Category Archives: BAX

Germacrone

Cancer: Breast, stomach

Action: Cell-cycle arrest

Traditional medicinal herbs are an untapped source of potential pharmaceutical compounds. Germacrone is a natural product isolated from Rhizoma curcuma longa (L.).

Breast Cancer

Germacrone has been investigated for its inhibition on the proliferation of breast cancer cell lines. Germacrone treatment significantly inhibited cell proliferation, increased lactate dehydrogenase (LDH) release, and induced mitochondrial membrane potential (ΔΨ m) depolarization in both MCF-7 and MDA-MB-231 cells in a dose-dependent manner. Germacrone induced MDA-MB-231 and MCF-7 cell-cycle arrest at the G0/G1 and G2/M phases respectively and induced MDA-MB-231 cell apoptosis.

In addition, germacrone treatment induced caspase-3, 7, 9, PARP cleavage. It was therefore concluded that germacrone inhibited the proliferation of breast cancer cell lines by inducing cell-cycle arrest and apoptosis through mitochondria-mediated caspase pathway. These results might provide some molecular basis for the anti-tumor activity of Rhizoma curcuma (Zhong et al., 2011).

Stomach Cancer

Germacrone, contained in zedoary oil from Rhizoma curcuma, significantly decreased the cell viability of AGS cells (P < 0.01) and MGC 803 cells (P < 0.01), and the inhibitory effects were attenuated by elevated concentrations of FBS. At high concentrations (>=90 mug/mL), zedoary oil killed GES-1 cells. At low concentrations (<=60 mug/mL), zedoary oil was less inhibitory toward gastric cancer cell lines. In AGS cells, zedoary oil inhibited cell proliferation in a dose- and time-dependent manner, with decreased PCNA protein expression in the zedoary oil-treated cells, and arrested the cell-cycle at S, G2/M and G0/G1 stages after treatment for 6–48 hours. At concentrations of 30, 60 and 90 mug/mL, which resulted in significant inhibition of proliferation and cell-cycle arrest, zedoary oil induced cell apoptosis.

Zedoary oil up-regulated the ratio of Bax/Bcl-2 protein expression (P < 0.01). Zedoary oil which contains germacrone was hence found to inhibit AGS cell proliferation through cell-cycle arrest and cell apoptosis promotion, which are related to Bax/Bcl-2 protein expression.

References

Shi H, Tan B, Ji G, et al. (2013). Zedoary oil (Ezhu You) inhibits proliferation of AGS cells. Chin Med, 8(1):13.


Zhong Z, Chen X, Tan W, et al. (2011). Germacrone inhibits the proliferation of breast cancer cell lines by inducing cell-cycle arrest and promoting apoptosis. Eur J Pharmacol, 667(1-3):50-55. doi:10.1016/j.ejphar.2011.03.041.

Gentianaceae

Cancer: Prostate, breast, lung, pancreatic

Action: Causes cell-cycle arrest

Gentianaceae is a naturally occurring alkaloid isolated from Sophora flavescens (Aiton).

Prostate Cancer; AR-

Gentianaceae has shown anti-proliferative properties in a number of types of cancer, including breast, gastric, lung and pancreatic tumors. Gentianaceae was also found to promote apoptosis and inhibit invasion of cancer cells.

The anti-tumor effects of gentianaceae were evaluated on androgen-independent PC-3 prostate cancer cells. The effects of gentianaceae on cell-cycle progression and apoptosis of PC-3 cells were tested. Gentianaceae-treated PC-3 cells underwent G0/G1 cell-cycle arrest. There was a significant reduction in the number of S phase and G2/M phase cells in the treated group when compared to untreated cells.

There was also an increase in the number of necrotic cells in the gentianaceae-treated group when compared to untreated cells. Gentianaceae treatment resulted in increased levels of caspase-3 and Bax and decreased levels of Bcl-2. The data suggest that gentianaceae inhibits the proliferation of androgen-independent prostate cancer cells by causing G0/G1 cell-cycle arrest and promoting apoptosis. Gentianaceae-induced apoptosis was mediated by down-regulation of Bcl-2/Bax ratios and up-regulation of caspase-3 levels (Zhang et al., 2012).

Reference

Zhang P, Wang Z, Chong T, Ji Z. (2012). Matrine inhibits proliferation and induces apoptosis of the androgen “American Typewriter”; “American Typewriter”;‑ independent prostate cancer cell line PC-3. Mol Med Report, 5(3):783-7. doi: 10.3892/mmr.2011.701.

Geniposide –Penta-acetyl Geniposide (Ac)5GP

Cancers:
Glioma, melanoma, liver, hepatocarcinogenesis, hepatoma, prostate, cervical

Action: Cytostatic, induces apoptosis

Gardenia, the fruit of Gardenia jasminoides Ellis, has been widely used to treat liver and gall bladder disorders in Chinese medicine. It has been shown recently that geniposide, the main ingredient of Gardenia fructus , exhibits anti-tumor effect.

Hepatocarcinogenesis, Glioma

It has been demonstrated that (Ac)5GP plays more potent roles than geniposide in chemoprevention. (Ac)5GP decreased DNA damage and hepatocarcinogenesis, induced by aflatoxin B1 (AFB1), by activating the phase II enzymes glutathione S-transferase (GST) and GSH peroxidase (GSH-Px). It reduced the growth and development of inoculated C6 glioma cells, especially in pre-treated rats. In addition to the preventive effect, (Ac)5GP exerts its actions on apoptosis and growth arrest.

Treatment of (Ac)5GP caused DNA fragmentation of glioma cells. (Ac)5GP induced sub- G1 peak through the activation of apoptotic cascades PKCdelta/JNK/Fas/caspase8 and caspase 3. It arrested the cell-cycle at G0/ G1 by inducing the expression of p21, thus suppressing the cyclin D1/cdk4 complex formation and the phosphorylation of E2F.

Data from in vivo experiments indicated that (Ac)5GP is not harmful to the liver, heart and kidney. (Ac)5GP is strongly suggested to be an anti-tumor agent for development in the future (Peng, Huang, & Wang, 2005).

Induces Apoptosis

Previous studies have demonstrated the apoptotic cascades protein kinase C (PKC) delta/c-Jun NH2-terminal kinase (JNK)/Fas/caspases induced by penta-acetyl geniposide [(Ac)5GP]. However, the upstream signals mediating PKCdelta activation have not yet been clarified. Ceramide, mainly generated from the degradation of sphingomyelin, was hypothesized upstream above PKCdelta in (Ac)5GP-transduced apoptosis.

After investigation, (Ac)5GP was shown to activate neutral sphingomyelinase (N-SMase) immediately, with its maximum at 15 min. The NGF and p75 enhanced by (Ac)5GP was inhibited when combined with GW4869, the N-SMase inhibitor, indicating NGF/p75 as the downstream signals of N-SMase/ceramide. To evaluate whether N-SMase is involved in (Ac)5GP-transduced apoptotic pathway, cells were treated with (Ac)5GP, alone or combined with GW4869. It was demonstrated that N-SMase inhibition blocked FasL expression and caspase 3 activation. Similarly, p75 antagonist peptide attenuated the FasL/caspase 3 expression. It indicated that N-SMase activation is pivotal in (Ac)5GP-mediated apoptosis.

SMase and NGF/p75 are suggested to mediate upstream above PKCdelta, thus transducing FasL/caspase cascades in (Ac)5GP-induced apoptosis (Peng, Huang, Hsu, & Wang, 2006).

Glioma

Penta-acetyl geniposide [(Ac)(5)GP], an acetylated geniposide product from Gardenia fructus, has been known to have hepato-protective properties and recent studies have revealed its anti-proliferative and apoptotic effect on C6 glioma cells. The anti-metastastic effect of (Ac)(5)GP in the rat neuroblastoma line C6 glioma cells were investigated.

Further (Ac)(5)GP also exerted an inhibitory effect on phosphoinositide 3-kinase (PI3K) protein expression, phosphorylation of extracellular signal-regulated kinases 1 and 2 (ERK1/2) and inhibition of activation of transcription factor nuclear factor kappa B (NF-kappaB), c-Fos, c-Jun.

Findings suggest (Ac)(5)GP is highly likely to be an inhibiting cancer migration agent to be further developed in the future (Huang et al., 2009).

Melanoma

A new iridoid glycoside, 10-O-(4'-O-methylsuccinoyl) geniposide, and two new pyronane glycosides, jasminosides Q and R, along with nine known iridoid glycosides, and two known pyronane glycosides, were isolated from a MeOH extract of Gardeniae Fructus, the dried ripe fruit of Gardenia jasminoides (Rubiaceae).

The structures of new compounds were elucidated on the basis of extensive spectroscopic analyzes and comparison with literature. Upon evaluation of these compounds on the melanogenesis in B16 melanoma cells induced with α-melanocyte-stimulating hormone (α-MSH), three compounds, i.e., 6-O-p-coumaroylgeniposide (3), 7, and 6'-O-sinapoyljasminoside (12), exhibited inhibitory effects with 21.6-41.0 and 37.5-47.7% reduction of melanin content at 30 and 50 µM, respectively, with almost no toxicity to the cells (83.7-106.1% of cell viability at 50 µM) (Akisha et al., 2012).

Hepatoma, Prostate Cancer, Cervical Cancer

Genipin is a metabolite of geniposide isolated from an extract of Gardenia fructus. Some observations suggested that genipin could induce cell apoptosis in hepatoma cells and PC3 human prostate cancer cells. Genipin could remarkably induce cytotoxicity in HeLa cells and inhibit its proliferation. Induction of the apoptosis by genipin was confirmed by analysis of DNA fragmentation and induction of sub-G(1) peak through flow cytometry.

The results also showed that genipin-treated HeLa cells cycle was arrested at G(1) phase. Western blot analysis revealed that the phosphorylated c-Jun NH(2)-terminal kinase (JNK) protein, phospho-Jun protein, p53 protein and bax protein significantly increased in a dose-dependent manner after treatment of genipin for 24 hours; the activation of JNK may result in the increase of the p53 protein level; the increase of the p53 protein led to the accumulation of bax protein; and bax protein further induced cell apoptotic death eventually (Cao et al., 2010).

References

Akihisa T, Watanabe K, Yamamoto A, et al. (2012). Melanogenesis inhibitory activity of monoterpene glycosides from Gardeniae Fructus. Chemistry & Biodiversity, 9(8), 1490-9. doi: 10.1002/cbdv.201200030.


Cao H, Feng Q, Xu W, et al. (2010). Genipin induced apoptosis associated with activation of the c-Jun NH2-terminal kinase and p53 protein in HeLa cells. Biol Pharm Bull, 33(8):1343-8.


Huang HP, Shih YW, Wu CH, et al. (2009). Inhibitory effect of penta-acetyl geniposide on C6 glioma cells metastasis by inhibiting matrix metalloproteinase-2 expression involved in both the PI3K and ERK signaling pathways. Chemico-biological Interactions, 181(1), 8-14. doi: 10.1016/j.cbi.2009.05.009.


Peng CH, Huang CN, Hsu SP, Wang CJ. (2006). Penta-acetyl geniposide induce apoptosis in C6 glioma cells by modulating the activation of neutral sphingomyelinase-induced p75 nerve growth factor receptor and protein kinase Cdelta pathway. Molecular Pharmacology, 70(3), 997-1004.


Peng CH, Huang CN, Wang CJ. (2005). The anti-tumor effect and mechanisms of action of penta-acetyl geniposide. Current Cancer Drug Targets, 5(4), 299-305.

Diosgenin

Cancer: Breast, colon, prostate, leukemia, stomach

Action: HER-2, apoptosis, chemo-enhancing

Diosgenin is a plant-derived steroid isolated from Trigonella foenum-graecum (L.).

Breast Cancer; Chemo-enhancing

Diosgenin preferentially inhibited proliferation and induced apoptosis in HER2-overexpressing cancer cells. Furthermore, diosgenin inhibited the phosphorylation of Akt and mTOR, and enhanced phosphorylation of JNK.

The use of pharmacological inhibitors revealed that the modulation of Akt, mTOR and JNK phosphorylation was required for diosgenin-induced FAS suppression. Finally, it was shown that diosgenin could enhance paclitaxel-induced cytotoxicity in HER2-overexpressing cancer cells. These results suggested that diosgenin has the potential to advance as chemo-preventive or chemotherapeutic agent for cancers that overexpress HER2 (Chiang et al., 2007).

Colon Cancer

On 24 hours exposure to diosgenin, MTT cytotoxicity activity reduced by ³50% was achieved at the higher concentrations (i.e., ³80 µmol/L). However, compared with the control, 20 to 60 µmol/L diosgenin reduced the MTT activity only by 5% to 30%. Diosgenin caused a significant time-dependent and dose-dependent decrease in the proliferation of HT-29 cells. Twenty four hours exposure to diosgenin (20 to 100 µmol/L) inhibited cell proliferation compared with untreated cell growth. The in vitro experiment results indicated that diosgenin inhibits cell growth and induces apoptosis in the HT-29 human colon cancer cell line in a dose-dependent manner.

Furthermore, diosgenin induces apoptosis in HT-29 cells at least in part by inhibition of bcl-2 and by induction of caspase-3 protein expression (Raju et al., 2004).

Breast Cancer

The electrochemical behavior of breast cancer cells was studied on a graphite electrode by cyclic voltammetry (CV) and potentiometric stripping analysis (PSA) in unexposed and diosgenin exposed cells. In both cases, only one oxidative peak at approximately +0.75 V was observed. The peak area in PSA was used to study the growth of the cells and the effect of diosgenin on MCF-7 cells. The results showed that diosgenin can effectively inhibit the viability and proliferation of the breast cancer cells (Li et al., 2005).

Leukemia

Cell viability was assessed via an MTT assay. Apoptosis was investigated in terms of nuclear morphology, DNA fragmentation, and phosphatidylserine externalization. Cell cycle analysis was performed via PI staining and flow cytometry (FCM). Western blotting and immunofluorescence methods were used to determine the levels of p53, cell-cycle-related proteins and Bcl-2 family members. Cell cycle analysis showed that diosgenin caused G2/M arrest independently of p53. The levels of cyclin B1 and p21Cip1/Waf1 were decreased, whereas cdc2 levels were increased. The anti-apoptotic Bcl-2 and Bcl-xL proteins were down-regulated, whereas the pro-apoptotic Bax was upregulated.

Diosgenin was hence found to inhibit K562 cell proliferation via cell-cycle G2/M arrest and apoptosis, with disruption of Ca2+ homeostasis and mitochondrial dysfunction playing vital roles (Liu et al., 2005).

In recent years, Akt signaling has gained recognition for its functional role in more aggressive, therapy-resistant malignancies. As it is frequently constitutively active in cancer cells, several drugs are being investigated for their ability to inhibit Akt signaling. Diosgenin (fenugreek), a dietary compound, was examined for its action on Akt signaling and its downstream targets on estrogen receptor positive (ER+) and estrogen receptor negative (ER-) breast cancer (BCa) cells. Additionally, in vivo tumor studies indicate diosgenin significantly inhibits tumor growth in both MCF-7 and MDA-231 xenografts in nude mice. Thus, these results suggest that diosgenin might prove to be a potential chemotherapeutic agent for the treatment of BCa (Srinivasan et al., 2009).

Leukemia, Stomach Cancer

Protodioscin (PD) was purified from fenugreek (Trigonella foenumgraecum L.) and identified by mass spectrometry, and 1H- and 13C-NMR. The effects of PD on cell viability in human leukemia HL-60 and human stomach cancer KATO III cells were investigated. PD displayed strong growth-inhibitory effect against HL-60 cells, but weak growth-inhibitory effect on KATO III cells.

These findings suggest that growth inhibition by PD of HL-60 cells results from the induction of apoptosis by this compound in HL-60 cells (Hibasami et al., 2003).

References

Chiang CT, Way TD, Tsai SJ, Lin JK. (2007). Diosgenin, a naturally occurring steroid, suppresses fatty acid synthase expression in HER2-overexpressing breast cancer cells through modulating Akt, mTOR and JNK phosphorylation. FEBS letters, 581(30), 5735-42. doi:     10.1016/j.febslet.2007.11.021.


Hibasami H, Moteki H, Ishikawa K, et al. (2003). Protodioscin isolated from fenugreek (Trigonella foenumgraecum L.) induces cell death and morphological change indicative of apoptosis in leukemic cell line H-60, but not in gastric cancer cell line KATO III. Int J Mol Med, 11(1):23-6.


Li J, Liu X, Guo M, et al. (2005). Electrochemical Study of Breast Cancer Cells MCF-7 and Its Application in Evaluating the Effect of Diosgenin. Analytical Sciences, 21(5), 561. doi:10.2116/analsci.21.561


Liu MJ, Wang Z, Ju Y, Wong RNS, Wu QY. (2005). Diosgenin induces cell-cycle arrest and apoptosis in human leukemia K562 cells with the disruption of Ca2+ homeostasis. Cancer Chemotherapy and Pharmacology, 55(1), 79-90, doi: 10.1007/s00280-004-0849-3


Raju J, Patlolla JMR, Swamy MV, Rao CV. (2004). Diosgenin, a Steroid Saponin of Trigonella foenum graecum (Fenugreek), Inhibits Azoxymethane-Induced Aberrant Crypt Foci Formation in F344 Rats and Induces Apoptosis in HT-29 Human Colon Cancer Cells. Cancer Epidemiol Biomarkers Prev, 13; 1392.


