Next Article in Journal
Estimation of ALU Repetitive Elements in Plasma as a Cost-Effective Liquid Biopsy Tool for Disease Prognosis in Breast Cancer
Next Article in Special Issue
Applications of Urinary Extracellular Vesicles in the Diagnosis and Active Surveillance of Prostate Cancer
Previous Article in Journal
Genome-Wide Analysis of lncRNA-mRNA Co-Expression Networks in CD133+/CD44+ Stem-like PDAC Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CAR-T-Derived Extracellular Vesicles: A Promising Development of CAR-T Anti-Tumor Therapy

by
Sara Pagotto
1,2,
Pasquale Simeone
2,3,
Davide Brocco
4,
Giulia Catitti
2,3,
Domenico De Bellis
2,3,
Simone Vespa
2,3,
Natalia Di Pietro
1,2,
Lisa Marinelli
4,
Antonio Di Stefano
4,
Serena Veschi
4,
Laura De Lellis
4,
Fabio Verginelli
2,4,
Francesco Kaitsas
5,
Manuela Iezzi
2,6,
Assunta Pandolfi
1,2,
Rosa Visone
1,2,
Nicola Tinari
1,2,
Ignazio Caruana
7,
Mauro Di Ianni
2,3,
Alessandro Cama
4,*,
Paola Lanuti
2,3,* and
Rosalba Florio
4
add Show full author list remove Hide full author list
1
Department of Medical, Oral and Biotechnological Sciences, “G.d’Annunzio” University of Chieti-Pescara, 66100 Chieti, Italy
2
Center for Advanced Studies and Technology (CAST), “G.d’Annunzio” University of Chieti-Pescara, 66100 Chieti, Italy
3
Department of Medicine and Aging Sciences, “G.d’Annunzio” University of Chieti-Pescara, 66100 Chieti, Italy
4
Department of Pharmacy, “G.d’Annunzio” University of Chieti-Pescara, 66100 Chieti, Italy
5
Sacred Heart Catholic University, 00168 Rome, Italy
6
Department of Neuroscience, Imaging and Clinical Sciences, “G.d’Annunzio” University of Chieti-Pescara, 66100 Chieti, Italy
7
Department of Pediatric Haemaology, Oncology and Stem Cell Transplantation, University Hospital Würzburg, 97080 Würzburg, Germany
*
Authors to whom correspondence should be addressed.
Cancers 2023, 15(4), 1052; https://doi.org/10.3390/cancers15041052
Submission received: 31 December 2022 / Revised: 26 January 2023 / Accepted: 3 February 2023 / Published: 7 February 2023

Abstract

:

Simple Summary

In this review we aim to address the potential of extracellular vesicles stemming from chimeric antigen receptor T (CAR-T) lymphocytes as therapeutic agents in tumors. We underlined how CAR-T-lymphocytes, representing one of the new frontiers of immunotherapy for the fight against refractory neoplastic diseases, demonstrated their potential effectiveness in cancer. However, the presence of physical barriers that prevent the entry of CAR-T and other immune effector cells, the hostile microenvironment that hampers persistence and activity of immune cells, as well as tumor heterogeneity resulted in their variable or low efficacy against solid tumors. The application of CAR-T-derived extracellular vesicles as therapeutic agents may improve the homing of CAR-T effector functions through their facilitated diffusion within solid tumors and at the same time might circumvent some of the adverse effects that are induced by the cellular counterpart.

Abstract

Extracellular vesicles (EVs) are a heterogenous population of plasma membrane-surrounded particles that are released in the extracellular milieu by almost all types of living cells. EVs are key players in intercellular crosstalk, both locally and systemically, given that they deliver their cargoes (consisting of proteins, lipids, mRNAs, miRNAs, and DNA fragments) to target cells, crossing biological barriers. Those mechanisms further trigger a wide range of biological responses. Interestingly, EV phenotypes and cargoes and, therefore, their functions, stem from their specific parental cells. For these reasons, EVs have been proposed as promising candidates for EV-based, cell-free therapies. One of the new frontiers of cell-based immunotherapy for the fight against refractory neoplastic diseases is represented by genetically engineered chimeric antigen receptor T (CAR-T) lymphocytes, which in recent years have demonstrated their effectiveness by reaching commercialization and clinical application for some neoplastic diseases. CAR-T-derived EVs represent a recent promising development of CAR-T immunotherapy approaches. This crosscutting innovative strategy is designed to exploit the advantages of genetically engineered cell-based immunotherapy together with those of cell-free EVs, which in principle might be safer and more efficient in crossing biological and tumor-associated barriers. In this review, we underlined the potential of CAR-T-derived EVs as therapeutic agents in tumors.

Graphical Abstract

1. Introduction

Communication has a central role in all biological functions and its evolutionary history dates to the earliest forms of living organisms on Earth. One of the mechanisms that is used by cells to communicate is based on the release of membrane-surrounded particles, named extracellular vesicles (EVs), stemming from almost all living prokaryotic and eukaryotic cell types. It was shown that their releasing processes have been conserved throughout evolution [1], and recent studies underlined their role as biological carriers of messages delivered through biological fluids and acting as key players of the intercellular crosstalk [2,3,4]. EVs carry proteins, lipids, and different types of RNAs and DNAs [5,6,7,8], which are transferred to target cells, influencing their functions [9]. It is also known that EVs are distributed in all body fluids, such as peripheral blood, saliva, tears, cerebrospinal fluid, and urine [10,11,12,13,14,15,16,17]. In the last decades, EVs have been involved in many physiological and pathological processes, therefore, attracting considerable attention both as disease biomarkers and as drug delivery vehicles [18,19,20,21,22,23].
It was demonstrated that EVs can cross different biological barriers, such as the blood-brain barrier (BBB) and tumor physical barriers [1]. Furthermore, compared to other nanoparticles, such as synthetic particles, nanomaterials, or lipid transfection agents, EVs show a low grade of cytotoxicity and immunogenicity [24,25,26,27] and, therefore, have great potential as drug delivery agents for cell-free therapies in several contexts [28,29,30].
Among genetically engineered cell-based therapies, chimeric antigen receptor (CAR)-T cells emerged as an innovative and powerful therapeutic tool against malignant cells [31]. CAR-T cells are genetically engineered T cells expressing recombinant receptors consisting of an extracellular target-recognition domain, generally arising from single-chain variable fragments (scFv), a transmembrane domain, and intracellular tyrosine-based activation motifs (ITAMS). By their extracellular domain, CAR-T cells recognize specific molecules on tumor cell membranes, while the intracellular domain triggers the T cell receptor (TCR) signaling activation, causing the death of target cells without the involvement of the major histocompatibility complex (MHC) molecules [32,33,34].
Besides the successful application of CAR-T cells in the treatment of B cell malignancies, their efficacy in other neoplastic diseases and largely in solid tumors is still limited. Several studies have underlined how the immunosuppressive tumor microenvironment (TME) that is present in these tumors negatively impacts CAR-T cell activity and long-term persistence, inducing their differentiation, exhaustion, and/or anergy [35,36,37,38]. The TME also represents a physical barrier that is difficult to penetrate by lymphocytes and CAR-T cells, reducing their anti-tumor activities [39,40]. As mentioned above, EVs do not present this limitation, being able to cross different biological barriers, including those that are present in tumors. Therefore, it is conceivable that CAR-T-derived EVs may circumvent the barriers that are encountered by cell-based approaches and may represent promising therapeutic agents in the treatment of refractory malignancies.
This review describes the results and the impact of CAR-T cells in the treatment of several malignancies, then it examines the EV biological functions and focuses on the potential of CAR-T-derived EVs as therapeutic agents for both hematological and solid tumors.

2. Chimeric Antigen Receptor (CAR)-T Cell, Structure, and Evolution

CARs are synthetic chimeric proteins that typically consist of four major components: (1) an extracellular target-binding domain, in most cases in the form of a scFv, which is derived from a monoclonal antibody that confers target antigen specificity [41]; (2) a hinge region, that provides flexibility to allow the antigen-binding domain to access the targeted epitope [42]; (3) a transmembrane domain that is derived from natural proteins including CD3ζ, CD4, CD8α, or CD28, which is an ɑ-helix within cell membranes that anchors the transgenic protein to the cellular membrane linking the extracellular antigen-binding domain to the intracellular cytoplasmic domain [43]; and (4) an intracellular signaling domain, which generally consists of one or more costimulatory domains fused to a CD3ζ or a γ chain of the high-affinity IgE Fc receptor (FcεRI) (Figure 1A). When expressed by T lymphocytes, the scFv determines the CAR antigen specificity and allows T cells to engage the antigen that is expressed by tumor cells in an MHC-independent manner. After binding to the target, cross-linked CAR molecules activate the endo-domain signaling, thus inducing the lysis of the engaged target cells through granzyme-B and perforin pathways [44,45]. Although their fundamental modular structure remained similar since their inception in the late 1980s [46], the improvement of the design of CAR-T cells has been the focus of recent years, resulting in five distinct generations according to the shape of their intracellular domain (Figure 1B,C).
The first-generation of CARs contained a single CD3ζ-chain intracellular domain, however this generation of CARs revealed two major drawbacks: the inability to achieve a high enough level of antitumor activity (insufficient IL-2 production) and a short lifespan for their lack of proliferation (Figure 1B) [47].
The second generation of CAR-T cells displayed a better immune response by leveraging the power of the co-stimulatory signaling domain (molecules such as CD28, CD134 or OX-40, CD137 or 4-1BB, etc., which are fused with CD3ζ), enhancing cytokine release, cytotoxicity, and their in vivo half-lives (Figure 1B) [48,49]. Subsequently, the third generation of CARs has been designed by combining multiple co-stimulatory signaling to improve the function of CAR-T cells (CD3ζ-CD28-OX40 or CD3ζ-CD28-41BB); however, the enhanced efficacy in comparison to second-generation CAR-T cells is still unclear (Figure 1B) [50,51,52]. The fourth generation of CARs, instead, is a second/third-generation construct including a T cell response element similar to the nuclear factor of the activated T cell (NFAT) that is capable of inducing the production of cytokines as IL-12, IL-13, and GM-CSF upon activation, thus providing benefits to T cells or the other elements of the immune system (Figure 1B) [40,53,54]. The fifth generation of CAR-T cells contains a supplementary intracellular domain of cytokine receptors (e.g., receptor β-chain fragment) with a binding site for the transcription factors such as STAT-3/5 (Figure 1B). The antigen-specific activation of this receptor simultaneously triggers three synergistic signals: the CAR-T cells activation, the memory T cell generation, and the reactivation-stimulation of the immune system [55].
Nevertheless, it must be underlined that CAR-T cell products (already used or under testing) are predominant autologous. Recently, to overcome the limitation that was observed with autologous and allogeneic CAR-T cell products, universal CAR-T (UCAR-T) cells have been developed. UCAR-T cells share the same killing mechanism as CAR-T, but they are characterized by different cell manufacturing processes and a better applicability [56]. UCAR-T cells are produced by a method that allows them to effectively abolish graft-versus-host disease (GvHD) by genetically disrupting the HLA Class I loci of the allogeneic T cells and/or TCR gene, by targeting genomic sequences in the constant regions of the endogenous α, or β subunits of the TCR or disrupting HLA-A locus of MHC gene complex. In this way, T cells do not recognize allogeneic antigens. For these purposes, zinc finger nuclease [57,58], transcription activator-like effector nuclease (TALEN) [59], and the CRISPR/Cas9 system [60,61,62] are the applied gene-editing techniques for obtaining UCAR-T cells. Many preclinical and clinical trials using UCAR-T have been activated worldwide [56,63]. Many of these studies produced UCAR-T that were directed against hematological malignancies, and CD19 is the most used target, even if CD5, CD7, CD20, CD22, and BCMA have been also used for these purposes [56,64]. Furthermore, GD2, NK group 2 member D ligand (NKG2DL) and mesothelin have also been tested for solid tumors [56,65,66].

3. Overview of CAR-T Cell Therapies in Hematological Malignancies

CAR-T cells represent one of the most revolutionary therapies for certain hematological malignancies, such as B cell acute lymphoblastic leukemia (B-ALL), diffuse large B cell lymphoma (DLBCL), multiple myeloma (MM), mantle cell lymphoma (MCL), and follicular lymphoma (FL). The US Food and Drug Administration (FDA) approved the first anti-CD19 CAR-T cell therapy for B cell malignancies in 2017. So far, another five CAR-T cell therapies have been approved, four of them targeting the CD19 antigen, and two targeting the B cell maturation antigen (BCMA) [67,68,69]. Although CAR-T cell therapies achieved outstanding outcomes in B-cell malignancies, disease relapse can be observed after treatment. The mechanisms that are responsible for CAR-T resistance and thus for disease recurrence include the antigen loss [70], the inadvertent transduction of tumor blast with CAR19 [71], and T cell exhaustion [38,72]. The most commonly observed adverse effects after CAR-T treatment are cytokine release syndrome (CRS) and immune effector cell-associated neurotoxicity syndrome (ICANS) whose management has been recently standardized [73].

3.1. Anti CD19-Approved CAR-T Cell Therapies

The first CAR-T cell product that was approved by the FDA (August 2017) was tisagenlecleucel (Kymriah™). It is an autologous CD19-targeted CAR-T cell product that is used for the pediatric and young adult treatment of relapsed/refractory (R/R) B-ALL [74]. More recently (May 2022), tisagenlecleucel was also approved by the FDA for the treatment of adult patients with R/R follicular lymphoma after two or more lines of systemic therapies [75]. FDA approval was based on the results of the global Phase 2 ELARA trial (NCT03568461), involving 94 patients. The results of this study showed an overall response rate (ORR) of 86%, even if complete responses occurred in 49% of patients (Grade ≥ 3, in 0%) neurological events in 37.1% (Grade ≥ 3, in 3%) and ICANS in 4.1% (Grade ≥ 3, in 1%). No treatment-related deaths were registered [75].
The axicabtagene ciloleucel (Yescarta™) is a cell product that is based on CD19 CAR-T cells and it was FDA approved in 2017 for the treatment of adult patients with DLBCL and with primary mediastinal large B cell lymphoma. Its approval was based on the results of the ZUMA-1 multicenter trial [67]. In this study, 111 patients received CD19 CAR-T cells; the ORR was 82%, and the complete response rate was 54%. Notably, cytokine release syndrome occurred in 90% of the treated patients (Grade ≥ 3, in 9%), while neurologic toxicity was evidenced in 78% of cases (Grade ≥ 3, in 25%).
In 2020, the FDA approved the use of brexucabtagene autoleucel (Tecartus™), a third generation CAR-T cell drug (CD19/FMC63), for treating R/R MCL patients [67]. Later, in October 2021, the FDA approved the use of Tecartus™, also for the treatment of adult patients (older than 26) that were affected by R/R B-ALL. In the pivotal Phase 2 ZUMA-3 trial (NCT02614066), 39 out of 55 patients (71%) had complete remission or complete remission with incomplete hematologic recovery, with 31 (56%) patients reaching complete remission. The most common adverse events of Grade 3 or higher were anemia (49%) and pyrexia (36%). CRS of Grade 3 or higher occurred in 13 (24%) patients, and neurological events of Grade 3 or higher occurred in 14 (25%) patients [76].
In 2021 (February), lisocabtagene maraleucel (liso-cel, Breyanzi™), a cell product based on CD19 CAR-T cells equipped with a 4-1BB costimulatory domain, was also approved by the FDA for R/R large B cell lymphoma treatment [77]. The approval was based on the results of the TRANSCEND NHL001 (NCT02631044) trial involving 268 patients [78]. The ORR was 73%, with the occurrence of CRS (42% of patients, Grade ≥ 3 in 4%) of treated patients. Neurologic toxicities occurred in 35% (95/268) of patients (Grade ≥ 3 in 12%, 31/268 patients).

3.2. Anti-BCMA Approved CAR-T Cell Therapies

Among the anti BCMA CAR-T therapies that have been approved by the FDA for hematological malignancies, the idecabtagene vicleucel (Abecma™) must be cited. It is the first anti-BCMA CAR-T cell therapy that was approved by the FDA (March 2021), used for the treatment of R/R MM after fourth-line or more therapies [69,79]. It was reported that, among patients treated with idecabtagene vicleucel (n = 128), 73% of them achieved an ORR, while for 28% of them a stringent complete response rate (sCR) was observed. Of note, CRS occurred in 85% of treated patients with mainly Grade 1–2. Neurotoxicity was observed in 18% of patients (3% Grade 3; no Grade 4 or 5). Recently, the results of a two-years follow-up of the administration of ciltacabtagene autoleucel (Carvykti™), a differentiated CAR-T therapy with two BCMA-targeting single-domain antibodies for the treatment of adult patients affected by R/R MM were published [80]. Interestingly, the ORR was 97.9% (the total number of enrolled patients = 97), with most patients (83%) achieving a complete response at 27.7 month follow-up. Among those patients, the CRS occurred in 92 patients (95%), with only 5% of Grade 3 or higher. Furthermore, ICANS was observed in 16% of the patients (mainly Grade 1–2). In 2022, the FDA approved the ciltacabtagene autoleucel immunotherapy for the treatment of adult patients with R/R MM after four or more prior lines of therapy [81].

4. CAR-T Cells in Solid Tumors

Given that CAR-T-based therapies have been used successfully to treat B cell malignancies [82], this has generated great interest in the extension of the CAR-T cell approach to treat solid tumors. Nevertheless, variable or low efficacy in this context has been demonstrated so far [83]. This is mainly associated with the nature of these malignancies, including the presence of important barriers that can prevent the entry and persistence of CAR-T and other effector cells of our immune system, the hostile microenvironment, as well as the tumor heterogeneity [84,85,86]. Therefore, a close association between the specificity of the CAR and its safety, and its efficacy may be considered when new CAR-T cells are developed against solid tumors [87,88,89].
Besides well documented side effects, CAR-T cells have attracted great interest for the treatment of tumors of different origins.

