Next Article in Journal
Hepatic, Extrahepatic and Extracellular Vesicle Cytochrome P450 2E1 in Alcohol and Acetaminophen-Mediated Adverse Interactions and Potential Treatment Options
Previous Article in Journal
A Multi-Strain Probiotic Formulation Improves Intestinal Barrier Function by the Modulation of Tight and Adherent Junction Proteins
Previous Article in Special Issue
Roads to Stat3 Paved with Cadherins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Central Roles of STAT3-Mediated Signals in Onset and Development of Cancers: Tumorigenesis and Immunosurveillance

1
Division of Molecular Psychoimmunology, Institute for Genetic Medicine, Hokkaido University, Sapporo 060-0815, Japan
2
Department of Molecular Biology, Graduate School of Medicine, Hokkaido University, Sapporo 060-8638, Japan
3
Department of Immunology, Graduate School of Pharmaceutical Sciences, Hokkaido University, Sapporo 060-0812, Japan
4
Department of Hematology, International University of Health and Welfare, Narita 286-8686, Japan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2022, 11(16), 2618; https://doi.org/10.3390/cells11162618
Submission received: 22 July 2022 / Revised: 12 August 2022 / Accepted: 20 August 2022 / Published: 22 August 2022
(This article belongs to the Special Issue STAT3: Role in Cancer and Stem Cells)

Abstract

:
Since the time of Rudolf Virchow in the 19th century, it has been well-known that cancer-associated inflammation contributes to tumor initiation and progression. However, it remains unclear whether a collapse of the balance between the antitumor immune response via the immunological surveillance system and protumor immunity due to cancer-related inflammation is responsible for cancer malignancy. The majority of inflammatory signals affect tumorigenesis by activating signal transducer and activation of transcription 3 (STAT3) and nuclear factor-κB. Persistent STAT3 activation in malignant cancer cells mediates extremely widespread functions, including cell growth, survival, angiogenesis, and invasion and contributes to an increase in inflammation-associated tumorigenesis. In addition, intracellular STAT3 activation in immune cells causes suppressive effects on antitumor immunity and leads to the differentiation and mobilization of immature myeloid-derived cells and tumor-associated macrophages. In many cancer types, STAT3 does not directly rely on its activation by oncogenic mutations but has important oncogenic and malignant transformation-associated functions in both cancer and stromal cells in the tumor microenvironment (TME). We have reported a series of studies aiming towards understanding the molecular mechanisms underlying the proliferation of various types of tumors involving signal-transducing adaptor protein-2 as an adaptor molecule that modulates STAT3 activity, and we recently found that AT-rich interactive domain-containing protein 5a functions as an mRNA stabilizer that orchestrates an immunosuppressive TME in malignant mesenchymal tumors. In this review, we summarize recent advances in our understanding of the functional role of STAT3 in tumor progression and introduce novel molecular mechanisms of cancer development and malignant transformation involving STAT3 activation that we have identified to date. Finally, we discuss potential therapeutic strategies for cancer that target the signaling pathway to augment STAT3 activity.

1. Introduction

The importance of chronic inflammation in the mechanism of cancer has been well-established [1]. Inflammation plays a crucial role in almost all aspects of the tumorigenic process [2]. The role of inflammation in tumorigenesis has been extensively investigated, and recent lines of evidence provide a possible links between inflammation and tumor recurrence and metastasis [2,3]. However, the effects of inflammatory signaling on tumorigenesis remain elusive.
In 1994, signal transducer and activation of transcription 3 (STAT3) was discovered as a transcription factor involved in interleukin-6 (IL-6)-induced hepatic acute phase responses by Kishimoto and Akira’s group and Darnell’s group [4,5]. IL-6, the best-known protumor cytokine, and its family of cytokines, including IL-11, IL-27, IL-31, cardiotrophin-1, ciliary neurotrophic factor, leukemia inhibitory factor (LIF), and oncostatin M (OSM), are involved in crucial physiological and/or pathological processes, such as cell growth, survival, differentiation, energy metabolism, angiogenesis, migration, invasion, metastasis, inflammation, and autoimmune diseases [6,7,8,9,10]. The IL-6 family cytokines, excluding IL-31, can transduce intracellular signals linking with the Janus kinase (JAK)-STAT3 pathway, the Src homology 2 (SH2)-containing protein tyrosine phosphatase-2 (SHIP2)-Ras-Raf-MEK-extracellular signal-regulated kinase (ERK) pathway, and the phosphoinositide 3-kinase (PI3K)-Akt pathway mediated by the activation of shared signal-transducing receptor component glycoprotein 130 (gp130, IL6ST) [6,7,8,9,10]. In these signaling pathways, STAT3 is considered to be a key signaling molecule of the IL-6-gp130 pathway because it acts as an oncogenic driver and plays an important role in mediating tumor-promoting inflammation [6,7,9,11,12]. Importantly, suppressor of cytokine signaling 3 (SOCS3) is induced by STAT3 and is postulated to modulate the primary negative regulation of the gp130-mediated signaling pathway [13,14].
Besides the IL-6 family, activation of cellular STAT3 is also triggered by hepatocyte growth factor receptor, c-MET, epidermal growth factor (EGF) receptor (EGFR) [15,16], and Src family kinases [11,12]. Furthermore, it has been reported that G-protein-coupled receptors, such as sphingosine-1-phosphate receptor 1, stimulate STAT3 via JAK and Src family kinases [17,18] and that Toll-like receptors (TLRs), such as TLR9 and TLR4, are considered to play crucial roles in inflammation via the activation of the JAK-STAT3 pathway [19,20,21]. MicroRNAs (miRNAs) such as miR-17-5p, miR-20a, miR-124, and miR-551b-3p have emerged as key modulators of cancer biology, and some of these miRNAs have been shown to be pivotal in the regulation of the JAK-STAT3 pathway [22,23,24,25,26] (Figure 1).
Moreover, STAT3 not only contributes functionally to promoting tumor cell proliferation, survival, invasion, angiogenesis, and immune evasion, but it has been recently indicated to play crucial roles in the inflammation associated with tumorigenesis, obesity and metabolic syndrome, the cancer stemness pathway, and premetastatic niche formation [11,12,27,28,29,30,31,32]. Furthermore, STAT3 has been shown to be involved in the formation of immunosuppressive tumor microenvironments (TMEs) via regulating not only immune cells but also cancer-associated fibroblasts (CAFs) and endothelial cells.
We previously performed a series of studies analyzing the roles of signal-transducing adaptor protein-2 (STAP-2) in the proliferation of several types of cancer cells via acting as an adaptor molecule that modulates STAT3 activity [33] and recently found that AT-rich interactive domain-containing protein 5A (ARID5A) functions as an RNA-binding molecule that stabilizes mRNAs such as those of indoleamine 2,3-dioxygenase 1 (IDO1), C-C motif chemokine ligand 2 (CCL2), and STAT3, resulting in the induction of an immunosuppressive TME in malignant tumors [34]. In this review, we provide an overview of the recent findings regarding the intrinsic and extrinsic roles of STAT3 during tumor progression. We further introduce the novel molecular mechanisms that we have identified to date involving STAT3 activation, cancer development, and malignant transformation. Finally, we address the potential therapeutic strategies against malignant tumors by targeting the signaling pathway to augment STAT3 activity.

2. STAT3 Signal in Cancer Cells

IL-6 is a well-established tumor-promoting cytokine among the IL-6 family of cytokines, which activates multiple STAT3-mediated tumor initiation and progression pathways [11,12,35,36]. For example, the IL-6/STAT3 axis enhances the transcriptional activation of various molecular targets that are crucial for cell cycle progression and survival (e.g., cyclin D1, myc, Bcl2-like 1, survivin, and miR-21) and angiogenesis (e.g., hypoxia-inducible factor 1α, vascular endothelial growth factor (VEGF), and matrix metalloproteinases (MMPs, e.g., MMP2, MMP7, and MMP9)) [37,38,39]. In the late stages of cancer, IL-6/STAT3 may promote the gain of invasive activity and the metastatic dissemination of cancer cells by inducing epithelial–mesenchymal transition (EMT) transcription factors (EMT-TFs), such as SNAI1 and TWIST [40]. Notably, the EMT program in cancer biology has been implicated to facilitate not only cell motility and invasiveness but to also possibly be involved in cancer stem cell (CSC) status and the resistance to anticancer drugs via epithelial–mesenchymal plasticity [41,42,43]. Consistently, IL-6/STAT3 signaling in cancerous EMT also results in the acquisition of cancer stemness in cancer cells; the self-renewal and population expansion of CSCs requires STAT3 in cooperation with stem-cell-associated transcription factors, such as NANOG [44]. In addition, IL-6/STAT3 signaling is crucial for the shift from non-CSCs to CSCs by upregulating the expression of Oct4 [45]. These functions of IL-6/STAT3 signaling ultimately lead to the development of several multidrug-resistant and malignant phenotypes [44,45] (Figure 1).

2.1. Pancreatic Cancer

High serum levels of IL-6 have been associated with poor overall survival prognosis in patients with highly malignant pancreatic cancer [46], and increased activity of IL-6/STAT3-mediated signaling has been reported to be associated with poor prognosis in patients with pancreatic ductal adenocarcinoma (PDAC) after resection [47]. The activation of STAT3 in PDAC has been reported in patient-derived clinical specimens and pancreatic cancer cells [48] and is a prognostic risk factor [49]. IL-6 also induces a mesenchymal phenotype in human pancreatic cancer cells via STAT3 activation and SNAI1 induction [50]. Interestingly, in a mouse model, STAT3 is involved in the reprogramming of acinar-to-ductal metaplasia (ADM), which is triggered by the sustained exocrine-tissue-specific expression of pancreatic and duodenal homeobox 1 (Pdx1), which is a pancreatic-progenitor-specific transcription factor [51]. ADM transdifferentiation occurs in chronic pancreatitis via STAT3 and is associated with pancreatic intraepithelial neoplasia (PanIN), which is a necessary step for the generation of neoplastic precursor lesions [52,53]. In a PDAC mouse model driven by KRAS [52,53], for instance, it has been reported that pancreatic epithelial cells bearing the constitutively active KRAS mutation KRASG12D trigger inflammation activation by recruiting immune cells, particularly myeloid cells, that facilitate the production of IL-6 and soluble IL6R (sIL6R) and, in turn, activate STAT3 via IL-6 trans-signaling through the binding of IL-6 to the soluble form of IL6R and the subsequent binding of IL-6 and sIL6R complexes to gp130-expressing cells [52]. Dysregulated STAT3 activation due to the homozygous loss of SOCS3 in the pancreas leads to the accelerated progression of PanIN and the onset of PDAC [52]. It has also been shown that the activation of KRAS increases cytokine levels, such as IL-6 and IL-11 in epithelial cells, which subsequently drives STAT3 activation in an autocrine manner, and that STAT3-triggerd MMP7 is necessary for tumor progression but not for tumor onset, which might be regulated by other STAT3 targets [53].
Because of the TME of PDAC, in which the low vascular density results in severe hypoxia and limited nutrient utilization, PDAC cells are known to have increased autophagy to rewire their metabolism to survive and maintain metabolic homeostasis in harsh environments [54,55]. In the mouse model of PDAC caused by KRAS mutations, increased levels of autophagy are required for IL-6-induced STAT3 activation. Mechanistically, the receptor for advanced glycation products promotes the IL-6-driven activation of STAT3 signaling in mitochondria, providing a bridge between autophagy and the IL-6-STAT3 signaling pathway [56].

2.2. Colorectal Cancer

Increased levels of IL-6 and sIL6R in the circulating blood and intestine in patients with inflammatory bowel disease are primary risk factors for colitis-associated cancer (CAC) [57]. The serum levels of IL-6 in patients with colorectal cancer (CRC) are correlated with the malignant tumor grade, and high IL-6 levels (≥10 pg/mL) are an independent indicator of poor prognosis [58]. Controversially, although it has also been shown that the IL-6 levels in the serum correlate with disease progression in CRC patients, the IL-6 level is not an independent prognostic marker [59]. On the other hand, the activation of STAT3 has been reported to associate with poor outcomes in CRC patients [60,61,62]. The expression of both IL6R and gp130 have been observed in epithelial cells of the intestine and in immune cells, and the release of membrane-bound IL6R has been detected in the serum with the progression of CAC [63]. sIL6R released within the TME can induce STAT3 activation in gp130-expressing cells by trans-signaling [64].
Compositional changes in the microbiota are associated with a predisposition to the development of colorectal tumors. It has been demonstrated that a high intake of dietary fat and meat is associated with a high risk of colorectal cancer, which may result from diet-induced differences in the microbiota composition and metabolic activities [65]. It has also been shown, using a CAC mouse model created by the injection of azoxymethane (AOM) followed by treatment with dextran sulfate sodium salt (DSS), that apoptosis-associated speck-like protein containing a caspase recruitment domain or NOD-like receptor family pyrin domain containing 6 performs important functions in CAC progression. Furthermore, an interesting finding has been observed in that wild-type mice cohabitating with mice lacking these inflammasome genes are more vulnerable to the initiation of CAC [66]. Mechanistically, IL-18-induced changes in the microbiota induce CC-chemokine ligand 5-driven inflammation, which accelerates epithelial cell proliferation through the regional activation of the IL-6/STAT3 pathway, eventually resulting in cancer formation [66].
Although the signal transduction of IL-6 is crucial for STAT3 activation in CRC initiation and development [63,67,68,69], the ablation of STAT3 in intestinal enterocytes has more significant effects on mucosal damage and regeneration, tumor growth, and proliferation than the lack of IL-6 in the CAC model induced by AOM and DSS [68], indicating that other cytokines involved in STAT3 activation, such as EGF family growth factors, IL-11, and IL-22, as well as hormones, such as leptin, may drive the activation of STAT3 in inflammation-induced CRC cells.
Sporadic CRC in colorectal adenomatosis Apc (Min/+) mice, a commonly used animal model bearing numerous adenomatous polyps reflecting familial adenomatosis of the colon based on heterozygosity for Apc truncation mutations, is also essential for IL-6 signaling [70,71]. Apc (Min/+) mice lacking STAT3 had a reduced occurrence of and suppressed the growth of early adenomas [72]. However, STAT3 deficiency promoted late tumor progression and led to the formation of invasive and metastatic carcinomas via the enhancement of carcinoembryonic antigen-related cellular adhesion molecule 1, which is involved in intercellular adhesion [72]. Conversely, it has also been suggested that STAT3 does not affect tumorigenesis, but the downregulation of Snail1 inhibits the transition from adenoma to cancer in Apc (Min/+) mice [73]. Additionally, IL-11 has been shown to correlate more strongly than IL-6 with increased STAT3 activation in human CRC specimens [70]. Subsequently, it has been demonstrated that IL-11/STAT3-mediated signaling functions as a stronger promoter of the progression of sporadic and inflammation-associated CRC than IL-6/STAT3 signaling in the progression of sporadic and inflammation-associated CRC progression, suggesting that IL-11/STAT3 signaling is a promising therapeutic target for the cure of CRCs [70,74]. Notably, the results of the possible tumor-suppressive roles of STAT3 in a CRC mouse model require further investigation regarding the underlying molecular mechanisms and consistency with clinical observations.
IL-6, together with transforming growth factor beta (TGF-β), induces the generation of Th17 cells, and Th17 cells and other cells producing IL-17A trigger sporadic CRC in mice and humans [75,76,77]. The “Th17 gene expression profiling” in stage I to II CRCs is correlated with a significant reduction in disease-free survival [77]. The product of human colonic bacterium, enterotoxigenic bacteroides fragilis, substantially induced CRC onset via STAT3 activation in Th17 cells [78]. In inflammation-associated colon cancer, increased TLR4 expression in intestinal epithelial cells leads to the activation of STAT3, which promotes the growth of CRC in vivo [21]. Furthermore, TLR4/STAT3 signaling has been shown to correlate with the clinical stage in human colorectal adenocarcinoma [21].

2.3. Prostate Cancer

The IL-6 levels in serum are increased in patients with castration-resistant or untreated metastatic prostate cancer and are associated with poor outcomes and resistance to treatment with chemotherapy [79]. Serum sIL6R levels have also been shown to be associated with the progression and metastasis of prostate cancer [79]. Serum IL-11 is a potential tumor biomarker for advanced prostate cancer [80]. The augmented expression of IL-11R and the activation of STAT3 have been observed in human prostate cancer [81,82], indicating IL-11R as a promising therapeutic target against human androgen-resistant and advanced prostate cancer [82]. However, the activation status of STAT3 has been reported to be inversely associated with the progression of distant metastases in prostate cancer [83], whereas conflicting reports suggest that it is an effective prognostic marker for prostate cancer [84]. Therefore, further evaluation is warranted. Furthermore, IL-6/STAT3 signaling has been implicated in the conversion from androgen-sensitive to androgen-resistant prostate cancer via the recruitment of myeloid-derived suppressor cells (MDSCs) [85,86]. Using a mouse model of prostate cancer, androgen deprivation has been shown to activate nuclear factor-κB (NF-κB) and STAT3 signaling in prostate cancer cells via leukocyte infiltration, which triggers androgen-dependent tumor cell death and consequently promotes androgen-independent survival. However, cytokines that exclusively activate STAT3 signaling in this environment have not yet been identified [87]. NF-κB activation in prostate cancer cell lines leads to the increased production of IL-6, which contributes to docetaxel resistance [88]. Thus, treatment by the simultaneous inhibition of NF-κB and IL-6/STAT3 signaling is a possible therapeutic strategy to improve the response to chemotherapy and radiation in prostate cancer. STAT3 has been demonstrated to directly bind to androgen receptors and to transcriptionally augment androgen-receptor-targeted genes, even upon a lack of high doses of androgen [89]. In contrast, the silencing of androgen receptor expression enhances CSC-like traits in prostate cancer via IL-6/STAT3 signaling [32]. In addition, blocking the JAK-STAT3 axis suppresses tumor onset and the self-renewal of prostate CSC-like cells [90].

2.4. Breast Cancer

Serum IL-6 levels in breast cancer patients have been reported to correlate with a poor prognosis and metastasis [91,92]. In contrast, local intratumoral autocrine/paracrine IL-6 signaling is crucial for regulating breast cancer cell proliferation, metastasis, and cancer stem cell self-renewal [93,94]. Augmentation of the IL-6-meditaed inflammatory loop induces resistance to trastuzumab, a HER2-targeted therapy used for HER2-positive breast cancer, by expanding the CSC population [95]. IL-6/STAT3 signaling is required for the maintenance of breast CSCs and tumor growth [31]. In particular, the IL-6/STAT3 pathway was found to be preferentially active in CD44+CD24 breast cancer cells, which have stem-cell-like characteristics, compared with other tumor cell types, and the inhibition of JAK2 decreased their number and blocked the growth of xenografts [31]. In addition, high levels of IL-6 were associated with resistance to paclitaxel in patients with malignant breast cancer [96]. However, the activation of STAT3 does not appear to be an independent marker of breast cancer prognosis [97,98]. Notably, the upregulated expression of IL-11 and the gp130-STAT3 pathway are implicated in the bone metastasis of breast cancer [99]. Although the activation of the IL-6/STAT3 signal has been primarily identified as being necessary for the proliferation of several types of CSCs, tumor-derived erythropoietin, mainly released under hypoxic conditions, also activates the JAK2–STAT3 axis in breast CSCs and promotes self-renewal [100].

2.5. Head and Neck Cancer

Increased expression levels of IL-6 and its receptor have been shown to contribute to poor prognosis in patients with head and neck cancer (HNSCC) [101,102]. Consistent with these findings, STAT3 signaling was found to be hyperactivated in HNSCC and to lead to poor outcomes, but STAT3 mutations are rarely detected [35,103]. Mutations in protein tyrosine phosphatase receptors (PTPRs), such as PTPRT and PTPRD, appear to frequently occur in HNSCC, indicating one cause of the STAT3 hyperactivation in HNSCC [104,105]. STAT3 signaling is a crucial pathway for the regulation of gene expression that promotes cell proliferation and survival as well as for the expression of growth factors and cytokines (such as IL-6, IL-10, VEGF, and TGFβ) that drive immune suppression [35].
EGFR, which acts upstream of the STAT3 signaling pathway, is overexpressed in 80% to 90% of HNSCC tumors and is linked to an overall decrease in survival and progression-free survival [106,107]. This finding led to the approval of the anti-EGFR monoclonal antibody cetuximab for the treatment of HNSCC. In addition, other receptor tyrosine kinases, such as HER2 and MET, are overexpressed in HNSCCs, and their overexpression may be associated with the resistance of HNSCCs to EGFR-targeted drugs that act via the activation of STAT3 and its gene targets [108,109,110].
To date, it has been well-documented that EMT is commonly involved in the acquisition of invasiveness and metastatic potential in malignant HNSCC tumors [111,112]. Mechanistically, IL-6 induces EMT changes in HNSCC cells via the activation of STAT3 signaling [113]. Additionally, cytokines and growth factors in the TME, particularly IL-6, EGF, and hepatocyte growth factor, suppress anoikis by activating tumor cell signaling pathways, including the RAS-MAPK, PI3K-mechanistic target of rapamycin kinase, and STAT3 pathways [114,115,116]. Notably, anoikis suppressors in the TME are produced by infiltrating immune cells, CAFs, endothelial cells, and tumor cells themselves [114], suggesting highly complicated crosstalk between the various cell types that contribute to metastasis in HNSCC.

2.6. Lung Cancer

Lung cancer is the leading cause of cancer-associated deaths worldwide, and the most common type of lung cancer is non-small-cell lung cancer (NSCLC), accounting for 85% of all lung cancer cases [117]. The STAT3-activating cytokine IL-6 is upregulated in the serum and exhaled breath condensate of NSCLC patients and correlates with a higher risk of metastasis and chemotherapy resistance [118,119,120,121,122,123,124,125]. Increased IL-11 expression has also been detected in the serum, tumors, and exhaled breath condensate of NSCLC patients and is associated with a higher risk of metastasis [126,127]. A high expression level of OSM is associated with poor outcomes in patients with NSCLC and enhances the EMT of NSCLC cells [128]. In addition, a sustained activation of STAT3 occurs in more than 50% of NSCLC patients [129,130], and its increased expression leads to low-grade tumor differentiation, lymph node metastasis, clinical stage progression, and drug resistance [131,132,133]. Mutations in receptor tyrosine kinases, such as EGFR, and Src family proteins have been associated with the constitutive activation of STAT3 in NSCLC [133,134], and STAT3 activation has been associated with lymph node metastasis and clinical stage progression and is an independent prognostic factor of NSCLC [135,136]. To date, the tumor-promoting functions mediated by STAT3 signaling in NSCLC have been well-documented to promote cell survival, angiogenesis, drug resistance, cancer cell stemness, and cancer immune evasion [117]. As a result, highly increased STAT3 expression enhanced the proliferation, survival, and radioresistance of NSCLC cells [132], whereas dominant-negative STAT3 resulted in the suppression of human lung cancer cell proliferation and invasive potential [137].
JAK-STAT3 signaling occurs during the early adaptive response to EGFR-tyrosine kinase inhibitor (TKI) therapy in EGFR-mutant NSCLC and may occur together with the downstream signaling of NF-κB activation [138]. In preclinical NSCLC models, such as patient-derived tumor xenograft models and cell lines, response rates to EGFR TKI therapy were improved by the addition of JAK or STAT3 inhibitors [138,139,140,141]. IL-6 autocrine signaling by tumor cells enhanced the activation of the JAK-STAT3 signaling pathway, whereas the addition of neutralizing anti-IL-6 antibodies reduced tumor growth in a mouse model [134,142]. Nevertheless, early clinical trials showed only a 5% response rate to treatment with the JAK inhibitor ruxolitinib in combination with erlotinib in patients who showed cancer progression during their prior treatment with erlotinib, suggesting that treatment with these drug combinations is not able to reverse previously established drug resistance [143]. Because early adaptive activation of JAK-STAT3 signaling was observed in preclinical models in response to EGFR TKI treatment [134,141], a combination of a JAK and/or STAT3 inhibitor and an EGFR TKI may be necessary for therapeutic efficacy [133,137]. Therefore, the JAK inhibitor INCB39110 has been investigated for its use as a treatment in combination with the third-generation EGFR KI osimertinib in patients with the EGFR-T790M mutation, which is a secondary site mutation in which methionine is substituted for threonine at position 790 that is found in more than 50% of patients with acquired resistance to EGFR TKIs, such as erlotinib and gefitinib. [144]. The coactivation of STAT3 and Yes1 associated transcriptional regulator (YAP1) has also been associated with the promotion of tumor cell survival after EGFR TKI treatment, and the co-inhibition of EGFR, STAT3, and Src-YAP1 signaling demonstrates a more effective synergistic effect than the single use of an EGFR TKI [139].

3. STAT3 in Cells of the TME

The TME of most cancers is rich in immune cells, immunosuppressive, and often affected by the complex immunomodulatory actions of IL-6 family cytokines. IL-6/STAT3 signaling in Th1 cells has the most obvious effect on the TME by suppressing cell-mediated antitumor immunity, whereas chronic inflammation contributes to the promotion of tumor progression and the dysregulation of angiogenesis and affects the recruitment, retention, and infiltration of leukocytes as well as immune responses via the activation of IL-6/STAT3 signaling in multifaceted innate and adaptive immune cells and nonimmune cells, such as cancer-associated fibroblasts and endothelial cells. In particular, during tumor initiation, the IL-6/STAT3 signal promotes the generation of pathogenic Th17 cells and MDSCs, suppressing antigen-presenting dendritic cells and antitumor cytotoxic CD8+ T cells and promoting regulatory T (Treg)-cell activity and tumor-associated macrophage phenotype switching from tumorigenic M1-type traits to immunosuppressive M2-type traits. Similarly, IL-11/STAT3 signaling facilitates inflammation-associated tumorigenesis in the gastrointestinal tract and polarizes T cells and macrophages to a more immunosuppressive phenotype [11,145,146,147]. In addition to its effects on tumor-associated immune cells, the enhanced activity of STAT3 via IL-6 family cytokines on CAFs is of great interest regarding its indirect tumor-promoting effects (Figure 1).

3.1. Immune Cells

STAT3 promotes immunosuppressive effects on the functions of CD8+ effector cells. The ablation of STAT3 in hematopoietic cell lineages facilitates antitumor immunity to inhibit tumor proliferation and metastasis via enhancing the functions of CD8+ T cells, natural killer (NK) cells, dendritic cells, and neutrophils in a murine model of melanoma, suggesting that STAT3 signaling may suppress tumor immune surveillance systems [148]. In addition, in a mouse transplant model, a loss of STAT3 signaling in the hematopoietic compartment facilitated the recruitment of tumor-infiltrating effector T cells and attenuated the infiltration of Treg cells. Consistently, the activation of STAT3 has been shown to be crucial for restricting the recruitment and activation of CD8+ T cells that are required to prevent the progression of melanoma [149,150]. Similarly, constitutively active STAT3 in human CD4+ T cells suppresses antitumor immunity by blocking the production of granzyme B, tumor necrosis factor (TNF), interferon gamma (IFNγ), IL-13, and other inflammatory cytokines [151].
IL-6/STAT3 signaling in CD4+ T cells is crucial for the differentiation of Th17 cells via the expression of RAR-related orphan receptor gamma [152,153,154]. Th17 cells comprise approximately 5% of the CD4+ T cells in PDACs. The role of Th17 cells in the TME is also context-dependent. In PDAC, IL-17 secretion from γδ T cells and Th17 cells may enhance antitumor immunity [155]. However, early in PDAC carcinogenesis, IL-17 has a direct mitogenic effect on KRAS-mutation-induced PanIN cells expressing IL-17R [156]. Although the effects of distinct T-cell subsets depend on the underlying immune context of the tumor due to various physiological conditions and environments and may be altered during the tumor progression of PDAC, the regulation of the differentiation and function of T cells in PDAC TMEs plays a crucial role in tumor immunity.
STAT3 has been demonstrated to be a transcriptional activator of the immune checkpoint molecules cytotoxic T-lymphocyte-associated protein 4 (CTLA4), programmed cell death 1 (PDCD1, also known as PD-1), and CD274 (also known as PD-L1) in T cells. Consistently, the promoter region of the PD-1 gene contains STAT3 binding sites, and PD-1 expression is promoted in response to signaling via the T-cell receptor/nuclear factor of activated T cells, IL-6/STAT3, and IL-12/STAT4 [157]. Notably, it has been demonstrated that PD-1 signaling is activated via STAT3 on CD4+ T cells and promotes collagen production by fibroblasts in pulmonary fibrosis, indicating the role of STAT3 in immunosuppression and tumor-promoting responses in the TME [158].
It has been well-documented that IL-6 inhibits the TGF-β-induced generation of Treg cells [159] via the IL-6/STAT3-mediated direct suppression of forkhead box P3 (FOXP3), which is a key transcriptional regulator of Treg cell differentiation on naive T cells [160]. Importantly, these inhibitory effects of IL-6/STAT3 signaling are restricted to inducible Treg cells and have no effects on the differentiation and function of natural Treg cells [161]. Interestingly, in the TME, IL-10-mediated STAT3 activation promotes Treg cell differentiation and enhances CTLA4 expression [162]. STAT3 can also activate FOXP3 gene expression, resulting in the promotion of Treg differentiation [163,164]. Therefore, STAT3 can regulate the peripheral immunity and tolerance of effector T cells in the TME, indicating that STAT3 is a potential therapeutic target to suppress the formation of an immunosuppressive TME.
The existence of M2 macrophages in the TME has been shown to be associated with poor outcomes in patients with most types of solid tumors [165]. Excess STAT3 activation promotes the polarization of M2 macrophages and increases the expression levels of arginase-1 (ARG-1), Fos-related antigen 1, TGF-β, IL-10, and VEGF-a, which are M2-associated markers [166]. M2-polarized macrophages have been implicated in the promotion of tumor growth of melanoma and Lewis lung cancer [166]. M2 macrophages can also facilitate STAT3 signaling in breast cancer cells to promote tumor proliferation [167], and STAT3 activation in M2-polarized macrophages can activate STAT3 signaling in ovarian cancer cells to promote their growth via the production of IL-6 and IL-10 [168]. In the TME, the Th2 cell cytokine IL-4 can also be activated by macrophages to promote their growth and induce STAT3-dependent cathepsin secretion by macrophages, supporting the development of pancreatic neuroendocrine tumors [169]. The TLR-induced increase in the activation of STAT3 has been shown to increase PD-L1 levels in M2 macrophages [170].
MDSCs, which are derived from pathologically activated neutrophils and monocytes, have been implicated in the promotion of immunosuppressive effects on antitumor immune cells in the TME [171]. STAT3 promotes the differentiation and expansion of MDSCs, thereby enhancing their ability to suppress effector T cells and promote the differentiation of Treg cells, which enable the augmentation of tumor formation [172]. The knockdown of STAT3 in MDSCs derived from patients with prostate cancer reduced the immunosuppressive functions of MDSCs against effector T cells by STAT3 [173]. Notably, a mouse model of acute colitis bearing the Y757F point mutation in murine gp130 (gp130Y757F/Y757F), which abrogates the SOCS3- and/or SHIP2-mediated negative feedback loop of the IL-6/STAT3 signal, resulting in the hyperactivation of STAT3 and has been shown to be resistant to colorectal damage and weight loss. This effect was shown to be produced by a small number of STAT3-induced granulocytic MDSCs (gMDSCs; also known as polymorphonuclear MDSCs [171]) with high expression of Arg1 and anti-inflammatory Th2 cytokines, such as IL-10 and TGF-β, [174], indicating that STAT3 promotes a precancerous host defense response during ulcerative colitis. In addition, gMDSCs are enriched by the local presence of IL-11, which activates STAT3 in CRC [175].
To limit the destructive responses of neutrophils, STAT3 necessarily acts as a negative regulator of neutrophil functions, suppresses the production of antitumor Th1 cytokines, such as IL-1, TNF, and IFNγ [176], and causes the unresponsiveness of neutrophils to chemotaxis by CXC-chemokine receptor 2 ligands [177,178]. Although neutrophils are necessary to limit the destructive inflammatory effects on the host, they may also facilitate tumorigenesis. In fact, the abrogation of STAT3 in neutrophils enhanced the cytolytic activity of neutrophils and induced tumor regression [148]. Similarly, STAT3 deletion in NK cells increased the expression of cytotoxic factors, such as perforin and granzyme B, and the activation of the NK cell marker CD226 [179]. On the other hand, the mutation-driven excessive STAT3 activity in NK cells from patients with chronic lymphoproliferative disorders of NK cells, T-cell large granular lymphocytic leukemias [180], and NK/T-cell lymphomas of the nasal type [181] promotes lymphomagenesis and provides a tumor-promoting TME. Thus, it is clear that the dysregulation of STAT3 in innate immune cells augments cancer cell proliferation and inhibits antitumor responses via the immunosurveillance system.

3.2. Non-Immune Cells

Fibroblasts exist in every solid organ to maintain their morphology and function by depositing extracellular matrix proteins and secreting soluble factors [182]. Histological similarities, such as mesenchymal morphology, are maintained among fibroblasts in various organs, but their genomic landscapes differ depending on the organs in which they are located [183]. It has been demonstrated that some fibroblasts contribute to tumor initiation, progression, and metastasis [184]. STAT3 signaling in CAFs [185] and tumor cells [185,186] may induce stromal remodeling of the TME characterized by fibrogenesis, a dysregulated organization of the ECM, and fibroblast contractility, which promote tumor cell motility, invasive activity, and resistance to chemical and immunological therapies [187]. However, the constitutive activation of STAT3 via the Y757F point mutation in pg130 in a mouse pulmonary fibrosis model not only promotes fibrosis in the absence of the TGF-β signaling molecule SMAD3, which is well-known to be crucial in the pathogenesis of lung fibrosis [188], but also results in desmoplasia formation and epithelial stiffness, which enhance tumorigenesis in PDAC mouse models [186]. CAFs assist tumor growth and dissemination through the production of factors such as EGF, IL-6, TGF-β, and VEGF, which promote tumor cell proliferation and angiogenesis [189,190]. The protumorigenic characteristics of these fibroblasts are partially modulated by STAT3 activity via the induction of various cytokines, including LIF [191]. For example, STAT3 acts as a downstream signaling molecule of the focal adhesion kinase-Src-JAK2 axis in CAFs, leading to the expression of CCL2 and immunosuppression. Mechanistic studies have shown that CAFs promote the growth of murine hepatocellular carcinomas by resulting in the mobilization of MDSCs [192]. Factors produced by CAFs can promote STAT3 signaling in other cell types, support intercellular communication between immune cells within the TME, and induce immunosuppression [193,194].

4. Promising Target Molecules in STAT3-Associated Tumors

4.1. STAP-2

Signal-transducing adaptor protein-2 (STAP-2) was originally identified as a c-Fms/macrophage colony-stimulating factor receptor-binding protein containing pleckstrin-homology (PH), SH2, and proline-rich domains. Interestingly, STAP-2 levels were strongly induced in the liver by lipopolysaccharide (LPS) stimulation and in isolated hepatocytes by IL-6 stimulation. Consistently, in STAP-2-deleted hepatocytes, the acute phase responses induced by IL-6 and the tyrosine phosphorylation levels of STAT3 were significantly decreased. Moreover, STAP-2 binds to activated STAT3 via a YXXQ motif in the C-terminal region, indicating that STAP-2 is an adaptor molecule that modulates STAT3 activity [195].
Breast tumor kinase (BRK, also known as protein tyrosine kinase 6), which is related to the Src family of tyrosine kinases, is overexpressed in approximately 85% of invasive ductal carcinomas [196]. STAP-2, which was the first BRK substrate to be identified [33], binds to BRK via its PH domain and contributes to the activation of STAT3 [197], and as the PH domain of STAP-2 is essential for the translocation of STAP-2 to the plasma membrane after EGF treatment, the binding ability of the PH domain of STAP-2 to BRK is considered to be an important biological characteristic [33]. Thus, the PH domain of STAP-2 may have a biological function in altering the subcellular localization of BRK and promoting its activation; STAP-2 acts as a scaffold protein that facilitates the interaction between BRK and STAT3. Taken together, the experiments using deletion mutants suggest that STAP-2 modulates multiple events, i.e., the binding of STAP-2 to BRK, the activation of BRK, and the subsequent promotion of the tyrosine phosphorylation of STAT3. Thus, STAP-2 functions in concert with BRK to promote breast cancer cell proliferation. As both BRK and STAP-2 expression are high in breast cancer cells, their coupling may promote the abnormal activation of STAT3. These data may provide insights into the molecular mechanisms and implications of the BRK/STAP-2/STAT3 interaction and may provide clues for the development of novel therapies for breast cancer.
STAP-2 enhances signaling via EGFR through its protein stabilization, leading to higher tumorigenesis in prostate cancer cells [198]. STAP-2 promotes EGFR signaling through a two-step process: the stabilization of EGFR and the subsequent activation of STAT3 through their direct interaction. In addition to EGFR, IL-6 signaling also activates STAT3, and IL6R blockade significantly inhibits tumor progression [199]. A knockdown of STAP-2 may inhibit prostate cancer cell growth by synergistically inhibiting EGFR and IL6R signaling. As described above, STAP-2 can bind to BRK and contributes to promoting the BRK-modulated activation of STAT3 and STAT5 [197,200]. In particular, BRK augments EGFR signaling by decreasing the casitas B-lineage lymphoma (CBL)-enhanced ubiquitination of EGFR [201]. The expression of cell-surface EGFR is important for the activation of RAS and ERK, and STAP-2 prevents the reduction in surface expression levels of EGFR, even after EGF treatment. Another biological mechanism of STAP-2 in prostate cancer cells is to inhibit the CBL-enhanced ubiquitination of EGFR and to restore EGFR [198]. A knockdown of STAP-2 inhibits the cell growth of prostate cancer cells [198]. STAP-2 is also a potent regulator of EGFR activation in prostate cancer cells. STAP-2 does not bind to EGFR K721A, a dimer-formation-deficient mutant, suggesting that STAP-2 interacts with EGFR after dimer formation [198]. Furthermore, STAP-2 acts to stabilize wild-type EGFR after EGF stimulation but not inactive EGFR mutants [198]. Indeed, the inhibition of tumor cell growth by STAP-2 knockdown occurs under EGFR-activated conditions but not under EGFR-inactivated conditions [198]. Notably, gefitinib treatment fails to further inhibit the cell proliferation of STAP-2-silenced prostate cancer cells [198]. The distinct regulatory mechanisms for EGFR surface expression in gefitinib-treated and STAP-2 knockdown cells demonstrates that the inhibition of STAP-2 function can destabilize both wild-type EGFR and gefitinib-resistant self-activated EGFR [198]. Therefore, STAP-2 inhibitors may have the potential to be effective anticancer agents against gefitinib-resistant prostate cancer (Figure 2).

4.2. ARID5A

Recent studies have demonstrated the crucial roles of Arid5a in inflammation, autoimmunity, and cancer [34,202]. Arid5a was identified as an RNA-binding protein that directly binds to a stem–loop element in the 3′-untranslated regions (UTRs) of IL-6 to stabilize IL-6 mRNA and augments IL-6 expression. Stimulation with LPS, IL-1β, or IL-6 leads to its expression in macrophages and embryonic fibroblasts. Arid5a has been shown to exacerbate symptoms in LPS-treated mice and experimental autoimmune encephalomyelitis (EAE) mice, accompanied with an increase in IL-6 levels [203]. The binding element of Arid5a in the IL-6 3′ UTR coincides with that of Regnase-1, resulting in Arid5a counteracting the destabilizing function of Regnase-1 on Il6 mRNA. Notably, in untreated rheumatoid arthritis (RA) patients, the expression of ARID5a in CD4+ T cells is enhanced, whereas treatment with the anti-IL6R antibody tocilizumab results in a decrease in the expression of Arid5a [204], indicating that the IL-6-ARID5A axis may be involved in the pathogenesis of RA. Importantly, Arid5a is deeply involved in the Th17-polarized differentiation of naïve CD4+ T cells via the stabilization of STAT3 mRNA, resulting in the development of EAE in mice [205].
Recently, the expression levels of Arid5a have been reported to be significantly increased in mesenchymal tumor subtypes of PDAC and CRC, such as the quasi-mesenchymal and consensus molecular subtype 4 subtypes, respectively [206]. In cells derived from the PDAC mouse model KPC (Pdx1-cre, KRASG12D, and p53R172H), the abrogation of Arid5a was shown to significantly downregulate EMT-TFs and EMT markers, such as Zeb1, Zeb2, Snai1, Snai2, Twist2, Acta2, and Itgb1, compared with WT KPC cells, whereas the expression of the representative epithelial marker E-cadherin (Cdh1 gene) was substantially increased [206]. Additionally, an ingenuity pathway analysis demonstrated that signaling pathways linked to EMT and metastasis, including the regulation of EMT by growth factors/development, IL-8, OSM, and stemness signals, were diminished upon the loss of Arid5a. In addition, the signaling pathways of IL-6, STAT3, and JAK/STAT were downregulated in Arid5a-deficient KPC cells [206]. Moreover, Arid5a expression was enhanced in in vitro EMT models induced by IL-6 and TGF-β stimulation [206], indicating the involvement of Arid5a in inducing the mesenchymal cell properties of PDAC. In agreement with these findings, a recent report indicated that the IL-6-Arid5a axis enhances the invasive and metastatic activities of breast cancer cells. Mechanistically, the Arid5a induced by IL-6 functions as a transcription factor, increasing the expression levels of the long non-coding RNA AU021063. Subsequently, AU021063 functions to activate breast cancer cell invasion and metastasis through the stabilization of tribbles homolog 3 mRNA and the activation of the Mek/Erk signaling pathway [207].
It has been shown that Arid5a enables mesenchymal tumor models of PDAC and CRC to facilitate immune evasiveness via promoting the tumor infiltration of immunosuppressive granulocytic MDSCs and Tregs and reducing the recruitment and activation of antitumor effector T cells [206]. Mechanistically, Arid5a functions as a dual regulator, leading to the formation of immunosuppressive TMEs in malignant tumors and triggering the metabolic reprograming and recruitment of suppressive immune cells. First, Arid5a promotes the inhibitory effect of Ido1 on effector CD4+/CD8+ T cells via the stabilization of Ido1 mRNA by binding to its 3′-UTR and by reducing the intratumoral tryptophan concentration [208,209]. Additionally, Ido1 expression in tumor tissues promotes Treg differentiation/activation by catabolizing tryptophan to produce kynurenine and ultimately activating aryl hydrocarbon receptors (AhR) [210,211]. AhR activation results in the extensive infiltration of gMDSCs to the TME [212]. Second, Arid5a upregulates chemokine Ccl2 expression in the TME via the stabilization of its mRNA and then Ccl2 enhances the infiltration of immunosuppressor cells, such as Tregs and gMDSCs [213,214,215,216], to the TME [206] (Figure 3).

5. Perspectives

In tumorigenesis, chronic inflammation and metabolic changes associated with genetic mutations in normal cells enable transformed cells to escape the immunological defense mechanisms of the tissue and to reprogram the functions of endogenous signaling mechanisms and cell populations in surrounding tissues, thereby disrupting the homeostatic balance of the entire organism and creating neoplasms within the organism. STAT3 plays a central role in this entire process. To date, a large amount of effort has been put into developing STAT3 inhibitors that both directly and indirectly target STAT3, including SH2 domain inhibitors and DNA-binding domain inhibitors, and JAK kinase inhibitors and Src inhibitors, respectively, and integrating STAT3-based combination immunotherapies [12,35] (Figure 1, Table 1 and Table 2). In particular, since its approval in 2009, tocilizumab has been used to inhibit IL-6/STAT3 signaling in autoimmune diseases, such as rheumatoid arthritis, caused by the overexpression of IL-6 and acute inflammatory diseases caused by cytokine storms associated with chimeric antigen receptor T-cell therapy and SARS-CoV-2 infection, and it has shown high therapeutic efficacy against various immune diseases. On the other hand, few effective therapies that target STAT3 signaling for the treatment of cancer in clinical practice have been developed [6,7,8,9,10] (Table 1 and Table 2). As mentioned above, cancer is a complex interplay of diverse cell populations that result in malignant transformation. Therefore, whereas analyses of the expression and function of molecules associated with STAT3 activation can assess the local and steady-state malignant potential of a cancer, they are insufficient to predict the stage or detailed course of a cancer. Furthermore, it is clear that in addition to STAT3 signaling various groups of molecules are involved in cancer development. The mode of interaction between these molecules also needs further study. In the future, it will be essential to perform spatiotemporal gene expression analyses to analyze multiple cell populations, improve technologies for the detection of aging and inflammation using artificial intelligence, and introduce mathematical analysis technology to integrate these technologies. Furthermore, it is also necessary to enhance the convergence of life science, physical science, engineering science, and computational science to create the next generation of cancer therapeutics.

Author Contributions

All authors summarized the literature, wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants-in-aid from the Ministry of Education, Science, Sports and Culture of Japan to SH (grant no. 22K07203) and TM (grant no. 19H03364).

Data Availability Statement

Not applicable.

Acknowledgments

We thank H.A. Popiel for his critical reading of the manuscript.

Conflicts of Interest

The authors declare no competing interests in association with this study.

References

  1. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Shalapour, S.; Karin, M. Pas de Deux: Control of Anti-tumor Immunity by Cancer-Associated Inflammation. Immunity 2019, 51, 15–26. [Google Scholar] [CrossRef] [PubMed]
  3. Bruni, D.; Angell, H.K.; Galon, J. The immune contexture and Immunoscore in cancer prognosis and therapeutic efficacy. Nat. Rev. Cancer 2020, 20, 662–680. [Google Scholar] [CrossRef]
  4. Akira, S.; Nishio, Y.; Inoue, M.; Wang, X.J.; Wei, S.; Matsusaka, T.; Yoshida, K.; Sudo, T.; Naruto, M.; Kishimoto, T. Molecular cloning of APRF, a novel IFN-stimulated gene factor 3 p91-related transcription factor involved in the gp130-mediated signaling pathway. Cell 1994, 77, 63–71. [Google Scholar] [CrossRef]
  5. Zhong, Z.; Wen, Z.; Darnell, J.E., Jr. Stat3: A STAT family member activated by tyrosine phosphorylation in response to epidermal growth factor and interleukin-6. Science 1994, 264, 95–98. [Google Scholar] [CrossRef] [PubMed]
  6. Jones, S.A.; Jenkins, B.J. Recent insights into targeting the IL-6 cytokine family in inflammatory diseases and cancer. Nat. Rev. Immunol. 2018, 18, 773–789. [Google Scholar] [CrossRef]
  7. Kang, S.; Tanaka, T.; Narazaki, M.; Kishimoto, T. Targeting Interleukin-6 Signaling in Clinic. Immunity 2019, 50, 1007–1023. [Google Scholar] [CrossRef]
  8. Murakami, M.; Kamimura, D.; Hirano, T. Pleiotropy and Specificity: Insights from the Interleukin 6 Family of Cytokines. Immunity 2019, 50, 812–831. [Google Scholar] [CrossRef] [Green Version]
  9. Hirano, T. IL-6 in inflammation, autoimmunity and cancer. Int. Immunol. 2021, 33, 127–148. [Google Scholar] [CrossRef]
  10. Kishimoto, T.; Kang, S. IL-6 Revisited: From Rheumatoid Arthritis to CAR T Cell Therapy and COVID-19. Annu. Rev. Immunol. 2022, 40, 323–348. [Google Scholar] [CrossRef]
  11. Yu, H.; Lee, H.; Herrmann, A.; Buettner, R.; Jove, R. Revisiting STAT3 signalling in cancer: New and unexpected biological functions. Nat. Rev. Cancer 2014, 14, 736–746. [Google Scholar] [CrossRef] [PubMed]
  12. Huynh, J.; Chand, A.; Gough, D.; Ernst, M. Therapeutically exploiting STAT3 activity in cancer–using tissue repair as a road map. Nat. Rev. Cancer 2019, 19, 82–96. [Google Scholar] [CrossRef] [PubMed]
  13. Kubo, M.; Hanada, T.; Yoshimura, A. Suppressors of cytokine signaling and immunity. Nat. Immunol. 2003, 4, 1169–1176. [Google Scholar] [CrossRef] [PubMed]
  14. Nicholson, S.E.; De Souza, D.; Fabri, L.J.; Corbin, J.; Willson, T.A.; Zhang, J.G.; Silva, A.; Asimakis, M.; Farley, A.; Nash, A.D.; et al. Suppressor of cytokine signaling-3 preferentially binds to the SHP-2-binding site on the shared cytokine receptor subunit gp130. Proc. Natl. Acad. Sci. USA 2000, 97, 6493–6498. [Google Scholar] [CrossRef] [Green Version]
  15. Boccaccio, C.; Ando, M.; Tamagnone, L.; Bardelli, A.; Michieli, P.; Battistini, C.; Comoglio, P.M. Induction of epithelial tubules by growth factor HGF depends on the STAT pathway. Nature 1998, 391, 285–288. [Google Scholar] [CrossRef]
  16. Quesnelle, K.M.; Boehm, A.L.; Grandis, J.R. STAT-mediated EGFR signaling in cancer. J. Cell. Biochem. 2007, 102, 311–319. [Google Scholar] [CrossRef]
  17. Lee, H.; Deng, J.; Kujawski, M.; Yang, C.; Liu, Y.; Herrmann, A.; Kortylewski, M.; Horne, D.; Somlo, G.; Forman, S.; et al. STAT3-induced S1PR1 expression is crucial for persistent STAT3 activation in tumors. Nat. Med. 2010, 16, 1421–1428. [Google Scholar] [CrossRef] [Green Version]
  18. Xin, H.; Lu, R.; Lee, H.; Zhang, W.; Zhang, C.; Deng, J.; Liu, Y.; Shen, S.; Wagner, K.U.; Forman, S.; et al. G-protein-coupled receptor agonist BV8/prokineticin-2 and STAT3 protein form a feed-forward loop in both normal and malignant myeloid cells. J. Biol. Chem. 2013, 288, 13842–13849. [Google Scholar] [CrossRef] [Green Version]
  19. 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] [Green Version]
  20. Kortylewski, M.; Kujawski, M.; Herrmann, A.; Yang, C.; Wang, L.; Liu, Y.; Salcedo, R.; Yu, H. Toll-like receptor 9 activation of signal transducer and activator of transcription 3 constrains its agonist-based immunotherapy. Cancer Res. 2009, 69, 2497–2505. [Google Scholar] [CrossRef] [Green Version]
  21. 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] [Green Version]
  22. Iliopoulos, D.; Hirsch, H.A.; Struhl, K. An epigenetic switch involving NF-κB, Lin28, Let-7 MicroRNA, and IL6 links inflammation to cell transformation. Cell 2009, 139, 693–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Guo, L.; Chen, C.; Shi, M.; Wang, F.; Chen, X.; Diao, D.; Hu, M.; Yu, M.; Qian, L.; Guo, N. Stat3-coordinated Lin-28-let-7-HMGA2 and miR-200-ZEB1 circuits initiate and maintain oncostatin M-driven epithelial-mesenchymal transition. Oncogene 2013, 32, 5272–5282. [Google Scholar] [CrossRef] [PubMed]
  24. Sugimura, K.; Miyata, H.; Tanaka, K.; Hamano, R.; Takahashi, T.; Kurokawa, Y.; Yamasaki, M.; Nakajima, K.; Takiguchi, S.; Mori, M.; et al. Let-7 expression is a significant determinant of response to chemotherapy through the regulation of IL-6/STAT3 pathway in esophageal squamous cell carcinoma. Clin. Cancer Res. 2012, 18, 5144–5153. [Google Scholar] [CrossRef] [Green Version]
  25. Navarro, A.; Diaz, T.; Martinez, A.; Gaya, A.; Pons, A.; Gel, B.; Codony, C.; Ferrer, G.; Martinez, C.; Montserrat, E.; et al. Regulation of JAK2 by miR-135a: Prognostic impact in classic Hodgkin lymphoma. Blood 2009, 114, 2945–2951. [Google Scholar] [CrossRef]
  26. Du, L.; Subauste, M.C.; DeSevo, C.; Zhao, Z.; Baker, M.; Borkowski, R.; Schageman, J.J.; Greer, R.; Yang, C.R.; Suraokar, M.; et al. miR-337-3p and its targets STAT3 and RAP1A modulate taxane sensitivity in non-small cell lung cancers. PLoS ONE 2012, 7, e39167. [Google Scholar] [CrossRef]
  27. Priceman, S.J.; Kujawski, M.; Shen, S.; Cherryholmes, G.A.; Lee, H.; Zhang, C.; Kruper, L.; Mortimer, J.; Jove, R.; Riggs, A.D.; et al. Regulation of adipose tissue T cell subsets by Stat3 is crucial for diet-induced obesity and insulin resistance. Proc. Natl. Acad. Sci. USA 2013, 110, 13079–13084. [Google Scholar] [CrossRef] [Green Version]
  28. Deng, J.; Liu, Y.; Lee, H.; Herrmann, A.; Zhang, W.; Zhang, C.; Shen, S.; Priceman, S.J.; Kujawski, M.; Pal, S.K.; et al. S1PR1-STAT3 signaling is crucial for myeloid cell colonization at future metastatic sites. Cancer Cell 2012, 21, 642–654. [Google Scholar] [CrossRef] [Green Version]
  29. Park, E.J.; Lee, J.H.; Yu, G.Y.; He, G.; Ali, S.R.; Holzer, R.G.; Osterreicher, C.H.; Takahashi, H.; Karin, M. Dietary and genetic obesity promote liver inflammation and tumorigenesis by enhancing IL-6 and TNF expression. Cell 2010, 140, 197–208. [Google Scholar] [CrossRef] [Green Version]
  30. Carro, M.S.; Lim, W.K.; Alvarez, M.J.; Bollo, R.J.; Zhao, X.; Snyder, E.Y.; Sulman, E.P.; Anne, S.L.; Doetsch, F.; Colman, H.; et al. The transcriptional network for mesenchymal transformation of brain tumours. Nature 2010, 463, 318–325. [Google Scholar] [CrossRef]
  31. Marotta, L.L.; Almendro, V.; Marusyk, A.; Shipitsin, M.; Schemme, J.; Walker, S.R.; Bloushtain-Qimron, N.; Kim, J.J.; Choudhury, S.A.; Maruyama, R.; et al. The JAK2/STAT3 signaling pathway is required for growth of CD44+CD24- stem cell-like breast cancer cells in human tumors. J. Clin. Investig. 2011, 121, 2723–2735. [Google Scholar] [CrossRef] [PubMed]
  32. Schroeder, A.; Herrmann, A.; Cherryholmes, G.; Kowolik, C.; Buettner, R.; Pal, S.; Yu, H.; Müller-Newen, G.; Jove, R. Loss of androgen receptor expression promotes a stem-like cell phenotype in prostate cancer through STAT3 signaling. Cancer Res. 2014, 74, 1227–1237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Matsuda, T.; Oritani, K. STAP-2 adaptor protein regulates multiple steps of immune and inflammatory responses. Biol. Pharm. Bull. 2021, 44, 895–901. [Google Scholar] [CrossRef]
  34. Hashimoto, S.; Kishimoto, T. Roles of RNA-binding proteins in immune diseases and cancer. Semin. Cancer Biol. 2022; in press. [Google Scholar] [CrossRef]
  35. Johnson, D.E.; O’Keefe, R.A.; Grandis, J.R. Targeting the IL-6/JAK/STAT3 signalling axis in cancer. Nat. Rev. Clin. Oncol. 2018, 15, 234–248. [Google Scholar] [CrossRef] [PubMed]
  36. Li, N.; Grivennikov, S.I.; Karin, M. The unholy trinity: Inflammation, cytokines, and STAT3 shape the cancer microenvironment. Cancer Cell 2011, 19, 429–431. [Google Scholar] [CrossRef] [Green Version]
  37. Löffler, D.; Brocke-Heidrich, K.; Pfeifer, G.; Stocsits, C.; Hackermüller, J.; Kretzschmar, A.K.; Burger, R.; Gramatzki, M.; Blumert, C.; Bauer, K.; et al. Interleukin-6 dependent survival of multiple myeloma cells involves the Stat3-mediated induction of microRNA-21 through a highly conserved enhancer. Blood 2007, 110, 1330–1333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Jenkins, B.J.; Grail, D.; Nheu, T.; Najdovska, M.; Wang, B.; Waring, P.; Inglese, M.; McLoughlin, R.M.; Jones, S.A.; Topley, N.; et al. Hyperactivation of Stat3 in gp130 mutant mice promotes gastric hyperproliferation and desensitizes TGF-beta signaling. Nat. Med. 2005, 11, 845–852. [Google Scholar] [CrossRef]
  39. Taniguchi, K.; Karin, M. IL-6 and related cytokines as the critical lynchpins between inflammation and cancer. Semin. Immunol. 2014, 26, 54–74. [Google Scholar] [CrossRef]
  40. Sullivan, N.J.; Sasser, A.K.; Axel, A.E.; Vesuna, F.; Raman, V.; Ramirez, N.; Oberyszyn, T.M.; Hall, B.M. Interleukin-6 induces an epithelial-mesenchymal transition phenotype in human breast cancer cells. Oncogene 2009, 28, 2940–2947. [Google Scholar] [CrossRef] [Green Version]
  41. Nieto, M.A. Epithelial Plasticity: A Common Theme in Embryonic and Cancer Cells. Science 2013, 342, 1234850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Lu, W.; Kang, Y. Epithelial-Mesenchymal Plasticity in Cancer Progression and Metastasis. Dev. Cell 2019, 49, 361–374. [Google Scholar] [CrossRef]
  43. Lambert, A.W.; Weinberg, R.A. Linking Emt Programmes to Normal and Neoplastic Epithelial Stem Cells. Nat. Rev. Cancer 2021, 21, 325–338. [Google Scholar] [CrossRef] [PubMed]
  44. Gawlik-Rzemieniewska, N.; Bednarek, I. The role of NANOG transcriptional factor in the development of malignant phenotype of cancer cells. Cancer Biol. Ther. 2016, 17, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kim, S.Y.; Kang, J.W.; Song, X.; Kim, B.K.; Yoo, Y.D.; Kwon, Y.T.; Lee, Y.J. Role of the IL-6-JAK1-STAT3-Oct-4 pathway in the conversion of non-stem cancer cells into cancer stem-like cells. Cell Signal. 2013, 25, 961–969. [Google Scholar] [CrossRef] [Green Version]
  46. Mitsunaga, S.; Ikeda, M.; Shimizu, S.; Ohno, I.; Furuse, J.; Inagaki, M.; Higashi, S.; Kato, H.; Terao, K.; Ochiai, A. Serum levels of IL-6 and IL-1beta can predict the efficacy of gemcitabine in patients with advanced pancreatic cancer. Br. J. Cancer 2013, 108, 2063–2069. [Google Scholar] [CrossRef] [Green Version]
  47. Denley, S.M.; Jamieson, N.B.; McCall, P.; Oien, K.A.; Morton, J.P.; Carter, C.R.; Edwards, J.; McKay, C.J. Activation of the IL-6R/Jak/stat pathway is associated with a poor outcome in resected pancreatic ductal adenocarcinoma. J. Gastrointestin. Surg. 2013, 17, 887–898. [Google Scholar] [CrossRef]
  48. Scholz, A.; Heinze, S.; Detjen, K.M.; Peters, M.; Welzel, M.; Hauff, P.; Schirner, M.; Wiedenmann, B.; Rosewicz, S. Activated signal transducer and activator of transcription 3 (STAT3) supports the malignant phenotype of human pancreatic cancer. Gastroenterology 2003, 125, 891–905. [Google Scholar] [CrossRef]
  49. Huang, C.; Huang, R.; Chang, W.; Jiang, T.; Huang, K.; Cao, J.; Sun, X.; Qiu, Z. The expression and clinical significance of pSTAT3, VEGF and VEGF-C in pancreatic adenocarcinoma. Neoplasma 2012, 59, 52–61. [Google Scholar] [CrossRef] [Green Version]
  50. Huang, C.; Yang, G.; Jiang, T.; Zhu, G.; Li, H.; Qiu, Z. The effects and mechanisms of blockage of STAT3 signaling pathway on IL-6 inducing EMT in human pancreatic cancer cells in vitro. Neoplasma 2011, 58, 396–405. [Google Scholar] [CrossRef] [Green Version]
  51. Miyatsuka, T.; Kaneto, H.; Shiraiwa, T.; Matsuoka, T.A.; Yamamoto, K.; Kato, K.; Nakamura, Y.; Akira, S.; Takeda, K.; Kajimoto, Y.; et al. Persistent expression of PDX-1 in the pancreas causes acinar-to-ductal metaplasia through Stat3 activation. Genes Dev. 2006, 20, 1435–1440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lesina, M.; Kurkowski, M.U.; Ludes, K.; Rose-John, S.; Treiber, M.; Klöppel, G.; Yoshimura, A.; Reindl, W.; Sipos, B.; Akira, S.; et al. Stat3/Socs3 activation by IL-6 transsignaling promotes progression of pancreatic intraepithelial neoplasia and development of pancreatic cancer. Cancer Cell 2011, 19, 456–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Fukuda, A.; Wang, S.C.; Morris, J.P., 4th; Folias, A.E.; Liou, A.; Kim, G.E.; Akira, S.; Boucher, K.M.; Firpo, M.A.; Mulvihill, S.J.; et al. Stat3 and MMP7 contribute to pancreatic ductal adenocarcinoma initiation and progression. Cancer Cell 2011, 19, 441–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yang, S.; Wang, X.; Contino, G.; Liesa, M.; Sahin, E.; Ying, H.; Bause, A.; Li, Y.; Stommel, J.M.; Dell’antonio, G.; et al. Pancreatic cancers require autophagy for tumor growth. Genes Dev. 2011, 25, 717–729. [Google Scholar] [CrossRef] [Green Version]
  55. Kamphorst, J.J.; Nofal, M.; Commisso, C.; Hackett, S.R.; Lu, W.; Grabocka, E.; Vander Heiden, M.G.; Miller, G.; Drebin, J.A.; Bar-Sagi, D.; et al. Human pancreatic cancer tumors are nutrient poor and tumor cells actively scavenge extracellular protein. Cancer Res. 2015, 75, 544–553. [Google Scholar] [CrossRef] [Green Version]
  56. Kang, R.; Loux, T.; Tang, D.; Schapiro, N.E.; Vernon, P.; Livesey, K.M.; Krasinskas, A.; Lotze, M.T.; Zeh, H.J., 3rd. The expression of the receptor for advanced glycation endproducts (RAGE) is permissive for early pancreatic neoplasia. Proc. Natl. Acad. Sci. USA 2012, 109, 7031–7036. [Google Scholar] [CrossRef] [Green Version]
  57. Rose-John, S.; Mitsuyama, K.; Matsumoto, S.; Thaiss, W.M.; Scheller, J. Interleukin-6 trans-signaling and colonic cancer associated with inflammatory bowel disease. Curr. Pharm. Des. 2009, 15, 2095–2103. [Google Scholar] [CrossRef]
  58. Belluco, C.; Nitti, D.; Frantz, M.; Toppan, P.; Basso, D.; Plebani, M.; Lise, M.; Jessup, J.M. Interleukin-6 blood level is associated with circulating carcinoembryonic antigen and prognosis in patients with colorectal cancer. Ann. Surg. Oncol. 2000, 7, 133–138. [Google Scholar] [CrossRef]
  59. Chung, Y.C.; Chang, Y.F. Serum interleukin-6 levels reflect the disease status of colorectal cancer. J. Surg. Oncol. 2003, 83, 222–226. [Google Scholar] [CrossRef]
  60. Kusaba, T.; Nakayama, T.; Yamazumi, K.; Yakata, Y.; Yoshizaki, A.; Inoue, K.; Nagayasu, T.; Sekine, I. Activation of STAT3 is a marker of poor prognosis in human colorectal cancer. Oncol. Rep. 2006, 15, 1445–1451. [Google Scholar] [CrossRef] [Green Version]
  61. Morikawa, T.; Baba, Y.; Yamauchi, M.; Kuchiba, A.; Nosho, K.; Shima, K.; Tanaka, N.; Huttenhower, C.; Frank, D.A.; Fuchs, C.S.; et al. STAT3 expression, molecular features, inflammation patterns, and prognosis in a database of 724 colorectal cancers. Clin. Cancer Res. 2011, 17, 1452–1462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Jin, C.; Wang, A.; Chen, J.; Liu, X.; Wang, G. Relationship between expression and prognostic ability of PTEN, STAT3 and VEGF-C in colorectal cancer. Exp. Ther Med. 2012, 4, 633–639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Becker, C.; Fantini, M.C.; Schramm, C.; Lehr, H.A.; Wirtz, S.; Nikolaev, A.; Burg, J.; Strand, S.; Kiesslich, R.; Huber, S.; et al. TGF-beta suppresses tumor progression in colon cancer by inhibition of IL-6 trans-signaling. Immunity 2004, 21, 491–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Rose-John, S. IL-6 trans-signaling via the soluble IL-6 receptor: Importance for the pro-inflammatory activities of IL-6. Int. J. Biol. Sci. 2012, 8, 1237–1247. [Google Scholar] [CrossRef] [PubMed]
  65. Sepich-Poore, G.D.; Zitvogel, L.; Straussman, R.; Hasty, J.; Wargo, J.A.; Knight, R. The microbiome and human cancer. Science 2021, 371, eabc4552. [Google Scholar] [CrossRef]
  66. Hu, B.; Elinav, E.; Huber, S.; Strowig, T.; Hao, L.; Hafemann, A.; Jin, C.; Wunderlich, C.; Wunderlich, T.; Eisenbarth, S.C.; et al. Microbiota-induced activation of epithelial IL-6 signaling links inflammasome-driven inflammation with transmissible cancer. Proc. Natl. Acad. Sci. USA 2013, 110, 9862–9867. [Google Scholar] [CrossRef] [Green Version]
  67. Becker, C.; Fantini, M.C.; Wirtz, S.; Nikolaev, A.; Lehr, H.A.; Galle, P.R.; Rose-John, S.; Neurath, M.F. IL-6 signaling promotes tumor growth in colorectal cancer. Cell Cycle 2005, 4, 217–220. [Google Scholar] [CrossRef]
  68. Grivennikov, S.; Karin, E.; Terzic, J.; Mucida, D.; Yu, G.Y.; Vallabhapurapu, S.; Scheller, J.; Rose-John, S.; Cheroutre, H.; Eckmann, L.; et al. IL-6 and Stat3 are required for survival of intestinal epithelial cells and development of colitis-associated cancer. Cancer Cell 2009, 15, 103–113. [Google Scholar] [CrossRef] [Green Version]
  69. Bollrath, J.; Phesse, T.J.; von Burstin, V.A.; Putoczki, T.; Bennecke, M.; Bateman, T.; Nebelsiek, T.; Lundgren-May, T.; Canli, O.; Schwitalla, S.; et al. gp130-mediated Stat3 activation in enterocytes regulates cell survival and cell-cycle progression during colitis-associated tumorigenesis. Cancer Cell 2009, 15, 91–102. [Google Scholar] [CrossRef] [Green Version]
  70. Putoczki, T.L.; Thiem, S.; Loving, A.; Busuttil, R.A.; Wilson, N.J.; Ziegler, P.K.; Nguyen, P.M.; Preaudet, A.; Farid, R.; Edwards, K.M.; et al. Interleukin-11 is the dominant IL-6 family cytokine during gastrointestinal tumorigenesis and can be targeted therapeutically. Cancer Cell 2013, 24, 257–271. [Google Scholar] [CrossRef] [Green Version]
  71. Fenton, J.I.; Hursting, S.D.; Perkins, S.N.; Hord, N.G. Interleukin-6 production induced by leptin treatment promotes cell proliferation in an Apc (Min/+) colon epithelial cell line. Carcinogenesis 2006, 27, 1507–1515. [Google Scholar] [CrossRef] [Green Version]
  72. Musteanu, M.; Blaas, L.; Mair, M.; Schlederer, M.; Bilban, M.; Tauber, S.; Esterbauer, H.; Mueller, M.; Casanova, E.; Kenner, L.; et al. Stat3 is a negative regulator of intestinal tumor progression in Apc(Min) mice. Gastroenterology 2010, 138, 1003–1011.e5. [Google Scholar] [CrossRef]
  73. Lee, J.; Kim, J.C.; Lee, S.E.; Quinley, C.; Kim, H.; Herdman, S.; Corr, M.; Raz, E. Signal transducer and activator of transcription 3 (STAT3) protein suppresses adenoma-to- carcinoma transition in Apcmin/+ mice via regulation of Snail-1 (SNAI) protein stability. J. Biol. Chem. 2012, 287, 18182–18189. [Google Scholar] [CrossRef] [Green Version]
  74. Ernst, M.; Putoczki, T.L. Targeting IL-11 signaling in colon cancer. Oncotarget 2013, 4, 1860–1861. [Google Scholar] [CrossRef] [Green Version]
  75. Dong, C. TH17 cells in development: An updated view of their molecular identity and genetic programming. Nat. Rev. Immunol. 2008, 8, 337–348. [Google Scholar] [CrossRef]
  76. Grivennikov, S.I.; Wang, K.; Mucida, D.; Stewart, C.A.; Schnabl, B.; Jauch, D.; Taniguchi, K.; Yu, G.Y.; Osterreicher, C.H.; Hung, K.E.; et al. Adenoma-linked barrier defects and microbial products drive IL-23/IL-17-mediated tumour growth. Nature 2012, 491, 254–258. [Google Scholar] [CrossRef] [Green Version]
  77. Tosolini, M.; Kirilovsky, A.; Mlecnik, B.; Fredriksen, T.; Mauger, S.; Bindea, G.; Berger, A.; Bruneval, P.; Fridman, W.H.; Pagès, F.; et al. Clinical impact of different classes of infiltrating T cytotoxic and helper cells (Th1, th2, treg, th17) in patients with colorectal cancer. Cancer Res. 2011, 71, 1263–1271. [Google Scholar] [CrossRef] [Green Version]
  78. Wu, S.; Rhee, K.J.; Albesiano, E.; Rabizadeh, S.; Wu, X.; Yen, H.R.; Huso, D.L.; Brancati, F.L.; Wick, E.; McAllister, F.; et al. A human colonic commensal promotes colon tumorigenesis via activation of T helper type 17T cell responses. Nat. Med. 2009, 15, 1016–1022. [Google Scholar] [CrossRef]
  79. Nguyen, D.P.; Li, J.; Tewari, A.K. Inflammation and prostate cancer: The role of interleukin 6 (IL-6). BJU Int. 2014, 113, 986–992. [Google Scholar] [CrossRef]
  80. Furuya, Y.; Nishio, R.; Junicho, A.; Nagakawa, O.; Fuse, H. Serum interleukin-11 in patients with benign prostatic hyperplasia and prostate cancer. Int. Urol. Nephrol. 2005, 37, 69–71. [Google Scholar] [CrossRef]
  81. Campbell, C.L.; Jiang, Z.; Savarese, D.M.; Savarese, T.M. Increased expression of the interleukin-11 receptor and evidence of STAT3 activation in prostate carcinoma. Am. J. Pathol. 2001, 158, 25–32. [Google Scholar] [CrossRef] [Green Version]
  82. Zurita, A.J.; Troncoso, P.; Cardo-Vila, M.; Logothetis, C.J.; Pasqualini, R.; Arap, W. Combinatorial screenings in patients: The interleukin-11 receptor alpha as a candidate target in the progression of human prostate cancer. Cancer Res. 2004, 64, 435–439. [Google Scholar] [CrossRef] [Green Version]
  83. Torres-Roca, J.F.; DeSilvio, M.; Mora, L.B.; Khor, L.Y.; Hammond, E.; Ahmad, N.; Jove, R.; Forman, J.; Lee, R.J.; Sandler, H.; et al. Activated STAT3 as a correlate of distant metastasis in prostate cancer: A secondary analysis of Radiation Therapy Oncology Group 86-10. Urology 2007, 69, 505–509. [Google Scholar] [CrossRef]
  84. Liu, X.; He, Z.; Li, C.H.; Huang, G.; Ding, C.; Liu, H. Correlation analysis of JAK-STAT pathway components on prognosis of patients with prostate cancer. Pathol. Oncol. Res. 2012, 18, 17–23. [Google Scholar] [CrossRef]
  85. Tam, L.; McGlynn, L.M.; Traynor, P.; Mukherjee, R.; Bartlett, J.M.; Edwards, J. Expression levels of the JAK/STAT pathway in the transition from hormone-sensitive to hormone-refractory prostate cancer. Br. J. Cancer 2007, 97, 378–383. [Google Scholar] [CrossRef]
  86. Wu, C.T.; Hsieh, C.C.; Lin, C.C.; Chen, W.C.; Hong, J.H.; Chen, M.F. Significance of IL-6 in the transition of hormone-resistant prostate cancer and the induction of myeloid-derived suppressor cells. J. Mol. Med. 2012, 90, 1343–1355. [Google Scholar] [CrossRef]
  87. Ammirante, M.; Luo, J.L.; Grivennikov, S.; Nedospasov, S.; Karin, M. B-cell-derived lymphotoxin promotes castration-resistant prostate cancer. Nature 2010, 464, 302–305. [Google Scholar] [CrossRef] [Green Version]
  88. Domingo-Domenech, J.; Oliva, C.; Rovira, A.; Codony-Servat, J.; Bosch, M.; Filella, X.; Montagut, C.; Tapia, M.; Campás, C.; Dang, L.; et al. Interleukin 6, a nuclear factor-kappaB target, predicts resistance to docetaxel in hormone-independent prostate cancer and nuclear factor-kappaB inhibition by PS-1145 enhances docetaxel antitumor activity. Clin. Cancer Res. 2006, 12, 5578–5586. [Google Scholar] [CrossRef] [Green Version]
  89. Matsuda, T.; Junicho, A.; Yamamoto, T.; Kishi, H.; Korkmaz, K.; Saatcioglu, F.; Fuse, H.; Muraguchi, A. Cross-talk between signal transducer and activator of transcription 3 and androgen receptor signaling in prostate carcinoma cells. Biochem. Biophys. Res. Commun. 2001, 283, 179–187. [Google Scholar] [CrossRef]
  90. Kroon, P.; Berry, P.A.; Stower, M.J.; Rodrigues, G.; Mann, V.M.; Simms, M.; Bhasin, D.; Chettiar, S.; Li, C.; Li, P.K.; et al. JAK- STAT blockade inhibits tumor initiation and clonogenic recovery of prostate cancer stem-like cells. Cancer Res. 2013, 73, 5288–5298. [Google Scholar] [CrossRef] [Green Version]
  91. Zhang, G.J.; Adachi, I. Serum interleukin-6 levels correlate to tumor progression and prognosis in metastatic breast carcinoma. Anticancer Res. 1999, 19, 1427–1432. [Google Scholar] [PubMed]
  92. Salgado, R.; Junius, S.; Benoy, I.; Van Dam, P.; Vermeulen, P.; Van Marck, E.; Huget, P.; Dirix, L.Y. Circulating interleukin-6 predicts survival in patients with metastatic breast cancer. Int. J. Cancer 2003, 103, 642–646. [Google Scholar] [CrossRef] [PubMed]
  93. Lim, B.; Woodward, W.A.; Wang, X.; Reuben, J.M.; Ueno, N.T. Inflammatory breast cancer biology: The tumour microenvironment is key. Nat. Rev. Cancer 2018, 18, 485–499. [Google Scholar] [CrossRef]
  94. Dethlefsen, C.; Hojfeldt, G.; Hojman, P. The role of intratumoral and systemic IL-6 in breast cancer. Breast Cancer Res. Treat. 2013, 138, 657–664. [Google Scholar] [CrossRef] [PubMed]
  95. Korkaya, H.; Kim, G.I.; Davis, A.; Malik, F.; Henry, N.L.; Ithimakin, S.; Quraishi, A.A.; Tawakkol, N.; D’Angelo, R.; Paulson, A.K.; et al. Activation of an IL6 inflammatory loop mediates trastuzumab resistance in HER2+ breast cancer by expanding the cancer stem cell population. Mol. Cell 2012, 47, 570–584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Rincon, M.; Broadwater, G.; Harris, L.; Crocker, A.; Weaver, D.; Dressler, L.; Berry, D.; Sutton, L.; Michaelson, R.; Messino, M.; et al. Interleukin-6, multidrug resistance protein-1 expression and response to paclitaxel in women with metastatic breast cancer: Results of cancer and leukemia group B trial 159806. Breast Cancer Res. Treat. 2006, 100, 301–308. [Google Scholar] [CrossRef]
  97. Widschwendter, A.; Tonko-Geymayer, S.; Welte, T.; Daxenbichler, G.; Marth, C.; Doppler, W. Prognostic significance of signal transducer and activator of transcription 1 activation in breast cancer. Clin. Cancer Res. 2002, 8, 3065–3074. [Google Scholar]
  98. Sato, T.; Neilson, L.M.; Peck, A.R.; Liu, C.; Tran, T.H.; Witkiewicz, A.; Hyslop, T.; Nevalainen, M.T.; Sauter, G.; Rui, H. Signal transducer and activator of transcription-3 and breast cancer prognosis. Am. J. Cancer Res. 2011, 1, 347–355. [Google Scholar]
  99. Ren, L.; Wang, X.; Dong, Z.; Liu, J.; Zhang, S. Bone metastasis from breast cancer involves elevated IL-11 expression and the gp130/STAT3 pathway. Med. Oncol. 2012, 30, 634. [Google Scholar] [CrossRef]
  100. Zhou, B.; Damrauer, J.S.; Bailey, S.T.; Hadzic, T.; Jeong, Y.; Clark, K.; Fan, C.; Murphy, L.; Lee, C.Y.; Troester, M.A.; et al. Erythropoietin promotes breast tumorigenesis through tumor-initiating cell self- renewal. J. Clin. Investig. 2014, 124, 553–563. [Google Scholar] [CrossRef] [Green Version]
  101. Karakasheva, T.A.; Lin, E.W.; Tang, Q.; Qiao, W.; Waldron, T.; Soni, M.; Klein-Szanto, A.J.; Sahu, V.; Basu, D.; Ohashi, S.; et al. IL-6 mediates cross-talk between tumor cells and activated fibroblasts in the tumor microenvironment. Cancer Res. 2018, 78, 4957–4970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Tsai, M.S.; Chen, W.C.; Lu, C.H.; Chen, M.F. The prognosis of head and neck squamous cell carcinoma related to immunosuppressive tumor microenvironment regulated by IL-6 signaling. Oral. Oncol. 2019, 91, 47–55. [Google Scholar] [CrossRef] [PubMed]
  103. Geiger, J.L.; Grandis, J.R.; Bauman, J.E. The STAT3 pathway as a therapeutic target in head and neck cancer: Barriers and innovations. Oral. Oncol. 2016, 56, 84–92. [Google Scholar] [CrossRef] [Green Version]
  104. Lui, V.W.; Peyser, N.D.; Ng, P.K.; Hritz, J.; Zeng, Y.; Lu, Y.; Li, H.; Wang, L.; Gilbert, B.R.; General, I.J.; et al. Frequent mutation of receptor protein tyrosine phosphatases provides a mechanism for STAT3 hyperactivation in head and neck cancer. Proc. Natl. Acad. Sci. USA 2014, 111, 1114–1119. [Google Scholar] [CrossRef] [Green Version]
  105. Peyser, N.D.; Du, Y.; Li, H.; Lui, V.; Xiao, X.; Chan, T.A.; Grandis, J.R. Loss-of-function PTPRD mutations lead to increased STAT3 activation and sensitivity to STAT3 inhibition in head and neck cancer. PLoS ONE 2015, 10, e0135750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Rubin Grandis, J.; Melhem, M.F.; Gooding, W.E.; Day, R.; Holst, V.A.; Wagener, M.M.; Drenning, S.D.; Tweardy, D.J. Levels of TGF-alpha and EGFR protein in head and neck squamous cell carcinoma and patient survival. J. Natl. Cancer Inst. 1998, 90, 824–832. [Google Scholar] [CrossRef] [Green Version]
  107. Zhu, X.; Zhang, F.; Zhang, W.; He, J.; Zhao, Y.; Chen, X. Prognostic role of epidermal growth factor receptor in head and neck cancer: A meta-analysis. J. Surg. Oncol. 2013, 108, 387–397. [Google Scholar] [CrossRef]
  108. Alsahafi, E.; Amelio, I.; Raulf, N.; Lucarelli, P.; Sauter, T.; Tavassoli, M. Clinical update on head and neck cancer: Molecular biology and ongoing challenges. Cell Death Dis. 2019, 10, 540. [Google Scholar] [CrossRef] [Green Version]
  109. Madoz-Gúrpide, J.; Zazo, S.; Chamizo, C.; Cassado, V.; Caramés, C.; Gavín, E.; Cristóbal, I.; García-Foncillas, J.; Rojo, F. Activation of MET pathway predicts poor outcome to cetuximab in patients with recurrent or metastatic head and neck cancer. J. Transl. Med. 2015, 13, 282. [Google Scholar] [CrossRef]
  110. Arnold, L.; Enders, J.; Thomas, S.M. Activated HGF-c-Met axis in head and neck cancer. Cancers 2017, 9, 169. [Google Scholar] [CrossRef] [Green Version]
  111. Chung, C.H.; Parker, J.S.; Ely, K.; Carter, J.; Yi, Y.; Murphy, B.A.; Ang, K.K.; EI-Naggar, A.K.; Zanation, A.M.; Cmelak, A.J.; et al. Gene expression profiles identify epithelial-to-mesenchymal transition and activation of nuclear factor-kappaB signaling as characteristics of a high-risk head and neck squamous cell carcinoma. Cancer Res. 2006, 66, 8210–8218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Schipper, J.H.; Frixen, U.H.; Behrens, J.; Unger, A.; Jahnke, K.; Birchmeier, W. E-cadherin expression in squamous cell carcinomas of head and neck: Inverse correlation with tumor dedifferentiation and lymph node metastasis. Cancer Res. 1991, 51, 6328–6337. [Google Scholar] [PubMed]
  113. Yadav, A.; Kumar, B.; Datta, J.; Teknos, T.N.; Kumar, P. IL-6 promotes head and neck tumor metastasis by inducing epithelial-mesenchymal transition via the JAK-STAT3-SNAIL signaling pathway. Mol. Cancer Res. 2011, 9, 1658–1667. [Google Scholar] [CrossRef] [Green Version]
  114. Peltanova, B.; Raudenska, M.; Masarik, M. Effect of tumor microenvironment on pathogenesis of the head and neck squamous cell carcinoma: A systematic review. Mol. Cancer 2019, 18, 63. [Google Scholar] [CrossRef]
  115. Zeng, Q.; Chen, S.; You, Z.; Yang, F.; Carey, T.E.; Saims, D.; Wang, C.Y. Hepatocyte growth factor inhibits anoikis in head and neck squamous cell carcinoma cells by activation of ERK and Akt signaling independent of NFκB. J. Biol. Chem. 2002, 277, 25203–25208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Neiva, K.G.; Zhang, Z.; Miyazawa, M.; Warner, K.A.; Karl, E.; Nör, J.E. Cross talk initiated by endothelial cells enhances migration and inhibits anoikis of squamous cell carcinoma cells through STAT3/Akt/ERK signaling. Neoplasia 2009, 11, 583–593. [Google Scholar] [CrossRef] [PubMed]
  117. Parakh, S.; Ernst, M.; Poh, A.R. Multicellular Effects of STAT3 in Non-small Cell Lung Cancer: Mechanistic Insights and Therapeutic Opportunities. Cancers 2021, 13, 6228. [Google Scholar] [CrossRef]
  118. Barrera, L.; Montes-Servín, E.; Barrera, A.; Ramírez-Tirado, L.A.; Salinas-Parra, F.; Bañales-Méndez, J.L.; Sandoval-Ríos, M.; Arrieta, Ó. Cytokine profile determined by data-mining analysis set into clusters of non-small-cell lung cancer patients according to prognosis. Ann. Oncol. 2015, 26, 428–435. [Google Scholar] [CrossRef]
  119. Pine, S.R.; Mechanic, L.E.; Enewold, L.; Chaturvedi, A.K.; Katki, H.A.; Zheng, Y.-L.; Bowman, E.D.; Engels, E.A.; Caporaso, N.E.; Harris, C.C. Increased levels of circulating interleukin 6, interleukin 8, C-reactive protein, and risk of lung cancer. J. Natl. Cancer Inst. 2011, 103, 1112–1122. [Google Scholar] [CrossRef]
  120. Ujiie, H.; Tomida, M.; Akiyama, H.; Nakajima, Y.; Okada, D.; Yoshino, N.; Takiguchi, Y.; Tanzawa, H. Serum hepatocyte growth factor and interleukin-6 are effective prognostic markers for non-small cell lung cancer. Anticancer Res. 2012, 32, 3251–3258. [Google Scholar]
  121. Song, L.; Smith, M.A.; Doshi, P.; Sasser, K.; Fulp, W.; Altiok, S.; Haura, E.B. Antitumorefficacyoftheanti-interleukin-6 (IL-6) antibody siltuximab in mouse xenograft models of lung cancer. J. Thorac. Oncol. 2014, 9, 974–982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Carpagnano, G.E.; Resta, O.; Foschino-Barbaro, M.P.; Gramiccioni, E.; Carpagnano, F. Interleukin-6 is increased in breath condensate of patients with non-small cell lung cancer. Int. J. Biol. Markers 2002, 17, 141–145. [Google Scholar] [CrossRef] [PubMed]
  123. Yamaguchi, T.; Kimura, H.; Yokota, S.; Yamamoto, Y.; Hashimoto, T.; Nakagawa, M.; Ito, M.; Ogura, T. Effect of IL-6 elevation in malignant pleural effusion on hyperfibrinogenemia in lung cancer patients. Jpn. J. Clin. Oncol. 2000, 30, 53–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Chang, C.H.; Hsiao, C.F.; Yeh, Y.M.; Chang, G.C.; Tsai, Y.H.; Chen, Y.M.; Huang, M.S.; Chen, H.L.; Li, Y.J.; Yang, P.C.; et al. Circulating interleukin-6 level is a prognostic marker for survival in advanced nonsmall cell lung cancer patients treated with chemotherapy. Int. J. Cancer 2013, 132, 1977–1985. [Google Scholar] [CrossRef]
  125. Silva, E.M.; Mariano, V.S.; Pastrez, P.R.A.; Pinto, M.C.; Castro, A.G.; Syrjanen, K.J.; Longatto-Filho, A. High systemic IL-6 is associated with worse prognosis in patients with non-small cell lung cancer. PLoS ONE 2017, 12, e0181125. [Google Scholar] [CrossRef]
  126. Zhao, M.; Liu, Y.; Liu, R.; Qi, J.; Hou, Y.; Chang, J.; Ren, L. Upregulation of IL-11, an IL-6 Family Cytokine, Promotes Tumor Progression and Correlates with Poor Prognosis in Non-Small Cell Lung Cancer. Cell Physiol. Biochem. 2018, 45, 2213–2224. [Google Scholar] [CrossRef]
  127. Wu, J.; Chen, J.; Lv, X.; Yang, Q.; Yao, S.; Zhang, D.; Chen, J. Clinical value of serum and exhaled breath condensate inflammatory factor IL-11 levels in non-small cell lung cancer: Clinical value of IL-11 in non-small cell lung cancer. Int. J. Biol. Markers 2021, 36, 64–76. [Google Scholar] [CrossRef]
  128. Hosoda, H.; Izumi, H.; Tukada, Y.; Takagiwa, J.; Chiaki, T.; Yano, M.; Arai, H. Plasma hepatocyte growth factor elevation may be associated with early metastatic disease in primary lung cancer patients. Ann. Thorac. Cardiovasc. Surg. 2012, 18, 1–7. [Google Scholar] [CrossRef] [Green Version]
  129. Haura, E.B.; Zheng, Z.; Song, L.; Cantor, A.; Bepler, G. Activated epidermal growth factor receptor-Stat-3 signaling promotes tumor survival in vivo in non-small cell lung cancer. Clin. Cancer Res. 2005, 11, 8288–8294. [Google Scholar] [CrossRef] [Green Version]
  130. Sánchez-Ceja, S.G.; Reyes-Maldonado, E.; Vázquez-Manríquez, M.E.; López-Luna, J.J.; Belmont, A.; Gutiérrez-Castellanos, S. Differential expression of STAT5 and Bcl-xL, and high expression of Neu and STAT3 in non-small-cell lung carcinoma. Lung Cancer 2006, 54, 163–168. [Google Scholar] [CrossRef]
  131. Yin, Z.; Zhang, Y.; Li, Y.; Lv, T.; Liu, J.; Wang, X. Prognostic significance of STAT3 expression and its correlation with chemoresis-tance of non-small cell lung cancer cells. Acta Histochem. 2012, 114, 151–158. [Google Scholar] [CrossRef] [PubMed]
  132. Yin, Z.J.; Jin, F.G.; Liu, T.G.; Fu, E.Q.; Xie, Y.H.; Sun, R.L. Overexpression of STAT3 potentiates growth, survival, and radioresistance of non-small-cell lung cancer (NSCLC) cells. J. Surg. Res. 2011, 171, 675–683. [Google Scholar] [CrossRef] [PubMed]
  133. Lee, H.J.; Zhuang, G.; Cao, Y.; Du, P.; Kim, H.J.; Settleman, J. Drug resistance via feedback activation of Stat3 in oncogene-addicted cancer cells. Cancer Cell 2014, 26, 207–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kim, S.M.; Kwon, O.J.; Hong, Y.K.; Kim, J.H.; Solca, F.; Ha, S.J.; Soo, R.A.; Christensen, J.G.; Lee, J.H.; Cho, B.C. Activation of IL-6R/JAK1/STAT3 signaling induces de novo resistance to irreversible EGFR inhibitors in non-small cell lung cancer with T790M resistance mutation. Mol. Cancer 2012, 11, 2254–2264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Jiang, R.; Jin, Z.; Liu, Z.; Sun, L.; Wang, L.; Li, K. Correlation of activated STAT3 expression with clinicopathologic features in lung adenocarcinoma and squamous cell carcinoma. Mol. Diagn. Ther. 2011, 15, 347–352. [Google Scholar] [CrossRef]
  136. Ai, T.; Wang, Z.; Zhang, M.; Zhang, L.; Wang, N.; Li, W.; Song, L. Expression and prognostic relevance of STAT3 and cyclin D1 in non-small cell lung cancer. Int. J. Biol. Markers 2012, 27, 132–138. [Google Scholar] [CrossRef]
  137. Xu, G.; Zhang, C.; Zhang, J. Dominant negative STAT3 suppresses the growth and invasion capability of human lung cancer cells. Mol. Med. Rep. 2009, 2, 819–824. [Google Scholar]
  138. Blakely, C.M.; Pazarentzos, E.; Olivas, V.; Asthana, S.; Yan, J.J.; Tan, I.; Hrustanovic, G.; Chan, E.; Lin, L.; Neel, D.S.; et al. NF-κB-activating complex engaged in response to EGFR oncogene inhibition drives tumor cell survival and residual disease in lung cancer. Cell Rep. 2015, 11, 98–110. [Google Scholar] [CrossRef] [Green Version]
  139. Gao, S.P.; Chang, Q.; Mao, N.; Daly, L.A.; Vogel, R.; Chan, T.; Liu, S.H.; Bournazou, E.; Schori, E.; Zhang, H.; et al. JAK2 inhibition sensitizes resistant EGFR-mutant lung adenocarcinoma to tyrosine kinase inhibitors. Sci. Signal. 2016, 9, ra33. [Google Scholar] [CrossRef] [Green Version]
  140. Chaib, I.; Karachaliou, N.; Pilotto, S.; Codony Servat, J.; Cai, X.; Li, X.; Drozdowskyj, A.; Servat, C.C.; Yang, J.; Hu, C.; et al. Co-activation of STAT3 and YES- associated protein 1 (YAP1) pathway in EGFR-mutant NSCLC. J. Natl. Cancer Inst. 2017, 109, djx014. [Google Scholar] [CrossRef]
  141. Shou, J.; You, L.; Yao, J.; Xie, J.; Jing, J.; Jing, Z.; Jiang, L.; Sui, X.; Pan, H.; Han, W. Cyclosporine A sensitizes human non- small cell lung cancer cells to gefitinib through inhibition of STAT3. Cancer Lett. 2016, 379, 124–133. [Google Scholar] [CrossRef]
  142. Song, L.; Rawal, B.; Nemeth, J.A.; Haura, E.B. JAK1 activates STAT3 activity in non-small-cell lung cancer cells and IL-6 neutralizing antibodies can suppress JAK1-STAT3 signaling. Mol. Cancer Ther. 2011, 10, 481–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Yu, H.A.; Perez, L.; Chang, Q.; Gao, S.P.; Kris, M.G.; Riely, G.J.; Bromberg, J. A phase 1/2 trial of ruxolitinib and erlotinib in patients with EGFR-mutant lung adenocarcinomas with acquired resistance to erlotinib. J. Thorac. Oncol. 2017, 12, 102–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Rotow, J.; Bivona, T.G. Understanding and targeting resistance mechanisms in NSCLC. Nat. Rev. Cancer 2017, 17, 637–658. [Google Scholar] [CrossRef] [PubMed]
  145. Hong, D.; Kurzrock, R.; Kim, Y.; Woessner, R.; Younes, A.; Nemunaitis, J.; Fowler, N.; Zhou, T.; Schmidt, J.; Jo, M.; et al. AZD9150, a next-generation antisense oligonucleotide inhibitor of STAT3 with early evidence of clinical activity in lymphoma and lung cancer. Sci. Transl. Med. 2015, 7, 314ra185. [Google Scholar] [CrossRef] [Green Version]
  146. Xiang, M.; Kim, H.; Ho, V.T.; Walker, S.R.; Bar-Natan, M.; Anahtar, M.; Liu, S.; Toniolo, P.A.; Kroll, Y.; Jones, N.; et al. Gene expression-based discovery of atovaquone as a STAT3 inhibitor and anticancer agent. Blood 2016, 128, 1845–1853. [Google Scholar] [CrossRef] [Green Version]
  147. Yuan, Z.L.; Guan, Y.J.; Chatterjee, D.; Chin, Y.E. Stat3 dimerization regulated by reversible acetylation of a single lysine residue. Science 2005, 307, 269–273. [Google Scholar] [CrossRef]
  148. Kortylewski, M.; Kujawski, M.; Wang, T.; Wei, S.; Zhang, S.; Pilon-Thomas, S.; Niu, G.; Kay, H.; Mulé, J.; Kerr, W.G.; et al. Inhibiting Stat3 signaling in the hematopoietic system elicits multicomponent antitumor immunity. Nat. Med. 2005, 11, 1314–1321. [Google Scholar] [CrossRef]
  149. Herrmann, A.; Priceman, S.J.; Swiderski, P.; Kujawski, M.; Xin, H.; Cherryholmes, G.A.; Zhang, W.; Zhang, C.; Lahtz, C.; Kowolik, C.; et al. CTLA4 aptamer delivers STAT3 siRNA to tumor-associated and malignant T cells. J. Clin. Investig. 2014, 124, 2977–2987. [Google Scholar] [CrossRef] [Green Version]
  150. Yue, C.; Shen, S.; Deng, J.; Priceman, S.J.; Li, W.; Huang, A.; Yu, H. STAT3 in CD8+T cells inhibits their tumor accumulation by downregulating CXCR3/CXCL10 Axis. Cancer Immunol. Res. 2015, 3, 864–870. [Google Scholar] [CrossRef] [Green Version]
  151. Schmetterer, K.G.; Neunkirchner, A.; Wojta-Stremayr, D.; Leitner, J.; Steinberger, P.; Pickl, W.F. STAT3 governs hyporesponsiveness and granzyme B-dependent suppressive capacity in human CD4+ T cells. FASEB J. 2015, 29, 759–771. [Google Scholar] [CrossRef] [PubMed]
  152. Veldhoen, M.; Hocking, R.J.; Atkins, C.J.; Locksley, R.M.; Stockinger, B. TGFbeta in the context of an inflammatory cytokine milieu supports de novo differentiation of IL-17-producing T cells. Immunity 2006, 24, 179–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Nishihara, M.; Ogura, H.; Ueda, N.; Tsuruoka, M.; Kitabayashi, C.; Tsuji, F.; Aono, H.; Ishihara, K.; Huseby, E.; Betz, U.A.; et al. IL-6-gp130-STAT3 in T cells directs the development of IL-17+ Th with a minimum effect on that of Treg in the steady state. Int. Immunol. 2007, 19, 695–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Harris, T.J.; Grosso, J.F.; Yen, H.R.; Xin, H.; Kortylewski, M.; Albesiano, E.; Hipkiss, E.L.; Getnet, D.; Goldberg, M.V.; Maris, C.H.; et al. Cutting edge: An in vivo requirement for STAT3 signaling in Th17 development and Th17-dependent autoimmunity. J. Immunol. 2007, 179, 4313–4317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Gnerlich, J.L.; Mitchem, J.B.; Weir, J.S.; Sankpal, N.V.; Kashiwagi, H.; Belt, B.A.; Porembka, M.R.; Herndon, J.M.; Eberlein, T.J.; Goedegebuure, P.; et al. Induction of Th17 cells in the tumor microenvironment improves survival in a murine model of pancreatic cancer. J. Immunol. 2010, 185, 4063–4071. [Google Scholar] [CrossRef] [Green Version]
  156. McAllister, F.; Bailey, J.M.; Alsina, J.; Nirschl, C.J.; Sharma, R.; Fan, H.; Rattigan, Y.; Roeser, J.C.; Lankapalli, R.H.; Zhang, H.; et al. Oncogenic Kras activates a hematopoietic-to-epithelial IL-17 signaling axis in preinvasive pancreatic neoplasia. Cancer Cell 2014, 25, 621–637. [Google Scholar] [CrossRef] [Green Version]
  157. Austin, J.W.; Lu, P.; Majumder, P.; Ahmed, R.; Boss, J.M. STAT3, STAT4, NFATc1, and CTCF regulate PD-1 through multiple novel regulatory regions in murine T cells. J. Immunol. 2014, 192, 4876–4886. [Google Scholar] [CrossRef]
  158. Celada, L.J.; Kropski, J.A.; Herazo-Maya, J.D.; Luo, W.; Creecy, A.; Abad, A.T.; Chioma, O.S.; Lee, G.; Hassell, N.E.; Shaginurova, G.I.; et al. PD-1 up-regulation on CD4+ T cells promotes pulmonary fibrosis through STAT3-mediated IL-17A and TGF-β1 production. Sci. Transl. Med. 2018, 10, eaar8356. [Google Scholar] [CrossRef] [Green Version]
  159. Bettelli, E.; Carrier, Y.; Gao, W.; Korn, T.; Strom, T.B.; Oukka, M.; Weiner, H.L.; Kuchroo, V.K. Reciprocal developmental pathways for the generation of pathogenic effector Th17 and regulatory T cells. Nature 2006, 441, 235–238. [Google Scholar] [CrossRef]
  160. Korn, T.; Mitsdoerffer, M.; Croxford, A.L.; Awasthi, A.; Dardalhon, V.A.; Galileos, G.; Vollmar, P.; Stritesky, G.L.; Kaplan, M.H.; Waisman, A.; et al. IL-6 controls Th17 immunity in vivo by inhibiting the conversion of conventional T cells into Foxp3+ regulatory T cells. Proc. Natl. Acad. Sci. USA 2008, 105, 18460–18465. [Google Scholar] [CrossRef] [Green Version]
  161. Fujimoto, M.; Nakano, M.; Terabe, F.; Kawahata, H.; Ohkawara, T.; Han, Y.; Ripley, B.; Serada, S.; Nishikawa, T.; Kimura, A.; et al. The influence of excessive IL-6 production in vivo on the development and function of Foxp3+ regulatory T cells. J. Immunol. 2011, 186, 32–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Hsu, P.; Santner-Nanan, B.; Hu, M.; Skarratt, K.; Lee, C.H.; Stormon, M.; Wong, M.; Fuller, S.J.; Nanan, R. IL-10 potentiates differentiation of human induced regulatory T cells via STAT3 and Foxo1. J. Immunol. 2015, 195, 3665–3674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zorn, E.; Nelson, E.A.; Mohseni, M.; Porcheray, F.; Kim, H.; Litsa, D.; Bellucci, R.; Raderschall, E.; Canning, C.; Soiffer, R.J.; et al. IL-2 regulates FOXP3 expression in human CD4+CD25+ regulatory T cells through a STAT-dependent mechanism and induces the expansion of these cells in vivo. Blood 2006, 108, 1571–1579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Kortylewski, M.; Xin, H.; Kujawski, M.; Lee, H.; Liu, Y.; Harris, T.; Drake, C.; Pardoll, D.; Yu, H. Regulation of the IL-23 and IL-12 balance by Stat3 signaling in the tumor microenvironment. Cancer Cell 2009, 15, 114–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Gentles, A.J.; Newman, A.M.; Liu, C.L.; Bratman, S.V.; Feng, W.; Kim, D.; Nair, V.S.; Xu, Y.; Khuong, A.; Hoang, C.D.; et al. The prognostic landscape of genes and infiltrating immune cells across human cancers. Nat. Med. 2015, 21, 938–945. [Google Scholar] [CrossRef]
  166. Giurisato, E.; Xu, Q.; Lonardi, S.; Telfer, B.; Russo, I.; Pearson, A.; Finegan, K.G.; Wang, W.; Wang, J.; Gray, N.S.; et al. Myeloid ERK5 deficiency suppresses tumor growth by blocking protumor macrophage polarization via STAT3 inhibition. Proc. Natl. Acad. Sci. USA 2018, 115, E2801–E2810. [Google Scholar] [CrossRef] [Green Version]
  167. Yang, J.; Liao, D.; Chen, C.; Liu, Y.; Chuang, T.H.; Xiang, R.; Markowitz, D.; Reisfeld, R.A.; Luo, Y. Tumor-associated macrophages regulate murine breast cancer stem cells through a novel paracrine EGFR/Stat3/Sox-2 signaling pathway. Stem Cells 2013, 31, 248–258. [Google Scholar] [CrossRef]
  168. Takaishi, K.; Komohara, Y.; Tashiro, H.; Ohtake, H.; Nakagawa, T.; Katabuchi, H.; Takeya, M. Involvement of M2-polarized macrophages in the ascites from advanced epithelial ovarian carcinoma in tumor progression via Stat3 activation. Cancer Sci. 2010, 101, 2128–2136. [Google Scholar] [CrossRef]
  169. Yan, D.; Wang, H.W.; Bowman, R.L.; Joyce, J.A. STAT3 and STAT6 signaling pathways synergize to promote cathepsin secretion from macrophages via IRE1α activation. Cell Rep. 2016, 16, 2914–2927. [Google Scholar] [CrossRef] [Green Version]
  170. Wölfle, S.J.; Strebovsky, J.; Bartz, H.; Sähr, A.; Arnold, C.; Kaiser, C.; Dalpke, A.H.; Heeg, K. PD-L1 expression on tolerogenic APCs is controlled by STAT-3. Eur. J. Immunol. 2011, 41, 413–424. [Google Scholar] [CrossRef]
  171. Veglia, F.; Sanseviero, E.; Gabrilovich, D.I. Myeloid-derived suppressor cells in the era of increasing myeloid cell diversity. Nat. Rev. Immunol. 2021, 21, 485–498. [Google Scholar] [CrossRef] [PubMed]
  172. Vasquez-Dunddel, D.; Pan, F.; Zeng, Q.; Gorbounov, M.; Albesiano, E.; Fu, J.; Blosser, R.L.; Tam, A.J.; Bruno, T.; Zhang, H.; et al. STAT3 regulates arginase-I in myeloid-derived suppressor cells from cancer patients. J. Clin. Investig. 2013, 123, 1580–1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Hossain, D.M.; Pal, S.K.; Moreira, D.; Duttagupta, P.; Zhang, Q.; Won, H.; Jones, J.; D’Apuzzo, M.; Forman, S.; Kortylewski, M. TLR9-targeted STAT3 silencing abrogates immunosuppressive activity of myeloid- derived suppressor cells from prostate cancer patients. Clin. Cancer Res. 2015, 21, 3771–3782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Dabritz, J.; Judd, L.M.; Chalinor, H.V.; Menheniott, T.R.; Giraud, A.S. Altered gp130 signalling ameliorates experimental colitis via myeloid cell-specific STAT3 activation and myeloid-derived suppressor cells. Sci. Rep. 2016, 6, 20584. [Google Scholar] [CrossRef]
  175. Sumida, K.; Ohno, Y.; Ohtake, J.; Kaneumi, S.; Kishikawa, T.; Takahashi, N.; Taketomi, A.; Kitamura, H. IL-11 induces differentiation of myeloid-derived suppressor cells through activation of STAT3 signalling pathway. Sci. Rep. 2015, 5, 13650. [Google Scholar] [CrossRef]
  176. Takeda, K.; Clausen, B.E.; Kaisho, T.; Tsujimura, T.; Terada, N.; Förster, I.; Akira, S. Enhanced Th1 activity and development of chronic enterocolitis in mice devoid of Stat3 in macrophages and neutrophils. Immunity 1999, 10, 39–49. [Google Scholar] [CrossRef] [Green Version]
  177. Panopoulos, A.D.; Zhang, L.; Snow, J.W.; Jones, D.M.; Smith, A.M.; El Kasmi, K.C.; Liu, F.; Goldsmith, M.A.; Link, D.C.; Murray, P.J.; et al. STAT3 governs distinct pathways in emergency granulopoiesis and mature neutrophils. Blood 2006, 108, 3682–3690. [Google Scholar] [CrossRef] [Green Version]
  178. Zhang, H.; Nguyen-Jackson, H.; Panopoulos, A.D.; Li, H.S.; Murray, P.J.; Watowich, S.S. STAT3 controls myeloid progenitor growth during emergency granulopoiesis. Blood 2010, 116, 2462–2471. [Google Scholar] [CrossRef] [Green Version]
  179. Gotthardt, D.; Putz, E.M.; Straka, E.; Kudweis, P.; Biaggio, M.; Poli, V.; Strobl, B.; Müller, M.; Sexl, V. Loss of STAT3 in murine NK cells enhances NK cell-dependent tumor surveillance. Blood 2014, 124, 2370–2379. [Google Scholar] [CrossRef]
  180. Jerez, A.; Clemente, M.J.; Makishima, H.; Koskela, H.; Leblanc, F.; Peng Ng, K.; Olson, T.; Przychodzen, B.; Afable, M.; Gomez-Segui, I.; et al. STAT3 mutations unify the pathogenesis of chronic lymphoproliferative disorders of NK cells and T cell large granular lymphocyte leukemia. Blood 2012, 120, 3048–3057. [Google Scholar] [CrossRef]
  181. Küçük, C.; Jiang, B.; Hu, X.; Zhang, W.; Chan, J.K.; Xiao, W.; Lack, N.; Alkan, C.; Williams, J.C.; Avery, K.N.; et al. Activating mutations of STAT5B and STAT3 in lymphomas derived from gammadelta-T or NK cells. Nat. Commun. 2015, 6, 6025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Lynch, M.D.; Watt, F.M. Fibroblast Heterogeneity: Implications for Human Disease. J. Clin. Investig. 2018, 128, 26–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Rinn, J.L.; Bondre, C.; Gladstone, H.B.; Brown, P.O.; Chang, H.Y. Anatomic Demarcation by Positional Variation in Fibroblast Gene Expression Programs. PLoS Genet. 2006, 2, e119. [Google Scholar] [CrossRef] [PubMed]
  184. Kalluri, R. The Biology and Function of Fibroblasts in Cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef]
  185. Sanz-Moreno, V.; Gaggioli, C.; Yeo, M.; Albrengues, J.; Wallberg, F.; Viros, A.; Hooper, S.; Mitter, R.; Féral, C.C.; Cook, M.; et al. ROCK and JAK1 signaling cooperate to control actomyosin contractility in tumor cells and stroma. Cancer Cell 2011, 20, 229–245. [Google Scholar] [CrossRef] [Green Version]
  186. Laklai, H.; Miroshnikova, Y.A.; Pickup, M.W.; Collisson, E.A.; Kim, G.E.; Barrett, A.S.; Hill, R.C.; Lakins, J.N.; Schlaepfer, D.D.; Mouw, J.K.; et al. Genotype tunes pancreatic ductal adenocarcinoma tissue tension to induce matricellular fibrosis and tumor progression. Nat. Med. 2016, 22, 497–505. [Google Scholar] [CrossRef] [Green Version]
  187. Nagathihalli, N.S.; Castellanos, J.A.; Shi, C.; Beesetty, Y.; Reyzer, M.L.; Caprioli, R.; Chen, X.; Walsh, A.J.; Skala, M.C.; Moses, H.L.; et al. Signal transducer and activator of transcription 3, mediated remodeling of the tumor microenvironment results in enhanced tumor drug delivery in a mouse model of pancreatic cancer. Gastroenterology 2015, 149, 1932–1943. [Google Scholar] [CrossRef] [Green Version]
  188. O’Donoghue, R.J.; Knight, D.A.; Richards, C.D.; Prêle, C.M.; Lau, H.L.; Jarnicki, A.G.; Jones, J.; Bozinovski, S.; Vlahos, R.; Thiem, S.; et al. Genetic partitioning of interleukin-6 signalling in mice dissociates Stat3 from Smad3-mediated lung fibrosis. EMBO Mol. Med. 2012, 4, 939–951. [Google Scholar] [CrossRef]
  189. Augsten, M. Cancer-associated fibroblasts as another polarized cell type of the tumor microenvironment. Front. Oncol. 2014, 4, 62. [Google Scholar] [CrossRef]
  190. Xing, F.; Saidou, J.; Watabe, K. Cancer associated fibroblasts (CAFs) in tumor microenvironment. Front. Biosci. 2010, 15, 166–179. [Google Scholar] [CrossRef] [Green Version]
  191. Albrengues, J.; Bertero, T.; Grasset, E.; Bonan, S.; Maiel, M.; Bourget, I.; Philippe, C.; Herraiz Serrano, C.; Benamar, S.; Croce, O.; et al. Epigenetic switch drives the conversion of fibroblasts into proinvasive cancer- associated fibroblasts. Nat. Commun. 2015, 6, 10204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Yang, X.; Lin, Y.; Shi, Y.; Li, B.; Liu, W.; Yin, W.; Dang, Y.; Chu, Y.; Fan, J.; He, R. FAP promotes immunosuppression by cancer-associated fibroblasts in the tumor microenvironment via STAT3-CCL2 signaling. Cancer Res. 2016, 76, 4124–4135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Cheng, Y.; Li, H.; Deng, Y.; Tai, Y.; Zeng, K.; Zhang, Y.; Liu, W.; Zhang, Q.; Yang, Y. Cancer-associated fibroblasts induce PDL1+ neutrophils through the IL6-STAT3 pathway that foster immune suppression in hepatocellular carcinoma. Cell Death Dis. 2018, 9, 422. [Google Scholar] [CrossRef] [PubMed]
  194. Tao, L.; Huang, G.; Wang, R.; Pan, Y.; He, Z.; Chu, X.; Song, H.; Chen, L. Cancer-associated fibroblasts treated with cisplatin facilitates chemoresistance of lung adenocarcinoma through IL-11/IL-11R/STAT3 signaling pathway. Sci. Rep. 2016, 6, 38408. [Google Scholar] [CrossRef] [PubMed]
  195. Minoguchi, M.; Minoguchi, S.; Aki, D.; Joo, A.; Yamamoto, T.; Yumioka, T.; Matsuda, T.; Yoshimura, A. STAP-2/BKS, an adaptor/docking protein, modulates STAT3 activation in acute-phase response through its YXXQ motif. J. Biol. Chem. 2003, 278, 11182–11189. [Google Scholar] [CrossRef] [Green Version]
  196. Brauer, P.M.; Tyner, A.L. Building a better understanding of the intra- cellular tyrosine kinase PTK6—BRK by BRK. Biochim. Biophys Acta 2010, 1806, 66–73. [Google Scholar]
  197. Ikeda, O.; Sekine, Y.; Mizushima, A.; Nakasuji, M.; Miyasaka, Y.; Yamamoto, C.; Muromoto, R.; Nanbo, A.; Oritani, K.; Yoshimura, A.; et al. Interactions of STAP-2 with Brk and STAT3 participate in cell growth of human breast cancer cells. J. Biol. Chem. 2010, 285, 38093–38103. [Google Scholar] [CrossRef] [Green Version]
  198. Kitai, Y.; Iwakami, M.; Saitoh, K.; Togi, S.; Isayama, S.; Sekine, Y.; Muromoto, R.; Kashiwakura, J.I.; Yoshimura, A.; Oritani, K.; et al. STAP-2 protein promotes prostate cancer growth by enhancing epidermal growth factor receptor stabilization. J. Biol. Chem. 2017, 292, 19392–19399. [Google Scholar] [CrossRef] [Green Version]
  199. Heo, T.-H.; Wahler, J.; Suh, N. Potential therapeutic implications of IL-6/IL-6R/gp130-targeting agents in breast cancer. Oncotarget 2016, 7, 15460–15473. [Google Scholar] [CrossRef] [Green Version]
  200. Grimley, P.M.; Dong, F.; Rui, H. Stat5a and Stat5b: Fraternal twins of signal transduction and transcriptional activation. Cytokine Growth Factor Rev. 1999, 10, 131–157. [Google Scholar] [CrossRef]
  201. Li, X.; Lu, Y.; Liang, K.; Hsu, J.M.; Albarracin, C.; Mills, G.B.; Hung, M.C.; Fan, Z. Brk/PTK6 sustains activated EGFR signaling through inhibiting EGFR degradation and transactivating EGFR. Oncogene 2012, 31, 4372–4383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Nyati, K.K.; Zaman, M.M.-U.; Sharma, P.; Kishimoto, T. Arid5a, an RNA-binding protein in immune regulation: RNA stability, inflammation, and autoimmunity. Trends Immunol. 2020, 41, 255–268. [Google Scholar] [CrossRef] [PubMed]
  203. Masuda, K.; Ripley, B.; Nishimura, R.; Mino, T.; Takeuchi, O.; Shioi, G.; Kiyonari, H.; Kishimoto, T. Arid5a controls IL-6 mRNA stability, which contributes to elevation of IL-6 level in vivo. Proc. Natl. Acad. Sci. USA 2013, 110, 9409–9414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Saito, Y.; Kagami, S.; Sanayama, Y.; Ikeda, K.; Suto, A.; Kashiwakuma, D.; Furuta, S.; Iwamoto, I.; Nonaka, K.; Ohara, O.; et al. AT-rich-interactive domain-containing protein 5A functions as a negative regulator of retinoic acid receptor-related orphan nuclear receptor γt-induced Th17 cell differentiation. Arthritis Rheumatol. 2014, 66, 1185–1194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Masuda, K.; Ripley, B.; Nyati, K.K.; Dubey, P.K.; Zaman, M.M.; Hanieh, H.; Higa, M.; Yamashita, K.; Standley, D.M.; Mashima, T.; et al. Arid5a regulates naive CD4+ T cell fate through selective stabilization of Stat3 mRNA. J. Exp. Med. 2016, 213, 605–619. [Google Scholar] [CrossRef] [PubMed]
  206. Parajuli, G.; Tekguc, M.; Wing, J.B.; Hashimoto, A.; Okuzaki, D.; Hirata, T.; Sasaki, A.; Itokazu, T.; Handa, H.; Sugino, H.; et al. Arid5a Promotes Immune Evasion by Augmenting Tryptophan Metabolism and Chemokine Expression. Cancer Immunol. Res. 2021, 9, 862–876. [Google Scholar] [CrossRef]
  207. Nyati, K.K.; Hashimoto, S.; Singh, S.K.; Tekguc, M.; Metwally, H.; Liu, Y.C.; Okuzaki, D.; Gemechu, Y.; Kang, S.; Kishimoto, T. The novel long noncoding RNA AU021063, induced by IL-6/Arid5a signaling, exacerbates breast cancer invasion and metastasis by stabilizing Trib3 and activating the Mek/Erk pathway. Cancer Lett. 2021, 520, 295–306. [Google Scholar] [CrossRef]
  208. Edinger, A.L.; Thompson, C.B. Antigen-presenting cells control T cell proliferation by regulating amino acid availability. Proc. Natl. Acad. Sci. USA 2002, 99, 1107–1109. [Google Scholar] [CrossRef] [Green Version]
  209. Munn, D.H.; Shafizadeh, E.; Attwood, J.T.; Bondarev, I.; Pashine, A.; Mellor, A.L. Inhibition of T cell proliferation by macrophage tryptophan catabolism. J. Exp. Med. 1999, 189, 1363–1372. [Google Scholar] [CrossRef]
  210. Mezrich, J.D.; Fechner, J.H.; Zhang, X.; Johnson, B.P.; Burlingham, W.J.; Bradfield, C.A. An interaction between kynurenine and the aryl hydrocarbon receptor can generate regulatory T cells. J. Immunol. 2010, 185, 3190–3198. [Google Scholar] [CrossRef] [Green Version]
  211. Nguyen, N.T.; Kimura, A.; Nakahama, T.; Chinen, I.; Masuda, K.; Nohara, K.; Fujii-Kuriyama, Y.; Kishimoto, T. Aryl hydrocarbon receptor negatively regulates dendritic cell immunogenicity via a kynurenine-dependent mechanism. Proc. Natl. Acad. Sci. USA 2010, 107, 19961–19966. [Google Scholar] [CrossRef] [Green Version]
  212. Neamah, W.H.; Singh, N.P.; Alghetaa, H.; Abdulla, O.A.; Chatterjee, S.; Busbee, P.B.; Nagarkatti, M.; Nagarkatti, P. Ahr activation leads to massive mobilization of myeloid-derived suppressor cells with immunosuppressive activity through regulation of Cxcr2 and microrna mir-150-5p and mir-543-3p that target anti-inflammatory genes. J. Immunol. 2019, 203, 1830–1844. [Google Scholar] [CrossRef] [PubMed]
  213. Fein, M.R.; He, X.Y.; Almeida, A.S.; Bružas, E.; Pommier, A.; Yan, R.; Eberhardt, A.; Fearon, D.T.; Van Aelst, L.; Wilkinson, J.E.; et al. Cancer cell Ccr2 orchestrates suppression of the adaptive immune response. J. Exp. Med. 2020, 217, e20181551. [Google Scholar] [CrossRef] [PubMed]
  214. Huang, B.; Pan, P.Y.; Li, Q.; Sato, A.I.; Levy, D.E.; Bromberg, J.; Divino, C.M.; Chen, S.H. Gr-1+Cd115+ Immature Myeloid Suppressor Cells Mediate the Development of Tumor-Induced T Regulatory Cells and T-Cell Anergy in Tumor-Bearing Host. Cancer Res. 2006, 66, 1123–1131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Huang, B.; Lei, Z.; Zhao, J.; Gong, W.; Liu, J.; Chen, Z.; Liu, Y.; Li, D.; Yuan, Y.; Zhang, G.M.; et al. Ccl2/Ccr2 Pathway Mediates Recruitment of Myeloid Suppressor Cells to Cancers. Cancer Lett. 2007, 252, 86–92. [Google Scholar] [CrossRef]
  216. Li, B.H.; Garstka, M.A.; Li, Z.F. Chemokines and Their Receptors Promoting the Recruitment of Myeloid-Derived Suppressor Cells into the Tumor. Mol. Immunol 2020, 117, 201–215. [Google Scholar] [CrossRef]
  217. Liu, Z.; Wang, H.; Guan, L.; Lai, C.; Yu, W.; Lai, M. LL1, a novel and highly selective STAT3 inhibitor, displays anti-colorectal cancer activities in vitro and in vivo. Br. J. Pharmacol. 2020, 177, 298–313. [Google Scholar] [CrossRef]
  218. Chen, X.; Pan, L.; Wei, J.; Zhang, R.; Yang, X.; Song, J.; Bai, R.Y.; Fu, S.; Pierson, C.R.; Finlay, J.L.; et al. LLL12B, a small molecule STAT3 inhibitor, induces growth arrest, apoptosis, and enhances cisplatin- mediated cytotoxicity in medulloblastoma cells. Sci. Rep. 2021, 11, 6517. [Google Scholar] [CrossRef]
  219. Siddiquee, K.; Zhang, S.; Guida, W.C.; Blaskovich, M.A.; Greedy, B.; Lawrence, H.R.; Yip, M.L.; Jove, R.; McLaughlin, M.M.; Lawrence, N.J.; et al. Selective chemical probe inhibitor of Stat3, identified through structure-based virtual screening, induces antitumor activity. Proc. Natl. Acad. Sci. USA 2007, 104, 7391–7396. [Google Scholar] [CrossRef] [Green Version]
  220. Siddiquee, K.A.; Gunning, P.T.; Glenn, M.; Katt, W.P.; Zhang, S.; Schrock, C.; Sebti, S.M.; Jove, R.; Hamilton, A.D.; Turkson, J. An Oxazole-Based Small-Molecule Stat3 Inhibitor Modulates Stat3 Stability and Processing and Induces Antitumor Cell Effects. ACS Chem. Biol. 2007, 2, 787–798. [Google Scholar] [CrossRef]
  221. Zhang, X.; Sun, Y.; Pireddu, R.; Yang, H.; Urlam, M.K.; Lawrence, H.R.; Guida, W.C.; Lawrence, N.J.; Sebti, S.M. A Novel Inhibitor of STAT3 Homodimerization Selectively Suppresses STAT3 Activity and Malignant Transformation. Cancer Res. 2013, 73, 1922–1933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Akiyama, Y.; Nonomura, C.; Ashizawa, T.; Iizuka, A.; Kondou, R.; Miyata, H.; Sugino, T.; Mitsuya, K.; Hayashi, N.; Nakasu, Y.; et al. The anti-tumor activity of the STAT3 inhibitor STX-0119 occurs via promotion of tumor-infiltrating lymphocyte accumulation in temozolomide-resistant glioblastoma cell line. Immunol. Lett. 2017, 190, 20–25. [Google Scholar] [CrossRef] [PubMed]
  223. Song, H.; Wang, R.; Wang, S.; Lin, J. A low-molecular-weight compound discovered through virtual database screening inhibits Stat3 function in breast cancer cells. Proc. Natl. Acad. Sci. USA 2005, 102, 4700–4705. [Google Scholar] [CrossRef] [Green Version]
  224. Schust, J.; Sperl, B.; Hollis, A.; Mayer, T.U.; Berg, T. Stattic: A small-molecule inhibitor of STAT3 activation and dimerization. Chem. Biol. 2006, 13, 1235–1242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Nishisaka, F.; Taniguchi, K.; Tsugane, M.; Hirata, G.; Takagi, A.; Asakawa, N.; Kurita, A.; Takahashi, H.; Ogo, N.; Shishido, Y.; et al. Antitumor activity of a novel oral signal transducer and activator of transcription 3 inhibitor YHO-1701. Cancer Sci. 2020, 111, 1774–1784. [Google Scholar] [CrossRef] [Green Version]
  226. Turkson, J.; Ryan, D.; Kim, J.S.; Zhang, Y.; Chen, Z.; Haura, E.; Laudano, A.; Sebti, S.; Hamilton, A.D.; Jove, R. Phosphotyrosyl peptides block Stat3-mediated DNA binding activity, gene regulation, and cell transformation. J. Biol. Chem. 2001, 276, 45443–45455. [Google Scholar] [CrossRef] [Green Version]
  227. Turkson, J.; Zhang, S.; Palmer, J.; Kay, H.; Stanko, J.; Mora, L.B.; Sebti, S.; Yu, H.; Jove, R. Inhibition of constitutive signal transducer and activator of transcription 3 activation by novel platinum complexes with potent antitumor activity. Mol. Cancer Ther. 2004, 3, 1533–1542. [Google Scholar] [CrossRef]
  228. Liang, M.; Zhan, F.; Zhao, J.; Li, Q.; Wuyang, J.; Mu, G.; Li, D.; Zhang, Y.; Huang, X. CPA-7 influences immune profile and elicits anti-prostate cancer effects by inhibiting activated STAT3. BMC Cancer 2016, 16, 504. [Google Scholar] [CrossRef] [Green Version]
  229. Huang, W.; Dong, Z.; Chen, Y.; Wang, F.; Wang, C.J.; Peng, H.; He, Y.; Hangoc, G.; Pollok, K.; Sandusky, G.; et al. Small-molecule inhibitors targeting the DNA-binding domain of STAT3 suppress tumor growth, metastasis and STAT3 target gene expression in vivo. Oncogene 2016, 35, 783–792. [Google Scholar] [CrossRef]
  230. Nagel-Wolfrum, K.; Buerger, C.; Wittig, I.; Butz, K.; Hoppe-Seyler, F.; Groner, B. The interaction of specific peptide aptamers with the DNA binding domain and the dimerization domain of the transcription factor Stat3 inhibits transactivation and induces apoptosis in tumor cells. Mol. Cancer Res. 2004, 2, 170–182. [Google Scholar] [CrossRef]
  231. Kobayashi, A.; Tanizaki, Y.; Kimura, A.; Ishida, Y.; Nosaka, M.; Toujima, S.; Kuninaka, Y.; Minami, S.; Ino, K.; Kondo, T. AG490, a Jak2 inhibitor, suppressed the progression of murine ovarian cancer. Eur. J. Pharmacol. 2015, 766, 63–75. [Google Scholar] [CrossRef] [PubMed]
  232. Derenzini, E.; Lemoine, M.; Buglio, D.; Katayama, H.; Ji, Y.; Davis, R.E.; Sen, S.; Younes, A. The JAK inhibitor AZD1480 regulates proliferation and immunity in Hodgkin lymphoma. Blood Cancer J. 2011, 1, e46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Murakami, T.; Takigawa, N.; Ninomiya, T.; Ochi, N.; Yasugi, M.; Honda, Y.; Kubo, T.; Ichihara, E.; Hotta, K.; Tanimoto, M.; et al. Effect of AZD1480 in an epidermal growth factor receptor-driven lung cancer model. Lung Cancer 2014, 83, 30–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Wilson, G.S.; Tian, A.; Hebbard, L.; Duan, W.; George, J.; Li, X.; Qiao, L. Tumoricidal Effects of the JAK Inhibitor Ruxolitinib (INC424) on Hepatocellular Carcinoma In Vitro. Cancer Lett. 2013, 341, 224–230. [Google Scholar] [CrossRef]
  235. Lo, M.C.; Peterson, L.F.; Yan, M.; Cong, X.; Hickman, J.H.; Dekelver, R.C.; Niewerth, D.; Zhang, D.E. JAK inhibitors suppress t(8;21) fusion protein-induced leukemia. Leukemia 2013, 27, 2272–2279. [Google Scholar] [CrossRef] [Green Version]
  236. Horiguchi, A.; Asano, T.; Kuroda, K.; Sato, A.; Asakuma, J.; Ito, K.; Hayakawa, M.; Sumitomo, M.; Asano, T. STAT3 inhibitor WP1066 as a novel therapeutic agent for renal cell carcinoma. Br. J. Cancer 2010, 102, 1592–1599. [Google Scholar] [CrossRef] [Green Version]
  237. Buerger, C.; Nagel-Wolfrum, K.; Kunz, C.; Wittig, I.; Butz, K.; Hoppe-Seyler, F.; Groner, B. Sequence-specific Peptide Aptamers, Interacting with the Intracellular Domain of the Epidermal Growth Factor Receptor, Interfere with Stat3 Activation and Inhibit the Growth of Tumor Cells. J. Biol. Chem. 2003, 278, 37610–37621. [Google Scholar] [CrossRef] [Green Version]
  238. Ge, H.; Liu, H.; Fu, Z.; Sun, Z. Therapeutic and Preventive Effects of an Epidermal Growth Factor Receptor Inhibitor on Oral Squamous Cell Carcinoma. J. Int. Med. Res. 2012, 40, 455–466. [Google Scholar] [CrossRef] [Green Version]
  239. Chang, A.Y.; Wang, M. Molecular mechanisms of action and potential biomarkers of growth inhibition of dasatinib (BMS-354825) on hepatocellular carcinoma cells. BMC Cancer 2013, 13, 267. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Multifaceted roles of STAT3 in physiological and pathological processes, and inhibitors targeting STAT3 signaling. STAT3 is activated by specific cytokines, growth factors, etc., and contributes to multiple physiological functions by regulating its target genes as a transcription factor. Representative STAT3 inhibitors are classified as those that target STAT3 directly (e.g., SH2 domain and DNA-binding domain inhibitors) and indirectly (e.g., JAK kinase and IL-6 inhibitors).
Figure 1. Multifaceted roles of STAT3 in physiological and pathological processes, and inhibitors targeting STAT3 signaling. STAT3 is activated by specific cytokines, growth factors, etc., and contributes to multiple physiological functions by regulating its target genes as a transcription factor. Representative STAT3 inhibitors are classified as those that target STAT3 directly (e.g., SH2 domain and DNA-binding domain inhibitors) and indirectly (e.g., JAK kinase and IL-6 inhibitors).
Cells 11 02618 g001
Figure 2. Functional roles of STAP-2 as a promising target molecule in STAT3-associated tumors. Stimulation by EGF or other molecules induces the phosphorylation of STAT3 and BRK, and STAP-2 is also phosphorylated and interacts with STAT3 and BRK as a scaffold protein. Subsequently, STAT3 translocates to the nucleus, where it regulates target genes and contributes to proliferation, etc.
Figure 2. Functional roles of STAP-2 as a promising target molecule in STAT3-associated tumors. Stimulation by EGF or other molecules induces the phosphorylation of STAT3 and BRK, and STAP-2 is also phosphorylated and interacts with STAT3 and BRK as a scaffold protein. Subsequently, STAT3 translocates to the nucleus, where it regulates target genes and contributes to proliferation, etc.
Cells 11 02618 g002
Figure 3. Arid5a mediates immune evasion. Arid5a is induced by LPS, IL-1β, or IL-6, directly binds to a stem–loop element in the 3′-untranslated regions of target genes to stabilize their mRNA, and augments their expression. Arid5a induces immune evasion by contributing to metabolic reprogramming, the upregulation of immunosuppressive chemokines, and the induction of mesenchymal properties through RNA stabilization.
Figure 3. Arid5a mediates immune evasion. Arid5a is induced by LPS, IL-1β, or IL-6, directly binds to a stem–loop element in the 3′-untranslated regions of target genes to stabilize their mRNA, and augments their expression. Arid5a induces immune evasion by contributing to metabolic reprogramming, the upregulation of immunosuppressive chemokines, and the induction of mesenchymal properties through RNA stabilization.
Cells 11 02618 g003
Table 1. Preclinical studies on STAT3 inhibitors.
Table 1. Preclinical studies on STAT3 inhibitors.
ActionInhibitor/
Compound
Mechanism of ActionCancer TypeRef.
Direct inhibitorsLL1SH2 domain inhibitorCRC, NSCLC[217]
LLL12BSH2 domain inhibitorMedulloblastoma[218]
S3I-201SH2 domain inhibitorBreast cancer, liver cancer[219]
S3I-M2001SH2 domain inhibitorBreast cancer[220]
S31-1757SH2 domain inhibitorBreast cancer, lung cancer[221]
STX-0119SH2 domain inhibitorGlioblastoma[222]
STA-21SH2 domain inhibitorBreast cancer[223]
StatticSH2 domain inhibitorBreast cancer, HNSCC[224]
YHO-1701SH2 domain inhibitorHNSCC, NSCLC[225]
PY*LKTKSH2 domain inhibitorNIH3T3/v-Src or v-Ras[226]
CPA-1DNA-binding domain inhibitorBreast cancer, colon cancer, melanoma[227]
CPA-7DNA-binding domain inhibitorProstate cancer, breast cancer, colon cancer, melanoma[227,228]
inS3-54A18DNA-binding domain inhibitorNSCLC[229]
DBD-1DNA-binding domain inhibitorMelanoma, myeloma[230]
Indirect inhibitorsAG490JAK inhibitorOvarian cancer, pancreatic cancer[231]
AZD1480JAK inhibitorLymphoma, lung cancer[232,233]
RuxolitinibJAK inhibitorHepatocellular carcinom[234]
TG101209JAK2 inhibitorLeukemia[235]
WP1066JAK inhibitorRenal cell carcinoma[236]
KDI1RTK inhibitorVulval and breast cancer[237]
PD153035RTK inhibitorOral squamous carcinoma[238]
DasatinibSrc inhibitorSynovial sarcoma, hepatocellular carcinoma, glioma, prostate cancer[239]
Abbreviations: CRC—colorectal cancer; HNSCC—head and neck squamous cell carcinoma; NSCLC—non-small cell lung carcinoma.
Table 2. STAT3 inhibitors being tested in clinical trials.
Table 2. STAT3 inhibitors being tested in clinical trials.
ActionInhibitor/CompoundTypeCancer TypePhaseNCT Number
Direct inhibitorsBBI608 (FDA approved)Small moleculesAdvanced malignanciesI/IINCT01775423
CRCIIINCT01830621
C188-9Small moleculesBC, CRC, HNSCC, HCC, NSCLC, GAC, melanoma, advanced cancerINCT03195699
OPB-31121Small moleculesadvanced cancer, solid tumorSINCT00955812
HCCI/IINCT01406574
OPB-51602Small moleculesMalignant solid tumorsINCT01184807
Hematological malignanciesINCT01344876
Nasopharyngeal carcinomaINCT02058017
OPB-111077Small moleculesAcute myeloid leukemiaINCT03197714
Advanced HCCINCT01942083
AZD-9150OligonucleotidesLymphomaI/IINCT01563302
Indirect inhibitorsAZD-1480JAK1/2Solid tumorsINCT01112397
CYT387JAK1/2MyelofibrosisI/IINCT02101268
PMF, post-PV, post-ET MFIIINCT03427866
Ruxolitinib (FDA approved)JAK1/2MyelofibrosisIIINCT03427866
LY2784544JAK2Myeloproliferative neoplasmsIINCT01594723
SB1518JAK2 MyelofibrosisIIINCT02055781
Siltuximab (FDA approved)IL-6RMultiple myelomaIINCT03315026
Tocilitizumab (FDA approved)IL-6RHCCI/IINCT02997956
CombinationsAZD9150, durvalumab (anti-PD-L1)Direct inhibitors and ICBNSCLCIINCT03334617
PC, CRC, NSCLCIINCT02983578
Advanced solid tumors, metastatic HNSCCI/IINCT02499328
Diffuse large B-cell lymphomaINCT02549651
BBI608, nivolumab (anti-PD-1)Direct inhibitors and ICBMetastatic CRCIINCT03647839
BBI608, pembrolizumab (anti-PD-1)Direct inhibitors and ICBMetastatic CRCI/IINCT02851004
Apatinib, SHR-1210 (anti-PD-1)Indirect inhibitors and ICBMelanomaIINCT03955354
Bevacizumab, atezolizumab (anti-PD-L1)Indirect inhibitors and ICBUnresectable HCCIIINCT03434379
Dasatinib, Ipilimumab (anti-CTLA-4)Indirect inhibitors and ICBGIST, stage III/IV soft tissue sarcomaINCT01643278
Dasatinib, nivolumab (anti-PD-1)Indirect inhibitors and ICBPhiladelphia chromosome positive ALLINCT02819804
Ruxolitinib, pembrolizumab (anti-PD-1)Indirect inhibitors and ICBHematological malignanciesIINCT04016116
Metastatic stage IV TNBCINCT03012230
Abbreviations: ALL—acute lymphoblastic leukemie; BC—breast cancer; CRC—colorectal cancer; GAC—gastric adenocarcinoma; GIST—gastrointestinal stromal tumor; HCC—hepatocellular carcinoma; HNSCC—head and neck squamous cell carcinoma; NSCLC—non-small cell lung carcinoma; PC—pancreatic cancer; PMF—primary myelofibrosis; Post-PV—post polycythemia vera; Post-ET MF—post-essential thrombocythemia myelofibrosis; TNBC—triple negative breast cancer.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hashimoto, S.; Hashimoto, A.; Muromoto, R.; Kitai, Y.; Oritani, K.; Matsuda, T. Central Roles of STAT3-Mediated Signals in Onset and Development of Cancers: Tumorigenesis and Immunosurveillance. Cells 2022, 11, 2618. https://doi.org/10.3390/cells11162618

AMA Style

Hashimoto S, Hashimoto A, Muromoto R, Kitai Y, Oritani K, Matsuda T. Central Roles of STAT3-Mediated Signals in Onset and Development of Cancers: Tumorigenesis and Immunosurveillance. Cells. 2022; 11(16):2618. https://doi.org/10.3390/cells11162618

Chicago/Turabian Style

Hashimoto, Shigeru, Ari Hashimoto, Ryuta Muromoto, Yuichi Kitai, Kenji Oritani, and Tadashi Matsuda. 2022. "Central Roles of STAT3-Mediated Signals in Onset and Development of Cancers: Tumorigenesis and Immunosurveillance" Cells 11, no. 16: 2618. https://doi.org/10.3390/cells11162618

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