Next Article in Journal
A Machine Learning Approach to Identify Key Residues Involved in Protein–Protein Interactions Exemplified with SARS-CoV-2 Variants
Previous Article in Journal
Hair Thickness Growth Effect of Adenosine Complex in Male-/Female-Patterned Hair Loss via Inhibition of Androgen Receptor Signaling
Previous Article in Special Issue
Combinatory Nanovesicle with siRNA-Loaded Extracellular Vesicle and IGF-1 for Osteoarthritis Treatments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Vesicles as Next-Generation Diagnostics and Advanced Therapy Medicinal Products

by
Agnieszka Stawarska
1,*,
Magdalena Bamburowicz-Klimkowska
1,
Elise Runden-Pran
2,
Maria Dusinska
2,
Mihaela Roxana Cimpan
3,
Ivan Rios-Mondragon
3 and
Ireneusz P. Grudzinski
1
1
Department of Toxicology and Food Science, Faculty of Pharmacy, Medical University of Warsaw, Banacha Str. 1, 02-097 Warsaw, Poland
2
Health Effects Laboratory, Department of Environmental Chemistry, Norwegian Institute for Air Research, 2007 Kjeller, Norway
3
Biomaterials—Department of Clinical Dentistry, Faculty of Medicine, University of Bergen, Årstadveien Str. 19, 5009 Bergen, Norway
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(12), 6533; https://doi.org/10.3390/ijms25126533
Submission received: 26 April 2024 / Revised: 10 June 2024 / Accepted: 11 June 2024 / Published: 13 June 2024
(This article belongs to the Special Issue Roles and Function of Extracellular Vesicles in Diseases 2.0)

Abstract

:
Extracellular vesicles (EVs) hold great promise for clinical application as new diagnostic and therapeutic modalities. This paper describes major GMP-based upstream and downstream manufacturing processes for EV large-scale production, also focusing on post-processing technologies such as surface bioengineering and uploading studies to yield novel EV-based diagnostics and advanced therapy medicinal products. This paper also focuses on the quality, safety, and efficacy issues of the bioengineered EV drug candidates before first-in-human studies. Because clinical trials involving extracellular vesicles are on the global rise, this paper encompasses different clinical studies registered on clinical-trial register platforms, with varying levels of advancement, highlighting the growing interest in EV-related clinical programs. Navigating the regulatory affairs of EVs poses real challenges, and obtaining marketing authorization for EV-based medicines remains complex due to the lack of specific regulatory guidelines for such novel products. This paper discusses the state-of-the-art regulatory knowledge to date on EV-based diagnostics and medicinal products, highlighting further research and global regulatory needs for the safe and reliable implementation of bioengineered EVs as diagnostic and therapeutic tools in clinical settings. Post-marketing pharmacovigilance for EV-based medicinal products is also presented, mainly addressing such topics as risk assessment and risk management.

1. Introduction

By the term “extracellular vesicles” (EVs), according to the currently accepted nomenclature, we mean heterogeneous vesicles of cellular origin, surrounded by a lipid bilayer, incapable of self-replication (not containing a functional nucleus), and which are produced by most cells through various mechanisms [1]. So far, EVs have been detected in plants [2,3,4], bacteria [5,6,7,8], fungi [6,9], in vitro cultures of eukaryotic cells, and biological samples obtained from humans and animals [10,11,12,13]. Established in 2011, the International Society of Extracellular Vesicles (ISEV) has updated the EV nomenclature. Currently, three types of EVs are distinguished, namely (i) microvesicles (MVs) or ectosomes, (ii) apoptotic bodies, and (iii) exosomes (Figure 1) [14]. MVs refer to EVs in the range of 100–1000 nm that are formed at and directly secreted from the cell membrane through a complex mechanism consisting of exfoliation or budding under normal circumstances or in response to different stimuli [15]. Several proteins have been identified as specific biomarkers of MVs, including CD40, ribosylation factor ADP 6 (ARF6), selectin, phosphatidylserine, and members of the Rho family of GTPases [16]. Apoptotic bodies (1000–5000 nm) are membrane-bound cellular remains released when programmed cell death is induced and contain DNA fragments that do not encode RNA and cellular organelles [17]. Annexin V and histones are reported as specific proteins of apoptotic bodies [18,19,20]. Exosomes are EVs with a size range of 30–150 nm, which are heterogeneous in both size and cargo composition. Therefore, exosomes have diverse physiological roles depending on the donor and target cells [21]. Exosomes have been reported to contain several types of specific surface markers, such as tetraspanins (CD9, CD63, and CD81), heat shock proteins (Hsp70 and Hsp90), MVB synthesis proteins, and membrane transporters and fusion proteins (annexins and flotillins) [22,23]. Although the origins of each type of vesicle and the corresponding markers have been determined, available technologies are not able to provide pure isolates of each type of EV since there is some overlap of physicochemical properties and molecular markers. This is why the ISEV, in its latest position paper issued in 2023 (MISEV2023), recommends the use of the generic term “EVs” instead of inconsistently mentioned and sometimes misleading terms, such as “exosomes” and “ectosomes”, which are associated with difficult-to-determine biogenesis pathways [1].
During the last decade, huge interest has arisen in the studies of EV subclasses, referred to here as exosomes, that are secreted by various cell types and are composed of lipids, proteins, and nucleic acid cargos [24]. Note that exosomes are formed through the inward budding of the endosomal membrane, resulting in the formation of intraluminal vesicles (ILVs) within multivesicular bodies (MVBs). Both endosomal sorting complex required for transport (ESCRT) and non-ESCRT systems are involved in these processes [25]. Interestingly, MVBs can then fuse with the plasma membrane, releasing the ILVs into the extracellular environment, and they are further referred to as exosomes [25]. After being released, EVs affect recipient cells at the molecular level through several mechanisms, which were described in details by Krylova at al [26]. This may include soluble signaling, which involves the proteolytic cleavage of ligands from the exosome surface, or alternative splicing, whereas juxtacrine signaling requires the assembly of ligands and receptors on the exosome and target cell surface. EVs can interact with surface receptors on recipient cells, triggering signal transduction pathways. This interaction can lead to changes in cellular behavior, such as proliferation, differentiation, or apoptosis. Then, EVs can fuse with the plasma membrane of recipient cells, directly delivering their cargo into the cytoplasm. This allows for the rapid and efficient transfer of functional molecules. Moreover, recipient cells can internalize EVs through endocytosis. Once inside, the EVs can be processed in endosomes and lysosomes, releasing their contents into the cytoplasm, affecting intracellular signaling pathways.
Exosomes play important roles in cell-to-cell communication and are thought to function in a variety of physiological and pathological processes, including immune modulation, tissue regeneration, and cancer progression [27]. Exosomes can transfer a variety of molecular cargos, including proteins, lipids, and nucleic acids, altering the phenotype and behavior of recipient cells [28]. The protein content of exosomes is highly diverse, and it can vary depending on the cell type that secretes them [29,30,31,32]. Lipids in exosomes include phospholipids, cholesterol, and sphingolipids [33]. The nucleic acids in exosomes are primarily composed of RNA species, including messenger RNA, microRNA, and other small non-coding RNA species [33]. Exosomes can cross biological barriers, such as the blood–brain barrier, and have the ability to deliver their cargo to specific target cells [28,33,34]. Due to their ability to transfer biologically active molecules, exosomes have garnered significant interest in recent years as potential therapeutic agents or diagnostic tools [35,36] (Figure 2).
Since EVs circulate in almost all types of bodily fluids, such as blood, urine, saliva, broncho-alveolar fluid, breast milk, and semen, the collection of which is of a non-invasive nature, they also have huge potential as biomarkers of many diseases. These include tumors [22,37,38,39], bone-related diseases [13], amyotrophic lateral sclerosis [40], traumatic brain injury [41], lung disease [42], liver disease [43,44], inflammatory eyes disease [45], helminthic infection [46], autoimmune diseases [47], and kidney disease [48]. It has been shown that the analysis of the composition of EVs (proteins, miRNA, mRNA, and lipids) can be useful not only in diagnosing but also in assessing the progression of the disease [22,49,50,51,52,53,54,55]. Depending on their origin, the EVs can modulate the morphology and metabolism of the target cells, which can lead to pathogenesis. EVs, notably exosomes, serve as distinctive messengers facilitating communication between tumors and host organisms. They are extensively investigated as endogenous nanoscale transporters carrying various molecules in both physiological and pathological conditions. Research indicates that EVs play a pivotal role in processes such as thrombosis, angiogenesis, vascular dysfunction, and others, influencing the progression of various hematologic diseases, including cancers [56]. For example, in hematological malignancies, EVs originating from cancer cells assume a critical role in facilitating intercellular communication. They accomplish this by transferring genetic materials, proteins, and other as-yet-unidentified molecules between cancer cells and the bone-marrow microenvironment. This transfer plausibly influences processes such as cell transformation, proliferation, and ultimately the progression of malignancy [57]. Due to the ability of EVs to deliver bioactive molecules and cross biological barriers, EVs are increasingly being considered as potential therapeutic agents [22,58,59,60,61,62] (Figure 2).

2. Extracellular Vesicles Loaded with Diagnostic and Therapeutic Cargos

The use of extracellular vesicles (EVs) as new biological transporters for different diagnostic or therapeutic moieties requires the large-scale production of cellular mass with Good Manufacturing Practice (GMP) regulations [63]. In most of the recent studies, the source of EVs was mesenchymal stromal cells, adipose tissue-derived stem cells, and human embryonic kidney cells [64]. However, the EV samples could also be collected from some bacterial cultures, plant cells, and even bovine milk [65]. Although obtaining EVs from human fluids, such as blood and plasma, does not require specific cell-culture technologies, the large-scale production of EVs from a specific cellular mass will need standardized cell-culture approaches following GMP protocols [64] (Figure 3).
Manufacturing EVs includes the so-called upstream and downstream processes [66]. The upstream processes are related to cell banking and cell expansion for the production of conditioned media [67,68]. It should be emphasized that setups based on cell culture in conventional 2D flask systems are limited to the production of small number of EVs, and this is a problem for commercial purposes. Therefore, 3D cell culture systems have been developed. These include stirred-tank, vertical wheel, rocker, and hollow-fiber bioreactors [69]. Notably, the hollow-fiber bioreactor is widely used for EV production. This system is composed of thousands of parallel hollow fibers whose membrane supports cell attachment and the transport of small molecules (nutrients) from the fiber lumen to the extra-capillary space where the cell mass resides [70]. Furthermore, flow control systems continuously deliver fresh or recirculated cell culture medium to the lumen of the fibers to support cell growth, while the harvesting of EV-conditioned medium from the extra-capillary space is performed at specific time intervals. This unique construction and operation allow for the production of a large cell mass and high number of EVs [71]. The optimal production of a large cell mass in hollow-fiber bioreactors should take into account the quality control of parameters such as the pH; temperature; osmolarity; hypoxia; and nutrients added to the medium, including glucose and ion contents [72]. Therefore, the upscaling of hollow-fiber bioreactors from preclinical studies to manufacturing scale requires numerous iterative tests to keep the cell mass phenotype acceptable while producing EV batches on a large scale (Supplementary Figure S1). Evidence from recent studies indicates that the large-cell mass production in hollow-fiber bioreactors could be applied for the manufacturing of EVs following GMP standards [73].
Creating therapeutic EVs involves several key steps, each critical to ensure the quality and efficacy of the final biological product. Quality control (QC) measures and standards in EV production encompass rigorous assessment of source cells, isolation and purification methods, EV characterization, functional assays, stability testing, and regulatory compliance. By implementing these QC measures, manufacturers can ensure consistency, safety, and efficacy of EV-based products for biomedical applications. The ongoing monitoring and continuous improvement of QC processes are essential to maintain product quality and meet regulatory standards. A detailed diagram of the EVs’ production, including the purpose, process description, and quality control at each stage, is presented in Table 1 and Figure 3.
Scaling up EV production for clinical and commercial use presents challenges in consistency, efficiency, and cost-effectiveness. Optimizing culture systems, such as bioreactors, and engineering source cells can enhance proliferation and EV secretion. Improved isolation methods, like tangential flow filtration and automation, can increase throughput and purity. Standardized protocols and real-time monitoring ensure quality and safety. Cost-saving measures, novel technologies, and regulatory compliance are essential for successful upscaling. Technology transfer plans and personnel training facilitate seamless production. Collaboration is key to advancing EV-based therapies and diagnostics efficiently and safely.
There are different downstream processes used for the large-scale production of EVs, such as (i) differential ultracentrifugation with density-gradient centrifugation, (ii) filtration paired with centrifugation and immunoaffinity capture, (iii) Size-Exclusion Chromatography (SEC), (iv) polymer-based precipitation, and (v) microfluidics technologies [33,36,74]. All of these methods have their own advantages and disadvantages (Table 2 and Figure 3). Nowadays, differential ultracentrifugation-based techniques are the gold standard for EV isolation. In this case, EV isolation is based on density and size, with the possibility of sequential separations of large volumes of samples. However, this technique is time-consuming; it requires large starting volumes, characterized by high susceptibility to operator-based variability; and exosomes may be damaged by high-speed centrifugation. Filtration—in particular, ultrafiltration—for EV isolation is an operational, simple, high-throughput, cost- and time-effective separation process. Unlike batch-wise ultracentrifugation, ultrafiltration can be operated in continuous mode, which makes this downstream separation technique suitable for large-scale production. However, potential EV degradation and lysis due to the shear forces and the clogging and trapping of exosomes on to the filter membrane (and therefore exosome loss) are risks that should be considered when applying filtration. High-purity EVs with no disturbances in their structure, integrity, and biological activity are obtained using SEC. This method offers high reproducibility and can be operated in pseudo-continuous mode with a high degree of automation. Microfluidics-based techniques and immune-affinity capture methods based on affinity, size, and density principles give highly specific and pure fraction of exosomes with fast sample processing, high efficiency, and high portability. Most methods have their usefulness depending on the application, but it should be remembered that each of the methods to a greater or lesser extent may not be suitable for large-scale EV purification [74]. After isolation, EVs should be thoroughly characterized by multiple methods to validate the isolation method. For exosomes, the most common characterization methods can be classified as marker-based transmembrane proteins (CD63, CD9, and CD81) or intraluminal proteins (TSG101 and ALIX), biophysical (dynamic light scattering, flow cytometry, nanoparticle tracking analysis, and tunable resistance pulse sensing), or imaging-based (atomic force microscopy and electron microscopy) [36]. EV characterization also includes cargo profiling by characterizing proteins, nucleic acids, lipids, and metabolites, using omics approaches such as proteomics, genomics, and lipidomics to ensure cargo consistency.
Due to the characteristic features of EVs, including low immunogenicity, innate stability, and the ability to cross biological barriers, the usefulness of EVs as therapeutic carriers to deliver various types of cargo molecules to the target cancer cells began to be investigated [75,76,77]. Comparing exosomes to synthetic liposomes having a similar structure of the lipid bilayer, we see that bio-derived exosomes are characterized by an increased ability to load biological molecules and greater efficiency in reaching target cells and delivering therapeutic agents when administered intravenously [78]. When using EVs to deliver active molecules, many approaches have been explored to load EVs with various drugs and other bioactive compounds. In general, there are two different approaches to load cargos into EVs, defined as (i) pre-loading and (ii) post-loading methods [33]. In pre-loading, the parental cells are engineered to produce bioactive molecules that are encapsulated into EVs by the natural sorting process during EV maturation. Thus, EV-producing cells secrete EVs loaded with the bioactive molecule of interest. In post-loading, drugs and bioactive molecules are directly loaded into EVs after EV isolation.
The gold standard for the uploading of exogenous materials into EVs is electroporation, where an electric pulse temporally generates pores in the lipid bilayer, and exogenous cargo can diffuse into the inner space of EVs. Recently, numerous nanomaterials, drugs, and nucleic acids have been electroporated into EVs. Pan et al. (2020) used exosomes isolated from the urine of gastric cancer patients for tumor diagnosis and therapy [79]. In this study, gold nanoparticles coated with polymethacrylates, in conjunction with chlorin-loaded exosomes, were used for cancer imaging and photodynamic therapy [79]. Chemotherapeutic drugs, such as paclitaxel and doxorubicin, have been loaded via electroporation into macrophage-derived EVs [80] and immature dendritic cell-derived EVs [81], respectively. Electroporation is also utilized to directly load nucleic acids into exosomes. Alvarez-Erviti et al. (2011) obtained siRNA-loaded exosomes, which possessed the ability to knock down β-secretase 1 in mouse brains [82], while in murine dendritic cell-derived exosomes loaded with shRNA, they allowed for a decrease in α-synuclein aggregation and alleviated dopaminergic neuron loss in PD models [83]. Sonication is a physical method that uses ultrasounds to weaken the integrity of the EV membrane, which facilitates the loading of exosomal cargo, including drugs [84], proteins [85], and nanomaterials [86]. Extrusion is a physical procedure that involves compressing a mixture of EVs, especially sized 30–150 nm, and cargo in an extruder to induce the recombination of the EV membrane so that the membrane collapses and uniformly mixes with the cargo [36]. Fuhrmann et al. [87] extruded the exosomal samples to load porphyrins. Haney et al. [85] used this strategy and limited cycles of rapid freezing and thawing to produce catalase-loaded exosomes. However, repeated freeze–thaw could inactivate proteins and induce exosome aggregation [34]. The abovementioned physical methods allow EVs to be loaded with a higher efficiency than simple cargo and EVs’ co-incubation. Nevertheless, recent reports demonstrate that many types of cargoes can be successfully loaded into EVs using this simple method. Paclitaxel and doxorubicin have shown enhanced chemotherapeutic effects after being loaded into prostate cancer cell-derived [88] and mesenchymal stromal/stem cell-derived exosomes [89]. Gong et al. applied this strategy to nucleic acids packaging and produced miRNA-loaded macrophage-derived exosomes [89,90]. Gold nanoparticles-labeled exosomes for noninvasive in vivo neuroimaging and tracking were obtained by Batzer et al. [91]. Encapsulation of withaferin A, anthocyanidins, curcumin, paclitaxel, docetaxel, and catalase into EVs has been achieved by incubating those compounds with EVs in the presence of organic solvents, such as ethanol, or detergents, such as saponin, to create pores on the exosomal surface, thus leading to an increase in membrane permeability [85,88,92].
EVs arming can also be achieved via genetic engineering, which involves the modification of the exosome-producing cells to produce exosomes with desired properties or cargoes. One common approach is to modify the EV-producing cells to overexpress a specific protein or peptide of interest, which can then be selectively packaged into the exosomes. This can be achieved by introducing a plasmid or viral vector containing the gene of interest into the cells and selecting for cells that stably express the gene [33,74,93]. Alternatively, CRISPR-Cas9 technology can be used to knock out or knock in specific genes to manipulate the cargo of the exosomes [94,95]. Another approach is to modify the exosome-producing cells to express a chimeric protein that contains a targeting moiety, such as an antibody or peptide, fused to a protein that directs the protein to the exosome membrane [36]. This can result in the specific packaging of the targeting moiety into the exosomes and selective targeting of specific cells or tissues. In addition to cargo modification, genetic engineering can also be used to modify the properties of the exosomes themselves, such as their size, shape, or surface charge. For example, the exosome-producing cells can be modified to overexpress specific membrane proteins or lipids that can affect the properties of the exosomes [34]. Genetic engineering of exosomes can provide a powerful approach for functionalizing exosomes with specific cargoes or properties and for achieving selective targeting of specific cells or tissues. However, it is important to note that genetic modification can affect the safety and efficacy of the exosomes and may require careful evaluation and regulation before their use in clinical applications [28,96].

