Next Article in Journal
Pre-Exposure Gene Expression in Baboons with and without Pancytopenia after Radiation Exposure
Next Article in Special Issue
The Role and Molecular Mechanism of Non-Coding RNAs in Pathological Cardiac Remodeling
Previous Article in Journal
Evaluating Complex Mixtures in the Zebrafish Embryo by Reconstituting Field Water Samples: A Metal Pollution Case Study
Previous Article in Special Issue
Current Insights into Long Non-Coding RNAs (LncRNAs) in Prostate Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exosomes: From Garbage Bins to Promising Therapeutic Targets

1
Department of Experimental Therapeutics, The University of Texas MD Anderson Cancer Center, Houston, TX 77030, USA
2
Department of Pharmacology and Toxicology, Faculty of Pharmacy, The University of Al-Azhar, Cairo 11754, Egypt
3
Department of Medical Biology, Faculty of Medicine, The University of Gaziantep, Gaziantep 27310, Turkey
4
Centro de Investigación y Estudios Avanzados del IPN, Unidad Monterrey, Apodaca NL CP 66600, Mexico
5
Center for RNA Interference and Non-Coding RNA, The University of Texas MD Anderson Cancer Center, Houston, TX 77030, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(3), 538; https://doi.org/10.3390/ijms18030538
Submission received: 14 December 2016 / Revised: 25 February 2017 / Accepted: 27 February 2017 / Published: 2 March 2017
(This article belongs to the Collection Regulation by Non-coding RNAs)

Abstract

:
Intercellular communication via cell-released vesicles is a very important process for both normal and tumor cells. Cell communication may involve exosomes, small vesicles of endocytic origin that are released by all types of cells and are found in abundance in body fluids, including blood, saliva, urine, and breast milk. Exosomes have been shown to carry lipids, proteins, mRNAs, non-coding RNAs, and even DNA out of cells. They are more than simply molecular garbage bins, however, in that the molecules they carry can be taken up by other cells. Thus, exosomes transfer biological information to neighboring cells and through this cell-to-cell communication are involved not only in physiological functions such as cell-to-cell communication, but also in the pathogenesis of some diseases, including tumors and neurodegenerative conditions. Our increasing understanding of why cells release exosomes and their role in intercellular communication has revealed the very complex and sophisticated contribution of exosomes to health and disease. The aim of this review is to reveal the emerging roles of exosomes in normal and pathological conditions and describe the controversial biological role of exosomes, as it is now understood, in carcinogenesis. We also summarize what is known about exosome biogenesis, composition, functions, and pathways and discuss the potential clinical applications of exosomes, especially as biomarkers and novel therapeutic agents.

Graphical Abstract

1. Introduction

Exosomes are membrane-derived nanovesicles of about 30–100 nm released by several types of cells, including mast cells, dendritic cells, B lymphocytes, neurons, adipocytes, endothelial cells, and epithelial cells [1]. Notably, tumor cells have been shown to produce and secrete exosomes in greater numbers than normal cells [2]. Exosomes have been found in numerous body fluids, including blood, amniotic fluid, urine, malignant ascites, cerebrospinal fluid, breast milk, saliva, lymph, and bile, under both healthy and morbid conditions [3,4,5].
Exosomes were first observed three decades ago by Pan and Johnstone while studying the maturation process of reticulocytes into erythrocytes. They noted that vesicles, later named “exosomes”, were shed from cultured monolayer cells and retained the transferrin receptor and many membrane-associated proteins [6,7]. Since it was believed that these vesicles were simply removing unnecessary proteins and other molecules from the releasing cells, exosomes were first thought to function only as cellular garbage disposals [8].
It was not until the mid-1990s that exosomes were shown to have an immunological function [9]. Since then, numerous studies have identified exosomes as a means of intercellular communication that play a role in normal physiological or biologically important processes, such as lactation, inflammation, cell proliferation, immune response, and neuronal function [10,11,12]. They are implicated as well in the pathogenesis of thrombosis, diabetes, and atherosclerosis, and also in the development and progression of diseases such as liver disease, neurodegenerative diseases [13,14,15], and, recently, cancer [16].

2. Biogenesis

Although details of the underlying mechanism remain incompletely defined, several processes have recently been shown to have a regulatory role in exosome biogenesis [17].
Exosomes are considered a distinct vesicle population that differs from microvesicles by size. Exosomes are defined as vesicles in the range of 30–100 nm, while microvesicles are defined as vesicles in the range of 100–1000 nm. Despite this clear distinction, however, the terms “exosome” and “microvesicle” have been used interchangeably in many published reports [18].

2.1. Formation

In general, exosome biogenesis consists of two steps, the inward budding of membranous vesicles of endosomes and their release into a structure known as a multivesicular body (MVB). The formation of MVBs occurs during the maturation of early endosomes into late endosomes with the accumulation of intraluminal vesicles [19]. After maturation, MVBs are directed for fusion with either the lysosome, where their cargo will undergo lysosomal degradation, or the plasma membrane, where their contents will be released into the extracellular space (Figure 1). When MVBs undergo this process, transmembrane proteins are incorporated into the invaginating membrane, maintaining a topological orientation similar to that of the plasma membrane [1,20].

2.2. Composition

The composition of exosomes differs from cell type to cell type. According to the most recent version of the exosome content database, Exocarta (Version 4), exosomes from various organisms and various cell types have been characterized as containing 4563 proteins, 194 lipids, 1639 mRNAs, and 764 miRNAs [21]. The protein content largely depends on the exosome’s cellular origin and is generally enriched for certain molecules, including targeting and fusion proteins (e.g., tetraspanins, lactadherin, and intergrins), cytoplasmic enzymes (e.g., GAPDH, peroxidases, pyruvate kinases, and lactate dehydrogenase), chaperones (e.g., heat shock proteins Hsp60, Hsp70, Hsp90, and the small HSPs), membrane trafficking proteins (e.g., Rab proteins, ARF GTPases, and annexins), proteins involved in MVB formation (e.g., ALIX, TSG101, and clathrin), cytoskeletal proteins (e.g., actin and tubulin), signal transduction proteins (e.g., protein kinases and heterotrimeric G proteins) (Figure 2) [22].
Exosome-specific protein conformation may be subject to the cell type or tissue birthplace from which it originates and may differ according to the physiological changes and stimulation that the cell underwent. For example, antigen-presenting cell-derived exosomes are enriched in antigen-presenting molecules, including major histocompatibility class (MHC)-I and -II complexes, as well as co-stimulatory molecules [23]. Tumor-derived exosomes usually contain tumor antigens in addition to certain immunosuppressive proteins such as FasL, TRAIL, or TGF-β [24]. Very relevant is the fact that exosomes also contain proteins involved in cell signaling pathways, such as the Notch ligand Δ-like 4 [25], Wnt-β-catenin signaling proteins [26], and some proteins involved in intercellular cell signaling, such as interleukins [27].
The main components of exosomes are lipids. Exosomes are enriched in cholesterol, diglycerides, glycerophospholipids, phospholipids, and sphingolipids or glycosylceramides (including sphingomyelin and ceramide) [28]. Besides these lipids, bioactive lipids, such as prostaglandins and leukotrienes, and enzymes activated in lipid metabolism, such as phospholipase C, are also found in exosomes [28,29]. In this way, exosomes function as lipid carriers, allowing the transport of the bioactive lipids they carry to a recipient cell [30]. Excitingly, exosomal content such as the fatty acid docosahexaenoic acid and lysophosphatidylcholine can enhance dendritic cells’ antigenic capacity [31]. In contrast, the exosomes that contain high levels of prostaglandin PGE2 are involved in tumor immune evasion and the promotion of tumor growth [30].
In addition to proteins and lipids, exosomes also contain functional RNA molecules, including mRNAs and other non-coding RNAs such as miRNAs and lncRNAs [32,33]. These exosomal RNAs, in particular the miRNAs, have been shown to function in the recipient cells [34,35]. Even though the effects of other exosomal loads on receiver cells cannot be ignored, miRNAs are important players in the key functions in this process. Sometimes, exosomal pathways can eliminate tumor-suppressor miRNAs that block metastatic progression [36,37]. Nevertheless, a very large body of evidence in the literature clearly indicates the tumor-promoting role of exosomal miRNAs [38,39,40].
Numerous studies have revealed that the amount of RNA contained in exosomes differs significantly from the amount of RNA present in the parental cell and that the exosomal RNA apparently lacks ribosomal RNA [41,42]. Curiously, however, exosomal RNA content in cancer patients is comparable to that in the original tumor; thus the scientific community has become interested in the potential of the exosomal miRNA profile as a diagnostic tool for cancer [3,43]. Notably, large quantities of miRNAs have been detected in tumors but are not present or are present at very low levels in the exosomes released by the parental cells [44,45]. These results indicate that some miRNAs might be preferentially directed for secretion. However, the processing involved in the selection, packaging, and release of these exosomal miRNAs is not understood, and whether these miRNAs could be used as reliable markers of disease is still under consideration. Collectively, the structure of exosomes is recognized to carry a variety of important proteins, lipids, and genetic materials, and cohesively they interact to guide intercellular communications in healthy and disease states [46].

2.3. Cargo Sorting

The mechanisms underlying sorting of proteins and lipids into exosomes are largely unknown, although some of the potential mechanisms that have been suggested involve heteromeric protein complexes (e.g., endosomal sorting complex required for transport (ESCRT)) and also associated proteins such as programmed cell death 6 interacting protein (also recognized as ALIX) and tumor susceptibility gene 101 protein (TSG101) [47,48]. ESCRT proteins, including ESCRT-I, ESCRT-II, and ESCRT-III, are need for cargo selection and the inward budding process (away from the cytoplasm). Some of the components of ESCRT, such as vacuolar protein sorting protein 31 (VPS31), vacuolar protein sorting protein 4B (VPS4B), and TSG101, have been found in endosome-like plasma membrane domains that generate exosomes [49]. Tumor cell exosomes have been shown to contain syndecan, syntenin, and ALIX. Down-modulation of any of these proteins reduced exosomal release, and production of syndecan-, syntenin-, and ALIX-containing exosomes was dependent on the normal functioning of the ESCRT machinery proteins [50].
The secretion of syntenin into exosomes is driven by syndecan, and this process induced heparin sulfate clustering. Overexpression of the enzyme heparanase cleaves the heparan sulfate, causing a noticeable increase in the secretion of exosomes. Moreover, heparanase has also been shown to alter exosome protein composition, which was demonstrated by increase of the levels of syndecan-1, Vascular endothelial growth factor (VEGF), and hepatocyte growth factor (HGF) [51]. The small integral membrane protein of the lysosome/late endosome has been shown to be secreted excessively in exosomes, and its overexpression increases exosome release and exosomal accumulation of ALIX and CD63 [52].
In addition to ESCRT, which recognizes ubiquitylated proteins, other ESCRT-independent mechanisms operate to generate exosomes [53]. These unconventional ESCRT-independent pathways seem to be driven by the presence of certain lipids, such as ceramides and lysobisphosphatidic acid [54,55]. Lipid-metabolizing enzymes, including sphingomyelinase, the enzyme that hydrolyzes sphingomyelin into ceramide, and phospholipase D, which hydrolyzes phosphatidylcholine to generate choline and phosphatidic acid, were shown to regulate exosome secretion [56,57]. Another sphingomyelin metabolite, sphyngosine-1-phosphate (S1P), was shown to play a key role in exosome biogenesis. Silencing of S1P1 receptors impairs the formation of CD63-, CD81-, or flotillin-positive exosomes [58].
Finally, ATP-binding cassette transporter A3, which works like a transporter for phosphatidylcholines, has a role in exosome production [59]. Remarkably, the ESCRT-independent sphingomyelinase pathway produces exosomes enriched in tetraspanins, proteins that contain transmembrane domains that may also be involved in endosomal sorting pathways [60]. Moreover, tetraspanins CD9, CD63, and CD81 are involved in exosome biogenesis and protein loading [61,62]. The signals that control the switch between the two mechanisms remain unknown.
As described, the molecular mechanisms that regulate the loading of proteins into exosomes have been extensively studied. However, the mechanisms by which RNA molecules are sorted into exosomes have remained unknown until recently. It has been shown that specific sequence motifs, such as GGAG present in microRNAs (miRNAs), are involved in this sorting and regulate the localization of miRNA molecules into exosomes [63]. The heterogeneous nuclear ribonucleoprotein A2B1 (hnRNPA2B1) has been shown to bind specifically to exosomal miRNAs through the recognition of GGAG motifs and to control their loading into exosomes. Furthermore, it has been reported that the hnRNPA2B1 loaded into exosomes is sumoylated and that this sumoylation controls the binding of hnRNPA2B1 to miRNAs [64]. In another study, it was found that the RNA-binding protein Y-box protein I (YBX1) binds to miR-223 and is also required for the sorting of this miRNA in the cell-free reaction. Furthermore, YBX1 plays an important role in the secretion of miRNAs in exosomes by HEK293T cells [65]. Kosaka and colleagues have shown that inhibition of sphingomyelinase expression reduced the number of exosomal miRNAs [66]. It has been recognized recently that there is a possible correlation between AGO2, a miRNA-induced silencing complex protein, and exosomal miRNA sorting [67,68]. Silencing of AGO2 has been shown to decrease the abundance of the preferentially exported miRNAs in exosomes [69].
Several reports have shown that exosomes act as transport vesicles for functional long non-coding RNAs (lncRNAs) such as TUC339, ROR, MALAT1, HOTAIR, and GAS5, which may induce cancer-like phenotypes and increase chemoresistance within the recipient cells [33,70,71]. The mechanism for loading lncRNAs into exosomes is currently unknown. Recent observations suggested that specific RNA-binding proteins such as ELAVL1 may play an important role in directing lncRNAs for exosomal transport [72].

