Next Article in Journal
Conventional Cancer Therapies Can Accelerate Malignant Potential of Cancer Cells by Activating Cancer-Associated Fibroblasts in Esophageal Cancer Models
Previous Article in Journal
Exploring the Expression of the «Dark Matter» of the Genome in Mesothelioma for Potentially Predictive Biomarkers for Prognosis and Immunotherapy
Previous Article in Special Issue
MMP-9 as Prognostic Marker for Brain Tumours: A Comparative Study on Serum-Derived Small Extracellular Vesicles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Vesicles and Epidermal Growth Factor Receptor Activation: Interplay of Drivers in Cancer Progression

1
Department of Medical and Surgical Sciences (DIMEC), University of Bologna, 40138 Bologna, Italy
2
IRCCS Azienda Ospedaliero-Universitaria di Bologna, 40138 Bologna, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2023, 15(11), 2970; https://doi.org/10.3390/cancers15112970
Submission received: 24 March 2023 / Revised: 12 May 2023 / Accepted: 25 May 2023 / Published: 30 May 2023

Abstract

:

Simple Summary

Extracellular vesicles (EVs), consisting of microvesicles and exosomes, serve as messengers for intercellular communication by transporting proteins and nucleic acids. In solid cancers of epithelial origin, Epidermal Growth Factor Receptor (EGFR) plays a pivotal role as a driver. In vitro studies conducted on EGFR-dependent solid tumours revealed the significant correlation between EGFR and EVs production, leading to the dissemination of EGFR itself and related molecules along with inducing cell proliferation, modifying the tumour microenvironment, facilitating metastases, and conferring resistance to treatments. Recently, liquid biopsy approaches started to exploit the interplay of EGFR and EVs in delivering proteins, RNAs, and DNAs via blood/plasma of EGFR-dependent tumour patients to evaluate their possible roles and applications as candidate biomarkers in diagnosing and monitoring tumour progression and therapy efficacy.

Abstract

Extracellular vesicles (EVs) are of great interest to study the cellular mechanisms of cancer development and to diagnose and monitor cancer progression. EVs are a highly heterogeneous population of cell derived particles, which include microvesicles (MVs) and exosomes (EXOs). EVs deliver intercellular messages transferring proteins, lipids, nucleic acids, and metabolites with implications for tumour progression, invasiveness, and metastasis. Epidermal Growth Factor Receptor (EGFR) is a major driver of cancer. Tumour cells with activated EGFR could produce EVs disseminating EGFR itself or its ligands. This review provides an overview of EVs (mainly EXOs and MVs) and their cargo, with a subsequent focus on their production and effects related to EGFR activation. In particular, in vitro studies performed in EGFR-dependent solid tumours and/or cell cultures will be explored, thus shedding light on the interplay between EGFR and EVs production in promoting cancer progression, metastases, and resistance to therapies. Finally, an overview of liquid biopsy approaches involving EGFR and EVs in the blood/plasma of EGFR-dependent tumour patients will also be discussed to evaluate their possible application as candidate biomarkers.

1. Introduction

Extracellular vesicles (EVs) are a highly heterogeneous population of cellular-derived phospholipid bilayer membrane entities. These particles can be isolated from various human biofluids [1,2,3] and deliver complex messages by transferring cytosolic proteins, lipids, RNA, DNA, and metabolites inside and on their surface [4,5,6,7,8].
Since the first hypothesis in 1946 by Chargaff and West [9], EVs were described decades later from platelets in 1967 [10] and then observed as matrix vesicles [11]. Other independent studies reported the release of membrane vesicles from rectal adenoma microvillus cells [12], virus-like particles in human cell cultures and bovine serum [13,14] and seminal plasma [15]. In the early 1980s, tumour-originating fragments were also described, reporting the fusion of multivesicular bodies (MVBs) with the plasma membrane, leading to the release of vesicles, later named exosomes (EXOs) [16,17,18]. Finally, in 1996, Raposo et al. documented that EVs could stimulate adaptive immune responses [19].
A milestone in EVs research was the discovery of cargo RNA encapsulated within EVs, suggesting a function as a means for intercellular transfer of genetically encoded messages [20,21,22] and, more recently, cargo DNA and the EV-DNA-protein complex [23,24,25]. EVs were then isolated from most cells and body fluids, and the research field is constantly growing.
The EV population comprises different subtypes and is currently categorized based on their biogenesis and size into two main groups: large/medium EVs and small EVs (sEVs). In general, large/medium EVs include apoptotic bodies and microvesicles (MVs), while sEVs include EXOs [3,4,6,26]. The list of EV subtypes is constantly growing, including exospheres, migrasomes, small EV clusters, and the recently discovered exomeres, supermeres, and chromatimeres [27,28,29,30]. In addition to these classes, some cancer-specific subtypes of EVs have been identified: oncosomes (~100–400 nm) produced by non-transformed cells, whose contents can determine oncogenic effects, and large oncosomes (~1–10 µm) derived from malignant cells [5].
In oncology, EVs were reported as essential for tumour initiation, progression, and metastasis. Malignant cells were shown to transfer bioactive molecules to the neighbouring stroma, interfering with signalling and regulation of gene expression in the recipient cell and creating a favourable tumour microenvironment (TME) for cancer progression and metastatic spread [31,32,33,34]. EVs also mediate epithelial-to-mesenchymal transition (EMT), invasion, angiogenesis, immune modulation, and migration, and communicate over long distances, influencing normal distant cells and establishing a pre-metastatic niche [7,35,36]. EVs’ roles in metastatic growth and the removal of chemotherapeutics from cancer cells have also been extensively studied [32,37,38].
In this review, we will primarily focus on EXOs and MVs. First, we will provide an overview of EVs, their biogenesis mechanisms, and the differences in light of EGFR. Next, we will examine and analyse the interplay between EVs and EGFR. We will explore two main perspectives: namely, the molecular mechanisms of EV release consequent to EGFR activation, and, conversely, EGFR activation in relation to EV release and capture. Finally, the use of EVs as diagnostic tools and putative biomarkers in EGFR-dependent tumours will be described.

2. EGFR and EV Biogenesis

EGFR (or ERBB1/HER1) is a prototype transmembrane tyrosine kinase receptor activated by several growth factors. Upon ligand binding, conformational changes induce the monomeric “activator kinase” to form homo- or heterodimers with other family members (HER2, HER3, and HER4) [39]. Subsequently, the “receiver kinase” trans-phosphorylates specific tyrosine residues on the activator C-lobe, which, in turn, stimulates a cascade of effector proteins. Some of them culminate with the activation of several intracellular pathways like the Ras/Raf/Mitogen-activated protein kinases (MAPK) pathway, the phosphatidylinositol 3-kinase (PI3K)/AKT8 virus oncogene cellular homolog (AKT)/Mammalian target of rapamycin (mTOR) pathway, and the signal transducer and activator of transcription (STAT) [40]. All of them influence specific and important cellular functions: proliferation, survival, angiogenesis, inflammation, and metabolism [41]. On the contrary, for fine regulation of this receptor, other effectors trigger the negative route through CBL and E3 ubiquitin ligase for EGFR sorting and degradation in the lysosome. Moreover, recent findings have attributed structural function not only to EGFR monomers and dimers, but also to other alternative oligomers both in the basal and ligand-bound states, suggesting the complexity of EGFR signalling [42,43]. Ubiquitously expressed in different tissue types, EGFR is also one of the receptors found to be altered in several cancers, particularly breast, non-small cell lung cancer (NSCLC), glioblastoma (GBM), head and neck squamous cell carcinoma (HNSCC), ovarian, and melanoma [44,45,46,47,48,49]. EGFR signalling can also induce the loading of phosphorylated EGFR as cargo in sEVs, as well as the release of EVs [50,51].
The MVs group comprises vesicles of different sizes (~150 nm–1 µm) that directly arise from the outward budding of the plasma membrane and are involved in intercellular communication, signalling pathway activation, and cell invasion by cell-independent matrix proteolysis [4,26]. Tumour-derived microvesicles (TMVs) and oncosomes originating from cancer cells are also classified as MVs and are able to transfer bioactive molecules (namely nucleic acids, lipids, and proteins) to recipient cells, promoting cancer progression, drug resistance, and providing diagnostic markers [33,52]. MVs, being derived from the plasma membrane, carry a portion of their parent cell’s membrane content when released into the extracellular space. This allows them to reflect the cellular composition and characteristics of their origin. A combination of factors induces the formation of MVs, such as the redistribution of phospholipids and the contraction of the actin-myosin machinery [53]. ADP-ribosylation factor 6 (ARF6) leads to the activation of phospholipase D, which recruits the extracellular signal-regulated kinase (ERK) to the plasma membrane. The subsequent phosphorylation and activation of the myosin light chain kinase finally trigger the release of MVs [32,33,35,36,54]. Within this mechanism, EGFR is known to interact with ARF6 and phospholipase D2 through MAPK, also inducing MV shedding [55,56]. As a consequence, ARF6 was shown to be involved in EGFR-dependent tumours inducing invasion and EMT and linked to prognosis, further suggesting that a deregulation of MV biogenesis may contribute to cancer progression [57,58,59].
EXOs (~40–150 nm) are intraluminal vesicles (ILVs) contained in MVBs, which are released into the extracellular environment upon fusion with the plasma membrane [5,60,61]. Unlike MVs and despite sharing common intracellular mechanisms and sorting machinery, EXOs are formed through a distinct and complex intracellular regulatory process that determines their composition and functions [8,61].
The best-described mechanism of EXO formation is driven by the endosomal sorting complex required for transport (ESCRT), which is composed of four complexes (ESCRT-0, -I, -II, and -III) and approximately thirty associated proteins. The membrane is reorganized, becoming highly enriched in tetraspanins, mainly CD9 and CD63, while ESCRT 0 recognizes ubiquitinated proteins on the outside [61,62]. The presence of different stimuli, such as phosphatidylinositol 3-phosphate (PIP3), hepatocyte growth factor-regulated tyrosine kinase substrate, ubiquitination of endocytosed proteins, and/or the curved membrane topology, induce ESCRT I and II to initiate and drive the intraluminal membrane budding [63,64]. ESCRT III completes the process of ILV formation through programmed cell death 6-interacting protein (ALIX) and tumour susceptibility gene 101 (TSG101), which by interacting with proteins and cellular components, create a physical force that causes the vesicle scission [8,26,52,65]. Then, EXOs are released by the fusion of the MVB with the plasma membrane. RAB GTPases, including RAB11 and RAB35, or RAB27A and RAB27B, as well as SNARE proteins, e.g., vesicle-associated membrane protein 7 (VAMP7), have been implicated in the membrane fusion [26,27]. Figure 1 summarizes the biogenesis of MVs and EXOs.
In the case of EGFR, ubiquitination and phosphorylation are key factors for the sorting of the endocytosed receptor, driving it to lysosomal degradation or recycling back to the cell surface [66]. Low levels of phosphorylation drive EGFR to clathrin-mediated endocytosis and recycling, while at high levels, EGFR is ubiquitinated and recognized by the ESCRT complex, which drives it into ILVs and possibly to lysosomes for degradation. However, EGFR mutations can lead to increased auto-phosphorylation and MAPK signalling in some cancers (e.g., NSCLC), displaying abnormal ubiquitination and an increased rate of internalization and recycling [67,68].
EXOs may also form without the ESCRT complex machinery pathway, mainly involving ceramide subdomain-mediated curvature and tetraspanin family members (CD9, CD63, and CD81). Syntenin 1 has also been shown to be involved in the packaging of cargo into ILVs, as has ARF6, which can aid ILV budding, sorting of molecules, and driving MVBs and ILVs to lysosomes for degradation [35,69]. The presence of specific surface proteins in MVBs and ILVs, including EGFR, along with GTPase RAS-related protein RAB7A, HSP70−HSP90 organizing protein complexes, and members of the SNARE complex, including VAMP7 and syntaxin 7 and 8, results in the degradation of ILV contents, including EGFR itself [36,54]. On the other hand, RABs, actin, and SNARE proteins also mediate the fusion of MVBs with the cell membrane, and the subsequent release of EXOs [31,36,54,70]. Interestingly, the role of ARF6 was shown to be specifically involved in sorting and trafficking of EGFR toward degradation [71], while in aggressiveness in brain tumour-derived sEVs, Annexin1 is involved in inward budding EGFR-associated MVB [72].

3. EVs and EGFR Content as a Biomarker

The contents of vesicles, including MVs and EXOs, can vary depending on several factors, such as the cell type, biogenesis pathway, and physiological or pathological conditions. In general, all EVs are loaded with proteins, lipids, and nucleic acids. The different types of cargo can also be specific per vesicle and cell type, and extensive research has been carried out to characterize the content of EVs. This resulted in the assembly of several datasets from EVs studies, and nowadays, two databases are publicly available: Vesiclepedia (www.microvesicles.org (accessed on 22 March 2023)) [73] and ExoCarta (www.exocarta.org (accessed on 22 March 2023)) [3,65,74].
The methods used for isolation and characterization of EVs may yield different outcomes [6,54,75,76,77]. Ultracentrifugation, polymeric precipitation, filtration, immunoaffinity isolation, and microfluidics techniques are the most commonly used methods that are not necessarily mutually exclusive, and combinations are recommended [76,78,79]. For example, sedimentation by ultracentrifugation procedures, depending on the density or ‘‘cargo’’, may lead to the presence of some small EVs in the large vesicle pellet, while some larger particles may remain in the upper part of the tube [80]. Differently, filter-based methods with low-molecular-weight filters (0.2 µm), employed to enrich for the smaller EV fraction, may lead to deformation or breakup of larger vesicles or platelets, thereby potentially skewing results [78].
This situation is further complicated by EV heterogeneity. The current purification methods may be unable to identify EVs based on their subtype or biogenesis, due to overlapping sizes and a lack of well-established genetic markers [4,26]. Furthermore, different subpopulations of MVBs coexist simultaneously in cells. For example, in HeLa cells two distinct EV populations have been reported after stimulation with EGF [81]: a population of CD63-positive endosomes containing EGFR and another subset of MVBs still positive for CD63 but negative for EGFR cargo.
Despite the difficulties in standardization, mandatory for a possible clinical application, EVs are considered a good source for tumour molecular profiling compared with cell-free nucleic acids because EVs protect nucleic acids from degradation [5,82]. Moreover, tumour cells release EVs with an altered cargo profile, influencing intercellular communication during metastatic spreading [31,83]. This is also reflected by qualitative and quantitative changes in sEV populations, which were identified in the blood of cancer patients, corroborating a function of EVs as diagnostic and prognostic markers [52,78,84,85,86,87,88]. In line with this, increased MV levels have been detected in the plasma of patients suffering from gastric, lung, breast, pancreatic, colorectal, and prostate cancers, as well as in hematologic malignancies [33,89,90,91,92,93,94].
Hereafter, the review will focus on EXO cargo, namely, proteins, RNAs, and DNA found within EXOs and MVs, with a focus on EGFR-dependent tumours. EV-EGFR as a biomarker in body fluids will also be discussed.

3.1. Proteins

Proteomics approaches in primary cell cultures or biofluids yielded catalogues of proteins in different types of EVs. Notably, the proteomic profiles obtained were significantly influenced by the isolation methods used for EVs, resulting in variable homogeneity of EVs and subfractions [79]. Furthermore, the same cell type may secrete different subgroups of vesicles depending on environmental factors (e.g., oxygen tension) or an activating stimulus [3]. Despite the partial overlapping results, some proteins can be considered pan-EV markers (i.e., common for most EVs), while other proteins and post-translational modifications are specific for vesicle localization, cellular origin, and mechanism of secretion. As expected, either cytoskeletal proteins and plasma membrane proteins or proteins involved in vesicle trafficking are highly abundant in EVs and are currently used as markers of EV subpopulations (Table 1).
In addition to the various proteins involved in biogenesis (e.g., ESCRT complex, ALIX, TSG101), sorting (e.g., glycan binding, tetraspanins) [3,26,65,96,97,98], and fusion with membrane (GTPase RAS-related protein, SNARE complex, ARF6, VAMP7, and syntaxin 7 and 8) [52,69] usually used to characterize EV populations, EVs contain specific stress proteins (heat shock proteins; HSP70−HSP90), antigen presentation complex (MHC I and MHC II) and transmembrane proteins (lysosomal-associated membrane protein, LAMP1; transferrin receptor, TfR), including the membrane receptor EGFR.
Li et al. [99] analysed EVs secreted by nasopharyngeal carcinoma (NPC) cells and detected an EXO subpopulation enriched for membrane CD9 and CD63 and non-membrane ALIX and TSG101 proteins. Subsequently, the authors evaluated the role of EGFR-rich EVs both in vivo (tumour tissues) and in vitro (high and low metastatic NPC cell lines). Additionally, they found that EGFR was overexpressed in metastatic NPC cell lines and patients. Furthermore, EGFR-rich EVs secreted by metastatic NPC cell lines induced cellular up-regulation of EGFR and abnormal signalling activation, which, in turn, are responsible for EMT and shorter survival in xenografted mice treated with secreted EGFR-rich EVs [99]. The importance of EXO-EGFR and related proteins was also reported in lung cancer and EMT. Jouda et al. [100], showed that EXO-EGFR from EGFR mutated NSCLC patients can induce the activation of the PI3K/AKT/ mTOR pathway in A549 cells. The authors also demonstrated an increase in vimentin, nuclear factor erythroid 2-related factor 2, and P-cadherin, suggesting that EXOs may act as mediators of EMT and tumour invasion.
The oncogenic role of EGFR in triggering the release of EXO-like extracellular vesicles and the evaluation of phosphoprotein content both in vivo and in vitro was also reported [101]. Both wild-type and mutant EGFR were detected in EVs isolated in the plasma of patients affected by GBM and in GBM xenograft mice, reflecting the mutations found in the tumour. After highlighting the ability of GBM-EVs to induce the release of EGFR and phospho-EGFR, the authors also reported similar findings in additional cancer cell lines, including A431 (epidermal), MDA-MB-231 (breast), BxPC3, PANC-1 (pancreatic), PC-3 (prostate), HCT116, DLD-1, and CaCo-2 (colon). In A431 cells, TGFα treatment boosted phosphorylation of both membrane EGFR and EV-engulfed EGFR. On the other hand, panEGFR inhibitors reduced phosphorylation only on the membrane located EGFR, but intriguingly increased the pEGFR activation recovered as cargo in the EV. Remarkably, the inhibition of both the EXO biogenesis and the apoptotic caspase cascade reduced the EVs release, suggesting a possible shared mechanism between EXO biogenesis and apoptosis. Thus, in EGFR-dependent cancer cells, blocking the oncogenic signals may lead to cellular stress that stimulates both exosomal and apoptotic pathways [101].

3.2. RNA

A milestone in EV research was the discovery of cargo RNA, suggesting its function as a means for intercellular transfer of genetically encoded messages.
The RNAs initially discovered and most studied in EVs are messenger RNA (mRNA) and microRNA (miRNA) [5,6,102]. The role of EV-RNAs is also emerging in liquid biopsy, as a promising non-invasive source for molecular testing, complementing and improving the use of circulating tumour cells and cell-free DNA (cfDNA) analysis [33,102]. In addition, enrichment analysis observed that several miRNAs were almost undetectable in colorectal cancer cell lysates but were found at high levels in the respective cell-derived MV and EXO [103]. Usually, the absence or minor amounts of ribosomal 18S and 28S in EVs are reported. However, with the rapid advances in deep sequencing or Next Generation Sequencing (NGS) techniques, fragments of rRNA, as well as long and short non-coding RNA and tRNA fragments, have been found in EVs [1,65]. Currently, the Vesiclepedia database includes over 27,000 and 10,000 entries mRNAs and miRNAs, respectively [73]. The general consensus is that EVs protect RNA from digestion in the extracellular environment. Furthermore, different ribonucleoproteins, as well as high- and low-density lipoproteins, can stably associate with RNA species [3,65,82].
In oncology, EV-RNAs are studied both as potential biomarkers in blood/plasma/serum or as emergent treatment strategies for malignant neoplasms. Table 2 reports RNAs studied in cell cultures or as biomarkers. For example, EV-miRNA Let-7b-3p, miR-150-3p, miR-145-3p, and miR-139-3p were suggested as biomarkers in the plasma of colon, breast, or lung cancer patients [104,105]. In line with this, Ogata-Kawata et al. [106] revealed that miR-23a, miR-1246, and miR-21 differ between EVs from CRC patients and controls. Furthermore, miR-486 was significantly enriched in serum EV-miRNA of CRC patients and associated with shorter survival and liver metastases [107,108,109]. Furthermore, Cha et al. [110] evaluated eight mRNA markers (MYC, VEGF, CDX2, CD133, CEA, CK19, EpCAM, and CD24) extracted from plasma EVs, showing statistically significant differences between healthy subjects and CRC patients for the combination of VEGF and CD133. Differently, Krug et al. [111] demonstrated that the measure of circulating tumour DNA (ct-DNA) and EXO-RNA was more sensitive than ct-DNA alone to detect EGFR mutations in advanced NSCLC. Nevertheless, whether enough RNA molecules are shuttled to EV to functionally influence the recipient cell remains to be proven.
In NSCLC, Xia et al. [112] suggested the possible use of EXO-miRNA as markers to identify patients with wild-type or mutated EGFR. In particular, the authors highlighted 96 EXO-miRNAs differentially expressed between the serum of healthy and NSCLC patients. Interestingly, 40 of them were also different between patients with mutated EGFR, including miR-142-5p, -592, -217, -451b, and -150, already reported as putative biomarkers. Moreover, the authors searched for EXO-miR-260 and -miR-1169 in cancer cell lines showing the same trend and expression as in the serum of patients, suggesting a correlation between downregulation of EXO-miR-260 and mutated EGFR, while upregulation of EXO-mir-1169 may identify patients with wild-type EGFR. Finally, EVs were also reported to mediate the transfer of long noncoding RNAs implicated in the pathological process of lung pulmonary fibrosis [113], which correlated with the presence of missense polymorphisms in members of the EGFR family [114].
Table 2. Some of the RNAs found in MVs and EXOs and studied in cell culture or as liquid biopsies biomarkers.
Table 2. Some of the RNAs found in MVs and EXOs and studied in cell culture or as liquid biopsies biomarkers.
Target/CargoRNA TypeEVsCancer TypeCell Line/TissueReferences
let-7-b3p, miR-150-3p, miR-145-3p, miR-139-3pmiRNAsEVColon/Breast/LungPlasma[105]
miR-486, miR-548cmiRNAEXOColorectalSerum[107,108,109]
miR-1246, miR-21, miR-23amiRNAEXOColorectalSerum/plasma [106]
miR-92amiRNAEXOColorectalSerum[115]
MYC, VEGF, CDX2, CD133, CEA, CK19, EpCAM, CD24mRNAEXO/MVColorectalPlasma[110]
miR-1246, -23a, -200c, -203a, -19a, 7641miRNAEXO/MVColorectalLM1863 [116]
BCAR4lncRNAEXOColorectalSerum[117]
KRTAP5-4, MAGEA3mRNAEXOColorectalSerum[117]
miR-142-5p, -592, -217, -451b, -150, -260, -1169miRNAEXONSCLCSerum[112]
linc00152lncRNAEXOGastric cancerPlasma[118]
let-7amiRNAEXOBreast HCC70[119]
miR-145miRNAEXOBreastT-47D[120]
miR-214miRNAEXOGastric SGC7901[121]
KRASG12D (target)siRNA EXOPancreasPANC-1, BxPC-3, Capan-1, MIA PaCa-2[122]
TPD52 (target)siRNAEXOBreastSKBR3[123]
miR-494-3pmiRNAEXONSCLCPlasma/NCI-H1975, HCC827[124]
miR-6087, miR-99b-5p, miR-7641, miR-378a-3p, miR-25-5p, miR-1293, miR-184, miR-3913-5pmiRNAEXONSCLCPlasma/H1975[125]
miR-26a/bmiRNAEXOGastric SGC7901[126]
m, messenger; mi, micro; EXO, exosome; MV, microvesicle; NSCLC, non-small cell lung cancer; sEV, small extracellular vesicle.

3.3. DNA

Despite most of the EV literature focusing on proteins and RNAs, numerous studies have recently reported the presence of DNA either associated with the surface of EVs or within their lumen [4]. Table 3 reports DNA genes studied in EVs. In 2013, one of the first studies reported the presence of genomic and mitochondrial DNA in human plasma EVs, and in 2014, mutant KRAS and TP53 DNAs were found in serum EXOs collected from cancer patients [24,127]. Moreover, recent findings have shown that horizontal transfer of mitochondrial DNA can occur through EVs. In detail, mitochondrial DNA carried within EVs has been found to promote the reactivation of dormant cancer stem-like cells, resulting in resistance to endocrine therapy [128].
Traditionally, it was believed that EV-DNA was the passive result of cell death by apoptosis or necrosis. However, recent studies have suggested an interesting hypothesis about the capability of EVs to also store and remove damaged DNA, potentially contributing to the maintenance of DNA integrity and homeostasis [129]. Additionally, Montermini et al. reported the EXO-DNA release as a response to cellular stress [101]. However, the biological role of EV-DNA requires further investigation.
Table 3. DNA found in MVs and EXOs.
Table 3. DNA found in MVs and EXOs.
DNAEVCancer/Normal Cell Line/TissueReference
BCR/ABLEVChronic Myelogenous Leukaemia/Healthy Plasma/VSMC, HEK293, K562[23]
KRAS, TP53EXOPancreatic cancerSerum/Panc-1, T3M-4 [24]
APC, KRAS, TP53, PIK3CAMV, EXOColorectal cancerTissue biopsy[130]
TGFBR2EXOColorectal cancerHCT116[131]
HRASEXO-likeintestinal epithelial cells and fibroblastsIEC-18, RAS-3, RAT-1[132]
EGFREXO-likeepidermoid cancerA431 [101]
EGFREXONSCLCPlasma and tissue biopsy [111]
EGFR T790MEXONSCLCPlasma [133]
KRASEXOPancreatic cancerPlasma [134]
KRASG12D, TP53R273HEXOPancreatic cancerPlasma [135]
BRAFV600EEVMelanomaPlasma [136]
TP53, MLH1, PTENMV, EXOProstate cancerPlasma/LNCaP, PC-3, RC92a/hTERT[137]
MYCL and S-EVProstate cancerPlasma/PC3, U87, U2OS-H2B-GFP[138]
DROSHA, LIG4, MACROD2, SATB1, RASSF6, BIRC2EXOOvarian cancerPlasma/OVCAR-8, OVCAR-5, FTE[139]
APC, Adenomatous Polyposis Coli; BCR/ABL, break-point cluster region v-abl Abelson murine leukaemia viral oncogene; BIRC2; baculoviral IAP repeat containing 2, BRAF, B-Raf proto-oncogene serine/threonine kinase; DROSHA, drosha ribonuclease III; EXO, exosome; HRAS, HRas proto-oncogene, GTPase; KRAS, Kirsten rat sarcoma virus; l-EV, large extracellular vesicle; LIG4, DNA ligase 4; MACROD2, mono-ADP ribosylhydrolase 2; MLH1, mutL homolog 1; MV, microvesicle; PIK3CA, phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha; PTEN, phosphatase and tensin homolog; RASSF6, Ras association domain family member 6; SATB1, SATB homeobox 1; s-EV, small extracellular vesicle; TGFBR2, transforming growth factor beta receptor 2; TP53, tumour protein p53; VSMC, vascular smooth muscle cells.
Reports have also shown that EV-DNA secretion varies under different conditions, leading to changes in the recipient cells. Lee et al. reported that full-length oncogenic H-RAS DNA was transferred in vitro via EVs, changing the behaviour of the recipient cells [132]. By comparing the emission of rat intestinal cells (IEC-18) and h-RAS mutated cells (RAS-3), the authors reported the increased excretion of EXO-like vesicles in non-transformed fibroblast (RAT-1), allowing the uptake of the oncogenic h-RAS with a concomitant mitogenic response. Nevertheless, whether the tumour EXO-gDNA could be horizontally transmitted with the consequent changes in the recipient cells is still controversial.
EV-DNA was found in different biological fluids and cell cultures, suggesting it is a promising candidate for liquid biopsy [29,134,140,141,142]. In cancer patients, EV-DNA is being investigated in several applications, mainly as a biomarker or to monitor disease and treatment response in liquid biopsy [139,143]. Moreover, the search by NGS for common hotspot mutations (BRAF, EGFR, and KRAS) showed a greater sensitivity if compared to circulating free DNA or circulating tumour DNA. The authors further claimed that cfDNA in plasma may be mainly present in EXOs [141].
In 2017, Allenson and colleagues found a KRAS mutation in EXOs of pancreatic ductal adenocarcinoma and associated it with disease free survival [134]. Moreover, the detection rates of mutant KRAS in EXOs were higher in EXOs as compared to cfDNA. Therefore, EV-DNA is now also considered a promising liquid biopsy approach for diagnosis, prognosis, and monitoring of treatment responses, and studies have reported its use in patients affected by pancreatic, ovarian, pheochromocytoma, and colorectal cancer [133,144,145,146,147]. In line with this, recent studies have focused on the role and usefulness of EXOs-DNA in liquid biopsy to detect mutations related to EGFR [148,149]. EGFR mutations in EXOs of malignant pleural effusion of lung adenocarcinoma patients reported 100% agreement for EGFR genotyping with the primary tumour [149]. Moreover, by analysing EV-DNA, it was possible to detect three additional T790 mutant cases in patients sensitive to EGFR-TKI and two more cases in patients with acquired resistance, suggesting that liquid biopsy using EV-DNA is promising for the detection of low-rate EGFR mutations in lung cancer. The utility of EV-DNA from liquid biopsies in lung adenocarcinoma was also confirmed by Park et al. in bronchial wash [150]. The T790M mutation rate was high, confirming that bronchial wash-derived EVs can be usefully analysed to detect DNA EGFR mutations, thus implying that EV-DNA analysis may provide valuable information for the management and treatment of lung cancer.

3.4. EV-EGFR in Body Fluids

In cell–cell communication, EVs have been shown to reflect the composition of their host cells [151,152]. Probably due to physiological alteration, it has been estimated that the blood of cancer patients contains around 4000 trillion EXOs, twice as much as compared to normal human blood, which makes them a potential non-invasive biomarker for different cancer screenings [153,154]. Many studies are focusing mostly on miRNAs and protein content analysis of circulating EVs. For example, in lung adenocarcinoma and lung squamous cell carcinoma, researchers are correlating specific EXO miRNAs with stages and histological subtypes [155]. As a matter of fact, EXOs express a high level of miRNAs, which are known to influence mRNA stability and TME [156]. The potential diagnostic power of such molecules has been shown in the literature: miR-139-5p, miR200b-5p, miR-378a, miR-126, and miR-379 allow for discrimination between healthy and lung cancer smokers, while other specific EXO miRNAs are useful to distinguish between lung adenocarcinomas and granulomas [157]. Recently, much attention has been concentrated on the presence of EXO-EGFR at DNA, RNA, and protein levels as diagnostic, prognostic, and therapeutic markers. Indeed, early detection of the EGFR mutation profile would lead to a prompt treatment schedule. An exhaustive table on the evaluation of EV-EGFR in liquid biopsy in different tumour types is provided by Frawley at al. [7]. Using a microfluidic electrochemical immunosensor system, Ortega et al. indirectly detected the EXO EGFR level for the diagnoses of 20 healthy patients and 30 breast cancer patients [158]. With a different flow cytometry-based approach, Wang et al. compared the EV EGFR content in several sample sera to identify glioma patients in correlation with tumour grade [159]. For diagnostic purposes, the first FDA approved liquid biopsy test was the Cobas EGFR Mutation Test v2® (Roche Diagnostics Inc., Basilea, Switzerland), which analyses the EGFR gene mutations in cfDNA [160]. Based on the study performed by Fernando et al., 93% of cfDNA derives from EXOs [141]. However, such a test showed only 58% sensitivity and 80% specificity. Subsequently, further studies suggested that combining the mutational profiles of exosomal RNA/DNA and cfDNA would improve both sensitivity and specificity [133].

4. Role of EGFR on Cargo Trafficking and Modulation of the Tumour Microenvironment

EVs are generated by the endolysosomal pathway and released by several cell types, including tumour cells. EVs resemble the phenotypic features of the cell of origin and carry different bioactive molecules, cytokines, and receptors, including the ERBB receptor family [161,162]. Zanetti-Domingues et al. reported a central role for EGFR receptor signalling in EV trafficking and its loading as cargo [35,69]. Several studies demonstrated a high abundance of circulating EXOs in the plasma of cancer patients compared to healthy ones [154,163]. While in the latter, EXO contributes to maintaining stable homeostasis, in the former, EXO release is associated with tumour progression, metastasis, and drug resistance, and is able to deliver active cargo to recipient cells in an autocrine, paracrine, or endocrine manner [31,70,151,164]. Indeed, EGFR members and their ligands are expressed in sEVs released by cancer cells [7,165]. Through a process known as “horizontal oncogene transmission”, Al-Nedawi et al. demonstrated, in 2008, the presence of a constitutively activated mutant isoform of EGFRvIII in U373 GBM cell-derived EVs, which appeared to be functional once internalized by the recipient cells, increasing MAPK, AKT, PDK1, and RAF signalling as well as VEGF expression [34]. In addition, subsequent studies showed a different proteome composition in the sEV driven by the mutant EGFR form compared with the wild-type receptor, influencing a pro-tumoral microenvironment [166].
An altered sEV proteome was also found in NSCLC, contributing to resistance to anti-EGFR therapies [167]. In vitro culture of PC9 cells, harbouring the EGFR-T790M mutation and insensitive to gefitinib treatment, released sEVs containing specific molecules of the AKT/mTOR pathway, influencing the drug sensitivity of the recipient cells but also enhancing proliferation and invasion. Moreover, sEVs secreted by gefitinib-treated PC9, can induce autophagy in autologous cells, decreasing cisplatin activity and allowing recipient cells to survive through an antagonistic interaction [168].
EXOs derived from patients with EGFR-positive lung adenocarcinoma caused, in recipient cells, vimentin overexpression, extracellular matrix degradation, and activation of PI3K/AKT/mTOR, a well-known pathway involved in the EMT process [100]. The link between EGFR-mediated EVs biogenesis and EMT was also demonstrated by Fujiwara et al. in HSC-3 oral squamous cell carcinoma; cells under EGF stimulation induced both EMT and increased EVs secretion, which, in turn, led to the expression of EMT markers in recipient cells [50,169]. The same phenomenon was seen in HNSCC under TGF-α stimulation [170]. Stromal cells and TME are also another target affected by EV release in order to sustain neoplastic features, especially where inflammation is one of the major malignant risk factors, like in colorectal cancer. EGFR was shown to be overexpressed in inflammatory bowel disease (IBD)-associated colorectal adenocarcinoma, and indeed, an increased level of EGFR was found in IBD mice-derived colorectal EVs which were also internalized by NIH3T3 fibroblast cells [171,172]. NIH3T3, in turn, demonstrated an increased protein level of EGFR after EVs incorporation associated with high ERK phosphorylation, increased proliferation, and KI67 expression [125,171]. Thus, EVs are the principal mediator of tumour–TME communication, interacting not only with fibroblasts, but also with other important recipient cells, such as endothelial or immune cells, shaping them in favour of tumour growth. The final effect is translated into pro-angiogenic and immune suppression signalling [35]. In particular, EVs influence T regulatory cells (Tregs) and myeloid-derived suppressor cells (MDSCs). For example, it has been demonstrated in NSCLC that EGFR+ sEVs induce tolerogenic dendritic cells, which, in turn, suppress T cell activation [173].
Thus, cancer cell-derived EVs, depending on their specific patterns or cargo, can reprogram long-distance and organ-specific cells to promote growth and invasion, form a pre-metastatic niche, modulate TME, and confer drug resistance (Figure 1).

5. EV-EGFR in Resistance to Therapy and as Vehicles of Treatment

5.1. Resistance to Therapy

Recently, EV-EGFR and related proteins have also been described as involved in resistance to therapies. Choi et al. [167] demonstrated significant differences in the proteome profiles of EVs released by sensitive NSCLC (PC9) compared to gefitinib-resistant cells (PC9R) due to an acquired EGFR mutation (T790M). In particular, the AKT/mTOR signalling pathway was hypothesised as a key component actively involved in gefitinib resistance. The importance of EVs in therapy resistance was also reported in breast cancer. HER2-positive EVs were shown to bind to monoclonal antibody-based drugs (trastuzumab), thereby reducing therapy bioavailability to their target tumour cells [94,174,175]. A different mechanism was shown by Martinez et al. [176], highlighting the increased levels of the immunosuppressive proteins transforming growth factor beta 1 and programmed death-ligand 1 in EVs in HER2-overexpressing drug-resistant cells and the serum of patients who do not respond to HER2-targeted drug treatments.
EXO-miRNAs were also associated with resistance to treatments against EGFR. A potential role of miR-494-3p as a marker of osimertinib resistance in NSCLC was reported by Kaźmierczak et al. [124] by analysing EXO-RNA from plasma sampled at baseline and after disease progression in patients with mutated EGFR T790M NSCLC. The expression of EXO miR-494-3p was significantly increased in progression samples compared with baseline samples. Furthermore, the authors confirmed that miR-494-3p was significantly upregulated in two EGFR-mutant NSCLC cell lines refractory to osimertinib, demonstrating a partial role of miR-494-3p in conferring resistance to treatment in vitro. Li et al. [125], analysing sensitive and resistant NSCLC cell lines, demonstrated that EXO-derived miRNAs may be involved in the mechanisms of resistance. Levels of EXO-miR-6087, miR-99b-5p, and miR-7641 were upregulated, and miR-378a-3p, miR-25-5p, and miR-1293 were downregulated in sensitive cells and linked to the RAS-MAPK signalling pathway, a member of the bypass pathways involved in drug resistance in EGFR-TKI therapy for NSCLC. Furthermore, analysing blood samples from 64 NSCLC patients with EGFR mutations (on exons 19 and 21), the authors reported that the upregulation of EXO-miR-184 and -3913-5p may indicate the development of drug resistance in patients.
Despite EGFR being early linked to oncogenesis in GBM, therapies exploiting it as a target are still not as efficient as in other types of cancer (e.g., NSCLC) [177]. Chi et al. evaluated the RNA content of tumour-derived EVs in patients unresponsive to dacomitinib [178]. The authors found 32 mRNAs differentially expressed between sensitive and non-sensitive patients, including LAMTOR2, an activator of MAPK and mTOR signalling, and CSF1, which encodes macrophage colony-stimulating factor. Tamizolide resistant GBM cells were shown to induce resistance in sensitive cells by transferring miR-1238 through EXOs [179]. miR-1238 reduced the activity of caveolin-1, a tumour suppressor gene, also through its interaction with EGFR. Regarding the role of EXOs in promoting metastasis, Zhang et al. [126] reported that EGFR delivered via EXOs may regulate TME, promoting gastric cancer liver metastasis. EGFR was present in EXOs of gastric adenocarcinoma cells and enriched in EXOs derived from gastric cancer patients serum, but not in EXOs of healthy human serum. The levels of EGFR in both cancer cells and EXOs were strongly decreased by transfection of siRNA targeting EGFR, confirming the importance of EGFR in facilitating protein transfer of EGFR itself. The authors then demonstrated that gastric cell-EXOs significantly promote HGF and EGFR expression in mixed liver cells while suppressing miR-26a/b expression, suggesting the regulation of the microenvironment of the liver to prepare favourable conditions for future metastasis.
Furthermore, EV-DNA was reported to be involved in resistance to therapy. Crow et al. isolated EXO-gDNA from ovarian cancer cells resistant to carboplatin treatment [180]. The authors further demonstrated that tumour-derived EXOs induced an EMT phenotype in sensitive cells, leading to the development of subpopulations of platinum-refractory cells.

5.2. Vehicles for Treatment

The use of EVs as cargo vehicles, mainly for miRNA or siRNA, has been explored as a potential treatment option for cancer. This approach involves using EVs to deliver therapeutic RNA molecules to cancer cells [181]. For example, EXOs containing the tumour suppressor miRNA let-7a were able to successfully inhibit EGFR-expressing cancer cells in vitro and decrease tumour growth in xenograft breast cancer mice in vivo [119]. Interestingly, the authors modified EXOs with GE11 peptide, which binds membrane EGFR more efficiently than EGF, but with lower mitogenic capability. They proved in vitro and in vivo that the amount of surface EGFR dictated drug delivery for GE11-EXOs in different breast cancer cell lines. Indeed, the delivery of let-7 miRNA to cancer cells was able to impair cell cycle progression and reduced cell division. Similar inhibitory effects were reported for EXO-mir targeting HER2 [120]. In this case, the authors used mesenchymal stem cell-derived EXOs as carriers of the oncosuppressor miR-145 in T-47D cells, confirming the tumour-suppressive effects of miR-145. In particular, miR-MSC-EXO demonstrated its role in both inducing apoptosis and inhibiting metastasis by downregulating MMP9 and HER2 and upregulating TP53. A different miRNA was tested by Wang et al. with EXOs transporting anti-miR-214. In this case, a link with chemoresistance was evidenced in gastric cancer cells in vitro, showing that EXOs, as vehicles of anti-miR-214, enhanced cell death and inhibited migration in cisplatin-resistant cells [121]. In 2017, EXOs transporting siRNA targeting the oncogenic mutated GTPase KRASG12D were used to treat pancreatic cancer, yielding suppression of tumour growth and extended survival in mice [122]. In 2019, Limoni et al. obtained promising results in vitro with EXOs containing TPD52 siRNA in HER2-expressing SKBR3 breast cancer cells [123].

6. Conclusions

Although there are several challenges associated with analysing EVs, such as the numerous sources of secretion into body fluids, the lack of standardized isolation and purification protocols, and the absence of a universally agreed-upon marker signature to identify the origin of EVs, their potential use is still significant and promising. An update in the field of Minimal Information for Studies of Extracellular Vesicles has been provided recently by the International Society for Extracellular Vesicles (ISEV). However, more extensive studies are needed to establish a low cost and accurate method for future clinical applications [182,183,184].
EGFR-dependent tumours through the EV cargo of specific proteins, mRNAs, or miRNAs can modulate long-distance and organ-specific cells by promoting growth and invasion, regulating TME, and conferring drug resistance.
Notably, recent studies demonstrated the transfer of EGFR and associated molecules from different cancer cells, activating pro-tumorigenic pathways in recipient cells. The process of transferring biological signals between cells by EV is similar to the hormonal communication and opens new perspectives in the evaluation of the biological effects of such transfer. Whether and how tumour cells sort and transfer cargo in EVs is still an unanswered question. On the other hand, EVs present in liquid biopsies, such as serum, plasma, CSF, and pleural effusions, offer a crucial and untapped opportunity for the development of non-invasive biomarkers that can revolutionize the diagnosis, prognosis, and therapy monitoring not just for EGFR-dependent tumours, but also for a wider range of cancers and diseases.

Author Contributions

Writing—original draft preparation, E.F. and D.R.; tables, figures, and editing, F.B., V.G., F.P., C.G., R.S.K., A.M., D.M.F., M.M. and M.S.; writing—review, G.D. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MIUR, PRIN2017, grant number 2017TATYMP_002, and Fondazione Carisbo “Ricerca medica e alta tecnologia 2022” and Fondazione del Monte 2021. GD was supported by AIRC (grant number MFAG 24684).

Acknowledgments

We are grateful to Rossella Solmi for her continuous support and assistance in writing.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Crescitelli, R.; Lässer, C.; Lötvall, J. Isolation and characterization of extracellular vesicle subpopulations from tissues. Nat. Protoc. 2021, 16, 1548–1580. [Google Scholar] [CrossRef] [PubMed]
  2. Raposo, G.; Stoorvogel, W. Extracellular vesicles: Exosomes, microvesicles, and friends. J. Cell Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef]
  3. Yáñez-Mó, M.; Siljander, P.R.M.; Andreu, Z.; Bedina Zavec, A.; Borràs, F.E.; Buzas, E.I.; Buzas, K.; Casal, E.; Cappello, F.; Carvalho, J.; et al. Biological properties of extracellular vesicles and their physiological functions. J. Extracell. Vesicles 2015, 4, 27066. [Google Scholar] [CrossRef] [PubMed]
  4. Tsering, T.; Li, M.; Chen, Y.; Nadeau, A.; Laskaris, A.; Abdouh, M.; Bustamante, P.; Burnier, J. V EV-ADD, a Database for EV-associated DNA in human liquid biopsy samples. J. Extracell. Vesicles 2022, 11, e12270. [Google Scholar] [CrossRef] [PubMed]
  5. Desmond, B.J.; Dennett, E.R.; Danielson, K.M. Circulating extracellular vesicle microRNA as diagnostic biomarkers in early colorectal cancer—A review. Cancers 2019, 12, 52. [Google Scholar] [CrossRef]
  6. Hill, A.F.; Pegtel, D.M.; Lambertz, U.; Leonardi, T.; O’Driscoll, L.; Pluchino, S.; Ter-Ovanesyan, D.; Nolte-‘t Hoen, E.N.M. ISEV position paper: Extracellular vesicle RNA analysis and bioinformatics. J. Extracell. Vesicles 2013, 2, 22859. [Google Scholar] [CrossRef]
  7. Frawley, T.; Piskareva, O. Extracellular vesicle dissemination of epidermal growth factor receptor and ligands and its role in cancer progression. Cancers 2020, 12, 3200. [Google Scholar] [CrossRef]
  8. Colombo, M.; Raposo, G.; Théry, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell Dev. Biol. 2014, 30, 255–289. [Google Scholar] [CrossRef]
  9. Chargaff, E.; West, R. The biological significance of the thromboplastic protein of blood. J. Biol. Chem. 1946, 166, 189–197. [Google Scholar] [CrossRef]
  10. Wolf, P. The nature and significance of platelet products in human plasma. Br. J. Haematol. 1967, 13, 269–288. [Google Scholar] [CrossRef]
  11. Anderson, H.C. Vesicles associated with calcification in the matrix of epiphyseal cartilage. J. Cell Biol. 1969, 41, 59–72. [Google Scholar] [CrossRef] [PubMed]
  12. De Broe, M.; Wieme, R.; Roels, F. Membrane fragments with koinozymic properties released from villous adenoma of the rectum. Lancet 1975, 306, 1214–1215. [Google Scholar] [CrossRef] [PubMed]
  13. Benz, E.W.; Moses, H.L. Brief communication: Small, virus-like particles detected in bovine sera by electron microscopy2. J. Natl. Cancer Inst. 1974, 52, 1931–1934. [Google Scholar] [CrossRef]
  14. Dalton, A.J. Microvesicles and vesicles of multivesicular bodies versus “virus-like” particles. J. Natl. Cancer Inst. 1975, 54, 1137–1148. [Google Scholar] [CrossRef]
  15. Ronquist, G.; Brody, I.; Gottfries, A.; Stegmayr, B. An Mg2+ and Ca2+-stimulated adenosine triphosphatase in human prostatic fluid: Part I. Andrologia 2009, 10, 261–272. [Google Scholar] [CrossRef] [PubMed]
  16. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle formation during reticulocyte maturation: Association of plasma membrane activities with released vesicles (exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  17. Pan, B.T.; Johnstone, R.M. Fate of the transferrin receptor during maturation of sheep reticulocytes in vitro: Selective externalization of the receptor. Cell 1983, 33, 967–978. [Google Scholar] [CrossRef]
  18. Dvorak, H.F.; Quay, S.C.; Orenstein, N.S.; Dvorak, A.M.; Hahn, P.; Bitzer, A.M.; Carvalho, A.C. Tumor shedding and coagulation. Science 1981, 212, 923–924. [Google Scholar] [CrossRef]
  19. Raposo, G.; Nijman, H.W.; Stoorvogel, W.; Liejendekker, R.; Harding, C.V.; Melief, C.J.; Geuze, H.J. B Lymphocytes secrete antigen-presenting vesicles. J. Exp. Med. 1996, 183, 1161–1172. [Google Scholar] [CrossRef]
  20. Ratajczak, J.; Miekus, K.; Kucia, M.; Zhang, J.; Reca, R.; Dvorak, P.; Ratajczak, M.Z. Embryonic stem cell-derived microvesicles reprogram hematopoietic progenitors: Evidence for horizontal transfer of mRNA and protein delivery. Leukemia 2006, 20, 847–856. [Google Scholar] [CrossRef]
  21. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed]
  22. Skog, J.; Würdinger, T.; van Rijn, S.; Meijer, D.H.; Gainche, L.; Curry, W.T.; Carter, B.S.; Krichevsky, A.M.; Breakefield, X.O. Glioblastoma microvesicles transport RNA and proteins that promote tumour growth and provide diagnostic biomarkers. Nat. Cell Biol. 2008, 10, 1470–1476. [Google Scholar] [CrossRef] [PubMed]
  23. Cai, J.; Han, Y.; Ren, H.; Chen, C.; He, D.; Zhou, L.; Eisner, G.M.; Asico, L.D.; Jose, P.A.; Zeng, C. Extracellular vesicle-mediated transfer of donor genomic DNA to recipient cells is a novel mechanism for genetic influence between cells. J. Mol. Cell Biol. 2013, 5, 227–238. [Google Scholar] [CrossRef] [PubMed]
  24. Kahlert, C.; Melo, S.A.; Protopopov, A.; Tang, J.; Seth, S.; Koch, M.; Zhang, J.; Weitz, J.; Chin, L.; Futreal, A.; et al. Identification of double-stranded genomic DNA spanning all chromosomes with mutated KRAS and P53 DNA in the serum exosomes of patients with pancreatic cancer. J. Biol. Chem. 2014, 289, 3869–3875. [Google Scholar] [CrossRef]
  25. Ghanam, J.; Chetty, V.K.; Anchan, S.; Reetz, L.; Yang, Q.; Rideau, E.; Liu, X.; Lieberwirth, I.; Wrobeln, A.; Hoyer, P.; et al. Extracellular vesicles transfer chromatin-like structures that induce non-mutational dysfunction of P53 in bone marrow stem cells. Cell Discov. 2023, 9, 12. [Google Scholar] [CrossRef]
  26. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  27. Zhang, Q.; Jeppesen, D.K.; Higginbotham, J.N.; Graves-Deal, R.; Trinh, V.Q.; Ramirez, M.A.; Sohn, Y.; Neininger, A.C.; Taneja, N.; McKinley, E.T.; et al. Supermeres are functional extracellular nanoparticles replete with disease biomarkers and therapeutic targets. Nat. Cell Biol. 2021, 23, 1240. [Google Scholar] [CrossRef]
  28. Valcz, G.; Buzás, E.I.; Kittel, Á.; Krenács, T.; Visnovitz, T.; Spisák, S.; Török, G.; Homolya, L.; Zsigrai, S.; Kiszler, G.; et al. En bloc release of MVB-like small extracellular vesicle clusters by colorectal carcinoma cells. J. Extracell. Vesicles 2019, 8, 1596668. [Google Scholar] [CrossRef]
  29. Malkin, E.Z.; Bratman, S.V. Bioactive DNA from extracellular vesicles and particles. Cell Death Dis. 2020, 11, 584. [Google Scholar] [CrossRef]
  30. Ma, L.; Li, Y.; Peng, J.; Wu, D.; Zhao, X.; Cui, Y.; Chen, L.; Yan, X.; Du, Y.; Yu, L. Discovery of the migrasome, an organelle mediating release of cytoplasmic contents during cell migration. Cell Res. 2015, 25, 24–38. [Google Scholar] [CrossRef]
  31. Seibold, T.; Waldenmaier, M.; Seufferlein, T.; Eiseler, T. Small extracellular vesicles and metastasis—Blame the messenger. Cancers 2021, 13, 4380. [Google Scholar] [CrossRef] [PubMed]
  32. Maacha, S.; Bhat, A.A.; Jimenez, L.; Raza, A.; Haris, M.; Uddin, S.; Grivel, J.C. Extracellular vesicles-mediated intercellular communication: Roles in the tumor microenvironment and anti-cancer drug resistance. Mol. Cancer 2019, 18, 55. [Google Scholar] [CrossRef] [PubMed]
  33. Menck, K.; Sivaloganathan, S.; Bleckmann, A.; Binder, C. Microvesicles in cancer: Small size, large potential. Int. J. Mol. Sci. 2020, 21, 5373. [Google Scholar] [CrossRef]
  34. Al-Nedawi, K.; Meehan, B.; Micallef, J.; Lhotak, V.; May, L.; Guha, A.; Rak, J. Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles derived from tumour cells. Nat. Cell Biol. 2008, 10, 619–624. [Google Scholar] [CrossRef]
  35. Zanetti-Domingues, L.C.; Bonner, S.E.; Martin-Fernandez, M.L.; Huber, V. Mechanisms of action of EGFR tyrosine kinase receptor incorporated in extracellular vesicles. Cells 2020, 9, 2505. [Google Scholar] [CrossRef] [PubMed]
  36. Bebelman, M.P.; Smit, M.J.; Pegtel, D.M.; Baglio, S.R. Biogenesis and function of extracellular vesicles in cancer. Pharmacol. Ther. 2018, 188, 1–11. [Google Scholar] [CrossRef]
  37. Rodrigues, P.; Melim, C.; Veiga, F.; Figueiras, A. An overview of exosomes in cancer therapy: A small solution to a big problem. Processes 2020, 8, 1561. [Google Scholar] [CrossRef]
  38. Han, L.; Lam, E.W.F.; Sun, Y. Extracellular vesicles in the tumor microenvironment: Old stories, but new tales. Mol. Cancer 2019, 18, 59. [Google Scholar] [CrossRef]
  39. Purba, E.R.; Saita, E.I.; Akhouri, R.R.; Öfverstedt, L.G.; Wilken, G.; Skoglund, U.; Maruyama, I.N. Allosteric activation of preformed EGF receptor dimers by a single ligand binding event. Front. Endocrinol. 2022, 13, 1042787. [Google Scholar] [CrossRef]
  40. Yarden, Y.; Sliwkowski, M.X. Untangling the ErbB signalling network. Nat. Rev. Mol. Cell Biol. 2001, 2, 127–137. [Google Scholar] [CrossRef]
  41. Chen, J.; Zeng, F.; Forrester, S.J.; Eguchi, S.; Zhang, M.Z.; Harris, R.C. Expression and function of the epidermal growth factor receptor in physiology and disease. Physiol. Rev. 2016, 96, 1025–1069. [Google Scholar] [CrossRef]
  42. Kaplan, M.; Narasimhan, S.; de Heus, C.; Mance, D.; van Doorn, S.; Houben, K.; Popov-Čeleketić, D.; Damman, R.; Katrukha, E.A.; Jain, P.; et al. EGFR dynamics change during activation in native membranes as revealed by NMR. Cell 2016, 167, 1241–1251.e11. [Google Scholar] [CrossRef]
  43. Needham, S.R.; Roberts, S.K.; Arkhipov, A.; Mysore, V.P.; Tynan, C.J.; Zanetti-Domingues, L.C.; Kim, E.T.; Losasso, V.; Korovesis, D.; Hirsch, M.; et al. EGFR oligomerization organizes kinase-active dimers into competent signalling platforms. Nat. Commun. 2016, 7, 13307. [Google Scholar] [CrossRef]
  44. Uribe, M.L.; Marrocco, I.; Yarden, Y. EGFR in cancer: Signaling mechanisms, drugs, and acquired resistance. Cancers 2021, 13, 2748. [Google Scholar] [CrossRef] [PubMed]
  45. Marrocco, I.; Romaniello, D.; Vaknin, I.; Drago-Garcia, D.; Oren, R.; Uribe, M.L.; Belugali Nataraj, N.; Ghosh, S.; Eilam, R.; Salame, T.; et al. Upfront admixing antibodies and EGFR inhibitors preempts sequential treatments in lung cancer models. EMBO Mol. Med. 2021, 13, e13144. [Google Scholar] [CrossRef] [PubMed]
  46. Maron, R.; Schechter, B.; Nataraj, N.B.; Ghosh, S.; Romaniello, D.; Marrocco, I.; Noronha, A.; Carvalho, S.; Yarden, Y.; Sela, M. Inhibition of a pancreatic cancer model by cooperative pairs of clinically approved and experimental antibodies. Biochem. Biophys. Res. Commun. 2019, 513, 219–225. [Google Scholar] [CrossRef] [PubMed]
  47. Romaniello, D.; Gelfo, V.; Pagano, F.; Ferlizza, E.; Sgarzi, M.; Mazzeschi, M.; Morselli, A.; Miano, C.; D’Uva, G.; Lauriola, M. Senescence-associated reprogramming induced by interleukin-1 impairs response to EGFR neutralization. Cell. Mol. Biol. Lett. 2022, 27, 20. [Google Scholar] [CrossRef]
  48. Romaniello, D.; Mazzeo, L.; Mancini, M.; Marrocco, I.; Noronha, A.; Kreitman, M.; Srivastava, S.; Ghosh, S.; Lindzen, M.; Salame, T.M.; et al. A combination of approved antibodies overcomes resistance of lung cancer to osimertinib by blocking bypass pathways. Clin. Cancer Res. 2018, 24, 5610–5621. [Google Scholar] [CrossRef]
  49. Romaniello, D.; Marrocco, I.; Nataraj, N.B.; Ferrer, I.; Drago-Garcia, D.; Vaknin, I.; Oren, R.; Lindzen, M.; Ghosh, S.; Kreitman, M.; et al. Targeting Her3, a catalytically defective receptor tyrosine kinase, prevents resistance of lung cancer to a third-generation EGFR kinase inhibitor. Cancers 2020, 12, 2394. [Google Scholar] [CrossRef]
  50. Fujiwara, T.; Eguchi, T.; Sogawa, C.; Ono, K.; Murakami, J.; Ibaragi, S.; Asaumi, J.; Calderwood, S.K.; Okamoto, K.; Kozaki, K. Carcinogenic epithelial-mesenchymal transition initiated by oral cancer exosomes is inhibited by anti-EGFR antibody cetuximab. Oral Oncol. 2018, 86, 251–257. [Google Scholar] [CrossRef]
  51. Sanderson, M.P.; Keller, S.; Alonso, A.; Riedle, S.; Dempsey, P.J.; Altevogt, P. Generation of novel, secreted epidermal growth factor receptor (EGFR/ErbB1) isoforms via metalloprotease-dependent ectodomain shedding and exosome secretion. J. Cell. Biochem. 2008, 103, 1783–1797. [Google Scholar] [CrossRef]
  52. Soltész, B.; Buglyó, G.; Németh, N.; Szilágyi, M.; Pös, O.; Szemes, T.; Balogh, I.; Nagy, B. The role of exosomes in cancer progression. Int. J. Mol. Sci. 2021, 23, 8. [Google Scholar] [CrossRef]
  53. Akers, J.C.; Gonda, D.; Kim, R.; Carter, B.S.; Chen, C.C. Biogenesis of extracellular vesicles (EV): Exosomes, microvesicles, retrovirus-like vesicles, and apoptotic bodies. J. Neurooncol. 2013, 113, 1–11. [Google Scholar] [CrossRef] [PubMed]
  54. Minciacchi, V.R.; Freeman, M.R.; Di Vizio, D. Extracellular vesicles in cancer: Exosomes, microvesicles and the emerging role of large oncosomes. Semin. Cell Dev. Biol. 2015, 40, 41–51. [Google Scholar] [CrossRef] [PubMed]
  55. Muralidharan-Chari, V.; Clancy, J.; Plou, C.; Romao, M.; Chavrier, P.; Raposo, G.; D’Souza-Schorey, C. ARF6-regulated shedding of tumor cell-derived plasma membrane microvesicles. Curr. Biol. 2009, 19, 1875–1885. [Google Scholar] [CrossRef] [PubMed]
  56. Tricarico, C.; Clancy, J.; D’Souza-Schorey, C. Biology and biogenesis of shed microvesicles. Small GTPases 2017, 8, 220–232. [Google Scholar] [CrossRef] [PubMed]
  57. Oka, S.; Uramoto, H.; Shimokawa, H.; Yamada, S.; Tanaka, F. Epidermal growth factor receptor-GEP100-Arf6 axis affects the prognosis of lung adenocarcinoma. Oncology 2014, 86, 263–270. [Google Scholar] [CrossRef]
  58. Sabe, H.; Hashimoto, S.; Morishige, M.; Ogawa, E.; Hashimoto, A.; Nam, J.-M.; Miura, K.; Yano, H.; Onodera, Y. The EGFR-GEP100-Arf6-AMAP1 Signaling pathway specific to breast cancer invasion and metastasis. Traffic 2009, 10, 982–993. [Google Scholar] [CrossRef]
  59. Matsumoto, Y.; Sakurai, H.; Kogashiwa, Y.; Kimura, T.; Matsumoto, Y.; Shionome, T.; Asano, M.; Saito, K.; Kohno, N. Inhibition of epithelial–mesenchymal transition by cetuximab via the EGFR-GEP100-Arf6-AMAP1 pathway in head and neck cancer. Head Neck 2017, 39, 476–485. [Google Scholar] [CrossRef]
  60. Eddama, M.M.R.; Gurung, R.; Fragkos, K.; Lorgelly, P.; Cohen, R.; Loizidou, M.; Clapp, L. The role of microvesicles as biomarkers in the screening of colorectal neoplasm. Cancer Med. 2022, 11, 2957–2968. [Google Scholar] [CrossRef]
  61. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef] [PubMed]
  62. Pols, M.S.; Klumperman, J. Trafficking and function of the tetraspanin CD63. Exp. Cell Res. 2009, 315, 1584–1592. [Google Scholar] [CrossRef]
  63. Raiborg, C.; Stenmark, H. The ESCRT machinery in endosomal sorting of ubiquitylated membrane proteins. Nature 2009, 458, 445–452. [Google Scholar] [CrossRef] [PubMed]
  64. Tamai, K.; Tanaka, N.; Nakano, T.; Kakazu, E.; Kondo, Y.; Inoue, J.; Shiina, M.; Fukushima, K.; Hoshino, T.; Sano, K.; et al. Exosome secretion of dendritic cells is regulated by Hrs, an ESCRT-0 protein. Biochem. Biophys. Res. Commun. 2010, 399, 384–390. [Google Scholar] [CrossRef] [PubMed]
  65. Abels, E.R.; Breakefield, X.O. Introduction to extracellular vesicles: Biogenesis, RNA cargo selection, content, release, and uptake. Cell. Mol. Neurobiol. 2016, 36, 301–312. [Google Scholar] [CrossRef] [PubMed]
  66. Bakker, J.; Spits, M.; Neefjes, J.; Berlin, I. The EGFR odyssey—From activation to destruction in space and time. J. Cell Sci. 2017, 130, 4087–4096. [Google Scholar] [CrossRef]
  67. Gajiwala, K.S.; Feng, J.; Ferre, R.; Ryan, K.; Brodsky, O.; Weinrich, S.; Kath, J.C.; Stewart, A. insights into the aberrant activity of mutant EGFR kinase domain and drug recognition. Structure 2013, 21, 209–219. [Google Scholar] [CrossRef]
  68. Kim, Y.; Li, Z.; Apetri, M.; Luo, B.; Settleman, J.E.; Anderson, K.S. Temporal resolution of autophosphorylation for normal and oncogenic forms of EGFR and differential effects of gefitinib. Biochemistry 2012, 51, 5212–5222. [Google Scholar] [CrossRef]
  69. Zanetti-Domingues, L.C.; Bonner, S.E.; Iyer, R.S.; Martin-Fernandez, M.L.; Huber, V. Cooperation and interplay between EGFR signalling and extracellular vesicle biogenesis in cancer. Cells 2020, 9, 2639. [Google Scholar] [CrossRef]
  70. Zhang, Y.; Liu, Y.; Liu, H.; Tang, W.H. Exosomes: Biogenesis, biologic function and clinical potential. Cell Biosci. 2019, 9, 19. [Google Scholar] [CrossRef]
  71. Ghossoub, R.; Lembo, F.; Rubio, A.; Gaillard, C.B.; Bouchet, J.; Vitale, N.; Slavík, J.; Machala, M.; Zimmermann, P. Syntenin-ALIX exosome biogenesis and budding into multivesicular bodies are controlled by ARF6 and PLD2. Nat. Commun. 2014, 5, 3477. [Google Scholar] [CrossRef] [PubMed]
