Category Archives: Anti-metastatic

Carnosol

Cancer: Breast, prostate, skin, colon, leukemia, stomach

Action: Anti-inflammatrory, anti-angiogenic

Carnosol is found in certain Mediterranean meats, fruits, vegetables, and olive oil. In particular, it is sourced from rosemary (Rosmarinus officinalis (L.)) and desert sage (Salvia pachyphylla (Epling ex Munz)).

Prostate Cancer, Breast Cancer, Skin Cancer, Colon Cancer, Leukemia

One agent, carnosol, has been evaluated for anti-cancer property in prostate, breast, skin, leukemia, and colon cancer with promising results. These studies have provided evidence that carnosol targets multiple deregulated pathways associated with inflammation and cancer that include nuclear factor kappa B (NFκB), apoptotic related proteins, phosphatidylinositol-3-kinase (PI3 K)/Akt, androgen and estrogen receptors, as well as molecular targets. In addition, carnosol appears to be well tolerated in that it has a selective toxicity towards cancer cells versus non-tumorigenic cells and is well tolerated when administered to animals.

This mini-review reports on the pre-clinical studies that have been performed to date with carnosol describing mechanistic, efficacy, and safety/tolerability studies as a cancer chemoprevention and anti-cancer agent (Johnson, 2011).

Literature evidence from animal and cell culture studies demonstrates the anti-cancer potential of rosemary extract, carnosol, carnosic acid, ursolic acid, and rosmarinic acid to suppress the development of tumors in several organs including the colon, breast, liver, stomach, as well as melanoma and leukemia cells (Ngo et al., 2011).

Anti-inflammatory

Treatment with retinoic acid (RA) or carnosol, two structurally unrelated compounds with anti-cancer properties, inhibited phorbol ester (PMA)-mediated induction of activator protein-1 (AP-1) activity and cyclooxygenase-2 (COX-2) expression in human mammary epithelial cells. Treatment with carnosol but not RA blocked increased binding of AP-1 to the COX-2 promoter. Carnosol but not RA inhibited the activation of PKC, ERK1/2, p38, and c-Jun NH2-terminal kinase mitogen-activated protein kinase. Overexpressing c-Jun but not CBP/p300 reversed the suppressive effect of carnosol on PMA-mediated stimulation of COX-2 promoter activity.

Carnosol inhibited the induction of COX-2 by blocking PKC signaling and thereby the binding of AP-1 to the CRE of the COX-2 promoter. Taken together, these results show that small molecules can block the activation of COX-2 transcription by distinct mechanisms (Subbaramaiah, 2002).

Breast Cancer

Two rosemary components, carnosol and ursolic acid, appear to be partly responsible for the anti-tumorigenic activity of rosemary. Supplementation of diets for 2 weeks with rosemary extract (0.5% by wt) but not carnosol (1.0%) or ursolic acid (0.5%) resulted in a significant decrease in the in vivo formation of rat mammary DMBA-DNA adducts, compared to controls. When injected intraperitoneally (i.p.) for 5 days at 200 mg/kg body wt, rosemary and carnosol, but not ursolic acid, significantly inhibited mammary adduct formation by 44% and 40%, respectively, compared to controls. Injection of this dose of rosemary and carnosol was associated with a significant 74% and 65% decrease, respectively, in the number of DMBA-induced mammary adenocarcinomas per rat, compared to controls. Ursolic acid injection had no effect on mammary tumorigenesis.

Therefore, carnosol is one rosemary constituent that can prevent DMBA-induced DNA damage and tumor formation in the rat mammary gland, and, thus, has potential for use as a breast cancer chemopreventative agent (Singletary et al., 1996).

Anti-angiogenic

The anti-angiogenic activity of carnosol and carnosic acid could contribute to the chemo-preventive, anti-tumoral and anti-metastatic activities of rosemary extracts and suggests that there is potential in the treatment of other angiogenesis-related malignancies (L-pez-JimŽnez et al., 2013).

References:

Johnson JJ. (2011). Carnosol: A promising anti-cancer and anti-inflammatory agent. Cancer Letters, 305(1):1-7. doi:10.1016/j.canlet.2011.02.005.


L-pez-JimŽnez A, Garc'a-Caballero M, Medina Mç, Quesada AR. (2013). Anti-angiogenic properties of carnosol and carnosic acid, two major dietary compounds from rosemary. Eur J Nutr, 52(1):85-95. doi: 10.1007/s00394-011-0289-x.


Ngo SN, Williams DB, Head RJ. (2011). Rosemary and cancer prevention: preclinical perspectives. Crit Rev Food Sci Nutr, 51(10):946-54. doi: 10.1080/10408398.2010.490883.


Singletary K, MacDonald C & Wallig M. (1996). Inhibition by rosemary and carnosol of 7,12-dimethylbenz[a]anthracene (DMBA)-induced rat mammary tumorigenesis and in vivo DMBA-DNA adduct formation. Cancer Letters, 104(1):43-8. doi: 10.1016/0304-3835(96)04227-9


Subbaramaiah K, Cole PA, Dannenberg AJ. (2002). Retinoids and Carnosol Suppress Cyclooxygenase-2 Transcription by CREB-binding Protein/p300-dependent and -independent Mechanisms. Cancer Res, 62:2522

Shikonin

Cancer: Sarcoma-180, lung, melanoma, leukemia

Action: Anti-inflammatory, inhibits angiogenesis, MDR

Shiunko is a Kampo herbal ointment often used for the treatment of burns in Japan. It is mainly isolated from the root of Lithospermum erythrorhizon (Siebold & Zuccarini), which had been used for treating tumors and inflammation in China since the 5th century. The naphthoquinone pigment shikonin is the most important pharmacologically active substance in the dried root of Lithospermum erythrorhizon. In traditional Chinese medicine root extracts of Lithospermum erythrorhizon have been used to treat macular eruption, measles, sore throat, carbuncles, and burns (Chen et al., 2002). The anti-tumor effect of shikonin was first evidenced by its activity against murine sarcoma-180 (Sankawa et al., 1977).

Melanoma

It has been reported that shikonin, the main chemical ingredient of L. erythrorhizon is a novel inhibitor of angiogenesis. Angiogenesis is critical for tumor growth and inflammation. It inhibited tumor necrosis factor-alpha-induced and B16 melanoma-induced angiogenesis in mice and normal developmental angiogenesis in the yolk-sac membranes of chick embryos. Shikonin also inhibited proliferation and migration of endothelial cells in culture and network formation by endothelial cells on Matrigel in vitro. The dose-responsive study suggests that the mechanism of this inhibitory effect on angiogenesis involves the prevention of network formation by endothelial cells via blocking integrin alpha v beta 3 expression (Hisa et al., 1998).

Anti-inflammatory

Shikonin also reported to exert anti-inflammatory and anti-cancer effects both in vitro and in vivo. It has been found that proteasome was a molecular target of shikonin in tumor cells, but whether shikonin targets macrophage proteasome needs to be investigated. Consistently, shikonin accumulated IκB-α, an inhibitor of NF-κB, and ubiquitinated proteins in rat primary macrophage cultures, demonstrating that the proteasome is a target of shikonin under inflammatory conditions.

Shikonin also induced macrophage cell apoptosis and cell death. These results demonstrate for the first time that proteasome inhibition by shikonin contributes to its anti-inflammatory effect. The novel finding about macrophage proteasome as a target of shikonin suggests that this medicinal compound has great potential to be developed into an anti-inflammatory agent (Lu et al., 2011).

Leukemia, MDR

Shikonin has a strong cytotoxic effect on a wide variety of cancer cell lines, especially different types of leukemia and several known MDR cell lines. Microarray-based gene expression analysis of U937 leukemia cells suggested that the cytotoxicity of shikonin is based on the disruption of normal mitochondrial function, overproduction of ROS, inhibition of cytoskeleton formation, and finally induction of cell-cycle arrest and apoptosis. These effects were validated using in vitro cell culture experiments exploiting the specific natural fluorescence of shikonin and thereby identifying the possible primary cellular mechanism of shikonin's cytotoxicity (Wiench et al., 2012).

Lung Cancer

To better understand the anti-metastatic role of shikonin in lung cancer, the effect of shikonin on lung cancer cell proliferation was investigated, as well as its adhesion to extracellular matrices (ECM), migration and invasion in non-small-cell lung cancer A549 cells. Taken together, findings provide new evidence that shikonin suppresses lung cancer invasion and metastasis by inhibiting integrin β1 expression and the ERK1/2 signaling pathway. Integrin β1 facilitates cancer cell adhesion, migration and metastasis by activating intracellular signaling pathways including the ERK and PI3K signaling pathways, and it is in this way that shikonin exerts its anti-cancer activity (Wang et al., 2013).

MDR

Numerous previous studies have proven that shikonin and its analogs not only are highly tumoricidal but also can bypass drug-transporter and apoptotic defect mediated drug resistance. Cancer drug resistance is a major obstacle for the success of chemotherapy. Since most clinical anti-cancer drugs could induce drug resistance, it is desired to develop candidate drugs that are highly efficacious but incompetent to induce drug resistance. Shikonin was investigated for its ability as an inducer of cancer drug resistance. Different cell lines (K562, MCF-7, and a MDR cell line K562/Adr), after repeatedly treated with shikonin for 18 months, were assayed for drug resistance and gene expression profiling. After an 18-month treatment, cells only developed a mere 2-fold resistance to shikonin and a marginal resistance to cisplatin and paclitaxel, without cross-resistance to shikonin analogs and other anti-cancer agents. These merits make shikonin and its analogs potential candidates for cancer therapy with the advantages of avoiding induction of drug resistance and bypassing existing drug resistance (Wu et al., 2013).

References

Chen X, Yang L, Oppenheim JJ, Howard OMZ. (2002). Cellular pharmacology studies of shikonin derivatives. Phytotherapy Research, 16(3):199–209.


Hisa T, Kimura Y, Takada K, Suzuki F, Takigawa M. (1998). Shikonin, an ingredient of Lithospermum erythrorhizon, inhibits angiogenesis in vivo and in vitro. Anti-cancer Res, 18(2A):783-90.


Lu L, Qin A, Huang H, et al. (2011). Shikonin extracted from medicinal Chinese herbs exerts anti-inflammatory effect via proteasome inhibition. Eur J Pharmacol. 658(2–3):242–247.


Sankawa U, Ebizuka Y, Miyazaki T, et al. (1977). Anti-tumor activity of shikonin and its derivatives. Chemical and Pharmaceutical Bulletin, 25(9):2392–2395.


Wang H, Wu C, Wan S, et al. (2013). Shikonin attenuates lung cancer cell adhesion to extracellular matrix and metastasis by inhibiting integrin β 1 expression and the ERK1/2 signaling pathway. Toxicology, 308:104-12. doi: 10.1016/j.tox.2013.03.015. Epub 2013 Apr 4.


Wiench B, Eichhorn T, Malte Paulsen M, Efferth T. (2012). Shikonin Directly Targets Mitochondria and Causes Mitochondrial Dysfunction in Cancer Cells. Evidence-Based Complementary and Alternative Medicine, 2012:726025. doi:10.1155/2012/726025


Wu H, Xie J, Pan Q, et al. (2013). Anti-cancer agent shikonin is an incompetent inducer of cancer drug resistance. PLoS One, 8(1):e52706. doi: 10.1371/journal.pone.0052706.

Costunolide and Dehydrocostus Lactone

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

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

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

Breast cancer; Anti-metastatic

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

Breast Cancer

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

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

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

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

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

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

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

ER- Breast Cancer

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

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

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

Bladder Cancer

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

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

Sarcomas; MDR

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

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

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

Leukemia, Lung Cancer

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

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

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

Prostate Cancer

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

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

Gastric Cancer, Prostate Cancer

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

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

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

Anti-metastatic

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

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

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

Ovarian Cancer, MDR

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

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

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

Anti-inflammatory, Anti-oxidant, Mediates Apoptosis

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

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

References

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

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


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


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


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


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


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


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


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


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


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


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


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

Evodiamine

Cancer: Pancreatic, gastric, breast; ER+, ER-, lung

Action: Inhibits NF- κB, inhibits metastasis, increases intracellular ROS, apoptosis, cell-cycle arrest, anti-cancer, MDR

Evodiamine, a naturally occurring indole alkaloid, is one of the main bioactive ingredients of Evodia rutaecarpa [(Juss.) Benth.] (alkaloidal component of the extract). With respect to the pharmacological actions of evodiamine, more attention has been paid to beneficial effects in insults involving cancer, obesity, nociception, inflammation, cardiovascular diseases, Alzheimer's disease, infectious diseases and thermo-regulative effects. Evodiamine has evolved a superior ability to bind various proteins (Yu et al., 2013). Evodiamine exhibits anti-proliferative, anti-metastatic, and apoptotic activities.

Anti-cancer, MDR

Evodiamine possesses anti-anxiety, anti-obesity, anti-nociceptive, anti-inflammatory, anti-allergic, and anti-cancer effects. As well, it has thermoregulation, protection of myocardial ischemia-reperfusion injury and vessel-relaxing activities (Kobayashi, 2003; Shin et al., 2007; Ko et al., 2007; Ji, 2011). Evodiamine exhibits anti-cancer activities both in vitro and in vivo by inducing cell-cycle arrest or apoptosis, and inhibiting angiogenesis, invasion, and metastasis in a variety of cancer cell lines (Ogasawara et al., 2001; Ogasawara et al., 2002; Fei et al., 2003; Shyu et al., 2006). It presents anti-cancer potentials at micromolar concentrations and even at the nanomolar level in some cell lines in vitro (Lee et al., 2006; Wang, Li, & Wang, 2010). Evodiamine also stimulates autophagy, which serves as a survival function (Yang et al., 2008). Compared with other compounds, evodiamine is less toxic to normal human cells, such as human peripheral blood mononuclear cells (Fei et al., 2003; Zhang et al., 2004). It also inhibits the proliferation of adriamycin-resistant human breast cancer NCI/ADR-RES cells both in vitro and in Balb-c/nude mice (Liao et al., 2005).

Lung Cancer, Cell-cycle Arrest

Evodiamine (10  mg/kg) administrated orally twice daily significantly inhibits   tumor growth (Liao et al., 2005). Moreover, treatment with 10 mg/kg evodiamine from the 6th day after tumor inoculation into mice reduces lung metastasis and does not affect the body weight of mice during the experimental period (Ogasawara et al., 2001).

Cell-cycle Arrest

Evodiamine inhibits TopI enzyme, forms the DNA covalent complex with a similar concentration to that of irinotecan, and induces DNA damage (Chan et al., 2009; Tsai et al., 2010; Dong et al., 2010). However, TopI may not be the main target of this compound. Cancer cells treated with evodiamine exhibit G 2 / M phase arrest (Kan et al., 2004; Huang et al., 2004; Liao et al., 2005) rather than S phase arrest, which is not consistent with the mechanism of classic TopI inhibitors, such as irinotecan. Therefore, other targets aside from TopI may also be important for realizing the anti-cancer potentials of evodiamine. This statement is supported by the fact that evodiamine has effects on tubulin polymerization (Huang et al., 2004).

Increases Intracellular ROS, Apoptosis

Exposure to evodiamine rapidly increases intracellular ROS followed by an onset of mitochondrial depolarization (Yang et al., 2007). The generation of ROS and nitric oxide acts in synergy and triggers mitochondria-dependent apoptosis (Yang et al., 2008). Evodiamine also induces caspase-dependent and caspase-independent apoptosis, down-regulates Bcl-2 expression, and up-regulates Bax expression in some cancer cells (Zhang et al., 2003; Lee et al., 2006). The phosphatidylinositol 3-kinase/Akt/caspase and Fas ligand (Fas-L)/NF-κB signaling pathways might account for evodiamine-induced cell death. Moreover, these signals could be increased by the ubiquitin-proteasome pathway (Wang, Li, & Wang, 2010).

Inhibits Metastasis

Evodiamine has a marked inhibitory activity on tumor cell migration in vitro. When evodiamine at 10 mg/kg was administered into mice from the 6th day after tumor inoculation, the number of tumor nodules in lungs was decreased by 48% as compared to control. The inhibition rate was equivalent to that produced by cisplatin. Results suggest that evodiamine may be regarded as a promising agent in tumor metastasis therapy (Ogasawara et al., 2005).

Inhibits NF-κB

Evodiamine inhibited tumor necrosis factor (TNF)-induced Akt activation and its association with IKK. This down-regulation potentiated the apoptosis induced by cytokines and chemotherapeutic agents and suppressed TNF-induced invasive activity. Overall, these results indicate that evodiamine inhibits both constitutive and induced NF-κB activation and NF-κB-regulated gene expression (Takada et al., 2005).

Breast Cancer

Endocrine sensitivity, assessed by the expression of estrogen receptor (ER), has long been the predict factor to guide therapeutic decisions. Tamoxifen has been the most successful hormonal treatment in endocrine-sensitive breast cancer. However, in estrogen-insensitive cancer tamoxifen showed less effectiveness than in estrogen-sensitive cancer. It is interesting to develop new drugs against both hormone-sensitive and insensitive tumor. In this present study Wang et al. (2013) examined anti-cancer effects of evodiamine extracted from the Chinese herb, Evodiae fructus, in estrogen-dependent and -independent human breast cancer cells, MCF-7 and MDA-MB-231 cells, respectively.

Breast Cancer; ER+, ER-

The expression of ER α and β in protein and mRNA levels was down-regulated by evodiamine according to data from immunoblotting and RT-PCR analysis. Overall, results indicate that evodiamine mediates degradation of ER and induces caspase-dependent pathway leading to inhibition of proliferation of breast cancer cell lines. It suggests that evodiamine may in part mediate through ER-inhibitory pathway to inhibit breast cancer cell proliferation.

Evodiamine (10 mg/kg) significantly reduced tumor growth and pulmonary metastasis. In vitro, evodiamine inhibited cell migration and invasion abilities through down-regulation of MMP-9, urokinase-type plasminogen activator (uPA) and uPAR expression. Evodiamine-induced G0/G1 arrest and apoptosis were associated with a decrease in Bcl-2, cyclin D1 and cyclin-dependent kinase 6 (CDK6) expression and an increase in Bax and p27Kip1 expression (Du et al., 201).

Gastric Cancer

A study by Rasul et al. (2012) was conducted to investigate the synchronized role of autophagy and apoptosis in evodiamine-induced cytotoxic activity on SGC-7901 human gastric adenocarcinoma cells and further to elucidate the underlying molecular mechanisms. Evodiamine significantly inhibited the proliferation of SGC-7901 cells and induced G2/M phase cell-cycle arrest.

Evodiamine-induced autophagy is partially involved in the death of SGC-7901 cells which was confirmed by using the autophagy inhibitor 3-methyladenine (3-MA). Evodiamine has therapeutic potential against cancers.

Pancreatic Cancer

In vitro application of the combination therapy triggered significantly higher frequency of pancreatic cancer cells apoptosis, inhibited the activities of PI3K, Akt, PKA, mTOR and PTEN, and decreased the activation of NF-κB and expression of NF- κB-regulated products. Evodiamine can augment the therapeutic effect of gemcitabine in pancreatic cancer through direct or indirect negative regulation of the PI3K/Akt pathway (Wei et al., 2012).

