Next Article in Journal
Cerebrospinal Fluid Biomarkers in Childhood Leukemias
Next Article in Special Issue
Integrin α10-Antibodies Reduce Glioblastoma Tumor Growth and Cell Migration
Previous Article in Journal
MINDIN Exerts Protumorigenic Actions on Primary Prostate Tumors via Downregulation of the Scaffold Protein NHERF-1
Previous Article in Special Issue
Satellitosis, a Crosstalk between Neurons, Vascular Structures and Neoplastic Cells in Brain Tumours; Early Manifestation of Invasive Behaviour
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role and Therapeutic Targeting of JAK/STAT Signaling in Glioblastoma

1
Department of Neuro-Oncology, University of Texas M.D. Anderson Cancer Center, 1515 Holcombe Blvd., Houston, TX 77030, USA
2
Department of Neurosurgery, University of Texas M.D. Anderson Cancer Center, 1515 Holcombe Blvd., Houston, TX 77030, USA
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(3), 437; https://doi.org/10.3390/cancers13030437
Submission received: 31 December 2020 / Revised: 19 January 2021 / Accepted: 21 January 2021 / Published: 24 January 2021
(This article belongs to the Special Issue Targeted Therapies for the Treatment of Glioblastoma)

Abstract

:

Simple Summary

Glioblastoma is one of the most treatment-refractory human malignancies, and despite techniques that have allowed scientists and clinicians to better understand the molecular underpinnings of resistance, little progress has been made in improving the survival of patients with glioblastoma. We posit that dysregulated Janus kinase/signal transducer and activator of transcription (JAK/STAT) signaling represents one major hub of tumorigenesis and resistance to medical therapies and that clinical study of its targeted inhibition is warranted, as well as highlighting the lessons learned from historical investigation going forward.

Abstract

Glioblastoma remains one of the deadliest and treatment-refractory human malignancies in large part due to its diffusely infiltrative nature, molecular heterogeneity, and capacity for immune escape. The Janus kinase/signal transducer and activator of transcription (JAK/STAT) signaling pathway contributes substantively to a wide variety of protumorigenic functions, including proliferation, anti-apoptosis, angiogenesis, stem cell maintenance, and immune suppression. We review the current state of knowledge regarding the biological role of JAK/STAT signaling in glioblastoma, therapeutic strategies, and future directions for the field.

1. Introduction

Despite decades of intense study, the prognosis for patients with glioblastoma (GBM) remains near universally poor, with inevitable therapeutic resistance and subsequent recurrence despite multimodality therapies. As advances in other solid and liquid cancers continue to be made, there remains an urgent unmet need for GBM therapeutics. Treatment resistance arises from a wide variety of mechanisms, including the blood–brain barrier (BBB), inter- and intra-tumoral heterogeneity, and a profoundly immunosuppressive tumor microenvironment (TME). As the mechanisms underlying these barriers have been and continue to be elucidated, a number of crucial oncogenic signaling pathways have been discovered that contribute redundantly in promoting tumorigenesis, disease recurrence, and confounding therapeutic strategies. Increasing evidence demonstrates the importance of Janus kinase (JAK)/signal transducer and activator of transcription (STAT) signaling as a pivotal molecular hub active in critical microenvironmental cellular populations, such as glioma cells, reactive astrocytes, and stromal and immune cells, that drives not only aggressive growth, invasion, treatment resistance, and cancer cell stemness but also tumor-mediated immunosuppression. Herein, we review the present knowledge to date of JAK/STAT signaling in promoting these activities, the historical approaches taken to target this pathway, and the future directions.

2. Physiologic JAK/STAT Signaling

The STAT family of transcription factors is comprised of seven proteins—STAT1, STAT2, STAT3, STAT4, STAT-5a, STAT-5b, and STAT6—which reside in the cytoplasm and are activated by phosphorylation as a downstream consequence of a number of signaling pathways, including cytokines, growth factors, or non-receptor tyrosine kinases [1]. In the classical JAK-mediated pathway, cytokine binding of its cognate receptor leads to receptor dimerization followed by docking of JAK and consequent phosphorylation of the receptor’s cytoplasmic tail. STAT proteins are then recruited via their SH2 domains to the activated receptor where tyrosine phosphorylation occurs, STAT hetero- or homodimerization ensues, and activated STAT then undergoes translocation to the nucleus to bind DNA elements such as promoters or enhancers to both directly and indirectly regulate transcription of associated genes (Figure 1) [2]. Although tyrosine phosphorylation of STAT is the most important activating step, STAT can be phosphorylated on serine residues to modulate their activity.
The negative regulation of STAT signaling is mediated through a number of mechanisms acting both upstream and downstream of STAT activation, such as the suppressors of cytokine signaling (SOCS) proteins that inhibit JAK activity via binding to their SH2 domains [3]. The protein inhibitors of activated STAT (PIASs) proteins are another group of proteins that can bind activated STAT and prevent DNA binding, thereby inhibiting downstream transcriptional programs [4,5,6]. Protein tyrosine phosphatases (PTPs), such as Src homology region 2-containing protein tyrosine phosphatase 2 (SHP2), inactivate STAT molecules via dephosphorylation [7,8].
Of the aforementioned seven STAT family members, pathogenic activation of STAT1, STAT3, and STAT5 have been studied in malignant glioma and are the main focus of this review.

3. Biological Principles in JAK/STAT Signaling in Glioblastoma Cells

3.1. STAT1 and STAT5 in GBM

The role of STAT1 in malignant glioma is evolving. Historically, STAT1 was believed to play a tumor suppressor role; however, more recent studies support a more nuanced view [9,10,11]. For example, exogenous overexpression of STAT1 was initially described to decrease proliferation and migration and increase apoptosis in glioma cells [12]. STAT1 signaling is also a negative regulator of HIF-1α-mediated CXCR4 and VEGF gene expression [9,13]. In direct contrast, a more recent study showed that IL-8 secreted by glioblastoma and other microenvironmental immune cells promotes glioma migration, invasion, and mesenchymal transition via the STAT1/HIF-1α/Snail cascade [14]. Consistently, investigators have found upregulated expression of STAT1 signaling genes to be associated with poor prognosis in pathways involving N-Myc or IGFBP-3 [11,14,15].
The complexity of the role of STAT1 in the GBM TME is further confounded by the therapeutic context. For example, STAT1 signaling is essential to induce the effects of oncolytic virotherapy by prompting proinflammatory and apoptotic effects through interferon [16]. Interestingly, STAT1 is expressed in the cytoplasm of reactive astrocytes at the leading or invasive edge of most GBMs as well as in microglia, although the mechanistic implications of this are unclear since it is not the activated phosphorylated form or in a transcriptionally active nuclear location [17]. On balance, it appears to be that the net effect of STAT1 signaling in glioblastoma is context dependent, and this axis may therefore be difficult to therapeutically modulate in a predictable fashion.
STAT5 signaling has been linked to tumorigenesis in GBM. STAT5 is mainly concentrated at the tumor leading edge and plays a role in proliferation and invasion [18,19]. Indeed, knockdown of STAT5 leads to impaired migration in epidermal growth factor variant III (EGFRvIII)-bearing GBM, likely as a result of decreased MMP-1 [20,21]. Granulocyte macrophage colony-stimulating factor (GM-CSF) secreted by glioma cells also activates STAT5 signaling in myeloid derived suppressor cells (MDSCs) to induce Bcl2a1 expression and downregulate IRF8 transcription, thereby inhibiting apoptosis and promoting proliferation, respectively [22,23]. Overall, STAT5 appears to be a potential therapeutic target in GBM, although a dearth of confirmatory in vivo data currently precludes this application.

3.2. STAT3 Dysregulation in GBM

Of all the STAT family members in GBM, STAT3 is the most well-characterized in its breadth of oncogenic activity and immune suppressive role. As gain-of-function STAT3 mutations have not been described in GBM, aberrant STAT3 signaling in malignant glioma occurs predominantly as a result of dysregulated upstream events that ultimately promote proliferation; neovascularization; resistance to apoptosis; and immune escape through downstream targets, such as Bcl-xL, Bcl2l1, Bcl-2, cyclin D1, and c-Myc (Figure 2) [24,25]. Constitutive activation of EGFR signaling—for instance, via EGFR amplification, which occurs in 60% of GBMs—leads to dysregulated STAT3 signaling in part via inhibition of nuclear phosphatases by TRIM59 and constitutive nuclear translocation of STAT3 by RanBP6 [26,27,28,29,30,31]. Basic fibroblast growth factor (bFGF), PDGF, c-MET, and PKCε can also activate STAT3 signaling [32,33]. Elevated expression of IL-6 within the TME by a number of cellular populations, including reactive astrocytes, also leads to increased STAT3 signaling through JAK signaling [34,35].
A lack of negative STAT3 regulation also contributes to its constitutive activation in gliomas. Indeed, epigenetic silencing of the PTPRD gene encoding the receptor protein tyrosine phosphatase delta—an inactivating dephosphorylator of STAT3—is commonly found in GBM as a consequence of CpG hypermethylation on chromosome 9p [36]. Additionally, SOCS1 and SOCS3 promoter hypermethylation–which inactivates these genes—occurs in 24% and 35% of GBM, respectively, and is associated with poor prognosis [37]. SOCS3 knockdown results in an increase in EGFR-related signaling pathways [37,38,39]. Recently characterized, the non-receptor tyrosine kinase bone marrow and X-linked (BMX) is also an inhibitor of SOCS3 and is highly upregulated in nearly 90% of patient-derived glioma stem cells (GSCs) [40]. Finally, overexpression of TGF-β-related protein Smad6 in nearly 90% of GBM tissues promotes the ubiquitination and subsequent degradation of the STAT3 inhibitor PIAS3 [41,42,43]. As a consequence of these aforementioned mechanisms, the reported frequency of phospho-Tyr-STAT3 positivity in human glioma samples ranges up to 60% and has been significantly correlated with histologic grade, EGFRvIII positivity, aggressive behavior, and poor prognosis [44,45,46,47,48,49]. Recent work has affirmed the negative prognostic significance of upregulation of JAK/STAT gene targets—e.g., cytokines, cytokine receptors, and JAKs—in the classical subtype of GBM [50].
Adding a layer of complexity to the role of STAT3 in gliomagenesis is that the role of STAT3 may be to a certain degree genetically determined. In one notable in vitro study, de la Iglesia et al. demonstrated that in GBMs harboring inactivating mutations of PTEN, consequent AKT activation led to downregulation of cytokine receptor leukemia inhibitory factor receptor-β and inhibition of STAT3 signaling, thereby leading to IL-8-induced proliferation and invasiveness [51]. A companion study by de la Iglesia et al. further found that in EGFRvIII-positive GBMs, STAT3 complexed with EGFRvIII in the nucleus to activate its tumorigenic activity [52]. In other words, in PTEN-deficient GBM (found in ~35% of GBM), STAT3 may be tumor suppressive rather than oncogenic [53]. These findings may have significant implications for the clinical study of STAT3 inhibition in GBM. However, because of the known complexity of STAT3 dysregulation, and because an association between PTEN status and response to STAT3 inhibition has not been shown, it would be premature to exclude such a significant proportion of patients with GBM from a clinical trial of a STAT3 inhibitor. PTEN status will likely need to be considered in the correlative analysis of outcome in any clinical trial involving STAT3 inhibition in GBM. In summary, the net balance favors aberrant STAT3 pathway activation as primarily tumorigenic in GBM.

3.3. Consequences of Dysregulated STAT3 Signaling in GBM

STAT3 activation underlies a host of tumorigenic cellular pathways, the foremost of which is in transcriptionally regulating glioma stem cells (GSCs)—a critically treatment-resistant population of cells that give rise to recurrent disease [54,55]. Indeed, activated STAT3 has been associated with Notch signaling in GBMs that is involved in regulating stem cell genes such as OCT4, SOC2, and NANOG [56,57,58]. Aberrant cytokine signaling also promotes GSC self-renewal. One such cytokine is transforming growth factor (TGF)-β, which is highly overexpressed in glioblastoma where it induces expression of the cytokine leukemia inhibitory factor (LIF) to consequently activate JAK/STAT3 to prevent differentiation of the GSCs [59]. Another mediator is IL-6, which is secreted by a multitude of microenvironmental cells that is critical for GSC survival [60,61]. miR-30 is another factor highly expressed in GSCs that binds and inhibits SOCS3 to further promote STAT3 activation [62]. STAT3 also promotes the transcription of miR-182-5p, which binds and inhibits the tumor suppressor protocadherin-8 [63,64]. Finally, STAT3 and NF-κB are preferentially bound and methylated by the enhancer of zeste homolog 2 (EZH2) in GSCs, enhancing their activation and self-renewal [65,66]. These are summarized in Figure 3.
The multi-faceted tumorigenic roles of STAT3 in human glioma cells has been well established by both molecular and pharmacologic inhibition. Targeted inhibition of STAT3 in GSCs not only triggers apoptosis but also leads to decreased proliferation, multipotency, and neurosphere formation [67,68,69]. RNAi-mediated knockdown of STAT3 in human U251 glioma cells led to increased apoptosis via the suppression of transcriptional downstream targets of STAT3, such as Bcl-xL, Mcl-1, and survivin [70,71,72]. WP1066, a small-molecule STAT3 inhibitor, similarly induced apoptosis in U87-MG cells and murine xenografts [73].
Angiogenesis is another hallmark of cancer that has been directly linked to STAT3. In histopathologic examination, phosphorylated STAT3 preferentially localizes to tumor endothelial cells along with vascular endothelial growth factor receptor-2 (VEGFR-2), suggesting an autocrine feed-forward loop promoting the hallmark neovascularization seen in GBM [74]. Indeed, tissue hypoxia dose-dependently increases STAT3 phosphorylation and consequent angiogenesis in human GBM cell lines by stabilizing HIF-1α to enable its transcription of VEGF [75]. GBM also responds to hypoxia by STAT3-mediated upregulation of key proteins involved in motility and invasiveness, such as the matrix metalloproteinases (MMPs-2, -3 and -9), focal adhesion kinase, fascin-1, and TWIST [76,77,78,79,80]. Additionally, activated STAT3 promotes the epithelial–mesenchymal transition (EMT) and invasive behavior [81,82].

3.4. Consequences of Dysregulated STAT3 Signaling on the Immune Microenvironment

With the surging enthusiasm for immunotherapy for other solid malignancies, immune therapeutic targeting of STAT3 is a compelling strategy for modulating tumor-mediated immune suppression, especially for a cancer like GBM that has in general not been responsive to immune checkpoint inhibitor monotherapies [83,84]. Indeed, GBMs have been shown to express a number of tumor-specific antigens—e.g., EGFRvIII, IL13Rα2, and HER2—that may be recognized by the host immune system [85,86,87,88], yet there is compelling evidence that glioma-mediated immune suppression prevents immune recognition and effector responses. For example, GBM is notable for a paucity of T cells and an enrichment of myeloid cells such macrophages and microglia, which may be even more evident in the transcriptionally defined GBM mesenchymal and classical subtypes [89,90]. As will be discussed, a substantial body of literature points to the role of aberrant STAT3 signaling in mediating dysfunction of both the innate and adaptive components of the immune system in GBM [91,92].

3.4.1. STAT3 Activation Generates an Immunosuppressive Cytokine Milieu

In GBM, interactions between tumor cells, reactive astrocytes, and microglia lead to high levels of IL-10 and TGF-β, which promotes a positive feedback loop of STAT3 signaling [93,94]. STAT3-driven HIF-1α transcription also leads to the secretion of immunosuppressive cytokines, such as galectin-3, CCL-2, and CSF. These events, combined with additional STAT3-driven elaboration of such factors as IL-4, IL-6, and GM-CSF, further promotes immunosuppressive crosstalk [55,95,96,97,98,99,100]. Anti-sense oligonucleotide-mediated STAT3 blockade in murine models of melanoma and breast and colon cancer show upregulation of proinflammatory cytokines, such as CXCL10, RANTES, TNF-α, and IFN-β [101].

3.4.2. STAT3 Activation Impairs the Innate Immune System

Preferential activation of the STAT3 pathway by the permissive cytokine milieu has profound effects on the innate immune system components of the TME, which are largely mediated by tumor-associated macrophages/microglia (TAMs), myeloid-derived suppressor cells (MDSCs), and other cells [102]. Natural killer (NK) cells in the setting of STAT3 activation, for instance, have impaired cytotoxicity [92]. STAT3 activation in GBM-resident TAMs, which comprise the largest population of infiltrative cells, leads to polarization toward an immunosuppressive phenotype that secretes IL-10 and TGFβ1 and are impaired in their ability to mediate phagocytosis [103,104,105,106,107,108]. These same TAMs also lack co-stimulatory molecules such as CD80 necessary for T cell activation and secrete IL-23, which induces Tregs to adopt a more immunosuppressive phenotype [109,110].
STAT3 activation also inhibits maturation of dendritic cells and their ability to express key molecules necessary for effective antigen presentation (e.g., MHC class II) and T cell activation (e.g., IL-12) [92,111]. Similar to its effect in TAMs, STAT3 also inhibits the expression of co-stimulatory molecules CD80 and CD86 on dendritic cells that are necessary for induction of T cell activation and proliferation.
Various cytokines—but particularly IL-6—are associated with the infiltration of MDSCs in the TME, which are positively correlated with glioma grade and have been shown to exert immune suppressive effects against T and NK cells through expression of enzymes such as arginase that trigger T cell arrest and apoptosis [112,113,114,115,116,117]. MDSCs also express IFN-α, which signals through interferon receptor type 1 (IFNAR1) to activate JAK1/STAT1 signaling, thereby upregulating expression of co-inhibitory molecule programmed death ligand-1 (PD-L1) [114]. STAT3 inhibition not only promotes the maturation of tumor-infiltrating dendritic cells to express the costimulatory molecules, such as CD80 and CD86, necessary for T cell activation, but also induces a shift in tumor immune cell composition toward less MDSCs and immunosuppressive macrophages/microglia [118,119,120,121].

3.4.3. STAT3 Activity Impairs the Adaptive Immune System

Constitutive STAT3 activation also has significant effects on the adaptive immune system. STAT3 activation alters the immune cell composition of the microenvironment, promoting decreased infiltration of effector CD4+ and CD8+ T cells and an increased proportion of Tregs [121]. Activation of STAT3 in CD8+ T cells reduces their secretion of IFN-γ, which further inhibits T cell activation [122]. STAT3 cooperates with transcription factor Forkhead box P3 (FOXP3) to promote the differentiation of immunosuppressive Tregs [123,124,125,126,127]. Indirectly through innate immunity, TAMs inhibit T cell proliferation via secretion of TGF-β [105]. STAT3 has also been shown to play an important role in regulating PD-L1 expression by antigen-presenting dendritic cells and in upregulating the expression of T cell coinhibitory molecule B7-H4 on GSCs and TAMs [95].
In a pivotal early study, Kortylewski et al. demonstrated that hematopoietic cell-specific inhibition of STAT3 in tumor-bearing mice led to significantly enhanced functional activity of T, NK, and dendritic cells, resulting in antitumor immunity and growth inhibition [92]. Hussain et al. similarly demonstrated that STAT3 inhibition of the immune cells from glioblastoma patients promotes not only expression of co-stimulatory molecules on microenvironmental TAMs and proinflammatory cytokines but also the expansion of effector T cells over immunosuppressive Tregs [122].
In summary, STAT3 signaling constitutes a crucial hub through which the tumor microenvironment is shaped and driven toward immunosuppression.

4. STAT3-Mediated Resistance to Therapeutic Modalities

At present, the standard-of-care for patients with GBM consists of maximum safe resection followed by the “Stupp regimen”: alkylating chemotherapy temozolomide (TMZ) combined with radiation, followed by adjuvant TMZ. TMZ mediates its cytotoxic effects via the formation of O6-methylguanine adducts on DNA, leading to replication-associated double-stranded DNA breaks and apoptosis [128,129]. A number of studies have shown that STAT3 signaling is involved in chemoresistance. Kohsaka et al. demonstrated that STAT3 inhibits the degradation of the enzyme O(6)methyl-guanine DNA methyltransferase (MGMT) [130], whose expression is associated with TMZ resistance [131]. Additionally, findings from Wang et al. [132] and Lee et al. [133] indicate that inhibition of STAT3 sensitizes glioma cells to TMZ. Another study by Li et al. linked a reduced level of the STAT3 inhibitory miR-519a—which normally mediates proapoptotic and autophagic responses to chemotherapy—to TMZ resistance in GBM cells [134]. They further showed that STAT3 inhibition strongly decreased the IC50 of TMZ and increased TMZ-induced apoptosis while upregulating Bcl-2 expression and downregulating Bax expression. Finally, they showed that STAT3 inhibition leads to increased TMZ-induced G0-G1 arrest and decreased cyclin D1 expression compared to TMZ alone. It should be noted that these findings are somewhat contradicted by recent work from Heynckes et al. [135], which found that recurrent GBM tissue from Stupp regimen-treated patients demonstrated decreased phospho-STAT3 expression compared to their original tumors. Furthermore, the authors found that treatment of two IFNγ-stimulated patient-derived GBM cell lines with TMZ led to decreased JAK/STAT pathway genes including STAT1 and STAT3, although this was not seen in the commercial LN229 GBM cell line. The generalizability of the authors’ primary conclusion that TMZ treatment leads to inhibition of JAK/STAT signaling in recurrent GBMs is uncertain. Notably, this study had methodological limitations, including small sample sizes (three cell lines and fifteen tissue samples) and—most critically—the absence of microenvironmental JAK/STAT expression analysis. Nevertheless, the authors’ findings suggest that the role of aberrant STAT3 signaling in underlying chemoresistance is nuanced and further point to the fact that the optimal role and timing of STAT3 inhibition in combination with conventional therapeutic strategies remains to be determined.
There is also evidence that STAT3 is involved in mediating resistance to radiotherapy, which induces cell death largely via single- and double-stranded DNA breaks and oxidative damage [136]. Rath et al. found that cytokine crosstalk between astrocytes and GSCs in vitro led to transcriptional upregulation of STAT3 target genes and radioresistance in the latter [137]. Radiation combined with pharmacological inhibition of STAT3 in the corresponding orthotopic murine xenografts led to decreased tumor size and prolonged survival, suggesting that STAT3 contributes to the baseline radioresistance of GSCs. More recently, Yu et al. elucidated that this may involve the protein regulator of chromosome condensation 2 (RCC2), which signals through STAT3 to activate transcription of DNA methyltransferase 1 to lead to silencing of tumor suppressor genes [138]. Radiation induces the secretion of exosomes by glioma cells containing proteins involved in numerous crucial signaling pathways including JAK/STAT, as well as proteins such as ribophorin II, which signal through STAT3 to promote anti-apoptosis via Mcl1 [139,140]. Studies of GBM cells with acquired radioresistance are notable for decreased SOCS3 and increased Forkhead box protein M1 (FOXM1), the latter of which interacts with STAT3 to increase transcription of DNA repair genes such as MRE11 and RAD51 [141,142]. Finally, irradiated GBM cells have been shown to acquire increased invasive, migratory, and mesenchymal properties via the upregulation of intercellular adhesion molecule (ICAM-1) through activated STAT3/NF-κB and Slug signaling [143,144]. In summary, STAT3 signaling appears to be involved in promoting intrinsic as well as acquired chemo- and radio-resistance.
There is substantial literature that supports the role of STAT3 activation in mediating resistance to targeted therapy. The reliance on several dysregulated—and often redundant—signaling pathways, including EGFR, PIK3CA, PDGF, and NF-κB, has been well-characterized in glioblastoma, and cross-talk between these have likely contributed in large part to the historical failures of targeted therapy [53,145,146]. Indeed, compensatory STAT3 upregulation has been demonstrated in numerous studies of other receptor tyrosine kinases (e.g., EGFR, MEK, HER2) in other solid tumors [147,148,149]. Treatment of GBM with anti-VEGF monoclonal antibody bevacizumab for instance—currently the only FDA-approved targeted treatment for glioblastoma and one that has not been shown to have survival benefit—eventually induces STAT3 activation likely via the hypoxia response, leading to the expression of stemness and invasive markers such as nestin and ICAM-1, respectively [150]. The combination of STAT3 and VEGF inhibition with AZD1480 and cediranib, respectively, led to significantly decreased angiogenesis and tumor volume in a murine model [150,151]. In a more recent study of acquired MET inhibitor resistance in GBM, Cruickshanks et al. found compensatory upregulation of a number of other signaling pathways, including MAPK, PI3K, and STAT [152]. Co-inhibition of STAT3 and MET restored sensitivity to apoptosis.

Rationale for Combination Therapy

From the preceding discussion, it is clear that the compensatory mechanisms of resistance to conventional, targeted, and immunotherapeutic strategies against GBM will likely thwart therapies based on a single vulnerability [153]. Consequently, one of the major tasks in neuro-oncology is leveraging continually evolving molecular knowledge to design rational therapeutic combinations. One recently published study of STAT3 inhibition and radiation in a syngeneic immune-competent glioma mouse model found that the combination led to immunologic reprogramming of the TME, with increased dendritic cell–T cell interaction and antigen presentation. This was associated, moreover, with significantly improved animal survival, indicating that a fully functional immune response was required for mediating the therapeutic effects of STAT3 inhibition [154]. Clinical investigators have also recognized that STAT3 inhibition as monotherapy is unlikely to demonstrate an improvement in survival for the vast majority of GBM patients, and, as such, the results of trials investigating STAT3 inhibition with conventional radiation (NCT03514069) and chemotherapy (NCT02315534) are eagerly awaited.
With regard to its well-characterized immunomodulatory effects, combining STAT3 inhibition with other immuno-oncologic strategies also represents an exciting prospect. For example, one preclinical study has shown improved effector functions of adoptively transferred CD8+ T cells in combination with STAT3 inhibition in myeloid cells [155]. Another study demonstrated that administration of a STAT3-targeting miR-124 in combination with a T cell co-stimulatory aptamer was able to increase CD4+ and CD8+ T cells and decrease macrophage trafficking into the TME in a genetically engineered murine glioma model [156]. Subsequent immunotherapeutic efforts focusing on further optimizing tumor-specific cytotoxicity and immune microenvironmental modulation are clearly warranted. The combination of STAT3 inhibition with other immunotherapies, such as adoptive cell transfer, tumor vaccination, oncolytic virotherapy, and immune checkpoint inhibition, are rational. Trials investigating the safety and efficacy of STAT3 and PD-1 blockade (e.g., NCT02851004, NCT02467361, NCT02983578, and NCT03421353) in various solid malignancies are underway, and there is hope that insights gained from these may inform trial design for GBM.

5. JAK/STAT Axis-Targeting Therapies

The aforementioned convergence of multiple oncogenic and immunosuppressive cellular pathways on the JAK/STAT signaling axis makes it an attractive molecular target. Indeed, Stechishin et al. found that targeted JAK2/STAT3 inhibition in orthotopic GBM xenografts regardless of molecular strategy led to increased cytotoxicity of GSCs regardless of MGMT promoter methylation status [157]. To date, a vast number of agents ranging from antisense oligonucleotides and repurposed drugs, JAK1/2 to direct STAT3 inhibitors have been the subject of investigation in numerous cancers. As the BBB presents a unique anatomic barrier to drug delivery in contrast to other malignancies, identifying high-potency, specific, BBB-penetrant molecules of favorable bioavailability is the foremost priority. A discussion of those studied in malignant glioma follows.
Targeting aberrant upstream IL-6/IL-6R signaling is one potential avenue of JAK/STAT blockade. Treatment with IL-6 pathway blockade via its receptor (IL-6R, tocilizumab) or binding soluble IL-6 (siltuximab) has been shown to inhibit glioma growth in vitro and reduce the expression of coinhibitory molecules such as PD-L1 on infiltrative myeloid cells [158,159]. A number of clinical trials investigating the efficacy of these agents in other solid malignancies are underway (e.g., NCT03135171, NCT04258150 NCT04524871, NCT03424005, and NCT04191421), although it should be noted that as far as potential therapies for GBM, these large monoclonal antibodies have not been found to attain therapeutic concentrations in the CSF after systemic administration [160].
A number of repurposed pharmacologic agents have been found to have STAT3 inhibitory activity; however, off-target effects due to lack of specificity and questionable CNS penetrance have limited their utility. Atovaquone is an anti-malarial drug FDA-approved for pneumocystis pneumonia that was found to have STAT3 inhibitory effects; notably, it appears to be poorly bioavailable in the CNS [161,162]. Arsenic trioxide (ATO) was shown to reduce STAT3 activation via JAK inhibition and induce apoptosis and stemness of GSCs. Despite encouraging safety data, a phase II trial combining ATO with radiation and temozolomide for newly diagnosed malignant glioma did not demonstrate a survival benefit [163,164,165]. Sorafenib, a multi-kinase (Raf, VEGFR2, and PDGFR-β) inhibitor with STAT3 inhibitory activity FDA approved for other solid malignancies, was shown to inhibit the proliferation of GBM in vitro, likely through its effects on AKT and MAPK [166,167]. Subsequent clinical trials combining sorafenib with mTOR inhibition, temozolomide, or radiation for patients with newly diagnosed or recurrent GBM did not demonstrate a survival benefit [168,169,170].

5.1. Nutraceutical Inhibition of STAT3

A staggering number of natural compounds have been found to have STAT3 inhibitory activity in vitro, but the clinical utility of these in their native forms for GBM is largely hindered by unfavorable pharmacokinetic properties including low potency, unacceptable toxicity, rapid metabolism, and/or poor BBB penetrance. Plant/fruit-derived resveratrol, for instance, is capable of inhibiting proliferation and invasion of glioma cells in a STAT3-dependent fashion in vitro and when administered intrathecally in vivo. Sadly, it is rapidly metabolized when given systemically [171,172]. To circumvent this, Jhaveri et. al., encapsulated resveratrol within transferrin-bearing liposomes targeting GBM cells [173]. This strategy was insufficient in that no overall significant growth inhibition after intravenous administration was seen. Earlier efforts with liposomal modification of ursolic acid met with similar lack of anti-proliferative and anti-angiogenic activity [174]. What appears to be increasingly clear is that the chief utility of natural STAT3-inhibiting compounds lies in serving as the structural basis for synthesizing novel and more effective small molecules. KYZ3, for instance, a synthetic derivative of quinoid diterpene cryptotanshinone found in dried roots, demonstrated cytotoxic activity in a murine model of triple-negative breast cancer [175]. The various challenges associated with a selected list of natural compounds are summarized in Table 1.

5.2. Pharmaceutical Inhibition of STAT3: JAK Inhibitors

JAK inhibition, although historically employed in the treatment of myeloproliferative disorders, represents another avenue of STAT3 pathway targeting in malignant glioma. As summarized in Table 2, a number of agents have encouraging preclinical evidence of antitumoral efficacy in glioma cells or stem cells, although confirmatory in vivo studies are pending (e.g., G6 and SAR317461) [211,212]. One agent is JSI-124, a JAK2 inhibitor with NF-κB pathway activating capability based on the structure of cucurbitacin, which has been shown to inhibit proliferation of glioblastoma in vitro and also promote the maturation and T-cell-activating capability of dendritic cells isolated from the spleens of tumor-bearing mice, leading to improved cytotoxicity and growth inhibition with subsequent dendritic cell vaccination [209,213,214,215]. To date, only one JAK inhibitor of potential benefit for GBM patients—AZD1480—has advanced to a phase I clinical trial, where unusual dose-limiting neuropsychiatric toxicities halted further development [216,217]. The most recent JAK inhibitors of translational relevance are pacritinib and ruxolitinib. Both are orally administered, BBB-penetrant, highly potent inhibitors of either JAK1/JAK2 (ruxolitinib) or JAK2 (pacritinib), with demonstrated GSC-specific cytotoxicity and chemosensitizing properties [218]. Interestingly, pacritinib was also shown to decrease the amount of miR-21-enriched exosomes released from tumor-associated macrophages, reducing an exogenous source of STAT3 upregulation in glioma cells [94]. Ruxolitinib not only decreased the invasiveness of GBM cells but also was found to be able to inhibit the compensatory JAK1/STAT1 signaling that limits the efficacy of oncolytic virotherapy [219,220,221]. Ruxolitinib is currently being evaluated in a phase I trial for patients with newly diagnosed MGMT-unmethylated malignant glioma in combination with radiation and temozolomide (NCT03514069).

5.3. Pharmaceutical Inhibition of STAT3: Peptide and Non-Peptide STAT3 Mimetics

Peptidomimetics such as PY*LKTK were among the earliest in vitro STAT3 inhibiting strategies, designed to bind the SH2 domain and thereby prevent dimerization and DNA binding [222]. Concerns about in vivo stability, immunogenicity, and low potency led to the design of non-peptide mimetics, such as LLL12 and STX-0119, which have been found to induce apoptosis in GBM cell lines or TMZ-resistant xenografts via inhibition of downstream STAT3 targets, such as survivin Bcl-2 and Bcl-xL [223,224,225,226]. In spite of these, moderate potency (e.g., IC50 15–44 µM for STX-0119), poor bioavailability, and/or unclear BBB penetrance limit the utility of these agents at this time.

5.4. Pharmaceutical Inhibition of STAT3: Oligonucleotide-Based Strategies

Oligonucleotides represent another strategy of direct STAT3 modulation. In a proof-of-concept experiment, Gu et al. designed a decoy oligodeoxynucleotide comprised of a STAT3-specific DNA cis-element and transfected it into human GBM cells U251 and A172, finding specific blockade of STAT3 activation with subsequent cell-cycle arrest and apoptosis mediated by decreases in mRNA levels of c-Myc, cyclin D1, and Bcl-xL [71]. Kortylewski et al. conjugated a STAT3-specific small interfering (siRNA) to a Toll-like receptor agonist CpG and found it was able to be internalized by tumor-associated macrophages and dendritic cells to mediate STAT3 inhibition [227]. MicroRNA (miRNA)-based strategies have also shown promise. Having initially found STAT3-inhibitory miR-124 expression to be significantly downregulated in gliomas compared to normal brain, Wei et al. demonstrated that exogenously administered miR-124 induced secretion of proinflammatory cytokines from GSCs, promoted CD8+ T cell effector with decreased Treg function, and led to immune-mediated growth inhibition in a murine model of glioma [228].
As oligonucleotides are susceptible to degradation by circulating nucleases, subsequent groups have developed various modifications to optimize stability and drug delivery. Yaghi et al. encapsulated miR-124 duplexes within lipid nanoparticles and demonstrated efficient uptake by immune cells with subsequent reduction in activated STAT3 and increased survival in a murine model of glioma [229]. Linder et al. complexed anti-STAT3 siRNA with a polyethylenimine (PEI) and phospholipid 1,2-dipalmitoyl-sn-glycero-3-phosphocholine liposomal conjugate and found it to mediate STAT3 inhibition in glioma cells, although in vivo no significant growth inhibition was observed, and tumoral STAT3 inhibition was quite heterogeneous upon histopathologic examination [230]. PEI-siRNA directed against survivin has also shown some efficacy, albeit in subcutaneously implanted gliomas in mice [231].
Overall, the major challenges for nanoparticle-based delivery of oligonucleotides include ensuring adequate tissue distribution in tumors with a high degree of vascular heterogeneity and minimizing nonspecific uptake by non-tumor or stromal cells. Further study is certainly warranted to optimize these approaches before clinical application.

5.5. Pharmaceutical Inhibition of STAT3: Direct STAT3 Inhibitors

A tremendous number of molecules have been synthesized to inhibit STAT3, usually acting by binding the SH2 domain required for interacting with phosphorylated tyrosine residues on receptors and STAT dimerization. Of these, only a handful have advanced to clinical trials for GBM, for reasons that will be discussed.
Stattic was one of the earliest SH2-binding STAT3 inhibitors, and was found to exert anti-proliferative and radiosensitizing effects on GSCs, although a lack of specificity and in vivo efficacy data has prevented it from moving forward [232,233,234,235]. AG490 is a caffeic acid derivative able to inhibit STAT3 phosphorylation in vitro to decrease the invasiveness of GBM cells, whose low pharmacologic potency has also hindered further development [73,236]. Promisingly, SH-4-54 is small molecule STAT3/STAT5 inhibitor based on salicylic acid with excellent BBB penetration that was found to potently induce apoptosis in GSCs and inhibit growth of gliomas in a murine xenograft model, although it should be noted that these were subcutaneously implanted tumors [237,238]. Another study by Cui et al. showed that SH-4-54 could induce apoptosis in TMZ-resistant GBM cells by inducing the translocation of mitochondrial STAT3 [239]. This led to the activation of mitochondrial STAT3, negative regulation of mitochondrial-encoded genes, and abnormal oxidative phosphorylation.
Attempts to improve the pharmacokinetic parameters of small molecules such as AG490 led to the development of subsequent compounds such as WP1193 [240] and one of the best-studied novel STAT3 inhibitors, WP1066. WP1066 demonstrated not only favorable BBB penetrance and proapoptotic effects in GBM cell lines but also the important capability of upregulating immune costimulatory molecule expression and the release of proinflammatory cytokines from TAMs. Indeed, this molecule was among the first to document the modulating effects of STAT3 on immune cell composition and effector function in the TME [73,122,241]. It has recently been shown to have radiosensitizing effects in GSCs [232]. It is currently the subject of study in a phase I clinical trial of patients with recurrent glioblastoma (NCT01904123) and pediatric patients with brain tumors (NCT04334863). Napabucasin is another recently characterized orally bioavailable small-molecule STAT3 inhibitor capable of inducing cell cycle arrest, apoptosis, and the reduction of markers of stemness, leading to improved survival in an orthotopic murine glioma model [242,243]. A phase I/II study of napabucasin in combination with TMZ for patients with recurrent glioblastoma has been completed, and the results are eagerly awaited (NCT02315534). The above discussed pharmacologic inhibitors of JAK/STAT signaling are summarized in Table 2.

6. STAT3-Related Biomarkers

In view of the tremendous intra- and inter-tumoral heterogeneity of GBM, identifying patients most likely to benefit from targeted STAT3 inhibition is a critically important challenge. There is yet to be a consensus, for instance, on whether expression of phospho-Tyr705 STAT3, phospho-Ser727 STAT3, or both represents maximally abnormal STAT3 pathway activation with its concomitantly poorest prognosis [47,244]. Interestingly, recent work by Tan et al. integrating the TCGA subtypes with STAT3-related gene expression data defined “STAT3-high” and “STAT3-low” gene signatures intended to assist with patient stratification [245]. STAT3-high tumors were enriched for classical/mesenchymal subtyping, IDHwt, and 1p/19q-non-codeleted status, while STAT3-low tumors were enriched for proneural subtypes with IDHmut and 1p/19q-codeleted status. Cells isolated from patients with STAT3-high gene signatures had lower IC50s upon treatment with STAT3 inhibitors such as Stattic or WP1066 relative to those with STAT3-low gene signatures. Of note, STAT3-low tumors (i.e., STAT3 inhibition non-responders) upregulated insulin-like growth factor binding protein 2 (IGFBP2) and insulin growth factor-1R (IGF-1R) in response to STAT3 inhibition, and the combination inhibition of IGF-1R and STAT3 in orthotopic xenografts led to re-sensitization to STAT3 inhibition. Interestingly, their functionally determined gene signature outperformed conventional immunohistochemically determined phospho-STAT3 expression in identifying STAT3 inhibition responders, although an important methodological limitation to note is that these were validated in severe combined immunodeficiency (SCID) mice who by definition lack an antitumoral immune response. In addition, it is also notable that the authors did not comment on whether PTEN mutation was associated with a “STAT3-low” phenotype, as might be expected based on work by de Iglesia et al. [51]. Future correlative studies based on molecular data of patients being treated with STAT3 inhibitors are warranted to refine the criteria by which patients should be considered for STAT3 inhibition.

7. Conclusions

As increasingly sophisticated genomic, transcriptomic, proteomic, and bioinformatic analyses continue to propel the field of oncology into the molecular era, glioblastoma continues to represent a glaringly unmet need. Rational therapeutic strategies need to account not only for anatomic barriers such as the BBB but also the tremendous intra- and intertumoral heterogeneity that characterize this disease. As the preclinical studies have shown, JAK/STAT signaling is tremendously complex, and, while on balance, constitutive activation tends to promote tumor proliferation, angiogenesis, and immune escape, the effects of targeting upstream or downstream effectors are not always predictable. Nevertheless, it is our position that, on balance, there is sufficient evidence of the importance of dysregulated JAK/STAT signaling as an important driver of gliomagenesis and treatment resistance that human study continues to be warranted for this deadly disease. Because it is unlikely that strategies based on one or two key molecular vulnerabilities will be generalizable to the entire patient population, continued efforts to define and validate biomarkers to help stratify patients appropriate for JAK/STAT combination therapy are crucial to advancing our understanding of the disease.

Author Contributions

Conceptualization, A.B.H. and A.O. Writing—original draft preparation, A.O., D.F. Writing—review and editing, A.B.H., M.O., D.F. All authors have read and agreed to the published version of the manuscript.

Funding

The article was supported by NIH grant R01CA120813.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The article was supported by the Ben and Catherine Ivy Foundation.

Conflicts of Interest

A.O., M.O., and D.F.—no conflict of interest to disclose. A.B.H. discloses stock or other ownership in Caris Life Science; consultancy or advisory role for Caris Life Science and WCG Oncology; research funding, drug, or devices from Celularity, Carthera, Codiak Bioscience, and Moleculin; and patents, royalties, or other intellectual property from Celldex Therapeutics, and DNAtrix.

References

  1. Reich, N.C.; Liu, L. Tracking STAT nuclear traffic. Nat. Rev. Immunol. 2006, 6, 602–612. [Google Scholar] [CrossRef] [PubMed]
  2. Swiatek-Machado, K.; Kaminska, B. STAT signaling in glioma cells. Adv. Exp. Med. Biol. 2013, 986, 189–208. [Google Scholar] [CrossRef] [PubMed]
  3. Sasaki, A.; Yasukawa, H.; Suzuki, A.; Kamizono, S.; Syoda, T.; Kinjyo, I.; Sasaki, M.; Johnston, J.A.; Yoshimura, A. Cytokine-inducible SH2 protein-3 (CIS3/SOCS3) inhibits Janus tyrosine kinase by binding through the N-terminal kinase inhibitory region as well as SH2 domain. Genes Cells 1999, 4, 339–351. [Google Scholar] [CrossRef] [PubMed]
  4. Yoshimura, A.; Ito, M.; Chikuma, S.; Akanuma, T.; Nakatsukasa, H. Negative Regulation of Cytokine Signaling in Immunity. Cold Spring Harb. Perspect. Biol. 2018, 10. [Google Scholar] [CrossRef]
  5. Shuai, K.; Liu, B. Regulation of gene-activation pathways by PIAS proteins in the immune system. Nat. Rev. Immunol. 2005, 5, 593–605. [Google Scholar] [CrossRef]
  6. O’Shea, J.J.; Gadina, M.; Schreiber, R.D. Cytokine signaling in 2002: New surprises in the Jak/Stat pathway. Cell 2002, 109, S121–S131. [Google Scholar] [CrossRef] [Green Version]
  7. Xu, D.; Qu, C.-K. Protein tyrosine phosphatases in the JAK/STAT pathway. Front. Biosci. 2008, 13, 4925–4932. [Google Scholar] [CrossRef] [Green Version]
  8. Roccograndi, L.; Binder, Z.A.; Zhang, L.; Aceto, N.; Zhang, Z.; Bentires-Alj, M.; Nakano, I.; Dahmane, N.; O’Rourke, D.M. SHP2 regulates proliferation and tumorigenicity of glioma stem cells. J. Neurooncol. 2017, 135, 487–496. [Google Scholar] [CrossRef]
  9. Hiroi, M.; Mori, K.; Sakaeda, Y.; Shimada, J.; Ohmori, Y. STAT1 represses hypoxia-inducible factor-1-mediated transcription. Biochem. Biophys. Res. Commun. 2009, 387, 806–810. [Google Scholar] [CrossRef]
  10. Chen, Z.; Mou, L.; Pan, Y. CXCL8 Promotes Glioma Progression By Activating The JAK/STAT1/HIF-1α/Snail Signaling Axis. OncoTargets Ther. 2019, 12, 8125–8138. [Google Scholar] [CrossRef] [Green Version]
  11. Thota, B.; Arimappamagan, A.; Kandavel, T.; Shastry, A.H.; Pandey, P.; Chandramouli, B.A.; Hegde, A.S.; Kondaiah, P.; Santosh, V. STAT-1 expression is regulated by IGFBP-3 in malignant glioma cells and is a strong predictor of poor survival in patients with glioblastoma. J. Neurosurg. 2014, 121, 374–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ju, H.; Li, X.; Li, H.; Wang, X.; Wang, H.; Li, Y.; Dou, C.; Zhao, G. Mediation of multiple pathways regulating cell proliferation, migration, and apoptosis in the human malignant glioma cell line U87MG via unphosphorylated STAT1: Laboratory investigation. J. Neurosurg. 2013, 118, 1239–1247. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Y.; Jin, G.; Zhang, J.; Mi, R.; Zhou, Y.; Fan, W.; Cheng, S.; Song, W.; Zhang, B.; Ma, M.; et al. Overexpression of STAT1 suppresses angiogenesis under hypoxia by regulating VEGF-A in human glioma cells. Biomed. Pharmacother. 2018, 104, 566–575. [Google Scholar] [CrossRef] [PubMed]
  14. Duarte, C.W.; Willey, C.D.; Zhi, D.; Cui, X.; Harris, J.J.; Vaughan, L.K.; Mehta, T.; McCubrey, R.O.; Khodarev, N.N.; Weichselbaum, R.R.; et al. Expression signature of IFN/STAT1 signaling genes predicts poor survival outcome in glioblastoma multiforme in a subtype-specific manner. PLoS ONE 2012, 7, e29653. [Google Scholar] [CrossRef]
  15. Meng, D.; Chen, Y.; Yun, D.; Zhao, Y.; Wang, J.; Xu, T.; Li, X.; Wang, Y.; Yuan, L.; Sun, R.; et al. High expression of N-myc (and STAT) interactor predicts poor prognosis and promotes tumor growth in human glioblastoma. Oncotarget 2015, 6, 4901–4919. [Google Scholar] [CrossRef] [Green Version]
  16. Rajaraman, S.; Canjuga, D.; Ghosh, M.; Codrea, M.C.; Sieger, R.; Wedekink, F.; Tatagiba, M.; Koch, M.; Lauer, U.M.; Nahnsen, S.; et al. Measles Virus-Based Treatments Trigger a Pro-Inflammatory Cascade and a Distinctive Immunopeptidome in Glioblastoma. Mol. Ther. Oncolytics 2018, 12. [Google Scholar] [CrossRef] [Green Version]
  17. Haybaeck, J.; Obrist, P.; Schindler, C.U.; Spizzo, G.; Doppler, W. STAT-1 expression in human glioblastoma and peritumoral tissue. Anticancer Res. 2007, 27, 3829–3835. [Google Scholar]
  18. Cao, S.; Wang, C.; Zheng, Q.; Qiao, Y.; Xu, K.; Jiang, T.; Wu, A. STAT5 regulates glioma cell invasion by pathways dependent and independent of STAT5 DNA binding. Neurosci. Lett. 2011, 7, 228–233. [Google Scholar] [CrossRef]
  19. Liang, Q.C.; Xiong, H.; Zhao, Z.W.; Jia, D.; Li, W.X.; Qin, H.Z.; Deng, J.P.; Gao, L.; Zhang, H.; Gao, G.D. Inhibition of transcription factor STAT5b suppresses proliferation, induces G1 cell cycle arrest and reduces tumor cell invasion in human glioblastoma multiforme cells. Cancer Lett. 2009, 273, 164–171. [Google Scholar] [CrossRef]
  20. Roos, A.; Dhruv, H.D.; Peng, S.; Inge, L.J.; Tuncali, S.; Pineda, M.; Millard, N.; Mayo, Z.; Eschbacher, J.M.; Loftus, J.C.; et al. EGFRvIII-Stat5 Signaling Enhances Glioblastoma Cell Migration and Survival. Mol. Cancer Res. 2018, 16, 1185–1195. [Google Scholar] [CrossRef] [Green Version]
  21. Korzus, E.; Nagase, H.; Rydell, R.; Travis, J. The mitogen-activated protein kinase and JAK-STAT signaling pathways are required for an oncostatin M-responsive element-mediated activation of matrix metalloproteinase 1 gene expression. J. Biol. Chem. 1997, 10, 1188–1196. [Google Scholar] [CrossRef] [Green Version]
  22. Medina-Echeverz, J.; Haile, L.A.; Zhao, F.; Gamrekelashvili, J.; Ma, C.; Métais, J.Y.; Dunbar, C.E.; Kapoor, V.; Manns, M.P.; Korangy, F.; et al. IFN-γ regulates survival and function of tumor-induced CD11b+ Gr-1high myeloid derived suppressor cells by modulating the anti-apoptotic molecule Bcl2a1. Eur. J. Immunol. 2014, 44, 2457–2467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Waight, J.D.; Netherby, C.; Hensen, M.L.; Miller, A.; Hu, Q.; Liu, S.; Bogner, P.N.; Farren, M.R.; Lee, K.P.; Liu, K.; et al. Myeloid-derived suppressor cell development is regulated by a STAT/IRF-8 axis. J. Clin. Investig. 2013, 123, 4464–4478. [Google Scholar] [CrossRef] [PubMed]
  24. Puram, S.V.; Yeung, C.M.; Jahani-Asl, A.; Lin, C.; de la Iglesia, N.; Konopka, G.; Jackson-Grusby, L.; Bonni, A. STAT3-iNOS Signaling Mediates EGFRvIII-Induced Glial Proliferation and Transformation. J. Neurosci. 2012, 6, 7806–7818. [Google Scholar] [CrossRef]
  25. Bromberg, J.F.; Wrzeszczynska, M.H.; Devgan, G. Stat3 as an oncogene. Cell 1999, 98, 295–303. [Google Scholar] [CrossRef] [Green Version]
  26. Lo, H.W.; Cao, X.; Zhu, H.; Ali-Osman, F. Constitutively activated STAT3 frequently coexpresses with epidermal growth factor receptor in high-grade gliomas and targeting STAT3 sensitizes them to Iressa and alkylators. Clin. Cancer Res. 2008, 14, 6042–6054. [Google Scholar] [CrossRef] [Green Version]
  27. Shao, H.; Cheng, H.Y.; Cook, R.G.; Tweardy, D.J. Identification and characterization of signal transducer and activator of transcription 3 recruitment sites within the epidermal growth factor receptor. Cancer Res. 2003, 15, 3923–3930. [Google Scholar]
  28. Kim, J.E.; Patel, M.; Ruzevick, J.; Jackson, C.M.; Lim, M. STAT3 activation in glioblastoma: Biochemical and therapeutic implications. Cancers 2014, 6, 376–395. [Google Scholar] [CrossRef] [Green Version]
  29. Sang, Y.; Li, Y.; Song, L.; Alvarez, A.A.; Zhang, W.; Lv, D.; Tang, J.; Liu, F.; Chang, Z.; Hatakeyama, S.; et al. TRIM59 Promotes Gliomagenesis by Inhibiting TC45 Dephosphorylation of STAT3. Cancer Res. 2018, 78, 1792–1804. [Google Scholar] [CrossRef] [Green Version]
  30. Oldrini, B. EGFR feedback-inhibition by Ran-binding protein 6 is disrupted in cancer. Nat. Commun. 2017, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  31. Heimberger, A.B.; Hlatky, R.; Suki, D.; Yang, D.; Weinberg, J.; Gilbert, M.; Sawaya, R.; Aldape, K. Prognostic effect of epidermal growth factor receptor and EGFRvIII in glioblastoma multiforme patients. Clin. Cancer Res. 2005, 11, 1462–1466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wu, J.; Feng, X.; Zhang, B.; Li, J.; Xu, X.; Liu, J.; Wang, X.; Wang, J.; Tong, X. Blocking the bFGF/STAT3 interaction through specific signaling pathways induces apoptosis in glioblastoma cells. J. Neurooncol. 2014, 120, 33–41. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, Y.; Li, Z.; Zhang, C.; Zhang, S.; Ji, Y.; Chen, F. Knockdown of PKCε expression inhibits growth, induces apoptosis and decreases invasiveness of human glioma cells partially through Stat3. J. Mol. Neurosci. 2015, 55, 21–31. [Google Scholar] [CrossRef] [PubMed]
  34. Weissenberger, J.; Loeffler, S.; Kappeler, A.; Kopf, M.; Lukes, A.; Afanasieva, T.A.; Aguzzi, A.; Weis, J. IL-6 is required for glioma development in a mouse model. Oncogene 2004, 23, 3308–3316. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, Q.; Li, G.; Li, R.; Shen, J.; He, Q.; Deng, L.; Zhang, C.; Zhang, J. IL-6 promotion of glioblastoma cell invasion and angiogenesis in U251 and T98G cell lines. J. Neurooncol. 2010. [Google Scholar] [CrossRef]
  36. Veeriah, S.; Brennan, C.; Meng, S.; Singh, B.; Fagin, J.A.; Solit, D.B.; Paty, P.B.; Rohle, D.; Vivanco, I.; Chmielecki, J.; et al. The tyrosine phosphatase PTPRD is a tumor suppressor that is frequently inactivated and mutated in glioblastoma and other human cancers. Proc. Natl. Acad. Sci. USA 2009, 9, 9435–9440. [Google Scholar] [CrossRef] [Green Version]
  37. Martini, M.; Pallini, R.; Luongo, G.; Cenci, T.; Lucantoni, C.; Larocca, L.M. Prognostic relevance of SOCS3 hypermethylation in patients with glioblastoma multiforme. Int. J. Cancer 2008, 15, 2955–2960. [Google Scholar] [CrossRef]
  38. Lindemann, C.; Hackmann, O.; Delic, S. SOCS3 promoter methylation is mutually exclusive to EGFR amplification in gliomas and promotes glioma cell invasion through STAT3 and FAK activation. Acta Neuropathol. 2011, 122, 241–251. [Google Scholar] [CrossRef]
  39. Zhou, H.; Miki, R.; Eeva, M.; Fike, F.M.; Seligson, D.; Yang, L.; Yoshimura, A.; Teitell, M.A.; Jamieson, C.A.; Cacalano, N.A. Reciprocal regulation of SOCS 1 and SOCS3 enhances resistance to ionizing radiation in glioblastoma multiforme. Clin. Cancer Res. 2007, 13, 2344–2353. [Google Scholar] [CrossRef] [Green Version]
  40. Shi, Y.; Guryanova, O.A.; Zhou, W. Ibrutinib inactivates BMX-STAT3 in glioma stem cells to impair malignant growth and radioresistance. Sci. Transl. Med. 2018, 10. [Google Scholar] [CrossRef] [Green Version]
  41. Brantley, E.C.; Nabors, L.B.; Gillespie, G.Y.; Choi, Y.H.; Palmer, C.A.; Harrison, K.; Roarty, K.; Benveniste, E.N. Loss of protein inhibitors of activated STAT-3 expression in glioblastoma multiforme tumors: Implications for STAT-3 activation and gene expression. Clin. Cancer Res. 2008, 14, 4694–4704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Chung, C.D.; Liao, J.; Liu, B.; Rao, X.; Jay, P.; Berta, P.; Shuai, K. Specific inhibition of Stat3 signal transduction by PIAS3. Science 1997, 278, 1803–1805. [Google Scholar] [CrossRef] [PubMed]
  43. Jiao, J.; Zhang, R.; Li, Z.; Yin, Y.; Fang, X.; Ding, X.; Cai, Y.; Yang, S.; Mu, H.; Zong, D.; et al. Nuclear Smad6 promotes gliomagenesis by negatively regulating PIAS3-mediated STAT3 inhibition. Nat. Commun. 2018, 9, 2504. [Google Scholar] [CrossRef] [PubMed]
  44. Alvarez, J.V.; Mukherjee, N.; Chakravarti, A.; Robe, P.; Zhai, G.; Chakladar, A.; Loeffler, J.; Black, P.; Frank, D.A. A STAT3 Gene Expression Signature in Gliomas is Associated with a Poor Prognosis. Transl. Oncogenomics 2007, 2, 99–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Birner, P.; Toumangelova-Uzeir, K.; Natchev, S.; Guentchev, M. STAT3 tyrosine phosphorylation influences survival in glioblastoma. J. Neurooncol. 2010, 100, 339–343. [Google Scholar] [CrossRef] [PubMed]
  46. Tu, Y.; Zhong, Y.; Fu, J.; Cao, Y.; Fu, G.; Tian, X.; Wang, B. Activation of JAK/STAT signal pathway predicts poor prognosis of patients with gliomas. Med. Oncol. 2011, 28, 15–23. [Google Scholar] [CrossRef] [PubMed]
  47. Abou-Ghazal, M.; Yang, D.S.; Qiao, W.; Reina-Ortiz, C.; Wei, J.; Kong, L.Y.; Fuller, G.N.; Hiraoka, N.; Priebe, W.; Sawaya, R.; et al. The incidence, correlation with tumor-infiltrating inflammation, and prognosis of phosphorylated STAT3 expression in human gliomas. Clin. Cancer Res. 2008, 15, 8228–8235. [Google Scholar] [CrossRef] [Green Version]
  48. Mizoguchi, M.; Betensky, R.A.; Batchelor, T.T.; Bernay, D.C.; Louis, D.N.; Nutt, C.L. Activation of STAT3, MAPK, and AKT in malignant astrocytic gliomas: Correlation with EGFR status, tumor grade, and survival. J. Neuropathol. Exp. Neurol. 2006, 65, 1181–1188. [Google Scholar] [CrossRef] [Green Version]
  49. Leventoux, N.; Augustus, M.; Azar, S.; Riquier, S.; Villemin, J.P.; Guelfi, S.; Falha, L.; Bauchet, L.; Gozé, C.; Ritchie, W.; et al. Transformation Foci in IDH1-mutated Gliomas Show STAT3 Phosphorylation and Downregulate the Metabolic Enzyme ETNPPL, a Negative Regulator of Glioma Growth. Sci. Rep. 2020, 26, 5504. [Google Scholar] [CrossRef]
  50. Park, A.K.; Kim, P.; Ballester, L.Y.; Esquenazi, Y.; Zhao, Z. Subtype-specific signaling pathways and genomic aberrations associated with prognosis of glioblastoma. Neuro-Oncology 2019, 1, 59–70. [Google Scholar] [CrossRef] [Green Version]
  51. De la Iglesia, N.; Konopka, G.; Lim, K.-L.; Nutt, C.L.; Bromberg, J.F.; Frank, D.A.; Mischel, P.S.; Louis, D.N.; Bonni, A. Deregulation of a STAT3-interleukin 8 signaling pathway promotes human glioblastoma cell proliferation and invasiveness. J. Neurosci. 2008, 28, 5870–5878. [Google Scholar] [CrossRef] [PubMed]
  52. De la Iglesia, N.; Konopka, G.; Puram, S.V.; Chan, J.A.; Bachoo, R.M.; You, M.J.; Levy, D.E.; Depinho, R.A.; & Bonni, A. Identification of a PTEN-regulated STAT3 brain tumor suppressor pathway. Genes Dev. 2008, 22, 449–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Cancer Genome Atlas Research Network. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 2008, 455, 1061–1068. [Google Scholar] [CrossRef] [PubMed]
  54. Bao, S.; Wu, Q.; McLendon, R.E.; Hao, Y.; Shi, Q.; Hjelmeland, A.B.; Dewhirst, M.W.; Bigner, D.D.; Rich, J.N. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef]
  55. Li, J.; Shen, J.; Wang, Z.; Xu, H.; Wang, Q.; Chai, S.; Fu, P.; Huang, T.; Anas, O.; Zhao, H.; et al. ELTD1 facilitates glioma proliferation, migration and invasion by activating JAK/STAT3/HIF-1α signaling axis. Sci. Rep. 2019, 25, 13904. [Google Scholar] [CrossRef] [Green Version]
  56. Chen, X.; Xu, H.; Yuan, P. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell 2008, 133, 1106–1117. [Google Scholar] [CrossRef] [Green Version]
  57. Guryanova, O.A.; Wu, Q.; Cheng, L. Nonreceptor tyrosine kinase BMX maintains self-renewal and tumorigenic potential of glioblastoma stem cells by activating STAT3. Cancer Cell 2011, 19, 498–511. [Google Scholar] [CrossRef] [Green Version]
  58. Cheng, W.; Zhang, C.; Ren, X.; Jiang, Y.; Han, S.; Liu, Y.; Cai, J.; Li, M.; Wang, K.; Liu, Y.; et al. Bioinformatic analyses reveal a distinct Notch activation induced by STAT3 phosphorylation in the mesenchymal subtype of glioblastoma. J. Neurosurg. 2017, 126, 249–259. [Google Scholar] [CrossRef]
  59. Peñuelas, S.; Anido, J.; Prieto-Sánchez, R.M.; Folch, G.; Barba, I.; Cuartas, I.; García-Dorado, D.; Poca, M.A.; Sahuquillo, J.; Baselga, J.; et al. TGF-beta increases glioma-initiating cell self-renewal through the induction of LIF in human glioblastoma. Cancer Cell 2009, 7, 315–327. [Google Scholar] [CrossRef] [Green Version]
  60. Hossain, A.; Gumin, J.; Gao, F. Mesenchymal Stem Cells Isolated From Human Gliomas Increase Proliferation and Maintain Stemness of Glioma Stem Cells through the IL-6/gp130/STAT3 Pathway. Stem Cells 2015, 33, 2400–2415. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, H.; Lathia, J.D.; Wu, Q.; Wang, J.; Li, Z.; Heddleston, J.M.; Eyler, C.E.; Elderbroom, J.; Gallagher, J.; Schuschu, J.; et al. Targeting interleukin 6 signaling suppresses glioma stem cell survival and tumor growth. Stem Cells 2009, 27, 2393–2404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Che, S.; Sun, T.; Wang, J.; Jiao, Y.; Wang, C.; Meng, Q.; Qi, W.; Yan, Z. miR-30 overexpression promotes glioma stem cells by regulating Jak/STAT3 signaling pathway. Tumour. Biol. 2015, 36, 6805–6811. [Google Scholar] [CrossRef] [PubMed]
  63. Xue, J.; Zhou, A.; Wu, Y.; Morris, S.A.; Lin, K.; Amin, S.; Verhaak, R.; Fuller, G.; Xie, K.; Heimberger, A.B.; et al. miR-182-5p Induced by STAT3 Activation Promotes Glioma Tumorigenesis. Cancer Res. 2016, 76, 4293–4304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zong, Z.; Pang, H.; Yu, R.; Jiao, Y. PCDH8 inhibits glioma cell proliferation by negatively regulating the AKT/GSK3β/β-catenin signaling pathway. Oncol. Lett. 2017, 14, 3357–3362. [Google Scholar] [CrossRef] [Green Version]
  65. Kim, E.; Kim, M.; Woo, D.H. Phosphorylation of EZH2 activates STAT3 signaling via STAT3 methylation and promotes tumorigenicity of glioblastoma stem-like cells. Cancer Cell 2013, 23, 839–852. [Google Scholar] [CrossRef] [Green Version]
  66. Liu, H.; Sun, Y.; Qi, X. EZH2 Phosphorylation Promotes Self-Renewal of Glioma Stem-Like Cells Through NF-κB Methylation. Front. Oncol. 2019, 9. [Google Scholar] [CrossRef] [Green Version]
  67. Li, G.H.; Wei, H.; Lv, S.Q.; Ji, H.; Wang, D.L. Knockdown of STAT3 expression by RNAi suppresses growth and induces apoptosis and differentiation in glioblastoma stem cells. Int. J. Oncol. 2010, 37, 103–110. [Google Scholar]
  68. Sherry, M.M.; Reeves, A.; Wu, J.K.; Cochran, B.H. STAT3 is required for proliferation and maintenance of multipotency in glioblastoma stem cells. Stem Cells 2009, 27, 2383–2392. [Google Scholar] [CrossRef] [Green Version]
  69. Villalva, C.; Martin-Lannerée, S.; Cortes, U.; Cortes, U.; Dkhissi, F.; Wager, M.; Le Corf, A.; Tourani, J.M.; Dusanter-Fourt, I.; Turhan, A.G.; et al. STAT3 is essential for the maintenance of neurosphere-initiating tumor cells in patients with glioblastomas: A potential for targeted therapy? Int. J. Cancer 2011, 15, 826–838. [Google Scholar] [CrossRef]
  70. Chen, F.; Xu, Y.; Luo, Y.; Zheng, D.; Song, Y.; Yu, K.; Li, H.; Zhang, L.; Zhong, W.; Ji, Y. Down-regulation of Stat3 decreases invasion activity and induces apoptosis of human glioma cells. J. Mol. Neurosci. 2010, 40, 353–359. [Google Scholar] [CrossRef]
  71. Gu, J.; Li, G.; Sun, T.; Su, Y.; Zhang, X.; Shen, J.; Tian, Z.; Zhang, J. Blockage of the STAT3 signaling pathway with a decoy oligonucleotide suppresses growth of human malignant glioma cells. J. Neurooncol. 2008, 89, 9–17. [Google Scholar] [CrossRef] [PubMed]
  72. Weissenberger, J.; Priester, M.; Bernreuther, C.; Rakel, S.; Glatzel, M.; Seifert, V.; Kögel, D. Dietary curcumin attenuates glioma growth in a syngeneic mouse model by inhibition of the JAK1,2/STAT3 signaling pathway. Clin. Cancer Res. 2010, 16, 5781–5795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Iwamaru, A.; Szymanski, S.; Iwado, E.; Aoki, H.; Yokoyama, T.; Fokt, I.; Hess, K.; Conrad, C.; Madden, T.; Sawaya, R.; et al. A novel inhibitor of the STAT3 pathway induces apoptosis in malignant glioma cells both in vitro and in vivo. Oncogene 2007, 26, 2435–2444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Schaefer, L.K.; Ren, Z.; Fuller, G.N.; Schaefer, T.S. Constitutive activation of Stat3alpha in brain tumors: Localization to tumor endothelial cells and activation by the endothelial tyrosine kinase receptor (VEGFR-2). Oncogene 2002, 21, 2058–2065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Kang, S.H.; Yu, M.O.; Park, K.J.; Chi, S.G.; Park, D.H.; Chung, Y.G. Activated STAT3 regulates hypoxia-induced angiogenesis and cell migration in human glioblastoma. Neurosurgery 2010, 67, 1386–1395. [Google Scholar] [CrossRef]
  76. Couto, M.; Coelho-Santos, V.; Santos, L.; Fontes-Ribeiro, C.; Silva, A.P.; Gomes CMF. The interplay between glioblastoma and microglia cells leads to endothelial cell monolayer dysfunction via the interleukin-6-induced JAK2/STAT3 pathway. J. Cell Physiol. 2019, 234, 19750–19760. [Google Scholar] [CrossRef]
  77. Li, R.; Li, G.; Deng, L.; Liu, Q.; Dai, J.; Shen, J.; Zhang, J. IL-6 augments the invasiveness of U87MG human glioblastoma multiforme cells via up-regulation of MMP-2 and fascin-1. Oncol. Rep. 2010, 23, 1553–1559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Zheng, Q.; Han, L.; Dong, Y.; Tian, J.; Huang, W.; Liu, Z.; Jia, X.; Jiang, T.; Zhang, J.; Li, X.; et al. JAK2/STAT3 targeted therapy suppresses tumor invasion via disruption of the EGFRvIII/JAK2/STAT3 axis and associated focal adhesion in EGFRvIII-expressing glioblastoma. Neuro-Oncology 2014, 16, 1229–1243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Xu, C.H.; Liu, Y.; Xiao, L.M.; Chen, L.K.; Zheng, S.Y.; Zeng, E.M.; Li, D.H.; Li, Y.P. Silencing microRNA-221/222 cluster suppresses glioblastoma angiogenesis by suppressor of cytokine signaling-3-dependent JAK/STAT pathway. J. Cell Physiol. 2019, 234, 22272–22284. [Google Scholar] [CrossRef] [PubMed]
  80. Hyun Hwang, J.; Smith, C.A.; Salhia, B.; Rutka, J.T. The Role of Fascin in the Migration and Invasiveness of Malignant Glioma Cells. Neoplasia 2008, 10, 149–159. [Google Scholar] [CrossRef] [Green Version]
  81. Ji, P.; Wang, L.; Liu, J.; Mao, P.; Li, R.; Jiang, H.; Lou, M.; Xu, M.; Yu, X. Knockdown of RPL34 inhibits the proliferation and migration of glioma cells through the inactivation of JAK/STAT3 signaling pathway. J. Cell Biochem. 2019, 120, 3259–3267. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, P.; Chen, F.Z.; Jia, Q.B.; Hu, D.F. Upregulation of microRNA-133a and downregulation of connective tissue growth factor suppress cell proliferation, migration, and invasion in human glioma through the JAK/STAT signaling pathway. IUBMB Life 2019, 71, 1857–1875. [Google Scholar] [CrossRef] [PubMed]
  83. Reardon, D.A.; Brandes, A.A.; Omuro, A.; Mulholland, P.; Lim, M.; Wick, A.; Baehring, J.; Ahluwalia, M.S.; Roth, P.; Bähr, O.; et al. Effect of Nivolumab vs Bevacizumab in Patients With Recurrent Glioblastoma: The CheckMate 143 Phase 3 Randomized Clinical Trial. JAMA Oncol. 2020, 6, 1003. [Google Scholar] [CrossRef] [PubMed]
  84. Phase III CheckMate-548 Trial of Opdivo Fails Endpoint. Available online: https://www.thepharmaletter.com/article/phase-iii-checkmate-58-trial-of-opdivo-fails-endpoint (accessed on 23 December 2020).
  85. Johnson, L.A.; Scholler, J.; Ohkuri, T.; Kosaka, A.; Patel, P.R.; McGettigan, S.E.; Nace, A.K.; Dentchev, T.; Thekkat, P.; Loew, A.; et al. Rational development and characterization of humanized anti–EGFR variant III chimeric antigen receptor T cells for glioblastoma. Sci. Transl. Med. 2015, 7, 275ra22. [Google Scholar] [CrossRef] [Green Version]
  86. Ahmed, N.; Salsman, V.S.; Kew, Y.; Shaffer, D.; Powell, S.; Zhang, Y.J.; Grossman, R.G.; Heslop, H.E.; Gottschalk, S. HER2-specific T cells target primary Glioblastoma stem cells and induce regression of autologous experimental tumors. Clin. Cancer Res. 2010, 16, 474–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Bielamowicz, K.; Fousek, K.; Byrd, T.T.; Samaha, H.; Mukherjee, M.; Aware, N.; Wu, M.F.; Orange, J.S.; Sumazin, P.; Man, T.K.; et al. Trivalent CAR T cells overcome interpatient antigenic variability in glioblastoma. Neuro-Oncology 2018, 20, 506–518. [Google Scholar] [CrossRef]
  88. Brown, C.E.; Starr, R.; Aguilar, B.; Shami, A.F.; Martinez, C.; D’Apuzzo, M.; Barish, M.E.; Forman, S.J.; Jensen, M.C. Stem-like tumor initiating cells isolated from IL13Rα2-expressing gliomas are targeted and killed by IL13-zetakine redirected T cells. Clin. Cancer Res. 2012, 18, 2199–2209. [Google Scholar] [CrossRef] [Green Version]
  89. Luoto, S.; Hermelo, I.; Vuorinen, E.M. Computational Characterization of Suppressive Immune Microenvironments in Glioblastoma. Cancer Res. 2018, 78, 5574–5585. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, Q.; Hu, X.; Hu, B. Tumor Evolution of Glioma Intrinsic Gene Expression Subtype Associates with Immunological Changes in the Microenvironment. Cancer Biol. 2016. [Google Scholar] [CrossRef]
  91. Yu, H.; Kortylewski, M.; Pardoll, D. Crosstalk between cancer and immune cells: Role of STAT3 in the tumour microenvironment. Nat. Rev. Immunol. 2007, 7, 41–51. [Google Scholar] [CrossRef]
  92. Kortylewski, M.; Kujawski, M.; Wang, T. Inhibiting STAT3 signaling in the hematopoietic system elicits multicomponent antitumor immunity. Nat. Med. 2005, 11, 1314–1321. [Google Scholar] [CrossRef] [PubMed]
  93. Henrik Heiland, D.; Ravi, V.M.; Behringer, S.P.; Frenking, J.H.; Wurm, J.; Joseph, K.; Garrelfs, N.; Strähle, J.; Heynckes, S.; Grauvogel, J.; et al. Tumor-associated reactive astrocytes aid the evolution of immunosuppressive environment in glioblastoma. Nat. Commun. 2019, 10, 2541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Chuang, H.-Y.; Su, Y.-K.; Liu, H.-W.; Chen, C.H.; Chiu, S.C.; Cho, D.Y.; Lin, S.Z.; Chen, Y.S.; Lin, C.M. Preclinical Evidence of STAT3 Inhibitor Pacritinib Overcoming Temozolomide Resistance via Downregulating miR-21-Enriched Exosomes from M2 Glioblastoma-Associated Macrophages. J. Clin. Med. 2019, 8, 959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Yao, Y.; Ye, H.; Qi, Z.; Mo, L.; Yue, Q.; Baral, A.; Hoon, D.; Vera, J.C.; Heiss, J.D.; Chen, C.C.; et al. B7-H4(B7x)-Mediated Cross-talk between Glioma-Initiating Cells and Macrophages via the IL6/JAK/STAT3 Pathway Lead to Poor Prognosis in Glioma Patients. Clin. Cancer Res. 2016, 1, 2778–2790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Chen, X.; Eksioglu, E.A.; Zhou, J.; Zhang, L.; Djeu, J.; Fortenbery, N.; Epling-Burnette, P.; Van Bijnen, S.; Dolstra, H.; Cannon, J.; et al. Induction of myelodysplasia by myeloid-derived suppressor cells. J. Clin. Investig. 2013, 123, 4595–4611. [Google Scholar] [CrossRef]
  97. Jeon, S.-B.; Yoon, H.J.; Chang, C.Y.; Koh, H.S.; Jeon, S.-H.; Park, E.J. Galectin-3 exerts cytokine-like regulatory actions through the JAK-STAT pathway. J. Immunol. 2010, 185, 7037–7046. [Google Scholar] [CrossRef] [Green Version]
  98. Ko, H.J.; Kim, Y.J. Signal transducer and activator of transcription proteins: Regulators of myeloid-derived suppressor cell-mediated immunosuppression in cancer. Arch. Pharm Res. 2016, 39, 1597–1608. [Google Scholar] [CrossRef]
  99. Saraiva, M.; O’Garra, A. The regulation of IL-10 production by immune cells. Nat. Rev. Immunol. 2010, 10, 170–181. [Google Scholar] [CrossRef] [Green Version]
  100. Yu, H.; Liu, Y.; McFarland, B.C.; Deshane, J.S.; Hurst, D.R.; Ponnazhagan, S.; Benveniste, E.N.; Qin, H. SOCS3 Deficiency in Myeloid Cells Promotes Tumor Development: Involvement of STAT3 Activation and Myeloid-Derived Suppressor Cells. Cancer Immunol. Res. 2015, 727–740. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, T.; Niu, G.; Kortylewski, M.; Burdelya, L.; Shain, K.; Zhang, S.; Bhattacharya, R.; Gabrilovich, D.; Heller, R.; Coppola, D.; et al. Regulation of the innate and adaptive immune responses by Stat-3 signaling in tumor cells. Nat. Med. 2004, 10, 48–54. [Google Scholar] [CrossRef]
  102. Cheng, F.; Wang, H.-W.; Cuenca, A.; Huang, M.; Ghansah, T.; Brayer, J.; Kerr, W.G.; Takeda, K.; Akira, S.; Schoenberger, S.P.; et al. A critical role for Stat3 signaling in immune tolerance. Immunity 2003, 19, 425–436. [Google Scholar] [CrossRef] [Green Version]
  103. Wang, N.; Liang, H.; Zen, K. Molecular mechanisms that influence the macrophage m1-m2 polarization balance. Front. Immunol. 2014, 5, 614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A.; et al. Macrophage plasticity, polarization, and function in health and disease. J. Cell Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef] [PubMed]
  105. Wu, A.; Wei, J.; Kong, L.Y.; Wang, Y.; Priebe, W.; Qiao, W.; Sawaya, R.; Heimberger, A.B. Glioma cancer stem cells induce immunosuppressive macrophages/microglia. Neuro-Oncology 2010, 12, 1113–1125. [Google Scholar] [CrossRef] [PubMed]
  106. Huettner, C.; Czub, S.; Kerkau, S.; Roggendorf, W.; Tonn, J.C. Interleukin 10 is expressed in human gliomas in vivo and increases glioma cell proliferation and motility in vitro. Anticancer Res. 1997, 17, 3217–3224. [Google Scholar]
  107. O’Farrell, A.M.; Liu, Y.; Moore, K.W.; Mui, A.L. IL-10 inhibits macrophage activation and proliferation by distinct signaling mechanisms: Evidence for Stat3-dependent and -independent pathways. EMBO J. 1998, 16, 1006–1018. [Google Scholar] [CrossRef] [PubMed]
  108. Lang, R.; Patel, D.; Morris, J.J.; Rutschman, R.L.; Murray, P.J. Shaping gene expression in activated and resting primary macrophages by IL-10. J. Immunol. 2002, 169, 2253–2263. [Google Scholar] [CrossRef] [Green Version]
  109. 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, 3, 114–123. [Google Scholar] [CrossRef] [Green Version]
  110. Hussain, S.F.; Yang, D.; Suki, D.; Aldape, K.; Grimm, E.; Heimberger, A.B. The role of human glioma-infiltrating microglia/macrophages in mediating antitumor immune responses. Neuro-Oncology 2006, 8, 261–279. [Google Scholar] [CrossRef] [Green Version]
  111. Molavi, O.; Ma, Z.; Hamdy, S.; Lavasanifar, A.; Samuel, J. Immunomodulatory and anticancer effects of intra-tumoral co-delivery of synthetic lipid A adjuvant and STAT3 inhibitor, JSI-124. Immunopharmacol. Immunotoxicol. 2009, 31, 214–221. [Google Scholar] [CrossRef]
  112. Poholek, C.H.; Raphael, I.; Wu, D.; Revu, S.; Rittenhouse, N.; Uche, U.U.; Majumder, S.; Kane, L.P.; Poholek, A.C.; McGeachy, M.J. Noncanonical STAT3 activity sustains pathogenic Th17 proliferation and cytokine response to antigen. J. Exp. Med. 2020, 217. [Google Scholar] [CrossRef] [PubMed]
  113. Yu, H.; Pardoll, D.; Jove, R. STATs in cancer inflammation and immunity: A leading role for STAT3. Nat. Rev. Cancer 2009, 9, 798–809. [Google Scholar] [CrossRef] [PubMed]
  114. Xiao, W.; Klement, J.D.; Lu, C.; Ibrahim, M.L.; Liu, K. IFNAR1 Controls Autocrine Type I IFN Regulation of PD-L1 Expression in Myeloid-Derived Suppressor Cells. J. Immunol. 2018, 1, 264–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Rodriguez, P.C.; Quiceno, D.G.; Zabaleta, J.; Ortiz, B.; Zea, A.H.; Piazuelo, M.B.; Delgado, A.; Correa, P.; Brayer, J.; Sotomayor, E.M.; et al. Arginase I production in the tumor microenvironment by mature myeloid cells inhibits T-cell receptor expression and antigen-specific T-cell responses. Cancer Res. 2004, 15, 5839–5849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Rajappa, P.; Cobb, W.S.; Vartanian, E.; Huang, Y.; Daly, L.; Hoffman, C.; Zhang, J.; Shen, B.; Yanowitch, R.; Garg, K.; et al. Malignant Astrocytic Tumor Progression Potentiated by JAK-mediated Recruitment of Myeloid Cells. Clin. Cancer Res. 2017, 15, 3109–3119. [Google Scholar] [CrossRef] [Green Version]
