Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
pISSN 2288-6982 l eISSN 2288-7105 Biodesign MINI REVIEW P 01-17 Natural products targeting STAT3 signaling pathways in cancer cells Yena Jin1,2, Younghwan Kim1, Yu-Jin Lee1, Dong Cho Han1,2 and Byoung-Mog Kwon1,2,* 1 Laboratory of Chemical Biology and Genomics, Korea Research Institute of Bioscience and Biotechnology, 2University of Science and Technology in Korea, 125 Gwahakro Yoosunggu, Daejeon 34141, Republic of Korea. *Correspondence: [email protected] Cancer is a leading cause of death worldwide. The roles and significance of Signal Transducer and Activator of Transcription 3 (STAT3) in human cancers have been extensively studied. STAT3 is a promising therapeutic target for cancer drug discovery. Herbal medicine-derived secondary metabolites efficiently impact all cancer hallmarks; therefore, natural products are a major source of new cancer drug development. Many natural products can inhibit STAT signaling pathways in cancer cells by targeting different pathways. In this review, we describe the structures of STAT3 inhibitors isolated from herbal medicines and discuss their targeting molecules and modes of action. INTRODUCTION Despite scientific and technological advancements that have led to tremendous progress in treating and preventing cancer, cancer is a leading cause of death in countries of all income levels. In 2012, an estimated 14.1 million new cancer cases and 8.2 million cancer deaths occurred worldwide (Torre et al., 2016). Dramatically improving our understanding of the genomic aberrations of tumors has allowed for the development of new drugs directed at specific targets; one example is the tyrosine kinase inhibitor imatinib used to treat blood cancers. One strategy for developing cancer drugs is to target drivers of cancer cell proliferation, metastasis, and survival, as well as cancer stem cells (Gonda and Ramsay, 2015; Medema, 2013). As shown Figure 1, Signal Transducer and Activator of Transcription 3 (STAT3) proteins play crucial roles in cellular signaling pathways that regulate diverse biological processes, including cell proliferation, differentiation, survival, inflammatory response, immunity, cancer stem cells, and angiogenesis (Yu et al., 2009). Seven members of the STAT family (STAT 1, 2, 3, 4, 5a, 5b, and 6) have been identified in the mammalian system; these members are activated by signals from cytokines and growth factor receptors and are translocated to the nucleus to regulate target gene transcription (Bromberg et al., 2000). STAT3 has received particular attention owing to the constitutive activation of STAT3 in a variety of tumors, including melanoma, and pancreatic, lung, colorectal, ovarian, endometrial, cervical, breast, brain, renal, prostate cancers, as well as head and neck squamous cell carcinoma (HNSCC), glioma, lymphoma, and leukemia (Yuan et al., 2015). Many studies have shown that aberrant STAT3 activation contributes to cell proliferation, differentiation, migration, and survival, as well as cancer stem support (Xiong et al., 2014). This unforeseen STAT3 activation could partially account for the drug resistance that often follows bdjn.org the primary targeted therapies and chemotherapies (Spitzner et al., 2014). Cragg et al reported that nature is a valuable source for potential chemotherapeutic agents (Cragg et al., 2014). Herbal medicine has played a vital role in disease treatment since prehistoric times (Assefa et al., 2010). It has been reported that 48.6% of antitumor medicines are derived, either directly or indirectly, from natural products (Cragg et al., 2009). Nutraceuticals can modulate inflammatory pathways and thus affect tumor survival, proliferation, invasion, angiogenesis, and metastasis (Gupta et al., 2010). This review focuses on STAT3 inhibitors isolated from herbal medicines and their target molecules and modes of action for the modulation of STAT3 activity. STAT3 IN CANCER Cancer is a class of diseases characterized by deregulated cell proliferation, cell invasion, and cell capacity to expand to secondary sites of the body. Cancer progression is a multistep and complex process that begins with abnormal cells with malignant potential or neoplastic characteristics and continues with tumor growth, stromal invasion, and metastasis. These events reply not only on intrinsic tumor characteristics but also on tumor microenvironment, which includes the surrounding and supportive stroma, humoral factors, different effectors of the immune system, and the vasculature (Harris and McCormick, 2010). STAT3 is crucial in tumor cell proliferation, invasion, and migration, and is capable of inducing the epithelial-mesenchymal transition (EMT), regulating the tumor microenvironment and promoting cancer stem cell (CSC) self-renewal and differentiation which all benefit the progression of cancer (Siveen et al., 2014). Therefore, constitutively activated STAT3 may be a therapeutic Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 1 Natural products targeting STAT3 signaling pathways in cancer cells FIGURE 1 I JAK/STAT3 signaling pathways. JAK/STAT3 pathways are activated through the binding of recruited cytokines or growth factors. Activated STAT3 forms a homo- or heterodimer and translocates from the cytoplasm to the nucleus where it binds DNA and coactivators and induces gene transcription. target for the treatment of human cancers. Table 1 outlines pre-clinical studies of pSTAT3 expression in various human tumors (Wang et al., 2012). There was a significant correlation between higher pSTAT3 and worse tumor outcomes, such as breast cancers exhibiting invasiveness, and head and neck as well as gastric cancers exhibiting nodal metastasis and/or existing in late clinical stages. As shown in Figure 1, STAT3 is activated by cytokines or growth factors (Debnath et al., 2012). STAT3 molecules form homodimers after phosphorylation of the tyrosine-705 residue that translocates into the nucleus. These activated STAT3 dimers promote proliferation primarily by stimulating the transcription of key cancer genes linked to tumor cell proliferation, such as cyclin D1, cyclin B and cdc2, which are involved in the regulation of the cell cycle. Activated STAT3 contributes to malignancy by preventing apoptosis pathways, which involves the increased expression of several anti-apoptotic proteins, such as survivin and members of the B-cell lymphoma (Bcl) family (Bcl-xL, Bcl-2 and TABLE 1 I Constitutive STAT3 activation and oncogenesis 2 Cancer type Constitutively activated STAT3 Mediators (or regulators) Breast cancer Human cell lines EGF/EGFR, Src, or JAK Endometrial and cervical cancer Human cell lines Down-stream regulation and target genes Cell cycle progression Anti-apoptosis by caspase-3 Genes: Bcl-XL, survivin and Mcl-1 Lung cancer Human cell lines EGF/EGFR, IL-6 and HGF, followed by Src, or mutant EGFR or CXCL12/CXCR4-JAK2 Multiple myeloma Human cell lines IL-6/JAK Ovarian cancer Human cell lines Genes: Bcl-XL, cyclin D1 Sarcomas Human cell lines Proliferation and anti-apoptosis by caspases Head and neck cancer Human HNSCC lines primary cultures of oral keratinocytes IL-6/gp130, TGF-α Or EGFR Proliferation, tumor growth Genes: cyclin D1, Bcl-2 and Bcl-XL Prostate cancer Human cell lines In vivo mouse models JAK1 or JAK2 Proliferation, tumor growth, anti-apoptosis by caspase-3 Melanoma Human cell lines In vivo mouse models Src, but not EGFR or JAK Proliferation and anti-apoptosis Genes: Bcl-XL, Mcl-1 Hepatocellular carcinoma Human cell lines In vivo mouse models TGF-α, or dysfunctional TGF-β signaling Proliferation, Genes: cyclin A, c-jun, c-fos, and c-myc Pancreatic cancer Human cell lines In vivo mouse models Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign Anti-apoptosis proliferation Anti-apoptosis , Gene: Bcl-XL cyclin D1 and p-STAT3: mutually exclusive events Tumor growth, angiogenesis and metastasis Gene: VEGF bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon Mcl-1), which are also known as direct target genes of STAT3 (Yu and Jove, 2004). STAT3 has also been found to be involved in colorectal cancer cell invasion and migration through the regulation of gene expression such as E-cadherin, VEGF, and matrix metalloproteinases (MMPs) (Wendt et al., 2014). CSCs, have been identified in a range of different types of cancer, and CSCs have a high potential for metastasis; in addition, owing to their resistance to cancer therapy, they often cause relapse after treatment (Visvader and Lindeman, 2012). It has been reported that STAT3 plays essential roles in maintaining cancer stemness and tumorigenic potential (Fouse and Costello, 2013). INHIBITION OF STAT3 ACTIVITY signaling (SOCS). Therefore, STAT activity could be modulated by a variety of methods including blocking ligand-receptor binding, inhibiting upstream kinases, and activating or inducing the expression of negative modulators, as well as indirect modulations such as reactive oxygen species (ROS). Many STAT3 inhibitors have been identified using a variety of techniques, including screens of large compound libraries (Schust and Berg, 2004), computer assisted virtual screens using the STAT3 crystal structure (Siddiquee et al., 2007), STAT3 inhibitors isolated from herbal medicines (Arumuggam et al., 2015), fragment-based design for a novel STAT3 inhibitor (Yu et al., 2013), and STAT3-binding peptides (Dhanik et al., 2012). The various targets for the modulation of STAT3 activity are summarized in Figure 2. Herbal medicines have been used for the treatment of cancer and vinca alkaloids such as vinblastine and vincristine were the first plant-derived anti-cancer drugs (Cragg et al., 2009). Many of the recently reported natural compounds exhibiting anti-tumor activities inhibit the activity of tumor-promoting proteins such as phosphatidylinositol-3-kinase (PI3K), B-cell lymphoma 2 (Bcl-2), and STAT3 (Chinembiri et al., 2014). Many natural products can inhibit STAT3 activity, and several natural or dietary compounds have been shown to possess in vitro and/or in vivo inhibitory effects against STAT3 (Arumuggam et al., 2015). In the present review, we focus on the STAT3 inhibitors isolated from herbal medicines, and discuss the targets and mode of actions of the STAT3 inhibitors. As previously mentioned, small molecule inhibitors targeting various members of the STAT3 signaling Since the first reported as an oncogene in 1995, STAT3 has been observed to be activated in many types of cancer (Watson and Miller, 1995). STAT3 plays major roles in tumor progression and metastasis by modulating the tumor microenvironment. Furthermore, JAK2/STAT3 signaling is important in tumorassociated immune cells and cancer stem-like cells (Yu et al., 2014). Therefore, STAT3 is a promising therapeutic target for cancer drug discovery (Yuan et al., 2015). As shown in Figure 1, STAT3 is activated through the binding of cytokines or growth factors to cell surface receptors. Cytokines, such as the interleukins IL-6, IL-10, and IL-11, as well as growth factors such as EGF, fibroblast growth factor (FGF), and vascular endothelial growth factor (VEGF), can activate the tyrosine phosphorylation cascade. Once ligands bind to their corresponding receptors, the receptors form a dimer complex. The ligand receptors recruit janus kinases (JAKs). JAKs lead to their activation via phosphorylation which in turn phosphorylates the cytoplasmic tyrosine residues on the receptors that serve as a dock for the SH2 domain of STAT3. STAT3 becomes activated (pSTAT3) through the phosphorylation of its Tyr705 residue located within its Src Homology 2 (SH2) domain. The activation of STAT3 triggers pSTAT3 to form a dimer via the interaction of the pTyr705 of one monomer and the SH2 domain of another. The activated dimer dissociates from the receptor and subsequently translocates from the cytoplasm to the nucleus where it binds specific DNA sequences and induces transcription. In addition to JAKs, STAT3 can be activated by nonreceptor tyrosine kinases such as Src and ABL. Activated STAT3 is negatively regulated by phosphatase, FIGURE 2 I Targets for STAT3 modulation. There are different approaches to inhibiting STAT3 such as Src homology region 2 domainsignaling, including by the direct interactions of small molecules with the proteins and the inhibition containing phosphatase-1 (SHP-1) and of upstream kinases that activate STAT3. STAT3 activity can also be modulated through negative regulators, such as phosphatases. SHP-2, and suppressors of cytokine bdjn.org Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 3 Natural products targeting STAT3 signaling pathways in cancer cells pathways have been employed to disrupt STAT3 signaling and activity. We briefly discuss selected inhibitors that target the STAT3-SH2 domain, the upstream kinase of STAT3, and phosphatase, which induces dephosphorylation of STAT3. STAT3 inhibitors modulating STAT3 activity via indirect pathways such as the induction of ROS are also discussed. DIRECTLY TARGETING STAT3 PROTEIN STAT3 exists in the cytoplasm as a monomer. STAT3 molecules form homodimers after phosphorylation of the Try-705 residue that translocates into the nucleus. These activated STAT3 dimers regulate the expression of cell cycle-regulatory proteins and genes associated with cell survival. Therefore, the best approach may be to inhibit STAT3 protein directly. The three domains of STAT3 (Levy and Darnell Jr, 2002), including the NH2-terminal, DNA-binding and SH2 domains, were identified as selective targets for the development of STAT3 inhibitors based on the understanding of the structure and function of STAT3. The most popular target is the SH2 domain because it is necessary for the recruitment of STAT3 to any activated receptor. In addition, inhibition of this target can block STAT3 dimerization, and consequently inhibit nuclear translocation that leads to the regulation of the expression of STAT3 target genes. Structurebased drug design and computational docking have been used to discover small-molecule inhibitors and many synthetic inhibitors of the STAT3 SH2 domain have been reported; some SH2 domain inhibitors are under clinical study (Zhao et al., 2016). Many natural products have been introduced as STAT3 inhibitors via blocking SH2 binding, inhibiting upst-ream kinases, or expressing phosphatase (Yang et al., 2013). In this section, we describe the STAT3 SH2 inhibi tors isolated from herbal medicines. Multiple novel small molecule inhibitors targeting the STAT3 SH2 domain have been identified from natural products (Figure 3). Cryptotanshinone (CTS), which is the one of the major representative components isolated from the root of Salvia FIGURE 3 I Structure of STAT3-SH2 domain inhibitors. 4 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign miltiorrhiza Bunge (Danshen), has been reported to have various pharmacological effects. An antitumor activity of CTS has also been reported in several studies (Chen et al., 2013). Cryptotanshinone and tanshinone IIA are well-known active components of Danshen and they have anti-tumor activities (Akaberi et al., 2015). It was also reported that tanshinone IIA induced apoptosis in human hepatocellular carcinoma cells. CTS was identified as a potent STAT3 inhibitor. CTS inhibited STAT3 Tyr705 phosphorylation in DU145 prostate cancer cells as well as the growth of the cells. CTS down-regulates the expression of STAT3 downstream target proteins such as cyclin D1, survivin, and Bcl-xL in DU145 cells. CTS only inhibited STAT3 Tyr705 phosphorylation and the activity of STAT3 upstream kinases was not affected (Shin et al., 2009). Computational modeling showed that CTS can bind to the STAT3 SH2 domain. Modeling also showed that cryptotanshinone bound to the SH2 domain via a number of hydrogen bonds with nearby residues, including Arg609 and Ile634 (Shin et al., 2009). The computationally proposed mechanism of CTS was confirmed by localization experiments conducted with confocal microscopy, suggesting that CTS bound to STAT3 molecules directly. When the cells were treated with CTS, most STAT3 was localized in the cytoplasm, as was CTS. This co-localization of CTS with STAT3 implied that CTS binds directly to STAT3. Because STAT3 Tyr705 phosphorylation is related to STAT3 dimerization, whether CTS inhibited STAT3 dimerization was assessed using native PAGE. CTS inhibited STAT3 dimerization in a time-dependent manner. Six hours after CTS treatment, dimerized STAT3 was significantly decreased. It was further confirmed that CTS decreased the amount of STAT3 dimers using an electrophoretic mobility shift assay (EMSA). The results of the experiment indicated that the DNA-binding activity of STAT3 decreased after a 3-hour treatment and was almost completely lost after a 6-hour treatment. These data suggest that CTS binds to STAT3 monomers, thereby blocking dimerization, and inhibiting STAT3 transcriptional regulatory activity. Therefore, the antitumor effects of CTS might be mediated by blocking STAT dimerization through the inhibition of the STAT3 SH2 domain and then inhibiting the expression of STAT3-target genes such as Bcl-xL, survivin, and cyclin D1. The down-regulation of these proteins inhibited cell cycle progression and led to the inhibition of cell growth. The down-regulation of cyclin D1 expression by CTS was induced by the arrest in the G1 phase of the cells. Additionally, the suppression of Bcl-xL expression by CTS appeared to induce cell death. Based on these data, CTS is a promising molecule for the development of STAT3 SH2 domain inhibitors. Alantolactone, a major component of Inula helenium, has been reported to possess various pharmacological activities, such as anti-inflammation and antitumor activities (Rasul et al., 2013). Alantolactone inhibited STAT3 phosphorylation at tyrosine 705 in a dose- and time dependent manner; however, this compound did not inhibit STAT3 phosphorylation at serine 727 in MDA- bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon MB-231 cells. The total STAT3 protein levels were not affected by alantolactone. Because STAT3 phosphorylation at tyrosine 705 leads to STAT3 dimerization and then nuclear translocation, consistent with the inhibition of STAT3 phosphorylation, alantolactone inhibited STAT3 translocation to the nucleus and also blocked the DNA-binding activity of STAT3. The STAT3 inhibition activity of alantolactone was tested in several breast cancer cell lines, including MDA-MB-231, MCF-10A, and MCF7. The results demonstrated that alantolactone significantly inhibited STAT3 phosphorylation in MDA-MB-231 cells, which are STAT3 activated tumor cells. As previously mentioned, the STAT family includes STAT1, STAT3, STAT5, and STAT6, and alantolactone did not inhibit the phosphorylation of STAT1 and only slightly inhibited STAT5 and STAT6 phosphorylation; however, STAT3 phosphorylation was almost completely inhibited in alantolactone-treated breast cancer cells, suggesting that alantolactone selectively inhibits STAT3 activation. Upstream kinases, such as JAK1/2, AKT, and EGFR, were not inhibited by alantolactone. Therefore, a structure-based molecular-docking study was performed and it was found that alantolactone could interact with the STAT3 SH2 domain. (Chun et al., 2015). The result suggests that alantolactone suppressed the inducible and constitutively activated STAT3 and blocked the nuclear translocation and the DNA-binding of STAT3 in MDAMB-231 cells by the inhibition of the STAT3 SH2 domain. It was reported that alantolactone also inhibited migration, invasion, adhesion, and colony formation in MDA-MB-231 cells and the compound significantly suppressed the growth of human breast cancer cell xenograft tumors. These results support the potential of alantolactone as a STAT3 inhibitor as well as a potential agent against triple-negative breast cancers. Scoparone (6,7-dimethoxycoumarin) is a coumarin-type natural product. More than 300 coumarins have been isolated from plants (Hoult and Paya, 1996). Scoparone is isolated from the Chinese herb Artemisia capillaris (yin chin) and has been used for the treatment of neonatal jaundice in Asia. Scoparone possesses several other biological properties, including anticoagulant, hypolipidemic, vasorelaxant, anti-oxidant, and antiinflammatory effects (Hung and Kuo, 2013). The inhibition of the transcriptional activity of the nuclear factor kappa-light-chainenhancer of activated B cells (NF-κB) appears to be responsible for its anti-inflammatory activity. It was recently reported that scoparone inhibited the growth of DU145 prostate-cancer cells by the inhibition of STAT3 activity (Kim et al., 2013). Scoparone repressed both constitutively and IL-6-induced activity of STAT3. Consistent with this finding, scoparone decreased the nuclear accumulation of STAT3, but did not reduce the phosphorylation of JAK2 or Src, the major upstream kinases responsible for STAT3 activation. Therefore, scoparone suppressed the transcription of STAT3 target genes such as cyclin D1, c-Myc, survivin, Bcl-2, and Socs3, which was confirmed by Western blot and quantitative real-time PCR analyses. Because scoparone down- bdjn.org regulates the STAT3-target genes, scoparone inhibits the growth of DU145 cells and induces apoptosis. Furthermore, scoparone treatment suppressed anchorage-independent growth in soft agar as well as the tumor growth of DU145 xenografts in nude mice, concomitant with a reduction in STAT3 phosphorylation. Because scoparone inhibits STAT3 activity and does not affect the activity of JAK2 or Src, it was postulated that scoparone might regulate STAT3 activity by directly binding to STAT3 through the SH2 domain, which was confirmed by a structurebased molecular modeling and docking study. The refined model suggested that two oxygen atoms within the lactone ring of scoparone might form hydrogen bonds with amino acid residues within the SH2 domain of STAT3 (Kim et al., 2013). These results suggest that scoparone inhibits the phosphorylation of STAT3 by binding to the SH2 domain of STAT3. Furthermore scoparone suppressed the growth of STAT3-actavted DU145 cells more potently than that of STAT3-negative PC-3 cells. Therefore, these results suggest that STAT3 is a novel molecular target for scoparone and that scoparone represents an anticancer drug candidate for the treatment of cancers in which the STAT3 signaling pathways are constitutively active tumor. A variety of STAT3 inhibitors block STAT phosphorylation, dimerization, and translocation to the nucleus, and inhibit STAT3 target gene expression. In this section, we discuss a few SH2 domain binders isolated from herbal medicines. To date, many natural products that target the SH2 domain and inhibit STAT3 dimerization have been reported, including cardamonin (2,4-dihydroxy-6-methoxychalcone), a class of chalcone compounds isolated from the spicy plant Alpinia conchigera Griff (Zingiberaceae); garcinol and polyisoprenylated benzophenone, isolated from the dried rind of the fruit Garcinia indica, plumbagin (5-hydroxy-2-methyl-1,4-naphthoquinone) isolated from the root of the plant Plumbago zeylanica; withaferin A originally isolated from the plant Withania somnifera; and tectochrysin (flavonoid compound) isolated from A. oxyphylla Miquel (Kamran et al., 2013). These compounds selectively inhibited STAT3 activity and suppressed the expression of STAT3 target genes such as cyclin D1, survivin, Bcl2, VEGF, and Bcl-xL, which led to inhibited proliferation and angiogenesis, and induced apoptosis in a variety of tumor cells. MODULATION OF STAT3 ACTIVITY BY KINASE INHIBITION The phosphorylation of JAK via the activation of cytokine receptors activates the cytoplasmic STAT proteins via phosphorylation. Additionally, the activation of STATs is also modulated by several growth factors such as EGF, HGF, and PDGF. The activated STATs form a dimer, which shifts to the nucleus, binds to specific DNA response elements, and regulates the genes related to cell growth, differentiation, and angiogenesis. The activation of the JAK2/STAT3 signaling pathway in human tumor tissues was recently confirmed, which therefore may provide a novel therapeutic strategy for human Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 5 Natural products targeting STAT3 signaling pathways in cancer cells cancer. In this section, we discuss natural products that inhibit STAT3 activity by the inhibition of JAK2 kinase (Yuan et al, 2015). Most JAK2/STAT3 pathway inhibitors isolated from herbal medicines inhibit STAT3 phosphorylation by the modulation of multiple kinases. In this review, we discuss JAK2/STAT3 pathway inhibitors that target JAK2 such as brevilin A and thymoquinone, as well as JAK2/STAT3 pathway inhibitors, such as cucurbitacin analogues and 4-methoxydalbergione (Figure 4). Brevilin A is a pseudoguaiane sesquiterpene isolated from Litsea glutinosa. Brevilin A was identified as a STAT3 pathway inhibitor through the cell-based luciferase assay against 1,440 natural products. The human prostate carcinoma DU145 and breast cancer MDA-MB-468 cell lines are well known tumor cells with constitutively activated STAT3 (Chen et al., 2013). Brevilin A inhibited STAT3 activity in a dose- and time-dependent manner in both DU145 and MDA-MB-468. Brevilin A also blocks JAK-JH1 tyrosine kinase activity in vitro. To test JAK-specific inhibition, the phosphorylation levels of p65, AKT and glycogen synthase kinase 3 beta (GSK-3β) were analyzed in brevilin A-treated tumor cells. Interestingly, brevilin A did not inhibit the phosphorylation of these proteins, indicating that brevilin A is a specific JAK kinas inhibitor. Brevilin A repressed both constitutive and IL6-induced phosphorylation of STAT3 at Try705 in a time- and dose-dependent manner. The inhibition of STAT3 activity blocks the translocation of STAT3 into the nucleus and then the downregulated expression of STAT3 target genes, e.g., c-Myc and cyclin D1. After the DU145 and MDA-MB-468 cells were treated with brevilin A for 24 and 48 hours, both c-Myc and cyclin D1 expression levels were reduced. Increased cleaved PARP was also observed, indicating that brevilin A induced DU145 and MDA-MB-468 apoptosis after 24 hours of treatment (Chen et al., 2013). These results are consistent with reports indicating that blocking STAT3 activity led to cell growth inhibition in DU145 (Shin et al., 2009) and MD-AMB-468 cells (Siddiquee et al., 2007). Furthermore brevilin A has exhibited preferential cell growth inhibition of DU145 and MDA-MB-468 cells, the growth of which is dependent on STAT3 signaling, over that of human nontransformed telomerase-immortalized fibroblasts BJ cells. JAKs family members contain tyrosine Janus homology domain 1 (JH1 domain), which is the tyrosine kinase domain and usually exhibits constitutive enzymatic activity. Because brevilin A inhibits JAK2 phosphorylation in tumor cells, to find the molecular target of brevilin A, JAK2-JH1 was over-expressed in HEK239T cells, and it was observed that STAT3 Tyr705 phosphorylation was increased. Brevilin A exhibited a significant inhibition of pSTAT3, induced by the overexpression of the JAK2JH1 domain. Interestingly, the compound did not inhibit pSTAT3 at Tyr705 in c-Src-overexpressed HEK293T cells compared with a known Src inhibitor, PD-18097. In addition, brevilin A did not inhibit phosphorylation in HEK293T cells with overexpressed JAK1-JH1 or JAK3-JH1 domains. Based on these results, brevilin A modulates the JAK2/STAT3 signal pathway via the inhibition of 6 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign FIGURE 4 I Structure of JAK2-STAT3 inhibitors. JAK2 kinase activity. Thus, these findings provide evidence that brevilin A is a selective JAK/STAT inhibitor and may act as a potential drug for targeting diseases caused by JAK/STAT abnormalities. Cucurbitacin B is one of several naturally occurring cucurbitacins that constitute a group of oxygenated triterpenes, which are characterized by the tetracyclic cucurbitacin nucleus skeleton and are present in many plants, such as β-glucosides (Chen et al., 2005). Cucurbitacins are essential herbs for a large number of traditional Chinese medicines and the bitter principles of Cucurbitaceae. Cucurbitacin B, a member of the cucurbitacins, is isolated from Trichosanthes kirilowii Maximowicz (Cucurbitaceae family). This active ingredient has antiproliferative effects on several types of malignancies and is known to be a dual inhibitor of the activation of both JAK2 and STAT3 in some malignancies (Thoennissen et al., 2009). JSI-124, a cucurbitacin analogue, blocked the activation of STAT3 in several human cancer cell lines that contain high levels of constitutively activated tyrosine-phosphorylated STAT3 such as A549, MDA-MB-468, and MDA-MB-231 cells. Subsequently the compound inhibited STAT3 DNA-binding activity and STAT3dependent gene expression. JSI-124 also decreased the levels of tyrosine-phosphorylated JAK2, but it did not decrease those of Src. JSI-124 was a highly selective JAK/STAT3 inhibitor because it did not inhibit other tumor survival pathways, such as those mediated by AKT, pERK1/2, or JNK (Blaskovich et al., 2003). Because JSI-124 suppresses the cellular levels of pSTAT3 and pJAK2 but not those of pERK1/2, pJNK, and pAKT, it is suggested that a STAT3 tyrosine kinase is a possible molecular target for JSI-124. Consistent with a direct inhibition of the enzymatic activity of a tyrosine kinase is the fact that the suppression of the STAT3 phosphotyrosine levels was rapid (i.e., was observed as early as 30 min and was complete after only bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon 2 hours of treatment). There are two well-characterized STAT3 tyrosine kinases, JAK and Src kinase. JSI-124 did not inhibit the kinase activities of JAK1, JAK2, and Src in an in vitro kinase assay as confirmed by in vitro kinase assays in which JAK2 and JAK1 enzymatic activities were inhibited by AG490, a known JAK inhibitor, but were not inhibited by JSI-124. Similarly, Src kinase activity was inhibited in vitro by the known Src kinase inhibitor PD180970 but was not inhibited by JSI-124, indicating that Src kinase is not a target. Although STAT3 is not known as a direct target of JSI-124, the ability of JSI-124 to increase mouse survival, to inhibit the growth of human and murine tumors and oncogene- transformed NIH 3T3 tumors in mice with high levels of constitutively activated STAT3 and to not inhibit the growth of those tumors with low levels of activated STAT3 further validates the interference of STAT3 signaling as a sound approach to cancer chemotherapy; therefore, JSI-124 is a good candidate for developing tumor therapeutics targeting JAK/STAT-activated tumors. 4-Methoxydalbergione (4-MD) is isolated from the dried heartwoods of Dalbergia odorifera and Dalbergia odorifera is mainly distributed in China; its heartwood is used for treating blood disorders, ischemia, swelling, necrosis, and rheumatic pain in China and Korea (Beldjoudi et al., 2003). In addition, it has been reported that 4-methoxydalbergione has anti-proliferative and apoptotic effects in human osteosarcoma cells. The inhibitory effects of 4-MD were compared between aggressively growing osteosarcoma (MG63) and mildly growing osteosarcoma (U-2-OS) cells (Park et al., 2016). 4-MD exhibited significant inhibition effects in a concentration-dependent manner in both MG63 and U-2-OS cells. However, 4-MD inhibited cell growth more strongly in MG63 cells than in U-2- OS cells. To determine whether the 4-MD-induced growth inhibition of osteosarcoma cells was associated with the induction of apoptosis, the cells were treated with 4-MD and were assessed using two apoptosis assays, the Annexin V-FITC and TUNEL assays. 4-MD inhibited the constitutive phosphorylation of JAK2 and STAT3 in a doseand time-dependent manner, and 30 μM MD completely inhibited pSTAT3 and pJAK2. In a time-course study, it was found that the inhibition of JAK2 phosphorylation occurred within 1 hour after the treatment, whereas the inhibition of STAT3 phosphorylation occurred at 3 hours after the treatment. These results suggest that the inhibition of STAT3 Tyr705 phosphorylation by 4-MD is caused by inhibited JAK2 activity. Because the nuclear translocation of STAT3 is mandatory for its oncogenic functions, whether 4-MD could suppress the nuclear translocation of STAT3 was determined. 4-MD inhibited the translocation of STAT3 to the nucleus in osteosarcoma cells. 4-MD also down-regulated the expression of STAT3 target proteins such as Bcl-2, Bcl- xL, and survivin. These results indicate that JAK2/STAT3 signaling is involved in the anti-osteosarcoma effect of 4-MD and that 4-MD induced apoptosis by the down-regulation of anti-apoptotic proteins. The in vivo antitumor activity of 4-MD was assessed bdjn.org in a xenograft mouse model; 4-MD significantly decreased tumor weight by 22.25±11.46% compared with that of the control. To determine whether growth inhibition and apoptosis were responsible for the observed antitumor activity of 4-MD, immunohistochemistry for proliferation, STAT3 signaling, and anti-apoptotic proteins were used to assess xenograft tissue. The results showed that 4-MD effectively suppressed the expression of anti-apoptotic molecule (e.g., survivin) and therapeutic target molecules (e.g., pSTAT3) in tumor tissues. Many natural products, including curcumin, inhibit the JAK/ STAT pathways by modulating the activity of multiple-kinases. Mitogen-activated protein kinases (MAPK) signaling cascades, including ERK1/2, c-Jun N-terminal kinase (JNK) and p38 MAPK as well as cAMP response element-binding protein (CREB), are also important for the activation of STAT3 (Steelman, et al., 2004). Thus, the inhibitory activity of 4-MD against a variety of kinases was examined. In 4-MD-treated tumor cells, the activity of ERK1/2, JNK and p38 MAPK, as well as that of CREB was significantly reduced in a dose-dependent manner. To determine the downstream consequences of these effects on MAPK and CREB, the effects of 4-MD on the expression of PTEN (phosphatase and tensin homolog deleted on chromosome ten) were examined. The results showed a concentrationdependent increase in PTEN in osteosarcoma cells treated with 4-MD compared with that in the untreated cells. These results suggested that 4-MD can regulate the activity of CREB and MAPK, consequently increasing the expression of PTEN and inhibiting JAK2, leading to STAT3 inactivation and osteosarcoma cell growth. Thymoquinone (TQ), a compound isolated from black seed oil (Nigella sativa), has been reported to possess anti-inflammatory and anticancer activities (Amin and Hosseinzadeh, 2016). TQ inhibited the constitutive phosphorylation, nuclear localization and reporter gene activity of STAT3 (Kundu et al., 2014). TQ down-regulated the expression of STAT3 target gene products, such as survivin, c-Myc, cyclin-D1, and cyclin-D2, and enhanced the expression of cell cycle inhibitory proteins p27 and p21. Several upstream kinases, such as EGFR tyrosine kinase, JAK2, and Src, are known to regulate STAT3 activation. The effect of TQ on the constitutive activation of these kinases in HCT116 cells was examined. Treatment with TQ markedly diminished the phosphorylation of EGFR at the tyrosine-1173 residue, as well as the phosphorylation of JAK2, and Src kinase. Although JAK2 can directly phosphorylate STAT3, it has been reported that JAK2 and Src function as upstream kinases to EGFR, which transmits activating signals to STAT3. The pretreatment of cells with AG490, a JAK2 inhibitor, attenuated the phosphorylation of EGFR and STAT3. Moreover, when tumor cells were treated with PP2, a Src kinase inhibitor, the tyrosine phosphorylation of EGFR and STAT3 was completely inhibited in HCT116 cells. The treatment of cells with gefitinib, an EGFR antagonist, also blocked STAT3 phosphorylation without affecting the levels of phosphorylated Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 7 Natural products targeting STAT3 signaling pathways in cancer cells JAK2 and Src. Collectively, these results revealed that TQ induced apoptosis in HCT116 cells by blocking STAT3 signaling via the inhibition of JAK2- and Src-mediated phosphorylation of EGFR tyrosine kinase. It was also found that TQ inhibited both constitutive and IL-6-inducible STAT3 phosphorylation which correlated with the inhibition of c-Src and JAK2 activation in multiple myeloma cells. TQ-treated multiple myeloma cells also exhibited down-regulated expression of STAT3-regulated genes, such as cyclin D1, Bcl-2, Bcl-xL, survivin, Mcl-1 and VEGF. Therefore, TQ induced the accumulation of cells in the sub-G1 phase, inhibited proliferation and induced apoptosis, as indicated by PARP cleavage. TQ is a good example of a STAT3 inhibitor acting through the regulation of upstream kinases activity. INHIBITION OF STAT3 ACTIVITY BY PHOSPHATASES STAT3 phosphorylation is also tightly regulated by dephosphorylation, which is medicated by the STAT3 protein tyrosine phosphatases (PTPs) SHP-1, SHP-2, TC-PTP, and PTPRT (Table 2, Böhmer and Friedrich, 2014). MEG2 also dephosphorylates STAT3 at Tyr705 via a direct interaction (Tremblay, 2013). Most natural products targeting phosphatases regulate the expression of SHP1 and/or SHP-2. In this section, we discuss the modes of action of natural products regulating the expression of SHP-1 and/or SHP-2 (Figure 5). SHP-1 was first identified in hematopoietic cells and is predominantly expressed in hematopoietic and epithelial cells. SHP-1 belongs to a family of non-receptor protein TABLE 2 I PTP family members with relevance for regulation of JAK-STAT signaling Systematic name (also gene name) Common synonyms Subfamily PTPN1 PTP1B Nontransmembrane PTP family 1 435 aa PTP for JAK2 and TYK2, STAT6, and several RTKs; can activate SRC-family kinases; knockout mice are resistant to high-fat diet, and exhibit increased insulin and leptin sensitivity; promotes tumorigenesis in mouse models of ERB2/neu-driven breast cancer PTPN2 TCPTP NT1 415 aa (48 kDa, TCPTP48); 387 (45 kDa, TCPTP45) PTP for several STATs, JAK1, and JAK3, and several RTKs; knockout mice die between 3-5 week of age from anemia and systemic inflammation; inactivating mutations found in ALL patients Hematopoietic (595 aa), epithelial (597 aa), long form (SHP-1L, 624 aa) 2 SH2 domains, negative regulation by SH2-domaincatalytic-domain interaction, C-terminus with regulatory function (lipid binding, pTyr residues) contains NLS; cytoplasmic to nuclear translocation observed; negative regulator of cytokine, growth factors and immunoreceptor signaling (e.g., Epo, CSF-1, BCR); substrates receptors, adaptor proteins, JAKs 593 aa 2 SH2 domains, negative regulation by SH2-domaincatalytic-domain interaction, C-terminus with regulatory function (pTyr residues); largely cytoplasmic; knockout embryonically lethal; positive regulator of Ras signaling downstream of several RTKs and cytokines (e.g., PFGFRβ, IL-6) by dephosphorylation of inhibitory molecules; activating mutations in Noonan Syndrome and JMML 1304 aa Transmembrane molecule with two intracellular PTP domains, membrane-proximal domain carries most of PTP activity; activates SRC-family kinases (e.g., LCK, FYN) by dephosphorylating their C-terminal inhibitory phosphosite; also reported as PTP for JAK1; knockout mice exhibit severe combined immunodeficiency; inactivating mutations in a subset of T-ALL patients PTPN6 PTPN11 PTPRC 8 SHP-1, SH-PTP1, HCP, PTP1C SHP-2, SH-PTP2, PTP1D CD45, LCA NT2 Protein, alternate gene products NT2 Receptorlike PTP family 1/6 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign Regulation, specific biochemical features, and functions bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon FIGURE 5 I Structure of STAT3 inhibitors through SHP-1 and/or SHP-2 expression. tyrosine phosphatases (PTPs); it is involved in hematopoietic signaling processes and has also been reported to function as a tumor suppressor during tumor progression. In contrast to the predominant expression of SHP-1 in hematopoietic cells, SHP-2 is a ubiquitously expressed enzyme that appears to be involved in multiple signaling pathways downstream of a variety of growth factors and cytokines. SHP-1 contains two SH2 domains, a catalytic PTP domain and a C-terminal tail. Notably, its phosphatase activity is highly dependent on its structural variability. For example, the closed-form chemical structure of SHP-1 is assembled by the N-SH2 domain and protrudes into the catalytic domain to directly block the entrance into the active site, and the highly mobile C-SH2 domain is believed to act as an antenna to search for the phosphopeptide activator. As previously mentioned, SHP-1 and SHP-2 have very similar structures and regulate the function of the JAK family of tyrosine kinases via their SH2 domains. It has been demonstrated that the hyperphosphorylation of JAK kinases in cells is frequently associated with a lack of functional SHP-1 or SHP-2, indicating that SHP-1 and SHP-2 are critical negative regulators of the JAK2/STAT3 signaling pathway (Bohmer and Friedrich, 2014). 10-Acetoxychavicol acetate (ACA) is isolated from the rhizomes of the commonly used ethno-medicinal plant Languas galangal (Zingiberaceae). ACA exhibits antioxidant and antiinflammatory activities via the suppression of xanthine oxidase, superoxide anion generation, and inducible nitric oxide synthase expression (Murakami and Ohigashi, 2007). ACA has also been reported to exhibit antitumor activity against multiple myeloma by the inhibition of nuclear factor kappa B (Ito et al., 2005). ACA induces tumor apoptosis by enhancing caspase-3 activity and inhibits the proliferation and angiogenesis of prostate tumor growth. bdjn.org ACA has been shown to suppress the phosphorylation of STAT3 (at tyrosine 705 and serine 727 sites) in a dose- and time-dependent manner, and the maximum inhibition was obtained at 5–10 μM in MDA-MB-231 breast cancer cells that expressed constitutively active STAT3. ACA also suppressed the phosphorylation of JAK2 and Src kinase in MDA-MB-231 cells in a dose-dependent manner. ACA suppressed IL-6-induced STAT3 phosphorylation in a concentration dependent manner. These results indicate that ACA has the ability to block constitutive and induced STAT3 activation in cancer cells. After being activated by IL-6, activated STAT3 forms dimers via a phosphotyrosineSH2 domain interaction, and the dimer then translocates into the nucleus. IL-6 induced the accumulation of STAT3 in the nucleus, whereas pretreatment with ACA markedly decreased the accumulation of STAT3 in the nucleus. The translocation of STAT3 dimers to the nucleus results in specific DNA binding to the promoters of target genes and thereby induces the expression of target genes. There are various target genes downstream of STAT3 that participate in the induction of antiapoptotic programs (e.g., survivin, Bcl-xL, and Mcl-1), protooncogene stabilization (e.g., c-Myc and cyclin D1), angiogenesis (e.g., VEGF and hypoxia inducible factor), and invasiveness (e.g., MMP-2 and MMP-9). The analysis of the target gene products of STAT3 revealed that ACA preferentially down-regulated the expression of matrix metalloproteinase such as MMP-2 and MMP-9. In contrast, the expression levels of anti-apoptotic and proliferative genes, such as cyclin D1, survivin, and Bcl2, were only slightly affected. As confirmed by zymography, the enzymatic activity of MMP-9 was also decreased by ACA; in addition, ChIP assay results indicated that the direct interaction between STAT3 and MMP-2/MMP-9 was significantly reduced by ACA. Together, these results suggest that ACA specifically represses the activities of the STAT3 downstream targets MMP- Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 9 Natural products targeting STAT3 signaling pathways in cancer cells 2 and MMP-9 and inhibits cancer cell migration and metastasis (Wang et al., 2014). Because ACA inhibited the phosphorylation of STAT3 (at tyrosine 705 and serine 727 sites), phosphorylation of JAK2 (at tyrosine 1,007 and 1,008 sites), and Src kinase (at tyrosine 416 site) in MDA-MB-231 cells, phosphatases are the best target molecule candidate of ACA for the inhibition of JAK2/STAT3 pathway. When sodium vanadate, a broad-acting tyrosine phosphatase inhibitor, was treated with ACA in breast cancer cells, the ACA-induced inhibition of STAT3 was rescued without affecting the basal level of STAT3 phosphorylation. The involvement of tyrosine phosphatases in the process of ACAmediated STAT3 inhibition has also been suggested. Many PTPs have been implicated in STAT3 regulation, including SHP-1, SHP2, TC-PTP, PTEN, PTP-1D, CD45, and PTP-e. Among these, ACA induced the expression of SHP-1 at the transcriptional level in a concentration-dependent manner. Immunoblot analysis further confirmed that ACA markedly up-regulated SHP-1 expression (Wang et al., 2014). The silencing of SHP-1 using siRNA reverses the inhibition of tumor cell migration by ACA and does not affect cancer cell viability. Therefore, SHP-1 is a target molecule for the anti-metastasis effect of ACA. The results showed that ACA regulated STAT3 activity by the expression of SHP-1 and then suppressed tumor metastasis without affecting cancer cell viability. Therefore, ACA is a potential compound for inhibition of tumor metastasis through the modulation of the STAT3 signaling pathway. Guggulsterone (GS), isolated from Commiphora mukul and used to treat obesity, diabetes, hyperlipidemia, atherosclerosis, and osteoarthritis, has recently been shown to antagonize the farnesoid X receptor and decrease the expression of bile acid– activated genes. GS also inhibits the proliferation and induces the apoptosis of a wide variety of human tumor cell types, including leukemia, head and neck carcinoma, multiple myeloma and melanoma, and breast, lung prostate and ovarian carcinoma (Almazari and Surh, 2013). GS suppressed STAT3 activation in U266 cells, and the inhibition was time-dependent with maximum inhibition occurring at approximately 4 hours and without effects on the expression of STAT3 protein. GS also inhibited IL-6induced pSTAT3. GS induced the dephosphorylation of STAT3, which was gradually reversed by washing GS out of the cells. The reversal was complete within 16 hours, and STAT3 protein levels did not change. As a known STAT3 inhibitor, GS also abrogates the translocation of STAT3 to the nucleus, the DNA-binding ability of STAT3, and the STAT3-dependent expression of genes such as cyclin D1, Bcl-2, Mcl-1, and VEGF. Because JAK2 is one of the main kinases involved, GS-treated whole-cell lysates were immunoprecipitated with anti-JAK2 antibodies and were then subjected to an immunocomplex kinase assay using GST-JAK2 as a substrate. It was found that GS suppressed the constitutive phosphorylation of JAK2 in a time-dependent manner. GS also suppressed the constitutive phosphorylation of c-Src kinase (Ahn 10 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign et al., 2008). However, the inhibition of STAT3 Tyr705 was not due to an inhibition of JAK2 activity because the inhibition of JAK2 phosphorylation occurred after the dephosphorylation of STAT3 (Ahn et al., 2008). These results suggest that the inhibition of STAT3 Tyr705 phosphorylation by GS is not caused by upstream kinases. Treatment of U266 cells with the broad tyrosine phosphatase inhibitor sodium pervanadate reversed the GS-induced inhibition of STAT3 phosphorylation. Tyrosine phosphatases might be involved in the GS-induced inhibition of STAT3 activation because protein tyrosine phosphatases have been involved in the modulation of STAT3 activity, and a PTP inhibitor rescued the GS-induced inhibition of STAT3 phosphorylation. Therefore, whether GS can modulate the expression of SHP-1 in U266 cells were examined. GS induced the expression of SHP-1 protein in a time-dependent manner, with maximum expression occurring after 120 to 240 minutes. In conclusion, GS induced the apoptosis of human MM cells by the inhibition of STAT3, which occurred by the expression of SHP-1. 5-Hydroxy-2-methyl-1,4-naphthoquinone, also known as plumbagin, an analogue of vitamin K3, is isolated from the Plumbaginaceae, Droseraceae, Ancestrocladaceae, and Dioncophyllaceae families. The root of chitrak (Plumbago zeylanica), a major source of plumbagin, has been used in Indian medicine as an antiatherogenic, cardiotonic, hepatoprotective, and neuroprotective agent. The active principle, plumbagin, is also isolated along with a series of structurally related naphthoquinones from the roots, leaves, bark, and wood of Juglans regia (also known as the English walnut, Persian walnut, and California walnut), Juglans cinerea (butternut and white walnut), and Juglans nigra (Padbye et al., 2012). Plumbagin has been shown to exert anticancer effects against a wide variety of tumor cells, including breast cancer, lung cancer, ovarian cancer, acute promyelocytic leukemia, melanoma, and prostate cancer (Padbye et al., 2012). Plumbagin inhibited the constitutive phosphorylation of STAT3 in U266 cells, with maximum inhibition occurring at 5 μM (Sandur et al., 2010). The inhibition was time dependent, with maximum inhibition occurring at 4 hours. Whether plumbagin modulates the phosphorylation of STAT3 at the serine 727 residue was also examined. Interestingly, plumbagin also inhibited the serine phosphorylation of STAT3 in a dose-dependent manner. Plumbagin had no effect on the expression of STAT3 protein under these conditions. EMSA assays of plumbagin-treated U266 cells showed that the treatment decreased STAT3 DNA-binding activity in a doseand time-dependent manner. These results suggests that this compound blocks the dimerization of STAT3 and inhibits its translocation to the nucleus. Therefore, plumbagin downregulated the expression of STAT3 target genes such as antiapoptotic proteins Bcl-xL, cell cycle regulator protein, cyclin D1, and VEGF. Plumbagin suppressed the constitutive phosphorylation of bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon JAK1 at tyrosine 1022/1023 and JAK2 at tyrosine 1007/1008. The expression levels of JAK1 and JAK2 remained unchanged cell types, including oral squamous carcinoma, colon carcinoma, lung carcinoma, pancreatic adenocarcinoma, breast carcinoma, under the same conditions. The relative activity of plumbagin compared with that of AG490 (JAK2 inhibitor) in the suppression of STAT3 phosphorylation was examined. At 5 μM, plumbagin was found to be more effective than AG490 was at 100 μM. By immune complex kinase assays, plumbagin was found to affect JAK2 activity in U266 cells, which indicated that plumbagin also suppressed the activity of JAK2. To monitor the effect of PTP on STAT3 tyrosine phosphorylation, sodium pervanadate (a broad-acting PTP inhibitor) and plumbagin were treated in U266 cells. The PTP inhibitor prevented the plumbagin-induced inhibition of STAT3 activation, suggesting that PTPs are involved in the plumbagin-induced inhibition of STAT3 activation. SHP-1 is an important negative regulator of JAK/STAT signaling in leukemias and lymphomas. Whether plumbagin can modulate the expression of SHP-1 in U266 cells was examined. Cells were incubated with different concentrations of plumbagin for 4 hours; whole-cell extracts were prepared and examined for SHP-1 protein using Western blot analysis. Plumbagin induced the expression of SHP-1 protein in U266 cells. The suppression of SHP-1 expression by small interfering RNA (siRNA) would abrogate the inhibitory effect of plumbagin on STAT3 activation; plumbagin failed to suppress STAT-3 activation in cells treated with SHP-1 siRNA. These siRNA results corroborate the previous evidence of the critical role of SHP-1 in the suppression of STAT-3 phosphorylation by plumbagin (Sandur et al., 2010). Because plumbagin down-regulated the expression of cyclin D1, the gene critical for cell proliferation and the cell cycle, it inhibited cell proliferation of of a dose-dependent manner and caused a significant accumulation of the cell population in the sub-G1 phase. The treatment of U266 cells with plumbagin induced a caspase-3-dependent cleavage of a 118-kDa PARP protein into an 87-kDa fragment. In addition, the overexpression of constitutively active STAT3 can rescue plumbagin-induced apoptosis. Antitumor activity of vitamin K3 analogues has been reported; juglone and 1,4-naphthoquinone can modulate NF-κB and leukemia, by inducing apoptosis and exhibiting significant chemopreventive effects (Joo and Jetten, 2010). Lee et al reported that FOH suppressed both constitutive and inducible STAT3 activation in MM. FOH specifically blocked STAT3 phosphorylation at Tyr705 but had no effect on STAT3 phosphorylation at Ser-727 (Lee et al., 2015). It is well known that JAK1/2 and c-Src have also been implicated in STAT3 activation. FOH inhibited the activation of constitutively active JAK1/2 and c-Src activation in MM cells. It was reported that pervanadatetreatment in U266 cells reversed STAT3 dephosphorylation induced by FOH, indicating that the PTP plays a major role in the dephosphorylation of STAT3. Interestingly, FOH induced SHP2 protein expression that correlated with a down-regulation in STAT3 phosphorylation in MM cells, and transfection with SHP2 siRNA reversed the observed STAT3 inhibitory effect of FOH. The expression of several STAT3-target gene products such as cyclin D1, Bcl-2, Bcl-2, and survivin was suppressed by FOH. The inhibition of cyclin D1 expression by FOH correlated with suppressed proliferation and the accumulation of cells in the subG1 phase of the cell cycle. In addition, down-regulation of the expression of Bcl-2 and Bcl-xl could contribute to the ability of FOH to induce apoptosis in MM cells. It was observed that FOH induced PARP cleavage by the activation of caspase-8, -9, and -3. FOH significantly suppressed MM growth in nude mice, down-regulated the expression of pJAK1/2, p-Src, p-STAT3, and various STAT3-regulated gene products in tumor tissues, and increased the levels of caspase-3 in the FOH treatment groups. Overall, these results demonstrate that FOH inhibits both inducible and constitutive STAT3 activation through the induction of tyrosine phosphatase, which may account for its antitumor effects observed in vivo. FOH is promising for potential applications in the treatment of cancer and other diseases. activation, and unlike plumbagin, shikonin, which is an analogue of vitamin K, is known to inhibit protein tyrosine phosphates. Thus, this finding suggests that not all analogues of vitamin K exhibit anti-tumor activities by the modulation of different targets. In conclusion, plumbagin is a selective inhibitor of the STAT3 pathways and is a useful vitamin K analogue for STAT3 modulation and for developing tumor therapeutics targeting JAKSTAT activated tumors with fewer side effects. Farnesol (FOH) is an isoprenoid, and isoprenoids are one class of phytochemicals that inhibit tumor cell proliferation and differentiation and induce apoptosis (Burke et al., 1997). FOH is used to treat obesity, diabetes, hyperlipidemia, and atherosclerosis. Additionally, FOH exhibits anticancer potential by suppressing the proliferation of a wide variety of human tumor bdjn.org Capillarisin (CPS) is isolated from Artemisia capillaries Thunb (Compositae), which has been used to treat various human diseases including liver cirrhosis, liver cancer, jaundice and cholecystitis in Asian countries, such as Korea, China, and Japan (Han et al., 2013). CPS has been reported to be a potent inhibitor of the NF-κB activation pathway and to have suppressed the phosphorylation of STAT3 at Tyr705, but not Ser727, without an effect on the expression of STAT3 protein in U266 cells (Lee et al., 2014). CPS substantially reduced STAT5 activation in MM cells, without affecting total STAT5 levels, as demonstrated by Western blot analysis. CPS also blocked the IL-6 induced phosphorylation of STAT3 and subsequently suppressed the DNA binding activity of STAT3. EMSA analysis of U266 cells showed that CPS decreased STAT3 DNA-binding activity in a dose- and time-dependent manner. Therefore, CPS suppresses the nuclear translocation of STAT3. CPS inhibits the activation of upstream kinases involved in the STAT3 signaling cascade, such as JAK1/2 and Src kinase, in a dose-dependent manner in MM cells. Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 11 Natural products targeting STAT3 signaling pathways in cancer cells To determine which target molecules are involved in CPSmediated STAT3 inactivation, a broad-acting tyrosine phosphatase inhibitor (sodium pervanadate) was used to treat U266 cells, and it was observed that the CPS-induced inhibition of STAT3 and JAK1/2 phosphorylation was rescued. These data suggest that tyrosine phosphatases are involved in the CPSinduced inhibition of JAK1/2 and STAT3 activation in MM cells. Thus, it was examined whether CPS affects the expression of SHP-1, SHP-2, and PTEN which are nontransmembrane PTPs expressed abundantly in hematopoietic cells. Treatment with CPS led to an increased expression of SHP-1 and SHP2, but not PTEN at the protein level. CPS also substantially enhanced mRNA levels of both SHP-1 and SHP-2 in U266 cells. Transfection with SHP-1/2 siRNAs reversed the inhibition of STAT3 and JAK1 activation by CPS. It was also found that CPS failed to suppress JAK1 and STAT3 activation in cells treated with SHP-1/2 siRNAs, thereby implicating the pivotal role of these two phosphatases in the STAT3/JAK1 inhibitory effects of CPS. The expression of several STAT3-target gene products was suppressed by CPS. These products included proliferative (cyclin D1) and anti-apoptotic (Bcl-2, Bcl-xl, survivin, and IAP-1) gene products. Thus, the down-regulation of the expression of Bclxl could contribute to the ability of CPS to induce apoptosis in MM cells. The down-regulation of cyclin D1 expression by CPS correlated with its significant anti-proliferative effects observed in MM cells. CPS was the first compound found to regulate the expression of SHP-1/2; therefore, there is sufficient rationale for investigating the therapeutic efficacy of STAT3 inhibitors isolated from herbal medicines and CPS is a potential lead molecule for the development of JAK/STAT3 pathway blockers via the expression of SHP-1/2 (Lee et al., 2014). Nimbolide (NL), is a terpenoid limonoid isolated from Azadirachta indica leaves and it has been found to exhibit diverse pharmacological activities such as anti-feedant, anti-malarial, anti-HIV, antimicrobial, and significant anti-cancer responses. The diverse anticancer effects of NL have been reported in multiple tumor types including those of breast, colorectal, brain, and liver cancer (Elumalai and Arunakaran, 2014). Both constitutive pSTAT3 in DU145 and IL-6-stimulated pSTAT3 levels in LNCaP were substantially reduced upon NL treatment. In DU145 cells, both JAK1 and JAK2 phosphorylation levels were reduced by NL, whereas Src phosphorylation was not affected. NL also reduced the phosphorylation of both JAK1 and JAK2 stimulated by IL-6 in LNCaP cells. These data indicate that the STAT3 inhibition caused by NL treatment may be caused by the down-regulation of upstream activation through JAK1 and JAK2 kinases. In addition, NL blocked STAT3 from binding to DNA and thereby reduced the transcription of STAT3 target genes. ROS are maintained at low levels by cellular anti-oxidative systems under normal physiological conditions and this low level of ROS can promote both cell survival and proliferation. However, when ROS accumulate at elevated, non-physiological concentrations, apoptotic cell death can be caused through MODULATION OF STAT3 ACTIVITY THROUGH ROS Reactive oxygen species (ROS) are critical for the metabolic and signal transduction pathways associated with cell growth and apoptosis. Cancer cells exhibit increased ROS levels compared with those of normal cells because of their accelerated metabolism. The high ROS levels in cancer cells renders these cells more susceptible to oxidative stress–induced cell death, which can be exploited for selective cancer therapy (Gorrini et al., 2013). Many known natural compounds, such as hydroxycinnamaldehyde (Han et al., 2004), curcumin, EGCG, piperlogumin (Raj et al., 2011), and resveratrol, generate ROS to destroy cancer cells (Sulliva and Chandel, 2014). It was reported that JAK/STAT-dependent signaling is regulated by redox-related pathways (Duhe, 2013). Recent reports suggest that ROS may inhibit JAK2 phosphorylation via posttranslational modification of its critical cysteine residues as well as STAT3 through PTP activation. In this section, we discuss natural products (Figure 6) that inhibit the JAK2/STAT3 pathway through ROS generation in tumor cells. 12 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign FIGURE 6 I Structure of STAT3 inhibitors that act through ROS modulation. bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon damage of DNA or proteins. N-acetyl-L-cysteine (NAC) and glutathione (GSH) are two well-known thiol-related antioxidants. Interestingly, pre-treatment with these two antioxidants substantially rescued STAT3 inhibition caused by NL treatment, thereby indicating that ROS may be involved in the inhibitory effect of NL on STAT3 in PCa cells. Furthermore, NAC/GSH pretreatment was also found to significantly reduce the cellular apoptosis induced by NL, as observed by flow cytometry and Western blot analysis, further demonstrating that oxidative stress may also contribute to the observed pro-apoptotic effects of NL. It was observed that ROS levels were increased by NL-treatment in DU145 cells. The GSH/GSSG system is one of the major intracellular antioxidant systems. The ratio of GSH to oxidized glutathione (GSSG) is an indicator of cellular oxidative stress. NL significantly decreased GSH and increased the GSSG/GSH ratio in DU145 cells, thereby indicating that the exposure of the cells to NL resulted in an imbalance of the GSH/GSSG system. To further investigate the role of GSH/GSSG imbalance in mediating NL-induced oxidative stress, both a GSH synthesis blocker, buthionine sulfoximine (BSO), and a GSH prodrug, NAC, were employed. It was found that pretreatment with NAC significantly prevented NL-induced ROS production, whereas BSO enhanced ROS production. Therefore, the GSH/GSSG imbalance in mediating NL-induced oxidative stress was caused by the down regulation of GSH. Furthermore, NL increased the level of H2O2 in a time-dependent manner; increased H2O2 production was also prevented by NAC and was enhanced by BSO. These results demonstrate that H2O2 is the major ROS induced by NL in PCa cells. In addition, BSO enhanced NLinduced cellular apoptosis when applied in combination with NL, and the increased apoptosis was also attenuated by NAC pretreatment in PCa cells. Glutathione reductase (GR), an enzyme involved in the reduction of GSSG to GSH, was inhibited by NL in a time-dependent manner in NL-treated cells. These results demonstrate that NL directly suppresses GR and GSH synthesis by inhibiting their activity and increasing ROS in tumor cell. It has also been reported that ROS can directly oxidize the cysteine residues of JAK2 in the catalytic domain, and are therefore able to suppress its kinase activity as well as downstream STAT3 activation (Duhe, 2013). NL-induced oxidative stress inhibited the phosphorylation of JAK2 and STAT3 and also significantly reduced PCa cell viability, supporting the notion that the oxidization of JAK2 may contribute to ROS-mediated STAT3 inhibition. NL also inhibited tumor progression and metastasis in a mouse prostate model. These findings suggest that NL could be a potent anticancer agent against various malignancies related to STAT3 activation. 6-Shogaol (6SG), an active compound in ginger (Zingiber officinale), exhibit anti-proliferative, antimetastatic, and proapoptotic activities. Ginger (Zingiber officinale Roscoe, Zingiberaceae) is one of the most commonly used dietary seasonings in the world (Marx et al., 2015). Ginger has long been bdjn.org used for the treatment of human diseases, such as the common cold, nausea, arthritis, migraines, and hypertension (Semwal et al., 2015). Ginger contains approximately 1.0-3.0% volatile oils and a number of pungent compounds. Gingerols are the most abundant compounds in fresh roots. Shogaols, the dehydrated form of gingerols, are found in only small quantities in the fresh root (Semwal et al., 2015). Previous animal studies have conclusively shown that ginger and its active phenolic constituents such as 6G/6SG can inhibit skin tumor formation and abrogate the initiation as well as progression of colon, liver, and breast cancer. Among various compounds in ginger, 6SG has been intensively studied for their anti-tumor potential in several tumor cell lines (Wang et al., 2014). 6SG inhibited constitutive STAT3 activation in a variety of STAT3-activated human cancer cell lines, such as MDA-MB-231, DU145, SCC4, A549, and HepG2 cells. 6SG clearly inhibited STAT3 activation in both MDA-MB-231 and DU145 cells with 30 μM 6SG compared with different types of tumor cells. 6SG also suppressed the phosphorylation and nuclear translocation of STAT3 in MDA-MB-231 and DU145 cells (Kim et al., 2015). The expression of STAT3 remained unchanged under the same conditions. 6SG also suppressed the phosphorylation of JAK2 in both MDA-MB-231 and DU145 cells. Bcl-2, Bcl-xl, IAP1, and survivin have been implicated in cell survival, and act downstream of the JAK/STAT signaling cascade; 6SG downregulates the expression of anti-apoptotic gene products in a time-dependent manner. Previous studies have reported that ROS induces the MAPK and AKT cascades. Interestingly, 6SG induces the activation of MAPK cascades including JNK, p38 MAPK, and ERK, in MDA-MB-231 cells. 6SG substantially up-regulated the phosphorylation of all three JNK, p38, and ERK MAP kinases in a time-dependent manner. Therefore, the potential role of ROS in 6SG mediated ERK, JNK, and p38 MAPK activation in MDA-MB-231 cells was explored. Whole-cell lysates of MDAMB-231 cells treated with 6SG for 8 hours with or without NAC pretreatment were analyzed for pJNK, p-p38, and pERK. NAC pretreatment drastically prevented the cleavage of PARP in MDA-MB-231 cells treated with 6SG for 24 hours. These results indicate that increasing ROS levels may mediate the effects of 6SG on PARP cleavage and apoptosis in MDA-MB-231 cells. To confirm the generation of ROS by 6SG, MDAMB- 231 cells were treated with 6SG in the presence or absence of the antioxidant NAC and stained with DCFHDA. Treatment with 6SG resulted in a large increase in DCF fluorescence intensity compared that observed for untreated control cells. NAC pretreatment clearly abrogated the observed effect of 6SG on ROS production. Although it is unclear how 6SG generates ROS in cancer cells, collectively, these results indicate that 6SG can inhibit the constitutive STAT3 signaling cascade and induce ROS-mediated JNK, p38 MAPK, and ERK activation. Moreover, 6SG induces apoptosis through the down-modulation of gene products that mediate tumor cell survival and proliferation in human breast cancer. Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 13 Natural products targeting STAT3 signaling pathways in cancer cells Sugiol is an abietane-type diterpene compound and is isolated from the roots of Salvia prionitisHance (Labiatae), Salvia miltiorrhiza Bunge, and Salvia viridis (Liu et al., 1995). Many studies have investigated the biological activities of sugiol, including its modest anti-tumor, anti-inflammatory, antimicrobial, and aldose reductase inhibitory activities (Saijo et al., 2015). Sugiol has also been reported to exhibit very weak antitumor activity against SW620, MDA-MB-231, and NCI-H23 cells (Son et al., 2005). However, sugiol selectively inhibited the growth of DU145 cells, which exhibit constitutively active STAT3, compared with other, STAT3-independent other cells such as LNCap, PC3, and non-tumorigenic MCF10A cells (Jung et al., 2015). These data suggest that sugiol primarily suppresses the proliferation of STAT3-dependent cancer cells, where STAT3-dependent cancer cells are those in which STAT3 is activated and the proliferation rate of the cancer cells is decreased by knockdown of STAT3. Sugiol dramatically decreased STAT3 phosphorylation at Try705, whereas the total amount of STAT3 protein remained nearly unchanged under these conditions in DU145 cells. The phosphorylation of STAT1 and STAT5 was not inhibited by sugiol under these conditions. Sugiol inhibited the expression of STAT3 target genes and the cell cycle regulators cyclin D1, cyclin A, survivin, VEGF, Mcl-1, and Bcl-xL, in DU145 cells. Because sugiol down-regulated cyclin D1 and A, cell cycle arrest at the G1/S stage was induced by the compound in DU145 cells. JAK and Src family proteins are the most well-known upstream kinases that phosphorylate STAT3 tyrosine residues; however, treatment with up to 50 µM sugiol did not affect JAK family proteins, such as JAK2, JAK3, and TYK2. An in vitro kinase assay was also performed for several upstream kinases of STAT3, including EGFR, FGFR, JAK2, JAK3, Lyn, and Src, to examine the effect of sugiol on these tyrosine kinases. At 10 µM, sugiol did not inhibit the activity of these kinases in an in vitro assay, which suggests that the inhibition of STAT3 activation is independent of upstream kinases. Transketolase (TKT) was identified as a target molecule of sugiol by affinity chromatography. TKT activity was decreased by sugiol in a dose-dependent manner in vitro. STAT3 phosphorylation was reduced in DU145 cells transfected with TKT siRNA. TKT siRNA knockdown increases ROS levels because TKT downregulates intracellular GSH levels. Sugiol induced a significant time-dependent decrease in GSH levels, and NAC and GSH treatment prevented sugiol-induced GSH depletion. Therefore, sugiol inhibits TKT activity and then reduces GSH levels, which leads to the generation of ROS in tumor cells. These data suggest that the interaction of sugiol with transketolase induced ROS generation and STAT3 dephosphorylation, which inhibited cell proliferation (Jung et al., 2015). As previously mentioned, ROS induce the MAPK and AKT cascades. When DU145 cells were treated with sugiol, the phosphorylation of ERK and AKT, but not JNK or p38, were increased and then gradually decreased in a time-dependent 14 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign manner. The treatment of cells with ROS scavengers NAC or GSH abrogated sugiol-induced STAT3 dephosphorylation and ERK phosphorylation, but not AKT phosphorylation. This result indicates that sugiol-induced ROS inhibit STAT3 activity and promote ERK activity. It was reported that ERK phosphorylates STAT3 on Ser727 and recruits tyrosine phosphatase (PTPase) or MAP kinase phosphatase (MKP) proteins, which dephosphorylate Tyr705 in cardiomyocytes (Booz et al., 2003). Notably, sugiol increased Ser727 phosphorylation but decreased Tyr705 phosphorylation in DU145 cells, as previously reported and the events were abrogated by pretreatment with the ERK inhibitor U0126. These results suggest that STAT3 was inactivated because of sugiolinduced ERK activation, and this process was induced by the direct interaction between STAT3 and p-ERK. Booz et al. proposed that ERK serves as a scaffolding protein to recruit a phosphatase to STAT3. Pretreatment with the broad-spectrum tyrosine phosphatase inhibitor pervanadate prevented the sugiolinduced inhibition of STAT3 activation, and it was found that the silencing of only MEG2 expression by siRNA abrogated the inhibitory effects of sugiol on STAT3 activation. The MEG2 overexpression inhibited STAT3 phosphorylation, which suggests that MEG2 is a critical phosphatase involved in the suppression of STAT3 phosphorylation by sugiol. Sugiol inhibits TKT activity and induces the dephosphorylation of STAT3 at Tyr-705 through ROS-mediated ERK activation, and it is responsible for the antitumor effect of sugiol in DU145 prostate cells (Jung et al., 2015). Sugiol exhibits a novel mechanism in inhibiting STAT3 activity via ROS and phosphatase. Therefore, sugiol is a good probe with which to elucidate the STAT3 signal pathways. CONCLUDING REMARKS Evidence indicating that STAT3 is an oncogene and an important target for therapy continues to accumulate; therefore, these important findings re-establish STAT3 as a desirable therapeutic target for human cancer treatment. However, despite intensive efforts, STAT3 has remained frustratingly elusive as a target for cancer therapy. Several natural or dietary compounds have been shown to possess in vitro and/or in vivo inhibitory effects against STAT3. However, non-specific effects in targeting STAT3, off-target effects, variable bioavailability and moderate efficacy constitute disadvantages of natural products as effective chemopreventive agents. In this review, we described specific or selective STAT3 inhibitors isolated from herbal medicines including STAT3SH2 domain inhibitors, up-stream kinase inhibitors of STAT3, STAT3 inhibitors by the regulation of PTP, and STAT3 modulators through the increase of ROS levels in cancer cells. A large number of STAT3 inhibitors are currently undergoing clinical trials, which demonstrates that with technological advances and the continuous improvement of all approaches, progress is being made. Most inhibitors described in this review will require further research and development if they are to bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon become the next generation of cancer therapies. Therefore, more STAT3 inhibitors may be isolated from herbal medicines for the development of cancer treatments, with less toxicity and greater activity. ACKNOWLEDGEMENTS This work was supported by the KRIBB Research Initiative Program, the Bio-Synergy Research Project (2012M3A9C404877), the Bio & Medical Technology Development Program (2015M3A9B5030311), the Foreign Plant Extract Library Program (2011-00497), funded by the Ministry of Science, ICT& Future Planning. AUTHOR INFORMATION The Authors declare no potential conflicts of interest. Original Submission: Feb 20, 2016 Revised Version Received: Mar 18, 2016 Accepted: Mar 19, 2016 REFERENCES Akaberi, M., Mehri, S., and Iranshahi, M. (2015). Multiple pro-apoptotic targets of abietane diterpenoids from Salvia species. Fitoterapia 100, 118132. derivatives through enhancing SHP-1 phosphatase activity. Eur J Med Chem 55, 220-227. Chen, W., Yin Lu, Y., Chen, G., and Huang, S. (2013).Molecular evidence of cryptotanshinone for treatment and prevention of human cancer. Anticancer Agents Med Chem 13, 979-987. Chen, X., Du, Y., Nan, J., Zhang, X., Qin, X., Wang, Y., Hou, J., Wang, Q., and Yang J. (2013). Brevilin A, a novel natural product, inhibits janus kinase activity and blocks STAT3 signaling in cancer cells. PLoS ONE 8, e63697. Chen, J.C., Chiu, M.H., Nie, R.L., Cordell, G.A., and Uiu, S.X. (2005). Cucurbitacins and cucurbitane glycosides: structures and biological activities. Nat Prod Rep 22, 386-399. Chinembiri, T.N, du Plessis, L.H., Gerber, M., Hamman, J.H., and du Plessis, J. (2014). Review of natural compounds for potential skin cancer treatment. Molecules 19, 11679-11722 Chun, J., Li, R.J., Cheng, M.S., and Kim, Y.S. (2015). Alantolactone selectively suppresses STAT3 activation and exhibits potent anticancer activity in MDA-MB-231 cells. Cancer Letters 357, 393–403. Cragg, G.M., Grothaus, P.G., and Newman, D.J. (2009). Impact of natural products on developing new anti-cancer agents Chem Rev 109, 30123043. Cragg, G.M., Grothaus, P.G., and Newman, D.J. (2014). New horizons for old drugs and drug leads. J Nat Prod 77, 703-723. Debnath, B.P., Xu, S., and Neamati, N. (2012). Small molecule inhibitors of signal transducer and activator of transcription 3 (Stat3). J Med Chem 55, 6645-6668. Dhanik, A., McMurray, J.S., and Kavraki, L.E. (2012). Binding modes of peptidomimetics designed toiInhibit STAT3. PLoS ONE 7, e51603. Ahn, K.S., Sethi, G., Sung, B., Goel, A., Ralhan, R., and Aggarwal, B.B. (2008). Guggulsterone, a farnesoid X receptor antagonist, inhibits constitutive and inducible STAT3 activation through induction of a protein tyrosine phosphatase SHP-1. Cancer Res 68, 4406-4415. Duhé, R.J. (2013) Redox regulation of Janus kinase. JAK-STAT, 2, e26141. Almazari, J., and Surh, Y.J. (2013). Cancer chemopreventive and therapeutic potential of guggulsterone. Top Curr Chem 329, 35-60. Fouse, S.D. and Costello, J. F. (2013). Cancer stem cells activate STAT3 the EZ Way. Cancer Cell 23, 711-713. Amni, B., and Hosseinzadeh, H. (2016). Black Cumin (Nigella sativa) and its active constituent, thymoquinone: An overview on the analgesic and anti-inflammatory effects. Planta Med 82, 8-16. Gonda, T.J. and Ramsay, R.G. (2015). Directly targeting transcriptional dysregulation in cancer. Nature Rev Cancer 15, 686-694. Arumuggam, N., Bhowmick, N.A., and Rupasinghe, H.P.V. (2015). Phytochemicals targeting JAK/STAT signaling and IDO expression in cancer. Phytother Res 29, 805-817. Assefa, B., Glatzel, G., and Buchmann, C. (2010). Ethnomedicinal uses of Hagenia abyssinica (Bruce) J.F. Gmel. among rural communities of Ethiopia. J Ethnobiol Ethnomed 6, 20. Beldjoudi, N., Mambu, L., Labaïed, M., Grellier, P., Ramanitrahasimbola, D., Rasoanaivo, P., Martin, M.T., and Frappier, F. (2003). Flavonoids from Dalbergia louvelii and their antiplasmodial Activity. J Nat Prod 66, 14471450. Blaskovich, M.A., Sun, J., Cantor, A., Turkson, J., Jove, R., and Sebti, S. M. (2003). Discovery of JSI-124 (Cucurbitacin I), a selective janus kinase/ signal transducer and activator of transcription 3 signaling pathway inhibitor with potent antitumor activity against human and murine cancer cells in mice. Cancer Res 63, 1270-1279. Böhmer, F.D., and Friedrich, K. (2014). Protein tyrosine phosphatases as wardens of STAT signaling. JAK-STAT 3, e28087. Booz, G.W., Day, J.N., and Baker, K.M. (2003).Angiotensin II effects on STAT3 phosphorylation in cardiomyocytes: evidence for Erk-dependent Tyr705 dephosphorylation, Basic Res Cardiol 98, 33-38. Bromberg, J. and Darnell Jr. J.E. (2000). The role of STATs in transcriptional control and their impact on cellular function. Oncogene 19, 2468-2473. Burke, Y.D., Stark, M.J., Roach, S.L., Sen, S.E., and Crowell, P.L. (1997). Inhibition of pancreatic cancer growth by the dietary isoprenoids farnesol and geraniol. Lipids 32,151-156. Chan, K.T., Li, K., Liu, S.L., Chu, K.H., Toh, M., and Xie, W.D. (2010). Cucurbitacin B inhibits STAT3 and the Raf/MEK/ERK pathway in leukemia cell line K562. Cancer Lett 289 (2010) 46–52 Chen, K.F., Tai, W.T., Hsu, C.Y., Huang, J.W., Liu, C.Y., Chen, P.J., Kim, I., and Shiau, C.W. (2012). Blockade of STAT3 activation by sorafenib bdjn.org Elumalai, P., and Arunakaran, J. (2014). Review on molecular and chemopreventive potential of nimbolide in cancer. Genomics Inform 12, 156-164. Gorrini, C., Harris, I.S., and Mak, T.W. (2013). Modulation of oxidative stress as an anticancer strategy. Nature Rev Drug Discov 12, 931-947. Gupta, S.C., Kim, J.H., Prasad, S., and Aggarwal, B.B. (2010). Regulation of survival, proliferation, invasion, angiogenesis, and metastasis of tumor cells through modulation of inflammatory pathways by nutraceuticals. Cancer Metastasis Rev 29, 405-434. Han, D.C., Lee, M.Y., Shin, K.D., Jeon, S.B., Kim, J. M., Son, K.H., Kim, H.C., Kim, H.M., and Kwon, B.M. (2004). 2'-benzoyloxycinnamaldehyde induces apoptosis in human carcinoma via reactive oxygen species. J Biol Chem 279, 6911-6920. Han, J.M., Kim, H.G., Choi, M.K., Lee, J.S., Wang, J.H., Park, H.J., Son, S.W., Qwang, S.Y., and Son, C.G. (2013). Artemisia capillaris extract protects against bile duct ligation-induced liver fibrosis in rats. Exp Toxicol Pathol 65, 837-844. Harris, T.J.R. and McCormick, F. (2010). The molecular pathology of cancer. Nature Rev Clin Oncol 7, 251-265. Hoult, J.R.S. and Paya, M. (1996). Pharmacological and biochemical actions of simple coumarins: Natural products with therapeutic potential. Gen Pharmacol 27, 713-722. Hung H.Y. and Kuo, S.C. (2013). Recent studies and progression of Yin Chen Hao (茵陳蒿 Yīn Chén Hāo), a long-term used traditional chinese medicine. J Tradit Complement Med 3, 2-6. Ito, K., Nakazato, T., Xian, M.J., Yamada, T., Hozumi, N., Murakami, A, Ohigashi, H., Ikeda, Y., and Kizaki, M. (2005). 10-acetoxychavicol acetate is a novel nuclear factor kappa B inhibitor with significant activity against multiple myeloma in vitro and in vivo. Cancer Res 65, 4417-4424. Joo, J.H., Jetten, A.M. (2010). Molecular mechanisms involved in farnesolinduced apoptosis. Cancer Lett 287, 123-135. Jung, S.N., Shin, D.S., Kim, H.N., Jeon, Y.J., Yun, J., Lee, Y.J., Kang, J.S., Han, D. C., and Kwon, B.M. (2015). Sugiol inhibits STAT3 activity via regulation of transketolase and ROS-mediated ERK activation in DU145 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 15 Natural products targeting STAT3 signaling pathways in cancer cells prostate carcinoma cells. Biochem Pharmacol 97, 38-50. Kamran, M.Z., Patil, P., and Gude, R.P. (2013). Role of STAT3 in cancer metastasis and translational advances. BioMed Res Intl 2013, 421821. Kim, J.K., Kim, J.Y., Kim, H.J., Park, K.G., Harris, R.A., Cho, W.J., Lee, J.T., and Lee, I.K. (2013) Scoparone exerts anti-tumor activity against DU145 prostate cancer cells viainhibition of STAT3 activity. PLoS ONE 8, e80391. Kim, S.M., Kim, C., Bae, H., Lee, J.H., Baek, S.H., Nam, D., Chung, W.S., Shim, B.S., Lee, S.G., Kim, S.H., Sethi, G., and Ahn, K.S. (2015). 6-Shogaol exerts anti-proliferative and pro-apoptotic effects through the modulation of STAT3 and MAPKs signaling pathways. Mol Carcinog 54, 1132-1146. Kundu, J., Choi, B.Y., Jeong, C.H., Kundu, J.K, and Chun, K.S. (2014). Thymoquinone induces apoptosis in human colon cancerHCT116 cells through inactivation of STAT3 by blockingJAK2- and Src‑mediated phosphorylation of EGF receptor tyrosine kinase. Oncol Rep 32, 821-828. Lee, J.H., Chiang S.Y., Nam, D., Chung, W.S., Lee, J., Na, Y.S., Sethi, G., and Ahn, K.S. (2014). Capillarisin inhibits constitutive and inducible STAT3 activation through induction of SHP-1 and SHP-2 tyrosine phosphatases. Cancer Lett 345, 140-148. Levy, D.E., Darnell Jr, J.E. (2002). Stats: Transcriptional control and biological impact. Nature Rev Mol Cell Biol 3, 651-662. Siddiquee, K., Zhang, S., Guida, W.C., Blaskovich, M.A., Greedy, B., Lawrence, H.R., Yip, M.L., Jove, R., McLaughlin, M.M., and Lawrence, N.J. (2007). Selective chemical probe inhibitor of Stat3, identified through structure-based virtual screening, induces antitumor activity. Proc Natl Acad Sci USA 104, 7391-7396. Siveen, K.S., Sikka, S., Surana, R., Dai, X., Zhang, J., Kumar, A.P., Tan, B.K.H., Sethi, G., and Bishayee, A. (2014). Targeting the STAT3 signaling pathway in cancer: Role of synthetic and natural inhibitors. Biochim Biophy Acta 1845, 136-154. Son, K.H., Oh, H.M., Choi, S.K., Han, D.C., and Kwon, B.M. (2005). Antitumor abietane diterpenes from the cones of Sequoia sempervirens. Bioorg Med Chem Lett 15, 2019-2021. Spitzner, M., Ebner, R., Wolff, H.A., Ghadimi, B.M., Wienands, J., and Grade, M. (2014). STAT3: A novel molecular mediator of resistance to chemoradiotherapy. Cancers 6, 1986-2011. Steelman, L.S., Pohnert, S.C., Shelton, J.G., Franklin, R.A., Bertrand, F.E., and McCubrey, J.A. (2004). JAK/STAT, Raf/MEK/ ERK, PI3K/Akt and BCRABL in cell cycle progression and leukemogenesis. Leukemia 18,189-218. Sullivan, L.B., and Chandel, N.S. (2014). Mitochondrial reactive oxygen species and cancer. Cancer Metab 2, 17. Liu, J., Zapp, J., and Becker, H. (1995). Comparative phytochemical investigation of Salvia miltiorrhiza and Salvia triloba. Planta Med 61, 453455. Thoennissen, N.H., Iwanski, G.B., Doan, N.B., Okamoto, R., Lin, P., Abbassi, S., Song, J.H., Yin, D., Melvin Toh, M., Xie, W.D. Said, J.W., and Koeffler, P. (2009). Cucurbitacin B induces apoptosis by inhibition of the JAK/STAT pathway and potentiates antiproliferative effects of gemcitabine on pancreatic cancer cells. Cancer Res 69, 5876-5884. Marx, W., McKavanagh, D., McCarthy, A.L., Bird, R., Ried, K., Chan, A., and Isenring, L. (2015). The effect of Ginger (Zingiber officinale) on platelet aggregation: A systematic literature review. Plos One 10, e0141119. Torre, L.A., Siegel, R.L., Ward, E.M., and Jemal, A. (2016). Global cancer incidence and mortality rates and trends-an update. Cancer Epidemiol Biomarkers Prev 25, 16-17. Medema, J.P. (2013). Cancer stem cells: The challenges ahead. Nature Cell Biol 15, 338-344. Tremblay, M.L. (2013). On the role of tyrosine phosphatases as negative regulators of STAT signaling in breast cancers: new findings and future perspectives. Breast Cancer Res 15, 312. Murakami, A., and Hajime Ohigashi, H. (2007). Targeting NOX, INOS and COX-2 in inflammatory cells: Chemoprevention using food phytochemicals. Int J Cancer 121, 2357-2363. Padhye, S., Dandawate, P., Yusufi, M., Ahmad, A., and Sarkar, F.H. (2012). Perspectives on medicinal properties of plumbagin and its analogs. Med Res Rev 32, 1131-1158. Park, K.R., Yun, H.M., Quang, T.H., Oh, H., Lee, D.S., Auh, Q.S., and Kim, E.C. (2016). 4-Methoxydalbergione suppresses growth and induces apoptosis in human osteosarcoma cells in vitro and in vivo xenograft model through down-regulation of the JAK2/STAT3 pathway Oncotarget in press. Padbye,S., Dandawate, P., Yusufi, M., Abmad, A., and Sarkar, F.H. (2010). Perspectives on medicinal properties of plumbagin and its analogs. Med Res Rev 32, 1131-1158. Raj, L., Ide, T., Gurkar, A.U., Foley, M., Schenone, M., Li, X., Tolliday, N.J., Golub, T.R., Carr, S.A., Shamji, A.F., Stern, A.M., Mandinova, A., Schreiber, S.L., and Lee, S.W. (2011). Selective killing of cancer cells by a small molecule targeting the stress response to ROS. Nature 475, 231-234. Visvader, J.E., and Lindeman, G.J. (2012). Cancer stem cells: Current status and evolving complexities. Cell Stem Cell 10, 717-728. Wang, J., Zhang, L., Chen, G., Zhang, J., Li, Z., Lu, W., Liu, M., and Pang, X. (2014). Small molecule 10-acetoxychavicol acetate suppresses breast tumor metastasis by regulating the SHP-1/STAT3/MMPs signaling pathway. Breast Cancer Res Treat 148, 279-289. Wang, S., Zhang, C., Yang, G., and Yang, Y. (2014). Biological properties of 6-gingerol: a brief review. Nat Prod Commun 9, 1027-1030. Wang, X., Crowe, P.J., Goldstein, D., and Yang, J.L. (2012). STAT3 inhibition, a novel approach to enhancing targeted therapy in human cancers. Int J Oncol 41, 1181-1191. Watson, C.J., and Miller, W.R. (1995). Elevated levels of members of the STAT family of transcription factors in breast carcinoma nuclear extracts. Br J Cancer 71, 840-844. Wendt, M.K., Balanis, N., Carlin, C.R., and Schiemann, W.P. (2014). STAT3 and epithelial–mesenchymal transitions in carcinomas. JAK-STAT 3, e28975. Rasul, A., Khan, M., Ali, M., Li, J., and Li, X. (2013). Targeting apoptosis pathways in cancer with alantolactone and isoalantolactone. The Scientific World J 2013, 248532. Xiong, A., Yang, Z., Shen, Y., Zhou, J., and Shen, Q. (2014). Transcription factor STAT3 as a novel molecular target for cancer prevention. Cancers 6, 926-957. Saijo, H., Kofujita, H., Takahashi, K., and Ashitani, T. (2015). Antioxidant activity and mechanism of the abietane-type diterpene ferruginol. Nat Prod Res 29, 1739-1743. Yang, J., Cai, X., Lu, W., Hu, C., Xu, X., Yu, Q., and Cao, P. (2013). Evodiamine inhibits STAT3 signaling by inducing phosphatase shatterproof 1 in hepatocellular carcinoma cells. Cancer Letters 328, 243-251. Sandur, S.K., Pandey, M.K., Sung, B., and Aggarwal, B.B. (2010). 5-Hydroxy-2-methyl-1,4-naphthoquinone, a Vitamin K3 analogue, suppresses STAT3 activation pathway through induction of protein tyrosine phosphatase, SHP-1: Potential role in chemosensitization. Mol Cancer Res 8, 107-118 Yu, H., and Jove, R. (2004). The STATS of cancer-New molecular targets come of age. Nature Rev Cancer 4, 97-105. Schust, J., and Berg, T. (2004) A high-throughput fluorescence polarization assay for signal transducer and activator of transcription 3. Anal Biochem 330, 114-118. Yu, H., Pardoll, D., and Jove, R. (2009). STATs in cancer inflammation and immunity: a leading role for STAT3. Nature Rev Cancer 9, 798-809. Semwal, R.B., Semwal, D.K., Combrinck, S., and Viljoen A.M. (2015). Gingerols and shogaols: Important nutraceutical principles from ginger. Phytochemistry 117, 554-568. Shin, D.S., Kim, H.N., Shin, K.D. Yoon, Y.J., Kim, S.J., Han, D.C., and Kwon, B.M. (2009). Cryptotanshinone inhibits constitutive STAT3 function through blocking the dimerization in DU145 prostate cancer cells. Cancer Res 69, 193-202. 16 Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign Yu, H., Lee, H., Herrmann, A., Buettner, R., and Jove, R. (2014). Revisiting STAT3 signalling in cancer: new and unexpected biological functions. Nature Rev Cancer 14, 736-746. Yu, W., Xiao, H., Lin, J., and Li, C. (2013). Discovery of novel STAT3 small molecule inhibitors via in silico site-directed fragment-based drug design. J Med Chem 56, 4402-4412. Yuan, J., Zhang, F., and Niu, R. (2015). Multiple regulation pathways and pivotal biological functions of STAT3 in cancer. Sci Rep 5, 17663. You, M., Yu, D.H., and Feng, G.S. (1999). Shp-2 tyrosine phosphatase functions as a negative regulator of the interferon-stimulated Jak/STAT bdjn.org Yena Jin, Younghwan Kim, Yu-Jin Lee, Dong Cho Han and Byoung-Mog Kwon pathway. Mol Cell Biol 19, 2416-2424. Zhang, J., Ahn, K.S., Kim, C., Shanmugam, M.K., Siveen, K.S., Arfuso, F., Samy, R.P., Deivasigamani, S., Lim, L.H., Wang, L., Goh, B.C., Kumar, A.P., Hui, K.M., and Sethi, G. (2106). Nimbolide-induced oxidative stress abrogates STAT3 signaling cascade and inhibits tumor growth in bdjn.org transgenic adenocarcinoma of mouse prostate model. Antioxid Redox Signal in press. Zhao, C., Li, H., Lin, H.J., Yang, S., Lin, J., and Liang, G. (2016). Feedback activation of STAT3 as a cancer drug-resistance mechanism. Trends Pharmcol Sci 37, 47-61. Biodesign l Vol.4 l No.1 l Mar 30, 2016 © 2016 Biodesign 17