Skip to main content

Bioengineered exosomal-membrane-camouflaged abiotic nanocarriers: neurodegenerative diseases, tissue engineering and regenerative medicine

Abstract

A bio-inspired strategy has recently been developed for camouflaging nanocarriers with biomembranes, such as natural cell membranes or subcellular structure-derived membranes. This strategy endows cloaked nanomaterials with improved interfacial properties, superior cell targeting, immune evasion potential, and prolonged duration of systemic circulation. Here, we summarize recent advances in the production and application of exosomal membrane-coated nanomaterials. The structure, properties, and manner in which exosomes communicate with cells are first reviewed. This is followed by a discussion of the types of exosomes and their fabrication methods. We then discuss the applications of biomimetic exosomes and membrane-cloaked nanocarriers in tissue engineering, regenerative medicine, imaging, and the treatment of neurodegenerative diseases. Finally, we appraise the current challenges associated with the clinical translation of biomimetic exosomal membrane-surface-engineered nanovehicles and evaluate the future of this technology.

Background

Nanomaterials have the potential to be used to diagnose and treat various human diseases due to their unique ability to deliver therapeutic bioactive molecules to target sites [1,2,3,4]. Treatment strategies utilizing nanomaterials have demonstrated improved efficacy and safety when compared to conventional therapies [5,6,7,8,9]. Despite the many potential applications of nanoparticles (NPs) in medicine, their clinical use is limited owing to their poor biocompatibility and inability to cross biological barriers. Due to their foreign nature, abiotic nanomaterials are rapidly cleared by the body’s mononuclear phagocyte system, resulting in a short duration of systemic circulation and reduced delivery efficacy to target sites [10].

To circumvent these hurdles, recent studies have focused on camouflaging abiotic NPs with biological cell membranes, such as those of red blood cells [11], white blood cells [12], platelets [13], stem cells [14], or cancer cells [15], to improve in vivo interactions and biofunctionality. This involves functionalizing the surface of NPs with a cell membrane via top-down approaches [12, 16]. This promising cell-mimicking approach enables NPs to acquire the inherent biological properties of progenitor cell membranes. By covering NPs with a natural cell membrane, the antigenic profile and interfacial properties of the progenitor cell can be faithfully preserved and transferred to the abiotic NPs [4, 17].

Another variety of biomimetic and nature-inspired technologies uses membranes of subcellular structures. Recently, exosomal membranes have attracted considerable interest for use as nanomaterial coatings [18, 19]. Exosomes are produced by cells and have optimal nanoscale sizes. Exosome membranes are more biomimetic than synthetic membranes when used to coat NPs. Membrane extraction from exosomes does not require aggressive techniques such as extrusion or sonication, which are often employed for cell membrane extraction and nanovesicle derivation. Exosomes are excellent intercellular messengers optimized for intercellular communication and interaction [20, 21]. For these reasons, coating NPs with the membranes of naturally secreted exosomes offers many advantages over the use of natural cell membranes. These advantages include intrinsic targeting, cell-specific uptake, prolonged systemic circulation, enhanced biocompatibility, stability, and immune evasion. Exosomal membrane-coated NPs have demonstrated exciting results, improving therapeutic efficacy and reducing off-target toxicity in healthy tissues (Fig. 1) [18, 19, 21].

Fig. 1
figure 1

Exosomal-membrane-coated nanosystems are promising nanotechnological tools for biomedical applications. Depiction of the biological benefits of exosomal-membrane-coated nanosystems, including intrinsic tissue-targeting and tissue-specific accumulation features, prolonged blood circulation, and enhanced biocompatibility, stability, and immune evasion abilities, as well as their applications in biomedical settings

The present review provides an overview of the most recent research and current advances in exosomal-membrane-coated nanomaterials. We first give an overview of the composition, mechanisms of biogenesis, and biological functions of natural exosomes, then discuss the fabrication of exosomal-membrane-functionalized NPs. Next, we review the biomedical applications of these biomimetic nanostructures, including tissue regeneration and the diagnosis and treatment of neurodegenerative diseases (Fig. 1). Finally, we discuss the major challenges for successfully implementing this technology in clinical settings and our perspectives on the future of this emerging biomimetic coating approach.

Exosomes: structure, properties, and cell communication

According to the Minimal Information for Studies of Extracellular Vesicles (MISEV) 2018 guidelines proposed by the International Society for Extracellular Vesicles (ISEV), extracellular vesicles (EVs) are defined as natural, cell-secreted vesicles bound by a phospholipid bilayer that are unable to replicate as they do not contain a functional nucleus [22].

Exosomes are a small subtype of cell-secreted EVs of endocytic origin, ranging from 30 to 150 nm in size [23, 24]. They function as mediators of cell–cell communication by delivering a wide range of biological components, such as proteins, lipids, and nucleic acids, to neighboring and distant cells. Exosomes are thus important messengers in intercellular communication [25]. Because of their ability to transport biomolecules between surrounding and distant cells, exosomes can mediate short- and long-distance cell–cell communication and influence various physiological and pathological functions of recipient cells [20].

Structure and physiology of exosomes

Similar to synthetic liposomes, exosomes have an amphiphilic structure consisting of an aqueous core surrounded by a phospholipid bilayer [26]. As shown in Fig. 2, exosomes are mainly composed of a diverse set of proteins [27], lipids [28], and nucleic acids [29]. The biological contents of an exosome resemble the composition of the cell that secreted it. As a result, exosome composition is directly related to the physiopathological status of their progenitor cells and can change in response to changes in physiological and pathological conditions [20, 30].

Fig. 2
figure 2

Exosome structure and composition. Exosomes are enriched with lipid rafts, nucleic acids, and proteins. The latter include adhesion molecules, tetraspanins (e.g., CD9, CD63, CD81, and CD82), proteins responsible for membrane transport and fusion (e.g., annexin and Rab GTPase), proteins involved in MVB biogenesis (e.g., Alix and Tsg101), heat shock proteins (e.g., Hsp70 and Hsp90), and cytoskeletal proteins (e.g., actin and tubulin). MHC class I and class II proteins may also be found in exosomes. Alix apoptosis-linked gene-2 interacting protein X, DNA deoxyribonucleic acid, Hsp heat shock protein, MHC major histocompatibility complex, miRNA microRNA, mRNA messenger RNA, MVB multivesicular body, Tsg101 tumor susceptibility gene 101

Exosomes are enriched in multiple proteins, both within and on their surface membranes. These proteins include adhesion molecules (e.g., integrins) [31], proteins responsible for membrane transport and fusion (e.g., annexins and Rab GTPases) [32], cytoskeletal proteins (e.g., actin and tubulin) [33], heat shock proteins (Hsps, e.g., Hsp70 and Hsp90) [34, 35], and proteins involved in the biogenesis of multivesicular bodies (MVBs), such as apoptosis-linked gene-2 interacting protein X (Alix) and tumor susceptibility gene 101 (Tsg101) [36]. Lysosomal proteins [e.g., lysosome-associated membrane glycoprotein 2b (Lamp2b)] [37] and surface tetraspanins (e.g., CD9, CD63, CD81, and CD82) [30, 38, 39] are also present in exosomes. The tetraspanins CD9 and CD81 facilitate direct membrane fusion between exosomes and target cells [40]. The tetraspanins CD55 and CD59 offer protection against complement membrane attacks [41]. Some of the above-mentioned proteins (CD9, CD63, CD81, Alix, Tsg101, and Hsp70) are often considered exosomal markers [30, 42]. Expression of the “self-marker” CD47 in some subsets of exosomes, a “don’t eat me” signal, avoids immune phagocytic clearance and increases the stability of exosomes in systemic circulation [43, 44]. Exosomes may also contain major histocompatibility complex (MHC) class I and II proteins that are responsible for antigen presentation [39].

The composition of exosomal phospholipid bilayers resembles that of their progenitor cells. The phospholipid bilayer is abundant in lipid rafts (submicroscopic membrane microdomains), which are rich in ceramides, cholesterol, sphingolipids, and phosphoglycerides. They are responsible for regulating cargo sorting into MVBs, exosome formation, rigidity, and structure [45, 46]. The lipid composition of exosomal membranes not only enables them to fuse directly with the plasma membranes of recipient cells, but also increases the physicochemical stability of exosomes in the extracellular environment [28]. This protects the exosomal cargo from degradation to ensure its integrity until it is distributed to target cells [18, 47].

In addition to their protein and lipid compositions, exosomes are carriers of a wide range of genetic materials that can be transmitted to neighboring and distant cells. These genetic materials include RNA molecules [e.g., messenger RNA (mRNA) and microRNA (miRNA)] and deoxyribonucleic acid (DNA) molecules (e.g., mitochondrial DNA and chromosomal DNA) [18, 47, 48]. Exosomal components and their main biofunctions are summarized in Table 1 [32, 34, 39, 40, 43, 44, 47,48,49,50,51,52,53].

Table 1 Common exosomal components and their main biofunctions

Exosome biogenesis

Exosome biogenesis typically involves: 1) invagination of the plasma membrane by inward budding, 2) accumulation of intraluminal vesicles within MVBs by inward budding of the MVB membrane, 3) fusion of MVBs with the plasma membrane, and 4) release of intraluminal vesicles as exosomes into the extracellular space upon fusion of the MVB with the plasma membrane [18, 47].

Exosome formation begins with invagination of the plasma membrane by inward budding, forming early endosomes. These structures then undergo a sequence of alterations to form late endosomes, which are also known as MVBs, and which are characterized by the presence of several intraluminal vesicles in their luminal space. Intraluminal vesicles are formed by inward budding of the MVB membrane [30, 56]. Once MVBs containing several intraluminal vesicles are formed, they can have one of two different fates: 1) degradation by fusion of the MVB with a lysosome, or 2) exocytosis through fusion of the MVB with the plasma membrane, leading to the release of intraluminal vesicles as exosomes into the extracellular space (Fig. 3) [30, 38, 57]. Secretion of exosomes into the extracellular environment through exocytosis is dependent on soluble N-ethylmaleimide-sensitive fusion attachment protein receptors and Rab GTPases such as Rab-27a, RAB-11, and Rab-31 [38, 58, 59].

Fig. 3
figure 3

Exosome biogenesis and cell–cell communication. Biogenesis of exosomes begins with inward budding of the plasma membrane to form an early endosome. A second inward budding of the early endosome generates a multivesicular body containing intraluminal vesicles. During the second inward budding, exosomes are loaded with their cargo (mRNAs, non-coding RNAs, proteins, and DNA fragments). Exosomal biogenesis can occur through both ESCRT-dependent and ESCRT-independent pathways. and The multivesicular body can fuse either to a lysosome for the degradation of its components or to the plasma membrane for secretion. and The multivesicular body ultimately fuses with the plasma membrane to release its intraluminal vesicles into the extracellular space as exosomes. The released exosomes may be taken up by target cells through receptor-mediated endocytosis, direct fusion with the recipient plasma membrane, phagocytosis and macropinocytosis, or ⑩ caveolin- and clathrin-mediated endocytosis. ESCRT endosomal sorting complex required for transport

Mechanisms of exosome biogenesis

The most reported mechanism for the formation of intraluminal vesicles within MVBs involves the “endosomal sorting complex required for transport” (ESCRT). This complex comprises four protein complexes (ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III) that function cooperatively to promote exosome biogenesis [18, 38, 47]. The ESCRT-dependent mechanism is initiated by the sequestration of ubiquitinated proteins by ESCRT-0, which subsequently recruits ESCRT-I and ESCRT-II. Both ESCRT-I and ESCRT-II are responsible for invagination of the MVB membrane. ESCRT-III causes the scission of inward budding vesicles [60], resulting in the formation of intraluminal vesicles [30].

Once the concomitant inhibition of all four ESCRT complexes has been shown not to suppress the formation of MVB, alternative ESCRT-independent mechanisms for MVB formation and exosome biogenesis have been suggested [61]. One of the proposed ESCRT-independent mechanisms is dependent on ceramides. A study conducted using mouse oligodendroglial cell lines showed that the secretion of exosomes did not require the ESCRT machinery, but was instead dependent on sphingomyelinase, an enzyme that catalyzes the production of ceramides [53]. The ESCRT-independent mechanism also appears to depend on the tetraspanin CD63, which is abundant in exosomes. It has been shown that CD63 plays an important role in mediating intraluminal vesicle formation [39, 62, 63].

Cell–cell communication in physiological and pathophysiological processes

Initially, exosomes were perceived as a means by which cells discharged unwanted or unnecessary materials, and were thus regarded as cellular waste. Today, it is generally accepted that exosomes serve an additional function by communicating with proximal and distal cells to reprogram those cells [38, 43]. Cell–cell communication is crucial for homeostasis. Exosomes released by healthy and diseased cells function as important mediators of intercellular communication because previously enclosed biomolecules can be delivered to neighboring and distant cells [18, 26].

Once exosomes are released into the extracellular space, they are internalized by recipient cells, which then undergo phenotypic and behavioral changes. Three mechanisms have been proposed for the cellular internalization of exosomes: 1) direct fusion of exosomes with the cell membrane, 2) interaction with cell-surface receptors (ligand-receptor interactions), and 3) uptake of exosomes through endocytosis. The latter includes caveolin-mediated endocytosis, clathrin-mediated endocytosis, lipid-raft-mediated endocytosis, phagocytosis, and macropinocytosis (Fig. 3) [64, 65].

Exosomes provide an important mechanism for short- and long-distance cellular communication. They play a key role in physiological processes such as tissue repair, cell proliferation, blood coagulation, and immune surveillance [38]. Each exosomal source cell can impart specific biofunctionalities that can be employed when developing exosome-based therapies. Recent studies have reported an important role of exosomes in immunomodulation. Immune-cell-derived exosomes can trigger potent immune responses because of their antigen presentation capabilities. Because of the expression of MHC molecules on their surface, exosomes derived from B lymphocytes can present antigens to CD4+ and CD8+ T cells to induce strong immune responses [66]. Exosomes derived from T cells can retain the immunostimulatory and tumor growth inhibitory effects of their progenitor cells [67,68,69,70]. Exosomes secreted from macrophages are endowed with intrinsic tropism towards inflammatory and tumorous tissues [67, 71,72,73,74]. Mesenchymal stem cell (MSC)-derived exosomes can be derived from adipose tissue [75,76,77], bone marrow [78,79,80,81], umbilical cord [82, 83], and human placenta [84], and are most commonly used for tissue regeneration and wound-healing applications [76, 77, 85]. They also exhibit important immunomodulatory properties [86].

In addition to their physiological functions, exosomes also play a vital role in various pathological processes. Prior studies have documented the contribution of exosomes to the spread and progression of neurodegenerative [87], cardiovascular [88], and malignant diseases [89]. Tumor cell-derived exosomes exhibit properties that are similar to those of their parent cells. These exosomes transport tumor antigens to modulate the tumor microenvironment and facilitate tumor dissemination [90]. Exosomes derived from tumor cells are involved in tumor development, tumor cell proliferation, the generation of pre-metastatic niches, the promotion of tumor angiogenesis, and tumor immunosuppression [91,92,93,94,95,96,97,98,99,100,101]. This is achieved by suppressing the activity of natural killer cells, differentiating dendritic cells (DCs), and activating T lymphocytes [90]. Table 2 [65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83, 91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114] shows the exosomes source, biofunctionality and biomedical applications.