4.1. Ovarian Cancer

Relevant efforts in the study of CAR-T cell efficacy have been focused on ovarian cancer. Tumor-associated glycoprotein 72 (TAG72), being highly expressed on the ovarian cancer cell surface, has been used as a target of CAR-T cell therapy in such a context. Those CAR-T cells demonstrated significant cytotoxic functions and relevant cytokine production, increasing the lifespan of treated mice [90]. It has been shown that CAR-T cells that were developed using MUC16 as a CAR induce the inhibition of ovarian cancer progression and the suppression of malignant cells [91]. Therefore, those CAR-T cells have been proposed for ovarian cancer treatment [91]. It has been also demonstrated that the growth of SKOV3 cells that express human epidermal growth factor receptor 2 (Her2)/neu was suppressed by Her2 CAR-T cells [91]. Additional CARs, such as 5T4 and FRα, were successfully used against ovarian cancer cells [92,93].

4.2. Pancreatic Cancer

Pancreatic cancer (PC) is characterized by a heterogeneous microenvironment, contributing to disease progression and resistance to chemotherapy. Many target candidates for CAR-T cell therapies against PC are currently under investigation in several clinical trials, among them are mucin 1 (MUC1), Claudin 18.2, prostate stem cell antigen (PSCA), prominin 1 (PROM1), epidermal growth factor receptor (EGFR), and mesothelin (MSLN) [94]. In particular, MSLN was largely studied as a target candidate, given that it is a cell surface glycoprotein that is expressed on nearly all PDAC tumors (85–100%), with low expression on mesothelial cells in the pleura, peritoneum, and pericardium [95]. It is known that the efficacy of anti-mesothelin CAR-T cells is dependent on the abundance of mesothelin epitopes [96]. Those CAR-T cells were effective in vitro, against mesothelin-positive cells, and also in vivo in immunodeficient mice with ovarian cancer-expressing mesothelin (OVCAR-8), in pancreatic cancer (KLM-1), and in patient-derived tumor mesothelioma xenografts (NCI-Meso63) [96]. Recently, CEACAM7 has been identified as a new target for PC treatment due to its relevant surface expression on primary human PC tumor cells, as compared to its poor expression in normal tissues [97]. In addition, CEACAM6, together with CD318 and TSPAN8, were identified as promising targets for CAR-T cell immunotherapy against PC [94].
However, even though important results were observed in preclinical studies, MSLN-CAR clinical trials showed low anti-tumor efficacy, possibly due to the heterogeneity of MSLN expression on tumor cells, as well as limited tumor penetration, or the lack of T cell persistence and exhaustion. Instead, CAR-T cells targeting CEACAM7 show significant remission of late-stage patient-derived PC xenograft tumors. Due to its restricted tissue expression, CEACAM7-directed CAR-T cells may offer a higher safety margin than commonly used targets such as CEACAM5 and HER2 with broader systemic expression [97].

4.3. Breast Cancer

Breast cancer cells are characterized by the modified expressions of several molecules that could be candidates as potential targets for CAR-T cell therapies. In a Phase I study (NCT03545815), it was shown that the combined administration of MSLN CAR-T, together with programmed cell death protein 1 (PD-1) blockade, strongly increased the cytotoxicity and persistence of T cells within the tumor [98,99]. In addition, it is reported that the use of epithelial cell adhesion molecule, EpCAM CAR-T cells, could foster cytokine release (i.e.,: interferon-γ, IL-2, and IL-6), which in turn induces strong apoptotic effects on cancer cells [100].

4.4. Hepatocellular Carcinoma

Hepatocellular carcinoma (HCC) is the most common liver cancer, and its incidence and mortality rates are overlapping. Multitargeted CAR-T cells have been developed against HCC, and among them, Glypican-3 (GPC3), alpha-fetoprotein (AFP), c-MET, and Mucin 1 NKG2DL have been used as targets to develop CAR-T cells [101,102,103]. Interestingly, it was demonstrated that second and third-generation GPC3 CAR-T were effective against HepG2 and Huh-7 tumor cell lines, both in vitro and in vivo [101]. Therefore, two different Phase I trials using autologous second generation 41BBζ GPC3 CAR-T in adults with advanced HCC were performed [NCT 02395250; NCT03146234], and those studies are included in the 22 clinical trials investigating the use of CAR-T for HCC treatment.

4.5. Glioblastoma

Glioblastoma (GBM) is the most common and aggressive malignant primary brain tumor in adults. Current treatment options usually involve surgery followed by chemotherapy or radiotherapy; however, life expectancy still remains extremely short (one year) [104]. Similar to other solid tumors, CAR-T cell therapies against GBM still face several challenges, such as tumor heterogeneity, tumor immunosuppressive microenvironment, and CAR-T cell homing and persistence. Several GBM-specific targets have already been identified including EGFRvIII, HER2, erythropoietin-producing hepatocellular carcinoma A2 (EphA2), GD2, or IL-13Rα2 [105,106,107,108,109]. Variant III (EGFRvIII) is the most common gene mutation of epidermal growth factor receptor (EGFR) and it is present in about 52% of glioma cells, but not in normal tissues [105]. A clinical study of CAR-T-targeting EGFRvIII was effective against GBM cells, showing, however, a dose- and time-dependent release of cytokines, with a relevant induced glioma cell toxicity in patients with recurrent GBM [105]. The same study also demonstrated that CAR-T cells induced the antigenic loss, making the tumor resistant to the treatment. More recently, a bold trivalent CAR strategy using a single lentiviral construct expressing three individual CARs directed against IL-13Rα2, HER, and EphA2 was developed to address the glioma antigenic escape issue [105]. Notably, IL13Rα2 is commonly expressed in more than 75% of GBMs, playing a crucial role in GBM invasiveness and progression [110]. As a matter of fact, human IL13Rα2 CAR-T therapy improves the GBM immune microenvironment and induces the activation of host immune cells [111]. The potential of HER2 as a new target for GBM has been also investigated, given that HER2 is overexpressed in GBM, as in many other human cancer types [112]. Furthermore, an ongoing clinical trial is studying the safety and efficacy of CAR-T cells targeting HER2 in subjects with progressive recurrent or refractory HER2-positive primary central nervous system (CNS) tumors or HER2-positive tumor metastatic to the CNS after standard intervention (NCT02442297).

4.6. Prostate Cancer

Prostate cancer (PCa) represents the second most frequently diagnosed malignancy among all male malignant tumors [113]. There are three main targets of CAR-T therapy for treating prostate tumors that have been studied, including prostate-specific membrane antigen (PSMA), PSCA, and EpCAM [114]. Among them, it is known that PSMA is overexpressed in prostate cancer and the endothelium of tumor neo-vasculature, even if its expression was also evidenced in many normal tissues [115]. It has been also shown that CAR-T cells targeting PSMA display significant anti-tumor effects on PSMA-positive tumor cells in vitro and in vivo [116]. Recently, studies that were carried out on prostate cancer cells using second-generation anti-PSMA CAR-T cells exhibited high cytotoxicity in in vitro models, even if the same results were not obtained when they were used in vivo. This lack of in vivo efficacy might be due to the effects that were exerted by the dynamic inhibitory tumor microenvironment. Furthermore, prostate stem cell antigen (PSCA) is a tumor-related molecule that is expressed in 90% of prostate cancer tissues, but also in 60–70% of normal prostate tissues. Currently, two Phase I/II clinical trials are underway to evaluate the efficacy and safety of these CAR-T cells targeting PSCA in patients with advanced prostate cancer (NCT03873805, NCT02744287) [114]. In addition, EpCAM is another antigen that is expressed by many types of human epithelial carcinomas, such as lung, breast, and prostate. However, even if the expression of EpCAM in prostate cancer is variable, one clinical trial evaluating the safety and efficacy of EpCAM CAR-T cells in patients with EpCAM-positive cancer is ongoing (NCT03013712) [114].

4.7. Renal Cancer

It has been reported that the antigen carboxy-anhydrase-IX (CA-IX), a metalloprotease that is involved in carbon dioxide hydration, is expressed by different types of renal cancers. Therefore it has been proposed as a novel CAR-T cell therapy target [117,118]. It is also known that CA-IX is expressed by several normal tissues (i.e., small intestine epithelium, gastric mucosa epithelium, duodenum, and biliary tree), but with low/moderate expression [119]. Notably, the expression of CA-IX may increase in a wide range of tissues under hypoxic conditions [120]. At present, a first generation of CA-IX CAR-T cells have been produced and they demonstrated cytotoxic functions [121].

4.8. Gastric Cancer

Bi-specific Trop2/PD-L1 CAR-T cells have been developed to fight gastric cancer cells. These CAR-T cells appeared able to induce the inhibition of gastric cancer growth upon their intra-tumoral injection [122]. It has been also demonstrated that MSLN CAR-T cells are effective in inducing gastric cancer cell death and inhibition of tumor growth [123]. Other antigens have been targeted with CAR-T cells in gastric cancer, including folate receptor 1 (FOLR1), claudin 18.2, HER2, and NKG2D [124,125,126,127]. More recently, it has been shown that ICAM-1 CAR-T cells alone or in combination with CAR activation-dependent interleukin (IL)-12 release or paclitaxel, improve the outcome of advanced gastric cancer patients expressing ICAM-1 [128].

4.9. Colorectal Cancer

It is known that colorectal cancer cells prominently express CD133, TAG-72, NKG2D, and Guanylate Cyclase 2C, and specific CAR-T cells redirected versus these antigens have been generated [89,129,130,131]. Recently, it has been demonstrated that mesenchymal stem cells (MSCs) engineered to release IL-7 and IL-12 increase the anti-tumor activity of CAR-T cells with specificity for the carcinoembryonic antigen (CEA), acting against colorectal carcinoma cells and altering the inflammatory action of Th2 in the tumor milieu [132]. CAR-T cells using DCLK1 as a target are also effective against primary and metastatic colon cancer cells [133].

4.10. Lung Cancer

Many different studies have demonstrated that specific CAR-T cells can be effective against lung cancer. CAR-T cells directed against the receptor tyrosine kinase-like orphan receptor 1 (ROR1) showed a strong anti-tumor activity in human lung cancer A549 cell lines. It has been demonstrated that those ROR1 CAR-T cells are able to infiltrate the tumor tissues and eradicate several layers of tumor cells [134]. Moreover, EGFRvIII CAR-T cells have been produced and tested. Those cells strongly kill A549-EGFRvIII cells through the expression of some key molecules, such as TNF-α, IFN-γ, granzyme B, and perforin. Notably, EGFRvIII CAR-T cells were able to reduce the A549-EGFRvIII cell metastasis in mice, with a significant extension of mouse survival without any side effects [135]. Some other molecules have been successfully used as CARs to treat lung cancer. Among them, EphA2, MSLN, mucin-1, and PSCA must be cited [136,137,138]. CAR-T cells using PD-L1 as a CAR against non-small cell lung carcinoma (NSCLC) exerted anti-tumor cytotoxic effects against PD-L1-high and EGFR-mut NSCLC, leading to the recovery of PD-L1+ NSCLC patients [139].
Regarding small cell lung cancer treatment, an attractive target is represented also by delta-like 3 (DLL3). It has been shown that DLL3 CAR-T cells alone or in combination with PD-1 inhibitors are able to kill DLL3+ tumor cells [140].
All the identified and tested CARs to produce CAR T cells against solid tumors are summarized in Table 1 and Figure 2.

5. Structure and Biogenesis of Extracellular Vesicles

Conventionally, EVs have been classified based on their biogenesis processes that, in turn, tends to affect their dimension [144,145]. There are three main EV subtypes that are usually described: exosomes, microvesicles, and apoptotic bodies. More recently, additional morphologically and structurally distinct EV subtypes have been described; among them “exomeres” that are large macromolecular complexes and small non-membrane- bound nanoparticles with diameters smaller than 50 nm, have been included under the “EV” umbrella term [146]; supermeres (a morphologically distinct EV population displaying an in vivo markedly greater uptake compared with small extracellular vesicles and exomeres) [147]; and migrasomes (vesicular structures that mediate a cell migration mechanism called migracytocis) [148] have been also identified (Figure 3). Little is known about these newly described EV subtypes and they are still not included in the most recent ISEV MISEV guidelines. Conversely, exosomes, microvesicles, and apoptotic bodies have been largely studied.
Exosomes, the smallest EV subtype, display diameters ranging from 30–50 to 150 nm [149,150]. They are released by the fusion of multivesicular endosomes or bodies (MVE or MVB) with the plasma membrane [151,152], either through mechanisms that are dependent on the proteins of the “endosomal sorting complex required for transport” (ESCRT) or ESCRT-independent [153,154]. The sorting of the cargo within the exosome, which can be itself either ESCR- dependent or ESCRT-independent, is apparently highly specific and it is drowned by the formation of membrane microdomains, which are generated by the clustering of the lipids and the membrane-associated proteins. These membrane microdomains participate in the recruitment of soluble molecules, such as RNAs and cytosolic proteins that will be further sorted within exosomes [2]. In more detail, in the ESCRT-dependent pathway, the ESCRT machinery involves, as a first step, the recruitment of the ESCRT-0 and ESCRT-I subunits that, at the limit of MVB membranes, induce the clustering of ubiquitylated transmembrane cargoes at the microdomain levels. The ESCRT-II is further recruited to induce the budding and the fission of these microdomain-cargo clusters [155]. The inhibition of the ESCRT components affects the exosome release and/or the sorting of their cargoes [156]. As mentioned, the exosome release may also be drawn by an ESCRT-independent pathway, given that it has been also demonstrated that the formation of the MBVs persist even upon the depletion of four ESCRT complexes [157]. Among the ESCRT-independent mechanisms of exosome biogenesis, the pathway based on the ceramide presence must be cited. In this context, the ceramide presence is associated with the formation of membrane subdomains, through the intervention of the neutral Type II sphingomyelinase that is able to induce the hydrolysis of sphingomyelin to ceramide [158]. Furthermore, the proteins of the tetraspanin family have been also involved in the ESCRT-independent endosomal sorting. Among tetraspanins, CD9, CD81, and CD82 have been associated with the cargo sorting of exosomes [159,160], while CD63 has been involved in the endosomal sorting processes that are associated with the genesis of exosomes from the melanocytes, melanoma cells, and fibroblasts of patients affected by the Down syndrome [154,161,162,163]. If the exosome biogenesis is drawn by ESCRT-dependent or ESCRT-independent mechanisms, this could depend on the parental cell releasing the exosomes or on the cargo within the generating exosomes [159,160].
Microvesicles (or microparticles), displaying diameters ranging from 100 to 1000 nm, are directly released by the budding of the plasma membranes, and for these reasons, they are also known as shedding vesicles, ectovesicles, or ectosomes [164].
Apoptotic bodies are the largest EVs (0.1–5 µm) that are released during the latest apoptosis stages [165,166], exposing the phosphatidylserine and carrying caspases 3 and 7 and their substrates (ROCK1 and PANX1) [167].
However, more recently, the International Society of Extracellular Vesicles (ISEV) established that the aforementioned classification produced confusion, given that many recent findings have demonstrated that EV subtypes overlap in diameters, as well as in functions and cargoes [168]. For these reasons, the ISEV recommended the use of the “extracellular vesicles” umbrella term for all EV subtypes, identifying small EVs as those with diameters smaller than 200 nm and medium/large as the EVs with diameters larger than 200 nm [169].
It is known that EVs are released in physiological and pathophysiological conditions, and their characteristics are influenced by cell topography, environmental factors, and external stimuli [170,171].
For these reasons, body fluids are a heterogeneous mixture of EVs that are released by different parental cells. In peripheral blood samples, EVs stemming from leukocytes, erythrocytes, platelets, endothelial cells, and even from cancer stem cells in cancer patients have been described [13,14,20,21,22,23,172,173].
It has been also demonstrated that EVs carry specific biomolecules that are related to the features, the functions, and the actual status of their parental cells. For these reasons, by studying their phenotypes, it is possible to identify the compartment from which they originated [14,17,169,172,174,175]. Many EV phenotypes have been identified in body fluids as well as in cell culture media [14,176]. For all these reasons, EVs dynamically reflect the actual status of any compartment of the body, therefore, they have been proposed as reliable biomarkers for liquid biopsy [11,173,177,178,179,180]. On the other hand, EVs are emerging as signalosome elements that are able to mediate adaptive responses, locally and systemically, with relevant implications in physiological and pathophysiological events in vivo [10]. These functions are determined by the EV cargoes and are addressed to target cells by specific surface-exposed ligands and receptors, and able to mediate the EV binding to target cells or to the extracellular matrix. Such an interaction with target cells or EV internalization further triggers intracellular signaling pathways within the recipient cells [10]. Moreover, EVs change target cell phenotypes, transferring active receptors, such as EGFRvIII and CCR5 [181,182]. The EV uptake by the recipient cells may occur through phagocytosis or membrane fusion [183,184]. EVs transfer their cargoes to the target cells through micropinocytosis [185,186,187]. It has been also shown that the delivering of molecules by the EVs are more effective on target cells than the corresponding soluble form of the same molecule, given that plasma membranes protect EV cargoes from the circulating enzyme degradation activities [188,189]. The interaction with the target cell produces different effects that are related to the phenotype and to the features of their parental cells, as well as to the mechanisms that induced EV release [10].
In addition to phagocytosis, membrane fusion, and micropinocytosis, there are a variety of ways that are used by EVs to interact with the recipient cells, such as macropinocytosis and lipid-raft-mediated uptake [2,190,191]. It has been reported that EVs also use the cargo transfer that is mediated by tunnelling nanotubes to modify the biology of target cells [192]. The clathrin/caveolin-mediated endocytosis is the most used process of EV internalization by target cells [2,190]. When delivered within the target cell, the EV cargo exits from the endosome pathway into the cell’s cytoplasm before the endosome’s maturation and acidification, avoiding consequent degradation [193]. This escape from the endosomal pathway is crucial for functional delivery of the EV content [194]. It is currently unknown whether EVs have evolved unique methods or mechanisms similar to viruses and bacteria to escape from this fate [195]. However, the EV cargo can be degraded, recycled within the cell, or re-released as intact vesicles into the extracellular space. It has also been established that the pattern of EV surface proteins is strictly related to their cargo fate [183,196,197,198]. In any case, the EV uptake often leads to lysosomal degradation for the recycling or the elimination of molecules from the recipient cells [199]. Using these mechanisms, EVs transfer many different types of biomolecules to target cells, including nucleic acids, proteins, and lipids. The subcellular localization of cargo molecules delivered by EVs and the compartments where these molecules carry out their functions depending on the cargo type and the target cell. The direct investigation of the fate of EV cargo within intracellular exosomes is a technically challenging task and many studies typically employ the labelling of EV membrane lipids and proteins to this end. Despite recent progress in understanding the efficiency of cargo loading, the direct evidence of cargo release is still limited [200].

