Next Article in Journal
Histopathological Spectrum and Molecular Characterization of Liver Tumors in the Setting of Fontan-Associated Liver Disease
Next Article in Special Issue
Haploinsufficiency of Adenomatous Polyposis Coli Coupled with Kirsten Rat Sarcoma Viral Oncogene Homologue Activation and P53 Loss Provokes High-Grade Glioblastoma Formation in Mice
Previous Article in Journal
Common Misconceptions about Diet and Breast Cancer: An Unclear Issue to Dispel
Previous Article in Special Issue
Integrating Multi-Omics Analysis for Enhanced Diagnosis and Treatment of Glioblastoma: A Comprehensive Data-Driven Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Glioma–Immune Cell Crosstalk in Tumor Progression

by
Mahmoud Elguindy
,
Jacob S. Young
,
Isha Mondal
,
Rongze O. Lu
and
Winson S. Ho
*
Department of Neurological Surgery, University of California, San Francisco, CA 94143, USA
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(2), 308; https://doi.org/10.3390/cancers16020308
Submission received: 19 November 2023 / Revised: 21 December 2023 / Accepted: 6 January 2024 / Published: 11 January 2024
(This article belongs to the Special Issue Signaling Pathways in Gliomas)

Abstract

:

Simple Summary

The environment surrounding primary glioma brain tumors plays a critical role in tumor development and progression. Immune cells are important constituents of this tumor microenvironment. Classically, immune activation is thought to halt tumor progression; however, recent evidence has shown that glioma cells can interact with immune cells to convert them into tumor-supporting cells. These reprogrammed immune cells allow glioma cells to evade the anti-tumor effects of the immune system and can even promote tumor growth. Understanding the mechanisms that result in this altered immune tumor microenvironment may pave the way to new therapies for brain tumors. This review aims to provide an updated view of the mechanisms underlying glioma–immune cell interactions and novel therapies being developed to combat glioma immune evasion.

Abstract

Glioma progression is a complex process controlled by molecular factors that coordinate the crosstalk between tumor cells and components of the tumor microenvironment (TME). Among these, immune cells play a critical role in cancer survival and progression. The complex interplay between cancer cells and the immune TME influences the outcome of immunotherapy and other anti-cancer therapies. Here, we present an updated view of the pro- and anti-tumor activities of the main myeloid and lymphocyte cell populations in the glioma TME. We review the underlying mechanisms involved in crosstalk between cancer cells and immune cells that enable gliomas to evade the immune system and co-opt these cells for tumor growth. Lastly, we discuss the current and experimental therapeutic options being developed to revert the immunosuppressive activity of the glioma TME. Knowledge of the complex interplay that elapses between tumor and immune cells may help develop new combination treatments able to overcome tumor immune evasion mechanisms and enhance response to immunotherapies.

1. Introduction

Glioma development and progression occur in concert with continuous alterations in the surrounding tumor microenvironment (TME). The TME is a heterogeneous milieu of cells consisting of cancer cells and nonmalignant cells, including stromal cells, fibroblasts, endothelial cells, and immune cells integrated in a complex extracellular matrix [1]. Recent studies have focused on investigating the role of the TME in tumor development and how these cells interact with each other and with the malignant cellular component [1,2,3]. There is increasing evidence that glioma cells can functionally sculpt the TME through secretion of cytokines, chemokines, and other factors that ultimately reprogram components of the TME to promote tumor survival and progression [1,4,5,6].
Immune cells are important constituents of the TME and play key roles in surveillance against the development and advancement of glioma. However, growing evidence has shown that glioma cells can also co-opt innate immune cells (macrophages, neutrophils, dendritic cells (DCs), myeloid-derived suppressor cells, and natural killer (NK) cells) as well as adaptive immune cells (T and B cells) to promote tumor progression [1,6,7,8]. A plethora of mechanisms regulated by tumor-associated cells enable the immune escape of glioma cells and potentiate an immune-suppressive TME. Understanding these altered crosstalk interactions between glioma and immune cells may help develop new treatments to overcome tumor immune evasion.
In this review, we summarize the current knowledge on glioma crosstalk with various immune cells in the TME that contribute to the reprogramming of the TME and promote tumor progression and clinical strategies targeting these interactions.

2. Myeloid Cells

2.1. Tumor-Associated Macrophages and Microglia (TAMs)

Macrophages in the brain consist of both bone marrow-derived macrophages (BMDMs) and resident macrophages of the central nervous system (CNS), termed microglia. Studies have revealed that microglia originate from yolk sac progenitors during embryogenesis and have distinct transcriptional states depending on their topological distribution in the human brain [9,10,11]. While the exact differences in microglia and BMDMs remain to be elucidated, these macrophage types modulate immune responses through pathogen phagocytosis and antigen presentation and also function in wound healing and tissue repair. Simplistically, macrophages are dichotomized into M1 pro-inflammatory (i.e., anti-tumor) and M2 alternatively activated, anti-inflammatory (i.e., pro-tumor) phenotypes. However, more recent evidence has suggested that macrophages display a complex profile of both M1 and M2 polarization markers and likely exist in a continuum rather than in two distinct states [12]. In gliomas, tumor-associated macrophages (TAMs) are the most abundant population of TME immune cells, making up to 50% of total live cells in gliomas [13]. TAMs release a wide array of growth factors and cytokines in response to factors produced by cancer cells, playing a critical tumor-supportive role in the TME. This direct, pro-tumorigenic effect of TAMs is illustrated in mouse models of glioma where perturbation of microglia/macrophage function attenuates glioma proliferation [14,15,16]. Moreover, high macrophage infiltration in gliomas correlates with a negative prognosis [17], similar to other tumor types [18], further establishing their role in cancer progression. Recent studies have begun to try to distinguish between microglia and BMDMs in glioma formation and progression. Lineage-tracing experiments in mouse glioblastoma models have suggested that 85% of TAMs are BMDMs, which are predominantly localized in the perivascular areas of the tumor, while microglia are peri-tumoral [19]; however, other studies have suggested that the majority of TAMs are intrinsic microglia [20]. Interestingly, studies comparing the type of TAM (i.e., microglia vs. BMDM) in IDH-mutant and wild-type tumors have suggested that IDH-mutant TAMs were composed primarily of microglia, while IDH-wild-type tumors had more BMDM infiltration [12,21]. In astrocytomas, increased infiltration of TAMs derived from BMDM was associated with higher tumor grade and reduced survival [22]. More in-depth studies are needed to accurately distinguish microglia from BMDMs and further elucidate their distinct roles within specific types of gliomas. Given the current lack of clear evidence on the differences in roles between microglia and BMDMs in glioma, our connotation of TAMs in this review denotes macrophages without distinction between these two myeloid cell populations.

2.1.1. Glioma Effect on TAMs

Gliomas help recruit TAMs to the TME by secreting various chemokines, including monocyte chemoattractant proteins (MCP-1/CCL2 and MCP-3) [23], C-X-C motif chemokine 12 (CXCL12) [24], colony-stimulating factor 1 (CSF-1) [25], lysyl oxidase (LOX) [26], and glial cell-derived neurotrophic factor (GDNF) [27] (Figure 1). Genomic alterations in gliomas can lead to upregulation of these chemokines and increased recruitment of TAMs. For example, neurofibromatosis-1 (NF1) deficiency in gliomas is associated with increased infiltration of TAMs [28]. Chen et al. also demonstrated that genomic alterations in the PTEN-phosphatidylinositol 3-kinase (PI3K) pathway, commonly mutated in glioblastoma (GBM) cells, lead to enhanced recruitment of TAMs. Mechanistically, this effect was mediated through the activation of Yes-associated protein 1 (YAP1), which transcriptionally upregulates LOX expression and secretion from tumor cells. Subsequently, LOX functions as a potent macrophage chemoattractant via activation of the β1 integrin-proline-rich tyrosine kinase 2 pathway (PYK2) in macrophages [26]. Additionally, amplification of epidermal growth factor receptor (EGFR) and its truncation mutant variant (EGFRvIII) induces MCP-1 expression and secretion, promoting TAM recruitment [29].
Gliomas can also reprogram TAMs in the TME toward an immunosuppressive, tumor-promoting phenotype. Glioma-derived CSF-1 not only acts as chemoattractant but also induces a shift of microglia and macrophages toward a pro-tumor phenotype, and blocking CSF-1 receptor (CSF-1R) inhibits alternative activation of TAMs and attenuates glioma progression [30]. However, gliomas can acquire resistance to CSF-1R inhibition mediated by changes in the TME that lead to increased secretion of interleukin (IL)-4, a potent inducer of pro-tumor, alternatively activated TAMs, from CD8+ T cells and other cell types [31]. Upregulation of IL-4/IL-4R signaling in TAMs results in their increased secretion of insulin-like growth factor (IGF-1), which, upon binding to IGF-1 receptor (IGF-1R) on tumor cells, upregulates the PI3K pathway, promoting tumor growth and malignancy [31]. Accordingly, halting this feedforward loop of TAM recruitment and alternative activation through a combination of CSF-1R and IGF-1R inhibition leads to a further reduction in tumor progression in mice [31]. In addition to CSF1 and IL-4-mediated TAM reprogramming, the chemokine CXCL16 released from tumor cells has also been shown to bind to CXCR6 on microglia cells and drive them toward a pro-tumor phenotype [32]. Another feedforward mechanism involves glioma-derived MCP-1/CCL2, which triggers the release of IL-6 from microglia, a ligand for signal and activator of transcription 3 (STAT3) pathway in gliomas [33]. STAT3 induces a variety of transcriptional factors that propagate tumorigenesis, including maintaining the cancer stem cell state and invasiveness and upregulation of immunosuppressive factors. In glioma stem cells (GSCs), STAT3 activation leads to increased secretion of macrophage inhibitory cytokine-1 (MIC-1), IL-10, IL-4, and transforming growth factor β (TGF-β), which inhibit the phagocytic activity of macrophages and induce them to become immunosuppressive [34,35,36,37]. In addition to reprogramming of TAMs, glioma cells can upregulate the antiphagocytic “don’t eat me” surface protein CD47, which binds to its receptor signal-regulatory protein alpha (SIRP⍺) on TAMs and inhibits their phagocytic activity [38,39]. Blocking this myeloid checkpoint CD47-SIRP⍺ axis using anti-CD47 antibody therapy was shown to limit tumor growth in both adult and pediatric gliomas [40,41] and also promoted M1 activation of TAMs [42].

2.1.2. TAMs’ Effect on Glioma

A number of TAM-secreted factors, including stress-inducible protein 1 (STI1) [43], epidermal growth factor (EGF) [25], TGF-β [44,45], and matrix metallopeptidase-2 (MMP-2) [46], have been shown to enhance proliferation and invasiveness of glioma cells (Figure 1). The release of cytokines, including IL-12, IL-1B, CCL8, and IL-6 by TAMs, has also been shown to promote glioma stem cell renewal [13,47,48]. In addition, Shi et al. identified pleiotrophin as a TAM-secreted factor that activates AKT signaling in glioma cells and promotes tumor growth and stem cell renewal [49]. Glioma-induced polarization of macrophages also promotes macrophage secretion of osteopontin (SPP1) [26] and the expression of vascular endothelial growth factor (VEGF) in a receptor for advanced glycation endproduct (RAGE)-dependent manner [50], both of which promote glioma cell survival and angiogenesis.
TAMs also play important roles in creating an immunosuppressive environment for glioma immune evasion. TAM expression of chemokines such as CCL2, 5, 20, and 22 increase recruitment of regulatory T cells (Tregs), which subsequently inhibit the activity of effector T cells, natural killer (NK) cells, and antigen-presenting cells (APCs) [51,52,53]. Microglia also express and secrete Fas-ligand (FasL), which binds to Fas receptors and induces the apoptosis of invading T cells in gliomas [54,55]. Additionally, recently described overexpression of indoleamine 2,3-dioxgynase 1 and tryptophan 2,3-dioxygenase 2 (IDO1/TDO2) in GBM promotes immunosuppression through the production of tryptophan metabolite L-Kynurenine (Kyn) [56]. Kyn can interact with aryl hydrocarbon receptor (AHR) on TAMs and other immune cells to repress pro-inflammatory cytokine release by TAMs and drive CD39 expression in TAMs, which contributes to T cell dysfunction while promoting Treg generation in GBM [56,57].
Recent spatial transcriptomic analysis has also highlighted distinct regions within glioma tumors that are enriched in myeloid and lymphoid cells and display mesenchymal-like transcriptional signatures [58], suggesting that these immune cells also genetically alter tumor cells. Indeed, evidence from Hara et al. has shown that macrophage-derived oncostatin can interact with receptors on glioblastoma cells and induce a mesenchymal-like transcriptional program through STAT3 activation [59]. Additionally, studies by Gangoso et al. have demonstrated that macrophage infiltration and interferon (IFN)-γ signaling induce a transcriptional and epigenetic reconfiguration in GSCs toward a mesenchymal-like profile, which leads to further recruitment of TAMs and immune evasion [60]. Correspondingly, this acquired mesenchymal-like state has been shown to be associated with poor prognosis and tumor recurrence [61]. Taken together, these studies highlight the bidirectional cellular interdependence between glioma cells and TAMs that help establish an immunosuppressive TME and promote tumorigenesis.

