Glioblastoma (GBM) contains self-renewing GBM stem cells (GSC) potentially amenable to immunologic targeting, but chimeric antigen receptor (CAR) T-cell therapy has demonstrated limited clinical responses in GBM. Here, we interrogated molecular determinants of CAR-mediated GBM killing through whole-genome CRISPR screens in both CAR T cells and patient-derived GSCs. Screening of CAR T cells identified dependencies for effector functions, including TLE4 and IKZF2. Targeted knockout of these genes enhanced CAR antitumor efficacy. Bulk and single-cell RNA sequencing of edited CAR T cells revealed transcriptional profiles of superior effector function and inhibited exhaustion responses. Reciprocal screening of GSCs identified genes essential for susceptibility to CAR-mediated killing, including RELA and NPLOC4, the knockout of which altered tumor–immune signaling and increased responsiveness of CAR therapy. Overall, CRISPR screening of CAR T cells and GSCs discovered avenues for enhancing CAR therapeutic efficacy against GBM, with the potential to be extended to other solid tumors.

Significance:

Reciprocal CRISPR screening identified genes in both CAR T cells and tumor cells regulating the potency of CAR T-cell cytotoxicity, informing molecular targeting strategies to potentiate CAR T-cell antitumor efficacy and elucidate genetic modifications of tumor cells in combination with CAR T cells to advance immuno-oncotherapy.

This article is highlighted in the In This Issue feature, p. 995

Glioblastoma (GBM) ranks as one of the most lethal of human cancers, with current therapy offering only palliation. Standard-of-care therapy consisting of maximal surgical resection followed by combined radiation and chemotherapy extends median survival by less than three months. The activation of antitumor immune responses may provide new opportunities to augment tumor control. As such, immunotherapies have been extensively investigated with positive results in preclinical studies, yet broad antitumor efficacy has not occurred in patients (1). The adoptive transfer of chimeric antigen receptor (CAR) engineered T cells has shown promising clinical activity in a subset of cancers, particularly B-cell malignancies (2, 3). To target GBM, CAR T cells have been engineered to recognize selected tumor antigens and have demonstrated cytolytic activity against GBM cells, including GBM stem cells (GSC; refs. 4–6). In patients with GBM, CAR T-cell therapies have shown early evidence of activity, clinical feasibility, and safety (7–10). However, the overall outcomes of CAR T-cell treatment remain unsatisfactory, prompting efforts to enhance the antitumor potency of GBM-targeting CAR T cells (11, 12). The functional potentiation of CAR T cells, though attractive due to the modifiable nature of these cells, requires a comprehensive understanding of the molecular events regulating CAR T-cell activation, exhaustion, and tumor-induced immune suppression (11, 13). Aside from CAR recognition of tumor antigens, the complicated and dynamic interaction between CAR T cells and their target tumor cells remains poorly characterized. Thus, there is a need for new strategies to further enhance CAR T-cell potency.

Gene editing using CRISPR/Cas9 is a promising approach to enhance cancer immunotherapy (14). Directed CRISPR/Cas9 gene knockout (KO) of checkpoint and other immune-regulatory receptors has shown utility for adoptive T-cell therapy (15, 16); however, this approach has focused on a limited set of known pathways. By contrast, large CRISPR-KO screens are an effective platform for unbiased target discovery and have been successfully used to identify genes in tumor cells which when deleted synergize with various types of immunotherapeutics (17–19). CRISPR screens in T cells identified modulators of T-cell receptor (TCR) activation in response to stimulation with CD3/CD28 agonistic beads, viruses, or tumor cells (20–22). Although CAR constructs are synthetic TCR-like receptors incorporating CD3ζ and costimulatory domains, the molecular events are not identical between TCR and CAR T-cell activation signaling pathways (23). Thus, unbiased CRISPR screen presents an attractive strategy for discovering key regulators of CAR–tumor interaction.

GSCs represent a potentially important cellular target in GBM, as they have been linked to therapeutic resistance, invasion into normal brain, promotion of angiogenesis, and immune modulation (24, 25). We hypothesized that systematic interrogation of molecular regulation of CAR T-cell efficacy against GBM could be optimized by screening both CAR T cells and GBM cells, thereby informing the interplay between a cell-based therapy and its target population. Here, we developed a robust method for performing whole-genome CRISPR-KO screens in both GBM cells and human CAR T cells. Using our well-established CAR T-cell platform targeting the tumor-associated surface marker IL13 receptor α2 (IL13Rα2; refs. 7, 8, 26), we identified novel CAR T cell– and tumor-intrinsic targets that substantially improved CAR T-cell cytotoxicity against GSCs both in vitro and in vivo. Targeted genetic modification of identified hits in CAR T cells potentiated their long-term activation, cytolytic activity, and in vivo antitumor function against GSCs, demonstrating that CRISPR screen on CAR T cells leads to the discovery of key targets for augmenting CAR T-cell therapeutic potency. In parallel, KO of identified targets in GSCs sensitized them to CAR-mediated killing both in vitro and in vivo, revealing potential avenues for combinatorial inhibitor treatment to augment CAR T-cell efficacy. Our findings represent a feasible and highly effective approach to discovering key targets that mediate effective tumor eradication using CAR T cells.

Genome-Wide CRISPR Screening of CAR T Cells Identifies Essential Regulators of Effector Activity

The fitness of CAR T-cell products correlates with clinical responses (27, 28), indicating that key regulators of CAR T-cell function can be targeted to potentiate therapeutic efficacy. T-cell exhaustion resulting from chronic tumor exposure limits CAR T-cell antitumor responses (29). To identify the essential regulators of T-cell functional activity in an unbiased manner, we performed genome-wide CRISPR screen adapting our previously developed in vitro tumor rechallenge assay, which differentiates CAR T-cell potency in the setting of high tumor burden and reflects in vivo antitumor activity (30, 31). IL13Rα2-targeted CAR T cells from two healthy human donors were lentivirally transduced to express the Brunello short-guide RNA (sgRNA) library (32) and the CAR construct and then electroporated with Cas9 protein. CAR T cells harboring CRISPR-mediated knockouts were recursively exposed to an excess amount of PBT030-2 GSCs (Fig. 1A), an IDH1 wild-type patient-derived GSC line that highly expresses IL13Rα2 (33). After tumor stimulation, CAR T cells were sorted from coculture and classified into subsets based on expression of the inhibitory receptor PD-1, which is associated with T-cell exhaustion (Fig. 1A). To identify gene knockouts that augment CAR T antitumor activity, we identified sgRNAs enriched in the less-exhausted PD-1–negative versus PD-1–positive CAR T-cell compartments (Fig. 1B; Supplementary Fig. S1A; Supplementary Table S1). To eliminate targets that nonspecifically impaired CAR T-cell proliferation or viability, we excluded sgRNAs depleted (β < −1) in CAR T-cells after 72-hour coculture with GSCs or 72-hour monoculture (Fig. 1C). Two hundred twenty genes were common hits in both T-cell donors (Fig. 1D). Many of these 220 genes are induced by the IL4 receptor (IL4R), which suppresses T-cell activity (34), as well as type I IFN, NFAT, TCF4, and JAK1/2, which all play complex roles in T-cell activation and mediate T-cell exhaustion and inhibition under some circumstances (refs. 35–38; Fig. 1E). In addition, these genes were enriched for pathways that contribute to T-cell exhaustion, including nuclear receptor transcription and cholesterol responses (refs. 39, 40; Supplementary Fig. S1B). In contrast, genes preferentially depleted in PD-1–positive cells included pathways associated with T-cell activation, including amide metabolism and NFκB signaling (41, 42), as well as negative regulation of oxidative stress–induced cell death (Fig. 1E). Together, these data verify that our screen identified genes involved in T-cell effector activity, providing candidate genes that can be modulated to prevent exhaustion and enhance effector function of CAR T cells.

Figure 1.

CRISPR/Cas9 screen in CAR T cells cocultured with GSCs. A, Overview of screen design. CAR T cells were transduced with a whole-genome CRISPR/Cas9 library and cocultured with GSCs, followed by a GSC rechallenge after 48 hours. At the conclusion of the screen (24 hours after the rechallenge), CAR T cells were sorted for PD-1 positivity, and PD-1+ or PD-1 CAR T cells were sequenced separately to identify enriched and depleted guides. B, Screen results in two replicates of independent donors with genes ordered alphabetically on the x-axis. The MAGECK β-value for each gene comparing PD-1 versus PD-1+ is plotted on the y-axis. Genes enriched in PD-1 cells at a β-value >1 are in blue or red, and genes with a β-value of < −1 (enriched in PD-1+ cells) are in green or purple. C, Plot of hits from B to exclude genes that are depleted following coculture of CAR T cells with GSCs (β-value < −1 on the y-axis) or in monoculture (β-value < −1 on the x-axis). Genes in blue or red are not depleted in either condition. D, Venn diagram illustrating common hits for depleted genes in two distinct T-cell donors. E, Ingenuity Pathway Analysis of master regulators (top five based on P values) of 220 overlapping genes in two T-cell donors. F, Common hits ranked by β-value in a combined model for PD-1 versus PD-1+ CAR T cells. Labeled hits were selected for validation.

Figure 1.

CRISPR/Cas9 screen in CAR T cells cocultured with GSCs. A, Overview of screen design. CAR T cells were transduced with a whole-genome CRISPR/Cas9 library and cocultured with GSCs, followed by a GSC rechallenge after 48 hours. At the conclusion of the screen (24 hours after the rechallenge), CAR T cells were sorted for PD-1 positivity, and PD-1+ or PD-1 CAR T cells were sequenced separately to identify enriched and depleted guides. B, Screen results in two replicates of independent donors with genes ordered alphabetically on the x-axis. The MAGECK β-value for each gene comparing PD-1 versus PD-1+ is plotted on the y-axis. Genes enriched in PD-1 cells at a β-value >1 are in blue or red, and genes with a β-value of < −1 (enriched in PD-1+ cells) are in green or purple. C, Plot of hits from B to exclude genes that are depleted following coculture of CAR T cells with GSCs (β-value < −1 on the y-axis) or in monoculture (β-value < −1 on the x-axis). Genes in blue or red are not depleted in either condition. D, Venn diagram illustrating common hits for depleted genes in two distinct T-cell donors. E, Ingenuity Pathway Analysis of master regulators (top five based on P values) of 220 overlapping genes in two T-cell donors. F, Common hits ranked by β-value in a combined model for PD-1 versus PD-1+ CAR T cells. Labeled hits were selected for validation.

Close modal

CRISPR Screening Empowers Discovery of Targets That Enhance CAR T-cell Cytotoxic Potency

We interrogated the 220 targets enriched in PD-1–negative cells common between two T-cell donors (Supplementary Table S2), focusing on four representative genes identified in the top third of hits, which have not been previously explored for their role in enhancing CAR T-cell function. These included the high-ranking hits: eukaryotic translation initiation factor 5A-1 (EIF5A), transcription factor transducin like enhancer of split 4 (TLE4), Ikaros family zinc finger protein 2 (IKZF2), and transmembrane protein 184B (TMEM184B; Fig. 1F). Most sgRNAs targeting these genes (≥2 out of 4 in both replicates) were enriched in PD-1–negative CAR T cells (Supplementary Fig. S1C). To verify the function of these targets by CRISPR-mediated KO on CAR T cells, we leveraged the challenging in vitro killing assay (CAR:tumor = 1:40), confirming that targeting TLE4, IKZF2, TMEM184B, or EIF5A improved in vitro killing potency of CAR T cells against GSCs, as well as their expansion potential, although sgEIF5A-3 effects were more modest (Fig. 2A and B). Mechanistically, KO of these genes reduced PD-1 expression on CAR T cells following tumor stimulation (Supplementary Fig. S2A). We and others have shown that CAR T-cell exhaustion is associated with coexpression of PD-1, LAG3, and TIM3 (43, 44). All four KOs reduced CAR T-cell exhaustion, TLE4-KO and IKZF2-KO most effectively (Fig. 2C). KO of these genes minimally affected initial CAR T-cell activation upon tumor cell recognition (Supplementary Fig. S2B), suggesting that these KOs improved T-cell fitness and long-term function instead of initial activation. Targeted KOs did not affect the expression and stability of the CAR in T cells (Supplementary Fig. S2C and S2D). As validation, we performed independent studies with a HER2-targeted CAR model that also demonstrated improvements in CAR killing and expansion, suggesting that genetic screens of CAR T cells may yield broadly effective molecular strategies (Supplementary Fig. S2E and S2F).

Figure 2.

Targets on CAR T cells improve effector potency and alter transcriptional profiles. A, Killing of CAR T cells with TLE4-, IKZF2-, TMEM184B-, or EIF5A-KO against cocultured GSCs (E:T = 1:40, 48 hours). B, Expansion of CAR T cells with different knockouts in coculture with GSCs (E:T = 1:40, 48 hours). C, CAR T cells with targeted KOs of specific genes were cocultured with PBT030-2 cells (E:T = 1:4) for 48 hours, and rechallenged with tumor cells against (E:T = 1:8) for 24 hours, and then analyzed for the expression of exhaustion markers. A–C, *, P < 0.05; **, P < 0.01; ***, P < 0.001 compared with CAR T cells transduced with nontargeting sgRNA (black) using unpaired Student t tests. D and E, Unsupervised clustering of ssGSEA scores comparing TLE4-KO (C) or IKZF2-KO (D) versus sgCONT CAR T cells for the signatures of selected T-cell populations (left) or immune and functional pathways (right). F, Left, box plot of genes involved in apoptotic signaling from RNA-seq data in sgCONT (blue) versus sgTLE4 (red). Right, Reactome network of genes downregulated following TLE4 KO that are involved in apoptotic signaling. G, Left, box plot of genes involved in AP-1 signaling from RNA-seq data in control (blue) vs. TLE4-KO (red) cells. Right, Reactome network of genes upregulated with TLE4-KO that are linked to FOS. Increasing node size and fill hue are proportional to node degree. H, Histogram of log2 fold change of gene expression (comparing TLE4-KO vs. control) for 250 genes previously shown to be upregulated with JUN overexpression. I, Left, box plot of genes involved in cytokine receptor signaling from RNA-seq data in control (blue) versus IKZF2-KO (red) cells. Right, Reactome network of genes upregulated with IKZF2-KO that are linked to a gene in the cytokine receptor signaling pathway (labeled in red). Increasing node size and fill hue are proportional to node degree. J, Left, box plot of genes in the NFAT pathways from RNA-seq data in control (blue) versus IKZF2-KO (red) cells. Right, Reactome network of genes upregulated with IKZF2-KO that are linked to upregulated genes in the NFAT pathway (labeled in red).

Figure 2.

Targets on CAR T cells improve effector potency and alter transcriptional profiles. A, Killing of CAR T cells with TLE4-, IKZF2-, TMEM184B-, or EIF5A-KO against cocultured GSCs (E:T = 1:40, 48 hours). B, Expansion of CAR T cells with different knockouts in coculture with GSCs (E:T = 1:40, 48 hours). C, CAR T cells with targeted KOs of specific genes were cocultured with PBT030-2 cells (E:T = 1:4) for 48 hours, and rechallenged with tumor cells against (E:T = 1:8) for 24 hours, and then analyzed for the expression of exhaustion markers. A–C, *, P < 0.05; **, P < 0.01; ***, P < 0.001 compared with CAR T cells transduced with nontargeting sgRNA (black) using unpaired Student t tests. D and E, Unsupervised clustering of ssGSEA scores comparing TLE4-KO (C) or IKZF2-KO (D) versus sgCONT CAR T cells for the signatures of selected T-cell populations (left) or immune and functional pathways (right). F, Left, box plot of genes involved in apoptotic signaling from RNA-seq data in sgCONT (blue) versus sgTLE4 (red). Right, Reactome network of genes downregulated following TLE4 KO that are involved in apoptotic signaling. G, Left, box plot of genes involved in AP-1 signaling from RNA-seq data in control (blue) vs. TLE4-KO (red) cells. Right, Reactome network of genes upregulated with TLE4-KO that are linked to FOS. Increasing node size and fill hue are proportional to node degree. H, Histogram of log2 fold change of gene expression (comparing TLE4-KO vs. control) for 250 genes previously shown to be upregulated with JUN overexpression. I, Left, box plot of genes involved in cytokine receptor signaling from RNA-seq data in control (blue) versus IKZF2-KO (red) cells. Right, Reactome network of genes upregulated with IKZF2-KO that are linked to a gene in the cytokine receptor signaling pathway (labeled in red). Increasing node size and fill hue are proportional to node degree. J, Left, box plot of genes in the NFAT pathways from RNA-seq data in control (blue) versus IKZF2-KO (red) cells. Right, Reactome network of genes upregulated with IKZF2-KO that are linked to upregulated genes in the NFAT pathway (labeled in red).

Close modal

TLE4 is a transcriptional corepressor of multiple genes encoding inflammatory cytokines (45) and IKZF2 is upregulated in exhausted T cells (37, 46, 47), supporting potential roles in inhibiting CAR T-cell function. To elucidate molecular mechanisms underlying the regulation of CAR T-cell activity, we compared the transcriptomes of CAR T cells with individual knockouts against cells transduced with nontargeted sgRNA (sgCONT). TLE4-KO in CAR T cells upregulated critical regulators of T-cell activation, including the transcription factor EGR1, which promotes Th1 cell differentiation (48), and the metabolic regulator BCAT, which mediates metabolic fitness in activated T cells (ref. 49; Supplementary Fig. S2G). IKZF2-KO in CAR T cells upregulated proinflammatory cytokines and pathways, including CXCL8, CCL3, and CCL4 (50–52), as well as EGR1, similar to TLE4-KO (Supplementary Fig. S2H). We next compared transcriptional profiles of TLE4-KO or IKZF2-KO CAR T cells to the signatures of known T-cell subsets and pathways (35, 53, 54). TLE4-KO or IKZF2-KO induced molecular signatures representing activation over memory T cells, together with key T-cell activation signaling pathways (TCR signaling, T-cell activation, AP-1, and ZAP; Fig. 2D and E; Supplementary Fig. S2I and S2J). T-cell activation characteristics in TLE4-KO or IKZF2-KO cells were uncoupled from exhaustion (Fig. 2D and E), suggesting retention of CAR T-cell function. TLE4-KO cells downregulated an apoptosis signature (Fig. 2D and F) and upregulated AP-1 signaling, which maintains CAR T-cell function (ref. 55; Fig. 2G). In particular, the AP-1 family transcription factor FOS was enriched after TLE4-KO, together with many of its downstream targets (Fig. 2G). As overexpression of the AP-1 family member JUN prevents CAR T-cell exhaustion (55), we investigated whether TLE4-KO phenocopied transcriptional changes of JUN overexpression, revealing that genes upregulated with TLE4-KO overlapped with genes with upregulated following JUN overexpression (Fig. 2H). IKZF2-KO upregulated pathways involving interactions between cytokines and their receptors, as well as NFAT signaling, which regulates key molecular signals following T-cell activation (ref. 56; Fig. 2I and J). As EGR1 was upregulated after IKZF2-KO, many genes in these pathways were likely downstream targets (Fig. 2I and J).

Whole-transcriptome analyses following TMEM184B or EIF5A KO revealed convergence of altered pathways, similar to those induced by TLE4-KO or IKZF2-KO, including the upregulation of BCAT1, EGR1, and IL17RB (Supplementary Fig. S3A and S3B) and the acquisition of memory or effector over naïve T-cell signatures (Supplementary Fig. S3C and S3D). However, targeting TMEM184B or EIF5A did not enrich for cytokine secretion and response pathways in CAR T cells (Supplementary Fig. S3E and S3F), which were found in TLE4-KO or IKZF2-KO CAR T cells. As these cytokines (CCL3 and CCL4) maintain T-cell function during chronic viral infection and in the tumor microenvironment (57, 58), our results indicate that TMEM184B-KO and EIF5A-KO CAR T cells might be prone to terminal effector differentiation and subsequent exhaustion, thereby compromising their overall functional capability despite their potent in vitro cytotoxicity. Overall, KO of these genes in CAR T cells also maintained transcriptional profiles of T-cell activation, which are associated with effector potency.

Targeting TLE4 and IKZF2 Modifies CAR T Subsets Associated with Effector Potency

To determine the impact of TLE4-KO or IKZF2-KO on specific subpopulations of CAR T cells, we performed comparative single-cell RNA sequencing (scRNA-seq) on KO and control CAR T cells with or without stimulation by tumor cells. Comparing TLE4-KO cells with control CAR T cells by unbiased clustering of pooled data identified 10 different clusters, the distribution of which was greatly influenced by stimulation (Fig. 3AC; Supplementary Fig. S4A). CD4+ and CD8+ CAR T cells were well delineated (Supplementary Fig. S4B). Stimulated cells downregulated naïve/memory-related markers (e.g., IL7R and CCR7) and upregulated activation-related markers (e.g., MKI67 and GZMB; Supplementary Fig. S4C). Stimulation enriched clusters 0, 1, and 10 (showing high expression of activation or exhaustion markers) and depleted clusters 3, 5, 7, and 9 (expressing naïve/memory markers; Fig. 3D). The impact of TLE4-KO on the overall distribution CAR T cells was more significant after stimulation; notably, cluster 8 was depleted after stimulation only in control, but not in TLE4-KO cells (Fig. 3C and D). This cluster represented a subset of CD4+ T cells expressing multiple costimulatory molecules, including CD28, ICOS, CD86, and TNFRSF4 (OX40), as well as the cytokine IL2 (Fig. 3D; Supplementary Fig. S4D). Although no proliferative activity was detected in this cluster (indicated by low Ki-67), preservation of this cluster in TLE4-KO cells was maintained post-stimulation (Fig. 3D). In TLE4-KO cells, cluster 8 also showed expression of the immune stimulatory cytokine CCL3 (Fig. 3E), costimulatory molecule TNFRSF4 (Fig. 3F), and AP-1 transcription factors FOS and JUN (Supplementary Fig. S4E and S4F), which were minimally expressed in control cells. Cluster 10 was an activated CD4+ subset expressing multiple cytokines, including IL2 and TNF, and this cluster displayed greater post-stimulation expansion in TLE4-KO cells (Fig. 3D). In the clusters with activation signatures (0, 1, and 10), TLE4-KO upregulated IFNG, BCAT, GZMB, CCL3, and CCL4 (Fig. 3G and H; Supplementary Fig. S4G–S4I). Combining the transcriptome readouts from all single cells revealed that tumor stimulation in TLE4-KO cells induced T-cell stimulatory and cytotoxic factors (e.g., GZMB, CCL3, CCL4, and IFNG) to a greater degree than control CAR T cells (Supplementary Fig. S4G–S4J). Taken together, the enhanced cytotoxicity of TLE-KO CAR T cells could result from the preservation of specific T-cell subsets after tumor stimulation.

Figure 3.

The effect of TLE4-KO on CAR T-cell subpopulations. A, UMAP projection of single-cell RNA-seq of control and TLE4-KO CAR T cells both before and after stimulation with GSCs. Cluster assignments for the overall population are shown. B, Cluster composition of unstimulated versus unstimulated control or TLE4-KO cell populations. C, Population distribution of control and TLE4-KO CAR T cells before and after stimulation. D, Characterization of clusters based upon cell proliferation. Top, violin plot of MKI67 expression. Middle, dot plot of CD4 versus CD8A expression wherein larger dots indicate a higher proportion of cells with expression and red versus blue fill indicates higher expression. Heat map, scaled expression of T-cell markers including costimulatory, activation, naïve, exhaustion, and Treg cell markers as well as AP-1 signaling. Bottom, proportion of cells in each cluster understimulated versus unstimulated conditions in control (blue) or TLE4-KO (red) populations. Positive values indicate increase in cluster occupancy following stimulation. E–H, Expression of CCL3 (E), TNFRSF4 (F), IFNG (G), and BCAT1 (H) in control or TLE4-KO CAR T cells superimposed on the UMAP projection.

Figure 3.

The effect of TLE4-KO on CAR T-cell subpopulations. A, UMAP projection of single-cell RNA-seq of control and TLE4-KO CAR T cells both before and after stimulation with GSCs. Cluster assignments for the overall population are shown. B, Cluster composition of unstimulated versus unstimulated control or TLE4-KO cell populations. C, Population distribution of control and TLE4-KO CAR T cells before and after stimulation. D, Characterization of clusters based upon cell proliferation. Top, violin plot of MKI67 expression. Middle, dot plot of CD4 versus CD8A expression wherein larger dots indicate a higher proportion of cells with expression and red versus blue fill indicates higher expression. Heat map, scaled expression of T-cell markers including costimulatory, activation, naïve, exhaustion, and Treg cell markers as well as AP-1 signaling. Bottom, proportion of cells in each cluster understimulated versus unstimulated conditions in control (blue) or TLE4-KO (red) populations. Positive values indicate increase in cluster occupancy following stimulation. E–H, Expression of CCL3 (E), TNFRSF4 (F), IFNG (G), and BCAT1 (H) in control or TLE4-KO CAR T cells superimposed on the UMAP projection.

