Next Article in Journal
nc886, a Non-Coding RNA, Is a New Biomarker and Epigenetic Mediator of Cellular Senescence in Fibroblasts
Next Article in Special Issue
Adipose Stromal/Stem Cell-Derived Extracellular Vesicles: Potential Next-Generation Anti-Obesity Agents
Previous Article in Journal
Shaping of Monocyte-Derived Dendritic Cell Development and Function by Environmental Factors in Rheumatoid Arthritis
Previous Article in Special Issue
Mesenchymal Stem Cell-Derived Extracellular Vesicle: A Promising Alternative Therapy for Osteoporosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Potential of Mesenchymal Stem Cells (MSCs) and MSC-Derived Extracellular Vesicles for the Treatment of Spinal Cord Injury

1
Department of Orthopedic Surgery, Hanil General Hospital, 308 Uicheon-ro, Dobong-gu, Seoul 01450, Korea
2
Department of Laboratory Animal Center, Daegu-Gyeongbuk Medical Innovation Foundation (DGMIF), Daegu 41061, Korea
3
Department of Biomaterials Science, Life and Industry Convergence Institute, Pusan National University, Miryang 50463, Korea
4
Efficacy Evaluation Team, Food Science R&D Center, KolmarBNH CO., LTD, 61Heolleungro 8-gil, Seocho-gu, Seoul 06800, Korea
5
Cellexobio, Co. Ltd., Daegu 42415, Korea
6
New Drug Development Center, Osong Medical Innovation Foundation, Chungbuk 28160, Korea
7
Department of Orthopedic Surgery, Yeungnam University College of Medicine, Yeungnam University Medical Center, 170 Hyonchung-ro, Namgu, Daegu 42415, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work as the first author.
Int. J. Mol. Sci. 2021, 22(24), 13672; https://doi.org/10.3390/ijms222413672
Submission received: 18 November 2021 / Revised: 14 December 2021 / Accepted: 18 December 2021 / Published: 20 December 2021

Abstract

:
Spinal cord injury (SCI) is a life-threatening condition that leads to permanent disability with partial or complete loss of motor, sensory, and autonomic functions. SCI is usually caused by initial mechanical insult, followed by a cascade of several neuroinflammation and structural changes. For ameliorating the neuroinflammatory cascades, MSC has been regarded as a therapeutic agent. The animal SCI research has demonstrated that MSC can be a valuable therapeutic agent with several growth factors and cytokines that may induce anti-inflammatory and regenerative effects. However, the therapeutic efficacy of MSCs in animal SCI models is inconsistent, and the optimal method of MSCs remains debatable. Moreover, there are several limitations to developing these therapeutic agents for humans. Therefore, identifying novel agents for regenerative medicine is necessary. Extracellular vesicles are a novel source for regenerative medicine; they possess nucleic acids, functional proteins, and bioactive lipids and perform various functions, including damaged tissue repair, immune response regulation, and reduction of inflammation. MSC-derived exosomes have advantages over MSCs, including small dimensions, low immunogenicity, and no need for additional procedures for culture expansion or delivery. Certain studies have demonstrated that MSC-derived extracellular vesicles (EVs), including exosomes, exhibit outstanding chondroprotective and anti-inflammatory effects. Therefore, we reviewed the principles and patho-mechanisms and summarized the research outcomes of MSCs and MSC-derived EVs for SCI, reported to date.

1. Introduction

Spinal cord injury (SCI) is a life-threatening, devastating injury to the spinal cord, leading to temporary or permanent changes to the cord, accompanied by partial or complete loss of motor, sensory, and autonomic functions [1,2]. SCI frequently leads to paralysis called paraplegia or quadriplegia, with sensory dysfunction below the injury level [3]. It generally induces severe psychological and economic burdens on patients and healthcare systems [4,5], and can negatively affect the majority of basic bodily functions, such as breathing, bowel and bladder function, hormone release, and sexual function, because of the loss of connection between the brain and the peripheral nervous system [6]. It is estimated that the prevalence and incidence of SCI worldwide is 236–4187 per million people, with up to 770,000 new cases per year and is more common in males below 30 years of age [7,8,9].
Initial mechanical insult caused by physical forces such as contusion, compression, transection, or stretching of the spinal column generally causes spinal cord disruption and results in primary injury [10,11,12,13,14]. Primary injury is defined as immediate mechanical injury to the spinal cord, which is an irreversible process [15,16]. A primary injury is followed by a cascade of secondary injuries, exacerbating the condition of the injured spinal cord [17,18]. Secondary injury occurs within minutes of the primary mechanical injury, manifested as local vascular damage, subsequent progressive hemorrhage, and ischemia, edema, thrombosis, ionic changes, oxidative stress through the release of free radicals, lipid peroxidation, excitotoxicity, and cell death promoted by apoptosis and cell necrosis [2]. Furthermore, the inflammatory reaction and gliosis hyperplasia, following the formation of an inhibitory environment and scar formation, interfere with axonal regeneration and limit the therapeutic potential [19,20].
Despite recent clinical advances in SCI management showing some improvement in patients’ quality of life, recovery from SCI remains substantially limited [21,22]. Pathologic targets for the treatment of SCI can be divided into three broad categories. First, surgical decompression and the removal of mechanical compressing material of the spinal cord at the initial trauma [2,23,24,25,26,27]. Second, anti-inflammatory treatments for the level around the injured spinal cord. There are some considerable inflammatory and anti-inflammatory mechanisms and factors, from the initial spinal cord injury to the late chronic phase [28,29,30]. Third, axonal regeneration at the site of the spinal cord injury, set as the ultimate goal for the treatment of spinal cord injuries [6]. With the initial SCI, macrophages are intensely infiltrated into the damaged lesion, and contribute to form a cavity of injury (COI) around the injured site, which cut off the neuronal regeneration [31,32,33,34]. Axonal regeneration is disturbed by scar [35] and the COI lesions that filled with fluid, because the axons do not have an ability to cross the liquid content in the COI without any bridge-like structures to cross it [36]. In addition, arachnoiditis, a granulomatous infiltration around the damaged spinal cord, contributes to the formation of a mature scar that does not contain astrocytes or other glial cells [20]. The prognosis of SCI patients remains abysmal, the mortality rate remains high, and life expectancy is significantly shortened [37].
Stem cell transplantation therapy for damaged spinal cords is a promising therapeutic strategy for replacing the damaged neuronal cells and creating an environment conducive to repair [3]. Cell therapies show neuroprotective and regenerative potential in SCI with diverse targets and stimulative responses, including regulation of inflammatory reactions, nutritional support, and promotion of neuronal plasticity [38]. Several types of stem cells have been tested, or are currently being tested, clinically for SCI treatment [39]. The majority of the experimental and clinical trials to treat SCI used mesenchymal stem cells (MSCs) isolated from bone marrow (BM-MSCs), umbilical cord MSCs (U-MSCs), and adipose tissue MSCs (AD-MSCs). Known mechanisms of MSCs to treat SCI include suppression of inflammation to limit secondary injury, secretion of paracrine factors that protect the remaining axons and promote axonal regeneration, and differentiation of MSCs into nerve cells to replace the damaged nerve cells [40,41]. MSC synthesis of neurotrophic and angiogenic factors promotes neuronal survival and regeneration. Furthermore, high biosafety and immunomodulation of MSCs make them the most promising cell type for SCI regenerative therapy [42]. However, despite these promising results of MSCs in SCI therapy, certain studies have reported that MSCs have numerous drawbacks and that their therapeutic properties are more likely to be due to their paracrine action [43].
In this review, we introduce the cutting-edge status of SCI treatment with MSCs and MSC-derived EVs, focusing on the potential therapeutic mechanisms with experimental and clinical trial results. Furthermore, we discuss the prospects, current limitations, and challenges of MSCs and MSC-derived EVs, along with hopes for future promising therapeutic methods for SCI.

2. Pathophysiology of Spinal Cord Injury

The pathological process of SCI can be divided into two consecutive phases: primary and secondary injury [44,45]. Primary injury occurs immediately at the time of injury, and secondary injury begins within minutes of the primary injury [46] (Figure 1).
Primary injury is an immediate mechanical injury caused by physical force that causes irreversible damage to the spinal cord [10,11,15,16]. The initial mechanical force leads to the rupture of the axonal membranes of the spinal cord and the release of inhibitory materials from the myelin sheath, including neurite outgrowth inhibitor protein A, myelin-associated glycoprotein, oligodendrocyte myelin glycoprotein, and chondroitin sulfate proteoglycan, all of which are powerful inhibitory factors for axonal regeneration [47,48,49,50,51,52].
Secondary injury is delayed and progressive, presenting within minutes of mechanical insult. In addition to mechanical damage, such as cord hemorrhage, swelling, ischemia, and blood-spinal cord barrier (BSCB) disruption, inflammatory cells release inflammatory cytokines due to BSCB destruction [53,54,55]. Secondary injury negatively affects cell survival in the damaged neuronal tissues and also the surrounding tissue, causing an enlargement of the lesion into the adjacent spinal cord segments in rostro-caudal directions [13]. Secondary injury includes electrolyte shifts, free radical generation, and release of toxic compounds and excitatory amino acids that trigger cell necrosis and apoptosis at the injured site [56,57,58,59,60,61,62,63,64,65]. Furthermore, proinflammatory cytokines and chemokines, such as interleukin (IL)-1β, IL-6, and tumor necrosis factor (TNF)-α, promote the differentiation of neural stem/progenitor cells into astroglia, resulting in the formation of scar tissue [66,67,68]. Tissue necrosis and cavity and scar formation, combined with axonal degeneration, ultimately impede functional recovery [69,70]. In the early phase of secondary injury, the glial scar plays a positive role in the injury site through BSCB regeneration and by limiting inflammation and toxic compounds and removing debris. However, at the later phase of injury, glial and fibrotic scars, along with axonal growth inhibitors, interfere with neuronal regeneration [71,72].
The BSCB plays a role in maintaining the normal function of the nervous system, and its unique properties and functions are regulated by neurovascular unit cells [73]. The BSCB is composed of the basement membrane, pericytes, capillary endothelial cells, and astrocyte foot processes [74]. The blood vessels at the site of injury are destroyed immediately after SCI, and the BSCB far from the injury area is permanently destroyed [75]. Loss of barrier integrity leads to increased permeability and the inflow of toxic materials into the injured spinal cord, resulting in edema and death of neuronal cells [76]. Therefore, ensuring that the integrity of the BSCB is uncompromised might be a potential target for SCI treatment. Pericytes, as a part of the neurovascular unit, are essential for the formation, maintenance of integrity, and function of the microvessels and BSCB. Pericytes have the ability to secure the stability of microvessels via three possible mechanisms: promoting the expression of endothelial tight junction proteins, regulating vesicle transport and body flow across cells, and moderating the tightness connection arrangement [77].
Neuroinflammation is characterized by the activation of local resident immune cells, and this activation is arbitrated by a protein complex-inflammasome called the nucleotide-binding domain-like receptor protein 3 (NLRP3) inflammasome. This inflammasome plays a very important role in SCI secondary injuries [78]. The NLRP3 inflammasome is located in the cytoplasm and is involved in the regulation of natural immune reactions [79,80]. Animal experimental models have shown that the NLRP3 inflammasome may be triggered and up-regulated following SCI and that inhibition of the NLRP3 inflammasome promotes functional recovery after SCI [81,82,83,84,85].
In addition, both classic and alternative complement pathways in the local immune response can be activated after SCI [86]. Activation of these pathways may exacerbate inflammatory reactions in the secondary injury process. Complements C1q and C3 are known to be related to the NF-κB signaling pathway [87,88], and secondary injury in SCI is regulated by NF-κB [84]. Thus, inhibiting NF-κB may be a possible mechanism for minimizing the inflammatory reaction and promoting functional recovery after SCI.
Macrophages are also involved in immune regulation in SCI secondary injuries. The polarization of macrophages determines their role in the inflammatory process [89]. The CD68+ phenotype is known as the pro-inflammatory macrophage, induced by TNF-α, IFN-γ, and IL-6 [89], while the CD163+ phenotype, known as the anti-inflammatory macrophage, is induced by IL-10. At the initial SCI site, myelin damage induces the infiltration of numerous macrophages into the site of necrosis via chemotaxis. The prevalence of CD68+/CD163- macrophages, which are the pro-inflammatory phenotypes around the COI, exhibit the severe inflammation and further contribute to progressive spinal cord destruction through beyond 16 weeks from the initial SCI [20,90]. Persistence of CD68+/CD163- macrophages showed an ongoing severe inflammatory state of the SCI lesion, and the gradual decline of them indicates a progressive increase in the anti-inflammatory process.
Astrocytes have the ability to hinder or promote recovery of the central nervous system (CNS); thus, they play a very important role in the SCI process [91,92,93,94,95]. Two phenotypes of reactive astrocytes, A1 and A2 astrocytes, pre-present and are induced by neuroinflammation and ischemia. A1 astrocytes, generally formed immediately after SCI induced by IL-1α, TNF-α, and C1q, have neurotoxic effects on myelin, synapses, and neurons that can lead to neuronal and oligodendrocyte death [96]. Contrarily, A2 astrocytes exert a protective function by up-regulating the expression of certain neurotrophic factors [91]. Therefore, selective inhibition of A1 astrocytes may be a potential SCI treatment strategy. Furthermore, it has recently been shown that reactive astrocytes eliminate red blood cells (RBCs) around SCI lesions through phagocytosis. This mechanism, named the astrocytic erythrophagocytosis, is considered to contribute to the rapid removal of scattered RBCs around the injured site to prevent macrophage aggregation and associated destructive inflammation.
Recently, microRNAs (miRNAs) have been shown to be involved in tissue injury and regenerative processes, and several miRNAs have attracted attention as potential targets for SCI treatment. miRNAs are endogenous non-coding RNAs with a length of 20–24 nucleotides that cause translational inhibition and degradation of these target messenger RNAs (mRNAs) [97,98]. miRNA-21 expression increases in the injury of various tissues and organs. It reduces neuronal apoptosis by promoting the activation of the phosphatase and tensin homolog-protein kinase B (Akt) signaling pathway [99] and regulating the expression of apoptosis-related proteins such as Bax, Bcl-2, caspase-9, and caspase-3 [100,101]. miRNA-133b also plays a key role in neuronal differentiation, growth, and apoptosis [102,103,104]. Yu et al. showed that reduced miRNA-133b expression reduces neuronal axonal regeneration and does not help recover motor function [105]. It has recently been shown that miRNA-126 promotes functional recovery after SCI. Hu et al. reported that miRNA-126 expression decreases after SCI, whereas increasing miRNA-126 levels appear to reduce inflammation and promote angiogenesis and functional recovery [106].