References

Chan ALF, Chang WS, Chen LM et al. (2009). Evodiamine stabilizes topoisomerase I-DNA cleavable complex to inhibit topoisomerase I activity. Molecules, (14):4:1342–1352.


Dong G, Sheng C, Wang CS, et al. (2010). Selection of evodiamine as a novel topoisomerase i inhibitor by structure-based virtual screening and hit optimization of evodiamine derivatives as anti-tumor agents. Journal of Medicinal Chemistry, 53(21):7521–7531.


Du J, Wang XF, Zhou QM, et al. (2013). Evodiamine induces apoptosis and inhibits metastasis in MDA “American Typewriter”; “American Typewriter”;‑ MB-231 human breast cancer cells in vitro and in vivo. Oncol Rep, 30(2):685-94. doi: 10.3892/or.2013.2498.


Fei XF, Wang BX, T. Li TJ et al. (2003). Evodiamine, a constituent of Evodiae Fructus, induces anti-proliferating effects in tumor cells. Cancer Science, 94(1):92–98.


Huang YC, Guh JH, Teng CM. (2004). Induction of mitotic arrest and apoptosis by evodiamine in human leukemic T-lymphocytes. Life Sciences, 75(1):35–49.


Ji YB. (2011). Active Ingredients of Traditional Chinese Medicine: Pharmacology and Application. People's Medical Publishing House Co., LTD. Connecticut USA


Kan SF, Huang WJ, Lin LC, Wang PS. (2004). Inhibitory effects of evodiamine on the growth of human prostate cancer cell line LNCaP. International Journal of Cancer, 110(5):641–651.


Ko HC, Wang YH, Liou KT et al. (2007). Anti-inflammatory effects and mechanisms of the ethanol extract of Evodia rutaecarpa and its bioactive components on neutrophils and microglial cells. European Journal of Pharmacology, 555(2-3):211–217.


Kobayashi Y. (2003). The nociceptive and anti-nociceptive effects of evodiamine from fruits of Evodia rutaecarpa in mice. Planta Medica, 69(5):425–428.


Lee TJ, Kim EJ, Kim S et al. (2006). Caspase-dependent and caspase-independent apoptosis induced by evodiamine in human leukemic U937 cells. Molecular Cancer Therapeutics, 5(9):2398–2407.


Liao CH, Pan SL, Guh JH et al. (2005). Anti-tumor mechanism of evodiamine, a constituent from Chinese herb Evodiae fructus, in human multiple-drug resistant breast cancer NCI/ADR-RES cells in vitro and in vivo. Carcinogenesis, 26(5):968–975.


Ogasawara M, Matsubara T, Suzuki H. (2001). Inhibitory effects of evodiamine on in vitro invasion and experimental lung metastasis of murine colon cancer cells. Biological and Pharmaceutical Bulletin, 24(8):917–920.


Ogasawara M, Matsunaga T, Takahashi S, Saiki I, Suzuki H. (2002). Anti-invasive and metastatic activities of evodiamine. Biological and Pharmaceutical Bulletin, 25(11):1491–1493.


Rasul A, Yu B, Zhong L, et al. (2012). Cytotoxic effect of evodiamine in SGC-7901 human gastric adenocarcinoma cells via simultaneous induction of apoptosis and autophagy. Oncol Rep, 27(5):1481-7. doi: 10.3892/or.2012.1694


Shin YW, Bae EA, Cai XF, Lee JJ, and Kim DH. (2007). In vitro and in vivo antiallergic effect of the fructus of Evodia rutaecarpa and its constituents, Biological and Pharmaceutical Bulletin, 30(1):197–199, 2007.


Shyu KG, Lin S, Lee CC et al. (2006). Evodiamine inhibits in vitro angiogenesis: implication for anti-tumorgenicity. Life Sciences, 78(19):2234–2243.


Takada Y, Kobayashi Y, Aggarwal BB. (2005). Evodiamine Abolishes Constitutive and Inducible NF- κB Activation by Inhibiting IκBα Kinase Activation, Thereby Suppressing NF-κ B-regulated Antiapoptotic and Metastatic Gene Expression, Up-regulating Apoptosis, and Inhibiting Invasion. The Journal of Biological Chemistry, 280:17203-17212. doi: 10.1074/jbc.M500077200.


Tsai HP, Lin LW, Lai ZY et al. (2010). Immobilizing topoisomerase I on a surface plasmon resonance biosensor chip to screen for inhibitors. Journal of Biomedical Science, 17(1):49.


Wang C, Li S, Wang MW. (2010). Evodiamine-induced human melanoma A375-S2 cell death was mediated by PI3K/Akt/caspase and Fas-L/NF- κ B signaling pathways and augmented by ubiquitin-proteasome inhibition. Toxicology in Vitro, 24(3):898–904.


Wang KL, Hsia SM, Yeh JY, et al. (2013). Anti-Proliferative Effects of Evodiamine on Human Breast Cancer Cells. PLoS One, 8(6):e67297.


Wei WT, Chen H, Wang ZH, et al. (2012). Enhanced anti-tumor efficacy of gemcitabine by evodiamine on pancreatic cancer via regulating PI3K/Akt pathway. Int J Biol Sci, 8(1):1-14.


Yu H, Jin H, Gong W, Wang Z, Liang H. (2013). Pharmacological actions of multi-target-directed evodiamine. Molecules, 18(2):1826-43. doi: 10.3390/molecules18021826.


Yang J, Wu LJ, Tashino SI, et al. (2007). Critical roles of reactive oxygen species in mitochondrial permeability transition in mediating evodiamine-induced human melanoma A375-S2 cell apoptosis. Free Radical Research, 41(10):1099–1108.


Zhang Y, Wu LJ, Tashiro SI, Onodera S, Ikejima T. (2003). Intracellular regulation of evodiamine-induced A375-S2 cell death. Biological and Pharmaceutical Bulletin, 26(11):1543–1547.


Zhang Y, Zhang QH, Wu LJ, et al. (2004). Atypical apoptosis in L929 cells induced by evodiamine isolated from Evodia rutaecarpa. Journal of Asian Natural Products Research, 6(1):19–27.

Apigenin

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

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

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

Induces Apoptosis, Anti-metastatic

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

Chemo-sensitizer, Induces Apoptosis

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

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

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

Breast Cancer, Prostate Cancer

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

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

Ovarian Cancer

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

Pancreatic Cancer

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

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

Induces Apoptosis, Inhibits Angiogenesis and Metastasis.

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

Lung Cancer

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

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

Induces Apoptosis

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

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

Colorectal Cancer

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

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

References

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

 

 

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

 

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

 

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

 

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

 


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

 

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

 

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

 

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

Andrographolide

Cancer: Leukemia, colorectal, lung

Action: Immunomodulatory,anti-inflammatory,anti-metastatic

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

Immunomodulatory Activity

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

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

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

Anti-inflammatory and Leukemic Therapies

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

Anti-inflammatory

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

Lung Cancer Metastasis

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

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

Colorectal Cancer

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

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

Anti-inflammatory, Induces Apoptosis

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

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

References

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


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


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


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


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


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


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


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

Honokiol (See also Injectables)

Cancer:
Lung, breast, prostate, leukemia, colorectal., esophageal., ovarian, myeloma, pancreatic, stomach, uterine

Action: Anti-angiogenic, chemo-sensitizer, multi-drug resistance reversal., anti-inflammatory, anxiolytic, anti-depressant, inhibits VEGF, anti-metastatic, synergistic effects with other cancer treatments

Honokiol is a phenolic compound purified from plants of the Magnolia genus, including Magnolia officinalis (Rehder & Wilson) and Magnolia grandiflora (L.), that exhibits anti-cancer effects in experimental models with various types of cancer cells, including esophageal., ovarian, breast, and lung cancer, as well as myeloma and leukemia. It is speculated that this compound causes cancer cell death in part through targeting mitochondria (Munroe et al., 2007; Chen et al., 2009; Fried & Arbiser, 2009).

Inhibits Angiogenesis, MDR, Anti-inflammatory, Inhibits VEGF

Honokiol is one of two dominant biphenolic compounds isolated from Magnolia spp. bark, and is the most widely researched active constituent of the bark. In vivo studies suggest that honokiol's greatest value is in its multiple anti-cancer actions. In vitro research suggests honokiol has potential to enhance current anti-cancer regimens by inhibiting angiogenesis, promoting apoptosis, providing direct cytotoxic activity, down-regulating cancer cell signaling pathways, regulating genetic expression, enhancing the effects of specific chemotherapeutic agents, radio-sensitizing cancer cells to radiation therapy, and inhibiting multi-drug resistance.

Honokiol also shows potential in preventive health by reducing inflammation and oxidative stress, providing neurological protection, and regulating glucose; in mental illness by its effects against anxiety and depression; and in helping regulate stress response signaling. Its anti-microbial effects demonstrate potential for partnering with anti-viral/antibiotic therapy, and treating secondary infections.

Honokiol may occupy a distinct therapeutic niche because of its unique characteristics: the ability to cross the blood brain barrier (BBB) and blood cerebrospinal fluid barrier (BCSFB), high systemic bioavailability, and its actions on a multiplicity of signaling pathways and genomic activity. There is a need for research on honokiol to progress to human studies and on into clinical use.

The preclinical research on honokiol's broad-ranging capabilities shows its potential as a therapeutic compound for numerous solid and hematological cancers, including its effectiveness in combating multi-drug resistance (MDR) and its synergy with other anti-cancer therapies. Research thus far shows no toxicity or serious adverse effects in animal models.

Honokiol has also been shown to inhibit spread of cancer cells through the lymph system by inhibiting one of the primary pathways involved in growth stimulation related to vascular endothelial growth factor (VEGF) (Wen et al., 2009).