2.2. Neutrophils

Neutrophils account for up to 70% of circulating leukocytes and are classically the first responders to acute inflammation [62]. Upon tissue damage or infection, epithelial cells secrete neutrophil-homing chemokines, inducing them to extravasate from circulation and enter the damaged tissue where they perform various functions, including phagocytosis of invading microorganisms, secretion of inflammatory cytokines (e.g., tumor necrosis factor alpha (TNF⍺), IL-4, IL-8), and formation of neutrophil extracellular traps (NETs) [62]. Current cancer literature suggests an important role for neutrophils within the TME [63]. Studies on various cancer types have suggested that neutrophil expansion, both in the TME and systemically, is associated with poor prognosis [63]. In tumors, neutrophils are dichotomized into N1 (tumor-suppressive or high-density neutrophils (HDNs)) and N2 (tumor-promoting or low-density neutrophils (LDNs)) phenotypes. In many cancer types, LDNs, which exhibit a more immature phenotype, predominate in the circulation and may contribute to cancer progression and metastasis [64].
In the context of glioma, little mechanistic work has been performed on neutrophils or other granulocytes until recently. Early work showing that higher-grade gliomas have higher circulating and tumor-infiltrating neutrophil levels has suggested that neutrophils play an important role in glioma TME, and glioma-derived factors may affect these cell types [65]. Further assessment of the relationship between neutrophils and prognosis has demonstrated that an increased proportion of neutrophils to lymphocytes in peripheral blood is associated with poor prognosis in glioma patients [66]. Moreover, neutrophil-specific DNA methylation associated with dexamethasone treatment (NDMI score) was recently shown to be an accurate marker for patient survival, inversely correlating with prognosis and positively correlating with an immunosuppressive TME [67]. Gliomas can induce peripheral neutrophilia through the secretion of granulocyte-colony stimulating factor (G-CSF) and granulocyte-macrophage colony-stimulating factor (GM-CSF) [68]. Recent work by Lad et al. has also demonstrated that GM-CSF in the glioblastoma secretome enhances the longevity of peripheral blood neutrophils [69]. In addition to enhancing granulopoiesis, gliomas can recruit neutrophils to the TME through expression of high levels of IL-8, under stimulation of IL-1, TNF⍺, and high-mobility group box 1 (HMGB1) derived from NETs [70,71]. Additionally, in in vitro co-cultures, glioma-derived IL-6 and IL-8 were shown to extend neutrophil survival, suggesting glioma–neutrophil interactions [72]. Further studies have identified RAGE expressed on glioma tissues as a binding receptor for neutrophils, which subsequently activates NF-κB signaling and further promotes neutrophil infiltration [71].
With respect to the effect of neutrophils on gliomas, work by Liang et al. has shown neutrophils can promote the mesenchymal transformation of gliomas via induction of S100A4 expression within glioma cells [73]. In contrast, other reports have also shown that neutrophils recruited during the early stages of glioma development exert an anti-tumor effect in mice through increased production of reactive oxygen species (ROS) [74]. However, this anti-tumor effect was lost with tumor progression. Interestingly, Lad et al. recently showed that a subset of TANs are recruited from skull marrow and uniquely differentiate into APC-like cells that activate T-cells and suppress tumor growth [69]. A more detailed understanding of neutrophils and other granulocytes and signals that pivot neutrophils to become immunosuppressive may hold promise for a better understanding of the reprogramming system of the TME and new potential immunotherapy targets.

2.3. Dendritic Cells (DCs)

Dendritic cells (DCs) are specialized antigen-presenting cells (APCs) that play key roles in the initiation of anti-tumor immune responses by linking innate and adaptive immunity [75]. DCs are not typically found in normal brain parenchyma but are present in vascular-rich compartments such as the choroid plexus and meninges [76,77]. In response to a variety of stimuli, including exposure to pathogens, nucleic acids, and type I interferons, DCs undergo activation and maturation, during which they acquire potent T-cell stimulatory ability [78]. While the specific role of DCs in the setting of gliomas is still being elucidated, current studies suggest an interplay between DCs, tumor cells, and TME. Studies on primary brain tumors and metastatic brain tumors have shown that DCs can recognize and traffic tumor antigens to tumor-draining deep cervical lymph nodes to elicit T cell-mediated responses [79,80,81]. They also can produce chemokines such as CCL9 and CCL10 that can recruit cytotoxic T lymphocytes (CTLs) into the tumor [81,82]. Furthermore, upon stimulation by T cell-produced IFN-γ, DCs can subsequently boost the anti-tumor activity of T and NK cells through the production of cytokines such as IL-12 [83]. It has been noted that the lymphatic migration of DCs to and from tumors and to draining cervical lymph nodes may be a critical aspect underlying response to immunotherapies, such as immune checkpoint inhibitors. Song et al. demonstrated that enhancement of lymphatic vasculature in the brain through VEGF-C expression synergizes with immune checkpoint blockade to elicit a robust adaptive immune response and improve survival outcomes in mouse glioma models [84]. Moreover, studies from Hu et al. have shown that VEGF-C-augmentation of lymphatics results in increased DC migration from the CNS to cervical lymph nodes in a CCL21/CCR7-dependent manner, and this DC migration is critical for the survival benefit seen with combined VEGF-C and immune checkpoint blockade therapy [85]. These findings highlight the essential role the lymphatic vasculature and DCs play in mediating anti-tumor responses.
Studies in other tumor types have identified various molecules found in the TME and expressed by tumor cells that inhibit DC activation and drive DCs toward an immunosuppressive, regulatory phenotype, including vascular endothelial growth factor (VEGF), prostaglandin E2 (PGE2), IL-10, and CSF-1 [86,87,88]. These regulatory DCs can then promote Treg activation and downregulate the recruitment and activity of CTLs by secreting cytokines such as IL-10 and TGF-β [89,90]. Nrf2, a redox-sensitive transcription factor expressed in DCs and whose overexpression positively correlates with glioma progression [91], was recently shown to suppress DC maturation and consequently result in decreased T cell activation [92]. Moreover, recent work on DC differentiation in IDH wild-type and mutant gliomas has also shown that wild-type tumors exhibit higher infiltration of DCs, while production of R-2-hydroxyglutarate (R-2-HG) in IDH-mutant gliomas impairs dendritic cell differentiation and maturation and suppresses major histocompatibility complex (MHC I/II) presentation, resulting in reduced T-cell activation [93]. Further investigation is needed into the specific roles of DC in gliomas and the complex interplay between DCs, tumor cells, and other immune cells.

2.4. Myeloid-Derived Suppressor Cells (MDSCs)

Myeloid-derived suppressor cells (MDSCs) are a heterogeneous population of myeloid-origin TME cells comprising progenitor and immature macrophages, granulocytes, and DCs. In patients with gliomas, peripheral blood MDSCs are elevated compared to non-tumor samples [94]. Additionally, the intratumoral density of MDSCs increases with glioma progression and correlates with patient survival [95,96]. MDSCs have been shown to be powerful inhibitors of anti-tumor immune responses through various mechanisms discussed below.
A variety of inflammatory mediators can induce MDSC expansion and recruitment to the TME, including tumor-derived IL-6, IL-8, IL-10, CSF-1, CCL2, CXCL2, PGE2, and TGF-β [97]. Gliomas can also increase intracellular levels of S100A8/9 (calprotectin) in MDSCs, an important marker of MDSC development and activation [98]. In addition, hypoxic conditions in the TME can alter the metabolism of MDSCs toward fatty acid oxidation, prompting upregulation of arginase I (Arg1) and nitric oxidase in MDSCs, which aid in their ability to suppress the immune system [99].
MDSCs present in the TME contribute to immunosuppression through suppression of T cells, NK cells, macrophages, and DCs functions and induction of Tregs [100,101,102]. While the exact roles MDSCs play in gliomas have yet to be fully elucidated, work in other tumor types has shown that MDSCs can induce oxidative stress by producing nitric oxide (NO), which can have several downstream effects. NO can upregulate Notch signaling and IL-6 signaling in tumor cells, leading to prolonged STAT3 activation and promoting cancer cell stemness [103]. In NK cells, MDSC-released NO reduces the cytotoxicity of NK cells and inhibits their ability to secrete IFN-γ and TNF⍺ [104]. Oxidative stress also potently blocks T-cell proliferation and activation by nitration/nitrosylation of chemokines and T-cell receptors [105]. TGF-β release by MDSCs has also been shown to induce Treg differentiation, NK cell anergy, and M2 macrophage polarization [106,107]. MDSCs are known to deplete essential metabolites in the TME, including arginine depletion promoted by Arg1 activity, which alters T cell receptor formation and induces T cell proliferation arrest [108,109]. Further work is needed to investigate the plethora of mechanisms through which MDSCs modulate the immune system in gliomas.

3. Lymphocytes

T cells, including CD8+ cytotoxic T lymphocytes (CTLs), regulatory CD4+ (Treg), and conventional CD4+ T cells, constitute 1–5% of the total glioma cellular content [110,111]. CTLs are considered the most potent anti-tumor cells, and many emerging immunotherapies have focused on enhancing CTL activity [112,113]. Upon priming and activation by antigens bound to major histocompatibility locus (MHC) I proteins expressed on APCs, CD8+ T cells differentiate into CTLs. CTLs can then release perforin- and granzyme-containing granules, resulting in an efficient anti-tumoral attack with direct destruction of target cells. CD4+ T helper 1 (Th-1) cells also mediate an anti-tumor response through the secretion of proinflammatory cytokines such as IL-2, TNF⍺, and IFN-γ, which promotes not only T-cell priming and CTL cytotoxicity but also the anti-tumoral activity of macrophages and NK cells [114,115]. Although debated in gliomas, the presence of tumor-infiltrating CD8+ T cells and Th-1 cytokines in tumors correlates with a favorable prognosis in terms of overall survival and disease-free survival in many malignancies [116].
A hallmark of tumor cells is the evasion of this lymphocyte-induced immune attack using two main strategies: avoiding immune recognition and instigating an immunosuppressive TME, including impairing the effector functions of anti-tumor CD4+ and CD8+ T cells and promoting immunosuppressive Treg and Th17 cells [117]. Below, we discuss the tumor–lymphocyte interactions that are involved in glioma immune evasion.

3.1. T cells

3.1.1. Reprogramming of T Cells

Gliomas can reprogram T cells into a dysfunctional state through multiple mechanisms (Figure 2). First, gliomas and tumor-associated cells, including lymphocytes, can express immune checkpoint molecules such as programmed death receptor Ligand-1 (PD-L1) [118,119,120], which binds to Programmed Death Receptor-1 (PD-1) on T-cells, blocking T-cell differentiation and activation. Notably, the expression of PD-L1 in gliomas is heterogeneous and relatively low, and evidence suggests that it is primarily expressed in lymphocytes [120] and myeloid cells in gliomas [121]. PD-L1 expression was noted to be lower in IDH-mutant, low-grade gliomas (LGG) [119], which also had reduced expression of T-cell lymphocyte-associated genes compared to IDH-wild-type, high-grade gliomas (HGG) [122]. Further investigation is needed to elucidate the mechanisms contributing to these differences in LGG versus HGG checkpoint molecule expression. Second, the glioma TME can recruit and promote Tregs to secrete immunosuppressive cytokines, as discussed below, which further suppress effector T cell function. Third, various factors secreted in the TME, such as NO by MDSCs and IL-10 and TGF-β from Tregs, induce T cell tolerance, whereby T cells do not respond to antigen stimulation due to the presence of immunosuppressive cytokines or signals [105,123]. Ultimately, T cell exhaustion may occur from persistent antigen stimulation of naïve T cells, which is characterized by reduced expression of effector molecules with increased expression of checkpoint molecules, including PD-1 and cytotoxic T-lymphocyte-associated protein 4 (CTLA-4) [124,125].

3.1.2. Evasion of Lymphocyte Surveillance

Gliomas, like many other tumors [126], can evade immune surveillance by downregulation of MHC-I expression [127]. MHC-I downregulation reduces the number of tumor-infiltrating lymphocytes and correlates with a worse prognosis [126]. In general, tumors may reduce MHC-I antigen expression through genomic defects perturbing heavy a-chain (i.e., HLA-encoded) or light chain (i.e., B2-microglobulin, B2M) gene expression or transcriptional (e.g., downregulation of type I and II IFN pathways, loss of transcriptional regulators, DNA hypermethylation or histone deacetylases) and post-transcriptional (e.g., microRNAs) silencing [128]. While the exact mechanisms underlying MHC-I downregulation in gliomas remain to be elucidated, a recent paper by Mondal et al. identified an intimate relationship between type I IFN signaling and MHC-I expression in glioma [129]. Specifically, they found that inhibition of tumor-expressed protein phosphatase-2A (PP2A), a serine-threonine phosphatase involved in DNA damage response and inhibition of which had previously been shown to enhance tumor immunity [130,131], leads to the accumulation of cytosolic double-stranded DNA and consequent induction of the cGAS-STING pathway. STING activation then induces phosphorylation of interferon regulatory factor 3 (IRF3) to promote type I IFN production and downstream JAK/STAT1 activity, leading to upregulation of MHC-I expression.

3.1.3. Recruitment of Tregs

Cancer cells and the TME simultaneously suppress anti-tumor T cell function through the recruitment of Tregs. Patients with malignant gliomas have an increased proportion of immunosuppressive Tregs and reduced CD4+ T cell number [132]. Glioma cells can recruit Tregs to the TME through the secretion of CCL22 [133,134] and other factors [135]. Indoleamine 2,3-dioxygnase 1 (IDO1) expression in gliomas has been suggested to be another important Treg recruiter, potentially via chemokine induction, and IDO upregulation is associated with a decrease in survival in glioma patients [136].
Tregs are known to exert immunosuppression through effects on numerous TME cell types [137] via various mechanisms, including consumption of IL-2, thereby suppressing T cell activation [138]; production of immunosuppressive cytokines such as IL-10, IL-35, and TGF-β [133,139,140]; conversion of ATP to adenosine, which hinders T cell activation [141]; CTLA-4-mediated suppression of APC function [142]; and secretion of granzyme or perforin to destroy effector cells [143]. While the induction of Treg activity is important for immune evasion by gliomas and correlates with glioma progression [144], the contribution of each of these mechanisms in gliomas has yet to be elucidated.

3.2. Natural Killer (NK) Cells

NK cells are circulatory, innate lymphoid cells recognized for their cytotoxic and cell-mediated killing functions. NK cells are important regulators of tumor surveillance, as illustrated by a correlation between low NK cell activity and increased cancer risk [145]. NKs employ death receptor-mediated apoptosis and perforin/granzyme-mediated cytotoxicity to target tumor cells and limit primary tumor growth [51].
Limited studies have examined the direct effect of glioma on NK function. Nonetheless, extrapolation of evidence from studies on other tumor types suggests that several factors in the glioma TME limit NK cell activity. TGF-β has been shown to negatively regulate NK cytotoxicity and proliferation [146]. TGF-β was also shown to downregulate the expression of NKG2D activating receptors on NK cells obtained from peripheral blood in GBM patients [147]. Additionally, PGE2 secreted by tumor cells can signal through a G-protein coupled receptor EP4 to inhibit IFN-γ production and anti-tumor activity of NK cells in mice [148]. IDO-production of Kyn by gliomas has also been shown to downregulate the expression of activating NKG2D and NKp46 receptors on NK cells [56,57,149]. Adenosine [150] and lactate [151] metabolites present in the TME have also been shown to reduce NK cytotoxicity and maturation.

