Next Article in Journal
Hemostatic Alginate/Nano-Hydroxyapatite Composite Aerogel Loaded with Tranexamic Acid for the Potential Protection against Alveolar Osteitis
Next Article in Special Issue
Membrane Vesicles Derived from Gut Microbiota and Probiotics: Cutting-Edge Therapeutic Approaches for Multidrug-Resistant Superbugs Linked to Neurological Anomalies
Previous Article in Journal
Activation of Somatostatin-Expressing Neurons in the Lateral Septum Improves Stress-Induced Depressive-like Behaviors in Mice
Previous Article in Special Issue
Extracellular Vesicles: A New Frontier for Cardiac Repair
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exosomes as CNS Drug Delivery Tools and Their Applications

1
College of Pharmacy, Nankai University, Tongyan Road, Haihe Education Park, Tianjin 300350, China
2
College of Life Sciences, Nankai University, Weijin Road, Nankai District, Tianjin 300350, China
3
State Key Laboratory of Medicinal Chemical Biology, Nankai University, Tongyan Road, Haihe Education Park, Tianjin 300350, China
*
Authors to whom correspondence should be addressed.
Pharmaceutics 2022, 14(10), 2252; https://doi.org/10.3390/pharmaceutics14102252
Submission received: 23 September 2022 / Revised: 13 October 2022 / Accepted: 19 October 2022 / Published: 21 October 2022
(This article belongs to the Special Issue Advances of Membrane Vesicles in Drug Delivery Systems)

Abstract

:
Central nervous system (CNS) diseases threaten the health of people all over the world. However, due to the structural and functional particularities of the brain and spinal cord, CNS-targeted drug development is rather challenging. Exosomes are small cellular vesicles with lipid bilayers that can be secreted by almost all cells and play important roles in intercellular communication. The advantages of low immunogenicity, the ability to cross the blood-brain barrier, and the flexibility of drug encapsulation make them stand out among CNS drug delivery tools. Herein, we reviewed the research on exosomes in CNS drug delivery over the past decade and outlined the impact of the drug loading mode, administration route, and engineered modification on CNS targeting. Finally, we highlighted the problems and prospects of exosomes as CNS drug delivery tools.

1. Introduction

The drug delivery system (DDS) can comprehensively regulate the distribution of drugs in the organism in space, time, and dose. It has multiple roles: (i) drug targeting; (ii) slow and controlled drug release; (iii) enhancing drug stability and regulating drug metabolism time; (iv) promoting drug absorption and passage through biological barriers. Modern drug delivery technology started with the advent of extended-release capsule technology in 1952, followed by oral and transdermal formulations, long-acting injectables, pegylated liposomes, drug-polymer complexes, antibody-drug conjugates, and nano-delivery systems over 70 years [1].
Delivery of drugs to the CNS has been a challenging subject to overcome. In the brain, the main biological barrier is the blood-brain barrier (BBB). Various pathways have been developed to overcome the poor penetration of the BBB [2,3,4,5,6]. However, so far, it remains a bottleneck in CNS drug development [7]. To overcome the BBB, lots of drug delivery systems have been developed, such as adeno-associated viral (AAV) vectors [8], erythrocyte membrane-encapsulated nanocarriers [9], cell-based delivery nanocarriers [10], extracellular vesicles (EVs) [11], injectable hydrogels [12], and immunomodulators [13]. Among them, EVs performed surprisingly well.
EVs are small membranous vesicles released into the extracellular matrix by almost all cells and are widely present in various body fluids [14]. It was initially thought that EVs were cell-secreted garbage bags as a cellular self-cleaning mechanism. Nevertheless, as research progressed, EVs were found to stably carry important signaling molecules such as proteins, mRNAs, miRNAs, and lipids, which have important roles in intercellular communication [15,16], disease diagnosis [17,18,19], disease development [20,21], and disease therapy [22,23]. EVs are a highly heterogeneous group of vesicles, which are broadly classified into three categories according to their biogenesis, size, and biophysical property: exosomes, microvesicles, and apoptotic bodies [14,24,25,26]. Among them, exosomes have received the most attention in recent decades. Exosomes are approximately 40–100 nm in diameter and float at a density of 1.13–1.19 g∙mL−1 in sucrose density gradient solutions [27,28,29,30]. The biogenesis process of exosomes is shown in Figure 1. It is worth noting that, according to the MISEV 2018 guidelines, it is difficult to fully purify and distinguish the above EVs due to methodological difficulties in isolation, and the term “exosomes”, even though widely used, is suggested to be replaced by the term “small extracellular vesicles (sEVs)” instead [31].
Considering the important function that exosomes contain diverse active molecules (Figure 2) and are capable of mediating cellular communications, they were vigorously developed as DDSs for the treatment of various diseases [32]. While liposomes have been used maturely in DDSs, we still cannot ignore their limitations as well as other synthetic nanoparticles, such as toxicity, immunogenicity, and short half-life due to rapid clearance by the mononuclear phagocytic system (MPS) [33]. Although PEG modification can improve the above defects to a certain extent, it may cause the acceleration of the blood clearance (ABC) phenomenon after repeated multiple dosing [34,35,36]. In addition, studies have shown that exposure to chemicals and cosmetics in daily life could increase PEG antibodies in the human body [37]. In contrast, when exosomes were used in DDSs, they had better biocompatibility and lower immunogenicity due to characteristics such as an endogenous biological origin [38], and longer circulation times [39]. Moreover, exosomes from different sources had different targeting capabilities and roles in crossing biological barriers [40,41].
Since the focus of this review was the study on exosomes as CNS drug delivery tools, all authors reached a consensus on article screening. More specifically, the review was performed in PubMed and Web of Science databases using the keywords (exosomes/extracellular vesicles/CNS/brain/spinal cord/drug delivery/therapeutic molecules). At the same time, articles of potential interest should include at least three of the above keywords in their title or abstract.

2. The Roles of Exosomes in the CNS

The dense BBB protects the brain from various pathogenic microorganisms but is also the biggest obstacle to treating CNS disease. Thus, designing a safe, effective, and specific DDS to cross the BBB is the focus of current research. Numerous drug delivery technologies have been devoted to crossing the BBB [42,43,44], among which exosomes exhibited more unique roles and advantages.

2.1. The Blood-Brain Barrier and Its Challenges in the Treatment

The BBB consists of brain capillary endothelial cells (BCECs), the basement membrane of pericytes, and astrocyte end-feet [45,46,47]. In addition, the presence of enzymes and efflux pumps (e.g., P-glycoproteins and multi-drug resistance proteins (MRPs)) restricts non-specific translocation [48]. The current approaches to facilitate the passage of therapeutic compounds across the BBB can be summarized in the following three methods: (i) invasive methods; (ii) pharmacological methods; (iii) physiological methods [49]. Intracerebral or intrathecal injection directly opens the blood-brain barrier, which is highly invasive and makes it easy to cause tissue trauma or inflammation. The pharmacological approach emphasizes drug modification. However, it is not sufficient and requires the assistance of delivery tools to maximize the drug’s effect. In the physiological approach, various nano-delivery systems including exosomes are used to cross the BBB.

2.2. Multiple Functions of CNS Cell-Derived Exosomes

Microglia, astrocytes, oligodendrocytes, and neurons can actively secrete exosomes [50,51,52,53] and play important roles in intercellular communications [54]. In the CNS, exosomes propagate signals not only over short distances within cells but also throughout the brain via the cerebrospinal fluid (CSF) [55]. Therefore, exosomes are involved in the maintenance of the normal physiological state of the CNS as well as in the development of pathology [56,57].
As the most fundamental structural and functional unit of the CNS, neurons play important roles in maintaining neural homeostasis. It was found that exosomes have astonishing efficacy in neurogenesis and repair [58]. Additionally, they can mediate intra-neural communication and make it better [59]. The delivery of miR-132 to vascular endothelial cells by neuronal exosomes could maintain cerebral vasculature integrity [60]. Similarly, neuronal-derived extracellular vesicles could also detect insulin/mTOR pathway changes in the brain of Down syndrome (DS) patients [61].
Microglia are important neural immune cells with active immune sensing functions [62]. Furthermore, exosomes secreted by microglia are actively involved in physiological and pathological processes in the CNS. For example, after traumatic brain injury, upregulation of miR-124-3p in microglia exosomes suppressed neuronal inflammation and promoted nerve growth via transfer to neurons [63]. Microglia could sense abnormal signals and convert from the inflammation-suppressing M2 type to the pro-inflammatory M1 type in pathologic conditions such as trauma and inflammation. Nevertheless, the involvement of exosomes may reverse this process, as exosomes from mesenchymal stem cells attenuated CNS inflammation and demyelination in EAE rats by regulating microglia polarization [64].
One important mediator of cell-to-cell communication is miRNA. For example, exosomes could transfer miR-133b from pluripotent mesenchymal stromal cells (MSC) to neuronal cells to promote nerve growth [65]. MSC-origin exosomes contained miRNAs that regulate neuroinflammation [66]. Exosomes secreted from glia and neurons mediated Alzheimer’s disease (AD) development via miRNA delivery. Changes in exosomal miRNAs in the peripheral blood or cerebrospinal fluid of patients were important guides for the early diagnosis of AD. More information on the roles of exosomal miRNAs in the diagnosis and treatment of AD have been reported [67].
Astrocytes are involved in the BBB formation, which may help regulate the ionic microenvironment surrounding neurons. Nevertheless, it has been shown that increasing the release of their exosomes can mitigate β-amyloid-induced neurotoxicity [68], offering novel ideas for the treatment of AD. However, the release of astrocyte exosomes is not always beneficial, and it has been shown that the microenvironment induced by astrocyte exosomal microRNA induced PTEN loss and brain metastatic growth of tumors [69].
Neural stem cells, in contrast to the aforementioned CNS cells, have the capacity for self-renewal and can differentiate into neurons, astrocytes, oligodendrocytes, etc. Neural stem cell exosomes held great promise in the treatment of neurodegenerative diseases [70], e.g., human neural stem cell exosomes have a protective effect against Parkinson’s disease [71] and could ameliorate AD [72]. Alongside this, extracellular vesicles derived from human neural stem cells have been shown to have good efficacy in the functional recovery of strokes in mice [73].

3. Strategies for Drug Loading into Exosomes

Although naked exosomes have therapeutic effects, they are limited. Drug-loaded exosomes can synergize the effects of drugs and exosomes, which have become a hotspot in DDS research.
The drug delivery system consists of two main components: one is the “outer shell”, as well as the drug delivery vehicle, which in this review refers to as exosomes. The other part is the “content”, which is also called the therapeutic drug. Currently, almost 20% of exosome-related topics are focused on drug loading, demonstrating its importance.
In contrast to the nano-formulations already developed, exosomes are loaded with active molecules more flexibly and diversely. Synthetic drug delivery vehicles like nanomaterials can be loaded with drugs during the synthesis. Exosomes, on the other hand, seem to be more receptive to the idea of loading active molecules before or after their secretion.

3.1. Pre-Secretory Loading

The idea of loading drugs before exosome secretion, i.e., at the biogenesis stage of exosomes, is to manipulate exosome-generating progenitor cells, which is also called cell engineering. The donor cells are capable of secreting exosomes containing the drugs.
Drugs can enter cells by transfection, e.g., transfecting HEK293T cells with circSCMH1 plasmid [74] can efficiently encapsulate them into secreted exosomes, and targeted delivery to the brain can enhance neuronal plasticity and inhibit glial reactivity, thus effectively ameliorating stroke symptoms in mice and rhesus monkeys. Ge et al. [75] transfected miR-124-3p mimics into microglia for 6 h, and exosomes containing miR-124-3p were extracted and isolated 48 h after the culture medium was changed. Intravenous injection of the above exosomes alleviated the neurodegenerative lesions in repetitive mild traumatic brain injury (rmTBI) mice.
Drugs can also be co-incubated with donor cells to gain access to the cell interior. For instance, macrophage RAW264.7 cells co-cultured with high doses of curcumin at 37 °C for 24 h were isolated by ultracentrifugation to produce exosomes containing curcumin and were effective in alleviating symptoms of middle cerebral artery occlusion (MCAO) in vivo [76].
The method of drug loading by manipulating donor cells before exosome secretion is simple to operate. However, this strategy has high requirements for the compatibility between parental cells and a drug [11], and a variety of uncontrolled factors such as cell status and culture conditions ultimately lead to unstable or low loading efficiency [77,78]. In conclusion, a lack of knowledge of the complex intracellular metabolic pathways is the main reason. A better understanding of the mechanisms by which therapeutic molecules are transported in cells and localized to exosomes may improve loading efficiency.