2.4. Release

The release of exosomes into the extracellular environment requires the transport and docking of MVBs as well as their fusion with the plasma membrane [73]. Many proteins have been implicated in the secretion of exosomes, but the precise mechanism of vesicle release remains elusive and is likely to vary among different cells. It has been proposed that exosome release is a Ca2+-dependent [74] and pH-dependent [75] process.
In some tumor cells, however, exosome release depends on the Rab GTPase family, whose members, such as RAB11, RAB27A, and RAB31, are important regulators of membrane trafficking [76,77,78]. Cells with mutant forms of these proteins release fewer exosomes [79]. Rab family member proteins also have been reported to play a role in exosome secretion, specifically Rab35, which regulates exosome secretion by interacting with the GTPase-activating protein TBC1 domain family member 10A–C [80].
The transcription factor p53 has been shown to be involved in exosome release. Activation of p53 through irradiation resulted in the release of greater numbers of exosomes [81]. A p53-regulated gene, TSAP6, was shown to enhance exosome production in cells undergoing a p53 response to stress [82]. Moreover, Lespagnol and colleagues provided direct evidence that exosome production is severely compromised in TSAP6-null mice [83].
Other studies have described different mechanisms of exosome secretion that involve the soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) protein YKT6 [26]. Briefly, the cytoskeleton and the contractile machinery of the cell move, attracting the opposing membranes with the assistance of the SNARE complex before pinching off the membrane connection and releasing the vesicle into the extracellular space [84,85]. Other proteins involved in SNARE disassembly are vSNARE VAMP7 and ATPase N-ethylmaleimide-sensitive factor, which have been reported to stimulate exocytosis of acetylcholinesterase-containing exosomes in the K562 human leukemia cell line [86].

2.5. Uptake

It is still controversial whether exosome uptake is cell type-specific [87] and whether it involves membrane fusion or endocytosis [75,88]. Moreover, exosome uptake may be clathrin-dependent or clathrin-independent [73].
Exosome uptake has been shown to occur via clathrin-mediated endocytosis [89], lipid raft-mediated endocytosis [90], heparin sulfate proteoglycans-dependent endocytosis [91], or phagocytosis [87]. Alternatively, exosomes could be internalized by direct fusion with the plasma membrane [75] or through binding to the surface of a recipient cell through exosomal adhesion molecules phosphatidylserine/lysophosphatidylcholine, and cellular receptors (e.g., LFA1, TIM1, and TIM4) [6].

3. Diversity in Exosome Function

There is no doubt that exosomes are involved in many physiological functions and processes, both normal and pathological. Originally, exosomes were described as a mechanism for elimination of excessive proteins or undesirable molecules from the cell [6]. It has been shown that exosomes are secreted to discard membrane proteins, such as transferrin receptors, that have become useless in mature red blood cells. Thus, exosomes were long considered a process whereby cells get rid of undesirable proteins and molecules, making the exosomes a compartment for cellular garbage transport and disposal [6,8]. In the last decade, the exosomes’ role as mediators of cellular communication has emerged, and we now have evidence revealing that exosomes control both normal physiological processes, such as immune response and lactation [10], and the expansion and progression of diseases, such as neurodegenerative diseases [15,92] and especially cancer [16]. Exosomes carry out a diverse range of functions and sometimes have opposing effects on the recipient cells depending on their tissue of origin and molecular content [93,94,95].
Here, we discuss in detail what is known about the functions of exosomes in normal and pathological conditions.

3.1. Bioactive Roles of Exosomes in Maintenance of Normal Physiology

Exosomes participate in the maintenance of normal physiology, for example, stem cell maintenance and tissue repair [96]. Exosomes have been implicated as morphogen transporters during development and differentiation. They are released by donor cells and spread through the adjacent tissue at different concentrations, enabling cell–cell communication [97,98]. Importantly, several reports have implicated exosomes in stem cell maintenance and plasticity, indicating that stem cell-derived exosomes have a pivotal role in tissue regeneration following injury [99,100]. Exosomes have also been implicated in cell phenotype modulation and tissue regeneration; for example, exosomes derived from hepatic stem cells can promote hepatocyte regeneration [101]. Exosomes also are involved in converting the hematopoietic stem cell phenotype into a liver cell phenotype [102] and in shifting the bone marrow cell transcriptome toward a lung phenotype in vivo [103].
Exosomes also have a role in tissue homeostasis, as in wound healing. For example, a recent study showed that, after injury, epithelial cells increased the number of exosomes transferring TGFβ1 mRNA, stimulating fibroblast differentiation through the repair and renewal of tissues subsequent to parenchymal damages [104].
Exosomes display a wide variety of immunomodulatory properties. To sustain strong immunostimulatory activity between mature dendritic cells and B lymphocytes that bind tightly to follicular dendritic cells and whose function is presenting antigen-MHC-II complexes to T lymphocytes, exosome release is necessary to maintain this communication [105]. Furthermore, the effects of immune activation can be mediated by exosome-promoted proliferation and survival of hematopoietic stem cells and activation of natural killer cells [31].
Exosomes have been found to have anti-inflammatory functions. Exosomes released from dendritic cells overexpressing IL-4 or IL-10 suppressed delayed-type hypersensitivity reactions in an MHC-II-dependent manner in a mouse model. These exosomes also suppressed the onset and reduced the severity of collagen-induced arthritis [106,107]. Moreover, FasL on the exosomes was also shown to be important for the suppression of delayed-type hypersensitivity reactions [107]. Dendritic cells treated with IL-4 and IL-10 have shown promise in the treatment of inflammatory and autoimmune diseases [108,109]. Plasma exosomes of mice immunized to a specific antigen were shown to have anti-inflammatory functions in the delayed-type hypersensitivity reaction model similar to those of dendritic cell exosomes, suggesting their relevance in vivo [110]. Exosomes may also have a beneficial role in sepsis, through increased phagocytosis of apoptotic cells [111].
Centrally, in addition to classical synaptic neurotransmission, neurons communicate via the secretion of exosomes and exosome-like vesicles that can contribute to a range of neurobiological functions, including synaptic plasticity [112].
Despite the importance of these findings, a better characterization of exosomes and understanding of their effects are needed if we are to further improve their application in the fields of regenerative medicine and immunotherapy. Most studies in these areas were conducted with non-physiological concentrations of exosomes, whereas in vivo investigations of exosome-induced mechanisms are hampered by the lack of insight into their biogenesis.

3.2. Pathological Roles of Exosomes in Spreading of Disease

The best understood role of exosomes in disease is their role in tumor biology (Figure 3). One of the hallmarks of cancer cells is that they react with their microenvironment; they can communicate and exchange information by secreted growth factors, cytokines, chemokines, and small molecular mediators (e.g., nucleotides) [113,114].
As very crucial cell-to-cell messenger mediators of communication, exosomes could be notably affecting a recipient cell if they transfer as cargo a specific molecule such as mRNA or non-coding RNA that can alter the gene expression or production of proteins in the recipient cell.
In the following paragraphs, we discuss individually the roles of exosomes in diverse mechanisms, such as metastasis, angiogenesis, hypoxia, and immune escape, which collectively support tumor progression.

3.2.1. Invasion, Metastasis, and Angiogenesis

Because exosomes carry genomic and proteomic materials known to mediate these hallmarks of cancer [32], it has been hypothesized that exosomes secreted by tumor cells have a role in the growth and spread of tumor cells. Indeed, many studies have demonstrated such potential in tumor-derived exosomes. For example, McCready and colleagues demonstrated that Hsp90α-containing exosomes isolated from an invasive cancer cell line could enhance cell migration via activation of plasmin, but the effect was abrogated if an anti-Hsp90 antibody was added to the exosomes [115]. Furthermore, tumor-derived exosomes have been described as having the capacity to establish a pre-metastatic niche with generation of a suitable microenvironment in distant and specific metastatic sites [116,117].
Proteomic analysis of exosomes secreted by human mesothelioma detected the presence of strong angiogenic factors that can increase angiogenesis and vessel density in the neighborhood of tumor cells [118]. In their study on melanoma-derived exosomes, Hood et al. [119] described the pro-angiogenic potential of such nanovesicles, which rapidly stimulated endothelial signaling, important for tissue matrices remodeling and endothelial angiogenesis. The same group reported later that the exosomes of the melanoma cell home to sentinel lymph nodes, which enforces coordinated molecular signals that recruit melanoma cells, inducing extracellular matrix deposition and angiogenesis in the lymph nodes [117]. Consistent with these observations, it has been shown that exosomes from highly metastatic melanoma donor cells augmented the metastatic conduct of primary tumors by continuously “teaching” bone marrow progenitors through the receptor tyrosine kinase MET [120].
One of the major factors that can be involved in the concession of pro-angiogenic activity to tumor exosomes is exemplified by tetraspanins, which are constitutively augmented in exosomes and have been found to contribute to exosome-mediated angiogenesis [121]. The same group that reported that finding later demonstrated that tumor exosomes are directed to non-transformed cells in pre-metastatic niches and organs. This modulates pre-metastatic organ cells, predominantly through transferred miRNA; thus miRNA from a metastasizing tumor arranges or prepares pre-metastatic organ stroma cells for hosting tumor cells [122].
Exosomes derived from a pancreatic tumor cell line overexpressing tetraspanin 8 (Tspan8 or D6.1A) are able to promote tumor growth by their capacity to induce angiogenesis both in vitro and in vivo [123]. Furthermore, exosomal Tspan8 contributes to the selective recruitment of proteins and mRNA into exosomes, including CD106 and CD49d; these two receptors are involved in the binding and internalization of exosomes by endothelial cells. Once the exosomes are internalized, induction of several angiogenesis-related genes, including VEGF and VEGFR2, is observed in combination with enhanced endothelial cell proliferation and migration and maturation of endothelial cell progenitors [124]. There is also evidence that exosomes from cancerous cells, where the Notch ligand Δ-like 4 was incorporated and where the Δ-like 4 protein was transferred into the cell membrane of host endothelial cells, result in inhibition of Notch signaling and the switch of the endothelial cell phenotype toward tip cells phenotype. The reverted phenotype appears to have a crucial role in vascular development and angiogenesis [25].

3.2.2. Hypoxia

Tumor hypoxia has emerged as a key factor in tumor progression and is associated with poor prognosis and chemoresistance [125]. During hypoxia, exosomes are secreted by tumor cells with increases in angiogenic factors and metastatic potential; this suggests that tumor cells are able to adapt to a hypoxic microenvironment by the secretion of exosomes to promote angiogenesis or facilitate metastasis to a more appropriate microenvironment [126].
There is evidence in several studies involving various cancer models for enhanced exosome release under hypoxic conditions. Borges and colleagues showed in a kidney model that TGF-β1–containing exosomes released by injured epithelial cells can mediate tissue regenerative responses and activation of fibroblasts. These findings strongly suggest the utility of exosome-targeted therapies to control tissue fibrosis [104]. Consistent with this finding was the observation that exosomes derived from hypoxic leukemia cells enhance angiogenic activity in endothelial cells [127]. Similarly, in a highly malignant squamous cell carcinoma model, hypoxic tumor cells modulate their microenvironment and facilitate angiogenesis and metastasis through exosomal secretion of certain proteins [126].
Kucharzewska and colleagues showed in a highly malignant glioblastoma model that exosomes reflect the hypoxic status of glioma cells and mediate hypoxia-dependent activation of vascular cells during tumor development [128]. King and colleagues reported that breast cancer cells grown under hypoxic conditions release more exosomes into their microenvironment via activation of HIF-1α to promote their own survival and invasion [129]. Another study showed that exosomal miR-135b shed from hypoxic multiple myeloma cells enhanced angiogenesis by targeting factor-inhibiting HIF-1 [39].

3.2.3. Cancer Exosomes and Immune Modulation

Tumor-derived exosomes have been reported both to stimulate and to suppress immune response. A significant collection of studies has demonstrated that exosomes can transport antigens such as MHC-I and MHC-II and carcinoembryonic antigen (CEA) from tumor cells to antigen-presenting dendritic cells [130,131,132,133]. The primary dendritic cells, cytotoxic T lymphocytes, induce an immune antitumor response and allow inhibition of tumor growth through MHC-I molecules [130,134]. Similarly, in an ex vivo human model system, dendritic cells pulsed with exosomes derived from malignant effusions proved an effective source of tumor antigens for cross-presentation to CD8+ cytotoxic T cells [131]. Moreover, exosomes obtained by stimulating dendritic cells may sensitize adjacent dendritic cells, thereby inducing the immune response [135]. Moreover, surface Hsp70-positive tumor-derived exosomes stimulate natural killer cell activity [136,137]. As result, natural killer cells initiate apoptosis in tumors through granzyme B [136,137].
Despite this, however, an alternative view suggests that exosomes have immunosuppressive effects and assist cancers in immune evasion. For example, tumor-derived exosomes express death ligands such as FasL and TRAIL or high amounts of galectin-9, which can promote T cell apoptosis [138,139,140]. Chalmin and colleagues showed that tumor-derived exosomes activate myeloid-derived suppressor cells and exert TGF-β1–mediated suppressive activity on T cells [141]. In addition, tumor-derived exosomes were shown to promote tumor growth by suppressing natural killer cell function [142,143,144]. Tumor-derived exosomes can also support and expand the immunosuppressive function of regulatory T cells [145,146]. Furthermore, tumor-derived exosomes can block the maturation of dendritic cells and macrophages in vivo and in vitro [147]. Nonetheless, the evidence that exosomes mediate increases or decreases of immunoregulatory functions and proposals that exosomes be administered as immunotherapy must be carefully inspected before translation to further clinical applications.
Growing evidence links tumor metastasis with chronic inflammatory processes and dysregulated activity of various immune cells [148]. Chow and colleagues demonstrate that breast cancer-derived exosomes trigger NF-κB signaling and promote inflammatory cytokine production through Toll-like receptors on macrophages [149]. A similar effect was observed with exosomes derived from malignant ascites of ovarian cancer patients [150]. Another important study showed that miRNAs in cancer-released exosomes can bind as ligands to Toll-like receptors and induce pro-metastatic inflammatory responses [151].

