyclin D1, survivin, and Bcl-xL.

Cryptotanshinone can suppress Bcl-2 expression and augment Fas sensitivity in DU145 prostate cancer cells. Park et al. (2010) show that JNK and p38 MAPK act upstream of Bcl-2 expression in Fas-treated DU145 cells, and that cryptotanshinone significantly blocked activation of these kinases. Moreover, cryptotanshinone sensitized several tumor cells to a broad range of anti-cancer agents. Collectively, the data suggest that cryptotanshinone has therapeutic potential in the treatment of human prostate cancer (Park et al., 2010).

Cryptotanshinone was colocalized with STAT3 molecules in the cytoplasm and inhibited the formation of STAT3 dimers. Computational modeling showed that cryptotanshinone could bind to the SH2 domain of STAT3. These results suggest that cryptotanshinone is a potent anti-cancer agent targeting the activation STAT3 protein. It is the first report that cryptotanshinone has anti-tumor activity through the inhibition of STAT3 (Shin et al., 2009).

Prostate Cancer; Androgen Receptor Positive

Anti-androgens to reduce or prevent androgens binding to androgen receptor (AR) are widely used to suppress AR-mediated PCa growth; however, the androgen depletion therapy is only effective for a short period of time. Xu et al., (2012) found that cryptotanshinone (CTS), with a structure similar to dihydrotestosterone (DHT), can effectively inhibit the DHT-induced AR transactivation and prostate cancer cell growth. Their results indicated that 0.5 µM CTS effectively suppresses the growth of AR-positive PCa cells, but has little effect on AR negative PC-3 cells and non-malignant prostate epithelial cells.

Furthermore, data indicated that CTS could modulate AR transactivation and suppress the DHT-mediated AR target genes expression in both androgen responsive PCa LNCaP cells and castration resistant CWR22rv1 cells. The mechanistic studies indicate that CTS functions as an AR inhibitor to suppress androgen/AR-mediated cell growth and PSA expression by blocking AR dimerization and the AR-coregulator complex formation.

Furthermore, they showed that CTS effectively inhibits CWR22Rv1 cell growth and expressions of AR target genes in the xenograft animal model. The previously un-described mechanisms of CTS may explain how CTS inhibits the growth of PCa cells and help us to establish new therapeutic concepts for the treatment of PCa.

Breast Cancer, Cervical Cancer, Leukemia, Hepatocellular Carcinoma

The three tanshinone derivatives, tanshinone I, tanshinone IIA, and cryptotanshinone, exhibited significant in vitro cytotoxicity against several human carcinoma cell lines (Wang et al., 2007).

Tanshinone I was found to inhibit the growth and invasion of breast cancer cells both in vitro and in vivo through regulation of adhesion molecules including ICAM-1 and VCAM-1 (Nizamutdinova et al., 2008), and induce apoptosis of leukemia cells by interfering with the mitochondrial transmembrane potential (ΔΨm), increasing the expression of Bax, as well as activating caspase-3 (Liu et al., 2010). Tanshinone IIA has been reported to inhibit the growth of cervical cancer cells through disrupting the assembly of microtubules, and induces G2/M phase arrest and apoptosis (Pan et al., 2010).

This compound can also inhibit invasion and metastasis of hepatocellular carcinoma (HCC) cells both in vitro and in vivo, by suppressing the expression of the metalloproteinases, MMP2 and MMP9 and interfering with the NFκB signaling pathway (Xu et al., 2009).

Breast Cancer

Cryptotanshione was reported to induce cell-cycle arrest at the G1-G0 phase, which was accompanied by the inhibition of cyclin D1 expression, retinoblastoma (Rb) protein phosphorylation, and of the rapamycin (mTOR) signaling pathway (Chen et al., 2010).

Hepato-protective Effect

Cryptotanshinone (20 or 40mg/kg) was orally administered 12 and 1h prior to GalN (700mg/kg)/LPS (10µg/kg) injection. The increased mortality and TNF- α levels by GalN/LPS were declined by cryptotanshinone pre-treatment. In addition, cryptotanshinone attenuated GalN/LPS-induced apoptosis, characterized by the blockade of caspase-3, -8, and -9 activation, as well as the release of cytochrome c from the mitochondria. Furthermore, cryptotanshinone significantly inhibited the activation of NF-κB and suppressed the production of pro-inflammatory cytokines.

These findings suggest that the hepato-protective effect of cryptotanshinone is likely to be associated with its anti-apoptotic activity and the down-regulation of MAPKs and NF-κB associated at least in part with suppressing TAK1 phosphorylation (Jin et al., 2013).

References

Chen W, Luo Y, Liu L, Zhou H, Xu B, Han X, Shen T, Liu Z, Lu Y, Huang S. (2010). Cryptotanshinone Inhibits Cancer Cell Proliferation by Suppressing Mammalian Target of Rapamycin–Mediated Cyclin D1 Expression and Rb Phosphorylation. Cancer Prev Res (Phila), 3(8):1015-25. doi: 10.1158/1940-6207.CAPR-10-0020. Epub 2010 Jul 13.

Jin DZ, Yina LL, Jia XQ, Zhu XZ. (2006). Cryptotanshinone inhibits cyclooxygenase-2 enzyme activity but not its expression. European Journal of Pharmacology, 549(1-3):166-72. doi:10.1016/j.ejphar.2006.07.055

Jin VQ, Jiang S, Wu YL, et al. (2013). Hepato-protective effect of cryptotanshinone from Salvia miltiorrhiza in d-galactosamine/lipopolysaccharide-induced fulminant hepatic failure. Phytomedicine. doi:10.1016/j.phymed.2013.07.016

Liu JJ, Liu WD, Yang HZ, et al. (2010). Inactivation of PI3k/Akt signaling pathway and activation of caspase-3 are involved in tanshinone I-induced apoptosis in myeloid leukemia cells in vitro. Ann Hematol, 89:1089–1097. doi: 10.1007/s00277-010-0996-z.

Nizamutdinova IT, Lee GW, Lee JS, et al. (2008). Tanshinone I suppresses growth and invasion of human breast cancer cells, MDA-MB-231, through regulation of adhesion molecules. Carcinogenesis, 29(10):1885-1892. doi:10.1093/carcin/bgn151

Pan TL, Hung YC, Wang PW, et al. (2010). Functional proteomic and structural insights into molecular targets related to the growth-inhibitory effect of tanshinone IIA on HeLa cells. Proteomics,10:914–929.

Park IJ, Kim MJ, Park OJ, et al. (2010). Cryptotanshinone sensitizes DU145 prostate cancer cells to Fas(APO1/CD95)-mediated apoptosis through Bcl-2 and MAPK regulation. Cancer Lett, 298:88–98. doi: 10.1016/j.canlet.2010.06.006.

Qiu F, Jiang J, Ma Ym, et al. (2013). Opposite Effects of Single-Dose and Multidose Administration of the Ethanol Extract of Danshen on CYP3A in Healthy Volunteers. Evidence-Based Complementary and Alternative Medicine, 2013(2013) http://dx.doi.org/10.1155/2013/730734

Shin DS, Kim HN, Shin KD, et al. (2009). Cryptotanshinone Inhibits Constitutive Signal Transducer and Activator of Transcription 3 Function through Blocking the Dimerization in DU145 Prostate Cancer Cells. Cancer Research, 69:193. doi: 10.1158/0008-5472.CAN-08-2575

Wang X, Morris-Natschke SL, Lee KH. (2007). New developments in the chemistry and biology of the bioactive constituents of Tanshen. Med Res Rev, 27:133–148. doi: 10.1002/med.20077.

Xu D, Lin TH, Li S, Da J, et al. (2012). Cryptotanshinone suppresses androgen receptor-mediated growth in androgen dependent and castration resistant prostate cancer cells. Cancer Lett, 316(1):11-22. doi: 10.1016/j.canlet.2011.10.006.