Table 2 Source of exosomes, biofunctionality and biomedical applications

Exosomes versus liposomes as drug delivery systems: a comparative overview

Liposomes and exosomes, biological and highly complex liposomal forms, are remarkably similar in terms of diameter and phospholipid bilayer structure, which resembles that of cell membranes [117, 118]. A distinctive feature of exosomes is their complex surface repertoire, which is responsible for enhancing cell-specific targeting and uptake [118,119,120]. Both of these amphiphilic vesicles are promising delivery mechanisms for both hydrophobic and hydrophilic drugs [18, 121].

Liposomes are lipid-based drug delivery systems of a synthetic nature with well-documented therapeutic benefits [121]. However, concerns related to the lack of specific-cell targeting, inability to cross biological barriers, rapid elimination from blood circulation, and immunogenicity have triggered the search for more “biologically inert” approaches [117, 122,123,124].

The idea of harnessing exosomes as drug delivery systems stems from the role natural exosomes play in intercellular communication. Naturally occurring exosomes have emerged as a more complex and biocompatible alternative to liposomes for drug delivery [119]. Some attributes of exosomes that make them more ideal than liposomes for drug delivery include enhanced biocompatibility, non-immunogenicity, and intrinsic cell-specific targeting. The latter is ascribed to the ability of exosomes to preserve the surface membrane composition and intrinsic targeting properties of their progenitor cells [19, 125, 126]. Exosomes secreted by specific cell types exhibit intrinsic cell tropism, which favors their uptake by target cells via well-established mechanisms [44, 119, 120, 127, 128]. Another advantage of natural exosomes as drug delivery systems is their optimal nanoscale size, which facilitates their penetration through biological barriers, such as the blood–brain barrier [129]. In addition, some subsets of exosomes are capable of evading immune recognition and clearance owing to the presence of the “self-marker” CD47 on their surface [130]. Evasion of immune surveillance increases exosomes’ duration of systemic circulation and protects their cargo from degradation [18, 44, 126]. Natural exosomes and synthetic liposomes as advanced drug delivery systems are shown in Table 3 [16, 117,118,119,120,121,122, 125,126,127, 129,130,131,132].

Table 3 Natural exosomes versus synthetic liposomes as advanced drug delivery systems

A close comparison of the biodistribution and pharmacokinetic profiles of liposomes and exosomes is unfortunately very limited and controversial [119]. A great deal of evidence has suggested that natural exosomes are rapidly eliminated from the bloodstream [135, 136], and similar to liposomes, they suffer non-specific accumulation in the liver [137,138,139]. Despite these comparable clearance rates [117], the pharmacokinetic benefits of some subsets of exosomes over liposomes have been strongly supported, with one study showing superior blood circulation of exosomes when compared to liposomes [132]. Bloodstream exosomes were detected 24 h after administration in vivo, which was ascribed to the privileged immunological features of exosomes conferred by innately surface-expressed CD47 [132]. Thus, the in vivo pharmacokinetics of exosomes appear to be related to their membrane protein profiles [133, 139].

In conclusion, when compared to liposomes, exosomes surface-enriched in CD47 can substantially reduce immune clearance; however, further evidence is required. There is still much to be discovered regarding the in vivo fate of exosomes, and the extent to which CD47 expression can shield exosomes from immune recognition and clearance should be further investigated [117]. Despite the site-specific targeting and pharmacokinetic superiority of natural exosomes over liposomes, the complexity and heterogeneity of intra-exosomal contents and the low production and isolation yields of exosomes remain challenging issues for clinical translation [120, 134].

Types of exosomes

Current exosome-based therapeutic platforms include natural exosomes and artificial exosomes (exosome-like NPs). Natural exosomes are endogenous cell-secreted nanovesicles that carry functional biomolecules from their progenitor cells. In addition to their endogenous cargo, exogenous therapeutics can also be loaded into naturally occurring exosomes either by modifying exosome progenitor cells (transfection of progenitor cells) or by loading exosomes directly with specific cargo [20]. Artificial exosomes are synthetic counterparts engineered to possess superior biopharmaceutical acceptability.

Natural exosomes

Natural exosomes have attracted considerable attention owing to their potential diagnostic and therapeutic applications. The physiopathological status of their progenitor cells has a significant impact on the cargo content of natural exosomes, highlighting the interest in using exosomes as biomarkers for pathological conditions [140]. Exosomes can be found in biological fluids such as blood [141], saliva [142], urine [143], and ascites [144], and have been used to non-invasively diagnose a wide range of human diseases [19, 145].

Exosomes carrying exogenous therapeutic cargo

Exosomes have been used experimentally as nanocarriers to deliver various therapeutic cargoes, such as anti-cancer drugs [146, 147], therapeutic proteins [148], nucleic acids [149, 150], and nanomaterials [151], for the treatment of various human diseases. These include cardiovascular and neurodegenerative diseases, wound healing, and cancer applications, with the latter being one of the most researched areas in exosomal therapy—the loading of exosomes with chemotherapeutic molecules has received considerable attention [152]. Exosomes were engineered to express surface ligands that could bind to specific molecules overexpressed by tumor cells to achieve greater accumulation at tumor sites. To further enhance the ability of exosomes to actively target tumor sites, immature DCs were genetically modified to express Lamp2b, an exosomal surface protein that interacts with αV integrin overexpressed by tumor cells [152]. Engineered doxorubicin-loaded exosomes enabled more efficient delivery of doxorubicin to breast cancer cells (95.4%), which was significantly higher than that of non-engineered doxorubicin-loaded exosomes (35.0%). This resulted in improved antitumor performance in vivo [152].

Exosome-based platforms have also been shown to be effective tools for neurodegenerative disease therapy, such as for Parkinson’s disease (PD) [153] and Alzheimer’s disease (AD) [154, 155]. As already mentioned, exosomes can penetrate the blood–brain barrier for effective brain-targeted drug delivery. For instance, both in vitro and in vivo experiments have shown that dopamine-loaded blood exosomes can effectively cross the blood–brain barrier for the targeted delivery of dopamine to the brain. Dopamine-loaded exosomes not only increased drug bioaccumulation in neuronal cells by more than 15-fold, but also reduced systemic toxicity and improved therapeutic efficacy against PD [153]. Regarding therapies targeting AD, several compounds have been encapsulated in natural exosomes, namely curcumin [154] and quercetin [155]. Exosomal drug loading resulted in 2.5-fold-higher brain accumulation in vivo when compared to free quercetin [155]. Similarly, an in vitro study has demonstrated the superior blood–brain barrier crossing ability of curcumin-loaded exosomes when compared to free drugs (60% and 15%, respectively) [154]. Both exosome-based platforms significantly improve cognitive dysfunction and alleviate AD symptoms by suppressing tau protein phosphorylation, thus showing promising results for AD therapy [154, 155].

Exosomal therapy has also recently been applied in wound-healing applications [156]. MSCs have attracted considerable interest for their ability to accelerate wound healing by stimulating cell proliferation and angiogenesis. MSC-derived exosomes have emerged as a novel, cell-free strategy for wound-healing applications because of their progenitor cell-related, tissue-regenerating properties. For instance, in a recent study, a miR-155 inhibitor was loaded into natural exosomes to yield an exosome-based system with synergistic effects on diabetic wound healing and closure [156]. The superior diabetic wound-healing effects of the loaded exosomes were clearly demonstrated in vitro and in vivo, yielding enhanced collagen deposition, re-epithelialization, and angiogenesis [156].

Artificial exosomes: biomimetic exosome-like nanomaterials

Despite the promising potential of naturally cell-secreted exosomes as drug delivery systems, their clinical use is hindered by the reduced number of exosomes naturally secreted by most cells, poor production and isolation yields, a lack of standardized methods for exosome isolation and purification, and low encapsulation efficiency [18, 157]. To overcome these limitations, extensive efforts have been devoted to the study of bio-inspired exosome-like NPs (Fig. 4). These artificial exosomes include: 1) exosome-mimetic nanovesicles, 2) synthetic exosome-like NPs, 3) hybrid exosome-like nanovesicles, and 4) exosomal membrane-coated NPs. A comparative analysis of natural exosomes and different strategies used to yield artificial exosomes is presented in Table 4 [156,157,158,159,160,161,162,163,164,165,166].

Fig. 4
figure 4

Approaches to constructing biomimetic exosome-like nanoparticles. Approaches to constructing biomimetic exosome-like NPs, which can be used as an alternative to natural exosomes, include: a generation of exosome-mimetic nanovesicles through direct extrusion of progenitor cells through porous membranes or by forcing them to move through microfluidic devices (a, a top-down approach). Preparation of synthetic exosome-like NPs using a synthetic strategy inspired by natural exosomes that involves self-assembly of synthetic phospholipid bilayers with specific antibodies, peptides, or surface proteins from biomembranes (b, a bottom-up approach). Fabrication of hybrid exosome-like nanovesicles by fusion of nanovesicles (natural exosomes or cell membrane nanovesicles) with synthetic liposomes through a top-down approach (c), or by fusion of two or more different biomembranes to generate hybrid membranes that incorporate multiple functionalities of different membrane types, and generation of exosomal-membrane- or cell membrane-coated NPs by coating NP cores with biomembranes (exosomal membranes or cell membrane nanovesicles, respectively) via top-down approaches (d). NP nanoparticle

Table 4 Comparison between natural exosomes and the different strategies used to yield artificial exosomes

Cell membrane derivation approach

Scalability remains a significant challenge for the clinical application of natural exosomes. The preparation of exosome-mimetic nanovesicles by the direct disassembly of progenitor cells through top-down approaches (i.e., the disintegration of complex and large molecules into less complex and smaller units) is an effective approach for the stable and scalable production of artificial exosomes. These artificial exosomes help address the low production yield of natural exosomes [158, 160]. This strategy involves the direct extrusion of progenitor cells through porous membranes, and is the most commonly used approach [162, 169, 170]. Alternatively, artificial exosomes can also be produced by forcing cells to move through the microchannels of microfluidic devices [171, 172] (Fig. 4a). The resulting cell-derived nanovesicles have the membrane surface composition and intrinsic targeting features of natural exosomes. For example, doxorubicin-loaded exosome-mimetic nanovesicles have been formed by extruding doxorubicin-loaded monocytes/macrophages through membrane filters. The resulting exosome-mimetic nanovesicles were similar in size and morphology and contained surface protein markers similar to those of natural exosomes. The production yield of nanovesicles was 100-fold higher than that of exosomes. After in vivo administration, exosome-mimetic nanovesicles accumulated efficiently in tumor tissues and inhibited tumor growth [173].

Applications for exosome-mimetic nanovesicles in wound-healing [174] and regenerative medicine [175] have also been reported. Recently, human umbilical MSCs have been repeatedly extruded through porous membranes to generate MSC-derived exosome-mimetic nanovesicles. These nanovesicles were more effective than MSC-derived exosomes in promoting wound-healing by stimulating dermal fibroblast proliferation [174]. Hepatocyte-derived exosomes play a prominent role in liver regeneration [175]. To address the low production yield of natural exosomes, hepatocytes were extruded through porous membranes to produce exosome-mimetic nanovesicles with a 100-fold higher production yield than exosomes. The resulting nanovesicles effectively stimulated liver cell proliferation and regeneration [175].

Synthetic approach

The preparation of synthetic exosome-like NPs using bottom-up approaches has been used to address the heterogeneity and safety concerns of natural exosomes. These bottom-up approaches involve building large and complex molecules by assembling small and less-complex units [158]. Synthetic exosome-like NPs are synthetic constructs inspired by natural exosomes (Fig. 4b). These NPs only include the essential components of natural exosomes. Their preparation involves the assembly of synthetic phospholipid bilayers (e.g., liposomes) that mimic the lipid composition and size of natural exosomes [176, 177]. For instance, exosome-like liposomes (with a lipid composition mimicking that of natural exosomes) were produced and efficiently used as carriers of curcumin for AD therapy, with an encapsulation efficiency of 94% [176]. These biomimetic exosome-like NPs increase curcumin stability and brain distribution, enhancing its neuroprotective effects against AD-related oxidative stress [176]. In addition, the assembled phospholipid bilayers can be subsequently functionalized with specific antibodies [178], peptides [179], and proteins [180], or coupled with membrane proteins extracted from cell membranes [164]. For example, leukocyte-mimicking liposomes (leukosomes) have been designed by incorporating membrane proteins extracted from leukocytes into synthetic liposomes. Leukosomes showed ninefold greater accumulation at melanoma sites than liposomes, meaning that they delivered doxorubicin more efficiently. This resulted in a more targeted therapy with superior antitumor efficacy [181]. In another study, proteins extracted from cancer-cell membranes were incorporated into synthetic liposomes to yield biomimetic liposomes for triple-negative breast cancer therapy. These biomimetic liposomes were coupled with surface-bound elastase to destroy the tumor extracellular matrix and facilitate drug and cytotoxic T cell infiltration. Elastase-bound biomimetic liposomes showed tumor-targeting capability, fostering the accumulation of chemotherapeutics at tumor sites [182].

Apart from modifying liposomes with membrane proteins extracted from cell membranes, surface proteins can also be incorporated into phospholipid bilayers using a cell-free protein synthesis technique [180]. For example, connexin 43 (Cx43)-embedded liposome-coated chitosan NPs have been synthesized using exosome-mimicking phospholipid bilayers. Exosome-like liposomes were used to deliver small interfering RNAs (siRNAs) targeting vascular endothelial growth factor (VEGF) to glioblastoma cells [180]. The chitosan NPs were first loaded with VEGF siRNA through electrostatic interactions and subsequently camouflaged with exosome-mimicking membranes. Cx43 integration improved glioblastoma cell delivery efficiency via Cx43-mediated gap-function channels, resulting in a 30% reduction in VEGF expression [180]. In a similar effort to enhance the cell internalization efficiency of liposomes, exosome-mimicking liposomes were created to combine the advantages of both entities [131]. The cell uptake efficiency of exosomes was 3-fold higher than that of liposomes due to the exosomes’ leveraging of cell internalization mechanisms [131].

Hybrid approach

Hybrid exosome-like nanovesicles have been prepared using top-down approaches to combine the biological functions of natural exosomes with the pharmaceutical benefits of nanomaterials [158]. The fabrication of hybrid exosome-like nanovesicles involves the fusion of the membranes of nanovesicles (natural exosomes or cell membrane nanovesicles) with synthetic liposomes (Fig. 4c), thus combining the benefits of exosomes and liposomes [165, 183, 184]. For example, exosome-liposome hybrids were prepared through membrane fusion of Raw264.7 cell-derived exosomes with synthetic liposomes using a freeze–thaw method. The cell internalization efficiency of Raw264.7 cell-derived exosome-liposome hybrids was almost 2-fold higher than that of natural exosomes [185].

Another variation of the hybrid approach involves the fusion of two or more biomembranes to create a hybrid membrane that incorporates the functionalities of each membrane. One of these hybrid systems was produced by fusing platelet membranes with membranes of bone marrow MSC-derived EVs for the treatment of ischemic heart disease. This hybrid system combines the intrinsic injured vasculature-targeting ability of platelets with the pro-angiogenic functions of EVs. The hybrid nanovesicles showed 1.8-fold higher accumulation in ischemic heart areas than unmodified EVs [186]. In another study, MSC-derived exosomes were fused with the platelet membrane via extrusion to yield a hybrid system for the treatment of myocardial infarction (MI). The hybrid nanovesicles were readily taken up by endothelial cells and cardiomyocytes because of their inherent ability to target injured vasculature, resulting in improved cardiac function in vivo [187].

Membrane-coated approach

Nanoscale materials are nanotechnological tools that are well-suited for drug delivery. Nanotechnology-based drug delivery systems can enhance the therapeutic and safety goals of conventional therapies, improving the diagnosis and treatment of various human diseases. However, despite the promising potential of nanomaterials as drug delivery systems, some drawbacks hinder their clinical translation [188] (Fig. 5).

Fig. 5
figure 5

Depiction of the main advantages (left) and disadvantages (right) of current nanotechnology-based drug delivery systems. ↑ indicates enhancement, ↓ indicates reduction. MOF metal organic framework, NP nanoparticle

To overcome some of the aforementioned drawbacks of nanoscale materials, recent studies have focused on coating nanomaterials with various types of biological membranes to produce biomimetic carriers. This helps to improve the interfacial properties of NPs, endowing them with prolonged systemic circulation and enhanced biocompatibility, immune evasion, and tissue specificity [12]. The membranes used for coating have included natural cell membranes and subcellular structures, such as membranes derived from exosomes (Fig. 4d).

Exosomal-membrane-coated NPs combine the advantages of endogenous exosomes (enhanced biocompatibility, reduced clearance by the mononuclear phagocyte system, and tissue specificity) with the pharmaceutical benefits of nanomaterials (higher drug-loading ability, easy scalability, greater flexibility to undergo surface modification, and controlled drug release) while overcoming their limitations [18, 19]. Exosomal-membrane-coated NPs are generated by coating the inner core of an NP with an exosomal membrane using a top-down approach. Thus, the inherent biological features of the exosomal membrane can be preserved and transferred to the NP [18, 19]. Hence, surface-engineering via exosomal-membrane-coated nanosystems offers substantial benefits over non-coated nanomaterials by extending their systemic half-life and enhancing tissue specificity [189]. For instance, the uptake of exosomal-membrane-coated metal–organic framework NPs by macrophages was reported to be only 30% of that of uncoated NPs [190]. In another study, exosomal membrane functionalization improved targeted accumulation in homotypic murine 4T1 breast tumors by 3.1-fold when compared to their non-coated counterparts [191].

Fabrication: engineering of exosomal-membrane-coated NPs

As shown in Fig. 6, the preparation of exosomal-membrane-coated nanosystems typically comprises three steps: 1) extraction of exosomal membranes through hypotonic treatment of exosomes, 2) selection and synthesis of the NP inner core, and 3) coating of the synthesized NP core with the extracted exosomal membrane to form a core–shell nanostructure [192].

Fig. 6
figure 6

Three steps involved in synthesizing exosomal-membrane-coated nanosystems. extraction of exosomal membranes through hypotonic treatment of exosomes that were previously isolated from cell-culture supernatants via ultracentrifugation, selection and fabrication of the NP inner core, and coating of the synthesized NP inner core with the extracted exosomal membrane via top-down approaches to obtain a core–shell nanostructure. MOF metal organic framework, NP nanoparticle, NK natural killer

Exosomal membrane extraction

The preparation of exosomal-membrane-coated NPs requires the extraction of the membrane through hypotonic treatment of exosomes. This treatment removes intravesicular components while leaving the surface membrane proteins intact. Surface membrane proteins play important roles in cell recognition, signaling, and communication [125].

Natural exosomes were collected from the cell-culture supernatant via ultracentrifugation (differential centrifugation). In line with MISEV guidelines (2015), this is the most commonly used and reliable method for isolating exosomes from cell-culture supernatants [22, 38]. Several techniques have been proposed. However, there are currently no standardized methods for exosome isolation. Once isolated, exosomes should be analyzed and characterized. According to MISEV guidelines, several techniques must be employed for the characterization of isolated exosomes. These include transmission electron microscopy to analyze surface morphology, NP-tracking analysis for size, and Western blotting for the detection of exosomal surface proteins [22]. The exosomal membranes are then extracted by resuspending the collected exosome pellets in a hypotonic lysis buffer containing a protease inhibitor cocktail. The lysate is then ultracentrifuged to remove intravesicular contents and isolate the exosomal membrane. Finally, the membrane-rich fraction is washed with isotonic buffers, such as phosphate-buffered saline, to collect purified exosomal membranes [190, 193].

Nanoparticle inner core selection and synthesis

The next step involves the selection and preparation of the NP inner core. Different nanomaterials have been used, ranging from organic cores to inorganic cores. Regardless of the NP composition and cell membrane types, it is essential to ensure that the nano-sized inner core has a negative zeta potential to facilitate electrostatic repulsion between the negatively charged NP surface and negatively charged membrane components [194]. This facilitates the correct orientation of the exosomal membrane around the NP core [195]. Cationic NP cores may hamper the coating process, as strong electrostatic interactions can lead to the unwanted bridging of the membrane structures and NP core materials, as described previously [196]. Other relevant parameters related to the NP core and NP core-membrane interfacial interactions that warrant further investigation are the NP’s core size, the surface curvature of the phospholipid bilayer, the impact on the sidedness of the membranes, and the completeness of the membrane coating.

Both organic and inorganic NP cores have been explored for the assembly of exosomal-membrane-coated NPs. The natural physicochemical properties of nanomaterial cores are related to their functionality. For instance, organic NPs are known for their biocompatibility, biodegradability, and high drug-loading capacity, and have mostly been employed in drug and gene delivery approaches. Liposomes [189] and poly(lactic-co-glycolic acid) (PLGA) NPs [197] have been used as NP cores and further coated with exosomal membranes. Poly(caprolactone) and human serum albumin NPs have also been used in this way [198]. In addition, inorganic-based nanomaterials such as mesoporous silica NPs [191], gold NPs (Au NPs) [199, 200], and iron oxide NPs [199] have been investigated for this purpose. Despite the bottlenecks associated with a lack of biodegradability, reduced biocompatibility, and toxicity when compared to organic-based nanoplatforms, inorganic NP cores have interesting properties for use in exosomal-membrane-coating approaches. The most commonly used cores are metallic NPs that exhibit intrinsic photothermal activity [191, 201]. Other strategies using metallic NP cores enable unique imaging features in exosomal-membrane-coated systems [199, 200], as well as magnetic properties for magnetic guidance-enhanced targeted migration [202]. The overall exosomal-membrane coating of inorganic NP cores is associated with a further increase in the biocompatibility of the nanosystem and introduces a more facile and tunable method for surface functionalization, as the traditional processes for ligand attachment onto the surface of inorganic NPs can be complex and system-specific.

Coating the nanoparticle core with an exosomal membrane

Extracted exosomal membranes can then be used to camouflage the NP core. This can be achieved using different coating methods similar to those used for camouflaging NPs with natural cell membranes [21]. So far, different strategies have been reported for assembling exosomal-membrane-coated NPs. These include physical extrusion through porous membranes, sonication, direct incubation of NPs with living cells, direct incubation of NPs with isolated cell-secreted exosomes, as well as microfluidic sonication-based techniques.

Co-extrusion/sonication

Co-extrusion through porous membranes followed by sonication is the most extensively used approach for assembling exosomal-membrane-coated NPs. Physical extrusion, also known as co-extrusion, was the first reported coating method, and is commonly used to prepare synthetic liposomes. In this method, the NP inner core and purified exosomal membrane are combined and co-extruded through porous membranes to produce exosomal-membrane-coated NPs [197, 199, 203]. The disruptive mechanical forces induced by physical extrusion can disrupt the exosomal membrane’s structure, enabling it to reassemble around the NP surface to form a core–shell nanostructure [197, 199, 203]. Another approach used to coat the NP core with the exosomal membrane is sonication. In this approach, both the NP and the purified exosomal membrane are exposed to similarly disruptive forces that are generated by ultrasonic energy, resulting in the spontaneous formation of a core–shell nanostructure [201, 204]. This approach has the advantage of losing less material when compared to physical extrusion [205, 206].

Direct incubation of NPs with cells or exosomes

Although physical extrusion and sonication are widely used to camouflage NPs with exosomal membranes, these approaches are labor-intensive and time-consuming. There is also the possibility of damaging the protein integrity of exosomal membranes using these techniques [18]. Because surface membrane proteins are critical for the biological functions of exosomes, damaging their integrity adversely affects the biological properties of these biomimetic nanosystems [207]. To prevent damage, non-disruptive coating techniques have been adopted to coat NPs with exosomal membranes. One of these approaches is based on the direct incubation of NPs with living cells to enable cells to secrete NP-containing exosomes. This strategy takes advantage of the exosome biogenesis pathway to encapsulate NPs in the exosomal membrane [200]. In another approach, exosomal-membrane-coated NPs are produced by direct incubation of NPs with pre-collected exosomes [208, 209].

Microfluidic sonication method

To overcome the limitations of physical extrusion and sonication, a microfluidic sonication-based coating technique was recently proposed for the design of core–shell PLGA NPs in a single continuous manner. This technique utilizes ultrasonication to coat PLGA NPs with several types of biological membranes, including lipid, exosomal, and cancer-cell membranes [44]. The membranes of the exosomes and cancer cells were isolated from A549 human lung carcinoma cells. They were used to coat PLGA NPs using a microfluidic sonication approach. The exosomal-membrane-coated PLGA NPs showed 1.0- and 5.5-fold-higher accumulation at A549 tumor sites when compared to cancer-cell-membrane- and lipid-membrane-coated NPs, respectively. These improved results were attributed to the homotypic targeting ability of exosomes and reduced immune uptake by monocytes/macrophages [44].

The microfluidic sonication approach was used in a subsequent study for coating PLGA NPs with MDA-MB-231 cell (an epithelial, human breast cancer cell line)-derived exosomal membranes that were functionalized with AS1411 aptamers [210]. Because of the exosomal membrane coating, the biomimetic nanosystem exhibited a systemic circulation duration that was 3.5-fold longer than that of AS1411-modified lipid-PLGA NPs. In addition, owing to the specific binding of AS1411 aptamers to nucleolin, a nucleolar protein that is overexpressed on the membrane of some cancer cells, the NPs demonstrated 1.59-fold-higher accumulation in tumors when compared to exosomal-membrane-coated NPs without AS1411 functionalization [210].

Biomedical applications in tissue engineering and neurodegenerative diseases

The following sections highlight some studies that employ exosomal-membrane-coated nanosystems for biomedical applications. Figure 7 summarizes the biomedical applications of exosomal-membrane-coated NPs for tissue engineering and regenerative medicine, as well as the diagnosis and treatment of neurodegenerative diseases.

Fig. 7
figure 7

Biomedical applications of exosomal-membrane-coated nanosystems. a Tissue engineering and regenerative medicine (e.g., skin regeneration and wound-healing applications). b Neurological disorders (e.g., neuroimaging and treatment of Alzheimer’s disease and Parkinson’s disease). CNS central nervous system, MI myocardial infarction

Tissue engineering and regenerative medicine

The purpose of tissue engineering and regenerative medicine is to generate viable human tissues and organs to replace diseased or damaged ones or to induce their regeneration in vivo [211, 212].

MSCs are multipotent cells that are promising for treating inflammatory diseases and cutaneous wounds owing to their multipotent differentiation and immunosuppressive and regenerative properties [18]. The therapeutic effects of MSCs on skin regeneration and wound-healing appear to be related to their ability to promote angiogenesis, enhance collagen synthesis and re-epithelialization, and accelerate skin regeneration and wound closure. Recently, MSC-derived exosomes have been investigated for skin regeneration and wound-healing as they can maintain the functional properties of their progenitor cells [202].

The wound-healing effects of MSC-derived exosomes were investigated in vivo by camouflaging superparamagnetic iron oxide NPs (Fe3O4 NPs) with MSC-derived exosomal membranes. This was achieved by the direct incubation of Fe3O4 NPs with MSCs. MSC-secreted exosomes contain exogenous NPs via the exosome biogenesis pathway (Fig. 8a) [202]. Because of the restricted ability of MSC-derived exosomes to target wounded skin sites, magnetic guidance was employed to efficiently deliver exosomal-membrane-coated Fe3O4 NPs to injured skin. Owing to their magnetic properties, the Fe3O4 cores enhanced the targeting ability of MSC-derived exosomes to the wounded skin sites of mice after intravenous (IV) injection. Treatment with exosomal-membrane-coated Fe3O4 NPs using magnetic guidance enhanced collagen synthesis and re-epithelialization, accelerated wound closure, and reduced scar formation, which resulted in the up-regulation of skin healing-associated proteins, such as cyclin A2, cyclin D1, VEGFA, and C-X-C motif chemokine 12. In summary, the pro-angiogenic effects of the exosome-mimicking nanosystem were 2-fold higher than those of non-coated NPs, leading to a significant reduction in the area of the injured skin after 3 and 5 weeks [202].

Fig. 8
figure 8

Exosome-based nanosystems for tissue engineering and regenerative medicine applications. a Preparation of MSC-derived exosomal-membrane-coated Fe3O4 NPs by coating the NP inner core with an exosomal membrane via the exosome biogenesis pathway (Reproduced with permission; copyright BioMed Central Ltd. (2020) [202]). b Schematic of the exosome spray method and fabrication process (Reproduced with permission; copyright American Chemical Society (2021) [214]). IV intravenous, MI myocardial infarction, MSC mesenchymal stem cell, NPs nanoparticles

Exosome-based therapy is a novel method for restoring bone defects without the use of cells. This therapeutic regime is based on cell–cell communication mediated by exosomes for the transfer of genetic materials and critical proteins. When compared to routine methods for bone defect restoration that require cell transplantation, cell-free exosome-based therapy is advantageous in reducing cell accumulation within the organ (e.g., the liver). Other benefits include an intrinsic homing effect, considerable chemical and physical stability, and low immunogenicity. Human adipose-derived stem cells (hASCs) undergo rapid osteogenic differentiation both in vitro and in vivo. The use of hASC-derived exosomes further accelerates angiogenesis and enables cation transfer and incorporation of cations into bone defects. In addition, hASC-derived exosomes have been shown to enhance the proliferation, migration, and osteogenic differentiation of MSCs both in vitro and in vivo. These properties render hASC-derived exosomes a suitable candidate and a promising alternative for future clinical trials [213].

As mentioned in the previous section, strategies involving cell transplantation have considerable drawbacks, such as their risks of tumorigenesis and immunogenicity. Therefore, an acellular approach has emerged based on the stem cell-derived secretome and its associated exosomes. Recently, scientists have investigated minimally invasive, sprayable cardiac patches based on MSC-derived exosomes [214]. These patches were used to create the product “exosome spray” (EXOS) by combining with a fibrin sealant to generate gelation properties (Fig. 8b). The fibrin scaffold was approved by the US Food and Drug Administration and was characterized using scanning electron microscopy. This invention can be used as an alternative to open surgery, which can result in severe physical trauma. EXOS increases the retention of MSC-derived exosomes in the heart even after MI, and increases the uptake of these exosomes by cardiomyocytes. This increased uptake leads to a reduction in cell apoptosis and an increase in the cell proliferation rate. In vivo experiments showed that EXOS reduced infarct size, improved cardiac function, preserved viable cardiac tissue cells, and increased ventricular wall thickness. Additional experiments have shown that EXOS is capable of improving angiomyogenesis after MI [214].

A recent study has shown that natural exosomes can mediate both the expression and transfer of genetic materials and vital proteins after their secretion from cells [215]. This is promising for different tissue engineering applications, as well as for promoting gene expression within targeted organs and tissues. The delivery of genetic materials and vital proteins to targeted organs and tissues may be enhanced by decorating the surface of exosomes with siRNA. The inclusion of siRNA on the exosomal surface protects the cargo, increases the targeted delivery ratio, and decreases off-target effects. This approach may be beneficial for myocardial regeneration. Scientists have found that genetically decorated exosomes derived from bone marrow stromal cells considerably improve tube formation from human umbilical vein endothelial cells. This strategy inhibits the proliferation of T cells in vitro and in vivo [216].

Neurodegenerative diseases

The most serious challenge in the diagnosis and treatment of neurodegenerative disorders is the difficulty of drug delivery systems in crossing the blood–brain barrier and targeting neuronal cells. Exosomal-membrane-coated NPs have been used to image and treat neurodegenerative disorders. For instance, Au NPs have been functionalized with neuron-targeting exosomes derived from genetically engineered human embryonic kidney cells (HEK293T). Coating Au NPs with exosomal membranes enhances their penetration through the blood–brain barrier and improves their accumulation in neuronal cells [217]. Exosomal membranes have been conjugated with neuron-targeting ligands such as the rabies virus glycoprotein (RVG) peptide. This combination improves the brain-targeting ability of exosomes because of the specific binding of the RVG peptide to acetylcholine receptors expressed by neuronal cells. To harness the potential benefits of such a strategy, exosome-producing HEK293T cells were transfected to produce exosomes with RVG peptides on their surfaces. The modified exosomes were then isolated from the cell-culture supernatant and used to coat Au NPs. The ability of Lamp2b-RVG and glycosylation-stabilized peptide-decorated Au NPs to penetrate the blood–brain barrier and specifically target brain cells was demonstrated both in vitro and in vivo after IV injection in a mouse model by bioluminescent imaging of the mouse brain. When compared to Au NPs coated with non-RVG-targeted exosomes, Au NPs coated with RVG-targeted exosomes were more efficacious in crossing the blood–brain barrier and accumulated more abundantly in brain cells. In vitro, a penetration rate of 20% across the blood–brain barrier was achieved 24 h after incubation, which was considerably higher than that of non-RVG-targeted exosome-coated Au NPs. This study reveals a promising approach to overcoming the challenge of crossing the blood–brain barrier, and pioneers the development of effective diagnostic and treatment strategies for various brain diseases [217].

AD is the most prevalent form of dementia worldwide. Globally, the number of people affected by this neurodegenerative disease is expected to increase considerably over the next few decades. AD is characterized by gradual memory loss and cognitive decline. Impairment of daily tasks occurs when patients lose their autonomy entirely [218]. The accumulation of amyloid beta peptides and hyperphosphorylated tau proteins in memory-associated areas of the brains of patients with AD results in the formation of amyloid plaques and neurofibrillary tangles, respectively. These aggregates are considered the two major histopathological hallmarks of the later stages of AD [219]. Cadmium (Cd) toxicity is associated with an increase in amyloid beta and phosphorylated tau protein levels, both of which are associated with AD. Furthermore, Cd exposure may contribute to AD due to degenerative brain alterations [220]. In a recent study investigating the potential of NPs and exosomes to ameliorate neurological disorders, copper sulfide NPs and MSC-derived exosomes were co-delivered to rats in a Cd-induced neurological disorder model [221]. Improved anticholinesterase, antioxidant, and anti-inflammatory responses were observed after IV injection of MSC-derived exosomes and copper sulfide NPs. Histological evaluations revealed that treatment with MSC-derived exosomes and copper sulfide NPs decreased the toxic effects of Cd on brain tissue and reduced degenerative alterations originating from neuronal disorders [221]. Future investigations should evaluate the applications of exosomal-membrane-coated NPs in the delivery of biomaterials, drugs, and genes.

PD is also a common neurodegenerative disease. It is characterized by the progressive loss of dopaminergic neurons, resulting in a dopamine deficit. In dopaminergic neurons, there is an abnormal accumulation of α-synuclein (α-syn). This protein is encoded by the synuclein alpha (SNCA) gene, which is the main component of Lewy bodies. Lewy body dementia is the most typical pathological manifestation of PD, and causes problems in thinking, movement, behavior, and mood [222]. Recently, a method was developed for reducing the expression and cytotoxicity of α-syn aggregates in dopaminergic neurons and delaying the progression of PD. This approach is based on the use of exosomal-membrane-coated NPs [223]. In this study, a biomimetic core–shell nanosystem was developed by co-loading phenylboronic acid-poly[2-(dimethylamino)ethyl acrylate] NPs with curcumin and siRNA targeting SNCA. The assembly core was subsequently coated with RVG-modified exosomal membranes derived from immature DCs. The biomimetic core–shell nanosystem effectively crossed the blood–brain barrier and targeted dopaminergic neurons. The loaded drugs were released into dopaminergic neurons in a reactive oxygen species-responsive manner to synergistically down-regulated α-syn synthesis and reduce existing α-syn aggregates. siRNA targeting SNCA inhibited α-syn aggregation by reducing α-syn synthesis, whereas curcumin directly reduced existing α-syn aggregates. Due to the synergistic effects of both drugs, the biomimetic nanosystem was more effective than its non-coated counterparts in clearing α-syn aggregates in dopaminergic neurons and in reducing SNCA mRNA expression (a 64% reduction was achieved). With the demonstration of improved neuronal repair and motor behavior in vivo, the biomimetic core–shell nanosystem has the potential for being used to effectively treat PD [223].

Clinical translation and regulation

Exosome-based systems should be included under the designations “investigative medicinal products” (Europe) and “investigative new drugs” (US) [103]. Regarding the characterization and quality control of exosome-based products, the ISEV offers useful guidelines and explanations regarding 1) nomenclature, 2) exosome collection and pre-treatment, 3) exosome separation and purification, 4) exosome characterization, and 5) recommendations on functional studies to be performed [22].

Increasing knowledge of exosome functionality and biological roles has provided pivotal opportunities for the application of exosome-mimicking nanosystems in tissue repair, wound-healing, and the management of neurodegenerative diseases, among other applications. Despite significant advances in the development of next-generation exosome-based therapies, significant challenges prevent the leverage of these therapies in clinical settings, including the need for extensive and robust characterization, issues concerning large-scale production and reproducibility of exosome-related biomaterials (including exosomal-membrane-based nanosystems), standardization of manufacturing protocols, and the necessity to better understand the biodistribution and targeting features of exosomes [224].

Despite the excitement of exosomal-membrane-coating nanotechnology as a novel field of research, these challenges pose a significant hurdle for human clinical applications. Studies reporting the therapeutic potential of exosomal-membrane-coated NPs in tissue repair, wound-healing, and neurodegenerative diseases have not yet been scaled up to human clinical trials, being restricted to in vitro and in vivo mouse models.

Challenges and future perspectives

Exosomal-membrane-coated NPs are emergent and promising nature-inspired delivery systems for biomedical applications. Although significant progress has been made in the field of exosomal-membrane-coating nanotechnology, this is a relatively new technological approach, and research in this area is still in its infancy. Enormous challenges currently hinder the implementation of exosomal-membrane-coated NPs in clinical settings, including 1) complex intra-exosomal composition, 2) heterogeneity, 3) reproducibility, 4) the lack of standardized methods for exosome isolation and purification, 5) the difficulty of large-scale manufacturing, 6) the lack of agreement over the ideal coating method, and 7) the high risk that the coating techniques may compromise the biological functions of natural exosomes and their safety profiles [18, 21]. Another critical issue faced by scientists is the current lack of understanding of the biogenesis, composition, and biological function of natural exosomes. To design exosomal-membrane-coated NPs more efficiently and safely, future research should focus on clarifying the complex composition, biological functionalities, and intrinsic targeting abilities of natural exosomes [18].

One major challenge when using natural, cell-secreted exosomes is exosome isolation and purity [22]. Various methods have been proposed for exosome isolation, including differential ultracentrifugation, density gradient ultracentrifugation, size-exclusion chromatography, and affinity/immunoaffinity capture [225]. All of these approaches have their own advantages and drawbacks, and thus far, there has been no standardization of the best isolation technique. Exosomes can be isolated from cell-culture supernatants or biological fluids such as plasma and serum. Each source has specific features that must be considered when isolating exosomes [22]. If exosomes are isolated from cells, one aspect to consider is the risk of isolation. Apart from cell-secreted exosomes, contaminant vesicles derived from fetal bovine serum (FBS) are often added to cell cultures. Precautions must be taken when using exosome-free FBS or bovine serum albumin instead of FBS [225], as exosomes isolated from plasma or serum are notorious for being contaminated with non-EV proteins (albumin and globulins) and non-EV lipidic structures (chylomicrons and lipoproteins), which can form non-EV particles [22]. Plasma is recommended over serum owing to the platelet EVs that are released during coagulation [226, 227]. Co-isolation of non-EV contaminants represents a major challenge for proper exosome isolation and analysis. Detailed information is needed regarding the isolation samples and their handling, namely storage and analytical procedures [22].

Different techniques have been investigated for coating NPs with exosomal membranes, with sonication and physical extrusion through porous membranes being the two most frequently used techniques. Another challenge for the clinical implementation of exosomal-membrane-coated NPs is related to the potential of coating methods to damage the integrity of the exosomal membrane’s structure and reduce its protein integrity. This may compromise the biological functions of natural exosomes and induce immunogenicity [18]. Exosomes contain a diverse set of proteins, some of which are responsible for their biological functions, whereas others may induce immune responses. Hence, manipulation of the exosomal membrane may modify the surface composition and orientation of these proteins. Such undue modifications may trigger immune responses and induce immunogenicity [18, 19]. There is an urgent need to develop new, non-disruptive coating techniques that do not adversely affect the protein integrity of the exosomal membrane or the efficacy and safety of biomimetic nanoplatforms [18].

Another major challenge is the lack of standardization regarding the best method for coating NPs with exosomal membranes [21]. It is generally accepted that the ideal coating method depends on the NP and cell types. Accordingly, studies should be performed using different types of NPs, progenitor cells, and coating methods to evaluate which encapsulation method is most favorable for a particular scenario [21].

The reduced number of exosomes naturally secreted by most cells and the current lack of standardized protocols for exosome isolation and purification represent major challenges for the successful implementation of natural exosomes in clinical settings and exosome production at a clinical scale [21]. Similar to natural exosomes, the clinical-scale production of membrane-coated NPs remains a significant obstacle. To circumvent the large-scale manufacturing challenges of these biomimetic NPs, approaches normally used to produce exosomes on a large scale, such as the generation of cell-derived nanovesicles using extrusion through porous membranes, have recently been employed [18]. In a recent effort to prepare exosome-mimetic nanovesicles to encapsulate NPs, magnetic MSC-derived nanovesicles have been used to camouflage iron oxide NPs for the treatment of ischemic strokes. Iron oxide NPs were encapsulated in MSC-derived nanovesicles by extruding MSCs treated with iron oxide NPs through porous membranes. The final exosome-mimetic nanovesicles exhibited a 5.1-fold higher accumulation at sites of ischemic brain injury in a mouse model after IV injection and magnetic guidance when compared to those administered without an external magnetic field. The nanovesicles were capable of inducing angiogenesis, demonstrating anti-apoptotic and anti-inflammatory characteristics, substantially reducing infarct volume, and enhancing motor function [228].

Another concern is the safety profile of exosomal-membrane-coated NPs. Because these NPs contain biological materials, their quality control is of high importance. Therefore, stringent investigation of the immunogenicity profiles and potential side effects of exosomal-membrane-coated NPs should be determined prior to their translation into clinical settings [18, 19, 21]. In the future, to reduce potential undesirable immune responses and ensure the biosafety of these biomimetic nanoplatforms, the development of personalized therapy that utilizes the patient’s own exosomes to camouflage NPs should be investigated [21].

Conclusions

The use of exosomal membranes to camouflage nanomaterials for biomedical applications is an attractive and promising technological approach because of their enhanced biocompatibility, non-immunogenicity, immune evasion abilities, prolonged blood circulation, intrinsic tissue-specific homing features, and cell-specific uptake [21]. Despite the enormous potential of exosomal-membrane-coated NPs for the targeted delivery of therapeutic and imaging molecules to sites of interest, this is a relatively new technological approach. Major challenges must be addressed before clinical translation can come to fruition. In recent years, research on this biomimetic approach is expected to continue to grow, which will enable the development of promising next-generation bioinspired nanosystems for a variety of biomedical applications with the potential to revolutionize the diagnosis and treatment of human diseases.

Availability of data and materials

Not applicable.

Abbreviations

AD:

Alzheimer’s disease

Cd:

Cadmium

Cx43:

Connexin 43

DC:

Dendritic cell

DNA:

Deoxyribonucleic acid

ESCRT:

Endosomal sorting complex required for transport

EV:

Extracellular vesicle

EXOS:

“Exosome spray”

FBS:

Fetal bovine serum

hASC:

Human adipose-derived stem cell

Hsps:

Heat shock proteins

ISEV:

International Society for Extracellular Vesicles

IV:

Intravenous

Lamp2b:

Lysosome-associated membrane glycoprotein 2b

MI:

Myocardial infarction

miRNA:

MicroRNA

MISEV:

Minimal Information for Studies of Extracellular Vesicles

mRNA:

Messenger RNA

MSC:

Mesenchymal stem cell

MVB:

Multivesicular body

NP:

Nanoparticle

PD:

Parkinson’s disease

PLGA:

Poly(lactic-co-glycolic acid)

RNA:

Ribonucleic acid

RVG:

Rabies virus glycoprotein

siRNA:

Small interfering RNA

VEGF:

Vascular endothelial growth factor

α-syn:

α-Synuclein

References

  1. Wang Z, Li R, Zhang J. On-demand drug delivery of triptolide and celastrol by poly (lactic-co-glycolic acid) nanoparticle/triglycerol monostearate-18 hydrogel composite for rheumatoid arthritis treatment. Adv Compos Hybrid Mater. 2022;5:2921–35.

    Article  CAS  Google Scholar 

  2. Ran F, Li C, Hao Z, Zhang X, Dai L, Si C, et al. Combined bactericidal process of lignin and silver in a hybrid nanoparticle on E. coli. Adv Compos Hybrid Mater. 2022;5:1841–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Ali Baig AB, Rathinam V, Ramya V. Facile fabrication of Zn-doped SnO2 nanoparticles for enhanced photocatalytic dye degradation performance under visible light exposure. Adv Compos Hybrid Mater. 2021;4:114–26.

    Article  CAS  Google Scholar 

  4. Shao M, Lopes D, Lopes J, Yousefiasl S, Macário-Soares A, Peixoto D, et al. Exosome membrane-coated nanosystems: exploring biomedical applications in cancer diagnosis and therapy. Matter. 2023;6(3):761–99.

    Article  CAS  Google Scholar 

  5. Islamipour Z, Zare EN, Salimi F, Ghomi M, Makvandi P. Biodegradable antibacterial and antioxidant nanocomposite films based on dextrin for bioactive food packaging. J Nanostruct Chem. 2022;12:991–1006.

    Article  CAS  Google Scholar 

  6. Movagharnezhad N, Ehsanimehr SS, Najafi MP. Synthesis of poly (N-vinylpyrrolidone)-grafted-magnetite bromoacetylated cellulose via ATRP for drug delivery. Mater Chem Horiz. 2022;1(2):89–98.

    Google Scholar 

  7. Heidari G, Hassanpour M, Nejaddehbashi F, Sarfjoo MR, Yousefiasl S, Sharifi E, et al. Biosynthesized nanomaterials with antioxidant and antimicrobial properties. Mater Chem Horiz. 2022;1(1):35–48.

    Google Scholar 

  8. Rabiee N, Bagherzadeh M, Ghadiri AM, Kiani M, Ahmadi S, Jajarmi V, et al. Calcium-based nanomaterials and their interrelation with chitosan: optimization for pCRISPR delivery. J Nanostruct Chem. 2022;12(5):919–32.

    Article  CAS  Google Scholar 

  9. Panthi G, Ranjit R, Khadka S, Gyawali KR, Kim HY, Park M. Characterization and antibacterial activity of rice grain-shaped ZnS nanoparticles immobilized inside the polymer electrospun nanofibers. Adv Compos Hybrid Mater. 2020;3(1):8–15.

    Article  CAS  Google Scholar 

  10. Banskota S, Yousefpour P, Chilkoti A. Cell-based biohybrid drug delivery systems: the best of the synthetic and natural worlds. Macromol Biosci. 2017. https://doi.org/10.1002/mabi.201600361.

    Article  PubMed  Google Scholar 

  11. Xia Q, Zhang Y, Li Z, Hou X, Feng N. Red blood cell membrane-camouflaged nanoparticles: a novel drug delivery system for antitumor application. Acta Pharm Sin B. 2019;9(4):675–89.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Lopes J, Lopes D, Pereira-Silva M, Peixoto D, Veiga F, Hamblin MR, et al. Macrophage cell membrane-cloaked nanoplatforms for biomedical applications. Small Methods. 2022;6(8):e2200289.

    Article  PubMed  Google Scholar 

  13. Han H, Bártolo R, Li J, Shahbazi MA, Santos HA. Biomimetic platelet membrane-coated nanoparticles for targeted therapy. Eur J Pharm Biopharm. 2022;172:1–15.

    Article  PubMed  Google Scholar 

  14. Ferreira-Faria I, Yousefiasl S, Macario-Soares A, Pereira-Silva M, Peixoto D, Zafar H, et al. Stem cell membrane-coated abiotic nanomaterials for biomedical applications. J Control Release. 2022;351:174–97.

    Article  CAS  PubMed  Google Scholar 

  15. Pereira-Silva M, Santos AC, Conde J, Hoskins C, Concheiro A, Alvarez-Lorenzo C, et al. Biomimetic cancer cell membrane-coated nanosystems as next-generation cancer therapies. Expert Opin Drug Deliv. 2020;17(11):1515–8.

    Article  PubMed  Google Scholar 

  16. Chugh V, Vijaya Krishna K, Pandit A. Cell membrane-coated mimics: a methodological approach for fabrication, characterization for therapeutic applications, and challenges for clinical translation. ACS Nano. 2021;15(11):17080–123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ravasco JMJM, Paiva-Santos AC, Conde J. Technological challenges of biomembrane-coated top-down cancer nanotherapy. Nat Rev Bioeng. 2023;1(3):156–8.

    Article  Google Scholar 

  18. Lu M, Huang Y. Bioinspired exosome-like therapeutics and delivery nanoplatforms. Biomaterials. 2020;242:119925.

    Article  CAS  PubMed  Google Scholar 

  19. Zhang X, Zhang H, Gu J, Zhang J, Shi H, Qian H, et al. Engineered extracellular vesicles for cancer therapy. Adv Mater. 2021;33(14):e2005709.

    Article  PubMed  Google Scholar 

  20. Li SP, Lin ZX, Jiang XY, Yu XY. Exosomal cargo-loading and synthetic exosome-mimics as potential therapeutic tools. Acta Pharmacol Sin. 2018;39(4):542–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Fathi P, Rao L, Chen X. Extracellular vesicle-coated nanoparticles. View. 2020;2(2):20200187.

    Article  Google Scholar 

  22. Thery C, Witwer KW, Aikawa E, Alcaraz MJ, Anderson JD, Andriantsitohaina R, et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): a position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J Extracell Vesicles. 2018;7(1):1535750.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Lotvall J, Hill AF, Hochberg F, Buzas EI, Di Vizio D, Gardiner C, et al. Minimal experimental requirements for definition of extracellular vesicles and their functions: a position statement from the International Society for extracellular vesicles. J Extracell Vesicles. 2014;3:26913.

    Article  PubMed  Google Scholar 

  24. Chen Z, Yang L, Cui Y, Zhou Y, Yin X, Guo J, et al. Cytoskeleton-centric protein transportation by exosomes transforms tumor-favorable macrophages. Oncotarget. 2016;7(41):67387–402.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Thakur A, Parra DC, Motallebnejad P, Brocchi M, Chen HJ. Exosomes: small vesicles with big roles in cancer, vaccine development, and therapeutics. Bioact Mater. 2021;10:281–94.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Kooijmans SAA, De Jong OG, Schiffelers RM. Exploring interactions between extracellular vesicles and cells for innovative drug delivery system design. Adv Drug Deliv Rev. 2021;173:252–78.

    Article  CAS  PubMed  Google Scholar 

  27. Naryzhny S, Volnitskiy A, Kopylov A, Zorina E, Kamyshinsky R, Bairamukov V, et al. Proteome of glioblastoma-derived exosomes as a source of biomarkers. Biomedicines. 2020;8(7):216.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Skotland T, Hessvik NP, Sandvig K, Llorente A. Exosomal lipid composition and the role of ether lipids and phosphoinositides in exosome biology. J Lipid Res. 2019;60(1):9–18.

    Article  CAS  PubMed  Google Scholar 

  29. Jeppesen DK, Fenix AM, Franklin JL, Higginbotham JN, Zhang Q, Zimmerman LJ, et al. Reassessment of exosome composition. Cell. 2019;177(2):428-45.e18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Zhang Y, Liu Y, Liu H, Tang WH. Exosomes: biogenesis, biologic function and clinical potential. Cell Biosci. 2019;9:19.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Choi DS, Kim DK, Kim YK, Gho YS. Proteomics of extracellular vesicles: exosomes and ectosomes. Mass Spectrom Rev. 2015;34(4):474–90.

    Article  CAS  PubMed  Google Scholar 

  32. Phuyal S, Hessvik NP, Skotland T, Sandvig K, Llorente A. Regulation of exosome release by glycosphingolipids and flotillins. FEBS J. 2014;281(9):2214–27.

    Article  CAS  PubMed  Google Scholar 

  33. Garcia NA, Ontoria-Oviedo I, Gonzalez-King H, Diez-Juan A, Sepulveda P. Glucose starvation in cardiomyocytes enhances exosome secretion and promotes angiogenesis in endothelial cells. PLoS One. 2015;10(9):e0138849.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Lauwers E, Wang YC, Gallardo R, Van Der Kant R, Michiels E, Swerts J, et al. Hsp90 mediates membrane deformation and exosome release. Mol Cell. 2018;71(5):689-702.e9.

    Article  CAS  PubMed  Google Scholar 

  35. Graziano F, Iacopino DG, Cammarata G, Scalia G, Campanella C, Giannone AG, et al. The triad Hsp60-miRNAs-extracellular vesicles in brain tumors: assessing its components for understanding tumorigenesis and monitoring patients. Appl Sci. 2021;11(6):2867.

    Article  CAS  Google Scholar 

  36. Larios J, Mercier V, Roux A, Gruenberg J. ALIX- and ESCRT-III–dependent sorting of tetraspanins to exosomes. J Cell Biol. 2020;219(3):e201904113.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Ferreira JV, da Rosa SA, Ramalho J, Máximo Carvalho C, Cardoso MH, Pintado P, et al. LAMP2A regulates the loading of proteins into exosomes. Sci Adv. 2022;8(12):eam1140.

    Article  Google Scholar 

  38. Yang B, Chen Y, Shi J. Exosome biochemistry and advanced nanotechnology for next-generation theranostic platforms. Adv Mater. 2019;31(2):e1802896.

    Article  PubMed  Google Scholar 

  39. Liang Y. Engineering exosomes for targeted drug delivery. Theranostics. 2021;11(7):3183–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Rana S, Yue S, Stadel D, Zöller M. Toward tailored exosomes: the exosomal tetraspanin web contributes to target cell selection. Int J Biochem Cell Biol. 2012;44(9):1574–84.

    Article  CAS  PubMed  Google Scholar 

  41. Clayton A, Harris CL, Court J, Mason MD, Morgan BP. Antigen-presenting cell exosomes are protected from complement-mediated lysis by expression of CD55 and CD59. Eur J Immunol. 2003;33(2):522–31.

    Article  CAS  PubMed  Google Scholar 

  42. Ke W, Afonin KA. Exosomes as natural delivery carriers for programmable therapeutic nucleic acid nanoparticles (NANPs). Adv Drug Deliv Rev. 2021;176:113835.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Datta B, Paul D, Pal U, Rakshit T. Intriguing biomedical applications of synthetic and natural cell-derived vesicles: a comparative overview. ACS Appl Bio Mater. 2021;4(4):2863–85.

    Article  CAS  PubMed  Google Scholar 

  44. Liu C, Zhang W, Li Y, Chang J, Tian F, Zhao F, et al. Microfluidic sonication to assemble exosome membrane-coated nanoparticles for immune evasion-mediated targeting. Nano Lett. 2019;19(11):7836–44.

    Article  CAS  PubMed  Google Scholar 

  45. Gurung S, Perocheau D, Touramanidou L, Baruteau J. The exosome journey: from biogenesis to uptake and intracellular signalling. Cell Commun Signal. 2021;19(1):47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Record M, Carayon K, Poirot M, Silvente-Poirot S. Exosomes as new vesicular lipid transporters involved in cell–cell communication and various pathophysiologies. Biochim Biophys Acta. 2014;1841(1):108–20.

    Article  CAS  PubMed  Google Scholar 

  47. Gangadaran P, Ahn BC. Extracellular vesicle- and extracellular vesicle mimetics-based drug delivery systems: new perspectives, challenges, and clinical developments. Pharmaceutics. 2020;12(5):442.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Valadi H, Ekström K, Bossios A, Sjöstrand M, Lee JJ, Lötvall JO. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol. 2007;9(6):654–9.

    Article  CAS  PubMed  Google Scholar 

  49. Hurwitz SN, Cheerathodi MR, Nkosi D, York SB, Meckes DG. Tetraspanin CD63 bridges autophagic and endosomal processes to regulate exosomal secretion and intracellular signaling of epstein-barr virus LMP1. J Virol. 2018;92(5):e01969-e2017.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Ostrowski M, Carmo NB, Krumeich S, Fanget I, Raposo G, Savina A, et al. Rab27a and Rab27b control different steps of the exosome secretion pathway. Nat Cell Biol. 2010;12(1):19–30.

    Article  CAS  PubMed  Google Scholar 

  51. Shimizu A, Sawada K, Kobayashi M, Yamamoto M, Yagi T, Kinose Y, et al. Exosomal CD47 plays an essential role in immune evasion in ovarian cancer. Mol Cancer Res. 2021;19(9):1583–95.

    Article  CAS  PubMed  Google Scholar 

  52. Gauvreau ME, Côté MH, Bourgeois-Daigneault MC, Rivard LD, Xiu F, Brunet A, et al. Sorting of MHC class II molecules into exosomes through a ubiquitin-independent pathway. Traffic. 2009;10(10):1518–27.

    Article  CAS  PubMed  Google Scholar 

  53. Trajkovic K, Hsu C, Chiantia S, Rajendran L, Wenzel D, Wieland F, et al. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science. 2008;319(5867):1244–7.

    Article  CAS  PubMed  Google Scholar 

  54. Subra C, Laulagnier K, Perret B, Record M. Exosome lipidomics unravels lipid sorting at the level of multivesicular bodies. Biochimie. 2007;89(2):205–12.

    Article  CAS  PubMed  Google Scholar 

  55. Menck K, Sönmezer C, Worst TS, Schulz M, Dihazi GH, Streit F, et al. Neutral sphingomyelinases control extracellular vesicles budding from the plasma membrane. J Extracell Vesicles. 2017;6(1):1378056.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Buschow SI, Van Balkom BW, Aalberts M, Heck AJ, Wauben M, Stoorvogel W. MHC class II-associated proteins in B-cell exosomes and potential functional implications for exosome biogenesis. Immunol Cell Biol. 2010;88(8):851–6.

    Article  CAS  PubMed  Google Scholar 

  57. Verweij FJ, Bebelman MP, Jimenez CR, Garcia-Vallejo JJ, Janssen H, Neefjes J, et al. Quantifying exosome secretion from single cells reveals a modulatory role for GPCR signaling. J Cell Biol. 2017;217(3):1129–42.

    Article  Google Scholar 

  58. Chen YD, Fang YT, Cheng YL, Lin CF, Hsu LJ, Wang SY, et al. Exophagy of annexin A2 via RAB11, RAB8A and RAB27A in IFN-gamma-stimulated lung epithelial cells. Sci Rep. 2017;7(1):5676.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Zheng Y, Campbell EC, Lucocq J, Riches A, Powis SJ. Monitoring the Rab27 associated exosome pathway using nanoparticle tracking analysis. Exp Cell Res. 2013;319(12):1706–13.

    Article  CAS  PubMed  Google Scholar 

  60. Colombo M, Moita C, Van Niel G, Kowal J, Vigneron J, Benaroch P, et al. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J Cell Sci. 2013;126(Pt 24):5553–65.

    CAS  PubMed  Google Scholar 

  61. Stuffers S, Sem Wegner C, Stenmark H, Brech A. Multivesicular endosome biogenesis in the absence of ESCRTs. Traffic. 2009;10(7):925–37.

    Article  CAS  PubMed  Google Scholar 

  62. Schorey JS, Cheng Y, Singh PP, Smith VL. Exosomes and other extracellular vesicles in host-pathogen interactions. EMBO Rep. 2015;16(1):24–43.

    Article  CAS  PubMed  Google Scholar 

  63. Vanniel G, Charrin S, Simoes S, Romao M, Rochin L, Saftig P, et al. The tetraspanin CD63 regulates ESCRT-independent and -dependent endosomal sorting during melanogenesis. Dev Cell. 2011;21(4):708–21.

    Article  CAS  Google Scholar 

  64. Horibe S, Tanahashi T, Kawauchi S, Murakami Y, Rikitake Y. Mechanism of recipient cell-dependent differences in exosome uptake. BMC Cancer. 2018;18(1):47.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Svensson KJ, Christianson HC, Wittrup A, Bourseau-Guilmain E, Lindqvist E, Svensson LM, et al. Exosome uptake depends on ERK1/2-heat shock protein 27 signaling and lipid Raft-mediated endocytosis negatively regulated by caveolin-1. J Biol Chem. 2013;288(24):17713–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Tan S, Wu T, Zhang D, Zhang Z. Cell or cell membrane-based drug delivery systems. Theranostics. 2015;5(8):863–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Lu J, Wu J, Tian J, Wang S. Role of T cell-derived exosomes in immunoregulation. Immunol Res. 2018;66(3):313–22.

    Article  CAS  PubMed  Google Scholar 

  68. Cai Z, Yang F, Yu L, Yu Z, Jiang L, Wang Q, et al. Activated T cell exosomes promote tumor invasion via fas signaling pathway. J Immunol. 2012;188(12):5954–61.

    Article  CAS  PubMed  Google Scholar 

  69. Zhang H, Xie Y, Li W, Chibbar R, Xiong S, Xiang J. CD4+ T cell-released exosomes inhibit CD8+ cytotoxic T-lymphocyte responses and antitumor immunity. Cell Mol Immunol. 2011;8(1):23–30.

    Article  PubMed  Google Scholar 

  70. Fu W, Lei C, Liu S, Cui Y, Wang C, Qian K, et al. CAR exosomes derived from effector CAR-T cells have potent antitumour effects and low toxicity. Nat Commun. 2019;10(1):4355.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Zhao C, Song W, Ma J, Wang N. Macrophage-derived hybrid exosome-mimic nanovesicles loaded with black phosphorus for multimodal rheumatoid arthritis therapy. Biomater Sci. 2022;10(23):6731–9.

    Article  CAS  PubMed  Google Scholar 

  72. Baek S, Jeon M, Jung HN, Lee W, Hwang JE, Lee JS, et al. M1 macrophage-derived exosome-mimetic nanovesicles with an enhanced cancer targeting ability. ACS Appl Bio Mater. 2022;5(6):2862–9.

    Article  CAS  PubMed  Google Scholar 

  73. Zhao Y, Zheng Y, Zhu Y, Zhang Y, Zhu H, Liu T. M1 Macrophage-derived exosomes loaded with gemcitabine and deferasirox against chemoresistant pancreatic cancer. Pharmaceutics. 2021;13(9):1493.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Hazrati A, Soudi S, Malekpour K, Mahmoudi M, Rahimi A, Hashemi SM, et al. Immune cells-derived exosomes function as a double-edged sword: role in disease progression and their therapeutic applications. Biomarker Res. 2022;10(1):30.

    Article  Google Scholar 

  75. Ceccarelli S, Pontecorvi P, Anastasiadou E, Napoli C, Marchese C. Immunomodulatory effect of adipose-derived stem cells: the cutting edge of clinical application. Front Cell Dev Biol. 2020;8:236.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Sadeghian-Nodoushan F, Nikukar H, Soleimani M, Jalali-Jahromi A, Hosseinzadeh S, Khojasteh A. A smart magnetic hydrogel containing exosome promotes osteogenic commitment of human adipose-derived mesenchymal stem cells. Iran J Basic Med Sci. 2022;25(9):1123–31.

    PubMed  PubMed Central  Google Scholar 

  77. Zhang Y, Li Y, Wang Q, Zheng D, Feng X, Zhao W, et al. Attenuation of hepatic ischemia-reperfusion injury by adipose stem cell-derived exosome treatment via ERK1/2 and GSK-3β signaling pathways. Int J Mol Med. 2022;49(2):13.

    Article  CAS  PubMed  Google Scholar 

  78. Wang J, Li M, Jin L, Guo P, Zhang Z, Zhanghuang C, et al. Exosome mimetics derived from bone marrow mesenchymal stem cells deliver doxorubicin to osteosarcoma in vitro and in vivo. Drug Deliv. 2022;29(1):3291–303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Xie X, Yang X, Wu J, Tang S, Yang L, Fei X, et al. Exosome from indoleamine 2,3-dioxygenase-overexpressing bone marrow mesenchymal stem cells accelerates repair process of ischemia/reperfusion-induced acute kidney injury by regulating macrophages polarization. Stem Cell Res Ther. 2022;13(1):367.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Won Lee G, Thangavelu M, Joung Choi M, Yeong Shin E, Sol Kim H, Seon Baek J, et al. Exosome mediated transfer of miRNA-140 promotes enhanced chondrogenic differentiation of bone marrow stem cells for enhanced cartilage repair and regeneration. J Cell Biochem. 2020;121(7):3642–52.

    Article  PubMed  Google Scholar 

  81. Dong J, Li L, Fang X, Zang M. Exosome-encapsulated microRNA-127-3p released from bone marrow-derived mesenchymal stem cells alleviates osteoarthritis through regulating CDH11-mediated Wnt/β-catenin pathway. J Pain Res. 2021;14:297–310.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Chang YH, Wu KC, Ding DC. Chondrogenic potential of human umbilical cord mesenchymal stem cells cultured with exosome-depleted fetal bovine serum in an osteoarthritis mouse model. Biomedicines. 2022;10(11):2773.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Wang H, Liu S, Zhang W, Liu M, Deng C. Human umbilical cord mesenchymal stem cell-derived exosome repairs endometrial epithelial cells injury induced by hypoxia via regulating miR-663a/CDKN2A axis. Oxid Med Cell Longev. 2022;2022:3082969.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Dehghani L, Khojasteh A, Soleimani M, Oraee-Yazdani S, Keshel SH, Saadatnia M, et al. Safety of intraparenchymal injection of allogenic placenta mesenchymal stem cells derived exosome in patients undergoing decompressive craniectomy following malignant middle cerebral artery infarct, a pilot randomized clinical trial. Int J Prev Med. 2022;13:7.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Ma J, Yong L, Lei P, Li H, Fang Y, Wang L, et al. Advances in microRNA from adipose-derived mesenchymal stem cell-derived exosome: focusing on wound healing. J Mater Chem B. 2022;10(46):9565–77.

    Article  CAS  PubMed  Google Scholar 

  86. Lee BC, Kang I, Yu KR. Therapeutic features and updated clinical trials of mesenchymal stem cell (MSC)-derived exosomes. J Clin Med. 2021;10(4):711.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Miyoshi E, Bilousova T, Melnik M, Fakhrutdinov D, Poon WW, Vinters HV, et al. Exosomal tau with seeding activity is released from Alzheimer’s disease synapses, and seeding potential is associated with amyloid beta. Lab Invest. 2021;101(12):1605–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Yang J, Yu X, Xue F, Li Y, Liu W, Zhang S. Exosomes derived from cardiomyocytes promote cardiac fibrosis via myocyte-fibroblast cross-talk. Am J Transl Res. 2018;10(12):4350–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Paskeh MDA, Entezari M, Mirzaei S, Zabolian A, Saleki H, Naghdi MJ, et al. Emerging role of exosomes in cancer progression and tumor microenvironment remodeling. J Hematol Oncol. 2022;15(1):83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Kim H, Kim EH, Kwak G, Chi SG, Kim SH, Yang Y. Exosomes: cell-derived nanoplatforms for the delivery of cancer therapeutics. Int J Mol Sci. 2020;22(1):14.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Pritchard A, Tousif S, Wang Y, Hough K, Khan S, Strenkowski J, et al. Lung tumor cell-derived exosomes promote M2 macrophage polarization. Cells. 2020;9(5):1303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Du S, Qian J, Tan S, Li W, Liu P, Zhao J, et al. Tumor cell-derived exosomes deliver TIE2 protein to macrophages to promote angiogenesis in cervical cancer. Cancer Lett. 2022;529:168–79.

    Article  CAS  PubMed  Google Scholar 

  93. Kim SM, Yang Y, Oh SJ, Hong Y, Seo M, Jang M. Cancer-derived exosomes as a delivery platform of CRISPR/Cas9 confer cancer cell tropism-dependent targeting. J Control Release. 2017;266:8–16.

    Article  CAS  PubMed  Google Scholar 

  94. Chow A, Zhou W, Liu L, Fong MY, Champer J, Van Haute D, et al. Macrophage immunomodulation by breast cancer-derived exosomes requires Toll-like receptor 2-mediated activation of NF-κB. Sci Rep. 2014;4(1):5750.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Qin W, Wang L, Tian H, Wu X, Xiao C, Pan Y, et al. CAF-derived exosomes transmitted Gremlin-1 promotes cancer progression and decreases the sensitivity of hepatoma cells to sorafenib. Mol Carcinog. 2022;61(8):764–75.

    Article  CAS  PubMed  Google Scholar 

  96. Li X, Li X, Zhang B, He B. The role of cancer stem cell-derived exosomes in cancer progression. Stem Cells Int. 2022;2022:9133658.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Ruivo CF, Bastos N, Adem B, Batista I, Duraes C, Melo CA, et al. Extracellular vesicles from pancreatic cancer stem cells lead an intratumor communication network (EVNet) to fuel tumour progression. Gut. 2022;71(10):2043.

    Article  CAS  PubMed  Google Scholar 

  98. Zhou S, Lan Y, Li Y, Li Z, Pu J, Wei L. Hypoxic tumor-derived exosomes induce M2 macrophage polarization via PKM2/AMPK to promote lung cancer progression. Cell Transpl. 2022;31:09636897221106998.

    Article  Google Scholar 

  99. Zhang B, Yeo RWY, Lai RC, Sim EWK, Chin KC, Lim SK. Mesenchymal stromal cell exosome–enhanced regulatory T-cell production through an antigen-presenting cell–mediated pathway. Cytotherapy. 2018;20(5):687–96.

    Article  CAS  PubMed  Google Scholar 

  100. Larssen P, Veerman RE, Akpinar GG, Hiltbrunner S, Karlsson MCI, Gabrielsson S. Allogenicity boosts extracellular vesicle-induced antigen-specific immunity and mediates tumor protection and long-term memory in vivo. J Immunol. 2019;203(4):825–34.

    Article  CAS  PubMed  Google Scholar 

  101. Matsuzaka Y, Yashiro R. Regulation of extracellular vesicle-mediated immune responses against antigen-specific presentation. Vaccines. 2022;10(10):1691.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Admyre C, Bohle B, Johansson SM, Focke-Tejkl M, Valenta R, Scheynius A, et al. B cell-derived exosomes can present allergen peptides and activate allergen-specific T cells to proliferate and produce TH2-like cytokines. J Allergy Clin Immunol. 2007;120(6):1418–24.

    Article  CAS  PubMed  Google Scholar 

  103. Asadi K, Amini A, Gholami A. Mesenchymal stem cell-derived exosomes as a bioinspired nanoscale tool toward next-generation cell-free treatment. J Drug Deliv Sci Technol. 2022;77:103856.

    Article  CAS  Google Scholar 

  104. Rani S, Ritter T. The exosome: a naturally secreted nanoparticle and its application to wound healing. Adv Mater. 2016;28(27):5542–52.

    Article  CAS  PubMed  Google Scholar 

  105. Vonk LA, Van Dooremalen SF, Liv N, Klumperman J, Coffer PJ, Saris DB, et al. Mesenchymal stromal/stem cell-derived extracellular vesicles promote human cartilage regeneration in vitro. Theranostics. 2018;8(4):906.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Chen P, Zheng L, Wang Y, Tao M, Xie Z, Xia C, et al. Desktop-stereolithography 3D printing of a radially oriented extracellular matrix/mesenchymal stem cell exosome bioink for osteochondral defect regeneration. Theranostics. 2019;9(9):2439–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Haraszti RA, Miller R, Stoppato M, Sere YY, Coles A, Didiot MC, et al. Exosomes produced from 3D cultures of MSCs by tangential flow filtration show higher yield and improved activity. Mol Ther. 2018;26(12):2838–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Fierabracci A, Del Fattore A, Luciano R, Muraca M, Teti A, Muraca M. Recent advances in mesenchymal stem cell immunomodulation: the role of microvesicles. Cell Transpl. 2015;24(2):133–49.

    Article  Google Scholar 

  109. Harrell CR, Jovicic N, Djonov V, Arsenijevic N, Volarevic V. Mesenchymal stem cell-derived exosomes and other extracellular vesicles as new remedies in the therapy of inflammatory diseases. Cells. 2019;8(12):1605.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Shi Y, Wang Y, Li Q, Liu K, Hou J, Shao C, et al. Immunoregulatory mechanisms of mesenchymal stem and stromal cells in inflammatory diseases. Nat Rev Nephrol. 2018;14(8):493–507.

    Article  CAS  PubMed  Google Scholar 

  111. Lee M, Ban JJ, Yang S, Im W, Kim M. The exosome of adipose-derived stem cells reduces β-amyloid pathology and apoptosis of neuronal cells derived from the transgenic mouse model of Alzheimer’s disease. Brain Res. 2018;1691:87–93.

    Article  CAS  PubMed  Google Scholar 

  112. Xing X, Li Z, Yang X, Li M, Liu C, Pang Y, et al. Adipose-derived mesenchymal stem cells-derived exosome-mediated microRNA-342-5p protects endothelial cells against atherosclerosis. Aging. 2020;12(4):3880–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Zeng CY, Xu J, Liu X, Lu YQ. Cardioprotective roles of endothelial progenitor cell-derived exosomes. Front Cardiovasc Med. 2021;8:717536.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Bai S, Yin Q, Dong T, Dai F, Qin Y, Ye L, et al. Endothelial progenitor cell-derived exosomes ameliorate endothelial dysfunction in a mouse model of diabetes. Biomed Pharmacother. 2020;131:110756.

    Article  CAS  PubMed  Google Scholar 

  115. Yoon EJ, Choi Y, Kim TM, Choi EK, Kim YB, Park D. The neuroprotective effects of exosomes derived from TSG101-overexpressing human neural stem cells in a stroke model. Int J Mol Sci. 2022;23(17):9532.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Qian C, Wang Y, Ji Y, Chen D, Wang C, Zhang G, et al. Neural stem cell-derived exosomes transfer miR-124-3p into cells to inhibit glioma growth by targeting FLOT2. Int J Oncol. 2022;61(4):115.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Smyth T, Kullberg M, Malik N, Smith-Jones P, Graner MW, Anchordoquy TJ. Biodistribution and delivery efficiency of unmodified tumor-derived exosomes. J Control Release. 2015;199:145–55.

    Article  CAS  PubMed  Google Scholar 

  118. Shafiei M, Ansari MNM, Razak SIA, Khan MUA. A comprehensive review on the applications of exosomes and liposomes in regenerative medicine and tissue engineering. Polymers. 2021;13(15):2529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. van der Koog L, Gandek TB, Nagelkerke A. Liposomes and extracellular vesicles as drug delivery systems: a comparison of composition, pharmacokinetics, and functionalization. Adv Healthc Mater. 2022;11(5):e2100639.

    Article  PubMed  Google Scholar 

  120. Srivastava A, Rathore S, Munshi A, Ramesh R. Organically derived exosomes as carriers of anticancer drugs and imaging agents for cancer treatment. Semin Cancer Biol. 2022;86:80–100.

    Article  CAS  PubMed  Google Scholar 

  121. Cheng X, Yan H, Pang S, Ya M, Qiu F, Qin P, et al. Liposomes as multifunctional nano-carriers for medicinal natural products. Front Chem. 2022;10:963004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Saleh AF, Lazaro-Ibanez E, Forsgard MA, Shatnyeva O, Osteikoetxea X, Karlsson F, et al. Extracellular vesicles induce minimal hepatotoxicity and immunogenicity. Nanoscale. 2019;11(14):6990–7001.

    Article  CAS  PubMed  Google Scholar 

  123. Zhu X, Badawi M, Pomeroy S, Sutaria DS, Xie Z, Baek A, et al. Comprehensive toxicity and immunogenicity studies reveal minimal effects in mice following sustained dosing of extracellular vesicles derived from HEK293T cells. J Extracell Vesicles. 2017;6(1):1324730.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Vakhshiteh F, Rahmani S, Ostad SN, Madjd Z, Dinarvand R, Atyabi F. Exosomes derived from miR-34a-overexpressing mesenchymal stem cells inhibit in vitro tumor growth: a new approach for drug delivery. Life Sci. 2021;266:118871.

    Article  CAS  PubMed  Google Scholar 

  125. He Z, Zhang Y, Feng N. Cell membrane-coated nanosized active targeted drug delivery systems homing to tumor cells: a review. Mater Sci Eng C Mater Biol Appl. 2020;106:110298.

    Article  CAS  PubMed  Google Scholar 

  126. Xu N, Guo R, Yang X, Li N, Yu J, Zhang P. Exosomes-mediated tumor treatment: one body plays multiple roles. Asian J Pharm Sci. 2021;17(3):385–400.

    Article  PubMed  PubMed Central  Google Scholar 

  127. Smyth TJ, Redzic JS, Graner MW, Anchordoquy TJ. Examination of the specificity of tumor cell derived exosomes with tumor cells in vitro. Biochim Biophys Acta. 2014;1838(11):2954–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Costa Verdera H, Gitz-Francois JJ, Schiffelers RM, Vader P. Cellular uptake of extracellular vesicles is mediated by clathrin-independent endocytosis and macropinocytosis. J Control Release. 2017;266:100–8.

    Article  CAS  PubMed  Google Scholar 

  129. Osorio-Querejeta I, Carregal-Romero S, Ayerdi-Izquierdo A, Mager I, Nash LA, Wood M, et al. MiR-219a-5p enriched extracellular vesicles induce OPC differentiation and EAE improvement more efficiently than liposomes and polymeric nanoparticles. Pharmaceutics. 2020;12(2):186.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Ben XY, Wang YR, Zheng HH, Li DX, Ren R, Ni PL, et al. Construction of exosomes that overexpress CD47 and evaluation of their immune escape. Front Bioeng Biotechnol. 2022;10:936951.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Lu M, Zhao X, Xing H, Xun Z, Zhu S, Lang L, et al. Comparison of exosome-mimicking liposomes with conventional liposomes for intracellular delivery of siRNA. Int J Pharm. 2018;550(1–2):100–13.

    Article  CAS  PubMed  Google Scholar 

  132. Kamerkar S, Lebleu VS, Sugimoto H, Yang S, Ruivo CF, Melo SA, et al. Exosomes facilitate therapeutic targeting of oncogenic KRAS in pancreatic cancer. Nature. 2017;546(7659):498–503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Lai CP, Mardini O, Ericsson M, Prabhakar S, Maguire C, Chen JW, et al. Dynamic biodistribution of extracellular vesicles in vivo using a multimodal imaging reporter. ACS Nano. 2014;8(1):483–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Yang Z, Shi J, Xie J, Wang Y, Sun J, Liu T, et al. Large-scale generation of functional mRNA-encapsulating exosomes via cellular nanoporation. Nat Biomed Eng. 2020;4(1):69–83.

    Article  CAS  PubMed  Google Scholar 

  135. Lázaro-Ibáñez E, Faruqu FN, Saleh AF, Silva AM, Tzu-Wen Wang J, Rak J, et al. Selection of fluorescent, bioluminescent, and radioactive tracers to accurately reflect extracellular vesicle biodistribution in vivo. ACS Nano. 2021;15(2):3212–27.

    Article  PubMed  PubMed Central  Google Scholar 

  136. 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, physicochemical properties, and pharmacokinetics. Eur J Pharm Sci. 2017;96:316–22.

    Article  CAS  PubMed  Google Scholar 

  137. Wiklander OP, Nordin JZ, O’loughlin A, Gustafsson Y, Corso G, Mager I, et al. Extracellular vesicle in vivo biodistribution is determined by cell source, route of administration and targeting. J Extracell Vesicles. 2015;4:26316.

    Article  PubMed  Google Scholar 

  138. Takahashi Y, Nishikawa M, Shinotsuka H, Matsui Y, Ohara S, Imai T, et al. Visualization and in vivo tracking of the exosomes of murine melanoma B16-BL6 cells in mice after intravenous injection. J Biotechnol. 2013;165(2):77–84.

    Article  CAS  PubMed  Google Scholar 

  139. Choi H, Choi Y, Yim HY, Mirzaaghasi A, Yoo JK, Choi C. Biodistribution of exosomes and engineering strategies for targeted delivery of therapeutic exosomes. Tissue Eng Regen Med. 2021;18(4):499–511.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Fuhrmann G, Herrmann IK, Stevens MM. Cell-derived vesicles for drug therapy and diagnostics: opportunities and challenges. Nano Today. 2015;10(3):397–409.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Abdel-Haq H. Blood exosomes as a tool for monitoring treatment efficacy and progression of neurodegenerative diseases. Neural Regen Res. 2019;14(1):72–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Zlotogorski-Hurvitz A, Dayan D, Chaushu G, Korvala J, Salo T, Sormunen R, et al. Human saliva-derived exosomes: comparing methods of isolation. J Histochem Cytochem. 2015;63(3):181–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. He L, Zhu D, Wang J, Wu X. A highly efficient method for isolating urinary exosomes. Int J Mol Med. 2019;43(1):83–90.

    CAS  PubMed  Google Scholar 

  144. Mitra A, Yoshida-Court K, Solley TN, Mikkelson M, Yeung CLA, Nick A, et al. Extracellular vesicles derived from ascitic fluid enhance growth and migration of ovarian cancer cells. Sci Rep. 2021;11(1):9149.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Li Z, Wang Y, Ding Y, Repp L, Kwon GS, Hu Q. Cell-based delivery systems: emerging carriers for immunotherapy. Adv Funct Mater. 2021;31(23):2100088.

    Article  CAS  Google Scholar 

  146. Gong C, Tian J, Wang Z, Gao Y, Wu X, Ding X, et al. Functional exosome-mediated co-delivery of doxorubicin and hydrophobically modified microRNA 159 for triple-negative breast cancer therapy. J Nanobiotechnol. 2019;17(1):93.

    Article  Google Scholar 

  147. Li YJ, Wu JY, Wang JM, Hu XB, Cai JX, Xiang DX. Gemcitabine loaded autologous exosomes for effective and safe chemotherapy of pancreatic cancer. Acta Biomater. 2020;101:519–30.

    Article  CAS  PubMed  Google Scholar 

  148. Yuan D, Zhao Y, Banks WA, Bullock KM, Haney M, Batrakova E, et al. Macrophage exosomes as natural nanocarriers for protein delivery to inflamed brain. Biomaterials. 2017;142:1–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Zhang Q, Zhang H, Ning T, Liu D, Deng T, Liu R, et al. Exosome-delivered c-Met siRNA could reverse chemoresistance to cisplatin in gastric cancer. Int J Nanomed. 2020;15:2323–35.

    Article  CAS  Google Scholar 

  150. Lou G, Song X, Yang F, Wu S, Wang J, Chen Z, et al. Exosomes derived from miR-122-modified adipose tissue-derived MSCs increase chemosensitivity of hepatocellular carcinoma. J Hematol Oncol. 2015;8:122.

    Article  PubMed  PubMed Central  Google Scholar 

  151. Betzer O, Perets N, Angel A, Motiei M, Sadan T, Yadid G, et al. In vivo neuroimaging of exosomes using gold nanoparticles. ACS Nano. 2017;11(11):10883–93.

    Article  CAS  PubMed  Google Scholar 

  152. Tian Y, Li S, Song J, Ji T, Zhu M, Anderson GJ, et al. A doxorubicin delivery platform using engineered natural membrane vesicle exosomes for targeted tumor therapy. Biomaterials. 2014;35(7):2383–90.

    Article  CAS  PubMed  Google Scholar 

  153. Qu M, Lin Q, Huang L, Fu Y, Wang L, He S, et al. Dopamine-loaded blood exosomes targeted to brain for better treatment of Parkinson’s disease. J Control Release. 2018;287:156–66.

    Article  CAS  PubMed  Google Scholar 

  154. Wang H, Sui H, Zheng Y, Jiang Y, Shi Y, Liang J, et al. Curcumin-primed exosomes potently ameliorate cognitive function in AD mice by inhibiting hyperphosphorylation of the Tau protein through the AKT/GSK-3beta pathway. Nanoscale. 2019;11(15):7481–96.

    Article  CAS  PubMed  Google Scholar 

  155. Qi Y, Guo L, Jiang Y, Shi Y, Sui H, Zhao L. Brain delivery of quercetin-loaded exosomes improved cognitive function in AD mice by inhibiting phosphorylated Tau-mediated neurofibrillary tangles. Drug Deliv. 2020;27(1):745–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Gondaliya P, Sayyed AA, Bhat P, Mali M, Arya N, Khairnar A, et al. Mesenchymal stem cell-derived exosomes loaded with miR-155 inhibitor ameliorate diabetic wound healing. Mol Pharm. 2022;19(5):1294–308.

    Article  CAS  PubMed  Google Scholar 

  157. Parodi A, Molinaro R, Sushnitha M, Evangelopoulos M, Martinez JO, Arrighetti N, et al. Bio-inspired engineering of cell- and virus-like nanoparticles for drug delivery. Biomaterials. 2017;147:155–68.

    Article  CAS  PubMed  Google Scholar 

  158. Li YJ, Wu JY, Liu J, Xu W, Qiu X, Huang S, et al. Artificial exosomes for translational nanomedicine. J Nanobiotechnol. 2021;19(1):242.

    Article  Google Scholar 

  159. Wang X, Zhao X, Zhong Y, Shen J, An W. Biomimetic exosomes: a new generation of drug delivery system. Front Bioeng Biotechnol. 2022;10:865682.

    Article  PubMed  PubMed Central  Google Scholar 

  160. Jia X, Tang J, Yao C, Yang D. Recent progress of extracellular vesicle engineering. ACS Biomater Sci Eng. 2021;7(9):4430–8.

    Article  CAS  PubMed  Google Scholar 

  161. Kim YS, Kim JY, Cho R, Shin DM, Lee SW, Oh YM. Adipose stem cell-derived nanovesicles inhibit emphysema primarily via an FGF2-dependent pathway. Exp Mol Med. 2017;49(1):e284.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Wu JY, Li YJ, Hu XB, Huang S, Luo S, Tang T, et al. Exosomes and biomimetic nanovesicles-mediated anti-glioblastoma therapy: a head-to-head comparison. J Control Release. 2021;336:510–21.

    Article  CAS  PubMed  Google Scholar 

  163. Guo P, Busatto S, Huang J, Morad G, Moses MA. A facile magnetic extrusion method for preparing endosome-derived vesicles for cancer drug delivery. Adv Funct Mater. 2021;31(44):2008326.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Zhang KL, Wang YJ, Sun J, Zhou J, Xing C, Huang G, et al. Artificial chimeric exosomes for anti-phagocytosis and targeted cancer therapy. Chem Sci. 2019;10(5):1555–61.

    Article  CAS  PubMed  Google Scholar 

  165. Rayamajhi S, Nguyen TDT, Marasini R, Aryal S. Macrophage-derived exosome-mimetic hybrid vesicles for tumor targeted drug delivery. Acta Biomater. 2019;94:482–94.

    Article  CAS  PubMed  Google Scholar 

  166. Li L, He D, Guo Q, Zhang Z, Ru D, Wang L, et al. Exosome-liposome hybrid nanoparticle codelivery of TP and miR497 conspicuously overcomes chemoresistant ovarian cancer. J Nanobiotechnology. 2022;20(1):50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Ni J, Mi Y, Wang B, Zhu Y, Ding Y, Ding Y, et al. Naturally equipped urinary exosomes coated poly (2-ethyl-2-oxazoline)-poly (D, L-lactide) nanocarriers for the pre-clinical translation of breast cancer. Bioengineering. 2022;9(8):363.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Holay M, Zhou J, Park JH, Landa I, Ventura CJ, Gao W, et al. Organotropic targeting of biomimetic nanoparticles to treat lung disease. Bioconjug Chem. 2022;33(4):586–93.

    Article  CAS  PubMed  Google Scholar 

  169. Yang Z, Xie J, Zhu J, Kang C, Chiang C, Wang X, et al. Functional exosome-mimic for delivery of siRNA to cancer: in vitro and in vivo evaluation. J Control Release. 2016;243:160–71.

    Article  CAS  PubMed  Google Scholar 

  170. Choo YW, Kang M, Kim HY, Han J, Kang S, Lee JR, et al. M1 macrophage-derived nanovesicles potentiate the anticancer efficacy of immune checkpoint inhibitors. ACS Nano. 2018;12(9):8977–93.

    Article  CAS  PubMed  Google Scholar 

  171. Jo W, Jeong D, Kim J, Cho S, Jang SC, Han C, et al. Microfluidic fabrication of cell-derived nanovesicles as endogenous RNA carriers. Lab Chip. 2014;14(7):1261–9.

    Article  CAS  PubMed  Google Scholar 

  172. Yoon J, Jo W, Jeong D, Kim J, Jeong H, Park J. Generation of nanovesicles with sliced cellular membrane fragments for exogenous material delivery. Biomaterials. 2015;59:12–20.

    Article  CAS  PubMed  Google Scholar 

  173. Jang SC. Bioinspired exosome-mimetic nanovesicles for targeted delivery of chemotherapeutics to malignant tumors. ACS Nano. 2013;7(9):7698–710.

    Article  CAS  PubMed  Google Scholar 

  174. Zhu J, Liu Z, Wang L, Jin Q, Zhao Y, Du A, et al. Exosome mimetics-loaded hydrogel accelerates wound repair by transferring functional mitochondrial proteins. Front Bioeng Biotechnol. 2022;10:866505.

    Article  PubMed  PubMed Central  Google Scholar 

  175. Wu JY, Ji AL, Wang ZX, Qiang GH, Qu Z, Wu JH, et al. Exosome-mimetic nanovesicles from hepatocytes promote hepatocyte proliferation in vitro and liver regeneration in vivo. Sci Rep. 2018;8(1):2471.

    Article  PubMed  PubMed Central  Google Scholar 

  176. Fernandes M, Lopes I, Magalhaes L, Sarria MP, Machado R, Sousa JC, et al. Novel concept of exosome-like liposomes for the treatment of Alzheimer’s disease. J Control Release. 2021;336:130–43.

    Article  CAS  PubMed  Google Scholar 

  177. Sakai-Kato K, Yoshida K, Takechi-Haraya Y, Izutsu KI. Physicochemical characterization of liposomes that mimic the lipid composition of exosomes for effective intracellular trafficking. Langmuir. 2020;36(42):12735–44.

    Article  CAS  PubMed  Google Scholar 

  178. Li K, Chang S, Wang Z, Zhao X, Chen D. A novel micro-emulsion and micelle assembling method to prepare DEC205 monoclonal antibody coupled cationic nanoliposomes for simulating exosomes to target dendritic cells. Int J Pharm. 2015;491(1–2):105–12.

    Article  CAS  PubMed  Google Scholar 

  179. De Miguel D, Basanez G, Sanchez D, Malo PG, Marzo I, Larrad L, et al. Liposomes decorated with Apo2L/TRAIL overcome chemoresistance of human hematologic tumor cells. Mol Pharm. 2013;10(3):893–904.

    Article  PubMed  Google Scholar 

  180. Lu M, Zhao X, Xing H, Liu H, Lang L, Yang T, et al. Cell-free synthesis of connexin 43-integrated exosome-mimetic nanoparticles for siRNA delivery. Acta Biomater. 2019;96:517–36.

    Article  CAS  PubMed  Google Scholar 

  181. Molinaro R, Martinez JO, Zinger A, De Vita A, Storci G, Arrighetti N, et al. Leukocyte-mimicking nanovesicles for effective doxorubicin delivery to treat breast cancer and melanoma. Biomater Sci. 2020;8(1):333–41.

    Article  CAS  PubMed  Google Scholar 

  182. Li YJ, Wu JY, Hu XB, Ding T, Tang T, Xiang DX. Biomimetic liposome with surface-bound elastase for enhanced tumor penetration and chemo-immumotherapy. Adv Healthc Mater. 2021;10(19):e2100794.

    Article  PubMed  Google Scholar 

  183. Jhan YY, Prasca-Chamorro D, Palou Zuniga G, Moore DM, Arun Kumar S, Gaharwar AK, et al. Engineered extracellular vesicles with synthetic lipids via membrane fusion to establish efficient gene delivery. Int J Pharm. 2020;573:118802.

    Article  CAS  PubMed  Google Scholar 

  184. Sun L, Fan M, Huang D, Li B, Xu R, Gao F, et al. Clodronate-loaded liposomal and fibroblast-derived exosomal hybrid system for enhanced drug delivery to pulmonary fibrosis. Biomaterials. 2021;271:120761.

    Article  CAS  PubMed  Google Scholar 

  185. Sato YT, Umezaki K, Sawada S, Mukai SA, Sasaki Y, Harada N, et al. Engineering hybrid exosomes by membrane fusion with liposomes. Sci Rep. 2016;6:21933.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Li Q, Song Y, Wang Q, Chen J, Gao J, Tan H, et al. Engineering extracellular vesicles with platelet membranes fusion enhanced targeted therapeutic angiogenesis in a mouse model of myocardial ischemia reperfusion. Theranostics. 2021;11(8):3916–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Hu S, Wang X, Li Z, Zhu D, Cores J, Wang Z, et al. Platelet membrane and stem cell exosome hybrid enhances cellular uptake and targeting to heart injury. Nano Today. 2021;39:101210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Liu H, Miao Z, Zha Z. Cell membrane-coated nanoparticles for immunotherapy. Chinese Chem Lett. 2022;33(4):1673–80.

    Article  CAS  Google Scholar 

  189. Qiao L, Hu S, Huang K, Su T, Li Z, Vandergriff A, et al. Tumor cell-derived exosomes home to their cells of origin and can be used as Trojan horses to deliver cancer drugs. Theranostics. 2020;10(8):3474–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Cheng G, Li W, Ha L, Han X, Hao S, Wan Y, et al. Self-assembly of extracellular vesicle-like metal-organic framework nanoparticles for protection and intracellular delivery of biofunctional proteins. J Am Chem Soc. 2018;140(23):7282–91.

    Article  CAS  PubMed  Google Scholar 

  191. Tian R, Wang Z, Niu R, Wang H, Guan W, Chang J. Tumor exosome mimicking nanoparticles for tumor combinatorial chemo-photothermal therapy. Front Bioeng Biotechnol. 2020;8:1010.

    Article  PubMed  PubMed Central  Google Scholar 

  192. Zheng Y, Li M, Weng B, Mao H, Zhao J. Exosome-based delivery nanoplatforms: next-generation theranostic platforms for breast cancer. Biomater Sci. 2022;10(7):1607–25.

    Article  CAS  PubMed  Google Scholar 

  193. Wang K, Ye H, Zhang X, Wang X, Yang B, Luo C, et al. An exosome-like programmable-bioactivating paclitaxel prodrug nanoplatform for enhanced breast cancer metastasis inhibition. Biomaterials. 2020;257:120224.

    Article  CAS  PubMed  Google Scholar 

  194. Hu H, Yang C, Zhang F, Li M, Tu Z, Mu L, et al. A versatile and robust platform for the scalable manufacture of biomimetic nanovaccines. Adv Sci. 2021;8(15):2002020.

    Article  CAS  Google Scholar 

  195. Jimenez-Jimenez C, Manzano M, Vallet-Regi M. Nanoparticles coated with cell membranes for biomedical applications. Biology. 2020;9(11):406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Luk BT, Jack Hu CM, Fang RH, Dehaini D, Carpenter C, Gao W, et al. Interfacial interactions between natural RBC membranes and synthetic polymeric nanoparticles. Nanoscale. 2014;6(5):2730–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Li S, Wu Y, Ding F, Yang J, Li J, Gao X, et al. Engineering macrophage-derived exosomes for targeted chemotherapy of triple-negative breast cancer. Nanoscale. 2020;12(19):10854–62.

    Article  CAS  PubMed  Google Scholar 

  198. Zhai Y, Su J, Ran W, Zhang P, Yin Q, Zhang Z, et al. Preparation and application of cell membrane-camouflaged nanoparticles for cancer therapy. Theranostics. 2017;7(10):2575–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Bose RJC, Uday Kumar S, Zeng Y, Afjei R, Robinson E, Lau K, et al. Tumor cell-derived extracellular vesicle-coated nanocarriers: an efficient theranostic platform for the cancer-specific delivery of anti-miR-21 and imaging agents. ACS Nano. 2018;12(11):10817–32.

    Article  PubMed  PubMed Central  Google Scholar 

  200. Sancho-Albero M, Encabo-Berzosa MDM, Beltran-Visiedo M, Fernandez-Messina L, Sebastian V, Sanchez-Madrid F, et al. Efficient encapsulation of theranostic nanoparticles in cell-derived exosomes: leveraging the exosomal biogenesis pathway to obtain hollow gold nanoparticle-hybrids. Nanoscale. 2019;11(40):18825–36.

    Article  CAS  PubMed  Google Scholar 

  201. Liu Q, Fan T, Zheng Y, Yang SL, Yu Z, Duo Y, et al. Immunogenic exosome-encapsulated black phosphorus nanoparticles as an effective anticancer photo-nanovaccine. Nanoscale. 2020;12(38):19939–52.

    Article  CAS  PubMed  Google Scholar 

  202. Li X, Wang Y, Shi L, Li B, Li J, Wei Z, et al. Magnetic targeting enhances the cutaneous wound healing effects of human mesenchymal stem cell-derived iron oxide exosomes. J Nanobiotechnology. 2020;18(1):113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Zhang C, Song J, Lou L, Qi X, Zhao L, Fan B, et al. Doxorubicin-loaded nanoparticle coated with endothelial cells-derived exosomes for immunogenic chemotherapy of glioblastoma. Bioeng Transl Med. 2021;6(3):e10203.

    Article  CAS  PubMed  Google Scholar 

  204. Xiong F, Ling X, Chen X, Chen J, Tan J, Cao W, et al. Pursuing specific chemotherapy of orthotopic breast cancer with lung metastasis from docking nanoparticles driven by bioinspired exosomes. Nano Lett. 2019;19(5):3256–66.

    Article  CAS  PubMed  Google Scholar 

  205. Fang RH, Kroll AV, Gao W, Zhang L. Cell membrane coating nanotechnology. Adv Mater. 2018;30(23):e1706759.

    Article  PubMed  PubMed Central  Google Scholar 

  206. Choi B, Park W, Park SB, Rhim WK, Han DK. Recent trends in cell membrane-cloaked nanoparticles for therapeutic applications. Methods. 2020;177:2–14.

    Article  CAS  PubMed  Google Scholar 

  207. Chen Y, Cheng K. Advances of biological-camouflaged nanoparticles delivery system. Nano Res. 2020;13(10):2617–24.

    Article  CAS  Google Scholar 

  208. Wang G, Hu W, Chen H, Shou X, Ye T, Xu Y. Cocktail strategy based on NK cell-derived exosomes and their biomimetic nanoparticles for dual tumor therapy. Cancers. 2019;11(10):1560.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Li J, Kong J, Ma S, Li J, Mao M, Chen K, et al. Exosome-coated 10B carbon dots for precise boron neutron capture therapy in a mouse model of glioma in situ. Adv Funct Mater. 2021;31(24):2100969.

    Article  CAS  Google Scholar 

  210. Han Z, Lv W, Li Y, Chang J, Zhang W, Liu C, et al. Improving tumor targeting of exosomal membrane-coated polymeric nanoparticles by conjugation with aptamers. ACS Appl Bio Mater. 2020;3(5):2666–73.

    Article  CAS  PubMed  Google Scholar 

  211. Yousefiasl S, Manoochehri H, Makvandi P, Afshar S, Salahinejad E, Khosraviyan P, et al. Chitosan/alginate bionanocomposites adorned with mesoporous silica nanoparticles for bone tissue engineering. J Nanostruct Chem. 2022. https://doi.org/10.1007/s40097-022-00507-z.

    Article  Google Scholar 

  212. Sood A, Bhaskar R, Won SY, Seok YJ, Kumar A, Han SS. Disulfide bond-driven hyaluronic acid/sericin nanoparticles for wound-healing application. J Nanostruct Chem. 2022. https://doi.org/10.1007/s40097-022-00505-1.

    Article  Google Scholar 

  213. Li W, Liu Y, Zhang P, Tang Y, Zhou M, Jiang W, et al. Tissue-engineered bone immobilized with human adipose stem cells-derived exosomes promotes bone regeneration. ACS Appl Mater Interfaces. 2018;10(6):5240–54.

    Article  CAS  PubMed  Google Scholar 

  214. Yao J, Huang K, Zhu D, Chen T, Jiang Y, Zhang J, et al. A minimally invasive exosome spray repairs heart after myocardial infarction. ACS Nano. 2021;15(7):11099–111.

    Article  CAS  PubMed  Google Scholar 

  215. Rabiee N, Yaraki MT, Garakani SM, Garakani SM, Ahmadi S, Lajevardi A, et al. Recent advances in porphyrin-based nanocomposites for effective targeted imaging and therapy. Biomaterials. 2020;232:119707.

    Article  CAS  PubMed  Google Scholar 

  216. Teng X, Chen L, Chen W, Yang J, Yang Z, Shen Z. Mesenchymal stem cell-derived exosomes improve the microenvironment of infarcted myocardium contributing to angiogenesis and anti-inflammation. Cell Physiol Biochem. 2015;37(6):2415–24.

    Article  CAS  PubMed  Google Scholar 

  217. Khongkow M, Yata T, Boonrungsiman S, Ruktanonchai UR, Graham D, Namdee K. Surface modification of gold nanoparticles with neuron-targeted exosome for enhanced blood-brain barrier penetration. Sci Rep. 2019;9(1):8278.

    Article  PubMed  PubMed Central  Google Scholar 

  218. Soares Martins T, Trindade D, Vaz M, Campelo I, Almeida M, Trigo G, et al. Diagnostic and therapeutic potential of exosomes in Alzheimer’s disease. J Neurochem. 2021;156(2):162–81.

    Article  CAS  PubMed  Google Scholar 

  219. Jia L, Zhu M, Kong C, Pang Y, Zhang H, Qiu Q, et al. Blood neuro-exosomal synaptic proteins predict Alzheimer’s disease at the asymptomatic stage. Alzheimers Dement. 2021;17(1):49–60.

    Article  CAS  PubMed  Google Scholar 

  220. Bakulski KM, Hu H, Park SK. Lead, cadmium and Alzheimer’s disease. In: Martin CR, Preedy VR, editors. Genetics, neurology, behavior, and diet in dementia. Academic Press; 2020. p. 813–30.

    Chapter  Google Scholar 

  221. Zaazaa AM, Abd El-Motelp BA, Ali NA, Youssef AM, Sayed MA, Mohamed SH. Stem cell-derived exosomes and copper sulfide nanoparticles attenuate the progression of neurodegenerative disorders induced by cadmium in rats. Heliyon. 2021;8(1):e08622.

    Article  PubMed  PubMed Central  Google Scholar 

  222. Meade RM, Fairlie DP, Mason JM. Alpha-synuclein structure and Parkinson’s disease: lessons and emerging principles. Mol Neurodegener. 2019;14(1):29.

    Article  PubMed  PubMed Central  Google Scholar 

  223. Liu L, Li Y, Peng H, Liu R, Ji W, Shi Z, et al. Targeted exosome coating gene-chem nanocomplex as “nanoscavenger” for clearing a-synuclein and immune activation of Parkinson’s disease. Sci Adv. 2020;6(50):eaba3967.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Lee JY, Kim HS. Extracellular vesicles in regenerative medicine: potentials and challenges. Tissue Eng Regen Med. 2021;18(4):479–84.

    Article  PubMed  PubMed Central  Google Scholar 

  225. Szatanek R, Baran J, Siedlar M, Baj-Krzyworzeka M. Isolation of extracellular vesicles: determining the correct approach. Int J Mol Med. 2015;36(1):11–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Zhang X, Takeuchi T, Takeda A, Mochizuki H, Nagai Y. Comparison of serum and plasma as a source of blood extracellular vesicles: Increased levels of platelet-derived particles in serum extracellular vesicle fractions alter content profiles from plasma extracellular vesicle fractions. PLoS ONE. 2022;17(6):e0270634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Coumans FAW, Brisson AR, Buzas EI, Dignat-George F, Drees EEE, El-Andaloussi S, et al. Methodological guidelines to study extracellular vesicles. Circ Res. 2017;120(10):1632–48.

    Article  CAS  PubMed  Google Scholar 

  228. Kim HY, Kim TJ, Kang L, Kim YJ, Kang MK, Kim J, et al. Mesenchymal stem cell-derived magnetic extracellular nanovesicles for targeting and treatment of ischemic stroke. Biomaterials. 2020;243:119942.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by grants from Fundação para a Ciência e Tecnologia (FCT) (SFRH/BD/148771/2019, 2021.05914.BD, PTDC/BTM-MAT/4738/2020), and also from the European Research Council—ERC Starting Grant (848325).

Author information

Authors and Affiliations

Authors

Contributions

DL and JL conducted literature searches, drafted the manuscript, and prepared figures and tables. MPS, DP, and JC drafted, edited, and revised the manuscript. NR, OM, ZHG, and XDW edited and revised the manuscript. PM contributed to figure designs. FV, ACPS, and PM commented, edited, and provided substantial improvements. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Xiang-Dong Wang, Pooyan Makvandi or Ana Cláudia Paiva-Santos.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

JC is a co-founder and shareholder of TargTex S.A. The remaining authors declare no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lopes, D., Lopes, J., Pereira-Silva, M. et al. Bioengineered exosomal-membrane-camouflaged abiotic nanocarriers: neurodegenerative diseases, tissue engineering and regenerative medicine. Military Med Res 10, 19 (2023). https://doi.org/10.1186/s40779-023-00453-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40779-023-00453-z

Keywords