4. Exosome-Based Diagnostics and Therapeutics

4.1. Exosomes as Therapeutic Target

Given that exosome levels are often elevated in correlation with greater severity of different types of cancer [4,152,153], one therapeutic strategy would involve reducing circulating exosomes to normal levels to prevent poor outcomes. With this perspective, many ongoing studies are designed to modulate exosome production either by acting on processes regulating their formation and/or release or by inhibiting their interaction with target cells through specific targeting of their components (Figure 4) [154].

4.1.1. Inhibition of Exosome Formation

Various cellular components are known to be crucial for the formation of exosomes. For example, components of ESCRT are known to be involved in formation of MVBs and intraluminal vesicles [155]. ESCRTs are composed of approximately 30 proteins that assemble into four complexes (ESCRT-0, -I, -II and -III) with associated proteins (VPS4, VTA1, and ALIX) [19]. ESCRT-0, -I, -II, and -III are conserved from yeast to mammals [19]. Several studies have linked the ESCRT-0 protein hepatocyte growth factor-regulated tyrosine kinase substrate (HGS, also known as HRS) to exosome secretion by showing reduced exosome release in HRS-depleted dendritic cells, HEK293 cells [26,156], and tumor cells [157]. Tumor cell exosomes have been shown to contain syndecan, syntenin, and ALIX; overexpression of syntenin induced increases in the ALIX-dependent release of exosomes [50,158].
ESCRT-independent mechanisms of exosome formation have also been described. These mechanisms involve the sphingolipid ceramide or tetraspanins. Small-molecule inhibitors of sphingomyelinase, the enzyme generating ceramide from sphingomyelin, or amiloride attenuate endosomal sorting and exosome production, thereby leading to reduction in tumor growth [56,159].
Furthermore, tetraspanins are enriched in the internal vesicles of MVBs and in exosomes [160]. Expression of tetraspanin Tspan8 could modify both the mRNA content and the protein composition of exosomes secreted by rat pancreatic adenocarcinoma cells [124].
Alternatively, the formation of exosomes may be controlled by specific signaling pathways triggered by Ras homolog family member A [161] or ADP-ribosylation factor 6 [162]. Targeting these pathways may have direct therapeutic significance.

4.1.2. Inhibition of Exosome Release

Several proteins, often including small GTPases of the Rab family, have been implicated in the secretion of exosomes. Rab27a and Rab27b, as well as their effector proteins, are important regulators of exosome release [77,163]. Interestingly, silencing Rab27a by RNA interference disrupted exosome-dependent and -independent mechanisms that modify the tumor microenvironment and can also reduce tumor growth and metastasis [164]. Other Rab proteins such as Rab11 and Rab35 might serve as alternative targets for impairing the release of exosomes by inhibiting the docking of MVBs with the plasma membrane [80,165]. There is also evidence of the involvement of lipids in the regulation of exosome secretion. Downregulation of the diacylglycerol kinase α protein resulted in inhibition of the secretion of Fas ligand-containing exosomes [166].
The final step of exosome release requires the fusion of MVBs with the plasma membrane. This process is mediated by a membrane-bridging SNARE complex machinery that might include VAMP7 [86]. The R-SNARE protein Ykt6 was reported to be involved in the secretion of exosomes carrying the morphogen Wnt from HEK293 cells [26], but further studies are required to confirm details.
Increased intracellular Ca2+ stimulates exosomes secretion, and dimethyl amiloride, an inhibitor of voltage-gated Ca2+ channels, decreases stimulated exosome release [74,141]. Finally, cellular microenvironmental pH plays an important role in exosome secretion, as its modification using proton pump inhibitors significantly suppressed exosome secretion [167]. Still, further research regarding inhibition of exosome release is missing. Since exosomes are implicated in intercellular communications, and maintaining normal cellular physiology this represent the most important limitations to be used as therapeutic strategy, due to the potential toxicity and other side-effects.

4.1.3. Inhibition of Exosome Uptake

Cells appear to take up exosomes by a variety of endocytic pathways, including clathrin-dependent endocytosis and clathrin-independent pathways such as macropinocytosis and phagocytosis [87,89,168]. It has been shown that treatment of exosomes with proteinase K significantly reduced uptake by ovarian cancer cells. These results indicate that surface proteins on exosomes may serve as receptors for uptake [20]. The uptake of tumor-derived exosomes seems to be mediated by surface phosphatidylserine, which can be blocked with diannexin [169]. Heparan sulfate proteoglycans have been suggested to function as internalizing receptors of cancer cell-derived exosomes. This uptake pathway seems to be important, because treatment with heparin significantly inhibited exosome-mediated stimulation of cancer cell migration [91]. Heparin also inhibited oncogenic EGFRvIII mRNA transfer by interfering with the binding of extracellular vesicles into recipient cells [170]. Moreover, exosome internalization could be inhibited by the knockdown of dynamin2, which is necessary for clathrin- and caveolin-dependent endocytosis [87].
Together, these data support the hypothesis that inhibition of exosome biogenesis, release, or uptake mechanisms may have beneficial effects in the treatment of cancer.

4.2. Exosome Removal as a Therapeutic Adjuvant in Cancer

Removal of exosomes from the entire circulatory system has been proposed as a novel strategy to treat cancer. Using an affinity plasmapheresis platform, the biotechnology company Aethlon Medical Inc. (San Diego, CA, USA) has established an adjunct therapeutic approach that decreases systemic secretion of HER2-positive exosomes by tumors and inhibits progression of HER2-positive breast tumors [171].

4.3. Exosomes as Cancer Immunotherapy

Tumor-derived exosomes carry antigens and have been used as a source of specific stimulus for the immune response against tumors [130]. In contrast, dendritic cell-derived exosomes have the structure necessary to induce very potent antigen-specific immune responses [172]. Previous studies showed that both tumor-derived and dendritic cell-derived exosomes stimulate tumor antigen-specific CD8+ cytotoxic T lymphocyte responses; moreover, these exosomes can induce antitumor immunity in experimental animal models and human clinical trials for colorectal, metastatic skin, and non-small cell lung cancers [173,174]. In mice, tumor-derived exosomes can serve as an antigen delivery system and can prevent autologous tumor development in a CD4+ and CD8+ T cell-dependent manner [130]. In phase I clinical trials, patients with advanced non-small cell lung cancer or metastatic melanoma vaccinated with dendritic cell-derived exosomes exhibited antitumor immune responses and tumor regression [175,176]. Recent evidence has shown that tumor exosome-loaded dendritic cells efficiently produce an antitumor reaction against autologous tumor cells in patients with malignant glioma [177].
Ascites-derived exosomes are believed to be as efficient as those derived from dendritic cells or tumors in sensitizing dendritic cells and prime cytotoxic T lymphocytes, which kill autologous tumor cells in vitro [131]. Furthermore, the exosomes derived from malignant effusions of ovarian cancer patients have been prepared and are ready for clinical trials [178,179]. Moreover, Dai and colleagues found that ascites-derived exosomes from patients with colorectal cancer in combination with granulocyte-macrophage colony-stimulating factor can efficiently induce potent CEA-specific antitumor immunity in patients with advanced colorectal cancer (Figure 4A) [180].

4.4. Exosomes as Cancer Diagnostic and Prognostic Markers

Several lines of evidence show that exosomes are present in many biologic body fluids; exosomes might therefore be considered accessible diagnostic biomarkers that hold great potential for detection of many pathological conditions, including cancer [181].
Over the past few years, several studies have addressed the potential use of exosomes as biomarkers in cancer, as their level in circulating blood correlates with prognosis [43,182,183]. Furthermore, miRNAs loaded into exosomes have been suggested as diagnostic and prognostic indicators for ovarian cancer, lung cancer, colon cancer, prostate cancer, and breast cancer [3,184,185,186,187,188]. In addition to miRNAs, exosomal lncRNAs from cancer patients have been defined as novel tumor biomarkers (Figure 4B) [189,190].

4.5. Exosomes as a Drug Delivery System

Exosomes offer distinct advantages as gene therapy delivery vectors as they comprise cellular membranes with multiple adhesive proteins on their surface [191]. Furthermore, their small size and flexibility enables them to cross major biological barriers such as the blood–brain barrier [192]. In contrast to established liposome formulations, exosomes are naturally occurring secreted membrane vesicles with lower toxicity; that they are very well tolerated in the body can be inferred from their ubiquitous presence in biological fluids. Their potential utility in drug delivery is implied by their intrinsic homing capacity [193]. For instance, exosomes derived from melanoma home preferentially to sentinel lymph nodes [117], this homing capacity can be used as targeted delivery system for drugs.
Exosomes are naturally adapted for the transport and intracellular delivery of proteins, mRNAs, miRNAs, various non-coding RNAs, mitochondrial DNA, and genomic DNA [32,194,195]. This makes them a valuable tool for the therapeutic delivery of siRNAs [196,197], miRNAs [198,199], and shRNAs [200]. In addition to transfer of interfering RNAs, other types of therapeutic cargo such as lipophilic small molecules can be loaded into these particles. Anti-inflammatory agent curcumin [201,202] and anticancer agents doxorubicin [203] and paclitaxel [204] have been loaded into exosomes or exosome-like vesicles. Exosome-based drug delivery systems may have precedence in the treatment of cancer owing to their endogenous origin, which minimizes their immunogenicity and toxicity. For instance, some studies have shown that the efficacy of doxorubicin loaded into exosomes was greatly enhanced over that delivered by other delivery systems and caused significantly fewer adverse effects on major organ systems, especially the heart, implying that delivery via exosomes might decrease the major downside of this chemotherapeutic drug (Figure 4A) [203,205].

5. Conclusions and Future Perspectives

Recent analyses of the composition and biogenesis of exosomes indicate that tumor cells secrete exosomes that can both block tumor growth by promoting antitumor immune responses and induce tumor growth by attenuating antitumor immunity or promoting angiogenesis and/or metastases to distant tissues or organs. The past decade has witnessed a renewed research interest in exosomes, mostly as a result of the demonstration of their immuno-stimulating effects in vivo. The propensity for these controversial effects is contingent upon the type and state of the host cells, the type and state of the recipient cells, and the microenvironment in which these interactions take place. Even though these studies have prompted the clinical application of exosomes, they have not addressed the mechanisms of biogenesis or cargo sorting or established the physiological relevance of the exosome payload. Ongoing advances in the analysis of the formation of multivesicular compartments will probably unravel the mechanisms of exosome generation, which will allow deepening of our understanding of the exact characteristics of exosomes and their functional role in cancer pathogenesis. Uncovering the physiological role of the entirely new mode of cell–cell communication mediated by exosomes may provide us the tools to further improve anticancer therapeutics and cancer diagnostics.