Srinivasan S, Koduru S, Kumar R, et al. (2009). Diosgenin targets Akt-mediated prosurvival signaling in human breast cancer cells. International Journal of Cancer, 125(4), 961–967. doi: 10.1002/ijc.24419

Dietary Flavones

Cancer:
Prostate, colorectal., breast, pancreatic, bladder, ovarian, leukemia, liver, glioma, osteosarcoma, melanoma

Action: Anti-inflammatory, TAM resistance, cancer stem cells, down-regulate COX-2, apoptosis, cell-cycle arrest, anti-angiogenic, chemo-sensitzer, adramycin (ADM) resistance

Sulforaphane, Phenethyl isothiocyanate (PEITC), quercetin, epicatechin, catechin, Luteolin, apigenin

Anti-inflammatory

The anti-inflammatory activities of celery extracts, some rich in flavone aglycones and others rich in flavone glycosides, were tested on the inflammatory mediators tumor necrosis factor α (TNF-α) and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) in lipopolysaccharide-stimulated macrophages. Pure flavone aglycones and aglycone-rich extracts effectively reduced TNF-α production and inhibited the transcriptional activity of NF-κB, while glycoside-rich extracts showed no significant effects.

Celery diets with different glycoside or aglycone contents were formulated and absorption was evaluated in mice fed with 5% or 10% celery diets. Relative absorption in vivo was significantly higher in mice fed with aglycone-rich diets as determined by HPLC-MS/MS (where MS/MS is tandem mass spectrometry). These results demonstrate that deglycosylation increases absorption of dietary flavones in vivo and modulates inflammation by reducing TNF-α and NF-κB, suggesting the potential use of functional foods rich in flavones for the treatment and prevention of inflammatory diseases (Hostetler et al., 2012).

Colorectal Cancer

Association between the 6 main classes of flavonoids and the risk of colorectal cancer was examined using data from a national prospective case-control study in Scotland, including 1,456 incident cases and 1,456 population-based controls matched on age, sex, and residence area.

Dietary, including flavonoid, data were obtained from a validated, self-administered food frequency questionnaire. Risk of colorectal cancer was estimated using conditional logistic regression models in the whole sample and stratified by sex, smoking status, and cancer site and adjusted for established and putative risk factors.

The significant dose-dependent reductions in colorectal cancer risk that were associated with increased consumption of the flavonols quercetin, catechin, and epicatechin, remained robust after controlling for overall fruit and vegetable consumption or for other flavonoid intake. The risk reductions were greater among nonsmokers, but no interaction beyond a multiplicative effect was present.

This was the first of several a priori hypotheses to be tested in this large study and showed strong and linear inverse associations of flavonoids with colorectal cancer risk (Theodoratou et al., 2007).

Anti-angiogenic, Prostate Cancer

Luteolin is a common dietary flavonoid found in fruits and vegetables. The anti-angiogenic activity of luteolin was examined using in vitro, ex vivo, and in vivo models. Angiogenesis, the formation of new blood vessels from pre-existing vascular beds, is essential for tumor growth, invasion, and metastasis; hence, examination of this mechanism of tumor growth is essential to understanding new chemo-preventive targets. In vitro studies using rat aortic ring assay showed that luteolin at non-toxic concentrations significantly inhibited microvessel sprouting and proliferation, migration, invasion and tube formation of endothelial cells, which are key events in the process of angiogenesis. Luteolin also inhibited ex vivo angiogenesis as revealed by chicken egg chorioallantoic membrane assay (CAM) and matrigel plug assay.

Pro-inflammatory cytokines such as IL-1β, IL-6, IL-8, and TNF-α level were significantly reduced by the treatment of luteolin in PC-3 cells. Luteolin (10 mg/kg/d) significantly reduced the volume and the weight of solid tumors in prostate xenograft mouse model, indicating that luteolin inhibited tumorigenesis by targeting angiogenesis. Moreover, luteolin reduced cell viability and induced apoptosis in prostate cancer cells, which were correlated with the down-regulation of AKT, ERK, mTOR, P70S6K, MMP-2, and MMP-9 expressions.

Taken together, these findings demonstrate that luteolin inhibits human prostate tumor growth by suppressing vascular endothelial growth factor receptor 2-mediated angiogenesis (Pratheeshkumar et al., 2012).

Pancreatic Cancer; Chemo-sensitizer

The potential of dietary flavonoids apigenin (Api) and luteolin (Lut) were assessed in their ability to enhance the anti-proliferative effects of chemotherapeutic drugs on BxPC-3 human pancreatic cancer cells; additionally, the molecular mechanism of the action was probed.

Simultaneous treatment with either flavonoid (0,13, 25 or 50µM) and chemotherapeutic drugs 5-fluorouracil (5-FU, 50µM) or gemcitabine (Gem, 10µM) for 60 hours resulted in less-than-additive effect (p<0.05). Pre-treatment for 24 hours with 13µM of 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-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 Api or Lut hence effectively aid in the anti-proliferative activity of chemotherapeutic drugs (Johnson et al., 2013).

Breast Cancer; Chemo-sensitizer, Tamoxifen

The oncogenic molecules in human breast cancer cells are inhibited by luteolin treatment and it was found that the level of cyclin E2 (CCNE2) mRNA was higher in tumor cells than in normal paired tissue samples as assessed using real-time reverse-transcriptase polymerase chain reaction (RT-PCR) analysis (n=257).

Combined treatment with 4-OH-TAM and luteolin synergistically sensitized the TAM-R cells to 4-OH-TAM. These results 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).

Breast Cancer

Consumers of higher levels of Brassica vegetables, particularly those of the genus Brassica (broccoli, Brussels sprouts and cabbage), reduce their susceptibility to cancer at a variety of organ sites. Brassica vegetables contain high concentrations of glucosinolates that can be hydrolyzed by the plant enzyme, myrosinase, or intestinal microflora to isothiocyanates, potent inducers of cytoprotective enzymes and inhibitors of carcinogenesis. Oral administration of either the isothiocyanate, sulforaphane, or its glucosinolate precursor, glucoraphanin, inhibits mammary carcinogenesis in rats treated with 7,12-dimethylbenz[a]anthracene. To determine whether sulforaphane exerts a direct chemo-preventive action on animal and human mammary tissue, the pharmacokinetics and pharmacodynamics of a single 150 µmol oral dose of sulforaphane were evaluated in the rat mammary gland.

Sulforaphane metabolites were detected at concentrations known to alter gene expression in cell culture. Elevated cytoprotective NAD(P)H:quinone oxidoreductase (NQO1) and heme oxygenase-1 (HO-1) gene transcripts were measured using quantitative real-time polymerase chain reaction. An observed 3-fold increase in NQO1 enzymatic activity, as well as 4-fold elevated immunostaining of HO-1 in rat mammary epithelium, provide strong evidence of a pronounced pharmacodynamic action of sulforaphane. In a subsequent pilot study, eight healthy women undergoing reduction mammoplasty were given a single dose of a broccoli sprout preparation containing 200 µmol of sulforaphane. Following oral dosing, sulforaphane metabolites were readily measurable in human breast tissue enriched for epithelial cells. These findings provide a strong rationale for evaluating the protective effects of a broccoli sprout preparation in clinical trials of women at risk for breast cancer (Cornblatt et al., 2007).

In a proof of principle clinical study, the presence of disseminated tumor cells (DTCs) was demonstrated in human breast tissue after a single dose of a broccoli sprout preparation containing 200 µmol of sulforaphane. Together, these studies demonstrate that sulforaphane distributes to the breast epithelial cells in vivo and exerts a pharmacodynamic action in these target cells consistent with its mechanism of chemo-protective efficacy.

Such efficacy, coupled with earlier randomized clinical trials revealing the safety of repeated doses of broccoli sprout preparations , supports further evaluation of broccoli sprouts in the chemoprevention of breast and other cancers (Cornblatt et al., 2007).

CSCs

Recent research into the effects of sulforaphane on cancer stem cells (CSCs) has drawn a great deal of interest. CSCs are suggested to be responsible for initiating and maintaining cancer, and to contribute to recurrence and drug resistance. A number of studies have indicated that sulforaphane may target CSCs in different types of cancer through modulation of NF- κB, SHH, epithelial-mesenchymal transition and Wnt/β-catenin pathways. Combination therapy with sulforaphane and chemotherapy in preclinical settings has shown promising results (Li et al., 2013).

Anti-inflammatory

Sulforaphane has been found to down-regulate COX-2 expression in human bladder transitional cancer T24 cells at both transcriptional- and translational levels. Cyclooxygenase-2 (COX-2) overexpression has been associated with the grade, prognosis and recurrence of transitional cell carcinoma (TCC) of the bladder. Sulforaphane (5-20 microM) induced nuclear translocation of NF-kappaB and reduced its binding to the COX-2 promoter, a key mechanism for suppressing COX-2 expression by sulforaphane. Moreover, sulforaphane increased expression of p38 and phosphorylated-p38 protein. Taken together, these data suggest that p38 is essential in sulforaphane-mediated COX-2 suppression and provide new insights into the molecular mechanisms of sulforaphane in the chemoprevention of bladder cancer (Shan et al., 2009).

Bladder Cancer

An aqueous extract of broccoli sprouts potently inhibits the growth of human bladder carcinoma cells in culture and this inhibition is almost exclusively due to the isothiocyanates. Isothiocyanates are present in broccoli sprouts as their glucosinolate precursors and blocking their conversion to isothiocyanates abolishes the anti-proliferative activity of the extract.

Moreover, the potency of isothiocyanates in the extract in inhibiting cancer cell growth was almost identical to that of synthetic sulforaphane, as judged by their IC50 values (6.6 versus 6.8 micromol/L), suggesting that other isothiocyanates in the extract may be biologically similar to sulforaphane and that nonisothiocyanate substances in the extract may not interfere with the anti-proliferative activity of the isothiocyanates. These data show that broccoli sprout isothiocyanate extract is a highly promising substance for cancer prevention/treatment and that its anti-proliferative activity is exclusively derived from isothiocyanates (Tang et al., 2006).

Ovarian Cancer

Sulforaphane is an extract from the mustard family recognized for its anti-oxidation abilities, phase 2 enzyme induction, and anti-tumor activity. The cell-cycle arrest in G2/M by sulforaphane and the expression of cyclin B1, Cdc2, and the cyclin B1/CDC2 complex in PA-1 cells using Western blotting and co-IP Western blotting. The anti-cancer effects of dietary isothiocyanate sulforaphane on ovarian cancer were investigated using cancer cells line PA-1.

Sulforaphane -treated cells accumulated in metaphase by CDC2 down-regulation and dissociation of the cyclin B1/CDC2 complex.

These findings suggest that, in addition to the known effects on cancer prevention, sulforaphane may also provide anti-tumor activity in established ovarian cancer (Chang et al., 2013).

Leukemia Stem Cells

Isolated leukemia stem cells (LSCs) showed high expression of Oct4, CD133, β-catenin, and Sox2 and imatinib (IM) resistance. Differentially, CD34(+)/CD38(-) LSCs demonstrated higher BCR-ABL and β-catenin expression and IM resistance than CD34(+)/CD38(+) counterparts. IM and sulforaphane (SFN) combined treatment sensitized CD34(+)/CD38(-) LSCs and induced apoptosis, shown by increased caspase 3, PARP, and Bax while decreased Bcl-2 expression. Mechanistically, imatinib (IM) and sulforaphane (SFN) combined treatment resensitized LSCs by inducing intracellular reactive oxygen species (ROS). Importantly, β-catenin-silenced LSCs exhibited reduced glutathione S-transferase pi 1 (GSTP1) expression and intracellular GSH level, which led to increased sensitivity toward IM and sulforaphane.

It was hence demonstrated that IM and sulforaphane combined treatment effectively eliminated CD34(+)/CD38(-) LSCs. Since SFN has been shown to be well tolerated in both animals and human, this regimen could be considered for clinical trials (Lin et al., 2012).

DCIS Stem Cells

A miR-140/ALDH1/SOX9 axis has been found to be critical to basal cancer stem cell self-renewal and tumor formation in vivo, suggesting that the miR-140 pathway may be a promising target for preventive strategies in patients with basal-like Ductal Carcinoma in Situ (DCIS). The dietary compound sulforaphane has been found to decrease Transcription factor SOX-9 and Acetaldehyde dehydrogenases (ALDH1), and thereby reduced tumor growth in vivo (Li et al., 2013).

Glioma, Prostate Cancer, Colon Cancer, Breast Cancer, Liver Cancer

Phenethyl isothiocyanate (PEITC), a natural dietary isothiocyanate, inhibits angiogenesis. The effects of PEITC were examined under hypoxic conditions on the intracellular level of the hypoxia inducible factor (HIF-1α) and extracellular level of the vascular endothelial growth factor (VEGF) in a variety of human cancer cell lines. Gupta et al., (2013) observed that PEITC suppressed the HIF-1α accumulation during hypoxia in human glioma U87, human prostate cancer DU145, colon cancer HCT116, liver cancer HepG2, and breast cancer SkBr3 cells. PEITC treatment also significantly reduced the hypoxia-induced secretion of VEGF.

Suppression of HIF-1α accumulation during treatment with PEITC in hypoxia was related to PI3K and MAPK pathways.

Taken together, these results suggest that PEITC inhibits the HIF-1α expression through inhibiting the PI3K and MAPK signaling pathway and provide a new insight into a potential mechanism of the anti-cancer properties of PEITC.

Breast Cancer Metastasis

Breast tumor metastasis is a leading cause of cancer-related deaths worldwide. Breast tumor cells frequently metastasize to brain and initiate severe therapeutic complications. The chances of brain metastasis are further elevated in patients with HER2 overexpression. The MDA-MB-231-BR (BR-brain seeking) breast tumor cells stably transfected with luciferase were injected into the left ventricle of mouse heart and the migration of cells to brain was monitored using a non-invasive IVIS bio-luminescent imaging system.

Results demonstrate that the growth of metastatic brain tumors in PEITC treated mice was about 50% less than that of control. According to Kaplan Meir's curve, median survival of tumor-bearing mice treated with PEITC was prolonged by 20.5%. Furthermore, as compared to controls, we observed reduced HER2, EGFR and VEGF expression in the brain sections of PEITC treated mice. These results demonstrate the anti-metastatic effects of PEITC in vivo in a novel breast tumor metastasis model and provides the rationale for further clinical investigation (Gupta et al., 2013).

Osteosarcoma, Melanoma

Phenethyl isothiocyanate (PEITC) has been found to induce apoptosis in human osteosarcoma U-2 OS cells. The following end points were determined in regard to human malignant melanoma cancer A375.S2 cells: cell morphological changes, cell-cycle arrest, DNA damage and fragmentation assays and morphological assessment of nuclear change, reactive oxygen species (ROS) and Ca2+ generations, mitochondrial membrane potential disruption, and nitric oxide and 10-N-nonyl acridine orange productions, expression and activation of caspase-3 and -9, B-cell lymphoma 2 (Bcl-2)-associated X protein (Bax), Bcl-2, poly (adenosine diphosphate-ribose) polymerase, and cytochrome c release, apoptosis-inducing factor and endonuclease G. PEITC

It was therefore concluded that PEITC-triggered apoptotic death in A375.S2 cells occurs through ROS-mediated mitochondria-dependent pathways (Huang et al., 2013).

Prostate Cancer

The glucosinolate-derived phenethyl isothiocyanate (PEITC) has recently been demonstrated to reduce the risk of prostate cancer (PCa) and inhibit PCa cell growth. It has been shown that p300/CBP-associated factor (PCAF), a co-regulator for the androgen receptor (AR), is upregulated in PCa cells through suppression of the mir-17 gene. Using AR-responsive LNCaP cells, the inhibitory effects of PEITC were observed on the dihydrotestosterone-stimulated AR transcriptional activity and cell growth of PCa cells.

Expression of PCAF was upregulated in PCa cells through suppression of miR-17. PEITC treatment significantly decreased PCAF expression and promoted transcription of miR-17 in LNCaP cells. Functional inhibition of miR-17 attenuated the suppression of PCAF in cells treated by PEITC. Results indicate that PEITC inhibits AR-regulated transcriptional activity and cell growth of PCa cells through miR-17-mediated suppression of PCAF, suggesting a new mechanism by which PEITC modulates PCa cell growth (Yu et al., 2013).

Bladder Cancer; Adramycin (ADM) Resistance

The role of PEITC on ADM resistance reversal of human bladder carcinoma T24/ADM cells has been examined, including an increased drug sensitivity to ADM, cell apoptosis rates, intracellular accumulation of Rhodamine-123 (Rh-123), an increased expression of DNA topoisomerase II (Topo-II), and a decreased expression of multi-drug resistance gene (MDR1), multi-drug resistance-associated protein (MRP1), bcl-2 and glutathione s transferase π (GST-π). The results indicated that PEITC might be used as a potential therapeutic strategy to ADM resistance through blocking Akt and activating MAPK pathway in human bladder carcinoma (Tang et al., 2013).

Breast Cancer; Chemo-enhancing

The synergistic effect between paclitaxel (taxol) and phenethyl isothiocyanate (PEITC) on the inhibition of breast cancer cells has been examined. Two drug-resistant breast cancer cell lines, MCF7 and MDA-MB-231, were treated with PEITC and taxol. Cell growth, cell-cycle, and apoptosis were examined.

The combination of PEITC and taxol significantly decreased the IC50 of PEITC and taxol over each agent alone. The combination also increased apoptosis by more than 2-fold over each single agent in both cell lines. A significant increase of cells in the G2/M phases was detected. Taken together, these results indicated that the combination of PEITC and taxol exhibits a synergistic effect on growth inhibition in breast cancer cells. This combination deserves further study in vivo (Liu et al., 2013).

References

Chang CC, Hung CM, Yang YR, Lee MJ, Hsu YC. (2013). Sulforaphane induced cell-cycle arrest in the G2/M phase via the blockade of cyclin B1/CDC2 in human ovarian cancer cells. J Ovarian Res, 6(1):41. doi: 10.1186/1757-2215-6-41


Cornblatt BS, Ye LX, Dinkova-Kostova AT, et al. (2007). Preclinical and clinical evaluation of sulforaphane for chemoprevention in the breast. Carcinogenesis, 28(7):1485-1490. doi: 10.1093/carcin/bgm049


Gupta B, Chiang L, Chae K, Lee DH. (2013). Phenethyl isothiocyanate inhibits hypoxia-induced accumulation of HIF-1 α and VEGF expression in human glioma cells. Food Chem, 141(3):1841-6. doi: 10.1016/j.foodchem.2013.05.006.