4. Immunotherapies

4.1. Cytokine and Molecular Therapies

Several cytokines fused to a specific tumor-associated epitope or delivered via oncolytic viruses have been investigated. IL2, IL12, and TNF⍺ were studied by Weiss et al., which demonstrated, in mice, that the delivery of each of these cytokines independently promoted CD4+ and CD8+ infiltration and reduced tumor volume, with the administration of IL-12 showing the highest cure rate [152]. Administration of TNF⍺ to human GBM patients increased tumor necrosis and tumor-infiltrating CD4+ and CD8+ T cells. However, the use of TNF⍺ therapy is limited due to systemic toxicities associated with the high levels needed to achieve efficacy [153]. Other cytokine/chemokine inhibitors have also been trialed with limited efficacy, such as a CSF1R inhibitor, which showed no improvement in progression-free or overall survival in the PLX3397 trial [154]. CCR2 inhibition targeting MDSC infiltration in gliomas did not show efficacy alone, but in combination with PD-1 blockade, was able to extend survival in murine glioma models [155]. A recent study utilizing recombinant IL-7, a cytokine that is required for T cell development, in GBM mouse models showed IL-7 treatment could expand CD8+ T cells in peripheral blood and the TME and improve overall survival in mice [156]. Other promising avenues of regulating cytokine levels in the TME include modulating intracellular pathways involved in immune suppression. For example, studies have shown that PP2A inhibition or activation of the STING pathway increases type I IFN signaling in TAMs, leading to increased activation of macrophages and CD8+ T cells and a subsequent reduction in tumor volume in mice [129,157].

4.2. Immune Checkpoint Inhibitors

Despite the notable success immune checkpoint inhibitors have shown in many cancer types, the same efficacy has not been observed in gliomas/glioblastomas. Unfortunately, three phase III trials assessing nivolumab, an anti-PD-1 therapy, alone or in combination with temozolomide or radiotherapy, failed to demonstrate a survival advantage in patients with GBM [158,159,160]. In addition, in a phase I trial, anti-PD-1 therapy in combination with anti-CTLA-4 therapy did not improve overall survival [161]. Nonetheless, there are some patients that show durable radiographic responses and prolonged survival following PD-1 blockade [162], suggesting that certain factors, such as PD-L1/PD-1 expression and patient-specific glioma-TME interactions, influence the efficacy of these treatments [163,164]. Indeed, recent studies looking into molecular and genetic factors associated with immune checkpoint inhibitor sensitivity identified patients with defects in replication stress response, a pathway important for maintaining genome stability and integrity during DNA replication, were highly sensitive to immune checkpoint blockade [165]. Moreover, results from Cloughesy et al. and Schapler et al. demonstrating that neoadjuvant PD-1 blockade can enhance the anti-tumor immune response [162,166] and improve overall and progression-free survival in patients with recurrent GBM [162], suggest that the timing of immune checkpoint blockade may also be an important factor affecting efficacy. Considering that current standard therapies for glioma patients are also immunomodulatory—such as temozolomide, which induces lymphopenia, which was shown to enhance antitumor activity associated with immunotherapies [167], and steroids, which are known to be immunosuppressive—further work is needed to investigate the optimal timing and order of immunotherapies given in these combination settings to achieve synergistic immune responses. Combination therapies with other immune checkpoint inhibitors, such as anti-CD47, which has shown promising results in adult and pediatric gliomas [40,41], may also be of benefit. In addition, further characterization of the molecular and immune landscape of immune checkpoint inhibitor-sensitive gliomas may allow for a more tailored and efficacious approach for utilizing these therapies as well as yield new targets that may be adapted to increase sensitivity to immune checkpoint blockade.

4.3. CAR T-Cell Therapy

CAR T therapy has emerged as a promising form of immunotherapy for various cancers. CAR T therapy involves engineering a patient’s own T cells to express a receptor specifically targeting antigens on tumor cells. Several trials using CAR T therapy in GBM have been completed or are currently active [168]. Early studies in mouse glioma models showed encouraging results for EGFRvIII-targeting CAR T cells, which led to cures in all mice with a sustainable response [169]. Subsequent studies on human patients demonstrated that while this therapy was safe and initially led to an anti-tumor response, disease recurrence, target-antigen loss, and escape were observed [170,171]. More recent work has focused on refining these CAR T cells to target multiple antigens, such as synNotch-CAR T targeting EGFRvIII and a brain-specific glycoprotein [172], to improve the specificity and persistence of these cells. Other studies have examined HER2 [173] and IL-13Ra2 [174], both highly expressed in GBM tissues, targeting CAR T cells in glioblastoma with promising initial results. For diffuse midline gliomas, the identification of a targetable, glioma-specific H3.3K27 mutation epitope [175] and the newly developed CAR-T targeting GD2 [176], a disialoganglioside highly expressed in H3K27M-mutated gliomas, hold significant promise for the use CAR T-based immunotherapy for patients with these tumors. More studies are also underway to develop modified CAR T cells with enhanced cytotoxicity and more durable responses [177,178]. Additional CAR T therapies for GBM and their potential limitations have been recently and comprehensively reviewed by Luksik et al. [168].

4.4. Oncolytic Viruses and Vaccines

Oncolytic viruses targeting tumor cells not only result in oncolysis of cancer cells but also augment host immune responses and promote proinflammatory pathways in the TME. Several oncolytic viruses have been and are currently being tested in gliomas [179]. DNX-2401, an adenovirus designed to be tumor-selective, has shown promising results in a phase I clinical trial on recurrent malignant glioma, with 20% of patients surviving greater than 3 years post-treatment and 12% showing at least a 95% reduction in tumor volume [180]. DNX-2401 was also recently tested in 12 pediatric diffuse intrinsic pontine glioma patients and was able to elicit immune-mediated anti-glioma response [181]. Trials examining engineered herpes simplex virus type 1 (HSV-1) oncolytic viruses have also reported encouraging results. Tumor-targeting HSV-1 G207 was shown to induce a strong immune response, with an increased number of tumor-infiltrating lymphocytes and probable improved survival in pediatric patients with high-grade gliomas [182]. Most recently, seminal work by Ling et al. has also demonstrated an HSV-1 oncolytic virus, CAN-3110, improved survival in recurrent GBM patients with HSV-1 seropositivity and dramatically increased tumor-infiltrating lymphocytes [183].
Vaccine-based therapies can elicit an anti-tumor immune response by introducing T cells to immunogenic tumor-specific or overexpressed antigens. Rindopepimut is an EGFRvIII-targeting peptide vaccine that was studied in a phase III clinical trial for newly diagnosed GBM patients who had undergone maximal resection and radiation therapy with temozolomide [184]. The trial showed that Rindopepimut alone did not improve overall survival and suggested that combination therapy with other immunotherapies may be of benefit. SurVaxM is another vaccine targeting a surviving peptide, an anti-apoptotic protein highly expressed in GBM cells. SurVaxM therapy was associated with higher median overall survival (86.6 weeks), and an ongoing clinical trial is evaluating this treatment in combination with pembrolizumab for recurrent GBM [185]. Other vaccine trials, including dendritic cell vaccines where DCs are generated to present antigens targeting GBM cells (e.g., HER2, IL13Ra2), have shown mixed results and are discussed in a comprehensive review by Sener et al. [186]

5. Conclusions and Perspectives

Cancer is no longer simply regarded as a genetic disease that leads to the formation of tumorigenic cells. Instead, we have come to appreciate the “ecological” aspect of this disease, where tumor development is intimately linked to reciprocal interactions with non-cancerous cells in their surrounding microenvironment [187]. Spatiotemporal interactions between these cells lead to adaptations and evolution of cancer as it progresses from initiation to immune evasion and invasion. The studies discussed above illustrate how gliomas sculpt an immune suppressive microenvironment that not only enables immune evasion but also promotes tumor growth and invasiveness. The infiltration of multiple immune cell types, including TAMs and Tregs, and their effect on neighboring cells through direct interactions and secreted cytokines, highlights the multidirectional, cellular interdependence the heterogenous TME exhibits. This complexity and multitude of immunosuppressive mechanisms within the TME may explain the limited efficacy of current immunotherapies.
Patient selection remains an important aspect to consider in the realm of immunotherapies. In this regard, it is important to note that the reciprocal interactions between glioma cells and the TME are continuously evolving from tumor initiation to progression and recurrence [188,189]. While the specific mechanisms underlying the evolution of glioma tumors remain to be solved, spatial transcriptomic and proteomic evidence has shown that glioma cells are dynamically reorganized into distinct regions that display unique molecular profiles or lineage states (e.g., proneural, mesenchymal, and classical) and are constantly adapting to evolving stresses from the surrounding TME [58,190]. Thus, the characterization of gliomas in patients cannot be a singular event and requires multiple regions and time points to be analyzed. With regard to surgery, this implicates the need for multiple tissue samples and serial biopsies and resections when treating patients with recurrent gliomas to best characterize their tumor-specific TME alterations and adapt immunotherapies based on this. In addition, recent studies have begun to examine changes in the immune TME associated with standard-of-care GBM therapies temozolomide and radiation [191]. Understanding how these treatments affect the TME and how gliomas may acquire resistance to these therapies through immune cell adaptations may help develop more efficacious combination treatments.
While significant progress has been made in characterizing the components of the TME, the majority of current studies, as highlighted above, have focused on the TAM population of immune cells, given their abundance in the TME. Further in-depth mechanistic studies on the role of other immune cell types in glioma biology are needed. Many fundamental aspects of immune cell–glioma interactions remain incompletely understood, such as the molecular changes involved in macrophage and DC reprogramming and NK cell suppression, the temporal alterations occurring in T cells over the course of glioma progression that lead to T cell exhaustion, the relationship between glioma mutations and immune cell recruitment and function, and, perhaps most importantly, how we may utilize genomic and transcriptomic analyses as approaches for precision medicine to develop more tailored and effective immune-based therapies. Defining these molecular underpinnings of glioma-TME crosstalk may lead to a new “magic bullet” therapeutic approach that interrupts the glioma ecosystem.