3. Mesenchymal Stem Cells for the Potential Treatment of Spinal Cord Injury

Stem cell transplantation therapies in SCI show neuroprotective and regenerative potential with different targets and responses [38]. Among the various stem cells currently available, certain inherent properties of MSCs are advantageous over other stem cells in the research; besides, MSCs are easier to harvest and isolate, and they have low immunogenicity [107,108,109]. Moreover, MSCs have fewer ethical considerations than other types of stem cells, such as embryonic stem cells [110]. The core capabilities of MSCs, such as homing, proliferation, differentiation, secretion, and immunomodulatory abilities, are of considerable interest for SCI treatment [111]. The International Society for Cellular Therapy position statement defined MSCs as cells that (1) adhere to plastic in culture conditions; (2) express CD105, CD73, and CD90, but not CD45, CD34, CD14, CD11b, CD79alpha, CD19, and HLA-DR surface molecules; and (3) are able to differentiate into osteoblasts, adipocytes, and chondroblasts in vitro [112].
MSCs are also known to have the ability to differentiate into neural cells and express neuronal markers [113,114,115,116] through specific procedures. Traditionally, the regenerative potential of MSCs is thought to be due to cell plasticity [117,118]. Although MSCs have the ability to differentiate into various neural and glial cells, most of their effects are based on their paracrine action [119]. MSCs produce and release a broad range of bioactive molecules, called secretomes. Proteomic analysis of secretomes revealed that they contain trophic factors and cytokines, such as growth factors, immunomodulators, and antioxidants [120]. Therefore, paracrine factors from MSCs have diverse functions, including anti-inflammatory, anti-apoptotic, extracellular matrix modulatory, and neuroprotective actions, by protective action against fibrosis, apoptosis, and oxidative damage [121].
Certain molecules, including vascular endothelial growth factor (VEGF), hepatocyte growth factor (HGF), insulin-like growth factor-1 (IGF-1), stanniocalcin-1, transforming growth factor-β (TGF-β), and granulocyte-macrophage colony-stimulating factor, promote the survival of damaged neurons and oligodendrocytes [122,123]. They also stimulate angiogenesis along with placental growth factor, monocyte chemoattractant protein-1, basic fibroblast growth factor (bFGF), and IL-6 [124]. Stimulation of neural cell proliferation and regeneration is mediated by glial cell-derived neurotrophic factor, brain-derived neurotrophic factor (BDNF), and nerve growth factor (NGF) [125]. MSCs express their immunomodulatory actions through cell-to-cell contact and the secretion of IL-10, TGF-β, PGE-2, galectin-1, indolamine 2,3 dioxygenase (IDO), and HLA-G [122,126,127,128].
By modulating inflammation, MSCs reduce neural damage to the remaining and surrounding unaffected tissues from secondary injury. MSCs can also improve neurite growth by improving the environment of the extracellular matrix by inhibiting gliosis [129]. In addition, antioxidant properties, direct cell fusion, mitochondrial transfer, and production of MSC microvesicles have been reported to exert their therapeutic effect [130,131,132].
MSCs are available from different tissues, including BM-MSCs, U-MSCs, AD-MSCs, neural stem cells, neural progenitor cells, embryonic stem cells, and induced pluripotent stem cells [38]. Among them, BM-MSCs, U-MSCs, and AD-MSCs, which have undergone the most studies and clinical trials, will be described in more detail in the subsequent sections.

3.1. Bone Marrow Mesenchymal Stem Cells

The bone marrow is the most popular source of MSCs. BM-MSCs are partially differentiated progenitor cells present in adult bone marrow. They are considered pluripotent, capable of differentiating into neurons and glial cells, and are involved in continuous hematopoiesis and bone regeneration [133,134]. However, further studies revealed that BM-MSC therapy is mainly involved in cell fusion and transdifferentiation, instead of cell differentiation. The introduction of BM-MSCs to the injury site showed a beneficial role in recovery from SCI by reducing the inflammatory reactions and astroglial scarring density [135,136], improving the microenvironment of the injury site, enhancing the nutritional support, and reducing the BSCB leakage [137]. Therefore, BM-MSCs might have diverse treatment potential for SCI because of their reduced immunogenicity and improved availability [38,138]. Interestingly, these beneficial effects occurred similarly when BM-MSCs were administered locally into the spinal cord cavity [139], intrathecally [140], or systemically [141,142]. In an animal experimental model of the BM-MSC intravenous graft, functional recovery of SCI was achieved through the expansion of neurotrophic factors, including NGF, BDNF, and VEGF, which are key regulators of neuronal differentiation, initiation, and maintenance of angiogenesis [142,143,144].
Jeon et al. performed a phase I trial in which BM-MSCs were administered into the intramedullary space (8 × 106 cells) and intradural space (4 × 107 cells) in 10 patients with SCI. Long-term follow-up of the patients showed that three patients with American Spinal Injury Association (ASIA) impairment scale (AIS) grade B improved their motor power of the upper extremities with better activities of daily living [145]. Furthermore, Saito et al. confirmed that significant improvement was observed in two patients with AIS grades B and C following BM-MSC therapy [146]. El-Kheir et al. reported an improvement in AIS grade in 17 out of 50 BM-MSC-treated patients combined with physiotherapy [147]. Karamouzine et al. administered BM-MSCs into 11 patients with AIS grade A and found that five patients had their AIS grade improved to C [148]. Several studies have reported improvement in AIS grade in chronic SCI patients who received MSCs through the intraspinal route [149,150,151]. In contrast, Pal et al. reported that 30 BM-MSC-treated patients did not show any improvement or conversion in their AIS grades [152].

3.2. Umbilical Cord-Derived Mesenchymal Stem Cells

U-MSCs have several unique characteristics for use in SCI treatment, including ease of sourcing, excellent in vitro expansion, and fast proliferation. Furthermore, U-MSCs show low immunogenicity because they express very low or no expression of human leukocyte antigen typing, which is associated with immune rejection [128,153]. To avoid immune rejection, these cells utilize several additional mechanisms, including modulation of dendritic cell and T-cell functions and induction of regulatory T-cells [154].
U-MSCs have the ability to develop into a homogeneous population that expresses neural markers and develop neural phenotypic features [155]. These cells can be differentiated into multiple cell types, including neural-like and glial-like cells [156,157,158,159]. U-MSCs can be collected noninvasively, and their usage has not been hampered by ethical issues. They showed improved motor function and alleviated allodynia and hyperalgesia after SCI, via protection of neurons from apoptosis [160], inhibition of glial scar formation via regulation of metalloproteinase-2 [161], attenuation of ischemic damage of the spinal cord [162], and decreased reactive astrocytes in animal experiments [163,164,165,166].
Liu et al. found that 13 out of 22 patients showed improvement in the AIS quality of life in most patients with incomplete SCI [167]. In addition, Cheng et al. found that 7 out of 10 patients who received cell therapy demonstrated improvement in sensation, motion, muscle tension, and self-care ability [168]. Kang et al. reported improved motor function in the lower limb and expanded the atrophied spinal cord after injection of U-MSCs into the subarachnoid, intradural, or extradural space of the spinal cord in patients with compressed fractures [169].

3.3. Adipose-Derived Mesenchymal Stem Cells

Adipose tissue is distributed ubiquitously in the body and can be easily collected using minimally invasive techniques, such as liposuction or simple surgical interventions, and contains more somatic stem cells compared to bone marrow [170,171]. AD- and BM-MSCs share certain characteristics, such as cell morphology and expression of cell surface antigens. However, the rates of proliferation and multilineage capabilities markedly differ [172]. More somatic stem cells are contained in the adipose tissue than in bone marrow, which makes AD-MSCs a good MSC material for SCI treatment [170,171].
BM-MSCs are characterized by slow proliferation and higher osteogenic and chondrogenicity. In contrast, AD-MSCs exhibit higher proliferative activity and secrete higher levels of IGF-1, VEGF-D, and IL-8 [173]. Furthermore, the secretion levels of VEGF-A, angiogenin, bFGF, NGF, stem cell-derived factor-1, and HGF from BM-MSCs are comparable to those of AD-MSCs [174]. According to these findings, AD-MSCs tend to promote angiogenesis stimulation more strongly [174].
Intravenous AD-MSC administration improved hindlimb motor function via angiogenesis activation and up-regulation of extracellular signal-regulated kinase and Akt [175]. AD-MSCs also contribute to cell survival and tissue repair by increasing the expression of beta3-tubulin, BDNF, and ciliary neurotrophic factor [176]. The inflammatory response can also be down-regulated by the administration of AD-MSCs, which is mediated by blocking the infiltration of ED1 macrophages and attenuating Notch1 signaling [174,177,178]. In addition, the intrinsic ability of AD-MSCs to transdifferentiate into neuron/motoneuron-like cells may reduce cavitation and immune suppression by inhibiting astrocyte reactivation and the secretion of anti-inflammatory factors [179].
Ra et al. observed toxicity and tumorigenicity following intravenous injection of human AD-MSC in eight male patients with chronic SCI. There were no serious transplant-related adverse events in all patients during the 3-month follow-up period [180]. Hur et al. reported that 10 out of 14 SCI patients exhibited improvement in their sensory function, five patients experienced improvement of motor function, and two patients had improved voluntary anal contraction after administration of AD-MSCs [181]. Bydon et al. reported that treatment of an SCI patient with 100 million autologous AD-MSCs showed improvement in ASIA motor and sensory scores as well as improvement in the quality of life [182].

4. Why Should We Pay Attention to Extracellular Vesicles over Mesenchymal Stem Cells as a Therapeutic Source for Spinal Cord Injury?

MSCs have the potential to regenerate injured tissues or control the immunologic cascade; however, they also have significant limitations, particularly in view of carrying out clinical studies and developing therapeutic agents in real clinical practice. First, MSCs have a survival issue after cell implantation [183]. The longevity of MSCs may be driven by insufficient environment, cell niche, survival of MSCs, and poor intercellular communication between the cells. In particular, certain researchers have demonstrated the paradoxical period after implantation that pro-inflammatory activity surpasses anti-inflammatory activity in some phases [184]. Within this period, MSCs cannot survive sufficiently with reduced function. Second, MSCs are significantly heterogeneous due to the diversity of donor condition, type, differentiation capacity, and other factors between cells. In addition, MSCs are severely sensitive to the environment, resulting in negative effects on disease modulation, such as severe inflammation and active osteoarthritis [185,186]. Finally, the entire manufacturing process for MSCs, including ex vivo expansion, isolation technique, and cultivation method, has not yet been standardized; undetermined factors can affect the senescence and loss of capacity of implanted MSCs. To overcome these limitations of MSCs, extracellular vesicles (EVs), also called exosomes, have recently emerged as a novel source in the field of regenerative and anti-inflammatory medicine. Hence, as an alternative material in regenerative medicine, EVs should be considered for further research and development.

