Category Archives: p53

Decursin

Cancer: Prostate, breast, fibrosarcoma, sarcoma

Action: MDR, inflammation, anti-cancer, angiogenesis

Decursin is isolated from Angelica gigas (Nakai).

Angelica gigas NAKAI is used to treat dysmenorrhea, amenorrhea, menopause, abdominal pain, injuries, migraine, and arthritis. The physicochemical and toxicological characterization of compounds in A. gigas NAKAI, decursin, decursinol angelate, diketone decursin, ether decursin, epoxide decursin and oxim decursin, have been extensively studied (Mahat et al., 2012).

Sarcoma; Anti-cancer

The in vivo anti-tumor activities of decursinol angelate (1) and decursin (2) isolated from the roots of Angelica gigas were investigated. These two compounds, when administered consecutively for 9 days at 50 and 100 mg/kg i.p. in mice, caused a significant increase in the life span and a significant decrease in the tumor weight and volume of mice inoculated with Sarcoma-180 tumor cells. These results suggest that decursinol angelate (1) and decursin (2) from A. gigas have anti-tumor activities (Lee et al., 2003).

Fibrosarcoma

Decursin and related coumarin compounds in herbal extracts have a number of biological activities against inflammation, angiogenesis and cancer. The human fibrosarcoma cell line, HT1080, was treated with TNFα (tumor necrosis factor α) in the presence or absence of CSL-32. Treatment of HT1080 cells with a derivative of decursin (CSL-32) inhibited their proliferation, without affecting cell viability, and TNF α-induced expression of pro-inflammatory mediators, such as MMP-9 (matrix metalloproteinase-9) and IL-8 (interleukin-8) (Lee et al., 2012).

Prostate Cancer

Androgen and androgen receptor (AR) signaling are crucial for the genesis of prostate cancer (PCa), which can often develop into androgen-ligand-independent diseases that are lethal to patients. As current chemotherapy is largely ineffective for PCa and has serious toxic side-effects, a collaborative effort has been initiated to identify and develop novel, safe and naturally occurring agents that target AR signaling from Oriental medicinal herbs for the chemoprevention and treatment of PCa. The discovery of decursin from an Oriental formula containing Korean Angelica gigas Nakai (Dang Gui) root as a novel anti-androgen/AR agent has been highlighted and the mechanisms to account for the specific anti-AR actions have been identified: rapid block of AR nuclear translocation, inhibition of binding of 5-dihydrotestesterone to AR, and increased proteasomal degradation of AR protein. Structure-activity analyzes reveal a critical requirement of the side-chain on decursin or its structural isomer decursinol angelate for anti-AR, cell-cycle arrest and pro-apoptotic activities.

This work demonstrates the feasibility of using activity-guided fractionation in cell culture assays combined with mechanistic studies to identify novel anti-androgen/AR agents from complex herbal mixtures (Lu et al., 2007).

MDR

Combination cancer therapy is one of the attractive approaches to overcome drug resistance of cancer cells. In the present study, Jang et al (2013) investigated the synergistic effect of decursin from Angelica gigas and doxorubicin on the induction of apoptosis in three human multiple myeloma cells. The combined treatment reduced mitochondrial membrane potential., suppressed the phosphorylation of JAK2, STAT3, and Src, activated SHP-2, and attenuated the expression of cyclind-D1 and survivin in U266 cells.

Overall, the combination treatment of decursin and doxorubicin can enhance apoptotic activity via mTOR and/or STAT3 signaling pathway in multiple myeloma cells.

Breast Cancer

Decursin significantly reduced protein expression and enzymatic activity of Pin1 in MDA-MB-231 cells. Kim et al (2013) found that decursin treatment enhanced the p53 expression level and failed to down-regulate Pin1 in the cells transfected with p53 siRNA, indicating the importance of p53 in the decursin-mediated Pin1 inhibition in MDA-MB-231 cells. Decursin stimulated association between peptidyl-prolyl cis/trans isomerase Pin1 to p53. Moreover, decursin facilitated p53 transcription in MDA-MB-231 cells. Overall, the study suggests the potential of decursin as an attractive cancer therapeutic agent for breast cancer by targeting Pin1.

References

Jang J, Jeong SJ, Kwon HY, Jung JH, et al. (2013). Decursin and Doxorubicin Are in Synergy for the Induction of Apoptosis via STAT3 and/or mTOR Pathways in Human Multiple Myeloma Cells. Evid Based Complement Alternat Med. 2013:506324. doi: 10.1155/2013/506324.

Kim JH, Jung JH, Kim SH, Jeong SJ. (2013). Decursin Exerts Anti-cancer Activity in MDA-MB-231 Breast Cancer Cells Via Inhibition of the Pin1 Activity and Enhancement of the Pin1/p53 Association.Phytother Res. doi: 10.1002/ptr.4986.

Lee S, Lee YS, Jung SH, et al. (2003). Anti-tumor activities of decursinol angelate and decursin from Angelica gigas. Arch Pharm Res, 26(9):727-30.

Lee SH, Lee JH, Kim EJ, et al. (2012). A novel derivative of decursin, CSL-32, blocks migration and production of inflammatory mediators and modulates PI3K and NF- κB activities in HT1080 cells. Cell Biol Int, 36(7):683-8. doi: 10.1042/CBI20110257.

Lu JX, Kim SH, Jiang C, Lee JJ, Guo JM. (2007). Oriental herbs as a source of novel anti-androgen and prostate cancer chemo-preventive agents. Acta Pharmacologica Sinica, 28, 1365–1372. doi:10.1111/j.1745-7254.2007.00683.x

Mahat B, Chae JW, Baek IH, et al. (2012). Physicochemical characterization and toxicity of decursin and their derivatives from Angelica gigas. Biol Pharm Bull, 35(7):1084-90.

Thymoquinone

Cancer: Osteosarcoma, pancreatic, colorectal., lung, liver, melanoma, breast

Action: Anti-inflammatory

For centuries, the black seed (Nigella sativa (L.)) herb and oil have been used in Asia, Middle East and Africa to promote health and fight disease. Thymoquinone (TQ) is the major phytochemical constituent of Nigella sativa (L.) oil extract. Phytochemical compounds are emerging as a new generation of anti-cancer agents with limited toxicity in cancer patients.

Osteosarcoma

The anti-proliferative and pro-apoptotic effects of TQ were evaluated in two human osteosarcoma cell lines with different p53 mutation status. TQ decreased cell survival dose-dependently and, more significantly, in p53-null MG63 cells (IC(50) = 17 muM) than in p53-mutant MNNG/HOS cells (IC(50) = 38 muM). Cell viability was reduced more selectively in MG63 tumor cells than in normal human osteoblasts.

It was therefore suggested that the resistance of MNNG/HOS cells to drug-induced apoptosis is caused by the up-regulation of p21(WAF1) by the mutant p53 (transcriptional activity was shown by p53 siRNA treatment) which induces cell-cycle arrest and allows repair of DNA damage.

Collectively, these findings show that TQ induces p53-independent apoptosis in human osteosarcoma cells. As the loss of p53 function is frequently observed in osteosarcoma patients, these data suggest the potential clinical usefulness of TQ for the treatment of these malignancies (Roepke et al., 2007).

Pancreatic Ductal Adenocarcinoma

Inflammation has been identified as a significant factor in the development of solid tumor malignancies. It has recently been shown that thymoquinone (Tq) induces apoptosis and inhibited proliferation in PDA cells. The effect of Tq on the expression of different pro-inflammatory cytokines and chemokines was analyzed by real-time polymerase chain reaction (PCR). Tq dose- and time-dependently significantly reduced PDA cell synthesis of MCP-1, TNF-alpha, interleukin (IL)-1beta and Cox-2. Tq also inhibited the constitutive and TNF-alpha-mediated activation of NF-kappaB in PDA cells and reduced the transport of NF-kappaB from the cytosol to the nucleus. Our data demonstrate previously undescribed anti-inflammatory activities of Tq in PDA cells, which are paralleled by inhibition of NF-kappaB. Tq as a novel inhibitor of pro-inflammatory pathways provides a promising strategy that combines anti-inflammatory and pro-apoptotic modes of action (Chehl et al., 2009).

Lung cancer, Hepatoma, Melanoma, Colon Cancer, Breast Cancer

The potential impact of thymoquinone (TQ) was investigated on the survival., invasion of cancer cells in vitro, and tumor growth in vivo. Exposure of cells derived from lung (LNM35), liver (HepG2), colon (HT29), melanoma (MDA-MB-435), and breast (MDA-MB-231 and MCF-7) tumors to increasing TQ concentrations resulted in a significant inhibition of viability through the inhibition of Akt phosphorylation leading to DNA damage and activation of the mitochondrial-signaling pro-apoptotic pathway. Administration of TQ (10 mg/kg/i.p.) for 18 days inhibited the LNM35 tumor growth by 39% (P < 0.05). Tumor growth inhibition was associated with significant increase in the activated caspase-3. In this context, it has been demonstrated that TQ treatment resulted in a significant inhibition of HDAC2 proteins. In view of the available experimental findings, it is contended that thymoquinone and/or its analogues may have clinical potential as an anti-cancer agent alone or in combination with chemotherapeutic drugs such as cisplatin (Attoub et al., 2012).

Colon Cancer

It was reported that TQ inhibits the growth of colon cancer cells which was correlated with G1 phase arrest of the cell-cycle. Furthermore, TUNEL staining and flow cytometry analysis indicate that TQ triggers apoptosis in a dose- and time-dependent manner. These results indicate that TQ is anti-neoplastic and pro-apoptotic against colon cancer cell line HCT116. The apoptotic effects of TQ are modulated by Bcl-2 protein and are linked to and dependent on p53. Our data support the potential for using the agent TQ for the treatment of colon cancer (Gali-Muhtasib et al., 2004).

References

Attoub S, Sperandio O, Raza H, et al. (2012). Thymoquinone as an anti-cancer agent: evidence from inhibition of cancer cells viability and invasion in vitro and tumor growth in vivo. Fundam Clin Pharmacol, 27(5):557-569. doi: 10.1111/j.1472-8206.2012.01056.x


Chehl N, Chipitsyna G, Gong Q, Yeo CJ, Arafat HA. (2009). Anti-inflammatory effects of the Nigella sativa seed extract, thymoquinone, in pancreatic cancer cells. HPB (Oxford), 11(5):373-81. doi: 10.1111/j.1477-2574.2009.00059.x.


Gali-Muhtasib H, Diab-Assaf M, Boltze C, et al. (2004). Thymoquinone extracted from black seed triggers apoptotic cell death in human colorectal cancer cells via a p53-dependent mechanism. Int J Oncol, 25(4):857-66


Roepke M, Diestel A, Bajbouj K, et al. (2007). Lack of p53 augments thymoquinone-induced apoptosis and caspase activation in human osteosarcoma cells. Cancer Biol Ther, 6(2):160-9.

Thymoquinone

Cancer: Osteosarcoma, pancreatic, colorectal., lung, liver, melanoma, breast

Action: Anti-inflammatory

For centuries, the black seed (Nigella sativa (L.)) herb and oil have been used in Asia, Middle East and Africa to promote health and fight disease. Thymoquinone (TQ) is the major phytochemical constituent of Nigella sativa (L.) oil extract. Phytochemical compounds are emerging as a new generation of anti-cancer agents with limited toxicity in cancer patients.

Osteosarcoma

The anti-proliferative and pro-apoptotic effects of TQ were evaluated in two human osteosarcoma cell lines with different p53 mutation status. TQ decreased cell survival dose-dependently and, more significantly, in p53-null MG63 cells (IC(50) = 17 muM) than in p53-mutant MNNG/HOS cells (IC(50) = 38 muM). Cell viability was reduced more selectively in MG63 tumor cells than in normal human osteoblasts.

It was therefore suggested that the resistance of MNNG/HOS cells to drug-induced apoptosis is caused by the up-regulation of p21(WAF1) by the mutant p53 (transcriptional activity was shown by p53 siRNA treatment) which induces cell-cycle arrest and allows repair of DNA damage.

Collectively, these findings show that TQ induces p53-independent apoptosis in human osteosarcoma cells. As the loss of p53 function is frequently observed in osteosarcoma patients, these data suggest the potential clinical usefulness of TQ for the treatment of these malignancies (Roepke et al., 2007).

Pancreatic Ductal Adenocarcinoma

Inflammation has been identified as a significant factor in the development of solid tumor malignancies. It has recently been shown that thymoquinone (Tq) induces apoptosis and inhibited proliferation in PDA cells. The effect of Tq on the expression of different pro-inflammatory cytokines and chemokines was analyzed by real-time polymerase chain reaction (PCR). Tq dose- and time-dependently significantly reduced PDA cell synthesis of MCP-1, TNF-alpha, interleukin (IL)-1beta and Cox-2. Tq also inhibited the constitutive and TNF-alpha-mediated activation of NF-kappaB in PDA cells and reduced the transport of NF-kappaB from the cytosol to the nucleus. Our data demonstrate previously undescribed anti-inflammatory activities of Tq in PDA cells, which are paralleled by inhibition of NF-kappaB. Tq as a novel inhibitor of pro-inflammatory pathways provides a promising strategy that combines anti-inflammatory and pro-apoptotic modes of action (Chehl et al., 2009).

Lung cancer, Hepatoma, Melanoma, Colon Cancer, Breast Cancer

The potential impact of thymoquinone (TQ) was investigated on the survival., invasion of cancer cells in vitro, and tumor growth in vivo. Exposure of cells derived from lung (LNM35), liver (HepG2), colon (HT29), melanoma (MDA-MB-435), and breast (MDA-MB-231 and MCF-7) tumors to increasing TQ concentrations resulted in a significant inhibition of viability through the inhibition of Akt phosphorylation leading to DNA damage and activation of the mitochondrial-signaling pro-apoptotic pathway. Administration of TQ (10 mg/kg/i.p.) for 18 days inhibited the LNM35 tumor growth by 39% (P < 0.05). Tumor growth inhibition was associated with significant increase in the activated caspase-3. In this context, it has been demonstrated that TQ treatment resulted in a significant inhibition of HDAC2 proteins. In view of the available experimental findings, it is contended that thymoquinone and/or its analogues may have clinical potential as an anti-cancer agent alone or in combination with chemotherapeutic drugs such as cisplatin (Attoub et al., 2012).

Colon Cancer

It was reported that TQ inhibits the growth of colon cancer cells which was correlated with G1 phase arrest of the cell-cycle. Furthermore, TUNEL staining and flow cytometry analysis indicate that TQ triggers apoptosis in a dose- and time-dependent manner. These results indicate that TQ is anti-neoplastic and pro-apoptotic against colon cancer cell line HCT116. The apoptotic effects of TQ are modulated by Bcl-2 protein and are linked to and dependent on p53. Our data support the potential for using the agent TQ for the treatment of colon cancer (Gali-Muhtasib et al., 2004).

References

Attoub S, Sperandio O, Raza H, et al. (2012). Thymoquinone as an anti-cancer agent: evidence from inhibition of cancer cells viability and invasion in vitro and tumor growth in vivo. Fundam Clin Pharmacol, 27(5):557-569. doi: 10.1111/j.1472-8206.2012.01056.x


Chehl N, Chipitsyna G, Gong Q, Yeo CJ, Arafat HA. (2009). Anti-inflammatory effects of the Nigella sativa seed extract, thymoquinone, in pancreatic cancer cells. HPB (Oxford), 11(5):373-81. doi: 10.1111/j.1477-2574.2009.00059.x.


Gali-Muhtasib H, Diab-Assaf M, Boltze C, et al. (2004). Thymoquinone extracted from black seed triggers apoptotic cell death in human colorectal cancer cells via a p53-dependent mechanism. Int J Oncol, 25(4):857-66


Roepke M, Diestel A, Bajbouj K, et al. (2007). Lack of p53 augments thymoquinone-induced apoptosis and caspase activation in human osteosarcoma cells. Cancer Biol Ther, 6(2):160-9.

Oridonin

Cancer: Prostate

Action: Growth arrest, autophagy

To investigate the mechanism of oridonin (ORI)-induced autophagy in prostate cancer PC-3 cells, PC-3 cells cultured in vitro were treated with ORI, and the inhibitory ratio of ORI on PC-3 cells was assayed by 3-4,5- dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide. After ORI treatment, the proliferation of PC-3 cells was inhibited significantly in a concentration and time-dependent manner. SEM examination revealed cellular shrinkage and disappearance of surface microvilli in ORI-treated cells. Under TEM examination, the nuclei exhibited chromatin condensation and the appearance of a large number of autophagosomes with double-membrane structure in cytoplasm. AO staining showed the existence of AVOs. The expression of LC3 and the mRNA level of beclin 1 was increased by ORI. Furthermore, autophagy inhibitor 3-methyladenine reversed the increase of beclin 1 mRNA. The growth of PC-3 cells was inhibited, and autophagy was induced by ORI, indicating ORI may have a potential antitumor effect.

Source
Ye LH, Li WJ, Jiang XQ, et al. Study on the autophagy of prostate cancer PC-3 cells induced by oridonin. Anat Rec (Hoboken). 2012 Mar;295(3):417-22. doi: 10.1002/ar.21528.

 

Cancer: Multiple myeloma

Action: Inhibits proliferation and induces apoptosis

This study was purposed to investigate the antitumor effect of oridonin on human multiple myeloma cell line U266 The results showed that the oridonin obviously inhibited the growth of U266 cell in dose-and time-dependent manners. As for morphological changes, characteristic apoptotic cells presented in U266 cells treated with 10 µmol/L oridonin for 24 hours. The apoptotic rate of U266 cells increased in dose and time dependent manners; after treatment of U266 cells with oridonin the mRNA levels of FGFR3, BCL2, CCND1 and MYC as well as the their protein levels decreased. Occasionally, the oridonin up-regulated the protein levels of P53 in the same manner. It is concluded that the oridonin can exert its anti-tumor effect by inhibiting proliferation and inducing apoptosis of U266 cell in dose dependent and time dependent manners, that maybe give the clues about new program of target therapy for multiple myeloma.

Source:

Duan HQ, Li MY, Gao L, et al. Mechanism concerning antitumor effect of oridonin on multiple myeloma cell line U266. Zhongguo Shi Yan Xue Ye Xue Za Zhi. 2014 Apr;22(2):364-9. doi: 10.7534/j.issn.1009-2137.2014.02.018.

Cancer: Multiple myeloma

Action: Induces apoptosis and autophagy

Exposure to oridonin (1-64 μmol/L) inhibited the proliferation of RPMI8266 cells in a concentration-dependent manner with an IC(50) value of 6.74 μmol/L. Exposure to oridonin (7 μmol/L) simultaneously induced caspase 3-mediated apoptosis and Beclin 1-dependent autophagy of RPMI8266 cells. Both the apoptosis and autophagy were time-dependent, and apoptosis was the main effector pathway of cell death. Exposure to oridonin (7 μmol/L) increased intracellular ROS and reduced SIRT1 nuclear protein in a time-dependent manner.

Oridonin simultaneously induces apoptosis and autophagy of human multiple myeloma RPMI8266 cells via regulation of intracellular ROS generation and SIRT1 nuclear protein. The cytotoxicity of oridonin is mainly mediated through the apoptotic pathway, whereas the autophagy protects the cells from apoptosis.

Source

Zeng R, Chen Y, Zhao S, Cui GH.Autophagy counteracts apoptosis in human multiple myeloma cells exposed to oridonin in vitro via regulating intracellular ROS and SIRT1. Acta Pharmacol Sin. 2012 Jan;33(1):91-100. doi: 10.1038/aps.2011.143.

Cancer: Prostate, acute promyelocytic leukemia, breast, non-small-cell lung (NSCL), Ehrlich ascites, P388 lymphocytic leukemia, colorectal., ovarian, esphageal

Action: Chemoresistance, Ara-C, VP-16 

Cancer cell arises in part through the acquisition of apoptotic resistance. Leukemia cells resistant to chemotherapy-induced apoptosis have been found to be sensitive to oridonin, a natural agent with potent anticancer activity. Weng et al., (2014) compared the response of human leukemia cells with oridonin and the antileukemia drugs Ara-C and VP-16. Compared with HL60 cells, K562 and K562/ADR cells displayed resistance to apoptosis stimulated by Ara-C and VP-16 but sensitivity to oridonin. Mechanistic investigations revealed that oridonin upregulated BIM-S by diminishing the expression of miR-17 and miR-20a, leading to mitochondria-dependent apoptosis. In contrast, neither Ara-C nor VP-16 could reduce miR-17 and miR-20a expression or could trigger BIM-S–mediated apoptosis.

Notably, silencing miR-17 or miR-20a expression by treatment with microRNA (miRNA; miR) inhibitors or oridonin restored sensitivity of K562 cells to VP-16. Synergistic effects of oridonin and VP-16 were documented in cultured cells as well as mouse tumor xenograft assays. Inhibiting miR-17 or miR-20a also augmented the proapoptotic activity of oridonin. Taken together, our results identify a miRNA-dependent mechanism underlying the anticancer effect of oridonin and provide a rationale for its combination with chemotherapy drugs in addressing chemoresistant leukemia cells.

Reference

Weng Hy, Huang Hl, Dong B, et al. Inhibition of miR-17 and miR-20a by Oridonin Triggers Apoptosis and Reverses Chemoresistance by Derepressing BIM-S. Cancer Res; 74(16); 1–11. doi: 10.1158/0008-5472.CAN-13-1748

Action: Induces apoptosis

Oridonin is a tetracycline diterpenoid isolated from the plant Rabdosia rubescens (RR) [(Hemsl.). Hara (Lamiaceae)] (dong ling cao) is a Chinese medicinal herb used widely in provinces including Henan. The aerial parts of RR and other species of the same genus has been reported to have the functions of clearing “heat” and “toxicity”, nourishing “yin”, removing “blood stasis”, and relieving swelling. RR has been used to treat stomach-ache, sore throat and cough.

Gastric Cancer, Esophageal Cancer, Liver Cancer, Prostate Cancer

RR and its extracts have been shown to be able to suppress disease progress, reduce tumor burden, alleviate syndrome and prolong survival in patients with gastric carcinoma, esophageal., liver and prostate cancers (Tang & Eisenbrand, 1992). Interestingly, other Isodon plants including Isodon japonicus Hara (IJ) and I. trichocarpus (IT) are also applied as home remedies for similar disorders in Japan and Korea.

Induces Apoptosis

These reports suggest that Isodon plants should have at least one essential anti-tumor component. In the 1970s, a bitter tetracycline diterpenoid compound, oridonin, was isolated from RR, IJ, and IT separately, and was shown to be a potent apoptosis inducer in a variety of cancer cells (Fujita et al., 1970; Fujita et al., 1976; Henan Medical Institute, 1978; Fujita et al., 1988).

Anti-cancer

There is currently research being undertaken regarding the relationship between the chemical structure/modifications and the molecular mechanisms underlying its anti-cancer activity, such as suppression of tumor proliferation and induction of tumor cell death, and the cell signal transduction in anti-cancer activity of oridonin (Zhang et al., 2010).

Prostate Cancer, Breast Cancer, NSCLC, Leukemia, Glioblastoma

Oridonin has been found to effectively inhibit the proliferation of a wide variety of cancer cells including those from prostate (LNCaP, DU145, PC3), breast (MCF-7, MDA-MB231), non-small-cell lung (NSCL) (NCI-H520, NCI-H460, NCI-H1299) cancers, acute promyelocytic leukemia (NB4), and glioblastoma multiforme (U118, U138).

Oridonin induced apoptosis and G0/G1 cell-cycle arrest in LNCaP prostate cancer cells. In addition, expression of p21waf1 was induced in a p53-dependent manner. Taken together, oridonin inhibited the proliferation of cancer cells via apoptosis and cell-cycle arrest with p53 playing a central role in several cancer types which express the wild-type p53 gene. Oridonin may be a novel, adjunctive therapy for a large variety of malignancies (Ikezoe et al., 2003).

Breast Cancer; Anti-metastatic

According to the flow cytometric analysis, oridonin suppressed MCF-7 cell growth by cell-cycle arrest at the G2/M phase and caused accumulation of MDA-MB-231 cells in the Sub-G1 phase. The induced apoptotic effect of oridonin was further confirmed by a morphologic characteristics assay and TUNEL assay. Meanwhile, oridonin significantly suppressed MDA-MB-231 cell migration and invasion, decreased MMP-2/MMP-9 activation and inhibited the expression of Integrin β1 and FAK. In conclusion, oridonin inhibited growth and induced apoptosis in breast cancer cells, which might be related to DNA damage and activation of intrinsic or extrinsic apoptotic pathways. Moreover, oridonin also inhibited tumor invasion and metastasis in vitro possibly via decreasing the expression of MMPs and regulating the Integrin β1/FAK pathway in MDA-MB-231 cells (Wang et al., 2013).

Gastric Cancer

The inhibitory effect of oridonin on gastric cancer HGC-27 cells was detected using the 3-(4, 5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide assay. After treated with oridonin (0, 1.25, 2.5, 5 and 10 µg/mL), HGC-27 cells were collected for anexin V-phycoerythrin and 7-amino-actinomycin D double staining and tested by flow cytometric analysis, and oridonin- induced apoptosis in HGC-27 cells was detected.

Oridonin significantly inhibited the proliferation of HGC-27 cells in a dose- and time-dependent manner. The inhibition rates of HGC-27 treated with four different concentrations of oridonin for 24 h (1.25, 2.5, 5 and 10 µg/mL) were 1.78% ± 0.36%, 4.96% ± 1.59%, 10.35% ± 2.76% and 41.6% ± 4.29%, respectively, which showed a significant difference (P < 0.05. Cells treated with oridonin showed typical apoptotic features with acridine orange/ethidium bromide staining. After treatment with oridonin, the cells became round, shrank, and developed small buds around the nuclear membrane while forming apoptotic bodies. However, the change in the release of LDH caused by necrosis was insignificant, suggesting that the major cause of oridonin-induced HGC-27 cell death was apoptosis. Flow cytometric analysis also revealed that oridonin induced significant apoptosis compared with the controls (P < 0.05).

Apoptosis of HGC-27 induced by oridonin may be associated with differential expression of Apaf-1, caspase-3 and cytochrome c, which are highly dependent upon the mitochondrial pathway (Sun et al., 2012).

Ehrlich Ascites, Leukemia

Oridonin has been found to also increase lifespan of mice bearing Ehrlich ascites or P388 lymphocytic leukemia. Oridonin triggered apoptosis in more than 50% of t(8;21) leukemic cells in vitro at concentration of 2 M or higher accompanied by degradation of AE oncoprotein, and showed significant anti-leukemia efficacies with low adverse effects in vivo. These data suggest possible beneficial effects for patients with t(8;21) acute myeloid leukemia (AML) (Zhou et al., 2007).

Prostate Cancer, Breast Cancer, Ovarian Cancer

Oridonin exhibited anti-proliferative activity toward all cancer cell lines tested, with an IC50 estimated by the MTT cell viability assay ranging from 5.8+/-2.3 to 11.72+/-4.8 microM. The increased incidence of apoptosis, identified by characteristic changes in cell morphology, was seen in tumor lines treated with oridonin. Notably, at concentrations that induced apoptosis among tumor cells, oridonin failed to induce apoptosis in cultures of normal human fibroblasts. Oridonin up-regulated p53 and Bax and down-regulated Bcl-2 expression in a dose-dependent manner and its absorption spectrum was measured in the presence and absence of double stranded (ds) DNA. Oridonin inhibits cancer cell growth in a cell-cycle specific manner and shifts the balance between pro- and anti-apoptotic proteins in favor of apoptosis. The present data suggest that further studies are warranted to assess the potential of oridonin in cancer prevention and/or treatment (Chen et al., 2005).

Ovarian Cancer Stem Cells; Chemotherapy Resistance

Oridonin was 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 types of compounds in the treatment of ovarian cancer, strategically designed, large scale studies are warranted (Chen et al., 2012).

Colorectal Cancer

Oridonin induced potent growth inhibition, cell-cycle arrest, apoptosis, senescence and colony-forming inhibition in three colorectal cancer cell lines in a dose-dependent manner in vitro. Daily i.p. injection of oridonin (6.25, 12.5 or 25 mg/kg) for 28 days significantly inhibited the growth of SW1116 s.c. xenografts in BABL/C nude mice.

Oridonin possesses potent in vitro and in vivo anti-colorectal cancer activities that correlated with induction of histone hyperacetylation and regulation of pathways critical for maintaining growth inhibition and cell-cycle arrest. Therefore, oridonin may represent a novel therapeutic option in colorectal cancer treatment as it has been shown to induce apoptosis and senescence of colon cancer cells in vitro and in vivo (Gao et al., 2010).

Colon Cancer; Apoptosis

Oridonin increased intracellular hydrogen peroxide levels and reduced the glutathione content in a dose-dependent manner. N-acetylcysteine, a reactive oxygen species scavenger, not only blocked the oridonin-induced increase in hydrogen peroxide and glutathione depletion, but also blocked apoptosis and senescence induced by oridonin.

Moreover, exogenous catalase could inhibit the increase in hydrogen peroxide and apoptosis induced by oridonin, but not the glutathione depletion and senescence. Furthermore, thioredoxin reductase (TrxR) activity was reduced by oridonin in vitro and in cells, which may cause the increase in hydrogen peroxide. In conclusion, the increase in hydrogen peroxide and glutathione depletion account for oridonin-induced apoptosis and senescence in colorectal cancer cells, and TrxR inhibition is involved in this process.

Given the importance of TrxR as a novel cancer target in colon cancer, oridonin would be a promising clinical candidate (Gao et al., 2012).

Prostate Cancer; Apoptosis

Oridonin (ORI) could inhibit the proliferation and induce apoptosis in various cancer cell lines. After ORI treatment, the proliferations of human prostate cancer (HPC) cell lines PC-3 and LNCaP were inhibited in a concentration and time-dependent manner. ORI induced cell-cycle arrest at the G2/M phase. Autophagy occurred before the onset of apoptosis and protected cancer cells in ORI-treated HPC cells. P21 was involved in ORI-induced autophagy and apoptosis (Li et al., 2012).

References

Chen S, Gao J, Halicka HD, et al. (2005). The cytostatic and cytotoxic effects of oridonin (Rubescenin), a diterpenoid from Rabdosia rubescens, on tumor cells of different lineage. Int J Oncol, 26(3):579-88.

 

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.

 

Fujita E, Fujita T, Katayama H, Shibuya M. (1970). Terpenoids. Part XV. Structure and absolute configuration of oridonin isolated from Isodon japonicus trichocarpus. J Chem Soc (Chem Comm), 21:1674–1681

 

Fujita E, Nagao Y, Node M, et al. (1976). Anti-tumor activity of the Isodon diterpenoids: structural requirements for the activity. Experientia, 32:203–206.

 

Fujita T, Takeda Y, Sun HD, et al. (1988). Cytotoxic and anti-tumor activities of Rabdosia diterpenoids. Planta Med, 54:414–417.

 

Henan Medical Institute, Henan Medical College, Yunnan Institute of Botany. (1978). Oridonin–a new anti-tumor subject. Chin Science Bull, 23:53–56.

 

Ikezoe T, Chen SS, Tong XJ, et al. (2003). Oridonin induces growth inhibition and apoptosis of a variety of human cancer cells. Int J Oncol, 23(4):1187-93.

 

Gao FH, Hu XH, Li W, Liu H, et al. (2010). Oridonin induces apoptosis and senescence in colorectal cancer cells by increasing histone hyperacetylation and regulation of p16, p21, p27 and c-myc. BMC Cancer, 10:610. doi: 10.1186/1471-2407-10-610.

 

Gao FH, Liu F, Wei W, et al. (2012). Oridonin induces apoptosis and senescence by increasing hydrogen peroxide and glutathione depletion in colorectal cancer cells. Int J Mol Med, 29(4):649-55. doi: 10.3892/ijmm.2012.895.

 

Li X, Li X, Wang J, Ye Z, Li JC. (2012) Oridonin up-regulates expression of P21 and induces autophagy and apoptosis in human prostate cancer cells. Int J Biol Sci. 2012;8(6):901-12. doi: 10.7150/ijbs.4554.

 

Sun KW, Ma YY, Guan TP, et al. (2012). Oridonin induces apoptosis in gastric cancer through Apaf-1, cytochrome c and caspase-3 signaling pathway. World J Gastroenterol, 18(48):7166-74. doi: 10.3748/wjg.v18.i48.7166.

 

Tang W, Eisenbrand G. (1992). Chinese drugs of plant origin: chemistry, pharmacology, and use in traditional and modern medicine. Berlin: Springer-Verlag, 817–847.

 

Wang S, Zhong Z, Wan J, et al. (2013). Oridonin induces apoptosis, inhibits migration and invasion on highly-metastatic human breast cancer cells. Am J Chin Med, 41(1):177-96. doi: 10.1142/S0192415X13500134.

 

Zhang Wj, Huang Ql, Hua Z-C. (2010). Oridonin: A promising anti-cancer drug from China. Frontiers in Biology, 5(6):540-545.

 

Zhou G-B, Kang H, Wang L, et al. (2007). Oridonin, a diterpenoid extracted from medicinal herbs, targets AML1-ETO fusion protein and shows potent anti-tumor activity with low adverse effects on t(8;21) leukemia in vitro and in vivo. Blood, 109(8):3441-3450.

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.

Decursin

Cancer: Prostate, breast, fibrosarcoma, sarcoma

Action: MDR, inflammation, anti-cancer, angiogenesis

Decursin is isolated from Angelica gigas (Nakai).

Angelica gigas NAKAI is used to treat dysmenorrhea, amenorrhea, menopause, abdominal pain, injuries, migraine, and arthritis. The physicochemical and toxicological characterization of compounds in A. gigas NAKAI, decursin, decursinol angelate, diketone decursin, ether decursin, epoxide decursin and oxim decursin, have been extensively studied (Mahat et al., 2012).

Sarcoma; Anti-cancer

The in vivo anti-tumor activities of decursinol angelate (1) and decursin (2) isolated from the roots of Angelica gigas were investigated. These two compounds, when administered consecutively for 9 days at 50 and 100 mg/kg i.p. in mice, caused a significant increase in the life span and a significant decrease in the tumor weight and volume of mice inoculated with Sarcoma-180 tumor cells. These results suggest that decursinol angelate (1) and decursin (2) from A. gigas have anti-tumor activities (Lee et al., 2003).

Fibrosarcoma

Decursin and related coumarin compounds in herbal extracts have a number of biological activities against inflammation, angiogenesis and cancer. The human fibrosarcoma cell line, HT1080, was treated with TNFα (tumor necrosis factor α) in the presence or absence of CSL-32. Treatment of HT1080 cells with a derivative of decursin (CSL-32) inhibited their proliferation, without affecting cell viability, and TNF α-induced expression of pro-inflammatory mediators, such as MMP-9 (matrix metalloproteinase-9) and IL-8 (interleukin-8) (Lee et al., 2012).

Prostate Cancer

Androgen and androgen receptor (AR) signaling are crucial for the genesis of prostate cancer (PCa), which can often develop into androgen-ligand-independent diseases that are lethal to patients. As current chemotherapy is largely ineffective for PCa and has serious toxic side-effects, a collaborative effort has been initiated to identify and develop novel, safe and naturally occurring agents that target AR signaling from Oriental medicinal herbs for the chemoprevention and treatment of PCa. The discovery of decursin from an Oriental formula containing Korean Angelica gigas Nakai (Dang Gui) root as a novel anti-androgen/AR agent has been highlighted and the mechanisms to account for the specific anti-AR actions have been identified: rapid block of AR nuclear translocation, inhibition of binding of 5-dihydrotestesterone to AR, and increased proteasomal degradation of AR protein. Structure-activity analyzes reveal a critical requirement of the side-chain on decursin or its structural isomer decursinol angelate for anti-AR, cell-cycle arrest and pro-apoptotic activities.

This work demonstrates the feasibility of using activity-guided fractionation in cell culture assays combined with mechanistic studies to identify novel anti-androgen/AR agents from complex herbal mixtures (Lu et al., 2007).

MDR

Combination cancer therapy is one of the attractive approaches to overcome drug resistance of cancer cells. In the present study, Jang et al (2013) investigated the synergistic effect of decursin from Angelica gigas and doxorubicin on the induction of apoptosis in three human multiple myeloma cells. The combined treatment reduced mitochondrial membrane potential., suppressed the phosphorylation of JAK2, STAT3, and Src, activated SHP-2, and attenuated the expression of cyclind-D1 and survivin in U266 cells.

Overall, the combination treatment of decursin and doxorubicin can enhance apoptotic activity via mTOR and/or STAT3 signaling pathway in multiple myeloma cells.

Breast Cancer

Decursin significantly reduced protein expression and enzymatic activity of Pin1 in MDA-MB-231 cells. Kim et al (2013) found that decursin treatment enhanced the p53 expression level and failed to down-regulate Pin1 in the cells transfected with p53 siRNA, indicating the importance of p53 in the decursin-mediated Pin1 inhibition in MDA-MB-231 cells. Decursin stimulated association between peptidyl-prolyl cis/trans isomerase Pin1 to p53. Moreover, decursin facilitated p53 transcription in MDA-MB-231 cells. Overall, the study suggests the potential of decursin as an attractive cancer therapeutic agent for breast cancer by targeting Pin1.

References

Jang J, Jeong SJ, Kwon HY, Jung JH, et al. (2013). Decursin and Doxorubicin Are in Synergy for the Induction of Apoptosis via STAT3 and/or mTOR Pathways in Human Multiple Myeloma Cells. Evid Based Complement Alternat Med. 2013:506324. doi: 10.1155/2013/506324.

Kim JH, Jung JH, Kim SH, Jeong SJ. (2013). Decursin Exerts Anti-cancer Activity in MDA-MB-231 Breast Cancer Cells Via Inhibition of the Pin1 Activity and Enhancement of the Pin1/p53 Association.Phytother Res. doi: 10.1002/ptr.4986.

Lee S, Lee YS, Jung SH, et al. (2003). Anti-tumor activities of decursinol angelate and decursin from Angelica gigas. Arch Pharm Res, 26(9):727-30.

Lee SH, Lee JH, Kim EJ, et al. (2012). A novel derivative of decursin, CSL-32, blocks migration and production of inflammatory mediators and modulates PI3K and NF- κB activities in HT1080 cells. Cell Biol Int, 36(7):683-8. doi: 10.1042/CBI20110257.

Lu JX, Kim SH, Jiang C, Lee JJ, Guo JM. (2007). Oriental herbs as a source of novel anti-androgen and prostate cancer chemo-preventive agents. Acta Pharmacologica Sinica, 28, 1365–1372. doi:10.1111/j.1745-7254.2007.00683.x

Mahat B, Chae JW, Baek IH, et al. (2012). Physicochemical characterization and toxicity of decursin and their derivatives from Angelica gigas. Biol Pharm Bull, 35(7):1084-90.

Camptothecin (CPT) & 10-hydroxycamptothecin (HCPT)

Cancer: Breast, colon

Action: Cytostatic

Breast Cancer

Recently, natural product DNA topoisomerase I inhibitors 10-hydroxycamptothecin (HCPT) and camptothecin (CPT) have been shown to have therapeutic effects in both in vitro and in vivo models of human breast cancer. After evaluation, the apoptotic pathways were characterized in vitro and in vivo in the human breast cancer cell lines MCF-7 and MDA-MB-468.

The elevation of p53 protein levels in MCF-7 cells treated with CPT was significantly inhibited by preincubation with DNA breaks inhibitor aphidicolin, while the elevation of p21WAF1/CIP1 protein levels was not inhibited. The elevation of p21WAF1/CIP1 in MDA-MB-468 cells treated with CPT was not inhibited by aphidicolin.

Results suggest that treatment with HCPT and CPT results in increased levels of p21WAF1/CIP1 protein and mRNA, and that they induce apoptosis in human breast cancer cells through both p53-dependent and -independent pathways. Findings may be significant in further understanding the mechanisms of actions of camptothecins in the treatment of human cancers (Liu & Zhang, 1998).

Colon Cancer

10-HCPT significantly repressed the proliferation of Colo 205 cells at a relatively low concentration (5-20 nM). Flow cytometry analysis and western blot and apoptosis assays demonstrated that low-dose 10-HCPT arrested Colo 205 cells in the G2 phase of the cell-cycle and triggered apoptosis through a caspase-3-dependent pathway.

Moreover, following oral administration at doses of 2.5-7.5 mg/kg/2 days, significant suppression of tumor growth by 10-HCPT was observed in mouse xenografts. No acute toxicity was observed after an oral challenge of 10-HCPT in BALB/c-nude mice every 2 days.

Results suggest that a relatively low dose of 10-HCPT (p.o.) is able to inhibit the growth of colon cancer, facilitating the development of a new protocol of human trials with this anti-cancer drug (Ping et al., 2006).

References

Liu, W., & Zhang, R. (1998). Up-regulation of p21WAF1/CIP1 in human breast cancer cell lines MCF-7 and MDA-MB-468 undergoing apoptosis induced by natural product anti-cancer drugs 10-hydroxycamptothecin and camptothecin through p53-dependent and independent pathways. International Journal of Oncology, 12(4), 793-804.


Ping, Y.H., Lee, H.C., Lee, J.Y., et al. (2006). Anti-cancer effects of low-dose 10-hydroxycamptothecin in human colon cancer. Oncology Reports, 15(5), 1273-9.

Artesunate

Cancer: Colon, esophageal., pancreatic, ovarian, multiple myeloma and diffuse large B-cell lymphoma, osteosarcoma, lung, breast, skin, leukemia/lymphoma

Action: Anti-metastatic, MDR, radio-sensitizer

Pulmonary Adenocarcinomas

Artesunate exerts anti-proliferative effects in pulmonary adenocarcinomas. It mediates these anti-neoplastic effects by virtue of activating Bak (Zhou et al., 2012). At the same time, it down-regulates epidermal growth factor receptor expression. This results in augmented non-caspase dependent apoptosis in the adenocarcinoma cells. Artesunate mediated apoptosis is time as well as dose-dependent. Interestingly, AIF and Bim play significant roles in this Bak-dependent accentuated apoptosis (Ma et al., 2011). Adenosine triphosphate (ATP)-binding cassette subfamily G member 2 (ABCG2) expression is also attenuated while transcription of matrix metallopeptidase 7 (MMP-7) is also down-regulated (Zhao et al., 2011). In addition, arsenuate enhances the radio-sensitization of lung carcinoma cells. It mediates this effect by down-regulating cyclin B1 expression, resulting in augmented G2/M phase arrest (Rasheed et al., 2010).

Breast Cancer

Similarly, artesunate exhibits anti-neoplastic effects in breast carcinomas. Artesunate administration is typically accompanied by attenuated turnover as well as accentuated peri-nuclear localization of autophagosomes in the breast carcinoma cells. Mitochondrial outer membrane permeability is typically augmented. As a result, artesunate augments programmed cellular decline in breast carcinoma cells (Hamacher-Brady et al., 2011).

Skin Cancer

Artesunate also exerts anti-neoplastic effects in skin malignancies. It mediates these effects by up-regulating p21. At the same time it down-regulates cyclin D1 (Jiang et al., 2012).

Colon Cancer

Artemisunate significantly inhibited both the invasiveness and anchorage independence of colon cancer SW620 cells in a dose-dependent manner. The protein level of intercellular adhesion molecule 1 (ICAM-1) was down-regulated as relative to the control group.

Artemisunate could potentially inhibit invasion of the colon carcinoma cell line SW620 by down-regulating ICAM-1 expression (Fan, Zhang, Yao & Li, 2008).

Multi-drug resistance; Colon Cancer

A profound cytotoxic action of the antimalarial., artesunate (ART), was identified against 55 cancer cell lines of the U.S. National Cancer Institute (NCI). The 50% inhibition concentrations (IC50 values) for ART correlated significantly to the cell doubling times (P = 0.00132) and the portion of cells in the G0/G1 (P = 0.02244) or S cell-cycle phases (P = 0.03567).

Efferth et al., (2003) selected mRNA expression data of 465 genes obtained by microarray hybridization from the NCI data-base. These genes belong to different biological categories (drug resistance genes, DNA damage response and repair genes, oncogenes and tumor suppressor genes, apoptosis-regulating genes, proliferation-associated genes, and cytokines and cytokine-associated genes). The constitutive expression of 54 of 465 (=12%) genes correlated significantly to the IC50 values for ART. Hierarchical cluster analysis of these 12 genes allowed the differentiation of clusters with ART-sensitive or ART-resistant cell lines (P = 0.00017).

Multi-drug-resistant cells differentially expressing the MDR1, MRP1, or BCRP genes were not cross-resistant to ART. ART acts via p53-dependent and- independent pathways in isogenic p53+/+ p21WAF1/CIP1+/+, p53-/- p21WAF1/CIP1+/+, and p53+/+ p21WAF1/CIP1-/- colon carcinoma cells.

Multi-drug resistance; Esophageal Cancer

The present study aimed to investigate the correlation between ABCG2 expression and the MDR of esophageal cancer and to estimate the therapeutic benefit of down-regulating ABCG2 expression and reversing chemoresistance in esophageal cells using artesunate (ART).

ART is a noteworthy antimalarial agent, particularly in severe and drug-resistant cancer cases, as ART is able to reverse drug resistance. ART exerted profound anti-cancer activity. The mechanism for the reversal of multi-drug resistance by ART in esophageal carcinoma was analyzed using cellular experiments, but still remains largely unknown (Liu, Zuo, & Guo, 2013).

Pancreatic Cancer

The combination of triptolide and artesunate could inhibit pancreatic cancer cell line growth, and induce apoptosis, accompanied by expression of HSP 20 and HSP 27, indicating important roles in the synergic effects. Moreover, tumor growth was decreased with triptolide and artesunate synergy. Results indicated that triptolide and artesunate in combination at low concentrations can exert synergistic anti-tumor effects in pancreatic cancer cells with potential clinical applications (Liu & Cui, 2013).

Ovarian Cancer

Advanced-stage ovarian cancer (OVCA) has a unifocal origin in the pelvis. Molecular pathways associated with extrapelvic OVCA spread are also associated with metastasis from other human cancers and with overall patient survival. Such pathways represent appealing therapeutic targets for patients with metastatic disease.

Pelvic and extrapelvic OVCA implants demonstrated similar patterns of signaling pathway expression and identical p53 mutations.

However, Marchion et al. (2013) identified 3 molecular pathways/cellular processes that were differentially expressed between pelvic and extrapelvic OVCA samples and between primary/early-stage and metastatic/advanced or recurrent ovarian, oral., and prostate cancers. Furthermore, their expression was associated with overall survival from ovarian cancer (P = .006), colon cancer (1 pathway at P = .005), and leukemia (P = .05). Artesunate-induced TGF-WNT pathway inhibition impaired OVCA cell migration.

Multiple Myeloma, B-cell Lymphoma

Findings indicate that artesunate is a potential drug for treatment of multiple myeloma and diffuse large B-cell lymphoma (DLBCL) at doses of the same order as currently in use for treatment of malaria without serious adverse effects. Artesunate treatment efficiently inhibited cell growth and induced apoptosis in cell lines. Apoptosis was induced concomitantly with down-regulation of MYC and anti-apoptotic Bcl-2 family proteins, as well as with cleavage of caspase-3. The IC50 values of artesunate in cell lines varied between 0.3 and 16.6 µm. Furthermore, some primary myeloma cells were also sensitive to artesunate at doses around 10 µm. Concentrations of this order are pharmacologically relevant as they can be obtained in plasma after intravenous administration of artesunate for malaria treatment (Holien et al., 2013).

Osteosarcoma, Leukemia/Lymphoma

Artesunate inhibits growth and induces apoptosis in human osteosarcoma HOS cell line in vitro and in vivo (Xu et al. 2011). ART alone or combined with chemotherapy drugs could inhibit the proliferation of B/T lymphocytic tumor cell lines as well ALL primary cells in vitro, probably through the mechanism of apoptosis, which suggest that ART is likely to be a potential drug in the treatment of leukemia/lymphoma (Zeng et al., 2009).

References

Efferth, T., Sauerbrey, A., Olbrich, A., et al. (2003) Molecular modes of action of artesunate in tumor cell lines. Mol Pharmacol, 64(2):382-94.


Fan, Y., Zhang, Y.L., Yao, G.T., & Li, Y.K. (2008). Inhibition of Artemisunate on the invasion of human colon cancer line SW620. Lishizzhen Medicine and Materia Medica Research, 19(7), 1740-1741.


Hamacher-Brady, A., Stein, H.A., Turschner, S., et al. (2011). Artesunate activates mitochondrial apoptosis in breast cancer cells via iron-catalyzed lysosomal reactive oxygen species production. J Biol Chem. 2011;286(8):6587–6601. doi: 10.1074/jbc.M110.210047.


Holien, T., Olsen, O.E., Misund, K., et al. (2013). Lymphoma and myeloma cells are highly sensitive to growth arrest and apoptosis induced by artesunate. Eur J Haematol, 91(4):339-46. doi: 10.1111/ejh.12176.


Jiang, Z., Chai, J., Chuang, H.H., et al. (2012). Artesunate induces G0/G1 cell-cycle arrest and iron-mediated mitochondrial apoptosis in A431 human epidermoid carcinoma cells. Anti-cancer Drugs, 23(6):606–613. doi: 10.1097/CAD.0b013e328350e8ac.


Liu, L., Zuo, L.F., Guo, J.W. (2013). Reversal of Multi-drug resistance by the anti-malaria drug artesunate in the esophageal cancer Eca109/ABCG2 cell line. Oncol Lett, 6(5):1475-1481.


Liu, Y. & Cui, Y.F. (2013). Synergism of cytotoxicity effects of triptolide and artesunate combination treatment in pancreatic cancer cell lines. Asian Pac J Cancer Prev, 14(9):5243-8.


Ma, H., Yaom Q., Zhang, A.M., et al. (2011). The effects of artesunate on the expression of EGFR and ABCG2 in A549 human lung cancer cells and a xenograft model. Molecules, 16(12):10556–10569. doi: 10.3390/molecules161210556.


Marchion, D.C., Xiong, Y., Chon, H.S., et al. (2013). Gene expression data reveal common pathways that characterize the unifocal nature of ovarian cancer. Am J Obstet Gynecol, S0002-9378(13)00827-2. doi: 10.1016/j.ajog.2013.08.004.


Rasheed, S.A., Efferth, T., Asangani, I.A., Allgayer, H. (2010). First evidence that the antimalarial drug artesunate inhibits invasion and in vivo metastasis in lung cancer by targeting essential extracellular proteases. Int J Cancer, 127(6):1475–1485. doi: 10.1002/ijc.25315.


Xu, Q., Li, Z.X., Peng, H.Q., et al. (2011). Artesunate inhibits growth and induces apoptosis in human osteosarcoma HOS cell line in vitro and in vivo. J Zhejiang Univ-Sci B (Biomed & Biotechnol), 12(4):247–255. doi: 10.1631/jzus.B1000373.


Zhao, Y., Jiang, W., Li, B., et al. (2011). Artesunate enhances radiosensitivity of human non-small-cell lung cancer A549 cells via increasing no production to induce cell-cycle arrest at G2/M phase. Int Immunopharmacol, 11(12):2039–2046. doi: 10.1016/j.intimp.2011.08.017.


Zeng, Y., Ni, X., Meng, W.T., Wen, Q., Jia, Y.Q. (2009). Inhibitive effect of artesunate on human lymphoblastic leukemia/lymphoma cells. Sichuan Da Xue Xue Bao Yi Xue Ban, 40(6):1038-43.


Zhou, C., Pan, W., Wang, X.P., Chen, T.S. (2012). Artesunate induces apoptosis via a bak-mediated caspase-independent intrinsic pathway in human lung adenocarcinoma cells. J Cell Physiol, 227(12):3778–3786. doi: 10.1002/jcp.24086.

Multi-drug resistance

Multi-drug resistance in cancer chemotherapy refers to the ability of cancer cells to survive from treatment of a wide range of drugs (Meszaros et al., 2009).

In addition to the MDR induced by drugs in early exposure, the MDR cancer cells may subsequently develop cross-resistance to several unexposed and structurally unrelated chemotherapeutic agents (Biedler et al., 1970).

How to tackle the MDR cells in chemotherapy is a pressing issue in cancer treatments. Verapamil was the first known Pgp inhibitor to increase the intracellular concentration of anti-cancer agents in MDR cells by binding to Pgp and inhibiting the Pgp-mediated efflux (Twentyman, 1992). It was believed that anti-cancer drug resistance could be reversed by drug efflux inhibition. Researchers developed and tested a range of Pgp inhibitors to improve the pharmacological effects of chemotherapy in cancer patients (Tsuruo et al., 1981; Stewart et al., 2000; Toppmeyer et al., 2002).

Mechanisms of MDR include decreased uptake of drugs, alterations in cellular pathways and increased active efflux of drugs (Gottesman, 2002; La Porta, 2007; Watson, 1991).

Overexpression of ATP-binding cassette (ABC) transporters is one of the most common mechanisms. Overexpression of the three major ABC transporters, i.e. P-glycoprotein (Pgp), multi-drug resistance-associated protein 1 (MRP1) and breast cancer resistance protein (BCRP/ABCG2), is frequently observed in cancer cell lines selected with chemotherapeutic drugs (Szakacs et al., 2006) and critical to clinical drug resistance (Leonard, 2003).

Fractions from 17 clinically used anti-tumor traditional Chinese medicinal herbs were tested for their potential to restore the sensitivity of MCF-7/ADR and A549/Taxol cells to a known anti-neoplastic agent. Five herbs, Curcuma wenyujin, Chrysanthemum indicum, Salvia chinensis, Ligusticum chuanxiong Hort. and Cassia tora L., could sensitize these resistant cancer cells at a non-toxic concentration (10  µg mL–1), and markedly increased doxorubicin accumulation in MCF-7/ADR cells, which necessitates further investigations into the active ingredients of these herbs and their underlying mechanisms (Yang et al., 2011).

Natural sources are a fertile ground to find novel drugs with activity against MDR cancer cells. In some countries, especially China, traditional herbal medicines are often used together with mainstream chemotherapeutic agents. The clinically used traditional Chinese herbs for the treatment of tumor can be classified into four categories based on the theory of Traditional Chinese Medicine (TCM): drugs (CH group) for 'Clearing away Heat and Toxins', drugs (PB group) for 'Promoting Blood Flow to Remove Stasis', drugs for 'Invigoration' and toxic drugs. Drugs for 'Invigoration' have indirect anti-neoplastic action by enhancing an organism's immunity and have been used clinically to minimise radiotherapy- and chemotherapy-induced toxicity (Fu & Chen, 2008; Chai, To, Lin, 2010).

Some of the recent findings on the circumvention of ABC transporters-mediated MDR by various ingredients and extracts of CM and their formulae, based on whether the MDR reversal involved Pgp alteration, are reviewed below.

Saponins

Ginsenosides are the major active components from Panax ginseng (Renshen). Ginsenosides are mainly triterpenoid dammarane derivatives. Several ginsenosides, namely Rg1, Rg3, Re, Rc and Rd inhibited drug efflux (Kim et al., 2003). A combination of purified saponins containing Rb1, Rb2, Rc, Rd, Re and Rg1 reversed MDR whereas individual ginsenosides did not produce any effect (Park et al., 2006). Ginsenosides reversed MDR of several chemotherapeutic drugs such as homoharringtonine, cytarabine, doxorubicin and etoposide in K562/VCR and in a dose-dependent manner in K562/DOX (Gao et al., 2004).

Pgp expression decreased but bcl-2 expression remained the same (Wang, 2003). Rb1 reversed MDR of harringtonolide and vincristine in K562/HHT and HL60/VCR cell lines respectively (Shi et al. , 2005).

Panax notoginseng (Sanqi) total saponins reversed MDR of doxorubicin in MCF-7/DOX and K562/VCR cell lines. The mechanism may be related to the decrease of Pgp expression (Si & Tien, 2005; Liu, Liu, & Fang, 2008).

Rg3, one of the active ginsenosides from Panax ginseng, restored the sensitivity of resistant KBV20 cell line to various anti-cancer drugs, including vincristine, doxorubicin, etoposide and colchicine in a time-and dose-dependent manner. This ginsenoside competitively inhibited the binding of substrate drugs to Pgp and its binding affinity to Pgp was remarkably higher than that of verapamil. In contrast to the dose-dependent effects in vitro, Rg3 increased animal life span in an in vivo MDR model in a dose-independent manner (Kim et al., 2003).

Flavonoids

Quercetin is one of the most widely distributed flavonoids in natural products including Chinese medicinal herbs such as Sophora japonica (Huai). Quercetin inhibited the binding of heat shock factor at the MDR1 promoter, thereby decreasing MDR1 transcription and reducing Pgp expression (Kim et al., 1998). Quercetin also inhibited the overexpression of Pgp mediated by arsenite (Kioka et al., 1992). In HL-60/DOX and K562/DOX cell lines, quercetin enhanced the anti-cancer sensitivity to daunorubicin and decreased Pgp expression (Cai et al., 2004; Cai et al., 2005). MDR reversal effect of quercetin was probably mediated by its action on mitochondrial membrane potential and the induction of apoptosis. Furthermore, quercetin derivatives rather than quercetin itself reversed MDR (Kothan et al., 2004). Quercetin increased the sensitivity of Pgp-overexpressing KBV1 cell line towards vinblastine and paclitaxel in a dose-dependent manner. Among many active flavonoids, quercetin was less potent than kaempferol but more effective than genistein and daidzein in reversing MDR. Genistein and daidzein had no effect on Pgp expression (Limtrakul, Khantamat, & Pintha, 2005).

Although quercetin may be a potential MDR reversing agent, lethal drug-drug interaction between quercetin and digoxin has been reported. Quercetin (40 mg/kg) elevated the peak blood concentration of digoxin and caused sudden death of tested animals (Wang et al., 2004).

Paeonol is a weak calcium channel blocker isolated from the root of Paeonia suffruticosa (Mudan). In K562/DOX cell line, paeonol showed positive MDR reversal effect towards doxorubicin, daunorubicin, vincristine and vinblastine without modulating Pgp expression [100]. In parental K562 cells, paeonol induced apoptosis in a time-and dose-dependent manner (Sun et al., 2004).

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 & Amiji, 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 (Um et al., 2008). 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 (Um et al., 2008). Bisdemethoxycurcumin modified from curcumin resulted in greater inhibition of Pgp expression (Limtrakul, Anuchapreeda, & Buddhasukh, 2004). Tetrahydrocurcumin, the major metabolite of curcumin, inhibited all three major ABC transporters (Limtrakul et al., 2007). Curcumin induced atypical and caspase-independent cell death in MDR cells (Piwocka, Bielak-Mijewska, & Sikora, 2002). In leukaemic cells collected from 78 childhood leukaemia patients, curcumin reduced Pgp expression (Anuchapreeda et al., 2006). A specialized nanoemulsion of curcumin is better than conventional solution form drugs in enhancing the efficiency of drug delivery into the cells, down-regulating Pgp expression, inhibiting the NFκB pathway and promoting apoptotic response (Choi et al., 2008).

Other Compounds

Schizandrins, the active constituents of Schisandra chinensis (Wuweizi), were investigated for their MDR reversal effects. Schizandrin A was the most potent in reversing MDR by enhancing apoptosis and down-regulating Pgp and total protein kinase C expression. The crude extract of Schisandra chinensis reversed the resistance against vincristine in vivo (Huang et al., 2008). Deoxyschizandrin and γ-schizandrin, among the nine dibenzo[a,c]cyclooctadiene lignans examined, enhanced intracellular drug concentration and induced cell-cycle arrest at the G2/M phase when combined with sub-toxic dosages of doxorubicin (Slaninová et al., 2009). Gomisin A, on the other hand, altered Pgp-substrate interaction by binding to Pgp simultaneously with substrates (Wan et al., 2006).

Formulae – injections (See Injectables)

'Shengmai Injection', consisting of Panax ginseng and Ophiopogon japonicus (Maidong), down-regulated Pgp expression in peripheral blood lymphocyte membrane. When used together with oxaliplatin, 5-fluorouracil or folinic acid, the injection prolonged the survival rate of colon cancer patients (Cao et al., 2005). The injection also enhanced the efficacy of tamoxifen and nifedipine in combination therapy (Lin et al., 2002).

'KLT Injection' consisting of the extract of Coix lacryma-jobi (Yiyi) enhanced the anti-cancer activities of paclitaxel and docetaxel and reversed MDR in a dose-dependent manner (Dong, Zheng, & Lu, 2002).

Formulae – powders

'Shenghe Powder', consisting of Panax ginseng, Scorophularia ningpoensis (Xuanshen) and Atractylodes macrocephala (Baizhu), increased the intracellular concentration of vincristine in resistant SGC-7901/VCR cell line, possibly due to the induction of apoptosis and down-regulation of Pgp and bcl-2 expression (Wang et al., 2007).

'Modified Sanwubai Powder', consisting of herbs such as Croton tiglium (Badou), Platycodon grandiflorum (Jiegeng) and Fritillaria thunbergii, induced apoptosis in SGC-7901 cell line and down-regulated the gene expressions of p53, bcl-2, rasP21CD44 and Pgp (Xu et al., 2005).

Formulae – others

Three herbal extracts used to treat diseases other than cancer, namely Ams-11, Fw-13 and Tul-17, greatly enhanced the efficacy of vincristine both in vitro and in vivo and reversed MDR in a dose-dependent manner. Tul-17 inhibited Pgp expression (Qu et al., 2006).

Oil emulsion from Brucea javanica (Yadanzi) reversed MDR when used together with other chemotherapeutic drugs such as vincristine, doxorubicin, cisplatin, mitomycin C, 5-fluorouracil or etoposide, probably due to down-regulation of Pgp expression or inhibition of TOPO II or both (Yu, Wu, Zhang, 2001).

'Sangeng Mixture Decoction', consisting of Reynoutria japonica (Huzhang), Actinidia arguta (Mihouligen) and Geum aleppicum (Shuiyangmeigen), reversed MDR of doxorubicin via down-regulation of Pgp expression (Feng et al., 2003).

FFTLG, a formula containing Actinidia arguta, reversed MDR in K562/DOX cell line by increasing the intracellular doxorubicin concentration (Guo, Xie, Feng, 2002).

R1, consisting of Ligusticum chuanxiong, Curcuma longa and Millettia dielsiana (Jixueteng), enhanced the anti-cancer activities of doxorubicin in MCF-7/DOX via down-regulation of Pgp expression (Chen et al., 2003; Lin, 2007).

Formulae

'Ganli Injection', consisting of matrine and tetramethylpyazine hydrochloride, reversed MDR by increasing the sensitivity of 5-fluorouracil and the intracellular concentration of doxorubicin in BEL-7402/5-FU cell line (Gu et al., 2007).

'Bushen Huayu Jiedu Formula', consisting of Cinnamomum cassia (Rougui), Psoralea corylifolia (Buguzhi) and Rheum palmatum, was tested in A549/DDP cell line and S180 tumor-bearing mice. In vitro, the formula significantly increased the intracellular concentration of cisplatin at high doses and inhibited the activity of calcium channel and LRP-56 expression at both high and low doses. In vivo, the formula improved the serum concentration, reduced the inflow and the release of Ca2+ and inhibited the LRP gene expression (Cao et al., 2004; Cao et al., 2008).

Four CM formulae, namely Glycyrrhiza glabra (GLYC), Hedyotis diffusa (OLEN), a formula consisting of 15 herbs including Cistanche deserticola (Roucongrong), Rabdosia rubescens (Donglingcao) and Zanthoxylum nitidum (Liangmianzhen) (SPES), and a formula consisting of eight herbs including Serenoa repens (Juyezhong), Scutellaria baicalensis (Huangqin), Panax ginseng and Glycyrrhiza glabra (PC-SPES) were cytotoxic to cancer cell lines in a dose-dependent manner. SPES, PC-SPES, OLEN decreased the bcl-2 gene expression and were pro-apoptotic, while GLYC was pro-necrotic without altering the over-expression of bcl-2 in MDR cells. Furthermore, OLEN, SPES and PC-SPES exhibited similar pharmacological effects to etoposide and vincristine (Sadava et al., 2002).

References

Anuchapreeda S, Thanarattanakorn P, Sittipreechacharn S, et al. (2006). Inhibitory effect of curcumin on MDR1 gene expression in patient leukemic cells. Arch Pharm Res, 29(10):866-873

Biedler JL, Riehm H. (1970). Cellular resistance to actinomycin D in Chinese hamster cells in vitro: cross-resistance, radioautographic, and cytogenetic studies. Cancer Res, 30:1174-1184.

Cai X, Chen FY, Han JY, et al. (2004). Restorative effect of quercetin on subcellular distribution of daunorubicin in Multi-drug-resistant leukemia cell lines K562/ADM and HL-60/ADM. Chin J Cancer, 23(12):1611-1615.

Cai X, Chen FY, Han JY, et al. (2005). Reversal of Multi-drug resistance of HL-60 adriamycin resistant leukemia cell line by quercetin and its mechanisms. Chin J Oncol, 27(6):326-329.

Cao CM, Ding XD, Wang XH, Liu P. (2005). Clinical study of shengmai injection in its reversing MDR effect in late phase colon carcinoma patients. Shandong J Tradit Chin Med, 24(9):529-532.

Cao Y, Zhang D, Zheng GJ, Yang Y, Zhang J. (2004). Study on drug resistance reversion and mechanism of bushen huayu jiedu formula in lung cancer cells of drug resistance. Shandong J Trad Chin Med, 23(2):100-104.

Cao Y, Xia Q, Meng H, Zhong A. (2008). Pharmacological effects of serum containing chinese medicine bushen huayu jiedu compound recipe in lung cancer drug-resistance cells. Chin J Integr Med, 14(1):46-50.

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.

Choi BH, Kim CG, Lim Y, Shin SY, Lee YH. (2008). Curcumin down-regulates the Multi-drug resistance mdr1b gene by inhibiting the PI3K/Akt/NF kappa B pathway. Cancer Lett, 259(1):111-118.

Chen XY, Liu JT. (2003). Study on tumor cells' Multi-drug resistance and its reversion by Chinese herbs. J Chin Integr Med, 1(3):221-225.

Dong QH, Zheng S, Lu QH. (2002). Study on the effect of Kanglaite injection on MDR human leukemia cell lines. J Pract Oncol, 17(1):24-26.

Feng ZQ, Guo Y, Zhu NX, et al. (2003). The experiment of SANGENG mixture decoction on reversing Multi-drug resistance. Bull Chin Cancer, 12(6):370-371.

Fu YL and Chen T. (2008). Research of traditional Chinese medicine in the treatment of leukemia: current status. Journal of Chinese Integrative Medicine, 6: 867–872.

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.

Gao RL, Lin XJ, Qian XD, Chen XH, Niu YP. (2004). Effect of ginsenosides and panaxatriol extracted from ginseng on inhibition of proliferation, inducing apoptosis and cytotoxic drug sensitivity in leukemic cells. J Chin Med Res, 4(2):97-99.

Gottesman MM. (2002). Mechanisms of cancer drug resistance. Annu Rev Med, 53:615-617.

Gu W, Zhai XF, Zhang YN, Ling CQ. (2007). In vitro study of Ganli injection on reversing acquired Multi-drug resistance of human hepatocellular carcinoma. Chin Trad Herbal Drugs, 38(6):871-874.

Guo Y, Xie CS, Feng ZQ. (2002). The study of effects on accumulation and efflux of intracellular adrimycine with FFTLG for Multi-drug-resistant cell lines K562/ADR and K562/VCR in vitro. Chin J Mod Appl Pharm, 19(4):268-272.

Huang M, Jin J, Sun H, Liu GT. (2008). Reversal of P-glycoprotein-mediated Multi-drug resistance of cancer cells by five schizandrins isolated from the Chinese herb Fructus Schizandrae. Cancer Chemother Pharmacol, 62(6):1015-1026.

Kim SH, Yeo GS, Lim YS, et al. (1998). Suppression of Multi-drug resistance via inhibition of heat shock factor by quercetin in MDR cells. Exp Mol Med, 30(2):87-92.

Kim SW, Kwon HY, Chi DW, (2003). Reversal of P-glycoprotein-mediated Multi-drug resistance by ginsenoside Rg3. Biochem Pharmacol, 65(1):75-82.

Kioka N, Hosokawa N, Komano T, et al. (1992). Quercetin, a bioflavonoid, inhibits the increase of human Multi-drug resistance gene (MDR1) expression caused by arsenite. FEBS Lett, 301(3):307-309.

Kothan S, Dechsupa S, Leger G, et al. (2004). Spontaneous mitochondrial membrane potential change during apoptotic induction by quercetin in K562 and K562/adr cells. Can J Physiol Pharm, 82(12):1084-1090.

La Porta CAM. (2007). Drug resistance in melanoma: new perspectives. Curr Med Chem, 14(4):387-91.

Leonard GD, Fojo T, Bates SE. (2003). The role of ABC transporters in clinical practice. Oncologist, 8(5):411-424.

Li Y, Wang ZZ, Yu TF (2005). In vitro study on the reversal of Multi-drug resistance (MDR) in HL60/VCR cell line with ginsenoside – Rb1. J Radioimmunol, 18(5):362-365.

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

Limtrakul P, Chearwae W, Shukla S, Phisalphong C, Ambudkar SV. (2007). Modulation of function of three ABC drug transporters, P-glycoprotein (ABCB1), mitoxantrone resistance protein (ABCG2) and Multi-drug resistance protein 1 (ABCC1) by tetrahydrocurcumin, a major metabolite of curcumin. Mol Cell Biochem, 296(1-2):85-95.

Limtrakul P, Khantamat O, Pintha K. (2005). Inhibition of P-glycoprotein function and expression by kaempferol and quercetin. J Chemother, 17(1):86-95.

Lin SY, Qin ZQ, Xu YF, Wu LC, Liu LM. (2002). Significance of combination shenmal injection, tamoxifen and nifedipine in treatment stage IV non-small-cell lung cancer. Chin J Cancer Prev Treat, 9(4):454-455.

Lin GF. (2007). Current situation in studying the reversing effect in Multi-drug-resistant in cancer cells by TCM. China Pharm, 16(10):63-64.

Liu LL, Liu YE, Fang GT. (2008). Reversal effect of Panax notoginseng saponins on Multi-drug resistance breast cancer cell MCF/ADM. Lishizhen Med Materia Med Res, 19(4):954-956.

Meszaros A, Balogh G. (2009). Multiple Drug Resistance. Gazelle Distribution, Lancaster, UK.

Park JD, Kim DS, Kwon HY, et al. (1996). Effects of ginseng saponin on modulation of Multi-drug resistance. Arch Pharm Res, 19(3):213-218.

Wang L. (2003). Reversing drug resistance of human erythroleukemia cell line K562/DOX on DOX by total saponings Panax ginseng. J Chongqing Med Univ, 28(4):424-427, 435.

Piwocka K, Bielak-Mijewska A, Sikora E. (2002). Curcumin induces caspase-3-independent apoptosis in human Multi-drug-resistant cells. Ann N Y Acad Sci, 973:250-254.

Qu Y, Liu SQ, Zhang XB, et al. (2006). Reversal of P-glycoprotein mediated Multi-drug resistance by traditional Chinese medicines. Nat Prod Res Dev, 18(6):932-936.

Sadava D, Ahn J, Zhan M, et al. (2002). Effects of four Chinese herbal extracts on drug-sensitive and Multi-drug-resistant small-cell lung carcinoma cells. Cancer Chemother Pharmacol, 49(4):261-266.

Shi XK, Zhang YJ, Zhao CJ. (1999). The reverse effects of ginseng saponin Rb1 on Multi-drug resistance of human leukemic cell line K562/HHT in vitro. Acta Acad Med Militaris Tertiae, 21(11):825-827.

Si YQ, Tien TD (2005). The Multi-drug-resistant reversing effect of total Panax notoginseng saponins studied in K562/VCR cell line. Chin J Tradit Med Sci Technol, 12(5):292-294.

Slaninová I, Brezinová L, Koubíková L, Slanina J. (2009). Dibenzocyclooctadiene lignans overcome drug resistance in lung cancer cells–study of structure-activity relationship. Toxicol In Vitro, 23(6):1047-1054.

Stewart A, Steiner J, Mellows G, et al. (2000). Phase I trial of XR9576 in healthy volunteers demonstrates modulation of Pglycoprotein in CD56+ lymphocytes after oral and intravenous administration. Clin Cancer Res, 6(11):4186-4191.

Sun GP, Wang H, Shen YX, et al. (2004). Study on effects of paeonol in inhibiting growth of K562 and inducing its apoptosis. Chin Pharmacol Bull, 20(5):550-552.

Szakacs G, Paterson JK, Ludwig JA, Booth-Genthe C, Gottesman MM. (2006). Targeting Multi-drug resistance in cancer. Nat Rev Drug Discov, 5(3):219-234.

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.

Toppmeyer D, Seidman AD, Pollak M, et al. (2002). Safety and efficacy of the Multi-drug resistance inhibitor Incel (biricodar; VX-710) in combination with paclitaxel for advanced breast cancer refractory to paclitaxel. Clin Cancer Res, 8(3):670-678.

Tsuruo T, Iida H, Tsukagoshi S, Sakurai Y. (1981). Overcoming of vincristine resistance in P388 leukemia in vivo and in vitro through enhanced cytotoxicity of vincristine and vinblastine by verapamil. Cancer Res, 41(5):1967-1972.

Twentyman PR. (1992). Cyclosporins as drug resistance modifiers. Biochem Pharmacol, 43(1):109-117.

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.

Wan CK, Zhu GY, Shen XL, et al. (2006). Gomisin A alters substrate interaction and reverses P-glycoprotein-mediated Multi-drug resistance in HepG2-DR cells. Biochem Pharmacol, 72(7):824-837.

Wang Y, Chao PL, Hsiu SL, Wen K, Hou Y. (2004). Lethal quercetin-digoxin interaction in pigs. Life Sci, 74(10):1191-1197.

Wang J, Xia Y, Wang H, Hou Z. (2007). Chinese herbs of Shenghe powder reverse Multi-drug resistance of gastric carcinoma SGC-7901. Integr Cancer Ther, 6(4):400-404

Watson JV. (1991). Introduction to Flow Cytometry Cambridge. Cambridge University Press, Cambridge, UK.

Xu L, Wang MY, Xu DQ, Zhou CX. (2005). Experimental study on modified 'Sanwubai Powder' in affecting multi-drug resistance gene expression of tumor. Shanghai J Tradit Chin Med, 39(8):59-60.

Yang L, Wei DD, Chen Z, et al. (2011). Reversal effects of traditional Chinese herbs on Multi-drug resistance in cancer cells. Natural Product Research (Formerly Natural Product Letters), 25(19):1885-1889. doi: 10.1080/14786419.2010.541395

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.

Yu LF, Wu YL, Zhang YP. (2001). Reversal of drug resistance in the vincristine-resistant human gastric cancer cell lines MKN28/VCR by emulsion of seminal oil of Brucea Javanica. World Chin J Digestol, 9(4):376-378.

Baicalin & Baicalein

Cancer:
Myeloma, liver, colorectal., breast, prostate, oral., hepatoma, ovarian

Action: Anti-cancer, cardiovascular disease, cytostatic, cardio-protective against Doxorubicin, anti-inflammatory, angiogenesis

Baicalin and baicalein are naturally occurring flavonoids that are found in the roots and leaves of some Chinese medicinal plants (including Scutellaria radix, Scutellaria rivularis (Benth.); Scutellaria baicalensis (Georgi) and Scutellaria lateriflora (L.)) are thought to have anti-oxidant activity and possible anti-angiogenic, anti-cancer, anxiolytic, anti-inflammatory and neuroprotective activities. In particular, Scutellaria baicalensis is one of the most popular and multi-purpose herbs used in China traditionally for treatment of inflammation, hypertension, cardiovascular diseases, and bacterial and viral infections (Ye et al., 2002; Zhang et al., 2011a).

Anti-cancer

Accumulating evidence demonstrates that Scutellaria also possesses potent anti-cancer activities. The bioactive components of Scutellaria have been confirmed to be flavones, wogonin, baicalein and baicalin. These phytochemicals are not only cytostatic but also cytotoxic to various human tumor cell lines in vitro and inhibit tumor growth in vivo. Most importantly, they show almost no or minor toxicity to normal epithelial and normal peripheral blood and myeloid cells. The anti-tumor functions of these flavones are largely due to their abilities to scavenge oxidative radicals, to attenuate NF-kappaB activity, to inhibit several genes important for regulation of the cell-cycle, to suppress COX-2 gene expression and to prevent viral infections (Li, 2008).

Multiple Myeloma

In the search for a more effective adjuvant therapy to treat multiple myeloma (MM), Ma et al. (2005) investigated the effects of the traditional Chinese herbal medicines Huang-Lian-Jie-Du-Tang (HLJDT), Gui-Zhi-Fu-Ling-Wan (GZFLW), and Huang-Lian-Tang (HLT) on the proliferation and apoptosis of myeloma cells. HLJDT inhibited the proliferation of myeloma cell lines and the survival of primary myeloma cells, especially MPC-1- immature myeloma cells, and induced apoptosis in myeloma cell lines via a mitochondria-mediated pathway by reducing mitochondrial membrane potential and activating caspase-9 and caspase-3.

Further experiments confirmed that Scutellaria radix was responsible for the suppressive effect of HLJDT on myeloma cell proliferation, and the baicalein in Scutellaria radix showed strong growth inhibition and induction of apoptosis in comparison with baicalin or wogonin. Baicalein as well as baicalin suppressed the survival in vitro of MPC-1- immature myeloma cells rather than MPC-1+ myeloma cells from myeloma patients.

Baicalein inhibited the phosphorylation of IkB-alpha, which was followed by decreased expression of the IL-6 and XIAP genes and activation of caspase-9 and caspase-3. Therefore, HLJDT and Scutellaria radix have an anti-proliferative effect on myeloma cells, especially MPC-1- immature myeloma cells, and baicalein may be responsible for the suppressive effect of Scutellaria radix by blocking IkB-alpha degradation (Ma, 2005).

Hepatoma

The effects of the flavonoids from Scutellaria baicalensis Georgi (baicalein, baicalin and wogonin) in cultured human hepatoma cells (Hep G2, Hep 3B and SK-Hep1) were compared by MTT assay and flow cytometry. All three flavonoids dose-dependently decreased the cell viabilities accompanying the collapse of mitochondrial membrane potential and the depletion of glutathione content. However, the influence of baicalein, baicalin or wogonin on cell-cycle progression was different.

All three flavonoids resulted in prominent increase of G2/M population in Hep G2 cells, whereas an accumulation of sub G1 (hypoploid) peak in Hep 3B cells was observed. In SK-Hep1 cells, baicalein and baicalin resulted in a dramatic boost in hypoploid peak, but wogonin mainly in G1 phase accumulation. These data, together with the previous findings in other hepatoma cell lines, suggest that baicalein, baicalin and wogonin might be effective candidates for inducing apoptosis or inhibiting proliferation in various human hepatoma cell lines (Chang, 2002).

Long dan xie gan tang (pinyin) is one of the most commonly used herbal formulas by patients with chronic liver disease in China. Accumulated anecdotal evidence suggests that Long dan tang may have beneficial effects in patients with hepatocellular carcinoma. Long dan tang is comprised of five herbs: Gentiana root, Scutellaria root, Gardenia fruit, Alisma rhizome, and Bupleurum root. The cytotoxic effects of compounds from the five major ingredients isolated from the above plants, i.e. gentiopicroside, baicalein, geniposide, alisol B acetate and saikosaponin-d, were investigated, respectively, on human hepatoma Hep3B cells..

Interestingly, baicalein by itself induced an increase in H(2)O(2) generation and the subsequent NF-kappaB activation; furthermore, it effectively inhibited the transforming growth factor-beta(1) (TGF-beta(1))-induced caspase-3 activation and cell apoptosis. Results suggest that alisol B acetate and saikosaponin-d induced cell apoptosis through the caspase-3-dependent and -independent pathways, respectively. Instead of inducing apoptosis, baicalein inhibits TGF-beta(1)-induced apoptosis via increase in cellular H(2)O(2) formation and NF-kappaB activation in human hepatoma Hep3B cells (Chou, Pan, Teng & Guh, 2003).

Ovarian Cancer

Ovarian cancer is one of the primary causes of death for women all through the Western world. Two kinds of ovarian cancer (OVCAR-3 and CP-70) cell lines and a normal ovarian cell line (IOSE-364) were selected to be investigated in the inhibitory effect of baicalin and baicalein on cancer cells. Largely, baicalin and baicalein inhibited ovarian cancer cell viability in both ovarian cancer cell lines with LD50 values in the range of 45-55 µM for baicalin and 25-40 µM for baicalein. On the other hand, both compounds had fewer inhibitory effects on normal ovarian cells viability with LD50 values of 177 µM for baicalin and 68 µM for baicalein.

Baicalin decreased expression of VEGF (20 µM), cMyc (80 µM), and NFkB (20 µM); baicalein decreased expression of VEGF (10 µM), HIF-1α (20 µM), cMyc (20 µM), and NFkB (40 µM). Therefore baicalein is more effective in inhibiting cancer cell viability and expression of VEGF, HIF-1α, cMyc, and NFκB in both ovarian cancer cell lines. It seems that baicalein inhibited cancer cell viability through the inhibition of cancer promoting genes expression including VEGF, HIF-1α, cMyc, and NFκB.

Overall, this study showed that baicalein and baicalin significantly inhibited the viability of ovarian cancer cells, while generally exerting less of an effect on normal cells. They have potential for chemoprevention and treatment of ovarian cancers (Chen, 2013).

Breast Cancer

Baicalin was found to be a potent inhibitor of mammary cell line MCF-7 and ductal breast epithelial tumor cell line T-47D proliferation, as well as having anti-proliferative effects on other cancer types such as the human head and neck cancer epithelial cell lines CAL-27 and FaDu. Overall, baicalin inhibited the proliferation of human breast cancer cells and CAL-27 and FaDu cells with effective potency (Franek, 2005).

Breast Cancer, Cell Invasion

The effect of Baicalein on cell viability of the human breast cancer MDA-MB-231 cell line was tested by MTT. 50, 100 µmol·L-1 of Baicalein inhibited significantly cell invasion(P0.01) and migration(P0.01) compared with control groups. The inhibitory rates were 50% and 77% in cell migration and 15% and 44% in cell invasion, respectively. 50 µmol·L-1 of Baicalein significantly inhibited the level of MMP 2 expression. 100 µmol·L-1 of Baicalein significantly inhibited the level of MMP 9 and uPA expressions.

Baicalein inhibits invasion and migration of MDA-MB-231 cells. The mechanisms may be involved in the direct inhibition of cell invasion and migration abilities, and the inhibition of MMP 2, MMP 9, and uPA expressions (Wang et al., 2010).

The proliferation of MDA-MB-231 cell line human breast adenocarcinoma was inhibited by baicalin in a dose-and time-dependent manner and the IC50 was 151 µmol/L. The apoptotic rate of the baicalin-treated MDA-MB-231 cells increased significantly at 48 hours. Flow cytometer analysis also revealed that most of the baicalin-treated MDA-MB-231 cells were arrested in the G2/M phase. Typically apoptotic characteristics such as condensed chromatin and apoptotic bodies were observed after being treated with baicalin for 48 hours.

The results of RT-PCR showed that the expression of bax was up-regulated; meanwhile, the expression of bcl-2 was down-regulated. Baicalin could inhibit the proliferation of MDA-MB-231 cells through apoptosis by regulating the expression of bcl-2, bax and intervening in the process of the cell-cycle (Zhu et al., 2008).

Oral Cancer

As an aryl hydrocarbon receptor (AhR) ligand, baicalein at high concentrations blocks AhR-mediated dioxin toxicity. Because AhR had been reported to play a role in regulating the cell-cycle, it is suspected that the anti-cancer effect of baicalein is associated with AhR. The molecular mechanism involved in the anti-cancer effect of baicalein in oral cancer cells HSC-3 has been investigated, including whether such an effect would be AhR-mediated. Results revealed that baicalein inhibited cell proliferation and increased AhR activity in a dose-dependent manner. Cell-cycle was arrested at the G1 phase and the expression of CDK4, cyclin D1, and phosphorylated retinoblastoma (pRb) was decreased.

When cells were pre-treated with LiCl, the inhibitor of GSK-3β, the decrease of cyclin D1 was blocked and the reduction of pRb was recovered. The data indicates that in HSC-3 the reduction of pRb is mediated by baicalein both through activation of AhR and facilitation of cyclin D1 degradation, which causes cell-cycle arrest at the G1 phase, and results in the inhibition of cell proliferation (Cheng, 2012).

Anti-inflammatory

Baicalin has also been examined for its effects on LPS-induced nitric oxide (NO) production and iNOS and COX-2 gene expressions in RAW 264.7 macrophages. The results indicated that baicalin inhibited LPS-induced NO production in a concentration-dependent manner without a notable cytotoxic effect on these cells. The decrease in NO production was consistent with the inhibition by baicalin of LPS-induced iNOS gene expression (Chen, 2001)

Angiogenesis Modulation

The modulation of angiogenesis is one possible mechanism by which baicalin may act in the treatment of cardiovascular diseases. This may be elucidated by investigating the effects of baicalin on the expression of vascular endothelial growth factor (VEGF), a critical factor for angiogenesis. The effects of baicalin and an extract of S. baicalensis on VEGF expression were tested in several cell lines. Both agents induced VEGF expression in all cells without increasing expression of hypoxia-inducible factor-1alpha (HIF-1alpha).

Their ability to induce VEGF expression was suppressed once ERRalpha expression was knocked down by siRNA, or ERRalpha-binding sites were deleted in the VEGF promoter. It was also found that both agents stimulated cell migration and vessel sprout formation from the aorta. These results therefore implicate baicalin and S. baicalensis in angiogenesis by inducing VEGF expression through the activation of the ERRalpha pathway (Zhang, 2011b).

Colon Cancer

The compounds of baicalein and wogonin, derived from the Chinese herb Scutellaria baicalensis, were studied for their effect in suppressing the viability of HT-29 human colon cancer cells. Following treatment with baicalein or wogonin, several apoptotic events were observed, including DNA fragmentation, chromatin condensation and increased cell-cycle arrest at the G1 phase. Baicalein and wogonin decreased Bcl-2 expression, whereas the expression of Bax was increased in a dose-dependent manner when compared to the control.

The results indicated that baicalein induced apoptosis via Akt activation, in a p53-dependent manner, in HT-29 colon cancer cells. Baicalein may serve as a chemo-preventive, or therapeutic, agent for HT-29 colon cancer (Kim et al., 2012).

Cardio-protective

The cardiotoxicity of doxorubicin limits its clinical use in the treatment of a variety of malignancies. Previous studies suggest that doxorubicin-associated cardiotoxicity is mediated by reactive oxygen species (ROS)-induced apoptosis. Baicalein attenuated phosphorylation of JNK induced by doxorubicin. Co-treatment of cardiomyocytes with doxorubicin and JNK inhibitor SP600125 (10 µM; 24 hours) reduced JNK phosphorylation and enhanced cell survival., suggesting that the baicalein protection against doxorubicin cardiotoxicity was mediated by JNK activation. Baicalein adjunct treatment confers anti-apoptotic protection against doxorubicin-induced cardiotoxicity without compromising its anti-cancer efficacy (Chang et al., 2011).

Prostate Cancer

There are four compounds capable of inhibiting prostate cancer cell proliferation in Scutellaria baicalensis: baicalein, wogonin, neobaicalein, and skullcapflavone. Comparisons of the cellular effects induced by the entire extract versus the four-compound combination produced comparable cell-cycle changes, levels of growth inhibition, and global gene expression profiles (r(2) = 0.79). Individual compounds exhibited anti-androgenic activities with reduced expression of the androgen receptor and androgen-regulated genes. In vivo, baicalein (20 mg/kg/d p.o.) reduced the growth of prostate cancer xenografts in nude mice by 55% at 2 weeks compared with placebo and delayed the average time for tumors to achieve a volume of approximately 1,000 mm(3) from 16 to 47 days (P < 0.001).

Most of the anti-cancer activities of S. baicalensis can be recapitulated with four purified constituents that function in part through inhibition of the androgen receptor signaling pathway (Bonham et al., 2005)

Cancer: Acute lymphocytic leukemia, lymphoma and myeloma

Action: Cell-cycle arrest, induces apoptosis

Scutellaria baicalensis (S.B.) is a widely used Chinese herbal medicine. S.B inhibited the growth of acute lymphocytic leukemia (ALL), lymphoma and myeloma cell lines by inducing apoptosis and cell cycle arrest at clinically achievable concentrations. The anti-proliferative effectwas associated with mitochondrial damage, modulation of the Bcl family of genes, increased level of the CDK inhibitor p27KIP1 and decreased level of c-myc oncogene. HPLC analysis of S.B. showed it contains 21% baicalin and further studies confirmed it was the major anti-cancer component of S.B. Thus, Scutellaria baicalensis should be tested in clinical trials for these hematopoietic malignancies (Kumagai et al., 2007).

References

Bonham M, Posakony J, Coleman I, Montgomery B, Simon J, Nelson PS. (2005). Characterization of chemical constituents in Scutellaria baicalensis with antiandrogenic and growth-inhibitory activities toward prostate carcinoma. Clin Cancer Res, 11(10):3905-14.


Chang WH Chen CH Lu FJ. (2002). Different Effects of Baicalein, Baicalin and Wogonin on Mitochondrial Function, Glutathione Content and cell-cycle Progression in Human Hepatoma Cell Lines. Planta Med, 68(2):128-32. doi: 10.1055/s-2002-20246


Chang WT, Li J, Huang HH, et al. (2011). Baicalein protects against doxorubicin-induced cardiotoxicity by attenuation of mitochondrial oxidant injury .and JNK activation. J Cell Biochem. doi: 10.1002/jcb.23201.


Chen J, Li Z, Chen AY, Ye X, et al. (2013). Inhibitory effect of baicalin and baicalein on ovarian cancer cells. Int J Mol Sci, 14(3):6012-25. doi: 10.3390/ijms14036012.


Chen YC, Shen SC, Chen LG, Lee TJ, Yang LL. (2001). Wogonin, baicalin, and baicalein inhibition of inducible nitric oxide synthase and cyclooxygenase-2 gene expressions induced by nitric oxide synthase inhibitors and lipopolysaccharide. Biochem Pharmacol,61(11):1417-27. doi:10.1016/S0006-2952(01)00594-9


Cheng YH, Li LA, Lin P, et al. (2012). Baicalein induces G1 arrest in oral cancer cells by enhancing the degradation of cyclin D1 and activating AhR to decrease Rb phosphorylation. Toxicol Appl Pharmacol, 263(3):360-7. doi: 10.1016/j.taap.2012.07.010.


Chou CC, Pan SL, Teng CM, & Guh JH. (2003). Pharmacological evaluation of several major ingredients of Chinese herbal medicines in human hepatoma Hep3B cells. European Journal of Pharmaceutical Sciences, 19(5), 403-12.


Franek KJ, Zhou Z, Zhang WD, Chen WY. (2005). In vitro studies of baicalin alone or in combination with Salvia miltiorrhiza extract as a potential anti-cancer agent. Int J Oncol, 26(1):217-24.


Kim SJ, Kim HJ, Kim HR, et al. (2012). Anti-tumor actions of baicalein and wogonin in HT-29 human colorectal cancer cells. Molecular Medicine Reports, 6(6):1443-1449. doi: 10.3892/mmr.2012.1085.


Li-Weber M. (2009). New therapeutic aspects of flavones: The anti-cancer properties of Scutellaria and its main active constituents Wogonin, Baicalein and Baicalin. Cancer Treat Rev, 35(1):57-68. doi: 10.1016/j.ctrv.2008.09.005.


Ma Z, Otsuyama K, Liu S, et al. (2005). Baicalein, a component of Scutellaria radix from Huang-Lian-Jie-Du-Tang (HLJDT), leads to suppression of proliferation and induction of apoptosis in human myeloma cells. Blood, 105(8):3312-8. doi:10.1182/blood-2004-10-3915.


Wang Xf, Zhou Qm, Su Sb. (2010). Experimental study on Baicalein inhibiting the invasion and migration of human breast cancer cells. Zhong Guo Yao Li Xue Tong Bao, 26(6): 745-750.


Zhang XW, Li WF, Li WW, et al. (2011a). Protective effects of the aqueous extract of Scutellaria baicalensis against acrolein-induced oxidative stress in cultured human umbilical vein endothelial cells. Pharm Biol, 49(3): 256–261. doi:10.3109/13880209.2010.501803.


Ye F, Xui L, Yi J, Zhang, W, Zhang DY. (2002). Anti-cancer activity of Scutellaria baicalensis and its potential mechanism. J Altern Complement Med, 8(5):567-72.


Zhang K, Lu J, Mori T, et al. (2011b). Baicalin increases VEGF expression and angiogenesis by activating the ERR{alpha}/PGC-1{alpha} pathway.[J]. Cardiovascular Research, 89(2):426-435.


Zhu Gq, Tang Lj, Wang L, Su Jj, et al. (2008). Study on Baicalin Induced Apoptosis of Human Breast Cancer Cell Line MDA-MB-231. An Hui Zhong Yi Xue Yuan Xue Bao, 27(2):20-23

Kumagai T, et al. (2007) Scutellaria baicalensis, a herbal medicine: Anti-proliferative and apoptotic activity against acute lymphocytic leukemia, lymphoma and myeloma cell lines. Leukemia Research 31 (2007) 523-530

Emodin (See also Aloe-Emodin)

Cancer:
Breast, colon, liver, chemotherapy, myeloma, oral., pancreatic, hepatocellular carcinoma, lung, leukemia

Action: MDR-1, cell-cycle arrest

Emodin is an active natural anthraquinone derivative component of a traditional Chinese and Japanese medicine isolated from the root and rhizomes of Rheum palmatum L., Senna obtusifolia [(L.) H.S.Irwin & Barneby], Fallopia japonica [Houtt. (Ronse Decr.)], Kalimeris indica (L.) Sch.Bip., Ventilago madraspatana (Gaertn.), Rumex nepalensis (Spreng.), Fallopia multiflora [(Thunb.) Haraldson], Cassia occidentalis [(L.) Link], Senna siamea [(Lam.) Irwin et Barneby] and Acalypha australis (L.).

Aloe-emodin is an active natural anthraquinone derivative, and is found in the roots and rhizomes of numerous Chinese medicinal herbs (including Rheum palmatum L) and exhibits anti-cancer effects on many types of human cancer cell lines.

Administration of rhubarb (Emodin) can effectively reverse severe acute pancreatitis (SAP) by regulating the levels of IL-15 and IL-18 (Yu & Yang, 2013).

Pancreatic Cancer

Emodin is a tyrosine kinase inhibitor that has an inhibitory effect on mammalian cell-cycle modulation in specific oncogene-overexpressing cells. Recently, there has been great progress in the preclinical study of the anti-cancer mechanisms of emodin. A recent study revealed that emodin has therapeutic effects on pancreatic cancer through various anti-tumor mechanisms. Notably, the therapeutic efficacy of emodin in combination with chemotherapy was found to be higher than the comparable single chemotherapeutic regime, and the combination therapy also exhibited fewer side-effects (Wei et al., 2013).

Hepatocellular Carcinoma, Pancreatic, Breast, Colorectal and Lung Cancers, and Leukemia

Emodin is found as an active ingredient in different Chinese herbs including Rheum palmatum and Polygonam multiflorum, and has diuretic, vasorelaxant, anti-bacterial., anti-viral., anti-ulcerogenic, anti-inflammatory, and anti-cancer effects. The anti-inflammatory effects of emodin have been exhibited in various in vitro as well as in vivo models of inflammation including pancreatitis, arthritis, asthma, atherosclerosis and glomerulonephritis. As an anti-cancer agent, emodin has been shown to suppress the growth of various tumor cell lines including hepatocellular carcinoma, pancreatic, breast, colorectal., leukemia, and lung cancers. Emodin is a pleiotropic molecule capable of interacting with several major molecular targets including NF-κB, casein kinase II, HER2/neu, HIF-1α, AKT/mTOR, STAT3, CXCR4, topoisomerase II, p53, p21, and androgen receptors which are involved in inflammation and cancer (Shrimali et al., 2013).

Hepatocellular Carcinoma

It has been found that emodin induces apoptotic responses in the human hepatocellular carcinoma cell lines (HCC) Mahlavu, PLC/PRF/5 and HepG2. The addition of emodin to these three cell lines led to inhibition of growth in a time-and dose-dependent manner. Emodin generated reactive oxygen species (ROS) in these cells which brought about a reduction of the intracellular mitochondrial transmembrane potential (ΔΨ m), followed by the activation of caspase–9 and caspase–3, leading to DNA fragmentation and apoptosis.

Preincubation of hepatoma cell lines with the hydrogen peroxide-scavenging enzyme, catalase (CAT) and cyclosporin A (CsA), partially inhibited apoptosis. These results demonstrate that enhancement of generation of ROS, DeltaPsim disruption and caspase activation may be involved in the apoptotic pathway induced by emodin (Jing et al., 2002).

Colon Cancer

In in vitro study, emodin induced cell morphological changes, decreased the percentage of viability, induced G2/M phase arrest and increased ROS and Ca(2+) productions as well as loss of mitochondrial membrane potential (ΔΨ(m)) in LS1034 cells. Emodin-triggered apoptosis was also confirmed by DAPI staining and these effects are concentration-dependent.

In in vivo study, emodin effectively suppressed tumor growth in tumor nude mice xenografts bearing LS1034. Overall, the potent in vitro and in vivo anti-tumor activities of emodin suggest that it might be developed for treatment of colon cancer in the future (Ma et al., 2012).

Myeloid Leukemia

It has been shown that emodin significantly induces cytotoxicity in the human myeloma cells through the elimination of myeloid cell leukemia 1 (Mcl-1). Emodin inhibited interleukin-6–induced activation of Janus-activated kinase 2 (JAK2) and phosphorylation of signal transducer and activator of transcription 3 (STAT3), followed by the decreased expression of Mcl-1. Activation of caspase-3 and caspase-9 was triggered by emodin, but the expression of other anti-apoptotic Bcl-2 family members, except Mcl-1, did not change in the presence of emodin. To clarify the importance of Mcl-1 in emodin-induced apoptosis, the Mcl-1 expression vector was introduced into the human myeloma cells by electroporation. Induction of apoptosis by emodin was almost abrogated in Mcl-1–overexpressing myeloma cells as the same level as in parental cells, which were not treated with emodin. Emodin therefore inhibits interleukin-6–induced JAK2/STAT3 pathway selectively and induces apoptosis in myeloma cells via down-regulation of Mcl-1, which is a good target for treating myeloma. Taken together, these results show emodin as a new potent anti-cancer agent for the treatment of multiple myeloma patients (Muto et al., 2007).

Breast Cancer; Block HER-2

The mechanism by which emodin prevents breast cancer is unknown; however the product of the HER-2/neu proto-oncogene, HER2 has been proposed to be involved. The product of the HER-2/neu proto-oncogene, HER2, is the second member of the human epidermal growth factor receptor (HER) family of tyrosine kinase receptors and has been suggested to be a ligand orphan receptor. Amplification of the HER2 gene and overexpression of the HER2 protein induces cell transformation and has been demonstrated in 10% to 40% of human breast cancer. HER2 overexpression has been suggested to associate with tumor aggressiveness, prognosis and responsiveness to hormonal and cytotoxic agents in breast cancer patients. These findings indicate that HER2 is an appropriate target for tumor-specific therapies.

A number of approaches have been investigated: (1) a humanized monoclonal antibody against HER2, rhuMAbHER2 (trastuzumab), which is already approved for clinical use in the treatment of patients with metastatic breast cancer; (2) tyrosine kinase inhibitors, such as emodin, which block HER2 phosphorylation and its intracellullar signaling; (3) active immunotherapy, such as vaccination; and (4) heat shock protein (Hsp) 90-associated signal inhibitors, such as radicicol derivatives, which induce degradation of tyrosine kinase receptors, such as HER2 (Kurebayashi, 2001).

MDR

The effects of emodin on the nucleoside transport and multi-drug resistance in cancer cells has also been investigated. Nucleoside transport inhibition was determined by thymidine incorporation assay. The cytotoxicity to cancer cells was determined by MTT assay. The pump efflux activity and the expression of P glycoprotein were examined by flow cytometric assay. Emodin was active in the inhibition of nucleoside transport, with an IC 50 value of 9 9 µmol·L -1. Emodin markedly enhanced the cytotoxicity of 5 FU, MMC and MTX against human hepatoma BEL 7402 cells and partly reversed the multi-drug resistance in human breast cancer MCF 7/Adr cells.

Emodin inhibited P-gp pump efflux activity and reduced the expression of P gp in MCF 7/Adr cells. These findings provide a biological basis for the application of emodin as a biochemical modulator to potentiate the effects of anti-tumor drugs and reverse the multi-drug resistance in cancer cells (Jiang et al., 2009).

Cell-cycle Arrest

Large quantities of emodin were isolated from the roots of Rheum emodi and a library of novel emodin derivatives 2–15 were prepared to evaluate their anti-proliferative activities against HepG2, MDA-MB-231 and NIH/3T3 cells lines. The derivatives 3 and 12 strongly inhibited the proliferation of HepG2 and MDA-MB-231 cancer cell line with an IC50 of 5.6, 13.03 and 10.44, 5.027, respectively, which is comparable to marketed drug epirubicin (III). The compounds 3 and 12 were also capable of inducing cell-cycle arrest and caspase dependent apoptosis in HepG2 cell lines and exhibit DNA intercalating activity. These emodin derivatives hold promise for developing safer alternatives to the marketed epirubicin (Narender et al., 2013).

Cell-cycle Arrest; MDR1 & AZT

3'-azido-3'-deoxythymidine (AZT) and emodin altered the cell-cycle distribution and led to an accumulation of cells in S phase. Meanwhile, the expression of MDR1 mRNA/p-gp protein was markedly decreased. These results show a synergistic growth-inhibitory effect of AZT and emodin in K562/ADM cells, which is achieved through S phase arrest. MDR1 might ultimately be responsible for these phenomena (Chen et al., 2013).

References

Chen P, Liu Y, Sun Y, et al. (2013). AZT and emodin exhibit synergistic growth-inhibitory effects on K562/ADM cells by inducing S phase cell-cycle arrest and suppressing MDR1 mRNA/p-gp protein expression. Pharm Biol.


Garg AK, Buchholz TA, Aggarwal BB. (2005). Chemo-sensitization and Radiosensitization of Tumors by Plant Polyphenols. Antioxid Redox Signal., 7(11-12):1630-47.


Jiang XF & Zhen YS. (1999). Reversal of Multi-drug resistance by emodin in cancer cells. Acta Pharmaceutica Sinica, 1999-03.


Jing X, Ueki N, Cheng J, Imanishi H, Hada T. (2002). Induction of apoptosis in hepatocellular carcinoma cell lines by emodin. Cancer Science, 93(8):874–882.


Kurebayashi J. (2001). Biological and clinical significance of HER2 overexpression in breast cancer. Breast Cancer, 8(1):45-51


Ma YS, Weng SW, Lin MW, et al. (2012). Anti-tumor effects of emodin on LS1034 human colon cancer cells in vitro and in vivo: Roles of apoptotic cell death and LS1034 tumor xenografts model. Food Chem Toxicol, 50(5): 1271–1278. doi: 10.1016/j.fct.2012.01.033.


Muto A, Hori M, Sasaki Y, et al. (2007). Emodin has a cytotoxic activity against human multiple myeloma as a Janus-activated kinase 2 inhibitor. Mol Cancer Ther. doi: 10.1158/1535-7163.MCT-06-0605.


Narender T, Sukanya P, Sharma K, et al. (2013). Preparation of novel anti-proliferative emodin derivatives and studies on their cell-cycle arrest, caspase dependent apoptosis and DNA binding interaction. Phytomedicine, 20(10):890-896.


Shrimali D, Shanmugam MK, Kumar AP, et al. (2013). Targeted abrogation of diverse signal transduction cascades by emodin for the treatment of inflammatory disorders and cancer. Cancer Lett:S0304-3835(13)00598-3. doi: 10.1016/j.canlet.2013.08.023.


Wei WT, Lin SZ, Liu DL, Wang ZH. (2013). The distinct mechanisms of the anti-tumor activity of emodin in different types of cancer (Review). Oncol Rep. doi: 10.3892/or.2013.2741.


Yu XW, Yang RZ. (2013). Effects of crude rhubarb on serum IL-15 and IL-18 levels in patients with severe acute pancreatitis. An Hui Yi Xue, 34(3): 285-287.

Costunolide and Dehydrocostus Lactone

Cancers:
Breast, cervical., lung, ovarian, bladder, leukemia, prostate, gastric

Action: Anti-inflammatory, pro-oxidative, MDR, lymphangiogenesis inhibitor, anti-metastasis, mediates apoptosis, anti-metastatic

Components of Saussurea lappa Clarke, Vladimiria souliei (Franchet) Lingelsheim (Compositae)

Breast cancer; Anti-metastatic

It was found that costunolide inhibited the growth and telomerase activity of MCF-7 and MDA-MB-231 cells in a concentration- and time-dependent manner. The expression of hTERT mRNA was also inhibited but hTR mRNA was not. In addition, the bindings of transcription factors in hTERT promoters were significantly decreased in both cells by the treatment of costunolide. These results suggest that costunolide inhibited the growth of both MCF-7 and MDA-MB-231 cells and this effect was mediated at least in part by a significant reduction in telomerase activity (Choi et al., 2005).

Breast Cancer

Costunolide has been demonstrated to suppress tumor growth and metastases of MDA-MB-231 highly metastatic human breast cancer cells via inhibiting TNF-α induced NF-kB activation. Costunolide also inhibited MDA-MB-231 tumor growth and metastases without affecting body weights in the in vivo mouse orthotopic tumor growth assays.

In addition, costunolide inhibited in vitro TNF-α induced invasion and migration of MDA-MB-231 cells. Costunolide further suppressed TNF-α induced NF-kB signaling activation, resulting in a reduced expression of MMP-9, a well-known NF-kB-dependent gene to mediate breast cancer cell growth and metastases. Taken together, these results suggest that SLC and its derivative costunolide suppress breast cancer growth and metastases by inhibiting TNF-α induced NF-k B activation, suggesting that costunolide as well as SLC may be promising anti-cancer drugs, especially for metastatic breast cancer (Choi et al., 2013).

Several Chinese herbs, namely, Herba Taraxaci Mongolici (Pu Gong Ying), Radix Glycyrrhizae Uralensis (Gan Cao), Radix Bupleuri (Chai Hu), Radix Aucklandiae Lappae/ Radix Aucklandiae Lappae (Mu Xiang), Fructus Trichosanthis (Gua Lou) and Rhizoma Dioscoreae Bulbiferae (Huang Yao Zi) are frequently used in complex traditional Chinese medicine formulas for breast hyperplasia and breast tumor therapy.

The pharmacological effects of these Chinese herbs are all described as 'clearing heat-toxin and resolving masses' in traditional use. A bioactivity-oriented screening platform, which was based on a human breast cancer MCF-7 cellular model was developed to rapidly screen the 6 Chinese herbs. Two potential anti-breast cancer compounds, which were costunolide (Cos) and dehydrocostus lactone (Dehy), were identified in Radix Aucklandiae Lappae.

Combination of the two compounds showed a synergism on inhibiting the proliferation of MCF-7 cells in vitro, which exhibits a potential application prospect for breast cancer therapy. This bioactivity-oriented screening strategy is rapid, economical., reliable and specific for screening potential anti-breast cancer compounds in traditional Chinese medicines (Peng et al., 2013).

Dehydrocostuslactone (DHE) suppresses the expression of cyclin D, cyclin A, cyclin-dependent kinase 2, and cdc25A and increases the amount of p53 and p21, resulting in G(0)/G(1)-S phase arrest in MCF-7 cells. In contrast, DHE caused S-G(2)/M arrest by increasing p21 expression and chk1 activation and inhibiting cyclin A, cyclin B, cdc25A, and cdc25C expression in MDA-MB-231 cells. Reduction of SOCS-1 and SOCS-3 expression by small interfering RNA inhibits DHE-mediated signal transducer and activator of transcription-3 inhibition, p21 up-regulation, and cyclin-dependent kinase 2 blockade, supporting the hypothesis that DHE inhibits cell-cycle progression and cell death through SOCS-1 and SOCS-3.

Significantly, animal studies have revealed a 50% reduction in tumor volume after a 45-day treatment period. Taken together, this study provides new insights into the molecular mechanism of the DHE action that may contribute to the chemoprevention of breast cancer (Kuo et al., 2009).

ER- Breast Cancer

Costunolide induced apoptosis through the extrinsic pathway, including the activation of Fas, caspase-8, caspase-3, and degradation of PARP. However, it did not have the same effect on the intrinsic pathway as revealed by analysis of mitochondrial membrane potential (Δψ m) with JC-1 dye and expression of Bcl2 and Bax proteins level.

Furthermore, costunolide induced cell-cycle arrest in the G2/M phase via decrease in Cdc2, cyclin B1 and increase in p21WAF1 expression, independent of p53 pathway in p53-mutant MDA-MB-231 cells, and increases Cdc2-p21WAF1 binding/

Through this study it was confirmed that costunolide induces G2/M cell-cycle arrest and apoptotic cell death via extrinsic pathway in MDA-MB-231 cells, suggesting that it could be a promising anti-cancer drug especially for ER negative breast cancer (Choi et al., 2012).

Bladder Cancer

Costunolide, a member of sesquiterpene lactone family, possesses potent anti-cancer properties. The effects of costunolide were investigated on the cell viability and apoptosis in human bladder cancer T24 cells. Treatment of T24 cells with costunolide resulted in a dose-dependent inhibition of cell viability and induction of apoptosis, which was associated with the generation of ROS and disruption of mitochondrial membrane potential (Δψm).

These effects were significantly blocked when the cells were pre-treated with N-acetyl- cysteine (NAC), a specific ROS inhibitor. Exposure of T24 cells to costunolide was also associated with increased expression of Bax, down-regulation of Bcl-2, and of   survivin and significant activation of caspase-3, and its downstream target PARP. These findings provide the rationale for further in vivo and clinical investigation of costunolide against human bladder cancer (Rasul et al., 2013).

Sarcomas; MDR

Human soft tissue sarcomas represent a rare group of malignant tumors that frequently exhibit chemotherapeutic resistance and increased metastatic potential following unsuccessful treatment.

The effects on cell proliferation, cell-cycle distribution, apoptosis induction, and ABC transporter expression were analyzed. Cells treated with costunolide showed no changes in cell-cycle, little in caspase 3/7 activity, and low levels of cleaved caspase-3 after 24 and 48 hours. Dehydrocostus lactone caused a significant reduction of cells in the G1 phase and an increase of cells in the S and G2/M phase. Moreover, it led to enhanced caspase 3/7 activity, cleaved caspase-3, and cleaved PARP indicating apoptosis induction.

These data demonstrate that dehydrocostus lactone affects cell viability, cell-cycle distribution and ABC transporter expression in soft tissue sarcoma cell lines. Furthermore, it led to caspase 3/7 activity as well as caspase-3 and PARP cleavage, which are indicators of apoptosis. Therefore, this compound may be a promising lead candidate for the development of therapeutic agents against drug-resistant tumors (Kretschmer et al., 2013).

Leukemia, Lung Cancer

Costunolide, an active compound isolated from the stem bark of Magnolia sieboldii, has been found to induce apoptosis via reactive oxygen species (ROS) and Bcl-2-dependent mitochondrial permeability transition in human leukemia cells. Mitogen-activated protein kinases (MAPKs) were investigated for their involvement in the costunolide-induced apoptosis in human promonocytic leukemia U937 cells.

Treatment with costunolide resulted in the significant activation of c-Jun N-terminal kinase (JNK), but not of extracellular-signal-related kinase (ERK1/2) or p38. In vitro kinase assays showed that JNK activity was low in untreated cells but increased dramatically after 30 minutes of costunolide treatment. U937 cells co-treated with costunolide and sorbitol, a JNK activator, exhibited higher levels of cell death. In addition, inhibition of the JNK pathway using a dominant-negative mutation of c-jun and JNK inhibitor SP600125, significantly prevented costunolide-induced apoptosis.

Furthermore, pre-treatment with the anti-oxidant NAC (N-acetyl-L-cysteine) blocked the costunolide-stimulated activation of JNK while the overexpression of Bcl-2 failed to reverse JNK activation. These results indicate that costunolide-induced JNK activation acts downstream of ROS but upstream of Bcl-2, and suggest that ROS-mediated JNK activation plays a key role in costunolide-induced apoptosis. Moreover, the administration of costunolide (intraperitoneally once a day for 7 days) significantly suppressed tumor growth and increased survival in 3LL Lewis lung carcinoma-bearing model (Choi et al., 2009).

Prostate Cancer

Several pharmacological and biochemical assays were used to characterize the apoptotic-signaling pathways of costunolide in prostate cancer cells. Costunolide showed effective anti-proliferative activity against hormone dependent (LNCaP) and independent (PC-3 and DU-145) prostate cancer cells (ATCC¨) by sulforhodamine B assay, clonogenic test and flow cytometric analysis of carboxyfluorescein succinimidyl ester labeling. In PC-3 cells data showed that costunolide induced a rapid overload of nuclear Ca(2+), DNA damage response and ATR phosphorylation.

This indicated the crucial role of intracellular Ca(2+) mobilization and thiol depletion but not of reactive oxygen species production in apoptotic signaling. Data suggest that costunolide induces the depletion of intracellular thiols and overload of nuclear Ca(2+) that cause DNA damage and p21 up-regulation. The association of p21 with the cyclin dependent kinase 2/cyclin E complex blocks cyclin dependent kinase 2 activity and inhibits Rb phosphorylation, leading to G1 arrest of the cell-cycle and subsequent apoptotic cell death in human prostate cancer cells (Hsu et al., 2011).

Gastric Cancer, Prostate Cancer

Radix Aucklandiae Lappae/Saussurea lappa has been used in Chinese traditional medicine for the treatment of abdominal pain, tenesmus, nausea, and cancer; previous studies have shown that S. lappa also induces G(2) growth arrest and apoptosis in gastric cancer cells. The effects of hexane extracts of S. lappa (HESLs) on the migration of DU145 and TRAMP-C2 prostate cancer cells were investigated.

The active compound, dehydrocostus lactone (DHCL), in fraction 7 dose-dependently inhibited the basal and EGF-induced migration of prostate cancer cells. HESL and DHCL reduced matrix metalloproteinase (MMP)-9 and tissue inhibitor of metalloproteinase (TIMP)-1 secretion but increased TIMP-2 levels in both the absence and presence of EGF. These results demonstrate that the inhibition of MMP-9 secretion and the stimulation of TIMP-2 secretion contribute to reduced migration of DU145 cells treated with HESL and DHCL.

This indicates that HESL containing its active principle, DHCL, has potential as an anti-metastatic agent for the treatment of prostate cancer (Kim et al., 2012).

Anti-metastatic

Lymphangiogenesis inhibitors from crude drugs used in Japan and Korea were investigated for their impact on metastasis. The three crude drugs Saussureae Radix, Psoraleae Semen and Aurantti Fructus Immaturus significantly inhibited the proliferation of temperature-sensitive rat lymphatic endothelial (TR-LE) cells in vitro.

Among isolated compounds, several compounds; costunolide, dehydrocostus lactone, psoracorylifol D, bavachinin, bakuchiol, showed an inhibitory effect on the proliferation and the capillary-like tube formation of TR-LE cells. In addition, all compounds showed selective inhibition of the proliferation of TR-LE cells compared to Hela and Lewis lung carcinoma (LLC) cells.

These compounds might offer clinical benefits as lymphangiogenesis inhibitors and may be good candidates for novel anti-cancer and anti-metastatic agents (Jeong et al., 2013).

Ovarian Cancer, MDR

The apoptosis-inducing effect of costunolide, a natural sesquiterpene lactone, was studied in platinum-resistant human ovarian cancer cells relative to cisplatin.

The MTT assay for cell viability, PI staining for cell-cycle profiling, and annexin V assay for apoptosis analysis were performed. Costunolide induced apoptosis of platinum-resistant cells in a time and dose-dependent manner and suppressed tumor growth in the SKOV3 (PT)-bearing mouse model. In addition, costunolide triggered the activation of caspase-3, caspase-8, and caspase-9. Pre-treatment with caspase inhibitors neutralized the pro-apoptotic activity of costunolide. We further demonstrated that costunolide induced a significant increase in intracellular reactive oxygen species (ROS). Moreover, costunolide synergized with cisplatin to induce cell death in platinum-resistant ovarian cancer cells.

Data suggests that costunolide, alone or in combination with cisplatin, may be of therapeutic potential in platinum-resistant ovarian cancers (Yang, Kim, Lee, & Choi, 2011).

Anti-inflammatory, Anti-oxidant, Mediates Apoptosis

Cheon et al. (2013) found that costunolide significantly inhibited RANKL-induced BMM differentiation into osteoclasts in a dose-dependent manner without causing cytotoxicity. Costunolide did not regulate the early signaling pathways of RANKL, including the mitogen-activated protein kinase and NF-κB pathways.

However, costunolide suppressed nuclear factor of activated T-cells, cytoplasmic 1 (NFATc1) expression via inhibition of c-Fos transcriptional activity without affecting RANKL-induced c-Fos expression. The inhibitory effects of costunolide were rescued by overexpression of constitutively active (CA)-NFATc1. Taken together, these results suggest that costunolide inhibited RANKL-induced osteoclast differentiation by suppressing RANKL-mediated c-Fos transcriptional activity.

References

Cheon YH, Song MJ, Kim JY, Kwak SC, Park JH, Lee CH, Kim JJ, Kim JY, Choi MK, Oh J, Kim YC, Yoon KH., Kwak HB, Lee MS. (2013). Costunolide inhibits osteoclast differentiation by suppressing c-Fos transcriptional activity. Phytotherapy, July, (6). doi: 10.1002/ptr.5034.

Choi SH, Im E, Kang HK, et al. (2005). Inhibitory effects of costunolide on the telomerase activity in human breast carcinoma cells. Cancer Lett, 227(2):153-62.


Choi JH, Lee KT. (2009). Costunolide-induced apoptosis in human leukemia cells: involvement of c-jun N-terminal kinase activation. Biol Pharm Bull, 32(10):1803-8.


Choi YK, Seo HS, Choi HS, et al. (2012). Induction of Fas-mediated extrinsic apoptosis, p21WAF1-related G2/M cell-cycle arrest and ROS generation by costunolide in estrogen receptor-negative breast cancer cells, MDA-MB-231. Mol Cell Biochem, 363(1-2):119-28. doi: 10.1007/s11010-011-1164-z.


Choi YK, Cho S-G, Woo S-M, et al. (2013). Saussurea lappa Clarke-Derived Costunolide Prevents TNF α-Induced Breast Cancer Cell Migration and Invasion by Inhibiting NF-κ B Activity. Evidence-Based Complementary and Alternative Medicine. doi:10.1155/2013/936257.


Hsu JL, Pan SL, Ho YF, Het al. (2011). Costunolide induces apoptosis through nuclear calcium2+ overload and DNA damage response in human prostate cancer. The Journal of Urology, 185(5):1967-74. doi: 10.1016/j.juro.2010.12.091.


Jeong D, Watari K, Shirouzu T, et al. (2013). Studies on lymphangiogenesis inhibitors from Korean and Japanese crude drugs. Biol Pharm Bull, 36(1):152-7.


Kim EJ, Hong JE, Lim SS, et al. (2012). The hexane extract of Saussurea lappa and its active principle, dehydrocostus lactone, inhibit prostate cancer cell migration. J Med Food, 15(1):24-32. doi: 10.1089/jmf.2011.1735.


Kretschmer N, Rinner B, Stuendl N, et al. (2012). Effect of costunolide and dehydrocostus lactone on cell-cycle, apoptosis, and ABC transporter expression in human soft tissue sarcoma cells. Planta Med, 78(16):1749-56. doi: 10.1055/s-0032-1315385.


Kuo PL, Ni WC, Tsai EM, Hsu YL. (2009). Dehydrocostuslactone disrupts signal transducers and activators of transcription 3 through up-regulation of suppressor of cytokine signaling in breast cancer cells. Mol Cancer Ther, 8(5):1328-39. doi: 10.1158/1535-7163.MCT-08-0914.


Peng ZX, Wang Y, Gu X, Wen YY, Yan C. (2013). A platform for fast screening potential anti-breast cancer compounds in traditional Chinese medicines. Biomed Chromatogr. doi: 10.1002/bmc.2990.


Rasul A, Bao R, Malhi M, et al. (2013). Induction of apoptosis by costunolide in bladder cancer cells is mediated through ROS generation and mitochondrial dysfunction. Molecules, 18(2):1418-33. doi: 10.3390/molecules18021418.


Yang YI, Kim JH, Lee KT, & Choi JH. (2011). Costunolide induces apoptosis in platinum-resistant human ovarian cancer cells by generating reactive oxygen species. Gynecologic Oncology, 123(3), 588-96. doi: 10.1016/j.ygyno.2011.08.031.

Luteolin

Cancer: Colorectal., ovarian, pancreatic

Action: Anti-inflammatory, immunomodulatory, radio-sensitizer, 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, which 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, 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.

Colon Cancer

Luteolin inhibited cyclin-dependent kinase (CDK)4 and CDK2 activity, resulting in G1 arrest with a concomitant decrease of phosphorylation of retinoblastoma protein. 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 also 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. 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.

Radio-protective

The aqueous extract of Perilla frutescens 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 Perilla 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 macrophages with luteolin, luteolin-7-glucoside, quercetin, and the isoflavonoid genistein inhibited both the LPS-stimulated TNF-α and interleukin-6 release, whereas eriodictyol and hesperetin only inhibited TNF-α release. 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-α release, respectively. Moreover, luteolin inhibited LPS-induced phosphorylation of Akt. Treatment of macrophages with LPS resulted in increased IκB-α phosphorylation and reduced the levels of IκB-α. Pre-treatment of cells with luteolin abolished the effects of LPS on IκB-α.

Xagorari et al. (2001) concluded that luteolin inhibits protein tyrosine phosphorylation, nuclear factor-κB-mediated gene expression and pro-inflammatory cytokine production in murine macrophages.

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. Consistent with this finding, luteolin did not inhibit LPS-induced degradation of IκB-α but inhibited JNK phosphorylation.

Luteolin consumption reduced LPS-induced IL-6 in plasma 4 hours after injection. Furthermore, luteolin decreased the induction of IL-6 mRNA by LPS in the hippocampus but not in the cortex or cerebellum. 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).

Immunostimulatory and Anti-inflammatory

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. Lut had a marked inhibitory effect on the inflammatory exudation, but did not affect the number of leucocytes. 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

Luteolin dose-dependently inhibited the expression and production of those inflammatory genes and mediators in macrophages stimulated with lipopolysaccharide (LPS). Semi-quantitative reverse-transcription polymerase chain reaction (RT-PCR) assay further confirmed the suppression of LPS-induced TNF- α, IL-6, iNOS and COX-2 gene expression by luteolin at a transcriptional level. Luteolin also reduced the DNA binding activity of nuclear factor-kappa B (NF-κB) in LPS-activated macrophages.

In addition, luteolin significantly inhibited the LPS-induced DNA binding activity of activating protein-1 (AP-1). It was also found that luteolin attenuated the LPS-mediated protein kinase B (Akt) and IKK phosphorylation, as well as reactive oxygen species (ROS) production. In sum, these data suggest that, by blocking NF-κB 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).

Pancreatic Cancer; Chemo-enhancing

Simultaneous treatment or pre-treatment (0, 6, 24 and 42h) of flavonoids and chemotherapeutic drugs and various concentrations (0-50µM) were assessed using the MTS cell proliferation assay. Pre-treatment for 24 hours with 13µM of either Apigenin or Luteolin, followed by Gem for 36 h 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, 36 hours). Lut (15µM, 24 hours) pre-treatment followed by Gem (10µM, 36h), significantly decreased protein expression of nuclear GSK-3β and 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

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. The critical role of c-Jun NH(2)-terminal kinase (JNK) was identified in regulation of p53 protein stability: luteolin activates JNK, and JNK then stabilizes p53 via phosphorylation, leading to reduced ubiquitination and proteasomal degradation.

An in vivo nude mice xenograft model confirmed that luteolin enhanced the cancer therapeutic activity of cisplatin via p53 stabilization and accumulation. In summary, data from this study reveal a novel molecular mechanism involved in the anti-cancer effects of luteolin and support its potential clinical application as a chemo-sensitizer in cancer therapy (Shi et al., 2007).

Breast Cancer; Chemo-sensitzer

Luteolin is a flavonoid that has been identified in many plant tissues and exhibits chemo-preventive or chemo-sensitizing properties against human breast cancer. However, the oncogenic molecules in human breast cancer cells that are inhibited by luteolin treatment have not been identified.

Relatively high levels of cyclin E2 (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- κ B 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.

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.

Ellagic Acid

Cancer:
Pancreatic, prostate, ovarian, breast, bladder, lymphoma, oral., melanoma

Action: Anti-cancer, induces apoptosis, promoted ROS and Ca2+ productions

Ellagic acid (EA) is a polyphenol compound widely found in fruits such as berries, walnuts, pecans, pomegranate, cranberries, and longan. It is well known to have a free radical scavenging activity and has been approved in Japan as an 'existing food additive' for anti-oxidative purposes (HHLW, 1996). In vitro evidence revealed that 100µM EA represented little toxic effect on human normal cells (Losso et al., 2004; Larrosa et al., 2006). A subchronic toxicity study further demonstrated that orally feeding EA (9.4, 19.1, 39.1g/kg b.w., resp.) could not induce mortality or treatment-related clinical signs throughout the experimental period on F344 rats (Tasaki et al., 2008), indicating the low toxicity of EA to mammalians. Furthermore, EA exhibits potent anti-cancer and anti-carcinogenesis activities towards breast, colorectal., oral., prostate (Losso et al., 2004; Larrosa et al., 2006; Malik et al., 2011), pancreatic (Edderkaoui et al., 2008), bladder (Li et al., 2005), neuroblastoma (Fjaeraa et al., 2009), melanoma (Kim et al., 2009), and lymphoma cells (Mishra et al., 2011).

Pancreatic Cancer

Edderkaoui et al. (2008) show that ellagic acid, a polyphenolic compound in fruits and berries, at concentrations 10 to 50 mmol/L stimulates apoptosis in human pancreatic adenocarcinoma cells. Ellagic acid stimulates the mitochondrial pathway of apoptosis associated with mitochondrial depolarization, cytochrome C release, and the downstream caspase activation. Ellagic acid does not directly affect mitochondria. Ellagic acid dose-dependently decreased NF-kappa B binding activity.

Furthermore, inhibition of NF-kappa B activity using IkB wild type plasmid prevented the effect of ellagic acid on apoptosis.

Pancreatic Cancer (PANC-1) cells were injected subcutaneously into Balb c nude mice, and tumor-bearing mice were treated with ellagic acid (EA). Treatment of PANC-1 xenografted mice with EA resulted in significant inhibition in tumor growth which was associated with suppression of cell proliferation and caspase-3 activation, and induction of PARP cleavage. EA also reversed epithelial to mesenchymal transition by up-regulating E-cadherin and inhibiting the expression of Snail, MMP-2 and MMP-9.

These data suggest that EA can inhibit pancreatic cancer growth, angiogenesis and metastasis by suppressing Akt, Shh and Notch pathways. In view of the fact that EA could effectively inhibit human pancreatic cancer growth by suppressing Akt, Shh and Notch pathways, our findings suggest that the use of EA would be beneficial for the management of pancreatic cancer (Zhao et al., 2013).

Ovarian Cancer

Ovarian carcinoma ES-2 and PA-1 cells were treated with EA (10~100  µ M) and assessed for viability, cell-cycle, apoptosis, anoikis, autophagy, and chemosensitivity to doxorubicin and their molecular mechanisms. EA inhibited cell proliferation in a dose- and time-dependent manner by arresting both cell lines at the G1 phase of the cell-cycle, which were from elevating p53 and Cip1/p21 and decreasing cyclin D1 and E levels. EA also induced caspase-3-mediated apoptosis by increasing the Bax :  Bcl-2 ratio and restored anoikis in both cell lines.

The enhancement of apoptosis and/or inhibition of autophagy in these cells by EA assisted the chemotherapy efficacy. The results indicated that EA is a potential novel chemoprevention and treatment assistant agent for human ovarian carcinoma Chung et al., 2013).

Prostate Cancer; AR+

In the present study, Pitchakarn et al. (2013) investigated anti-invasive effects of ellagic acid (EA) in androgen-independent human (PC-3) and rat (PLS10) prostate cancer cell lines in vitro. The results indicated that non-toxic concentrations of EA significantly inhibited the motility and invasion of cells examined in migration and invasion assays. They found that EA significantly reduced proteolytic activity of collagenase/gelatinase secreted from the PLS-10 cell line. Collagenase IV activity was also concentration-dependently inhibited by EA. These results demonstrated that EA has an ability to inhibit invasive potential of prostate cancer cells through action on protease activity.

Breast Cancer

The role of estrogen (E2) in regulation of cell proliferation and breast carcinogenesis is well-known. Recent reports have associated several miRNAs with estrogen receptors in breast cancers. Investigation of the regulatory role of miRNAs is critical for understanding the effect of E2 in human breast cancer, as well as developing strategies for cancer chemoprevention.

In this study Munagala et al. (2013) used the well-established ACI rat model that develops mammary tumors upon E2 exposure and identified a 'signature' of 33 significantly modulated miRNAs during the process of mammary tumorigenesis. Several of these miRNAs were altered as early as 3 weeks after initial E2 treatment and their modulation persisted throughout the mammary carcinogenesis process, suggesting that these molecular changes are early events. This is the first systematic study examining the changes in miRNA expression associated with E2 treatment in ACI rats as early as 3weeks until tumor time point. The effect of a chemo-preventive agent, ellagic acid in reversing miRNAs modulated during E2-mediated mammary tumorigenesis is also established. These observations provide mechanistic insights into the new molecular events behind the chemo-preventive action of ellagic acid and treatment of breast cancer.

Bladder Cancer

To investigate the effects of ellagic acid on the growth inhibition of TSGH8301 human bladder cancer cells in vitro, cells were incubated with various doses of ellagic acid for different time periods. Results indicated that ellagic acid induced morphological changes, decreased the percentage of viable cells through the induction of G0/G1 phase arrest and apoptosis, and also showed that ellagic acid promoted ROS and Ca2+ productions and decreased the level of ΔΨm and promoted activities of caspase-9 and -3.

On the basis of these observations, Ho et al (2013) suggest that ellagic acid induced cytotoxic effects for causing a decrease in the percentage of viable cells via G0/G1 phase arrest and induction of apoptosis in TSGH8301 cells.

Lymphoma

Protein Kinase C (PKC) isozymes are key components involved in cell proliferation and their over activation leads to abnormal tumor growth. PKC follows signaling pathway by activation of downstream gene NF-kB and early transcription factor c-Myc. Over activation of NF-kB and c-Myc gene are also linked with unregulated proliferation of cancer cells.

Therefore any agent which can inhibit the activation of Protein kinase C, NF-kB and c-Myc may be useful in reducing cancer progression. The role of ellagic acid was tested in regulation of tumor suppressor gene Transforming growth factor-β1 (TGF-β1). DL mice were treated with three different doses (40, 60 and 80 mg/kg body weight) of ellagic acid. Ascites cells of mice were used for the experiments. Ellagic acid administration to DL mice decreased oxidative stress by reducing lipid peroxidation.

The anti-carcinogenic action of ellagic acid was also confirmed by up-regulation of TGF-β1 and down-regulation of c-Myc. Lymphoma prevention by ellagic acid is further supported by decrease in cell proliferation, cell viability, ascites fluid accumulation and increase in life span of DL mice. All these findings suggest that ellagic acid prevents the cancer progression by down- regulation of PKC signaling pathway leading to cell proliferation (Mishra et al., 2013).

References

Chung YC, Lu LC, Tsai MH, et al. (2013). The inhibitory effect of ellagic Acid on cell growth of ovarian carcinoma cells. Evid Based Complement Alternat Med, 2013(2013):306705. doi: 10.1155/2013/306705.


Edderkaoui M, Odinokova I, Ohno I, et al. (2008). Ellagic acid induces apoptosis through inhibition of nuclear factor κ B in pancreatic cancer cells. World Journal of Gastroenterology, 14(23):3672–3680.


Fjaeraa C, NŒnberg E. (2009). Effect of ellagic acid on proliferation, cell adhesion and apoptosis in SH-SY5Y human neuroblastoma cells. Biomedicine and Pharmacotherapy, 63(4):254–261.


HHLW (Ministry of Health, Labor and Welfare of Japan). (1996). List of Existing Food Additives, Notification No. 120 of the Ministry of Health and Welfare.


Ho CC, Huang AC, Yu CS, Lien JC, et al. (2013). Ellagic acid induces apoptosis in tsgh8301 human bladder cancer cells through the endoplasmic reticulum stress- and mitochondria-dependent signaling pathways. Environ Toxicol. doi: 10.1002/tox.21857.


Kim S, Liu Y, Gaber MW, Bumgardner JD, Haggard WO, Yang Y. (2009). Development of chitosan-ellagic acid films as a local drug delivery system to induce apoptotic death of human melanoma cells. Journal of Biomedical Materials Research, 90(1):145–155.


Larrosa M, Tomás-Barberán FA, Espín JC. (2006). The dietary hydrolysable tannin punicalagin releases ellagic acid that induces apoptosis in human colon adenocarcinoma Caco-2 cells by using the mitochondrial pathway. Journal of Nutritional Biochemistry, 17(9):611–625.


Li TM, Chen GW, Su CC, et al. (2005). Ellagic acid induced p53/p21 expression, G1 arrest and apoptosis in human bladder cancer T24 cells. Anti-cancer Research, 25(2 A):971–979.


Losso JN, Bansode RR, Trappey A, II, Bawadi HA, Truax R. (2004). In vitro anti-proliferative activities of ellagic acid. Journal of Nutritional Biochemistry, 15(11):672–678.


Mishra S, Vinayak M. (2013). Ellagic acid checks lymphoma promotion via regulation of PKC signaling pathway. Mol Biol Rep, 40(2):1417-28. doi: 10.1007/s11033-012-2185-8.


Malik A, Afaq S, Shahid M, Akhtar K, Assiri A. (2011). Influence of ellagic acid on prostate cancer cell proliferation: a caspase-dependent pathway. Asian Pacific Journal of Tropical Medicine, 4(7):550–555.


Mishra S, Vinayak M. (2011). Anti-carcinogenic action of ellagic acid mediated via modulation of oxidative stress regulated genes in Dalton lymphoma bearing mice. Leukemia and Lymphoma, 52(11):2155–2161.


Munagala R, Aqil F, Vadhanam MV, Gupta RC. (2013). MicroRNA 'signature' during estrogen-mediated mammary carcinogenesis and its reversal by ellagic acid intervention. Cancer Lett, S0304-3835(13)00462-X. doi: 10.1016/j.canlet.2013.06.012.


Pitchakarn P, Chewonarin T, Ogawa K, et al. (2013). Ellagic Acid inhibits migration and invasion by prostate cancer cell lines. Asian Pac J Cancer Prev, 14(5):2859-63.


Tasaki M, Umemura T, Maeda M, et al. (2008). Safety assessment of ellagic acid, a food additive, in a subchronic toxicity study using F344 rats. Food and Chemical Toxicology, 46(3):1119–1124.


Zhao M, Tang SN, Marsh JL, et al. (2013). Ellagic acid inhibits human pancreatic cancer growth in Balb c nude mice. Cancer Letters, 337(2):210–217

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.

Parthenolide

Cancer:
Myeloma, ovarian adenocarcinoma, breast, acute myelogenous leukemia, epithelial-to-mesenchymal transition (EMT)

Action: Anti-cancer, anti-inflammatory, inhibits NF-κB, promotes apoptosis

Inhibits NF-κB & Promotes Apoptosis

Parthenolide, a sesquiterpene lactone derived from the leaves of feverfew (Tanacetum parthenium (L.)), is considered a main bioactive component of this herb. Feverfew has been used orally or as an infusion for the treatment of migraine, arthritis, fever, and stomachache. Besides its anti-inflammatory and anti-migraine properties, parthenolide also shows anti-cancer activities in a variety of cell lines. It contains an alpha-methylene-gamma-lactone ring and an epoxide moiety which are able to interact with nucleophilic sites of biologically important molecules.

Parthenolide modulates multiple targets, thereby contributing to its various in vitro and in vivo effects. Inhibition of NF-kappaB activity, constitutive in many types of cancers via either interaction with IKK or more directly with the p65 subunit of NF-kappaB, is considered one of the main mechanisms of its action. In addition, inhibition of STAT and MAP kinase activities and the induction of sustained JNK activity as well as p53 activity via influencing MDM2 and HDAC1 levels lead to an increased susceptibility of cancer cells to chemo- and radio- therapy. At the epigenetic level, parthenolide reduces HDAC1 level and, by inhibiting DNMT2 activity, induces global hypomethylation of DNA, which can restore the expressions of some suppressor genes.

Moreover, this compound reduces the cellular level of GSH in cancer cells, followed by ROS accumulation and apoptosis. A unique property of parthenolide is its ability to induce cell death mainly in cancer cells, while sparing healthy ones and it also protects normal cells from UVB and oxidative stress. More remarkably, it seems to have the potential to target some cancer stem cells. Its wide array of biological activity and low toxicity make parthenolide a very promising drug with multi-pharmacological potential, largely dependent on the cellular context (Koprowska et al., 2010).

Multiple Myeloma

It has been shown that parthenolide is a potent anti-MM-CSC agent. Multiple myeloma (MM) is an incurable plasma cell malignancy where nearly all patients succumb to a relapse. The current preclinical models of MM target the plasma cells, constituting the bulk of the tumor, leaving the cancer stem cells to trigger a relapse. It demonstrated preferential toxicity toward MM-CSCs over non-tumorigenic MM cells. Addition of the bone marrow stromal compartment abrogated andrographolide activity while having no effect on parthenolide cytoxicity. It hence has anti-CSC activity in myeloma, suggesting that it has the potential to improve the survival of patients with MM by eliminating the relapse-causing MM-CSCs (Gunn et al., 2011).

Acute Myelogenous Leukemia

Parthenolide (PTL), a naturally occurring small molecule, induces robust apoptosis in primary human acute myelogenous leukemia (AML) cells and blast crisis CML (bcCML) cells while sparing normal hematopoietic cells. Furthermore, analysis of progenitor cells using in vitro colony assays, as well as stem cells using the non-obese diabetic/severe combined immunodeficient (NOD/SCID) xenograft model, show that PTL also preferentially targets AML progenitor and stem cell populations.

Notably, in comparison to the standard chemotherapy drug cytosine arabinoside (Ara-C), PTL is much more specific to leukemia cells. The molecular mechanism of PTL-mediated apoptosis is strongly associated with inhibition of nuclear factor κB (NF-κB), pro-apoptotic activation of p53, and increased reactive oxygen species (ROS) (Guzman, et al., 2005).

PTL is known to be a potent inhibitor of NF-κB (Bork et al., 1997). The mechanism of NF-κB down-regulation appears to occur via inhibition of the IKK complex. Guzman et al. (2005) observed strong inhibition of NF-κB in primary AML cells and speculate that this activity contributes to the efficacy of PTL. However, previous genetic studies using a dominant-negative repressor of NF-κB activity have shown that inhibition of NF-κB alone is not sufficient to mediate the robust cell death observed with PTL. Rather, blockade of the NF-κB pathway appears to sensitize primary AML cells to death and induces a relatively slow spontaneous apoptosis (~50% cell death in 36 h) (Guzman et al., 2002). Similarly, studies by Romano et al. (2000) showed that treatment of primary AML blasts with NF-κB decoy oligonucleotides was not sufficient to induce a strong apoptotic response.

Consequently, PTL must be affecting other pathways relevant to AML-specific survival. One such pathway appears to be mediated by the activity of p53. PTL induced rapid up-regulation of p53 protein with concomitant phosphorylation on serine (Woynarowski & Konopa, 1981).

Ovarian Carcinoma

Results suggest that parthenolide may induce apoptotic cell death in ovarian carcinoma cell lines by activating the mitochondrial pathway and the caspase-8- and Bid-dependent pathways. The apoptotic effect of parthenolide appears to be mediated by the formation of reactive oxygen species and the depletion of GSH. Parthenolide might be beneficial in the treatment of epithelial ovarian adenocarcinoma and combination therapy (Kwak et al., 2013).

Epithelial-to-Mesenchymal Transition (EMT)

Detyrosinated tubulin, a post-translational modification of α-tubulin and a hallmark of stable microtubules, has gained recent attention given its association with tumor progression, invasiveness, and chemoresistance. Also, epithelial-to-mesenchymal transition (EMT) promotes tubulin detyrosination through tubulin tyrosine ligase (TTL) suppression. Given the induction of EMT associated with inflammation and cancer progression, Whipple et al. (2013) tested anti-inflammatory nuclear factor-kappaB (NF-κB) inhibitors on a panel of human breast carcinoma cells to examine their effects on detyrosinated tubulin to identify more specific tubulin-directed anti-cancer treatments.

Breast Cancer

Sesquiterpene lactones, parthenolide and costunolide, selectively decrease detyrosinated tubulin independent of their inhibition of NF-κB. Live-cell scoring of suspended cells treated with parthenolide and costunolide show reduction in the frequency of microtentacles and inhibition of reattachment. Selective targeting of detyrosinated tubulin with parthenolide and costunolide can reduce McTN frequency and inhibit tumor cell reattachment. These actions are independent of their effects on NF-κB inhibition, presenting a novel anti-cancer property and therapeutic opportunity to selectively target a stable subset of microtubules in circulating tumor cells to reduce metastatic potential with less toxicity in breast cancer patients (Whipple et al., 2013).

References

Bork PM, Schmitz ML, Kuhnt M, Escher C, Heinrich M. (1997). Sesquiterpene lactone containing Mexican Indian medicinal plants and pure sesquiterpene lactones as potent inhibitors of transcription factor NF-kappaB. FEBS Lett, 402:85-90.


Gunn EJ, Williams JT, Huynh DT, et al. (2011). The natural products parthenolide and andrographolide exhibit anti-cancer stem cell activity in multiple myeloma. Leuk Lymphoma, 52(6):1085-97.


Guzman ML, Rossi RM, Karnischky L, et al. (2005). The sesquiterpene lactone parthenolide induces apoptosis of human acute myelogenous leukemia stem and progenitor cell. Blood, 105(11): 4163–4169. doi: 10.1182/blood-2004-10-4135.


Guzman ML, Swiderski CF, Howard DS, et al. (2002). Preferential induction of apoptosis for primary human leukemic stem cells. Proc Natl Acad Sci U S A, 99:16220-16225.


Koprowska K, Czyz M. (2010). Molecular mechanisms of parthenolide's action: Old drug with a new face. Postepy Hig Med Dosw, 64:100-14.


Kwak SW, Park ES, Lee CS. (2013). Parthenolide induces apoptosis by activating the mitochondrial and death receptor pathways and inhibits FAK-mediated cell invasion. Mol Cell Biochem.


Romano MF, Lamberti A, Bisogni R, et al. (2000). Enhancement of cytosine arabinoside-induced apoptosis in human myeloblastic leukemia cells by NF-kappa B/Rel-specific decoy oligodeoxynucleotides. Gene Ther, 7:1234-1237.


Whipple RA, Vitolo MI, Boggs AE, et al. (2013). Parthenolide and costunolide reduce microtentacles and tumor cell attachment by selectively targeting detyrosinated tubulin independent from NF- κ B inhibition. Breast Cancer Res,15(5):R83.


Woynarowski JM, Konopa J. (1981). Inhibition of DNA biosynthesis in HeLa cells by cytotoxic and anti-tumor sesquiterpene lactones. Mol Pharmacol,19:97-102.

Oridonin

Cancer: Prostate, acute promyelocytic leukemia, breast, non-small-cell lung (NSCL), Ehrlich ascites, P388 lymphocytic leukemia, colorectal., ovarian, esphageal

Action: Induces apoptosis

Oridonin is a tetracycline diterpenoid isolated from the plant Rabdosia rubescens (RR) [(Hemsl.). Hara (Lamiaceae)] (dong ling cao) is a Chinese medicinal herb used widely in provinces including Henan. The aerial parts of RR and other species of the same genus has been reported to have the functions of clearing “heat” and “toxicity”, nourishing “yin”, removing “blood stasis”, and relieving swelling. RR has been used to treat stomach-ache, sore throat and cough.

Gastric Cancer, Esophageal Cancer, Liver Cancer, Prostate Cancer

RR and its extracts have been shown to be able to suppress disease progress, reduce tumor burden, alleviate syndrome and prolong survival in patients with gastric carcinoma, esophageal., liver and prostate cancers (Tang & Eisenbrand, 1992). Interestingly, other Isodon plants including Isodon japonicus Hara (IJ) and I. trichocarpus (IT) are also applied as home remedies for similar disorders in Japan and Korea.

Induces Apoptosis

These reports suggest that Isodon plants should have at least one essential anti-tumor component. In the 1970s, a bitter tetracycline diterpenoid compound, oridonin, was isolated from RR, IJ, and IT separately, and was shown to be a potent apoptosis inducer in a variety of cancer cells (Fujita et al., 1970; Fujita et al., 1976; Henan Medical Institute, 1978; Fujita et al., 1988).

Anti-cancer

There is currently research being undertaken regarding the relationship between the chemical structure/modifications and the molecular mechanisms underlying its anti-cancer activity, such as suppression of tumor proliferation and induction of tumor cell death, and the cell signal transduction in anti-cancer activity of oridonin (Zhang et al., 2010).

Prostate Cancer, Breast Cancer, NSCLC, Leukemia, Glioblastoma

Oridonin has been found to effectively inhibit the proliferation of a wide variety of cancer cells including those from prostate (LNCaP, DU145, PC3), breast (MCF-7, MDA-MB231), non-small-cell lung (NSCL) (NCI-H520, NCI-H460, NCI-H1299) cancers, acute promyelocytic leukemia (NB4), and glioblastoma multiforme (U118, U138).

Oridonin induced apoptosis and G0/G1 cell-cycle arrest in LNCaP prostate cancer cells. In addition, expression of p21waf1 was induced in a p53-dependent manner. Taken together, oridonin inhibited the proliferation of cancer cells via apoptosis and cell-cycle arrest with p53 playing a central role in several cancer types which express the wild-type p53 gene. Oridonin may be a novel, adjunctive therapy for a large variety of malignancies (Ikezoe et al., 2003).

Breast Cancer; Anti-metastatic

According to the flow cytometric analysis, oridonin suppressed MCF-7 cell growth by cell-cycle arrest at the G2/M phase and caused accumulation of MDA-MB-231 cells in the Sub-G1 phase. The induced apoptotic effect of oridonin was further confirmed by a morphologic characteristics assay and TUNEL assay. Meanwhile, oridonin significantly suppressed MDA-MB-231 cell migration and invasion, decreased MMP-2/MMP-9 activation and inhibited the expression of Integrin β1 and FAK. In conclusion, oridonin inhibited growth and induced apoptosis in breast cancer cells, which might be related to DNA damage and activation of intrinsic or extrinsic apoptotic pathways. Moreover, oridonin also inhibited tumor invasion and metastasis in vitro possibly via decreasing the expression of MMPs and regulating the Integrin β1/FAK pathway in MDA-MB-231 cells (Wang et al., 2013).

Gastric Cancer

The inhibitory effect of oridonin on gastric cancer HGC-27 cells was detected using the 3-(4, 5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide assay. After treated with oridonin (0, 1.25, 2.5, 5 and 10 µg/mL), HGC-27 cells were collected for anexin V-phycoerythrin and 7-amino-actinomycin D double staining and tested by flow cytometric analysis, and oridonin- induced apoptosis in HGC-27 cells was detected.

Oridonin significantly inhibited the proliferation of HGC-27 cells in a dose- and time-dependent manner. The inhibition rates of HGC-27 treated with four different concentrations of oridonin for 24 h (1.25, 2.5, 5 and 10 µg/mL) were 1.78% ± 0.36%, 4.96% ± 1.59%, 10.35% ± 2.76% and 41.6% ± 4.29%, respectively, which showed a significant difference (P < 0.05. Cells treated with oridonin showed typical apoptotic features with acridine orange/ethidium bromide staining. After treatment with oridonin, the cells became round, shrank, and developed small buds around the nuclear membrane while forming apoptotic bodies. However, the change in the release of LDH caused by necrosis was insignificant, suggesting that the major cause of oridonin-induced HGC-27 cell death was apoptosis. Flow cytometric analysis also revealed that oridonin induced significant apoptosis compared with the controls (P < 0.05).

Apoptosis of HGC-27 induced by oridonin may be associated with differential expression of Apaf-1, caspase-3 and cytochrome c, which are highly dependent upon the mitochondrial pathway (Sun et al., 2012).

Ehrlich Ascites, Leukemia

Oridonin has been found to also increase lifespan of mice bearing Ehrlich ascites or P388 lymphocytic leukemia. Oridonin triggered apoptosis in more than 50% of t(8;21) leukemic cells in vitro at concentration of 2 M or higher accompanied by degradation of AE oncoprotein, and showed significant anti-leukemia efficacies with low adverse effects in vivo. These data suggest possible beneficial effects for patients with t(8;21) acute myeloid leukemia (AML) (Zhou et al., 2007).

Prostate Cancer, Breast Cancer, Ovarian Cancer

Oridonin exhibited anti-proliferative activity toward all cancer cell lines tested, with an IC50 estimated by the MTT cell viability assay ranging from 5.8+/-2.3 to 11.72+/-4.8 microM. The increased incidence of apoptosis, identified by characteristic changes in cell morphology, was seen in tumor lines treated with oridonin. Notably, at concentrations that induced apoptosis among tumor cells, oridonin failed to induce apoptosis in cultures of normal human fibroblasts. Oridonin up-regulated p53 and Bax and down-regulated Bcl-2 expression in a dose-dependent manner and its absorption spectrum was measured in the presence and absence of double stranded (ds) DNA. Oridonin inhibits cancer cell growth in a cell-cycle specific manner and shifts the balance between pro- and anti-apoptotic proteins in favor of apoptosis. The present data suggest that further studies are warranted to assess the potential of oridonin in cancer prevention and/or treatment (Chen et al., 2005).

Ovarian Cancer Stem Cells; Chemotherapy Resistance

Oridonin was 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 types of compounds in the treatment of ovarian cancer, strategically designed, large scale studies are warranted (Chen et al., 2012).

Colorectal Cancer

Oridonin induced potent growth inhibition, cell-cycle arrest, apoptosis, senescence and colony-forming inhibition in three colorectal cancer cell lines in a dose-dependent manner in vitro. Daily i.p. injection of oridonin (6.25, 12.5 or 25 mg/kg) for 28 days significantly inhibited the growth of SW1116 s.c. xenografts in BABL/C nude mice.

Oridonin possesses potent in vitro and in vivo anti-colorectal cancer activities that correlated with induction of histone hyperacetylation and regulation of pathways critical for maintaining growth inhibition and cell-cycle arrest. Therefore, oridonin may represent a novel therapeutic option in colorectal cancer treatment as it has been shown to induce apoptosis and senescence of colon cancer cells in vitro and in vivo (Gao et al., 2010).

Colon Cancer; Apoptosis

Oridonin increased intracellular hydrogen peroxide levels and reduced the glutathione content in a dose-dependent manner. N-acetylcysteine, a reactive oxygen species scavenger, not only blocked the oridonin-induced increase in hydrogen peroxide and glutathione depletion, but also blocked apoptosis and senescence induced by oridonin.

Moreover, exogenous catalase could inhibit the increase in hydrogen peroxide and apoptosis induced by oridonin, but not the glutathione depletion and senescence. Furthermore, thioredoxin reductase (TrxR) activity was reduced by oridonin in vitro and in cells, which may cause the increase in hydrogen peroxide. In conclusion, the increase in hydrogen peroxide and glutathione depletion account for oridonin-induced apoptosis and senescence in colorectal cancer cells, and TrxR inhibition is involved in this process.

Given the importance of TrxR as a novel cancer target in colon cancer, oridonin would be a promising clinical candidate (Gao et al., 2012).

Prostate Cancer; Apoptosis

Oridonin (ORI) could inhibit the proliferation and induce apoptosis in various cancer cell lines. After ORI treatment, the proliferations of human prostate cancer (HPC) cell lines PC-3 and LNCaP were inhibited in a concentration and time-dependent manner. ORI induced cell-cycle arrest at the G2/M phase. Autophagy occurred before the onset of apoptosis and protected cancer cells in ORI-treated HPC cells. P21 was involved in ORI-induced autophagy and apoptosis (Li et al., 2012).

References

Chen S, Gao J, Halicka HD, et al. (2005). The cytostatic and cytotoxic effects of oridonin (Rubescenin), a diterpenoid from Rabdosia rubescens, on tumor cells of different lineage. Int J Oncol, 26(3):579-88.


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.


Fujita E, Fujita T, Katayama H, Shibuya M. (1970). Terpenoids. Part XV. Structure and absolute configuration of oridonin isolated from Isodon japonicus trichocarpus. J Chem Soc (Chem Comm), 21:1674–1681


Fujita E, Nagao Y, Node M, et al. (1976). Anti-tumor activity of the Isodon diterpenoids: structural requirements for the activity. Experientia, 32:203–206.


Fujita T, Takeda Y, Sun HD, et al. (1988). Cytotoxic and anti-tumor activities of Rabdosia diterpenoids. Planta Med, 54:414–417.


Henan Medical Institute, Henan Medical College, Yunnan Institute of Botany. (1978). Oridonin–a new anti-tumor subject. Chin Science Bull, 23:53–56.


Ikezoe T, Chen SS, Tong XJ, et al. (2003). Oridonin induces growth inhibition and apoptosis of a variety of human cancer cells. Int J Oncol, 23(4):1187-93.


Gao FH, Hu XH, Li W, Liu H, et al. (2010). Oridonin induces apoptosis and senescence in colorectal cancer cells by increasing histone hyperacetylation and regulation of p16, p21, p27 and c-myc. BMC Cancer, 10:610. doi: 10.1186/1471-2407-10-610.


Gao FH, Liu F, Wei W, et al. (2012). Oridonin induces apoptosis and senescence by increasing hydrogen peroxide and glutathione depletion in colorectal cancer cells. Int J Mol Med, 29(4):649-55. doi: 10.3892/ijmm.2012.895.


Li X, Li X, Wang J, Ye Z, Li JC. (2012) Oridonin up-regulates expression of P21 and induces autophagy and apoptosis in human prostate cancer cells. Int J Biol Sci. 2012;8(6):901-12. doi: 10.7150/ijbs.4554.


Sun KW, Ma YY, Guan TP, et al. (2012). Oridonin induces apoptosis in gastric cancer through Apaf-1, cytochrome c and caspase-3 signaling pathway. World J Gastroenterol, 18(48):7166-74. doi: 10.3748/wjg.v18.i48.7166.


Tang W, Eisenbrand G. (1992). Chinese drugs of plant origin: chemistry, pharmacology, and use in traditional and modern medicine. Berlin: Springer-Verlag, 817–847.


Wang S, Zhong Z, Wan J, et al. (2013). Oridonin induces apoptosis, inhibits migration and invasion on highly-metastatic human breast cancer cells. Am J Chin Med, 41(1):177-96. doi: 10.1142/S0192415X13500134.


Zhang Wj, Huang Ql, Hua Z-C. (2010). Oridonin: A promising anti-cancer drug from China. Frontiers in Biology, 5(6):540-545.


Zhou G-B, Kang H, Wang L, et al. (2007). Oridonin, a diterpenoid extracted from medicinal herbs, targets AML1-ETO fusion protein and shows potent anti-tumor activity with low adverse effects on t(8;21) leukemia in vitro and in vivo. Blood, 109(8):3441-3450.

Gambogic acid

Cancer:
Leukemia, metastatic breast, osteocarcinoma, glioblastoma, breast, lung, liver

Action: Anti-cancer, tumor apoptosis

Gambogic acid (GA) is the principal active ingredient of gamboges which is the resin from various Garcinia species including Garcinia hanburyi (Hook. F.) and Garcinia morella (Panthong et al., 2007). GA is a natural product with potent apoptotic activity. GA has various biological effects, such as anti-inflammatory, analgesic and anti-pyretic as well as anti-cancer activities.

Tumor Apoptosis, Osteocarcinoma, Glioblastoma, Breast, Lung, Liver

GA binding to Transferrin receptor (TfR) induces a unique signal leading to rapid apoptosis of tumor cells. (Kasibhatla et al., 2005; Gu et al., 2008). GA enhances p53 protein level through inhibition of mdm2 oncogene expression and thereby hampers p53 harboring tumor growth. GA could increase the chemotherapeutic effect of cisplatin in human osteosarcoma treatment through inducing the cell-cycle arrest and promoting cell apoptosis (Zhao et al., 2013).

In vitro and in vivo studies have demonstrated its potential as an excellent cytotoxicity against a variety of malignant tumors, including glioblastoma, as well as cancers of the breast, lung and liver. GA is currently investigated in clinical trials in China (Qi et al., 2008).

Leukemia

Gambogic acid (GA) has been found to inhibit the proliferation of Jurkat leukemia cells with 50% inhibitory concentration values of 1.51±0.09 (24 hours), 0.98±0.13 (48 hours), and 0.67±0.12 µmol/L (72 hours). GA was able to induce apoptosis of Jurkat cells. Treated by GA, the expression of DIO-1 was up-regulated, and that of Bcl-2 and NF-κB was down-regulated, leading to the activation of pro-caspase 3. GA induced the translocation of DIO-1 to the nucleus. GA suppressed the proliferation of Jurkat cells by apoptosis induction. DIO-1 triggered early-stage cell death in GA-treated Jurkat cells (Wang et al., 2008).

Metastatic Cancer

Patients with advanced or metastatic cancer who had not received any effective routine conventional treatment, or who had failed to respond to the existing conventional treatment, were randomly assigned to receive either 45 mg/m(2) gambogic acid intravenously from days 1– 5 of a 2-week cycle (Group A), or 45 mg/m(2) every other day for a total of 5 times during a 2-week cycle (Group B). The ORRs were 14.3% in Group A and 0% in Group B. It was not possible to analyze the significant difference because one of the values was zero. The disease control rates (DCRs) were 76.2% in Group A and 61.5% in Group B (P = 0.0456). The observed adverse reactions were mostly Grades I and II, and occurred in most patients after administration of the trial drug. There was no significant difference in the incidence of adverse reactions between the two arms.

The preliminary results of this phase IIa exploratory study suggest that gambogic acid has a favorable safety profile when administered at 45 mg/m(2). The DCR was greater in patients receiving gambogic acid on days 1–5 of a 2-week cycle, but the incidence of adverse reactions was similar irrespective of the administration schedule (Chi et al., 2013).

References

Chi Y, Zhan XK, Yu H, et al. (2013). An open-labeled, randomized, multicenter phase IIa study of gambogic acid injection for advanced malignant tumors. Chin Med J, 126(9):1642-6.


Gu H, Wang X, Rao S, et al. (2008). Gambogic acid mediates apoptosis as a p53 inducer through down-regulation of mdm2 in wild-type p53-expressing cancer cells. Mol Cancer Ther, 7:3298–3305. doi: 10.1158/1535-7163.MCT-08-0212.


Kasibhatla S, Jessen KA, Maliartchouk S, et al. (2005). A role for transferrin receptor in triggering apoptosis when targeted with gambogic acid. Proc Natl Acad Sci, 102:12095–12100. doi: 10.1073/pnas.0406731102.


Panthong A, Norkaew P, Kanjanapothi D, et al. (2007). Anti-inflammatory, analgesic and anti-pyretic activities of the extract of gamboge from Garcinia hanburyi Hook f. J Ethnopharmacol, 111:335–340. doi: 10.1016/j.jep.2006.11.038.


Qi Q, Gu H, Yang Y, et al. (2008). Involvement of matrix metalloproteinase 2 and 9 in gambogic acid induced suppression of MDA-MB-435 human breast carcinoma cell lung metastasis. J Mol Med, 86:1367–1377. doi: 10.1007/s00109-008-0398-z.


Wang Y, Chen Y, Chen Z, et al. (2008). Gambogic acid induces death inducer-obliterator 1-mediated apoptosis in Jurkat T cells. Acta Pharmacologica Sinica, 29:349–354. doi:10.1111/j.1745-7254.2008.00762.x.


Zhao W, You CC, Zhuang JP, et al. (2013). Viability inhibition effect of gambogic acid combined with cisplatin on osteosarcoma cells via mitochondria-independent apoptotic pathway. Mol Cell Biochem.

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.