  117. 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] [Green Version]
  118. Fujita, M.; Zhu, X.; Sasaki, K.; Ueda, R.; Low, K.L.; Pollack, I.F.; Okada, H. Inhibition of STAT3 promotes the efficacy of adoptive transfer therapy using type-1 CTLs by modulation of the immunological microenvironment in a murine intracranial glioma. J. Immunol. 2008, 15, 2089–2098. [Google Scholar] [CrossRef] [Green Version]
  119. Farren, M.R.; Carlson, L.M.; Netherby, C.S.; Lindner, I.; Li, P.K.; Gabrilovich, D.I.; Abrams, S.I.; Lee, K.P. Tumor-induced STAT3 signaling in myeloid cells impairs dendritic cell generation by decreasing PKCβII abundance. Sci. Signal. 2014, 18, ra16. [Google Scholar] [CrossRef] [Green Version]
  120. Chang, Q.; Bournazou, E.; Sansone, P.; Berishaj, M.; Gao, S.P.; Daly, L.; Wels, J.; Theilen, T.; Granitto, S.; Zhang, X.; et al. The IL-6/JAK/Stat3 feed-forward loop drives tumorigenesis and metastasis. Neoplasia 2013, 15, 848–862. [Google Scholar] [CrossRef] [Green Version]
  121. Zhang, L.; Alizadeh, D.; Van Handel, M.; Kortylewski, M.; Yu, H.; Badie, B. Stat3 inhibition activates tumor macrophages and abrogates glioma growth in mice. Glia 2009, 57, 1458–1467. [Google Scholar] [CrossRef]
  122. Hussain, S.F.; Kong, L.Y.; Jordan, J.; Conrad, C.; Madden, T.; Fokt, I.; Priebe, W.; Heimberger, A.B. A novel small molecule inhibitor of signal transducers and activators of transcription 3 reverses immune tolerance in malignant glioma patients. Cancer Res. 2007, 67, 9630–9636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. 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] [PubMed] [Green Version]
  124. Andaloussi, A.E.; Lesniak, M.S. An increase in CD4+CD25+FOXP3+ regulatory T cells in tumor-infiltrating lymphocytes of human glioblastoma multiforme. Neuro-Oncology 2006, 8, 234–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. 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]
  126. Pallandre, J.-R.; Brillard, E.; Créhange, G.; Radlovic, A.; Remy-Martin, J.P.; Saas, P.; Rohrlich, P.S.; Pivot, X.; Ling, X.; Tiberghien, P.; et al. Role of STAT3 in CD4+CD25+FOXP3+ Regulatory Lymphocyte Generation: Implications in Graft-versus-Host Disease and Antitumor Immunity. J. Immunol. 2007, 179, 7593–7604. [Google Scholar] [CrossRef] [Green Version]
  127. Wei, J.; Wu, A.; Kong, L.Y.; Wang, Y.; Fuller, G.; Fokt, I.; Melillo, G.; Priebe, W.; Heimberger, A.B. Hypoxia potentiates glioma-mediated immunosuppression. PLoS ONE 2011, 20, e16195. [Google Scholar] [CrossRef] [Green Version]
  128. Roos, W.; Batista, L.; Naumann, S.; Wick, W.; Weller, M.; Menck, C.F.; Kaina, B. Apoptosis in malignant glioma cells triggered by the temozolomide-induced DNA lesion O6-methylguanine. Oncogene 2007, 26, 186–197. [Google Scholar] [CrossRef] [Green Version]
  129. Fu, D.; Calvo, J.; Samson, L. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nat. Rev. Cancer 2012, 12, 104–120. [Google Scholar] [CrossRef] [Green Version]
  130. Kohsaka, S.; Wang, L.; Yachi, K.; Mahabir, R.; Narita, T.; Itoh, T.; Tanino, M.; Kimura, T.; Nishihara, H.; Tanaka, S. STAT3 Inhibition Overcomes Temozolomide Resistance in Glioblastoma by Downregulating MGMT Expression. Mol. Cancer Ther. 2012, 1289–1299. [Google Scholar] [CrossRef] [Green Version]
  131. Kitange, G.J.; Carlson, B.L.; Schroeder, M.A. Induction of MGMT expression is associated with temozolomide resistance in glioblastoma xenografts. Neuro-Oncology 2009, 11, 281–291. [Google Scholar] [CrossRef] [Green Version]
  132. Wang, Y.; Chen, L.; Bao, Z.; Li, S.; You, G.; Yan, W.; Shi, Z.; Liu, Y.; Yang, P.; Zhang, W.; et al. Inhibition of STAT3 reverses alkylator resistance through modulation of the AKT and β-catenin signaling pathways. Oncol. Rep. 2011, 26, 1173–1180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Lee, E.S.; Ko, K.K.; Joe, Y.A.; Kang, S.G.; Hong, Y.K. Inhibition of STAT3 reverses drug resistance acquired in temozolomide-resistant human glioma cells. Oncol. Lett. 2011, 2, 115–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Li, H.; Chen, L.; Li, J.; Zhou, Q.; Huang, A.; Liu, W.W.; Wang, K.; Gao, L.; Qi, S.T.; Lu, Y.T. miR-519a enhances chemosensitivity and promotes autophagy in glioblastoma by targeting STAT3/Bcl2 signaling pathway. J. Hematol. Oncol. 2018, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Heynckes, S.; Daka, K.; Franco, P.; Gaebelein, A.; Frenking, J.H.; Doria-Medina, R.; Mader, I.; Delev, D.; Schnell, O.; Heiland, D.H. Crosslink between Temozolomide and PD-L1 immune-checkpoint inhibition in glioblastoma multiforme. BMC Cancer 2019, 19, 117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Sato, N.; Mizumoto, K.; Nakamura, M. Radiation-induced centrosome overduplication and multiple mitotic spindles in human tumor cells. Exp. Cell Res. 2000, 255, 321–326. [Google Scholar] [CrossRef]
  137. Rath, B.H.; Wahba, A.; Camphausen, K.; Tofilon, P.J. Coculture with astrocytes reduces the radiosensitivity of glioblastoma stem-like cells and identifies additional targets for radiosensitization. Cancer Med. 2015, 4, 1705–1716. [Google Scholar] [CrossRef]
  138. Yu, H.; Zhang, S.; Ibrahim, A.N.; Wang, J.; Deng, Z.; Wang, M. RCC2 promotes proliferation and radio-resistance in glioblastoma via activating transcription of DNMT1. Biochem. Biophys. Res. Commun. 2019, 27, 999–1006. [Google Scholar] [CrossRef]
  139. Mrowczynski, O.D.; Madhankumar, A.B.; Sundstrom, J.M. Exosomes impact survival to radiation exposure in cell line models of nervous system cancer. Oncotarget 2018, 9, 36083–36101. [Google Scholar] [CrossRef]
  140. Li, C.; Ran, H.; Song, S.; Liu, W.; Zou, W.; Jiang, B.; Zhao, H.; Shao, B. Overexpression of RPN2 suppresses radiosensitivity of glioma cells by activating STAT3 signal transduction. Mol. Med. 2020, 13, 43. [Google Scholar] [CrossRef]
  141. Ventero, M.P.; Fuentes-Baile, M.; Quereda, C.; Perez-Valeciano, E.; Alenda, C.; Garcia-Morales, P.; Saceda, M. Radiotherapy resistance acquisition in Glioblastoma. Role of SOCS1 and SOCS3. PLoS ONE 2019, 14, e0212581. [Google Scholar] [CrossRef] [Green Version]
  142. Maachani, U.B.; Shankavaram, U.; Kramp, T.; Tofilon, P.J.; Camphausen, K.; Tandle, A.T. FOXM1 and STAT3 interaction confers radioresistance in glioblastoma cells. Oncotarget 2016, 22, 77365–77377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Zhong, C.; Tao, B.; Chen, Y. B7-H3 Regulates Glioma Growth and Cell Invasion Through a JAK2/STAT3/Slug-Dependent Signaling Pathway. Onco Targets Ther. 2020, 13, 2215–2224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Lin, J.-C.; Tsai, J.-T.; Chao, T.-Y.; Ma, H.-I.; Liu, W.-H. The STAT3/Slug Axis Enhances Radiation-Induced Tumor Invasion and Cancer Stem-like Properties in Radioresistant Glioblastoma. Cancers 2018, 10, 512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Robe, P.A.; Martin, D.H.; Nguyen-Khac, M.T.; Artesi, M.; Deprez, M.; Albert, A.; Vanbelle, S.; Califice, S.; Bredel, M.; Bours, V. Early termination of ISRCTN45828668, a phase 1/2 prospective, randomized study of sulfasalazine for the treatment of progressing malignant gliomas in adults. BMC Cancer 2009, 9, 372. [Google Scholar] [CrossRef] [Green Version]
  146. Atkinson, G.P.; Nozell, S.E.; Benveniste, E.T. NF-kappaB and STAT3 signaling in glioma: Targets for future therapies. Expert Rev. Neurother. 2010, 10, 575–586. [Google Scholar] [CrossRef] [Green Version]
  147. Zhao, C.; Li, H.; Lin, H.-J.; Yang, S.; Lin, J.; Liang, G. Feedback Activation of STAT3 as a Cancer Drug-Resistance Mechanism. Trends Pharmacol. Sci. 2016, 37, 47–61. [Google Scholar] [CrossRef]
  148. Gao, L.; Li, F.; Dong, B.; Zhang, J.; Rao, Y.; Cong, Y.; Mao, B.; Chen, X. Inhibition of STAT3 and ErbB2 suppresses tumor growth, enhances radiosensitivity, and induces mitochondria-dependent apoptosis in glioma cells. Int. J. Radiat. Oncol. Biol. Phys 2010, 15, 1223–1231. [Google Scholar] [CrossRef]
  149. Jensen, K.V.; Hao, X.; Aman, A.; Luchman, H.A.; Weiss, S. EGFR blockade in GBM brain tumor stem cells synergizes with JAK2/STAT3 pathway inhibition to abrogate compensatory mechanisms in vitro and in vivo. Neuro-Oncol. Adv. 2020, 2, vdaa020. [Google Scholar] [CrossRef] [Green Version]
  150. De Groot, J.; Liang, J.; Kong, L.-Y.; Wei, J.; Piao, Y.; Fuller, G.; Qiao, W.; Heimberger, A.B. Modulating Antiangiogenic Resistance by Inhibiting the Signal Transducer and Activator of Transcription 3 Pathway in Glioblastoma. Oncotarget 2012, 3, 1036–1048. [Google Scholar] [CrossRef] [Green Version]
  151. Batchelor, T.T.; Reardon, D.A.; de Groot, J.F.; Wick, W.; Weller, M. Antiangiogenic therapy for glioblastoma: Current status and future prospects. Clin. Cancer Res. 2014, 20, 5612–5619. [Google Scholar] [CrossRef] [Green Version]
  152. Cruickshanks, N.; Zhang, Y.; Hine, S.; Gibert, M.; Yuan, F.; Oxford, M.; Grello, C.; Pahuski, M.; Dube, C.; Guessous, F.; et al. Discovery and Therapeutic Exploitation of Mechanisms of Resistance to MET Inhibitors in Glioblastoma. Clin. Cancer Res. 2019, 25, 663–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wong, R.A.; Luo, X.; Lu, M.; An, Z.; Haas-Kogan, D.A.; Phillips, J.J.; Shokat, K.M.; Weiss, W.A.; Fan, Q.W. Cooperative Blockade of PKCα and JAK2 Drives Apoptosis in Glioblastoma. Cancer Res. 2020, 15, 709–718. [Google Scholar] [CrossRef] [PubMed]
  154. Ott, M.; Kassab, C.; Marisetty, A.; Hashimoto, Y.; Wei, J.; Zamler, D.; Leu, J.S.; Tomaszowski, K.H.; Sabbagh, A.; Fang, D.; et al. Radiation with STAT3 Blockade Triggers Dendritic Cell-T cell Interactions in the Glioma Microenvironment and Therapeutic Efficacy. Clin. Cancer Res. 2020, 26, 4983–4994. [Google Scholar] [CrossRef] [PubMed]
  155. Herrmann, A.; Kortylewski, M.; Kujawski, M.; Zhang, C.; Reckamp, K.; Armstrong, B.; Wang, L.; Kowolik, C.; Deng, J.; Figlin, R.; et al. Targeting Stat3 in the myeloid compartment drastically improves the in vivo antitumor functions of adoptively transferred T cells. Cancer Res. 2010, 70, 7455–7464. [Google Scholar] [CrossRef] [Green Version]
  156. Kong, L.-Y.; Wei, J.; Fuller, G.N.; Schrand, B.; Gabrusiewicz, K.; Zhou, S.; Rao, G.; Calin, G.; Gilboa, E.; Heimberger, A.B. Tipping a favorable CNS intratumoral immune response using immune stimulation combined with inhibition of tumor-mediated immune suppression. Oncoimmunology 2016, 5, e1117739. [Google Scholar] [CrossRef] [Green Version]
  157. Stechishin, O.D.; Luchman, H.A.; Ruan, Y.; Blough, M.D.; Nguyen, S.A.; Kelly, J.J.; Cairncross, J.G.; Weiss, S. On-target JAK2/STAT3 inhibition slows disease progression in orthotopic xenografts of human glioblastoma brain tumor stem cells. Neuro-Oncology 2013, 15, 198–207. [Google Scholar] [CrossRef] [Green Version]
  158. Lamano, J.B.; Lamano, J.B.; Li, Y.D.; DiDomenico, J.D.; Choy, W.; Veliceasa, D.; Oyon, D.E.; Fakurnejad, S.; Ampie, L.; Kesavabhotla, K.; et al. Glioblastoma-Derived IL6 Induces Immunosuppressive Peripheral Myeloid Cell PD-L1 and Promotes Tumor Growth. Clin. Cancer Res. 2019, 25, 3643–3657. [Google Scholar] [CrossRef] [Green Version]
  159. Kudo, M.; Jono, H.; Shinriki, S.; Yano, S.; Nakamura, H.; Makino, K.; Hide, T.; Muta, D.; Ueda, M.; Ota, K.; et al. Antitumor effect of humanized anti–interleukin-6 receptor antibody (tocilizumab) on glioma cell proliferation: Laboratory investigation. J. Neurosurg. 2009, 111, 219–225. [Google Scholar] [CrossRef] [Green Version]
  160. Nellan, A.; McCully, C.M.L.; Cruz Garcia, R.; Jayaprakash, N.; Widemann, B.C.; Lee, D.W.; Warren, K.E. Improved CNS exposure to tocilizumab after cerebrospinal fluid compared to intravenous administration in rhesus macaques. Blood 2018, 132, 662–666. [Google Scholar] [CrossRef]
  161. 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]
  162. Takabe, H.; Warnken, Z.N.; Zhang, Y.; Davis, D.A.; Smyth, H.; Kuhn, J.G.; Weitman, S.; Williams Iii, R.O. A Repurposed Drug for Brain Cancer: Enhanced Atovaquone Amorphous Solid Dispersion by Combining a Spontaneously Emulsifying Component with a Polymer Carrier. Pharmaceutics 2018, 10, 60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Grimm, S.A.; Marymont, M.; Chandler, J.P.; Muro, K.; Newman, S.B.; Levy, R.M.; Jovanovic, B.; McCarthy, K.; Raizer, J.J. Phase I study of arsenic trioxide and temozolomide in combination with radiation therapy in patients with malignant gliomas. J. Neurooncol. 2012, 110, 237–243. [Google Scholar] [CrossRef] [PubMed]
  164. Kumthekar, P.; Grimm, S.; Chandler, J.; Mehta, M.; Marymont, M.; Levy, R.; Muro, K.; Helenowski, I.; McCarthy, K.; Fountas, L.; et al. A phase II trial of arsenic trioxide and temozolomide in combination with radiation therapy for patients with malignant gliomas. J. Neurooncol. 2017, 133, 589–594. [Google Scholar] [CrossRef] [PubMed]
  165. Sun, H.; Zhang, S. Arsenic trioxide regulates the apoptosis of glioma cell and glioma stem cell via down-regulation of stem cell marker Sox2. Biochem. Biophys. Res. Commun. 2011, 410, 692–697. [Google Scholar] [CrossRef] [PubMed]
  166. Yang, F.; Brown, C.; Buettner, R. Sorafenib induces growth arrest and apoptosis of human glioblastoma cells through the dephosphorylation of signal transducers and activators of transcription 3. Mol. Cancer Ther. 2010, 9, 953–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Carra, E.; Barbieri, F.; Marubbi, D.; Pattarozzi, A.; Favoni, R.E.; Florio, T.; Daga, A. Sorafenib selectively depletes human glioblastoma tumor-initiating cells from primary cultures. Cell Cycle 2013, 12, 491–500. [Google Scholar] [CrossRef] [Green Version]
  168. Hottinger, A.F.; Ben Aissa, A.; Espeli, V. Phase I study of sorafenib combined with radiation therapy and temozolomide as first-line treatment of high-grade glioma. Br. J. Cancer 2014, 110, 2655–2661. [Google Scholar] [CrossRef]
  169. Lee, E.Q.; Kuhn, J.; Lamborn, K.R.; Abrey, L.; DeAngelis, L.M.; Lieberman, F.; Robins, H.I.; Chang, S.M.; Yung, W.K.; Drappatz, J.; et al. Phase I/II study of sorafenib in combination with temsirolimus for recurrent glioblastoma or gliosarcoma: North American Brain Tumor Consortium study 05-02. Neuro-Oncology 2012, 14, 1511–1518. [Google Scholar] [CrossRef] [Green Version]
  170. Zustovich, F.; Landi, L.; Lombardi, G.; Porta, C.; Galli, L.; Fontana, A.; Amoroso, D.; Galli, C.; Andreuccetti, M.; Falcone, A.; et al. Sorafenib plus daily low-dose temozolomide for relapsed glioblastoma: A phase II study. Anticancer Res. 2013, 33, 3487–3494. [Google Scholar] [CrossRef]
  171. Gambini, J.; Inglés, M.; Olaso, G.; Lopez-Grueso, R.; Bonet-Costa, V.; Gimeno-Mallench, L.; Mas-Bargues, C.; Abdelaziz, K.M.; Gomez-Cabrera, M.C.; Vina, J.; et al. Properties of Resveratrol: In Vitro and In Vivo Studies about Metabolism, Bioavailability, and Biological Effects in Animal Models and Humans. Oxid Med. Cell Longev. 2015, 2015, 837042. [Google Scholar] [CrossRef] [Green Version]
