Next Article in Journal
Hierarchical Structure and Properties of the Bone at Nano Level
Next Article in Special Issue
Hypothermic Preservation of Adipose-Derived Mesenchymal Stromal Cells as a Viable Solution for the Storage and Distribution of Cell Therapy Products
Previous Article in Journal
Droplet Microfluidics Enables Tracing of Target Cells at the Single-Cell Transcriptome Resolution
Previous Article in Special Issue
Identification of the Factor That Leads Human Mesenchymal Stem Cell Lines into Decellularized Bone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Promise to Reality: Bioengineering Strategies to Enhance the Therapeutic Potential of Extracellular Vesicles

by
Miguel de Almeida Fuzeta
1,2,
Pedro P. Gonçalves
1,2,
Ana Fernandes-Platzgummer
1,2,
Joaquim M. S. Cabral
1,2,
Nuno Bernardes
1,2,* and
Cláudia L. da Silva
1,2,*
1
iBB–Institute for Bioengineering and Biosciences and Department of Bioengineering, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
2
Associate Laboratory i4HB–Institute for Health and Bioeconomy at Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
*
Authors to whom correspondence should be addressed.
Bioengineering 2022, 9(11), 675; https://doi.org/10.3390/bioengineering9110675
Submission received: 8 October 2022 / Revised: 31 October 2022 / Accepted: 3 November 2022 / Published: 10 November 2022
(This article belongs to the Special Issue Mesenchymal Stem Cells in Regenerative Medicine)

Abstract

:
Extracellular vesicles (EVs) have been the focus of great attention over the last decade, considering their promising application as next-generation therapeutics. EVs have emerged as relevant mediators of intercellular communication, being associated with multiple physiological processes, but also in the pathogenesis of several diseases. Given their natural ability to shuttle messages between cells, EVs have been explored both as inherent therapeutics in regenerative medicine and as drug delivery vehicles targeting multiple diseases. However, bioengineering strategies are required to harness the full potential of EVs for therapeutic use. For that purpose, a good understanding of EV biology, from their biogenesis to the way they are able to shuttle messages and establish interactions with recipient cells, is needed. Here, we review the current state-of-the-art on EV biology, complemented by representative examples of EVs roles in several pathophysiological processes, as well as the intrinsic therapeutic properties of EVs and paradigmatic strategies to produce and develop engineered EVs as next-generation drug delivery systems.

Graphical Abstract

1. Introduction

All cells share the ability to secrete extracellular vesicles (EVs), phospholipid bilayer membrane structures enclosing a portion of their own cytoplasm [1]. These vesicles are able to transfer their cargo of biomolecules, including proteins, lipids, and nucleic acids, triggering alterations in recipient cells [2,3,4]. For this reason, EVs play essential roles in intercellular communication, being associated with multiple physiological and pathological processes.
Considering their natural ability to shuttle messages between cells, researchers and companies have shifted their attention for the promising use of EVs as natural therapeutic systems. In fact, EVs are able to mediate some of the therapeutic effects from their cells of origin [5,6]. Therefore, EVs could be used in the substitution of their cell of origin, as a cell-free therapy triggering an equivalent therapeutic effect, but avoiding the complexity and safety concerns associated with cellular therapies.
Furthermore, EVs have been explored as drug delivery vehicles for the treatment of several diseases, through the loading of therapeutic molecules and shuttling them to target cells and tissues. Their small size and resemblance to the cell membrane makes EVs ideal candidates to cross biological barriers, while offering high biocompatibility to target cells [7,8,9]. Some EVs show inherent targeting ability and display tropism towards particular cells or tissues [10,11,12]. Due to their biological origin, EVs present generally low immunogenicity and toxicity, allowing to overcome safety issues associated with synthetic nanocarriers [13,14,15,16]. Thus, EVs have emerged as promising drug delivery systems (DDS), presenting advantageous features that may allow them to outperform synthetic nanocarriers. In fact, EVs have been recently described to deliver functional RNA more efficiently than state-of-the-art synthetic RNA nanocarriers [17].
In addition to their ability to pack therapeutic molecules, EVs can be further engineered to enhance their targeting capacity towards specific tissues. This has been achieved either by the genetic engineering of parental cells in order to express a targeting moiety fused to an EV transmembrane protein, or by anchoring targeting ligands to the surface of EVs after their isolation. Here, we review recent and notable developments in the field, allowing the engineering of EVs to enhance their therapeutic potential, elucidating the pros and cons of each strategy. This is preceded by a contextualization of the biology of EVs and their roles in major pathophysiological processes. We also review the most recent strategies employed to address the critical issues in upstream and downstream processing of EVs to their scalable manufacturing. These platforms are expected to support the large numbers and high purity of EVs needed as the field moves more intensively from pre-clinical to clinical studies towards the validation of EVs as key players in therapeutic applications.

2. EV Biology

The identification of EVs can be traced back to as early as 1946, when they were described as pro-coagulant particles in plasma [18], and later in the 1960s described as “platelet-dust” and as cartilage matrix vesicles associated with bone calcification [19,20]. A major breakthrough occurred in 1983, when a mechanism for the release of transferrin receptors from maturing red blood cells through vesicles was described [21,22]. These vesicles were later named “exosomes” in 1987 [23].
For some time, EVs were only considered to be a means to remove unwanted material from the cell. However, the field of EVs was revolutionized in 1996 when exosomes were shown to play a role in antigen presentation, opening an entirely new discussion that EVs might play a role in the transfer of biological information between cells [2]. This was later consolidated in 2006 and 2007, when EVs were shown to contain RNA (miRNA and mRNA) that could be delivered to recipient cells, changing their behavior [3,4]. Since then, EVs have emerged as relevant players in intercellular communication, mainly through their ability to transfer a cargo of biomolecules, including proteins, lipids, and nucleic acids, which trigger alterations on recipient cells (Figure 1).
The term EVs was proposed in 2011 to define all the different types of extracellular membrane structures [1]. However, EVs actually comprise a highly heterogeneous group. Depending on their biogenesis, EVs are broadly categorized either as exosomes, or microvesicles [9]. Exosomes are generated through the endosomal pathway [9,24]. Endocytosis at the cell membrane leads to the formation of early endosomes. During endosome maturation into late endosomes, there is an inward budding of endosomes resulting in the accumulation of intraluminal vesicles (ILV), which leads to the formation of multivesicular bodies (MVB), also named multivesicular endosomes (MVE). Upon the fusion of MVE with cell membrane, ILV are released to the extracellular space originating exosomes, which generally display 50–150 nm in diameter. Microvesicles are formed by the outward budding of the plasma membrane, ranging in size from 50 nm to 1 µm in diameter, or even higher [9]. Exosomes and microvesicles show overlapping properties, such as size, density, and molecular composition, making it challenging to distinguish different co-isolated EV subpopulations [25]. Additionally, the composition of EVs may differ among different secreting cells.
The complexity of EVs is further increased when we consider other structures, such as apoptotic bodies released from cells undergoing apoptosis, which can span over a large size range (from 100 nm to 5 µm in diameter) [25], or the recently identified mitovesicles from mitochondrial origin [26]. Interestingly, recent studies revealed that EVs could even transport whole organelles inside them, such as mitochondria, thus enabling their functional trafficking between cells [27,28]. Both the trafficking of mitovesicles, or whole mitochondria shuttled inside EVs, seem to be associated with the processes of mitochondrial regulation and dysfunction, namely in neurodegenerative disorders. Moreover, EVs display physical characteristics (i.e., size and density) similar to other secreted non-vesicular nanoparticles, such as lipoproteins of various densities [29] and the recently identified exomeres [30].

2.1. EV Biogenesis

Cargo incorporated in exosomes originate from endocytosis at the plasma membrane or are directly targeted to early endosomes via the biosynthetic pathway, from the trans-Golgi network [9]. These sorting processes are regulated by various Rab GTPases. The formation of ILV can be regulated by the endosomal sorting complex required for transport (ESCRT), a family of proteins that associate in a stepwise manner at the membrane of MVE [9,25]. Firstly, ESCRT-0 and ESCRT-I subunits cluster membrane-associated cargo in microdomains of the limiting membrane of MVE. The tumor susceptibility gene 101 protein (TSG101) is one of the main ESCRT-I components, being used as an EV protein marker. This is followed by ESCRT-II-mediated recruitment of ESCRT-III that performs budding and fission of this microdomain into the MVE lumen.
Although ESCRT-III is required for fission of ILVs, cargo clustering and membrane budding can be either ESCRT-dependent or ESCRT-independent [9]. The latter can rely on syntenin and the ESCRT accessory protein ALG-2 interacting protein X (ALIX), which links cargo to ESCRT-III [9,31]. ESCRT-independent biogenesis is aided by lipids such as ceramide, which allows the generation of membrane subdomains imposing a spontaneous curvature on the membranes [32,33]. Additionally, proteins of the tetraspanin family (e.g., CD63, CD81 and CD9) have been shown to regulate ESCRT-independent cargo sorting to exosomes [9,34,35]. Some tetraspanins also show the potential to form microdomains and induce budding.
Mature MVE can follow a degradative route by fusion with lysosomes or autophagosomes. Alternatively, MVE are transported along microtubules to the plasma membrane. At this stage, MVE fuse with the plasma membrane leading to exosome release in a process mediated by Rab GTPases (e.g., Rab27A/B, Rab35), actin and SNARE (soluble N-ethylmaleimide-sensitive fusion attachment protein receptor) proteins [9,25,36,37,38].
Microvesicle biogenesis shares several mechanisms common to exosome biogenesis. This includes the formation of microdomains (in this case in the plasma membrane) where specific lipids and cargo are clustered, as well as a similar role of ESCRT machinery and ceramide in vesicle formation [9]. However, microdomain formation is followed by the translocation of lipids between leaflets of the plasma membrane, a process unique to microvesicle formation. This process is mediated by Ca2+-dependent enzymes (e.g., translocases, scramblases and calpain), rearranging the asymmetry of membrane phospholipids in a way that causes the physical bending of the membrane, favoring membrane budding [25,39]. The most significant examples are the exposition of phosphatidylserine (PS) and phosphatidylethanolamine (PE) from the inner leaflet to the cell surface.

2.2. EV Interaction with Recipient Cells

After being released into the extracellular space, EVs are able to interact with cells either close-by or far away, triggering phenotypic changes in these cells. EV binding to recipient cells can be mediated by tetraspanins, integrins, proteoglycans, lectins, lipids (e.g., PS) and extracellular matrix components (e.g., fibronectin and laminin) [9]. After binding to a recipient cell, EVs can follow multiple routes to deliver their message. EVs can elicit changes in recipient cells by simply binding to specific surface receptors, triggering signaling pathways (e.g., antigen presentation), but without delivering any EV cargo [9].
EVs can also be internalized through multiple EV uptake pathways [40] (Figure 2). EVs can undergo clathrin-mediated endocytosis, through the formation of a clathrin coat in a portion of cell membrane surrounding the EV to be internalized [41]. This clathrin coat promotes membrane deformation, which results in membrane invagination and formation of a bud that surrounds the EV that then pinches off, separating itself from the membrane [42]. Once in the cytosol, this internalized vesicle undergoes clathrin un-coating. EVs can also be internalized by clathrin-independent endocytosis, such as caveolin-mediated endocytosis, involving the formation of cave-like invaginations in the plasma membrane named caveolae, which become internalized into the cell (similarly to clathrin-mediated endocytosis) [40,43]. Caveolin-1 is required for the formation of caveolae, which are also rich in cholesterol and sphingolipids. Clathrin-independent endocytosis may also occur via lipid rafts [44]. These plasma membrane microdomains have altered phospholipid composition, being more tightly packed and consequently less fluid, but float freely in the plasma membrane. Lipid rafts can be found in invaginations formed by caveolin-1 or in planar regions of the plasma membrane associated with flotillins [40]. However, lipid raft-mediated endocytosis of EVs seems to be caveolae-independent [41,44].
Alternatively, EV uptake can happen through non-specific processes such as phagocytosis and macropinocytosis [40,45,46]. Phagocytosis involves the formation of invaginations surrounding material to be internalized, with or without the formation of enveloping membrane extensions [47]. Although this process is generally used to internalize larger particles, such as cells, it has been observed to be used to take up EVs. Phosphatidylinositol-3-kinase (PI3K) plays an important role in this process [48]. In macropinocytosis, membrane extensions are formed surrounding a portion of extracellular fluid and fuse back with the plasma membrane internalizing that portion of extracellular content [40,47]. This requires Na+/H+ exchanger activity and is dependent on actin, cholesterol and the rac 1 GTPase. Both of these uptake mechanisms seem to be triggered (at least partially) by PS present on the outer leaflet of EV membranes.
Internalized EVs follow the endosomal pathway, eventually reaching MVE. At this stage, EVs can follow different fates [9,25]. They can be recycled back to the plasma membrane and released to the extracellular space. MVE can fuse with the lysosome leading to the degradation of the contents of internalized EVs, which can still be a relevant source of metabolites for the host cell. Alternatively, EVs may undergo endosomal escape, through back fusion with the limiting membrane of MVE, releasing their contents to the cytoplasm of the recipient cell. EVs may also fuse directly with the plasma membrane, releasing their cargo directly into the cytosol of the recipient cell. Intraluminal material released by internalized EVs includes nucleic acids (miRNA, mRNA), proteins and lipids, which are able to trigger alterations in the recipient cell. However, little is known about how EV cargo is unpackaged and delivered to the designated site of action, either the cytoplasm or the nucleus. Intracellular delivery routes are being investigated and may include direct transfer into the endoplasmic reticulum [49] or the nucleus [50,51].

3. EVs in Intercellular Communication

3.1. Physiological Roles of EVs

Given their ability to elicit changes in recipient cells, EVs have been implicated in numerous physiological processes [52,53]. In fertilization, EVs secreted from the egg promote sperm-egg fusion in a tetraspanin CD9-dependent process that was observed in mice [54]. Later, microvesicles released by early embryo cells promote trophoblast migration and implantation in the uterus through the JNK and FAK pathways, which are activated by microvesicle cargo proteins laminin and fibronectin [55]. EVs have been implicated in development by carrying key morphogen molecules, such as Wnt proteins (e.g., Wingless) and Sonic Hedgehog [56,57,58]. Mating behavior can also be altered by EVs, since exosomes secreted by the reproductive glands of male Drosophila melanogaster interact with female reproductive tract epithelium and inhibit the re-mating of females [59].
EVs are important in the nervous system, being secreted by neurons and glial cells alike to mediate intercellular communication [60]. Neuron-derived EVs have multiple relevant roles at synapses, such as promoting synaptic growth at the neuromuscular junction and regulating postsynaptic retrograde signaling [61,62,63]. Oligodendrocyte-derived EVs are able to promote neuronal viability and increase neuron firing rate [64]. EVs also play a role in the peripheral nervous system, where Schwann cells are able to secrete EVs to promote axon regeneration [65].
EVs play a relevant part in the regulation of immune responses through exchange among multiple types of immune cells. EVs play a crucial role in major histocompatibility complex (MHC) class II antigen presentation, since dendritic cells (DC) secrete exosomes carrying peptide-containing MHCII that stimulate naïve CD4+ T cells [66,67]. DC-derived exosomes were also able to differentiate T helper cells toward a T helper type 1 (Th1) phenotype and enhance immunogenicity in vivo [68]. In the opposite direction, T cell-derived exosomes are able to transport miRNA to antigen-presenting cells, modulating their mRNA expression levels [69]. Exosomes were also found to transfer miRNA between DC in vivo, modulating gene expression in the recipient cell [70]. Another study revealed that exosome-mediated miRNA transfer from T regulatory (Treg) cells to Th1 cells was able to reduce inflammatory responses of recipient Th1 cells [71].
Physiological tissue regeneration processes are also supported by EVs. Endothelial cell-derived EVs were able to reduce atherosclerotic lesion formation when delivered to smooth muscle cells [72]. The authors of this study observed that alterations in endothelial cells, which were previously described to be triggered by blood flow-induced shear stress, led to the enrichment of endothelial cell-derived EVs in specific miRNA molecules that had atheroprotective effects after delivery to smooth muscle cells. In a kidney injury model, injured epithelial cells secreted exosomes that activated fibroblasts to initiate tissue regenerative responses and fibrosis mediated by exosomal transforming growth factor (TGF)-β1 mRNA [73].

3.2. EVs in Pathological Processes

In addition to their relevant role under normal physiological conditions, EVs have been associated with multiple pathological processes [52,53,74]. Numerous studies reveal tumor-derived EVs as relevant mediators of intercellular communication within the tumor microenvironment (TME), which is composed of multiple non-tumorigenic cells able to collectively support tumor growth and progression, such as endothelial cells, fibroblasts, and immune cells, among others [75].
Under hypoxic conditions (1% O2), glioblastoma multiforme (GBM)-derived exosomes amplified the activation of ERK1/2 MAPK, PI3K/Akt and FAK pathways in endothelial cells, compared to normoxic conditions, resulting in increased endothelial cell sprouting [76]. GBM-derived EVs were also found to skew monocyte-to-macrophage differentiation to a tumor-supportive M2-type macrophage phenotype [77]. Conversely, lymph node macrophages were able to suppress tumor growth by absorbing tumor-derived EVs and preventing their interaction with pro-tumorigenic B cells [78]. Tumor-derived EVs are also able to suppress anti-tumor adaptive immunity. Tumor-derived EVs induced apoptosis of CD8+ T cells and were also able to alter the differentiation of CD4+ T cells into a state that suppresses cytotoxic T cell activity, contributing to tumor escape from the immune system [79].
Tumor-derived EVs also promote tumor invasion and metastasis. They help establishing pre-metastatic niches, by interacting with normal cells at the metastatic sites. Melanoma-derived exosomes recruited bone marrow progenitor cells to future sites of metastasis and re-educated them toward a vasculogenesis supporting phenotype, enhancing tumor invasion and metastasis in vivo [80]. This re-education effect was mediated by a tyrosine kinase receptor differentially expressed in exosomes from highly metastatic melanoma cells compared to less aggressive ones.
In another study, exosomes from pancreatic cancer cells induced the formation of pre-metastatic niche in the liver of mice [81]. These exosomes transferred migration inhibitory factor (MIF) to Kupffer cells in the liver that secreted TGF-β, which subsequently increased fibronectin production by hepatic stellate cells. Fibronectin enhanced the recruitment and retention of bone marrow-derived macrophages, establishing an environment favorable for metastasis [81].
Remarkably, the formation of pre-metastatic niches was found to be promoted by exosomes from different tumor types that targeted specific organs, depending on integrins displayed on their membrane [82]. Exosomes expressing integrins α6β4 and α6β1 were able to bind specifically to fibroblasts and epithelial cells in the lung, mediating lung tropism, while exosomes containing integrin αvβ5 bound to Kupffer cells, leading to liver tropism.
EVs derived from non-tumorigenic cells can also support tumor growth under specific circumstances. For instance, exosomes from astrocytes were found to support tumor growth in brain metastatic breast cancer in vivo. Astrocyte-derived exosomes mediated miRNA transfer to metastatic tumor cells, reducing the levels of a target mRNA encoding for the tumor suppressor PTEN [83]. Decreased PTEN levels triggered and increased secretion of CCL2 chemokine by metastatic tumor cells, resulting in recruitment of IBA1-expressing myeloid cells that enhanced proliferation and reduced apoptosis of metastatic tumor cells.
In addition to their roles as mediators in tumor progression, tumor-derived EVs also provide a way to eliminate chemotherapeutic agents from cancer cells, enabling chemotherapy resistance. The microvesicle-mediated release of gemcitabine was identified as a key factor for resistance to this drug in human pancreatic cancer cells, both in vitro and in vivo [84]. Moreover, just as tumor-stromal interactions mediated by EVs play a relevant role in tumor progression, stromal cell-derived EVs can also mediate resistance to therapy. Exosomes derived from stromal cells were able to mediate miRNA transfer to ovarian cancer cells, increasing their chemoresistance to paclitaxel [85].
EVs are also involved in the cell-cell transport of pathogenic proteins associated with neurodegenerative diseases, such as the prion protein (PrP) abnormal isoform PrPSC in prion disease, β-amyloid in Alzheimer’s disease and α-synuclein in Parkinson’s disease [86,87]. However, the relevance of EV-mediated versus EV-independent spread and propagation of these proteins in disease progression is still unclear. Contrastingly, some studies have described the natural beneficial effects of EVs in these pathologies, namely in the clearance of β-amyloid peptides [88,89].
In cardiovascular diseases, EVs have also been found to mediate the cross-talk between different cell types in the heart, with implications in disease progression. The secretion of cardiac fibroblast-derived exosomes triggered gene expression alterations in cardiomyocytes, leading to increased pathological cardiac hypertrophy, which contributes to heart failure [90]. In another study, macrophage-derived exosomes transferred miRNA to cardiac fibroblasts, suppressing fibroblast proliferation and promoting fibroblast inflammation during cardiac injury in mice [91].
In multiple infectious diseases, viruses are able to take advantage of EVs to transfer their genetic material between infected and non-infected cells in the substitution of direct interaction between viruses and target cells [92]. For example, exosomes derived from human hepatoma cells infected with hepatitis C virus were able to transmit the infection to naïve hepatoma cells [93]. Moreover, this exosome-mediated transmission was partially resistant to antibody neutralization.
Considering the numerous roles of EVs in disease progression, EVs have been extensively studied as novel biomarkers for disease [94,95,96], as well as targets for new therapeutic strategies [97,98].

4. EVs as Reconfigurable Natural Therapeutic Systems

4.1. EVs as Intrinsically Therapeutic Agents

Given their ability to participate in intercellular communication, conveying messages from their cells of origin to target recipient cells, EVs have innate therapeutic potential, particularly interesting for tissue regeneration. EVs are able to mediate some of the therapeutic effects from their cells of origin by carrying lipids, proteins, and genetic material (mRNA and miRNA), and transferring this cargo to target cells, or by triggering signaling pathways through cell surface interactions.
EVs derived from stem and progenitor cells have gained particular interest due to numerous therapeutic properties attributed to them, which include: immunomodulatory capacity (mainly by reducing inflammation) [99,100,101,102]; suppressing apoptosis and stimulating cell proliferation [103,104]; promoting angiogenesis [105,106]; stimulating wound repair [107,108]; and recruiting and reprograming cells for tissue regeneration [3]. Among the most studied EV-secreting cells with therapeutic properties we can find mesenchymal stromal cells (MSC), embryonic stem cells, induced pluripotent stem cells (iPSC), cardiac progenitor cells, and DC [109].
In particular, a growing body of evidence indicates that many of the therapeutic features of MSC, currently under study in numerous clinical trials, are exerted in a paracrine manner and mediated by EVs. The paracrine activity of MSC was initially observed in mice and pig models of myocardial infarction, where conditioned medium from MSC cultures limited infarct size and improved heart function [110,111,112,113]. This was followed by similar evidence supporting the paracrine activity of MSC in other organs [114]. In subsequent studies, EVs secreted by MSC were described as the mediators of these paracrine trophic activities, reducing myocardial ischemia/reperfusion injury [5] and also allowing improved recovery from acute kidney injury [6,115] in mice.
Numerous studies followed, reporting the different therapeutic activities of EVs derived from MSC and other cells. MSC-derived EVs (MSC-EVs) allowed improved recovery from stroke in mice, by promoting neuronal survival and angioneurogenesis [116]. Human bone marrow MSC-EVs allowed a better recovery from traumatic brain injury in mice [117] and improved the recovery from acute spinal cord injury in rats [100]. Attenuated inflammation upon EV treatment supported a better recovery in both studies.
In the context of MSC-EVs, vesicles obtained from different tissue sources showed therapeutic potential against hepatic indications. Human umbilical cord matrix MSC-EVs ameliorated liver fibrosis in mice by inactivating the TGF-β1/Smad signaling pathway, and inhibiting epithelial-to-mesenchymal transition (EMT) in hepatocytes [118]. Bone marrow MSC-EVs reduced hepatic injury in mice, improving their survival [119]. Reduction in hepatocyte apoptosis was proposed to be mediated by the lncRNA Y-RNA-1 carried by EVs. Human umbilical cord matrix MSC-EVs also inhibited pulmonary infiltration of macrophages and suppressed the production of pro-inflammatory and pro-proliferative factors in a murine model of pulmonary hypertension [99].
The large number of preclinical studies using EVs has already been translated into a few clinical trials. The safety and efficacy of MSC-EVs have been evaluated in clinical trials for the treatment of type 1 diabetes (NCT02138331), macular holes (NCT03437759) and chronic kidney disease, with positive safety and efficacy results in the latter [120].
More recently, MSC-EVs have been proposed for the treatment of coronavirus disease-19 (COVID-19), aiming to reduce dysregulated immune responses and the cytokine storm associated with respiratory pathological states of this disease [121]. The rationale for using MSC-EVs is based on previously mentioned observations of inflammatory attenuation in several pathological conditions, and supported by studies in relevant lung disease models, including lung injury [122,123]. However, the mechanisms behind the beneficial effects of EVs are not fully elucidated yet. Some phase I/II and even phase III clinical trials have already been registered in different countries for the use of EVs for the treatment of COVID-19, most of them using EVs derived from MSC either administered intravenously (e.g., NCT04798716, NCT05354141) or by inhalation (e.g., NCT04602442, NCT04276987) (clinicaltrials.gov, accessed on 15 September 2022 using the search term “(extracellular vesicles OR exosomes) AND COVID-19”).

4.2. EVs as Drug Delivery Systems

In addition to their use as innate therapeutic products mainly in the context of regenerative medicine, EVs are also promising vehicles for drug delivery to treat numerous conditions. Given their small size and the ability to shuttle messages to other cells in virtually any site in the organism eliciting a functional response, EVs can be regarded as nature’s nanocarriers. In fact, EVs comprise numerous features that make them appealing for the development of novel DDS, even outperforming synthetic nanocarriers (e.g., lipid nanoparticles [17]) or viral delivery platforms (e.g., adeno-associated virus (AAVs) [124]) in certain aspects.
By using EVs, we can take advantage of endogenous cellular machinery to produce the desired therapeutic cargo and sorting it inside EVs. Additionally, EVs have the ability to overcome biological barriers, namely tissue barriers (e.g., blood-brain barrier (BBB)), cellular barriers (by different EV uptake mechanisms) and intracellular barriers, exerting functional effects on target cells [7,125,126]. Due to their biological origin, EVs are generally low immunogenic and are not toxic, contrarily to some synthetic nanocarriers, with different studies revealing minimal hepatoxicity and overall toxicity of human-derived EVs both in vitro and in mouse models [13,15,127,128]. On the other hand, EVs are non-replicative and non-mutagenic, relieving some of the safety concerns associated with cell therapies. Therefore, EVs lie in a sweet spot between synthetic nanocarriers and cell therapies, presenting exciting opportunities for developing next-generation DDS with increased efficacy and lower side-effects.
EVs are able to carry different cargos, including small molecules such as the natural compound curcumin or chemotherapeutic drugs (e.g., doxorubicin and paclitaxel), as well as delivering proteins and different RNA molecules (e.g., siRNA, miRNA, and mRNA) [7,129,130,131,132,133]. In this way, EVs can deliver therapeutic molecules in a more efficient and selective manner to target diseased cells and tissues, while minimizing their side effects, as well as protecting cargo from degradation (particularly relevant for RNA molecules). For example, doxorubicin-loaded EVs showed similar cytotoxicity to the free drug in both in vitro and in vivo models of breast cancer, but with reduced cardiotoxicity [130].
Therapeutic cargo can be loaded into EVs by two different strategies, either exogenously, by inserting cargo directly into EVs after EV production and isolation, or endogenously, where therapeutic cargo is loaded into EVs at the moment of EV biogenesis [126,134]. Several techniques have been applied in order to accomplish the exogenous loading of EVs including direct incubation, electroporation, sonication, saponin, freeze/thaw cycles, or extrusion [134]. For example, curcumin was loaded into EVs through direct incubation (e.g., mixing at 22 °C for 5 min) in several studies, yielding diverse positive therapeutic outcomes such as improved bioavailability and anti-inflammatory effect of this drug in a mouse model of inflammation [129], as well as enhanced tumor growth inhibition both in vitro and in vivo, compared to free curcumin [135]. Electroporation has been applied in numerous studies, for example to load EVs with therapeutic siRNA with positive outcomes in mouse models of Alzheimer’s disease [7] and pancreatic ductal adenocarcinoma (PDAC) [14], or for loading small molecules such as doxorubicin, with improved outcomes in different in vivo cancer models [130,136,137].
A few studies compared the efficiency of the exogenous loading of EVs using different techniques. The hydrophobicity of small molecules can influence their loading into EVs, since hydrophobic porphyrins were loaded in EVs from different cells simply by direct incubation, while hydrophilic porphyrins benefited significantly from active loading techniques such as electroporation, extrusion, and especially saponin treatment [138]. In another study, exosomes were loaded with catalase and tested in in vitro and in vivo models of Parkinson’s disease [139]. Different loading techniques were tested (incubation, saponin treatment, freeze/thaw cycles, sonication, and extrusion), revealing improved loading efficiency, sustained release, and catalase preservation upon active loading, especially using sonication, extrusion, or saponin treatment. Sonication also improved exosome loading with paclitaxel and yielded positive therapeutic outcome in multidrug resistant (MDR) cancer cells [131]. However, active loading techniques induce the temporary disruption of the EV membrane that can lead to a loss of EV content, or altered morphology, and may also induce the aggregation of EVs or their cargo, as previously reported [140,141].
Additionally, EVs can be fused with liposomes previously synthesized to carry therapeutic cargo, thereby creating EV-liposome hybrids carrying this cargo [142]. MSC-EVs fused with liposomes loaded with a chemotherapeutic compound increased the drug delivery efficiency when compared with the free drug or the drug-loaded liposome in cancer in vitro models [143]. EVs may even be combined with biomaterials to improve their functional delivery to target tissues. For example, the injection of hydrogels of multiple formulations carrying EVs allowed sustained EV release over time, enhanced local EV retention and improved therapeutic outcomes in different pathological models, both in vitro and in vivo [144].
The endogenous loading of EVs can be achieved by taking advantage of the natural sorting machinery of cells for the production and/or loading of cargo into EVs. Cells can be loaded with a cargo by direct incubation, which is then sorted and released inside EVs. MSC incubated with paclitaxel incorporated this chemotherapeutic agent and released it inside EVs [145]. These paclitaxel-loaded MSC-EVs were able to inhibit tumor cell proliferation in vitro.
Alternatively, endogenous loading of EVs can be achieved by the genetic modification of parental cells to express desired RNA molecules or proteins, which will then be loaded into EVs. For example, adipose tissue-derived MSC were genetically modified to express miR-122, previously reported to reduce drug resistance in hepatocellular carcinoma [146]. Modified MSC secreted EVs packaging miR-122, which increased antitumor efficacy of chemotherapy on hepatocellular carcinoma in vivo.
A new system was developed to achieve protein loading into EVs using an optically reversible protein-protein interaction module [147]. The authors used a photoreceptor cryptochrome 2 (CRY2) and the CRY-interacting protein (CIBN), which bind under blue light illumination. CRY2 was fused with a cargo protein and CIBN was conjugated with the EV transmembrane protein CD9. As a result, the transient docking of CRY2-conjugated cargo proteins with CD9-conjugated CIBN was observed in the generated exosomes in the presence of blue light. When the blue light was removed, the proteins detached and the cargos were released into the intraluminal space of exosomes, allowing cargo proteins to be delivered to recipient cells both in vitro and in vivo. This strategy was used for the delivery of super-repressor IκBα to relieve sepsis-associated organ damage and reducing mortality in mice [148].
In addition to their unique drug loading abilities, EVs also exhibit intrinsic targeting properties that can be valuable for drug delivery, since the protein and lipid composition of EVs can influence cell/organ tropism [12]. As previously mentioned, depending on the integrins displayed on their membrane, EVs can show tropism towards lung or liver in pre-metastatic niche formation [10]. Another example is the involvement of PS in EV recognition and uptake by macrophages [11].
Still, EVs can be engineered in order to improve specificity to target cells. Akin to the techniques used for endogenous drug loading, parental cells can be genetically modified to express a targeting moiety fused to an EV transmembrane protein. The first example relied on the fusion of lysosome-associated membrane protein 2 (Lamp2b), abundant on the surface of EVs, with the rabies viral glycoprotein (RVG) peptide that binds to the acetylcholine receptor [7,149]. In the context of Alzheimer’s disease, this strategy allowed EVs to target neurons, oligodendrocytes, and microglia, and functionally deliver electroporated siRNA to the brain in mice. The fusion of targeting moieties with Lamp2b has been used to endow EVs with targeting capacity in several studies, including an αv integrin-targeting iRGD peptide to target tumor cells and tumor-associated vascular endothelium for doxorubicin delivery [150], and a fragment of interleukin 3 (IL3) to target IL3 receptors on chronic myeloid leukemia (CML) cells for delivery of a chemotherapeutic agent or siRNA [151].
Different types of targeting moieties and transmembrane proteins have been used (Table 1). For example, EV-producing cells have been modified to express recombinant anti-epidermal growth factor receptor (EGFR) nanobodies fused to glycosylphosphatidylinositol (GPI)-anchoring peptides [152]. Since EVs are enriched in GPI, EVs were enriched in GPI linked nanobodies displayed on their surface. This allowed EVs to target specifically EGFR-expressing tumor cells.
Alternatively, targeting ligands can be exposed on the surface of EVs after EV isolation, avoiding the challenging genetic engineering of producer cells. Several different strategies have been applied in this context (Table 2). A recombinant protein was developed, consisting of an anti-EGFR nanobody fused to the C1C2 domain of lactadherin, which binds to PS present on the surface of EVs, directing these modified EVs to EGFR-positive cancer cells [156]. In another study, EVs were modified in order to target neuropilin-1 (NRP-1), which is a transmembrane glycoprotein overexpressed in glioma cells and the tumor vascular endothelium [157]. For this purpose, an NRP-1-targeting peptide was conjugated with the surface of EVs by click chemistry.
The promising results obtained in preclinical studies using EV-based DDS for numerous clinical indications, prompted their use in pioneering clinical trials over the past few years [109]. Plant-derived EVs loaded with curcumin have been used for the treatment of colon cancer (NCT01294072). Tumor-derived EVs loaded with chemotherapeutic agents (NCT01854866) and EVs derived from malignant pleural effusion loaded with methotrexate (NCT02657460) are being studied for the treatment of malignant ascites and pleural effusion. Other clinical trials are testing the use of allogeneic MSC-EVs enriched in miR-124 for the treatment of acute ischemic stroke (NCT03384433), or loaded with KRASG12D siRNA for the treatment of metastatic pancreatic cancer (NCT03608631).

5. Production and Isolation of EVs

5.1. Upstream Processing

One of the main challenges faced in the field of EVs is the low yield obtained during manufacturing [166]. This renders basic scientific research, pre-clinical studies, and clinical trials more challenging, leading to a significant delay in the progress of the field in order to harness the full potential of EVs in clinical settings [167]. The whole bioprocessing pipeline, from upstream to downstream processing, needs to be considered for EV manufacturing optimization. Upstream processing comprises cell culture leading to EV production, thus requiring the appropriate choice of both the cell culture medium used and the culture vessel, which can range from small vessels such as T-flasks up to 500 L bioreactors, depending on the chosen approach. This step influences the characteristics of the final product and therefore should be carefully selected, while having in mind a specific application.
Progress is already being made towards optimizing the upstream conditions of EV production through the adaptation of already well-established technologies in other fields to produce functional and safe EVs in a cost-effective way. The scaling-up or scaling-out of such platforms is expected to decrease the costs associated with EV manufacturing [168]. The simplest scaling option is to adopt the use of multi-layer flasks (e.g., hyperflasks, Cellstacks, Cell Factories) instead of the use of conventional T-flasks [168]. These systems present several layers for cell growth, decreasing the manipulation time needed, in comparison to conventional T-flasks and reducing lab footprint. For example, hyperflasks were used for the cultivation of bone marrow-derived MSC and the EVs isolated from the conditioned medium (i.e., containing EVs) were shown to significantly suppress the symptoms of graft vs. host disease (GvHD) in a murine model [169]. However, culture monitoring is challenging when using these planar systems, which do not lead to cost savings in the whole process, namely in what concerns the quantity of cell culture medium used [170].
Bioreactors represent promising scale-up or scale-out options for EV manufacturing, which, under optimized conditions, can generate higher cell and EV yields in a more cost-effective way (e.g., lower need for consumables such as culture medium), compared to static systems [168]. In fact, some bioreactor systems were already explored for EV production.
Hollow-fiber bioreactors allow adherent cells to grow attached to the fibers composed of a permeable membrane, while the cell culture medium flows across the membrane. In a recent study, it was observed that umbilical cord matrix MSC-EVs produced in a hollow-fiber system displayed enhanced osteochondral regeneration activity when compared to 2D culture flasks, leading to superior repair activity of cartilage defects in vivo [170]. Another study revealed that umbilical cord matrix-derived MSC expanded in hollow-fiber bioreactors produced EVs with enhanced renoprotective effects in comparison to EVs produced by cells grown in conventional 2D cultures [171].
Stirred-tank bioreactors can be used as a scalable option that provides homogenization of the cell culture medium through mechanical stirring. Microcarrier technology can be used in combination with these bioreactors for expanding adherent cells and providing a high surface-area-to-volume ratio [166]. Similar to hollow-fibers, stirred-tank bioreactors can enhance EV productivity as high as 100-fold in comparison to conventional 2D culture flasks [172].
In the context of scalable platforms available for EV production, our group recently employed a single-use Vertical-WheelTM bioreactor for the production of human MSC-EVs [173]. This system provides a gentler and more homogenous mixing in comparison to the stirred-tank configuration, while also making use of microcarriers for adherent cell culture [174]. Our work revealed an average yield improvement of 6-fold referred to the final EV concentration in the culture medium, in comparison to culture flasks [173].
The use of different culture media during conditioning periods for EV obtention also impacts EV productivity and can potentially render the process more efficient. For instance, the use of non-supplemented OptiMEM (i.e., optimized formulation of Eagle’s Minimal Essential Medium (MEM)) led to an increase in the production of HEK293T-derived EVs, in comparison to OptiMEM supplemented with 10% fetal bovine serum (FBS), DMEM (i.e., Dulbecco’s Modified Eagle’s Medium) supplemented with 10% FBS, or DMEM supplemented with 10% EV-depleted FBS, among others [175].
In order to circumvent the low yields associated with EV production from cell culture supernatants (i.e., conditioned media), the production of EV-like particles has also been employed. This is achieved through the complete disruption of the cellular membrane leading to the release of self-assembled particles, reaching production yields that can be as high as 100-fold compared to normal EV release [176]. The most common approach to achieve this is extrusion, whereby cells are filtered through sequentially smaller pores [168].
Growing adherent cells as 3D cell aggregates instead of monolayers is another viable option to increase EV production yields. The formation of 3D MSC aggregates provides a microenvironment that better mimics in vivo conditions leading to the maintenance of MSC phenotype and innate MSC properties, which could explain the higher yields attained with this strategy. Moreover, 3D scaffolds into which cells can adhere and proliferate can also be used to promote the production of EVs and even to enhance their functional activity. This was demonstrated using bone marrow-derived MSC seeded into collagen scaffolds, which produced 2-fold more EVs compared to 2D culture conditions and enhanced the regenerative capacity of these EVs in traumatic brain injury mouse models [177].
Importantly, clinical trials employing bioreactors for EV production are already ongoing. Recently, Codiak Biosciences developed exoSTING, a therapeutic EV-based product in which EVs are loaded with stimulators of the interferon gene for the treatment of multiple cancers [178]. This strategy is currently under clinical testing in a phase 1/2 trial (NCT04592484).
Of note, there is still a gap in our knowledge concerning how the upstream conditions influence the characteristics and the quantity of produced EVs. For instance, in the context of the manufacturing of monoclonal antibodies, the optimization of upstream conditions was able to lead to improvements of 10–100-fold in product titers over the years [166]. Therefore, by adjusting operational parameters in the process such as the composition of culture media, oxygen concentration, materials used, as well as the cell passage, level of confluency, and viability, one can expect substantial improvements in EV yields.

5.2. Downstream Processing

The downstream processing of EVs, which comprises all the steps following upstream processing to recover and isolate EVs from cell culture conditioned medium and eventually purify them, also deserves special attention considering the particular characteristics of EVs. Over the last years, the emergence of EVs as promising tools for the development of new therapeutic options, as well as valuable structures to understand organismal pathophysiological mechanisms, has prompted the development of EV isolation methods.
However, currently available isolation methods still present considerable limitations, including poor standardization, the alteration of physicochemical properties of the isolated EV population, and the modification of EV cargo profiles. For instance, the miRNA profile in exosomes obtained from blood serum was reported to be different depending on the isolation method used, when comparing a precipitation-based isolation method with ultracentrifugation [179]. Likewise, the purity of EV samples largely depends on the isolation method used, and a compromise between the purity of a given sample and the amount of recovered EVs has to be made [180,181].
Isolation methods available can be categorized based on the principle used for the isolation of EVs from other particles present in a body fluid or cell culture conditioned medium. These methods can be generally divided into ultracentrifugation-based methods, precipitation, size-based, and microfluidics, among others.

5.2.1. Ultracentrifugation-Based Methods

Ultracentrifugation methods can be divided into two distinct techniques: differential ultracentrifugation, and density gradient centrifugation. Differential ultracentrifugation is based on differences in the size and density of particles. It subjects the samples to multiple low-speed centrifugations to remove dead cells and debris, which correspond to the larger biological particles present in the sample. The low centrifugation speed periods are followed by high-speed centrifugations of 100,000× g (or more) in order to pellet the smaller EVs and allow the collection of these particles [182,183]. Differential ultracentrifugation is the most employed primary method of EV isolation, with more than 75% of participants reporting its use in a worldwide survey performed by the International Society for Extracellular Vesicles (ISEV) in 2020 [184]. It can be considered a method with intermediate recovery and specificity [185] and is capable of processing large volumes of fluid [186]. However, it is highly laborious and time-consuming, presenting the possibility of causing damage to EV structure, which can affect their performance for a specific application [183]. This method also presents low reproducibility and may lead to protein aggregation [182,183]. Nonetheless, numerous pathological conditions were already shown to benefit from interventions based on the administration of MSC-EVs isolated through differential ultracentrifugation. These include bone disorders [187], skeletal muscle [188] and spinal cord injuries [189,190], as well as cardiac [191] and liver [192] pathologies.
Density gradient ultracentrifugation is a low recovery, high specificity method [185] that consists in the ultracentrifugation of particles in a biocompatible medium, sucrose, or iodixanol, with a gradient of concentrations that allows the isolation of EVs based on their density [182]. Although this method provides higher purities than differential ultracentrifugation [182,183], it is also time-consuming and the duration and force of the centrifugation can affect the quality of EVs. In addition, the media used may negatively impact the therapeutic activity of the isolated EVs [193]. Furthermore, this method is largely instrument-dependent, and residual contaminants are often co-isolated with EVs, requiring highly trained technicians to be performed well [182,183].

5.2.2. Precipitation-Based Methods

Precipitation-based methods aim to lower the hydration levels of EVs using polymers, such as polyethylene glycol (PEG), or other synthetic hydrophilic polymers prone to interact with water molecules [183]. This is a high recovery and cost-effective method that is instrument-independent and can be used for large samples [182]. This method uses lower-speed centrifugation cycles in comparison to differential ultracentrifugation [183], thus not being as mechanically harsh. In fact, precipitation-based methods, in opposition to ultracentrifugation techniques, may preserve the proteins naturally bound to the surface of EVs, which is important for the downstream analysis of EV function [194]. However, this approach lacks specificity, providing a high contamination of the sample and leading to the precipitation of non-vesicular components, such as lipids, nucleic acids, and proteins [183]. Additionally, the precipitation reagents also influence the biological activity of the isolated EVs [195].

5.2.3. Size-Based Methods

Size-based methods consist mainly of filtration-based separation methods and size exclusion chromatography (SEC). Ultrafiltration uses a filter membrane with a defined size exclusion limit to filtrate the particles present in a sample. It is less costly and faster than ultracentrifugation, and it is easy to operate presenting good portability, and providing high purity of isolated particles [183]. MSC-EVs isolated through this method have shown angiogenic stimulatory activity [196], a property that can be harnessed for the treatment of a wide spectrum of conditions, as well as therapeutic activity in other diseases such as osteoarthritis [197]. However, low recovery, membrane clogging and EV trapping are disadvantages usually associated to this method [182]. Tangential flow filtration (TFF) is another filtration-based technique that attenuates the membrane clogging faced in ultrafiltration due to the tangential flow of fluid across the membrane surface instead of cross-flow. TFF is more efficient and gentler than differential centrifugation, providing a higher recovery and the isolation of fewer single macromolecules and aggregates [198]. Additionally, it allows the processing of large volumes of fluid and is a time-efficient, scalable, reproducible, and robust technique [198]. This method has already been used to isolate human adipose tissue MSC-EVs, that upon intra-articular injection attenuated osteoarthritis progression and prevented cartilage degeneration in mouse models [199]. Additionally, MSC-EVs isolated through TFF displayed the ability to improve alveolarization and vascularization in animal models of bronchopulmonary dysplasia, reducing the damage induced by hypoxia [200]. This further demonstrates that EVs isolated through TFF can maintain their biological therapeutic activity.
Size exclusion chromatography (SEC) allows the separation of particles based on their interaction with the pores present in the stationary phase. Sample particles with higher dimensions will not be retained within such pores and, as a result, will be eluted first (i.e., before smaller particles). This represents a gentle separation method of EVs that maintains the structural integrity and biological activity of the vesicles, which can be illustrated by the fact that SEC isolated MSC-EVs, but not ultracentrifugation isolated MSC-EVs, maintained their immunomodulatory activity, impairing T cell proliferation [201]. Moreover, this technique is easy to operate and provides a higher purity than differential ultracentrifugation [183]. Despite these advantages, SEC may suffer from pore blockage, it is a time-consuming method, and presents high equipment costs. Additionally, SEC involves sample dilution, thus usually requiring a concentration step afterward, which can further impose yield losses on the process.

5.2.4. Microfluidic-Based Methods

Microfluidic-based methods utilize microscale devices to isolate EVs. These methods can be based on EV-specific surface markers, such as microfluidic-based immunoaffinity, which rely on surfaces coated with antibodies for isolation. For example, a microfluidic device functionalized with antibodies against the tetraspanin protein, CD63, was able to efficiently isolate circulating EVs, being well suited for exosome-based diagnostics [202]. Beads coated with capture antibodies can also be employed, allowing to obtain higher EV yields [203]. In the future, this methodology may even allow isolating MSC-EV subpopulations showing increased therapeutic effects; however, the identification of such subsets is still challenging [193]. Microfluidic devices for EV isolation can also be based on EV size and density, including approaches such as membrane filtration, nano-wire-based traps, nano-sized deterministic lateral displacement, and acoustic isolations [204]. Overall, these methods are simple and efficient, as well as easy to automate, and they provide higher sensitivity than differential centrifugation.

5.2.5. Other Methods

Other relevant methods based on different EV properties are available. For instance, anion exchange chromatography is a method that takes advantage of the negative surface charge of EVs for their isolation. This method presents a similar yield, morphology, and size in comparison to differential ultracentrifugation [205]. MSC-EVs purified through this technique demonstrated the ability to prevent the onset of type 1 diabetes and uveoretinitis [206], showing the maintenance of the biological activity of the isolated MSC-EVs. Nevertheless, anion exchange chromatography may co-isolate particles with a negative net charge such as proteins. Thus, combining it with another technique such as SEC may provide greater specificity.
Apart from the microfluidic approach to the immunoaffinity isolation principle, other approaches based on chromatography are available. This method is based on the interaction of the EV surface proteins, such as CD9 or CD63 with antibodies present in the chromatography column. This provides a high sensitivity and high specificity towards specific EV subpopulations. Additionally, it is a gentle approach that conserves EV morphology [183]. Nevertheless, this is associated with low EV recovery and high costs, not being suitable for processing large volumes [183].

5.2.6. Combination of EV Isolation Methods

More than 45% of researchers included in a worldwide ISEV survey in 2020 reported the use of not a single method for the isolation of EVs, but a combination of techniques [184]. This approach is particularly beneficial because it allows obtaining higher purities than single isolation methods, and certain combinations of methods were already applied in studies with different applications for EVs.
TFF combined with SEC is becoming the preferred approach, in comparison to the typically used ultracentrifugation. TFF and SEC, which feature easy scalability and are able to comply with good manufacturing practices (GMP), were already studied in the context of an EV-based strategy that intends to utilize HEK293-derived engineered EVs carrying an immunostimulatory cytokine fusion protein, the heterodimeric IL-15/lactadherin, for a targeted cancer immunotherapy approach [207]. Importantly, such a strategy can be combined with bioreactor cultivation at upstream as a scalable and efficient method to produce purified bioactive EVs, which has the potential to be translated into industrial EV production, allowing the processing of large volumes (in the scale of liters of conditioned medium), without significant changes in size and morphology. A combined TFF and SEC approach was already employed in the isolation of MDA-MB-231-derived EVs to demonstrate the function of an innovative platform that permits the study of EV-mediated RNA delivery, the CRISPR-operated stoplight system for functional intercellular RNA exchange [208]. Additionally, SEC may also be combined with ultracentrifugation to improve the separation of EVs from lipoproteins, providing an increased purity [209].
Currently, in the field of EVs, there is still no isolation method that can provide a high recovery of EVs combined with a high specificity [185]. As such, a compromise is always required between these two variables when choosing a suitable downstream strategy. This factor, combined with overall unsatisfactory efficiencies and high associated costs, are still the main bottlenecks that undermine large-scale EV production [183]. Other factors, such as the type of biological raw materials from where EVs are isolated, the starting volume to be processed, as well as the quantity of EVs desired and their attributes, must also be considered when choosing an isolation method or a combination thereof, which should aim to be highly efficient, scalable, and GMP compliant.

6. Future Perspectives

In spite of significant advances made in the field of EVs over the last decade, the translation of EV-based therapies into clinical settings still faces several challenges ahead (Figure 3). Some of these are common to the challenges faced by cell therapies (previously reviewed by our group) [210] due to the cellular origin of EVs, while others are specific to the nature of EVs.
Establishing efficient but safe and reproducible methods for EV drug loading, and to engineer EV targeting, is still challenging and will certainly continue to be the focus of further research. The methods used for exogenous EV loading show low efficiencies, while genetic modification of EV-secreting cells for vesicle modification (either for drug loading or targeting purposes) is still troublesome and difficult in primary cells. Additionally, novel strategies have been developed over the last few years aiming to modify the surface of EVs after their isolation with significant progresses, a trend expected to continue in the following years.
The selection of appropriate EV-secreting cells and the tissue of origin of the cells needs to be carefully considered since each cell type will originate EVs with different properties (e.g., proteins and RNA packaged inside them). In addition, each cell source presents different features, including the availability and easiness to isolate from human tissue sources, cell expansion ability, and EV secretion capacity.
Appropriate cell culture conditions need to be implemented in order to assure reproducibility and compliance with GMP guidelines. Scalable processes need to be implemented for EV production in order to achieve the production of high numbers of EVs required for clinical application. Bioreactors of different geometries and configurations offer several options in order to achieve this goal, as reviewed in Section 5.1. The use of serum-/xeno(geneic)-free (S/XF) culture conditions, will also be advantageous for clinical translation. Several cell culture platforms have been tested in pioneering studies aiming to achieve a scalable production of EVs, including hollow-fiber bioreactors [211,212] and spinner systems [172], among others.
In the last 10 years, our group established different platforms for the scalable expansion of human MSC from different sources under S/XF conditions [213,214,215,216]. Recently, we further developed this work and established a platform for the scalable production of human MSC-EVs under S/XF conditions by combining the Vertical-WheelTM bioreactor system with a human platelet lysate culture supplement [173]. All of these studies allowed a significant improvement of EV production yields. Still, further work needs to be developed in order establish platforms able to reproducibly manufacture EVs at even larger scales and at a competitive cost, amenable to clinical translation. The optimization of bioreactor operation parameters, such as agitation, oxygen concentration, temperature, pH, and culture medium formulation, will be essential to maximize EV productivity. Thus, further work addressing the impact of these parameters on EV production is deemed necessary.
Similarly, EV isolation processes will need to assure reproducibility, GMP compliance, and scalability, whilst balancing suitable EV purity and yields, as reviewed in Section 5.2. Novel strategies with promising application for scalable EV isolation include combining SEC with bind-elute chromatography, whereby smaller contaminants penetrate beads and bind to its core, allowing EVs to flow through at high recovery yields [217]. Alternatively, asymmetric depth-filtration allows the immobilization of EVs at the surface and within the depth of a porous medium, while smaller particles are eluted, followed by EV recovery upon reversion of the carrier flow through the filter [218]. In the future, EV manufacturing could even benefit from integrated production and isolation processes under continuous operation in order to improve productivity and reduce costs, similar to other biopharmaceutical products (e.g., monoclonal antibodies) [219,220,221,222].
Envisaging their use as off-the-shelf products, appropriate storage conditions for EV products must be clearly defined and their stability must be assured. Strategies may be implemented in order to prolong the stability and shelf-life of EVs, such as the addition of trehalose and human albumin to EV suspensions. Trehalose, in particular, is a natural sugar that stabilizes proteins, cell membranes, and liposomes, decreases intracellular ice formation during freezing and prevents protein aggregation, being widely used in the food and drug industry, which recently revealed the capacity to prevent aggregation and cryodamage of EVs [223,224], as well as improving the short-term and long-term storage of EVs when combined with human albumin [225]. Nevertheless, only a few studies to date have addressed the storage of EV products. Further work will be required, studying EV stability and potency after storage, in conditions mimicking potentially subsequent clinical trials, in order to appropriately translate pre-clinical studies into the clinical scenario.
The translation of EV products to clinical practice will require establishing standardized identity criteria and potency assays for EVs, allowing cross comparison between different laboratories, thereby supporting the development and validation of EV-based therapies and their progression to clinical testing. Motivated by this need, a consortium of researchers recently established identity criteria including quantifiable metrics for MSC-EVs [226], and also presented requirements for the development of standardized potency tests for the therapeutic application of these EVs [227]. It is important to clearly define the mechanisms of action of EV-based therapeutics for clinical translation, and these should be reflected in suitable potency assays. However, a full elucidation of the therapeutic mechanism of EVs is challenging, since it will be multifaceted and vary between disease models.
Afterwards, appropriate preclinical models must be selected to characterize the safety and toxicology of therapeutic EVs, as well as their pharmacokinetic and pharmacodynamic profiles [109]. Information from these studies will be relevant to determine proper doses for clinical studies, which will be challenging given the current heterogeneity of EV preparations and their different therapeutic potency, depending on the targeted disease. Importantly, pre-clinical studies and current clinical trials indicate that EVs are generally safe and well tolerated [13,14,15,16,128].
To conclude, there is a long road ahead for the application of EV therapeutics in the clinical setting, as the field of EVs is still at its infancy. Nevertheless, it is already clear that EVs will likely give rise to relevant new therapeutic solutions, given their unique set of characteristics compared with synthetic nanocarriers, as well as chemical and biological products, including cell therapies.

Author Contributions

Conceptualization, M.d.A.F., N.B. and C.L.d.S.; writing—original draft preparation, M.d.A.F. and P.P.G.; writing—review and editing, A.F.-P., J.M.S.C., N.B. and C.L.d.S.; funding acquisition, A.F.-P., J.M.S.C. and C.L.d.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financed by national funds from FCT—Fundação para a Ciência e a Tecnologia, I.P., in the scope of the project UIDB/04565/2020 and UIDP/04565/2020 of the Research Unit Institute for Bioengineering and Biosciences—iBB and the project LA/P/0140/2020 of the Associate Laboratory Institute for Health and Bioeconomy—i4HB, and through the project PTDC/EQU-EQU/31651/2017. MAF acknowledges FCT for the PhD fellowship (PD/BD/128328/2017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. György, B.; Szabó, T.G.; Pásztói, M.; Pál, Z.; Misják, P.; Aradi, B.; László, V.; Pállinger, É.; Pap, E.; Kittel, Á.; et al. Membrane Vesicles, Current State-of-the-Art: Emerging Role of Extracellular Vesicles. Cell. Mol. Life Sci. 2011, 68, 2667–2688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. 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]
  3. Ratajczak, J.; Miekus, K.; Kucia, M.; Zhang, J.; Reca, R.; Dvorak, P.; Ratajczak, M.Z. Embryonic Stem Cell-Derived Microvesicles Reprogram Hematopoietic Progenitors: Evidence for Horizontal Transfer of MRNA and Protein Delivery. Leukemia 2006, 20, 847–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, J.O. Exosome-Mediated Transfer of MRNAs and MicroRNAs Is a Novel Mechanism of Genetic Exchange between Cells. Nat. Cell. Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lai, R.C.; Arslan, F.; Lee, M.M.; Sze, N.S.K.; Choo, A.; Chen, T.S.; Salto-Tellez, M.; Timmers, L.; Lee, C.N.; El Oakley, R.M.; et al. Exosome Secreted by MSC Reduces Myocardial Ischemia/Reperfusion Injury. Stem. Cell. Res. 2010, 4, 214–222. [Google Scholar] [CrossRef] [Green Version]
  6. Bruno, S.; Grange, C.; Collino, F.; Deregibus, M.C.; Cantaluppi, V.; Biancone, L.; Tetta, C.; Camussi, G. Microvesicles Derived from Mesenchymal Stem Cells Enhance Survival in a Lethal Model of Acute Kidney Injury. PLoS ONE 2012, 7, e33115. [Google Scholar] [CrossRef]
  7. Alvarez-Erviti, L.; Seow, Y.; Yin, H.; Betts, C.; Lakhal, S.; Wood, M.J.A. Delivery of SiRNA to the Mouse Brain by Systemic Injection of Targeted Exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef]
  8. EL Andaloussi, S.; Lakhal, S.; Mäger, I.; Wood, M.J.A. Exosomes for Targeted SiRNA Delivery across Biological Barriers. Adv. Drug Deliv. Rev. 2013, 65, 391–397. [Google Scholar] [CrossRef]
  9. Van Niel, G.; D’Angelo, G.; Raposo, G. Shedding Light on the Cell Biology of Extracellular Vesicles. Nat. Rev. Mol. Cell. Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  10. Ghoroghi, S.; Mary, B.; Asokan, N.; Goetz, J.G.; Hyenne, V. Tumor extracellular vesicles drive metastasis (it’s a long way from home). FASEB Bioadv. 2021, 3, 930–943. [Google Scholar] [CrossRef]
  11. Matsumoto, A.; Takahashi, Y.; Nishikawa, M.; Sano, K.; Morishita, M.; Charoenviriyakul, C.; Saji, H.; Takakura, Y. Role of Phosphatidylserine-Derived Negative Surface Charges in the Recognition and Uptake of Intravenously Injected B16BL6-Derived Exosomes by Macrophages. J. Pharm. Sci. 2017, 106, 168–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Murphy, D.E.; Jong, O.G.D.; Brouwer, M.; Wood, M.J.; Lavieu, G.; Schiffelers, R.M.; Vader, P. Extracellular Vesicle-Based Therapeutics: Natural versus Engineered Targeting and Trafficking. Exp. Mol. Med. 2019, 51, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhu, X.; Badawi, M.; Pomeroy, S.; Sutaria, D.S.; Xie, Z.; Baek, A.; Jiang, J.; Elgamal, O.A.; Mo, X.; Perle, K.L.; et al. Comprehensive Toxicity and Immunogenicity Studies Reveal Minimal Effects in Mice Following Sustained Dosing of Extracellular Vesicles Derived from HEK293T Cells. J. Extracell. Vesicles 2017, 6, 1324730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kamerkar, S.; Lebleu, V.S.; Sugimoto, H.; Yang, S.; Ruivo, C.F.; Melo, S.A.; Lee, J.J.; Kalluri, R. Exosomes Facilitate Therapeutic Targeting of Oncogenic KRAS in Pancreatic Cancer. Nature 2017, 546, 498–503. [Google Scholar] [CrossRef] [Green Version]
  15. Saleh, A.F.; Lázaro-Ibáñez, E.; Forsgard, M.A.M.; Shatnyeva, O.; Osteikoetxea, X.; Karlsson, F.; Heath, N.; Ingelsten, M.; Rose, J.; Harris, J.; et al. Extracellular Vesicles Induce Minimal Hepatotoxicity and Immunogenicity. Nanoscale 2019, 11, 6990–7001. [Google Scholar] [CrossRef]
  16. Guo, M.; Wu, F.; Hu, G.; Chen, L.; Xu, J.; Xu, P.; Wang, X.; Li, Y.; Liu, S.; Zhang, S.; et al. Autologous Tumor Cell–Derived Microparticle-Based Targeted Chemotherapy in Lung Cancer Patients with Malignant Pleural Effusion. Sci. Transl. Med. 2019, 11, 1–16. [Google Scholar] [CrossRef]
  17. Murphy, D.E.; Jong, O.G.d.; Evers, M.J.W.; Nurazizah, M.; Schiffelers, R.M.; Vader, P. Natural or Synthetic RNA Delivery: A Stoichiometric Comparison of Extracellular Vesicles and Synthetic Nanoparticles. Nano Lett. 2021, 21, 1888–1895. [Google Scholar] [CrossRef]
  18. Chargaff, E.; West, R. The Biological Significance of the Thromboplastic Protein of Blood. J. Biol. Chem. 1946, 166, 189–197. [Google Scholar] [CrossRef]
  19. Wolf, P. The Nature and Significance of Platelet Products in Human Plasma. Br. J. Haematol. 1967, 13, 269–288. [Google Scholar] [CrossRef]
  20. Anderson, H.C. Vesicles Associated with Calcification in the Matrix of Epiphyseal Cartilage. J. Cell. Biol. 1969, 41, 59–72. [Google Scholar] [CrossRef]
  21. Pan, B.T.; Johnstone, R.M. Fate of the Transferrin Receptor during Maturation of Sheep Reticulocytes in vitro: Selective Externalization of the Receptor. Cell 1983, 33, 967–978. [Google Scholar] [CrossRef]
  22. Harding, C.; Heuser, J.; Stahl, P. Receptor-Mediated Endocytosis of Transferrin and of the Transferrin Receptor in Rat Reticulocytes Recycling. J. Cell. Biol. 1983, 97, 329–339. [Google Scholar] [CrossRef] [PubMed]
  23. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle Formation during Reticulocyte Maturation. Association of Plasma Membrane Activities with Released Vesicles (Exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  24. Mathivanan, S.; Ji, H.; Simpson, R.J. Exosomes: Extracellular Organelles Important in Intercellular Communication. J. Proteom. 2010, 73, 1907–1920. [Google Scholar] [CrossRef] [PubMed]
  25. Mathieu, M.; Martin-Jaular, L.; Lavieu, G.; Théry, C. Specificities of Secretion and Uptake of Exosomes and Other Extracellular Vesicles for Cell-to-Cell Communication. Nat. Cell. Biol. 2019, 21, 9–17. [Google Scholar] [CrossRef]
  26. D’Acunzo, P.; Pérez-gonzález, R.; Kim, Y.; Hargash, T.; Miller, C.; Alldred, M.J.; Erdjument-bromage, H.; Penikalapati, S.C.; Pawlik, M.; Saito, M.; et al. Mitovesicles Are a Novel Population of Extracellular Vesicles of Mitochondrial Origin Altered in Down Syndrome. Sci. Adv. 2021, 7, eabe5085. [Google Scholar] [CrossRef] [PubMed]
  27. Peruzzotti-Jametti, L.; Bernstock, J.D.; Willis, C.M.; Manferrari, G.; Rogall, R.; Fernandez-Vizarra, E.; Williamson, J.C.; Braga, A.; van den Bosch, A.; Leonardi, T.; et al. Neural Stem Cells Traffic Functional Mitochondria via Extracellular Vesicles. PLoS Biol. 2021, 19, e3001166. [Google Scholar] [CrossRef]
  28. Zorova, L.D.; Kovalchuk, S.I.; Popkov, V.A.; Chernikov, V.P.; Zharikova, A.A.; Khutornenko, A.A.; Zorov, S.D.; Plokhikh, K.S.; Zinovkin, R.A.; Evtushenko, E.A.; et al. Do Extracellular Vesicles Derived from Mesenchymal Stem Cells Contain Functional Mitochondria? Int. J. Mol. Sci. 2022, 23, 7408. [Google Scholar] [CrossRef]
  29. Karimi, N.; Cvjetkovic, A.; Jang, S.C.; Crescitelli, R.; Hosseinpour Feizi, M.A.; Nieuwland, R.; Lötvall, J.; Lässer, C. Detailed Analysis of the Plasma Extracellular Vesicle Proteome after Separation from Lipoproteins. Cell. Mol. Life Sci. 2018, 75, 2873–2886. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, H.; Freitas, D.; Kim, H.S.; Fabijanic, K.; Li, Z.; Chen, H.; Mark, M.T.; Molina, H.; Martin, A.B.; Bojmar, L.; et al. Identification of Distinct Nanoparticles and Subsets of Extracellular Vesicles by Asymmetric Flow Field-Flow Fractionation. Nat. Cell. Biol. 2018, 20, 332–343. [Google Scholar] [CrossRef]
  31. 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]
  32. Trajkovic, K.; Hsu, C.; Chiantia, S.; Rajendran, L.; Wenzel, D.; Wieland, F.; Schwille, P.; Brügger, B.; Simons, M. Ceramide Triggers Budding of Exosome Vesicles into Multivesicular Endosomes. Science 2008, 319, 1244–1247. [Google Scholar] [CrossRef] [PubMed]
  33. Ira; Johnston, L.J. Sphingomyelinase Generation of Ceramide Promotes Clustering of Nanoscale Domains in Supported Bilayer Membranes. Biochim. Biophys. Acta Biomembr. 2008, 1778, 185–197. [Google Scholar] [CrossRef] [PubMed]
  34. Buschow, S.I.; Nolte-’t Hoen, E.N.M.; van Niel, G.; Pols, M.S.; ten Broeke, T.; Lauwen, M.; Ossendorp, F.; Melief, C.J.M.; Raposo, G.; Wubbolts, R.; et al. MHC II In Dendritic Cells Is Targeted to Lysosomes or t Cell-Induced Exosomes via Distinct Multivesicular Body Pathways. Traffic 2009, 10, 1528–1542. [Google Scholar] [CrossRef]
  35. 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] [Green Version]
  36. 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] [Green Version]
  37. Hsu, C.; Morohashi, Y.; Yoshimura, S.I.; Manrique-Hoyos, N.; Jung, S.Y.; Lauterbach, M.A.; Bakhti, M.; Grønborg, M.; Möbius, W.; Rhee, J.S.; 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] [Green Version]
  38. Jahn, R.; Scheller, R.H. SNAREs—Engines for Membrane Fusion. Nat. Rev. Mol. Cell. Biol. 2006, 7, 631–643. [Google Scholar] [CrossRef]
  39. Piccin, A.; Murphy, W.G.; Smith, O.P. Circulating Microparticles: Pathophysiology and Clinical Implications. Blood Rev. 2007, 21, 157–171. [Google Scholar] [CrossRef]
  40. Mulcahy, L.A.; Pink, R.C.; Carter, D.R.F. Routes and Mechanisms of Extracellular Vesicle Uptake. J. Extracell. Vesicles 2014, 3, 24641. [Google Scholar] [CrossRef]
  41. 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] [Green Version]
  42. Kirchhausen, T. Clathrin. Annu. Rev. Biochem. 2000, 69, 699–727. [Google Scholar] [CrossRef] [PubMed]
  43. Nanbo, A.; Kawanishi, E.; Yoshida, R.; Yoshiyama, H. Exosomes Derived from Epstein-Barr Virus-Infected Cells Are Internalized via Caveola-Dependent Endocytosis and Promote Phenotypic Modulation in Target Cells. J. Virol. 2013, 87, 10334–10347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Svensson, K.J.; Christianson, H.C.; Wittrup, A.; Bourseau-Guilmain, E.; Lindqvist, E.; Svensson, L.M.; Mörgelin, 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] [Green Version]
  45. 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]
  46. Costa Verdera, H.; Gitz-Francois, J.J.; Schiffelers, R.M.; Vader, P. Cellular Uptake of Extracellular Vesicles Is Mediated by Clathrin-Independent Endocytosis and Macropinocytosis. J. Control. Release 2017, 266, 100–108. [Google Scholar] [CrossRef]
  47. Doherty, G.J.; McMahon, H.T. Mechanisms of Endocytosis. Annu. Rev. Biochem. 2009, 78, 857–902. [Google Scholar] [CrossRef] [Green Version]
  48. Stephens, L.; Ellson, C.; Hawkins, P. Roles of PI3Ks in Leukocyte Chemotaxis and Phagocytosis. Curr. Opin. Cell. Biol. 2002, 14, 203–213. [Google Scholar] [CrossRef]
  49. Heusermann, W.; Hean, J.; Trojer, D.; Steib, E.; von Bueren, S.; Graff-Meyer, A.; Genoud, C.; Martin, K.; Pizzato, N.; Voshol, J.; et al. Exosomes Surf on Filopodia to Enter Cells at Endocytic Hot Spots, Traffic within Endosomes, and Are Targeted to the ER. J. Cell Biol. 2016, 213, 173–184. [Google Scholar] [CrossRef] [Green Version]
  50. Santos, M.F.; Rappa, G.; Karbanová, J.; Kurth, T.; Corbeil, D.; Lorico, A. VAMP-Associated Protein-A and Oxysterol-Binding Protein–Related Protein 3 Promote the Entry of Late Endosomes into the Nucleoplasmic Reticulum. J. Biol. Chem. 2018, 293, 13834–13848. [Google Scholar] [CrossRef]
  51. Corbeil, D.; Santos, M.F.; Karbanová, J.; Kurth, T.; Rappa, G.; Lorico, A. Uptake and Fate of Extracellular Membrane Vesicles: Nucleoplasmic Reticulum-Associated Late Endosomes as a New Gate to Intercellular Communication. Cells 2020, 9, 1931. [Google Scholar] [CrossRef] [PubMed]
  52. Lo Cicero, A.; Stahl, P.D.; Raposo, G. Extracellular Vesicles Shuffling Intercellular Messages: For Good or for Bad. Curr. Opin. Cell Biol. 2015, 35, 69–77. [Google Scholar] [CrossRef] [PubMed]
  53. Maas, S.L.N.; Breakefield, X.O.; Weaver, A.M. Extracellular Vesicles: Unique Intercellular Delivery Vehicles. Trends Cell Biol. 2017, 27, 172–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Miyado, K.; Yoshida, K.; Yamagata, K.; Sakakibara, K.; Okabe, M.; Wang, X.; Miyamoto, K.; Akutsu, H.; Kondo, T.; Takahashi, Y.; et al. The Fusing Ability of Sperm Is Bestowed by CD9-Containing Vesicles Released from Eggs in Mice. Proc. Natl. Acad. Sci. USA 2008, 105, 12921–12926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Desrochers, L.M.; Bordeleau, F.; Reinhart-King, C.A.; Cerione, R.A.; Antonyak, M.A. Microvesicles Provide a Mechanism for Intercellular Communication by Embryonic Stem Cells during Embryo Implantation. Nat. Commun. 2016, 7, 11958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. 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]
  57. Beckett, K.; Monier, S.; Palmer, L.; Alexandre, C.; Green, H.; Bonneil, E.; Raposo, G.; Thibault, P.; Le Borgne, R.; Vincent, J.P. Drosophila S2 Cells Secrete Wingless on Exosome-like Vesicles but the Wingless Gradient Forms Independently of Exosomes. Traffic 2013, 14, 82–96. [Google Scholar] [CrossRef] [Green Version]
  58. Vyas, N.; Walvekar, A.; Tate, D.; Lakshmanan, V.; Bansal, D.; Cicero, A.L.; Raposo, G.; Palakodeti, D.; Dhawan, J. Vertebrate Hedgehog Is Secreted on Two Types of Extracellular Vesicles with Different Signaling Properties. Sci. Rep. 2014, 4, 7357. [Google Scholar] [CrossRef] [Green Version]
  59. Corrigan, L.; Redhai, S.; Leiblich, A.; Fan, S.J.; Perera, S.M.W.; Patel, R.; Gandy, C.; Mark Wainwright, S.; Morris, J.F.; Hamdy, F.; et al. BMP-Regulated Exosomes from Drosophila Male Reproductive Glands Reprogram Female Behavior. J. Cell Biol. 2014, 206, 671–688. [Google Scholar] [CrossRef] [Green Version]
  60. Gassama, Y.; Favereaux, A. Emerging Roles of Extracellular Vesicles in the Central Nervous System: Physiology, Pathology, and Therapeutic Perspectives. Front. Cell. Neurosci. 2021, 15, 626043. [Google Scholar] [CrossRef]
  61. Fauré, J.; Lachenal, G.; Court, M.; Hirrlinger, J.; Chatellard-Causse, C.; Blot, B.; Grange, J.; Schoehn, G.; Goldberg, Y.; Boyer, V.; et al. Exosomes Are Released by Cultured Cortical Neurones. Mol. Cell. Neurosci. 2006, 31, 642–648. [Google Scholar] [CrossRef] [PubMed]
  62. Korkut, C.; Ataman, B.; Ramachandran, P.; Ashley, J.; Barria, R.; Gherbesi, N.; Budnik, V. Trans-Synaptic Transmission of Vesicular Wnt Signals through Evi/Wntless. Cell 2009, 139, 393–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Korkut, C.; Li, Y.; Koles, K.; Brewer, C.; Ashley, J.; Yoshihara, M.; Budnik, V. Regulation of Postsynaptic Retrograde Signaling by Presynaptic Exosome Release. Neuron 2013, 77, 1039–1046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Frühbeis, C.; Fröhlich, D.; Kuo, W.P.; Amphornrat, J.; Thilemann, S.; Saab, A.S.; Kirchhoff, F.; Möbius, W.; Goebbels, S.; Nave, K.A.; et al. Neurotransmitter-Triggered Transfer of Exosomes Mediates Oligodendrocyte-Neuron Communication. PLoS Biol. 2013, 11, e1001604. [Google Scholar] [CrossRef] [Green Version]
  65. Lopez-Verrilli, M.A.; Picou, F.; Court, F.A. Schwann Cell-Derived Exosomes Enhance Axonal Regeneration in the Peripheral Nervous System. Glia 2013, 61, 1795–1806. [Google Scholar] [CrossRef] [PubMed]
  66. Théry, C.; Duban, L.; Segura, E.; Væron, P.; Lantz, O.; Amigorena, S. Indirect Activation of Naïve CD4+ T Cells by Dendritic Cell-Derived Exosomes. Nat. Immunol. 2002, 3, 1156–1162. [Google Scholar] [CrossRef]
  67. Segura, E.; Nicco, C.; Lombard, B.; Véron, P.; Raposo, G.; Batteux, F.; Amigorena, S.; Théry, C. ICAM-1 on Exosomes from Mature Dendritic Cells Is Critical for Efficient Naive T-Cell Priming. Blood 2005, 106, 216–223. [Google Scholar] [CrossRef] [Green Version]
  68. Qazi, K.R.; Gehrmann, U.; Jordö, E.D.; Karlsson, M.C.I.; Gabrielsson, S. Antigen-Loaded Exosomes Alone Induce Thl-Type Memory through a B Cell Dependent Mechanism. Blood 2009, 113, 2673–2683. [Google Scholar] [CrossRef] [Green Version]
  69. Mittelbrunn, M.; Gutiérrez-Vázquez, C.; Villarroya-Beltri, C.; González, S.; Sánchez-Cabo, F.; González, M.Á.; Bernad, A.; Sánchez-Madrid, F. Unidirectional Transfer of MicroRNA-Loaded Exosomes from T Cells to Antigen-Presenting Cells. Nat. Commun. 2011, 2, 282. [Google Scholar] [CrossRef] [Green Version]
  70. Montecalvo, A.; Larregina, A.T.; Shufesky, W.J.; Stolz, D.B.; Sullivan, M.L.G.; Karlsson, J.M.; Baty, C.J.; Gibson, G.A.; Erdos, G.; Wang, Z.; et al. Mechanism of Transfer of Functional MicroRNAs between Mouse Dendritic Cells via Exosomes. Blood 2012, 119, 756–766. [Google Scholar] [CrossRef]
  71. Okoye, I.S.; Coomes, S.M.; Pelly, V.S.; Czieso, S.; Papayannopoulos, V.; Tolmachova, T.; Seabra, M.C.; Wilson, M.S. MicroRNA-Containing T-Regulatory-Cell-Derived Exosomes Suppress Pathogenic T Helper 1 Cells. Immunity 2014, 41, 89–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hergenreider, E.; Heydt, S.; Tréguer, K.; Boettger, T.; Horrevoets, A.J.G.; Zeiher, A.M.; Scheffer, M.P.; Frangakis, A.S.; Yin, X.; Mayr, M.; et al. Atheroprotective Communication between Endothelial Cells and Smooth Muscle Cells through MiRNAs. Nat. Cell Biol. 2012, 14, 249–256. [Google Scholar] [CrossRef] [PubMed]
  73. Borges, F.T.; Melo, S.A.; Özdemir, B.C.; Kato, N.; Revuelta, I.; Miller, C.A.; Gattone, V.H.; 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. 2013, 24, 385–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Shah, R.; Patel, T.; Freedman, J.E. Circulating Extracellular Vesicles in Human Disease. N. Engl. J. Med. 2018, 379, 958–966. [Google Scholar] [CrossRef]
  75. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  76. Kucharzewska, P.; Christianson, H.C.; Welch, J.E.; Svensson, K.J.; Fredlund, E.; Ringnér, M.; Mörgelin, 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] [Green Version]
  77. De Vrij, J.; Niek Maas, S.L.; Kwappenberg, K.M.C.; Schnoor, R.; Kleijn, A.; Dekker, L.; Luider, T.M.; De Witte, L.D.; Litjens, M.; Van Strien, M.E.; et al. Glioblastoma-Derived Extracellular Vesicles Modify the Phenotype of Monocytic Cells. Int. J. Cancer 2015, 137, 1630–1642. [Google Scholar] [CrossRef] [Green Version]
  78. Pucci, F.; Garris, C.; Lai, C.P.; Newton, A.; Pfirschke, C.; Engblom, C.; Alvarez, D.; Sprachman, M.; Evavold, C.; Magnuson, A.; et al. SCS Macrophages Suppress Melanoma by Restricting Tumor-Derived Vesicle-B Cell Interactions. Science 2016, 352, 242–246. [Google Scholar] [CrossRef] [Green Version]
  79. Wieckowski, E.U.; Visus, C.; Szajnik, M.; Szczepanski, M.J.; Storkus, W.J.; Whiteside, T.L. Tumor-Derived Microvesicles Promote Regulatory T Cell Expansion and Induce Apoptosis in Tumor-Reactive Activated CD8 + T Lymphocytes. J. Immunol. 2009, 183, 3720–3730. [Google Scholar] [CrossRef] [Green Version]
  80. Peinado, H.; Alečković, M.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; García-Santos, G.; Ghajar, C.M.; 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]
  81. Costa-Silva, B.; Aiello, N.M.; Ocean, A.J.; Singh, S.; Zhang, H.; Thakur, B.K.; Becker, A.; Hoshino, A.; Mark, M.T.; Molina, H.; et al. Pancreatic Cancer Exosomes Initiate Pre-Metastatic Niche Formation in the Liver. Nat. Cell Biol. 2015, 17, 816–826. [Google Scholar] [CrossRef] [PubMed]
  82. Hoshino, A.; Costa-Silva, B.; Shen, T.L.; Rodrigues, G.; Hashimoto, A.; Tesic Mark, M.; Molina, H.; Kohsaka, S.; Di Giannatale, A.; Ceder, S.; et al. Tumour Exosome Integrins Determine Organotropic Metastasis. Nature 2015, 527, 329–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhang, L.; Zhang, S.; Yao, J.; Lowery, F.J.; Zhang, Q.; Huang, W.C.; Li, P.; Li, M.; Wang, X.; Zhang, C.; et al. Microenvironment-Induced PTEN Loss by Exosomal MicroRNA Primes Brain Metastasis Outgrowth. Nature 2015, 527, 100–104. [Google Scholar] [CrossRef] [Green Version]
  84. Muralidharan-Chari, V.; Kohan, H.G.; Asimakopoulos, A.G.; Sudha, T.; Sell, S.; Kannan, K.; Boroujerdi, M.; Davis, P.J.; Mousa, S.A. Microvesicle Removal of Anticancer Drugs Contributes to Drug Resistance in Human Pancreatic Cancer Cells. Oncotarget 2016, 7, 50365–50379. [Google Scholar] [CrossRef] [PubMed]
  85. Au Yeung, C.L.; Co, N.N.; Tsuruga, T.; Yeung, T.L.; Kwan, S.Y.; Leung, C.S.; Li, Y.; Lu, E.S.; Kwan, K.; Wong, K.K.; et al. Exosomal Transfer of Stroma-Derived MiR21 Confers Paclitaxel Resistance in Ovarian Cancer Cells through Targeting APAF1. Nat. Commun. 2016, 7, 11150. [Google Scholar] [CrossRef] [Green Version]
  86. Vella, L.J.; Hill, A.F.; Cheng, L. Focus on Extracellular Vesicles: Exosomes and Their Role in Protein Trafficking and Biomarker Potential in Alzheimer’s and Parkinson’s Disease. Int. J. Mol. Sci. 2016, 17, 173. [Google Scholar] [CrossRef] [Green Version]
  87. Quek, C.; Hill, A.F. The Role of Extracellular Vesicles in Neurodegenerative Diseases. Biochem. Biophys. Res. Commun. 2017, 483, 1178–1186. [Google Scholar] [CrossRef]
  88. Yuyama, K.; Sun, H.; Mitsutake, S.; Igarashi, Y. Sphingolipid-Modulated Exosome Secretion Promotes Clearance of Amyloid-β by Microglia. J. Biol. Chem. 2012, 287, 10977–10989. [Google Scholar] [CrossRef] [Green Version]
  89. Falker, C.; Hartmann, A.; Guett, I.; Dohler, F.; Altmeppen, H.; Betzel, C.; Schubert, R.; Thurm, D.; Wegwitz, F.; Joshi, P.; et al. Exosomal Cellular Prion Protein Drives Fibrillization of Amyloid Beta and Counteracts Amyloid Beta-Mediated Neurotoxicity. J. Neurochem. 2016, 137, 88–100. [Google Scholar] [CrossRef] [Green Version]
  90. Lyu, L.; Wang, H.; Li, B.; Qin, Q.; Qi, L.; Nagarkatti, M.; Nagarkatti, P.; Janicki, J.S.; Wang, X.L.; Cui, T. A Critical Role of Cardiac Fibroblast-Derived Exosomes in Activating Renin Angiotensin System in Cardiomyocytes. J. Mol. Cell. Cardiol. 2015, 89, 268–279. [Google Scholar] [CrossRef]
  91. Wang, C.; Zhang, C.; Liu, L.; Xi, A.; Chen, B.; Li, Y.; Du, J. Macrophage-Derived Mir-155-Containing Exosomes Suppress Fibroblast Proliferation and Promote Fibroblast Inflammation during Cardiac Injury. Mol. Ther. 2017, 25, 192–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Raab-Traub, N.; Dittmer, D.P. Viral Effects on the Content and Function of Extracellular Vesicles. Nat. Rev. Microbiol. 2017, 15, 559–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ramakrishnaiah, V.; Thumann, C.; Fofana, I.; Habersetzer, F.; Pan, Q.; De Ruiter, P.E.; Willemsen, R.; Demmers, J.A.A.; Raj, V.S.; Jenster, G.; et al. Exosome-Mediated Transmission of Hepatitis C Virus between Human Hepatoma Huh7.5 Cells. Proc. Natl. Acad. Sci. USA 2013, 110, 13109–13113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Katsuda, T.; Kosaka, N.; Ochiya, T. The Roles of Extracellular Vesicles in Cancer Biology: Toward the Development of Novel Cancer Biomarkers. Proteomics 2014, 14, 412–425. [Google Scholar] [CrossRef] [PubMed]
  95. Thompson, A.G.; Gray, E.; Heman-Ackah, S.M.; Mäger, I.; Talbot, K.; El Andaloussi, S.; Wood, M.J.; Turner, M.R. Extracellular Vesicles in Neurodegenerative Disease-Pathogenesis to Biomarkers. Nat. Rev. Neurol. 2016, 12, 346–357. [Google Scholar] [CrossRef]
  96. Szabo, G.; Momen-Heravi, F. Extracellular Vesicles in Liver Disease and Potential as Biomarkers and Therapeutic Targets. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 455–466. [Google Scholar] [CrossRef]
  97. Vader, P.; Breakefield, X.O.; Wood, M.J.A. Extracellular Vesicles: Emerging Targets for Cancer Therapy. Trends Mol. Med. 2014, 20, 385–393. [Google Scholar] [CrossRef] [Green Version]
  98. Urabe, F.; Kosaka, N.; Ito, K.; Kimura, T.; Egawa, S.; Ochiya, T. Extracellular Vesicles as Biomarkers and Therapeutic Targets for Cancer. Am. J. Physiol. Cell Physiol. 2020, 318, C29–C39. [Google Scholar] [CrossRef]
  99. Lee, C.; Mitsialis, S.A.; Aslam, M.; Vitali, S.H.; Vergadi, E.; Konstantinou, G.; Sdrimas, K. Exosomes Mediate the Cytoprotective Action of Mesenchymal Stromal Cells on Hypoxia-Induced Pulmonary Hypertension. Circulation 2012, 126, 2601–2611. [Google Scholar] [CrossRef] [Green Version]
  100. Ruppert, K.A.; Nguyen, T.T.; Prabhakara, K.S.; Toledano Furman, N.E.; Srivastava, A.K.; Harting, M.T.; Cox, C.S.; Olson, S.D. Human Mesenchymal Stromal Cell-Derived Extracellular Vesicles Modify Microglial Response and Improve Clinical Outcomes in Experimental Spinal Cord Injury. Sci. Rep. 2018, 8, 480. [Google Scholar] [CrossRef]
  101. Song, Y.; Dou, H.; Li, X.; Zhao, X.; Li, Y.; Liu, D.; Ji, J.; Liu, F.; Ding, L.; Ni, Y.; et al. Exosomal MiR-146a Contributes to the Enhanced Therapeutic Efficacy of Interleukin-1β-Primed Mesenchymal Stem Cells Against Sepsis. Stem Cells 2017, 35, 1208–1221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Seo, Y.; Kim, H.S.; Hong, I.S. Stem Cell-Derived Extracellular Vesicles as Immunomodulatory Therapeutics. Stem Cells Int. 2019, 2019, 10–15. [Google Scholar] [CrossRef] [PubMed]
  103. 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] [Green Version]
  104. Gatti, S.; Bruno, S.; Deregibus, M.C.; Sordi, A.; Cantaluppi, V.; Tetta, C.; Camussi, G. Microvesicles Derived from Human Adult Mesenchymal Stem Cells Protect against Ischaemia-Reperfusion-Induced Acute and Chronic Kidney Injury. Nephrol. Dial. Transplant. 2011, 26, 1474–1483. [Google Scholar] [CrossRef] [Green Version]
  105. Bian, S.; Zhang, L.; Duan, L.; Wang, X.; Min, Y.; Yu, H. Extracellular Vesicles Derived from Human Bone Marrow Mesenchymal Stem Cells Promote Angiogenesis in a Rat Myocardial Infarction Model. J. Mol. Med. 2014, 92, 387–397. [Google Scholar] [CrossRef]
  106. Vrijsen, K.R.; Maring, J.A.; Chamuleau, S.A.J.; Verhage, V.; Mol, E.A.; Deddens, J.C.; Metz, C.H.G.; Lodder, K.; van Eeuwijk, E.C.M.; van Dommelen, S.M.; et al. Exosomes from Cardiomyocyte Progenitor Cells and Mesenchymal Stem Cells Stimulate Angiogenesis Via EMMPRIN. Adv. Healthc. Mater. 2016, 5, 2555–2565. [Google Scholar] [CrossRef]
  107. Zhang, B.; Wang, M.; Gong, A.; Zhang, X.; Wu, X.; Zhu, Y.; Shi, H.; Wu, L.; Zhu, W.; Qian, H.; et al. HucMSC-Exosome Mediated-Wnt4 Signaling Is Required for Cutaneous Wound Healing. Stem Cells 2015, 33, 2158–2168. [Google Scholar] [CrossRef]
  108. Fang, S.; Xu, C.; Zhang, Y.; Xue, C.; Yang, C.; Bi, H.; Qian, X.; Wu, M.; Ji, K.; Zhao, Y.; et al. Umbilical Cord-Derived Mesenchymal Stem Cell-Derived Exosomal MicroRNAs Suppress Myofibroblast Differentiation by Inhibiting the Transforming Growth Factor-β/SMAD2 Pathway During Wound Healing. Stem Cells Transl. Med. 2016, 5, 1425–1439. [Google Scholar] [CrossRef]
  109. Wiklander, O.P.B.; Brennan, M.; Lötvall, J.; Breakefield, X.O.; Andaloussi, S.E.L. Advances in Therapeutic Applications of Extracellular Vesicles. Sci. Transl. Med. 2019, 11, 492. [Google Scholar] [CrossRef]
  110. Gnecchi, M.; He, H.; Liang, O.D.; Melo, L.G.; Morello, F.; Mu, H.; Noiseux, N.; Zhang, L.; Pratt, R.E.; Ingwall, J.S.; et al. Paracrine Action Accounts for Marked Protection of Ischemic Heart by Akt-Modified Mesenchymal Stem Cells. Nat. Med. 2005, 11, 367–368. [Google Scholar] [CrossRef]
  111. Gnecchi, M.; He, H.; Noiseux, N.; Liang, O.D.; Zhang, L.; Morello, F.; Mu, H.; Melo, L.G.; Pratt, R.E.; Ingwall, J.S.; et al. Evidence Supporting Paracrine Hypothesis for Akt-Modified Mesenchymal Stem Cell-Mediated Cardiac Protection and Functional Improvement. FASEB J. 2006, 20, 661–669. [Google Scholar] [CrossRef] [PubMed]
  112. Timmers, L.; Kiang, S.; Arslan, F.; Armstrong, J.S.; Hoefer, I.E.; Doevendans, P.A.; Piek, J.J.; Menshawe, R.; Oakley, E.; Choo, A.; et al. Reduction of Myocardial Infarct Size by Human Mesenchymal Stem Cell Conditioned Medium. Stem Cell Res. 2008, 1, 129–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Timmers, L.; Kiang, S.; Hoefer, I.E.; Arslan, F.; Chai, R.; Van Oorschot, A.A.M.; Jose, M.; Strijder, C.; Kwan, S.; Choo, A.; et al. Human Mesenchymal Stem Cell-Conditioned Medium Improves Cardiac Function Following Myocardial Infarction. Stem Cell Res. 2011, 6, 206–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Bi, B.; Schmitt, R.; Israilova, M.; Nishio, H.; Cantley, L.G. Stromal Cells Protect against Acute Tubular Injury via an Endocrine Effect. J. Am. Soc. Nephrol. 2007, 18, 2486–2496. [Google Scholar] [CrossRef] [Green Version]
  115. Bruno, S.; Grange, C.; Deregibus, M.C.; Calogero, R.A.; Saviozzi, S.; Collino, F.; Morando, L.; Busca, A.; Falda, M.; Bussolati, B.; et al. Mesenchymal Stem Cell-Derived Microvesicles Protect against Acute Tubular Injury. J. Am. Soc. Nephrol. 2009, 20, 1053–1067. [Google Scholar] [CrossRef] [Green Version]
  116. Doeppner, T.R.; Herz, J.; Görgens, A.; Schlechter, J.; Ludwig, A.-K.; Radtke, S.; de Miroschedji, K.; Horn, P.A.; Giebel, B.; Hermann, D.M. Extracellular Vesicles Improve Post-Stroke Neuroregeneration and Prevent Postischemic Immunosuppression. Stem Cells Transl. Med. 2015, 4, 1131–1143. [Google Scholar] [CrossRef] [Green Version]
  117. Kim, D.; Nishida, H.; An, S.Y.; Shetty, A.K.; Bartosh, T.J.; Prockop, D.J. Chromatographically Isolated CD63 + CD81 + Extracellular Vesicles from Mesenchymal Stromal Cells Rescue Cognitive Impairments after TBI. Proc. Natl. Acad. Sci. USA 2016, 113, 170–175. [Google Scholar] [CrossRef] [Green Version]
  118. Li, T.; Yan, Y.; Wang, B.; Qian, H.; Zhang, X.; Shen, L.; Wang, M.; Zhou, Y.; Zhu, W.; Li, W.; et al. Exosomes Derived from Human Umbilical Cord Mesenchymal Stem Cells Alleviate Liver Fibrosis. Stem Cells Dev. 2013, 22, 845–854. [Google Scholar] [CrossRef] [Green Version]
  119. Haga, H.; Yan, I.K.; Takahashi, K.; Matsuda, A.; Patel, T. Extracellular Vesicles from Bone Marrow-Derived Mesenchymal Stem Cells Improve Survival from Lethal Hepatic Failure in Mice. Stem Cells Transl. Med. Med. 2017, 6, 1262–1272. [Google Scholar] [CrossRef]
  120. Nassar, W.; El-Ansary, M.; Sabry, D.; Mostafa, M.A.; Fayad, T.; Kotb, E.; Temraz, M.; Saad, A.N.; Essa, W.; Adel, H. Umbilical Cord Mesenchymal Stem Cells Derived Extracellular Vesicles Can Safely Ameliorate the Progression of Chronic Kidney Diseases. Biomater. Res. 2016, 20, 21. [Google Scholar] [CrossRef]
  121. Börger, V.; Weiss, D.J.; Anderson, J.D.; Borràs, F.E.; Bussolati, B.; Carter, D.R.F.; Dominici, M.; Falcón-Pérez, J.M.; Gimona, M.; Hill, A.F.; et al. International Society for Extracellular Vesicles and International Society for Cell and Gene Therapy Statement on Extracellular Vesicles from Mesenchymal Stromal Cells and Other Cells: Considerations for Potential Therapeutic Agents to Suppress Coronavirus Disease-19. Cytotherapy 2020, 22, 482–485. [Google Scholar] [CrossRef] [PubMed]
  122. Monsel, A.; Zhu, Y.G.; Gennai, S.; Hao, Q.; Hu, S.; Rouby, J.J.; Rosenzwajg, M.; Matthay, M.A.; Lee, J.W. Therapeutic Effects of Human Mesenchymal Stem Cell-Derived Microvesicles in Severe Pneumonia in Mice. Am. J. Respir. Crit. Care. Med. 2015, 192, 324–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Morrison, T.J.; Jackson, M.V.; Cunningham, E.K.; Kissenpfennig, A.; McAuley, D.F.; O’Kane, C.M.; Krasnodembskaya, A.D. Mesenchymal Stromal Cells Modulate Macrophages in Clinically Relevant Lung Injury Models by Extracellular Vesicle Mitochondrial Transfer. Am. J. Respir. Crit. Care. Med. 2017, 196, 1275–1286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Yom-Tov, N.; Guy, R.; Offen, D. Extracellular Vesicles over Adeno-Associated Viruses: Advantages and Limitations as Drug Delivery Platforms in Precision Medicine. Adv. Drug Deliv. Rev. 2022, 190, 114535. [Google Scholar] [CrossRef]
  125. 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. 2011, 19, 1769–1779. [Google Scholar] [CrossRef]
  126. Elsharkasy, O.M.; Nordin, J.Z.; Hagey, D.W.; de Jong, O.G.; Schiffelers, R.M.; Andaloussi, S.E.L.; Vader, P. Extracellular Vesicles as Drug Delivery Systems: Why and How? Adv. Drug Deliv. Rev. 2020, 159, 332–343. [Google Scholar] [CrossRef]
  127. Batrakova, E.V.; Kim, M.S. Using Exosomes, Naturally-Equipped Nanocarriers, for Drug Delivery. J. Control. Release 2015, 219, 396–405. [Google Scholar] [CrossRef] [Green Version]
  128. Rodrigues, S.C.; Cardoso, R.M.S.; Gomes, C.F.; Duarte, F.V.; Freire, P.C.; Neves, R.; Simoes-Correia, J. Toxicological Profile of Umbilical Cord Blood-Derived Small Extracellular Vesicles. Membranes 2021, 11, 647. [Google Scholar] [CrossRef]
  129. 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. 2010, 18, 1606–1614. [Google Scholar] [CrossRef]
  130. Toffoli, G.; Hadla, M.; Corona, G.; Caligiuri, I.; Palazzolo, S.; Semeraro, S.; Gamini, A.; Canzonieri, V.; Rizzolio, F. Exosomal Doxorubicin Reduces the Cardiac Toxicity of Doxorubicin. Nanomedicine 2015, 10, 2963–2971. [Google Scholar] [CrossRef]
  131. Kim, M.S.; Haney, M.J.; Zhao, Y.; Mahajan, V.; Deygen, I.; Klyachko, N.L.; Inskoe, E.; Piroyan, A.; Sokolsky, M.; Okolie, O.; et al. Development of Exosome-Encapsulated Paclitaxel to Overcome MDR in Cancer Cells. Nanomedicine 2016, 12, 655–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Kugeratski, F.G.; McAndrews, K.M.; Kalluri, R. Multifunctional Applications of Engineered Extracellular Vesicles in the Treatment of Cancer. Endocrinology 2021, 162, bqaa250. [Google Scholar] [CrossRef] [PubMed]
  133. Kooijmans, S.A.A.; de Jong, O.G.; Schiffelers, R.M. Exploring Interactions between Extracellular Vesicles and Cells for Innovative Drug Delivery System Design. Adv. Drug Deliv. Rev. 2021, 173, 252–278. [Google Scholar] [CrossRef] [PubMed]
  134. Vader, P.; Mol, E.A.; Pasterkamp, G.; Schiffelers, R.M. Extracellular Vesicles for Drug Delivery. Adv. Drug Deliv. Rev. 2016, 106, 148–156. [Google Scholar] [CrossRef]
  135. Aqil, F.; Munagala, R.; Jeyabalan, J.; Agrawal, A.K.; Gupta, R. Exosomes for the Enhanced Tissue Bioavailability and Efficacy of Curcumin. AAPS J. 2017, 19, 1691–1702. [Google Scholar] [CrossRef]
  136. 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]
  137. Hadla, M.; Palazzolo, S.; Corona, G.; Caligiuri, I.; Canzonieri, V.; Toffoli, G.; Rizzolio, F. Exosomes Increase the Therapeutic Index of Doxorubicin in Breast and Ovarian Cancer Mouse Models. Nanomedicine 2016, 11, 2431–2441. [Google Scholar] [CrossRef]
  138. Fuhrmann, G.; Serio, A.; Mazo, M.; Nair, R.; Stevens, M.M. Active Loading into Extracellular Vesicles Significantly Improves the Cellular Uptake and Photodynamic Effect of Porphyrins. J. Control. Release 2015, 205, 35–44. [Google Scholar] [CrossRef]
  139. Haney, M.J.; Klyachko, N.L.; Zhao, Y.; Gupta, R.; Plotnikova, E.G.; He, Z.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as Drug Delivery Vehicles for Parkinson’s Disease Therapy. J. Control. Release 2015, 207, 18–30. [Google Scholar] [CrossRef] [Green Version]
  140. Kooijmans, S.A.A.; Stremersch, S.; Braeckmans, K.; De Smedt, S.C.; Hendrix, A.; Wood, M.J.A.; Schiffelers, R.M.; Raemdonck, K.; Vader, P. Electroporation-Induced SiRNA Precipitation Obscures the Efficiency of SiRNA Loading into Extracellular Vesicles. J. Control. Release 2013, 172, 229–238. [Google Scholar] [CrossRef]
  141. Reshke, R.; Taylor, J.A.; Savard, A.; Guo, H.; Rhym, L.H.; Kowalski, P.S.; Trung, M.T.; Campbell, C.; Little, W.; Anderson, D.G.; et al. Reduction of the Therapeutic Dose of Silencing RNA by Packaging It in Extracellular Vesicles via a Pre-MicroRNA Backbone. Nat. Biomed. Eng. 2020, 4, 52–68. [Google Scholar] [CrossRef] [PubMed]
  142. Sato, Y.T.; Umezaki, K.; Sawada, S.; Mukai, S.A.; Sasaki, Y.; Harada, N.; Shiku, H.; Akiyoshi, K. Engineering Hybrid Exosomes by Membrane Fusion with Liposomes. Sci. Rep. 2016, 6, 21933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Piffoux, M.; Silva, A.K.A.; Wilhelm, C.; Gazeau, F.; Tareste, D. Modification of Extracellular Vesicles by Fusion with Liposomes for the Design of Personalized Biogenic Drug Delivery Systems. ACS Nano 2018, 12, 6830–6842. [Google Scholar] [CrossRef] [PubMed]
  144. Chabria, Y.; Duffy, G.P.; Lowery, A.J.; Dwyer, R.M. Hydrogels: 3D Drug Delivery Systems for Nanoparticles and Extracellular Vesicles. Biomedicines 2021, 9, 1694. [Google Scholar] [CrossRef] [PubMed]
  145. Pascucci, L.; Coccè, V.; Bonomi, A.; Ami, D.; Ceccarelli, P.; Ciusani, E.; Viganò, L.; Locatelli, A.; Sisto, F.; Doglia, S.M.; et al. Paclitaxel Is Incorporated by Mesenchymal Stromal Cells and Released in Exosomes That Inhibit in Vitro Tumor Growth: A New Approach for Drug Delivery. J. Control. Release 2014, 192, 262–270. [Google Scholar] [CrossRef]
  146. Lou, G.; Song, X.; Yang, F.; Wu, S.; Wang, J.; Chen, Z.; Liu, Y. Exosomes Derived from MIR-122-Modified Adipose Tissue-Derived MSCs Increase Chemosensitivity of Hepatocellular Carcinoma. J. Hematol. Oncol. 2015, 8, 122. [Google Scholar] [CrossRef] [Green Version]
  147. Yim, N.; Ryu, S.W.; Choi, K.; Lee, K.R.; Lee, S.; Choi, H.; Kim, J.; Shaker, M.R.; Sun, W.; Park, J.H.; et al. Exosome Engineering for Efficient Intracellular Delivery of Soluble Proteins Using Optically Reversible Protein-Protein Interaction Module. Nat. Commun. 2016, 7, 12277. [Google Scholar] [CrossRef] [Green Version]
  148. Choi, H.; Kim, Y.; Mirzaaghasi, A.; Heo, J.; Kim, Y.N.; Shin, J.H.; Kim, S.; Kim, N.H.; Cho, E.S.; Yook, J.I.; et al. Exosome-Based Delivery of Super-Repressor IκBα Relieves Sepsis-Associated Organ Damage and Mortality. Sci. Adv. 2020, 6, eaaz6980. [Google Scholar] [CrossRef] [Green Version]
  149. El-Andaloussi, S.; Lee, Y.; Lakhal-Littleton, S.; Li, J.; Seow, Y.; Gardiner, C.; Alvarez-Erviti, L.; Sargent, I.L.; Wood, M.J.A. Exosome-Mediated Delivery of SiRNA in Vitro and in Vivo. Nat. Protoc. 2012, 7, 2112–2126. [Google Scholar] [CrossRef]
  150. Zheng, Y.; Li, M.; Weng, B.; Mao, H.; Zhao, J. Exosome-based delivery nanoplatforms: Next-generation theranostic platforms for breast cancer. Biomater. Sci. 2022, 10, 1607–1624. [Google Scholar] [CrossRef]
  151. Bellavia, D.; Raimondo, S.; Calabrese, G.; Forte, S.; Cristaldi, M.; Patinella, A.; Memeo, L.; Manno, M.; Raccosta, S.; Diana, P.; et al. Interleukin 3- Receptor Targeted Exosomes Inhibit in Vitro and in Vivo Chronic Myelogenous Leukemia Cell Growth. Theranostics 2017, 7, 1333–1345. [Google Scholar] [CrossRef] [PubMed]
  152. Kooijmans, S.A.A.; Aleza, C.G.; Roffler, S.R.; van Solinge, W.W.; Vader, P.; Schiffelers, R.M. Display of GPI-Anchored Anti-EGFR Nanobodies on Extracellular Vesicles Promotes Tumour Cell Targeting. J. Extracell. Vesicles 2016, 5, 31053. [Google Scholar] [CrossRef] [PubMed]
  153. Yang, J.; Zhang, X.; Chen, X.; Wang, L.; Yang, G. Exosome Mediated Delivery of MiR-124 Promotes Neurogenesis after Ischemia. Mol. Nucleic Acids 2017, 7, 278–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Mentkowski, K.I.; Lang, J.K. Exosomes Engineered to Express a Cardiomyocyte Binding Peptide Demonstrate Improved Cardiac Retention in Vivo. Sci. Rep. 2019, 9, 10041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Ohno, S.I.; Takanashi, M.; Sudo, K.; Ueda, S.; Ishikawa, A.; Matsuyama, N.; Fujita, K.; Mizutani, T.; Ohgi, T.; Ochiya, T.; et al. Systemically Injected Exosomes Targeted to EGFR Deliver Antitumor Microrna to Breast Cancer Cells. Mol. Ther. 2013, 21, 185–191. [Google Scholar] [CrossRef] [Green Version]
  156. Kooijmans, S.A.A.; Gitz-Francois, J.J.J.M.; Schiffelers, R.M.; Vader, P. Recombinant Phosphatidylserine-Binding Nanobodies for Targeting of Extracellular Vesicles to Tumor Cells: A Plug-and-Play Approach. Nanoscale 2018, 10, 2413–2426. [Google Scholar] [CrossRef] [Green Version]
  157. Jia, G.; Han, Y.; An, Y.; Ding, Y.; He, C.; Wang, X.; Tang, Q. Biomaterials NRP-1 Targeted and Cargo-Loaded Exosomes Facilitate Simultaneous Imaging and Therapy of Glioma in Vitro and in Vivo. Biomaterials 2018, 178, 302–316. [Google Scholar] [CrossRef]
  158. Kooijmans, S.A.A.; Fliervoet, L.A.L.; Van Der Meel, R.; Fens, M.H.A.M.; Heijnen, H.F.G.; Van Bergen En Henegouwen, P.M.P.; Vader, P.; Schiffelers, R.M. PEGylated and Targeted Extracellular Vesicles Display Enhanced Cell Specificity and Circulation Time. J. Control. Release 2016, 224, 77–85. [Google Scholar] [CrossRef]
  159. Wang, Y.; Chen, X.; Tian, B.; Liu, J.; Yang, L.; Zeng, L.; Chen, T.; Hong, A. Nucleolin-Targeted Extracellular Vesicles as a Versatile Platform for Biologics Delivery to Breast Cancer. Theranostics 2017, 7, 1360–1372. [Google Scholar] [CrossRef]
  160. Tamura, R.; Uemoto, S.; Tabata, Y. Augmented Liver Targeting of Exosomes by Surface Modification with Cationized Pullulan. Acta Biomater. 2017, 57, 274–284. [Google Scholar] [CrossRef]
  161. Pi, F.; Binzel, D.W.; Lee, T.J.; Li, Z.; Sun, M.; Rychahou, P.; Li, H.; Haque, F.; Wang, S.; Croce, C.M.; et al. Nanoparticle Orientation to Control RNA Loading and Ligand Display on Extracellular Vesicles for Cancer Regression. Nat. Nanotechnol. 2018, 13, 82–89. [Google Scholar] [CrossRef] [PubMed]
  162. Tian, T.; Zhang, H.X.; He, C.P.; Fan, S.; Zhu, Y.L.; Qi, C.; Huang, N.P.; Xiao, Z.D.; Lu, Z.H.; Tannous, B.A.; et al. Surface Functionalized Exosomes as Targeted Drug Delivery Vehicles for Cerebral Ischemia Therapy. Biomaterials 2018, 150, 137–149. [Google Scholar] [CrossRef] [PubMed]
  163. Ye, Z.; Zhang, T.; He, W.; Jin, H.; Liu, C.; Yang, Z.; Ren, J. Methotrexate-Loaded Extracellular Vesicles Functionalized with Therapeutic and Targeted Peptides for the Treatment of Glioblastoma Multiforme. ACS Appl. Mater. Interfaces 2018, 10, 12341–12350. [Google Scholar] [CrossRef] [PubMed]
  164. Gao, X.; Ran, N.; Dong, X.; Zuo, B.; Yang, R.; Zhou, Q.; Moulton, H.M.; Seow, Y.; Yin, H.F. Anchor Peptide Captures, Targets, and Loads Exosomes of Diverse Origins for Diagnostics and Therapy. Sci. Transl. Med. 2018, 10, eaat0195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Pham, T.C.; Jayasinghe, M.K.; Pham, T.T.; Yang, Y.; Wei, L.; Usman, W.M.; Chen, H.; Pirisinu, M.; Gong, J.; Kim, S.; et al. Covalent Conjugation of Extracellular Vesicles with Peptides and Nanobodies for Targeted Therapeutic Delivery. J. Extracell. Vesicles 2021, 10, e12057. [Google Scholar] [CrossRef] [PubMed]
  166. Paganini, C.; Capasso Palmiero, U.; Pocsfalvi, G.; Touzet, N.; Bongiovanni, A.; Arosio, P. Scalable Production and Isolation of Extracellular Vesicles: Available Sources and Lessons from Current Industrial Bioprocesses. Biotechnol. J. 2019, 14, 1800528. [Google Scholar] [CrossRef] [PubMed]
  167. El Andaloussi, S.; Mäger, I.; Breakefield, X.O.; Wood, M.J.A. Extracellular Vesicles: Biology and Emerging Therapeutic Opportunities. Nat. Publ. Group 2013, 12, 347–357. [Google Scholar] [CrossRef]
  168. Grangier, A.; Branchu, J.; Volatron, J.; Piffoux, M.; Gazeau, F.; Wilhelm, C.; Silva, A.K.A. Technological Advances towards Extracellular Vesicles Mass Production. Adv. Drug Deliv. Rev. 2021, 176, 113843. [Google Scholar] [CrossRef]
  169. Madel, R.J.; Börger, V.; Dittrich, R.; Bremer, M.; Tertel, T.; Ngo, N.; Phuong, T.; Baba, H.A.; Kordelas, L.; Buer, J.; et al. Independent Human Mesenchymal Stromal Cell-Derived Extracellular Vesicle Preparations Differentially Affect Symptoms in an Advanced Murine Graft-versus-Host-Disease Model. bioRxiv 2020. [Google Scholar] [CrossRef]
  170. Patel, D.B.; Gray, K.M.; Santharam, Y.; Lamichhane, T.N.; Stroka, K.M.; Jay, S.M. Impact of Cell Culture Parameters on Production and Vascularization Bioactivity of Mesenchymal Stem Cell-Derived Extracellular Vesicles. Bioeng. Transl. Med. 2017, 2, 170–179. [Google Scholar] [CrossRef]
  171. Cao, J.; Wang, B.; Tang, T.; Lv, L.; Ding, Z.; Li, Z.; Hu, R.; Wei, Q.; Shen, A.; Fu, Y.; et al. Three-Dimensional Culture of MSCs Produces Exosomes with Improved Yield and Enhanced Therapeutic Efficacy for Cisplatin-Induced Acute Kidney Injury. Stem Cell Res. 2020, 11, 206. [Google Scholar] [CrossRef] [PubMed]
  172. Haraszti, R.A.; Miller, R.; Stoppato, M.; Sere, Y.Y.; Coles, A.; Didiot, M.C.; Wollacott, R.; Sapp, E.; Dubuke, M.L.; Li, X.; et al. Exosomes Produced from 3D Cultures of MSCs by Tangential Flow Filtration Show Higher Yield and Improved Activity. Mol. Ther. 2018, 26, 2838–2847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. de Almeida Fuzeta, M.; Bernardes, N.; Oliveira, F.D.; Costa, A.C.; Fernandes-Platzgummer, A.; Farinha, J.P.; Rodrigues, C.A.V.; Jung, S.; Tseng, R.J.; Milligan, W.; et al. Scalable Production of Human Mesenchymal Stromal Cell-Derived Extracellular Vesicles Under Serum-/Xeno-Free Conditions in a Microcarrier-Based Bioreactor Culture System. Front. Cell. Dev. Biol. 2020, 8, 553444. [Google Scholar] [CrossRef] [PubMed]
  174. Croughan, M.S.; Giroux, D.; Fang, D.; Lee, B. Novel Single-Use Bioreactors for Scale-Up of Anchorage-Dependent Cell Manufacturing for Cell Therapies. Stem. Cell Manuf. 2016, 105–139. [Google Scholar] [CrossRef]
  175. Bost, J.P.; Saher, O.; Hagey, D.; Mamand, D.R.; Liang, X.; Zheng, W.; Corso, G.; Gustafsson, O.; Görgens, A.; Smith, C.E.; et al. Growth Media Conditions Influence the Secretion Route and Release Levels of Engineered Extracellular Vesicles. Adv. Healthc. Mater. 2022, 11, e2101658. [Google Scholar] [CrossRef]
  176. Lu, M.; Huang, Y. Bioinspired Exosome-like Therapeutics and Delivery Nanoplatforms. Biomaterials 2020, 242, 119925. [Google Scholar] [CrossRef]
  177. Zhang, Y.; Chopp, M.; Zhang, Z.G.; Katakowski, M.; Xin, H.; Qu, C.; Ali, M.; Mahmood, A.; Xiong, Y. Systemic Administration of Cell-Free Exosomes Generated by Human Bone Marrow Derived Mesenchymal Stem Cells Cultured under 2D and 3D Conditions Improves Functional Recovery in Rats after Traumatic Brain Injury. Neurochem. Int. 2017, 111, 69–81. [Google Scholar] [CrossRef]
  178. Jang, S.C.; Economides, K.D.; Moniz, R.J.; Sia, C.L.; Lewis, N.; McCoy, C.; Zi, T.; Zhang, K.; Harrison, R.A.; Lim, J.; et al. ExoSTING, an Extracellular Vesicle Loaded with STING Agonists, Promotes Tumor Immune Surveillance. Commun. Biol. 2021, 4, 497. [Google Scholar] [CrossRef]
  179. Rekker, K.; Saare, M.; Roost, A.M.; Kubo, A.L.; Zarovni, N.; Chiesi, A.; Salumets, A.; Peters, M. Comparison of Serum Exosome Isolation Methods for MicroRNA Profiling. Clin. Biochem. 2014, 47, 135–138. [Google Scholar] [CrossRef]
  180. Stam, J.; Bartel, S.; Bischoff, R.; Wolters, J.C. Isolation of Extracellular Vesicles with Combined Enrichment Methods. J. Chromatogr. B 2021, 1169, 122604. [Google Scholar] [CrossRef]
  181. Brennan, K.; Martin, K.; FitzGerald, S.P.; O’Sullivan, J.; Wu, Y.; Blanco, A.; Richardson, C.; Mc Gee, M.M. A Comparison of Methods for the Isolation and Separation of Extracellular Vesicles from Protein and Lipid Particles in Human Serum. Sci. Rep. 2020, 10, 1039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Zhang, M.; Jin, K.; Gao, L.; Zhang, Z.; Li, F.; Zhou, F.; Zhang, L. Methods and Technologies for Exosome Isolation and Characterization. Small Methods 2018, 2, 1800021. [Google Scholar] [CrossRef]
  183. Zhang, F.; Guo, J.; Zhang, Z.; Duan, M.; Wang, G.; Qian, Y.; Zhao, H.; Yang, Z.; Jiang, X. Application of Engineered Extracellular Vesicles for Targeted Tumor Therapy. J. Biomed. Sci. 2022, 29, 14. [Google Scholar] [CrossRef] [PubMed]
  184. Royo, F.; Théry, C.; Falcón-Pérez, J.M.; Nieuwland, R.; Witwer, K.W. Methods for Separation and Characterization of Extracellular Vesicles: Results of a Worldwide Survey Performed by the ISEV Rigor and Standardization Subcommittee. Cells 2020, 9, 1955. [Google Scholar] [CrossRef]
  185. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal Information for Studies of Extracellular Vesicles 2018 (MISEV2018): A Position Statement of the International Society for Extracellular Vesicles and Update of the MISEV2014 Guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [Green Version]
  186. Zhao, Z.; Wijerathne, H.; Godwin, A.K.; Soper, S.A. Isolation and Analysis Methods of Extracellular Vesicles (EVs). Extracell. Vesicles Circ. Nucl. Acids. 2021, 2, 80. [Google Scholar] [CrossRef]
  187. Otsuru, S.; Desbourdes, L.; Guess, A.J.; Hofmann, T.J.; Relation, T.; Kaito, T.; Dominici, M.; Iwamoto, M.; Horwitz, E.M. Extracellular Vesicles Released from Mesenchymal Stromal Cells Stimulate Bone Growth in Osteogenesis Imperfecta. Cytotherapy 2018, 20, 62–73. [Google Scholar] [CrossRef]
  188. Sicco, C.L.; Reverberi, D.; Balbi, C.; Ulivi, V.; Principi, E.; Pascucci, L.; Becherini, P.; Bosco, M.C.; Varesio, L.; Franzin, C.; et al. Mesenchymal Stem Cell-Derived Extracellular Vesicles as Mediators of Anti-Inflammatory Effects: Endorsement of Macrophage Polarization. Stem Cells Transl. Med. 2017, 6, 1018–1028. [Google Scholar] [CrossRef]
  189. Han, T.; Song, P.; Wu, Z.; Xiang, X.; Liu, Y.; Wang, Y.; Fang, H.; Niu, Y.; Shen, C. MSC Secreted Extracellular Vesicles Carrying TGF-Beta Upregulate Smad 6 Expression and Promote the Regrowth of Neurons in Spinal Cord Injured Rats. Stem. Cell Rev. Rep. 2022, 18, 1078. [Google Scholar] [CrossRef]
  190. Lu, Y.; Zhou, Y.; Zhang, R.; Wen, L.; Wu, K.; Li, Y.; Yao, Y.; Duan, R.; Jia, Y. Bone Mesenchymal Stem Cell-Derived Extracellular Vesicles Promote Recovery Following Spinal Cord Injury via Improvement of the Integrity of the Blood-Spinal Cord Barrier. Front. Neurosci. 2019, 13, 209. [Google Scholar] [CrossRef]
  191. Luther, K.M.; Haar, L.; McGuinness, M.; Wang, Y.; Lynch IV, T.L.; Phan, A.; Song, Y.; Shen, Z.; Gardner, G.; Kuffel, G.; et al. Exosomal MiR-21a-5p Mediates Cardioprotection by Mesenchymal Stem Cells. J. Mol. Cell. Cardiol. 2018, 119, 125–137. [Google Scholar] [CrossRef]
  192. Mardpour, S.; Hassani, S.N.; Mardpour, S.; Sayahpour, F.; Vosough, M.; Ai, J.; Aghdami, N.; Hamidieh, A.A.; Baharvand, H. Extracellular Vesicles Derived from Human Embryonic Stem Cell-MSCs Ameliorate Cirrhosis in Thioacetamide-Induced Chronic Liver Injury. J. Cell. Physiol. 2018, 233, 9330–9344. [Google Scholar] [CrossRef]
  193. Varderidou-Minasian, S.; Lorenowicz, M.J. Mesenchymal Stromal/Stem Cell-Derived Extracellular Vesicles in Tissue Repair: Challenges and Opportunities. Theranostics 2020, 10, 5979. [Google Scholar] [CrossRef] [PubMed]
  194. McBride, J.D.; Rodriguez-Menocal, L.; Guzman, W.; Candanedo, A.; Garcia-Contreras, M.; Badiavas, E.v. Bone Marrow Mesenchymal Stem Cell-Derived CD63+ Exosomes Transport Wnt3a Exteriorly and Enhance Dermal Fibroblast Proliferation, Migration, and Angiogenesis In Vitro. Stem Cells Dev. 2017, 26, 1384–1398. [Google Scholar] [CrossRef]
  195. Paolini, L.; Zendrini, A.; di Noto, G.; Busatto, S.; Lottini, E.; Radeghieri, A.; Dossi, A.; Caneschi, A.; Ricotta, D.; Bergese, P. Residual Matrix from Different Separation Techniques Impacts Exosome Biological Activity. Sci. Rep. 2016, 6, 23550. [Google Scholar] [CrossRef] [Green Version]
  196. Xue, C.; Shen, Y.; Li, X.; Li, B.; Zhao, S.; Gu, J.; Chen, Y.; Ma, B.; Wei, J.; Han, Q.; et al. Exosomes Derived from Hypoxia-Treated Human Adipose Mesenchymal Stem Cells Enhance Angiogenesis Through the PKA Signaling Pathway. Stem Cells Dev. 2018, 27, 456–465. [Google Scholar] [CrossRef]
  197. Zhu, Y.; Wang, Y.; Zhao, B.; Niu, X.; Hu, B.; Li, Q.; Zhang, J.; Ding, J.; Chen, Y.; Wang, Y. Comparison of Exosomes Secreted by Induced Pluripotent Stem Cell-Derived Mesenchymal Stem Cells and Synovial Membrane-Derived Mesenchymal Stem Cells for the Treatment of Osteoarthritis. Stem Cell Res. Ther. 2017, 8, 64. [Google Scholar] [CrossRef] [Green Version]
  198. Busatto, S.; Vilanilam, G.; Ticer, T.; Lin, W.L.; Dickson, D.W.; Shapiro, S.; Bergese, P.; Wolfram, J. Tangential Flow Filtration for Highly Efficient Concentration of Extracellular Vesicles from Large Volumes of Fluid. Cells 2018, 7, 273. [Google Scholar] [CrossRef] [Green Version]
  199. Woo, C.H.; Kim, H.K.; Jung, G.Y.; Jung, Y.J.; Lee, K.S.; Yun, Y.E.; Han, J.; Lee, J.; Kim, W.S.; Choi, J.S.; et al. Small Extracellular Vesicles from Human Adipose-Derived Stem Cells Attenuate Cartilage Degeneration. J. Extracell. Vesicles 2020, 9, 1735249. [Google Scholar] [CrossRef] [Green Version]
  200. Porzionato, A.; Zaramella, P.; Dedja, A.; Guidolin, D.; van Wemmel, K.; Macchi, V.; Jurga, M.; Perilongo, G.; de Caro, R.; Baraldi, E.; et al. Intratracheal Administration of Clinical-Grade Mesenchymal Stem Cell-Derived Extracellular Vesicles Reduces Lung Injury in a Rat Model of Bronchopulmonary Dysplasia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2019, 316, L6–L19. [Google Scholar] [CrossRef]
  201. Monguió-Tortajada, M.; Roura, S.; Gálvez-Montón, C.; Pujal, J.M.; Aran, G.; Sanjurjo, L.; Franquesa, M.; Sarrias, M.R.; Bayes-Genis, A.; Borràs, F.E. Nanosized UCMSC-Derived Extracellular Vesicles but Not Conditioned Medium Exclusively Inhibit the Inflammatory Response of Stimulated T Cells: Implications for Nanomedicine. Theranostics 2017, 7, 270. [Google Scholar] [CrossRef]
  202. Kanwar, S.S.; Dunlay, C.J.; Simeone, D.M.; Nagrath, S. Microfluidic Device (ExoChip) for on-Chip Isolation, Quantification and Characterization of Circulating Exosomes. Lab. Chip. 2014, 14, 1891–1900. [Google Scholar] [CrossRef]
  203. Shao, H.; Chung, J.; Lee, K.; Balaj, L.; Min, C.; Carter, B.S.; Hochberg, F.H.; Breakefield, X.O.; Lee, H.; Weissleder, R. Chip-Based Analysis of Exosomal MRNA Mediating Drug Resistance in Glioblastoma. Nat. Commun. 2015, 6, 6999. [Google Scholar] [CrossRef] [Green Version]
  204. Guo, S.C.; Tao, S.C.; Dawn, H. Microfluidics-Based on-a-Chip Systems for Isolating and Analysing Extracellular Vesicles. J. Extracell. Vesicles 2018, 7, 1508271. [Google Scholar] [CrossRef] [Green Version]
  205. Heath, N.; Grant, L.; de Oliveira, T.M.; Rowlinson, R.; Osteikoetxea, X.; Dekker, N.; Overman, R. Rapid Isolation and Enrichment of Extracellular Vesicle Preparations Using Anion Exchange Chromatography. Sci. Rep. 2018, 8, 5730. [Google Scholar] [CrossRef] [Green Version]
  206. Shigemoto-Kuroda, T.; Oh, J.Y.; Kim, D.K.; Jeong, H.J.; Park, S.Y.; Lee, H.J.; Park, J.W.; Kim, T.W.; An, S.Y.; Prockop, D.J.; et al. MSC-Derived Extracellular Vesicles Attenuate Immune Responses in Two Autoimmune Murine Models: Type 1 Diabetes and Uveoretinitis. Stem Cell Rep. 2017, 8, 1214. [Google Scholar] [CrossRef] [Green Version]
  207. Watson, D.C.; Yung, B.C.; Bergamaschi, C.; Chowdhury, B.; Bear, J.; Stellas, D.; Morales-Kastresana, A.; Jones, J.C.; Felber, B.K.; Chen, X.; et al. Scalable, CGMP-Compatible Purification of Extracellular Vesicles Carrying Bioactive Human Heterodimeric IL-15/Lactadherin Complexes. J. Extracell. Vesicles 2018, 7, 1442088. [Google Scholar] [CrossRef]
  208. de Jong, O.G.; Murphy, D.E.; Mäger, I.; Willms, E.; Garcia-Guerra, A.; Gitz-Francois, J.J.; Lefferts, J.; Gupta, D.; Steenbeek, S.C.; van Rheenen, J.; et al. A CRISPR-Cas9-Based Reporter System for Single-Cell Detection of Extracellular Vesicle-Mediated Functional Transfer of RNA. Nat. Commun. 2020, 11, 1113. [Google Scholar] [CrossRef] [Green Version]
  209. You, Y.; Muraoka, S.; Jedrychowski, M.P.; Hu, J.; McQuade, A.K.; Young-Pearse, T.; Aslebagh, R.; Shaffer, S.A.; Gygi, S.P.; Blurton-Jones, M.; et al. Human Neural Cell Type-Specific Extracellular Vesicle Proteome Defines Disease-Related Molecules Associated with Activated Astrocytes in Alzheimer’s Disease Brain. J. Extracell. Vesicles 2022, 11, e12183. [Google Scholar] [CrossRef]
  210. de Almeida Fuzeta, M.; Branco, A.D.M.; Fernandes-Platzgummer, A.; da Silva, C.L.; Cabral, J.M.S. Addressing the Manufacturing Challenges of Cell-Based Therapies. Adv. Biochem. Eng. Biotechnol. 2019, 171, 225–278. [Google Scholar]
  211. Watson, D.C.; Bayik, D.; Srivatsan, A.; Bergamaschi, C.; Valentin, A.; Niu, G.; Bear, J.; Monninger, M.; Sun, M.; Morales-Kastresana, A.; et al. Efficient Production and Enhanced Tumor Delivery of Engineered Extracellular Vesicles. Biomaterials 2016, 105, 195–205. [Google Scholar] [CrossRef]
  212. Mendt, M.; Kamerkar, S.; Sugimoto, H.; McAndrews, K.M.; Wu, C.C.; Gagea, M.; Yang, S.; Blanko, E.V.R.; Peng, Q.; Ma, X.; et al. Generation and Testing of Clinical-Grade Exosomes for Pancreatic Cancer. JCI Insight 2018, 3, e99263. [Google Scholar] [CrossRef]
  213. dos Santos, F.; Campbell, A.; Fernandes-Platzgummer, A.; Andrade, P.Z.; Gimble, J.M.; Wen, Y.; Boucher, S.; Vemuri, M.C.; da Silva, C.L.; Cabral, J.M.S. A Xenogeneic-Free Bioreactor System for the Clinical-Scale Expansion of Human Mesenchymal Stem/Stromal Cells. Biotechnol. Bioeng. 2014, 116, 1116–1127. [Google Scholar] [CrossRef]
  214. Carmelo, J.G.; Fernandes-Platzgummer, A.; Diogo, M.M.; da Silva, C.L.; Cabral, J.M.S. A Xeno-Free Microcarrier-Based Stirred Culture System for the Scalable Expansion of Human Mesenchymal Stem/Stromal Cells Isolated from Bone Marrow and Adipose Tissue. Biotechnol. J. 2015, 10, 1235–1247. [Google Scholar] [CrossRef]
  215. de Soure, A.M.; Fernandes-Platzgummer, A.; Moreira, F.; Lilaia, C.; Liu, S.H.; Ku, C.P.; Huang, Y.F.; Milligan, W.; Cabral, J.M.S.; da Silva, C.L. Integrated Culture Platform Based on a Human Platelet Lysate Supplement for the Isolation and Scalable Manufacturing of Umbilical Cord Matrix-Derived Mesenchymal Stem/Stromal Cells. J. Tissue Eng. Regen. Med. 2017, 11, 1630–1640. [Google Scholar] [CrossRef]
  216. de Sousa Pinto, D.; Bandeiras, C.; de Almeida Fuzeta, M.; Rodrigues, C.A.V.; Jung, S.; Hashimura, Y.; Tseng, R.J.; Milligan, W.; Lee, B.; Ferreira, F.C.; et al. Scalable Manufacturing of Human Mesenchymal Stromal Cells in the Vertical-Wheel Bioreactor System: An Experimental and Economic Approach. Biotechnol. J. 2019, 14, 1800716. [Google Scholar] [CrossRef]
  217. Corso, G.; Mäger, I.; Lee, Y.; Görgens, A.; Bultema, J.; Giebel, B.; Wood, M.J.A.; Nordin, J.Z.; El Andaloussi, S. Reproducible and Scalable Purification of Extracellular Vesicles Using Combined Bind-Elute and Size Exclusion Chromatography. Sci. Rep. 2017, 7, 11561. [Google Scholar] [CrossRef]
  218. Chernyshev, V.S.; Chuprov-Netochin, R.N.; Tsydenzhapova, E.; Svirshchevskaya, E.V.; Poltavtseva, R.A.; Merdalimova, A.; Yashchenok, A.; Keshelava, A.; Sorokin, K.; Keshelava, V.; et al. Asymmetric Depth-Filtration: A Versatile and Scalable Method for High-Yield Isolation of Extracellular Vesicles with Low Contamination. Extracell. Vesicles 2022, 11, e12256. [Google Scholar] [CrossRef]
  219. Castilho, L.R.; Anspach, F.B.; Deckwer, W.D. An Integrated Process for Mammalian Cell Perfusion Cultivation and Product Purification Using a Dynamic Filter. Biotechnol. Prog. 2002, 18, 776–781. [Google Scholar] [CrossRef]
  220. Godawat, R.; Konstantinov, K.; Rohani, M.; Warikoo, V. End-to-End Integrated Fully Continuous Production of Recombinant Monoclonal Antibodies. J. Biotechnol. 2015, 213, 13–19. [Google Scholar] [CrossRef]
  221. Carvalho, R.J.; Castilho, L.R. Tools Enabling Continuous and Integrated Upstream and Downstream Processes in the Manufacturing of Biologicals. In Continuous Biomanufacturing—Innovative Technologies and Methods; Subramanian, G., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2018; pp. 31–68. [Google Scholar]
  222. Gronemeyer, P.; Thiess, H.; Zobel-Roos, S.; Ditz, R.; Strube, J. Integration of Upstream and Downstream in Continuous Biomanufacturing. In Continuous Biomanufacturing—Innovative Technologies and Methods; Subramanian, G., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2017; pp. 481–510. [Google Scholar]
  223. Bosch, S.; De Beaurepaire, L.; Allard, M.; Mosser, M.; Heichette, C.; Chrétien, D.; Jegou, D.; Bach, J.M. Trehalose Prevents Aggregation of Exosomes and Cryodamage. Sci. Rep. 2016, 6, 36162. [Google Scholar] [CrossRef]
  224. Jain, N.K.; Roy, I. Effect of Trehalose on Protein Structure. Protein Sci. 2009, 18, 24–36. [Google Scholar] [CrossRef]
  225. Görgens, A.; Corso, G.; Hagey, D.W.; Jawad Wiklander, R.; Gustafsson, M.O.; Felldin, U.; Lee, Y.; Bostancioglu, R.B.; Sork, H.; Liang, X.; et al. Identification of Storage Conditions Stabilizing Extracellular Vesicles Preparations. J. Extracell. Vesicles 2022, 11, e12238. [Google Scholar] [CrossRef]
  226. Witwer, K.W.; Van Balkom, B.W.M.; Bruno, S.; Choo, A.; Dominici, M.; Gimona, M.; Hill, A.F.; De Kleijn, D.; Koh, M.; Lai, R.C.; et al. Defining Mesenchymal Stromal Cell (MSC)-Derived Small Extracellular Vesicles for Therapeutic Applications. J. Extracell. Vesicles 2019, 8, 1609206. [Google Scholar] [CrossRef] [Green Version]
  227. Gimona, M.; Brizzi, M.F.; Choo, A.B.H.; Dominici, M.; Davidson, S.M.; Grillari, J.; Hermann, D.M.; Hill, A.F.; de Kleijn, D.; Lai, R.C.; et al. Critical Considerations for the Development of Potency Tests for Therapeutic Applications of Mesenchymal Stromal Cell-Derived Small Extracellular Vesicles. Cytotherapy 2021, 23, 373–380. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of an extracellular vesicle (EV) and its biological composition. EVs are composed by a phospholipid bilayer membrane enclosing intraluminal fluid with cytoplasmic origin. They contain biomolecules from their cell of origin, which include, other than lipids, several types of proteins (e.g., involved in cell adhesion, as well as other transmembrane and intraluminal proteins with various functions) and nucleic acids (e.g., mRNA and miRNA). Figure created with BioRender.com.
Figure 1. Schematic representation of an extracellular vesicle (EV) and its biological composition. EVs are composed by a phospholipid bilayer membrane enclosing intraluminal fluid with cytoplasmic origin. They contain biomolecules from their cell of origin, which include, other than lipids, several types of proteins (e.g., involved in cell adhesion, as well as other transmembrane and intraluminal proteins with various functions) and nucleic acids (e.g., mRNA and miRNA). Figure created with BioRender.com.
Bioengineering 09 00675 g001
Figure 2. Interaction of extracellular vesicles (EVs) with recipient cells. EVs bind to the surface of recipient cells in a process that can be mediated by several molecules, being able to elicit functional changes without entering the cell. EVs may also be internalized by recipient cells through different uptake routes, which include clathrin- or caveolin-mediated endocytosis, endocytosis mediated by lipid rafts, phagocytosis or macropinocytosis. Internalized vesicles follow the endosomal pathway, being sorted into early endosomes and proceeding to MVE. Then, EVs can follow different routes: they can be recycled back to the plasma membrane and released; EVs can fuse with the limiting membrane of MVE releasing their contents to the cytoplasm of the recipient cell; or MVE may fuse with lysosomes leading to EV degradation. In alternative to EV uptake, EVs may also fuse directly with the plasma membrane of the recipient cell, releasing their cargo directly into the cytosol. MVE—multivesicular endosomes. Figure created with BioRender.com.
Figure 2. Interaction of extracellular vesicles (EVs) with recipient cells. EVs bind to the surface of recipient cells in a process that can be mediated by several molecules, being able to elicit functional changes without entering the cell. EVs may also be internalized by recipient cells through different uptake routes, which include clathrin- or caveolin-mediated endocytosis, endocytosis mediated by lipid rafts, phagocytosis or macropinocytosis. Internalized vesicles follow the endosomal pathway, being sorted into early endosomes and proceeding to MVE. Then, EVs can follow different routes: they can be recycled back to the plasma membrane and released; EVs can fuse with the limiting membrane of MVE releasing their contents to the cytoplasm of the recipient cell; or MVE may fuse with lysosomes leading to EV degradation. In alternative to EV uptake, EVs may also fuse directly with the plasma membrane of the recipient cell, releasing their cargo directly into the cytosol. MVE—multivesicular endosomes. Figure created with BioRender.com.
Bioengineering 09 00675 g002
Figure 3. Key factors for the translation of extracellular vesicle (EV)-based therapeutics into clinical settings. Figure created with BioRender.com.
Figure 3. Key factors for the translation of extracellular vesicle (EV)-based therapeutics into clinical settings. Figure created with BioRender.com.
Bioengineering 09 00675 g003
Table 1. Overview of strategies to engineer targeting of extracellular vesicles (EVs) through genetic modification of producing cells to express a targeting moiety fused to an EV transmembrane protein.
Table 1. Overview of strategies to engineer targeting of extracellular vesicles (EVs) through genetic modification of producing cells to express a targeting moiety fused to an EV transmembrane protein.
EV Transmembrane ProteinTargeting MoietyTargetPurposeRefs.
Lamp2bRVG peptideAcetylcholine receptor in neuronsDeliver BACE1 siRNA for Alzheimer’s disease treatment[7,149]
RVG peptideAcetylcholine receptor in neuronsDeliver miR-124 to promote neurogenesis after stroke[153]
iRGD peptideαv integrin-positive breast cancer cellsDeliver doxorubicin[136]
IL3 fragment IL3 receptor in CML cellsDeliver Imatinib or BCR-ABL siRNA [151]
Cardiomyocyte specific peptideCardiomyocytesTarget cardiomyocytes[154]
PDGFRGE11 peptideEGFR-expressing cancer cellsDeliver let-7a miRNA[155]
GPI-anchoring peptideAnti-EGFR nanobodyEGFR-expressing cancer cellsTarget cancer cells[152]
Lamp2b—lysosome-associated membrane protein 2; RVG—rabies viral glycoprotein; BACE1—beta secretase 1; IL3—interleukin 3; CML—chronic myeloid leukemia; PDGFR—platelet-derived growth factor receptor; EGFR—epidermal growth factor receptor; GPI—glycosylphosphatidylinositol.
Table 2. Overview of strategies to engineer extracellular vesicles (EVs) targeting by anchoring a targeting moiety to EVs after its production and isolation.
Table 2. Overview of strategies to engineer extracellular vesicles (EVs) targeting by anchoring a targeting moiety to EVs after its production and isolation.
Linkage MethodTargeting MoietyTargetPurposeRefs.
Post-insertion of phospholipid (DMPE)-PEG fusion molecules in EV membranesAnti-EGFR nanobody (conj. with DMPE-PEG)EGFR-expressing cancer cellsTarget cancer cells[158]
Membrane anchoring cholesterolAS1411 DNA aptamer (conj. with cholesterol)Nucleolin on breast cancer cellsDeliver let-7 miRNA or VEGF siRNA[159]
Electrostatic interaction between cationized pullulan and EVsCationized pullulan (a polysaccharide polymer)Hepatocyte asialoglycoprotein receptorsTarget injured liver[160]
C1C2 domain of lactadherin binding to PS present on EV membraneAnti-EGFR nanobody (conj. with C1C2)EGFR-expressing cancer cellsTarget cancer cells[156]
Membrane anchoring cholesterolRNA aptamers or folate (conj. with cholesterol)PSMA, EGFR or folate receptor on prostate, breast, or colorectal cancers, respectivelyDeliver survivin-targeting siRNA[161]
Click chemistry reactionc(RGDyK) peptideIntegrin αvβ3 in reactive cerebral vascular endothelial cells after ischemiaDeliver curcumin to stroke lesions[162]
ApoA-I mimetic peptide interaction with phospholipids on EV membraneLDL peptideLDL receptor on GBM cellsDelivery of KLA peptide and methotrexate[163]
CP05 peptide binding to CD63 present on EV membraneMuscle targeting peptide M12 (conj. with CP05)MuscleDeliver PMO to muscle for Duchenne muscular dystrophy treatment[164]
Click chemistry reactionNRP-1 targeting peptide (RGE)NRP-1 in glioma cells and tumor vascular endotheliumDeliver SPIONs and curcumin for imaging and therapy of glioma[157]
Covalent bond by protein ligating enzymes Sortase A or OaAEP1 ligaseEGFR-targeting peptide or nanobodies targeting EGFR or HER2Cancer cells expressing EGFR or HER2Deliver paclitaxel or mRNA[165]
DMPE—1,2-Dimyristoyl-sn-glycero-3-phosphoethanolamine; PEG—polyethylene glycol; EGFR—epidermal growth factor receptor; VEGF—vascular endothelial growth factor; PS—phosphatidylserine; PSMA—Prostate-specific membrane antigen; LDL—low-density lipoprotein; GBM—glioblastoma multiforme; PMO—phosphorodiamidate morpholino oligomer; NRP-1—neuropilin-1; SPION—superparamagnetic iron oxide nanoparticles; HER2—human epidermal growth factor receptor 2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

de Almeida Fuzeta, M.; Gonçalves, P.P.; Fernandes-Platzgummer, A.; Cabral, J.M.S.; Bernardes, N.; da Silva, C.L. From Promise to Reality: Bioengineering Strategies to Enhance the Therapeutic Potential of Extracellular Vesicles. Bioengineering 2022, 9, 675. https://doi.org/10.3390/bioengineering9110675

AMA Style

de Almeida Fuzeta M, Gonçalves PP, Fernandes-Platzgummer A, Cabral JMS, Bernardes N, da Silva CL. From Promise to Reality: Bioengineering Strategies to Enhance the Therapeutic Potential of Extracellular Vesicles. Bioengineering. 2022; 9(11):675. https://doi.org/10.3390/bioengineering9110675

Chicago/Turabian Style

de Almeida Fuzeta, Miguel, Pedro P. Gonçalves, Ana Fernandes-Platzgummer, Joaquim M. S. Cabral, Nuno Bernardes, and Cláudia L. da Silva. 2022. "From Promise to Reality: Bioengineering Strategies to Enhance the Therapeutic Potential of Extracellular Vesicles" Bioengineering 9, no. 11: 675. https://doi.org/10.3390/bioengineering9110675

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