Skip to main content

Reprogramming T cell differentiation and exhaustion in CAR-T cell therapy

Abstract

T cell differentiation is a highly regulated, multi-step process necessary for the progressive establishment of effector functions, immunological memory, and long-term control of pathogens. In response to strong stimulation, as seen in severe or chronic infections or cancer, T cells acquire a state of hypo-responsiveness known as exhaustion, limiting their effector function. Recent advances in autologous chimeric antigen receptor (CAR)-T cell therapies have revolutionized the treatment of hematologic malignancies by taking advantage of the basic principles of T cell biology to engineer products that promote long-lasting T cell response. However, many patients’ malignancies remain unresponsive to treatment or are prone to recur. Discoveries in T cell biology, including the identification of key regulators of differentiation and exhaustion, offer novel opportunities to have a durable impact on the fate of CAR-T cells after infusion. Such next-generation CAR-T cell therapies and their clinical implementation may result in the next leap forward in cancer treatment for selected patients. In this context, this review summarizes the foundational principles of T cell differentiation and exhaustion and describes how they can be utilized and targeted to further improve the design and efficacy of CAR-T cell therapies.

Introduction

T cells are an essential component of the adaptive immune system, playing a critical role in recognizing and eliminating infected and malignant cells [1]. They differentiate into distinct subsets with specialized functions to enable effective immune surveillance and response. However, prolonged exposure to antigens or chronic infections can lead to T cell exhaustion, which is characterized by diminished functionality and decreased ability to clear target cells [1].

Chimeric antigen receptor (CAR)-T cell therapy is a promising immunotherapeutic approach that involves genetically modifying T cells to express a CAR that recognizes a specific antigen on the surface of cancer cells [2]. CAR-T cell therapy has demonstrated remarkable efficacy in treating hematologic malignancies, with several CAR-T cell therapies currently available for clinical use [3]. In recent years, the US Food and Drug Administration (FDA) has approved six such therapies: Kymriah (tisagenlecleucel), Yescarta (axicabtagene ciloleucel), Tecartus (brexucabtagene autoleucel), Breyanzi (lisocabtagene maraleucel), Carvykti (ciltacabtagene autoleucel), and Abecma (idecabtagene vicleucel). Despite unprecedented benefits to countless patients, autologous CAR-T therapy has not worked equally for all patients with hematologic malignancies. In B cell acute lymphoblastic leukemia (B-ALL) and non-Hodgkin lymphoma (NHL), 50% and 56% of patients, respectively, experienced relapse from CAR-T cell treatment at 12-month follow-up [4, 5], and a cell therapy remains to be approved for chronic lymphocytic leukemia (CLL), where only 26% of patients present with a durable response [6]. CAR-T cell therapy has also not yet shown clinical benefit in a broader range of indications, particularly in solid tumors [7].

The path to broader clinical success of CAR-T cell therapy is likely to involve a series of technological advancements, perhaps even tailored to individual indications, and will certainly incorporate process improvements and engineering techniques meant to control the composition of the product, possibly along the memory-effector and exhaustion axes. Proper understanding of the mechanisms of T cell differentiation and exhaustion and the methods used to manipulate them will be crucial for developing the next generation of CAR-T cell therapies with antitumor efficacy in solid tumors [8]. In this review, we will summarize T cell differentiation and exhaustion processes and the role of T cell biology in CAR-T therapy. We will also provide insights into new therapeutic opportunities.

Role of T cell differentiation during CAR-T product manufacturing and its impact on clinical outcomes

CAR-T products generated from cancer patients vary widely in terms of differentiation status; such variation is likely a manifestation of differences in the quality of the patients’ starting T cell material [9,10,11,12,13,14]. In the last couple of years, transcriptomic signatures based on the differential expression of hundreds of memory-, effector-, and dysfunction-associated genes have been used to characterize cell therapies and correlate product characteristics with clinical response. Some of the largest clinical studies published to date have revealed that the clinical efficacy of CAR-T therapy strongly positively correlates with signatures of memory and negatively correlates with signatures of effector function or exhaustion. A study of 41 patients with advanced CLL treated with autologous CAR-T therapy (Kymriah) identified high memory, low effector, and low exhaustion gene scores as determinants of response [12]. Another study of 71 patients with B-ALL and Hodgkin lymphoma (HL) treated with Kymriah identified a network, or regulon, of target genes upregulated by the master regulator of memory function T cell factor-1 (TCF-1) as a strong predictor of response [12, 13]. In a more recent study of 12 acute lymphoblastic leukemia (ALL) patients, the authors used CITE-seq to determine that the frequency of early memory T cells in the CAR-T products of patients treated with Kymriah was predictive of response [15]. In a study of 24 patients with large B cell lymphoma (LBCL) treated with Yescarta, memory CD8 + T cells were significantly more frequent in CAR-T cell products from patients with continued complete response (CR) than those from patients with partial response/progressive disease (PR/PD); the opposite was true for exhausted CD8 + T cells [11]. Intriguingly, an independent study of 32 patients with LBCL found a statistical correlation between a population of memory-like CD8 + T cells and response for patients treated with Kymriah but not for those treated with Yescarta [16]. Finally, in a study of 54 melanoma patients treated with tumor-infiltrating lymphocyte (TIL) therapy, TIL products showing high expression of memory-associated genes and low expression of granzyme A (GZMA) and interferon gamma (IFN-γ) correlated with response [17]. Patients with a complete response received a considerably higher dose of neoantigen-specific TILs with memory characteristics than patients who did not have a response.

While it is now generally well established that CAR-T cell composition has a direct impact on the activity of CAR-T cell therapy in the clinic, a clear definition of the product attributes associated with response remains lacking. While initial attempts to define memory among CAR-T cells relied on historical cell surface markers associated with memory in the peripheral blood, such as L-selectin (CD62L) and chemokine receptor type 7 (CCR7), the direct association between these markers and response has remained elusive and study-dependent. Many studies have not been able to identify a correlation between conventional homing markers and clinical outcome [12, 18,19,20]. More specifically, CCR7 and CD62L in isolation were not associated with response [12, 20]. When a correlation has been reported, its significance has been relatively weak (p = 0.0464 [21], p = 0.0317 for central memory T cell (Tcm) [15], and p = 0.0327 [22]). At least four clinical trials with a common aim to generate a less-differentiated product from CD62L or CCR7 enrichment in the starting leukapheresis material have yet to report improved persistence and/or response in the clinic (NCT01318317; NCT01815749; NCT02062359). In one study, CAR-T cells generated from CD62L-purified T cells did not show improved efficacy compared to CAR-T cells made from non-enriched T cells (NCT01865617) [23]. Besides enrichment, alterations to the manufacturing process designed to bolster memory in the end product have not yet resulted in clinical activity superior to that of conventional methods (NCT03318900; NCT03274219; NCT01087294).

Shortening the manufacturing process with the intent to improve stemness and potency of the CAR-T product is a tantalizing concept that has seen recent clinical success [24]. A CD19-directed CAR-T therapeutic modality (YTB323), manufactured in less than 2 days, has shown favorable efficacy in 20 patients with B cell lymphoma when treated at a 25-fold lower dose than a more conventional CAR-T therapy [NCT03960840]. Here again, however, correlation between CCR7 expression in the product and response did not appear to be statistically significant. The cause of the discrepancy between the importance of memory in CAR-T product composition and the lack of improved efficacy in products enriched for CD62L or CCR7 may be multifactorial. Of note, the inflammatory cytokine interleukin 12 (IL-12) can induce high levels of CD62L expression despite skewing T cells toward terminal effector differentiation [25]. Furthermore, T cells can be induced in vitro to re-express both CD62L and CCR7 upon T cell receptor (TCR), interleukin-2 (IL-2), or interleukin-21 (IL-21) stimulation [26]. CD27 is a member of the tumor necrosis factor receptor family, which once activated supports memory formation by promoting IL-2-independent survival by maintaining the expression of interleukin-7 receptor-α (IL-7RA) [27, 28]. CD27 was shown in two large studies to correlate with efficacy when co-expressed with CD45RO [12] or CCR7 [11]; thus, CD27 might represent a better predictive tool than the homing markers CD62L and CCR7, when taken in isolation.

Pharmacokinetic parameters, memory composition, and patient outcomes

Pharmacokinetic parameters such as peak CAR-T cell level in patients, or maximum concentration (Cmax), along with area under the curve (AUC), have been demonstrated to strongly associate with response, irrespective of treatment or indications [4, 12, 22, 29]. However, to date, the connection between product composition and these predictive pharmacokinetic parameters has not been well established. A recent study may shed new light on this fundamentally important question [30]. This study, which used mathematical algorithms trained on existing clinical data to model pharmacokinetic behaviors and predict response based on product characteristics, first found that transcriptomic analysis of the pre-infusion products of two CAR-T products (Yescarta and Kymriah) in three separate indications resulted in more accurate predictions than standard flow-cytometry-based immunophenotyping [30]. Importantly, the authors found that CΑR-T products that were associated with short responses were characterized by deficient proliferative and functional capacity. These qualities are typical of Τ cell exhaustion and terminal differentiation, even within similar memory and effector cell populations. Additionally, CAR-T expansion after product infusion, which drives Cmax, is representative of memory T cell proliferative capacity [22].

In other words, CAR-T product effectiveness is likely dictated not only by population frequencies but also by the cells’ proliferative capacity, a feature intrinsically associated with the early differentiation of T cells. Given the importance of the composition of CAR-T products for clinical efficacy, it becomes essential to understand T cell differentiation in a broader biological context.

T cell differentiation during acute response to antigens

T cell differentiation is a highly regulated process initiated the moment a T cell encounters its cognate antigen. During the ensuing expansion phase, the immune response is dominated by effector T cells (Teff) that mediate antigen clearance, followed by a contraction phase where Teff give way to long-lived memory T cells (Tmem) [31]. Tmem cells uniformly retain high developmental and proliferative potential compared to Teff cells, which is facilitated by their ability to metabolically recruit additional capacities (e.g., high spare respiratory capacity) [31]. Memory T cells that home to secondary lymphoid organs or circulate in the periphery can be identified by enhanced expression of IL-7RA, which allows them to undergo homeostatic proliferation in the absence of TCR stimulation. Depending on their localization, memory T cells also express high levels of homing molecules, such as CD62L, CCR7 in lymphoid organs, CX3CR1 in the periphery, and CD103 and Hobit in tissues [32]. As T cells undergo the process of differentiation, they progressively acquire hallmarks of effector function, such as increased cytotoxic activity and cytokine production, while progressively losing their ability to proliferate (Fig. 1).

Fig. 1
figure 1

CD8 + T cell differentiation model in acute and chronic antigenic stimulation settings. Peerless NFAT appears to be one of the key upstream determinants of the differentiation path taken by T cells. “Created with BioRender.com”. SLEC: short-lived effector cells

T cell differentiation and exhaustion during chronic antigen response

The above-described T cell differentiation, including the formation of memory T cell subsets, is driven by the eradication of the antigen and thus the resolution of the infection. When a disease and its associated antigens persist at high levels, as seen in chronic infections or cancer, T cells undergo an alternative T cell differentiation path commonly known as T cell exhaustion. Exhausted T cells (Tex) are defined by sustained expression of inhibitory receptors, also known as checkpoint inhibitors, such as programmed cell death protein 1 (PD-1), lymphocyte activation gene 3 (LAG-3), 2B4, and cytotoxic T-lymphocyte-associated antigen 4 (CTLA-4); a diminished ability to produce effector cytokines such as IFN-γ and tumor necrosis factor (TNF); and a reduced ability to proliferate [33]. They are also characterized by high activity of TCR-responsive transcription factors, including thymocyte selection-associated high-mobility group box (TOX), B cell-activating transcription factor (BATF), interferon regulatory factor 4 (IRF4), and nuclear factor of activated T cell (NFAT) [34]. This phenotypic and transcriptional profile has recently been extended to demonstrate that Tex cells fundamentally differ from regular T cells based on a distinct epigenetic landscape [35,36,37]. T cells in chronic infection and in tumors acquire exhaustion-specific patterns of chromatin accessibility [38,39,40,41]. Tmem cells have an open chromatin structure specifically enriched for binding sites to basic leucine zipper, interferon regulatory factors, and T-box transcription factors (TFs). In contrast, Tex cells (like Teff cells) harbor an open chromatin structure around effector genes; however, Tex cells are uniquely enriched for binding sites to NFAT and nuclear receptor subfamily 4 group A (NR4A) [42], and effector genes like IFN-γ show high levels of negative DNA methylation in line with their hypofunctional state [42, 43]. Importantly, this pattern is stably imprinted and cannot be reversed by the blockade of immune checkpoint inhibitors such as with the anti-PD-1 antibodies used in the clinic [43, 44].

Similar to effector T cell differentiation during acute stimulation, under chronic stimulation T cells undergo sequential differentiation toward terminal exhaustion (Fig. 1). An early form of dysfunctional T cell, herein referred to as precursor exhausted T cells (Tpex), was first identified in a model of chronic viral infection [45]. In mice, Tpex cells are detected within the first week after the initial challenge, indicating that commitment to a hypo-responsive state takes up to 7 days of continuous TCR stimulation [46]. Similar to peripheral Tmem cells, Tpex cells express TCF-1, retain a self-renewing capacity, and can differentiate into effector-like Tex cells. However, Tpex cells also express TFs associated with differentiation, such as TOX, nuclear factor of activated T cells 2 (NFATC2), and BATF [47], and several checkpoint inhibitors, such as PD-1 and T cell immunoreceptor with Ig and ITIM domain (TIGIT) [48]. Tpex are also limited in their proliferation and cytokine production capacity [48]. In line with other types of Tex cells, Tpex cells are epigenetically locked in an exhausted state, without the ability to return to a more physiological state [44]. In human cancer, higher levels of TCF-1 + Tpex cells in tumors have been linked to longer patient survival and improved therapeutic outcomes and are essential for the long-term maintenance of T cell response [49,50,51,52,53]. Similarly, Tpex cells have been shown to play an important role in the therapeutic activity of antibodies blocking checkpoint inhibitors [49,50,51, 53, 54], highlighting the importance of this T cell subset for immunity in both humans and mice.

The continuous presence of antigens leads to the emergence of terminally differentiated and exhausted T cells (Texterm). Trajectory analyses support a model whereby Texterm cells are derived from Tpex cells [55]. Texterm cells are identified by the high expression of eomesodermin (EOMES), TOX, CD69, and checkpoint inhibitors such as PD-1, LAG-3, or T cell immunoglobulin domain and mucin domain 3 (TIM-3) and low levels of T-BET [56]. During progression toward late dysfunction, traditional effector properties such as the capacity to produce interleukin-2 (IL-2), TNF, and IFN-γ fade away and eventually disappear entirely [57]. Texterm cells correlate with more severe disease in HIV and are enriched in lung tumor tissue compared with blood [58]. Interestingly, in an autochthonous model of lung cancer, most intratumoral T cells eventually differentiate into Texterm cells, and this shift coincides with progression of the tumor from a “hot,” or T cell inflamed, microenvironment to a “cold,” non-T cell inflamed, state [59], suggesting that, given enough time, T cells continually exposed to TCR stimulation terminally differentiate and become inoperant.

Regulation of T cell differentiation during acute and chronic conditions

T cell differentiation during acute and chronic stimulation alike is a gradual process that is tightly regulated by epigenetic and transcriptional modulators, whose abundance and timing of expression is key to proper fate commitment. Interestingly, although chronic and acute stimulation results in vastly different outcomes, the types of modulators involved in each process overlap considerably.

Transcriptional regulation of differentiation in the context of acute antigenic challenge

In the early stages of an infection, most T cells differentiate toward an effector function (short-lived effector cells, or SLECs), whereas a subset of activated T cells (memory precursor effector cells, or MPECs) are poised to become long-term, self-renewing memory T cells, protecting the host from recurring infections. At single-cell resolution, the transcriptional profile of CD8 + T cells at the onset of TCR engagement shows striking divergence and reveals two distinct subpopulations along a memory-effector axis after the first division [60,61,62]. The canonical Wnt-signaling pathway is important for the maturation and homeostasis of peripheral memory T cells [63], and the Wnt-dependent factor TCF-1 and its functional homolog lymphoid enhancer-binding factor-1 (LEF-1) are key regulators in the formation of memory and the inhibition of effector differentiation [64, 65] (Fig. 2A). TCF-1 utilizes both transcriptional regulation and histone deacetylation via its intrinsic HDAC domain to direct T cell fate [66, 67]. EOMES acts directly downstream of TCF-1 [65] and is critical to maintain memory, in part because it promotes sustained expression of interleukin-2 receptor subunit beta (IL-2RB), thus supporting interleukin-15 (IL-15) and IL-2 signaling and continued proliferation [68, 69]. Forkhead O transcription factor 1 (FOXO1) enforces stem cell-like properties and represses T-BET, IFN-γ, and granzyme B (GZMB) effector functions [70,71,72,73], and its continuous activity is necessary for the maintenance of memory in both acute and chronic conditions [74,75,76]. FOXO1 acts upstream of TCF-1, as it directly binds and upregulates expression of the TF [70, 75]. TCF-1 and FOXO1 act in synergy by promoting expression of the pro-memory and pro-survival genes EOMES, IL-7RA, CD62L, CCR7, and B cell leukemia 2 (BCL-2) [71, 73, 75]. Inversely, FOXO1 is inhibited by mammalian target of rapamycin complex 1 (mTORC1), which is in line with the opposing effects of FOXO1 and mammalian target of rapamycin (mTOR) signaling on CD8 + T cell differentiation [71]. BACH2, another downstream target of FOXO1, establishes a stem-like transcriptional program at the single-cell level [77] and contributes to memory formation by restricting the access of JUN family TFs to the regulatory elements of TCR-induced genes [78]. Interestingly, BACH2 also enacts an epigenetic program of memory, with reduced chromatin accessibility at regions open in terminally exhausted CD8 + T cells and genomic regions controlled by TFs such as runt-domain transcription factors (RUNX) and BATF. BACH2 promotes inhibitor of DNA-binding 3 (ID3) and B cell lymphoma 6 (BCL-6) expression but suppresses the expression of killer cell lectin-like receptor subfamily G member 1 (KLRG1) and B-lymphocyte-induced maturation protein 1 (BLIMP-1) [79].

Fig. 2
figure 2

Graded expression of key regulators of T cell differentiation and exhaustion in acute and chronic conditions. A: During acute TCR stimulation, the following factors are highest among Tmem cells and gradually decrease in expression during the process of effector differentiation: TCF-1 [80], BACH2 [77], and FOXO1 [75]. Expression of EOMES [81], RUNX3 [82], and BATF-IRF4 [83, 84] is low among Tmem cells, peaks among early Teff cells and recedes in late Teff cells. Finally, the following factors gradually increase in expression and contribute to effector T cell differentiation: T-BET [85] and BLIMP-1 [86]. B: During chronic TCR stimulation, the following factors are highest among Tpex cells and gradually decrease in expression along the course of exhaustion: TCF-1 [35, 87], TOX [87, 88], and FOXO1 [35, 73, 76]. While TCF-1 remains low, expression of TOX and FOXO1 gradually increases as T cells become terminally exhausted. On the other hand, expression of T-BET [85, 87, 89], BATF-IRF4 [35, 90], and NR4A [91] is at its lowest in Tpex cells and is induced during chronic TCR engagement. While T-BET and BATF-IRF4 expression peaks among intermediate exhausted T cells and subsequently recedes, NR4A expression remains high and peaks at the Texterm cell stage. Expression of EOMES [87, 89, 92] and BLIMP-1 [93] also peaks at the terminal exhausted stage, but their gradual increase in expression starts later, at the intermediate exhausted stage

Likewise, downstream differentiation is tightly regulated by its own set of transcriptional and epigenetic modulators. In CD8 + T cells, the interleukin-21 (IL-21)- and interleukin-12 (IL-12)-induced TF BATF regulates chromatin accessibility and promotes the upregulation of key transcriptional regulators of effector differentiation such as runt-domain transcription factor 3 (RUNX3), T-BET, and BLIMP-1 and, together with IRF4 and Jun, upregulates the IL-12 and IL-2 signaling pathways [83, 94, 95]. On the other hand, BATF represses expression of the effector molecules IFN-γ and GZMB. Hence, BATF augments the propagation of inflammatory signals while restraining expression of downstream effector molecules, thus acting as a gateway in the differentiation process. RUNX3, another BATF-induced TF [83], is essential for long-term immunity [96]. RUNX3 governs chromatin accessibility to a broad number of cis-regulatory regions and upregulates IRF4 and BLIMP-1, while it mediates downregulation of the memory TFs BACH2 and TCF-1. Similar to BATF, RUNX3 curtails terminal differentiation by limiting T-BET expression, thus ensuring timely expression of regulators of differentiation. The inflammatory cytokine IL-12 directs further differentiation by inducing expression of the effector TFs T-BET and BLIMP-1 and downregulating factors essential for memory-like TCF-1 and IL-7RA [80, 97]. T-BET favors the induction of a terminally differentiated state [97] and, together with BLIMP-1, promotes the expression of effector molecules such as granzymes while repressing memory-associated factors such as TCF-1, IL-7RA, and CD62L [98]. BLIMP-1 has a negative effect on proliferation, and its expression correlates with a greater level of apoptosis after restimulation [99, 100]. Type I interferons induce expression of the NOTCH family of receptors and, together with interleukin-2 receptor (IL-2R), mTOR, and T-BET, form a positive feedback module of differentiation that integrates signals from various sources [101]. Altogether, it is the graded expression of competing sets of transcription factors that controls CD8 + T cell differentiation and fate commitment toward memory versus effector T cell function.

Transcriptional regulation of differentiation in the context of chronic antigen stimulation

During chronic antigen stimulation, the induction of exhaustion is independent of inflammatory signals and other environmental factors; instead, the degree of exhaustion is directly linked to the quantity of antigen present as well as the duration/frequency of TCR stimulation [40, 102]. It should be noted that, contrary to exhaustion, other forms of T cell dysfunction can be induced by the tumor microenvironment, and factors such as tumor-associated hypoxia, acidity, and altered lipid metabolism can impair CAR-T cell functionality, while the presence of immunosuppressive cell populations can curtail their activity. This important topic has been extensively discussed elsewhere [103,104,105,106]. The balance between NFAT cells and activator protein 1 (AP-1) dictates the outcome of the branching point between physiological versus hypo-responsive differentiation, and chronic TCR stimulation drives this imbalance, resulting in “partnerless” NFAT [107], which directs a transcriptional and epigenetic program of exhaustion [42, 108] (Fig. 1). Partnerless NFAT induces a transcriptional program characterized by continuous expression of the repressive transcription factors early growth response 2 (EGR2), zinc finger protein 2 (IKZF2), IRF4, TOX, and NR4A and other exhaustion-associated factors [108,109,110,111,112]. Downstream of these initiating events, many of the same regulatory factors found during acute stimulation play a similar role in the chronic setting (Fig. 2B). For instance, TCF-1 and FOXO1 both play a central role in the generation and maintenance of Tpex cells [64, 65, 73, 74, 113, 114]. FOXO1 directly promotes expression of PD-1, which, in a positive feedback loop, indirectly promotes expression of FOXO1 via inhibition of mTOR signaling [73]. Likewise, BACH2 is required for long-term immunity in chronic viral infection [115], while BLIMP-1 antagonizes memory formation in both chronic (Tpex) and acute (Tcm) conditions [93, 99]. BLIMP-1 expression is highest in the highly differentiated Texterm [93]. BATF together with IRF4 cooperates to establish exhaustion [116]. The NFAT-induced regulator TOX is highly expressed among Tex cells in both humans and mice, particularly in Tpex and Texterm cells [87, 88, 111, 117]. TOX directly contributes to the open chromatin structure and continuously high levels of expression of multiple checkpoint inhibitors, including PD-1 [20, 87, 88, 111]. TOX binds to chromatin remodeling factors and thus contributes to the exhaustion-linked epigenetic landscape of Tex cells [88, 111]. The role of TOX in T cell dysfunction may be time dependent, since short-term expression of TOX is not sufficient to induce terminal exhaustion [88, 111]. In sequence with TOX, the NR4A family of TFs contributes to T cell exhaustion by orchestrating a genome-wide, exhaustion-linked alteration of the epigenetic and transcriptomic landscape in CD8 + TILs and CAR-T cells [112, 118]. Like TOX, NR4A TFs directly bind to and upregulate PD-1 gene expression [112]. NR4A1 represses effector gene expression by inhibiting AP-1 function, and its ablation enhances antitumor immunity [91]. Interestingly, NR4A1 plays a central role in general tolerance since its ablation exacerbates autoimmunity in a model of induced colitis [91].

Controlling CAR-T differentiation via cell engineering

The discrepancy between the clinical benefit of high memory composition in CAR-T cells and the absence of demonstrated improvement of CAR-T cells generated under memory-skewing conditions is puzzling. One confounding factor is that markers typically used as surrogates for memory, such as CD62L or CCR7, might not be appropriate markers for in vitro expanded CAR-T cells. However, it is also quite plausible that to increase efficacy over current CAR-T therapies and to demonstrate durable clinical benefit a regulatory network of memory needs to be stably engineered to retain memory characteristics post-infusion. Accordingly, we will focus this section on engineering approaches meant to stably reprogram CAR-T cells’ differentiation and exhaustion states. Note that other recent reviews have addressed the question of next-generation CAR-T enhancements more broadly, beyond manipulation of differentiation and exhaustion [119, 120].

Factors involved in T cell memory or dysfunction that have been evaluated in the context of adoptive cell therapy are summarized in Table 1. One precursor study revealed that constitutive expression of BCL-6, a master regulator of T follicular helper CD4 + T cell differentiation and repressor of BLIMP-1 [121], can increase homeostatic proliferation and the maintenance of memory T cells, which increases their frequency over time [122]. The potential to directly modulate transcription factors in CAR-T cells specifically was more recently demonstrated via ectopic expression of the AP-1 transcription factor cJUN [123]. Using three different CAR constructs against as many targets, Lynn and colleagues demonstrated that ectopically expressed cJUN improves the antitumor activity as well as persistence of CAR-T cells in vivo. cJUN may possibly compete for chromatin binding with NR4A3, which is known to repress the development of memory T cells [124]. In contrast, a subsequent mouse study failed to show any benefit of ectopic expression of cJUN; instead, co-expression of cJUN with BATF reduced the advantage conferred by ectopic expression of BATF alone [84]. However, in a subsequent study using exhaustion-inducing conditions (i.e., low effector-to-target ratios) BATF deletion, not overexpression, improved the antitumor immunity conferred by CAR-T cells, both in vitro and in vivo [125]. Alternatively, the improved antitumor activity associated with cJUN overexpression might be due to an improved ability of T cells to recognize low levels of antigens [126]. The master transcriptional regulator of central memory, TCF-1, is another interesting choice for CAR-T cell therapies. Murine T cells expressing TCF-1 ectopically show greater control of chronic viral infection and vastly out-compete wild-type T cells 30–40 days post-infection [67]. In a B16 melanoma model, adoptive transfer of TCF-1-positive T cells mediated greater tumor control, lower degrees of differentiation and exhaustion, and consequently greater cytokine production than their wild-type counterparts [65]. Natural killer T (NKT) cells are a type of αβ T cells endowed with innate immune properties. LEF-1, a functional homolog to TCF-1, promotes expansion when ectopically expressed in NKT cells transduced with an anti-GD2 CAR construct and improves antitumor efficacy in vivo [127]. In a chronic model of viral infection, deletion of FOXO1 leads to a decrease in memory T cells and loss of viral control over time [73], making FOXO1 a prime candidate to improve the stemness of a CAR-T product. Inducible expression of a constitutively active form of FOXO1, as well as TCF-1, after adoptive transfer of CAR-T cells leads to increased persistence compared to control CAR-Ts [128]. BACH2, another master regulator of stemness, has not been investigated in the context of CAR-T cells; however, in a model of chronic viral infection, enforced expression of BACH2 in adoptively transferred CD8 + T cells promotes the establishment of a memory transcriptional program and induced prolonged persistence of T cells in the absence of differentiation or exhaustion [79]. The same study revealed that ectopic expression of another TF implicated in the regulation of survival and stemness, SRY-Box transcription factor 4 (SOX4), increases memory recall response [79]; however, its potential role as an oncogene severely limits its use in cell therapy [129]. In an adoptive transfer model, enforced expression of RUNX3 improves the overall survival of B16-bearing mice and increases the accumulation of T cells over time within the tumor environment, possibly by limiting the degree of terminal differentiation/exhaustion [130]. Interestingly, RUNX3 also enforces a tissue-resident memory phenotype and function and promotes localization to the small intestine endothelium in mice, a distinctive feature that could be exploited for the treatment of colorectal cancer, for instance [130].

Table 1 List of factors involved in T cell memory or dysfunction evaluated in the context of adoptive cell transfer

The suppression of factors driving the exhaustion of T cells is another approach that has shown promising results in preclinical models. Recently, deletion of DNA methyltransferase 3A (DNMT3A) was shown to preserve the antitumor activity of CAR-T cells during prolonged tumor exposure [132]. This was accompanied by a lower methylation profile; increased expression of TCF-1, LEF-1, and CCR7; sustained cytokine secretion; and cytolytic activity upon repeat antigenic challenge. CAR-T cells depleted of DNMT3A were able to control a tumor rechallenge 148 days after the initial tumor challenge, which is indicative of long-term memory. Interestingly, a retrospective analysis of CAR-T cells in CLL revealed a significantly lower expression score for genes targeted by DNMT3A in the products from non-responders, compared to all other groups, in support of the negative role of this factor in CAR-T antitumor potency [12, 132]. Another epigenetic regulator of differentiation, histone lysine methyltransferase (SUV39H1), is involved in the silencing of stem/memory genes, and its depletion increased survival and long-term memory in a model of bacterial infection [133].

Two downstream targets of NFAT have also received recent attention: the NR4A family of orphan nuclear receptors and the TOX family of (high-mobility group box) DNA-binding proteins. Deletion of NR4A1 [91] and all three NR4A family members [118] confers robust antitumor responses to CD8 + tumor-infiltrating T lymphocytes by restoring AP-1 function. Likewise, ablation of TOX and TOX2 in adoptively transferred T cells confers almost complete immunity to an aggressive model of melanoma [112]. An important caveat here is that, under chronic conditions, TOX is required for the long-term maintenance of immunity [88, 111, 134], which makes the applicability of TOX deletion for CAR-T cell therapy questionable, especially in the context of chronic antigenic exposure, which may occur within the microenvironment of epithelial tumors. Likewise, while heterozygous ablation of BLIMP-1 increases expression of CD62L and TCF-1 on T cells [93, 140], complete abrogation of BLIMP-1 results in reduced cytotoxic activity and loss of immunity in a chronic model of viral infection [93]. This issue can be circumvented by co-down-regulation of NR4A3, a compensatory mechanism upregulated upon BLIMP-1 deletion [141], or by targeting a positive regulator of BLIMP-1 expression, hematopoietic progenitor kinase-1 (HPK1) [137]. Two groups found that a member of the chromatin remodeling cBAF complex, AT-rich interactive domain-containing protein 1A (ARID1A), was a negative determinant of Tmem cell fate and promoted the acquisition of exhaustion-associated features, respectively [135, 136]. Pharmacological or genetic ablation of ARID1A resulted, in both cases, in improved in vivo antitumor function. BTG anti-proliferation factor 1 (BTG1), a cell stress regulator, has been found to be essential for establishing exhaustion in a murine model of chronic viral infection [142]. Interestingly, it is also associated with exhaustion among CAR-NKT cells in the clinic, and BTG1 downregulation in these cells improves expansion and in vivo antitumor functionality [138].

Functional ablation of PD-1 in human CAR-T cells confers improved antitumor activity in various solid tumor xenograft models [143,144,145]. During chronic viral infection, complete PD-1 removal initially increases the proliferation of antiviral T cells during the expansion phase [139]; however, during the ensuing contraction phase, PD-1 ablation precipitates the natural decline of antiviral T cells, ultimately resulting in fewer memory cells [139, 146]. These data bring to light the duality of T cell dysfunction in cell therapy, increasing the risk for tumor immune escape due to T cell hypo-functionality but also maintaining long-term immunity in conditions of chronic stimulation. The right balance may depend on the relative tumor burden [22], with a high tumor burden leading to more prolonged antigen exposure and a more profound exhaustion state of the cell therapy product [30].

Collectively, the above data suggest that in patients with solid tumors or hematologic malignancies alike, CAR-T cells would be more successful if used earlier in the course of the disease. Instead of being used as primary salvage therapy in high-risk patients, they should be developed as maintenance therapy for patients whose disease has responded to initial therapy and who have a decreased tumor volume.

Conclusions

Despite profound clinical successes in defined and relatively small patient populations, the full potential of engineered T cell therapies across the larger oncology landscape has yet to be realized. Numerous innovative approaches are being leveraged to deliver a new generation of medicines with the potential to overcome many of the hurdles that are believed to be responsible for the limited efficacy of CAR-T cell therapy related to premature T cell exhaustion, particularly in solid tumors. The modulation of gene transcription programs responsible for differentiation and exhaustion in T cells, via ectopic expression or ablation of epigenetic regulators or transcription factors, is a promising approach to provide more durable antitumor efficacy than has been possible with unmodified CAR-T cells. Through these genetic manipulations, a slew of next-generation CAR-T cells have been designed to engage tumor targets for a longer duration while counteracting the negative effect of terminal differentiation and/or exhaustion, showing improved efficacy in preclinical models. Because of the stable nature of transcriptional modifications, these functional memory “armored” T cell therapies have the potential to provide a more durable clinical benefit than previous approaches, and perhaps even overcome the limitations of functional cell persistence that may be responsible for failures in treating solid tumors. At the forefront of this new generation of therapeutics, a first-of-its-kind clinical study is investigating the antitumor potential of LYL797, a ROR1-targeted CAR-T cell therapy armored with cJun (NCT05274451), while a second study is using mesothelin-targeted CAR-T cells armored with a dominant-negative form of the checkpoint receptor PD-1 (NCT04577326).

Future perspectives

Besides using blood-derived T cells as starting material in the manufacture of CAR-T cells, other cell subsets may offer distinct, unique advantages. The use of CD4 + T cells, innate-like T lymphocytes, as well as cord blood-derived hematopoietic stem cells to generate next-generation CAR-based therapeutic modalities is particularly tantalizing.

CD4 + T cells

Despite extensive research into the functionality of CD8 + T cells, mounting evidence suggests that other T cell subsets, in particular CD4 + T cells, may also contribute to the efficacy of CAR-T or TCR-T therapies. When combined with CD8 + T cells, CD4 + T cells with an under-differentiated phenotype dramatically improved the therapeutic activity of adoptively transferred T cells in preclinical tumor models [147, 148]. In addition to providing helper functions, CD4 + CAR-T cells can elicit potent cytotoxic activity, are less susceptible to activation-induced cell death, and express lower levels of inhibitory immune checkpoint receptors compared to CD8 + CAR-T cells [148, 149]. Interestingly, native cytolytic CD4 + T cells may be important for antitumor control in human bladder cancer [150, 151]. Altering T cell polarization toward helper T cells is another tantalizing approach to enhance the activity of CAR-T therapies. GATA3, a master regulator of Th2 T cells, is positively correlated with long-term persistence of CAR-T cells in humans [13]. A composite score of activity along the IL-6/STAT3 pathway, a key pathway of Th2 differentiation among CAR-T cells, has been correlated with positive clinical responses [12]. Finally, CAR-T cells generated under Th9-polarizing conditions have also been explored as a means of improving therapeutic activity. Th9 T cells express high levels of IL9, TNFA, and IL-2 and lower levels of IFNG; they also exhibit a less-differentiated phenotype compared to conventional CAR-T cells after prolonged culture in vitro [152]. In a humanized mouse model, these Th9-polarized CAR-T cells have demonstrated a tangible improvement in efficacy compared to conventional CAR-T cells [152].

Innate-like T lymphocytes

As an alternative to conventional CD4 + or CD8 + T cells, a new class of engineered cell therapies exploiting the favorable biological features of innate-like T lymphocytes is in development. γδ T cells and NKT cells express an invariant TCR and may be engineered to express CARs to target specific tumor-associated antigens [153,154,155,156]. Unlike CAR-T cells, NKT and γδ T cells express highly functional activating NK receptors that endow them with the ability to kill tumor cells in the absence of expression of CAR- or TCR-targeting antigens [157]. This feature enhances the potential of engineered invariant T cells to effectively eradicate heterogeneous tumors expressing variable levels of CAR-targeting antigens and to mitigate antigen escape as a resistance mechanism. Other investigators have found that CAR- or TCR-engineered NKT cells may modulate the immunosuppressive microenvironment in syngeneic mouse tumor models by minimizing the presence of myeloid-derived suppressor cells and enhancing the presence of pro-inflammatory macrophages [158]. Through the cross-priming of dendritic cells, adoptively transferred engineered NKT cells have also been demonstrated to enhance the antitumor efficacy of host CD8 + cytotoxic T cells, inducing durable protection against tumor rechallenge [159]. Thus, NKT cells have the potential to overcome many of the persistent inhibitory signals in the tumor microenvironment that promote exhaustion and limit the functional activity of CAR-T cells in cancer types that have proven to be unsusceptible to treatment. To further enhance the ability of NKT and T cells to avoid exhaustion in the face of chronic antigen stimulation or inhibitory signals from the tumor microenvironment, IL-15 armoring has been demonstrated to promote the self-renewal of progenitor exhausted cells [160, 161].

Cord blood-derived hematopoietic stem cells

A fundamental challenge of autologous CAR-T cell therapy, only partially addressed by the next-generation advancements reviewed here, is driven by the vast patient-to-patient differences in the quality of blood-derived T cells [12]. Similarly, non-genetically modified tumor-infiltrating lymphocytes, expanded in vitro for therapeutic application, have been found to be typically hypofunctional and differentiated [162]. Healthy donor-derived, allogeneic, “off-the-shelf” cell therapies are now being developed to alleviate this specific issue [163]. An important challenge of off-the-shelf CAR-T therapies manufactured from blood-derived mature T cells is the need to generate large batches of cells via extensive ex vivo expansion, which may limit their proliferative capacity and functionality in patients. New manufacturing techniques have been deployed to address this potential issue and allow CAR-T or CAR-NKT cells to maximize their capacity for expansion and persistent tumor cell killing in vivo [155, 164]. To this end, the use of umbilical cord blood-derived CD34 + hematopoietic stem cells transduced with a non-alloreactive iNKT TCR as starting material presents several unique advantages [155]. It alleviates the need to suppress expression of the endogenous TCR in an allogeneic context; furthermore, it allows for the large expansion of cells during manufacture, without leading to terminal differentiation of the end product, since a majority of the expansion phase occurs prior to the emergence of mature, functional T cells.

Although the technological advances cited here have the potential to rectify premature exhaustion and inadequate antitumor efficacy of cell therapies, definitive validation data from clinical studies are likely years away. Other strategies to curb the negative influence of the tumor microenvironment, tumor-associated hypoxia, acidity, altered lipid metabolism, and metabolomics are fields of ongoing research to advance CAR-T cell functionality. Thus, the study of additional T cell-enhancing approaches in the clinic is needed to ensure that the next generation of therapies becomes highly effective and potentially curative drugs for a wider population of patients than is currently being served by the approved autologous T cell products in hematological malignancies and emerging data from solid tumors.

Availability of data and materials

Not applicable.

Abbreviations

ALL:

Acute lymphoblastic leukemia

AP-1:

Activator protein 1

ARID1A:

AT-rich interactive domain-containing protein 1A

AUC:

Area under the curve

B-ALL:

B cell acute lymphoblastic leukemia

BATF:

B cell-activating transcription factor

BCL-2:

B cell leukemia 2

BCL-6:

B cell lymphoma 6

BLIMP-1:

B-lymphocyte-induced maturation protein 1

BTG1:

BTG anti-proliferation factor-1

CAR-T:

Chimeric antigen receptor T cell

CCR7:

Chemokine receptor type-7

CD103:

Cluster of differentiation-103

CD62L:

L-selectin

CLL:

Chronic lymphocytic leukemia

Cmax:

Maximum concentration

CR:

Complete response

CTLA-4:

Cytotoxic T-lymphocyte-associated antigen-4

CX3CR1:

Motif chemokine receptor-1

DNMT3A:

DNA methyltransferase 3A

EGR2:

Early growth response 2

EOMES:

Eomesodermin

FDA:

Food and Drug Administration

FOXO1:

Forkhead O transcription factors-1

GATA3:

GATA-binding protein-3

GvHD:

Graft-versus-host disease

GZMA:

Granzyme A

GZMB:

Granzyme B

HPK1:

Hematopoietic progenitor kinase-1

ID3:

Inhibitor of DNA-binding-3

IFN-γ:

Interferon gamma

IKZF2:

Zinc finger protein 2

IL-12:

Interleukin-12

IL-15:

Interleukin-15

IL-2:

Interleukin-2

IL-21:

Interleukin-21

IL-2R:

Interleukin-2 receptor

IL-2RB:

Interleukin-2 receptor subunit beta

IL-6:

Interleukin-6

IL-7RA:

Interleukin-7 receptor-Α

IL9:

Interleukin-9

IRF4:

Interferon regulatory factor 4

IRFs:

Interferon regulatory factors

KLRG1:

Killer cell lectin-like receptor subfamily G member 1

LAG-3:

Lymphocyte activation gene 3

LBCL:

Large B cell lymphoma

LEF-1:

Lymphoid enhancer-binding factor-1

MPEC:

Memory precursor effector cell

mTOR:

Mammalian target of rapamycin

mTORC1:

Mammalian target of rapamycin complex 1

NFAT:

Nuclear factor of activated T cell

NFATC2:

Nuclear factor of activated T cells 2

NHL:

Non-Hodgkin lymphoma

NKT:

Natural killer T cells

NR4A:

Nuclear receptor-4A

NR4A1:

Nuclear receptor-4A1

NR4A3:

Nuclear receptor-4A3

NR4As:

Nuclear receptor subfamily 4 group A

PD:

Progressive disease

PD-1:

Programmed cell death protein 1

PR:

Partial response

RUNX:

Runt-domain transcription factors

RUNX3:

Runt-domain transcription factors-3

SOX4:

SRY-box transcription factor 4

SLEC:

Short-lived effector cell

STAT3:

Signal transducer and activator of transcription-3

SUV39H1:

Histone lysine methyltransferase

TCF-1:

T cell factor-1

Tcm:

Central memory T cell

TCR:

T cell receptor

Teff:

Effector T cells

Teff:

Effector T cells

Tex:

Exhausted T cells

Texterm:

Terminally differentiated and exhausted T cells

TF:

Transcription factor

TIGIT:

T cell immunoreceptor with Ig and ITIM domain

TIL:

Tumor-infiltrating lymphocytes

TIM-3:

T cell immunoglobulin domain and mucin domain 3

Tmem:

Memory T cells

TNF:

Tumor necrosis factor

TNFA:

Tumor necrosis factor alpha

TOX:

Thymocyte selection-associated high-mobility group box

Tpex:

Precursor exhausted T cells

References

  1. Tsimberidou AM, et al. T-cell receptor-based therapy: an innovative therapeutic approach for solid tumors. J Hematol Oncol. 2021;14(1):102.

    CAS  PubMed  PubMed Central  Google Scholar 

  2. June CH, et al. CAR T cell immunotherapy for human cancer. Science. 2018;359(6382):1361–5.

    CAS  PubMed  Google Scholar 

  3. Gattinoni L, et al. T memory stem cells in health and disease. Nat Med. 2017;23(1):18–27.

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Neelapu SS, et al. Axicabtagene ciloleucel CAR T-cell therapy in refractory large B-cell lymphoma. N Engl J Med. 2017;377(26):2531–44.

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Maude SL, et al. Tisagenlecleucel in children and young adults with B-Cell lymphoblastic leukemia. N Engl J Med. 2018;378(5):439–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Porter DL, et al. Chimeric antigen receptor T cells persist and induce sustained remissions in relapsed refractory chronic lymphocytic leukemia. Sci Transl Med. 2015;7(303):303ra139.

    PubMed  PubMed Central  Google Scholar 

  7. Guzman G, et al. CAR-T therapies in solid tumors: opportunities and challenges. Curr Oncol Rep. 2023;25(5):479–89.

    PubMed  PubMed Central  Google Scholar 

  8. Norelli M, et al. Clinical pharmacology of CAR-T cells: linking cellular pharmacodynamics to pharmacokinetics and antitumor effects. Biochim Biophys Acta. 2016;1865(1):90–100.

    CAS  PubMed  Google Scholar 

  9. Stadtmauer EA, et al. Long-term safety and activity of NY-ESO-1 SPEAR T cells after autologous stem cell transplant for myeloma. Blood Adv. 2019;3(13):2022–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Sommer C, et al. Preclinical evaluation of allogeneic CAR T cells targeting BCMA for the treatment of multiple myeloma. Mol Ther. 2019;27(6):1126–38.

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Deng Q, 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(12):1878–87.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Fraietta JA, et al. Determinants of response and resistance to CD19 chimeric antigen receptor (CAR) T cell therapy of chronic lymphocytic leukemia. Nat Med. 2018;24(5):563–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Chen GM, et al. Integrative bulk and single-cell profiling of premanufacture T-cell populations reveals factors mediating long-term persistence of CAR T-cell therapy. Cancer Discov. 2021;11(9):2186–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Bai Z, et al. Single-cell multiomics dissection of basal and antigen-specific activation states of CD19-targeted CAR T cells. J Immunother Cancer. 2021;9:5.

    Google Scholar 

  15. Bai Z, et al. Single-cell antigen-specific landscape of CAR T infusion product identifies determinants of CD19-positive relapse in patients with ALL. Sci Adv. 2022;8(23):eabj2820.

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Haradhvala NJ, et al. Distinct cellular dynamics associated with response to CAR-T therapy for refractory B cell lymphoma. Nat Med. 2022;28(9):1848–59.

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Krishna S, et al. Stem-like CD8 T cells mediate response of adoptive cell immunotherapy against human cancer. Science. 2020;370(6522):1328–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Rossi J, et al. Preinfusion polyfunctional anti-CD19 chimeric antigen receptor T cells are associated with clinical outcomes in NHL. Blood. 2018;132(8):804–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. D’Angelo SP, et al. Antitumor activity associated with prolonged persistence of adoptively transferred NY-ESO-1 (c259)T cells in synovial sarcoma. Cancer Discov. 2018;8(8):944–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Sekine T, et al. TOX is expressed by exhausted and polyfunctional human effector memory CD8(+) T cells. Sci Immunol. 2020;5(49):eaba7918.

    CAS  PubMed  Google Scholar 

  21. Guo Y, et al. Phase I study of chimeric antigen receptor-modified T cells in patients with EGFR-positive advanced biliary tract cancers. Clin Cancer Res. 2018;24(6):1277–86.

    CAS  PubMed  Google Scholar 

  22. Locke FL, et al. Tumor burden, inflammation, and product attributes determine outcomes of axicabtagene ciloleucel in large B-cell lymphoma. Blood Adv. 2020;4(19):4898–911.

    PubMed  PubMed Central  Google Scholar 

  23. Sheih A, et al. Clonal kinetics and single-cell transcriptional profiling of CAR-T cells in patients undergoing CD19 CAR-T immunotherapy. Nat Commun. 2020;11(1):219.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Dickinson MJ, et al. A novel autologous CAR-T therapy, YTB323, with preserved T-cell stemness shows enhanced CAR T-cell efficacy in preclinical and early clinical development. Cancer Discov. 2023;13(9):1982–97.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Lisiero DN, et al. Enhanced sensitivity to IL-2 signaling regulates the clinical responsiveness of IL-12-primed CD8(+) T cells in a melanoma model. J Immunol. 2011;186(9):5068–77.

    CAS  PubMed  Google Scholar 

  26. van Leeuwen EM, et al. Functional re-expression of CCR7 on CMV-specific CD8+ T cells upon antigenic stimulation. Int Immunol. 2005;17(6):713–9.

    PubMed  Google Scholar 

  27. Hendriks J, Xiao Y, Borst J. CD27 promotes survival of activated T cells and complements CD28 in generation and establishment of the effector T cell pool. J Exp Med. 2003;198(9):1369–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Carr JM, et al. CD27 mediates interleukin-2-independent clonal expansion of the CD8+ T cell without effector differentiation. Proc Natl Acad Sci U S A. 2006;103(51):19454–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Raje N, et al. Anti-BCMA CAR T-cell therapy bb2121 in relapsed or refractory multiple myeloma. N Engl J Med. 2019;380(18):1726–37.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Kirouac DC, et al. Deconvolution of clinical variance in CAR-T cell pharmacology and response. Nat Biotechnol. 2023. https://doi.org/10.1038/s41587-023-01687-x.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Williams MA, Bevan MJ. Effector and memory CTL differentiation. Annu Rev Immunol. 2007;25:171–92.

    CAS  PubMed  Google Scholar 

  32. Behr FM, et al. Armed and ready: transcriptional regulation of tissue-resident memory CD8 T cells. Front Immunol. 2018;9:1770.

    PubMed  PubMed Central  Google Scholar 

  33. Wherry EJ, Kurachi M. Molecular and cellular insights into T cell exhaustion. Nat Rev Immunol. 2015;15(8):486–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Schietinger A, Greenberg PD. Tolerance and exhaustion: defining mechanisms of T cell dysfunction. Trends Immunol. 2014;35(2):51–60.

    CAS  PubMed  Google Scholar 

  35. Chen Y, et al. BATF regulates progenitor to cytolytic effector CD8(+) T cell transition during chronic viral infection. Nat Immunol. 2021;22(8):996–1007.

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Kallies A, Zehn D, Utzschneider DT. Precursor exhausted T cells: key to successful immunotherapy? Nat Rev Immunol. 2020;20(2):128–36.

    CAS  PubMed  Google Scholar 

  37. Zander R, et al. Tfh-cell-derived interleukin 21 sustains effector CD8(+) T cell responses during chronic viral infection. Immunity. 2022;55(3):475–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Philip M, et al. Chromatin states define tumour-specific T cell dysfunction and reprogramming. Nature. 2017;545(7655):452–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Sen DR, et al. The epigenetic landscape of T cell exhaustion. Science. 2016;354(6316):1165–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Schietinger A, et al. Tumor-specific T cell dysfunction is a dynamic antigen-driven differentiation program initiated early during tumorigenesis. Immunity. 2016;45(2):389–401.

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Charmoy M, et al. PD-1(+) Tcf1(+) CD8(+) T cells from established chronic infection can form memory while retaining a stableimprint of persistent antigen exposure. Cell Rep. 2021;36(10):109672.

    CAS  PubMed  Google Scholar 

  42. Scott-Browne JP, et al. Dynamic changes in chromatin accessibility occur in CD8(+) T cells responding to viral infection. Immunity. 2016;45(6):1327–40.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Ghoneim HE, et al. De novo epigenetic programs inhibit PD-1 blockade-mediated T cell rejuvenation. Cell. 2017;170(1):142–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Pauken KE, et al. Epigenetic stability of exhausted T cells limits durability of reinvigoration by PD-1 blockade. Science. 2016;354(6316):1160–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Utzschneider DT, et al. T cells maintain an exhausted phenotype after antigen withdrawal and population reexpansion. Nat Immunol. 2013;14(6):603–10.

    CAS  PubMed  Google Scholar 

  46. Kane H, et al. Longitudinal analysis of invariant natural killer T cell activation reveals a cMAF-associated transcriptional state of NKT10 cells. Elife. 2022;11:e765868.

    Google Scholar 

  47. Galletti G, et al. Two subsets of stem-like CD8(+) memory T cell progenitors with distinct fate commitments in humans. Nat Immunol. 2020;21(12):1552–62.

    PubMed  PubMed Central  Google Scholar 

  48. Gonzalez NM, et al. Schrodinger’s T Cells: molecular insights into stemness and exhaustion. Front Immunol. 2021;12:725618.

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Im SJ, et al. Defining CD8+ T cells that provide the proliferative burst after PD-1 therapy. Nature. 2016;537(7620):417–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Siddiqui I, et al. Intratumoral Tcf1(+)PD-1(+)CD8(+) T Cells with stem-like properties promote tumor control in response to vaccination and checkpoint blockade immunotherapy. Immunity. 2019;50(1):195–211.

    CAS  PubMed  Google Scholar 

  51. Sade-Feldman M, et al. Defining T cell states associated with response to checkpoint immunotherapy in melanoma. Cell. 2018;175(4):998–1013.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Brummelman J, et al. High-dimensional single cell analysis identifies stem-like cytotoxic CD8(+) T cells infiltrating human tumors. J Exp Med. 2018;215(10):2520–35.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Miller BC, et al. Subsets of exhausted CD8(+) T cells differentially mediate tumor control and respond to checkpoint blockade. Nat Immunol. 2019;20(3):326–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Kurtulus S, et al. Checkpoint blockade immunotherapy induces dynamic changes in PD-1(-)CD8(+) tumor-infiltrating T cells. Immunity. 2019;50(1):181–94.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. van der Leun AM, Thommen DS, Schumacher TN. CD8(+) T cell states in human cancer: insights from single-cell analysis. Nat Rev Cancer. 2020;20(4):218–32.

    PubMed  PubMed Central  Google Scholar 

  56. Dolina JS, et al. CD8(+) T cell exhaustion in cancer. Front Immunol. 2021;12:715234.

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Thommen DS, et al. A transcriptionally and functionally distinct PD-1(+) CD8(+) T cell pool with predictive potential in non-small-cell lung cancer treated with PD-1 blockade. Nat Med. 2018;24(7):994–1004.

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Bengsch B, et al. Epigenomic-guided mass cytometry profiling reveals disease-specific features of exhausted CD8 T cells. Immunity. 2018;48(5):1029–45.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Cui C, et al. Neoantigen-driven B cell and CD4 T follicular helper cell collaboration promotes anti-tumor CD8 T cell responses. Cell. 2021;184(25):6101–18.

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Kakaradov B, et al. Early transcriptional and epigenetic regulation of CD8(+) T cell differentiation revealed by single-cell RNA sequencing. Nat Immunol. 2017;18(4):422–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Arsenio J, et al. Early specification of CD8+ T lymphocyte fates during adaptive immunity revealed by single-cell gene-expression analyses. Nat Immunol. 2014;15(4):365–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Plumlee CR, et al. Environmental cues dictate the fate of individual CD8+ T cells responding to infection. Immunity. 2013;39(2):347–56.

    CAS  PubMed  Google Scholar 

  63. Chae WJ, Bothwell ALM. Canonical and non-canonical Wnt signaling in immune cells. Trends Immunol. 2018;39(10):830–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Jeannet G, et al. Essential role of the Wnt pathway effector Tcf-1 for the establishment of functional CD8 T cell memory. Proc Natl Acad Sci U S A. 2010;107(21):9777–82.

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Zhou X, et al. Differentiation and persistence of memory CD8(+) T cells depend on T cell factor 1. Immunity. 2010;33(2):229–40.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Xing S, et al. Tcf1 and Lef1 transcription factors establish CD8(+) T cell identity through intrinsic HDAC activity. Nat Immunol. 2016;17(6):695–703.

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Shan Q, et al. Ectopic Tcf1 expression instills a stem-like program in exhausted CD8(+) T cells to enhance viral and tumor immunity. Cell Mol Immunol. 2021;18(5):1262–77.

    CAS  PubMed  Google Scholar 

  68. Banerjee A, et al. Cutting edge: The transcription factor eomesodermin enables CD8+ T cells to compete for the memory cell niche. J Immunol. 2010;185(9):4988–92.

    CAS  PubMed  Google Scholar 

  69. Intlekofer AM, et al. Effector and memory CD8+ T cell fate coupled by T-bet and eomesodermin. Nat Immunol. 2005;6(12):1236–44.

    CAS  PubMed  Google Scholar 

  70. Hess Michelini R, et al. Differentiation of CD8 memory T cells depends on Foxo1. J Exp Med. 2013;210(6):1189–200.

    PubMed  Google Scholar 

  71. Rao RR, et al. Transcription factor Foxo1 represses T-bet-mediated effector functions and promotes memory CD8(+) T cell differentiation. Immunity. 2012;36(3):374–87.

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Delpoux A, et al. FOXO1 constrains activation and regulates senescence in CD8 T cells. Cell Rep. 2021;34(4):108674.

    CAS  PubMed  Google Scholar 

  73. Staron MM, et al. The transcription factor FoxO1 sustains expression of the inhibitory receptor PD-1 and survival of antiviral CD8(+) T cells during chronic infection. Immunity. 2014;41(5):802–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Utzschneider DT, et al. Active maintenance of T cell memory in acute and chronic viral infection depends on continuous expression of FOXO1. Cell Rep. 2018;22(13):3454–67.

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Kim MV, et al. The transcription factor Foxo1 controls central-memory CD8+ T cell responses to infection. Immunity. 2013;39(2):286–97.

    CAS  PubMed  Google Scholar 

  76. Delpoux A, et al. Continuous activity of Foxo1 is required to prevent anergy and maintain the memory state of CD8(+) T cells. J Exp Med. 2018;215(2):575–94.

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Yao C, et al. BACH2 enforces the transcriptional and epigenetic programs of stem-like CD8(+) T cells. Nat Immunol. 2021;22(3):370–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Roychoudhuri R, et al. BACH2 regulates CD8(+) T cell differentiation by controlling access of AP-1 factors to enhancers. Nat Immunol. 2016;17(7):851–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Hu G, Chen J. A genome-wide regulatory network identifies key transcription factors for memory CD8(+) T-cell development. Nat Commun. 2013;4:2830.

    PubMed  Google Scholar 

  80. Danilo M, et al. Suppression of Tcf1 by inflammatory cytokines facilitates effector CD8 T cell differentiation. Cell Rep. 2018;22(8):2107–17.

    CAS  PubMed  Google Scholar 

  81. McLane LM, et al. Differential localization of T-bet and Eomes in CD8 T cell memory populations. J Immunol. 2013;190(7):3207–15.

    CAS  PubMed  Google Scholar 

  82. Wang D, et al. The transcription factor Runx3 establishes chromatin accessibility of cis-regulatory landscapes that drive memory cytotoxic T lymphocyte formation. Immunity. 2018;48(4):659–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Kurachi M, et al. The transcription factor BATF operates as an essential differentiation checkpoint in early effector CD8+ T cells. Nat Immunol. 2014;15(4):373–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Seo H, et al. BATF and IRF4 cooperate to counter exhaustion in tumor-infiltrating CAR T cells. Nat Immunol. 2021;22(8):983–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Kao C, et al. Transcription factor T-bet represses expression of the inhibitory receptor PD-1 and sustains virus-specific CD8+ T cell responses during chronic infection. Nat Immunol. 2011;12(7):663–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Shin HM, et al. Epigenetic modifications induced by Blimp-1 Regulate CD8(+) T cell memory progression during acute virus infection. Immunity. 2013;39(4):661–75.

    CAS  PubMed  Google Scholar 

  87. Beltra JC, et al. Developmental relationships of four exhausted CD8(+) T cell subsets reveals underlying transcriptional and epigenetic landscape control mechanisms. Immunity. 2020;52(5):825–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Scott AC, et al. TOX is a critical regulator of tumour-specific T cell differentiation. Nature. 2019;571(7764):270–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Paley MA, et al. Progenitor and terminal subsets of CD8+ T cells cooperate to contain chronic viral infection. Science. 2012;338(6111):1220–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Quigley M, et al. Transcriptional analysis of HIV-specific CD8+ T cells shows that PD-1 inhibits T cell function by upregulating BATF. Nat Med. 2010;16(10):1147–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Liu X, et al. Genome-wide analysis identifies NR4A1 as a key mediator of T cell dysfunction. Nature. 2019;567(7749):525–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Li J, et al. High levels of eomes promote exhaustion of anti-tumor CD8(+) T Cells. Front Immunol. 2018;9:2981.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Shin H, et al. A role for the transcriptional repressor Blimp-1 in CD8(+) T cell exhaustion during chronic viral infection. Immunity. 2009;31(2):309–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Xin G, et al. A critical role of IL-21-induced BATF in sustaining CD8-T-cell-mediated chronic viral control. Cell Rep. 2015;13(6):1118–24.

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Kuroda S, et al. Basic leucine zipper transcription factor, ATF-like (BATF) regulates epigenetically and energetically effector CD8 T-cell differentiation via Sirt1 expression. Proc Natl Acad Sci U S A. 2011;108(36):14885–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Milner JJ, et al. Erratum: Runx3 programs CD8(+) T cell residency in non-lymphoid tissues and tumours. Nature. 2018;554(7692):392.

    CAS  PubMed  Google Scholar 

  97. Joshi NS, et al. Inflammation directs memory precursor and short-lived effector CD8(+) T cell fates via the graded expression of T-bet transcription factor. Immunity. 2007;27(2):281–95.

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Xin A, et al. A molecular threshold for effector CD8(+) T cell differentiation controlled by transcription factors Blimp-1 and T-bet. Nat Immunol. 2016;17(4):422–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Rutishauser RL, et al. Transcriptional repressor Blimp-1 promotes CD8(+) T cell terminal differentiation and represses the acquisition of central memory T cell properties. Immunity. 2009;31(2):296–308.

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Kallies A, et al. Transcriptional repressor Blimp-1 is essential for T cell homeostasis and self-tolerance. Nat Immunol. 2006;7(5):466–74.

    CAS  PubMed  Google Scholar 

  101. Backer RA, et al. A central role for Notch in effector CD8(+) T cell differentiation. Nat Immunol. 2014;15(12):1143–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Utzschneider DT, et al. High antigen levels induce an exhausted phenotype in a chronic infection without impairing T cell expansion and survival. J Exp Med. 2016;213(9):1819–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Lopez-Cantillo G, et al. CAR-T Cell performance: how to improve their persistence? Front Immunol. 2022;13:878209.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Zhang M, et al. Optimization of metabolism to improve efficacy during CAR-T cell manufacturing. J Transl Med. 2021;19(1):499.

    PubMed  PubMed Central  Google Scholar 

  105. Yan C, et al. Exhaustion-associated cholesterol deficiency dampens the cytotoxic arm of antitumor immunity. Cancer Cell. 2023;41(7):1276–93.

    CAS  PubMed  Google Scholar 

  106. Jaccard A, et al. Reductive carboxylation epigenetically instructs T cell differentiation. Nature. 2023;621(7980):849–56.

    CAS  PubMed  Google Scholar 

  107. Atsaves V, et al. AP-1 transcription factors as regulators of immune responses in cancer. Cancers. 2019;11(7):1037.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Martinez GJ, et al. The transcription factor NFAT promotes exhaustion of activated CD8(+) T cells. Immunity. 2015;42(2):265–78.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Mognol GP, et al. Exhaustion-associated regulatory regions in CD8(+) tumor-infiltrating T cells. Proc Natl Acad Sci U S A. 2017;114(13):E2776–85.

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Wagle MV, et al. Antigen-driven EGR2 expression is required for exhausted CD8(+) T cell stability and maintenance. Nat Commun. 2021;12(1):2782.

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Khan O, et al. TOX transcriptionally and epigenetically programs CD8(+) T cell exhaustion. Nature. 2019;571(7764):211–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Seo H, et al. TOX and TOX2 transcription factors cooperate with NR4A transcription factors to impose CD8(+) T cell exhaustion. Proc Natl Acad Sci U S A. 2019;116(25):12410–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Utzschneider DT, et al. T cell factor 1-expressing memory-like CD8(+) T cells sustain the immune response to chronic viral infections. Immunity. 2016;45(2):415–27.

    CAS  PubMed  Google Scholar 

  114. Wu T, et al. The TCF1-Bcl6 axis counteracts type I interferon to repress exhaustion and maintain T cell stemness. Sci Immunol. 2016;1(6):eaai8593.

    PubMed  PubMed Central  Google Scholar 

  115. Roychoudhuri R, et al. The transcription factor BACH2 promotes tumor immunosuppression. J Clin Invest. 2016;126(2):599–604.

    PubMed  PubMed Central  Google Scholar 

  116. Man K, 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(6):1129–41.

    CAS  PubMed  Google Scholar 

  117. Blank CU, et al. Defining “T cell exhaustion.” Nat Rev Immunol. 2019;19(11):665–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Chen J, et al. NR4A transcription factors limit CAR T cell function in solid tumours. Nature. 2019;567(7749):530–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Labanieh L, Mackall CL. CAR immune cells: design principles, resistance and the next generation. Nature. 2023;614(7949):635–48.

    CAS  PubMed  Google Scholar 

  120. Rafiq S, Hackett CS, Brentjens RJ. Engineering strategies to overcome the current roadblocks in CAR T cell therapy. Nat Rev Clin Oncol. 2020;17(3):147–67.

    PubMed  Google Scholar 

  121. Choi J, Crotty S. Bcl6-mediated transcriptional regulation of follicular helper T cells (T(FH)). Trends Immunol. 2021;42(4):336–49.

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Ichii H, et al. Role for Bcl-6 in the generation and maintenance of memory CD8+ T cells. Nat Immunol. 2002;3(6):558–63.

    CAS  PubMed  Google Scholar 

  123. Lynn RC, et al. c-Jun overexpression in CAR T cells induces exhaustion resistance. Nature. 2019;576(7786):293–300.

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Odagiu L, et al. Role of the orphan nuclear receptor NR4A family in T-cell biology. Front Endocrinol. 2020;11:624122.

    Google Scholar 

  125. Zhang X, et al. Depletion of BATF in CAR-T cells enhances antitumor activity by inducing resistance against exhaustion and formation of central memory cells. Cancer Cell. 2022;40(11):1407–22.

    CAS  PubMed  Google Scholar 

  126. Heitzeneder S, et al. GPC2-CAR T cells tuned for low antigen density mediate potent activity against neuroblastoma without toxicity. Cancer Cell. 2022;40(1):53–69.

    CAS  PubMed  Google Scholar 

  127. Ngai H, et al. LEF1 drives a central memory program and supports antitumor activity of natural killer T cells. Cancer Immunol Res. 2023;11(2):171–83.

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Smole A, et al. Expression of inducible factors reprograms CAR-T cells for enhanced function and safety. Cancer Cell. 2022;40(12):1470–87.

    CAS  PubMed  Google Scholar 

  129. Moreno CS. SOX4: the unappreciated oncogene. Semin Cancer Biol. 2020;67(Pt 1):57–64.

    CAS  PubMed  Google Scholar 

  130. Milner JJ, et al. Runx3 programs CD8(+) T cell residency in non-lymphoid tissues and tumours. Nature. 2017;552(7684):253–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Van Kaer L. LEF1 creates memories in iNKT cells that potentiate antitumor immunity. Cancer Immunol Res. 2023;11(2):144.

    PubMed  Google Scholar 

  132. Prinzing B, et al. Deleting DNMT3A in CAR T cells prevents exhaustion and enhances antitumor activity. Sci Transl Med. 2021;13(620):eabh0272.

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Pace L, et al. The epigenetic control of stemness in CD8(+) T cell fate commitment. Science. 2018;359(6372):177–86.

    CAS  PubMed  Google Scholar 

  134. Alfei F, et al. TOX reinforces the phenotype and longevity of exhausted T cells in chronic viral infection. Nature. 2019;571(7764):265–9.

    CAS  PubMed  Google Scholar 

  135. Guo A, et al. cBAF complex components and MYC cooperate early in CD8(+) T cell fate. Nature. 2022;607(7917):135–41.

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Belk JA, et al. Genome-wide CRISPR screens of T cell exhaustion identify chromatin remodeling factors that limit T cell persistence. Cancer Cell. 2022;40(7):768–86.

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Si J, et al. Hematopoietic progenitor kinase1 (HPK1) mediates T cell dysfunction and is a druggable target for T cell-based immunotherapies. Cancer Cell. 2020;38(4):551–66.

    CAS  PubMed  Google Scholar 

  138. Heczey A, et al. Anti-GD2 CAR-NKT cells in relapsed or refractory neuroblastoma: updated phase 1 trial interim results. Nat Med. 2023;29:1379–88.

    CAS  PubMed  Google Scholar 

  139. Odorizzi PM, et al. Genetic absence of PD-1 promotes accumulation of terminally differentiated exhausted CD8+ T cells. J Exp Med. 2015;212(7):1125–37.

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Behr FM, et al. Blimp-1 rather than hobit drives the formation of tissue-resident memory CD8(+) T cells in the lungs. Front Immunol. 2019;10:400.

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Jung IY, et al. BLIMP1 and NR4A3 transcription factors reciprocally regulate antitumor CAR T cell stemness and exhaustion. Sci Transl Med. 2022;14(670):eabn7336.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Giles JR, et al. Shared and distinct biological circuits in effector, memory and exhausted CD8(+) T cells revealed by temporal single-cell transcriptomics and epigenetics. Nat Immunol. 2022;23(11):1600–13.

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Kawalekar OU, et al. Distinct signaling of coreceptors regulates specific metabolism pathways and impacts memory development in CAR T cells. Immunity. 2016;44(2):380–90.

    CAS  PubMed  Google Scholar 

  144. Rupp LJ, et al. CRISPR/Cas9-mediated PD-1 disruption enhances anti-tumor efficacy of human chimeric antigen receptor T cells. Sci Rep. 2017;7(1):737.

    PubMed  PubMed Central  Google Scholar 

  145. Cherkassky L, et al. Human CAR T cells with cell-intrinsic PD-1 checkpoint blockade resist tumor-mediated inhibition. J Clin Invest. 2016;126(8):3130–44.

    PubMed  PubMed Central  Google Scholar 

  146. Kalia V, et al. Metabolic regulation by PD-1 signaling promotes long-lived quiescent CD8 T cell memory in mice. Sci Transl Med. 2021;13(615):eaba6006.

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Sommermeyer D, et al. Chimeric antigen receptor-modified T cells derived from defined CD8+ and CD4+ subsets confer superior antitumor reactivity in vivo. Leukemia. 2016;30(2):492–500.

    CAS  PubMed  Google Scholar 

  148. Wang D, et al. Glioblastoma-targeted CD4+ CAR T cells mediate superior antitumor activity. JCI Insight. 2018;3:10.

    Google Scholar 

  149. Teijeira A, et al. Metabolic consequences of T-cell costimulation in anticancer immunity. Cancer Immunol Res. 2019;7(10):1564–9.

    CAS  PubMed  Google Scholar 

  150. Cachot A, et al. Tumor-specific cytolytic CD4 T cells mediate immunity against human cancer. Sci Adv. 2021;7(9):1612.

    Google Scholar 

  151. Oh DY, et al. Intratumoral CD4(+) T cells mediate anti-tumor cytotoxicity in human bladder cancer. Cell. 2020;181(7):1612–25.

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Liu L, et al. Enhanced CAR-T activity against established tumors by polarizing human T cells to secrete interleukin-9. Nat Commun. 2020;11(1):5902.

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Heczey A, et al. Invariant NKT cells with chimeric antigen receptor provide a novel platform for safe and effective cancer immunotherapy. Blood. 2014;124(18):2824–33.

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Nishimoto KP, et al. Allogeneic CD20-targeted gammadelta T cells exhibit innate and adaptive antitumor activities in preclinical B-cell lymphoma models. Clin Transl Immunology. 2022;11(2):e1373.

    CAS  PubMed  PubMed Central  Google Scholar 

  155. Li YR, et al. Development of allogeneic HSC-engineered iNKT cells for off-the-shelf cancer immunotherapy. Cell Rep Med. 2021;2(11):100449.

    CAS  PubMed  PubMed Central  Google Scholar 

  156. Zhu Y, et al. Development of hematopoietic stem cell-engineered invariant natural killer T cell therapy for cancer. Cell Stem Cell. 2019;25(4):542–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Brennan PJ, Brigl M, Brenner MB. Invariant natural killer T cells: an innate activation scheme linked to diverse effector functions. Nat Rev Immunol. 2013;13(2):101–17.

    CAS  PubMed  Google Scholar 

  158. Delfanti G, et al. TCR-engineered iNKT cells induce robust antitumor response by dual targeting cancer and suppressive myeloid cells. Sci Immunol. 2022;7(74):eabn6563.

    CAS  PubMed  Google Scholar 

  159. Simonetta F, et al. Allogeneic CAR invariant natural killer T cells exert potent antitumor effects through host CD8 T-cell cross-priming. Clin Cancer Res. 2021;27(21):6054–64.

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Lee J, et al. IL-15 promotes self-renewal of progenitor exhausted CD8 T cells during persistent antigenic stimulation. Front Immunol. 2023;14:1117092.

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Kohli K, et al. IL-15 mediated expansion of rare durable memory T cells following adoptive cellular therapy. J Immunother Cancer. 2021. https://doi.org/10.1136/jitc-2020-002232.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Restifo NP, Dudley ME, Rosenberg SA. Adoptive immunotherapy for cancer: harnessing the T cell response. Nat Rev Immunol. 2012;12(4):269–81.

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Yang Y, Jacoby E, Fry TJ. Challenges and opportunities of allogeneic donor-derived CAR T cells. Curr Opin Hematol. 2015;22(6):509–15.

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Engels B, et al. Preservation of T-cell stemness with a novel expansionless CAR-T manufacturing process, which reduces manufacturing time to less than two days. Drives Enhanc CAR-T Cell Efficacy Blood. 2021;138(Supplement 1):2848–2848.

    Google Scholar 

Download references

Funding

This work was supported in part by the National Institutes of Health/National Cancer Institute award number P30 CA016672 (University of Texas MD Anderson Cancer Center). This work was also supported by Mr. and Mrs. Steven Mckenzie’s Endowment and donor funds from Jamie’s Hope and Mrs. and Mr. James Ritter for Dr. Tsimberidou’s Personalized Medicine Program (Initiative for Molecular Profiling and Advanced Cancer Therapy, IMPACT).

Author information

Authors and Affiliations

Authors

Contributions

YB, AMT, and MB wrote the main manuscript text. YB and MB prepared figures. All authors edited, reviewed, and approved the manuscript.

Corresponding author

Correspondence to Apostolia M. Tsimberidou.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

The authors permit the Journal of Hematology and Oncology to publish this work.

Competing interests

A.M.T declares receipt of clinical trial research funding (to The University of Texas MD Anderson Cancer Center) from OBI Pharma, Agenus, IMMATICS, Tachyon, Novocure, Parker Institute for Cancer Immunotherapy, Orionis, Tempus, and Tvardi; fees for consulting; or advisory roles for Avstera Therapeutics, Bioeclipse, BrYet, Diaccurate, Macrogenics, NEX-I and VinceRx. Y.B is Employee of Appia Bio. J.D is Employee of Appia Bio. The remaining authors declare no relevant conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bulliard, Y., Andersson, B.S., Baysal, M.A. et al. Reprogramming T cell differentiation and exhaustion in CAR-T cell therapy. J Hematol Oncol 16, 108 (2023). https://doi.org/10.1186/s13045-023-01504-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13045-023-01504-7

Keywords