Close modal

Comparison between IKZF2-KO cells and control CAR T cells identified 10 clusters using unbiased clustering of pooled data (Fig. 4A). In parallel with the comparisons between TLE4-KO versus control cells, we observed a dramatic change in cluster distribution and gene expression after stimulation, with moderate changes from IKZF2-KO (Fig. 4B and C; Supplementary Fig. S5A–S5D). Given a role for IKZF2 in regulatory T cells (Treg; ref. 59), cluster 0, characterized by Treg signatures (e.g., CLTA4, FOXP3, and IL2RA), was reduced in IKZF2-KO cells (Fig. 4BD). Cluster 10 was induced after stimulation, enriched in IKZF2-KO cells, and expressed elevated levels of the AP-1 signaling molecules FOS and JUN (Fig. 4BD). These cells expressed a limited repertoire of cytokines beyond TNF, but had medium-to-high levels of Ki-67, high expression of EGR1 and IL8, and exclusively expressed CXCL10 and CCND1 (Fig. 4DF; Supplementary Fig. S5D). Upregulated genes in cluster 10 were enriched for transcriptional regulation by ATF3 and JUN (Supplementary Fig. S5E and S5F). This subset contained a very limited number of cells and was present only upon stimulation, potentially explaining the lack of differential expression of FOS and JUN in bulk RNA-seq analysis in IKZF2-KO cells. Cluster 2 was also expanded after stimulation in both IKZF2-KO and control cells (Fig. 4D; Supplementary Fig. S5A). However, induction of activation-associated genes in this cluster, including IFNG, CCL3, and CCL4, was more robust in IKZF2-KO versus control cells upon tumor stimulation (Fig. 4G and H; Supplementary Fig. S5G). In IKZF2-KO cells, CCL3 was expressed at higher levels in clusters 0, 1, and 9 (Fig. 4G). As a result, IKZF2-KO cells exhibited an augmented responsiveness to tumor stimulation, illustrated by the upregulation of activation-associated cytokines (Supplementary Fig. S5H). Overall, scRNA-seq analysis revealed that TLE4-KO or IKZF2-KO resulted in the preservation or expansion of certain CAR T-cell subsets after tumor stimulation. These cellular subsets displayed transcriptional signatures of T-cell cytotoxicity and/or immune stimulation, providing some underlying mechanisms of their superior effector function against tumor cells.

Figure 4.

IKZF2 regulates CAR T-cell subpopulations. A, UMAP projection of scRNA-seq of control and IKZF2-KO CAR T cells both before and after stimulation with GSCs. Top, cluster assignments for the overall population. B, Cluster composition of unstimulated versus unstimulated control or IKZF2-KO cell populations. C, Population distribution of control and IKZF2-KO CAR T cells before and after stimulation. D, Characterization of clusters based upon cell proliferation. Top, violin plot of MKI67 expression. Middle, dot plot of CD4 versus CD8A expression wherein larger dots indicate a higher proportion of cells with expression and red versus blue fill indicates higher expression. Heat map, scaled expression of T-cell markers including costimulatory, activation, naïve, exhaustion, and Treg cell markers as well as AP-1 signaling. Bottom, proportion of cells in each cluster understimulated versus unstimulated conditions in sgCONT (blue) or sgIKZF2 (red) populations. Positive values indicate increase in cluster occupancy following stimulation. E, Expression of CXCL10 and CCND1 across clusters (violin plot). F, Expression of top upregulated genes in bulk RNA-seq for sgIKZF2 versus sgCONT across single-cell clusters. G and H, Expression of IFNG (G) and CCL3 (H) in sgCONT or sgIKZF2 CAR T cells superimposed on the UMAP projection.

Figure 4.

IKZF2 regulates CAR T-cell subpopulations. A, UMAP projection of scRNA-seq of control and IKZF2-KO CAR T cells both before and after stimulation with GSCs. Top, cluster assignments for the overall population. B, Cluster composition of unstimulated versus unstimulated control or IKZF2-KO cell populations. C, Population distribution of control and IKZF2-KO CAR T cells before and after stimulation. D, Characterization of clusters based upon cell proliferation. Top, violin plot of MKI67 expression. Middle, dot plot of CD4 versus CD8A expression wherein larger dots indicate a higher proportion of cells with expression and red versus blue fill indicates higher expression. Heat map, scaled expression of T-cell markers including costimulatory, activation, naïve, exhaustion, and Treg cell markers as well as AP-1 signaling. Bottom, proportion of cells in each cluster understimulated versus unstimulated conditions in sgCONT (blue) or sgIKZF2 (red) populations. Positive values indicate increase in cluster occupancy following stimulation. E, Expression of CXCL10 and CCND1 across clusters (violin plot). F, Expression of top upregulated genes in bulk RNA-seq for sgIKZF2 versus sgCONT across single-cell clusters. G and H, Expression of IFNG (G) and CCL3 (H) in sgCONT or sgIKZF2 CAR T cells superimposed on the UMAP projection.

Close modal

Genome-Wide Screening of GSCs Identified Genes Mediating Resistance to CAR T Cells

Augmenting the efficacy of CAR T cells against GBM can be approached by studying T cells themselves, as above, which may inform targeted KOs in addition to CAR engineering for enhancing CAR activity. Reciprocal screening of GBM cells, especially GSCs, potentially informs interactions with CAR T cells to predict clinical responsiveness to CAR T-cell therapy. To identify potential genes in GSCs that promote resistance to CAR-mediated cytotoxicity, we performed genome-wide CRISPR screens on two independent patient-derived GSC lines (PBT030-2 and PBT036), both derived from primary GBM tumors with high expression of IL13Rα2 (33). To identify tumor cell targets that rendered GBM cells more susceptible to T-cell immunotherapy, we subjected GSCs to two rounds of coculture with IL13Rα2-targeted CAR T cells (Fig. 5A; Supplementary Fig. S6A). We identified sgRNAs that were enriched (β > 1) or depleted (β < −1) in the surviving GSCs compared with GSCs in monoculture for the same amount of time (Fig. 5B). The genes with sgRNAs depleted in coculture (β < −1) represented targets that promoted CAR killing upon KO (Fig. 5C). To exclude sgRNAs that nonspecifically targeted essential genes for GSC survival, we removed gene hits that were depleted in GSCs after 48-hour culture without CAR T cells (Fig. 5C). A total of 228 CAR-modulating genes were identified as hits in either GSC line, with only four overlapping targets common to both lines (Fig. 5D). Enriched pathways included tumor immune modulation, such as MHC I antigen presentation, IL1 signaling, and NFκB activation (Fig. 5E), indicating that sgRNAs depleted in surviving GSCs targeted genes responsible for resistance to T-cell killing.

Figure 5.

CRISPR/Cas9 screen in GSCs cocultured with CAR T cells. A, Overview of screen design. GSCs were transduced with a whole-genome CRISPR/Cas9 library and subjected to two rounds of CAR T-cell killing (total E:T = 1:1). GSCs were then extracted, and libraries were prepared and sequenced to identify enriched and depleted guides. B, Results of the screen in each GSC model. Genes are ordered alphabetically on the x-axis and by MAGECK β score on the y-axis comparing coculture versus untreated GSCs. Genes in purple or green are enriched at β > 1 (sgRNAs targeting genes that impair GSC killing by CAR T cells) and those in red or blue are depleted at β < −1 (sgRNAs targeting genes that promote GSC killing by CAR T cells). C, Plot of depleted genes for each model ordered alphabetically on the x-axis by MAGECK β score on the y-axis comparing untreated day 3 vs. day 0. Points in gray are depleted at β < −1 (sgRNAs targeting the gene impair GSC survival). The remaining points in red or blue indicate genes for which KO does not affect GSC survival. D, Venn diagram illustrating common hits for depleted genes in two models. E, ClueGO plot of Gene Ontology (GO) and Reactome pathways enriched in the union of hits for both models. F, Log2 fold change of normalized counts for each sgRNA targeting common CRISPR screen hits comparing coculture to day 0.

Figure 5.

CRISPR/Cas9 screen in GSCs cocultured with CAR T cells. A, Overview of screen design. GSCs were transduced with a whole-genome CRISPR/Cas9 library and subjected to two rounds of CAR T-cell killing (total E:T = 1:1). GSCs were then extracted, and libraries were prepared and sequenced to identify enriched and depleted guides. B, Results of the screen in each GSC model. Genes are ordered alphabetically on the x-axis and by MAGECK β score on the y-axis comparing coculture versus untreated GSCs. Genes in purple or green are enriched at β > 1 (sgRNAs targeting genes that impair GSC killing by CAR T cells) and those in red or blue are depleted at β < −1 (sgRNAs targeting genes that promote GSC killing by CAR T cells). C, Plot of depleted genes for each model ordered alphabetically on the x-axis by MAGECK β score on the y-axis comparing untreated day 3 vs. day 0. Points in gray are depleted at β < −1 (sgRNAs targeting the gene impair GSC survival). The remaining points in red or blue indicate genes for which KO does not affect GSC survival. D, Venn diagram illustrating common hits for depleted genes in two models. E, ClueGO plot of Gene Ontology (GO) and Reactome pathways enriched in the union of hits for both models. F, Log2 fold change of normalized counts for each sgRNA targeting common CRISPR screen hits comparing coculture to day 0.

Close modal

Knockout of RELA or NPLOC4 Sensitizes GSCs to CAR-Mediated Antitumor Activity

Next, we sought to confirm and further characterize the function of common top hits whose deletion promoted CAR killing (Fig. 5D). V-Rel reticuloendotheliosis viral oncogene homolog A (RELA) and nuclear protein localization protein 4 homolog (NPLOC4) were selected for further validation as all sgRNAs targeting these two genes in the screen showed depletion in GSCs cocultured with CAR (Fig. 5F). As expected from our selection process, CRISPR-mediated KO of either RELA or NPLOC4 caused limited reduction in the growth of GSCs in vitro compared with GSCs transduced with control nontargeted sgRNAs (sgCONT; Supplementary Fig. S6B). When cocultured with CAR T cells in a challenging in vitro model at low T-cell ratios [effector:target (E:T) = 1:40; 48 hours], RELA or NPLOC4 KO in GSCs increased susceptibility to CAR T cell–mediated killing (Fig. 6A), which was also associated with increased expansion of CAR T cells (Fig. 6B). Thus, KO of either RELA or NPLOC4 in GSCs enhanced the cytotoxic and proliferative potency of CAR T cells.

Figure 6.

RELA or NPLOC4 disruption improves CAR T-cell killing of GSCs. A, CAR T-cell killing of GSCs (E:T = 1:40, 48 hours) with CRISPR-mediated KO of RELA or NPLOC4. B, CAR T-cell expansion in coculture with GSCs (E:T = 1:40, 48 hours) with CRISPR-mediated KO of RELA or NPLOC4. A and B, *, P < 0.05; **, P < 0.01; ***, P < 0.001 compared with GSCs transduced with nontargeting sgRNA (black) using unpaired Student t tests. C, RNA-seq of GSCs following RELA-KO plotted as –log10 FDR (y-axis) versus log2 fold change of RELA-KO versus control (x-axis). Blue or red points are genes with < −1.5 or >1.5 fold change, respectively, at an FDR of <0.05. D, Reactome network of genes downregulated following RELA-KO. Only genes linked in the Reactome database to at least one other gene are shown. Node size and color saturation are proportional to node degree. Activating interactions are indicated by arrowheads; dotted lines indicate predicted interactions. E, Pathway enrichment of genes in the Reactome network of downregulated genes in C. F, RNA-seq of GSCs following NPLOC4-KO plotted as –log10 FDR (y-axis) versus log2 fold change of NPLOC4-KO versus control (x-axis). Blue or red points are genes with < −1.5 or >1.5 fold change, respectively, at an FDR of <0.05. G, Reactome network of genes downregulated following NPLOC4-KO. H, Pathway enrichment of genes in the Reactome network of downregulated genes in F.

Figure 6.

RELA or NPLOC4 disruption improves CAR T-cell killing of GSCs. A, CAR T-cell killing of GSCs (E:T = 1:40, 48 hours) with CRISPR-mediated KO of RELA or NPLOC4. B, CAR T-cell expansion in coculture with GSCs (E:T = 1:40, 48 hours) with CRISPR-mediated KO of RELA or NPLOC4. A and B, *, P < 0.05; **, P < 0.01; ***, P < 0.001 compared with GSCs transduced with nontargeting sgRNA (black) using unpaired Student t tests. C, RNA-seq of GSCs following RELA-KO plotted as –log10 FDR (y-axis) versus log2 fold change of RELA-KO versus control (x-axis). Blue or red points are genes with < −1.5 or >1.5 fold change, respectively, at an FDR of <0.05. D, Reactome network of genes downregulated following RELA-KO. Only genes linked in the Reactome database to at least one other gene are shown. Node size and color saturation are proportional to node degree. Activating interactions are indicated by arrowheads; dotted lines indicate predicted interactions. E, Pathway enrichment of genes in the Reactome network of downregulated genes in C. F, RNA-seq of GSCs following NPLOC4-KO plotted as –log10 FDR (y-axis) versus log2 fold change of NPLOC4-KO versus control (x-axis). Blue or red points are genes with < −1.5 or >1.5 fold change, respectively, at an FDR of <0.05. G, Reactome network of genes downregulated following NPLOC4-KO. H, Pathway enrichment of genes in the Reactome network of downregulated genes in F.

