IMR Press / FBL / Volume 28 / Issue 9 / DOI: 10.31083/j.fbl2809206
Open Access Review
Present Status and Advances in Chimeric Antigen Receptor T Cell Therapy for Glioblastoma
Shuchang Zhou1,2,†Han Sun1,2,†Sun Il Choi1,2,*Jinlong Yin1,2,*
Show Less
1 Henan Key Laboratory of Brain Targeted Bio-Nanomedicine, School of Life Sciences & School of Pharmacy, Henan University, 475004 Kaifeng, Henan, China
2 Henan-Macquarie University Joint Centre for Biomedical Innovation, School of Life Sciences, Henan University, 475004 Kaifeng, Henan, China
*Correspondence: sunil.choi@hotmail.com (Sun Il Choi); jlyin@henu.edu.cn (Jinlong Yin)
These authors contributed equally.
Front. Biosci. (Landmark Ed) 2023, 28(9), 206; https://doi.org/10.31083/j.fbl2809206
Submitted: 25 February 2023 | Revised: 4 April 2023 | Accepted: 7 April 2023 | Published: 15 September 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Adoptive chimeric antigen receptor (CAR) T cells designed to recognize specific tumor antigens have shown promising results in cancer therapy. While CAR T cell therapy has demonstrated notable clinical effectiveness for hematologic disease, efforts to develop therapies for solid tumors, including glioblastoma (GBM), have been hampered by heterogeneity, an immunosuppressive tumor microenvironment, and difficulty in trafficking. Several specific tumor antigens, such as IL13Rα2, EGFRvIII, and HER2, have been attempted in clinical trials; however, limited efficacy has been observed. In this review, we discuss the current status of CAR T therapy for GBM in clinical trials and highlight the potential target antigens for CAR T cells. Additionally, we summarize the mechanisms used to enhance their efficacy and explore the challenges and future prospects of CAR T cell therapy for GBM.

Keywords
CAR T
glioblastoma
immunotherapy
cancer therapy
cell therapy
1. Introduction

Glioblastoma (GBM) is a primary malignant tumor of the central nervous system (CNS) that is the most common and aggressive with World Health Organization (WHO) grade IV tumors [1, 2]. Standard therapies for GBM include microsurgical resection followed by adjuvant chemotherapy and radiotherapy. However, despite such aggressive therapy and intense research efforts, the overall survival (OS) rate has remained at 15–18 months [1, 2]. Therefore, new therapeutic approaches in GBM are urgently warranted [3].

Immunotherapy that uses natural defense mechanisms was first attempted as adoptive cell transfer using autologous lymphocytes in patients with hematological cancer [4] and now could revolutionize the treatment of even solid cancers. However, these approaches lack specificity for cancer cells, and to address this limitation, the exploration of genetically engineered T cell receptors (TCR) to identify specific target cancer-associated antigens ultimately generated chimeric antigen receptor (CAR) T cell therapies [5]. CAR T cell therapy has shown the potential to prolong patient survival and induce remission, even in cases where standard treatment has failed [6]. However, since the first CAR was reported in 1993, it has taken decades for CAR T cells targeting CD19 to be approved by the Food and Drug Administration (FDA) as a therapy for patients with relapsed or refractory B cell acute lymphoblastic leukemia, and investigators are continually seeking to extend the therapeutic effects of CAR T cells [7, 8]. Although recent clinical and ongoing studies have yielded some promising results, the progress of CAR T cell therapy for GBM remains challenging, with only suboptimal efficacy achieved to date [9].

Solid tumors such as GBM display a high antigen heterogeneity compared to liquid tumors. This allows tumor cells to evade recognition by CAR T cells, typically encode specificity toward a single antigenic target, and thus cannot identify all cancer cells in a tumor. In addition, expanding T cell specificity to multiple antigens can increase the risk of on-target, off-tumor toxicity. Moreover, physical barriers, such as collagen-rich stroma surrounding tumor cells, can prevent T cell infiltration in solid tumors. Furthermore, the CAR T cells administered to patients with GBM encounter significant challenges in a highly immunosuppressive TME with molecular, cellular, and metabolic profiles that eventually create exhaustion and dysfunction of T cells. In GBM therapy, CAR T cells also need to penetrate the brain tissue protected by the blood-brain barrier (BBB). Early-phase clinical studies of CAR T cell therapy for patients with GBM have demonstrated safety but did not achieve sufficient antitumor activity. Despite this, some CAR T cells have shown tumor control, and complete responses have been reported [9]. In this review, we summarize the targets for completed and ongoing clinical trials in GBM therapy (Fig. 1) (Tables 1,2, Ref. [10, 11, 12, 13, 14]) and discuss the challenges and prospects of CAR T application (Fig. 2) to guide subsequent research for the treatment of GBM.

Fig. 1.

Several targetable tumor-associated antigens for GBM CAR T cell therapy. The figure was created with BioRender.com.

Fig. 2.

Summary of challenges and solutions for CAR T cell-based therapeutic in GBM. The figure was created with BioRender.com.

Table 1.Completed chimeric antigen receptor (CAR) T cell-based clinical trials in patients with glioblastoma (GBM).
Molecular target Clinical trial identifier and title Study phase CAR T cell dosage Sponsor/site Response (Median OS) Ref.
IL13Rα2 NCT00730613 I - intravenous infusion City of Hope Medical Center 11 months [10]
Cellular adoptive immunotherapy using genetically modified T-lymphocytes in treating patients with recurrent or refractory high-grade malignant glioma - up to 108 CAR T cells on days 1, 3, and 5 for 2 weeks treatment repeated after 3 weeks National Cancer Institute (NCI)
NCT01082926 I - intravenous infusion City of Hope Medical Center 19.7 month [11]
Phase I study of cellular immunotherapy for recurrent/refractory malignant glioma using intratumoral infusions of GRm13z40-2, an allogeneic CD8+ cytolitic T-cell line genetically modified to express the il 13-Zetakine and hytk and to be resistant to glucocorticoids, in combination with interleukin-2 - 1 × 108 CAR T cells with IL-2 twice per week for 2 weeks National Cancer Institute (NCI)
EGFRvIII NCT02209376 I - intravenous infusion University of Pennsylvania 8 months [12]
Autologous T cells redirected to EGFRVIII-with a chimeric antigen receptor in patients with EGFRVIII + GBM - 1.75 × 108 –5 × 108 CAR T cells University of California, San Francisco
NCT01454596 I/II - 6.3 × 106 to 2.6 × 1010 EGFRvIII-CAR T cells National Cancer Institute (NCI) 6.9 months [13]
CAR T Cell receptor immunotherapy targeting EGFRvIII for patients with malignant gliomas expressing EGFRvIII
HER2 NCT01109095 I - 1 × 106/m2 to 1 × 108/m2 autologous HER2-CAR VSTs Baylor College of Medicine 11.1 months (95% CI, 4.1–27.2 months) from the first T cell infusion [14]
CMV-specific cytotoxic T lymphocytes expressing CAR targeting HER2 in patients with GBM The Methodist Hospital Research Institute 24.5 months (95% CI, 17.2–34.6 months) from diagnosis
Center for Cell and Gene Therapy, Baylor College of Medicine
Table 2.Ongoing CAR T cell-based clinical trials in patients with GBM.
Molecular target Clinical trial identifier and title Study phase CAR T cell dosage Sponsor/site
NKG2DL NCT04270461 I - intravenous administration or hepatic portal artery injection Jiujiang University Affiliated Hospital KAEDI
NKG2D-based CAR T cells immunotherapy for patients with r/r NKG2DL+ solid tumors - NKG2D-CAR T cells: 1–10 × 106/kg, 70 mL/time
B7-H3 NCT04077866 I/II - three intratumoral or intracerebroventricular injections of CAR T cells at two doses in between temozolomide cycles Second Affiliated Hospital Zhejiang
B7-H3 CAR T for recurrent or refractory glioblastoma Ningbo Yinzhou People’s Hospital
Huizhou Municipal Central Hospital
BoYuan RunSheng Pharma Co., Ltd.
NCT04385173 I - three intratumoral or Intracerebroventricular injections of B7-H3 CAR T with 1–2 weeks of intervals between cycles of temozolomide treatment Second Affiliated Hospital, School of Medicine, Zhejiang University
Pilot study of B7-H3 CAR T in treating patients with recurrent and refractory GBM BoYuan RunSheng Pharma Co., Ltd.
GD2 NCT04196413 I - rolling-6 dose escalation design test GD2-CAR T cells in subjects with H3K27M-mutant DIPG Crystal Mackall, MD
GD2 CAR T cells in diffuse intrinsic pontine gliomas (DIPG) & spinal diffuse midline glioma (DMG) California Institute for Regenerative Medicine (CIRM)
Cure Search National Institutes of Health (NIH)
CD147 NCT04045847 Early I - intracavity injection Xijing Hospital
CD147-CAR T cells in patients with recurrent malignant glioma - every 7 days for 3 weeks
MMP2 NCT04214392 I - each treatment cycle begins with one or two CAR T cell infusions (one at each catheter site) and lasts for 1 week City of Hope Medical Center, National Cancer Institute (NCI)
CAR T cells with a chlorotoxin tumor-targeting domain for the treatment of MMP2+ recurrent or progressive GBM
2. Targets of CAR T Cell Therapy in Completed and Ongoing Clinical Trials for GBM
2.1 IL13Rα2

GBM tumors exhibit high expression of interleukin 13 receptor alpha 2 (IL13Rα2) in >75% of tumor tissue, with little to no expression in the surrounding normal brain tissue, and was the first target of CAR T cell therapy for GBM [15, 16]. IL13Rα2 has a greater binding affinity for IL-13 than the IL13Rα1 receptor in GBM tumors, effectively sequestering IL-13 from IL13Rα1. In healthy brain tissue, IL-13 binds with the low-affinity monomer and then forms a heterodimer complex with IL4Rα, followed by activation of STAT6 to induce proapoptosis activity [17, 18]. Due to specificity and high affinity for GBM tumors, IL13Rα2 is considered a desirable target in the first CAR T therapy for GBM.

