Category Archives: TRAIL

Wogonin

Cancer:
Breast, lung (NSCLC), gallbladder carcinoma, osteosarcoma, colon, cervical

Action: Neuro-protective, anti-lymphangiogenesis, anti-angiogenic, anti-estrogenic, chemo-sensitizer, pro-oxidative, hypoxia-induced drug resistance, anti-metastatic, anti-tumor, anti-inflammatory

Wogonin is a plant monoflavonoid isolated from Scutellaria rivularis (Benth.) and Scutellaria baicalensis (Georgi).

Breast Cancer; ER+ & ER-

Effects of wogonin were examined in estrogen receptor (ER)-positive and -negative human breast cancer cells in culture for proliferation, cell-cycle progression, and apoptosis. Cell growth was attenuated by wogonin (50-200 microM), independently of its ER status, in a time- and concentration-dependent manner. Apoptosis was enhanced and accompanied by up-regulation of PARP and Caspase 3 cleavages as well as pro-apoptotic Bax protein. Akt activity was suppressed and reduced phosphorylation of its substrates, GSK-3beta and p27, was observed. Suppression of Cyclin D1 expression suggested the down-regulation of the Akt-mediated canonical Wnt signaling pathway.

ER expression was down-regulated in ER-positive cells, while c-ErbB2 expression and its activity were suppressed in ER-negative SK-BR-3 cells. Wogonin feeding to mice showed inhibition of tumor growth of T47D and MDA-MB-231 xenografts by up to 88% without any toxicity after 4 weeks of treatment. As wogonin was effective both in vitro and in vivo, our novel findings open the possibility of wogonin as an effective therapeutic and/or chemo-preventive agent against both ER-positive and -negative breast cancers, particularly against the more aggressive and hormonal therapy-resistant ER-negative types (Chung et al., 2008).

Neurotransmitter Action

Kim et al. (2011) found that baicalein and wogonin activated the TREK-2 current by increasing the opening frequency (channel activity: from 0.05 ± 0.01 to 0.17 ± 0.06 in baicalein treatment and from 0.03 ± 0.01 to 0.29 ± 0.09 in wogonin treatment), while leaving the single-channel conductance and mean open time unchanged. Baicalein continuously activated TREK-2, whereas wogonin transiently activated TREK-2. Application of baicalein and wogonin activated TREK-2 in both cell attached and excised patches, suggesting that baicalein and wogonin may modulate TREK-2 either directly or indirectly with different mechanisms. These results suggest that baicalein- and wogonin-induced TREK-2 activation help set the resting membrane potential of cells exposed to pathological conditions and thus may give beneficial effects in neuroprotection.

Anti-metastasic

The migration and invasion assay was used to evaluate the anti-metastasis effect of wogonin. Wogonin at the dose of 1–10 µM, which did not induce apoptosis, significantly inhibited the mobility and invasion activity of human gallbladder carcinoma GBC-SD cells. In addition, the expressions of matrix metalloproteinase (MMP)-2, MMP-9 and phosphorylated extracellular regulated protein kinase 1/2 (ERK1/2) but not phosphorylated Akt were dramatically suppressed by wogonin in a concentration-dependent manner. Furthermore, the metastasis suppressor maspin was confirmed as the downstream target of wogonin.

These findings suggest that wogonin inhibits cell mobility and invasion by up-regulating the metastasis suppressor maspin. Together, these data provide novel insights into the chemo-protective effect of wogonin, a main active ingredient of Chinese medicine Scutellaria baicalensis (Dong et al., 2011).

Anti-tumor and Anti-metastatic

Kimura & Sumiyoshi (2012) examined the effects of wogonin isolated from Scutellaria baicalensis roots on tumor growth and metastasis using a highly metastatic model in osteosarcoma LM8-bearing mice. Wogonin (25 and 50mg/kg, twice daily) reduced tumor growth and metastasis to the lung, liver and kidney, angiogenesis (CD31-positive cells), lymphangiogenesis (LYVE-1-positive cells), and TAM (F4/80-positive cell) numbers in the tumors of LM8-bearing mice. Wogonin (10–100µM) also inhibited increases in IL-1β production and cyclooxygenase (COX)-2 expression induced by lipopolysaccharide in THP-1 macrophages. The anti-tumor and anti-metastatic actions of wogonin may be associated with the inhibition of VEGF-C-induced lymphangiogenesis through a reduction in VEGF-C-induced VEGFR-3 phosphorylation by the inhibition of COX-2 expression and IL-1β production in Tumor-associated macrophages (TAMs).

Anti-inflammatory

Wogonin extracted from Scutellariae baicalensis and S. barbata is a cell-permeable and orally available flavonoid that displays anti-inflammatory properties. Wogonin is reported to suppress the release of NO by iNOS, PGE2 by COX-2, pro-inflammatory cytokines, and MCP-1 gene expression and NF-kB activation (Chen et al., 2008).

Hypoxia-Induced Drug Resistance (MDR)

Hypoxia-induced drug resistance is a major obstacle in the development of effective cancer therapy. The reversal abilities of wogonin on   hypoxia resistance were examined and the underlying mechanisms discovered. MTT assay revealed that hypoxia increased maximal 1.71-, 2.08-, and 2.15-fold of IC50 toward paclitaxel, ADM, and DDP in human colon cancer cell lines HCT116, respectively. Furthermore, wogonin showed strong reversal potency in HCT116 cells in hypoxia and the RF reached 2.05. Hypoxia-inducible factor-1α (HIF-1α) can activate the expression of target genes involved in glycolysis. Wogonin decreased the expression of glycolysis-related proteins (HKII, PDHK1, LDHA), glucose uptake, and lactate generation in a dose-dependent manner.

In summary, wogonin could be a good candidate for the development of a new multi-drug resistance (MDR) reversal agent and its reversal mechanism probably is due to the suppression of HIF-1α expression via inhibiting PI3K/Akt signaling pathway (Wang et al., 2013).

NSCLC

Wogonin, a flavonoid originated from Scutellaria baicalensis Georgi, has been shown to enhance TRAIL-induced apoptosis in malignant cells in in vitro studies. In this study, the effect of a combination of TRAIL and wogonin was tested in a non-small-cell lung cancer xenografted tumor model in nude mice. Consistent with the in vitro study showing that wogonin sensitized A549 cells to TRAIL-induced apoptosis, wogonin greatly enhanced TRAIL-induced suppression of tumor growth, accompanied with increased apoptosis in tumor tissues as determined by TUNEL assay.

The down-regulation of these antiapoptotic proteins was likely mediated by proteasomal degradation that involved intracellular reactive oxygen species (ROS), because wogonin robustly induced ROS accumulation and ROS scavengers butylated hydroxyanisole (BHA) and N-acetyl-L-cysteine (NAC) and the proteasome inhibitor MG132 restored the expression of these antiapoptotic proteins in cells co-treated with wogonin and TRAIL.

These results show for the first time that wogonin enhances TRAIL's anti-tumor activity in vivo, suggesting this strategy has an application potential for clinical anti-cancer therapy (Yang et al., 2013).

Colon Cancer

Following treatment with baicalein or wogonin, several apoptotic events were observed, including DNA fragmentation, chromatin condensation and increased cell-cycle arrest in the G1 phase. Baicalein and wogonin decreased Bcl-2 expression, whereas the expression of Bax was increased in a dose-dependent manner compared with the control. Furthermore, the induction of apoptosis was accompanied by an inactivation of phosphatidylinositol 3-kinase (PI3K)/Akt in a dose-dependent manner.

The administration of baicalein to mice resulted in the inhibition of the growth of HT-29 xenografts without any toxicity following 5 weeks of treatment. The results indicated that baicalein induced apoptosis via Akt activation in a p53-dependent manner in the HT-29 colon cancer cells and that it may serve as a chemo-preventive or therapeutic agent for HT-29 colon cancer (Kim et al., 2012).

Breast

The involvement of insulin-like growth factor-1 (IGF-1) and estrogen receptor α (ERα) in the inhibitory effect of wogonin on the breast adenocarcinoma growth was determined. Moreover, the effect of wogonin on the angiogenesis of chick chorioallantoic membrane (CAM) was also investigated. The results showed wogonin and ICI182780 both exhibited a potent ability to blunt IGF-1-stimulated MCF-7 cell growth. Either of wogonin and ICI182780 significantly inhibited ERα and p-Akt expressions in IGF-1-treated cells. The inhibitory effect of wogonin showed no difference from that of ICI182780 on IGF-1-stimulated expressions of ERα and p-Akt. Meanwhile, wogonin at different concentrations showed significant inhibitory effect on CAM angiogenesis.

These results suggest the inhibitory effect of wogonin on breast adenocarcinoma growth via inhibiting IGF-1-mediated PI3K-Akt pathway and regulating ERα expression. Furthermore, wogonin has a strong anti-angiogenic effect on CAM model (Ma et al., 2012).

Chemoresistance; Cervical Cancer, NSCLC

Chemoresistance to cisplatin is a major limitation of cisplatin-based chemotherapy in the clinic. The combination of cisplatin with other agents has been recognized as a promising strategy to overcome cisplatin resistance. Previous studies have shown that wogonin (5,7-dihydroxy-8-methoxyflavone), a flavonoid isolated from the root of the medicinal herb Scutellaria baicalensis Georgi, sensitizes cancer cells to chemotheraputics such as etoposide, adriamycin, tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and TNF.

In this study, the non-small-cell lung cancer cell line A549 and the cervical cancer cell line HeLa were treated with wogonin or cisplatin individually or in combination. It was found for the first time that wogonin is able to sensitize cisplatin-induced apoptosis in both A549 cells and HeLa cells as indicated by the potentiation of activation of caspase-3, and cleavage of the caspase-3 substrate PARP in wogonin and cisplatin co-treated cells.

Results provided important new evidence supporting the potential use of wogonin as a cisplatin sensitizer for cancer therapy (He et al., 2012).

References

Chen LG, Hung LY, Tsai KW, et al. (2008). Wogonin, a bioactive flavonoid in herbal tea, inhibits inflammatory cyclooxygenase-2 gene expression in human lung epithelial cancer cells. Mol Nutr Food Res. 52:1349-1357.


Chung H, Jung YM, Shin DH, et al. (2008). Anti-cancer effects of wogonin in both estrogen receptor-positive and -negative human breast cancer cell lines in vitro and in nude mice xenografts. Int J Cancer, 122(4):816-22.


Dong P, Zhang Y, Gu J, et al. (2011). Wogonin, an active ingredient of Chinese herb medicine Scutellaria baicalensis, inhibits the mobility and invasion of human gallbladder carcinoma GBC-SD cells by inducing the expression of maspin. J Ethnopharmacol, 137(3):1373-80. doi: 10.1016/j.jep.2011.08.005.


He F, Wang Q, Zheng XL, et al. (2012). Wogonin potentiates cisplatin-induced cancer cell apoptosis through accumulation of intracellular reactive oxygen species. Oncology Reports, 28(2), 601-605. doi: 10.3892/or.2012.1841.


Kim EJ, Kang D, Han J. (2011). Baicalein and wogonin are activators of rat TREK-2 two-pore domain K+ channel. Acta Physiologica, 202(2):185–192. doi: 10.1111/j.1748-1716.2011.02263.x.


Kim SJ, Kim HJ, Kim HR, et al. (2012). Anti-tumor actions of baicalein and wogonin in HT-29 human colorectal cancer cells. Mol Med Rep, 6(6):1443-9. doi: 10.3892/mmr.2012.1085.


Kimura Y & Sumiyoshi M. (2012). Anti-tumor and anti-metastatic actions of wogonin isolated from Scutellaria baicalensis roots through anti-lymphangiogenesis. Phytomedicine, 20(3-4):328-336. doi:10.1016/j.phymed.2012.10.016


Ma X, Xie KP, Shang F, et al. (2012). Wogonin inhibits IGF-1-stimulated cell growth and estrogen receptor α expression in breast adenocarcinoma cell and angiogenesis of chick chorioallantoic membrane. Sheng Li Xue Bao, 64(2):207-12.


Wang H, Zhao L, Zhu LT, et al. (2013). Wogonin reverses hypoxia resistance of human colon cancer HCT116 cells via down-regulation of HIF-1α and glycolysis, by inhibiting PI3K/Akt signaling pathway. Mol Carcinog. doi: 10.1002/mc.22052.


Yang L, Wang Q, Li D, et al. (2013). Wogonin enhances anti-tumor activity of tumor necrosis factor-related apoptosis-inducing ligand in vivo through ROS-mediated down-regulation of cFLIPL and IAP proteins. Apoptosis, 18(5):618-26. doi: 10.1007/s10495-013-0808-8.

Resveratrol 98%

Cancer:
Breast, lymphoma, breast, gastric, colorectal, esophageal, prostate, pancreatic, leukemia, skin, lung

Action: Chemoprevention, anti-inflammatory, MDR, chemotherapy-induced cytotoxicity, radio-sensitizer, enhances chemo-sensitivity

Resveratrol (RSV) is a phytoalexin found in food products including berries and grapes, as well as plants (including Fallopia japonica (Houtt.), Gnetum cleistostachyum (C. Y. Cheng), Vaccinium arboretum (Marshall), Vaccinium angustifolium (Aiton) and Vaccinium corymbosum (L.)

Although resveratrol is ubiquitous in nature, it is found in a limited number of edible substances, most notably in grapes. In turn, due to the peculiar processing methodology, resveratrol is found predominantly in red wines. Thus, resveratrol received intense and immediate attention. A large number of resveratrol anti-cancer activities were reported, affecting all the steps of cancerogenesis, namely initiation, promotion, and progression. Thereafter, an exponential number of reports on resveratrol accumulated and, so far, more than 5,000 studies have been published (Borriello et al., 2014).

Up to the end of 2011, more than 50 studies analyzed the effect of resveratrol as an anti-cancer compound in animal models of different cancers, including skin cancer (non-melanoma skin cancer and melanoma); breast, gastric, colorectal, esophageal, prostate, and pancreatic cancers; hepatoma, neuroblastoma, fibrosarcoma, and leukemia (Ahmad et al., 2004; Hayashibara et al., 2002; Pozo-Guisado et al., 2005; Mohan et al., 2006; Tang et al., 2006). In general, these preclinical studies suggest a positive activity of the molecule in lowering the progression of cancer, reducing its dimension, and decreasing the number of metastases (Vang et al., 2011).

Breast

Resveratrol was shown to have cancer chemo-preventive activity in assays representing three major stages of carcinogenesis. It has been found to mediate anti-inflammatory effects and inhibit cyclooxygenase and hydroperoxidase functions (anti-promotion activity). It has also been found to inhibit the development of pre-neoplastic lesions in carcinogen-treated mouse mammary glands in culture and inhibited tumorigenesis in a mouse skin cancer model (Jang et al., 1997).

In addition, resveratrol, a partial ER agonist itself, acts as an ER antagonist in the presence of estrogen leading to inhibition of human breast cancer cells (Lu et al., 1999).

Besides chemo-preventive effects, resveratrol appears to exhibit therapeutic effects against cancer itself. Limited data in humans have revealed that RSV is pharmacologically safe (Aggarwal et al., 2004).

Chemotherapy-Induced Cytotoxicity

RSV markedly enhanced Dox-induced cytotoxicity in MCF-7/adr and MDA-MB-231 cells. Treatment with a combination of RSV and Dox significantly increased the cellular accumulation of Dox by down-regulating the expression levels of ATP-binding cassette (ABC) transporter genes, MDR1, and MRP1. Further in vivo experiments in the xenograft model revealed that treatment with a combination of RSV and Dox significantly inhibited tumor volume by 60%, relative to the control group.

These results suggest that treatment with a combination of RSV and Dox would be a helpful strategy for increasing the efficacy of Dox by promoting an intracellular accumulation of Dox and decreasing multi-drug resistance in human breast cancer cells (Kim et al., 2013).

Radio-sensitizer/Lung Cancer

Previous studies indicated that resveratrol (RV) may sensitize tumor cells to chemotherapy and ionizing radiation (IR). However, the mechanisms by which RV increases the radiation sensitivity of cancer cells have not been well characterized. Here, we show that RV treatment enhances IR-induced cell killing in non-small-cell lung cancer (NSCLC) cells through an apoptosis-independent mechanism. Further studies revealed that the percentage of senescence-associated β-galactosidase (SA-β-gal)-positive senescent cells was markedly higher in cells treated with IR in combination with RV compared with cells treated either with IR or RV alone, suggesting that RV treatment enhances IR-induced premature senescence in lung cancer cells.

Collectively, these results demonstrate that RV-induced radio-sensitization is associated with significant increase of ROS production, DNA-DSBs and senescence induction in irradiated NSCLC cells, suggesting that RV treatment may sensitize lung cancer cells to radiotherapy via enhancing IR-induced premature senescence (Luo et al., 2013).

Lymphoma

Ko et al. (2011) examined the effects of resveratrol on the anaplastic large-cell lymphoma (ALCL) cell line SR-786. Resveratrol inhibited growth and induced cellular differentiation, as demonstrated by morphological changes and elevated expression of T cell differentiation markers CD2, CD3, and CD8. Resveratrol also triggered cellular apoptosis, as demonstrated by morphological observations, DNA fragmentation, and cell-cycle analyzes. Further, the surface expression of the death receptor Fas/CD95 was increased by resveratrol treatment. Our data suggest that resveratrol may have potential therapeutic value for ALCL.

Skin Cancer

Treatment with combinations of resveratrol and black tea polyphenol (BTP) also decreased expression of proliferating cell nuclear antigen in mouse skin tissues/tumors than their solitary treatments as determined by immunohistochemistry. In addition, histological and cell death analysis also confirmed that resveratrol and BTP treatment together inhibits cellular proliferation and markedly induces apoptosis. Taken together, results for the first time lucidly illustrate that resveratrol and BTP in combination impart better suppressive activity than either of these agents alone and accentuate that development of novel combination therapies/chemo-prevention using dietary agents will be more beneficial against cancer (George et al., 2011).

Prostate Cancer

Resveratrol-induced ROS production, caspase-3 activity and apoptosis were inhibited by N-acetylcysteine. Bax was a major pro-apoptotic gene mediating the effects of resveratrol as Bax siRNA inhibited resveratrol-induced apoptosis. Resveratrol enhanced the apoptosis-inducing potential of TRAIL, and these effects were inhibited by either dominant negative FADD or caspase-8 siRNA. The combination of resveratrol and TRAIL enhanced the mitochondrial dysfunctions during apoptosis. These properties of resveratrol strongly suggest that it could be used either alone or in combination with TRAIL for the prevention and/or treatment of prostate cancer (Shankar et al., 2007).

Breast Cancer

Scarlatti et al. (2008) demonstrate that resveratrol acts via multiple pathways to trigger cell death, induces caspase-dependent and caspase-independent cell death in MCF-7 casp-3 cells, induces only caspase-independent cell death in MCF-7vc cells, and stimulates macroautophagy. Using BECN1 and hVPS34 (human vacuolar protein sorting 34) small interfering RNAs, they demonstrated that resveratrol activates Beclin 1-independent autophagy in both cell lines, whereas cell death via this uncommon form of autophagy occurs only in MCF-7vc cells. They also show that this variant form of autophagic cell death is blocked by the expression of caspase-3, but not by its enzymatic activity. In conclusion, this study reveals that non-canonical autophagy induced by resveratrol can act as a caspase-independent cell death mechanism in breast cancer cell.

References

Aggarwal BB, Bhardwaj A, Aggarwal RS et al. (2004). Role of Resveratrol in Prevention and Therapy of Cancer: Preclinical and Clinical Studies. Anti-cancer Research, 24(5A): 2783-2840.


Ahmad KA, Clement MV, Hanif IM, et al (2004). Resveratrol inhibits drug-induced apoptosis in human leukemia cells by creating an intracellular milieu nonpermissive for death execution. Cancer Res, 64:1452–1459


Borriello A, Bencivenga D, Caldarelli I, et al. (2014). Resveratrol: from basic studies to bedside. Cancer Treat Res, 159:167-84. doi: 10.1007/978-3-642-38007-5_10.


George J, Singh M, Srivastava AK, et al (2011). Resveratrol and black tea polyphenol combination synergistically suppress mouse skin tumors growth by inhibition of activated MAPKs and p53. PLoS ONE, 6:e23395


Hayashibara T, Yamada Y, Nakayama S, et al (2002). Resveratrol induces down-regulation in survivin expression and apoptosis in HTLV-1-infected cell lines: a prospective agent for adult T cell leukemia chemotherapy. Nutr Cancer, 44:193–201


Jang M, Cai L, Udeani GO, et al. (1997). Cancer Chemo-preventive Activity of Resveratrol, a Natural Product Derived from Grapes. Science, 275(5297):218-220.


Kim TH, Shin YJ, Won AJ, et al. (2013). Resveratrol enhances chemosensitivity of doxorubicin in Multi-drug-resistant human breast cancer cells via increased cellular influx of doxorubicin. Biochim Biophys Acta, S0304-4165(13)00463-7. doi: 10.1016/j.bbagen.2013.10.023.


Ko YC, Chang CL, Chien HF, et al (2011). Resveratrol enhances the expression of death receptor Fas/CD95 and induces differentiation and apoptosis in anaplastic large-cell lymphoma cells. Cancer Lett, 309:46–53


Lu R, Serrero G. (1999). Resveratrol, a natural product derived from grape, exhibits antiestrogenic activity and inhibits the growth of human breast cancer cells. Journal of Cellular Physiology, 179(3):297-304.


Luo H, Wang L, Schulte BA, et al. (2013). Resveratrol enhances ionizing radiation-induced premature senescence in lung cancer cells. Int J Oncol, 43(6):1999-2006. doi: 10.3892/ijo.2013.2141.


Mohan J, Gandhi AA, Bhavya BC, et al. (2006). Caspase-2 triggers Bax-Bak-dependent and – independent cell death in colon cancer cells treated with resveratrol. J Biol Chem, 281:17599–17611


Pozo-Guisado E, Merino JM, Mulero-Navarro S, et al. (2005). Resveratrol-induced apoptosis in MCF-7 human breast cancer cells involves a caspase-independent mechanism with down-regulation of Bcl-2 and NF-kappaB. Int J Cancer, 115:74–84.


Scarlatti F, Maffei R, Beau I, et al (2008). Role of non-canonical Beclin 1-independent autophagy in cell death induced by resveratrol in human breast cancer cells. Cell Death Differ, 8:1318–1329


Shankar S, Siddiqui I, Srivastava RK. (2007). Molecular mechanisms of resveratrol (3,4,5- trihydroxy-trans-stilbene) and its interaction with TNF-related apoptosis inducing ligand (TRAIL) in androgen-insensitive prostate cancer cells. Mol Cell Biochem, 304:273–285


Tang HY, Shih A, Cao HJ, et al. (2006). Resveratrol-induced cyclooxygenase-2 facilitates p53-dependent apoptosis in human breast cancer cells. Mol Cancer Ther, 5:2034–2042


Vang O, Ahmad N, Baile CA, et al. (2011). What is new for an old molecule? Systematic review and recommendations on the use of resveratrol. PLoS ONE, 6:e19881

Periplocin

Cancer: Lung, colorectal, leukemia

Action: Apoptosis-inducing, cytostatic effect

Apoptosis

The anti-tumor component of Cortex periplocae is periplocin. Periplocin is one of the cardenolides isolated from cortex periplocae which is used for treatment of rheumatoid arthritis and reinforcement of bones and tendons in traditional medicine.

Periplocin has been reported to inhibit many cell lines, including MCF-7, TE-13, QG-56, SMMC-7721, T24, Hela, K562, TE-13 and Eca-109 cells. Studies have shown that periplocin reduces the expression of survivin, an inhibitor of apoptosis. It also releases caspases-3 and -7 from complexes and thereby increases their activities, ultimately inducing tumor cell apoptosis (Zhao et al., 2009).

Lung Cancer

The anti-tumor activity of periplocin was investigated in lung cancer cells both in vitro and in vivo, and its anti-cancer mechanism was explored. Periplocin inhibited the growth of lung cancer cells and induced their apoptosis in a time- and dose-dependent manner by cell-cycle arrest in G0/G1 phase. Periplocin exhibited anti-tumor activity both in human (A549) and mouse (LL/2) lung cancer xenograft models. Immunohistochemical analysis revealed that intratumoral angiogenesis was significantly suppressed.

Furthermore, anti-cancer activity mediated by periplocin was associated with decreased level of phosphorylated AKT and ERK both in vitro and in vivo, which are important for cell growth and survival. Moreover, periplocin induced apoptosis by down-regulating Bcl-2 and up-regulating Bax, leading to activation of caspase-3 and caspase-9.

These findings suggest that periplocin could inhibit the growth of lung cancer both in vitro and in vivo, which could be attributed to the inhibition of proliferation and the induction of apoptosis signaling pathways, such as AKT and ERK. These observations provide further evidence on the anti-tumor effect of periplocin, and it may be of importance to further explore its potential role as a therapeutic agent for cancer (Lu et al., 2010).

Colorectal Carcinomas

The Wnt/beta-catenin signaling pathway plays an important role in the development and progression of human cancers, especially in colorectal carcinomas. Periplocin extracted from cortex periplocae (CPP) significantly inhibited the proliferation of SW480 cells in a time-and dose-dependent manner (P<0.01). CPP (0.5 microg/mL) also caused G0/G1 cell-cycle arrest of SW480 cells and induced cell apoptosis (P<0.05). Compared to untreated control cells, after the treatment with CPP, the protein levels of beta-catenin in total cell lysates, cytosolic extracts, and nuclear extracts were reduced (P<0.01); the binding activity of the TCF complex in nucleus to its specific DNA binding site was suppressed; mRNAs of the downstream target genes survivin, c-myc and cyclin D1 were decreased (P<0.01) while beta-catenin mRNA remained unchanged.

CPP could significantly inhibit the proliferation of SW480 cells, which may be through down-regulating the Wnt/beta-catenin signaling pathway (Du et al., 2009).

Pro-apoptotic and Cytostatic Effect/Leukemia

Cardenoliddes are steroid glycosides which are known to exert cardiotonic effects by inhibiting the Na(+)/K(+)-ATPase. Several of these compounds have been shown also to possess anti-tumor potential. The aim of the present work was the characterization of the tumor cell growth inhibition activity of four cardenolides, isolated from Periploca graeca L., and the mechanisms underlying such an effect.

The pro-apoptotic and cytostatic effect of the compounds was tested in U937 (monocytic leukemia) and PC3 (prostate adenocarcinoma). Characterization of apoptosis and cell-cycle impairment was obtained by cytofluorimetry and WB. Periplocymarin and periplocin were the most active compounds, periplocymarin being more effective than the reference compound ouabain. The reduction of cell number by these two cardenolides was due in PC3 cells mainly to the activation of caspase-dependent apoptotic pathways, while in U937 cells to the induction of cell-cycle impairment without extensive cell death. Interestingly, periplocymarin, at cytostatic but non-cytotoxic doses, was shown to sensitize U937 cells to TRAIL. Taken together, these data outline that cardiac glycosides are promising anti-cancer drugs and contribute to the identification of new natural cardiac glycosides to obtain chemically modified non-cardioactive/low toxic derivatives with enhanced anti-cancer potency (Bloise et al., 2009).

References

Bloise E, Braca A, De Tommasi N, Belisario MA. (2009). Pro-apoptotic and cytostatic activity of naturally occurring cardenolides. Cancer Chemother Pharmacol, 64(4):793-802. doi: 10.1007/s00280-009-0929-5.


Du YY, Liu X, Shan BE. (2009). Periplocin extracted from cortex periplocae induces apoptosis of SW480 cells through inhibiting the Wnt/beta-catenin signaling pathway. Ai Zheng, 28(5):456-60.


Lu ZJ, Zhou Y, Song Q, et al. (2010). Periplocin inhibits growth of lung cancer in vitro and in vivo by blocking AKT/ERK signaling pathways. Cell Physiol Biochem, 26(4-5):609-18. doi: 10.1159/000322328.


Zhao LM, Ai J, Zhang Q, et al. (2009). Periplocin (a sort of ethanol from Cortex periplocae) induces apoptosis of esophageal carcinoma cells by influencing expression of related genes. Tumor (Chin), 29:1025-1030.

Luteolin

Cancer: Colorectal., pancreatic, ovarian, breast

Action: Anti-inflammatory, radio-protective, TAM chemo-sensitizer

Luteolin is a flavonoid found in many plants and foods, including Terminalia chebula (Retz.), Prunella vulgaris (L.) and Perilla frutescens [(L.) Britton].

Luteolin is contained in Ocimum sanctum L. or Ocimum tenuiflorum L, commonly known as Holy Basil in English or Tulsi in various Indian languages; it is an important medicinal plant in the various traditional and folk systems of medicine in Southeast Asia. Scientific studies have shown it to possess anti-inflammatory, anti-analgesic, anti-pyretic, anti-diabetic, hepato-protective, hypolipidemic, anti-stress, and immunomodulatory activities. It has been found to prevent chemical-induced skin, liver, oral., and lung cancers and mediates these effects by increasing the anti-oxidant activity, altering the gene expressions, inducing apoptosis, and inhibiting angiogenesis and metastasis.

Radio-protective

The aqueous extract of Tulsi 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. The chemo-preventive and radio-protective properties of Tulsi emphasize aspects that warrant future research to establish its activity and utility in cancer prevention and treatment (Baliga et al., 2013).

Anti-inflammatory

Pre-treatment of RAW 264.7 with luteolin, luteolin-7-glucoside, quercetin, and the isoflavonoid genistein inhibited both the LPS-stimulated TNF-αand interleukin-6 release, whereas eriodictyol and hesperetin only inhibited TNF-αrelease. From the compounds tested luteolin and quercetin were the most potent in inhibiting cytokine production with an IC50 of less than 1 and 5 µM for TNF-αrelease, respectively. Pre-treatment of the cells with luteolin attenuated LPS-induced tyrosine phosphorylation of many discrete proteins. Luteolin inhibited LPS-induced phosphorylation of Akt. Treatment of macrophages with LPS resulted in increased IκB-αphosphorylation and reduced the levels of IκB-α. It was concluded that luteolin inhibits protein tyrosine phosphorylation, nuclear factor-κB-mediated gene expression and pro-inflammatory cytokine production in murine macrophages (Xagorari et al., 2001).

Luteolin (Lut) possesses significant anti-inflammatory activity in well established models of acute and chronic inflammation, such as xylene-induced ear edema in mice (ED50= 107 mg/ kg), carrageenin-induced swellingof the ankle, acetic acid-induced pleurisy and croton oil-induced gaseous pouch granuloma in rats. Its combined immunostimulatory and anti-inflammatory activity, and inhibitory effect upon immediate hypersensitive response provide the pharmacologic bases for the beneficial effects of Lut in the treatment of chronic bronchitis (Chen et al., 1986).

Anti-inflammatory; Lung

Luteolin dose-dependently inhibited the expression and production of nitric oxide (NO) and prostaglandin E2 (PGE2), as well as the expression of inducible NO synthase (iNOS), cyclooxygenase-2 (COX-2), tumor necrosis factor-alpha (TNF-α), and interleukin-6 (IL-6). Luteolin also reduced the DNA binding activity of nuclear factor-kappa B (NF-κB) in LPS-activated macrophages. Moreover, luteolin blocked the degradation of IκB-α and nuclear translocation of NF-κB p65 subunit.

In sum, these data suggest that, by blocking NF-κ>B and AP-1 activation, luteolin acts to suppress the LPS-elicited inflammatory events in mouse alveolar macrophages, and this effect was mediated, at least in part, by inhibiting the generation of reactive oxygen species. These observations suggest a possible therapeutic application of this agent for treating inflammatory disorders in the lung (Chen et al., 2007).

Anti-inflammatory; Neuroinflammation

Pre-treatment of primary murine microglia and BV-2 microglial cells with luteolin inhibited LPS-stimulated IL-6 production at both the mRNA and protein levels. Whereas luteolin had no effect on the LPS-induced increase in NF-κB DNA binding activity, it markedly reduced AP-1 transcription factor binding activity. To determine whether luteolin might have similar effects in vivo, mice were provided drinking water supplemented with luteolin for 21 days and then they were injected i.p. with LPS. Luteolin consumption reduced LPS-induced IL-6 in plasma 4 hours after injection. Taken together, these data suggest luteolin inhibits LPS-induced IL-6 production in the brain by inhibiting the JNK signaling pathway and activation of AP-1 in microglia. Thus, luteolin may be useful for mitigating neuroinflammation (Jang et al., 2008).

Colon Cancer

Activities of CDK4 and CDK2 decreased within 2 hours after luteolin treatment, with a 38% decrease in CDK2 activity (P < 0.05) observed in cells treated with 40 µmol/l luteolin. Luteolin inhibited CDK2 activity in a cell-free system, suggesting that it directly inhibits CDK2.

tLuteolin promoted G2/M arrest at 24 hours post-treatment  by down-regulating cyclin B1 expression and inhibiting cell division cycle (CDC)2 activity. Luteolin promoted apoptosis with increased activation of caspases 3, 7, and 9 and enhanced poly(ADP-ribose) polymerase cleavage and decreased expression of p21CIP1/WAF1, survivin, Mcl-1, Bcl-xL, and Mdm-2. Decreased expression of these key antiapoptotic proteins could contribute to the increase in p53-independent apoptosis that was observed in HT-29 cells. Lim et al., (2007) demonstrated that luteolin promotes both cell-cycle arrest and apoptosis in the HT-29 colon cancer cell line, providing insight about the mechanisms underlying its anti-tumorigenic activities.

Pancreatic Cancer; Chemotherapy

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

Pre-treatment of cells with 11-19µM of either flavonoid for 24 hours resulted in 59-73% growth inhibition when followed by Gem (10µM, 36h). Lut (15µM, 24h) Pre-treatment followed by Gem (10µM, 36h), significantly decreased protein expression of nuclear GSK-3βand NF-κB p65 and increased pro-apoptotic cytosolic cytochrome c. Pre-treatment of human pancreatic cancer cells BxPC-3 with low concentrations of Lut effectively aid in the anti-proliferative activity of chemotherapeutic drugs (Johnson et al., 2013).

Ovarian Cancer

Luteolin has been found to repress NF-kappaB (NF-κ>B, a pro-inflammatory transcription factor) and inhibit pro-inflammatory cytokines such as TNF-αand IL-6. Additionally, it has been shown to stabilize p53 protein, sensitize TRAIL (TNF receptor apoptosis-inducing ligand) induced apoptosis, and prevent or delay chemotherapy-resistance.

Recent studies further indicate that luteolin potently inhibits VEGF production and suppresses 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. Unlike NSAIDS (non-steroid anti-inflammatory drugs), well documented clinical data for phyto-active compounds are lacking. In order to evaluate objectively the potential benefit of these compounds in the treatment of ovarian cancer, strategically designed, large scale studies are warranted (Chen et al., 2012).

Chemo-sensitizer

The sensitization effect of luteolin on cisplatin-induced apoptosis is p53 dependent, as such effect is only found in p53 wild-type cancer cells but not in p53 mutant cancer cells. Moreover, knockdown of p53 by small interfering RNA made p53 wild-type cancer cells resistant to luteolin and cisplatin. Second, Shi et al., (2007) observed a significant increase of p53 protein level in luteolin-treated cancer cells without increase of p53 mRNA level, indicating the possible effect of luteolin on p53 posttranscriptional regulation.

In summary, data from this study reveal a novel molecular mechanism involved in the anti-cancer effect of luteolin and support its potential clinical application as a chemo-sensitizer in cancer therapy.

Breast Cancer; TAM Chemo-sensitizer

This study found that the level of cyclin E2 (CCNE2) mRNA was higher in tumor cells (4.89-fold, (∗)P=0.005) than in normal paired tissue samples as assessed using real-time reverse-transcriptase polymerase chain reaction (RT-PCR) analysis (n=257). Further, relatively high levels of CCNE2 protein expression were detected in tamoxifen-resistant (TAM-R) MCF-7 cells.

These results showed that the level of CCNE2 protein expression was specifically inhibited in luteolin-treated (5µM) TAM-R cells, either in the presence or absence of 4-OH-TAM (100nM). Combined treatment with 4-OH-TAM and luteolin synergistically sensitized the TAM-R cells to 4-OH-TAM. The results of this study suggest that luteolin can be used as a chemo-sensitizer to target the expression level of CCNE2 and that it could be a novel strategy to overcome TAM resistance in breast cancer patients (Tu et al., 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.


Chen CY, Peng WH, Tsai KD and Hsu SL. (2007). Luteolin suppresses inflammation-associated gene expression by blocking NF-κB and AP-1 activation pathway in mouse alveolar macrophages. Life Sciences, 81(23-24):1602-1614. doi:10.1016/j.lfs.2007.09.028


Chen MZ, Jin WZ, Dai LM, Xu SY. (1986). Effect of luteolin on inflammation and immune function. Chinese Journal of Pharmacology and Toxicology, 1986-01.


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.


Jang S, Kelley KW, Johnson RW. (2008). Luteolin reduces IL-6 production in microglia by inhibiting JNK phosphorylation and activation of AP-1. PNAS, 105(21):7534-7539


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, S0278-6915(13)00491-2. doi: 10.1016/j.fct.2013.07.036.


Lim DY, Jeong Y, Tyner Al., Park JHY. (2007). Induction of cell-cycle arrest and apoptosis in HT-29 human colon cancer cells by the dietary compound luteolin. Am J Physiol Gastrointest Liver Physiol, 292: G66-G75. doi:10.1152/ajpgi.00248.2006.


Shi R, Huang Q, Zhu X, et al. (2007). Luteolin sensitizes the anti-cancer effect of cisplatin via c-Jun NH2-terminal kinase-mediated p53 phosphorylation and stabilization. Molecular Cancer Therapeutics, 6(4):1338-1347. doi: 10.1158/1535-7163.MCT-06-0638.


Tu SH, Ho CT, Liu MF, et al. (2013). Luteolin sensitizes drug-resistant human breast cancer cells to tamoxifen via the inhibition of cyclin E2 expression. Food Chem, 141(2):1553-61. doi: 10.1016/j.foodchem.2013.04.077.


Xagorari A, Papapetropoulos A, Mauromatis A, et al. (2001). Luteolin inhibits an endotoxin-stimulated phosphorylation cascade and pro-inflammatory cytokine production in macrophages. JPET, 296(1):181-187.

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.

Andrographolide

Cancer: Leukemia, colorectal, lung

Action: Immunomodulatory,anti-inflammatory,anti-metastatic

Andrographolide (Andro), a diterpenoid lactone isolated from a traditional herbal medicine Andrographis paniculata [(Burm. f.) Wall. Ex Nees], is known to possess multiple pharmacological activities. Andrographolide has been shown to exhibit antioxidative, anti-cancer, anti-inflammatory, anti-diabetes, and anti-aging properties (Trivedi et al., 2007; Chao et al., 2010).

Immunomodulatory Activity

The immunomodulatory activity of HN-02, an extract containing a mixture of andrographolides, was evaluated at 1.0, 1.5, and 2.5 mg/kg on different in vivo and in vitro experimental models. It was also found that HN-02 treatment stimulated phagocytosis in mice. A significant increase in total WBC count and relative weight of spleen and thymus was observed in mice during 30 days of treatment with HN-02.

The present experimental findings demonstrate that HN-02 has the ability to enhance immune function, possibly through modulation of immune responses altered during antigen interaction, and to reverse the immunosuppression induced by CYP (Naik, 2009).

The ethanol extract and purified diterpene andrographolides of Andrographis paniculata (Acanthaceae) induced significant stimulation of antibody and delayed type hypersensitivity (DTH) response to sheep red blood cells (SRBC) in mice. The plant preparations also stimulated non-specific immune response of the animals measured in terms of macrophage migration index (MMI) phagocytosis of Escherichia coli and proliferation of splenic lymphocytes. The stimulation of both antigen specific and non-specific immune response was, however, of lower order with andrographolide than with the ethanol extract, suggesting that substance(s) other than andrographolide present in the extract may also be contributing towards immunostimulation (Puri, 1993)

Anti-inflammatory and Leukemic Therapies

Andrographolide has been shown to attenuate MMP-9 expression, with its main mechanism likely involving the NF-κB signal pathway. These results provide new opportunities for the development of new anti-inflammatory and leukemic therapies. This activity was shown in a study in which andrographolide (1–50µM) exhibited concentration-dependent inhibition of MMP-9 activation, induced by either tumor necrosis factor-α (TNF-α), or lipopolysaccharide (LPS), in THP-1cells.

Anti-inflammatory

Lee et al (2012) found that andrographolide could significantly inhibit the degradation of inhibitor-κB-α (IκB-α) induced by TNF-α. They used electrophoretic mobility shift assay and reporter gene detection to show that andrographolide also markedly inhibited NF-signaling, anti-translocation and anti-activation. These results provide new opportunities for the development of new anti-inflammatory and leukemic therapies.

Lung Cancer Metastasis

Andrographolide is known to have the potential to be developed as a chemotherapeutic agent, in particular in the treatment of lung cancer. In order to understand the anti-cancer properties of andrographolide, its effect on migration and invasion in human lung cancer A549 cells was examined. The results of the wound-healing assay and the in vitro transwell assay revealed that andrographolide inhibited dose-dependently the migration and invasion of A549 cells under non-cytotoxic concentrations.

These results indicated that andrographolide exerted an inhibitory effect on the activity and the mRNA and protein levels of MMP-7, but not MMP-2 or MMP-9. The andrographolide-inhibited MMP-7 expression or activity appeared to occur via activator protein-1 (AP-1) because its DNA binding activity was suppressed by andrographolide. Additionally, the transfection of Akt over-expression vector (Akt1 cDNA) to A549 cells could result in an increase expression of MMP-7 concomitantly with a marked induction on cell invasion. These findings suggested that the inhibition on MMP-7 expression by andrographolide may be through suppression on PI3K/Akt/AP-1 signaling pathway, which in turn leads to the reduced invasiveness of the cancer cells (Lee, 2010).

Colorectal Cancer

Andrographolide has also been shown to have potent anti-cancer activity against human colorectal carcinoma Lovo cells by inhibiting cell-cycle progression. To further investigate the mechanism for the anti-cancer properties of andrographolide, it was used to examine the effect on migration and invasion of Lovo cells. The results of wound-healing assay and in vitro transwell assay revealed that andrographolide inhibited dose-dependently the migration and invasion of Lovo cells under non-cytotoxic concentrations.

The down-regulation of MMP-7 appeared to be via the inactivation of activator protein-1 (AP-1) since the treatment with andrographolide suppressed the nuclear protein level of AP-1, which was accompanied by a decrease in DNA-binding level of the factor. Taken together, these results indicate that andrographolide reduces the MMP-7-mediated cellular events in Lovo cells, and provide a new mechanism for its anti-cancer activity (Shi, 2009)

Anti-inflammatory, Induces Apoptosis

Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is an important member of the tumor necrosis factor subfamily with great potential in cancer therapy; additionally andrographolide is known to possess potent anti-inflammatory and anti-cancer activities which may be attributed to its action on TRAIL. It has been shown that pre-treatment with andrographolide significantly enhances TRAIL-induced apoptosis in various human cancer cell lines, including those TRAIL-resistant cells.

Pre-treatment with an anti-oxidant (N-acetylcysteine) or a c-Jun NH(2)-terminal kinase inhibitor (SP600125) effectively prevented andrographolide-induced p53 activation and DR4 up-regulation and eventually blocked the andrographolide-induced sensitization on TRAIL-induced apoptosis. Taken together, these results present a novel anti-cancer effect of andrographolide and support its potential application in cancer therapy to overcome TRAIL resistance (Zhou, 2008).

References

Chao HP, Kuo CD, Chiu JH, Fu SL. (2010). Andrographolide exhibits anti-invasive activity against colon cancer cells via inhibition of MMP2 activity. Planta Medica, 76(16):1827–1833. doi: 10.1055/s-0030-1250039.


Lee WR, Chung CL, Hsiao CJ, et al. (2012). Suppression of matrix metalloproteinase-9 expression by andrographolide in human monocytic THP-1 cells via inhibition of NF- κB activation. Phytomedicine, 19(3):270-277. doi: 10.1016/j.phymed.2011.11.012


Lee YC, Lin HH, Hsu CH, et al. (2010). Inhibitory effects of andrographolide on migration and invasion in human non-small-cell lung cancer A549 cells via down-regulation of PI3K/Akt signaling pathway. Eur J Pharmacol, 632(1-3):23-32. doi: 10.1016/j.ejphar.2010.01.009.


Naik SR, Hule A. (2009). Evaluation of Immunomodulatory Activity of an Extract of Andrographolides from Andographis paniculata. Planta Med, 75(8):785-91. doi: 10.1055/s-0029-1185398.


Puri A, Saxena R, Saxena RP, et al. (1993). Immunostimulant agents from Andrographis paniculata. J Nat Prod, 56(7):995-9.


Shi MD, Lin HH, Chiang TA, et al. (2009). Andrographolide could inhibit human colorectal carcinoma Lovo cells migration and invasion via down-regulation of MMP-7 expression. Chem Biol Interact, 180(3):344-52. doi: 10.1016/j.cbi.2009.04.011.


Trivedi NP, Rawal UM, Patel BP. (2007). Hepato-protective effect of andrographolide against hexachlorocyclohexane- induced oxidative injury. Integrative Cancer Therapies, 6(3):271–280. doi: 10.1177/1534735407305985.


Zhou J, Lu GD, Ong CS, Ong CN, Shen HM. (2008). Andrographolide sensitizes cancer cells to TRAIL-induced apoptosis via p53-mediated death receptor 4 up-regulation. Mol Cancer Ther, 7(7):2170-80. doi: 10.1158/1535-7163.MCT-08-0071.

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.

Curcumin

Cancer: Colorectal., prostate, pancreatic

Action: MDR, chemo-preventive activity, anti-inflammatory, attenuation of immune suppression

Chemo-preventive Activity

Curcumin is a naturally occurring, dietary polyphenolic phytochemical that is under preclinical trial evaluation for cancer-preventive drug development. It is derived from the rhizome of Curcuma longa L. and has both anti-oxidant and anti-inflammatory properties; it inhibits chemically-induced carcinogenesis in the skin, forestomach, and colon when it is administered during initiation and/or postinitiation stages. Chemo-preventive activity of curcumin is observed when it is administered prior to, during, and after carcinogen treatment as well as when it is given only during the promotion/progression phase (starting late in premalignant stage) of colon carcinogenesis (Kawamori et al., 1999)

Anti-inflammatory

With respect to inflammation, in vitro, it inhibits the activation of free radical-activated transcription factors, such as nuclear factor κB (NFκB) and AP-1, and reduces the production of pro-inflammatory cytokines such as tumor necrosis factor-α (TNFα), interleukin-1β (IL-1β), and interleukin-8 (Chan et al., 1998)

Prostate Cancer

In addition, NF-kappaB and AP-1 may play a role in the survival of prostate cancer cells, and curcumin may abrogate their survival mechanisms (Mukhopadhyay et al., 2001).

Pancreatic Cancer

In patients suffering from pancreatic cancer, orally-administered curcumin was found to be well-tolerated and despite limited absorption, had a reasonable impact on biological activity in some patients. This was attributed to its potent nuclear factor-kappaB (NF-kappaB) and tumor-inhibitory properties, against advanced pancreatic cancer (Dhillon et al., 2008)

MDR

Curcumin, the major component in Curcuma longa (Jianghuang), inhibited the transport activity of all three major ABC transporters, i.e. Pgp, MRP1 and ABCG2 (Ganta et al., 2009).

Curcumin reversed MDR of doxorubicin or daunorubicin in K562/DOX cell line and decreased Pgp expression in a time-dependent manner (Chang et al., 2006). Curcumin enhanced the sensitivity to vincristine by the inhibition of Pgp in SGC7901/VCR cell line (Tang et al., 2005). Moreover, curcumin was useful in reversing MDR associated with a decrease in bcl-2 and survivin expression but an increase in caspase-3 expression in COC1/DDP cell line (Ying et al., 2007).

The cytotoxicity of vincristine and paclitaxel were also partially restored by curcumin in resistant KBV20C cell line. Curcumin derivatives reversed MDR by inhibiting Pgp efflux (Um et al., 2008). A chlorine substituent at the meta-or para-position on benzamide improved MDR reversal [72]. Bisdemethoxycurcumin modified from curcumin resulted in greater inhibition of Pgp expression (Limtrakul et al., 2004).

Attenuation of Immune Suppression

Curcumin (a chalcone) exhibited toxicity to human neural stem cells (hNSCs). Although oridonin (a diterpene) showed a null toxicity toward hNSCs, it repressed the enzymatic function only marginally in contrast to its potent cytotoxicity in various cancer cell lines. While the mode of action of the enzyme-polyphenol complex awaits to be investigated, the sensitivity of enzyme inhibition was compared to the anti-proliferative activities toward three cancer cell lines.

The IC50s obtained from both sets of the experiments indicate that they are in the vicinity of micromolar concentration with the enzyme inhibition slightly more active.

These results suggest that attenuation of immune suppression via inhibition of IDO-1 enzyme activity may be one of the important mechanisms of polyphenols in chemoprevention or combinatorial cancer therapy (Chen et al., 2012).

Cancer Stem Cells

In cancers that appear to follow the stem cell model, pathways such as Wnt, Notch and Hedgehog may be targeted with natural compounds such as curcumin or drugs to reduce the risk of initiation of new tumors. Disease progression of established tumors could also potentially be inhibited by targeting the tumorigenic stem cells alone, rather than aiming to reduce overall tumor size.

Cancer treatments could be evaluated by assessing stem cell markers before and after treatment. Targeted stem cell specific treatment of cancers may not result in 'complete' or 'partial' responses radiologically, as stem cell targeting may not reduce the tumor bulk, but eliminate further tumorigenic potential. These changes are discussed using breast, pancreatic, and lung cancer as examples (Reddy et al., 2011).

Multiple Cancer Effects; Cell-signaling

Curcumin has been shown to interfere with multiple cell signaling pathways, including cell-cycle (cyclin D1 and cyclin E), apoptosis (activation of caspases and down-regulation of anti-apoptotic gene products), proliferation (HER-2, EGFR, and AP-1), survival (PI3K/AKT pathway), invasion (MMP-9 and adhesion molecules), angiogenesis (VEGF), metastasis (CXCR-4) and inflammation (NF- κB, TNF, IL-6, IL-1, COX-2, and 5-LOX).

The activity of curcumin reported against leukemia and lymphoma, gastrointestinal cancers, genitourinary cancers, breast cancer, ovarian cancer, head and neck squamous cell carcinoma, lung cancer, melanoma, neurological cancers, and sarcoma reflects its ability to affect multiple targets (Anand et al., 2008).

Anti-inflammatory; Cell-signaling

Curcumin, a liposoluble polyphenolic pigment isolated from the rhizomes of Curcuma longa L. (Zingiberaceae), is another potential candidate for new anti-cancer drug development. Curcumin has been reported to influence many cell-signaling pathways involved in tumor initiation and proliferation. Curcumin inhibits COX-2 activity, cyclin D1 and MMPs overexpresion, NF-kB, STAT and TNF-alpha signaling pathways and regulates the expression of p53 tumor suppressing gene.

Curcumin is well-tolerated but has a reduced systemic bioavailability. Polycurcumins (PCurc 8) and curcumin encapsulated in biodegradable polymeric nanoparticles showed higher bioavailability than curcumin together with a significant tumor growth inhibition in both in vitro and in vivo studies (Cretu et al., 2012). Curcumin also sensitizes tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis through reactive oxygen species-mediated up-regulation of death receptor 5 (DR5) (Jung et al., 2005).

Curcumin and bioavailability

Curcumin, a major constituent of the spice turmeric, suppresses expression of the enzyme cyclooxygenase 2 (Cox-2) and has cancer chemo-preventive properties in rodents. It possesses poor systemic availability. Marczylo et al. (2007) explored whether formulation with phosphatidylcholine increases the oral bioavailability or affects the metabolite profile of curcumin. Their results suggest that curcumin formulated with phosphatidylcholine furnishes higher systemic levels of parent agent than unformulated curcumin.

Curcuminoids are poorly water-soluble compounds and to overcome some of the drawbacks of curcuminoids, Aditya et al. (2012) explored the potential of liposomes for the intravenous delivery of curcuminoids. The curcuminoids-loaded liposomes were formulated from phosphatidylcholine (soy PC). Curcumin/curcuminoids were encapsulated in phosphatidylcholine vesicles with high yields. Vesicles in the size range around 200 nm were selected for stability and cell experiments. Liposomal curcumin were found to be twofold to sixfold more potent than corresponding curcuminoids. Moreover, the mixture of curcuminoids was found to be more potent than pure curcumin in regard to the anti-oxidant and anti-inflammatory activities (Basnet et al., 2012). Results suggest that the curcumin-phosphatidylcholine complex improves the survival rate by increasing the anti-oxidant activity (Inokuma et al., 2012). Recent clinical trials on the effectiveness of phosphatidylcholine formulated curcumin in treating eye diseases have also shown promising results, making curcumin a potent therapeutic drug candidate for inflammatory and degenerative retinal and eye diseases (Wang et al., 2012). Data demonstrate that treatment with curcumin dissolved in sesame oil or phosphatidylcholine curcumin improves the peripheral neuropathy of R98C mice by alleviating endoplasmic reticulum stress, by reducing the activation of unfolded protein response (Patzk- et al., 2012).

References

Aditya NP, Chimote G, Gunalan K, et al. (2012). Curcuminoids-loaded liposomes in combination with arteether protects against Plasmodium berghei infection in mice. Exp Parasitol, 131(3):292-9. doi: 10.1016/j.exppara.2012.04.010.


Anand P, Sundaram C, Jhurani S, Kunnumakkara AB, Aggarwal BB. (2008). Curcumin and cancer: An 'old-age' disease with an 'age-old' solution. Cancer Letters, 267(1):133–164. doi: 10.1016/j.canlet.2008.03.025.


Basnet P, Hussain H, Tho I, Skalko-Basnet N. (2012). Liposomal delivery system enhances anti-inflammatory properties of curcumin. J Pharm Sci, 101(2):598-609. doi: 10.1002/jps.22785.


Chan MY, Huang HI, Fenton MR, Fong D. (1998). In Vivo Inhibition of Nitric Oxide Synthase Gene Expression by Curcumin, a Cancer-preventive Natural Product with Anti-Inflammatory Properties. Biochemical Pharmacology, 55(12), 1955-1962.


Chang HY, Pan KL, Ma FC, et al. (2006). The study on reversing mechanism of Multi-drug resistance of K562/DOX cell line by curcumin and erythromycin. Chin J Hem, 27(4):254-258.


Chen SS, Corteling R, Stevanato L, Sinden J. (2012). Polyphenols Inhibit Indoleamine 3,5-Dioxygenase-1 Enzymatic Activity — A Role of Immunomodulation in Chemoprevention. Discovery Medicine.


Cre ţ u E, Trifan A, Vasincu A, Miron A. (2012). Plant-derived anti-cancer agents – curcumin in cancer prevention and treatment. Rev Med Chir Soc Med Nat Iasi, 116(4):1223-9.


Dhillon N, Aggarwal BB, Newman RA, et al. (2008). Phase II trial of curcumin in patients with advanced pancreatic cancer. Clin Cancer Res,14(14):4491-9. doi: 10.1158/1078-0432.CCR-08-0024.


Ganta S, Amiji M. (2009). Coadministration of paclitaxel and curcumin in nanoemulsion formulations To overcome Multi-drug resistance in tumor cells. Mol Pharm, 6(3):928-939. doi: 10.1021/mp800240j.


Inokuma T, Yamanouchi K, Tomonaga T, et al. (2012). Curcumin improves the survival rate after a massive hepatectomy in rats. Hepatogastroenterology, 59(119):2243-7. doi: 10.5754/hge10650.


Jung EM, Lim JH, Lee TJ, et al. (2005). Curcumin sensitizes tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis through reactive oxygen species-mediated up-regulation of death receptor 5 (DR5). Carcinogenesis, 26(11):1905-1913.


Kawamori T, Lubet R, Steele V E, et al. (1999). Chemo-preventive Effect of Curcumin, a Naturally Occurring Anti-Inflammatory Agent, during the Promotion/Progression Stages of Colon Cancer. Cancer Research, 59(3), 597-601.


Limtrakul P, Anuchapreeda S, Buddhasukh D. (2004). Modulation of human Multi-drug resistance MDR-1 gene by natural curcuminoids. BMC Cancer, 4:13.


Marczylo TH, Verschoyle RD, Cooke DN, et al. (2007). Comparison of systemic availability of curcumin with that of curcumin formulated with phosphatidylcholine. Cancer Chemother Pharmacol, 60(2):171-7.


Mukhopadhyay A, Bueso-Ramos C, Chatterjee D, Pantazis P, & Aggarwal., B. B. (2001). Curcumin downregulates cell survival mechanisms in human prostate cancer cell lines. Oncogene, 20(52), 7597-7609.


Patzk- A, Bai Y, Saporta MA, et al. (2012). Curcumin derivatives promote Schwann cell differentiation and improve neuropathy in R98C CMT1B mice. Brain, 135(Pt 12):3551-66. doi: 10.1093/brain/aws299.


Reddy RM, Kakarala M, Wicha MS. (2011). Clinical trial design for testing the stem cell model for the prevention and treatment of cancer. Cancers (Basel), 3(2):2696-708. doi: 10.3390/cancers3022696.


Tang XQ, Bi H, Feng JQ, Cao JG. (2005). Effect of curcumin on Multi-drug resistance in resistant human gastric carcinoma cell line SGC7901/VCR. Acta Pharmacol Sin, 26(8):1009-1016.


Um Y, Cho S, Woo HB, et al. (2008). Synthesis of curcumin mimics with Multi-drug resistance reversal activities. Bioorg Med Chem,16(7):3608-3615.


Wang LL, Sun Y, Huang K, Zheng L. (2012). Curcumin, a potential therapeutic candidate for retinal diseases. Mol Nutr Food Res, 57(9):1557-68. doi: 10.1002/mnfr.201200718.


Ying HC, Zhang SL, Lv J. (2007). Drug-resistant reversing effect of curcumin on COC1/DDP and its mechanism. J Mod Oncol, 15(5):604-607.