Close modal

RELA (also known as p65) is an NFκB subunit that regulates critical downstream effectors of immunosuppressive pathways in tumors (60, 61). NPLOC4 mediates nuclear pore transport of proteins, but its role in cancer or immune modulation remains unclear. To elucidate the mechanism by which these genes mediate GSC sensitivity to CAR T-cell killing, we performed in-depth characterization of GSCs harboring KO of each gene. The increased sensitivity was not a result of alterations in target antigen expression on GSCs (Supplementary Fig. S6C). CAR T cells induced PD-L1 in GSCs, which was not altered by depletion of either RELA or NPLOC4 (Supplementary Fig. S6D). Likewise, CAR T cells cocultured with GSCs transduced with sgCONT, sgRELA, or sgNPLOC4 did not show differences in activation after stimulation as indicated by markers CD69 and CD137, or exhaustion measured by levels of exhaustion markers, including PD-1, LAG3, and TIM3 (ref. 30; Supplementary Fig. S6E). Whole-transcriptome analysis of GSCs after RELA-KO showed downregulation of immunosuppressive cytokines, including CXCL3, CCL20, and IL32 (Fig. 6C), all of which suppress antitumor immune responses (62, 63). Downregulated genes were highly enriched for known direct transcriptional targets of RELA, and RELA-KO reduced NFκB signaling, as well as the immunosuppressive effectors of TNF responsiveness and IL10 signaling (Fig. 6D and E; Supplementary Fig. S6F). Targeting NPLOC4 in GSCs downregulated genes mediating rearrangement of extracellular matrix (ECM), including proteoglycans, integrins, and collagens (Fig. 6FH). Reactome analysis revealed the involvement of specific tumorigenic factors, such as EGFR and PDGFA (Fig. 6G). Pathways downregulated after NPLOC4 depletion were highly enriched for ECM remodeling and cell adhesion (Fig. 6H; Supplementary Fig. S6G). Although tumor ECM remodeling has been reported to suppress antitumor immune responses by preventing T-cell trafficking into the tumors, ECM-associated factors may directly repress T-cell activity (64, 65). To interrogate NPLOC4 interactions, we performed immune precipitation followed by mass spectrometry (IP/MS), revealing that NPLOC4 bound multiple targets in immune-related pathways (IL1, Fc receptor, antigen presentation), WNT signaling, and protein synthesis/degradation pathways (Supplementary Fig. S7A). These mechanisms may regulate the immune-related profiles of GSCs, where NPLOC4-KO led to the upregulation of immune stimulatory cytokines (Supplementary Fig. S7B and S7C). Together, we found that the tumor-intrinsic regulators RELA and NPLOC4 mediate GBM resistance to CAR T-cell cytotoxicity via mechanisms distinct from induction of CAR T-cell exhaustion.

CRISPR Screening Identified Targets with Functional and Clinical Relevance in GSCs

Next, we used an orthotopic intracranial patient-derived xenograft (PDX) model to evaluate whether modulating the identified targets on GSCs enhanced the antitumor function of CAR T cells in a preclinical setting. Established GBM PDXs were treated with CAR T cells delivered intracranially into the tumors, mimicking our clinical trial design of CAR T-cell administration to patients with GBMs (7, 8). First, we used CAR T cells without CRISPR KO to treat control, RELA-KO, or NPLOC4-KO tumors. A limited number of CAR T cells (50,000/mouse) completely eradicated xenografts derived from RELA-KO or NPLOC4-KO GSCs, whereas the same CAR T cells were only partially effective against tumors established with sgCONT GSCs (Fig. 7A and B; Supplementary Fig. S8). These results suggest that tumors with low expression of RELA and/or NPLOC4 are more sensitive to CAR T therapy.

Figure 7.

Functional and clinical relevance of targets on GSCs and CAR T cells. A and B, Kaplan–Meier survival curves comparing mouse survival for RELA (A) or NPLOC4 (B) KO with nontargeting controls. Tumors were established by orthotopically implanting 2 × 105 PBT030-2 GSCs, and treated after 8 days with 5 × 104 CAR T cells. P values were shown comparing each group with “sgCONT + CAR” group using the log-rank test. C, Left, correlation of RELA expression with immune and T-cell signatures in TCGA GBM RNA-seq data. Right, scatter plot of lymphocyte infiltration signature score versus RELA expression by tumor from TCGA GBM RNA-seq data. D, Left, correlation of NPLOC4 expression with immune and T-cell signatures in TCGA GBM RNA-seq data. Right, scatter plot of NPLOC4 expression versus wound-healing signature score by tumor from TCGA GBM RNA-seq data. C and D,P values were calculated as Pearson correlation coefficients. E and F, Kaplan–Meier curves demonstrating prolonged survival in an intracranial xenograft model of GBM treated with TLE4-KO (E) or IKZF2-KO (F) CAR T cells (blue) compared with nontargeting control (black). Tumors were established by orthotopically implanting 2 × 105 PBT030-2 GSCs, and treated after 8 days with 2 × 104 CAR T cells. P values were shown comparing each group with the “CAR sgCONT” group using the log-rank test. G, Fold change after stimulation of genes significantly upregulated (FDR < 0.05, log2 fold change > 1) following IKZF2-KO after tumor stimulation, in an independent data set of clinical CAR T-cell products from patients with CLL. H, Fold change after stimulation of genes enriched in cluster 10 (as shown in Fig. 3A). I and J, Fold change after stimulation of genes significantly upregulated (FDR < 0.05, log2 fold change > 1) following IKZF2 (I) or TLE4 (J) KO. G–J, CAR T cells were stratified by response of the patient from which they were derived—complete responders or nonresponders—to CAR therapy and the log2 fold change of stimulated versus mock-stimulated gene expression was plotted; P values were calculated by unpaired Student t tests.

Figure 7.

Functional and clinical relevance of targets on GSCs and CAR T cells. A and B, Kaplan–Meier survival curves comparing mouse survival for RELA (A) or NPLOC4 (B) KO with nontargeting controls. Tumors were established by orthotopically implanting 2 × 105 PBT030-2 GSCs, and treated after 8 days with 5 × 104 CAR T cells. P values were shown comparing each group with “sgCONT + CAR” group using the log-rank test. C, Left, correlation of RELA expression with immune and T-cell signatures in TCGA GBM RNA-seq data. Right, scatter plot of lymphocyte infiltration signature score versus RELA expression by tumor from TCGA GBM RNA-seq data. D, Left, correlation of NPLOC4 expression with immune and T-cell signatures in TCGA GBM RNA-seq data. Right, scatter plot of NPLOC4 expression versus wound-healing signature score by tumor from TCGA GBM RNA-seq data. C and D,P values were calculated as Pearson correlation coefficients. E and F, Kaplan–Meier curves demonstrating prolonged survival in an intracranial xenograft model of GBM treated with TLE4-KO (E) or IKZF2-KO (F) CAR T cells (blue) compared with nontargeting control (black). Tumors were established by orthotopically implanting 2 × 105 PBT030-2 GSCs, and treated after 8 days with 2 × 104 CAR T cells. P values were shown comparing each group with the “CAR sgCONT” group using the log-rank test. G, Fold change after stimulation of genes significantly upregulated (FDR < 0.05, log2 fold change > 1) following IKZF2-KO after tumor stimulation, in an independent data set of clinical CAR T-cell products from patients with CLL. H, Fold change after stimulation of genes enriched in cluster 10 (as shown in Fig. 3A). I and J, Fold change after stimulation of genes significantly upregulated (FDR < 0.05, log2 fold change > 1) following IKZF2 (I) or TLE4 (J) KO. G–J, CAR T cells were stratified by response of the patient from which they were derived—complete responders or nonresponders—to CAR therapy and the log2 fold change of stimulated versus mock-stimulated gene expression was plotted; P values were calculated by unpaired Student t tests.

Close modal

To further dissect the roles of RELA and NPLOC4 in immune modulation in GBM, we analyzed 41 GSC samples and found that high RELA- or NPLOC4-expressing GSCs showed enrichment in immune-suppression signatures (Supplementary Fig. S9A and S9B). Interrogating The Cancer Genome Atlas (TCGA) GBM data set revealed that RELA and NPLOC4 both positively correlated with TGFβ signaling, a key pathway mediating immune suppression in GBMs and many other types of tumors (66). RELA was also positively correlated with immunosuppressive regulatory T-cell signatures and negatively correlated with the signatures of antitumor T-cell responses (lymphocyte infiltration, TCR richness, Th1 and CD8 T cells; Fig. 7C). NPLOC4 was negatively correlated with the immune stimulatory IFNγ responses (Fig. 7D). The infiltration signature of CD4+ and CD8+ T cells in GBM inversely correlated with RELA or NPLOC4 expression (Supplementary Fig. S9C and S9D). These results suggest that high expression of RELA and NPLOC4 in GBM is indicative of a more suppressive tumor immune microenvironment and repressed antitumor T-cell responses.

CRISPR Screening Identified Targets with Functional and Clinical Relevance in CAR T Cells

We next evaluated the molecular targets identified in our CAR T-cell screen in vivo, with the goal of establishing clinically translatable strategies to improve CAR T-cell function. The antitumor function of different CAR T cells was tested against tumors without CRISPR knockouts, with a further limited CAR T-cell dose (20,000/mouse) showing enhanced survival benefit as compared with the control CAR T cells that failed to achieve long-term tumor eradication (Fig. 7E and F). Consistent with improved maintenance of T-cell effector activity and decreased exhaustion, targeting either TLE4 or IKZF2 augmented in vivo antitumor activity of CAR T cells against PDXs, as measured by extension of survival in tumor-bearing mice (Fig. 7E and F; Supplementary Fig. S10). Depletion of TMEM184B or EIF5A in CAR T cells showed a trend toward improved efficacy in increasing the survival of tumor-bearing mice (Supplementary Fig. S11A and S11B). Therefore, these targets on GSCs and CAR T cells can be exploited to advance the efficacy of CAR therapy against established GBM tumors.

We then investigated whether the CAR T-cell targets indicate the potency of clinical therapeutic products. We mapped upregulated genes in IKZF2-KO CAR T cells compared with control CAR T cells after tumor stimulation, with the transcriptomes of CAR T-cell products from patients with chronic lymphocytic leukemia (CLL) achieving complete responses (CR) or no responses (NR; ref. 27). Supporting our results, these genes were induced to a greater degree after CAR stimulation in the products from patients achieving CR (Fig. 7G). Similarly, genes enriched in cluster 10, whose expansion was induced by tumor stimulation and further augmented with TLE4-KO, were also highly expressed in the products from patients with CR (Fig. 7H). Further, both TLE4-KO and IKZF2-KO led to gene upregulation similar to comparisons of products from patients with CR and NR (Fig. 7I and J).

To further understand how TLE4 and IKZF2 contribute to the function of clinical CAR T-cell products, we analyzed scRNA-seq from 24 patient-derived CD19-CAR T-cell products (67). An unbiased clustering of the scRNA-seq data revealed that IKZF2 expression was highly enriched in cluster 7 (Supplementary Fig. S12A and S12B), overlapping with key markers of immune-suppressive Tregs (CTLA4, FOXP3, IL2RA; Supplementary Fig. S12C and S12D). Cluster 7 was more frequently detected in patients with progressive disease (PD) than those with CR (Supplementary Fig. S12E). In these same cells, cluster 11 represented exhausted T cells, as indicated by the markers TOX and TOX2 (Supplementary Fig. S13A–S13C; ref. 37). This cluster showed low expression of TLE4-repressed genes, indicating high TLE4 activity (Supplementary Fig. S13D and S13E). Further, TLE4 was upregulated in CAR T cells undergoing extended ex vivo culture (Supplementary Fig. S13F), a process associated with impaired effector function (68). Together, these observations establish that the targets identified from CRISPR screening have clinical implications for both tumor immunoreactivity and CAR T-cell functional potency (Supplementary Fig. S14).

T cell–based therapies may offer several advantages in GBM therapy. T cell–based therapies, especially when delivered into the cerebrospinal fluid (CSF), traffic to multifocal tumor populations within the central nervous system (CNS; refs. 8, 69–71), thus overcoming challenges associated with the blood–brain barrier that limit the CNS penetration of most pharmacologic agents. T-cell therapies compensate for cellular plasticity within brain tumors more effectively than traditional pharmacologic agents. GBMs display striking intratumoral heterogeneity, and tumor cells readily compensate for targeted agents against specific molecular targets. With T-cell therapy targeting different antigens, personalized treatments based on the antigen expression profile of individual tumors may be designed. T cell–based therapies induce secondary responses that augment endogenous antitumor responses. Adoptive cell transfer, especially CAR T therapies, has been investigated in clinical trials for patients with GBM, but efficacy has been restricted to limited cases (11). Our focus on CAR T cells was prompted not only by the potential value for clinical translation, but also because our findings inform a broader understanding of T-cell function in brain tumor biology.

Previous genetic screens used to identify interactions between immune cells and tumor cells have largely focused on the tumor cells (18, 19, 29), as these cells are easier to manipulate genetically. Screens on tumor-reactive mouse T cells have also been reported (20, 72, 73) given the establishment of Cas9-knock-in mouse strain (74), as well as the convenience to acquire large numbers of these cells. Here, we interrogated both the human CAR T-cell and tumor cell compartments. The screening strategy on CAR T cells was greatly facilitated by the development of the nonviral Cas9 expression system in primary human T cells (21). Here, the screening on tumor cells was performed on two independent GSCs, displaying a relatively narrow range of shared molecular targets involved in mediating responses to CAR T cells in our studies, which might be a consequence of subtype difference between these GSC lines (33). The screening identified both rational targets (RELA/p65) and novel targets (NPLOC4) in immune regulation, which were not restricted to a specific GBM molecular subclass. NPLOC4 displayed unexpected associations with GBM-targeting immune cell activity, as NPLOC4-KO in GSCs led to enhanced potency of CAR T cells and increased cytokine production in GSCs, although the detailed mechanism awaits further investigation. In the analyses of GSC models and the TCGA database, high RELA and NPLOC4 expression was associated with immuno-suppressive signatures. More specifically, higher expression of RELA and NPLOC4 in GBMs correlated with low infiltration of both CD4+ and CD8+ T cells, indicating that targeting these genes may confer immune modulatory effect and enhance antitumor T-cell responses in GBMs.

The assay used for CRISPR screening in T cells is crucial for reliable readouts and is required for its sensitivity to differentiate effective versus noneffective therapies. Although the in vivo antitumor efficacy in mouse models has been the standard to evaluate the functional quality of T cells in adoptive transfer, the utilization of this system in screening has been controversial. Tumor-infiltrating T cells harvested after the injection of therapeutic cells display signatures of tumor reactivity (72) or, conversely, T-cell exhaustion (40). The differential results appear model-dependent, leading to mixed interpretation of the results. The coculture assays that we used in this study identified key regulators by creating challenging screening environments. For the screening on GSCs, two rounds of short-term (24 h) killing with a relatively large number of T cells (total E:T = 1:1) were performed and GSCs were harvested immediately after the second round of killing, minimizing the effect of knocking out genes essential for the GSC growth. For the screening on CAR T cells, a repetitive challenge assay was used with an excessive number of GSCs (total E:T = 1:12), which we have shown to induce CAR T-cell exhaustion (30). The screen was performed by comparing a less exhausted (PD-1 negative) with a more exhausted (PD-1 positive) subset, informing prioritization for maintenance of recursive killing function, while reducing the noise from tumor cell or T-cell growth. The screening was performed with two independent CAR T-cell donors, and the relatively small proportion of overlapping hits between the two donors was expected and consistent with previous studies (21, 75), due to the variation in T-cell populations between individuals. The target validation was done with different T-cell donors and CAR platforms; therefore, the discovered immunotherapy targets may be generalizable to multiple CAR designs. Although we validated four representative genes, the screening on CAR T cells resulted in more than 200 potential targets involved in critical pathways of T-cell biology and activation, offering additional targets for future investigation of CAR refinement. One limitation of our approach, however, is the exclusion of apoptosis pathways in tumor cells due to their critical role in tumor cell growth, which have been demonstrated as important regulators of CAR T-cell–mediated tumor killing as well as tumor-induced CAR T-cell exhaustion (29).

T-cell exhaustion has been considered one of the major hurdles for reducing CAR T-cell potency (76–78). Blocking/KO of inhibitory receptors is being rigorously investigated to augment CAR activity or other tumor-targeting T cells (29, 79, 80). T-cell exhaustion is a feedback mechanism after activation, occurring upon recursive exposure to antigens in the contexts of chronic infection or the tumor microenvironment (77, 81), compromising their antitumor potency (78). Here, we observed that TLE4-KO or IKZF2-KO resulted in unstimulated CAR T cells to express transcriptional profiles of activation, while prohibiting exhaustion. The AP-1 family transcription factors FOS and JUN, which were induced after both TLE4-KO and IKZF2-KO, provide a possible mechanism by which CAR T-cell fitness was protected. The protein c-JUN forms homodimers or c-FOS/c-JUN heterodimers to initiate transcription of proinflammatory cytokines, and heterodimers with other cofactors (including BATF, IRF4, JUNB, and JUND) induce inhibitory receptors or suppress transcriptional activity of c-JUN (82–85). FOS was more upregulated than suppressive cofactors after TLE4-KO, therefore driving T-cell activation together with a protection from exhaustion, which was reminiscent of the effect after expressing c-JUN in CAR T cells with tonic signaling (55). In IKZF2-KO cells, however, the uncoupling of activation from exhaustion signatures was likely influenced by the upregulation of cytokines CCL3 and CCL4, which inversely correlated with PD-1 expression during T-cell exhaustion (86). Both TLE4-KO and IKZF2-KO in CAR T cells upregulated essential regulators for Th1 cell differentiation (BCAT and EGR1, respectively), consistent with a previously identified role of this T-cell population in mediating antitumor immunity (87, 88). Consequently, targeted KOs in CAR T cells enhanced not only killing, but also expansion potential, which is correlated with clinical responses (89). Although it remains unresolved if these KOs potentiate CAR activity in immune-competent settings, our results have revealed the feasibility that CAR T cells can be modified for their activation/exhaustion signals to achieve functional improvement in clinically relevant models. Consistent with these findings, we explored public databases of scRNA-seq on patient-derived CAR T-cell products and discovered that high IKZF2 expression and TLE4 activity were associated with other suppressive/exhaustion signatures of CAR T cells as well as poor clinical responses.

Single-cell analyses reveal subset composition within a mixed cell sample, such as CAR T cells, in which minority populations serve critical roles. scRNA-seq revealed that CAR activation, rather than genetic modification of CAR T cells (TLE4-KO or IKZF2-KO), resulted in a major cluster switch, which is consistent with the observation that TLE-KO or IKZF2-KO in monoculture CAR T cells did not dramatically alter transcriptional profiles, as suggested by bulk RNA-seq. Following tumor challenge, KO of targeted genes upregulated T-cell activation markers and proinflammatory cytokines across different clusters, especially IFNG and CCL3, which showed similar induction by both TLE-KO and IKZF2-KO. Further, after CAR activation, TLE4-KO maintained a specific cluster, which existed preactivation, and IKZF2-KO led to the emergence of a new cluster. The transcriptional signature of these clusters (expression of several costimulation molecules and cytokines) indicated their critical role in mediating effector function of CAR T cells. Therefore, the superior functions of TLE4-KO or IKZF2-KO CAR T cells were likely the result of a generally elevated activation state, as well as the stimulatory effect from critical subsets. Our scRNA-seq results also suggested the existence of Treg-like populations, the expansion of which was seen after CAR activation and can be reduced by IKZF2-KO. The suppressive function of these cells still requires further investigation, but these results indicate the potential of enhancing CAR function through inhibiting differentiation toward Treg-like cells. Both TLE4-KO and IKZF2-KO CAR T cells appear to modify specific CD4+ T-cell subsets, which supports our previous observation that CD4+ CAR T cells play a critical role in mediating potent effector function (30).

In summary, whole-genome CRISPR KO screens identified a limited set of key targets on GSCs that promote resistance to CAR killing. Modifying these genes before CAR T-cell infusions, with genetic or chemical approaches, might be a potential strategy to enhance CAR T-cell antitumor function and also guide patient selection. The spectrum of identified CAR T targets in our screen was broad, but manipulating expression of several of these targets appeared to regulate shared transcriptional programs, including cytokines and AP-1 regulation. Given the exciting advances in CAR T engineering with logic gates, it should be possible to interrogate multiple molecular targets within CAR T cells to determine potential synergy or additional benefit. Although our screens and evaluations were mainly focused on GSCs, the results are also expected to be consistent non-GSC populations, as CAR-mediated cytotoxicity has been demonstrated not to bias between antigen-positive GSCs and more differentiated tumor cells (4–6). Although further investigation is needed to demonstrate the function of these targets in T cells harboring other CAR constructs, our results found that most of the targets selected for deeper studies regulate core pathways in immune cell function, suggesting that these targets likely have broad regulatory function in immune cell antitumor function in various contexts. Future studies will also permit interrogation of critical targets in an immune-intact environment, but our current focus was to interrogate molecular regulators directly involved in CAR T killing of the therapeutically resistant human GSC population. Overall, the bidirectional screening provides a platform to robustly identify targets with the potential toward clinical benefit in CAR T and other adoptive cell transfer therapies.

Lentiviral Transduction on GSCs

GSCs were acquired from patient specimens at City of Hope under protocols approved by the Institutional Review Board (IRB), and maintained as tumorspheres in GSC media as previously described (4, 90). GSC lines used in this study to test CAR T-cell function are IDH1/2 wild-type. The sgRNA library and single-targeted sgRNA lentiviral plasmids (containing a puromycin-resistant gene) for GSC transduction were purchased from Addgene (#73179 and #52961, respectively). Lentiviral particles were generated as previously described (91). For lentiviral transduction, GSC tumorspheres were dissociated into single cells using Accutase (Innovative Cell Technologies), resuspended in GSC media, and lentivirus was added at a 1:50 v/v ratio. GSCs were then washed once after 12 hours, resuspended in fresh GSC media, and cultured for three days. To ensure that only transduced cells were expanded for further assays, GSCs were selected by puromycin (Thermo Fisher Scientific) for seven continuous days, with a 1:10,000 v/v ratio into GSC media.

Lentiviral Transduction on Primary Human T Cells

Naïve and memory T cells were isolated from healthy donors at City of Hope under protocols approved by the IRB (26, 30). The constructs of IL13Rα2-targeted and HER2-targeted CARs were described in previous studies (8, 26, 92). Procedures of CAR-only transduction on primary human T cells were previously described (44). The sgRNA library and single-targeted sgRNA lentiviral plasmids for T-cell transduction were purchased from Addgene (#73179 and #52961, respectively). All sgRNA plasmids contain a puromycin-resistant gene. Dual transduction of CAR and sgRNA was performed using modification of previously reported procedures (21). In brief, primary T cells were stimulated with Dynabeads Human T expander CD3/CD28 (Invitrogen; T cells:beads = 1:2) for 24 hours and transduced with sgRNA lentivirus (1:250 v/v ratio). Cells were washed after 6 hours and then transduced with CAR lentivirus [multiplicity of infection (MOI) = 0.5]. Four days after CAR transduction, CD3/CD28 beads were removed and cells were resuspended in Lonza electroporation buffer P3 (Lonza, #V4XP-3032; 2 × 108 cells/mL). Cas9 protein (MacroLab; 40 μmol/L stock) was then added to the cell suspension (1:10 v/v ratio), and electroporation was performed using a 4D-Nucleofactor Core Unit (Lonza, #AAF-1002B). Cells were recovered in prewarmed X-VIVO 15 media (Lonza) for 30 minutes before proceeding to ex vivo expansion. All T-cell transduction and ex vivo expansion experiments were performed in X-VIVO 15 containing 10% FBS, 50 U/mL recombinant human IL2 (rhIL2), and 0.5 ng/mL rhIL15, at 6 × 105 cells/mL. To ensure that only sgRNA-transduced cells were expanded, puromycin (1:10,000 v/v ratio) was added to the media three days after electroporation, and puromycin selection was performed for six continuous days before CAR T cells were used for further assays. CRISPR screening was performed on two independent donors, and two other donors were used to generate IL13Rα2-targeted and HER2-targeted CARs, respectively.

CRISPR Screening on GSCs

GSCs transduced with the CRISPR KO library were dissociated into single cells, and cocultured with CAR T cells at an E:T ratio of 1:2 in culture plates precoated with Matrigel. After 24 hours, the media containing CAR T cells and tumor debris were removed, and the same number of CAR T cells were added in fresh media. Twenty-four hours after the second CAR T-cell addition, the media were removed and the remaining GSCs were washed with PBS and harvested. Genomic DNA was isolated from the remaining GSCs after coculture with CAR T cells, as well as GSCs harvested before coculture and GSCs after monoculture for 48 hours.

CRISPR Screening on CAR T Cells

T cells transduced with CAR and the CRISPR KO library were cocultured with GSCs at an E:T ratio of 1:4 in culture plates precoated with Matrigel. After 48 hours, CAR T cells were rechallenged by GSCs doubling the number of the initial coculture. Twenty-four hours after the rechallenge, the coculture was harvested and stained with fluorescence-conjugated antibodies against human CD45 (BD Biosciences; catalog no. 340665, RRID:AB_400075), PD-1 (BioLegend; catalog no. 329922, RRID:AB_10933429), and IL13 (BioLegend; catalog no. 501914, RRID:AB_2616746). Different subsets were sorted using an Aria SORP (BD Biosciences): total CAR T cells (CD45+, IL13+), PD-1+ CAR T cells (CD45+, IL13+, PD-1+), and PD-1 CAR T cells (CD45+, IL13+, PD-1). Genomic DNA was isolated from the sorted subsets of cells, as well as CAR T cells harvested before coculture and CAR T cells after monoculture for 72 hours.

CRISPR/Cas9 Screen Analysis

FASTQ files were trimmed to 20-bp CRISPR guide sequences using BBDuk from the BBMap (https://jgi.doe.gov/data-and-tools/bbtools; RRID:SCR_016965) toolkit, and quality control was performed using FastQC (RRID:SCR_014583, https://www.bioinformatics.babraham.ac.uk/projects/fastqc/). FASTQs were aligned to the library and processed into counts using the MAGECK-VISPR “count” function (https://bitbucket.org/liulab/mageck-vispr/src/master/). β values were calculated using a Maximum Likelihood Estimation model generated independently for each comparison. Nontargeting sgRNAs were used to derive a null distribution to determine P values.

In Vitro Cytotoxicity and Flow Cytometry Assays

For in vitro cytotoxicity test, CAR T cells were cocultured with GSCs at an E:T ratio of 1:40. After 48 hours of coculture, the numbers of CAR T cells and GSCs were evaluated by flow cytometry. Flow cytometry assays were performed on GSCs, CAR T cells from monoculture or coculture with procedures described previously (30). For coculture, anti-CD45 (BD Biosciences; catalog no. 340665, RRID:AB_400075) staining was used to distinguish GSCs with T cells, and CAR T cells were identified by anti-IL13 (BioLegend; catalog no. 501914, RRID:AB_2616746) staining. Other antibodies used for flow cytometry are listed as follows: PD-L1 (Thermo Fisher Scientific, catalog no. 17-5983-42, RRID:AB_10597586), TIM3 (Thermo Fisher Scientific; catalog no. 17-3109-42, RRID:AB_1963622), LAG3 (Thermo Fisher Scientific; catalog no. 12-2239-41, RRID:AB_2572596), PD-1 (BioLegend; catalog no. 329922, RRID:AB_10933429), CD69 (BD Biosciences; catalog no. 340560, RRID:AB_400523), CD137 (BD Biosciences; catalog no. 555956, RRID:AB_396252), and IL13Rα2 (BioLegend; catalog no. 354404, RRID:AB_11218789). All samples were analyzed via a Macsquant Analyzer (Miltenyi Biotec) and processed via FlowJo v10 (RRID:SCR_008520).

RNA-seq Analysis

Total mRNA from GSCs or CAR T cells was isolated and purified by the RNeasy Mini Kit (Qiagen Inc.) and sequenced with Illumina protocols on a HiSeq 2500 to generate 50-bp reads. Trim Galore (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/; RRID:SCR_011847) was used to trim adaptors and remove low-quality reads. Reads were quantified against Gencode v29 using Salmon (RRID:SCR_017036, https://combine-lab.github.io/salmon/) with correction for fragment-level GC bias, positional bias, and sequence-specific bias. Transcripts were summarized to gene level and processed to transcripts per million (TPM) using the R/Bioconductor (https://www.bioconductor.org/) package DESeq2 (RRID:SCR_000154, https://bioconductor.org/packages/release/bio-c/html/DESeq2.html). Comparisons were performed using contrasts in DESeq2 followed by Benjamini–Hochberg adjustment to correct for false discovery rate (FDR).

Gene Set Enrichment Analysis

ClueGO gene-set enrichment plots were generated using the ClueGO plugin (http://apps.cytoscape.org/apps/cluego, RRID:SCR_005748) for GO BP, KEGG, or Reactome gene sets and visualized in Cytoscape v3.7.2 (https://cytoscape.org/). Gene set enrichment analysis (GSEA; RRID:SCR_003199) plots were generated from preranked lists using the mean β value as the ranking metric. Reactome networks were created using the Reactome FI plugin (https://reactome.org/tools/reactome-fiviz) with network version 2018 and visualized in Cytoscape. Networks were clustered using a built-in network clustering algorithm, which utilizes spectral partition-based network clustering, and node layout and color were determined by module assignment. GSEA plots from RNA-seq data were generated from preranked lists. Weighting metrics for preranked lists were generated using the DESeq2 results from the gene knockdown versus nontargeting control and applying the formula: −log10(FDR) × log2(fold change). ssGSEA scores for specific immune or functional pathways were generated using the ssGSEA function from the R/Bioconductor package GSVA (https://bioconductor.org/packages/release/bioc/html/GSVA.html; ref. 93) and plotted using pheatmap (https://cran.r-project.org/web/packages/pheatmap/). ChEA enrichments were performed using Enrichr (https://amp.pharm.mssm.edu/Enrichr/). Bar plots for positive or negative gene set enrichments were performed using Metascape (https://metascape.org/gp/index.html) for significantly upregulated or downregulated genes (FDR <0.05 and log2 fold change >1 or < −1).

Reactome Networks and KEGG Pathways

Reactome networks were derived from RNA-seq data using the Cytoscape Reactome FI plugin (RRID:SCR_003032). A gene list of upregulated (FDR <0.05 and log2 fold change >1) or downregulated (FDR <0.05 and log2 fold change < −1) genes plus the target gene (as KO by CRISPR/Cas9 would not be detected by RNA-seq) was input into Reactome FI, and all genes with at least one edge were included in the network plot. Node color (light to dark) and size (small to large) are proportional to node degree. Pathway enrichment was performed on this network of genes using the Reactome FI enrichment option. Box plots for genes from selected pathways were generated using RNA-seq TPM data. KEGG pathway visualizations were generated using the R/Bioconductor package pathview (https://www.bioconductor.org/packages/release/bioc/html/pathview.html) for selected pathways, and genes were colored on the basis of the log2 fold change KO versus control.

scRNA-seq

scRNA-seq files were processed using the Cell Ranger workflow (https://support.10xgenomics.com/single-cell-gene-expression/software/overview/welcome). FASTQ files were generated using the Cell Ranger “mkfastq” command with default parameters. FASTQs were aligned to the hg19 genome build using the “count” function and aggregated using the default Cell Ranger “aggr” parameters with normalization performed by subsampling wells to equalize read depth across cells. Downstream analyses were performed using the R/Bioconductor package Seurat (https://satijalab.org/seurat/; ref. 94). Specifically, data sets of stimulated and unstimulated cells in KO or control populations were merged using the “FindIntegrationAnchors” Seurat function. Clustering was performed using UMAP using PCA for dimensional reduction and a resolution of 0.6 from 1 to 20 dimensions. Dead cell clusters were determined by high expression of mitochondrial genes and removed. Samples were then reclustered. Clusters with similar CD4 or CD8, Ki-67, and marker expression, determined using the “FindAllMarkers” function that were proximal on the UMAP projection, were merged. All plots for gene expression were generated using normalized data from the default parameters of the “NormalizeData” function. Gene expression was visualized on the UMAP projection using the “FeaturePlot” function with a maximum cutoff or gene expression determined on a gene-by-gene basis.

Functional Analysis of CAR T Cells in Orthotopic GBM Models

All mouse experiments were performed using protocols approved by the City of Hope Institutional Animal Care and Use Committee. Orthotopic GBM models were generated using 6- to 8-week-old NOD/SCID/IL2R−/− (NSG) mice (IMSR; catalog no. JAX:005557, RRID:IMSR_JAX:005557), as previously described (95). Briefly, ffLuc-transduced GSCs (1 × 105/mouse) were stereotactically implanted (intracranially) into the right forebrain of NSG mice. Randomization was performed after 8 days of tumor injection based on bioluminescence signal, and mice were then treated intracranially with CAR T cells (2 × 104 or 5 × 104/mouse as indicated for each experiment). To ensure statistical power, all treatment groups included ≥6 animals. Mice were monitored by the Department of Comparative Medicine at City of Hope for survival and any symptoms related to tumor progression, with euthanasia applied according to the American Veterinary Medical Association Guidelines. Studies were done in both male and female animals. Investigators were not blinded for randomization and treatment.

TCGA Data Analysis

Analysis of genes in the TCGA data set was performed using RNA-seq TCGA GBM data. Immune infiltration signatures were previously reported (96). GSEA plots for each gene in the context of TCGA GBM data were generated using the normalized gene expression as a continuous phenotype.

CAR T-cell Responder Analysis

Gene sets derived from TLE4-KO or IKZF2-KO were analyzed in the context of CAR T-cell nonresponders versus responders from a previous report on patients with CLL (27). Genes upregulated in bulk RNA-seq of CAR T cells following KO of TLE4 or IKZF2 (FDR <0.05 and log2 fold change > 1) were plotted by their fold change expression in stimulated versus unstimulated CAR T cells for responders or nonresponders. Fold change was calculated using DESeq2 for stimulated versus unstimulated cells independently for each group (nonresponder or complete responder). Cluster 10–enriched genes in the TLE4-KO and control sc-seq data, identified by the “FindAllMarkers” function in Seurat subsetted for overexpressed genes, were plotted similarly. Genes upregulated (>0.4 log2 fold change of normalized counts) in sc-seq for IKZF2-KO versus control in stimulated CAR T cells were plotted similarly.

Statistical Analysis

CAR T-cell functional data (tumor killing, expansion, survival of tumor-bearing mice) were analyzed via GraphPad Prism. Group means ± SEM were plotted. Methods of P value calculations are indicated in figure legends.

Genomic Data Reporting and Sharing

Both bulk RNA-seq and scRNA-seq data sets have been deposited to the Gene Expression Omnibus (GSE163409) and will be accessible to the public at the time of publication.

D. Wang reports grants from NCI during the conduct of the study; in addition, D. Wang has a patent for 63/117439 pending. R.C. Gimple reports grants from NIH during the conduct of the study. M.H. Lorenzini reports personal fees from Kite, a Gilead company, outside the submitted work. B. Badie reports grants from NIH and grants and personal fees from Mustang Therapeutics during the conduct of the study; in addition, B. Badie has a patent for CAR T-cell delivery pending. C.E. Brown reports grants from Ivy Foundation, CIRM, and grants from NIH during the conduct of the study; personal fees from Mustang Bio outside the submitted work; in addition, C.E. Brown has a patent for 63/117439 pending. J.N. Rich reports grants and personal fees from Synchronicity and other support from Function Oncology outside the submitted work. No disclosures were reported by the other authors.

D. Wang: Conceptualization, data curation, formal analysis, validation, investigation, visualization, methodology, writing–original draft, writing–review and editing. Q. Wu: Validation and methodology. L.J.Y. Kim: Software, formal analysis, and visualization. Z. Qiu: Software, formal analysis, and visualization. P. Lin: Methodology. M.H. Lorenzini: Writing–original draft. B. Badie: Resources. S.J. Forman: Resources and supervision. Q. Xie: Conceptualization, data curation, formal analysis, supervision, validation, investigation, visualization, methodology, writing–original draft, writing–review and editing. C.E. Brown: Conceptualization, resources, data curation, supervision, funding acquisition, validation, investigation, visualization, writing–original draft, project administration, writing–review and editing. J.N. Rich: Conceptualization, resources, data curation, supervision, funding acquisition, investigation, visualization, methodology, project administration, writing–review and editing. B.C. Prager: Data curation, software, formal analysis, validation, investigation, visualization, methodology, writing–original draft, writing–review and editing. R.C. Gimple: Data curation, validation, and methodology. B. Aguilar: Data curation, validation, investigation, and visualization. D. Alizadeh: Validation, investigation, and visualization. H. Tang: Data curation, software, formal analysis, and methodology. D. Lv: Data curation, investigation. R. Starr: Data curation, validation, and methodology. A. Brito: Data curation and methodology.

This study was supported by the following funding sources: NIH: CA234923 (to D. Wang), CA217066 (to B.C. Prager), CA217065 (to R.C. Gimple), CA203101 (to L.J.Y. Kim), CA236500 (to C.E. Brown), CA238662, CA197718, NS203434 (to J.N. Rich); Ivy Foundation (to C.E. Brown); and Westlake Education Foundation (to Q. Xie). C.E. Brown receives royalty payments from Mustang Bio. Patent associated with this study has been submitted (application serial number: 63/117439).

The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

1.
Lim
M
,
Xia
Y
,
Bettegowda
C
,
Weller
M
. 
Current state of immunotherapy for glioblastoma
.
Nat Rev Clin Oncol
2018
;
15
:
422
42
.
2.
Maude
SL
,
Laetsch
TW
,
Buechner
J
,
Rives
S
,
Boyer
M
,
Bittencourt
H
, et al
Tisagenlecleucel in children and young adults with B-cell lymphoblastic leukemia
.
N Engl J Med
2018
;
378
:
439
48
.
3.
Neelapu
SS
,
Locke
FL
,
Bartlett
NL
,
Lekakis
LJ
,
Miklos
DB
,
Jacobson
CA
, et al
Axicabtagene ciloleucel CAR T-cell therapy in refractory large B-cell lymphoma
.
N Engl J Med
2017
;
377
:
2531
44
.
4.
Brown
CE
,
Starr
R
,
Aguilar
B
,
Shami
AF
,
Martinez
C
,
D'Apuzzo
M
, et al
Stem-like tumor-initiating cells isolated from IL13Ralpha2 expressing gliomas are targeted and killed by IL13-zetakine-redirected T Cells
.
Clin Cancer Res
2012
;
18
:
2199
209
.
5.
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
.
Clin Cancer Res
2010
;
16
:
474
85
.
6.
Morgan
RA
,
Johnson
LA
,
Davis
JL
,
Zheng
Z
,
Woolard
KD
,
Reap
EA
, et al
Recognition of glioma stem cells by genetically modified T cells targeting EGFRvIII and development of adoptive cell therapy for glioma
.
Hum Gene Ther
2012
;
23
:
1043
53
.
7.
Brown
CE
,
Badie
B
,
Barish
ME
,
Weng
L
,
Ostberg
JR
,
Chang
WC
, 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
72
.
8.
Brown
CE
,
Alizadeh
D
,
Starr
R
,
Weng
L
,
Wagner
JR
,
Naranjo
A
, et al
Regression of glioblastoma after chimeric antigen receptor T-cell therapy
.
N Engl J Med
2016
;
375
:
2561
9
.
9.
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 Oncol
2017
;
3
:
1094
101
.
10.
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
.
Sci Transl Med
2017
;
9
:
eaaa0984
.
11.
Akhavan
D
,
Alizadeh
D
,
Wang
D
,
Weist
MR
,
Shepphird
JK
,
Brown
CE
. 
CAR T cells for brain tumors: lessons learned and road ahead
.
Immunol Rev
2019
;
290
:
60
84
.
12.
Chuntova
P
,
Downey
KM
,
Hegde
B
,
Almeida
ND
,
Okada
H
. 
Genetically engineered T-cells for malignant glioma: overcoming the barriers to effective immunotherapy
.
Front Immunol
2018
;
9
:
3062
.
13.
Lim
WA
,
June
CH
. 
The principles of engineering immune cells to treat cancer
.
Cell
2017
;
168
:
724
40
.
14.
Simeonov
DR
,
Marson
A
. 
CRISPR-based tools in immunity
.
Annu Rev Immunol
2019
;
37
:
571
97
.
15.
Stadtmauer
EA
,
Fraietta
JA
,
Davis
MM
,
Cohen
AD
,
Weber
KL
,
Lancaster
E
, et al
CRISPR-engineered T cells in patients with refractory cancer
.
Science
2020
;
367
:
eaba7365
.
16.
Tang
N
,
Cheng
C
,
Zhang
X
,
Qiao
M
,
Li
N
,
Mu
W
, et al
TGF-beta inhibition via CRISPR promotes the long-term efficacy of CAR T cells against solid tumors
.
JCI Insight
2020
;
5
:
e133977
.
17.
Crowther
MD
,
Dolton
G
,
Legut
M
,
Caillaud
ME
,
Lloyd
A
,
Attaf
M
, et al
Genome-wide CRISPR-Cas9 screening reveals ubiquitous T cell cancer targeting via the monomorphic MHC class I-related protein MR1
.
Nat Immunol
2020
;
21
:
178
85
.
18.
Manguso
RT
,
Pope
HW
,
Zimmer
MD
,
Brown
FD
,
Yates
KB
,
Miller
BC
, et al
In vivo CRISPR screening identifies Ptpn2 as a cancer immunotherapy target
.
Nature
2017
;
547
:
413
8
.
19.
Patel
SJ
,
Sanjana
NE
,
Kishton
RJ
,
Eidizadeh
A
,
Vodnala
SK
,
Cam
M
, et al
Identification of essential genes for cancer immunotherapy
.
Nature
2017
;
548
:
537
42
.
20.
Dong
MB
,
Wang
G
,
Chow
RD
,
Ye
L
,
Zhu
L
,
Dai
X
, et al
Systematic immunotherapy target discovery using genome-scale in vivo CRISPR screens in CD8 T cells
.
Cell
2019
;
178
:
1189
204
.
21.
Shifrut
E
,
Carnevale
J
,
Tobin
V
,
Roth
TL
,
Woo
JM
,
Bui
CT
, et al
Genome-wide CRISPR screens in primary human T cells reveal key regulators of immune function
.
Cell
2018
;
175
:
1958
71
.
22.
Wei
J
,
Long
L
,
Zheng
W
,
Dhungana
Y
,
Lim
SA
,
Guy
C
, et al
Targeting REGNASE-1 programs long-lived effector T cells for cancer therapy
.
Nature
2019
;
576
:
471
6
.
23.
Harris
DT
,
Hager
MV
,
Smith
SN
,
Cai
Q
,
Stone
JD
,
Kruger
P
, et al
Comparison of T cell activities mediated by human TCRs and CARs that use the same recognition domains
.
J Immunol
2018
;
200
:
1088
100
.
24.
Lathia
JD
,
Mack
SC
,
Mulkearns-Hubert
EE
,
Valentim
CL
,
Rich
JN
. 
Cancer stem cells in glioblastoma
.
Genes Dev
2015
;
29
:
1203
17
.
25.
Prager
BC
,
Xie
Q
,
Bao
S
,
Rich
JN
. 
Cancer stem cells: the architects of the tumor ecosystem
.
Cell Stem Cell
2019
;
24
:
41
53
.
26.
Brown
CE
,
Aguilar
B
,
Starr
R
,
Yang
X
,
Chang
WC
,
Weng
L
, et al
Optimization of IL13Ralpha2-targeted chimeric antigen receptor T cells for improved anti-tumor efficacy against glioblastoma
.
Mol Ther
2018
;
26
:
31
44
.
27.
Fraietta
JA
,
Lacey
SF
,
Orlando
EJ
,
Pruteanu-Malinici
I
,
Gohil
M
,
Lundh
S
, et al
Determinants of response and resistance to CD19 chimeric antigen receptor (CAR) T cell therapy of chronic lymphocytic leukemia
.
Nat Med
2018
;
24
:
563
71
.
28.
Rossi
J
,
Paczkowski
P
,
Shen
YW
,
Morse
K
,
Flynn
B
,
Kaiser
A
, et al
Preinfusion polyfunctional anti-CD19 chimeric antigen receptor T cells are associated with clinical outcomes in NHL
.
Blood
2018
;
132
:
804
14
.
29.
Singh
N
,
Lee
YG
,
Shestova
O
,
Ravikumar
P
,
Hayer
KE
,
Hong
SJ
, et al
Impaired death receptor signaling in leukemia causes antigen-independent resistance by inducing CAR T-cell dysfunction
.
Cancer Discov
2020
;
10
:
552
67
.
30.
Wang
D
,
Aguilar
B
,
Starr
R
,
Alizadeh
D
,
Brito
A
,
Sarkissian
A
, et al
Glioblastoma-targeted CD4+ CAR T cells mediate superior antitumor activity
.
JCI Insight
2018
;
3
:
e99048
.
31.
Wang
D
,
Starr
R
,
Alizadeh
D
,
Yang
X
,
Forman
SJ
,
Brown
CE
. 
In vitro tumor cell rechallenge for predictive evaluation of chimeric antigen receptor T cell antitumor function
.
J Vis Exp
2019
;
144
:
10.3791/59275
.
32.
Doench
JG
,
Fusi
N
,
Sullender
M
,
Hegde
M
,
Vaimberg
EW
,
Donovan
KF
, et al
Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9
.
Nat Biotechnol
2016
;
34
:
184
91
.
33.
Brown
CE
,
Warden
CD
,
Starr
R
,
Deng
X
,
Badie
B
,
Yuan
YC
, et al
Glioma IL13Ralpha2 is associated with mesenchymal signature gene expression and poor patient prognosis
.
PLoS One
2013
;
8
:
e77769
.
34.
Silva-Filho
JL
,
Caruso-Neves
C
,
Pinheiro
AAS
. 
IL-4: an important cytokine in determining the fate of T cells
.
Biophys Rev
2014
;
6
:
111
8
.
35.
Crawford
A
,
Angelosanto
JM
,
Kao
C
,
Doering
TA
,
Odorizzi
PM
,
Barnett
BE
, et al
Molecular and transcriptional basis of CD4(+) T cell dysfunction during chronic infection
.
Immunity
2014
;
40
:
289
302
.
36.
Martinez
GJ
,
Pereira
RM
,
Aijo
T
,
Kim
EY
,
Marangoni
F
,
Pipkin
ME
, et al
The transcription factor NFAT promotes exhaustion of activated CD8(+) T cells
.
Immunity
2015
;
42
:
265
78
.
37.
Khan
O
,
Giles
JR
,
McDonald
S
,
Manne
S
,
Ngiow
SF
,
Patel
KP
, et al
TOX transcriptionally and epigenetically programs CD8(+) T cell exhaustion
.
Nature
2019
;
571
:
211
8
.
38.
Dong
Y
,
Li
X
,
Yu
Y
,
Lv
F
,
Chen
Y
. 
JAK/STAT signaling is involved in IL-35-induced inhibition of hepatitis B virus antigen-specific cytotoxic T cell exhaustion in chronic hepatitis B
.
Life Sci
2020
;
252
:
117663
.
39.
Ma
X
,
Bi
E
,
Lu
Y
,
Su
P
,
Huang
C
,
Liu
L
, et al
Cholesterol induces CD8(+) T cell exhaustion in the tumor microenvironment
.
Cell Metab
2019
;
30
:
143
56
.
40.
Chen
J
,
Lopez-Moyado
IF
,
Seo
H
,
Lio
CJ
,
Hempleman
LJ
,
Sekiya
T
, et al
NR4A transcription factors limit CAR T cell function in solid tumours
.
Nature
2019
;
567
:
530
4
.
41.
Singh
U
,
Shamran
H
,
Singh
N
,
Guan
HB
,
Mishra
M
,
Price
RL
, et al
Blocking fatty acid amide hydrolase reduces T cell activation and attenuates experimental colitis
.
J Immunol
2015
;
194
:
196.3
.
42.
Schmitz
ML
,
Krappmann
D
. 
Controlling NF-kappa B activation in T cells by costimulatory receptors
.
Cell Death Differ
2006
;
13
:
834
42
.
43.
Eyquem
J
,
Mansilla-Soto
J
,
Giavridis
T
,
van der Stegen
SJ
,
Hamieh
M
,
Cunanan
KM
, et al
Targeting a CAR to the TRAC locus with CRISPR/Cas9 enhances tumour rejection
.
Nature
2017
;
543
:
113
7
.
44.
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
.
Sci Transl Med
2020
;
12
:
eaaw2672
.
45.
Bandyopadhyay
S
,
Valdor
R
,
Macian
F
. 
Tle4 regulates epigenetic silencing of gamma interferon expression during effector T helper cell tolerance
.
Mol Cell Biol
2014
;
34
:
233
45
.
46.
Naluyima
P
,
Lal
KG
,
Costanzo
MC
,
Kijak
GH
,
Gonzalez
VD
,
Blom
K
, et al
Terminal effector CD8 T cells defined by an IKZF2(+)IL-7R(−) transcriptional signature express FcgammaRIIIA, expand in HIV infection, and mediate potent HIV-specific antibody-dependent cellular cytotoxicity
.
J Immunol
2019
;
203
:
2210
21
.
47.
Sowell
RT
,
Kaech
SM
. 
Probing the diversity of T cell dysfunction in cancer
.
Cell
2016
;
166
:
1362
4
.
48.
Shin
HJ
,
Lee
JB
,
Park
SH
,
Chang
J
,
Lee
CW
. 
T-bet expression is regulated by EGR1-mediated signaling in activated T cells
.
Clin Immunol
2009
;
131
:
385
94
.
49.
Ananieva
EA
,
Patel
CH
,
Drake
CH
,
Powell
JD
,
Hutson
SM
. 
Cytosolic branched chain aminotransferase (BCATc) regulates mTORC1 signaling and glycolytic metabolism in CD4+ T cells
.
J Biol Chem
2014
;
289
:
18793
804
.
50.
Hess
C
,
Means
TK
,
Autissier
P
,
Woodberry
T
,
Altfeld
M
,
Addo
MM
, et al
IL-8 responsiveness defines a subset of CD8 T cells poised to kill
.
Blood
2004
;
104
:
3463
71
.
51.
Trifilo
MJ
,
Bergmann
CC
,
Kuziel
WA
,
Lane
TE
. 
CC chemokine ligand 3 (CCL3) regulates CD8(+)-T-cell effector function and migration following viral infection
.
J Virol
2003
;
77
:
4004
14
.
52.
Kiniry
BE
,
Hunt
PW
,
Hecht
FM
,
Somsouk
M
,
Deeks
SG
,
Shacklett
BL
. 
Differential expression of CD8(+) T cell cytotoxic effector molecules in blood and gastrointestinal mucosa in HIV-1 infection
.
J Immunol
2018
;
200
:
1876
88
.
53.
Duraiswamy
J
,
Ibegbu
CC
,
Masopust
D
,
Miller
JD
,
Araki
K
,
Doho
GH
, et al
Phenotype, function, and gene expression profiles of programmed death-1(hi) CD8 T cells in healthy human adults
.
J Immunol
2011
;
186
:
4200
12
.
54.
Wherry
EJ
,
Ha
SJ
,
Kaech
SM
,
Haining
WN
,
Sarkar
S
,
Kalia
V
, et al
Molecular signature of CD8+ T cell exhaustion during chronic viral infection
.
Immunity
2007
;
27
:
670
84
.
55.
Lynn
RC
,
Weber
EW
,
Sotillo
E
,
Gennert
D
,
Xu
P
,
Good
Z
, et al
c-Jun overexpression in CAR T cells induces exhaustion resistance
.
Nature
2019
;
576
:
293
300
.
56.
Muller
MR
,
Rao
A
. 
NFAT, immunity and cancer: a transcription factor comes of age
.
Nat Rev Immunol
2010
;
10
:
645
56
.
57.
Myers
LM
,
Tal
MC
,
Torrez Dulgeroff
LB
,
Carmody
AB
,
Messer
RJ
,
Gulati
G
, et al
A functional subset of CD8(+) T cells during chronic exhaustion is defined by SIRPalpha expression
.
Nat Commun
2019
;
10
:
794
.
58.
Maimela
NR
,
Liu
S
,
Zhang
Y
. 
Fates of CD8+ T cells in tumor microenvironment
.
Comput Struct Biotechnol J
2019
;
17
:
1
13
.
59.
Bhairavabhotla
R
,
Kim
YC
,
Glass
DD
,
Escobar
TM
,
Patel
MC
,
Zahr
R
, et al
Transcriptome profiling of human FoxP3+ regulatory T cells
.
Hum Immunol
2016
;
77
:
201
13
.
60.
Nishio
H
,
Yaguchi
T
,
Sugiyama
J
,
Sumimoto
H
,
Umezawa
K
,
Iwata
T
, et al
Immunosuppression through constitutively activated NF-kappa B signalling in human ovarian cancer and its reversal by an NF-kappa B inhibitor
.
Br J Cancer
2014
;
110
:
2965
74
.
61.
Yamamoto
Y
,
Gaynor
RB
. 
Therapeutic potential of inhibition of the NF-kappaB pathway in the treatment of inflammation and cancer
.
J Clin Invest
2001
;
107
:
135
42
.
62.
Liao
W
,
Overman
MJ
,
Boutin
AT
,
Shang
X
,
Zhao
D
,
Dey
P
, et al
KRAS-IRF2 axis drives immune suppression and immune therapy resistance in colorectal cancer
.
Cancer Cell
2019
;
35
:
559
72
.
63.
Li
WM
,
Liu
HR
. 
CCL20-CCR6 cytokine network facilitate Treg activity in advanced grades and metastatic variants of hepatocellular carcinoma
.
Scand J Immunol
2016
;
83
:
33
7
.
64.
Do
Y
,
Nagarkatti
PS
,
Nagarkatti
M
. 
Role of CD44 and hyaluronic acid (HA) in activation of alloreactive and antigen-specific T cells by bone marrow-derived dendritic cells
.
J Immunother
2004
;
27
:
1
12
.
65.
Jankowska
KI
,
Williamson
EK
,
Roy
NH
,
Blumenthal
D
,
Chandra
V
,
Baumgart
T
, et al
Integrins modulate T cell receptor signaling by constraining actin flow at the immunological synapse
.
Front Immunol
2018
;
9
:
25
.
66.
Batlle
E
,
Massague
J
. 
Transforming growth factor-beta signaling in immunity and cancer
.
Immunity
2019
;
50
:
924
40
.
67.
Deng
Q
,
Han
G
,
Puebla-Osorio
N
,
Ma
MCJ
,
Strati
P
,
Chasen
B
, et al
Characteristics of anti-CD19 CAR T cell infusion products associated with efficacy and toxicity in patients with large B cell lymphomas
.
Nat Med
2020
;
26
:
1878
87
.
68.
Ghassemi
S
,
Nunez-Cruz
S
,
O'Connor
RS
,
Fraietta
JA
,
Patel
PR
,
Scholler
J
, et al
Reducing ex vivo culture improves the antileukemic activity of chimeric antigen receptor (CAR) T cells
.
Cancer Immunol Res
2018
;
6
:
1100
9
.
69.
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
.
Nat Med
2018
;
24
:
572
9
.
70.
Theruvath
J
,
Sotillo
E
,
Mount
CW
,
Graef
CM
,
Delaidelli
A
,
Heitzeneder
S
, et al
Locoregionally administered B7-H3-targeted CAR T cells for treatment of atypical teratoid/rhabdoid tumors
.
Nat Med
2020
;
26
:
712
9
.
71.
Donovan
LK
,
Delaidelli
A
,
Joseph
SK
,
Bielamowicz
K
,
Fousek
K
,
Holgado
BL
, et al
Locoregional delivery of CAR T cells to the cerebrospinal fluid for treatment of metastatic medulloblastoma and ependymoma
.
Nat Med
2020
;
26
:
720
31
.
72.
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
.
Nat Biotechnol
2019
;
37
:
1302
13
.
73.
Arvanitis
CD
,
Ferraro
GB
,
Jain
RK
. 
The blood-brain barrier and blood-tumour barrier in brain tumours and metastases
.
Nat Rev Cancer
2020
;
20
:
26
41
.
74.
Platt
RJ
,
Chen
S
,
Zhou
Y
,
Yim
MJ
,
Swiech
L
,
Kempton
HR
, et al
CRISPR-Cas9 knockin mice for genome editing and cancer modeling
.
Cell
2014
;
159
:
440
55
.
75.
Henriksson
J
,
Chen
X
,
Gomes
T
,
Ullah
U
,
Meyer
KB
,
Miragaia
R
, et al
Genome-wide CRISPR screens in T helper cells reveal pervasive crosstalk between activation and differentiation
.
Cell
2019
;
176
:
882
96
.
76.
Brown
CE
,
Mackall
CL
. 
CAR T cell therapy: inroads to response and resistance
.
Nat Rev Immunol
2019
;
19
:
73
4
.
77.
Wherry
EJ
,
Kurachi
M
. 
Molecular and cellular insights into T cell exhaustion
.
Nat Rev Immunol
2015
;
15
:
486
99
.
78.
Long
AH
,
Haso
WM
,
Shern
JF
,
Wanhainen
KM
,
Murgai
M
,
Ingaramo
M
, et al
4-1BB costimulation ameliorates T cell exhaustion induced by tonic signaling of chimeric antigen receptors
.
Nat Med
2015
;
21
:
581
90
.
79.
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
.
J Clin Invest
2016
;
126
:
3130
44
.
80.
Rafiq
S
,
Yeku
OO
,
Jackson
HJ
,
Purdon
TJ
,
van Leeuwen
DG
,
Drakes
DJ
, et al
Targeted delivery of a PD-1-blocking scFv by CAR-T cells enhances anti-tumor efficacy in vivo
.
Nat Biotechnol
2018
;
36
:
847
56
.
81.
Schietinger
A
,
Philip
M
,
Krisnawan
VE
,
Chiu
EY
,
Delrow
JJ
,
Basom
RS
, et al
Tumor-specific T cell dysfunction is a dynamic antigen-driven differentiation program initiated early during tumorigenesis
.
Immunity
2016
;
45
:
389
401
.
82.
Man
K
,
Gabriel
SS
,
Liao
Y
,
Gloury
R
,
Preston
S
,
Henstridge
DC
, et al
Transcription factor IRF4 promotes CD8(+) T cell exhaustion and limits the development of memory-like T cells during chronic infection
.
Immunity
2017
;
47
:
1129
41
.
83.
Li
P
,
Spolski
R
,
Liao
W
,
Wang
L
,
Murphy
TL
,
Murphy
KM
, et al
BATF-JUN is critical for IRF4-mediated transcription in T cells
.
Nature
2012
;
490
:
543
6
.
84.
Meixner
A
,
Karreth
F
,
Kenner
L
,
Wagner
EF
. 
JunD regulates lymphocyte proliferation and T helper cell cytokine expression
.
EMBO J
2004
;
23
:
1325
35
.
85.
Chiu
R
,
Angel
P
,
Karin
M
. 
Jun-B differs in its biological properties from, and is a negative regulator of, c-Jun
.
Cell
1989
;
59
:
979
86
.
86.
Wei
F
,
Zhong
S
,
Ma
Z
,
Kong
H
,
Medvec
A
,
Ahmed
R
, et al
Strength of PD-1 signaling differentially affects T-cell effector functions
.
Proc Natl Acad Sci U S A
2013
;
110
:
E2480
9
.
87.
Jiao
S
,
Subudhi
SK
,
Aparicio
A
,
Ge
Z
,
Guan
B
,
Miura
Y
, et al
Differences in tumor microenvironment dictate T helper lineage polarization and response to immune checkpoint therapy
.
Cell
2019
;
179
:
1177
90
.
88.
Tran
E
,
Turcotte
S
,
Gros
A
,
Robbins
PF
,
Lu
YC
,
Dudley
ME
, et al
Cancer immunotherapy based on mutation-specific CD4+ T cells in a patient with epithelial cancer
.
Science
2014
;
344
:
641
5
.
89.
Lee
DW
,
Kochenderfer
JN
,
Stetler-Stevenson
M
,
Cui
YK
,
Delbrook
C
,
Feldman
SA
, et al
T cells expressing CD19 chimeric antigen receptors for acute lymphoblastic leukaemia in children and young adults: a phase 1 dose-escalation trial
.
Lancet
2015
;
385
:
517
28
.
90.
Bao
S
,
Wu
Q
,
McLendon
RE
,
Hao
Y
,
Shi
Q
,
Hjelmeland
AB
, et al
Glioma stem cells promote radioresistance by preferential activation of the DNA damage response
.
Nature
2006
;
444
:
756
60
.
91.
Xie
Q
,
Wu
TP
,
Gimple
RC
,
Li
Z
,
Prager
BC
,
Wu
Q
, et al
N(6)-methyladenine DNA modification in glioblastoma
.
Cell
2018
;
175
:
1228
43
.
92.
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
.
Clin Cancer Res
2018
;
24
:
95
105
.
93.
Hanzelmann
S
,
Castelo
R
,
Guinney
J
. 
GSVA: gene set variation analysis for microarray and RNA-seq data
.
BMC Bioinformatics
2013
;
14
:
7
.
94.
Stuart
T
,
Butler
A
,
Hoffman
P
,
Hafemeister
C
,
Papalexi
E
,
Mauck
WM
 III
, et al
Comprehensive integration of single-cell data
.
Cell
2019
;
177
:
1888
902
.
95.
Brown
CE
,
Starr
R
,
Martinez
C
,
Aguilar
B
,
D'Apuzzo
M
,
Todorov
I
, et al
Recognition and killing of brain tumor stem-like initiating cells by CD8+ cytolytic T cells
.
Cancer Res
2009
;
69
:
8886
93
.
96.
Thorsson
V
,
Gibbs
DL
,
Brown
SD
,
Wolf
D
,
Bortone
DS
,
Ou Yang
TH
, et al
The immune landscape of cancer
.
Immunity
2018
;
48
:
812
30
.