  72. Mallawaaratchy, D.M.; Hallal, S.; Russell, B.; Ly, L.; Ebrahimkhani, S.; Wei, H.; Christopherson, R.I.; Buckland, M.E.; Kaufman, K.L. Comprehensive proteome profiling of glioblastoma-derived extracellular vesicles identifies markers for more aggressive disease. J. Neurooncol. 2017, 131, 233–244. [Google Scholar] [CrossRef] [PubMed]
  73. Vesiclepedia: Home—Extracellular Vesicles Database. Available online: http://microvesicles.org/ (accessed on 22 March 2023).
  74. ExoCarta. Available online: http://www.exocarta.org/ (accessed on 22 March 2023).
  75. Mateescu, B.; Kowal, E.J.K.; van Balkom, B.W.M.; Bartel, S.; Bhattacharyya, S.N.; Buzás, E.I.; Buck, A.H.; de Candia, P.; Chow, F.W.N.; Das, S.; et al. Obstacles and opportunities in the functional analysis of extracellular vesicle RNA—An ISEV position paper. J. Extracell. Vesicles 2017, 6, 1286095. [Google Scholar] [CrossRef]
  76. Witwer, K.W.; Buzás, E.I.; Bemis, L.T.; Bora, A.; Lässer, C.; Lötvall, J.; Nolte-‘t Hoen, E.N.; Piper, M.G.; Sivaraman, S.; Skog, J.; et al. Standardization of sample collection, isolation and analysis methods in extracellular vesicle research. J. Extracell. Vesicles 2013, 2, 20360. [Google Scholar] [CrossRef] [PubMed]
  77. Brennan, K.; Martin, K.; FitzGerald, S.P.; O’Sullivan, J.; Wu, Y.; Blanco, A.; Richardson, C.; Mc Gee, M.M. A Comparison of methods for the isolation and separation of extracellular vesicles from protein and lipid particles in human serum. Sci. Rep. 2020, 10, 1039. [Google Scholar] [CrossRef]
  78. Théry, C.; Clayton, A.; Amigorena, S.; Raposo, G.; Clayton, A. Isolation and characterization of exosomes from cell culture supernatants. Curr. Protoc. Cell Biol. 2006, 30, 22. [Google Scholar] [CrossRef]
  79. Jablonska, J.; Pietrowska, M.; Ludwig, S.; Lang, S.; Thakur, B.K. Challenges in the isolation and proteomic analysis of cancer exosomes—Implications for translational research. Proteomes 2019, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  80. Ismail, N.; Wang, Y.; Dakhlallah, D.; Moldovan, L.; Agarwal, K.; Batte, K.; Shah, P.; Wisler, J.; Eubank, T.D.; Tridandapani, S.; et al. Macrophage microvesicles induce macrophage differentiation and miR-223 transfer. Blood 2013, 121, 984–995. [Google Scholar] [CrossRef]
  81. White, I.J.; Bailey, L.M.; Aghakhani, M.R.; Moss, S.E.; Futter, C.E. EGF Stimulates annexin 1-dependent inward vesiculation in a multivesicular endosome subpopulation. EMBO J. 2006, 25, 7600759. [Google Scholar] [CrossRef]
  82. González-Masiá, J.A.; García-Olmo, D.; García-Olmo, D.C. Circulating nucleic acids in plasma and serum (CNAPS): Applications in oncology. OncoTargets Ther. 2013, 6, 819–832. [Google Scholar] [CrossRef]
  83. Logozzi, M.; Spugnini, E.; Mizzoni, D.; Di Raimo, R.; Fais, S. Extracellular acidity and increased exosome release as key phenotypes of malignant tumors. Cancer Metastasis Rev. 2019, 38, 93–101. [Google Scholar] [CrossRef] [PubMed]
  84. Arbelaiz, A.; Azkargorta, M.; Krawczyk, M.; Santos-Laso, A.; Lapitz, A.; Perugorria, M.J.; Erice, O.; Gonzalez, E.; Jimenez-Agüero, R.; Lacasta, A.; et al. Serum extracellular vesicles contain protein biomarkers for primary sclerosing cholangitis and cholangiocarcinoma. Hepatology 2017, 66, 1125–1143. [Google Scholar] [CrossRef]
  85. Cumba Garcia, L.M.; Peterson, T.E.; Cepeda, M.A.; Johnson, A.J.; Parney, I.F. Isolation and analysis of plasma-derived exosomes in patients with glioma. Front. Oncol. 2019, 9, 651. [Google Scholar] [CrossRef]
  86. Mousavi, S.; Moallem, R.; Hassanian, S.M.; Sadeghzade, M.; Mardani, R.; Ferns, G.A.; Khazaei, M.; Avan, A. Tumor-derived exosomes: Potential biomarkers and therapeutic target in the treatment of colorectal cancer. J. Cell. Physiol. 2019, 234, 12422–12432. [Google Scholar] [CrossRef]
  87. Rashed, M.H.; Bayraktar, E.; Helal, G.K.; Abd-Ellah, M.; Amero, P.; Chavez-Reyes, A.; Rodriguez-Aguayo, C. Exosomes: From garbage bins to promising therapeutic targets. Int. J. Mol. Sci. 2017, 18, 538. [Google Scholar] [CrossRef] [PubMed]
  88. Shao, Y.; Shen, Y.; Chen, T.; Xu, F.; Chen, X.; Zheng, S. The functions and clinical applications of tumor-derived exosomes. Oncotarget 2016, 7, 60736–60751. [Google Scholar] [CrossRef]
  89. Baran, J.; Baj-Krzyworzeka, M.; Weglarczyk, K.; Szatanek, R.; Zembela, M.; Barbasz, J.; Czupryna, A.; Szczepanik, A.; Zembala, M. Circulating tumour-derived microvesicles in plasma of gastric cancer patients. Cancer Immunol. Immunother. 2010, 59, 841–850. [Google Scholar] [CrossRef]
  90. Zhao, L.; Bi, Y.; Kou, J.; Shi, J.; Piao, D. Phosphatidylserine exposing-platelets and microparticles promote procoagulant activity in colon cancer patients. J. Exp. Clin. Cancer Res. 2016, 35, 54. [Google Scholar] [CrossRef] [PubMed]
  91. Galindo-Hernandez, O.; Villegas-Comonfort, S.; Candanedo, F.; González-Vázquez, M.C.; Chavez-Ocaña, S.; Jimenez-Villanueva, X.; Sierra-Martinez, M.; Salazar, E.P. Elevated concentration of microvesicles isolated from peripheral blood in breast cancer patients. Arch. Med. Res. 2013, 44, 208–214. [Google Scholar] [CrossRef]
  92. Caivano, A.; Laurenzana, I.; De Luca, L.; La Rocca, F.; Simeon, V.; Trino, S.; D’Auria, F.; Traficante, A.; Maietti, M.; Izzo, T.; et al. High serum levels of extracellular vesicles expressing malignancy-related markers are released in patients with various types of hematological neoplastic disorders. Tumor Biol. 2015, 36, 9739–9752. [Google Scholar] [CrossRef]
  93. Kharmate, G.; Hosseini-Beheshti, E.; Caradec, J.; Chin, M.Y.; Tomlinson Guns, E.S. Epidermal growth factor receptor in prostate cancer derived exosomes. PLoS ONE 2016, 11, 0154967. [Google Scholar] [CrossRef]
  94. König, L.; Kasimir-Bauer, S.; Bittner, A.K.; Hoffmann, O.; Wagner, B.; Santos Manvailer, L.F.; Kimmig, R.; Horn, P.A.; Rebmann, V. Elevated levels of extracellular vesicles are associated with therapy failure and disease progression in breast cancer patients undergoing neoadjuvant chemotherapy. Oncoimmunology 2018, 7, e1376153. [Google Scholar] [CrossRef]
  95. Teng, F.; Fussenegger, M. Shedding light on extracellular vesicle biogenesis and bioengineering. Adv. Sci. 2021, 8, 2003505. [Google Scholar] [CrossRef]
  96. Escrevente, C.; Grammel, N.; Kandzia, S.; Zeiser, J.; Tranfield, E.M.; Conradt, H.S.; Costa, J. Sialoglycoproteins and N-Glycans from secreted exosomes of ovarian carcinoma cells. PLoS ONE 2013, 8, 0078631. [Google Scholar] [CrossRef] [PubMed]
  97. Hosseini-Beheshti, E.; Pham, S.; Adomat, H.; Li, N.; Tomlinson Guns, E.S. Exosomes as biomarker enriched microvesicles: Characterization of exosomal proteins derived from a panel of prostate cell lines with distinct AR phenotypes. Mol. Cell. Proteomics 2012, 11, 863. [Google Scholar] [CrossRef] [PubMed]
  98. Choi, D.S.; Park, J.O.; Jang, S.C.; Yoon, Y.J.; Jung, J.W.; Choi, D.Y.; Kim, J.W.; Kang, J.S.; Park, J.; Hwang, D.; et al. Proteomic analysis of microvesicles derived from human colorectal cancer ascites. Proteomics 2011, 11, 2745–2751. [Google Scholar] [CrossRef] [PubMed]
  99. Li, F.; Zhao, X.; Sun, R.; Ou, J.; Huang, J.; Yang, N.; Xu, T.; Li, J.; He, X.; Li, C.; et al. EGFR-rich extracellular vesicles derived from highly metastatic nasopharyngeal carcinoma cells accelerate tumour metastasis through PI3K/AKT pathway-suppressed ROS. J. Extracell. Vesicles 2020, 10, e12003. [Google Scholar] [CrossRef] [PubMed]
  100. Jouida, A.; O’Callaghan, M.; Mc Carthy, C.; Fabre, A.; Nadarajan, P.; Keane, M.P. Exosomes from EGFR-mutated adenocarcinoma induce a hybrid EMT and MMP9-dependant tumor invasion. Cancers 2022, 14, 3776. [Google Scholar] [CrossRef] [PubMed]
  101. Montermini, L.; Meehan, B.; Garnier, D.; Lee, W.J.; Lee, T.H.; Guha, A.; Al-Nedawi, K.; Rak, J. Inhibition of oncogenic epidermal growth factor receptor kinase triggers release of exosome-like extracellular vesicles and impacts their phosphoprotein and DNA content. J. Biol. Chem. 2015, 290, 24534–24546. [Google Scholar] [CrossRef]
  102. Fabbiano, F.; Corsi, J.; Gurrieri, E.; Trevisan, C.; Notarangelo, M.; D’Agostino, V.G. RNA packaging into extracellular vesicles: An orchestra of RNA-binding proteins? J. Extracell. Vesicles 2020, 10, e12043. [Google Scholar] [CrossRef]
  103. Chen, M.; Xu, R.; Rai, A.; Suwakulsiri, W.; Izumikawa, K.; Ishikawa, H.; Greening, D.W.; Takahashi, N.; Simpson, R.J. Distinct shed microvesicle and exosome microRNA signatures reveal diagnostic markers for colorectal cancer. PLoS ONE 2019, 14, e0210003. [Google Scholar] [CrossRef]
  104. Ferlizza, E.; Solmi, R.; Sgarzi, M.; Ricciardiello, L.; Lauriola, M. The roadmap of colorectal cancer screening. Cancers 2021, 13, 1101. [Google Scholar] [CrossRef] [PubMed]
  105. Min, L.; Zhu, S.; Chen, L.; Liu, X.; Wei, R.; Zhao, L.; Yang, Y.; Zhang, Z.; Kong, G.; Li, P.; et al. Evaluation of circulating small extracellular vesicles derived miRNAs as Biomarkers of early colon cancer: A comparison with plasma total miRNAs. J. Extracell. Vesicles 2019, 8, 1643670. [Google Scholar] [CrossRef]
  106. Ogata-Kawata, H.; Izumiya, M.; Kurioka, D.; Honma, Y.; Yamada, Y.; Furuta, K.; Gunji, T.; Ohta, H.; Okamoto, H.; Sonoda, H.; et al. Circulating exosomal microRNAs as biomarkers of colon cancer. PLoS ONE 2014, 9, e92921. [Google Scholar] [CrossRef] [PubMed]
  107. Yan, S.; Han, B.; Gao, S.; Wang, X.; Wang, Z.; Wang, F.; Zhang, J.; Xu, D.; Sun, B. Exosome-encapsulated microRNAs as circulating biomarkers for colorectal cancer. Oncotarget 2017, 8, 60149–60158. [Google Scholar] [CrossRef]
  108. Peng, Z.Y.; Gu, R.H.; Yan, B. Downregulation of exosome-encapsulated miR-548c-5p is associated with poor prognosis in colorectal cancer. J. Cell. Biochem. 2019, 120, 1457–1463. [Google Scholar] [CrossRef]
  109. Liu, C.; Eng, C.; Shen, J.; Lu, Y.; Yoko, T.; Mehdizadeh, A.; Chang, G.J.; Rodriguez-Bigas, M.A.; Li, Y.; Chang, P.; et al. Serum exosomal miR-4772-3p is a predictor of tumor recurrence in stage II and III colon cancer. Oncotarget 2016, 7, 76250–76260. [Google Scholar] [CrossRef]
  110. Cha, B.S.; Park, K.S.; Park, J.S. Signature mRNA markers in extracellular vesicles for the accurate diagnosis of colorectal cancer. J. Biol. Eng. 2020, 14, 4. [Google Scholar] [CrossRef] [PubMed]
  111. Krug, A.K.; Enderle, D.; Karlovich, C.; Priewasser, T.; Bentink, S.; Spiel, A.; Brinkmann, K.; Emenegger, J.; Grimm, D.G.; Castellanos-Rizaldos, E.; et al. Improved EGFR mutation detection using combined exosomal RNA and circulating tumor DNA in NSCLC patient plasma. Ann. Oncol. 2018, 29, 700–706. [Google Scholar] [CrossRef]
  112. Xia, J.; Luo, M.; Dai, L.; Wang, L.; Wang, L.; Zhu, J. Serum exosomal microRNAs as predictive markers for EGFR mutations in non–small-cell lung cancer. J. Clin. Lab. Anal. 2021, 35, e23743. [Google Scholar] [CrossRef]
  113. Chen, L.; Yang, Y.; Yue, R.; Peng, X.; Yu, H.; Huang, X. Exosomes derived from hypoxia-induced alveolar epithelial cells stimulate interstitial pulmonary fibrosis through a HOTAIRM1-dependent mechanism. Lab. Investig. 2022, 102, 935–944. [Google Scholar] [CrossRef] [PubMed]
  114. Martinelli, M.; Pacilli, A.M.G.; Rivetti, S.; Lauriola, M.; Fasano, L.; Carbonara, P.; Mattei, G.; Valentini, I.; Scapoli, L.; Solmi, R. A role for epidermal growth factor receptor in idiopathic pulmonary fibrosis onset. Mol. Biol. Rep. 2011, 38, 4613–4617. [Google Scholar] [CrossRef]
  115. Matsumura, T.; Sugimachi, K.; Iinuma, H.; Takahashi, Y.; Kurashige, J.; Sawada, G.; Ueda, M.; Uchi, R.; Ueo, H.; Takano, Y.; et al. Exosomal microRNA in serum is a novel biomarker of recurrence in human colorectal cancer. Br. J. Cancer 2015, 113, 275–281. [Google Scholar] [CrossRef] [PubMed]
  116. Ji, H.; Chen, M.; Greening, D.W.; He, W.; Rai, A.; Zhang, W.; Simpson, R.J. Deep sequencing of RNA from three different extracellular vesicle (EV) subtypes released from the human LIM1863 colon cancer cell line uncovers distinct miRNA-enrichment signatures. PLoS ONE 2014, 9, e110314. [Google Scholar] [CrossRef] [PubMed]
  117. Dong, L.; Lin, W.; Qi, P.; Xu, M.; Wu, X.; Ni, S.; Huang, D.; Weng, W.; Tan, C.; Sheng, W.; et al. Circulating long RNAs in serum extracellular vesicles: Their characterization and potential application as biomarkers for diagnosis of colorectal cancer. Cancer Epidemiol. Biomarkers Prev. 2016, 25, 1158–1166. [Google Scholar] [CrossRef]
  118. Li, Q.; Shao, Y.; Zhang, X.; Zheng, T.; Miao, M.; Qin, L.; Wang, B.; Ye, G.; Xiao, B.; Guo, J. Plasma long noncoding RNA protected by exosomes as a potential stable biomarker for gastric cancer. Tumor Biol. 2015, 36, 2007–2012. [Google Scholar] [CrossRef]
  119. Ohno, S.I.; Takanashi, M.; Sudo, K.; Ueda, S.; Ishikawa, A.; Matsuyama, N.; Fujita, K.; Mizutani, T.; Ohgi, T.; Ochiya, T.; et al. Systemically injected exosomes targeted to EGFR deliver antitumor microRNA to breast cancer cells. Mol. Ther. 2013, 21, 185–191. [Google Scholar] [CrossRef]
  120. Sheykhhasan, M.; Kalhor, N.; Sheikholeslami, A.; Dolati, M.; Amini, E.; Fazaeli, H. Exosomes of mesenchymal stem cells as a proper vehicle for transfecting MiR-145 into the breast cancer cell line and its effect on metastasis. Biomed Res. Int. 2021, 2021, 5516078. [Google Scholar] [CrossRef]
  121. Wang, X.; Zhang, H.; Bai, M.; Ning, T.; Ge, S.; Deng, T.; Liu, R.; Zhang, L.; Ying, G.; Ba, Y. Exosomes serve as nanoparticles to deliver anti-miR-214 to reverse chemoresistance to cisplatin in gastric cancer. Mol. Ther. 2018, 26, 774–783. [Google Scholar] [CrossRef]
  122. Kamerkar, S.; LeBleu, V.S.; Sugimoto, H.; Yang, S.; Ruivo, C.F.; Melo, S.A.; Lee, J.J.; Kalluri, R. Exosomes facilitate therapeutic targeting of oncogenic KRAS in pancreatic cancer. Nature 2017, 546, 498–503. [Google Scholar] [CrossRef]
  123. Limoni, S.K.; Moghadam, M.F.; Moazzeni, S.M.; Gomari, H.; Salimi, F. Engineered exosomes for targeted transfer of siRNA to HER2 positive breast cancer cells. Appl. Biochem. Biotechnol. 2019, 187, 352–364. [Google Scholar] [CrossRef]
  124. Kaźmierczak, D.; Eide, I.J.Z.; Gencheva, R.; Lai, Y.; Lewensohn, R.; Tsakonas, G.; Grundberg, O.; de Petris, L.; McGowan, M.; Brustugun, O.T.; et al. Elevated expression of miR-494-3p is associated with resistance to osimertinib in EGFR T790M-positive non-small cell lung cancer. Transl. Lung Cancer Res. 2022, 11, 722–734. [Google Scholar] [CrossRef]
  125. Li, X.; Chen, C.; Wang, Z.; Liu, J.; Sun, W.; Shen, K.; Lv, Y.; Zhu, S.; Zhan, P.; Lv, T.; et al. Elevated exosome-derived miRNAs predict osimertinib resistance in non-small cell lung cancer. Cancer Cell Int. 2021, 21, 428. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, H.; Deng, T.; Liu, R.; Bai, M.; Zhou, L.; Wang, X.; Li, S.; Wang, X.; Yang, H.; Li, J.; et al. Exosome-delivered EGFR regulates liver microenvironment to promote gastric cancer liver metastasis. Nat. Commun. 2017, 8, 15016. [Google Scholar] [CrossRef] [PubMed]
  127. Cai, J.; Guan, W.; Tan, X.; Chen, C.; Li, L.; Wang, N.; Zou, X.; Zhou, F.; Wang, J.; Pei, F.; et al. SRY Gene transferred by extracellular vesicles accelerates atherosclerosis by promotion of leucocyte adherence to endothelial cells. Clin. Sci. 2015, 129, 259–269. [Google Scholar] [CrossRef] [PubMed]
  128. Sansone, P.; Savini, C.; Kurelac, I.; Chang, Q.; Amato, L.B.; Strillacci, A.; Stepanova, A.; Iommarini, L.; Mastroleo, C.; Daly, L.; et al. Packaging and transfer of mitochondrial DNA via exosomes regulate escape from dormancy in hormonal therapy-resistant breast cancer. Proc. Natl. Acad. Sci. USA 2017, 114, E9066–E9075. [Google Scholar] [CrossRef] [PubMed]
  129. Takahashi, A.; Okada, R.; Nagao, K.; Kawamata, Y.; Hanyu, A.; Yoshimoto, S.; Takasugi, M.; Watanabe, S.; Kanemaki, M.T.; Obuse, C.; et al. Exosomes maintain cellular homeostasis by excreting harmful DNA from cells. Nat. Commun. 2017, 8, 15287. [Google Scholar] [CrossRef]
  130. Thakur, K.; Singh, M.S.; Feldstein-Davydova, S.; Hannes, V.; Hershkovitz, D.; Tsuriel, S. Extracellular vesicle-derived DNA vs. cfDNA as a biomarker for the detection of colon cancer. Genes 2021, 12, 1171. [Google Scholar] [CrossRef]
  131. Fricke, F.; Lee, J.; Michalak, M.; Warnken, U.; Hausser, I.; Suarez-Carmona, M.; Halama, N.; Schnölzer, M.; Kopitz, J.; Gebert, J. TGFBR2-dependent alterations of exosomal cargo and functions in DNA mismatch repair-deficient HCT116 colorectal cancer cells. Cell Commun. Signal. 2017, 15, 14. [Google Scholar] [CrossRef]
  132. Lee, T.H.; Chennakrishnaiah, S.; Audemard, E.; Montermini, L.; Meehan, B.; Rak, J. Oncogenic Ras-driven cancer cell vesiculation leads to emission of double-stranded DNA capable of interacting with target cells. Biochem. Biophys. Res. Commun. 2014, 451, 295–301. [Google Scholar] [CrossRef]
  133. Castellanos-Rizaldos, E.; Grimm, D.G.; Tadigotla, V.; Hurley, J.; Healy, J.; Neal, P.L.; Sher, M.; Venkatesan, R.; Karlovich, C.; Raponi, M.; et al. Exosome-based detection of EGFR T790M in plasma from non–small cell lung cancer patients. Clin. Cancer Res. 2018, 24, 2944–2950. [Google Scholar] [CrossRef] [PubMed]
  134. Allenson, K.; Castillo, J.; San Lucas, F.A.; Scelo, G.; Kim, D.U.; Bernard, V.; Davis, G.; Kumar, T.; Katz, M.; Overman, M.J.; et al. High prevalence of mutant KRAS in circulating exosome-derived DNA from early-stage pancreatic cancer patients. Ann. Oncol. 2017, 28, 741. [Google Scholar] [CrossRef]
  135. Yang, S.; Che, S.P.Y.; Kurywchak, P.; Tavormina, J.L.; Gansmo, L.B.; Correa de Sampaio, P.; Tachezy, M.; Bockhorn, M.; Gebauer, F.; Haltom, A.R.; et al. Detection of mutant KRAS and TP53 DNA in circulating exosomes from healthy individuals and patients with pancreatic cancer. Cancer Biol. Ther. 2017, 18, 158. [Google Scholar] [CrossRef] [PubMed]
  136. Zocco, D.; Bernardi, S.; Novelli, M.; Astrua, C.; Fava, P.; Zarovni, N.; Carpi, F.M.; Bianciardi, L.; Malavenda, O.; Quaglino, P.; et al. Isolation of extracellular vesicles improves the detection of mutant DNA from plasma of metastatic melanoma patients. Sci. Rep. 2020, 10, 15745. [Google Scholar] [CrossRef] [PubMed]
  137. Lázaro-Ibáñez, E.; Sanz-Garcia, A.; Visakorpi, T.; Escobedo-Lucea, C.; Siljander, P.; Ayuso-Sacido, Á.; Yliperttula, M. Different gDNA content in the subpopulations of prostate cancer extracellular vesicles: Apoptotic bodies, microvesicles, and exosomes. Prostate 2014, 74, 1379–1390. [Google Scholar] [CrossRef]
  138. Vagner, T.; Spinelli, C.; Minciacchi, V.R.; Balaj, L.; Zandian, M.; Conley, A.; Zijlstra, A.; Freeman, M.R.; Demichelis, F.; De, S.; et al. Large extracellular vesicles carry most of the tumour DNA circulating in prostate cancer patient plasma. J. Extracell. Vesicles 2018, 7, 1505403. [Google Scholar] [CrossRef]
  139. Yokoi, A.; Villar-Prados, A.; Oliphint, P.A.; Zhang, J.; Song, X.; DeHoff, P.; Morey, R.; Liu, J.; Roszik, J.; Clise-Dwyer, K.; et al. Mechanisms of nuclear content loading to exosomes. Sci. Adv. 2019, 5, eaax8849. [Google Scholar] [CrossRef]
  140. García-Silva, S.; Gallardo, M.; Peinado, H. DNA-loaded extracellular vesicles in liquid biopsy: Tiny players with big potential? Front. Cell Dev. Biol. 2021, 8, 622579. [Google Scholar] [CrossRef]
  141. Fernando, M.R.; Jiang, C.; Krzyzanowski, G.D.; Ryan, W.L. New evidence that a large proportion of human blood plasma cell-free DNA is localized in exosomes. PLoS ONE 2017, 12, e0183915. [Google Scholar] [CrossRef]
  142. Jin, Y.; Chen, K.; Wang, Z.; Wang, Y.; Liu, J.; Lin, L.; Shao, Y.; Gao, L.; Yin, H.; Cui, C.; et al. DNA in serum extracellular vesicles is stable under different storage conditions. BMC Cancer 2016, 16, 753. [Google Scholar] [CrossRef]
  143. Choi, D.; Montermini, L.; Jeong, H.; Sharma, S.; Meehan, B.; Rak, J. Mapping subpopulations of cancer cell-derived extracellular vesicles and particles by nano-flow cytometry. ACS Nano 2019, 13, 10499–10511. [Google Scholar] [CrossRef]
  144. Bernard, V.; Kim, D.U.; San Lucas, F.A.; Castillo, J.; Allenson, K.; Mulu, F.C.; Stephens, B.M.; Huang, J.; Semaan, A.; Guerrero, P.A.; et al. Circulating nucleic acids are associated with outcomes of patients with pancreatic cancer. Gastroenterology 2019, 156, 108–118.e4. [Google Scholar] [CrossRef] [PubMed]
  145. Keserű, J.S.; Soltész, B.; Lukács, J.; Márton, É.; Szilágyi-Bónizs, M.; Penyige, A.; Póka, R.; Nagy, B. Detection of cell-free, exosomal and whole blood mitochondrial DNA copy number in plasma or whole blood of patients with serous epithelial ovarian cancer. J. Biotechnol. 2019, 298, 76–81. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, L.; Li, Y.; Guan, X.; Zhao, J.; Shen, L.; Liu, J. Exosomal double-stranded DNA as a biomarker for the diagnosis and preoperative assessment of pheochromocytoma and paraganglioma. Mol. Cancer 2018, 17, 128. [Google Scholar] [CrossRef]
  147. Lucchetti, D.; Zurlo, I.V.; Colella, F.; Ricciardi-Tenore, C.; Di Salvatore, M.; Tortora, G.; De Maria, R.; Giuliante, F.; Cassano, A.; Basso, M.; et al. Mutational status of plasma exosomal KRAS predicts outcome in patients with metastatic colorectal cancer. Sci. Rep. 2021, 11, 22686. [Google Scholar] [CrossRef] [PubMed]
  148. Li, X.; Wang, Q.; Wang, R. Roles of exosome genomic DNA in colorectal cancer. Front. Pharmacol. 2022, 13, 923232. [Google Scholar] [CrossRef]
  149. Lee, J.S.; Hur, J.Y.; Kim, I.A.; Kim, H.J.; Choi, C.M.; Lee, J.C.; Kim, W.S.; Lee, K.Y. Liquid biopsy using the supernatant of a pleural effusion for EGFR genotyping in pulmonary adenocarcinoma patients: A comparison between cell-free DNA and extracellular vesicle-derived DNA. BMC Cancer 2018, 18, 1236. [Google Scholar] [CrossRef]
  150. Park, J.; Lee, C.; Eom, J.S.; Kim, M.-H.; Cho, Y.-K. Detection of EGFR mutations using bronchial washing-derived extracellular vesicles in patients with non-small-cell lung carcinoma. Cancers 2020, 12, 2822. [Google Scholar] [CrossRef]
  151. Zduriencikova, M.; Gronesova, P.; Cholujova, D.; Sedlak, J. Potential biomarkers of exosomal cargo in endocrine signaling. Endocr. Regul. 2015, 49, 141–150. [Google Scholar] [CrossRef]
  152. Carretero-González, A.; Otero, I.; Carril-Ajuria, L.; de Velasco, G.; Manso, L. Exosomes: Definition, role in tumor development and clinical implications. Cancer Microenviron. 2018, 11, 13–21. [Google Scholar] [CrossRef]
  153. 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.; et al. Glypican-1 identifies cancer exosomes and detects early pancreatic cancer. Nature 2015, 523, 177–182. [Google Scholar] [CrossRef]
  154. Caradec, J.; Kharmate, G.; Hosseini-Beheshti, E.; Adomat, H.; Gleave, M.; Guns, E. Reproducibility and efficiency of serum-derived exosome extraction methods. Clin. Biochem. 2014, 47, 1286–1292. [Google Scholar] [CrossRef]
  155. Jin, X.; Chen, Y.; Chen, H.; Fei, S.; Chen, D.; Cai, X.; Liu, L.; Lin, B.; Su, H.; Zhao, L.; et al. Evaluation of tumor-derived exosomal miRNA as potential diagnostic biomarkers for early-stage non-small cell lung cancer using next-generation sequencing. Clin. Cancer Res. 2017, 23, 5311–5319. [Google Scholar] [CrossRef] [PubMed]
  156. Xu, Z.; Wang, Z.; Sun, H.; Xin, H. Evaluation of exosomal miRNA in blood as a potential diagnostic biomarker for human non-small cell lung cancer. Med. Sci. Monit. 2020, 26, e924721. [Google Scholar] [CrossRef] [PubMed]
  157. Cazzoli, R.; Buttitta, F.; Di Nicola, M.; Malatesta, S.; Marchetti, A.; Rom, W.N.; Pass, H.I. MicroRNAs derived from circulating exosomes as noninvasive biomarkers for screening and diagnosing lung cancer. J. Thorac. Oncol. 2013, 8, 1156–1162. [Google Scholar] [CrossRef]
  158. Ortega, F.G.; Piguillem, S.V.; Messina, G.A.; Tortella, G.R.; Rubilar, O.; Jiménez Castillo, M.I.; Lorente, J.A.; Serrano, M.J.; Raba, J.; Fernández Baldo, M.A. EGFR detection in extracellular vesicles of breast cancer patients through immunosensor based on silica-chitosan nanoplatform. Talanta 2019, 194, 243–252. [Google Scholar] [CrossRef]
  159. Wang, H.; Jiang, D.; Li, W.; Xiang, X.; Zhao, J.; Yu, B.; Wang, C.; He, Z.; Zhu, L.; Yang, Y. Evaluation of serum extracellular vesicles as noninvasive diagnostic markers of glioma. Theranostics 2019, 9, 5347–5358. [Google Scholar] [CrossRef]
  160. Cobas EGFR Mutation Test. Available online: https://www.fda.gov/drugs/resources-information-approved-drugs/cobas-egfr-mutation-test-v2 (accessed on 23 February 2023).
  161. Jouida, A.; McCarthy, C.; Fabre, A.; Keane, M.P. Exosomes: A new perspective in EGFR-mutated lung cancer. Cancer Metastasis Rev. 2021, 40, 589–601. [Google Scholar] [CrossRef] [PubMed]
  162. Kalluri, R. The biology and function of fibroblasts in cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef]
  163. Rabinowits, G.; Gerçel-Taylor, C.; Day, J.M.; Taylor, D.D.; Kloecker, G.H. Exosomal microRNA: A diagnostic marker for lung cancer. Clin. Lung Cancer 2009, 10, 42–46. [Google Scholar] [CrossRef]
  164. Liu, J.; Ren, L.; Li, S.; Li, W.; Zheng, X.; Yang, Y.; Fu, W.; Yi, J.; Wang, J.; Du, G. The biology, function, and applications of exosomes in cancer. Acta Pharm. Sin. B 2021, 11, 2783–2797. [Google Scholar] [CrossRef]
  165. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  166. Choi, D.; Montermini, L.; Kim, D.-K.; Meehan, B.; Roth, F.P.; Rak, J. The impact of oncogenic EGFRvIII on the proteome of extracellular vesicles released from glioblastoma cells. Mol. Cell. Proteomics 2018, 17, 1948–1964. [Google Scholar] [CrossRef] [PubMed]
  167. Choi, D.Y.; You, S.; Jung, J.H.; Lee, J.C.; Rho, J.K.; Lee, K.Y.; Freeman, M.R.; Kim, K.P.; Kim, J. Extracellular vesicles shed from gefitinib-resistant nonsmall cell lung cancer regulate the tumor microenvironment. Proteomics 2014, 14, 1845–1856. [Google Scholar] [CrossRef] [PubMed]
  168. Li, X.-Q.; Liu, J.-T.; Fan, L.-L.; Liu, Y.; Cheng, L.; Wang, F.; Yu, H.-Q.; Gao, J.; Wei, W.; Wang, H.; et al. Exosomes derived from gefitinib-treated EGFR-mutant lung cancer cells alter cisplatin sensitivity via up-regulating autophagy. Oncotarget 2016, 7, 24585–24595. [Google Scholar] [CrossRef]
  169. Fujiwara, T.; Eguchi, T.; Sogawa, C.; Ono, K.; Murakami, J.; Ibaragi, S.; Asaumi, J.; Okamoto, K.; Calderwood, S.K.; Kozaki, K. Anti-EGFR antibody cetuximab is secreted by oral squamous cell carcinoma and alters EGF-driven mesenchymal transition. Biochem. Biophys. Res. Commun. 2018, 503, 1267–1272. [Google Scholar] [CrossRef]
  170. Garnier, D.; Magnus, N.; Lee, T.H.; Bentley, V.; Meehan, B.; Milsom, C.; Montermini, L.; Kislinger, T.; Rak, J. Cancer cells induced to express mesenchymal phenotype release exosome-like extracellular vesicles carrying tissue factor. J. Biol. Chem. 2012, 287, 43565–43572. [Google Scholar] [CrossRef]
  171. Hasegawa, K.; Kuwata, K.; Yoshitake, J.; Shimomura, S.; Uchida, K.; Shibata, T. Extracellular vesicles derived from inflamed murine colorectal tissue induce fibroblast proliferation via epidermal growth factor receptor. FEBS J. 2021, 288, 1906–1917. [Google Scholar] [CrossRef]
  172. Svrcek, M.; Cosnes, J.; Tiret, E.; Bennis, M.; Parc, Y.; Fléjou, J.-F. Expression of epidermal growth factor receptor (EGFR) is frequent in inflammatory bowel disease (IBD)-associated intestinal cancer. Virchows Arch. 2007, 450, 243–244. [Google Scholar] [CrossRef]
  173. Huang, S.; Li, Y.; Zhang, J.; Rong, J.; Ye, S. Epidermal growth factor receptor-containing exosomes induce tumor-specific regulatory t cells. Cancer Investig. 2013, 31, 330–335. [Google Scholar] [CrossRef]
  174. Kasimir-Bauer, S.; Bittner, A.K.; König, L.; Reiter, K.; Keller, T.; Kimmig, R.; Hoffmann, O. Does primary neoadjuvant systemic therapy eradicate minimal residual disease? analysis of disseminated and circulating tumor cells before and after therapy. Breast Cancer Res. 2016, 18, 20. [Google Scholar] [CrossRef] [PubMed]
  175. Ciravolo, V.; Huber, V.; Ghedini, G.C.; Venturelli, E.; Bianchi, F.; Campiglio, M.; Morelli, D.; Villa, A.; Mina, P.D.; Menard, S.; et al. Potential role of HER2-overexpressing exosomes in countering trastuzumab-based therapy. J. Cell. Physiol. 2012, 227, 658–667. [Google Scholar] [CrossRef] [PubMed]
  176. Martinez, V.G.; O’Neill, S.; Salimu, J.; Breslin, S.; Clayton, A.; Crown, J.; O’Driscoll, L. Resistance to HER2-targeted anti-cancer drugs is associated with immune evasion in cancer cells and their derived extracellular vesicles. Oncoimmunology 2017, 6, 1362530. [Google Scholar] [CrossRef]
  177. Westphal, M.; Maire, C.L.; Lamszus, K. EGFR as a target for glioblastoma treatment: An unfulfilled promise. CNS Drugs 2017, 31, 723–735. [Google Scholar] [CrossRef]
  178. Chi, A.S.; Cahill, D.P.; Reardon, D.A.; Wen, P.Y.; Mikkelsen, T.; Peereboom, D.M.; Wong, E.T.; Gerstner, E.R.; Dietrich, J.; Plotkin, S.R.; et al. Exploring predictors of response to dacomitinib in EGFR -amplified recurrent glioblastoma. JCO Precis. Oncol. 2020, 4, 593–613. [Google Scholar] [CrossRef]
  179. Yin, J.; Zeng, A.; Zhang, Z.; Shi, Z.; Yan, W.; You, Y. Exosomal transfer of miR-1238 contributes to temozolomide-resistance in glioblastoma. EBioMedicine 2019, 42, 238–251. [Google Scholar] [CrossRef]
  180. Crow, J.; Atay, S.; Banskota, S.; Artale, B.; Schmitt, S.; Godwin, A.K. Exosomes as mediators of platinum resistance in ovarian cancer. Oncotarget 2017, 8, 11917–11936. [Google Scholar] [CrossRef] [PubMed]
  181. Kugeratski, F.G.; McAndrews, K.M.; Kalluri, R. Multifunctional applications of engineered extracellular vesicles in the treatment of cancer. Endocrinology 2021, 162, bqaa250. [Google Scholar] [CrossRef]
  182. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the international society for extracellular vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef]
  183. Snyder, O.L.; Campbell, A.W.; Christenson, L.K.; Weiss, M.L. Improving reproducibility to meet minimal information for studies of extracellular vesicles 2018 guidelines in nanoparticle tracking analysis. J. Vis. Exp. 2021, 177, e63059. [Google Scholar] [CrossRef]
  184. Witwer, K.W.; Goberdhan, D.C.I.; O’Driscoll, L.; Théry, C.; Welsh, J.A.; Blenkiron, C.; Buzás, E.I.; Di Vizio, D.; Erdbrügger, U.; Falcón-Pérez, J.M.; et al. Updating MISEV: Evolving the minimal requirements for studies of extracellular vesicles. J. Extracell. Vesicles 2021, 10, e12182. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biogenesis of microvesicles (MVs) and exosomes (EXOs) from tumour cells and effects of EV-EGFR on recipient cells. MVs directly arise from the outward budding of the plasma membrane. EXO biogenesis involves invagination of the plasma membrane, internalizing EGFR. The early endosome interacts with the Golgi apparatus and involves complex machinery (endosomal sorting complex required for transport, ESCRT, programmed cell death 6-interacting protein, ALIX, tumour susceptibility gene 101,TSG101) and different tetraspanins (CD9, CD63, and CD81), leading to a second inward invagination and the formation of the intraluminal vesicles (ILVs) contained in multi-vesicular bodies (MVBs). EXOs are then released into the extracellular milieu upon fusion with the plasma membrane.
Figure 1. Biogenesis of microvesicles (MVs) and exosomes (EXOs) from tumour cells and effects of EV-EGFR on recipient cells. MVs directly arise from the outward budding of the plasma membrane. EXO biogenesis involves invagination of the plasma membrane, internalizing EGFR. The early endosome interacts with the Golgi apparatus and involves complex machinery (endosomal sorting complex required for transport, ESCRT, programmed cell death 6-interacting protein, ALIX, tumour susceptibility gene 101,TSG101) and different tetraspanins (CD9, CD63, and CD81), leading to a second inward invagination and the formation of the intraluminal vesicles (ILVs) contained in multi-vesicular bodies (MVBs). EXOs are then released into the extracellular milieu upon fusion with the plasma membrane.
Cancers 15 02970 g001
Table 1. Some of the most commonly found proteins in MVs and EXOs [3,26,33,52,65,95].
Table 1. Some of the most commonly found proteins in MVs and EXOs [3,26,33,52,65,95].
Biological Role MVsEXOs
Membrane Organizers—Adhesion Molecules—Membrane Receptor Tetraspanins: CD9, CD81, CD82Tetraspanins: CD9, CD81, CD63, TSPAN6, TSPAN8, CD151, CD37, CD53,
Integrins, selectinsFlotilin 1 and 2, A33, EpCAM, CD147, Integrin α and β, P-selectin
Biogenesis and SortingARF6, RAB11, ALIX, TSG101, ERK, PLD, VPS4, ESCRT-I, -III, LGALS4Protein kinases, β-catenin, 14-3-3, G proteins, ALIX, TSG101, syntenin, ubiquitin, clathrin, VPS32, VPS4, ESCRT-0-III, LGALS3BP
Cytoskeletal ProteinsActin, tubulin, moesinTubulin, moesin
Cell Type SpecificMHC-I, CD14, HSP70-90MHC-I-II, TFR, WNT, CD86, HSP20-27-60-70-84-90
Cancer Associated and Growth Factorsc-Met, Caveolin-1, EGFR, EpCAM, EMMPRIN, MUC1, FAK, HepPar1 TNF-α, TGF-β, TRAIL, GPC1
A33, cell surface A33 antigen; ALIX, programmed cell death-6 interacting protein; ARF6, ADP-ribosylation factor 6; CD, cluster of differentiation; c-Met, hepatocyte growth factor receptor; EGFR, epidermal growth factor receptor; EMMPRIN, extracellular matrix metalloproteinase inducer; EpCAM, epithelial cell adhesion molecule; ERK, extracellular signal-regulated kinase; ESCRT, endosomal sorting complex required for transport; FAK, focal adhesion kinase; HSP, heat shock protein; LGALS, lectin galactoside-binding soluble; MHC, major histocompatibility complex; MUC1, Mucin short variant S1; PLD, phospholipase-D; RAB, Ras-related protein; TfR, transferrin receptor; TGF, transforming growth factor; TNF, tumour necrosis factor; TRAIL, TNF-related apoptosis-inducing ligand; TSG101, tumour susceptibility gene 101; TSPAN, tetraspanin; VPS4, vacuolar protein sorting.
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

Ferlizza, E.; Romaniello, D.; Borrelli, F.; Pagano, F.; Girone, C.; Gelfo, V.; Kuhre, R.S.; Morselli, A.; Mazzeschi, M.; Sgarzi, M.; et al. Extracellular Vesicles and Epidermal Growth Factor Receptor Activation: Interplay of Drivers in Cancer Progression. Cancers 2023, 15, 2970. https://doi.org/10.3390/cancers15112970

AMA Style

Ferlizza E, Romaniello D, Borrelli F, Pagano F, Girone C, Gelfo V, Kuhre RS, Morselli A, Mazzeschi M, Sgarzi M, et al. Extracellular Vesicles and Epidermal Growth Factor Receptor Activation: Interplay of Drivers in Cancer Progression. Cancers. 2023; 15(11):2970. https://doi.org/10.3390/cancers15112970

Chicago/Turabian Style

Ferlizza, Enea, Donatella Romaniello, Francesco Borrelli, Federica Pagano, Cinzia Girone, Valerio Gelfo, Rikke Sofie Kuhre, Alessandra Morselli, Martina Mazzeschi, Michela Sgarzi, and et al. 2023. "Extracellular Vesicles and Epidermal Growth Factor Receptor Activation: Interplay of Drivers in Cancer Progression" Cancers 15, no. 11: 2970. https://doi.org/10.3390/cancers15112970

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