Inhibits Angiogenesis, Gastric Cancer

A 2012 in vivo study in PLoS One showed that honokiol, by inhibiting angiogenic pathways such as STAT-3, dampened peritoneal dissemination of gastric cancer in mice (5 mg/kg delivered intraperitoneally) (Liu et al., 2012).    

Induces Apoptosis; Leukemia

Honokiol induces cell apoptosis in several cell lines, such as leukemia cell lines HL-60, colon cancer cell lines RKO, lung cancer cell lines A549 and CH27 (Hirano et al., 1994; Wang et al., 2004; Hibasami et al., 1998; Konoshima et al., 1991;Yang et al., 2002; Kong et al., 2005). It also has remarkable in vivo anti-tumor activities in tumor mouse models (Bai et al., 2003). Honokiol has demonstrated potent anti-angiogenic and anti-tumor properties against aggressive angiosarcoma by blocking of VEGF-induced VEGF receptor 2 autophosphorylation (Konoshima et al., 1991; Yang et al., 2002).

MDR

Honokiol has also been found to down-regulate the expression of P-glycoprotein at mRNA and protein levels in MCF-7/ADR, a human breast MDR cancer cell line. The down-regulation of P-glycoprotein is accompanied with a partial recovery of the intracellular drug accumulation (Xu et al., 2006).

Prostate Cancer

In addition, it has been shown that prostate cancer cells that failed to respond to hormone withdrawal responded to honokiol-induced apoptosis. It was found to significantly induce death in cells surrounding primary and metastatic prostate cancers, the prostate stromal fibroblasts, marrow stromal cells, and bone marrow-associated endothelial cells. Honokiol is hence a promising nontoxic agent that could be used as an adjuvant with low-dose docetaxel for the treatment of hormone-refractory prostate cancer and its distant bone metastases (Shigemura et al., 2007).

Anti-metastatic

Honokiol inhibited the activity of MMP-9, which may be responsible, in part, for the inhibition of tumor cell invasiveness (Nagase et al., 2001).

Breast Cancer

The development of more targeted and low toxic drugs from traditional Chinese medicines for breast cancer are needed due to most of the anti-breast cancer drugs often being limited because of drug resistance and serious adverse reactions. Results have shown that honokiol inhibited the rate of breast cancer MDA-MB-231 cell growth (Nagalingam et al., 2012).

Synergistic Effects with Other Cancer Treatments

One of the most promising benefits of honokiol is its ability to synergize with other cancer treatments. Clinical trials are desperately needed to validate the potential synergy that has been demonstrated in vitro and in vivo.

Chemotherapy

• A 2013 in vitro study published in the International Journal of Oncology showed that honokiol synergized chemotherapy drugs in Multi-drug-resistant breast cancer (Tian et al., 2013). A 2011 in vitro study published in PLoS One found that honokiol enhanced the apoptotic effects of the anti-cancer drug gemcitabine against pancreatic cancer (Arora et al., 2011).

• In vivo research published in Oncology Letters in 2011 found honokiol enhanced the action of cisplatin against colon cancer (Cheng et al., 2011).

• A 2010 in vitro study from the Journal of Biological Regulators and Homeostatic Agents showed that honokiol resensitized cancer cells to doxorubicin in Multi-drug-resistant uterine cancer (Angelini et al., 2010).

• A 2010 in vitro study published in Toxicology Mechanisms and Methods showed honokiol performed synergistically with the drug imatinib against human leukemia cells (Wang et al., 2010).

• 2008 in vivo research published in the International Journal of Gynecological Cancer showed honokiol to potentiate the activity of cisplatin in murine models of ovarian cancer (Liu et al., 2008).

• 2005 in vitro research published in Blood showed honokiol enhanced the cytotoxicity induced by fludarabine, cladribine, and chlorambucil, indicating it is a potent inducer of apoptosis in B-CLL cells (Battle et al., 2005).

Radiation treatment

• 2012 in vitro research published in Molecular Cancer Therapeutics showed that honokiol was able to sensitize cancer cells to radiation treatments (Ponnurangam et al., 2012).

• A 2011 in vitro study published in American Journal of Physiology Gastrointestinal and Liver Physiology showed honokiol sensitized treatment-resistant colon cancer cells to radiation therapy (He et al., 2011).

Inhibition of multi-drug resistance

Honokiol has been shown to interact with genes that are involved with mechanisms of drug efflux, thus reversing MDR in experimental models. The exact mechanisms of action in this regard are thought to be related to effects of blocking of NF-kB activity, but other mechanisms may also be involved (Xu et al., 2006).

References

Angelini A, Di Ilio C, Castellani ML, Conti P, Cuccurullo F. (2010). Modulation of Multi-drug resistance p-glycoprotein activity by flavonoids and honokiol in human doxorubicin-resistant sarcoma cells (MES-SA/DX-5): Implications for natural sedatives as chemosensitizing agents in cancer therapy. Journal of Biological Regulators & Homeostatic Agents, 24(2). 197-205.


Arora S, Bhardwaj A, Srivastava SK, et al. (2011). Honokiol arrests Cell-cycle, induces apoptosis, and potentiates the cytotoxic effect of gemcitabine in human pancreatic cancer cells. PLoS One, 6(6), e21573. doi: 10.1371/journal.pone.0021573.


Bai X, Cerimele F, Ushio-Fukai M, et al. (2003). Honokiol, a small molecular weight natural product, inhibits angiogenesis in vitro and tumor growth in vivo. J Biol Chem, 278: 35501–7.


Battle TE, Arbiser J, Frank DA. (2005). The natural product honokiol induces caspase-dependent apoptosis in B-cell chronic lymphocytic leukemia (B-CLL) cells. Blood, 106(2), 690-697.


Chen G, Izzo J, Demizu Y, et al. (2009). Different redox states in malignant and nonmalignant esophageal epithelial cells and differential cytotoxic responses to bile acid and honokiol. Antioxid. Redox Signal., 11(5):1083–1095


Cheng N, Xia T, Han Y, et al. (2001). Synergistic anti-tumor effects of liposomal honokiol combined with cisplatin in colon cancer models. Oncology Letters, 2(5), 957-962.


Eliaz I. (2013). Honokiol research review: A promising extract with multiple applications. Natural Medicine Journal., 5(7).


Fried LE, Arbiser JL. (2009). Honokiol, a multifunctional anti-angiogenic and anti-tumor agent. Antioxid. Redox Signal., 1(5):1139–1148. doi: 10.1089/ARS.2009.2440.


He Z, Subramaniam D, Ramalingam S, et al. (2011). Honokiol radiosensitizes colorectal cancer cells: enhanced activity in cells with mismatch repair defects. American Journal of Physiology: Gastrointest and Liver Physiology, 301(5):G929-937.


Hibasami H, Achiwa Y, Katsuzaki H, et al. (1998). Honokiol induces apoptosis in human lymphoid leukemia Molt 4B cells. Int J Mol Med, 2:671–3.


Hirano T, Gotoh M, Oka K. (1994). Natural flavonoids and lignans are potent cytostatic agents against human leukemic HL-60 cells. Life Sci, 55:1061–9.


Hou X, Yuan X, Zhang B, Wang S, Chen Q. (2013). Screening active anti-breast cancer compounds from Cortex Magnolia officinalis by 2D LC-MS. J Sep Sci, 36(4):706-12. doi: 10.1002/jssc.201200896.


Kong ZL, Tzeng SC, Liu YC. (2005). Cytotoxic neolignans: an SAR study. Bioorg Med Chem Lett, 15: 163–6.


Konoshima T, Kozuka M, Tokuda H, et al. (1991). Studies on inhibitors of skin tumor promotion. IX. Neolignans from Magnolia officinalis. J Nat Prod, 54: 816–22.


Liu Y, Chen L, He X, et al. (2010). Enhancement of therapeutic effectiveness by combining liposomal honokiol with cisplatin in ovarian carcinoma. International Journal of Gynecological Cancer, 18(4), 652-659.


Liu SH, Wang KB, Lan KH, et al. (2012). Calpain/SHP-1 interaction by honokiol dampening peritoneal dissemination of gastric cancer in nu/nu mice. PLoS One, 7(8):e43711.


Munroe ME, Arbiser JL, Bishop GA. (2007). Honokiol, a natural plant product, inhibits inflammatory signals and alleviates inflammatory arthritis. J. Immunol., 179(2):753–763


Nagalingam A, Arbiser JL, Bonner MY, Saxena NK, Sharma D. (2012). Honokiol activates AMP-activated protein kinase in breast cancer cells via an LKB1-dependent pathway and inhibits breast carcinogenesis. Breast Cancer Research, 14:R35 doi:10.1186/bcr3128


Nagase H, Ikeda K, Sakai Y. (2001). Inhibitory Effect of Magnolol and Honokiol from Magnolia obovata on Human Fibrosarcoma HT-1080 Invasiveness in vitro. Planta Med, 67(8): 705-708. DOI: 10.1055/s-2001-18345


Ponnurangam S, Mammen JM, Ramalingam S, et al. (2012). Honokiol in combination with radiation targets notch signaling to inhibit colon cancer stem cells. Molecular Cancer Therapeutics, 11(4), 963-972. doi: 10.1371/journal.pone.0043711.


Shigemura K, Arbiser JL, Sun SY, et al. (2007). Honokiol, a natural plant product, inhibits the bone metastatic growth of human prostate cancer cells. Cancer, 109(7), 1279-1289.


Tian W, Deng Y, Li L, et al. (2013). Honokiol synergizes chemotherapy drugs in Multi-drug-resistant breast cancer cells via enhanced apoptosis and additional programmed necrotic death. International Journal of Oncology, 42(2), 721-732. doi: 10.3892/ijo.2012.1739.


