Next Article in Journal
Potent Anticancer Activity of CXCR4-Targeted Nanostructured Toxins in Aggressive Endometrial Cancer Models
Next Article in Special Issue
Non-Coding RNAs Derived from Extracellular Vesicles Promote Pre-Metastatic Niche Formation and Tumor Distant Metastasis
Previous Article in Journal
Sequential and Hybrid PET/MRI Acquisition in Follow-Up Examination of Glioblastoma Show Similar Diagnostic Performance
Previous Article in Special Issue
Exosomes from EGFR-Mutated Adenocarcinoma Induce a Hybrid EMT and MMP9-Dependant Tumor Invasion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tumor-Derived Extracellular Vesicles in Cancer Immunoediting and Their Potential as Oncoimmunotherapeutics

by
Meysam Najaflou
1,2,
Mehdi Shahgolzari
3,
Ahmad Yari Khosroushahi
1,2,* and
Steven Fiering
4,5,*
1
Department of Medical Nanotechnology, Faculty of Advanced Medical Science, Tabriz University of Medical Sciences, Tabriz 51666-14766, Iran
2
Drug Applied Research Center, Tabriz University of Medical Sciences, Tabriz 51666-14766, Iran
3
Dental Research Center, Hamadan University of Medical Sciences, Hamadan 65175-4171, Iran
4
Department of Microbiology and Immunology, Geisel School of Medicine at Dartmouth, Hanover, NH 03755, USA
5
Dartmouth Cancer Center, Geisel School of Medicine at Dartmouth and Dartmouth-Hitchcock Medical Center, Lebanon, NH 03756, USA
*
Authors to whom correspondence should be addressed.
Submission received: 11 November 2022 / Revised: 16 December 2022 / Accepted: 20 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue Extracellular Vesicle in Cancer Biology)

Abstract

:

Simple Summary

Tumor cell-derived extracellular vesicles (TEVs) are an important means of tumor communication with, and manipulation of, the patient’s physiology. TEVs influence the local tumor environment as well as the systemic conditions of the patient. Progressive changes in tumor interactions with the host immune system are defined as “immunoediting”. Here, we summarize TEV effects on the immune system during the stages of cancer immunoediting and outline the molecular and cellular characteristics of interactions that result in complete tumor regression versus tumor immune escape and progression. Generally, the cargo profile of TEVs naturally changes during immunoediting toward immunosuppression while different cell stress or treatment conditions can inhibit this process or even reverse it to immunostimulation by altering the TEVs cargos. Therefore, understanding potential immunotherapeutic properties and how they can be manipulated to treat cancer should be considered a new research approach in oncoimmunotherapy.

Abstract

The tumor microenvironment (TME) within and around a tumor is a complex interacting mixture of tumor cells with various stromal cells, including endothelial cells, fibroblasts, and immune cells. In the early steps of tumor formation, the local microenvironment tends to oppose carcinogenesis, while with cancer progression, the microenvironment skews into a protumoral TME and the tumor influences stromal cells to provide tumor-supporting functions. The creation and development of cancer are dependent on escape from immune recognition predominantly by influencing stromal cells, particularly immune cells, to suppress antitumor immunity. This overall process is generally called immunoediting and has been categorized into three phases; elimination, equilibrium, and escape. Interaction of tumor cells with stromal cells in the TME is mediated generally by cell-to-cell contact, cytokines, growth factors, and extracellular vesicles (EVs). The least well studied are EVs (especially exosomes), which are nanoparticle-sized bilayer membrane vesicles released by many cell types that participate in cell/cell communication. EVs carry various proteins, nucleic acids, lipids, and small molecules that influence cells that ingest the EVs. Tumor-derived extracellular vesicles (TEVs) play a significant role in every stage of immunoediting, and their cargoes change from immune-activating in the early stages of immunoediting into immunosuppressing in the escape phase. In addition, their cargos change with different treatments or stress conditions and can be influenced to be more immune stimulatory against cancer. This review focuses on the emerging understanding of how TEVs affect the differentiation and effector functions of stromal cells and their role in immunoediting, from the early stages of immunoediting to immune escape. Consideration of how TEVs can be therapeutically utilized includes different treatments that can modify TEV to support cancer immunotherapy.

1. Introduction

Roughly 19.3 million new cancer cases are expected to be diagnosed in a year across the world. Before the recent establishment of immune-based cancer therapies, the immune system was generally considered to play a minor role in cancer biology. However, immunotherapies, specifically checkpoint blockade antibodies (CPB), are now established therapy for many cancers, with extensive ongoing research into new higher-impact immunotherapies [1,2]. It is now accepted that the immune system does recognize and attempts to eliminate cancer and an important question in oncology is: how do cancer cells evade the immune system? There is considerable information, but the process is complex and variable between tumor types and between patients, so the understanding is incomplete. The immune system acts as a double-edged that can control and eliminate tumors through or can help tumors progress through immunosuppression and support of angiogenesis [3,4].
The TME contains tumor cells, stromal cells including stromal fibroblasts, endothelial cells, and immune cells that can make up more than half of the overall tumor mass. These cells influence tumor cells and each other through complex interactions such as extracellular matrix (ECM) secretion, soluble factors, cell–cell communication, cytokines, chemokines, and inflammatory mediators [5,6,7]. Extracellular vesicles (EVs) are an important means of intercellular communication and include vesicles up to 1 μm with plasma membrane origin, and smaller lipid bilayer vesicles (30–100 nm), which according to the International Society for Extracellular Vesicles are called small extracellular vesicles (exosomes), that are cup-shaped or doughnut-shaped [8,9]. EVs carry and deliver membrane and cytosolic components, including proteins, lipids, and nucleic acids [10,11]. The physiological and pathological function of EVs depends on their contents and ability to deliver their cargoes. Like other secreted biologically active components, vesicular-based cell-to-cell communication does not require cell contact and can act over long distances [8]. Their internalization in target cells can be through direct fusion with the plasma membrane, endocytosis, phagocytosis, or through ligands on their surface binding to receptors on other cells [8]. Their compositions are associated with endosome biogenesis and parental cell type since EVs with different origins contain unique subsets of components with different cell type-associated functions [10].
The process of tumors interacting with the immune system has been characterized by three phases, elimination, equilibrium, and escape [12,13,14]. During the elimination phase, small tumors not clinically recognized are eliminated without awareness by anyone. During the equilibrium phase, tumors are held in check by immune pressure but not eliminated, and tumors in this stage are also generally not recognized clinically. Lastly, in the escape phase, tumors grow large and are clinically recognized as cancer [12]. It should be noted that most information about cancer comes from tumors in the escape phase, since before that they are small, not identified in humans, and generally hard to study in animal models.
During the elimination phase, the innate and adaptive immune systems work together to identify and eliminate transformed cells that have evaded genetic mechanisms that suppress malignancies. Tumor cells that survive the elimination phase replicate and enable the tumor to reach the equilibrium phase. Both innate and adaptive immunity appears to play a major role in limiting tumor progression in this phase and tumors are controlled, but not eliminated by the immune system. During equilibrium, the tumor is sufficiently immunosuppressive to avoid elimination but is unable to significantly expand. Ultimately in the escape phase, tumors with more robust immunosuppressive mechanisms overcome the immune system via different mechanisms, grow until they are clinically detectable, locally invade and generate metastases [12,13]. In each of the immunoediting phases, cancer-secreted factors interact with the immune system. Some of them can help the tumor grow by suppressing the functional immune cells, whereas others stimulate the immune cells to respond against cancer Figure 1 [12,13,15]. As noted below, EVs play a role in all three stages, and this role generally goes from an antitumor effect in the elimination phase to a protumor effect in the escape phase.
EVs from malignant cells are important mediators of malignant cell communication in the TME and beyond and may support cancer metastasis, angiogenesis, therapy resistance, and immunoregulation leading to resistance to immune surveillance [16,17,18]. Recent studies show that cancer cells use EVs to communicate with one another and with stromal and normal cells [8,10]. The cancer-mediated immunoregulation mechanisms and factors they influence include the expression of surface molecules such as PD-L1 by cancer cells and the recruitment of immunosuppressive cell types such as Tregs and myeloid-derived suppressor cells (MDSC). This review discusses the effects of TEV on the immune system in each phase of cancer immunoediting, their roles in immunoregulation in the TME, and their potential use in cancer immunotherapy.

2. Tumor-Derived Extracellular Vesicles and Tumor-Supportive Cells

During the initial phases of tumor formation, the local microenvironment has mostly anticancer effects [14], but as tumors progress, the healthy microenvironment changes into the TME, immune protection is lost, tumor growth continues, and the stromal cells are influenced by cancer cells to support tumorigenic functions [19,20,21,22]. Along with tumor growth, immunoediting proceeds step by step with the formation of the TME. Tumors that successfully escape the equilibrium phase alter the surrounding healthy microenvironment by manipulating surrounding cancer-associated fibroblasts (CAFs), Tumor-associated macrophages (TAMs), MDSCs, and other immune cells and by recruiting new TME cells that further suppress the immune response via cytokines/chemokines, cell-to-cell contact, growth factors, and EVs [4,23].

2.1. TEVs Modulate Macrophage Activity in TME

Macrophages are complex and plastic cells that adopt a range of phenotypes from strongly immune suppressive to strongly immune stimulatory depending on the environmental signals they receive. While there are no clear distinctions on this continuum, for convenience they are often divided into M1 subtypes with pro-inflammatory properties that express cytokines such as IFN-γ, TNF-α, CXCL-10, IL-12, and high amounts of nitric oxide (NOS), and M2 subtypes with an anti-inflammatory function that release anti-inflammatory cytokines such as IL-4, IL-10, IL-13 and express high amounts of arginase-1(ARG1), scavenger receptors, and mannose receptor [24]. TAMs, begin to shift phenotypic from M1 to M2 through macrophage polarization with exposure to tumor-derived factors and TEVs in the TME and hypoxic conditions and act as a bridge between the adaptive and innate immune systems [25,26]. The M2 cells manifest local supportive functions for the tumor [26,27]. Cancer-derived extracellular vesicles increase M2 polarization by activating signaling pathways such as STAT3, p38MAPK, NF-κB, ERK1/2, and PI3K/AKT [28,29,30] and reprogram M1 tending cells into the cancer-promoting M2 end of the macrophage phenotype spectrum. Table 1 outlines a variety of data on specific activities of TEV in mediating immunosuppression.
The phosphorylated STATs, besides supporting M2 polarization, augment the secretion of generally immune suppressive cytokines such as IL-6 and upregulate PD-L1 that directly suppresses effector T cells through PD-1 on the T cells [28,31,32,33]. Breast cancer cell-derived vesicular gp130 stimulates bone marrow-derived macrophages (BMDMs) to secrete IL-6 by transferring gp130 into BMDMs which results in phosphorylation of STAT3 causing macrophage polarization and IL-6 secretion [28]. Similar to gp130, vesicular Anx II coupling with STAT3 stimulates other signaling pathways in M2 polarization including the p38MAPK, and NF-κB pathways in macrophages, leading to augmented IL-6 and cancer progression [31]. Besides proteins, EVs deliver microRNAs such as miR-222, miR-29a-3p, and miR-146a-5p that also stimulate the STAT3 and the NF-B signaling pathways leading to M2 polarization [34]. Other microRNAs, miR-106b and miR-934, when transferred to macrophages via TEV, activate the PI3K/AKT signaling pathway, which also stimulates macrophages toward M2 polarization [30,35].
Table 1. Effect of tumor-derived extracellular vesicles on the macrophages in the tumor microenvironment.
Table 1. Effect of tumor-derived extracellular vesicles on the macrophages in the tumor microenvironment.
Cancer TypeCellular SourceVesicular CargoThe Main ResultRefs.
Breast cancerMCF10A
MCF10AT
MCF10CA1a
MDA-MB-231
Anx IIActivated NF-B, p38MAPK, and STAT3 pathways in macrophages, leading to increased IL-6 and TNF-α secretion [31]
C57BL/6 EO771gp130Caused macrophages to shift from a normal to a polarized phenotype such as TAM via activation of the IL-6 response pathway and STAT3.[28]
4T1miR-125b-1-3p, miR-100-5p, and miR-183-5pInhibited the expression of PPP2CA, which could promote the release of pro-inflammatory cytokines such as IL-1b, IL-6, and TNF-a from macrophages stimulating tumor invasion.[20]
MDA-MB-231Vesicular CD63 proteinPolarized and activated macrophages, in which CD206 (a marker for M2) was expressed more than NOS2 (a marker for M1).[36]
Prostate cancerPC3miRNA Let-7bProstate-derived extracellular vesicles had more miRNA Let-7b than cellular miRNA Let-7b can lead to macrophage polarization.[37]
Lung cancerA549Vesicular cargoesAltered transcriptomic and bioenergetic profiles of macrophages, forced them to polarize to an M2 phenotype.[38]
NCI-H1437
NCI-H1792
NCI-H2087
miR-103aPolarized monocytes toward immunosuppressive M2-type macrophages.[39]
A549
H1299
Vesicular cargoesEnhanced the levels of MMP2, MMP9 CD163, TNF-, IL-8, IL-6, and IL-10 and decreased expression of iNOS which led macrophages to exhibit a dual M1/M2 phenotype[40]
A549
H1299
Vesicular PRPS2Induced M2 polarization and led to drug resistance of cancer cells.[41]
Hepatocellular carcinoma (HCC)PLC/PRF/5Long non-coding RNAs (lncRNA) TUC339Caused macrophage polarization to be more immunosuppressive.[42]
Hepa1-6
H22
miR-146a-5pEnhanced M2 polarization by triggering NF-B signaling and producing pro-inflammatory proteins[34]
Colorectal cancer(CRC)DLD-1miR-145Induced M2 polarization via upregulation of IL-10 and downregulation of HDAC11.[43]
Blood samples from CRC patients

HCT116
HT29
miR-106bContributed to M2 polarization of macrophages via significant increase in the miR-106b level in macrophages. It directly suppressed programmed cell death 4 (PDCD4) at a post-transcription level that led to an activated PI3Kγ, AKT, and mTOR signaling cascade.[35]
Blood samples from CRC patients

HCT-8
LoVo
HT-29
Caco-2
miR-934Induced M2 macrophage polarization by activating the PI3K/AKT signaling pathway and downregulating PTEN.[30]
CT-26
SW620
Cytoskeleton-centric proteinsIn macrophages, caused cytoskeleton reorganization via promoting elongation and F-actin polarization.[44]
Blood samples from CRC patients

HCT116
DLD-1
HT29
miR-1246Reprogrammed macrophages into the cancer-promoting state after macrophage uptake.[45]
Blood samples from CRC patients

DLD1
HCT116
Lovo
SW480
SW620
HT29
CaR-1
RKO
Colo205
Colo320DM
miR-203Promoted M2 polarization, which modulated liver metastasis of colon cancer cells.[46]
Epithelial ovarian cancerSKOV3 miR-21-3p, miR-181d-5p, and miR-125b-5p Promoted M2 macrophage polarization results in epithelial ovarian cancer cell proliferation and migration under hypoxic circumstances.[47]
GlioblastomaGSC20
GSC276
U87
Vesicular cargoesThe presence of phospho-STAT3 in TEVs switched monocytes toward the tumor-supportive M2 phenotype[33]
U87MG
SBN19
U251
FasL, TRAIL, CTLA-4, CD39, and CD73Promoted M2 polarization by activating the NF-κB pathway in macrophages[48]
U251Vesicular cargoInduced M2 polarization leading to tumor growth via promoting TAM Arginase-1+ exosome secretion[49]
Oral squamous cell carcinomaSCC-9
CAL-27
miR-29a-3pTargeted macrophages directly, and activated p-STAT1 to promote M2 expression [32]
Cal-27CMTM6 Delivered CMTM6 to macrophages and induced M2-like macrophage polarization by activating ERK1/2 signaling[29]
Ovarian cancerBlood samples from overian cancer patients

Skov3
A2780
miR-222Induced M2 polarization of macrophages by activating STAT3 pathway[50]

2.2. TEVs Modulate Fibroblast Activity in TME

In the normal state, fibroblasts are activated during wound healing to help in the process by creating an extracellular matrix (ECM), which serves as a scaffold for other cells [51]. Cancer-associated fibroblasts (CAFs) resemble myofibroblasts and often make up the majority of the cancer stroma [51]. Unlike normal fibroblasts (NFs), CAFs create an excessive ECM and secrete pro-invasive molecules such as ECM-degrading proteases. Hence, CAFs support ECM remodeling, and invasion by producing various kinds of cytokines, growth factors, chemokines, and matrix-degradable enzymes [52,53,54]. The molecular mechanisms that convert normal fibroblasts (NFs) to CAFs in TME are not fully understood. MiRNAs have a major function in the transition and activation of fibroblasts, as evidenced by the fact that dysregulation and disruption of miR-1, 206, 31,214, 155, and 31 secretion leads to the differentiation of NFs to CAFs through modulating FOXO3a, vascular endothelial growth factor (VEGF), and CCL2 signaling [55,56]. Recent studies show that crucial miRNAs in TEVs promote the differentiation of NFs into CAFs [21,57,58,59] (Table 2). Ovarian cancer vesicular miR-630 transformed NFs into CAFs by activating the NF-κB and inhibiting the KLF6 pathway [60]. In another study, lung cancer vesicular miR-210 activated the JAK2/STAT3 pathway, and ten-eleven translocation 2 (TET2) promoted the transformation of NFs into CAF [61]. The JAK2/STAT3 pathway activated by miR-210 resulted in increased expression of some pro-angiogenic factors such as FGF2, MMP9, and VEGF. In addition, breast cancer vesicular proteins such as survivin and ITGB4 converted NFs into myofibroblasts by increasing superoxide dismutase 1 (SOD1) and lactate in CAFs in a BNIP3L-dependent manner [62]. In bladder cancer, the vesicular TGF-β protein activates the TGF-β pathway and triggers CAF differentiation by SMAD pathway activation [19,63].