Acknowledgments

We thank Kathryn Hale, scientific editor, Department of Scientific Publications, The University of Texas MD Anderson Cancer Center, for critical reading and editing of the manuscript; and Leticia Evans, Program Coordinator, Department of Experimental Therapeutics, The University of Texas MD Anderson Cancer Center, for reading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thery, C.; Zitvogel, L.; Amigorena, S. Exosomes: Composition, biogenesis and function. Nat. Rev. Immunol. 2002, 2, 569–579. [Google Scholar] [PubMed]
  2. Jenjaroenpun, P.; Kremenska, Y.; Nair, V.M.; Kremenskoy, M.; Joseph, B.; Kurochkin, I.V. Characterization of RNA in exosomes secreted by human breast cancer cell lines using next-generation sequencing. PeerJ 2013, 1, e201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Taylor, D.D.; Gercel-Taylor, C. MicroRNA signatures of tumor-derived exosomes as diagnostic biomarkers of ovarian cancer. Gynecol. Oncol. 2008, 110, 13–21. [Google Scholar] [CrossRef] [PubMed]
  4. Simpson, R.J.; Lim, J.W.; Moritz, R.L.; Mathivanan, S. Exosomes: Proteomic insights and diagnostic potential. Expert Rev. Proteom. 2009, 6, 267–283. [Google Scholar] [CrossRef] [PubMed]
  5. Gallo, A.; Tandon, M.; Alevizos, I.; Illei, G.G. The majority of microRNAs detectable in serum and saliva is concentrated in exosomes. PLoS ONE 2012, 7, e30679. [Google Scholar] [CrossRef] [PubMed]
  6. Pan, B.T.; Teng, K.; Wu, C.; Adam, M.; Johnstone, R.M. Electron microscopic evidence for externalization of the transferrin receptor in vesicular form in sheep reticulocytes. J. Cell Biol. 1985, 101, 942–948. [Google Scholar] [CrossRef] [PubMed]
  7. 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] [PubMed]
  8. Johnstone, R.M. The jeanne manery-fisher memorial lecture 1991. Maturation of reticulocytes: Formation of exosomes as a mechanism for shedding membrane proteins. Biochem. Cell Biol. 1992, 70, 179–190. [Google Scholar] [CrossRef] [PubMed]
  9. 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] [PubMed]
  10. Admyre, C.; Johansson, S.M.; Qazi, K.R.; Filen, J.J.; Lahesmaa, R.; Norman, M.; Neve, E.P.; Scheynius, A.; Gabrielsson, S. Exosomes with immune modulatory features are present in human breast milk. J. Immunol. 2007, 179, 1969–1978. [Google Scholar] [CrossRef] [PubMed]
  11. Simhadri, V.R.; Reiners, K.S.; Hansen, H.P.; Topolar, D.; Simhadri, V.L.; Nohroudi, K.; Kufer, T.A.; Engert, A.; Pogge von Strandmann, E. Dendritic cells release HLA-B-associated transcript-3 positive exosomes to regulate natural killer function. PLoS ONE 2008, 3, e3377. [Google Scholar] [CrossRef] [PubMed]
  12. Gu, Y.; Li, M.; Wang, T.; Liang, Y.; Zhong, Z.; Wang, X.; Zhou, Q.; Chen, L.; Lang, Q.; He, Z.; et al. Lactation-related microRNA expression profiles of porcine breast milk exosomes. PLoS ONE 2012, 7, e43691. [Google Scholar] [CrossRef] [PubMed]
  13. Azevedo, L.C.; Janiszewski, M.; Pontieri, V.; Pedro Mde, A.; Bassi, E.; Tucci, P.J.; Laurindo, F.R. Platelet-derived exosomes from septic shock patients induce myocardial dysfunction. Crit. Care 2007, 11, R120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Masyuk, A.I.; Masyuk, T.V.; Larusso, N.F. Exosomes in the pathogenesis, diagnostics and therapeutics of liver diseases. J. Hepatol. 2013, 59, 621–625. [Google Scholar] [CrossRef] [PubMed]
  15. Vella, L.J.; Sharples, R.A.; Nisbet, R.M.; Cappai, R.; Hill, A.F. The role of exosomes in the processing of proteins associated with neurodegenerative diseases. Eur. Biophys. J. 2008, 37, 323–332. [Google Scholar] [CrossRef] [PubMed]
  16. Rak, J.; Guha, A. Extracellular vesicles—Vehicles that spread cancer genes. Bioessays News Rev. Mol. Cell. Dev. Biol. 2012, 34, 489–497. [Google Scholar] [CrossRef] [PubMed]
  17. Colombo, M.; Raposo, G.; Thery, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell Dev. Biol. 2014, 30, 255–289. [Google Scholar] [CrossRef] [PubMed]
  18. Van der Pol, E.; Boing, A.N.; Harrison, P.; Sturk, A.; Nieuwland, R. Classification, functions, and clinical relevance of extracellular vesicles. Pharmacol. Rev. 2012, 64, 676–705. [Google Scholar] [CrossRef] [PubMed]
  19. Hanson, P.I.; Cashikar, A. Multivesicular body morphogenesis. Annu. Rev. Cell Dev. Biol. 2012, 28, 337–362. [Google Scholar] [CrossRef] [PubMed]
  20. Escrevente, C.; Keller, S.; Altevogt, P.; Costa, J. Interaction and uptake of exosomes by ovarian cancer cells. BMC Cancer 2011, 11, 108. [Google Scholar] [CrossRef] [PubMed]
  21. Mathivanan, S.; Fahner, C.J.; Reid, G.E.; Simpson, R.J. Exocarta 2012: Database of exosomal proteins, RNA and lipids. Nucleic Acids Res. 2012, 40, D1241–D1244. [Google Scholar] [CrossRef] [PubMed]
  22. Thery, C.; Ostrowski, M.; Segura, E. Membrane vesicles as conveyors of immune responses. Nat. Rev. Immunol. 2009, 9, 581–593. [Google Scholar] [CrossRef] [PubMed]
  23. Mignot, G.; Roux, S.; Thery, C.; Segura, E.; Zitvogel, L. Prospects for exosomes in immunotherapy of cancer. J. Cell. Mol. Med. 2006, 10, 376–388. [Google Scholar] [CrossRef] [PubMed]
  24. Clayton, A.; Mason, M.D. Exosomes in tumour immunity. Curr. Oncol. 2009, 16, 46–49. [Google Scholar] [CrossRef] [PubMed]
  25. Sheldon, H.; Heikamp, E.; Turley, H.; Dragovic, R.; Thomas, P.; Oon, C.E.; Leek, R.; Edelmann, M.; Kessler, B.; Sainson, R.C.; et al. New mechanism for notch signaling to endothelium at a distance by Δ-like 4 incorporation into exosomes. Blood 2010, 116, 2385–2394. [Google Scholar] [CrossRef] [PubMed]
  26. Gross, J.C.; Chaudhary, V.; Bartscherer, K.; Boutros, M. Active wnt proteins are secreted on exosomes. Nat. Cell Biol. 2012, 14, 1036–1045. [Google Scholar] [CrossRef] [PubMed]
  27. Hasegawa, H.; Thomas, H.J.; Schooley, K.; Born, T.L. Native IL-32 is released from intestinal epithelial cells via a non-classical secretory pathway as a membrane-associated protein. Cytokine 2011, 53, 74–83. [Google Scholar] [CrossRef] [PubMed]
  28. Record, M.; Carayon, K.; Poirot, M.; Silvente-Poirot, S. Exosomes as new vesicular lipid transporters involved in cell-cell communication and various pathophysiologies. Biochim. Biophys. Acta 2014, 1841, 108–120. [Google Scholar] [CrossRef] [PubMed]
  29. Subra, C.; Grand, D.; Laulagnier, K.; Stella, A.; Lambeau, G.; Paillasse, M.; de Medina, P.; Monsarrat, B.; Perret, B.; Silvente-Poirot, S.; et al. Exosomes account for vesicle-mediated transcellular transport of activatable phospholipases and prostaglandins. J. Lipid Res. 2010, 51, 2105–2120. [Google Scholar] [CrossRef] [PubMed]
  30. Xiang, X.; Poliakov, A.; Liu, C.; Liu, Y.; Deng, Z.B.; Wang, J.; Cheng, Z.; Shah, S.V.; Wang, G.J.; Zhang, L.; et al. Induction of myeloid-derived suppressor cells by tumor exosomes. Int. J. Cancer 2009, 124, 2621–2633. [Google Scholar] [CrossRef] [PubMed]
  31. Pitt, J.M.; Charrier, M.; Viaud, S.; Andre, F.; Besse, B.; Chaput, N.; Zitvogel, L. Dendritic cell-derived exosomes as immunotherapies in the fight against cancer. J. Immunol. 2014, 193, 1006–1011. [Google Scholar] [CrossRef] [PubMed]
  32. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, 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]
  33. Gezer, U.; Ozgur, E.; Cetinkaya, M.; Isin, M.; Dalay, N. Long non-coding RNAs with low expression levels in cells are enriched in secreted exosomes. Cell Biol. Int. 2014, 38, 1076–1079. [Google Scholar] [CrossRef] [PubMed]
  34. Skog, J.; Wurdinger, T.; van Rijn, S.; Meijer, D.H.; Gainche, L.; Sena-Esteves, M.; Curry, W.T., Jr.; 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]
  35. Kogure, T.; Lin, W.L.; Yan, I.K.; Braconi, C.; Patel, T. Intercellular nanovesicle-mediated microRNA transfer: A mechanism of environmental modulation of hepatocellular cancer cell growth. Hepatology 2011, 54, 1237–1248. [Google Scholar] [CrossRef] [PubMed]
  36. Ohshima, K.; Inoue, K.; Fujiwara, A.; Hatakeyama, K.; Kanto, K.; Watanabe, Y.; Muramatsu, K.; Fukuda, Y.; Ogura, S.; Yamaguchi, K.; et al. Let-7 microRNA family is selectively secreted into the extracellular environment via exosomes in a metastatic gastric cancer cell line. PLoS ONE 2010, 5, e13247. [Google Scholar] [CrossRef] [PubMed]
  37. Ostenfeld, M.S.; Jeppesen, D.K.; Laurberg, J.R.; Boysen, A.T.; Bramsen, J.B.; Primdal-Bengtson, B.; Hendrix, A.; Lamy, P.; Dagnaes-Hansen, F.; Rasmussen, M.H.; et al. Cellular disposal of mir23b by rab27-dependent exosome release is linked to acquisition of metastatic properties. Cancer Res. 2014, 74, 5758–5771. [Google Scholar] [CrossRef] [PubMed]
  38. Fabbri, M.; Paone, A.; Calore, F.; Galli, R.; Croce, C.M. A new role for microRNAs, as ligands of toll-like receptors. RNA Biol. 2013, 10, 169–174. [Google Scholar] [CrossRef] [PubMed]
  39. Umezu, T.; Tadokoro, H.; Azuma, K.; Yoshizawa, S.; Ohyashiki, K.; Ohyashiki, J.H. Exosomal mir-135b shed from hypoxic multiple myeloma cells enhances angiogenesis by targeting factor-inhibiting HIF-1. Blood 2014, 124, 3748–3757. [Google Scholar] [CrossRef] [PubMed]
  40. Zhou, W.; Fong, M.Y.; Min, Y.; Somlo, G.; Liu, L.; Palomares, M.R.; Yu, Y.; Chow, A.; O'Connor, S.T.; Chin, A.R.; et al. Cancer-secreted mir-105 destroys vascular endothelial barriers to promote metastasis. Cancer Cell 2014, 25, 501–515. [Google Scholar] [CrossRef] [PubMed]
  41. Mittelbrunn, M.; Gutierrez-Vazquez, C.; Villarroya-Beltri, C.; Gonzalez, S.; Sanchez-Cabo, F.; Gonzalez, M.A.; Bernad, A.; Sanchez-Madrid, F. Unidirectional transfer of microRNA-loaded exosomes from T cells to antigen-presenting cells. Nat. Commun. 2011, 2, 282. [Google Scholar] [CrossRef] [PubMed]
  42. Bellingham, S.A.; Coleman, B.M.; Hill, A.F. Small RNA deep sequencing reveals a distinct miRNA signature released in exosomes from prion-infected neuronal cells. Nucleic Acids Res. 2012, 40, 10937–10949. [Google Scholar] [CrossRef] [PubMed]
  43. Rabinowits, G.; Gercel-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] [PubMed]
  44. Pigati, L.; Yaddanapudi, S.C.; Iyengar, R.; Kim, D.J.; Hearn, S.A.; Danforth, D.; Hastings, M.L.; Duelli, D.M. Selective release of microRNA species from normal and malignant mammary epithelial cells. PLoS ONE 2010, 5, e13515. [Google Scholar] [CrossRef] [PubMed]
  45. Jaiswal, R.; Luk, F.; Gong, J.; Mathys, J.M.; Grau, G.E.; Bebawy, M. Microparticle conferred microrna profiles—Implications in the transfer and dominance of cancer traits. Mol. Cancer 2012, 11, 37. [Google Scholar] [CrossRef] [PubMed]
  46. Azmi, A.S.; Bao, B.; Sarkar, F.H. Exosomes in cancer development, metastasis, and drug resistance: A comprehensive review. Cancer Metastasis Rev. 2013, 32, 623–642. [Google Scholar] [CrossRef] [PubMed]
  47. Raiborg, C.; Stenmark, H. The escrt machinery in endosomal sorting of ubiquitylated membrane proteins. Nature 2009, 458, 445–452. [Google Scholar] [CrossRef] [PubMed]
  48. Nabhan, J.F.; Hu, R.; Oh, R.S.; Cohen, S.N.; Lu, Q. Formation and release of arrestin domain-containing protein 1-mediated microvesicles (ARMMS) at plasma membrane by recruitment of TSG101 protein. Proc. Natl. Acad. Sci. USA 2012, 109, 4146–4151. [Google Scholar] [CrossRef] [PubMed]
  49. Booth, A.M.; Fang, Y.; Fallon, J.K.; Yang, J.M.; Hildreth, J.E.; Gould, S.J. Exosomes and HIV gag bud from endosome-like domains of the T cell plasma membrane. J. Cell Biol. 2006, 172, 923–935. [Google Scholar] [CrossRef] [PubMed]
  50. Baietti, M.F.; Zhang, Z.; Mortier, E.; Melchior, A.; Degeest, G.; Geeraerts, A.; Ivarsson, Y.; Depoortere, F.; Coomans, C.; Vermeiren, E.; et al. Syndecan-syntenin-alix regulates the biogenesis of exosomes. Nat. Cell Biol. 2012, 14, 677–685. [Google Scholar] [CrossRef] [PubMed]