Xu YX, Feng T, Li R, Liu ZC. (2009). Tanshinone II-A inhibits invasion and metastasis of human hepatocellular carcinoma cells in vitro and in vivo. Tumori, 95:789–795.

Blueberin

Cancer: Colon, prostate, cervical., breast

Action: Anti-inflammatory, blood sugar regulation

Blueberin is isolated from Vaccinium arctostaphylos (L.).

Colon Cancer

Research has shown that diets rich in phenolic compounds such as those associated with blueberries such as blueberin may be associated with lower risks of several chronic diseases including cancer.

To probe this effect, the bioactivities of various components of blueberries were investigated and their potential anti-proliferation and apoptosis induction effects were investigated using two colon cancer cell lines, HT-29 and Caco-2. Polyphenols in three blueberry cultivars, Briteblue, Tifblue, and Powderblue, were extracted and freeze-dried. The extracts were further separated into phenolic acids, tannins, flavonols, and anthocyanins using an HLB cartridge and LH20 column. The phenolic acid fraction showed relatively lower bioactivities with 50% inhibition at 1000 µg/mL. The greatest anti-proliferation effect among all four fractions was from the anthocyanin fractions. Both HT-29 and Caco-2 cell growth was significantly inhibited by >50% by the anthocyanin fractions at concentrations of 15−50 µg/mL. Anthocyanin fractions also resulted in 2−7 times increase in DNA fragmentation, indicating the induction of apoptosis. The effective dosage levels are close to the reported range of anthocyanin concentrations in rat plasma. These findings suggest that blueberry intake may reduce colon cancer risk (Yi, 2005).

Prostate Cancer; AR+, AR-

The role of polyphenol fractions from both wild and cultivated blueberry fruit was probed in the inhibitory effects on the proliferation of LNCaP, an androgen-sensitive prostate cancer cell line, and DU145, a more aggressive androgen insensitive prostate cancer cell line. When 20µg/ml of a wild blueberry polyphenol fraction was added to LNCaP media, growth was inhibited to 11% of control with an IC50 of 13.3µg/ml. Two similar polyphenol-rich fractions from cultivated blueberries at the same concentration inhibited LNCaP growth to 57% and 26% of control with an IC50 of 22.7 and 5.8µg/ml, respectively. Differences in cell growth inhibition of LNCaP and DU145 cell lines by blueberry fractions rich in polyphenols indicate that blueberry proanthocyanidins have an effect primarily on androgen-dependent growth of prostate cancer cells. Possible molecular mechanisms for growth inhibition are reviewed (Schmidt, 2006).

Prostate Cancer

The mechanism(s) by which three flavonoid-enriched fractions from lowbush blueberry (Vaccinium angustifolium) down-regulate matrix metalloproteinase (MMP) activity in DU145 human prostate cancer cells were investigated. Regulation of MMPs is crucial to regulate extracellular matrix (ECM) proteolysis which is important in metastasis. Findings indicate that blueberry flavonoids may use multiple mechanisms in down-regulating MMP activity in these cells (Matchett, 2005).

Cervical Cancer, Breast Cancer

Blueberin, extracted with hexane, 50% hexane/ethyl acetate, ethyl acetate, ethanol, and 70% acetone/water at ambient temperature was tested for in vitro anti-cancer activity on cervical and breast cancer cell lines. Ethanol extracts strongly inhibited CaSki and SiHa cervical cancer cell lines and MCF-7 and T47-D breast cancer cell lines. An unfractionated aqueous extract of raspberry and the ethanol extract of blueberry significantly inhibited mutagenesis by both direct-acting and metabolically activated carcinogens (Wedge et al., 2001).

Anti-inflammatory

The reduction of fasting glucose was correlated with the reduction of serum CRP in the Blueberin group whereas in the Placebo group CRP levels were not significantly reduced. Furthermore, the Blueberin also significantly reduced the levels of plasma enzymes ALT, AST and GGT, indicating that, in addition to anti-diabetes effects, the Blueberin also possess pharmacologically relevant anti-inflammatory properties (Abidov et al., 2006).

References

Abidov M, Ramazanov A, Jimenez Del Rio M, Chkhikvishvili I. (2006). Effect of Blueberin on fasting glucose, C-reactive protein and plasma aminotransferases, in female volunteers with diabetes type 2: double-blind, placebo controlled clinical study. Georgian Med News, (141):66-72.

Matchett MD, MacKinnon, L, Sweeney MI, Gottschall-Pass KT, Hurta, RAR. (2006). Inhibition of matrix metalloproteinase activity in DU145 human prostate cancer cells by flavonoids from lowbush blueberry (Vaccinium angustifolium): possible roles for protein kinase C and mitogen-activated protein-kinase-mediated events. The Journal of Nutritional Biochemistry. doi: 10.1016/j.jnutbio.2005.05.014.

Schmidt BM, Erdman Jr JW, Lila MA. (2006). Differential effects of blueberry proanthocyanidins on androgen sensitive and insensitive human prostate cancer cell lines. Cancer Letters, 231(2):240-246. doi: 10.1021/jf049238n.

Wedge DE, Meepagala KM, Magee JB, et al. (2001). Anti-carcinogenic Activity of Strawberry, Blueberry, and Raspberry Extracts to Breast and Cervical Cancer Cells. Journal of Medicinal Food, 4(1):49-51. doi: 10.1089/10966200152053703.

Yi W, Fischer J, Krewer G, Akoh C. (2005). Phenolic Compounds from Blueberries Can Inhibit Colon Cancer Cell Proliferation and Induce Apoptosis. J. Agric. Food Chem, 53(18):7320–7329. doi: 10.1021/jf051333o.

Apigenin

Cancer:
Breast, gastrointestinal., prostate, ovarian, pancreatic

Action: Anti-proliferative effect, induces apoptosis, chemo-sensitizer

Apigenin (4′,5,7-trihydroxyflavone, 5,7-dihydroxy-2-(4-hydroxyphenyl)-4H-1-benzopyran-4-one) is a flavonoid found in many fruits, vegetables, and herbs, the most abundant sources being the leafy herb parsley and dried flowers of chamomile. Present in dietary sources as a glycoside, it is cleaved in the gastrointestinal lumen to be absorbed and distributed as apigenin itself. For this reason, the epithelium of the gastrointestinal tract is exposed to higher concentrations of apigenin than tissues at other locations. This would also be true for epithelial cancers of the gastrointestinal tract. There is evidence that the actions of apigenin might hinder the ability of gastrointestinal cancers to progress and spread.

Induces Apoptosis, Anti-metastatic

Apigenin has been shown to inhibit cell growth, sensitize cancer cells to elimination by apoptosis, and hinder the development of blood vessels to serve the growing tumor. It also has actions that alter the relationship of the cancer cells with their microenvironment. Apigenin is able to reduce cancer cell glucose uptake, inhibit remodeling of the extracellular matrix, inhibit cell adhesion molecules that participate in cancer progression, and oppose chemokine signaling pathways that direct the course of metastasis into other locations. As such, apigenin may provide some additional benefit beyond existing drugs in slowing the emergence of metastatic disease (Lefort, 2013).

Chemo-sensitizer, Induces Apoptosis

Choi & Kim (2009) investigated the effects of combined treatment with 5-fluorouracil and apigenin on proliferation and apoptosis, as well as the underlying mechanism, in human breast cancer MDA-MB-453 cells. The MDA-MB-453 cells, which have been shown to overexpress ErbB2, were resistant to 5-fluorouracil; 5-fluorouracil exhibited a small dose-dependent anti-proliferative effect, with an IC50 of 90 microM. Interestingly, combined treatment with apigenin significantly decreased the resistance. Cellular proliferation was significantly inhibited in cells exposed to 5-fluorouracil at its IC50 and apigenin (5, 10, 50 and 100 microM), compared with proliferation in cells exposed to 5-fluorouracil alone.

This inhibition in turn led to apoptosis, as evidenced by an increased number of apoptotic cells and the activation of caspase-3. Moreover, compared with 5-fluorouracil alone, 5-fluorouracil in combination with apigenin at concentrations >10 microM exerted a pro-apoptotic effect via the inhibition of Akt expression.