2.3. TEVs Effect on MDSC Formation in TME

MDSCs normally protect the host from the damaging consequences of excessive immune activation in pathological conditions such as wound healing, but in the TME, MDSCs promote angiogenesis, invasion, and metastasis, as well as block antitumor immunity [68]. MDSCs can generate robust immunosuppressive responses via numerous pathways such as the release of reactive oxygen species (ROS), NO via iNOS, arginine depletion by arginase, secretion of immunosuppressive cytokines such as IL-10 and TGF-β, and stimulation of apoptosis of immune effector cells via the Fas ligand pathway [68,69,70,71]. Therefore, MDSCs are critical mediators in helping cancers evade the immune system. Immature myeloid cells (IMCs) can fail to differentiate under several pathologic conditions (infection, inflammation, and cancer) and develop the features of dysfunctional myeloid cells that include MDSCs through a variety of mechanisms involving numerous substances that accumulate in the TME. Several growth factors and interleukins such as GM-CSF and interleukin IL-6 promote the differentiation of IMCs into MDSCs via activating the STAT-3 signaling pathway [72,73]. MDSCs are a diverse category of IMCs with immunosuppressive characteristics and activities [69]. It also was reported that immature natural killer (NK) cells, can be converted to MDSC [74].
Various tumor-derived factors within or on TEVs surface induce MDSCs in vitro, including IL-1β, IL-6, IL-10, prostaglandin E2 (PGE2), TGF-β, stem cell factor (SCF), and VEGF [72,75] (Table 3). The most important vesicular mediators involved in the differentiation of IMCs into MDSCs include PD-L1, PGE2, TGF-β, and HSP70 [75,76,77,78]. STAT pathways participate since various TEV can differentiate bone marrow myeloid cells into MDSCs by activating STATs [79,80].
Among vesicular cargoes, vesicular PD-L1 enhanced MDSC and M2 formation in breast cancer and glioblastoma and stimulated MDSCs and nonclassical monocyte (NCM) differentiation [76,77]. In breast cancer, TEVs carrying PGE2 and TGF-β switched the differentiation of IMCs into MDSC and also stimulated MDSC expression of Cox2, IL-6, VEGF, and arginase-1 [75]. Vesicular miR-181a and miR-9 stimulated MDSC generation by inhibiting SOCS3 and PIAS3 (two major regulators in the JAK/STAT signaling pathway’s negative feedback loop) [80], and mIR-1246 in glioblastoma-derived EVs induced activation and differentiation of MDSCs via specificity phosphatase 3 (DUSP3) in an ERK-dependent manner [81].
Table 3. Effect of TEVs on immature myeloid cell differentiation to MDSC in the TME.
Table 3. Effect of TEVs on immature myeloid cell differentiation to MDSC in the TME.
Cancer TypeCellular SourceVesicular CargoThe Main ResultRefs.
Breast cancer4T1 tumor model in BALB/c micePGE2 and TGF-βInduced the differentiation of IMCs to MDSC expressing IL-6, Cox2, VEGF, and arginase-1.[75]
MCF-7
4T1
MDA-MB-231
PD-L1+Boosted tumor growth and accumulation of MDSCs and M2 in the TME.[76]
4T1Vesicular cargoesDifferentiated bone marrow cells into MDSCs [79]
4T1 tumor-bearing mice plasma

4T1
miR-181a and miR-9 Stimulated MDSC differentiation by inhibiting SOCS3 and PIAS3 (regulators of the JAK/STAT signaling pathway).[80]
Gastric cancerMKN-28
MKN-45
SGC-7901
Vesicular cargoIncreased frequency of MDSC, and decreased CD8+ T and NK cells.[22]
Renal cancerRenCa HSP 70Antigen-specific immunosuppression effect on CTL[78]
GlioblastomaBlood samples from glioma patientsmiR-1246Induced MDSCs via specificity phosphatase 3 (DUSP3) and ERK-dependent manner.[81]
P3
G422
GL261
U87
miR-29aIncreased MDSCs via interaction with high-mobility group box transcription factor 1 (Hbp1) and protein kinase cAMP-dependent type I regulatory subunit alpha (Prkar1a).[82]
Blood samples from glioma patient
Human astrocytes supernatant
Vesicular cargoActing on MDSC, reduced T-cell immune response in an indirect manner.[83]
Blood samples from glioma patientPD-L1Induced immunosuppressive monocytes, including MDSCs and nonclassical monocytes.[77]
Lung cancer (LC)95D
H292
H358
miR-21aInduced MDSC expression by downregulation of the PDCD4 protein.[84]

3. TEVs-Mediated Communication between Tumor and Immune Cells

TEV early in tumor development can stimulate antitumor immunity. The interaction between immune cells and cancer in TME is categorized into seven potential steps (Figure 2) [4] which is called the “cancer-immunity cycle”. In the right conditions, TEV from tumor cells can also support antitumor immunity. TEVs contain and transfer TAs and damage-associated molecular patterns (DAMPs) to innate immune cells, especially dendritic cells (Step 1) [85,86]. Tumor-derived EVs are a source of shared TAs for CTL cross-priming.
Dendritic cells (DCs) respond to TEVs carrying DAMPs and TAs, mature, and migrate to lymph nodes (Steps 1–2). The tumor antigen is cross-presented on MHC class I (MHC-I) in the lymph nodes where it activates naive CD8 T cells (Step 3). The activated effector T cells go to the tumor site (Step 4), penetrate the tumor tissue (Step 5), identify cancer cells by tumor antigens presented on MHC-I (Step 6), then attack and kill them (Step 7). One or more of these stages may be disrupted in many cancer patients, resulting in ineffective immune responses to cancer. Disruption at any stage of this cycle is caused by cancer cells and their secreted factors, including via TEVs [86,87]. This disruption and immune system suppression block antitumor immunity and support cancer progression.

3.1. Elimination Phase TEV Involvement

This phase has not been directly detected in vivo in humans since it occurs with very small tumors. The innate and adaptive immune systems collaborate to identify and eliminate tumors that have evaded intrinsic tumor suppressor mechanisms in developing tumors [14]. Cancer immunosurveillance is proposed to remove newly generated neoplastic cells that have the potential to develop tumors. The processes of how the immune system is alerted of the presence of primary tumor cells remain unknown. Among the possibilities, the generation of neoantigens by abnormal cells within the created inflammatory environment (via immune cell subsets, recognition molecules, and effector cytokines) results in the detection of nascent cancers, and the traditional warning signals such as IFNs are likely involved [88,89,90]. T cells are the primary immune cells that identify and eliminate tumor cells [88,91]. However, B cells and their antibodies also seem to play a role in recognizing and removing these cells [92]. IFN-γ has a direct anti-proliferative impact on tumors through the STAT1 pathway and causes the release of cytokines such as CXCL9, 10, and 11 that increase immune activation by recruiting effector T cells [93,94]. IFN alpha and beta (type I IFNs) also play an important role in activating CD103+ DCs to cross-present tumor antigens [15,95].
Physical characteristics of the tumor environment such as hypoxia can cause tumor cell death, potentially resulting in the release of DAMPs such as Heat shock proteins (HSPs) and high mobility group box 1 (HMGB1), which act as ligands for Toll-like receptors on innate immune cells [85]. EVs can carry TAs, interferon, and DAMPs that stimulate immunological responses against tumors [96,97]. EVs carried CEA and HER2 TAAs that triggered immune responses and improved anti-tumor responses in vivo [98]. The release of tumor antigens and EVs can be altered under various TME situations. For example, an acidic microenvironment, quite common for tumors, increased the number of secreted EVs [99].
TEVs can play a key role in NK cell activation, DC maturation, and CD8+ effector T-cell development [100,101]. TEVs may also carry surface proteins derived from cancer cells which promote the uptake of TEVs by DCs. There are reports supporting LFA-1/CD54 and mannose-rich C-type lectin receptor interactions as enabling TEV uptake by DCs [102,103]. Uptake of TEVs by DCs enhanced DC expression of co-stimulatory receptors such as CD80, CD86, and also MHC II expression and boosted interferon and cytokine production along with DC maturation [104,105,106]. Breast cancer cells generated EVs that convey dsDNA to DCs, causing IFN alpha and beta expression in a STING-dependent manner and elevation of costimulatory molecules in DCs [107].
Furthermore, TEVs carry molecules that promoted CD8+ T-cell activation and enhanced tumor cytotoxic T lymphocyte (CTL) responses in vivo in mice [105,106,108]. EVs generated from brain tumors were delivered to mice on days 7 and 14 post-tumor inoculation, stimulating antibody production and T-cell activation. Antitumor antibodies and T cells present at the time of tumor inoculation appear to have caused enough tumor cell death to generate further T-cell antitumor response [109].

Tumor-Derived Immunostimulatory Vesicular DAMPs

During an immunogenic cell death (ICD), cancer cells release danger signals (DAMPs) raising the immunogenicity of dying cancer cells [85,110,111,112]. ICD is more immune stimulatory than necrosis which can suppress immunological responses [113] and necrosis generally does not strongly stimulate CD8+ T-cell-dependent immune responses [114]. DAMPs are secreted as a result of endoplasmic reticulum (ER) stress induced by mitochondrial ROS, membrane-lipid peroxidation, and ER-directed ROS generation [115,116,117]. DAMPs can also be released during necroptosis, pyroptosis, and ferroptosis [118,119,120]. EVs from cancer cells can carry DAMPs including HSPs, HMGB1, histones, ATP, vesicular RNAs, and cell-free DNA inside or on the surface [121,122,123,124,125,126]. Interestingly, EVs with surface-bound HSP70 stimulate more helper T cells (Th1) and CTL than TEVs with cytoplasmic HSP70 inside EVs [126]. Hsp70-enriched TEVs elicited significant CD4+ Th1 immune responses and promoted the production of MHC class II molecules on antigen-presenting cells, leading to the elimination of cancer cells [127]. CD94+ NK cells in the presence of TEVs possessing membrane HSP70 released granzyme B [126,128] and expressed stimulating receptors such as the NKG2D, CD69, and NKp44 while also down-regulating inhibitory receptor CD94 [129].

3.2. Equilibrium Phase TEV Involvement

Molecular processes that initiate immune-mediated cancer dormancy/control, i.e., the equilibrium phase (EqP), are not well understood in part because this phase is hard to model and has been minimally characterized in humans [130]. Not surprisingly, when overall mechanisms are poorly understood, there is not much known about the involvement of EVs in the equilibrium phase. In the equilibrium phase, the adaptive effector functions and the resistance of the tumor are in a dynamic balance. There are clear indications that tumors in the escape phase having metastasized, can return to equilibrium following chemotherapy and be dormant for many years before relapse. This occurs in particular with metastatic breast tumors where metastatic cells stop proliferating but survive in a quiescent state [131]. The role, if any that the immune system plays in maintaining this dormancy is not clear.
In the EqP, TEVs may suppress different adaptive immune cell types through various mechanisms such as inhibiting effector cells such as CD8+ T cells and NK cells, suppressing DC maturation and activation, increasing M2 and TAM immune suppressive polarization, and stimulating CAF differentiation [64,132,133]. However, as we noted previously, TEV can also mediate tumor-suppressing signals. TEVs containing miR-23b derived from mesenchymal bone marrow cancer stem cells (CSC) can induce cancer dormancy via downregulation of the MARCKS gene that mediates breast cancer cells’ differentiation into CSCs through the Wnt-β-catenin pathway [134,135].
Considering PD-L1 and IFN-γ in the EqP of tumors is of interest for understanding the involvement of TEV and highlighting the complexity of molecular interactions. While IFN-γ supports CD8 T-cell effector function, IFN-γ stimulation also increases the quantity of PD-L1 on melanoma-released EVs that in turn suppressed the effector function of CD8+ T cells [136]. IFN-γ induced tumor dormancy when the interferon-gamma receptor 1 (IFNGR1) expression level was low but resulted in tumor elimination when it was high [137]. GW4869 treatment or Rab27a knockdown can inhibit vesicular-PD-L1 secretion, and significantly augment anti-PD-L1 therapeutic efficacy in 4T1 tumor growth [138]. Animal studies have shown that TEVs can also impair the production of interferons as well as decrease innate immune activity via EGFR- and MEKK2- dependent pathways [139].

3.3. Escape Phase TEV Involvement

Clinically recognized tumors have generally moved from equilibrium to escape. In the equilibrium phase, genome instability and accumulation of mutations in cancer cells over time leads to selection for low immunogenicity, expression of immune suppressive ligands, and escape from the immune system [140]. Tumors can eventually overcome antitumor immunity through mechanisms already mentioned, including tumor antigen editing, loss of MHC I expression, and expression of immune inhibitors such as PD-L1 [141,142,143] or suppressive mediators such as IL-10 [144], TGF-β [145], and TRAIL decoy receptors [146,147]. Recruitment and activation of immune-suppressing cells such as Tregs also contribute to escape [148].

3.3.1. Effect of TEVs on Dendritic Cells

Maturation of DCs requires inflammation-related stimuli which stimulate the expression of co-stimulatory molecules such as CD86, CD80, and CD40. TEVs can modify or block the differentiation of immature myeloid cells (IMC) to DC or divert the DCs maturation from IMC to MDSC or M2 macrophage (Table 2 and Table 3) by interacting with bone marrow IMC and inducing the production of IL-6, and decreasing expression of CD83 and CD86, as reported for breast cancer, murine mammary adenocarcinoma, and melanoma [149,150]. TEVs also can disrupt DC maturation and T-cell immune response with HLA-G-associated mechanisms in renal cancer [133] (Table 4). Some vesicular proteins such as MALAT1 directly interact with DCs and induce DC autophagy, which decreases DC-mediated T-cell activation [151]. Furthermore, TEV-treated DCs were ineffective at inducing CD4+ T-cell proliferation and activation but promoted differentiation into Treg [152]. TEVs fatty acids can create immunologically dysfunctional DCs by increasing intracellular lipid content by activating the peroxisome proliferator-activated receptor (PPAR) resulting in extra fatty acid oxidation (FAO) which shifts the DCs’ metabolism toward oxidative phosphorylation of mitochondria and the disruption of the function of DCs [153,154,155]. It was reported that human prostate cancer-derived extracellular vesicles purified from cultured cells contained PGE2 and triggered the expression of CD73 and CD39 on DCs in vitro, resulting in the generation of adenosine from ATP and inhibition of TNF-α and IL-12-production which reduced T-cell activation [156].
HSP72 and HSP105 on the membrane of TEVs interact with TLR2 and TLR4 on DCs which induced IL-6 secretion by DCs that increased STAT3-dependent MMP-9 transcription activity in cancer cells resulting in tumor invasion [161]. Galectin-9 on glioblastoma-derived EVs binds to the TIM3 DCs receptor and inhibits antigen presentation by DCs, leading to disrupted antitumor immune responses of cytotoxic T cells [142]. Important DC receptors such as Tim-3 and galectin-9 [157] and SIRPα as the ligand for CD47 were up-regulated on the tumor cells’ membranes and derived TEV [143,164]. TLR4 on the DCs decreased after treatment with pancreatic cancer-derived vesicular miR-203 resulting in reduced expression of cytokines such as TNF-α and IL-12, subsequently reducing DC maturation and Th1 differentiation [125]. Besides the vesicular proteins, vesicular miRs also affect DC’s function. For example, miR-212-3p transferred to DCs by pancreatic cancer-derived extracellular vesicles suppressed regulatory factor X-associated protein (RFXAP), decreased MHC II expression, and reduced antigen presentation by DCs [165]. Table 4 summarizes reports of TEV impacts on DC.

3.3.2. Effect of TEVs on T Cells

TEVs have a broad array of mechanisms by which they impact T cells. TEVs modify antitumor response by reducing T-cell viability, proliferation, and effector activities [166,167,168]. TEVs can disrupt T-cell effector function indirectly by blocking APC maturation [142,151,152] or directly by inhibiting activated CD8+ T-cell function, inducing CD8+ T-cell death through pro-apoptotic molecules (galectin-group proteins and FasL), promoting Treg expansion, and inducing T-cell exhaustion [169,170]. PD-L1 enriched glioblastoma-derived EVs perhaps surprisingly suppress monocytes rather than T-cells [77]. Nasopharyngeal carcinoma-derived vesicular galectin-9 induced apoptosis in CD4+ T cells via interaction with Tim-3 [171], as well as impairing T-cell function by interaction with TIM3 receptor on DCs in glioblastoma [142]. TEVs can carry pro-apoptotic Bax that induces apoptosis in CD8+T cells [172] and downregulates JAK3 expression which blocks CD8+ T-cell activation [167,173]. In Treg cell activation, both CD45 negative and positive EVs derived from plasma in head and neck cancer induced Treg differentiation of CD4 cells, but CD45(-) EVs also reduced CD8+ T-cell activation due to their higher adenosine concentrations [174]. EVs generated from multiple myeloma reduced the viability of CD4+ T cells and boosted the proliferation of Treg cells [175].
Vesicular PD-L1 promotes CD8+ T-cell apoptosis via PD-1/PD-L1 and PD-L1/CD80 signaling pathways [176], blocks T-cell activation in the draining lymph node in TRAMP- C2 prostate cancer mouse model [177,178], and reduces the proliferation of CD8+ T cells by decreasing IL- 2 and IFN-γ in the TME [136]. FasL on the TEVs decreased T-cell receptor (TCR) and CD3ζ expression in T cells leading to T-cell apoptosis [179], and melanoma-derived vesicular TNF downregulates TCR via redox signaling in T cells [180].
Pancreatic cancer cell EVs can stimulate p38 MAP kinase signaling in T lymphocytes that causes ER stress, which triggers the PERK–eIF2–ATF4–CHOP signaling cascade resulting in T-cell death [181]. Vesicular microRNAs in the serum of patients with nasopharyngeal carcinoma influenced T-cell differentiation and activation through suppression of the MAPK1 signaling pathway [182], while EVs with a high amount of miR-24–3 reduced CD4+ and CD8+ T-cell proliferation by targeting FGF11 [183]. In addition, mesothelioma cells’ EVs carrying TGF-β decreased proliferative response to IL-2 in T effector cells, but not in T-reg cells [184].
Vesicular galectin-1 plays a role in the induction of T-cell suppression [185]. TEVs also can induce T-cell exhaustion, by carrying inhibitory molecules, including PD-L1, CTLA- 4, TIM3, LAG3, and TIGIT [186,187]. miR-146a-5p and 14-3-3ζ in HCC-derived EVs induced T-cell exhaustion via activating M2-macrophages by inhibiting transcription factor SALL4 [30,188]. EVs carrying circRNA-002178 from patients’ serum with lung adenocarcinoma could boost PD-L1 production by sponging miR-34 in cancer cells, leading to CD8+T-cell exhaustion in vitro [132].
In addition, cancer patients’ plasma TEVs can prevent the activation of Th1 and Th17 lymphocytes and change them to immunosuppressive Treg phenotype cells [167,182]. The mutant KRAS gene is involved in the NSCLC-generated EVs-mediated transition of naive CD4+ T cells towards a FoxP3+ T-reg phenotype in a cytokine-independent manner in an NSCLC xenograft mouse model [189]. Table 5 summarizes reports on TEV suppressive effects on T cells.

3.3.3. Effect of TEVs on NK Cells

NK cells play an important role in cancer immunosurveillance by expressing death-inducing ligands such as FasL, TRAIL and JAK/STAT pathway [197,198]. However, like most immune cells, the activation of NK cells is controlled by a complex balance of activating and inhibiting signals. Tumor cells trigger several activating receptors, such as NKG2D, natural cytotoxicity receptors (NCRs), and DNAX accessory molecule-1 (DNAM-1/CD226) [199]. Vesicular NKG2D, TGF-β, and MICA*008 suppress or downregulate the expression of NKG2D in both NK and CD8+ T cells resulting in decreasing cytotoxicity of these cells by reducing the expression of cytotoxic molecules [200,201,202,203,204,205,206].

4. Potentials of EVs in Cancer Therapy

TEVs have both immunostimulation and immunosuppression effects [96,97,161], but the potential has not yet been clinically utilized. Admittedly, TEV-focused therapy will be challenging. One challenge to using EVs in cancer immunotherapy is developing a system that provides uniform reagents for clinical use. However, a variety of preclinical studies illustrate the therapeutic potential of TEVs.
TEVs from tumors in the elimination phase stimulated immune cell responses against cancer development [101,104,128], while, not surprisingly, by the time tumor progression occurs, TEVs tend to suppress immune cells and support tumor immune escape. Three general cancer immunotherapy approaches involving TEV can be conceived: I) inhibition of TEV secretion II) increasing the immunostimulatory factors on TEVs’ surfaces, and III) using EVs as carriers in cancer vaccines.
To block the secretion of TEVs, the factors involved in their secretion, such as endosomal sorting complexes required for transport machinery (ESCRT), soluble NSF attachment protein receptor (SNARE), and Rab proteins (Rab11, Rab 27a, Rab 27b, and Rab 35) could be suppressed using drugs including Y27632, Imipramine, Calpeptin, Manumycin A, D-Pantethine, GW4869, and Simvastatin [207,208,209].
Cells under stress produce more immunostimulatory molecules on TEVs and secrete more EVs, which can be caused by treatment-induced stress [210]. Thus some cancer therapies can increase the production of immunostimulatory vesicular factors and this may help to disrupt and even overcome the process of usual tumor immunoediting [211,212]. Opposing increased immunostimulation from tumor EVs, treating with immune checkpoint inhibitors can boost the secretion of immunosuppressive EVs [136].
Hyperthermia is a useful cancer treatment and heat or other stress can modulate TEVs. Heat-stressed B lymphoma cells’ EVs possess more IL-6 and IL-17 stimulating molecules such as HSP90, HSP60, HSP70, CD40, and CD86, which can turn Tregs into Th 17 cells [211,213]. Heat stress boosts MHC-I expression on tumor cells [106] and generates TEVs equipped with chemokines such as CCL2,3,4,5, and CCL20 that functionally activate DC and T cells more strongly against tumors [214] thus stimulating “self-vaccination” [215]. Irradiated mouse breast cancer cells’ TEVs transmit dsDNA to DCs and induce DC to overexpress costimulatory molecules as well as STING-dependent type I IFN [107] and irradiated melanoma cells’ TEVs contained DC activation DAMPs such as HSP70, HMGB, and other stress-related proteins [124]. EVs from Melphalan (a genotoxic drug) treated myeloma cells can boost NK cell IFN-γ production via activating the NF-κB pathway in a TLR2/Hsp70-dependent manner [123]. IFN-γ treated cancer cells secreted a high amount of immune stimulatory TEVs that enhance the number of M1 macrophages by improving their capacity to ingest TEVs and promoting antibody production against cancer cells [216] as well as reducing Tregs and suppressing the expression of PD-L1, VEGF receptor 2, and IDO-1 [217]. In another strategy, modification of the vesicular contents by non-stress methods such as a lentiviral vector encoding two B7 costimulatory molecules (CD80, CD86) increased CD86 and CD80 expression in DCs and induced proliferation of CD4+ T cells, Th1 cytokine secretion, and CTL response [218].
TEVs have been studied as vaccine carriers and employed as immunogens for DC loading. Their immunogenicity in boosting DC-driven anti-tumor immunity was greater than tumor lysate, and they increased splenocyte proliferation and IL-2 release in mouse leukemia and melanoma cancer models [219,220]. In different syngeneic mouse models with large tumors, TEVs equipped with HMGB1 augment DC immunogenicity and elicit long-lasting antitumor immunity and tumor suppression [221]. Overall, EVs from a variety of cell types, including immune cells such as DCs and cancer cells, have the potential as a cancer vaccine and cancer immunotherapeutic [222] such as for colon cancer [2].
TEVs have potential use as drug carriers since they have an affinity for ingestion by cancer cells, are biocompatible and non-toxic with long half-lives in circulation, and their potential has been evaluated [223,224,225]. Some studies carried doxorubicin and in comparison to free doxorubicin, they boosted the therapeutic efficacy [224,225]. Additionally, TEVs carrying doxorubicin and paclitaxel crossed the blood–brain barrier (BBB) as part of in vivo studies [223]. This strategy of using TEVs to carry chemotherapy drugs as a cancer treatment has clinical potential and needs further study.

5. Conclusions

TEVs cargos are not static during cancer development; they change as tumors evolve and are stressed for various reasons. TEVs modulate immunostimulating or immunosuppressing effects against cancer cells by modifying immune cells during the tumor immunoediting phases [12]. TEVs play an immunostimulatory role in the early stages of immunoediting [8,12,96,101], are more immunosuppressive in the escape phase, and finally, at the late stages, they are more uniformly immune suppressive and play a major role in cancer immune escape [19,156,168]. Normally. the cargo profile of TEVs naturally changes in tumor immunoediting toward immunosuppression, while different cell stress or treatment conditions can inhibit this process or even reverse it to immunostimulation by altering the TEVs cargos profile. Therefore, to benefit from the therapeutic effects of EVs, the secretion of immunosuppression TEVs could be inhibited by disrupting the normal process of immunoediting. Generating EVs to apply therapeutically could be achieved by stimulating the release of immune-activating vesicular cargoes in vitro by using suitable treatment methods and subsequently using EVs in vivo as adjuvant therapy. Since TEV’s very diverse cargo profiles depend on the cancer cell condition, understanding the immunotherapeutic properties of TEVs to utilize against cancer should be considered a new research line in oncoimmunotherapy.

Author Contributions

Conceptualization, M.N.; writing - original draft, M.N.; writing—review and editing, S.F. and A.Y.K. and M.N. and M.S.; supervision, A.Y.K. and S.F.; Funding acquisition, S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. 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]
  3. Hussain, S.P.; Amstad, P.; Raja, K.; Ambs, S.; Nagashima, M.; Bennett, W.P.; Shields, P.G.; Ham, A.-J.; Swenberg, J.A.; Marrogi, A.J. Increased p53 mutation load in noncancerous colon tissue from ulcerative colitis: A cancer-prone chronic inflammatory disease. Cancer Res. 2000, 60, 3333–3337. [Google Scholar] [PubMed]
  4. Quail, D.F.; Joyce, J.A. Microenvironmental regulation of tumor progression and metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef]
  5. Hanahan, D.; Coussens, L.M. Accessories to the crime: Functions of cells recruited to the tumor microenvironment. Cancer Cell 2012, 21, 309–322. [Google Scholar] [CrossRef] [Green Version]
  6. Meurette, O.; Mehlen, P. Notch Signaling in the Tumor Microenvironment. Cancer Cell 2018, 34, 536–548. [Google Scholar] [CrossRef] [Green Version]
  7. Danenberg, E.; Bardwell, H.; Zanotelli, V.R.; Provenzano, E.; Chin, S.-F.; Rueda, O.M.; Green, A.; Rakha, E.; Aparicio, S.; Ellis, I.O. Breast tumor microenvironment structures are associated with genomic features and clinical outcome. Nat. Genet. 2022, 54, 660–669. [Google Scholar] [CrossRef]
  8. Robbins, P.D.; Morelli, A.E. Regulation of immune responses by extracellular vesicles. Nat. Rev. Immunol. 2014, 14, 195–208. [Google Scholar] [CrossRef] [Green Version]
  9. 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. 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]
  10. Kalluri, R. The biology and function of exosomes in cancer. J. Clin. Investig. 2016, 126, 1208–1215. [Google Scholar] [CrossRef]
  11. Melo, S.A.; Luecke, L.B.; Kahlert, C.; Fernandez, A.F.; Gammon, S.T.; Kaye, J.; LeBleu, V.S.; Mittendorf, E.A.; Weitz, J.; Rahbari, N. Glypican-1 identifies cancer exosomes and detects early pancreatic cancer. Nature 2015, 523, 177–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. O’Donnell, J.S.; Teng, M.W.; Smyth, M.J. Cancer immunoediting and resistance to T cell-based immunotherapy. Nat. Rev. Clin. Oncol. 2019, 16, 151–167. [Google Scholar] [CrossRef] [PubMed]
  13. Dunn, G.P.; Bruce, A.T.; Ikeda, H.; Old, L.J.; Schreiber, R.D. Cancer immunoediting: From immunosurveillance to tumor escape. Nat. Immunol. 2002, 3, 991–998. [Google Scholar] [CrossRef] [PubMed]
  14. Schreiber, R.D.; Old, L.J.; Smyth, M.J. Cancer immunoediting: Integrating immunity’s roles in cancer suppression and promotion. Science 2011, 331, 1565–1570. [Google Scholar] [CrossRef] [Green Version]
  15. von Locquenghien, M.; Rozalén, C.; Celià-Terrassa, T. Interferons in cancer immunoediting: Sculpting metastasis and immunotherapy response. J. Clin. Investig. 2021, 131. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, X.-G.; Zhou, C.-F.; Zhang, Y.-M.; Yan, R.-M.; Wei, W.-F.; Chen, X.-J.; Yi, H.-Y.; Liang, L.-J.; Fan, L.-s.; Liang, L. Cancer-derived exosomal miR-221-3p promotes angiogenesis by targeting THBS2 in cervical squamous cell carcinoma. Angiogenesis 2019, 22, 397–410. [Google Scholar] [CrossRef] [PubMed]
  17. Dorayappan, K.D.P.; Wanner, R.; Wallbillich, J.J.; Saini, U.; Zingarelli, R.; Suarez, A.A.; Cohn, D.E.; Selvendiran, K. Hypoxia-induced exosomes contribute to a more aggressive and chemoresistant ovarian cancer phenotype: A novel mechanism linking STAT3/Rab proteins. Oncogene 2018, 37, 3806–3821. [Google Scholar] [CrossRef]
  18. Lopatina, T.; Koni, M.; Grange, C.; Cedrino, M.; Femminò, S.; Lombardo, G.; Favaro, E.; Brizzi, M.F. IL-3 signalling in the tumour microenvironment shapes the immune response via tumour endothelial cell-derived extracellular vesicles. Pharm. Res. 2022, 179, 106206. [Google Scholar] [CrossRef]
  19. Yeon, J.H.; Jeong, H.E.; Seo, H.; Cho, S.; Kim, K.; Na, D.; Chung, S.; Park, J.; Choi, N.; Kang, J.Y. Cancer-derived exosomes trigger endothelial to mesenchymal transition followed by the induction of cancer-associated fibroblasts. Acta Biomater. 2018, 76, 146–153. [Google Scholar] [CrossRef]
  20. Guo, J.; Duan, Z.; Zhang, C.; Wang, W.; He, H.; Liu, Y.; Wu, P.; Wang, S.; Song, M.; Chen, H. Mouse 4T1 breast cancer cell–derived exosomes induce proinflammatory cytokine production in macrophages via miR-183. J. Immunol. 2020, 205, 2916–2925. [Google Scholar] [CrossRef]
  21. Zhou, M.; Wang, S.; Liu, D.; Zhou, J. LINC01915 Facilitates the Conversion of Normal Fibroblasts into Cancer-Associated Fibroblasts Induced by Colorectal Cancer-Derived Extracellular Vesicles through the miR-92a-3p/KLF4/CH25H Axis. ACS Biomater. Sci. Eng. 2021, 7, 5255–5268. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, J.; Wu, S.; Zheng, X.; Zheng, P.; Fu, Y.; Wu, C.; Lu, B.; Ju, J.; Jiang, J. Immune suppressed tumor microenvironment by exosomes derived from gastric cancer cells via modulating immune functions. Sci. Rep. 2020, 10, 14749. [Google Scholar] [CrossRef] [PubMed]
  23. Guan, H.; Peng, R.; Fang, F.; Mao, L.; Chen, Z.; Yang, S.; Dai, C.; Wu, H.; Wang, C.; Feng, N. Tumor-associated macrophages promote prostate cancer progression via exosome-mediated miR-95 transfer. J. Cell Physiol. 2020, 235, 9729–9742. [Google Scholar] [CrossRef] [PubMed]
  24. Gordon, S. Alternative activation of macrophages. Nat. Rev. Immunol. 2003, 3, 23–35. [Google Scholar] [CrossRef] [PubMed]
  25. Jiang, M.; Li, X.; Zhang, J.; Lu, Y.; Shi, Y.; Zhu, C.; Liu, Y.; Qin, B.; Luo, Z.; Du, Y. Dual Inhibition of Endoplasmic Reticulum Stress and Oxidation Stress Manipulates the Polarization of Macrophages under Hypoxia to Sensitize Immunotherapy. ACS Nano 2021, 15, 14522–14534. [Google Scholar] [CrossRef]
  26. Mantovani, A.; Marchesi, F.; Malesci, A.; Laghi, L.; Allavena, P. Tumour-associated macrophages as treatment targets in oncology. Nat. Rev. Clin. Oncol. 2017, 14, 399–416. [Google Scholar] [CrossRef]
  27. Scott, E.M.; Jacobus, E.J.; Lyons, B.; Frost, S.; Freedman, J.D.; Dyer, A.; Khalique, H.; Taverner, W.K.; Carr, A.; Champion, B.R. Bi-and tri-valent T cell engagers deplete tumour-associated macrophages in cancer patient samples. J. Immunother. Cancer 2019, 7, 1–18. [Google Scholar] [CrossRef]
  28. Ham, S.; Lima, L.G.; Chai, E.P.Z.; Muller, A.; Lobb, R.J.; Krumeich, S.; Wen, S.W.; Wiegmans, A.P.; Möller, A. Breast cancer-derived exosomes alter macrophage polarization via gp130/STAT3 signaling. Front. Immunol. 2018, 9, 871. [Google Scholar] [CrossRef] [Green Version]
  29. Pang, X.; Wang, S.-s.; Zhang, M.; Jiang, J.; Fan, H.-y.; Wu, J.-s.; Wang, H.-f.; Liang, X.-h.; Tang, Y.-l. OSCC cell-secreted exosomal CMTM6 induced M2-like macrophages polarization via ERK1/2 signaling pathway. Cancer Immunol. Immunother. 2021, 70, 1015–1029. [Google Scholar] [CrossRef]
  30. Zhao, S.; Mi, Y.; Guan, B.; Zheng, B.; Wei, P.; Gu, Y.; Zhang, Z.; Cai, S.; Xu, Y.; Li, X. Tumor-derived exosomal miR-934 induces macrophage M2 polarization to promote liver metastasis of colorectal cancer. J. Hematol. Oncol. 2020, 13, 156. [Google Scholar] [CrossRef]
  31. Maji, S.; Chaudhary, P.; Akopova, I.; Nguyen, P.M.; Hare, R.J.; Gryczynski, I.; Vishwanatha, J.K. Exosomal annexin II promotes angiogenesis and breast cancer metastasis. Mol. Cancer Res. 2017, 15, 93–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Cai, J.; Qiao, B.; Gao, N.; Lin, N.; He, W. Oral squamous cell carcinoma-derived exosomes promote M2 subtype macrophage polarization mediated by exosome-enclosed miR-29a-3p. Am. J. Physiol.-Cell Physiol. 2019, 316, C731–C740. [Google Scholar] [CrossRef] [PubMed]
  33. Gabrusiewicz, K.; Li, X.; Wei, J.; Hashimoto, Y.; Marisetty, A.L.; Ott, M.; Wang, F.; Hawke, D.; Yu, J.; Healy, L.M. 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]
  34. Yin, C.; Han, Q.; Xu, D.; Zheng, B.; Zhao, X.; Zhang, J. SALL4-mediated upregulation of exosomal miR-146a-5p drives T-cell exhaustion by M2 tumor-associated macrophages in HCC. Oncoimmunology 2019, 8, e1601479. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, C.; Dou, R.; Wei, C.; Liu, K.; Shi, D.; Zhang, C.; Liu, Q.; Wang, S.; Xiong, B. Tumor-derived exosomal microRNA-106b-5p activates EMT-cancer cell and M2-subtype TAM interaction to facilitate CRC metastasis. Mol. Ther. 2021, 29, 2088–2107. [Google Scholar] [CrossRef]
  36. Piao, Y.J.; Kim, H.S.; Hwang, E.H.; Woo, J.; Zhang, M.; Moon, W.K. Breast cancer cell-derived exosomes and macrophage polarization are associated with lymph node metastasis. Oncotarget 2018, 9, 7398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Costanzi, E.; Romani, R.; Scarpelli, P.; Bellezza, I. Extracellular Vesicles-Mediated Transfer of miRNA Let-7b from PC3 Cells to Macrophages. Genes 2020, 11, 1495. [Google Scholar] [CrossRef]
  38. Pritchard, A.; Tousif, S.; Wang, Y.; Hough, K.; Khan, S.; Strenkowski, J.; Chacko, B.K.; Darley-Usmar, V.M.; Deshane, J.S. Lung tumor cell-derived exosomes promote M2 macrophage polarization. Cells 2020, 9, 1303. [Google Scholar] [CrossRef]
  39. 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] [Green Version]
  40. Chen, L.; Yang, Y.; Huang, C.; Cao, P.; Wu, Q.; Chen, S.; Chen, F. Human THP-1 macrophages activated by exosomes derived from lung adenocarcinoma cells promote lung cancer cell invasion. Chin. J. Cell Mol. Immunol. 2019, 35, 967–972. [Google Scholar]
  41. Liu, G.; Luo, Y.; Hou, P. PRPS2 Enhances Resistance to Cisplatin via Facilitating Exosomes-mediated Macrophage M2 Polarization in Non-small Cell Lung Cancer. Immunol. Investig. 2021, 51, 1–14. [Google Scholar] [CrossRef] [PubMed]
  42. Li, X.; Lei, Y.; Wu, M.; Li, N. Regulation of macrophage activation and polarization by HCC-derived exosomal lncRNA TUC339. Int. J. Mol. Sci. 2018, 19, 2958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Shinohara, H.; Kuranaga, Y.; Kumazaki, M.; Sugito, N.; Yoshikawa, Y.; Takai, T.; Taniguchi, K.; Ito, Y.; Akao, Y. Regulated polarization of tumor-associated macrophages by mir-145 via colorectal cancer–derived extracellular vesicles. J. Immunol. 2017, 199, 1505–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Chen, Z.; Yang, L.; Cui, Y.; Zhou, Y.; Yin, X.; Guo, J.; Zhang, G.; Wang, T.; He, Q.-Y. Cytoskeleton-centric protein transportation by exosomes transforms tumor-favorable macrophages. Oncotarget 2016, 7, 67387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. 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] [PubMed] [Green Version]
  46. Takano, Y.; Masuda, T.; Iinuma, H.; Yamaguchi, R.; Sato, K.; Tobo, T.; Hirata, H.; Kuroda, Y.; Nambara, S.; Hayashi, N. Circulating exosomal microRNA-203 is associated with metastasis possibly via inducing tumor-associated macrophages in colorectal cancer. Oncotarget 2017, 8, 78598. [Google Scholar] [CrossRef] [Green Version]
  47. 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]
  48. Azambuja, J.H.; Ludwig, N.; Yerneni, S.; Rao, A.; Braganhol, E.; Whiteside, T.L. Molecular profiles and immunomodulatory activities of glioblastoma-derived exosomes. Neuro-Oncol. Adv. 2020, 2, vdaa056. [Google Scholar] [CrossRef]
  49. Azambuja, J.H.; Ludwig, N.; Yerneni, S.S.; Braganhol, E.; Whiteside, T.L. Arginase-1+ exosomes from reprogrammed macrophages promote glioblastoma progression. Int. J. Mol. Sci. 2020, 21, 3990. [Google Scholar] [CrossRef]
  50. Ying, X.; Wu, Q.; Wu, X.; Zhu, Q.; Wang, X.; Jiang, L.; Chen, X.; Wang, X. Epithelial ovarian cancer-secreted exosomal miR-222-3p induces polarization of tumor-associated macrophages. Oncotarget 2016, 7, 43076. [Google Scholar] [CrossRef] [Green Version]
  51. Kalluri, R. The biology and function of fibroblasts in cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, Y.-f.; Yu, Z.-l.; Lv, M.-y.; Cai, Z.-r.; Zou, Y.-f.; Lan, P.; Wu, X.-j.; Gao, F. Cancer-associated fibroblasts impact the clinical outcome and treatment response in colorectal cancer via immune system modulation: A comprehensive genome-wide analysis. Mol. Med. 2021, 27, 139. [Google Scholar] [CrossRef] [PubMed]
  53. Pelon, F.; Bourachot, B.; Kieffer, Y.; Magagna, I.; Mermet-Meillon, F.; Bonnet, I.; Costa, A.; Givel, A.-M.; Attieh, Y.; Barbazan, J. Cancer-associated fibroblast heterogeneity in axillary lymph nodes drives metastases in breast cancer through complementary mechanisms. Nat. Commun. 2020, 11, 404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yang, J.; Shi, X.; Yang, M.; Luo, J.; Gao, Q.; Wang, X.; Wu, Y.; Tian, Y.; Wu, F.; Zhou, H. Glycolysis reprogramming in cancer-associated fibroblasts promotes the growth of oral cancer through the lncRNA H19/miR-675-5p/PFKFB3 signaling pathway. Int. J. Oral Sci. 2021, 13, 12. [Google Scholar] [CrossRef] [PubMed]
  55. Mitra, A.K.; Zillhardt, M.; Hua, Y.; Tiwari, P.; Murmann, A.E.; Peter, M.E.; Lengyel, E. MicroRNAs reprogram normal fibroblasts into cancer-associated fibroblasts in ovarian cancer. Cancer Discov. 2012, 2, 1100–1108. [Google Scholar] [CrossRef] [Green Version]
  56. Shen, H.; Yu, X.; Yang, F.; Zhang, Z.; Shen, J.; Sun, J.; Choksi, S.; Jitkaew, S.; Shu, Y. Reprogramming of normal fibroblasts into cancer-associated fibroblasts by miRNAs-mediated CCL2/VEGFA signaling. PLoS Genet. 2016, 12, e1006244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Motohara, T.; Masuda, K.; Morotti, M.; Zheng, Y.; El-Sahhar, S.; Chong, K.Y.; Wietek, N.; Alsaadi, A.; Karaminejadranjbar, M.; Hu, Z. An evolving story of the metastatic voyage of ovarian cancer cells: Cellular and molecular orchestration of the adipose-rich metastatic microenvironment. Oncogene 2019, 38, 2885–2898. [Google Scholar] [CrossRef] [Green Version]
  58. Baroni, S.; Romero-Cordoba, S.; Plantamura, I.; Dugo, M.; D’ippolito, E.; Cataldo, A.; Cosentino, G.; Angeloni, V.; Rossini, A.; Daidone, M. Exosome-mediated delivery of miR-9 induces cancer-associated fibroblast-like properties in human breast fibroblasts. Cell Death Dis. 2016, 7, e2312. [Google Scholar] [CrossRef] [Green Version]
  59. Wang, C.; Wang, Y.; Chang, X.; Ba, X.; Hu, N.; Liu, Q.; Fang, L.; Wang, Z. Melanoma-derived exosomes endow fibroblasts with an invasive potential via miR-21 target signaling pathway. Cancer Manag. Res. 2020, 12, 12965. [Google Scholar] [CrossRef]
  60. Cui, Y.; Wang, D.; Xie, M. Tumor-derived extracellular vesicles promote activation of carcinoma-associated fibroblasts and facilitate invasion and metastasis of ovarian cancer by carrying miR-630. Front. Cell Dev. Biol. 2021, 9, 1576. [Google Scholar] [CrossRef]
  61. Fan, J.; Xu, G.; Chang, Z.; Zhu, L.; Yao, J. miR-210 transferred by lung cancer cell-derived exosomes may act as proangiogenic factor in cancer-associated fibroblasts by modulating JAK2/STAT3 pathway. Clin. Sci. 2020, 134, 807–825. [Google Scholar] [CrossRef] [PubMed]
  62. Li, K.; Liu, T.; Chen, J.; Ni, H.; Li, W. Survivin in breast cancer–derived exosomes activates fibroblasts by up-regulating SOD1, whose feedback promotes cancer proliferation and metastasis. J. Biol. Chem. 2020, 295, 13737–13752. [Google Scholar] [CrossRef] [PubMed]
  63. Goulet, C.R.; Bernard, G.; Tremblay, S.; Chabaud, S.; Bolduc, S.; Pouliot, F. Exosomes induce fibroblast differentiation into cancer-associated fibroblasts through TGFβ signaling. Mol. Cancer Res. 2018, 16, 1196–1204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sung, J.S.; Kang, C.W.; Kang, S.; Jang, Y.; Chae, Y.C.; Kim, B.G.; Cho, N.H. ITGB4-mediated metabolic reprogramming of cancer-associated fibroblasts. Oncogene 2020, 39, 664–676. [Google Scholar] [CrossRef] [PubMed]
  65. Rai, A.; Greening, D.W.; Chen, M.; Xu, R.; Ji, H.; Simpson, R.J. Exosomes derived from human primary and metastatic colorectal cancer cells contribute to functional heterogeneity of activated fibroblasts by reprogramming their proteome. Proteomics 2019, 19, 1800148. [Google Scholar] [CrossRef] [PubMed]
  66. Hu, T.; Hu, J. Melanoma-derived exosomes induce reprogramming fibroblasts into cancer-associated fibroblasts via Gm26809 delivery. Cell Cycle 2019, 18, 3085–3094. [Google Scholar] [CrossRef] [PubMed]
  67. Mito, I.; Takahashi, H.; Kawabata-Iwakawa, R.; Horikawa, M.; Ida, S.; Tada, H.; Matsuyama, T.; Misawa, K.; Takeda, S.; Chikamatsu, K. Tumor-derived exosomes elicit cancer-associated fibroblasts shaping inflammatory tumor microenvironment in head and neck squamous cell carcinoma. Oral Oncol. 2023, 136, 106270. [Google Scholar] [CrossRef]
  68. Gabrilovich, D.I.; Nagaraj, S. Myeloid-derived suppressor cells as regulators of the immune system. Nat. Rev. Immunol. 2009, 9, 162–174. [Google Scholar] [CrossRef] [Green Version]
  69. Groth, C.; Hu, X.; Weber, R.; Fleming, V.; Altevogt, P.; Utikal, J.; Umansky, V. Immunosuppression mediated by myeloid-derived suppressor cells (MDSCs) during tumour progression. Br. J. Cancer 2019, 120, 16–25. [Google Scholar] [CrossRef] [Green Version]
  70. Alicea-Torres, K.; Sanseviero, E.; Gui, J.; Chen, J.; Veglia, F.; Yu, Q.; Donthireddy, L.; Kossenkov, A.; Lin, C.; Fu, S. Immune suppressive activity of myeloid-derived suppressor cells in cancer requires inactivation of the type I interferon pathway. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  71. Ferrer, G.; Jung, B.; Chiu, P.Y.; Aslam, R.; Palacios, F.; Mazzarello, A.N.; Vergani, S.; Bagnara, D.; Chen, S.-S.; Yancopoulos, S. Myeloid-derived suppressor cell subtypes differentially influence T-cell function, T-helper subset differentiation, and clinical course in CLL. Leukemia 2021, 35, 3163–3175. [Google Scholar] [CrossRef] [PubMed]
  72. Lechner, M.G.; Liebertz, D.J.; Epstein, A.L. Characterization of cytokine-induced myeloid-derived suppressor cells from normal human peripheral blood mononuclear cells. J. Immunol. 2010, 185, 2273–2284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Jiang, M.; Chen, J.; Zhang, W.; Zhang, R.; Ye, Y.; Liu, P.; Yu, W.; Wei, F.; Ren, X.; Yu, J. Interleukin-6 trans-signaling pathway promotes immunosuppressive myeloid-derived suppressor cells via suppression of suppressor of cytokine signaling 3 in breast cancer. Front. Immunol. 2017, 8, 1840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Park, Y.-J.; Song, B.; Kim, Y.-S.; Kim, E.-K.; Lee, J.-M.; Lee, G.-E.; Kim, J.-O.; Kim, Y.-J.; Chang, W.-S.; Kang, C.-Y. Tumor microenvironmental conversion of natural killer cells into myeloid-derived suppressor cells. Cancer Res. 2013, 73, 5669–5681. [Google Scholar] [CrossRef] [Green Version]
  75. Xiang, X.; Poliakov, A.; Liu, C.; Liu, Y.; Deng, Z.b.; Wang, J.; Cheng, Z.; Shah, S.V.; Wang, G.J.; Zhang, L. Induction of myeloid-derived suppressor cells by tumor exosomes. Int. J. Cancer 2009, 124, 2621–2633. [Google Scholar] [CrossRef] [Green Version]
  76. Wang, Y.; Goliwas, K.F.; Severino, P.E.; Hough, K.P.; Van Vessem, D.; Wang, H.; Tousif, S.; Koomullil, R.P.; Frost, A.R.; Ponnazhagan, S. Mechanical strain induces phenotypic changes in breast cancer cells and promotes immunosuppression in the tumor microenvironment. Lab. Investig. 2020, 100, 1503–1516. [Google Scholar] [CrossRef]
  77. Himes, B.T.; Peterson, T.E.; de Mooij, T.; Garcia, L.M.C.; Jung, M.-Y.; Uhm, S.; Yan, D.; Tyson, J.; Jin-Lee, H.J.; Parney, D. The role of extracellular vesicles and PD-L1 in glioblastoma-mediated immunosuppressive monocyte induction. Neuro-Oncology 2020, 22, 967–978. [Google Scholar] [CrossRef]
  78. 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, 1–14. [Google Scholar] [CrossRef]
  79. Liu, Q.W.; Chen, Y.; Li, J.Y.; Xiao, L.; Zhang, W.J.; Zhao, J.L.; Gu, H.C.; Wu, H.Y.; Zuo, G.S.L.; Deng, K.Y. Bone marrow cells are differentiated into MDSCs by BCC-Ex through down-regulating the expression of CXCR4 and activating STAT3 signalling pathway. J. Cell. Mol. Med. 2021, 25, 5497–5510. [Google Scholar] [CrossRef]
  80. 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]
  81. Qiu, W.; Guo, X.; Li, B.; Wang, J.; Qi, Y.; Chen, Z.; Zhao, R.; Deng, L.; Qian, M.; Wang, S. Exosomal miR-1246 from glioma patient body fluids drives the differentiation and activation of myeloid-derived suppressor cells. Mol. Ther. 2021, 29, 3449–3464. [Google Scholar] [CrossRef] [PubMed]
  82. Guo, X.; Qiu, W.; Wang, J.; Liu, Q.; Qian, M.; Wang, S.; Zhang, Z.; Gao, X.; Chen, Z.; Guo, Q. Glioma exosomes mediate the expansion and function of myeloid-derived suppressor cells through microRNA-29a/Hbp1 and microRNA-92a/Prkar1a pathways. Int. J. Cancer 2019, 144, 3111–3126. [Google Scholar] [CrossRef] [PubMed]
  83. Domenis, R.; Cesselli, D.; Toffoletto, B.; Bourkoula, E.; Caponnetto, F.; Manini, I.; Beltrami, A.P.; Ius, T.; Skrap, M.; Di Loreto, C. Systemic T cells immunosuppression of glioma stem cell-derived exosomes is mediated by monocytic myeloid-derived suppressor cells. PLoS ONE 2017, 12, e0169932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhang, X.; Li, F.; Tang, Y.; Ren, Q.; Xiao, B.; Wan, Y.; Jiang, S. miR-21a in exosomes from Lewis lung carcinoma cells accelerates tumor growth through targeting PDCD4 to enhance expansion of myeloid-derived suppressor cells. Oncogene 2020, 39, 6354–6369. [Google Scholar] [CrossRef] [PubMed]
  85. Krysko, D.V.; Garg, A.D.; Kaczmarek, A.; Krysko, O.; Agostinis, P.; Vandenabeele, P. Immunogenic cell death and DAMPs in cancer therapy. Nat. Rev. Cancer 2012, 12, 860–875. [Google Scholar] [CrossRef]
  86. Guo, Y.; Wang, S.-Z.; Zhang, X.; Jia, H.-R.; Zhu, Y.-X.; Zhang, X.; Gao, G.; Jiang, Y.-W.; Li, C.; Chen, X. In situ generation of micrometer-sized tumor cell-derived vesicles as autologous cancer vaccines for boosting systemic immune responses. Nat. Commun. 2022, 13, 1–20. [Google Scholar] [CrossRef]
  87. Hiam-Galvez, K.J.; Allen, B.M.; Spitzer, M.H. Systemic immunity in cancer. Nat. Rev. Cancer 2021, 21, 345–359. [Google Scholar] [CrossRef]
  88. Łuksza, M.; Sethna, Z.M.; Rojas, L.A.; Lihm, J.; Bravi, B.; Elhanati, Y.; Soares, K.; Amisaki, M.; Dobrin, A.; Hoyos, D. Neoantigen quality predicts immunoediting in survivors of pancreatic cancer. Nature 2022, 606, 389–395. [Google Scholar] [CrossRef]
  89. DuPage, M.; Mazumdar, C.; Schmidt, L.M.; Cheung, A.F.; Jacks, T. Expression of tumour-specific antigens underlies cancer immunoediting. Nature 2012, 482, 405–409. [Google Scholar] [CrossRef] [Green Version]
  90. Lakatos, E.; Williams, M.J.; Schenck, R.O.; Cross, W.C.; Househam, J.; Zapata, L.; Werner, B.; Gatenbee, C.; Robertson-Tessi, M.; Barnes, C.P. Evolutionary dynamics of neoantigens in growing tumors. Nat. Genet. 2020, 52, 1057–1066. [Google Scholar] [CrossRef]
  91. Rao, S.; Gharib, K.; Han, A. Cancer immunosurveillance by T cells. Int. Rev. Cell Mol. Biol. 2019, 342, 149–173. [Google Scholar] [PubMed]
  92. Hu, Q.; Hong, Y.; Qi, P.; Lu, G.; Mai, X.; Xu, S.; He, X.; Guo, Y.; Gao, L.; Jing, Z. Atlas of breast cancer infiltrated B-lymphocytes revealed by paired single-cell RNA-sequencing and antigen receptor profiling. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef] [PubMed]
  93. Cole, K.E.; Strick, C.A.; Paradis, T.J.; Ogborne, K.T.; Loetscher, M.; Gladue, R.P.; Lin, W.; Boyd, J.G.; Moser, B.; Wood, D.E. Interferon–inducible T cell alpha chemoattractant (I-TAC): A novel Non-ELR CXC Chemokine with potent activity on activated T cells through selective high affinity binding to CXCR3. J. Exp. Med. 1998, 187, 2009–2021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Yan, Y.; Zheng, L.; Du, Q.; Yazdani, H.; Dong, K.; Guo, Y.; Geller, D.A. Interferon regulatory factor 1 (IRF-1) activates anti-tumor immunity via CXCL10/CXCR3 axis in hepatocellular carcinoma (HCC). Cancer Lett. 2021, 506, 95–106. [Google Scholar] [CrossRef] [PubMed]
  95. Dunn, G.P.; Bruce, A.T.; Sheehan, K.C.; Shankaran, V.; Uppaluri, R.; Bui, J.D.; Diamond, M.S.; Koebel, C.M.; Arthur, C.; White, J.M. A critical function for type I interferons in cancer immunoediting. Nat. Immunol. 2005, 6, 722–729. [Google Scholar] [CrossRef]
  96. Jan, A.T.; Rahman, S.; Khan, S.; Tasduq, S.A.; Choi, I. Biology, pathophysiological role, and clinical implications of exosomes: A critical appraisal. Cells 2019, 8, 99. [Google Scholar] [CrossRef] [Green Version]
  97. Yang, M.-q.; Du, Q.; Varley, P.R.; Goswami, J.; Liang, Z.; Wang, R.; Li, H.; Stolz, D.B.; Geller, D.A. Interferon regulatory factor 1 priming of tumour-derived exosomes enhances the antitumour immune response. Br. J. Cancer 2018, 118, 62–71. [Google Scholar] [CrossRef] [Green Version]
  98. Hartman, Z.C.; Wei, J.; Glass, O.K.; Guo, H.; Lei, G.; Yang, X.-Y.; Osada, T.; Hobeika, A.; Delcayre, A.; Le Pecq, J.-B. Increasing vaccine potency through exosome antigen targeting. Vaccine 2011, 29, 9361–9367. [Google Scholar] [CrossRef] [Green Version]
  99. Logozzi, M.; Capasso, C.; Di Raimo, R.; Del Prete, S.; Mizzoni, D.; Falchi, M.; Supuran, C.T.; Fais, S. Prostate cancer cells and exosomes in acidic condition show increased carbonic anhydrase IX expression and activity. J. Enzym. Inhib. Med. Chem. 2019, 34, 272–278. [Google Scholar] [CrossRef] [Green Version]
  100. Shi, S.; Rao, Q.; Zhang, C.; Zhang, X.; Qin, Y.; Niu, Z. Dendritic cells pulsed with exosomes in combination with PD-1 antibody increase the efficacy of sorafenib in hepatocellular carcinoma model. Transl. Oncol. 2018, 11, 250–258. [Google Scholar] [CrossRef]
  101. Wolfers, J.; Lozier, A.; Raposo, G.; Regnault, A.; Thery, C.; Masurier, C.; Flament, C.; Pouzieux, S.; Faure, F.; Tursz, T. Tumor-derived exosomes are a source of shared tumor rejection antigens for CTL cross-priming. Nat. Med. 2001, 7, 297–303. [Google Scholar] [CrossRef] [PubMed]
  102. 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]
  103. Segura, E.; Guérin, C.; Hogg, N.; Amigorena, S.; Théry, C. CD8+ dendritic cells use LFA-1 to capture MHC-peptide complexes from exosomes in vivo. J. Immunol. 2007, 179, 1489–1496. [Google Scholar] [CrossRef] [Green Version]
  104. Liu, H.; Chen, L.; Peng, Y.; Yu, S.; Liu, J.; Wu, L.; Zhang, L.; Wu, Q.; Chang, X.; Yu, X. Dendritic cells loaded with tumor derived exosomes for cancer immunotherapy. Oncotarget 2018, 9, 2887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ren, G.; Wang, Y.; Yuan, S.; Wang, B. Dendritic cells loaded with HeLa-derived exosomes simulate an antitumor immune response. Oncol. Lett. 2018, 15, 6636–6640. [Google Scholar] [CrossRef] [PubMed]
  106. Dai, S.; Wan, T.; Wang, B.; Zhou, X.; Xiu, F.; Chen, T.; Wu, Y.; Cao, X. More efficient induction of HLA-A* 0201-restricted and carcinoembryonic antigen (CEA)–specific CTL response by immunization with exosomes prepared from heat-stressed CEA-positive tumor cells. Clin. Cancer Res. 2005, 11, 7554–7563. [Google Scholar] [CrossRef] [Green Version]
  107. Diamond, J.M.; Vanpouille-Box, C.; Spada, S.; Rudqvist, N.-P.; Chapman, J.R.; Ueberheide, B.M.; Pilones, K.A.; Sarfraz, Y.; Formenti, S.C.; Demaria, S. Exosomes shuttle TREX1-sensitive IFN-stimulatory dsDNA from irradiated cancer cells to DCs. Cancer Immunol. Res. 2018, 6, 910–920. [Google Scholar] [CrossRef] [Green Version]
  108. Hao, S.; Bai, O.; Yuan, J.; Qureshi, M.; Xiang, J. Dendritic cell-derived exosomes stimulate stronger CD8+ CTL responses and antitumor immunity than tumor cell-derived exosomes. Cell Mol. Immunol. 2006, 3, 205–211. [Google Scholar]
  109. Graner, M.W.; Alzate, O.; Dechkovskaia, A.M.; Keene, J.D.; Sampson, J.H.; Mitchell, D.A.; Bigner, D.D. Proteomic and immunologic analyses of brain tumor exosomes. FASEB J. 2009, 23, 1541–1557. [Google Scholar] [CrossRef] [Green Version]
  110. Garg, A.; Martin, S.; Golab, J.; Agostinis, P. Danger signalling during cancer cell death: Origins, plasticity and regulation. Cell Death Differ. 2014, 21, 26–38. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, G.; Liu, Z.; Ding, H.; Zhou, Y.; Doan, H.A.; Sin, K.W.T.; Zhu, Z.J.; Flores, R.; Wen, Y.; Gong, X. Tumor induces muscle wasting in mice through releasing extracellular Hsp70 and Hsp90. Nat. Commun. 2017, 8, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Wang, Z.; Yang, C.; Li, L.; Jin, X.; Zhang, Z.; Zheng, H.; Pan, J.; Shi, L.; Jiang, Z.; Su, K. Tumor-derived HMGB1 induces CD62L dim neutrophil polarization and promotes lung metastasis in triple-negative breast cancer. Oncogenesis 2020, 9, 1–17. [Google Scholar] [CrossRef] [PubMed]
  113. Sachet, M.; Liang, Y.Y.; Oehler, R. The immune response to secondary necrotic cells. Apoptosis 2017, 22, 1189–1204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Gamrekelashvili, J.; Ormandy, L.A.; Heimesaat, M.M.; Kirschning, C.J.; Manns, M.P.; Korangy, F.; Greten, T.F. Primary sterile necrotic cells fail to cross-prime CD8+ T cells. Oncoimmunology 2012, 1, 1017–1026. [Google Scholar] [CrossRef] [Green Version]
  115. Panaretakis, T.; Kepp, O.; Brockmeier, U.; Tesniere, A.; Bjorklund, A.C.; Chapman, D.C.; Durchschlag, M.; Joza, N.; Pierron, G.; Van Endert, P. Mechanisms of pre-apoptotic calreticulin exposure in immunogenic cell death. EMBO J. 2009, 28, 578–590. [Google Scholar] [CrossRef] [Green Version]
  116. Turubanova, V.D.; Balalaeva, I.V.; Mishchenko, T.A.; Catanzaro, E.; Alzeibak, R.; Peskova, N.N.; Efimova, I.; Bachert, C.; Mitroshina, E.V.; Krysko, O. Immunogenic cell death induced by a new photodynamic therapy based on photosens and photodithazine. J. Immunother. Cancer 2019, 7, 1–13. [Google Scholar] [CrossRef]
  117. Garg, A.D.; Krysko, D.V.; Vandenabeele, P.; Agostinis, P. The emergence of phox-ER stress induced immunogenic apoptosis. Oncoimmunology 2012, 1, 786–788. [Google Scholar] [CrossRef] [Green Version]
  118. Choi, M.E.; Price, D.R.; Ryter, S.W.; Choi, A.M. Necroptosis: A crucial pathogenic mediator of human disease. JCI Insight 2019, 4, e128834. [Google Scholar] [CrossRef] [Green Version]
  119. Hou, L.; Yang, Z.; Wang, Z.; Zhang, X.; Zhao, Y.; Yang, H.; Zheng, B.; Tian, W.; Wang, S.; He, Z. NLRP3/ASC-mediated alveolar macrophage pyroptosis enhances HMGB1 secretion in acute lung injury induced by cardiopulmonary bypass. Lab. Investig. 2018, 98, 1052–1064. [Google Scholar] [CrossRef]
  120. Yang, D.; He, Y.; Muñoz-Planillo, R.; Liu, Q.; Núñez, G. Caspase-11 requires the pannexin-1 channel and the purinergic P2X7 pore to mediate pyroptosis and endotoxic shock. Immunity 2015, 43, 923–932. [Google Scholar] [CrossRef] [Green Version]
  121. Li, W.; Deng, M.; Loughran, P.A.; Yang, M.; Lin, M.; Yang, C.; Gao, W.; Jin, S.; Li, S.; Cai, J. LPS induces active HMGB1 release from hepatocytes into exosomes through the coordinated activities of TLR4 and caspase-11/GSDMD signaling. Front. Immunol. 2020, 11, 229. [Google Scholar] [CrossRef] [PubMed]
  122. Nair, R.R.; Mazza, D.; Brambilla, F.; Gorzanelli, A.; Agresti, A.; Bianchi, M.E. LPS-challenged macrophages release microvesicles coated with histones. Front. Immunol. 2018, 9, 1463. [Google Scholar] [CrossRef] [PubMed]
  123. Vulpis, E.; Cecere, F.; Molfetta, R.; Soriani, A.; Fionda, C.; Peruzzi, G.; Caracciolo, G.; Palchetti, S.; Masuelli, L.; Simonelli, L. Genotoxic stress modulates the release of exosomes from multiple myeloma cells capable of activating NK cell cytokine production: Role of HSP70/TLR2/NF-kB axis. Oncoimmunology 2017, 6, e1279372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Jella, K.; Nasti, T.; Li, Z.; Lawson, D.; Ahmed, R.; Dynan, W.; Khan, M. Post-irradiated tumor-derived exosomes lead to melanoma tumor growth delay, potentially mediated by death associated molecular pattern (damps) proteins. Int. J. Radiat. Oncol. Biol. Phys. 2018, 102, S155. [Google Scholar] [CrossRef] [Green Version]
  125. 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] [PubMed]
  126. Xie, Y.; Bai, O.; Zhang, H.; Yuan, J.; Zong, S.; Chibbar, R.; Slattery, K.; Qureshi, M.; Wei, Y.; Deng, Y. Membrane-bound HSP70-engineered myeloma cell-derived exosomes stimulate more efficient CD8+ CTL-and NK-mediated antitumour immunity than exosomes released from heat-shocked tumour cells expressing cytoplasmic HSP70. J. Cell Mol. Med. 2010, 14, 2655–2666. [Google Scholar] [CrossRef] [Green Version]
  127. Cho, J.-a.; Lee, Y.-S.; Kim, S.-H.; Ko, J.-K.; Kim, C.-W. MHC independent anti-tumor immune responses induced by Hsp70-enriched exosomes generate tumor regression in murine models. Cancer Lett. 2009, 275, 256–265. [Google Scholar] [CrossRef]
  128. 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] [Green Version]
  129. Lv, L.-H.; Wan, Y.-L.; Lin, Y.; Zhang, W.; Yang, M.; Li, G.-L.; Lin, H.-M.; Shang, C.-Z.; Chen, Y.-J.; Min, J. Anticancer drugs cause release of exosomes with heat shock proteins from human hepatocellular carcinoma cells that elicit effective natural killer cell antitumor responses in vitro. J. Biol. Chem. 2012, 287, 15874–15885. [Google Scholar] [CrossRef] [Green Version]
  130. Koebel, C.M.; Vermi, W.; Swann, J.B.; Zerafa, N.; Rodig, S.J.; Old, L.J.; Smyth, M.J.; Schreiber, R.D. Adaptive immunity maintains occult cancer in an equilibrium state. Nature 2007, 450, 903–907. [Google Scholar] [CrossRef]
  131. Neophytou, C.M.; Kyriakou, T.-C.; Papageorgis, P. Mechanisms of metastatic tumor dormancy and implications for cancer therapy. Int. J. Mol. Sci. 2019, 20, 6158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Wang, J.; Zhao, X.; Wang, Y.; Ren, F.; Sun, D.; Yan, Y.; Kong, X.; Bu, J.; Liu, M.; Xu, S. circRNA-002178 act as a ceRNA to promote PDL1/PD1 expression in lung adenocarcinoma. Cell Death Dis. 2020, 11, 1–11. [Google Scholar] [CrossRef] [PubMed]
  133. Grange, C.; Tapparo, M.; Tritta, S.; Deregibus, M.C.; Battaglia, A.; Gontero, P.; Frea, B.; Camussi, G. Role of HLA-G and extracellular vesicles in renal cancer stem cell-induced inhibition of dendritic cell differentiation. BMC Cancer 2015, 15, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ono, M.; Kosaka, N.; Tominaga, N.; Yoshioka, Y.; Takeshita, F.; Takahashi, R.-u.; Yoshida, M.; Tsuda, H.; Tamura, K.; Ochiya, T. Exosomes from bone marrow mesenchymal stem cells contain a microRNA that promotes dormancy in metastatic breast cancer cells. Sci. Signal. 2014, 7, ra63. [Google Scholar] [CrossRef]
  135. Sandiford, O.A.; Donnelly, R.J.; Markos, H.; Burgmeyer, L.M.; Sinha, G.; Pamarthi, S.H.; Sherman, L.S.; Ferrer, A.I.; DeVore, D.E.; Patel, S.A. Mesenchymal Stem Cell–Secreted Extracellular Vesicles Instruct Stepwise Dedifferentiation of Breast Cancer Cells into Dormancy at the Bone Marrow Perivascular Region. Cancer Res. 2021, 81, 1567–1582. [Google Scholar] [CrossRef] [PubMed]
  136. Chen, G.; Huang, A.C.; Zhang, W.; Zhang, G.; Wu, M.; Xu, W.; Yu, Z.; Yang, J.; Wang, B.; Sun, H. Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature 2018, 560, 382–386. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, Y.; Liang, X.; Yin, X.; Lv, J.; Tang, K.; Ma, J.; Ji, T.; Zhang, H.; Dong, W.; Jin, X. Blockade of IDO-kynurenine-AhR metabolic circuitry abrogates IFN-γ-induced immunologic dormancy of tumor-repopulating cells. Nat. Commun. 2017, 8, 1–15. [Google Scholar] [CrossRef]
  138. 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. 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]
  139. Gao, L.; Wang, L.; Dai, T.; Jin, K.; Zhang, Z.; Wang, S.; Xie, F.; Fang, P.; Yang, B.; Huang, H. Tumor-derived exosomes antagonize innate antiviral immunity. Nat. Immunol. 2018, 19, 233–245. [Google Scholar] [CrossRef]
  140. Bakhoum, S.F.; Ngo, B.; Laughney, A.M.; Cavallo, J.-A.; Murphy, C.J.; Ly, P.; Shah, P.; Sriram, R.K.; Watkins, T.B.; Taunk, N.K. Chromosomal instability drives metastasis through a cytosolic DNA response. Nature 2018, 553, 467–472. [Google Scholar] [CrossRef] [Green Version]
  141. Dhatchinamoorthy, K.; Colbert, J.D.; Rock, K.L. Cancer Immune Evasion Through Loss of MHC Class I Antigen Presentation. Front. Immunol. 2021, 12, 469. [Google Scholar] [CrossRef] [PubMed]
  142. 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, 1–16. [Google Scholar] [CrossRef] [PubMed]
  143. Willingham, S.B.; Volkmer, J.-P.; Gentles, A.J.; Sahoo, D.; Dalerba, P.; Mitra, S.S.; Wang, J.; Contreras-Trujillo, H.; Martin, R.; Cohen, J.D. The CD47-signal regulatory protein alpha (SIRPa) interaction is a therapeutic target for human solid tumors. Proc. Natl. Acad. Sci. USA 2012, 109, 6662–6667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mannino, M.H.; Zhu, Z.; Xiao, H.; Bai, Q.; Wakefield, M.R.; Fang, Y. The paradoxical role of IL-10 in immunity and cancer. Cancer Lett. 2015, 367, 103–107. [Google Scholar] [CrossRef]
  145. Flavell, R.A.; Sanjabi, S.; Wrzesinski, S.H.; Licona-Limón, P. The polarization of immune cells in the tumour environment by TGFβ. Nat. Rev. Immunol. 2010, 10, 554–567. [Google Scholar] [CrossRef] [Green Version]
  146. Pardoll, D.M. The blockade of immune checkpoints in cancer immunotherapy. Nat. Rev. Cancer 2012, 12, 252–264. [Google Scholar] [CrossRef] [Green Version]
  147. Dhani, S.; Zhao, Y.; Zhivotovsky, B. A long way to go: Caspase inhibitors in clinical use. Cell Death Dis. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  148. Malchow, S.; Leventhal, D.S.; Nishi, S.; Fischer, B.I.; Shen, L.; Paner, G.P.; Amit, A.S.; Kang, C.; Geddes, J.E.; Allison, J.P. Aire-dependent thymic development of tumor-associated regulatory T cells. Science 2013, 339, 1219–1224. [Google Scholar] [CrossRef] [Green Version]
  149. Yu, S.; Liu, C.; Su, K.; Wang, J.; Liu, Y.; Zhang, L.; Li, C.; Cong, Y.; Kimberly, R.; Grizzle, W.E. Tumor exosomes inhibit differentiation of bone marrow dendritic cells. J. Immunol. 2007, 178, 6867–6875. [Google Scholar] [CrossRef] [Green Version]
  150. Maus, R.L.; Jakub, J.W.; Nevala, W.K.; Christensen, T.A.; Noble-Orcutt, K.; Sachs, Z.; Hieken, T.J.; Markovic, S.N. Human melanoma-derived extracellular vesicles regulate dendritic cell maturation. Front. Immunol. 2017, 8, 358. [Google Scholar] [CrossRef] [Green Version]
  151. Liu, Y.; Yin, Z.; Lu, P.; Ma, Y.; Luo, B.; Xiang, L.; Zhang, W.; He, Y.; Liang, X. Lung carcinoma cells secrete exosomal MALAT1 to inhibit dendritic cell phagocytosis, inflammatory response, costimulatory molecule expression and promote dendritic cell autophagy via AKT/mTOR pathway. OncoTargets Ther. 2020, 13, 10693. [Google Scholar] [CrossRef] [PubMed]
  152. 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] [PubMed]
  153. Yin, X.; Zeng, W.; Wu, B.; Wang, L.; Wang, Z.; Tian, H.; Wang, L.; Jiang, Y.; Clay, R.; Wei, X. PPARα inhibition overcomes tumor-derived exosomal lipid-induced dendritic cell dysfunction. Cell Rep. 2020, 33, 108278. [Google Scholar] [CrossRef] [PubMed]
  154. Zhao, F.; Xiao, C.; Evans, K.S.; Theivanthiran, T.; DeVito, N.; Holtzhausen, A.; Liu, J.; Liu, X.; Boczkowski, D.; Nair, S. Paracrine Wnt5a-β-catenin signaling triggers a metabolic program that drives dendritic cell tolerization. Immunity 2018, 48, 147–160.e147. [Google Scholar] [CrossRef] [Green Version]
  155. Wculek, S.K.; Khouili, S.C.; Priego, E.; Heras-Murillo, I.; Sancho, D. Metabolic control of dendritic cell functions: Digesting information. Front. Immunol. 2019, 10, 775. [Google Scholar] [CrossRef] [Green Version]
  156. 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]
  157. Gao, J.; Qiu, X.; Li, X.; Fan, H.; Zhang, F.; Lv, T.; Song, Y. Expression profiles and clinical value of plasma exosomal Tim-3 and Galectin-9 in non-small cell lung cancer. Biochem. Biophys. Res. Commun. 2018, 498, 409–415. [Google Scholar] [CrossRef]
  158. Haderk, F.; Schulz, R.; Iskar, M.; Cid, L.L.; Worst, T.; Willmund, K.V.; Schulz, A.; Warnken, U.; Seiler, J.; Benner, A. Tumor-derived exosomes modulate PD-L1 expression in monocytes. Sci. Immunol. 2017, 2, eaah5509. [Google Scholar] [CrossRef] [Green Version]
  159. Maus, R.L.; Jakub, J.W.; Hieken, T.J.; Nevala, W.K.; Christensen, T.A.; Sutor, S.L.; Flotte, T.J.; Markovic, S.N. Identification of novel, immune-mediating extracellular vesicles in human lymphatic effluent draining primary cutaneous melanoma. Oncoimmunology 2019, 8, e1667742. [Google Scholar] [CrossRef] [Green Version]
  160. Yang, C.; Kim, S.-H.; Bianco, N.R.; Robbins, P.D. Tumor-derived exosomes confer antigen-specific immunosuppression in a murine delayed-type hypersensitivity model. PLoS ONE 2011, 6, e22517. [Google Scholar] [CrossRef] [Green Version]
  161. Shen, Y.; Guo, D.; Weng, L.; Wang, S.; Ma, Z.; Yang, Y.; Wang, P.; Wang, J.; Cai, Z. Tumor-derived exosomes educate dendritic cells to promote tumor metastasis via HSP72/HSP105-TLR2/TLR4 pathway. Oncoimmunology 2017, 6, e1362527. [Google Scholar] [CrossRef] [PubMed]
  162. Prieto, D.; Sotelo, N.; Seija, N.; Sernbo, S.; Abreu, C.; Durán, R.; Gil, M.; Sicco, E.; Irigoin, V.; Oliver, C. S100-A9 protein in exosomes from chronic lymphocytic leukemia cells promotes NF-κB activity during disease progression. Blood J. Am. Soc. Hematol. 2017, 130, 777–788. [Google Scholar] [CrossRef] [PubMed]
  163. Cheng, P.; Corzo, C.A.; Luetteke, N.; Yu, B.; Nagaraj, S.; Bui, M.M.; Ortiz, M.; Nacken, W.; Sorg, C.; Vogl, T. Inhibition of dendritic cell differentiation and accumulation of myeloid-derived suppressor cells in cancer is regulated by S100A9 protein. J. Exp. Med. 2008, 205, 2235–2249. [Google Scholar] [CrossRef] [PubMed]
  164. Kaur, S.; Elkahloun, A.G.; Arakelyan, A.; Young, L.; Myers, T.G.; Otaizo-Carrasquero, F.; Wu, W.; Margolis, L.; Roberts, D.D. CD63, MHC class 1, and CD47 identify subsets of extracellular vesicles containing distinct populations of noncoding RNAs. Sci. Rep. 2018, 8, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. 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. [Google Scholar] [CrossRef] [Green Version]
  166. Thommen, D.S.; Schumacher, T.N. T cell dysfunction in cancer. Cancer Cell 2018, 33, 547–562. [Google Scholar] [CrossRef] [Green Version]
  167. 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]
  168. Huber, V.; Fais, S.; Iero, M.; Lugini, L.; Canese, P.; Squarcina, P.; Zaccheddu, A.; Colone, M.; Arancia, G.; Gentile, M. Human colorectal cancer cells induce T-cell death through release of proapoptotic microvesicles: Role in immune escape. Gastroenterology 2005, 128, 1796–1804. [Google Scholar] [CrossRef]
  169. Abusamra, A.J.; Zhong, Z.; Zheng, X.; Li, M.; Ichim, T.E.; Chin, J.L.; Min, W.-P. Tumor exosomes expressing Fas ligand mediate CD8+ T-cell apoptosis. Blood Cells Mol. Dis. 2005, 35, 169–173. [Google Scholar] [CrossRef]
  170. 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 γδ T cell equilibrium via tumor-derived exosomes. Oncogene 2019, 38, 2830–2843. [Google Scholar] [CrossRef]
  171. Klibi, J.; Niki, T.; Riedel, A.; Pioche-Durieu, C.; Souquere, S.; Rubinstein, E.; Le Moulec, S.; Guigay, J.; Hirashima, M.; Guemira, F. Blood diffusion and Th1-suppressive effects of galectin-9–containing exosomes released by Epstein-Barr virus–infected nasopharyngeal carcinoma cells. Blood J. Am. Soc. Hematol. 2009, 113, 1957–1966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Kim, J.W.; Wieckowski, E.; Taylor, D.D.; Reichert, T.E.; Watkins, S.; Whiteside, T.L. Fas ligand–positive membranous vesicles isolated from sera of patients with oral cancer induce apoptosis of activated T lymphocytes. Clin. Cancer Res. 2005, 11, 1010–1020. [Google Scholar] [CrossRef] [PubMed]
  173. Taylor, D.; Gercel-Taylor, C. Tumour-derived exosomes and their role in cancer-associated T-cell signalling defects. Br. J. Cancer 2005, 92, 305–311. [Google Scholar] [CrossRef] [PubMed]
  174. 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. Immune suppressive effects of plasma-derived exosome populations in head and neck cancer. Cancers 2020, 12, 1997. [Google Scholar] [CrossRef]
  175. Shao, Q.; Deng, L.; Liu, H.; Liu, Z.; Chen, J.; Jiang, F.; Yan, S.; Fu, R. Involvement of MM cell-derived exosomes in T lymphocytes immune responses. Oncol. Lett. 2020, 20, 1. [Google Scholar] [CrossRef]
  176. 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. 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] [Green Version]
  177. Poggio, M.; Hu, T.; Pai, C.-C.; Chu, B.; Belair, C.D.; Chang, A.; Montabana, E.; Lang, U.E.; Fu, Q.; Fong, L. Suppression of exosomal PD-L1 induces systemic anti-tumor immunity and memory. Cell 2019, 177, 414–427.e413. [Google Scholar] [CrossRef] [Green Version]
  178. Zhou, K.; Guo, S.; Li, F.; Sun, Q.; Liang, G. Exosomal PD-L1: New insights into tumor immune escape mechanisms and therapeutic strategies. Front. Cell Dev. Biol. 2020, 8, 569219. [Google Scholar] [CrossRef]
  179. Taylor, D.D.; Gerçel-Taylor, Ç.; Lyons, K.S.; Stanson, J.; Whiteside, T.L. T-cell apoptosis and suppression of T-cell receptor/CD3-ζ by Fas ligand-containing membrane vesicles shed from ovarian tumors. Clin. Cancer Res. 2003, 9, 5113–5119. [Google Scholar]
  180. Söderberg, A.; Barral, A.M.; Söderström, M.; Sander, B.; Rosén, A. Redox-signaling transmitted in trans to neighboring cells by melanoma-derived TNF-containing exosomes. Free Radic. Biol. Med. 2007, 43, 90–99. [Google Scholar] [CrossRef]
  181. Shen, T.; Huang, Z.; Shi, C.; Pu, X.; Xu, X.; Wu, Z.; Ding, G.; Cao, L. Pancreatic cancer-derived exosomes induce apoptosis of T lymphocytes through the p38 MAPK-mediated endoplasmic reticulum stress. FASEB J. 2020, 34, 8442–8458. [Google Scholar] [CrossRef] [PubMed]
  182. Ye, S.-b.; Li, Z.-L.; Luo, D.-h.; Huang, B.-j.; Chen, Y.-S.; Zhang, X.-s.; Cui, J.; Zeng, Y.-x.; Li, J. Tumor-derived exosomes promote tumor progression and T-cell dysfunction through the regulation of enriched exosomal microRNAs in human nasopharyngeal carcinoma. Oncotarget 2014, 5, 5439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. 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.; Zhang, X.S. 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] [PubMed]
  184. Clayton, A.; Mitchell, J.P.; 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] [Green Version]
  185. 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, 1–15. [Google Scholar] [CrossRef]
  186. Jiang, Y.; Li, Y.; Zhu, B. T-cell exhaustion in the tumor microenvironment. Cell Death Dis. 2015, 6, e1792. [Google Scholar] [CrossRef] [Green Version]
  187. McLane, L.M.; Abdel-Hakeem, M.S.; Wherry, E.J. CD8 T cell exhaustion during chronic viral infection and cancer. Annu. Rev. Immunol. 2019, 37, 457–495. [Google Scholar] [CrossRef] [Green Version]
  188. Wang, X.; Shen, H.; Zhangyuan, G.; Huang, R.; Zhang, W.; He, Q.; Jin, K.; Zhuo, H.; Zhang, Z.; Wang, J. 14-3-3ζ delivered by hepatocellular carcinoma-derived exosomes impaired anti-tumor function of tumor-infiltrating T lymphocytes. Cell Death Dis. 2018, 9, 1–14. [Google Scholar] [CrossRef] [Green Version]
  189. Kalvala, A.; Wallet, P.; Yang, L.; Wang, C.; Li, H.; Nam, A.; Nathan, A.; Mambetsariev, I.; Poroyko, V.; Gao, H. 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]
  190. 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]
  191. Czystowska-Kuzmicz, M.; Sosnowska, A.; Nowis, D.; Ramji, K.; Szajnik, M.; Chlebowska-Tuz, J.; Wolinska, E.; Gaj, P.; Grazul, M.; Pilch, Z. Small extracellular vesicles containing arginase-1 suppress T-cell responses and promote tumor growth in ovarian carcinoma. Nat. Commun. 2019, 10, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. 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] [Green Version]
  193. Yamada, N.; Kuranaga, Y.; Kumazaki, M.; Shinohara, H.; Taniguchi, K.; Akao, Y. Colorectal cancer cell-derived extracellular vesicles induce phenotypic alteration of T cells into tumor-growth supporting cells with transforming growth factor-β1-mediated suppression. Oncotarget 2016, 7, 27033. [Google Scholar] [CrossRef]
  194. Hellwinkel, J.E.; Redzic, J.S.; Harland, T.A.; Gunaydin, D.; Anchordoquy, T.J.; Graner, M.W. Glioma-derived extracellular vesicles selectively suppress immune responses. Neuro-Oncology 2015, 18, 497–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Mrizak, D.; Martin, N.; Barjon, C.; Jimenez-Pailhes, A.-S.; Mustapha, R.; Niki, T.; Guigay, J.; Pancré, V.; de Launoit, Y.; Busson, P. Effect of nasopharyngeal carcinoma-derived exosomes on human regulatory T cells. J. Natl. Cancer Inst. 2015, 107, dju363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Rong, L.; Li, R.; Li, S.; Luo, R. Immunosuppression of breast cancer cells mediated by transforming growth factor-β in exosomes from cancer cells. Oncol. Lett. 2016, 11, 500–504. [Google Scholar] [CrossRef] [Green Version]
  197. Dianat-Moghadam, H.; Mahari, A.; Heidarifard, M.; Parnianfard, N.; Pourmousavi-Kh, L.; Rahbarghazi, R.; Amoozgar, Z. NK cells-directed therapies target circulating tumor cells and metastasis. Cancer Lett. 2021, 497, 41–53. [Google Scholar] [CrossRef]
  198. Bottos, A.; Gotthardt, D.; Gill, J.W.; Gattelli, A.; Frei, A.; Tzankov, A.; Sexl, V.; Wodnar-Filipowicz, A.; Hynes, N.E. Decreased NK-cell tumour immunosurveillance consequent to JAK inhibition enhances metastasis in breast cancer models. Nat. Commun. 2016, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  199. Kruse, P.H.; Matta, J.; Ugolini, S.; Vivier, E. Natural cytotoxicity receptors and their ligands. Immunol. Cell Biol. 2014, 92, 221–229. [Google Scholar] [CrossRef]
  200. Hedlund, M.; Stenqvist, A.-C.; Nagaeva, O.; Kjellberg, L.; Wulff, M.; Baranov, V.; Mincheva-Nilsson, L. Human placenta expresses and secretes NKG2D ligands via exosomes that down-modulate the cognate receptor expression: Evidence for immunosuppressive function. J. Immunol. 2009, 183, 340–351. [Google Scholar] [CrossRef] [Green Version]
  201. Ashiru, O.; Boutet, P.; Fernández-Messina, L.; Agüera-González, S.; Skepper, J.N.; Valés-Gómez, 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] [PubMed] [Green Version]
  202. Labani-Motlagh, A.; Israelsson, P.; Ottander, U.; Lundin, E.; Nagaev, I.; Nagaeva, O.; Dehlin, E.; Baranov, V.; Mincheva-Nilsson, L. Differential expression of ligands for NKG2D and DNAM-1 receptors by epithelial ovarian cancer-derived exosomes and its influence on NK cell cytotoxicity. Tumor Biol. 2016, 37, 5455–5466. [Google Scholar] [CrossRef]
  203. Lundholm, M.; Schröder, M.; Nagaeva, O.; Baranov, V.; Widmark, A.; Mincheva-Nilsson, L.; Wikström, P. Prostate tumor-derived exosomes down-regulate NKG2D expression on natural killer cells and CD8+ T cells: Mechanism of immune evasion. PLoS ONE 2014, 9, e108925. [Google Scholar] [CrossRef] [PubMed]
  204. Clayton, A.; Mitchell, J.P.; 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]
  205. López-Cobo, S.; Campos-Silva, C.; Moyano, A.; Oliveira-Rodríguez, M.; Paschen, A.; Yáñez-Mó, M.; Blanco-López, M.C.; Valés-Gómez, M. Immunoassays for scarce tumour-antigens in exosomes: Detection of the human NKG2D-Ligand, MICA, in tetraspanin-containing nanovesicles from melanoma. J. Nanobiotechnol. 2018, 16, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Admyre, C.; Johansson, S.M.; Qazi, K.R.; Filén, J.-J.; Lahesmaa, R.; Norman, M.; Neve, E.P.; Scheynius, A.; Gabrielsson, S. Exosomes with immune modulatory features are present in human breast milk. J. Immunol. 2007, 179, 1969–1978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Catalano, M.; O’Driscoll, L. Inhibiting extracellular vesicles formation and release: A review of EV inhibitors. J. Extracell. Vesicles 2020, 9, 1703244. [Google Scholar] [CrossRef] [Green Version]
  208. Kulshreshtha, A.; Singh, S.; Ahmad, M.; Khanna, K.; Ahmad, T.; Agrawal, A.; Ghosh, B. Simvastatin mediates inhibition of exosome synthesis, localization and secretion via multicomponent interventions. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  209. Datta, A.; Kim, H.; Lal, M.; McGee, L.; Johnson, A.; Moustafa, A.A.; Jones, J.C.; Mondal, D.; Ferrer, M.; Abdel-Mageed, A.B. Manumycin A suppresses exosome biogenesis and secretion via targeted inhibition of Ras/Raf/ERK1/2 signaling and hnRNP H1 in castration-resistant prostate cancer cells. Cancer Lett. 2017, 408, 73–81. [Google Scholar] [CrossRef]
  210. Huda, M.N.; Nafiujjaman, M.; Deaguero, I.G.; Okonkwo, J.; Hill, M.L.; Kim, T.; Nurunnabi, M. Potential use of exosomes as diagnostic biomarkers and in targeted drug delivery: Progress in clinical and preclinical applications. ACS Biomater. Sci. Eng. 2021, 7, 2106–2149. [Google Scholar] [CrossRef]
  211. Chen, W.; Wang, J.; Shao, C.; Liu, S.; Yu, Y.; Wang, Q.; Cao, X. Efficient induction of antitumor T cell immunity by exosomes derived from heat-shocked lymphoma cells. Eur. J. Immunol. 2006, 36, 1598–1607. [Google Scholar] [CrossRef] [PubMed]
  212. Podolska, M.J.; Shan, X.; Janko, C.; Boukherroub, R.; Gaipl, U.S.; Szunerits, S.; Frey, B.; Muñoz, L.E. Graphene-induced hyperthermia (GIHT) combined with radiotherapy fosters immunogenic cell death. Front. Oncol. 2021, 11, 664615. [Google Scholar] [CrossRef] [PubMed]
  213. Guo, D.; Chen, Y.; Wang, S.; Yu, L.; Shen, Y.; Zhong, H.; Yang, Y. Exosomes from heat-stressed tumour cells inhibit tumour growth by converting regulatory T cells to Th17 cells via IL-6. Immunology 2018, 154, 132–143. [Google Scholar] [CrossRef] [PubMed]
  214. Chen, T.; Guo, J.; Yang, M.; Zhu, X.; Cao, X. Chemokine-containing exosomes are released from heat-stressed tumor cells via lipid raft-dependent pathway and act as efficient tumor vaccine. J. Immunol. 2011, 186, 2219–2228. [Google Scholar] [CrossRef] [Green Version]
  215. Toraya-Brown, S.; Fiering, S. Local tumour hyperthermia as immunotherapy for metastatic cancer. Int. J. Hyperth. 2014, 30, 531–539. [Google Scholar] [CrossRef]
  216. Guo, G.; Tan, Z.; Liu, Y.; Shi, F.; She, J. The therapeutic potential of stem cell-derived exosomes in the ulcerative colitis and colorectal cancer. Stem Cell Res. Ther. 2022, 13, 1–18. [Google Scholar] [CrossRef]
  217. Shi, X.; Sun, J.; Li, H.; Lin, H.; Xie, W.; Li, J.; Tan, W. Antitumor efficacy of interferon-γ-modified exosomal vaccine in prostate cancer. Prostate 2020, 80, 811–823. [Google Scholar] [CrossRef]
  218. Hu, W.; Huang, F.; Ning, L.; Hao, J.; Wan, J.; Hao, S. Enhanced immunogenicity of leukemia-derived exosomes via transfection with lentiviral vectors encoding costimulatory molecules. Cell Oncol. 2020, 43, 889–900. [Google Scholar] [CrossRef]
  219. Gu, X.; Erb, U.; Büchler, M.W.; Zöller, M. Improved vaccine efficacy of tumor exosome compared to tumor lysate loaded dendritic cells in mice. Int. J. Cancer 2015, 136, E74–E84. [Google Scholar] [CrossRef]
  220. Lee, Y.S.; Kim, S.H.; Cho, J.A.; Kim, C.W. Introduction of the CIITA gene into tumor cells produces exosomes with enhanced anti-tumor effects. Exp. Mol. Med. 2011, 43, 281–290. [Google Scholar] [CrossRef] [Green Version]
  221. Zuo, B.; Qi, H.; Lu, Z.; Chen, L.; Sun, B.; Yang, R.; Zhang, Y.; Liu, Z.; Gao, X.; You, A. Alarmin-painted exosomes elicit persistent antitumor immunity in large established tumors in mice. Nat. Commun. 2020, 11, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Thakur, A.; Parra, D.C.; Motallebnejad, P.; Brocchi, M.; Chen, H.J. Exosomes: Small vesicles with big roles in cancer, vaccine development, and therapeutics. Bioact. Mater. 2022, 10, 281–294. [Google Scholar] [CrossRef] [PubMed]
  223. Yang, T.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S. Exosome delivered anticancer drugs across the blood-brain barrier for brain cancer therapy in Danio rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef] [PubMed]
  224. Yong, T.; Zhang, X.; Bie, N.; Zhang, H.; Zhang, X.; Li, F.; Hakeem, A.; Hu, J.; Gan, L.; Santos, H.A. Tumor exosome-based nanoparticles are efficient drug carriers for chemotherapy. Nat. Commun. 2019, 10, 1–16. [Google Scholar]
  225. Qiao, L.; Hu, S.; Huang, K.; Su, T.; Li, Z.; Vandergriff, A.; Cores, J.; Dinh, P.-U.; Allen, T.; Shen, D. Tumor cell-derived exosomes home to their cells of origin and can be used as Trojan horses to deliver cancer drugs. Theranostics 2020, 10, 3474. [Google Scholar] [CrossRef]
Figure 1. Cancer immunoediting and involved effector molecules and receptors: (a) During the elimination phase, the innate and adaptive immune systems work together to identify and eliminate transformed cells that have evaded genetic malignant cell suppression mechanisms. The generation of tumor antigens and extracellular vesicles by cancer cells trigger immunological responses against tumors that result in the detection and elimination of nascent cancer cells by the immune system (b) if they survive the elimination phase, tumors may reach the equilibrium phase. Adaptive and innate immunity limits tumor progression in this phase, therefore tumor growth is reduced, and surviving tumor cells are kept under control and possibly kept dormant by the immune system (c) Ultimately in the escape phase, tumors overcome the immune system by immune suppression through recruiting immunosuppressive cells, and suppressing effector immune cells via different mechanisms.
Figure 1. Cancer immunoediting and involved effector molecules and receptors: (a) During the elimination phase, the innate and adaptive immune systems work together to identify and eliminate transformed cells that have evaded genetic malignant cell suppression mechanisms. The generation of tumor antigens and extracellular vesicles by cancer cells trigger immunological responses against tumors that result in the detection and elimination of nascent cancer cells by the immune system (b) if they survive the elimination phase, tumors may reach the equilibrium phase. Adaptive and innate immunity limits tumor progression in this phase, therefore tumor growth is reduced, and surviving tumor cells are kept under control and possibly kept dormant by the immune system (c) Ultimately in the escape phase, tumors overcome the immune system by immune suppression through recruiting immunosuppressive cells, and suppressing effector immune cells via different mechanisms.
Cancers 15 00082 g001
Figure 2. TEV in the cancer-immunity cycle: (1) Release of tumor antigens (TAs) along with tumor-derived extracellular vesicles (TEVs) that carry TA + DAMPs from dying cancer cells; (2) Presentation of TAs on the major histocompatibility complex (MHC) by dendritic cells; (3) T-cell receptor recognition of TAs on the MHC, leading to T-cell activation; (4) Migration of activated T cells to the tumors; (5) T-cell infiltration into the tumor; (6) Recognition of cancer antigens within the tumor; (7) Attack and killing of tumor cells.
Figure 2. TEV in the cancer-immunity cycle: (1) Release of tumor antigens (TAs) along with tumor-derived extracellular vesicles (TEVs) that carry TA + DAMPs from dying cancer cells; (2) Presentation of TAs on the major histocompatibility complex (MHC) by dendritic cells; (3) T-cell receptor recognition of TAs on the MHC, leading to T-cell activation; (4) Migration of activated T cells to the tumors; (5) T-cell infiltration into the tumor; (6) Recognition of cancer antigens within the tumor; (7) Attack and killing of tumor cells.
Cancers 15 00082 g002
Table 2. Effect of tumor-derived extracellular vesicles on fibroblast differentiation in the tumor microenvironment.
Table 2. Effect of tumor-derived extracellular vesicles on fibroblast differentiation in the tumor microenvironment.
Cancer TypeCellular SourceVesicular CargoThe Main ResultRefs.
Triple-negative breast cancerMDA-MB-231
BT-20
MDA-MB-453
MCF7
BT-474
SK-BR-3
ITGB4 proteinsEnhanced mitophagy and lactate generation in CAFs in a BNIP3L-dependent manner.[64]
TNBC MDA-MB-231
MDA-MB-468
miR-9Influenced the properties of NFs and promoted the switch to CAF, thereby leading to tumor growth.[58]
Ovarian cancerA2780
SKVO3
miR-630Raised amounts of α-SMA and FAP in NFs resulted in the differentiation of NFs into CAFs via inhibiting KLF6 and activating the NFκB pathway.[60]
Bladder cancerRT4
T24
SW1710
TGF-βTriggered the differentiation of fibroblasts to CAFs by SMAD pathway activation[63]
Colorectal cancerSW620
SW480
Vesicular cargoActivated normal quiescent fibroblasts (α-SMA−, CAV+) into CAF-like fibroblasts (α-SMA+, CAV−) with pro-proliferative and pro-angiogenic features[65]
Lung cancer (LC)A549
H460
miR-210Promoted the NFs transferring into CAFs via activating JAK2/STAT3 pathway, and ten-eleven translocation 2 (TET2)[61]
MelanomaB16BL6 eTGF-βTriggered TGF-β signaling in HUVECs and differentiated them into CAFs[19]
B16F0 Gm26809Stimulated conversion of fibroblast NIH3T3 cells into CAFs [66]
B16-F10miR-21 Stimulated invasiveness of fibroblasts by increasing matrix metalloprotein (MMP) and down-regulation of tissue inhibitor of metalloproteinase 3 (TIMP3) expressions.[59]
Breast cancerMCF-7
MDA-MB-231
Vesicular survivinConverted NFs into myofibroblasts by upregulating SOD1 and increased proliferation, EMT, and stemness.[62]
Head and neck squamous cell carcinoma (HNSCC)SAS
HSC-3
Vesicular cargoConvert normal fibroblasts into CAF-like cells and raised fibroblast proliferation, migration and activation of 11 signaling pathways (IL-6- and IL-17-related signaling)[67]
Table 4. Effect of the tumor-derived extracellular vesicles on Dendritic cells.
Table 4. Effect of the tumor-derived extracellular vesicles on Dendritic cells.
Cancer Type.Cellular SourceVesicular CargoThe Main ResultRefs.
Prostate cancerDU145PGE2Triggered the expression of CD73 and then CD39 on DCs, resulting in inhibition of TNFα- and IL-12-production via an ATP-dependent manner[156]
NSCLCBlood samples from NSCLC patientsGalectin-9 and Tim-3 Interacted with TIM-3 on DCs [157]
Renal cancerCD105+ CSCs
CD105 TCs
HLA-GDisrupted maturation of DCs and T-cell immune responses[133]
GlioblastomaCSF samples from glioma patients

GL261
U87MG
U118 MG
Galectin-9Inhibited antigen recognition, processing, and presentation by interacting with TIM-3 on DCs[142]
Ascites of glioma patientsPD-L1Impaired DCs maturation via formation of immunosuppressive monocytes[77]
Blood samples from glioma patients

GSC20
GSC267
GSC17
MEC-1
Vesicular cargoSkewed monocytes toward an immune suppressive phenotype and induced programmed PD-L1 expression on monocytes through STAT3 phosphorylation and TLR7-dependent manner[33,158]
MelanomaSKMEL28
A375
C32TG
S100, A8/A9Inhibited DCs maturation and reduced expression of CD83, CD86, Th1 polarizing chemokines (such as Flt3L, IL-15), and migration chemokines (MIP-1α and MIP-1β)[150]
lymphatic fluid sample of melanoma patients

ATCC
S100A9Inhibited DCs maturation and prepared metastatic niche in lymph nodes[159]
B16-F0TGF-β1Increased mRNA levels of IL-4 and TGF-β1 which inhibited DCs’ maturation[160]
Blood samples from melanoma patients

B16-F0
HSP72 and HSP105Induced secretion of IL-6 from DCs via TLR4- and TLR2-dependent manner activating STAT3-dependent MMP 9 activity[161]
lymphocytic leukemiaBlood samples from CLL patientsS100A8/A9CD83, CD86, IL-12, and IL-15 expressions were all downregulated via activating the NFκB pathway[162,163]
lung carcinomaLLCPD-L1Myeloid precursor cells were unable to differentiate into CD11c+ DCs in the presence of vesicular PD-L1 and resulted in DCs death[152]
LLC
A549
MALAT1 Inhibited DC function and T-cell proliferation and increased DC autophagy via AKT/mTOR Pathway[151]
Breast cancerMDA-MB-231
TS/A
Vesicular cargoInhibited the development of myeloid precursor cells into DCs by increasing IL-6 production and reducing CD83 and CD86 expression[149]
4T1PD-L1Myeloid precursor cells were unable to differentiate into CD11c+ DCs in the presence of vesicular PD-L1 and resulted in DC death[152]
Blood samples from melanoma patients

4T1
HSP72 and HSP105Promoted DCs to IL-6 secretion in a TLR2- and TLR4-dependent manner which activated STAT3-dependent MMP 9 activity[161]
Table 5. Effect of the tumor-derived extracellular vesicles on T cells.
Table 5. Effect of the tumor-derived extracellular vesicles on T cells.
Cancer TypeCellular SourceVesicular CargoMechanism of ActionRefs.
Ovarian cancerAscites of ovarian patients

OVCAR3
SKOV3
AD10
TGF-β1,
IL-10
Increased IL-10, FasL, TGF-β1, CTLA-4, which promoted Treg proliferation, suppressor activity, and Treg cell survival.[190]
Blood samples from ovarian patients

Ascites of ovarian patients

OVCAR-3
AD10
A2780
Skov3
CaOv-3
MDAH2774
OvCa-14
OVP-10
Arginase-1Inhibited antigen-specific T-cell proliferation[191]
Prostate cancerPleural fluid samples of malignant pleural mesothelioma patients

DU145
PC3
PGE2T-cell inhibition was mediated through the adenosine A2A receptor[192]
DU145
PC3
TGF-β1Skewed IL-2 responses in T cells and suppressed cytotoxicity[184]
MelanomaBlood samples from melanoma patients

Blood samples from melanoma tumor-bearing mice

WM1552C
WM35
WM793
WM902B
UACC-903
1205Lu
WM9
WM164
PD-L1Suppressed the function of CD8 T cells [136]
Colorectal cancerBlood samples from colorectal patients

SW403
CRC28462
1869col
FasL, TRAILInduced T-cell apoptosis [168]
DLD-1
WiDr
TGF-β1Induced differentiation of T cells to Treg-like cells via the TGF-β pathway while inactivating the SAPK signaling pathway[193]
Caco-2Galectin- 1Induced suppressor phenotype in human CD8+ T cells [185]
Head and neck
cancer
Tu167
SCC0209
HN60
Galectin- 1Induced suppressor phenotype in human CD8+ T cells [185]
Blood samples from HNSCC patientsVesicular cargoInduced apoptosis in CD8+ T cells by converting CD4+ T cells to Treg[174]
GlioblastomaBlood samples from glioma patients

UPN933
E3-2
E6-5
Vesicular cargoDeactivated T cells by FasL-dependent mechanisms and inhibit secretion of IL-2 [194]
Nasopharyngeal cancer (NPC)Blood samples from NPC patients

Blood samples from NPC tumor-bearing mice

C15
C17
Galectin- 9Induced huge apoptosis in T cells via membrane receptor Tim-3[171]
Blood samples from NPC patients

C15
C17
CCL20Facilitated Treg recruitment and expansion that increased secretion of immunosuppressive cytokines (IL10, TGFB1)[195]
Blood samples from NPC patients

TW03
C666
CNE2
miR- 24–3pBlocked T-cell proliferation and Th1 and Th17 differentiation and promoted Treg induction via dephosphorylating ERK, STAT1, and STAT3 by reducing IL-2, IFNγ, and IL-17 secretion and phosphorylating STAT5 with increasing IL-6, IL-1β, and IL-10 secretion[182,183]
Oral squamous cell carcinoma(OSCC)SCC-9
SCC-4
CAL-27
HSP70Altered development and cytotoxicity of T cells in an HSP70-dependent way via miR-21/PTEN/PD-L1 regulatory pathway[170]
Blood samples from OSCC patients

PCI-13
FasLInduced apoptotic pathways in T cells through triggering caspase-3 cleavage, the release of cytochrome c that led to disrupting mitochondrial membrane, and decreased TCR-ζ chain production[172]
Breast cancerMCF7 CD73, CD39Inhibited T cells via the adenosine A2A receptor[192]
BT-474
MDA-MB-231
TGF-β1
Suppressed T-cell proliferation[196]
Lung cancerBlood samples from lung cancer patients

A549
PC9
95D
circRNA-
002178
Enhanced PDL1 expression led to induced T-cell exhaustion[132]
Hepatocellular
Carcinoma (HCC)
Blood samples from HCC patients

MHCC97H
14- 3- 3ζInhibited the functions of T cells against cancer in the HCC microenvironment[188]
Hepa1-6
H22
SALL4/miR-146a- 5pT cells were exhausted by reducing IFN-γ and TNF-α expression while increasing the expression of inhibitory receptors such as PD-1 and CTLA-4[34]
Pancreatic cancerBxPC-3
tdTomato-BxPC-3
Vesicular cargoInduced ER stress-mediated apoptosis via activating the p38 MAP kinase signaling[181]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Najaflou, M.; Shahgolzari, M.; Khosroushahi, A.Y.; Fiering, S. Tumor-Derived Extracellular Vesicles in Cancer Immunoediting and Their Potential as Oncoimmunotherapeutics. Cancers 2023, 15, 82. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15010082

AMA Style

Najaflou M, Shahgolzari M, Khosroushahi AY, Fiering S. Tumor-Derived Extracellular Vesicles in Cancer Immunoediting and Their Potential as Oncoimmunotherapeutics. Cancers. 2023; 15(1):82. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15010082

Chicago/Turabian Style

Najaflou, Meysam, Mehdi Shahgolzari, Ahmad Yari Khosroushahi, and Steven Fiering. 2023. "Tumor-Derived Extracellular Vesicles in Cancer Immunoediting and Their Potential as Oncoimmunotherapeutics" Cancers 15, no. 1: 82. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15010082

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