Category Archives: anti-apoptotic

Sanguinarine (See also chelerythrine)

Cancer:
Prostate, bladder, breast, colon, melanoma, leukemia

Action: Pro-oxidative, anti-inflammatory, apoptosis induction

AR+/AR- Prostate Cancer

Sanguinarine, a benzophenanthridine alkaloid derived from the bloodroot plant Sanguinaria canadensis (L.), has been shown to possess anti-microbial, anti-inflammatory, anti-cancer and anti-oxidant properties. It has been shown that sanguinarine possesses strong anti-proliferative and pro-apoptotic properties against human epidermoid carcinoma A431 cells and immortalized human HaCaT keratinocytes. Employing androgen-responsive human prostate carcinoma LNCaP cells and androgen-unresponsive human prostate carcinoma DU145 cells, the anti-proliferative properties of sanguinarine against prostate cancer were also examined.

The mechanism of the anti-proliferative effects of sanguinarine against prostate cancer were examined by determining the effect of sanguinarine on critical molecular events known to regulate the cell-cycle and the apoptotic machinery.

A highlight of this study was the fact that sanguinarine induced growth-inhibitory and anti-proliferative effects in human prostate carcinoma cells irrespective of their androgen status. To our knowledge, this is the first study showing the involvement of cyclin kinase inhibitor-cyclin-cyclin-dependent kinase machinery during cell-cycle arrest and apoptosis of prostate cancer cells by sanguinarine. These results suggest that sanguinarine may be developed as an agent for the management of prostate cancer (Adhami et al., 2004).

Breast Cancer

The effects of this compound were examined on reactive oxygen species (ROS) production and its association with apoptotic tumor cell death using a human breast carcinoma MDA-MB-231 cell line. Cytotoxicity was evaluated by trypan blue exclusion methods. Apoptosis was detected using DAPI staining, agarose gel electrophoresis and flow cytometer. The expression levels of proteins were determined by Western blot analyzes and caspase activities were measured using colorimetric assays.

These observations clearly indicate that ROS is involved in the early molecular events in the sanguinarine-induced apoptotic pathway. Data suggests that sanguinarine-induced ROS are key mediators of MMP collapse, which leads to the release of cytochrome c followed by caspase activation, culminating in apoptosis (Choi, Kim, Lee & Choi, 2008).

Leukemia

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. Sanguinarine and chelerythrine produced a dose-dependent increase in DNA damage and cytotoxicity in both primary mouse spleen cells and L1210 cells. Chelidonine did not show a significant cytotoxicity or damage DNA in both cell types, but completely arrested growth of L1210 cells.

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).

T-lymphoblastic Leukemia

Apoptogenic and DNA-damaging effects of chelidonine (CHE) and sanguinarine (SAN), two structurally related benzophenanthridine alkaloids isolated from Chelidonium majus, were compared. Both alkaloids induced apoptosis in human acute T-lymphoblastic leukemia MT-4 cells. Apoptosis induction by CHE and SAN in these cells were 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.

The involvement of the mitochondria in apoptosis induction by both alkaloids was supported by cytochrome C elevation in cytosol, with an accompanying decrease in cytochrome C content in the mitochondrial fraction. At the same time, two alkaloids under study differed drastically in their cell-cycle phase-specific effects, since only CHE arrested MT-4 cells at the G2/M phase. It was previously demonstrated, that CHE, in contrast to SAN, does not interact directly with DNA. (Philchenkov, Kaminskyy, Zavelevich, & Stoika, 2008).

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), however 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).

Colorectal Cancer

The effects of sanguinarine, a benzophenanthridine alkaloid, was examined on reactive oxygen species (ROS) production, and the association of these effects with apoptotic cell death, in a human colorectal cancer HCT-116 cell line. Sanguinarine generated ROS, followed by a decrease in mitochondrial membrane potential (MMP), activation of caspase-9 and -3, and down-regulation of anti-apoptotic proteins, such as Bcl2, XIAP and cIAP-1. Sanguinarine also promoted the activation of caspase-8 and truncation of Bid (tBid).

Observations clearly indicate that ROS, which are key mediators of Egr-1 activation and MMP collapse, are involved in the early molecular events in the sanguinarine-induced apoptotic pathway acting in HCT-116 cells (Han, Kim, Yoo, & Choi, 2013).

Bladder Cancer

Although the effects of sanguinarine, a benzophenanthridine alkaloid, on the inhibition of some kinds of cancer cell growth have been established, the underlying mechanisms are not completely understood. This study investigated possible mechanisms by which sanguinarine exerts its anti-cancer action in cultured human bladder cancer cell lines (T24, EJ, and 5637). Sanguinarine treatment resulted in concentration-response growth inhibition of the bladder cancer cells by inducing apoptosis.

Taken together, the data provide evidence that sanguinarine is a potent anti-cancer agent, which inhibits the growth of bladder cancer cells and induces their apoptosis through the generation of free radicals (Han et al., 2013).

Melanoma

Sanguinarine is a natural isoquinoline alkaloid derived from the root of Sanguinaria canadensis and from other poppy fumaria species, and is known to have a broad spectrum of pharmacological properties. Current study has found that sanguinarine, at low micromolar concentrations, showed a remarkably rapid killing activity against human melanoma cells. Sanguinarine disrupted the mitochondrial transmembrane potential (ΔΨ m), released cytochrome C and Smac/DIABLO from mitochondria to cytosol, and induced oxidative stress. Thus, pre-treatment with the thiol anti-oxidants NAC and GSH abrogated the killing activity of sanguinarine. Collectively, data suggests that sanguinarine is a very rapid inducer of human melanoma caspase-dependent cell death that is mediated by oxidative stress (Burgeiro, Bento, Gajate, Oliveira, & Mollinedo, 2013).

References

Adhami YM, Aziz MH, Reagan-Shaw SR, et al. (2004). Sanguinarine causes cell-cycle blockade and apoptosis of human prostate carcinoma cells via modulation of cyclin kinase inhibitor-cyclin-cyclin-dependent kinase machinery. Mol Cancer Ther, 3:933


Burgeiro A, Bento AC, Gajate C, Oliveira PJ, Mollinedo F. (2013). Rapid human melanoma cell death induced by sanguinarine through oxidative stress. European Journal of Pharmacology, 705(1-3), 109-18. doi: 10.1016/j.ejphar.2013.02.035.


Choi WY, Kim GY, Lee WH, Choi YH. (2008). Sanguinarine, a benzophenanthridine alkaloid, induces apoptosis in MDA-MB-231 human breast carcinoma cells through a reactive oxygen species-mediated mitochondrial pathway. Chemotherapy, 54(4), 279-87. doi: 10.1159/000149719.


Han MH, Kim GY, Yoo YH, Choi YH. (2013). Sanguinarine induces apoptosis in human colorectal cancer HCT-116 cells through ROS-mediated Egr-1 activation and mitochondrial dysfunction. Toxicology Letters, 220(2), 157-66. doi: 10.1016/j.toxlet.2013.04.020.


Han MH, Park C, Jin CY, et al. (2013). Apoptosis induction of human bladder cancer cells by sanguinarine through reactive oxygen species-mediated up-regulation of early growth response gene-1. PLoS One, 8(5), e63425. doi: 10.1371/journal.pone.0063425.


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.


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.

Rosmarinic Acid

Cancer: Leukemia

Action: Anti-oxidative, MDR

Leukemia

Because tumor necrosis factor-alpha (TNF-alpha) is well known to induce inflammatory responses, its clinical use is limited in cancer treatment. Rosmarinic acid (RA), a naturally occurring polyphenol flavonoid, has been reported to inhibit TNF-alpha-induced NF-kappaB activation in human dermal fibroblasts. Investigation found that RA treatment significantly sensitizes TNF-alpha-induced apoptosis in human leukemia U937 cells through the suppression of nuclear transcription factor-kappaB (NF-kappaB) and reactive oxygen species (ROS). This inhibition was correlated with suppression of NF-kappaB-dependent anti-apoptotic proteins (IAP-1, IAP-2, and XIAP). RA treatment also normalized TNF-alpha-induced ROS generation. Additionally, ectopic Bcl-2 expressing U937 reversed combined treatment-induced cell death, cytochrome c release into cytosol, and collapse of mitochondrial potential.

Results demonstrated that RA inhibits TNF-alpha-induced ROS generation and NF-kappaB activation, and enhances TNF-alpha-induced apoptosis (Moon, Kim, Lee, Choi, & Kim, 2010).

MDR

The intracellular accumulation of adriamycin, rhodamine123 (Rh123), and the expression of P-glycoprotein (P-gp) were assayed by flow cytometry. The influence of RA on the transcription of MDR1 gene was determined by reverse transcription-polymerase chain reaction. The results showed that RA could reverse the MDR of SGC7901/Adr cells, increase the intracellular accumulation of Adr and Rh123, and decrease the transcription of MDR1 gene and the expression of P-gp in SGC7901/Adr cells (Li et al., 2013).

Anti-cancer

Rosmarinic acid (RA), one of the major components of polyphenol, possesses attractive remedial features. Supplementation with RA significantly reduced the formation of aberrant crypt foci (ACF) and ACF multiplicity in 1,2-dimethylhydrazine (DMH) treated rats. Moreover RA supplementation prevented the alterations in circulatory anti-oxidant enzymes and colonic bacterial enzymes activities. Overall, results showed that all three doses of RA inhibited carcinogenesis, though the effect of the intermediary dose of 5 mg/kg b.w. was more pronounced (Karthikkumar et al., 2012).

References

Karthikkumar V, Sivagami G, Vinothkumar R, Rajkumar D, Nalini N. (2012). Modulatory efficacy of rosmarinic acid on premalignant lesions and anti-oxidant status in 1,2-dimethylhydrazine induced rat colon carcinogenesis. Environ Toxicol Pharmacol, 34(3):949-58. doi: 10.1016/j.etap.2012.07.014.


Li FR, Fu YY, Jiang DH, et al. (2013). Reversal effect of rosmarinic acid on Multi-drug resistance in SGC7901/Adr cell. J Asian Nat Prod Res, 15(3):276-85. doi: 10.1080/10286020.2012.762910.


Moon DO, Kim MO, Lee JD, Choi YH, Kim GY. (2010). Rosmarinic acid sensitizes cell death through suppression of TNF-alpha-induced NF-kappaB activation and ROS generation in human leukemia U937 cells. Cancer Letters, 288(2), 183-191. doi: 10.1016/j.canlet.2009.06.033.

RG3 (See also Ginsenosides)

Cancer: Glioblastoma, prostate, breast, colon

Action: Anti-angiogenesis, MDR, enhances chemotherapy, MDR, enhanced paclitaxel absorption, anti-metastatic

RG3 is a ginsenoside isolated from red ginseng (Panax ginseng (L.)), after being peeled, heated, and dried.

Angiosuppressive Activity

Aberrant angiogenesis is an essential step for the progression of solid tumors. Thus anti-angiogenic therapy is one of the most promising approaches to control tumor growth.

Rg3 was found to inhibit the proliferation of human umbilical vein endothelial cells (HUVEC) with an IC50 of 10 nM in Trypan blue exclusion assay.