5. Overview and Characteristics of Extracellular Vesicles

EVs are lipid bilayer vesicles derived from cells, serum, or other biological fluids (Figure 2). These are involved in biological signal transduction between cells and are emerging as mediators of disease therapeutics, diagnostic biomarkers, and drug delivery systems because of their ability to regulate various biological processes [187,188,189]. EVs are cell-derived vesicles that include cell-derived genetic materials and possess biological functional activity. EV cargo consists of bioactive molecules, including mRNAs, miRNAs, DNA, lipids, proteins, and metabolites. There are four types of EVs, categorized by their size and composition. Among them, exosomes are 50–200 nm in diameter and have a cup-shaped lipid bilayer membrane structure. Exosome membranes are enriched in cholesterol, ceramide, and sphingolipids, which are secreted through the budding of intraluminal vesicles and multivesicular bodies (MVBs) [190,191]. Exosomes are released by fusing MVBs with the cell membrane. However, another extracellular vesicle, microvesicles, is released from the cell through the outer bud of the cell membrane or apoptotic cell membrane [192].
Stem cell therapy and cell transplantation have been extensively studied and used as cell therapies in human clinical trials. In particular, MSCs are commonly used as cell-based therapeutics owing to their regenerative and immunosuppressive effects [193]. Currently, more than 600 clinical trials using MSCs are available at www.clinicaltrials.gov (5 November 2021) [194]. According to clinical trial information, MSCs are used to treat SCI, osteoarthritis, knee cartilage damage, and cancer. Damaged tissue regeneration is possible through the paracrine action of MSCs, and MSCs have been shown to be effective in healing and regenerating damaged tissues. However, there are certain disadvantages of systemically administered MSCs, including that they remain in the tissue for a long time to cause an immune response and are expensive and difficult to store as cell therapy, which need to be overcome. In recent years, MSC-based therapies have undergone multiple paradigm shifts to address these issues [194]. MSC-derived EVs are important because they show biological changes and regenerative effects as therapeutic agents for diseases and are safe as cell-free therapeutic agents. Currently, a clinical approach is underway to utilize MSC-derived EVs as a treatment for bone loss, diabetes mellitus type 1, Alzheimer’s disease, and sepsis (http://clinicaltrials.gov, accessed on 5 November 2021).
In contrast, the genetic material contained in EVs can be analyzed and used as a biomarker for disease diagnosis. miRNAs are packaged inside exosomes, and miRNAs from these secreted exosomes can be ideal biomarkers because of their high stability and degradation resistance. miRNAs are non-coding RNAs of 21–25 nucleotides, which regulate cellular responses after mRNA transcription [195]. Disease-specific miRNAs can be identified by comparing and analyzing the expression patterns of exosomal miRNAs from healthy individuals and exosomal miRNAs from patients. Circulating miRNAs are biomarkers for neurodegenerative diseases or cancers and can diagnose diseases early and observe prognosis [196,197]. For example, the let-7 miRNA family is abundant in the CNS and is involved in neurogenesis and increased cerebrospinal fluid in patients with Alzheimer’s disease [198,199].
The EVs use is one of the potentially promising approaches for delivering therapeutic agents to the CNS. EVs are miniscule and lipophilic vesicles that can be used in regenerative medicine, or to deliver drugs for anti-cancer, anti-inflammatory, and immune modulation across the blood-brain barrier (BBB) [200]. Effective treatment has become possible by modifying the surface of EVs or loading and delivering therapeutic substances to EVs. In addition, studies have reported the use of EVs as a drug delivery system to deliver therapeutic agents such as amyloid, miRNA-124, miRNA-133, siRNA, paclitaxel, and doxorubicin [201,202,203,204,205].

6. Research Trials Using Extracellular Vesicles for Spinal Cord Injury

To date, certain studies have evaluated the therapeutic efficacy of EVs for SCI in animal models. Numerous studies have demonstrated that EVs can pass through the BBB, reach the target lesion of the injured spinal cord, and positively affect the lesion. Guo et al. showed that intranasal EVs therapy could partly improve structural and electrophysiological function and, most importantly, significantly elicit functional recovery in rats with complete cord injury [206]. Kim HY et al. also demonstrated that accumulated EV-like nanovesicles enhanced blood vessel formation, attenuated inflammation and apoptosis in the injured spinal cord and consequently improved spinal cord function [207]. In addition, Zhong D, et al. reported that EVs could enhance the angiogenic activities in the injured spinal cord, with sufficient VEGF-A in the EVs, accelerated microvascular regeneration, reduced spinal cord cavity formation, and improved functional scores using the Basso mouse scale [208]. Together, these findings suggest that EVs possess many affirmative factors for the regeneration of injured spinal cords and, in the near future, may be a novel therapeutic agent for SCI in humans. However, in the literature and clinical trial registries, there have been no reports of clinical trials with EVs for SCI. For clinical trials, certain issues of the EVs, such as optimal reference for manufacturing and large quantity production, have to be proven entirely.

7. Conclusions and Future Perspectives

Functional recovery after SCI is considerably limited because of the very low plasticity and weak regenerative capacity of the CNS. Moreover, owing to the complex and long-term pathological process of SCI, recovery of the injured spinal cord is hampered by various factors. To date, SCI treatment is an unresolved challenge and there remains no effective strategy to restore the lost functions.
The various direct and indirect pathways for the regeneration of injured tissues possessed by MSCs and MSC-derived EVs have promising potential for SCI treatment. A series of small-scale patient trials revealed that MSC transplantation yielded better outcomes than the traditional treatments such as rehabilitation, including improvements in movement, sensation, and quality of life. Recently, MSC-derived EVs have become very popular in the field of regenerative medicine because they have various abilities to repair damaged tissues. Thus, MSC-derived EVs can be a good alternative material to overcome the inherent limitations of MSCs.
Certainly, various issues related to MSC-derived EVs, such as tissue sources, isolation, purification, and amplification, must be addressed first. To generate MSC-derived EVs in high yield and purity without affecting the biological activity of exosomes, it is necessary to establish a fast, inexpensive, and simple standardized isolation technique and purification procedure. In addition, along with the production of MSC-derived EVs, future trials to establish clinical effectiveness and safety in human applications should be conducted. Future studies to establish a comprehensive theoretical basis for the clinical application of MSC-derived EVs will provide a direction and hope for the clinical treatment of SCI.

Author Contributions

Conceptualization, G.W.L. and M.-S.S.; Methodology, G.-U.K. and G.W.L.; Software, G.W.L.; Validation, G.W.L., S.Y.Y. and M.-S.S.; Formal Analysis, G.-U.K. and S.-E.S.; Investigation, G.-U.K., S.-E.S., K.-K.K., J.-H.C., S.L., M.S. and J.-H.L.; Resources, G.W.L.; Data Curation, S.-E.S., S.-K.K., Y.I.K. and G.W.L.; Writing—Original Draft Preparation, G.-U.K. and S.-E.S.; Writing—Review and Editing, G.W.L. and M.-S.S.; Visualization, G.-U.K. and S.-E.S.; Supervision, G.W.L. and M.-S.S.; Project Administration, G.W.L.; Funding Acquisition, G.W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Bio & Medical Technology Development Program of the National Research Foundation (NRF) funded by the Ministry of Science & ICT (2019M3E5D1A02068105). This study was also supported by Daegu Metropolitan and support by Daegu-Gyeongbuk Medical Innovation Foundation (2021 Daegu Medi-Startup Program).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

We would like to thank Choi Young for assisting Figure 1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aziz, I.; Ramli, M.D.C.; Zain, N.S.M.; Sanusi, J. Behavioral and Histopathological Study of Changes in Spinal Cord Injured Rats Supplemented withSpirulina platensis. Evid. Based Complement. Altern. Med. 2014, 2014, 871657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sykova, E.; Cizkova, D.; Kubinova, S. Mesenchymal Stem Cells in Treatment of Spinal Cord Injury and Amyotrophic Lateral Sclerosis. Front. Cell Dev. Biol. 2021, 9, 695900. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, W.-C.; Liu, W.-F.; Bai, Y.-Y.; Zhou, Y.-Y.; Zhang, Y.; Wang, C.-M.; Lin, S.; He, H.-F. Transplantation of mesenchymal stem cells for spinal cord injury: A systematic review and network meta-analysis. J. Transl. Med. 2021, 19, 178. [Google Scholar] [CrossRef] [PubMed]
  4. Bhat, I.A.; Sivanarayanan, T.B.; Somal, A.; Pandey, S.; Bharti, M.K.; Panda, B.S.K.; Indu, B.; Verma, M.; Anand, J.; Sonwane, A.; et al. An allogenic therapeutic strategy for canine spinal cord injury using mesenchymal stem cells. J. Cell. Physiol. 2019, 234, 2705–2718. [Google Scholar] [CrossRef]
  5. Vismara, I.; Papa, S.; Rossi, F.; Forloni, G.; Veglianese, P. Current Options for Cell Therapy in Spinal Cord Injury. Trends Mol. Med. 2017, 23, 831–849. [Google Scholar] [CrossRef]
  6. Liau, L.L.; Looi, Q.H.; Chia, W.C.; Subramaniam, T.; Ng, M.H.; Law, J.X. Treatment of spinal cord injury with mesenchymal stem cells. Cell Biosci. 2020, 10, 112. [Google Scholar] [CrossRef] [PubMed]
  7. Fehlings, M.G.; Singh, A.; Tetreault, L.; Kalsi-Ryan, S.; Nouri, A. Global prevalence and incidence of traumatic spinal cord injury. Clin. Epidemiol. 2014, 6, 309–331. [Google Scholar] [CrossRef] [Green Version]
  8. Kumar, R.; Lim, J.; Mekary, R.A.; Rattani, A.; Dewan, M.C.; Sharif, S.Y.; Osorio-Fonseca, E.; Park, K.B. Traumatic Spinal Injury: Global Epidemiology and Worldwide Volume. World Neurosurg. 2018, 113, e345–e363. [Google Scholar] [CrossRef]
  9. Witiw, C.D.; Fehlings, M.G. Acute Spinal Cord Injury. J. Spin. Disord. Tech. 2015, 28, 202–210. [Google Scholar] [CrossRef]
  10. Morales, I.I.; Toscano-Tejeida, D.; Ibarra, A. Non Pharmacological Strategies to Promote Spinal Cord Regeneration: A View on Some Individual or Combined Approaches. Curr. Pharm. Des. 2016, 22, 720–727. [Google Scholar] [CrossRef]
  11. Tyler, J.Y.; Xu, X.-M.; Cheng, J.-X. Nanomedicine for treating spinal cord injury. Nanoscale 2013, 5, 8821–8836. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, J.; Pearse, D.D. Therapeutic Hypothermia in Spinal Cord Injury: The Status of Its Use and Open Questions. Int. J. Mol. Sci. 2015, 16, 16848–16879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ramer, L.M.; Ramer, M.S.; Steeves, J.D. Setting the stage for functional repair of spinal cord injuries: A cast of thousands. Spinal Cord 2005, 43, 134–161. [Google Scholar] [CrossRef] [PubMed]
  14. Rosenzweig, E.; McDonald, J.W. Rodent models for treatment of spinal cord injury: Research trends and progress toward useful repair. Curr. Opin. Neurol. 2004, 17, 121–131. [Google Scholar] [CrossRef] [PubMed]
  15. Kumar, H.; Ropper, A.E.; Lee, S.-H.; Han, I. Propitious Therapeutic Modulators to Prevent Blood-Spinal Cord Barrier Disruption in Spinal Cord Injury. Mol. Neurobiol. 2016, 54, 3578–3590. [Google Scholar] [CrossRef]
  16. Carlson, G.D.; Gorden, C. Current developments in spinal cord injury research. Spine J. 2002, 2, 116–128. [Google Scholar] [CrossRef]
  17. Hayta, E.; Elden, H. Acute spinal cord injury: A review of pathophysiology and potential of non-steroidal anti-inflammatory drugs for pharmacological intervention. J. Chem. Neuroanat. 2018, 87, 25–31. [Google Scholar] [CrossRef]
  18. Tator, C.H.; Fehlings, M.G. Review of the secondary injury theory of acute spinal cord trauma with emphasis on vascular mechanisms. J. Neurosurg. 1991, 75, 15–26. [Google Scholar] [CrossRef] [PubMed]
  19. Fehlings, M.G.; Hawryluk, G.W. Scarring after spinal cord injury. J. Neurosurg. Spine 2010, 13, 165–167. [Google Scholar] [CrossRef] [PubMed]
  20. Kwiecien, J.M.; Dabrowski, W.; Dąbrowska-Bouta, B.; Sulkowski, G.; Oakden, W.; Kwiecien-Delaney, C.J.; Yaron, J.R.; Zhang, L.; Schutz, L.; Marzec-Kotarska, B.; et al. Prolonged inflammation leads to ongoing damage after spinal cord injury. PLoS ONE 2020, 15, e0226584. [Google Scholar] [CrossRef] [Green Version]
  21. Wilson, J.R.; Forgione, N.; Fehlings, M.G. Emerging therapies for acute traumatic spinal cord injury. Can. Med. Assoc. J. 2013, 185, 485–492. [Google Scholar] [CrossRef] [Green Version]
  22. Dvorak, M.F.; Noonan, V.K.; Fallah, N.; Fisher, C.G.; Finkelstein, J.; Kwon, B.K.; Rivers, C.S.; Ahn, H.; Paquet, J.; Tsai, E.; et al. The Influence of Time from Injury to Surgery on Motor Recovery and Length of Hospital Stay in Acute Traumatic Spinal Cord Injury: An Observational Canadian Cohort Study. J. Neurotrauma 2015, 32, 645–654. [Google Scholar] [CrossRef]
  23. Burke, J.F.; Yue, J.K.; Ngwenya, L.B.; Winkler, E.A.; Talbott, J.F.; Pan, J.Z.; Ferguson, A.R.; Beattie, M.S.; Bresnahan, J.C.; Haefeli, J.; et al. Ultra-Early (<12 Hours) Surgery Correlates With Higher Rate of American Spinal Injury Association Impairment Scale Conversion After Cervical Spinal Cord Injury. Neurosurgery 2019, 85, 199–203. [Google Scholar] [CrossRef] [PubMed]
  24. Fehlings, M.G.; Vaccaro, A.; Wilson, J.R.; Singh, A.; Cadotte, D.W.; Harrop, J.S.; Aarabi, B.; Shaffrey, C.; Dvorak, M.; Fisher, C.; et al. Early versus Delayed Decompression for Traumatic Cervical Spinal Cord Injury: Results of the Surgical Timing in Acute Spinal Cord Injury Study (STASCIS). PLoS ONE 2012, 7, e32037. [Google Scholar] [CrossRef]
  25. Wutte, C.; Klein, B.; Becker, J.; Mach, O.; Panzer, S.; Strowitzki, M.; Maier, D.; Grassner, L. Earlier Decompression (<8 h) Results in Better Neurological and Functional Outcome after Traumatic Thoracolumbar Spinal Cord Injury. J. Neurotrauma 2019, 36, 2020–2027. [Google Scholar] [CrossRef]
  26. Rath, N.; Balain, B. Spinal cord injury—The role of surgical treatment for neurological improvement. J. Clin. Orthop. Trauma 2017, 8, 99–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Ahuja, C.S.; Wilson, J.R.; Nori, S.; Kotter, M.R.N.; Druschel, C.; Curt, A.; Fehlings, M.G. Traumatic spinal cord injury. Nat. Rev. Dis. Primers 2017, 3, 17018. [Google Scholar] [CrossRef]
  28. Streijger, F.; So, K.; Manouchehri, N.; Gheorghe, A.; Okon, E.B.; Chan, R.M.; Ng, B.; Shortt, K.; Sekhon, M.S.; Griesdale, D.E.; et al. A Direct Comparison between Norepinephrine and Phenylephrine for Augmenting Spinal Cord Perfusion in a Porcine Model of Spinal Cord Injury. J. Neurotrauma 2018, 35, 1345–1357. [Google Scholar] [CrossRef] [PubMed]
  29. Hawryluk, G.; Whetstone, W.; Saigal, R.; Ferguson, A.; Talbott, J.; Bresnahan, J.; Dhall, S.; Pan, J.; Beattie, M.; Manley, G. Mean Arterial Blood Pressure Correlates with Neurological Recovery after Human Spinal Cord Injury: Analysis of High Frequency Physiologic Data. J. Neurotrauma 2015, 32, 1958–1967. [Google Scholar] [CrossRef] [Green Version]
  30. Ryken, T.C.; Hurlbert, R.J.; Hadley, M.N.; Aarabi, B.; Dhall, S.S.; Gelb, D.E.; Rozzelle, C.J.; Theodore, N.; Walters, B. The Acute Cardiopulmonary Management of Patients with Cervical Spinal Cord Injuries. Neurosurgery 2013, 72, 84–92. [Google Scholar] [CrossRef] [Green Version]
  31. Seki, T.; Fehlings, M.G. Mechanistic insights into posttraumatic syringomyelia based on a novel in vivo animal model. J. Neurosurg. Spine 2008, 8, 365–375. [Google Scholar] [CrossRef]
  32. Kwiecien, J.M.; Jarosz, B.; Oakden, W.; Klapec, M.; Stanisz, G.J.; Delaney, K.H.; Kotlinska-Hasiec, E.; Janik, R.; Rola, R.; Dabrowski, W. An in vivo model of anti-inflammatory activity of subdural dexamethasone following the spinal cord injury. Neurol. Neurochir. Polska 2016, 50, 7–15. [Google Scholar] [CrossRef] [PubMed]
  33. Kwiecien, J.M.; Jarosz, B.; Urdzikova, L.M.; Rola, R.; Dabrowski, W. Original article Subdural infusion of dexamethasone inhibits leukomyelitis after acute spinal cord injury in a rat model. Folia Neuropathol. 2015, 1, 41–51. [Google Scholar] [CrossRef] [Green Version]
  34. Hu, R.; Zhou, J.; Luo, C.; Lin, J.; Wang, X.; Li, X.; Bian, X.; Li, Y.; Wan, Q.; Yu, Y.; et al. Glial scar and neuroregeneration: Histological, functional, and magnetic resonance imaging analysis in chronic spinal cord injury. J. Neurosurg. Spine 2010, 13, 169–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Sofroniew, M.V. Molecular dissection of reactive astrogliosis and glial scar formation. Trends Neurosci. 2009, 32, 638–647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kwiecien, J.M. Original article Cellular mechanisms of white matter regeneration in an adult dysmyelinated rat model. Folia Neuropathol. 2013, 3, 189–202. [Google Scholar] [CrossRef]
  37. Jain, N.B.; Ayers, G.D.; Peterson, E.N.; Harris, M.B.; Morse, L.R.; O’Connor, K.C.; Garshick, E. Traumatic Spinal Cord Injury in the United States, 1993-2012. JAMA 2015, 313, 2236–2243. [Google Scholar] [CrossRef] [PubMed]
  38. Huang, L.; Fu, C.; Xiong, F.; He, C.; Wei, Q. Stem Cell Therapy for Spinal Cord Injury. Cell Transplant. 2021, 30, 963689721989266. [Google Scholar] [CrossRef]
  39. Trounson, A.; McDonald, C. Stem Cell Therapies in Clinical Trials: Progress and Challenges. Cell Stem Cell 2015, 17, 11–22. [Google Scholar] [CrossRef] [Green Version]
  40. Nori, S.; Ahuja, C.S.; Fehlings, M.G. Translational Advances in the Management of Acute Spinal Cord Injury: What is New? What is Hot? Neurosurgery 2017, 64, 119–128. [Google Scholar] [CrossRef]
  41. Ruzicka, J.; Urdzikova, L.M.; Gillick, J.; Amemori, T.; Romanyuk, N.; Karova, K.; Zaviskova, K.; Dubisova, J.; Kubinova, S.; Murali, R.; et al. A Comparative Study of Three Different Types of Stem Cells for Treatment of Rat Spinal Cord Injury. Cell Transplant. 2017, 26, 585–603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Mukhamedshina, Y.O.; Gracheva, O.A.; Mukhutdinova, D.M.; Chelyshev, Y.A.; Rizvanov, A. Mesenchymal stem cells and the neuronal microenvironment in the area of spinal cord injury. Neural Regen. Res. 2019, 14, 227–237. [Google Scholar] [CrossRef]
  43. Liu, W.-Z.; Ma, Z.-J.; Li, J.-R.; Kang, X.-W. Mesenchymal stem cell-derived exosomes: Therapeutic opportunities and challenges for spinal cord injury. Stem Cell Res. Ther. 2021, 12, 102. [Google Scholar] [CrossRef]
  44. McDonald, J.W.; Sadowsky, C. Spinal-cord injury. Lancet 2002, 359, 417–425. [Google Scholar] [CrossRef]
  45. Tator, C.H. Update on the Pathophysiology and Pathology of Acute Spinal Cord Injury. Brain Pathol. 1995, 5, 407–413. [Google Scholar] [CrossRef]
  46. Alizadeh, A.; Dyck, S.M.; Karimi-Abdolrezaee, S. Traumatic Spinal Cord Injury: An Overview of Pathophysiology, Models and Acute Injury Mechanisms. Front. Neurol. 2019, 10, 282. [Google Scholar] [CrossRef] [Green Version]
  47. Snow, D.M.; Beller, J.A. Proteoglycans: Road Signs for Neurite Outgrowth. Neural Regen. Res. 2014, 9, 343–355. [Google Scholar] [CrossRef] [PubMed]
  48. Li, M.; Shibata, A.; Li, C.; Braun, P.E.; McKerracher, L.; Roder, J.; Kater, S.B.; David, S. Myelin-associated glycoprotein inhibits neurite/axon growth and causes growth cone collapse. J. Neurosci. Res. 1996, 46, 404–414. [Google Scholar] [CrossRef]
  49. Domeniconi, M.; Cao, Z.; Spencer, T.; Sivasankaran, R.; Wang, K.C.; Nikulina, E.; Kimura, N.; Cai, H.; Deng, K.; Gao, Y.; et al. Myelin-Associated Glycoprotein Interacts with the Nogo66 Receptor to Inhibit Neurite Outgrowth. Neuron 2002, 35, 283–290. [Google Scholar] [CrossRef] [Green Version]
  50. Schwab, M.E. Nogo and axon regeneration. Curr. Opin. Neurobiol. 2004, 14, 118–124. [Google Scholar] [CrossRef]
  51. Fournier, A.E.; Grandpre, T.; Strittmatter, S. Identification of a receptor mediating Nogo-66 inhibition of axonal regeneration. Nat. Cell Biol. 2001, 409, 341–346. [Google Scholar] [CrossRef]
  52. Moreno-Flores, M.T.; Avila, J. The Quest to Repair the Damaged Spinal Cord. Recent Patents CNS Drug Discov. 2012, 1, 55–63. [Google Scholar] [CrossRef] [Green Version]
  53. Oudega, M. Molecular and cellular mechanisms underlying the role of blood vessels in spinal cord injury and repair. Cell Tissue Res. 2012, 349, 269–288. [Google Scholar] [CrossRef] [PubMed]
  54. Ulndreaj, A.; Chio, J.C.; Ahuja, C.S.; Fehlings, M.G. Modulating the immune response in spinal cord injury. Expert Rev. Neurother. 2016, 16, 1127–1129. [Google Scholar] [CrossRef] [Green Version]
  55. Nakamura, M.; Houghtling, R.A.; MacArthur, L.; Bayer, B.M.; Bregman, B.S. Differences in cytokine gene expression profile between acute and secondary injury in adult rat spinal cord. Exp. Neurol. 2003, 184, 313–325. [Google Scholar] [CrossRef]
  56. Liu, M.; Wu, W.; Li, H.; Li, S.; Huang, L.-T.; Yang, Y.-Q.; Sun, Q.; Wang, C.-X.; Yu, Z.; Hang, C.-H. Necroptosis, a novel type of programmed cell death, contributes to early neural cells damage after spinal cord injury in adult mice. J. Spinal Cord Med. 2014, 38, 745–753. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, Y.; Wang, H.; Tao, Y.; Zhang, S.; Wang, J.; Feng, X. Necroptosis inhibitor necrostatin-1 promotes cell protection and physiological function in traumatic spinal cord injury. Neuroscience 2014, 266, 91–101. [Google Scholar] [CrossRef]
  58. Li, S.; Stys, P.K. Mechanisms of Ionotropic Glutamate Receptor-Mediated Excitotoxicity in Isolated Spinal Cord White Matter. J. Neurosci. 2000, 20, 1190–1198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Dong, H.; Fazzaro, A.; Xiang, C.; Korsmeyer, S.J.; Jacquin, M.F.; McDonald, J.W. Enhanced Oligodendrocyte Survival after Spinal Cord Injury in Bax-Deficient Mice and Mice with Delayed Wallerian Degeneration. J. Neurosci. 2003, 23, 8682–8691. [Google Scholar] [CrossRef]
  60. Crowe, M.J.; Bresnahan, J.C.; Shuman, S.L.; Masters, J.N.; Beattie, M.S. Apoptosis and delayed degeneration after spinal cord injury in rats and monkeys. Nat. Med. 1997, 3, 73–76. [Google Scholar] [CrossRef]
  61. Beattie, M.S.; Farooqui, A.A.; Bresnahan, J.C. Review of Current Evidence for Apoptosis After Spinal Cord Injury. J. Neurotrauma 2000, 17, 915–925. [Google Scholar] [CrossRef] [PubMed]
  62. Schanne, F.A.X.; Kane, A.B.; Young, E.E.; Farber, J.L. Calcium Dependence of Toxic Cell Death: A Final Common Pathway. Science 1979, 206, 700–702. [Google Scholar] [CrossRef]
  63. Garcia, E.; Aguilar-Cevallos, J.; Silva-Garcia, R.; Ibarra, A. Cytokine and Growth Factor Activation In Vivo and In Vitro after Spinal Cord Injury. Mediat. Inflamm. 2016, 2016, 9476020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gazdic, M.; Volarevic, V.; Harrell, C.R.; Fellabaum, C.; Jovicic, N.; Arsenijevic, N.; Stojkovic, M. Stem Cells Therapy for Spinal Cord Injury. Int. J. Mol. Sci. 2018, 19, 1039. [Google Scholar] [CrossRef] [Green Version]
  65. Fehlings, M.G.; Nakashima, H.; Nagoshi, N.; Chow, D.S.L.; Grossman, R.G.; Kopjar, B. Rationale, design and critical end points for the Riluzole in Acute Spinal Cord Injury Study (RISCIS): A randomized, double-blinded, placebo-controlled parallel multi-center trial. Spinal Cord 2015, 54, 8–15. [Google Scholar] [CrossRef] [Green Version]
  66. Kwon, B.K.; Streijger, F.; Fallah, N.; Noonan, V.K.; Bélanger, L.M.; Ritchie, L.; Paquette, S.J.; Ailon, T.; Boyd, M.C.; Street, J.; et al. Cerebrospinal Fluid Biomarkers To Stratify Injury Severity and Predict Outcome in Human Traumatic Spinal Cord Injury. J. Neurotrauma 2017, 34, 567–580. [Google Scholar] [CrossRef]
  67. Biglari, B.; Swing, T.; Child, C.; Büchler, A.; Westhauser, F.; Bruckner, T.; Ferbert, T.; Gerner, H.J.; Moghaddam, A. A pilot study on temporal changes in IL-1β and TNF-α serum levels after spinal cord injury: The serum level of TNF-α in acute SCI patients as a possible marker for neurological remission. Spinal Cord 2015, 53, 510–514. [Google Scholar] [CrossRef] [PubMed]
  68. Leal-Filho, M.B. Spinal cord injury: From inflammation to glial scar. Surg. Neurol. Int. 2011, 2, 112. [Google Scholar] [CrossRef] [Green Version]
  69. Bradbury, E.J.; Burnside, E.R. Moving beyond the glial scar for spinal cord repair. Nat. Commun. 2019, 10, 3879. [Google Scholar] [CrossRef]
  70. Silver, J.; Miller, J.H. Regeneration beyond the glial scar. Nat. Rev. Neurosci. 2004, 5, 146–156. [Google Scholar] [CrossRef]
  71. Orr, M.B.; Gensel, J.C. Spinal Cord Injury Scarring and Inflammation: Therapies Targeting Glial and Inflammatory Responses. Neurother. 2018, 15, 541–553. [Google Scholar] [CrossRef] [Green Version]
  72. Yuan, Y.-M.; He, C. The glial scar in spinal cord injury and repair. Neurosci. Bull. 2013, 29, 421–435. [Google Scholar] [CrossRef] [Green Version]
  73. Bartanusz, V.; Jezova, D.; Alajajian, B.; Digicaylioglu, M. The blood-spinal cord barrier: Morphology and Clinical Implications. Ann. Neurol. 2011, 70, 194–206. [Google Scholar] [CrossRef] [PubMed]
  74. Reinhold, A.K.; Rittner, H.L. Barrier function in the peripheral and central nervous system—A review. Pflügers Arch. Eur. J. Physiol. 2017, 469, 123–134. [Google Scholar] [CrossRef] [PubMed]
  75. Figley, S.A.; Khosravi, R.; Legasto, J.M.; Tseng, Y.-F.; Fehlings, M.G. Characterization of Vascular Disruption and Blood–Spinal Cord Barrier Permeability following Traumatic Spinal Cord Injury. J. Neurotrauma 2014, 31, 541–552. [Google Scholar] [CrossRef] [Green Version]
  76. Hu, J.; Yu, Q.; Xie, L.; Zhu, H. Targeting the blood-spinal cord barrier: A therapeutic approach to spinal cord protection against ischemia-reperfusion injury. Life Sci. 2016, 158, 1–6. [Google Scholar] [CrossRef]
  77. Jo, D.H.; Kim, J.H.; Heo, J.-I.; Kim, J.H.; Cho, C.-H. Interaction between pericytes and endothelial cells leads to formation of tight junction in hyaloid vessels. Mol. Cells 2013, 36, 465–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Zhou, K.; Shi, L.; Wang, Y.; Chen, S.; Zhang, J. Recent Advances of the NLRP3 Inflammasome in Central Nervous System Disorders. J. Immunol. Res. 2016, 2016, 9238290. [Google Scholar] [CrossRef]
  79. Martinon, F.; Mayor, A.; Tschopp, J. The Inflammasomes: Guardians of the Body. Annu. Rev. Immunol. 2009, 27, 229–265. [Google Scholar] [CrossRef] [Green Version]
  80. Davis, B.K.; Wen, H.; Ting, J.P.-Y. The Inflammasome NLRs in Immunity, Inflammation, and Associated Diseases. Annu. Rev. Immunol. 2011, 29, 707–735. [Google Scholar] [CrossRef] [Green Version]
  81. Jiang, W.; Li, M.; He, F.; Zhou, S.; Zhu, L. Targeting the NLRP3 inflammasome to attenuate spinal cord injury in mice. J. Neuroinflamm. 2017, 14, 207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Jiang, W.; Huang, Y.; Han, N.; He, F.; Li, M.; Bian, Z.; Liu, J.; Sun, T.; Zhu, L. Quercetin suppresses NLRP3 inflammasome activation and attenuates histopathology in a rat model of spinal cord injury. Spinal Cord 2016, 54, 592–596. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, G.; Shi, D.; Zhi, Z.; Ao, R.; Yu, B. Melatonin ameliorates spinal cord injury by suppressing the activation of inflammasomes in rats. J. Cell. Biochem. 2019, 120, 5183–5192. [Google Scholar] [CrossRef]
  84. Zendedel, A.; Mönnink, F.; Hassanzadeh, G.; Zaminy, A.; Ansar, M.M.; Habib, P.; Slowik, A.D.; Kipp, M.; Beyer, C. Estrogen Attenuates Local Inflammasome Expression and Activation after Spinal Cord Injury. Mol. Neurobiol. 2018, 55, 1364–1375. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, W.; Huang, Y.; He, F.; Liu, J.; Li, M.; Sun, T.; Ren, W.; Hou, J.; Zhu, L. Dopamine D1 Receptor Agonist A-68930 Inhibits NLRP3 Inflammasome Activation, Controls Inflammation, and Alleviates Histopathology in a Rat Model of Spinal Cord Injury. Spine 2016, 41, E330–E334. [Google Scholar] [CrossRef]
  86. Anderson, A.J.; Robert, S.; Huang, W.; Young, W.; Cotman, C.W. Activation of complement pathways after contusion-induced spinal cord injury. J. Neurotrauma 2004, 21, 1831–1846. [Google Scholar] [CrossRef] [PubMed]
  87. Lin, Z.; Lin, H.; Li, W.; Huang, Y.; Dai, H. Complement Component C3 Promotes Cerebral Ischemia/Reperfusion Injury Mediated by TLR2/NFκB Activation in Diabetic Mice. Neurochem. Res. 2018, 43, 1599–1607. [Google Scholar] [CrossRef]
  88. Fraser, D.A.; Arora, M.; Bohlson, S.S.; Lozano, E.; Tenner, A.J. Generation of inhibitory NFkappaB complexes and phosphorylated cAMP response element-binding protein correlates with the anti-inflammatory activity of complement protein C1q in human monocytes. J. Biol. Chem. 2007, 282, 7360–7367. [Google Scholar] [CrossRef] [Green Version]
  89. Guerrero, A.R.; Uchida, K.; Nakajima, H.; Watanabe, S.; Nakamura, M.; Johnson, W.E.; Baba, H. Blockade of interleukin-6 signaling inhibits the classic pathway and promotes an alternative pathway of macrophage activation after spinal cord injury in mice. J. Neuroinflamm. 2012, 9, 40. [Google Scholar] [CrossRef] [Green Version]
  90. Kigerl, K.; Gensel, J.C.; Ankeny, D.P.; Alexander, J.K.; Donnelly, D.J.; Popovich, P.G. Identification of Two Distinct Macrophage Subsets with Divergent Effects Causing either Neurotoxicity or Regeneration in the Injured Mouse Spinal Cord. J. Neurosci. 2009, 29, 13435–13444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Liddelow, S.A.; Barres, B.A. Reactive Astrocytes: Production, Function, and Therapeutic Potential. Immunity 2017, 46, 957–967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Martinez, F.O.; Gordon, S. The M1 and M2 paradigm of macrophage activation: Time for reassessment. F1000Prime Rep. 2014, 6, 13. [Google Scholar] [CrossRef] [Green Version]