6. T-Lymphocytes-Derived Extracellular Vesicles

Extracellular vesicles that are secreted by T-lymphocytes modulate the activity of immune responses regulating the activity of immune cells [201]. Leukocyte-derived EVs have been implicated in many immune response functions, such as the recognition and the removal of pathogens and harmful substances, and inflammatory responses [202,203]. In such a context, great interest has been focused on T-lymphocyte-derived EVs. Several T cell subsets, such as cytotoxic T-lymphocytes (CTL), T helper (Th), and T regulatory cells (Tregs) may release active EVs within the extracellular milieu [187,204]. Stimulation conditions (i.e., the presence of co-stimulatory molecules) participate in the regulation of the release of different EV subtypes [205]. Although a large proportion of cargo proteins is also common in EVs that are produced by other cell types [206,207], they express specific T cell receptors and adhesion molecules [201,207]. The study of EVs stemming from T-lymphocytes revealed that they express different typical markers. As described below, the release of EVs from T-lymphocytes is highly influenced by lipid metabolism and signaling. For instance, silencing nSMase2, or inhibiting its activity, produces a significant decrease of ceramide production with a consequent inhibition of EV release by T cells [208]. Furthermore, aSMase plays a pivotal role in the secretion mechanisms of CTL-derived EVs [209]. In addition, diacylglycerol kinase α (DGKα), lipid kinase that controls T cell activation, has been involved in EV secretion by T-lymphocytes [210,211,212,213], and its inhibition increases EV release [211,212].
It was shown that EVs that are released from Tregs inhibit CD8+ T cell responses [214]. Furthermore, Treg-derived EVs modify dendritic functions, induce an increase in IL-10, and a decrease in IL-6 production, participating in the inhibition of immune reaction in tissues through the transfer of miR-150-5p and miR-142-3p [215].
It is worth noting that the activation of T-lymphocytes induces the release of EVs with proapoptotic functions due to FasL and Apo2 expression [216]. FasL and Apo2L on EV surfaces bind to target cells, interacting with their respective receptors, thus inducing apoptosis through the caspase pathway activation [217]. It was also shown that EVs that are released by Th cells, enriched in TCR, may participate with the immunological synapses that are formed by B lymphocytes that act as antigen presenting cells, stimulating pMHC-II signaling [218,219]. Furthermore, CTL-derived EVs express CD3, CD8, and TCR [212,220,221]. It is also interesting to note that EVs that are released by CTL and natural killer cells (NK) express some other functional molecules, such as perforin and granzyme A and B [222].
An RFA schematic representation of immune functions of T-lymphocyte-derived EVs is shown in Figure 4.

7. Chimeric Antigen Receptor (CAR)-T Cell-Derived Extracellular Vesicles

CAR-T EVs present many potential advantages when compared with the CAR-T parental cells. First, CAR-T EVs are stable, but they have a limited life span and are incapable of proliferating, designating an immunotherapy product that is associated with low collateral toxicity risks (e.g.,: low incidence of CRS) [143]. In such a context, it is known that CAR-T-induced CRS, an uncontrolled release of cytokines by CAR-T cells that is one of the most serious and frequent complications that is associated with CAR-T cell infusion, occurring approximately 10 days after the CAR-T cell administration [223]. In line with this lower toxicity potential, CAR-T EVs can be used as cell-free immunotherapy agents and, therefore, they should not be considered advance therapy medicinal products, which may simplify regulatory agency approval. Furthermore, considering their low immunogenicity in heterologous infusions, it is conceivable that they might be used in a third-party setting as off-the-shelf product. Then, differently from the parental cells, EVs may easily cross tumor-barriers, as demonstrated by the presence of tumor-derived EVs in body fluids [172,223], even in tumors that are characterized by strong fibrotic reaction [23].
Some studies also underlined how cell-free EV therapies could be implemented in terms of EV production and functionality by appropriately stimulating the producer cells [143]. Lastly, in the case of autologous CAR production from leukemia patients with circulating disease, CAR-T EV therapies may have the advantage of not incurring the risk of re-infusing tumor cells into the patient.
In this regard, it has been reported that residual tumor cells may be incidentally transduced with CAR, which may lead to CAR-T cell therapy resistance [71].
The expression of TCR and proapoptotic molecules (granzymes A and B, perforin, Apo2L, and FasL) by EVs that are derived from T-lymphocytes confer them cytotoxic functions and antigen specificity [220,222]. For these reasons, it was hypothesized that CAR-T cell-derived EVs might be potent vectors delivering proapoptotic messages to cancer cells. However, the surface expression of the CAR molecule on CAR-T cell-derived EVs is mandatory to specifically induce cancer cell death as it occurs for their parental CAR-T cells [82,223].
The production of EVs expressing the CAR can be optimized both in terms of increase of EV release and induction of CAR expression on EV surfaces. In more detail, it is known that TCR activation boosts the production of CTL-derived EVs [221], while higher levels of CAR expression in Evs can be obtained using different antigen stimulation strategies, such as cells beads that are coated with the recombinant CAR target antigen, or cells expressing the CAR antigen [143].
A carefully designed proof-of-principle study demonstrated that CAR-T EVs with EGFR and HER-2 specific CAR efficiently and specifically kill HER2+ and EGFR+ cancer cells in mouse xenograft models, without impacting cells that do not express those molecules [143]. Moreover, in that study, PD-L1 was shown to inhibit CAR-T cells, but not CAR-T EVs both in vitro and in vivo, suggesting a better resilience of EVs to tumor immune evasion strategies. Finally, the same study demonstrated that differently from CAR-T cells, CAR-T EVs did not elicit cytokine release syndrome in mouse models [143]. Altogether high target specificity, insensitivity to PD-L1 immunosuppression, and a lack of cytokine release syndrome induction appear promising advantages of CAR-T EVs on their cellular counterpart. Similar results have been obtained using EVs that are derived from mesothelin (MSLN)-targeted CAR-T cells, showing their efficiency in targeting MSLN-positive and triple-negative breast cancer cells via secretion of perforin and granzyme B. More importantly, the study demonstrated a marked antitumor effect with low toxicity in vivo, in both BT-549 and MDA231-MSLN xenograft breast tumor models [142].
Recently, it was shown that CAR-T carrying a CAR construct driving the transcription of RN7SL1, a non-coding RNA that induces the stimulation of IFN genes, secrete EVs delivering RN7LS1 [224]. Those CAR-T EVs improve immune activation against tumor cells, transferring RN7SL1 to endogenous immune cells, but not to tumor cells. RN7SL1 recipient cells, which are T, myeloid, and dendritic cells, activate IFN-dependent inflammatory responses, improving the immunostimulatory effects of dendritic and myeloid cells that end up in activation of CD8 T cells against tumor cells. These activated immune cells, in turn, orchestrate solid tumor rejection, even when tumors lack adequate expression of neoantigens [224]. Therefore, these data show that RN7SL1-containing CAR-T EVs synergize with immune cells to increase their efficacy against cancer cells.
In summary, the use of EVs derived from CAR-T cells to treat solid tumors may allow overcoming of some issues that are related to CAR T cell treatments with consequent advantages, such as the ability of EVs to easily cross the physical tumor barriers and to avoid the negative impact of the immunosuppressive tumor microenvironment that largely alters CAR-T cell functionality [35,36,37,38]. Furthermore, due to their limited life span, their inability to proliferate, and possibly, to their lower content of cytokines related to CRS and to neurotoxicity, as well as to their slower induction of apoptotic effects [141], EVs were shown to be associated to lower collateral toxicity risks as compared to CAR-T cells (e.g., low incidence of CRS) [143]. In addition, EVs are not considered advanced therapy medicinal products, which may simplify regulatory agency approval. Of note, considering their low immunogenicity in heterologous infusions it is conceivable that they might be used in a third-party setting as off-the-shelf products. Finally, CAR-T EV therapies may have the advantage of not incurring the risk of re-infusing tumor cells into the patient, given that it has been demonstrated that, together with CAR-T, residual tumor cells may be incidentally transduced with the CAR, which may lead to CAR-T cell therapy resistance [71].
An intriguing extension of this field of research concerns the antitumor potential of CAR-NK-derived EVs. Considering that CTL and NK cells use the same cytotoxic effectors (i.e., Fas L, granzymes, and perforins) [225,226,227,228], it may be speculated that also CAR-NK EVs might have a great potential as therapeutic agents. Furthermore, in vitro NK cell expansion increases EV release [229], as also demonstrated for T cells, which may provide an efficient strategy to produce Evs on large scale, particularly useful for EV preclinical and clinical applications [229]. However, NK cell recognition of targets is not dependent on antigen specificity, but it is related to the integration of different signals that are associated with many NK receptors, requiring a deep analysis of NK functions before planning any experimental design. This is probably the leading cause of the scarce development of CAR-NK, which may also interfere with the development of CAR-NK EVs therapies [229,230]. Nevertheless, given that the use of CAR-NK cells presents some advantages such as greater safety and more efficient EV production, both CAR-NKs and CAR-NK EVs appear interesting fields of research for the development of safer and more efficient antitumor therapeutic strategies.

8. Extracellular Vesicles as Drug Delivery Systems

EVs share many characteristics with synthetic nanocarriers. Both systems allow proper drug delivery by controlling the space and the time distribution of drugs, minimizing the related off-target side toxicity. The optimization of drug delivery systems (DDSs) has greatly advanced over the last few decades and sophisticated synthetic nanocarriers have been engineered to resemble biological structures and improve the circulation, targeting, and responsiveness of a wide range of therapeutics including small molecules, peptides and nucleic acids. Even if they display some advantages related to low production costs, easy upscale production, and favorable physico-chemical properties in terms of shape, size, and surface properties, they suffer from a poor in vitro-in vivo correlation (IVIVC) and unpredictable nano-bio interactions [231,232]. Starting from this evidence, published data suggest that there are still many issues in the drug delivery field for a satisfactory targeting specificity. These challenges related to the physical interactions with different biological environments may be overcome by biological or bio-derived drug delivery systems. Cell-derived EV-based carrier systems are well integrated in this frame. They combine the unique functionalities of natural materials and multifunctional design of synthetic nanomaterials. Since they are generated by an endogenous source, the natural cargo (miRNA, DNA, proteins, and lipids) composition, and their natural involvement in the intercellular communication, they have gained considerable interest and were proposed as a novel generation of delivery systems, potentially useful for solving the well-known unsatisfactory response to conventional treatments. It has been reported that EVs, due to their inherent targeting characteristics, are able to establish advantageous interactions with target cells more than liposomes, the synthetic homologous nanostructures [233].
Indeed, given that EVs home target tissues crossing the biological barriers, the EV drug delivery system, together with their pharmacokinetic profile, may result in a higher therapeutic efficacy when compared to synthetic liposomes. In such a context, the biodistribution of EVs is crucial both in terms of efficacy and safety, and, for these reasons, it was analyzed using different mouse models, using fluorescent labeled EVs, or luciferase tagged vesicles. In this way it has been demonstrated that EVs accumulate in the liver, spleen, kidney, and in gastrointestinal tract and that the biodistribution could be influenced by the route of administration, and by the EV origin and size [234,235,236]. On the other hand, it is known that liver and spleen are the main accumulation organs of liposomes.
Moreover, EVs have been further engineered to improve their delivery capabilities. They have been modified by introducing specific ligands, stimuli-responsive moieties, and immune evasive factors, resulting in carrier systems which simultaneously take advantages from biological and exogenous features. In a recent study, different cell types were genetically engineered to enrich specific siRNAs in EVs, improving the functionality of in vivo siRNA delivery, significantly reducing siRNA therapeutic doses with respect to those typically that are delivered via lipid nanoparticles [237]. Despite the EVs advantages in targeting, safety, and pharmacokinetics over the synthetic nanocarriers, many challenges that are related to the upscaling production, isolation, purification, characterization, and storage, need to be faced for EV clinical translation [238].
The above-mentioned systems are classified as biological and chemical methods and refer to carrier-mediated strategies. Further approaches that are used for the intracellular delivery of cargoes include physical techniques based on a membrane-disruption-mediated mechanism. Usually, the carried-based delivery exploits a fusion or endocytic entry route, while the membrane disruption approaches involve different entry pathways including permeabilization and direct penetration processes.
EVs have shown great potential for the intracellular delivery of different cargoes including both synthetic and biological molecules. Examples of loading mechanisms, applicable in a post-EV isolation stage mix electroporation, sonication, cycles of freeze-thaw, extrusion, and the employment of pore forming agents. It is important to highlight that the above-mentioned techniques are useful for improving the loading, but they may be responsible for different negative effects that result in a loss of the EV membrane integrity, as well as aggregate production and cargo impurity. Haney et al. observed that, for protein catalase, the loading efficiency decreased in the following order: sonication > extrusion > surfactant treatment > freeze-thaw treatment > incubation, with the higher and the lower values of 26 ± 1.2% and 4.9 ± 0.5%, respectively. Similarly, in other studies, sonication, compared to other methods, guaranteed the higher loading of chemotherapeutics and hollow gold nanomaterials [239]. The exosomes cargo loading efficiency is limited when compared to synthetic liposomes [240]. Nevertheless, many investigations are focused on the technical optimization of these loading procedures, even if guidelines for the EV loading and standardized procedures are still lacking. Indeed, EV heterogeneity hampers the evaluation of the loading performance through the entrapment efficiency and loading capacity, parameters that are commonly used to assess the suitability of a drug-carrier system [241,242].
The generation of an off-the-shelf EV-based product is emerging as a new frontier of drug delivery systems. Furthermore, EV stability remains the major issue for their implementation in clinical settings. Interestingly, Görgens et al. recently carried out an in-depth analysis of EV stability on different storage buffer formulations, studying how they affect their surface integrity and biological activity. EVs that were obtained from cell culture supernatants were stored up to 2 years at different temperatures and with different methods. They found that a freezing solution containing the HEPES buffer, albumin, and trehalose obtained the best EV preservation for long-term storage and for other downstream analyses even with several freeze-thaw cycles [243].

9. Conclusions

The application of CAR-T Evs as therapeutic agents might allow us to obtain a cell-free product that possibly results in more controllable adverse effects. It could also prevent the CRS syndrome onset and related complications. Furthermore, the use of CAR-T Evs may improve the traffic of CAR-T effector functions by the infiltration of solid tumors, given that Evs may more easily cross physical tumor barriers. Furthermore, even if some recent reports demonstrated that EV treatments should be added to CAR-T cell-based therapies [142], further studies are needed to clarify which scheme is more suitable between the substitution of CAR-T cell-based therapies with CAR-T Evs or the complementary use of CAR-T together with their Evs. Strategies to optimize the in vitro production of bioactive CAR-T Evs are also needed. Finally, strategies to increase the EV homing efficiency to their specific targets must be optimized to reduce the EV dispersion.

Author Contributions

Conceptualization, S.P., A.C., P.L., M.D.I., I.C. and R.F.; data curation, F.K., S.V. (Simone Vespa), and D.B.; writing—original draft preparation, S.P., R.F., S.V. (Serena Veschi), P.S., G.C., D.D.B., L.D.L., F.V., L.M. and N.D.P.; writing—review and editing, A.C., P.L., M.D.I. and I.C.; supervision, A.P., A.D.S., M.I., R.V. and N.T.; project administration, A.C., P.L., M.D.I., I.C. and R.F.; funding acquisition, A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Italian Ministry of University and Research (MUR), National Center for Gene Therapy and Drugs based on RNA Technology (CN00000041) and framework of PON RI 2014/2020, I.1—“Innovative PhDs with industrial characterization”, grant number DOT1353889-1 (Domenico De Bellis PhD fellowship).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simeone, P.; Bologna, G.; Lanuti, P.; Pierdomenico, L.; Guagnano, M.T.; Pieragostino, D.; Del Boccio, P.; Vergara, D.; Marchisio, M.; Miscia, S.; et al. Extracellular Vesicles as Signaling Mediators and Disease Biomarkers across Biological Barriers. Int. J. Mol. Sci. 2020, 21, 2514. [Google Scholar] [CrossRef] [PubMed]
  2. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef] [PubMed]
  3. Zaborowski, M.P.; Balaj, L.; Breakefield, X.O.; Lai, C.P. Extracellular Vesicles: Composition, Biological Relevance, and Methods of Study. Bioscience 2015, 65, 783–797. [Google Scholar] [CrossRef] [PubMed]
  4. Catitti, G.; De Bellis, D.; Vespa, S.; Simeone, P.; Canonico, B.; Lanuti, P. Extracellular Vesicles as Players in the Anti-Inflammatory Inter-Cellular Crosstalk Induced by Exercise Training. Int. J. Mol. Sci. 2022, 23, 14098. [Google Scholar] [CrossRef]
  5. Guescini, M.; Genedani, S.; Stocchi, V.; Agnati, L.F. Astrocytes and Glioblastoma cells release exosomes carrying mtDNA. J. Neural Transm. 2010, 117, 1. [Google Scholar] [CrossRef]
  6. Guescini, M.; Guidolin, D.; Vallorani, L.; Casadei, L.; Gioacchini, A.M.; Tibollo, P.; Battistelli, M.; Falcieri, E.; Battistin, L.; Agnati, L.F.; et al. C2C12 myoblasts release micro-vesicles containing mtDNA and proteins involved in signal transduction. Exp. Cell Res. 2010, 316, 1977–1984. [Google Scholar] [CrossRef]
  7. Balaj, L.; Lessard, R.; Dai, L.; Cho, Y.-J.; Pomeroy, S.L.; Breakefield, X.O.; Skog, J. Tumour microvesicles contain retrotransposon elements and amplified oncogene sequences. Nat. Commun. 2011, 2, 180. [Google Scholar] [CrossRef]
  8. Kahlert, C.; Melo, S.A.; Protopopov, A.; Tang, J.; Seth, S.; Koch, M.; Zhang, J.; Weitz, J.; Chin, L.; Futreal, A.; et al. Identification of double-stranded genomic DNA spanning all chromosomes with mutated KRAS and p53 DNA in the serum exosomes of patients with pancreatic cancer. J. Biol. Chem. 2014, 289, 3869–3875. [Google Scholar] [CrossRef]
  9. Lee, T.H.; Chennakrishnaiah, S.; Meehan, B.; Montermini, L.; Garnier, D.; D’Asti, E.; Hou, W.; Magnus, N.; Gayden, T.; Jabado, N.; et al. Barriers to horizontal cell transformation by extracellular vesicles containing oncogenic H-ras. Oncotarget 2016, 7, 51991–52002. [Google Scholar] [CrossRef]
  10. Yáñez-Mó, M.; Siljander, P.R.-M.; Andreu, Z.; Zavec, A.B.; Borràs, F.E.; Buzas, E.I.; Buzas, K.; Casal, E.; Cappello, F.; Carvalho, J.; et al. Biological properties of extracellular vesicles and their physiological functions. J. Extracell. Vesicles 2015, 4, 27066. [Google Scholar] [CrossRef] [Green Version]
  11. Falasca, K.; Lanuti, P.; Ucciferri, C.; Pieragostino, D.; Cufaro, M.C.; Bologna, G.; Federici, L.; Miscia, S.; Pontolillo, M.; Auricchio, A.; et al. Circulating extracellular vesicles as new inflammation marker in HIV infection. AIDS 2021, 35, 595–604. [Google Scholar] [CrossRef]
  12. Rossi, C.; Cicalini, I.; Cufaro, M.C.; Agnifili, L.; Mastropasqua, L.; Lanuti, P.; Marchisio, M.; De Laurenzi, V.; Del Boccio, P.; Pieragostino, D. Multi-Omics Approach for Studying Tears in Treatment-Naïve Glaucoma Patients. Int. J. Mol. Sci. 2019, 20, 4029. [Google Scholar] [CrossRef]
  13. Brocco, D.; Lanuti, P.; Pieragostino, D.; Cufaro, M.C.; Simeone, P.; Bologna, G.; Di Marino, P.; De Tursi, M.; Grassadonia, A.; Irtelli, L.; et al. Phenotypic and Proteomic Analysis Identifies Hallmarks of Blood Circulating Extracellular Vesicles in NSCLC Responders to Immune Checkpoint Inhibitors. Cancers 2021, 13, 585. [Google Scholar] [CrossRef]
  14. Marchisio, M.; Simeone, P.; Bologna, G.; Ercolino, E.; Pierdomenico, L.; Pieragostino, D.; Ventrella, A.; Antonini, F.; Del Zotto, G.; Vergara, D.; et al. Flow Cytometry Analysis of Circulating Extracellular Vesicle Subtypes from Fresh Peripheral Blood Samples. Int. J. Mol. Sci. 2020, 22, 48. [Google Scholar] [CrossRef]
  15. Pieragostino, D.; Cicalini, I.; Lanuti, P.; Ercolino, E.; di Ioia, M.; Zucchelli, M.; Zappacosta, R.; Miscia, S.; Marchisio, M.; Sacchetta, P.; et al. Enhanced release of acid sphingomyelinase-enriched exosomes generates a lipidomics signature in CSF of Multiple Sclerosis patients. Sci. Rep. 2018, 8, 3071. [Google Scholar] [CrossRef]
  16. Pieragostino, D.; Lanzini, M.; Cicalini, I.; Cufaro, M.C.; Damiani, V.; Mastropasqua, L.; De Laurenzi, V.; Nubile, M.; Lanuti, P.; Bologna, G.; et al. Tear proteomics reveals the molecular basis of the efficacy of human recombinant nerve growth factor treatment for Neurotrophic Keratopathy. Sci. Rep. 2022, 12, 1229. [Google Scholar] [CrossRef]
  17. Pieragostino, D.; Lanuti, P.; Cicalini, I.; Cufaro, M.C.; Ciccocioppo, F.; Ronci, M.; Simeone, P.; Onofrj, M.; van der Pol, E.; Fontana, A.; et al. Proteomics characterization of extracellular vesicles sorted by flow cytometry reveals a disease-specific molecular cross-talk from cerebrospinal fluid and tears in multiple sclerosis. J. Proteom. 2019, 204, 103403. [Google Scholar] [CrossRef]
  18. Matsuzaka, Y.; Yashiro, R. Extracellular Vesicles as Novel Drug-Delivery Systems through Intracellular Communications. Membranes 2022, 12, 550. [Google Scholar] [CrossRef]
  19. Van Niel, G.; Carter, D.R.F.; Clayton, A.; Lambert, D.W.; Raposo, G.; Vader, P. Challenges and directions in studying cell-cell communication by extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2022, 23, 369–382. [Google Scholar] [CrossRef]
  20. Brocco, D.; Simeone, P.; Buca, D.; Di Marino, P.; De Tursi, M.; Grassadonia, A.; De Lellis, L.; Martino, M.T.; Veschi, S.; Iezzi, M.; et al. Blood Circulating CD133+ Extracellular Vesicles Predict Clinical Outcomes in Patients with Metastatic Colorectal Cancer. Cancers 2022, 14, 1357. [Google Scholar] [CrossRef]
  21. Serafini, F.L.; Delli Pizzi, A.; Simeone, P.; Giammarino, A.; Mannetta, C.; Villani, M.; Izzi, J.; Buca, D.; Catitti, G.; Chiacchiaretta, P.; et al. Circulating Extracellular Vesicles: Their Role in Patients with Abdominal Aortic Aneurysm (AAA) Undergoing EndoVascular Aortic Repair (EVAR). Int. J. Mol. Sci. 2022, 23, 6015. [Google Scholar] [CrossRef] [PubMed]
  22. Catitti, G.; Cufaro, M.C.; De Bellis, D.; Cicalini, I.; Vespa, S.; Tonelli, F.; Miscia, G.; Secondi, L.; Simeone, P.; De Laurenzi, V.; et al. Extracellular Vesicles in Regenerative Processes Associated with Muscle Injury Recovery of Professional Athletes Undergoing Sub Maximal Strength Rehabilitation. Int. J. Mol. Sci. 2022, 23, 14913. [Google Scholar] [CrossRef] [PubMed]
  23. Brocco, D.; De Bellis, D.; Di Marino, P.; Simeone, P.; Grassadonia, A.; De Tursi, M.; Grottola, T.; Di Mola, F.F.; Di Gregorio, P.; Zappacosta, B.; et al. High Blood Concentration of Leukocyte-Derived Extracellular Vesicles Is Predictive of Favorable Clinical Outcomes in Patients with Pancreatic Cancer: Results from a Multicenter Prospective Study. Cancers 2022, 14, 4748. [Google Scholar] [CrossRef] [PubMed]
  24. Klyachko, N.L.; Arzt, C.J.; Li, S.M.; Gololobova, O.A.; Batrakova, E. V Extracellular Vesicle-Based Therapeutics: Preclinical and Clinical Investigations. Pharmaceutics 2020, 12, 1171. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, D.; Tai, P.W.L.; Gao, G. Adeno-associated virus vector as a platform for gene therapy delivery. Nat. Rev. Drug Discov. 2019, 18, 358–378. [Google Scholar] [CrossRef]
  26. Kauffman, K.J.; Dorkin, J.R.; Yang, J.H.; Heartlein, M.W.; DeRosa, F.; Mir, F.F.; Fenton, O.S.; Anderson, D.G. Optimization of Lipid Nanoparticle Formulations for mRNA Delivery in Vivo with Fractional Factorial and Definitive Screening Designs. Nano Lett. 2015, 15, 7300–7306. [Google Scholar] [CrossRef]
  27. Kanasty, R.; Dorkin, J.R.; Vegas, A.; Anderson, D. Delivery materials for siRNA therapeutics. Nat. Mater. 2013, 12, 967–977. [Google Scholar] [CrossRef]
  28. Kibria, G.; Ramos, E.K.; Wan, Y.; Gius, D.R.; Liu, H. Exosomes as a Drug Delivery System in Cancer Therapy: Potential and Challenges. Mol. Pharm. 2018, 15, 3625–3633. [Google Scholar] [CrossRef]
  29. Barile, L.; Vassalli, G. Exosomes: Therapy delivery tools and biomarkers of diseases. Pharmacol. Ther. 2017, 174, 63–78. [Google Scholar] [CrossRef]
  30. Duan, L.; Xu, L.; Xu, X.; Qin, Z.; Zhou, X.; Xiao, Y.; Liang, Y.; Xia, J. Exosome-mediated delivery of gene vectors for gene therapy. Nanoscale 2021, 13, 1387–1397. [Google Scholar] [CrossRef]
  31. Mirzaei, H.R.; Mirzaei, H.; Lee, S.Y.; Hadjati, J.; Till, B.G. Prospects for chimeric antigen receptor (CAR) γδ T cells: A potential game changer for adoptive T cell cancer immunotherapy. Cancer Lett. 2016, 380, 413–423. [Google Scholar] [CrossRef]
  32. Watanabe, K.; Kuramitsu, S.; Posey, A.D.; June, C.H. Expanding the Therapeutic Window for CAR T Cell Therapy in Solid Tumors: The Knowns and Unknowns of CAR T Cell Biology. Front. Immunol. 2018, 9, 2486. [Google Scholar] [CrossRef]
  33. Curran, K.J.; Pegram, H.J.; Brentjens, R.J. Chimeric antigen receptors for T cell immunotherapy: Current understanding and future directions. J. Gene Med. 2012, 14, 405–415. [Google Scholar] [CrossRef]
  34. Strohl, W.R.; Naso, M. Bispecific T-Cell Redirection versus Chimeric Antigen Receptor (CAR)-T Cells as Approaches to Kill Cancer Cells. Antibodies 2019, 8, 41. [Google Scholar] [CrossRef]
  35. Tumino, N.; Weber, G.; Besi, F.; Del Bufalo, F.; Bertaina, V.; Paci, P.; Quatrini, L.; Antonucci, L.; Sinibaldi, M.; Quintarelli, C.; et al. Polymorphonuclear myeloid-derived suppressor cells impair the anti-tumor efficacy of GD2.CAR T-cells in patients with neuroblastoma. J. Hematol. Oncol. 2021, 14, 191. [Google Scholar] [CrossRef]
  36. Caforio, M.; Sorino, C.; Caruana, I.; Weber, G.; Camera, A.; Cifaldi, L.; De Angelis, B.; Del Bufalo, F.; Vitale, A.; Goffredo, B.M.; et al. GD2 redirected CAR T and activated NK-cell-mediated secretion of IFNγ overcomes MYCN-dependent IDO1 inhibition, contributing to neuroblastoma cell immune escape. J. Immunother. Cancer 2021, 9, e001502. [Google Scholar] [CrossRef]
  37. Canzonetta, C.; Pelosi, A.; Di Matteo, S.; Veneziani, I.; Tumino, N.; Vacca, P.; Munari, E.; Pezzullo, M.; Theuer, C.; De Vito, R.; et al. Identification of neuroblastoma cell lines with uncommon TAZ+/mesenchymal stromal cell phenotype with strong suppressive activity on natural killer cells. J. Immunother. Cancer 2021, 9, e001313. [Google Scholar] [CrossRef]
  38. Titov, A.; Kaminskiy, Y.; Ganeeva, I.; Zmievskaya, E.; Valiullina, A.; Rakhmatullina, A.; Petukhov, A.; Miftakhova, R.; Rizvanov, A.; Bulatov, E. Knowns and Unknowns about CAR-T Cell Dysfunction. Cancers 2022, 14, 1078. [Google Scholar] [CrossRef]
  39. Marofi, F.; Motavalli, R.; Safonov, V.A.; Thangavelu, L.; Yumashev, A.V.; Alexander, M.; Shomali, N.; Chartrand, M.S.; Pathak, Y.; Jarahian, M.; et al. CAR T cells in solid tumors: Challenges and opportunities. Stem Cell Res. Ther. 2021, 12, 81. [Google Scholar] [CrossRef]
  40. Caruana, I.; Savoldo, B.; Hoyos, V.; Weber, G.; Liu, H.; Kim, E.S.; Ittmann, M.M.; Marchetti, D.; Dotti, G. Heparanase promotes tumor infiltration and antitumor activity of CAR-redirected T lymphocytes. Nat. Med. 2015, 21, 524–529. [Google Scholar] [CrossRef] [Green Version]
  41. Gross, G.; Waks, T.; Eshhar, Z. Expression of immunoglobulin-T-cell receptor chimeric molecules as functional receptors with antibody-type specificity. Proc. Natl. Acad. Sci. USA 1989, 86, 10024–10028. [Google Scholar] [CrossRef] [PubMed]
  42. Hudecek, M.; Sommermeyer, D.; Kosasih, P.L.; Silva-Benedict, A.; Liu, L.; Rader, C.; Jensen, M.C.; Riddell, S.R. The nonsignaling extracellular spacer domain of chimeric antigen receptors is decisive for in vivo antitumor activity. Cancer Immunol. Res. 2015, 3, 125–135. [Google Scholar] [CrossRef] [PubMed]
  43. Bridgeman, J.S.; Hawkins, R.E.; Bagley, S.; Blaylock, M.; Holland, M.; Gilham, D.E. The optimal antigen response of chimeric antigen receptors harboring the CD3zeta transmembrane domain is dependent upon incorporation of the receptor into the endogenous TCR/CD3 complex. J. Immunol. 2010, 184, 6938–6949. [Google Scholar] [CrossRef] [PubMed]
  44. Caruana, I.; Diaconu, I.; Dotti, G. From monoclonal antibodies to chimeric antigen receptors for the treatment of human malignancies. Semin. Oncol. 2014, 41, 661–666. [Google Scholar] [CrossRef]
  45. Benmebarek, M.-R.; Karches, C.H.; Cadilha, B.L.; Lesch, S.; Endres, S.; Kobold, S. Killing Mechanisms of Chimeric Antigen Receptor (CAR) T Cells. Int. J. Mol. Sci. 2019, 20, 1283. [Google Scholar] [CrossRef]
  46. Kuwana, Y.; Asakura, Y.; Utsunomiya, N.; Nakanishi, M.; Arata, Y.; Itoh, S.; Nagase, F.; Kurosawa, Y. Expression of chimeric receptor composed of immunoglobulin-derived V regions and T-cell receptor-derived C regions. Biochem. Biophys. Res. Commun. 1987, 149, 960–968. [Google Scholar] [CrossRef]
  47. Eshhar, Z.; Waks, T.; Gross, G.; Schindler, D.G. Specific activation and targeting of cytotoxic lymphocytes through chimeric single chains consisting of antibody-binding domains and the gamma or zeta subunits of the immunoglobulin and T-cell receptors. Proc. Natl. Acad. Sci. USA 1993, 90, 720–724. [Google Scholar] [CrossRef]
  48. Cappell, K.M.; Kochenderfer, J.N. A comparison of chimeric antigen receptors containing CD28 versus 4-1BB costimulatory domains. Nat. Rev. Clin. Oncol. 2021, 18, 715–727. [Google Scholar] [CrossRef]
  49. Roselli, E.; Faramand, R.; Davila, M.L. Insight into next-generation CAR therapeutics: Designing CAR T cells to improve clinical outcomes. J. Clin. Investig. 2021, 131, e142030. [Google Scholar] [CrossRef]
  50. Till, B.G.; Jensen, M.C.; Wang, J.; Qian, X.; Gopal, A.K.; Maloney, D.G.; Lindgren, C.G.; Lin, Y.; Pagel, J.M.; Budde, L.E.; et al. CD20-specific adoptive immunotherapy for lymphoma using a chimeric antigen receptor with both CD28 and 4-1BB domains: Pilot clinical trial results. Blood 2012, 119, 3940–3950. [Google Scholar] [CrossRef] [Green Version]
  51. Holzinger, A.; Barden, M.; Abken, H. The growing world of CAR T cell trials: A systematic review. Cancer Immunol. Immunother. 2016, 65, 1433–1450. [Google Scholar] [CrossRef]
  52. Ramos, C.A.; Rouce, R.; Robertson, C.S.; Reyna, A.; Narala, N.; Vyas, G.; Mehta, B.; Zhang, H.; Dakhova, O.; Carrum, G.; et al. In Vivo Fate and Activity of Second- versus Third-Generation CD19-Specific CAR-T Cells in B Cell Non-Hodgkin’s Lymphomas. Mol. Ther. 2018, 26, 2727–2737. [Google Scholar] [CrossRef]
  53. Chmielewski, M.; Abken, H. TRUCKs: The fourth generation of CARs. Expert Opin. Biol. Ther. 2015, 15, 1145–1154. [Google Scholar] [CrossRef]
  54. Buechner, J.; Caruana, I.; Künkele, A.; Rives, S.; Vettenranta, K.; Bader, P.; Peters, C.; Baruchel, A.; Calkoen, F.G. Chimeric Antigen Receptor T-Cell Therapy in Paediatric B-Cell Precursor Acute Lymphoblastic Leukaemia: Curative Treatment Option or Bridge to Transplant? Front. Pediatr. 2021, 9, 784024. [Google Scholar] [CrossRef]
  55. Kagoya, Y.; Tanaka, S.; Guo, T.; Anczurowski, M.; Wang, C.-H.; Saso, K.; Butler, M.O.; Minden, M.D.; Hirano, N. A novel chimeric antigen receptor containing a JAK-STAT signaling domain mediates superior antitumor effects. Nat. Med. 2018, 24, 352–359. [Google Scholar] [CrossRef]
  56. Depil, S.; Duchateau, P.; Grupp, S.A.; Mufti, G.; Poirot, L. “Off-the-shelf” allogeneic CAR T cells: Development and challenges. Nat. Rev. Drug Discov. 2020, 19, 185–199. [Google Scholar] [CrossRef]
  57. Torikai, H.; Reik, A.; Liu, P.-Q.; Zhou, Y.; Zhang, L.; Maiti, S.; Huls, H.; Miller, J.C.; Kebriaei, P.; Rabinovich, B.; et al. A foundation for universal T-cell based immunotherapy: T cells engineered to express a CD19-specific chimeric-antigen-receptor and eliminate expression of endogenous TCR. Blood 2012, 119, 5697–5705. [Google Scholar] [CrossRef]
  58. Torikai, H.; Reik, A.; Soldner, F.; Warren, E.H.; Yuen, C.; Zhou, Y.; Crossland, D.L.; Huls, H.; Littman, N.; Zhang, Z.; et al. Toward eliminating HLA class I expression to generate universal cells from allogeneic donors. Blood 2013, 122, 1341–1349. [Google Scholar] [CrossRef]
  59. Poirot, L.; Philip, B.; Schiffer-Mannioui, C.; Le Clerre, D.; Chion-Sotinel, I.; Derniame, S.; Potrel, P.; Bas, C.; Lemaire, L.; Galetto, R.; et al. Multiplex Genome-Edited T-cell Manufacturing Platform for “Off-the-Shelf” Adoptive T-cell Immunotherapies. Cancer Res. 2015, 75, 3853–3864. [Google Scholar] [CrossRef]
  60. Ren, J.; Liu, X.; Fang, C.; Jiang, S.; June, C.H.; Zhao, Y. Multiplex Genome Editing to Generate Universal CAR T Cells Resistant to PD1 Inhibition. Clin. Cancer Res. 2017, 23, 2255–2266. [Google Scholar] [CrossRef] [Green Version]
  61. Georgiadis, C.; Preece, R.; Nickolay, L.; Etuk, A.; Petrova, A.; Ladon, D.; Danyi, A.; Humphryes-Kirilov, N.; Ajetunmobi, A.; Kim, D.; et al. Long Terminal Repeat CRISPR-CAR-Coupled “Universal” T Cells Mediate Potent Anti-leukemic Effects. Mol. Ther. 2018, 26, 1215–1227. [Google Scholar] [CrossRef] [PubMed]
  62. Knott, G.J.; Doudna, J.A. CRISPR-Cas guides the future of genetic engineering. Science 2018, 361, 866–869. [Google Scholar] [CrossRef] [PubMed]
  63. Morgan, M.A.; Büning, H.; Sauer, M.; Schambach, A. Use of Cell and Genome Modification Technologies to Generate Improved “Off-the-Shelf” CAR T and CAR NK Cells. Front. Immunol. 2020, 11, 1965. [Google Scholar] [CrossRef] [PubMed]
  64. Cooper, M.L.; DiPersio, J.F. Chimeric antigen receptor T cells (CAR-T) for the treatment of T-cell malignancies. Best Pract. Res. Clin. Haematol. 2019, 32, 101097. [Google Scholar] [CrossRef]
  65. Feldmann, A.; Arndt, C.; Koristka, S.; Berndt, N.; Bergmann, R.; Bachmann, M.P. Conventional CARs versus modular CARs. Cancer Immunol. Immunother. 2019, 68, 1713–1719. [Google Scholar] [CrossRef]
  66. Martínez Bedoya, D.; Dutoit, V.; Migliorini, D. Allogeneic CAR T Cells: An Alternative to Overcome Challenges of CAR T Cell Therapy in Glioblastoma. Front. Immunol. 2021, 12, 640082. [Google Scholar] [CrossRef]
  67. Neelapu, S.S.; Locke, F.L.; Bartlett, N.L.; Lekakis, L.J.; Miklos, D.B.; Jacobson, C.A.; Braunschweig, I.; Oluwole, O.O.; Siddiqi, T.; Lin, Y.; et al. Axicabtagene Ciloleucel CAR T-Cell Therapy in Refractory Large B-Cell Lymphoma. N. Engl. J. Med. 2017, 377, 2531–2544. [Google Scholar] [CrossRef]
  68. Maude, S.L.; Laetsch, T.W.; Buechner, J.; Rives, S.; Boyer, M.; Bittencourt, H.; Bader, P.; Verneris, M.R.; Stefanski, H.E.; Myers, G.D.; et al. Tisagenlecleucel in Children and Young Adults with B-Cell Lymphoblastic Leukemia. N. Engl. J. Med. 2018, 378, 439–448. [Google Scholar] [CrossRef]
  69. Sharma, P.; Kanapuru, B.; George, B.; Lin, X.; Xu, Z.; Bryan, W.W.; Pazdur, R.; Theoret, M.R. FDA Approval Summary: Idecabtagene Vicleucel for Relapsed or Refractory Multiple Myeloma. Clin. Cancer Res. 2022, 28, 1759–1764. [Google Scholar] [CrossRef]
  70. Grupp, S.A.; Kalos, M.; Barrett, D.; Aplenc, R.; Porter, D.L.; Rheingold, S.R.; Teachey, D.T.; Chew, A.; Hauck, B.; Wright, J.F.; et al. Chimeric antigen receptor-modified T cells for acute lymphoid leukemia. N. Engl. J. Med. 2013, 368, 1509–1518. [Google Scholar] [CrossRef] [Green Version]
  71. Ruella, M.; Xu, J.; Barrett, D.M.; Fraietta, J.A.; Reich, T.J.; Ambrose, D.E.; Klichinsky, M.; Shestova, O.; Patel, P.R.; Kulikovskaya, I.; et al. Induction of resistance to chimeric antigen receptor T cell therapy by transduction of a single leukemic B cell. Nat. Med. 2018, 24, 1499–1503. [Google Scholar] [CrossRef]
  72. Kloss, C.C.; Lee, J.; Zhang, A.; Chen, F.; Melenhorst, J.J.; Lacey, S.F.; Maus, M.V.; Fraietta, J.A.; Zhao, Y.; June, C.H. Dominant-Negative TGF-β Receptor Enhances PSMA-Targeted Human CAR T Cell Proliferation And Augments Prostate Cancer Eradication. Mol. Ther. 2018, 26, 1855–1866. [Google Scholar] [CrossRef]
  73. Yakoub-Agha, I.; Chabannon, C.; Bader, P.; Basak, G.W.; Bonig, H.; Ciceri, F.; Corbacioglu, S.; Duarte, R.F.; Einsele, H.; Hudecek, M.; et al. Management of adults and children undergoing chimeric antigen receptor T-cell therapy: Best practice recommendations of the European Society for Blood and Marrow Transplantation (EBMT) and the Joint Accreditation Committee of ISCT and EBMT (JACIE). Haematologica 2020, 105, 297–316. [Google Scholar] [CrossRef]
  74. Schuster, S.J.; Bishop, M.R.; Tam, C.S.; Waller, E.K.; Borchmann, P.; McGuirk, J.P.; Jäger, U.; Jaglowski, S.; Andreadis, C.; Westin, J.R.; et al. Tisagenlecleucel in Adult Relapsed or Refractory Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2019, 380, 45–56. [Google Scholar] [CrossRef]
  75. Fowler, N.H.; Dickinson, M.; Dreyling, M.; Martinez-Lopez, J.; Kolstad, A.; Butler, J.; Ghosh, M.; Popplewell, L.; Chavez, J.C.; Bachy, E.; et al. Tisagenlecleucel in adult relapsed or refractory follicular lymphoma: The phase 2 ELARA trial. Nat. Med. 2022, 28, 325–332. [Google Scholar] [CrossRef]
  76. Shah, B.D.; Ghobadi, A.; Oluwole, O.O.; Logan, A.C.; Boissel, N.; Cassaday, R.D.; Leguay, T.; Bishop, M.R.; Topp, M.S.; Tzachanis, D.; et al. KTE-X19 for relapsed or refractory adult B-cell acute lymphoblastic leukaemia: Phase 2 results of the single-arm, open-label, multicentre ZUMA-3 study. Lancet 2021, 398, 491–502. [Google Scholar] [CrossRef]
  77. Abramson, J.S.; Palomba, M.L.; Gordon, L.I.; Lunning, M.A.; Wang, M.; Arnason, J.; Mehta, A.; Purev, E.; Maloney, D.G.; Andreadis, C.; et al. Lisocabtagene maraleucel for patients with relapsed or refractory large B-cell lymphomas (TRANSCEND NHL 001): A multicentre seamless design study. Lancet 2020, 396, 839–852. [Google Scholar] [CrossRef]
  78. Ogasawara, K.; Dodds, M.; Mack, T.; Lymp, J.; Dell’Aringa, J.; Smith, J. Population Cellular Kinetics of Lisocabtagene Maraleucel, an Autologous CD19-Directed Chimeric Antigen Receptor T-Cell Product, in Patients with Relapsed/Refractory Large B-Cell Lymphoma. Clin. Pharmacokinet. 2021, 60, 1621–1633. [Google Scholar] [CrossRef]
  79. Munshi, N.C.; Anderson, L.D.; Shah, N.; Madduri, D.; Berdeja, J.; Lonial, S.; Raje, N.; Lin, Y.; Siegel, D.; Oriol, A.; et al. Idecabtagene Vicleucel in Relapsed and Refractory Multiple Myeloma. N. Engl. J. Med. 2021, 384, 705–716. [Google Scholar] [CrossRef]
  80. Martin, T.; Usmani, S.Z.; Berdeja, J.G.; Agha, M.; Cohen, A.D.; Hari, P.; Avigan, D.; Deol, A.; Htut, M.; Lesokhin, A.; et al. Ciltacabtagene Autoleucel, an Anti-B-cell Maturation Antigen Chimeric Antigen Receptor T-Cell Therapy, for Relapsed/Refractory Multiple Myeloma: CARTITUDE-1 2-Year Follow-Up. J. Clin. Oncol. 2022, JCO.22.00842. [Google Scholar] [CrossRef]
  81. Berdeja, J.G.; Madduri, D.; Usmani, S.Z.; Jakubowiak, A.; Agha, M.; Cohen, A.D.; Stewart, A.K.; Hari, P.; Htut, M.; Lesokhin, A.; et al. Ciltacabtagene autoleucel, a B-cell maturation antigen-directed chimeric antigen receptor T-cell therapy in patients with relapsed or refractory multiple myeloma (CARTITUDE-1): A phase 1b/2 open-label study. Lancet 2021, 398, 314–324. [Google Scholar] [CrossRef] [PubMed]
  82. June, C.H.; O’Connor, R.S.; Kawalekar, O.U.; Ghassemi, S.; Milone, M.C. CAR T cell immunotherapy for human cancer. Science 2018, 359, 1361–1365. [Google Scholar] [CrossRef] [PubMed]
  83. Chan, J.D.; Lai, J.; Slaney, C.Y.; Kallies, A.; Beavis, P.A.; Darcy, P.K. Cellular networks controlling T cell persistence in adoptive cell therapy. Nat. Rev. Immunol. 2021, 21, 769–784. [Google Scholar] [CrossRef]
  84. Pellegrino, M.; Del Bufalo, F.; De Angelis, B.; Quintarelli, C.; Caruana, I.; de Billy, E. Manipulating the Metabolism to Improve the Efficacy of CAR T-Cell Immunotherapy. Cells 2020, 10, 14. [Google Scholar] [CrossRef] [PubMed]
  85. Caruana, I.; Simula, L.; Locatelli, F.; Campello, S. T lymphocytes against solid malignancies: Winning ways to defeat tumours. Cell Stress 2018, 2, 200–212. [Google Scholar] [CrossRef]
  86. Vinci, M.; Burford, A.; Molinari, V.; Kessler, K.; Popov, S.; Clarke, M.; Taylor, K.R.; Pemberton, H.N.; Lord, C.J.; Gutteridge, A.; et al. Functional diversity and cooperativity between subclonal populations of pediatric glioblastoma and diffuse intrinsic pontine glioma cells. Nat. Med. 2018, 24, 1204–1215. [Google Scholar] [CrossRef]
  87. Min, I.M.; Shevlin, E.; Vedvyas, Y.; Zaman, M.; Wyrwas, B.; Scognamiglio, T.; Moore, M.D.; Wang, W.; Park, S.; Park, S.; et al. CAR T Therapy Targeting ICAM-1 Eliminates Advanced Human Thyroid Tumors. Clin. Cancer Res. 2017, 23, 7569–7583. [Google Scholar] [CrossRef]
  88. Park, S.; Shevlin, E.; Vedvyas, Y.; Zaman, M.; Park, S.; Hsu, Y.-M.S.; Min, I.M.; Jin, M.M. Micromolar affinity CAR T cells to ICAM-1 achieves rapid tumor elimination while avoiding systemic toxicity. Sci. Rep. 2017, 7, 14366. [Google Scholar] [CrossRef]
  89. Magee, M.S.; Abraham, T.S.; Baybutt, T.R.; Flickinger, J.C.; Ridge, N.A.; Marszalowicz, G.P.; Prajapati, P.; Hersperger, A.R.; Waldman, S.A.; Snook, A.E. Human GUCY2C-Targeted Chimeric Antigen Receptor (CAR)-Expressing T Cells Eliminate Colorectal Cancer Metastases. Cancer Immunol. Res. 2018, 6, 509–516. [Google Scholar] [CrossRef]
  90. Murad, J.P.; Kozlowska, A.K.; Lee, H.J.; Ramamurthy, M.; Chang, W.-C.; Yazaki, P.; Colcher, D.; Shively, J.; Cristea, M.; Forman, S.J.; et al. Effective Targeting of TAG72+ Peritoneal Ovarian Tumors via Regional Delivery of CAR-Engineered T Cells. Front. Immunol. 2018, 9, 2268. [Google Scholar] [CrossRef]
  91. Chekmasova, A.A.; Rao, T.D.; Nikhamin, Y.; Park, K.J.; Levine, D.A.; Spriggs, D.R.; Brentjens, R.J. Successful eradication of established peritoneal ovarian tumors in SCID-Beige mice following adoptive transfer of T cells genetically targeted to the MUC16 antigen. Clin. Cancer Res. 2010, 16, 3594–3606. [Google Scholar] [CrossRef]
  92. Owens, G.L.; Sheard, V.E.; Kalaitsidou, M.; Blount, D.; Lad, Y.; Cheadle, E.J.; Edmondson, R.J.; Kooner, G.; Gilham, D.E.; Harrop, R. Preclinical Assessment of CAR T-Cell Therapy Targeting the Tumor Antigen 5T4 in Ovarian Cancer. J. Immunother. 2018, 41, 130–140. [Google Scholar] [CrossRef]
  93. Zuo, S.; Wen, Y.; Panha, H.; Dai, G.; Wang, L.; Ren, X.; Fu, K. Modification of cytokine-induced killer cells with folate receptor alpha (FRα)-specific chimeric antigen receptors enhances their antitumor immunity toward FRα-positive ovarian cancers. Mol. Immunol. 2017, 85, 293–304. [Google Scholar] [CrossRef]
  94. Schäfer, D.; Tomiuk, S.; Küster, L.N.; Al Rawashdeh, W.; Henze, J.; Tischler-Höhle, G.; Agorku, D.J.; Brauner, J.; Linnartz, C.; Lock, D.; et al. Identification of CD318, TSPAN8 and CD66c as target candidates for CAR T cell based immunotherapy of pancreatic adenocarcinoma. Nat. Commun. 2021, 12, 1453. [Google Scholar] [CrossRef]
  95. Cutmore, L.C.; Brown, N.F.; Raj, D.; Chauduri, S.; Wang, P.; Maher, J.; Wang, Y.; Lemoine, N.R.; Marshall, J.F. Pancreatic Cancer UK Grand Challenge: Developments and challenges for effective CAR T cell therapy for pancreatic ductal adenocarcinoma. Pancreatology 2020, 20, 394–408. [Google Scholar] [CrossRef]
  96. Tomar, S.; Zhang, J.; Khanal, M.; Hong, J.; Venugopalan, A.; Jiang, Q.; Sengupta, M.; Miettinen, M.; Li, N.; Pastan, I.; et al. Development of Highly Effective Anti-Mesothelin hYP218 Chimeric Antigen Receptor T Cells With Increased Tumor Infiltration and Persistence for Treating Solid Tumors. Mol. Cancer Ther. 2022, 21, 1195–1206. [Google Scholar] [CrossRef]
  97. Raj, D.; Nikolaidi, M.; Garces, I.; Lorizio, D.; Castro, N.M.; Caiafa, S.G.; Moore, K.; Brown, N.F.; Kocher, H.M.; Duan, X.; et al. CEACAM7 Is an Effective Target for CAR T-cell Therapy of Pancreatic Ductal Adenocarcinoma. Clin. Cancer Res. 2021, 27, 1538–1552. [Google Scholar] [CrossRef]