Rg3 (1-10(3) nM) also dose-dependently suppressed the capillary tube formation of HUVEC on the Matrigel in the presence or absence of 20 ng/ml vascular endothelial growth factor (VEGF). The Matrix metalloproteinases (MMPs), such as MMP-2 and MMP-9, which play an important role in the degradation of basement membrane in angiogenesis and tumor metastasis present in the culture supernatant of Rg3-treated aortic ring culture were found to decrease in their gelatinolytic activities. Taken together, these data underpin the anti-tumor properties of Rg3 through its angiosuppressive activity (Yue et al., 2006).

Glioblastoma

Rg3 has been reported to exert anti-cancer activities through inhibition of angiogenesis and cell proliferation. The mechanisms of apoptosis by ginsenoside Rg3 were related with the MEK signaling pathway and reactive oxygen species. Our data suggest that ginsenoside Rg3 is a novel agent for the chemotherapy of glioblastoma multiforme (GBM) (Choi et al., 2013).

Sin, Kim, & Kim (2012) report that chronic treatment with Rg3 in a sub-lethal concentration induced senescence-like growth arrest in human glioma cells. Rg3-induced senescence was partially rescued when the p53/p21 pathway was inactivated. Data indicate that Rg3 induces senescence-like growth arrest in human glioma cancer through the Akt and p53/p21-dependent signaling pathways.

MDR/Enhanced Paclitaxel Absorption

The penetration of paclitaxel through the Caco-2 monolayer from the apical side to the basal side was facilitated by 20(s)-ginsenoside Rg3 in a concentration-dependent manner. Rg3 also inhibited P-glycoprotein (P-gp), and the maximum inhibition was achieved at 80 µM (p < 0.05). The relative bioavailability (RB)% of paclitaxel with 20(s)-ginsenoside Rg3 was 3.4-fold (10 mg/kg) higher than that of the control. Paclitaxel (20 mg/kg) co-administered with 20(s)-ginsenoside Rg3 (10 mg/kg) exhibited an effective anti-tumor activity with the relative tumor growth rate (T/C) values of 39.36% (p <0.05).

The results showed that 20(s)-ginsenoside Rg3 enhanced the oral bioavailability of paclitaxel in rats and improved the anti-tumor activity in nude mice, indicating that oral co-administration of paclitaxel with 20(s)-ginsenoside Rg3 could provide an effective strategy in addition to the established i.v. route (Yang et al., 2012).

Prostate Cancer

The anti-proliferation effect of Rg3 on prostate cancer cells has been well reported. Rg3 treatment triggered the activation of p38 MAPK; and SB202190, a specific inhibitor of p38 MAPK, antagonized the Rg3-induced regulation of AQP1 and cell migration, suggesting a crucial role for p38 in the regulation process. Rg3 effectively suppresses migration of PC-3M cells by down-regulating AQP1 expression through p38 MAPK pathway and some transcription factors acting on the AQP1 promoter (Pan et al., 2012).

Enhances Chemotherapy

The clinical use of cisplatin (cis-diamminedichloroplatinum II) has been limited by the frequent emergence of cisplatin-resistant cell populations and numerous other adverse effects. Therefore, new agents are required to improve the therapy and health of cancer patients. Oral administration of ginsenoside Rg3 significantly inhibited tumor growth and promoted the anti-neoplastic efficacy of cisplatin in mice inoculated with CT-26 colon cancer cells. In addition, Rg3 administration remarkably inhibited cisplatin-induced nephrotoxicity, hepatotoxicity and oxidative stress.

Rg3 promotes the efficacy of cisplatin by inhibiting HO-1 and NQO-1 expression in cancer cells and protects the kidney and liver against tissue damage by preventing cisplatin-induced intracellular ROS generation (Lee et al., 2012).

Colon Cancer

Rg3-induced apoptosis in HT-29 cells is mediated via the AMPK signaling pathway, and that 20(S)-Rg3 is capable of inducing apoptosis in colon cancer. Rg3-treated cells displayed several apoptotic features, including DNA fragmentation, proteolytic cleavage of poly (ADP-ribose) polymerase (PARP) and morphological changes. 20(S)-Rg3 down-regulated the expression of anti-apoptotic protein B-cell CLL/lymphoma 2 (Bcl2), up-regulated the expression of pro-apoptotic protein of p53 and Bcl-2-associated X protein (Bax), and caused the release of mitochondrial cytochrome c, PARP, caspase-9 and caspase-3 (Yuan et al., 2010).

Anti-metastatic

Studies have linked Rg3 with anti-metastasis of cancer in vivo and in vitro and the CXC receptor 4 (CXCR4) is a vital molecule in migration and homing of cancer to the docking regions. At a dosage without obvious cytotoxicity, Rg3 treatment elicits a weak CXCR4 stain color, decreases the number of migrated cells in CXCL12-elicited chemotaxis and reduces the width of the scar in wound healing and Rg3 is a new CXCR4 inhibitor (Chen et al., 2011).

References

Chen XP, Qian LL, Jiang H, Chen JH. (2011). Ginsenoside Rg3 inhibits CXCR4 expression and related migrations in a breast cancer cell line. Int J Clin Oncol, 16(5):519-23. doi: 10.1007/s10147-011-0222-6.


Choi YJ, Lee HJ, Kang DW, et al. (2013). Ginsenoside Rg3 induces apoptosis in the U87MG human glioblastoma cell line through the MEK signaling pathway and reactive oxygen species. Oncol Rep. doi: 10.3892/or.2013.2555.


Lee CK, Park KK, Chung AS, Chung WY. (2012). Ginsenoside Rg3 enhances the chemosensitivity of tumors to cisplatin by reducing the basal level of nuclear factor erythroid 2-related factor 2-mediated heme oxygenase-1/NAD(P)H quinone oxidoreductase-1 and prevents normal tissue damage by scavenging cisplatin-induced intracellular reactive oxygen species. Food Chem Toxicol, 50(7):2565-74. doi: 10.1016/j.fct.2012.01.005.


Pan XY, Guo H, Han J, et al. (2012). Ginsenoside Rg3 attenuates cell migration via inhibition of aquaporin 1 expression in PC-3M prostate cancer cells. Eur J Pharmacol, 683(1-3):27-34. doi: 10.1016/j.ejphar.2012.02.040.


Sin S, Kim SY, Kim SS. (2012). Chronic treatment with ginsenoside Rg3 induces Akt-dependent senescence in human glioma cells. Int J Oncol., 41(5):1669-74. doi: 10.3892/ijo.2012.1604.


Yang LQ, Wang B, Gan H, et al. (2012). Enhanced oral bioavailability and anti-tumor effect of paclitaxel by 20(s)-ginsenoside Rg3 in vivo. Biopharm Drug Dispos., 33(8):425-36. doi: 10.1002/bdd.1806.


Yuan HD, Quan HY, Zhang Y, et al. (2010). 20(S)-Ginsenoside Rg3-induced apoptosis in HT-29 colon cancer cells is associated with AMPK signaling pathway. Mol Med Rep., 3(5):825-31. doi: 10.3892/mmr.2010.328.


Yue PY, Wong DY, Wu PK, et al. (2006). The angiosuppressive effects of 20 (R)-ginsenoside Rg3. Biochem Pharmacol, 72(4):437-45.

Phenolics

Cancer: Prostate

Action: Chemo-preventive, anti-oxidant, modulate insulin-like growth factor-I (IGF-I)

Natural phenolic compounds play an important role in cancer prevention and treatment. Phenolic compounds from medicinal herbs and dietary plants include phenolic acids, flavonoids, tannins, stilbenes, curcuminoids, coumarins, lignans, quinones, and others. Various bioactivities of phenolic compounds are responsible for their chemo-preventive properties (e.g. anti-oxidant, anti-carcinogenic, or anti-mutagenic and anti-inflammatory effects) and also contribute to their inducing apoptosis by arresting cell-cycle, regulating carcinogen metabolism and ontogenesis expression, inhibiting DNA binding and cell adhesion, migration, proliferation or differentiation, and blocking signaling pathways. A review by Huang et al., (2010) covers the most recent literature to summarize structural categories and molecular anti-cancer mechanisms of phenolic compounds from medicinal herbs and dietary plants (Huang, Cai, & Zhang., 2010).

Phenolics are compounds possessing one or more aromatic rings bearing one or more hydroxyl groups with over 8,000 structural variants, and generally are categorized as phenolic acids and analogs, flavonoids, tannins, stilbenes, curcuminoids, coumarins, lignans, quinones, and others based on the number of phenolic rings and of the structural elements that link these rings (Fresco et al., 2006).

Phenolic Acids

Phenolic acids are a major class of phenolic compounds, widely occurring in the plant kingdom.   Predominant phenolic acids include hydroxybenzoic acids (e.g. gallic acid, p-hydroxybenzoic acid, protocatechuic acid, vanillic acid, and syringic acid) and hydroxycinnamic acids (e.g. ferulic acid, caffeic acid, p-coumaric acid, chlorogenic acid, and sinapic acid). Natural phenolic acids, either occurring in the free or conjugated forms, usually appear as esters or amides.

Due to their structural similarity, several other polyphenols are considered as phenolic acid analogs such as capsaicin, rosmarinic acid, gingerol, gossypol, paradol, tyrosol, hydroxytyrosol, ellagic acid, cynarin, and salvianolic acid B (Fresco et al., 2006; Han et al., 2007).

Gallic acid is widely distributed in medicinal herbs, such as Barringtonia racemosa, Cornus officinalis, Cassia auriculata, Polygonum aviculare, Punica granatum, Rheum officinale, Rhus chinensis, Sanguisorba officinalis, and Terminalia chebula as well as dietary spices, for example, thyme and clove. Other hydroxybenzoic acids are also ubiquitous in medicinal herbs and dietary plants (spices, fruits, vegetables).

For example, Dolichos biflorus, Feronia elephantum, and Paeonia lactiflora contain hydroxybenzoic acid; Cinnamomum cassia, Lawsonia inermis, dill, grape, and star anise possess protocatechuic acid; Foeniculum vulgare, Ipomoea turpethum, and Picrorhiza scrophulariiflora have vanillic acid; Ceratostigma willmottianum and sugarcane straw possess syringic acid (Cai et al., 2004; Shan et al., 2005; Sampietro & Vattuone, 2006; Stagos et al., 2006; Surveswaran et al., 2007).

Ferulic, caffeic, and p-coumaric acid are present in many medicinal herbs and dietary spices, fruits, vegetables, and grains (Cai et al., 2004). Wheat bran is a good source of ferulic acids. Free, soluble-conjugated, and bound ferulic acids in grains are present in the ratio of 0.1:1:100. Red fruits (blueberry, blackberry, chokeberry, strawberry, red raspberry, sweet cherry, sour cherry, elderberry, black currant, and red currant) are rich in hydroxycinnamic acids (caffeic, ferulic, p-coumaric acid) and p-hydroxybenzoic, ellagic acid, which contribute to their anti-oxidant activity (Jakobek et al., 2007).

Chlorogenic acids are the ester of caffeic acids and are the substrate for enzymatic oxidation leading to browning, particularly in apples and potatoes. Chlorogenic acid is a major phenolic acid from medicinal plants especially in the species of Apocynaceae and Asclepiadaceae (Huang et al., 2007).

Salvianolic acid B is a major water-soluble polyphenolic acid extracted from Radix salviae miltiorrhizae, which is a common herbal medicine clinically used as an anti-oxidant agent for thousands of years in China. There are 9 activated phenolic hydroxyl groups that may be responsible for the release of active hydrogen to block lipid peroxidation reaction. Rosmarinic acid is an anti-oxidant phenolic compound, which is found in many dietary spices such as mint, sweet basil, oregano, rosemary, sage, and thyme.