3. Surface Bioengineered Extracellular Vesicles as Targeted Delivery Systems

To achieve targeted delivery, the surface of EVs can be modified by functionalization with specific ligands or antibodies that bind to specific receptors on the target cells. Here are some common methods for exosome surface functionalization. A basic method is covalent conjugation, which involves attaching specific ligands or antibodies to the surface of exosomes by forming covalent bonds between functional groups on the exosome surface and the ligand or antibody [33,97]. Then comes the non-covalent conjugation, which involves adsorption of the ligands or antibodies onto the surface of exosomes through non-covalent interactions, such as electrostatic interactions or hydrophobic interactions [33,97]. Another method involves lipid membrane modifications resulting in incorporating modified lipids with specific ligands or antibodies into the exosome membrane, which can then be recognized by target cells [36,98]. More recently, genetic engineering was used where genetically modified cells were enabled to produce EVs that express specific ligands or antibodies on the surface of EVs [33,98]. The choice of method for EV surface functionalization depends on the specific application and the properties of the ligands or antibodies being used. Successful functionalization can improve the specificity and efficiency of exosome-based therapeutics, making them promising tools for targeted drug delivery and personalized medicine [33]. Covalent conjugation is a commonly used method to achieve this because it involves the formation of a stable chemical bond between the surface of the exosome and the molecule of interest. The molecule of interest is usually modified with a functional group such as a thiol (-SH), amine (-NH2), or carboxyl (-COOH) group that can react with a complementary functional group on the exosome surface [99]. One common approach for covalent conjugation is to use maleimide chemistry. Maleimide groups react specifically with thiols to form a stable thioether bond. Thus, molecules with thiol groups, such as cysteine residues in proteins, can be conjugated to maleimide-functionalized exosomes [100]. Another approach is to use azide–alkyne cycloaddition or click chemistry, which produces no alterations in exosome size and function [28,98]. Covalent conjugation can be used to target exosomes to specific cells or tissues by attaching targeting moieties such as antibodies or peptides that recognize cell surface receptors [33,97]. Additionally, it can be used to deliver therapeutic molecules such as small interfering RNA (siRNA) or chemotherapeutic drugs to specific cells or tissues [82,101]. However, it is important to note that covalent conjugation can affect the biological properties of exosomes and may alter their biodistribution and clearance in vivo [98]. Therefore, careful characterization and evaluation of the functionalized exosomes is essential before their use in biomedical applications.
Non-covalent conjugation is an alternative method for functionalizing EVs that does not involve the formation of a stable chemical bond between the surface of the exosome and the molecule of interest. Instead, non-covalent conjugation relies on weak, reversible interactions such as electrostatic interactions, hydrophobic interactions, or hydrogen bonding [28,102]. One common approach for non-covalent conjugation is to use the liposome fusion. Note that liposomes are artificial vesicles made of phospholipids that can fuse with the exosome membrane and deliver their cargo into the exosome lumen or surface. The liposomes can be functionalized with various molecules, such as proteins, peptides, or nucleic acids, which can then be transferred to the exosome surface or lumen [97,98]. Another approach is to use a biotin–streptavidin conjugation. Biotinylated molecules can bind to streptavidin-coated exosomes via biotin–streptavidin interactions, which are one of the strongest non-covalent interactions in nature. Streptavidin can also be conjugated to various molecules, such as antibodies, peptides, or nucleic acids, which can then be delivered to the exosome surface or lumen [33,103]. Non-covalent conjugation has several advantages over covalent conjugation. It is a gentle method that does not require harsh chemical reactions, which can affect the biological properties of exosomes. It also allows for reversible binding, which can facilitate the release of the cargo at the target site. However, non-covalent conjugation can also have lower efficiency and specificity than covalent conjugation, and the cargo may dissociate from the exosome under certain conditions. In summary, both covalent and non-covalent conjugation are valuable methods for functionalizing exosomes, and the choice of method depends on the specific application and the properties of the cargo and the exosome [98,104].
EVs subclassed as exosomes are surrounded by a lipid bilayer membrane that contains various lipids and proteins, and modifications of this membrane can be used to functionalize exosomes. One approach for membrane modification is to insert lipids with functional groups or ligands into the exosome membrane. One common type of modified lipid is polyethylene glycol (PEG)-modified lipids, which can improve the stability and circulation time of exosomes in vivo by reducing their clearance by the immune system [76,97]. Another approach is to insert lipids with targeting moieties or ligands, such as antibodies or peptides, into the exosome membrane to achieve selective targeting of specific cells or tissues. This can be achieved by coupling the targeting moiety to a lipid molecule that can be incorporated into the exosome membrane, such as phosphatidylcholine or cholesterol. The targeting moiety can then bind to specific receptors on the target cells or tissues. In addition to lipid modification, the exosome membrane can also be modified by attaching proteins or peptides to the surface of the exosomes. This can be achieved by fusing the protein or peptide to a lipid anchor that can insert into the exosome membrane [33,34,74]. The modification of the exosome membrane can provide a versatile approach for functionalizing exosomes with various molecules and achieving selective targeting of specific cells or tissues. However, it is important to note that membrane modification can affect the biological properties of exosomes and may alter their interactions with cells and tissues. Therefore, careful characterization and evaluation of the modified exosomes is essential before their use in biomedical applications [74,95,98].
Functionalized EVs have shown great potential for both cancer treatment and diagnosis. Recent studied evidenced that functionalized exosomes can be used to deliver therapeutic agents, such as small-molecule drugs [89,105,106], nucleic acids [82,83,101,107,108], or proteins [85], to cancer cells. EVs can also carry a mixture of different components, such as human vascular endothelial growth factor A (VEGF-A) and human bone morphogenetic protein 2 (BMP-2) mRNAs within a customized injectable PEGylated poly (glycerol sebacate) acrylate (PEGS-A) hydrogel for bone tissue regeneration [109]. In addition to their abovementioned uses, cargo EVs can be engineered to express IL6 signal transducer, which has the ability to inhibit the IL6 trans-signaling pathway and thus have anti-inflammatory effects [110]. By modifying the exosome surface with targeting moieties, such as antibodies or peptides, the exosomes can selectively bind to cancer cells and deliver their cargo, resulting in enhanced efficacy and reduced toxicity. Functionalized exosomes can also be used for immunotherapy by delivering antigens or immune modulators to stimulate the immune system against cancer cells (Figure 2) [111]. Note that exosomes can be isolated from various body fluids, including blood, urine, and cerebrospinal fluid, and their cargo can be analyzed to identify biomarkers for cancer diagnosis and monitoring. By modifying the exosome surface with targeting moieties, such as aptamers or antibodies, exosomes can be selectively captured and analyzed to detect cancer-specific biomarkers, such as proteins or nucleic acids [112,113]. Functionalized exosomes can also be used for imaging by loading them with contrast agents or fluorescent dyes to visualize cancer cells and tumors [35,91]. Studies evidence that functionalized exosomes can be used in combination with other cancer therapies, such as chemotherapy or radiation therapy, to enhance their efficacy and reduce their side effects [34,74,84,97,114]. By delivering therapeutic agents to cancer cells and stimulating the immune system, functionalized exosomes can enhance the tumor-killing effect of other cancer therapies and improve patient outcomes.
Due to the possibility of their functionalization, exosomes are increasingly used for targeting for various therapeutic purposes, and thus various possibilities of their applications in cancer therapy and diagnostics are being investigated [24,33,115]. Among other things, EVs are suitable candidates for improving the targeting of chemotherapeutic drugs. At present, chemotherapeutic-loaded EVs are described as internalizing in tumor cells to induce cell death [116], improving the cytotoxic effect of paclitaxel on LNCaP and PC3 cells [88], inhibiting and improving malignant U-87 cells growth in a dose-dependent manner [117], enhancing antiproliferative and anti-inflammatory activity in CFPAC-1 cells [116], and highly efficient targeting to αv-integrin-positive breast cancer cells in vitro and high specific delivery to tumor tissue without overt toxicity in a MDA-MB-231 cancer mouse model [81]. Moreover, nucleic acid-loaded EVs allowed for a reduction of cell proliferation and invasion and an increase in apoptosis and cell-cycle arrest, along with inhibited growth, in xenograft tumors in vivo [101]; the delivery of siRNA molecules and induction/downregulation of TPD52 gene expression in SKBR3 breast cancer cells [107]; and decreased rates of cell migration and proliferation due to exosomes releasing miR26a from selectively bound HepG2 cells [118].

4. Clinical Trials on Extracellular Vesicles

EVs have been used, among other things, in the treatment of cancer [22,62,119,120,121], neurodegenerative diseases [122,123,124], cardiovascular disease [125,126,127], kidney disease [48], lung disease [42], and in regenerative medicine [128]. There are two different uses of EVs. In the first case, the specific biological properties of EVs, including their role in intercellular communication, are used to target a given tissue and reduce pathological signals or mimic the natural repair process. These properties are associated with a large number of proteins and lipids on the surface of the EVs. In the second case, EVs serve as carriers for delivering therapeutic agents to their destinations [129,130]. In recent years, the term “hybrid EV system” or “biohybrid” has been introduced to refer to the formula of nanoscale drug delivery systems consisting of (synthetic) conventional components and EVs. The use of biohybrid systems improves the arrival at the destination and the effectiveness of loading the drug [131].
Clinical studies on EVs subclassed as exosomes in cancer diagnostics have attracted significant attention due to the potential for the early detection and precise identification of various types of cancers. The research focuses on utilizing exosomes to identify cancer-specific biomarkers, providing valuable information for accurate diagnosis. Notable findings include the identification of cancer-specific biomarkers through the isolation and analysis of exosomes from cancer patients [132,133]. Exosomes show promise in regard to enabling the early detection of tumors and facilitating timely intervention and improved prognosis [134]. The potential for differentiating between various types of cancers using exosomal biomarkers is a key area of investigation, aiming to establish panels of biomarkers for improved diagnostic accuracy [135]. Additionally, investigations into the prognostic value of exosomal biomarkers connect specific exosome profiles to disease progression and prognosis, guiding treatment strategies and informing patients about the likely course of their cancer [136,137,138,139]. Research efforts have expanded to explore the clinical utility of exosomal biomarkers across various cancer types, enhancing the applicability of exosome-based diagnostics [136,140]. In the clinical trials.gov database, the most registered research concerns the diagnostic use of exosomes, which is illustrated in Figure 4. Several examples of research at various levels of advancement are provided: terminated trials—NCT04960956 Glycosylation of Exosomes in Prostate and Urothelial Carcinoma and NCT04357717 ExoDx Prostate Evaluation in Prior Negative Prostate Biopsy Setting; completed ones—NCT04394572 Identification of New Diagnostic Protein Markers for Colorectal Cancer and NCT05101655 Construction of Microfluidic Exosome Chip for Diagnosis of Lung Metastasis of Osteosarcoma; and with active recruiting—NCT05854030 A Study on Serum Exosomal miRNA Predicting the Effective and Prognosis of Immunotherapy Combined With Chemotherapy in Pulmonary Squamous Carcinoma.
In the context of cancer gene therapy, clinical studies on exosomes have emerged as a promising path for delivering genetic materials to target cells. Exosomes, with their unique properties, serve as valuable carriers for genetic materials, potentially modulating gene expression and improving therapeutic outcomes. Notable findings include the exploration of exosomes as carriers for delivering genetic materials, protecting and transporting therapeutic cargo to specific target cells [141,142,143]. Clinical research focuses on their potential in delivering RNA-based therapies, such as small interfering RNA (siRNA) or microRNA, for targeted gene silencing in cancer cells [144], with an example from the clinicaltrial.gov base—NCT03608631 Exosomes in Treating Participants With Metastatic Pancreas Cancer With KrasG12D Mutation. Investigations have also explored exosomes as carriers for CRISPR/Cas9 gene-editing tools, allowing for precise genome modifications [145,146]. The potential to enhance therapeutic efficacy by improving the stability and targeted delivery of genetic materials using exosomes is an area of active exploration. Ongoing clinical research is dedicated to translating exosome-based gene therapies from preclinical studies to human trials, evaluating safety and efficacy in cancer patients [147].
In the area of cancer drug delivery, clinical research on exosomes has become a promising field, exploring their potential to transport therapeutic agents to specific cells or tissues. Exosomes, being natural carriers released by cells, offer unique advantages in drug delivery. Research investigates modifying exosomes to carry targeting ligands or antibodies, minimizing side effects, and enhancing the specificity of drug delivery [95,98]. Studies have explored the use of exosomes as carriers for delivering anticancer drugs, protecting therapeutic payloads from degradation, and enhancing their targeted delivery to cancer cells [95]. Clinical research has also entered into modifying exosomes to carry specific targeting ligands or antibodies, allowing for the precise delivery of therapeutic drugs to cancer cells, while minimizing off-target effects [148]. The use of exosomes to deliver drugs with the aim of minimizing side effects is a focus of ongoing research, aiming to enhance the therapeutic index of anticancer medications [95]. Clinical studies have explored loading exosomes with multiple therapeutic agents for co-delivery, aiming to enhance treatment efficacy through synergistic effects [149]. Investigations have also explored exosomes’ potential to overcome multidrug resistance in cancer cells, encapsulating drugs and delivering them to resistant cells [148,150]. The significant promise of clinical research on exosomes in cancer drug delivery lies in the potential to develop more targeted and effective cancer treatments, minimizing side effects, and improving overall therapeutic outcomes, what the research NCT04453046 Hemopurifier Plus Pembrolizumab in Head and Neck Cancer and NCT01294072Study Investigating the Ability of Plant Exosomes to Deliver Curcumin to Normal and Colon Cancer Tissue may prove in the future.
In cancer neurology applications, clinical research on exosomes brings to light the complex interplay between exosomes and the nervous system in the context of cancer-associated neurological conditions. Understanding these mechanisms may open opportunities for diagnostic biomarkers and innovative therapeutic approaches for neurological diseases linked to cancer. Research in this area highlights the active participation of exosomes in mediating communication within the nervous system, playing a role in the transfer of various signaling molecules between different cell types [151]. Studies have focused on understanding the role of exosomes in neurological diseases associated with cancer, such as Alzheimer’s disease and multiple sclerosis [152,153,154]. Investigations explore the potential of exosomal biomarkers in identifying and understanding neurological conditions associated with cancer, offering specific molecular cargo as diagnostic or prognostic indicators [136,152,155]. Clinical research has also explored the therapeutic potential of exosomes in treating neurological disorders associated with cancer, considering the engineering of exosomes to deliver therapeutic agents to affected areas of the nervous system [156]. The role of exosomes in allowing communication between cancer cells in the brain microenvironment during metastasis is also under investigation [157,158,159].
In the context of cancer regulation of inflammatory processes, clinical research on exosomes has brought significant interest due to their involvement in modulating immune responses and inflammation within the tumor microenvironment (NCT01159288 Trial of a Vaccination with Tumor Antigen-loaded Dendritic Cell-derived Exosomes (CSET 1437)). This research highlights the complex interplay between exosomes, inflammation, and cancer progression, offering insights into potential therapeutic strategies for modulating the inflammatory microenvironment in cancer. Studies have investigated how exosomes released by cancer cells can influence immune responses and inflammatory processes, carrying immunomodulatory factors that affect the activity of immune cells [160]. Research focuses on understanding the role of exosomes in promoting inflammation within the tumor microenvironment, carrying pro-inflammatory signals that contribute to a pro-tumorigenic inflammatory milieu [161,162]. Investigations explore how exosomes contribute to the activation or suppression of inflammatory signaling pathways, influencing the overall inflammatory state [161]. Clinical research has also undertaken the potential anti-inflammatory properties of exosomes derived from certain cell types, carrying factors that contribute to resolving inflammation within the tumor microenvironment or modulating immune responses toward an anti-tumorigenic state [163,164]. Ongoing studies explore the possibility of targeting exosomes as a therapeutic strategy to control inflammation in cancer, modulating their release or content to regulate the inflammatory microenvironment within tumors [165].
The EU Clinical Trials Register (https://www.clinicaltrialsregister.eu/ctr-search/search, accessed on 12 January 2024) includes 11 studies, at various levels of advancement, in which exosomes were also taken into account. However, exosomes are not the main subject of interest in these studies. They are isolated and characterized as an additional diagnostic element in the course of effectiveness and safety tests in the treatment of various cancers, mainly prostate (EudraCT Numbers 2018-004458-14, 2015-000270-36, and 2015-001361-27), pancreas (EudraCT Numbers 2015-004860-12 and 2017-003621-15), and kidney (EudraCT Numbers 2011-006009-85 and 2018-001201-93). The information provided indicates a substantial presence of exosome-related clinical trials on https://clinicaltrials.gov/. Out of the 146 registered trials, 66 are interventional, and 80 are observational. Phase 2 trials are the most numerous, with 21 trials, followed by phase 1 trials with 15. Additionally, there are two studies in early phase 1 and phase 3 each. Notably, 31 studies have the status “not applicable,” and a significant portion of 76 studies focuses on examining exosomes for their utility as diagnostic and prognostic markers in cancer. This suggests a diverse and active exploration of exosomes in various phases and contexts within the realm of clinical research, particularly emphasizing their potential diagnostic and prognostic applications in cancer (see Figure 4). Most studies report lung cancer, mainly NSCLC, which accounts for 19% of all described trials. Subsequently, prostate cancer occurs in 14.23% of cases; pancreatic cancer in 8.8%; and breast cancer, primarily with triple-negative characteristics, constitutes 7.5% of cases. Gastric cancer occurs in 5.4%, and head and neck cancer in 4.8% of cases. Other types of cancers collectively make up 40.2% of all cases.
A phase I trial examined the safety and efficacy of exosome-based delivery of paclitaxel (exoPTX) in patients. The findings revealed that exoPTX significantly reduced tumor size with fewer side effects compared to traditional paclitaxel treatment. The overall response rate (ORR) was higher in the exosome group, and patients experienced less severe neuropathy and gastrointestinal toxicity, highlighting the potential of exosome-mediated drug delivery [148]. A pilot study explored exosomes loaded with KRAS G12D siRNA in patients. The treatment effectively silenced the KRAS mutation, resulting in tumor shrinkage in a subset of patients. This approach showed a higher disease control rate compared to conventional chemotherapy, underscoring the potential of exosome-based RNA interference (RNAi) therapies [166]. EV-based therapies have also been investigated for their potential in treating neurological disorders. A clinical trial investigated the use of MSC-derived exosomes in patients with ischemic stroke. The study found that exosome-treated patients had significantly better neurological recovery, as measured by the National Institutes of Health Stroke Scale (NIHSS) and modified Rankin Scale (mRS), compared to those receiving standard post-stroke care. The researchers attributed these improvements to reduced brain inflammation and enhanced neurogenesis [167]. EV-based therapies generally exhibit a superior safety profile compared to traditional treatments. Their natural origin and ability to specifically target disease sites minimize off-target effects and systemic toxicity. For instance, the reduced side effects observed with exosome-mediated paclitaxel delivery underscore this advantage. Exosomes offer enhanced targeted delivery mechanisms, which improve the efficacy of therapeutic agents. This is particularly evident in cancer therapy, where exosomes can deliver chemotherapeutic drugs or genetic material directly to tumor cells. This targeted approach not only improves treatment outcomes but also reduces adverse effects compared to conventional chemotherapy. In regenerative medicine, exosome therapy often outperforms traditional methods by promoting cellular repair and modulating immune responses more effectively. Clinical trials have shown that exosome treatments can achieve better functional recovery and organ repair compared to existing therapies, highlighting their regenerative potential.
EV-based therapies bring promise to healthcare but also raise ethical concerns and risks, particularly concerning patient consent and data privacy, safety, and equitable access. Addressing ethical issues and mitigating potential risks associated with EV-based therapies requires a multifaceted approach involving informed-consent processes, robust data-privacy measures, regulatory compliance, transparent communication, and ethical oversight. By prioritizing patient autonomy, confidentiality, and safety, stakeholders can uphold ethical principles and promote trust in the development and implementation of EV-based therapies.

5. Regulatory Affairs of Extracellular Vesicles

In light of business assessments, EVs have great potential for commercialization both toward new diagnostic and therapeutic products. Exosomes’ global market size was estimated to be ca. > USD 250 million in 2022, reaching ca. USD 3.2 billion in 2032 (CARG 2023-2032 29.9%) [168]. Interestingly, the value of the cancer segment for this market is estimated at being over USD 1.2 billion in 2032. Note that EVs-based diagnostics and therapies hold immense market potential and economic impact, driven by advances in biotechnology, personalized medicine, and regenerative medicine. Despite regulatory challenges and market competition, EV-based products are poised to revolutionize healthcare delivery, improve patient outcomes, and stimulate economic growth in the biotechnology sector. Continued investment in research, development, and commercialization efforts is essential to realize the full potential of EV-based technologies and address unmet medical needs effectively.
As it was presented, EVs constitute a promising therapeutic option in many disease areas, including cancer; nevertheless, obtaining the marketing authorization for EV-based medicines in the EU would be a complex process and could be challenging, as currently there are no specific European guidelines for such products. More recently, the U.S. Food and Drug Administration (FDA) announced a work to be done on new guidelines addressing the manufacturing of EV-based products, specifically in regard to mesenchymal stromal cell (MSC)-derived EVs [169]. In a recent announcement, the U.S. FDA also postulated to overcome some limitations related to large-scale EV production and to provide industrial guidelines to meet the demand for EVs to be involved for therapeutic produces [170]. It is a general consensus that the EV-based products fall within the definition of a medicinal product [171]: “Any substance or combination of substances presented for treating or preventing disease in human beings. Any substance or combination of substances which may be administered to human beings with a view to making a medical diagnosis or to restoring, correcting or modifying physiological functions in human beings is likewise considered a medicinal product” (Directive 2001/83/ec) [172]. Also, they are generally considered biological medicinal products, following the definition of the European Medicines Agency (EMA), that this is a medicine whose active substance is made by a living organism [173]. Nevertheless, depending on their manufacturing process, the molecules they carry, and the intended therapeutic use, the EVs may also be classified as biotechnological products or advanced therapy medicinal products (ATMPs).
Currently, a majority of EV-based products being developed or used in clinical trials come from native cells (e.g., mesenchymal stromal stem cells, dendritic cells, and genetically unchanged cells). If the cells carry active substances not intended to introduce any gene modification in the patient’s body, they would most possibly be classified as biological products, even if the methods of priming or stimulation of the naïve cells were applied to produce the EVs. If bioengineering methods are used to obtain the cells producing the EVs (e.g., the cells are immortalized or other genetic modifications are made), but the molecules carried by the vesicles do not fall under the definition of ATMP, the products should most probably be classified as biotechnological products. On the other hand, if the intended therapeutic purpose of the EV-based medicine involves the modification in the recipients’ genome, the product should be classified as a gene therapy medicinal product (GTMP, being part of ATMP; Figure 2) [171,174].
To better understand the differentiation between ATMP and non-ATMP, the scientific recommendations on the classification of such products published by the EMA Committee for Advanced Therapies (CAT) can be looked at. In the list of products published in February 2024, there are, e.g., two medicinal products which contain EVs:
  • Human umbilical cord mesenchymal stem cells derived, PEGylated exosomes carrying recombinant hTERT mRNA and proteins;
  • Conditioned medium (secretome) from expanded donor bone-marrow mesenchymal stem cells containing cytokines, growth factors, proteins, and extracellular vesicles.
The first one is classified as GTMP, based on the facts that the active substance is a recombinant nucleic acid administered to human beings with a view to regulating a genetic sequence and that its therapeutic effect relates directly to the recombinant nucleic acid sequences it contains. The second product is classified as not ATMP on the basis that it does not contain or consist of recombinant nucleic acids, nor does it contain cells or tissues (Scientific Recommendations on Classification of ATMP, 2024) [175].
In conclusion, in the case of bioengineered EVs, the most appropriate classification would be a biotechnological product or ATMP, depending on its mechanism of action. The issue of classification presented above is important from the point of view of the scope of developmental studies to be performed. It also has implications on the procedural levels, i.e., if the marketing authorization application should be proceeded at the centralized level, by the EMA, or could be proceeded locally. In the case that the product is classified as biologic, the registration can proceed nationally in the Member State selected or through the decentralized procedure in some selected Member States. If it is a biotechnological product or the GTMP, the MA application should be filed to the EMA.
All in all, regardless of the registration procedure selected, the applicant for the MA for the medicinal product in the European Union is required to demonstrate the appropriate quality of the product applied for, its safety for the targeted group of patients, and its efficacy in the proposed indication(s). These are done through the elaborated set of quality and non-clinical studies, and clinical trials, the scope of which is strictly related with the specificity of the medicinal product applied for. This is particularly applicable to such complex products as bioengineered EVs. It should be carefully defined on a case-by-case basis, keeping in mind the provenance of the EVs (“producer cells”), the techniques applied in their manufacturing, the molecules they carry (either in the vesicles or on their surface), and, last but not least, the planned therapeutic action.
As already mentioned, currently there are no specific guidelines issued by the EMA for EV-based medicinal products. Since these products are biologicals, it has been postulated that following the current set of EMA guidelines for biological medicinal products would be sufficient to assure appropriate quality of the EV-based medicinal product (Supplementary Table S1) [173]. There are also guidelines and recommendations prepared by the scientific community involved in the field, e.g., Minimum Information for Studies of Extracellular Vesicles (MISEV) [1], or guidelines prepared by Extracellular Vesicle Translation to Clinical Perspectives—EVOLVE France [171]. The quality documentation should contain such general information as the nomenclature, structure, and general properties of EV-based active substances defined early as active pharmaceutical ingredients [176]. The quality aspects which should be taken into account involve, e.g., establishing a repeatable, validated manufacturing process; the selection of appropriate process controls; and the definition of release criteria, especially for quality, purity and defining chemical and biological impurities (including microbial and viral contamination or TSE agents), potency, and process evaluation and validation. The specifications should address the traits of the origin cells, as well as possible modifications of the EVs and their contents [173,174,177]. The validation of the analytical procedure used for quality control (QC), as well as EV batch analysis, including justification and specification methods used, is a critical step in QS management for EV-based ATMPs. To date, there is no specific information on EV references or EV standards used for specific EV subclasses. For the small- and medium-sized EVs, it would be recommended to also take into account general requirements for nanomedicines (https://www.ema.europa.eu/en/human-regulatory-overview/research-and-development/scientific-guidelines/multidisciplinary-guidelines/multidisciplinary-nanomedicines, accessed on 6 March 2024). Specific care should be taken for stability studies addressing EV-based investigational medicinal products (EV-based IMPs). Note that pristine EVs could behave in different way as compared to bioengineered EVs composed of different added moieties such as genes; proteins; and others, e.g., surface-decorating agents [178]. Therefore, the storage, description, and composition of EV-based IMPs matter, including the pharmaceutical and biotechnological development and the description of their manufacture based on GMP standards, batch formulae, control of excipients—especially of human or animal origin—if any are used in the formulation, and finally activities related to facilities and equipment used in the manufacturing process and applied contained closure system. It should be emphasized that the validation of the analytical procedure for all methods and tests used in a quality-control system will be required in laboratory-based standards and certifications. This should be especially considered for all novel methods and assays developed for EV characterization [179].
In general, the application of the risk-based approach, as per the “Guideline on quality, non-clinical and clinical requirements for investigational advanced therapy medicinal products in clinical trials—Scientific guideline” (EMA/CAT/852602/2018) [180] and “ICH guideline Q9 on quality risk management” (EMA/CHMP/ICH/24235/2006) [181], is recommended when working on the manufacturing process and controls. To date, nonclinical documentation for EV-based medicinal products, as required in the Common Technical Document (CTD) [182,183], should provide information on non-clinical models, the general outline of the non-clinical development, and the timing of non-clinical studies (Figure 5).
It is obvious that the non-clinical development pathway for EV-based medicinal products, especially EV-based ATMPs, is significantly different for other medicinal products based on simple chemical agents. In general, the non-clinical data should provide information on the efficacy and safety of the biological active dose; therefore, in vivo animal studies should be carefully planned to ensure the generation of robust data, while considering the 3R (reduction, replacement, and refinement) rules. The use of 3D cultures, especially advanced organoids and lab-on-chip systems, should be involved in such studies [184]. In preclinical pharmacology studies, the dose level for proof of concept should allow for the estimation of the biologically effective dose (BED) to help establish a clinical effecting dose (CED). It is also important to establish the BED with an acceptably safety level of the EV-based product. Therefore, toxicology studies supported with pharmacokinetic profiling should be applied as minimum non-clinical data requirements before first-in-human studies. In non-clinical safety studies, such as systemic toxicity, a primary focus before clinical setups, the need for additional toxicity studies, such as genotoxicity, tumorigenicity, reproductive, and developmental toxicity, as well as immunotoxicity studies, should be determined on a case-by-case basis, considering the risk of the particular class of EV-based products in clinical scenarios. Therefore, general toxicity studies should provide information for the estimation of the safe starting dose and dosing regimen and identify relevant safety concerns. Standard genotoxicity assays are generally not appropriated for EV-based ATMPs; however, the specific toxicological concerns, such as genome modifications due to insertional mutagenesis, should be taken into considerations. It seems that standard lifetime rodent carcinogenicity studies will not be required; however, the tumorigenic and teratogenic potential should be tested using in vitro models for neoplasm signals, oncogenic activation, or the cell proliferation index. It is generally accepted that immune responses of the innate and adaptive systems should be examined in different EV-treated models. In general, repeated-toxicity studies will support multiple-dose regiments in human subjects. This will allow us to collect the data for long-term adverse effects, if any.
Importantly, EMA strongly recommends discussing the development plan of all biological medicinal products with the agency through the procedure of Scientific Advice. Having in mind the complexity of the EV-based therapeutic products and the room for interpretation of the currently available guidelines, it would be reasonable to enter into discussion with the authorities early in the development.
The range of required non-clinical studies depends on the intended purpose of the therapeutic EV and its classification. In general, the recommendations of “Strategies to identify and mitigate risks for first-in-human and early clinical trials with investigational medicinal products—Scientific guideline” (EMEA/CHMP/SWP/28367/07) [185] should be followed; however, in case of products to be used, e.g., in cancer, the “ICH S9 guideline on non-clinical evaluation for anticancer pharmaceuticals” (EMA/CHMP/ICH/646107/2008) [186] would be applicable. In the case of EV-based products, which will be classified as biotechnological, the “ICH S6 (R1) guideline on preclinical safety evaluation of biotechnology-derived pharmaceuticals” (CPMP/ICH/302/95) [187] should be applied. If the developed product falls under the definition of the GTMP, the applicable guidelines on quality, non-clinical and clinical requirements for advanced-therapy medicinal products should also be taken into account (Supplementary Table S1). Importantly, the product tested in non-clinical studies, especially the toxicology studies, should be representative of that, which is intended to be used in clinical trials. If possible, it is recommended to use the formulation of the product, which is similar to that intended for human subjects. As an exception rather than the rule, in some early pharmacokinetic (PK) or pharmacodynamic (PD) studies, a more preliminary formulation of the product can be used. The non-clinical studies should bring convincing proof that the developed medicinal product is safe to be administered to human subjects (safety pharmacology and toxicology studies) and that it likely to exert the intended therapeutic effect (PK and PD studies). Currently, a majority of EV-based products are of allogenic origin; therefore, the assessment of the immunogenic potential of the products is of particular importance. The pivotal safety studies (safety pharmacology, toxicology, and immunogenicity) are required to be performed in the Good Laboratory Practice (GLP) certified facility(-ies). The extent of the PD studies to be performed depends on the planned therapeutic indication. If possible, the demonstration of the proof of concept with the use of a suitable animal model of the disease should be executed. These studies do not need to be performed in a GLP-certified facility; however, the compliance with GLP recommendations would be highly beneficial.
In terms of clinical development, general principles that apply to any investigational medical product established in the “Regulation (EU) No 536/2014 of the European Parliament and of the Council of 16 April 2014 on clinical trials on medicinal products for human use, and repealing Directive 2001/20/EC Text with EEA relevance” (Regulation 536/2014) [188] should be considered. The scope of the clinical development program would depend on the planned therapeutic use of the EV-based product. Given the complex nature of the products, the scientific community recommends to follow the guidelines for investigational advanced therapy medicinal products, as per the “Guideline on quality, non-clinical and clinical requirements for investigational advanced therapy medicinal products in clinical trials—Scientific guideline” (EMA/CAT/852602/2018) [180], especially at the early stages of the development. In later stages (phase II/III), the target disease-specific guidelines should be followed, which can be found on the EMA website. Generally, the aim of the clinical program would be to determine the safety and efficacy of the developed product, the dosing regimen recommended for therapy, the spectrum of possible adverse events, and the limitations of the products use (e.g., in pregnancy, while breast-feeding). For exploratory clinical trials, especially for the First-in-Human (FIH) studies, the primary objective should be focused on safety and the tolerability effects. Therefore, the design of the exploratory clinical studies of EV-based ATMPs should involve extended or permanent adverse effects, e.g., immunosuppressive or immunostimulatory effects, integration into the genome, or even malignant transformation. Since the EVs are derived from human cells/tissues and are mostly composed of naturally occurring substances, their safety is considered rather high. Nevertheless, as it was mentioned before that the potential for immunogenicity should definitely be examined (in case of allogenic provenance of EVs). Of note, given the biologic source of the EVs, there might be limitations of current techniques for their tracking and imaging within cells and tissues. Therefore, the determination of their pharmacokinetics and biodistribution may be challenging. For data purposes, the labeling of EVs using different dyes should be used to study cellular uptakes (in vitro) or in vivo biodistribution using specific dyes and modalities [180].
The legal requirements and principle regulatory guidelines to be taken into account during the development; the non-clinical studies and clinical trials of EV-based medicinal products, as per the scientific community recommendations [171,173,174]; and the authors’ expertise are described in Supplementary Table S1.
The applicable guidelines may differ depending on the specific characteristics of the developed product, and the scope should be adjusted on a case-by-case basis. Also, the list is not exhaustive; as each product has its individual particulars, other specific guidelines may also apply. As presented in the text, for bioengineered EVs, the guidelines for biotechnological products and, in some cases, ATMP are generally applicable. The ATMP guidelines, which may be followed but are not mandatory for non-ATMP EV-based products are marked with an asterisk (*). Nevertheless, it is generally recommended by the scientific community to take into account the recommendations for ATMPs in the development of EVs, even if the product is not strictly classified as an ATMP. The scope of clinical guidelines depends on the planned indication of the product. In Supplementary Table S1, guidelines addressing general considerations for clinical trials are presented, plus the guidelines related to potential cancer therapy. The design of clinical studies for EV-based medicinal products, including EV-based ATMPs, deserves specific considerations. Because the data from non-clinical, pharmaco-dynamic, pharmacokinetic, and toxicity studies may be limited to the specific human situation and diseases, this may hamper, among other things, the prediction of a safety starting dose for FIH trials (EMEA/CHMP/SWP/28367/07 Rev. 1). Therefore, the risk assessment in pre- and post-marketing surveillance should take into consideration for numerous factors collected in non-clinical and clinical phases. It seems reasonable to conclude that the pharmacovigilance program is optional for all new EV-based medical products released on global markets, and this could be encompassed by the Marketing Authorization Holder in a dossier presenting the strategy for post-marketing activities [189,190]. This strategy requires us to develop a Pharmacovigilance System Mater File (FSMF) and the Risk Management Plan (RMP). The legal basis for regulatory approval in the European Union to address numerous regulations is presented in Supplementary Table S1. The EV-based medicinal products provide new possibilities for restoring, correcting, or modifying cellular and molecular functions, or making some novel diagnosis. At the same time, because of their novelty, complexity, and specificity, they may produce new risks to patients. The risks for EVs identified in non-clinical and clinical studies should be encompassed within risk management plans, connecting both risk communication and risk-mitigation systems [189,191]. This may include risks to patients in relation to the quality, safety, storage, and distribution of the EV product; risks to patients in relation to specific diseases and mode of treatments; and risks to patients in relation to the environment, especially when EV-based GMO products are developed. Therefore, more regulatory standards are still required to identify such risks and provide a global regulatory platform for the effective mitigation of their adverse effects to patients. In other words, the detection of the risk due to EV-based medicinal products should start early in non-clinical and quality studies before FIH, and novel strategies and technologies, such as safety-by-design methods, lab-on-a-chip, and even artificial intelligence, should be applied in future research and development programs following some regulatory guidelines.

6. Future Perspectives

Exosomes, with their unique properties and potential applications, have garnered interest not only in biomedicine but also in related fields, such as bioengineering, materials science, and pharmacology. Bioengineers explore methods to modify exosomes for targeted drug delivery, improved stability, and controlled release. Techniques like surface functionalization, lipid modification, and encapsulation enable the customization of exosomes with specific properties for biomedical uses. Exosomes inspire the design of biomimetic nanoparticles, mimicking their structure and function for drug delivery and tissue engineering. Synthetic vesicles and liposomes developed by bioengineers resemble exosomes’ membrane composition and surface markers, enhancing biocompatibility and targeting. Materials scientists design and optimize biomaterial-based platforms for the efficient isolation and purification of exosomes from complex biological samples. Nanomaterials, microfluidic devices, and functionalized surfaces offer precise control over exosome capture and separation for downstream applications. Researchers study interactions between exosomes and various materials, like nanoparticles, polymers, and scaffolds, to understand their effect on exosome stability, cargo delivery, and therapeutic efficacy. Material properties such as surface chemistry, topography, and mechanical properties influence exosome binding, uptake, and intracellular trafficking in recipient cells. Pharmacologists explore exosomes as natural drug delivery vehicles for therapeutic compounds like small molecules, nucleic acids, and biologics. Exosome-mediated drug delivery offers benefits such as enhanced bioavailability, prolonged circulation, and targeted delivery to specific tissues or cells. Pharmacologists investigate exosome-mediated intercellular communication’s role in modulating physiological processes, disease progression, and drug response. Exosomes act as carriers of signaling molecules, cytokines, and growth factors, influencing cellular signaling pathways and pharmacological responses in recipient cells.
Extracellular vesicles as membrane-bound particles secreted by cells that hold significant promise in shaping the future of diagnostics and ATMPs. As our comprehension of EVs deepens, their potential to revolutionize these fields becomes increasingly evident. In the field of diagnostics, EVs serve as carriers of a diverse molecular cargo reflective of their cell of origin, enclosing proteins, nucleic acids, and lipids. This inherent characteristic renders them invaluable biomarkers for a spectrum of diseases. Emerging diagnostic methodologies involve isolating EVs from bodily fluids like blood or urine and scrutinizing their cargo for disease-specific signatures. Diseases, ranging from cancer to neurodegenerative disorders and infectious diseases, stand to benefit from such EV-based diagnostic approaches. By facilitating earlier disease detection, enabling personalized medicine strategies, and facilitating real-time treatment response monitoring, EV-based diagnostics promise to usher in a new era of healthcare (Figure 2).
Simultaneously, EVs hold immense potential in the domain of ATMPs. Compared to traditional gene-based or cell-based therapies, the EV medicinal products present a compelling alternative owing to their diminished immunogenicity, smaller size, and remarkable ability to traverse biological barriers, including the formidable blood–brain barrier. Furthermore, EVs can be tailored to carry therapeutic payloads such as drugs, nucleic acids, or proteins with precision to target cells. Future ATMPs may render EVs useful as versatile drug delivery vehicles to address innumerable conditions, including cancer, neurological disorders, and inflammatory diseases. By offering safer, more efficacious treatment options with reduced side effects compared with conventional therapies, EV-based therapies hold the promise of significantly enhancing patient outcomes.
Despite these prospects, the full realization of EVs’ clinical potential necessitates concerted efforts in further research and development. More studies should be performed on standardization methods used for upstream and downstream processes in EV production based on large-scale manufacturing. Quality, safety, and efficacy programs need a global regulatory approval, especially for those EVs used as advanced therapy medicinal products in rare diseases, cancers, and regenerative medicines. For preclinical and clinical safety studies, both toxicology and risk management should be defined for EV-based medicines. Challenges representative of scalability, standardization of methodologies, and regulatory approval frameworks must be diligently addressed to facilitate widespread adoption in clinical settings. Nonetheless, the paths of EVs in diagnostics and ATMPs promise transformative changes in healthcare, characterized by early disease interception, tailored therapeutic interventions, and overall improved patient well-being. Future research directions and emerging trends in extracellular vesicle (EV) research offer promising avenues for expanding our understanding of EV biology and harnessing EVs’ therapeutic and diagnostic potential. Precision medicine holds promise in exploring novel EV biomarkers for early disease detection and treatment-response prediction, while personalized EV-based therapies tailored to individual patient profiles could revolutionize treatment strategies. Advanced technologies, such as engineering EVs with customizable cargo and targeting ligands and improving imaging and detection techniques for real-time visualization and tracking of EVs, offer exciting opportunities for innovation. Bioproduction and manufacturing optimization are essential for scaling up EV production for clinical-grade applications, while standardized protocols and quality-control measures ensure product consistency and safety. Clinical translation and commercialization efforts, supported by collaboration between academia, industry, and regulatory bodies, are critical for bringing EV-based diagnostics and therapies to patients. Effective collaboration fosters a synergistic ecosystem that accelerates innovation and improves patient outcomes in diverse fields of healthcare.

7. Conclusions

Extracellular vesicles emerge as promising candidates in shaping the future of diagnostics and ATMPs. Clinical research on EVs represents a multidimensional approach with profound implications for cancer diagnostics, gene therapy, drug delivery, neurology applications, and the regulation of inflammatory processes. The cumulative knowledge from these studies holds the potential to initiate a new era of diagnostic and therapeutic strategies, offering innovative solutions in the treatment of diverse conditions associated with cancer. Their ability to serve as carriers of diverse molecular cargo reflective of their cell of origin positions them as invaluable biomarkers for a wide array of diseases, offering opportunities for earlier disease detection and personalized medicine approaches. Moreover, EVs exhibit significant potential in the area of ATMPs, presenting an alternative to traditional gene-based and cell-based therapies due to their advantageous properties, such as reduced immunogenicity and the ability to cross biological barriers. Despite these promising prospects, addressing challenges related to scalability, standardization, and regulatory approval is imperative to realize the full clinical potential of EVs. Nevertheless, the path of EVs in diagnostics and ATMPs suggests transformative changes in healthcare, including early disease interception, tailored therapeutic interventions, and improved patient outcomes.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms25126533/s1.

Author Contributions

Conceptualization, A.S. and I.P.G.; methodology, A.S.; software, A.S.; validation, A.S., M.B.-K. and I.P.G.; formal analysis, A.S.; investigation, A.S.; resources, A.S.; data curation, A.S.; writing—original draft preparation, A.S., M.B.-K., E.R.-P., M.D., M.R.C. and I.P.G.; writing—review and editing, A.S., M.B.-K., E.R.-P., M.D., M.R.C., I.R.-M. and I.P.G.; visualization, A.S.; supervision, I.P.G.; project administration, I.P.G.; funding acquisition, A.S. and I.P.G. All authors have read and agreed to the published version of the manuscript.

Funding

The work was financially supported by the Programme “Applied Research” through TEPCAN project granted under the Norwegian Financial Mechanisms 2014–2021/POLNOR 2019 (EEA and Norway Grants), Thematic areas: Welfare, health and care (NCBR Funding No. NOR/POLNOR/TEPCAN/0057/2019-00).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figure 1, Figure 2, Figure 3 and Figure 5 were created with BioRenderTM (license agreement).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Welsh, J.A.; Goberdhan, D.C.I.; O’Driscoll, L.; Buzas, E.I.; Blenkiron, C.; Bussolati, B.; Cai, H.; Di Vizio, D.; Driedonks, T.A.P.; Erdbrügger, U.; et al. Minimal information for studies of extracellular vesicles (MISEV2023): From basic to advanced approaches. J. Extracell. Vesicles 2024, 13, e12404. [Google Scholar] [CrossRef]
  2. Feng, J.; Xiu, Q.; Huang, Y.; Troyer, Z.; Li, B.; Zheng, L. Plant-Derived Vesicle-Like Nanoparticles as Promising Biotherapeutic Tools: Present and Future. Adv. Mater. 2023, 35, e2207826. [Google Scholar] [CrossRef] [PubMed]
  3. Karamanidou, T.; Tsouknidas, A. Plant-Derived Extracellular Vesicles as Therapeutic Nanocarriers. Int. J. Mol. Sci. 2021, 23, 191. [Google Scholar] [CrossRef] [PubMed]
  4. Lian, M.Q.; Chng, W.H.; Liang, J.; Yeo, H.Q.; Lee, C.K.; Belaid, M.; Tollemeto, M.; Wacker, M.G.; Czarny, B.; Pastorin, G. Plant-derived extracellular vesicles: Recent advancements and current challenges on their use for biomedical applications. J. Extracell. Vesicles 2022, 11, e12283. [Google Scholar] [CrossRef]
  5. Briaud, P.; Carroll, R.K. Extracellular Vesicle Biogenesis and Functions in Gram-Positive Bacteria. Infect. Immun. 2020, 88, e00433-20. [Google Scholar] [CrossRef]
  6. Brown, L.; Wolf, J.M.; Prados-Rosales, R.; Casadevall, A. Through the wall: Extracellular vesicles in Gram-positive bacteria, mycobacteria and fungi. Nat. Rev. Microbiol. 2015, 13, 620–630. [Google Scholar] [CrossRef]
  7. Dean, S.N.; Thakur, M.; Spangler, J.R. Extracellular vesicle production in Gram-positive bacteria. Microb. Biotechnol. 2022, 15, 1055–1057. [Google Scholar] [CrossRef]
  8. Mehaffy, C.; Ryan, J.M.; Kruh-Garcia, N.A.; Dobos, K.M. Extracellular Vesicles in Mycobacteria and Tuberculosis. Front. Cell. Infect. Microbiol. 2022, 12, 912831. [Google Scholar] [CrossRef]
  9. Liebana-Jordan, M.; Brotons, B.; Falcon-Perez, J.M.; Gonzalez, E. Extracellular Vesicles in the Fungi Kingdom. Int. J. Mol. Sci. 2021, 22, 7221. [Google Scholar] [CrossRef]
  10. Jackson, K.K.; Mata, C.; Marcus, R.K. A rapid capillary-channeled polymer (C-CP) fiber spin-down tip approach for the isolation of plant-derived extracellular vesicles (PDEVs) from 20 common fruit and vegetable sources. Talanta 2023, 252, 123779. [Google Scholar] [CrossRef]
  11. Du, S.; Guan, Y.; Xie, A.; Yan, Z.; Gao, S.; Li, W.; Rao, L.; Chen, X.; Chen, T. Extracellular vesicles: A rising star for therapeutics and drug delivery. J. Nanobiotechnology 2023, 21, 231. [Google Scholar] [CrossRef] [PubMed]
  12. Baranyai, T.; Herczeg, K.; Onódi, Z.; Voszka, I.; Módos, K.; Marton, N.; Nagy, G.; Mäger, I.; Wood, M.J.; El Andaloussi, S.; et al. Isolation of Exosomes from Blood Plasma: Qualitative and Quantitative Comparison of Ultracentrifugation and Size Exclusion Chromatography Methods. PLoS ONE 2015, 10, e0145686. [Google Scholar] [CrossRef] [PubMed]
  13. Huber, J.; Longaker, M.T.; Quarto, N. Circulating and extracellular vesicle-derived microRNAs as biomarkers in bone-related diseases. Front. Endocrinol. 2023, 14, 1168898. [Google Scholar] [CrossRef] [PubMed]
  14. Teng, F.; Fussenegger, M. Shedding Light on Extracellular Vesicle Biogenesis and Bioengineering. Adv. Sci. 2020, 8, 2003505. [Google Scholar] [CrossRef] [PubMed]
  15. Yáñez-Mó, M.; Siljander, P.R.; Andreu, Z.; Zavec, A.B.; Borràs, F.E.; Buzas, E.I.; Buzas, K.; Casal, E.; Cappello, F.; Carvalho, J.; et al. Biological properties of extracellular vesicles and their physiological functions. J. Extracell. Vesicles 2015, 4, 27066. [Google Scholar] [CrossRef] [PubMed]
  16. Bian, X.; Xiao, Y.T.; Wu, T.; Yao, M.; Du, L.; Ren, S.; Wang, J. Microvesicles and chemokines in tumor microenvironment: Mediators of intercellular communications in tumor progression. Mol. Cancer 2019, 18, 50. [Google Scholar] [CrossRef] [PubMed]
  17. Akers, J.C.; Gonda, D.; Kim, R.; Carter, B.S.; Chen, C.C. Biogenesis of extracellular vesicles (EV): Exosomes, microvesicles, retrovirus-like vesicles, and apoptotic bodies. J. Neurooncol. 2013, 113, 1–11. [Google Scholar] [CrossRef] [PubMed]
  18. Crescitelli, R.; Lässer, C.; Szabó, T.G.; Kittel, A.; Eldh, M.; Dianzani, I.; Buzás, E.I.; Lötvall, J. Distinct RNA profiles in subpopulations of extracellular vesicles: Apoptotic bodies, microvesicles and exosomes. J. Extracell. Vesicles 2013, 2, 20677. [Google Scholar] [CrossRef] [PubMed]
  19. Turiák, L.; Misják, P.; Szabó, T.G.; Aradi, B.; Pálóczi, K.; Ozohanics, O.; Drahos, L.; Kittel, A.; Falus, A.; Buzás, E.I.; et al. Proteomic characterization of thymocyte-derived microvesicles and apoptotic bodies in BALB/c mice. J. Proteom. 2011, 74, 2025–2033. [Google Scholar] [CrossRef]
  20. Wu, D.; Ingram, A.; Lahti, J.H.; Mazza, B.; Grenet, J.; Kapoor, A.; Liu, L.; Kidd, V.J.; Tang, D. Apoptotic release of histones from nucleosomes. J. Biol. Chem. 2002, 277, 12001–12008. [Google Scholar] [CrossRef]
  21. Gurung, S.; Perocheau, D.; Touramanidou, L.; Baruteau, J. The exosome journey: From biogenesis to uptake and intracellular signalling. Cell Commun. Signal. 2021, 19, 47. [Google Scholar] [CrossRef]
  22. Urabe, F.; Kosaka, N.; Ito, K.; Kimura, T.; Egawa, S.; Ochiya, T. Extracellular vesicles as biomarkers and therapeutic targets for cancer. Am. J. Physiol. Cell Physiol. 2020, 318, C29–C39. [Google Scholar] [CrossRef] [PubMed]
  23. Yoshioka, Y.; Konishi, Y.; Kosaka, N.; Katsuda, T.; Kato, T.; Ochiya, T. Comparative marker analysis of extracellular vesicles in different human cancer types. J. Extracell. Vesicles 2013, 2, 20424. [Google Scholar] [CrossRef] [PubMed]
  24. Song, Y.; Kim, Y.; Ha, S.; Sheller-Miller, S.; Yoo, J.; Choi, C.; Park, C.H. The emerging role of exosomes as novel therapeutics: Biology, technologies, clinical applications, and the next. Am. J. Reprod. Immunol. 2021, 85, e13329. [Google Scholar] [CrossRef] [PubMed]
  25. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Thery, C.; Raposo, G. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126, 5553–5565. [Google Scholar] [CrossRef] [PubMed]
  26. Krylova, S.V.; Feng, D.R. The Machinery of Exosomes: Biogenesis, Release, and Uptake. Int. J. Mol. Sci. 2023, 24, 1337. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, Z.G.; He, Z.Y.; Liang, S.; Yang, Q.; Cheng, P.; Chen, A.M. Comprehensive proteomic analysis of exosomes derived from human bone marrow, adipose tissue, and umbilical cord mesenchymal stem cells. Stem Cell Res. Ther. 2020, 11, 511. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, H.Z.; Wang, L.Y.; Zeng, X.L.; Schwarz, H.; Nanda, H.S.; Peng, X.S.; Zhou, Y.B. Exosomes, a New Star for Targeted Delivery. Front. Cell Dev. Biol. 2021, 9, 751079. [Google Scholar] [CrossRef] [PubMed]
  29. Qiu, H.; Shi, S.S.; Wang, S.C.; Peng, H.; Ding, S.J.; Wang, L.C. Proteomic Profiling Exosomes from Vascular Smooth Muscle Cell. Proteom. Clin. Appl. 2018, 12, 1700097. [Google Scholar] [CrossRef]
  30. Li, D.M.; Zhang, W.; Chen, X.L.; Ling, H.S.; Xie, P.; Chen, Z.H.; Adili, A.; Chen, Z.X.; Yang, F.Q.; Zhang, C.Y.; et al. Proteomic profiling of MIN6 cell-derived exosomes. J. Proteom. 2020, 224, 103841. [Google Scholar] [CrossRef]
  31. Wang, Y.T.; Shi, T.J.; Srivastava, S.; Kagan, J.; Liu, T.; Rodland, K.D. Proteomic Analysis of Exosomes for Discovery of Protein Biomarkers for Prostate and Bladder Cancer. Cancers 2020, 12, 2335. [Google Scholar] [CrossRef] [PubMed]
  32. Wei, H.; Green, E.; Ball, L.; Fan, H.K.; Lee, J.N.F.; Strange, C.; Wang, H.J. Proteomic Analysis of Exosomes Secreted from Human Alpha-1 Antitrypsin Overexpressing Mesenchymal Stromal Cells. Biology 2022, 11, 9. [Google Scholar] [CrossRef] [PubMed]
  33. Ferreira, D.; Moreira, J.N.; Rodrigues, L.R. New advances in exosome-based targeted drug delivery systems. Crit. Rev. Oncol. Hematol. 2022, 172, 103628. [Google Scholar] [CrossRef] [PubMed]
  34. Fu, S.Y.; Wang, Y.; Xia, X.H.; Zheng, J.L.C. Exosome engineering: Current progress in cargo loading and targeted delivery. Nanoimpact 2020, 20, 100261. [Google Scholar] [CrossRef]
  35. Qian, R.J.; Jing, B.P.; Jiang, D.W.; Gai, Y.K.; Zhu, Z.Y.; Huang, X.J.; Gao, Y.; Lan, X.L.; An, R. Multi-antitumor therapy and synchronous imaging monitoring based on exosome. Eur. J. Nucl. Med. Mol. Imaging 2022, 49, 2668–2681. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, J.; Chen, D.; Ho, E.A. Challenges in the development and establishment of exosome-based drug delivery systems. J. Control. Release 2021, 329, 894–906. [Google Scholar] [CrossRef] [PubMed]
  37. Lee, Y.; Ni, J.; Beretov, J.; Wasinger, V.C.; Graham, P.; Li, Y. Recent advances of small extracellular vesicle biomarkers in breast cancer diagnosis and prognosis. Mol. Cancer 2023, 22, 33. [Google Scholar] [CrossRef] [PubMed]
  38. Kosaka, N.; Yoshioka, Y.; Fujita, Y.; Ochiya, T. Versatile roles of extracellular vesicles in cancer. J. Clin. Investig. 2016, 126, 1163–1172. [Google Scholar] [CrossRef] [PubMed]
  39. Schou, A.S.; Nielsen, J.E.; Askeland, A.; Jørgensen, M.M. Extracellular vesicle-associated proteins as potential biomarkers. Adv. Clin. Chem. 2020, 99, 1–48. [Google Scholar]
  40. Barbo, M.; Ravnik-Glavač, M. Extracellular Vesicles as Potential Biomarkers in Amyotrophic Lateral Sclerosis. Genes 2023, 14, 325. [Google Scholar] [CrossRef]
  41. Karnati, H.K.; Garcia, J.H.; Tweedie, D.; Becker, R.E.; Kapogiannis, D.; Greig, N.H. Neuronal Enriched Extracellular Vesicle Proteins as Biomarkers for Traumatic Brain Injury. J. Neurotrauma 2019, 36, 975–987. [Google Scholar] [CrossRef] [PubMed]
  42. Kubo, H. Extracellular Vesicles in Lung Disease. Chest 2018, 153, 210–216. [Google Scholar] [CrossRef] [PubMed]
  43. Muñoz-Hernández, R.; Rojas, Á.; Gato, S.; Gallego, J.; Gil-Gómez, A.; Castro, M.J.; Ampuero, J.; Romero-Gómez, M. Extracellular Vesicles as Biomarkers in Liver Disease. Int. J. Mol. Sci. 2022, 23, 16217. [Google Scholar] [CrossRef] [PubMed]
  44. Thietart, S.; Rautou, P.E. Extracellular vesicles as biomarkers in liver diseases: A clinician’s point of view. J. Hepatol. 2020, 73, 1507–1525. [Google Scholar] [CrossRef]
  45. Louie, H.H.; Mugisho, O.O.; Chamley, L.W.; Rupenthal, I.D. Extracellular Vesicles as Biomarkers and Therapeutics for Inflammatory Eye Diseases. Mol. Pharm. 2023, 20, 23–40. [Google Scholar] [CrossRef] [PubMed]
  46. Mu, Y.; McManus, D.P.; Gordon, C.A.; Cai, P. Parasitic Helminth-Derived microRNAs and Extracellular Vesicle Cargos as Biomarkers for Helminthic Infections. Front. Cell. Infect. Microbiol. 2021, 11, 708952. [Google Scholar] [CrossRef] [PubMed]
  47. Xu, K.; Liu, Q.; Wu, K.; Liu, L.; Zhao, M.; Yang, H.; Wang, X.; Wang, W. Extracellular vesicles as potential biomarkers and therapeutic approaches in autoimmune diseases. J. Transl. Med. 2020, 18, 432. [Google Scholar] [CrossRef] [PubMed]
  48. Grange, C.; Bussolati, B. Extracellular vesicles in kidney disease. Nat. Rev. Nephrol. 2022, 18, 499–513. [Google Scholar] [CrossRef]
  49. Shao, H.; Chung, J.; Balaj, L.; Charest, A.; Bigner, D.D.; Carter, B.S.; Hochberg, F.H.; Breakefield, X.O.; Weissleder, R.; Lee, H. Protein typing of circulating microvesicles allows real-time monitoring of glioblastoma therapy. Nat. Med. 2012, 18, 1835–1840. [Google Scholar] [CrossRef]
  50. Akoto, T.; Saini, S. Role of Exosomes in Prostate Cancer Metastasis. Int. J. Mol. Sci. 2021, 22, 3528. [Google Scholar] [CrossRef]
  51. Esfandyari, S.; Elkafas, H.; Chugh, R.M.; Park, H.S.; Navarro, A.; Al-Hendy, A. Exosomes as Biomarkers for Female Reproductive Diseases Diagnosis and Therapy. Int. J. Mol. Sci. 2021, 22, 2165. [Google Scholar] [CrossRef] [PubMed]
  52. Lakshmi, S.; Hughes, T.A.; Priya, S. Exosomes and exosomal RNAs in breast cancer: A status update. Eur. J. Cancer 2021, 144, 252–268. [Google Scholar] [CrossRef] [PubMed]
  53. Sorop, A.; Constantinescu, D.; Cojocaru, F.; Dinischiotu, A.; Cucu, D.; Dima, S.O. Exosomal microRNAs as Biomarkers and Therapeutic Targets for Hepatocellular Carcinoma. Int. J. Mol. Sci. 2021, 22, 4997. [Google Scholar] [CrossRef] [PubMed]
  54. Thind, A.; Wilson, C. Exosomal miRNAs as cancer biomarkers and therapeutic targets. J. Extracell. Vesicles 2016, 5, 31292. [Google Scholar] [CrossRef] [PubMed]
  55. Qiu, S.; Xie, L.; Lu, C.; Gu, C.; Xia, Y.; Lv, J.; Xuan, Z.; Fang, L.; Yang, J.; Zhang, L.; et al. Gastric cancer-derived exosomal miR-519a-3p promotes liver metastasis by inducing intrahepatic M2-like macrophage-mediated angiogenesis. J. Exp. Clin. Cancer Res. 2022, 41, 296. [Google Scholar] [CrossRef]
  56. Litwińska, Z.; Łuczkowska, K.; Machaliński, B. Extracellular vesicles in hematological malignancies. Leuk. Lymphoma 2019, 60, 29–36. [Google Scholar] [CrossRef] [PubMed]
  57. Khalife, J.; Sanchez, J.F.; Pichiorri, F. Extracellular Vesicles in Hematological Malignancies: From Biomarkers to Therapeutic Tools. Diagnostics 2020, 10, 1065. [Google Scholar] [CrossRef] [PubMed]
  58. Wiklander, O.P.B.; Brennan, M.; Lötvall, J.; Breakefield, X.O.; El Andaloussi, S. Advances in therapeutic applications of extracellular vesicles. Sci. Transl. Med. 2019, 11, eaav8521. [Google Scholar] [CrossRef] [PubMed]
  59. Herrmann, I.K.; Wood, M.J.A.; Fuhrmann, G. Extracellular vesicles as a next-generation drug delivery platform. Nat. Nanotechnol. 2021, 16, 748–759. [Google Scholar] [CrossRef]
  60. Mead, B.; Tomarev, S. Extracellular vesicle therapy for retinal diseases. Prog. Retin. Eye Res. 2020, 79, 100849. [Google Scholar] [CrossRef]
  61. Takahashi, Y.; Takakura, Y. Extracellular vesicle-based therapeutics: Extracellular vesicles as therapeutic targets and agents. Pharmacol. Ther. 2023, 242, 108352. [Google Scholar] [CrossRef] [PubMed]
  62. Weng, Z.; Zhang, B.; Wu, C.; Yu, F.; Han, B.; Li, B.; Li, L. Therapeutic roles of mesenchymal stem cell-derived extracellular vesicles in cancer. J. Hematol. Oncol. 2021, 14, 136. [Google Scholar] [CrossRef]
  63. Lau, H.C.; Han, D.W.; Park, J.; Lehner, E.; Kals, C.; Arzt, C.; Bayer, E.; Auer, D.; Schally, T.; Grasmann, E.; et al. GMP-compliant manufacturing of biologically active cell-derived vesicles produced by extrusion technology. J. Extracell. Biol. 2022, 1, e70. [Google Scholar] [CrossRef]
  64. Estes, S.; Konstantinov, K.; Young, J.D. Manufactured extracellular vesicles as human therapeutics: Challenges, advances, and opportunities. Curr. Opin. Biotechnol. 2022, 77, 102776. [Google Scholar] [CrossRef] [PubMed]
  65. Munir, J.; Lee, M.; Ryu, S. Exosomes in Food: Health Benefits and Clinical Relevance in Diseases. Adv. Nutr. 2020, 11, 687–696. [Google Scholar] [CrossRef]
  66. Syromiatnikova, V.; Prokopeva, A.; Gomzikova, M. Methods of the Large-Scale Production of Extracellular Vesicles. Int. J. Mol. Sci. 2022, 23, 10522. [Google Scholar] [CrossRef]
  67. Kink, J.A.; Bellio, M.A.; Forsberg, M.H.; Lobo, A.; Thickens, A.S.; Lewis, B.M.; Ong, I.M.; Khan, A.; Capitini, C.M.; Hematti, P. Large-scale bioreactor production of extracellular vesicles from mesenchymal stromal cells for treatment of acute radiation syndrome. Stem Cell Res. Ther. 2024, 15, 72. [Google Scholar] [CrossRef]
  68. Pincela Lins, P.M.; Pirlet, E.; Szymonik, M.; Bronckaers, A.; Nelissen, I. Manufacture of extracellular vesicles derived from mesenchymal stromal cells. Trends Biotechnol. 2023, 41, 965–981. [Google Scholar] [CrossRef] [PubMed]
  69. Chen, Y.S.; Lin, E.Y.; Chiou, T.W.; Harn, H.J. Exosomes in clinical trial and their production in compliance with good manufacturing practice. Tzu Chi Med. J. 2020, 32, 113–120. [Google Scholar] [CrossRef]
  70. Hou, Y.; Mi, K.; Sun, L.; Zhou, K.; Wang, L.; Zhang, L.; Liu, Z.; Huang, L. The Application of Hollow Fiber Cartridge in Biomedicine. Pharmaceutics 2022, 14, 1485. [Google Scholar] [CrossRef]
  71. Ng, C.Y.; Kee, L.T.; Al-Masawa, M.E.; Lee, Q.H.; Subramaniam, T.; Kok, D.; Ng, M.H.; Law, J.X. Scalable Production of Extracellular Vesicles and Its Therapeutic Values: A Review. Int. J. Mol. Sci. 2022, 23, 7986. [Google Scholar] [CrossRef] [PubMed]
  72. Tantakitti, F.; Pata, S.; Laopajon, W.; Kasinrerk, W.; Hidalgo-Bastida, A. Production of High-Value Proteins under Stringent Cost Constraints—The Case of Hollow Fiber Technology for Cell Culture. Processes 2023, 11, 889. [Google Scholar] [CrossRef]
  73. Jakl, V.; Ehmele, M.; Winkelmann, M.; Ehrenberg, S.; Eiseler, T.; Friemert, B.; Rojewski, M.T.; Schrezenmeier, H. A novel approach for large-scale manufacturing of small extracellular vesicles from bone marrow-derived mesenchymal stromal cells using a hollow fiber bioreactor. Front. Bioeng. Biotechnol. 2023, 11, 1107055. [Google Scholar] [CrossRef] [PubMed]
  74. Joshi, B.S.; Ortiz, D.; Zuhorn, I.S. Converting extracellular vesicles into nanomedicine: Loading and unloading of cargo. Materials Today Nano 2021, 16, 100148. [Google Scholar] [CrossRef]
  75. Akuma, P.; Okagu, O.D.; Udenigwe, C.C. Naturally Occurring Exosome Vesicles as Potential Delivery Vehicle for Bioactive Compounds. Front. Sustain. Food Syst. 2019, 3, 23. [Google Scholar] [CrossRef]
  76. Cao, X.H.; Liang, M.X.; Wu, Y.; Yang, K.; Tang, J.H.; Zhang, W. Extracellular vesicles as drug vectors for precise cancer treatment. Nanomedicine 2021, 16, 1519–1537. [Google Scholar] [CrossRef]
  77. Somiya, M.; Yoshioka, Y.; Ochiya, T. Drug delivery application of extracellular vesicles; insight into production, drug loading, targeting, and pharmacokinetics. Aims Bioeng. 2017, 4, 73–92. [Google Scholar] [CrossRef]
  78. Song, H.; Liu, B.; Dong, B.; Xu, J.; Zhou, H.; Na, S.; Liu, Y.Y.; Pan, Y.X.; Chen, F.Y.; Li, L.; et al. Exosome-Based Delivery of Natural Products in Cancer Therapy. Front. Cell Dev. Biol. 2021, 9, 650426. [Google Scholar] [CrossRef] [PubMed]
  79. Pan, S.J.; Pei, L.J.; Zhang, A.M.; Zhang, Y.H.; Zhang, C.L.; Huang, M.; Huang, Z.C.; Liu, B.; Wang, L.R.; Ma, L.J.; et al. Passion fruit-like exosome-PMA/Au-BSA@Ce6 nanovehicles for real-time fluorescence imaging and enhanced targeted photodynamic therapy with deep penetration and superior retention behavior in tumor. Biomaterials 2020, 230, 119606. [Google Scholar] [CrossRef]
  80. Vader, P.; Breakefield, X.O.; Wood, M.J.A. Extracellular vesicles: Emerging targets for cancer therapy. Trends Mol. Med. 2014, 20, 385–393. [Google Scholar] [CrossRef]
  81. Tian, Y.H.; Li, S.P.; Song, J.; Ji, T.J.; Zhu, M.T.; Anderson, G.J.; Wei, J.Y.; Nie, G.J. A doxorubicin delivery platform using engineered natural membrane vesicle exosomes for targeted tumor therapy. Biomaterials 2014, 35, 2383–2390. [Google Scholar] [CrossRef]
  82. Alvarez-Erviti, L.; Seow, Y.Q.; Yin, H.F.; Betts, C.; Lakhal, S.; Wood, M.J.A. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef]
  83. Izco, M.; Blesa, J.; Schleef, M.; Schmeer, M.; Porcari, R.; Al-Shawi, R.; Ellmerich, S.; de Toro, M.; Gardiner, C.; Seow, Y.; et al. Systemic Exosomal Delivery of shRNA Minicircles Prevents Parkinsonian Pathology. Mol. Ther. 2019, 27, 2111–2122. [Google Scholar] [CrossRef]
  84. Li, Y.J.; Wu, J.Y.; Wang, J.M.; Hu, X.B.; Cai, J.X.; Xiang, D.X. Gemcitabine loaded autologous exosomes for effective and safe chemotherapy of pancreatic cancer. Acta Biomater. 2020, 101, 519–530. [Google Scholar] [CrossRef]
  85. Haney, M.J.; Klyachko, N.L.; Zhaoa, Y.L.; Gupta, R.; Plotnikova, E.G.; He, Z.J.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J. Control. Release 2015, 207, 18–30. [Google Scholar] [CrossRef]
  86. Sancho-Albero, M.; Rubio-Ruiz, B.; Perez-Lopez, A.M.; Sebastian, V.; Martin-Duque, P.; Arruebo, M.; Santamaria, J.; Unciti-Broceta, A. Cancer-derived exosomes loaded with ultrathin palladium nanosheets for targeted bioorthogonal catalysis. Nat. Catal. 2019, 2, 864–872. [Google Scholar] [CrossRef]
  87. Fuhrmann, G.; Serio, A.; Mazo, M.; Nair, R.; Stevens, M.M. Active loading into extracellular vesicles significantly improves the cellular uptake and photodynamic effect of porphyrins. J. Control. Release 2015, 205, 35–44. [Google Scholar] [CrossRef]
  88. Saari, H.; Lazaro-Ibanez, E.; Viitala, T.; Vuorimaa-Laukkanen, E.; Siljander, P.; Yliperttula, M. Microvesicle- and exosome-mediated drug delivery enhances the cytotoxicity of Paclitaxel in autologous prostate cancer cells. J. Control. Release 2015, 220, 727–737. [Google Scholar] [CrossRef]
  89. Gong, C.A.; Tian, J.; Wang, Z.; Gao, Y.; Wu, X.; Ding, X.Y.; Qiang, L.; Li, G.R.; Han, Z.M.; Yuan, Y.F.; et al. Functional exosome-mediated co-delivery of doxorubicin and hydrophobically modified microRNA 159 for triple-negative breast cancer therapy. J. Nanobiotechnol. 2019, 17, 93. [Google Scholar] [CrossRef] [PubMed]
  90. del Pozo-Acebo, L.; de las Hazas, M.C.L.; Tome-Carneiro, J.; Gil-Cabrerizo, P.; San-Cristobal, R.; Busto, R.; Garcia-Ruiz, A.; Davalos, A. Bovine Milk-Derived Exosomes as a Drug Delivery Vehicle for miRNA-Based Therapy. Int. J. Mol. Sci. 2021, 22, 1105. [Google Scholar] [CrossRef] [PubMed]
  91. Betzer, O.; Perets, N.; Ange, A.; Motiei, M.; Sadan, T.; Yadid, G.; Offen, D.; Popovtzer, R. In Vivo Neuroimaging of Exosomes Using Gold Nanoparticles. ACS Nano 2017, 11, 10883–10893. [Google Scholar] [CrossRef]
  92. Munagala, R.; Aqil, F.; Jeyabalan, J.; Gupta, R.C. Bovine milk-derived exosomes for drug delivery. Cancer Lett. 2016, 371, 48–61. [Google Scholar] [CrossRef]
  93. Li, X.; Corbett, A.L.; Taatizadeh, E.; Tasnim, N.; Little, J.P.; Garnis, C.; Daugaard, M.; Guns, E.; Hoorfar, M.; Li, I.T.S. Challenges and opportunities in exosome research-Perspectives from biology, engineering, and cancer therapy. APL Bioeng. 2019, 3, 011503. [Google Scholar] [CrossRef]
  94. Busatto, S.; Iannotta, D.; Walker, S.A.; Di Marzio, L.; Wolfram, J. A Simple and Quick Method for Loading Proteins in Extracellular Vesicles. Pharmaceuticals 2021, 14, 356. [Google Scholar] [CrossRef]
  95. Liang, Y.J.; Duan, L.; Lu, J.P.; Xia, J. Engineering exosomes for targeted drug delivery. Theranostics 2021, 11, 3183–3195. [Google Scholar] [CrossRef]
  96. Xi, X.M.; Chen, M.; Xia, S.J.; Lu, R. Drug loading techniques for exosome-based drug delivery systems. Pharmazie 2021, 76, 61–67. [Google Scholar]
  97. Donoso-Quezada, J.; Ayala-Mar, S.; Gonzalez-Valdez, J. State-of-the-art exosome loading and functionalization techniques for enhanced therapeutics: A review. Crit. Rev. Biotechnol. 2020, 40, 804–820. [Google Scholar] [CrossRef]
  98. Sadeghi, S.; Tehrani, F.R.; Tahmasebi, S.; Shafiee, A.; Hashemi, S.M. Exosome engineering in cell therapy and drug delivery. Inflammopharmacology 2023, 31, 145–169. [Google Scholar] [CrossRef]
  99. Chen, D. Potential of Surface Modified Exosomes for Extended Intravitreal Drug Delivery. Master’s Thesis, University of Waterloo, Waterloo, ON, Canada, 2021. [Google Scholar]
  100. Nan, W.B.; Zhang, C.; Wang, H.; Chen, H.L.; Ji, S.L. Direct Modification of Extracellular Vesicles and Its Applications for Cancer Therapy: A Mini-Review. Front. Chem. 2022, 10, 910341. [Google Scholar] [CrossRef] [PubMed]
  101. Cooper, J.M.; Wiklander, P.B.O.; Nordin, J.Z.; Al-Shawi, R.; Wood, M.J.; Vithlani, M.; Schapira, A.H.V.; Simons, J.P.; El-Andaloussi, S.; Alvarez-Erviti, L. Systemic Exosomal siRNA Delivery Reduced Alpha-Synuclein Aggregates in Brains of Transgenic Mice. Mov. Disord. 2014, 29, 1476–1485. [Google Scholar] [CrossRef] [PubMed]
  102. Luo, Z.W.; Li, F.X.Z.; Liu, Y.W.; Rao, S.S.; Yin, H.; Huang, J.; Chen, C.Y.; Hu, Y.; Zhang, Y.; Tan, Y.J.; et al. Aptamer-functionalized exosomes from bone marrow stromal cells target bone to promote bone regeneration. Nanoscale 2019, 11, 20884–20892. [Google Scholar] [CrossRef] [PubMed]
  103. Hood, J.L. Post isolation modification of exosomes for nanomedicine applications. Nanomedicine 2016, 11, 1745–1756. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, C.L.; Xu, M.C.; Fan, Q.Z.; Li, C.H.; Zhou, X.Y. Therapeutic potential of exosome-based personalized delivery platform in chronic inflammatory diseases. Asian J. Pharm. Sci. 2023, 18, 100772. [Google Scholar] [CrossRef] [PubMed]
  105. Wei, H.X.; Chen, J.Y.; Wang, S.L.; Fu, F.H.; Zhu, X.; Wu, C.Y.; Liu, Z.J.; Zhong, G.X.; Lin, J.H. A Nanodrug Consisting Of Doxorubicin And Exosome Derived From Mesenchymal Stem Cells For Osteosarcoma Treatment In Vitro. Int. J. Nanomed. 2019, 14, 8603–8610. [Google Scholar] [CrossRef] [PubMed]
  106. Lennaard, A.J.; Mamand, D.R.; Wiklander, R.J.; El Andaloussi, S.; Wiklander, O.P.B. Optimised Electroporation for Loading of Extracellular Vesicles with Doxorubicin. Pharmaceutics 2022, 14, 38. [Google Scholar] [CrossRef] [PubMed]
  107. Limoni, S.K.; Moghadam, M.F.; Moazzeni, S.M.; Gomari, H.; Salimi, F. Engineered Exosomes for Targeted Transfer of siRNA to HER2 Positive Breast Cancer Cells. Appl. Biochem. Biotechnol. 2019, 187, 352–364. [Google Scholar] [CrossRef] [PubMed]
  108. Kim, G.; Kim, M.; Lee, Y.; Byun, J.W.; Hwang, D.W.; Lee, M. Systemic delivery of microRNA-21 antisense oligonucleotides to the brain using T7-peptide decorated exosomes. J. Control. Release 2020, 317, 273–281. [Google Scholar] [CrossRef] [PubMed]
  109. Ma, Y.F.; Sun, L.L.; Zhang, J.J.; Chiang, C.L.; Pan, J.J.; Wang, X.Y.; Kwak, K.J.; Li, H.; Zhao, R.L.; Rima, X.Y.; et al. Exosomal mRNAs for Angiogenic-Osteogenic Coupled Bone Repair. Adv. Sci. 2023, 10, 2302622. [Google Scholar] [CrossRef] [PubMed]
  110. Conceiça, M.; Forcina, L.; Wiklander, O.P.B.; Gupta, D.; Nordin, J.Z.; Vrellaku, B.; McClorey, G.; Mäger, I.; Görgens, A.; Lundin, P.; et al. Engineered extracellular vesicle decoy receptor-mediated modulation of the IL6 trans-signalling pathway in muscle. Biomaterials 2021, 266, 120435. [Google Scholar] [CrossRef]
  111. Yao, Y.; Fu, C.M.; Zhou, L.; Mi, Q.S.; Jiang, A.M. DC-Derived Exosomes for Cancer Immunotherapy. Cancers 2021, 13, 3667. [Google Scholar] [CrossRef]
  112. Mohammadi, M.; Zargartalebi, H.; Salahandish, R.; Aburashed, R.; Yong, K.W.; Sanati-Nezhad, A. Emerging technologies and commercial products in exosome-based cancer diagnosis and prognosis. Biosens. Bioelectron. 2021, 183, 113176. [Google Scholar] [CrossRef] [PubMed]
  113. Li, J.; Li, Y.R.; Li, P.L.; Zhang, Y.; Du, L.T.; Wang, Y.S.; Zhang, C.P.; Wang, C.X. Exosome detection via surface-enhanced Raman spectroscopy for cancer diagnosis. Acta Biomater. 2022, 144, 1–14. [Google Scholar] [CrossRef]
  114. Wang, Y.; Guo, M.; Lin, D.; Liang, D.; Zhao, L.; Zhao, R.; Wang, Y. Docetaxel-loaded exosomes for targeting non-small cell lung cancer: Preparation and evaluation in vitro and in vivo. Drug Deliv. 2021, 28, 1510–1523. [Google Scholar] [CrossRef] [PubMed]
  115. Dimik, M.; Abeysinghe, P.; Logan, J.; Mitchell, M. The exosome: A review of current therapeutic roles and capabilities in human reproduction. Drug Deliv. Transl. Res. 2023, 13, 473–502. [Google Scholar] [CrossRef] [PubMed]
  116. Altanerova, U.; Jakubechova, J.; Benejova, K.; Petra, P.; Pesta, M.; Pitule, P.; Ondrej, T.; Kausitz, J.; Zduriencikova, M.; Repiska, V.; et al. Prodrug suicide gene therapy for cancer targeted intracellular by mesenchymal stem cell exosomes. Int. J. Cancer 2019, 144, 897–908. [Google Scholar] [CrossRef] [PubMed]
  117. Yang, T.Z.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S.H. Exosome Delivered Anticancer Drugs Across the Blood-Brain Barrier for Brain Cancer Therapy in Danio Rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef]
  118. Liang, G.F.; Kan, S.; Zhu, Y.L.; Feng, S.Y.; Feng, W.P.; Gao, S.G. Engineered exosome-mediated delivery of functionally active miR-26a and its enhanced suppression effect in HepG2 cells. Int. J. Nanomed. 2018, 13, 585–599. [Google Scholar] [CrossRef]
  119. Lee, H.; Park, H.; Noh, G.J.; Lee, E.S. pH-responsive hyaluronate-anchored extracellular vesicles to promote tumor-targeted drug delivery. Carbohydr. Polym. 2018, 202, 323–333. [Google Scholar] [CrossRef]
  120. Marar, C.; Starich, B.; Wirtz, D. Extracellular vesicles in immunomodulation and tumor progression. Nat. Immunol. 2021, 22, 560–570. [Google Scholar] [CrossRef]
  121. Araujo-Abad, S.; Saceda, M.; de Juan Romero, C. Biomedical application of small extracellular vesicles in cancer treatment. Adv. Drug Deliv. Rev. 2022, 182, 114117. [Google Scholar] [CrossRef]
  122. Galieva, L.R.; James, V.; Mukhamedshina, Y.O.; Rizvanov, A.A. Therapeutic Potential of Extracellular Vesicles for the Treatment of Nerve Disorders. Front. Neurosci. 2019, 13, 163. [Google Scholar] [CrossRef] [PubMed]
  123. Yang, X.; Ma, Y.; Xie, H.; Dong, S.; Rao, G.; Yang, Z.; Zhang, J.; Wu, Q. Extracellular Vesicles in the Treatment of Parkinson’s Disease: A Review. Curr. Med. Chem. 2021, 28, 6375–6394. [Google Scholar] [CrossRef]
  124. Reed, S.L.; Escayg, A. Extracellular vesicles in the treatment of neurological disorders. Neurobiol. Dis. 2021, 157, 105445. [Google Scholar] [CrossRef] [PubMed]
  125. Han, C.; Yang, J.; Sun, J.; Qin, G. Extracellular vesicles in cardiovascular disease: Biological functions and therapeutic implications. Pharmacol. Ther. 2022, 233, 108025. [Google Scholar] [CrossRef]
  126. Huang, C.; Neupane, Y.R.; Lim, X.C.; Shekhani, R.; Czarny, B.; Wacker, M.G.; Pastorin, G.; Wang, J.W. Extracellular vesicles in cardiovascular disease. Adv. Clin. Chem. 2021, 103, 47–95. [Google Scholar]
  127. Giró, O.; Jiménez, A.; Pané, A.; Badimon, L.; Ortega, E.; Chiva-Blanch, G. Extracellular vesicles in atherothrombosis and cardiovascular disease: Friends and foes. Atherosclerosis 2021, 330, 61–75. [Google Scholar] [CrossRef] [PubMed]
  128. Lee, J.R.; Park, B.W.; Kim, J.; Choo, Y.W.; Kim, H.Y.; Yoon, J.K.; Kim, H.; Hwang, J.W.; Kang, M.; Kwon, S.P.; et al. Nanovesicles derived from iron oxide nanoparticles-incorporated mesenchymal stem cells for cardiac repair. Sci. Adv. 2020, 6, eaaz0952. [Google Scholar] [CrossRef]
  129. Armstrong, J.P.K.; Stevens, M.M. Strategic design of extracellular vesicle drug delivery systems. Adv. Drug Deliv. Rev. 2018, 130, 12–16. [Google Scholar] [CrossRef] [PubMed]
  130. Sun, D.; Zhuang, X.; Xiang, X.; Liu, Y.; Zhang, S.; Liu, C.; Barnes, S.; Grizzle, W.; Miller, D.; Zhang, H.G. A novel nanoparticle drug delivery system: The anti-inflammatory activity of curcumin is enhanced when encapsulated in exosomes. Mol. Ther. 2010, 18, 1606–1614. [Google Scholar] [CrossRef]
  131. Ou, Y.H.; Liang, J.; Czarny, B.; Wacker, M.G.; Yu, V.; Wang, J.W.; Pastorin, G. Extracellular Vesicle (EV) biohybrid systems for cancer therapy: Recent advances and future perspectives. Semin. Cancer Biol. 2021, 74, 45–61. [Google Scholar] [CrossRef]
  132. Sidhom, K.; Obi, P.O.; Saleem, A. A Review of Exosomal Isolation Methods: Is Size Exclusion Chromatography the Best Option? Int. J. Mol. Sci. 2020, 21, 6466. [Google Scholar] [CrossRef] [PubMed]
  133. Zhou, X.B.; Zhang, H.M.; Feng, H.; Meng, X.K.; Li, S.Y.; Cui, Y.Y.; Zhao, Y.N.; Chen, Y.B.; Tong, T. Glypican-1 in serum-derived exosomes as a potential biomarker in liquid biopsy of non-small cell lung cancer. Minerva Biotecnol. 2019, 31, 60–67. [Google Scholar] [CrossRef]
  134. Melo, S.A.; Luecke, L.B.; Kahlert, C.; Fernandez, A.F.; Gammon, S.T.; Kaye, J.; LeBleu, V.S.; Mittendorf, E.A.; Weitz, J.; Rahbari, N.; et al. Glypican-1 identifies cancer exosomes and detects early pancreatic cancer. Nature 2015, 523, 177–182. [Google Scholar] [CrossRef] [PubMed]
  135. He, C.J.; Zheng, S.; Luo, Y.; Wang, B. Exosome Theranostics: Biology and Translational Medicine. Theranostics 2018, 8, 237–255. [Google Scholar] [CrossRef] [PubMed]
  136. Deng, Y.; Sun, Z.W.; Wang, L.; Wang, M.H.; Yang, J.; Li, G.X. Biosensor-based assay of exosome biomarker for early diagnosis of cancer. Front. Med. 2022, 16, 157–175. [Google Scholar] [CrossRef]
  137. Ran, Z.H.; Wu, S.B.; Ma, Z.; Chen, X.W.; Liu, J.; Yang, J.C. Advances in exosome biomarkers for cervical cancer. Cancer Med. 2022, 11, 4966–4978. [Google Scholar] [CrossRef] [PubMed]
  138. Salciccia, S.; Frisenda, M.; Bevilacqua, G.; Gobbi, L.; Bucca, B.; Moriconi, M.; Viscuso, P.; Gentilucci, A.; Mariotti, G.; Cattarino, S.; et al. Exosome Analysis in Prostate Cancer: How They Can Improve Biomarkers’ Performance. Curr. Issues Mol. Biol. 2023, 45, 6085–6096. [Google Scholar] [CrossRef] [PubMed]
  139. Kalluri, R. The biology and function of exosomes in cancer. J. Clin. Investig. 2016, 126, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  140. Liu, J.Y.; Ren, L.W.; Li, S.; Li, W.; Zheng, X.J.; Yang, Y.H.; Fu, W.Q.; Yi, J.; Wang, J.H.; Du, G.H. The biology, function, and applications of exosomes in cancer. Acta Pharm. Sin. B 2021, 11, 2783–2797. [Google Scholar] [CrossRef]
  141. Munoz, J.L.; Bliss, S.A.; Greco, S.J.; Ramkissoon, S.H.; Ligon, K.L.; Rameshwar, P. Delivery of Functional Anti-miR-9 by Mesenchymal Stem Cell-derived Exosomes to Glioblastoma Multiforme Cells Conferred Chemosensitivity. Mol. Ther.-Nucleic Acids 2013, 2, e126. [Google Scholar] [CrossRef]
  142. El Andaloussi, S.; Lakhal, S.; Mäger, I.; Wood, M.J.A. Exosomes for targeted siRNA delivery across biological barriers. Adv. Drug Deliv. Rev. 2013, 65, 391–397. [Google Scholar] [CrossRef]
  143. Steinbichler, T.B.; Dudás, J.; Skvortsov, S.; Ganswindt, U.; Riechelmann, H.; Skvortsova, I.I. Therapy resistance mediated by exosomes. Mol. Cancer 2019, 18, 58. [Google Scholar] [CrossRef] [PubMed]
  144. Wahlgren, J.; Karlson, T.D.; Brisslert, M.; Sani, F.V.; Telemo, E.; Sunnerhagen, P.; Valadi, H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012, 40, e130. [Google Scholar] [CrossRef] [PubMed]
  145. Kim, S.M.; Yang, Y.; Oh, S.J.; 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. [Google Scholar] [CrossRef] [PubMed]
  146. Duan, L.; Ouyang, K.; Wang, J.H.; Xu, L.M.; Xu, X.; Wen, C.I.; Xie, Y.X.; Liang, Y.J.; Xia, J. Exosomes as Targeted Delivery Platform of CRISPR/Cas9 for Therapeutic Genome Editing. ChemBioChem 2021, 22, 3360–3368. [Google Scholar] [CrossRef] [PubMed]
  147. Mendt, M.; Kamerkar, S.; Sugimoto, H.; McAndrews, K.M.; Wu, C.C.; Gagea, M.; Yang, S.J.; Blanko, E.V.R.; Peng, Q.; Ma, X.Y.; et al. Generation and testing of clinical-grade exosomes for pancreatic cancer. JCI Insight 2018, 3, e99263. [Google Scholar] [CrossRef] [PubMed]
  148. Kim, M.S.; Haney, M.J.; Zhao, Y.; Mahajan, V.; Deygen, I.; Klyachko, N.L.; Inskoe, E.; Piroyan, A.; Sokolsky, M.; Okolie, O.; et al. Development of exosome-encapsulated paclitaxel to overcome MDR in cancer cells. Nanomed.-Nanotechnol. Biol. Med. 2016, 12, 655–664. [Google Scholar] [CrossRef] [PubMed]
  149. Liang, G.F.; Zhu, Y.L.; Ali, D.J.; Tian, T.; Xu, H.T.; Si, K.; Sun, B.; Chen, B.A.; Xiao, Z.D. Engineered exosomes for targeted co-delivery of miR-21 inhibitor and chemotherapeutics to reverse drug resistance in colon cancer. J. Nanobiotechnol. 2020, 18, 10. [Google Scholar] [CrossRef] [PubMed]
  150. Butreddy, A.; Kommineni, N.; Dudhipala, N. Exosomes as Naturally Occurring Vehicles for Delivery of Biopharmaceuticals: Insights from Drug Delivery to Clinical Perspectives. Nanomaterials 2021, 11, 1481. [Google Scholar] [CrossRef]
  151. Frühbeis, C.; Fröhlich, D.; Kuo, W.P.; Amphornrat, J.; Thilemann, S.; Saab, A.S.; Kirchhoff, F.; Möbius, W.; Goebbels, S.; Nave, K.A.; et al. Neurotransmitter-Triggered Transfer of Exosomes Mediates Oligodendrocyte-Neuron Communication. PLoS Biol. 2013, 11, e1001604. [Google Scholar] [CrossRef]
  152. Jiang, Y.; Wang, F.B.; Wang, K.; Zhong, Y.Q.; Wei, X.F.; Wang, Q.F.; Zhang, H. Engineered Exosomes: A Promising Drug Delivery Strategy for Brain Diseases. Curr. Med. Chem. 2022, 29, 3111–3124. [Google Scholar] [CrossRef] [PubMed]
  153. Ahmed, S.; Paramasivam, P.; Kamath, M.; Sharma, A.; Rome, S.; Murugesan, R. Genetic Exchange of Lung-Derived Exosome to Brain Causing Neuronal Changes on COVID-19 Infection. Mol. Neurobiol. 2021, 58, 5356–5368. [Google Scholar] [CrossRef] [PubMed]
  154. Iranpanah, A.; Kooshki, L.; Moradi, S.Z.; Saso, L.; Fakhri, S.; Khan, H. The Exosome-Mediated PI3K/Akt/mTOR Signaling Pathway in Neurological Diseases. Pharmaceutics 2023, 15, 1006. [Google Scholar] [CrossRef] [PubMed]
  155. Dong, X.Y.; Zheng, D.M.; Nao, J.F. Circulating Exosome microRNAs as Diagnostic Biomarkers of Dementia. Front. Aging Neurosci. 2020, 12, 580199. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, Y.; Bi, J.Y.; Huang, J.Y.; Tang, Y.N.; Du, S.Y.; Li, P.Y. Exosome: A Review of Its Classification, Isolation Techniques, Storage, Diagnostic and Targeted Therapy Applications. Int. J. Nanomed. 2020, 15, 6917–6934. [Google Scholar] [CrossRef] [PubMed]
  157. Hoshino, A.; Costa-Silva, B.; Shen, T.L.; Rodrigues, G.; Hashimoto, A.; Mark, M.T.; Molina, H.; Kohsaka, S.; Di Giannatale, A.; Ceder, S.; et al. Tumour exosome integrins determine organotropic metastasis. Nature 2015, 527, 329. [Google Scholar] [CrossRef]
  158. Martinez-Espinosa, I.; Serrato, J.A.; Ortiz-Quintero, B. The Role of Exosome-Derived microRNA on Lung Cancer Metastasis Progression. Biomolecules 2023, 13, 1574. [Google Scholar] [CrossRef] [PubMed]
  159. Xu, Z.H.; Miao, Z.W.; Jiang, Q.Z.; Gan, D.X.; Wei, X.G.; Xue, X.Z.; Li, J.Q.; Zheng, F.; Qin, X.X.; Fang, W.G.; et al. Brain microvascular endothelial cell exosome-mediated S100A16 up-regulation confers small-cell lung cancer cell survival in brain. FASEB J. 2019, 33, 1742–1757. [Google Scholar] [CrossRef]
  160. Awadasseid, A.; Wu, Y.L.; Zhang, W. Extracellular Vesicles (Exosomes) as Immunosuppressive Mediating Variables in Tumor and Chronic Inflammatory Microenvironments. Cells 2021, 10, 2533. [Google Scholar] [CrossRef]
  161. Li, X.; Liu, Y.; Zheng, S.S.; Zhang, T.Y.; Wu, J.; Sun, Y.; Zhang, J.Z.; Liu, G.Y. Role of exosomes in the immune microenvironment of ovarian cancer. Oncol. Lett. 2021, 21, 377. [Google Scholar] [CrossRef]
  162. Ricklefs, F.L.; Alayo, Q.; Krenzlin, H.; Mahmoud, A.B.; Speranza, M.C.; Nakashima, H.; Hayes, J.L.; Lee, K.; Balaj, L.; Passaro, C.; et al. Immune evasion mediated by PD-L1 on glioblastoma-derived extracellular vesicles. Sci. Adv. 2018, 4, eaar2766. [Google Scholar] [CrossRef] [PubMed]
  163. Khan, N.A.; Asim, M.; Biswas, K.H.; Alansari, A.N.; Saman, H.; Sarwar, M.Z.; Osmonaliev, K.; Uddin, S. Exosome nanovesicles as potential biomarkers and immune checkpoint signaling modulators in lung cancer microenvironment: Recent advances and emerging concepts. J. Exp. Clin. Cancer Res. 2023, 42, 221. [Google Scholar] [CrossRef] [PubMed]
  164. Kwantwi, L.B. Exosome-mediated crosstalk between tumor cells and innate immune cells: Implications for cancer progression and therapeutic strategies. J. Cancer Res. Clin. Oncol. 2023, 149, 9487–9503. [Google Scholar] [CrossRef] [PubMed]
  165. Song, J.Y.; Song, B.Q.; Yuan, L.J.; Yang, G.D. Multiplexed strategies toward clinical translation of extracellular vesicles. Theranostics 2022, 12, 6740–6761. [Google Scholar] [CrossRef] [PubMed]
  166. Kamerkar, S.; LeBleu, V.S.; Sugimoto, H.; Yang, S.J.; Ruivo, C.F.; Melo, S.A.; Lee, J.J.; Kalluri, R. Exosomes facilitate therapeutic targeting of oncogenic KRAS in pancreatic cancer. Nature 2017, 546, 498. [Google Scholar] [CrossRef] [PubMed]
  167. Xin, H.Q.; Li, Y.; Cui, Y.S.; Yang, J.J.; Zhang, Z.G.; Chopp, M. Systemic administration of exosomes released from mesenchymal stromal cells promote functional recovery and neurovascular plasticity after stroke in rats. J. Cereb. Blood Flow Metab. 2013, 33, 1711–1715. [Google Scholar] [CrossRef]
  168. Exosomes Market Report; Global Market Insights (GMI): Bellevue, WA, USA, 2023.
  169. Jay, S.M.; Bauer, S.R. Engineering Quality Control into Biomanufacturing of Extracellular Vesicle-Based Products; U.S. Food and Drug Administration: Silver Spring, MD, USA, 2023.
  170. Jay, S.; Bentley, W.; Sung, K.; Snodderly, K. Applying Additive Manufacturing for Continuous Production of Extracellular Vesicle Product; U.S. Food and Drug Administration: Silver Spring, MD, USA, 2023.
  171. Silva, A.K.A.; Morille, M.; Piffoux, M.; Arumugam, S.; Mauduit, P.; Larghero, J.; Bianchi, A.; Aubertin, K.; Blanc-Brude, O.; Noël, D.; et al. Development of extracellular vesicle-based medicinal products: A position paper of the group “Extracellular Vesicle translatiOn to clinicaL perspectiVEs—EVOLVE France”. Adv. Drug Deliv. Rev. 2021, 179, 114001. [Google Scholar] [CrossRef] [PubMed]
  172. Directive 2001/83/EC of the European Parliament and of the Council of 6 November 2001 on the Community Code Relating to Medicinal Products for Human Use. 2001. Available online: https://www.ema.europa.eu/en/documents/regulatory-procedural-guideline/directive-200183ec-european-parliament-and-council-6-november-2001-community-code-relating-medicinal-products-human-use_en.pdf (accessed on 6 March 2024).
  173. Lener, T.; Gimona, M.; Aigner, L.; Börger, V.; Buzas, E.; Camussi, G.; Chaput, N.; Chatterjee, D.; Court, F.A.; Del Portillo, H.A.; et al. Applying extracellular vesicles based therapeutics in clinical trials—An ISEV position paper. J. Extracell. Vesicles 2015, 4, 30087. [Google Scholar] [CrossRef] [PubMed]
  174. Muthu, S.; Bapat, A.; Jain, R.; Jeyaraman, N.; Jeyaraman, M. Exosomal therapy-a new frontier in regenerative medicine. Stem Cell Investig. 2021, 8, 7. [Google Scholar] [CrossRef]
  175. Scientific Recommendations on Classification of Advanced Therapy Medicinal Products EMA/140033/2021. 2024. Available online: https://www.ema.europa.eu/en/human-regulatory-overview/marketing-authorisation/advanced-therapies-marketing-authorisation/scientific-recommendations-classification-advanced-therapy-medicinal-products (accessed on 25 April 2024).
  176. Guideline on Quality, Non-Clinical and Clinical Requirements for Investigational Advanced Therapy Medicinal Products in Clinical Trials. EMA/CAT/123573/2024 Committee for Advanced Therapies (CAT). 2024. Available online: https://www.ema.europa.eu/en/guideline-quality-non-clinical-clinical-requirements-investigational-advanced-therapy-medicinal-products-clinical-trials-scientific-guideline (accessed on 25 April 2024).
  177. Cheng Ke, K.R. Guidelines for clinical translation and commercialization of extracellular vesicles and exosomes based therapeutics. Extracell. Vesicle 2023, 2, 100029. [Google Scholar] [CrossRef]
  178. Ruzycka-Ayoush, M.; Nowicka, A.M.; Kowalczyk, A.; Gluchowska, A.; Targonska, A.; Mosieniak, G.; Sobczak, K.; Donten, M.; Grudzinski, I.P. Exosomes derived from lung cancer cells: Isolation, characterization, and stability studies. Eur. J. Pharm. Sci. 2023, 181, 106369. [Google Scholar] [CrossRef] [PubMed]
  179. Kowalczyk, A.; Gajda-Walczak, A.; Ruzycka-Ayoush, M.; Targonska, A.; Mosieniak, G.; Glogowski, M.; Szumera-Cieckiewicz, A.; Prochorec-Sobieszek, M.; Bamburowicz-Klimkowska, M.; Nowicka, A.M.; et al. Parallel SPR and QCM-D Quantitative Analysis of CD9, CD63, and CD81 Tetraspanins: A Simple and Sensitive Way to Determine the Concentration of Extracellular Vesicles Isolated from Human Lung Cancer Cells. Anal. Chem. 2023, 95, 9520–9530. [Google Scholar] [CrossRef] [PubMed]
  180. Guideline on Quality, Non-Clinical and Clinical Requirements for Investigational Advanced Therapy Medicinal Products in Clinical Trials—Scientific guideline EMA/CAT/852602/2018. 2018. Available online: https://www.ema.europa.eu/en/guideline-quality-non-clinical-clinical-requirements-investigational-advanced-therapy-medicinal-products-clinical-trials-scientific-guideline (accessed on 6 March 2024).
  181. ICH Q9 Quality Risk Management—Scientific Guideline EMA/CHMP/ICH/24235/2006. 2006. Available online: https://www.ema.europa.eu/en/ich-q9-quality-risk-management-scientific-guideline (accessed on 6 March 2024).
  182. EMA/CPMP/ICH/2887/1999. ICH Guideline M4 (R4) on Common Technical Document (CTD) for the Registration of Pharmaceuticals for Human Use—Organisation of CTD 2021. Available online: https://www.ema.europa.eu/system/files/documents/scientific-guideline/m4_step_5_ctd_for_the_registration_of_pharmaceuticals_for_human_use_-_organisation_of_ctd-en.pdf (accessed on 25 April 2024).
  183. Jordan, D. An overview of the Common Technical Document (CTD) regulatory dossier. Med. Writ. 2014, 23, 101–105. [Google Scholar] [CrossRef]
  184. Elje, E.; Mariussen, E.; McFadden, E.; Dusinska, M.; Rundén-Pran, E. Different Sensitivity of Advanced Bronchial and Alveolar Mono- and Coculture Models for Hazard Assessment of Nanomaterials. Nanomaterials 2023, 13, 407. [Google Scholar] [CrossRef]
  185. Guideline on Strategies to Identify and Mitigate Risks for First-in-Human and Early Clinical Trials with Investigational Medicinal Products EMEA/CHMP/SWP/28367/07. 2017. Available online: https://www.ema.europa.eu/en/documents/scientific-guideline/guideline-strategies-identify-and-mitigate-risks-first-human-and-early-clinical-trials-investigational-medicinal-products-revision-1_en.pdf (accessed on 6 March 2024).
  186. ICH S9 Non-Clinical Evaluation for Anticancer Pharmaceuticals—Scientific Guideline EMA/CHMP/ICH/646107/2008. 2013. Available online: https://www.ema.europa.eu/en/ich-s9-non-clinical-evaluation-anticancer-pharmaceuticals-scientific-guideline (accessed on 6 March 2024).
  187. ICH S6 (R1) Preclinical Safety Evaluation of Biotechnology-Derived Pharmaceuticals—Scientific Guideline CPMP/ICH/302/95. 1997. Available online: https://www.ema.europa.eu/en/ich-s6-r1-preclinical-safety-evaluation-biotechnology-derived-pharmaceuticals-scientific-guideline (accessed on 6 March 2024).
  188. Regulation (EU) No 536/2014 of the European Parliament and of the Council of 16 April 2014 on Clinical Trials on Medicinal Products for Human Use, and Repealing Directive 2001/20/EC Text with EEA Relevance Regulation 536/2014. 2014. Available online: https://eur-lex.europa.eu/eli/reg/2014/536/ (accessed on 6 March 2024).
  189. Aimaletdinov, A.M.; Gomzikova, M.O. Tracking of Extracellular Vesicles’ Biodistribution: New Methods and Approaches. Int. J. Mol. Sci. 2022, 23, 11312. [Google Scholar] [CrossRef]
  190. Guideline on the Risk-Based Approach According to Annex I, Part IV of Directive 2001/83/EC Applied to Advanced Therapy Medicinal Products. EMA/CAT/CPWP/686637/2011 Committee for Advanced Therapies (CAT) 11 February 2013. Available online: https://www.ema.europa.eu/en/documents/scientific-guideline/draft-guideline-quality-non-clinical-clinical-requirements-investigational-advanced-therapy-medicinal-products-clinical-trials-second-version_en.pdf (accessed on 6 March 2024).
  191. Guideline on Safety and Efficacy Follow-Up and Risk Management of Advanced Therapy Medicinal Products. EMEA/149995/2008 rev.1, Committee for Medicinal Products for Human Use (CHMP) January 2018. Available online: https://pink.citeline.com/-/media/supporting-documents/pink-sheet/2018/02/atmp_risk_mgt_guide_wc500242959.pdf (accessed on 6 March 2024).
Figure 1. Classification of extracellular vesicles—three types of EVs: apoptotic bodies, microvesicles, and exosomes. Abbreviations: EVs—extracellular vesicles; MVB—multivesicular bodies.
Figure 1. Classification of extracellular vesicles—three types of EVs: apoptotic bodies, microvesicles, and exosomes. Abbreviations: EVs—extracellular vesicles; MVB—multivesicular bodies.
Ijms 25 06533 g001
Figure 2. Extracellular vesicles (EVs) as personalized biological drug products. Abbreviations: GMP—Good Manufacturing Practice.
Figure 2. Extracellular vesicles (EVs) as personalized biological drug products. Abbreviations: GMP—Good Manufacturing Practice.
Ijms 25 06533 g002
Figure 3. Overall EVs’ GMP-based manufacturing workflow. Abbreviations: EVs—extracellular vesicles; SEC—Size-Exclusion Chromatography; GMP—Good Manufacturing Practice; TEM—Transmission Electron Microscopy; NTA—Nanoparticle Tracking Analysis; WB—Western Blot.
Figure 3. Overall EVs’ GMP-based manufacturing workflow. Abbreviations: EVs—extracellular vesicles; SEC—Size-Exclusion Chromatography; GMP—Good Manufacturing Practice; TEM—Transmission Electron Microscopy; NTA—Nanoparticle Tracking Analysis; WB—Western Blot.
Ijms 25 06533 g003
Figure 4. Clinical trials using exosomes in the treatment and diagnosis of cancer in the clinicaltrials.gov database.
Figure 4. Clinical trials using exosomes in the treatment and diagnosis of cancer in the clinicaltrials.gov database.
Ijms 25 06533 g004
Figure 5. The CTD triangle. Abbreviations: CTD—Common Technical Document.
Figure 5. The CTD triangle. Abbreviations: CTD—Common Technical Document.
Ijms 25 06533 g005
Table 1. Simplified EVs’ manufacturing process.
Table 1. Simplified EVs’ manufacturing process.
ObjectiveProcess 1Process 2Quality Control
Cell culturingGrow the cells that will produce the EVsSelect a cell typeCulture these cells in flasks or bioreactors under controlled conditionsAuthentication, purity assessment
EVs productionStimulate the cells to produce and release EVsOnce cells reach a certain density, induce exosome production through specific stimuliAllow cells to secrete EVs into the culture medium over a period of timeAuthentication, purity assessment
Harvesting EVsCollect the EV-rich culture mediumCollect the culture medium containing the EVsRemove the cells from the medium through centrifugation or filtrationAuthentication, purity assessment
EVs isolationPurify the EVsPerform differential centrifugationUse ultracentrifugation or Size-Exclusion Chromatography to isolate EVs based on size and densityMethod validation, contaminant removal
EVs characterizationEnsure the EVs meet quality and functional standardsAnalyze the size, concentration, and surface markers of the EVs using NTA and flow cytometryConfirm the presence of specific proteins or RNASize distribution, quantification, cargo profiling
EVs packaging and storagePrepare the EVs for storage and distributionSterile filter the EVs preparation to ensure it is free from contaminantsPackage the EVs in sterile vials and store at −80 °C or in liquid nitrogenBiological activity, safety evaluation, storage conditions
Abbreviations: EVs—extracellular vesicles; NTA—Nanoparticle Tracking Analysis; RNA—ribonucleic acid.
Table 2. Comparison of different EV isolation techniques, taking into account the advantages and disadvantages of each method.
Table 2. Comparison of different EV isolation techniques, taking into account the advantages and disadvantages of each method.
MethodProcessProsCons
Differential centrifugationSequentially spins samples at increasing speeds to remove cells, debris, and larger vesicles, eventually pelleting EVs
-
Widely used
-
Large volumes
-
No special equipment
-
Time-consuming
-
Limited specificity; co-isolate other particles
-
Potential aggregation
UltracentrifugationUses very high-speed centrifugation to pellet EVs based on their density
-
Highly effective for small EVs
-
High purity with density gradient
-
Expensive
-
Potential EV damage
-
Time-intensive
SECSeparates EVs based on size as they pass through a column with porous beads
-
Gentle on EVs
-
High purity
-
Reproducible and scalable
-
Limited sample volume
-
Specialized chromatography columns
-
Time-consuming
Polymer-based precipitationUses polymers (e.g., polyethylene glycol) to precipitate EVs from solution
-
Simple and quick
-
No special equipment
-
Suitable for a wide range of sample volumes
-
Co-precipitates contaminants
-
Lower purity
-
Needs further purification
UltrafiltrationUses membrane filters to separate EVs based on size
-
Rapid
-
Scalable
-
No centrifugation.
-
Suitable for processing large volumes
-
Membrane clogging
-
Potential EVs damage
-
Multiple steps for purity
Immunoaffinity captureUses antibodies specific to EV surface markers to capture EVs from a mixture
-
High specificity
-
Can isolate subpopulations
-
High purity
-
Expensive
-
Small volumes
-
Requires knowledge of markers
MicrofluidicsUtilizes microfluidic devices to isolate EVs based on size, charge, or other properties
-
High precision
-
Rapid
-
Integrates with analytics
-
Specialized equipment
-
Limited scalability
-
Complex development
Abbreviations: EVs—extracellular vesicles; SEC—Size-Exclusion Chromatography.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stawarska, A.; Bamburowicz-Klimkowska, M.; Runden-Pran, E.; Dusinska, M.; Cimpan, M.R.; Rios-Mondragon, I.; Grudzinski, I.P. Extracellular Vesicles as Next-Generation Diagnostics and Advanced Therapy Medicinal Products. Int. J. Mol. Sci. 2024, 25, 6533. https://doi.org/10.3390/ijms25126533

AMA Style

Stawarska A, Bamburowicz-Klimkowska M, Runden-Pran E, Dusinska M, Cimpan MR, Rios-Mondragon I, Grudzinski IP. Extracellular Vesicles as Next-Generation Diagnostics and Advanced Therapy Medicinal Products. International Journal of Molecular Sciences. 2024; 25(12):6533. https://doi.org/10.3390/ijms25126533

Chicago/Turabian Style

Stawarska, Agnieszka, Magdalena Bamburowicz-Klimkowska, Elise Runden-Pran, Maria Dusinska, Mihaela Roxana Cimpan, Ivan Rios-Mondragon, and Ireneusz P. Grudzinski. 2024. "Extracellular Vesicles as Next-Generation Diagnostics and Advanced Therapy Medicinal Products" International Journal of Molecular Sciences 25, no. 12: 6533. https://doi.org/10.3390/ijms25126533

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop