Next Article in Journal
Trans Species RNA Activity: Sperm RNA of the Father of an Autistic Child Programs Glial Cells and Behavioral Disorders in Mice
Next Article in Special Issue
Effect of Essential Oil Components on the Activity of Steroidogenic Cytochrome P450
Previous Article in Journal
Cardiovascular Biomarkers in Cardio-Oncology: Antineoplastic Drug Cardiotoxicity and Beyond
Previous Article in Special Issue
Inhibition of Cell Proliferation and Cell Death by Apigetrin through Death Receptor-Mediated Pathway in Hepatocellular Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Terpenoid-Mediated Targeting of STAT3 Signaling in Cancer: An Overview of Preclinical Studies

1
Center for Global Health Research, Saveetha Medical College and Hospitals, Saveetha Institute of Medical and Technical Sciences, Saveetha University, Chennai 602105, India
2
University Centre for Research and Development, Chandigarh University, Gharuan, Mohali 140413, India
3
Department of Chemistry, University Institute of Sciences, Chandigarh University, Gharuan, Mohali 140413, India
4
Department of Biotechnology, Parul Institute of Applied Sciences and Research and Development Cell, Parul University, Vadodara 391760, India
*
Author to whom correspondence should be addressed.
Biomolecules 2024, 14(2), 200; https://doi.org/10.3390/biom14020200
Submission received: 8 January 2024 / Revised: 1 February 2024 / Accepted: 5 February 2024 / Published: 7 February 2024

Abstract

:
Cancer has become one of the most multifaceted and widespread illnesses affecting human health, causing substantial mortality at an alarming rate. After cardiovascular problems, the condition has a high occurrence rate and ranks second in terms of mortality. The development of new drugs has been facilitated by increased research and a deeper understanding of the mechanisms behind the emergence and advancement of the disease. Numerous preclinical and clinical studies have repeatedly demonstrated the protective effects of natural terpenoids against a range of malignancies. Numerous potential bioactive terpenoids have been investigated in natural sources for their chemopreventive and chemoprotective properties. In practically all body cells, the signaling molecule referred to as signal transducer and activator of transcription 3 (STAT3) is widely expressed. Numerous studies have demonstrated that STAT3 regulates its downstream target genes, including Bcl-2, Bcl-xL, cyclin D1, c-Myc, and survivin, to promote the growth of cells, differentiation, cell cycle progression, angiogenesis, and immune suppression in addition to chemotherapy resistance. Researchers viewed STAT3 as a primary target for cancer therapy because of its crucial involvement in cancer formation. This therapy primarily focuses on directly and indirectly preventing the expression of STAT3 in tumor cells. By explicitly targeting STAT3 in both in vitro and in vivo settings, it has been possible to explain the protective effect of terpenoids against malignant cells. In this study, we provide a complete overview of STAT3 signal transduction processes, the involvement of STAT3 in carcinogenesis, and mechanisms related to STAT3 persistent activation. The article also thoroughly summarizes the inhibition of STAT3 signaling by certain terpenoid phytochemicals, which have demonstrated strong efficacy in several preclinical cancer models.

1. Introduction

The process of cancer advancement is intricate and multifaceted, starting with aberrant cells that have the potential to become malignant or neoplastic. Subsequently, the tumor grows, stromal invasion occurs, and metastasis occurs. This occurrence depends not only on tumor-intrinsic effects but also on the tumor microenvironment, which includes surrounding and supporting stroma, humoral factors, various immune system effectors, and vasculature [1,2]. The discovery of the Janus kinase (JAK)-signal transducer and activator of transcription (STAT) pathway occurred initially within the framework of downstream signaling mediated by interferon-α (IFNα), interferon-γ (IFNγ), and interleukin-6 (IL-6) [3,4]. Based on available data, STAT3 and STAT5, two of the seven members of the STAT protein family, are crucial for advancing cancer. They are essential transcription factors that control the expression of a wide variety of genes [5,6], contributing to tumor growth. In addition, they are vital for transducing signals from many receptor and non-receptor tyrosine kinases that are commonly active in cancer cells. Although STAT3 and STAT5 both play a role in the survival and multiplication of tumor cells, STAT3 stands out as a potentially effective target for cancer treatment because it served as an important player in the recruitment of stromal cells such as immune cells to the tumor microenvironment, which supports the growth of tumors [7,8].
In addition to being an oncogene and transcription activator, STAT3 is essential for tumor cell proliferation, invasion, and migration. It can induce the epithelial–mesenchymal transition (EMT), regulate the tumor microenvironment, and support cancer stem cell (CSC) self-renewal and differentiation, all of which are beneficial to the advancement of cancer [9,10]. STAT3 is commonly found with stimulated activity in most human carcinoma cell lines and tumor tissues. Previous research has demonstrated that STAT3 can also control the expression of genes by epigenetic alteration. For example, unphosphorylated STAT31 can control the architecture of the chromatin, and acetylated STAT3 can contribute to the DNA methylation that silences tumor-suppressor genes [11]. Numerous findings have demonstrated the pivotal function of STAT3 in developing cancer, rendering it a prime candidate for cancer treatment [12].
Recently, numerous unforeseen functions of the JAK-STAT3 pathway in cancer have surfaced, along with the fundamental processes via which this route is triggered and carries out its carcinogen-promoting impacts. Historically, cytokines and growth factors were assumed to be the primary activators of the JAK-STAT3 pathway. Toll-like receptors (TLRs) such as TLR9 and TLR4 have been found in several recent investigations to be significant activators of the JAK–STAT3 pathway [13,14]. In turn, STAT3 stimulates the production of specific TLRs in cancerous cells, accelerating the growth of tumors [15,16].
Several studies have found a link between nutrition and cancer therapeutic efficacy. Using phytochemicals from dietary and medicinal plants appropriately has been demonstrated to reduce cancer mortality by affecting the activation of many carcinogenic molecules [17,18]. An increasingly popular and feasible method of lessening the effects of cancer is natural cancer chemoprevention, which focuses on phytochemicals, vitamins, and minerals as a safe and affordable alternative to anticancer medication. Unlike pharmaceutical drugs, which are mono-target molecules, herbal treatments contain multitarget substances, meaning they can regulate the start and spread of cancer [19,20]. Recent research on these natural chemicals has revealed that they target several cellular signaling pathways, including NF-kB, MAPK, Wnt, Akt, Notch, p53, AR, and ER pathways, to demonstrate their multifaceted impact on cancer cells. This implies that the prevention and treatment of tumor formation may benefit from using these natural substances, either on their own or in conjunction with traditional therapeutic drugs [21]. Terpenoids are the largest class of natural compounds and have been shown to have about 25,000 different chemical structures. They have potential uses in the chemical, fragrance, medicinal, and flavoring industries [22]. Terpenoids are classified into subclasses based on their structures: monoterpenoids, sesquiterpenoids, diterpenoids, triterpenoids, and tetraterpenoids. Terpenoids have garnered significant attention due to their potential health advantages, including the possibility of having anticancer properties and other relevant pharmacological activities [23]. This review article intends to investigate numerous naturally occurring terpenoids that can reduce the JAK/STAT signaling pathway, thereby decreasing cancer cell development caused by abnormal JAK/STAT signaling. We have compiled and evaluated the most recent studies to find putative therapeutic and chemopreventive terpenoids with known molecular targets.

2. Terpenoid-Mediated Cancer Chemoprevention

Several experimental findings have suggested that a diet high in fruits and vegetables was frequently linked to lower cancer risk. There have been investigations of effective phytochemicals inhibiting cancer signaling pathways. Consequently, the scientific and medical communities are very interested in the bioactivities of natural phytochemicals and their potential to prevent cancer [24,25]. A wide range of potential chemopreventive bioactives is being studied. The term “chemoprevention” was initially used by Sporn et al. to describe the use of natural or synthetic chemicals [26]. According to a recently developed definition, chemopreventive agents can inhibit the molecular pathways contributing to the genesis and spread of cancer. A potential link between dietary ingredients and carcinogenesis has been suggested by the diverse eating habits that have been increasingly responsible for either lower or greater incidences of cancer development in ethnic groups in various nations [27]. Cancers of the breast, prostate, and gastrointestinal system, for example, have been observed to be less common in the Asian population. Drug research can benefit from novel approaches provided by natural ingredients [28]. One of the most practical methods of controlling cancer may be phytochemicals like terpenoids through cancer chemoprevention. Terpenoids (also referred to as terpenes) are easily derived from vegetables, fruits, spices, teas, herbs, and medicinal plants and have been shown to reduce experimental carcinogenesis in several organs in preclinical models. Terpenoids are categorized into monoterpenes (C10), sesquiterpenes (C15), diterpenes (C20), sesterterpenes (C25), triterpenes (C30), tetraterpenes (C40), and polyterpenes based on the number of building units. Terpenoids, sometimes referred to as isoprenoids, are a family of naturally occurring chemicals derived from plants that are thought to be the most diversified and have a variety of crucial physiological roles [29,30]. Currently, terpenoids have more than 25,000 different chemical structures and potential uses in the chemical and fragrance industries. According to recent reports, the mechanisms underlying terpenoids’ potential to prevent cancer may involve a combination of immune-boosting, anti-inflammatory, antioxidant, and hormone modulation effects, as well as have an impact on drug-metabolizing enzyme expression, cell cycle progression and differentiation, induction of apoptosis, and suppression of proliferation and angiogenesis [31]. These mechanisms would therefore play a role in the initial and secondary modification stages of cancer development. Even though terpenes as dietary supplements have demonstrated enormous promise for cancer prevention in the past ten years, more thorough preclinical and translational research is still required [32]. We have thus investigated the chemopreventive mechanism of this class of natural chemicals by focusing on the targets of the STAT3 signaling pathway.

3. STAT3 Signaling Cascade and Tumorigenesis

The term “signal transducer and activator of transcription” comes from the cytoplasmic inducible transcription factors comprising signal transducer and activator of transcription (STAT) proteins. These proteins relay extracellular signals from the cell surface to the nucleus to express the target genes [33,34]. Since their discovery, the STAT protein family of transcription factors has been thoroughly researched in relation to cancer [35]. The seven members of the STAT family are STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, and STAT6 [36,37]. The 750–850 amino acid range of these proteins varies, and their functional role is to mediate signals from growth factors (GFs) and cytokine receptors on the cell membrane to the nucleus [38]. Extracellular signaling proteins (growth factors and cytokines) frequently activate STAT3 signaling when they bind precisely to the appropriate receptors on the cell surface. Once activated by phosphorylation, STAT3 forms homodimers, moves into the nucleus, and attaches to DNA to regulate the expression of downstream genes, boosting cell proliferation and survival, invasion, angiogenesis, chemotherapeutic resistance, and immunological escape [39]. One important STAT3 upstream regulator is interleukin 6 (IL-6). When IL-6 binds to the IL-6 receptor (IL-6R), a heterohexameric complex is formed consisting of IL-6, IL-6R, and the membrane-spanning protein IL-6R subunit (gp130, also known as IL-6Rβ). Then, STAT3 and Janus kinases (JAKs) are recruited and activated. Afterward, pJAKs tyrosine phosphorylate the cytoplasmic portion of gp130 [40]. By identifying and binding to the phosphorylated tyrosine docking site on phosphorylated gp130, STAT3’s SH2 domain attracts STAT3 to the area around the active JAK enzyme. Following this, Tyr705 on STAT3 becomes phosphorylated, encouraging the SH2 domain to bind with Tyr705 to create a head-to-tail dimer. Following its dissociation from cell surface receptors and translocation into the nucleus, the pSTAT3 dimer controls the expression of several STAT3 target genes [41]. A positive feedback loop is created when cancer cells’ overactive STAT3 signaling pathway increases IL-6 synthesis. STAT3 can be activated by the non-receptor tyrosine kinase Src in addition to JAKs [42]. STAT3 can control the expression of genes linked to cancer development, and its chronic activation has been extensively studied in various human cancers [43] (Figure 1). Takeda et al. showed that selective loss of STAT3 in mice resulted in an early lethal phenotype during embryogenesis, and tissue-specific ablation showed decreased oncogenesis and increased inflammation [44]. Numerous genes that control antiapoptosis (Bcl-2, Bcl-xL, survivin, and Mcl-1) and cell cycle progression (cyclin D1, c-Myc, and Pim1/2), angiogenesis (HGF, VEGF, bFGF, IL-17, and IL-23), cell proliferation (IL-6, CXCL12, and HIF-1α), metastasis (MMP-1/2/3/9, fascin, vimentin, and ICAM-1), and immune suppression and inflammation (IL-10, IL-23, TGFβ, and COX2) have been demonstrated to be upregulated by STAT3 [45]. Apart from the upregulation of prosurvival proteins, STAT3 can potentially decrease the expression of growth suppressor transcription factors, including p53 and nedin [46].
Furthermore, inconsistent chemotherapeutic activity is caused by chemoresistance arising from epithelial–mesenchymal transition (EMT) and cancer stem cells (CSCs) [47]. pSTAT3 has been discovered to be a significant contributor to CSCs and EMT, and inhibiting the STAT3 pathway has been shown to revert chemoresistance. To initiate the activation of EMT-related genes, activated STAT3 in the form of pSTAT3 forms a homodimer and is translocated into the nucleus. Numerous downstream STAT3 targets, including ZEB1, TWIST, SNAIL, FOXM1, and SLUG, facilitate the expression of EMT mediators and ABC transporters, resulting in chemoresistance [48,49,50].

4. Terpenoids: Modulators of STAT3 Signaling

Phytochemicals harm cancer cells via various methods, including signaling pathway alteration and the initiation of apoptosis. Anti-cancer drugs work by preventing the emergence of carcinogenic species and carcinogens from interacting with cells, which delays the development of tumors [51,52]. The signaling cascades most strongly linked to tumor progression are mitogen-activated protein kinase (MAPK), NF-kB, and activator of transcription proteins (STAT). The metabolic pathways in different types of cancer are controlled by modulating the signaling pathways, which can be overactivated or inhibited. In addition to promoting angiogenesis, enhanced glycolysis, and metastasis, the modifications further contribute to the development and spread of cancer [53]. For the treatment of malignancies, signaling pathways and modified enzymes and factors are significant targets to block or stimulate [54]. Terpenoids are among the bioactive phytocompounds shown to inhibit the JAK/STAT pathway through various methods. There are multiple sites of action in the JAK/STAT pathway that phytochemicals can target [55,56]. These phytochemicals can inhibit this signaling pathway by lowering the quantities of growth hormones or cytokines that activate the JAK/STAT protein [57]. Terpenes can also function by inhibiting JAK phosphorylation before STAT activation. It is possible to regulate JAK/STAT signaling by blocking the translocation of STAT dimer from the cytoplasm into the nucleus and STAT dimerization. The ultimate purpose of the signaling pathway is to prevent STAT-DNA binding, which directly suppresses JAK/STAT-regulated gene transcription [58,59]. This group of phytoconstituents can block the JAK/STAT pathway by preventing STAT3 from being phosphorylated. The phosphorylation of JAK1, JAK2, and c-Src in diverse cancer cell lines was also suppressed by various terpenoids [60] (Table 1) (Figure 2). Subsequent sections address the function of different terpenoids in the JAK/STAT signaling pathway in different types of cancer. In the following sections, we examine the experimental evidence for the ability of various terpenoids derived from multiple plants to modulate the JAK/STAT system in preclinical cancer models.

