Next Article in Journal
Schober Test and Its Modifications Revisited—What Are We Actually Measuring? Computerized Tomography-Based Analysis
Next Article in Special Issue
Cellular and Molecular Players in the Tumor Microenvironment of Renal Cell Carcinoma
Previous Article in Journal
The Effect of Upadacitinib on Lipid Profile and Cardiovascular Events: A Meta-Analysis of Randomized Controlled Trials
Previous Article in Special Issue
Pathways of Angiogenic and Inflammatory Cytokines in Multiple Myeloma: Role in Plasma Cell Clonal Expansion and Drug Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Vesicles and Their Roles in the Tumor Immune Microenvironment

1
Myeloma Research Group, Australian Centre for Blood Diseases, Central Clinical School, Monash University—Alfred Health, Melbourne, VIC 3004, Australia
2
Malignant Haematology and Stem Cell Transplantation, Department of Haematology, Alfred Hospital, Melbourne, VIC 3004, Australia
3
Department of Clinical Hematology, Monash University, Melbourne, VIC 3004, Australia
*
Authors to whom correspondence should be addressed.
J. Clin. Med. 2022, 11(23), 6892; https://doi.org/10.3390/jcm11236892
Submission received: 16 September 2022 / Revised: 15 November 2022 / Accepted: 15 November 2022 / Published: 22 November 2022
(This article belongs to the Special Issue Immune Microenvironment in Solid and Hematological Malignancies)

Abstract

:
Tumor cells actively incorporate molecules (e.g., proteins, lipids, RNA) into particles named extracellular vesicles (EVs). Several groups have demonstrated that EVs can be transferred to target (recipient) cells, making EVs an important means of intercellular communication. Indeed, EVs are able to modulate the functions of target cells by reprogramming signaling pathways. In a cancer context, EVs promote the formation of a supportive tumor microenvironment (TME) and (pre)metastatic niches. Recent studies have revealed that immune cells, tumor cells and their secretome, including EVs, promote changes in the TME and immunosuppressive functions of immune cells (e.g., natural killer, dendritic cells, T and B cells, monocytes, macrophages) that allow tumor cells to establish and propagate. Despite the growing knowledge on EVs and on their roles in cancer and as modulators of the immune response/escape, the translation into clinical practice remains in its early stages, hence requiring improved translational research in the EVs field. Here, we comprehensively review the current knowledge and most recent research on the roles of EVs in tumor immune evasion and immunosuppression in both solid tumors and hematological malignancies. We also highlight the clinical utility of EV-mediated immunosuppression targeting and EV-engineering. Importantly, we discuss the controversial role of EVs in cancer biology, current limitations and future perspectives to further the EV knowledge into clinical practice.

1. Introduction

1.1. Extracellular Vesicles

Extracellular vesicles (EVs) are particles delimited by a lipid bilayer, typically described as small EVs (e.g., exosomes, ~100 nm in diameter) and medium-large EVs (shed microvesicles, 200 nm to ~1300 nm) [1,2]. EVs are released from all cell types into the extracellular space and play important roles in both physiological and pathological processes including cancer [1,3,4,5,6,7].
Cells actively incorporate factors (e.g., proteins, lipids, RNA) into EVs which can be transferred to target (recipient) cells, making EVs an important means of intercellular communication [8,9,10,11]. Indeed, EVs are able to modulate the function of recipient cells by reprogramming signaling pathways. In cancer, EVs are able to promote the formation of a supportive tumor microenvironment (TME) and (pre)metastatic niches [8,9,12,13,14,15,16]. Importantly, EVs protect their molecular content (cargo) from degradation in the extracellular space and can be collected (isolated, enriched and purified) for downstream applications/analyses from biofluids such as blood and urine, making them ideal candidates for liquid biopsy [7,17,18]. In addition, EVs express common surface markers with their cell of origin, designating them as ideal candidates not only for biomarker discovery and diagnostics but also for therapeutics in cancer [7,8,17].

1.2. The Tumor Microenvironment

The existence of dynamic interactions of cancer cells with the (tumor) microenvironment is well established. The TME consists of cellular components including fibroblasts, endothelial cells, immune cells (e.g., macrophages, T and B lymphocytes, dendritic cells-DC), non-cellular components including soluble factors (e.g., cytokines, growth factors, chemokines), extracellular matrix (ECM) proteins (laminin, fibronectin and collagen) and EVs. The TME is essential for stimulation of the heterogeneity of cancer cell, clonal evolution and to promote multidrug resistance ultimately favoring cancer progression and metastasis formation. “In the specialized TME, cells are protected from apoptotic stimuli and may, therefore, actively promote cancer progression as well as treatment failure” [19,20,21,22,23,24].

1.3. EVs and the Tumor Immune Microenvironment

Accumulating evidence has confirmed immune evasion to be one of the hallmarks of tumor progression both in solid and blood cancers [25,26]. Recent studies have established immunosuppressive activities of immune cells that allow tumor cells to establish, proliferate and disseminate [27]. Tumor cells are capable of evading immune cells by releasing immunosuppressive molecules (e.g., interleukin IL-10, transforming growth factor β—TGF-β) or by loss of adhesion molecules such as ICAM-1. Tumor cells are able to develop resistance to apoptosis by overexpressing BCL-2 and other anti-apoptosis molecules, or immune inhibitory receptor ligands such as programmed death ligand 1 (PD-L1), Fas ligand and tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL) [28]. Furthermore, tumor antigens are heterogeneous and have high mutation rates making immune cells unable to recognize and kill tumor cells. Cancer immunotherapy, by enhancing the immune system function in cancer patients to effectively kill tumor cells, has recently emerged as a novel successful therapeutic strategy [29,30,31,32,33].
The ability of cancer cells to use EVs to establish an ideal TME for disease progression [34,35,36,37,38,39,40] has generated particular interest. We have recently demonstrated, in our model of multiple myeloma, a hematological malignancy, that in response to their interaction with tumor-derived EVs (TEVs), human naïve stromal cells acquire pro-proliferative and pro-migratory phenotypes, with subsequent increase in tumor cells’ adhesion to pre-treated stromal cells [18]. We also demonstrated, for the first time in myeloma, that the depletion of albumin in EVs derived from blood plasma can improve protein detection by mass spectrometry. Importantly, by using this optimized proteomic approach, we demonstrated that TEVs collected from plasma of patients with myeloma are enriched in proteins associated with cell migration and cell adhesion among other biological processes [6,18]. Unpublished data indicate that TEVs collected from plasma of patients with myeloma are also enriched in factors which may induce immune suppression and tumor-promoting inflammation. This is consistent with a recent work by Lyden et al. [7] describing the proteome profile of TEVs derived from solid cancers. The authors describe molecules specific to TEVs, such as S100A4, S100A13, BSG, LGALS9, which may induce immune suppression and tumor-promoting inflammation. Lyden et al. [8,9] have previously demonstrated that TEVs are transferred to recipient cells at distant sites and by releasing their content TEVs can generate pro-inflammatory (pre)metastatic niches that support future metastasis.
TEVs have been shown to induce immunosuppression in several other cancer models [41,42,43,44,45]. Buzas [12] has recently provided a broad overview on the roles of EVs in physio-pathological mechanisms of innate and adaptive immunity, including inflammation, antigen presentation and the development and activation of B cells and T cells. The author highlights important progress related to several pathological conditions (e.g., in antimicrobial defence and in allergic, autoimmune responses) with a brief summary on the role of EVs in anti-tumor immune responses.
In this review, we discuss the current knowledge derived from analyses of these particles and their roles in tumor immunosuppression. We also discuss the clinical utility of EVs in immunotherapy and the limitations of their applicability in clinical settings. Methodological difficulties and inconsistencies in rigor and reporting have hampered the translation of research findings into clinical practice. Recently, significant progress has been made in techniques for the isolation, enrichment and characterization of EVs, but further improvement is required, particularly when attributing specific functions to EVs versus other factors present in the extracellular space that are often co-isolated. Data described in this review confirm the role of EVs as key modulators of immune functions in the TME, hence a better understanding of cellular and molecular mechanisms governing the EV-mediated interactions may provide important insights for developing new strategies to disrupt the cell-to-cell communication within the TME and inform new therapeutic strategies to fight solid and blood cancers.

2. EVs and Tumor Immune Suppression

Recent exiting observations have suggested that EVs derived from tumor cells express PDL1, a natural ligand for programmed death 1 (PD1) [46]. Tumor-derived PDL1 can suppress the functions of immune cells by directly binding with PD1 on their surface, hence promoting tumor immune escape [47,48,49]. Blocking the interaction between PD1 and PDL1 has been of revolutionary significance in cancer therapy, resulting in approvals for PD1/PDL1 blocking agents from the US Food and Drug Administration (FDA) for the treatment of several cancers [50]. A recent review [51] has extensively described the role of PDL1 in EV-mediated tumor immune evasion. As PDL1-positive EVs can impair immune activities and support tumor growth in a similar way as tumor-derived PDL1, targeting EV-PDL1 may augment the anti-tumor memory response and overcome resistance to PD1 blockade. In addition, EV-PDL1 has been shown to serve as early indicator of clinical benefit in melanoma [52]. As biomarker, EV-PDL1 may allow for rational implementation of therapies to minimize toxicity and maximize clinical benefit.
In this section, we will describe TEV interactions with immune cells which activate downstream signaling resulting in molecular and functional alterations of the recipient cell. These interactions have a profound impact on cancer development, progression, metastasis and emergence of drug resistance. Delineating the cell types, molecules and functions present in the TME will support the development of more effective treatments. A summary of these mechanisms can be found in Table 1 and Figure 1.

2.1. EVs and Tumor Immune Suppression: Role of T and B Cells

As recently described in a review by Vergani et al. [43] “tumor cells are able to escape recognition by T cells by modifying their immunogenic profile, for example by antigen loss or accumulation of mutations in genes involved in immune recognition” [118]. Tumor cells can acquire the expression of molecules with ‘don’t eat me’ function (i.e., CD47), leading to inhibition of their phagocytosis by macrophages [119]. Tumor cells can acquire immunosuppressive factors to directly kill T cells (i.e., FasL, TRAIL) and can overexpress immune checkpoint molecules (ICP), e.g., PD-L1, thereby affecting the functions of activated CD8+ T cells and promoting T cell exhaustion [120]. Immune suppressive molecules enriched in TEVs can be released into the circulation exerting immune suppressive functions also at distant (metastatic) sites from the primary tumor site [43]. In a study by Dou X et al. [53], small cell lung cancer (SCLC) cells expressing high levels of PD-L1 released TEVs overexpressing PD-L1. The latter induced downregulation of the T-cell activation marker CD69 on CD8+ T cells and also modulated the release of various cytokines from stimulated cells. Importantly, the inhibition of CD8+ T cell activation mediated by EVs was significantly reversed by an anti-PDL1 blocking antibody. The authors also performed a retrospective study showing that TEV-PDL1 was an independent prognostic factor and significantly correlated with progression-free survival in SCLC patients. These results corroborate findings by Theodoraki et al. [121], where EV-PDL1 can serve as a diagnostic biomarker for predicting the effectiveness of therapy, as well as a new strategy to enhance T-cell-mediated immunotherapy against SCLC cancers. TEV-PDL1 has also been described in melanoma, glioblastoma, prostate cancer, NSCLC, head and neck cancers, breast and gastric cancers [47,59,60,61,62,64].
In addition to T cells suppression mediated by EV-PDL1, other mechanisms have been described [65,67,68,69,70,72]. For instance, Czystowska-Kuzmicz et al. [48] have confirmed that ovarian cancer cells use TEVs to deliver the metabolic checkpoint factor ARG1 to DCs in LNs, hence inhibiting antigen-specific T cell proliferation. The authors observed that overexpression of ARG1 in mouse ovarian cancer cells is associated with accelerated tumor progression and that the latter can be inhibited by an arginase inhibitor. Gupta P et al. [54] describe TEVs as contributors of ovarian cancer progression mediated by sphingosine kinase-1 (SPHK1). In this previously unrecognized mechanism, TEV-SPHK1 upregulate S1P levels in the TME, where S1P induces T cell exhaustion and is also able to upregulate PD-L1 expression through E2F1-mediated transcription. Importantly, the SPHK1 inhibitor PF543 improved T cell-mediated cytotoxicity. The authors also show positive effects when combining PF543 with an anti-PD-1 antibody on tumor burden and metastasis vs. PF543 alone in vivo. Hence, improved immune checkpoint inhibition by targeting SPHK1/S1P signaling has potential clinical application in ovarian cancer. Xie F et al. [55] observed that breast cancer cells secrete TEV enriched in active TGF-β type II receptor (TβRII) which transfer to CD8+ T cells induced the activation of transforming growth factor-beta signaling protein 3 (SMAD3). SMAD3 cooperates with transcription factor TCF1 to promote CD8+ T cell exhaustion. This results in failure of immunotherapy. These findings identify a new possible targetable mechanism by which breast cancer cells induce TEV-mediated T cell exhaustion and negatively affect anti-tumor immunity.
Vignard et al. [56] found that miRNAs enriched in melanoma-EVs (e.g., miR-122, miR149, miR-3187-3p, miR-181a/b, miR-498) downregulate T-cell responses and cytotoxic activity by decreasing T-cell receptor (TCR) signaling and granzyme B and cytokine secretions. These observations suggest that miRNAs in melanoma-derived EVs promote tumor immune evasion and may represent a therapeutic target. Leukemic cells derived from patients with acute myeloid leukemia (AML) carrying nucleo-phosmin (NPM1) secrete EVs enriched in miR-19a-3p. The latter targets SLC6A8-mediated creatine import and promotes immunosuppressive activities of CD8+ T cells. These observations indicate that TEV-miR-19a-3p might be a promising therapeutic target for AML carrying NPM1 mutations [57].
Both the significance of humoral immune responses and the processes through which tumor associated antigens (TAAs) are recognized by the immune system and trigger a humoral response remain undefined to date. As evidenced by the production of auto-antibodies against TAAs, B cell-associated auto-immune responses are found in several tumor types. It is suggested that tumor cells express certain TAAs in their TEVs in order to divert the humoral immune response away from the tumor, thereby preventing its effective detection and elimination [74,122]. Capello et al. [74] observed that pancreatic ductal adenocarcinoma (PDAC)-EVs expose several TAAs and bind auto-antibodies present in PDAC patient blood, exerting a broad decoy-like function. Hence, complement-dependent cytotoxicity and potentially antibody-dependent cell-mediated cytotoxicity are inhibited. Similarly, it has been shown that “TEV mediate the sequestration of anti-CD20 [123] and anti-Her2 [124] therapeutic antibodies impairing antibody-based cytotoxicity directed at cancer cells and promoting resistance to cell surface directed therapeutic antibodies. These observations might explain the limited success of EV-based cancer immunotherapies and provide new insights into TEV-mediated immune escape processes with further investigation warranted”.

2.2. EVs and Tumor Immune Suppression: Role of Tregs

Tumor sites and draining LNs are also embedded with regulatory T cells (Tregs) [125], specialized immune suppressive cellular subsets, which compete with antigen specific CD8+ T cells [125,126]. Increased Treg/Tconv (conventional T cells) and Treg/CD8+ T cell ratios at the tumor site have often been described and are associated with poor prognosis in several cancers [127]. As the depletion of Treg cells significantly reduces tumor burden [128,129,130], strategies targeting receptors (CCR4, CD25, CTLA-4, CCR8) preferentially expressed on tumor infiltrating Tregs [125,128,131] are currently being evaluated in clinical trials.
TEV have been shown to induce the suppression of anti-tumor immune responses by also modulating the actions of Treg [58,70,75,116,132,133,134]. Indeed, TEV containing TGF-β1 induce Treg cells [134]. Ning et al. [135] demonstrated that TEVs derived from colorectal cancer (CRC) cells containing miR-208b promoted Treg expansion by targeting programmed cell death factor 4 (PDCD4). Treg expansion mediated by TEV-miR-208b resulted in tumor growth and oxaliplatin resistance. These findings highlight a potential role of TEV-miR-208b as a novel target for immunotherapy. CD73 + γδT1 cells have been found to be the predominant Treg population in breast cancer [71], with their prevalence in circulation found to be related to tumor burden. Breast cancer-derived TEV could transmit long non-coding (lnc)RNA SNHG16 to Vδ1 T cells. SNHG16 served as a competing endogenous RNA by sponging miR-16–5p, which led to the de-repression of its target gene SMAD5 and resulted in potentiation of the TGF-β1/SMAD5 pathway to upregulate CD73 expression in Vδ1 T cells. These observations clarify the significance of CD73 + Vδ1 Tregs in breast cancer and identify a new potential target (Treg subpopulation or TEV) with potential benefit for patients with breast cancer. TEV isolated from supernatants of head and neck squamous (HNSCC) cancer cells and ovarian cancer cells but not normal cells induced the generation and enhanced expansion of human Treg [136]. When Treg were co-incubated with TEVs, overexpression of FasL, IL-10, TGF-β1, CTLA-4, granzyme B and perforin was observed as well as inhibition of T cell proliferation. Purified Treg were resistant to TEV-mediated apoptosis relative to other T cells. The ability of TEV to promote Treg expansion was counteracted by neutralizing antibodies specific for TGF-β1 and/or IL-10. In line with these observations, other researchers have shown that EVs from the plasma of patients with advanced stages of HNSCCs exhibit immune suppression functions through the synthesis of adenosine, expansion of Treg and induction of immunosuppressive phenotype of CD8+ T cells mediated by galectin-1 [59,137] and also by inducing apoptosis of CD8+ T cells and suppression of CD4+ T cell proliferation [46]. These EVs were shown to carry immunosuppressive proteins, including PD-1, PD-L1, Fas, FasL, CTLA-4, TRAIL, CD73, COX2, TGFβ-LAP [46]. These data can explain the high speed of HNSCC cancer formation and its high rate of cancer recurrence and inform TEVs as potential therapeutic targets to prevent the dysfunction of T cells and promote anti-tumor immune responses.
TEV from nasopharyngeal carcinoma were shown to transfer miRNA (miR-24-3p; miR-891a; miR-106a-5p; miR-20a-5p) to promote downregulation of ERK/STAT1/STAT3 phosphorylation with a shift of T cells towards Treg phenotypes [138].
TEV interacting with immune cells create an immunoinhibitory microenvironment favoring immune escape also in blood cancers. Chronic and acute myeloid leukemia evade immune system surveillance and induce immunosuppression by expanding pro-leukemic Foxp3+ Tregs. High levels of immunosuppressive Tregs predict inferior response to chemotherapy, leukemia relapse and shorter survival [139]. In a recent study Swatler et al. [140] identified leukemic TEVs as drivers of effector pro-leukemic Tregs which expressed CD39, CCR8, CD30, TNFR2, CCR4, TIGIT, IL21R. Furthermore, the authors observed that leukemic EVs shuttled co-stimulatory ligand 4-1BBL/CD137L regulating expression of receptors such as CD30 and TNFR2 and ultimately promoting suppressive functions and effector phenotype of Tregs. The authors describe a Rab27a-dependent secretion of EVs and suggest that targeting of Rab27a-dependent secretion of leukemic EVs may be a viable therapeutic approach in myeloid leukemia.

2.3. EVs and Tumor Immune Suppression: Role of Natural Killer Cells

As described in a review by Carlsten and Järås [141] natural killer (NK) cells, large granular lymphocytes, are involved in immune defense against cells infected by viruses and tumor cells. The anti-tumor actions of NK cells are based on signals from activating and inhibitory cell surface receptors. NK cells are also able to lyse target cells via antibody-dependent cellular cytotoxicity which represents a critical mode of action of several therapeutic anti-cancer antibodies [142,143]. Cancer-associated mechanisms often restrain the proper activities of NK cells, leading to inadequate disease control and cancer progression. NKG2D, the classic activating NK receptor, render tumor cells susceptible to NK cell-mediated cytolysis when expressed at high levels. However, cancer cells are able to evade NKG2D-mediated immunosurveillance by shedding NKG2D ligand (NKG2DL), hence with reduced expression levels of NKG2D [144].
As shown by several groups [145,146,147], soluble NKG2DL (e.g., MICA, MICB, ULBP1-2) and TEV expressing NKG2D-L trigger the reduction of NK cell surface NKG2D.
Whiteside et al. [148] have demonstrated for the first time in 2011 inhibition of NK cell-mediated lysis of leukemic K-562 cells by TEVs isolated from patient serum. They further demonstrated that circulating TEVs from AML patients targeted purified healthy donor-derived NK cells directly resulting in down-regulation of NKG2D. TEV were shown to carry TGF-β1 which antibody-mediated neutralization inhibited suppressive activities of AML-TEVs, confirming that TEV-TGF-β1 promoted NK-cell dysfunction [149,150,151]. These observations suggest that persistently elevated circulating levels of biologically functional AML-TEVs impair immune responses and contribute to leukemia progression. AML-TEVs were also shown to bind to therapeutic NK-92 cells with activation of intracellular signaling leading to a lack of therapeutic efficiency during adoptive cell therapy [149]. Furthermore, chemotherapy significantly increase the secretion of AML-TEVs, thereby potentially contributing to therapy resistance as confirmed by other groups [152,153].
TEVs carrying miR-23a have been shown to reduce expression of lysosomal-associated membrane protein 1 (CD107a/LAMP1), an NK cell activation marker, in leukemia and lung adenocarcinoma. TGF-β and miR-23a enriched in hypoxia-induced TEV induce NK cell suppression by downregulating CD107a and NKG2D in vitro [154,155]. The immunosuppressive role of EVs is also described in solid tumors such as PDAC, lung cancer and hepatocellular carcinoma [81,82,156], with a better understanding of metastasis formation, cancer progression and drug resistance mediated by NK cell dysfunction.

2.4. EVs and Tumor Immune Suppression: Role of MDSCs

Myeloid-derived suppressor cells (MDSCs) are a heterogeneous population of immature myeloid cells with suppressive functions, containing myeloid progenitor cells and granulocyte precursors, macrophages and DCs. Elevated MDSCs levels in the blood of cancer patients are associated with inhibited T cell proliferation and reduced IFNγ production. The proportion of MDSCs in blood correlates with tumor burden and is inversely associated with immune checkpoint inhibitors (ICIs) response [90,91,116,132,133,157,158]. TEV have been shown to activate MDSCs [79] and to upregulate PD-L1, Cox2, IL-6, VEGF, arginase-1, NF-kB, TLR4 expression on MDSCs [90,91]. Furthermore, TEV-miRNA are able to alter myeloid cells to myeloid-derived suppressor cells or tumor-associated M2 macrophages (TAMs) and promote the malignant behaviour of cancers [76,77,78,80,89,159]. Melanoma-derived EVs, enriched in HSP86, induced the generation of PD-L1 + CD11b + Gr1 + MDSCs that suppressed T cell functions [90]. Another recent work suggests the presence of hepatoma-derived growth factor (HDGF) in the culture medium of human multiple myeloma (MM) cell lines, cell lysates and TEVs. The authors demonstrated that HDGF acts in an autocrine fashion activating the AKT pathway in MM cells, hence maintaining proliferation and playing a significant role in metabolism. The authors also describe a potential paracrine role where HDGF affects cells in the TME including macrophages and immature monocytes. HDGF induces macrophages’ polarization to an M1-like phenotype and alters naïve CD14+ monocytes to functionally suppressive MDSCs [116].

2.5. EVs and Tumor Immune Suppression: Role of Antigen Presenting Cells

DCs, the major type of antigen presenting cells (APCs), are involved in both innate and adaptive immunity. Their main function is to process antigens and present them to T cells initiating the adaptive immune response. Immature DCs (iDC) express high levels of chemokine receptors (CCRs) and low levels of co-stimulatory factors, capture and process antigens and migrate to LNs where they acquire antigen presenting capabilities. Here, DCs secrete cytokines and express high levels of co-stimulatory factors (e.g., CD80, CD86). This differentiation process is critical for DCs to become APCs and activate adaptive immune responses. In cancer pathophysiology, DCs orchestrate the immune reactions in the tumor and TEVs have been shown to promote immune dysfunction in DCs both in solid and blood cancer models [88,160,161].
Binding of galectin 9 (LGALS9) expressed on EVs derived from glioblastoma multiforme (GBM) cells to the TIM3 receptor of DCs affects antigen recognition, processing and presentation, leading to inhibition of cytotoxic T cell-mediated anti-tumor responses [85]. In a recent study, de Paula Silva et al. [86] observed significant effects on DCs function when DCs were treated with TEVs enriched from squamous cell carcinoma (SCC). The authors describe reduced inflammatory response and expression of lipopolysaccharide (LPS) response in addition to reduced chemokine and chemokine-mediated signaling and cytokine-cytokine receptor interaction. The authors identified EV-miRNAs as mediators of DC dysfunction (by targeting genes such as tumor necrosis factor-α—TNF-α, IL-12, IL-14), including inhibition of co-stimulatory molecules, antigen processing and presentation, DC maturation, migration, reduced production of inflammatory cytokines and increased DC death. In PDAC, EVs carrying miR-203 promote tumorigenesis by inhibition of the functions of DCs. EV-miR203 downregulates Toll-like receptor 4 (TLR4), IL-12 and TNF-α expression, hence preventing antigen presentation by DCs [84]. Similarly, EVs enriched from pancreatic cancer cells contain miR-212-3p which, transferred by DCs, downregulates the expression of regulatory factor X-associated protein (RFXAP). The latter decreases MHC class II expression with associated inactivation of CD4+ T cells, contributing to the generation of an immunotolerant TME in PDAC [83]. Targeting EV-miRNAs may represent a new strategy to improve treatment outcomes in patients with cancer.
Salimu et al. [87] have demonstrated that EVs containing the lipid mediator prostaglandin E2 (PGE2) derived from prostate cancer cells induced DCs to express the immunosuppressive marker CD73 on their surface. This significantly impacts on cytokine production and T cell activation. DCs expressing both CD73 and CD39 are able to hydrolyze ATP to adenosine in the TME. Adenosine has direct tumor-promoting, angiogenic and metastasis-inducing effects and is also a powerful inhibitor of anti-tumor immune effector cells in the TME. It also impairs maturation and function of DCs. “Although TEV have been reported to generate anti-tumor immune responses in several murine tumor models, the results from Salimu et al. strongly support the notion that, even if antigen is delivered by TEV to DC, TEV immunosuppressive properties override the potential antigen-delivery function. Indeed, a clinical trial with TEV as a cancer vaccine in patients with colorectal cancer [162] found no clinical benefit, reflecting the immunosuppressive effects of TEV”.

2.6. EVs and Tumor Immune Suppression: Role of Monocytes and Macrophages

Monocytes, an innate immune cell population, are able to either suppress or promote anti-tumor immunity depending on the context. Monocytes can differentiate into TAMs and DCs, promote angiogenesis, remodel the extracellular matrix, kill tumor cells and recruit lymphocytes.
Macrophages are the most abundant immune cells in TME and can polarize from M1 to M2 cells (TAMs) depending on physiological or pathological conditions. “While M1 macrophages (induced by T helper type 1-like cytokines such as IFN-γ) produce pro-inflammatory cytokines, chemokines and reactive nitrogen/oxygen intermediates and are involved in antimicrobial and tumoricidal activity, M2 TAMs (induced by IL-4 and IL-13) show anti-inflammatory and tumor-promoting activities. TAM represent the majority of macrophages present in the TME and a high number of TAM in various solid tumors is associated with poor prognosis and metastasis formation” [163,164].
The transition from conventional pro-inflammatory M1 to immunosuppressive M2 TAMs is a highly dynamic process regulated by several cellular signaling pathways in the TME. Accumulating evidence has established that TEVs are involved in this dynamic polarization form M1 to a M2 phenotype in cancer [80,96,98,101,103,105,108,109,164]. For instance, RNA sequencing of TEVs derived from chronic lymphocytic leukemia (CLL) revealed a high enrichment of non-coding Y RNA hY4 which transfer increased the release of cytokines [e.g., C-C motif chemokine ligand 2 (CCL2), CCL4, IL-6] and the expression of PDL1 expression in circulating monocytes via TLR7 signaling [49]. Pharmacologic inhibition of endosomal TLR signaling resulted in a substantially reduced activation of monocytes in vitro and attenuated CLL development in vivo.
Huber et al. [89] reported that melanoma-derived EVs enriched in miRNA (miR-155, miR-125b, miR-100, miR-146a, miR-146b, let-7e, miR-125a and miR-99b) conferred immunosuppressive properties to monocytes derived from healthy donors. The authors also observed that baseline levels of circulating EV-miRNA were predictive of ICIs resistance. TEVs from glioblastoma-derived stem cells (GSC) have also been shown to induce an immunosuppressive phenotype by promoting monocyte differentiation into M2 TAMs and enhancing the levels of PD-L1 [114]. Similarly, TEVs from gastric cancer-mediated monocyte differentiation into PD1 + TAMs with M2 properties in vitro and in vivo [115]. Notably, PD1 + TAM correlated with poor prognosis in gastric cancer and suppressed CD8+ T cell functions. In pancreatic cancer patients, immunosuppressive circulating CD14 + HLA-DRlo/neg monocytes were increased when compared to healthy controls. Importantly, TEV mediated the down-regulation of HLA-DR in monocytes and induced arginase expression and ROS [112]. The overexpression of the EMT transcriptional factor Snail in HNSCC cells enhanced the production of TEV-miR-21 which, transferred to monocytes, induced the downregulation of M1 markers and the increase of M2 markers [113]. Knockdown of miR-21 in HNSCC cells decreased tumor growth, M2 infiltration and angiogenesis in an in vivo model. Interestingly, a high expression of miR-21 correlated with increased SNAI1 and with M2 polarization in HNSCC patient samples.
Popēna et al. [104] have shown that macrophages exposed to colorectal cancer cell-derived TEVs displayed increased CXCL10 secretion and levels of the surface marker CD14, while monocytes displayed increased CXCL10, TNF-α and IL-1β secretion. Noteworthy, monocytes from patients with advanced cancer secreted significantly more TNFα than monocytes from patients at an early stage of the disease. Furthermore, an increase in the frequency of CD14 + CD169 + macrophages is associated with the development and progression of colorectal cancer and a higher level of both macrophages’ phenotypes (TAMs and M1). Elevated levels of CXCL10 in the serum of patients with colorectal cancer correlated with liver metastasis and poor survival. In this model TEV promote the polarization of inactive (M0) macrophages to both M1 and TAM with secretion of CXCL10, IL-6, IL-23 (M1) or IL-10 (TAM), supporting the notion that increased levels of circulating CXCL10 and IL-6 are associated with advanced stage of disease and that IL-6 is an independent prognostic marker of poor survival. The authors have also analysed the internalization of TEV by monocytes and macrophages and suggest that the endocytosis of TEVs occurs via phagocytosis and endocytic pathways dependent on dynamin. By elucidating such mechanisms, new therapeutic avenues may be informed.
Lung cancer-derived TEVs can deliver miR-21 and miR-29a to macrophage activating TLR8 and the NFKB pathway and leading to the secretion of IL6 and TFNA with associated pro-tumor inflammatory environment [110]. Similarly, miR-21 from neuroblastoma-TEV could again trigger TLR8 in monocytes, which led to upregulation of miR-155 in those cells. miR-155 could then be passaged back to the tumor cells via EVs downregulating the miR-155 target telomeric repeat binding factor 1 (TERF1), a telomerase inhibitor, with subsequent upregulation of telomerase which promotes neuroblastoma drug resistance [117]. Gerloff et al. [97] showed for the first time that miR-125b-5p transferred by cutaneous melanoma-derived EVs induces a tumor-promoting TAM phenotype in macrophages.
Recently, Wang & Gao [100] have shown that pancreatic cancer cell-derived EVs transferred miR-155-5p to macrophages and then promoted polarization to TAM phenotype. The authors describe a previously unrecognized tumor immune evasion-promoting function of TEV-miR-155-5p which suppressed EHF resulting in activation of the Akt/NF-κB signaling. This study suggested that the miR-155-5p/EHF/Akt/NF-κB axis can be exploited to prevent cancer immune evasion triggered by therapies.
Hypoxia, an intrinsic property of several cancers, induces increased secretion of EVs often enriched in immunosuppressive proteins and miRNA. Indeed, TEV-mediated polarization of macrophages to TAMs in a PDL1/HIF-1α-dependent manner in a non-small-cell lung cancer model was shown [102]. Hypoxic TEVs isolated from melanoma cells promoted TAM polarization in infiltrating macrophages in a syngeneic mouse model of macrophage infiltration [95]. In addition, miR-103a and Let-7a, enriched in TEVs at higher levels under hypoxic conditions, have been shown to promote TAM polarization in lung, skin and melanoma cancers [93,95]. Epithelial ovarian cancer (EOC) cells secrete TEVs that are able to promote macrophage polarization under hypoxic conditions by transferring miRNAs (miR-21-3p, miR-125b-5p, miR-181d-5p), hence promoting tumor proliferation and metastasis [111].
Moreover, there is evidence that TEVs transfer growth factor receptors to leukocytes impairing the antiviral immunity of patients with cancer. Lung cancer-derived TEVs can transfer activated EGFR molecules to macrophages activating MEK kinase 2 (MEKK2), which negatively regulates the antiviral immune response. This mechanism may explain the immunocompromised status of cancer patients [94].

3. Immune Cells-Derived EVs

EVs produced by immune cells can participate in tumor immune responses [44]. EVs derived by T cells have been shown to be involved in invasion and metastasis processes of several cancers. For example, EVs isolated from activated CD8+ T cells express Fas and promote melanoma and lung cancer cells invasion via the Fas/FasL pathway [165].
Plasma of patients with HNSCC contained a high number of CD3+ T cell-derived EVs vs. healthy donors. These EVs showed high expression of CD15s, expressed by Treg, suggesting that the EVs increase may be correlated to the expansion of this immune suppressive cell subtype [166].
In a recent work, Chiou et al. [167] have shown that EVs derived from activated T cells are enriched in small RNA fragments (tRNA, tRF) that inhibit the activation of T cells. Hence, T cells may selectively secrete EV-tRF that can inhibit their activation. EVs enriched in tRFs may be found in circulation and may represent biomarkers of ongoing immune activation during anti-tumor immune responses.
NK EVs have been shown to be enriched in the tumor suppressor miR-186, which was downregulated in MYCN-amplified neuroblastoma cell lines and in tumor lesions from high-risk patients vs. low-risk patients [168]. Fenselau & Ostrand-Rosenberg [169] observed that MDSC-EVs expressed molecules such as S100A8 and S100A9 and displayed tumor-promoting activities. In a breast cancer mouse model, MDSCs derived from the primary tumor area were shown to secrete more EVs than those from bone marrow or spleen [170]. MDSC-EVs inhibited cytotoxic T cell and M1 macrophage functions, thus amplifying the actions of MDSCs in the TME. These findings have also been observed in an in vitro model of colorectal cancer [171]. Exhausted T cells isolated from lesions surgically removed from HCC patients can secrete EVs, as demonstrated by Wang et al. [172]. Exhausted T cell EVs can be taken up by non-exhausted CD8 T cells and are able to induce exhaustion and reduced IFNγ and IL-2 cytokine production.
TAM-EVs induce colorectal cancer cells migration and invasion and provide significant plasticity of BRG1 expression, a key factor promoting the colorectal cancer metastasis. This dynamic and reciprocal cross-talk between colorectal cancer cells and TAM provides a new opportunity for the treatment of metastatic colorectal cancer. TAM-EVs facilitated CD8+ T cell exhaustion via the miR-21-5p/YOD1/YAP/β-catenin axis in HCC as described by Pu et al. [173]. TAM-EVs promote invasion and metastasis of esophageal cancer through the lncRNA AFAP1-AS1/miRNA-26a/ATF2 axis as shown by Mi et al. [174]. These observations provide novel insight on the roles of macrophage-derived EVs carrying lncRNAs in tumor pathogenesis [175]. Overall these findings enhance our knowledge on the complex interplay mediated by EVs in the tumor microenvironment and enable the identification of new targets with potential benefit for cancer patients.