Gupta P, Adkins C, Lockman P, Srivastava SK. (2013). Metastasis of Breast Tumor Cells to Brain Is Suppressed by Phenethyl Isothiocyanate in a Novel In Vivo Metastasis Model. PLoS One, 8(6):e67278. doi:10.1371/journal.pone.0067278


Hostetler G, Riedl K, Cardenas H, et al. (2012). Flavone deglycosylation increases their anti-inflammatory activity and absorption. Molecular Nutrition & Food Research, 56(4):558-569. doi: 10.1002/mnfr.201100596


Huang SH, Hsu MH, Hsu SC, et al. (2013). Phenethyl isothiocyanate triggers apoptosis in human malignant melanoma A375.S2 cells through reactive oxygen species and the mitochondria-dependent pathways. Hum Exp Toxicol. doi: 10.1177/0960327113491508


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, 60:83-91. doi: 10.1016/j.fct.2013.07.036.


Li Q, Yao Y, Eades G, Liu Z, Zhang Y, Zhou Q. (2013). Down-regulation of miR-140 promotes cancer stem cell formation in basal-like early stage breast cancer. Oncogene. doi: 10.1038/onc.2013.226.


Li Y, Zhang T. (2013). Targeting cancer stem cells with sulforaphane, a dietary component from broccoli and broccoli sprouts. Future Oncol, 9(8):1097-103. doi: 10.2217/fon.13.108.


Lin LC, Yeh CT, Kuo CC, et al. (2012). Sulforaphane potentiates the efficacy of imatinib against chronic leukemia cancer stem cells through enhanced abrogation of Wnt/ β-catenin function. J Agric Food Chem, 60(28):7031-9. doi: 10.1021/jf301981n.


Liu K, Cang S, Ma Y, Chiao JW. (2013). Synergistic effect of paclitaxel and epigenetic agent phenethyl isothiocyanate on growth inhibition, cell-cycle arrest and apoptosis in breast cancer cells. Cancer Cell Int, 13(1):10. doi: 10.1186/1475-2867-13-10.


Pratheeshkumar P, Son YO, Budhraja A, et al. (2012). Luteolin inhibits human prostate tumor growth by suppressing vascular endothelial growth factor receptor 2-mediated angiogenesis. PLoS One, 7(12):52279. doi: 10.1371/journal.pone.0052279.


Tang K, Lin Y, Li LM. (2013). The role of phenethyl isothiocyanate on bladder cancer ADM resistance reversal and its molecular mechanism. Anat Rec (Hoboken), 296(6):899-906. doi: 10.1002/ar.22677.


Tang L, Zhang Y, Jobson HE, et al. (2006). Potent activation of mitochondria-mediated apoptosis and arrest in S and M phases of cancer cells by a broccoli sprout extract. Mol Cancer Ther, 5(4):935-44. doi: 10.1158/1535-7163.MCT-05-0476


Theodoratou E, Kyle J, Cetnarskyj R, et al. (2007). Dietary flavonoids and the risk of colorectal cancer. Cancer Epidemiol Biomarkers Prev,16(4):684-93.


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.


Shan Y, Wu K, Wang W, et al. (2009). Sulforaphane down-regulates COX-2 expression by activating p38 and inhibiting NF-kappaB-DNA-binding activity in human bladder T24 cells. Int J Oncol, 34(4):1129-34.


Yu C, Gong AY, Chen D, et al. (2013). Phenethyl isothiocyanate inhibits androgen receptor-regulated transcriptional activity in prostate cancer cells through suppressing PCAF. Mol Nutr Food Res. doi: 10.1002/mnfr.201200810.

Cryptotanshinone (See also Tanshinone)

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

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

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

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

Anti-proliferative Agent

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

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

Anti-inflammatory; COX-2, PGE2

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

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

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

Prostate Cancer

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

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

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

Prostate Cancer; Androgen Receptor Positive

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

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

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

Breast Cancer, Cervical Cancer, Leukemia, Hepatocellular Carcinoma

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

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

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

Breast Cancer

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

Hepato-protective Effect

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

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

References

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

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

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

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

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

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

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

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

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

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

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

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

Corosolic acid

Cancer:
Myeloid leukemia, cervical., glioblastoma, gastric, sarcoma

Action: Immunosuppressive activity

Corosolic Acid is isolated from Lagerstroemia speciosa [(L.) Pers.] and Crataegus pinnatifida var. psilosa (C. K. Schneider).

Sarcoma; Immunosuppressive Activity

The results from an in vivo study showed that Corosolic acid (CA) administration did not suppress the tumor proliferation index, but significantly impaired subcutaneous tumor development and lung metastasis.

CA administration inhibited signal transducer and activator of transcription-3 (Stat3) activation and increased in the number of infiltrating lymphocytes in tumor tissues. Ex vivo analysis demonstrated that a significant immunosuppressive effect of MDSC in tumor-bearing mice was abrogated and the mRNA expressions of cyclooxygenase-2 and CCL2 in MDSC were significantly decreased by CA administration.

Furthermore, CA enhanced the anti-tumor effects of adriamycin and cisplatin in vitro. Since Stat3 is associated with tumor progression not only in osteosarcoma, but also in other malignant tumors, these findings indicate that CA might be widely useful in anti-cancer therapy by targeting the immunosuppressive activity of MDSC and through its synergistic effects with anti-cancer agents (Horlad et al., 2013).

Cervical Cancer

Xu et al. (2009) investigated the response of human cervix adenocarcinoma HeLa cells to Corosolic acid (CRA) treatment. These results showed that CRA significantly inhibited cell viability in both a dose- and a time-dependent manner. CRA treatment induced S cell-cycle arrest and caused apoptotic death in HeLa cells. It was found that CRA increased in Bax/Bcl-2 ratios by up-regulating Bax expression, disrupted mitochondrial membrane potential and triggered the release of cytochrome c from mitochondria into the cytoplasm.

These results, taken together, indicate CRA could have strong potentials for clinical application in treating human cervix adenocarcinoma and improving cancer chemotherapy.

Glioblastoma

Tumor-associated macrophages (TAMs) of M2 phenotype promote tumor proliferation and are associated with a poor prognosis in patients with glioblastoma.

The natural compounds possessing inhibitory effects on M2 polarisation in human monocyte-derived macrophages were investigated. Among 130 purified natural compounds examined, corosolic acid significantly inhibited the expression of CD163, one of the phenotype markers of M2 macrophages, as well as suppressed the secretion of IL-10, one of the anti-inflammatory cytokines preferentially produced by M2 macrophages, thus suggesting that corosolic acid suppresses M2 polarisation of macrophages.

Furthermore, corosolic acid inhibited the proliferation of glioblastoma cells, U373 and T98G, and the activation of Signal transducer and activator of transcription-3 (STAT3) and Nuclear Factor-kappa B (NF-κB), in both human macrophages and glioblastoma cells. These results indicate that corosolic acid suppresses the M2 polarisation of macrophages and tumor cell proliferation by inhibiting both STAT3 and NF-κB activation. Therefore, corosolic acid may be a new tool for tumor prevention and therapy (Fujiwara et al., 2010).

Gastric Cancer

Corosolic acid (CRA) suppresses HER2 expression, which in turn promotes cell-cycle arrest and apoptotic cell death of gastric cancer cells, providing a rationale for future clinical trials of CRA in the treatment of HER2-positive gastric cancers. CRA combined with adriamycin and 5-fluorouracil enhanced this growth inhibition, but not with docetaxel and paclitaxel (Lee et al., 2010).

Leukemia

Corosolic acid displayed about the same potent cytotoxic activity as ursolic acid against several human cancer cell lines. In addition, the compound displayed antagonistic activity against the phorbol ester-induced morphological modification of K-562 leukemic cells, indicating the suppression of protein kinase C (PKC) activity by the cytotoxic compound (Ahn et al., 1998).

References

Ahn KS, Hahm MS, Park EJ, Lee HK, Kim IH. (1998). Corosolic acid isolated from the fruit of Crataegus pinnatifida var. psilosa is a protein kinase C inhibitor as well as a cytotoxic agent. Planta Med, 64(5):468-70.


Fujiwara Y, Komohara Y, Ikeda T, Takeya M. (2010). Corosolic acid inhibits glioblastoma cell proliferation by suppressing the activation of signal transducer and activator of transcription-3 and nuclear factor-kappa B in tumor cells and tumor-associated macrophages. Cancer Science. doi: 10.1111/j.1349-7006.2010.01772.x


Horlad H, Fujiwara Y, Takemura K, et al. (2013). Corosolic acid impairs tumor development and lung metastasis by inhibiting the immunosuppressive activity of myeloid-derived suppressor cells. Molecular Nutrition & Food Research, 57(6):1046-1054. doi: 10.1002/mnfr.201200610


Lee MS, Cha EY, Thuong PT, et al. (2010). Down-regulation of human epidermal growth factor receptor 2/neu oncogene by corosolic acid induces cell-cycle arrest and apoptosis in NCI-N87 human gastric cancer cells. Biol Pharm Bull, 33(6):931-7.


Xu YF, Ge RL, Du J, et al. (2009). Corosolic acid induces apoptosis through mitochondrial pathway and caspases activation in human cervix adenocarcinoma HeLa cells. Cancer Letters, 284(2):229-237. doi:10.1016/j.canlet.2009.04.028.

Betulin and Betulinic acid

Cancer:
Neuroblastoma, medulloblastoma, glioblastoma, colon, lung, oesophageal, leukemia, melanoma, pancreatic, prostate, breast, head & neck, myeloma, nasopharyngeal, cervical, ovarian, esophageal squamous carcinoma

Action: Anti-angiogenic effects, induces apoptosis, anti-oxidant, cytotoxic and immunomodifying activities

Betulin is a naturally occurring pentacyclic triterpene found in many plant species including, among others, in Betula platyphylla (white birch tree), Betula X caerulea [Blanch. (pro sp.)], Betula cordifolia (Regel), Betula papyrifera (Marsh.), Betula populifolia (Marsh.) and Dillenia indica L . It has anti-retroviral., anti-malarial., and anti-inflammatory properties, as well as a more recently discovered potential as an anti-cancer agent, by inhibition of topoisomerase (Chowdhury et al., 2002).

Betulin is found in the bark of several species of plants, principally the white birch (Betula pubescens ) (Tan et al., 2003) from which it gets its name, but also the ber tree (Ziziphus mauritiana ), selfheal (Prunella vulgaris ), the tropical carnivorous plants Triphyophyllum peltatum and Ancistrocladus heyneanus, Diospyros leucomelas , a member of the persimmon family, Tetracera boiviniana , the jambul (Syzygium formosanum ) (Zuco et al., 2002), flowering quince (Chaenomeles sinensis ) (Gao et al., 2003), rosemary (Abe et al., 2002) and Pulsatilla chinensis (Ji et al., 2002).

Anti-cancer, Induces Apoptosis

The in vitro characterization of the anti-cancer activity of betulin in a range of human tumor cell lines (neuroblastoma, rhabdomyosarcoma-medulloblastoma, glioma, thyroid, breast, lung and colon carcinoma, leukaemia and multiple myeloma), and in primary tumor cultures isolated from patients (ovarian carcinoma, cervical carcinoma and glioblastoma multiforme) was carried out to probe its anti-cancer effect. The remarkable anti-proliferative effect of betulin in all tested tumor cell cultures was demonstrated. Furthermore, betulin altered tumor cell morphology, decreased their motility and induced apoptotic cell death. These findings demonstrate the anti-cancer potential of betulin and suggest that it may be applied as an adjunctive measure in cancer treatment (Rzeski, 2009).

Lung Cancer

Betulin has also shown anti-cancer activity on human lung cancer A549 cells by inducing apoptosis and changes in protein expression profiles. Differentially expressed proteins explained the cytotoxicity of betulin against human lung cancer A549 cells, and the proteomic approach was thus shown to be a potential tool for understanding the pharmacological activities of pharmacophores (Pyo, 2009).

Esophageal Squamous Carcinoma

The anti-tumor activity of betulin was investigated in EC109 cells. With the increasing doses of betulin, the inhibition rate of EC109 cell growth was increased, and their morphological characteristics were changed significantly. The inhibition rate showed dose-dependent relation.

Leukemia

Betulin hence showed potent inhibiting effects on EC109 cells growth in vitro (Cai, 2006).

A major compound of the methanolic extract of Dillenia indica L. fruits, betulinic acid, showed significant anti-leukaemic activity in human leukaemic cell lines U937, HL60 and K562 (Kumar, 2009).

Betulinic acid effectively induces apoptosis in neuroectodermal and epithelial tumor cells and exerts little toxicity in animal trials. It has been shown that betulinic acid induced marked apoptosis in 65% of primary pediatric acute leukemia cells and all leukemia cell lines tested. When compared for in vitro efficiency with conventionally used cytotoxic drugs, betulinic acid was more potent than nine out of 10 standard therapeutics and especially efficient in tumor relapse. In isolated mitochondria, betulinic acid induced release of both cytochrome c and Smac. Taken together, these results indicated that betulinic acid potently induces apoptosis in leukemia cells and should be further evaluated as a future drug to treat leukemia (Ehrhardt, 2009).

Multiple Myeloma

The effect of betulinic acid on the induction apoptosis of human multiple myeloma RPMI-8226 cell line was investigated. The results showed that within a certain concentration range (0, 5, 10, 15, 20 microg/ml), IC50 of betulinic acid to RPMI-8226 at 24 hours was 10.156+/-0.659 microg/ml, while the IC50 at 48 hours was 5.434+/-0.212 microg/ml, and its inhibiting effect on proliferation of RPMI-8226 showed both a time-and dose-dependent manner.

It is therefore concluded that betulinic acid can induce apoptosis of RPMI-8226 within a certain range of concentration in a time- and dose-dependent manner. This phenomenon may be related to the transcriptional level increase of caspase 3 gene and decrease of bcl-xl. Betulinic acid also affects G1/S in cell-cycle which arrests cells at phase G0/G1 (Cheng, 2009).

Anti-angiogenic Effects, Colorectal Cancer

Betulinic acid isolated from Syzygium campanulatum Korth (Myrtaceae) was found to have anti-angiogenic effects on rat aortic rings, matrigel tube formation, cell proliferation and migration, and expression of vascular endothelial growth factor (VEGF). The anti-tumor effect was studied using a subcutaneous tumor model of HCT 116 colorectal carcinoma cells established in nude mice. Anti-angiogenesis studies showed potent inhibition of microvessels outgrowth in rat aortic rings, and studies on normal and cancer cells did not show any significant cytotoxic effect.

In vivo anti-angiogenic study showed inhibition of new blood vessels in chicken embryo chorioallantoic membrane (CAM), and in vivo anti-tumor study showed significant inhibition of tumor growth due to reduction of intratumor blood vessels and induction of cell death. Collectively, these results indicate betulinic acid as an anti-angiogenic and anti-tumor candidate (Aisha, 2013).

Nasopharyngeal Carcinoma Melanoma, Leukemia, Lung, Colon, Breast,Prostate, Ovarian Cancer

Betulinic acid is an effective and potential anti-cancer chemical derived from plants. Betulinic acid can kill a broad range of tumor cell lines, but has no effect on untransformed cells. The chemical also kills melanoma, leukemia, lung, colon, breast, prostate and ovarian cancer cells via induction of apoptosis, which depends on caspase activation. However, no reports are yet available about the effects of betulinic acid on nasopharyngeal carcinoma (NPC), a widely spread malignancy in the world, especially in East Asia.

In a study, Liu & Luo (2012) showed that betulinic acid can effectively kill CNE2 cells, a cell line derived from NPC. Betulinic acid-induced CNE2 apoptosis was characterized by typical apoptosis hallmarks: caspase activation, DNA fragmentation, and cytochrome c release.

These observations suggest that betulinic acid may serve as a potent and effective anti-cancer agent in NPC treatment. Further exploration of the mechanism of action of betulinic acid could yield novel breakthroughs in anti-cancer drug discovery.

Cervical Carcinoma

Betulinic acid has shown anti-tumor activity in some cell lines in previous studies. Its anti-tumor effect and possible mechanisms were investigated in cervical carcinoma U14 tumor-bearing mice. The results showed that betulinic acid (100 mg/kg and 200 mg/kg) effectively suppressed tumor growth in vivo. Compared with the control group, betulinic acid significantly improved the levels of IL-2 and TNF-alpha in tumor-bearing mice and increased the number of CD4+ lymphocytes subsets, as well as the ratio of CD4+/CD8+ at a dose of 200 mg/kg.

Furthermore, treatment with betulinic acid induced cell apoptosis in a dose-dependent manner in tumor-bearing mice, and inhibited the expression of Bcl-2 and Ki-67 protein while upregulating the expression of caspase-8 protein. The mechanisms by which BetA exerted anti-tumor effects might involve the induction of tumor cell apoptosis. This process is also related to improvement in the body's immune response (Wang, 2012).

Anti-oxidant, Cytotoxic and Immunomodifying Activities

Betulinic acid exerted cytotoxic activity through dose-dependent impairment of viability and mitochondrial activity of rat insulinoma m5F (RINm5F) cells. Decrease of RINm5F viability was mediated by nitric oxide (NO)-induced apoptosis. Betulinic acid also potentiated NO and TNF-α release from macrophages therefore enhancing their cytocidal action. The rosemary extract developed more pronounced anti-oxidant, cytotoxic and immunomodifying activities, probably due to the presence of betulinic acid (Kontogianni, 2013).

Pancreatic Cancer

Lamin B1 is a novel therapeutic target of Betulinic Acid in pancreatic cancer. The role and regulation of lamin B1 (LMNB1) expression in human pancreatic cancer pathogenesis and betulinic acid-based therapy was investigated. Lamin proteins are thought to be involved in nuclear stability, chromatin structure and gene expression. Elevation of circulating LMNB1 marker in plasma could detect early stages of HCC patients, with 76% sensitivity and 82% specificity. Lamin B1 is a clinically useful biomarker for early stages of HCC in tumor tissues and plasma (Sun, 2010).