  172. Xue, S.; Xiao-Hong, S.; Lin, S.; Jie, B.; Li-Li, W.; Jia-Yao, G.; Shun, S.; Pei-Nan, L.; Mo-Li, W.; Qian, W.; et al. Lumbar puncture-administered resveratrol inhibits STAT3 activation, enhancing autophagy and apoptosis in orthotopic rat glioblastomas. Oncotarget 2016, 7, 75790–75799. [Google Scholar] [CrossRef] [PubMed]
  173. Jhaveri, A.; Deshpande, P.; Pattni, B.; Torchilin, V. Transferrin-targeted, resveratrol-loaded liposomes for the treatment of glioblastoma. J. Control. Release 2018, 277, 89–101. [Google Scholar] [CrossRef] [PubMed]
  174. Rocha, T.G.R.; de Lopes, S.C.A.; Cassali, G.D.; Ferreira, E.; Veloso, E.S.; Leite, E.A.; Tavitian, B. Evaluation of Antitumor Activity of Long-Circulating and pH-Sensitive Liposomes Containing Ursolic Acid in Animal Models of Breast Tumor and Gliosarcoma. Integr. Cancer Ther. 2016, 15, 512–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Zhang, W.; Yu, W.; Cai, G.; Zhu, J.; Zhang, C.; Li, S.; Guo, J.; Yin, G.; Chen, C.; Kong, L. A new synthetic derivative of cryptotanshinone KYZ3 as STAT3 inhibitor for triple-negative breast cancer therapy. Cell Death Dis. 2018, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  176. Verdura, S.; Cuyàs, E.; Llorach-Parés, L.; Pérez-Sánchez, A.; Micol, V.; Nonell-Canals, A.; Joven, J.; Valiente, M.; Sánchez-Martínez, M.; Bosch-Barrera, J.; et al. Silibinin is a direct inhibitor of STAT3. Food Chem. Toxicol. 2018, 116 Pt B, 161–172. [Google Scholar] [CrossRef] [PubMed]
  177. Elhag, R.; Mazzio, E.A.; Soliman, K.F.A. The Effect of Silibinin in Enhancing Toxicity of Temozolomide and Etoposide in p53 and PTEN-mutated Resistant Glioma Cell Lines. Anticancer Res. 2015, 35, 1263–1269. [Google Scholar]
  178. Wang, C.; He, C.; Lu, S.; Wang, X.; Wang, L.; Liang, S.; Wang, X.; Piao, M.; Cui, J.; Chi, G.; et al. Autophagy activated by silibinin contributes to glioma cell death via induction of oxidative stress-mediated BNIP3-dependent nuclear translocation of AIF. Cell Death Dis. 2020, 11, 1–16. [Google Scholar] [CrossRef]
  179. Bosch-Barrera, J.; Sais, E.; Cañete, N.; Marruecos, J.; Cuyàs, E.; Izquierdo, A.; Porta, R.; Haro, M.; Brunet, J.; Pedraza, S.; et al. Response of brain metastasis from lung cancer patients to an oral nutraceutical product containing silibinin. Oncotarget 2016, 7, 32006–32014. [Google Scholar] [CrossRef] [Green Version]
  180. Flaig, T.W.; Glodé, M.; Gustafson, D.; van Bokhoven, A.; Tao, Y.; Wilson, S.; Su, L.J.; Li, Y.; Harrison, G.; Agarwal, R.; et al. A study of high-dose oral silybin-phytosome followed by prostatectomy in patients with localized prostate cancer. Prostate 2010, 70, 848–855. [Google Scholar] [CrossRef]
  181. Lu, L.; Li, C.; Li, D.; Wang, Y.; Zhou, C.; Shao, W.; Peng, J.; You, Y.; Zhang, X.; Shen, X. Cryptotanshinone inhibits human glioma cell proliferation by suppressing STAT3 signaling. Mol. Cell Biochem. 2013, 381, 273–282. [Google Scholar] [CrossRef]
  182. Cai, Y.; Zhang, W.; Chen, Z.; Shi, Z.; He, C.; Chen, M. Recent insights into the biological activities and drug delivery systems of tanshinones. Int. J. Nanomed. 2016, 11, 121–130. [Google Scholar] [CrossRef] [Green Version]
  183. Wang, X.; Yu, Z.; Wang, C.; Cheng, W.; Tian, X.; Huo, X.; Wang, Y.; Sun, C.; Feng, L.; Xing, J.; et al. Alantolactone, a natural sesquiterpene lactone, has potent antitumor activity against glioblastoma by targeting IKKβ kinase activity and interrupting NF-κB/COX-2-mediated signaling cascades. J. Exp. Clin. Cancer Res. 2017, 36. [Google Scholar] [CrossRef] [PubMed]
  184. Khan, M.; Yi, F.; Rasul, A.; Li, T.; Wang, N.; Gao, H.; Gao, R.; Ma, T. Alantolactone induces apoptosis in glioblastoma cells via GSH depletion, ROS generation, and mitochondrial dysfunction. IUBMB Life 2012, 64, 783–794. [Google Scholar] [CrossRef] [PubMed]
  185. Matias, D.; Balça-Silva, J.; Dubois, L.G.; Pontes, B.; Ferrer, V.P.; Rosário, L.; do Carmo, A.; Echevarria-Lima, J.; Sarmento-Ribeiro, A.B.; Lopes, M.C.; et al. Dual treatment with shikonin and temozolomide reduces glioblastoma tumor growth, migration and glial-to-mesenchymal transition. Cell Oncol. (Dordr.) 2017, 40, 247–261. [Google Scholar] [CrossRef] [PubMed]
  186. Zhao, Q.; Kretschmer, N.; Bauer, R.; Efferth, T. Shikonin and its derivatives inhibit the epidermal growth factor receptor signaling and synergistically kill glioblastoma cells in combination with erlotinib. Int. J. Cancer 2015, 137, 1446–1456. [Google Scholar] [CrossRef] [PubMed]
  187. Miao, Z.; Yu, F.; Ren, Y.; Yang, J. d,l-Sulforaphane Induces ROS-Dependent Apoptosis in Human Gliomablastoma Cells by Inactivating STAT3 Signaling Pathway. Int. J. Mol. Sci. 2017, 18, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Karmakar, S.; Weinberg, M.S.; Banik, N.L.; Patel, S.J.; Ray, S.K. Activation of multiple molecular mechanisms for apoptosis in human malignant glioblastoma T98G and U87MG cells treated with sulforaphane. Neuroscience 2006, 141, 1265–1280. [Google Scholar] [CrossRef]
  189. Bijangi-Vishehsaraei, K.; Saadatzadeh, M.R.; Wang, H.; Nguyen, A.; Kamocka, M.M.; Cai, W.; Cohen-Gadol, A.A.; Halum, S.L.; Sarkaria, J.N.; Pollok, K.E.; et al. Sulforaphane suppresses the growth of glioblastoma cells, glioblastoma stem cell–like spheroids, and tumor xenografts through multiple cell signaling pathways. J. Neurosurg. 2017, 127, 1219–1230. [Google Scholar] [CrossRef]
  190. Kim, B.; Lee, K.Y.; Park, B. Crocin Suppresses Constitutively Active STAT3 through Induction of Protein Tyrosine Phosphatase SHP-1. J. Cell Biochem. 2017, 118, 3290–3298. [Google Scholar] [CrossRef]
  191. Colapietro, A.; Mancini, A.; Vitale, F.; Martellucci, S.; Angelucci, A.; Llorens, S.; Mattei, V.; Gravina, G.L.; Alonso, G.L.; Festuccia, C. Crocetin Extracted from Saffron Shows Antitumor Effects in Models of Human Glioblastoma. Int. J. Mol. Sci. 2020, 21, 423. [Google Scholar] [CrossRef] [Green Version]
  192. Lautenschläger, M.; Sendker, J.; Hüwel, S.; Galla, H.J.; Brandt, S.; Düfer, M.; Riehemann, K.; Hensel, A. Intestinal formation of trans-crocetin from saffron extract (Crocus sativus L.) and in vitro permeation through intestinal and blood brain barrier. Phytomedicine 2015, 22, 36–44. [Google Scholar] [CrossRef] [PubMed]
  193. Wu, N.; Liu, J.; Zhao, X.; Yan, Z.; Jiang, B.; Wang, L.; Cao, S.; Shi, D.; & Lin, X. Cardamonin induces apoptosis by suppressing STAT3 signaling pathway in glioblastoma stem cells. Tumour Biol. 2015, 36, 9667–9676. [Google Scholar] [CrossRef] [PubMed]
  194. Jaiswal, S.; Sharma, A.; Shukla, M.; Lal, J. Gender-related pharmacokinetics and bioavailability of a novel anticancer chalcone, cardamonin, in rats determined by liquid chromatography tandem mass spectrometry. J. Chromatogr. B 2015, 986–987, 23–30. [Google Scholar] [CrossRef] [PubMed]
  195. Zhou, T.; Yang, Y.; Zhang, H.; Che, Y.; Wang, W.; Lv, H.; Li, J.; Wang, Y.; Hou, S. Serenoa Repens Induces Growth Arrest, Apoptosis and Inactivation of STAT3 Signaling in Human Glioma Cells. Technol. Cancer Res. Treat. 2015, 14, 729–736. [Google Scholar] [CrossRef] [PubMed]
  196. Wei, M.; Ma, R.; Huang, S.; Liao, Y.; Ding, Y.; Li, Z.; Guo, Q.; Tan, R.; Zhang, L.; Zhao, L. Oroxylin A increases the sensitivity of temozolomide on glioma cells by hypoxia-inducible factor 1α/hedgehog pathway under hypoxia. J. Cell Physiol. 2019, 234, 17392–17404. [Google Scholar] [CrossRef]
  197. Zou, M.; Hu, C.; You, Q.; Zhang, A.; Wang, X.; Guo, Q. Oroxylin A induces autophagy in human malignant glioma cells via the mTOR-STAT3-Notch signaling pathway. Mol. Carcinog. 2015, 54, 1363–1375. [Google Scholar] [CrossRef] [PubMed]
  198. Ren, G.; Chen, H.; Zhang, M.; Yang, N.; Yang, H.; Xu, C.; Li, J.; Ning, C.; Song, Z.; Zhou, S.; et al. Pharmacokinetics, tissue distribution and excretion study of Oroxylin, A.; Oroxylin A 7-O-glucuronide and Oroxylin A sodium sulfonate in rats after administration of Oroxylin, A. Fitoterapia 2020, 142, 104480. [Google Scholar] [CrossRef]
  199. Michaud-Levesque, J.; Bousquet-Gagnon, N.; Béliveau, R. Quercetin abrogates IL-6/STAT3 signaling and inhibits glioblastoma cell line growth and migration. Exp. Cell Res. 2012, 1, 925–935. [Google Scholar] [CrossRef]
  200. Fujiwara, Y.; Komohara, Y.; Kudo, R.; Tsurushima, K.; Ohnishi, K.; Ikeda, T.; Takeya, M. Oleanolic acid inhibits macrophage differentiation into the M2 phenotype and glioblastoma cell proliferation by suppressing the activation of STAT3. Oncol. Rep. 2011, 26, 1533–1537. [Google Scholar] [CrossRef]
  201. Guo, G.; Yao, W.; Zhang, Q.; Bo, Y. Oleanolic acid suppresses migration and invasion of malignant glioma cells by inactivating MAPK/ERK signaling pathway. PLoS ONE 2013, 8, e72079. [Google Scholar] [CrossRef]
  202. Jiang, Q.; Yang, X.; Du, P.; Zhang, H.; Zhang, T. Dual strategies to improve oral bioavailability of oleanolic acid: Enhancing water-solubility, permeability and inhibiting cytochrome P450 isozymes. Eur. J. Pharm. Biopharm. 2016, 99, 65–72. [Google Scholar] [CrossRef] [PubMed]
  203. Perry, M.-C.; Demeule, M.; Régina, A.; Moumdjian, R.; Béliveau, R. Curcumin inhibits tumor growth and angiogenesis in glioblastoma xenografts. Mol. Nutr. Food Res. 2010, 54, 1192–1201. [Google Scholar] [CrossRef] [PubMed]
  204. Wang, W.H.; Shen, C.Y.; Chien, Y.C.; Chang, W.S.; Tsai, C.W.; Lin, Y.H.; Hwang, J.J. Validation of Enhancing Effects of Curcumin on Radiotherapy with F98/FGT Glioblastoma-Bearing Rat Model. Int. J. Mol. Sci. 2020, 19, 4385. [Google Scholar] [CrossRef] [PubMed]
  205. Dei Cas, M.; Ghidoni, R. Dietary Curcumin: Correlation between Bioavailability and Health Potential. Nutrients 2019, 11, 2147. [Google Scholar] [CrossRef] [Green Version]
  206. Cho, H.J.; Park, J.H.; Nam, J.H.; Chang, Y.C.; Park, B.; Hoe, H.S. Ascochlorin Suppresses MMP-2-Mediated Migration and Invasion by Targeting FAK and JAK-STAT Signaling Cascades. J. Cell Biochem. 2018, 119, 300–313. [Google Scholar] [CrossRef]
  207. Magae, J.; Tsuruga, M.; Maruyama, A.; Furukawa, C.; Kojima, S.; Shimizu, H.; Ando, K. Relationship between peroxisome proliferator-activated receptor-γ activation and the ameliorative effects of ascochlorin derivatives on type II diabetes. J. Antibiot. 2009, 62, 365–369. [Google Scholar] [CrossRef] [Green Version]
  208. Ge, W.; Chen, X.; Han, F.; Liu, Z.; Wang, T.; Wang, M.; Chen, Y.; Ding, Y.; Zhang, Q. Synthesis of Cucurbitacin B Derivatives as Potential Anti-Hepatocellular Carcinoma Agents. Molecules 2018, 23, 3345. [Google Scholar] [CrossRef] [Green Version]
  209. Premkumar, D.R.; Jane, E.P.; Pollack, I.F. Cucurbitacin-I inhibits Aurora kinase A.; Aurora kinase B and survivin, induces defects in cell cycle progression and promotes ABT-737-induced cell death in a caspase-independent manner in malignant human glioma cells. Cancer Biol. Ther. 2015, 16, 233–243. [Google Scholar] [CrossRef] [Green Version]
  210. Almeida, L.; Vaz-da-Silva, M.; Falcão, A.; Soares, E.; Costa, R.; Loureiro, A.I.; Fernandes-Lopes, C.; Rocha, J.F.; Nunes, T.; Wright, L.; et al. Pharmacokinetic and safety profile of trans-resveratrol in a rising multiple-dose study in healthy volunteers. Mol. Nutr. Food Res. 2009, 53 (Suppl. 1), S7–S15. [Google Scholar] [CrossRef]
  211. Baskin, R.; Park, S.O.; Keserű, G.M.; Bisht, K.S.; Wamsley, H.L.; Sayeski, P.P. The Jak2 small molecule inhibitor, G6, reduces the tumorigenic potential of T98G glioblastoma cells in vitro and in vivo. PLoS ONE 2014, 27, e105568. [Google Scholar] [CrossRef] [Green Version]
  212. Mukthavaram, R.; Ouyang, X.; Saklecha, R. Effect of the JAK2/STAT3 inhibitor SAR317461 on human glioblastoma tumorspheres. J. Transl. Med. 2015, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Nefedova, Y.; Nagaraj, S.; Rosenbauer, A.; Muro-Cacho, C.; Sebti, S.M.; Gabrilovich, D.I. Regulation of dendritic cell differentiation and antitumor immune response in cancer by pharmacologic-selective inhibition of the janus-activated kinase 2/signal transducers and activators of transcription 3 pathway. Cancer Res. 2005, 15, 9525–9535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. McFarland, B.C.; Gray, G.K.; Nozell, S.E.; Hong, S.W.; Benveniste, E.N. Activation of the NF-κB pathway by the STAT3 inhibitor JSI-124 in human glioblastoma cells. Mol. Cancer Res. 2013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Su, Y.; Li, G.; Zhang, X.; Gu, J.; Zhang, C.; Tian, Z.; Zhang, J. JSI-124 Inhibits Glioblastoma Multiforme Cell Proliferation through G(2)/M Cell Cycle Arrest and Apoptosis Augment. Cancer Biol. Ther. 2008, 1243–1249. [Google Scholar] [CrossRef] [Green Version]
  216. Plimack, E.R.; Lorusso, P.M.; McCoon, P.; Tang, W.; Krebs, A.D.; Curt, G.; Eckhardt, S.G. AZD1480: A phase I study of a novel JAK2 inhibitor in solid tumors. Oncologist 2013, 18, 819–820. [Google Scholar] [CrossRef] [Green Version]
  217. McFarland, B.C.; Ma, J.Y.; Langford, C.P.; Gillespie, G.Y.; Yu, H.; Zheng, Y.; Nozell, S.E.; Huszar, D.; Benveniste, E.N. Therapeutic Potential of AZD1480 for the Treatment of Human Glioblastoma. Mol. Cancer Ther. 2011, 2384–2393. [Google Scholar] [CrossRef] [Green Version]
  218. Jensen, K.V.; Cseh, O.; Aman, A.; Weiss, S.; Luchman, H.A. The JAK2/STAT3 inhibitor pacritinib effectively inhibits patient-derived GBM brain tumor initiating cells in vitro and when used in combination with temozolomide increases survival in an orthotopic xenograft model. PLoS ONE 2017, 18, e0189670. [Google Scholar] [CrossRef] [Green Version]
  219. Delen, E.; Doganlar, O.; Doganlar, Z.B.; Delen, O. Inhibition of the Invasion of Human Glioblastoma U87 Cell Line by Ruxolitinib: A Molecular Player of miR-17 and miR-20a Regulating JAK/STAT Pathway. Turk. Neurosurg. 2020, 30, 182–189. [Google Scholar] [CrossRef]
  220. Haile, W.B.; Gavegnano, C.; Tao, S.; Jiang, Y.; Schinazi, R.F.; Tyor, W.R. The Janus kinase inhibitor ruxolitinib reduces HIV replication in human macrophages and ameliorates HIV encephalitis in a murine model. Neurobiol. Dis. 2016, 92 Pt B, 137–143. [Google Scholar] [CrossRef] [Green Version]
  221. Kurokawa, C.; Iankov, I.D.; Anderson, S.K.; Aderca, I.; Leontovich, A.A.; Maurer, M.J.; Oberg, A.L.; Schroeder, M.A.; Giannini, C.; Greiner, S.M.; et al. Constitutive Interferon Pathway Activation in Tumors as an Efficacy Determinant Following Oncolytic Virotherapy. J. Natl. Cancer Inst. 2018, 110, 1123–1132. [Google Scholar] [CrossRef] [Green Version]
  222. 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] [PubMed] [Green Version]
  223. Fuh, B.; Sobo, M.; Cen, L. LLL-3 inhibits STAT3 activity, suppresses glioblastoma cell growth and prolongs survival in a mouse glioblastoma model. Br. J. Cancer 2009, 100, 106–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Ball, S.; Li, C.; Li, P.K.; Lin, J. The small molecule, LLL12, inhibits STAT3 phosphorylation and induces apoptosis in medulloblastoma and glioblastoma cells. PLoS ONE 2011, 19, e18820. [Google Scholar] [CrossRef] [PubMed]
  225. Ashizawa, T.; Akiyama, Y.; Miyata, H.; Iizuka, A.; Komiyama, M.; Kume, A.; Omiya, M.; Sugino, T.; Asai, A.; Hayashi, N.; et al. Effect of the STAT3 inhibitor STX-0119 on the proliferation of a temozolomide-resistant glioblastoma cell line. Int. J. Oncol. 2014, 45, 411–418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. 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]