Author Contributions

M.E., J.S.Y., I.M., R.O.L. and W.S.H. contributed to the conceptualization, design, writing, review, and editing of this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Matthew Larson Pediatric Brain Tumor Foundation (W.S.H.); the Meghan Rose Bradley Foundation (W.S.H.); a US National Institutes of Health (NIH) grant (No. R01NS131545) (W.S.H.); and a US National Institutes of Health (NIH) grant (No. R01NS126501) (R.O.L.).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Quail, D.F.; Joyce, J.A. The Microenvironmental Landscape of Brain Tumors. Cancer Cell 2017, 31, 326–341. [Google Scholar] [CrossRef]
  2. Jain, S.; Rick, J.W.; Joshi, R.S.; Beniwal, A.; Spatz, J.; Gill, S.; Chang, A.C.; Choudhary, N.; Nguyen, A.T.; Sudhir, S.; et al. Single-cell RNA sequencing and spatial transcriptomics reveal cancer-associated fibroblasts in glioblastoma with protumoral effects. J. Clin. Investig. 2023, 133, e147087. [Google Scholar] [CrossRef] [PubMed]
  3. Clavreul, A.; Etcheverry, A.; Chassevent, A.; Quillien, V.; Avril, T.; Jourdan, M.L.; Michalak, S.; Francois, P.; Carre, J.L.; Mosser, J.; et al. Isolation of a new cell population in the glioblastoma microenvironment. J. Neurooncol. 2012, 106, 493–504. [Google Scholar] [CrossRef] [PubMed]
  4. Sharma, P.; Aaroe, A.; Liang, J.; Puduvalli, V.K. Tumor microenvironment in glioblastoma: Current and emerging concepts. Neurooncol. Adv. 2023, 5, vdad009. [Google Scholar] [CrossRef] [PubMed]
  5. Klemm, F.; Maas, R.R.; Bowman, R.L.; Kornete, M.; Soukup, K.; Nassiri, S.; Brouland, J.P.; Iacobuzio-Donahue, C.A.; Brennan, C.; Tabar, V.; et al. Interrogation of the Microenvironmental Landscape in Brain Tumors Reveals Disease-Specific Alterations of Immune Cells. Cell 2020, 181, 1643–1660.e1617. [Google Scholar] [CrossRef]
  6. Friebel, E.; Kapolou, K.; Unger, S.; Nunez, N.G.; Utz, S.; Rushing, E.J.; Regli, L.; Weller, M.; Greter, M.; Tugues, S.; et al. Single-Cell Mapping of Human Brain Cancer Reveals Tumor-Specific Instruction of Tissue-Invading Leukocytes. Cell 2020, 181, 1626–1642.e1620. [Google Scholar] [CrossRef] [PubMed]
  7. Grabowski, M.M.; Sankey, E.W.; Ryan, K.J.; Chongsathidkiet, P.; Lorrey, S.J.; Wilkinson, D.S.; Fecci, P.E. Immune suppression in gliomas. J. Neurooncol. 2021, 151, 3–12. [Google Scholar] [CrossRef] [PubMed]
  8. Menna, G.; Mattogno, P.P.; Donzelli, C.M.; Lisi, L.; Olivi, A.; Della Pepa, G.M. Glioma-Associated Microglia Characterization in the Glioblastoma Microenvironment through a ‘Seed-and Soil’ Approach: A Systematic Review. Brain Sci. 2022, 12, 718. [Google Scholar] [CrossRef]
  9. Hoeffel, G.; Chen, J.; Lavin, Y.; Low, D.; Almeida, F.F.; See, P.; Beaudin, A.E.; Lum, J.; Low, I.; Forsberg, E.C.; et al. C-Myb(+) erythro-myeloid progenitor-derived fetal monocytes give rise to adult tissue-resident macrophages. Immunity 2015, 42, 665–678. [Google Scholar] [CrossRef]
  10. Bottcher, C.; Schlickeiser, S.; Sneeboer, M.A.M.; Kunkel, D.; Knop, A.; Paza, E.; Fidzinski, P.; Kraus, L.; Snijders, G.J.L.; Kahn, R.S.; et al. Human microglia regional heterogeneity and phenotypes determined by multiplexed single-cell mass cytometry. Nat. Neurosci. 2019, 22, 78–90. [Google Scholar] [CrossRef]
  11. Masuda, T.; Sankowski, R.; Staszewski, O.; Bottcher, C.; Amann, L.; Sagar; Scheiwe, C.; Nessler, S.; Kunz, P.; van Loo, G.; et al. Spatial and temporal heterogeneity of mouse and human microglia at single-cell resolution. Nature 2019, 566, 388–392. [Google Scholar] [CrossRef]
  12. Muller, S.; Kohanbash, G.; Liu, S.J.; Alvarado, B.; Carrera, D.; Bhaduri, A.; Watchmaker, P.B.; Yagnik, G.; Di Lullo, E.; Malatesta, M.; et al. Single-cell profiling of human gliomas reveals macrophage ontogeny as a basis for regional differences in macrophage activation in the tumor microenvironment. Genome Biol. 2017, 18, 234. [Google Scholar] [CrossRef]
  13. Xuan, W.; Lesniak, M.S.; James, C.D.; Heimberger, A.B.; Chen, P. Context-Dependent Glioblastoma-Macrophage/Microglia Symbiosis and Associated Mechanisms. Trends Immunol. 2021, 42, 280–292. [Google Scholar] [CrossRef] [PubMed]
  14. Daginakatte, G.C.; Gutmann, D.H. Neurofibromatosis-1 (Nf1) heterozygous brain microglia elaborate paracrine factors that promote Nf1-deficient astrocyte and glioma growth. Hum. Mol. Genet. 2007, 16, 1098–1112. [Google Scholar] [CrossRef] [PubMed]
  15. Pong, W.W.; Higer, S.B.; Gianino, S.M.; Emnett, R.J.; Gutmann, D.H. Reduced microglial CX3CR1 expression delays neurofibromatosis-1 glioma formation. Ann. Neurol. 2013, 73, 303–308. [Google Scholar] [CrossRef]
  16. Bettinger, I.; Thanos, S.; Paulus, W. Microglia promote glioma migration. Acta Neuropathol. 2002, 103, 351–355. [Google Scholar] [CrossRef] [PubMed]
  17. Roggendorf, W.; Strupp, S.; Paulus, W. Distribution and characterization of microglia/macrophages in human brain tumors. Acta Neuropathol. 1996, 92, 288–293. [Google Scholar] [CrossRef] [PubMed]
  18. Bingle, L.; Brown, N.J.; Lewis, C.E. The role of tumour-associated macrophages in tumour progression: Implications for new anticancer therapies. J. Pathol. 2002, 196, 254–265. [Google Scholar] [CrossRef]
  19. Chen, Z.; Feng, X.; Herting, C.J.; Garcia, V.A.; Nie, K.; Pong, W.W.; Rasmussen, R.; Dwivedi, B.; Seby, S.; Wolf, S.A.; et al. Cellular and Molecular Identity of Tumor-Associated Macrophages in Glioblastoma. Cancer Res. 2017, 77, 2266–2278. [Google Scholar] [CrossRef]
  20. Muller, A.; Brandenburg, S.; Turkowski, K.; Muller, S.; Vajkoczy, P. Resident microglia, and not peripheral macrophages, are the main source of brain tumor mononuclear cells. Int. J. Cancer 2015, 137, 278–288. [Google Scholar] [CrossRef] [PubMed]
  21. Haddad, A.F.; Young, J.S.; Oh, J.Y.; Okada, H.; Aghi, M.K. The immunology of low-grade gliomas. Neurosurg. Focus 2022, 52, E2. [Google Scholar] [CrossRef] [PubMed]
  22. Venteicher, A.S.; Tirosh, I.; Hebert, C.; Yizhak, K.; Neftel, C.; Filbin, M.G.; Hovestadt, V.; Escalante, L.E.; Shaw, M.L.; Rodman, C.; et al. Decoupling genetics, lineages, and microenvironment in IDH-mutant gliomas by single-cell RNA-seq. Science 2017, 355, eaai8478. [Google Scholar] [CrossRef] [PubMed]
  23. Platten, M.; Kretz, A.; Naumann, U.; Aulwurm, S.; Egashira, K.; Isenmann, S.; Weller, M. Monocyte chemoattractant protein-1 increases microglial infiltration and aggressiveness of gliomas. Ann. Neurol. 2003, 54, 388–392. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, S.C.; Hong, J.H.; Hsueh, C.; Chiang, C.S. Tumor-secreted SDF-1 promotes glioma invasiveness and TAM tropism toward hypoxia in a murine astrocytoma model. Lab. Investig. 2012, 92, 151–162. [Google Scholar] [CrossRef]
  25. Coniglio, S.J.; Eugenin, E.; Dobrenis, K.; Stanley, E.R.; West, B.L.; Symons, M.H.; Segall, J.E. Microglial stimulation of glioblastoma invasion involves epidermal growth factor receptor (EGFR) and colony stimulating factor 1 receptor (CSF-1R) signaling. Mol. Med. 2012, 18, 519–527. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, P.; Zhao, D.; Li, J.; Liang, X.; Li, J.; Chang, A.; Henry, V.K.; Lan, Z.; Spring, D.J.; Rao, G.; et al. Symbiotic Macrophage-Glioma Cell Interactions Reveal Synthetic Lethality in PTEN-Null Glioma. Cancer Cell 2019, 35, 868–884.e866. [Google Scholar] [CrossRef]
  27. Ku, M.C.; Wolf, S.A.; Respondek, D.; Matyash, V.; Pohlmann, A.; Waiczies, S.; Waiczies, H.; Niendorf, T.; Synowitz, M.; Glass, R.; et al. GDNF mediates glioblastoma-induced microglia attraction but not astrogliosis. Acta Neuropathol. 2013, 125, 609–620. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Q.; Hu, B.; Hu, X.; Kim, H.; Squatrito, M.; Scarpace, L.; deCarvalho, A.C.; Lyu, S.; Li, P.; Li, Y.; et al. Tumor Evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with Immunological Changes in the Microenvironment. Cancer Cell 2017, 32, 42–56.e46. [Google Scholar] [CrossRef]
  29. An, Z.; Knobbe-Thomsen, C.B.; Wan, X.; Fan, Q.W.; Reifenberger, G.; Weiss, W.A. EGFR Cooperates with EGFRvIII to Recruit Macrophages in Glioblastoma. Cancer Res. 2018, 78, 6785–6794. [Google Scholar] [CrossRef]
  30. Pyonteck, S.M.; Akkari, L.; Schuhmacher, A.J.; Bowman, R.L.; Sevenich, L.; Quail, D.F.; Olson, O.C.; Quick, M.L.; Huse, J.T.; Teijeiro, V.; et al. CSF-1R inhibition alters macrophage polarization and blocks glioma progression. Nat. Med. 2013, 19, 1264–1272. [Google Scholar] [CrossRef]
  31. Quail, D.F.; Bowman, R.L.; Akkari, L.; Quick, M.L.; Schuhmacher, A.J.; Huse, J.T.; Holland, E.C.; Sutton, J.C.; Joyce, J.A. The tumor microenvironment underlies acquired resistance to CSF-1R inhibition in gliomas. Science 2016, 352, aad3018. [Google Scholar] [CrossRef] [PubMed]
  32. Lepore, F.; D’Alessandro, G.; Antonangeli, F.; Santoro, A.; Esposito, V.; Limatola, C.; Trettel, F. CXCL16/CXCR6 Axis Drives Microglia/Macrophages Phenotype in Physiological Conditions and Plays a Crucial Role in Glioma. Front. Immunol. 2018, 9, 2750. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, J.; Sarkar, S.; Cua, R.; Zhou, Y.; Hader, W.; Yong, V.W. A dialog between glioma and microglia that promotes tumor invasiveness through the CCL2/CCR2/interleukin-6 axis. Carcinogenesis 2012, 33, 312–319. [Google Scholar] [CrossRef] [PubMed]
  34. 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 Oncol. 2010, 12, 1113–1125. [Google Scholar] [CrossRef] [PubMed]
  35. Komohara, Y.; Horlad, H.; Ohnishi, K.; Fujiwara, Y.; Bai, B.; Nakagawa, T.; Suzu, S.; Nakamura, H.; Kuratsu, J.; Takeya, M. Importance of direct macrophage-tumor cell interaction on progression of human glioma. Cancer Sci. 2012, 103, 2165–2172. [Google Scholar] [CrossRef]
  36. Nusblat, L.M.; Carroll, M.J.; Roth, C.M. Crosstalk between M2 macrophages and glioma stem cells. Cell Oncol. 2017, 40, 471–482. [Google Scholar] [CrossRef]
  37. Ye, X.Z.; Xu, S.L.; Xin, Y.H.; Yu, S.C.; Ping, Y.F.; Chen, L.; Xiao, H.L.; Wang, B.; Yi, L.; Wang, Q.L.; et al. Tumor-associated microglia/macrophages enhance the invasion of glioma stem-like cells via TGF-beta1 signaling pathway. J. Immunol. 2012, 189, 444–453. [Google Scholar] [CrossRef]
  38. Hutter, G.; Theruvath, J.; Graef, C.M.; Zhang, M.; Schoen, M.K.; Manz, E.M.; Bennett, M.L.; Olson, A.; Azad, T.D.; Sinha, R.; et al. Microglia are effector cells of CD47-SIRPalpha antiphagocytic axis disruption against glioblastoma. Proc. Natl. Acad. Sci. USA 2019, 116, 997–1006. [Google Scholar] [CrossRef]
  39. Jaiswal, S.; Jamieson, C.H.; Pang, W.W.; Park, C.Y.; Chao, M.P.; Majeti, R.; Traver, D.; van Rooijen, N.; Weissman, I.L. CD47 is upregulated on circulating hematopoietic stem cells and leukemia cells to avoid phagocytosis. Cell 2009, 138, 271–285. [Google Scholar] [CrossRef]
  40. Willingham, S.B.; Volkmer, J.P.; Gentles, A.J.; Sahoo, D.; Dalerba, P.; Mitra, S.S.; Wang, J.; Contreras-Trujillo, H.; Martin, R.; Cohen, J.D.; et al. The CD47-signal regulatory protein alpha (SIRPa) interaction is a therapeutic target for human solid tumors. Proc. Natl. Acad. Sci. USA 2012, 109, 6662–6667. [Google Scholar] [CrossRef]
  41. Gholamin, S.; Mitra, S.S.; Feroze, A.H.; Liu, J.; Kahn, S.A.; Zhang, M.; Esparza, R.; Richard, C.; Ramaswamy, V.; Remke, M.; et al. Disrupting the CD47-SIRPalpha anti-phagocytic axis by a humanized anti-CD47 antibody is an efficacious treatment for malignant pediatric brain tumors. Sci. Transl. Med. 2017, 9. [Google Scholar] [CrossRef]