Gossypol, a polyphenolic aldehyde, derived from the seeds of the cotton plant (genus Gossypium, family Malvaceae), has contraceptive activity and can cause hypokalemia in some men. Gingerol, a phenolic substance, is responsible for the spicy taste of ginger.

Polyphenols

Polyphenols are a structural class of mainly natural, organic chemicals characterized by the presence of large multiples of phenol structural units. The number and characteristics of these phenol structures underlie the unique physical, chemical, and biological (metabolic, toxic, therapeutic, etc.) properties of particular members of the class. They may be broadly classified as phenolic acids, flavonoids, stilbenes, and lignans (Manach et al., 2004).

Initial evidence on cancer came from epidemiologic studies suggesting that a diet that includes regular consumption of fruits and vegetables (rich in polyphenols) significantly reduces the risk of many cancers.

Polyphenolic cancer action can be attributed not only to their ability to act as anti-oxidants but also to their ability to interact with basic cellular mechanisms. Such interactions include interference with membrane and intracellular receptors, modulation of signaling cascades, interaction with the basic enzymes involved in tumor promotion and metastasis, interaction with oncogenes and oncoproteins, and, finally, direct or indirect interactions with nucleic acids and nucleoproteins. These actions involve almost the whole spectrum of basic cellular machinery – from the cell membrane to signaling cytoplasmic molecules and to the major nuclear components – and provide insights into their beneficial health effects (Kampa et al., 2007).

Polyphenols and Copper

Anti-cancer polyphenolic nutraceuticals from fruits, vegetables, and spices are generally recognized as anti-oxidants, but can be pro-oxidants in the presence of copper ions. Through multiple assays, Khan et al. (2013) show that polyphenols luteolin, apigenin, epigallocatechin-3-gallate, and resveratrol are able to inhibit cell proliferation and induce apoptosis in different cancer cell lines. Such cell death is prevented to a significant extent by cuprous chelator neocuproine and reactive oxygen species scavengers. We also show that normal breast epithelial cells, cultured in a medium supplemented with copper, become sensitized to polyphenol-induced growth inhibition.

Since the concentration of copper is significantly elevated in cancer cells, their results strengthen the idea that an important anti-cancer mechanism of plant polyphenols is mediated through intracellular copper mobilization and reactive oxygen species generation leading to cancer cell death. Moreover, this pro-oxidant chemo-preventive mechanism appears to be a mechanism common to several polyphenols with diverse chemical structures and explains the preferential cytotoxicity of these compounds toward cancer cells.

IGF-1; Prostate Cancer

The ability of polyphenols from tomatoes and soy (genistein, quercetin, kaempferol, biochanin A, daidzein and rutin) were examined for their ability to modulate insulin-like growth factor-I (IGF-I)–induced in vitro proliferation and apoptotic resistance in the AT6.3 rat prostate cancer cell line. IGF-I at 50 µg/L in serum-free medium produced maximum proliferation and minimized apoptosis. Genistein, quercetin, kaempferol and biochanin A exhibited dose-dependent inhibition of growth with a 50% inhibitory concentration (IC50) between 25 and 40 µmol/L, whereas rutin and daidzein were less potent with an IC50 of >60 µmol/L. Genistein and kaempferol potently induced G2/M cell-cycle arrest.

Genistein, quercetin, kaempferol and biochanin A, but not daidzein and rutin, counteracted the anti-apoptotic effects of IGF-I. Human prostate epithelial cells grown in growth factor-supplemented medium were also sensitive to growth inhibition by polyphenols. Genistein, biochanin A, quercetin and kaempferol reduced the insulin receptor substrate-1 (IRS-1) content of AT6.3 cells and prevented the down-regulation of IGF-I receptor β in response to IGF-I binding.

Several polyphenols suppressed phosphorylation of AKT and ERK1/2, and more potently inhibited IRS-1 tyrosyl phosphorylation after IGF-I exposure. In summary, polyphenols from soy and tomato products may counteract the ability of IGF-I to stimulate proliferation and prevent apoptosis via inhibition of multiple intracellular signaling pathways involving tyrosine kinase activity (Wang et al., 2003).

Flavonoids

Flavonoids have been linked to reducing the risk of major chronic diseases including cancer because they have powerful anti-oxidant activities in vitro, being able to scavenge a wide range of reactive species (e.g. hydroxyl radicals, peroxyl radicals, hypochlorous acid, and superoxide radicals) (Hollman & Katan, 2000).

Flavonoids are a group of more than 4,000 phenolic compounds that occur naturally in plants (Ren et al., 2003). These compounds commonly have the basic skeleton of phenylbenzopyrone structure (C6-C3-C6) consisting of 2 aromatic rings (A and B rings) linked by 3 carbons that are usually in an oxygenated central pyran ring, or C ring (12). According to the saturation level and opening of the central pyran ring, they are categorized mainly into flavones (basic structure, B ring binds to the 2 position), flavonols (having a hydroxyl group at the 3 position), flavanones (dihydroflavones) and flavanonols (dihydroflavonols; 2–3 bond is saturated), flavanols (flavan-3-ols and flavan-3,4-diols; C-ring is 1-pyran), anthocyanins (anthocyanidins; C-ring is 1-pyran, and 1–2 and 3–4 bonds are unsaturated), chalcones (C-ring is opened), isoflavonoids (mainly isoflavones; B ring binds to the 3 position), neoflavonoids (B ring binds to the 4-position), and biflavonoids (dimer of flavones, flavonols, and flavanones) (Iwashina, 2000; Cai et al., 2004; Cai et al., 2006; Ren et al., 2003)

Tannins

Tannins are natural, water-soluble, polyphenolic compounds with molecular weight ranging from 500 to 4,000, usually classified into 2 classes: hydrolysable tannins (gallo- and ellagi-tannins) and condensed tannins (proanthocyanidins) (Cai et al., 2004).

The former are complex polyphenols, which can be degraded into sugars and phenolic acids through either pH changes or enzymatic or nonenzymatic hydrolysis. The basic units of hydrolysable tannins of the polyster type are gallic acid and its derivatives (Fresco et al., 2006). Tannins are commonly found combined with alkaloids, polysaccharides, and proteins, particularly the latter (Han et al., 2007).

Stilbenes

Stilbenes are phenolic compounds displaying 2 aromatic rings linked by an ethane bridge, structurally characterized by the presence of a 1,2-diarylethene nucleus with hydroxyls substituted on the aromatic rings. They are distributed in higher plants and exist in the form of oligomers and in monomeric form (e.g. resveratrol, oxyresveratrol) and as dimeric, trimeric, and polymeric stilbenes or as glycosides.

The well-known compound, trans-resveratrol, a phytoalexin produced by plants, is the member of this chemical famil most abundant in the human diet (especially rich in the skin of red grapes), possessing a trihydroxystilben skeleton (Han et al., 2007). There are monomeric stilbenes in 4 species of medicinal herbs, that is, trans-resveratrol in root of Polygonum cuspidatum, Polygonum multiflorum, and P. lactiflora; piceatannol in root of P. multiflorum; and oxyresveratrol in fruit of Morus alba (Cai et al., 2006).

It was reported that dimeric stilbenes and stilbene glycosides were identified from these species (Xiao et al., 2002). In addition, 40 stilbene oligomers were isolated from 6 medicinal plant species (Shorea hemsleyana, Vatica rassak, Vatica indica, Hopea utilis, Gnetum parvifolium, and Kobresia nepalensis). Other stilbenes that have recently been identified in dietary sources, such as piceatannol and its glucoside (usually named astringin) and pterostilbene, are also considered as potential chemo-preventive agents. These and other in vitro and in vivo studies provide a rationale in support of the use of stilbenes as phytoestrogens to protect against hormone-dependent tumors (Athar et al., 2007).

Curcuminoids

Curcuminoids are ferulic acid derivatives, which contain 2 ferulic acid molecules linked by a methylene with a β -diketone structure in a highly conjugated system. Curcuminoids and ginerol analogues are natural phenolic compounds from plants of the family Zingiberaceae. Curcuminoids include 3 main chemical compounds: curcumin, demethoxycurcumin, and bisdemethoxycurcumin (Cai et al., 2006). All 3 curcuminoids impart the characteristic yellow color to turmeric, particularly to its rhizome, and are also major yellow pigments of mustard. Curcuminoids containing Curcuma longa (turmeric) and ginerol analogues containing Zingiber officinale (ginger) are not only used as Chinese traditional medicines but also as natural color agents or ordinary spices.

In addition, curcuminoids with anti-oxidant properties have been isolated from various Curcuma or Zingiber species, such as the Indian medicinal herb Curcuma xanthorrhiza.

Coumarins

Coumarins are lactones obtained by cyclization of cis-ortho-hydroxycinnamic acid, belonging to the phenolics with the basic skeleton of C6+ C3. This precursor is formed through isomerization and hydroxylation of the structural analogs trans-hydroxycinnamic acid and derivatives. Coumarins are present in plants in the free form and as glycosides. In general, coumarins are characterized by great chemical diversity, mainly differing in the degree of oxygenation of their benzopyrane moiety.

In nature, most coumarins are C7-hydroxylated (Fresco et al., 2006; Cai et al., 2006). Major coumarin constituents included simple hydroxylcoumarins (e.g. aesculin, esculetin, scopoletin, and escopoletin), furocoumarins and isofurocoumarin (e.g. psoralen and isopsoralen from Psoralea corylifolia), pyranocoumarins (e.g. xanthyletin, xanthoxyletin, seselin, khellactone, praeuptorin A), bicoumarins, dihydro-isocoumarins (e.g. bergenin), and others (e.g. wedelolactone from Eclipta prostrata) (Shan et al., 2005).

Plants, fruits, vegetables, olive oil, and beverages (coffee, wine, and tea) are all dietary sources of coumarins; for example, seselin from fruit of Seseli indicum, khellactone from fruit of Ammi visnaga, and praeuptorin A from Peucedanum praeruptorum (Sonnenberg et al., 1995). In previous studies, it was found that coumarins occurred in the medicinal herbs Umbelliferae, Asteraceae, Convolvulaceae, Leguminosae, Magnoliaceae, Oleaceae, Rutaceae, and Ranunculaceae, such as simple coumarins from A. annua, furocoumarins (5-methoxyfuranocoumarin) from Angelica sinensis, pyranocoumarins from Citrus aurantium, and isocoumarins from Agrimonia pilosa. Coumarins have also been detected in some Indian medicinal plants (e.g. Toddalia aculeata, Murraya exotica, Foeniculum vulgare, and Carum copticum) and dietary spices (e.g. cumin and caraway). In addition, coumestans, derivatives of coumarin, including coumestrol, a phytoestrogen, are found in a variety of medicinal and dietary plants such as soybeans and Pueraria mirifica (Chansakaow et al., 2000).

Lignans

Lignans are also derived from cis-o-hydroxycinnamic acid and are dimers (with 2 C6-C3 units) resulting from tail–tail linkage of 2 coniferl or sinapyl alcohol units (Cai et al., 2007). Lignans are mainly present in plants in the free form and as glycosides in a few (Fresco et al., 2006). Main lignan constituents are lignanolides (e.g. arctigenin, arctiin, secoisolariciresinol, and matairesinol from Arctium lappa), cyclolignanolides (e.g. chinensin from Polygala tenuifolia), bisepoxylignans (e.g. forsythigenol and forsythin from Forsythia suspensa), neolignans (e.g. magnolol from Cedrus deodara and Magnolia officinalis), and others (e.g. schizandrins, schizatherins, and wulignan from Schisandra chinensis; pinoresinol from Pulsatilla chinensis; and furofuran lignans from Cuscuta chinensis) (Surveswaran et al., 2007).

The famous tumor therapy drug podophyllotoxin (cyclolignanolide) was first identified in Podophyllum peltatum, which Native Americans used to treat warts, and also found in a traditional medicinal plant Podophyllum emodi var. chinense (Efferth et al., 2007). Two new lignans (podophyllotoxin glycosides) were isolated from the Chinese medicinal plant, Sinopodophyllum emodi (Zhao et al., 2002). Different lignans (e.g. cubebin, hinokinin, yatein, and isoyatein) were identified from leaves, berries, and stalks of Piper cubeba L. (Piperaceae), an Indonesian medicinal plant (Elfahmi et al., 2007).

Milder et al. (2005) established a lignan database from Dutch plant foods by quantifying lariciresinol, pinoresinol, secoisolariciresinol, and matairesinol in 83 solid foods and 26 beverages commonly consumed in The Netherlands. They reported that flaxseed (mainly secoisolariciresinol), sesame seeds, and Brassica vegetables (mainly pinoresinol and lariciresinol) contained unexpectedly high levels of lignans. Sesamol, sesamin, and their glucosides are also good examples of this type of compound, which comes from sesame oil and sunflower oil.

Quinones

Natural quinones in medicinal plants fall into 4 categories: anthraquinones, phenanthraquinones, naphthoquinones, and benzoquinones (Cai et al., 2004). Anthraquinones are the largest class of natural quinones and occur more widely in medicinal and dietary plants than other natural quinones (Cai et al., 2006). The hydroxyanthraquinones normally have 1 to 3 hydroxyl groups on the anthraquinone structure. Previous investigation found that quinones were distributed in 12 species of medicinal herbs from 9 families such as Polygalaceae, Rubiaceae, Boraginaceae, Labiatae, Leguminosae, Myrsinaceae, and so forth (Surveswaran et al., 2007).

For example, high content benzoquinones and derivatives (embelin, embelinol, embeliaribyl ester, embeliol) are found in Indian medicinal herb Embelia ribes; naphthoquinones (shikonin, alkannan, and acetylshikonin) come from Lithospermum erythrorhizon and juglone comes from Juglans regia; phenanthraquinones (tanshinone I, II A, and II B ) were detected in Salvia miltiorrhiza; denbinobin was detected in Dendrobium nobile; and many anthraquinones and their glycosides (e.g. rhein, emodin, chrysophanol, aloe-emodin, physcion, purpurin, pseudopurpurin, alizarin, munjistin, emodin-glucoside, emodin-malonyl-glucoside, etc.) were identified in the rhizomes and roots from P. cuspidatum (also in leaves), P. multiflorum, and R. officinale in the Polygalaceae and Rubia cordifolia in the Rubiaceae (Surveswaran et al., 2007; Huang et al., 2008). In addition, some naphthoquinones were isolated from maize (Zea mays L.) roots (Luthje et al., 1998).

References:

Athar M, Back JH, Tang XW, et al. (2007). Resveratrol: a review of preclinical studies for human cancer prevention. Toxicol Appl Pharm, 224:274–283.


Cai YZ, Luo Q, Sun M and Corke H. (2004). Anti-oxidant activity and phenolic compounds of 112 traditional Chinese medicinal plants associated with anticancer. Life Sci, 74:2157–2184.


Cai YZ, Sun M, Xing J, Luo Q and Corke H. (2006). Structure-radical scavenging activity relationships of phenolic compounds from traditional Chinese medicinal plants. Life Sci, 78:2872–2888.


Chansakaow S, Ishikawa T, Seki H, et al. (2000). Identification of deoxymiroestrol as the actual rejuvenating principle of 'Kwao Keur', Pueraria mirifica. J. Nat. Prod, 63(2):173–5. doi:10.1021/np990547v.


Efferth T, Li P CH, Konkimalla V and Kaina B. (2007). From traditional Chinese medicine to rational cancer therapy. Trends Mol Med, 13:353–361.


Elfahmi, Ruslan K, Batterman S, et al. (2007). Lignan profile of Piper cubeba, an Indonesian medicinal plant. Biochem Syst Ecol, 35:397–402.


Fresco P, Borges F, Diniz C and Marques M PM. (2006). New insights on the anti-cancer properties of dietary polyphenols. Med Res Rev, 26:747–766.


Han XZ, Shen T and Lou HX. (2007). Dietary polyphenols and their biological significance. Int J Mol Sci, 8:950–988


Hollman P and Katan M B. (2000). Flavonols, flavones, and flavanols—nature, occurrence, and dietary burden. J Sci Food Agric, 80:1081–1093.


Huang WY, Cai YZ, Xing J, Corke H and Sun M. (2007). A potential anti-oxidant resource: endophytic fungi isolated from traditional Chinese medicinal plants. Econ Bot, 61:14–30.


Huang WY, Cai YZ, Xing J, Corke H and Sun M. (2008). Comparative analysis of bioactivities of four Polygonum species. Planta Med, 74:43–49.


Huang WH, Cai YZ, Zhang Y. (2010). Natural Phenolic Compounds From Medicinal Herbs and Dietary Plants: Potential Use for Cancer Prevention. Nutrition and Cancer, 62(1):1–20 doi: 10.1080/01635580903191585


Iwashina T. (2000). The structure and distribution of the flavonoids in plants. J Plant Res, 113:287–299.


Jakobek L, Seruga M, Novak I and Medvidovic-Kosanovic M. (2007). Flavonols, phenolic acids, and anti-oxidant activity of some red fruits. Deut Lebensm-Runsch, 103:369–378.


Kampa M, Nifli AP, Notas G, Castanas E. (2007). Polyphenols and cancer cell growth. Rev Physiol Biochem Pharmacol, 159:79-113.


Khan HY, Zubair H, Faisal M, et al. (2013). Plant polyphenol induced cell death in human cancer cells involves mobilization of intracellular copper ions and reactive oxygen species generation: A mechanism for cancer chemo-preventive action. Mol Nutr Food Res. doi: 10.1002/mnfr.201300417.


Luthje S, Van Gestelen P, Cordoba-Pedregosa MC, et al. (1998). Quinones in plant plasma membranes—a missing link?. Protoplasma, 205:43–51.


Manach C, Scalbert A, Morand C, RŽmŽsy C, JimŽnez L. (2004). Polyphenols: food sources and bioavailability. Am J Clin Nutr, 79: 727–47.


Milder I, Arts I, van de Putte B, Venema DP and Hollman P. (2005). Lignan contents of Dutch plant foods: a database including lariciresinol, pinoresinol, secoisolariciresinol and matairesinol. Brit J Nutr, 93:393–402.


Ren WY, Qiao ZH, Wang HW, Zhu L and Zhang L. (2003). Flavonoids: promising anti-cancer agents. Med Res Rev, 23:519–534.


Sampietro DA and Vattuone MA. (2006). Sugarcane straw and its phytochemicals as growth regulators of weed and crop plants. Plant Growth Regul, 48: 21–27.


Shan B, Cai YZ, Sun M and Corke H. (2005). Anti-oxidant capacity of 26 spice extracts and characterization of their phenolic constituents. J Agric Food Chem, 53:7749–7759.


Sonnenberg H, Kaloga M, Eisenbac N and Fromming KK. (1995). Isolation and characterization of an angular-type dihydropyranocoumaringlycoside from the fruits of Ammi visnaga (L) Lam (Apiaceae). Zeitschrift Natur C-A J BioSci, 50: 729–731.


Stagos D, Kazantzoglou, G, Theofanidou, D, Kakalopoulou, G, Magiatis, P. (2006). Activity of grape extracts from Greek varieties of Vitis vinifera against mutagenicity induced by bleomycin and hydrogen peroxide in Salmonella typhimurium strain TA102. Mutat Res-Gen Tox En, 609:165–175.


Surveswaran S, Cai YZ, Corke H and Sun M. (2007). Systematic evaluation of natural phenolic anti-oxidants from 133 Indian medicinal plants. Food Chem, 102:938–953.


Wang SH, DeGroff VL, Clinton SK. (2003). Tomato and Soy Polyphenols Reduce Insulin-Like Growth Factor-I–Stimulated Rat Prostate Cancer Cell Proliferation and Apoptotic Resistance In Vitro via Inhibition of Intracellular Signaling Pathways Involving Tyrosine Kinase. J. Nutr, 133(7):2367-2376


Xiao K, Xuan LJ, Xu YM, Bai D, Zhong DX. (2002). Dimeric stilbene glycosides from Polygonum cuspidatum. Eur J Org Chem, 3:564–568.


Zhao C, Nagatsu A, Hatano K, Shirai N, Kato S. (2003). New lignan glycosides from Chinese medicinal plant, Sinopodophyllum emodi. Chem Pharm Bull, 51:255–261.

Paeoniflorin

Cancer: Hepatocellular carcinoma, colorectal, liver

Action: Radio-protective, ameliorated myelosuppression, MDR

Radio-protective

The radio-protective effect of paeoniflorin (PF), a main bioactive component in the traditional Chinese herb peony, on irradiated thymocytes and the possible mechanisms of protection have been investigated. Ionizing radiation can induce DNA damage and cell death by generating reactive oxygen species (ROS).

It was found 60Co γ-ray irradiation increased cell death and DNA fragmentation in a dose-dependent manner while increasing intracellular ROS. Pre-treatment of thymocytes with PF (50–200 µg/ml) reversed this tendency and attenuated irradiation-induced ROS generation. Hydroxyl-scavenging action of PF in vitro was detected through electron spin resonance assay. Several anti-apoptotic characteristics of PF, including the ability to diminish cytosolic Ca2+ concentration, inhibit caspase-3 activation, and up-regulate Bcl-2 and down-regulate Bax in 4 Gy-irradiated thymocytes, were determined.

Extracellular regulated kinase (ERK), c-Jun NH2-terminal kinase (JNK), and p38 kinase, were activated by 4 Gy irradiation, with their activation partly blocked by pre-treatment of cells with PF. The presence of ERK inhibitor PD98059, JNK inhibitor SP600125 and p38 inhibitor SB203580 decreased cell death in 4 Gy-irradiated thymocytes. These results suggest PF protects thymocytes against irradiation-induced cell damage by scavenging ROS and attenuating the activation of the mitogen-activated protein kinases (Li et al., 2007).

Liver Cancer

Prostaglandin E2 (PGE2) has been shown to play an important role in tumor development and progression. PGE2 mediates its biological activity by binding any one of four prostanoid receptors (EP1 through EP4). Paeoniflorin, a monoterpene glycoside, significantly inhibited the proliferation of HepG2 and SMMC-7721 cells stimulated by butaprost at multiple time points (24, 48, and 72 hours). Paeoniflorin induced apoptosis in HepG2 and SMMC-7721 cells, which was quantified by annexin-V and propidium iodide staining. Our results indicate that the expression of the EP2 receptor and Bcl-2 was significantly increased, whereas that of Bax and cleaved caspase-3 was decreased in HepG2 and SMMC-7721 cells.

Paeoniflorin, which may be a promising agent in the treatment of liver cancer, induced apoptosis in hepatocellular carcinoma cells by down-regulating EP2 expression and also increased the Bax-to-Bcl-2 ratio, thus up-regulating the activation of caspase-3 (Hu et al., 2013).

Colorectal Cancer

Results showed that positive cells of Proliferating Cell Nuclear Antigen (PCNA) in paeoniflorin (PF) and docetaxel-treated group was decreased to 30% and 15% respectively, compared with control group of tumors. But apoptosis cells in docetaxel treated groups studied by TUNEL is increased to 40 ± 1.2% and 30 ± 1.5% respectively, compared with 24 ± 2.3% in negative control. Furthermore, the efficiency of tumor-bearing mice treated by PF was superior to docetaxel in vivo. Overall, PF may be an effective chemo-preventive agent against colorectal cancer HT29 (Wang et al., 2012).

Ameliorates Myelosuppression

The administration of paeoniflorin and albiflorin (CPA) extracted from Paeonia radix, significantly ameliorated myelosuppression in all cases. For the X-ray irradiated mice and the chemotherapy treated mice and rabbits, high dosages of CPA resulted in the recovery of, respectively, 94.4%, 95.3% and 97.7% of hemoglobin content; 67.7%, 92.0% and 94.3% of platelet numbers; 26.8%, 137.1% and 107.3% of white blood cell counts; as well as a reversal in the reduction of peripheral differential white blood cell counts.

There was also a recovery of 50.9%, 146.1% and 92.3%, respectively, in the animals' relative spleen weight. Additionally, a recovery of 35.7% and 87.2% respectively in the number of bone marrow nucleated cells was observed in the radio- and chemo -therapy-treated mice. Bone marrow white blood cell counts also resumed to normal levels (Xu et al., 2011).

MDR

Studies have shown that NF-κB activation may play an essential role in the development of chemotherapy resistance in carcinoma cells. Paeonißorin, a principal bioactive component of the root of Paeonia lactißora, has been reported to exhibit various pharmacological effects. In the present study, Fanh et al. (2012) reported for the first time that paeoniflorin at non-toxic concentrations may effectively modulate multi-drug resistance (MDR) of the human gastric cancer cell line SGC7901/vincristine (VCR) via the inhibition of NF-κB activation and, at least partly, by subsequently down-regulating its target genes MDR1, BCL-XL and BCL-2.

References

Fang S, Zhu W, Zhang Y, Shu Y, Liu P. (2012). Paeoniflorin modulates Multi-drug resistance of a human gastric cancer cell line via the inhibition of NF- κB activation. Mol Med Rep, 5(2):351-6. doi: 10.3892/mmr.2011.652.


Hu S, Sun W, Wei W, et al. (2013). Involvement of the prostaglandin E receptor EP2 in paeoniflorin-induced human hepatoma cell apoptosis. Anti-cancer Drugs, 24(2):140-9. doi: 10.1097/CAD.0b013e32835a4dac.


Li CR, Zhou Z, Zhu D, et al. (2007). Protective effect of paeoniflorin on irradiation-induced cell damage involved in modulation of reactive oxygen species and the mitogen-activated protein kinases. The International Journal of Biochemistry & Cell Biology, 39(2):426–438


Wang H, Zhou H, Wang CX, et al. (2012). Paeoniflorin inhibits growth of human colorectal carcinoma HT 29 cells in vitro and in vivo. Food Chem Toxicol, 50(5):1560-7. doi: 10.1016/j.fct.2012.01.035.


Xu W, Zhou L, Ma X, et al. (2011). Therapeutic effects of combination of paeoniflorin and albiflorin from Paeonia radix on radiation and chemotherapy-induced myelosuppression in mice and rabbits. Asian Pac J Cancer Prev, 12(8):2031-7.

Nomilin

Cancer: Melanoma, breast cancer

Action: Anti-angiogenic

Nomilin is a triterpenoid present in common edible citrus fruits (Citrus grandis [(L.) Osb.], Citrus unshiu [(Swingle) Marcow.] and Citrus reticulata (Blanco)) with putative anti-cancer properties.

Melanoma

Nomilin possess anti-metastatic action, inducing metastasis in C57BL/6 mice through the lateral tail vein using highly metastatic B16F-10 melanoma cells. Administration of nomilin inhibited tumor nodule formation in the lungs (68%) and markedly increased the survival rate of the metastatic tumor–bearing animals. Nomilin showed an inhibition of tumor cell invasion and activation of matrix metalloproteinases. Treatment with nomilin induced apoptotic response.

Nomilin treatment also exhibited a down-regulated Bcl-2 and cyclin-D1 expression and up-regulated p53, Bax, caspase-9, caspase-3, p21, and p27 gene expression in B16F-10 cells. Pro-inflammatory cytokine production and gene expression were found to be down-regulated in nomilin-treated cells. The study also reveals that nomilin could inhibit the activation and nuclear translocation of anti-apoptotic transcription factors such as nuclear factor (NF)-κB, CREB, and ATF-2 in B16F-10 cells (Pratheeshkumar et al., 2011).

Breast Cancer; ER+

A panel of 9 purified limonoids, including limonin, nomilin, obacunone, limonexic acid (LNA), isolimonexic acid (ILNA), nomilinic acid glucoside (NAG), deacetyl nomilinic acid glucoside (DNAG), limonin glucoside (LG) and obacunone glucoside (OG) as well as 4 modified compounds such as limonin methoxime (LM), limonin oxime (LO), defuran limonin (DL), and defuran nomilin (DN), were screened for their cytotoxicity on estrogen receptor (ER)-positive (MCF-7) or ER-negative (MDA-MB-231) human breast cancer cells. Findings indicated that the citrus limonoids may have potential for the prevention of estrogen-responsive breast cancer (MCF-7) via caspase-7 dependent pathways (Lin et al., 2013).

Blocks Angoigenesis

Nomilin significantly inhibited tumor-directed capillary formation. Serum pro-inflammatory cytokines such as IL-1β, IL-6, TNF-α and GM-CSF and also serum NO levels were significantly reduced by the treatment of nomilin. Administration of nomilin significantly reduced the serum level of VEGF, a pro-angiogenic factor and increased the anti-angiogenic factors IL-2 and TIMP-1. Nomilin significantly retarded endothelial cell proliferation, migration, invasion and tube formation. These data clearly demonstrate the anti-angiogenic potential of nomilin by down-regulating the activation of MMPs, production of VEGF, NO and pro-inflammatory cytokines as well as up-regulating IL-2 and TIMP (Pratheeshkumar et al., 2011).

References

Kim J, Jayaprakasha GK, Patil BS. (2013). Limonoids and their anti-proliferative and anti-aromatase properties in human breast cancer cells. Food Funct, 4(2):258-65. doi: 10.1039/c2fo30209h.


Pratheeshkumar P, Raphael TJ & Kuttan G. (2011). Nomilin Inhibits Metastasis via Induction of Apoptosis and Regulates the Activation of Transcription Factors and the Cytokine Profile in B16F-10 Cells. Integr Cancer Ther. doi: 10.1177/1534735411403307


Pratheeshkumar P, Kuttan G. (2011). Nomilin inhibits tumor-specific angiogenesis by down-regulating VEGF, NO and pro-inflammatory cytokine profile and also by inhibiting the activation of MMP-2 and MMP-9. Eur J Pharmacol, 668(3):450-8. doi: 10.1016/j.ejphar.2011.07.029.

Magnolol

Cancer:
Bladder, breast, colon, prostate, glioblastoma, ovarian, leukemia, lung

Action: Anti-inflammatory, apoptosis, inhibits angiogenesis, anti-metastatic

Magnolol (Mag), an active constituent isolated from the Chinese herb hou po (Magnolia officinalis (Rehder & Wilson)) has long been used to suppress inflammatory processes. It has anti-cancer activity in colon, hepatoma, and leukemia cell lines.

Anti-inflammatory

Magnolol (Mag) suppressed IL-6-induced promoter activity of cyclin D1 and monocyte chemotactic protein (MCP)-1 for which STAT3 activation plays a role. Pre-treatment of ECs with Mag dose-dependently inhibited IL-6-induced Tyr705 and Ser727 phosphorylation in STAT3 without affecting the phosphorylation of JAK1, JAK2, and ERK1/2. Mag pre-treatment of these ECs dose-dependently suppressed IL-6-induced promoter activity of intracellular cell adhesion molecule (ICAM)-1 that contains functional IL-6 response elements (IREs).

In conclusion, our results indicate that Mag inhibits IL-6-induced STAT3 activation and subsequently results in the suppression of downstream target gene expression in ECs. These results provide a therapeutic basis for the development of Mag as an anti-inflammatory agent for vascular disorders including atherosclerosis (Chen et al., 2006).

Bladder Cancer; Inhibits Angiogenesis

In the present study, Chen et al. (2013) demonstrated that magnolol significantly inhibited angiogenesis in vitro and in vivo, evidenced by the attenuation of hypoxia and vascular endothelial growth factor (VEGF)-induced tube formation of human umbilical vascular endothelial cells, vasculature generation in chicken chorioallantoic membrane, and Matrigel plug.

In hypoxic human bladder cancer cells (T24), treatment with magnolol inhibited hypoxia-stimulated H2O2 formation, HIF-1α induction including mRNA, protein expression, and transcriptional activity as well as VEGF secretion. Interestingly, magnolol also acts as a VEGFR2 antagonist, and subsequently attenuates the downstream AKT/mTOR/p70S6K/4E-BP-1 kinase activation both in hypoxic T24 cells and tumor tissues. As expected, administration of magnolol greatly attenuated tumor growth, angiogenesis and the protein expression of HIF-1α, VEGF, CD31, a marker of endothelial cells, and carbonic anhydrase IX, an endogenous marker for hypoxia, in the T24 xenograft mouse model.

Collectively, these findings strongly indicate that the anti-angiogenic activity of magnolol is, at least in part, mediated by suppressing HIF-1α/VEGF-dependent pathways, and suggest that magnolol may be a potential drug for human bladder cancer therapy.

Colon Cancer; Induces Apoptosis

Emerging evidence has suggested that activation of AMP-activated protein kinase (AMPK), a potential cancer therapeutic target, is involved in apoptosis in colon cancer cells. However, the effects of magnolol on human colon cancer through activation of AMPK remain unexplored.

Magnolol displayed several apoptotic features, including propidium iodide labeling, DNA fragmentation, and caspase-3 and poly(ADP-ribose) polymerase cleavages. Park et al. (2012) showed that magnolol induced the phosphorylation of AMPK in dose- and time-dependent manners.

Magnolol down-regulated expression of the anti-apoptotic protein Bcl2, up-regulated expression of pro-apoptotic protein p53 and Bax, and caused the release of mitochondrial cytochrome c. Magnolol-induced p53 and Bcl2 expression was abolished in the presence of compound C. Magnolol inhibited migration and invasion of HCT-116 cells through AMPK activation. These findings demonstrate that AMPK mediates the anti-cancer effects of magnolol through apoptosis in HCT-116 cells.

Ovarian Cancer

Treatment of HER-2 overexpressing ovarian cancer cells with magnolol down-regulated the HER-2 downstream PI3K/Akt signaling pathway, and suppressed the expression of downstream target genes, vascular endothelial growth factor (VEGF), matrix metalloproteinase 2 (MMP2) and cyclin D1. Consistently, magnolol-mediated inhibition of MMP2 activity could be prevented by co-treatment with epidermal growth factor. Migration assays revealed that magnolol treatment markedly reduced the motility of HER-2 overexpressing ovarian cancer cells. These findings suggest that magnolol may act against HER-2 and its downstream PI3K/Akt/mTOR-signaling network, thus resulting in suppression of HER-2mediated transformation and metastatic potential in HER-2 overexpressing ovarian cancers. These results provide a novel mechanism to explain the anti-cancer effect of magnolol (Chuang et al., 2011).

Lung Cancer

Magnolol has been found to inhibit cell growth, increase lactate dehydrogenase release, and modulate cell cycle in human lung carcinoma A549 cells. Magnolol induced the activation of caspase-3 and cleavage of Poly-(ADP)-ribose polymerase, and decreased the expression level of nuclear factor-κB/Rel A in the nucleus. In addition, magnolol inhibited basic fibroblast growth factor-induced proliferation and capillary tube formation of human umbilical vein endothelial cells. These data indicate that magnolol is a potential candidate for the treatment of human lung carcinoma (Seo et al., 2011).

Prostate Cancer; Anti-metastatic

Matrix metalloproteinases (MMPs) are enzymes involved in various steps of metastasis development. The objective of this study was to study the effects of magnolol on cancer invasion and metastasis using PC-3 human prostate carcinoma cells. Magnolol inhibited cell growth in a dose-dependent manner. In an invasion assay conducted in Transwell chambers, magnolol showed 33 and 98% inhibition of cancer cell at 10 microM and 20 microM concentrations, respectively, compared to the control. The protein and mRNA levels of both MMP-2 and MMP-9 were down-regulated by magnolol treatment in a dose-dependent manner.

These results demonstrate the anti-metastatic properties of magnolol in inhibiting the adhesion, invasion, and migration of PC-3 human prostate cancer cells (Hwang et al., 2010).

Glioblastoma Cancer

Magnolol has been found to concentration-dependently (0-40 microM) decrease the cell number in a cultured human glioblastoma cancer cell line (U373) and arrest the cells at the G0/G1 phase of the cell-cycle.

Pre-treatment of U373 with p21/Cip1 specific antisense oligodeoxynucleotide prevented the magnolol-induced increase of p21/Cip1 protein levels and the decrease of DNA synthesis. Magnolol at a concentration of 100 microM induced DNA fragmentation in U373. These findings suggest the potential applications of magnolol in the treatment of human brain cancers (Chen et al. 2011).

Inhibits Angiogenesis

Magnolol inhibited VEGF-induced Ras activation and subsequently suppressed extracellular signal-regulated kinase (ERK), phosphatidylinositol-3-kinase (PI3K)/Akt and p38, but not Src and focal adhesion kinase (FAK). Interestingly, the knockdown of Ras by short interfering RNA produced inhibitory effects that were similar to the effects of magnolol on VEGF-induced angiogenic signaling events, such as ERK and Akt/eNOS activation, and resulted in the inhibition of proliferation, migration, and vessel sprouting in HUVECs.

In combination, these results demonstrate that magnolol is an inhibitor of angiogenesis and suggest that this compound could be a potential candidate in the treatment of angiogenesis-related diseases (Kim et al., 2013).

References

Chen LC, Liu YC, Liang YC, Ho YS, Lee WS. (2009). Magnolol inhibits human glioblastoma cell proliferation through up-regulation of p21/Cip1. J Agric Food Chem, 57(16):7331-7. doi: 10.1021/jf901477g.


Chen MC, Lee CF, Huang WH, Chou TC. (2013). Magnolol suppresses hypoxia-induced angiogenesis via inhibition of HIF-1 α /VEGF signaling pathway in human bladder cancer cells. Biochem Pharmacol, 85(9):1278-87. doi: 10.1016/j.bcp.2013.02.009.


Chen SC, Chang YL, Wang DL, Cheng JJ. (2006). Herbal remedy magnolol suppresses IL-6-induced STAT3 activation and gene expression in endothelial cells. Br J Pharmacol, 148(2): 226–232. doi: 10.1038/sj.bjp.0706647


Chuang TC, Hsu SC, Cheng YT, et al. (2011). Magnolol down-regulates HER2 gene expression, leading to inhibition of HER2-mediated metastatic potential in ovarian cancer cells. Cancer Lett, 311(1):11-9. doi: 10.1016/j.canlet.2011.06.007.


Hwang ES, Park KK. (2010). Magnolol suppresses metastasis via inhibition of invasion, migration, and matrix metalloproteinase-2/-9 activities in PC-3 human prostate carcinoma cells. Biosci Biotechnol Biochem, 74(5):961-7.


Kim KM, Kim NS, Kim J, et al. (2013). Magnolol Suppresses Vascular Endothelial Growth Factor-Induced Angiogenesis by Inhibiting Ras-Dependent Mitogen-Activated Protein Kinase and Phosphatidylinositol 3-Kinase/Akt Signaling Pathways. Nutr Cancer.


Park JB, Lee MS, Cha EY, et al. (2012). Magnolol-induced apoptosis in HCT-116 colon cancer cells is associated with the AMP-activated protein kinase signaling pathway. Biol Pharm Bull, 35(9):1614-20.


Seo JU, Kim MH, Kim HM, Jeong HJ. (2011). Anti-cancer potential of magnolol for lung cancer treatment. Arch Pharm Res, 34(4):625-33. doi: 10.1007/s12272-011-0413-8.

Gypenosides

Cancer: Leukemia, colorectal., oral., esophageal

Action: Apoptosis,inhibits cell proliferation and migration

Gypenosides (Gyp), found in Gynostemma pentaphyllum Makino [(Thunb) Makino], have been used as folk medicine for centuries and have exhibited diverse pharmacological effects, including anti-leukemia effects in vitro and in vivo.

Gyp have been used to examine effects on cell viability, cell-cycle, and induction of apoptosis in vitro. They were administered in the diet to mice injected with WEHI-3 cells in vivo. Gyp inhibited the growth of WEHI-3 cells. These effects were associated with the induction of G0/G1 arrest, morphological changes, DNA fragmentation, and increased sub-G1 phase. Gyp promoted the production of reactive oxygen species, increased Ca2+ levels, and induced the depolarization of the mitochondrial membrane potential.

The effects of Gyp were dose- and time-dependent. Moreover, Gyp increased levels of the pro-apoptotic protein Bax, reduced levels of the anti-apoptotic proteins Bcl-2, and stimulated release of cytochrome c, AIF (apoptosis-inducing factor), and Endo G (endonuclease G) from mitochondria. The levels of GADD153, GRP78, ATF6-α, and ATF4-α were increased by Gyp, resulting in ER (endoplasmic reticular) stress in WEHI-3 cells. Oral consumption of Gyp increased the survival rate of mice injected with WEHI-3 cells used as a mouse model of leukemia.

Results of these experiments provide new information on understanding mechanisms of Gyp-induced effects on cell-cycle arrest and apoptosis in vitro and in an in vivo animal model (Hsu et al., 2011).

Inhibits Cell Proliferation and Migration

Results indicated that Gypenosides (Gyp) inhibited cell proliferation and migration in SW620 and Eca-109 cells in dose- and time-dependent manner. Gyp elevated intracellular ROS level, decreased the Δψ m, and induced apoptotic morphology such as cell shrinkage and chromatin condensation, suggesting oxidative stress and mitochondria-dependent cell apoptosis that might be involved in Gyp-induced cell viability loss in SW620 and Eca-109 cells. The findings indicate Gyp may have valuable application in clinical colon cancer and esophageal cancer treatments (Yan et al., 2013).

Gyp-induced cell death occurs through caspase-dependent and caspase-independent apoptotic signaling pathways, and the compound reduced tumor size in a xenograft nu/nu mouse model of oral cancer.

Gyp induced morphological changes, decreased the percentage of viable cells, caused G0/G1 phase arrest, and triggered apoptotic cell death in SAS cells. Cell-cycle arrest induced by Gyp was associated with apoptosis. The production of ROS, increased intracellular Ca(2+) levels, and the depolarization of ΔΨ(m) were observed. Gyp increased levels of the pro-apoptotic protein Bax but inhibited the levels of the anti-apoptotic proteins Bcl-2 and Bcl-xl. Gyp also stimulated the release of cytochrome c and Endo G. Translocation of GADD153 to the nucleus was stimulated by Gyp. Gyp in vivo attenuated the size and volume of solid tumors in a murine xenograft model of oral cancer (Lu et al., 2012).

Cell-cycle Arrest

Lin et al. (2011) have shown that gypenosides (Gyp) induced cell-cycle arrest and apoptosis in many human cancer cell lines. In the present study the effects of Gyp on cell morphological changes and viability, cell-cycle arrest and induction of apoptosis in vitro and effects on Gyp in an in vivo murine xenograft model were demonstrated. Results indicated that Gyp induced morphological changes, decreased cell viability, induced G0/G1 arrest, DNA fragmentation and apoptosis (sub-G1 phase) in HL-60 cells. Gyp increased reactive oxygen species production and Ca(2+) levels but reduced mitochondrial membrane potential in a dose- and time-dependent manner.

Oral consumption of Gyp reduced tumor size of HL-60 cell xenograft mode mice in vivo. These results provide new information on understanding mechanisms by which Gyp induces cell-cycle arrest and apoptosis in vitro and in vivo (Lin et al., 2011).

References

Hsu HY, Yang JS, Lu KW, et al. (2011). An Experimental Study on the Anti-leukemia Effects of Gypenosides In Vitro and In Vivo. Integr Cancer Ther, 10(1):101-12. doi: 10.1177/1534735410377198.


Lin JJ, Hsu HY, Yang JS, et al. (2011). Molecular evidence of anti-leukemia activity of gypenosides on human myeloid leukemia HL-60 cells in vitro and in vivo using a HL-60 cells murine xenograft model. Phytomedicine,18(12):1075-85. doi: 10.1016/j.phymed.2011.03.009.


Lu KW, Chen JC, Lai TY, et al. (2012). Gypenosides suppress growth of human oral cancer SAS cells in vitro and in a murine xenograft model: the role of apoptosis mediated by caspase-dependent and caspase-independent pathways. Integr Cancer Ther, 11(2):129-40. doi: 10.1177/1534735411403306.


Yan H, Wang X, Wang Y, Wang P, Xiao Y. (2013). Antiproliferation and anti-migration induced by gypenosides in human colon cancer SW620 and esophageal cancer Eca-109 cells. Hum Exp Toxicol.

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

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.

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.

Angelicin

Cancer: Leukemia, colon, ER+ Ovarian

Action: Apoptotic, anti-cancer

Angelicin is a furanocoumarin. It can be found in Bituminaria bituminosa and is structurally related to psoralens, a well-known chemical class of photosensitizers used for its anti-proliferative activity in treatment of different skin diseases.

Induces Apoptosis

The cellular cytotoxicity of angelicin was examined by cell viability assay, DNA fragmentation by DNA ladder assay, and activation of caspases and Bcl-2 family proteins by Western blot analyzes. The results suggest that angelicin increased cellular cytotoxicity in a dose- and time-dependent manner with IC(50) of 49.56 µM at 48 hours of incubation.

In addition, angelicin dose-dependently downregulated the expression of anti-apoptotic proteins including Bcl-2, Bcl-xL, and Mcl-1 suggesting the involvement of the intrinsic mitochondria-mediated apoptotic pathway which did not participate in Fas/FasL-induced caspase-8-mediated extrinsic, MAP kinases, and PI3K/AKT/GSK-3β pathway.

Taken together, these data indicate that angelicin is an effective apoptosis-inducing natural compound of human SH-SY5Y neuroblastoma cells which suggests that this compound may have a role in future therapies for human neuroblastoma cancer (Rahman et al., 2012).

Anti-cancer

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. Angelicin isolated from Aurantti Fructus Immaturus showed selective inhibition of the proliferation of TR-LE cells (Jeong et al., 2013). Angelicin, isolated from Bituminaria morisiana was subjected to cytotoxicity screening against a panel of human cancer cells (Leonti et al., 2010).

References

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.


Leonti M, Casu L, Gertsch J, et al. (2010). A pterocarpan from the seeds of Bituminaria morisiana. J Nat Med. 64(3):354-7. doi: 10.1007/s11418-010-0408-7.


Rahman MA, Kim NH, Yang H, Huh SO. (2012). Angelicin induces apoptosis through intrinsic caspase-dependent pathway in human SH-SY5Y neuroblastoma cells. Mol Cell Biochem, 369(1-2):95-104. doi: 10.1007/s11010-012-1372-1.

Qingkailing

Cancer: Leukemia, sarcoma

Action: Antibiotic, anti-apoptotic, anti-inflammatory, neuro-protective, pro-apoptotic, immunomodulating, MMPs regulation

Anti-inflammatory and Immunomodulating

Qingkailing and Shuanghuanglian (SHHL) are two commonly used Chinese herbal preparations with reported anti-inflammatory activity. The effects of these two preparations on the capacity of staphylococcal toxic shock syndrome toxin 1 (TSST-1), to stimulate the production of cytokines (IL-1β, IL-6, TNF-α, IFN-γ) and chemokines (MIP-1α, MIP-1β and MCP-1) by peripheral blood mononuclear cell (PBMC), was tested. Their effect on LPS-stimulated NF-κB transcriptional activity in a THP-1 cell line, and on human monocyte chemotactic response to chemoattractants, was also evaluated.

The results suggested that the pharmacological basis for the anti-inflammatory effects of Qingkailing and SHHL is the result of suppression of NF-κB regulated gene transcription, leading to suppressed production of pro-inflammatory cytokines and chemokines. Interference with leukocyte chemotaxis also contributes to the anti-inflammatory and immunomodulating effects of these medicinals. Identification of the responsible components in these two herbal preparations may yield compounds suitable for structural modification into potent novel drugs (Chen et al., 2002).

Leukemia

The MTT assay, cell morphology, DNA gel electrophoresis, and flow-cytometry were utilized to study the apoptotic effect of Qingkailing, and its active compounds, on the human acute promyelocytic leukemia (HL-60) cell line.

Qingkailing and its active compounds, Baicalin and hyodeoxycholic acid, exhibited strong cytotoxicity in inhibiting HL-60 cells, while Bezoar cholic acid showed a weaker effect. Apoptosis could be induced after being treated for 6 h by the former two compounds, displaying a typical apoptosis peak under flow-cytometry, but could not be induced by the latter.

Qingkailing could induce apoptosis in leukemia cells in vitro, which could serve as a mechanism of Qingkailing in the treatment of acute promyelocytic leukemia (Chen, Dong, & Zhang, 2001).

Qingkailing injection could prevent the decrease of MMP induced by injury of hypoxia-hypoglycemia-reoxygenation, stabilize MMP, inhibit cell apoptosis, and protect hippocampal neurons (Tsing, 2006).

Matrix Metalloproteinases (MMPs) Regulation

Matrix metalloproteinases (MMPs) play vital roles in many pathological conditions, including cancer, cardiovascular disease, arthritis and inflammation. Modulating MMP activity may therefore be a useful therapeutic approach in treating these diseases. Qingkailing is a popular Chinese anti-inflammatory formulation used to treat symptoms such as rheumatoid arthritis, acute hypertensive cerebral hemorrhage, hepatitis and upper respiratory tract infection.

One of the components of Qingkailing, Fructus gardeniae, strongly inhibits MMP activity. The IC50 values for the primary herbal extract and water extract against MMP-16 were 32 and 27 µg/ml, respectively. In addition, the herbal extracts influenced HT1080 human fibrosarcoma cell growth and morphology.

These data may provide molecular mechanisms for the therapeutic effects of Qingkailing and herbal medicinal Fructus gardenia (Yang et al., 2008).

Sources

Chen X, Howard OM, Yang X, Wang L, Oppenheim JJ, Krakauer T. (2002). Effects of Shuanghuanglian and Qingkailing, two multi-components of traditional Chinese medicinal preparations, on human leukocyte function. Life Sciences, 70(24), 2897-2913.


Chen ZT, Dong Q, Zhang L. (2001). Study on the effect of Qingkailing injection and its active principle in inducing cell apoptosis in human acute promyelocytic leukemia. Chinese Journal of Integrated Traditional and Western Medicine, 21(11), 840-842.


Tsing H. (2006). Influences of Qingkailing Injection on neuron apoptosis and mitochondrial membrane potential. Journal of Beijing University of Traditional Chinese Medicine, 2006(2), R285.5.


Yang JG, Shen YH, Hong Y, Jin FH, Zhao SH, Wang MC, Shi XJ,   Fang XX. (2008). Stir-baked Fructus gardeniae (L.) extracts inhibit matrix metalloproteinases and alter cell morphology. Journal of Ethnopharmacology, 117(2), 285-289.

Qingkailing

Cancer: Leukemia, sarcoma

Action: Antibiotic, anti-apoptotic, anti-inflammatory, neuro-protective, pro-apoptotic, immunomodulating, MMPs regulation

Anti-inflammatory and Immunomodulating

Qingkailing and Shuanghuanglian (SHHL) are two commonly used Chinese herbal preparations with reported anti-inflammatory activity. The effects of these two preparations on the capacity of staphylococcal toxic shock syndrome toxin 1 (TSST-1), to stimulate the production of cytokines (IL-1β, IL-6, TNF-α, IFN-γ) and chemokines (MIP-1α, MIP-1β and MCP-1) by peripheral blood mononuclear cell (PBMC), was tested. Their effect on LPS-stimulated NF-κB transcriptional activity in a THP-1 cell line, and on human monocyte chemotactic response to chemoattractants, was also evaluated.

The results suggested that the pharmacological basis for the anti-inflammatory effects of Qingkailing and SHHL is the result of suppression of NF-κB regulated gene transcription, leading to suppressed production of pro-inflammatory cytokines and chemokines. Interference with leukocyte chemotaxis also contributes to the anti-inflammatory and immunomodulating effects of these medicinals. Identification of the responsible components in these two herbal preparations may yield compounds suitable for structural modification into potent novel drugs (Chen et al., 2002).

Leukemia

The MTT assay, cell morphology, DNA gel electrophoresis, and flow-cytometry were utilized to study the apoptotic effect of Qingkailing, and its active compounds, on the human acute promyelocytic leukemia (HL-60) cell line.

Qingkailing and its active compounds, Baicalin and hyodeoxycholic acid, exhibited strong cytotoxicity in inhibiting HL-60 cells, while Bezoar cholic acid showed a weaker effect. Apoptosis could be induced after being treated for 6 h by the former two compounds, displaying a typical apoptosis peak under flow-cytometry, but could not be induced by the latter.

Qingkailing could induce apoptosis in leukemia cells in vitro, which could serve as a mechanism of Qingkailing in the treatment of acute promyelocytic leukemia (Chen, Dong, & Zhang, 2001).

Qingkailing injection could prevent the decrease of MMP induced by injury of hypoxia-hypoglycemia-reoxygenation, stabilize MMP, inhibit cell apoptosis, and protect hippocampal neurons (Tsing, 2006).

Matrix Metalloproteinases (MMPs) Regulation

Matrix metalloproteinases (MMPs) play vital roles in many pathological conditions, including cancer, cardiovascular disease, arthritis and inflammation. Modulating MMP activity may therefore be a useful therapeutic approach in treating these diseases. Qingkailing is a popular Chinese anti-inflammatory formulation used to treat symptoms such as rheumatoid arthritis, acute hypertensive cerebral hemorrhage, hepatitis and upper respiratory tract infection.

One of the components of Qingkailing, Fructus gardeniae, strongly inhibits MMP activity. The IC50 values for the primary herbal extract and water extract against MMP-16 were 32 and 27 µg/ml, respectively. In addition, the herbal extracts influenced HT1080 human fibrosarcoma cell growth and morphology.

These data may provide molecular mechanisms for the therapeutic effects of Qingkailing and herbal medicinal Fructus gardenia (Yang et al., 2008).

Sources

Chen X, Howard OM, Yang X, Wang L, Oppenheim JJ, Krakauer T. (2002). Effects of Shuanghuanglian and Qingkailing, two multi-components of traditional Chinese medicinal preparations, on human leukocyte function. Life Sciences, 70(24), 2897-2913.


Chen ZT, Dong Q, Zhang L. (2001). Study on the effect of Qingkailing injection and its active principle in inducing cell apoptosis in human acute promyelocytic leukemia. Chinese Journal of Integrated Traditional and Western Medicine, 21(11), 840-842.


Tsing H. (2006). Influences of Qingkailing Injection on neuron apoptosis and mitochondrial membrane potential. Journal of Beijing University of Traditional Chinese Medicine, 2006(2), R285.5.


Yang JG, Shen YH, Hong Y, Jin FH, Zhao SH, Wang MC, Shi XJ,   Fang XX. (2008). Stir-baked Fructus gardeniae (L.) extracts inhibit matrix metalloproteinases and alter cell morphology. Journal of Ethnopharmacology, 117(2), 285-289.

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.

Kanglaite injection (KLT)

Cancer: Lung, stomach, liver, kidney, breast, nasopharynx, esophagus, pancreas, colon-rectum, ovarian, prostate, lymphoma, leukemia

Action: Anti-tumoral, immunomodular, chemotherapy support, radiation support

Ingredients: yi yi ren (Coix Lacryma-jobi seed oil, CLSO).

Indications: primary NSCLC and primary liver cancer, which are not suitable for surgery, of qi and yin deficiency, lingering “Dampness due to Spleen deficiency types”. It has synergic effect when combined with radiotherapy or chemotherapy. It has certain anti-cachexia and analgesic effects for middle or late-stage tumor patients.

Dosage and usage:

Slow intravenous drip: 200 ml, once daily, 21 days as a course of treatment with 3-5 days interval.

When combined with radiotherapy or chemotherapy, the dosage can be reduced according to the practical conditions. (Drug Information Reference in Chinese, 2000. See end).

Invented by the famous pharmacological professor, Prof. Li Dapeng, Kanglaite Injection (KLT) has been listed by the Chinese government as a “State Basic Drug”, a “State Basic Medical Insurance Drug” and a “State Key New Drug”.

Based on pre-clinical studies at John Hopkins University, USA, tumor-inhibitive rate of KLT on transplanted breast carcinoma induced by cell strain MDA-MB-231 was over 50%. KLT could inhibit the expression of COX2 of the strain in vitro and act as an inhibitor of fatty acid synthase.

The broad ranged basic studies in China also revealed KLT different mechanisms such as inducing cancer cell apoptosis, inhibiting angiogenesis, reversing MDR and regulating gene expression of Fas/Apo-1 and Bcl-2.

Both Chinese and overseas clinical experiences have shown that KLT has proven effect in the treatment of cancers mainly at the sites of lung, breast, liver, nasopharynx, esophagus, stomach, pancreas, kidney, colon-rectum, ovary and prostate. This agent is also applied in the treatment of malignant lymphoma and acute leukemia. KLT has brought great benefits to over 500,000 cancer patients in more than 2,000 big or medium hospitals in China since 1997.

The year 1995 witnessed KLT patent certificates granted from China and the USA. In August 1997 the phase III clinical study was successfully completed and the injection was officially launched in China after final approval from the Ministry of Public Health.

Doctors in America carried out a phase 1 study of Kanglaite in 2003. They gave it to 16 people who had different types of cancer including lung, prostate and oesophageal cancers. The results showed people did not have many side-effects but the effect on their cancer varied. Some people showed no response, and their cancers continued to grow. But in others, the cancer stopped growing for a few months.

Standard treatment course for KLT is 200 ml (2 bottles) per day via intravenous drip x 42 days (84 bottles). There is a break for 4-5 days after 21 days. Clinical experiences in China and Russia suggest 2 treatment courses for those with late stage advanced and metastatic tumors for better therapeutic effect and evident prolongation of life (Conti, n.d.).

A consecutive cohort of 60 patients was divided into two groups, the experimental group receiving Kanglaite” Injection combined with chemotherapy and the control group receiving chemotherapy alone. After more than two courses of treatment, efficacy, quality of life and side-effects were evaluated. The response rate and KPS score of the experimental group were significantly improved as compared with those of the control group(P<0.05). In addition, gastrointestinal reactions and bone marrow suppression were significantly lower than in the control group(P<0.05). Kanglaite” Injection enhanced efficacy and reduced the side-effects of chemotherapy, improving quality of life of gastric cancer patients (Zhan et al., 2012).

Lung Cancer

C57BL/6 mice with Lewis lung carcinoma were divided into four groups: the control group (C), cisplatin group (1 mg/kg, DDP), low KLT group (6.25 ml/kg body weight [L]), and high KLT group (12.5 ml/kg body weight [H]). T cell proliferation was determined by the MTT assay. Nuclear factor-kappa B (NF-κB), inhibitor kappa B alpha

(IκBα), IκB kinase (IKK) and epidermal growth factor receptor (EGFR) levels were measured by western blotting. An enzyme-linked immunosorbent assay was used to analyze the expression of interleukin-2 (IL-2).

Intraperitoneal KLT significantly inhibited the growth of Lewis lung carcinoma, and the spleen index was significantly higher in the L and H groups than in the C group. KLT stimulated T cell proliferation in a dose-dependent manner. Treatment with KLT at either 6.25 or 12.5 ml/kg decreased the level of NF-κB in the nucleus in a dose-dependent manner, and KLT markedly decreased the expression of IκBα, IKK and EGFR in the cytoplasm of tumor cells and overall. IL-2 was significantly increased in the supernatant of splenocytes in the H group.

These results demonstrate that KLT has pronounced anti-tumor and immunostimulatory activities in C57BL/6 mice with Lewis lung carcinoma. These may affect the regulation of NF-κB/IκB expression, in addition to cytokines such as IL-2 and EGFR. Further work needs to investigate the relevant signaling pathway effects, but our findings suggest that KLT may be a promising anti-tumor drug for clinical use (Pan et al., 2012).

Skin Keratinocytes

Ultraviolet (UV) radiation plays an important role in the pathogenesis of skin photoaging. Depending on the wavelength of UV, the epidermis is affected primarily by UVB. One major characteristic of photoaging is the dehydration of the skin. Membrane-inserted water channels (aquaporins) are involved in this process. In this study we demonstrated that UVB radiation induced aquaporin-3 (AQP3) down-regulation in cultured human skin keratinocytes. Kanglaite is a mixture consisting of extractions of Coix Seed, which is an effective anti-neoplastic agent and can inhibit the activities of protein kinase C and NF-κB. We demonstrated that Kanglaite inhibited UVB-induced AQP3 down-regulation of cultured human skin keratinocytes. Our findings provide a potential new agent for anti-photoaging (Shan et al., 2012).

Hepatocellular Carcinoma

KLT produced an obvious time and dose-dependent inhibitory effect on HepG2 cells, and marked apoptosis was detected by FCM. The protein of Fas increased by 11.01%, 18.71%, 28.71% and 37.15%; the protein of FasL increased by 1.49%, 1.91%, 3.27% and 3.38% in comparison with the control (P<0.05). Real-time fluorescent quantitative RT-PCR showed that treating HepG2 cells with KLT caused the up-regulation of Fas and FasL mRNA. KLT inhibits HepG2 growth by inducing apoptosis, which may be mediated through activation of the Fas/FasL pathway (Lu et al., 2009).

Glomerular Nephritis

MTT, telomere repeat amplification protocol (TRAP), ELISA, PAGE and silver-stain were applied to detect the growth rate and telomerase activity of mesengial cell (MC) after stimulation of Kang Lai Te (KLT) and IL-1. The growth rate of MC was enhanced by IL-1 stimulation, which was accompanied with a reduction of the activity of telomerase. Adversely, the growth rate of MC was reduced by KLT, which was accompanied with an enhancement of activity of telomerase. Moreover, the growth rate of MC and the activity of telomerase were both inhibited by the combinative use of IL-1 and KLT without any influence from the sequence of their administration. KLT could inhibit proliferation and telomerase activity of MC with or without pre-stimulation with IL-1. KLT might be useful to prevent and treat glomerular nephritis related to MC proliferation (Hu et al., 2005).

Lung Metastasis

To screen the differential expression genes of Kanglaite in anti-tumor metastasis mRNA was extracted and purified from the lung of the mouse with LA795 lung metastasis, and hybridized respectively on 4 096-gene chip. cDNA microarray was scanned for the fluorescent signals and analyzing difference expression. Twenty-seven differential expressed genes were obtained.

Among these genes, 25 were up-regulated and 2 were down-regulated. Twelve of them were Mus musculus cDNA clone. Six genes related with genesis, development and metastasis of tumor. cDNA microarray for analysis of gene expression patterns is a powerful method to identify differential expressed genes. In this study, 6 genes are thought to be associated genes of Kanglaite in anti-tumor metastasis (Wu et al., 2003).

Lung Cancer; Chemo Side Effects

Sixteen reports were included in the meta-analysis. The quality of 16 studies was low. Pooling data of 5 studies indicated that the effect of Kanglaite+NP (Vinorelbine+Cisplatin) was better than NP with RR 1.46, 95% Confidence Interval 1.13 to 1.91. Pooling data of 3 studies of MVP (Mitomycin+Vindsine+ Cisplatin) plus Kanglaite indicated that the effect was better with RR 1.84, 95%CI 1.22 to 2.76. Pooling data of 2 studies showed that the effect of GP (Gemcitabine+Cisplatin) plus Kanglaite was better than GP with RR 1.63, 95%CI 1.09 to 2.43.

Fourteen studies revealed that Kanglaite may reduce the side-effects induced by regular treatment. Ten studies showed regular treatment plus Kanglaite can stabilize/improve quality of life (Zhu et al., 2009).

Apoptosis

Some studies show Kanglaite could inhibit some anti-apoptotic genes and activate some pro-apoptotic genes. Its injection solution is one of the new anti-cancer medicines that can significantly inhibit various kinds of tumor cells, so it has become the core of research into how to further explore KLT injection to promote tumor cell apoptosis by impacting on related genes (Lu et al., 2008).

References

Conti, M. (n.d.). Anti-cancer Chinese herbal kanglaite. Cancer Evolution. Retrieved from: http://www.cancerevolution.info/cancer-therapies/alternative-therapies/83-anticancer-chinese-herbal-kanglaite.html.


Hu, Y,H., Liang, W.K. Gong, Z.F. Xu,Q.L. Zou. (2005). The effect of kanglaite injection (KLT) on the proliferation and telomerase activity of rat mesangial cells. Zhongguo Zhong Yao Za Zhi, 30(6):450-453.


Lu, Y., Li, C.S., Dong, Q. (2008) Chinese herb related molecules of cancer-cell-apoptosis: a mini-review of progress between Kanglaite injection and related genes. J Exp Clin Cancer Res, 27:31. doi: 10.1186/1756-9966-27-31.


Lu, Y., L.Q. Wu, Q. Dong,C.S. Li. (2009). Experimental study on the effect of Kang-Lai-Te induced apoptosis of human hepatoma carcinoma cell HepG2. Hepatobiliary Pancreat Dis Int, 8(3):267-272.


Pan, P.,Y. Wu,Z.Y. Guo,R. et al. (2012). Anti-tumor activity and immunomodulatory effects of the intraperitoneal administration of Kanglaite in vivo in Lewis lung carcinoma. J Ethnopharmacol, 143(2):680-685.


Shan, S.J., Xiao T., Chen J., et al. (2012). Kanglaite attenuates UVB-induced down-regulation of aquaporin-3 in cultured human skin keratinocytes. Int J Mol Med, 29(4):625-629.


Wu, Y., Yang Y., Wu D. (2003). Study on the gene expression patterns of Kanglaite in anti-lung metastasis of LA795 mouse. Zhongguo Fei Ai Za Zhi, 6(6):473-476.


Zhan, Y.P., Huang X.E., Cao J. (2012). Clinical safety and efficacy of Kanglaite(R) (Coix Seed Oil) injection combined with chemotherapy in treating patients with gastric cancer. Asian Pac J Cancer Prev, 13(10):5319-5321.


Zhu, L.Z. Yang, S. Wang, Y. Tang. (2009). Kanglaite for Treating Advanced Non-small-cell Lung Cancer: A Systematic Review. Zhongguo Fei Ai Za Zhi, 12(3):208-215.

Artesunate

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

Action: Anti-metastatic, MDR, radio-sensitizer

Pulmonary Adenocarcinomas

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

Breast Cancer

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

Skin Cancer

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

Colon Cancer

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

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

Multi-drug resistance; Colon Cancer

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

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

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

Multi-drug resistance; Esophageal Cancer

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

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

Pancreatic Cancer

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

Ovarian Cancer

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

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

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

Multiple Myeloma, B-cell Lymphoma

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

Osteosarcoma, Leukemia/Lymphoma

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

References

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


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


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


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


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


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


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


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


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


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


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


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


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


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

Baicalin & Baicalein

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

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

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

Anti-cancer

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

Multiple Myeloma

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

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

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

Hepatoma

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

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

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

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

Ovarian Cancer

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

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

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

Breast Cancer

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

Breast Cancer, Cell Invasion

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

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

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

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

Oral Cancer

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

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

Anti-inflammatory

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

Angiogenesis Modulation

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

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

Colon Cancer

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

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

Cardio-protective

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

Prostate Cancer

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

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

Cancer: Acute lymphocytic leukemia, lymphoma and myeloma

Action: Cell-cycle arrest, induces apoptosis

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

References

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


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


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


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


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


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


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


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


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


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


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


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


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


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


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


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

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

Emodin (See also Aloe-Emodin)

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

Action: MDR-1, cell-cycle arrest

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

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

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

Pancreatic Cancer

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

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

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

Hepatocellular Carcinoma

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

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

Colon Cancer

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

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

Myeloid Leukemia

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

Breast Cancer; Block HER-2

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

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

MDR

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

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

Cell-cycle Arrest

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

Cell-cycle Arrest; MDR1 & AZT

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

References

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


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


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


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


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


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


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


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


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


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


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