Next Article in Journal
Effects of the In Situ Growth of CNTs on Ti-Coated Diamond Surfaces on the Mechanical Properties of Diamond/Aluminum Composites
Next Article in Special Issue
Physiochemical Characterization of Lipidic Nanoformulations Encapsulating the Antifungal Drug Natamycin
Previous Article in Journal
Optimizing Antimicrobial Efficacy: Investigating the Impact of Zinc Oxide Nanoparticle Shape and Size
Previous Article in Special Issue
Recent Uses of Lipid Nanoparticles, Cell-Penetrating and Bioactive Peptides for the Development of Brain-Targeted Nanomedicines against Neurodegenerative Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Engineered and Mimicked Extracellular Nanovesicles for Therapeutic Delivery

by
Verena Poinsot
*,
Nathalie Pizzinat
and
Varravaddheay Ong-Meang
Inserm, CNRS, Faculté de Santé, Université Toulouse III—Paul Sabatier, I2MC U1297, 31432 Toulouse, France
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(7), 639; https://doi.org/10.3390/nano14070639
Submission received: 6 March 2024 / Revised: 2 April 2024 / Accepted: 3 April 2024 / Published: 6 April 2024
(This article belongs to the Special Issue Advances in Pharmaceutical Applications of Lipid-Based Nanoparticles)

Abstract

:
Exosomes are spherical extracellular nanovesicles with an endosomal origin and unilamellar lipid-bilayer structure with sizes ranging from 30 to 100 nm. They contain a large range of proteins, lipids, and nucleic acid species, depending on the state and origin of the extracellular vesicle (EV)-secreting cell. EVs’ function is to encapsulate part of the EV-producing cell content, to transport it through biological fluids to a targeted recipient, and to deliver their cargos specifically within the aimed recipient cells. Therefore, exosomes are considered to be potential biological drug-delivery systems that can stably deliver their cargo into targeted cells. Various cell-derived exosomes are produced for medical issues, but their use for therapeutic purposes still faces several problems. Some of these difficulties can be avoided by resorting to hemisynthetic approaches. We highlight here the uses of alternative exosome-mimes involving cell-membrane coatings on artificial nanocarriers or the hybridization between exosomes and liposomes. We also detail the drug-loading strategies deployed to make them drug-carrier systems and summarize the ongoing clinical trials involving exosomes or exosome-like structures. Finally, we summarize the open questions before considering exosome-like disposals for confident therapeutic delivery.

1. Introduction

Besides the apoptotic bodies, endogenous cell-membrane-based vesicles are secreted by most cell types and are called extracellular vesicles (EVs). EVs adopt various shapes and sizes, from one micrometer to several nanometers. Described for the first time about 30 years ago [1], exosomes are the nano-sized members of this vast family. These are spherical nanovesicles with an endosomal origin and unilamellar lipid bilayer structure [2] with sizes ranging from 30 to 100 nm. Exosomes were first identified as garbage bags [3], but their capability to stimulate adaptive immune responses has been rapidly observed [4]. The biogenesis isolation and characterization methods have been widely reviewed (e.g., [5,6]).
EVs were found to contain a large range of proteins, lipids, and nucleic acids species [7,8,9] in their aqueous core, but also membrane proteins, MHC (major histocompatibility complex) molecules, and integrins embedded in the phospholipid membrane. Moreover, their content depends on the state and origin of the EV-secreting cell [10]. Consequently, EVs are demonstrated to be heterogeneous in both content and functionality [11,12]. The EVs’ function is to encapsulate part of the EV-producing cell content, to transport it through biological fluids, and to a targeted recipient. Indeed, exosomes were demonstrated to be able to target (even at distance) cells that are different from parent cells (usually called recipient cells). After targeting, these nanovesicles are internalized to deliver their cargos specifically within the recipient cells. This fate implies excellent biocompatibility, low toxicity, and immunogenicity. It produced an increasing interest in exploiting EVs for diagnostic and therapeutic purposes, and even clinical trials [13,14,15,16,17]. It can be noticed that, until the pure spherical nanosizing of the exosomes cannot be proved, exosomes should be called EVs or nanosized EVs [18].
In line with their vesicular structure, EVs are permeable and quite dynamic systems. However, the knowledge on how EVs can transfer their content to recipient cells remains limited but central for the further development of EV-inspired therapies. The review of Kooijmans et al. [19] gives the landscape and latest insights in critical effectors in the EV–cell interface.
Structurally, EVs are biological members of the lipidic nanoparticles family, as their formation involves all lipid-type pathways, lipid molecular types, and lipid transporter metabolism [20]. Lipids—with approximately 280 molecular species—constitute the second component in the exosome membrane [21]. Lipids and lipid rafts are also one of the main components of exosomes obtained from donor cells [22,23]. Even if exosomes are formed by the endosome’s membrane invagination, the lipid composition of exosomes differs when compared to the parent-cell membrane. Exosomes contain the CD36 in their cargo, which acts as fatty acid translocase [24], allowing individual nutrition and exogenous lipids to modify their lipid structure. Various factors influence the abundance and composition of these lipids, including the producer-cell line, cell type, and the physiological stage of the parent cell [25] (Figure 1).
The lipidomics array demonstrated this parent-cell specificity, as ceramide and gangliosides are the dominant lipids in exosomes derived from neuronal and glial cells [26]. In contrast, cholesterol, sphingomyelin, and ceramides are enriched in exosomes deriving from mouse neuroblastoma [27]. Endosomal sorting complex required for transport (ESCRT)-dependent and -independent mechanisms are proposed to explain this statement. ESCRT is composed of four complexes [28]. The ESCRT-0 complex recognizes and sequesters ubiquitinated transmembrane proteins in the endosomal membrane. The ESCRT-I and -II complexes are implicated in membrane-bud formation with sorted cargos. ESCRT-III drives vesicle ablation [29]. Therefore, ESCRT-I and -II determine the lipid content of exosomes, but the lipid raft-mediated cargo loading is ESCRT-independent. These domains are enriched in cholesterol, sphingolipids, and GPI-anchored proteins that are involved in the protein sorting [30]. Phospholipase D2 [31] and diacylglycerol kinase [32] are involved in the maturation of endosomes, especially the formation of the intra-lumiminal vesicles [33]. The sphingomyelinase and bis-monoacylglycerol phosphate play key roles in exosome biogenesis, as well as budding and release [34]. An increased amount of cell ether lipids also facilitates exosome release by increasing the late endosome fusion with the plasma membrane. Finally, the phosphoinositides, PIP3, PI (3,5), and P2 also help the formation of exosomes by inhibiting the PI3K/Akt (a negative regulator of the exosome secretion signaling pathway) [35,36]. The amount of cholesterol in the late endosomes is decisive in their fate. Indeed, cholesterol-rich vesicles fuse with the cell plasma membrane and its content released into the extracellular space, while cholesterol-free vesicles are directed to the lysosome for destruction [31].
Exosomes are involved in lipid synthesis, transportation, and degradation [37]. Referring to the proteins, the amount of lipids in exosomes is four times higher compared to the original cells content [38]. Moreover, their lipid composition differs from the one of the parent cell [21,39]. In accordance with their formation and release mechanisms, the exosomes are globally enriched with sphingomyelins, phosphatidylserins, gangliosides, glycosphingolipids, phosphatidic acid, phosphatidylcholins, phosphatidylethanolamins, and cholesterol [21,40]. However, these were reported to be impoverished in diacylglycerol, relative to the membranes of their cells of origin [41,42,43].
In addition to their lipidic structure, exosomes are biological vehicles that directly transfer lipids (including arachidonic acid and eicosanoid) [44] and contain various lipid transporters. Exosomes also contain a variety of enzymes involved in lipid metabolism, including phospholipases D, A2 [45], and oxilipin synthases (i.e., 5- and 15-lipoxygenase, leukotriene A4 hydrolase, and leukotriene C4 synthase, PGD synthase) [46,47].
Exosomes significantly impact the immune system. Indeed, tumor cells actively release exosomes that interplay with the immune system and the cancer development itself. Immunosuppressive molecules present in tumor-derived exosomes enable cancer cells to evade the immune system by impairing the function of immune cells such as the cytotoxic T lymphocytes or NK cells, allowing them to spread and proliferate. Conversely, many tumor-derived exosomes contain tumoral antigens that may be delivered to dendritic cells and stimulate T-cell-mediated antitumor immune response [48]. Similarly, innate immune cell-derived exosomes take part in antitumor immune responses and their properties could be enhanced to target and eliminate cancer cells [49]. Moreover, several vaccine studies have demonstrated that infected cells secrete exosomes containing multiple antigens from the hosted pathogen, helping to overcome pathogen escape from the immune system [50].
Liposomes are lipidic nanoparticles (LNPs) with spherical hollows and a lipid bilayer structure, allowing to encapsulate a variety of drugs within the lipid membrane or into the cavity (e.g., [51,52,53,54]). Even if they demonstrate good biological compatibility, increasing their stability, retention time in blood, and the efficiency of the drug delivery was necessary [55,56,57,58]. The use of PEGylated lipids, pH-sensitive cationic lipids, enabled their evaluation in a clinical trial [59].
LNPs and EVs allow two different targeting modes: the active and passive ones [60,61,62]. Active targeting can be achieved for the liposomes by conjugating specific ligands to the LNP surface, while EVs are biosynthesized with a large toolbox allowing them to bind to specific target cell receptors, mediating their inclusion. Although LNPs offer many advantages for drug-delivery systems, their poor stability in circulation causes drug release in non-target organs [63]. Therefore, exosomes are considered as drug-delivery systems, biologically produced, that can stably deliver their cargo into targeted cells.
Various cell types have been used to produce exosomes for therapeutic issues, including cancer model cell lines (e.g., [64,65]) and immature dendritic cells [66]. Unfortunately, their use for therapeutic purposes faces several problems. First, for all types of investigated cells, a scaling-up of the exosome production yield is required for clinical translatability [67,68]. Until now, mesenchymal stem cells (MSC) are the quantitatively best exosome-producing cells [69]. Secondly, their biogenesis and secretion are fine-regulated processes affected by numerous factors, like pH, calcium concentration, and cellular state [7]. Consequently, their production remains unreproducible in amount and composition. Finally, due to their size, density, and presence of membrane proteins, their isolation from the free proteins (with potentially undesired immunogenicity or activity) of the surroundings is a real challenge. Therefore, nanosized-EVs cannot be easily used in clinical trials as high-performance drug carriers, and developing synthetic exosome mimes or engineered biological nanoEVs remains a suitable strategy.
In this review, we first summarize the ongoing clinical trials involving exosomes. Then, we report on therapeutic assays involving engineered exosomes, including their biological modifications and the drug-loading strategies used to transform exosomes into therapeutic carriers. We also highlight the uses of the alternative exosome mimes including the cell-membrane coating on artificial nanocarriers and the hybridization between exosomes and liposomes. Indeed, these were extensively investigated over the last decade regarding their potential as drug-delivery systems, due to their upgraded biocompatibility, biodegradability, low-immunogenicity, and long-term blood circulation. Finally, we conclude this work with the opened questions to address before considering exosome-like disposals for confident therapeutic delivery.

2. Clinical Trials Based on Exosomes

The effects of exosomes on targeted pathologies have been reviewed elsewhere [70]. Therefore, this topic will not be detailed here. However, these studies led to several clinical trials that are interesting to summarize. The first reported trials started in the year 2000, but this work only focuses on the ongoing assays (Table 1).
Technological advances, based on exosomes and applied in the field of diagnosis and prognosis, focused principally on cancer pathologies, but their potential translation and clinical settings remain under scope. Few overviews were published on this topic [71,72,73] and some clinical trials linked to this topic can be found in Table 1.
ClinicalTrials is a governmental online site (https://www.clinicaltrials.gov/—accessed on 18 February 2024) listing authorized therapeutic trials. Searching for the keyword “exosomes”, a list of 361 studies was obtained, beginning in 1999. These clinical studies are either observational, establishing the diagnostic and therapeutic character of exosomes used on human samples (serum, biopsy, etc.), or interventional (phase 2 or 3), monitoring their effects along the treatment in humans.
WHO is another website (https://www.who.int/data/stories/world-health-statistics-2023—accessed on 18 February 2024) giving access to worldwide statistics for health. Concerning the year 2023, it reports that “between 2000 and 2019, the number of deaths caused by non-communicable diseases increased by more than a third, from 31 million lives lost to 41 million lives lost, or nearly 3 out of every 4 deaths worldwide”. Additionally, it indicates that “The four major non-communicable diseases—cardiovascular disease (17.9 million deaths), cancer (9.3 million deaths), chronic respiratory disease (4.1 million deaths) and diabetes (2.0 million deaths)—collectively killed around 33.3 million people in 2019”. The breakdown of the 361 studies found on ClinicalTrials does not mirror these statistics as 141 of the trials aim at various cancers (39%), 39 concern chronic diseases such as diabetes (6.6%), hypertension (1%), or obesity (3%), and 25 concern cardiorespiratory diseases (6.9%). The table below presents the most recent clinical trials involving exosomes. Close to 56% of the 57 trials are theragnostic studies.

3. Artificially Loaded Exosomes

Drugs are bioactive compounds that often present low water solubility, poor organ targeting specificity, fast metabolization (degradation), easy accumulation in healthy tissues (toxicity), and poor ability to penetrate cells. Consequently, improving their delivery requests the development of reasonable carriers. As evoked in the introduction, exosomes are circulating extracellular nanovesicles derived from biological cells with a membrane structure resembling the one of the parent cells. Hence, they demonstrate low immunogenicity and high stability in the blood circulation. Consequently, the protection of the drugs and their addressing by exosomes have been considered for years. Numerous works reported that exosomes effectively protected the loaded drug from the environment, increasing their efficiency, but also enhanced their targeting capacity, and recipient cells fusing ability. It has also been noticed that, thanks to their unique membrane structure and size, exosomes easily pass through the biological barrier [74] and can escape some immunity responses [4,48]. However, introducing ingredients into pre-existing and stable structures is not simple and several strategies have been proposed.

3.1. Loading Methods

Developing an efficient and reproducible method to embed the desired therapeutic ingredient into exosomes without disturbing their biological integrity and identity is a major concern to achieve an exosome-based drug-delivery system. Two types of loading are under the scope of this paper: (1) biologically-engineered exosome production, where the ingredient is encoded by the cell that produces the exosomes’ natural cargo; and (2) the physicochemical post-loading of exosomes [75]. In this work, we reviewed only the second loading type (Figure 2).
The most employed method for small hydrophobic molecules and the easiest is the co-incubation of the drug and the exosomes at room temperature. However, co-incubation is time-consuming, has very low loading ratios (around 1%), and is not applicable to hydrophilic compounds.
Only a few types of drugs can be passively loaded into exosomes, implying that the most employed loading methods are based on a supplementary physical or chemical force, affecting the exosomes and therapeutic cargo integrities.
Sonication and electroporation constitute alternatives to co-incubation. The sonication principle is based on the weakening of the exosome-membrane integrity by exposure to electromagnetic waves, facilitating the drug entrance. This method is compatible with hydrophilic drugs and allows a close to 30% loading ratio. The drawbacks of sonication are exosome aggregation processes and membrane disruption, both affecting the immune tolerance of the generated exosomes.
Electroporation is a rapid and high-loading yield (up to 60%) technique. This time, electric pulses are transitory, creating micropores in the lipid bi-membrane, allowing the entrance of the chemicals. The disadvantages are significant losses of membrane integrity and possible protein destruction.
Another efficient, but less tested, method is extrusion where the drug and the exosomes have to pass together through a 100 to 400 nm membrane. During this process, the exosome membrane is destroyed and reformed, including the proximal drug molecules. All these methods are detailed and recently reviewed elsewhere [76].
The protocols that were developed are summarized in Table 2. It should be noted that the loading efficiencies reported are not fully robust. Even if authors make their best efforts to measure the drug content, the outcomes are highly variable, mostly due to the difficulty of quantifying the exosomes and of assessing if the drug is encapsulated or absorbed.

3.2. Loaded Exosomes Preclinical Applications

The diversity of the performed studies reflects the diversity of the ingredients that are encapsulated within the exosomes. Therefore, we choose to describe the published works in terms of the large families of compounds they transport.

3.2.1. Loading for Tracking and Labelling Purposes

The marking and tracking of exosomes are necessary for setting up therapeutic applications. Optical imaging is currently the most used imaging method for tracking molecules and cells, and fluorescence microscopy is used for studying the physiological activities of living cells. For such purposes, the exosomes need to be labeled with fluorescent probes, and the physiological changes in individual cells can be monitored. Several methods of fluorescent labeling for exosomes were described. First, lipophilic organic dyes can insert the lipid exosome bilayer [77,78,79,80]. Alternatively, green fluorescent proteins can be genetically encoded in the parent cell [81]. Another approach, high-specific immunofluorescence labeling, consists of labelling the antibody with an organic dye before the exosome interaction [82].
Finally, exosome mimes based on trackable solid core nanoparticles with a cell-membrane shell were developed (see the paragraph exosome mimes).