  93. Sofroniew, M.V. Astrogliosis. Cold Spring Harb. Perspect. Biol. 2014, 7, a020420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Anderson, M.A.; Burda, J.E.; Ren, Y.; Ao, Y.; O’Shea, T.M.; Kawaguchi, R.; Coppola, G.; Khakh, B.S.; Deming, T.J.; Sofroniew, M.V. Astrocyte scar formation aids central nervous system axon regeneration. Nature 2016, 532, 195–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef] [PubMed]
  97. Bartel, D.P. MicroRNAs: Genomics, Biogenesis, Mechanism, and Function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  98. Akcakaya, P.; Akçakaya, P.; Ekelund, S.; Kolosenko, I.; Caramuta, S.; Özata, D.M.; Xie, H.; Lindforss, U.; Olivecrona, H.; Lui, W.-O. miR-185 and miR-133b deregulation is associated with overall survival and metastasis in colorectal cancer. Int. J. Oncol. 2011, 39, 311–318. [Google Scholar] [CrossRef] [Green Version]
  99. Xu, G.; Ao, R.; Zhi, Z.; Jia, J.; Yu, B. miR-21 and miR-19b delivered by hMSC-derived EVs regulate the apoptosis and differentiation of neurons in patients with spinal cord injury. J. Cell Physiol. 2019, 234, 10205–10217. [Google Scholar] [CrossRef]
  100. Han, Z.; Chen, F.; Ge, X.; Tan, J.; Lei, P.; Zhang, J. miR-21 alleviated apoptosis of cortical neurons through promoting PTEN-Akt signaling pathway in vitro after experimental traumatic brain injury. Brain Res. 2014, 1582, 12–20. [Google Scholar] [CrossRef]
  101. Liu, N.-K.; Wang, X.-F.; Lu, Q.-B.; Xu, X.-M. Altered microRNA expression following traumatic spinal cord injury. Exp. Neurol. 2009, 219, 424–429. [Google Scholar] [CrossRef] [Green Version]
  102. Xia, C.; Cai, Y.; Lin, Y.; Guan, R.; Xiao, G.; Yang, J. MiR-133b-5p regulates the expression of the heat shock protein 70 during rat neuronal cell apoptosis induced by the gp120 V3 loop peptide. J. Med. Virol. 2016, 88, 437–447. [Google Scholar] [CrossRef]
  103. Lu, X.C.; Zheng, J.Y.; Tang, L.J.; Huang, B.S.; Li, K.; Tao, Y.; Yu, Z.J.; Zhu, R.L.; Li, S.; Li, L.X. MiR-133b Promotes Neurite Outgrowth by Targeting RhoA Expression. Cell. Physiol. Biochem. 2015, 35, 246–258. [Google Scholar] [CrossRef]
  104. Heyer, M.P.; Pani, A.K.; Smeyne, R.J.; Kenny, P.J.; Feng, G. Normal midbrain dopaminergic neuron development and function in miR-133b mutant mice. J. Neurosci. 2012, 32, 10887–10894. [Google Scholar] [CrossRef] [Green Version]
  105. Yu, Y.-M.; Gibbs, K.M.; Davila, J.; Campbell, N.; Sung, S.; Todorova, T.I.; Otsuka, S.; Sabaawy, H.E.; Hart, R.P.; Schachner, M. MicroRNA miR-133b is essential for functional recovery after spinal cord injury in adult zebrafish. Eur. J. Neurosci. 2011, 33, 1587–1597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hu, J.; Zeng, L.; Huang, J.; Wang, G.; Lu, H. miR-126 promotes angiogenesis and attenuates inflammation after contusion spinal cord injury in rats. Brain Res. 2015, 1608, 191–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Groh, M.E.; Maitra, B.; Szekely, E.; Koç, O.N. Human mesenchymal stem cells require monocyte-mediated activation to suppress alloreactive T cells. Exp. Hematol. 2005, 33, 928–934. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, W.; Zhang, F.; Shi, H.; Tan, R.; Han, S.; Ye, G.; Pan, S.; Sun, F.; Liu, X. Comparisons of Rabbit Bone Marrow Mesenchymal Stem Cell Isolation and Culture Methods In Vitro. PLoS ONE 2014, 9, e88794. [Google Scholar] [CrossRef]
  109. Stewart, M.C.; Stewart, A.A. Mesenchymal Stem Cells: Characteristics, Sources, and Mechanisms of Action. Veter Clin. N. Am. Equine Pr. 2011, 27, 243–261. [Google Scholar] [CrossRef]
  110. Volarevic, V.; Markovic, B.S.; Gazdic, M.; Volarevic, A.; Jovicic, N.; Arsenijevic, N.; Armstrong, L.; Djonov, V.; Lako, M.; Stojkovic, M. Ethical and Safety Issues of Stem Cell-Based Therapy. Int. J. Med. Sci. 2018, 15, 36–45. [Google Scholar] [CrossRef] [Green Version]
  111. Ren, Z.; Qi, Y.; Sun, S.; Tao, Y.; Shi, R. Mesenchymal Stem Cell-Derived Exosomes: Hope for Spinal Cord Injury Repair. Stem Cells Dev. 2020, 29, 1467–1478. [Google Scholar] [CrossRef]
  112. Dominici, M.; Le Blanc, K.; Mueller, I.; Slaper-Cortenbach, I.; Marini, F.C.; Krause, D.S.; Deans, R.J.; Keating, A.; Prockop, D.J.; Horwitz, E.M. Minimal criteria for defining multipotent mesenchymal stromal cells. The International Society for Cellular Therapy position statement. Cytotherapy 2006, 8, 315–317. [Google Scholar] [CrossRef] [PubMed]
  113. Ullah, I.; Subbarao, R.B.; Rho, G.J. Human mesenchymal stem cells—Current trends and future prospective. Biosci. Rep. 2015, 35, 2. [Google Scholar] [CrossRef]
  114. Park, H.-W.; Cho, J.-S.; Park, C.-K.; Jung, S.J.; Park, C.-H.; Lee, S.-J.; Oh, S.B.; Park, Y.-S.; Chang, M.-S. Directed Induction of Functional Motor Neuron-Like Cells from Genetically Engineered Human Mesenchymal Stem Cells. PLoS ONE 2012, 7, e35244. [Google Scholar] [CrossRef]
  115. Tropel, P.; Platet, N.; Platel, J.; Noël, D.; Albrieux, M.; Benabid, A.; Berger, F. Functional Neuronal Differentiation of Bone Marrow-Derived Mesenchymal Stem Cells. Stem. Cells 2006, 24, 2868–2876. [Google Scholar] [CrossRef]
  116. Mezey, E.; Chandross, K.J.; Harta, G.; Maki, R.A.; McKercher, S.R. Turning Blood into Brain: Cells Bearing Neuronal Antigens Generated in Vivo from Bone Marrow. Science 2000, 290, 1779–1782. [Google Scholar] [CrossRef] [Green Version]
  117. Qi, X.; Shao, M.; Peng, H.; Bi, Z.; Su, Z.; Li, H. In vitro differentiation of bone marrow stromal cells into neurons and glial cells and differential protein expression in a two-compartment bone marrow stromal cell/neuron co-culture system. J. Clin. Neurosci. 2010, 17, 908–913. [Google Scholar] [CrossRef]
  118. Sanchez-Ramos, J.; Song, S.; Cardozo-Pelaez, F.; Hazzi, C.; Stedeford, T.; Willing, A.; Freeman, T.; Saporta, S.; Janssen, W.; Patel, N.A.; et al. Adult Bone Marrow Stromal Cells Differentiate into Neural Cells in Vitro. Exp. Neurol. 2000, 164, 247–256. [Google Scholar] [CrossRef] [Green Version]
  119. Lim, W.L.; Liau, L.L.; Ng, M.H.; Chowdhury, S.R.; Law, J.X. Current Progress in Tendon and Ligament Tissue Engineering. Tissue Eng. Regen. Med. 2019, 16, 549–571. [Google Scholar] [CrossRef]
  120. Shin, S.; Lee, J.; Kwon, Y.; Park, K.-S.; Jeong, J.-H.; Choi, S.-J.; Bang, S.I.; Chang, J.W.; Lee, C. Comparative Proteomic Analysis of the Mesenchymal Stem Cells Secretome from Adipose, Bone Marrow, Placenta and Wharton’s Jelly. Int. J. Mol. Sci. 2021, 22, 845. [Google Scholar] [CrossRef] [PubMed]
  121. Marconi, S.; Castiglione, G.; Turano, E.; Bissolotti, G.; Angiari, S.; Farinazzo, A.; Constantin, G.; Bedogni, G.; Bedogni, A.; Bonetti, B. Human Adipose-Derived Mesenchymal Stem Cells Systemically Injected Promote Peripheral Nerve Regeneration in the Mouse Model of Sciatic Crush. Tissue Eng. Part A 2012, 18, 1264–1272. [Google Scholar] [CrossRef]
  122. Uccelli, A.; Moretta, L.; Pistoia, V. Mesenchymal stem cells in health and disease. Nat. Rev. Immunol. 2008, 8, 726–736. [Google Scholar] [CrossRef] [PubMed]
  123. Rehman, J.; Traktuev, D.; Li, J.; Merfeld-Clauss, S.; Temm-Grove, C.J.; Bovenkerk, J.E.; Pell, C.L.; Johnstone, B.H.; Considine, R.V.; March, K.L. Secretion of Angiogenic and Antiapoptotic Factors by Human Adipose Stromal Cells. Circulation 2004, 109, 1292–1298. [Google Scholar] [CrossRef]
  124. Sorrell, J.M.; Baber, M.A.; Caplan, A.I. Influence of Adult Mesenchymal Stem Cells onIn VitroVascular Formation. Tissue Eng. Part A 2009, 15, 1751–1761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hofer, H.R.; Tuan, R.S. Secreted trophic factors of mesenchymal stem cells support neurovascular and musculoskeletal therapies. Stem Cell Res. Ther. 2016, 7, 131. [Google Scholar] [CrossRef] [Green Version]
  126. Lim, J.; Razi, Z.R.M.; Law, J.X.; Nawi, A.M.; Idrus, R.B.H.; Chin, T.G.; Mustangin, M.; Ng, M.H. Mesenchymal Stromal Cells from the Maternal Segment of Human Umbilical Cord is Ideal for Bone Regeneration in Allogenic Setting. Tissue Eng. Regen. Med. 2018, 15, 75–87. [Google Scholar] [CrossRef]
  127. Avanzini, M.A.; Bernardo, M.E.; Cometa, A.M.; Perotti, C.; Zaffaroni, N.; Novara, F.; Visai, L.; Moretta, A.; Del Fante, C.; Villa, R.; et al. Generation of mesenchymal stromal cells in the presence of platelet lysate: A phenotypic and functional comparison of umbilical cord blood- and bone marrow-derived progenitors. Haematologica 2009, 94, 1649–1660. [Google Scholar] [CrossRef]
  128. Weiss, M.L.; Anderson, C.; Medicetty, S.; Seshareddy, K.B.; Weiss, R.J.; VanderWerff, I.; Troyer, D.; McIntosh, K.R. Immune properties of human umbilical cord Wharton’s jelly-derived cells. Stem Cells 2008, 26, 2865–2874. [Google Scholar] [CrossRef] [PubMed]
  129. Zhao, Y.; Tang, F.; Xiao, Z.; Han, G.; Wang, N.; Yin, N.; Chen, B.; Jiang, X.; Yun, C.; Han, W.; et al. Clinical Study of NeuroRegen Scaffold Combined with Human Mesenchymal Stem Cells for the Repair of Chronic Complete Spinal Cord Injury. Cell Transplant. 2017, 26, 891–900. [Google Scholar] [CrossRef]
  130. Hafez, P.; Chowdhury, S.R.; Jose, S.; Law, J.X.; Ruszymah, B.H.I.; Ramzisham, A.R.M.; Ng, M.H. Development of an In Vitro Cardiac Ischemic Model Using Primary Human Cardiomyocytes. Cardiovasc. Eng. Technol. 2018, 9, 529–538. [Google Scholar] [CrossRef] [PubMed]
  131. Da Costa Gonçalves, F.; Grings, M.; Nunes, N.S.; Pinto, F.O.; Garcez, T.N.A.; Visioli, F.; Leipnitz, G.; Paz, A.H. Antioxidant properties of mesenchymal stem cells against oxidative stress in a murine model of colitis. Biotechnol. Lett. 2017, 39, 613–622. [Google Scholar] [CrossRef]
  132. Kemp, K.; Mallam, E.; Hares, K.; Witherick, J.; Scolding, N.; Wilkins, A. Mesenchymal Stem Cells Restore Frataxin Expression and Increase Hydrogen Peroxide Scavenging Enzymes in Friedreich Ataxia Fibroblasts. PLoS ONE 2011, 6, e26098. [Google Scholar] [CrossRef]
  133. Hofstetter, C.P.; Schwarz, E.J.; Hess, D.; Widenfalk, J.; El Manira, A.; Prockop, D.J.; Olson, L. Marrow stromal cells form guiding strands in the injured spinal cord and promote recovery. Proc. Natl. Acad. Sci. USA 2002, 99, 2199–2204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Charbord, P. Bone Marrow Mesenchymal Stem Cells: Historical Overview and Concepts. Hum. Gene Ther. 2010, 21, 1045–1056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Matyas, J.J.; Stewart, A.N.; Goldsmith, A.; Nan, Z.; Skeel, R.L.; Rossignol, J.; Dunbar, G.L. Effects of bone-marrow-derived MSC transplantation on functional recovery in a rat model of spinal cord injury: Comparisons of transplant locations and cell concentrations. Cell Transplant. 2017, 26, 1472–1482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Han, D.; Wu, C.; Xiong, Q.; Zhou, L.; Tian, Y. Anti-inflammatory Mechanism of Bone Marrow Mesenchymal Stem Cell Transplantation in Rat Model of Spinal Cord Injury. Cell Biophys. 2015, 71, 1341–1347. [Google Scholar] [CrossRef]
  137. Matsushita, T.; Lankford, K.L.; Arroyo, E.J.; Sasaki, M.; Neyazi, M.; Radtke, C.; Kocsis, J.D. Diffuse and persistent blood–spinal cord barrier disruption after contusive spinal cord injury rapidly recovers following intravenous infusion of bone marrow mesenchymal stem cells. Exp. Neurol. 2015, 267, 152–164. [Google Scholar] [CrossRef]
  138. Kim, C.; Kim, H.J.; Lee, H.; Lee, H.; Lee, S.J.; Lee, S.T.; Yang, S.R.; Chung, C.K. Mesenchymal Stem Cell Transplantation Promotes Functional Recovery through MMP2/STAT3 Related Astrogliosis after Spinal Cord Injury. Int. J. Stem. Cells 2019, 12, 331–339. [Google Scholar] [CrossRef] [Green Version]
  139. Nandoe, R.D.S.; Hurtado, A.; Levi, A.D.O.; Grotenhuis, J.A.; Oudega, M. Bone Marrow Stromal Cells for Repair of the Spinal Cord: Towards Clinical Application. Cell Transplant. 2006, 15, 563–577. [Google Scholar] [CrossRef] [Green Version]
  140. Cížková, D.; Novotna, I.; Slovinska, L.; Vanický, I.; Jergová, S.; Rosocha, J.; Radoňak, J. Repetitive Intrathecal Catheter Delivery of Bone Marrow Mesenchymal Stromal Cells Improves Functional Recovery in a Rat Model of Contusive Spinal Cord Injury. J. Neurotrauma 2011, 28, 1951–1961. [Google Scholar] [CrossRef]
  141. Osaka, M.; Honmou, O.; Murakami, T.; Nonaka, T.; Houkin, K.; Hamada, H.; Kocsis, J.D. Intravenous administration of mesenchymal stem cells derived from bone marrow after contusive spinal cord injury improves functional outcome. Brain Res. 2010, 1343, 226–235. [Google Scholar] [CrossRef] [PubMed]
  142. Čížková, D.; Rosocha, J.; Vanický, I.; Jergová, S.; Cizek, M. Transplants of Human Mesenchymal Stem Cells Improve Functional Recovery After Spinal Cord Injury in the Rat. Cell. Mol. Neurobiol. 2006, 26, 1165–1178. [Google Scholar] [CrossRef] [PubMed]
  143. Liu, F.; Xuan, A.; Chen, Y.; Zhang, J.; Xu, L.; Yan, Q.; Long, D. Combined effect of nerve growth factor and brain-derived neurotrophic factor on neuronal differentiation of neural stem cells and the potential molecular mechanisms. Mol. Med. Rep. 2014, 10, 1739–1745. [Google Scholar] [CrossRef] [Green Version]
  144. Hoeben, A.; Landuyt, B.; Highley, M.S.; Wildiers, H.; Van Oosterom, A.T.; De Bruijn, E.A. Vascular endothelial growth factor and angiogenesis. Pharmacol. Rev. 2004, 56, 549–580. [Google Scholar] [CrossRef]
  145. Jeon, S.R.; Park, J.H.; Lee, J.H.; Kim, D.Y.; Kim, H.S.; Sung, I.Y.; Choi, G.H.; Jeon, M.H.; Kim, G.G. Treatment of Spinal Cord Injury with Bone Marrow-Derived, Cultured Autologous Mesenchymal Stem Cells. Tissue Eng. Regen Med. 2010, 7, 316–322. [Google Scholar]
  146. Saito, F.; Nakatani, T.; Iwase, M.; Maeda, Y.; Murao, Y.; Suzuki, Y.; Fukushima, M.; Ide, C. Administration of cultured autologous bone marrow stromal cells into cerebrospinal fluid in spinal injury patients: A pilot study. Restor. Neurol. Neurosci. 2012, 30, 127–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. El-Kheir, W.A.; Gabr, H.; Awad, M.R.; Ghannam, O.; Barakat, Y.; Farghali, H.A.M.A.; El Maadawi, Z.M.; Ewes, I.; Sabaawy, H.E. Autologous Bone Marrow-Derived Cell Therapy Combined with Physical Therapy Induces Functional Improvement in Chronic Spinal Cord Injury Patients. Cell Transplant. 2014, 23, 729–745. [Google Scholar] [CrossRef] [Green Version]
  148. Karamouzian, S.; Nematollahi-Mahani, S.N.; Nakhaee, N.; Eskandary, H. Clinical safety and primary efficacy of bone marrow mesenchymal cell transplantation in subacute spinal cord injured patients. Clin. Neurol. Neurosurg. 2012, 114, 935–939. [Google Scholar] [CrossRef]
  149. Jiang, P.-C.; Xiong, W.-P.; Wang, G.; Ma, C.; Yao, W.-Q.; Kendell, S.F.; Mehling, B.M.; Yuan, X.-H.; Wu, D.-C. A clinical trial report of autologous bone marrow-derived mesenchymal stem cell transplantation in patients with spinal cord injury. Exp. Ther. Med. 2013, 6, 140–146. [Google Scholar] [CrossRef] [Green Version]
  150. Dai, G.; Liu, X.; Zhang, Z.; Yang, Z.; Dai, Y.; Xu, R. Transplantation of autologous bone marrow mesenchymal stem cells in the treatment of complete and chronic cervical spinal cord injury. Brain Res. 2013, 1533, 73–79. [Google Scholar] [CrossRef]
  151. Mendonça, M.V.P.; Larocca, T.F.; Souza, B.S.D.F.; Villarreal, C.F.; Silva, L.F.M.; Matos, A.C.; Novaes, M.A.; Bahia, C.M.P.; Martinez, A.C.D.O.M.; Kaneto, C.M.; et al. Safety and neurological assessments after autologous transplantation of bone marrow mesenchymal stem cells in subjects with chronic spinal cord injury. Stem Cell Res. Ther. 2014, 5, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Pal, R.; Venkataramana, N.K.; Bansal, A.; Balaraju, S.; Jan, M.; Chandra, R.; Dixit, A.; Rauthan, A.; Murgod, U.; Totey, S. Ex vivo-expanded autologous bone marrow-derived mesenchymal stromal cells in human spinal cord injury/paraplegia: A pilot clinical study. Cytotherapy 2009, 11, 897–911. [Google Scholar] [CrossRef] [PubMed]
  153. Liau, L.; Ruszymah, B.; Ng, M.; Law, J. Characteristics and clinical applications of Wharton’s jelly-derived mesenchymal stromal cells. Curr. Res. Transl. Med. 2020, 68, 5–16. [Google Scholar] [CrossRef] [PubMed]
  154. Barry, F.P.; Murphy, J.M.; English, K.; Mahon, B.P. Immunogenicity of Adult Mesenchymal Stem Cells: Lessons from the Fetal Allograft. Stem Cells Dev. 2005, 14, 252–265. [Google Scholar] [CrossRef] [PubMed]
  155. Goodwin, H.; Bicknese, A.; Chien, S.-N.; Bogucki, B.; Oliver, D.; Quinn, C.; Wall, D. Multilineage differentiation activity by cells isolated from umbilical cord blood: Expression of bone, fat, and neural markers. Biol. Blood Marrow Transplant. 2001, 7, 581–588. [Google Scholar] [CrossRef] [Green Version]
  156. Sabapathy, V.; Sundaram, B.; Sreelakshmi, V.M.; Mankuzhy, P.; Kumar, S. Human Wharton’s Jelly Mesenchymal Stem Cells Plasticity Augments Scar-Free Skin Wound Healing with Hair Growth. PLoS ONE 2014, 9, e93726. [Google Scholar] [CrossRef]
  157. Nagamura-Inoue, T.; He, H. Umbilical cord-derived mesenchymal stem cells: Their advantages and potential clinical utility. World J. Stem. Cells 2014, 6, 195–202. [Google Scholar] [CrossRef]
  158. Kim, D.-W.; Staples, M.; Shinozuka, K.; Pantcheva, P.; Kang, S.-D.; Borlongan, C.V. Wharton’s Jelly-Derived Mesenchymal Stem Cells: Phenotypic Characterization and Optimizing Their Therapeutic Potential for Clinical Applications. Int. J. Mol. Sci. 2013, 14, 11692–11712. [Google Scholar] [CrossRef] [Green Version]
  159. De Girolamo, L.; Lucarelli, E.; Alessandri, G.; Avanzini, M.A.; Bernardo, M.E.; Biagi, E.; Brini, A.T.; D’Amico, G.; Fagioli, F.; Ferrero, I.; et al. Mesenchymal stem/stromal cells: A new “cells as drugs” paradigm. Efficacy and critical aspects in cell therapy. Curr. Pharm. Des. 2013, 19, 2459–2473. [Google Scholar] [CrossRef] [Green Version]
  160. Dasari, V.R.; Veeravalli, K.K.; Tsung, A.J.; Gondi, C.S.; Gujrati, M.; Dinh, D.H.; Rao, J.S. Neuronal apoptosis is inhibited by cord blood stem cells after spinal cord injury. J. Neurotrauma 2009, 26, 2057–2069. [Google Scholar] [CrossRef]
  161. Veeravalli, K.K.; Dasari, V.R.; Tsung, A.J.; Dinh, D.H.; Gujrati, M.; Fassett, D.; Rao, J.S. Human umbilical cord blood stem cells upregulate matrix metalloproteinase-2 in rats after spinal cord injury. Neurobiol. Dis. 2009, 36, 200–212. [Google Scholar] [CrossRef]
  162. Judas, G.I.; Ferreira, S.G.; Simas, R.; Sannomiya, P.; Benício, A.; Da Silva, L.F.F.; Moreira, L.F.P. Intrathecal injection of human umbilical cord blood stem cells attenuates spinal cord ischaemic compromise in rats. Interact. Cardiovasc. Thorac. Surg. 2014, 18, 757–762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Kaner, T.; Karadag, T.; Cirak, B.; Erken, H.A.; Karabulut, A.; Kiroglu, Y.; Akkaya, S.; Acar, F.; Coskun, E.; Genc, O.; et al. The effects of human umbilical cord blood transplantation in rats with experimentally induced spinal cord injury. J. Neurosurg. Spine 2010, 13, 543–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Dasari, V.R.; Spomar, D.G.; Li, L.; Gujrati, M.; Rao, J.S.; Dinh, D.H. Umbilical Cord Blood Stem Cell Mediated Downregulation of Fas Improves Functional Recovery of Rats after Spinal Cord Injury. Neurochem. Res. 2008, 33, 134–149. [Google Scholar] [CrossRef] [PubMed]
  165. Cui, B.; Li, E.; Yang, B.; Wang, B. Human umbilical cord blood-derived mesenchymal stem cell transplantation for the treatment of spinal cord injury. Exp. Ther. Med. 2014, 7, 1233–1236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Hu, S.-L.; Luo, H.-S.; Li, J.-T.; Xia, Y.-Z.; Li, L.; Zhang, L.-J.; Meng, H.; Cui, G.-Y.; Chen, Z.; Wu, N.; et al. Functional recovery in acute traumatic spinal cord injury after transplantation of human umbilical cord mesenchymal stem cells. Crit. Care Med. 2010, 38, 2181–2189. [Google Scholar] [CrossRef] [PubMed]
  167. Liu, J.; Han, D.; Wang, Z.; Xue, M.; Zhu, L.; Yan, H.; Zheng, X.; Guo, Z.; Wang, H. Clinical analysis of the treatment of spinal cord injury with umbilical cord mesenchymal stem cells. Cytotherapy 2013, 15, 185–191. [Google Scholar] [CrossRef]
  168. Cheng, H.; Liu, X.; Hua, R.; Dai, G.; Wang, X.; Gao, J.; An, Y. Clinical observation of umbilical cord mesenchymal stem cell transplantation in treatment for sequelae of thoracolumbar spinal cord injury. J. Transl. Med. 2014, 12, 253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Kang, K.-S.; Kim, S.; Oh, Y.; Yu, J.; Kim, K.-Y.; Park, H.; Song, C.-H.; Han, H.J. A 37-year-old spinal cord-injured female patient, transplanted of multipotent stem cells from human UC blood, with improved sensory perception and mobility, both functionally and morphologically: A case study. Cytotherapy 2005, 7, 368–373. [Google Scholar] [CrossRef] [PubMed]
  170. Zuk, P.A.; Zhu, M.; Mizuno, H.; Huang, J.; Futrell, J.W.; Katz, A.J.; Benhaim, P.; Lorenz, H.P.; Hedrick, M.H. Multilineage Cells from Human Adipose Tissue: Implications for Cell-Based Therapies. Tissue Eng. 2001, 7, 211–228. [Google Scholar] [CrossRef] [Green Version]
  171. De Ugarte, D.A.; Morizono, K.; Elbarbary, A.; Alfonso, Z.; Zuk, P.A.; Zhu, M.; Dragoo, J.L.; Ashjian, P.; Thomas, B.; Benhaim, P.; et al. Comparison of Multi-Lineage Cells from Human Adipose Tissue and Bone Marrow. Cells Tissues Organs 2003, 174, 101–109. [Google Scholar] [CrossRef] [PubMed]
  172. Danisovic, L.; Varga, I.; Polák, S.; Ulicná, M.; Hlavacková, L.; Böhmer, D.; Vojtassák, J. Comparison of in vitro chondrogenic potential of human mesenchymal stem cells derived from bone marrow and adipose tissue. Gen. Physiol. Biophys. 2009, 28, 56–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Hsiao, S.; Asgari, A.; Lokmic, Z.T.; Sinclair, R.; Dusting, G.J.; Lim, S.Y.; Dilley, R.J. Comparative Analysis of Paracrine Factor Expression in Human Adult Mesenchymal Stem Cells Derived from Bone Marrow, Adipose, and Dermal Tissue. Stem Cells Dev. 2012, 21, 2189–2203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Ahmadian Kia, N.; Bahrami, A.R.; Ebrahimi, M.; Matin, M.M.; Neshati, Z.; Almohaddesin, M.R.; Aghdami, N.; Bidkhori, H.R. Comparative analysis of chemokine receptor’s expression in mesenchymal stem cells derived from human bone marrow and adipose tissue. J. Mol. Neurosci. 2011, 44, 178–185. [Google Scholar] [CrossRef] [PubMed]
  175. Ohta, Y.; Hamaguchi, A.; Ootaki, M.; Watanabe, M.; Takeba, Y.; Iiri, T.; Matsumoto, N.; Takenaga, M. Intravenous infusion of adipose-derived stem/stromal cells improves functional recovery of rats with spinal cord injury. Cytotherapy 2017, 19, 839–848. [Google Scholar] [CrossRef] [Green Version]
  176. Aras, Y.; Sabanci, P.A.; Kabatas, S.; Duruksu, G.; Subasi, C.; Erguven, M.; Karaoz, E. The Effects of Adipose Tissue-Derived Mesenchymal Stem Cell Transplantation During the Acute and Subacute Phases Following Spinal Cord Injury. Turk. Neurosurg. 2015, 26, 127–139. [Google Scholar] [CrossRef] [Green Version]
  177. Zhou, Z.; Tian, X.; Mo, B.; Xu, H.; Zhang, L.; Huang, L.; Yao, S.; Huang, Z.; Wang, Y.; Xie, H.; et al. Adipose mesenchymal stem cell transplantation alleviates spinal cord injury-induced neuroinflammation partly by suppressing the Jagged1/Notch pathway. Stem. Cell Res. Ther. 2020, 11, 212. [Google Scholar] [CrossRef]
  178. Leu, S.; Lin, Y.-C.; Yuen, C.-M.; Yen, C.-H.; Kao, Y.-H.; Sun, C.-K.; Yip, H.-K. Adipose-derived mesenchymal stem cells markedly attenuate brain infarct size and improve neurological function in rats. J. Transl. Med. 2010, 8, 63. [Google Scholar] [CrossRef] [Green Version]
  179. Gao, S.; Guo, X.; Zhao, S.; Jin, Y.; Zhou, F.; Yuan, P.; Cao, L.; Wang, J.; Qiu, Y.; Sun, C.; et al. Differentiation of human adipose-derived stem cells into neuron/motoneuron-like cells for cell replacement therapy of spinal cord injury. Cell Death Dis. 2019, 10, 597. [Google Scholar] [CrossRef]
  180. Ra, J.C.; Shin, I.S.; Kim, S.H.; Kang, S.K.; Kang, B.C.; Lee, H.Y.; Kim, Y.J.; Jo, J.Y.; Yoon, E.J.; Choi, H.J.; et al. Safety of Intravenous Infusion of Human Adipose Tissue-Derived Mesenchymal Stem Cells in Animals and Humans. Stem Cells Dev. 2011, 20, 1297–1308. [Google Scholar] [CrossRef]
  181. Hur, J.W.; Cho, T.-H.; Park, D.-H.; Lee, J.-B.; Park, J.-Y.; Chung, Y.-G. Intrathecal transplantation of autologous adipose-derived mesenchymal stem cells for treating spinal cord injury: A human trial. J. Spinal Cord Med. 2015, 39, 655–664. [Google Scholar] [CrossRef] [PubMed]
  182. Bydon, M.; Dietz, A.B.; Goncalves, S.; Moinuddin, F.M.; Alvi, M.A.; Goyal, A.; Yolcu, Y.; Hunt, C.L.; Garlanger, K.L.; Del Fabro, A.S.; et al. CELLTOP Clinical Trial: First Report From a Phase 1 Trial of Autologous Adipose Tissue–Derived Mesenchymal Stem Cells in the Treatment of Paralysis Due to Traumatic Spinal Cord Injury. Mayo Clin. Proc. 2020, 95, 406–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Heldring, N.; Mäger, I.; Wood, M.J.; Le Blanc, K.; Andaloussi, S.E. Therapeutic Potential of Multipotent Mesenchymal Stromal Cells and Their Extracellular Vesicles. Hum. Gene Ther. 2015, 26, 506–517. [Google Scholar] [CrossRef]
  184. Kretlow, J.D.; Jin, Y.-Q.; Liu, W.; Zhang, W.J.; Hong, T.-H.; Zhou, G.; Baggett, L.S.; Mikos, A.G.; Cao, Y. Donor age and cell passage affects differentiation potential of murine bone marrow-derived stem cells. BMC Cell Biol. 2008, 9, 60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Kim, G.; Shon, O.-J.; Seo, M.-S.; Choi, Y.; Park, W.; Lee, G. Mesenchymal Stem Cell-Derived Exosomes and Their Therapeutic Potential for Osteoarthritis. Biology 2021, 10, 285. [Google Scholar] [CrossRef]
  186. Lee, M.J.; Kim, J.; Kim, M.Y.; Bae, Y.-S.; Ryu, S.H.; Lee, T.G.; Kim, J.H. Proteomic Analysis of Tumor Necrosis Factor-α-Induced Secretome of Human Adipose Tissue-Derived Mesenchymal Stem Cells. J. Proteome Res. 2010, 9, 1754–1762. [Google Scholar] [CrossRef]
  187. Familtseva, A.; Jeremic, N.; Tyagi, S.C. Exosomes: Cell-created drug delivery systems. Mol. Cell. Biochem. 2019, 459, 1–6. [Google Scholar] [CrossRef]
  188. Barile, L.; Vassalli, G. Exosomes: Therapy delivery tools and biomarkers of diseases. Pharmacol. Ther. 2017, 174, 63–78. [Google Scholar] [CrossRef] [Green Version]
  189. Rani, S.; Ryan, A.E.; Griffin, M.D.; Ritter, T. Mesenchymal Stem Cell-derived Extracellular Vesicles: Toward Cell-free Therapeutic Applications. Mol. Ther. 2015, 23, 812–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Sarko, D.K.; McKinney, C.E. Exosomes: Origins and Therapeutic Potential for Neurodegenerative Disease. Front. Neurosci. 2017, 11, 82. [Google Scholar] [CrossRef] [Green Version]
  191. Van der Pol, E.; Böing, A.N.; Harrison, P.; Sturk, A.; Nieuwland, R. Classification, functions, and clinical relevance of extracellular vesicles. Pharmacol. Rev. 2012, 64, 676–705. [Google Scholar] [CrossRef] [Green Version]
  192. Akers, J.C.; Gonda, D.; Kim, R.; Carter, B.S.; Chen, C.C. Biogenesis of extracellular vesicles (EV): Exosomes, microvesicles, retrovirus-like vesicles, and apoptotic bodies. J. Neuro-Oncol. 2013, 113, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Musiał-Wysocka, A.; Kot, M.; Majka, M. The Pros and Cons of Mesenchymal Stem Cell-Based Therapies. Cell Transplant. 2019, 28, 801–812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Lee, G.W.; Seo, M.-S.; Kang, K.-K.; Oh, S.-K. Epidural fat-derived mesenchymal stem cell: First report of epidural fat-derived mesenchymal stem cell. Asian Spine J. 2019, 13, 361. [Google Scholar] [CrossRef] [Green Version]
  195. Wahid, F.; Shehzad, A.; Khan, T.; Kim, Y.Y. MicroRNAs: Synthesis, mechanism, function, and recent clinical trials. Biochim. Biophys. Acta (BBA) Bioenerg. 2010, 1803, 1231–1243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Wang, L.; Zhang, L. Circulating Exosomal miRNA as Diagnostic Biomarkers of Neurodegenerative Diseases. Front. Mol. Neurosci. 2020, 13, 53. [Google Scholar] [CrossRef] [Green Version]
  197. Chen, J.-J.; Zhao, B.; Zhao, J.; Li, S. Potential Roles of Exosomal MicroRNAs as Diagnostic Biomarkers and Therapeutic Application in Alzheimer’s Disease. Neural Plast. 2017, 2017, 7027380. [Google Scholar] [CrossRef] [PubMed]
  198. Winkler, C.W.; Taylor, K.G.; Peterson, K.E. Location Is Everything: Let-7b microRNA and TLR7 Signaling Results in a Painful TRP. Sci. Signal. 2014, 7, pe14. [Google Scholar] [CrossRef] [PubMed]
  199. Lehmann, S.M.; Krüger, C.; Park, B.; Derkow, K.; Rosenberger, K.; Baumgart, J.; Trimbuch, T.; Eom, G.; Hinz, M.; Kaul, D.; et al. An unconventional role for miRNA: Let-7 activates Toll-like receptor 7 and causes neurodegeneration. Nat. Neurosci. 2012, 15, 827–835. [Google Scholar] [CrossRef]
  200. René, C.; Parks, R. Delivery of Therapeutic Agents to the Central Nervous System and the Promise of Extracellular Vesicles. Pharmaceutics 2021, 13, 492. [Google Scholar] [CrossRef]
  201. Yang, T.; Martin, P.; Fogarty, B.; Brown, A.; Schurman, K.; Phipps, R.; Yin, V.P.; Lockman, P.; Bai, S. Exosome Delivered Anticancer Drugs Across the Blood-Brain Barrier for Brain Cancer Therapy in Danio Rerio. Pharm. Res. 2015, 32, 2003–2014. [Google Scholar] [CrossRef] [PubMed]
  202. Didiot, M.-C.; Hall, L.M.; Coles, A.H.; Haraszti, R.A.; Godinho, B.M.; Chase, K.; Sapp, E.; Ly, S.; Alterman, J.F.; Hassler, M.R.; et al. Exosome-mediated Delivery of Hydrophobically Modified siRNA for Huntingtin mRNA Silencing. Mol. Ther. 2016, 24, 1836–1847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Xin, H.; Wang, F.; Li, Y.; Lu, Q.-E.; Cheung, W.L.; Zhang, Y.; Zhang, Z.G.; 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 Transplant. 2017, 26, 243–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Yang, J.; Zhang, X.; Chen, X.; Wang, L.; Yang, G. Exosome Mediated Delivery of miR-124 Promotes Neurogenesis after Ischemia. Mol. Ther. Nucleic Acids 2017, 7, 278–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Yuyama, K.; Sun, H.; Sakai, S.; Mitsutake, S.; Okada, M.; Tahara, H.; Furukawa, J.-I.; Fujitani, N.; Shinohara, Y.; Igarashi, Y. Decreased Amyloid-β Pathologies by Intracerebral Loading of Glycosphingolipid-enriched Exosomes in Alzheimer Model Mice. J. Biol. Chem. 2014, 289, 24488–24498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Guo, S.; Perets, N.; Betzer, O.; Ben-Shaul, S.; Sheinin, A.; Michaelevski, I.; Popovtzer, R.; Offen, D.; Levenberg, S. Intranasal Delivery of Mesenchymal Stem Cell Derived Exosomes Loaded with Phosphatase and Tensin Homolog siRNA Repairs Complete Spinal Cord Injury. ACS Nano 2019, 13, 10015–10028. [Google Scholar] [CrossRef]
  207. Kim, H.Y.; Kumar, H.; Jo, M.-J.; Kim, J.; Yoon, J.-K.; Lee, J.-R.; Kang, M.; Choo, Y.W.; Song, S.Y.; Kwon, S.P.; et al. Therapeutic Efficacy-Potentiated and Diseased Organ-Targeting Nanovesicles Derived from Mesenchymal Stem Cells for Spinal Cord Injury Treatment. Nano Lett. 2018, 18, 4965–4975. [Google Scholar] [CrossRef]
  208. Zhong, D.; Cao, Y.; Li, C.-J.; Li, M.; Rong, Z.-J.; Jiang, L.; Guo, Z.; Lu, H.-B.; Hu, J.-Z. Neural stem cell-derived exosomes facilitate spinal cord functional recovery after injury by promoting angiogenesis. Exp. Biol. Med. 2020, 245, 54–65. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram for damage stages and responses in spinal cord injury.
Figure 1. Schematic diagram for damage stages and responses in spinal cord injury.
Ijms 22 13672 g001
Figure 2. Transmission electron microscopy (TEM) image of MSC-derived EV morphology.
Figure 2. Transmission electron microscopy (TEM) image of MSC-derived EV morphology.
Ijms 22 13672 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, G.-U.; Sung, S.-E.; Kang, K.-K.; Choi, J.-H.; Lee, S.; Sung, M.; Yang, S.Y.; Kim, S.-K.; Kim, Y.I.; Lim, J.-H.; et al. Therapeutic Potential of Mesenchymal Stem Cells (MSCs) and MSC-Derived Extracellular Vesicles for the Treatment of Spinal Cord Injury. Int. J. Mol. Sci. 2021, 22, 13672. https://doi.org/10.3390/ijms222413672

AMA Style

Kim G-U, Sung S-E, Kang K-K, Choi J-H, Lee S, Sung M, Yang SY, Kim S-K, Kim YI, Lim J-H, et al. Therapeutic Potential of Mesenchymal Stem Cells (MSCs) and MSC-Derived Extracellular Vesicles for the Treatment of Spinal Cord Injury. International Journal of Molecular Sciences. 2021; 22(24):13672. https://doi.org/10.3390/ijms222413672

Chicago/Turabian Style

Kim, Gang-Un, Soo-Eun Sung, Kyung-Ku Kang, Joo-Hee Choi, Sijoon Lee, Minkyoung Sung, Seung Yun Yang, Seul-Ki Kim, Young In Kim, Ju-Hyeon Lim, and et al. 2021. "Therapeutic Potential of Mesenchymal Stem Cells (MSCs) and MSC-Derived Extracellular Vesicles for the Treatment of Spinal Cord Injury" International Journal of Molecular Sciences 22, no. 24: 13672. https://doi.org/10.3390/ijms222413672

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

Article Metrics

Back to TopTop