4. EV-Mediated Immune Activation

TEV are able to induce immunoinhibitory signaling but also anti-tumor immunity [176,177]. TEVs carry TAAs such as MUC1 [178], HSP-70, damage-associated molecular patterns (DAMPs), which are transferred to immune cells and induce anti-tumor responses as shown in Figure 1 [44,179,180,181]. As such TEVs may be utilized as vaccine adjuvants and components of anti-tumor vaccines.
Ortiz-Bonilla CJ et al. [182] observed that priming with bladder cancer-derived TEV prevented tumor growth in mice. Pro-inflammatory factors were shown to be enriched in TEV using cytokine array analyses. The effect of these factors might have contributed to enhanced TME infiltration of immune cells in tumors primed with EVs.
Montfort et al. [183] observed that EVs produced by melanoma cells expressing sphingomyelinase (nSMase)-2 promoted the expression of IL12, CXCL9 and CCL19 by DCs in vitro, with associated CD4+ T cell activation and reduced proportion of Tregs. Lymphoma and melanoma-derived TEVs have been found to mediate and enhance DC-based anti-tumor immunity [179,184]. Glioma-TEVs reduce the proportion of Treg cells at the tumor site, attenuating immune escape [185]. Immune cells-derived EVs can also directly carry out anti-tumor functions. CD8+ T cell EVs are cytotoxic and can directly kill tumor cells or deplete mesenchymal tumor stromal cells [186]. DC-EVs enriched from patients with hepatocellular carcinoma reduce Treg cells and stimulate CD8+ T lymphocytes at the tumor site [187].
NK-EVs derived from IL-2 dependent NK-92 cells display anti-tumor activity against solid and blood cancer cells [188]. Neuroblastoma-derived EVs promote secretion of NK-EVs with enhanced cytotoxic activity [189]. In this line, NK-derived EV miR-186 inhibited neuroblastoma growth and immune escape mechanisms [168].
Zhou et al. [190] showed that miR-765, a negative regulator of proteolipid protein 2 (PLP2), was downregulated in an in vivo model of human uterine corpus endometrial cancer (UCEC), thus promoting tumor progression and epithelial mesenchymal transition (EMT). On the contrary, mir-765 was overexpressed in CD45RO-CD8+ T cells and their EVs. Treatment with these EVs reduced tumor growth via regulation of the miR-765/PLP2 axis. These observations open up new directions for the development and implementation of adjuvant therapies based on patient-derived EVs aimed at preventing cancer progression and also support the potential application of TEV in personalized medicine.
While largely displaying immunosuppressive functions in mouse cancer models [191], TAM have also been shown to display immunostimulatory activities in cancer. Cianciaruso et al. [192] observed that TAM-EVs enriched from bladder cancer tumors displayed immunostimulatory properties at both molecular and biological levels. The authors detected positive regulators of the immune response such as pattern recognition receptors (e.g., STING) and several proteins involved in TLR signaling. Analyses of the content of TAM-EVs revealed a signature associated with immunostimulatory M1-like profiles of TAM. The cargo included members of the DOCK (“dedicator of cytokinesis”) family, factors involved in intracellular signaling networks, chemotaxis and immunity, mediating the activation and recruitment of lymphocytes. Furthermore, TAM-EVs promoted proliferation of T cells and IFNγ production ex vivo, whilst EVs enriched from TAM-depleted tumors were not able to alter T cell functions.
Further studies are required to elucidate the several mechanisms involved in TEV-mediated stimulation vs. suppression of immune cells. The TME in which TEVs exert their functions may provide the context for their immunosuppressive vs. anti-tumor activity.

5. Translational Applications: Targeting EV-Mediated Tumor Immune Suppression

EVs have been exploited as therapeutic tools and delivery vehicles since the pioneering observation by Zitovgel et al. [193] that DC-derived EVs pulsed with TAAs induced CD8+ T cell responses and promoted the eradication of established tumors in mice.
EVs have great potential for drug development and as therapeutic delivery systems due to properties such as biocompatibility and stability [194]. The expression of CD47 enables EVs to avoid immune rejection via ‘don’t eat me’ signaling, contributing to the prolonged circulation time of EVs in comparison to cell-based or free drug therapies. In addition, EVs are better suited to long-term storage with limited loss of function. Importantly, ligands expressed on the surface of EVs by engaging cell receptors can activate signaling pathways in both tumor and TME compartments [195].
Adhesion molecules such as CD44, CD54 and integrins, facilitate the interaction between EVs and recipient cells (e.g., immune cells, tumor cells). Preferential uptake of TEVs by a specific cell type may be driven by the expression of specific adhesion molecules [43]. For example, integrins α6β4 and αvβ5 expressed by breast and pancreatic TEVs determined their exclusive uptake by lung fibroblasts or liver macrophages, contributing to metastasis formation [8]. Targeting these specific interactions may represent a new efficacious approach.
TEVs can also be engineered to express factors for preferential uptake by immune cells and associated with immunostimulatory activity. For example, Khani et al. [196] demonstrated DC-mediated suppression of breast cancer by engineering TEVs with miR-142 and Let7i. We have recently described the utility of EVs engineered with RNA as a strategy to re-activate the immune system against cancers including myeloma [197]. Although EVs carrying RNAs, including microRNAs, hold great promise as an innovative approach in nanomedicine to fight cancer, their use in clinical practice is still limited [198,199,200,201].
Depletion of suppressive factors such as TGF-β in TEVs has also triggered increased uptake by DCs, thereby enhancing anti-tumor immune responses. TGF-β1 was silenced using a shRNA strategy, promoting the release of TGF-β1-depleted TEVs from leukemic cells with improved activation of the immune system directed against leukemic cells [202,203].
As EVs can transfer molecules to recipient cells, ongoing translational research is focused on developing and optimizing strategies to deliver a variety of payloads with EVs for therapeutic interventions [204]. HELA-Exos, a formulation of tumor cell-specific EVs, has been recently optimized to specifically transfer the TLR3 agonist Hiltonol and the immunogenic cell death (ICD) inducer ELANE into breast cancer cells. HELA-Exos has been shown to enhance the immunogenicity of breast cancer cells and indirectly activate tumor-infiltrating DCs in both a mouse xenograft model and patient-derived tumor organoids [205]. An EV-based immune checkpoint blockade that antagonizes the interaction between CD47 and signal regulatory protein alpha (SIRPalpha) to block the “don’t eat me” signal CD47 at the tumor site has been described. EVs harboring SIRPα variants (SIRPα-exosomes) were sufficient to induce remarkably augmented phagocytosis of tumor cells by macrophages and increased CD8+ T cell TME infiltration [206]. EVs can also be loaded with PH20 hyaluronidase to break down high-molecular-weight hyaluronan (HA) in the TME. The resulting oligo-HA induces DC maturation via TLR4 activation and elicits a more potent anti-tumor response. Moreover, combining EV-PH20 with anti-PD-L1 antibody provides potent tumor-specific CD8+ T cell immune responses as well as prominent tumor growth inhibition both in syngeneic and spontaneous breast cancer models [207]. A platform designed to cross-link tumor cells with T cells and induce a strong immune response (synthetic multivalent antibodies retargeted exosome—SMART-Exosome) was able to effectively kill tumor cells both in vitro and in vivo [208]. EVs have been engineered to deliver the STING agonist cyclic GMP-AMP (iExoSTINGa). Selective targeting of the STING pathway in APCs with iExoSTINGa was associated with superior potency compared with STINGa alone in suppressing tumor growth in melanoma. Moreover, iExoSTINGa showed superior uptake of STINGa by DCs when compared to STINGa alone, hence leading to increased accumulation of activated CD8+ T cells and an anti-tumor immune response [209].
Immune cell-derived EVs have potential for cancer immunotherapy. DC-derived EVs are able to stimulate tumor-specific immune responses when loaded with TAAs (e.g., melanoma-associated antigen 3-MAGE-A3, alpha-fetoprotein-AFP) or IFN-γ [187,210,211]. The addition of hyaluronic acid (HA), 3-(diethylamino) propylamine, mono-phosphoryl lipid A and MUC1 induces uptake of macrophage-derived EVs by DCs, the release of TAAs in the endocytic compartment and subsequent improved antigen presentation and T cell activation [212]. EVs generated from NK cell membranes are enriched in FasL and TNF, making them cytotoxic to cancer [213,214].
TEV may abrogate beneficial effects of anti-tumor immune therapies. TEVs derived from B-cell lymphoma were shown to bind complement, hence protecting tumor cells from complement-dependent cytolysis in vivo [123]. Adoptive NK-92 cells transfer to relapsed AML patients has not provided significant therapeutic benefits, presumably because the transferred cells encountered circulating TEV carrying an immunosuppressive cargo [149]. As mentioned above, EVs, by enriching immunotherapeutic targets such as CD20 [44123], mediate sequestration of therapeutic antibodies, thus interfering with antibody-based therapies and reducing binding of these antibodies to tumor cells with associated inhibition of cytotoxicity directed at tumor cells. “A phase I clinical trial is currently investigating depletion of circulating EVs via a proprietary hemo-purifier device (https://clinicaltrials.gov/ct2/show/NCT04453046, accessed on 15 November 2022), neutralizing their immunosuppressive and drug resistance effects”.
Overcoming EV-mediated immune escape by blocking TEVs’ release or inhibiting their uptake may represent an additional therapeutic approach in cancer. For example, heparin has been shown to inhibit cellular uptake of TEVs with subsequent reduction of tumor cell migration, adhesion and inhibition of tumor growth and metastasis in several cancer models [215,216,217]. Further evaluation of heparin-based TEV-targeting may provide new opportunities for inhibiting immune escape.

6. Limitations and Future Perspectives

As discussed elsewhere [6,197], “a lack of standardised methodologies (e.g., EV purification, quantification, labelling) and data reporting limits inter-study comparisons and the translation of EVs from bench to bedside”. This may explain the discrepancy between the large amount of published research on the role of EVs in the tumor immune microenvironment and the few applications in clinical trial settings. Ongoing research aims to develop and optimize methodologies for isolation and purification of EVs and EV-subtypes from biological samples and for accurate quantification and content analysis. Our group has demonstrated the critical importance of samples collection and preparation for EV analyses and data interpretation [6,18,197,218]. Unpublished data inform the utility of critical evaluation of pre-analytical factors for EV-RNA isolation and preparation. “The amount and type of starting material, collection tube types, protocols for RNA isolation/preparation, representing some of the critical factors to take into consideration when working and reporting on EVs” [6].
Growing knowledge suggests dissimilar functions of factors associated with EVs vs. their soluble (non-vesicular) counterparts [44]. “EV-associated Hsp70, for example, can contribute to radiotherapy resistance in tumors [219], while immunization with non-vesicular Hsp70 is associated with more positive effects [220]”. Hence functions may not be correctly attributed to EVs or soluble factors. In fact, bio-physical properties of EVs impose challenges for their isolation/purification. Furthermore, other (co-)factors often co-isolate negatively affecting the enrichment of highly purified EVs and downstream applications (e.g., proteomics, genomics—‘omics’) and data interpretation [6,218]. Understanding the specific contribution of EVs versus other components of the tumor-associated secretome (e.g., cytokines, chemokines, growth factors) would be of critical importance in developing and implementing new promising strategies targeting single versus combined components of the secretome to improve the survival of patients with cancer [221]. As demonstrated by us and others [18,222], EVs can be isolated/purified from biofluids and cell culture supernatants [1], although “a consensus on the optimal source (i.e., plasma vs. serum) and the standardization of pre-analytical factors and reporting are still lacking”. In a recent report, we show for the first time in myeloma that the depletion of human serum albumin (a highly abundant protein in blood) from small EV preparations obtained from blood plasma of myeloma patients, improves the detection of proteins when using mass-spectrometry based proteomics [6,18]. Several methods have been developed for improved EVs’ isolation/purification from different sources and for surface and internal cargo modifications [6,18,223,224,225], with position statements regularly published to provide scientists with protocols for EV isolation/purification, analyses, accurate reporting of pre-analytical variables and methods [6,197,226]. As described in this review, circulating EVs play a role as diagnostic markers and as predictive markers of therapeutic response to immunotherapy. As recently summarized by Yu et al. [227], characteristics of EVs such as their stability and their mirroring of their parental cell in terms of composition, plus the technical capability to extract low levels of signal from background noise, makes them an intriguing proposition for use as liquid biomarkers (i.e., liquid biopsy) [218,227]. However, current methods for enrichment of EVs from complex biofluids (e.g., plasma/serum) are not able to define the cell or tissue of origin, although demonstrating the source of circulating factors (e.g., EVs, cell-free RNA and DNA) is of critical importance. “Protocols tailored to minimize the activation and release of platelet-derived EVs (the most abundant EVs in blood) have been optimized for platelet-free plasma preparation. These protocols should be widely adopted and accurately reported as discussed elsewhere” [6]. The expression levels of the TAA on EVs can be inconsistent on EVs posing additional challenges in detection, purification, analyses and reproducibility. Whiteside et al. [228] have developed a method to separate tumor from non-tumor derived EVs in blood to investigate their composition and potential application as cancer biomarkers. Of note, the study of non-tumor plasma EV fractions of each patient provided insight into changes in EV composition occurring during tumor development and progression at systemic level. Prote/gen-omic strategies may provide important insights into specific cargo enrichment and the source of EVs. A recent study by Lyden et al. [7] provided a comprehensive analysis of blood-derived EVs enriched for unique proteomic signatures in solid tumors, confirming that protein packaging reflects cancer biology and is heterogeneous across 16 types of cancer. The ‘gold standard’ ultracentrifugation was utilized for EV enrichment, although known to require high starting volumes, specialized equipment and long processing time. Furthermore, ultracentrifugation determines the co-isolation of molecules that may affect downstream applications/analyses and data interpretation, making this method unsuitable in clinical settings.
“Although not demonstrating major therapeutic benefits, trials utilizing EVs have indicated that EV-based therapy is well tolerated and clinically accomplishable” [229]. The lack of standardized EV enrichment/isolation techniques, storage methods, appropriate quality controls and inter-study comparisons, amongst other factors, has hindered further translation of EVs into clinical settings. As described in this review, EVs have been engineered for targeted drug delivery and improved efficacy of anti-cancer drugs. Engineering strategies would benefit from improved isolation methods of sub-populations of EVs and a better understanding of EV biodistribution in both pre-clinical and clinical settings [229]. Furthermore, “exploitation of allogeneic EV sources for development of off-the-shelf products would provide valuable (scalable) resource for immune-therapy applications” [230]. Finally, although beyond the scope of this review, high precision and individualized approaches, however challenging, are of critical importance and need to be tested in well-designed clinical trials. Improved outcomes of clinical trials may be achieved by better patient stratification strategies, for example by accessing to cutting-edge technologies and applying growing knowledge based on the use of fully annotated registries [https://www.mrdr.net.au/, accessed on 15 November 2022] and cytogenetic analysis or based on analyses of circulating tumor cells, circulating cell-free nucleic acids, proteins and EVs [218,231]. In this regard, we also highlight the importance of biobanks (https://www.mrdr.net.au/biobank-myeloma-1000-project/, accessed on 15 November 2022), which enable large-scale analyses for the identification of specific diseases biomarkers starting from biological or digital material with fully annotated clinical and biological data. These features are essential for improving personalized medical approaches, where effective biomarker identification is a critical step for diagnosis, prognosis and therapy prediction. The generation of biobanks for specific EV-end use would be beneficial for the progress of EV-based research.
In conclusion, a growing number of clinical trials involving EVs is ongoing, although no EV-based immunotherapies are currently approved. The biological role of TEVs is controversial with some molecules expressed on TEV displaying immune activating properties while others are immune suppressive. Further investigation of EVs’ function is warranted. In-depth analyses of the TME and TEV composition and functions will improve our understanding of EV-based immune therapies failures leading to the development of rational therapeutic application of EVs to cure cancer.

Author Contributions

Project conceptualization, manuscript preparation: A.R. Modifications of the initial draft and manuscript finalization: T.K. and A.S. All authors (A.R., T.K. and A.S.) contributed to the review article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by Monash University (Melbourne, Australia)—Postgraduate Publication Award (AR).

Conflicts of Interest

No potential conflict of interest disclosed.

Abbreviations

A
AMLacute myeloid leukaemia
APCsantigen presenting cells
ARG1arginase-1
B
BAG6large proline-rich protein BAG6
C
CCL2C-C motif chemokine ligand 2
CCRschemokine receptors
CLLchronic lymphocytic leukemia
CRCcolorectal cancer
D
DAMPsdamage-associated molecular patterns
DCdendritic cells
E
eTregeffector Treg
EGFRepidermal growth factor receptor
EMTepithelial mesenchymal transition
EOCepithelial ovarian cancer
ECMextracellular matrix
EVsExtracellular Vesicles
F
FAS-Ltumor necrosis factor ligand superfamily, member 6
FDAFood and Drug Administration
G
LGALS9galectin 9
GBMglioblastoma multiforme
gp100glycoprotein 100
H
HNSCChead and neck squamous
HDGFhepatoma-derived growth factor
HSP70heat shock protein 70
HAhyaluronan
HAhyaluronic acid
I
iDCImmature DCs
ICPimmune checkpoint proteins
ICDimmunogenic cell death
IFN-βinterferon-β
L
LGALS9galectin 9
LPSlipopolysaccharide
lnclong non-coding
LNslymph nodes
CD107a/LAMP1lysosomal-associated membrane protein 1
M
MEKK2MEK kinase 2
MART1melanoma-associated antigen 3-MAGE-A3
MIFmacrophage migration inhibitory factor
MUC1mucin-1
MMmultiple myeloma
MDSCsMyeloid-derived suppressor cells
N
NKnatural killer
NKG2DLNKG2D ligand
NPM1nucleophosmin
P
PDACpancreatic ductal adenocarcinoma
PDCD4programmed cell death factor 4
PD1programmed death 1
PD-L1programmed death ligand 1
PGE2prostaglandin E2
PLP2proteolipid protein 2
R
RFXAPregulatory factor X-associated protein
Tregsregulatory T cells
S
SIRPalphasignal regulatory protein alpha
SCLCsmall cell lung cancer
sNMaseneutral sphingomyelinase
nSMasesphingomyelinase
SPHK1sphingosine kinase-1
SPHK1sphingosine kinase 1
SCCsquamous cell carcinoma
T
TCRT-cell receptor
TERF1telomeric repeat binding factor 1
TβRIITGF-β type II receptor
TLR4Toll-like receptor 4
TGF-βtransforming growth factor β
SMAD3transforming growth factor-beta signaling protein 3
TAAstumor associated antigens
TMETumor Microenvironment
TNFtumor necrosis factor
TRAILtumor necrosis factor (TNF)-related apoptosis-inducing ligand
TAMstumor-associated macrophages
TEVstumor-derived EVs

References

  1. 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] [Green Version]
  2. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef]
  3. Wang, S.E. Extracellular vesicles in cancer therapy. Semin. Cancer Biol. 2022, 86, 296–309. [Google Scholar] [CrossRef]
  4. Reale, A.; Khong, T.; Xu, R.; Savvidou, I.; Lim, S.; Carmichael, I.; Greening, D.W.; Spencer, A. Understanding the role of extracellular vesicles in multiple myeloma. In Hematologic Malignancies; Copland, B., Ed.; Nova Science Publishers, Inc.: Hauppauge, NY, USA, 2020; pp. 217–261. [Google Scholar]
  5. Pietrowska, M.; Wlosowicz, A.; Gawin, M.; Widlak, P. MS-Based Proteomic Analysis of Serum and Plasma: Problem of High Abundant Components and Lights and Shadows of Albumin Removal. Adv. Exp. Med. Biol. 2019, 1073, 57–76. [Google Scholar] [CrossRef]
  6. Reale, A.; Khong, T.; Xu, R.; Chen, M.; Mithraprabhu, S.; Bingham, N.; Spencer, A.; Greening, D.W. Human Plasma Extracellular Vesicle Isolation and Proteomic Characterization for the Optimization of Liquid Biopsy in Multiple Myeloma. Methods Mol. Biol. 2021, 2261, 151–191. [Google Scholar] [CrossRef] [PubMed]
  7. Hoshino, A.; Kim, H.S.; Bojmar, L.; Gyan, K.E.; Cioffi, M.; Hernandez, J.; Zambirinis, C.P.; Rodrigues, G.; Molina, H.; Heissel, S.; et al. Extracellular Vesicle and Particle Biomarkers Define Multiple Human Cancers. Cell 2020, 182, 1044–1061.e18. [Google Scholar] [CrossRef]
  8. Hoshino, A.; Costa-Silva, B.; Shen, T.L.; Rodrigues, G.; Hashimoto, A.; Tesic Mark, M.; Molina, H.; Kohsaka, S.; Di Giannatale, A.; Ceder, S.; et al. Tumour exosome integrins determine organotropic metastasis. Nature 2015, 527, 329–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Peinado, H.; Alečković, M.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; García-Santos, G.; Ghajar, C.; et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat. Med. 2012, 18, 883–891. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, J.; Hendrix, A.; Hernot, S.; Lemaire, M.; De Bruyne, E.; Van Valckenborgh, E.; Lahoutte, T.; De Wever, O.; Vanderkerken, K.; Menu, E. Bone marrow stromal cell-derived exosomes as communicators in drug resistance in multiple myeloma cells. Blood 2014, 124, 555–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Roccaro, A.M.; Sacco, A.; Maiso, P.; Azab, A.K.; Tai, Y.T.; Reagan, M.; Azab, F.; Flores, L.M.; Campigotto, F.; Weller, E.; et al. BM mesenchymal stromal cell-derived exosomes facilitate multiple myeloma progression. J. Clin. Investig. 2013, 123, 1542–1555. [Google Scholar] [CrossRef] [Green Version]
  12. Buzas, E.I. The roles of extracellular vesicles in the immune system. Nat. Rev. Immunol. 2022. [Google Scholar] [CrossRef] [PubMed]
  13. Hao, Q.; Wu, Y.; Wu, Y.; Wang, P.; Vadgama, J.V. Tumor-Derived Exosomes in Tumor-Induced Immune Suppression. Int. J. Mol. Sci. 2022, 23, 1461. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, R.; Rai, A.; Chen, M.; Suwakulsiri, W.; Greening, D.W.; Simpson, R.J. Extracellular vesicles in cancer—Implications for future improvements in cancer care. Nat. Rev. Clin. Oncol. 2018, 15, 617–638. [Google Scholar] [CrossRef] [PubMed]
  15. Fontana, F.; Carollo, E.; Melling, G.E.; Carter, D.R.F. Extracellular Vesicles: Emerging Modulators of Cancer Drug Resistance. Cancers 2021, 13, 749. [Google Scholar] [CrossRef] [PubMed]
  16. Uchihara, T.; Miyake, K.; Yonemura, A.; Komohara, Y.; Itoyama, R.; Koiwa, M.; Yasuda, T.; Arima, K.; Harada, K.; Eto, K.; et al. Extracellular Vesicles from Cancer-Associated Fibroblasts Containing Annexin A6 Induces FAK-YAP Activation by Stabilizing beta1 Integrin, Enhancing Drug Resistance. Cancer Res. 2020, 80, 3222–3235. [Google Scholar] [CrossRef] [PubMed]
  17. Tutrone, R.; Donovan, M.J.; Torkler, P.; Tadigotla, V.; McLain, T.; Noerholm, M.; Skog, J.; McKiernan, J. Clinical utility of the exosome based ExoDx Prostate(IntelliScore) EPI test in men presenting for initial Biopsy with a PSA 2–10 ng/mL. Prostate Cancer Prostatic Dis. 2020, 23, 607–614. [Google Scholar] [CrossRef] [PubMed]
  18. Reale, A.; Carmichael, I.; Xu, R.; Mithraprabhu, S.; Khong, T.; Chen, M.; Fang, H.; Savvidou, I.; Ramachandran, M.; Bingham, N.; et al. Human myeloma cell- and plasma-derived extracellular vesicles contribute to functional regulation of stromal cells. Proteomics 2021, 21, e2000119. [Google Scholar] [CrossRef] [PubMed]
  19. Desantis, V.; Savino, F.D.; Scaringella, A.; Potenza, M.A.; Nacci, C.; Frassanito, M.A.; Vacca, A.; Montagnani, M. The Leading Role of the Immune Microenvironment in Multiple Myeloma: A New Target with a Great Prognostic and Clinical Value. J. Clin. Med. 2022, 11, 2513. [Google Scholar] [CrossRef]
  20. Weiss, F.; Lauffenburger, D.; Friedl, P. Towards targeting of shared mechanisms of cancer metastasis and therapy resistance. Nat. Rev. Cancer 2022, 22, 157–173. [Google Scholar] [CrossRef] [PubMed]
  21. Wu, F.; Yang, J.; Liu, J.; Wang, Y.; Mu, J.; Zeng, Q.; Deng, S.; Zhou, H. Signaling pathways in cancer-associated fibroblasts and targeted therapy for cancer. Signal Transduct. Target. Ther. 2021, 6, 218. [Google Scholar] [CrossRef]
  22. Morana, O.; Wood, W.; Gregory, C.D. The Apoptosis Paradox in Cancer. Int. J. Mol. Sci. 2022, 23, 1328. [Google Scholar] [CrossRef] [PubMed]
  23. Singleton, D.C.; Macann, A.; Wilson, W.R. Therapeutic targeting of the hypoxic tumour microenvironment. Nat. Rev. Clin. Oncol. 2021, 18, 751–772. [Google Scholar] [CrossRef] [PubMed]
  24. Valkenburg, K.C.; de Groot, A.E.; Pienta, K.J. Targeting the tumour stroma to improve cancer therapy. Nat. Rev. Clin. Oncol. 2018, 15, 366–381. [Google Scholar] [CrossRef]
  25. Pittet, M.J.; Michielin, O.; Migliorini, D. Clinical relevance of tumour-associated macrophages. Nat. Rev. Clin. Oncol. 2022, 19, 402–421. [Google Scholar] [CrossRef] [PubMed]
  26. Bruni, D.; Angell, H.K.; Galon, J. The immune contexture and Immunoscore in cancer prognosis and therapeutic efficacy. Nat. Rev. Cancer 2020, 20, 662–680. [Google Scholar] [CrossRef]
  27. Kalaora, S.; Nagler, A.; Wargo, J.A.; Samuels, Y. Mechanisms of immune activation and regulation: Lessons from melanoma. Nat. Rev. Cancer 2022, 22, 195–207. [Google Scholar] [CrossRef]
  28. Labani-Motlagh, A.; Ashja-Mahdavi, M.; Loskog, A. The Tumor Microenvironment: A Milieu Hindering and Obstructing Antitumor Immune Responses. Front. Immunol. 2020, 11, 940. [Google Scholar] [CrossRef]
  29. Kraehenbuehl, L.; Weng, C.H.; Eghbali, S.; Wolchok, J.D.; Merghoub, T. Enhancing immunotherapy in cancer by targeting emerging immunomodulatory pathways. Nat. Rev. Clin. Oncol. 2022, 19, 37–50. [Google Scholar] [CrossRef]
  30. Couzin-Frankel, J. Breakthrough of the year 2013. Cancer immunotherapy. Science 2013, 342, 1432–1433. [Google Scholar] [CrossRef] [Green Version]
  31. Wolchok, J. Putting the Immunologic Brakes on Cancer. Cell 2018, 175, 1452–1454. [Google Scholar] [CrossRef]
  32. Koury, J.; Lucero, M.; Cato, C.; Chang, L.; Geiger, J.; Henry, D.; Hernandez, J.; Hung, F.; Kaur, P.; Teskey, G.; et al. Immunotherapies: Exploiting the Immune System for Cancer Treatment. J. Immunol. Res. 2018, 2018, 9585614. [Google Scholar] [CrossRef] [PubMed]
  33. Murciano-Goroff, Y.R.; Warner, A.B.; Wolchok, J.D. The future of cancer immunotherapy: Microenvironment-targeting combinations. Cell Res. 2020, 30, 507–519. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, S.; Mi, Y.; Zheng, B.; Wei, P.; Gu, Y.; Zhang, Z.; Xu, Y.; Cai, S.; Li, X.; Li, D. Highly-metastatic colorectal cancer cell released miR-181a-5p-rich extracellular vesicles promote liver metastasis by activating hepatic stellate cells and remodelling the tumour microenvironment. J. Extracell. Vesicles 2022, 11, e12186. [Google Scholar] [CrossRef] [PubMed]
  35. Sirkisoon, S.R.; Wong, G.L.; Aguayo, N.R.; Doheny, D.L.; Zhu, D.; Regua, A.T.; Arrigo, A.; Manore, S.G.; Wagner, C.; Thomas, A.; et al. Breast cancer extracellular vesicles-derived miR-1290 activates astrocytes in the brain metastatic microenvironment via the FOXA2-->CNTF axis to promote progression of brain metastases. Cancer Lett. 2022, 540, 215726. [Google Scholar] [CrossRef]
  36. Giusti, I.; Di Francesco, M.; D’Ascenzo, S.; Palmerini, M.G.; Macchiarelli, G.; Carta, G.; Dolo, V. Ovarian cancer-derived extracellular vesicles affect normal human fibroblast behavior. Cancer Biol. Ther. 2018, 19, 722–734. [Google Scholar] [CrossRef] [Green Version]
  37. Fang, T.; Lv, H.; Lv, G.; Li, T.; Wang, C.; Han, Q.; Yu, L.; Su, B.; Guo, L.; Huang, S.; et al. Tumor-derived exosomal miR-1247-3p induces cancer-associated fibroblast activation to foster lung metastasis of liver cancer. Nat. Commun. 2018, 9, 191. [Google Scholar] [CrossRef] [Green Version]
  38. Forder, A.; Hsing, C.Y.; Trejo Vazquez, J.; Garnis, C. Emerging Role of Extracellular Vesicles and Cellular Communication in Metastasis. Cells 2021, 10, 3429. [Google Scholar] [CrossRef]
  39. Faict, S.; Muller, J.; De Veirman, K.; De Bruyne, E.; Maes, K.; Vrancken, L.; Heusschen, R.; De Raeve, H.; Schots, R.; Vanderkerken, K.; et al. Exosomes play a role in multiple myeloma bone disease and tumor development by targeting osteoclasts and osteoblasts. Blood Cancer J. 2018, 8, 105. [Google Scholar] [CrossRef] [Green Version]
  40. Menu, E.; Vanderkerken, K. Exosomes in Multiple Myeloma: From bench to bedside. Blood 2022. [Google Scholar] [CrossRef]
  41. Wang, J.; De Veirman, K.; Faict, S.; Frassanito, M.A.; Ribatti, D.; Vacca, A.; Menu, E. Multiple myeloma exosomes establish a favourable bone marrow microenvironment with enhanced angiogenesis and immunosuppression. J. Pathol. 2016, 239, 162–173. [Google Scholar] [CrossRef]
  42. Yamaguchi, T.; Kawamoto, E.; Gaowa, A.; Park, E.J.; Shimaoka, M. Remodeling of Bone Marrow Niches and Roles of Exosomes in Leukemia. Int. J. Mol. Sci. 2021, 22, 1881. [Google Scholar] [CrossRef]
  43. Vergani, E.; Daveri, E.; Vallacchi, V.; Bergamaschi, L.; Lalli, L.; Castelli, C.; Rodolfo, M.; Rivoltini, L.; Huber, V. Extracellular vesicles in anti-tumor immunity. Semin. Cancer Biol. 2022, 86, 64–79. [Google Scholar] [CrossRef] [PubMed]
  44. Marar, C.; Starich, B.; Wirtz, D. Extracellular vesicles in immunomodulation and tumor progression. Nat. Immunol. 2021, 22, 560–570. [Google Scholar] [CrossRef]
  45. Kugeratski, F.G.; Kalluri, R. Exosomes as mediators of immune regulation and immunotherapy in cancer. FEBS J. 2021, 288, 10–35. [Google Scholar] [CrossRef]
  46. Ludwig, S.; Floros, T.; Theodoraki, M.N.; Hong, C.S.; Jackson, E.K.; Lang, S.; Whiteside, T.L. Suppression of Lymphocyte Functions by Plasma Exosomes Correlates with Disease Activity in Patients with Head and Neck Cancer. Clin. Cancer Res. 2017, 23, 4843–4854. [Google Scholar] [CrossRef] [Green Version]
  47. Ricklefs, F.L.; Alayo, Q.; Krenzlin, H.; Mahmoud, A.B.; Speranza, M.C.; Nakashima, H.; Hayes, J.L.; Lee, K.; Balaj, L.; Passaro, C.; et al. Immune evasion mediated by PD-L1 on glioblastoma-derived extracellular vesicles. Sci. Adv. 2018, 4, eaar2766. [Google Scholar] [CrossRef] [Green Version]
  48. Czystowska-Kuzmicz, M.; Sosnowska, A.; Nowis, D.; Ramji, K.; Szajnik, M.; Chlebowska-Tuz, J.; Wolinska, E.; Gaj, P.; Grazul, M.; Pilch, Z.; et al. Small extracellular vesicles containing arginase-1 suppress T-cell responses and promote tumor growth in ovarian carcinoma. Nat. Commun. 2019, 10, 3000. [Google Scholar] [CrossRef] [Green Version]
  49. Haderk, F.; Schulz, R.; Iskar, M.; Cid, L.L.; Worst, T.; Willmund, K.V.; Schulz, A.; Warnken, U.; Seiler, J.; Benner, A.; et al. Tumor-derived exosomes modulate PD-L1 expression in monocytes. Sci. Immunol. 2017, 2, eaah5509. [Google Scholar] [CrossRef] [Green Version]
  50. Twomey, J.D.; Zhang, B. Cancer Immunotherapy Update: FDA-Approved Checkpoint Inhibitors and Companion Diagnostics. AAPS J. 2021, 23, 39. [Google Scholar] [CrossRef]
  51. Daassi, D.; Mahoney, K.M.; Freeman, G.J. The importance of exosomal PDL1 in tumour immune evasion. Nat. Rev. Immunol. 2020, 20, 209–215. [Google Scholar] [CrossRef]
  52. Chen, G.; Huang, A.C.; Zhang, W.; Zhang, G.; Wu, M.; Xu, W.; Yu, Z.; Yang, J.; Wang, B.; Sun, H.; et al. Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature 2018, 560, 382–386. [Google Scholar] [CrossRef]
  53. Dou, X.; Hua, Y.; Chen, Z.; Chao, F.; Li, M. Extracellular vesicles containing PD-L1 contribute to CD8+ T-cell immune suppression and predict poor outcomes in small cell lung cancer. Clin. Exp. Immunol. 2022, 207, 307–317. [Google Scholar] [CrossRef] [PubMed]
  54. Gupta, P.; Kadamberi, I.P.; Mittal, S.; Tsaih, S.W.; George, J.; Kumar, S.; Vijayan, D.K.; Geethadevi, A.; Parashar, D.; Topchyan, P.; et al. Tumor Derived Extracellular Vesicles Drive T Cell Exhaustion in Tumor Microenvironment through Sphingosine Mediated Signaling and Impacting Immunotherapy Outcomes in Ovarian Cancer. Adv. Sci. 2022, 9, e2104452. [Google Scholar] [CrossRef] [PubMed]
  55. Xie, F.; Zhou, X.; Su, P.; Li, H.; Tu, Y.; Du, J.; Pan, C.; Wei, X.; Zheng, M.; Jin, K.; et al. Breast cancer cell-derived extracellular vesicles promote CD8(+) T cell exhaustion via TGF-beta type II receptor signaling. Nat. Commun. 2022, 13, 4461. [Google Scholar] [CrossRef]
  56. Vignard, V.; Labbe, M.; Marec, N.; Andre-Gregoire, G.; Jouand, N.; Fonteneau, J.F.; Labarriere, N.; Fradin, D. MicroRNAs in Tumor Exosomes Drive Immune Escape in Melanoma. Cancer Immunol. Res. 2020, 8, 255–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Peng, M.; Ren, J.; Jing, Y.; Jiang, X.; Xiao, Q.; Huang, J.; Tao, Y.; Lei, L.; Wang, X.; Yang, Z.; et al. Tumour-derived small extracellular vesicles suppress CD8+ T cell immune function by inhibiting SLC6A8-mediated creatine import in NPM1-mutated acute myeloid leukaemia. J. Extracell. Vesicles 2021, 10, e12168. [Google Scholar] [CrossRef]
  58. Wieckowski, E.U.; Visus, C.; Szajnik, M.; Szczepanski, M.J.; Storkus, W.J.; Whiteside, T.L. Tumor-derived microvesicles promote regulatory T cell expansion and induce apoptosis in tumor-reactive activated CD8+ T lymphocytes. J. Immunol. 2009, 183, 3720–3730. [Google Scholar] [CrossRef] [Green Version]
  59. Theodoraki, M.N.; Hoffmann, T.K.; Jackson, E.K.; Whiteside, T.L. Exosomes in HNSCC plasma as surrogate markers of tumour progression and immune competence. Clin. Exp. Immunol. 2018, 194, 67–78. [Google Scholar] [CrossRef] [Green Version]
  60. Yang, Y.; Li, C.W.; Chan, L.C.; Wei, Y.; Hsu, J.M.; Xia, W.; Cha, J.H.; Hou, J.; Hsu, J.L.; Sun, L.; et al. Exosomal PD-L1 harbors active defense function to suppress T cell killing of breast cancer cells and promote tumor growth. Cell Res. 2018, 28, 862–864. [Google Scholar] [CrossRef] [Green Version]
  61. Poggio, M.; Hu, T.; Pai, C.C.; Chu, B.; Belair, C.D.; Chang, A.; Montabana, E.; Lang, U.E.; Fu, Q.; Fong, L.; et al. Suppression of Exosomal PD-L1 Induces Systemic Anti-tumor Immunity and Memory. Cell 2019, 177, 414–427.e13. [Google Scholar] [CrossRef]
  62. Kim, D.H.; Kim, H.; Choi, Y.J.; Kim, S.Y.; Lee, J.E.; Sung, K.J.; Sung, Y.H.; Pack, C.G.; Jung, M.K.; Han, B.; et al. Exosomal PD-L1 promotes tumor growth through immune escape in non-small cell lung cancer. Exp. Mol. Med. 2019, 51, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cordonnier, M.; Nardin, C.; Chanteloup, G.; Derangere, V.; Algros, M.P.; Arnould, L.; Garrido, C.; Aubin, F.; Gobbo, J. Tracking the evolution of circulating exosomal-PD-L1 to monitor melanoma patients. J. Extracell. Vesicles 2020, 9, 1710899. [Google Scholar] [CrossRef] [Green Version]
  64. Zhang, M.; Fan, Y.; Che, X.; Hou, K.; Zhang, C.; Li, C.; Wen, T.; Wang, S.; Cheng, Y.; Liu, Y.; et al. 5-FU-Induced Upregulation of Exosomal PD-L1 Causes Immunosuppression in Advanced Gastric Cancer Patients. Front. Oncol. 2020, 10, 492. [Google Scholar] [CrossRef] [Green Version]
  65. Wang, X.; Shen, H.; Zhangyuan, G.; Huang, R.; Zhang, W.; He, Q.; Jin, K.; Zhuo, H.; Zhang, Z.; Wang, J.; et al. 14-3-3zeta delivered by hepatocellular carcinoma-derived exosomes impaired anti-tumor function of tumor-infiltrating T lymphocytes. Cell Death Dis. 2018, 9, 159. [Google Scholar] [CrossRef] [Green Version]
  66. Shenoy, G.N.; Loyall, J.; Berenson, C.S.; Kelleher, R.J., Jr.; Iyer, V.; Balu-Iyer, S.V.; Odunsi, K.; Bankert, R.B. Sialic Acid-Dependent Inhibition of T Cells by Exosomal Ganglioside GD3 in Ovarian Tumor Microenvironments. J. Immunol. 2018, 201, 3750–3758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Shenoy, G.N.; Loyall, J.; Maguire, O.; Iyer, V.; Kelleher, R.J., Jr.; Minderman, H.; Wallace, P.K.; Odunsi, K.; Balu-Iyer, S.V.; Bankert, R.B. Exosomes Associated with Human Ovarian Tumors Harbor a Reversible Checkpoint of T-cell Responses. Cancer Immunol. Res. 2018, 6, 236–247. [Google Scholar] [CrossRef] [Green Version]
  68. Sharma, P.; Diergaarde, B.; Ferrone, S.; Kirkwood, J.M.; Whiteside, T.L. Melanoma cell-derived exosomes in plasma of melanoma patients suppress functions of immune effector cells. Sci. Rep. 2020, 10, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Zhou, J.; Yang, Y.; Wang, W.; Zhang, Y.; Chen, Z.; Hao, C.; Zhang, J. Melanoma-released exosomes directly activate the mitochondrial apoptotic pathway of CD4(+) T cells through their microRNA cargo. Exp. Cell Res. 2018, 371, 364–371. [Google Scholar] [CrossRef]
  70. Kalvala, A.; Wallet, P.; Yang, L.; Wang, C.; Li, H.; Nam, A.; Nathan, A.; Mambetsariev, I.; Poroyko, V.; Gao, H.; et al. Phenotypic Switching of Naive T Cells to Immune-Suppressive Treg-Like Cells by Mutant KRAS. J. Clin. Med. 2019, 8, 1726. [Google Scholar] [CrossRef] [Green Version]
  71. Ni, C.; Fang, Q.Q.; Chen, W.Z.; Jiang, J.X.; Jiang, Z.; Ye, J.; Zhang, T.; Yang, L.; Meng, F.B.; Xia, W.J.; et al. Breast cancer-derived exosomes transmit lncRNA SNHG16 to induce CD73+gammadelta1 Treg cells. Signal Transduct. Target. Ther. 2020, 5, 41. [Google Scholar] [CrossRef] [PubMed]
  72. Clayton, A.; Al-Taei, S.; Webber, J.; Mason, M.D.; Tabi, Z. Cancer exosomes express CD39 and CD73, which suppress T cells through adenosine production. J. Immunol. 2011, 187, 676–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ye, L.; Zhang, Q.; Cheng, Y.; Chen, X.; Wang, G.; Shi, M.; Zhang, T.; Cao, Y.; Pan, H.; Zhang, L.; et al. Tumor-derived exosomal HMGB1 fosters hepatocellular carcinoma immune evasion by promoting TIM-1(+) regulatory B cell expansion. J. Immunother. Cancer 2018, 6, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Capello, M.; Vykoukal, J.V.; Katayama, H.; Bantis, L.E.; Wang, H.; Kundnani, D.L.; Aguilar-Bonavides, C.; Aguilar, M.; Tripathi, S.C.; Dhillon, D.S.; et al. Exosomes harbor B cell targets in pancreatic adenocarcinoma and exert decoy function against complement-mediated cytotoxicity. Nat. Commun. 2019, 10, 254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Yin, Y.; Cai, X.; Chen, X.; Liang, H.; Zhang, Y.; Li, J.; Wang, Z.; Chen, X.; Zhang, W.; Yokoyama, S.; et al. Tumor-secreted miR-214 induces regulatory T cells: A major link between immune evasion and tumor growth. Cell Res. 2014, 24, 1164–1180. [Google Scholar] [CrossRef] [PubMed]
  76. Chalmin, F.; Ladoire, S.; Mignot, G.; Vincent, J.; Bruchard, M.; Remy-Martin, J.P.; Boireau, W.; Rouleau, A.; Simon, B.; Lanneau, D.; et al. Membrane-associated Hsp72 from tumor-derived exosomes mediates STAT3-dependent immunosuppressive function of mouse and human myeloid-derived suppressor cells. J. Clin. Investig. 2010, 120, 457–471. [Google Scholar] [CrossRef]
  77. Gao, Y.; Xu, H.; Li, N.; Wang, H.; Ma, L.; Chen, S.; Liu, J.; Zheng, Y.; Zhang, Y. Renal cancer-derived exosomes induce tumor immune tolerance by MDSCs-mediated antigen-specific immunosuppression. Cell Commun. Signal 2020, 18, 106. [Google Scholar] [CrossRef]
  78. Jiang, M.; Zhang, W.; Zhang, R.; Liu, P.; Ye, Y.; Yu, W.; Guo, X.; Yu, J. Cancer exosome-derived miR-9 and miR-181a promote the development of early-stage MDSCs via interfering with SOCS3 and PIAS3 respectively in breast cancer. Oncogene 2020, 39, 4681–4694. [Google Scholar] [CrossRef]
  79. Guo, X.; Qiu, W.; Liu, Q.; Qian, M.; Wang, S.; Zhang, Z.; Gao, X.; Chen, Z.; Xue, H.; Li, G. Immunosuppressive effects of hypoxia-induced glioma exosomes through myeloid-derived suppressor cells via the miR-10a/Rora and miR-21/Pten Pathways. Oncogene 2018, 37, 4239–4259. [Google Scholar] [CrossRef]
  80. Wang, X.; Luo, G.; Zhang, K.; Cao, J.; Huang, C.; Jiang, T.; Liu, B.; Su, L.; Qiu, Z. Hypoxic Tumor-Derived Exosomal miR-301a Mediates M2 Macrophage Polarization via PTEN/PI3Kgamma to Promote Pancreatic Cancer Metastasis. Cancer Res. 2018, 78, 4586–4598. [Google Scholar] [CrossRef] [Green Version]
  81. Zhao, J.; Schlosser, H.A.; Wang, Z.; Qin, J.; Li, J.; Popp, F.; Popp, M.C.; Alakus, H.; Chon, S.H.; Hansen, H.P.; et al. Tumor-Derived Extracellular Vesicles Inhibit Natural Killer Cell Function in Pancreatic Cancer. Cancers 2019, 11, 874. [Google Scholar] [CrossRef]
  82. Chang, W.A.; Tsai, M.J.; Hung, J.Y.; Wu, K.L.; Tsai, Y.M.; Huang, Y.C.; Chang, C.Y.; Tsai, P.H.; Hsu, Y.L. miR-150-5p-Containing Extracellular Vesicles Are a New Immunoregulator That Favor the Progression of Lung Cancer in Hypoxic Microenvironments by Altering the Phenotype of NK Cells. Cancers 2021, 13, 6252. [Google Scholar] [CrossRef] [PubMed]
  83. Ding, G.; Zhou, L.; Qian, Y.; Fu, M.; Chen, J.; Chen, J.; Xiang, J.; Wu, Z.; Jiang, G.; Cao, L. Pancreatic cancer-derived exosomes transfer miRNAs to dendritic cells and inhibit RFXAP expression via miR-212-3p. Oncotarget 2015, 6, 29877–29888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhou, M.; Chen, J.; Zhou, L.; Chen, W.; Ding, G.; Cao, L. Pancreatic cancer derived exosomes regulate the expression of TLR4 in dendritic cells via miR-203. Cell Immunol. 2014, 292, 65–69. [Google Scholar] [CrossRef]
  85. Wang, M.; Cai, Y.; Peng, Y.; Xu, B.; Hui, W.; Jiang, Y. Exosomal LGALS9 in the cerebrospinal fluid of glioblastoma patients suppressed dendritic cell antigen presentation and cytotoxic T-cell immunity. Cell Death Dis. 2020, 11, 896. [Google Scholar] [CrossRef]
  86. De Paula Silva, E.; Marti, L.C.; Andreghetto, F.M.; de Sales, R.O.; Hoberman, M.; Dos Santos Dias, B.; Diniz, L.F.A.; Dos Santos, A.M.; Moyses, R.A.; Curioni, O.A.; et al. Extracellular vesicles cargo from head and neck cancer cell lines disrupt dendritic cells function and match plasma microRNAs. Sci. Rep. 2021, 11, 18534. [Google Scholar] [CrossRef] [PubMed]
  87. Salimu, J.; Webber, J.; Gurney, M.; Al-Taei, S.; Clayton, A.; Tabi, Z. Dominant immunosuppression of dendritic cell function by prostate-cancer-derived exosomes. J. Extracell. Vesicles 2017, 6, 1368823. [Google Scholar] [CrossRef] [Green Version]
  88. Ning, Y.; Shen, K.; Wu, Q.; Sun, X.; Bai, Y.; Xie, Y.; Pan, J.; Qi, C. Tumor exosomes block dendritic cells maturation to decrease the T cell immune response. Immunol. Lett. 2018, 199, 36–43. [Google Scholar] [CrossRef]
  89. Huber, V.; Vallacchi, V.; Fleming, V.; Hu, X.; Cova, A.; Dugo, M.; Shahaj, E.; Sulsenti, R.; Vergani, E.; Filipazzi, P.; et al. Tumor-derived microRNAs induce myeloid suppressor cells and predict immunotherapy resistance in melanoma. J. Clin. Investig. 2018, 128, 5505–5516. [Google Scholar] [CrossRef] [Green Version]
  90. Fleming, V.; Hu, X.; Weller, C.; Weber, R.; Groth, C.; Riester, Z.; Huser, L.; Sun, Q.; Nagibin, V.; Kirschning, C.; et al. Melanoma Extracellular Vesicles Generate Immunosuppressive Myeloid Cells by Upregulating PD-L1 via TLR4 Signaling. Cancer Res. 2019, 79, 4715–4728. [Google Scholar] [CrossRef]
  91. Xiang, X.; Poliakov, A.; Liu, C.; Liu, Y.; Deng, Z.B.; Wang, J.; Cheng, Z.; Shah, S.V.; Wang, G.J.; Zhang, L.; et al. Induction of myeloid-derived suppressor cells by tumor exosomes. Int. J. Cancer 2009, 124, 2621–2633. [Google Scholar] [CrossRef]
  92. Li, L.; Cao, B.; Liang, X.; Lu, S.; Luo, H.; Wang, Z.; Wang, S.; Jiang, J.; Lang, J.; Zhu, G. Microenvironmental oxygen pressure orchestrates an anti- and pro-tumoral gammadelta T cell equilibrium via tumor-derived exosomes. Oncogene 2019, 38, 2830–2843. [Google Scholar] [CrossRef] [PubMed]
  93. Hsu, Y.L.; Hung, J.Y.; Chang, W.A.; Jian, S.F.; Lin, Y.S.; Pan, Y.C.; Wu, C.Y.; Kuo, P.L. Hypoxic Lung-Cancer-Derived Extracellular Vesicle MicroRNA-103a Increases the Oncogenic Effects of Macrophages by Targeting PTEN. Mol. Ther. 2018, 26, 568–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Gao, L.; Wang, L.; Dai, T.; Jin, K.; Zhang, Z.; Wang, S.; Xie, F.; Fang, P.; Yang, B.; Huang, H.; et al. Tumor-derived exosomes antagonize innate antiviral immunity. Nat. Immunol. 2018, 19, 233–245. [Google Scholar] [CrossRef] [PubMed]
  95. Park, J.E.; Dutta, B.; Tse, S.W.; Gupta, N.; Tan, C.F.; Low, J.K.; Yeoh, K.W.; Kon, O.L.; Tam, J.P.; Sze, S.K. Hypoxia-induced tumor exosomes promote M2-like macrophage polarization of infiltrating myeloid cells and microRNA-mediated metabolic shift. Oncogene 2019, 38, 5158–5173. [Google Scholar] [CrossRef]
  96. Cooks, T.; Pateras, I.S.; Jenkins, L.M.; Patel, K.M.; Robles, A.I.; Morris, J.; Forshew, T.; Appella, E.; Gorgoulis, V.G.; Harris, C.C. Mutant p53 cancers reprogram macrophages to tumor supporting macrophages via exosomal miR-1246. Nat. Commun. 2018, 9, 771. [Google Scholar] [CrossRef] [Green Version]
  97. Gerloff, D.; Lutzkendorf, J.; Moritz, R.K.C.; Wersig, T.; Mader, K.; Muller, L.P.; Sunderkotter, C. Melanoma-Derived Exosomal miR-125b-5p Educates Tumor Associated Macrophages (TAMs) by Targeting Lysosomal Acid Lipase A (LIPA). Cancers 2020, 12, 464. [Google Scholar] [CrossRef] [Green Version]
  98. Linton, S.S.; Abraham, T.; Liao, J.; Clawson, G.A.; Butler, P.J.; Fox, T.; Kester, M.; Matters, G.L. Tumor-promoting effects of pancreatic cancer cell exosomes on THP-1-derived macrophages. PLoS ONE 2018, 13, e0206759. [Google Scholar] [CrossRef] [Green Version]
  99. Dai, E.; Han, L.; Liu, J.; Xie, Y.; Kroemer, G.; Klionsky, D.J.; Zeh, H.J.; Kang, R.; Wang, J.; Tang, D. Autophagy-dependent ferroptosis drives tumor-associated macrophage polarization via release and uptake of oncogenic KRAS protein. Autophagy 2020, 16, 2069–2083. [Google Scholar] [CrossRef]
  100. Wang, S.; Gao, Y. Pancreatic cancer cell-derived microRNA-155-5p-containing extracellular vesicles promote immune evasion by triggering EHF-dependent activation of Akt/NF-kappaB signaling pathway. Int. Immunopharmacol. 2021, 100, 107990. [Google Scholar] [CrossRef]
  101. Wolf-Dennen, K.; Gordon, N.; Kleinerman, E.S. Exosomal communication by metastatic osteosarcoma cells modulates alveolar macrophages to an M2 tumor-promoting phenotype and inhibits tumoricidal functions. Oncoimmunology 2020, 9, 1747677. [Google Scholar] [CrossRef]
  102. Liu, Y.; Lu, M.; Chen, J.; Li, S.; Deng, Y.; Yang, S.; Ou, Q.; Li, J.; Gao, P.; Luo, Z.; et al. Extracellular vesicles derived from lung cancer cells exposed to intermittent hypoxia upregulate programmed death ligand 1 expression in macrophages. Sleep Breath. 2022, 26, 893–906. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, D.; Wang, X.; Si, M.; Yang, J.; Sun, S.; Wu, H.; Cui, S.; Qu, X.; Yu, X. Exosome-encapsulated miRNAs contribute to CXCL12/CXCR4-induced liver metastasis of colorectal cancer by enhancing M2 polarization of macrophages. Cancer Lett. 2020, 474, 36–52. [Google Scholar] [CrossRef] [PubMed]
  104. Popena, I.; Abols, A.; Saulite, L.; Pleiko, K.; Zandberga, E.; Jekabsons, K.; Endzelins, E.; Llorente, A.; Line, A.; Riekstina, U. Effect of colorectal cancer-derived extracellular vesicles on the immunophenotype and cytokine secretion profile of monocytes and macrophages. Cell Commun. Signal 2018, 16, 17. [Google Scholar] [CrossRef] [Green Version]
  105. Ham, S.; Lima, L.G.; Chai, E.P.Z.; Muller, A.; Lobb, R.J.; Krumeich, S.; Wen, S.W.; Wiegmans, A.P.; Moller, A. Breast Cancer-Derived Exosomes Alter Macrophage Polarization via gp130/STAT3 Signaling. Front. Immunol. 2018, 9, 871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. He, C.; Hua, W.; Liu, J.; Fan, L.; Wang, H.; Sun, G. Exosomes derived from endoplasmic reticulum-stressed liver cancer cells enhance the expression of cytokines in macrophages via the STAT3 signaling pathway. Oncol. Lett. 2020, 20, 589–600. [Google Scholar] [CrossRef] [PubMed]
  107. Yao, X.; Tu, Y.; Xu, Y.; Guo, Y.; Yao, F.; Zhang, X. Endoplasmic reticulum stress-induced exosomal miR-27a-3p promotes immune escape in breast cancer via regulating PD-L1 expression in macrophages. J. Cell Mol. Med. 2020, 24, 9560–9573. [Google Scholar] [CrossRef] [PubMed]
  108. Chow, A.; Zhou, W.; Liu, L.; Fong, M.Y.; Champer, J.; Van Haute, D.; Chin, A.R.; Ren, X.; Gugiu, B.G.; Meng, Z.; et al. Macrophage immunomodulation by breast cancer-derived exosomes requires Toll-like receptor 2-mediated activation of NF-kappaB. Sci. Rep. 2014, 4, 5750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wu, L.; Zhang, X.; Zhang, B.; Shi, H.; Yuan, X.; Sun, Y.; Pan, Z.; Qian, H.; Xu, W. Exosomes derived from gastric cancer cells activate NF-kappaB pathway in macrophages to promote cancer progression. Tumour. Biol. 2016, 37, 12169–12180. [Google Scholar] [CrossRef]
  110. Fabbri, M.; Paone, A.; Calore, F.; Galli, R.; Gaudio, E.; Santhanam, R.; Lovat, F.; Fadda, P.; Mao, C.; Nuovo, G.J.; et al. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc. Natl. Acad. Sci. USA 2012, 109, E2110–E2116. [Google Scholar] [CrossRef] [Green Version]
  111. Chen, X.; Zhou, J.; Li, X.; Wang, X.; Lin, Y.; Wang, X. Exosomes derived from hypoxic epithelial ovarian cancer cells deliver microRNAs to macrophages and elicit a tumor-promoted phenotype. Cancer Lett. 2018, 435, 80–91. [Google Scholar] [CrossRef]
  112. Javeed, N.; Gustafson, M.P.; Dutta, S.K.; Lin, Y.; Bamlet, W.R.; Oberg, A.L.; Petersen, G.M.; Chari, S.T.; Dietz, A.B.; Mukhopadhyay, D. Immunosuppressive CD14(+)HLA-DR(lo/neg) monocytes are elevated in pancreatic cancer and "primed" by tumor-derived exosomes. Oncoimmunology 2017, 6, e1252013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Hsieh, C.H.; Tai, S.K.; Yang, M.H. Snail-overexpressing Cancer Cells Promote M2-Like Polarization of Tumor-Associated Macrophages by Delivering MiR-21-Abundant Exosomes. Neoplasia 2018, 20, 775–788. [Google Scholar] [CrossRef] [PubMed]
  114. Gabrusiewicz, K.; Li, X.; Wei, J.; Hashimoto, Y.; Marisetty, A.L.; Ott, M.; Wang, F.; Hawke, D.; Yu, J.; Healy, L.M.; et al. Glioblastoma stem cell-derived exosomes induce M2 macrophages and PD-L1 expression on human monocytes. Oncoimmunology 2018, 7, e1412909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, F.; Li, B.; Wei, Y.; Zhao, Y.; Wang, L.; Zhang, P.; Yang, J.; He, W.; Chen, H.; Jiao, Z.; et al. Tumor-derived exosomes induce PD1(+) macrophage population in human gastric cancer that promotes disease progression. Oncogenesis 2018, 7, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Hoelzinger, D.B.; Quinton, S.J.; Walters, D.K.; Vardam-Kaur, T.; Tschumper, R.C.; Borges da Silva, H.; Jelinek, D.F. Extracellular vesicle proteomic analysis leads to the discovery of HDGF as a new factor in multiple myeloma biology. Blood Adv. 2022, 6, 3458–3471. [Google Scholar] [CrossRef] [PubMed]
  117. Challagundla, K.B.; Wise, P.M.; Neviani, P.; Chava, H.; Murtadha, M.; Xu, T.; Kennedy, R.; Ivan, C.; Zhang, X.; Vannini, I.; et al. Exosome-mediated transfer of microRNAs within the tumor microenvironment and neuroblastoma resistance to chemotherapy. J. Natl. Cancer Inst. 2015, 107, djv135. [Google Scholar] [CrossRef] [Green Version]
  118. Mpakali, A.; Stratikos, E. The Role of Antigen Processing and Presentation in Cancer and the Efficacy of Immune Checkpoint Inhibitor Immunotherapy. Cancers 2021, 13, 134. [Google Scholar] [CrossRef]
  119. Imbert, P.R.C.; Saric, A.; Pedram, K.; Bertozzi, C.R.; Grinstein, S.; Freeman, S.A. An Acquired and Endogenous Glycocalyx Forms a Bidirectional “Don’t Eat” and “Don’t Eat Me” Barrier to Phagocytosis. Curr. Biol. 2021, 31, 77–89.e5. [Google Scholar] [CrossRef]
  120. Wang, J.; Zeng, H.; Zhang, H.; Han, Y. The role of exosomal PD-L1 in tumor immunotherapy. Transl. Oncol. 2021, 14, 101047. [Google Scholar] [CrossRef]
  121. Theodoraki, M.N.; Yerneni, S.; Gooding, W.E.; Ohr, J.; Clump, D.A.; Bauman, J.E.; Ferris, R.L.; Whiteside, T.L. Circulating exosomes measure responses to therapy in head and neck cancer patients treated with cetuximab, ipilimumab, and IMRT. Oncoimmunology 2019, 8, 1593805. [Google Scholar] [CrossRef]
  122. Kato, T.; Fahrmann, J.F.; Hanash, S.M.; Vykoukal, J. Extracellular Vesicles Mediate B Cell Immune Response and Are a Potential Target for Cancer Therapy. Cells 2020, 9, 1518. [Google Scholar] [CrossRef] [PubMed]
  123. Aung, T.; Chapuy, B.; Vogel, D.; Wenzel, D.; Oppermann, M.; Lahmann, M.; Weinhage, T.; Menck, K.; Hupfeld, T.; Koch, R.; et al. Exosomal evasion of humoral immunotherapy in aggressive B-cell lymphoma modulated by ATP-binding cassette transporter A3. Proc. Natl. Acad. Sci. USA 2011, 108, 15336–15341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ciravolo, V.; Huber, V.; Ghedini, G.C.; Venturelli, E.; Bianchi, F.; Campiglio, M.; Morelli, D.; Villa, A.; Della Mina, P.; Menard, S.; et al. Potential role of HER2-overexpressing exosomes in countering trastuzumab-based therapy. J. Cell Physiol. 2012, 227, 658–667. [Google Scholar] [CrossRef] [PubMed]
  125. Togashi, Y.; Shitara, K.; Nishikawa, H. Regulatory T cells in cancer immunosuppression—Implications for anticancer therapy. Nat. Rev. Clin. Oncol. 2019, 16, 356–371. [Google Scholar] [CrossRef]
  126. De Simone, M.; Arrigoni, A.; Rossetti, G.; Gruarin, P.; Ranzani, V.; Politano, C.; Bonnal, R.J.P.; Provasi, E.; Sarnicola, M.L.; Panzeri, I.; et al. Transcriptional Landscape of Human Tissue Lymphocytes Unveils Uniqueness of Tumor-Infiltrating T Regulatory Cells. Immunity 2016, 45, 1135–1147. [Google Scholar] [CrossRef] [Green Version]
  127. Saito, T.; Nishikawa, H.; Wada, H.; Nagano, Y.; Sugiyama, D.; Atarashi, K.; Maeda, Y.; Hamaguchi, M.; Ohkura, N.; Sato, E.; et al. Two FOXP3(+)CD4(+) T cell subpopulations distinctly control the prognosis of colorectal cancers. Nat. Med. 2016, 22, 679–684. [Google Scholar] [CrossRef]
  128. Kidani, Y.; Nogami, W.; Yasumizu, Y.; Kawashima, A.; Tanaka, A.; Sonoda, Y.; Tona, Y.; Nashiki, K.; Matsumoto, R.; Hagiwara, M.; et al. CCR8-targeted specific depletion of clonally expanded Treg cells in tumor tissues evokes potent tumor immunity with long-lasting memory. Proc. Natl. Acad. Sci. USA 2022, 119, e2114282119. [Google Scholar] [CrossRef]
  129. Zhang, A.; Ren, Z.; Tseng, K.F.; Liu, X.; Li, H.; Lu, C.; Cai, Y.; Minna, J.D.; Fu, Y.X. Dual targeting of CTLA-4 and CD47 on Treg cells promotes immunity against solid tumors. Sci. Transl. Med. 2021, 13, eabg8693. [Google Scholar] [CrossRef]
  130. Klages, K.; Mayer, C.T.; Lahl, K.; Loddenkemper, C.; Teng, M.W.; Ngiow, S.F.; Smyth, M.J.; Hamann, A.; Huehn, J.; Sparwasser, T. Selective depletion of Foxp3+ regulatory T cells improves effective therapeutic vaccination against established melanoma. Cancer Res. 2010, 70, 7788–7799. [Google Scholar] [CrossRef] [Green Version]
  131. Shitara, K.; Nishikawa, H. Regulatory T cells: A potential target in cancer immunotherapy. Ann. N. Y. Acad. Sci. 2018, 1417, 104–115. [Google Scholar] [CrossRef]
  132. Watanabe, T. Realization of Osteolysis, Angiogenesis, Immunosuppression, and Drug Resistance by Extracellular Vesicles: Roles of RNAs and Proteins in Their Cargoes and of Ectonucleotidases of the Immunosuppressive Adenosinergic Noncanonical Pathway in the Bone Marrow Niche of Multiple Myeloma. Cancers 2021, 13, 2969. [Google Scholar] [CrossRef] [PubMed]
  133. Zebrowska, A.; Widlak, P.; Whiteside, T.; Pietrowska, M. Signaling of Tumor-Derived sEV Impacts Melanoma Progression. Int. J. Mol. Sci. 2020, 21, 5066. [Google Scholar] [CrossRef] [PubMed]
  134. Clayton, A.; Mitchell, J.P.; Court, J.; Mason, M.D.; Tabi, Z. Human tumor-derived exosomes selectively impair lymphocyte responses to interleukin-2. Cancer Res. 2007, 67, 7458–7466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Ning, T.; Li, J.; He, Y.; Zhang, H.; Wang, X.; Deng, T.; Liu, R.; Li, H.; Bai, M.; Fan, Q.; et al. Exosomal miR-208b related with oxaliplatin resistance promotes Treg expansion in colorectal cancer. Mol. Ther. 2021, 29, 2723–2736. [Google Scholar] [CrossRef]
  136. Szajnik, M.; Czystowska, M.; Szczepanski, M.J.; Mandapathil, M.; Whiteside, T.L. Tumor-derived microvesicles induce, expand and up-regulate biological activities of human regulatory T cells (Treg). PLoS ONE 2010, 5, e11469. [Google Scholar] [CrossRef] [Green Version]
  137. Maybruck, B.T.; Pfannenstiel, L.W.; Diaz-Montero, M.; Gastman, B.R. Tumor-derived exosomes induce CD8(+) T cell suppressors. J. Immunother. Cancer 2017, 5, 65. [Google Scholar] [CrossRef]
  138. Ye, S.B.; Zhang, H.; Cai, T.T.; Liu, Y.N.; Ni, J.J.; He, J.; Peng, J.Y.; Chen, Q.Y.; Mo, H.Y.; Jun, C.; et al. Exosomal miR-24-3p impedes T-cell function by targeting FGF11 and serves as a potential prognostic biomarker for nasopharyngeal carcinoma. J. Pathol. 2016, 240, 329–340. [Google Scholar] [CrossRef]
  139. Szczepanski, M.J.; Szajnik, M.; Czystowska, M.; Mandapathil, M.; Strauss, L.; Welsh, A.; Foon, K.A.; Whiteside, T.L.; Boyiadzis, M. Increased frequency and suppression by regulatory T cells in patients with acute myelogenous leukemia. Clin. Cancer Res. 2009, 15, 3325–3332. [Google Scholar] [CrossRef] [Green Version]
  140. Swatler, J.; Turos-Korgul, L.; Brewinska-Olchowik, M.; De Biasi, S.; Dudka, W.; Le, B.V.; Kominek, A.; Cyranowski, S.; Pilanc, P.; Mohammadi, E.; et al. 4-1BBL-containing leukemic extracellular vesicles promote immunosuppressive effector regulatory T cells. Blood Adv. 2022, 6, 1879–1894. [Google Scholar] [CrossRef]
  141. Carlsten, M.; Jaras, M. Natural Killer Cells in Myeloid Malignancies: Immune Surveillance, NK Cell Dysfunction, and Pharmacological Opportunities to Bolster the Endogenous NK Cells. Front. Immunol. 2019, 10, 2357. [Google Scholar] [CrossRef]
  142. Myers, J.A.; Miller, J.S. Exploring the NK cell platform for cancer immunotherapy. Nat. Rev. Clin. Oncol. 2021, 18, 85–100. [Google Scholar] [CrossRef] [PubMed]
  143. Rea, D.; Henry, G.; Khaznadar, Z.; Etienne, G.; Guilhot, F.; Nicolini, F.; Guilhot, J.; Rousselot, P.; Huguet, F.; Legros, L.; et al. Natural killer-cell counts are associated with molecular relapse-free survival after imatinib discontinuation in chronic myeloid leukemia: The IMMUNOSTIM study. Haematologica 2017, 102, 1368–1377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Hilpert, J.; Grosse-Hovest, L.; Grunebach, F.; Buechele, C.; Nuebling, T.; Raum, T.; Steinle, A.; Salih, H.R. Comprehensive analysis of NKG2D ligand expression and release in leukemia: Implications for NKG2D-mediated NK cell responses. J. Immunol. 2012, 189, 1360–1371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Boissel, N.; Rea, D.; Tieng, V.; Dulphy, N.; Brun, M.; Cayuela, J.M.; Rousselot, P.; Tamouza, R.; Le Bouteiller, P.; Mahon, F.X.; et al. BCR/ABL oncogene directly controls MHC class I chain-related molecule A expression in chronic myelogenous leukemia. J. Immunol. 2006, 176, 5108–5116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Clayton, A.; Mitchell, J.P.; Court, J.; Linnane, S.; Mason, M.D.; Tabi, Z. Human tumor-derived exosomes down-modulate NKG2D expression. J. Immunol. 2008, 180, 7249–7258. [Google Scholar] [CrossRef] [Green Version]
  147. Ashiru, O.; Boutet, P.; Fernandez-Messina, L.; Aguera-Gonzalez, S.; Skepper, J.N.; Vales-Gomez, M.; Reyburn, H.T. Natural killer cell cytotoxicity is suppressed by exposure to the human NKG2D ligand MICA*008 that is shed by tumor cells in exosomes. Cancer Res. 2010, 70, 481–489. [Google Scholar] [CrossRef] [Green Version]
  148. Szczepanski, M.J.; Szajnik, M.; Welsh, A.; Whiteside, T.L.; Boyiadzis, M. Blast-derived microvesicles in sera from patients with acute myeloid leukemia suppress natural killer cell function via membrane-associated transforming growth factor-beta1. Haematologica 2011, 96, 1302–1309. [Google Scholar] [CrossRef] [Green Version]
  149. Hong, C.S.; Sharma, P.; Yerneni, S.S.; Simms, P.; Jackson, E.K.; Whiteside, T.L.; Boyiadzis, M. Circulating exosomes carrying an immunosuppressive cargo interfere with cellular immunotherapy in acute myeloid leukemia. Sci. Rep. 2017, 7, 14684. [Google Scholar] [CrossRef] [Green Version]
  150. Hong, C.S.; Muller, L.; Whiteside, T.L.; Boyiadzis, M. Plasma exosomes as markers of therapeutic response in patients with acute myeloid leukemia. Front. Immunol. 2014, 5, 160. [Google Scholar] [CrossRef] [Green Version]
  151. Hong, C.S.; Funk, S.; Muller, L.; Boyiadzis, M.; Whiteside, T.L. Isolation of biologically active and morphologically intact exosomes from plasma of patients with cancer. J. Extracell. Vesicles 2016, 5, 29289. [Google Scholar] [CrossRef]
  152. Hong, C.S.; Jeong, E.; Boyiadzis, M.; Whiteside, T.L. Increased small extracellular vesicle secretion after chemotherapy via upregulation of cholesterol metabolism in acute myeloid leukaemia. J. Extracell. Vesicles 2020, 9, 1800979. [Google Scholar] [CrossRef] [PubMed]
  153. Binder, H.M.; Maeding, N.; Wolf, M.; Cronemberger Andrade, A.; Vari, B.; Krisch, L.; Gomes, F.G.; Blochl, C.; Muigg, K.; Poupardin, R.; et al. Scalable Enrichment of Immunomodulatory Human Acute Myeloid Leukemia Cell Line-Derived Extracellular Vesicles. Cells 2021, 10, 3321. [Google Scholar] [CrossRef] [PubMed]
  154. Berchem, G.; Noman, M.Z.; Bosseler, M.; Paggetti, J.; Baconnais, S.; Le Cam, E.; Nanbakhsh, A.; Moussay, E.; Mami-Chouaib, F.; Janji, B.; et al. Hypoxic tumor-derived microvesicles negatively regulate NK cell function by a mechanism involving TGF-beta and miR23a transfer. Oncoimmunology 2016, 5, e1062968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Graner, M.W.; Schnell, S.; Olin, M.R. Tumor-derived exosomes, microRNAs, and cancer immune suppression. Semin. Immunopathol. 2018, 40, 505–515. [Google Scholar] [CrossRef]
  156. Zhang, P.F.; Gao, C.; Huang, X.Y.; Lu, J.C.; Guo, X.J.; Shi, G.M.; Cai, J.B.; Ke, A.W. Cancer cell-derived exosomal circUHRF1 induces natural killer cell exhaustion and may cause resistance to anti-PD1 therapy in hepatocellular carcinoma. Mol. Cancer 2020, 19, 110. [Google Scholar] [CrossRef]
  157. Yang, Y.; Li, C.; Liu, T.; Dai, X.; Bazhin, A.V. Myeloid-Derived Suppressor Cells in Tumors: From Mechanisms to Antigen Specificity and Microenvironmental Regulation. Front. Immunol. 2020, 11, 1371. [Google Scholar] [CrossRef]
  158. Rodriguez, P.C.; Ochoa, A.C. Arginine regulation by myeloid derived suppressor cells and tolerance in cancer: Mechanisms and therapeutic perspectives. Immunol. Rev. 2008, 222, 180–191. [Google Scholar] [CrossRef] [Green Version]
  159. Tian, X.; Shen, H.; Li, Z.; Wang, T.; Wang, S. Tumor-derived exosomes, myeloid-derived suppressor cells, and tumor microenvironment. J. Hematol. Oncol. 2019, 12, 84. [Google Scholar] [CrossRef] [Green Version]
  160. Tucci, M.; Passarelli, A.; Mannavola, F.; Felici, C.; Stucci, L.S.; Cives, M.; Silvestris, F. Immune System Evasion as Hallmark of Melanoma Progression: The Role of Dendritic Cells. Front. Oncol. 2019, 9, 1148. [Google Scholar] [CrossRef] [Green Version]
  161. Wculek, S.K.; Cueto, F.J.; Mujal, A.M.; Melero, I.; Krummel, M.F.; Sancho, D. Dendritic cells in cancer immunology and immunotherapy. Nat. Rev. Immunol. 2020, 20, 7–24. [Google Scholar] [CrossRef]
  162. Dai, S.; Wei, D.; Wu, Z.; Zhou, X.; Wei, X.; Huang, H.; Li, G. Phase I clinical trial of autologous ascites-derived exosomes combined with GM-CSF for colorectal cancer. Mol. Ther. 2008, 16, 782–790. [Google Scholar] [CrossRef] [PubMed]
  163. Kumari, N.; Choi, S.H. Tumor-associated macrophages in cancer: Recent advancements in cancer nanoimmunotherapies. J. Exp. Clin. Cancer Res. 2022, 41, 68. [Google Scholar] [CrossRef] [PubMed]
  164. Christofides, A.; Strauss, L.; Yeo, A.; Cao, C.; Charest, A.; Boussiotis, V.A. The complex role of tumor-infiltrating macrophages. Nat. Immunol. 2022, 23, 1148–1156. [Google Scholar] [CrossRef] [PubMed]
  165. Cai, Z.; Yang, F.; Yu, L.; Yu, Z.; Jiang, L.; Wang, Q.; Yang, Y.; Wang, L.; Cao, X.; Wang, J. Activated T cell exosomes promote tumor invasion via Fas signaling pathway. J. Immunol. 2012, 188, 5954–5961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Beccard, I.J.; Hofmann, L.; Schroeder, J.C.; Ludwig, S.; Laban, S.; Brunner, C.; Lotfi, R.; Hoffmann, T.K.; Jackson, E.K.; Schuler, P.J.; et al. Immune Suppressive Effects of Plasma-Derived Exosome Populations in Head and Neck Cancer. Cancers 2020, 12, 1997. [Google Scholar] [CrossRef] [PubMed]
  167. Chiou, N.T.; Kageyama, R.; Ansel, K.M. Selective Export into Extracellular Vesicles and Function of tRNA Fragments during T Cell Activation. Cell Rep. 2018, 25, 3356–3370.e4. [Google Scholar] [CrossRef] [Green Version]
  168. Neviani, P.; Wise, P.M.; Murtadha, M.; Liu, C.W.; Wu, C.H.; Jong, A.Y.; Seeger, R.C.; Fabbri, M. Natural Killer-Derived Exosomal miR-186 Inhibits Neuroblastoma Growth and Immune Escape Mechanisms. Cancer Res. 2019, 79, 1151–1164. [Google Scholar] [CrossRef]
  169. Fenselau, C.; Ostrand-Rosenberg, S. Molecular cargo in myeloid-derived suppressor cells and their exosomes. Cell Immunol. 2021, 359, 104258. [Google Scholar] [CrossRef]
  170. Rashid, M.H.; Borin, T.F.; Ara, R.; Piranlioglu, R.; Achyut, B.R.; Korkaya, H.; Liu, Y.; Arbab, A.S. Critical immunosuppressive effect of MDSCderived exosomes in the tumor microenvironment. Oncol. Rep. 2021, 45, 1171–1181. [Google Scholar] [CrossRef]
  171. Lan, J.; Sun, L.; Xu, F.; Liu, L.; Hu, F.; Song, D.; Hou, Z.; Wu, W.; Luo, X.; Wang, J.; et al. M2 Macrophage-Derived Exosomes Promote Cell Migration and Invasion in Colon Cancer. Cancer Res. 2019, 79, 146–158. [Google Scholar] [CrossRef]
  172. Wang, X.; Shen, H.; He, Q.; Tian, W.; Xia, A.; Lu, X.J. Exosomes derived from exhausted CD8+ T cells impaired the anticancer function of normal CD8+ T cells. J. Med. Genet. 2019, 56, 29–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Pu, J.; Xu, Z.; Nian, J.; Fang, Q.; Yang, M.; Huang, Y.; Li, W.; Ge, B.; Wang, J.; Wei, H. M2 macrophage-derived extracellular vesicles facilitate CD8+T cell exhaustion in hepatocellular carcinoma via the miR-21-5p/YOD1/YAP/beta-catenin pathway. Cell Death Discov. 2021, 7, 182. [Google Scholar] [CrossRef] [PubMed]
  174. Mi, X.; Xu, R.; Hong, S.; Xu, T.; Zhang, W.; Liu, M. M2 Macrophage-Derived Exosomal lncRNA AFAP1-AS1 and MicroRNA-26a Affect Cell Migration and Metastasis in Esophageal Cancer. Mol. Ther. Nucleic Acids 2020, 22, 779–790. [Google Scholar] [CrossRef]
  175. Hu, W.; Liu, C.; Bi, Z.-Y.; Zhou, Q.; Zhang, H.; Li, L.-L.; Zhang, J.; Zhu, W.; Song, Y.-Y.Y.; Zhang, F.; et al. Comprehensive landscape of extracellular vesicle-derived RNAs in cancer initiation, progression, metastasis and cancer immunology. Mol. Cancer 2020, 19, 102. [Google Scholar] [CrossRef]
  176. Yang, P.; Peng, Y.; Feng, Y.; Xu, Z.; Feng, P.; Cao, J.; Chen, Y.; Chen, X.; Cao, X.; Yang, Y.; et al. Immune Cell-Derived Extracellular Vesicles—New Strategies in Cancer Immunotherapy. Front. Immunol. 2021, 12, 771551. [Google Scholar] [CrossRef] [PubMed]
  177. Andrews, D.W.; Judy, K.D.; Scott, C.B.; Garcia, S.; Harshyne, L.A.; Kenyon, L.; Talekar, K.; Flanders, A.; Atsina, K.B.; Kim, L.; et al. Phase Ib Clinical Trial of IGV-001 for Patients with Newly Diagnosed Glioblastoma. Clin. Cancer Res. 2021, 27, 1912–1922. [Google Scholar] [CrossRef] [PubMed]
  178. Dionisi, M.; De Archangelis, C.; Battisti, F.; Rahimi Koshkaki, H.; Belleudi, F.; Zizzari, I.G.; Ruscito, I.; Albano, C.; Di Filippo, A.; Torrisi, M.R.; et al. Tumor-Derived Microvesicles Enhance Cross-Processing Ability of Clinical Grade Dendritic Cells. Front. Immunol. 2018, 9, 2481. [Google Scholar] [CrossRef] [Green Version]
  179. Ma, J.; Wei, K.; Zhang, H.; Tang, K.; Li, F.; Zhang, T.; Liu, J.; Xu, P.; Yu, Y.; Sun, W.; et al. Mechanisms by Which Dendritic Cells Present Tumor Microparticle Antigens to CD8(+) T Cells. Cancer Immunol. Res. 2018, 6, 1057–1068. [Google Scholar] [CrossRef] [Green Version]
  180. Dassler-Plenker, J.; Reiners, K.S.; van den Boorn, J.G.; Hansen, H.P.; Putschli, B.; Barnert, S.; Schuberth-Wagner, C.; Schubert, R.; Tuting, T.; Hallek, M.; et al. RIG-I activation induces the release of extracellular vesicles with antitumor activity. Oncoimmunology 2016, 5, e1219827. [Google Scholar] [CrossRef]
  181. Gastpar, R.; Gehrmann, M.; Bausero, M.A.; Asea, A.; Gross, C.; Schroeder, J.A.; Multhoff, G. Heat shock protein 70 surface-positive tumor exosomes stimulate migratory and cytolytic activity of natural killer cells. Cancer Res. 2005, 65, 5238–5247. [Google Scholar] [CrossRef]
  182. Ortiz-Bonilla, C.J.; Uccello, T.P.; Gerber, S.A.; Lord, E.M.; Messing, E.M.; Lee, Y.F. Bladder Cancer Extracellular Vesicles Elicit a CD8 T Cell-Mediated Antitumor Immunity. Int. J. Mol. Sci. 2022, 23, 2904. [Google Scholar] [CrossRef]
  183. Montfort, A.; Bertrand, F.; Rochotte, J.; Gilhodes, J.; Filleron, T.; Milhes, J.; Dufau, C.; Imbert, C.; Riond, J.; Tosolini, M.; et al. Neutral Sphingomyelinase 2 Heightens Anti-Melanoma Immune Responses and Anti-PD-1 Therapy Efficacy. Cancer Immunol. Res. 2021, 9, 568–582. [Google Scholar] [CrossRef] [PubMed]
  184. Chen, Z.; You, L.; Wang, L.; Huang, X.; Liu, H.; Wei, J.Y.; Zhu, L.; Qian, W. Dual effect of DLBCL-derived EXOs in lymphoma to improve DC vaccine efficacy in vitro while favor tumorgenesis in vivo. J. Exp. Clin. Cancer Res. 2018, 37, 190. [Google Scholar] [CrossRef] [Green Version]
  185. Scholl, J.N.; de Fraga Dias, A.; Pizzato, P.R.; Lopes, D.V.; Moritz, C.E.J.; Jandrey, E.H.F.; Souto, G.D.; Colombo, M.; Rohden, F.; Sevigny, J.; et al. Characterization and antiproliferative activity of glioma-derived extracellular vesicles. Nanomedicine 2020, 15, 1001–1018. [Google Scholar] [CrossRef] [PubMed]
  186. Seo, N.; Shirakura, Y.; Tahara, Y.; Momose, F.; Harada, N.; Ikeda, H.; Akiyoshi, K.; Shiku, H. Activated CD8(+) T cell extracellular vesicles prevent tumour progression by targeting of lesional mesenchymal cells. Nat. Commun. 2018, 9, 435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Lu, Z.; Zuo, B.; Jing, R.; Gao, X.; Rao, Q.; Liu, Z.; Qi, H.; Guo, H.; Yin, H. Dendritic cell-derived exosomes elicit tumor regression in autochthonous hepatocellular carcinoma mouse models. J. Hepatol. 2017, 67, 739–748. [Google Scholar] [CrossRef] [PubMed]
  188. Enomoto, Y.; Li, P.; Jenkins, L.M.; Anastasakis, D.; Lyons, G.C.; Hafner, M.; Leonard, W.J. Cytokine-enhanced cytolytic activity of exosomes from NK Cells. Cancer Gene Ther. 2022, 29, 734–749. [Google Scholar] [CrossRef]
  189. Shoae-Hassani, A.; Hamidieh, A.A.; Behfar, M.; Mohseni, R.; Mortazavi-Tabatabaei, S.A.; Asgharzadeh, S. NK Cell-derived Exosomes From NK Cells Previously Exposed to Neuroblastoma Cells Augment the Antitumor Activity of Cytokine-activated NK Cells. J. Immunother. 2017, 40, 265–276. [Google Scholar] [CrossRef]
  190. Zhou, W.J.; Zhang, J.; Xie, F.; Wu, J.N.; Ye, J.F.; Wang, J.; Wu, K.; Li, M.Q. CD45RO(-)CD8(+) T cell-derived exosomes restrict estrogen-driven endometrial cancer development via the ERbeta/miR-765/PLP2/Notch axis. Theranostics 2021, 11, 5330–5345. [Google Scholar] [CrossRef]
  191. DeNardo, D.G.; Ruffell, B. Macrophages as regulators of tumour immunity and immunotherapy. Nat. Rev. Immunol. 2019, 19, 369–382. [Google Scholar] [CrossRef]
  192. Cianciaruso, C.; Beltraminelli, T.; Duval, F.; Nassiri, S.; Hamelin, R.; Mozes, A.; Gallart-Ayala, H.; Ceada Torres, G.; Torchia, B.; Ries, C.H.; et al. Molecular Profiling and Functional Analysis of Macrophage-Derived Tumor Extracellular Vesicles. Cell Rep. 2019, 27, 3062–3080.e11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Zitvogel, L.; Regnault, A.; Lozier, A.; Wolfers, J.; Flament, C.; Tenza, D.; Ricciardi-Castagnoli, P.; Raposo, G.; Amigorena, S. Eradication of established murine tumors using a novel cell-free vaccine: Dendritic cell-derived exosomes. Nat. Med. 1998, 4, 594–600. [Google Scholar] [CrossRef] [PubMed]
  194. Hadla, M.; Palazzolo, S.; Corona, G.; Caligiuri, I.; Canzonieri, V.; Toffoli, G.; Rizzolio, F. Exosomes increase the therapeutic index of doxorubicin in breast and ovarian cancer mouse models. Nanomedicine 2016, 11, 2431–2441. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, D.X.; Vu, L.T.; Ismail, N.N.; Le, M.T.N.; Grimson, A. Landscape of extracellular vesicles in the tumour microenvironment: Interactions with stromal cells and with non-cell components, and impacts on metabolic reprogramming, horizontal transfer of neoplastic traits, and the emergence of therapeutic resistance. Semin. Cancer Biol. 2021, 74, 24–44. [Google Scholar] [CrossRef] [PubMed]
  196. Khani, A.T.; Sharifzad, F.; Mardpour, S.; Hassan, Z.M.; Ebrahimi, M. Tumor extracellular vesicles loaded with exogenous Let-7i and miR-142 can modulate both immune response and tumor microenvironment to initiate a powerful anti-tumor response. Cancer Lett. 2021, 501, 200–209. [Google Scholar] [CrossRef]
  197. Reale, A.; Khong, T.; Mithraprabhu, S.; Spencer, A. Translational Potential of RNA Derived From Extracellular Vesicles in Multiple Myeloma. Front. Oncol. 2021, 11, 718502. [Google Scholar] [CrossRef]
  198. Desantis, V.; Saltarella, I.; Lamanuzzi, A.; Melaccio, A.; Solimando, A.G.; Mariggio, M.A.; Racanelli, V.; Paradiso, A.; Vacca, A.; Frassanito, M.A. MicroRNAs-Based Nano-Strategies as New Therapeutic Approach in Multiple Myeloma to Overcome Disease Progression and Drug Resistance. Int. J. Mol. Sci. 2020, 21, 3084. [Google Scholar] [CrossRef]
  199. Burnett, J.C.; Rossi, J.J. RNA-based therapeutics: Current progress and future prospects. Chem. Biol. 2012, 19, 60–71. [Google Scholar] [CrossRef] [Green Version]
  200. Giacobino, C.; Canta, M.; Fornaguera, C.; Borros, S.; Cauda, V. Extracellular Vesicles and Their Current Role in Cancer Immunotherapy. Cancers 2021, 13, 2280. [Google Scholar] [CrossRef]
  201. Zhang, L.; Sun, M.; He, Z.; Sun, J.; Li, H.; Luo, Q. Multi-functional extracellular vesicles: Potentials in cancer immunotherapy. Cancer Lett. 2022, 551, 215934. [Google Scholar] [CrossRef]
  202. Huang, F.; Wan, J.; Hu, W.; Hao, S. Enhancement of Anti-Leukemia Immunity by Leukemia-Derived Exosomes Via Downregulation of TGF-beta1 Expression. Cell Physiol. Biochem. 2017, 44, 240–254. [Google Scholar] [CrossRef] [PubMed]
  203. Huang, F.; Wan, J.; Hao, S.; Deng, X.; Chen, L.; Ma, L. TGF-beta1-silenced leukemia cell-derived exosomes target dendritic cells to induce potent anti-leukemic immunity in a mouse model. Cancer Immunol. Immunother. 2017, 66, 1321–1331. [Google Scholar] [CrossRef] [PubMed]
  204. Dang, X.T.T.; Kavishka, J.M.; Zhang, D.X.; Pirisinu, M.; Le, M.T.N. Extracellular Vesicles as an Efficient and Versatile System for Drug Delivery. Cells 2020, 9, 2191. [Google Scholar] [CrossRef] [PubMed]
  205. Huang, L.; Rong, Y.; Tang, X.; Yi, K.; Qi, P.; Hou, J.; Liu, W.; He, Y.; Gao, X.; Yuan, C.; et al. Engineered exosomes as an in situ DC-primed vaccine to boost antitumor immunity in breast cancer. Mol. Cancer 2022, 21, 45. [Google Scholar] [CrossRef]
  206. Koh, E.; Lee, E.J.; Nam, G.H.; Hong, Y.; Cho, E.; Yang, Y.; Kim, I.S. Exosome-SIRPalpha, a CD47 blockade increases cancer cell phagocytosis. Biomaterials 2017, 121, 121–129. [Google Scholar] [CrossRef]
  207. Hong, Y.; Kim, Y.K.; Kim, G.B.; Nam, G.H.; Kim, S.A.; Park, Y.; Yang, Y.; Kim, I.S. Degradation of tumour stromal hyaluronan by small extracellular vesicle-PH20 stimulates CD103(+) dendritic cells and in combination with PD-L1 blockade boosts anti-tumour immunity. J. Extracell. Vesicles 2019, 8, 1670893. [Google Scholar] [CrossRef] [Green Version]
  208. Cheng, Q.; Shi, X.; Han, M.; Smbatyan, G.; Lenz, H.J.; Zhang, Y. Reprogramming Exosomes as Nanoscale Controllers of Cellular Immunity. J. Am. Chem. Soc. 2018, 140, 16413–16417. [Google Scholar] [CrossRef]
  209. McAndrews, K.M.; Che, S.P.Y.; LeBleu, V.S.; Kalluri, R. Effective delivery of STING agonist using exosomes suppresses tumor growth and enhances antitumor immunity. J. Biol. Chem. 2021, 296, 100523. [Google Scholar] [CrossRef]
  210. Chen, S.; Lv, M.; Fang, S.; Ye, W.; Gao, Y.; Xu, Y. Poly(I:C) enhanced anti-cervical cancer immunities induced by dendritic cells-derived exosomes. Int. J. Biol. Macromol. 2018, 113, 1182–1187. [Google Scholar] [CrossRef]
  211. Zhao, Z.; McGill, J.; Gamero-Kubota, P.; He, M. Microfluidic on-demand engineering of exosomes towards cancer immunotherapy. Lab Chip 2019, 19, 1877–1886. [Google Scholar] [CrossRef]
  212. Lee, H.; Park, H.; Noh, G.J.; Lee, E.S. pH-responsive hyaluronate-anchored extracellular vesicles to promote tumor-targeted drug delivery. Carbohydr. Polym. 2018, 202, 323–333. [Google Scholar] [CrossRef] [PubMed]
  213. Zhu, L.; Kalimuthu, S.; Gangadaran, P.; Oh, J.M.; Lee, H.W.; Baek, S.H.; Jeong, S.Y.; Lee, S.W.; Lee, J.; Ahn, B.C. Exosomes Derived From Natural Killer Cells Exert Therapeutic Effect in Melanoma. Theranostics 2017, 7, 2732–2745. [Google Scholar] [CrossRef]
  214. Zhu, L.; Gangadaran, P.; Kalimuthu, S.; Oh, J.M.; Baek, S.H.; Jeong, S.Y.; Lee, S.W.; Lee, J.; Ahn, B.C. Novel alternatives to extracellular vesicle-based immunotherapy—Exosome mimetics derived from natural killer cells. Artif. Cells Nanomed. Biotechnol. 2018, 46, S166–S179. [Google Scholar] [CrossRef] [Green Version]
  215. Atai, N.A.; Balaj, L.; van Veen, H.; Breakefield, X.O.; Jarzyna, P.A.; Van Noorden, C.J.; Skog, J.; Maguire, C.A. Heparin blocks transfer of extracellular vesicles between donor and recipient cells. J. Neurooncol. 2013, 115, 343–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Wu, X.; Kang, M.; Wang, D.; Zhu, M.; Hu, Y.; Zhang, Y.; Deng, C.; Chen, J.; Teng, L. Heparan sulfate analogues regulate tumor-derived exosome formation that attenuates exosome functions in tumor processes. Int. J. Biol. Macromol. 2021, 187, 481–491. [Google Scholar] [CrossRef] [PubMed]
  217. Sento, S.; Sasabe, E.; Yamamoto, T. Application of a Persistent Heparin Treatment Inhibits the Malignant Potential of Oral Squamous Carcinoma Cells Induced by Tumor Cell-Derived Exosomes. PLoS ONE 2016, 11, e0148454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Mithraprabhu, S.; Chen, M.; Savvidou, I.; Reale, A.; Spencer, A. Liquid biopsy: An evolving paradigm for the biological characterisation of plasma cell disorders. Leukemia 2021, 35, 2771–2783. [Google Scholar] [CrossRef]
  219. Ostheimer, C.; Gunther, S.; Bache, M.; Vordermark, D.; Multhoff, G. Dynamics of Heat Shock Protein 70 Serum Levels As a Predictor of Clinical Response in Non-Small-Cell Lung Cancer and Correlation with the Hypoxia-Related Marker Osteopontin. Front. Immunol. 2017, 8, 1305. [Google Scholar] [CrossRef] [Green Version]
  220. McNulty, S.; Colaco, C.A.; Blandford, L.E.; Bailey, C.R.; Baschieri, S.; Todryk, S. Heat-shock proteins as dendritic cell-targeting vaccines—Getting warmer. Immunology 2013, 139, 407–415. [Google Scholar] [CrossRef]
  221. Madden, E.C.; Gorman, A.M.; Logue, S.E.; Samali, A. Tumour Cell Secretome in Chemoresistance and Tumour Recurrence. Trends Cancer 2020, 6, 489–505. [Google Scholar] [CrossRef]
  222. Xu, R.; Greening, D.W.; Chen, M.; Rai, A.; Ji, H.; Takahashi, N.; Simpson, R.J. Surfaceome of Exosomes Secreted from the Colorectal Cancer Cell Line SW480: Peripheral and Integral Membrane Proteins Analyzed by Proteolysis and TX114. Proteomics 2019, 19, e1700453. [Google Scholar] [CrossRef] [PubMed]
  223. Veerman, R.E.; Teeuwen, L.; Czarnewski, P.; Gucluler Akpinar, G.; Sandberg, A.; Cao, X.; Pernemalm, M.; Orre, L.M.; Gabrielsson, S.; Eldh, M. Molecular evaluation of five different isolation methods for extracellular vesicles reveals different clinical applicability and subcellular origin. J. Extracell. Vesicles 2021, 10, e12128. [Google Scholar] [CrossRef] [PubMed]
  224. Crescitelli, R.; Lasser, C.; Lotvall, J. Isolation and characterization of extracellular vesicle subpopulations from tissues. Nat. Protoc. 2021, 16, 1548–1580. [Google Scholar] [CrossRef] [PubMed]
  225. Rai, A.; Fang, H.; Fatmous, M.; Claridge, B.; Poh, Q.H.; Simpson, R.J.; Greening, D.W. A Protocol for Isolation, Purification, Characterization, and Functional Dissection of Exosomes. Methods Mol. Biol. 2021, 2261, 105–149. [Google Scholar] [CrossRef] [PubMed]
  226. Erdbrugger, U.; Blijdorp, C.J.; Bijnsdorp, I.V.; Borras, F.E.; Burger, D.; Bussolati, B.; Byrd, J.B.; Clayton, A.; Dear, J.W.; Falcon-Perez, J.M.; et al. Urinary extracellular vesicles: A position paper by the Urine Task Force of the International Society for Extracellular Vesicles. J. Extracell. Vesicles 2021, 10, e12093. [Google Scholar] [CrossRef] [PubMed]
  227. Yu, D.; Li, Y.; Wang, M.; Gu, J.; Xu, W.; Cai, H.; Fang, X.; Zhang, X. Exosomes as a new frontier of cancer liquid biopsy. Mol. Cancer 2022, 21, 56. [Google Scholar] [CrossRef] [PubMed]
  228. Ludwig, N.; Hong, C.S.; Ludwig, S.; Azambuja, J.H.; Sharma, P.; Theodoraki, M.N.; Whiteside, T.L. Isolation and Analysis of Tumor-Derived Exosomes. Curr. Protoc. Immunol. 2019, 127, e91. [Google Scholar] [CrossRef] [PubMed]
  229. Claridge, B.; Lozano, J.; Poh, Q.H.; Greening, D.W. Development of Extracellular Vesicle Therapeutics: Challenges, Considerations, and Opportunities. Front. Cell Dev. Biol. 2021, 9, 734720. [Google Scholar] [CrossRef]
  230. Rohde, E.; Pachler, K.; Gimona, M. Manufacturing and characterization of extracellular vesicles from umbilical cord-derived mesenchymal stromal cells for clinical testing. Cytotherapy 2019, 21, 581–592. [Google Scholar] [CrossRef]
  231. Harrison, S.J.; Perrot, A.; Alegre, A.; Simpson, D.; Wang, M.C.; Spencer, A.; Delimpasi, S.; Hulin, C.; Sunami, K.; Facon, T.; et al. Subgroup analysis of ICARIA-MM study in relapsed/refractory multiple myeloma patients with high-risk cytogenetics. Br. J. Haematol. 2021, 194, 120–131. [Google Scholar] [CrossRef]
Figure 1. Tumor-derived EVs regulate immune system cells’ functions. TEVs can promote tumor progression by suppression of innate and adaptive immune cells, as indicated in the left green panel. In fact, TEVs hamper effective antigen cross presentation in DCs, contribute to T cell dysfunction through check point inhibition, polarize mononuclear cells towards an immunosuppressive phenotype, induce immunosuppressive properties of MDSCs, promote Treg/Breg expansion and reduce NK cells cytotoxicity activity. TEVs can also promote tumor recognition enhancing anti-tumor immune functions, as shown in the right blue panel. During a functional immune response, TEVs deliver TAAs to NK cells to stimulate their cytotoxic function or to DCs to indirectly stimulate cytotoxic T cells, to polarize mononuclear cells towards an immunosuppressive phenotype. Macrophages can actively uptake TEVs from circulation promoting cancer cell recognition. Examples of molecules involved in these mechanisms are also indicated. (EVs, tumor-derived extracellular vesicles; TAAs, tumor-associated antigens; NK, natural killer; Treg, regulatory T cells; MDSC, myeloid-derived suppressor cells; DCs, dendritic cells; sNMase, neutral sphingomyelinase; HSP70, heat shock protein 70; MUC1, mucin-1; BAG6, large proline-rich protein BAG6; PD-L1, programmed death-ligand 1; FAS-L, tumor necrosis factor ligand superfamily, member 6; ARG1, arginase-1; TGF-β, transforming growth factor beta; LGALS9, galectin 9; MIF, macrophage migration inhibitory factor; SPHK1, sphingosine kinase 1, PGE2, prostaglandin E2). Figure generated with the aid of BioRender.com (adapted from “Pro- and Anti-Tumor Immune Cells in the Tumor Microenvironment biorender”, accessed on 15 November 2022).
Figure 1. Tumor-derived EVs regulate immune system cells’ functions. TEVs can promote tumor progression by suppression of innate and adaptive immune cells, as indicated in the left green panel. In fact, TEVs hamper effective antigen cross presentation in DCs, contribute to T cell dysfunction through check point inhibition, polarize mononuclear cells towards an immunosuppressive phenotype, induce immunosuppressive properties of MDSCs, promote Treg/Breg expansion and reduce NK cells cytotoxicity activity. TEVs can also promote tumor recognition enhancing anti-tumor immune functions, as shown in the right blue panel. During a functional immune response, TEVs deliver TAAs to NK cells to stimulate their cytotoxic function or to DCs to indirectly stimulate cytotoxic T cells, to polarize mononuclear cells towards an immunosuppressive phenotype. Macrophages can actively uptake TEVs from circulation promoting cancer cell recognition. Examples of molecules involved in these mechanisms are also indicated. (EVs, tumor-derived extracellular vesicles; TAAs, tumor-associated antigens; NK, natural killer; Treg, regulatory T cells; MDSC, myeloid-derived suppressor cells; DCs, dendritic cells; sNMase, neutral sphingomyelinase; HSP70, heat shock protein 70; MUC1, mucin-1; BAG6, large proline-rich protein BAG6; PD-L1, programmed death-ligand 1; FAS-L, tumor necrosis factor ligand superfamily, member 6; ARG1, arginase-1; TGF-β, transforming growth factor beta; LGALS9, galectin 9; MIF, macrophage migration inhibitory factor; SPHK1, sphingosine kinase 1, PGE2, prostaglandin E2). Figure generated with the aid of BioRender.com (adapted from “Pro- and Anti-Tumor Immune Cells in the Tumor Microenvironment biorender”, accessed on 15 November 2022).
Jcm 11 06892 g001
Table 1. TEV-mediated tumor immune suppression.
Table 1. TEV-mediated tumor immune suppression.
Cancer Type (EV-Source)Functional MoleculesRecipient/Target CellsFunctional Effect in Recipient CellsReferences
Small cell lung cancer PD-L1T cellsInhibition of T cell activation [53]
Ovarian cancer ARG1T cellsSuppression of T cells proliferation[48]
Ovarian cancer SPHK1T cellsPromotion of T cell exhaustion[54]
Breast cancer TβRIIT cellsCD8+ T cell exhaustion via SMAD3 and TCF1[55]
MelanomamiR-3187-3p, miR-498, miR-122, miR149, miR-181a/bT cellsInhibition of T cell activation[56]
LeukemiamiR-19a-3pT cellsSuppression of T cells immune function[57]
Melanoma and squamous cell carcinoma of head and neckFasLT cellsApoptosis of CD8+ T cells and Treg expansion[58]
Head and neck cancerPD-L1CD8+ T cellsDecreased CD8+ T cell activation[59]
Glioblastoma cancer PD-L1T cellsSuppressed T cell activation and proliferation[47]
Breast cancerPD-L1T cellsImpaired activation and cancer killing potential of T cells[60]
Prostate and melanomaPD-L1CD8+ T cellsSuppressed T cell activity[61]
Non small cell lung cancerPD-L1CD8+ T cellsImmunosuppressive properties[62]
MelanomaPD-L1T cellsSuppressed T cell functions[52,63]
Gastric cancerPD-L1T cellsApoptosis of T cells; reduced activation of peripheral blood mononuclear cells[64]
Hepatocellular carcinoma14-3-3ζT cellsImmunosuppressive phenotype[65]
Ovarian cancerGanglioside GD3T cellsDecreased T cell activation[66]
Ovarian cancer T cellsSuppressed T cell functions[67]
Melanoma CD8+ T cells and NK cellsEnhanced apoptosis and suppressed proliferation and activation of CD8+ T cells; decreased NKG2D in NK cells[68]
Head and neck cancer T cells and NK cellsEnhanced apoptosis of CD8+ T cells; suppressed proliferation of CD4+ T cells; decreased NKG2D in NK cells[46]
Melanoma cancermiRNAsCD4+ T cellsEnhanced apoptosis of CD4+ T cells[69]
Non-small cell lung cancermutant KRAS DNACD4+ T cellsConversion of naïve CD4+ T cells into Treg-like cells[70]
Breast cancerlncRNA SNHG16T cellsInduction of CD73+γδ1 Tregs[71]
Bladdre, colorectal, prostate, breast cancerCD39, CD73T cellsHydrolysis of ATP and generation of adenosine[72]
Hepatocellular carcinomaHMGB1B cellsExpansion of Bregs via TLR2/4-MAPK signaling pathway[73]
Pancreatic cancer TAAsB cellsInhibition of complement-dependent and antibody-dependent cell-mediated cytotoxicity[74]
Lung cancermiR-214TregSuppression of PTEN; Treg expansion[75]
Colorectal cancermiR-208bTregTreg expansion by targeting PDCD4, tumour growth and drug resistance[76]
Breast cancerlncRNA SNHG16Vδ1 T cellsPotentiation of the TGF-β1/SMAD5 pathway to upregulate CD73 expression in Vδ1 T cells[77]
Ovarian and head and neck squamous cancers TregIncreased FasL, IL-10, TGF-β1, CTLA-4, granzyme B and perforin expression and Treg-mediated stronger suppression of T cell proliferation[78]
Nasopharyngeal carcinomamiR-24-3p; miR-891a; miR-106a-5p; miR-20a-5p; miR-1908TregDownregulated ERK/STAT1/STAT3 phosphorylation with a shift of T cells towards Treg phenotype[79]
Leukemia4-1BBL/CD137LTregElevated expression of effector/tumour Treg markers (CD39, CCR8, CD30, TNFR2, CCR4, TIGIT, IL21R)[80]
Pancreatic ductal adenocarcinomaTGF-β1NK cellsDown-regulation of NKG2D, CD107a, TNF-α, INF-γ, CD71, CD98; impaired glucose uptake ability; attenuated NK cell cytotoxic activity[81]
Hypoxic lung adenocarcinoma cellsmiR-150-5pNK cellsDown-regulation of CD226 and functional repression of NK cells[82]
Pancreatic cancermiR-212-3pDCsInduced immune tolerance via RFXAP[83]
Pancreatic cancermiR-203DCsInhibition of antigen presentation via TLR4/TNF-α/IL-12 [84]
Glioblastoma multiforme LGALS9DCsInhibition of antigen recognition, processing and presentation[85]
Oral and oropharyngeal squamous cell carcinomamiR-17-5p, miR-21, miR-16, miR-24, miR-181a, miR-23bDCsImpaired differentiation and maturation of mono-DCs[86]
Prostate cancerPGE2DCsDisrupted cytokine production with inhibition of T cell activation, and increased secretion of adenosine with impaired DC functions and direct pro-tumour effects[87]
Chronic lymphocytic leukaemia and breast cancer DCsInhibited DC maturation[88]
Melanoma miR-155, miR-125b, miR-100, miR-146a, miR-146b, let-7e, miR-125a, and miR-99bMDSCsInduced immunosuppressive properties[89]
Melanoma HSP86MDSCsTLR4 and NFkB activation on MDSCs, generation of PD-L1+CD11b+Gr1+ MDSCs that suppress T cell functionality[90]
Thymoma, mammary carcinoma, and colon carcinomaHsp72MDSCsImmunosuppressive signaling via TLR2/STAT3 axis[76]
Renal cancerHSP70MDSCsMDSC proliferation and activation via TLR2 signaling to promote tumour growth and immunosuppression[77]
Breast cancermiR-9, miR-181aMDSCsActivated JAK/STAT signaling in eMDSCs through the targeting of SOCS3 and PIAS3[78]
Breast cancerTGF-β1, PGE2MDSCsMDSC accumulation and accelerated tumour growth[91]
Hypoxic glioma cancer cellsmiR-10a, miR-21MDSCsPromoted MDSCs expansion and activation through miR-10a/Rora/IκBα/NF-κB and miR-21/Pten/PI3K/AKT pathways[79]
Oral squamous cell carcinomamiR-21MDSCs and γδ T cellsInhibited γδ T cell functions through MDSCs[92]
Hypoxic pancreatic cancer cellsmiR-301aMacrophagesM2 polarization via PTEN/PI3Kgamma[80]
Lung cancermiR-103aMacrophagesM2 polarization[93]
Lung cancerEGFRMacrophagesLower host innate antiviral immunity through MEKK2/IRF3 axis[94]
Melanoma, lung and squamous skin cancersLet-7aMacrophagesIncreased OXPHOS activity and TAM via AKT/mTOR[95]
p53-mutant cancer cellsmiR-1246MacrophagesPromoted TAM phenotype[96]
MelanomamiR-125b-5pMacrophagesPromoted survival via LIPA[97]
Hypoxic pancreatic cancer cellsmiR-301a-3pMacrophagesInduced M2 phenotype via PTEN/PI3K signaling[80]
Pancreatic cancerArachidonic acidMacrophagesPromoted M2 phenotype[98]
Pancreatic cancer cells undergoing ferroptosisKRASG12D proteinMacrophagesPromoted M2 polarization through STAT3-dependent fatty acid oxidation[99]
Pancreatic cancermiR-155-5pMacrophagesTAM formation via EHF/Akt/NF-kB axis[100]
Metastatic osteosarcoma cells MacrophagesInduced M2 polarization and impaired phagocytosis, efferocytosis, and macrophage-dependent tumour cell killing[101]
Non-small cell lung cancer MacrophagesInduced M2 polarization and via PD-L1/HIF1α[102]
Colorectal cancermiR-25-3p, miR-130b-3p, miR-425-5pMacrophagesInduced M2 polarization through suppression of PTEN and activation of PI3K/Akt signaling and contributed to the establishment of liver metastasis[103]
Colorectal cancer MacrophagesMixed M1/M2 secretion pattern[104]
Breast cancergp130MacrophagesPromoted activation of STAT3 signaling, and enhanced the levels of protumourigenic cytokines and the survival of macrophages[105]
Liver cancer cells undergoing ER stress MacrophagesPromoted secretion of IL-6, MCP-1, IL-10 and TNF-α in macrophages through STAT3 signaling[106]
Breast cancer cells undergoing ER stressmiR-27a-3pMacrophagesIncrease in PD-L1 expression in macrophages[107]
Breast cancer MacrophagesActivation of NF-κB pathway in macrophages, and enhanced levels of IL-6, TNF-α, GCSF, and CCL2 in a TLR2-dependent manner[108]
Gastric cancer MacrophagesActivation of NF-κB and expression of the proinflammatory factors IL-6, TNF-α, and CCL2 in macrophages to promote tumour progression[109]
Lung cancermiR-21, miR-29aMacrophagesPro-metastatic inflammatory response by serving as ligands of TLR receptors in macrophages; NF-κB activation and increasing the secretion of IL-6 and TNF-α[110]
Hypoxic ovarian cancer cellsmiR-21-3p, miR-125b-5p, miR-181d-5pMacrophagesDifferentiation into TAM[111]
Chronic lymphocytic leukemia (CLL)Non coding Y RNA hY4MonocytesIncreased PD-L1[49]
Pancreatic cancer MonocytesDecreased HLA-DR expression, induced arginase and ROS[112]
Snail-expressing head and neck squamous cell carcinomamiR-21MonocytesM2 polarization[113]
Glioblastoma stem cells MonocytesM2 polarization and enhanced PD-L1 expression[114]
Gastric cancer MonocytesMonocyte differentiation into PD-1+ TAMs, which suppress CD8+ T cell functions[115]
Multiple myelomaHDGFMonocytesMDSCs with suppressive functions[116]
NeuroblastomamiR-21MonocytesTrigger TLR8 with increased miR-155 which transfer to tumour cells leads to drug resistance[117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Reale, A.; Khong, T.; Spencer, A. Extracellular Vesicles and Their Roles in the Tumor Immune Microenvironment. J. Clin. Med. 2022, 11, 6892. https://doi.org/10.3390/jcm11236892

AMA Style

Reale A, Khong T, Spencer A. Extracellular Vesicles and Their Roles in the Tumor Immune Microenvironment. Journal of Clinical Medicine. 2022; 11(23):6892. https://doi.org/10.3390/jcm11236892

Chicago/Turabian Style

Reale, Antonia, Tiffany Khong, and Andrew Spencer. 2022. "Extracellular Vesicles and Their Roles in the Tumor Immune Microenvironment" Journal of Clinical Medicine 11, no. 23: 6892. https://doi.org/10.3390/jcm11236892

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