Download PDF
Review  |  Open Access  |  23 Apr 2022

Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases

Views: 2653 |  Downloads: 680 |  Cited:   1
Extracell Vesicles Circ Nucleic Acids 2022;3:102-17.
10.20517/evcna.2022.08 |  © The Author(s) 2022.
Author Information
Article Notes
Cite This Article

Abstract

Mesenchymal stem cells-derived extracellular vesicles (MSC-EVs) have noticeably attracted clinicians’ attention in treating ocular diseases. As the paracrine factor of MSCs and an alternative for cell-free therapies, MSC-EVs can be conveniently dropped over the ocular surface or diffused through the retina upon intravitreal injection, without increasing the risks of cellular rejection and tumor formation. For clinical translation, a standardized and scalable production, as well as reprogramming the MSC-EVs, are highly encouraged. This review aims to assess the potential approaches for EV production and functional modification, in addition to summarizing the worldwide clinical trials initiated for various physiological systems and the specific biochemical effects of MSC-EVs on the therapy of eye diseases. Recent advances in the therapy of ocular diseases based on MSC-EVs are reviewed, and the associated challenges and prospects are discussed as well.

Keywords

Mesenchymal stem cells, extracellular vesicles, ophthalmic diseases, therapy

INTRODUCTION

Mesenchymal stem cells (MSCs) are a heterogeneous population of stromal stem cells that can be isolated from various tissues, including bone marrow, adipose tissue, umbilical cords, and even urine[1]. Recently, MSCs have been extensively recognized as an experimental and therapeutic tool spanning from physiological regulation to organ remodeling, due to their superiority in low antigenicity and tumorigenicity[2-3]. When the term “Mesenchymal Stem Cell” was searched on ClinicalTrials.gov, more than 1000 clinical trials could be retrieved, which were directly associated with various diseases, suggesting the great potential of the mentioned term in regenerative medicine.

The roles of MSCs may include straight differentiation into target cells to replace injured tissues and generate various bioactive substances including extracellular vesicles (EVs), especially nano-sized exosomes[4,5]. EVs are generally classified as exosomes (30-150 nm) formed inwardly during the maturation of multiple vesicle endosomes[6-7], microvesicles (50-1000 nm) directly shed from plasma membrane, and apoptotic bodies (1000-5000 nm) released by dying cells[8,9] [Figure 1A]. However, due to the overlapping size range and the lack of specific markers, current “exosome” preparations are a mixture of EVs with undefined biogenesis origin and undetermined purity. According to the MISEV2018 position paper from ISEV[10], in this review, we use the term “MSC-EVs” to describe the MSC-derived exosomal preparations. EVs are released from living cells and can be found in almost all body fluids, including blood, urine, breast milk, tears, saliva, vitreous fluid, and aqueous humor[11-14]. The top two most studied body fluids are still blood and urine observed in 143 clinical trials [www.clinicaltrials.gov (Accessed: August 2021)] [Figure 1B]. These nanoparticles carry plenty of bioactive molecules, such as proteins, lipids, RNAs [messenger RNAs (mRNAs), circular RNAs (circRNAs), small RNAs (sRNAs), long non-coding RNAs (lncRNAs)], and DNAs [genomic DNA (gDNA), complementary DNA (cDNA), and mitochondrial DNA (mDNA)][15-19] that are delivered to recipient cells mediating intercellular response. Compared with MSCs, EV-based therapeutics have shown unique advantages, including cell-free therapy, large-scale EV production, low immunogenicity, and high bioavailability, making these vesicles possible drugs in treating various diseases (e.g., eye diseases).

Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases

Figure 1. Illustration of biogenesis of EVs and MSC-EVs for ocular therapy. (A) EVs mainly include microvesicles, apoptotic bodies, and exosomes based on their biogenesis. (B) EV sources and the proportion of EV sources derived from 143 clinical trials [www.clinicaltrials.gov (Accessed: August 2021)]. (C) MSC-EVs, which are used for ocular therapy, are associated with the mechanisms of cell proliferation, angiogenesis, immunoregulation, and miRNA-dependent regulation. EVs: Extracellular vesicles; MSC-EVs: mesenchymal stem cells-derived EVs; ILVs: intraluminal vesicles; MVB: multivesicular body; RGC: retinal ganglion cells; VEGF-A: vascular endothelial growth factor A; M2: macrophage of type 2; T cells: thymus-dependent lymphocyte; Th cells: helper T cells; miRNA: microRNA.

The human eye has a localized array of surface molecules and cytokines, and it is a sensory organ that reacts with visible light and enables us to use visual information for various purposes[20-22]. The intercellular signaling pathway is critical to maintaining the homeostasis of the intraocular microenvironment. EVs from both non-immune and immune cells play important roles in immune regulation[23]. At present, for the visual system, researchers have mainly concentrated on the application of EV-based therapeutics in a variety of ocular diseases, such as physical injuries, immune-related diseases, and other eye diseases. These nanoparticles migrate to injured or inflammatory sites, releasing genetic materials or proteins to repair damage by participating in signaling pathways[24-29]. As a new model of cell-free therapy, EVs have been evaluated in preliminary clinical trials and have shown great efficacy[30]. In addition, EV-associated products have also been applied in the treatment of ocular diseases[31,32]. With the continuous exploration of the physical and chemical properties and functions of the EVs, these naturally occurring nanoparticles have been feasibly applied to clinical medicine.

The present review aims to assess the biological characteristics of MSC-EVs and consider novel methods for EV isolation. More importantly, the translational application of MSC-EVs in eye diseases and the current challenges are discussed.

MSC-EVs

Secretion of the cells in the form of EVs was traditionally considered as unimportant waste material, cellular “garbage bags”, or dust particles, while it was later found that these nano-vesicles are vital messengers and participate in diverse physiological and pathological processes, such as bone tissue regeneration[33], tumor defense[34], nerve signal transmission[35], endothelial cell migration[36], and immune tolerance[37], as summarized in Table 1. The paracrine effect of MSCs was first described by Haynesworth et al.[51], who reported the synthesis and secretion of various growth factors, chemokines, and cytokines by MSCs. In 2009, Bruno et al.[52] demonstrated that microvesicles derived from MSCs may activate a proliferative program in surviving tubular cells after injury via a horizontal transfer of mRNAs. MSC-EVs were first fractionated with a particle size of 55-65 nm by high-performance liquid chromatography. In total, 857 related proteins and 151 microRNAs (miRNAs) of MSC origin have been detected by mass spectrometry, antibody array technology, and microarray analysis[53,54]. Besides, among the functional elements of these EVs, the roles of miRNAs in EV-based therapeutics have been widely investigated[55-57]. We mainly summarize several mechanisms achieved by recent studies, including proliferation, angiogenesis, immunoregulation, and miRNA-dependent activity [Figure 1C].

Table 1

Summary of the mechanisms and applications of the MSC-derived EV therapies

ApplicationDisease modelEV sourceDoseEffector molecule
Tissue regenerationNephrectomyUMSCs10 μgProteins: ANG-1, etc[38]
Renal ischemia reperfusionhP-MSCs100 μgmiRNA let-7a-5p[39]
Calvarial bone defectBMSCs5 × 108 particlesProtein:
BMP2[40]
Tumor defenseMetastatic lung nodulesAD-MSCs2.32 × 109 particlesmiR-101[41]
Gastric cancerUMSCs64 μgProtein:
L-PGDS[42]
LCBMSCs50 μgProtein: caspase 3[43]
Nerve injurySciatic nerve transectionUMSCs100 μgProtein:
IL-10, etc[44]
Brain injuryBMSCs200 μL (unknown)miR-140-5p[45]
ADBMSCs30 μgmiR-29c-3p[46]
Endothelial cell migrationAngiogenesisOM-MSCs50 μgmiR-612[47]
Myocardial ischemiaDPSC3.5 × 1010 particlesmiR-4732-3p[48]
immunoregulationKnee osteoarthritisSMSC5 μL (unknown)miR-31[49]
Chronic asthmaUMSCs40 μgmiR-146a-5p[50]

Preparation and modification of EVs

The important role of EVs in the diagnosis, prognosis, and regeneration of diseases promotes the development of EV isolation techniques. As EVs are nanoparticles and originate from a complex fluid environment, obtaining homogeneous and high purity therapeutic EVs remains a great challenge. The current methods for isolating EVs are mainly based on physical (size and density) and chemical (affinity) properties, as well as immunoaffinity chromatography (combining the use of liquid chromatography with the specific binding of antibodies)[58-60]. However, the most ubiquitous adopted method for EV preparation is still dominated by ultracentrifugation (UC). Methods based on micromachining technology, due to label-free processing, cost-effectiveness, and amenability to automation, have emerged as a promising method for label-free EV separation. Inglis et al.[61] designed and implemented theoretical models for the critical particle size of fractionation in deterministic lateral displacement (DLD) separation arrays, aiming to provide a theory and experimental measurements for critical conditions. Wunsch et al.[62] applied this technique to the true nanoscales, where they could function in EV separation, such as exosomes. This study revealed a potential for the on-chip sorting of these nanoparticles. For fast EV enrichment, a technology that integrates 1024 nanoscale DLD (nano-DLD) arrays on a single chip allows parallel processing to reach 900 μL/h[63]. Moreover, compared with other methods, including UC, UC plus density gradient, size-exclusion chromatography, and co-precipitation, the chip showed a superior efficiency. Recently, our team reported a novel exosome detection method via the ultrafast-isolation system (EXODUS) that allowed automated label-free purification of exosomes from various biofluids[64]. We also reported a size-based EV isolation tool, namely ExoTIC, to efficiently isolate EVs from small sample volumes, providing an analytical tool for preclinical studies[65]. These techniques are advantageous for the standardized preparation of MSC-EVs and can accelerate the clinical translation of MSC-EVs.

The massive production of MSC-EVs is another research hotspot for the cell-free treatment model. To expand the clinical translation of MSC-EVs, the methods used for large-scale production of EVs with a good manufacturing practice (GMP) level are necessary. To date, the efficiency of EVs has been improved by changing the culture method of donor cells, such as three-dimensional (3D) environmental culture[66-68]. Natural extracellular matrix and 3D biological scaffolds were used for cell attachment, cell growth, and production of functional EVs[69]. Using the 3D spheroid culture method based on photolithography and micro-pattern technologies, gene expression profiles of MSCs were confirmed with a high differentiation efficiency[70]. Cone et al.[71] assessed the potential therapeutic effects of EVs from a 3D culture of bone marrow-derived MSCs (BMSCs) in an Alzheimer’s disease (AD) model, and it was revealed that intranasally administration of MSC-EVs ameliorated pathology and cognitive deficits of AD. Mend et al.[72] reported a bioreactor-based and clinical-grade production of engineered exosomes with the ability to target oncogenic KRAS. The clinical-grade product was tested in multiple in vitro and in vivo experiments to confirm the feasibility of various therapies for human diseases.

In addition to natural EV agents, the development of different modifications of MSC-EVs may provide new approaches for gene therapy and drug delivery. Exogenous nucleic acids, such as miRNA, siRNA, DNA carrier, and DNA probe, are loaded into EVs by electroporation, accompanied with favorable biocompatibility and biostability. For instance, the engineered MSC-EVs can serve as a promising anti-osteoporosis therapy via loading Shn3 gene-targeted siRNA[73]. Angiopep-2 (Ang) is a ligand that binds specifically to the lipoprotein receptor-related protein 1 receptor and improves the high efficiency of transport across the blood-brain barrier (BBB)[74-76]. Several scholars designed a multifunctional exosome-mimetics decorated with Ang and load docetaxel for anti-glioblastoma therapy[77]. This personalized approach also achieved the purpose of targeted therapy.

Overall, with the advances of nanomedicine in molecular cell biology, pharmaceutical science, and nano-engineering[78], higher requirements for engineering transformation of MSC-EVs are demanded, especially standardized production and storage of MSC-EVs.

EV-based therapeutics

EVs have been extensively studied in clinical trials. A statistical analysis of 143 EV-dependent clinical trials was performed, and significant conclusions were obtained, as shown in Figure 2 [www.clinicaltrials.gov (Accessed: August 2021)]. The majority of clinical studies are conducted in the United States and China, and respiratory, tumor, and gland-related diseases were research hotspots [Figure 2A and B]. Based on research purposes, we divide all studies into four groups, which are followed by diagnosis, monitoring, treatment, and mechanisms [Figure 2C]. Then, we calculate the percentages of EVs involved in 108 studies that are mainly related to exosomal RNAs and proteins [Figure 2D]. Treatment-related research accounted for 15% of all items, which are mostly derived from MSCs [Figure 2E]. As shown in Figure 2F, most clinical trials are still in the infancy stage. At present, therapeutic vesicles are widely used in cardiovascular and cerebrovascular diseases, respiratory diseases, neurological diseases, cancer, and bone regeneration by affecting cell cycle arrest or apoptosis[79-84]. As a good example of the application of EVs in bone regeneration, osteoarthritis (OA) is a joint degenerative disease characterized by synovial inflammation and articular cartilage damage. The treatment of OA mainly depends on surgery and drugs. Several studies have shown that EVs maintain homeostasis and improve the severity of osteoarthritis pathologically through local and distal intercellular and intracellular signaling pathways[84-86]. In a rat model of glucocorticoid-induced femoral head necrosis, human umbilical cord-derived MSC-EVs (UMSC-EVs) could reduce the apoptosis of bone cells through the miR-21-PTEN-AKT signaling pathway[87].

Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases

Figure 2. The progress of 143 clinical trials involving EVs worldwide [www.clinicaltrials.gov (Accessed: August 2021)]. (A) The number of studies on EVs by physiological structure. There are 22 categories according to the body structure in 143 studies. (B) The number of studies on EVs is classified according to country and project progress status. According to country classification, 27 countries participated in the 143 studies on exosomes, and there are nine research statuses in 143 studies. (C) The percentage is presented based on the research purpose. There are four different directions of EV-based research, including diagnosis, treatment, monitoring, and mechanism. (D) The contents of EVs were evaluated in 108 studies. Others were used to illustrate an unclear description. (E) The proportion of cell sources used for the treatment with EVs in 42 studies. There are nine therapeutic sources of EVs, mainly originating from MSCs. (F) According to the definition presented by the US Food and Drug Administration (FDA), the number of studies in each clinical trial stage is shown. “Not Applicable” is used to describe trials without FDA-defined phases. EVs: Extracellular vesicles; MSCs: mesenchymal stem cells.

MSC-EV-BASED THERAPY FOR OPHTHALMIC DISEASES

Corneal disease

Corneal disease is the major cause of vision loss, which may be caused by several clinical conditions, including physical trauma, chemical burns, infections, limbal stem cell defects, age-related degeneration, and corneal malnutrition[26]. Although corneal transplantation has made significant progress in the past decade, there are still problems, such as few donors, immune rejection, and long-term use of immunosuppressant agents[88,89]. The role of MSCs in corneal regeneration therapy can be directly attributed to cell replacement[90] and delivering targeted biomolecules[91-94]. Several scholars attempted to incorporate hydrogel with exogenous recombinant human stromal cell-derived factor-1 alpha for corneal epithelial regeneration[95]. EVs have a promising prospect of therapeutic applications, as they inherit parental cell-derived molecules. Thus, MSC-EVs have also been applied in the therapy of corneal disease.

Many studies have confirmed the therapeutic efficiency of MSC-EVs for eye diseases including corneal and retinal models in vitro and in vivo, as presented in Table 2. Overall, tissue sources of MSCs for ocular diseases mainly originate from human corneal stroma, bone marrow, umbilical cord, and adipose tissues. In-depth analysis and generalization of the mechanism of MSC-EVs for corneal disease can be summarized in the following aspects: (1) MSC-EVs enhance the proliferation of human corneal epithelial cells (HCECs) and promote the migration of HCECs after corneal disease[96-98]; (2) MSC-EVs reduce scar formation, neovascularization, and hemorrhage after corneal disease[96,97,99]; (3) MSC-derived products decrease the levels of inflammatory cytokines, such as interleukin (IL)-1β, IL-6, and IL-10[98]; (4) MSC-EV-based treatment can inhibit neutrophil infiltration and polarize M2 macrophage infiltration[98,100]; and (5) MSC-EVs depress the expression levels of fibrotic genes (Col3a1 and Acta2) and serve as a delivery vehicle for miRNA in blocking corneal scarring blocking scarring and initiating regeneration after wound healing[100].

Table 2

Studies on the therapy of ocular diseases using MSC-EVs

PositionEV sourceAdministration route/doseResultsEffector molecule
CorneaBMSCsViscoelastic gel carrier/unclearEnhance HCECs proliferation and wound healing; reduce scar formation, neovascularization, and hemorrhageUnclear[96]
BMSCsCo-culture/unclearInduce proliferation and migration of damaged HCECs; inhibit cell apoptosisUnclear[97]
ADSCsTopical administration/unclearPromote proliferation and migration of HCECs, reduce inflammatory cytokine levels, polarize infiltrating macrophages toward M2Unclear[98]
CSSCsEVs drop/5.0 × 106 particlesAccelerate wound healingUnclear[99]
CSSCsTopical fifibrin
gel/1 × 107 particles
Decreased expression of fibrotic genes Col3a1 and Acta2, blocked neutrophil infiltrationmiRNA[100]
ADSCsCo-culture/1.61 × 1010 particlesToxicological testingUnclear[101]
BMSCsCo-culture/unclearfacilitate wound healingUnclear[102]
RetinaUMSCsIV/2.5 μgInhibition of MCP-1MCP-1[28]
BMSCsIV/3 × 109 particlesThrough miRNA dependent mechanismsmiRNA[56]
UMSCsTail vein/55 μgMiR-126 expression and downregulating the HMGB1 signaling pathwaymiR-126[103]
ADSCsIS/unclearDelivering microRNA-222 acts as mediators in retinal tissue repairmiRNA-222[104]
BMSCsIV/4 × 109 particlesReduce neuroinflammation and neuronal apoptosisUnclear[105]
BMSCsTail vein/30 μgInhibit activation of antigen-presenting cells and suppress the development of Th1 and Th17 cellsUnclear[106]
UMSCsIV/0.05 μgAmeliorate retinal injury via downregulation of VEGF-AUnclear[107]
UMSCsIV/1 × 109 particlesPromoting the RGCs survival and glia cells activationUnclear[108]
BMSCsIV/1 × 109 particlesPreserving RGC numbers and protecting against axonal degenerationUnclear[109]
ES-MSCsIO/15 μg
Improved Brn3a+ RGCs survival and improved cognitive visual behaviorUnclear[110]

EV-encapsulated natural lipid bilayers were considered as a good carrier to protect miRNAs from degradation. The differences in certain miRNA or miRNA expressions of EVs showed the diversity of receptor phenotypic regulation by non-coding RNA[111-113]. To elucidate the molecular mechanisms of EVs in the treatment of corneal wounds, researchers have performed numerous cutting-edge experiments. Using small-interfering RNAs (siRNAs) to knock down the mRNA of ESCRT protein Alix resulted in a reduction of 85% of EV miRNA; thus, EVs lacking miRNA lost their regeneration function[100]. The finding indicates that miRNA is a key adjustable molecule for EVs to exert restorative effects. Some studies have concentrated on the exosomal miRNA functions (i.e., regulating angiogenesis and anti-fibrotic immunosuppressive agents)[114,115], suggesting that miRNAs play an important role in maintaining homeostasis. Moreover, the miRNA expression profiles are variable in different cell types, which were mainly reflected in the number and category[116].

The therapeutic potential of MSCs can be related to cultivation conditions and cellular microenvironment. The influences of two-dimensional (2D) and 3D culture conditions on the therapeutic efficacy of MSC secretomes on corneal wound healing were studied with in vitro cell and organ culture experiments[84]. Notably, the secretomes from the MSC 3D environment facilitate wound healing in corneal fibroblast cells and enhance epithelialization. Ha et al.[101] conducted a toxicological evaluation of exosomes derived from human adipose tissue-derived mesenchymal stem cells (ADSCs), and the eye irritation test suggested that ADSC exosomes are safely used as a topical treatment.

Altogether, MSC-EV-dependent therapeutic molecules can regulate intercellular signaling pathways, and engineered EVs may be an emerging agent for corneal diseases.

Retinal diseases

Retinal ganglion cell (RGC) death is the irreversible endpoint of optic neuropathy. Glaucoma is a group of progressive optic neuropathies characterized by the gradual disappearance of RGCs. Under an in vitro strict culture, MSCs can induce differentiation into neuroectodermal cells, including neuronal cells[117]. MSCs used for the treatment of glaucoma mainly contribute to producing neurotrophins, differentiation into functional RGCs, and interaction with TM (trabecular meshwork), thereby reducing the intraocular pressure of glaucoma[118]. The capacity of MSC-EVs for neuroprotection and immunomodulation in the treatment of retinal diseases is mainly due to miRNA-dependent and inflammatory responses[119].

Studies have proven that miR-34a-5p, miR-126, and miR-222 affect the progression of retinal damage through diverse mechanisms[103,104,120]. For instance, in a cell model of diabetic retinopathy, MSC-derived exosomal lncRNA SNHG7 suppresses endothelial-mesenchymal transition and tube formation by negatively regulating miRNA[120]. miR-126 has been reported as an endothelial cell-restricted miRNA that mediates inflammation and vascular development[103]. HMGB1, one of the target genes of miR-126, has high expression levels in various inflammatory and autoimmune diseases[121]. Co-culture of MSC-EVs with high expressions of miR-126 and human retinal endothelial cells was found to significantly reduce the level of HMGB1 protein and improve retinal inflammation caused by hyperglycemia in diabetic rats[103]. A previous study showed that MSC-EVs are endocytosed by retinal neurons, retinal ganglion cells, and microglia as biomaterials for neuroprotective and regenerative therapy of retinal disorders[105]. Moreover, in a clinical trial, Zhang et al.[122] also proved that MSC-EV therapy may be an advantageous and safe method for improving visual outcomes after surgery for refractory macular holes.

With the advent of induced pluripotent stem cells (iPSCs), tremendous progress has been made in stem cell biology and regenerative medicine. Human iPSCs are widely used in animal modeling, drug discovery, and cell therapy development[123]. Besides MSC-EV-based therapies, iPSCs can also be used as an unlimited source for retinal degenerative diseases[124]. Retinal pigment epithelial cells derived from iPS (iPS-RPE) replace damaged or diseased cells and promote the healing and repairing of eye tissues[125]. Studies have shown that iPS-RPE plays an effective role in delaying photoreceptor degeneration by stably surviving in a degraded ocular environment and releasing neuroprotective factors, such as the pigment epithelium-derived factor[126]. To obtain an adequate source of cells, Reichman et al.[127] developed a two-step culture system to effectively differentiate iPSCs into retinal cells and achieved large-scale production and storage of hiPSCs-derived retinal cells and tissues. The development of iPSCs is expected to be another novel approach to treat retinal diseases in the future.

Other ocular diseases

The immunoregulatory effects of MSC-EVs have been reported in a variety of experimental models, such as rheumatoid arthritis[128], neurodegenerative disorders[129], and inflammatory bowel disease[130]. Studies have demonstrated that the anti-inflammatory effect of MSC-EVs is closely associated with regulating the activity of macrophages[131-133], natural killer cells[134], B cells[135,136], and T cells[137,138]. Scholars also used this positive influence in the modeling of inflammatory-related eye diseases. Uveitis, an inflammatory disorder involving the pigmented vascular coat of the eyeball, can result in blindness in the absence of timely therapy. Similar to inflammatory eye disease, MSC-EVs suppress autoimmunity in models of experimental autoimmune uveoretinitis (EAU) by inhibiting the development of T cells[106]. Administering MSC into rodents with induced models of clinical diseases with an appropriate dose can result in the reversal of abnormalities for weeks thereafter. Zhang et al.[139] examined the long-term effects of BMSCs in a recurrent EAU model in rats. The results demonstrate that BMSCs significantly decreased responses of T helper 1 (Th1) and Th17 cells, suppressed the functions of antigen-presenting cells, and upregulated T regulatory cells. In the study of EAU in Lewis rats, MSCs showed an inhibitory effect on activation and maturation of dendritic cells via regulation of STAT1 and STAT6 phosphorylation[140]. In the subsequent studies using MSC-EVs on the same EAU models, it was found that administration of MSC-EVs could ameliorate uveitis similar to their parent cells[141]. Using in vitro experiments, the effects of MSC-EVs on immune-cell activation were assessed using allogeneic mixed lymphocyte reaction assays. Consistent with previous MSC-related findings, MSC-EVs simultaneously reduce the infiltration of T cells and the levels of inflammatory cytokines[106].

Sjögren’s syndrome (SS), a chronic multi-system autoimmune disease mainly involving the exocrine gland, causes dry mouth (hyposialia or even asialia) and dry eye (xerophthalmia)[142,143]. MSCs, as a therapeutic approach to treat SS, have been assessed in preclinical trials[144,145]. In an in vivo study, Xu et al.[146] proposed a novel therapeutic approach to alleviate diseases in patients with primary SS by infusing allogeneic UMSCs. These effects are nutritive, anti-inflammatory, anti-immunologic, and associated with the healing of abnormalities. Regarding EVs secreted by MSCs, MSC-EVs may be an ideal replacement for decreasing the pathogenesis of SS. Rui et al.[147] found that murine olfactory ecto-MSC-derived exosomes significantly improved impaired immunosuppressive function of myeloid-derived suppressor cells by administering MSC-EVs intravenously into mice with induced models of SS. Considering the limited expandability, significant donor variations, and safety concerns of MSC sources, it is essential to optimize a protocol that can be easily scaled up to produce standardized iPSC-MSCs, showing the same potential to prevent the progression of SS[148].

Taken together, a combination of MSC and MSC-EVs with emerging technologies may provide novel insight for into the therapy of eye diseases.

CHALLENGES AND PROSPECTS

Although MSC-EVs are regarded as a new treatment strategy, their affiliated clinical challenges are worthy of further assessment.

Firstly, obtaining an appropriate cell line is a prerequisite for collecting EVs. The existing research on eye diseases is summarized in Table 2. Therapeutic vesicles are mainly sourced from cells of adipose, bone marrow, umbilical cord, and cornea. However, there is a lack of research comparing MSCs from various sources for the treatment of eye diseases. The impurity of MSCs leads to the complexity of the contents of EVs, negatively influencing their performance. Therefore, to obtain high-quality products, the tissue source, identification, and functional testing of MSCs are required. Despite that, the mechanism of MSC-EVs for other diseases has been studied in detail [Table 1], EV contents such as protein and genetic cargo for the therapy of eye diseases remain unknown. Most studies still observe the curative effect by injecting intact MSC-EVs into model animals [Table 2]. Above all, the personalized design of EVs is also essential; to date, based on the pathogenesis and pathological process, various genetic or non-genetic engineering methods have been developed for producing EVs with specific biological characteristics[148-152]. To achieve precision treatment, it is necessary to master the knowledge of EVs related to the composition, identification, purification, and function of distinct cell origins and under various physiological statuses.

Another question is how to achieve a standardized and stable production of EVs as drugs with a GMP level. Studies showed that the heterogeneity of EVs derives from their size, contents, and cell origin[153]. Proteomics analysis of EVs revealed the heterogeneity of protein profiles, suggesting that there is an urgent need to optimize and standardize the purification method to obtain high-quality EVs[154]. Despite the emergence of many novel EV preparation techniques, a consensus on manufacturing the therapeutic vesicles from cell culture needs to be reached. Referring to the recent ISEV workshop position papers, there are some issues that should be considered[155]. First, the maximum cell death rate must be less than 5% to prevent dead cells from releasing particles unrelated to the therapeutic purpose that affect EV purity. Secondly, the detection of cellular microbial contamination such as mycoplasma and viruses must be performed to meet the requirements of standard level for clinical use. In addition, the standardized protocol of MSC-EV preparation from cultured cell-conditioned-medium should be automated, timesaving, and have a high recovery efficiency. Some operational considerations need to be noted, for example, avoiding repeated freeze–thaw samples and ensuring temperature control during EV separation to prevent the destruction of functional molecules in vesicles. The EXODUS platform is a promising tool that highly satisfies all demands for the collection of EVs in a large-volume culture medium[64]. Finally, the quality of the MSC-EV preparation should be evaluated by the size, morphology, specific markers, and detection of contaminants. The scientific storage and transportation conditions of MSC-EVs are important to ensure the efficacy.

Given the security of EVs, the mechanism-dependent and safety data of MSC-EVs mainly originate from preclinical in vitro and animal research. However, it is important to indicate whether the results of the application of MSC-EVs in animal experiments can be reliably used in human clinical trials. This relies on conducting a large number of clinical trials. For the treatment of eye diseases, the administration routes of EVs mainly involve injection, eye drops, and dressing. MSC-EVs serve as an ideal source for drug delivery, regardless of encapsulating in biomaterials or dissolving in liquid[156], maintaining the biological activity.

CONCLUSIONS

The main limitation of MSC therapy for optic neuritis is the difficulty of reaching the site of pathology in the optic nerve and retina. In the view that EVs can cross the BBB, while MSCs cannot, and deliver various therapeutic factors to the brain, MSC-EVs have been extensively tested as a beneficial treatment for the control of chronic inflammation of the central nervous system. Adequate EVs can be produced by large-scale expanding parental cells. For precision medicine, engineering and modification of EVs can improve the targeted drug delivery efficiency by overexpressing therapeutic molecules, such as miRNAs. Although additional advanced research is still required to explore the mechanism of EVs in the therapy of various eye diseases, it is undeniable that MSC-EVs have promising prospects in ocular repairing.

DECLARATIONS

Authors’ contributions

Conceptualized the manuscript: Liu F, Hu L

Wrote the first draft of the manuscript: Liu X

Supplied technical knowledge to support the manuscript throughout the revision process: Liu F, Hu L

Drew figures: Liu X

Contributed to the manuscript revision: Liu F, Hu L

All the authors read and approved the submitted version of the manuscript

Availability of data and materials

Not applicable.

Financial support and sponsorship

The work was supported by the research funds provided by the National Natural Science Foundation of China (21904098), the Zhejiang Provincial and Ministry of Health Research Fund for Medical Sciences (WKJ-ZJ-1910), the Wenzhou Basic Research Project (Y2020202), and the Wenzhou Institute of Biomaterials & Engineering (WIBEZD2017006-05).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2022.

REFERENCES

1. Tsiapalis D, O’Driscoll L. Mesenchymal stem cell derived extracellular vesicles for tissue engineering and regenerative medicine applications. Cells 2020;9:991.

2. Hsu TW, Lu YJ, Lin YJ, et al. Transplantation of 3D MSC/HUVEC spheroids with neuroprotective and proangiogenic potentials ameliorates ischemic stroke brain injury. Biomaterials 2021;272:120765.

3. Regmi S, Seo Y, Ahn JS, et al. Heterospheroid formation improves therapeutic efficacy of mesenchymal stem cells in murine colitis through immunomodulation and epithelial regeneration. Biomaterials 2021;271:120752.

4. Cocucci E, Meldolesi J. Ectosomes and exosomes: shedding the confusion between extracellular vesicles. Trends Cell Biol 2015;25:364-72.

5. Uccelli A, Moretta L, Pistoia V. Mesenchymal stem cells in health and disease. Nat Rev Immunol 2008;8:726-36.

6. Rani S, Ritter T. The exosome-A naturally secreted nanoparticle and its application to wound healing. Adv Mater 2016;28:5542-52.

7. Niel G, D’Angelo G, Raposo G. Shedding light on the cell biology of extracellular vesicles. Nat Rev Mol Cell Biol 2018;19:213-28.

8. Boriachek K, Islam MN, Möller A, et al. Biological functions and current advances in isolation and detection strategies for exosome nanovesicles. Small 2018;14:1702153.

9. Clayton A, Boilard E, Buzas EI, et al. Considerations towards a roadmap for collection, handling and storage of blood extracellular vesicles. J Extracell Vesicles 2019;8:1647027.

10. Théry C, Witwer KW, Aikawa E, 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.

11. Chen C, Zong S, Liu Y, et al. Profiling of exosomal biomarkers for accurate cancer identification: combining DNA-PAINT with machine- learning-based classification. Small 2019;15:e1901014.

12. Huang T, Song C, Zheng L, Xia L, Li Y, Zhou Y. The roles of extracellular vesicles in gastric cancer development, microenvironment, anti-cancer drug resistance, and therapy. Mol Cancer 2019;18:62.

13. Zhao Y, Weber SR, Lease J, et al. Liquid biopsy of vitreous reveals an abundant vesicle population consistent with the size and morphology of exosomes. Transl Vis Sci Technol 2018;7:6.

14. Li S, Li Y, Chen B, et al. ExoRBase: a database of circRNA, lncRNA and mRNA in human blood exosomes. Nucleic Acids Res 2018;46:D106-12.

15. Thakur BK, Zhang H, Becker A, et al. Double-stranded DNA in exosomes: a novel biomarker in cancer detection. Cell Res 2014;24:766-9.

16. Skotland T, Sandvig K, Llorente A. Lipids in exosomes: current knowledge and the way forward. Prog Lipid Res 2017;66:30-41.

17. Colombo M, Raposo G, Théry C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu Rev Cell Dev Biol 2014;30:255-89.

18. Barile L, Moccetti T, Marbán E, Vassalli G. Roles of exosomes in cardioprotection. Eur Heart J 2017;38:1372-9.

19. Hori J, Yamaguchi T, Keino H, Hamrah P, Maruyama K. Immune privilege in corneal transplantation. Prog Retin Eye Res 2019;72:100758.

20. London A, Benhar I, Schwartz M. The retina as a window to the brain-from eye research to CNS disorders. Nat Rev Neurol 2013;9:44-53.

21. Perez VL, Caspi RR. Immune mechanisms in inflammatory and degenerative eye disease. Trends Immunol 2015;36:354-63.

22. Stein-Streilein J. Immune regulation and the eye. Trends Immunol 2008;29:548-54.

23. Robbins PD, Morelli AE. Regulation of immune responses by extracellular vesicles. Nat Rev Immunol 2014;14:195-208.

24. Xu H, Jia S, Xu H. Potential therapeutic applications of exosomes in different autoimmune diseases. Clin Immunol 2019;205:116-24.

25. Mansoor H, Ong HS, Riau AK, Stanzel TP, Mehta JS, Yam GH. Current trends and future perspective of mesenchymal stem cells and exosomes in corneal diseases. Int J Mol Sci 2019;20:2853.

26. Klingeborn M, Dismuke WM, Bowes Rickman C, Stamer WD. Roles of exosomes in the normal and diseased eye. Prog Retin Eye Res 2017;59:158-77.

27. Liu S, Wang W, Ning Y, et al. Exosome-mediated miR-7-5p delivery enhances the anticancer effect of Everolimus via blocking MNK/eIF4E axis in non-small cell lung cancer. Cell Death Dis 2022;13:129.

28. Yu B, Shao H, Su C, et al. Exosomes derived from MSCs ameliorate retinal laser injury partially by inhibition of MCP-1. Sci Rep 2016;6:34562.

29. Batsali AK, Georgopoulou A, Mavroudi I, Matheakakis A, Pontikoglou CG, Papadaki HA. The role of bone marrow mesenchymal stem cell derived extracellular vesicles (MSC-EVs) in normal and abnormal hematopoiesis and their therapeutic potential. J Clin Med 2020;9:856.

30. Phinney DG, Pittenger MF. Concise review: MSC-derived exosomes for cell-free therapy. Stem Cells 2017;35:851-8.

31. Mead B, Tomarev S. Extracellular vesicle therapy for retinal diseases. Prog Retin Eye Res 2020;79:100849.

32. Anderson ME, Goldhaber J, Houser SR, Puceat M, Sussman MA. Embryonic stem cell-derived cardiac myocytes are not ready for human trials. Circ Res 2014;115:335-8.

33. Xie Y, Chen Y, Zhang L, Ge W, Tang P. The roles of bone-derived exosomes and exosomal microRNAs in regulating bone remodelling. J Cell Mol Med 2017;21:1033-41.

34. Di Pace AL, Tumino N, Besi F, et al. Characterization of human NK cell-derived exosomes: role of DNAM1 receptor in exosome-mediated cytotoxicity against tumor. Cancers (Basel) 2020;12:661.

35. Counil H, Krantic S. Synaptic activity and (neuro)inflammation in Alzheimer’s disease: could exosomes be an additional link? J Alzheimers Dis 2020;74:1029-43.

36. Xia X, Zhang L, Chi J, et al. Helicobacter pylori infection impairs endothelial function through an exosome-mediated mechanism. J Am Heart Assoc 2020;9:e014120.

37. Li C, Guo F, Wang X, et al. Exosome-based targeted RNA delivery for immune tolerance induction in skin transplantation. J Biomed Mater Res A 2020;108:1493-500.

38. Ko KW, Park SY, Lee EH, et al. Integrated bioactive scaffold with polydeoxyribonucleotide and stem-cell-derived extracellular vesicles for kidney regeneration. ACS Nano 2021;15:7575-85.

39. Zhang C, Shang Y, Chen X, et al. Supramolecular nanofibers containing arginine-glycine-aspartate (RGD) peptides boost therapeutic efficacy of extracellular vesicles in kidney repair. ACS Nano 2020;14:12133-47.

40. Huang CC, Kang M, Lu Y, et al. Functionally engineered extracellular vesicles improve bone regeneration. Acta Biomater 2020;109:182-94.

41. Zhang K, Dong C, Chen M, et al. Extracellular vesicle-mediated delivery of miR-101 inhibits lung metastasis in osteosarcoma. Theranostics 2020;10:411-25.

42. You B, Jin C, Zhang J, et al. MSC-derived extracellular vesicle-delivered L-PGDS inhibit gastric cancer progression by suppressing cancer cell stemness and STAT3 phosphorylation. Stem Cells Int 2022;2022:9668239.

43. Kalimuthu S, Gangadaran P, Li XJ, et al. In vivo therapeutic potential of mesenchymal stem cell-derived extracellular vesicles with optical imaging reporter in tumor mice model. Sci Rep 2016;6:30418.

44. Ma Y, Dong L, Zhou D, et al. Extracellular vesicles from human umbilical cord mesenchymal stem cells improve nerve regeneration after sciatic nerve transection in rats. J Cell Mol Med 2019;23:2822-35.

45. Qian Y, Li Q, Chen L, et al. Mesenchymal stem cell-derived extracellular vesicles alleviate M1 microglial activation in brain injury of mice with subarachnoid hemorrhage via microRNA-140-5p delivery. Int J Neuropsychopharmacol 2022;25:328-38.

46. Sha S, Shen X, Cao Y, Qu L. Mesenchymal stem cells-derived extracellular vesicles ameliorate Alzheimer’s disease in rat models via the microRNA-29c-3p/BACE1 axis and the Wnt/β-catenin pathway. Aging (Albany NY) 2021;13:15285-306.

47. Ge L, Xun C, Li W, et al. Extracellular vesicles derived from hypoxia-preconditioned olfactory mucosa mesenchymal stem cells enhance angiogenesis via miR-612. J Nanobiotechnology 2021;19:380.

48. Sánchez-Sánchez R, Gómez-Ferrer M, Reinal I, et al. MiR-4732-3p in extracellular vesicles from mesenchymal stromal cells is cardioprotective during myocardial ischemia. Front Cell Dev Biol 2021;9:734143.

49. Wang K, Li F, Yuan Y, et al. Synovial mesenchymal stem cell-derived EV-packaged miR-31 downregulates histone demethylase KDM2A to prevent knee osteoarthritis. Mol Ther Nucleic Acids 2020;22:1078-91.

50. Dong L, Wang Y, Zheng T, et al. Hypoxic hUCMSC-derived extracellular vesicles attenuate allergic airway inflammation and airway remodeling in chronic asthma mice. Stem Cell Res Ther 2021;12:4.

51. Haynesworth SE, Baber MA, Caplan AI. Cytokine expression by human marrow-derived mesenchymal progenitor cells in vitro: effects of dexamethasone and IL-1α. J Cell Physiol 1996;166:585-92.

52. Bruno S, Grange C, Deregibus MC, et al. Mesenchymal stem cell-derived microvesicles protect against acute tubular injury. J Am Soc Nephrol 2009;20:1053-67.

53. Lai RC, Tan SS, Teh BJ, et al. Proteolytic potential of the MSC exosome proteome: implications for an exosome-mediated delivery of therapeutic proteasome. Int J Proteomics 2012;2012:971907.

54. Chen TS, Lai RC, Lee MM, Choo AB, Lee CN, Lim SK. Mesenchymal stem cell secretes microparticles enriched in pre-microRNAs. Nucleic Acids Res 2010;38:215-24.

55. Yu T, Zhao C, Hou S, Zhou W, Wang B, Chen Y. Exosomes secreted from miRNA-29b-modified mesenchymal stem cells repaired spinal cord injury in rats. Braz J Med Biol Res 2019;52:e8735.

56. Borderud SP, Li Y, Burkhalter JE, Sheffer CE, Ostroff JS. Electronic cigarette use among patients with cancer: characteristics of electronic cigarette users and their smoking cessation outcomes. Cancer 2014;120:3527-35.

57. Xie K, Liu L, Chen J, Liu F. Exosomal miR-1246 derived from human umbilical cord blood mesenchymal stem cells attenuates hepatic ischemia reperfusion injury by modulating T helper 17/regulatory T balance. IUBMB Life 2019;71:2020-30.

58. Lee K, Shao H, Weissleder R, Lee H. Acoustic purification of extracellular microvesicles. ACS Nano 2015;9:2321-7.

59. Zhang H, Lyden D. Asymmetric-flow field-flow fractionation technology for exomere and small extracellular vesicle separation and characterization. Nat Protoc 2019;14:1027-53.

60. Lim J, Choi M, Lee H, et al. Direct isolation and characterization of circulating exosomes from biological samples using magnetic nanowires. J Nanobiotechnology 2019;17:1.

61. Inglis DW, Davis JA, Austin RH, Sturm JC. Critical particle size for fractionation by deterministic lateral displacement. Lab Chip 2006;6:655-8.

62. Wunsch BH, Smith JT, Gifford SM, et al. Nanoscale lateral displacement arrays for the separation of exosomes and colloids down to 20 nm. Nat Nanotechnol 2016;11:936-40.

63. Smith JT, Wunsch BH, Dogra N, et al. Integrated nanoscale deterministic lateral displacement arrays for separation of extracellular vesicles from clinically-relevant volumes of biological samples. Lab Chip 2018;18:3913-25.

64. Chen Y, Zhu Q, Cheng L, et al. Exosome detection via the ultrafast-isolation system: EXODUS. Nat Methods 2021;18:212-8.

65. Liu F, Vermesh O, Mani V, et al. The exosome total isolation chip. ACS Nano 2017;11:10712-23.

66. Panchalingam KM, Jung S, Rosenberg L, Behie LA. Bioprocessing strategies for the large-scale production of human mesenchymal stem cells: a review. Stem Cell Res Ther 2015;6:225.

67. Petry F, Smith JR, Leber J, Salzig D, Czermak P, Weiss ML. Manufacturing of human umbilical cord mesenchymal stromal cells on microcarriers in a dynamic system for clinical use. Stem Cells Int 2016;2016:4834616.

68. Rafiq QA, Coopman K, Nienow AW, Hewitt CJ. Systematic microcarrier screening and agitated culture conditions improves human mesenchymal stem cell yield in bioreactors. Biotechnol J 2016;11:473-86.

69. Huleihel L, Hussey GS, Naranjo JD, et al. Matrix-bound nanovesicles within ECM bioscaffolds. Sci Adv 2016;2:e1600502.

70. Wang W, Itaka K, Ohba S, et al. 3D spheroid culture system on micropatterned substrates for improved differentiation efficiency of multipotent mesenchymal stem cells. Biomaterials 2009;30:2705-15.

71. Cone AS, Yuan X, Sun L, et al. Mesenchymal stem cell-derived extracellular vesicles ameliorate Alzheimer’s disease-like phenotypes in a preclinical mouse model. Theranostics 2021;11:8129-42.

72. Mendt M, Kamerkar S, Sugimoto H, et al. Generation and testing of clinical-grade exosomes for pancreatic cancer. JCI Insight 2018;3:99263.

73. Cui Y, Guo Y, Kong L, et al. A bone-targeted engineered exosome platform delivering siRNA to treat osteoporosis. Bioact Mater 2022;10:207-21.

74. Song P, Song N, Li L, Wu M, Lu Z, Zhao X. Angiopep-2-modified carboxymethyl chitosan-based pH/reduction dual-stimuli-responsive nanogels for enhanced targeting glioblastoma. Biomacromolecules 2021;22:2921-34.

75. Xu H, Li C, Wei Y, et al. Angiopep-2-modified calcium arsenite-loaded liposomes for targeted and pH-responsive delivery for anti-glioma therapy. Biochem Biophys Res Commun 2021;551:14-20.

76. Liu X, Wu R, Li Y, et al. Angiopep-2-functionalized nanoparticles enhance transport of protein drugs across intestinal epithelia by self-regulation of targeted receptors. Biomater Sci 2021;9:2903-16.

77. Wu JY, Li YJ, Wang J, et al. Multifunctional exosome-mimetics for targeted anti-glioblastoma therapy by manipulating protein corona. J Nanobiotechnology 2021;19:405.

78. Tran PHL, Xiang D, Tran TTD, et al. Exosomes and nanoengineering: a match made for precision therapeutics. Adv Mater 2020;32:e1904040.

79. Du W, Zhang K, Zhang S, et al. Enhanced proangiogenic potential of mesenchymal stem cell-derived exosomes stimulated by a nitric oxide releasing polymer. Biomaterials 2017;133:70-81.

80. Willis GR, Fernandez-Gonzalez A, Anastas J, et al. Mesenchymal stromal cell exosomes ameliorate experimental bronchopulmonary dysplasia and restore lung function through macrophage immunomodulation. Am J Respir Crit Care Med 2018;197:104-16.

81. Ouyang X, Han X, Chen Z, Fang J, Huang X, Wei H. MSC-derived exosomes ameliorate erectile dysfunction by alleviation of corpus cavernosum smooth muscle apoptosis in a rat model of cavernous nerve injury. Stem Cell Res Ther 2018;9:246.

82. de Araujo Farias V, O’Valle F, Serrano-Saenz S, et al. Exosomes derived from mesenchymal stem cells enhance radiotherapy-induced cell death in tumor and metastatic tumor foci. Mol Cancer 2018;17:122.

83. Petrović A, Bogojević D, Korać A, et al. Oxidative stress-dependent contribution of HMGB1 to the interplay between apoptosis and autophagy in diabetic rat liver. J Physiol Biochem 2017;73:511-21.

84. Jin Z, Ren J, Qi S. Human bone mesenchymal stem cells-derived exosomes overexpressing microRNA-26a-5p alleviate osteoarthritis via down-regulation of PTGS2. Int Immunopharmacol 2020;78:105946.

85. Zhang X, Hsueh MF, Huebner JL, Kraus VB. TNF-α carried by Plasma extracellular vesicles predicts knee osteoarthritis progression. Front Immunol 2021;12:758386.

86. Ju C, Liu R, Zhang Y, et al. Exosomes may be the potential new direction of research in osteoarthritis management. Biomed Res Int 2019;2019:7695768.

87. Kuang MJ, Huang Y, Zhao XG, et al. Exosomes derived from Wharton’s jelly of human umbilical cord mesenchymal stem cells reduce osteocyte apoptosis in glucocorticoid-induced osteonecrosis of the femoral head in rats via the miR-21-PTEN-AKT signalling pathway. Int J Biol Sci 2019;15:1861-71.

88. Hancox Z, Heidari Keshel S, Yousaf S, Saeinasab M, Shahbazi MA, Sefat F. The progress in corneal translational medicine. Biomater Sci 2020;8:6469-504.

89. Qi X, Xie L, Cheng J, Zhai H, Zhou Q. Characteristics of immune rejection after allogeneic cultivated limbal epithelial transplantation. Ophthalmology 2013;120:931-6.

90. Karp JM, Leng Teo GS. Mesenchymal stem cell homing: the devil is in the details. Cell Stem Cell 2009;4:206-16.

91. Navas A, Magaña-Guerrero FS, Domínguez-López A, et al. Anti-inflammatory and anti-fibrotic effects of human amniotic membrane mesenchymal stem cells and their potential in corneal repair. Stem Cells Transl Med 2018;7:906-17.

92. Galindo S, Herreras JM, López-Paniagua M, et al. Therapeutic effect of human adipose tissue-derived mesenchymal stem cells in experimental corneal failure due to limbal stem cell niche damage. Stem Cells 2017;35:2160-74.

93. Lu X, Ru Y, Chu C, et al. Lentivirus-mediated IL-10-expressing bone marrow mesenchymal stem cells promote corneal allograft survival via upregulating lncRNA 003946 in a rat model of corneal allograft rejection. Theranostics 2020;10:8446-67.

94. Yang YH, Hsieh TL, Ji AT, et al. Stromal tissue rigidity promotes mesenchymal stem cell-mediated corneal wound healing through the transforming growth factor β signaling pathway. Stem Cells 2016;34:2525-35.

95. Tang Q, Luo C, Lu B, et al. Thermosensitive chitosan-based hydrogels releasing stromal cell derived factor-1 alpha recruit MSC for corneal epithelium regeneration. Acta Biomater 2017;61:101-13.

96. Fernandes-Cunha GM, Na KS, Putra I, et al. Corneal wound healing effects of mesenchymal stem cell secretome delivered within a viscoelastic gel carrier. Stem Cells Transl Med 2019;8:478-89.

97. Nuzzi R, Buono L, Scalabrin S, De Iuliis M, Bussolati B. Effect of stem cell-derived extracellular vesicles on damaged human corneal endothelial cells. Stem Cells Int 2021;2021:6644463.

98. Park GW, Heo J, Kang JY, et al. Topical cell-free conditioned media harvested from adipose tissue-derived stem cells promote recovery from corneal epithelial defects caused by chemical burns. Sci Rep 2020;10:12448.

99. Samaeekia R, Rabiee B, Putra I, et al. Effect of human corneal mesenchymal stromal cell-derived exosomes on corneal epithelial wound healing. Invest Ophthalmol Vis Sci 2018;59:5194-200.

100. Shojaati G, Khandaker I, Funderburgh ML, et al. Mesenchymal stem cells reduce corneal fibrosis and inflammation via extracellular vesicle-mediated delivery of miRNA. Stem Cells Transl Med 2019;8:1192-201.

101. Ha DH, Kim SD, Lee J, et al. Toxicological evaluation of exosomes derived from human adipose tissue-derived mesenchymal stem/stromal cells. Regul Toxicol Pharmacol 2020;115:104686.

102. Carter K, Lee HJ, Na KS, et al. Characterizing the impact of 2D and 3D culture conditions on the therapeutic effects of human mesenchymal stem cell secretome on corneal wound healing in vitro and ex vivo. Acta Biomater 2019;99:247-57.

103. Zhang W, Wang Y, Kong Y. Exosomes derived from mesenchymal stem cells modulate miR-126 to ameliorate hyperglycemia-induced retinal inflammation via targeting HMGB1. Invest Ophthalmol Vis Sci 2019;60:294-303.

104. Safwat A, Sabry D, Ragiae A, Amer E, Mahmoud RH, Shamardan RM. Adipose mesenchymal stem cells-derived exosomes attenuate retina degeneration of streptozotocin-induced diabetes in rabbits. J Circ Biomark 2018;7:1849454418807827.

105. Mathew B, Ravindran S, Liu X, et al. Mesenchymal stem cell-derived extracellular vesicles and retinal ischemia-reperfusion. Biomaterials 2019;197:146-60.

106. Shigemoto-Kuroda T, Oh JY, Kim DK, et al. MSC-derived extracellular vesicles attenuate immune responses in two autoimmune murine models: type 1 diabetes and uveoretinitis. Stem Cell Reports 2017;8:1214-25.

107. He GH, Zhang W, Ma YX, et al. Mesenchymal stem cells-derived exosomes ameliorate blue light stimulation in retinal pigment epithelium cells and retinal laser injury by VEGF-dependent mechanism. Int J Ophthalmol 2018;11:559-66.

108. Pan D, Chang X, Xu M, et al. UMSC-derived exosomes promote retinal ganglion cells survival in a rat model of optic nerve crush. J Chem Neuroanat 2019;96:134-9.

109. Mead B, Ahmed Z, Tomarev S. Mesenchymal stem cell-derived small extracellular vesicles promote neuroprotection in a genetic DBA/2J mouse model of glaucoma. Invest Ophthalmol Vis Sci 2018;59:5473-80.

110. Seyedrazizadeh SZ, Poosti S, Nazari A, et al. Extracellular vesicles derived from human ES-MSCs protect retinal ganglion cells and preserve retinal function in a rodent model of optic nerve injury. Stem Cell Res Ther 2020;11:203.

111. Ferguson SW, Nguyen J. Exosomes as therapeutics: the implications of molecular composition and exosomal heterogeneity. J Control Release 2016;228:179-90.

112. Chatterjee N, Rana S, Espinosa-Diez C, Anand S. MicroRNAs in cancer: challenges and opportunities in early detection, disease monitoring, and therapeutic agents. Curr Pathobiol Rep 2017;5:35-42.

113. Salehi M, Sharifi M. Exosomal miRNAs as novel cancer biomarkers: challenges and opportunities. J Cell Physiol 2018;233:6370-80.

114. Zimta AA, Baru O, Badea M, Buduru SD, Berindan-Neagoe I. The role of angiogenesis and pro-angiogenic exosomes in regenerative dentistry. Int J Mol Sci 2019;20:406.

115. Reza-Zaldivar EE, Hernández-Sapiéns MA, Minjarez B, Gutiérrez-Mercado YK, Márquez-Aguirre AL, Canales-Aguirre AA. Potential effects of MSC-derived exosomes in neuroplasticity in Alzheimer’s disease. Front Cell Neurosci 2018;12:317.

116. Ueno M, Asada K, Toda M, et al. Concomitant evaluation of a panel of exosome proteins and MiRs for qualification of cultured human corneal endothelial cells. Invest Ophthalmol Vis Sci 2016;57:4393-402.

117. Petrenko Y, Vackova I, Kekulova K, et al. A comparative analysis of multipotent mesenchymal stromal cells derived from different sources, with a focus on neuroregenerative potential. Sci Rep 2020;10:4290.

118. Osborne A, Sanderson J, Martin KR. Neuroprotective effects of human mesenchymal stem cells and platelet-derived growth factor on human retinal ganglion cells. Stem Cells 2018;36:65-78.

119. Harrell CR, Fellabaum C, Arsenijevic A, Markovic BS, Djonov V, Volarevic V. Therapeutic potential of mesenchymal stem cells and their secretome in the treatment of glaucoma. Stem Cells Int 2019;2019:7869130.

120. Cao X, Xue LD, Di Y, Li T, Tian YJ, Song Y. MSC-derived exosomal lncRNA SNHG7 suppresses endothelial-mesenchymal transition and tube formation in diabetic retinopathy via miR-34a-5p/XBP1 axis. Life Sci 2021;272:119232.

121. Tang ST, Wang F, Shao M, Wang Y, Zhu HQ. MicroRNA-126 suppresses inflammation in endothelial cells under hyperglycemic condition by targeting HMGB1. Vascul Pharmacol 2017;88:48-55.

122. Zhang X, Liu J, Yu B, Ma F, Ren X, Li X. Effects of mesenchymal stem cells and their exosomes on the healing of large and refractory macular holes. Graefes Arch Clin Exp Ophthalmol 2018;256:2041-52.

123. Shi Y, Inoue H, Wu JC, Yamanaka S. Induced pluripotent stem cell technology: a decade of progress. Nat Rev Drug Discov 2017;16:115-30.

124. Reichman S, Goureau O. Production of retinal cells from confluent human iPS cells. In: Turksen K, Nagy A, editors. Induced pluripotent stem (iPS) cells. New York: Springer; 2016. p. 339-51.

125. Jin ZB, Okamoto S, Mandai M, Takahashi M. Induced pluripotent stem cells for retinal degenerative diseases: a new perspective on the challenges. J Genet 2009;88:417-24.

126. Sun J, Mandai M, Kamao H, et al. Protective effects of human iPS-derived retinal pigmented epithelial cells in comparison with human mesenchymal stromal cells and human neural stem cells on the degenerating retina in rd1 mice. Stem Cells 2015;33:1543-53.

127. Reichman S, Slembrouck A, Gagliardi G, et al. Generation of storable retinal organoids and retinal pigmented epithelium from adherent human iPS cells in xeno-free and feeder-free conditions. Stem Cells 2017;35:1176-88.

128. Chang L, Kan L. Mesenchymal stem cell-originated exosomal circular RNA circFBXW7 attenuates cell proliferation, migration and inflammation of fibroblast-like synoviocytes by targeting miR-216a-3p/HDAC4 in rheumatoid arthritis. J Inflamm Res 2021;14:6157-71.

129. Riazifar M, Mohammadi MR, Pone EJ, et al. Stem cell-derived exosomes as nanotherapeutics for autoimmune and neurodegenerative disorders. ACS Nano 2019;13:6670-88.

130. Li Y, Ma K, Zhang L, Xu H, Zhang N. Human umbilical cord blood derived-mesenchymal stem cells alleviate dextran sulfate sodium-induced colitis by increasing regulatory T cells in mice. Front Cell Dev Biol 2020;8:604021.

131. Cao L, Xu H, Wang G, Liu M, Tian D, Yuan Z. Extracellular vesicles derived from bone marrow mesenchymal stem cells attenuate dextran sodium sulfate-induced ulcerative colitis by promoting M2 macrophage polarization. Int Immunopharmacol 2019;72:264-74.

132. Hyvärinen K, Holopainen M, Skirdenko V, et al. Mesenchymal stromal cells and their extracellular vesicles enhance the anti-inflammatory phenotype of regulatory macrophages by downregulating the production of interleukin (IL)-23 and IL-22. Front Immunol 2018;9:771.

133. Zhao H, Shang Q, Pan Z, et al. Exosomes from adipose-derived stem cells attenuate adipose inflammation and obesity through polarizing M2 macrophages and beiging in white adipose tissue. Diabetes 2018;67:235-47.

134. Fan Y, Herr F, Vernochet A, Mennesson B, Oberlin E, Durrbach A. Human fetal liver mesenchymal stem cell-derived exosomes impair natural killer cell function. Stem Cells Dev 2019;28:44-55.

135. Khare D, Or R, Resnick I, Barkatz C, Almogi-Hazan O, Avni B. Mesenchymal stromal cell-derived exosomes affect mRNA expression and function of B-lymphocytes. Front Immunol 2018;9:3053.

136. Adamo A, Brandi J, Caligola S, et al. Extracellular vesicles mediate mesenchymal stromal cell-dependent regulation of B cell PI3K-AKT signaling pathway and actin cytoskeleton. Front Immunol 2019;10:446.

137. Blazquez R, Sanchez-Margallo FM, de la Rosa O, et al. Immunomodulatory potential of human adipose mesenchymal stem cells derived exosomes on in vitro stimulated T cells. Front Immunol 2014;5:556.

138. Tian J, Zhu Q, Zhang Y, et al. Olfactory ecto-mesenchymal stem cell-derived exosomes ameliorate experimental colitis via modulating Th1/Th17 and treg cell responses. Front Immunol 2020;11:598322.

139. Zhang L, Zheng H, Shao H, et al. Long-term therapeutic effects of mesenchymal stem cells compared to dexamethasone on recurrent experimental autoimmune uveitis of rats. Invest Ophthalmol Vis Sci 2014;55:5561-71.

140. Dong L, Chen X, Shao H, Bai L, Li X, Zhang X. Mesenchymal stem cells inhibited dendritic cells via the regulation of STAT1 and STAT6 phosphorylation in experimental autoimmune uveitis. Curr Mol Med 2018;17:478-87.

141. Bai L, Shao H, Wang H, et al. Author correction: effects of mesenchymal stem cell-derived exosomes on experimental autoimmune uveitis. Sci Rep 2018;8:9889.

142. Sandhya P, Kurien BT, Danda D, Scofield RH. Update on pathogenesis of Sjogren’s syndrome. Curr Rheumatol Rev 2017;13:5-22.

143. Mavragani CP. Mechanisms and new strategies for primary Sjögren’s syndrome. Annu Rev Med 2017;68:331-43.

144. Yao G, Qi J, Liang J, et al. Mesenchymal stem cell transplantation alleviates experimental Sjögren’s syndrome through IFN-β/IL-27 signaling axis. Theranostics 2019;9:8253-65.

145. Shi B, Qi J, Yao G, et al. Mesenchymal stem cell transplantation ameliorates Sjögren’s syndrome via suppressing IL-12 production by dendritic cells. Stem Cell Res Ther 2018;9:308.

146. Xu J, Wang D, Liu D, et al. Allogeneic mesenchymal stem cell treatment alleviates experimental and clinical Sjögren syndrome. Blood 2012;120:3142-51.

147. Rui K, Hong Y, Zhu Q, et al. Olfactory ecto-mesenchymal stem cell-derived exosomes ameliorate murine Sjögren’s syndrome by modulating the function of myeloid-derived suppressor cells. Cell Mol Immunol 2021;18:440-51.

148. Hai B, Shigemoto-Kuroda T, Zhao Q, Lee RH, Liu F. Inhibitory effects of iPSC-MSCs and their extracellular vesicles on the onset of sialadenitis in a mouse model of Sjögren’s syndrome. Stem Cells Int 2018;2018:2092315.

149. Richter M, Vader P, Fuhrmann G. Approaches to surface engineering of extracellular vesicles. Adv Drug Deliv Rev 2021;173:416-26.

150. Colao IL, Corteling R, Bracewell D, Wall I. Manufacturing exosomes: a promising therapeutic platform. Trends Mol Med 2018;24:242-56.

151. Yao X, Wei W, Wang X, Chenglin L, Björklund M, Ouyang H. Stem cell derived exosomes: microRNA therapy for age-related musculoskeletal disorders. Biomaterials 2019;224:119492.

152. Bei HP, Hung PM, Yeung HL, Wang S, Zhao X. Bone-a-petite: engineering exosomes towards bone, osteochondral, and cartilage repair. Small 2021;17:e2101741.

153. Kalluri R, LeBleu VS. The biology, function, and biomedical applications of exosomes. Science 2020;367:eaau6977.

154. Kowal J, Arras G, Colombo M, et al. Proteomic comparison defines novel markers to characterize heterogeneous populations of extracellular vesicle subtypes. Proc Natl Acad Sci U S A 2016;113:E968-77.

155. Witwer KW, Buzás EI, Bemis LT, et al. Standardization of sample collection, isolation and analysis methods in extracellular vesicle research. J Extracell Vesicles 2013;2:20360.

156. Lai RC, Yeo RW, Tan KH, Lim SK. Exosomes for drug delivery - a novel application for the mesenchymal stem cell. Biotechnol Adv 2013;31:543-51.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Liu X, Hu L, Liu F. Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases. Extracell Vesicles Circ Nucleic Acids 2022;3:102-17. http://dx.doi.org/10.20517/evcna.2022.08

AMA Style

Liu X, Hu L, Liu F. Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases. Extracellular Vesicles and Circulating Nucleic Acids. 2022; 3(2): 102-17. http://dx.doi.org/10.20517/evcna.2022.08

Chicago/Turabian Style

Liu, Xiaoling, Liang Hu, Fei Liu. 2022. "Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases" Extracellular Vesicles and Circulating Nucleic Acids. 3, no.2: 102-17. http://dx.doi.org/10.20517/evcna.2022.08

ACS Style

Liu, X.; Hu L.; Liu F. Mesenchymal stem cell-derived extracellular vesicles for cell-free therapy of ocular diseases. Extracell. Vesicles. Circ. Nucleic. Acids. 2022, 3, 102-17. http://dx.doi.org/10.20517/evcna.2022.08

About This Article

Special Issue

© The Author(s) 2022. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
2653
Downloads
680
Citations
1
Comments
0
35

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 20 clicks
Like This Article 35 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Extracellular Vesicles and Circulating Nucleic Acids
ISSN 2767-6641 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/