  98. Cherkassky, L.; Morello, A.; Villena-Vargas, J.; Feng, Y.; Dimitrov, D.S.; Jones, D.R.; Sadelain, M.; Adusumilli, P.S. Human CAR T cells with cell-intrinsic PD-1 checkpoint blockade resist tumor-mediated inhibition. J. Clin. Investig. 2016, 126, 3130–3144. [Google Scholar] [CrossRef]
  99. Hu, W.; Zi, Z.; Jin, Y.; Li, G.; Shao, K.; Cai, Q.; Ma, X.; Wei, F. CRISPR/Cas9-mediated PD-1 disruption enhances human mesothelin-targeted CAR T cell effector functions. Cancer Immunol. Immunother. 2019, 68, 365–377. [Google Scholar] [CrossRef]
  100. Zhou, Y.; Wen, P.; Li, M.; Li, Y.; Li, X.-A. Construction of chimeric antigen receptor-modified T cells targeting EpCAM and assessment of their anti-tumor effect on cancer cells. Mol. Med. Rep. 2019, 20, 2355–2364. [Google Scholar] [CrossRef]
  101. Jiang, Z.; Jiang, X.; Chen, S.; Lai, Y.; Wei, X.; Li, B.; Lin, S.; Wang, S.; Wu, Q.; Liang, Q.; et al. Anti-GPC3-CAR T Cells Suppress the Growth of Tumor Cells in Patient-Derived Xenografts of Hepatocellular Carcinoma. Front. Immunol. 2016, 7, 690. [Google Scholar] [CrossRef] [PubMed]
  102. Guo, J.; Tang, Q. Recent updates on chimeric antigen receptor T cell therapy for hepatocellular carcinoma. Cancer Gene Ther. 2021, 28, 1075–1087. [Google Scholar] [CrossRef] [PubMed]
  103. Sun, B.; Yang, D.; Dai, H.; Liu, X.; Jia, R.; Cui, X.; Li, W.; Cai, C.; Xu, J.; Zhao, X. Eradication of Hepatocellular Carcinoma by NKG2D-Based CAR-T Cells. Cancer Immunol. Res. 2019, 7, 1813–1823. [Google Scholar] [CrossRef] [PubMed]
  104. Maggs, L.; Cattaneo, G.; Dal, A.E.; Moghaddam, A.S.; Ferrone, S. CAR T Cell-Based Immunotherapy for the Treatment of Glioblastoma. Front. Neurosci. 2021, 15, 662064. [Google Scholar] [CrossRef]
  105. Soler, D.C.; Kerstetter-Fogle, A.; McCormick, T.S.; Sloan, A.E. Using chimeric antigen receptor T-cell therapy to fight glioblastoma multiforme: Past, present and future developments. J. Neurooncol. 2022, 156, 81–96. [Google Scholar] [CrossRef]
  106. Chow, K.K.H.; Naik, S.; Kakarla, S.; Brawley, V.S.; Shaffer, D.R.; Yi, Z.; Rainusso, N.; Wu, M.-F.; Liu, H.; Kew, Y.; et al. T cells redirected to EphA2 for the immunotherapy of glioblastoma. Mol. Ther. 2013, 21, 629–637. [Google Scholar] [CrossRef]
  107. An, Z.; Hu, Y.; Bai, Y.; Zhang, C.; Xu, C.; Kang, X.; Yang, S.; Li, W.; Zhong, X. Antitumor activity of the third generation EphA2 CAR-T cells against glioblastoma is associated with interferon gamma induced PD-L1. Oncoimmunology 2021, 10, 1960728. [Google Scholar] [CrossRef]
  108. De Billy, E.; Pellegrino, M.; Orlando, D.; Pericoli, G.; Ferretti, R.; Businaro, P.; Ajmone-Cat, M.A.; Rossi, S.; Petrilli, L.L.; Maestro, N.; et al. Dual IGF1R/IR inhibitors in combination with GD2-CAR T-cells display a potent anti-tumor activity in diffuse midline glioma H3K27M-mutant. Neuro. Oncol. 2022, 24, 1150–1163. [Google Scholar] [CrossRef]
  109. Arnone, C.M.; Polito, V.A.; Mastronuzzi, A.; Carai, A.; Diomedi, F.C.; Antonucci, L.; Petrilli, L.L.; Vinci, M.; Ferrari, F.; Salviato, E.; et al. Oncolytic adenovirus and gene therapy with EphA2-BiTE for the treatment of pediatric high-grade gliomas. J. Immunother. Cancer 2021, 9, e001930. [Google Scholar] [CrossRef]
  110. Zhu, G.; Zhang, Q.; Zhang, J.; Liu, F. Targeting Tumor-Associated Antigen: A Promising CAR-T Therapeutic Strategy for Glioblastoma Treatment. Front. Pharmacol. 2021, 12, 661606. [Google Scholar] [CrossRef]
  111. Alizadeh, D.; Wong, R.A.; Gholamin, S.; Maker, M.; Aftabizadeh, M.; Yang, X.; Pecoraro, J.R.; Jeppson, J.D.; Wang, D.; Aguilar, B.; et al. IFNγ Is Critical for CAR T Cell-Mediated Myeloid Activation and Induction of Endogenous Immunity. Cancer Discov. 2021, 11, 2248–2265. [Google Scholar] [CrossRef]
  112. Zhang, C.; Burger, M.C.; Jennewein, L.; Genßler, S.; Schönfeld, K.; Zeiner, P.; Hattingen, E.; Harter, P.N.; Mittelbronn, M.; Tonn, T.; et al. ErbB2/HER2-Specific NK Cells for Targeted Therapy of Glioblastoma. J. Natl. Cancer Inst. 2016, 108, djv375. [Google Scholar] [CrossRef]
  113. Weimin, S.; Abula, A.; Qianghong, D.; Wenguang, W. Chimeric cytokine receptor enhancing PSMA-CAR-T cell-mediated prostate cancer regression. Cancer Biol. Ther. 2020, 21, 570–580. [Google Scholar] [CrossRef]
  114. Jiang, Y.; Wen, W.; Yang, F.; Han, D.; Zhang, W.; Qin, W. Prospect of Prostate Cancer Treatment: Armed CAR-T or Combination Therapy. Cancers 2022, 14, 967. [Google Scholar] [CrossRef]
  115. Lloyd, G.G.; Steel, J.M.; Young, R.J. Eating disorders and psychiatric morbidity in patients with diabetes mellitus. Psychother. Psychosom. 1987, 48, 189–195. [Google Scholar] [CrossRef]
  116. Junghans, R.P.; Ma, Q.; Rathore, R.; Gomes, E.M.; Bais, A.J.; Lo, A.S.Y.; Abedi, M.; Davies, R.A.; Cabral, H.J.; Al-Homsi, A.S.; et al. Phase I Trial of Anti-PSMA Designer CAR-T Cells in Prostate Cancer: Possible Role for Interacting Interleukin 2-T Cell Pharmacodynamics as a Determinant of Clinical Response. Prostate 2016, 76, 1257–1270. [Google Scholar] [CrossRef]
  117. Bagley, S.J.; O’Rourke, D.M. Clinical investigation of CAR T cells for solid tumors: Lessons learned and future directions. Pharmacol. Ther. 2020, 205, 107419. [Google Scholar] [CrossRef]
  118. Bui, M.H.T.; Seligson, D.; Han, K.; Pantuck, A.J.; Dorey, F.J.; Huang, Y.; Horvath, S.; Leibovich, B.C.; Chopra, S.; Liao, S.-Y.; et al. Carbonic anhydrase IX is an independent predictor of survival in advanced renal clear cell carcinoma: Implications for prognosis and therapy. Clin. Cancer Res. 2003, 9, 802–811. [Google Scholar]
  119. Yeku, O.; Li, X.; Brentjens, R.J. Adoptive T-Cell Therapy for Solid Tumors. Am. Soc. Clin. Oncol. Educ. book. Am. Soc. Clin. Oncol. Annu. Meet. 2017, 37, 193–204. [Google Scholar] [CrossRef]
  120. Tafreshi, N.K.; Lloyd, M.C.; Bui, M.M.; Gillies, R.J.; Morse, D.L. Carbonic anhydrase IX as an imaging and therapeutic target for tumors and metastases. Subcell. Biochem. 2014, 75, 221–254. [Google Scholar] [CrossRef]
  121. Li, J.; Li, W.; Huang, K.; Zhang, Y.; Kupfer, G.; Zhao, Q. Chimeric antigen receptor T cell (CAR-T) immunotherapy for solid tumors: Lessons learned and strategies for moving forward. J. Hematol. Oncol. 2018, 11, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Zhao, W.; Jia, L.; Zhang, M.; Huang, X.; Qian, P.; Tang, Q.; Zhu, J.; Feng, Z. The killing effect of novel bi-specific Trop2/PD-L1 CAR-T cell targeted gastric cancer. Am. J. Cancer Res. 2019, 9, 1846–1856. [Google Scholar] [PubMed]
  123. Lv, J.; Zhao, R.; Wu, D.; Zheng, D.; Wu, Z.; Shi, J.; Wei, X.; Wu, Q.; Long, Y.; Lin, S.; et al. Mesothelin is a target of chimeric antigen receptor T cells for treating gastric cancer. J. Hematol. Oncol. 2019, 12, 18. [Google Scholar] [CrossRef] [PubMed]
  124. Jiang, H.; Shi, Z.; Wang, P.; Wang, C.; Yang, L.; Du, G.; Zhang, H.; Shi, B.; Jia, J.; Li, Q.; et al. Claudin18.2-Specific Chimeric Antigen Receptor Engineered T Cells for the Treatment of Gastric Cancer. J. Natl. Cancer Inst. 2019, 111, 409–418. [Google Scholar] [CrossRef]
  125. Tao, K.; He, M.; Tao, F.; Xu, G.; Ye, M.; Zheng, Y.; Li, Y. Development of NKG2D-based chimeric antigen receptor-T cells for gastric cancer treatment. Cancer Chemother. Pharmacol. 2018, 82, 815–827. [Google Scholar] [CrossRef]
  126. Kim, M.; Pyo, S.; Kang, C.H.; Lee, C.O.; Lee, H.K.; Choi, S.U.; Park, C.H. Folate receptor 1 (FOLR1) targeted chimeric antigen receptor (CAR) T cells for the treatment of gastric cancer. PLoS ONE 2018, 13, e0198347. [Google Scholar] [CrossRef]
  127. Song, Y.; Tong, C.; Wang, Y.; Gao, Y.; Dai, H.; Guo, Y.; Zhao, X.; Wang, Y.; Wang, Z.; Han, W.; et al. Effective and persistent antitumor activity of HER2-directed CAR-T cells against gastric cancer cells in vitro and xenotransplanted tumors in vivo. Protein Cell 2018, 9, 867–878. [Google Scholar] [CrossRef]
  128. Jung, M.; Yang, Y.; McCloskey, J.E.; Zaman, M.; Vedvyas, Y.; Zhang, X.; Stefanova, D.; Gray, K.D.; Min, I.M.; Zarnegar, R.; et al. Chimeric Antigen Receptor T Cell Therapy Targeting ICAM-1 in Gastric Cancer. Mol. Ther. Oncolytics 2020, 18, 587–601. [Google Scholar] [CrossRef]
  129. Deng, X.; Gao, F.; Li, N.; Li, Q.; Zhou, Y.; Yang, T.; Cai, Z.; Du, P.; Chen, F.; Cai, J. Antitumor activity of NKG2D CAR-T cells against human colorectal cancer cells in vitro and in vivo. Am. J. Cancer Res. 2019, 9, 945–958. [Google Scholar]
  130. Wang, Y.; Chen, M.; Wu, Z.; Tong, C.; Dai, H.; Guo, Y.; Liu, Y.; Huang, J.; Lv, H.; Luo, C.; et al. CD133-directed CAR T cells for advanced metastasis malignancies: A phase I trial. Oncoimmunology 2018, 7, e1440169. [Google Scholar] [CrossRef]
  131. Hege, K.M.; Bergsland, E.K.; Fisher, G.A.; Nemunaitis, J.J.; Warren, R.S.; McArthur, J.G.; Lin, A.A.; Schlom, J.; June, C.H.; Sherwin, S.A. Safety, tumor trafficking and immunogenicity of chimeric antigen receptor (CAR)-T cells specific for TAG-72 in colorectal cancer. J. Immunother. Cancer 2017, 5, 22. [Google Scholar] [CrossRef]
  132. Hombach, A.A.; Geumann, U.; Günther, C.; Hermann, F.G.; Abken, H. IL7-IL12 Engineered Mesenchymal Stem Cells (MSCs) Improve A CAR T Cell Attack Against Colorectal Cancer Cells. Cells 2020, 9, 873. [Google Scholar] [CrossRef] [Green Version]
  133. Sureban, S.M.; Berahovich, R.; Zhou, H.; Xu, S.; Wu, L.; Ding, K.; May, R.; Qu, D.; Bannerman-Menson, E.; Golubovskaya, V.; et al. DCLK1 Monoclonal Antibody-Based CAR-T Cells as a Novel Treatment Strategy against Human Colorectal Cancers. Cancers 2019, 12, 54. [Google Scholar] [CrossRef]
  134. Wallstabe, L.; Göttlich, C.; Nelke, L.C.; Kühnemundt, J.; Schwarz, T.; Nerreter, T.; Einsele, H.; Walles, H.; Dandekar, G.; Nietzer, S.L.; et al. ROR1-CAR T cells are effective against lung and breast cancer in advanced microphysiologic 3D tumor models. JCI Insight 2019, 4, e126345. [Google Scholar] [CrossRef]
  135. Zhang, Z.; Jiang, J.; Wu, X.; Zhang, M.; Luo, D.; Zhang, R.; Li, S.; He, Y.; Bian, H.; Chen, Z. Chimeric antigen receptor T cell targeting EGFRvIII for metastatic lung cancer therapy. Front. Med. 2019, 13, 57–68. [Google Scholar] [CrossRef]
  136. Ye, L.; Lou, Y.; Lu, L.; Fan, X. Mesothelin-targeted second generation CAR-T cells inhibit growth of mesothelin-expressing tumors in vivo. Exp. Ther. Med. 2019, 17, 739–747. [Google Scholar] [CrossRef]
  137. Li, N.; Liu, S.; Sun, M.; Chen, W.; Xu, X.; Zeng, Z.; Tang, Y.; Dong, Y.; Chang, A.H.; Zhao, Q. Chimeric Antigen Receptor-Modified T Cells Redirected to EphA2 for the Immunotherapy of Non-Small Cell Lung Cancer. Transl. Oncol. 2018, 11, 11–17. [Google Scholar] [CrossRef]
  138. Wei, X.; Lai, Y.; Li, J.; Qin, L.; Xu, Y.; Zhao, R.; Li, B.; Lin, S.; Wang, S.; Wu, Q.; et al. PSCA and MUC1 in non-small-cell lung cancer as targets of chimeric antigen receptor T cells. Oncoimmunology 2017, 6, e1284722. [Google Scholar] [CrossRef]
  139. Liu, M.; Wang, X.; Li, W.; Yu, X.; Flores-Villanueva, P.; Xu-Monette, Z.Y.; Li, L.; Zhang, M.; Young, K.H.; Ma, X.; et al. Targeting PD-L1 in non-small cell lung cancer using CAR T cells. Oncogenesis 2020, 9, 72. [Google Scholar] [CrossRef]
  140. Chen, X.; Amar, N.; Zhu, Y.; Wang, C.; Xia, C.; Yang, X.; Wu, D.; Feng, M. Combined DLL3-targeted bispecific antibody with PD-1 inhibition is efficient to suppress small cell lung cancer growth. J. Immunother. Cancer 2020, 8, e000785. [Google Scholar] [CrossRef]
  141. Aharon, A.; Horn, G.; Bar-Lev, T.H.; Zagagi Yohay, E.; Waks, T.; Levin, M.; Deshet Unger, N.; Avivi, I.; Globerson Levin, A. Extracellular Vesicles Derived from Chimeric Antigen Receptor-T Cells: A Potential Therapy for Cancer. Hum. Gene Ther. 2021, 32, 1224–1241. [Google Scholar] [CrossRef] [PubMed]
  142. Yang, P.; Cao, X.; Cai, H.; Feng, P.; Chen, X.; Zhu, Y.; Yang, Y.; An, W.; Yang, Y.; Jie, J. The exosomes derived from CAR-T cell efficiently target mesothelin and reduce triple-negative breast cancer growth. Cell. Immunol. 2021, 360, 104262. [Google Scholar] [CrossRef] [PubMed]
  143. Fu, W.; Lei, C.; Liu, S.; Cui, Y.; Wang, C.; Qian, K.; Li, T.; Shen, Y.; Fan, X.; Lin, F.; et al. CAR exosomes derived from effector CAR-T cells have potent antitumour effects and low toxicity. Nat. Commun. 2019, 10, 4355. [Google Scholar] [CrossRef] [PubMed]
  144. Merchant, M.L.; Rood, I.M.; Deegens, J.K.J.; Klein, J.B. Isolation and characterization of urinary extracellular vesicles: Implications for biomarker discovery. Nat. Rev. Nephrol. 2017, 13, 731–749. [Google Scholar] [CrossRef] [PubMed]
  145. György, B.; Szabó, T.G.; Pásztói, M.; Pál, Z.; Misják, P.; Aradi, B.; László, V.; Pállinger, E.; Pap, E.; Kittel, A.; et al. Membrane vesicles, current state-of-the-art: Emerging role of extracellular vesicles. Cell. Mol. Life Sci. 2011, 68, 2667–2688. [Google Scholar] [CrossRef]
  146. Gandham, S.; Su, X.; Wood, J.; Nocera, A.L.; Alli, S.C.; Milane, L.; Zimmerman, A.; Amiji, M.; Ivanov, A.R. Technologies and Standardization in Research on Extracellular Vesicles. Trends Biotechnol. 2020, 38, 1066–1098. [Google Scholar] [CrossRef]
  147. Zhang, Q.; Jeppesen, D.K.; Higginbotham, J.N.; Graves-Deal, R.; Trinh, V.Q.; Ramirez, M.A.; Sohn, Y.; Neininger, A.C.; Taneja, N.; McKinley, E.T.; et al. Supermeres are functional extracellular nanoparticles replete with disease biomarkers and therapeutic targets. Nat. Cell Biol. 2021, 23, 1240–1254. [Google Scholar] [CrossRef]
  148. Yu, S.; Yu, L. Migrasome biogenesis and functions. FEBS J. 2022, 289, 7246–7254. [Google Scholar] [CrossRef]
  149. De Lellis, L.; Florio, R.; Di Bella, M.C.; Brocco, D.; Guidotti, F.; Tinari, N.; Grassadonia, A.; Lattanzio, R.; Cama, A.; Veschi, S. Exosomes as Pleiotropic Players in Pancreatic Cancer. Biomedicines 2021, 9, 275. [Google Scholar] [CrossRef]
  150. Clancy, J.W.; Boomgarden, A.C.; D’Souza-Schorey, C. Profiling and promise of supermeres. Nat. Cell Biol. 2021, 23, 1217–1219. [Google Scholar] [CrossRef]
  151. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  152. Colombo, M.; Raposo, G.; Théry, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell Dev. Biol. 2014, 30, 255–289. [Google Scholar] [CrossRef]
  153. Andreu, Z.; Yáñez-Mó, M. Tetraspanins in extracellular vesicle formation and function. Front. Immunol. 2014, 5, 442. [Google Scholar] [CrossRef]
  154. Van Niel, G.; Charrin, S.; Simoes, S.; Romao, M.; Rochin, L.; Saftig, P.; Marks, M.S.; Rubinstein, E.; Raposo, G. The tetraspanin CD63 regulates ESCRT-independent and -dependent endosomal sorting during melanogenesis. Dev. Cell 2011, 21, 708–721. [Google Scholar] [CrossRef]
  155. Tamai, K.; Tanaka, N.; Nakano, T.; Kakazu, E.; Kondo, Y.; Inoue, J.; Shiina, M.; Fukushima, K.; Hoshino, T.; Sano, K.; et al. Exosome secretion of dendritic cells is regulated by Hrs, an ESCRT-0 protein. Biochem. Biophys. Res. Commun. 2010, 399, 384–390. [Google Scholar] [CrossRef]
  156. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Théry, C.; Raposo, G. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126, 5553–5565. [Google Scholar] [CrossRef]
  157. Stuffers, S.; Sem Wegner, C.; Stenmark, H.; Brech, A. Multivesicular endosome biogenesis in the absence of ESCRTs. Traffic 2009, 10, 925–937. [Google Scholar] [CrossRef]
  158. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brügger, B.; Simons, M. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef]
  159. Buschow, S.I.; Nolte-’t Hoen, E.N.M.; van Niel, G.; Pols, M.S.; ten Broeke, T.; Lauwen, M.; Ossendorp, F.; Melief, C.J.M.; Raposo, G.; Wubbolts, R.; et al. MHC II in dendritic cells is targeted to lysosomes or T cell-induced exosomes via distinct multivesicular body pathways. Traffic 2009, 10, 1528–1542. [Google Scholar] [CrossRef]
  160. Venn, G.; Lauder, R.M.; Hardingham, T.E.; Muir, H. Effects of catabolic and anabolic cytokines on proteoglycan biosynthesis in young, old and osteoarthritic canine cartilage. Biochem. Soc. Trans. 1990, 18, 973–974. [Google Scholar] [CrossRef]
  161. Theos, A.C.; Truschel, S.T.; Tenza, D.; Hurbain, I.; Harper, D.C.; Berson, J.F.; Thomas, P.C.; Raposo, G.; Marks, M.S. A lumenal domain-dependent pathway for sorting to intralumenal vesicles of multivesicular endosomes involved in organelle morphogenesis. Dev. Cell 2006, 10, 343–354. [Google Scholar] [CrossRef] [PubMed]
  162. Van Niel, G.; Bergam, P.; Di Cicco, A.; Hurbain, I.; Lo Cicero, A.; Dingli, F.; Palmulli, R.; Fort, C.; Potier, M.C.; Schurgers, L.J.; et al. Apolipoprotein E Regulates Amyloid Formation within Endosomes of Pigment Cells. Cell Rep. 2015, 13, 43–51. [Google Scholar] [CrossRef] [PubMed]
  163. Gauthier, S.A.; Pérez-González, R.; Sharma, A.; Huang, F.-K.; Alldred, M.J.; Pawlik, M.; Kaur, G.; Ginsberg, S.D.; Neubert, T.A.; Levy, E. Enhanced exosome secretion in Down syndrome brain—A protective mechanism to alleviate neuronal endosomal abnormalities. Acta Neuropathol. Commun. 2017, 5, 65. [Google Scholar] [CrossRef] [PubMed]
  164. Lindenbergh, M.F.S.; Stoorvogel, W. Antigen Presentation by Extracellular Vesicles from Professional Antigen-Presenting Cells. Annu. Rev. Immunol. 2018, 36, 435–459. [Google Scholar] [CrossRef] [PubMed]
  165. Xu, X.; Lai, Y.; Hua, Z.-C. Apoptosis and apoptotic body: Disease message and therapeutic target potentials. Biosci. Rep. 2019, 39, BSR20180992. [Google Scholar] [CrossRef]
  166. Nagata, S. Apoptosis and Clearance of Apoptotic Cells. Annu. Rev. Immunol. 2018, 36, 489–517. [Google Scholar] [CrossRef]
  167. Calvo, V.; Izquierdo, M. Imaging Polarized Secretory Traffic at the Immune Synapse in Living T Lymphocytes. Front. Immunol. 2018, 9, 684. [Google Scholar] [CrossRef]
  168. Coumans, F.A.W.; Brisson, A.R.; Buzas, E.I.; Dignat-George, F.; Drees, E.E.E.; El-Andaloussi, S.; Emanueli, C.; Gasecka, A.; Hendrix, A.; Hill, A.F.; et al. Methodological Guidelines to Study Extracellular Vesicles. Circ. Res. 2017, 120, 1632–1648. [Google Scholar] [CrossRef]
  169. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef]
  170. Tauro, B.J.; Greening, D.W.; Mathias, R.A.; Mathivanan, S.; Ji, H.; Simpson, R.J. Two distinct populations of exosomes are released from LIM1863 colon carcinoma cell-derived organoids. Mol. Cell. Proteomics 2013, 12, 587–598. [Google Scholar] [CrossRef]
  171. Pallet, N.; Sirois, I.; Bell, C.; Hanafi, L.-A.; Hamelin, K.; Dieudé, M.; Rondeau, C.; Thibault, P.; Desjardins, M.; Hebert, M.-J. A comprehensive characterization of membrane vesicles released by autophagic human endothelial cells. Proteomics 2013, 13, 1108–1120. [Google Scholar] [CrossRef]
  172. Brocco, D.; Lanuti, P.; Simeone, P.; Bologna, G.; Pieragostino, D.; Cufaro, M.C.; Graziano, V.; Peri, M.; Di Marino, P.; De Tursi, M.; et al. Circulating Cancer Stem Cell-Derived Extracellular Vesicles as a Novel Biomarker for Clinical Outcome Evaluation. J. Oncol. 2019, 2019, 5879616. [Google Scholar] [CrossRef]
  173. Buca, D.; D’Antonio, F.; Buca, D.; Di Sebastiano, F.; Simeone, P.; Di Girolamo, R.; Bologna, G.; Vespa, S.; Catitti, G.; Liberati, M.; et al. Extracellular Vesicles in pregnancy: Their potential role as a liquid biopsy. J. Reprod. Immunol. 2022, 154, 103734. [Google Scholar] [CrossRef]
  174. Grande, R.; Dovizio, M.; Marcone, S.; Szklanna, P.B.; Bruno, A.; Ebhardt, H.A.; Cassidy, H.; Ní Áinle, F.; Caprodossi, A.; Lanuti, P.; et al. Platelet-Derived Microparticles From Obese Individuals: Characterization of Number, Size, Proteomics, and Crosstalk With Cancer and Endothelial Cells. Front. Pharmacol. 2019, 10, 7. [Google Scholar] [CrossRef]
  175. Lanuti, P.; Santilli, F.; Marchisio, M.; Pierdomenico, L.; Vitacolonna, E.; Santavenere, E.; Iacone, A.; Davì, G.; Romano, M.; Miscia, S. A novel flow cytometric approach to distinguish circulating endothelial cells from endothelial microparticles: Relevance for the evaluation of endothelial dysfunction. J. Immunol. Methods 2012, 380, 16–22. [Google Scholar] [CrossRef]
  176. Simeone, P.; Celia, C.; Bologna, G.; Ercolino, E.; Pierdomenico, L.; Cilurzo, F.; Grande, R.; Diomede, F.; Vespa, S.; Canonico, B.; et al. Diameters and Fluorescence Calibration for Extracellular Vesicle Analyses by Flow Cytometry. Int. J. Mol. Sci. 2020, 21, 7885. [Google Scholar] [CrossRef]
  177. Vaidyanathan, R.; Soon, R.H.; Zhang, P.; Jiang, K.; Lim, C.T. Cancer diagnosis: From tumor to liquid biopsy and beyond. Lab Chip 2018, 19, 11–34. [Google Scholar] [CrossRef]
  178. Buca, D.; Bologna, G.; D’Amico, A.; Cugini, S.; Musca, F.; Febbo, M.; D’Arcangelo, D.; Buca, D.; Simeone, P.; Liberati, M.; et al. Extracellular Vesicles in Feto-Maternal Crosstalk and Pregnancy Disorders. Int. J. Mol. Sci. 2020, 21, 2120. [Google Scholar] [CrossRef]
  179. Serafini, F.L.; Lanuti, P.; Delli Pizzi, A.; Procaccini, L.; Villani, M.; Taraschi, A.L.; Pascucci, L.; Mincuzzi, E.; Izzi, J.; Chiacchiaretta, P.; et al. Diagnostic Impact of Radiological Findings and Extracellular Vesicles: Are We Close to Radiovesicolomics? Biology 2021, 10, 1265. [Google Scholar] [CrossRef]
  180. Lucidi, A.; Buca, D.; Ronsini, C.; Tinari, S.; Bologna, G.; Buca, D.; Leombroni, M.; Liberati, M.; D’Antonio, F.; Scambia, G.; et al. Role of Extracellular Vesicles in Epithelial Ovarian Cancer: A Systematic Review. Int. J. Mol. Sci. 2020, 21, 8762. [Google Scholar] [CrossRef]
  181. Mack, M.; Kleinschmidt, A.; Brühl, H.; Klier, C.; Nelson, P.J.; Cihak, J.; Plachý, J.; Stangassinger, M.; Erfle, V.; Schlöndorff, D. Transfer of the chemokine receptor CCR5 between cells by membrane-derived microparticles: A mechanism for cellular human immunodeficiency virus 1 infection. Nat. Med. 2000, 6, 769–775. [Google Scholar] [CrossRef] [PubMed]
  182. Al-Nedawi, K.; Meehan, B.; Micallef, J.; Lhotak, V.; May, L.; Guha, A.; Rak, J. Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles derived from tumour cells. Nat. Cell Biol. 2008, 10, 619–624. [Google Scholar] [CrossRef] [PubMed]
  183. Christianson, H.C.; Svensson, K.J.; van Kuppevelt, T.H.; Li, J.-P.; Belting, M. Cancer cell exosomes depend on cell-surface heparan sulfate proteoglycans for their internalization and functional activity. Proc. Natl. Acad. Sci. USA 2013, 110, 17380–17385. [Google Scholar] [CrossRef] [PubMed]
  184. Morelli, A.E.; Larregina, A.T.; Shufesky, W.J.; Sullivan, M.L.G.; Stolz, D.B.; Papworth, G.D.; Zahorchak, A.F.; Logar, A.J.; Wang, Z.; Watkins, S.C.; et al. Endocytosis, intracellular sorting, and processing of exosomes by dendritic cells. Blood 2004, 104, 3257–3266. [Google Scholar] [CrossRef]
  185. Fitzner, D.; Schnaars, M.; van Rossum, D.; Krishnamoorthy, G.; Dibaj, P.; Bakhti, M.; Regen, T.; Hanisch, U.-K.; Simons, M. Selective transfer of exosomes from oligodendrocytes to microglia by macropinocytosis. J. Cell Sci. 2011, 124, 447–458. [Google Scholar] [CrossRef]
  186. Parolini, I.; Federici, C.; Raggi, C.; Lugini, L.; Palleschi, S.; De Milito, A.; Coscia, C.; Iessi, E.; Logozzi, M.; Molinari, A.; et al. Microenvironmental pH is a key factor for exosome traffic in tumor cells. J. Biol. Chem. 2009, 284, 34211–34222. [Google Scholar] [CrossRef]
  187. Théry, C.; Ostrowski, M.; Segura, E. Membrane vesicles as conveyors of immune responses. Nat. Rev. Immunol. 2009, 9, 581–593. [Google Scholar] [CrossRef]
  188. Tanaka, M.; Itai, T.; Adachi, M.; Nagata, S. Downregulation of Fas ligand by shedding. Nat. Med. 1998, 4, 31–36. [Google Scholar] [CrossRef]
  189. Schneider, P.; Holler, N.; Bodmer, J.L.; Hahne, M.; Frei, K.; Fontana, A.; Tschopp, J. Conversion of membrane-bound Fas(CD95) ligand to its soluble form is associated with downregulation of its proapoptotic activity and loss of liver toxicity. J. Exp. Med. 1998, 187, 1205–1213. [Google Scholar] [CrossRef]
  190. Mulcahy, L.A.; Pink, R.C.; Carter, D.R.F. Routes and mechanisms of extracellular vesicle uptake. J. Extracell. Vesicles 2014, 3, 24641. [Google Scholar] [CrossRef]
  191. Mathieu, M.; Martin-Jaular, L.; Lavieu, G.; Théry, C. Specificities of secretion and uptake of exosomes and other extracellular vesicles for cell-to-cell communication. Nat. Cell Biol. 2019, 21, 9–17. [Google Scholar] [CrossRef]
  192. Rustom, A.; Saffrich, R.; Markovic, I.; Walther, P.; Gerdes, H.-H. Nanotubular highways for intercellular organelle transport. Science 2004, 303, 1007–1010. [Google Scholar] [CrossRef] [Green Version]
  193. Oh, N.; Park, J.-H. Endocytosis and exocytosis of nanoparticles in mammalian cells. Int. J. Nanomedicine 2014, 9 (Suppl. S1), 51–63. [Google Scholar] [CrossRef]
  194. Hung, M.E.; Leonard, J.N. A platform for actively loading cargo RNA to elucidate limiting steps in EV-mediated delivery. J. Extracell. Vesicles 2016, 5, 31027. [Google Scholar] [CrossRef]
  195. O’Brien, K.; Ughetto, S.; Mahjoum, S.; Nair, A.V.; Breakefield, X.O. Uptake, functionality, and re-release of extracellular vesicle-encapsulated cargo. Cell Rep. 2022, 39, 110651. [Google Scholar] [CrossRef] [PubMed]
  196. Shukla, D.; Liu, J.; Blaiklock, P.; Shworak, N.W.; Bai, X.; Esko, J.D.; Cohen, G.H.; Eisenberg, R.J.; Rosenberg, R.D.; Spear, P.G. A novel role for 3-O-sulfated heparan sulfate in herpes simplex virus 1 entry. Cell 1999, 99, 13–22. [Google Scholar] [CrossRef]
  197. Antes, T.J.; Middleton, R.C.; Luther, K.M.; Ijichi, T.; Peck, K.A.; Liu, W.J.; Valle, J.; Echavez, A.K.; Marbán, E. Targeting extracellular vesicles to injured tissue using membrane cloaking and surface display. J. Nanobiotechnology 2018, 16, 61. [Google Scholar] [CrossRef]
  198. Hurwitz, S.N.; Meckes, D.G. Extracellular Vesicle Integrins Distinguish Unique Cancers. Proteomes 2019, 7, 14. [Google Scholar] [CrossRef] [PubMed]
  199. Tian, T.; Wang, Y.; Wang, H.; Zhu, Z.; Xiao, Z. Visualizing of the cellular uptake and intracellular trafficking of exosomes by live-cell microscopy. J. Cell. Biochem. 2010, 111, 488–496. [Google Scholar] [CrossRef]
  200. Gurung, S.; Perocheau, D.; Touramanidou, L.; Baruteau, J. The exosome journey: From biogenesis to uptake and intracellular signalling. Cell Commun. Signal. 2021, 19, 47. [Google Scholar] [CrossRef]
  201. Hazrati, A.; Soudi, S.; Malekpour, K.; Mahmoudi, M.; Rahimi, A.; Hashemi, S.M.; Varma, R.S. Immune cells-derived exosomes function as a double-edged sword: Role in disease progression and their therapeutic applications. Biomark. Res. 2022, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  202. Bekeschus, S.; Seebauer, C.; Wende, K.; Schmidt, A. Physical plasma and leukocytes—Immune or reactive? Biol. Chem. 2018, 400, 63–75. [Google Scholar] [CrossRef] [Green Version]
  203. Pugholm, L.H.; Bæk, R.; Søndergaard, E.K.L.; Revenfeld, A.L.S.; Jørgensen, M.M.; Varming, K. Phenotyping of Leukocytes and Leukocyte-Derived Extracellular Vesicles. J. Immunol. Res. 2016, 2016, 6391264. [Google Scholar] [CrossRef]
  204. Calvo, V.; Izquierdo, M. Inducible Polarized Secretion of Exosomes in T and B Lymphocytes. Int. J. Mol. Sci. 2020, 21, 2631. [Google Scholar] [CrossRef]
  205. Van der Vlist, E.J.; Arkesteijn, G.J.A.; van de Lest, C.H.A.; Stoorvogel, W.; Nolte-’t Hoen, E.N.M.; Wauben, M.H.M. CD4(+) T cell activation promotes the differential release of distinct populations of nanosized vesicles. J. Extracell. Vesicles 2012, 1, 18364. [Google Scholar] [CrossRef]
  206. Choi, D.-S.; Kim, D.-K.; Kim, Y.-K.; Gho, Y.S. Proteomics of extracellular vesicles: Exosomes and ectosomes. Mass Spectrom. Rev. 2015, 34, 474–490. [Google Scholar] [CrossRef]
  207. Ventimiglia, L.N.; Alonso, M.A. Biogenesis and Function of T Cell-Derived Exosomes. Front. Cell Dev. Biol. 2016, 4, 84. [Google Scholar] [CrossRef]
  208. Mittelbrunn, M.; Gutiérrez-Vázquez, C.; Villarroya-Beltri, C.; González, S.; Sánchez-Cabo, F.; González, M.Á.; Bernad, A.; Sánchez-Madrid, F. Unidirectional transfer of microRNA-loaded exosomes from T cells to antigen-presenting cells. Nat. Commun. 2011, 2, 282. [Google Scholar] [CrossRef]
  209. Herz, J.; Pardo, J.; Kashkar, H.; Schramm, M.; Kuzmenkina, E.; Bos, E.; Wiegmann, K.; Wallich, R.; Peters, P.J.; Herzig, S.; et al. Acid sphingomyelinase is a key regulator of cytotoxic granule secretion by primary T lymphocytes. Nat. Immunol. 2009, 10, 761–768. [Google Scholar] [CrossRef]
  210. Alonso, R.; Mazzeo, C.; Mérida, I.; Izquierdo, M. A new role of diacylglycerol kinase alpha on the secretion of lethal exosomes bearing Fas ligand during activation-induced cell death of T lymphocytes. Biochimie 2007, 89, 213–221. [Google Scholar] [CrossRef]
  211. Alonso, R.; Mazzeo, C.; Rodriguez, M.C.; Marsh, M.; Fraile-Ramos, A.; Calvo, V.; Avila-Flores, A.; Merida, I.; Izquierdo, M. Diacylglycerol kinase α regulates the formation and polarisation of mature multivesicular bodies involved in the secretion of Fas ligand-containing exosomes in T lymphocytes. Cell Death Differ. 2011, 18, 1161–1173. [Google Scholar] [CrossRef] [PubMed]
  212. Alonso, R.; Rodríguez, M.C.; Pindado, J.; Merino, E.; Mérida, I.; Izquierdo, M. Diacylglycerol kinase alpha regulates the secretion of lethal exosomes bearing Fas ligand during activation-induced cell death of T lymphocytes. J. Biol. Chem. 2005, 280, 28439–28450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Sanjuán, M.A.; Jones, D.R.; Izquierdo, M.; Mérida, I. Role of diacylglycerol kinase alpha in the attenuation of receptor signaling. J. Cell Biol. 2001, 153, 207–220. [Google Scholar] [CrossRef] [PubMed]
  214. Xie, Y.; Zhang, X.; Zhao, T.; Li, W.; Xiang, J. Natural CD8+25+ regulatory T cell-secreted exosomes capable of suppressing cytotoxic T lymphocyte-mediated immunity against B16 melanoma. Biochem. Biophys. Res. Commun. 2013, 438, 152–155. [Google Scholar] [CrossRef] [PubMed]
  215. Tung, S.L.; Boardman, D.A.; Sen, M.; Letizia, M.; Peng, Q.; Cianci, N.; Dioni, L.; Carlin, L.M.; Lechler, R.; Bollati, V.; et al. Regulatory T cell-derived extracellular vesicles modify dendritic cell function. Sci. Rep. 2018, 8, 6065. [Google Scholar] [CrossRef]
  216. Martínez-Lorenzo, M.J.; Anel, A.; Gamen, S.; Monlen, I.; Lasierra, P.; Larrad, L.; Piñeiro, A.; Alava, M.A.; Naval, J. Activated human T cells release bioactive Fas ligand and APO2 ligand in microvesicles. J. Immunol. 1999, 163, 1274–1281. [Google Scholar] [CrossRef]
  217. Krammer, P.H.; Arnold, R.; Lavrik, I.N. Life and death in peripheral T cells. Nat. Rev. Immunol. 2007, 7, 532–542. [Google Scholar] [CrossRef]
  218. Choudhuri, K.; Llodrá, J.; Roth, E.W.; Tsai, J.; Gordo, S.; Wucherpfennig, K.W.; Kam, L.C.; Stokes, D.L.; Dustin, M.L. Polarized release of T-cell-receptor-enriched microvesicles at the immunological synapse. Nature 2014, 507, 118–123. [Google Scholar] [CrossRef]
  219. Saliba, D.G.; Céspedes-Donoso, P.F.; Bálint, Š.; Compeer, E.B.; Korobchevskaya, K.; Valvo, S.; Mayya, V.; Kvalvaag, A.; Peng, Y.; Dong, T.; et al. Composition and structure of synaptic ectosomes exporting antigen receptor linked to functional CD40 ligand from helper T cells. Elife 2019, 8, e47528. [Google Scholar] [CrossRef]
  220. Peters, P.J.; Geuze, H.J.; Van der Donk, H.A.; Slot, J.W.; Griffith, J.M.; Stam, N.J.; Clevers, H.C.; Borst, J. Molecules relevant for T cell-target cell interaction are present in cytolytic granules of human T lymphocytes. Eur. J. Immunol. 1989, 19, 1469–1475. [Google Scholar] [CrossRef]
  221. Blanchard, N.; Lankar, D.; Faure, F.; Regnault, A.; Dumont, C.; Raposo, G.; Hivroz, C. TCR activation of human T cells induces the production of exosomes bearing the TCR/CD3/zeta complex. J. Immunol. 2002, 168, 3235–3241. [Google Scholar] [CrossRef]
  222. Peters, P.J.; Borst, J.; Oorschot, V.; Fukuda, M.; Krähenbühl, O.; Tschopp, J.; Slot, J.W.; Geuze, H.J. Cytotoxic T lymphocyte granules are secretory lysosomes, containing both perforin and granzymes. J. Exp. Med. 1991, 173, 1099–1109. [Google Scholar] [CrossRef]
  223. Tang, X.-J.; Sun, X.-Y.; Huang, K.-M.; Zhang, L.; Yang, Z.-S.; Zou, D.-D.; Wang, B.; Warnock, G.L.; Dai, L.-J.; Luo, J. Therapeutic potential of CAR-T cell-derived exosomes: A cell-free modality for targeted cancer therapy. Oncotarget 2015, 6, 44179–44190. [Google Scholar] [CrossRef]
  224. Johnson, L.R.; Lee, D.Y.; Eacret, J.S.; Ye, D.; June, C.H.; Minn, A.J. The immunostimulatory RNA RN7SL1 enables CAR-T cells to enhance autonomous and endogenous immune function. Cell 2021, 184, 4981–4995.e14. [Google Scholar] [CrossRef]
  225. Bossi, G.; Griffiths, G.M. Degranulation plays an essential part in regulating cell surface expression of Fas ligand in T cells and natural killer cells. Nat. Med. 1999, 5, 90–96. [Google Scholar] [CrossRef]
  226. Del Vecchio, F.; Martinez-Rodriguez, V.; Schukking, M.; Cocks, A.; Broseghini, E.; Fabbri, M. Professional killers: The role of extracellular vesicles in the reciprocal interactions between natural killer, CD8+ cytotoxic T-cells and tumour cells. J. Extracell. Vesicles 2021, 10, e12075. [Google Scholar] [CrossRef]
  227. Lugini, L.; Cecchetti, S.; Huber, V.; Luciani, F.; Macchia, G.; Spadaro, F.; Paris, L.; Abalsamo, L.; Colone, M.; Molinari, A.; et al. Immune surveillance properties of human NK cell-derived exosomes. J. Immunol. 2012, 189, 2833–2842. [Google Scholar] [CrossRef]
  228. Wu, C.-H.; Li, J.; Li, L.; Sun, J.; Fabbri, M.; Wayne, A.S.; Seeger, R.C.; Jong, A.Y. Extracellular vesicles derived from natural killer cells use multiple cytotoxic proteins and killing mechanisms to target cancer cells. J. Extracell. Vesicles 2019, 8, 1588538. [Google Scholar] [CrossRef]
  229. Wu, F.; Xie, M.; Hun, M.; She, Z.; Li, C.; Luo, S.; Chen, X.; Wan, W.; Wen, C.; Tian, J. Natural Killer Cell-Derived Extracellular Vesicles: Novel Players in Cancer Immunotherapy. Front. Immunol. 2021, 12, 658698. [Google Scholar] [CrossRef]
  230. Hu, W.; Wang, G.; Huang, D.; Sui, M.; Xu, Y. Cancer Immunotherapy Based on Natural Killer Cells: Current Progress and New Opportunities. Front. Immunol. 2019, 10, 1205. [Google Scholar] [CrossRef]
  231. Marschall, A.L.J. Targeting the Inside of Cells with Biologicals: Chemicals as a Delivery Strategy. BioDrugs 2021, 35, 643–671. [Google Scholar] [CrossRef] [PubMed]
  232. Witwer, K.W.; Wolfram, J. Extracellular vesicles versus synthetic nanoparticles for drug delivery. Nat. Rev. Mater. 2021, 6, 103–106. [Google Scholar] [CrossRef] [PubMed]
  233. Stewart, M.P.; Langer, R.; Jensen, K.F. Intracellular Delivery by Membrane Disruption: Mechanisms, Strategies, and Concepts. Chem. Rev. 2018, 118, 7409–7531. [Google Scholar] [CrossRef] [PubMed]
  234. Wiklander, O.P.B.; Nordin, J.Z.; O’Loughlin, A.; Gustafsson, Y.; Corso, G.; Mäger, I.; Vader, P.; Lee, Y.; Sork, H.; Seow, Y.; et al. Extracellular vesicle in vivo biodistribution is determined by cell source, route of administration and targeting. J. Extracell. Vesicles 2015, 4, 26316. [Google Scholar] [CrossRef]
  235. Gangadaran, P.; Li, X.J.; Lee, H.W.; Oh, J.M.; Kalimuthu, S.; Rajendran, R.L.; Son, S.H.; Baek, S.H.; Singh, T.D.; Zhu, L.; et al. A new bioluminescent reporter system to study the biodistribution of systematically injected tumor-derived bioluminescent extracellular vesicles in mice. Oncotarget 2017, 8, 109894–109914. [Google Scholar] [CrossRef]
  236. Zhang, H.; Freitas, D.; Kim, H.S.; Fabijanic, K.; Li, Z.; Chen, H.; Mark, M.T.; Molina, H.; Martin, A.B.; Bojmar, L.; et al. Identification of distinct nanoparticles and subsets of extracellular vesicles by asymmetric flow field-flow fractionation. Nat. Cell Biol. 2018, 20, 332–343. [Google Scholar] [CrossRef]
  237. Reshke, R.; Taylor, J.A.; Savard, A.; Guo, H.; Rhym, L.H.; Kowalski, P.S.; Trung, M.T.; Campbell, C.; Little, W.; Anderson, D.G.; et al. Reduction of the therapeutic dose of silencing RNA by packaging it in extracellular vesicles via a pre-microRNA backbone. Nat. Biomed. Eng. 2020, 4, 52–68. [Google Scholar] [CrossRef]
  238. Herrmann, I.K.; Wood, M.J.A.; Fuhrmann, G. Extracellular vesicles as a next-generation drug delivery platform. Nat. Nanotechnol. 2021, 16, 748–759. [Google Scholar] [CrossRef]
  239. Haney, M.J.; Klyachko, N.L.; Zhao, Y.; Gupta, R.; Plotnikova, E.G.; He, Z.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J. Control. Release 2015, 207, 18–30. [Google Scholar] [CrossRef]
  240. Malekian, F.; Shamsian, A.; Kodam, S.P.; Ullah, M. Exosome engineering for efficient and targeted drug delivery: Current status and future perspective. J. Physiol. 2022, in press. [Google Scholar] [CrossRef]
  241. Rankin-Turner, S.; Vader, P.; O’Driscoll, L.; Giebel, B.; Heaney, L.M.; Davies, O.G. A call for the standardised reporting of factors affecting the exogenous loading of extracellular vesicles with therapeutic cargos. Adv. Drug Deliv. Rev. 2021, 173, 479–491. [Google Scholar] [CrossRef]
  242. Chen, C.; Sun, M.; Wang, J.; Su, L.; Lin, J.; Yan, X. Active cargo loading into extracellular vesicles: Highlights the heterogeneous encapsulation behaviour. J. Extracell. Vesicles 2021, 10, e12163. [Google Scholar] [CrossRef]
  243. Görgens, A.; Corso, G.; Hagey, D.W.; Jawad Wiklander, R.; Gustafsson, M.O.; Felldin, U.; Lee, Y.; Bostancioglu, R.B.; Sork, H.; Liang, X.; et al. Identification of storage conditions stabilizing extracellular vesicles preparations. J. Extracell. Vesicles 2022, 11, e12238. [Google Scholar] [CrossRef]
Figure 1. Structure and evolution of CAR-T cells. (A) The CAR contains an extracellular binding domain (usually in the form of scFv), a hinge region, a transmembrane domain, and an intracellular signaling domain. (B) The first generation of CARs has only a single transduction intracellular domain such as the CD3ζ-chain, while the second generation includes the addition of one co-stimulatory molecule such as CD28, OX-40 (CD134), and 4-1BB (CD137). The third CAR generation, instead, is equipped with two different co-stimulatory molecules; while the fourth generation contains usually one costimulatory molecule and a transcription factor that is capable of inducing the production of specific cytokines (e.g., IL-12). Lastly, the fifth CAR generation has been developed based on the second CAR generation with the addition of an extra intracellular domain of cytokine receptors with a binding site for transcription factors, such as STAT-3/5. (C) Different intracellular domains of CAR-T cells.
Figure 1. Structure and evolution of CAR-T cells. (A) The CAR contains an extracellular binding domain (usually in the form of scFv), a hinge region, a transmembrane domain, and an intracellular signaling domain. (B) The first generation of CARs has only a single transduction intracellular domain such as the CD3ζ-chain, while the second generation includes the addition of one co-stimulatory molecule such as CD28, OX-40 (CD134), and 4-1BB (CD137). The third CAR generation, instead, is equipped with two different co-stimulatory molecules; while the fourth generation contains usually one costimulatory molecule and a transcription factor that is capable of inducing the production of specific cytokines (e.g., IL-12). Lastly, the fifth CAR generation has been developed based on the second CAR generation with the addition of an extra intracellular domain of cytokine receptors with a binding site for transcription factors, such as STAT-3/5. (C) Different intracellular domains of CAR-T cells.
Cancers 15 01052 g001
Figure 2. Schematic representation of the targets used to produce CAR-T cells to treat solid tumors.
Figure 2. Schematic representation of the targets used to produce CAR-T cells to treat solid tumors.
Cancers 15 01052 g002
Figure 3. Schematic representation of extracellular vesicle subtypes. Large extracellular vesicles include EV subtypes with diameters larger than 100–200 nm, such as apoptotic bodies, migrasomes, and microvescicles. Small extracellular vesicles include exosomes, exomeres, and supermeres (supernatant of exomeres), all characterized by diameters smaller than 200 nm. In some cases, large and small EV size overlap.
Figure 3. Schematic representation of extracellular vesicle subtypes. Large extracellular vesicles include EV subtypes with diameters larger than 100–200 nm, such as apoptotic bodies, migrasomes, and microvescicles. Small extracellular vesicles include exosomes, exomeres, and supermeres (supernatant of exomeres), all characterized by diameters smaller than 200 nm. In some cases, large and small EV size overlap.
Cancers 15 01052 g003
Figure 4. T-lymphocyte-derived extracellular vesicle involvement in immune responses.
Figure 4. T-lymphocyte-derived extracellular vesicle involvement in immune responses.
Cancers 15 01052 g004
Table 1. List of targets used to produce CAR-T cells to treat solid tumors.
Table 1. List of targets used to produce CAR-T cells to treat solid tumors.
Cancer TypeTargets Used for CAR-T Cell Production
Ovarian CancerTAG72 [90]; MUC16 [91]; HER2 * [91] [141]; 5T4 [92]; FRα [93]
Pancreatic CancerMSLN [95,96]; CEACAM7 [97]; CEACAM6, CD318, TSPAN8 [94]
Breast CancerMSLN * [98,99,142]; EpCAM [100]; HER2 * [141,143];
Hepatocellular CarcinomaGPC3 [101]; AFP [102]; NKG2DL [103]
GlioblastomaIL13Rα2 [110,111]; HER2 [112]; EphA2 [106,107]; EGFRvIII [105]
Prostate CancerPSMA [115,116]; PSCA [114]; EpCAM [114]
Renal CancerCA-IX [121]
Gastric CancerTrop2/PD-L1 [122]; MSLN [123]; FOLR1 [126]; claudin 18.2 [124];
HER2 [127]; NKG2D [125]; ICAM-1 [128]
Colorectal CancerCD133 [130]; TAG-72 [131]; NKG2D [129];
Guanylate Cyclase 2C [89]; CEA [132]; DCLK1 [133]
Lung CancerROR1 [134]; EGFRvIII [135]; EphA2 [137]; MSLN [136];
mucin-1 [138]; PSCA [138]; PD-L1 [139]; DLL3 [140]
* Common targets between CAR-T cells and CAR-T-derived EVs used in preclinical studies.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pagotto, S.; Simeone, P.; Brocco, D.; Catitti, G.; De Bellis, D.; Vespa, S.; Di Pietro, N.; Marinelli, L.; Di Stefano, A.; Veschi, S.; et al. CAR-T-Derived Extracellular Vesicles: A Promising Development of CAR-T Anti-Tumor Therapy. Cancers 2023, 15, 1052. https://doi.org/10.3390/cancers15041052

AMA Style

Pagotto S, Simeone P, Brocco D, Catitti G, De Bellis D, Vespa S, Di Pietro N, Marinelli L, Di Stefano A, Veschi S, et al. CAR-T-Derived Extracellular Vesicles: A Promising Development of CAR-T Anti-Tumor Therapy. Cancers. 2023; 15(4):1052. https://doi.org/10.3390/cancers15041052

Chicago/Turabian Style

Pagotto, Sara, Pasquale Simeone, Davide Brocco, Giulia Catitti, Domenico De Bellis, Simone Vespa, Natalia Di Pietro, Lisa Marinelli, Antonio Di Stefano, Serena Veschi, and et al. 2023. "CAR-T-Derived Extracellular Vesicles: A Promising Development of CAR-T Anti-Tumor Therapy" Cancers 15, no. 4: 1052. https://doi.org/10.3390/cancers15041052

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop