Skip to main content

Cell-free therapy based on extracellular vesicles: a promising therapeutic strategy for peripheral nerve injury

Abstract

Peripheral nerve injury (PNI) is one of the public health concerns that can result in a loss of sensory or motor function in the areas in which injured and non-injured nerves come together. Up until now, there has been no optimized therapy for complete nerve regeneration after PNI. Exosome-based therapies are an emerging and effective therapeutic strategy for promoting nerve regeneration and functional recovery. Exosomes, as natural extracellular vesicles, contain bioactive molecules for intracellular communications and nervous tissue function, which could overcome the challenges of cell-based therapies. Furthermore, the bioactivity and ability of exosomes to deliver various types of agents, such as proteins and microRNA, have made exosomes a potential approach for neurotherapeutics. However, the type of cell origin, dosage, and targeted delivery of exosomes still pose challenges for the clinical translation of exosome therapeutics. In this review, we have focused on Schwann cell and mesenchymal stem cell (MSC)-derived exosomes in nerve tissue regeneration. Also, we expressed the current understanding of MSC-derived exosomes related to nerve regeneration and provided insights for developing a cell-free MSC therapeutic strategy for nerve injury.

Introduction

Peripheral nerve injury (PNI) is a common neurological disorder in the clinic that seriously influences human health [1]. In these cases, patients endure neuropathic pain, which can result in dysfunction of the sensory and motor systems and also cause disability [2, 3]. Hence, the development of novel treatment strategies to enhance peripheral nerve repair post-injury, especially those that can accelerate axonal nerve regeneration, is necessary [4].

Several factors are involved in axonal outgrowth in peripheral nerve regeneration, such as the transformation of the phenotype of Schwann cells (SCs), the infiltration of immune cells, neurovascular regeneration, and also neuronal soma formation, which plays a key role in the initiation and control of axonal regeneration [5, 6]. On the other hand, one of the main challenges in the regeneration of peripheral nerves is the low speed of axon growth (only 1 mm per day) [7]. Although the peripheral nerve has the potential for self-regeneration, in several cases, such as the long length of the nerve defect, the long duration between injury and treatment limits its spontaneous self-regeneration [8]. Also, self-regeneration is often inadequate and might be prevented by scar formation [8]. One of the strategies in PNI is direct neurorrhaphy, but it can be applied only in cases with short gaps [9]. Another strategy for the treatment of large nerve defects is autologous nerve grafting, which has served as the gold standard [10]. However, autologous nerve grafts via microsurgical procedure are limited due to insufficient nerve sources, potential donor site dysfunction, size mismatch between donor nerves and nerve grafts, and other complications [11, 12].

In the field of peripheral nerve injury (PNI), several studies have been done in the past few years to find a way to replace the autologous nerve graft method in clinics with a new way to speed up axonal regeneration without harming other nerves [13]. One of the new therapeutic strategies in regenerative medicine is cell therapy. Although cell therapy has demonstrated beneficial effects on peripheral nerve tissue treatment and regeneration, it still has several drawbacks, such as a decreased survival rate of the engrafted cells, the low regenerative capacity of cells, tumorigenesis, immune-mediated rejection, the risk of capillary blockade during infusion, and ethical concerns that hinder the wide use of cells in the clinic [14, 15]. One of the alternative therapeutic strategies is the design of various tissue-engineered nerve guide conduits (NGCs) to provide mechanical, biological, and biochemical supports [16]. Although these synthetic NGCs with or without cells and growth factors (GFs) have been shown to be beneficial, the results of their use in the treatment of PNI remain far from ideal [17]. Exosome-based therapy with or without NGCs is now used in PNI as an alternative to cell therapy and tissue-engineered NGC alone.

Exosomes as extracellular vesicles

Johnstone et al. [18] showed in 1970 that exosomes are extracellular vesicles (EV) that carry different substrates. After that, several extracellular vehicles have been identified and can now be classified as apoptotic bodies, microvesicles (MVs), and exosomes (EXOs), depending on their biogenesis and size. According to this classification, EXOs are the smallest extracellular vesicles (30–100 nm) with a lipid bilayer membrane released by all types of cells, such as Schwann cells (SCs) and mesenchymal stromal cells (MSCs) [19]. Furthermore, EXOs are the key mediators of paracrine mechanisms, and the biogenesis of EXOs is the endocytic pathway. Recently, several studies have focused on EXOs, and they have demonstrated that different types of EXOs are released from certain types of cells that are associated with pathological and physiological conditions, such as neurodegenerative diseases, tumors, and tissue fibrosis [20]. On the other hand, EXOs have several cellular signaling molecules like DNA, proteins, lipids, mRNA, miRNA, lncRNA, and circRNA that mediate intercellular communication due to transferring these types of cargo [21,22,23].

The therapeutic effects of Schwann cells-derived exosome on PNI

Schwan cells(SCs), the glial cells of the PNS, are a critical factor for maintaining homeostasis in the nerves and facilitating the regeneration process of the PNI [24]. SCs provide the nutrition to support axonal regeneration, and SCs are the basic cell type that organizes the formation of myelin sheaths along the axon [25]. A chain of molecular and cellular reactions known as Wallerian degeneration (WD) was initiated in PNI. In this case, the peripheral glia (the SCs) were dedifferentiated into a non-myelinating cell type and proliferated to omit the endoneurial myelin and all debris [26,27,28]. Also, SCs secrete neurotrophic factors and specific cytokines [27]. Furthermore, miRNAs can be conveyed by EXOs from SCs to neurons to promote the regeneration of PNI. Indeed, the level of miRNA expression by SCs plays the main role in the nerve regeneration process [29, 30].

The results of several studies have demonstrated that miRNA can augment SC proliferation and axon myelination during development and injury [31]. SC-derived exosomes have been shown to be internalized by axons and enhance neurite outgrowth, and direct injection of SC-derived exosomes can improve axon growth following in vivo PNI [32]. Moreover, SC-derived exosomes can change the growth cone phenotype to a pro-regeneration morphology and decrease the activity of the GTPase RhoA, which plays a role in axon retraction and collapse of the growth cone [32].

Altogether, these studies have demonstrated that SCs can release EXOs, and these SC-derived exosomes have been illustrated to play an essential role in neurodegeneration, neurodevelopment, and neuroprotection [33, 34]. Also, a study illustrated that SC-derived exosomes can increase the axonal regeneration rate of dorsal root ganglion neurons in in vitro and in vivo investigations, indicating the role of SC-derived exosomes in axonal regeneration [32]. On the other hand, these promotive effects are dependent on the type of SC-derived exosomes released from various phenotypic SCs [35].

Recently, Lopez-Leal et al. [35] demonstrated that only the SC-derived exosomes secreted by repair SCs enhanced axonal regeneration after PNI, but they did not show these promotive effects in the differentiated SCs. Indeed, repair of SC-derived exosomes mediated the effect of promoting neurite growth from dorsal root ganglia explants by downregulating PTEN, activating PI3K pathways, and also transferring exosomal miRNA-21 (Fig. 1) [36].

Fig. 1
figure 1

Illustrating exosome contents that promote axon regeneration through the PI3/AKT signaling pathway. Schwann cell-derived exosomes released from different phenotypic Schwann cells, like mature SCs and repair SCs, carry different cargoes that influence their functions. For example, repair SC-derived exosomes exhibit axonal regeneration after nerve injury due to their containing miRNA-21. miRNA-21 can cause the downregulation of phosphatase and tensin homolog (PTEN) and consequently the activation of phosphoinositide 3-kinase (PI3K) in the neurons. Also, SC-derived exosomes can inhibit neuron apoptosis and increase cell viability. On the other hand, exosomes derived from mature myelinating SCs cannot promote axonal regeneration and also inhibit SC migration

According to the results of this study, it is obvious that the repair of SC-derived exosomes shuttled specific proteins and miRNAs that enhanced axonal regeneration and the survival of neurons, while the EXOs from differentiated SCs did not exhibit the promotive effect to enhance axonal regeneration due to the number of miRNAs that suppressed cell migration, such as miRNA-21 and miRNA-92a-3p miRNAs, so they could focus on myelination (the main function of differentiated SCs) (Fig. 1) [37]. The positive effects of SC-derived exosomes in the nerve studies are summarized in Table 1.

Table 1 Summary of the positive effects of SC-derived exosomes in the nerve injury studies

Therefore, it can be concluded that various phenotypes of SCs can transfer different exosomal cargoes that are required for specific functions. On the other hand, several studies have reported that SCs can promote cancer growth and dissemination in pancreatic cancer and melanoma. In these cases, tumor cells exploit SCs due to their exosomal cargo, which promotes higher proliferation and inhibits apoptosis [38,39,40,41]. As regards obtaining SC-derived exosomes, it is necessary to sacrifice normal nerve tissue to harvest SCs. This disadvantage is the main challenge for clinical translation [42]. Hence, we need further investigation to identify optimal EXO content and an efficient strategy to obtain SC-derived exosomes without the need to sacrifice normal nerve tissue or find an alternative cell type with similar efficiency to SCs [42].

The therapeutic effects of mesenchymal stromal cells-derived exosome on PNI

MSCs are adult multipotent stem cells that can be isolated from various human tissues (i.e., adipose, bone marrow, umbilical cord blood, dental pulp, etc.) [42]. MSCs have been identified as having multi-directional differentiation potential, high self-renewal ability, and low immunogenicity, so they are one of the most common potential off-the-shelf stem cells in cell therapy (Fig. 2) [47].

Fig. 2
figure 2

Novel strategies of MSC-derived exosomes for curing nerve injury. MSCs can be isolated from bone marrow, adipose tissue, endometrium tissues, the umbilical cord, and the dental pulp. Their exosomes can regulate nerve-related cellular functions. MSC-derived exosomes are able to modulate neuroinflammation and immune cell reactions, neuroprotection, angiogenesis, and axonal regrowth and remyelination

Several studies have shown that adult multipotent MSCs, which are similar to SCs, can help with functional recovery after PNI by encouraging the growth and survival of neurons. However, there are several drawbacks to MSC-based therapy, including its high cost, cellular phenotypic instability, and the risk of microinfarction caused by transported MSCs that become lodged in the pulmonary microvasculature [48, 49]. As a result, a new cell-free therapy with similar efficacy to that of MSCs must be developed for PNI.