IL13Rα2-specific targeting CAR T cells were first developed in 2004 using an extracellular altered IL-13 domain and CD3ζ intracellular part [19]. The first clinical trial involving IL13Rα2 CAR T cells was conducted in 2015, where IL13Rα2 CAR T cells were directly administered into the post-surgical cavity of three patients with GBM (NCT00730613) [10]. One patient from an ongoing trial reported a 7.5-month regression period with 11 months of median OS without severe harmfulness [10]. Consequently, monitoring the viability and movement of cytotoxic T cells engineered with a herpes simplex virus type 1 thymidine kinase (HSV1-TK) and IL-13 zetakine CAR in glioma cells achieved through safe noninvasive positron emission tomography (PET) imaging, providing potential optimization for immunotherapies in the clinical trials (NCT01082926) [11]. Despite the promising initial results in clinical trials, recurrence is a challenge for IL13Rα2 CAR T cell therapy, and efforts are being made to enhance their efficacy [20]. The next second-generation CARs were developed by adding a costimulatory 4-1BB domain, which is superior to the first-generation CARs [21]. These improved CARs have shown increased persistence and enhanced antitumor activity. Additionally, further advancements, such as the addition of activation-induced IL-15 expression to CAR T cells, can increase the proliferation and cytokine production in CAR T cells, leading to improved performance in treating GBM [22]. Moreover, targeting IL13Rα2-single domain to CD3ε subunit fusion, forming IL13Rα2 ε-TRuC-T cells, induced cytokine release, and lysed tumor cells in a U251-subcutaneous xenograft model [23]. Additionally, ongoing clinical trials evaluate the combination of IL13Rα2-CAR T with immune checkpoint inhibitors, such as nivolumab and ipilimumab (NCT04003649) [24].

2.2 EGFRvIII

Epidermal growth factor receptor (EGFR) is a receptor tyrosine kinase (RTK) commonly amplified and/or mutated in about half of GBM patients. Among EGFR mutations, the most common is the activated EGFR variant III mutant (EGFRvIII), produced by in-frame deletion of exons 2–7 of the coding sequence and found in approximately 45% of GBM patients [25]. EGFRvIII has been an attractive target because of the assumed role in tumorigenesis and because the selective expression on tumor tissue has been well-characterized [26]. In addition, EGFRvIII expression is specific to GBM and is not found in healthy tissues, so EGFRvIII has been assessed for good targets. Although inhibitors targeting EGFRvIII have been used in GBM patients, they have not shown significant clinical effectiveness because of signaling bypass, whereby inhibition of the EGFRvIII pathway activates other RTKs [27]. Consequently, EGFRvIII could be a viable and appealing target for CAR T cell therapy in GBM.

The first clinical trial with EGFRvIII CAR T cells was reported in 2017 that used a single dose of EGFRvIII CAR T cells that was intravenously administered to 10 patients with recurrent GBM (NCT02209376) and which exhibited safety and a restricted antitumor response with downregulated expression of the targeted antigen [12]. In another clinical trial, third-generation EGFRvIII CAR T cells were administered after lymphatic regression chemotherapy and intravenous IL-2 injections, but they did not show clinical success and were well-tolerated without any apparent toxicity (NCT01454596) [13]. Additionally, EGFRvIII CAR T cell preparation using humanized scFv has been proven safe but failed to show clinical benefit [28].

Further analysis of patients receiving a single whole-body dose of the second-generation EGFRvIII targeting CAR T cell formulation was conducted on tumor tissues to assess the antitumor immune response and CAR T cell effectiveness. These CAR T cells were shown to be present in some patient tissues, and increased frequency of autologous non-CAR T cells and immunosuppressive Tregs were also found in treated tumor tissues [12]. Immunohistochemical staining showed the increased expression of other soluble immunosuppressive molecules such as programmed cell death ligand 1 (PD-L1), indoleamine 2, 3-dioxygenase 1 (IDO1), and transforming growth factor-β (TGF-β), indicating that EGFRvIII CAR T cells produced immunosuppressive responses in tumors [29]. Recently, more studies focused on modifying anti-EGFRvIII CAR T cells. PDIA3-mutant EGFRvIII CAR T cells, where PDIA3 had been engineered to be associated with low cell invasiveness, survival, and metastasis, were more efficient in the antigen-specific killing of human GBM cells [30]. A CAR was designed using a mAb806-based binder, recognizing tumor-specific untethered EGFR; the mAb806 epitope is also exposed in the EGFRvIII variant transcript. Treatment with EGFR806-CAR T cells successfully showed an improvement in 50% of U87 orthotopic xenografts surviving to 90 days [31]. Flow cytometry analysis of CAR T infusion products from the first human trial also suggested that PD-1 predicted peripheral engraftment and progression-free survival in recurrent GBM (NCT02209376) [32].

2.3 HER2

Human epidermal growth factor receptor 2 (HER2), also known as ERBB2, is a member of the EGFR family that is overexpressed in approximately 15% of patients with GBM contributing to poor survival rate [33]. Aberrant HER2 signaling deregulates apoptosis and cell proliferation and increases the degree of aplasia in the GBM [34]. Although HER2 is overexpressed in numerous tumors, such as GBM, breast, and ovarian, it can be expressed in some healthy tissues, leading to safety issues.

The first reported use of HER2 CAR T cell therapy in metastatic colon cancer resulted in respiratory distress within 15 min of administration and death five days later [35]. In addition, analysis of serum samples after cell infusion showed elevated levels of cytokines, including interferon-γ (IFNγ), granulocyte macrophage-colony stimulating factor (GM-CSF), tumor necrosis factor-α (TNF-α), IL-6, and IL-10, indicating the occurrence of a cytokine storm [35]. Since then, efforts have been made to address safety concerns associated with CAR T cell therapy. Phase I clinical trials were conducted in 2017 using a second-generation HER2-directed treatment (NCT01109095), that utilized an FRP5-based exodomain and a CD28ζ endodomain-targeting CAR T cells. The intravenously injected HER2 CAR T cells showed that compared with the unmodified virus-specific T cells (VSTs), the HER2 CAR VSTs exerted a significant cytotoxic effect on the HER2-positive glioma cell line U373 and had no significant dose-limited toxicity and only weak side effects; however, two patients had seizures and headaches, which were probably related to the T cell infusion. Detection of circulating HER2 CAR T cells in the blood via qPCR after the final injection showed that more than half of the patient HER2 CAR T cells lasted more than six weeks [14]. In addition to T cells, HER2-specific CAR natural killer (NK) cells derived from the human NK cell line NK-92 specifically lysed GBM-derived cell lines and showed an in vivo antitumor activity in xenografts. Utilizing an NK cell line is currently under evaluation in a phase 1 clinical trial [36].

2.4 NKG2D Ligand

The NK group 2-member D (NKG2D) receptor, a C-type lectin-like homodimeric receptor, is expressed on various immune cells, including NK cells, CD8+ T cells, γδ T cells, and some autoreactive CD4+ T cells [37]. The ligand for NKG2D (NKG2DL) is frequently upregulated in stressed and malignant tissues and overexpressed in tumor cells, but is absent in healthy cells, allowing the immune system to discriminate between cancer cells and healthy tissue. In GBM, patient-derived samples and, interestingly, stem-like cells expressed NKG2DL and released NKG2DL in a soluble form through various mechanisms [38, 39]. Additionally, cyclic EZH2, encoded by EZH2-92aa, has been demonstrated to promote immunological escape from GBM by inhibiting the expression of surface NKG2D ligands [40].

In 2015, CAR T cells targeting NKG2D were first proposed but exhibited significant clinical toxicity in animals [41]. Since then, NKG2D CAR T cells have demonstrated practical anticancer effects in hepatocellular carcinoma, pancreatic, colorectal, and gastrointestinal tube adenocarcinomas [42, 43, 44, 45]. In addition, combined with radiotherapy, CAR T cells that express full-length murine NKG2D have been shown to considerably extend OS in orthotopic xenograft models that use murine GBM cells [46]. Moreover, human NKG2D CAR T cells eradicated human GBM and GBM-stem-like cells in vitro and a subcutaneous tumor model [39]. However, a phase 1 clinical trial for solid tumors, including GBM, aimed at evaluating the safety and clinical effects of NKG2D CAR T cells was withdrawn due to administrative reasons (NCT04270461).

2.5 B7-H3

B7-Homolog 3 protein (B7-H3, also known as CD276) is a member of the B7 family of immunoregulatory proteins that bind to receptors on lymphocytes that regulate the immune response [47]. B7-H3 is a related cell surface protein ligand that is overexpressed in tumor tissues, with high expression in >70% of GBM biopsy samples [48, 49, 50] but is not detectable in healthy brain tissue [51, 52, 53]. Knockdown of B7-H3 has been shown to decrease MYC (MYC proto-oncogene, BHLH transcription factor) expression and inhibit tumor growth in a GBM xenograft mice model [54].

The constructed B7-H3 CAR T cells have been examined for cytotoxic effects on GBM cells since 2019 [51, 55, 56]. In the enhanced B7-H3 CAR T cells, the internal stimulatory structural domains of CD28 or IL-7 receptors were fused to extracellular PD-1 to create PD-1 decoy receptors, which produced a more enduring anticancer activity [55]. In 2022, the results from a human trial of B7-H3 CAR T for multiple basal cell carcinoma demonstrated that intratumoral injection of these cells partially controlled tumor growth in patients with BCC with minor adverse events. However, the capability and safety of B7-H3 CAR T therapy require further investigation in a larger patient cohort [57]. In addition, an ongoing randomized, phase I/II study (NCT04077866) is currently evaluating the safety and performance of B7-H3 CAR T cells in patients with refractory or recurrent GBM between cycles of temozolomide [58]. Another ongoing phase I study (NCT04385173) is evaluating the safety and efficacy of B7-H3 CAR T in patients with GBM who have relapsed on temozolomide or have not responded to standard therapy.

2.6 GD2

Gangliosides (GD) are frequently expressed in normal tissue. However, disialoganglioside (GD2) is a specific type of GD that is mainly expressed in fetal tissue [59] and is also highly expressed in several cancer tissues, including melanoma, retinoblastoma, neuroblastoma, and GBM [60, 61]. In addition, GD2 has been identified as a cancer stem cell marker in breast cancer [62] and with the in vitro neurosphere formation capacity of GBM patient-derived cells [63].

In preclinical studies, CAR T cells targeting GD2 have demonstrated a solid ability to kill neuroblastoma cell lines in vitro as well as in subcutaneous xenografts [64] and patient-derived diffuse midline glioma orthotopic xenografts [65]. Moreover, genetically engineered CAR T cells with a truncated form of anti-GD2 (GD2 tCAR) exhibited enhanced specific targeting [60], and the bifunctional mesenchymal stem cells (MSCs), expressing GD2 tCAR and the proapoptotic agent tumor necrosis factor-related apoptosis-inducing ligand, showed potent antitumor activity against GD2-positive GBM cells. Furthermore, the preclinical efficacy of GD2-directed CAR T cells was recently demonstrated against H3K27M mutant GBM cells, providing a theoretical basis for the first human phase I clinical trial (NCT04196413) [66]. CAR T cells manufactured using a clinical retroviral vector and a GD2-specific CAR encoding an IL-15 transgene further improved tumor control and achieved a striking 50% bioluminescence imaging response rate to intracranial tumors [67].

2.7 CD147

Immunohistochemical analysis was used to evaluate the expression pattern of CD147 in GBM samples from 206 patients and healthy brain tissue samples from 36 individuals [68, 69]. CD147 expression was found to be higher in GBM samples than that in healthy tissues. Increased CD147 expression was related to a worse OS rate in patients with GBM. These findings show that CD147 is abundantly expressed in GBM [69]. CAR T cells and NK cells were transduced with a CD147-specific CAR, which detected the CD147 surface marker and targeted hepatocellular carcinoma (HCC) via a logic-gated (log) GPC3-synNotch mechanism. CD147-CAR did not induce significant targeted/tumor removal toxicity in a human CD147 transgenic mice model but did selectively kill double antigen (GPC3+CD147+)-positive HCC cells while failing to kill the single antigen (GPC3-CD147+)-positive HCC cells [70]. An early phase I clinical trial tested the safety, tolerability, and efficacy of CD147-specific CAR T cell therapy in patients with recurrent GBM (NCT04045847).

2.8 Chlorotoxin

The primary chlorotoxin (CLTX) can interact with isoforms of matrix metalloproteinase-2 (MMP-2) that were identified on the surface of GBM cells and are significantly elevated in GBM and associated malignancies [71]. CLTX peptide has the ability to bind to GBM specifically, and when combined with CAR T cells, it results in a potent anti-GBM activity that can target tumors that do not express other GBM-associated antigens. Importantly, CLTX CAR T cells were observed to specifically target cells expressing MMP-2 without showing any off-target effector activity against healthy cells or after transfer to mice [72]. A phase I study was initiated to treat MMP-2-positive relapsed or progressing GBM with T cells carrying CLTX CARs (NCT04214392).

3. Challenges and Solutions for CAR T Cell Therapy for GBM
3.1 Effective Targeting of Tumor Cells

Identifying tumor-specific antigens is a crucial factor in the development of CAR T cell therapy however, it is also challenging due to the heterogeneity of tumors, lack of tumor-specific antigens, and expression of shared target candidates in important healthy tissue, which increases the risk of off-target tumor toxicity.

Tumor antigens are classified into tumor-associated antigens (TAA) and tumor-specific antigens (TSA). Both tumor cells and healthy tissue cells have a high expression level of TAAs, albeit at a low level. Only tumor cells express TSAs, which makes them a perfect target for antigens [73]. As the discovery and screening of TSAs are challenging, the most used recognition target of CAR is TAA, so the diverse expression of TAA in different kinds of tumor cells may affect the recognition of cancer cells by CAR T cells and reduce the effect of CAR T therapy. Numerous potential targets have therefore been identified and yielded encouraging results in the preclinical setting (Table 3, Ref. [74, 75, 76, 77, 78, 79, 80, 81, 82]). Antigens such as carbonic anhydrase IX (CAIX) [74, 75], CSPG4 [76, 77], or TRPO [78, 79] that are primarily investigated in non-CNS tumors are upregulated in GBM and have therefore been considered as potential targets. Expression of antigens as stem-like cell surface markers, including CD133 [83], CD70 [80, 84], and EphA2 [81, 82], is enhanced by alkylating chemotherapy or radiotherapy typically provided as first-line therapy in GBM, thus making these antigens promising in treating recurrent disease. Other targets that have been primarily selected because of their relatively high expression in GBM tissue include heparan sulfate proteoglycan glypican-2 (GPC2) [85, 86] and TWEAK receptor Fn14 [87].

Table 3.Preclinical study of CAR T therapy for GBM.
Molecular target Characteristics Research advances Ref.
CAIX - a surrogate marker for hypoxia - in GBM mice model, direct intratumoral injection of anti-CAIX CAR T cells is effective with limiting side effects [74, 75]
- particularly upregulated in GBM
- associated with a poor patient outcome and survival rate
CSPG4 - related to cell proliferation and migration in vitro and in vivo - to prevent the emergence of antigen-negative clones, it is worth exploring strategies that combine CSPG4-CAR T cells with CAR T cells targeting IL13α2R, EGFRvIII, or HER2 [76, 77]
- as a marker for GBM stem cells (GSCs)
TROP2 - play a role in tumor growth, invasion, and metastasis of various solid tumors - CD27-based TROP2 CAR T cells showed higher antitumor activity, upregulating IL-7Rα expression while downregulating PD-1 expression [78, 79]
- highly expressed in GBM, but low expressed in normal brain
CD70 - type II transmembrane protein binding to CD27 - CD70 CAR T cells effectively induced anti-tumor responses against CD70+ gliomas both in vitro and in vivo [80]
- expressed on activated T cells and mature DCs
- overexpressed in primary and recurrent GBM, but not on tumor infiltrating T cells - no toxicity in xenograft and syngeneic models
- associated with poor survival in GBM
EphA2 - Eph family of RTKs - the third-generation of EphA2 CAR T cells obviously improved survival rate in tumor bearing mice in vivo, showing better antitumor activity [81, 82]
- associated with tumorigenesis, angiogenesis, invasion and metastasis
- highly expressed in GBM but only at low levels in normal brain tissue

Several approaches for overcoming the heterogeneity in tumor antigens highlight the complexity of developing effective CAR T cell therapies for GBM, thus increasing tumor cell subpopulations while targeting a single antigen. One is a simultaneous administration of each CAR T cell targeting a different GBM antigen. The other is dual-targeting, using CAR T cells expressing two scFvs to simplify the number of injections. However, clinical treatment with bispecific CARs or T cells expressing two CARs, such as in CD19/CD22- and CD19/CD20-targeting approaches, showed limited results in leukemia. Another method is engineering CARs, whose antigen-binding structural domain consists of a ligand binding various different antigens (called a universal CAR); the representative example is the T1E peptide that attaches to the EGFR family of receptors for solid tumors [88]. Furthermore, bispecific T cell engager (BiTE) expressed in T cells provides an opportunity to target multiple antigens and redirect bystander T cells to tumor cells. In addition to targeting various antigens, efforts to activate bystander T cells to identify tumor cells are being pursued, including the transgenic expression of cytokines (e.g., IL-18 or IL-36γ) or CD40L [89]. Likewise, engineered T cells that secrete the FLT3 ligand have been shown to promote epitope spreading [90], and intratumoral administration of anti-Fn14xCD3 BiTE has demonstrated antitumor effects in mouse GBM models [87].

As a switch-controlled CAR T, the synthetic notch (synNotch) CARs can induce the expression of a second CAR upon activation by an antigen-specific synNotch [91]. SynNotch-GD2-B7H3 [92] and SynNotch-HER2 [93] were designed to express the kill switch consistently. Recently, dual administration of synNotch CARs, (1) EGFRvIII-SynNotch CAR and EphA2 or IL13Rα2 tandem CAR [94] and (2) ALPPL2-synNotch CAR and melanoma cell adhesion molecule (MCAM), mesothelin, or HER2 CAR [95], can conserve energy by entering standby mode when not in direct contact with antigens. Moreover, drug-regulated CARs using FDA-approved molecules showed controlled effects and limited toxicity. The recent representative examples exhibit signal neutralization by an inhibitable protease CAR and versatile protease regulatable CAR [96]. Their activity levels remain stable during the entire treatment process, which overcomes the challenge of T cell depletion that has hampered conventional CAR T therapy.

3.2 Enhancing Tumor-Homing (Trafficking) and Penetration

CAR T cell therapy is a promising modality for cancer, but its effectiveness is often limited by poor infiltration of CAR T cells into tumors and the presence of a highly immunosuppressive tumor microenvironment (TME). The GBM TME consists of stromal cells, immune cells along with enzymes, growth factors, inflammatory cytokines, and pathogenic stimuli within the extracellular matrix (ECM), as well as highly proliferative malignant astrocytoma cells and likely cancer stem cells [97], all of which form discrete niches inside the tumor [98]. All these individual components of the TME can act as a physical barrier that limits the effectiveness of CAR T therapy. As a result, numerous studies have focused on developing strategies to improve the ability of infused CAR T cells to penetrate and traffic into the TME to enhance their antitumor activity. Overall, these approaches aim to overcome the challenges posed by the TME and improve the therapeutic potential of CAR T cell therapy.

In tumor evasion of immune responses, chemokine ligands, and receptors play different roles. However, low expression of chemokine receptors can reduce immune cell infiltration, contributing to tumor evasion of immune responses [99]. In recent preclinical studies, a strategy to improve the tumor-homing and antitumor efficacy of CAR T cells has been explored by overexpressing CCR2 (ligand; CCL2, CCL7, CCL8, CCL13, and CCL16) [100], CCR4 (ligand; CCL4, CCL3, CCL5, CCL17, and CCL22) [101], and CCR6 (ligand CCL20) [102, 103]. Thus, these approaches can attract CAR T cells to tumors with high expression of specific chemokines, leading to adequate tumor clearance. In addition, in parallel to the development of technologies to enable real-time imaging of CAR T cells in vivo allows for non-invasive monitoring of CAR T cell trafficking to tumor cells and facilitates the examination of novel interventions to enhance CAR T cell effects [104]. Moreover, combined treatment of oncolytic adenovirus and chemokine CXCL11 increased CAR T cell infiltration and reprogramed the immunosuppressive TME, thereby improving therapeutic efficacy [105]. These advances hold promise for improving the efficacy of CAR T cell therapy against solid tumors by enhancing their ability to penetrate the tumor microenvironment and target cancer cells.

In the case of GBM tumors, another physical barrier that hampers CAR T cell trafficking is the BBB. The BBB is a highly selective semipermeable boundary that separates circulating blood from the extracellular fluid in the brain and CNS. The BBB consists of endothelial cells of the capillary wall, astrocytes that wrap around the capillaries, and pericytes embedded in the basement membrane of the capillaries [105]. The BBB acts as a protective barrier to prevent the entry of potentially harmful substances, including immune cells, as well as CAR T cells. Therefore, strategies to enhance the ability of CAR T cells to cross the BBB and reach GBM cells are needed to improve the efficacy of CAR T cell therapy against GBM. CAR T cells can be transferred across the BBB in association with constitutively or inducible expressed cytokines, chemokines, or other molecules and can be actively utilized for CAR design. In addition, using in vitro BBB models, the iPSC-derived brain endothelial-like cell trans-well coculture model can be a valuable tool for deciphering the mechanism of CAR T-induced BBB destruction, concomitant toxicity, and effector functions of target cells after BBB [106].

Furthermore, another aspect of enhancing CAR T cell trafficking is the route of CAR T infusion. Although most studies of CAR T cell therapy have used intravenous injection, a more site-specific delivery method may be beneficial in the therapy of solid tumors, including those of breast, pleura, and liver [107] as well as brain [21, 66], by bypassing circumventing multiple barriers to CAR T cell trafficking as well as restricting on-target, off-tumor toxicity. In HER2 CAR T cell treatment for breast cancer brain metastasis, local intracranial administration demonstrated in vivo antitumor potency using three administration routes: intravenous, local intratumoral, and regional intraventricular [108]. Similarly, another study on the delivery route and therapeutic strength demonstrated that local intracranial local delivery of IL13Rα2-targeted CAR T cells induces improved anticancer effectiveness than intravenous route, with intraventricular infusions showing possible benefit over intracranial tumor infusions in a multifocal disease model [21]. Clinical studies have explored the local delivery approaches for CAR T cell therapy in GBM. In one study, a patient, who received CAR T cells transduced with anti-mesothelin mRNA by both intravenous and regional infusions, exhibited an evoked antitumor somatic response relating to multiple additional antigens [109]. In another clinical study, the first four patients with diffuse intrinsic pontine glioma/ H3K27M-mutated diffuse midline gliomas were treated with GD2-CAR T cells; those who were subsequently administered intracerebroventricularly exhibited clinical benefit [66]. These early results suggest that the ventricular infusion-related approach for CAR T therapy has the potential for treating GBM.

3.3 Proliferative Activity against CAR T Cell Exhaustion

CAR T cell exhaustion is a phenomenon that arises from prolonged exposure to antigen stimulation and an immunosuppressive TME, including various suppressive immune cells such as Treg cells, NK cells, myeloid-derived suppressor cells (MDSCs), and tumor-associated macrophages (TAMs). Maintaining CAR T cell effector function, persistence, and overcoming exhaustion remains a major challenge in achieving clinical potency [110]. Exhausted T cells have a low proliferation rate in response to antigenic stimuli (e.g., the response to antigenic stimuli), progressive loss of effector functions (e.g., reduced cytokine production and weakened killing effects), expression of multiple inhibitory receptors (e.g., PD-1, Tim3, and LAG3). Furthermore, altered metabolism from oxidative phosphorylation to glycolysis is also a feature of T cell exhaustion [111]. The comparison between monocyte-derived T cells of the dysfunctional and functional remission groups revealed a significant difference in the proportion of CD8+ T cells expressing PD-1 and LAG3 (lymphocyte activation gene 3, CD223). The findings imply that the initial cell phenotype before infusion may strongly correlate with treatment failure and short-term [112]. It has been demonstrated that introducing a dominant-negative PD-1 receptor can enhance the functional persistence of CAR T and prevent exhaustion [113]. CAR T cells targeting human CAIX and engineered to secrete human anti-PD-L1 antibodies exhibited significantly enhanced activity, elevated cytokine production, and greater uptake by immune cells, leading to a substantial reduction in tumor growth of clear cell renal cell carcinoma [114].

Treg cells negatively affect suppressor cytotoxic T cells through multiple mechanisms, including the secretion of immunosuppressive cytokines, competitively consuming IL-2, and the inhibition of antigen-presenting cells and obstructing T cell activation by cytotoxic T lymphocyte antigen 4 (CTLA4) [115]. To overcome the immunosuppressive effects of Treg cells, one group generated a BiTE that induced the expression of CAR targeting EGFRvIII and allowed the conversion of Treg cells into cytotoxic T cells [116]. In the GBM model, combining GD2-targeted CAR T cells and NKG2D CAR NK cells, NKG2D CAR NK cells selectively allowed MDSCs and restored impaired CAR T cell activity [117]. In addition, targeting MDSC induced the release of proinflammatory cytokines and chemokines, resulting in tumor regression and increased survival, compared to CAR T cell therapy alone [117]. Moreover, the inhibition of colony-stimulating factor 1 receptor (CSF1R) to deplete TAM and targeted ablation of PD-1 in myeloid progenitor cells revealed more potent antitumor activity and tumor growth suppression [118]. TAMs, the most abundant immune-infiltrating cells in the TME, prevent T cell-mediated antitumor immunity by boosting the secretion and recruitment of cytokines and enzymes, such as arginase one or IDO1, and by increasing the recruitment of Treg cells [119, 120]. In animal models, the survival rate was notably increased by CAR targeting CD70, presented on the putative M1 macrophages and CD4+ T cells. Therefore, targeting the CD70/CD27 axis could be an effective poly therapeutic strategy for treating both the GBM and the immune TEM [84].

Another potential approach is redirecting or circumventing the endogenous response to suppressive cytokines and inhibitory signaling pathways. Immunosuppressive cytokines, such as TGF-β, VEGF, IL-4, and IL-10, are frequently found in TME and can block T cell effector function directly and attract suppressor cell types. In addition, recruited immune cells can enhance the tolerogenic nature of the TME by producing immunosuppressive cytokines themselves [99]. Furthermore, TGF-β directly impacts the expression of chemokine receptors such as CXCR3 on T cells and blocks infiltration of T cells in the colorectal cancer model [121]. In solid tumor xenografts, knocking down TGF-βRII by CRIPSR-Cas9 in CAR T cells has been shown to enhance efficacy, reduce exhaustion, and prevent polarization to Treg cells [122]. Recently, phase 1 trials have reported that PSMA CAR T cells were optimally engineered with a TGFβRII to block TGF-β signaling, led to expansion in blood and tumor trafficking and improved antitumor immunity in patients with metastatic castration-resistant prostate cancer [123]. In a recent study, investigators conducted a genome-wide CRISPR knockout screen to identify the critical role of the interferon-γ receptor (IFNγR) signaling in the susceptibility of solid tumors to CAR T cells [124]. These findings underscore the potential significance of the IFNγR signaling pathway (IFNGR1, JAK1, or JAK2) in developing and optimizing CAR T cell therapies for solid tumors [124].

Engineering CAR T to secrete proinflammatory soluble factors can modify the TME, leading to an antitumor response. CAR T engineered to overexpress the IL-12β p40 subunit can stimulate the production of IL-23 by activated T cells, which promotes proliferation via STAT3 signaling [125]. In addition, EGFRvIII CAR T therapy combined with a single locally administered dose of IL-12 allowed for a durable antitumor response by promoting infiltration of proinflammatory CD4+ T cells, decreasing the number of Treg cells, and activating myeloid compartments [126]. In two recent studies, CD19-CAR T cells engineered to coexpress human orthogonal IL-2 and its receptor drove a complete response in lymphoma, even with substantially reduced CAR T cell doses, by selectively expanding and activating CAR T cells in vivo [127, 128]. In addition, cancer-derived extracellular vesicles have been investigated to induce phenotypical, functional, and transcriptional dysfunction of CAR T cells [129]. A recent report described a strategy that exploits the cooperate relationship between the transcription factors nuclear factor of activated T cells (NFAT) and activator protein 1 (AP-1) to improve the effector functions of T cells was reported. However, NFAT in the absence of AP-1 can led to negative feedback on T cell exhaustion. To address this issue, CAR T cells overexpressing a NEAT, basic leucine zipper ATF-like transcription factor (BATF), promoted the survival and expansion of tumor-infiltrating CAR T cells by interferon regulatory factor 4 (IRF4) [130]. Collectively, these approaches could potentially improve the efficacy of CAR T cell therapy by preventing T cell exhaustion and enhancing the persistence of CAR T cells within the TME.

4. Conclusions and Perspectives

CAR T cell holds a great promising therapeutic strategy for GBM, however, there are still several challenges that need to be considered. One of the major challenges is selecting a target antigen that can selectively eliminate cancer cells while sparing healthy brain tissue. Another challenge is the immune escape mechanisms applied to cancer cells, which can limit the effectiveness of CAR T cells. This includes downregulating the target antigen, inhibiting T cell function, and recruitment of immunosuppressive cells. Furthermore, optimizing CAR T cell delivery to the brain is another significant hurdle to overcome. Several approaches, such as local delivery, disruption of the BBB, and modification of CAR T cells, are being explored to improve CAR T cell delivery to the brain. In addition, developing combinatorial strategies for enhancing the CAR T cells’ efficacy and stimulating the innate immune system is essential. Combining CAR T cells with checkpoint inhibitors, vaccines, and other immunomodulatory agents is an approach that can enhance the activity of CAR T cells and solve immune escape mechanisms.

In the latest research, UniVec-CAR is a new approach that utilizes a single lentiviral vector that expresses accessory molecules to induce the production of an immunostimulatory cytokine, an antibody that ameliorated cytokine release syndrome, and regulatory T cell transcription factors [131]. Meanwhile, several recently published T cell immunization approaches can be actively utilized in CAR T research [9]. For example, after cisplatin chemotherapy, the production of CCL20 and IL-1β at the tumor site induced the recruitment and activation of innate lymphoid cells (ILC3) in the tumor and enhanced antitumor immunity [132]. In addition, using approved pharmacological inhibitors of cyclin-dependent kinases 4 and 6 (CDK4/6) promoted the phenotypic and functional acquisition of immune T cell memory, which is essential for the success of CAR T cell therapy. This approach has been incorporated into clinical trial therapy design to prevent the depletion of CAR T cells [133]. Accordingly, persevering and meticulous investigation is imperative to optimize the potency of CAR T cells and address the challenges associated with this new therapeutic strategy.

Additionally, beyond the conventional CAR T cells, CAR natural killer (CAR NK) and CAR macrophage (CAR M) are potential alternative approaches to CAR T cell therapy for GBM treatment, as they address some of their limitations [134]. Unlike CAR T cells, CAR NK cells do not require prior sensitization to tumor antigens and can identify and eliminate cancer cells without requiring major histocompatibility complex (MHC) matching. Furthermore, they have a low risk of causing graft versus host disease (GvHD) and CRS. A novel cytotoxic CAR NK cells from cord blood by transducing them with an anti-CD19 vector engineered to produce IL-15 for increased efficacy [135]. CAR M therapy involves genetically modifying macrophages to express a CAR on their surface, allowing them to target and kill cancer cells. Similar to CAR NK cells, they do not require MHC matching and are less likely to cause GvHD and CRS. The first published CAR M, which designed antigen-specific phagocytosis and activation of T cells, led to decreased tumor burden in mouse models and enhanced antitumor response in humanized mice, along with induction of a proinflammatory tumor microenvironment [136]. Although CAR NK and CAR M therapies offer several advantages over CAR T cell therapy, they are still in the early stages of development and clinical testing, and their effectiveness in treating GBM and overcoming the limitations of CAR T cell therapy is yet to be determined, and further research is needed.

Abbreviations

ALPPL2, alkaline phosphatase placental-like 2; AP-1, activator protein 1; B7-H3, B7-Homolog 3 protein; BATF, basic leucine zipper ATF-like transcription factor; BBB, blood-brain barrier; BiTE, bispecific T cell engager; CAIX, carbonic anhydrase IX; CAR, chimeric antigen receptor; CDK4/6, cyclin-dependent kinases 4 and 6; CLTX, chlorotoxin; synNotch, synthetic notch; CNS, central nervous system; CSF1R, colony stimulating factor 1 receptor; CTLA4, cytotoxic T lymphocyte antigen 4; ECM, extracellular matrix; EGFR, epidermal growth factor receptor; EGFRvIII, EGFR variant III; GBM, glioblastoma; GPC2, heparan sulfate proteoglycan glypican-2; HER2, human epidermal growth factor receptor 2; HCC, hepatocellular carcinoma; IDO1, indoleamine 2, 3-dioxygenase 1; IFNγ, interferon γ; IFNγR, interferon-γ receptor; I-L13Rα2, interleukin-13 receptor alpha 2; ILC3s, innate lymphoid cells; IRF4, interferon regulatory factor 4; LAG3, lymphocyte activation gene 3; MDSCs, myeloid-derived suppressor cells; NFAT, nuclear factor of activated T cells; NK, natural killer; NKG2D, natural killer group 2 member D; NKG2DL, NKG2D ligands; OS, overall survival; RTK, receptor tyrosine kinase; TAA, tumor-associated antigens; TCR, T cell receptors; TGF-β, transforming growth factor-β; TME, tumor microenvironment; GM-CSF, granulocyte macrophage-colony stimulating factor; TNF-α, tumor necrosis factor-α; VSTs, virus-specific T cells; Treg, regulatory T cells; TSA, tumor-specific antigens; PD-L1, programmed cell death ligand 1.

Author Contributions

SZ: Investigation, Writing – original draft, Writing – review & editing. HS: Investigation, Writing – original draft, Writing – review & editing. SIC: Conceptualization, Investigation, Visualization, Writing – original draft, Writing – review & editing. JY: Conceptualization, Supervision, Funding acquisition, Writing – review & editing. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 82173228.

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
McKinnon C, Nandhabalan M, Murray SA, Plaha P. Glioblastoma: clinical presentation, diagnosis, and management. BMJ (Clinical Research Ed.). 2021; 374: n1560.
[2]
Louis DN, Perry A, Wesseling P, Brat DJ, Cree IA, Figarella-Branger D, et al. The 2021 WHO Classification of Tumors of the Central Nervous System: a summary. Neuro-oncology. 2021; 23: 1231–1251.
[3]
Ostrom QT, Patil N, Cioffi G, Waite K, Kruchko C, Barnholtz-Sloan JS. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2013-2017. Neuro-oncology. 2020; 22: iv1–iv96.
[4]
Rosenberg SA, Packard BS, Aebersold PM, Solomon D, Topalian SL, Toy ST, et al. Use of tumor-infiltrating lymphocytes and interleukin-2 in the immunotherapy of patients with metastatic melanoma. A preliminary report. The New England Journal of Medicine. 1988; 319: 1676–1680.
[5]
Huang B, Li X, Li Y, Zhang J, Zong Z, Zhang H. Current Immunotherapies for Glioblastoma Multiforme. Frontiers in Immunology. 2021; 11: 603911.
[6]
Charrot S, Hallam S. CAR-T Cells: Future Perspectives. HemaSphere. 2019; 3: e188.
[7]
Turtle CJ, Hanafi LA, Berger C, Gooley TA, Cherian S, Hudecek M, et al. CD19 CAR-T cells of defined CD4+:CD8+ composition in adult B cell ALL patients. The Journal of Clinical Investigation. 2016; 126: 2123–2138.
[8]
Abramson JS, Palomba ML, Gordon LI, Lunning MA, Wang M, Arnason J, et al. Lisocabtagene maraleucel for patients with relapsed or refractory large B-cell lymphomas (TRANSCEND NHL 001): a multicentre seamless design study. Lancet (London, England). 2020; 396: 839–852.
[9]
Choi SI, Yin J. Prospective approaches to enhancing CAR T cell therapy for glioblastoma. Frontiers in Immunology. 2022; 13: 1008751.
[10]
Brown CE, Badie B, Barish ME, Weng L, Ostberg JR, Chang WC, et al. Bioactivity and Safety of IL13Rα2-Redirected Chimeric Antigen Receptor CD8+ T Cells in Patients with Recurrent Glioblastoma. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2015; 21: 4062–4072.
[11]
Keu KV, Witney TH, Yaghoubi S, Rosenberg J, Kurien A, Magnusson R, et al. Reporter gene imaging of targeted T cell immunotherapy in recurrent glioma. Science Translational Medicine. 2017; 9: eaag2196.
[12]
O’Rourke DM, Nasrallah MP, Desai A, Melenhorst JJ, Mansfield K, Morrissette JJD, 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. Science Translational Medicine. 2017; 9: eaaa0984.
[13]
Goff SL, Morgan RA, Yang JC, Sherry RM, Robbins PF, Restifo NP, et al. Pilot Trial of Adoptive Transfer of Chimeric Antigen Receptor-transduced T Cells Targeting EGFRvIII in Patients With Glioblastoma. Journal of Immunotherapy (Hagerstown, Md.: 1997). 2019; 42: 126–135.
[14]
Ahmed N, Brawley V, Hegde M, Bielamowicz K, Kalra M, Landi D, et al. HER2-Specific Chimeric Antigen Receptor-Modified Virus-Specific T Cells for Progressive Glioblastoma: A Phase 1 Dose-Escalation Trial. JAMA Oncology. 2017; 3: 1094–1101.
[15]
Brown CE, Warden CD, Starr R, Deng X, Badie B, Yuan YC, et al. Glioma IL13Rα2 is associated with mesenchymal signature gene expression and poor patient prognosis. PloS one. 2013; 8: e77769.
[16]
Brown CE, Alizadeh D, Starr R, Weng L, Wagner JR, Naranjo A, et al. Regression of Glioblastoma after Chimeric Antigen Receptor T-Cell Therapy. The New England Journal of Medicine. 2016; 375: 2561–2569.
[17]
Thaci B, Brown CE, Binello E, Werbaneth K, Sampath P, Sengupta S. Significance of interleukin-13 receptor alpha 2-targeted glioblastoma therapy. Neuro-oncology. 2014; 16: 1304–1312.
[18]
Sattiraju A, Solingapuram Sai KK, Xuan A, Pandya DN, Almaguel FG, Wadas TJ, et al. IL13RA2 targeted alpha particle therapy against glioblastomas. Oncotarget. 2017; 8: 42997–43007.
[19]
Kilian M, Bunse T, Wick W, Platten M, Bunse L. Genetically Modified Cellular Therapies for Malignant Gliomas. International Journal of Molecular Sciences. 2021; 22: 12810.
[20]
Krenciute G, Prinzing BL, Yi Z, Wu MF, Liu H, Dotti G, et al. Transgenic Expression of IL15 Improves Antiglioma Activity of IL13Rα2-CAR T Cells but Results in Antigen Loss Variants. Cancer Immunology Research. 2017; 5: 571–581.
[21]
Brown CE, Aguilar B, Starr R, Yang X, Chang WC, Weng L, et al. Optimization of IL13Rα2-Targeted Chimeric Antigen Receptor T Cells for Improved Anti-tumor Efficacy against Glioblastoma. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2018; 26: 31–44.
[22]
Zannikou M, Duffy JT, Levine RN, Seblani M, Liu Q, Presser A, et al. IL15 modification enables CAR T cells to act as a dual targeting agent against tumor cells and myeloid-derived suppressor cells in GBM. Journal for Immunotherapy of Cancer. 2023; 11: e006239.
[23]
Baeuerle PA, Ding J, Patel E, Thorausch N, Horton H, Gierut J, et al. Synthetic TRuC receptors engaging the complete T cell receptor for potent anti-tumor response. Nature Communications. 2019; 10: 2087.
[24]
Knudson KM, Hwang S, McCann MS, Joshi BH, Husain SR, Puri RK. Recent Advances in IL-13Rα2-Directed Cancer Immunotherapy. Frontiers in Immunology. 2022; 13: 878365.
[25]
Cancer Genome Atlas Research Network. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature. 2008; 455: 1061–1068.
[26]
An Z, Aksoy O, Zheng T, Fan QW, Weiss WA. Epidermal growth factor receptor and EGFRvIII in glioblastoma: signaling pathways and targeted therapies. Oncogene. 2018; 37: 1561–1575.
[27]
Greenall SA, Donoghue JF, Van Sinderen M, Dubljevic V, Budiman S, Devlin M, et al. EGFRvIII-mediated transactivation of receptor tyrosine kinases in glioma: mechanism and therapeutic implications. Oncogene. 2015; 34: 5277–5287.
[28]
Sampson JH, Choi BD, Sanchez-Perez L, Suryadevara CM, Snyder DJ, Flores CT, et al. EGFRvIII mCAR-modified T-cell therapy cures mice with established intracerebral glioma and generates host immunity against tumor-antigen loss. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2014; 20: 972–984.
[29]
Feldman L, Brown C, Badie B. Chimeric Antigen Receptor (CAR) T Cell Therapy for Glioblastoma. Neuromolecular Medicine. 2022; 24: 35–40.
[30]
Ye L, Park JJ, Dong MB, Yang Q, Chow RD, Peng L, et al. In vivo CRISPR screening in CD8 T cells with AAV-Sleeping Beauty hybrid vectors identifies membrane targets for improving immunotherapy for glioblastoma. Nature Biotechnology. 2019; 37: 1302–1313.
[31]
Ravanpay AC, Gust J, Johnson AJ, Rolczynski LS, Cecchini M, Chang CA, et al. EGFR806-CAR T cells selectively target a tumor-restricted EGFR epitope in glioblastoma. Oncotarget. 2019; 10: 7080–7095.
[32]
Tang OY, Tian L, Yoder T, Xu R, Kulikovskaya I, Gupta M, et al. PD1 Expression in EGFRvIII-Directed CAR T Cell Infusion Product for Glioblastoma Is Associated with Clinical Response. Frontiers in Immunology. 2022; 13: 872756.
[33]
Koka V, Potti A, Forseen SE, Pervez H, Fraiman GN, Koch M, et al. Role of Her-2/neu overexpression and clinical determinants of early mortality in glioblastoma multiforme. American Journal of Clinical Oncology. 2003; 26: 332–335.
[34]
Ahmed N, Salsman VS, Kew Y, Shaffer D, Powell S, Zhang YJ, et al. HER2-specific T cells target primary glioblastoma stem cells and induce regression of autologous experimental tumors. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2010; 16: 474–485.
[35]
Morgan RA, Yang JC, Kitano M, Dudley ME, Laurencot CM, Rosenberg SA. Case report of a serious adverse event following the administration of T cells transduced with a chimeric antigen receptor recognizing ERBB2. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2010; 18: 843–851.
[36]
Zhang C, Burger MC, Jennewein L, Genßler S, Schönfeld K, Zeiner P, et al. ErbB2/HER2-Specific NK Cells for Targeted Therapy of Glioblastoma. Journal of the National Cancer Institute. 2016; 108: djv375.
[37]
Dhar P, Wu JD. NKG2D and its ligands in cancer. Current Opinion in Immunology. 2018; 51: 55–61.
[38]
Flüh C, Chitadze G, Adamski V, Hattermann K, Synowitz M, Kabelitz D, et al. NKG2D ligands in glioma stem-like cells: expression in situ and in vitro. Histochemistry and Cell Biology. 2018; 149: 219–233.
[39]
Yang D, Sun B, Dai H, Li W, Shi L, Zhang P, et al. T cells expressing NKG2D chimeric antigen receptors efficiently eliminate glioblastoma and cancer stem cells. Journal for Immunotherapy of Cancer. 2019; 7: 171.
[40]
Zhong J, Yang X, Chen J, He K, Gao X, Wu X, et al. Circular EZH2-encoded EZH2-92aa mediates immune evasion in glioblastoma via inhibition of surface NKG2D ligands. Nature Communications. 2022; 13: 4795.
[41]
VanSeggelen H, Hammill JA, Dvorkin-Gheva A, Tantalo DGM, Kwiecien JM, Denisova GF, et al. T Cells Engineered With Chimeric Antigen Receptors Targeting NKG2D Ligands Display Lethal Toxicity in Mice. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2015; 23: 1600–1610.
[42]
Sun B, Yang D, Dai H, Liu X, Jia R, Cui X, et al. Eradication of Hepatocellular Carcinoma by NKG2D-Based CAR-T Cells. Cancer Immunology Research. 2019; 7: 1813–1823.
[43]
Tao K, He M, Tao F, Xu G, Ye M, Zheng Y, et al. Development of NKG2D-based chimeric antigen receptor-T cells for gastric cancer treatment. Cancer Chemotherapy and Pharmacology. 2018; 82: 815–827.
[44]
Deng X, Gao F, Li N, Li Q, Zhou Y, Yang T, et al. Antitumor activity of NKG2D CAR-T cells against human colorectal cancer cells in vitro and in vivo. American Journal of Cancer Research. 2019; 9: 945–958.
[45]
Wang J, Liu X, Ji J, Luo J, Zhao Y, Zhou X, et al. Orthotopic and Heterotopic Murine Models of Pancreatic Cancer Exhibit Different Immunological Microenvironments and Different Responses to Immunotherapy. Frontiers in Immunology. 2022; 13: 863346.
[46]
Weiss T, Schneider H, Silginer M, Steinle A, Pruschy M, Polić B, et al. NKG2D-Dependent Antitumor Effects of Chemotherapy and Radiotherapy against Glioblastoma. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2018; 24: 882–895.
[47]
Collins M, Ling V, Carreno BM. The B7 family of immune-regulatory ligands. Genome Biology. 2005; 6: 223.
[48]
Zhou WT, Jin WL. B7-H3/CD276: An Emerging Cancer Immunotherapy. Frontiers in Immunology. 2021; 12: 701006.
[49]
Picarda E, Ohaegbulam KC, Zang X. Molecular Pathways: Targeting B7-H3 (CD276) for Human Cancer Immunotherapy. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2016; 22: 3425–3431.
[50]
Zhang C, Zhang Z, Li F, Shen Z, Qiao Y, Li L, et al. Large-scale analysis reveals the specific clinical and immune features of B7-H3 in glioma. Oncoimmunology. 2018; 7: e1461304.
[51]
Nehama D, Di Ianni N, Musio S, Du H, Patané M, Pollo B, et al. B7-H3-redirected chimeric antigen receptor T cells target glioblastoma and neurospheres. EBioMedicine. 2019; 47: 33–43.
[52]
Du H, Hirabayashi K, Ahn S, Kren NP, Montgomery SA, Wang X, et al. Antitumor Responses in the Absence of Toxicity in Solid Tumors by Targeting B7-H3 via Chimeric Antigen Receptor T Cells. Cancer Cell. 2019; 35: 221–237.e8.
[53]
Majzner RG, Theruvath JL, Nellan A, Heitzeneder S, Cui Y, Mount CW, et al. CAR T Cells Targeting B7-H3, a Pan-Cancer Antigen, Demonstrate Potent Preclinical Activity Against Pediatric Solid Tumors and Brain Tumors. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2019; 25: 2560–2574.
[54]
Zhang J, Wang J, Marzese DM, Wang X, Yang Z, Li C, et al. B7H3 regulates differentiation and serves as a potential biomarker and theranostic target for human glioblastoma. Laboratory Investigation; a Journal of Technical Methods and Pathology. 2019; 99: 1117–1129.
[55]
Huang B, Luo L, Wang J, He B, Feng R, Xian N, et al. B7-H3 specific T cells with chimeric antigen receptor and decoy PD-1 receptors eradicate established solid human tumors in mouse models. Oncoimmunology. 2019; 9: 1684127.
[56]
Tang X, Zhao S, Zhang Y, Wang Y, Zhang Z, Yang M, et al. B7-H3 as a Novel CAR-T Therapeutic Target for Glioblastoma. Molecular Therapy Oncolytics. 2019; 14: 279–287.
[57]
Hu G, Li G, Wen W, Ding W, Zhou Z, Zheng Y, et al. Case report: B7-H3 CAR-T therapy partially controls tumor growth in a basal cell carcinoma patient. Frontiers in Oncology. 2022; 12: 956593.
[58]
Kontos F, Michelakos T, Kurokawa T, Sadagopan A, Schwab JH, Ferrone CR, et al. B7-H3: An Attractive Target for Antibody-based Immunotherapy. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2021; 27: 1227–1235.
[59]
Yanagisawa M, Yoshimura S, Yu RK. Expression of GD2 and GD3 gangliosides in human embryonic neural stem cells. ASN Neuro. 2011; 3: e00054.
[60]
Golinelli G, Grisendi G, Prapa M, Bestagno M, Spano C, Rossignoli F, et al. Targeting GD2-positive glioblastoma by chimeric antigen receptor empowered mesenchymal progenitors. Cancer Gene Therapy. 2020; 27: 558–570.
[61]
Nazha B, Inal C, Owonikoko TK. Disialoganglioside GD2 Expression in Solid Tumors and Role as a Target for Cancer Therapy. Frontiers in Oncology. 2020; 10: 1000.
[62]
Battula VL, Shi Y, Evans KW, Wang RY, Spaeth EL, Jacamo RO, et al. Ganglioside GD2 identifies breast cancer stem cells and promotes tumorigenesis. The Journal of Clinical Investigation. 2012; 122: 2066–2078.
[63]
Woo SR, Oh YT, An JY, Kang BG, Nam DH, Joo KM. Glioblastoma specific antigens, GD2 and CD90, are not involved in cancer stemness. Anatomy & Cell Biology. 2015; 48: 44–53.
[64]
Prapa M, Caldrer S, Spano C, Bestagno M, Golinelli G, Grisendi G, et al. A novel anti-GD2/4-1BB chimeric antigen receptor triggers neuroblastoma cell killing. Oncotarget. 2015; 6: 24884–24894.
[65]
Mount CW, Majzner RG, Sundaresh S, Arnold EP, Kadapakkam M, Haile S, et al. Potent antitumor efficacy of anti-GD2 CAR T cells in H3-K27M+ diffuse midline gliomas. Nature Medicine. 2018; 24: 572–579.
[66]
Majzner RG, Ramakrishna S, Yeom KW, Patel S, Chinnasamy H, Schultz LM, et al. GD2-CAR T cell therapy for H3K27M-mutated diffuse midline gliomas. Nature. 2022; 603: 934–941.
[67]
Gargett T, Ebert LM, Truong NTH, Kollis PM, Sedivakova K, Yu W, et al. GD2-targeting CAR-T cells enhanced by transgenic IL-15 expression are an effective and clinically feasible therapy for glioblastoma. Journal for Immunotherapy of Cancer. 2022; 10: e005187.
[68]
Li H, Xi Z, Dai X, Wu W, Li Y, Liu Y, et al. CD147 and glioma: a meta-analysis. Journal of Neuro-oncology. 2017; 134: 145–156.
[69]
Yang M, Yuan Y, Zhang H, Yan M, Wang S, Feng F, et al. Prognostic significance of CD147 in patients with glioblastoma. Journal of Neuro-oncology. 2013; 115: 19–26.
[70]
Tseng HC, Xiong W, Badeti S, Yang Y, Ma M, Liu T, et al. Efficacy of anti-CD147 chimeric antigen receptors targeting hepatocellular carcinoma. Nature Communications. 2020; 11: 4810.
[71]
Deshane J, Garner CC, Sontheimer H. Chlorotoxin inhibits glioma cell invasion via matrix metalloproteinase-2. The Journal of Biological Chemistry. 2003; 278: 4135–4144.
[72]
Wang D, Starr R, Chang WC, Aguilar B, Alizadeh D, Wright SL, et al. Chlorotoxin-directed CAR T cells for specific and effective targeting of glioblastoma. Science Translational Medicine. 2020; 12: eaaw2672.
[73]
Aparicio C, Belver M, Enríquez L, Espeso F, Núñez L, Sánchez A, et al. Cell Therapy for Colorectal Cancer: The Promise of Chimeric Antigen Receptor (CAR)-T Cells. International Journal of Molecular Sciences. 2021; 22: 11781.
[74]
Cui J, Zhang Q, Song Q, Wang H, Dmitriev P, Sun MY, et al. Targeting hypoxia downstream signaling protein, CAIX, for CAR T-cell therapy against glioblastoma. Neuro-oncology. 2019; 21: 1436–1446.
[75]
Huang BR, Liu YS, Lai SW, Lin HJ, Shen CK, Yang LY, et al. CAIX Regulates GBM Motility and TAM Adhesion and Polarization through EGFR/STAT3 under Hypoxic Conditions. International Journal of Molecular Sciences. 2020; 21: 5838.
[76]
Harrer DC, Dörrie J, Schaft N. CSPG4 as Target for CAR-T-Cell Therapy of Various Tumor Entities-Merits and Challenges. International Journal of Molecular Sciences. 2019; 20: 5942.
[77]
Pellegatta S, Savoldo B, Di Ianni N, Corbetta C, Chen Y, Patané M, et al. Constitutive and TNFα-inducible expression of chondroitin sulfate proteoglycan 4 in glioblastoma and neurospheres: Implications for CAR-T cell therapy. Science Translational Medicine. 2018; 10: eaao2731.
[78]
Chen H, Wei F, Yin M, Zhao Q, Liu Z, Yu B, et al. CD27 enhances the killing effect of CAR T cells targeting trophoblast cell surface antigen 2 in the treatment of solid tumors. Cancer Immunology, Immunotherapy: CII. 2021; 70: 2059–2071.
[79]
Hou B, Tang Y, Li W, Zeng Q, Chang D. Efficiency of CAR-T Therapy for Treatment of Solid Tumor in Clinical Trials: A Meta-Analysis. Disease Markers. 2019; 2019: 3425291.
[80]
Jin L, Ge H, Long Y, Yang C, Chang YE, Mu L, et al. CD70, a novel target of CAR T-cell therapy for gliomas. Neuro-oncology. 2018; 20: 55–65.
[81]
An Z, Hu Y, Bai Y, Zhang C, Xu C, Kang X, et al. Antitumor activity of the third generation EphA2 CAR-T cells against glioblastoma is associated with interferon gamma induced PD-L1. Oncoimmunology. 2021; 10: 1960728.
[82]
Yi Z, Prinzing BL, Cao F, Gottschalk S, Krenciute G. Optimizing EphA2-CAR T Cells for the Adoptive Immunotherapy of Glioma. Molecular Therapy. Methods & Clinical Development. 2018; 9: 70–80.
[83]
Zhu X, Prasad S, Gaedicke S, Hettich M, Firat E, Niedermann G. Patient-derived glioblastoma stem cells are killed by CD133-specific CAR T cells but induce the T cell aging marker CD57. Oncotarget. 2015; 6: 171–184.
[84]
Seyfrid M, Maich WT, Shaikh VM, Tatari N, Upreti D, Piyasena D, et al. CD70 as an actionable immunotherapeutic target in recurrent glioblastoma and its microenvironment. Journal for Immunotherapy of Cancer. 2022; 10: e003289.
[85]
Heitzeneder S, Bosse KR, Zhu Z, Zhelev D, Majzner RG, Radosevich MT, et al. GPC2-CAR T cells tuned for low antigen density mediate potent activity against neuroblastoma without toxicity. Cancer Cell. 2022; 40: 53–69.e9.
[86]
Bosse KR, Raman P, Zhu Z, Lane M, Martinez D, Heitzeneder S, et al. Identification of GPC2 as an Oncoprotein and Candidate Immunotherapeutic Target in High-Risk Neuroblastoma. Cancer Cell. 2017; 32: 295–309.e12.
[87]
Li G, Zhang Z, Cai L, Tang X, Huang J, Yu L, et al. Fn14-targeted BiTE and CAR-T cells demonstrate potent preclinical activity against glioblastoma. Oncoimmunology. 2021; 10: 1983306.
[88]
Davies DM, Foster J, Van Der Stegen SJC, Parente-Pereira AC, Chiapero-Stanke L, Delinassios GJ, et al. Flexible targeting of ErbB dimers that drive tumorigenesis by using genetically engineered T cells. Molecular Medicine (Cambridge, Mass.). 2012; 18: 565–576.
[89]
Li X, Daniyan AF, Lopez AV, Purdon TJ, Brentjens RJ. Cytokine IL-36γ improves CAR T-cell functionality and induces endogenous antitumor response. Leukemia. 2021; 35: 506–521.
[90]
Lai J, Mardiana S, House IG, Sek K, Henderson MA, Giuffrida L, et al. Adoptive cellular therapy with T cells expressing the dendritic cell growth factor Flt3L drives epitope spreading and antitumor immunity. Nature Immunology. 2020; 21: 914–926.
[91]
Roybal KT, Williams JZ, Morsut L, Rupp LJ, Kolinko I, Choe JH, et al. Engineering T Cells with Customized Therapeutic Response Programs Using Synthetic Notch Receptors. Cell. 2016; 167: 419–432.e16.
[92]
Moghimi B, Muthugounder S, Jambon S, Tibbetts R, Hung L, Bassiri H, et al. Preclinical assessment of the efficacy and specificity of GD2-B7H3 SynNotch CAR-T in metastatic neuroblastoma. Nature Communications. 2021; 12: 511.
[93]
Hernandez-Lopez RA, Yu W, Cabral KA, Creasey OA, Lopez Pazmino MDP, Tonai Y, et al. T cell circuits that sense antigen density with an ultrasensitive threshold. Science (New York, N.Y.). 2021; 371: 1166–1171.
[94]
Choe JH, Watchmaker PB, Simic MS, Gilbert RD, Li AW, Krasnow NA, et al. SynNotch-CAR T cells overcome challenges of specificity, heterogeneity, and persistence in treating glioblastoma. Science Translational Medicine. 2021; 13: eabe7378.
[95]
Hyrenius-Wittsten A, Su Y, Park M, Garcia JM, Alavi J, Perry N, et al. SynNotch CAR circuits enhance solid tumor recognition and promote persistent antitumor activity in mouse models. Science Translational Medicine. 2021; 13: eabd8836.
[96]
Labanieh L, Majzner RG, Klysz D, Sotillo E, Fisher CJ, Vilches-Moure JG, et al. Enhanced safety and efficacy of protease-regulated CAR-T cell receptors. Cell. 2022; 185: 1745–1763.e22.
[97]
Gilbertson RJ, Rich JN. Making a tumour’s bed: glioblastoma stem cells and the vascular niche. Nature Reviews. Cancer. 2007; 7: 733–736.
[98]
Hambardzumyan D, Bergers G. Glioblastoma: Defining Tumor Niches. Trends in Cancer. 2015; 1: 252–265.
[99]
Hou AJ, Chen LC, Chen YY. Navigating CAR-T cells through the solid-tumour microenvironment. Nature Reviews. Drug Discovery. 2021; 20: 531–550.
[100]
Richards RM, Sotillo E, Majzner RG. CAR T Cell Therapy for Neuroblastoma. Frontiers in Immunology. 2018; 9: 2380.
[101]
Vitanza NA, Johnson AJ, Wilson AL, Brown C, Yokoyama JK, Künkele A, et al. Locoregional infusion of HER2-specific CAR T cells in children and young adults with recurrent or refractory CNS tumors: an interim analysis. Nature Medicine. 2021; 27: 1544–1552.
[102]
Liu G, Rui W, Zheng H, Huang D, Yu F, Zhang Y, et al. CXCR2-modified CAR-T cells have enhanced trafficking ability that improves treatment of hepatocellular carcinoma. European Journal of Immunology. 2020; 50: 712–724.
[103]
Jin L, Tao H, Karachi A, Long Y, Hou AY, Na M, et al. CXCR1- or CXCR2-modified CAR T cells co-opt IL-8 for maximal antitumor efficacy in solid tumors. Nature Communications. 2019; 10: 4016.
[104]
Sakemura R, Can I, Siegler EL, Kenderian SS. In vivo CART cell imaging: Paving the way for success in CART cell therapy. Molecular Therapy Oncolytics. 2021; 20: 625–633.
[105]
Wang G, Zhang Z, Zhong K, Wang Z, Yang N, Tang X, et al. CXCL11-armed oncolytic adenoviruses enhance CAR-T cell therapeutic efficacy and reprogram tumor microenvironment in glioblastoma. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2023; 31: 134–153.
[106]
Huang J, Li YB, Charlebois C, Nguyen T, Liu Z, Bloemberg D, et al. Application of blood brain barrier models in pre-clinical assessment of glioblastoma-targeting CAR-T based immunotherapies. Fluids and Barriers of the CNS. 2022; 19: 38.
[107]
Kankeu Fonkoua LA, Sirpilla O, Sakemura R, Siegler EL, Kenderian SS. CAR T cell therapy and the tumor microenvironment: Current challenges and opportunities. Molecular Therapy Oncolytics. 2022; 25: 69–77.
[108]
Priceman SJ, Tilakawardane D, Jeang B, Aguilar B, Murad JP, Park AK, et al. Regional Delivery of Chimeric Antigen Receptor-Engineered T Cells Effectively Targets HER2+ Breast Cancer Metastasis to the Brain. Clinical Cancer Research: an Official Journal of the American Association for Cancer Research. 2018; 24: 95–105.
[109]
Rafiq S, Hackett CS, Brentjens RJ. Engineering strategies to overcome the current roadblocks in CAR T cell therapy. Nature Reviews. Clinical Oncology. 2020; 17: 147–167.
[110]
Gumber D, Wang LD. Improving CAR-T immunotherapy: Overcoming the challenges of T cell exhaustion. EBioMedicine. 2022; 77: 103941.
[111]
Ando M, Ito M, Srirat T, Kondo T, Yoshimura A. Memory T cell, exhaustion, and tumor immunity. Immunological Medicine. 2020; 43: 1–9.
[112]
Finney OC, Brakke HM, Rawlings-Rhea S, Hicks R, Doolittle D, Lopez M, et al. CD19 CAR T cell product and disease attributes predict leukemia remission durability. The Journal of Clinical Investigation. 2019; 129: 2123–2132.
[113]
Cherkassky L, Morello A, Villena-Vargas J, Feng Y, Dimitrov DS, Jones DR, et al. Human CAR T cells with cell-intrinsic PD-1 checkpoint blockade resist tumor-mediated inhibition. The Journal of Clinical Investigation. 2016; 126: 3130–3144.
[114]
Suarez ER, Chang DK, Sun J, Sui J, Freeman GJ, Signoretti S, et al. Chimeric antigen receptor T cells secreting anti-PD-L1 antibodies more effectively regress renal cell carcinoma in a humanized mouse model. Oncotarget. 2016; 7: 34341–34355.
[115]
Togashi Y, Shitara K, Nishikawa H. Regulatory T cells in cancer immunosuppression - implications for anticancer therapy. Nature Reviews. Clinical Oncology. 2019; 16: 356–371.
[116]
Choi BD, Yu X, Castano AP, Bouffard AA, Schmidts A, Larson RC, et al. CAR-T cells secreting BiTEs circumvent antigen escape without detectable toxicity. Nature Biotechnology. 2019; 37: 1049–1058.
[117]
Parihar R, Rivas C, Huynh M, Omer B, Lapteva N, Metelitsa LS, et al. NK Cells Expressing a Chimeric Activating Receptor Eliminate MDSCs and Rescue Impaired CAR-T Cell Activity against Solid Tumors. Cancer Immunology Research. 2019; 7: 363–375.
[118]
Strauss L, Mahmoud MAA, Weaver JD, Tijaro-Ovalle NM, Christofides A, Wang Q, et al. Targeted deletion of PD-1 in myeloid cells induces antitumor immunity. Science Immunology. 2020; 5: eaay1863.
[119]
Rodriguez-Garcia A, Palazon A, Noguera-Ortega E, Powell DJ, Jr, Guedan S. CAR-T Cells Hit the Tumor Microenvironment: Strategies to Overcome Tumor Escape. Frontiers in Immunology. 2020; 11: 1109.
[120]
Sterner RM, Kenderian SS. Myeloid cell and cytokine interactions with chimeric antigen receptor-T-cell therapy: implication for future therapies. Current Opinion in Hematology. 2020; 27: 41–48.
[121]
Gunderson AJ, Yamazaki T, McCarty K, Fox N, Phillips M, Alice A, et al. TGFβ suppresses CD8+ T cell expression of CXCR3 and tumor trafficking. Nature Communications. 2020; 11: 1749.
[122]
Tang N, Cheng C, Zhang X, Qiao M, Li N, Mu W, et al. TGF-β inhibition via CRISPR promotes the long-term efficacy of CAR T cells against solid tumors. JCI Insight. 2020; 5: e133977.
[123]
Narayan V, Barber-Rotenberg JS, Jung IY, Lacey SF, Rech AJ, Davis MM, et al. PSMA-targeting TGFβ-insensitive armored CAR T cells in metastatic castration-resistant prostate cancer: a phase 1 trial. Nature Medicine. 2022; 28: 724–734.
[124]
Larson RC, Kann MC, Bailey SR, Haradhvala NJ, Llopis PM, Bouffard AA, et al. CAR T cell killing requires the IFNγR pathway in solid but not liquid tumours. Nature. 2022; 604: 563–570.
[125]
Ma X, Shou P, Smith C, Chen Y, Du H, Sun C, et al. Interleukin-23 engineering improves CAR T cell function in solid tumors. Nature Biotechnology. 2020; 38: 448–459.
[126]
Agliardi G, Liuzzi AR, Hotblack A, De Feo D, Núñez N, Stowe CL, et al. Intratumoral IL-12 delivery empowers CAR-T cell immunotherapy in a pre-clinical model of glioblastoma. Nature Communications. 2021; 12: 444.
[127]
Aspuria PJ, Vivona S, Bauer M, Semana M, Ratti N, McCauley S, et al. An orthogonal IL-2 and IL-2Rβ system drives persistence and activation of CAR T cells and clearance of bulky lymphoma. Science Translational Medicine. 2021; 13: eabg7565.
[128]
Zhang Q, Hresko ME, Picton LK, Su L, Hollander MJ, Nunez-Cruz S, et al. A human orthogonal IL-2 and IL-2Rβ system enhances CAR T cell expansion and antitumor activity in a murine model of leukemia. Science Translational Medicine. 2021; 13: eabg6986.
[129]
Cox MJ, Lucien F, Sakemura R, Boysen JC, Kim Y, Horvei P, et al. Leukemic extracellular vesicles induce chimeric antigen receptor T cell dysfunction in chronic lymphocytic leukemia. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2021; 29: 1529–1540.
[130]
Seo H, González-Avalos E, Zhang W, Ramchandani P, Yang C, Lio CWJ, et al. BATF and IRF4 cooperate to counter exhaustion in tumor-infiltrating CAR T cells. Nature Immunology. 2021; 22: 983–995.
[131]
Smole A, Benton A, Poussin MA, Eiva MA, Mezzanotte C, Camisa B, et al. Expression of inducible factors reprograms CAR-T cells for enhanced function and safety. Cancer Cell. 2022; 40: 1470–1487.e7.
[132]
Bruchard M, Geindreau M, Perrichet A, Truntzer C, Ballot E, Boidot R, et al. Recruitment and activation of type 3 innate lymphoid cells promote antitumor immune responses. Nature Immunology. 2022; 23: 262–274.
[133]
Lelliott EJ, Kong IY, Zethoven M, Ramsbottom KM, Martelotto LG, Meyran D, et al. CDK4/6 Inhibition Promotes Antitumor Immunity through the Induction of T-cell Memory. Cancer Discovery. 2021; 11: 2582–2601.
[134]
Larson RC, Maus MV. Recent advances and discoveries in the mechanisms and functions of CAR T cells. Nature Reviews. Cancer. 2021; 21: 145–161.
[135]
Liu E, Tong Y, Dotti G, Shaim H, Savoldo B, Mukherjee M, et al. Cord blood NK cells engineered to express IL-15 and a CD19-targeted CAR show long-term persistence and potent antitumor activity. Leukemia. 2018; 32: 520–531.
[136]
Klichinsky M, Ruella M, Shestova O, Lu XM, Best A, Zeeman M, et al. Human chimeric antigen receptor macrophages for cancer immunotherapy. Nature Biotechnology. 2020; 38: 947–953.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top