Wang Y, Yang Z, Zhao X. (2010). Honokiol induces parapoptosis and apoptosis and exhibits schedule-dependent synergy in combination with imatinib in human leukemia cells. Toxicology Mechanisms and Methods, 20(5), 234-241. doi: 10.3109/15376511003758831.


Wang T, Chen F, Chen Z, et al. (2004). Honokiol induces apoptosis through p53-independent pathway in human colorectal cell line RKO. World J Gastroenterol, 10: 2205–8.


Wen J, Fu AF, Chen LJ, et al. (2009). Liposomal honokiol inhibits VEGF-D-induced lymphangiogenesis and metastasis in xenograft tumor model. International Journal of Cancer, 124(11), 2709-2718. doi: 10.1002/ijc.24244.


Xu D, Lu Q, Hu X. (2006). Down-regulation of P-glycoprotein expression in MDR breast cancer cell MCF-7/ADR by honokiol. Cancer Letters, 243(2), 274-280.


Yang SE, Hsieh MT, Tsai TH, Hsu SL. (2002). Down-modulation of Bcl-XL, release of cytochrome c and sequential activation of caspases during honokiol-induced apoptosis in human squamous lung cancer CH27 cells. Biochemical Pharmacology, 63(9), 1641-1651.

Source

Eliaz I. (2013). Honokiol research review: A promising extract with multiple applications. Natural Medicine Journal., 5(7). Retrieved from http://www.naturalmedicinejournal.com/article_content.asp?edition=1.

Ginsenoside (See also Rg3)

Cancer:
Breast, colorectal., brain, leukemia, acute myeloid leukemia (AML), melanoma, lung, glioblastoma, prostate, fibroblast carcinoma

Action: Multi-drug resistance, apoptosis, anti-cancer, chemotherapy sensitizer, CYP450 regulating, inhibits growth and metastasis, down-regulates MMP-9, enhances 5-FU, anti-inflammatory

Inhibits Growth and Metastasis

Ginsenosides, belonging to a group of saponins with triterpenoid dammarane skeleton, show a variety of pharmacological effects. Among them, some ginsenoside derivatives, which can be produced by acidic and alkaline hydrolysis, biotransformation and steamed process from the major ginsenosides in ginseng plant, perform stronger activities than the major primeval ginsenosides on inhibiting growth or metastasis of tumor, inducing apoptosis and differentiation of tumor and reversing multi-drug resistance of tumor. Therefore ginsenoside derivatives are promising as anti-tumor active compounds and drugs (Cao et al., 2012).

Ginsenoside content can vary widely depending on species, location of growth, and growing time before harvest. The root, the organ most often used, contains saponin complexes. These are often split into two groups: the Rb1 group (characterized by the protopanaxadiol presence: Rb1, Rb2, Rc and Rd) and the Rg1 group (protopanaxatriol: Rg1, Re, Rf, and Rg2). The potential health effects of ginsenosides include anti-carcinogenic, immunomodulatory, anti-inflammatory, anti-allergic, anti-atherosclerotic, anti-hypertensive, and anti-diabetic effects as well as anti-stress activity and effects on the central nervous system (Christensen, 2009).

Ginsenosides are considered the major pharmacologically active constituents, and approximately 12 types of ginsenosides have been isolated and structurally identified. Ginsenoside Rg3 was metabolized to ginsenoside Rh2 and protopanaxadiol by human fecal microflora (Bae et al., 2002). Ginsenoside Rg3 and the resulting metabolites exhibited potent cytotoxicity against tumor cell lines (Bae et al., 2002).

Screen-Shot-2014-03-28-at-11.53.41-am1

Ginseng Extracts (GE); Methanol-(alc-GE) or Water-extracted (w-GE) and ER+ Breast Cancer

Ginseng root extracts and the biologically active ginsenosides have been shown to inhibit proliferation of human cancer cell lines, including breast cancer. However, there are conflicting data that suggest that ginseng extracts (GEs) may or may not have estrogenic action, which might be contraindicated in individuals with estrogen-dependent cancers. The current study was designed to address the hypothesis that the extraction method of American ginseng (Panax quinquefolium) root will dictate its ability to produce an estrogenic response using the estrogen receptor (ER)-positive MCF-7 human breast cancer cell model. MCF-7 cells were treated with a wide concentration range of either methanol-(alc-GE) or water-extracted (w-GE) ginseng root for 6 days.

An increase in MCF-7 cell proliferation by GE indicated potential estrogenicity. This was confirmed by blocking GE-induced MCF-7 cell proliferation with ER antagonists ICI 182,780 (1 nM) and 4-hydroxytamoxifen (0.1 microM). Furthermore, the ability of GE to bind ERalpha or ERbeta and stimulate estrogen-responsive genes was examined. Alc-GE, but not w-GE, was able to increase MCF-7 cell proliferation at low concentrations (5-100 microg/mL) when cells were maintained under low-estrogen conditions. The stimulatory effect of alc-GE on MCF-7 cell proliferation was blocked by the ER antagonists ICI 182,780 or 4-hydroxyta-moxifen. At higher concentrations of GE, both extracts inhibited MCF-7 and ER-negative MDA-MB-231 cell proliferation regardless of media conditions.

These data indicate that low concentrations of alc-GE, but not w-GE, elicit estrogenic effects, as evidenced by increased MCF-7 cell proliferation, in a manner antagonized by ER antagonists, interactions of alc-GE with estrogen receptors, and increased expression of estrogen-responsive genes by alc-GE. Thus, discrepant results between different laboratories may be due to the type of GE being analyzed for estrogenic activity (King et al., 2006).

Anti-cancer

Previous studies suggested that American ginseng and notoginseng possess anti-cancer activities. Using a special heat-preparation or steaming process, the content of Rg3, a previously identified anti-cancer ginsenoside, increased significantly and became the main constituent in the steamed American ginseng. As expected, using the steamed extract, anti-cancer activity increased significantly. Notoginseng has a very distinct saponin profile compared to that of American ginseng. Steaming treatment of notoginseng also significantly increased anti-cancer effect (Wang et al., 2008).

Steam Extraction; Colorectal Cancer

After steaming treatment of American ginseng berries (100-120 ¡C for 1 h, and 120 ¡C for 0.5-4 h), the content of seven ginsenosides, Rg1, Re, Rb1, Rc, Rb2, Rb3, and Rd, decreased; the content of five ginsenosides, Rh1, Rg2, 20R-Rg2, Rg3, and Rh2, increased. Rg3, a previously identified anti-cancer ginsenoside, increased significantly. Two h of steaming at 120 ¡C increased the content of ginsenoside Rg3 to a greater degree than other tested ginsenosides. When human colorectal cancer cells were treated with 0.5 mg/mL steamed berry extract (120 ¡C 2 hours), the anti-proliferation effects were 97.8% for HCT-116 and 99.6% for SW-480 cells.

After staining with Hoechst 33258, apoptotic cells increased significantly by treatment with steamed berry extract compared with unheated extracts. The steaming of American ginseng berries hence augments ginsenoside Rg3 content and increases the anti-proliferative effects on two human colorectal cancer cell lines (Wang et al., 2006).

Glioblastoma

The major active components in red ginseng consist of a variety of ginsenosides including Rg3, Rg5 and Rk1, each of which has different pharmacological activities. Among these, Rg3 has been reported to exert anti-cancer activities through inhibition of angiogenesis and cell proliferation.

It is essential to develop a greater understanding of this novel compound by investigating the effects of Rg3 on a human glioblastoma cell line and its molecular signaling mechanism. The mechanisms of apoptosis by ginsenoside Rg3 were related with the MEK signaling pathway and reactive oxygen species. These data suggest that ginsenoside Rg3 is a novel agent for the chemotherapy of GBM (Choi et al., 2013).

Colon Cancer; Chemotherapy

Rg3 can inhibit the activity of NF-kappaB, a key transcriptional factor constitutively activated in colon cancer that confers cancer cell resistance to chemotherapeutic agents. Compared to treatment with Rg3 or chemotherapy alone, combined treatment was more effective (i.e., there were synergistic effects) in the inhibition of cancer cell growth and induction of apoptosis and these effects were accompanied by significant inhibition of NF-kappaB activity.

NF-kappaB target gene expression of apoptotic cell death proteins (Bax, caspase-3, caspase-9) was significantly enhanced, but the expression of anti-apoptotic genes and cell proliferation marker genes (Bcl-2, inhibitor of apoptosis protein (IAP-1) and X chromosome IAP (XIAP), Cox-2, c-Fos, c-Jun and cyclin D1) was significantly inhibited by the combined treatment compared to Rg3 or docetaxel alone.

These results indicate that ginsenoside Rg3 inhibits NF-kappaB, and enhances the susceptibility of colon cancer cells to docetaxel and other chemotherapeutics. Thus, ginsenoside Rg3 could be useful as an anti-cancer or adjuvant anti-cancer agent (Kim et al., 2009).

Prostate Cancer; Chemo-sensitizer

Nuclear factor-kappa (NF-kappaB) is also constitutively activated in prostate cancer, and gives cancer cells resistance to chemotherapeutic agents. Rg3 has hence also been found to increase susceptibility of prostate (LNCaP and PC-3, DU145) cells against chemotherapeutics; prostate cancer cell growth as well as activation of NF-kappaB was examined. It has been found that a combination treatment of Rg3 (50 microM) with a conventional agent docetaxel (5 nM) was more effective in the inhibition of prostate cancer cell growth and induction of apoptosis as well as G(0)/G(1) arrest accompanied with the significant inhibition of NF-kappaB activity, than those by treatment of Rg3 or docetaxel alone.

The combination of Rg3 (50 microM) with cisplatin (10 microM) and doxorubicin (2 microM) was also more effective in the inhibition of prostate cancer cell growth and NF-kappaB activity than those by the treatment of Rg3 or chemotherapeutics alone. These results indicate that ginsenoside Rg3 inhibits NF-kappaB, and enhances the susceptibility of prostate cancer cells to docetaxel and other chemotherapeutics. Thus, ginsenoside Rg3 could be useful as an anti-cancer agent (Kim et al., 2010).

Colon Cancer

Ginsenosides may not only be useful in themselves, but also for their downstream metabolites. Compound K (20-O-( β -D-glucopyranosyl)-20(S)-protopanaxadiol) is an active metabolite of ginsenosides and induces apoptosis in various types of cancer cells. This study investigated the role of autophagy in compound K-induced cell death of human HCT-116 colon cancer cells. Compound K activated an autophagy pathway characterized by the accumulation of vesicles, the increased positive acridine orange-stained cells, the accumulation of LC3-II, and the elevation of autophagic flux.

Compound K-provoked autophagy was also linked to the generation of intracellular reactive oxygen species (ROS); both of these processes were mitigated by the pre-treatment of cells with the anti-oxidant N-acetylcysteine.   Moreover, compound K activated the c-Jun NH2-terminal kinase (JNK) signaling pathway, whereas down-regulation of JNK by its specific inhibitor SP600125 or by small interfering RNA against JNK attenuated autophagy-mediated cell death in response to compound K.

Notably, compound K-stimulated autophagy as well as apoptosis was induced by disrupting the interaction between Atg6 and Bcl-2. Taken together, these results indicate that the induction of autophagy and apoptosis by compound K is mediated through ROS generation and JNK activation in human colon cancer cells (Kim et al., 2013b).

Lung Cancer; SCC

Korea white ginseng (KWG) has been investigated for its chemo-preventive activity in a mouse lung SCC model. N-nitroso-trischloroethylurea (NTCU) was used to induce lung tumors in female Swiss mice, and KWG was given orally. KWG significantly reduced the percentage of lung SCCs from 26.5% in the control group to 9.1% in the KWG group and in the meantime, increased the percentage of normal bronchial and hyperplasia. KWG was also found to greatly reduce squamous cell lung tumor area from an average of 9.4% in control group to 1.5% in the KWG group.

High-performance liquid chromatography/mass spectrometry identified 10 ginsenosides from KWG extracts, Rb1 and Rd being the most abundant as detected in mouse blood and lung tissue. These results suggest that KWG could be a potential chemo-preventive agent for lung SCC (Pan et al., 2013).

Leukemia

Rg1 was found to significantly inhibit the proliferation of K562 cells in vitro and arrest the cells in G2/M phase. The percentage of positive cells stained by SA-beta-Gal was dramatically increased (P < 0.05) and the expression of cell senescence-related genes was up-regulated. The observation of ultrastructure showed cell volume increase, heterochromatin condensation and fragmentation, mitochondrial volume increase, and lysosomes increase in size and number. Rg1 can hence induce the senescence of leukemia cell line K562 and play an important role in regulating p53-p21-Rb, p16-Rb cell signaling pathway (Cai et al., 2012).

Leukemia, Lymphoma

It has been found that Rh2 inhibits the proliferation of human leukemia cells concentration- and time-dependently with an IC(50) of ~38 µM. Rh2 blocked cell-cycle progression at the G(1) phase in HL-60 leukemia and U937 lymphoma cells, and this was found to be accompanied by the down-regulations of cyclin-dependent kinase (CDK) 4, CDK6, cyclin D1, cyclin D2, cyclin D3 and cyclin E at the protein level. Treatment of HL-60 cells with Rh2 significantly increased transforming growth factor- β (TGF- β ) production, and co-treatment with TGF- β neutralizing antibody prevented the Rh2-induced down-regulations of CDK4 and CDK6, up-regulations of p21(CIP1/WAF1) and p27(KIP1) levels and the induction of differentiation. These results demonstrate that the Rh2-mediated G(1) arrest and the differentiation are closely linked to the regulation of TGF- β production in human leukemia cells (Chung et al., 2012).

NSCLC

Ginsenoside Rh2, one of the components in ginseng saponin, has been shown to have anti-proliferative effect on human NSCLC cells and is being studied as a therapeutic drug for NSCLC. MicroRNAs (miRNAs) are small, non-coding RNA molecules that play a key role in cancer progression and prevention.

A unique set of changes in the miRNA expression profile in response to Rh2 treatment in the human NSCLC cell line A549 has been identified using miRNA microarray analysis. These miRNAs are predicted to have several target genes related to angiogenesis, apoptosis, chromatic modification, cell proliferation and differentiation. Thus, these results may assist in the better understanding of the anti-cancer mechanism of Rh2 in NSCLC (An et al., 2012).

Ginsenoside Concentrations

Ginsenosides, the major chemical composition of Chinese white ginseng (Panax ginseng C. A. Meyer), can inhibit tumor, enhance body immune function, prevent neurodegeneration. The amount of ginsenosides in the equivalent extraction of the nanoscale Chinese white ginseng particles (NWGP) was 2.5 times more than that of microscale Chinese white ginseng particles (WGP), and the extractions from NWGP (1000 microg/ml) reached a high tumor inhibition of 64% exposed to human lung carcinoma cells (A549) and 74% exposed to human cervical cancer cells (Hela) after 72 hours. Thia work shows that the nanoscale Chinese WGP greatly improves the bioavailability of ginsenosides (Ji et al., 2012).

Chemotherapy Side-effects

Pre-treatment with American ginseng berry extract (AGBE), a herb with potent anti-oxidant capacity, and one of its active anti-oxidant constituents, ginsenoside Re, was examined for its ability to counter cisplatin-induced emesis using a rat pica model. In rats, exposure to emetic stimuli such as cisplatin causes significant kaolin (clay) intake, a phenomenon called pica. We therefore measured cisplatin-induced kaolin intake as an indicator of the emetic response.

Rats were pre-treated with vehicle, AGBE (dose range 50–150 mg/kg, IP) or ginsenoside Re (2 and 5 mg/kg, IP). Rats were treated with cisplatin (3 mg/kg, IP) 30 min later. Kaolin intake, food intake, and body weight were measured every 24 hours, for 120 hours.

A significant dose-response relationship was observed between increasing doses of pre-treatment with AGBE and reduction in cisplatin-induced pica. Kaolin intake was maximally attenuated by AGBE at a dose of 100 mg/kg. Food intake also improved significantly at this dose (P<0.05). pre-treatment ginsenoside (5 mg/kg) also decreased kaolin intake >P<0.05). In vitro studies demonstrated a concentration-response relationship between AGBE and its ability to scavenge superoxide and hydroxyl.

Pre-treatment with AGBE and its major constituent, Re, hence attenuated cisplatin-induced pica, and demonstrated potential for the treatment of chemotherapy-induced nausea and vomiting. Significant recovery of food intake further strengthens the conclusion that AGBE may exert an anti-nausea/anti-emetic effect (Mehendale et al., 2005).

MDR

Because ginsenosides are structurally similar to cholesterol, the effect of Rp1, a novel ginsenoside derivative, on drug resistance using drug-sensitive OVCAR-8 and drug-resistant NCI/ADR-RES and DXR cells. Rp1 treatment resulted in an accumulation of doxorubicin or rhodamine 123 by decreasing MDR-1 activity in doxorubicin-resistant cells. Rp1 synergistically induced cell death with actinomycin D in DXR cells. Rp1 appeared to redistribute lipid rafts and MDR-1 protein.

Rp1 reversed resistance to actinomycin D by decreasing MDR-1 protein levels and Src phosphorylation with modulation of lipid rafts. Addition of cholesterol attenuated Rp1-induced raft aggregation and MDR-1 redistribution. Rp1 and actinomycin D reduced Src activity, and overexpression of active Src decreased the synergistic effect of Rp1 with actinomycin D. Rp1-induced drug sensitization was also observed with several anti-cancer drugs, including doxorubicin. These data suggest that lipid raft-modulating agents can be used to inhibit MDR-1 activity and thus overcome drug resistance (Yun et al., 2013).

Hypersensitized MDR Breast Cancer Cells to Paclitaxel

The effects of Rh2 on various tumor-cell lines for its effects on cell proliferation, induction of apoptosis, and potential interaction with conventional chemotherapy agents were investigated. Jia et al., (2004) showed that Rh2 inhibited cell growth by G1 arrest at low concentrations and induced apoptosis at high concentrations in a variety of tumor-cell lines, possibly through activation of caspases. The apoptosis induced by Rh2 was mediated through glucocorticoid receptors. Most interestingly, Rh2 can act either additively or synergistically with chemotherapy drugs on cancer cells. Particularly, it hypersensitized multi-drug-resistant breast cancer cells to paclitaxel.

These results suggest that Rh2 possesses strong tumor-inhibiting properties, and potentially can be used in treatments for multi-drug-resistant cancers, especially when it is used in combination with conventional chemotherapy agents.

MDR; Leukemia, Fibroblast Carcinoma

It was previously reported that a red ginseng saponin, 20(S)-ginsenoside Rg3 could modulate MDR in vitro and extend the survival of mice implanted with ADR-resistant murine leukemia P388 cells. A cytotoxicity study revealed that 120 microM of Rg3 was cytotoxic against a multi-drug-resistant human fibroblast carcinoma cell line, KB V20C, but not against normal WI 38 cells in vitro. 20 microM Rg3 induced a significant increase in fluorescence anisotropy in KB V20C cells but not in the parental KB cells. These results clearly show that Rg3 decreases the membrane fluidity thereby blocking drug efflux (Kwon et al., 2008).

MDR

Ginsenoside Rb1 is a representative component of panaxadiol saponins, which belongs to dammarane-type tritepenoid saponins and mainly exists in family araliaceae. It has been reported that ginsenoside Rb1 has diverse biological activities. The research development in recent decades on its pharmacological effects of cardiovascular system, anti-senility, reversing multi-drug resistance of tumor cells, adjuvant anti-cancer chemotherapy, and promoting peripheral nerve regeneration have been established (Jia et al., 2008).

Enhances Cyclophosphamide

Cyclophosphamide, an alkylating agent, has been shown to possess various genotoxic and carcinogenic effects, however, it is still used extensively as an anti-tumor agent and immunosuppressant in the clinic. Previous reports reveal that cyclophosphamide is involved in some secondary neoplasms.

C57BL/6 mice bearing B16 melanoma and Lewis lung carcinoma cells were respectively used to estimate the anti-tumor activity in vivo. The results indicated that oral administration of Rh(2) (5, 10 and 20 mg/kg body weight) alone has no obvious anti-tumor activity and genotoxic effect in mice, while Rh(2) synergistically enhanced the anti-tumor activity of cyclophosphamide (40 mg/kg body weight) in a dose-dependent manner.

Rh(2) decreased the micronucleus formation in polychromatic erythrocytes and DNA strand breaks in white blood cells in a dose-dependent way. These results suggest that ginsenoside Rh(2) is able to enhance the anti-tumor activity and decrease the genotoxic effect of cyclophosphamide (Wang, Zheng, Liu, Li, & Zheng, 2006).

Down-regulates MMP-9, Anti-metastatic

The effects of the purified ginseng components, panaxadiol (PD) and panaxatriol (PT), were examined on the expression of matrix metalloproteinase-9 (MMP-9) in highly metastatic HT1080 human fibrosarcoma cell line. A significant down-regulation of MMP-9 by PD and PT was detected by Northern blot analysis; however, the expression of MMP-2 was not changed by treatment with PD and PT. The results of the in vitro invasion assay revealed that PD and PT reduced tumor cell invasion through a reconstituted basement membrane in the transwell chamber. Because of the similarity of chemical structure between PD, PT and dexamethasone (Dexa), a synthetic glucocorticoid, we investigated whether the down-regulation of MMP-9 by PD and PT were mediated by the nuclear translocation of glucocorticoid receptor (GR). Increased GR in the nucleus of HT1080 human fibrosarcoma cells treated by PD and PT was detected by immunocytochemistry.

Western blot and gel retardation assays confirmed the increase of GR in the nucleus after treatment with PD and PT. These results suggest that GR-induced down-regulation of MMP-9 by PD and PT contributes to reduce the invasive capacity of HT1080 cells (Park et al., 1999).

Enhances 5-FU; Colorectal Cancer

Panaxadiol (PD) is the purified sapogenin of ginseng saponins, which exhibit anti-tumor activity. The possible synergistic anti-cancer effects of PD and 5-FU on a human colorectal cancer cell line, HCT-116, have been investigated.

The significant suppression on HCT-116 cell proliferation was observed after treatment with PD (25 microM) for 24 and 48 hours. Panaxadiol (25 microM) markedly (P < 0.05) enhanced the anti-proliferative effects of 5-FU (5, 10, 20 microM) on HCT-116 cells compared to single treatment of 5-FU for 24 and 48 hours.

Flow cytometric analysis on DNA indicated that PD and 5-FU selectively arrested cell-cycle progression in the G1 phase and S phase (P < 0.01), respectively, compared to the control condition. Combination use of 5-FU with PD significantly (P < 0.001) increased cell-cycle arrest in the S phase compared to that treated by 5-FU alone.

The combination of 5-FU and PD significantly enhanced the percentage of apoptotic cells when compared with the corresponding cell groups treated by 5-FU alone (P < 0.001). Panaxadiol hence enhanced the anti-cancer effects of 5-FU on human colorectal cancer cells through the regulation of cell-cycle transition and the induction of apoptotic cells (Li et al., 2009).

Colorectal Cancer

The possible synergistic anti-cancer effects of Panaxadiol (PD) and Epigallocatechin gallate (EGCG), on human colorectal cancer cells and the potential role of apoptosis in the synergistic activities, have been investigated.

Cell growth was suppressed after treatment with PD (10 and 20   µm) for 48   h. When PD (10 and 20   µm) was combined with EGCG (10, 20, and 30   µm), significantly enhanced anti-proliferative effects were observed in both cell lines. Combining 20   µm of PD with 20 and 30   µm of EGCG significantly decreased S-phase fractions of cells. In the apoptotic assay, the combination of PD and EGCG significantly increased the percentage of apoptotic cells compared with PD alone (p   <   0.01).

Data from this study suggested that apoptosis might play an important role in the EGCG-enhanced anti-proliferative effects of PD on human colorectal cancer cells (Du et al., 2013).

Colorectal Cancer; Irinotecan

Cell cycle analysis demonstrated that combining irinotecan treatment with panaxadiol significantly increased the G1-phase fractions of cells, compared with irinotecan treatment alone. In apoptotic assays, the combination of panaxadiol and irinotecan significantly increased the percentage of apoptotic cells compared with irinotecan alone (P<0.01). Increased activity of caspase-3 and caspase-9 was observed after treating with panaxadiol and irinotecan.

Data from this study suggested that caspase-3- and caspase-9-mediated apoptosis may play an important role in the panaxadiol enhanced anti-proliferative effects of irinotecan on human colorectal cancer cells (Du et al., 2012).

Anti-inflammatory

Ginsenoside Re inhibited IKK- β phosphorylation and NF- κ B activation, as well as the expression of pro-inflammatory cytokines, TNF- α and IL-1 β , in LPS-stimulated peritoneal macrophages, but it did not inhibit them in TNF- α – or PG-stimulated peritoneal macrophages. Ginsenoside Re also inhibited IRAK-1 phosphorylation induced by LPS, as well as IRAK-1 and IRAK-4 degradations in LPS-stimulated peritoneal macrophages.

Orally administered ginsenoside Re significantly inhibited the expression of IL-1 β and TNF- α on LPS-induced systemic inflammation and TNBS-induced colitis in mice. Ginsenoside Re inhibited colon shortening and myeloperoxidase activity in TNBS-treated mice. Ginsenoside Re reversed the reduced expression of tight-junction-associated proteins ZO-1, claudin-1, and occludin. Ginsenoside Re (20 mg/kg) inhibited the activation of NF- κ B in TNBS-treated mice. On the basis of these findings, ginsenoside Re may ameliorate inflammation by inhibiting the binding of LPS to TLR4 on macrophages (Lee et al., 2012).

Induces Apoptosis

Compound K activated an autophagy pathway characterized by the accumulation of vesicles, the increased positive acridine orange-stained cells, the accumulation of LC3-II, and the elevation of autophagic flux. Compound K activated the c-Jun NH2-terminal kinase (JNK) signaling pathway, whereas down-regulation of JNK by its specific inhibitor SP600125 or by small interfering RNA against JNK attenuated autophagy-mediated cell death in response to compound K. Compound K also provoked apoptosis, as evidenced by an increased number of apoptotic bodies and sub-G1 hypodiploid cells, enhanced activation of caspase-3 and caspase-9, and modulation of Bcl-2 and Bcl-2-associated X protein expression (Kim et al., 2013b).

Lung Cancer

AD-1, a ginsenoside derivative, concentration-dependently reduces lung cancer cell viability without affecting normal human lung epithelial cell viability. In A549 and H292 lung cancer cells, AD-1 induces G0/G1 cell-cycle arrest, apoptosis and ROS production. The apoptosis can be attenuated by a ROS scavenger – N-acetylcysteine (NAC). In addition, AD-1 up-regulates the expression of p38 and ERK phosphorylation. Addition of a p38 inhibitor, SB203580, suppresses the AD-1-induced decrease in cell viability. Furthermore, genetic silencing of p38 attenuates the expression of p38 and decreases the AD-1-induced apoptosis.

These data support development of AD-1 as a potential agent for lung cancer therapy (Zhang et al., 2013).

Pediatric AML

In this study, Chen et al. (2013) demonstrated that compound K, a major ginsenoside metabolite, inhibited the growth of the clinically relevant pediatric AML cell lines in a time- and dose-dependent manner. This growth-inhibitory effect was attributable to suppression of DNA synthesis during cell proliferation and the induction of apoptosis was accompanied by DNA double strand breaks. Findings suggest that as a low toxic natural reagent, compound K could be a potential drug for pediatric AML intervention and to improve the outcome of pediatric AML treatment.

Melanoma

Jeong et al. (2013) isolated 12 ginsenoside compounds from leaves of Panax ginseng and tested them in B16 melanoma cells. It significantly reduced melanin content and tyrosinase activity under alpha-melanocyte stimulating hormone- and forskolin-stimulated conditions. It significantly reduced the cyclic AMP (cAMP) level in B16 melanoma cells, and this might be responsible for the regulation down of MITF and tyrosinase. Phosphorylation of a downstream molecule, a cAMP response-element binding protein, was significantly decreased according to Western blotting and immunofluorescence assay. These data suggest that A-Rh4 has an anti-melanogenic effect via the protein kinase A pathway.

Leukemia

Rg1 can significantly inhibit the proliferation of leukemia cell line K562 in vitro and arrest the cells in G2/M phase. The percentage of positive cells stained by SA-beta-Gal was dramatically increased (P < 0.05) and the expression of cell senescence-related genes was up-regulated. The observation of ultrastructure showed cell volume increase, heterochromatin condensation and fragmentation, mitochondrial volume increase, and lysosomes increase in size and number (Cai et al., 2012).

Ginsenosides and CYP 450 Enzymes

In vitro experiments have shown that both crude ginseng extract and total saponins at high concentrations (.2000 mg/ml) inhibited CYP2E1 activity in mouse and human microsomes (Nguyen et al., 2000). Henderson et al. (1999) reported the effects of seven ginsenosides and two eleutherosides (active components of the ginseng root) on the catalytic activity of a panel of cDNA-expressed CYP isoforms (CYP1A2, CYP2C9, CYP2C19, CYP2D6, and CYP3A4) using 96-well plate fluorometrical assay.

Of the constituents tested, Ginsenoside Rd caused weak inhibitory activity against CYP3A4, CYP2D6, CYP2C19,and CYP2C9, but ginsenoside Re and ginsenoside Rf (200 mM) produced a 70% and 54%increase in the activity of CYP2C9 and CYP3A4, respectively. The authors suggested that the activating effects of ginsenosides on CYP2C9 and CYP3A4 might be due to a matrix effect caused by the test compound fluorescing at the same wavelength as the metabolite of the marker substrates. Chang et al. (2002) reported the effects of two types of ginseng extract and ginsenosides (Rb1, Rb2, Rc, Rd, Re, Rf, and Rg1) on CYP1 catalytic activities.

The ginseng extracts inhibited human recombinant CYP1A1, CYP1A2, and CYP1B1 activities in a concentration-dependent manner. Rb1, Rb2, Rc, Rd, Re, Rf, and Rg1 at low concentrations had no effect on CYP1 activities, but Rb1, Rb2, Rc, Rd, and Rf at a higher ginsenoside concentration (50 mg/ml) inhibited these activities. These results indicated that various ginseng extracts and ginsenosides inhibited CYP1 activity in an enzyme-selective and extract-specific manner (Zhou et al., 2003).

References

An IS, An S, Kwon KJ, Kim YJ, Bae S. (2012). Ginsenoside Rh2 mediates changes in the microRNA expression profile of human non-small-cell lung cancer A549 cells. Oncol Rep, 29(2):523-8. doi: 10.3892/or.2012.2136.



Bae EA, Han MJ, Choo MK et al. (2002). Metabolism of 20(S)- and 20(R)-ginsenoside R-g3 by human intestinal bacteria and its relation to in vitro biological activities. Biol. Pharm. Bull, 25:58–63.


Cai S, Zhou Y, Liu J, et al. (2012). Experimental study on human leukemia cell line K562 senescence induced by ginsenoside Rg1. Zhongguo Zhong Yao Za Zhi, 37(16):2424-8.


Cao M, Yu HS, Song XB, Ma BP. (2012) Advances in the study of derivatization of ginsenosides and their anti-tumor structure-activity relationship. Yao Xue Xue Bao, 47(7):836-43.


Chang TKH, Chen J, Benetton SA et al. (2002). In vitro effect of standardized ginseng extracts and individual ginsenosides on the catalytic activity of human CYP1A1, CYP1A2, and CYP1B1. Drug Metab. Dispos, 30:378–384.


Chen Y, Xu Y, Zhu Y, Li X. (2013). Anti-cancer effects of ginsenoside compound k on pediatric acute myeloid leukemia cells. Cancer Cell Int, 13(1):24. doi: 10.1186/1475-2867-13-24.


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, 30(3): 1362-1370. doi: 10.3892/or.2013.2555.


Christensen LP. (2009). Ginsenosides chemistry, biosynthesis, analysis, and potential health effects. Adv Food Nutr Res., 55:1-99. doi: 10.1016/S1043-4526(08)00401-4.


Chung KS, Cho SH, Shin JS, et al. (2013). Ginsenoside Rh2 induces Cell-cycle arrest and differentiation in human leukemia cells by upregulating TGF- β expression. Carcinogenesis, 34(2):331-40. doi: 10.1093/carcin/bgs341.


Du GJ, Wang CZ, Zhang ZY, et al. (2012) Caspase-mediated pro-apoptotic interaction of panaxadiol and irinotecan in human colorectal cancer cells. J Pharm Pharmacol, 64(5):727-34. doi: 10.1111/j.2042-7158.2012.01463.x.


Du GJ, Wang CZ, Qi LW, et al. (2013). The synergistic apoptotic interaction of panaxadiol and epigallocatechin gallate in human colorectal cancer cells. Phytother Res, 27(2):272-7. doi: 10.1002/ptr.4707.


Henderson GL, Harkey MR, Gershwin, ME, et al. (1999). Effects of ginseng components on c-DNA-expressed cytochrome P450 enzyme catalytic activity. Life Sci, PL209–PL214.


Jeong YM, Oh WK, Tran TL, et al. (2013). Aglycone of Rh4 inhibits melanin synthesis in B16 melanoma cells: possible involvement of the protein kinase A pathway. Biosci Biotechnol Biochem, 77(1):119-25.


Ji Y, Rao Z, Cui J, et al. (2012). Ginsenosides extracted from nanoscale Chinese white ginseng enhances anti-cancer effect. J Nanosci Nanotechnol, 12(8):6163-7.


Jia WW, Bu X, Philips D, et al. (2004). Rh2, a compound extracted from ginseng, hypersensitizes Multi-drug-resistant tumor cells to chemotherapy. Can J Physiol Pharmacol, 82(7):431-7.


Jia JM, Wang ZQ, Wu LJ, Wu YL. (2008). Advance of pharmacological study on ginsenoside Rb1. Zhongguo Zhong Yao Za Zhi, 33(12):1371-7.


Kim YJ, Yamabe N, Choi P, et al. (2013a) Efficient Thermal Deglycosylation of Ginsenoside Rd and Its Contribution to the Improved Anti-cancer Activity of Ginseng. J Agric Food Chem.


Kim AD, Kang KA, Kim HS, et al. (2013b). A ginseng metabolite, compound K, induces autophagy and apoptosis via generation of reactive oxygen species and activation of JNK in human colon cancer cells. Cell Death Dis, 4:e750. doi: 10.1038/cddis.2013.273.


Kim SM, Lee SY, Cho JS, et al. (2010). Combination of ginsenoside Rg3 with docetaxel enhances the susceptibility of prostate cancer cells via inhibition of NF-kappaB. Eur J Pharmacol, 631(1-3):1-9. doi: 10.1016/j.ejphar.2009.12.018.


Kim SM, Lee SY, Yuk DY, et al. (2009). Inhibition of NF-kappaB by ginsenoside Rg3 enhances the susceptibility of colon cancer cells to docetaxel. Arch Pharm Res, 32:755–765. doi: 10.1007/s12272-009-1515-4.


King ML, Adler SR, Murphy LL. (2006). Extraction-dependent effects of American ginseng (Panax quinquefolium) on human breast cancer cell proliferation and estrogen receptor activation. Integr Cancer Ther, 5(3):236-43.


Kwon HY, Kim EH, Kim SW, et al. (2008). Selective toxicity of ginsenoside Rg3 on Multi-drug-resistant cells by membrane fluidity modulation. Arch Pharm Res, 31(2):171-7.


Lee IA, Hyam SR, Jang SE, Han MJ, Kim DH. (2012). Ginsenoside Re ameliorates inflammation by inhibiting the binding of lipopolysaccharide to TLR4 on macrophages. J Agric Food Chem, 60(38):9595-602.


Li XL, Wang CZ, Mehendale SR, et al. (2009). Panaxadiol, a purified ginseng component, enhances the anti-cancer effects of 5-fluorouracil in human colorectal cancer cells. Cancer Chemother Pharmacol, 64(6):1097-104. doi: 10.1007/s00280-009-0966-0.


Mehendale S, Aung H, Wang A, et al. (2005). American ginseng berry extract and ginsenoside Re attenuate cisplatin-induced kaolin intake in rats. Cancer Chemotherapy and Pharmacology, 56(1):63-9. doi: 10.1007/s00280-004-0956-1.


Nguyen TD, Villard PH, Barlatier A et al. (2000). Panax vietnamensis protects mice against carbon tetrachloride-induced hepatotoxicity without any modification of CYP2E1 gene expression. Planta Med, 66:714–719.


Pan J, Zhang Q, Li K, et al. (2013). Chemoprevention of lung squamous cell carcinoma by ginseng. Cancer Prev Res (Phila), 6(6):530-9. doi: 10.1158/1940-6207.CAPR-12-0366.


Park MT, Cha HJ, Jeong JW, et al. (1999). Glucocorticoid receptor-induced down-regulation of MMP-9 by ginseng components, PD and PT contributes to inhibition of the invasive capacity of HT1080 human fibrosarcoma cells. Mol Cells, 9(5):476-83.


Wang CZ and Yuan CS. (2008). Potential Role of Ginseng in the Treatment of Colorectal Cancer. Am. J. Chin. Med, 36:1019. doi: 10.1142/S0192415X08006545


Wang Z, Zheng Q, Liu K, Li G, Zheng R. (2006). Ginsenoside Rh(2) enhances anti-tumor activity and decreases genotoxic effect of cyclophosphamide. Basic Clin Pharmacol Toxicol, 98(4):411-5.


Wang CZ, Zhang B, Song WX, et al. (2006). Steamed American ginseng berry: ginsenoside analyzes and anti-cancer activities. Journal of agricultural and food chemistry, 54(26):9936-42.


Yun UJ, Lee JH, Koo KH, et al. (2013). Lipid raft modulation by Rp1 reverses Multi-drug resistance via inactivating MDR-1 and Src inhibition. Biochem Pharmacol, 85(10):1441-53. doi: 10.1016/j.bcp.2013.02.025.


Zhang LH, Jia YL, Lin XX, et al. (2013). AD-1, a novel ginsenoside derivative, shows anti-lung cancer activity via activation of p38 MAPK pathway and generation of reactive oxygen species. Biochim Biophys Acta, 1830(8):4148-59. doi: 10.1016/j.bbagen.2013.04.008.


Zhou Sf, Gao Yh, Jiang Wq et al. (2003) Interactions of Herbs with Cytochrome P450. DRUG METABOLISM REVIEWS, 35(1):35–98.