Recent studies have demonstrated that the applied MSCs’ therapeutic activities are related to paracrine factors such as cytokines, proteins, and especially their EXOs [50]. Recently, EXOs have been identified as the main paracrine effectors of MSCs and can mediate cell-to-cell communication and maintain homeostatic and dynamic microenvironments for tissue regeneration [51, 52]. MSC-derived EXOs have the potential to activate PI3k/Akt, ERK, and STAT3 signaling pathways to induce the expression of several growth factors (GFs) like NGF, insulin-like growth factor-1 (IGF-1), and stromal-derived growth factor-1 (SDF-1) [53]. EXOs derived from MSCs can also activate Wnt/b-catenin and Notch signaling pathways. The Wnt signaling pathway is involved in the control of inflammation after being activated by injury [54,55,56]. Moreover, studies have demonstrated that exosomal miRNAs (miR-221, miR-218, miR-199b, miR-148a, and miR-135b) can promote neuronal differentiation, proliferation, and axonal outgrowth [57,58,59]. Also, a study has demonstrated that the miR-17-92 cluster can promote axonal outgrowth, neurogenesis, and functional recovery by activating the PI3K/protein kinase B/mechanistic target of the rapamycin/glycogen synthase kinase 3-β signaling pathway [60, 61].

In another study, Zhang et al. [62] also demonstrated that MSC-derived EXOs carry an elevated level of the miR-17-92 cluster, which can activate the PTEN/mTOR signaling pathway in recipient neuron cells. Several studies have demonstrated that MSCs with miRNA overexpression are better influenced by functional recovery in PNI situations than naive MSCs [63]. Indeed, the function of MSC-derived EXOs depends on the condition of the original cell that releases EXOs (like SC-derived EXOs), which is related to the miRNA content of EXOs and influences their biological function.

Furthermore, recent studies have shown that MSC-derived EXO can upregulate miRNA and promote angiogenesis [64]. MSC-derived EXOs are identified as the main immunomodulatory mediators due to their immunomodulatory proteins [65]. About this, several studies have demonstrated that MSC-derived EXOs have a positive immunomodulatory effect in various pathologic conditions due to their induction of high levels of anti-inflammatory cytokines like IL10 and TGF-β1 and their enhancement of the expression of IL1B, IL6, TNFA, and IL12P40 as proinflammatory factors [66]. Also, MSC-derived EXOs induce regulatory T cells (Tregs), which are recognized as immune tolerance agents [67]. The majority of these studies about the effects of MSC-derived exosomes on the nerve injuries are summarized in Table 2. Altogether, these outcomes show a superior potential role for MSC-derived EXOs and their miRNA in the regeneration of PNI (Fig. 2).

Table 2 Summary of the positive effects of MSCs-derived exosomes in the studies for PNI

MSC therapy versus MSC-derived exosomes in clinical practice

So far, MSC therapies have emerged as a powerful tool in clinical practice for the treatment of various diseases, such as neurodegenerative disease [68]. Although MSCs can differentiate into various cells like nerve cells and also regulate the microenvironment of the injured area to accelerate the regeneration of peripheral nerves [69], a few stem cell products derived from mesenchymal stromal cells and clinical trials have applied MSCs after nerve injury due to certain disadvantages of this application [70]. The lack of clinical usage of MSCs in peripheral nerve therapies may be due to several unavoidable drawbacks with regards to MSC therapy. One of these drawbacks is that the large diameter of these cells may lead to their aggravation in the lung after intravenous injection and, thus, infusion toxicity [71]. Furthermore, MSC injections may result in oncological complications. In addition, aging is another limitation [72]. Thus, the clinical application of MSCs raises some ethical and safety concerns since they are limited to nerve regeneration [69]. Therefore, the optimal approach for the use of derivatives of MSCs to take their advantages to repair PNI can be useful.

As mentioned, some of the observed beneficial effects of MSC therapies can be partly due to their paracrine action rather than the long-term engraftment of transplanted MSCs [73]. Also, studies have investigated whether MSC-derived extracellular vesicles like EXOs exert functions similar to those of MSCs by beneficially promoting peripheral nerve regeneration [74, 75]. So it can be concluded that MSC-derived exosomes, as cellular paracrine products may play a major role in recovery post-nerve injury [74]. It is widely accepted that MSC-derived EXOs play an essential role in the amelioration of disease [76]. On the other hand, the advantages of EXO therapies are related to their safety profile [68]. The safety profile of MSC-derived EXOs therapy is a critical consideration in its clinical application [77]. Current evidence suggests that EXOs have a proper safety profile, with low immunogenicity, permeability, easy storage (they can be lyophilized), and minimal adverse effects reported in clinical trials [78]. However, the manufacturing process of EXO-based clinical products needs to be standardized to ensure consistency in quality, safety, and efficacy [79]. Consequently, MSC-derived EXOs have been subject to much research interest in recent years.

Therapeutic and clinical application of exosomes in neurodegenerative disease

Angiogenesis and vascular regeneration

The vascular network plays a critical role in maintaining the microenvironment homeostasis of the peripheral nervous system through the supply of oxygen and nutrients, which are essential for the regeneration of the PNS [85]. Also, several studies have demonstrated that there is an interlinkage between nerve regeneration after injury and vascularization [86]. Furthermore, the vascular networks can provide tracks for SCs to migrate along, and endothelial cells (EC) secrete various bioactive agents that are conducive to neurite elongation [86, 87]. So, reconstruction of the vascular network following PNI is another purpose for the regeneration of the peripheral nerve. However, in current strategies of treatment such as nerve guide conduits and decellularized grafts, vascular regeneration is one of the main challenges [88, 89].

In tissue regeneration, reconstructing blood vessels and blood flow in injured and ischemic tissues is necessary [90,91,92]. Recently, several studies have illustrated that EXO can promote vascular regeneration as a key regulator. The study of Nooshabadi et al. [64] has shown that human endometrial MSC-derived EXOs have a positive effect on the angiogenesis process in a dose-dependent manner and can be applied in the treatment of vascular disease and wound healing. Zhang et al. [93] found that umbilical cord MSC-derived EXOs can promote vascular regeneration in ECs by activating the Wnt/β-catenin pathway. Also, Liu et al. [94] have shown that EXOs from induced pluripotent stem cell-derived mesenchymal stem cells enhance angiogenesis due to activating the PI3K/AKT signaling pathway in EC cells. Moreover, Gong et al. [95] demonstrated that proangiogenic miRNAs can be transferred within ECs through generated exosomes to enhance angiogenesis. Thus, it is reasonable to hypothesize that exosomes promote angiogenesis by at least three distinct mechanisms, including:

  1. 1.

    EXOs can promote EC survival and proliferation by upregulating Cyclin-D1 and downregulating p53, p21, and p27.

  2. 2.

    EXOs promote angiogenesis in EC cells by activating signaling pathways such as Wnt/β-catenin and PI3K/AKT.

  3. 3.

    EXOs contain proangiogenic miRNAs that are transferred between ECs.

On the other hand, studies have shown that EXOs control vascular regeneration after PNI to change how peripheral nerve regeneration happens. A study by Xin et al. [96] showed that the MSC-derived EXOs can enhance functional recovery, neurogenesis, neurite remodeling, and angiogenesis. Similar effects were found in another study, which showed that MSC-derived EXOs can promote neurogenesis and angiogenesis and reduce inflammation in rat models [97, 98]. All of the mentioned studies about the angiogenesis effects of MSC-derived exosomes are summarized in Table 3. In summary, all of these studies demonstrated that vascular regeneration and angiogenesis that are intermediated by EXOs are conducive to peripheral nerve regeneration, which can be a superior therapeutic strategy for PNI repair by facilitating angiogenesis and vascular regeneration.

Table 3 Summary of the angiogenesis effects of MSCs-derived

Axon outgrowth and regeneration

Recently, several studies have demonstrated that exosomes modulate axonal regeneration due to the transfer of specific exosome contents, such as protein, and microRNAs, from SCs to axons [103]. Also, multiple studies discussed how axonal regeneration can be promoted by EXOs, and most of these studies demonstrated that derived EXOs from various cell sources can promote axonal regeneration by impinging directly on the phosphatase and tensin homolog (PTEN), the mechanistic target of rapamycin (mTOR) signaling pathway [104, 105]. The PTEN-mTOR pathway is a key factor in axonal regeneration. Accordingly, EXOs have neurophysiological activities that can promote neurite outgrowth [104].

A previous study has demonstrated that EXOs derived from SCs can be internalized by axons and also increase axonal regeneration in in vitro and in vivo studies [32]. Mechanically, EXOs have the potential to change growth cone morphology to a pro-regenerative phenotype and can also decrease the activity of the GTPase RhoA, which is involved in axon retraction and growth cone collapse [32, 106]. Another main factor in EXOs that facilitate axon regeneration is microRNAs like miR-21. A study showed that expression of miR-21 can promote dorsal root ganglion (DRG) axon regrowth [107]. Also, miRNA can facilitate axon regeneration in peripheral nerves due to the knockout of Dicer, a key activator of the RNA-induced silencing complex (RISC) [75, 108].

The results of studies suggest that EXOs can be sent to recipient neurons as effective miRNAs to control the growth of axons. Furthermore, in another study, Buccan et al. [83] demonstrated the effect of MSC-derived EXOs on neurite outgrowth. Their results have shown that cultured dorsal root ganglia (DRG) neurons with MSC-derived EXOs increased the neurite outgrowth of the DRG neurons after co-culturing with EXOs due to the presence of growth factors like BDNF, FGF-1, GDNF, IGF-1, and NGF in the MSC-derived EXOs.

A recent study by Shariati et al. [84] demonstrated that human adipose stem cell (human ADSC)-derived exosomes penetrate into the target cells and increase viability, cell growth, and induce neural differentiation of PC12 cell lines due to the existence of growth factors, proteins, and miRNAs. Overall, evidence has demonstrated that MSC-derived EXOs have the potential to promote axonal regeneration through three main mechanisms (Fig. 3):

  1. 1.

    Transport of miRNAs to induce axonal outgrowth

  2. 2.

    Shuttle neurotrophic growth factors facilitate axon regeneration.

  3. 3.

    Impinging directly on the PTEN-mTOR pathway

Fig. 3
figure 3

Diagram illustrates how MSC-derived exosomes can regulate expression of miRNA and activate PI3K/Akt, which induce activation of the PI3/AKT signaling pathway in neural cells, leading to promotion of nerve regeneration

Neuroinflammation

Neuroinflammation is one of the key factors in recovery from PNI. When PNI has occurred, myelinating SCs are dedifferentiated and activated in the distal stump of the nerve. Dedifferentiated SCs begin to clear cell debris and residual injured myelin in a Wallerian degeneration (WD) process. On the other hand, the differentiated SCs release several chemokines and proinflammatory cytokines that lead to a neuroinflammatory response [109, 110]. The neuroinflammatory response leads to the accumulation of peripheral immune cells, like circulating macrophages, at the injury site [111]. Circulating macrophages are essential for the regeneration of axons due to the clearance of axonal and myelin debris because degenerated axon debris inhibits axonal growth in the later stages of WD [112, 113].

Despite the fact that neuroinflammation has a double-edged sword effect, neuroinflammation has some positive effects in the process of regeneration from nerve injury, but excessive inflammatory responses can not only be obstacles for nerve regeneration but also cause neuropathic pain. Hence, an appropriate level of neuroinflammation is the main target for PNI.

Multiple studies have reported that the main immunosuppressive effects of MSCs are related to the immunoregulatory properties of their secretome, such as EV exosomes [114, 115]. MSC-derived EXOs demonstrate their immunomodulatory effect via miRNAs. For example, a study demonstrated that miR‐21 in MSC-derived EXOs can modulate immunoreactions by diminishing signal transducers and activators of transcription 3 (STAT3) expression and inhibiting the nuclear factor kappa β (NF‐κβ) pathway [116]. Furthermore, miR‐181c, which is found in MSC-derived EXOs, plays a key role in reducing inflammation by reducing NF‐κβ activation and repressing the toll-like receptor 4 (TLR4) signaling pathway [117]. Another miRNA that exists in MSC-derived EXOs is miR‐21‐5p. It has been demonstrated that miR‐21‐5p diminished proinflammatory cytokines and increased M1 to M2 polarization in alveolar macrophages by inhibition of iNOS mRNA expression (Fig. 4) [118]. In the same way, miR‐326, miR‐182, miR‐17‐ 5p, miR‐140‐5p, miR‐9, and miR‐let7 that are found in MSC-derived EXOs can also reduce inflammation by suppressing proinflammatory cytokines [119].

Fig. 4
figure 4

MSC-derived exosomes exhibit immunomodulatory and anti-inflammatory effects, which decrease nerve tissue damage. MSC-derived exosomes have an immunomodulatory effect due to the interaction of exosomal miRNAs. MSC-derived exosomes are able to transform macrophages from the M0 and M1 phenotypes to the M2 phenotype. Also, they increase secretion of M2-related cytokines such as TGF-β and IL-10 and also decrease M1-related cytokine levels (IL-6, IL-12, and TNF-α)

Indeed, MSC-derived exosomes with specific miRNAs can induce the polarization of macrophages from the M1 to the M2 phenotype to promote nerve regeneration. In a nerve tissue injury situation, macrophages differentiate into M1 macrophages that can promote an inflammation response and also aggravate tissue damage. But MSC-derived exosomes induce the polarization of macrophages from the M1 to the M2 phenotype to promote nerve regeneration due to their specific miRNAs (Fig. 4). Collectively, we can conclude that MSC-derived EXOs modulate neuroinflammation to promote axonal outgrowth.

Exosomes ameliorate neuropathic pain

Neuropathic pain is a type of chronic pain that occurs as a result of a lesion or disease in the peripheral nervous system [120, 121]. Although the exact mechanisms of neuropathic pain as a chronic pain are poorly determined, studies have reported that neuropathic pain is mostly related to neuroinflammation [122]. Up until now, clinical treatment strategies for neuropathic pain have included physical, pharmacological, and interventional approaches, but none of them appear to be effective in controlling the condition [123]. Therefore, an effective clinical strategy is necessary for the treatment of neuropathic pain. Nowadays, exosome therapies represent a potential candidate for clinical neuropathic pain treatment due to their anti-neuroinflammation effects [124]. Clinical studies have suggested that EXOs can modulate immune responses and promote the healing process, thereby potentially alleviating inflammation and pain [125]. Indeed, EXOs can suppress the production of proinflammatory cytokines such as TNF-α,IL-1β, and PGE2 in tissue-injured areas and also stimulate the release of IL-10, leading to antinociceptive effects [126]. Exosomal miRNAs like miR-181c-5p, miR-216a-5p, and miR-126-3p have also been demonstrated to ameliorate neuropathic pain in sciatic nerve compression in in vivo studies [123, 127]. EXOs can ameliorate neuropathic pain by reducing proinflammatory cytokines and promoting neuronal proliferation and function [128].

Exosomes as nanocarriers

Exosomes are lipid bilayer-enclosed vesicles that originate from the internal budding of the late endosomal membrane and are secreted by all types of cells. Exosomes are therefore a natural cargo for cell–cell communication. This property attracted the attention of researchers to exosome-based gene or drug delivery systems. Exosome-based delivery systems have several advantages over other delivery systems due to the following reasons:

  1. 1.

    A variety of biological cargoes can be delivered by exosomes, like drugs, small RNAs, mRNAs, and proteins.

  2. 2.

    Natural capacity of exosomes to cross biological barriers, like the blood–brain barrier.

  3. 3.

    Exosomes can transfer into other tissues with no blood supply.

  4. 4.

    Exosomes can influence targeted tissue for a long period of time.

  5. 5.

    Exosomes are biocompatible and genetically engineered.

  6. 6.

    To avoid systemic toxicity, exosomes can be engineered as surface proteins to distinguish specific targeted tissues and avoid unwanted accumulation in surrounding tissues [129].

Exosomes transport genes or drugs by fusing with the cell membrane of the receptor in acidic environments. Indeed, studies have demonstrated that exosomes have the tetraspanin protein CD9 on their surfaces [130, 131], which can fuse with the target cell membrane to transport a specific gene or drug for therapy. Exosomes have thus been used as an effective carrier for in vivo drug or nucleic acid delivery by researchers. But several challenges have existed regarding the clinical use of exosomes as carriers, such as methods to transport drugs or nucleic acids into exosomes efficiently. Various methods have been applied to import proteins, drugs, and nucleic acids into exosomes, such as incubation, freeze–thaw cycles, sonication, extrusion, electroporation, thermal shock transfection, saponin-assisted loading, hypotonic dialysis, and the pH gradient method [132].

Although many clinical trials are being conducted to examine the therapeutic effect of exosomes in a range of clinical settings such as SARS-CoV-2 pneumonia, acute ischemic stroke, macular holes, and cerebrovascular disorders [133], only a few preclinical studies in peripheral nerve regeneration have been conducted. Yang et al. [134] showed that the implementation of nerve guidance conduits containing NT-3 mRNA-loaded ADSC-derived exosomes significantly improved gastrocnemius muscle function in a rat sciatic nerve defect model. Neurotrophin-3 (NT-3) concentration, a prominent neurotrophic factor in peripheral nerve regeneration, is not sufficient in the early phase of nerve injury. Therefore, in this study, ADSCs were transduced by NT-3 mRNA, and exosomes extracted from these cells were embedded in alginate hydrogel to build a nerve guidance conduit. Sustained release of NT-3 mRNA containing exosomes was detected at least after 2 weeks of NGC [134]. Notable nerve regeneration and functional recovery of the gastrocnemius in treated rats were observed.

Fan et al. [135] found that miR-146a-loaded MSC-derived exosomes helped improve diabetic peripheral neuropathy (DPN). Exosome injection increased both mechanical and thermal stimulus thresholds while decreasing nerve conduction velocity. miR-146a, an anti-inflammatory factor whose expression mediates dorsal root ganglion survival in DPN, was transfected into mouse bone marrow-derived mesenchymal stem cells (BMMSCs). Exosomes derived from transfected MSCs were injected once a week for four weeks to improve conduction velocity and thermal and mechanical stimulus threshold in treated groups of mice.

Singh et al. [136] fused BMMSC-derived exosomes with polypyrrole NPs (PpyNPs)-loaded liposomes via a 10-time freeze–thaw process. This hybrid could provide both chemical and electrical cues for nerve regeneration, as exosomes contain chemical ingredients and PpyNPs induce electrical conduction for nerve regeneration. Intramuscular injection of the hybrid notably normalized the compound muscle action potential and the nerve conduction velocity in DPN rats. Surprisingly, hyperglycemia and weight loss have been controlled in the treated group as a result of the paracrine effect of the hybrid injection.

Liu et al. [137] demonstrated the crosstalk between SCs and neurons in the peripheral nerve. The researchers discovered that miR-21 levels are lower in SC-derived exosomes from diabetic peripheral neuropathy rats, which may impair their ability to induce nerve regeneration. Exosomes derived from transduced SCs with miR-21-lentiviral vectors improved the cells’ ability for neurite growth induction in a high glucose condition compared with exosomes derived from glucose-exposed SCs. It has also been demonstrated that miR-21 exerts its effect partly through p-AKT signaling. The promising results encourage more investigation for exosome therapy to move forward.

Advantages of exosomes as nanocarriers in comparison to other synthetic vesicles

Since the 1990s, more than 50 man-made nanoparticles have been approved for use in clinical settings. All of these nanoparticles are simple two-layer lipids with a few extra ingredients [138]. Even though it seems unlikely that more complex nanocarriers with more of a natural biological structure and cargo could be made on a large scale, exosomes could be used to test the potential of these kinds of multifunctional drug carriers.

Besides, applying exosomes is likely to facilitate issues associated with drug loading and delivery, which substantially reduce the efficiency of nanoparticle production. Both loading exosomes with the cargo of interest and surface modification can be obtained through natural cellular processes in exosome production as well as cargo delivery via endocytosis/membrane fusion [139]. Naturally loaded exosomes can be produced through genetically engineered cells of origin to produce the desired molecules inside and/or on the cell membrane, bypassing cargo degradation and receptor implantation during nanoparticle synthesis [138].

Although nanomedicines mostly evoke fewer side effects in comparison to free drugs due to their less frequent encounter with non-targeted tissues, some polyethylene glycol (PEG)-conjugated nanoparticles can trigger inflammatory and rarely life-threatening reactions. Early-phase clinical trials have proven allogenic MSC-derived EVs safe [140, 141]. Similarly, there is a growing body of evidence on the safety of blood-derived EVs in blood transfusions that could be used to predict the safety profile of allogenic exosomes [138, 142]. However, due to the enormously heterogeneous cargo of exosomes, which may affect off-target results, caution should be exercised in generalizing this data [143].

The mechanical stiffness of exosomes is another advantage, as shown in a study using extracellular matrix-simulating hydrogels. EVs are superior to nanocarriers in both their tolerance of a stress-relaxing environment and their ability to cross biological barriers like the blood–brain barrier due to their surface proteins [133, 144]. Therefore, the immunocompatibility and organ-organelle tropism of exosomes may serve them as more efficient therapeutic agents.

Challenges of exosome therapy

Although the favorable results from exosome therapy in preclinical and clinical studies are encouraging, as an unprecedented therapeutic approach, there are some issues that need to be tackled before applying them in clinical settings. Firstly, the lack of GMP-compliant large-scale production techniques has hindered the transition of exosome therapy from preclinical to clinical studies. A variety of methods to propagate cell sources, from 3-D cell culture to bioreactors, have been applied, yet they demand more improvements to efficiently meet the need for a clinical dose of exosomes [145].

Secondly, in the absence of effective isolation techniques, exosomes are precipitated with other undesirable molecular contaminants, which impede the clinical translation of the therapeutic agents. Differential centrifugation followed by ultracentrifugation has been more frequently applied in preclinical studies [146] than the other currently available methods, including size-exclusion chromatography (SEC) density, gradient ultracentrifugation, precipitation, immunoaffinity-based capture [147], and microfluidics-based technologies. However, while the technique has proven effective, coprecipitation of non-exosome molecules, low efficiency, and impaired exosome structure remain issues to be settled [148].

Thirdly, by carrying a diverse range of molecules with synergic or additive effects, EVs may intensify the therapeutic effects, yet some concerns remain due to their potential oncogenic activity, particularly for those driven by stem cell proteins [149, 150]. In other words, the inability to precisely both characterize and quantify exosomes’ cargo as well as the inability to target them for a specific receptor raises off-target concerns.

Fourthly, the lack of techniques for precise quantification of exosomes poses pharmacokinetic challenges. Available methods of exosome measurement are highly sensitive and accurate and include nanoparticle tracking analysis, electron microscopy, surface plasmon resonance, flow cytometry, dynamic light scattering, tunable resistive pulse sensing, and single-particle reflectance imaging sensors. However, these techniques are quite costly and laborious, particularly when it comes to large-scale production.

Fifthly, optical imaging demonstrated exosomes’ rapid accumulation in the liver and spleen following intravenous injection, which represents their undesirable biodistribution and short half-life. The considerably lower half-life of exosomes than nanocarriers (60 min vs. several hours) is another shortcoming that needs to be overcome [149]. Therefore, a range of issues should be avoided, from the manufacturing of exosomes to accurately characterizing and quantifying them, in order to guarantee a safe and effective therapy with known possible side effects. Hence, the International Society for extracellular vesicles (ISEV), established in 2014, updated isolation and characterization methods in 2018 to further accurate and reliable EV isolation [151]. Furthermore, the EV-TRACK platform, developed in 2017, encourages authors to share their isolation and characterization techniques and receive advice on possible drawbacks to enhance more reproducible and concrete results [133].

Future perspectives and conclusions

When a peripheral nerve injury occurs, a series of complex events occur in the neuron’s cell body and in surrounding cells. Several factors are involved in the nerve regeneration process, such as inflammation, trophic factors, angiogenesis, and SCs. Due to the requirement to sacrifice a healthy tissue nerve, current Schwann cell-based therapy to regenerate peripheral nerves is not an ideal approach. Furthermore, the use of autologous SC exosomes to treat PNI does not overcome the obstacle of needing to sacrifice a functioning nerve to gain the SC-derived exosomes. On the other hand, MSCs have been shown to be efficacious in improving neurite outgrowth, and they are applied in PNI studies. MSC transplantation using nerve guide conduits has shown positive effects in animal models of nerve gap injuries but is still far from being widely accepted.

Recently, MSC-derived exosomes have been known as the main regulatory mediator that mediates tissue regeneration. MSC-derived exosomes have a therapeutic effect similar to MSCs and, due to several advantages over MSCs, can be used as a cell-free therapy to treat peripheral nerve injury instead of MSCs. MSC-derived exosomes play a pivotal role in mediating intercellular communication in the peripheral nerve microenvironment. Indeed, MSC-derived exosomes transfer genetic substrates such as miRNAs, neurotrophic factors, and proteins to axons to regulate axonal regrowth, as described in this study. Moreover, MSC exosome-based therapy can resolve the issues caused by stem cell transplantation. Hence, in the future, MSC exosome-based therapy will be a cell-free approach for regenerating PNI. Although several studies have shown that injecting MSC-derived exosomes into nerve stumps or supplementing nerve conduits for the treatment of peripheral nerve injury is effective and safe, further research is needed to determine the potential of MSC-derived exosomes for clinical application.

Availability of data and materials

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Abbreviations

PNI:

Peripheral nerve injury

SCs:

Schwann cell

MSCs:

Mesenchymal stem cells

NGCs:

Nerve guide conduits

GFs:

Growth factors

MVs:

Microvesicles

EXOs:

Exosomes

EV:

Extracellular vesicles

WD:

Wallerian degeneration

PTEN:

Phosphatase and tensin homolog

IGF-1:

Insulin-like growth factor-1

Tregs:

Regulatory T cells

mTOR:

Mechanistic target of rapamycin

DRG:

Dorsal root ganglion

RISC:

RNA-induced silencing complex

NT-3:

Neurotrophin-3

PpyNPs:

Polypyrrole NPs

DPN:

Diabetic peripheral neuropathy

SEC:

Size-exclusion chromatography

ISEV:

International Society for Extracellular Vesicles

References

  1. Lim E-MF, Nakanishi ST, Hoghooghi V, Eaton SE, Palmer AL, Frederick A, Stratton JA, Stykel MG, Whelan PJ, Zochodne DW. AlphaB-crystallin regulates remyelination after peripheral nerve injury. Proc Natl Acd Sci. 2017;114(9):E1707–16.

    CAS  Google Scholar 

  2. Panagopoulos GN, Megaloikonomos PD, Mavrogenis AF. The present and future for peripheral nerve regeneration. Orthopedics. 2017;40(1):e141–56.

    PubMed  Google Scholar 

  3. Zack-Williams SD, Butler PE, Kalaskar DM. Current progress in use of adipose derived stem cells in peripheral nerve regeneration. World J Stem Cells. 2015;7(1):51.

    PubMed  PubMed Central  Google Scholar 

  4. Kaiser R, Ullas G, Havránek P, Homolková H, Miletín J, Tichá P, Sukop A. Current concepts in peripheral nerve injury repair. Acta Chir Plast. 2017;59(2):85–91.

    CAS  PubMed  Google Scholar 

  5. Zhang H, Shao Z, Zhu Y, Shi L, Li Z, Hou R, Zhang C, Yao D. Toll-like receptor 4 (TLR4) expression affects Schwann cell behavior in vitro. Sci Rep. 2018. https://doi.org/10.1038/s41598-018-28516-5.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Caillaud M, Richard L, Vallat JM, Desmoulière A, Billet F. Peripheral nerve regeneration and intraneural revascularization. Neural Regener Res. 2019;14(1):24.

    CAS  Google Scholar 

  7. Lee SK, Wolfe SW. Peripheral nerve injury and repair. JAAOS-J Am Acad Orthop Surg. 2000;8(4):243–52.

    CAS  Google Scholar 

  8. Arslantunali D, Dursun T, Yucel D, Hasirci N, Hasirci VJ. Peripheral nerve conduits: technology update. Med Dev Evid Res. 2014;7:405.

    CAS  Google Scholar 

  9. Li R, Liu Z, Pan Y, Chen L, Zhang Z, Lu L. Peripheral nerve injuries treatment: a systematic review. Cell Biochem Biophys. 2014;68(3):449–54.

    CAS  PubMed  Google Scholar 

  10. Ray WZ, Mackinnon SE. Management of nerve gaps: autografts, allografts, nerve transfers, and end-to-side neurorrhaphy. Exp Neurol. 2010;223(1):77.

    PubMed  Google Scholar 

  11. Singh A, Asikainen S, Teotia AK, Shiekh PA, Huotilainen E, Qayoom I, Partanen J, Seppälä J, Kumar A. Biomimetic photocurable three-dimensional printed nerve guidance channels with aligned cryomatrix lumen for peripheral nerve regeneration. ACS Appl Mater Interfaces. 2018;10(50):43327–42.

    CAS  PubMed  Google Scholar 

  12. Lu S, Yang J, Yang B, Yin Z, Liu M, Yin L, Zheng W. Analysis and Design of Surgical Instrument Localization Algorithm. Comput Model Eng Sci. 2023;137(1):669.

    Google Scholar 

  13. Daly W, Yao L, Zeugolis D, Windebank A, Pandit A. A biomaterials approach to peripheral nerve regeneration: bridging the peripheral nerve gap and enhancing functional recovery. J R Soc Interface. 2012;9(67):202–21.

    CAS  PubMed  Google Scholar 

  14. Jarrige M, Frank E, Herardot E, Martineau S, Darle A, Benabides M, Domingues S, Chose O, Habeler W, Lorant J, Baldeschi C. The future of regenerative medicine: cell therapy using pluripotent stem cells and acellular therapies based on extracellular vesicles. Cells. 2021;10(2):240.

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Marks PW, Witten CM, Califf RM. Clarifying stem-cell therapy’s benefits and risks. N Engl J Med. 2017;376(11):1007–9.

    PubMed  Google Scholar 

  16. Quigley AF, Bulluss KJ, Kyratzis IL, Gilmore K, Mysore T, Schirmer KS, Kennedy EL, O’Shea M, Truong YB, Edwards SL, Peeters G. Engineering a multimodal nerve conduit for repair of injured peripheral nerve. J Neural Eng. 2013;10(1):016008.

    CAS  PubMed  Google Scholar 

  17. Magaz A, Faroni A, Gough JE, Reid AJ, Li X, Blaker JJ. Bioactive silk-based nerve guidance conduits for augmenting peripheral nerve repair. Adv Healthcare Mater. 2018;7(23):1800308.

    Google Scholar 

  18. Johnstone RM. Exosomes biological significance: a concise review. Blood Cells Mol Dis. 2006;36(2):315–21.

    CAS  PubMed  Google Scholar 

  19. Nawaz M, Fatima F, Vallabhaneni KC, Penfornis P, Valadi H, Ekström K, Kholia S, Whitt JD, Fernandes JD, Pochampally R, Squire JA. Extracellular vesicles: evolving factors in stem cell biology. Stem Cells Int. 2016;2016:1.

    Google Scholar 

  20. Dong R, Liu Y, Yang Y, Wang H, Xu Y, Zhang Z. MSC-derived exosomes-based therapy for peripheral nerve injury: a novel therapeutic strategy. BioMed Res Int. 2019;2019:1.

    Google Scholar 

  21. Li P, Kaslan M, Lee SH, Yao J, Gao Z. Progress in exosome isolation techniques. Theranostics. 2017;7(3):789.

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Blanc L, Vidal M. New insights into the function of Rab GTPases in the context of exosomal secretion. Small GTPases. 2018;9(1–2):95–106.

    CAS  PubMed  Google Scholar 

  23. Tian Y, Xiao H, Yang Y, Zhang P, Yuan J, Zhang W, Chen L, Fan Y, Zhang J, Cheng H, Deng T, Yang L, Wang W, Chen G, Wang P, Gong P, Niu X, Zhang X. Crosstalk between 5-methylcytosine and N(6)-methyladenosine machinery defines disease progression, therapeutic response and pharmacogenomic landscape in hepatocellular carcinoma. Mol Cancer. 2023;22(1):5.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Jessen KR, Mirsky R. The origin and development of glial cells in peripheral nerves. Nat Rev Neurosci. 2005;6(9):671–82.

    CAS  PubMed  Google Scholar 

  25. Kidd GJ, Ohno N, Trapp BD. Biology of Schwann cells. In: Peripheral nerve disorders. Elsevier; 2013. p. 55–79.

    Google Scholar 

  26. Monje PV, Soto J, Bacallao K, Wood PM. Schwann cell dedifferentiation is independent of mitogenic signaling and uncoupled to proliferation: role of cAMP and JNK in the maintenance of the differentiated state. J Biol Chem. 2010;285(40):31024–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Rotshenker S. Wallerian degeneration: the innate-immune response to traumatic nerve injury. J Neuroinflamm. 2011;8(1):1–14.

    Google Scholar 

  28. Xiahui Z, Weixiong P, Shenglian P, Zhe L, Rongrong P, Qinglai W. High glucose suppresses osteogenic differentiation and induces mitochondrial dysfunction in osteoblasts via Sirt1/Recql4 axis; a laboratory study using mouse cells. J Biol Regul Homeost Agents. 2022;36(4):889–99.

    Google Scholar 

  29. Arthur-Farraj PJ, Morgan CC, Adamowicz M, Gomez-Sanchez JA, Fazal SV, Beucher A, Razzaghi B, Mirsky R, Jessen KR, Aitman TJ. Changes in the coding and non-coding transcriptome and DNA methylome that define the Schwann cell repair phenotype after nerve injury. Cell Rep. 2017;20(11):2719–34.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Chang LW, Viader A, Varghese N, Payton JE, Milbrandt J, Nagarajan R. An integrated approach to characterize transcription factor and microRNA regulatory networks involved in Schwann cell response to peripheral nerve injury. BMC Genom. 2013;14(1):1–20.

    CAS  Google Scholar 

  31. Yi S, Yuan Y, Chen Q, Wang X, Gong L, Liu J, Gu X, Li S. Regulation of Schwann cell proliferation and migration by miR-1 targeting brain-derived neurotrophic factor after peripheral nerve injury. Sci Rep. 2016;6(1):1–10.

    Google Scholar 

  32. Lopez-Verrilli MA, Picou F, Court FA. Schwann cell-derived exosomes enhance axonal regeneration in the peripheral nervous system. Glia. 2013;61(11):1795–806.

    PubMed  Google Scholar 

  33. Kalani A, Tyagi A, Tyagi N. Exosomes: mediators of neurodegeneration, neuroprotection and therapeutics. Mol Neurobiol. 2014;49(1):590–600.

    CAS  PubMed  Google Scholar 

  34. Lai CP, Breakefield XO. Role of exosomes/microvesicles in the nervous system and use in emerging therapies. Front Physiol. 2012;3:228.

    CAS  PubMed  PubMed Central  Google Scholar 

  35. López-Leal R, Díaz-Viraqué F, Catalán RJ, Saquel C, Enright A, Iraola G, Court FA. Schwann cell reprogramming into repair cells increases miRNA-21 expression in exosomes promoting axonal growth. J Cell Sci. 2020;133(12):jcs239004.

    PubMed  Google Scholar 

  36. Gregorio CD, Díaz P, López-Leal R, Manque P. Purification of exosomes from primary Schwann cells, RNA extraction, and next-generation sequencing of exosomal RNAs. In: Schwann cells. Springer; 2018. p. 299–315.

    Google Scholar 

  37. Gurvich CT, Fitzgerald PB, Georgiou-Karistianis N, White OB. Saccadic impairment in schizophrenia with prominent negative symptoms. Neuroreport. 2008;19(14):1435–9.

    PubMed  Google Scholar 

  38. Deborde S, Omelchenko T, Lyubchik A, Zhou Y, He S, McNamara WF, Chernichenko N, Lee SY, Barajas F, Chen CH, Bakst RL. Schwann cells induce cancer cell dispersion and invasion. J Clin Investig. 2016;126(4):1538–54.

    PubMed  PubMed Central  Google Scholar 

  39. Shurin GV, Kruglov O, Ding F, Lin Y, Hao X, Keskinov AA, You Z, Lokshin AE, LaFramboise WA, Falo LD Jr, Shurin MR. Melanoma-induced reprogramming of schwann cell signaling aids tumor growth. Cancer Res. 2019;79(10):2736–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Lou J-S, Zhao L-P, Huang Z-H, Chen X-Y, Xu J-T, Tai WC-S, Tsim KWK, Chen Y-T, Xie T. Ginkgetin derived from Ginkgo biloba leaves enhances the therapeutic effect of cisplatin via ferroptosis-mediated disruption of the Nrf2/HO-1 axis in EGFR wild-type non-small-cell lung cancer. Phytomedicine. 2021;80:153370.

    CAS  PubMed  Google Scholar 

  41. Piwei H, Minghui W, Shufan J, Maolin H. c-Myc positively enhances rpl34 expression to regulate osteosarcoma cell proliferation. J Biol Regul Homeost Agents. 2022;36(5):1291–8.

    Google Scholar 

  42. Qing L, Chen H, Tang J, Jia X. Exosomes and their MicroRNA cargo: new players in peripheral nerve regeneration. Neurorehabil Neural Repair. 2018;32(9):765–76.

    PubMed  PubMed Central  Google Scholar 

  43. Cong M, Shen M, Wu X, Li Y, Wang L, He Q, Shi H, Ding F. Improvement of sensory neuron growth and survival via negatively regulating PTEN by miR-21-5p-contained small extracellular vesicles from skin precursor-derived Schwann cells. Stem Cell Res Therapy. 2021;12(1):1–15.

    Google Scholar 

  44. Wu X, Wang L, Cong M, Shen M, He Q, Ding F, Shi H. Extracellular vesicles from skin precursor-derived Schwann cells promote axonal outgrowth and regeneration of motoneurons via Akt/mTOR/p70S6K pathway. Ann Transl Med. 2020;8(24):1640.

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Sohn EJ, Park HT, Shin YK. Exosomes derived from differentiated Schwann cells inhibit Schwann cell migration via microRNAs. Neuroreport. 2020;31(7):515–22.

    CAS  PubMed  Google Scholar 

  46. Hyung S, Kim J, Yu C, Jung H, Hong J. Neuroprotective effect of glial cell-derived exosomes on neurons. Immunotherapy. 2019;5:156.

    Google Scholar 

  47. Yin K, Wang S, Zhao RC. Exosomes from mesenchymal stem/stromal cells: a new therapeutic paradigm. Biomarker Res. 2019;7(1):1–8.

    CAS  Google Scholar 

  48. Lou G, Chen Z, Zheng M, Liu Y. Mesenchymal stem cell-derived exosomes as a new therapeutic strategy for liver diseases. Exp Mol Med. 2017;49(6):e346–e346.

    PubMed  PubMed Central  Google Scholar 

  49. Lianbo Shao Y, Zhang BL, Wang J, Zhang Z, Zhang L, Xiao P, Meng Q, Geng Y, Xi-yong Y, Li Y. MiRNA-sequence indicates that mesenchymal stem cells and exosomes have similar mechanism to enhance cardiac repair. BioMed Res Int. 2017;2017:1–9. https://doi.org/10.1155/2017/4150705.

    Article  CAS  Google Scholar 

  50. Akyurekli C, Le Y, Richardson RB, Fergusson D, Tay J, Allan DS. A systematic review of preclinical studies on the therapeutic potential of mesenchymal stromal cell-derived microvesicles. Stem Cell Rev Rep. 2015;11(1):150–60.

    CAS  PubMed  Google Scholar 

  51. Wen D, Peng Y, Liu D, Weizmann Y, Mahato RI. Mesenchymal stem cell and derived exosome as small RNA carrier and Immunomodulator to improve islet transplantation. J Controll Releas. 2016;238:166–75.

    CAS  Google Scholar 

  52. Sarvar DP, Shamsasenjan K, Akbarzadehlaleh P. Mesenchymal stem cell-derived exosomes: new opportunity in cell-free therapy. Adv Pharmaceutical Bull. 2016;6(3):293.

    CAS  Google Scholar 

  53. Shabbir A, Cox A, Rodriguez-Menocal L, Salgado M, Badiavas EV. Mesenchymal stem cell exosomes induce proliferation and migration of normal and chronic wound fibroblasts, and enhance angiogenesis in vitro. Stem Cells Dev. 2015;24(14):1635–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Wang X, Jiao Y, Pan Y, Zhang L, Gong H, Qi Y, Wang M, Gong H, Shao M, Wang X, Jiang D. Fetal dermal mesenchymal stem cell-derived exosomes accelerate cutaneous wound healing by activating notch signaling. Stem Cells Int. 2019;2019:1–11. https://doi.org/10.1155/2019/2402916.

    Article  CAS  Google Scholar 

  55. Tianyi Y, Gao M, Yang P, Liu D, Wang D, Song F, Zhang X, Liu Y. Insulin promotes macrophage phenotype transition through PI3K/Akt and PPAR‐γ signaling during diabetic wound healing. J Cell Physiol. 2019;234(4):4217–31. https://doi.org/10.1002/jcp.27185.

    Article  CAS  Google Scholar 

  56. Zhao H, Ming T, Tang S, Ren S, Yang H, Liu M, Tao Q, Xu H. Wnt signaling in colorectal cancer: pathogenic role and therapeutic target. Mol Cancer. 2022;21(1):144.

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Ji-Feng X, Yang G, Pan XH, Zhang SJ, Zhao C, Qiu BS, Hai-Feng G, Hong JF, Li Cao Y, Chen BX, Bi Q, Wang YP. Altered microRNA expression profile in exosomes during osteogenic differentiation of human bone marrow-derived mesenchymal stem cells. PLoS ONE. 2014;9(12):e114627. https://doi.org/10.1371/journal.pone.0114627.

    Article  CAS  Google Scholar 

  58. Xin H, Li Y, Chopp M. Exosomes/miRNAs as mediating cell-based therapy of stroke. Front Cell Neurosci. 2014. https://doi.org/10.3389/fncel.2014.00377.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Zhu Y, Huang R, Wu Z, Song S, Cheng L, Zhu R. Deep learning-based predictive identification of neural stem cell differentiation. Nat Commun. 2021;12(1):2614.

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Xin H, Wang F, Li Y, Qing-E L, Cheung WL, Zhang Y, Zhang ZG, Chopp M. Secondary release of exosomes from astrocytes contributes to the increase in neural plasticity and improvement of functional recovery after stroke in rats treated with exosomes harvested from microRNA 133b-overexpressing multipotent mesenchymal stromal cells. Cell Transpl. 2017;26(2):243–57. https://doi.org/10.3727/096368916X693031.

    Article  Google Scholar 

  61. Xin H, Katakowski M, Wang F, Qian JY, Liu XS, Ali MM, Buller B, Zhang ZG, Chopp M. MicroRNA-17–92 cluster in exosomes enhance neuroplasticity and functional recovery after stroke in rats. Stroke. 2017;48(3):747–53. https://doi.org/10.1161/STROKEAHA.116.015204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhang Y, Chopp M, Liu XS, Katakowski M, Wang X, Tian X, David W, Zhang ZG. Exosomes derived from mesenchymal stromal cells promote axonal growth of cortical neurons. Mol Neurobiol. 2017;54(4):2659–73. https://doi.org/10.1007/s12035-016-9851-0.

    Article  CAS  PubMed  Google Scholar 

  63. Prathipati P, Nandi SS, Mishra PK. Stem cell-derived exosomes, autophagy, extracellular matrix turnover, and miRNAs in cardiac regeneration during stem cell therapy. Stem Cell Rev Rep. 2017;13(1):79–91. https://doi.org/10.1007/s12015-016-9696-y.

    Article  CAS  PubMed  Google Scholar 

  64. Nooshabadi VT, Verdi J, Ebrahimi-Barough S, Mowla J, Atlasi MA, Mazoochi T, Valipour E, Shafiei S, Ai J, Banafshe HR. Endometrial mesenchymal stem cell-derived exosome promote endothelial cell angiogenesis in a dose dependent manner: a new perspective on regenerative medicine and cell-free therapy. Arch Neurosci. 2019. https://doi.org/10.5812/ans.94041.

    Article  Google Scholar 

  65. Lai RC, Tan SS, Teh BJ, Sze SK, Arslan F, de Kleijn DP, Choo A, Lim SK. Proteolytic potential of the MSC exosome proteome: implications for an exosome-mediated delivery of therapeutic proteasome. Int J Proteom. 2012;2012:1–14. https://doi.org/10.1155/2012/971907.

    Article  CAS  Google Scholar 

  66. Zhou L, Liu Y, Sun H, Li H, Zhang Z, Hao P. Usefulness of enzyme-free and enzyme-resistant detection of complement component 5 to evaluate acute myocardial infarction. Sens Actuators B Chem. 2022;369:132315.

    CAS  Google Scholar 

  67. Zhang B, Yin Y, Lai RC, Tan SS, Choo ABH, Lim SK. Mesenchymal stem cells secrete immunologically active exosomes. Stem Cells and Dev. 2014;23(11):1233–44. https://doi.org/10.1089/scd.2013.0479.

    Article  CAS  Google Scholar 

  68. Mendt M, Rezvani K, Shpall E. Mesenchymal stem cell-derived exosomes for clinical use. Bone Marrow Transpl. 2019;54(S2):789–92. https://doi.org/10.1038/s41409-019-0616-z.

    Article  Google Scholar 

  69. Zhang R-C, Wen-Qi D, Zhang J-Y, Shao-Xia Y, Fang-Zhi L, Ding H-M, Cheng Y-B, Ren C, Geng D-Q. Mesenchymal stem cell treatment for peripheral nerve injury: a narrative review. Neural Regener Res. 2021;16(11):2170. https://doi.org/10.4103/1673-5374.310941.

    Article  CAS  Google Scholar 

  70. Lavorato A, Raimondo S, Boido M, Muratori L, Durante G, Cofano F, Vincitorio F, Petrone S, Titolo P, Tartara F, Vercelli A, Garbossa D. Mesenchymal stem cell treatment perspectives in peripheral nerve regeneration: systematic review. Int J Mol Sci. 2021;22(2):572. https://doi.org/10.3390/ijms22020572.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kim HJ, Park JS. Usage of human mesenchymal stem cells in cell-based therapy: advantages and disadvantages. Dev Reprod. 2017;21(1):1.

    PubMed  PubMed Central  Google Scholar 

  72. Mazini L, Rochette L, Admou B, Amal S, Malka G. Hopes and limits of adipose-derived stem cells (ADSCs) and mesenchymal stem cells (MSCs) in wound healing. Int J Mol Sci. 2020;21(4):1306. https://doi.org/10.3390/ijms21041306.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Gnecchi M, Danieli P, Malpasso G, Ciuffreda MC. Paracrine mechanisms of mesenchymal stem cells in tissue repair. In: Gnecchi M, editor. mesenchymal stem cells. Springer New York; 2016. p. 123–46.

    Google Scholar 

  74. Namini MS, Ebrahimi-Barough S, Ai J, Jahromi HK, Mikaeiliagah E, Azami M, Bahrami N, Lotfibakhshaiesh N, Saremi J, Shirian S. Tissue-engineered core-shell silk-fibroin/poly-l-lactic acid nerve guidance conduit containing encapsulated exosomes of human endometrial stem cells promotes peripheral nerve regeneration. ACS Biomater Sci Eng. 2023. https://doi.org/10.1021/acsbiomaterials.3c00157.

    Article  PubMed  Google Scholar 

  75. Mao Q, Nguyen PD, Shanti RM, Shi S, Shakoori P, Zhang Q, Le AD. Gingiva-derived mesenchymal stem cell-extracellular vesicles activate schwann cell repair phenotype and promote nerve regeneration. Tissue Eng Part A. 2019;25(11–12):887–900.

    CAS  PubMed  Google Scholar 

  76. Fan B, Li C, Szalad A, Wang L, Pan W, Zhang R, Chopp M, Zhang ZG, Liu XS. Mesenchymal stromal cell-derived exosomes ameliorate peripheral neuropathy in a mouse model of diabetes. Diabetologia. 2020;63:431–43.

    CAS  PubMed  Google Scholar 

  77. Xu M, Yang Q, Sun X, Wang Y. Recent advancements in the loading and modification of therapeutic exosomes. Fron Bioeng Biotechnol. 2020;8:586130.

    Google Scholar 

  78. Zhang Y, Bi J, Huang J, Tang Y, Du S, Li P. Exosome: a review of its classification, isolation techniques, storage, diagnostic and targeted therapy applications. Int J Nanomed. 2020;15:6917–34.

    CAS  Google Scholar 

  79. Lotfy A, AboQuella NM, Wang H. Mesenchymal stromal/stem cell (MSC)-derived exosomes in clinical trials. Stem Cell Res Therapy. 2023;14(1):1–18.

    Google Scholar 

  80. Rao F, Zhang D, Fang T, Lu C, Wang B, Ding X, Wei S, Zhang Y, Pi W, Xu H, Wang Y. Exosomes from human gingiva-derived mesenchymal stem cells combined with biodegradable chitin conduits promote rat sciatic nerve regeneration. Stem Cells Int. 2019. https://doi.org/10.1155/2019/2546367.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Wei JJ, Chen YF, Xue CL, Ma BT, Shen YM, Guan J, Bao XJ, Wu H, Han Q, Wang RZ, Zhao CH. Protection of nerve injury with exosome extracted from mesenchymal stem cell. Acta Acad Med Sin. 2016;38(1):33–6.

    Google Scholar 

  82. Zhang QZ, Chen C, Chang MB, Shanti RM, Cannady SB, O’Malley BW, Shi S, Le AD. Oral rehabilitation of patients sustaining orofacial injuries: the UPenn initiative. Adv Dental Res. 2019;30(2):50–6.

    CAS  Google Scholar 

  83. Bucan V, Vaslaitis D, Peck CT, Strauß S, Vogt PM, Radtke C. Effect of exosomes from rat adipose-derived mesenchymal stem cells on neurite outgrowth and sciatic nerve regeneration after crush injury. Mol Neurobiol. 2019;56(3):1812–24.

    CAS  PubMed  Google Scholar 

  84. Shariati Najafabadi S, Amirpour N, Amini S, Zare N, Kazemi M, Salehi H. Human adipose derived stem cell exosomes enhance the neural differentiation of PC12 cells. Mol Biol Rep. 2021;48(6):5033–43.

    CAS  PubMed  Google Scholar 

  85. Muangsanit P, Shipley RJ, Phillips JB. Vascularization strategies for peripheral nerve tissue engineering. Anatom Record. 2018;301(10):1657–67.

    Google Scholar 

  86. Cattin AL, Burden JJ, Van Emmenis L, Mackenzie FE, Hoving JJ, Calavia NG, Guo Y, McLaughlin M, Rosenberg LH, Quereda V, Jamecna D. Macrophage-induced blood vessels guide Schwann cell-mediated regeneration of peripheral nerves. Cell. 2015;162(5):1127–39.

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Li C, Lin L, Zhang L, Xu R, Chen X, Ji J, Li Y. Long noncoding RNA p21 enhances autophagy to alleviate endothelial progenitor cells damage and promote endothelial repair in hypertension through SESN2/AMPK/TSC2 pathway. Pharmacol Res. 2021;173:105920.

    CAS  PubMed  Google Scholar 

  88. Lovati AB, D’arrigo D, Odella S, Tos P, Geuna S, Raimondo S. Nerve repair using decellularized nerve grafts in rat models: a review of the literature. Front Cell Neurosci. 2018;12:427.

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Habre SB, Bond G, Jing XL, Kostopoulos E, Wallace RD, Konofaos P. The surgical management of nerve gaps: present and future. Ann Plast Surg. 2018;80(3):252–61.

    Google Scholar 

  90. Namini MS, Ebrahimi-Barough S, Daneshimehr F, Ai J. Interplay between angiogenesis and neurogenesis in nerve regeneration. In: Biomaterials for vasculogenesis and angiogenesis. Elsevier; 2022. p. 111–45.

    Google Scholar 

  91. Zeng Q, Bie B, Guo Q, Yuan Y, Han Q, Han X, Chen M, Zhang X, Yang Y, Liu M, Liu P, Deng H, Zhou X. Hyperpolarized Xe NMR signal advancement by metal-organic framework entrapment in aqueous solution. Proc Natl Acad Sci. 2020;117(30):17558–63.

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Wang Y, Zhai W, Cheng S, Li J, Zhang H. Surface-functionalized design of blood-contacting biomaterials for preventing coagulation and promoting hemostasis. Friction. 2023;11(8):1371–94.

    CAS  Google Scholar 

  93. Zhang B, Wu X, Zhang X, Sun Y, Yan Y, Shi H, Zhu Y, Wu L, Pan Z, Zhu W, Qian H. Human umbilical cord mesenchymal stem cell exosomes enhance angiogenesis through the Wnt4/β-catenin pathway. Stem Cells Transl Med. 2015;4(5):513–22.

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Liu X, Li Q, Niu X, Hu B, Chen S, Song W, Ding J, Zhang C, Wang Y. Exosomes secreted from human-induced pluripotent stem cell-derived mesenchymal stem cells prevent osteonecrosis of the femoral head by promoting angiogenesis. Int J Biol Sci. 2017;13(2):232.

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Gong M, Yu B, Wang J, Wang Y, Liu M, Paul C, Millard RW, Xiao DS, Ashraf M, Xu M. Mesenchymal stem cells release exosomes that transfer miRNAs to endothelial cells and promote angiogenesis. Oncotarget. 2017;8(28):45200.

    PubMed  PubMed Central  Google Scholar 

  96. Xin H, Li Y, Cui Y, Yang JJ, Zhang ZG, Chopp M. Systemic administration of exosomes released from mesenchymal stromal cells promote functional recovery and neurovascular plasticity after stroke in rats. J Cereb Blood Flow Metab. 2013;33(11):1711–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Zhang Y, Chopp M, Meng Y, Katakowski M, Xin H, Mahmood A, Xiong Y. Effect of exosomes derived from multipluripotent mesenchymal stromal cells on functional recovery and neurovascular plasticity in rats after traumatic brain injury. J Neurosurg. 2015;122(4):856–67.

    PubMed  PubMed Central  Google Scholar 

  98. Zhang K, Yang Y, Ge H, Wang J, Lei X, Chen X, Wan F, Feng H, Tan L. Neurogenesis and proliferation of neural stem/progenitor cells conferred by artesunate via FOXO3a/p27Kip1 axis in mouse stroke model. Mol Neurobiol. 2022;59(8):4718–29.

    CAS  PubMed  Google Scholar 

  99. Hu Y, Rao SS, Wang ZX, Cao J, Tan YJ, Luo J, Li HM, Zhang WS, Chen CY, Xie H. Exosomes from human umbilical cord blood accelerate cutaneous wound healing through miR-21-3p-mediated promotion of angiogenesis and fibroblast function. Theranostics. 2018;8(1):169.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  101. Heo JS, Kim S. Human adipose mesenchymal stem cells modulate inflammation and angiogenesis through exosomes. Sci Rep. 2022;12(1):1–11.

    Google Scholar 

  102. Tutuianu R, Rosca AM, Iacomi DM, Simionescu M, Titorencu I. Human mesenchymal stromal cell-derived exosomes promote in vitro wound healing by modulating the biological properties of skin keratinocytes and fibroblasts and stimulating angiogenesis. Int J Mol Sci. 2021;22(12):6239.

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Ching RC, Wiberg M, Kingham PJ. Schwann cell-like differentiated adipose stem cells promote neurite outgrowth via secreted exosomes and RNA transfer. Stem Cell Res Therapy. 2018;9(1):1–12.

    Google Scholar 

  104. Tang BL. Promoting axonal regeneration through exosomes: An update of recent findings on exosomal PTEN and mTOR modifiers. Brain Res Bull. 2018;143:123–31.

    CAS  PubMed  Google Scholar 

  105. Ma ZX, Liu Z, Xiong HH, Zhou ZP, Ouyang LS, Xie FK, Tang YM, Wu ZD, Feng Y. MicroRNAs: protective regulators for neuron growth and development. Neural Regener Res. 2023;18(4):734.

    CAS  Google Scholar 

  106. Lopez-Leal R, Court FA. Schwann cell exosomes mediate neuron–glia communication and enhance axonal regeneration. Cell Mol Neurobiol. 2016;36(3):429–36.

    CAS  PubMed  Google Scholar 

  107. Cong M, Shen M, Wu X, Li Y, Wang L, He Q, Shi H, Ding F. Improvement of sensory neuron growth and survival via negatively regulating PTEN by miR-21-5p-contained small extracellular vesicles from skin precursor-derived Schwann cells. Stem Cell Res Therapy. 2021;12:1–15.

    Google Scholar 

  108. Yu B, Zhou S, Qian T, Wang Y, Ding F, Gu X. Altered microRNA expression following sciatic nerve resection in dorsal root ganglia of rats. Acta Biochim Biophys Sin. 2011;43(11):909–15.

    CAS  PubMed  Google Scholar 

  109. Stoll G, Jander S, Myers RR. Degeneration and regeneration of the peripheral nervous system: from Augustus Waller’s observations to neuroinflammation. J Peripheral Nerv Syst. 2002;7(1):13–27.

    Google Scholar 

  110. Gaudet AD, Popovich PG, Ramer MS. Wallerian degeneration: gaining perspective on inflammatory events after peripheral nerve injury. J Neuroinflamm. 2011;8(1):1–13.

    Google Scholar 

  111. Lu L, Dong J, Liu Y, Qian Y, Zhang G, Zhou W, Zhao A, Ji G, Xu H. New insights into natural products that target the gut microbiota: effects on the prevention and treatment of colorectal cancer. Front Pharmacol. 2022;13:964793.

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Tofaris GK, Patterson PH, Jessen KR, Mirsky R. Denervated Schwann cells attract macrophages by secretion of leukemia inhibitory factor (LIF) and monocyte chemoattractant protein-1 in a process regulated by interleukin-6 and LIF. J Neurosci. 2002;22(15):6696–703.

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Liu Y, Dong T, Chen Y, Sun N, Liu Q, Huang Z, Yang Y, Cheng H, Yue K. Biodegradable and cytocompatible hydrogel coating with antibacterial activity for the prevention of implant-associated infection. ACS Appl Mater Interfaces. 2023;15(9):11507–19.

    CAS  PubMed  Google Scholar 

  114. Bruno S, Deregibus MC, Camussi G. The secretome of mesenchymal stromal cells: role of extracellular vesicles in immunomodulation. Immunol Lett. 2015;168(2):154–8.

    CAS  PubMed  Google Scholar 

  115. Wang Y, Zhai W, Zhang H, Cheng S, Li J. Injectable polyzwitterionic lubricant for complete prevention of cardiac adhesion. Macromol Biosci. 2023;23(4):2200554.

    CAS  Google Scholar 

  116. Das A, Ganesh K, Khanna S, Sen CK, Roy S. Engulfment of apoptotic cells by macrophages: a role of microRNA-21 in the resolution of wound inflammation. J Immunol. 2014;192(3):1120–9.

    CAS  PubMed  Google Scholar 

  117. Li T, Yan Y, Wang B, Qian H, Zhang X, Shen L, Wang M, Zhou Y, Zhu W, Li W, Xu W. Exosomes derived from human umbilical cord mesenchymal stem cells alleviate liver fibrosis. Stem Cells Dev. 2013;22(6):845–54.

    CAS  PubMed  Google Scholar 

  118. Cui GH, Wu J, Mou FF, Xie WH, Wang FB, Wang QL, Fang J, Xu YW, Dong YR. Exosomes derived from hypoxia‐preconditioned mesenchymal stromal cells ameliorate cognitive decline by rescuing synaptic dysfunction and regulating inflammatory responses in APP/PS1 mice. FASEB J. 2018;32(2):654–68. https://doi.org/10.1096/fj.201700600R.

    Article  CAS  PubMed  Google Scholar 

  119. Hajinejad M, Sahab‐Negah S. Neuroinflammation: the next target of exosomal microRNAs derived from mesenchymal stem cells in the context of neurological disorders. J Cell Physiol. 2021;236(12):8070–81. https://doi.org/10.1002/jcp.30495.

    Article  CAS  PubMed  Google Scholar 

  120. Nicholas M, Vlaeyen JWS, Rief W, Barke A, Aziz Q, Benoliel R, Cohen M, Evers S, Giamberardino MA, Goebel A, Korwisi B, Perrot S, Svensson P, Wang SJ, Treede RD. The IASP classification of chronic pain for ICD-11: chronic primary pain. Pain. 2019;160(1):28–37. https://doi.org/10.1097/j.pain.0000000000001390.

    Article  PubMed  Google Scholar 

  121. Bing H, Das P, Lv X, Shi M, Aa J, Wang K, Duan L, Gilbert JA, Nie Y, Xiao-Lei W. Effects of ‘healthy’ fecal microbiota transplantation against the deterioration of depression in fawn-hooded rats. mSystems. 2022. https://doi.org/10.1128/msystems.00218-22.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Sommer C, Leinders M, Üçeyler N. Inflammation in the pathophysiology of neuropathic pain. Pain. 2018;159(3):595–602.

    CAS  PubMed  Google Scholar 

  123. Zhang YU, Ye G, Zhao J, Chen Y, Kong L, Sheng C, Yuan L. Exosomes carried miR-181c-5p alleviates neuropathic pain in CCI rat models. Anais da Acad Brasileira de Ciencias. 2022. https://doi.org/10.1590/0001-3765202220210564.

    Article  Google Scholar 

  124. Cata JP, Uhelski ML, Gorur A, Dougherty PM. Nociception and pain: new roles for exosomes. Neuroscientist. 2022;28(4):349–63.

    CAS  PubMed  Google Scholar 

  125. D’Agnelli S, Gerra MC, Bignami E, Arendt-Nielsen L. Exosomes as a new pain biomarker opportunity. Mol Pain. 2020;16:1744806920957800.

    PubMed  PubMed Central  Google Scholar 

  126. Cata JP, Gorur A, Yuan X, Berg NK, Sood AK, Eltzschig HK. Role of micro-RNA for pain after surgery: narrative review of animal and human studies. Anesth Anal. 2020;130(6):1638.

    Google Scholar 

  127. Liu W, Rong Y, Wang J, Zhou Z, Ge X, Ji C, Jiang D, Gong F, Li L, Chen J, Zhao S. Exosome-shuttled miR-216a-5p from hypoxic preconditioned mesenchymal stem cells repair traumatic spinal cord injury by shifting microglial M1/M2 polarization. J Neuroinflamm. 2020;17(1):1–22.

    Google Scholar 

  128. Bagi HM, Ahmadi S, Tarighat F, Rahbarghazi R, Soleimanpour H. Interplay between exosomes and autophagy machinery in pain management: state of the art. Neurobiol Pain. 2022;12:100095. https://doi.org/10.1016/j.ynpai.2022.100095.

    Article  Google Scholar 

  129. Liang Y, Xu X, Li X, Xiong J, Li B, Duan L, Wang D, Xia J. Chondrocyte-targeted microRNA delivery by engineered exosomes toward a cell-free osteoarthritis therapy. ACS Appl Mater Interfaces. 2020;12(33):36938–47.

    CAS  PubMed  Google Scholar 

  130. Huang L, Chaoquan H, Hui Chao Y, Zhang YL, Hou J, Zhong X, He L, Li H, Chen H. Drug-resistant endothelial cells facilitate progression, EMT and chemoresistance in nasopharyngeal carcinoma via exosomes. Cell Signall. 2019;63:109385. https://doi.org/10.1016/j.cellsig.2019.109385.

    Article  CAS  Google Scholar 

  131. Riquelme JA, Takov K, Santiago-Fernández C, Rossello X, Lavandero S, Yellon DM, Davidson SM. Increased production of functional small extracellular vesicles in senescent endothelial cells. J Cell Mol Med. 2020;24(8):4871–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Xi XM, Xia SJ, Lu R. Drug loading techniques for exosome-based drug delivery systems. Die Pharmazie Int J Pharm Sci. 2021;76(2–3):61–7.

    CAS  Google Scholar 

  133. Herrmann IK, Wood MJ, Fuhrmann G. Extracellular vesicles as a next-generation drug delivery platform. Nat Nanotechnol. 2021;16(7):748–59.

    CAS  PubMed  Google Scholar 

  134. Yang Z, Yang Y, Xu Y, Jiang W, Shao Y, Xing J, Chen Y, Han Y. Biomimetic nerve guidance conduit containing engineered exosomes of adipose-derived stem cells promotes peripheral nerve regeneration. Stem Cell Res Therapy. 2021;12(1):1–14.

    Google Scholar 

  135. Fan B, Chopp M, Zhang ZG, Liu XS. Treatment of diabetic peripheral neuropathy with engineered mesenchymal stromal cell-derived exosomes enriched with microRNA-146a provide amplified therapeutic efficacy. Exp Neurol. 2021;341:113694.

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Singh A, Raghav A, Shiekh PA, Kumar A. Transplantation of engineered exosomes derived from bone marrow mesenchymal stromal cells ameliorate diabetic peripheral neuropathy under electrical stimulation. Bioactive Mater. 2021;6(8):2231–49.

    CAS  Google Scholar 

  137. Liu YP, Tian MY, Yang YD, Li H, Zhao TT, Zhu J, Mou FF, Cui GH, Guo HD, Shao SJ. Schwann cells-derived exosomal miR-21 participates in high glucose regulation of neurite outgrowth. iScience. 2022;25(10):105141.

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Walker S, Busatto S, Pham A, Tian M, Suh A, Carson K, Quintero A, Lafrence M, Malik H, Santana MX, Wolfram J. Extracellular vesicle-based drug delivery systems for cancer treatment. Theranostics. 2019;9(26):8001.

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Reshke R, Taylor JA, Savard A, Guo H, Rhym LH, Kowalski PS, Trung MT, Campbell C, Little W, Anderson DG, Gibbings D. Reduction of the therapeutic dose of silencing RNA by packaging it in extracellular vesicles via a pre-microRNA backbone. Nat Biomed Eng. 2020;4(1):52–68.

    CAS  PubMed  Google Scholar 

  140. Kou M, Huang L, Yang J, Chiang Z, Chen S, Liu J, Guo L, Zhang X, Zhou X, Xu X, Yan X. Mesenchymal stem cell-derived extracellular vesicles for immunomodulation and regeneration: a next generation therapeutic tool? Cell Death Dis. 2022;13(7):580.

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Shi MM, Yang QY, Monsel A, Yan JY, Dai CX, Zhao JY, Shi GC, Zhou M, Zhu XM, Li SK, Li P. Preclinical efficacy and clinical safety of clinical-grade nebulized allogenic adipose mesenchymal stromal cells-derived extracellular vesicles. J Extracell Vesicles. 2021;10(10):e12134.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Hao P, Li H, Zhou L, Sun H, Han J, Zhang Z. Serum metal ion-induced cross-linking of photoelectrochemical peptides and circulating proteins for evaluating cardiac ischemia/reperfusion. ACS Sensors. 2022;7(3):775–83.

    CAS  PubMed  Google Scholar 

  143. McVey MJ, Weidenfeld S, Maishan M, Spring C, Kim M, Tabuchi A, Srbely V, Takabe-French A, Simmons S, Arenz C, Semple JW. Platelet extracellular vesicles mediate transfusion-related acute lung injury by imbalancing the sphingolipid rheostat. Blood J Am Soc Hematol. 2021;137(5):690–701.

    CAS  Google Scholar 

  144. Li J, Zhao J, Tan T, Liu M, Zeng Z, Zeng Y, Zhang L, Fu C, Chen D, Xie T. Nanoparticle drug delivery system for glioma and its efficacy improvement strategies: a comprehensive review. Int J Nanomed. 2020;15:2563–82.

    CAS  Google Scholar 

  145. Mennan C, Garcia J, Roberts S, Hulme C, Wright K. A comprehensive characterisation of large-scale expanded human bone marrow and umbilical cord mesenchymal stem cells. Stem Cell Res Therapy. 2019;10(1):1–15.

    Google Scholar 

  146. Gudbergsson JM, Jønsson K, Simonsen JB, Johnsen KB. Systematic review of targeted extracellular vesicles for drug delivery–considerations on methodological and biological heterogeneity. J Control Releas. 2019;306:108–20.

    CAS  Google Scholar 

  147. Brennan CW, Verhaak RG, McKenna A, Campos B, Noushmehr H, Salama SR, Zheng S, Chakravarty D, Sanborn JZ, Berman SH, Beroukhim R. The somatic genomic landscape of glioblastoma. Cell. 2014;157(3):753.

    CAS  Google Scholar 

  148. Colao IL, Corteling R, Bracewell D, Wall I. Manufacturing exosomes: a promising therapeutic platform. Trends Mol Med. 2018;24(3):242–56.

    CAS  PubMed  Google Scholar 

  149. Kurian TK, Banik S, Gopal D, Chakrabarti S, Mazumder N. Elucidating methods for isolation and quantification of exosomes: a review. Mol Biotechnol. 2021;63(4):249–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Yang B, Li Y, Zheng W, Yin Z, Liu M, Yin L, Liu C. Motion prediction for beating heart surgery with GRU. Biomed Signal Process Control. 2023;83:104641.

    Google Scholar 

  151. Théry C, Witwer KW, Aikawa E, Alcaraz MJ, Anderson JD, Andriantsitohaina R, Antoniou A, Arab T, Archer F, Atkin‐Smith GK. 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. 2018;7(1):1535750.

Download references

Acknowledgements

Not applicable.

Funding

This study was not funded by any funding institute.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the investigation, conceptualization, and analysis of the information in this manuscript and were involved in the writing process. MS, FD, and HK contributed to methodology and writing—original draft. NB and JA performed reviewing and editing. VM prepared figures. SE contributed to supervisions, methodology, and reviewing and editing.

Corresponding authors

Correspondence to Hossein Kargar Jahromi or Somayeh Ebrahimi-Barough.

Ethics declarations

Ethics approval and consent to participate

This article does not contain any studies with human participants or animals performed by any of the authors.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Namini, M.S., Daneshimehr, F., Beheshtizadeh, N. et al. Cell-free therapy based on extracellular vesicles: a promising therapeutic strategy for peripheral nerve injury. Stem Cell Res Ther 14, 254 (2023). https://doi.org/10.1186/s13287-023-03467-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-023-03467-5

Keywords