3.2. Post-Secretory Loading

The post-secretory approach refers to the way that naive exosomes are extracted and isolated from donor cells, and then the purified exosomes are loaded with drugs by various methods in vitro. Some common approaches include co-incubation, electroporation, sonication, repeated freeze-thawing, extrusion, and so on (Figure 3).
Exosomes have lipid bilayers, so co-incubation of exosomes with hydrophobic compounds allows passive entry of drugs into exosomes. For example, aptamer F5R1 was co-incubated with exosomes derived from the human embryonic kidney (HEK) 293T cells at 37 °C for some time [79], and an aptamer was able to efficiently enter exosomes and reduce a-synuclein aggregation, exerting a therapeutic effect on PD. It is worth mentioning that drug saturation solutions can be used to improve the efficiency of drug entry in vitro. This idea was applied by Qu et al. [80] to the experiment of dopamine loading in mouse blood exosomes. The enormous concentration difference contributed to the passive penetration of dopamine into exosomes during co-incubation. Dopamine-loaded exosomes showed a PD therapeutic effect. Co-incubation loading efficiency is related to drug hydrophobicity, molecular weight, exosome size, the lipid content of exosomes, and so on [81]. Therefore, the appropriate mixing ratio must be mastered to achieve higher loading efficiency.
Electroporation enables the membranes of exosomes to temporarily create pores in the presence of an electric current, through which the drugs enter the exosomes. For example, Zhang et al. [82] loaded ZIKV-specific siRNA into exosomes derived from HEK 293T cells at 450 V and 100 mF. They could cross the BBB and placental barrier to treat fetal microcephaly caused by ZIKV infection. Electroporation charging efficiency can be affected by voltage, electrical capacitance, discharge time, and other conditions.
Sonication is also commonly used. Luo et al. [83] isolated exosomes from fresh bovine milk and encapsulated epicatechin gallate (ECG) into exosomes using ultrasonication with a loading efficiency of 25.96%. The ultrasonically loaded exosomes were anti-apoptotic and neuroprotective against Parkinson’s disease. The loading efficiency of sonication is high but may lead to the aggregation of exosomes.
Aside from the abovementioned modalities, chemical reactions have also been skillfully applied [84]. LJM-3064 potently inhibits the inflammatory response and reduces the areas of demyelinating lesions in CNS by chemically covalently binding to amine groups on exosome surfaces via EDC/NHS. There are also alternative drug loading methods using exosome transfection reagents, which can directly transfer nucleic acids including siRNAs, miRNAs, mRNAs, and even plasmid DNAs into isolated exosomes [85]. These novel ways provide new ideas that may expand the development of exosomal drug delivery.
The addition of some adjuvants like saponin will improve loading efficiency appropriately. For example, in loading catalase, the authors [86] investigated the drug loading efficiency with and without saponin separately when using co-incubation as a modality and demonstrated that the addition of saponin significantly increased the drug loading efficiency from 4.9% to 18.5%. It was worth mentioning that the study also explored comparing the differences in drug loading rates of four modalities loaded on the same drug, and it was found that the loading efficiencies of sonication and extrusion were higher, 26.1% and 22.2%, respectively. However, this may not apply to other research, and the optimal method still needs to be explored personally in the respective studies. In some cases, multiple methods may be used at the same time to improve the efficiency of drug loading. For example, to load curcumin into mouse embryonic stem cell (MESC)-derived exosomes, Anuradha et al. [87] first incubated curcumin with exosomes in a 1:4 ratio for 15 min at room temperature, followed by rapid freeze-thawing two to three times. Successfully loaded exosomes were capable of reducing vascular inflammation and promoting neurovascular recovery in mice with ischemia-reperfusion injury.
In conclusion, efficient drug loading remains one of the biggest challenges. Especially for small molecules or extremely hydrophobic molecules, although simple co-incubation is the most commonly used method, it no longer meets the requirements. Therefore, Haney et al. [88] explored the optimization of co-incubation conditions. The team found that when the pH of the co-incubation system was adjusted to 8.0, the drug loading rate increased significantly. The possible reason was that a pH close to pI reduced the molecule’s charge and increased its hydrophobicity. It is reasonable to assume that optimizing other incubation conditions such as appropriate temperature increase, shaking, longer incubation time, etc. may also improve the loading rate. In addition, other suitable methods can be directly used instead of co-incubation. For example, mild sonication could significantly increase the loading rates of small molecules such as doxorubicin [88], paclitaxel [88,89], and gemcitabine [90].
Drug loading after secretion is controlled and efficient. Disadvantages, however, include destruction of exosome integrity, facile aggregation, etc. The character of the drug and the properties of exosomes may all influence the loading efficiency [91]. Therefore, the choice of drug loading mode is ultimately based on experimental needs.

4. Drug Administration Affects the Efficiency of Exosomes into the CNS

In DDS, appropriate changes in administration modes may enhance the therapeutic effect. Exosomes biodistribution was determined by cellular origin, route of administration, and targeting [92,93]. Thus, intravenous [94,95,96], nasal [97,98,99,100], intraperitoneal [101,102,103], and local [104] injections have been used extensively.
When exosomes are used as drug delivery tools, the most common administration is intravenous injection. For example, brain endothelium-derived exosomes might reach the brain after intravenous injection to deliver doxorubicin for the treatment of brain cancer in zebrafish [105]. Nevertheless, intravenously injected exosomes have restricted CNS tropism. In addition to this, the half-life of intravenous exosomes is shorter. Therefore, most of them are still surface-modified. One of the most widely used is the modification of neuron-targeting peptides (RVG) to enhance their brain-targeting capacity.
Intranasal injection allows drug delivery to the brain via intranasal deposits, olfactory bulbs, trigeminal nerves, and the respiratory epithelium [106,107]. It is therefore widely used in brain disorders [108]. For example, exosomes containing cholesterol-modified AMO181a were capable of brain delivery and exerted therapeutic effects on ischemic brain injury when administered intranasally [98]. MSC-exosomes injected intranasally could cross the BBB and migrate to the injured area of the spinal cord, achieving recovery [109].
The local administration has also attracted attention. In a recent study, Han et al. [104] explored the use of an array of patches using autonomously designed gel microneedles loaded with 3D cultured MSC-exosomes to achieve in situ repair of exosomes in spinal cord injury. MSC-exosomes could reduce the neuroinflammatory response, promote polarization of microglia to M2 phenotype, and reduce glial scar formation to facilitate the repair of neurological injury. This successful exploration will also bring more researchers to think, innovate, and achieve breakthroughs concerning the exosomes administrations.

5. Engineering Exosomes Enhance the CNS Targeting and Therapeutic Efficacy

Exosomes inherit the characteristics of donor cells [110] and have certain passive targeting abilities, also known as a homing ability [111,112]. However, this natural homing ability is weak and may be affected by various factors such as injection methods, drug doses, and individual differences [92], thereby weakening or even losing the CNS targeting capacity of exosomes. Moreover, it has been demonstrated that intravenously injected naked exosomes were easily enriched in reticuloendothelial system (RES) organs such as the liver and it was extremely difficult for them to reach the CNS [92,93].
Engineered exosomes are endowed with desired targeting capabilities because they are decorated with various functional molecules on their surfaces [113,114,115,116]. Functional molecules can utilize specific ligand/receptor binding strategies to increase brain targeting capabilities [110,117]. Strategies for surface modification of engineered exosomes include genetic engineering and chemical modification [117,118].
Genetic engineering modification of exosomes is the fusion of the gene sequence of a target functional protein or peptide with a selected exosomal membrane protein, and then donor cells transfected with the above plasmids secrete engineered exosomes with targeting ligands on their surface [119]. Lamp2b-RVG is the most widely used hybrid plasmid for genetically modified brain-targeted exosome modifications. Lamp2b is a highly abundant membrane-localized protein on exosomes. Rabies virus glycoprotein (RVG) is the more recognized brain-targeting peptide, and RVG selectively targets neurons and brain endothelial cells by binding to the nicotinic acetylcholine receptor (nAChR) [120]. For example, Alvarez-Erviti et al. [118] genetically engineered Lamp2b-RVG plasmids and transfected autologous-derived dendritic cells to make them secrete engineered exosomes. Intravenous injection of the above exosomes specifically delivered siRNA to neurons, microglia, and oligodendrocytes in the brain, leading to specific gene knockdown and offering new hope for AD treatment. Genetic engineering strategies are currently widely used, but the procedures are complex and costly.
The chemical modification enables in vitro manipulation after exosome isolation. For example, stearoyl-RVG was inserted into the phospholipid bilayer of immature DC exosomes via hydrophobic interactions. The drug-loaded exosomes were able to scavenge α-synuclein aggregates and reduce cytotoxicity in PD neurons [121]. Han et al. [122] synthesized RVG-15-PEG-DSPE and inserted it onto the surface of exosomes containing doxorubicin to achieve the treatment of glioblastoma. In addition to RVG, other types of brain-targeting small molecules can also be engineered by chemical ligations. For example, to attach RGD peptides to the surface of exosomes, Tian et al. [123] first reacted naive exosomes isolated from MSC cells with DBCO-sulfo-NHS to produce EXO-DBCO, which was chemically reacted with the exposed amino groups on the surface of exosomes using NHS. Then, EXO-DBCO was reacted with the azide-modified cRGD in a click reaction, so the azide-modified cRGD was successfully attached to EXO-DBCO. Finally, the attachment of the targeted functional peptide cRGD to the surface of the exosomes by the chemical reaction was achieved (known as the bioorthogonal reaction) [124,125,126,127]. Jia et al. [128] also modified cargo-loaded exosomes with the glioma-targeting peptide RGE by click chemistry and finally obtained engineered exosomes with therapeutic effects on glioma. The chemical modification strategy is less costly and has higher linkage efficiency, but has more influencing factors such as temperature, solvent, etc.
Although current engineered exosomes focused on increasing targeting, Kojima et al. [129] pioneered the idea of increasing exosome release. Specifically, Kojima’s team constructed a triple hybrid plasmid with STEAP3 capable of participating in exosomes biogenesis, and syndecan-4 (SDC4) capable of supporting endosomal membrane outgrowth to form polycystic and L-aspartate oxidase (NadB) capable of promoting cellular metabolism by regulating the citric acid cycle. The combined expression of these genes significantly increased exosome production.
The aim of engineered exosomes is to overcome some deficiencies of naked exosomes and enable them to better serve as CNS drug delivery tools.

6. Research and Applications of Exosomal Drug Delivery Systems in CNS Diseases

As exosomes have been studied more intensively, numerous studies have been conducted on various CNS disease models to assess the efficacy of exosomal drug delivery systems and help their clinical translation.

6.1. Alzheimer’s Disease

Alzheimer’s disease (AD) is a progressive neurodegenerative disorder manifested by memory loss and cognitive decline [130]. There are no available therapeutic agents and the current drugs merely alleviate the associated symptoms. Neurogenic fibrillary tangles (NFT) composed of amyloid β (Aβ) plaques and abnormally phosphorylated tau protein (p-tau) are the major pathological hallmarks and AD is accompanied by neuronal death, synaptic loss, neuroinflammation, and brain injury [131,132,133].
Exosomes contain a wide variety of proteins, RNAs, etc., and therefore can mediate intercellular communication [134]. Because of this, exosomes have excellent performance in delivering RNA [135,136,137]. Small interfering RNAs (siRNAs) are among the many RNA drugs widely used, which result in gene silencing [138]. In recent years, siRNAs delivery technology has advanced rapidly, but extrahepatic and especially brain siRNA-targeted delivery has still been in the exploratory stage. Alvarez-Erviti et al. [118] made the first attempt to achieve intracerebral targeted delivery using exosomes as siRNA delivery vehicles. This study used self-derived dendritic cell exosomes to minimize immunogenicity, as well as RVG neuro-targeting peptides to modify exosomes. Suitable siRNAs were designed for the AD therapeutic target BACE1, resulting in a strong knockdown of the corresponding mRNA and proteins in mice.

6.2. Parkinson’s Disease

PD is the second most common neurodegenerative disorder worldwide, the major pathological hallmark being dopaminergic (DA) neuron degenerative death [139,140,141].
Catalase has been shown to have a protective effect on neurons. With this in mind, Haney et al. [86] explored the use of different methods for loading catalase into RAW264.7 cell exosomes. After intranasal administration, catalase-loaded exosomes were taken up by neurons and microglia of PD mice and exhibited significant neuroprotective effects. Likewise, Kojima et al. [129] used a mixed plasmid containing peroxidase mRNA along with RVG to transfect HEK-293T cells. By subcutaneous injection of the cells, the cells produced exosomes containing peroxidase mRNA in vivo and delivered it to the brain and alleviated the neurotoxicity and neuroinflammation caused by PD.
Alpha-synuclein aggregation is closely related to Lewy body formation and dopaminergic neuron death and plays an important role in the pathogenesis of PD. Thus, the reduction of α-synuclein protofibril aggregation is an idea for PD therapy. Cooper and colleagues [142] showed a significant reduction in α-Syn mRNA and protein levels in the brains of both normal and PD mice treated with RVG exosomes loaded with anti-α-Syn siRNA. This team targeted the reduction of α-synuclein aggregation again [143] using primary dendritic cell-derived, RVG-modified exosomes encapsulating the therapeutic molecule shRNA-MCs, which were specifically delivered to the brains of PD mice by intravenous injection and attenuated α-synuclein aggregation, reducing the loss of dopaminergic neurons and thus improving PD-related symptoms. It is noteworthy that the authors considered that shRNA size could affect drug delivery efficiency during exosome electroporation and specifically designed a smaller loop shape, i.e., shRNA-MCs, to achieve better loading efficiencies and therapeutic effects. This novel approach provided insights into the exosomal drug delivery systems for the treatment of other neurodegenerative disorders. Along similar lines, RVG-exosomes loaded with aptamer F5R1 were also able to deliver to neurons in vitro and in vivo, significantly reducing the pathological aggregations induced by α-synuclein pre-formed fibers (PFFs) [79].
Dopamine is also used for PD in addition to the treatment options mentioned above, but dopamine reaching the brain is problematic. Qu et al. [80] used blood exosomes to encapsulate dopamine and successfully delivered it to the striatum and substantia nigra of the brains by intravenous injection to ameliorate symptoms of PD mice.

6.3. Huntington’s Disease

Huntington’s disease (HD) is a hereditary neurodegenerative disorder [144]. Disease-causing genes, also known as Huntington’s genes, produce abnormal Huntington’s proteins which tend to adhere and aggregate, ultimately resulting in neuronal cell death [145,146]. Thus, silencing this gene is a strategy for HD therapy. Since exosomes have been widely used for siRNA delivery [147], Didiot et al. [148] modified siRNA hydrophobically to improve loading efficiency when incubated with exosomes. After unilateral infusion of exosomes into the mice striatum, exosomes carried siRNA for diffusion to the contralateral striatum and transported it to neurons, ultimately resulting in dose-dependent silencing of Huntingtin mRNA and protein in bilateral ventricles.

6.4. Stroke

A stroke is a general term for brain tissue damage caused by sudden rupture or blockage of cerebral blood vessels, including ischemic and hemorrhagic stroke [149,150,151]. It is a neurological disease with a high rate of disability worldwide [152,153].
Subarachnoid hemorrhage is a major reason for hemorrhagic stroke. In one study [154], miR-193b-3p was delivered to the brains of mice by intravenous injection of exosomes, which ameliorated inflammatory response through inhibition of HDAC3. Lastly, neurobehavioral disturbances, brain edema, BBB injury, and neurodegeneration were alleviated.
Compared with hemorrhagic stroke, ischemic stroke accounts for the vast majority of the total number of strokes [155] and the main causes are occlusion and stenosis of the associated arteries. MiRNAs were involved in regulating nervous system development and function [156,157] and their therapeutic potential has been demonstrated in a variety of CNS diseases [158,159]. MSC-exosomes showed superior effects in the treatment of various neurological disorders [160,161,162], as evidenced by neural repair and inhibition of neuroinflammation [95,163]. Therefore, several studies [164,165,166] have explored the use of MSC-exosome-encapsulated miRNAs as therapeutic molecules for the treatment of ischemic strokes. For example, Zhao et al. [164] transfected MSCs with miR-223-3p to generate miRNA-loaded exosomes. After intravenous injection, the volume of cerebral infarction was greatly reduced, and neurological function was markedly improved. Its protective effect was closely associated with an increased polarization of microglia towards the M2 phenotype. MiR-124 also biased neural progenitor cells toward neuronal lineage through exosomes, greatly contributing to neurogenesis after ischemia [165]. Similarly, miR-17-92-enriched exosomes improved neurological function and promoted neurogenesis [166].
Curcumin is a natural polyphenolic compound with a potent anti-inflammatory effectivity [167,168]. RGD peptides were bound to the surface of MSC-exosomes by bioorthogonal chemistry, and curcumin entered the exosomes after co-incubation. Drug-loaded exosomes targeted microglia, neurons, and astrocytes in the brain following intravenous injection and effectively inhibited inflammatory response and apoptosis in ischemic brain injury areas [123]. Whereas the aforementioned brain targeting of exosomes was RGD mediated, He et al. [76] took advantage of the targeted migration capability of mouse macrophage RAW264.7 exosomes themselves, loading curcumin as a therapeutic agent as well. After intravenous injection, the exosomes targeted ischemic areas and exerted powerful neuroprotective effects by eliminating ROS generation and mitigating BBB destruction. Exosomes from mouse embryonic stem cells (MESC) may also bind curcumin for stroke treatment [87].
Aside from the curcumin and miRNAs mentioned above, other active molecules such as circSCMH1 [74], recombinant human NGF mRNA [169], and pigment epithelium-derived factor (PEDF) [170] could also be coated with exosomes as therapeutic molecules for strokes treatment.

6.5. Brain or Spinal Cord Injury and Neuroinflammation

Brain or spinal cord injury is a neuroinflammatory condition that endangers human health. In spinal cord injury (SCI), MSC-exosomes acquired the homing property of its progenitor cells and were able to automatically penetrate the BBB to reach the injured spinal cord region. Taking advantage of this property, Guo et al. [109] repaired SCI after intranasal injection of MSC-exosomes containing phosphatase and tensin homolog siRNA (PTEN-siRNA), which effectively reduced neuroinflammation and glial proliferation. In addition to SCI, repetitive mild TBI (rmTBI) not only triggered neuroinflammation but may also have led to long-term neurodegenerative diseases like AD [171]. Microglia, as specific resident macrophages in the CNS, play key roles in the development and regression of neuroinflammation [172,173]. One study [75] reported that intravenous administration of microglia exosomes harboring miR-124-3p could target neurons and alleviate neurodegeneration. Multiple sclerosis (MS) is a classic neuroinflammatory disorder mainly characterized by inflammatory demyelinating lesions in the white matter of the CNS [174]. The LJM-3064 aptamer, which was chemically covalently bound to the surface of MSCs exosomes, significantly ameliorated neuroinflammation by reducing areas of demyelinating lesions in the spinal cord and brain [84]. Zhuang et al. [175] verified the therapeutic effect of the exosomes’ drug delivery system in three neuroinflammatory models: LPS-induced encephalitis models, experimental autoimmune encephalomyelitis (EAE) models, and brain-tumor-induced neuroinflammation. Exosomes loaded with curcumin or JSI124 (a signal transducer and activator of the transcription 3 inhibitor) could target microglia and induce microglia apoptosis to alleviate neuroinflammation after intranasal injection.

6.6. Brain Tumor

Gliomas account for 40% to 50% of brain tumors and are the most common primary intracranial tumors. Glioma has a high mortality, disability, and very poor prognosis [176]. A major reason why current drugs have difficulty in achieving the desired therapeutic effect is the difficulty in crossing the BBB to accumulate sufficient local drug concentrations at the tumor site. Exosomes, as drug delivery systems capable of crossing the blood-brain barrier, have played important roles in glioma treatment. For example, a recent study developed exosomes modified with Angiopep-2 and transcriptional peptide trans-activators and loaded doxorubicin to reach gliomas by crossing the BBB, not only improving the survival time of glioma mice by more than 2-fold but also greatly reducing the toxic effects of doxorubicin [177]. Similarly, Jia et al. [128] loaded curcumin and SPIONs (MRI contrast agents) into exosomes simultaneously by electroporation and then modified the exosomes with glioma-targeting peptides by click chemistry, ultimately showing dual antitumor effects.
In tumor therapy, overcoming drug resistance is also a major task. Glioblastoma multiforme (GBM) often exhibit chemo- and radio-resistance [178]. Yamada’s group found that the main reason for the resistance of GBM to the chemotherapeutic drug temozolomide was the upregulation of miR-9 in cells, which was involved in the expression of the P-glycoprotein, a drug efflux transporter. Given this, the authors used mesenchymal stem-cell-derived exosomes carrying anti-miR-9 to eventually reverse the expression of multidrug transporter proteins and sensitize GBM cells to temozolomide [179].
Apart from gliomas, brain tumors contain astrocytoma, meningioma, neurofibroma, and so on [180]. Because of the complex and precise structures in the brain, whatever type of cytopathic brain tumor, once compressed to any part of the brain, will cause damage to different functions in the body. For the treatment of brain tumors, exosome delivery systems continue to perform well. For example, brain endothelium-derived exosomes could deliver the anticancer drug doxorubicin via the BBB for the treatment of brain cancer in zebrafish [105].

6.7. Other Brain Diseases: Viral Infection; Drug Addiction

Drug addiction is highly damaging to individuals and society, mainly because of a series of withdrawal symptoms that occur when the drug stops stimulating the CNS. Nevertheless, there are very limited treatments available [181]. Among them, opioids mediate analgesic and addictive effects through MOR receptors. Thus, Liu’s team boldly attempted to use exosomes derived from RVG-modified HEK 293T cells to deliver MOR siRNA specifically to the mouse brain and downregulate MOR expression to mediate morphine addiction treatment [182].
Viruses can invade the organism and proliferate in the host cells through multiple pathways. Even some viruses invade the CNS, such as the ZIKA virus, which could cross the placenta and BBB to enter the fetal brain and cause fetal microcephaly [82]. Zhang et al. [183] designed an RVG-modified exosome derived from HEK 293T cells, which contained ZIKV-specific siRNA. Once intravenously injected, exosomes could cross the mice’s placental barrier from the maternal circulatory system to the fetal mice and cross the fetal BBB to target the head, thus protecting the fetal mice from ZIKV virus infection and alleviating neuroinflammation.
In summary, exosomes have been widely used as CNS drug delivery tools. They can be loaded with a variety of therapeutic molecules, including nucleic acids, proteins, and small molecule compounds. Exosomal drug delivery systems have shown good results in treating almost all CNS diseases, from neurodegenerative diseases to neuroinflammatory diseases, from brain tumors to brain viral infections. This greatly encouraged researchers to continue to explore and utilize exosomes as drug delivery tools for the treatment of CNS diseases. Additionally, the studies on exosomal drug delivery systems for CNS diseases are summarized in Table 1.

7. Conclusions and Perspectives

According to the latest database (http://www.exocarta.org/, accessed on 23 September 2022), 9769 proteins, 3408 mRNAs, and 2838 microRNAs have been identified in exosomes. These rich contents make the role of exosomes in intercellular communication, disease development, and therapy increasingly expansive. It brings light to exosomes as CNS drug delivery tools. To date, exosome studies have involved a number of species including mice, rats, zebrafishes, and monkeys (https://evtrack.org/, accessed on 23 September 2022), which will drive the clinical translation of exosomes.
As CNS drug delivery tools, exosomes have several characteristics compared to synthetic nanoparticles: low immunogenicity, non-toxicity, high cargo carrying, protective capabilities [184,185,186], and the ability to cross the blood-brain barrier [187,188,189]. To improve the efficiency of entering the CNS, exosome surface engineering modification has been widely used. It can increase the local concentration of drugs at the lesion site and reduce toxic side effects while improving therapeutic efficacy. However, the structural and functional stability of engineered exosomes and the mechanism of improving CNS targeting efficiency have not been thoroughly studied. Therefore, more research is needed to enhance the stability, security, and standardization of these strategies.
There are various types of active molecules encapsulated in exosomes, including nucleic acids (miRNAs, siRNAs, shRNAs, mRNAs, etc.), small molecule compounds (doxorubicin, dopamine, etc.), natural products (curcumin, paclitaxel, resveratrol, etc.), and proteins (catalase, pigment epithelium-derived factor, etc.). For different types of therapeutic drugs and exosomes, various loading methods exist, both before and after the secretion of exosomes. Genetic engineering, co-incubation, electroporation, sonication, extrusion, and repeated freeze-thawing have been employed. On this basis, we need to further investigate the mechanisms of exosome internalization to better understand how they work and thus select suitable drug loading methods for different types of therapeutic molecules. There has been increasing evidence that exosomes carrying the therapeutic molecules mentioned above have made breakthroughs in various CNS diseases such as brain tumors, neurodegenerative diseases, multiple scleroses, brain or spinal cord injuries, strokes, drug addictions, and viral infections. However, progress in the isolation and purification of exosomes as well as mechanistic studies are needed to fully realize the potential of exosomes in CNS drug delivery systems.
In conclusion, engineering modifications, administration routes, type of therapeutic molecules, and encapsulation modes are all important parameters of exosomes as CNS drug delivery tools (Figure 4) and affect their clinical transformation. The majority of patients with CNS diseases will benefit if we can pay more attention to the problems already present, deepen our understanding of unexplained mechanisms, and realize their maximal clinical potential.

Author Contributions

Writing—original draft preparation, K.S.; writing—review and editing, K.S. and X.Z.; visualization, W.Z. and F.Y.; supervision, W.Z. and H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Key R&D Program of China (2018YFA0507204) and the National Natural Science Foundation of China (22077068).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CNScentral nervous system
RVGrabies virus glycoprotein
DDSdrug delivery system
BBBblood-brain barrier
AAVadeno-associated virus
EVsextracellular vesicles
MPSmononuclear phagocytic system
PEGpolyethylene glycol
ABCacceleration of blood clearance
CSFcerebrospinal fluid
MSCmesenchymal stem cell
SCIspinal cord injury
TBItraumatic brain injury
MSmultiple sclerosis
EAEexperimental allergic encephalomyelitis
RESreticuloendothelial system
PDParkinson’s disease
ADAlzheimer’s disease

References

  1. Park, H.; Otte, A.; Park, K. Evolution of drug delivery systems: From 1950 to 2020 and beyond. J. Control. Release 2022, 342, 53–65. [Google Scholar] [CrossRef]
  2. Bally, M.B.; Harvie, P.; Wong, F.M.; Kong, S.; Wasan, E.K.; Reimer, D.L. Biological barriers to cellular delivery of lipid-based DNA carriers. Adv. Drug Deliv. Rev. 1999, 38, 291–315. [Google Scholar] [CrossRef]
  3. Begley, D.J. Delivery of therapeutic agents to the central nervous system: The problems and the possibilities. Pharmacol. Ther. 2004, 104, 29–45. [Google Scholar] [CrossRef]
  4. Deprez, J.; Lajoinie, G.; Engelen, Y.; De Smedt, S.C.; Lentacker, I. Opening doors with ultrasound and microbubbles: Beating biological barriers to promote drug delivery. Adv. Drug Deliv. Rev. 2021, 172, 9–36. [Google Scholar] [CrossRef]
  5. Henrich-Noack, P.; Nikitovic, D.; Neagu, M.; Docea, A.O.; Engin, A.B.; Gelperina, S.; Shtilman, M.; Mitsias, P.; Tzanakakis, G.; Gozes, I.; et al. The blood-brain barrier and beyond: Nano-based neuropharmacology and the role of extracellular matrix. Nanomedicine 2019, 17, 359–379. [Google Scholar] [CrossRef]
  6. Yang, R.; Wei, T.; Goldberg, H.; Wang, W.; Cullion, K.; Kohane, D.S. Getting Drugs Across Biological Barriers. Adv. Mater. 2017, 29, 1606596. [Google Scholar] [CrossRef]
  7. Wang, Y.I.; Abaci, H.E.; Shuler, M.L. Microfluidic blood-brain barrier model provides in vivo-like barrier properties for drug permeability screening. Biotechnol. Bioeng. 2017, 114, 184–194. [Google Scholar] [CrossRef]
  8. Chen, W.; Yao, S.; Wan, J.; Tian, Y.; Huang, L.; Wang, S.; Akter, F.; Wu, Y.; Yao, Y.; Zhang, X. BBB-crossing adeno-associated virus vector: An excellent gene delivery tool for CNS disease treatment. J. Control. Release 2021, 333, 129–138. [Google Scholar] [CrossRef]
  9. Castro, F.; Martins, C.; Silveira, M.J.; Moura, R.P.; Pereira, C.L.; Sarmento, B. Advances on erythrocyte-mimicking nanovehicles to overcome barriers in biological microenvironments. Adv. Drug Deliv. Rev. 2021, 170, 312–339. [Google Scholar] [CrossRef]
  10. Mendanha, D.; Vieira de Castro, J.; Ferreira, H.; Neves, N.M. Biomimetic and cell-based nanocarriers—New strategies for brain tumor targeting. J. Control. Release 2021, 337, 482–493. [Google Scholar] [CrossRef]
  11. Escude Martinez de Castilla, P.; Tong, L.; Huang, C.; Sofias, A.M.; Pastorin, G.; Chen, X.; Storm, G.; Schiffelers, R.M.; Wang, J.W. Extracellular vesicles as a drug delivery system: A systematic review of preclinical studies. Adv. Drug Deliv. Rev. 2021, 175, 113801. [Google Scholar] [CrossRef]
  12. Akhtar, A.; Andleeb, A.; Waris, T.S.; Bazzar, M.; Moradi, A.R.; Awan, N.R.; Yar, M. Neurodegenerative diseases and effective drug delivery: A review of challenges and novel therapeutics. J. Control. Release 2021, 330, 1152–1167. [Google Scholar] [CrossRef]
  13. Nash, A.; Aghlara-Fotovat, S.; Hernandez, A.; Scull, C.; Veiseh, O. Clinical translation of immunomodulatory therapeutics. Adv. Drug Deliv. Rev. 2021, 176, 113896. [Google Scholar] [CrossRef]
  14. Tkach, M.; Thery, C. Communication by Extracellular Vesicles: Where We Are and Where We Need to Go. Cell 2016, 164, 1226–1232. [Google Scholar] [CrossRef] [Green Version]
  15. Mathieu, M.; Martin-Jaular, L.; Lavieu, G.; Thery, C. Specificities of secretion and uptake of exosomes and other extracellular vesicles for cell-to-cell communication. Nat. Cell Biol. 2019, 21, 9–17. [Google Scholar] [CrossRef]
  16. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [Green Version]
  17. Chong, S.Y.; Lee, C.K.; Huang, C.; Ou, Y.H.; Charles, C.J.; Richards, A.M.; Neupane, Y.R.; Pavon, M.V.; Zharkova, O.; Pastorin, G.; et al. Extracellular Vesicles in Cardiovascular Diseases: Alternative Biomarker Sources, Therapeutic Agents, and Drug Delivery Carriers. Int. J. Mol. Sci. 2019, 20, 3272. [Google Scholar] [CrossRef] [Green Version]
  18. de Jong, O.G.; Kooijmans, S.A.A.; Murphy, D.E.; Jiang, L.; Evers, M.J.W.; Sluijter, J.P.G.; Vader, P.; Schiffelers, R.M. Drug Delivery with Extracellular Vesicles: From Imagination to Innovation. Acc. Chem. Res. 2019, 52, 1761–1770. [Google Scholar] [CrossRef] [Green Version]
  19. Pan, S.; Zhang, Y.; Natalia, A.; Lim, C.Z.J.; Ho, N.R.Y.; Chowbay, B.; Loh, T.P.; Tam, J.K.C.; Shao, H. Extracellular vesicle drug occupancy enables real-time monitoring of targeted cancer therapy. Nat. Nanotechnol. 2021, 16, 734–742. [Google Scholar] [CrossRef]
  20. Fan, J.; Wei, Q.; Koay, E.J.; Liu, Y.; Ning, B.; Bernard, P.W.; Zhang, N.; Han, H.; Katz, M.H.; Zhao, Z.; et al. Chemoresistance Transmission via Exosome-Mediated EphA2 Transfer in Pancreatic Cancer. Theranostics 2018, 8, 5986–5994. [Google Scholar] [CrossRef]
  21. Hutcheson, J.D.; Aikawa, E. Extracellular vesicles in cardiovascular homeostasis and disease. Curr. Opin. Cardiol. 2018, 33, 290–297. [Google Scholar] [CrossRef] [PubMed]
  22. Roefs, M.T.; Sluijter, J.P.G.; Vader, P. Extracellular Vesicle-Associated Proteins in Tissue Repair. Trends Cell Biol. 2020, 30, 990–1013. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, X.; Xie, F.; Wang, L.; Zhang, L.; Zhang, S.; Fang, M.; Zhou, F. The function and clinical application of extracellular vesicles in innate immune regulation. Cell. Mol. Immunol. 2020, 17, 323–334. [Google Scholar] [CrossRef] [PubMed]
  24. Gyorgy, B.; Szabo, T.G.; Pasztoi, M.; Pal, Z.; Misjak, P.; Aradi, B.; Laszlo, V.; Pallinger, E.; Pap, E.; Kittel, A.; et al. Membrane vesicles, current state-of-the-art: Emerging role of extracellular vesicles. Cell. Mol. Life Sci. 2011, 68, 2667–2688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. El Andaloussi, S.; Mager, I.; Breakefield, X.O.; Wood, M.J. Extracellular vesicles: Biology and emerging therapeutic opportunities. Nat. Rev. Drug Discov. 2013, 12, 347–357. [Google Scholar] [CrossRef]
  26. Shao, H.; Im, H.; Castro, C.M.; Breakefield, X.; Weissleder, R.; Lee, H. New Technologies for Analysis of Extracellular Vesicles. Chem. Rev. 2018, 118, 1917–1950. [Google Scholar] [CrossRef]
  27. Thery, C.; Zitvogel, L.; Amigorena, S. Exosomes: Composition, biogenesis and function. Nat. Rev. Immunol. 2002, 2, 569–579. [Google Scholar] [CrossRef]
  28. Pluchino, S.; Smith, J.A. Explicating Exosomes: Reclassifying the Rising Stars of Intercellular Communication. Cell 2019, 177, 225–227. [Google Scholar] [CrossRef] [Green Version]
  29. Willms, E.; Johansson, H.J.; Mager, I.; Lee, Y.; Blomberg, K.E.; Sadik, M.; Alaarg, A.; Smith, C.I.; Lehtio, J.; El Andaloussi, S.; et al. Cells release subpopulations of exosomes with distinct molecular and biological properties. Sci. Rep. 2016, 6, 22519. [Google Scholar] [CrossRef] [Green Version]
  30. Farooqi, A.A.; Desai, N.N.; Qureshi, M.Z.; Librelotto, D.R.N.; Gasparri, M.L.; Bishayee, A.; Nabavi, S.M.; Curti, V.; Daglia, M. Exosome biogenesis, bioactivities and functions as new delivery systems of natural compounds. Biotechnol. Adv. 2018, 36, 328–334. [Google Scholar] [CrossRef]
  31. Thery, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Vader, P.; Mol, E.A.; Pasterkamp, G.; Schiffelers, R.M. Extracellular vesicles for drug delivery. Adv. Drug Deliv. Rev. 2016, 106, 148–156. [Google Scholar] [CrossRef] [PubMed]
  33. Hua, S.; Wu, S.Y. The use of lipid-based nanocarriers for targeted pain therapies. Front. Pharmacol. 2013, 4, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Mima, Y.; Hashimoto, Y.; Shimizu, T.; Kiwada, H.; Ishida, T. Anti-PEG IgM Is a Major Contributor to the Accelerated Blood Clearance of Polyethylene Glycol-Conjugated Protein. Mol. Pharm. 2015, 12, 2429–2435. [Google Scholar] [CrossRef] [PubMed]
  35. Ishihara, T.; Takeda, M.; Sakamoto, H.; Kimoto, A.; Kobayashi, C.; Takasaki, N.; Yuki, K.; Tanaka, K.; Takenaga, M.; Igarashi, R.; et al. Accelerated blood clearance phenomenon upon repeated injection of PEG-modified PLA-nanoparticles. Pharm. Res. 2009, 26, 2270–2279. [Google Scholar] [CrossRef] [PubMed]
  36. Ishida, T.; Kashima, S.; Kiwada, H. The contribution of phagocytic activity of liver macrophages to the accelerated blood clearance (ABC) phenomenon of PEGylated liposomes in rats. J. Control. Release 2008, 126, 162–165. [Google Scholar] [CrossRef] [PubMed]
  37. Armstrong, J.K.; Hempel, G.; Koling, S.; Chan, L.S.; Fisher, T.; Meiselman, H.J.; Garratty, G. Antibody against poly(ethylene glycol) adversely affects PEG-asparaginase therapy in acute lymphoblastic leukemia patients. Cancer 2007, 110, 103–111. [Google Scholar] [CrossRef]
  38. van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  39. Batrakova, E.V.; Kim, M.S. Using exosomes, naturally-equipped nanocarriers, for drug delivery. J. Control. Release 2015, 219, 396–405. [Google Scholar] [CrossRef] [Green Version]
  40. Ha, D.; Yang, N.; Nadithe, V. Exosomes as therapeutic drug carriers and delivery vehicles across biological membranes: Current perspectives and future challenges. Acta Pharm. Sin. B 2016, 6, 287–296. [Google Scholar] [CrossRef]
  41. van der Meel, R.; Fens, M.H.; Vader, P.; van Solinge, W.W.; Eniola-Adefeso, O.; Schiffelers, R.M. Extracellular vesicles as drug delivery systems: Lessons from the liposome field. J. Control. Release 2014, 195, 72–85. [Google Scholar] [CrossRef] [PubMed]
  42. de Abreu, R.C.; Fernandes, H.; da Costa Martins, P.A.; Sahoo, S.; Emanueli, C.; Ferreira, L. Native and bioengineered extracellular vesicles for cardiovascular therapeutics. Nat. Rev. Cardiol. 2020, 17, 685–697. [Google Scholar] [CrossRef] [PubMed]
  43. Saeedi, M.; Eslamifar, M.; Khezri, K.; Dizaj, S.M. Applications of nanotechnology in drug delivery to the central nervous system. Biomed. Pharmacother. 2019, 111, 666–675. [Google Scholar] [CrossRef]
  44. Chen, Y.; Liu, L. Modern methods for delivery of drugs across the blood-brain barrier. Adv. Drug Deliv. Rev. 2012, 64, 640–665. [Google Scholar] [CrossRef]
  45. Zlokovic, B.V. The blood-brain barrier in health and chronic neurodegenerative disorders. Neuron 2008, 57, 178–201. [Google Scholar] [CrossRef] [Green Version]
  46. Lee, J.; Kim, S.E.; Moon, D.; Doh, J. A multilayered blood vessel/tumor tissue chip to investigate T cell infiltration into solid tumor tissues. Lab Chip 2021, 21, 2142–2152. [Google Scholar] [CrossRef]
  47. Deng, Z.; Sheng, Z.; Yan, F. Ultrasound-Induced Blood-Brain-Barrier Opening Enhances Anticancer Efficacy in the Treatment of Glioblastoma: Current Status and Future Prospects. J. Oncol. 2019, 2019, 2345203. [Google Scholar] [CrossRef] [Green Version]
  48. Saint-Pol, J.; Gosselet, F.; Duban-Deweer, S.; Pottiez, G.; Karamanos, Y. Targeting and Crossing the Blood-Brain Barrier with Extracellular Vesicles. Cells 2020, 9, 851. [Google Scholar] [CrossRef] [Green Version]
  49. Gabathuler, R. Approaches to transport therapeutic drugs across the blood-brain barrier to treat brain diseases. Neurobiol. Dis. 2010, 37, 48–57. [Google Scholar] [CrossRef]
  50. Faure, J.; Lachenal, G.; Court, M.; Hirrlinger, J.; Chatellard-Causse, C.; Blot, B.; Grange, J.; Schoehn, G.; Goldberg, Y.; Boyer, V.; et al. Exosomes are released by cultured cortical neurones. Mol. Cell. Neurosci. 2006, 31, 642–648. [Google Scholar] [CrossRef]
  51. Fruhbeis, C.; Frohlich, D.; Kuo, W.P.; Kramer-Albers, E.M. Extracellular vesicles as mediators of neuron-glia communication. Front. Cell. Neurosci. 2013, 7, 182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Potolicchio, I.; Carven, G.J.; Xu, X.; Stipp, C.; Riese, R.J.; Stern, L.J.; Santambrogio, L. Proteomic analysis of microglia-derived exosomes: Metabolic role of the aminopeptidase CD13 in neuropeptide catabolism. J. Immunol. 2005, 175, 2237–2243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Taylor, A.R.; Robinson, M.B.; Gifondorwa, D.J.; Tytell, M.; Milligan, C.E. Regulation of heat shock protein 70 release in astrocytes: Role of signaling kinases. Dev. Neurobiol. 2007, 67, 1815–1829. [Google Scholar] [CrossRef] [PubMed]
  54. Milane, L.; Singh, A.; Mattheolabakis, G.; Suresh, M.; Amiji, M.M. Exosome mediated communication within the tumor microenvironment. J. Control. Release 2015, 219, 278–294. [Google Scholar] [CrossRef]
  55. Zhang, T.; Ma, S.; Lv, J.; Wang, X.; Afewerky, H.K.; Li, H.; Lu, Y. The emerging role of exosomes in Alzheimer’s disease. Ageing Res. Rev. 2021, 68, 101321. [Google Scholar] [CrossRef]
  56. Ramirez, S.H.; Andrews, A.M.; Paul, D.; Pachter, J.S. Extracellular vesicles: Mediators and biomarkers of pathology along CNS barriers. Fluids Barriers CNS 2018, 15, 19. [Google Scholar] [CrossRef] [Green Version]
  57. Basso, M.; Bonetto, V. Extracellular Vesicles and a Novel Form of Communication in the Brain. Front. Neurosci. 2016, 10, 127. [Google Scholar] [CrossRef] [Green Version]
  58. Sharma, P.; Mesci, P.; Carromeu, C.; McClatchy, D.R.; Schiapparelli, L.; Yates, J.R., 3rd; Muotri, A.R.; Cline, H.T. Exosomes regulate neurogenesis and circuit assembly. Proc. Natl. Acad. Sci. USA 2019, 116, 16086–16094. [Google Scholar] [CrossRef] [Green Version]
  59. Budnik, V.; Ruiz-Canada, C.; Wendler, F. Extracellular vesicles round off communication in the nervous system. Nat. Rev. Neurosci. 2016, 17, 160–172. [Google Scholar] [CrossRef] [Green Version]
  60. Xu, B.; Zhang, Y.; Du, X.F.; Li, J.; Zi, H.X.; Bu, J.W.; Yan, Y.; Han, H.; Du, J.L. Neurons secrete miR-132-containing exosomes to regulate brain vascular integrity. Cell Res. 2017, 27, 882–897. [Google Scholar] [CrossRef]
  61. Perluigi, M.; Picca, A.; Montanari, E.; Calvani, R.; Marini, F.; Matassa, R.; Tramutola, A.; Villani, A.; Familiari, G.; Domenico, F.D.; et al. Aberrant crosstalk between insulin signaling and mTOR in young Down syndrome individuals revealed by neuronal-derived extracellular vesicles. Alzheimers Dement 2022, 18, 1498–1510. [Google Scholar] [CrossRef] [PubMed]
  62. Xiong, X.Y.; Liu, L.; Yang, Q.W. Functions and mechanisms of microglia/macrophages in neuroinflammation and neurogenesis after stroke. Prog. Neurobiol. 2016, 142, 23–44. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, S.; Ge, X.; Yu, J.; Han, Z.; Yin, Z.; Li, Y.; Chen, F.; Wang, H.; Zhang, J.; Lei, P. Increased miR-124-3p in microglial exosomes following traumatic brain injury inhibits neuronal inflammation and contributes to neurite outgrowth via their transfer into neurons. FASEB J. 2018, 32, 512–528. [Google Scholar] [CrossRef] [Green Version]
  64. Li, Z.; Liu, F.; He, X.; Yang, X.; Shan, F.; Feng, J. Exosomes derived from mesenchymal stem cells attenuate inflammation and demyelination of the central nervous system in EAE rats by regulating the polarization of microglia. Int. Immunopharmacol. 2019, 67, 268–280. [Google Scholar] [CrossRef]
  65. Xin, H.; Li, Y.; Buller, B.; Katakowski, M.; Zhang, Y.; Wang, X.; Shang, X.; Zhang, Z.G.; Chopp, M. Exosome-mediated transfer of miR-133b from multipotent mesenchymal stromal cells to neural cells contributes to neurite outgrowth. Stem Cells 2012, 30, 1556–1564. [Google Scholar] [CrossRef] [Green Version]
  66. Giunti, D.; Marini, C.; Parodi, B.; Usai, C.; Milanese, M.; Bonanno, G.; Kerlero de Rosbo, N.; Uccelli, A. Role of miRNAs shuttled by mesenchymal stem cell-derived small extracellular vesicles in modulating neuroinflammation. Sci. Rep. 2021, 11, 1740. [Google Scholar] [CrossRef]
  67. Cui, G.H.; Zhu, J.; Wang, Y.C.; Wu, J.; Liu, J.R.; Guo, H.D. Effects of exosomal miRNAs in the diagnosis and treatment of Alzheimer’s disease. Mech. Ageing Dev. 2021, 200, 111593. [Google Scholar] [CrossRef]
  68. Deng, Z.; Wang, J.; Xiao, Y.; Li, F.; Niu, L.; Liu, X.; Meng, L.; Zheng, H. Ultrasound-mediated augmented exosome release from astrocytes alleviates amyloid-beta-induced neurotoxicity. Theranostics 2021, 11, 4351–4362. [Google Scholar] [CrossRef]
  69. Zhang, L.; Zhang, S.; Yao, J.; Lowery, F.J.; Zhang, Q.; Huang, W.C.; Li, P.; Li, M.; Wang, X.; Zhang, C.; et al. Microenvironment-induced PTEN loss by exosomal microRNA primes brain metastasis outgrowth. Nature 2015, 527, 100–104. [Google Scholar] [CrossRef] [Green Version]
  70. Vogel, A.; Upadhya, R.; Shetty, A.K. Neural stem cell derived extracellular vesicles: Attributes and prospects for treating neurodegenerative disorders. eBioMedicine 2018, 38, 273–282. [Google Scholar] [CrossRef]
  71. Lee, E.J.; Choi, Y.; Lee, H.J.; Hwang, D.W.; Lee, D.S. Human neural stem cell-derived extracellular vesicles protect against Parkinson’s disease pathologies. J. Nanobiotechnol. 2022, 20, 198. [Google Scholar] [CrossRef] [PubMed]
  72. Apodaca, L.A.; Baddour, A.A.D.; Garcia, C., Jr.; Alikhani, L.; Giedzinski, E.; Ru, N.; Agrawal, A.; Acharya, M.M.; Baulch, J.E. Human neural stem cell-derived extracellular vesicles mitigate hallmarks of Alzheimer’s disease. Alzheimers Res. Ther. 2021, 13, 57. [Google Scholar] [CrossRef] [PubMed]
  73. Webb, R.L.; Kaiser, E.E.; Scoville, S.L.; Thompson, T.A.; Fatima, S.; Pandya, C.; Sriram, K.; Swetenburg, R.L.; Vaibhav, K.; Arbab, A.S.; et al. Human Neural Stem Cell Extracellular Vesicles Improve Tissue and Functional Recovery in the Murine Thromboembolic Stroke Model. Transl. Stroke Res. 2018, 9, 530–539. [Google Scholar] [CrossRef] [Green Version]
  74. Yang, L.; Han, B.; Zhang, Z.; Wang, S.; Bai, Y.; Zhang, Y.; Tang, Y.; Du, L.; Xu, L.; Wu, F.; et al. Extracellular Vesicle-Mediated Delivery of Circular RNA SCMH1 Promotes Functional Recovery in Rodent and Nonhuman Primate Ischemic Stroke Models. Circulation 2020, 142, 556–574. [Google Scholar] [CrossRef] [PubMed]
  75. Ge, X.; Guo, M.; Hu, T.; Li, W.; Huang, S.; Yin, Z.; Li, Y.; Chen, F.; Zhu, L.; Kang, C.; et al. Increased Microglial Exosomal miR-124-3p Alleviates Neurodegeneration and Improves Cognitive Outcome after rmTBI. Mol. Ther. 2020, 28, 503–522. [Google Scholar] [CrossRef] [Green Version]
  76. He, R.; Jiang, Y.; Shi, Y.; Liang, J.; Zhao, L. Curcumin-laden exosomes target ischemic brain tissue and alleviate cerebral ischemia-reperfusion injury by inhibiting ROS-mediated mitochondrial apoptosis. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 117, 111314. [Google Scholar] [CrossRef]
  77. Sutaria, D.S.; Badawi, M.; Phelps, M.A.; Schmittgen, T.D. Achieving the Promise of Therapeutic Extracellular Vesicles: The Devil is in Details of Therapeutic Loading. Pharm. Res. 2017, 34, 1053–1066. [Google Scholar] [CrossRef] [Green Version]
  78. Sutaria, D.S.; Jiang, J.; Elgamal, O.A.; Pomeroy, S.M.; Badawi, M.; Zhu, X.; Pavlovicz, R.; Azevedo-Pouly, A.C.P.; Chalmers, J.; Li, C.; et al. Low active loading of cargo into engineered extracellular vesicles results in inefficient miRNA mimic delivery. J. Extracell. Vesicles 2017, 6, 1333882. [Google Scholar] [CrossRef] [Green Version]
  79. Ren, X.; Zhao, Y.; Xue, F.; Zheng, Y.; Huang, H.; Wang, W.; Chang, Y.; Yang, H.; Zhang, J. Exosomal DNA Aptamer Targeting alpha-Synuclein Aggregates Reduced Neuropathological Deficits in a Mouse Parkinson’s Disease Model. Mol. Ther. Nucleic Acids 2019, 17, 726–740. [Google Scholar] [CrossRef] [Green Version]
  80. Qu, M.; Lin, Q.; Huang, L.; Fu, Y.; Wang, L.; He, S.; Fu, Y.; Yang, S.; Zhang, Z.; Zhang, L.; et al. Dopamine-loaded blood exosomes targeted to brain for better treatment of Parkinson’s disease. J. Control. Release 2018, 287, 156–166. [Google Scholar] [CrossRef]
  81. 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] [PubMed]
  82. Bogoch, I.I.; Brady, O.J.; Kraemer, M.U.G.; German, M.; Creatore, M.I.; Brent, S.; Watts, A.G.; Hay, S.I.; Kulkarni, M.A.; Brownstein, J.S.; et al. Potential for Zika virus introduction and transmission in resource-limited countries in Africa and the Asia-Pacific region: A modelling study. Lancet Infect. Dis. 2016, 16, 1237–1245. [Google Scholar] [CrossRef] [Green Version]
  83. Luo, S.; Sun, X.; Huang, M.; Ma, Q.; Du, L.; Cui, Y. Enhanced Neuroprotective Effects of Epicatechin Gallate Encapsulated by Bovine Milk-Derived Exosomes against Parkinson’s Disease through Antiapoptosis and Antimitophagy. J. Agric. Food Chem. 2021, 69, 5134–5143. [Google Scholar] [CrossRef] [PubMed]
  84. Hosseini Shamili, F.; Alibolandi, M.; Rafatpanah, H.; Abnous, K.; Mahmoudi, M.; Kalantari, M.; Taghdisi, S.M.; Ramezani, M. Immunomodulatory properties of MSC-derived exosomes armed with high affinity aptamer toward mylein as a platform for reducing multiple sclerosis clinical score. J. Control. Release 2019, 299, 149–164. [Google Scholar] [CrossRef] [PubMed]
  85. Hu, G.; Liao, K.; Niu, F.; Yang, L.; Dallon, B.W.; Callen, S.; Tian, C.; Shu, J.; Cui, J.; Sun, Z.; et al. Astrocyte EV-Induced lincRNA-Cox2 Regulates Microglial Phagocytosis: Implications for Morphine-Mediated Neurodegeneration. Mol. Ther. Nucleic Acids 2018, 13, 450–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Haney, M.J.; Klyachko, N.L.; Zhao, Y.; Gupta, R.; Plotnikova, E.G.; He, Z.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J. Control. Release 2015, 207, 18–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kalani, A.; Chaturvedi, P.; Kamat, P.K.; Maldonado, C.; Bauer, P.; Joshua, I.G.; Tyagi, S.C.; Tyagi, N. Curcumin-loaded embryonic stem cell exosomes restored neurovascular unit following ischemia-reperfusion injury. Int. J. Biochem. Cell Biol. 2016, 79, 360–369. [Google Scholar] [CrossRef] [Green Version]
  88. Haney, M.J.; Zhao, Y.; Jin, Y.S.; Li, S.M.; Bago, J.R.; Klyachko, N.L.; Kabanov, A.V.; Batrakova, E.V. Macrophage-Derived Extracellular Vesicles as Drug Delivery Systems for Triple Negative Breast Cancer (TNBC) Therapy. J. Neuroimmune Pharmacol. 2020, 15, 487–500. [Google Scholar] [CrossRef]
  89. Kim, M.S.; Haney, M.J.; Zhao, Y.; Mahajan, V.; Deygen, I.; Klyachko, N.L.; Inskoe, E.; Piroyan, A.; Sokolsky, M.; Okolie, O.; et al. Development of exosome-encapsulated paclitaxel to overcome MDR in cancer cells. Nanomedicine 2016, 12, 655–664. [Google Scholar] [CrossRef] [Green Version]
  90. 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]
  91. Lee, J.; Lee, H.; Goh, U.; Kim, J.; Jeong, M.; Lee, J.; Park, J.H. Cellular Engineering with Membrane Fusogenic Liposomes to Produce Functionalized Extracellular Vesicles. ACS Appl. Mater. Interfaces 2016, 8, 6790–6795. [Google Scholar] [CrossRef]
  92. Wiklander, O.P.; Nordin, J.Z.; O’Loughlin, A.; Gustafsson, Y.; Corso, G.; Mager, I.; Vader, P.; Lee, Y.; Sork, H.; Seow, Y.; et al. Extracellular vesicle in vivo biodistribution is determined by cell source, route of administration and targeting. J. Extracell. Vesicles 2015, 4, 26316. [Google Scholar] [CrossRef] [Green Version]
  93. Lai, C.P.; Mardini, O.; Ericsson, M.; Prabhakar, S.; Maguire, C.; Chen, J.W.; Tannous, B.A.; Breakefield, X.O. Dynamic biodistribution of extracellular vesicles in vivo using a multimodal imaging reporter. ACS Nano 2014, 8, 483–494. [Google Scholar] [CrossRef] [Green Version]
  94. Nakazaki, M.; Morita, T.; Lankford, K.L.; Askenase, P.W.; Kocsis, J.D. Small extracellular vesicles released by infused mesenchymal stromal cells target M2 macrophages and promote TGF-beta upregulation, microvascular stabilization and functional recovery in a rodent model of severe spinal cord injury. J. Extracell. Vesicles 2021, 10, e12137. [Google Scholar] [CrossRef]
  95. Riazifar, M.; Mohammadi, M.R.; Pone, E.J.; Yeri, A.; Lasser, C.; Segaliny, A.I.; McIntyre, L.L.; Shelke, G.V.; Hutchins, E.; Hamamoto, A.; et al. Stem Cell-Derived Exosomes as Nanotherapeutics for Autoimmune and Neurodegenerative Disorders. ACS Nano 2019, 13, 6670–6688. [Google Scholar] [CrossRef]
  96. Dasgupta, D.; Nakao, Y.; Mauer, A.S.; Thompson, J.M.; Sehrawat, T.S.; Liao, C.Y.; Krishnan, A.; Lucien, F.; Guo, Q.; Liu, M.; et al. IRE1A Stimulates Hepatocyte-Derived Extracellular Vesicles That Promote Inflammation in Mice With Steatohepatitis. Gastroenterology 2020, 159, 1487–1503.e17. [Google Scholar] [CrossRef]
  97. Thome, A.D.; Thonhoff, J.R.; Zhao, W.; Faridar, A.; Wang, J.; Beers, D.R.; Appel, S.H. Extracellular Vesicles Derived From Ex Vivo Expanded Regulatory T Cells Modulate In Vitro and In Vivo Inflammation. Front. Immunol. 2022, 13, 875825. [Google Scholar] [CrossRef]
  98. Kim, M.; Lee, Y.; Lee, M. Hypoxia-specific anti-RAGE exosomes for nose-to-brain delivery of anti-miR-181a oligonucleotide in an ischemic stroke model. Nanoscale 2021, 13, 14166–14178. [Google Scholar] [CrossRef]
  99. Moss, L.D.; Sode, D.; Patel, R.; Lui, A.; Hudson, C.; Patel, N.A.; Bickford, P.C. Intranasal delivery of exosomes from human adipose derived stem cells at forty-eight hours post injury reduces motor and cognitive impairments following traumatic brain injury. Neurochem. Int. 2021, 150, 105173. [Google Scholar] [CrossRef]
  100. Thomi, G.; Joerger-Messerli, M.; Haesler, V.; Muri, L.; Surbek, D.; Schoeberlein, A. Intranasally Administered Exosomes from Umbilical Cord Stem Cells Have Preventive Neuroprotective Effects and Contribute to Functional Recovery after Perinatal Brain Injury. Cells 2019, 8, 855. [Google Scholar] [CrossRef]
  101. Wang, D.; Xue, H.; Tan, J.; Liu, P.; Qiao, C.; Pang, C.; Zhang, L. Bone marrow mesenchymal stem cells-derived exosomes containing miR-539-5p inhibit pyroptosis through NLRP3/caspase-1 signalling to alleviate inflammatory bowel disease. Inflamm. Res. 2022, 71, 833–846. [Google Scholar] [CrossRef]
  102. Heidari, N.; Abbasi-Kenarsari, H.; Namaki, S.; Baghaei, K.; Zali, M.R.; Ghaffari Khaligh, S.; Hashemi, S.M. Adipose-derived mesenchymal stem cell-secreted exosome alleviates dextran sulfate sodium-induced acute colitis by Treg cell induction and inflammatory cytokine reduction. J. Cell. Physiol. 2021, 236, 5906–5920. [Google Scholar] [CrossRef]
  103. Tsai, S.C.; Yang, K.D.; Chang, K.H.; Lin, F.C.; Chou, R.H.; Li, M.C.; Cheng, C.C.; Kao, C.Y.; Chen, C.P.; Lin, H.C.; et al. Umbilical Cord Mesenchymal Stromal Cell-Derived Exosomes Rescue the Loss of Outer Hair Cells and Repair Cochlear Damage in Cisplatin-Injected Mice. Int. J. Mol. Sci. 2021, 22, 6664. [Google Scholar] [CrossRef]
  104. Han, M.; Yang, H.; Lu, X.; Li, Y.; Liu, Z.; Li, F.; Shang, Z.; Wang, X.; Li, X.; Li, J.; et al. Three-Dimensional-Cultured MSC-Derived Exosome-Hydrogel Hybrid Microneedle Array Patch for Spinal Cord Repair. Nano Lett. 2022, 22, 6391–6401. [Google Scholar] [CrossRef]
  105. Yang, T.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S. Exosome delivered anticancer drugs across the blood-brain barrier for brain cancer therapy in Danio rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef] [Green Version]
  106. Quintana, D.S.; Guastella, A.J.; Westlye, L.T.; Andreassen, O.A. The promise and pitfalls of intranasally administering psychopharmacological agents for the treatment of psychiatric disorders. Mol. Psychiatry 2016, 21, 29–38. [Google Scholar] [CrossRef]
  107. Erdo, F.; Bors, L.A.; Farkas, D.; Bajza, A.; Gizurarson, S. Evaluation of intranasal delivery route of drug administration for brain targeting. Brain Res. Bull. 2018, 143, 155–170. [Google Scholar] [CrossRef]
  108. Kou, L.; Bhutia, Y.D.; Yao, Q.; He, Z.; Sun, J.; Ganapathy, V. Transporter-Guided Delivery of Nanoparticles to Improve Drug Permeation across Cellular Barriers and Drug Exposure to Selective Cell Types. Front. Pharmacol. 2018, 9, 27. [Google Scholar] [CrossRef] [Green Version]
  109. Guo, S.; Perets, N.; Betzer, O.; Ben-Shaul, S.; Sheinin, A.; Michaelevski, I.; Popovtzer, R.; Offen, D.; Levenberg, S. Intranasal Delivery of Mesenchymal Stem Cell Derived Exosomes Loaded with Phosphatase and Tensin Homolog siRNA Repairs Complete Spinal Cord Injury. ACS Nano 2019, 13, 10015–10028. [Google Scholar] [CrossRef]
  110. Mulcahy, L.A.; Pink, R.C.; Carter, D.R. Routes and mechanisms of extracellular vesicle uptake. J. Extracell. Vesicles 2014, 3, 24641. [Google Scholar] [CrossRef]
  111. Qi, Z.; Yan, Z.; Wang, Y.; Ji, N.; Yang, X.; Zhang, A.; Li, M.; Xu, F.; Zhang, J. Ginsenoside Rh2 Inhibits NLRP3 Inflammasome Activation and Improves Exosomes to Alleviate Hypoxia-Induced Myocardial Injury. Front. Immunol. 2022, 13, 883946. [Google Scholar] [CrossRef]
  112. Bashyal, S.; Thapa, C.; Lee, S. Recent progresses in exosome-based systems for targeted drug delivery to the brain. J. Control. Release 2022, 348, 723–744. [Google Scholar] [CrossRef]
  113. Guo, Z.; Zhang, Y.; Xu, W.; Zhang, X.; Jiang, J. Engineered exosome-mediated delivery of circDIDO1 inhibits gastric cancer progression via regulation of MiR-1307-3p/SOCS2 Axis. J. Transl. Med. 2022, 20, 326. [Google Scholar] [CrossRef]
  114. Li, X.; Yu, Q.; Zhao, R.; Guo, X.; Liu, C.; Zhang, K.; Zhang, W.; Liu, J.; Yu, J.; Wang, S.; et al. Designer Exosomes for Targeted Delivery of a Novel Therapeutic Cargo to Enhance Sorafenib-Mediated Ferroptosis in Hepatocellular Carcinoma. Front. Oncol. 2022, 12, 898156. [Google Scholar] [CrossRef]
  115. Huang, X.; Wu, W.; Jing, D.; Yang, L.; Guo, H.; Wang, L.; Zhang, W.; Pu, F.; Shao, Z. Engineered exosome as targeted lncRNA MEG3 delivery vehicles for osteosarcoma therapy. J. Control. Release 2022, 343, 107–117. [Google Scholar] [CrossRef]
  116. Cui, Y.; Guo, Y.; Kong, L.; Shi, J.; Liu, P.; Li, R.; Geng, Y.; Gao, W.; Zhang, Z.; Fu, D. A bone-targeted engineered exosome platform delivering siRNA to treat osteoporosis. Bioact. Mater. 2022, 10, 207–221. [Google Scholar] [CrossRef]
  117. Terasawa, K.; Tomabechi, Y.; Ikeda, M.; Ehara, H.; Kukimoto-Niino, M.; Wakiyama, M.; Podyma-Inoue, K.A.; Rajapakshe, A.R.; Watabe, T.; Shirouzu, M.; et al. Lysosome-associated membrane proteins-1 and -2 (LAMP-1 and LAMP-2) assemble via distinct modes. Biochem. Biophys. Res. Commun. 2016, 479, 489–495. [Google Scholar] [CrossRef]
  118. Alvarez-Erviti, L.; Seow, Y.; Yin, H.; Betts, C.; Lakhal, S.; Wood, M.J. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nat. Biotechnol. 2011, 29, 341–345. [Google Scholar] [CrossRef]
  119. Liang, Y.; Duan, L.; Lu, J.; Xia, J. Engineering exosomes for targeted drug delivery. Theranostics 2021, 11, 3183–3195. [Google Scholar] [CrossRef]
  120. Phoolcharoen, W.; Prehaud, C.; van Dolleweerd, C.J.; Both, L.; da Costa, A.; Lafon, M.; Ma, J.K. Enhanced transport of plant-produced rabies single-chain antibody-RVG peptide fusion protein across an in cellulo blood-brain barrier device. Plant Biotechnol. J. 2017, 15, 1331–1339. [Google Scholar] [CrossRef]
  121. Liu, L.; Li, Y.; Peng, H.; Liu, R.; Ji, W.; Shi, Z.; Shen, J.; Ma, G.; Zhang, X. Targeted exosome coating gene-chem nanocomplex as “nanoscavenger” for clearing alpha-synuclein and immune activation of Parkinson’s disease. Sci. Adv. 2020, 6, eaba3967. [Google Scholar] [CrossRef] [PubMed]
  122. Han, M.; Xing, H.; Chen, L.; Cui, M.; Zhang, Y.; Qi, L.; Jin, M.; Yang, Y.; Gao, C.; Gao, Z.; et al. Efficient antiglioblastoma therapy in mice through doxorubicin-loaded nanomicelles modified using a novel brain-targeted RVG-15 peptide. J. Drug Target. 2021, 29, 1016–1028. [Google Scholar] [CrossRef] [PubMed]
  123. Tian, T.; Zhang, H.X.; He, C.P.; Fan, S.; Zhu, Y.L.; Qi, C.; Huang, N.P.; Xiao, Z.D.; Lu, Z.H.; Tannous, B.A.; et al. Surface functionalized exosomes as targeted drug delivery vehicles for cerebral ischemia therapy. Biomaterials 2018, 150, 137–149. [Google Scholar] [CrossRef] [PubMed]
  124. Jewett, J.C.; Bertozzi, C.R. Cu-free click cycloaddition reactions in chemical biology. Chem. Soc. Rev. 2010, 39, 1272–1279. [Google Scholar] [CrossRef]
  125. Bertozzi, C.R. A decade of bioorthogonal chemistry. Acc. Chem. Res. 2011, 44, 651–653. [Google Scholar] [CrossRef] [Green Version]
  126. Robinson, P.V.; de Almeida-Escobedo, G.; de Groot, A.E.; McKechnie, J.L.; Bertozzi, C.R. Live-Cell Labeling of Specific Protein Glycoforms by Proximity-Enhanced Bioorthogonal Ligation. J. Am. Chem. Soc. 2015, 137, 10452–10455. [Google Scholar] [CrossRef] [Green Version]
  127. Agarwal, P.; Beahm, B.J.; Shieh, P.; Bertozzi, C.R. Systemic Fluorescence Imaging of Zebrafish Glycans with Bioorthogonal Chemistry. Angew. Chem. Int. Ed. Engl. 2015, 54, 11504–11510. [Google Scholar] [CrossRef] [Green Version]
  128. Jia, G.; Han, Y.; An, Y.; Ding, Y.; He, C.; Wang, X.; Tang, Q. NRP-1 targeted and cargo-loaded exosomes facilitate simultaneous imaging and therapy of glioma in vitro and in vivo. Biomaterials 2018, 178, 302–316. [Google Scholar] [CrossRef]
  129. Kojima, R.; Bojar, D.; Rizzi, G.; Hamri, G.C.; El-Baba, M.D.; Saxena, P.; Auslander, S.; Tan, K.R.; Fussenegger, M. Designer exosomes produced by implanted cells intracerebrally deliver therapeutic cargo for Parkinson’s disease treatment. Nat. Commun. 2018, 9, 1305. [Google Scholar] [CrossRef] [Green Version]
  130. Oboudiyat, C.; Glazer, H.; Seifan, A.; Greer, C.; Isaacson, R.S. Alzheimer’s disease. Semin. Neurol. 2013, 33, 313–329. [Google Scholar] [CrossRef]
  131. Karran, E.; De Strooper, B. The amyloid hypothesis in Alzheimer disease: New insights from new therapeutics. Nat. Rev. Drug Discov. 2022, 21, 306–318. [Google Scholar] [CrossRef] [PubMed]
  132. Ismail, Z.; Creese, B.; Aarsland, D.; Kales, H.C.; Lyketsos, C.G.; Sweet, R.A.; Ballard, C. Psychosis in Alzheimer disease—Mechanisms, genetics and therapeutic opportunities. Nat. Rev. Neurol. 2022, 18, 131–144. [Google Scholar] [CrossRef] [PubMed]
  133. Leng, F.; Edison, P. Neuroinflammation and microglial activation in Alzheimer disease: Where do we go from here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef] [PubMed]
  134. Pitt, J.M.; Kroemer, G.; Zitvogel, L. Extracellular vesicles: Masters of intercellular communication and potential clinical interventions. J. Clin. Investig. 2016, 126, 1139–1143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zhang, D.; Lee, H.; Wang, X.; Rai, A.; Groot, M.; Jin, Y. Exosome-Mediated Small RNA Delivery: A Novel Therapeutic Approach for Inflammatory Lung Responses. Mol. Ther. 2018, 26, 2119–2130. [Google Scholar] [CrossRef] [Green Version]
  136. Wahlgren, J.; De, L.K.T.; Brisslert, M.; Vaziri Sani, F.; Telemo, E.; Sunnerhagen, P.; Valadi, H. Plasma exosomes can deliver exogenous short interfering RNA to monocytes and lymphocytes. Nucleic Acids Res. 2012, 40, e130. [Google Scholar] [CrossRef] [Green Version]
  137. Shtam, T.A.; Kovalev, R.A.; Varfolomeeva, E.Y.; Makarov, E.M.; Kil, Y.V.; Filatov, M.V. Exosomes are natural carriers of exogenous siRNA to human cells in vitro. Cell Commun. Signal. 2013, 11, 88. [Google Scholar] [CrossRef] [Green Version]
  138. Kim, B.; Park, J.H.; Sailor, M.J. Rekindling RNAi Therapy: Materials Design Requirements for In Vivo siRNA Delivery. Adv. Mater. 2019, 31, e1903637. [Google Scholar] [CrossRef] [Green Version]
  139. Balestrino, R.; Schapira, A.H.V. Parkinson disease. Eur. J. Neurol. 2020, 27, 27–42. [Google Scholar] [CrossRef]
  140. GBD 2015 Maternal Mortality Collaborators. Global, regional, and national levels of maternal mortality, 1990–2015: A systematic analysis for the Global Burden of Disease Study 2015. Lancet 2016, 388, 1775–1812. [Google Scholar] [CrossRef]
  141. Olanow, C.W.; Kieburtz, K.; Katz, R. Clinical approaches to the development of a neuroprotective therapy for PD. Exp. Neurol. 2017, 298, 246–251. [Google Scholar] [CrossRef] [PubMed]
  142. Cooper, J.M.; Wiklander, P.B.; Nordin, J.Z.; Al-Shawi, R.; Wood, M.J.; Vithlani, M.; Schapira, A.H.; 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] [Green Version]
  143. 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] [PubMed]
  144. Ross, C.A.; Tabrizi, S.J. Huntington’s disease: From molecular pathogenesis to clinical treatment. Lancet Neurol. 2011, 10, 83–98. [Google Scholar] [CrossRef]
  145. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. The Huntington’s Disease Collaborative Research Group. Cell 1993, 72, 971–983. [CrossRef]
  146. Tabrizi, S.J.; Estevez-Fraga, C.; van Roon-Mom, W.M.C.; Flower, M.D.; Scahill, R.I.; Wild, E.J.; Munoz-Sanjuan, I.; Sampaio, C.; Rosser, A.E.; Leavitt, B.R. Potential disease-modifying therapies for Huntington’s disease: Lessons learned and future opportunities. Lancet Neurol. 2022, 21, 645–658. [Google Scholar] [CrossRef]
  147. Zhang, Y.; Liu, Q.; Zhang, X.; Huang, H.; Tang, S.; Chai, Y.; Xu, Z.; Li, M.; Chen, X.; Liu, J.; et al. Recent advances in exosome-mediated nucleic acid delivery for cancer therapy. J. Nanobiotechnol. 2022, 20, 279. [Google Scholar] [CrossRef]
  148. Didiot, M.C.; Hall, L.M.; Coles, A.H.; Haraszti, R.A.; Godinho, B.M.; Chase, K.; Sapp, E.; Ly, S.; Alterman, J.F.; Hassler, M.R.; et al. Exosome-mediated Delivery of Hydrophobically Modified siRNA for Huntingtin mRNA Silencing. Mol. Ther. 2016, 24, 1836–1847. [Google Scholar] [CrossRef] [Green Version]
  149. Ganguly, K.; Khanna, P.; Morecraft, R.J.; Lin, D.J. Modulation of neural co-firing to enhance network transmission and improve motor function after stroke. Neuron 2022, 110, 2363–2385. [Google Scholar] [CrossRef]
  150. Debette, S.; Markus, H.S. Stroke Genetics: Discovery, Insight Into Mechanisms, and Clinical Perspectives. Circ. Res. 2022, 130, 1095–1111. [Google Scholar] [CrossRef]
  151. Kelly, P.J.; Kavanagh, E.; Murphy, S. Stroke: New Developments and Their Application in Clinical Practice. Semin. Neurol. 2016, 36, 317–323. [Google Scholar] [CrossRef] [PubMed]
  152. Owolabi, M.O.; Thrift, A.G.; Mahal, A.; Ishida, M.; Martins, S.; Johnson, W.D.; Pandian, J.; Abd-Allah, F.; Yaria, J.; Phan, H.T.; et al. Primary stroke prevention worldwide: Translating evidence into action. Lancet Public Health 2022, 7, e74–e85. [Google Scholar] [CrossRef]
  153. Zhao, L.; Li, J.; Kalviainen, R.; Jolkkonen, J.; Zhao, C. Impact of drug treatment and drug interactions in post-stroke epilepsy. Pharmacol. Ther. 2022, 233, 108030. [Google Scholar] [CrossRef]
  154. Lai, N.; Wu, D.; Liang, T.; Pan, P.; Yuan, G.; Li, X.; Li, H.; Shen, H.; Wang, Z.; Chen, G. Systemic exosomal miR-193b-3p delivery attenuates neuroinflammation in early brain injury after subarachnoid hemorrhage in mice. J. Neuroinflamm. 2020, 17, 74. [Google Scholar] [CrossRef]
  155. Writing Group, M.; Mozaffarian, D.; Benjamin, E.J.; Go, A.S.; Arnett, D.K.; Blaha, M.J.; Cushman, M.; Das, S.R.; de Ferranti, S.; Despres, J.P.; et al. Heart Disease and Stroke Statistics-2016 Update: A Report From the American Heart Association. Circulation 2016, 133, e38–e360. [Google Scholar] [CrossRef]
  156. Lemaire, Q.; Raffo-Romero, A.; Arab, T.; Van Camp, C.; Drago, F.; Forte, S.; Gimeno, J.P.; Begard, S.; Colin, M.; Vizioli, J.; et al. Isolation of microglia-derived extracellular vesicles: Towards miRNA signatures and neuroprotection. J. Nanobiotechnol. 2019, 17, 119. [Google Scholar] [CrossRef] [PubMed]
  157. Leidinger, P.; Backes, C.; Deutscher, S.; Schmitt, K.; Mueller, S.C.; Frese, K.; Haas, J.; Ruprecht, K.; Paul, F.; Stahler, C.; et al. A blood based 12-miRNA signature of Alzheimer disease patients. Genome Biol. 2013, 14, R78. [Google Scholar] [CrossRef] [Green Version]
  158. Hamzei Taj, S.; Kho, W.; Riou, A.; Wiedermann, D.; Hoehn, M. MiRNA-124 induces neuroprotection and functional improvement after focal cerebral ischemia. Biomaterials 2016, 91, 151–165. [Google Scholar] [CrossRef]
  159. Upadhya, R.; Madhu, L.N.; Attaluri, S.; Gitai, D.L.G.; Pinson, M.R.; Kodali, M.; Shetty, G.; Zanirati, G.; Kumar, S.; Shuai, B.; et al. Extracellular vesicles from human iPSC-derived neural stem cells: miRNA and protein signatures, and anti-inflammatory and neurogenic properties. J. Extracell. Vesicles 2020, 9, 1809064. [Google Scholar] [CrossRef]
  160. Mathew, B.; Ravindran, S.; Liu, X.; Torres, L.; Chennakesavalu, M.; Huang, C.C.; Feng, L.; Zelka, R.; Lopez, J.; Sharma, M.; et al. Mesenchymal stem cell-derived extracellular vesicles and retinal ischemia-reperfusion. Biomaterials 2019, 197, 146–160. [Google Scholar] [CrossRef]
  161. Perets, N.; Hertz, S.; London, M.; Offen, D. Intranasal administration of exosomes derived from mesenchymal stem cells ameliorates autistic-like behaviors of BTBR mice. Mol. Autism 2018, 9, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Thomi, G.; Surbek, D.; Haesler, V.; Joerger-Messerli, M.; Schoeberlein, A. Exosomes derived from umbilical cord mesenchymal stem cells reduce microglia-mediated neuroinflammation in perinatal brain injury. Stem Cell Res. Ther. 2019, 10, 105. [Google Scholar] [CrossRef] [PubMed]
  163. Williams, A.M.; Dennahy, I.S.; Bhatti, U.F.; Halaweish, I.; Xiong, Y.; Chang, P.; Nikolian, V.C.; Chtraklin, K.; Brown, J.; Zhang, Y.; et al. Mesenchymal Stem Cell-Derived Exosomes Provide Neuroprotection and Improve Long-Term Neurologic Outcomes in a Swine Model of Traumatic Brain Injury and Hemorrhagic Shock. J. Neurotrauma 2019, 36, 54–60. [Google Scholar] [CrossRef] [PubMed]
  164. Zhao, Y.; Gan, Y.; Xu, G.; Hua, K.; Liu, D. Exosomes from MSCs overexpressing microRNA-223-3p attenuate cerebral ischemia through inhibiting microglial M1 polarization mediated inflammation. Life Sci. 2020, 260, 118403. [Google Scholar] [CrossRef]
  165. Yang, J.; Zhang, X.; Chen, X.; Wang, L.; Yang, G. Exosome Mediated Delivery of miR-124 Promotes Neurogenesis after Ischemia. Mol. Ther. Nucleic Acids 2017, 7, 278–287. [Google Scholar] [CrossRef] [Green Version]
  166. Xin, H.; Katakowski, M.; Wang, F.; Qian, J.Y.; Liu, X.S.; Ali, M.M.; Buller, B.; Zhang, Z.G.; Chopp, M. MicroRNA cluster miR-17-92 Cluster in Exosomes Enhance Neuroplasticity and Functional Recovery After Stroke in Rats. Stroke 2017, 48, 747–753. [Google Scholar] [CrossRef] [Green Version]
  167. Lu, L.; Qi, S.; Chen, Y.; Luo, H.; Huang, S.; Yu, X.; Luo, Q.; Zhang, Z. Targeted immunomodulation of inflammatory monocytes across the blood-brain barrier by curcumin-loaded nanoparticles delays the progression of experimental autoimmune encephalomyelitis. Biomaterials 2020, 245, 119987. [Google Scholar] [CrossRef]
  168. Yuan, R.; Li, Y.; Han, S.; Chen, X.; Chen, J.; He, J.; Gao, H.; Yang, Y.; Yang, S.; Yang, Y. Fe-Curcumin Nanozyme-Mediated Reactive Oxygen Species Scavenging and Anti-Inflammation for Acute Lung Injury. ACS Cent. Sci. 2022, 8, 10–21. [Google Scholar] [CrossRef]
  169. Yang, J.; Wu, S.; Hou, L.; Zhu, D.; Yin, S.; Yang, G.; Wang, Y. Therapeutic Effects of Simultaneous Delivery of Nerve Growth Factor mRNA and Protein via Exosomes on Cerebral Ischemia. Mol. Ther. Nucleic Acids 2020, 21, 512–522. [Google Scholar] [CrossRef]
  170. Huang, X.; Ding, J.; Li, Y.; Liu, W.; Ji, J.; Wang, H.; Wang, X. Exosomes derived from PEDF modified adipose-derived mesenchymal stem cells ameliorate cerebral ischemia-reperfusion injury by regulation of autophagy and apoptosis. Exp. Cell Res. 2018, 371, 269–277. [Google Scholar] [CrossRef]
  171. Edwards, G., 3rd; Moreno-Gonzalez, I.; Soto, C. Amyloid-beta and tau pathology following repetitive mild traumatic brain injury. Biochem. Biophys. Res. Commun. 2017, 483, 1137–1142. [Google Scholar] [CrossRef] [PubMed]
  172. Yim, A.; Smith, C.; Brown, A.M. Osteopontin/secreted phosphoprotein-1 harnesses glial-, immune-, and neuronal cell ligand-receptor interactions to sense and regulate acute and chronic neuroinflammation. Immunol. Rev. 2022, 311, 224–233. [Google Scholar] [CrossRef] [PubMed]
  173. Spiteri, A.G.; Wishart, C.L.; Pamphlett, R.; Locatelli, G.; King, N.J.C. Microglia and monocytes in inflammatory CNS disease: Integrating phenotype and function. Acta Neuropathol. 2022, 143, 179–224. [Google Scholar] [CrossRef]
  174. Kaufmann, M.; Schaupp, A.L.; Sun, R.; Coscia, F.; Dendrou, C.A.; Cortes, A.; Kaur, G.; Evans, H.G.; Mollbrink, A.; Navarro, J.F.; et al. Identification of early neurodegenerative pathways in progressive multiple sclerosis. Nat. Neurosci. 2022, 25, 944–955. [Google Scholar] [CrossRef]
  175. Zhuang, X.; Xiang, X.; Grizzle, W.; Sun, D.; Zhang, S.; Axtell, R.C.; Ju, S.; Mu, J.; Zhang, L.; Steinman, L.; et al. Treatment of brain inflammatory diseases by delivering exosome encapsulated anti-inflammatory drugs from the nasal region to the brain. Mol. Ther. 2011, 19, 1769–1779. [Google Scholar] [CrossRef] [PubMed]
  176. Omuro, A.; DeAngelis, L.M. Glioblastoma and other malignant gliomas: A clinical review. JAMA 2013, 310, 1842–1850. [Google Scholar] [CrossRef] [PubMed]
  177. Zhu, Z.; Zhai, Y.; Hao, Y.; Wang, Q.; Han, F.; Zheng, W.; Hong, J.; Cui, L.; Jin, W.; Ma, S.; et al. Specific anti-glioma targeted-delivery strategy of engineered small extracellular vesicles dual-functionalised by Angiopep-2 and TAT peptides. J. Extracell. Vesicles 2022, 11, e12255. [Google Scholar] [CrossRef]
  178. Yamada, R.; Nakano, I. Glioma stem cells: Their role in chemoresistance. World Neurosurg. 2012, 77, 237–240. [Google Scholar] [CrossRef]
  179. 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]
  180. Shah, V.; Kochar, P. Brain Cancer: Implication to Disease, Therapeutic Strategies and Tumor Targeted Drug Delivery Approaches. Recent Pat. Anticancer Drug Discov. 2018, 13, 70–85. [Google Scholar] [CrossRef]
  181. Pouletty, P. Drug addictions: Towards socially accepted and medically treatable diseases. Nat. Rev. Drug Discov. 2002, 1, 731–736. [Google Scholar] [CrossRef] [PubMed]
  182. Liu, Y.; Li, D.; Liu, Z.; Zhou, Y.; Chu, D.; Li, X.; Jiang, X.; Hou, D.; Chen, X.; Chen, Y.; et al. Targeted exosome-mediated delivery of opioid receptor Mu siRNA for the treatment of morphine relapse. Sci. Rep. 2015, 5, 17543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Zhang, R.; Fu, Y.; Cheng, M.; Ma, W.; Zheng, N.; Wang, Y.; Wu, Z. sEVs(RVG) selectively delivers antiviral siRNA to fetus brain, inhibits ZIKV infection and mitigates ZIKV-induced microcephaly in mouse model. Mol. Ther. 2022, 30, 2078–2091. [Google Scholar] [CrossRef] [PubMed]
  184. Bunggulawa, E.J.; Wang, W.; Yin, T.; Wang, N.; Durkan, C.; Wang, Y.; Wang, G. Recent advancements in the use of exosomes as drug delivery systems. J. Nanobiotechnol. 2018, 16, 81. [Google Scholar] [CrossRef] [Green Version]
  185. Kooijmans, S.A.; Vader, P.; van Dommelen, S.M.; van Solinge, W.W.; Schiffelers, R.M. Exosome mimetics: A novel class of drug delivery systems. Int. J. Nanomed. 2012, 7, 1525–1541. [Google Scholar] [CrossRef] [Green Version]
  186. Ingato, D.; Lee, J.U.; Sim, S.J.; Kwon, Y.J. Good things come in small packages: Overcoming challenges to harness extracellular vesicles for therapeutic delivery. J. Control. Release 2016, 241, 174–185. [Google Scholar] [CrossRef] [Green Version]
  187. Aryani, A.; Denecke, B. Exosomes as a Nanodelivery System: A Key to the Future of Neuromedicine? Mol. Neurobiol. 2016, 53, 818–834. [Google Scholar] [CrossRef] [Green Version]
  188. El Andaloussi, S.; Lakhal, S.; Mager, I.; Wood, M.J. Exosomes for targeted siRNA delivery across biological barriers. Adv. Drug Deliv. Rev. 2013, 65, 391–397. [Google Scholar] [CrossRef]
  189. Grapp, M.; Wrede, A.; Schweizer, M.; Huwel, S.; Galla, H.J.; Snaidero, N.; Simons, M.; Buckers, J.; Low, P.S.; Urlaub, H.; et al. Choroid plexus transcytosis and exosome shuttling deliver folate into brain parenchyma. Nat. Commun. 2013, 4, 2123. [Google Scholar] [CrossRef]
Figure 1. Biogenesis of exosomes. Cells produce small vesicles by endocytosis, which fuse to form endosomes, and are subsequently accompanied by the entry of nucleic acids, proteins, etc. into the endosomes. The endosomes gradually evolve into multivesicular bodies (MVBs), which are subsequently released extracellularly to form exosomes. RER: rough endoplasmic reticulum; SER: smooth endoplasmic reticulum; MVB: multivesicular body; ESCRT: endosomal sorting complex required for transport.
Figure 1. Biogenesis of exosomes. Cells produce small vesicles by endocytosis, which fuse to form endosomes, and are subsequently accompanied by the entry of nucleic acids, proteins, etc. into the endosomes. The endosomes gradually evolve into multivesicular bodies (MVBs), which are subsequently released extracellularly to form exosomes. RER: rough endoplasmic reticulum; SER: smooth endoplasmic reticulum; MVB: multivesicular body; ESCRT: endosomal sorting complex required for transport.
Pharmaceutics 14 02252 g001
Figure 2. Rich contents and surface markers of exosomes. Exosomes are rich in cholesterol and sphingomyelin and have a lipid bilayer structure. Universally, exosomes contain tubulin, heat shock proteins, actin-binding proteins, ALIX, tetra membrane proteins, abundant nucleic acids, etc. CXCR4: CXC chemokine receptor 4; TCR: T cell receptor; HSP: heat shock proteins; ICAMs: intercellular cell adhesion molecules; TSG: tumor suppressor gene; lncRNA: Long non-coding RNA; miRNA: micro-RNA; cicRNA: circular RNA.
Figure 2. Rich contents and surface markers of exosomes. Exosomes are rich in cholesterol and sphingomyelin and have a lipid bilayer structure. Universally, exosomes contain tubulin, heat shock proteins, actin-binding proteins, ALIX, tetra membrane proteins, abundant nucleic acids, etc. CXCR4: CXC chemokine receptor 4; TCR: T cell receptor; HSP: heat shock proteins; ICAMs: intercellular cell adhesion molecules; TSG: tumor suppressor gene; lncRNA: Long non-coding RNA; miRNA: micro-RNA; cicRNA: circular RNA.
Pharmaceutics 14 02252 g002
Figure 3. Post-secretory loading methods of exosomes. Drug encapsulation strategies for isolated exosomes include sonication, electroporation, co-incubation, freeze-thaw cycles, extrusion, transfection kits, etc.
Figure 3. Post-secretory loading methods of exosomes. Drug encapsulation strategies for isolated exosomes include sonication, electroporation, co-incubation, freeze-thaw cycles, extrusion, transfection kits, etc.
Pharmaceutics 14 02252 g003
Figure 4. The research process of exosomes as CNS drug delivery tools. (A) Exosomes are isolated from cells or body fluids; (B) multiple drug loading modes of exosomes; (C) exosomes can be modified by genetic engineering and chemical reactions; (D) several administration routes of cargo-loading exosomes; (E) exosomes loaded with therapeutic molecules are used to treat a variety of CNS diseases. DC: dendritic cell; MSC: mesenchymal stem cell; CSF: cerebrospinal fluid; Exos: exosomes; PD: Parkinson’s disease; AD: Alzheimer’s disease.
Figure 4. The research process of exosomes as CNS drug delivery tools. (A) Exosomes are isolated from cells or body fluids; (B) multiple drug loading modes of exosomes; (C) exosomes can be modified by genetic engineering and chemical reactions; (D) several administration routes of cargo-loading exosomes; (E) exosomes loaded with therapeutic molecules are used to treat a variety of CNS diseases. DC: dendritic cell; MSC: mesenchymal stem cell; CSF: cerebrospinal fluid; Exos: exosomes; PD: Parkinson’s disease; AD: Alzheimer’s disease.
Pharmaceutics 14 02252 g004
Table 1. Preclinical studies on exosomes as DDS for CNS diseases treatment.
Table 1. Preclinical studies on exosomes as DDS for CNS diseases treatment.
DiseaseDonor CellTherapeutic MoleculeDrug Loading MethodModification StrategyAnimalAdministration RouteTargeted CellsRef.
Alzheimer’s diseaseself-derived dendritic cellsBACE1 siRNAelectroporationLamp2b-RVGMiceintravenousneurons, microglia, and oligodendrocytes[118]
Parkinson’s diseaseHEK-293T cellscatalase mRNAtransfectionLamp2b-RVGMicesubcutaneousunknown[129]
primary dendritic cellsa-Syn siRNAelectroporationLamp2b-RVGMiceintravenousunknown[142]
primary dendritic cellsshRNA-MCselectroporationLamp2b-RVGMiceintravenousunknown[143]
HEK-293T cellsaptamer F5R1co-incubationLamp2b-RVGMiceintraperitoneal microglia, neurons, and astrocytes[79]
mice blooddopamineco-incubationNoneMiceintravenousunknown[80]
RAW264.7catalaseco-incubation, freeze-thaw, sonication, or extrusionNoneMiceintranasalneurons and microglia[86]
Huntington’s diseaseglioblastoma U87 cellshsiRNAHTTco-incubationNoneMiceunilateral brain infusionneurons[148]
Strokemesenchymal stromal cells (MSC)curcuminco-incubationc(RGDyK) peptideMiceintravenousmicroglia, neurons, and astrocytes[123]
bone marrow mesenchymal stem cells (BMSC)miR-193b-3pelectroporationLamp2b-RVGMiceintravenousunknown[154]
HEK-293T cellscirc SCMH1transfectionLamp2b-RVGMice and rhesus monkeysintravenousmicroglia, neurons, and astrocytes[74]
mesenchymal stem cells (MSCs)miR-223-3ptransfectionNoneRatsintravenousunknown[164]
RAW264.7curcuminco-incubationNoneRatsintravenousneurons and endothelium cells[76]
mouse embryonic stem cells (MESCs)curcuminco-incubationNoneMiceintranasalastrocytes and neurons[87]
multipotent mesenchymal stromal cells (MSCs) miR-17-92transfectionNoneRatsintravenousunknown[166]
adipose-derived stem cells (ADSCs)PEDFtransfectionNoneRatsintravenousunknown[170]
HEK-293T cellsrecombinant human NGF mRNAtransfectionLamp2b-RVGMiceintravenousmicroglia, neurons, and astrocytes[169]
bone marrow mesenchymal stem cells (BMSC)miR-124electroporationLamp2b-RVGMiceintravenousneurons, astrocytes, and oligodendrocytes[165]
Repetitive mild traumatic brain injury (rmTBI)microgliamiR-124-3ptransfectionNoneMiceintravenousmicroglia, neurons, and astrocytes[75]
Spinal cord injury (SCI)mesenchymal stem cells (MSC)PTEN-siRNAco-incubationNoneRatsintranasalneurons[109]
Multiple sclerosisEL-4 cellscurcumin or JSI124co-incubationNoneMiceintranasalmicroglia[175]
mesenchymal stem cells (MSCs)LJM-3064 aptamerEDC/NHSNoneMiceintravenous unknown[84]
Brain tumorbrain endothelial cell (bEND.3) doxorubicinco-incubationNonezebrafishesintravenousunknown[105]
RAW264.7curcumin and SPIONselectroporationRGE-peptideMiceintravenousglioma[128]
ZIKV infectionHEK-293T cellsZIKV-specific
siRNA
electroporationLamp2b-RVGMiceintravenousmicroglia, neurons, and astrocytes[183]
Morphine addictionHEK-293T cellsMu (MOR) siRNAtransfectionLamp2b-RVGMiceintravenousneuro2A[182]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, K.; Zheng, X.; Jin, H.; Yu, F.; Zhao, W. Exosomes as CNS Drug Delivery Tools and Their Applications. Pharmaceutics 2022, 14, 2252. https://doi.org/10.3390/pharmaceutics14102252

AMA Style

Sun K, Zheng X, Jin H, Yu F, Zhao W. Exosomes as CNS Drug Delivery Tools and Their Applications. Pharmaceutics. 2022; 14(10):2252. https://doi.org/10.3390/pharmaceutics14102252

Chicago/Turabian Style

Sun, Ke, Xue Zheng, Hongzhen Jin, Fan Yu, and Wei Zhao. 2022. "Exosomes as CNS Drug Delivery Tools and Their Applications" Pharmaceutics 14, no. 10: 2252. https://doi.org/10.3390/pharmaceutics14102252

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