  51. Thompson, C.A.; Purushothaman, A.; Ramani, V.C.; Vlodavsky, I.; Sanderson, R.D. Heparanase regulates secretion, composition, and function of tumor cell-derived exosomes. J. Biol. Chem. 2013, 288, 10093–10099. [Google Scholar] [CrossRef] [PubMed]
  52. Zhu, H.; Guariglia, S.; Yu, R.Y.; Li, W.; Brancho, D.; Peinado, H.; Lyden, D.; Salzer, J.; Bennett, C.; Chow, C.W. Mutation of simple in charcot-marie-tooth 1c alters production of exosomes. Mol. Biol. Cell 2013, 24, 1619–1637. [Google Scholar] [CrossRef] [PubMed]
  53. Stuffers, S.; Sem Wegner, C.; Stenmark, H.; Brech, A. Multivesicular endosome biogenesis in the absence of escrts. Traffic 2009, 10, 925–937. [Google Scholar] [CrossRef] [PubMed]
  54. Matsuo, H.; Chevallier, J.; Mayran, N.; Le Blanc, I.; Ferguson, C.; Faure, J.; Blanc, N.S.; Matile, S.; Dubochet, J.; Sadoul, R.; et al. Role of lbpa and alix in multivesicular liposome formation and endosome organization. Science 2004, 303, 531–534. [Google Scholar] [CrossRef] [PubMed]
  55. Babst, M. MVB vesicle formation: Escrt-dependent, escrt-independent and everything in between. Curr. Opin. Cell Biol. 2011, 23, 452–457. [Google Scholar] [CrossRef] [PubMed]
  56. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brugger, B.; Simons, M. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef] [PubMed]
  57. Laulagnier, K.; Grand, D.; Dujardin, A.; Hamdi, S.; Vincent-Schneider, H.; Lankar, D.; Salles, J.P.; Bonnerot, C.; Perret, B.; Record, M. Pld2 is enriched on exosomes and its activity is correlated to the release of exosomes. FEBS Lett. 2004, 572, 11–14. [Google Scholar] [CrossRef] [PubMed]
  58. Kajimoto, T.; Okada, T.; Miya, S.; Zhang, L.; Nakamura, S. Ongoing activation of sphingosine 1-phosphate receptors mediates maturation of exosomal multivesicular endosomes. Nat. Commun. 2013, 4, 2712. [Google Scholar] [CrossRef] [PubMed]
  59. Aung, T.; Chapuy, B.; Vogel, D.; Wenzel, D.; Oppermann, M.; Lahmann, M.; Weinhage, T.; Menck, K.; Hupfeld, T.; Koch, R.; et al. Exosomal evasion of humoral immunotherapy in aggressive b-cell lymphoma modulated by ATP-binding cassette transporter a3. Proc. Natl. Acad. Sci. USA 2011, 108, 15336–15341. [Google Scholar] [CrossRef] [PubMed]
  60. Van Niel, G.; Charrin, S.; Simoes, S.; Romao, M.; Rochin, L.; Saftig, P.; Marks, M.S.; Rubinstein, E.; Raposo, G. The tetraspanin CD63 regulates ESCRT-independent and -dependent endosomal sorting during melanogenesis. Dev. Cell 2011, 21, 708–721. [Google Scholar] [CrossRef] [PubMed]
  61. Mazurov, D.; Barbashova, L.; Filatov, A. Tetraspanin protein CD9 interacts with metalloprotease CD10 and enhances its release via exosomes. FEBS J. 2013, 280, 1200–1213. [Google Scholar] [CrossRef] [PubMed]
  62. Perez-Hernandez, D.; Gutierrez-Vazquez, C.; Jorge, I.; Lopez-Martin, S.; Ursa, A.; Sanchez-Madrid, F.; Vazquez, J.; Yanez-Mo, M. The intracellular interactome of tetraspanin-enriched microdomains reveals their function as sorting machineries toward exosomes. J. Biol. Chem. 2013, 288, 11649–11661. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, J.; Li, S.; Li, L.; Li, M.; Guo, C.; Yao, J.; Mi, S. Exosome and exosomal microRNA: Trafficking, sorting, and function. Genom. Proteom. Bioinform. 2015, 13, 17–24. [Google Scholar] [CrossRef] [PubMed]
  64. Villarroya-Beltri, C.; Gutierrez-Vazquez, C.; Sanchez-Cabo, F.; Perez-Hernandez, D.; Vazquez, J.; Martin-Cofreces, N.; Martinez-Herrera, D.J.; Pascual-Montano, A.; Mittelbrunn, M.; Sanchez-Madrid, F. Sumoylated HNRNPA2B1 controls the sorting of miRNAs into exosomes through binding to specific motifs. Nat. Commun. 2013, 4, 2980. [Google Scholar] [CrossRef] [PubMed]
  65. Shurtleff, M.J.; Temoche-Diaz, M.M.; Karfilis, K.V.; Ri, S.; Schekman, R. Y-box protein 1 is required to sort micrornas into exosomes in cells and in a cell-free reaction. Elife 2016, 5, e19276. [Google Scholar] [CrossRef] [PubMed]
  66. Kosaka, N.; Iguchi, H.; Hagiwara, K.; Yoshioka, Y.; Takeshita, F.; Ochiya, T. Neutral sphingomyelinase 2 (nSMase2)-dependent exosomal transfer of angiogenic microRNAs regulate cancer cell metastasis. J. Biol. Chem. 2013, 288, 10849–10859. [Google Scholar] [CrossRef] [PubMed]
  67. Goldie, B.J.; Dun, M.D.; Lin, M.; Smith, N.D.; Verrills, N.M.; Dayas, C.V.; Cairns, M.J. Activity-associated miRNA are packaged in Map1b-enriched exosomes released from depolarized neurons. Nucleic Acids Res. 2014, 42, 9195–9208. [Google Scholar] [CrossRef] [PubMed]
  68. Melo, S.A.; Sugimoto, H.; O’Connell, J.T.; Kato, N.; Villanueva, A.; Vidal, A.; Qiu, L.; Vitkin, E.; Perelman, L.T.; Melo, C.A.; et al. Cancer exosomes perform cell-independent microRNA biogenesis and promote tumorigenesis. Cancer Cell 2014, 26, 707–721. [Google Scholar] [CrossRef] [PubMed]
  69. Guduric-Fuchs, J.; O’Connor, A.; Camp, B.; O’Neill, C.L.; Medina, R.J.; Simpson, D.A. Selective extracellular vesicle-mediated export of an overlapping set of microRNAs from multiple cell types. BMC Genom. 2012, 13, 357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Kogure, T.; Yan, I.K.; Lin, W.L.; Patel, T. Extracellular vesicle-mediated transfer of a novel long noncoding RNA tuc339: A mechanism of intercellular signaling in human hepatocellular cancer. Genes Cancer 2013, 4, 261–272. [Google Scholar] [CrossRef] [PubMed]
  71. Takahashi, K.; Yan, I.K.; Kogure, T.; Haga, H.; Patel, T. Extracellular vesicle-mediated transfer of long non-coding RNA ror modulates chemosensitivity in human hepatocellular cancer. FEBS Open Bio 2014, 4, 458–467. [Google Scholar] [CrossRef] [PubMed]
  72. Ahadi, A.; Brennan, S.; Kennedy, P.J.; Hutvagner, G.; Tran, N. Long non-coding RNAs harboring miRNA seed regions are enriched in prostate cancer exosomes. Sci. Rep. 2016, 6, 24922. [Google Scholar] [CrossRef] [PubMed]
  73. Van Niel, G.; Porto-Carreiro, I.; Simoes, S.; Raposo, G. Exosomes: A common pathway for a specialized function. J. Biochem. 2006, 140, 13–21. [Google Scholar] [CrossRef] [PubMed]
  74. Savina, A.; Furlan, M.; Vidal, M.; Colombo, M.I. Exosome release is regulated by a calcium-dependent mechanism in K562 cells. J. Biol. Chem. 2003, 278, 20083–20090. [Google Scholar] [CrossRef] [PubMed]
  75. Parolini, I.; Federici, C.; Raggi, C.; Lugini, L.; Palleschi, S.; de Milito, A.; Coscia, C.; Iessi, E.; Logozzi, M.; Molinari, A.; et al. Microenvironmental PH is a key factor for exosome traffic in tumor cells. J. Biol. Chem. 2009, 284, 34211–34222. [Google Scholar] [CrossRef] [PubMed]
  76. Savina, A.; Vidal, M.; Colombo, M.I. The exosome pathway in K562 cells is regulated by rab11. J. Cell Sci. 2002, 115, 2505–2515. [Google Scholar] [PubMed]
  77. Ostrowski, M.; Carmo, N.B.; Krumeich, S.; Fanget, I.; Raposo, G.; Savina, A.; Moita, C.F.; Schauer, K.; Hume, A.N.; Freitas, R.P.; et al. Rab27a and rab27b control different steps of the exosome secretion pathway. Nat. Cell Biol. 2010, 12, 19–30. [Google Scholar] [CrossRef] [PubMed]
  78. Bobrie, A.; Colombo, M.; Raposo, G.; Thery, C. Exosome secretion: Molecular mechanisms and roles in immune responses. Traffic 2011, 12, 1659–1668. [Google Scholar] [CrossRef] [PubMed]
  79. Fader, C.M.; Savina, A.; Sanchez, D.; Colombo, M.I. Exosome secretion and red cell maturation: Exploring molecular components involved in the docking and fusion of multivesicular bodies in K562 cells. Blood Cells Mol. Dis. 2005, 35, 153–157. [Google Scholar] [CrossRef] [PubMed]
  80. Hsu, C.; Morohashi, Y.; Yoshimura, S.; Manrique-Hoyos, N.; Jung, S.; Lauterbach, M.A.; Bakhti, M.; Gronborg, M.; Mobius, W.; Rhee, J.; et al. Regulation of exosome secretion by rab35 and its GTPase-activating proteins TBC1D10A-c. J. Cell Biol. 2010, 189, 223–232. [Google Scholar] [CrossRef] [PubMed]
  81. Yu, X.; Harris, S.L.; Levine, A.J. The regulation of exosome secretion: A novel function of the p53 protein. Cancer Res. 2006, 66, 4795–4801. [Google Scholar] [CrossRef] [PubMed]
  82. Amzallag, N.; Passer, B.J.; Allanic, D.; Segura, E.; Thery, C.; Goud, B.; Amson, R.; Telerman, A. Tsap6 facilitates the secretion of translationally controlled tumor protein/histamine-releasing factor via a nonclassical pathway. J. Biol. Chem. 2004, 279, 46104–46112. [Google Scholar] [CrossRef] [PubMed]
  83. Lespagnol, A.; Duflaut, D.; Beekman, C.; Blanc, L.; Fiucci, G.; Marine, J.C.; Vidal, M.; Amson, R.; Telerman, A. Exosome secretion, including the DNA damage-induced p53-dependent secretory pathway, is severely compromised in tsap6/steap3-null mice. Cell Death Differ. 2008, 15, 1723–1733. [Google Scholar] [CrossRef] [PubMed]
  84. Cocucci, E.; Racchetti, G.; Podini, P.; Meldolesi, J. Enlargeosome traffic: Exocytosis triggered by various signals is followed by endocytosis, membrane shedding or both. Traffic 2007, 8, 742–757. [Google Scholar] [CrossRef] [PubMed]
  85. Sudhof, T.C.; Rothman, J.E. Membrane fusion: Grappling with snare and SM proteins. Science 2009, 323, 474–477. [Google Scholar] [CrossRef] [PubMed]
  86. Fader, C.M.; Sanchez, D.G.; Mestre, M.B.; Colombo, M.I. Ti-vamp/vamp7 and vamp3/cellubrevin: Two v-snare proteins involved in specific steps of the autophagy/multivesicular body pathways. Biochim. Biophys. Acta 2009, 1793, 1901–1916. [Google Scholar] [CrossRef] [PubMed]
  87. Feng, D.; Zhao, W.L.; Ye, Y.Y.; Bai, X.C.; Liu, R.Q.; Chang, L.F.; Zhou, Q.; Sui, S.F. Cellular internalization of exosomes occurs through phagocytosis. Traffic 2010, 11, 675–687. [Google Scholar] [CrossRef] [PubMed]
  88. Del Conde, I.; Shrimpton, C.N.; Thiagarajan, P.; Lopez, J.A. Tissue-factor-bearing microvesicles arise from lipid rafts and fuse with activated platelets to initiate coagulation. Blood 2005, 106, 1604–1611. [Google Scholar] [CrossRef] [PubMed]
  89. Tian, T.; Zhu, Y.L.; Zhou, Y.Y.; Liang, G.F.; Wang, Y.Y.; Hu, F.H.; Xiao, Z.D. Exosome uptake through clathrin-mediated endocytosis and macropinocytosis and mediating mir-21 delivery. J. Biol. Chem. 2014, 289, 22258–22267. [Google Scholar] [CrossRef] [PubMed]
  90. Svensson, K.J.; Christianson, H.C.; Wittrup, A.; Bourseau-Guilmain, E.; Lindqvist, E.; Svensson, L.M.; Morgelin, M.; Belting, M. Exosome uptake depends on Erk1/2-heat shock protein 27 signaling and lipid raft-mediated endocytosis negatively regulated by caveolin-1. J. Biol. Chem. 2013, 288, 17713–17724. [Google Scholar] [CrossRef] [PubMed]
  91. Christianson, H.C.; Svensson, K.J.; van Kuppevelt, T.H.; Li, J.P.; Belting, M. Cancer cell exosomes depend on cell-surface heparan sulfate proteoglycans for their internalization and functional activity. Proc. Natl. Acad. Sci. USA 2013, 110, 17380–17385. [Google Scholar] [CrossRef] [PubMed]
  92. Bellingham, S.A.; Guo, B.B.; Coleman, B.M.; Hill, A.F. Exosomes: Vehicles for the transfer of toxic proteins associated with neurodegenerative diseases? Front. Physiol. 2012, 3, 124. [Google Scholar] [CrossRef] [PubMed]
  93. Malik, Z.A.; Kott, K.S.; Poe, A.J.; Kuo, T.; Chen, L.; Ferrara, K.W.; Knowlton, A.A. Cardiac myocyte exosomes: Stability, Hsp60, and proteomics. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H954–H965. [Google Scholar] [CrossRef] [PubMed]
  94. Ibrahim, A.; Marban, E. Exosomes: Fundamental biology and roles in cardiovascular physiology. Annu. Rev. Physiol. 2016, 78, 67–83. [Google Scholar] [CrossRef] [PubMed]
  95. Yang, C.; Robbins, P.D. The roles of tumor-derived exosomes in cancer pathogenesis. Clin. Dev. Immunol. 2011, 2011, 842849. [Google Scholar] [CrossRef] [PubMed]
  96. Lai, R.C.; Chen, T.S.; Lim, S.K. Mesenchymal stem cell exosome: A novel stem cell-based therapy for cardiovascular disease. Regen. Med. 2011, 6, 481–492. [Google Scholar] [CrossRef] [PubMed]
  97. Cadigan, K.M. Regulating morphogen gradients in the drosophila wing. Semin. Cell Dev. Biol. 2002, 13, 83–90. [Google Scholar] [CrossRef]
  98. Lakkaraju, A.; Rodriguez-Boulan, E. Itinerant exosomes: Emerging roles in cell and tissue polarity. Trends Cell Biol. 2008, 18, 199–209. [Google Scholar] [CrossRef] [PubMed]
  99. Camussi, G.; Deregibus, M.C.; Bruno, S.; Grange, C.; Fonsato, V.; Tetta, C. Exosome/microvesicle-mediated epigenetic reprogramming of cells. Am. J. Cancer Res. 2011, 1, 98–110. [Google Scholar] [PubMed]
  100. Ratajczak, M.Z.; Kucia, M.; Jadczyk, T.; Greco, N.J.; Wojakowski, W.; Tendera, M.; Ratajczak, J. Pivotal role of paracrine effects in stem cell therapies in regenerative medicine: Can we translate stem cell-secreted paracrine factors and microvesicles into better therapeutic strategies? Leukemia 2012, 26, 1166–1173. [Google Scholar] [CrossRef] [PubMed]
  101. Herrera, M.B.; Fonsato, V.; Gatti, S.; Deregibus, M.C.; Sordi, A.; Cantarella, D.; Calogero, R.; Bussolati, B.; Tetta, C.; Camussi, G. Human liver stem cell-derived microvesicles accelerate hepatic regeneration in hepatectomized rats. J. Cell. Mol. Med. 2010, 14, 1605–1618. [Google Scholar] [CrossRef] [PubMed]
  102. Jang, Y.Y.; Collector, M.I.; Baylin, S.B.; Diehl, A.M.; Sharkis, S.J. Hematopoietic stem cells convert into liver cells within days without fusion. Nat. Cell Biol. 2004, 6, 532–539. [Google Scholar] [CrossRef] [PubMed]
  103. Aliotta, J.M.; Sanchez-Guijo, F.M.; Dooner, G.J.; Johnson, K.W.; Dooner, M.S.; Greer, K.A.; Greer, D.; Pimentel, J.; Kolankiewicz, L.M.; Puente, N.; et al. Alteration of marrow cell gene expression, protein production, and engraftment into lung by lung-derived microvesicles: A novel mechanism for phenotype modulation. Stem Cells 2007, 25, 2245–2256. [Google Scholar] [CrossRef] [PubMed]
  104. Borges, F.T.; Melo, S.A.; Ozdemir, B.C.; Kato, N.; Revuelta, I.; Miller, C.A.; Gattone, V.H., 2nd; LeBleu, V.S.; Kalluri, R. TGF-β1-containing exosomes from injured epithelial cells activate fibroblasts to initiate tissue regenerative responses and fibrosis. J. Am. Soc. Nephrol. JASN 2013, 24, 385–392. [Google Scholar] [CrossRef] [PubMed]
  105. Chaput, N.; Thery, C. Exosomes: Immune properties and potential clinical implementations. Semin. Immunopathol. 2011, 33, 419–440. [Google Scholar] [CrossRef] [PubMed]
  106. Kim, S.H.; Lechman, E.R.; Bianco, N.; Menon, R.; Keravala, A.; Nash, J.; Mi, Z.; Watkins, S.C.; Gambotto, A.; Robbins, P.D. Exosomes derived from IL-10-treated dendritic cells can suppress inflammation and collagen-induced arthritis. J. Immunol. 2005, 174, 6440–6448. [Google Scholar] [CrossRef] [PubMed]
  107. Kim, S.H.; Bianco, N.R.; Shufesky, W.J.; Morelli, A.E.; Robbins, P.D. Effective treatment of inflammatory disease models with exosomes derived from dendritic cells genetically modified to express IL-4. J. Immunol. 2007, 179, 2242–2249. [Google Scholar] [CrossRef] [PubMed]
  108. Andre, F.; Escudier, B.; Angevin, E.; Tursz, T.; Zitvogel, L. Exosomes for cancer immunotherapy. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. ESMO 2004, 15, iv141–iv144. [Google Scholar]
  109. Tran, T.H.; Mattheolabakis, G.; Aldawsari, H.; Amiji, M. Exosomes as nanocarriers for immunotherapy of cancer and inflammatory diseases. Clin. Immunol. 2015, 160, 46–58. [Google Scholar] [CrossRef] [PubMed]
  110. Kim, S.H.; Bianco, N.R.; Shufesky, W.J.; Morelli, A.E.; Robbins, P.D. Mhc class ii+ exosomes in plasma suppress inflammation in an antigen-specific and Fas ligand/Fas-dependent manner. J. Immunol. 2007, 179, 2235–2241. [Google Scholar] [CrossRef] [PubMed]
  111. Miksa, M.; Wu, R.; Dong, W.; Das, P.; Yang, D.; Wang, P. Dendritic cell-derived exosomes containing milk fat globule epidermal growth factor-factor viii attenuate proinflammatory responses in sepsis. Shock 2006, 25, 586–593. [Google Scholar] [CrossRef] [PubMed]
  112. Chivet, M.; Hemming, F.; Pernet-Gallay, K.; Fraboulet, S.; Sadoul, R. Emerging role of neuronal exosomes in the central nervous system. Front. Physiol. 2012, 3, 145. [Google Scholar] [CrossRef] [PubMed]
  113. Majka, M.; Janowska-Wieczorek, A.; Ratajczak, J.; Ehrenman, K.; Pietrzkowski, Z.; Kowalska, M.A.; Gewirtz, A.M.; Emerson, S.G.; Ratajczak, M.Z. Numerous growth factors, cytokines, and chemokines are secreted by human CD34(+) cells, myeloblasts, erythroblasts, and megakaryoblasts and regulate normal hematopoiesis in an autocrine/paracrine manner. Blood 2001, 97, 3075–3085. [Google Scholar] [CrossRef] [PubMed]
  114. Ponsaerts, P.; Berneman, Z.N. Modulation of cellular behavior by exogenous messenger RNA. Leukemia 2006, 20, 767–769. [Google Scholar] [CrossRef] [PubMed]
  115. McCready, J.; Sims, J.D.; Chan, D.; Jay, D.G. Secretion of extracellular Hsp90α via exosomes increases cancer cell motility: A role for plasminogen activation. BMC Cancer 2010, 10, 294. [Google Scholar] [CrossRef] [PubMed]
  116. Jung, T.; Castellana, D.; Klingbeil, P.; Cuesta Hernandez, I.; Vitacolonna, M.; Orlicky, D.J.; Roffler, S.R.; Brodt, P.; Zoller, M. Cd44v6 dependence of premetastatic niche preparation by exosomes. Neoplasia 2009, 11, 1093–1105. [Google Scholar] [CrossRef] [PubMed]
  117. Hood, J.L.; San, R.S.; Wickline, S.A. Exosomes released by melanoma cells prepare sentinel lymph nodes for tumor metastasis. Cancer Res. 2011, 71, 3792–3801. [Google Scholar] [CrossRef] [PubMed]
  118. Hegmans, J.P.; Bard, M.P.; Hemmes, A.; Luider, T.M.; Kleijmeer, M.J.; Prins, J.B.; Zitvogel, L.; Burgers, S.A.; Hoogsteden, H.C.; Lambrecht, B.N. Proteomic analysis of exosomes secreted by human mesothelioma cells. Am. J. Pathol. 2004, 164, 1807–1815. [Google Scholar] [CrossRef]
  119. Hood, J.L.; Pan, H.; Lanza, G.M.; Wickline, S.A.; Consortium for Translational Research in Advanced Imaging and Nanomedicine. Paracrine induction of endothelium by tumor exosomes. Lab. Investig. J. Tech. Methods Pathol. 2009, 89, 1317–1328. [Google Scholar] [CrossRef] [PubMed]
  120. Peinado, H.; Aleckovic, M.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; Garcia-Santos, G.; Ghajar, C.; et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through met. Nat. Med. 2012, 18, 883–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Rana, S.; Claas, C.; Kretz, C.C.; Nazarenko, I.; Zoeller, M. Activation-induced internalization differs for the tetraspanins CD9 and tspan8: Impact on tumor cell motility. Int. J. Biochem. Cell Biol. 2011, 43, 106–119. [Google Scholar] [CrossRef] [PubMed]
  122. Rana, S.; Malinowska, K.; Zoller, M. Exosomal tumor microRNA modulates premetastatic organ cells. Neoplasia 2013, 15, 281–295. [Google Scholar] [CrossRef] [PubMed]
  123. Gesierich, S.; Berezovskiy, I.; Ryschich, E.; Zoller, M. Systemic induction of the angiogenesis switch by the tetraspanin d6.1a/co-029. Cancer Res. 2006, 66, 7083–7094. [Google Scholar] [CrossRef] [PubMed]
  124. Nazarenko, I.; Rana, S.; Baumann, A.; McAlear, J.; Hellwig, A.; Trendelenburg, M.; Lochnit, G.; Preissner, K.T.; Zoller, M. Cell surface tetraspanin tspan8 contributes to molecular pathways of exosome-induced endothelial cell activation. Cancer Res. 2010, 70, 1668–1678. [Google Scholar] [CrossRef] [PubMed]
  125. Alqawi, O.; Wang, H.P.; Espiritu, M.; Singh, G. Chronic hypoxia promotes an aggressive phenotype in rat prostate cancer cells. Free Radic. Res. 2007, 41, 788–797. [Google Scholar] [CrossRef] [PubMed]
  126. Park, J.E.; Tan, H.S.; Datta, A.; Lai, R.C.; Zhang, H.; Meng, W.; Lim, S.K.; Sze, S.K. Hypoxic tumor cell modulates its microenvironment to enhance angiogenic and metastatic potential by secretion of proteins and exosomes. Mol. Cell. Proteom. MCP 2010, 9, 1085–1099. [Google Scholar] [CrossRef] [PubMed]
  127. Tadokoro, H.; Umezu, T.; Ohyashiki, K.; Hirano, T.; Ohyashiki, J.H. Exosomes derived from hypoxic leukemia cells enhance tube formation in endothelial cells. J. Biol. Chem. 2013, 288, 34343–34351. [Google Scholar] [CrossRef] [PubMed]
  128. Kucharzewska, P.; Christianson, H.C.; Welch, J.E.; Svensson, K.J.; Fredlund, E.; Ringner, M.; Morgelin, M.; Bourseau-Guilmain, E.; Bengzon, J.; Belting, M. Exosomes reflect the hypoxic status of glioma cells and mediate hypoxia-dependent activation of vascular cells during tumor development. Proc. Natl. Acad. Sci. USA 2013, 110, 7312–7317. [Google Scholar] [CrossRef] [PubMed]
  129. King, H.W.; Michael, M.Z.; Gleadle, J.M. Hypoxic enhancement of exosome release by breast cancer cells. BMC Cancer 2012, 12, 421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wolfers, J.; Lozier, A.; Raposo, G.; Regnault, A.; Thery, C.; Masurier, C.; Flament, C.; Pouzieux, S.; Faure, F.; Tursz, T.; et al. Tumor-derived exosomes are a source of shared tumor rejection antigens for CTL cross-priming. Nat. Med. 2001, 7, 297–303. [Google Scholar] [CrossRef] [PubMed]
  131. Andre, F.; Schartz, N.E.; Movassagh, M.; Flament, C.; Pautier, P.; Morice, P.; Pomel, C.; Lhomme, C.; Escudier, B.; Le Chevalier, T.; et al. Malignant effusions and immunogenic tumour-derived exosomes. Lancet 2002, 360, 295–305. [Google Scholar] [CrossRef]
  132. Dai, S.; Zhou, X.; Wang, B.; Wang, Q.; Fu, Y.; Chen, T.; Wan, T.; Yu, Y.; Cao, X. Enhanced induction of dendritic cell maturation and hla-a*0201-restricted cea-specific CD8(+) CTL response by exosomes derived from il-18 gene-modified cea-positive tumor cells. J. Mol. Med. 2006, 84, 1067–1076. [Google Scholar] [CrossRef] [PubMed]
  133. Wang, D.G.; Sun, S.Z.; Wang, Z.G.; Wang, X. study of induction of tumor specific cytotoxic T lymphocyte by using tumor-derived exosome. Hua Xi Kou Qiang Yi Xue Za Zhi 2006, 24, 160–163. (In Chinese) [Google Scholar] [CrossRef] [PubMed]
  134. Yao, Y.; Chen, L.; Wei, W.; Deng, X.; Ma, L.; Hao, S. Tumor cell-derived exosome-targeted dendritic cells stimulate stronger CD8+ CTL responses and antitumor immunities. Biochem. Biophys. Res. Commun. 2013, 436, 60–65. [Google Scholar] [CrossRef] [PubMed]
  135. Hao, S.; Bai, O.; Li, F.; Yuan, J.; Laferte, S.; Xiang, J. Mature dendritic cells pulsed with exosomes stimulate efficient cytotoxic t-lymphocyte responses and antitumour immunity. Immunology 2007, 120, 90–102. [Google Scholar] [CrossRef] [PubMed]
  136. Gastpar, R.; Gehrmann, M.; Bausero, M.A.; Asea, A.; Gross, C.; Schroeder, J.A.; Multhoff, G. Heat shock protein 70 surface-positive tumor exosomes stimulate migratory and cytolytic activity of natural killer cells. Cancer Res. 2005, 65, 5238–5247. [Google Scholar] [CrossRef] [PubMed]
  137. Radons, J.; Multhoff, G. Immunostimulatory functions of membrane-bound and exported heat shock protein 70. Exerc. Immunol. Rev. 2005, 11, 17–33. [Google Scholar] [PubMed]
  138. Andreola, G.; Rivoltini, L.; Castelli, C.; Huber, V.; Perego, P.; Deho, P.; Squarcina, P.; Accornero, P.; Lozupone, F.; Lugini, L.; et al. Induction of lymphocyte apoptosis by tumor cell secretion of Fasl-bearing microvesicles. J. Exp. Med. 2002, 195, 1303–1316. [Google Scholar] [CrossRef] [PubMed]
  139. Huber, V.; Fais, S.; Iero, M.; Lugini, L.; Canese, P.; Squarcina, P.; Zaccheddu, A.; Colone, M.; Arancia, G.; Gentile, M.; et al. Human colorectal cancer cells induce T-cell death through release of proapoptotic microvesicles: Role in immune escape. Gastroenterology 2005, 128, 1796–1804. [Google Scholar] [CrossRef] [PubMed]
  140. Klibi, J.; Niki, T.; Riedel, A.; Pioche-Durieu, C.; Souquere, S.; Rubinstein, E.; Le Moulec, S.; Guigay, J.; Hirashima, M.; Guemira, F.; et al. Blood diffusion and th1-suppressive effects of galectin-9-containing exosomes released by epstein-barr virus-infected nasopharyngeal carcinoma cells. Blood 2009, 113, 1957–1966. [Google Scholar] [CrossRef] [PubMed]
  141. Chalmin, F.; Ladoire, S.; Mignot, G.; Vincent, J.; Bruchard, M.; Remy-Martin, J.P.; Boireau, W.; Rouleau, A.; Simon, B.; Lanneau, D.; et al. Membrane-associated Hsp72 from tumor-derived exosomes mediates stat3-dependent immunosuppressive function of mouse and human myeloid-derived suppressor cells. J. Clin. Investig. 2010, 120, 457–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Liu, C.; Yu, S.; Zinn, K.; Wang, J.; Zhang, L.; Jia, Y.; Kappes, J.C.; Barnes, S.; Kimberly, R.P.; Grizzle, W.E.; et al. Murine mammary carcinoma exosomes promote tumor growth by suppression of nk cell function. J. Immunol. 2006, 176, 1375–1385. [Google Scholar] [CrossRef] [PubMed]
  143. Clayton, A.; Mitchell, J.P.; Court, J.; Linnane, S.; Mason, M.D.; Tabi, Z. Human tumor-derived exosomes down-modulate NKG2D expression. J. Immunol. 2008, 180, 7249–7258. [Google Scholar] [CrossRef] [PubMed]
  144. Mincheva-Nilsson, L.; Baranov, V. Cancer exosomes and NKG2D receptor-ligand interactions: Impairing nkg2d-mediated cytotoxicity and anti-tumour immune surveillance. Semin. Cancer Biol. 2014, 28, 24–30. [Google Scholar] [CrossRef] [PubMed]
  145. Wieckowski, E.U.; Visus, C.; Szajnik, M.; Szczepanski, M.J.; Storkus, W.J.; Whiteside, T.L. Tumor-derived microvesicles promote regulatory t cell expansion and induce apoptosis in tumor-reactive activated CD8+ T lymphocytes. J. Immunol. 2009, 183, 3720–3730. [Google Scholar] [CrossRef] [PubMed]
  146. Wada, J.; Onishi, H.; Suzuki, H.; Yamasaki, A.; Nagai, S.; Morisaki, T.; Katano, M. Surface-bound TGF-β1 on effusion-derived exosomes participates in maintenance of number and suppressive function of regulatory T-cells in malignant effusions. Anticancer Res. 2010, 30, 3747–3757. [Google Scholar] [PubMed]
  147. Yu, S.; Liu, C.; Su, K.; Wang, J.; Liu, Y.; Zhang, L.; Li, C.; Cong, Y.; Kimberly, R.; Grizzle, W.E.; et al. Tumor exosomes inhibit differentiation of bone marrow dendritic cells. J. Immunol. 2007, 178, 6867–6875. [Google Scholar] [CrossRef] [PubMed]
  148. Quail, D.F.; Joyce, J.A. Microenvironmental regulation of tumor progression and metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef] [PubMed]
  149. Chow, A.; Zhou, W.; Liu, L.; Fong, M.Y.; Champer, J.; van Haute, D.; Chin, A.R.; Ren, X.; Gugiu, B.G.; Meng, Z.; et al. Macrophage immunomodulation by breast cancer-derived exosomes requires toll-like receptor 2-mediated activation of nf-kappab. Sci. Rep. 2014, 4, 5750. [Google Scholar] [CrossRef] [PubMed]
  150. Bretz, N.P.; Ridinger, J.; Rupp, A.K.; Rimbach, K.; Keller, S.; Rupp, C.; Marme, F.; Umansky, L.; Umansky, V.; Eigenbrod, T.; et al. Body fluid exosomes promote secretion of inflammatory cytokines in monocytic cells via toll-like receptor signaling. J. Biol. Chem. 2013, 288, 36691–36702. [Google Scholar] [CrossRef] [PubMed]
  151. Fabbri, M.; Paone, A.; Calore, F.; Galli, R.; Gaudio, E.; Santhanam, R.; Lovat, F.; Fadda, P.; Mao, C.; Nuovo, G.J.; et al. MicroRNAs bind to toll-like receptors to induce prometastatic inflammatory response. Proc. Natl. Acad. Sci. USA 2012, 109, E2110–E2116. [Google Scholar] [CrossRef] [PubMed]
  152. Silva, J.; Garcia, V.; Rodriguez, M.; Compte, M.; Cisneros, E.; Veguillas, P.; Garcia, J.M.; Dominguez, G.; Campos-Martin, Y.; Cuevas, J.; et al. Analysis of exosome release and its prognostic value in human colorectal cancer. Genes Chromosomes Cancer 2012, 51, 409–418. [Google Scholar] [CrossRef] [PubMed]
  153. Friel, A.M.; Corcoran, C.; Crown, J.; O’Driscoll, L. Relevance of circulating tumor cells, extracellular nucleic acids, and exosomes in breast cancer. Breast Cancer Res. Treat. 2010, 123, 613–625. [Google Scholar] [CrossRef] [PubMed]
  154. Raposo, G.; Stoorvogel, W. Extracellular vesicles: Exosomes, microvesicles, and friends. J. Cell Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef] [PubMed]
  155. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Thery, C.; Raposo, G. Analysis of escrt functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126, 5553–5565. [Google Scholar] [CrossRef] [PubMed]
  156. 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]
  157. Hoshino, D.; Kirkbride, K.C.; Costello, K.; Clark, E.S.; Sinha, S.; Grega-Larson, N.; Tyska, M.J.; Weaver, A.M. Exosome secretion is enhanced by invadopodia and drives invasive behavior. Cell Rep. 2013, 5, 1159–1168. [Google Scholar] [CrossRef] [PubMed]
  158. Friand, V.; David, G.; Zimmermann, P. Syntenin and syndecan in the biogenesis of exosomes. Biol. Cell Auspices Eur. Cell Biol. Organ. 2015, 107, 331–341. [Google Scholar] [CrossRef] [PubMed]
  159. Bianco, F.; Perrotta, C.; Novellino, L.; Francolini, M.; Riganti, L.; Menna, E.; Saglietti, L.; Schuchman, E.H.; Furlan, R.; Clementi, E.; et al. Acid sphingomyelinase activity triggers microparticle release from glial cells. EMBO J. 2009, 28, 1043–1054. [Google Scholar] [CrossRef] [PubMed]
  160. Escola, J.M.; Kleijmeer, M.J.; Stoorvogel, W.; Griffith, J.M.; Yoshie, O.; Geuze, H.J. Selective enrichment of tetraspan proteins on the internal vesicles of multivesicular endosomes and on exosomes secreted by human b-lymphocytes. J. Biol. Chem. 1998, 273, 20121–20127. [Google Scholar] [CrossRef] [PubMed]
  161. Li, B.; Antonyak, M.A.; Zhang, J.; Cerione, R.A. Rhoa triggers a specific signaling pathway that generates transforming microvesicles in cancer cells. Oncogene 2012, 31, 4740–4749. [Google Scholar] [CrossRef] [PubMed]
  162. 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]
  163. Hendrix, A.; de Wever, O. Rab27 GTPases distribute extracellular nanomaps for invasive growth and metastasis: Implications for prognosis and treatment. Int. J. Mol. Sci. 2013, 14, 9883–9892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Bobrie, A.; Krumeich, S.; Reyal, F.; Recchi, C.; Moita, L.F.; Seabra, M.C.; Ostrowski, M.; Thery, C. Rab27a supports exosome-dependent and -independent mechanisms that modify the tumor microenvironment and can promote tumor progression. Cancer Res. 2012, 72, 4920–4930. [Google Scholar] [CrossRef] [PubMed]
  165. Savina, A.; Fader, C.M.; Damiani, M.T.; Colombo, M.I. Rab11 promotes docking and fusion of multivesicular bodies in a calcium-dependent manner. Traffic 2005, 6, 131–143. [Google Scholar] [CrossRef] [PubMed]
  166. Alonso, R.; Mazzeo, C.; Rodriguez, M.C.; Marsh, M.; Fraile-Ramos, A.; Calvo, V.; Avila-Flores, A.; Merida, I.; Izquierdo, M. Diacylglycerol kinase alpha regulates the formation and polarisation of mature multivesicular bodies involved in the secretion of Fas ligand-containing exosomes in t lymphocytes. Cell Death Differ. 2011, 18, 1161–1173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Federici, C.; Petrucci, F.; Caimi, S.; Cesolini, A.; Logozzi, M.; Borghi, M.; D’Ilio, S.; Lugini, L.; Violante, N.; Azzarito, T.; et al. Exosome release and low ph belong to a framework of resistance of human melanoma cells to cisplatin. PLoS ONE 2014, 9, e88193. [Google Scholar]
  168. Fitzner, D.; Schnaars, M.; van Rossum, D.; Krishnamoorthy, G.; Dibaj, P.; Bakhti, M.; Regen, T.; Hanisch, U.K.; Simons, M. Selective transfer of exosomes from oligodendrocytes to microglia by macropinocytosis. J. Cell Sci. 2011, 124, 447–458. [Google Scholar] [CrossRef] [PubMed]
  169. Lima, L.G.; Chammas, R.; Monteiro, R.Q.; Moreira, M.E.; Barcinski, M.A. Tumor-derived microvesicles modulate the establishment of metastatic melanoma in a phosphatidylserine-dependent manner. Cancer Lett. 2009, 283, 168–175. [Google Scholar] [CrossRef] [PubMed]
  170. Atai, N.A.; Balaj, L.; van Veen, H.; Breakefield, X.O.; Jarzyna, P.A.; van Noorden, C.J.; Skog, J.; Maguire, C.A. Heparin blocks transfer of extracellular vesicles between donor and recipient cells. J. Neuro Oncol. 2013, 115, 343–351. [Google Scholar] [CrossRef] [PubMed]
  171. Marleau, A.M.; Chen, C.S.; Joyce, J.A.; Tullis, R.H. Exosome removal as a therapeutic adjuvant in cancer. J. Transl. Med. 2012, 10, 134. [Google Scholar] [CrossRef] [PubMed]
  172. Viaud, S.; Thery, C.; Ploix, S.; Tursz, T.; Lapierre, V.; Lantz, O.; Zitvogel, L.; Chaput, N. Dendritic cell-derived exosomes for cancer immunotherapy: What’s next? Cancer Res. 2010, 70, 1281–1285. [Google Scholar] [CrossRef] [PubMed]
  173. Hao, S.; Moyana, T.; Xiang, J. Review: Cancer immunotherapy by exosome-based vaccines. Cancer Biother. Radiopharm. 2007, 22, 692–703. [Google Scholar] [CrossRef] [PubMed]
  174. Tan, A.; de la Pena, H.; Seifalian, A.M. The application of exosomes as a nanoscale cancer vaccine. Int. J. Nanomed. 2010, 5, 889–900. [Google Scholar]
  175. Morse, M.A.; Garst, J.; Osada, T.; Khan, S.; Hobeika, A.; Clay, T.M.; Valente, N.; Shreeniwas, R.; Sutton, M.A.; Delcayre, A.; et al. A phase I study of dexosome immunotherapy in patients with advanced non-small cell lung cancer. J. Transl. Med. 2005, 3, 9. [Google Scholar] [CrossRef] [PubMed]
  176. Escudier, B.; Dorval, T.; Chaput, N.; Andre, F.; Caby, M.P.; Novault, S.; Flament, C.; Leboulaire, C.; Borg, C.; Amigorena, S.; et al. Vaccination of metastatic melanoma patients with autologous dendritic cell (DC) derived-exosomes: Results of thefirst phase i clinical trial. J. Transl. Med. 2005, 3, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Bu, N.; Wu, H.; Sun, B.; Zhang, G.; Zhan, S.; Zhang, R.; Zhou, L. Exosome-loaded dendritic cells elicit tumor-specific CD8+ cytotoxic t cells in patients with glioma. J. Neurooncol. 2011, 104, 659–667. [Google Scholar] [CrossRef] [PubMed]
  178. Navabi, H.; Croston, D.; Hobot, J.; Clayton, A.; Zitvogel, L.; Jasani, B.; Bailey-Wood, R.; Wilson, K.; Tabi, Z.; Mason, M.D.; et al. Preparation of human ovarian cancer ascites-derived exosomes for a clinical trial. Blood Cells Mol. Dis. 2005, 35, 149–152. [Google Scholar] [CrossRef] [PubMed]
  179. Runz, S.; Keller, S.; Rupp, C.; Stoeck, A.; Issa, Y.; Koensgen, D.; Mustea, A.; Sehouli, J.; Kristiansen, G.; Altevogt, P. Malignant ascites-derived exosomes of ovarian carcinoma patients contain CD24 and epcam. Gynecol. Oncol. 2007, 107, 563–571. [Google Scholar] [CrossRef] [PubMed]
  180. Dai, S.; Wei, D.; Wu, Z.; Zhou, X.; Wei, X.; Huang, H.; Li, G. Phase i clinical trial of autologous ascites-derived exosomes combined with gm-csf for colorectal cancer. Mol. Ther. 2008, 16, 782–790. [Google Scholar] [CrossRef] [PubMed]
  181. Herring, J.M.; McMichael, M.A.; Smith, S.A. Microparticles in health and disease. J. Vet. Intern. Med. 2013, 27, 1020–1033. [Google Scholar] [CrossRef] [PubMed]
  182. Li, J.; Sherman-Baust, C.A.; Tsai-Turton, M.; Bristow, R.E.; Roden, R.B.; Morin, P.J. Claudin-containing exosomes in the peripheral circulation of women with ovarian cancer. BMC Cancer 2009, 9, 244. [Google Scholar] [CrossRef] [PubMed]
  183. Lau, C.; Kim, Y.; Chia, D.; Spielmann, N.; Eibl, G.; Elashoff, D.; Wei, F.; Lin, Y.L.; Moro, A.; Grogan, T.; et al. Role of pancreatic cancer-derived exosomes in salivary biomarker development. J. Biol. Chem. 2013, 288, 26888–26897. [Google Scholar] [CrossRef] [PubMed]
  184. Rodriguez, M.; Silva, J.; Lopez-Alfonso, A.; Lopez-Muniz, M.B.; Pena, C.; Dominguez, G.; Garcia, J.M.; Lopez-Gonzalez, A.; Mendez, M.; Provencio, M.; et al. Different exosome cargo from plasma/bronchoalveolar lavage in non-small-cell lung cancer. Genes Chromosomes Cancer 2014, 53, 713–724. [Google Scholar] [CrossRef] [PubMed]
  185. 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]
  186. 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. Off. Publ. Int. Assoc. Study Lung Cancer 2013, 8, 1156–1162. [Google Scholar] [CrossRef] [PubMed]
  187. Huang, X.; Yuan, T.; Liang, M.; Du, M.; Xia, S.; Dittmar, R.; Wang, D.; See, W.; Costello, B.A.; Quevedo, F.; et al. Exosomal mir-1290 and mir-375 as prognostic markers in castration-resistant prostate cancer. Eur. Urol. 2015, 67, 33–41. [Google Scholar] [CrossRef] [PubMed]
  188. Eichelser, C.; Stuckrath, I.; Muller, V.; Milde-Langosch, K.; Wikman, H.; Pantel, K.; Schwarzenbach, H. Increased serum levels of circulating exosomal microRNA-373 in receptor-negative breast cancer patients. Oncotarget 2014, 5, 9650–9663. [Google Scholar] [CrossRef] [PubMed]
  189. Wang, J.; Zhou, Y.; Lu, J.; Sun, Y.; Xiao, H.; Liu, M.; Tian, L. Combined detection of serum exosomal mir-21 and hotair as diagnostic and prognostic biomarkers for laryngeal squamous cell carcinoma. Med. Oncol. 2014, 31, 148. [Google Scholar] [CrossRef] [PubMed]
  190. 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. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2015, 36, 2007–2012. [Google Scholar] [CrossRef] [PubMed]
  191. Vickers, K.C.; Remaley, A.T. Lipid-based carriers of microRNAs and intercellular communication. Curr. Opin. Lipidol. 2012, 23, 91–97. [Google Scholar] [CrossRef] [PubMed]
  192. Alvarez-Erviti, L.; Seow, Y.; Yin, H.; Betts, C.; Lakhal, S.; Wood, M.J. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef] [PubMed]
  193. Lai, R.C.; Yeo, R.W.; Tan, K.H.; Lim, S.K. Exosomes for drug delivery—A novel application for the mesenchymal stem cell. Biotechnol. Adv. 2013, 31, 543–551. [Google Scholar] [CrossRef] [PubMed]
  194. Thakur, B.K.; Zhang, H.; Becker, A.; Matei, I.; Huang, Y.; Costa-Silva, B.; Zheng, Y.; Hoshino, A.; Brazier, H.; Xiang, J.; et al. Double-stranded DNA in exosomes: A novel biomarker in cancer detection. Cell Res. 2014, 24, 766–769. [Google Scholar] [CrossRef] [PubMed]
  195. 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] [PubMed]
  196. Wahlgren, J.; Karlson, T.D.L.; Brisslert, M.; Vaziri Sani, F.; Telemo, E.; Sunnerhagen, P.; Valadi, H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012, 40, e130. [Google Scholar] [CrossRef] [PubMed]
  197. El Andaloussi, S.; Lakhal, S.; Mager, I.; Wood, M.J. Exosomes for targeted siRNA delivery across biological barriers. Adv. Drug Deliv. Rev. 2013, 65, 391–397. [Google Scholar] [CrossRef] [PubMed]
  198. Munoz, J.L.; Bliss, S.A.; Greco, S.J.; Ramkissoon, S.H.; Ligon, K.L.; Rameshwar, P. Delivery of functional anti-mir-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity. Mol. Ther. Nucleic Acids 2013, 2, e126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Ohno, S.; 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. J. Am. Soc. Gene Ther. 2013, 21, 185–191. [Google Scholar] [CrossRef] [PubMed]
  200. Pan, Q.; Ramakrishnaiah, V.; Henry, S.; Fouraschen, S.; de Ruiter, P.E.; Kwekkeboom, J.; Tilanus, H.W.; Janssen, H.L.; van der Laan, L.J. Hepatic cell-to-cell transmission of small silencing RNA can extend the therapeutic reach of RNA interference (RNAi). Gut 2012, 61, 1330–1339. [Google Scholar] [CrossRef] [PubMed]
  201. Sun, D.; Zhuang, X.; Xiang, X.; Liu, Y.; Zhang, S.; Liu, C.; Barnes, S.; Grizzle, W.; Miller, D.; Zhang, H.G. A novel nanoparticle drug delivery system: The anti-inflammatory activity of curcumin is enhanced when encapsulated in exosomes. Mol. Ther. J. Am. Soc. Gene Ther. 2010, 18, 1606–1614. [Google Scholar] [CrossRef] [PubMed]
  202. Zhuang, X.; Xiang, X.; Grizzle, W.; Sun, D.; Zhang, S.; Axtell, R.C.; Ju, S.; Mu, J.; Zhang, L.; Steinman, L.; et al. Treatment of brain inflammatory diseases by delivering exosome encapsulated anti-inflammatory drugs from the nasal region to the brain. Mol. Ther. J. Am. Soc. Gene Ther. 2011, 19, 1769–1779. [Google Scholar] [CrossRef] [PubMed]
  203. Tian, Y.; Li, S.; Song, J.; Ji, T.; Zhu, M.; Anderson, G.J.; Wei, J.; Nie, G. A doxorubicin delivery platform using engineered natural membrane vesicle exosomes for targeted tumor therapy. Biomaterials 2014, 35, 2383–2390. [Google Scholar] [CrossRef] [PubMed]
  204. Yang, T.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S. Exosome delivered anticancer drugs across the blood-brain barrier for brain cancer therapy in danio rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef] [PubMed]
  205. Tacar, O.; Sriamornsak, P.; Dass, C.R. Doxorubicin: An update on anticancer molecular action, toxicity and novel drug delivery systems. J. Pharm. Pharmacol. 2013, 65, 157–170. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Exosome biogenesis, cargo sorting, and release. Illustration of loading into exosomes of cargo such as nucleic acid and proteins. Endocytosis of the plasma membrane (A) results in the uptake of proteins, nucleic acids, and membrane-associated molecules, and formation of the early endosome (EE) (B); Upon transformation of the early endosome into the late endosome (LE) (C), exosomes are formed by inward budding of the late endosome/multivesicular body (MVB) with the content in a similar orientation as in the plasma membrane (D); Fusion of the MVB with the plasma membrane allows for the release of exosomes into the extracellular space (E); Alternatively, the MVB may fuse with the lysosome for degradation (F). ER: Endoplasmic reticulum.
Figure 1. Exosome biogenesis, cargo sorting, and release. Illustration of loading into exosomes of cargo such as nucleic acid and proteins. Endocytosis of the plasma membrane (A) results in the uptake of proteins, nucleic acids, and membrane-associated molecules, and formation of the early endosome (EE) (B); Upon transformation of the early endosome into the late endosome (LE) (C), exosomes are formed by inward budding of the late endosome/multivesicular body (MVB) with the content in a similar orientation as in the plasma membrane (D); Fusion of the MVB with the plasma membrane allows for the release of exosomes into the extracellular space (E); Alternatively, the MVB may fuse with the lysosome for degradation (F). ER: Endoplasmic reticulum.
Ijms 18 00538 g001
Figure 2. Molecular composition of exosomes. Exosomes are membrane-derived nanovesicles (30–100 nm in diameter) secreted from several cell types. They pack a variety of cellular components, including nucleic acids (e.g., DNA, mRNA, and miRNA), lipids (e.g., cholesterol and ceramide), mRNAs, membrane trafficking proteins (e.g., annexin, Rab 27, SNAP25), chaperones (e.g., Hsp70 and Hsp90), and various tissue-specific proteins involved in antigen presentation as integrins and tetraspanins (CD9, CD63, CD81, and CD82) as well as MHC-I and -II (Major Histocompatibility Complex).
Figure 2. Molecular composition of exosomes. Exosomes are membrane-derived nanovesicles (30–100 nm in diameter) secreted from several cell types. They pack a variety of cellular components, including nucleic acids (e.g., DNA, mRNA, and miRNA), lipids (e.g., cholesterol and ceramide), mRNAs, membrane trafficking proteins (e.g., annexin, Rab 27, SNAP25), chaperones (e.g., Hsp70 and Hsp90), and various tissue-specific proteins involved in antigen presentation as integrins and tetraspanins (CD9, CD63, CD81, and CD82) as well as MHC-I and -II (Major Histocompatibility Complex).
Ijms 18 00538 g002
Figure 3. Biological functions of exosomes in tumorigenesis. Exosomes released from tumor cells affect the local tumor microenvironment and are critically involved in tumor initiation, growth, progression, and metastasis by transferring oncogenic proteins and nucleic acids. (A) Exosomes travel to distant sites to promote generation of the pre-metastatic niche; (B) Angiogenesis is increased and endothelial and stromal cell differentiation is induced, leading to a pro-tumor environment; (C) Exosomes have immunosuppressive effects and assist cancers in immune evasion. Cytotoxic T cells are induced to apoptosis, while natural killer cell proliferation is impaired, and T-helper cells differentiate toward a T-regulatory cell phenotype; (D) Bone marrow-derived cells are recruited to tumor and pre-tumor tissue where they contribute to cancer development; (E) Exosomes are also responsible for the recruitment and activation of tumor-associated macrophages (TAMs) by promoting their polarization. TAMs support diverse phenotypes within the primary tumor, including growth, angiogenesis, and invasion, by secreting a plethora of pro-tumorigenic proteases, cytokines, and growth factors; (F) Exosomes can functionally modify fibroblasts by reprogramming these cells to cancer-associated fibroblasts (CAFs), which exhibit myofibroblastic differentiation. Red arrows indicates a negative contribution or repression and green arrows indicate an activation or positive function.
Figure 3. Biological functions of exosomes in tumorigenesis. Exosomes released from tumor cells affect the local tumor microenvironment and are critically involved in tumor initiation, growth, progression, and metastasis by transferring oncogenic proteins and nucleic acids. (A) Exosomes travel to distant sites to promote generation of the pre-metastatic niche; (B) Angiogenesis is increased and endothelial and stromal cell differentiation is induced, leading to a pro-tumor environment; (C) Exosomes have immunosuppressive effects and assist cancers in immune evasion. Cytotoxic T cells are induced to apoptosis, while natural killer cell proliferation is impaired, and T-helper cells differentiate toward a T-regulatory cell phenotype; (D) Bone marrow-derived cells are recruited to tumor and pre-tumor tissue where they contribute to cancer development; (E) Exosomes are also responsible for the recruitment and activation of tumor-associated macrophages (TAMs) by promoting their polarization. TAMs support diverse phenotypes within the primary tumor, including growth, angiogenesis, and invasion, by secreting a plethora of pro-tumorigenic proteases, cytokines, and growth factors; (F) Exosomes can functionally modify fibroblasts by reprogramming these cells to cancer-associated fibroblasts (CAFs), which exhibit myofibroblastic differentiation. Red arrows indicates a negative contribution or repression and green arrows indicate an activation or positive function.
Ijms 18 00538 g003
Figure 4. Exosome-based Diagnostics and Therapeutics. Exosomes hold a potential to be used as therapeutic or diagnostics tools. (A) Therapeutics: Exogenous or autologous exosomes can be isolated to deliver a desired payload in combination with chemotherapeutics, adjuvants of chemotherapy or as immunotherapy; (B) Diagnostics: Biomarkers can be determined to evaluate the expression of proteins, inflammation markers or nc-RNA present in exosomes from biological fluids.
Figure 4. Exosome-based Diagnostics and Therapeutics. Exosomes hold a potential to be used as therapeutic or diagnostics tools. (A) Therapeutics: Exogenous or autologous exosomes can be isolated to deliver a desired payload in combination with chemotherapeutics, adjuvants of chemotherapy or as immunotherapy; (B) Diagnostics: Biomarkers can be determined to evaluate the expression of proteins, inflammation markers or nc-RNA present in exosomes from biological fluids.
Ijms 18 00538 g004

Share and Cite

MDPI and ACS Style

H. Rashed, M.; Bayraktar, E.; K. Helal, G.; Abd-Ellah, M.F.; Amero, P.; Chavez-Reyes, A.; Rodriguez-Aguayo, C. Exosomes: From Garbage Bins to Promising Therapeutic Targets. Int. J. Mol. Sci. 2017, 18, 538. https://doi.org/10.3390/ijms18030538

AMA Style

H. Rashed M, Bayraktar E, K. Helal G, Abd-Ellah MF, Amero P, Chavez-Reyes A, Rodriguez-Aguayo C. Exosomes: From Garbage Bins to Promising Therapeutic Targets. International Journal of Molecular Sciences. 2017; 18(3):538. https://doi.org/10.3390/ijms18030538

Chicago/Turabian Style

H. Rashed, Mohammed, Emine Bayraktar, Gouda K. Helal, Mohamed F. Abd-Ellah, Paola Amero, Arturo Chavez-Reyes, and Cristian Rodriguez-Aguayo. 2017. "Exosomes: From Garbage Bins to Promising Therapeutic Targets" International Journal of Molecular Sciences 18, no. 3: 538. https://doi.org/10.3390/ijms18030538

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