It was found that lamin B1 was significantly down-regulated by BA treatment in pancreatic cancer in both in vitro culture and xenograft models. Overexpression of lamin B1 was pronounced in human pancreatic cancer and increased lamin B1 expression was directly associated with low grade differentiation, increased incidence of distant metastasis and poor prognosis of pancreatic cancer patients.

Furthermore, knockdown of lamin B1 significantly attenuated the proliferation, invasion and tumorigenicity of pancreatic cancer cells. Lamin B1 hence plays an important role in pancreatic cancer pathogenesis and is a novel therapeutic target of betulinic acid treatment (Li, 2013).

Multiple Myeloma, Prostate Cancer

The inhibition of the ubiquitin-proteasome system (UPS) of protein degradation is a valid anti-cancer strategy and has led to the approval of bortezomib for the treatment of multiple myeloma. However, the alternative approach of enhancing the degradation of oncoproteins that are frequently overexpressed in cancers is less developed. Betulinic acid (BA) is a plant-derived small molecule that can increase apoptosis specifically in cancer but not in normal cells, making it an attractive anti-cancer agent.

Results in prostate cancer suggest that BA inhibits multiple deubiquitinases (DUBs), which results in the accumulation of poly-ubiquitinated proteins, decreased levels of oncoproteins, and increased apoptotic cell death. In the TRAMP transgenic mouse model of prostate cancer, treatment with BA (10 mg/kg) inhibited primary tumors, increased apoptosis, decreased angiogenesis and proliferation, and lowered androgen receptor and cyclin D1 protein.

BA treatment also inhibited DUB activity and increased ubiquitinated proteins in TRAMP prostate cancer but had no effect on apoptosis or ubiquitination in normal mouse tissues. Overall, this data suggests that BA-mediated inhibition of DUBs and induction of apoptotic cell death specifically in prostate cancer but not in normal cells and tissues may provide an effective non-toxic and clinically selective agent for chemotherapy (Reiner, 2013).

Melanoma

Betulinic acid was recently described as a melanoma-specific inducer of apoptosis, and it was investigated for its comparable efficacy against metastatic tumors and those in which metastatic ability and 92-kD gelatinase activity had been decreased by introduction of a normal chromosome 6. Human metastatic C8161 melanoma cells showed greater DNA fragmentation and growth arrest and earlier loss of viability in response to betulinic acid than their non-metastatic C8161/neo 6.3 counterpart.

These effects involved induction of p53 without activation of p21WAF1 and were synergized by bromodeoxyuridine in metastatic Mel Juso, with no comparable responses in non-metastatic Mel Juso/neo 6 cells. These data suggest that betulinic acid exerts its inhibitory effect partly by increasing p53 without a comparable effect on p21WAF1 (Rieber, 1998).

As a result of bioassay–guided fractionation, betulinic acid has been identified as a melanoma-specific cytotoxic agent. In follow-up studies conducted with athymic mice carrying human melanomas, tumor growth was completely inhibited without toxicity. As judged by a variety of cellular responses, anti-tumor activity was mediated by the induction of apoptosis. Betulinic acid is inexpensive and available in abundant supply from common natural sources, notably the bark of white birch trees. The compound is currently undergoing preclinical development for the treatment or prevention of malignant melanoma (Pisha, 1995).

Betulinic acid strongly and consistently suppressed the growth and colony-forming ability of all human melanoma cell lines investigated. In combination with ionizing radiation the effect of betulinic acid on growth inhibition was additive in colony-forming assays.

Betulinic acid also induced apoptosis in human melanoma cells as demonstrated by Annexin V binding and by the emergence of cells with apoptotic morphology. The growth-inhibitory action of betulinic acid was more pronounced in human melanoma cell lines than in normal human melanocytes.

The properties of betulinic acid make it an interesting candidate, not only as a single agent but also in combination with radiotherapy. It is therefore concluded that the strictly additive mode of growth inhibition in combination with irradiation suggests that the two treatment modalities may function by inducing different cell death pathways or by affecting different target cell populations (Selzer, 2000).

Betulinic acid has been demonstrated to induce programmed cell death with melanoma and certain neuroectodermal tumor cells. It has been demonstrated currently that the treatment of cultured UISO-Mel-1 (human melanoma cells) with betulinic acid leads to the activation of p38 and stress activated protein kinase/c-Jun NH2-terminal kinase (a widely accepted pro-apoptotic mitogen-activated protein kinases (MAPKs)) with no change in the phosphorylation of extracellular signal-regulated kinases (anti-apoptotic MAPK). Moreover, these results support a link between the MAPKs and reactive oxygen species (ROS).

These data provide additional insight in regard to the mechanism by which betulinic acid induces programmed cell death in cultured human melanoma cells, and it likely that similar responses contribute to the anti-tumor effect mediated with human melanoma carried in athymic mice (Tan, 2003).

Glioma

Betulinic acid triggers apoptosis in five human glioma cell lines. Betulinic acid-induced apoptosis requires new protein, but not RNA, synthesis, is independent of p53, and results in p21 protein accumulation in the absence of a cell-cycle arrest. Betulinic acid-induced apoptosis involves the activation of caspases that cleave poly(ADP ribose)polymerase.

Betulinic acid induces the formation of reactive oxygen species that are essential for BA-triggered cell death. The generation of reactive oxygen species is blocked by BCL-2 and requires new protein synthesis but is unaffected by caspase inhibitors, suggesting that betulinic acid toxicity sequentially involves new protein synthesis, formation of reactive oxygen species, and activation of crm-A-insensitive caspases (Wolfgang, 1999).

Head and Neck Carcinoma

In two head and neck squamous carcinoma (HNSCC) cell lines betulinic acid induced apoptosis, which was characterized by a dose-dependent reduction in cell numbers, emergence of apoptotic cells, and an increase in caspase activity. Western blot analysis of the expression of various Bcl-2 family members in betulinic acid–treated cells showed, surprisingly, a suppression of the expression of the pro-apoptotic protein Bax but no changes in Mcl-1 or Bcl-2 expression.

These data clearly demonstrate for the first time that betulinic acid has apoptotic activity against HNSCC cells (Thurnher et al., 2003).

References

Abe F, Yamauchi T, Nagao T, et al. (2002). Ursolic acid as a trypanocidal constituent in rosemary. Biological & Pharmaceutical Bulletin, 25(11):1485–7. doi:10.1248/bpb.25.1485. PMID 12419966.


Aisha AF, Ismail Z, Abu-Salah KM, et al. (2013). Syzygium campanulatum korth methanolic extract inhibits angiogenesis and tumor growth in nude mice. BMC Complement Altern Med,13:168. doi: 10.1186/1472-6882-13-168.


Cai WJ, Ma YQ, Qi YM et al. (2006). Ai bian ji bian tu bian can kao wen xian ge shi    Carcinogenesis,Teratogenesis & Mutagenesis,18(1):16-8.


Cheng YQ, Chen Y, Wu QL, Fang J, Yang LJ. (2009). Zhongguo Shi Yan Xue Ye Xue Za Zhi, 17(5):1224-9.


Chowdhury AR, Mandal S, Mittra B, et al. (2002). Betulinic acid, a potent inhibitor of eukaryotic topoisomerase I: identification of the inhibitory step, the major functional group responsible and development of more potent derivatives. Medical Science Monitor, 8(7): BR254–65. PMID 12118187.


Ehrhardt H, Fulda S, FŸhrer M, Debatin KM & Jeremias I. (2004). Betulinic acid-induced apoptosis in leukemia cells. Leukemia, 18:1406–1412. doi:10.1038/sj.leu.2403406


Gao H, Wu L, Kuroyanagi M, et al. (2003). Anti-tumor-promoting constituents from Chaenomeles sinensis KOEHNE and their activities in JB6 mouse epidermal cells. Chemical & Pharmaceutical Bulletin, 51(11):1318–21. doi:10.1248/cpb.51.1318. PMID 14600382.


Ji ZN, Ye WC, Liu GG, Hsiao WL. (2002). 23-Hydroxybetulinic acid-mediated apoptosis is accompanied by decreases in bcl-2 expression and telomerase activity in HL-60 Cells. Life Sciences, 72(1):1–9. doi:10.1016/S0024-3205(02)02176-8. PMID 12409140.


Kontogianni VG, Tomic G, Nikolic I, et al. (2013). Phytochemical profile of Rosmarinus officinalis and Salvia officinalis extracts and correlation to their anti-oxidant and anti-proliferative activity. Food Chem,136(1):120-9. doi: 10.1016/j.foodchem.2012.07.091.


Kumar D, Mallick S, Vedasiromoni JR, Pal BC. (2010). Anti-leukemic activity of Dillenia indica L. fruit extract and quantification of betulinic acid by HPLC. Phytomedicine, 17(6):431-5.


Li L, Du Y, Kong X, et al. (2013). Lamin B1 Is a Novel Therapeutic Target of Betulinic Acid in Pancreatic Cancer. Clin Cancer Res, Epub July 9. doi: 10.1158/1078-0432.CCR-12-3630


Liu Y, Luo W. (2012). Betulinic acid induces Bax/Bak-independent cytochrome c release in human nasopharyngeal carcinoma cells. Molecules and cells, 33(5):517-524. doi: 10.1007/s10059-012-0022-5


Pisha E, Chai H, Lee I-S, et al. (1995). Discovery of betulinic acid as a selective inhibitor of human melanoma that functions by induction of apoptosis. Nature Medicine, 1:1046 – 1051. doi: 10.1038/nm1095-1046


Pyo JS, Roh SH, Kim DK, et al. (2009). Anti-Cancer Effect of Betulin on a Human Lung Cancer Cell Line: A Pharmacoproteomic Approach Using 2 D SDS PAGE Coupled with Nano-HPLC Tandem Mass Spectrometry. Planta Med, 75(2): 127-131. doi: 10.1055/s-0028-1088366


Reiner T, Parrondo R, de Las Pozas A, Palenzuela D, Perez-Stable C. (2013). Betulinic Acid Selectively Increases Protein Degradation and Enhances Prostate Cancer-Specific Apoptosis: Possible Role for Inhibition of Deubiquitinase Activity. PLoS One, 8(2):e56234. doi: 10.1371/journal.pone.0056234.


Rieber M & Strasberg-Rieber M. (1998). Induction of p53 without increase in p21WAF1 in betulinic acid-mediated cell death is preferential for human metastatic melanoma. DNA Cell Biol, 17(5):399–406. doi:10.1089/dna.1998.17.399.


Rzeski W, Stepulak A, Szymanski M, et al. (2009). Betulin Elicits Anti-Cancer Effects in Tumor Primary Cultures and Cell Lines In Vitro. Basic and Clinical Pharmacology and Toxicology, 105(6):425–432. doi: 10.1111/j.1742-7843.2009.00471.x


Selzer E, Pimentel E, Wacheck V, et al. (2000). Effects of Betulinic Acid Alone and in Combination with Irradiation in Human Melanoma Cells. Journal of Investigative Dermatology, 114:935–940; doi:10.1046/j.1523-1747.2000.00972.x


Sun S, Xu MZ, Poon RT, Day PJ, Luk JM. (2010). Circulating Lamin B1 (LMNB1) biomarker detects early stages of liver cancer in patients. J Proteome Res, 9(1):70-8. doi: 10.1021/pr9002118.


Tan YM, Yu R, Pezzuto JM. (2003). Betulinic Acid-induced Programmed Cell Death in Human Melanoma Cells Involves Mitogen-activated Protein Kinase Activation. Clin Cancer Res, 9:2866.


Thurnher D, Turhani D, Pelzmann M, et al. (2003). Betulinic acid: A new cytotoxic compound against malignant head and neck cancer cells. Head & Neck. 25(9):732–740. doi: 10.1002/hed.10231


Wang P, Li Q, Li K, Zhang X, et al. (2012). Betulinic acid exerts immunoregulation and anti-tumor effect on cervical carcinoma (U14) tumor-bearing mice. Pharmazie, 67(8):733-9.


Wick W, Grimmel C, Wagenknecht B, Dichgans J, Weller M. (1999). Betulinic Acid-Induced Apoptosis in Glioma Cells: A Sequential Requirement for New Protein Synthesis, Formation of Reactive Oxygen Species, and Caspase Processing. JPET, 289(3):1306-1312.


Zuco V, Supino R, Righetti SC, et al. (2002). Selective cytotoxicity of betulinic acid on tumor cell lines, but not on normal cells. Cancer Letters, 175(1): 17–25. doi:10.1016/S0304-3835(01)00718-2. PMID 11734332.

Berberine

Cancer:
Liver,leukemia, breast, prostate, epidermoid (squamous-cell carcinoma), cervical.,testicular, melanoma, lymphoma, hepatoma

Action: Radio-sensitizer, anti-inflammatory, cell-cycle arrest, angiogenesis, chemo-enhancing, anti-metastatic, anti-oxidative

Berberine is a major phytochemical component of the roots and bark of herbal plants such as Berberis, Hydrastis canadensis and Coptis chinensis. It has been implicated in the cytotoxic effects on multiple cancer cell lines.

Anti-inflammatory

Berberine is an isoquinoline alkaloid widely distributed in natural herbs, including Rhizoma Coptidis chinensis and Epimedium sagittatum (Sieb. et Zucc.), a widely prescribed Chinese herb (Chen et al., 2008). It has a broad range of bioactivities, such as anti-inflammatory, anti-bacterial., anti-diabetes, anti-ulcer, sedation, protection of myocardial ischemia-reperfusion injury, expansion of blood vessels, inhibition of platelet aggregation, hepato-protective, and neuroprotective effects (Lau et al., 2001; Yu et al., 2005; Kulkarni & Dhir, 2010; Han et al., 2011; Ji, 2011). Berberine has been used in the treatment of diarrhea, neurasthenia, arrhythmia, diabetes, and so forth (Ji, 2011).

Angiogenesis, Chemo-enhancing

Inhibition of tumor invasion and metastasis is an important aspect of berberine's anti-cancer activities (Tang et al., 2009; Ho et al., 2009). A few studies have reported berberine's inhibition of tumor angiogenesis (Jie et al., 2011; Hamsa & Kuttan, 2012). In addition, its combination with chemotherapeutic drugs or irradiation could enhance the therapeutic effects (Youn et al., 2008; Hur et al., 2009).

Cell-cycle Arrest

The potential molecular targets and mechanisms of berberine are rather complicated. Berberine interacts with DNA or RNA to form a berberine-DNA or a berberine-RNA complex, respectively (Islam & Kumar. 2009; Li et al., 2012). Berberine is also identified as an inhibitor of several enzymes, such as N-acetyltransferase (NAT), cyclooxygenase-2 (COX-2), and telomerase (Sun et al., 2009).

Other mechanisms of berberine are mainly related to its effect on cell-cycle arrest and apoptosis, including regulation of cyclin-dependent kinase (CDK) family of proteins (Sun et al., 2009; Mantena, Sharma, & Katiyar, 2006) and expression regulation of B-cell lymphoma 2 (Bcl-2) family of proteins (such as Bax, Bcl-2, and Bcl-xL) (Sun et al., 2009), and caspases (Eom et al., 2010; Mantena, Sharma, & Katiyar, 2006). Furthermore, berberine inhibits the activation of the nuclear factor κ-light-chain-enhancer of activated B cells (NF-κB) and induces the formation of intracellular reactive oxygen species (ROS) in cancer cells (Sun et al., 2009; Eom et al., 2010). Interestingly, these effects might be specific for cancer cells (Sun et al., 2009).

Several studies have shown that berberine has anti-cancer potential by interfering with the multiple aspects of tumorigenesis and tumor progression in both in vitro and in vivo experiments. These observations have been well summarized in recent reports (Sun et al., 2009; Tan et al., 2011). Berberine inhibits the proliferation of multiple cancer cell lines by inducing cell-cycle arrest at the G1 or G 2 / M phases and by apoptosis (Sun et al., 2009; Eom et al., 2010; Burgeiro et al., 2011). In addition, berberine induces endoplasmic reticulum stress (Chang et al., 1990; Eom et al., 2010) and autophagy (Wang et al., 2010) in cancer cells.

However, compared with clinically prescribed anti-cancer drugs, the cytotoxic potency of berberine is much lower, with an IC50 generally at 10 µM to 100 µM depending on the cell type and treatment duration in vitro (Sun et al., 2009). Besides, berberine also induces morphologic differentiation in human teratocarcinoma (testes) cells (Chang et al., 1990).

Anti-metastatic

The effect of berberine on invasion, migration, metastasis, and angiogenesis is mediated through the inhibition of focal adhesion kinase (FAK), NF-κB, urokinase-type plasminogen-activator (u-PA), matrix metalloproteinase 2 (MMP-2), and matrix metalloproteinase 9 (MMP-9) (Ho et al., 2009; Hamsa & Kuttan. (2011); reduction of Rho kinase-mediated Ezrin phosphorylation (Tang et al., 2009); reduction of the expression of COX-2, prostaglandin E, and prostaglandin E receptors (Singh et al., 2011); down-regulation of hypoxia-inducible factor 1 (HIF-1), vascular endothelial growth factor (VEGF), pro-inflammatory mediators (Jie et al., 2011; Hamsa & Kuttan, 2012).

Hepatoma, Leukaemia

The cytotoxic effects of Coptis chinensis extracts and their major constituents on hepatoma and leukaemia cells in vitro have been investigated. Four human liver cancer cell lines, namely HepG2, Hep3B, SK-Hep1 and PLC/PRF/5, and four leukaemia cell lines, namely K562, U937, P3H1 and Raji, were investigated. C. chinensis exhibited strong activity against SK-Hep1 (IC50 = 7 microg/mL) and Raji (IC50 = 4 microg/mL) cell lines. Interestingly, the two major compounds of C. chinensis, berberine and coptisine, showed a strong inhibition on the proliferation of both hepatoma and leukaemia cell lines. These results suggest that the C. chinensis extract and its major constituents berberine and coptisine possess active anti-hepatoma and anti-leukaemia activities (Lin, 2004).

Leukemia

The steady-state level of nucleophosmin/B23 mRNA decreased during berberine-induced (25 g/ml, 24 to 96 hours) apoptosis of human leukemia HL-60 cells. A decline in telomerase activity was also observed in HL-60 cells treated with berberine. A stable clone of nucleophosmin/B23 over-expressed in HL-60 cells was selected and found to be less responsive to berberine-induced apoptosis. About 35% to 63% of control vector–transfected cells (pCR3) exhibited morphological characteristics of apoptosis, while about 8% to 45% of nucleophosmin/B23-over-expressed cells (pCR3-B23) became apoptotic after incubation with 15 g/ml berberine for 48 to 96 hours.

These results indicate that berberine-induced apoptosis is associated with the down-regulation of nucleophosmin/B23 and telomerase activity. Nucleophosmin/B23 may play an important role in the control of the cellular response to apoptosis induction (Hsing, 1999).

Prostate Cancer

In vitro treatment of androgen-insensitive (DU145 and PC-3) and androgen-sensitive (LNCaP) prostate cancer cells with berberine inhibited cell proliferation and induced cell death in a dose-dependent (10-100 micromol/L) and time-dependent (24–72 hours) manner. Berberine significantly (P < 0.05-0.001) enhanced apoptosis of DU145 and LNCaP cells with induction of a higher ratio of Bax/Bcl-2 proteins, disruption of mitochondrial membrane potential., and activation of caspase-9, caspase-3, and poly(ADP-ribose) polymerase.

The effectiveness of berberine in checking the growth of androgen-insensitive, as well as androgen-sensitive, prostate cancer cells without affecting the growth of normal prostate epithelial cells indicates that it may be a promising candidate for prostate cancer therapy (Mantena, 2006).

In another study, the treatment of human prostate cancer cells (PC-3) with berberine-induced dose-dependent apoptosis; however, this effect of berberine was not seen in non-neoplastic human prostate epithelial cells (PWR-1E). Berberine-induced apoptosis was associated with the disruption of the mitochondrial membrane potential., release of apoptogenic molecules (cytochrome c and Smac/DIABLO) from mitochondria and cleavage of caspase-9,-3 and PARP proteins.

Berberine-induced apoptosis was blocked in the presence of the anti-oxidant, N-acetylcysteine, through the prevention of disruption of mitochondrial membrane potential and subsequently release of cytochrome c and Smac/DIABLO. Taken together, these results suggest that the berberine-mediated cell death of human prostate cancer cells is regulated by reactive oxygen species, and therefore suggests that berberine may be considered for further studies as a promising therapeutic candidate for prostate cancer (Meeran, 2008).

Breast Cancer

DNA microarray technology has been used to understand the molecular mechanism underlying the anti-cancer effect of berberine carcinogenesis in two human breast cancer cell lines, the ER-positive MCF-7 and ER-negative MDA-MB-231 cells; specifically, whether it affects the expression of cancer-related genes. Treatment of the cancer cells with berberine markedly inhibited their proliferation in a dose- and time-dependent manner. The growth-inhibitory effect was much more profound in MCF-7 cell line than that in MDA-MB-231 cells.

IFN-β is among the most important anti-cancer cytokines, and the up-regulation of this gene by berberine is, at least in part, responsible for its anti-proliferative effect. The results of this study implicate berberine as a promising extract for chemoprevention and chemotherapy of certain cancers (Kang, 2005).

Breast Cancer Metastasis

Berberine also inhibits the growth of Anoikis-resistant MCF-7 and MDA-MB-231 breast cancer cell lines by inducing cell-cycle arrest. Anoikis, or detachment-induced apoptosis, may prevent cancer progression and metastasis by blocking signals necessary for survival of localized cancer cells. Resistance to anoikis is regarded as a prerequisite for metastasis; however, little is known about the role of berberine in anoikis-resistance.

The anoikis-resistant cells have a reduced growth rate and are more invasive than their respective adherent cell lines. The effect of berberine on growth was compared to that of doxorubicine, which is a drug commonly used to treat breast cancer, in both the adherent and anoikis-resistant cell lines. Berberine promoted the growth inhibition of anoikis-resistant cells to a greater extent than doxorubicine treatment. Treatment with berberine-induced cell-cycle arrest at G0/G1 in the anoikis-resistant MCF-7 and MDA-MB-231 cells was compared to untreated control cells. These results reveal that berberine can efficiently inhibit growth by inducing cell-cycle arrest in anoikis-resistant MCF-7 and MDA-MB-231 cells. Further analysis of these phenotypes is essential for understanding the effect of berberine on anoikis-resistant breast cancer cells, which would be relevant for the therapeutic targeting of breast cancer metastasis (Kim, 2010).

Melanoma

Berberine inhibits melanoma cancer cell migration by reducing the expressions of cyclooxygenase-2, prostaglandin E2 and prostaglandin E2 receptors. The effects and associated molecular mechanism of berberine on human melanoma cancer cell migration using melanoma cell lines A375 and Hs294 were probed in an in vitro cell migration assay, indicating that over- expression of cyclo-oxygenase (COX)-2, its metabolite prostaglandin E2 (PGE2) and PGE2 receptors promote the migration of cells.

Moreover, berberine inhibited the activation of nuclear factor-kappa B (NF-kB), an up- stream regulator of COX-2, in A375 cells, and treatment of cells with caffeic acid phenethyl ester, an inhibitor of NF-kB, inhibited cell migration. Together, these results indicate that berberine inhibits melanoma cell migration, an essential step in invasion and metastasis, by inhibition of COX-2, PGE2 and PGE2 receptors (Sing, 2011).

Cell-cycle Arrest, Squamous-cell Carcinoma

The in vitro treatment of human epidermoid carcinoma A431 cells with berberine decreases cell viability and induces cell death in a dose (5-75 microM)- and time (12–72 hours)-dependent manner, which was associated with an increase in G(1) arrest. G(0)/G(1) phase of the cell-cycle is known to be controlled by cyclin dependent kinases (Cdk), cyclin kinase inhibitors (Cdki) and cyclins.

Pre-treatment of A431 cells with the pan-caspase inhibitor (z-VAD-fmk) significantly blocked the berberine-induced apoptosis in A431 cells confirmed that berberine-induced apoptosis is mediated through activation of caspase 3-dependent pathway.

Together, these results indicate berberine as a chemotherapeutic agent against human epidermoid carcinoma A431 (squamous-cell) cells in vitro; further in vivo studies are required to determine whether berberine could be an effective chemotherapeutic agent for the management of non-melanoma skin cancers (Mantena, 2006).

Cervical Cancer, Radio-sensitizer

Cervical cancer remains one of the major killers amongst women worldwide. In India, a cisplatin based chemo/radiotherapy regimen is used for the treatment of advanced cervical cancer. Evidence shows that most of the chemotherapeutic drugs used in current clinical practice are radio-sensitizers. Natural products open a new avenue for treatment of cancer, as they are generally tolerated at high doses. Animal studies have confirmed the anti-tumorigenic activity of natural products, such as curcumin and berberine.

Berberine is a natural chemo-preventive agent, extracted from Berberis aristata, which has been shown to suppress and retard carcinogenesis by inhibiting inflammation.

The combined therapy of cisplatin/berberine and radiotherapy produced up-regulation of pro-apoptotic proteins Bax and p73, while causing down regulation of the anti-apoptotic proteins Bcl-xL, COX-2, cyclin D1. This additionally was accompanied by increased activity of caspase-9 and caspase-3, and reduction in telomerase activity. Results demonstrated that the treatment combination of berberine/cisplatin had increased induction of apoptosis relative to cisplatin alone (Komal., Singh, & Deshwal., 2013).

Anti-oxidative; Breast, Liver and Colon Cancer

The effect of B. vulgaris extract and berberine chloride on cellular thiobarbituric acid reactive species (TBARS) formation (lipid peroxidation), diphenyle–alpha-picrylhydrazyl (DPPH) oxidation, cellular nitric oxide (NO) radical scavenging capability, superoxide dismutase (SOD), glutathione peroxidase (GPx), acetylcholinesterase (AChE) and alpha-gulcosidase activities were spectrophotometrically determined.

Barberry crude extract contains 0.6 mg berberine/mg crude extract. Barberry extract showed potent anti-oxidative capacity through decreasing TBARS, NO and the oxidation of DPPH that is associated with GPx and SOD hyperactivation. Both berberine chloride and barberry ethanolic extract were shown to have inhibitory effect on the growth of breast, liver and colon cancer cell lines (MCF7, HepG2 and CACO-2, respectively) at different incubation times starting from 24 hours up to 72 hours and the inhibitory effect increased with time in a dose-dependent manner.

This work demonstrates the potential of the barberry crude extract and its active alkaloid, berberine, for suppressing lipid peroxidation, suggesting a promising use in the treatment of hepatic oxidative stress, Alzheimer and idiopathic male factor infertility. As well, berberis vulgaris ethanolic extract is a safe non-toxic extract as it does not inhibit the growth of PBMC that can induce cancer cell death (Abeer et al., 2013).

Source:

Alkaloids Isolated from Natural Herbs as the Anti-cancer Agents. Evidence-Based Complementary and Alternative Medicine. Volume 2012 (2012) http://dx.doi.org/10.1155/2012/485042

References

Burgeiro A, Gajate C, Dakir EH, et al. (2011). Involvement of mitochondrial and B-RAF/ERK signaling pathways in berberine-induced apoptosis in human melanoma cells. Anti-Cancer Drugs, 22(6):507–518.


Chang KSS, Gao C, Wang LC. (1990). Berberine-induced morphologic differentiation and down-regulation of c-Ki-ras2 protooncogene expression in human teratocarcinoma cells. Cancer Letters, 55(2):103–108.


Chen J, ZHao H, Wang X, et al. (2008). Analysis of major alkaloids in Rhizoma coptidis by capillary electrophoresis-electrospray-time of flight mass spectrometry with different background electrolytes. Electrophoresis, 29(10):2135–2147.


Eom KS, Kim HJ, So HS, et al. (2010). Berberine-induced apoptosis in human glioblastoma T98G Cells Is mediated by endoplasmic reticulum stress accompanying reactive oxygen species and mitochondrial dysfunction. Biological and Pharmaceutical Bulletin, 33(10):1644–1649.


El-Wahab AEA, Ghareeb DA, et al. (2013). In vitro biological assessment of berberis vulgaris and its active constituent, berberine: anti-oxidants, anti-acetylcholinesterase, anti-diabetic and anti-cancer effects. BMC Complementary and Alternative Medicine, 13:218 doi:10.1186/1472-6882-13-218


Hamsa TP & Kuttan G. (2011). Berberine inhibits pulmonary metastasis through down-regulation of MMP in metastatic B16F-10 melanoma cells. Phytotherapy Research, 26(4):568–578.


Hamsa TP & Kuttan G. (2012). Anti-angiogenic activity of berberine is mediated through the down-regulation of hypoxia-inducible factor-1, VEGF, and pro-inflammatory mediators. Drug and Chemical Toxicology, 35(1):57–70.


Han J, Lin H, Huang W. (2011). Modulating gut microbiota as an anti-diabetic mechanism of berberine. Medical Science Monitor, 17(7):RA164–RA167.


Ho YT, Yang JS, Li TC, et al. (2009). Berberine suppresses in vitro migration and invasion of human SCC-4 tongue squamous cancer cells through the inhibitions of FAK, IKK, NF-κB, u-PA and MMP-2 and -9. Cancer Letters, 279(2):155–162.


Hur JM, Hyun MS, Lim SY, Lee WY, Kim D. (2009). The combination of berberine and irradiation enhances anti-cancer effects via activation of p38 MAPK pathway and ROS generation in human hepatoma cells. Journal of Cellular Biochemistry, 107(5):955–964.


Islam MM & Kumar GS. (2009). RNA-binding potential of protoberberine alkaloids: spectroscopic and calorimetric studies on the binding of berberine, palmatine, and coralyne to protonated RNA structures. DNA and Cell Biology, 28(12):637–650.


Ji JB. (2011). Active Ingredients of Traditional Chinese Medicine: Pharmacology and Application, People's Medical Publishing House Cp., LTD.


Jie S, Li H, Tian Y, et al. (2011). Berberine inhibits angiogenic potential of Hep G2 cell line through VEGF down-regulation in vitro. Journal of Gastroenterology and Hepatology, 26(1):179–185.


Kang JX, Liu J, Wang J, He C, Li FP. (2005). The extract of huanglian, a medicinal herb, induces cell growth arrest and apoptosis by up-regulation of interferon-β and TNF-α in human breast cancer cells. Carcinogenesis, 26(11):1934-1939. doi:10.1093/carcin/bgi154


Kim JB, Yu JH, Ko E, et al. (2010). The alkaloid Berberine inhibits the growth of Anoikis-resistant MCF-7 and MDA-MB-231 breast cancer cell lines by inducing cell-cycle arrest. Phytomedicine, 17(6):436-40. doi: 10.1016/j.phymed.2009.08.012.


Komal Singh M, & Deshwal VK. (2013). Natural plant product berberine/cisplatin based radiotherapy for cervical cancer: The new and effective method to treat cervical cancer. Global Journal of Research on Medicinal Plants and Indigenous Medicine, 2(5), 278-291.


Kulkarni SK & Dhir A. (2010). Berberine: a plant alkaloid with therapeutic potential for central nervous system disorders. Phytotherapy Research, 24(3):317–324.


Lau CW, X. Q. Yao XQ, et al. (2001). Cardiovascular actions of berberine. Cardiovascular Drug Reviews, 19(3):234–244.


Li, XL Hu XJ, Wang H, et al. (2012). Molecular spectroscopy evidence for berberine binding to DNA: comparative binding and thermodynamic profile of intercalation. Biomacromolecules, 13(3):873–880.


Lin CC, Ng LT, Hsu FF, Shieh DE, Chiang LC. (2004). Cytotoxic effects of Coptis chinensis and Epimedium sagittatum extracts and their major constituents (berberine, coptisine and icariin) on hepatoma and leukaemia cell growth. Clin Exp Pharmacol Physiol, 31(1-2):65-9.


Mantena SK, Sharma SD, Katiyar SK. (2006). Berberine, a natural product, induces G1-phase cell-cycle arrest and caspase-3-dependent apoptosis in human prostate carcinoma cells. Mol Cancer Ther, 5(2):296-308. doi: 10.1158/1535-7163.MCT-05-0448


Mantena SK, Sharma SD, Katiyar SK. (2006). Berberine inhibits growth, induces G1 arrest and apoptosis in human epidermoid carcinoma A431 cells by regulating Cdki–Cdk-cyclin cascade, disruption of mitochondrial membrane potential and cleavage of caspase 3 and PARP. Carcinogenesis, 27(10):2018-27. doi: 10.1093/carcin/bgl043


Meeran SM, Katiyar S & Katiyar SK. (2008). Berberine-induced apoptosis in human prostate cancer cells is initiated by reactive oxygen species generation. Toxicology and Applied Pharmacology, 229(1):33-43. doi:10.1016/j.taap.2007.12.027


Singh T, Vaid M, Katiyar N, et al. (2011). Berberine, an isoquinoline alkaloid, inhibits melanoma cancer cell migration by reducing the expressions of cyclooxygenase-2, prostaglandin E and prostaglandin E receptors. Carcinogenesis, 32(1):86–92.


Sun Y, Xun K, Wang Y, Chen X. (2009). A systematic review of the anti-cancer properties of berberine, a natural product from Chinese herbs. Anti-Cancer Drugs, 20(9):757–769.


Tan W, Lu J, Huang M, et al. (2011). Anti-cancer natural products isolated from chinese medicinal herbs. Chinese Medicine, 6(1):27.


Tang F, Wang D, Duan C, et al. (2009) Berberine inhibits metastasis of nasopharyngeal carcinoma 5-8F cells by targeting rho kinase-mediated ezrin phosphorylation at threonine 567. Journal of Biological Chemistry, 284(40):27456–27466.


Wang N, Feng Y, Zhu M et al. (2010). Berberine induces autophagic cell death and mitochondrial apoptosis in liver cancer cells: the cellular mechanism. Journal of Cellular Biochemistry, 111(6):1426–1436.


Wu HL, Hsu CY, Liu WH, Yung BYM. (1999). Berberine‐induced apoptosis of human leukemia HL‐60 cells is associated with down‐regulation of nucleophosmin/B23 and telomerase activity. International Journal of Cancer, 81(6):923–929.


Youn MJ, So HS, Cho HJ, et al. (2008). Berberine, a natural product, combined with cisplatin enhanced apoptosis through a mitochondria/caspase-mediated pathway in HeLa cells. Biological and Pharmaceutical Bulletin, 31(5):789–795.


Yu HH, Kim KJ, Cha JD, et al. (2005). Antimicrobial activity of berberine alone and in combination with ampicillin or oxacillin against methicillin-resistant Staphylococcus aureus. Journal of Medicinal Food, 8(4):454–461.

Berbamine

Cancer: Breast, leukemia, liver, neutropenia

Action: Anti-metastatic, chemo-sensitizer

Breast Cancer, Leukemia

Berbamine (BER), isolated from the Chinese herb Berberis amurensis and Berberis vulgaris (L.), selectively induces apoptosis in certain breast cancer and leukemia cell lines.

Studies have shown that berbamine suppresses the growth, migration and invasion in highly-metastatic human breast cancer cells by possibly inhibiting Akt and NF-kappaB signaling with their upstream target c-Met and downstream targets Bcl-2/Bax, osteopontin, VEGF, MMP-9 and MMP-2.

BER has synergistic effects with anti-cancer agents trichostatin A, celecoxib and carmofur on inhibiting the growth of MDA-MB-231 cells and reducing the ratio of Bcl-2/Bax and/or VEGF expressions in the cancer cells. These findings suggest that berbamine may have wide therapeutic and/or adjuvant therapeutic application in the treatment of human breast cancer and other cancers (Wang, 2009).

MDR, Leukemia stem cells

Previous studies have shown that berbamine selectively induces apoptosis of imatinib (IM)-resistant-Bcr/Abl-expressing leukemia cells from the K562 cell line and CML patients. Berbamine derivatives obtained by synthesis were found to have very high activity in vitro. Six of these exhibited consistent high anti-tumor activity for imatinib-resistant K562 leukemia cells. Their IC(50) values at 48h were 0.36-0.55 microM, whereas berbamine IC(50) value was 8.9 microM. Cell cycle analysis results showed that compound 3h could reduce G0/G1 cells. In particular, these compounds displayed potent inhibition of the cytoplasm-to-nucleus translocation of NF-kappaB p65 which plays a critical role in the survival of leukemia stem cells (Xie, 2009).

Liver Cancer, Leukemia

Meng et al. (2013) reported that berbamine and one of its derivatives, bbd24, potently suppressed liver cancer cell proliferation and induced cancer cell death by targeting Ca2+/calmodulin-dependent protein kinase II (CAMKII). Furthermore, berbamine inhibited the in vivo tumorigenicity of liver cancer cells in NOD/SCID mice and downregulated the self-renewal abilities of liver cancer-initiating cells. Berbamine inhibits proliferation and induces apoptosis of KU812 leukaemia cells by increasing Smad3 activity (Kapoor, 2012).

Chronic Myeloid Leukemia, Leukopenia

During imatinib therapy, many patients with chronic myeloid leukemia (CML) develop severe neutropenia, leading to treatment interruptions, and potentially compromising response to imatinib. Berbamine (a bisbenzylisoquinoline alkaloid) has been widely used in Asian countries for managing leukopenia associated with chemotherapy. With berbamine support, the time to achieve complete cytogenetic response was significantly shorter (median, 6.5 vs. 10 months, p = 0.007). There were no severe adverse events associated with berbamine treatment. In conclusion, the present study reveals the potential clinical value of berbamine in the treatment of CML with imatinib-induced neutropenia (Zhao et al., 2011).

References

Kapoor S. (2012). Emerging role of berbamine as an anti-cancer agent in systemic malignancies besides chronic myeloid leukemia. Zhejiang Univ Sci B, 13(9):761-2.


Meng Z, Li T, Ma X, et al. (2013). Berbamine Inhibits the Growth of Liver Cancer Cells and Cancer-Initiating Cells by Targeting Ca2+/Calmodulin-Dependent Protein Kinase II. Mol Cancer Ther.


Wang S, Liu Q, Zhang Y, et al. (2009). Suppression of growth, migration and invasion of highly-metastatic human breast cancer cells by berbamine and its molecular mechanisms of action. Mol Cancer, 8:81.


Xie J, Ma T, Gu Y, et al. (2009). Berbamine derivatives: A novel class of compounds for anti-leukemia activity. Eur J Med Chem, 44(8):3293-8. doi: 10.1016/j.ejmech.2009.02.018


Zhao Y, Tan Y, Wu G, et al. (2011). Berbamine overcomes imatinib-induced neutropenia and permits cytogenetic responses in Chinese patients with chronic-phase chronic myeloid leukemia. Int J Hematol, 94(2):156-62. doi: 10.1007/s12185-011-0887-7.

Aloe-emodin (See also Emodin)

Cancer:
Nasopharyngeal., ER α degradation, Lung, breast, oral., glioblastoma, liver cancer prevention

Action: Cytostatic, radio-sensitizing, chemo-sensitizing

Nasopharyngeal Carcinoma

Aloe-emodin (AE), a natural., biologically active compound from Aloe vera leaves has been shown to induce apoptosis in several cancer cell lines in vitro. Investigation showed that AE induced G2/M phase arrest by increasing levels of cyclin B1 bound to Cdc2, and also caused an increase in apoptosis of nasopharyngeal carcinoma (NPC) cells, which was characterized by morphological changes, nuclear condensation, DNA fragmentation, caspase-3 activation, cleavage of poly (ADP-ribose) polymerase (PARP) and increased sub-G(1) population. Treatment of NPC cells with AE also resulted in a decrease in Bcl-X(L) and an increase in Bax expression.

Collectively, results indicate that the caspase-8-mediated activation of the mitochondrial death pathway plays a critical role in AE-induced apoptosis of NPC cells (Lin et al., 2010).

Glioblastoma

Aloe emodin arrested the cell-cycle in the S phase and promoted the loss of mitochondrial membrane potential in glioblastoma U87 cells that indicated the early event of the mitochondria-induced apoptotic pathway. It plays an important role in the regulation of cell growth and death (Ismail et al., 2013).

Breast Cancer

The anthraquinones emodin and aloe-emodin are also abundant in the rhizome Rheum palmatum and can induce cytosolic estrogen receptor α (ER α) degradation; it primarily affected nuclear ER α distribution similar to the action of estrogen when protein degradation was blocked. In conclusion, our data demonstrate that emodin and aloe-emodin specifically suppress breast cancer cell proliferation by targeting ER α protein stability through distinct mechanisms (Huang et al., 2013).

Lung Cancer

Photoactivated aloe-emodin induced anoikis and changes in cell morphology, which were in part mediated through its effect on cytoskeleton in lung carcinoma H460 cells. The expression of protein kinase Cδ (PKCδ) was triggered by aloe-emodin and irradiation in H460 cells. Furthermore, the photoactivated aloe-emodin-induced cell death and translocation of PKCδ from the cytosol to the nucleus was found to be significantly inhibited by rottlerin, a PKCδ-selective inhibitor (Chang et al., 2012).

Oral Cancer; Radio-sensitizing, Chemo-sensitizing

The treatment of cancer with chemotherapeutic agents and radiation has two major problems: time-dependent development of tumor resistance to therapy (chemoresistance and radioresistance) and nonspecific toxicity toward normal cells. Many plant-derived polyphenols have been studied intensively for their potential chemo-preventive properties and are pharmacologically safe.

These compounds include genistein, curcumin, resveratrol, silymarin, caffeic acid phenethyl ester, flavopiridol, emodin, green tea polyphenols, piperine, oleandrin, ursolic acid, and betulinic acid. Recent research has suggested that these plant polyphenols might be used to sensitize tumor cells to chemotherapeutic agents and radiation therapy by inhibiting pathways that lead to treatment resistance. These agents have also been found to be protective from therapy-associated toxicities.

Treatment with aloe-emodin at 10 to 40 microM resulted in cell-cycle arrest at G2/M phase. The alkaline phosphatase (ALP) activity in KB cells increased upon treatment with aloe-emodin when compared to controls. This is one of the first studies to focus on the expression of ALP in human oral carcinomas cells treated with aloe-emodin. These results indicate that aloe-emodin has anti-cancer effect on oral cancer, which may lead to its use in chemotherapy and chemo-prevention of oral cancer (Xiao et al., 2007).

Liver Cancer Prevention

In Hep G2 cells, aloe-emodin-induced p53 expression and was accompanied by induction of p21 expression that was associated with a cell-cycle arrest in G1 phase. In addition, aloe-emodin had a marked increase in Fas/APO1 receptor and Bax expression. In contrast, with p53-deficient Hep 3B cells, the inhibition of cell proliferation of aloe-emodin was mediated through a p21-dependent manner that did not cause cell-cycle arrest or increase the level of Fas/APO1 receptor, but rather promoted aloe-emodin-induced apoptosis by enhancing expression of Bax.

These findings suggest that aloe-emodin may be useful in liver cancer prevention (Lian et al., 2005).

References

Chang WT, You BJ, Yang WH, et al. (2012). Protein kinase C delta-mediated cytoskeleton remodeling is involved in aloe-emodin-induced photokilling of human lung cancer cells. Anti-cancer Res, 32(9):3707-13.

Huang PH, Huang CY, Chen MC, et al. (2013). Emodin and Aloe-Emodin Suppress Breast Cancer Cell Proliferation through ER α Inhibition. Evid Based Complement Alternat Med, 2013:376123. doi: 10.1155/2013/376123.

Ismail S, Haris K, Abdul Ghani AR, et al. (2013). Enhanced induction of cell-cycle arrest and apoptosis via the mitochondrial membrane potential disruption in human U87 malignant glioma cells by aloe emodin. J Asian Nat Prod Res.

Lian LH, Park EJ, Piao HS, Zhao YZ, Sohn DH. (2005). Aloe Emodin‐Induced Apoptosis in Cells Involves a Mitochondria‐Mediated Pathway. Basic & Clinical Pharmacology & Toxicology, 96(6):495–502.

Lin, ML, Lu, YC, Chung, JG, et al. (2010). Aloe-emodin induces apoptosis of human nasopharyngeal carcinoma cells via caspase-8-mediated activation of the mitochondrial death pathway. Cancer Letters, 291(1), 46-58. doi: 10.1016/j.canlet.2009.09.016.

Xiao B, Guo J, Liu D, Zhang S. (2007). Aloe-emodin induces in vitro G2/M arrest and alkaline phosphatase activation in human oral cancer KB cells. Oral Oncol, 43(9):905-10.

Alisol B Acetate

Cancer:
Liver, melanoma, ovarian, sarcoma, gastric cancer

Action: Cytostatic, cytotoxic

Four prostane-type triterpenes were isolated from a methanol extract of Alismatis Rhizoma by bioassay-guided isolation using in vitro cytotoxic assay. The compounds were identified as alisol B 23-acetate (1), alisol C 23-acetate (2), alisol B (3), alisol A 24-acetate (4) by spectroscopic methods. Amongst the compounds, alisol B (3) showed significant cytotoxicity against SK-OV3, B16-F10, and HT1080 cancer cell lines with ED50 values of 7.5, 7.5, 4.9 microg/ml, respectively (Lee et al., 2001).

Hepatocellular Carcinoma

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, respectively, on human hepatoma Hep3B cells, were investigated.

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).

Gastric Cancer

The cytotoxic effect of alisol B acetate on SGC7901 cells was measured by MTT assay and phase-contrast and electron microscopy. Cell-cycle and mitochondrial transmembrane potential (Deltapsim) were determined by flow cytometry and Western blotting was used to detect the expression of apoptosis-regulated gene Bcl-2, Bax, Apaf-1, caspase-3, caspase-9, Akt, P-Akt and phosphatidylinositol 3-kinases (PI3K).

Alisol B acetate inhibited the proliferation of SGC7901 cell line in a time- and dose-dependent manner. Alisol B acetate exhibits an anti-proliferative effect in SGC7901 cells by inducing apoptosis. Apoptosis of SGC7901 cells involves mitochondria-caspase and PI3K/Akt dependent pathways (Xu, Zhao & Li, 2009).

References

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.

 

 

Lee S, Kho Y, Min B, et al. (2001). Cytotoxic triterpenoides from Alismatis rhizome. Archives of Pharmacal Research. 24(6), 524-526.

 

Xu YH, Zhao LJ, & Li Y. (2009). Alisol B acetate induces apoptosis of SGC7901 cells via mitochondrial and phosphatidylinositol 3-kinases/Akt signaling pathways.

 

World Journal of Gastroenterology, 15(23), 2870-2877.

Acetyl-keto-beta-boswellic acid (AKBA)

Cancer: Colorectal, prostate, gastric

Action: Anti-cancer

Apoptotic

Acetyl-keto-beta-boswellic acid (AKBA), a triterpenoid isolated from Boswellia carterri Birdw and Boswellia serrata, has been found to inhibit tumor cell growth and to induce apoptosis. Boswellic acids trigger apoptosis via a pathway dependent on caspase-8 activation, and independent of Fas/Fas ligand interaction in colon cancer HT-29 cells (Liu et al., 2002).

Colon Cancer

Although there is increasing evidence showing that boswellic acid might be a potential anti-cancer agent, the mechanisms involved in its action are unclear. It has been shown that acetyl-keto-beta-boswellic acid (AKBA) inhibits cellular growth in several colon cancer cell lines. Cell cycle analysis by flow cytometry showed that cells were arrested at the G1 phase after AKBA treatment.

These results demonstrate that AKBA inhibits cellular growth in colon cancer cells. These findings may have implications for the use of boswellic acids as potential anti-cancer agents in colon cancer (Liu et al., 2006).

AKBA significantly inhibited human colon adenocarcinoma growth, showing arrest of the cell-cycle in G1-phase and induction of apoptosis. AKBA administration in mice effectively delayed the growth of HT-29 xenografts without signs of toxicity (Yuan et al., 2013).

Gastric Cancer

AKBA exhibited anti-cancer activity in vitro and in vivo. With oral application in mice, AKBA significantly inhibited gastric cancer cells line SGC-7901 and MKN-45 xenografts without toxicity.

This effect might be associated with its roles in cell-cycle arrest and apoptosis induction. The results also showed activation of p21(Waf1/Cip1) and p53 in mitochondria and increased cleaved caspase-9, caspase-3, and PARP and Bax/Bcl-2 ratio after AKBA treatment. Upon AKBA treatment, β-catenin expression in nuclei was inhibited, and membrane β-catenin was activated (Zhang et al., 2013).

Prostate

The apoptotic effects and the mechanisms of action of AKBA were studied in LNCaP and PC-3 human prostate cancer cells. AKBA induced apoptosis in both cell lines at concentrations above 10 microg/mL. AKBA-induced apoptosis was correlated with the activation of caspase-3 and caspase-8 as well as with poly(ADP)ribose polymerase (PARP) cleavage.

AKBA treatment increased the levels of CAAT/enhancer binding protein homologous protein (CHOP) and activated a DR5 promoter reporter but did not activate a DR5 promoter reporter with the mutant CHOP binding site. These results suggest that AKBA induces apoptosis in prostate cancer cells through a DR5-mediated pathway, which probably involves the induced expression of CHOP (Lu et al., 2008).

References

Liu J-J, Nilsson A, Oredsson S, et al. (2002). Boswellic acids trigger apoptosis via a pathway dependent on caspase-8 activation but independent on Fas/Fas ligand interaction in colon cancer HT-29 cells. Carcinogenesis. 23(12): 2087–2093. doi:10.1093/carcin/23.12.2087.

 

 

Liu JJ, Huang B, Hooi SC. (2006). Acetyl-keto-beta-boswellic acid inhibits cellular proliferation through a p21-dependent pathway in colon cancer cells. Br J Pharmacol, 148(8):1099-107.

 

Lu M, Xia L, Hua H, Jing Y. (2008). Acetyl-keto-beta-boswellic acid induces apoptosis through a death receptor 5-mediated pathway in prostate cancer cells. Cancer Res, 68(4):1180-6. doi: 10.1158/0008-5472.CAN-07-2978.

 

Yuan Y, Cui SX, Wang Y, et al. (2013). Acetyl-11-keto-beta-boswellic acid (AKBA) prevents human colonic adenocarcinoma growth through modulation of multiple signaling pathways. Biochim Biophys Acta, 1830(10):4907-16. doi: 10.1016/j.bbagen.2013.06.039.

 

Zhang YS, Xie JZ, Zhong JL, et al. (2013) Acetyl-11-keto-β-boswellic acid (AKBA) inhibits human gastric carcinoma growth through modulation of the Wnt/β -catenin signaling pathway. Biochim Biophys Acta, 1830(6):3604-15. doi: 10.1016/j.bbagen.2013.03.003.

Oxymatrine or Compound Matrine (Ku Shen)

Cancer: Sarcoma, pancreatic, breast, liver, lung, oral., rectal., stomach, leukemia, adenoid cystic carcinoma

Action: Anti-inflammatory, anti-proliferative, chemo-sensitizer, chemotherapy support, cytostatic, radiation support, anti-angiogenesis

Ingredients: ku shen (Sophora flavescens), bai tu ling (Heterosmilax chinensis).

TCM functions: Clearing Heat, inducing diuresis, cooling Blood, removing Toxin, dispersing lumps and relieving pain (Drug Information Reference in Chinese: See end, 2000-12).

Indications: Pain and bleeding caused by cancer.

Dosage and usage:

Intramuscular injection: 2-4 ml each time, twice daily; intravenous drip: 12 ml mixed in 200 ml NaCl injection, once daily. The total amount of 200 ml administration makes up a course of treatment. 2-3 consecutive courses can be applied.

Anti-cancer

Oxymatrine, isolated from the dried roots of Sophora flavescens (Aiton), has a long history of use in traditional Chinese medicine to treat inflammatory diseases and cancer. Kushen alkaloids (KS-As) and kushen flavonoids (KS-Fs) are well-characterized components in kushen. KS-As containing oxymatrine, matrine, and total alkaloids have been developed in China as anti-cancer drugs. More potent anti-tumor activities were identified in KS-Fs than in KS-As in vitro and in vivo (Sun et al., 2012). The four major alkaloids in compound Ku Shen injection are matrine, sophoridine, oxymatrine and oxysophocarpine (Qi, Zhang, & Zhang, 2013).

Sarcoma

When a high dose was used, the tumor-inhibitory rate of oxymatrine was 31.36%, and the vascular density of S180 sarcoma was lower than that in the control group and the expression of VEGF and bFGF was down-regulated. Oxymatrine hence has an inhibitory effect on S180 sarcoma and strong inhibitory effects on angiogenesis. Its mechanism may be associated with the down-regulating of VEGF and bFGF expression (Kong et al., 2003).

T Cell Leukemia

Matrine, a small molecule derived from the root of Sophora flavescens AIT was demonstrated to be effective in inducing T cell anergy in human T cell leukemia Jurkat cells.

The results showed that passage of the cells, and concentration and stimulation time of ionomycin on the cells could influence the ability of T cell anergy induction.

The cells exposed to matrine showed markedly decreased mRNA expression of interleukin-2, an indicator of T cell anergy. Pre-incubation with matrine or ionomycin could also shorten extracellular signal-regulated kinase (ERK) and suppress c-Jun NH(2)-terminal kinase (JNK) expression on the anergic Jurkat cells when the cells were stimulated with anti-OKT-3 plus anti-CD28 antibodies. Thus, matrine is a strong candidate for further investigation as a T cell immunotolerance inducer (Li et al., 2010).

Osteosarcoma

Results showed that treatment with oxymatrine resulted in a significant inhibition of cell proliferation and DNA synthesis in a dose-dependent manner, which has been attributed to apoptosis. Oxymatrine considerably inhibited the expression of Bcl-2 whilst increasing that of Bax.

Oxymatrine significantly suppressed tumor growth in female BALB/C nude mice bearing osteosarcoma MNNG/HOS xenograft tumors. In addition, no evidence of drug-related toxicity was identified in the treated animals by comparing the body weight increase and mortality (Zhang et al., 2013).

Pancreatic Cancer

Oxymatrine decreased the expression of angiogenesis-associated factors, including nuclear factor κB (NF-κB) and vascular endothelial growth factor (VEGF). Finally, the anti-proliferative and anti-angiogenic effects of oxymatrine on human pancreatic cancer were further confirmed in pancreatic cancer xenograft tumors in nude mice (Chen et al., 2013).

Furthermore, oxymatrine treatment led to the release of cytochrome c and activation of caspase-3 proteins. Oxymatrine can induce apoptotic cell death of human pancreatic cancer, which might be attributed to the regulation of Bcl-2 and IAP families, release of mitochondrial cytochrome c and activation of caspase-3 (Ling et al., 2011).

Rectal Carcinoma

Eighty-four patients diagnosed with rectal carcinoma at the People”s Hospital of Yichun city in Jiangxi province from September 2006 to September 2011, were randomly divided into two groups: therapeutic group and control group. The patients in the therapeutic group were treated with compound matrine and intensity modulated radiation therapy (IMRT) (30 Gy/10 f/2 W), while the patients in control group were treated with IMRT.

The clinical effect and survival rate in the therapeutic group were significantly higher (47.6%) than those in the control group (21.4%). All patients were divided by improvement, stability, and progression of disease in accordance with Karnofsky Performance Scale (KPS). According to the KPS, 16 patients had improvement, 17 stabilized and 9 had disease progress in the therapeutic group.

However, the control group had 12 improvements, 14 stabilized, and 16 disease progress. Quality of life in the therapeutic group was higher than that in the control group by rank sum test. The level of sIL-2R and IL-8 in the therapeutic group was lower on the first and 14th day, post radiation, when compared to the control group. However, there was no significant difference on the first day and 14th day, between both experimental groups post therapy, according to the student test. Compound matrine can decrease the side-effects of IMRT, significantly inhibit sIL-2R and IL-8 in peripheral blood from radiation, and can improve survival quality in patients with rectal cancer (Yin et al., 2013).

Gastric Cancer

Seventy-six cases of advanced gastric cancer were collected from June 2010 to November 2011, and randomly divided into either an experimental group or control group. Patients in the two groups were treated with matrine injection combined with SP regimen, or SP regimen alone, respectively. The effectiveness rate of the experimental group and control group was 57.5% and 52.8% respectively.

The treatment of advanced gastric cancer with matrine injection, combined with the SP regimen, can significantly improve levels of white blood cells and hemoglobin, liver function, incidence of diarrhea and constipation, and neurotoxicity, to improve the quality of life in patients with advanced gastric cancer (Xia, 2013).

Adenoid Cystic Carcinoma

Adenoid cystic carcinoma (ACC-2) cells were cultured in vitro. MTT assay was used to measure the cell proliferative effect. Compound radix Sophorae flavescentis injection could inhibit the proliferation of ACC-2 cells in vitro, and the dosage effect relationship was significant (P < 0.01). Radix Sophorae flavescentis injection could enhance ACC-2 cells Caspase-3 protein expression (P < 0.05 or P < 0.01), in a dose-dependent manner. It also could effectively restrain human adenoid cystic carcinoma ACC-2 cells Caspases-3 protein expression, and induce apoptosis, inhibiting tumor cell proliferation (Shi & Hu, 2012).

Breast Cancer; Chemotherapy

A retrospective analysis of oncological data of 70 postoperative patients with breast cancer from January 2008 to August 2011 was performed. According to the treatment method, the patients were divided into a therapy group (n=35) or control group (n=35). Patients in the control group were treated with the taxotere, adriamycin and cyclophosphamide regimen (TAC). The therapy group was treated with a combination of TAC and sophora root injection. Improved quality of life and incidence of adverse events, before and after treatment, for 2 cycles (21 days for a cycle) were compared.

The improvement rate of total quality of life in the therapy group was higher than that of the control group (P < 0.05). The drop of white blood cells and platelets, gastrointestinal reaction, elevated SGPT, and the incidence of hair loss in the therapy group were lower than those of the control group (P < 0.05).

Sophora root injection combined with chemotherapy in treatment of breast cancer can enhance the effect of chemotherapy, reduce toxicity and side-effects, and improve quality of life (An, An, & Wu, 2012).

Lung cancer; Pleural Effusion

The therapeutic efficiency of Fufang Kushen Injection Liquid (FFKSIL), IL-2, α-IFN on lung cancer accompanied with malignancy pleural effusions, was observed.

One hundred and fifty patients with lung cancer, accompanied with pleural effusions, were randomly divided into treatment and control groups. The treatment group was divided into three groups: injected FFKSIL plus IL-2, FFKSIL plus α-tFN, and IL-2 plus α>-IFN, respectively. The control group was divided into three groups and injected FFKSIL, IL-2 and α>-IFN, respectively. The effective rate of FFKSIL, IL-2, and α-IFN in a combination was significantly superior to single pharmacotherapy. The effective rate of fufangkushen plus ct-IFN was highest. The effect of FFKSIL, IL-2, and α-IFN, in a combination, on lung cancer with pleural effusions was significantly better than single pharmacotherapy. Moreover, the effect of FFKSIL plus IL-2 or α-IFN had the greatest effect (Hu & Mei, 2012).

Gastric Cancer

Administration of FFKSIL significantly enhanced serum IgA, IgG, IgM, IL-2, IL-4 and IL-10 levels, decreased serum IL-6 and TNF-αlevels, lowered the levels of lipid peroxides and enhanced GSH levels and activities of GSH-dependent enzymes. Our results suggest that FFKSIL blocks experimental gastric carcinogenesis by protecting against carcinogen-induced oxidative damage and improving immunity activity (Zhou et al., 2012).

Colorectal Cancer; Chemotherapy

Eighty patients after colorectal cancer resection were randomly divided into two groups: 40 patients in the control group were treated with routine chemotherapy including 5-fluorouridine(5-FU), calcium folinate(CF) and oxaliplatin, and 40 patients in the experimental group were treated with the same chemotherapy regime combined with 20 mLád-1 compound Kushen injection, for 10d during chemotherapy. In the control group the numbers of CD3+,CD4+T cells,NK cells and CD4+/CD8+ ratio significantly declined relative to prior to chemotherapy (P < 0.05), while CD8+T lymphocyte number increased significantly. In the experimental group, there were no significant differences between the numbers of CD3+,CD4+,CD8+T cells ,NK cells, and CD4+/CD8+ ratio, before and after chemotherapy (P > 0.05).

Compound Kushen injection can improve the immunologic function of patients receiving chemotherapy after colorectal cancer resection (Chen, Yu, Yuan, & Yuan, 2009).

NSCLC; Chemotherapy

A total of 286 patients with advanced NSCLC were enrolled for study. The patients were treated with either compound Kushen injection in combination with NP (NVB + CBP) chemotherapy (vinorelbine and carboplatin, n = 144), or with NP (NVB + CBP) chemotherapy alone (n = 142). The following indicators were observed: levels of Hb, WBC, PLT and T cell subpopulations in blood, serum IgG level, short-term  efficacy, adverse effects and quality of life.

The gastrointestinal reactions and the myelosuppression in the combination chemotherapy group were alleviated when compared with the chemotherapy alone group, showing a significant difference (P < 0.05). CD (8)(+) cells were markedly declined in the combination chemotherapy group, and the CD (4)(+)/CD (8)(+) ratio showed an elevation trend in the chemotherapy alone group. The Karnofsky Performance Scale (KPS) scores and serum IgM and IgG levels were higher in the combination chemotherapy group than those in the chemotherapy alone group (P < 0.01 and P < 0.05).

The compound Kushen injection plus NP chemotherapy regimen showed better therapeutic effect, reduced adverse effects of chemotherapy and improved the quality of life in patients with stage III and IV NSCLC (Fan et al., 2010).

Lung Adenocarcinoma

Different concentrations of matrine injection could inhibit the growth of SPCA/I human lung adenocarcinoma cells. There was a positive correlation between the inhibition rate and the drug concentration. Different concentrations of matrine injection combined with anti-tumor drugs had a higher growth inhibition rate than anti-tumor drugs alone. Matrine injection has direct growth suppression effect on SPCA/I human lung adenocarcinoma cells and SS+ injection combined with anti-tumor drugs shows a significant synergistic effect on tumor cells (Zhu, Jiang, Lu, Guo, & Gan, 2008).

Liver Cancer

Fifty-seven patients with unresectable primary liver cancer were randomly divided into 2 groups. The treatment group with 27 cases was treated by TACE combined with composite Kushen injection, and the control group with 30 cases was treated by TACE alone. One, two, and three year survival rates of the treatment group were 67%, 48%, and 37% respectively, and those of control group were 53%, 37%, and 20% respectively. There were significant differences between both groups (P < 0.05).

Combined TACE with composite Kushen injection can increase the efficacy of patients with unresectable primary liver cancer (Wang & Cheng, 2009).

Chemotherapy

Ten RCTs were included in a meta-analysis, whose results suggest that compared with chemotherapy alone, the combination had a statistically significant benefit in healing efficacy and improving quality of life. As well,  the combination also had a statistically significant benefit in myelosuppression, white blood cell, hematoblast, liver function and in reducing the gastroenteric reaction, decreasing the of CD3, CD4, CD4/CD8, and NK cells (Huang et al., 2011).

Colorectal Cancer, NSCLC, Breast Cancer; Chemotherapy

Fufang kushen Injection might improve the efficacies of chemotherapy in patients with colorectal cancer, NSCLC and breast cancer.

The results of a meta-analysis of 33 studies of randomized controlled trials with a total of 2,897 patients demonstrated that the short-term efficacies in patients with colorectal cancer, NSCLC, and breast cancer receiving Fufangkushen Injection plus chemotherapy were significantly better than for those receiving chemotherapy alone. However the results for patients with gastric cancer on combined chemotherapy were not significantly different from those for patients on chemotherapy alone (Fang, Lin, & Fan, 2011).

References

An, A.J., An, G.W., & Wu, Y.C. (2012). Observation of compound recipe light yellow Sophora root injection combined with chemotherapy in treatment of 35 postoperative patients with breast cancer. Medical & Pharmaceutical Journal of Chinese People”s Liberation Army, 24(10), 43-46. doi: 10.3969/j.issn.2095-140X.2012.10.016.


Chen, G., Yu, B., Yuan, S.J., & Yuan, Q. (2009). Effects of compound Kushen injection on the immunologic function of patients after colorectal cancer resection. Evaluation and Analysis of Drug-Use in Hospitals of China, 2009(9), R735.3. doi: cnki:sun:yypf.0.2009-09-025.


Chen H, Zhang J, Luo J, et al. (2013). Anti-angiogenic effects of oxymatrine on pancreatic cancer by inhibition of the NF-κB-mediated VEGF signaling pathway. Oncol Rep, 30(2):589-95. doi: 10.3892/or.2013.2529.


Fan, C.X., Lin, C.L., Liang, L., Zhao, Y.Y., Liu, J., Cui, J., Yang, Q.M., Wang, Y.L., & Zhang, A.R. (2010). Enhancing effect of compound Kushen injection in combination with chemotherapy for patients with advanced non-small-cell lung cancer. Chinese Journal of Oncology, 32(4), 294-297.


Fang, L., Lin, N.M., Fan, Y. (2011). Short-term  efficacies of Fufangkushen Injection plus chemotherapy in patients with solid tumors: a meta-analysis of randomized trials. Zhonghua Yi Xue Za Zhi, 91(35):2476-81.


Hu, D.J., & Mei, X.D. (2012). Observing therapeutic efficiency of fufangkushen injection, IL-2, α-IFN on lung cancer accompanied with malignancy pleural effusions. Journal of Clinical Pulmonology, 17(10), 1844-1845.


Huang S, Fan W, Liu P, Tian J. (2011). Meta-analysis of compound matrine injection combined with cisplatin chemotherapy for advanced gastric cancer. Zhongguo Zhong Yao Za Zhi, 36(22):3198-202.


Kong, Q-Z., Huang, D-S., Huang, T. et al. (2003). Experimental study on inhibiting angiogenesis in mice S180 by injections of three traditional Chinese herbs. Chinese Journal of Hospital Pharmacy, 2003-11. doi: CNKI:SUN:ZGYZ.0.2003-11-002


Li T, Wong VK, Yi XQ, et al. (2010). Matrine induces cell anergy in human Jurkat T cells through modulation of mitogen-activated protein kinases and nuclear factor of activated T-cells signaling with concomitant up-regulation of anergy-associated genes expression. Biol Pharm Bull, 33(1):40-6.


Ling Q, Xu X, Wei X, et al. (2011). Oxymatrine induces human pancreatic cancer PANC-1 cells apoptosis via regulating expression of Bcl-2 and IAP families, and releasing of cytochrome c. J Exp Clin Cancer Res, 30:66. doi: 10.1186/1756-9966-30-66.


Qi, L., Zhang, J., Zhang, Z. (2013). Determination of four alkaloids in Compound Kushen Injection by high performance liquid chromatography with ionic liquid as mobile phase additive. Chinese Journal of Chromatography, 31(3): 249-253. doi: 10.3724/SP.J.1123.2012.10039.


Shi, B., & Xu, H. (2012). Effects of compound radix Sophorae flavescentis injection on proliferation, apoptosis and caspase-3 expression in adenoid cystic carcinoma ACC-2 cells. Chinese Pharmacological Bulletin, 5(10), 721-724.


Sun M, Cao H, Sun L, et al. (2012). Anti-tumor activities of kushen: literature review. Evid Based Complement Alternat Med, 2012:373219. doi: 10.1155/2012/373219.


Wang, H.M., & Cheng, X.M. (2009). Composite Ku Shen injection combined with hepatic artery embolism on unresectable primary liver cancer. Modern Journal of Integrated Traditional Chinese and Western Medicine, 18(2), 1334–1335.


Xia, G. (2013). Clinical observation of compound matrine injection combined with SP regimen in advanced gastric cancer. Journal of Liaoning Medical University, 2013(1), 37-38.


Yin, W.H., Sheng, J.W., Xia, H.M., Chen, J., Wu, Y.W., & Fan, H.Z. (2013). Study on the effect of compound matrine on the level of sIL-2R and IL-8 in peripheral blood cells of patients with rectal cancer to radiation. Global Traditional Chinese Medicine, 2013(2), 100-104.


Zhang Y, Sun S, Chen J, et al. (2013). Oxymatrine induces mitochondria dependent apoptosis in human osteosarcoma MNNG/HOS cells through inhibition of PI3K/Akt pathway. Tumor Biol.


Zhou, S-K., Zhang, R-L., Xu, Y-F., Bi, T-N. (2012) Anti-oxidant and Immunity Activities of Fufang Kushen Injection Liquid. Molecules 2012, 17(6), 6481-6490; doi:10.3390/molecules17066481


Zhu, M.Y., Jiang, Z.H., Lu, Y.W., Guo, Y., & Gan, J.J. (2008). Matrine and anti-tumor drugs in inhibiting the growth of human lung cancer cell line. Journal of Chinese Integrative Medicine, 6(2), 163-165. doi: 10.3736/jcim20080211.

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.

Isorhamnetin

Cancer:
Lung, colon, acute myeloid leukemia, T lymphoma, Ehrlich carcinoma, gastric, esophageal squamous cell, chronic myelogenous leukemia

Action: Dox-induced cardiotoxicity, anti-oxidant

Isorhamnetin, the anti-tumor component of Hippophae rhamnoides Linn, is also a member of the ßavonoid class of compounds. Its chemical name is 3,5,7-trihydroxy-2-(4-hydroxy-3-methoxyphenyl) chromen-4-one and its molecular formula is C16H12O7.

Lung Cancer

Isorhamnetin shows good inhibitory effects on human lung adenocarcinoma A549 cells, human colon cancer HT-29 cells, human chronic myeloid leukemia K562 cells, human acute myeloid leukemia HL-60 cells, mouse T lymphoma YAC-1 cells and mouse Ehrlich carcinoma. In terms of its mechanism of action, it seems that isorhamnetin simultaneously reduces the expression of Bcl-2 and increases the expression of Bax, which activates caspase-9 and its downstream factor caspase-3, thus resulting in cell death (Zhu et al. 2005).

Colorectal Cancer

It was demonstrated that isorhamnetin prevents colorectal tumorigenesis. Dietary isorhamnetin decreased mortality, tumor number, and tumor burden by 62%, 35%, and 59%, respectively. Magnetic resonance imaging, histopathology, and immunohistochemical analysis revealed that dietary isorhamnetin resolved the DSS-induced inflammatory response faster than control diet.

These observations suggest the chemo-protective effects of isorhamnetin in colon cancer are linked to its anti-inflammatory activities and its inhibition of oncogenic Src activity and consequential loss of nuclear β-catenin, activities that are dependent on CSK expression (Saud et al., 2013).

Gastric Cancer

The potential effects of isorhamnetin (IH), a 3'-O-methylated metabolite of quercetin, were investigated on the peroxisome proliferator-activated receptor γ (PPAR-γ) signaling cascade using proteomics technology platform, gastric cancer (GC) cell lines, and xenograft mice model.

It was observed that IH exerted a strong anti-proliferative effect and increased cytotoxicity in combination with chemotherapeutic drugs. IH also inhibited the migratory/invasive properties of gastric cancer cells, which could be reversed in the presence of PPAR-γ inhibitor.