Taken together, results suggest that 5-fluorouracil acts synergistically with apigenin inhibiting cell growth and inducing apoptosis via the down-regulation of ErbB2 expression and Akt signaling (Choi, 2009).

Breast Cancer, Prostate Cancer

Two flavonoids, genistein and apigenin, have been implicated as chemo-preventive agents against prostate and breast cancers; however, the mechanisms behind their respective cancer-protective effects may vary significantly. It was thought that the anti-proliferative action of these flavonoids on prostate (DU-145) and breast (MDA-MB-231) cancer cells expressing only estrogen receptor (ER) β is mediated by this ER subtype. It was found that both genistein and apigenin, although not 17β-estradiol, exhibited anti-proliferative effects and pro-apoptotic activities through caspase-3 activation in these two cell lines. In yeast transcription assays, both flavonoids displayed high specificity toward ERβ transactivation, particularly at lower concentrations.

However, in mammalian assay, apigenin was found to be more ERβ-selective than genistein, which has equal potency in inducing transactivation through ERα and ERβ. Small interfering RNA-mediated down-regulation of ERβ abrogated the anti-proliferative effect of apigenin in both cancer cells but did not reverse that of genistein. These results unveil that the anti-cancer action of apigenin is mediated, in part, by ERβ. The differential use of ERα and ERβ signaling for transaction between genistein and apigenin demonstrates the complexity of phytoestrogen action in the context of their anti-cancer properties (Mak, 2006).

Ovarian Cancer

Id1 (inhibitor of differentiation or DNA binding protein 1) contributes to tumorigenesis by stimulating cell proliferation, inhibiting cell differentiation and facilitating tumor neoangiogenesis. Elevated Id1 is found in ovarian cancers and its level correlates with the malignant potential of ovarian tumors. Therefore, Id1 is a potential target for ovarian cancer treatment. It has been demonstrated that apigenin inhibits proliferation and tumorigenesis of human ovarian cancer A2780 cells through Id1. Apigenin has been found to suppress the expression of Id1 through activating transcription factor 3 (ATF3). These results may elucidate a new mechanism underlying the inhibitory effects of apigenin on cancer cells (Li, 2009).

Pancreatic Cancer

Simultaneous treatment or pre-treatment (0, 6, 24 and 42 hours) of apigenin and chemotherapeutic drugs and various concentrations (0-50µM) were assessed using the MTS cell proliferation assay. Simultaneous treatment with apigenin (0,13, 25 or 50µM) and chemotherapeutic drugs 5-fluorouracil (5-FU, 50µM) or gemcitabine (Gem, 10µM) for 60 hours resulted in less-than-additive effect (p<0.05). Pre-treatment for 24 hours with 13µM of apigenin, followed by Gem for 36 hours was optimal to inhibit cell proliferation.

Pre-treatment of cells with 11-19µM of apigenin for 24 hours resulted in 59-73% growth inhibition when followed by Gem (10µM, 36h). Pre-treatment of human pancreatic cancer cells BxPC-3 with low concentrations of apigenin hence effectively aids in the anti-proliferative activity of chemotherapeutic drugs (Johnson, 2013).

Induces Apoptosis, Inhibits Angiogenesis and Metastasis.

Preclinical studies have also shown that Ocimum sanctum L. and some of the phytochemicals it contains (including apigenin) prevents chemical-induced skin, liver, oral., and lung cancers. These effects are thought to be mediated by increasing the anti-oxidant activity, altering gene expression, inducing apoptosis, and inhibiting angiogenesis and metastasis. The aqueous extract of Ocimum sanctum L. has been shown to protect mice against γ-radiation-induced sickness and mortality and to selectively protect the normal tissues against the tumoricidal effects of radiation. In particular, important phytochemicals like apigenin have also been shown to prevent radiation-induced DNA damage. This warrants its future research to establish its activity and utility in cancer prevention and treatment (Baliga, 2013).

Lung Cancer

Apigenin has been found to induce apoptosis and cell death in lung epithelium cancer (A549) cells with an IC50 value of 93.7 ± 3.7 µM for 48 hours treatment. Target identification investigations using A549 cells and in cell-free systems demonstrate that apigenin depolymerized microtubules and inhibited reassembly of cold depolymerized microtubules of A549 cells. Again apigenin inhibited polymerization of purified tubulin with an IC50 value of 79.8 ± 2.4 µM. Interestingly, apigenin also showed synergistic anti-cancer effects with another natural anti-tubulin agent, curcumin. Apigenin and curcumin synergistically induce cell death and apoptosis and also block cell-cycle progression at G2/M phase of A549 cells.

Understanding the mechanism of the synergistic effect of apigenin and curcumin could help to develop anti-cancer combination drugs from cheap and readily available nutraceuticals (Choudhury, 2013).

Induces Apoptosis

It has been shown that the dietary flavonoid apigenin binds and inhibits adenine nucleotide translocase-2 (ANT2), resulting in enhancement of Apo2L/TRAIL-induced apoptosis by up-regulation of DR5, making it a potential cancer therapeutic agent. Apigenin has been found to enhance Apo2L/TRAIL-induced apoptosis in cancer cells by inducing DR5 expression through binding ANT2. Similarly to apigenin, knockdown of ANT2 enhanced Apo2L/TRAIL-induced apoptosis by up-regulating DR5 expression at the post-transcriptional level.

Moreover, silencing of ANT2 attenuated the enhancement of Apo2L/TRAIL-induced apoptosis by apigenin. These results suggest that apigenin Up-regulates DR5 and enhances Apo2L/TRAIL-induced apoptosis by binding and inhibiting ANT2. ANT2 inhibitors like apigenin may hence contribute to Apo2L/TRAIL therapy (Oishi, 2013).

Colorectal Cancer

Apigenin has anti-proliferation, anti-invasion and anti-migration effects in three kinds of colorectal adenocarcinoma cell lines, namely SW480, DLD-1 and LS174T. Proteomic analysis with SW480 indicated that apigenin up-regulated the expression of transgelin (TAGLN) in mitochondria to exert its anti-tumor growth and anti-metastasis effects. Apigenin decreased the expression of MMP-9 in a dose-dependent manner. Transfection of three truncated forms of TAGLN and wild type has identified TAGLN as a repressor of MMP-9 expression.

This research provides direct evidence that apigenin inhibits tumor growth and metastasis both in vitro and in vivo. Apigenin up-regulates TAGLN and down-regulates MMP-9 expression through decreasing phosphorylation of Akt at Ser473 and in particular Thr308 to prevent cancer cell proliferation and migration (Chunhua, 2013).

References

Baliga MS, Jimmy R, Thilakchand KR, et al. (2013). Ocimum Sanctum L (Holy Basil or Tulsi) and Its Phytochemicals in the Prevention and Treatment of Cancer. Nutr Cancer, 65(1):26-35. doi: 10.1080/01635581.2013.785010.

 

 

Choi EJ, Kim GH. (2009). 5-Fluorouracil combined with apigenin enhances anti-cancer activity through induction of apoptosis in human breast cancer MDA-MB-453 cells. Oncol Rep, 22(6):1533-7.

 

Choudhury D, Ganguli A, Dastidar DG, et al. (2013). Apigenin shows synergistic anti-cancer activity with curcumin by binding at different sites of tubulin. Biochimie, 95(6):1297-309. doi: 10.1016/j.biochi.2013.02.010.

 

Chunhua L, Donglan L, Xiuqiong F, et al. (2013). Apigenin up-regulates transgelin and inhibits invasion and migration of colorectal cancer through decreased phosphorylation of AKT. J Nutr Biochem. doi: 10.1016/j.jnutbio.2013.03.006.

 

Johnson JL, Gonzalez de Mejia E. (2013). Interactions between dietary flavonoids apigenin or luteolin and chemotherapeutic drugs to potentiate anti-proliferative effect on human pancreatic cancer cells, in vitro. Food Chem Toxicol, 20:83-91. doi: 10.1016/j.fct.2013.07.036.

 


Lefort ƒC, Blay J. (2013). Apigenin and its impact on gastrointestinal cancers. Mol Nutr Food Res, 57(1):126-44. doi: 10.1002/mnfr.201200424.

 

Li ZD, Hu XW, Wang YT & Fang J. (2009). Apigenin inhibits proliferation of ovarian cancer A2780 cells through Id1. FEBS Letters, 583(12):1999-2003 doi:10.1016/j.febslet.2009.05.013.

 

Mak P, Leung YK, Tang WY, Harwood C & Ho SM. (2006). Apigenin suppresses cancer cell growth through ERβ. Neoplasia, 8(11):896–904.

 

Oishi M, Iizumi Y, Taniguchi T, et al. (2013). Apigenin Sensitizes Prostate Cancer Cells to Apo2L/TRAIL by Targeting Adenine Nucleotide Translocase-2. PLoS One, 8(2):e55922. doi: 10.1371/journal.pone.0055922.

Genistein (See also Daidzien)

Cancer:
Breast, kidney, prostate, renal., liver, endometrial., ovarian

Action: Anti-angiogenesis, cell-cycle arrest, cancer stem cells, VEGF, radiotherapy, sex hormone-binding globulin (SHBG), insulin-like growth factor-1 (IGF-1)

Genistein is a natural isoflavone phytoestrogen present in a number of plants, including soy, fava, and kudzu (Glycine max [(L.) Merr.], Vicia faba (L.), Pueraria lobata [(Willd.) Ohwi]).

Phytoestrogens

Phytoestrogens have been investigated at the epidemiological., clinical and molecular levels to determine their potential health benefits. The two major groups of phytoestrogens, isoflavones and lignans, are abundant in soy products and flax respectively, but are also present in a variety of other foods. It is thought that these estrogen-like compounds may protect against chronic diseases, such as hormone-dependent cancers, cardiovascular disease and osteoporosis (Stark & Madar, 2002).

S-Equol Production and Isoflavone Metabolism

S-Equol and Breast Cancer

Differences in ability to metabolize daidzein to equol might help explain inconsistent findings about isoflavones and breast cancer. Tseng et al. (2013) examined equol-producing status in relation to breast density, a marker of breast cancer risk, and evaluated whether an association of isoflavone intake with breast density differs by equol-producing status in a sample of Chinese immigrant women. In their sample, 30% were classified as equol producers. In adjusted linear regression models, equol producers had significantly lower mean dense tissue area (32.8 vs. 37.7 cm(2), P = 0.03) and lower mean percent breast density (32% vs. 35%, P = 0.03) than nonproducers. Significant inverse associations of isoflavone intake with dense area and percent density were apparent, but only in equol producers (interaction P = 0.05 for both).

Although these findings warrant confirmation in a larger sample, they offer a possible explanation for the inconsistent findings about soy intake and breast density and possibly breast cancer risk as well. The findings further suggest the importance of identifying factors that influence equol-producing status and exploring appropriate targeting of interventions.

S-Equol and Dietary Factors

S-(-)equol, an intestinally derived metabolite of the soy isoflavone daidzein, is proposed to enhance the efficacy of soy diets. Setchell et al. (2013) performed a comprehensive dietary analysis of 143 macro- and micro-nutrients in 159 healthy adults to determine whether the intake of specific nutrients favors equol production. Three-day diet records were collected and analyzed using Nutrition Data System for Research software and S-(-)equol was measured in urine by mass spectrometry.

Equol producers accounted for 29.6% of participants. No significant differences were observed for total protein, carbohydrate, fat, saturated fat, or fiber intakes between equol producers and nonproducers. However, principal component analysis revealed differences in several nutrients, including higher intakes of polyunsaturated fatty acids (P = 0.039), maltose (P = 0.02), and vitamins A (P = 0.01) and E (P = 0.035) and a lower intake of total cholesterol (P = 0.010) in equol producers.

Subtle differences in some nutrients may influence the ability to produce equol.

S-Equol and Dietary Factors; Fats

The soy isoflavones, daidzein and genistein, and the lignans, matairesinol and secoisolariciresinol, are phytoestrogens metabolized extensively by the intestinal microflora. Considerable important evidence is already available that shows extensive interindividual variation in isoflavone metabolism. There was a 16-fold variation in total isoflavonoid excretion in urine after the high-isoflavone treatment period. The variation in urinary equol excretion was greatest (664-fold), and subjects fell into two groups: poor equol excretors and good equol excretors (36%). A significant negative correlation was found between the proportion of energy from fat in the habitual diet and urinary equol excretion (r = -0.55; p = 0.012). Good equol excretors consumed less fat as percentage of energy than poor excretors (26 +/- 2.3% compared with 35 +/- 1.6%, p < 0.01) and more carbohydrate as percentage of energy than poor excretors (55 +/- 2.9% compared with 47 +/- 1.7%, p < 0.05).

It is suggested that the dietary fat intake decreases the capacity of gut microbial flora to synthesize equol (Rowland et al., 2000).

Isoflavones and Fermented Soy Foods

Serum concentrations of total isoflavones after 1–4 hours were significantly higher in the aglycone-rich fermented soybeans (Fsoy) group than in the glucoside-rich non-fermented soybeans (Soy) group. The Fsoy group showed significantly higher maximum concentration (Cmax: 2.79 ± 0.13 vs 1.74 ± 0.13 µmol L(-1) ) and area under the curve (AUC(0-24 h) : 23.78 ± 2.41 vs 19.95 ± 2.03 µmol day L(-1) ) and lower maximum concentration time (Tmax: 1.00 ± 0.00 vs 5.00 ± 0.67 h) compared with the Soy group. The cumulative urinary excretion of total isoflavones after 2 hours was significantly higher in the Fsoy group than in the Soy group. Individual isoflavones (daidzein, genistein and glycitein) showed similar trends to total isoflavones. Equol (a metabolite from daidzein) did not differ between the two groups.

The results of this study demonstrated that the isoflavones of aglycone-rich Fsoy were absorbed faster and in greater amounts than those of glucoside-rich Soy in postmenopausal Japanese women (Okabe et al., 2011).

Phytoestrogens and Breast Cancer; ER+/ER-, ER α /ER β

Dietary-derived Anti-angiogenic Compounds

Consumption of a plant-based diet can prevent the development and progression of chronic diseases that are associated with extensive neovascularization; however, little is known about the mechanisms. To determine whether prevention might be associated with dietary-derived angiogenesis inhibitors, the urine of healthy human subjects consuming a plant-based diet was fractionated and the fractions examined for their ability to inhibit the proliferation of vascular endothelial cells.

The isoflavonoid genistein was the most potent, and inhibited endothelial cell proliferation and in vitro angiogenesis at concentrations giving half-maximal inhibition of 5 and 150 microM, respectively. Genistein concentrations in urine of subjects consuming a plant-based diet are in the micromolar range, while those of subjects consuming a traditional Western diet are lower by a factor of > 30. The high excretion of genistein in urine of vegetarians and in addition to these results suggest that genistein may contribute to the preventive effect of a plant-based diet on chronic diseases, including solid tumors, by inhibiting neovascularization.

Thus, genistein may represent a member of a new class of dietary-derived anti-angiogenic compounds (Fotsis et al., 1993).

ERβ as a Down-regulator of ER+ Breast Cancer

The estrogen receptor (ER) isoform known as ERβ has become the focus of intense investigation as a potential drug target. The existence of clear-cut differences in ERβ and ERα expression suggests that tissues could be differentially targeted with ligands selective for either isoform (Couse et al., 1997; Enmark et al., 1997). In particular, the fact that ER β is widely expressed but not the primary estrogen receptor in, for example, the uterus (where estrogenic effects are mediated via ERα) (Harris, Katzenellenbogen, & Katzenellenbogen, 2002) opens up the possibility of targeting other tissues while avoiding certain classical estrogenic effects.

A major advance toward understanding how some phytoestrogens achieve modest ERβ selectivity was the X-ray structure determination of the ERβ ligand binding domain (LBD) complexed with genistein (GEN) (Pike et al., 1999), a 40-fold ERβ-selective ligand (Harris et al., 2002). This study clearly showed that there are only two residue substitutions in close proximity to GEN: ERα Leu384 is replaced by ER β Met336, and ERα Met421 is replaced by ER β Ile373.

ERbeta works as counter partner of ERalpha through inhibition of the transactivating function of ERalpha by heterodimerization, distinct regulation on several specific promoters by ERalpha or ERbeta, and ERbeta-specific regulated genes which are probably related to its anti-proliferative properties. Epidemiological studies of hormone replacement therapy and isoflavone (genistein) consumption indicate the possible contribution of ERbeta-specific signaling in breast cancer prevention. A selective estrogen receptor modulator, which works as an antagonist of ERalpha and an agonist of ERbeta, may be a promising chemo-preventive treatment (Saji, Hirose, & Toi, 2005).

Genistein and Apoptosis

The association between consumption of genistein containing soybean products and lower risk of breast cancer suggests a cancer chemo-preventive role for genistein. Consistent with this suggestion, exposing cultured human breast cancer cells to genistein inhibits cell proliferation, although this is not completely understood. To better understand how genistein works, the ability of genistein to induce apoptosis was compared in phenotypically dissimilar MCF-7 and MDA-MB-231 human breast cancer cells that express the wild-type and mutant p53 gene, respectively.

After 6 days of incubation with 50 microM genistein, MCF-7, but not MDA-MB-231 cells, showed morphological signs of apoptosis. Marginal proteolytic cleavage of poly-(ADP-ribose)-polymerase and significant DNA fragmentation were also detected in MCF-7 cells.

In elucidating these findings, it was determined that after 2 days of incubation with genistein, MCF-7, but not MDA-MB-231 cells, had significantly higher levels of p53. Accordingly, the expression of certain proteins modulated by p53 was also studied. Levels of p21 increased in both of the genistein-treated cell lines, suggesting that p21 gene expression was activated but in a p53-independent manner; whereas no significant changes in levels of the pro-apoptotic protein, Bax, were found. In MCF-7 cells, levels of the anti-apoptotic protein, Bcl-2, decreased slightly at 18–24 hours but then increased considerably after 48 hours. Hence, the Bax:Bcl-2 ratio initially increased but later decreased.

Data suggests that at the concentration tested, MCF-7 cells, in contrast to MDA-MB-231 cells, were sensitive to the induction of apoptosis by genistein. However, the roles of Bax and Bcl-2 are unclear (Xu & Loo, 2001).

Genistein Derivatives and Breast Cancer Inhibition

Genistein binds to estrogen receptors and stimulates growth at concentrations that would be achieved by a high soy diet, but inhibits growth at high experimental concentrations.

The estrogen receptor (ER) is a major target for the treatment of breast cancer cells. Genistein, a soy isoflavone, possesses a structure similar to estrogen and can both mimic and antagonize estrogen effects although at high concentrations it inhibits breast cancer cell proliferation. Hence, to enhance the anti-cancer activity of Genistein at lower concentrations, seven structurally modified derivatives of Genistein based on the structural requirements for an optimal anti-cancer effect were synthesised. Among those seven, three derivatives showed high anti-proliferative activity with IC(50) levels in the range of 1-2.5 µM, i.e., at much lower concentrations range than Genistein itself, in three ER-positive breast cancer cell lines (MCF-7, 21PT and T47D) studied. In our analysis, we noticed that at IC(50) concentrations, the MA-6, MA-8 and MA-19 Genistein derivatives induced apoptosis, inhibited ER-α messenger RNA expression and increased the ratio of ER-β to ER-α levels in a manner comparable to that of the parent compound Genistein.

Of note, these three modified Genistein derivatives exerted their effects at concentrations 10–15 times lower than the parent compound, decreasing the likelihood of significant ER- α pathway activation, which has been a concern for Genistein. Hence these compounds might play a useful role in breast cancer chemoprevention (Marik et al., 2011).

Genistein and ER α

To determine the effects of low-dose, long-term genistein exposure MCF-7 breast cancer cells were cultured in 10nM genistein for 10-12 weeks and investigated whether or not this long-term genistein treatment (LTGT) altered the expression of estrogen receptor alpha (ERalpha) and the activity of the PI3-K/Akt signaling pathway. This is known to be pivotal in the signaling of mitogens such as oestradiol (E(2)), insulin-like growth factor-1 (IGF-1) and epidermal growth factor (EGF). LTGT significantly reduced the growth promoting effects of E(2) and increased the dose-dependent growth-inhibitory effect of the PI3-K inhibitor, LY 294002, compared to untreated control MCF-7 cells.

This was associated with a significant decreased protein expression of total Akt and phosphorylated Akt but not ERalpha. Rapamycin, an inhibitor of one of the downstream targets of Akt, mammalian target of rapamycin (mTOR), also dose-dependently inhibited growth but the response to this drug was similar in LTGT and control MCF-7 cells. The protein expression of liver receptor homologue-1 (LRH1), an orphan nuclear receptor implicated in tumorigenesis was not affected by LTGT.

These results show that LTGT results in a down-regulation of the PI3-K/Akt signaling pathway and may be a mechanism through which genistein could offer protection against breast cancer (Anastasius et al., 2009).

Genistein and ER+/ER-

Genistein was found to cause a dose-dependent growth inhibition of the two hormone-sensitive cell lines T47D and ZR75.1 and of the two hormone-independent cell lines MDAMB-231 and BT20. Flow cytometric analysis of cells treated for 4 days with 15 and 30 M genistein showed a dose-dependent accumulation in the G2M phase of the cell-cycle. At the highest tested concentration, there was a 7-fold increase in the percentage of cells in G2M (63%) with respect to the control (9%) in the case of T47D cells and a 2.4-fold increase in the case of BT20. An intermediate 4-fold accumulation was observed in the case of MDAMB-231 and ZR75.1. The G2M arrest was coupled with a parallel depletion of the G0/G1 phase.

To understand the mechanism of action underlying the block in G2M induced by genistein, Cappelletti et al. (2000) investigated the expression and the activity of cyclins and of cyclin-dependent kinases specifically involved in the G2M transition. As expected, p34cdc-2 expression, monitored by Western blotting, was unaffected by genistein treatment in all cell lines. With the exception of the T47D cell line, we revealed an increase in the tyrosine phosphorylated form of p34, suggesting an inactivation of the p34cdc-2 catalytic activity consequent to treatment of cells with genistein. In fact, immunoprecipitates from genistein-treated MDAMB-231 and BT20 cells displayed a 4-fold decrease in kinase activity evaluated using the histone H1 as substrate.

Conversely, no variation in kinase activity was observed between treated and untreated ZR75.1 cells despite the increase in p34 phosphorylation. In cells treated with 30 M genistein, cyclin B1 (p62) increased 2.8-,8-and 103-fold, respectively, in BT20, MDAMB-231, and ZR75.1 cells, suggesting an accumulation of the p62, which is instead rapidly degraded in cycling cells. No effects were observed on cyclin expression in T47D cells.

We therefore conclude that genistein causes a G2M arrest in breast cancer cell lines, but that such growth arrest is not necessarily coupled with deregulation of the p34cdc-2/cyclin B1 complex only in all of the studied cell lines.

Genistein and ER+/ER-; MDR

Genistein is a potent inhibitor of the growth of the human breast carcinoma cell lines, MDA-468 (estrogen receptor negative), and MCF-7 and MCF-7-D-40 (estrogen receptor positive) (IC50 values from 6.5 to 12.0 µg/ml). The presence of the estrogen receptor is not required for the isoflavones to inhibit tumor cell growth (MDA-468 vs MCF-7 cells). In addition, the effects of genistein and biochanin A are not attenuated by over expression of the multi-drug resistance gene product (MCF-7-D40 vs MCF-7 cells (Peterson et al., 1991).

Studies have shown that genistein exerts multiple suppressive effects on both estrogen receptor positive (ER+) as well as estrogen receptor negative (ER-) human breast carcinoma lines suggesting that the mechanisms of these effects may be independent of ER pathways.

In the present study however Shao et al. (2000) provide evidence that in the ER+ MCF-7, T47D and 549 lines but not in the ER-MDA-MB-231 and MDA-MB-468 lines both presumed 'ER-dependent' and 'ER-independent' actions of genistein are mediated through ER pathways. Genistein's anti-proliferative effects are estrogen dependent in these ER+ lines, being more pronounced in estrogen-containing media and in the presence of exogenous 17-beta estradiol. Genistein also inhibits the expression of ER-downstream genes including pS2 and TGF-beta in these ER+ lines and this inhibition is also dependent on the presence of estrogen. Genistein inhibits estrogen-induced protein tyrosine kinase (PTK) activity. Genistein is only a weak transcriptional activator and actually decreases ERE-CAT levels induced by 17-beta estradiol in the ER+ lines.

Genistein also decreases steady state ER mRNA only in the presence of estrogen in the ER+ lines thereby manifesting another suppression of and through the ER pathway. Their observations resurrect the hypothesis that genistein functions as a 'good estrogen' in ER+ breast carcinomas. Since chemo-preventive effects of genistein would be targeted to normal ER-positive ductal-lobular cells of the breast, this 'good estrogen' action of genistein is most relevant to our understanding of chemoprevention.

Genistein and Concentration

The anti-proliferative activity of the isoflavones daidzein and genistein were investigated in three breast cancer cell lines with different patterns of estrogen receptor (ER) and c erbB 2 protein expression (ERα positive MCF 7 cells, c erbB 2 positive SK BR 3 cells and ERα/c erbB 2 positive ZR 75 1). After treatment at various concentrations (1 200 µM for 72 hours), the effect of daidzein and genistein on the proliferation of different cell types varied; these effects were found to be associated with ERα and c erbB 2 expression. Daidzein and genistein exhibited biphasic effects (stimulatory or inhibitory) on proliferation and ERα expression in MCF 7 cells. Although 1 µM daidzein significantly stimulated cell growth, ERα expression was unaffected. However, genistein showed marked increases in proliferation and ERα expression after exposure to <10 µM genistein.

Notably, the inhibition of cell proliferation by 200 µM genistein was greater compared to that by daidzein at the same concentration. Daidzein and genistein significantly inhibited proliferation of SK BR 3 and ZR 75 1 cells in a dose-dependent manner. In addition, ERα and c erbB 2 expression was reduced by daidzein and genistein in both SK BR 3 and ZR 75 1 cells in a dose-dependent manner. However, the effect of genistein was greater compared to that of daidzein.

In conclusion, the isoflavones daidzein and genistein showed anti breast cancer activity, which was associated with expression of the ERα and c erbB 2 receptors (Choi et al., 2013).

ER- α / ER β Receptors

Isoflavones are phytoestrogens that have been linked to both beneficial as well as adverse effects in relation to cell proliferation and cancer risks. The mechanisms that could be involved in this dualistic mode of action were investigated. One mechanism relates to the different ultimate cellular effects of activation of estrogen receptor (ER) α, promoting cell proliferation, and of ERβ, promoting apoptosis, with the major soy isoflavones genistein and daidzein activating especially ERβ.

A second mode of action includes the role of epigenetics, including effects of isoflavones on DNA methylation, histone modification and miRNA expression patterns. The overview presented reveals that we are only at the start of unraveling the complex underlying mode of action for effects of isoflavones, both beneficial or adverse, on cell proliferation and cancer risks. It is evident that whatever model system will be applied, its relevance to human tissues with respect to ERα and ERβ levels, co-repressor and co-activator characteristics as well as its relevance to human exposure regimens, needs to be considered and defined (Rietjens et al., 2013).

Genistein and ER+/ER-, ER- α / ER β Receptors

A novel mechanism of adipokine, adiponectin (APN) -mediated signaling that influences mammary epithelial cell proliferation, differentiation, and apoptosis to modify breast cancer risk has been identified. It was demonstrated that early dietary exposure to soy protein isolate induced mammary tissue APN production without corresponding effects on systemic APN levels. In estrogen receptor (ER)-negative MCF-10A cells, recombinant APN promoted lobuloalveolar differentiation by inhibiting oncogenic signal transducer and activator of transcription 3 activity.

In ER-positive HC11 cells, recombinant APN increased ERβ expression, inhibited cell proliferation, and induced apoptosis. Using the estrogen-responsive 4X-estrogen response element promoter-reporter construct to assess ER transactivation and small interfering RNA targeting of ERα and ERβ, Rahal et al. (2011) show that APN synergized with the soy phytoestrogen genistein to promote ERβ signaling in the presence of estrogen (17β-estradiol) and ERβ-specific agonist 2,3-bis(4-hydroxyphenyl)-propionitrile and to oppose ERα signaling in the presence of the ERα-specific agonist 4,4',4'-(4-propyl-(1H)-pyrazole-1,3,5-triyl)trisphenol.

The enhancement of ERβ signaling with APN + genistein co-treatments was associated with induction of apoptosis, increased expression of pro-apoptotic/prodifferentiation genes (Bad, p53, and Pten), and decreased anti-apoptotic (Bcl2 and survivin) transcript levels. These results suggest that mammary-derived APN can influence adjacent epithelial function by ER-dependent and ER-independent mechanisms that are consistent with reduction of breast cancer risk and suggest local APN induction by dietary factors as a targeted approach for promotion of breast health.

Genistein and Non-breast Cancer

Genistein Concentrations; Endometrial Cancer

The influence of two phytoestrogens (Genistein and Daidzein) on estrogen-related receptor-α in endometrial cancer cell line Ishikawa was investigated on the proliferation of the cells in this cell line. Ishikawa cells were incubated with different concentrations of Genistein and Daidzein (40, 20, 10, 5 µmol/L) for 24 hours or 48 hours, followed by Real-Time PCR for analyzing the expression of ERR-α mRNA in the cell line. MTT assay was then performed to evaluate the proliferation of Ishikawa cells.

The expression level of ERR-α mRNA in Ishikawa cells was higher than that of the control group after being dealt for 24 hours or 48 hours with Genistein, and the concentration 20 µmol/L was most effective. Nevertheless, this up-regulation was blocked when the cells were treated with 40 µmol/L Genistein. Lower concentration (5, 10 µmol/L) Genistein had depressant effect on proliferation of the cells, while higher concentrations (20, 40 µmol/L) had stimulant effect. After being treated with different concentrations of Daidzein, the expression of ERR- α mRNA in all experimental groups was significantly higher than that in the control group. In the 24 hour group, the concentration 40 µmol/L had most obvious effect; but in the 48 hour group, the concentration 20 µmol/L had most obvious effect, and this up-regulation was blocked when the concentration was elevated to 40 µmol/L.

Noticeably, all concentrations of Daidzein had depressant effect on the proliferation of Ishikawa cells in both 24 hour and 48 hour groups. In the 24 hour group, lower concentrations were more effective, but in the 48 hour group, concentration showed no significant effect. In lower concentrations, both Genistein and Daidzein have up-regulation effect on the expression of ERR-α, and block the proliferation of Ishikawa cells; but in higher concentrations, the up-regulation effect on ERR-α mRNA expression by these two phytoestrogens is not obvious. Genistein stimulates the proliferation of lshikawa cells in higher concentrations, while Daidzein suppresses the proliferation, especially in lower concentrations (Xin et al., 2009).

Genistein and VEGF; Ovarian Cancer

Genistein represses NF-kappaB (NF-κB), a pro-inflammatory transcription factor, and inhibits pro-inflammatory cytokines such as TNF-α and IL-6 in epithelial ovarian cancer. Additionally, it has been shown to stabilize p53 protein, sensitize TRAIL (TNF receptor apoptosis-inducing ligand) induce apoptosis, and prevent or delay chemotherapy-resistance. Recent studies further indicate that genistein potently inhibits VEGF production and suppresses ovarian cancer cell metastasis in vitro.

Based on widely published in vitro and mouse-model data, some anti-inflammatory phytochemicals appear to exhibit activity in modulating the tumor microenvironment. Specifically, apiegenin, baicalein, curcumin, EGCG, genistein, luteolin, oridonin, quercetin, and wogonin repress NF-kappaB (NF-κB, a pro-inflammatory transcription factor) and inhibit pro-inflammatory cytokines such as TNF-α and IL-6. Recent studies further indicate that apigenin, genistein, kaempferol, luteolin, and quercetin potently inhibit VEGF production and suppress ovarian cancer cell metastasis in vitro. Lastly, oridonin and wogonin were suggested to suppress ovarian CSCs as is reflected by down-regulation of the surface marker EpCAM (Chen, Michael, & Butler-Manuel, 2012).

Renal Cell Carcinoma, Prostate Cancer; Radiotherapy

The KCI-18 RCC cell line was generated from a patient with papillary renal cell carcinoma. Tumor cells metastasize from the primary renal tumor to the lungs, liver and mesentery mimicking the progression of RCC in humans. Treatment of established kidney tumors with genistein demonstrated a tendency to stimulate the growth of the primary kidney tumor and increase the incidence of metastasis to the mesentery lining the bowel. In contrast, when given in conjunction with kidney tumor irradiation, genistein significantly inhibited the growth and progression of established kidney tumors. These findings confirm the potentiation of radiotherapy by genistein in the orthotopic RCC model as previously shown in orthotopic models of prostate cancer. These studies in both RCC and prostate tumor models demonstrate that the combination of genistein with primary tumor irradiation is a more effective and safer therapeutic approach as the tumor growth and progression are inhibited both in the primary and metastatic sites (Gilda et al., 2007).

Cell-cycle Arrest

Genistein treatment increased Wee1 levels and decreased phospho-Wee1 (Ser 642). Moreover, genistein substantially decreased the Ser473 and Thr308 phosphorylation of Akt and up-regulated PTEN expression. Down-regulation of PTEN by siRNA in genistein-treated cells increased phospho-Wee1 (Ser642), whereas it decreased phospho-Cdc2 (Tyr15), resulting in decreased G2/M cell-cycle-arrest. Therefore, induction of G2/M cell-cycle arrest by genistein involved up-regulation of PTEN (Liu et al., 2013).

Cancer Stem Cells (CSCs)

Cancer stem cells (CSCs) are cells that exist within a tumor with a capacity for self-renewal and an ability to differentiate, giving rise to heterogeneous populations of cancer cells. These cells are increasingly being implicated in resistance to conventional therapeutics and have also been implicated in tumor recurrence. Several cellular signaling pathways including Notch, Wnt, phosphoinositide-3-kinase-Akt-mammalian target of rapamycin pathways, and known markers such as CD44, CD133, CD166, ALDH, etc. have been associated with CSCs.

Here, we have reviewed our current understanding of self-renewal pathways and factors that help in the survival of CSCs with special emphasis on those that have been documented to be modulated by well characterized natural agents such as curcumin, sulforaphane, resveratrol, genistein, and epigallocatechin gallate (Dandawate et al., 2013).

Genistein and Sex Hormone-binding Globulin (SHBG)

Studies have indicated a correlation between a high level of urinary lignans and isoflavonoid phytoestrogens, particularly genistein, and a low incidence of hormone-dependent cancers, such as breast and prostate cancer. Previously it has been observed that a vegetarian diet is associated with high plasma levels of sex hormone-binding globulin (SHBG), reducing clearance of sex hormones and probably risk of breast and prostate cancer. In the present study we investigated the in vitro effect of genistein on the production of SHBG by human hepatocarcinoma (Hep-G2) cells in culture and its effect on cell proliferation.

It has additionally been found that genistein not only significantly increases the SHBG production by Hep-G2 cells, but also suppresses the proliferation of those cancer cells already at a stage when SHBG production continues to be high. It is hence concluded that, in addition to the lignan enterolactone, the most abundant urinary isoflavonoid genistein stimulates SHBG production and inhibits Hep-G2 cancer cell proliferation (Mousavi et al., 1993).

Insulin-like Growth Factor-1 (IGF-1); Prostate Cancer

Elevated levels of insulin-like growth factor-1 (IGF-1) are associated with an increased risk of several different cancers, including prostate cancer. Inhibition of IGF-1 and the downstream signaling pathways mediated by the activation of the IGF-1 receptor (IGF-1R) may be involved in inhibiting prostate carcinogenesis. Genistein treatment caused a significant inhibition of IGF-1-stimulated cell growth. Flow cytometry analysis revealed that genistein significantly decreased the number of IGF-1-stimulated cells in the G0/G1 phase of the cell-cycle. In IGF-1-treated cells, genistein effectively inhibited the phosphorylation of IGF-1R and the phosphorylation of its downstream targets, such as Src, Akt, and glycogen synthase kinase-3β (GSk-3β). IGF-1 treatment decreased the levels of E-cadherin but increased the levels of β-catenin and cyclin D1.

However, genistein treatment greatly attenuated IGF-1-induced β-catenin signaling that correlated with increasing the levels of E-cadherin and decreasing cyclin D1 levels in PC-3 cells. In addition, genistein inhibited T-cell factor/lymphoid enhancer factor (TCF/LEF)-dependent transcriptional activity. These results showed that genistein effectively inhibited cell growth in IGF-1-stimulated PC-3 cells, possibly by inhibiting downstream of IGF-1R activation (Lee et al., 2012).

Sex Hormone-binding Globulin (SHBG); Hepatoma

Sex hormone-binding globulin (SHBG) is the main transport binding protein for sex steroid hormones in plasma and regulates their accessibility to target cells. Plasma SHBG is secreted by the liver under the control of hormones and nutritional factors. In the human hepatoma cell line (HepG2), thyroid and estrogenic hormones, and a variety of drugs including the anti-estrogen tamoxifen, the phytoestrogen, genistein and mitotane (Op'DDD) increase SHBG production and SHBG gene promoter activity. In contrast, monosaccharides (glucose or fructose) effectively decrease SHBG expression by inducing lipogenesis, which reduces hepatic HNF-4alpha levels, a transcription factor that plays a critical role in controlling the SHBG promoter. Interestingly, diminishing hepatic lipogenesis and free fatty acid liver biosynthesis also appear to be associated with the positive effects of thyroid hormones and PPARgamma antagonists on SHBG expression.

This mechanism provides a biological explanation for why SHBG is a sensitive biomarker of insulin resistance and the metabolic syndrome, and why low plasma SHBG levels are a risk factor for developing hyperglycemia and type 2 diabetes, especially in women (Pugeat et al., 2009).

Cancer: Pancreatic

Pancreatic cancer remains the fourth most common cause of cancer related death in the United States. Therefore, novel strategies for the prevention and treatment are urgently needed. Genistein is a prominent isoflavonoid found in soy products and has been proposed to be responsible for lowering the rate of pancreatic cancer in Asians. However, the molecular mechanism(s) by which genistein elicits its effects on pancreatic cancer cells has not been fully elucidated.

Wang et al., (2006) have previously shown that genistein induces apoptosis and inhibits the activation of nuclear factor kappaB (NF-kappaB) pathway. Moreover, Notch signaling is known to play a critical role in maintaining the balance between cell proliferation, differentiation and apoptosis, and thereby may contribute to the development of pancreatic cancer. Hence, in our study, they investigated whether there is any cross talk between Notch and NF-kappaB during genistein-induced apoptosis in BxPC-3 pancreatic cancer cells. They found that genistein inhibits cell growth and induces apoptotic processes in BxPC-3 pancreatic cancer cells.

This was partly due to inhibition of Notch-1 activity. BxPC-3 cells transfected with Notch-1 cDNA showed induction of NF-kappaB activity, and this was inhibited by genistein treatment. From these results, we conclude that the inhibition of Notch-1 and NF-kappaB activity and their cross talk provides a novel mechanism by which genistein inhibits cell growth and induces apoptotic processes in pancreatic cancer cells.

References

Anastasius N, Boston S, Lacey M, Storing N, Whitehead SA. (2009). Evidence that low-dose, long-term genistein treatment inhibits oestradiol-stimulated growth in MCF-7 cells by down-regulation of the PI3-kinase/Akt signaling pathway. J Steroid Biochem Mol Biol, 116(1-2):50-55.


Cappelletti V, Fioravanti L, Miodini P, Di Fronzo G J. (2000). Genistein blocks breast cancer cells in the G2M phase of the cell-cycle. Cell. Biochem, 79(4):594-600. doi: 10.1002/1097-4644(20001215)79:4<594::AID-JCB80>3.0.CO;2-4.


Chen SS, Michael A, Butler-Manuel SA. (2012). Advances in the treatment of ovarian cancer: a potential role of anti-inflammatory phytochemicals. Discov Med, 13(68):7-17.


Choi EJ, Kim GH. (2013). Anti-proliferative activity of daidzein and genistein may be related to ERα /c-erbB-2 expression in human breast cancer cells. Mol Med Rep, 7(3):781-4. doi: 10.3892/mmr.2013.1283.


Couse JF, Lindzey J, Grandien K, Gustafsson JA, Korach KS. (1997). Tissue distribution and quantitative analysis of estrogen receptor-alpha (ERalpha) and estrogen receptor-beta (ERbeta) messenger ribonucleic acid in the wild-type and ERalpha-knockout mouse. Endocrinology, 138(1997):4613–4621


Dandawate P, Padhye S, Ahmad A, Sarkar FH. (2013). Novel strategies targeting cancer stem cells through phytochemicals and their analogs. Drug Deliv Transl Res, 3(2):165-182.


Enmark E, Peltohuikko M, Grandien K, et al. (1997). Human estrogen receptor beta-gene structure, chromosomal localization, and expression pattern. J. Clin. Endocrinol. Metab, 82(1997):4258–4265.


Fotsis T, Pepper M, Adlercreutz H, et al. (1993). Genistein, a dietary-derived inhibitor of in vitro angiogenesis. Proc Natl Acad Sci, 90(7):2690-4.


Harris HA, Albert LM, Leathurby Y, et al. (2002). Evaluation of an estrogen receptor- β agonist in animal models of human disease. Endocrinology, 144(2003):4241–4249


Harris HA, Katzenellenbogen JA, Katzenellenbogen BS. (2002). Characterization of the biological roles of the estrogen receptors, ER alpha and ER beta, in estrogen target tissues in vivo through the use of an ER alpha-selective ligand. Endocrinology, 143(2002):4172–4177.


Hillman GG, Wang Y, Che M, et al. (2007). Progression of renal cell carcinoma is inhibited by genistein and radiation in an orthotopic model. BMC Cancer, 7:4. doi:10.1186/1471-2407-7-4.


Lee J, Ju J, Park S, et al. (2012). Inhibition of IGF-1 Signaling by Genistein: Modulation of E-Cadherin Expression and Down-regulation of β -Catenin Signaling in Hormone Refractory PC-3 Prostate Cancer Cells. Nutrition and Cancer, 64(1). doi:10.1080/01635581.2012.630161


Liu YL, Zhang GQ, Yang Y, et al. (2013). Genistein Induces G2/M Arrest in Gastric Cancer Cells by Increasing the Tumor Suppressor PTEN Expression. Nutr Cancer.


Marik R, Allu M, Anchoori R, et al. (2011). Potent genistein derivatives as inhibitors of estrogen receptor alpha-positive breast cancer. Cancer Biol Ther, 11(10):883-92.


Mousavi Y, Adlercreutz H. (1993). Genistein is an effective stimulator of sex hormone-binding globulin production in hepatocarcinoma human liver cancer cells and suppresses proliferation of these cells in culture. Steroids, 58(7):301-4.


Okabe Y, Shimazu T, Tanimoto H. (2011). Higher bioavailability of isoflavones after a single ingestion of aglycone-rich fermented soybeans compared with glucoside-rich non-fermented soybeans in Japanese postmenopausal women. J Sci Food Agric, 91(4):658-63. doi: 10.1002/jsfa.4228.


Peterson G, Barnes S. (1991). Genistein inhibition of the growth of human breast cancer cells: independence from estrogen receptors and the multi-drug resistance gene. Biochemical and Biophysical Research Communications, 179(1):661-667. doi:10.1016/0006-291X(91)91423-A.


Pike ACW, Brzozowski AM, Hubbard RE, et al. (1999). Structure of the ligand-binding domain of oestrogen receptor beta in the presence of a partial agonist and a full antagonist. EMBO J, 18(1999): 4608–4618


Pugeat M, Nader N, Hogeveen K, et al. (2010). Sex hormone-binding globulin gene expression in the liver: Drugs and the metabolic syndrome. Mol Cell Endocrinol, 316(1):53-9. doi: 10.1016/j.mce.2009.09.020.


Rahal OM, Simmen RC. (2011). Paracrine-Acting Adiponectin Promotes Mammary Epithelial Differentiation and Synergizes with Genistein to Enhance Transcriptional Response to Estrogen Receptor β Signaling. Endocrinology, 152(9):3409-21. doi: 10.1210/en.2011-1085.


Rietjens IM, Sotoca AM, Vervoort J, Louisse J. (2013). Mechanisms underlying the dualistic mode of action of major soy isoflavones in relation to cell proliferation and cancer risks. Mol Nutr Food Res, 57(1):100-13. doi: 10.1002/mnfr.201200439.


Rowland IR, Wiseman H, Sanders TA, Adlercreutz H, Bowey EA. (2000). Interindividual variation in metabolism of soy isoflavones and lignans: influence of habitual diet on equol production by the gut microflora. Nutr Cancer, 36(1):27-32.


Saji S, Hirose M, Toi M. (2005). Clinical significance of estrogen receptor beta in breast cancer. Cancer Chemother Pharmacol, 56(1):21-6.


Setchell KD, Brown NM, Summer S, et al. (2013). Dietary Factors Influence Production of the Soy Isoflavone Metabolite S-(-)Equol in Healthy Adults. J Nutr.


Shao ZM, Shen ZZ, Fontana JA, Barsky SH. (2000). Genistein's ER-dependent and independent actions are mediated through ER pathways in ER-positive breast carcinoma cell lines. Anti-cancer Res, 20(4):2409-16.


Stark A, Madar Z. (2002). Phytoestrogens: a review of recent findings. J Pediatr Endocrinol Metab, 15(5):561-72.


Tseng M, Byrne C, Kurzer MS, Fang CY. (2013). Equol-producing status, isoflavone intake, and breast density in a sample of u.s. Chinese women. Cancer Epidemiol Biomarkers Prev, 22(11):1975-83. doi: 10.1158/1055-9965.EPI-13-0593.


Xin Z, Siji L, Yan D, Weijuan X, Jie S, Qianyu W. (2009). Influence of Genistein and Daidzein on estrogen-related receptor- α in an Endometrial Carcinoma Cell Line. Tong Ji Da Xue Xue Bao (Yi Xue Ban), 30(4): 12-17.


Xu J, Loo G. (2001). Different effects of genistein on molecular markers related to apoptosis in two phenotypically dissimilar breast cancer cell lines. Journal of Cellular Biochemistry, 82(1), 78-88.

Wang Z, Zhang Y, Banerjee S, Li Y, Sarkar FH. (2006) Inhibition of nuclear factor kappab activity by genistein is mediated via Notch-1 signaling pathway in pancreatic cancer cells. Int J Cancer. 2006 Apr 15;118(8):1930-6.