  42. Zhang, M.; Hutter, G.; Kahn, S.A.; Azad, T.D.; Gholamin, S.; Xu, C.Y.; Liu, J.; Achrol, A.S.; Richard, C.; Sommerkamp, P.; et al. Anti-CD47 Treatment Stimulates Phagocytosis of Glioblastoma by M1 and M2 Polarized Macrophages and Promotes M1 Polarized Macrophages In Vivo. PLoS ONE 2016, 11, e0153550. [Google Scholar] [CrossRef] [PubMed]
  43. Carvalho da Fonseca, A.C.; Wang, H.; Fan, H.; Chen, X.; Zhang, I.; Zhang, L.; Lima, F.R.; Badie, B. Increased expression of stress inducible protein 1 in glioma-associated microglia/macrophages. J. Neuroimmunol. 2014, 274, 71–77. [Google Scholar] [CrossRef] [PubMed]
  44. Wesolowska, A.; Kwiatkowska, A.; Slomnicki, L.; Dembinski, M.; Master, A.; Sliwa, M.; Franciszkiewicz, K.; Chouaib, S.; Kaminska, B. Microglia-derived TGF-beta as an important regulator of glioblastoma invasion—An inhibition of TGF-beta-dependent effects by shRNA against human TGF-beta type II receptor. Oncogene 2008, 27, 918–930. [Google Scholar] [CrossRef]
  45. Angioni, R.; Sanchez-Rodriguez, R.; Viola, A.; Molon, B. TGF-beta in Cancer: Metabolic Driver of the Tolerogenic Crosstalk in the Tumor Microenvironment. Cancers 2021, 13, 401. [Google Scholar] [CrossRef]
  46. Markovic, D.S.; Glass, R.; Synowitz, M.; Rooijen, N.; Kettenmann, H. Microglia stimulate the invasiveness of glioma cells by increasing the activity of metalloprotease-2. J. Neuropathol. Exp. Neurol. 2005, 64, 754–762. [Google Scholar] [CrossRef] [PubMed]
  47. Kujawski, M.; Kortylewski, M.; Lee, H.; Herrmann, A.; Kay, H.; Yu, H. Stat3 mediates myeloid cell-dependent tumor angiogenesis in mice. J. Clin. Investig. 2008, 118, 3367–3377. [Google Scholar] [CrossRef]
  48. Zhang, X.; Chen, L.; Dang, W.Q.; Cao, M.F.; Xiao, J.F.; Lv, S.Q.; Jiang, W.J.; Yao, X.H.; Lu, H.M.; Miao, J.Y.; et al. CCL8 secreted by tumor-associated macrophages promotes invasion and stemness of glioblastoma cells via ERK1/2 signaling. Lab. Investig. 2020, 100, 619–629. [Google Scholar] [CrossRef]
  49. Shi, Y.; Ping, Y.F.; Zhou, W.; He, Z.C.; Chen, C.; Bian, B.S.; Zhang, L.; Chen, L.; Lan, X.; Zhang, X.C.; et al. Tumour-associated macrophages secrete pleiotrophin to promote PTPRZ1 signalling in glioblastoma stem cells for tumour growth. Nat. Commun. 2017, 8, 15080. [Google Scholar] [CrossRef]
  50. Chen, X.; Zhang, L.; Zhang, I.Y.; Liang, J.; Wang, H.; Ouyang, M.; Wu, S.; da Fonseca, A.C.C.; Weng, L.; Yamamoto, Y.; et al. RAGE expression in tumor-associated macrophages promotes angiogenesis in glioma. Cancer Res. 2014, 74, 7285–7297. [Google Scholar] [CrossRef]
  51. Ebner, F.; Brandt, C.; Thiele, P.; Richter, D.; Schliesser, U.; Siffrin, V.; Schueler, J.; Stubbe, T.; Ellinghaus, A.; Meisel, C.; et al. Microglial activation milieu controls regulatory T cell responses. J. Immunol. 2013, 191, 5594–5602. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, D.; Yang, L.; Yue, D.; Cao, L.; Li, L.; Wang, D.; Ping, Y.; Shen, Z.; Zheng, Y.; Wang, L.; et al. Macrophage-derived CCL22 promotes an immunosuppressive tumor microenvironment via IL-8 in malignant pleural effusion. Cancer Lett. 2019, 452, 244–253. [Google Scholar] [CrossRef]
  53. Korbecki, J.; Kojder, K.; Siminska, D.; Bohatyrewicz, R.; Gutowska, I.; Chlubek, D.; Baranowska-Bosiacka, I. CC Chemokines in a Tumor: A Review of Pro-Cancer and Anti-Cancer Properties of the Ligands of Receptors CCR1, CCR2, CCR3, and CCR4. Int. J. Mol. Sci. 2020, 21, 8412. [Google Scholar] [CrossRef] [PubMed]
  54. Badie, B.; Schartner, J.; Prabakaran, S.; Paul, J.; Vorpahl, J. Expression of Fas ligand by microglia: Possible role in glioma immune evasion. J. Neuroimmunol. 2001, 120, 19–24. [Google Scholar] [CrossRef]
  55. Ford, A.L.; Foulcher, E.; Lemckert, F.A.; Sedgwick, J.D. Microglia induce CD4 T lymphocyte final effector function and death. J. Exp. Med. 1996, 184, 1737–1745. [Google Scholar] [CrossRef] [PubMed]
  56. Takenaka, M.C.; Gabriely, G.; Rothhammer, V.; Mascanfroni, I.D.; Wheeler, M.A.; Chao, C.C.; Gutierrez-Vazquez, C.; Kenison, J.; Tjon, E.C.; Barroso, A.; et al. Control of tumor-associated macrophages and T cells in glioblastoma via AHR and CD39. Nat. Neurosci. 2019, 22, 729–740. [Google Scholar] [CrossRef] [PubMed]
  57. Campesato, L.F.; Budhu, S.; Tchaicha, J.; Weng, C.H.; Gigoux, M.; Cohen, I.J.; Redmond, D.; Mangarin, L.; Pourpe, S.; Liu, C.; et al. Blockade of the AHR restricts a Treg-macrophage suppressive axis induced by L-Kynurenine. Nat. Commun. 2020, 11, 4011. [Google Scholar] [CrossRef]
  58. Ravi, V.M.; Will, P.; Kueckelhaus, J.; Sun, N.; Joseph, K.; Salie, H.; Vollmer, L.; Kuliesiute, U.; von Ehr, J.; Benotmane, J.K.; et al. Spatially resolved multi-omics deciphers bidirectional tumor-host interdependence in glioblastoma. Cancer Cell 2022, 40, 639–655.e613. [Google Scholar] [CrossRef]
  59. Hara, T.; Chanoch-Myers, R.; Mathewson, N.D.; Myskiw, C.; Atta, L.; Bussema, L.; Eichhorn, S.W.; Greenwald, A.C.; Kinker, G.S.; Rodman, C.; et al. Interactions between cancer cells and immune cells drive transitions to mesenchymal-like states in glioblastoma. Cancer Cell 2021, 39, 779–792.e711. [Google Scholar] [CrossRef]
  60. Gangoso, E.; Southgate, B.; Bradley, L.; Rus, S.; Galvez-Cancino, F.; McGivern, N.; Guc, E.; Kapourani, C.A.; Byron, A.; Ferguson, K.M.; et al. Glioblastomas acquire myeloid-affiliated transcriptional programs via epigenetic immunoediting to elicit immune evasion. Cell 2021, 184, 2454–2470.e2426. [Google Scholar] [CrossRef]
  61. Phillips, H.S.; Kharbanda, S.; Chen, R.; Forrest, W.F.; Soriano, R.H.; Wu, T.D.; Misra, A.; Nigro, J.M.; Colman, H.; Soroceanu, L.; et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 2006, 9, 157–173. [Google Scholar] [CrossRef] [PubMed]
  62. Malech, H.L.; Deleo, F.R.; Quinn, M.T. The role of neutrophils in the immune system: An overview. Methods Mol. Biol. 2014, 1124, 3–10. [Google Scholar] [CrossRef] [PubMed]
  63. Coffelt, S.B.; Wellenstein, M.D.; de Visser, K.E. Neutrophils in cancer: Neutral no more. Nat. Rev. Cancer 2016, 16, 431–446. [Google Scholar] [CrossRef] [PubMed]
  64. Deniset, J.F.; Kubes, P. Recent advances in understanding neutrophils. F1000Research 2016, 5, 2912. [Google Scholar] [CrossRef]
  65. Fossati, G.; Ricevuti, G.; Edwards, S.W.; Walker, C.; Dalton, A.; Rossi, M.L. Neutrophil infiltration into human gliomas. Acta Neuropathol. 1999, 98, 349–354. [Google Scholar] [CrossRef] [PubMed]
  66. Wiencke, J.K.; Koestler, D.C.; Salas, L.A.; Wiemels, J.L.; Roy, R.P.; Hansen, H.M.; Rice, T.; McCoy, L.S.; Bracci, P.M.; Molinaro, A.M.; et al. Immunomethylomic approach to explore the blood neutrophil lymphocyte ratio (NLR) in glioma survival. Clin. Epigenetics 2017, 9, 10. [Google Scholar] [CrossRef]
  67. Wiencke, J.K.; Molinaro, A.M.; Warrier, G.; Rice, T.; Clarke, J.; Taylor, J.W.; Wrensch, M.; Hansen, H.; McCoy, L.; Tang, E.; et al. DNA methylation as a pharmacodynamic marker of glucocorticoid response and glioma survival. Nat. Commun. 2022, 13, 5505. [Google Scholar] [CrossRef] [PubMed]
  68. Nitta, T.; Sato, K.; Allegretta, M.; Brocke, S.; Lim, M.; Mitchell, D.J.; Steinman, L. Expression of granulocyte colony stimulating factor and granulocyte-macrophage colony stimulating factor genes in human astrocytoma cell lines and in glioma specimens. Brain Res. 1992, 571, 19–25. [Google Scholar] [CrossRef]
  69. Lad, B.M.; Beniwal, A.S.; Jain, S.; Shukla, P.; Jung, J.; Shah, S.S.; Yagnik, G.; Babikir, H.; Nguyen, A.T.; Gill, S.; et al. Glioblastoma induces the recruitment and differentiation of hybrid neutrophils from skull bone marrow. bioRxiv 2023. [Google Scholar] [CrossRef]
  70. Chen, Y.; Douglass, T.; Jeffes, E.W.; Xu, Q.; Williams, C.C.; Arpajirakul, N.; Delgado, C.; Kleinman, M.; Sanchez, R.; Dan, Q.; et al. Living T9 glioma cells expressing membrane macrophage colony-stimulating factor produce immediate tumor destruction by polymorphonuclear leukocytes and macrophages via a “paraptosis”-induced pathway that promotes systemic immunity against intracranial T9 gliomas. Blood 2002, 100, 1373–1380. [Google Scholar] [CrossRef]
  71. Zha, C.; Meng, X.; Li, L.; Mi, S.; Qian, D.; Li, Z.; Wu, P.; Hu, S.; Zhao, S.; Cai, J.; et al. Neutrophil extracellular traps mediate the crosstalk between glioma progression and the tumor microenvironment via the HMGB1/RAGE/IL-8 axis. Cancer Biol. Med. 2020, 17, 154–168. [Google Scholar] [CrossRef] [PubMed]
  72. Hor, W.S.; Huang, W.L.; Lin, Y.S.; Yang, B.C. Cross-talk between tumor cells and neutrophils through the Fas (APO-1, CD95)/FasL system: Human glioma cells enhance cell viability and stimulate cytokine production in neutrophils. J. Leukoc. Biol. 2003, 73, 363–368. [Google Scholar] [CrossRef] [PubMed]
  73. Liang, J.; Piao, Y.; Holmes, L.; Fuller, G.N.; Henry, V.; Tiao, N.; de Groot, J.F. Neutrophils promote the malignant glioma phenotype through S100A4. Clin. Cancer Res. 2014, 20, 187–198. [Google Scholar] [CrossRef] [PubMed]
  74. Magod, P.; Mastandrea, I.; Rousso-Noori, L.; Agemy, L.; Shapira, G.; Shomron, N.; Friedmann-Morvinski, D. Exploring the longitudinal glioma microenvironment landscape uncovers reprogrammed pro-tumorigenic neutrophils in the bone marrow. Cell Rep. 2021, 36, 109480. [Google Scholar] [CrossRef] [PubMed]
  75. Wculek, S.K.; Cueto, F.J.; Mujal, A.M.; Melero, I.; Krummel, M.F.; Sancho, D. Dendritic cells in cancer immunology and immunotherapy. Nat. Rev. Immunol. 2020, 20, 7–24. [Google Scholar] [CrossRef] [PubMed]
  76. Serot, J.M.; Bene, M.C.; Foliguet, B.; Faure, G.C. Monocyte-derived IL-10-secreting dendritic cells in choroid plexus epithelium. J. Neuroimmunol. 2000, 105, 115–119. [Google Scholar] [CrossRef]
  77. McMenamin, P.G. Distribution and phenotype of dendritic cells and resident tissue macrophages in the dura mater, leptomeninges, and choroid plexus of the rat brain as demonstrated in wholemount preparations. J. Comp. Neurol. 1999, 405, 553–562. [Google Scholar] [CrossRef]
  78. Luft, T.; Pang, K.C.; Thomas, E.; Hertzog, P.; Hart, D.N.; Trapani, J.; Cebon, J. Type I IFNs enhance the terminal differentiation of dendritic cells. J. Immunol. 1998, 161, 1947–1953. [Google Scholar] [CrossRef]
  79. D’Agostino, P.M.; Gottfried-Blackmore, A.; Anandasabapathy, N.; Bulloch, K. Brain dendritic cells: Biology and pathology. Acta Neuropathol. 2012, 124, 599–614. [Google Scholar] [CrossRef]
  80. Salmon, H.; Idoyaga, J.; Rahman, A.; Leboeuf, M.; Remark, R.; Jordan, S.; Casanova-Acebes, M.; Khudoynazarova, M.; Agudo, J.; Tung, N.; et al. Expansion and Activation of CD103(+) Dendritic Cell Progenitors at the Tumor Site Enhances Tumor Responses to Therapeutic PD-L1 and BRAF Inhibition. Immunity 2016, 44, 924–938. [Google Scholar] [CrossRef]
  81. Roberts, E.W.; Broz, M.L.; Binnewies, M.; Headley, M.B.; Nelson, A.E.; Wolf, D.M.; Kaisho, T.; Bogunovic, D.; Bhardwaj, N.; Krummel, M.F. Critical Role for CD103(+)/CD141(+) Dendritic Cells Bearing CCR7 for Tumor Antigen Trafficking and Priming of T Cell Immunity in Melanoma. Cancer Cell 2016, 30, 324–336. [Google Scholar] [CrossRef] [PubMed]
  82. Kastenmuller, W.; Brandes, M.; Wang, Z.; Herz, J.; Egen, J.G.; Germain, R.N. Peripheral prepositioning and local CXCL9 chemokine-mediated guidance orchestrate rapid memory CD8+ T cell responses in the lymph node. Immunity 2013, 38, 502–513. [Google Scholar] [CrossRef] [PubMed]
  83. Mittal, D.; Vijayan, D.; Putz, E.M.; Aguilera, A.R.; Markey, K.A.; Straube, J.; Kazakoff, S.; Nutt, S.L.; Takeda, K.; Hill, G.R.; et al. Interleukin-12 from CD103(+) Batf3-Dependent Dendritic Cells Required for NK-Cell Suppression of Metastasis. Cancer Immunol. Res. 2017, 5, 1098–1108. [Google Scholar] [CrossRef]
  84. Song, E.; Mao, T.; Dong, H.; Boisserand, L.S.B.; Antila, S.; Bosenberg, M.; Alitalo, K.; Thomas, J.L.; Iwasaki, A. VEGF-C-driven lymphatic drainage enables immunosurveillance of brain tumours. Nature 2020, 577, 689–694. [Google Scholar] [CrossRef]
  85. Zhou, C.; Ma, L.; Xu, H.; Huo, Y.; Luo, J. Meningeal lymphatics regulate radiotherapy efficacy through modulating anti-tumor immunity. Cell Res. 2022, 32, 543–554. [Google Scholar] [CrossRef]
  86. Zong, J.; Keskinov, A.A.; Shurin, G.V.; Shurin, M.R. Tumor-derived factors modulating dendritic cell function. Cancer Immunol. Immunother. 2016, 65, 821–833. [Google Scholar] [CrossRef] [PubMed]
  87. De Smedt, T.; Van Mechelen, M.; De Becker, G.; Urbain, J.; Leo, O.; Moser, M. Effect of interleukin-10 on dendritic cell maturation and function. Eur. J. Immunol. 1997, 27, 1229–1235. [Google Scholar] [CrossRef] [PubMed]
  88. Yamaguchi, Y.; Tsumura, H.; Miwa, M.; Inaba, K. Contrasting effects of TGF-beta 1 and TNF-alpha on the development of dendritic cells from progenitors in mouse bone marrow. Stem Cells 1997, 15, 144–153. [Google Scholar] [CrossRef]
  89. Audiger, C.; Rahman, M.J.; Yun, T.J.; Tarbell, K.V.; Lesage, S. The Importance of Dendritic Cells in Maintaining Immune Tolerance. J. Immunol. 2017, 198, 2223–2231. [Google Scholar] [CrossRef]
  90. Iberg, C.A.; Jones, A.; Hawiger, D. Dendritic Cells as Inducers of Peripheral Tolerance. Trends Immunol. 2017, 38, 793–804. [Google Scholar] [CrossRef]
  91. Tsai, W.C.; Hueng, D.Y.; Lin, C.R.; Yang, T.C.; Gao, H.W. Nrf2 Expressions Correlate with WHO Grades in Gliomas and Meningiomas. Int. J. Mol. Sci. 2016, 17, 722. [Google Scholar] [CrossRef]
  92. Wang, J.; Liu, P.; Xin, S.; Wang, Z.; Li, J. Nrf2 suppresses the function of dendritic cells to facilitate the immune escape of glioma cells. Exp. Cell Res. 2017, 360, 66–73. [Google Scholar] [CrossRef]
  93. Friedrich, M.; Hahn, M.; Michel, J.; Sankowski, R.; Kilian, M.; Kehl, N.; Gunter, M.; Bunse, T.; Pusch, S.; von Deimling, A.; et al. Dysfunctional dendritic cells limit antigen-specific T cell response in glioma. Neuro Oncol. 2023, 25, 263–276. [Google Scholar] [CrossRef]
  94. Raychaudhuri, B.; Rayman, P.; Ireland, J.; Ko, J.; Rini, B.; Borden, E.C.; Garcia, J.; Vogelbaum, M.A.; Finke, J. Myeloid-derived suppressor cell accumulation and function in patients with newly diagnosed glioblastoma. Neuro Oncol. 2011, 13, 591–599. [Google Scholar] [CrossRef] [PubMed]
  95. Gieryng, A.; Pszczolkowska, D.; Walentynowicz, K.A.; Rajan, W.D.; Kaminska, B. Immune microenvironment of gliomas. Lab. Investig. 2017, 97, 498–518. [Google Scholar] [CrossRef]
  96. Yi, L.; Xiao, H.; Xu, M.; Ye, X.; Hu, J.; Li, F.; Li, M.; Luo, C.; Yu, S.; Bian, X.; et al. Glioma-initiating cells: A predominant role in microglia/macrophages tropism to glioma. J. Neuroimmunol. 2011, 232, 75–82. [Google Scholar] [CrossRef] [PubMed]
  97. Mi, Y.; Guo, N.; Luan, J.; Cheng, J.; Hu, Z.; Jiang, P.; Jin, W.; Gao, X. The Emerging Role of Myeloid-Derived Suppressor Cells in the Glioma Immune Suppressive Microenvironment. Front. Immunol. 2020, 11, 737. [Google Scholar] [CrossRef] [PubMed]
  98. Gielen, P.R.; Schulte, B.M.; Kers-Rebel, E.D.; Verrijp, K.; Bossman, S.A.; Ter Laan, M.; Wesseling, P.; Adema, G.J. Elevated levels of polymorphonuclear myeloid-derived suppressor cells in patients with glioblastoma highly express S100A8/9 and arginase and suppress T cell function. Neuro Oncol. 2016, 18, 1253–1264. [Google Scholar] [CrossRef]
  99. Corzo, C.A.; Condamine, T.; Lu, L.; Cotter, M.J.; Youn, J.I.; Cheng, P.; Cho, H.I.; Celis, E.; Quiceno, D.G.; Padhya, T.; et al. HIF-1alpha regulates function and differentiation of myeloid-derived suppressor cells in the tumor microenvironment. J. Exp. Med. 2010, 207, 2439–2453. [Google Scholar] [CrossRef]
  100. Fu, W.; Hou, X.; Dong, L.; Hou, W. Roles of STAT3 in the pathogenesis and treatment of glioblastoma. Front. Cell Dev. Biol. 2023, 11, 1098482. [Google Scholar] [CrossRef]
  101. Kumar, V.; Patel, S.; Tcyganov, E.; Gabrilovich, D.I. The Nature of Myeloid-Derived Suppressor Cells in the Tumor Microenvironment. Trends Immunol. 2016, 37, 208–220. [Google Scholar] [CrossRef] [PubMed]
  102. Haist, M.; Stege, H.; Grabbe, S.; Bros, M. The Functional Crosstalk between Myeloid-Derived Suppressor Cells and Regulatory T Cells within the Immunosuppressive Tumor Microenvironment. Cancers 2021, 13, 210. [Google Scholar] [CrossRef]
  103. Peng, D.; Tanikawa, T.; Li, W.; Zhao, L.; Vatan, L.; Szeliga, W.; Wan, S.; Wei, S.; Wang, Y.; Liu, Y.; et al. Myeloid-Derived Suppressor Cells Endow Stem-like Qualities to Breast Cancer Cells through IL6/STAT3 and NO/NOTCH Cross-talk Signaling. Cancer Res. 2016, 76, 3156–3165. [Google Scholar] [CrossRef] [PubMed]
  104. Stiff, A.; Trikha, P.; Mundy-Bosse, B.; McMichael, E.; Mace, T.A.; Benner, B.; Kendra, K.; Campbell, A.; Gautam, S.; Abood, D.; et al. Nitric Oxide Production by Myeloid-Derived Suppressor Cells Plays a Role in Impairing Fc Receptor-Mediated Natural Killer Cell Function. Clin. Cancer Res. 2018, 24, 1891–1904. [Google Scholar] [CrossRef]
  105. Mazzoni, A.; Bronte, V.; Visintin, A.; Spitzer, J.H.; Apolloni, E.; Serafini, P.; Zanovello, P.; Segal, D.M. Myeloid suppressor lines inhibit T cell responses by an NO-dependent mechanism. J. Immunol. 2002, 168, 689–695. [Google Scholar] [CrossRef]
  106. Sinha, P.; Clements, V.K.; Bunt, S.K.; Albelda, S.M.; Ostrand-Rosenberg, S. Cross-talk between myeloid-derived suppressor cells and macrophages subverts tumor immunity toward a type 2 response. J. Immunol. 2007, 179, 977–983. [Google Scholar] [CrossRef] [PubMed]
  107. Li, H.; Han, Y.; Guo, Q.; Zhang, M.; Cao, X. Cancer-expanded myeloid-derived suppressor cells induce anergy of NK cells through membrane-bound TGF-beta 1. J. Immunol. 2009, 182, 240–249. [Google Scholar] [CrossRef]
  108. Rodriguez, P.C.; Quiceno, D.G.; Ochoa, A.C. L-arginine availability regulates T-lymphocyte cell-cycle progression. Blood 2007, 109, 1568–1573. [Google Scholar] [CrossRef]
  109. Rodriguez, P.C.; Zea, A.H.; DeSalvo, J.; Culotta, K.S.; Zabaleta, J.; Quiceno, D.G.; Ochoa, J.B.; Ochoa, A.C. L-arginine consumption by macrophages modulates the expression of CD3 zeta chain in T lymphocytes. J. Immunol. 2003, 171, 1232–1239. [Google Scholar] [CrossRef]
  110. Gonzalez-Tablas Pimenta, M.; Otero, A.; Arandia Guzman, D.A.; Pascual-Argente, D.; Ruiz Martin, L.; Sousa-Casasnovas, P.; Garcia-Martin, A.; Roa Montes de Oca, J.C.; Villasenor-Ledezma, J.; Torres Carretero, L.; et al. Tumor cell and immune cell profiles in primary human glioblastoma: Impact on patient outcome. Brain Pathol. 2021, 31, 365–380. [Google Scholar] [CrossRef]
  111. Chen, J.; Sinha, N.; Cobb, O.; Liu, C.; Ersen, A.; Phillips, J.J.; Tihan, T.; Gutmann, D.H.; Dahiya, S. Immune cell analysis of pilocytic astrocytomas reveals sexually dimorphic brain region-specific differences in T-cell content. Neurooncol. Adv. 2021, 3, vdab068. [Google Scholar] [CrossRef] [PubMed]
  112. Raskov, H.; Orhan, A.; Christensen, J.P.; Gogenur, I. Cytotoxic CD8(+) T cells in cancer and cancer immunotherapy. Br. J. Cancer 2021, 124, 359–367. [Google Scholar] [CrossRef] [PubMed]
  113. Durgeau, A.; Virk, Y.; Corgnac, S.; Mami-Chouaib, F. Recent Advances in Targeting CD8 T-Cell Immunity for More Effective Cancer Immunotherapy. Front. Immunol. 2018, 9, 14. [Google Scholar] [CrossRef] [PubMed]
  114. Kalams, S.A.; Walker, B.D. The critical need for CD4 help in maintaining effective cytotoxic T lymphocyte responses. J. Exp. Med. 1998, 188, 2199–2204. [Google Scholar] [CrossRef] [PubMed]
  115. Shankaran, V.; Ikeda, H.; Bruce, A.T.; White, J.M.; Swanson, P.E.; Old, L.J.; Schreiber, R.D. IFNgamma and lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature 2001, 410, 1107–1111. [Google Scholar] [CrossRef]
  116. Fridman, W.H.; Pages, F.; Sautes-Fridman, C.; Galon, J. The immune contexture in human tumours: Impact on clinical outcome. Nat. Rev. Cancer 2012, 12, 298–306. [Google Scholar] [CrossRef]
  117. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  118. Rahman, M.; Kresak, J.; Yang, C.; Huang, J.; Hiser, W.; Kubilis, P.; Mitchell, D. Analysis of immunobiologic markers in primary and recurrent glioblastoma. J. Neurooncol. 2018, 137, 249–257. [Google Scholar] [CrossRef]
  119. Berghoff, A.S.; Kiesel, B.; Widhalm, G.; Rajky, O.; Ricken, G.; Wohrer, A.; Dieckmann, K.; Filipits, M.; Brandstetter, A.; Weller, M.; et al. Programmed death ligand 1 expression and tumor-infiltrating lymphocytes in glioblastoma. Neuro Oncol. 2015, 17, 1064–1075. [Google Scholar] [CrossRef]
  120. Nduom, E.K.; Wei, J.; Yaghi, N.K.; Huang, N.; Kong, L.Y.; Gabrusiewicz, K.; Ling, X.; Zhou, S.; Ivan, C.; Chen, J.Q.; et al. PD-L1 expression and prognostic impact in glioblastoma. Neuro Oncol. 2016, 18, 195–205. [Google Scholar] [CrossRef]
  121. Antonios, J.P.; Soto, H.; Everson, R.G.; Moughon, D.; Orpilla, J.R.; Shin, N.P.; Sedighim, S.; Treger, J.; Odesa, S.; Tucker, A.; et al. Immunosuppressive tumor-infiltrating myeloid cells mediate adaptive immune resistance via a PD-1/PD-L1 mechanism in glioblastoma. Neuro Oncol. 2017, 19, 796–807. [Google Scholar] [CrossRef]
  122. Kohanbash, G.; Carrera, D.A.; Shrivastav, S.; Ahn, B.J.; Jahan, N.; Mazor, T.; Chheda, Z.S.; Downey, K.M.; Watchmaker, P.B.; Beppler, C.; et al. Isocitrate dehydrogenase mutations suppress STAT1 and CD8+ T cell accumulation in gliomas. J. Clin. Investig. 2017, 127, 1425–1437. [Google Scholar] [CrossRef] [PubMed]
  123. Sakaguchi, S.; Yamaguchi, T.; Nomura, T.; Ono, M. Regulatory T cells and immune tolerance. Cell 2008, 133, 775–787. [Google Scholar] [CrossRef]
  124. Woroniecka, K.; Chongsathidkiet, P.; Rhodin, K.; Kemeny, H.; Dechant, C.; Farber, S.H.; Elsamadicy, A.A.; Cui, X.; Koyama, S.; Jackson, C.; et al. T-Cell Exhaustion Signatures Vary with Tumor Type and Are Severe in Glioblastoma. Clin. Cancer Res. 2018, 24, 4175–4186. [Google Scholar] [CrossRef] [PubMed]
  125. Mohme, M.; Schliffke, S.; Maire, C.L.; Runger, A.; Glau, L.; Mende, K.C.; Matschke, J.; Gehbauer, C.; Akyuz, N.; Zapf, S.; et al. Immunophenotyping of Newly Diagnosed and Recurrent Glioblastoma Defines Distinct Immune Exhaustion Profiles in Peripheral and Tumor-infiltrating Lymphocytes. Clin. Cancer Res. 2018, 24, 4187–4200. [Google Scholar] [CrossRef] [PubMed]
  126. Cornel, A.M.; Mimpen, I.L.; Nierkens, S. MHC Class I Downregulation in Cancer: Underlying Mechanisms and Potential Targets for Cancer Immunotherapy. Cancers 2020, 12, 1760. [Google Scholar] [CrossRef] [PubMed]
  127. Zagzag, D.; Salnikow, K.; Chiriboga, L.; Yee, H.; Lan, L.; Ali, M.A.; Garcia, R.; Demaria, S.; Newcomb, E.W. Downregulation of major histocompatibility complex antigens in invading glioma cells: Stealth invasion of the brain. Lab. Investig. 2005, 85, 328–341. [Google Scholar] [CrossRef]
  128. Taylor, B.C.; Balko, J.M. Mechanisms of MHC-I Downregulation and Role in Immunotherapy Response. Front. Immunol. 2022, 13, 844866. [Google Scholar] [CrossRef]