3.2.2. RNA-Loaded EVs

Exosomes naturally carry nucleic acids, such as DNA and RNA, to target cells in both biological and pathogenic processes [8]. Therefore, several studies were conducted to deliver genetic materials to correct diseases altering gene expression and improving genetic therapy. In particular, the successful delivery of exogenous genetic material by exosomes has been highlighted and their post-delivery functionality was described as a valuable asset [83]. The delivery of siRNA to T-cells and monocytes using human exosomes was investigated in gene therapy purposes [84]. The use of endothelial exosomes to deliver siRNA was also under scope [85]. Using electroporation, exosomes isolated from endothelial cells were loaded with siRNA designed to silence a vector encoding for luciferase (pGL2). Their efficiency in delivering siRNAs was tested on luciferase-expressing endothelial cells. The significant lowering of luciferase expression indicated that endothelial exosomes can deliver exogenous functional agents to cells in vitro.
Another in vitro study investigated the exosome-delivery of siRNA against RAD51 (a eukaryote gene protein that assists in the repair of DNA double-strand breaks), in cancer cell lines. Exosomes were isolated from HeLa cells by ultracentrifugation, and chemically loaded with Alexa-fluor 488-labeled siRNA and then co-cultured with recipient cells (HeLa and HT1080 cells). The successful delivery efficiency of siRNA by exosomes was confirmed by confocal microscopy and flow cytometry. Western blot analysis showed a considerable reduction in RAD51 and RAD52 protein levels, indicating successful down-regulation of this specific gene [86].
Surface proteins, such as PDL-1 or CTLA-4 and FGL-1, expressed by immune and tumor cells, ensure the modulation of the immune microenvironment. Immune checkpoint therapy encounters significant clinical challenges, including low response rate to cancer treatment, acquired resistance, and immune-related adverse events. Inhibiting FGL1 and the immunosuppressive cytokine TGF-β1 is essential for colorectal cancer immunotherapy. Therefore, the authors constructed a modified erythrocyte exosome (targeting the cancer cells) and loaded them with siFGL1 and siTGF-β1. They reported a significant anti-tumor effect, both in vitro and in vivo [87].
The mutant form of the GTPase KRAS is a crucial driver of pancreatic cancer. The authors prepared exosomes derived from normal fibroblast-like mesenchymal cells. They demonstrated a better persistence of exosomes in the mice circulation compared to liposomes, probably thanks to a CD47-mediated protection from phagocytosis. Moreover, they loaded them with siRNA or short hairpin RNA specific to oncogenic KrasG12D and reported an increased overall survival in mouse models of pancreatic cancer [88].
EVs’ delivery of siRNA was also under the scope for spinal cord injury treatments. Bone marrow MSC-derived exosomes were electroporated to load them with siRNA silencing CTGF gene. The loaded bone marrow-MSC-derived exosomes were injected into the tail vein of spinal cord injured rats. These inhibited the expressions of CTGF gene but also produced several positive side effects like quenching inflammation, decreased neuronal apoptosis, and reactive scar formation [89].
The peptide CAQK was chemically anchored to the membranes of EVs isolated from induced neural stem cells, after which CCL2-siRNA was loaded through electroporation. These drug-delivery systems specifically delivered siRNA to the injured region and were taken up by target cells. A synergistic effect was noticed between the inherent functions of the stem cell-derived vesicles and the loaded siRNA. Indeed, these loaded and modified EVs limited the negative effects of the inflammatory response and neuronal injury [90].
For all these studies, siRNAs were not optimized for their nanovesicular delivery, except for one work where cholesterol-conjugated siRNAs (hydrophobic siRNAs) were evaluated concerning the conservation of the activity and the loading rate onto EVs. The commercial linker triethyl glycol was preferred due to the better loading of the conjugate within the nanovesicles. It is noticed that these siRNAs-cholesterol-conjugates presented a saturation level of around 3,000 siRNA copies per EV [91].
In addition to siRNA, the delivery of microRNAs (short non-coding RNAs, involved in post-transcriptional regulation of gene expression in multicellular organisms by affecting both the stability and translation of mRNA) was also investigated, as exosomes are known to carry and deliver miRNA to targeted cells naturally (recent, e.g., [92]).
Epidermal growth factor (EGF) and EGF receptor-specific peptide (GE11), which binds specifically to EGF receptor, were incorporated onto the surfaces of exosomes that carried miRNA lethal-7 gene (let-7a), to deliver let-7a to EGF receptor-expressing cancer tissue. The ability of GE11- or EGF-positive exosomes to bind to EGF receptor in various breast cancer cell lines, including HCC70, HCC1954, and MCF-7, was tested. miRNA let-7a was introduced into GE11 exosomes by lipofection. The GE11 EVs were labeled with cell-membrane PKH67 dye to track them within the recipient cells. The GE11 exosomes containing let-7a were intravenously injected into tumor-bearing mice, where they strongly inhibited the expression of HMGA2, indicating that exosomes successfully deliver their cargo to the target cells [93].
The vascular dysfunction of ovarian cancer contributes to chemotherapeutic resistance, and miR-484 was found to be down-regulated in both the cancer cells and the angiogenic endothelial cells. To enhance their tumor targeting, HEK293T cell-derived exosomes were modified to present the adhesion peptide RGD on their surface. In vivo, the injection of RGD-modified exosomes loaded with miR-484 induced vessel normalization and, in turn, sensitized the cancer cells to chemotherapy-induced apoptosis [94].
Similarly to siRNA, hydrophobic miRNA-conjugates were developed. Considering the low loading of rates miRNA within the nanovesicles, instead of engineering cholesterol-conjugates, it was proposed to conjugate miRNA with a cell-penetrating peptide. The YARA peptide, derived from human immunodeficiency virus, was covalently conjugated with miR-21-5p, and then incubated with mesenchymal stem or cancer cell-derived exosomes. The YARA-miR-21-5p demonstrated a significantly increased loading capacity and was rapidly delivered into the targeted fibroblast cells. Indeed, the loaded exosomes significantly enhanced the proliferation, migration, and invasion of human and mouse fibroblasts, which are vital steps in wound healing [95].
Another approach was to investigate the use of mi-RNA with drug co-delivery.
Triple-negative breast cancer corresponds to tumoral cells that lack human epidermal growth factor receptor 2, estrogen receptor, and progesterone receptor expressions. Therefore, only chemotherapy with platinum and untargeted chemotherapy are possible with inherent side effects. Increased binding to integrin αvβ3 was obtained through a continuous protein kinase C activation in monocyte-derived macrophages, allowing obtainment of exosomes with a high metalloprotein A15 presentation. These were co-loaded with Dox and the cholesterol-miR159-conjugate to induce synergistic therapeutic effects in MDA-MB-231 cells. In vivo, this nanovesicular miR159 and Dox delivery effectively silenced the TCF-7 gene without adverse effects [96].
Similar observations/conclusions were reported regarding the anticancer 5-fluorouracil when loaded with miR-21i into 5-FU-resistant colorectal cancer cell line HCT-1165FR-derived EVs, using electroporation. Systematic administration of the so obtained exosomes in tumor-bearing mice demonstrated an anti-tumor effect. Moreover, the combinational delivery of miR-21i and 5-FU with the engineered exosomes reversed drug resistance and significantly enhanced the death of 5-FU-resistant colon cancer cells, compared with the treatment with only miR-21i or 5-FU [97].
Finally, one recent publication reports the vectorization for glioblastoma treatment using the antisense miRNA oligonucleotides loaded into modified exosomes. To enhance glioblastoma targeting, the T7 peptide binding the transferrin receptor overexpressed on the surface of glioblastoma cells was used to decorate the exosomes membrane. The antisense against miR-21 (AMO-21) was loaded by electroporation. In vitro studies demonstrated an increased AMO-21 delivery to C6 glioblastoma cells and in vivo the induction of the expression of PDCD4 and PTEN in tumors, resulting in tumor size reduction [98].

3.2.3. Protein-Loaded EVs

As described for miRNAs or siRNAs, the nanoEVs can be loaded with proteins using the same methods (electroporation, freeze-thaw cycles, and/or sonication). The vectorization of several proteins, amino acids, peptides, or glycoproteins of therapeutic interest was investigated using modified nanoEVs.
Tripeptidyl-peptidase 1, also known as lysosomal pepstatin-insensitive protease (TTPI), was loaded into IC21 cell line (mouse peritoneal macrophage)-derived exosomes, using sonication. The EVs significantly increase the stability of TPPI against protease degradation and provide in vitro efficient TPP1 delivery to targeted neuronal ceroid lipofuscinose 2 cells. The majority of EV-TPP1 (≈70%) is delivered to target organelles and lysosomes. [99].
Nanovesicles were prepared from induced pluripotent stem cell-derived endothelial cells through an extrusion approach to function as exosome mimetics. Abundant membrane C-X-C motif chemokine receptor 4 allowed endothelial cell-targeting. Moreover, endothelial homology facilitated their accumulation in the cells. Dapagliflozin, promoting neovascularization in diabetic mice, was loaded into these bio-sourced artificial vesicles. The ability of induced pluripotent stem cells-endothelial cell nanovesicles to facilitate angiogenesis was conserved. Moreover, diabetic wound healing stimulated by HIF-1α/VEGFA pathway was demonstrated [100].
Interestingly, to increase their delivery on the injured zone, EVs were embedded into implantable biomaterials to investigate their potential for enzyme prodrug therapy. EVs derived from human mesenchymal stem cells were used as smart carriers for stabilizing enzymes in a hydrogel for a local conversion to active anti-inflammatory compounds. β-glucuronidase was co-incubated with MSC-derives EVs and incorporated into poly(vinyl alcohol) hydrogels to achieve site-specific release of curcumin (Cur) from its glucuronide precursor. The Cur formation was monitored using density-dependent color scanning electron microscopy imaging [101].
The efficiency of exosomes to deliver human insulin was studied on three different cell lines: hepatocarcinoma- (HepG2), dermal fibroblast- (HDFa), and pancreatic β (RIN-m)-cells. These cells were selected due to their glucose level regulation capacity. An autologous approach was developed, where the EVs derived from each cell line were loaded with human insulin by electroporation and internalized in cultures of their parent cells. The insulin internalization was monitored thanks to a fluorescein isothiocyanate labeling, and their glucose regulation effect was monitored by quantification of the glucose in the extracellular medium. Curiously and in contrast to HepG2 and HDFa, RIN-m cells were found to be insensitive to the loaded exosomes [102].
Electroporation was used to pack nanoEVs derived from donor platelets with a recombinant yes-associated protein 1. Yap1 is known to promote the stemness and differentiation potentials of tendon progenitor cells in vitro. These progenitor cells were exposed to the Yap1-loaded exosomes and differentiated toward tendon-like cells. They conserved their functions after long-term passage. The Yap1-exosomes were applied to progenitor tendon cells on a hydrogel to assist new tendon-like tissue formation in full-cut Achilles tendon defect [103].
A combination of freeze–thaw cycles and sonication was used to load human chorionic gonadotropin into exosomes isolated from uterine fluid. This hormone is a glycoprotein described to improve endometrial receptivity to implantation. The drug-release pattern and uptake of exosomes into the endometrial cells were evaluated and revealed that these loaded nanovesicles were functional [104].

3.2.4. Drug-Loaded EVs

The majority of recent publications in the loaded-exosome field are focused on the delivery of chemical drugs for several pathologies, but mostly aimed at cancer therapy.
One of precursor publications concerns a cell transfection with iRGD plasmid. iRGD is a 9-amino acid cyclic peptide (CRGDKGPDC) able to home onto tumor tissues and to bind and release doxorubicin (Dox). Exosomes were obtained from mouse immature dendritic cells that have been engineered by fusing the exosome membrane protein (Lamp2b) to αv integrin-specific iRGD peptide. Using electroporation, they were loaded with Dox. The last showed an efficient targeting and Dox delivery to αv integrin-positive breast cancer cells in vitro. When intravenously injected, these exosomes delivered Dox specifically to tumor tissues in BALB/c nude mice leading to the inhibition of tumor growth [68].
Later, Li et al. loaded one side Dox into exosomes isolated from A33-positive LIM1215 cells, and the other side coated the surface of superparamagnetic iron oxide nanoparticles with A33 antibodies. The interaction of the two produces a complex that demonstrated in vivo an excellent tumor targeting ability and the inhibition of the colon tumor growth [105].
Classically-activated M1 macrophages are known to suppress tumor growth. To enhance their tumor-suppressive effect, M1-exosomes were isolated from M1-macrophages and loaded with paclitaxel (Ptx), a taxoid chemotherapeutic drug, by a slight sonication. Mice bearing 4T1 breast-tumors were used to perform the therapeutic effect of Ptx-M1-exosomes in vivo, and their cytotoxicity was studied. The M1-exosomes provided a pro-inflammatory environment and exhibited an anti-tumor effect in vivo. When loaded with Ptx, the anti-tumor effects were increased [106].
Exploiting the capacity of exosomes to easily penetrate the blood–brain barrier, Ptx-loaded embryonic stem cell-derived exosomes were prepared to treat glioblastoma. The anti-glioblastoma effect of these embryonic stem cell exosomes was demonstrated. Then, exosomes were engineered by introduction of c(RGDyK) and loaded with Ptx. Their efficiency to deliver drug in glioblastoma mode was compared to the drug alone or the drug-loaded exosomes. The loaded engineered exosomes were revealed to be the best vehicles to target the glioblastoma cells [107].
To enhance pancreatic cancer therapy, surface-modified EVs were loaded with Ptx. For this purpose, human pancreatic cancer cell-derived EVs were collected. The functional ligand RGD, binding specifically to αvβ3 integrins highly expressed in pancreatic cancer cells, was conjugated to the EV surface. Ptx-loaded nanovesicles were intravenously injected into xenograft mice and produced a significant reduction in tumor size compared to the free Ptx. Interestingly, comparing the effects of RGD-modified and Ptx-loaded EVs derived from colon adenocarcinoma HT29 cells with those derived from pancreatic cancer cells, it was demonstrated that autologous EVs present more benefits [108].
Exosomes derived from human umbilical cord-derived mesenchymal stem cells promote the migration of neural stem cells NSCs in vitro and in vivo. Therefore, a collagen-scaffold was designed, including these nanovesicles to attract NSCs to enhance the therapy after complete spinal cord injury. A dual bio-specificity peptide allowed the effective retention of the exosomes within scaffolds. To complete the neural regeneration and reduce scar deposition, chemotherapy was proposed by loading these exosomes with Ptx during the extrusion process [109].
Gemcitabine (GEM), a nucleotide analog, is the first-line chemotherapeutic drug for pancreatic cancer treatment. Unfortunately, it is associated with important side effects when administered systemically. Therefore, it was proposed to encapsulate GEM within exosomes. The authors demonstrated that for pancreatic cancer, autologous exosomes (cancer cell-derived) facilitate the tumoral uptake of GEM and significantly increase its cytotoxic effect [110].
Immunotherapy in treating pancreatic ductal adenocarcinoma (like chemotherapy) does not provide satisfactory outcomes for patients due to the unique tumor microenvironment and low cancer immunogenicity. The retained approach was to reverse the tumor immunosuppression of M2-like tumor-associated macrophages through disruption of the dectin 1/galectin-9 axis. Bone marrow mesenchymal stem cell-derived exosomes were loaded with galectin-9 siRNA, by electroporation, and surface modified with oxaliplatin to assess whether they can act as an immunogenic cell death trigger. The authors found that these elicit tumor-suppressive macrophage polarization, cytotoxic T-lymphocyte recruitment, and Tregs down-regulation [111].
Ferroptosis is a form of cell death recently reported during exposure to erastin, characterized by an alteration of the cellular redox metabolism, resulting in massive lipid peroxidation in the plasma membrane. Exosomes obtained from HEK293T cells—transfected with CD47-overexpressing plasmid—were isolated and loaded with erastin and Bengal Red via sonication. Hepa1-6 cell xenograft C57BL/6 mice were injected with these engineered exosomes and the photodynamic therapy was conducted by a 532 nm laser irradiation. The CD47 surface functionalization allowed the exosomes to escape phagocytosis. In addition, the embedded erastin induced strong ferroptosis in vitro and in vivo in hepatocarcinoma cells after irradiation [112].
Chronic liver injury and liver failure caused by acute liver injury remains challenging to treat. The potential of quercetin (a flavonoid) and vitamin A loaded into adipose mesenchymal stem cell-derived exosomes was investigated to improve this fate. It was reported that vitamin A enhances the liver targeting of exosomes in acute liver injury in mice (induced by CCl4). The quercetin and vitamin A loaded mesenchymal stem cell exosomes reduced the rapid senescence-like response induced by acute liver injury [113].
Due to their high production capacity and biocompatibility, milk EVs have recently been considered to improve oral drug bioavailability [114,115,116].
The delivery to breast tissue and the anticancer activity in cell lines of free Cur and resveratrol (Rsv) were compared with encapsulated Cur and Res in milk-derived exosomes. The experiments were carried out on MCF-7 and MDA-MB-231 breast cancer and also on MCF-10A non-tumorigenic cells. The loaded milk-exosomes entered the cells primarily via clathrin-mediated endocytosis, avoiding ATP-binding cassette transporters (ABC). The milk exosomes were capable of protecting Cur and Rsv from metabolism and to deliver the polyphenols to the mammary [117].
The ability of exosomes to cross the blood–brain barrier and target the recipient cell was taken in account to develop a theragnostic strategy for glioma. Superparamagnetic iron oxide nanoparticles and Cur were placed into exosomes to confer imaging and therapeutic functions. After loading, the exosome membranes were chemically modified with neuropilin-1-targeted peptide (RGERPPR, RGE) to target glioma. When administered to glioma cells and orthotopic glioma models, these provided good results for imaging and therapy of glioma through magnetic flow hyperthermia and Cur chemotherapy [118].
However, many drugs have poor oral bioavailability, warranting the use of injections. The capacity of cow milk EVs and intestinal epithelial cell-derived EVs to deliver oral drugs was studied. EVs, fluorescently labelled and loaded with Cur were tested in an intestinal model (Caco-2). Epithelial cell-derived EVs showed notably higher cell uptake compared to cow milk EVs. However, the cell uptake was significantly higher in differentiated compared to undifferentiated cells for both types of EVs [119].
Finally, a singular work reported on the vectorization of antibiotics using nanoEVs. Staphylococcus aureus survivals inside phagocytes constitute bacterial reservoirs relatively protected from antibiotic treatments. Linezolid was loaded into exosomes produced by mouse RAW264.7 macrophages to overcome intracellular infections by pathogenic bacteria. This exosomal formulation of linezolid was effective in vitro and in vivo without evident cytotoxicity in macrophages [120].

4. Artificial Mimes for Exosomes

Despite the large number of studies carried out using exosomes isolated from cell cultures, one cannot ignore the major issues linked to their production and purification.
One of the major problems encountered for exosome-based drug-delivery systems is the low productivity of exosomes. Indeed, the typical reported production yields are maximally of 1 μg exosomal protein/mL of culture medium [121], whereas the typical therapeutical dose used was of 10–100 μg exosomal protein/mouse [122]. EVs are usually prepared by culturing the parent cells and seem to reach their upper production-limit after around 12 h of incubation (time depending on the parent cells) [123]. Bioreactors are reported to improve the yield of exosomes [124,125,126]. However, this quantitative improvement comes at the expense of quality, since larger vesicles are obtained. Another way to enhance the production yield is to apply stresses (hypoxia, low pH, drugs, …) on the parent cells. Unfortunately, it was demonstrated that cellular stresses changed the composition of exosome cargo and surface [127]. Moreover, these increased contaminants such as apoptotic bodies.
Conventionally, exosomes are isolated through ultracentrifugation-based methods, but due to their size and density, these are contaminated with proteins and are often fusing or aggregating under extreme physical constraints [121]. Therefore, other exosome-isolation methods have been explored, such as size exclusion chromatography (SEC), ultrafiltration (UF), or polymer-based precipitation (PP). More adapted for the large-scale production of exosomes, these also have disadvantages (UF—low purity, damages EVs by shear stress; SEC—low capacity, high dilution of samples; PP—low purity, severe EV aggregation) [128,129,130]. The recent technical evolution involves microfluidics [131].
Another issue comes from the observation of subpopulations within the nanoEVs, which differed in their composition and size [132]. Different compositions of proteins and RNAs in exosomes, dependent on the collection method, were reported [133,134]. This can be explained by enrichments in various subpopulations, contaminants or by the presence of the protein corona that is adsorbed on the nanoparticle surface along the purification/isolation procedure. One has to note that it is unclear whether the observed protein coronas are natively present on exosomes or are the result of the isolation procedure. In any case, they clearly change the properties of nanoparticles [135,136]. Therefore, optimization of the isolation method is important to preserve the properties of exosomes and to reduce the risk of side effects induced by contaminants or specific subpopulations.
Alternative solutions have been proposed considering the (hemi)synthetical production of exosome-like structures (ex vivo).
Lipid-based nanovesicles are difficult to obtain through spontaneous organization, especially when membrane proteins must be integrated. As the presence of membrane proteins is necessary for cellular targeting and internalization, innovative strategies were elaborated to include recognition elements on the surface of synthetic nanoparticles (Figure 3).
The most exploited liposome-alternative approach consists of the so-called cellular nanoparticles (cNPs). These are cell-membrane coated nanoparticles produced by a series of sonication and extrusion of cell membrane debris and vesicles. Indeed, this approach allows the cNPs to contain the lipids and proteins of the source cells and consequently exhibit several properties of the exosomes and parent cells. These have a potential variety of biomedical applications, including drug delivery, phototherapy, combination therapy, immune modulation, sensing, and diagnosis [137,138,139]. For example, it was reported that platelet-derived cNPs still interact with circulating tumor cells and bind to the vessels and tissues [140,141,142]. It was also observed that cNPs obtained from M1 macrophage membranes can repolarize tumor activated macrophages to an M1-like phenotype [106,143].
Interestingly, based on exosome’s spherical structures, the designed mimics often have the same shape. Convinced that rod structures are more appropriate to penetrate tissues and to be endocytosed, Zhang et al. developed cancer cell membrane-coated nanoparticles with the two shapes. Thanks to the caveolin-mediated pathway, the coated nanorods were more efficient for endocytosis and in the endoplasmic reticulum region accumulation compared to the spherical ones [144].

4.1. Cell-Membrane Covered Napoparticles

4.1.1. Single-Membrane Covering

Artificially produced bio-sourced nanoparticles have received growing attention in drug delivery. Because of anti-cancer drugs’ toxicity, the tumor-specific delivery enabled by the cNPs constitutes the most part of the published exosome-mime applications, as they allow the delivery of the drugs of interest at much lower concentrations to target sites, with limited side effects (see also loading part) (Figure 3).
As erythrocytes represent the most abundant blood circulating cells, these are the first biological source of reported cellular NP-coatings but are still under scope to ensure a ready clinical translation. Indeed, a higher membrane extraction rate must be obtained for a potential scaling-up, and the encapsulation efficiencies should be improved to provide batch-to-batch reproducibility [145]. Early in 2013, polymeric nanoparticles were masked using erythrocyte membranes [146] and were used to deliver drugs, especially Dox and Ptx [147,148,149]. Among them, the most surprising approach was the direct encapsulation of Ptx spherical crystals with erythrocyte membranes. Due to the high drug content, the authors achieved a stronger cytotoxicity and higher drug accumulation in tumor. Therefore, they claimed that it is the most effective Ptx formulation against tumor growth among all in tumor-bearing mice models [149].
Table 2. Drug-loading methods for nanoEVs found in the literature.
Table 2. Drug-loading methods for nanoEVs found in the literature.
Loaded MoleculesExosome Cell LineLoading MethodsMax Loading Efficiency/CapacityTargeted Pathology/FonctionRef.
siRNA MAPK1HeLa, HTB 177, human blood cellsElectroporation: 150 V-100 µFNSCancer therapy[84]
siRNA luciferase Primary endothelial cell mice Electroporation: 400 V-200 µF 25% Drug absorption [85]
siRNA RAD51 and RAD52HeLa, ascitesElectroporation: 700 V-350 µFNSDrug absorption[86]
siRNA FGL
siRNA TGF-β1
RAW 264.7Transfection with commercial reagent97%Cancer immunotherapy[87]
siRNA KRASG12DHuman MSCElectroporation: 400 V-150 µFNSPancreatic cancer[88]
siRNA CTGFRat MSCElectroporation: 400 V-200 to 500 µF33%Cancer therapy[89]
siRNA CCL2Induced neural SC, surface modified with CAQKElectroporation: 400 mV-125 μF (pulse time 10–15 ms)17%Spinal cord injury[90]
siRNA DNSL1; SND1; ENDD1; RINI; 5NTD; ENPP1; CN37Umbilical cord Wharton’s jelly mesenchymal stem cellsCo-incubation, 37 °C, 1 h18%Huntington’s disease[93]
miRNA 484HEK293TElectroporation: 700 V-150 µFNSMetastatic ovarian cancer[94]
miRNA 21-5p (conjugated to CPP)Mouse embryonic fibroblasts (MF)
Human primar dermal fibroblasts (HDFa)
Co-incubation, 37 °C, 8 h1600 copies/exoLoading optimization[95]
miRNA 159
doxorubicine
Human monocytes THP-1Co-incubation, 37 °C, 1 h 3016%Breast cancer[96]
miRNA 21 inhibitor
5-FU (5-fluorouracile)
Transfected HEK293T cellsElectroporation: 1000 V-15 µF, 10 ms3%Colorectal caecinima[97]
miRNA 21 antisensTransfected HEK293T cellsElectroporation: 400 mV-125 μF1.7%Glioblastoma[98]
TPP1-encoding pDNAMacrophagesSonication; saponin-permeationNSLate-infantile neuronal ceroid lipofuscinosis[99]
DapaglifloziniPS-derived ECsExtrusion28%Diabetic wound healing[100]
β-glucuronidaseHuman MSCsSaponin permeation: 0.1 mg/mL, 10 min, RTNSEV-based hydrogel-developments[101]
Human insulinHepatocellular carcinoma (HepG2)
Primary dermal fibroblasts (HDFa)
Co-incubation, RT, 1 h
Electroporation: 200 V-50 µF
3%
50%
Diabetes melitus treatment[102]
Yap1 proteinRat plasma plateletsElectroporation: 200 V-500 µF, 26 msNSAchilles tendon injury[103]
Human chorionic gonadotrophinHuman uterine fluidFreeze-thaw cycle and sonication40%Assisted reproductive technology[104]
DoxorubicinHuman breast cancer cells (MDA-MB-231, MCF 7)Electroporation: 350 V-150 µFNSTumor therapy[68]
DoxorubicinHuma CRC cells LIM1215Incubation: 5 min
Dialyse PBS overnight
9%Colorectal cancer[105]
PaclitaxelM1 macrophages RAW 264.7Sonication: 20% amplitude; 6 cycles of 30 sNSCancer therapy[106]
PaclitaxelESC H9 cellsIncubation: 2H; RTNSGlioblatoma[107]
PaclitaxelPANC-1 (pancreatic ductal carcinoma) cells; U937Co-incubation, RT, 1 h and sonication37%Pancreatic cancer[108]
PaclitaxelHuman umbilical cord-derived mesenchymal stem cellsExtrusion14%Spinal cord injury[109]
GemcitabinePancreatic cancer cellsIncubation: 2H; 37 °C
Sonication: 20% amplitude; 3 cycles of 30 s
12%Pancreatic cancer[111]
siRNA galectin-9
oxaliplatin
BM MSC cellsElectroporation: 400 V-125 µF13%Pancreatic cancer[112]
Erastin
rose bengal
CD47-overexpressing HEK293TSonication60%
84%
Hepatocarcinoma[113]
Vit A + quercetinMice MSCIncubation: 2 h; 22 °CVit A: 0.279 µg/mL exo
Que: 0.19 µg/mL exo
Acute liver injury[114]
Curcuma resveratrolRaw bovine milkIncubation: 1 h, 37 °C darkness
Sonication: 1 Hz; 150 W; 4 or 6 cycles of 20 s.
Electroporation: 400 V; 2 pulses
Cur 2 µg/0.5 ng prot exo;
Res 15 µg/0.8 ng prot exo
Breast cancer[119]
CurcumaCow skimmed milk Intestinal cells (Caco-2)Incubation: overnight; RT
Size exclusion chromatography
Cur:9%
Caco-2: 3%
Drug absorption[149]
NS: not specified, RT: room temperature.
Since immunotherapy remains a powerful cancer therapy, the first generation of cNPs were designed using immune-cell membranes.
Employing the “cell ghosts” (nanoghosts) production protocol, core-shell nanoparticles in the 100–200 nm size-range were obtained using serial extrusions of cellular membranes through a polycarbonate membrane. The coating of monocyte cell membrane on a biodegradable poly(latic-co-glycolic acid) core formerly loaded with Dox was reported, and the intracellular uptake of these cNPs into metastatic breast niches was tested using MCF-7 cells. Comparing FITC-loaded cNPs with FITC-loaded PLGA NP controls, a higher cellular uptake was observed for the first ones. No observable toxicity towards MCF-7 free cells could be seen [150].
The same core-shell disposal was reported, exploiting this time the property of inflammatory neutrophils to target the circulating tumoral cells and the metastatic niche. The neutrophils membrane associated with the poly(latic-co-glycolic acid) nanoparticles demonstrated a highly preserved binding capacity to 4T1 cell models in vitro, and a high circulating cancer cells capture efficiency in vivo. Moreover, these cNPs were found to have improved homing to the premetastatic niche. When loaded with carfilzomib, they depleted the tumoral cells circulating in the blood selectively [151]. Finally, similar cNPs were reported to target hepatocarcinoma cells. The drug-free coated nanoparticles were described as highly biocompatible with a cancer-cell specificity. When loaded with Cur, the cNPs were revealed to be effective cancer-cell killers with a potent anti-tumor activity in a murine H22 model [152].
In addition to the cancer studies, a single report on HIV has been published. T-cell-membrane-coated nanoparticles with a polymeric core have been designed. These still present the T-cell antigens that are critical for HIV binding (CD4 receptor and CCR5 or CXCR4 coreceptors). The cNPs demonstrated a selective binding with the HIV glycoprotein gp120, inhibiting the gp120-induced killing of bystander CD4+ T-cells by a decoy strategy [153].
In the literature, tumor-addressing is almost ensured through a cancer cell membrane encapsulation or core coating.
A mesoporous silica nanoparticle-supported PEGylated liposome core has been developed, enabling the coating with various cancer cell membranes. The moderate rigidity of these cNPs cores facilitated their penetration throughout multicellular spheroids in vitro, indicating a probable facilitated infiltration into the tumor. They were directly internalized by membrane fusion, and the PEGylated cores were released into the cytosol. Their ability to co-encapsulate Dox and mefuparib hydrochloride has also been tested [154].
Recently, in addition to the NP masking with glioma C6 cancer cell membranes, the retroenantiomer of quorum-sensing peptide (WSW)—which can efficiently cross the blood–brain barrier—has been inserted. The so-obtained nanosuspensions were not cleared by the immune system, could cross the blood–brain barrier, and targeted the tumor selectively [155]. Another way to cross the blood–brain barrier for the cNPs was followed by exploiting the penetration ability of metastatic 4T1cancer cells. Following this approach, a biomimetic nanoplatform was designed by masking a succinobucol-loaded pH-sensitive polymeric particle with a 4T1 cell membrane to promote the preferential targeting of cerebral ischemic lesions [156].
To obtain nanoparticles capable of targeting oral squamous cancer cells, membranes of mesenchyme stem cells obtained from dental pulp were used to cover metal-organic nanoparticles. As they present the CXCR2 receptor, these particles possessed the attempted specificity for the targeted cancer cells. Moreover, when loaded with Dox, the cNPs could induce CAL27 cell death in vitro and blocked CAL27 tumor growth in vivo [157].

4.1.2. Hybrid Membrane Covering

The production of nanovesicles derived from two different types of cells has been published, resulting in a novel cNP family called hybrid nanovesicles, that inherit the characteristics from the two source cells (Figure 3).
cNPs derived from monocytes were used to deliver Dox into SKOV-3 ovarian cancer cells. Using confocal microscopy and flow cytometry the authors proved that such mimetic disposals are more efficient than free Dox, by reducing side effects and increasing its cytotoxicity for the targeted cells. The comparison of exosome mimes with the natural exosomes produced by these cells demonstrated the conservation of the exosome markers profile (CD81, CD63) with a reproducible and 2.5 times more efficient production yield for the mimes [158].
Although chemotherapy is a promising strategy for bone invasion treatment, its clinical applications are limited by the lack of tumor-specific targeting and poor permeability in bone tissue. Therefore, developing a smart bone and cancer dual targeting drug-delivery platform is necessary. Spherical nanoparticles designed for bone and cancer targeting were covered with a head and neck squamous cell carcinoma WSU-HN6 cell (H) and red blood cell (RBC) hybrid membrane. These were modified by an oligopeptide of eight aspartate acid (Asp8) and exhibited the same surface proteins as those of WSU-HN6 and RBC. In vivo, these nanoparticles were mainly localized proximal to the cancer cells. Moreover, they showed effective cancer growth-inhibition properties when compared to other formulations [159].
Another idea was followed by preparing EV mimics presenting the macrophage signal regulatory protein alpha SIRPα (SαV-C-nanovesicles) to disrupt the CD47-SIRPα axis releasing a “do not eat me” signal to prevent the phagocytosis of cancer cells. As the authors aimed to repolarize TAMs towards the M1 phenotype simultaneously, they designed hybrid cell membrane nanovesicles displaying SIRPα variants and containing M2-to-M1 repolarization signals. These hybrid cell-membrane covered nanovesicles lead to potent tumor inhibition in a poorly immunogenic triple negative breast cancer model [160].

4.1.3. Membrane Covering of Organometallic Cores

Besides the structural function of the solid nanoparticles allowing for a fixed size and shape of the cNPs, some solid inner cores were also designed to play the therapeutic role.
In 2017, an early publication reported the possibility of using platelets as carriers for gold nanorods. These were loaded into platelet cells by electroporation. The resulting loaded-gold particles retain the long blood circulation and cancer-targeting characteristics of the unloaded cells. They also present the good photothermal property of the gold rods [161].
Later, the same group designed a platelet-cancer stem cell hybrid membrane-coated iron oxide magnetic nanoparticle. The platelet membrane conferred the immune evading ability, and the cancer stem cell membrane conferred the homotypic targeting capabilities, thanks to specific surface adhesion molecules. The hybrid covering of the magnetic particles enhanced the photothermal therapy of head and neck squamous cell carcinoma, inhibiting tumor growth in mice [162].
Coating upconverting particles—used for photodynamic therapy and imaging purposes—with macrophage membranes makes it possible to utilize the adhesion between macrophages and cancer cells for effective cancer targeting. Natural macrophage membranes, along with their associated membrane proteins, were reconstructed into vesicles and then coated onto synthetic upconverting NPs. The resulting macrophage membrane-camouflaged particles exhibited effective cancer-targeting capability inherited from the source cells and were further used for enhanced in vivo cancer imaging, with a good in vivo biocompatibility [163]. Indeed, the introduction Fe2O3 in iron metal nodes can effectively catalyze the Fenton reaction to produce hydroxyl radicals (OH) and overcome the hypoxic environment of tumor tissue by generating oxygen, improving tumor therapy. Such organometallic framework nanoparticles were camouflaged with erythrocyte membranes to enhance blood circulation and tissue residence time in the body. Moreover, the aptamer of the targeted AS1411 molecule was added to the surface to address the photosensitizers on the tumor domain. It was reported that the camouflage by the erythrocyte membrane could reduce side effects and improve the therapeutic effect of photo-dynamic therapy and chemo-dynamic therapy [164].
Even if the membrane camouflage remains a consensual approach, conjugates of small anticancer drug molecules with the polyzwitterion poly(2-(N-oxide-N,N-diethylamino)ethyl methacrylate) were recently designed, exhibiting a long blood-circulation half-lives and producing negligible interaction of the polyzwitterion with proteins and phospholipids. Adsorption of the polyzwitterion–drug conjugates to tumor endothelial cells and then to cancer cells favored their infiltration into tumors in mice. The simplicity and potency of the polyzwitterion–drug conjugates should facilitate the design of translational anticancer nanomedicines [165].

4.2. Exosome–Liposome Hybrids

As evoked above, the drawbacks of developing exosomes for drug-delivery purposes are related to their low biological production, the poor yield of the current isolation techniques, their limited colloidal stability at ambient environment, and their aggregation occurring during storage. The polymeric or metallic exosome-alternative nanoparticles present better stabilities and higher loading capacity but have immunogenicity, long circulation, and toxicity issues that were partially addressed through cell membrane covering strategies. Moreover, the limitations of conventional isolation techniques, such as ultrafiltration, ultracentrifugation, precipitation, microfluidics-based separation, and immune affinity-based separation, are numerous. Indeed, aggregation phenomenon, loss of integrity of exosomes due to mechanical stress, and low purity due to entrainment or adsorption effect of proteins or small hydrophobic molecules are hazards often reported [166,167].
To limit these drawbacks, fused liposome–exosomes are proposed alternatives for developing drug-delivery systems. Indeed, these artificial structures bring together the biological functions, targeting, and high loading capacities of exosomes, with the improved colloidal stability, and the easy surface modification of liposomes. Owing to their similar bilayer lipidic structures, exosomes and liposomes can easily fuse into hybrid exosomes (HEs) [168].
Despite the fact that HEs were supposed to alter the efficiency of the interactions with the cells—considering the modification of the membrane lipids composition—several studies demonstrated their high potential for DDS (Table 3).

4.2.1. Hybridization Techniques

The freeze–thaw method is a valuable reported method to fuse membranes with liposomes. Its principle is based on the temporary disruption of the plasma membrane by ice-crystal formation, allowing water-soluble molecules to penetrate the structure. A high fusion efficiency is reported. Usually, cell-derived exosomes are mixed with the liposomes in a 1:1 ratio. The mixture is then frozen in liquid nitrogen and thawed subsequently at room temperature. The duration and number of freeze–thaw cycles are variable (e.g., [169,170]). The disadvantage of this method is that the high-frequency freeze–thaw cycles affect the biological activity of pharmaceuticals and also alter the integrity of the exosome membrane.
Therefore, some authors preferred natural incubation, to preserve the vesicles and drugs, but the fusion efficiency under these conditions is relatively low and time-consuming (e.g., [171]).
The most frequently used technique remains extrusion, where exosomes and liposomes simultaneously pass through membrane pores under physical pressure to form mixed vesicles at the exit. The principal advantage resides in the size homogeneity of the hybrid vesicles. Usually, the two ingredients are vortexed and sonicated before the extrusion through 100–400 nm pore-sized polycarbonate ester films. The extrusion has to be repeated (up to 10 times) [172].

4.2.2. RNA-Loaded HEs

The limited drug-loading capacity of exosomes and the unclear intracellular transport mechanism may affect the transfection efficiency of exogenous nucleic acids. Therefore, the integration of artificially synthesized exogenous lipids into isolated exosomes to deliver siRNA was expected to provide a promising therapy for genetic diseases.
Considering that the complex components of tumor cell-derived exosomes contain key components of carcinogenesis, membrane fusion hybrid exosomes were engineered by fusing membranes of EVs derived from SK-Hep1 hepatic adenocarcinoma cells with phospholipids. Cyclin-dependent kinase 1 (CDK1)-siRNA (which inhibits the CDK1 gene to kill c-Myc-overexpressing hepatocarcinoma cells) was loaded using electroporation. Hybrid exosomes achieved stable siRNA encapsulation, precise tumor targeting, and effective gene silencing by the delivered CDK1-siRNA [173].
Osteoporosis is a public health problem linked to the aging population. In this context, NIH-3T3 cells were genetically engineered to enhance their CXCR4 expression. The CXCR4+ exosomes derived from these cells were fused with liposomes carrying the miRNA-188 inhibitor to produce hybrid nanoparticles. These HEs specifically accumulated in the bone marrow and released antagomir-188, thus promoting osteogenesis and inhibiting adipogenesis in bone marrow MSC [174].
Similarly, siRNA-AF647 was added during the hydration of the liposome film. Then cardiac progenitor cell-derived exosomes were mixed with the loaded liposomes, and hybrid exosomes were obtained through extrusion. The designed hybrid exosomes reduced the toxicity of liposomes and retained the gene silencing effects. Upon fusion, the properties of cardiac progenitor cell-derived exosomes were also preserved, such as the ability to activate endothelial signaling pathways and stimulate microvascular endothelial cell migration, which may have therapeutic potential for salvaging myocardial tissue upon infarctions [175].

4.2.3. Protein-Loaded HEs

The clustered, regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9) system has become essential for in vivo gene edition. Nevertheless, challenges persist in its clinical application. The CRISPR/Cas9 system is mainly delivered to the body through viral vectors, questioning its clinical application. The expression plasmid of Cas9 is much larger than the small nucleic acids usually packed into exosomes. The delivery of such large plasmids via hybrid exosomes was studied based on the high plasticity of liposomes. Large nucleic acids, including CRISPR/Cas9 expression vectors, could be encapsulated into hybrid exosomes derived from mesenchymal stem cells, opening new possibilities for genetic manipulations in vivo [172].

4.2.4. Drug-Loaded HEs

Hybrid nanovesicles obtained by fusing liposomes and tumor-derived extracellular nanovesicles (4T1 breast cancer and B16F10 melanoma cells) were successfully developed to deliver Dox for immune chemotherapeutic purposes. This engineered drug-delivery carrier obtained by extrusion techniques produced effective immunogenic cell death induction for the tumor microenvironment. The hybrid nanovesicles loaded with Dox exhibited synergistic antitumor effects on multiple tumor models [176]. Compared to the loaded tumor-derived exosomes, the lipid-modified exosomes displayed long-term colloidal steadiness and high encapsulation efficacy without any apparent burst release of drug under normal physiological conditions [177].
Rayamajhi et al. were the first to fuse mouse macrophage J774A.1-derived exosomes with liposomes and to load them with water-soluble chemotherapeutic drug Dox. Taking advantage of the tumor-targeting properties of macrophage membranes, the Dox-HEs demonstrated an enhanced targeting ability and cytotoxicity to J774A.1, K7M2, 4T1, and NIH/3T3 cells, when compared to Dox-loaded liposomes [178].
Considering the exosome capacity to cross the blood–brain barrier, HEs obtained from the sonication-induced fusion of salvianolic acid B-loaded blood exosomes with cryptotanshinone (CT)-loaded liposomes were proposed to treat glioma. On the one hand, the exosomes were isolated from rat serum, and the SAB encapsulation was obtained via electroporation. On the other hand, truncated Lyp-1 peptide CTP was included during DSPE/PEG-liposome formation. The tLyp-1 peptide enabled the capture and accumulation of the HEs in vitro in human glioblastoma astrocytoma cells (U87) and in deep tumor regions in vivo on BALB/c nude mice bearing U87-Luc xenograft tumors, resulting in intensive drug accumulation at glioma sites. As these HEs transported bi-therapeutic agents (SAB and CPT), synergistic anti-glioma mechanisms with two separate compartments of glioma, cancer cells and angiogenesis, were observed [179].
Pulmonary fibrosis is a progressive and fatal lung disease, but most drugs cannot effectively hinder its progression due to their poor organ targeting. Therefore, hybrid exosomes were designed by fusing chlodronate-loaded liposomes and fibroblast-derived exosomes. The loaded clodronate depleted hepatic macrophages to reduce the hepatic uptake and utilized the homing of autologous exosomes to deliver the anti-fibrotic drug Nintedanib (Nin). The authors reported the massive accumulation and penetration of HEs in pulmonary fibrosis, enhancing the suppressive effect of Nin on fibrosis [180].
Similarly, this team also developed HEs to increase, this time, the efficacy of the chlodronate inhibition of Kupffer cells and to deliver selectively Nin to liver fibroblasts [181].
Finally, another type of cancer bi-therapy combining a chemotherapeutic drug and a miRNA was proposed using HEs. Indeed, to circumvent ovarian cancer resistance to chemotherapy during chemotherapy, SKOV3-CDDP tumor cell-derived exosomes were fused with liposomes including tumor targeting peptides. The HEs were engineered to encapsulate triptolide and miRNA-497. Tumor targeting was ensured through homologous targeting, CD24-overexpession on the exosome membrane, and by the inclusion of cRGD peptide in the liposome part of the hybrid structure. It was reported that these hybrid nanoparticles effectively targeted tumor sites. In the tumor microenvironment, miR497 and tripotolide were rapidly released, inducing synergistically the apoptosis of ovarian cancer cells and the inhibition of the PI3K/AKT/mTOR signaling pathway [182].
Photothermal therapy (PTT) is a recent but widely used cancer therapy. It can destroy tumor cells through hyperthermia but also achieve effective controlled release of drugs. Hyperthermia is generated by a laser irradiation of the tumoral accumulated photothermal agents.
It was proposed to engineer CD47-overexpressing CT26 colorectal carcinoma cells to produce CD47-enriched EV membranes capable of targeting carcinoma tumors and ensuring safe circulation in blood. In parallel, a liposome containing indocyanine green as a photothermal agent and R837 (an imidazoquinoline amine) as an immune adjuvant was produced. Finally, the engineered EVs and liposomes were fused. The so obtained HEs were described as taking advantage of the two systems: the stable blood circulation, tumoral targeting, improved the macrophage-mediated phagocytosis of tumor cells (by blocking CD47 signaling) with a photothermal destructive capacity [159].
Following the same strategy, platelet-derived exosomes were combined with photothermal-sensitive DPPC/DPPG liposomes, encapsulating this time glucose oxidase (GOx) and ferric ammonium (FA). GOx was used to oxidize glucose to produce hydrogen peroxide and FA to catalyze cascade reactions, enhancing the chemotherapy efficacy. Moreover, the photothermal effect further enhanced the therapeutic effect. This hybrid system was reported to significantly reduce tumor size and prolong lifespan in treating primary or metastatic tumors [183].
Unfortunately, ovarian hyperthermic chemotherapy is not efficient within the peritonea, as the chemical agents are not capable of penetrating the large tumors. To overcome this issue, BALB-3T3 fibroblasts were genetically engineered to produce CD47-rich exosomes, and thermosensitive liposomes were loaded with granulocyte-macrophage colony stimulating factor or docetaxel. The two were freeze–thaw fused. It was reported that, after intravenous administration, the HEs accumulated in tumors and released their therapeutic ingredients. The intraperitoneal chemotherapy of ovarian cancer was described as accelerated under the hyperthermic condition. These hybrid exosomes effectively inhibited the tumor progression [184].
Table 3. Exosome hybridization and loading methods found in the literature.
Table 3. Exosome hybridization and loading methods found in the literature.
Exosome-Donor CellsLiposomesCargoFusion TechniqueApplicationRef.
RAW 264.7 macrophages
CMS7 sarcoma fibroblast
DOPC or DOPSFluorescent DMPEFreeze–thawSetting-up hybridation[170]
HEK293 T
embryon cells
Not mentionedCrispr/Cas9Co-incubationGene therapy[172]
A549 and 3C3 cancer cellsDOPC/DOPSAnti siRNA GFPMembrane extrusionSetting-up for gene delivery[173]
K-HEP1 cancer cells DPPCsiRNA CDK1Membrane extrusionHepatocarcinoma[174]
NIH-3T3 CD24++
cancer cells
DOTAP/cholesterolAnti miRNA 188Membrane extrusionOsteoporosis[175]
SKOv3 cancer cells
CP cardiomyoblasts
DLin-MC3-MA/DPPC/cholesterol/18:1-BiotinylCapPE/DMG-PEGsiRNA AF647or AF488Membrane extrusionCardiac injury[176]
4T1 breast cancer cells
B16F10 melanoma
DOPC/cholesterol/mPEG2000/DSPEDoxorubicinSonication and extrusionCancer immunotherapy[177]
J774A.1 macrophagesEgg-PC and cholesterolDoxorubicinMembrane extrusion [179]
Rat blood-serumDSPE-PEG2000-tLyp-1SAB in exosome
tLP1 and CTP
Membrane extrusionGlioma[180]
L-929 fibroblast
raw 264.7 macrophages
Cholesterol/DOPC/DSPE-PEG2000CLD and NinMembrane extrusionPulmonary fibrosis[181]
L-929 fibroblastCholesterol/DOPC/DSPE-PEG2000CLD and NinMembrane extrusionLiver fibrosis[182]
SCOV3-CDDP CD24++
cancer cells
DSPE-PEG1000-COOHcRGT, tripotolide and miRNA-497Membrane extrusionOvarian cancer[183]
Mouse blood plateletsThermosensitive liposome: DPPC/DPPG/cypateGlucoxidase and ferric ammoniumPEG mediationHepatocarcinoma[184]
3T3 CD24++
cancer cells
Thermosensitive liposome:
CFL/DPPC/MSPC/DSPE-PEG2000
GM-CSF and docetaxelFreeze–thawMetastatic peritoneal carcinoma[50]
DOPC—dioleoylphosphatidylcholine; DOPS—dioleoylphosphatidylserine; DPPC—dipalmitoylphosphatidylcholine; DOTAP—1,2-dioleoyl-3-trimethylammonium propane; DLin-MC3-MA—(6Z,9Z,28Z,31Z)-heptatriaconta-6,9,28,31-tetraen-19-yl 4-(dimethylamino)butanoate; DMG—dimethylglycine; PEG—polyethylenglycol; DSPE—disterarylphosphatidylethanolamine; DPPG—dipalmitoylphosphoglycerol; MSPC—myristoylstearoylphosphatidylcholine; CFL—cerasome forming lipid.

5. Perspectives

Since exosomes were identified quite early as high-potential drug-delivery vehicles, this review primarily focuses on recent nanovesicular loading strategies for nucleic acid, protein drugs, and small molecules. However, the last decade demonstrated that exosome-mediated drug-delivery methods suffer from low mass production, delivery rates, and high cost. One supplementary important issue remains that the specific functions of exosomes are not fully understood, making it challenging to ensure their long-term safety and efficacy. Moreover, one must remember that exosomes have different membrane and cargo compositions and biological functions, depending on source cells and environmental conditions, inducing reproducibility issues. For all these reasons, they cannot be considered as simply acting like empty carriers, especially when they are derived from circulating cancer or infected cells.
Therefore, the potential of exosome-mimes or analogues was studied more recently. Nevertheless, considering exosomes or exosome-like vesicles, the loading of exogen therapeutic agents is almost necessary. Unfortunately, gentle co-incubation is inefficient, and the alternative techniques partly damage the vesicles. Therefore, it is essential to ensure consistency of exosomes and drugs structure and function.
We especially focused on the recent developments of exosomes-mimes, because they open new lines considering the important issues that are required to be addressed before exosomes can be approved for clinical use. Although exosome-based therapeutics have succeeded in numerous studies, several hurdles also remain. We will announce them hereafter, to put into perspective the path that remains to be taken before their clinical application
Enhancing EV secretion is one of the challenges, and several approaches have been developed, such as culturing under turbulent fluids, under acidic conditions, under salty stress, etc. However, it is important to note that exosome composition is dependent on the environmental conditions. This questions the conservation of the attempted function under these near-to-lethal stresses.
Even if exosomes have intrinsic addressing capacity thanks to surface components and autologous recognition, important research energy was dedicated to further modifying their surface with supplementary targeting ability or extended circulating half-life. Such approaches are developed for EVs, but also for the exosome-mimicking particles. This time, it must be noticed that such approaches can potentially induce immune reactions, especially if these peptides can be systemically released. Consequently, enhancing the anchoring of such targeting peptides on exosomes is a real issue.
Another line of work would be to develop the embedding of exosomes and exosome-like structures in biocompatible hydrogels to control their release, extend their residence time at the applied site, and optimize cell targeting through proximity.
The synthetic exosomes that can mimic the characteristics of natural exosomes are under scope but have not been as extensively investigated as modified exosomes. Two promising approaches are yet to be developed: (i) the cell or exosome membrane coating of organic or metallic particles; and (ii) hybrid exosomes, taking the advantages from the exosomes and liposomes. These engineered exosomes can carry different therapeutic agents and present specific targeting fragments on their surface, providing possible pluripotent vectorization platforms. However, concerning the HEs, controlling the fusion of liposomes and cell debris, as well as preventing the fusion between liposomes or hybrid exosomes, remain important issues that need to be addressed.
Moreover, if hybrid structures are not easily cleared by the mononuclear phagocytic or reticuloendothelial systems, this allows the improvement of drug-delivery efficiency, bringing into question their long-term accumulation and linked toxicity. To our knowledge, no work has been published in this direction.
Finally, one major issue for the clinical validation of exosomes or exosome-like structures resides in the purity of these (loaded) carriers. The surface adsorption of medium proteins, modification agents, or drugs is almost kept under silence. Indeed, the simple PBS-washing or repeated extrusion sequences are often not efficient enough to break Van der Walls or ionic interactions. However, these can be disrupted during the in vivo circulation or along their storage. Therefore, the issues are two-fold: how can it be ensured that the desired drug quantity is actually delivered?; and can the immunogenicity or toxicity of the in vivo released ingredients and carriers be avoided? Again, even if these issues are often evoked in the field reviews, we found no published studies on these concerns.
Finally, less is known about lipid trafficking linked to exosomes. Applications connected with the recent publications on the lipid mediators implied in the inflammation regulation would benefit from more studies.
Started only ten years ago, the clinical use of exosome-like vehicles needs to address all these more technical and deep-research issues. However, considering their great potential, it is paramount to open the application field of such technologies. Indeed, most pre-clinical assays are focused on cancer therapy or gene edition. Nevertheless, exosomes are also promising candidates for vaccine development as it has been demonstrated in several vaccine studies that infected cells secrete exosomes containing multiple antigens from the pathogen.

Author Contributions

Conceptualization, V.P.; writing—original draft preparation, V.P. and V.O.-M.; writing—review and editing, N.P.; visualization, V.P. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by Inserm and University Toulouse III—Paul Sabatier.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

cNPs—cellular membrane coated nanoparticles; Cur—curcumin; DDS—drug-delivery system; Dox—doxorubicin; EGF—epidermal growth factor; ESCRT—endosomal sorting complex required for transport; EVs—extracellular vesicles; HNVs—hybrid nanovesicles; HE—hybrid exosomes; LNPs—lipid nanoparticles; MSC—mesenchymal stem cells; miRNA—microRNA; nanoEV—nanometric extracellular vesicles; Ptx—paclitaxel; siRNA—small interfering RNA.

References

  1. Johnstone, R.M.; Adam, M.; Hammond, J.; 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] [PubMed]
  2. Colombo, M.; Raposo, G.; Aroc, C.J.; Théry, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell. Dev. Bio. 2014, 30, 255–289. [Google Scholar] [CrossRef] [PubMed]
  3. Ristorcelli, E.; Beraud, E.; Mathieu, S.; Lombardo, D.; Verine, A. Essential role of Notch signaling in apoptosis of human pancreatic tumoral cells mediated by exosomal nanoparticles. Int. J. Cancer 2009, 125, 1016–1026. [Google Scholar] [CrossRef] [PubMed]
  4. Raposo, G.; Nijman, H.W.; Stoorvogel, W.; Liejendekker, R.; Harding, C.V.; Melief, C.; Geuze, H.J. B lymphocytes secrete antigen-presenting vesicles. J. Exp. Med. 1996, 183, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  5. Farooqi, A.A.; Desai, N.N.; Qureshi, M.Z.; Librelotto, D.R.N.; Gasparri, M.L.; Bishayee, A.; Nabavi, S.M.; Curti, V.; Daglia, M. Exosome biogenesis, bioactivities and functions as new delivery systems of natural compounds. Biotechnol. Adv. 2018, 36, 328–334. [Google Scholar] [CrossRef] [PubMed]
  6. Maeki, M.; Kimura, N.; Sato, Y.; Harashima, H.; Tokeshi, M. Advances in microfluidics for lipid nanoparticles and extracellular vesicles and applications in drug delivery systems. Adv. Drug Deliv. Rev. 2018, 128, 84–100. [Google Scholar] [CrossRef] [PubMed]
  7. 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] [PubMed]
  8. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, J.O. Exosome mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed]
  9. Kahlert, C.; Melo, S.A.; Protopopov, A.; Tang, J.; Seth, S.; Koch, M.; Zhang, J.; Weitz, J.; Chin, L.; Futreal, A.; et al. Identification of double-stranded genomic DNA spanning all chromosomes with mutated KRAS and p53 DNA in the serum exosomes of patients with pancreatic cancer. J. Biol. Chem. 2014, 289, 3869–3875. [Google Scholar] [CrossRef]
  10. de Jong, O.G.; Verhaar, M.C.; Chen, Y.; Vader, P.; Gremmels, H.; Posthuma, G.; Schiffelers, R.M.; Gucek, M.; van Balkom, B.W. Cellular stress conditions are reflected in the protein and RNA content of endothelial cell-derived exosomes. J. Extracell. Vesicles 2012, 1, 18396. [Google Scholar] [CrossRef]
  11. Willms, E.; Cabanas, C.; Mager, I.; Wood, M.J.A.; Vader, P. Extracellular vesicle heterogeneity: Subpopulations, isolation techniques, and diverse functions in cancer progression. Front. Immunol. 2018, 9, 738. [Google Scholar] [CrossRef] [PubMed]
  12. Greening, D.W.; Simpson, R.J. Understanding extracellular vesicle diversity—Current status. Expert Rev. Proteom. 2018, 15, 887–910. [Google Scholar] [CrossRef] [PubMed]
  13. O’Brien, K.; Breyne, K.; Ughetto, S.; Laurent, L.C.; Breakefield, X.O. RNA delivery by extracellular vesicles in mammalian cells and its applications. Nat. Rev. Mol. Cell Biol. 2020, 21, 585–606. [Google Scholar] [CrossRef] [PubMed]
  14. Wiklander, O.P.B.; Brennan, M.; Lötvall, J.; Breakefield, X.O.; El Andaloussi, S. Advances in therapeutic applications of extracellular vesicles. Sci. Transl. Med. 2019, 11, 8521. [Google Scholar] [CrossRef] [PubMed]
  15. Robbins, P.D.; Morelli, A.E. Regulation of immune responses by extracellular vesicles. Nat. Rev. Immunol. 2014, 14, 195–208. [Google Scholar] [CrossRef]
  16. Zhou, H.; Pisitkun, T.; Aponte, A.; Yuen, P.S.; Hoffert, J.D.; Yasuda, H.; Hu, X.; Chawla, L.; Shen, R.F.; Knepper, M.A.; et al. Exosomal Fetuin-A identified by proteomics: A novel urinary biomarker for detecting acute kidney injury. Kidney Int. 2006, 70, 1847–1857. [Google Scholar] [CrossRef]
  17. Miranda, K.C.; Bond, D.T.; McKee, M.; Skog, J.; Paunescu, T.G.; Da Silva, N.; Brown, D.; Russo, L.M. Nucleic acids within urinary exosomes/microvesicles are potential biomarkers for renal disease. Kidney Int. 2010, 78, 191–199. [Google Scholar] [CrossRef]
  18. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D. Minimal information for studies of extracellular vesicles, (MISEV 2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV 2014 guidelines. J. Extracell. Vesicles 2017, 7, 1535750. [Google Scholar] [CrossRef] [PubMed]
  19. 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]
  20. Record, M.; Silvente-Poirot, S.; Poirot, M.; Wakelam, M.O. Extracellular vesicles: Lipids as key components of their biogenesis and functions. J. Lipid Res. 2018, 59, 1316–1324. [Google Scholar] [CrossRef]
  21. Llorente, A.; Skotland, T.; Sylvänne, T.; Kauhanen, D.; Rog, T.; Orłowski, A.; Vattulainen, I.; Ekroos, K.; Sandvig, K.; Bo, C. Molecular lipidomics of exosomes released by PC-3 prostate cancer cells. Lipids 2013, 1831, 1302–1309. [Google Scholar] [CrossRef] [PubMed]
  22. Keerthikumar, S.; Chisanga, D.; Ariyaratne, D.; Al Saffar, H.; Anand, S.; Zhao, K.; Samuel, M.; Pathan, M.; Jois, M.; Chilamkurti, N.; et al. ExoCarta: A Web-Based Compendium of Exosomal Cargo. J. Mol. Biol. 2016, 428, 688–692. [Google Scholar] [CrossRef] [PubMed]
  23. Ristorcelli, E.; Beraud, E.; Verrando, P.; Villard, C.; Lafitte, D.; Sbarra, V.; Lombardo, D.; Verine, A. Human tumor nanoparticles induce apoptosis of pancreat cancer cells. FASEB J. 2008, 22, 3358–3369. [Google Scholar] [CrossRef] [PubMed]
  24. Garcia, N.; Gonzàlez-King, H.; Grueso, E.; Sànchez, R.; Martinez-Romero, A.; Jàvega, B.; O’connor, J.; Simons, P.; Handberg, A.; Sepúlveda, P. Circulating exosomes deliver free fatty acids from the bloodstream to cardiac cells: Possible role of CD36. PLoS ONE 2019, 14, e0217546. [Google Scholar] [CrossRef] [PubMed]
  25. Ghadami, S.; Dellinger, K. The lipid composition of extracellular vesicles: Applications in diagnostics and therapeutic delivery. Front. Mol. Biosci. 2023, 10, 1198044. [Google Scholar] [CrossRef] [PubMed]
  26. Czubowicz, K.; Jésko, H.; Wencel, P.; Lukiw, W.J.; Strosznajder, R.P. The role of ceramide and sphingosine-1-phosphate in Alzheimer’s disease and other neurodegenerative disorders. Mol. Neurobiol. 2019, 56, 5436–5455. [Google Scholar] [CrossRef] [PubMed]
  27. Yuyama, K.; Sun, H.; Sakai, S.; Mitsutake, S.; Okada, M.; Tahara, H.; Furukawa, J.-I.; Fujitani, N.; Shinohara, Y.; Igarashi, Y. Decreased amyloid-β pathologies by intracerebral loading of glycosphingolipid-enriched exosomes in Alzheimer model mice. J. Biol. Chem. 2014, 289, 24488–24498. [Google Scholar] [CrossRef] [PubMed]
  28. Hanson, P.I.; Cashikar, A. Multivesicular body morphogenesis. Annu. Rev. Cell Dev. Biol. 2012, 28, 337–362. [Google Scholar] [CrossRef] [PubMed]
  29. Aguirre, R.S.; Kulkarni, A.; Becker, M.W.; Lei, X.; Sarkar, S.; Ramanadham, S.; Phelps, E.A.; Nakayasu, E.S.; Sims, E.K.; Mirmira, R.G. Extracellular vesicles in β cell biology: Role of lipids in vesicle biogenesis, cargo, and intercellular signaling. Mol. Metab. 2022, 63, 101545. [Google Scholar] [CrossRef]
  30. Li, S.-P.; Lin, Z.-X.; Jiang, X.-Y.; Yu, X.-Y. Exosomal cargo-loading and synthetic exosome-mimics as potential therapeutic tools. Acta Pharm. Sin. 2018, 39, 542–551. [Google Scholar] [CrossRef]
  31. Egea-Jimenez, A.L.; Zimmermann, P. Phospholipase D and phosphatidic acid in the biogenesis and cargo loading of extracellular vesicles: Thematic review series: Exosomes and microvesicles: Lipids as key components of their biogenesis and functions. J. Lipid Res. 2018, 59, 1554–1560. [Google Scholar] [CrossRef] [PubMed]
  32. Alonso, R.; Mazzeo, C.; Rodríguez, M.C.; Marsh, M.; Fraile-Ramos, A.; Calvo, V.; Avila-Flores, A.; Merida, I.; Izquierdo, M. Diacylglycerol kinase α regulates the formation and polarisation of mature multivesicular bodies involved in the secretion of Fas ligand-containing exosomes in T lymphocytes. Cell Death Differ. 2011, 18, 1161–1173. [Google Scholar] [CrossRef] [PubMed]
  33. Ghossoub, R.; Lembo, F.; Rubio, A.; Gaillard, C.B.; Bouchet, J.; Vitale, N.; Slavík, J.; Machala, M.; Zimmermann, P.J. Syntenin-ALIX exosome biogenesis and budding into multivesicular bodies are controlled by ARF6 and PLD2. Nat. Commun. 2014, 5, 3477. [Google Scholar] [CrossRef] [PubMed]
  34. Chu, Z.; Witte, D.P.; Qi, X. Saposin C–LBPA interaction in late-endosomes/lysosomes. Exp. Cell Res. 2005, 303, 300–307. [Google Scholar] [CrossRef] [PubMed]
  35. Miranda, A.M.; Lasiecka, Z.M.; Xu, Y.; Neufeld, J.; Shahriar, S.; Simoes, S.; Chan, R.B.; Oliveira, T.G.; Small, S.A.; Di Paolo, G. Neuronal lysosomal dysfunction releases exosomes harboring APP C-terminal fragments and unique lipid signatures. Nat. Commun. 2018, 9, 16. [Google Scholar] [CrossRef] [PubMed]
  36. Young, L.N.; Cho, K.; Lawrence, R.; Zoncu, R.; Hurley, J.H. Dynamics and Architecture of NRBF2-Containing phosphatidylinositol 3-kinase Complex I of autophagy. Proc. Natl. Acad. Sci. USA 2016, 113, 8224–8229. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, W.; Zhu, N.; Yan, T.; Shi, Y.-N.; Chen, J.; Zhang, C.-J.; Xie, X.-J.; Liao, D.-F.; Qin, L. The crosstalk: Exosomes and lipid metabolism. Cell Commun. Signal 2020, 18, 119. [Google Scholar] [CrossRef] [PubMed]
  38. Hannafon, B.N.; Ding, W.-Q. Intercellular Communication By Exosome-Derived MicroRNAs In Cancer. Int. J. Mol. Sci. 2013, 14, 14240–14269. [Google Scholar] [CrossRef] [PubMed]
  39. Sandvig, K.; Llorente, A. Proteomic analysis of microvesicles released by the human prostate cancer cell line PC-3, M111.012914. Mol. Cell Proteom. 2012, 11, M111.012914–1–M111.012914-11. [Google Scholar] [CrossRef]
  40. Choi, D.S.; Kim, D.K.; Kim, Y.K.; Gho, Y.S. Proteomics, transcriptomics and lipidomics of exosomes and ectosomes. Proteomics 2013, 13, 1554–1571. [Google Scholar] [CrossRef]
  41. Laulagnier, K.; Motta, C.; Hamdi, S.; Roy, S.; Fauvelle, F.; Pageaux, J.-F.; Kobayashi, T.; Salles, J.-P.; Perret, B.; Bonnerot, C.; et al. Mast cell- and dendritic cell-derived exosomes display a specific lipid composition and an unusual membrane organization. Biochem. J. 2004, 380, 161–171. [Google Scholar] [CrossRef] [PubMed]
  42. Nishida-Aoki, N.; Izumi, Y.; Takeda, H.; Takahashi, M.; Ochiya, T.; Bamba, T. Lipidomic Analysis of Cells and Extracellular Vesicles from High- and Low-Metastatic Triple-Negative Breast Cancer. Metabolites 2020, 10, 67. [Google Scholar] [CrossRef] [PubMed]
  43. Haraszti, R.A.; Didiot, M.-C.; Sapp, E.; Leszyk, J.; Shaffer, S.A.; Rockwell, H.E.; Gao, F.; Narain, N.R.; DiFiglia, M.; Kiebish, M.A.; et al. High-resolution proteomic and lipidomic analysis of exosomes and microvesicles from different cell sources. J. Extra Vesicles 2016, 5, 32570. [Google Scholar] [CrossRef] [PubMed]
  44. Pizzinat, N.; Ong-Meang, V.; Bourgailh-Tortosa, F.; Blanzat, M.; Perquis, L.; Cussac, D.; Parini, A.; Poinsot, V. Extracellular vesicles of MSCs and cardiomyoblasts are vehicles for lipid mediators. Biochimie 2020, 178, 69–80. [Google Scholar] [CrossRef] [PubMed]
  45. Subra, C.; Grand, D.; Laulagnier, K.; Stella, A.; Lambeau, G.; Paillasse, M.; De Medina, P.; Monsarrat, B.; Perret, B.; Silvente-Poirot, B.; et al. Exosomes account for vesicle-mediated transcellular transport of activatable phospholipases and prostaglandins. J. Lip. Res. 2010, 51, 2105–2120. [Google Scholar] [CrossRef] [PubMed]
  46. Esser, J.; Gehrmann, U.; D’Alexandri, F.L.; Hidalgo-Estévez, A.M.; Wheelock, C.E.; Scheynius, A.; Gabrielsson, S.; Rådmark, O.J. Exosomes from human macrophages and dendritic cells contain enzymes for leukotriene biosynthesis and promote granulocyte migration. J. Allergy Clin. Immunol. 2010, 126, 1032–1040. [Google Scholar] [CrossRef]
  47. Ong-Meang, V.; Blanzat, M.; Savchenko, L.; Perquis, L.; Guardia, M.; Pizzinat, N.; Poinsot, V. Extracellular Vesicles Produced by the Cardiac Microenvironment Carry Functional Enzymes to Produce Lipid Mediators In Situ. Int. J. Mol. Sci. 2023, 24, 5866. [Google Scholar] [CrossRef] [PubMed]
  48. Khoushab, S.; Aghmiuni, M.H.; Esfandiari, N.; Sarvandani, M.R.R.; Rashidi, M.; Taheriazam, A.; Entezari, M.; Hashemi, M. Unlocking the Potential of Exosomes in Cancer Research: A Paradigm Shift in Diagnosis, Treatment, and Prevention. Pathol. Res. Pract. 2024, 255, 155214. [Google Scholar] [CrossRef] [PubMed]
  49. Jung, I.; Shin, S.; Baek, M.C.; Yea, K. Modification of immune cell-derived exosomes for enhanced cancer immunotherapy: Current advances and therapeutic applications. Exp. Mol. Med. 2024, 56, 19–31. [Google Scholar] [CrossRef]
  50. Huda, M.N.; Nurunnabi, M. Potential Application of Exosomes in Vaccine Development and Delivery. Pharm. Res. 2022, 39, 2635–2671. [Google Scholar] [CrossRef]
  51. Cheng, L.; Jin, C.; Lv, W.; Ding, Q.; Han, X. Developing a highly stable PLGA-mPEG nanoparticle loaded with cisplatin for chemotherapy of ovarian cancer. PLoS ONE 2011, 6, e25433. [Google Scholar] [CrossRef] [PubMed]
  52. Lajunen, T.; Viitala, L.; Kontturi, L.-S.; Laaksonen, T.; Liang, H.; Vuorimaa-Laukkanen, E.; Viitala, T.; Le Guével, X.; Yliperttula, M.; Murtomäki, L.; et al. Light induced cytosolic drug delivery from liposomes with gold nanoparticles. J. Control. Release 2015, 203, 85–98. [Google Scholar] [CrossRef] [PubMed]
  53. Tanaka, H.; Akita, H.; Ishida, R.; Tange, K.; Arai, M.; Kubo, K.; Harashima, H. Neutral biodegradable lipid-envelope-type nanoparticle using vitamin A-scaffold for nuclear targeting of plasmid DNA. Biomaterials 2014, 35, 1755–1761. [Google Scholar] [CrossRef] [PubMed]
  54. Thi, E.P.; Mire, C.E.; Lee, A.C.H.; Geisbert, J.B.; Zhou, J.Z.; Agans, K.N.; Snead, N.M.; Deer, D.J.; Barnard, T.R.; Fenton, K.A.; et al. Lipid nanoparticle siRNA treatment of Ebola-virus-Makona-infected nonhuman primates. Nature 2015, 521, 362–365. [Google Scholar] [CrossRef] [PubMed]
  55. Suk, J.S.; Xu, Q.; Kim, N.; Hanes, J.; Ensign, L.M. PEGylation as a strategy for improving nanoparticle-based drug and gene delivery. Adv. Drug Deliv. Rev. 2016, 99, 28–51. [Google Scholar] [CrossRef] [PubMed]
  56. Cheng, X.; Lee, R.J. The role of helper lipids in lipid nanoparticles (LNPs) designed for oligonucleotide delivery. Adv. Drug Deliv. Rev. 2016, 99, 129–137. [Google Scholar] [CrossRef] [PubMed]
  57. Allen, T.M.; Cullis, P.R. Liposomal drug delivery systems: From concept to clinical applications. Adv. Drug Deliv. Rev. 2013, 65, 36–48. [Google Scholar] [CrossRef] [PubMed]
  58. Dawidczyk, C.M.; Kim, C.; Park, J.H.; Russell, L.M.; Lee, K.H.; Pomper, M.G.; Searson, P.C. State-of-the-art in design rules for drug delivery platforms: Lessons learned from FDA-approved nanomedicines. J. Control. Release 2014, 187, 133–144. [Google Scholar] [CrossRef]
  59. Sercombe, L.; Veerati, T.; Moheimani, F.; Wu, S.Y.; Sood, A.K.; Hua, S. Advances and challenges of liposome assisted drug delivery. Front. Pharmacol. 2015, 6, 286. [Google Scholar] [CrossRef]
  60. Xu, X.; Ho, W.; Zhang, X.; Bertrand, N.; Farokhzad, O. Cancer nanomedicine: From targeted delivery to combination therapy. Trends Mol. Med. 2015, 21, 223–232. [Google Scholar] [CrossRef]
  61. Noble, G.T.; Stefanick, J.F.; Ashley, J.D.; Kiziltepe, T.; Bilgicer, B. Ligand-targeted liposome design: Challenges and fundamental considerations. Trends Biotechnol. 2014, 32, 32–45. [Google Scholar] [CrossRef]
  62. Yingchoncharoen, P.; Kalinowski, D.S.; Richardson, D.R. Lipid-based drug delivery systems in cancer therapy: What is available and what is yet to come. Pharmacol. Rev. 2016, 68, 701–787. [Google Scholar] [CrossRef] [PubMed]
  63. Bastos, N.; Ruivo, C.F.; da Silva, S.; Melo, S.A. Exosomes in cancer: Use them or target them? Semin. Cell Dev. Biol. 2017, 78, 13–21. [Google Scholar] [CrossRef]
  64. Moon, B.; Chang, S. Exosome as a Delivery Vehicle for Cancer Therapy. Cells 2022, 11, 316. [Google Scholar] [CrossRef]
  65. Mizrak, A.; Bolukbasi, M.F.; Ozdener, G.B.; Brenner, G.J.; Madlener, S.; Erkan, E.P.; Strobel, T.; Breakefield, X.O.; Saydam, O. Genetically engineered microvesicles carrying suicide mRNA/protein inhibit schwannoma tumor growth. Mol. Ther. 2013, 21, 101–108. [Google Scholar] [CrossRef] [PubMed]
  66. Yin, W.; Ouyang, S.; Li, Y.; Xiao, B.; Yang, H. Immature dendritic cell-derived exosomes: A promise subcellular vaccine for autoimmunity. Inflammation 2013, 36, 232–240. [Google Scholar] [CrossRef] [PubMed]
  67. El Andaloussi, S.; Mäger, I.; Breakefield, X.O.; Wood, M.J.A. Extracellular vesicles: Biology and emerging therapeutic opportunities. Nat. Rev. Drug Discov. 2013, 12, 347. [Google Scholar] [CrossRef]
  68. Tian, Y.H.; Li, S.P.; Song, J.; Ji, T.J.; Zhu, M.T.; Anderson, G.J.; Wei, J.Y.; Nie, G.J. A doxorubicin delivery platform using engineered natural membrane vesicle exosomes for targeted tumor therapy. Biomaterials 2014, 35, 2383–2390. [Google Scholar] [CrossRef]
  69. Yeo, R.W.; Lai, R.C.; Zhang, B.; Tan, S.S.; Yin, Y.; The, B.J.; Lim, S.K. Mesenchymal stem cell: An efficient mass producer of exosomes for drug delivery. Adv. Drug Deliv. Rev. 2013, 65, 336–341. [Google Scholar] [CrossRef]
  70. Rezaie, J.; Feghhi, M.; Etemadi, T.A. review on exosomes application in clinical trials: Perspective, questions, and challenges. Cell Commun. Signal. 2022, 20, 145. [Google Scholar] [CrossRef]
  71. Mohammadi, M.; Zargartalebi, H.; Salahandish, R.; Aburashed, R.; Yong, K.W.; Sanati-Nezhad, A. Emerging technologies and commercial products in exosome-based cancer diagnosis and prognosis. Biosens. Bioelectron. 2021, 183, 113176. [Google Scholar] [CrossRef] [PubMed]
  72. Urbanelli, L.; Buratta, S.; Sagini, K.; Ferrara, G.; Lanni, M.; Emiliani, C. Exosome-based strategies for Diagnosis and Therapy. Recent Patents CNS Drug Discov. 2015, 10, 10–27. [Google Scholar] [CrossRef] [PubMed]
  73. Makler, A.; Asghar, W. Exosomal biomarkers for cancer diagnosis and patient monitoring. Expert Rev. Mol. Diagn. 2020, 20, 387–400. [Google Scholar] [CrossRef] [PubMed]
  74. Yousefpour, P.; Chilkoti, A. Co-opting biology to deliver drugs. Biotechnol. Bioeng. 2014, 111, 1699–1716. [Google Scholar] [CrossRef] [PubMed]
  75. Sutaria, D.; Badawi, M.; Phelps, M.; Schmittgen, T. Achieving the promise of therapeutic extracellular vesicles: The devil is in details of therapeutic loading. Pharm. Res. 2017, 34, 1053–1066. [Google Scholar] [CrossRef] [PubMed]
  76. Donoso-Quezada, J.; Ayala-Mar, S.; González-Valdez, J. State-of-the-art exosome loading and functionalization techniques for enhanced therapeutics: A review. Crit. Rev. Biotechnol. 2020, 40, 804–820. [Google Scholar] [CrossRef] [PubMed]
  77. Pegtel, D.M.; Cosmopoulos, K.; Thorley-Lawson, D.A.; Eijndhoven, M.A.J.V.; Hopmans, E.S.; Lindenberg, J.L.; Gruijl, T.D.D.; Wuerdinger, T.; Middeldorp, J.M. Functional delivery of viral miRNA by exosomes. Proc. Natl. Acad. Sci. USA 2010, 107, 6328–6333. [Google Scholar] [CrossRef] [PubMed]
  78. Morelli, E.A.; Larregina, A.T.; Shufesky, W.J.; Sullivan, M.L.G.; Beer Stolz, D.; Papworth, G.D.; Zahorchak, A.F.; Logar, A.J.; Wang, Z.; Watkins, S.C.; et al. Endocytosis, intracellular sorting, and processing of exosomes by dendritic cells. Blood 2004, 104, 3257–3266. [Google Scholar] [CrossRef] [PubMed]
  79. Peinado, H.; Ale Kovi, M.A.; Lavotshkin, S.; Matei, I.; Costa-Silva, B.; Moreno-Bueno, G.; Hergueta-Redondo, M.; Williams, C.; Garcıà-Santos, G.; Ghajar, C.M. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat. Med. 2012, 18, 883–891. [Google Scholar] [CrossRef]
  80. Yukawa, H.; Suzuki, K.; Aoki, K.; Arimoto, T.; Yasui, T.; Kaji, N.; Ishikawa, T.; Ochiya, T.; Baba, Y. Imaging of angiogenesis of human umbilical vein endothelial cells by uptake of exosomes secreted from hepatocellular carcinoma cells. Sci. Rep. 2018, 8, 6765. [Google Scholar] [CrossRef]
  81. Suetsugu, A.; Honma, K.; Saji, S.; Moriwaki, H.; Ochiya, T.; Hoffman, R.M. Imaging exosome transfer from breast cancer cells to stroma at metastatic sites in orthotopic nude-mouse models. Adv. Drug Deliv. Rev. 2013, 65, 383–390. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, C.; Zong, S.; Wang, Z.; Lu, J.; Zhu, D.; Zhang, Y.; Cui, Y. Imaging and Intracellular Tracking of Cancer-Derived Exosomes Using Single-Molecule Localization-Based Super-Resolution Microscope. ACS Appl. Mater. Interfaces 2016, 8, 25825–25833. [Google Scholar] [CrossRef] [PubMed]
  83. Andaloussi, S.E.L.; 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] [PubMed]
  84. Wahlgren, J.; Karlson, T.D.L.; Brisslert, M.; Vaziri Sani, F.; Telemo, E.; Sunnerhagen, P.; Valadi, H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012, 40, e130. [Google Scholar] [CrossRef] [PubMed]
  85. He, J.; Banizs, A.B.; Huang, T.; Dryden, K.; Berr, S.S.; Stone, J.R.; Nakamoto, R.K.; Shi, W. In vitro evaluation of endothelial exosomes as carriers for small interfering ribonucleic acid delivery. Int. J. Nanomed. 2014, 9, 4223–4230. [Google Scholar] [CrossRef] [PubMed]
  86. Shtam, T.A.; Kovalev, R.A.; Varfolomeeva, E.Y.; Makarov, E.M.; Kil, Y.V.; Filatov, M.V. Exosomes are natural carriers of exogenous siRNA to human cells in vitro. Cell Commun. Signal. 2013, 11, 88. [Google Scholar] [CrossRef] [PubMed]
  87. Pei, X.; Zhang, X.; Zhang, L.; Yuan, M.; Sun, L.; Yu, F.; Wang, B.; Zhao, J.; He, H.; Yang, V.C. Targeted exosomes for co-delivery of siFGL1 and siTGF-β1 trigger combined cancer immunotherapy by remodeling immunosuppressive tumor microenvironment. Chem. Eng. J. 2021, 421, 129774. [Google Scholar] [CrossRef]
  88. 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]
  89. Huang, W.; Qu, M.; Li, L.; Liu, T.; Lin, M.; Yu, X. SiRNA in MSC derived exosomes silences CTGF gene for locomotor recovery in spinal cord injury rats. Stem Cell Res. Ther. 2021, 12, 334. [Google Scholar] [CrossRef]
  90. Rong, Y.; Wang, Z.; Tang, P.; Wang, J.; Ji, C.; Chang, J.; Zhu, Y.; Ye, W.; Bai, J.; Liu, W.; et al. Engineered extracellular vesicles for delivery of siRNA promoting targeted repair of traumatic spinal cord injury. Bioact. Mater. 2023, 23, 328–342. [Google Scholar] [CrossRef]
  91. Haraszt, R.A.; Miller, R.; Didiot, M.; Biscans, A.; Alterman, J.F.; Hassler, M.R.; Roux, L.; Echeverria, D.; Sapp, E.; DiFiglia, M.; et al. Optimized Cholesterol-siRNA Chemistry Improves Productive Loading onto Extracellular Vesicles. Mol. Ther. 2018, 26, 1973–1982. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, D.; Cai, G.; Liu, K.; Zhuang, Z.; Jia, K.; Pei, S.; Wang, X.; Wang, H.; Xu, S.; Cui, C.; et al. Microglia exosomal miRNA-137 attenuates ischemic brain injury through targeting Notch1. Aging 2021, 13, 4079–4095. [Google Scholar] [CrossRef]
  93. 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] [PubMed]
  94. Zhao, Z.; Shuang, T.; Gao, Y.; Lu, F.; Zhang, J.; He, W.; Qu, L.; Chen, B.; Hao, Q. Targeted delivery of exosomal miR-484 reprograms tumor vasculature for chemotherapy sensitization. Cancer Lett. 2022, 530, 45–58. [Google Scholar] [CrossRef]
  95. Hade, M.D.; Suire, C.N.; Suo, Z. An Effective Peptide-Based Platform for Efficient Exosomal Loading and Cellular Delivery of a microRNA. ACS Appl. Mater. Interfaces 2023, 15, 3851–3866. [Google Scholar] [CrossRef] [PubMed]
  96. Gong, C.; Tian, J.; Wang, Z.; Gao, Y.; Wu, X.; Ding, X.; Qiang, L.; Li, G.; Han, Z.; Yuan, Y.; et al. Functional exosome-mediated codelivery of doxorubicin and hydrophobically modified microRNA 159 for triple-negative breast cancer therapy. J. Nanobiotechnol. 2019, 17, 93. [Google Scholar] [CrossRef] [PubMed]
  97. Liang, G.; Zhu, Y.; Ali, D.J.; Tian, T.; Xu, H.; Si, K.; Sun, B.; Chen, B.; Xiao, Z. Engineered exosomes for targeted co-delivery of miR-21 inhibitor and chemotherapeutics to reverse drug resistance in colon cancer. J. Nanobiotechnol. 2020, 18, 10. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, G.; Kim, M.; Lee, Y.; Byun, J.W.; Hwang, D.W.; Lee, M. Systemic delivery of microRNA-21 antisense oligonucleotides to the brain using T7-peptide decorated exosomes. J. Control. Release 2020, 317, 273–281. [Google Scholar] [CrossRef] [PubMed]
  99. Haney, M.J.; Klyachko, N.L.; Harrison, E.B.; Zhao, Y.; Kabanov, A.V.; Batrakova, E.V. TPP1 Delivery to Lysosomes with Extracellular Vesicles and their Enhanced Brain Distribution in the Animal Model of Batten Disease. Adv. Health Mater. 2019, 8, e1801271. [Google Scholar] [CrossRef]
  100. Zhang, W.; Wang, L.; Guo, H.; Chen, L.; Huang, X. Dapagliflozin-Loaded Exosome Mimetics Facilitate Diabetic Wound Healing by HIF-1a-Mediated Enhancement of Angiogenesis. Adv. Healthc. Mater. 2022, 12, 2202751. [Google Scholar] [CrossRef]
  101. Fuhrmann, G.; Chandrawati, R.; Parmar, P.A.; Keane, T.J.; Maynard, S.A.; Bertazzo, S.; Stevens, M.M. Engineering Extracellular Vesicles with the Tools of Enzyme Prodrug Therapy. Adv. Mater. 2018, 30, 1706616. [Google Scholar] [CrossRef]
  102. Rodríguez-Morales, B.; Antunes-Ricardo, M.; González-Valdez, J. Exosome-Mediated Insulin Delivery for the Potential Treatment of Diabetes Mellitus. Pharmaceutics 2021, 13, 1870. [Google Scholar] [CrossRef] [PubMed]
  103. Lu, J.; Yang, X.; He, C.; Chen, Y.; Li, C.; Li, S.; Chen, Y.; Wu, Y.; Xiang, Z.; Kang, J.; et al. Rejuvenation of tendon stem/progenitor cells for functional tendon regeneration through platelet-derived exosomes loaded with recombinant Yap1. Acta Biomater. 2023, 161, 80–99. [Google Scholar] [CrossRef]
  104. Hajipour, H.; Farzadi, L.; Roshangar, L.; Latifi, Z.; Kahroba, H.; Shahnazi, V.; Hamdi, K.; Ghasemzadeh, A.; Fattahi, A.; Nouri, M. A human chorionic gonadotropin (hCG) delivery platform using engineered uterine exosomes to improve endometrial receptivity. Life Sci. 2021, 275, 119351. [Google Scholar] [CrossRef]
  105. Li, Y.; Gao, Y.; Gong, C.; Wang, Z.; Xia, Q.; Gu, F.; Hu, C.; Zhang, L.; Guo, H.; Gao, S. A33 antibody functionalized exosomes for targeted delivery of doxorubicin against colorectal cancer. Nanomedicine 2018, 14, 1973–1985. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, P.; Wang, H.; Huang, Q.; Peng, C.; Yao, L.; Chen, H.; Qiu, Z.; Wu, Y.; Wang, L.; Chen, W. Exosomes from M1-polarized macrophages enhance paclitaxel antitumor activity by activating macrophages-mediated inflammation. Theranostics 2019, 9, 1714–1727. [Google Scholar] [CrossRef]
  107. Zhu, Q.; Ling, X.; Yang, Y.; Zhang, J.; Li, Q.; Niu, X.; Hu, G.; Chen, B.; Li, H.; Wang, Y.; et al. Embryonic stem cells-derived exosomes endowed with targeting properties as chemotherapeutics delivery vehicles for glioblastoma therapy. Adv. Sci. 2019, 6, 1801899. [Google Scholar] [CrossRef] [PubMed]
  108. Al Faruque, H.; Choi, E.; Kim, J.; Kim, E. Enhanced effect of autologous EVs delivering paclitaxel in pancreatic cancer. J. Control. Release 2022, 347, 330–346. [Google Scholar] [CrossRef]
  109. Zhang, L.; Fan, C.; Hao, W.; Zhuang, Y.; Liu, X.; Zhao, Y.; Chen, B.; Xiao, Z.; Chen, Y.; Dai, J. NSCs Migration Promoted and Drug Delivered Exosomes-Collagen Scaffold via a Bio-Specific Peptide for One-Step Spinal Cord Injury Repair. Adv. Healthc. Mater. 2021, 10, 2001896. [Google Scholar] [CrossRef]
  110. Li, Y.-J.; Wu, J.-Y.; Wang, J.-M.; Hu, X.-B.; Cai, J.-X.; Xiang, D.-X. Gemcitabine loaded autologous exosomes for effective and safe chemotherapy of pancreatic cancer. Acta Biomater. 2020, 101, 519–530. [Google Scholar] [CrossRef]
  111. Zhou, W.; Zhou, Y.; Chen, X.; Ning, T.; Chen, H.; Guo, Q.; Zhang, Y.; Liu, P.; Zhang, Y.; Li, C.; et al. Pancreatic cancer-targeting exosomes for enhancing immunotherapy and reprogramming tumor microenvironment. Biomaterials 2021, 268, 120546. [Google Scholar] [CrossRef] [PubMed]
  112. Du, J.; Wan, Z.; Wang, C.; Lu, F.; Wei, M.; Wang, D.; Hao, Q. Designer exosomes for targeted and efficient ferroptosis induction in cancer via chemo-photodynamic therapy. Theranostics 2021, 11, 8185–8196. [Google Scholar] [CrossRef] [PubMed]
  113. Fang, J.; Liang, W. ASCs-derived exosomes loaded with vitamin A and quercetin inhibit rapid senescence-like response after acute liver injury. Biochem. Biophys. Res. Commun. 2021, 572, 125–130. [Google Scholar] [CrossRef] [PubMed]
  114. Timofeeva, A.M.; Paramonik, A.P.; Sedykh, S.S.; Nevinsky, G.A. Milk Exosomes: Next-Generation Agents for Delivery of Anticancer Drugs and Therapeutic Nucleic Acids. Int. J. Mol. Sci. 2023, 24, 10194. [Google Scholar] [CrossRef] [PubMed]
  115. Ngu, A.; Wang, S.; Wang, H.; Khanam, A.; Zempleni, J. Milk exosomes in nutrition and drug delivery. Am. J. Physiol. Physiol. 2022, 322, C865–C874. [Google Scholar] [CrossRef] [PubMed]
  116. Zhong, J.; Xia, B.; Shan, S.; Zheng, A.; Zhang, S.; Chen, J.; Liang, X.J. High-quality milk exosomes as oral drug delivery system. Biomaterials 2021, 277, 121126. [Google Scholar] [CrossRef] [PubMed]
  117. González-Sarrías, A.; Iglesias-Aguirre, C.E.; Cortés-Martín, A.; Vallejo, F.; Cattivelli, A.; del Pozo-Acebo, L.; Del Saz, A.; Hazas, M.C.L.d.L.; Dávalos, A.; Espín, J.C. Milk-derived exosomes as nanocarriers to deliver curcumin and resveratrol in breast tissue and enhance their anticancer activity. Int. J. Mol. Sci. 2022, 23, 2860. [Google Scholar] [CrossRef] [PubMed]
  118. Jia, G.; Han, Y.; An, Y.; Ding, Y.; He, C.; Wang, X.; Tang, Q. 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] [PubMed]
  119. Carobolante, G.; Mantaj, J.; Ferrari, E.; Vllasaliu, D. Cow milk and intestinal epithelial cell-derived extracellular vesicles as systems for enhancing oral drug delivery. Pharmaceutics 2020, 12, 226. [Google Scholar] [CrossRef]
  120. Yang, X.; Shi, G.; Guo, J.; Wang, C.; He, Y. Exosome-encapsulated antibiotic against intracellular infections of methicillin-resistant Staphylococcus aureus. Int. J. Nanomed. 2018, 13, 8095–8104. [Google Scholar] [CrossRef]
  121. Charoenviriyakul, C.; Takahashi, Y.; Morishita, M.; Matsumoto, A.; Nishikawa, M.; Takakura, Y. Cell type-specific and common characteristics of exosomes derived from mouse cell lines: Yield, physico-chemical properties, and pharmacokinetics. Eur. J. Pharm. Sci. 2017, 96, 316–322. [Google Scholar] [CrossRef] [PubMed]
  122. Willis, G.R.; Kourembanas, S.; Mitsialis, S.A. Toward exosome-based therapeutics: Isolation, heterogeneity, and fit-for-purpose potency. Front. Cardiovasc. Med. 2017, 4, 63. [Google Scholar] [CrossRef] [PubMed]
  123. Riches, A.; Campbell, E.; Borger, E.; Powis, S. Regulation of exosome release from mammary epithelial and breast cancer cells—A new regulatory pathway. Eur. J. Cancer 2014, 50, 1025–1034. [Google Scholar] [CrossRef] [PubMed]
  124. 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] [PubMed]
  125. Jeske, R.; Liu, C.; Duke, L.; Castro, M.L.C.; Muok, L.; Arthur, P.; Singh, M.; Jung, S.; Sun, L.; Li, Y. Upscaling human mesenchymal stromal cell production in a novel vertical-wheel bioreactor enhances extracellular vesicle secretion and cargo profile. Bioact. Mat. 2023, 25, 732–747. [Google Scholar] [CrossRef] [PubMed]
  126. 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]
  127. Harmati, M.; Tarnai, Z.; Decsi, G.; Kormondi, S.; Szegletes, Z.; Janovak, L.; Dekany, I.; Saydam, O.; Gyukity-Sebestyen, E.; Dobra, G.; et al. Stressors alter intercellular communication and exosome profile of nasopharyngeal carcinoma cells. J. Oral Pathol. Med. 2017, 46, 259–266. [Google Scholar] [CrossRef] [PubMed]
  128. Heinemann, M.L.; Ilmer, M.; Silva, L.P.; Hawke, D.H.; Recio, A.; Vorontsova, M.A.; Alt, E.; Vykoukal, J. Benchtop isolation and characterization of functional exosomes by sequential filtration. J. Chromatogr. A 2014, 1371, 125–135. [Google Scholar] [CrossRef]
  129. Sidhom, K.; Obi, P.O.; Saleem, A.A. Review of Exosomal Isolation Methods: Is Size Exclusion Chromatography the Best Option? Int. J. Mol. Sci. 2020, 21, 6466. [Google Scholar] [CrossRef]
  130. Weng, Y.; Sui, Z.; Shan, Y.; Hu, Y.; Chen, Y.; Zhang, L.; Zhang, Y. Effective isolation of exosomes with polyethylene glycol from cell culture supernatant for in-depth proteome profiling. Analyst 2016, 141, 4640–4646. [Google Scholar] [CrossRef]
  131. Wang, J.; Ma, P.; Kim, D.H.; Liu, B.-F.; Demirci, U. Towards microfluidic-based exosome isolation and detection for tumor therapy. Nano Today 2021, 37, 101066. [Google Scholar] [CrossRef] [PubMed]
  132. Willms, E.; Johansson, H.J.; Mäger, I.; Lee, Y.; Blomberg, K.E.M.; Sadik, M.; Alaarg, A.; Smith, C.E.; Lehtiö, J.; EL Andaloussi, S.; et al. Cells release subpopulations of exosomes with distinct molecular and biological properties. Sci. Rep. 2016, 6, 22519. [Google Scholar] [CrossRef] [PubMed]
  133. Van Deun, J.; Mestdagh, P.; Sormunen, R.; Cocquyt, V.; Vermaelen, K.; Vandesompele, J.; Bracke, M.; De Wever, O.; Hendrix, A. The impact of disparate isolation methods for extracellular vesicles on downstream RNA profiling. J. Extracell. Vesicles 2014, 3, 24858. [Google Scholar] [CrossRef] [PubMed]
  134. Kowal, J.; Arras, G.; Colombo, M.; Jouve, M.; Morath, J.P.; Primdal Bengtson, B.; Dingli, F.; Loew, D.; Tkach, M.; Thery, C. Proteomic comparison defines novel markers to characterize heterogeneous populations of extracellular vesicle subtypes. Proc. Natl. Acad. Sci. USA 2016, 113, E968–E977. [Google Scholar] [CrossRef] [PubMed]
  135. Lesniak, A.; Fenaroli, F.; Monopoli, M.P.; Aberg, C.; Dawson, K.A.; Salvati, A. Effects of the presence or absence of a protein corona on silica nanoparticle uptake and impact on cells. ACS Nano 2012, 6, 5845–5857. [Google Scholar] [CrossRef] [PubMed]
  136. Digiacomo, L.; Cardarelli, F.; Pozzi, D.; Palchetti, S.; Digman, M.A.; Gratton, E.; Capriotti, A.L.; Mahmoudi, M.; Caracciolo, G. An apolipoprotein-enriched biomolecular corona switches the cellular uptake mechanism and trafficking pathway of lipid nanoparticles. Nanoscale 2017, 9, 17254–17262. [Google Scholar] [CrossRef] [PubMed]
  137. Jang, S.C.; Kim, O.Y.; Yoon, C.M.; Choi, D.-S.; Roh, T.-Y.; Park, J.; Nilsson, J.; Lötvall, J.; Kim, Y.-K.; Gho, Y.S. BCorrection to bioinspired exosome-mimetic nanovesicles for targeted delivery of chemotherapeutics to malignant tumors. ACS Nano 2013, 7, 7698–7710. [Google Scholar] [CrossRef] [PubMed]
  138. Molinaro, R.; Corbo, C.; Martinez, J.O.; Taraballi, F.; Evangelopoulos, M.; Minardi, S.; Yazdi, I.K.; Zhao, P.; De Rosa, E.; Sherman, M.B.; et al. Biomimetic proteolipid vesicles for targeting inflamed tissues. Nat. Mater. 2016, 15, 1037–1046. [Google Scholar] [CrossRef]
  139. Zhang, P.; Chen, Y.; Zeng, Y.; Shen, C.; Li, R.; Guo, Z.; Li, S.; Zheng, Q.; Chu, C.; Wang, Z.; et al. Virus-mimetic nanovesicles as a versatile antigen-delivery system. Proc. Natl. Acad. Sci. USA 2015, 112, 6129–6138. [Google Scholar] [CrossRef]
  140. Hu, C.-M.J.; Fang, R.H.; Wang, K.-C.; Luk, B.T.; Thamphiwatana, S.; Dehaini, D.; Nguyen, P.; Angsantikul, P.; Wen, C.H.; Kroll, A.V.; et al. Nanoparticle biointerfacing by platelet membrane cloaking. Nature 2015, 526, 118–121. [Google Scholar] [CrossRef]
  141. Rao, L.; Meng, Q.; Huang, Q.; Wang, Z.; Yu, G.; Li, A.; Ma, W.; Zhang, N.; Guo, S.; Zhao, X.; et al. Platelet–leukocyte hybrid membrane-coated immunomagnetic beads for highly efficient and highly specific isolation of circulating tumor cells. Adv. Funct. Mater. 2018, 28, 1803531. [Google Scholar] [CrossRef]
  142. Jing, L.; Qu, H.; Wu, D.; Zhu, C.; Yang, Y.; Jin, X.; Zheng, J.; Shi, X.; Yan, X.; Wang, Y. Platelet-camouflaged nanococktail: Simultaneous inhibition of drug-resistant tumor growth and metastasis via a cancer cells and tumor vasculature dual-targeting strategy. Theranostics 2018, 8, 2683–2695. [Google Scholar] [CrossRef] [PubMed]
  143. Choo, Y.W.; Kang, M.; Kim, H.Y.; Han, J.; Kang, S.; Lee, J.-R.; Jeong, G.-J.; Kwon, S.P.; Song, S.Y.; Go, S.; et al. M1 macrophage-derived nanovesicles potentiate the anticancer efficacy of immune checkpoint inhibitors. ACS Nano 2018, 12, 8977–8993. [Google Scholar] [CrossRef] [PubMed]
  144. Zhang, W.; Yu, M.; Xi, Z.; Nie, D.; Dai, Z.; Wang, J.; Qian, K.; Weng, H.; Gan, Y.; Xu, L. Cancer cell membrane-camouflaged nanorods with endoplasmic reticulum targeting for improved antitumor therapy. ACS Appl. Mater. Interfaces 2019, 11, 46614–46625. [Google Scholar] [CrossRef]
  145. Yanga, J.; Wanga, F.; Lua, Y.; Qia, J.; Denga, L.; Sousac, F.; Sarmentoc, B.; Xua, X.; Cuia, W. Recent advance of erythrocyte-mimicking nanovehicles: From bench to bedside. J. Control. Rel. 2019, 314, 81–91. [Google Scholar] [CrossRef]
  146. Aryal, S.; Hu, C.M.; Fang, R.H.; Dehaini, D.; Carpenter, C.; Zhang, D.E.; Zhang, L. Erythrocyte membrane-cloaked polymeric nanoparticles for controlled drug loading and release. Nanomedicine 2013, 8, 1271–1280. [Google Scholar] [CrossRef]
  147. Luk, B.T.; Fang, R.H.; Hu, C.M.; Copp, J.A.; Thamphiwatana, S.; Dehaini, D.; Gao, W.; Zhang, K.; Li, S.; Zhang, L. Safe and Immunocompatible nanocarriers cloaked in RBC membranes for drug delivery to treat solid tumors. Theranostics 2016, 6, 1004–1011. [Google Scholar] [CrossRef]
  148. Fu, Q.; Lv, P.; Chen, Z.; Ni, D.; Zhang, L.; Yue, H.; Yue, Z.; Wei, W.; Ma, G. Programmed co-delivery of paclitaxel and doxorubicin boosted by camouflaging with erythrocyte membrane. Nanoscale 2015, 7, 4020–4030. [Google Scholar] [CrossRef] [PubMed]
  149. Zhai, Z.; Xu, P.; Yao, J.; Li, R.; Gong, L.; Yin, Y.; Lin, Z. Erythrocyte-mimicking paclitaxel nanoparticles for improving biodistributions of hydrophobic drugs to enhance antitumor efficacy. Drug Deliv. 2020, 27, 387–399. [Google Scholar] [CrossRef]
  150. Krishnamurthy, S.; Gnanasammandhan, M.K.; Xie, C.; Huang, K.; Cui, M.Y.; Chan, J.M. Monocyte cell membrane-derived nanoghosts for targeted cancer therapy. Nanoscale 2016, 8, 6981–6985. [Google Scholar] [CrossRef]
  151. Kang, T.; Zhu, Q.; Wei, D.; Feng, J.; Yao, J.; Jiang, T.; Song, Q.; Wei, X.; Chen, H.; Gao, X.; et al. Nanoparticles Coated with Neutrophil Membranes Can Effectively Treat Cancer Metastasis. ACS Nano 2017, 11, 1397–1411. [Google Scholar] [CrossRef] [PubMed]
  152. Xie, X.; Wang, H.; Williams, G.R.; Yang, Y.; Zheng, Y.; Wu, J.; Zhu, L.M. Erythrocyte membrane cloaked curcumin-loaded nanoparticles for enhanced chemotherapy. Pharmaceutics 2019, 11, 429. [Google Scholar] [CrossRef] [PubMed]
  153. Wei, X.; Zhang, G.; Ran, D.; Krishnan, N.; Fang, R.H.; Gao, W.; Spector, S.A.; Zhang, L. T-Cell-Mimicking nanoparticles can neutralize HIV infectivity. Adv. Mater. 2018, 30, e1802233. [Google Scholar] [CrossRef] [PubMed]
  154. Nie, D.; Dai, Z.; Li, J.; Yang, Y.; Xi, Z.; Wang, J.; Zhang, W.; Qian, K.; Guo, S.; Zhu, C.; et al. Cancer-cell membrane- coated nanoparticles with a yolk–shell structure augment cancer chemotherapy. Nano Lett. 2019, 20, 936–946. [Google Scholar] [CrossRef] [PubMed]
  155. Fan, Y.; Cui, Y.; Hao, W.; Chen, M.; Liu, Q.; Wang, Y.; Yang, M.; Li, Z.; Gong, W.; Song, S.; et al. Carrier-free highly drug-loaded biomimetic nanosuspensions encapsulated by cancer cell membrane based on homology and active targeting for the treatment of glioma. Bioact. Mater. 2021, 6, 4402–4414. [Google Scholar] [CrossRef] [PubMed]
  156. He, W.; Mei, Q.; Li, J.; Zhai, Y.; Chen, Y.; Wang, R.; Lu, E.; Zhang, X.-Y.; Zhang, Z.; Sha, X. Preferential targeting cerebral ischemic lesions with cancer cell-inspired nanovehicle for ischemic stroke treatment. Nano Lett. 2021, 21, 3033–3043. [Google Scholar] [CrossRef] [PubMed]
  157. Zhou, D.; Chen, X.; Bu, W.; Meng, L.; Wang, C.; Jin, N.; Chen, Y.; Ren, C.; Zhang, K.; Sun, H. Modification of Metal-Organic Framework Nanoparticles Using Dental Pulp Mesenchymal Stem Cell Membranes to Target Oral Squamous Cell Carcinoma. J Colloid Interface Sci. 2021, 601, 650–660. [Google Scholar] [CrossRef] [PubMed]
  158. Pisano, S.; Pierini, I.; Gu, J.; Gazze, A.; Francis, L.W.; Gonzalez, D.; Conlan, R.S.; Corradetti, B. Immune (cell) derived exosome mimetics (IDEM) as a treatment for ovarian cancer. Front. Cell Dev. Biol. 2020, 8, 553576. [Google Scholar] [CrossRef] [PubMed]
  159. Chen, H.; Deng, J.; Yao, X.; He, Y.; Li, H.; Jian, Z.; Tang, Y.; Zhang, X.; Zhang, J.; Dai, H. Bone-targeted erythrocyte-cancer hybrid membrane-camouflaged nanoparticles for enhancing photothermal and hypoxia-activated chemotherapy of bone invasion by OSCC. J. Nanobiotechnol. 2021, 19, 342. [Google Scholar] [CrossRef]
  160. Rao, L.; Wu, L.; Liu, Z.; Tian, R.; Yu, G.; Zhou, Z.; Yang, K.; Xiong, H.-G.; Zhang, A.; Yu, G.-T.; et al. Hybrid cellular membrane nanovesicles amplify macrophage immune responses against cancer recurrence and metastasis. Nat. Commun. 2020, 11, 4909. [Google Scholar] [CrossRef]
  161. Rao, L.; Bu, L.; Ma, L.; Wang, W.; Liu, H.; Wan, D.; Liu, J.; Li, A.; Guo, S.; Zhang, L.; et al. Platelet -Facilitated Photothermal Therapy of Head and Neck Squamous Cell Carcinoma. Angew. Chem. Int. 2018, 130, 998–1003. [Google Scholar] [CrossRef]
  162. Bu, L.; Rao, L.; Yu, G.; Chen, L.; Deng, W.; Liu, J.; Wu, H.; Meng, Q.; Guo, S.; Zhao, X.; et al. Cancer Stem Cell-Platelet Hybrid Membrane-Coated Magnetic Nanoparticles for Enhanced Photothermal Therapy of Head and Neck Squamous Cell Carcinoma. Adv. Funct. Mater. 2019, 29, 1807733. [Google Scholar] [CrossRef]
  163. Rao, L.; He, Z.; Meng, Q.-F.; Zhou, Z.; Bu, L.-L.; Guo, S.-S.; Liu, W.; Zhao, X.Z. Effective cancer targeting and imaging using macrophage membrane-camouflaged upconversion nanoparticles. J. Biomed. Mater. Res. A 2017, 105, 521–530. [Google Scholar] [CrossRef] [PubMed]
  164. Zhao, Y.; Wang, J.; Cai, X.; Ding, P.; Lv, H.; Pei, R. Metal–Organic Frameworks with Enhanced Photodynamic Therapy: Synthesis, Erythrocyte Membrane Camouflage, and Aptamer-Targeted Aggregation. ACS Appl. Mater. Interfaces 2020, 12, 23697–23706. [Google Scholar] [CrossRef] [PubMed]
  165. Chen, S.; Zhong, Y.; Fan, W.; Xiang, J.; Wang, G.; Zhou, Q.; Wang, J.; Geng, Y.; Sun, R.; Zhang, Z.; et al. Enhanced tumour penetration and prolonged circulation in blood of polyzwitterion–drug conjugates with cell-membrane affinity. Nat. Biomed. Eng. 2021, 5, 1019–1037. [Google Scholar] [CrossRef] [PubMed]
  166. van Dommelen, S.M.; Vader, P.; Lakhal, S.; Kooijmans, S.; van Solinge, W.W.; Wood, M.J.; Schiffelers, R.M. Microvesicles and exosomes: Opportunities for cell-derived membrane vesicles in drug delivery. J. Control. Release 2012, 161, 635–644. [Google Scholar] [CrossRef] [PubMed]
  167. Li, P.; Kaslan, M.; Lee, S.H.; Yao, J.; Gao, Z. Progress in exosome isolation techniques. Theranostics 2017, 7, 789–804. [Google Scholar] [CrossRef] [PubMed]
  168. Liu, A.; Yang, G.; Liu, Y.; Liu, T. Research progress in membrane fusion-based hybrid exosomes for drug delivery systems, Front. Bioeng. Biotechnol. 2022, 10, 939441. [Google Scholar] [CrossRef]
  169. 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]
  170. Cheng, L.; Zhang, X.; Tang, J.; Lv, Q.; Liu, J. Gene-engineered exosomes-thermosensitive liposomes hybrid nanovesicles by the blockade of CD47 signal for combined photothermal therapy and cancer immunotherapy. Biomaterials 2021, 275, 120964. [Google Scholar] [CrossRef]
  171. Lin, Y.; Wu, J.; Gu, W.; Huang, Y.; Tong, Z.; Huang, L.; Tan, J. Exosome-liposome hybrid nanoparticles deliver CRISPR/Cas9 system in MSCs. Adv. Sci. 2018, 5, 1700611. [Google Scholar] [CrossRef] [PubMed]
  172. Jhan, Y.-Y.; Prasca-Chamorro, D.; Zuniga, G.P.; Moore, D.M.; Kumar, S.A.; Gaharwar, A.K.; Bishop, C.J. Engineered extracellular vesicles with synthetic lipids via membrane fusion to establish efficient gene delivery. Int. J. Pharm. 2020, 573, 118802. [Google Scholar] [CrossRef] [PubMed]
  173. Zhou, X.; Miao, Y.; Wang, Y.; He, S.; Guo, L.; Mao, J.; Chen, M.; Yang, Y.; Zhang, X.; Gan, Y. Tumour-derived extracellular vesicle membrane hybrid lipid nanovesicles enhance siRNA delivery by tumour-homing and intracellular freeway transportation. J. Extracell. Vesicle 2022, 11, e12198. [Google Scholar] [CrossRef] [PubMed]
  174. Hu, Y.; Li, X.; Zhang, Q.; Gu, Z.; Luo, Y.; Guo, J.; Wang, X.; Jing, Y.; Chen, X.; Su, J. Exosome-guided bone targeted delivery of Antagomir-188 as an anabolic therapy for bone loss. Bioact. Mater. 2021, 6, 2905–2913. [Google Scholar] [CrossRef] [PubMed]
  175. Evers, M.J.W.; van de Wakker, S.I.; de Groot, E.M.; de Jong, O.G.; Gitz-François, J.J.J.; Seinen, C.S.; Sluijter, J.P.G.; Schiffelers, R.M.; Vader, P. Functional siRNA delivery by extracellular vesicle-liposome hybrid nanoparticles. Adv. Healthc. Mater. 2021, 11, 2101202. [Google Scholar] [CrossRef] [PubMed]
  176. Hu, M.; Zhang, J.; Kong, L.; Yu, Y.; Hu, Q.; Yang, T.; Wang, Y.; Tu, K.; Qiao, Q.; Qin, X.; et al. Immunogenic hybrid nanovesicles of liposomes and tumor-derived nanovesicles for cancer immunochemotherapy. ACS Nano 2021, 15, 3123–3138. [Google Scholar] [CrossRef] [PubMed]
  177. Yang, B.; Chen, Y.; Shi, J. Exosome biochemistry and advanced nanotechnology for next-generation theranostic platforms. Adv. Mater. 2019, 31, 1802896. [Google Scholar] [CrossRef] [PubMed]
  178. Rayamajhi, S.; Nguyen, T.D.T.; Marasini, R.; Aryal, S. Macrophage-derived exosome-mimetic hybrid vesicles for tumor targeted drug delivery. Acta Biomater. 2019, 94, 482–494. [Google Scholar] [CrossRef] [PubMed]
  179. Wang, R.; Wang, X.; Zhao, H.; Li, N.; Li, J.; Zhang, H.; Di, L. Targeted delivery of hybrid nanovesicles for enhanced brain penetration to achieve synergistic therapy of glioma. J. Control. Release 2024, 365, 331–347. [Google Scholar] [CrossRef]
  180. Sun, L.; Fan, M.; Huang, D.; Li, B.; Xu, R.; Gao, F.; Chen, Y. Clodronate-loaded liposomal and fibroblast-derived exosomal hybrid system for enhanced drug delivery to pulmonary fibrosis. Biomaterials 2021, 271, 120761. [Google Scholar] [CrossRef]
  181. Ji, K.; Fan, M.; Huang, D.; Sun, L.; Li, B.; Xu, R.; Zhang, J.; Shao, X.; Chen, Y. Clodronate-nintedanib-loaded exosome-liposome hybridization enhances the liver fibrosis therapy by inhibiting Kupffer cell activity. Biomater. Sci. 2022, 10, 702–713. [Google Scholar] [CrossRef] [PubMed]
  182. Li, L.; He, D.; Guo, Q.; Zhang, Z.; Ru, D.; Wang, L.; Gong, K.; Liu, F.; Duan, Y.; Li, H. Exosome-liposome hybrid nanoparticle codelivery of TP and miR497 conspicuously overcomes chemoresistant ovarian cancer. J. Nanobiotechnol. 2022, 20, 50. [Google Scholar] [CrossRef] [PubMed]
  183. Ma, Y.; Zhang, Y.; Han, R.; Li, Y.; Zhai, Y.; Qian, Z.; Gu, Y.; Li, S. A cascade synergetic strategy induced by photothermal effect based on platelet exosome nanoparticles for tumor therapy. Biomaterials 2022, 282, 121384. [Google Scholar] [CrossRef] [PubMed]
  184. Lv, Q.; Cheng, L.; Lu, Y.; Zhang, X.; Wang, Y.; Deng, J.; Zhou, J.; Liu, B. Thermosensitive exosome-liposome hybrid nanoparticle-mediated chemoimmunotherapy for improved treatment of metastatic peritoneal cancer. Adv. Sci. 2020, 7, 2000515. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the composition (families of proteins, lipids, and nucleic acids) and membrane orientation of EVs. Tetraspanins are commonly found in EVs and include CD63, CD81, and CD9. Each listed component may be present but not always simultaneously. Abbreviations: ESCRT, endosomal sorting complex required for transport; MHC, major histocompatibility complex. This figure is inspired by the one published by Colombo et al. [2].
Figure 1. Schematic representation of the composition (families of proteins, lipids, and nucleic acids) and membrane orientation of EVs. Tetraspanins are commonly found in EVs and include CD63, CD81, and CD9. Each listed component may be present but not always simultaneously. Abbreviations: ESCRT, endosomal sorting complex required for transport; MHC, major histocompatibility complex. This figure is inspired by the one published by Colombo et al. [2].
Nanomaterials 14 00639 g001
Figure 2. Methods for drug loading into exosomes, including co-incubation, saponin incubation, electroporation, freeze–thaw cycle, extrusion, and sonication. Different methods can be combined. The drug-loading efficiency of exosomes depends on the loading methods and conditions.
Figure 2. Methods for drug loading into exosomes, including co-incubation, saponin incubation, electroporation, freeze–thaw cycle, extrusion, and sonication. Different methods can be combined. The drug-loading efficiency of exosomes depends on the loading methods and conditions.
Nanomaterials 14 00639 g002
Figure 3. Exosome mimes are constituted by cell or EV membrane elements that are covering nanoparticles or that are mixed with liposome ingredients. These strategies allow to take advantage of 1) the shape and tracking-ability of the nanoparticles, or 2) of the loading capacity of liposomes, and 3) retains the exosome capacity to escape immune reactions and to target the recipient cells for drug delivery.
Figure 3. Exosome mimes are constituted by cell or EV membrane elements that are covering nanoparticles or that are mixed with liposome ingredients. These strategies allow to take advantage of 1) the shape and tracking-ability of the nanoparticles, or 2) of the loading capacity of liposomes, and 3) retains the exosome capacity to escape immune reactions and to target the recipient cells for drug delivery.
Nanomaterials 14 00639 g003
Table 1. Clinical trials running between 2011 and 2024 that involved exosomes in primary outcome measures. Source: classic.clinicaltrials.gov. Key word EXOSOME N = 171.
Table 1. Clinical trials running between 2011 and 2024 that involved exosomes in primary outcome measures. Source: classic.clinicaltrials.gov. Key word EXOSOME N = 171.
Pathology TypePathologyStudyTypeStart DateNCT Number
Blood pathologySepsisFunction of circulating exosomes in sepsis-induced immunosuppressionO15/04/2021NCT04979767
Brain diseaseAlzheimer’s diseaseSafety and efficacy evaluation of allogenic adipose MSC-Exos in patients with Alzheimer’s diseaseI01/07/2020NCT04388982
Brain diseaseParkinson’s diseaseLRRK2 and other novel exosome proteins in Parkinson’s disease: biomarkers associated with Parkinson’s disease susceptibility and/or progression in exosome-proteomes derivedO01/01/2013NCT01860118
CancerBladder cancerCharacterization of the protein profile in tumor derived exosomes from the gallbladder carcinoma patientsO01/01/2018NCT03581435
CancerBladder cancerUse of urinary exosome lncRNAs for preoperative diagnosis of lymphatic metastasis in patients with bladder cancerO01/06/2023NCT05270174
CancerBladder cancerUse of urine exosomal RNA for accurate diagnosis of urinary tract tumors and the development of kitsO01/06/2023NCT06193941
CancerBone cancerIdentification and characterization of miRNAs content in circulating tumor exosomes in bone metastasesO03/12/2018NCT03895216
CancerBrain cancerClinical relevance of detecting molecular abnormalities in glial tumor exosomes (blood sampling)I15/12/2023NCT06116903
CancerBreast cancerOmic technologies to track resistance to palbociclib in metastatic breast cancer: longitudinal changes in volatile organic components profile and exosomes according to response to treatmentI08/09/2020NCT04653740
CancerBreast cancerDevelopment of a horizontal data integration classifier for noninvasive early diagnosis of breast cancer (from different radiomics analyses on baseline mammograms and molecular analyses on peripheral blood (miRNA sequencing from exosomes) and urineI19/01/2021NCT04781062
CancerBreast cancerExosome as prognostic and predictive biomarker in early breast cancer patientsI01/05/2021NCT05955521
CancerBreast cancerFeasibility of exosome analysis in cerebrospinal fluid during the diagnostic workup of metastatic meningitis (Exo-LCR)I04/01/2023NCT05286684
CancerBreast cancerStudy to measure the expression of the HER2-HER3 dimer in tumor and blood (exosomes) samples from patients with HER2 positive breast cancer receiving HER2 targeted therapiesO20/12/2019NCT04288141
CancerCancerPilot study with the aim to quantify a stress protein in the blood and in the urine for the monitoring and early diagnosis of malignant solid tumors: concentration of HSP70 exosomes in the blood and urineI15/12/2015NCT02662621
CancerColon cancerStudy investigating the ability of plant exosomes to deliver curcumin to normal and colon cancer tissueI01/01/2011NCT01294072
CancerColorectal cancerContents of circulating extracellular vesicles: prognostic role of exosomes and their contents on the survival of colorectal cancer patientsO01/07/2020NCT04523389
CancerColorectal cancerIdentification in blood sample of new diagnostic protein markers derived from circulating tumor exosomes for colorectal cancerO07/01/2021NCT04394572
CancerEpithelial cancer
Bowel disease
Plant exosomes +/− curcumin to abrogate symptoms of inflammatory bowel diseaseI01/03/2018NCT04879810
CancerGastric cancerCirculating exosomes as potential prognostic and predictive biomarkers in advanced gastric cancer patients (“EXO-PPP study“): characterization of the molecular profile in tumor derived exosomes, correlation of plasma level and kinetics of gastric cancer derived exosomesO01/01/2013NCT01779583
CancerLiver cancerEarly plasmatic biomarkers of tumor response in high dose hypofractionated radiotherapy. Work Package 3: Immune response (analyses of immunological parameters, description of secreted markers and nanovesicles production, verification of the presence and evolution of activation markers and quantification of secreted exosomes)I16/09/2015NCT02439008
CancerLung cancerPrediction of immunotherapeutic effect of advanced non-small cell lung cancer (detection of the difference of miRNA expression profiles of exosomes in NSCLC patients before and after immunotherapy -pabolizumab, nafulizumab-)I08/06/2020NCT04427475
CancerLung cancerMolecular profiling of exosomes in tumor-draining vein of early-staged lung cancerI21/06/2021NCT04939324
CancerLung cancerCombined diagnosis of cancerous tissue and exosome in early lung cancerO20/05/2018NCT03542253
CancerLung cancerEarly diagnosis of lung cancer using blood plasma derived exosome (evaluating the possibility of distinguishing between normal and lung cancer patients through the analysis of lung cancer-specific exosomal protein)O09/04/2020NCT04529915
CancerLung cancerStudy of exosome EML4-ALK fusion in NSCLC clinical diagnosis and dynamic monitoring (the objective response rate of those NSCLC patients receiving ALK inhibitors treatment according to exosome ALK fusion diagnosis and fluorescence in situ hybridization examination)O01/08/2020NCT04499794
CancerLung cancerSerum exosomal miRNA predicting the therapeutic efficiency in lung squamous carcinomaO01/04/2022NCT05854030
CancerLung cancerExtracellular vesicles (isolate exosomes from bronchial washings) and particles as biomarkers of recurrence in non-small cell lung cancerO14/06/2022NCT05424029
CancerLymphoma, B-cellaggressive non-Hodgkin (B-NHL)Exosomes and immunotherapy in non-Hodgkin B-cell lymphomas (quantification of CD20 and PDL-1 in exosomes purified from cell cultures of DLBCL human cell lines and from healthy volunteers)I02/07/2019NCT03985696
CancerMelanomaAnalysis of circulating exosomes in plasma of melanoma patients (dosage of proteic biomarkers in circulating exosome)O01/03/2019NCT05744076
CancerOropharyngeal cancerExosome testing as a screening modality for human papillomavirus-positive oropharyngeal squamous cell carcinomaO25/02/2015NCT02147418
CancerOvarian cancerNon-coding RNA in the exosome of the epithelia ovarian cancer (expression of miRNA/lncRNA, expression of micro-RNA (miRNA) and long non-coding RNA (lncRNA) compared between high grade serous ovarian carcinoma group and control group)O10/11/2018NCT03738319
CancerPancreatic cancerTrial of ascorbic acid + triple therapy with nanoparticle paclitaxel protein bound + cisplatin + gemcitabine in patients with advanced stage IV metastatic pancreatic cancerI15/12/2017NCT03410030
CancerPancreatic cancerCirculating extracellular exosomal small RNA as potential biomarker for human pancreatic cancerI01/11/2020NCT04636788
CancerPancreatic canceriExosomes in Treating Participants With Metastatic Pancreas Cancer With KrasG12D Mutation (Mesenchymal Stromal Cells-derived Exosomes with KRAS G12D siRNA)I27/01/2021NCT03608631
CancerProstate cancerExosomal microRNA in predicting the aggressiveness of prostate cancer in Chinese patients by comparison of the differences in microRNA expression between non-prostate cancer subjects, pathologically insignificant and significant prostate cancer patientsO03/05/2018NCT03911999
CancerProstate cancerExoDx prostate evaluation in prior negative prostate biopsy settingO15/03/2020NCT04357717
CancerRectal cancerExosomes in rectal cancer (characterization of plasmatic exosomal biomarker levels in patients with locally advanced rectal cancer undergoing neoadjuvant chemoradiation therapy)O13/02/2018NCT03874559
CancerRenal cancerEvaluation of urinary exosomes presence from clear cell renal cell carcinomaO29/01/2020NCT04053855
CancerRenal cancerA companion diagnostic study to develop circulating exosomes as predictive biomarkers for the response to immunotherapy in renal cell carcinoma (blood and urine collection)O01/01/2023NCT05705583
CancerSarcomaStudy of blood exosomes in monitoring patients with sarcoma (EXOSARC)O19/11/2018NCT03800121
CancerSystemic autoimmune diseasesUrine miRNAs-exosomes to identify biomarkers for lupus nephritisO02/08/2020NCT04894695
Chronic diseaseDiabetes type 2Determine the levels of circulating extracellular vesicles released by human islets of langerhans in plasmaO01/12/2016NCT03106246
Chronic diseaseDiabetes type 2Mechanisms behind severe insulin resistance during pregnancy in women with glucose metabolic disorders (SIR-MET) maternal hormonal, inflammatory and metabolic markers in the blood, as well as the level, content and bioactivity of exosomes are studiedO01/05/2021NCT04924504
Chronic diseaseDiabetic retinopathyRole of the serum exosomal miRNA (sequencing) in diabetic retinopathyO01/07/2018NCT03264976
Chronic diseaseDiabetic retinopathyStudy on exosome changes from plasma, atrial fluid and vitreous fluid in patients with proliferative diabetic retinopathy by proteomic analysis of proteinsO01/01/2024NCT06198543
Chronic diseaseDiabetic retinopathyProteomic study of plasma exosomes in patients with diabetic retinopathyO02/01/2024NCT06188013
Chronic diseaseHypertensionNew biomarkers and difficult-to-treat hypertension; concentrations and variabilities of urinary exosomal sodium channels and plasma angiotensins in patients with difficult-to-treat arterial hypertensionO01/05/2016NCT03034265
Endothelial pathologyEndothelial dysfunctionCirculating exosomes and endothelial dysfunction in patients with obstructive sleep apneas hypopneas syndrome: miRNA contained in exosomes, comparison of exosome content between obese-OSA patients with endothelial dysfunction and withoutI14/01/2021NCT04459182
Genetic disaseDystrophic epidermolysis bullosaStudy to assess the effectiveness and safety of AGLE-102, an allogeneic derived MSC EVs product derived from normal donor on lesions in subjects with dystrophic epidermolysis bullosaI01/01/2024NCT04173650
Heart pathologyCardiac remodelingExosome as integrative tool for prognostic stratification of adverse cardiac remodeling in stemi patients: the MIRACLE Study (verify whether the profile of circulating plasma exosomes reflects cardiovascular magnetic resonance in acute ST-segment–elevation myocardial infarction)O03/05/2023NCT06070974
Heart pathologyAtrial fibrillationRole of exosomes derived from epicardial fat in atrial fibrillation: analysis of the differences in the quantity of exosomes derived epicardial fat biopsy in patients with and without atrial fibrillation.I21/01/2018NCT03478410
Infectious diseaseCOVID-19Intravenous infusion of CAP-1002 (cardiosphere-derived cells) in patients with COVID-19I15/11/2020NCT04623671
Infectious diseaseCOVID-19The use of exosomes, delivered intravenously-ARDOXSO™, for the treatment of acute respiratory distress syndrome or novel coronavirus pneumonia caused by COVID-19I2023-09NCT04798716
Infectious diseaseHIVInflammation, NK cells, antisense protein and exosomes, and correlation with immune response during HIV infectionI22/04/2022NCT05243381
Respiratory pathologyAcute respiratory distress syndromeScreening of differential miRNAs of inflammatory exosomes in plasma and alveolar lavage fluid of patients with sepsis complicated with acute respiratory distress syndromeO25/07/2022NCT05476029
Respiratory pathologyObstructive sleep apnea syndromeExosomes implication in PD1-PD-L1 activation in obstructive sleep apnea syndromeO03/03/2019NCT03811600
Respiratory pathologyAcute respiratory distress syndromeSafety and efficacy of EXO-CD24 in preventing clinical deterioration in patients with mild–moderate acute respiratory distress syndromeI04/07/2023NCT05947747
Respiratory pathologyRespiratory dysfunctionBlood circulating microRNAs as biomarkers of respiratory dysfunction in patients with refractory epilepsyI16/03/2018NCT03419000
Systemic autoimmune diseasesPsoriasisSafety and tolerability study of MSC exosome ointmentI08/03/2022NCT05523011
I: interventional, O: observational.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Poinsot, V.; Pizzinat, N.; Ong-Meang, V. Engineered and Mimicked Extracellular Nanovesicles for Therapeutic Delivery. Nanomaterials 2024, 14, 639. https://doi.org/10.3390/nano14070639

AMA Style

Poinsot V, Pizzinat N, Ong-Meang V. Engineered and Mimicked Extracellular Nanovesicles for Therapeutic Delivery. Nanomaterials. 2024; 14(7):639. https://doi.org/10.3390/nano14070639

Chicago/Turabian Style

Poinsot, Verena, Nathalie Pizzinat, and Varravaddheay Ong-Meang. 2024. "Engineered and Mimicked Extracellular Nanovesicles for Therapeutic Delivery" Nanomaterials 14, no. 7: 639. https://doi.org/10.3390/nano14070639

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