  227. Kortylewski, M.; Swiderski, P.; Herrmann, A.; Wang, L.; Kowolik, C.; Kujawski, M.; Lee, H.; Scuto, A.; Liu, Y.; Yang, C.; et al. In vivo delivery of siRNA to immune cells by conjugation to a TLR9 agonist enhances antitumor immune responses. Nat. Biotechnol. 2009, 27, 925–932. [Google Scholar] [CrossRef] [Green Version]
  228. Wei, J.; Wang, F.; Kong, L.Y.; Xu, S.; Doucette, T.; Ferguson, S.D.; Yang, Y.; McEnery, K.; Jethwa, K.; Gjyshi, O.; et al. miR-124 inhibits STAT3 signaling to enhance T cell-mediated immune clearance of glioma. Cancer Res. 2013, 1, 3913–3926. [Google Scholar] [CrossRef] [Green Version]
  229. Yaghi, N.K.; Wei, J.; Hashimoto, Y.; Kong, L.Y.; Gabrusiewicz, K.; Nduom, E.K.; Ling, X.; Huang, N.; Zhou, S.; Kerrigan, B.C.; et al. Immune modulatory nanoparticle therapeutics for intracerebral glioma. Neuro-Oncology 2017, 19, 372–382. [Google Scholar] [CrossRef] [Green Version]
  230. Linder, B.; Weirauch, U.; Ewe, A.; Uhmann, A.; Seifert, V.; Mittelbronn, M.; Harter, P.N.; Aigner, A.; Kögel, D. Therapeutic Targeting of Stat3 Using Lipopolyplex Nanoparticle-Formulated siRNA in a Syngeneic Orthotopic Mouse Glioma Model. Cancers 2019, 11, 333. [Google Scholar] [CrossRef] [Green Version]
  231. Hendruschk, S.; Wiedemuth, R.; Aigner, A.; Töpfer, K.; Cartellieri, M.; Martin, D.; Kirsch, M.; Ikonomidou, C.; Schackert, G.; Temme, A. RNA interference targeting survivin exerts antitumoral effects in vitro and in established glioma xenografts in vivo. Neuro-Oncology 2011, 13, 1074–1089. [Google Scholar] [CrossRef] [Green Version]
  232. Masliantsev, K.; Pinel, B.; Balbous, A.; Guichet, P.O.; Tachon, G.; Milin, S.; Godet, J.; Duchesne, M.; Berger, A.; Petropoulos, C.; et al. Impact of STAT3 phosphorylation in glioblastoma stem cells radiosensitization and patient outcome. Oncotarget 2017, 9, 3968–3979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Leidgens, V.; Proske, J.; Rauer, L.; Moeckel, S.; Renner, K.; Bogdahn, U.; Riemenschneider, M.J.; Proescholdt, M.; Vollmann-Zwerenz, A.; Hau, P.; et al. Stattic and metformin inhibit brain tumor initiating cells by reducing STAT3-phosphorylation. Oncotarget 2017, 8, 8250–8263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Debnath, B.; Xu, S.; Neamati, N. Small molecule inhibitors of signal transducer and activator of transcription 3 (Stat3) protein. J. Med. Chem. 2012, 55, 6645–6668. [Google Scholar] [CrossRef] [PubMed]
  235. Sanseverino, I.; Purificato, C.; Gauzzi, M.C.; Gessani, S. Revisiting the Specificity of Small Molecule Inhibitors: The Example of Stattic in Dendritic Cells. Chem. Biol. 2012, 19, 1213–1214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Senft, C.; Priester, M.; Polacin, M.; Schröder, K.; Seifert, V.; Kögel, D.; Weissenberger, J. Inhibition of the JAK-2/STAT3 signaling pathway impedes the migratory and invasive potential of human glioblastoma cells. J. Neurooncol. 2011, 101, 393–403. [Google Scholar] [CrossRef] [PubMed]
  237. Haftchenary, S.; Luchman, H.A.; Jouk, A.O.; Veloso, A.J.; Page, B.D.; Cheng, X.R.; Dawson, S.S.; Grinshtein, N.; Shahani, V.M.; Kerman, K.; et al. Potent Targeting of the STAT3 Protein in Brain Cancer Stem Cells: A Promising Route for Treating Glioblastoma. ACS Med. Chem. Lett. 2013, 4, 1102–1107. [Google Scholar] [CrossRef] [Green Version]
  238. Yue, P.; Lopez-Tapia, F.; Paladino, D.; Li, Y.; Chen, C.H.; Namanja, A.T.; Hilliard, T.; Chen, Y.; Tius, M.A.; Turkson, J. Hydroxamic acid and benzoic acid-based Stat3 inhibitors suppress human glioma and breast cancer phenotypes in vitro and in vivo. Cancer Res. 2016, 76, 652–663. [Google Scholar] [CrossRef] [Green Version]
  239. Cui, P.; Wei, F.; Hou, J.; Su, Y.; Wang, J.; Wang, S. STAT3 inhibition induced temozolomide-resistant glioblastoma apoptosis via triggering mitochondrial STAT3 translocation and respiratory chain dysfunction. Cell. Signal. 2020, 71, 109598. [Google Scholar] [CrossRef]
  240. Sai, K.; Wang, S.; Balasubramaniyan, V.; Conrad, C.; Lang, F.F.; Aldape, K.; Szymanski, S.; Fokt, I.; Dasgupta, A.; Madden, T.; et al. Induction of cell-cycle arrest and apoptosis in glioblastoma stem-like cells by WP1193, a novel small molecule inhibitor of the JAK2/STAT3 pathway. J. Neurooncol. 2012, 107, 487–501. [Google Scholar] [CrossRef]
  241. Ferrajoli, A.; Faderl, S.; Van, Q.; Koch, P.; Harris, D.; Liu, Z.; Hazan-Halevy, I.; Wang, Y.; Kantarjian, H.M.; Priebe, W.; et al. WP1066 disrupts Janus kinase-2 and induces caspase-dependent apoptosis in acute myelogenous leukemia cells. Cancer Res. 2007, 1, 11291–11299. [Google Scholar] [CrossRef] [Green Version]
  242. Han, D.; Yu, T.; Dong, N.; Wang, B.; Sun, F.; Jiang, D. Napabucasin, a novel STAT3 inhibitor suppresses proliferation, invasion and stemness of glioblastoma cells. J. Exp. Clin. Cancer Res. 2019, 5, 289. [Google Scholar] [CrossRef] [PubMed]
  243. Li, Y.; Rogoff, H.A.; Keates, S.; Gao, Y.; Murikipudi, S.; Mikule, K.; Leggett, D.; Li, W.; Pardee, A.B.; Li, C.J. Suppression of cancer relapse and metastasis by inhibiting cancer stemness. Proc. Natl. Acad. Sci. USA 2015, 112, 1839–1844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Wen, Z.; Zhong, Z.; Darnell, J.E., Jr. Maximal activation of transcription by Stat1 and Stat3 requires both tyrosine and serine phosphorylation. Cell 1995. [Google Scholar] [CrossRef] [Green Version]
  245. Tan, M.S.Y.; Sandanaraj, E.; Chong, Y.K. A STAT3-based gene signature stratifies glioma patients for targeted therapy. Nat. Commun. 2019, 10, 3601. [Google Scholar] [CrossRef]
Figure 1. Physiologic JAK-STAT signaling. The classical JAK/STAT signaling pathway begins with (A) recognition and binding of an extracellular cytokine by its respective receptor leading to (B) receptor dimerization, docking of JAK protein, JAK autophosphorylation, and phosphorylation of the receptor’s cytoplasmic tail. STAT proteins are then able to (C) bind via their SH2 domains to these activated receptors, undergoing the critical activating step of tyrosine phosphorylation that allows them to dimerize with other STAT proteins and (D) translocate to the nucleus to effect transcription of target genes. Abbreviations: JAK—Janus kinase; STAT—signal transducer and activator of transcription.
Figure 1. Physiologic JAK-STAT signaling. The classical JAK/STAT signaling pathway begins with (A) recognition and binding of an extracellular cytokine by its respective receptor leading to (B) receptor dimerization, docking of JAK protein, JAK autophosphorylation, and phosphorylation of the receptor’s cytoplasmic tail. STAT proteins are then able to (C) bind via their SH2 domains to these activated receptors, undergoing the critical activating step of tyrosine phosphorylation that allows them to dimerize with other STAT proteins and (D) translocate to the nucleus to effect transcription of target genes. Abbreviations: JAK—Janus kinase; STAT—signal transducer and activator of transcription.
Cancers 13 00437 g001
Figure 2. STAT3 pathway activation represents a focal point of tumorigenesis and immune escape. Aberrant STAT3 activation occurs as a result of several potential upstream and downstream regulatory events including growth factor receptor signaling (e.g., epidermal growth factor (EGFR), platelet-derived growth factor (PDGF), and c-MET)), inhibition of negative regulators of STAT3 (e.g., protein tyrosine phosphatases (PTPs), suppressors of cytokine signaling (SOCS), and protein inhibitor of activated STAT 3 (PIAS3)), and microenvironmental cytokine crosstalk between immune and glioma cells. STAT3 activation transcriptionally upregulates key genes involved in proliferation, stem cell self-renewal, angiogenesis, invasiveness, and formation of the immune microenvironment. The balance of microenvironmental cytokines favors the infiltration of immunosuppressive immune cell populations, including myeloid derived suppressor cells (MDSCs), tumor-associated macrophages/microglia (TAMs), and Tregs. These cytokines activate STAT3 signaling within immune cell populations to increase immunosuppressive macrophage polarization, decreased antigen presentation, and decreased T cell activation.
Figure 2. STAT3 pathway activation represents a focal point of tumorigenesis and immune escape. Aberrant STAT3 activation occurs as a result of several potential upstream and downstream regulatory events including growth factor receptor signaling (e.g., epidermal growth factor (EGFR), platelet-derived growth factor (PDGF), and c-MET)), inhibition of negative regulators of STAT3 (e.g., protein tyrosine phosphatases (PTPs), suppressors of cytokine signaling (SOCS), and protein inhibitor of activated STAT 3 (PIAS3)), and microenvironmental cytokine crosstalk between immune and glioma cells. STAT3 activation transcriptionally upregulates key genes involved in proliferation, stem cell self-renewal, angiogenesis, invasiveness, and formation of the immune microenvironment. The balance of microenvironmental cytokines favors the infiltration of immunosuppressive immune cell populations, including myeloid derived suppressor cells (MDSCs), tumor-associated macrophages/microglia (TAMs), and Tregs. These cytokines activate STAT3 signaling within immune cell populations to increase immunosuppressive macrophage polarization, decreased antigen presentation, and decreased T cell activation.
Cancers 13 00437 g002
Figure 3. STAT3 activation in glioma stem cells. A number of dysregulated extra- and intra-cellular signaling pathways lead to constitutive STAT3 activation in glioma stem cells, which are important progenitor cell populations that are capable of seeding recurrent disease. Chief among these is aberrant cytokine-related JAK/STAT signaling, which includes both autocrine (e.g., IL-6 and leukemia inhibitory factor (LIF)) and paracrine (e.g., IL-6, transforming growth factor (TGF)-β among others) factors. Intracellular regulators include miR-30 (which binds and inhibits the STAT3-inhibitory protein SOCS3) and enhancer of zeste homolog 2 (EZH2), which methylates STAT3 to promote its constitutive activation. The net effect is to promote self-renewal, prevent apoptosis and differentiation.
Figure 3. STAT3 activation in glioma stem cells. A number of dysregulated extra- and intra-cellular signaling pathways lead to constitutive STAT3 activation in glioma stem cells, which are important progenitor cell populations that are capable of seeding recurrent disease. Chief among these is aberrant cytokine-related JAK/STAT signaling, which includes both autocrine (e.g., IL-6 and leukemia inhibitory factor (LIF)) and paracrine (e.g., IL-6, transforming growth factor (TGF)-β among others) factors. Intracellular regulators include miR-30 (which binds and inhibits the STAT3-inhibitory protein SOCS3) and enhancer of zeste homolog 2 (EZH2), which methylates STAT3 to promote its constitutive activation. The net effect is to promote self-renewal, prevent apoptosis and differentiation.
Cancers 13 00437 g003
Table 1. Natural compounds with STAT3 inhibitory activity evaluated in glioma.
Table 1. Natural compounds with STAT3 inhibitory activity evaluated in glioma.
Natural
Compound
MechanismPreclinical Evidence of Efficacy in GliomaLimitationsReferences
SilibininSTAT3 inhibition, autophagy,
chemosensitization
In vitroPoor oral bioavailability, low potency[176,177,178,179,180]
CryptotanshinoneSTAT3 inhibitionIn vitroPoor bioavailability[181,182]
AlantolactoneSTAT3, NF-κB inhibitionIn vitroPoor bioavailability, rapid metabolism, low potency[183,184]
ShikoninSTAT3, EGFR inhibitionIn vitroPoor bioavailability, low potency [185,186]
SulforaphaneJAK2, STAT3, NF-κB inhibitionIn vitroPoor bioavailability, moderate potency[187,188,189]
CrocetinSTAT3 inhibition (SHP-1 induction)In vitroPoor bioavailability, limited BBB penetrance, low potency[190,191,192]
CardamoninSTAT3 inhibitionIn vitroPoor bioavailability[193,194]
Serenoa repens
(Saw palmetto)
STAT3 inhibitionIn vitroPoor bioavailability[195]
Oroxylin AmTOR, STAT3 inhibitionIn vitroPoor bioavailability, rapid metabolism, low potency[196,197,198]
QuercetinIL-6, STAT3 inhibitionIn vitroPoor bioavailability, rapid metabolism, low potency[199]
Oleanolic acidSTAT3 inhibition, IL-10 inhibitionIn vitroPoor bioavailability, rapid metabolism, low potency[200,201,202]
CucurminJAK1, JAK2, STAT3 inhibitionIn vitroPoor bioavailability, low potency[203,204,205]
AscochlorinFAK, STAT3 inhibitionIn vitroPoor bioavailability, low potency[206,207]
CucurbitacinJAK, STAT3, PI3K, MAPK inhibitionIn vitroPoor bioavailability, specificity, high toxicity[208,209]
ResveratrolSTAT3 inhibitionIn vitro
In vivo
Poor bioavailability, low potency[171,172,173,210]
Table 2. Pharmacologic inhibitors of STAT3 signaling investigated in glioma.
Table 2. Pharmacologic inhibitors of STAT3 signaling investigated in glioma.
AgentMechanismEvidence of Efficacy in GliomaNotesReferences
G6JAK2 In vitro In vivo studies lacking, therapeutic requirement for JAK2 overexpression[211]
SAR317461JAK2 In vitro In vivo studies lacking, compensatory autophagy[212]
AZD1480JAK1/JAK2 In vitro
In vivo
Unacceptable dose-limiting toxicities[150,216,217]
JSI-124JAK2 In vitro
In vivo
Anti-proliferative, immune modulatory[213,214,215]
PacritinibJAK2 In vitro
In vivo
BBB penetrant, chemosensitizing[94,218]
* RuxolitinibJAK1/JAK2 In vitro
In vivo
BBB penetrant, anti-proliferative, radiosensitizing, immune modulatory[219,220]
PY * LKTKSTAT3 In vitroIn vivo studies lacking, low potency[222]
LLL12STAT3 In vitro
In vivo
Potent, low solubility/poor bioavailability, unclear BBB penetrance[223,224]
STX-0119STAT3In vitro
In vivo
Minimal growth inhibition of GBM in mouse model [225,226]
AG490STAT3, JAK2In vitroLow potency, in vivo efficacy lacking[73,236]
StatticSTAT3, STAT1, STAT2In vitroSusceptible to intracellular modification, in vivo efficacy lacking, low specificity[232,233,234,235]
WP1193STAT3, JAK2In vitroIn vivo efficacy data lacking[240]
SH-4-54STAT3
STAT5
In vitro
In vivo
BBB penetrant, potent, specific, in vivo studies in subcutaneously implanted GBMs[237,238,239]
* Napabucasin (BBI608)STAT3In vitro
In vivo
Bioavailable, BBB penetrant[242,243]
* WP1066STAT3, JAK2 In vitro
In vivo
Bioavailable, BBB penetrant, immune modulatory[73,122,241]
* Currently being evaluated in clinical trials.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ou, A.; Ott, M.; Fang, D.; Heimberger, A.B. The Role and Therapeutic Targeting of JAK/STAT Signaling in Glioblastoma. Cancers 2021, 13, 437. https://doi.org/10.3390/cancers13030437

AMA Style

Ou A, Ott M, Fang D, Heimberger AB. The Role and Therapeutic Targeting of JAK/STAT Signaling in Glioblastoma. Cancers. 2021; 13(3):437. https://doi.org/10.3390/cancers13030437

Chicago/Turabian Style

Ou, Alexander, Martina Ott, Dexing Fang, and Amy B. Heimberger. 2021. "The Role and Therapeutic Targeting of JAK/STAT Signaling in Glioblastoma" Cancers 13, no. 3: 437. https://doi.org/10.3390/cancers13030437

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