Using molecular docking analysis, Ramachandran et al. (2013) demonstratd that IH formed interactions with seven polar residues and six nonpolar residues within the ligand-binding pocket of PPAR-γ that are reported to be critical for its activity and could competitively bind to PPAR-γ. IH significantly increased the expression of PPAR-γ in tumor tissues obtained from xenograft model of GC. Overall, these findings clearly indicate that anti-tumor effects of IH may be mediated through modulation of the PPAR-γ activation pathway in GC.

Cardiac-protective; Doxorubicin

Isorhamnetin is a natural anti-oxidant with obvious cardiac-protective effect. Its action against doxorubicin-induced cardotoxicity and underlying mechanisms were investigated. Doxorubicin (Dox) is an anthracycline antibiotic for cancer therapy with limited usage due to cardiotoxicity. The aim of this study is to investigate the possible protective effect of isorhamnetin against Dox-induced cardiotoxicity and its underlying mechanisms. In an in vivo investigation, rats were intraperitoneally (i.p.) administered with Dox to duplicate the model of Dox-induced chronic cardiotoxicity.

Daily pre-treatment with isorhamnetin (5 mg/kg, i.p.) for 7 days was found to reduce Dox-induced myocardial damage significantly, including the decline of cardiac index, decrease in the release of serum cardiac enzymes, and amelioration of heart vacuolation. In vitro studies on H9c2 cardiomyocytes, isorhamnetin was effective to reduce Dox-induced cell toxicity. Isorhamnetin also potentiated the anti-cancer activity of Dox in MCF-7, HepG2 and Hep2 cells. These findings indicated that isorhamnetin can be used as an adjuvant therapy for the long-term clinical use of Dox (Sun et al., 2013).

Chronic Myelogenous Leukemia

The isorhamnetin 3-o-robinobioside and its original extract, ethyl acetate extract, from Nitraria retusa leaves, were evaluated for their ability to induce anti-oxidant and anti-genotoxic effects in human chronic myelogenous leukemia cell line. They were shown to have a great anti-oxidant and anti-genotoxic potential on human chronic myelogenous leukemia cell line K562 (Boubaker et al., 2012).

Esophageal Cancer

The flavonol aglycone isorhamnetin shows anti-proliferative activity in a variety of cancer cells and it inhibits the proliferation of human esophageal squamous carcinoma Eca-109 cells in vitro (Shi et al., 2012).

References

Boubaker J, Ben Sghaier M, Skandrani I, et al. (2012). Isorhamnetin 3-O-robinobioside from Nitraria retusa leaves enhance anti-oxidant and anti-genotoxic activity in human chronic myelogenous leukemia cell line K562. BMC Complement Altern Med, 12:135. doi: 10.1186/1472-6882-12-135.


Ramachandran L, Manu KA, Shanmugam MK, et al. (2013). Isorhamnetin inhibits proliferation and invasion and induces apoptosis through the modulation of peroxisome proliferator-activated receptor γ activation pathway in gastric cancer. J Biol Chem, 288(26):18777. doi: 10.1074/jbc.A112.388702.


Saud SM, Young MR, Jones-Hall YL, et al. (2013). Chemo-preventive activity of plant flavonoid isorhamnetin in colorectal cancer is mediated by oncogenic Src and β -catenin. Cancer Res, 73:5473.


Shi C, Fan LY, Cai Z, Liu YY, Yang CL. (2012). Cellular stress response in Eca-109 cells inhibits apoptosis during early exposure to isorhamnetin. Neoplasma, 59(4):361-9. doi: 10.4149/neo_2012_047.


Sun J, Sun G, Meng X, et al. (2013). Isorhamnetin protects against doxorubicin-induced cardiotoxicity in vivo and in vitro. PLoS One, 8(5):e64526. doi: 10.1371/journal.pone.0064526.


Zhu L, Wang ZR, Zhou LM, et al. (2005). Effects and mechanisms of isorhamnetin on lung carcinoma. Space Med Med Eng (Chin), 18:381-383.

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

Quercetin

Cancer: Leukemia, prostate

Action: MDR, apoptosis-inducing

Quercetin is a plant-derived flavonol found in many fruits, vegetables, leaves and grains. It is also found in red wine.

MDR/ Apoptotic-inducing

Natural products from plants such as flavonoids are potential drugs to overcome multi-drug resistance (MDR) in cancer treatments. Quercetin exhibits cytotoxicity against erythroleukemic cells: IC50 are 11.0 +/- 2.0 micromol/L and 5.0 +/- 0.4 micromol/L for K562 and K562/adr, respectively. Quercetin induces cell death via apoptosis in both K562 and K562/adr cells and does not inhibit Pgp-mediated efflux of 99mTc-MIBI. Quercetin (10 micromol/L, 3 h) and etoposide (100 micromol/L, 24 hours) induce similar levels of apoptosis in K562 and K562/adr cells.

Quercetin induces an increase followed by a decrease in inner mitochondrial membrane potential   |DeltaPsim| value depending on its concentration. A decrease in the |DeltaPsim| value is associated with an increase in the percentage of early apoptotic cells.

It is clearly shown that quercetin results in a spontaneous DeltaPsim change during apoptotic induction. Therefore, quercetin is potentially an apoptotic-inducing agent, which reacts at the mitochondrial level (Kothan et al., 2004).

MDR

Expression of the MDR1 gene, which encodes P-glycoprotein, is increased under some stress conditions. It has been reported that quercetin, a bioflavonoid, inhibits the expression of heat-shock proteins. The effects of quercetin have been identified on the MDR1 gene expression in the human hepatocarcinoma cells line, HepG2. The increase of P-glycoprotein synthesis and MDR1 mRNA accumulation caused by exposure to arsenite were inhibited by quercetin. Although many drugs that prevent the P-glycoprotein function have been reported, this is the first report to describe the inhibition of MDR1 expression by a reagent (Kioka et al., 1992).

Leukemia

Leukemia cells were treated with quercetin, after which apoptosis, Mcl-1 expression, and Bax activation and translocation were evaluated. Quercetin-induced apoptosis was accompanied by Mcl-1 down-regulation and Bax conformational change and mitochondrial translocation that triggered cytochrome c release. In vivo administration of quercetin attenuated tumor growth in U937 xenografts. The TUNEL-positive apoptotic cells in tumor sections increased in quercetin-treated mice as compared with controls.

These data suggest that quercetin may be useful for the treatment of leukemia by preferentially inducing apoptosis in leukemia versus normal hematopoietic cells through a process involving Mcl-1 down-regulation, which, in turn, potentiates Bax activation and mitochondrial translocation, culminating in apoptosis (Cheng et al., 2010).

Prostate Cancer

The anti-angiogenic activity of quercetin was probed using ex vivo, in vivo and in vitro models. Angiogenesis is a crucial step in the growth and metastasis of cancers, since it enables the growing tumor to receive oxygen and nutrients. Quercetin (20 mg/kg/d) significantly reduced the volume and the weight of solid tumors in prostate xenograft mouse model, indicating that quercetin inhibited tumorigenesis by targeting angiogenesis.

Furthermore, quercetin reduced the cell viability and induced apoptosis in prostate cancer cells, which were correlated with the down-regulation of AKT, mTOR and P70S6K expressions. Collectively, these results suggest that quercetin inhibits tumor growth and angiogenesis by targeting VEGF-R2 regulated AKT/mTOR/P70S6K signaling pathway, and could be used as a potential drug candidate for cancer therapy (Pratheeshkumar et al., 2012).

References

Cheng SP, Gao N, Zhang Z, et al. (2010). Quercetin Induces Tumor-Selective Apoptosis through Down-regulation of Mcl-1 and Activation of Bax. Clin Cancer Res, 16(23):5679-91. doi: 10.1158/1078-0432.CCR-10-1565


Kioka N, Hosokawa N, Komano T, Hirayoshi K, Nagate K, Ueda K. (1992). Quercetin, a bioflavonoid, inhibits the increase of human Multi-drug resistance gene (< i> MDR1</i>) expression caused by arsenite. FEBS Lett, 301(3):307-9.


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 Pharmacol, 82(12):1084-90.


Pratheeshkumar P, Budhraja A, et al. (2012). Quercetin inhibits angiogenesis mediated human prostate tumor growth by targeting VEGFR- 2 regulated AKT/mTOR/P70S6K signaling pathways. PLoS One, 7(10):e47516. doi: 10.1371/journal.pone.0047516.

Chelerythrine, Chelidonine and Sanguinarine

Cancer:
Leukemia, oral squamous cell carcinoma, melanoma

Action: Cytotoxic, MDR, apoptosis-triggering, inhibits proliferation

Sanguinarine, chelerythrine and chelidonine are isoquinoline alkaloids derived from the greater celandine. They possess a broad spectrum of pharmacological activities. It has been shown that their anti-tumor activity is mediated via different mechanisms, which can be promising targets for anti-cancer therapy. This study focuses on the differential effects of these alkaloids upon cell viability, DNA damage, and nucleus integrity in mouse primary spleen and lymphocytic leukemic cells, L1210.

Data suggests that cytotoxic and DNA-damaging effects of chelerythrine and sanguinarine are more selective against mouse leukemic cells and primary mouse spleen cells, whereas chelidonine blocks proliferation of L1210 cells. The action of chelidonine on normal and tumor cells requires further investigation (Kaminsky, Lin, Filyak, & Stoika, 2008).

MDR

Cancer cells often develop multi-drug resistance (MDR) which is a multidimensional problem involving several mechanisms and targets. This study demonstrates that chelidonine, an alkaloid extract from Chelidonium majus, which contains protoberberine and benzo[c]phenanthridine alkaloids, has the ability to overcome MDR of different cancer cell lines through interaction with ABC-transporters, CYP3A4 and GST, by induction of apoptosis, and cytotoxic effects.

Chelidonine and the alkaloid extract inhibited P-gp/MDR1 activity in a concentration-dependent manner in Caco-2 and CEM/ADR5000 and reversed their doxorubicin resistance. In addition, chelidonine and the alkaloid extract inhibited the activity of the drug, modifying enzymes CYP3A4 and GST in a dose-dependent manner. The expression analysis identified a common set of regulated genes related to apoptosis, cell-cycle, and drug metabolism.

Results suggest that chelidonine is a promising compound for overcoming MDR and enhancing cytotoxicity of chemotherapeutics, especially against leukemia cells. Its efficacy needs to be confirmed in animal models (El-Readi, Eid, Ashour, Tahrani & Wink, 2013).

Induces Apoptosis, Leukemia

Sanguinarine, chelerythrine and chelidonine possess prominent apoptotic effects towards cancer cells. This study found that sanguinarine and chelerythrine induced apoptosis in human CEM T-leukemia cells, accompanied by an early increase in cytosolic cytochrome C that precedes caspases-8, -9 and -3 processing. Effects of sanguinarine and chelerythrine on mitochondria were confirmed by clear changes in morphology (3h), howerver chelidonine did not affect mitochondrial integrity. Sanguinarine and chelerythrine also caused marked DNA damage in cells after 1h, but a more significant increase in impaired cells occurred after 6h. Chelidonine induced intensive DNA damage in 15–20% cells after 24h.

Results demonstrated that rapid cytochrome C release in CEM T-leukemia cells exposed to sanguinarine or chelerythrine was not accompanied by changes in Bax, Bcl-2 and Bcl-X((L/S)) proteins in the mitochondrial fraction, and preceded activation of the initiator caspase-8 (Kaminskyy, Kulachkovskyy, & Stoika, 2008).

Induces Apoptosis

Chelerythrine, formerly identified as a protein kinase C inhibitor, has also been shown to inhibit the anti-apoptotic Bcl-2 family proteins. Chelerythrine initiates the rapid mitochondrial apoptotic death of H9c2 cardiomyoblastoma cells in a manner that is likely independent of the generation of ROS from mitochondria (Funakoshi et al., 2011).

Oral Cancer, Inhibits cell proliferation

The effects of benzo[c] phenanthridine alkaloids (QBA), known mainly as sanguinarine and chelerythrine, on the inhibition of some kinds of cancer cell proliferation have been established. Sanguinarine is a potential inhibitor of tumorigenesis which suggests that it may be valuable in the development of new anti-cancer drugs for the treatment of oral squamous cell carcinoma (OSCC) (Tsukamoto et al., 2011).

Apoptotic Effects; Melanoma

Mixtures of isoquinoline alkaloids containing protopine, chelidonine, sanguinarine, allocryptopine, and stylopine were applied to murine fibroblast NIH/3T3, mouse melanoma B16F10, and human breast cancer MCF7 cell cultures for 20 and 40 min, and the content of alkaloids in the cell media was measured by capillary electrophoresis (CE). CE separation of isoquinoline alkaloids was performed in 30 mM phosphate buffer (pH 2.5). As these alkaloids have native fluorescence, they were directly detected using the commercially available UV light-emitting diode without fluorescent derivatization. The results showed a differential ability of celandine alkaloids to penetrate into the normal and cancer cell interior, which was inversely proportional to their cytotoxic activity.

While the most effective transport of celandine alkaloids from the cell medium to the cell interior was observed for normal murine fibroblast NIH/3T3 cells (about 55% of total content), cytotoxicity tests demonstrated selective and profound apoptotic effects of a five-alkaloid combination in the mouse melanoma B16F10 cell line (Kulp & Bragina, 2013).

Leukemia

The methanol extract isolated from the greater celandine Chelidonium majus L. (CME) has a strong anti-oxidant potential and exerted the anti-proliferative activity via apoptosis on leukemia cells. CME, due to the presence of the isoquinoline alkaloids and the flavonoid components may play an important role in both cancer chemoprevention through its anti-oxidant activity and modern cancer chemotherapy as a cytotoxic and apoptosis-inducing agent (Nadova et al., 2008).

Apoptosis-inducing Activity

Apoptogenic and DNA-damaging effects of chelidonine (CHE) and sanguinarine (SAN), two structurally related benzophenanthridine alkaloids isolated from Chelidonium majus L. (Papaveraceae), were compared. Both alkaloids induced apoptosis in human acute T-lymphoblastic leukaemia MT-4 cells. Apoptosis induction by CHE and SAN in these cells was accompanied by caspase-9 and -3 activation and an increase in the pro-apoptotic Bax protein. An elevation in the percentage of MT-4 cells possessing caspase-3 in active form after their treatment with CHE or SAN was in parallel to a corresponding increase in the fraction of apoptotic cells. CHE, in contrast to SAN, does not interact directly with DNA.

This fact is in line with DNA-damaging effects of the alkaloids detected in the COMET assay. Nevertheless, apoptosis-inducing activity of CHE even slightly exceeded that of SAN (Philchenkov et al., 2008).

Chelidonium majus L. alkaloids chelidonine, sanguinarine, chelerythrine, protopine and allocryptopine were identified as major components of Ukrain. Apart from sanguinarine and chelerythrine, chelidonine turned out to be a potent inducer of apoptosis, triggering cell death at concentrations of 0.001 mM, while protopine and allocryptopine were less effective. Similar to Ukrain, apoptosis signaling of chelidonine involved Bcl-2 controlled mitochondrial alterations and caspase-activation (Habermehl et al., 2006).

References

El-Readi MZ, Eid S, Ashour ML, Tahrani A, & Wink M. (2013). Modulation of Multi-drug resistance in cancer cells by chelidonine and Chelidonium majus alkaloids. Phytomedicine, 20(3-4), 282-94. doi: 10.1016/j.phymed.2012.11.005.


Funakoshi T, Aki T, Nakayama H, et al. (2011). Reactive oxygen species-independent rapid initiation of mitochondrial apoptotic pathway by chelerythrine. Toxicol In Vitro, 25(8):1581-7. doi: 10.1016/j.tiv.2011.05.028.


Habermehl D, Kammerer B, Handrick R, et al. (2006). Pro-apoptotic activity of Ukrain is based on Chelidonium majus L. alkaloids and mediated via a mitochondrial death pathway. BMC Cancer, 6:14.


Kaminskyy V, Lin KW, Filyak Y, & Stoika R. (2008). Differential effect of sanguinarine, chelerythrine and chelidonine on DNA damage and cell viability in primary mouse spleen cells and mouse leukemic cells. Cell Biology International., 32(2), 271-277.


Kaminskyy V, Kulachkovskyy O,Stoika R. (2008). A decisive role of mitochondria in defining rate and intensity of apoptosis induction by different alkaloids. Toxicology Letters, 177(3), 168-81. doi: 10.1016/j.toxlet.2008.01.009.


Kulp M, Bragina O. (2013). Capillary electrophoretic study of the synergistic biological effects of alkaloids from Chelidonium majus L. in normal and cancer cells. Analytical and Bioanalytical Chemistry, 405(10), 3391-7. doi: 10.1007/s00216-013-6755-y.


Nadova S, Miadokova E, Alfoldiova L, et al. (2008). Potential anti-oxidant activity, cytotoxic and apoptosis-inducing effects of Chelidonium majus L. extract on leukemia cells. Neuro Endocrinol Lett, 29(5):649-52.


Philchenkov A., Kaminskyy V., Zavelevich M., Stoika R. (2008). Apoptogenic activity of two benzophenanthridine alkaloids from Chelidonium majus L. does not correlate with their DNA-damaging effects. Toxicology In Vitro, 22(2), 287-95.


Tsukamoto H, Kondo S, Mukudai Y, et al., (2011). Evaluation of anti-cancer activities of benzo[c]phenanthridine alkaloid sanguinarine in oral squamous cell carcinoma cell line. Anti-cancer Res, 31(9):2841-6.


Zhe C, Li-Juan W, Ming Hui W, et al. (2011). Mechanism governing reversal of Multi-drug resistance in human breast carcinoma cells by chelerythrine. Zhongguo Yi Xue Ke Xue Yuan Xue Bao, 33(1):45-50. doi: 10.3881/j.issn.1000-503X.2011.01.010.

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.