  129. Mondal, I.; Das, O.; Sun, R.; Gao, J.; Yu, B.; Diaz, A.; Behnan, J.; Dubey, A.; Meng, Z.; Eskandar, E.; et al. PP2Ac Deficiency Enhances Tumor Immunogenicity by Activating STING-Type I Interferon Signaling in Glioblastoma. Cancer Res. 2023, 83, 2527–2542. [Google Scholar] [CrossRef]
  130. Maggio, D.; Ho, W.S.; Breese, R.; Walbridge, S.; Wang, H.; Cui, J.; Heiss, J.D.; Gilbert, M.R.; Kovach, J.S.; Lu, R.O.; et al. Inhibition of protein phosphatase-2A with LB-100 enhances antitumor immunity against glioblastoma. J. Neurooncol. 2020, 148, 231–244. [Google Scholar] [CrossRef]
  131. Ho, W.S.; Sizdahkhani, S.; Hao, S.; Song, H.; Seldomridge, A.; Tandle, A.; Maric, D.; Kramp, T.; Lu, R.; Heiss, J.D.; et al. LB-100, a novel Protein Phosphatase 2A (PP2A) inhibitor, sensitizes malignant meningioma cells to the therapeutic effects of radiation. Cancer Lett. 2018, 415, 217–226. [Google Scholar] [CrossRef]
  132. Fecci, P.E.; Mitchell, D.A.; Whitesides, J.F.; Xie, W.; Friedman, A.H.; Archer, G.E.; Herndon, J.E., 2nd; Bigner, D.D.; Dranoff, G.; Sampson, J.H. Increased regulatory T-cell fraction amidst a diminished CD4 compartment explains cellular immune defects in patients with malignant glioma. Cancer Res. 2006, 66, 3294–3302. [Google Scholar] [CrossRef]
  133. Chang, A.L.; Miska, J.; Wainwright, D.A.; Dey, M.; Rivetta, C.V.; Yu, D.; Kanojia, D.; Pituch, K.C.; Qiao, J.; Pytel, P.; et al. CCL2 Produced by the Glioma Microenvironment Is Essential for the Recruitment of Regulatory T Cells and Myeloid-Derived Suppressor Cells. Cancer Res. 2016, 76, 5671–5682. [Google Scholar] [CrossRef]
  134. Humphries, W.; Wei, J.; Sampson, J.H.; Heimberger, A.B. The role of tregs in glioma-mediated immunosuppression: Potential target for intervention. Neurosurg. Clin. N. Am. 2010, 21, 125–137. [Google Scholar] [CrossRef] [PubMed]
  135. Crane, C.A.; Ahn, B.J.; Han, S.J.; Parsa, A.T. Soluble factors secreted by glioblastoma cell lines facilitate recruitment, survival, and expansion of regulatory T cells: Implications for immunotherapy. Neuro Oncol. 2012, 14, 584–595. [Google Scholar] [CrossRef] [PubMed]
  136. Wainwright, D.A.; Balyasnikova, I.V.; Chang, A.L.; Ahmed, A.U.; Moon, K.S.; Auffinger, B.; Tobias, A.L.; Han, Y.; Lesniak, M.S. IDO expression in brain tumors increases the recruitment of regulatory T cells and negatively impacts survival. Clin. Cancer Res. 2012, 18, 6110–6121. [Google Scholar] [CrossRef] [PubMed]
  137. Ward-Hartstonge, K.A.; Kemp, R.A. Regulatory T-cell heterogeneity and the cancer immune response. Clin. Transl. Immunol. 2017, 6, e154. [Google Scholar] [CrossRef] [PubMed]
  138. Thornton, A.M.; Shevach, E.M. CD4+CD25+ immunoregulatory T cells suppress polyclonal T cell activation in vitro by inhibiting interleukin 2 production. J. Exp. Med. 1998, 188, 287–296. [Google Scholar] [CrossRef]
  139. Collison, L.W.; Workman, C.J.; Kuo, T.T.; Boyd, K.; Wang, Y.; Vignali, K.M.; Cross, R.; Sehy, D.; Blumberg, R.S.; Vignali, D.A. The inhibitory cytokine IL-35 contributes to regulatory T-cell function. Nature 2007, 450, 566–569. [Google Scholar] [CrossRef]
  140. Jarnicki, A.G.; Lysaght, J.; Todryk, S.; Mills, K.H. Suppression of antitumor immunity by IL-10 and TGF-beta-producing T cells infiltrating the growing tumor: Influence of tumor environment on the induction of CD4+ and CD8+ regulatory T cells. J. Immunol. 2006, 177, 896–904. [Google Scholar] [CrossRef]
  141. Deaglio, S.; Dwyer, K.M.; Gao, W.; Friedman, D.; Usheva, A.; Erat, A.; Chen, J.F.; Enjyoji, K.; Linden, J.; Oukka, M.; et al. Adenosine generation catalyzed by CD39 and CD73 expressed on regulatory T cells mediates immune suppression. J. Exp. Med. 2007, 204, 1257–1265. [Google Scholar] [CrossRef] [PubMed]
  142. Wing, K.; Onishi, Y.; Prieto-Martin, P.; Yamaguchi, T.; Miyara, M.; Fehervari, Z.; Nomura, T.; Sakaguchi, S. CTLA-4 control over Foxp3+ regulatory T cell function. Science 2008, 322, 271–275. [Google Scholar] [CrossRef] [PubMed]
  143. Grossman, W.J.; Verbsky, J.W.; Barchet, W.; Colonna, M.; Atkinson, J.P.; Ley, T.J. Human T regulatory cells can use the perforin pathway to cause autologous target cell death. Immunity 2004, 21, 589–601. [Google Scholar] [CrossRef]
  144. Wainwright, D.A.; Dey, M.; Chang, A.; Lesniak, M.S. Targeting Tregs in Malignant Brain Cancer: Overcoming IDO. Front. Immunol. 2013, 4, 116. [Google Scholar] [CrossRef] [PubMed]
  145. Imai, K.; Matsuyama, S.; Miyake, S.; Suga, K.; Nakachi, K. Natural cytotoxic activity of peripheral-blood lymphocytes and cancer incidence: An 11-year follow-up study of a general population. Lancet 2000, 356, 1795–1799. [Google Scholar] [CrossRef]
  146. Trotta, R.; Dal Col, J.; Yu, J.; Ciarlariello, D.; Thomas, B.; Zhang, X.; Allard, J., 2nd; Wei, M.; Mao, H.; Byrd, J.C.; et al. TGF-beta utilizes SMAD3 to inhibit CD16-mediated IFN-gamma production and antibody-dependent cellular cytotoxicity in human NK cells. J. Immunol. 2008, 181, 3784–3792. [Google Scholar] [CrossRef]
  147. Crane, C.A.; Han, S.J.; Barry, J.J.; Ahn, B.J.; Lanier, L.L.; Parsa, A.T. TGF-beta downregulates the activating receptor NKG2D on NK cells and CD8+ T cells in glioma patients. Neuro Oncol. 2010, 12, 7–13. [Google Scholar] [CrossRef]
  148. Holt, D.; Ma, X.; Kundu, N.; Fulton, A. Prostaglandin E(2) (PGE(2)) suppresses natural killer cell function primarily through the PGE(2) receptor EP4. Cancer Immunol. Immunother. 2011, 60, 1577–1586. [Google Scholar] [CrossRef]
  149. Labadie, B.W.; Bao, R.; Luke, J.J. Reimagining IDO Pathway Inhibition in Cancer Immunotherapy via Downstream Focus on the Tryptophan-Kynurenine-Aryl Hydrocarbon Axis. Clin. Cancer Res. 2019, 25, 1462–1471. [Google Scholar] [CrossRef]
  150. Young, A.; Ngiow, S.F.; Gao, Y.; Patch, A.M.; Barkauskas, D.S.; Messaoudene, M.; Lin, G.; Coudert, J.D.; Stannard, K.A.; Zitvogel, L.; et al. A2AR Adenosine Signaling Suppresses Natural Killer Cell Maturation in the Tumor Microenvironment. Cancer Res. 2018, 78, 1003–1016. [Google Scholar] [CrossRef]
  151. Brand, A.; Singer, K.; Koehl, G.E.; Kolitzus, M.; Schoenhammer, G.; Thiel, A.; Matos, C.; Bruss, C.; Klobuch, S.; Peter, K.; et al. LDHA-Associated Lactic Acid Production Blunts Tumor Immunosurveillance by T and NK Cells. Cell Metab. 2016, 24, 657–671. [Google Scholar] [CrossRef] [PubMed]
  152. Weiss, T.; Puca, E.; Silginer, M.; Hemmerle, T.; Pazahr, S.; Bink, A.; Weller, M.; Neri, D.; Roth, P. Immunocytokines are a promising immunotherapeutic approach against glioblastoma. Sci. Transl. Med. 2020, 12, eabb2311. [Google Scholar] [CrossRef] [PubMed]
  153. Birocchi, F.; Cusimano, M.; Rossari, F.; Beretta, S.; Rancoita, P.M.V.; Ranghetti, A.; Colombo, S.; Costa, B.; Angel, P.; Sanvito, F.; et al. Targeted inducible delivery of immunoactivating cytokines reprograms glioblastoma microenvironment and inhibits growth in mouse models. Sci. Transl. Med. 2022, 14, eabl4106. [Google Scholar] [CrossRef] [PubMed]
  154. Butowski, N.; Colman, H.; De Groot, J.F.; Omuro, A.M.; Nayak, L.; Wen, P.Y.; Cloughesy, T.F.; Marimuthu, A.; Haidar, S.; Perry, A.; et al. Orally administered colony stimulating factor 1 receptor inhibitor PLX3397 in recurrent glioblastoma: An Ivy Foundation Early Phase Clinical Trials Consortium phase II study. Neuro Oncol. 2016, 18, 557–564. [Google Scholar] [CrossRef] [PubMed]
  155. Flores-Toro, J.A.; Luo, D.; Gopinath, A.; Sarkisian, M.R.; Campbell, J.J.; Charo, I.F.; Singh, R.; Schall, T.J.; Datta, M.; Jain, R.K.; et al. CCR2 inhibition reduces tumor myeloid cells and unmasks a checkpoint inhibitor effect to slow progression of resistant murine gliomas. Proc. Natl. Acad. Sci. USA 2020, 117, 1129–1138. [Google Scholar] [CrossRef] [PubMed]
  156. Campian, J.L.; Ghosh, S.; Kapoor, V.; Yan, R.; Thotala, S.; Jash, A.; Hu, T.; Mahadevan, A.; Rifai, K.; Page, L.; et al. Long-Acting Recombinant Human Interleukin-7, NT-I7, Increases Cytotoxic CD8 T Cells and Enhances Survival in Mouse Glioma Models. Clin. Cancer Res. 2022, 28, 1229–1239. [Google Scholar] [CrossRef] [PubMed]
  157. Ho, W.S.; Mondal, I.; Xu, B.; Das, O.; Sun, R.; Chiou, P.; Cai, X.; Tahmasebinia, F.; McFadden, E.; Wu, C.Y.; et al. PP2Ac/STRN4 negatively regulates STING-type I IFN signaling in tumor-associated macrophages. J. Clin. Investig. 2023, 133, e162139. [Google Scholar] [CrossRef] [PubMed]
  158. Nayak, L.; Molinaro, A.M.; Peters, K.; Clarke, J.L.; Jordan, J.T.; de Groot, J.; Nghiemphu, L.; Kaley, T.; Colman, H.; McCluskey, C.; et al. Randomized Phase II and Biomarker Study of Pembrolizumab plus Bevacizumab versus Pembrolizumab Alone for Patients with Recurrent Glioblastoma. Clin. Cancer Res. 2021, 27, 1048–1057. [Google Scholar] [CrossRef]
  159. Omuro, A.; Brandes, A.A.; Carpentier, A.F.; Idbaih, A.; Reardon, D.A.; Cloughesy, T.; Sumrall, A.; Baehring, J.; van den Bent, M.; Bahr, O.; et al. Radiotherapy combined with nivolumab or temozolomide for newly diagnosed glioblastoma with unmethylated MGMT promoter: An international randomized phase III trial. Neuro Oncol. 2023, 25, 123–134. [Google Scholar] [CrossRef] [PubMed]
  160. Lim, M.; Weller, M.; Idbaih, A.; Steinbach, J.; Finocchiaro, G.; Raval, R.R.; Ansstas, G.; Baehring, J.; Taylor, J.W.; Honnorat, J.; et al. Phase III trial of chemoradiotherapy with temozolomide plus nivolumab or placebo for newly diagnosed glioblastoma with methylated MGMT promoter. Neuro Oncol. 2022, 24, 1935–1949. [Google Scholar] [CrossRef] [PubMed]
  161. Omuro, A.; Vlahovic, G.; Lim, M.; Sahebjam, S.; Baehring, J.; Cloughesy, T.; Voloschin, A.; Ramkissoon, S.H.; Ligon, K.L.; Latek, R.; et al. Nivolumab with or without ipilimumab in patients with recurrent glioblastoma: Results from exploratory phase I cohorts of CheckMate 143. Neuro Oncol. 2018, 20, 674–686. [Google Scholar] [CrossRef] [PubMed]
  162. Cloughesy, T.F.; Mochizuki, A.Y.; Orpilla, J.R.; Hugo, W.; Lee, A.H.; Davidson, T.B.; Wang, A.C.; Ellingson, B.M.; Rytlewski, J.A.; Sanders, C.M.; et al. Neoadjuvant anti-PD-1 immunotherapy promotes a survival benefit with intratumoral and systemic immune responses in recurrent glioblastoma. Nat. Med. 2019, 25, 477–486. [Google Scholar] [CrossRef]
  163. Zhao, J.; Chen, A.X.; Gartrell, R.D.; Silverman, A.M.; Aparicio, L.; Chu, T.; Bordbar, D.; Shan, D.; Samanamud, J.; Mahajan, A.; et al. Immune and genomic correlates of response to anti-PD-1 immunotherapy in glioblastoma. Nat. Med. 2019, 25, 462–469. [Google Scholar] [CrossRef] [PubMed]
  164. Arrieta, V.A.; Dmello, C.; McGrail, D.J.; Brat, D.J.; Lee-Chang, C.; Heimberger, A.B.; Chand, D.; Stupp, R.; Sonabend, A.M. Immune checkpoint blockade in glioblastoma: From tumor heterogeneity to personalized treatment. J. Clin. Investig. 2023, 133, e163447. [Google Scholar] [CrossRef] [PubMed]
  165. McGrail, D.J.; Pilie, P.G.; Dai, H.; Lam, T.N.A.; Liang, Y.; Voorwerk, L.; Kok, M.; Zhang, X.H.; Rosen, J.M.; Heimberger, A.B.; et al. Replication stress response defects are associated with response to immune checkpoint blockade in nonhypermutated cancers. Sci. Transl. Med. 2021, 13, eabe6201. [Google Scholar] [CrossRef] [PubMed]
  166. Schalper, K.A.; Rodriguez-Ruiz, M.E.; Diez-Valle, R.; Lopez-Janeiro, A.; Porciuncula, A.; Idoate, M.A.; Inoges, S.; de Andrea, C.; Lopez-Diaz de Cerio, A.; Tejada, S.; et al. Neoadjuvant nivolumab modifies the tumor immune microenvironment in resectable glioblastoma. Nat. Med. 2019, 25, 470–476. [Google Scholar] [CrossRef] [PubMed]
  167. Sampson, J.H.; Aldape, K.D.; Archer, G.E.; Coan, A.; Desjardins, A.; Friedman, A.H.; Friedman, H.S.; Gilbert, M.R.; Herndon, J.E.; McLendon, R.E.; et al. Greater chemotherapy-induced lymphopenia enhances tumor-specific immune responses that eliminate EGFRvIII-expressing tumor cells in patients with glioblastoma. Neuro Oncol. 2011, 13, 324–333. [Google Scholar] [CrossRef]
  168. Luksik, A.S.; Yazigi, E.; Shah, P.; Jackson, C.M. CAR T Cell Therapy in Glioblastoma: Overcoming Challenges Related to Antigen Expression. Cancers 2023, 15, 1414. [Google Scholar] [CrossRef]
  169. Sampson, J.H.; Choi, B.D.; Sanchez-Perez, L.; Suryadevara, C.M.; Snyder, D.J.; Flores, C.T.; Schmittling, R.J.; Nair, S.K.; Reap, E.A.; Norberg, P.K.; et al. EGFRvIII mCAR-modified T-cell therapy cures mice with established intracerebral glioma and generates host immunity against tumor-antigen loss. Clin. Cancer Res. 2014, 20, 972–984. [Google Scholar] [CrossRef]
  170. O’Rourke, D.M.; Nasrallah, M.P.; Desai, A.; Melenhorst, J.J.; Mansfield, K.; Morrissette, J.J.D.; Martinez-Lage, M.; Brem, S.; Maloney, E.; Shen, A.; et al. A single dose of peripherally infused EGFRvIII-directed CAR T cells mediates antigen loss and induces adaptive resistance in patients with recurrent glioblastoma. Sci. Transl. Med. 2017, 9, eaaa0984. [Google Scholar] [CrossRef] [PubMed]
  171. Brown, C.E.; Alizadeh, D.; Starr, R.; Weng, L.; Wagner, J.R.; Naranjo, A.; Ostberg, J.R.; Blanchard, M.S.; Kilpatrick, J.; Simpson, J.; et al. Regression of Glioblastoma after Chimeric Antigen Receptor T-Cell Therapy. N. Engl. J. Med. 2016, 375, 2561–2569. [Google Scholar] [CrossRef] [PubMed]
  172. Choe, J.H.; Watchmaker, P.B.; Simic, M.S.; Gilbert, R.D.; Li, A.W.; Krasnow, N.A.; Downey, K.M.; Yu, W.; Carrera, D.A.; Celli, A.; et al. SynNotch-CAR T cells overcome challenges of specificity, heterogeneity, and persistence in treating glioblastoma. Sci. Transl. Med. 2021, 13, eabe7378. [Google Scholar] [CrossRef]
  173. Ahmed, N.; Brawley, V.; Hegde, M.; Bielamowicz, K.; Kalra, M.; Landi, D.; Robertson, C.; Gray, T.L.; Diouf, O.; Wakefield, A.; et al. HER2-Specific Chimeric Antigen Receptor-Modified Virus-Specific T Cells for Progressive Glioblastoma: A Phase 1 Dose-Escalation Trial. JAMA Oncol. 2017, 3, 1094–1101. [Google Scholar] [CrossRef] [PubMed]
  174. Brown, C.E.; Badie, B.; Barish, M.E.; Weng, L.; Ostberg, J.R.; Chang, W.C.; Naranjo, A.; Starr, R.; Wagner, J.; Wright, C.; et al. Bioactivity and Safety of IL13Ralpha2-Redirected Chimeric Antigen Receptor CD8+ T Cells in Patients with Recurrent Glioblastoma. Clin. Cancer Res. 2015, 21, 4062–4072. [Google Scholar] [CrossRef]
  175. Chheda, Z.S.; Kohanbash, G.; Okada, K.; Jahan, N.; Sidney, J.; Pecoraro, M.; Yang, X.; Carrera, D.A.; Downey, K.M.; Shrivastav, S.; et al. Novel and shared neoantigen derived from histone 3 variant H3.3K27M mutation for glioma T cell therapy. J. Exp. Med. 2018, 215, 141–157. [Google Scholar] [CrossRef]
  176. Majzner, R.G.; Ramakrishna, S.; Yeom, K.W.; Patel, S.; Chinnasamy, H.; Schultz, L.M.; Richards, R.M.; Jiang, L.; Barsan, V.; Mancusi, R.; et al. GD2-CAR T cell therapy for H3K27M-mutated diffuse midline gliomas. Nature 2022, 603, 934–941. [Google Scholar] [CrossRef] [PubMed]
  177. Schmidts, A.; Srivastava, A.A.; Ramapriyan, R.; Bailey, S.R.; Bouffard, A.A.; Cahill, D.P.; Carter, B.S.; Curry, W.T.; Dunn, G.P.; Frigault, M.J.; et al. Tandem chimeric antigen receptor (CAR) T cells targeting EGFRvIII and IL-13Ralpha2 are effective against heterogeneous glioblastoma. Neurooncol. Adv. 2023, 5, vdac185. [Google Scholar] [CrossRef]
  178. Muhammad, N.; Wang, R.; Li, W.; Zhang, Z.; Chang, Y.; Hu, Y.; Zhao, J.; Zheng, X.; Mao, Q.; Xia, H. A novel TanCAR targeting IL13Ralpha2 and EphA2 for enhanced glioblastoma therapy. Mol. Ther. Oncolytics 2022, 24, 729–741. [Google Scholar] [CrossRef]
  179. Hamad, A.; Yusubalieva, G.M.; Baklaushev, V.P.; Chumakov, P.M.; Lipatova, A.V. Recent Developments in Glioblastoma Therapy: Oncolytic Viruses and Emerging Future Strategies. Viruses 2023, 15, 547. [Google Scholar] [CrossRef]
  180. Lang, F.F.; Conrad, C.; Gomez-Manzano, C.; Yung, W.K.A.; Sawaya, R.; Weinberg, J.S.; Prabhu, S.S.; Rao, G.; Fuller, G.N.; Aldape, K.D.; et al. Phase I Study of DNX-2401 (Delta-24-RGD) Oncolytic Adenovirus: Replication and Immunotherapeutic Effects in Recurrent Malignant Glioma. J. Clin. Oncol. 2018, 36, 1419–1427. [Google Scholar] [CrossRef]
  181. Gallego Perez-Larraya, J.; Garcia-Moure, M.; Labiano, S.; Patino-Garcia, A.; Dobbs, J.; Gonzalez-Huarriz, M.; Zalacain, M.; Marrodan, L.; Martinez-Velez, N.; Puigdelloses, M.; et al. Oncolytic DNX-2401 Virus for Pediatric Diffuse Intrinsic Pontine Glioma. N. Engl. J. Med. 2022, 386, 2471–2481. [Google Scholar] [CrossRef]
  182. Friedman, G.K.; Johnston, J.M.; Bag, A.K.; Bernstock, J.D.; Li, R.; Aban, I.; Kachurak, K.; Nan, L.; Kang, K.D.; Totsch, S.; et al. Oncolytic HSV-1 G207 Immunovirotherapy for Pediatric High-Grade Gliomas. N. Engl. J. Med. 2021, 384, 1613–1622. [Google Scholar] [CrossRef]
  183. Ling, A.L.; Solomon, I.H.; Landivar, A.M.; Nakashima, H.; Woods, J.K.; Santos, A.; Masud, N.; Fell, G.; Mo, X.; Yilmaz, A.S.; et al. Clinical trial links oncolytic immunoactivation to survival in glioblastoma. Nature 2023, 623, 157–166. [Google Scholar] [CrossRef]
  184. Weller, M.; Butowski, N.; Tran, D.D.; Recht, L.D.; Lim, M.; Hirte, H.; Ashby, L.; Mechtler, L.; Goldlust, S.A.; Iwamoto, F.; et al. Rindopepimut with temozolomide for patients with newly diagnosed, EGFRvIII-expressing glioblastoma (ACT IV): A randomised, double-blind, international phase 3 trial. Lancet Oncol. 2017, 18, 1373–1385. [Google Scholar] [CrossRef] [PubMed]
  185. Fenstermaker, R.A.; Ciesielski, M.J.; Qiu, J.; Yang, N.; Frank, C.L.; Lee, K.P.; Mechtler, L.R.; Belal, A.; Ahluwalia, M.S.; Hutson, A.D. Clinical study of a survivin long peptide vaccine (SurVaxM) in patients with recurrent malignant glioma. Cancer Immunol. Immunother. 2016, 65, 1339–1352. [Google Scholar] [CrossRef] [PubMed]
  186. Sener, U.; Ruff, M.W.; Campian, J.L. Immunotherapy in Glioblastoma: Current Approaches and Future Perspectives. Int. J. Mol Sci. 2022, 23, 46. [Google Scholar] [CrossRef] [PubMed]
  187. Luo, W. Nasopharyngeal carcinoma ecology theory: Cancer as multidimensional spatiotemporal “unity of ecology and evolution” pathological ecosystem. Theranostics 2023, 13, 1607–1631. [Google Scholar] [CrossRef]
  188. Ceccarelli, M.; Barthel, F.P.; Malta, T.M.; Sabedot, T.S.; Salama, S.R.; Murray, B.A.; Morozova, O.; Newton, Y.; Radenbaugh, A.; Pagnotta, S.M.; et al. Molecular Profiling Reveals Biologically Discrete Subsets and Pathways of Progression in Diffuse Glioma. Cell 2016, 164, 550–563. [Google Scholar] [CrossRef]
  189. Varn, F.S.; Johnson, K.C.; Martinek, J.; Huse, J.T.; Nasrallah, M.P.; Wesseling, P.; Cooper, L.A.D.; Malta, T.M.; Wade, T.E.; Sabedot, T.S.; et al. Glioma progression is shaped by genetic evolution and microenvironment interactions. Cell 2022, 185, 2184–2199.e2116. [Google Scholar] [CrossRef]
  190. Miller, T.E.; El Farran, C.A.; Couturier, C.P.; Chen, Z.; D’Antonio, J.P.; Verga, J.; Villanueva, M.A.; Castro, L.N.G.; Tong, Y.E.; Saadi, T.A.; et al. Programs, Origins, and Niches of Immunomodulatory Myeloid Cells in Gliomas. bioRxiv 2023. [Google Scholar] [CrossRef]
  191. Yeo, A.T.; Rawal, S.; Delcuze, B.; Christofides, A.; Atayde, A.; Strauss, L.; Balaj, L.; Rogers, V.A.; Uhlmann, E.J.; Varma, H.; et al. Single-cell RNA sequencing reveals evolution of immune landscape during glioblastoma progression. Nat. Immunol. 2022, 23, 971–984. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The bidirectional cellular interactions between myeloid and glioma cells. Myeloid and glioma cells exhibit numerous crosstalk interactions. Gliomas secrete various factors that help recruit and reprogram various myeloid immune cells. In turn, these reprogrammed myeloid cells can also secrete factors that promote tumor growth, transcriptionally rewire glioma cells towards a mesenchymal-like profile, and further recruit and reconfigure other immune cells in the TME, allowing tumor cells to evade and suppress the immune system. See text for abbreviations. Up and down arrows indicate upregulation and downregulation of the associated molecules or signaling pathways, respectively. Font color of signaling pathways upregulated in the schematized glioma cell corresponds to the associated immune cell color. Created with BioRender.com.
Figure 1. The bidirectional cellular interactions between myeloid and glioma cells. Myeloid and glioma cells exhibit numerous crosstalk interactions. Gliomas secrete various factors that help recruit and reprogram various myeloid immune cells. In turn, these reprogrammed myeloid cells can also secrete factors that promote tumor growth, transcriptionally rewire glioma cells towards a mesenchymal-like profile, and further recruit and reconfigure other immune cells in the TME, allowing tumor cells to evade and suppress the immune system. See text for abbreviations. Up and down arrows indicate upregulation and downregulation of the associated molecules or signaling pathways, respectively. Font color of signaling pathways upregulated in the schematized glioma cell corresponds to the associated immune cell color. Created with BioRender.com.
Cancers 16 00308 g001
Figure 2. Immunosuppressive tumor microenvironment interactions limiting anti-tumor lymphocyte activity. Gliomas and other tumor-associated cells express various factors, including immune checkpoint proteins and cytokines, that result in dysfunctional, exhausted CD8+ T cells. In addition, gliomas sculpt the TME to promote recruitment of Treg cells, pro-tumor TAMs, and MDSCs, which further halt cytotoxic immune activation and promote tumor immune evasion. Up and down arrows illustrate upregulation and downregulation of the indicated molecules, respectively. Created with BioRender.com.
Figure 2. Immunosuppressive tumor microenvironment interactions limiting anti-tumor lymphocyte activity. Gliomas and other tumor-associated cells express various factors, including immune checkpoint proteins and cytokines, that result in dysfunctional, exhausted CD8+ T cells. In addition, gliomas sculpt the TME to promote recruitment of Treg cells, pro-tumor TAMs, and MDSCs, which further halt cytotoxic immune activation and promote tumor immune evasion. Up and down arrows illustrate upregulation and downregulation of the indicated molecules, respectively. Created with BioRender.com.
Cancers 16 00308 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Elguindy, M.; Young, J.S.; Mondal, I.; Lu, R.O.; Ho, W.S. Glioma–Immune Cell Crosstalk in Tumor Progression. Cancers 2024, 16, 308. https://doi.org/10.3390/cancers16020308

AMA Style

Elguindy M, Young JS, Mondal I, Lu RO, Ho WS. Glioma–Immune Cell Crosstalk in Tumor Progression. Cancers. 2024; 16(2):308. https://doi.org/10.3390/cancers16020308

Chicago/Turabian Style

Elguindy, Mahmoud, Jacob S. Young, Isha Mondal, Rongze O. Lu, and Winson S. Ho. 2024. "Glioma–Immune Cell Crosstalk in Tumor Progression" Cancers 16, no. 2: 308. https://doi.org/10.3390/cancers16020308

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