4.1. Monoterpene

Thymoquinone

Thymoquinone (TQ), an ingredient obtained from the medicinal plant Nigella sativa, has been used in medicine for almost 2000 years. A phytochemical found in black seed (Nigella sativa), thymoquinone is widely utilized in traditional medicine in the Middle East, the Mediterranean, South Asia, and Africa. Recent research has discovered that TQ has an adverse effect against many cancer subtypes and suppressed cellular growth in several cancer cell lines, including breast, colon, larynx, lung, myeloblastic leukemia, osteosarcoma, ovary, and pancreatic [126,127,128]. According to a study by Zhu et al., TQ inhibited the expression of multiple STAT3-regulated genes in gastric cancer HGC27 cells, including proliferative (cyclin D1), anti-apoptotic (Bcl-2, survivin), and angiogenic (VEGF) gene products [61]. Stimulation of colon cancer cells with TQ reduced nuclear localization, constitutive phosphorylation, and reporter gene activity of STAT3. TQ increased the expression of the cell cycle inhibitory proteins p27 and p21 and decreased the expression of STAT3 target gene products, including survivin, c-Myc, and cyclin-D1 and -D2. TQ treatment reduced the phosphorylation of upstream kinases, including Src kinase, EGFR tyrosine kinase, and Janus-activated kinase-2 (JAK2). Drug-induced tyrosine phosphorylation of JAK2 and Src attenuated EGFR and STAT3 tyrosine phosphorylation, but EGFR tyrosine kinase inhibitor therapy (gefitinib) suppressed STAT3 phosphorylation in HCT116 colon cancer cells without influencing JAK2 or Src phosphorylation [62]. Al-Rawashde et al. [63] explained that TQ significantly reduced cell proliferation and caused apoptosis by downregulating the expression levels of BCR ABL, JAK2, STAT3, STAT5A, and STAT5B targets in K562 leukemia cells. Another study found that in HL60 leukemia cells, TQ-mediated activation of apoptosis and reduced cell proliferation were linked to the downregulation of essential elements of the JAK/STAT (such as JAK2, STAT3, p-STAT3, STAT5, and p-STAT5) and PI3K/AKT (such as PI3K, p-PI3K, AKT, and p-AKT,) signaling pathways [64]. TQ significantly reduced both constitutive and inducible phosphorylation of STAT3 in breast tumor tissue but did not affect STAT5 in breast cancer. Additionally, it reduced the growth of tumors, downregulated downstream target proteins of STAT3 (such as Jak2, Src, and cyclin D1), and boosted caspase-3 and -9 apoptotic activity in breast cancer. Remarkably, the elevation of tumoral microRNA-125a-5p was concomitant with the synergism of doxorubicin anti-neoplastic action that TQ induced [65]. Moreover, TQ’s growth-inhibitory and apoptosis-inducing actions in renal cell carcinoma and melanoma were linked to the suppression of Jak2/STAT3 by downregulating the phosphorylated level of STAT3 [66,67]. Additional research revealed that by preventing the phosphorylation of the upstream kinase Src, TQ reduced the constitutive phosphorylation and DNA binding activity of signal transducer and activator of transcription-3 (STAT3) in skin cancer and myeloma. Furthermore, TQ administration reduced the expression of the STAT3 target gene products survivin and cyclin D1 [68,69].

4.2. Diterpene

4.2.1. Andrographolide

Andrographis paniculate (Burm.f.) Nees is a widely growing plant revered as a successful herbal remedy in many nations. Andrographolide is a novel diterpene lactone derived from A. paniculate [129]. A growing body of research has shown that AD has anti-inflammatory, anti-HIV, anti-influenza virus, and anti-cancer properties both in vivo and in vitro [130]. By directly and highly selectively binding to STAT3, andrographolide controlled autophagy and programmed death-ligand 1 (PD-L1) expression in NSCLC. Proteomics investigation revealed that andrographolide inhibited JAK2/STAT3 signaling by downregulating the phosphorylation status of STAT3, which in turn decreased the expression of tumor PD-L1 in NSCLC [70]. An additional investigation revealed a correlation between the constitutive activation level of STAT3 and the cancer cell’s ability to withstand doxorubicin-induced apoptosis. The short-term MTT assay and the long-term colony formation experiment revealed that andrographolide significantly increased doxorubicin-induced cell death in cancer cells, showing that andrographolide increases cancer cell sensitivity to doxorubicin primarily through STAT3 inhibition. Mechanistically, andrographolide inhibited IL-6-mediated STAT3 activation and concomitant nuclear transport, which was further associated through the repressed expression of Jak1/2 as well as the interaction between STAT3 and gp130 [71].

4.2.2. Cryptotanshinone

One of the main tanshinones identified from the well-known Chinese herb Salvia miltiorrhiza Bunge is called cryptotanshinone (CPT), and it is utilized extensively in conventional and modern medicine [131]. Numerous studies have suggested that CPT may limit the growth and metastasis of malignancies in vivo and the cell growth of cancer cells in vitro [132]. Numerous in vitro tests also revealed various ways in which CPT suppresses tumors. Overexpression of STAT3 has been linked to cancer, and CPT treatment slows the growth of cancer by blocking STAT3 by selectively decreasing STAT3 Tyr705 phosphorylation and its downstream target proteins, which include Bcl-xL, survivin, and cyclin D1 in prostate cancer [72]. Furthermore, ovarian cancer cells treated with CPT showed a significant reduction in lactate generation and glucose uptake. The mechanism by which CPT exhibited its anti-tumor impact involved focusing on the STAT3/SIRT3/HIF-1α signaling cascade both in vitro and in vivo. This pathway may be restored by introducing SIRT3 shRNA into ovarian cancer cells [73]. Ge et al. showed that by decreasing the activity of STAT3 and numerous upstream regulatory signaling pathways, CPT dramatically caused cell death and cell cycle arrest of the BxPC 3 human pancreatic cancer cells [74]. In esophageal cancer cells, CPT also reduced the expression of STAT3 and the activation of STAT3 mediated by IL-6. The mechanism behind the stimulation of apoptosis and prevention of cell proliferation in esophageal cancer cells has been attributed to a CPT-mediated downregulation of STAT3 [75]. Furthermore, cryptotanshinone significantly inhibited cancer cell proliferation, promoting cell death by reducing STAT3 phosphorylation at Tyr705 and impeding nuclear translocation. Cell proliferation was repressed due to the coordinated downregulation of P-AKT, cyclin D1, C-MYC, MEKK2, survivin, and HGF, which halted cell cycle progression at the G0/G1 phase [76]. Some studies have shown that CPT has a growth-inhibiting impact on glioma cells. CPT reduced the nuclear translocation of STAT3 and decreased the phosphorylation of Tyr705 but not Ser727. Cell cycle progression was severely halted in the G1/G0 phase after the STAT3-regulated proteins cyclin D1 and survivin were downregulated [77,78]. Additionally, CPT displayed growth-suppressive effects mediated by STAT3 in chronic myeloid leukemia models [79,80]. These findings suggest that CPT might be a potent antiproliferation medication for managing several types of carcinoma and that STAT3 signaling suppression may play a role in its mode of action.

4.2.3. Oridonin

Oridonin is a naturally occurring terpenoid used in many Chinese medicine formulations. It is derived from Isodon rubescens (Hemsl.) H.Hara, a plant used in traditional Chinese herbal medicine. Oridonin has several established anti-cancer properties, including the ability to fight off oral cancer, gastric cancer, nasopharyngeal carcinoma, esophageal cancer, ovarian cancer, leukemia, myeloma, and so forth [133,134,135,136]. Its primary mechanisms include suppression of migration and invasion, induction of apoptosis, inhibition of proliferation, and reversal of drug resistance [137,138]. Oridonin has significantly decreased the migration and invasion of BCPAP and TPC-1 thyroid cancer cells, as demonstrated by the Matrigel invasion experiment, transwell migration assay, and wound healing assay. An increasing body of research suggests that the epithelial–mesenchymal transition is related to the JAK2 (Janus kinase-2)/STAT3 (signal transducer and activator of transcription 3) signaling pathway. As predicted, when oridonin was administered to thyroid cancer TPC-1 and BCPAP cells, the protein levels of phosphorylated-JAK2 and phosphorylated-STAT3 were significantly decreased and also prevented EMT in these cells [81]. In vitro research using human nasopharyngeal cancer cell lines (CNE-2Z and HNE-1) demonstrated that oridonin markedly reduced cell invasion and migration. Western blotting results later showed that oridonin administration decreased the phosphorylation levels of AKT and STAT3. The findings are consistent with the hypothesis that decreased AKT and STAT3 protein activity is linked to oridonin’s anti-metastatic action [82]. Additionally, oridonin induced apoptosis and reduced the growth of U2OS osteosarcoma cells [83]. As demonstrated by DAPI staining and reduced expression of antiapoptotic proteins, the antiproliferative effects were mostly caused by the induction of apoptosis. Additionally, it was shown that oridonin significantly reduced the expression of MMP-2, -3, and -9 in osteosarcoma cells and, in addition, reduced the level of phosphorylation of p-STAT3 (Tyr 705) and p-STAT3 (Ser 727) [83].

4.3. Triterpene

4.3.1. Brusatol

Brusatol (Bru), a substance initially identified and extracted from Brucea sumatrana seeds in 1968, has long been used to treat amebiasis, malaria, and dysenteric illnesses. With the rapid advancement of medical chemistry, Bru has been reported to be a potent antitumor agent in several types of malignant cells, including lung cancer, pancreatic cancer, colorectal cancer, and leukemia [139]. Hall et al. (1979) revealed that Bru had a potent suppressive impact on tumor cell metabolism and proliferation of lymphocytic leukemia cells by significantly reducing c-myc synthesis [140]. The cell viability, migration, and invasion were all significantly reduced by Bru in Hep-2 laryngeal cancer cells. Bru also blocked Hep-2 cells in the S phase, resulting in cell death in Hep-2 laryngeal cancer cells. Bru further decreased the expression levels of markers related to the epithelial–mesenchymal transition (EMT). Western blotting data demonstrated that Bru could reduce JAK2/STAT3 protein level and phosphorylation status. The anti-tumor action of Bru on Hep-2 cells in vitro was indicated by in vivo tests, where Bru showed no observable toxicity and inhibited the growth of xenograft laryngeal tumors [84]. Bru-induced head and neck squamous cell carcinoma growth was similarly attributed to the disruption of the activation of STAT3 and upstream kinases such as JAK1, JAK2, and Src [85]. It also decreased nuclear STAT3 levels and the protein’s capacity to bind DNA [85]. Bru was shown to reduce proliferation and cause apoptosis in PATU-8988 and PANC-1 pancreatic cancer cells by lowering Bcl-2 expression and enhancing Bax and cleaved Caspase-3 expression. Possible pro-apoptotic signaling pathways are linked to the repression of NF-kB, STAT3, JNK, and p38 MAPK stimulation in pancreatic cancer cells [86]. In hepatocellular carcinoma HCCLM3 cell line and tumor tissues, Bru therapy decreased the levels of many mesenchymal markers while upregulating the expression level of epithelial markers such as occludin and E-cadherin. Furthermore, a decrease in the expression of the transcription factors Twist and Snail was noted. Because STAT3 signaling can control the expression of these two factors, the modulation of this pathway by Bru was also studied [87]. Bru inhibited the phosphorylation of STAT3Y705, while STAT3 knockdown with siRNA resulted in the restoration of epithelial markers. Bru (1 mg/kg) significantly reduced tumor burden in an orthotopic mouse model of liver cancer while decreasing lung metastasis [87].

4.3.2. Betulinic Acid

A pentacyclic triterpene of the lupane class, betulinic acid shares several biological properties with other triterpenoids, most notably anticancer properties. Recent studies on the cytotoxic and anticancer properties of betulinic acid have demonstrated that it promotes apoptosis in cancer cells while enhancing apoptosis throughout the mitochondrial route in healthy cells without harming them [141,142]. In multiple myeloma cells, betulinic acid suppressed constitutive stimulation of STAT3, Src kinase, JAK1, and JAK2. The downregulation of STAT3 activation caused by betulinic acid was reversed by pervanadate, indicating the participation of a protein tyrosine phosphatase (PTP). Moreover, betulinic acid increased PTP SHP-1 expression, and the inhibition of STAT3 activation by betulinic acid was eliminated by hindering the SHP-1 gene, preventing betulinic acid-induced apoptosis. Additionally, betulinic acid inhibited the expression of genes that are controlled by STAT3, including cyclin D1, survivin, bcl-xL, and bcl-2. A boost in the population of sub-G1 cells and an increase in caspase-3-mediated PARP cleavage suggested a correlation with an increase in apoptosis. In agreement with these findings, betulinic acid-induced apoptosis was considerably attenuated by constitutive active STAT3 overexpression [88]. According to Shin et al. [89], betulinic acid inhibited the phosphorylation level, DNA binding property, and nuclear accumulation of STAT3 in PC-3 prostate cancer cells caused by hypoxia. Furthermore, betulinic acid dramatically decreased cellular and secreted amounts of VEGF, a major angiogenic factor and a target gene of STAT3 produced by hypoxia. Notably, the chromatin immunoprecipitation (ChiP) experiment showed that BA prevented STAT3 and HIF-1α from attaching to the VEGF promoter. Moreover, BA therapy under hypoxia significantly decreased VEGF production, which was successfully increased by siRNA transfection-mediated STAT3 suppression.

4.3.3. Celastrol

Celastrol is a pentacyclic triterpenoid obtained from the traditional Chinese medicinal Tripterygium wilfordii and exhibits various pharmacological properties. Celastrol in particular has been shown in modern pharmacological studies to have significant broad-spectrum anticancer activities in the treatment of a variety of cancers, including lung cancer, liver cancer, colorectal cancer, hematological malignancies, gastric cancer, prostate cancer, renal cell carcinoma, breast cancer, osteosarcoma, glioma, cervical cancer, and ovarian cancer [143]. Celastrol suppressed both constitutive and stimulated STAT3 activation in hepatocellular carcinoma, and this inhibition was mediated by inhibiting the expression of upstream kinase c-Src and Janus-activated kinase-1 and -2. Celastrol inhibited STAT3 activity, suppressing several gene products involved in proliferation, survival, and angiogenesis such as cyclin D1, Bcl-2, Bcl-xL, survivin, and VEGF. Ultimately, celastrol prevented the formation of human HCC xenograft tumors in athymic nu/nu mice and reduced STAT3 activation in tumor tissues when administered intraperitoneally without causing any adverse effects [90]. Overall, these findings show that celastrol shows antiproliferative and proapoptotic effects in HCC by suppressing STAT3 signaling both in vitro and in vivo. Zhao et al. [91] demonstrated that celastrol inhibited the growth, proliferation, and metastasis of NSCLC cells. Celastrol caused a substantial increase in intracellular ROS, which activated the ER stress pathway, inhibited the P-STAT3 pathway, and ultimately resulted in cell death. These effects were countered by pre-treatment with N-acetyl-L-cysteine (NAC). Celastrol’s tumor suppressive effect was also demonstrated in an in vivo lung cancer model of HL60 cells. Furthermore, Kannaiyan and colleagues [92] revealed that celastrol suppressed the proliferation of multiple myeloma (MM) cell lines. It improved thalidomide and bortezomib’s apoptotic effects synergistically. Celastrol’s effects were mediated via phosphorylation of both p65 and IκBα, as well as constitutively active NF-κB produced by inhibition of IκBα kinase activity. Additionally, celastrol reduced the constitutive and IL6-induced stimulation of STAT3 in MM, which resulted in apoptosis, as demonstrated by an increase in the sub-G1-phase cell accumulation, pro-apoptotic protein expression, and caspase-3 activity.

4.3.4. Cucurbitacin B

Cucurbitacin B, or CuB, is a steroid with a unique bitter taste and cytotoxic characteristics found in plants from the Cucurbitaceae family and other plant families like Brassicaceae. CuB is a naturally occurring triterpenoid from the Cucurbitaceae family of plants. It has been discovered that there are more than 18 different types of cucurbitacin, with cucurbitacin B being a common kind [144]. CuB has been shown to activate apoptosis in various cancer types via the signaling pathways of Wnt/β-catenin, JAK/STAT, NF-κB, PI3K/AKT, and MAPK/ERK [145,146]. Cucurbitacin B induced growth inhibition and apoptosis in pancreatic cancer cells in a dose- and time-responsive manner [93]. This was connected to decreased expression of cyclin A, cyclin B1, and Bcl-XL, with subsequent activation of the caspase cascade; suppression of active JAK2, STAT3, and STAT5; and elevated levels of p21WAF1 even in cells without functioning p53. Notably, gemcitabine’s antiproliferative effects on pancreatic cancer cells were enhanced by the synergistic interaction of gemcitabine (pyrimidine nucleoside analog) and CuB. Additionally, through JAK2/STAT3 downregulation, CuB reduced the size of pancreatic cancer xenografts in athymic nude mice compared to controls without apparent drug toxicities. In a different study, the anticancer potential of CuB was demonstrated in PANC-1 pancreatic cancer cells through the suppression of Bcl-2, survival expression, and STAT3 activation [94]. According to Yar Saglam et al., CuB, both by itself and in conjunction with gefitinib (a selective inhibitor of EGFR), significantly inhibited the proliferation of colorectal cancer cells and triggered apoptosis. CuB administration alone or coupled with gefitinib reduced cyclin D1, pSTAT3 (Tyr705), pEGFR, Bcl-2, and BCL2L1 levels [95]. Moreover, it was found that in HepG2 hepatocellular carcinoma cells, cucurbitacin B reduced cell viability, induced cell cycle arrest, and formed apoptotic bodies. Protein expression analysis revealed that CuB decreased the expression of Bcl-2 and reduced the phosphorylation of STAT3. CuB also suppressed the growth of a HepG2 tumor in nude mice [96]. Subsequent research revealed that CuB markedly reduced gastric cancer cell growth while inducing apoptosis and cell cycle arrest. The expression of STAT3 and downstream targets, including cyclin D1, c-Myc, and Bcl-xL, markedly decreased with CuB therapy. CuB combined with cisplatin increased gastric cancer cell cytotoxicity, probably due to enhanced reduction in tyrosine phosphorylation of STAT3 (TYR-705). Moreover, the therapeutic efficacy of CuB in vivo was validated using a xenograft mouse model [97,98].

4.3.5. Cucurbitacin E

Cucurbitacin E (CuE), an oxygenated tetracyclic triterpenoid, is prevalent in the fruit pulps of Citrullus colocynthis. CuE has been shown to exhibit significant biological activities, such as antiviral, antitumor, and anti-inflammatory properties [147]. It is a type of triterpene that exhibits strong anticancer properties in many cancer cells [148]. Cucurbitacin E has induced dose- and time-dependent apoptosis, cell cycle arrest, and growth suppression in pancreatic cancer cells [99]. According to additional findings from Western blotting, cucurbitacin E therapy can decrease STAT3 phosphorylation while upregulating p53 expression in PANC-1 pancreatic cancer cells.
In the event of prostate cancer, on the other hand, CuE dramatically suppressed downstream protein kinases, including extracellular signal-regulated kinase and p38 mitogen-activated protein kinases, and blocked the vascular endothelial growth factor receptor (VEGFR) 2-mediated Janus kinase (Jak) 2-signal transducer and activator of transcription (STAT) 3 signaling pathway [100]. Additional research on bladder cancer cells revealed that cucurbitacin E-induced G2/M arrest might be associated with downregulation of pSTAT3, cyclin-dependent kinase 1 (CDK1), and cyclin B in human bladder cancer T24 Cells [101].

4.3.6. Cucurbitacin I

The natural tetracyclic triterpenoid cucurbitacin I (CuI), which is derived from the Cucurbitaceae and Cruciferae families, has been utilized in traditional medicine for its antipyretic, analgesic, anti-inflammatory, and antibacterial properties. According to recent results, CuI, one of 12 cucurbitacin derivatives, has a powerful anticancer effect against different cancer cells, such as lung cancer, breast cancer, prostate cancer, cervical cancer, colon cancer, gastric cancer, and hepatocellular carcinoma [149]. Numerous studies have revealed that STAT3 is an indirect molecular target of CuI’s anticarcinogenic action [150]. CuI causes cell apoptosis, inhibits the growth of pancreatic cancer, and stops the cell cycle in the G2/M phase [102]. Additionally, the compound significantly slowed down cell invasion potential and hindered the development of pancreatic tumor xenografts in nude mice. Furthermore, CuI-induced suppression of pancreatic cancer cell growth appears to include JAK2/STAT3 signaling pathway suppression because JAK2/STAT3 activators successfully prevented the inhibition [102]. Furthermore, it was demonstrated that administering CuI to HepG2 hepatocellular carcinoma cells decreases the number of proteins in the PI3K/AKT/mTOR, MAPK, and JAK2/STAT3 cascades. In varying doses, CuI also reduced the expression of the genes for MAPK, STAT3, mTOR, JAK2, and Akt [103]. Ni et al. found that CuI suppresses ERK activation and the downstream phosphorylation level of mTOR and STAT3, which may cause reduced viability of A549 lung cancer cells, impeded colony formation, and induced apoptosis [104]. Another study that examined the anticancer potential of various cucurbitacin compounds in different cancer models found that CuI inhibited the activation of STAT3 and JAK2 in A549 lung cancer cells [105]. Furthermore, cucurbitacin E and B inhibited the activation of both STAT3 and JAK2 but mildly activated apoptosis and suppressed tumor growth in lung cancer. Subsequent analysis revealed that glioblastoma cells treated with CuI had lower levels of phosphorylated-STAT3, accompanied by increased apoptosis and cell cycle arrest. In particular, CuI caused G2/M accumulation in glioblastoma cells by downregulating the downstream targets of STAT3 such as cyclin B1 and cdc2 levels [106].
Furthermore, this compound was found to lower the levels of phosphorylated STAT3 and Janus kinase-3 (JAK3), which may be a significant factor in the reduction of cell viability, cell cycle arrest, and activation of apoptosis in lymphomas. Alongside these modifications, there were notable decreases in several other downstream targets of STAT3, such as cyclin D3, mcl-1, bcl-2, and bcl-xL [107].

4.3.7. Ursolic Acid

The cyclosqualenoid family contains ursolic acid (UA), a ursane-type pentacyclic triterpene acid widely found in the leaves and berries of naturally occurring medicinal plants, including Vaccinium macrocarpon Air. (cranberry), Arctostaphylos uva-ursi (L.) Spreng (bearberry), Calluna vulgaris, Ocimum sanctum, and many more. Recent research has shown evidence of ursolic acid’s therapeutic benefits in various human disorders, including different kinds of malignancies caused by inflammation [151].
In A549 and H460 lung cancer cells, UA was explicitly recognized as suppressing STAT3 activity. Additionally, it prevented the growth of tumor spheres, migration, invasion, and angiogenesis. UA efficacy involves UA binding to the epidermal growth factor receptor (EGFR), lowering pEGFR levels, and blocking the downstream JAK2/STAT3 pathway. Moreover, UA decreased the expressions of MMPs, VEGF, and PD-L1, as well as the formation of STAT3/MMP2 and STAT3/PD-L1 complexes [108]. Additionally, the effects of UA on inhibiting the nuclear translocation of STAT3 in colorectal cancer cells were assessed, and it was found that in HCT116 and HT29 cells, UA substantially decreased the expression of JAK2 and STAT3 [109]. Moreover, it is commonly recognized that IL-6 treatment results in the phosphorylation of STAT3 [152]; consequently, the authors investigated whether UA could prevent the STAT3 phosphorylation that IL-6 produced. UA significantly decreased P-STAT3, which produces interleukin-6 (IL-6) in hepatocellular cancer cells. In these cell lines, UA reduced the expression of STAT3’s downstream target genes, including Bcl-2, Bcl-xl, and survivin. UA also inhibited colony formation, cell migration, and viability in liver cancer cells. In vivo, oral daily administration of UA inhibited both the development of HEPG2 tumors and STAT3 phosphorylation [110]. In the dorsolateral prostate (DLP) tissues of a TRAMP mouse model of prostate cancer, UA inhibited the activation of many pro-inflammatory mediators, such as NF-κB, STAT3, AKT, and IKKα/β phosphorylation. This was linked with a decrease in blood levels of TNF-α and IL-6. Furthermore, immunohistochemical analysis of tumor tissue sections showed that UA dramatically up-regulated the levels of caspase-3 as well as substantially downregulated the expression of cyclin D1 and COX-2 [111].

4.4. Sesquiterpene

4.4.1. Alantolactone

One of the most significant sesquiterpenes, alantolactone (ALT), is obtained in substantial quantities from the root of Inula helenium (elecampane). Some plants, including Inula species Aucklandia lappa, Rudbeckia subtomentosa, and Radix inulae, can also provide ALT, in addition to Inula helenium [153]. According to in vitro research [154], this substance prevents the growth of different cancer cells, including those found in breast, stomach, colon, pancreatic, and squamous cell lung malignancies. An increasing amount of data from recent years has indicated that ALT demonstrates anticancer potential by targeting the cell cycle, cell proliferation, invasion and metastasis, and the activation of apoptosis, as well as by blocking numerous signaling pathways implicated in the progression of cancer [155]. ALT inhibited the STAT3 phosphorylation and signaling pathway in prostate cancer cells after 72 h. Additionally, by upregulating the expression of p53 and downregulating the expression of SOX2, Oct-4, Nanog, CD133, and CD44, ALT was able to modify the stemness of prostate cancer cells [112]. Ahmad et al. [113] showed that, after suppressing the production of STAT3 and survivin, ALT reduced cell viability and enhanced cell death and apoptosis in human leukemia THP-1 cells. ALT was also shown to demonstrate potential efficacy against triple-negative breast cancer MDA-MB-231 cells by decreasing the STAT3 signaling pathway [114]. Alantolactone efficiently inhibited both constitutive and inducible STAT3 activation at tyrosine 705. The nucleus translocation of STAT3, its DNA-binding activity, and the expression of STAT3 target genes were all reduced by alantolactone. ALT significantly inhibited constitutively activated STAT3 in pancreatic cancer cells while having minimal effect on the EGFR pathway. Additionally, both in vitro and in vivo, the combination of alantolactone and an EGFR inhibitor such as erlotinib or afatinib showed a remarkable combinatorial anticancer effect against pancreatic cancer cells [115].

4.4.2. β-Caryophyllene Oxide

Natural bicyclic sesquiterpenes such as β-caryophyllene (BCP) and β-caryophyllene oxide (BCPO) are found in various plants worldwide, such as basil (Ocimum spp.), cinnamon (Cinnamomum spp.), black pepper (Piper nigrum L.), cloves (Syzygium aromaticum), cannabis (Cannabis sativa L.), and lavender (Lavandula angustifolia). Significant anticancer effects are possessed by both BCP and BCPO (BCP(O)), which inhibit the growth and multiplication of different cancer cells in vitro, such as cervical cancer HeLa cells, liver cancer HepG2 cells, leukemia AGS cells, gastric cancer SNU-1 cells, and stomach cancer SNU-16 cells [156]. CPO inhibited constitutive STAT3 stimulation in multiple myeloma, breast, and prostate cancer cell lines, with MM cells exhibiting significant dose- and time-dependent effects. The CPO-mediated STAT3 suppression was accomplished by blocking the activation of JAK1/2 and c-Src, two upstream kinases. Additionally, the down-regulation of STAT3 caused by CPO was reversed via vanadate therapy, indicating the participation of tyrosine phosphatase [116]. The authors observed that constitutive STAT3 activation was down-regulated in correlation with the tyrosine phosphatase SHP-1 expression that CPO promoted. Interestingly, CPO’s capacity to block STAT3 activation was eliminated with siRNA ablation of the SHP-1 gene. By blocking STAT3 activation, CPO reduced the ability of MM1.5 multiple myeloma tumor cell lines to grow, triggering apoptosis. The results indicate that CPO is a novel STAT3 signaling cascade blocker, which opens up an array of therapeutic options for various malignancies that have constitutively activated STAT3 [116].

4.4.3. Dihydroartemisinin

Artemisinin, a plant-derived antimalarial medicine, has a semisynthetic derivative called dihydroartemisinin (DHA). It is substantially more water soluble and has more antimalarial activity than other artemisinin derivatives [157]. According to recent research, DHA demonstrates anticancer effects by causing cell death, altering the cell cycle, and preventing tumor angiogenesis. DHA adversely affects several oncological signaling pathways, such as the JAK/STAT, Wnt/β-catenin, MAPK/ERK, and translationally controlled tumor protein (TCTP) cascades [158]. DHA suppressed breast cancer cell growth and promoted apoptosis in DDP (cisplatin)-resistant cells by inhibiting STAT3/DDA1 signaling. DDA1 (DET1- and DDB1-associated 1) knockdown increased DDP-resistant cell death, suppressed cell proliferation, enhanced G0/G1 phase arrest, and decreased cyclin D1 expression. Also, by targeting DDA1, STAT3 knockdown inhibited the growth of DDP-resistant cells, causing apoptosis and G0/G1 cell cycle arrest [117]. Additionally, DHA efficiently inhibited the phosphorylation of STAT3 and the inactivation of STAT3, leading to the downregulation of Mcl-1 and survivin, which improved the chemosensitivity of ABT-263 in lung cancer cells [118]. Subsequently, DHA and ABT-263 cotreatment substantially decreased xenograft development in nude mice. Jia et al. [119] also noticed that DHA exclusively blocked JAK2/STAT3 signaling in HNSCC head and neck cancer cells, exhibiting significant and specific inhibitory effects on STAT3 phosphorylation. Furthermore, DHA substantially decreased HNSCC growth in vivo (xenograft tumor model of Cal-27 cells) and in vitro (Fadu, Cal-27 and Hep-2), possibly by reducing cell migration and inducing death. Additionally, combined treatment of DHA and cisplatin decreased the growth of HNSCC tumors. A report by Wang et al. showed that DHA suppressed Jak2/STAT3 signaling, which in turn increased cell death and altered the expression of several apoptotic-related proteins in colon cancer cells [120]. Furthermore, DHA reduced the activity of the IL-6/STAT3 signaling pathway and prevented the downregulation of STAT3 and β-catenin protein expression mediated by miR-130b-3p. Moreover, the authors revealed that miR-130b-3p demonstrates direct targeting capability for STAT3, which reduces invasion by LSCC (laryngeal squamous cell carcinoma). DHA can prevent IL-6-induced EMT and invasion in LSCC by increasing miR-130b-3p expression, which reduces activation of the IL-6/STAT3 and β-catenin signaling pathways [121]. DHA also inhibited B16F10 melanoma cell growth in vitro as well as in vivo in BALB/C mice. In addition, DHA dose-dependently prevented melanoma invasion, migration, and colony formation. In the tumor and spleen, DHA promoted the level of expression of IFN-γ and decreased the levels of IL-10 and IL-6. Furthermore, by controlling the STAT3 pathway, DHA treatment significantly accelerated the mitochondrial apoptosis of melanoma [122].

4.4.4. Parthenolide

Parthenolide (PN) is a sesquiterpene lactone obtained from feverfew (Tanacetum parthenium), a medicinal herb. It has an α-methylene-γ-lactone ring and an epoxide group that can interact with the nucleophilic sites of biological molecules. Due to its comprehensive evaluation, PN has been used as an herbal medicine for millennia because of its anti-inflammatory and anti-migraine qualities. The potential of PN as an anticancer agent has garnered significant interest recently [159]. PN has been shown to suppress the STAT3 signaling pathway in SGC-7901/DPP gastric cancer cells, resulting in apoptosis and enhanced drug chemosensitivity. PN treatment inhibited cancer cell growth by inhibiting STAT3-mediated production of Bax, Bcl-2, Bcl-xL, and cyclin D1 [123]. PN was also found to be a potent inhibitor of JAKs in HepG2 liver cancer cells and MDA-MB-231 breast cancer cells. It inhibited the kinase activity of JAK2 by covalently altering its Cys178, Cys243, Cys335, and Cys480 residues. It also had comparable interactions with other JAKs. Moreover, PN suppressed the IL-6-induced migration of cancer cells and specifically halted the development of tumors expressing a constitutively active STAT3 [124].

4.4.5. γ-Tocotrienol

γ-Tocotrienol, a vitamin E superfamily member obtained from rice bran and palm oil, has garnered significant interest due to its putative anticarcinogenic and antiproliferative properties [160]. γ-Tocotrienol reduced both constitutive and inducible STAT3 activation while having no effect on STAT5. γ-Tocotrienol also prevented the activation of JAK1, JAK2, and Src, which are involved in the activation of STAT3. The down-regulation of STAT3 mediated by γ-tocotrienol was reversed by pervanadate, indicating the participation of a protein tyrosine phosphatase. In fact, γ-tocotrienol promoted the expression of a protein tyrosine phosphatase SHP-1, and its capacity to prevent STAT3 activation was eliminated upon SHP-1 gene deletion via small interfering RNA. Additionally, γ-tocotrienol suppressed the expression of gene products controlled by STAT3, such as cyclin D1, Bcl-2, Bcl-xL, survivin, Mcl-1, and VEGF. Lastly, γ-tocotrienol reduced proliferation, caused apoptosis, and significantly increased the apoptotic effects of chemotherapeutic drugs (paclitaxel and doxorubicin) in hepatocellular carcinoma HepG2 cells [125].

5. Conclusions and Future Perspectives

Although STAT3 expression is tightly controlled in normal cells, it is constitutively activated in various malignancies. The abnormal activation of STAT3 promotes tumor metastasis by promoting tumor cell proliferation, angiogenesis, migration, and invasion. Furthermore, evading immune surveillance is significantly aided by the activation of STAT3 signaling. Since numerous types of cancer are linked to STAT3 expression and function, natural phytocompounds have made this protein their main therapeutic target. Several epidemiological and preclinical investigations have explored the relationship between dietary terpenoids and the suppression of the transcription factor STAT3 in various cancer cell lines. Several terpenoid family chemicals have demonstrated strong evidence of their chemopreventive and chemoprotective properties. Among them, some compounds directly target STAT3, whereas others indirectly reduce its function. Furthermore, malignant cells with abnormal STAT3 activation exhibit increased cellular invasion, migration, differentiation, and growth. Natural terpenes have been thoroughly studied for their ability to inhibit the signaling arrays of unregulated STAT3 in cancer cells. With regard to these sixteen drugs, there is still a lack of information about antitumor potential by targeting the STAT3 signaling cascade in animal models, particularly in humans. In order to corroborate and clinically establish the chemopreventive nature of terpenoids, it can be recommended for additional studies including both humans and animals to be carried out. Based on this review, we suggest further research to identify more natural polyphenol candidates that can efficiently target STAT3. We predict that this review will provide new insight into future research on plant-derived terpenoids to treat various cancers by targeting STAT3 pathways.

Author Contributions

F.K.: conceptualization, writing—original draft preparation, resources, writing—review and editing. P.P.: conceptualization, writing—original draft preparation, resources. M.V.: formal analysis, writing—review and editing, visualization, supervision, project administration. T.K.U.: formal analysis, writing—review and editing, visualization, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Faubert, B.; Solmonson, A.; DeBerardinis, R.J. Metabolic reprogramming and cancer progression. Science 2020, 368, eaaw5473. [Google Scholar] [CrossRef] [PubMed]
  2. Guo, Z.; Li, Z.; Zhang, M.; Bao, M.; He, B.; Zhou, X. LncRNA FAS-AS1 upregulated by its genetic variation rs6586163 promotes cell apoptosis in nasopharyngeal carcinoma through regulating mitochondria function and Fas splicing. Sci. Rep. 2023, 13, 8218. [Google Scholar] [CrossRef] [PubMed]
  3. Taga, T.; Hibi, M.; Hirata, Y.; Yamasaki, K.; Yasukawa, K.; Matsuda, T.; Hirano, T.; Kishimoto, T. Interleukin-6 triggers the association of its receptor with a possible signal transducer, gp130. Cell 1989, 58, 573–581. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, X.; Li, J.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT signaling pathway: From bench to clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef] [PubMed]
  5. Yu, H.; Jove, R. The STATs of cancer--new molecular targets come of age. Nat. Rev. Cancer 2004, 4, 97–105. [Google Scholar] [CrossRef]
  6. Molenda, S.; Sikorska, A.; Florczak, A.; Lorenc, P.; Dams-Kozlowska, H. Oligonucleotide-Based Therapeutics for STAT3 Targeting in Cancer-Drug Carriers Matter. Cancers 2023, 15, 5647. [Google Scholar] [CrossRef]
  7. Herrmann, A.; Kortylewski, M.; Kujawski, M.; Zhang, C.; Reckamp, K.; Armstrong, B.; Wang, L.; Kowolik, C.; Deng, J.; Figlin, R.; et al. Targeting Stat3 in the myeloid compartment drastically improves the in vivo antitumor functions of adoptively transferred T cells. Cancer Res. 2010, 70, 7455–7464. [Google Scholar] [CrossRef]
  8. Wang, Y.; Shen, Y.; Wang, S.; Shen, Q.; Zhou, X. The role of STAT3 in leading the crosstalk between human cancers and the immune system. Cancer Lett. 2018, 415, 117–128. [Google Scholar] [CrossRef]
  9. Chen, R.Y.; Yen, C.J.; Liu, Y.W.; Guo, C.G.; Weng, C.Y.; Lai, C.H.; Wang, J.M.; Lin, Y.J.; Hung, L.Y. CPAP promotes angiogenesis and metastasis by enhancing STAT3 activity. Cell Death Differ. 2020, 27, 1259–1273. [Google Scholar] [CrossRef]
  10. Shen, Y.; Wang, X.; Liu, Y.; Singhal, M.; Gürkaşlar, C.; Valls, A.F.; Lei, Y.; Hu, W.; Schermann, G.; Adler, H.; et al. STAT3-YAP/TAZ signaling in endothelial cells promotes tumor angiogenesis. Sci. Signal. 2021, 14, eabj8393. [Google Scholar] [CrossRef]
  11. Lee, H.; Zhang, P.; Herrmann, A.; Yang, C.; Xin, H.; Wang, Z.; Hoon, D.S.; Forman, S.J.; Jove, R.; Riggs, A.D.; et al. Acetylated STAT3 is crucial for methylation of tumor-suppressor gene promoters and inhibition by resveratrol results in demethylation. Proc. Natl. Acad. Sci. USA 2012, 109, 7765–7769. [Google Scholar] [CrossRef] [PubMed]
  12. Tolomeo, M.; Cascio, A. The Multifaced Role of STAT3 in Cancer and Its Implication for Anticancer Therapy. Int. J. Mol. Sci. 2021, 22, 603. [Google Scholar] [CrossRef]
  13. Hossain, D.M.; Dos Santos, C.; Zhang, Q.; Kozlowska, A.; Liu, H.; Gao, C.; Moreira, D.; Swiderski, P.; Jozwiak, A.; Kline, J.; et al. Leukemia cell-targeted STAT3 silencing and TLR9 triggering generate systemic antitumor immunity. Blood 2014, 123, 15–25. [Google Scholar] [CrossRef] [PubMed]
  14. Eyking, A.; Ey, B.; Rünzi, M.; Roig, A.I.; Reis, H.; Schmid, K.W.; Gerken, G.; Podolsky, D.K.; Cario, E. Toll-like receptor 4 variant D299G induces features of neoplastic progression in Caco-2 intestinal cells and is associated with advanced human colon cancer. Gastroenterology 2011, 141, 2154–2165. [Google Scholar] [CrossRef] [PubMed]
  15. Tye, H.; Kennedy, C.L.; Najdovska, M.; McLeod, L.; McCormack, W.; Hughes, N.; Dev, A.; Sievert, W.; Ooi, C.H.; Ishikawa, T.O.; et al. STAT3-driven upregulation of TLR2 promotes gastric tumorigenesis independent of tumor inflammation. Cancer Cell 2012, 22, 466–478. [Google Scholar] [CrossRef] [PubMed]
  16. Fehri, E.; Ennaifer, E.; Bel Haj Rhouma, R.; Ardhaoui, M.; Boubaker, S. TLR9 and Glioma: Friends or Foes? Cells 2022, 12, 152. [Google Scholar] [CrossRef]
  17. Li, L.; Wang, S.; Zhou, W. Balance Cell Apoptosis and Pyroptosis of Caspase-3-Activating Chemotherapy for Better Antitumor Therapy. Cancers 2022, 15, 26. [Google Scholar] [CrossRef] [PubMed]
  18. Benvenuto, M.; Albonici, L.; Focaccetti, C.; Ciuffa, S.; Fazi, S.; Cifaldi, L.; Miele, M.T.; De Maio, F.; Tresoldi, I.; Manzari, V.; et al. Polyphenol-Mediated Autophagy in Cancer: Evidence of In Vitro and In Vivo Studies. Int. J. Mol. Sci. 2020, 21, 6635. [Google Scholar] [CrossRef]
  19. Pandey, P.; Khan, F. A mechanistic review of the anticancer potential of hesperidin, a natural flavonoid from citrus fruits. Nutr. Res. 2021, 92, 21–31. [Google Scholar] [CrossRef]
  20. Muhammad, N.; Usmani, D.; Tarique, M.; Naz, H.; Ashraf, M.; Raliya, R.; Tabrez, S.; Zughaibi, T.A.; Alsaieedi, A.; Hakeem, I.J.; et al. The Role of Natural Products and Their Multitargeted Approach to Treat Solid Cancer. Cells 2022, 11, 2209. [Google Scholar] [CrossRef]
  21. Rahman, M.M.; Sarker, M.T.; Alam Tumpa, M.A.; Yamin, M.; Islam, T.; Park, M.N.; Islam, M.R.; Rauf, A.; Sharma, R.; Cavalu, S.; et al. Exploring the recent trends in perturbing the cellular signaling pathways in cancer by natural products. Front. Pharmacol. 2022, 13, 950109. [Google Scholar] [CrossRef]
  22. Masyita, A.; Mustika Sari, R.; Dwi Astuti, A.; Yasir, B.; Rahma Rumata, N.; Emran, T.B.; Nainu, F.; Simal-Gandara, J. Terpenes and terpenoids as main bioactive compounds of essential oils, their roles in human health and potential application as natural food preservatives. Food Chem. X 2022, 13, 100217. [Google Scholar] [CrossRef]
  23. Ge, J.; Liu, Z.; Zhong, Z.; Wang, L.; Zhuo, X.; Li, J.; Jiang, X.; Ye, X.Y.; Xie, T.; Bai, R. Natural terpenoids with anti-inflammatory activities: Potential leads for anti-inflammatory drug discovery. Bioorg. Chem. 2022, 124, 105817. [Google Scholar] [CrossRef] [PubMed]
  24. Saini, R.K.; Keum, Y.S.; Daglia, M.; Rengasamy, K.R. Dietary carotenoids in cancer chemoprevention and chemotherapy: A review of emerging evidence. Pharmacol. Res. 2020, 157, 104830. [Google Scholar] [CrossRef] [PubMed]
  25. Delgado-Gonzalez, P.; Garza-Treviño, E.N.; de la Garza Kalife, D.A.; Quiroz Reyes, A.; Hernández-Tobías, E.A. Bioactive Compounds of Dietary Origin and Their Influence on Colorectal Cancer as Chemoprevention. Life 2023, 13, 1977. [Google Scholar] [CrossRef] [PubMed]
  26. Sporn, M.B.; Suh, N. Chemoprevention: An essential approach to controlling cancer. Nat. Rev. Cancer 2002, 2, 537–543. [Google Scholar] [CrossRef]
  27. Steward, W.P.; Brown, K. Cancer chemoprevention: A rapidly evolving field. Br. J. Cancer 2013, 109, 1–7. [Google Scholar] [CrossRef]
  28. Cai, Y.; Luo, Q.; Sun, M.; Corke, H. Antioxidant activity and phenolic compounds of 112 traditional Chinese medicinal plants associated with anticancer. Life Sci. 2004, 74, 2157–2184. [Google Scholar] [CrossRef]
  29. Withers, S.T.; Keasling, J.D. Biosynthesis and engineering of isoprenoid small molecules. Appl. Microbiol. Biotechnol. 2007, 73, 980–990. [Google Scholar] [CrossRef]
  30. Rabi, T.; Bishayee, A. Terpenoids and breast cancer chemoprevention. Breast Cancer Res. Treat. 2009, 115, 223–239. [Google Scholar] [CrossRef]
  31. Sheikh, I.; Sharma, V.; Tuli, H.S.; Aggarwal, D.; Sankhyan, A.; Vyas, P.; Sharma, A.K.; Bishayee, A. Cancer chemoprevention by flavonoids, dietary polyphenols and terpenoids. Biointerface Res. Appl. Chem. 2020, 11, 8502–8537. [Google Scholar] [CrossRef]
  32. Chen, Y.; Zhu, Z.; Chen, J.; Zheng, Y.; Limsila, B.; Lu, M.; Gao, T.; Yang, Q.; Fu, C.; Liao, W. Terpenoids from Curcumae Rhizoma: Their anticancer effects and clinical uses on combination and versus drug therapies. Biomed. Pharmacother. 2021, 138, 111350. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, S.; Zhao, Y.; Shen, F.; Long, D.; Yu, T.; Lin, X. Introduction of exogenous wild-type p53 mediates the regulation of oncoprotein 18/stathmin signaling via nuclear factor-κB in non-small cell lung cancer NCI-H1299 cells. Oncol. Rep. 2019, 41, 2051–2059. [Google Scholar] [CrossRef]
  34. Xue, C.; Yao, Q.; Gu, X.; Shi, Q.; Yuan, X.; Chu, Q.; Bao, Z.; Lu, J.; Li, L. Evolving cognition of the JAK-STAT signaling pathway: Autoimmune disorders and cancer. Signal. Transduct. Target. Ther. 2023, 8, 204. [Google Scholar] [CrossRef]
  35. Kwon, S. Molecular dissection of Janus kinases as drug targets for inflammatory diseases. Front. Immunol. 2022, 13, 1075192. [Google Scholar] [CrossRef]
  36. Turkson, J. STAT proteins as novel targets for cancer drug discovery. Expert. Opin. Ther. Targets 2004, 8, 409–422. [Google Scholar] [CrossRef] [PubMed]
  37. Tuli, H.S.; Sak, K.; Iqubal, A.; Garg, V.K.; Varol, M.; Sharma, U.; Chauhan, A.; Yerer, M.B.; Dhama, K.; Jain, M.; et al. STAT signaling as a target for intervention: From cancer inflammation and angiogenesis to non-coding RNAs modulation. Mol. Biol. Rep. 2022, 49, 8987–8999. [Google Scholar] [CrossRef]
  38. Rusek, M.; Smith, J.; El-Khatib, K.; Aikins, K.; Czuczwar, S.J.; Pluta, R. The Role of the JAK/STAT Signaling Pathway in the Pathogenesis of Alzheimer’s Disease: New Potential Treatment Target. Int. J. Mol. Sci. 2023, 24, 864. [Google Scholar] [CrossRef] [PubMed]
  39. Villarino, A.V.; Kanno, Y.; O’Shea, J.J. Mechanisms and consequences of Jak-STAT signaling in the immune system. Nat. Immunol. 2017, 18, 374–384. [Google Scholar] [CrossRef]
  40. Banerjee, S.; Biehl, A.; Gadina, M.; Hasni, S.; Schwartz, D.M. JAK-STAT Signaling as a Target for Inflammatory and Autoimmune Diseases: Current and Future Prospects. Drugs 2017, 77, 521–546, Erratum in Drugs 2017, 77, 939. [Google Scholar] [CrossRef]
  41. Xin, P.; Xu, X.; Deng, C.; Liu, S.; Wang, Y.; Zhou, X.; Ma, H.; Wei, D.; Sun, S. The role of JAK/STAT signaling pathway and its inhibitors in diseases. Int. Immunopharmacol. 2020, 80, 106210. [Google Scholar] [CrossRef]
  42. Gao, Q.; Liang, X.; Shaikh, A.S.; Zang, J.; Xu, W.; Zhang, Y. JAK/STAT Signal Transduction: Promising Attractive Targets for Immune, Inflammatory and Hematopoietic Diseases. Curr. Drug Targets 2018, 19, 487–500. [Google Scholar] [CrossRef]
  43. El-Tanani, M.; Al Khatib, A.O.; Aladwan, S.M.; Abuelhana, A.; McCarron, P.A.; Tambuwala, M.M. Importance of STAT3 signalling in cancer, metastasis and therapeutic interventions. Cell. Signal. 2022, 92, 110275. [Google Scholar] [CrossRef]
  44. Takeda, J.; Yoshida, K.; Nakagawa, M.M.; Nannya, Y.; Yoda, A.; Morishita, D.; Saiki, R.; Chiba, K.; Tanaka, H.; Shiraishi, Y.; et al. EPOR/JAK/STAT signaling pathway as therapeutic target of acute erythroid leukemia. Blood 2021, 138, 610. [Google Scholar] [CrossRef]
  45. Arshad, S.; Naveed, M.; Ullia, M.; Javed, K.; Butt, A.; Khawar, M.; Amjad, F. Targeting STAT-3 signaling pathway in cancer for development of novel drugs: Advancements and challenges. Genet. Mol. Biol. 2020, 43, e20180160. [Google Scholar] [CrossRef] [PubMed]
  46. Aigner, P.; Just, V.; Stoiber, D. STAT3 isoforms: Alternative fates in cancer? Cytokine 2019, 118, 27–34. [Google Scholar] [CrossRef]
  47. Singh, A.; Settleman, J. EMT, cancer stem cells and drug resistance: An emerging axis of evil in the war on cancer. Oncogene 2010, 29, 4741–4751. [Google Scholar] [CrossRef]
  48. Liu, W.H.; Chen, M.T.; Wang, M.L.; Lee, Y.Y.; Chiou, G.Y.; Chien, C.S.; Huang, P.I.; Chen, Y.W.; Huang, M.C.; Chiou, S.H.; et al. Cisplatin-selected resistance is associated with increased motility and stem-like properties via activation of STAT3/Snail axis in atypical teratoid/rhabdoid tumor cells. Oncotarget 2015, 6, 1750–1768. [Google Scholar] [CrossRef]
  49. Lin, J.C.; Tsai, J.T.; Chao, T.Y.; Ma, H.I.; Liu, W.H. The STAT3/Slug Axis Enhances Radiation-Induced Tumor Invasion and Cancer Stem-like Properties in Radioresistant Glioblastoma. Cancers 2018, 10, 512. [Google Scholar] [CrossRef] [PubMed]
  50. Shih, P.C.; Mei, K.C. Role of STAT3 signaling transduction pathways in cancer stem cell-associated chemoresistance. Drug Discov. Today 2021, 26, 1450–1458. [Google Scholar] [CrossRef] [PubMed]
  51. Tewari, D.; Patni, P.; Bishayee, A.; Sah, A.N.; Bishayee, A. Natural products targeting the PI3K-Akt-mTOR signaling pathway in cancer: A novel therapeutic strategy. Semin. Cancer Biol. 2022, 80, 1–17. [Google Scholar] [CrossRef]
  52. Huang, A.; Zhou, W. Mn-based cGAS-STING activation for tumor therapy. Chin. J. Cancer Res. 2023, 35, 19–43. [Google Scholar] [CrossRef]
  53. Yang, J.; Ren, B.; Yang, G.; Wang, H.; Chen, G.; You, L.; Zhang, T.; Zhao, Y. The enhancement of glycolysis regulates pancreatic cancer metastasis. Cell. Mol. Life Sci. 2020, 77, 305–321. [Google Scholar] [CrossRef]
  54. Park, J.H.; Pyun, W.Y.; Park, H.W. Cancer Metabolism: Phenotype, Signaling and Therapeutic Targets. Cells 2020, 9, 2308. [Google Scholar] [CrossRef]
  55. Wang, Z.; Hui, C.; Xie, Y. Natural STAT3 inhibitors: A mini perspective. Bioorg. Chem. 2021, 115, 105169. [Google Scholar] [CrossRef]
  56. Mohan, C.D.; Rangappa, S.; Preetham, H.D.; Chandra Nayaka, S.; Gupta, V.K.; Basappa, S.; Sethi, G.; Rangappa, K.S. Targeting STAT3 signaling pathway in cancer by agents derived from Mother Nature. Semin. Cancer Biol. 2022, 80, 157–182. [Google Scholar] [CrossRef] [PubMed]
  57. Siveen, K.S.; Sikka, S.; Surana, R.; Dai, X.; Zhang, J.; Kumar, A.P.; Tan, B.K.; Sethi, G.; Bishayee, A. Targeting the STAT3 signaling pathway in cancer: Role of synthetic and natural inhibitors. Biochim. Biophys. Acta 2014, 1845, 136–154. [Google Scholar] [CrossRef] [PubMed]
  58. Bose, S.; Banerjee, S.; Mondal, A.; Chakraborty, U.; Pumarol, J.; Croley, C.R.; Bishayee, A. Targeting the JAK/STAT Signaling Pathway Using Phytocompounds for Cancer Prevention and Therapy. Cells 2020, 9, 1451. [Google Scholar] [CrossRef] [PubMed]
  59. Yang, J.; Wang, L.; Guan, X.; Qin, J.J. Inhibiting STAT3 signaling pathway by natural products for cancer prevention and therapy: In vitro and in vivo activity and mechanisms of action. Pharmacol. Res. 2022, 182, 106357. [Google Scholar] [CrossRef] [PubMed]
  60. El-Baba, C.; Baassiri, A.; Kiriako, G.; Dia, B.; Fadlallah, S.; Moodad, S.; Darwiche, N. Terpenoids’ anti-cancer effects: Focus on autophagy. Apoptosis 2021, 26, 491–511. [Google Scholar] [CrossRef] [PubMed]
  61. Zhu, W.Q.; Wang, J.; Guo, X.F.; Liu, Z.; Dong, W.G. Thymoquinone inhibits proliferation in gastric cancer via the STAT3 pathway in vivo and in vitro. World J. Gastroenterol. 2016, 22, 4149–4159. [Google Scholar] [CrossRef] [PubMed]
  62. Kundu, J.; Choi, B.Y.; Jeong, C.H.; Kundu, J.K.; Chun, K.S. Thymoquinone induces apoptosis in human colon cancer HCT116 cells through inactivation of STAT3 by blocking JAK2- and Src mediated phosphorylation of EGF receptor tyrosine kinase. Oncol. Rep. 2014, 32, 821–828. [Google Scholar] [CrossRef] [PubMed]
  63. Al-Rawashde, F.A.; Wan Taib, W.R.; Ismail, I.; Johan, M.F.; Al-Wajeeh, A.S.; Al-Jamal, H.A.N. Thymoquinone Induces Downregulation of BCR-ABL/JAK/STAT Pathway and Apoptosis in K562 Leukemia Cells. Asian Pac. J. Cancer Prev. 2021, 22, 3959–3965. [Google Scholar] [CrossRef]
  64. Almajali, B.; Johan, M.F.; Al-Wajeeh, A.S.; Wan Taib, W.R.; Ismail, I.; Alhawamdeh, M.; Al-Tawarah, N.M.; Ibrahim, W.N.; Al-Rawashde, F.A.; Al-Jamal, H.A.N. Gene Expression Profiling and Protein Analysis Reveal Suppression of the C-Myc Oncogene and Inhibition JAK/STAT and PI3K/AKT/mTOR Signaling by Thymoquinone in Acute Myeloid Leukemia Cells. Pharmaceuticals 2022, 15, 307. [Google Scholar] [CrossRef] [PubMed]
  65. Atteia, H.H.; Arafa, M.H.; Mohammad, N.S.; Amin, D.M.; Sakr, A.T. Thymoquinone upregulates miR-125a-5p, attenuates STAT3 activation, and potentiates doxorubicin antitumor activity in murine solid Ehrlich carcinoma. J. Biochem. Mol. Toxicol. 2021, 35, e22924. [Google Scholar] [CrossRef] [PubMed]
  66. Raut, P.K.; Lee, H.S.; Joo, S.H.; Chun, K.S. Thymoquinone induces oxidative stress-mediated apoptosis through downregulation of Jak2/STAT3 signaling pathway in human melanoma cells. Food Chem. Toxicol. 2021, 157, 112604. [Google Scholar] [CrossRef]
  67. Chae, I.G.; Song, N.Y.; Kim, D.H.; Lee, M.Y.; Park, J.M.; Chun, K.S. Thymoquinone induces apoptosis of human renal carcinoma Caki-1 cells by inhibiting JAK2/STAT3 through pro-oxidant effect. Food Chem. Toxicol. 2020, 139, 111253. [Google Scholar] [CrossRef]
  68. Park, J.E.; Kim, D.H.; Ha, E.; Choi, S.M.; Choi, J.S.; Chun, K.S.; Joo, S.H. Thymoquinone induces apoptosis of human epidermoid carcinoma A431 cells through ROS-mediated suppression of STAT3. Chem. Biol. Interact. 2019, 312, 108799. [Google Scholar] [CrossRef]
  69. Li, F.; Rajendran, P.; Sethi, G. Thymoquinone inhibits proliferation, induces apoptosis and chemosensitizes human multiple myeloma cells through suppression of signal transducer and activator of transcription 3 activation pathway. Br. J. Pharmacol. 2010, 161, 541–554. [Google Scholar] [CrossRef]
  70. Wang, X.R.; Jiang, Z.B.; Xu, C.; Meng, W.Y.; Liu, P.; Zhang, Y.Z.; Xie, C.; Xu, J.Y.; Xie, Y.J.; Liang, T.L.; et al. Andrographolide suppresses non-small-cell lung cancer progression through induction of autophagy and antitumor immune response. Pharmacol. Res. 2022, 179, 106198. [Google Scholar] [CrossRef]
  71. Zhou, J.; Ong, C.N.; Hur, G.M.; Shen, H.M. Inhibition of the JAK-STAT3 pathway by andrographolide enhances chemosensitivity of cancer cells to doxorubicin. Biochem. Pharmacol. 2010, 79, 1242–1250. [Google Scholar] [CrossRef]
  72. Shin, D.S.; Kim, H.N.; Shin, K.D.; Yoon, Y.J.; Kim, S.J.; Han, D.C.; Kwon, B.M. Cryptotanshinone inhibits constitutive signal transducer and activator of transcription 3 function through blocking the dimerization in DU145 prostate cancer cells. Cancer Res. 2009, 69, 193–202. [Google Scholar] [CrossRef]
  73. Yang, Y.; Cao, Y.; Chen, L.; Liu, F.; Qi, Z.; Cheng, X.; Wang, Z. Cryptotanshinone suppresses cell proliferation and glucose metabolism via STAT3/SIRT3 signaling pathway in ovarian cancer cells. Cancer Med. 2018, 7, 4610–4618. [Google Scholar] [CrossRef] [PubMed]
  74. Ge, Y.; Yang, B.; Chen, Z.; Cheng, R. Cryptotanshinone suppresses the proliferation and induces the apoptosis of pancreatic cancer cells via the STAT3 signaling pathway. Mol. Med. Rep. 2015, 12, 7782–7788. [Google Scholar] [CrossRef] [PubMed]
  75. Ji, Y.; Liu, Y.; Xue, N.; Du, T.; Wang, L.; Huang, R.; Li, L.; Yan, C.; Chen, X. Cryptotanshinone inhibits esophageal squamous-cell carcinoma in vitro and in vivo through the suppression of STAT3 activation. OncoTargets Ther. 2019, 12, 883–896. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, Z.; Zhu, R.; Zheng, J.; Chen, C.; Huang, C.; Ma, J.; Xu, C.; Zhai, W.; Zheng, J. Cryptotanshinone inhibits proliferation yet induces apoptosis by suppressing STAT3 signals in renal cell carcinoma. Oncotarget 2017, 8, 50023–50033. [Google Scholar] [CrossRef] [PubMed]
  77. Lu, L.; Li, C.; Li, D.; Wang, Y.; Zhou, C.; Shao, W.; Peng, J.; You, Y.; Zhang, X.; Shen, X. Cryptotanshinone inhibits human glioma cell proliferation by suppressing STAT3 signaling. Mol. Cell. Biochem. 2013, 381, 273–282. [Google Scholar] [CrossRef] [PubMed]
  78. Lu, L.; Zhang, S.; Li, C.; Zhou, C.; Li, D.; Liu, P.; Huang, M.; Shen, X. Cryptotanshinone inhibits human glioma cell proliferation in vitro and in vivo through SHP-2-dependent inhibition of STAT3 activation. Cell Death Dis. 2017, 8, e2767. [Google Scholar] [CrossRef]
  79. Cheng, R.; Huang, Y.; Fang, Y.; Wang, Q.; Yan, M.; Ge, Y. Cryptotanshinone enhances the efficacy of Bcr-Abl tyrosine kinase inhibitors via inhibiting STAT3 and eIF4E signalling pathways in chronic myeloid leukaemia. Pharm. Biol. 2021, 59, 893–903. [Google Scholar] [CrossRef]
  80. Dong, B.; Liang, Z.; Chen, Z.; Li, B.; Zheng, L.; Yang, J.; Zhou, H.; Qu, L. Cryptotanshinone suppresses key onco-proliferative and drug-resistant pathways of chronic myeloid leukemia by targeting STAT5 and STAT3 phosphorylation. Sci. China Life Sci. 2018, 61, 999–1009. [Google Scholar] [CrossRef]
  81. Liu, W.; Wang, X.; Wang, L.; Mei, Y.; Yun, Y.; Yao, X.; Chen, Q.; Zhou, J.; Kou, B. Oridonin represses epithelial-mesenchymal transition and angiogenesis of thyroid cancer via downregulating JAK2/STAT3 signaling. Int. J. Med. Sci. 2022, 19, 965–974. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, W.; Huang, G.; Yang, Y.; Gao, R.; Zhang, S.; Kou, B. Oridonin inhibits epithelial-mesenchymal transition of human nasopharyngeal carcinoma cells by negatively regulating AKT/STAT3 signaling pathway. Int. J. Med. Sci. 2021, 18, 81–87. [Google Scholar] [CrossRef] [PubMed]
  83. Du, Y.; Zhang, J.; Yan, S.; Tao, Z.; Wang, C.; Huang, M.; Zhang, X. Oridonin inhibits the proliferation, migration and invasion of human osteosarcoma cells via suppression of matrix metalloproteinase expression and STAT3 signalling pathway. J. Buon 2019, 24, 1175–1180. [Google Scholar] [PubMed]
  84. Zhou, J.; Hou, J.; Wang, J.; Wang, J.; Gao, J.; Bai, Y. Brusatol inhibits laryngeal cancer cell proliferation and metastasis via abrogating JAK2/STAT3 signaling mediated epithelial-mesenchymal transition. Life Sci. 2021, 284, 119907. [Google Scholar] [CrossRef] [PubMed]
  85. Lee, J.H.; Rangappa, S.; Mohan, C.D.; Basappa; Sethi, G.; Lin, Z.X.; Rangappa, K.S.; Ahn, K.S. Brusatol, a Nrf2 Inhibitor Targets STAT3 Signaling Cascade in Head and Neck Squamous Cell Carcinoma. Biomolecules 2019, 9, 550. [Google Scholar] [CrossRef] [PubMed]
  86. Xiang, Y.; Ye, W.; Huang, C.; Lou, B.; Zhang, J.; Yu, D.; Huang, X.; Chen, B.; Zhou, M. Brusatol inhibits growth and induces apoptosis in pancreatic cancer cells via JNK/p38 MAPK/NF-κb/Stat3/Bcl-2 signaling pathway. Biochem. Biophys. Res. Commun. 2017, 487, 820–826. [Google Scholar] [CrossRef] [PubMed]
  87. Lee, J.H.; Mohan, C.D.; Deivasigamani, A.; Jung, Y.Y.; Rangappa, S.; Basappa, S.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Garg, M.; et al. Brusatol suppresses STAT3-driven metastasis by downregulating epithelial-mesenchymal transition in hepatocellular carcinoma. J. Adv. Res. 2020, 26, 83–94. [Google Scholar] [CrossRef] [PubMed]
  88. Pandey, M.K.; Sung, B.; Aggarwal, B.B. Betulinic acid suppresses STAT3 activation pathway through induction of protein tyrosine phosphatase SHP-1 in human multiple myeloma cells. Int. J. Cancer 2010, 127, 282–292. [Google Scholar] [CrossRef]
  89. Shin, J.; Lee, H.J.; Jung, D.B.; Jung, J.H.; Lee, H.J.; Lee, E.O.; Lee, S.G.; Shim, B.S.; Choi, S.H.; Ko, S.G.; et al. Suppression of STAT3 and HIF-1 alpha mediates anti-angiogenic activity of betulinic acid in hypoxic PC-3 prostate cancer cells. PLoS ONE 2011, 6, e21492. [Google Scholar] [CrossRef]
  90. Rajendran, P.; Li, F.; Shanmugam, M.K.; Kannaiyan, R.; Goh, J.N.; Wong, K.F.; Wang, W.; Khin, E.; Tergaonkar, V.; Kumar, A.P.; et al. Celastrol suppresses growth and induces apoptosis of human hepatocellular carcinoma through the modulation of STAT3/JAK2 signaling cascade in vitro and in vivo. Cancer Prev. Res. 2012, 5, 631–643. [Google Scholar] [CrossRef]
  91. Zhao, Z.; Wang, Y.; Gong, Y.; Wang, X.; Zhang, L.; Zhao, H.; Li, J.; Zhu, J.; Huang, X.; Zhao, C.; et al. Celastrol elicits antitumor effects by inhibiting the STAT3 pathway through ROS accumulation in non-small cell lung cancer. J. Transl. Med. 2022, 20, 525. [Google Scholar] [CrossRef]
  92. Kannaiyan, R.; Hay, H.S.; Rajendran, P.; Li, F.; Shanmugam, M.K.; Vali, S.; Abbasi, T.; Kapoor, S.; Sharma, A.; Kumar, A.P.; et al. Celastrol inhibits proliferation and induces chemosensitization through down-regulation of NF-κB and STAT3 regulated gene products in multiple myeloma cells. Br. J. Pharmacol. 2011, 164, 1506–1521, Erratum in Br. J. Pharmacol. 2012, 165, 540. [Google Scholar] [CrossRef]
  93. Thoennissen, N.H.; Iwanski, G.B.; Doan, N.B.; Okamoto, R.; Lin, P.; Abbassi, S.; Song, J.H.; Yin, D.; Toh, M.; Xie, W.D.; et al. Cucurbitacin B induces apoptosis by inhibition of the JAK/STAT pathway and potentiates antiproliferative effects of gemcitabine on pancreatic cancer cells. Cancer Res. 2009, 69, 5876–5884. [Google Scholar] [CrossRef]
  94. Zhang, M.; Sun, C.; Shan, X.; Yang, X.; Li-Ling, J.; Deng, Y. Inhibition of pancreatic cancer cell growth by cucurbitacin B through modulation of signal transducer and activator of transcription 3 signaling. Pancreas 2010, 39, 923–929. [Google Scholar] [CrossRef]
  95. Yar Saglam, A.S.; Alp, E.; Elmazoglu, Z.; Menevse, S. Treatment with cucurbitacin B alone and in combination with gefitinib induces cell cycle inhibition and apoptosis via EGFR and JAK/STAT pathway in human colorectal cancer cell lines. Hum. Exp. Toxicol. 2016, 35, 526–543. [Google Scholar] [CrossRef]
  96. Zhang, M.; Zhang, H.; Sun, C.; Shan, X.; Yang, X.; Li-Ling, J.; Deng, Y. Targeted constitutive activation of signal transducer and activator of transcription 3 in human hepatocellular carcinoma cells by cucurbitacin B. Cancer Chemother. Pharmacol. 2009, 63, 635–642. [Google Scholar] [CrossRef]
  97. Xie, Y.L.; Tao, W.H.; Yang, T.X.; Qiao, J.G. Anticancer effect of cucurbitacin B on MKN-45 cells via inhibition of the JAK2/STAT3 signaling pathway. Exp. Ther. Med. 2016, 12, 2709–2715. [Google Scholar] [CrossRef] [PubMed]
  98. Xu, J.; Chen, Y.; Yang, R.; Zhou, T.; Ke, W.; Si, Y.; Yang, S.; Zhang, T.; Liu, X.; Zhang, L.; et al. Cucurbitacin B inhibits gastric cancer progression by suppressing STAT3 activity. Arch. Biochem. Biophys. 2020, 684, 108314. [Google Scholar] [CrossRef] [PubMed]
  99. Sun, C.; Zhang, M.; Shan, X.; Zhou, X.; Yang, J.; Wang, Y.; Li-Ling, J.; Deng, Y. Inhibitory effect of cucurbitacin E on pancreatic cancer cells growth via STAT3 signaling. J. Cancer Res. Clin. Oncol. 2010, 136, 603–610. [Google Scholar] [CrossRef] [PubMed]
  100. Dong, Y.; Lu, B.; Zhang, X.; Zhang, J.; Lai, L.; Li, D.; Wu, Y.; Song, Y.; Luo, J.; Pang, X.; et al. Cucurbitacin E, a tetracyclic triterpenes compound from Chinese medicine, inhibits tumor angiogenesis through VEGFR2-mediated Jak2-STAT3 signaling pathway. Carcinogenesis 2010, 31, 2097–2104, Erratum in Carcinogenesis 2012, 33, 946. [Google Scholar] [CrossRef] [PubMed]
  101. Huang, W.W.; Yang, J.S.; Lin, M.W.; Chen, P.Y.; Chiou, S.M.; Chueh, F.S.; Lan, Y.H.; Pai, S.J.; Tsuzuki, M.; Ho, W.J.; et al. Cucurbitacin E Induces G2/M Phase Arrest through STAT3/p53/p21 Signaling and Provokes Apoptosis via Fas/CD95 and Mitochondria-Dependent Pathways in Human Bladder Cancer T24 Cells. Evid.-Based Complement. Altern. Med. 2012, 2012, 952762. [Google Scholar] [CrossRef]
  102. Xu, D.; Shen, H.; Tian, M.; Chen, W.; Zhang, X. Cucurbitacin I inhibits the proliferation of pancreatic cancer through the JAK2/STAT3 signalling pathway in vivo and in vitro. J. Cancer 2022, 13, 2050–2060. [Google Scholar] [CrossRef] [PubMed]
  103. Üremiş, N.; Üremiş, M.M.; Çiğremiş, Y.; Tosun, E.; Baysar, A.; Türköz, Y. Cucurbitacin I exhibits anticancer efficacy through induction of apoptosis and modulation of JAK/STAT3, MAPK/ERK, and AKT/mTOR signaling pathways in HepG2 cell line. J. Food Biochem. 2022, 46, e14333. [Google Scholar] [CrossRef]
  104. Ni, Y.; Wu, S.; Wang, X.; Zhu, G.; Chen, X.; Ding, Y.; Jiang, W. Cucurbitacin I induces pro-death autophagy in A549 cells via the ERK-mTOR-STAT3 signaling pathway. J. Cell. Biochem. 2018, 119, 6104–6112. [Google Scholar] [CrossRef] [PubMed]
  105. Sun, J.; Blaskovich, M.A.; Jove, R.; Livingston, S.K.; Coppola, D.; Sebti, S.M. Cucurbitacin Q: A selective STAT3 activation inhibitor with potent antitumor activity. Oncogene 2005, 24, 3236–3245, Erratum in Oncogene 2008, 27, 1344. [Google Scholar] [CrossRef]
  106. Su, Y.; Li, G.; Zhang, X.; Gu, J.; Zhang, C.; Tian, Z.; Zhang, J. JSI-124 inhibits glioblastoma multiforme cell proliferation through G(2)/M cell cycle arrest and apoptosis augment. Cancer Biol. Ther. 2008, 7, 1243–1249. [Google Scholar] [CrossRef]
  107. Shi, X.; Franko, B.; Frantz, C.; Amin, H.M.; Lai, R. JSI-124 (cucurbitacin I) inhibits Janus kinase-3/signal transducer and activator of transcription-3 signalling, downregulates nucleophosmin-anaplastic lymphoma kinase (ALK), and induces apoptosis in ALK-positive anaplastic large cell lymphoma cells. Br. J. Haematol. 2006, 135, 26–32. [Google Scholar] [CrossRef]
  108. Kang, D.Y.; Sp, N.; Lee, J.M.; Jang, K.J. Antitumor Effects of Ursolic Acid through Mediating the Inhibition of STAT3/PD-L1 Signaling in Non-Small Cell Lung Cancer Cells. Biomedicines 2021, 9, 297. [Google Scholar] [CrossRef] [PubMed]
  109. Kim, K.; Shin, E.A.; Jung, J.H.; Park, J.E.; Kim, D.S.; Shim, B.S.; Kim, S.H. Ursolic Acid Induces Apoptosis in Colorectal Cancer Cells Partially via Upregulation of MicroRNA-4500 and Inhibition of JAK2/STAT3 Phosphorylation. Int. J. Mol. Sci. 2018, 20, 114. [Google Scholar] [CrossRef]
  110. Liu, T.; Ma, H.; Shi, W.; Duan, J.; Wang, Y.; Zhang, C.; Li, C.; Lin, J.; Li, S.; Lv, J.; et al. Inhibition of STAT3 signaling pathway by ursolic acid suppresses growth of hepatocellular carcinoma. Int. J. Oncol. 2017, 51, 555–562. [Google Scholar] [CrossRef]
  111. Shanmugam, M.K.; Ong, T.H.; Kumar, A.P.; Lun, C.K.; Ho, P.C.; Wong, P.T.; Hui, K.M.; Sethi, G. Ursolic acid inhibits the initiation, progression of prostate cancer and prolongs the survival of TRAMP mice by modulating pro-inflammatory pathways. PLoS ONE 2012, 7, e32476. [Google Scholar] [CrossRef] [PubMed]
  112. Babaei, G.; Khadem Ansari, M.H.; Aziz, S.G.; Bazl, M.R. Alantolactone inhibits stem-like cell phenotype, chemoresistance and metastasis in PC3 cells through STAT3 signaling pathway. Res. Pharm. Sci. 2020, 15, 551–562. [Google Scholar] [CrossRef] [PubMed]
  113. Ahmad, B.; Gamallat, Y.; Su, P.; Husain, A.; Rehman, A.U.; Zaky, M.Y.; Bakheet, A.M.H.; Tahir, N.; Xin, Y.; Liang, W. Alantolactone induces apoptosis in THP-1 cells through STAT3, survivin inhibition, and intrinsic apoptosis pathway. Chem. Biol. Drug Des. 2021, 97, 266–272. [Google Scholar] [CrossRef] [PubMed]
  114. Chun, J.; Li, R.J.; Cheng, M.S.; Kim, Y.S. Alantolactone selectively suppresses STAT3 activation and exhibits potent anticancer activity in MDA-MB-231 cells. Cancer Lett. 2015, 357, 393–403. [Google Scholar] [CrossRef] [PubMed]
  115. Zheng, H.; Yang, L.; Kang, Y.; Chen, M.; Lin, S.; Xiang, Y.; Li, C.; Dai, X.; Huang, X.; Liang, G.; et al. Alantolactone sensitizes human pancreatic cancer cells to EGFR inhibitors through the inhibition of STAT3 signaling. Mol. Carcinog. 2019, 58, 565–576. [Google Scholar] [CrossRef] [PubMed]
  116. Kim, C.; Cho, S.K.; Kapoor, S.; Kumar, A.; Vali, S.; Abbasi, T.; Kim, S.H.; Sethi, G.; Ahn, K.S. β-Caryophyllene oxide inhibits constitutive and inducible STAT3 signaling pathway through induction of the SHP-1 protein tyrosine phosphatase. Mol. Carcinog. 2014, 53, 793–806. [Google Scholar] [CrossRef]
  117. Zhang, J.; Li, Y.; Wang, J.G.; Feng, J.Y.; Huang, G.D.; Luo, C.G. Dihydroartemisinin Affects STAT3/DDA1 Signaling Pathway and Reverses Breast Cancer Resistance to Cisplatin. Am. J. Chin. Med. 2023, 51, 445–459. [Google Scholar] [CrossRef]
  118. Yan, X.; Li, P.; Zhan, Y.; Qi, M.; Liu, J.; An, Z.; Yang, W.; Xiao, H.; Wu, H.; Qi, Y.; et al. Dihydroartemisinin suppresses STAT3 signaling and Mcl-1 and Survivin expression to potentiate ABT-263-induced apoptosis in Non-small Cell Lung Cancer cells harboring EGFR or RAS mutation. Biochem. Pharmacol. 2018, 150, 72–85. [Google Scholar] [CrossRef]
  119. Jia, L.; Song, Q.; Zhou, C.; Li, X.; Pi, L.; Ma, X.; Li, H.; Lu, X.; Shen, Y. Dihydroartemisinin as a Putative STAT3 Inhibitor, Suppresses the Growth of Head and Neck Squamous Cell Carcinoma by Targeting Jak2/STAT3 Signaling. PLoS ONE 2016, 11, e0147157. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, D.; Zhong, B.; Li, Y.; Liu, X. Dihydroartemisinin increases apoptosis of colon cancer cells through targeting Janus kinase 2/signal transducer and activator of transcription 3 signaling. Oncol. Lett. 2018, 15, 1949–1954. [Google Scholar] [CrossRef] [PubMed]
  121. Sun, Y.; Lu, X.; Li, H.; Li, X. Dihydroartemisinin inhibits IL-6-induced epithelial-mesenchymal transition in laryngeal squamous cell carcinoma via the miR-130b-3p/STAT3/β-catenin signaling pathway. J. Int. Med. Res. 2021, 49, 3000605211009494. [Google Scholar] [CrossRef]
  122. Yu, R.; Jin, L.; Li, F.; Fujimoto, M.; Wei, Q.; Lin, Z.; Ren, X.; Jin, Q.; Li, H.; Meng, F.; et al. Dihydroartemisinin inhibits melanoma by regulating CTL/Treg anti-tumor immunity and STAT3-mediated apoptosis via IL-10 dependent manner. J. Dermatol. Sci. 2020, 99, 193–202. [Google Scholar] [CrossRef] [PubMed]
  123. Li, H.; Lu, H.; Lv, M.; Wang, Q.; Sun, Y. Parthenolide facilitates apoptosis and reverses drug-resistance of human gastric carcinoma cells by inhibiting the STAT3 signaling pathway. Oncol. Lett. 2018, 15, 3572–3579. [Google Scholar] [CrossRef]
  124. Liu, M.; Xiao, C.; Sun, M.; Tan, M.; Hu, L.; Yu, Q. Parthenolide Inhibits STAT3 Signaling by Covalently Targeting Janus Kinases. Molecules 2018, 23, 1478. [Google Scholar] [CrossRef] [PubMed]
  125. Rajendran, P.; Li, F.; Manu, K.A.; Shanmugam, M.K.; Loo, S.Y.; Kumar, A.P.; Sethi, G. γ-Tocotrienol is a novel inhibitor of constitutive and inducible STAT3 signalling pathway in human hepatocellular carcinoma: Potential role as an antiproliferative, pro-apoptotic and chemosensitizing agent. Br. J. Pharmacol. 2011, 163, 283–298. [Google Scholar] [CrossRef] [PubMed]
  126. Darakhshan, S.; Bidmeshki Pour, A.; Hosseinzadeh Colagar, A.; Sisakhtnezhad, S. Thymoquinone and its therapeutic potentials. Pharmacol. Res. 2015, 95–96, 138–158. [Google Scholar] [CrossRef] [PubMed]
  127. Fatima Shad, K.; Soubra, W.; Cordato, D.J. The role of thymoquinone, a major constituent of Nigella sativa, in the treatment of inflammatory and infectious diseases. Clin. Exp. Pharmacol. Physiol. 2021, 48, 1445–1453. [Google Scholar] [CrossRef]
  128. Mir, P.A.; Mohi-Ud-Din, R.; Banday, N.; Maqbool, M.; Raza, S.N.; Farooq, S.; Afzal, S.; Mir, R.H. Anticancer Potential of Thymoquinone: A Novel Bioactive Natural Compound from Nigella sativa L. Anti-Cancer Agents Med. Chem. 2022, 22, 3401–3415. [Google Scholar] [CrossRef] [PubMed]
  129. Dai, Y.; Chen, S.R.; Chai, L.; Zhao, J.; Wang, Y.; Wang, Y. Overview of pharmacological activities of Andrographis paniculata and its major compound andrographolide. Crit. Rev. Food Sci. Nutr. 2019, 59 (Suppl. S1), S17–S29. [Google Scholar] [CrossRef]
  130. Vetvicka, V.; Vannucci, L. Biological properties of andrographolide, an active ingredient of Andrographis Paniculata: A narrative review. Ann. Transl. Med. 2021, 9, 1186. [Google Scholar] [CrossRef]
  131. Wang, X.; Yang, Y.; Liu, X.; Gao, X. Pharmacological properties of tanshinones, the natural products from Salvia miltiorrhiza. Adv. Pharmacol. 2020, 87, 43–70. [Google Scholar] [CrossRef]
  132. Wu, Y.H.; Wu, Y.R.; Li, B.; Yan, Z.Y. Cryptotanshinone: A review of its pharmacology activities and molecular mechanisms. Fitoterapia 2020, 145, 104633. [Google Scholar] [CrossRef] [PubMed]
  133. Liu, Z.; Ouyang, L.; Peng, H.; Zhang, W.Z. Oridonin: Targeting programmed cell death pathways as an anti-tumour agent. Cell Prolif. 2012, 45, 499–507. [Google Scholar] [CrossRef]
  134. Yao, Z.; Xie, F.; Li, M.; Liang, Z.; Xu, W.; Yang, J.; Liu, C.; Li, H.; Zhou, H.; Qu, L.H. Oridonin induces autophagy via inhibition of glucose metabolism in p53-mutated colorectal cancer cells. Cell Death Dis. 2017, 8, e2633. [Google Scholar] [CrossRef] [PubMed]
  135. Li, X.; Zhang, C.T.; Ma, W.; Xie, X.; Huang, Q. Oridonin: A Review of Its Pharmacology, Pharmacokinetics and Toxicity. Front. Pharmacol. 2021, 12, 645824. [Google Scholar] [CrossRef]
  136. Sobral, P.J.M.; Vicente, A.T.S.; Salvador, J.A.R. Recent advances in oridonin derivatives with anticancer activity. Front. Chem. 2023, 11, 1066280. [Google Scholar] [CrossRef] [PubMed]
  137. Abdullah, N.A.; Md Hashim, N.F.; Ammar, A.; Muhamad Zakuan, N. An Insight into the Anti-Angiogenic and Anti-Metastatic Effects of Oridonin: Current Knowledge and Future Potential. Molecules 2021, 26, 775. [Google Scholar] [CrossRef]
  138. Hu, X.; Huang, S.; Ye, S.; Jiang, J. The Natural Product Oridonin as an Anticancer Agent: Current Achievements and Problems. Curr. Pharm. Biotechnol. 2023. ahead of print. [Google Scholar] [CrossRef]
  139. Cai, S.J.; Liu, Y.; Han, S.; Yang, C. Brusatol, an NRF2 inhibitor for future cancer therapeutic. Cell Biosci. 2019, 9, 45. [Google Scholar] [CrossRef]
  140. Hall, I.H.; Lee, K.H.; Eigebaly, S.A.; Imakura, Y.; Sumida, Y.; Wu, R.Y. Antitumor agents. XXXIV: Mechanism of action of bruceoside A and brusatol on nucleic acid metabolism of P-388 lymphocytic leukemia cells. J. Pharm. Sci. 1979, 68, 883–887. [Google Scholar] [CrossRef]
  141. Jiang, W.; Li, X.; Dong, S.; Zhou, W. Betulinic acid in the treatment of tumour diseases: Application and research progress. Biomed. Pharmacother. 2021, 142, 111990. [Google Scholar] [CrossRef] [PubMed]
  142. Aswathy, M.; Vijayan, A.; Daimary, U.D.; Girisa, S.; Radhakrishnan, K.V.; Kunnumakkara, A.B. Betulinic acid: A natural promising anticancer drug, current situation, and future perspectives. J. Biochem. Mol. Toxicol. 2022, 36, e23206. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, C.; Dai, S.; Zhao, X.; Zhang, Y.; Gong, L.; Fu, K.; Ma, C.; Peng, C.; Li, Y. Celastrol as an emerging anticancer agent: Current status, challenges and therapeutic strategies. Biomed. Pharmacother. 2023, 163, 114882. [Google Scholar] [CrossRef] [PubMed]
  144. Dai, S.; Wang, C.; Zhao, X.; Ma, C.; Fu, K.; Liu, Y.; Peng, C.; Li, Y. Cucurbitacin B: A review of its pharmacology, toxicity, and pharmacokinetics. Pharmacol. Res. 2023, 187, 106587. [Google Scholar] [CrossRef] [PubMed]
  145. Garg, S.; Kaul, S.C.; Wadhwa, R. Cucurbitacin B and cancer intervention: Chemistry, biology and mechanisms (Review). Int. J. Oncol. 2018, 52, 19–37. [Google Scholar] [CrossRef] [PubMed]
  146. Zieniuk, B.; Pawełkowicz, M. Recent Advances in the Application of Cucurbitacins as Anticancer Agents. Metabolites 2023, 13, 1081. [Google Scholar] [CrossRef]
  147. Kumar, A.; Sharma, B.; Sharma, U.; Parashar, G.; Parashar, N.C.; Rani, I.; Ramniwas, S.; Kaur, S.; Haque, S.; Tuli, H.S. Apoptotic and antimetastatic effect of cucurbitacins in cancer: Recent trends and advancement. Naunyn Schmiedebergs Arch. Pharmacol. 2023, 396, 1867–1878. [Google Scholar] [CrossRef] [PubMed]
  148. Saeed, M.E.M.; Boulos, J.C.; Elhaboub, G.; Rigano, D.; Saab, A.; Loizzo, M.R.; Hassan, L.E.A.; Sugimoto, Y.; Piacente, S.; Tundis, R.; et al. Cytotoxicity of cucurbitacin E from Citrullus colocynthis against multidrug-resistant cancer cells. Phytomedicine 2019, 62, 152945. [Google Scholar] [CrossRef] [PubMed]
  149. Varela, C.; Melim, C.; Neves, B.G.; Sharifi-Rad, J.; Calina, D.; Mamurova, A.; Cabral, C. Cucurbitacins as potential anticancer agents: New insights on molecular mechanisms. J. Transl. Med. 2022, 20, 630. [Google Scholar] [CrossRef]
  150. Lin, X.; Farooqi, A.A. Cucurbitacin mediated regulation of deregulated oncogenic signaling cascades and non-coding RNAs in different cancers: Spotlight on JAK/STAT, Wnt/β-catenin, mTOR, TRAIL-mediated pathways. Semin. Cancer Biol. 2021, 73, 302–309. [Google Scholar] [CrossRef]
  151. Mlala, S.; Oyedeji, A.O.; Gondwe, M.; Oyedeji, O.O. Ursolic Acid and Its Derivatives as Bioactive Agents. Molecules 2019, 24, 2751. [Google Scholar] [CrossRef]
  152. Hirano, T. IL-6 in inflammation, autoimmunity and cancer. Int. Immunol. 2021, 33, 127–148. [Google Scholar] [CrossRef] [PubMed]
  153. Liu, X.; Bian, L.; Duan, X.; Zhuang, X.; Sui, Y.; Yang, L. Alantolactone: A sesquiterpene lactone with diverse pharmacological effects. Chem. Biol. Drug Des. 2021, 98, 1131–1145. [Google Scholar] [CrossRef] [PubMed]
  154. Cai, Y.; Gao, K.; Peng, B.; Xu, Z.; Peng, J.; Li, J.; Chen, X.; Zeng, S.; Hu, K.; Yan, Y. Alantolactone: A Natural Plant Extract as a Potential Therapeutic Agent for Cancer. Front. Pharmacol. 2021, 12, 781033. [Google Scholar] [CrossRef]
  155. Babaei, G.; Gholizadeh-Ghaleh Aziz, S.; Rajabi Bazl, M.; Khadem Ansari, M.H. A comprehensive review of anticancer mechanisms of action of Alantolactone. Biomed. Pharmacother. 2021, 136, 111231. [Google Scholar] [CrossRef] [PubMed]
  156. Fidyt, K.; Fiedorowicz, A.; Strządała, L.; Szumny, A. β-caryophyllene and β-caryophyllene oxide-natural compounds of anticancer and analgesic properties. Cancer Med. 2016, 5, 3007–3017. [Google Scholar] [CrossRef]
  157. Liu, X.; Cao, J.; Huang, G.; Zhao, Q.; Shen, J. Biological Activities of Artemisinin Derivatives Beyond Malaria. Curr. Top. Med. Chem. 2019, 19, 205–222. [Google Scholar] [CrossRef]
  158. Slezakova, S.; Ruda-Kucerova, J. Anticancer Activity of Artemisinin and its Derivatives. Anticancer. Res. 2017, 37, 5995–6003. [Google Scholar] [CrossRef]
  159. Sztiller-Sikorska, M.; Czyz, M. Parthenolide as Cooperating Agent for Anti-Cancer Treatment of Various Malignancies. Pharmaceuticals 2020, 13, 194. [Google Scholar] [CrossRef]
  160. Pang, K.L.; Mai, C.W.; Chin, K.Y. Molecular Mechanism of Tocotrienol-Mediated Anticancer Properties: A Systematic Review of the Involvement of Endoplasmic Reticulum Stress and Unfolded Protein Response. Nutrients 2023, 15, 1854. [Google Scholar] [CrossRef]
Figure 1. Various stimuli like growth factors and cytokines cause abnormal activation of the STAT3 signal transduction cascade. STAT3 phosphorylation causes nuclear localization, which stimulates STAT3 target genes implicated in cancer processes such as cell proliferation, angiogenesis, metastasis, invasion, and drug resistance.
Figure 1. Various stimuli like growth factors and cytokines cause abnormal activation of the STAT3 signal transduction cascade. STAT3 phosphorylation causes nuclear localization, which stimulates STAT3 target genes implicated in cancer processes such as cell proliferation, angiogenesis, metastasis, invasion, and drug resistance.
Biomolecules 14 00200 g001
Figure 2. Different terpene classes, including monoterpene, diterpene, triterpene, and sesquiterpene, target STAT3 signaling pathways by directly or indirectly influencing the phosphorylation status of STAT3 protein in various cancer cells.
Figure 2. Different terpene classes, including monoterpene, diterpene, triterpene, and sesquiterpene, target STAT3 signaling pathways by directly or indirectly influencing the phosphorylation status of STAT3 protein in various cancer cells.
Biomolecules 14 00200 g002
Table 1. Different classes of terpenes exhibiting their anticancer potential by downregulating key molecules of STAT3 signaling pathway.
Table 1. Different classes of terpenes exhibiting their anticancer potential by downregulating key molecules of STAT3 signaling pathway.
ClassPhytochemicalCancerModelMolecular TargetMechanismReferences
MonoterpeneThymoquinone
Biomolecules 14 00200 i001
Gastric cancerHGC27, BGC823, SGC7901 cells; xenograft tumor mouse modelJAK2, STAT3, SrcReduced tumor growth, induced apoptosis[61]
Colon cancerHCT116 cellsJAK2, STAT3, SrcInhibited cell proliferation, induced apoptosis[62]
LeukemiaK562 cellsJAK2, STAT3, STAT5Inhibited cell proliferation, induced apoptosis[63]
Acute myeloid leukemiaHL60 cellsc-Myc, PI3K, AKT, mTOR, JAK2, STAT3, STAT5a, STAT5bInhibited cell proliferation, induced apoptosis[64]
Breast cancerMice bearing solid Ehrlich tumorsSTAT3, caspase-3/9Attenuated tumor growth, induced apoptosis, chemosensitivity[65]
MelanomaSK-MEL-28 cells, SK-MEL-28 tumor xenograftsJak2, STAT3Induced apoptotic cell death[66]
Renal cell carcinomaCaki-1 cells, tumor xenograft miceJak2, STAT3, cyclin D2Attenuated tumor growth, induced apoptosis[67]
Skin cancerA431 cells, tumor xenograft miceSTAT3, Src, cyclin D1Attenuated tumor growth, induced apoptosis[68]
MyelomaU266 and RPMI 8226 cellsSTAT3, c-Src, JAK2Attenuated cell growth, induced apoptosis, G1 cell cycle arrest, chemosensitivity[69]
DiterpeneAndrographolide
Biomolecules 14 00200 i002
Non-small-cell lung cancer (NSCLC)H1975 and H1299 cell linesSTAT3, PARP, PD-L1, P62Inducted autophagy, antitumor immune response[70]
Multiple cancer cell typesHCT116, MDA-MB-231, HepG2, HeLa, TPC-1STAT3, Bcl-xl, cyclin D1Induced cell death, chemosensitivity[71]
Cryptotanshinone
Biomolecules 14 00200 i003
Prostate cancerDU145 cellsJAK2, STAT3Retarded cell proliferation, induced apoptosis[72]
Ovarian cancerHey and A2780 cellsHIF- 1α, STAT3Retarded cell proliferation and glucose metabolism[73]
Pancreatic cancer BxPC-3 cellsJAK2, STAT3, mTOR, AktRetarded cell proliferation, induced apoptosis[74]
Esophageal cancerEC109 and CAES17 cells, athymic nude miceJAK2, STAT3Retarded cell proliferation, induced apoptosis[75]
Renal cell carcinomaA498, 786-O, ACHNSTAT3, cyclin D1, Bcl-2Retarded cell proliferation, induced apoptosis[76]
GliomaT98G and U87 cellsSTAT3, cyclin D1, survivinSuppressed cell viability, induced cell cycle arrest and apoptosis[77]
GliomaC6, U251, T98G, U87; nude xenograft miceSTAT3, SHP2, cyclin D1Inhibited cell proliferation[78]
Chronic myeloid leukemiaK562-R cells, xenografts in nude miceSTAT3, eIF4EInhibited tumor growth, induced apoptosis[79]
Chronic myeloid leukemiaK562 cellsSTAT3, STAT5Suppressed key oncogenic proliferation, drug resistance[80]
Oridonin
Biomolecules 14 00200 i004
Thyroid cancerTPC-1 and BCPAP cellsJAK2, STAT3Inhibited metastatic, angiogenesis, and modulated EMT[81]
Nasopharyngeal carcinomaCNE-2Z and HNE-1 cellsAkt, STAT3Inhibited metastatic, angiogenesis, and modulated EMT[82]
OsteosarcomaU2OS cellsSTAT3, MMP-2, 3, 9Suppressed cell viability, induced apoptosis[83]
TriterpeneBrusatol
Biomolecules 14 00200 i005
Laryngeal cancerHep-2 cells, xenograft laryngeal tumorJAK2, STAT3Inhibited viability, migration, and invasion ability[84]
Head and neck squamous cell carcinomaUMSCC 47, UD SCC2, JMAR, Tu167, LN686, FaDuJAK1, JAK2, STAT3, SrcReduced cell growth, induced apoptosis[85]
Pancreatic cancerPANC-1 and PATU-8988 cellsNF-κb, Stat3Reduced cell growth, induced apoptosis[86]
Hepatocellular carcinomaHCCLM3 cell lineSTAT3Inhibited cell migration and invasion[87]
Betulinic acid
Biomolecules 14 00200 i006
Multiple myelomaU266 and MM.1S cellsJAK1, JAK2, STAT3, SrcInduced apoptotic cancer cell death[88]
Prostate cancerPC3HIF-1α, STAT3Exert anti-angiogenic activity[89]
Celastrol
Biomolecules 14 00200 i007
Hepatocellular carcinomaC3A, HepG2, Hep3B, PLC/PRF5; athymic nu/nu miceJAK1, JAK2, STAT3, SrcInhibited cell migration and invasion, induced apoptosis[90]
NSCLCH460, PC-9, H520, BEAS-2B, PC-9 cells; thymic BALB/c nude miceSTAT3, Bcl-2Reduced cell growth, proliferation, and metastasis[91]
Multiple myelomaU266, RPMI 8226IκBα kinase, STAT3Suppressed cell viability, induced cell cycle arrest and apoptosis[92]
Cucurbitacin B
Biomolecules 14 00200 i008
Pancreatic cancerMiaPaCa-2,
AsPC-1
JAK2, STAT3, and STAT5 activationCytotoxicity[93]
Pancreatic cancerPANC-1STAT3 Growth and cell cycle inhibition, induced apoptosis[94]
Colorectal cancerHT-29, HCT-116STAT3Growth and cell cycle inhibition, induced apoptosis[95]
Hepatocellular carcinomaHepG2STAT3 Cell cycle arrest, retarded cell growth[96]
Gastric cancerMKN-45STAT3Cell cycle arrest, retarded cell growth[97]
Gastric cancerSGC7901, BGC823, MGC803, MKN74; human gastric cancer xenograftSTAT3, c-Myc, Bcl-xLSuppressed invasion, induced apoptosis[98]
Cucurbitacin E
Biomolecules 14 00200 i009
Pancreatic cancerPANC-1STAT3Cell cycle arrest, retarded cell growth[99]
Prostate cancerPC-3, xenograftSTAT3, JAK2Inhibited angiogenesis, proliferation, survival, and migration[100]
Human bladder cancerT24 cellsSTAT3, CDK1, cyclin BInduced G(2)/M phase arrest and apoptosis[101]
Cucurbitacin I
Biomolecules 14 00200 i010
Pancreatic cancerASPC-1, BXPC-3, CFPAC-1, SW 1990; orthotopic xenograft miceSTAT3, JAK2Inhibited proliferation[102]
Hepatocellular carcinomaHepG2 cellsSTAT3, JAK2Induced antiproliferation and G2/M phase of cell cycle[103]
Lung cancerA549 cellsERK, mTOR, STAT3Decreased cell viability, inhibited colony formation, induced apoptosis[104]
Lung cancerA549; nude mouse tumor xenograft modelSTAT3Suppressed tumor growth, induced apoptosis[105]
GlioblastomaU251 and A172 cellsSTAT3, cyclin B1, cdc2Decreased cell viability, induced G2/M cell cycle arrest, induced apoptosis[106]
LymphomaALK+ ALCL cell linesSTAT3, JAK3, NPM-ALKDecreased cell viability, induced G2/M cell cycle arrest, induced apoptosis[107]
Ursolic Acid
Biomolecules 14 00200 i011
Lung cancerA549 and H460 cellsSTAT3, MMP2, PD-L1, VEGFAttenuated cell growth, invasion, migration, angiogenesis induced apoptosis, G1 cell cycle arrest, chemosensitivity[108]
Colorectal cancerHCT116 and HT29 cellsSTAT3Induced apoptotic cell death[109]
Hepatocellular carcinomaHep3B, HEPG2, SSMC-7721, and Huh7 cells; mouse xenograft tumor modelSTAT3Suppressed cell viability, cell migration, and colony formation[110]
Prostate cancerTRAMP miceNF-κB, STAT3, AKT, IKKα/βReduced tumor growth[111]
SesquiterpeneAlantolactone
Biomolecules 14 00200 i012
Prostate cancerPC3SOX2, Oct-4, Nanog, CD133, CD44, STAT3Antimetastatic potential[112]
LeukemiaTHP-1 cellsSTAT3, survivinDecreased cell viability, increased cell death and apoptosis[113]
Breast cancerMDA-MB-231 cellsSTAT3, MAPKs, NF-κBInhibition of migration, invasion, and adhesion[114]
Pancreatic cancerBxPC-3, AsPC1, and PANC-1 cellsSTAT3, Bcl-2Reduced cell growth, induced cell death[115]
β-Caryophyllene oxide
Biomolecules 14 00200 i013
Human multiple myeloma (MM) cell lines, human prostate carcinoma, human breast carcinoma U266,
MM1.S, DU145, MDAMB-231
JAK1, JAK2, STAT3, SrcReduced cell proliferation and invasion, induced apoptosis [116]
Dihydroartemisinin
Biomolecules 14 00200 i014
Breast cancerMDA-MB-231 cellsSTAT3, DDA1Repressed cell proliferation, induced apoptosis[117]
Lung cancerH1975, HCC827, H1650, H3255, A549, H727STAT3, Mcl-1, survivinBax-dependent apoptosis[118]
Head and neck CarcinomaFadu, Cal-27, and Hep-2 cells; xenograft modelJak2, STAT3Apoptosis induction, attenuation of cell migration[119]
Colon cancerHCT116Jak2, STAT3Suppressed cell viability, induced apoptosis[120]
Laryngeal squamous cell carcinomaAMC-HN-8 and Tu212 cellsIL-6/STAT3, β-cateninRepressed EMT and invasion[121]
MelanomaB16F10 cells; BALB/c xenograft miceSTAT3, IL-10, IL-6Induced apoptotic cell death[122]
Parthenolide
Biomolecules 14 00200 i015
Gastric cancerSGC-7901/DDP cellJAK2, STAT3, cyclin D1Decreased cell viability, induced G1 cell cycle arrest, induced apoptosis[123]
Multiple cancer subtypesHepG2, MDA-MB-231, MDA-MB-468, HCT116, HT-29, Lovo, NCI-H1299, Colo205, BGC, H460, and Du145 cellsJAK1, STAT3,Suppressed cellular growth and migration[124]
γ-Tocotrienol
Biomolecules 14 00200 i016
Hepatocellular carcinomaHepG2, C3A, SNU-387, and PLC/PRF5 cellsSTAT3, cyclin D1, Bcl-2, Bcl-xL, survivin, Mcl-1, VEGFInhibited proliferation, induced apoptosis, chemosensitivity[125]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, F.; Pandey, P.; Verma, M.; Upadhyay, T.K. Terpenoid-Mediated Targeting of STAT3 Signaling in Cancer: An Overview of Preclinical Studies. Biomolecules 2024, 14, 200. https://doi.org/10.3390/biom14020200

AMA Style

Khan F, Pandey P, Verma M, Upadhyay TK. Terpenoid-Mediated Targeting of STAT3 Signaling in Cancer: An Overview of Preclinical Studies. Biomolecules. 2024; 14(2):200. https://doi.org/10.3390/biom14020200

Chicago/Turabian Style

Khan, Fahad, Pratibha Pandey, Meenakshi Verma, and Tarun Kumar Upadhyay. 2024. "Terpenoid-Mediated Targeting of STAT3 Signaling in Cancer: An Overview of Preclinical Studies" Biomolecules 14, no. 2: 200. https://doi.org/10.3390/biom14020200

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop