Skip to main content

Potential pre-activation strategies for improving therapeutic efficacy of mesenchymal stem cells: current status and future prospects

Abstract

Mesenchymal stem cell (MSC)-based therapy has been considered as a promising approach targeting a variety of intractable diseases due to remarkable multiple effect of MSCs, such as multilineage differentiation, immunomodulatory property, and pro-regenerative capacity. However, poor engraftment, low survival rate of transplanted MSC, and impaired donor-MSC potency under host age/disease result in unsatisfactory therapeutic outcomes. Enhancement strategies, including genetic manipulation, pre-activation, and modification of culture method, have been investigated to generate highly functional MSC, and approaches for MSC pre-activation are highlighted. In this review, we summarized the current approaches of MSC pre-activation and further classified, analysed the scientific principles and main characteristics of these manipulations, and described the pros and cons of individual pre-activation strategies. We also discuss the specialized tactics to solve the challenges in this promising field so that it improves MSC therapeutic functions to serve patients better.

Introduction

In recent decades, MSCs in cell-based therapy have spanned across various diseases in experimental and clinical researches worldwide, exhibiting therapeutic efficacy over conventional treatments due to their distinctive biological properties [1,2,3,4,5]. They have isolated from perinatal tissues, such as umbilical cord, umbilical cord blood and placenta, and multiple biological tissues in adults, including bone marrow, adipose tissue, muscle, and lung [6, 7]. MSCs, as a kind of multipotent stromal cells, possess the potential for self-renewal and multilineage differentiation into adipocytes, muscles, chondrocytes, osteoblasts, and neuronal cells [8, 9]. In addition, increasing evidence has revealed that MSCs exert immunomodulation, reparative, and regenerative effects through high paracrine activity [10,11,12] (Fig. 1). More importantly, MSCs are immune privileged, which means allogeneic MSCs transplantation will not elicit inflammatory response, mainly due to their lack of class-II major histocompatibility complex (MHC-II) and costimulatory molecules [13, 14]. Their outstanding features jointly make MSCs the ideal seed cells in cell therapy after haematopoietic stem cells.

Fig. 1
figure 1

Isolation sources and action mechanisms of MSCs. The diagram illustrates various sources of isolated MSCs and the biological characteristics related to the therapeutic effect

However, MSCs from different individuals are heterogeneous in their biological effects. Moreover, the body's internal environment (sick or not, youth or old) affects the quality of the isolated stem cells. For example, adipose-derived MSCs (ASCs) isolated from obese and type 2 diabetes (T2D) individuals exhibit functional defectives, such as increased apoptosis, reduced immunosuppressive activities, and loss of stemness [15,16,17]. And MSCs from old donors show impairment of proliferation and differentiation, depression of immunoregulation, and reduced secretion of bioactive molecules [18]. In addition, once administered in the body, MSCs may undergo apoptosis shortly because they exposed to the harsh host microenvironment, including hypoxia, oxidative stress as well as chronic inflammation. It reported that only about 28% of the intravenously injected MSCs survived after one day, [19] and fewer than 1% of cells persisted more than a week [20, 21]. Even if transplanted in situ, most MSCs lose their biological activities within one week [22]. Besides, most of the infused MSCs trapped in the lung microvasculature instead of the target tissues [23]. These adverse conditions will cause various problems, such as low survival rates of the transplanted cells, poor migration and homing of MSCs, and limit the functionalities of the injected cells.

To achieve the desired therapeutic potential, it seems unreasonable to increase the dosage and frequency of transplanted MSCs, as this may increase the risk of pulmonary embolism and the cost [24]. Optimizing the potency and therapeutic benefits of MSCs is a top priority. Several strategies attempted to optimize stem cells were proposed, which roughly divided into two categories, namely genetic modification and non-genetic modification (pre-activation).

In terms of genetic modification, MSCs will produce or overexpress functional genes that enable them to resist hostile microenvironment and apoptosis, increase migration and homing, and enhance paracrine effects. Several studies suggest that gene transfected MSCs have better therapeutic potential than wild-type MSCs [25,26,27]. However, safety is the greatest barrier for the future clinical therapeutic use of genetically modified MSCs. It reported that viral expression systems can elicit immune and inflammatory responses in the host, and viral integration in the host genome poses a tumourigenic risk [28, 29]. Additionally, the therapeutic potential and long-term function improvements in genetically engineered MSCs need to fully elucidate. Therefore, the development of highly efficient non-genetic modification methods, collectively referred to here as pre-activation, is an alternative and operational way to improve the treatment outcome of MSCs.

MSCs can be pre-activated to achieve the desired function and reverse their inactivation because they can recognize the stimuli in the microenvironment and remember them [3, 30]. Reviewing the current literature, the pre-activation of MSCs is mainly based on the in vivo physiological microenvironment of MSCs survival and simulated in vitro, which is called "physiological microenvironment simulation pre-activation." Or in vitro adaptive regulation of MSCs is based on the pathological microenvironment of the disease, known as "pathological microenvironment simulation pre-activation." The primary goal of the review article is to provide specific methods involving in both types of pre-activation (Fig. 2).

Fig. 2
figure 2

An overview of currently available strategies for preconditioning MSCs to improve the beneficial therapeutic effects of MSCs

Methods

The most relevant English papers were selected by the distinct keywords, including mesenchymal stem cell and pretreatment, in the database of Google Scholar, MEDLINE, PubMed, and Embase. We set dates of searching from 2010 to 2021. An initial repeat assessment of selected articles was performed using endonote. Titles, abstracts, and full-text articles are further screened independently by two reviewers, and articles relevant to the topic of the current review were included. We also studied the references of the relevant papers, according to the needs of the writing.

Physiological microenvironment simulation pre-activation

The number of MSCs in primary culture is limited. It needs to expand in vitro to achieve therapeutically relevant cell numbers, and excessive ex vivo manipulation leads to senescence, decreased stemness, and impaired regenerative capacity [23]. Maintaining the “youthfulness” of MSCs in vitro is very important. Stem cells live in specific areas of tissues, named as stem cell niche. It is a multidimensional environment composed of both cellular and acellular components to manipulate stem cell proliferation, determine stem cell fate, and maintain stem cell homeostasis. The cellular and acellular components involved in a number of necessary clues, namely other cells (cell–cell interactions), physical elements (temperature, osmotic pressure, stretch, and electrical signals), chemical factors (PH, oxygen, nutrients, ionic strength, metabolites cytokines, and chemokines), and extracellular matrix (composition, structure, topology, and stiffness) [31,32,33]. In this regard, recreating the complex in vivo microenvironment in vitro would provide a powerful tool for MSC production and maintain their inherent biological properties. And hypoxic and three-dimensional (3D) culture is by far the most intensively studied.

Pre-activation with hypoxia

Under in vitro culture conditions, MSCs commonly exposed to an environment where the average oxygen tension is approximately 21% [34]. However, MSCs generally reside in a hypoxic microenvironment with physiological oxygen concentrations ranging from 1 to 11% in vivo [35, 36]. Several studies have illustrated that high oxygen concentration causes environmental stress in cultured MSCs, and then induces DNA damage and senescence [37, 38], and decreases their activities [39, 40]. Therefore, hypoxia is a crucial component of the physiological microenvironment for MSCs. Hypoxia-inducible factors (HIFs), especially HIF-1, are the crucial modulator of cellular response to hypoxia [41]. HIF-1 is a heterodimer containing two subunits, HIF-1α and HIF-1β. The lack of O2 allows oxygen labile protein HIF-1α accumulation and translocates into the nucleus, and then binds with HIF-1β to form the heterodimer, which further binds to a hypoxia-response element (HRE) in the target genes with co-activators such as CBP/p300 and then regulates the transcription of numerous genes [42,43,44].

Hypoxic pre-activation has multiple beneficial effects on MSCs. For example, a hypoxic culture environment maintains undifferentiated states of MSCs. Several reports showed that low oxygen tension increased the expression of multipotent stem cell markers (Oct4, Sox2, and Nanog) in ASCs without changing their surface markers and morphology [45, 46].

In addition, hypoxia facilitates the proliferation and survival of MSCs, leading to a higher expansion and more yield ASCs compared to normoxic state (20% O2) [45,46,47,48]. And the hypoxia priming seems to promote the mobilization of MSCs in vitro in migration assay [47, 49] and improve homing of MSCs in vivo [50, 51]. Lee et al. group demonstrated that hypoxia promoted the proliferation and migration potential of MSCs through the HIF-1α-GRP78-Akt signal axis [34]. The results of Rosová et al. showed that MSC cultured in hypoxia augmented the expression of hepatocyte growth factor (HGF) and its major receptor cMet, and HGF/cMet was the main signalling for MSC migration [51]. Another study found that hypoxic pre-activation promoted the migration of bone marrow-derived mesenchymal stem cells (BMSCs) by ways of increasing potassium Kv2.1 channel expression and FAK activities [50].

Besides, hypoxia priming could protect MSCs against the hostile microenvironment and maintain gene stability [37, 52]. The accumulation of HIF-1α under hypoxia activates normal cellular prion protein, which protects MSCs from oxidative stress-induced apoptosis via the activation of superoxide dismutase and catalase to inactivate cleaved caspase-3 [53]. Furthermore, when exposed to hypoxia, MSCs show a prolonged life span and avoid replicative senescence, and express fewer senescence-associated β-galactosidase compared to normoxic culture [53]. Hypoxia-conditioned MSCs exhibit gene stability with decreased DNA damage and less chromosomal aberration [37].

Moreover, the secretion profile of MSCs changes and the paracrine function is enhanced, after hypoxic culture. For instance, under hypoxia, the secretion of pro-angiogenic factors such as VEGF, HGF, and fibroblast growth factor-basic (bFGF) increased in MSCs. In contrast, anti-angiogenic factors such as thrombospondin-1 and plasminogen activator inhibitor-1 decreased in MSCs [3]. Furthermore, the immunosuppressive characteristics of MSCs improved under hypoxia with upregulation of anti-inflammatory factors including interleukin-6 (IL-6), IL-10, and indoleamine 2,3-dioxygenase (IDO) [54, 55]. And, hypoxia-cultivated MSCs inhibited the proliferation of CD4 and CD8 T lymphocytes and promoted the generation of Treg cells more effectively than MSCs exposed to normal oxygen [56, 57].

Significantly, numerous experimental studies found that MSCs pre-activated with hypoxia shows more prominent therapeutic effects than untreated MSCs. Hypoxia-pre-activated MSCs could obviously attenuate pulmonary oedema, alleviate pulmonary fibrosis, and improve lung function, compared to normoxic cultured MSCs in a pulmonary fibrosis model [58]. In addition, transplantation of hypoxic MSCs into a rat myocardial infarction model led to greater vascularization and smaller infarct size in injured sites compared to normoxic MSCs [59]. Furthermore, streptozotocin-induced diabetic mice received hypoxia-pre-activated MSCs caused significantly lower random and fasting blood glucose as well as improved oral glucose tolerance compared to normoxic MSCs-treated diabetic mice [40].

Therefore, compared with normoxic MSCs, MSCs pretreated with hypoxia present more favourable properties and hold better therapeutic potential. Additionally, hypoxic pre-activation has the advantages of being simple, low cost, easy popularization, and suitable for large-scale cell production. However, several issues need to be addressed before further application of hypoxic MSCs into the clinic. It is necessary to optimize the oxygen concentration in hypoxic pre-activation because the oxygen concentrations in physiological niches where MSCs live vary by tissue origin (1% to 7% in bone marrow, 10% to 15% in adipose tissue, and 1.5% to 5% in birth-associated tissues) is [58]. Moreover, the optimal duration for hypoxia also should be revealed. More importantly, biosafety is the most concerning issue of stem cell-based therapy and should be carefully determined before clinical application.

Pre-activation with 3D culture

Compared with the two-dimensional (2D) culture environment, the 3D culture system imitates the natural MSC microenvironment in vivo and provides enhanced cell–cell interactions or cell-ECM interaction, which can significantly improve the biological behaviours of MSCs, such as proliferation, immune regulation, and committed differentiation [60]. The 3D culture systems of MSCs, including multicellular spheroids, scaffolds, and hydrogels, have attracted more and more attention.

Spheroid culture

Spheroid culture, a comparably easy method for strengthening the biological activities of MSCs, has established [61]. Various methods have developed for the generation of MSC spheroid, such as the hanging drop technique, low-attachment approach as well as forced aggregation techniques [62]. With the development of biomaterials, scaffold-based culture platforms to generate MSC spheroid have developed.

The spheroid culture system can benefit the therapeutic potential of MSCs through increasing stemness and facilitating differentiation into different cell lineages [57, 63]. The results of Cheng et al. showed that chitosan film-based spheroid culture could dedifferentiate MSCs into more primitive state with downregulation of mesenchymal lineage markers CD29, CD90, and CD105 and upregulation of pluripotency-related markers Sox2, Oct4, Nanog, and SSEA-4 [45, 64]. Moreover, Zhang et al. employed a microgravity bioreactor to generate MSCs spheroids and found similar results [65]. Furthermore, spheroid-derived MSCs alter their differentiation preference and can transdifferentiate into non-mesenchymal lineage cells such as neural cells and hepatocytes [64, 65].

The spheroid culture system can benefit the therapeutic potential of MSCs by enhancing proliferation, migration, and homing efficiency. Compared to monolayer MSC culture, spheroid-derived MSCs exhibited higher proliferative activity [66]. The SDF-1/CXCR4 signalling pathway plays crucial role in the migration and engraftment of transplanted MSCs. Culturing of MSCs as spheroid restores the loss of CXCR4 expression caused by 2D culture [67]. Beside, increased expression of matrix metalloproteinases MMP-9 and MMP-13 has found in spheroid-derived MSCs, facilitating cellular invasion via the basement membrane [66].

The spheroid culture system can benefit the therapeutic potential of MSCs through promoting the secretion of therapeutic factors, including immunomodulatory and pro-angiogenic cytokines. Bartosh et al. reported that MSCs in a hanging drop model secreted higher levels of the anti-inflammatory factors such as tumour necrosis factor-inducible gene 6 protein (TSG-6) and stanniocalcin 1, and more effective suppressed lipopolysaccharides (LPS)-stimulated macrophage secretion of tumour necrosis factor-alpha (TNF-α) than control MSCs [68]. Several studies have demonstrated that other anti-inflammatory cytokines such as prostaglandin E2 (PGE-2), transforming growth factor-β1 (TGF-β1) as well as IL-6 also exhibited a greater levels in spheroid derived MSCs [69,70,71,72] than 2D cultured MSCs. Furthermore, the pro-angiogenic and pro-regenerative function enhanced in spheroid derived MSCs due to the significantly increased secretion of VEGF, HGF, bFGF, and angiogenin (ANG) [67, 73].

Spheroid culture optimized the biological properties of MSCs and enables them better therapeutic effects in vivo. Bartosh et al. proved that spheroid-derived MSCs were more effective than monolayer MSCs in suppressing inflammatory responses in a zymosan-induced peritonitis model of mouse. Spheroid-derived MSCs treated group showed substantially decreased neutrophil activity and proinflammatory molecules in serum [70]. Cheng et al. reported that spheroid-derived MSCs injected into the impaired healing wounds significantly promoted healing rates by increasing cellular engraftment and enhancing angiogenesis [66]. The study of Bhang et al. demonstrated that compared with the control group, more significant angiogenesis and less fibrosis were observed in the ischemic region in the spheroid-derived MSC transplantation group [74].

However, some disadvantages of the spheroid culture system limit the large-scale production of MSC spheroids for in vivo applications [75]. For example, the spheroid culture technique affects the spheroid size, and the variability of spheroid size impacts the therapeutic performance of MSCs. Therefore, it is necessary to discriminate the clinical needs, develop a reproducible spheroid culture system, and utilize animal models and clinical trials to confirm its safety and effectiveness before using spheroid-derived MSC in clinical practice.

Scaffold-based culture

Extracellular matrix (ECM) is one of the critical acellular components of stem cell niche, which can considered as the “soil” for stem cells. Cells in the niche are mechanically anchored to the ECM through transmembrane proteins known as integrins [76]. The primary function of ECM is to provide the necessary scaffold for cell growth and transduction of mechanical sensing signals to cells through cell-ECM interaction. And it also supplies essential nutrients and growth factors to cells [61, 77]. Therefore, it is necessary to build engineered niches to simulate native ECM in vivo. Recent advances in biomaterials have contributed to the development of artificial ECM culture systems for MSCs, and a variety of natural and synthetic biomaterials have emerged [61]. They should fulfil the properties of matrix mechanics, degradability, and biocompatibility. And the design principles of engineered ECM mainly depend on the native microenvironment of the stem cells type of interest or the desired phenotypic output. The existing engineered ECM could improve the biological properties of MSCs such as proliferation, homing, lineage differentiation, and paracrine [78,79,80,81].

Engineered ECM in the forms of scaffold, film, hydrogel, and sponge has been developed and used to amplify MSCs and enhance their biological properties [82,83,84]. Alginate has extensively used as a biocompatible carrier in tissue engineering. The work of Ewa-Choy et al. documented that the 3D alginate hydrogels created a microenvironment that facilitated the differentiation of ASCs into chondrogenic-like cells in ASCs and nasal chondrocytes co-cultured system. The efficiency of ASC-chondrocyte differentiation depended on alginate concentration [85]. Of note, the addition of specific inducers to the ECM scaffold will further increase the differentiation efficiency of MSCs into desired cell types. Several studies have shown that hydroxyapatite (HAp) is similar to pre-existing minerals during the bone remodelling process and possessing superior osteoinductive activity [86, 87]. BMSCs were implanted into the porous sponge constructed from duck’s feet-derived collagen (DC)/ with or without hydroxyapatite (HAp) to observe their osteogenic differentiation. Under the induction of an osteoconductive regulator dexamethasone, BMSCs in DC/HAp sponge tended to show higher proliferative activity and greater osteogenic differentiation [88].

The paracrine profile and immunomodulatory roles of MSCs seeded on engineered ECM scaffolds are also significantly altered. MSCs seeded in biomaterials showed up-regulation of anti-inflammatory regulators such as PGE2 and TSG-6, and down-regulation of anti-inflammatory regulators such as monocyte chemoattractant protein-1 (MCP-1), IL-6, and receptor activator of nuclear factor κ-B (NF-κB) ligand (RANKL) [60, 89]. Alginate-encapsulated MSCs attenuate TNF-α secretion and enhance PGE2 production more effectively than MSCs in 2D in an LPS-stimulated model of organotypic hippocampal slice culture [90]. In addition, ASCs cultured in alginate hydrogels showed significant inhibition on the proliferation of phytohaemagglutinin-stimulated peripheral blood mononuclear cells compared to monolayer culture [91]. Moreover, alginate hydrogel encapsulated MSCs promoted the conversion of macrophages to the anti-inflammatory M2 phenotype in vitro. And they played a similar immunomodulatory role in a rat model of spinal cord injury (SCI) as a greater percentage of the M2 subsets at the site of injury compared to control [92, 93].

Therefore, scaffold-based 3D culture endowed MSCs with more excellent biological activities and outstanding therapeutic efficacy than 2D culture. With people's understanding of biomaterials, their biological performance is also constantly expanding. Using only biological materials, biomimetic composite materials with multiple functions can be manufactured. For example, Jiao et al. developed a double-phase biomimetic procallus with gelatin-reduced graphene oxide (GOG) and photo-crosslinked gelatin hydrogel, which provide hypoxic microenvironment and mediated bidirectional differentiation of BMSCs to osteogenesis and angiogenesis, thereby promoting the regeneration of bone defects [94]. Therefore, a new generation of scaffold-based MSC culture techniques should not only simulate stem cell niches from multiple aspects simultaneously, provide as many stem cell niche components as possible, but also release bioactive molecules to target effector cells and activate "self-repair" mechanisms at damaged sites.

Pathological microenvironment simulation pre-activation

Since MSC enters the body, it will face the pathological microenvironment caused by the disease. There are a variety of destructive factors in this setting that can cause oxidative stress and apoptosis of transplanted cells, and significantly compromise the inherent therapeutic properties of MSCs. Researchers are committed to fine-tune the characteristics of the cells against hostile environments and suited for the targeted diseases. Based on the plasticity and memory ability of MSCs, cues in the trauma microenvironment, such as environmental factors (hypoxia), chemical factors (inflammatory factors and cytokines), will be the primary consideration for in vitro manipulation of MSCs [95]. In addition, new approaches have constantly explored, such as drugs, Traditional Chinese medicine and small molecule compounds have also become emerged for MSC priming [28, 96, 97]. Another concern is that the biological characteristics of endogenous MSCs in patients will change with the disease development, showing the loss and decline their function [98,99,100]. Therefore, it is meaningful to improve the biological activities of MSCs and enhance their therapeutic efficacy, especially to develop patient-customized therapeutic MSCs.

Pre-activation of MSCs with hypoxia

It has mentioned in the previous sections that hypoxia is an important component of the stem cell niches. Actually, MSCs always delivered into the injury site of ischaemia and hypoxia in animal experiments and clinical studies. Therefore, hypoxia is also a crucial component of the pathological setting [34]. And hypoxic culture in vitro can induce the memory of MSC to injury microenvironment for better therapeutic efficiency, which have been discussed in previous sections. However, there is a difference in oxygen concentrations between stem cell niches and damaged tissue. Furthermore, local oxygen concentrations in damaged tissues vary during different repair periods. Thus, the oxygen concentration of stem cell pretreatment was varied by purpose. For example, the in vitro culture of MSCs refers to physiological oxygen concentrations and further pretreated with pathologic oxygen concentrations prior to their in vivo application.

Pre-activation of MSCs with inflammatory factors

The pre-activation with inflammatory factors and cytokines is considered the most common means to mimic the inflammatory microenvironment in vivo and play a significant role in regulating the immunomodulatory function of stem cells [3]. In contrast with others, the proinflammatory cytokines such as TNF-α, INF-γ, and IL-1β are frequently observed in the traumatic microenvironment and are extensively studied for preactivating MSCs [3].

Pre-activation of MSCs with TNF-α

Increasing evidence suggests that appropriate priming of MSCs with disease-related stimuli improves their biological function and plays better therapeutic roles [101]. TNF-α is expressed in ischemic and injured tissues and commonly used to mimic the acute inflammatory environment [102]. Pre-activation of gingival tissue-derived MSCs (GMSCs) with TNF-α enhanced exosomal CD73 expression, which was essential for inducing anti-inflammatory M2 macrophage polarization [103]. Exosomes derived from TNF-α preactivated GMSC exhibited stronger anti-osteoclastogenic activity than control, thereby reducing periodontal bone resorption in a mice model of ligature-induced periodontitis [103]. Furthermore, TNF‐α- pre-activated MSCs showed improved proliferation, migration, and survival under H2O2‐-induced oxidative stress. And they exerted better endothelial protective functions through the massive secretion of HGF, VEGF and other cytokines than control MSCs [104]. Additionally, MSCs primed with TNF-α accelerated local vascularization of the injured sites in the ischemic hindlimb and cutaneous wound via secretion of pro-angiogenic cytokines, such as IL-6 and IL-8 [105, 106].

Pre-activation of MSCs with interferon (IFN)-γ

The proinflammatory cytokine IFN-γ is also a representative factor used for MSC pre-activation [107]. In response to IFN-γ, MSCs had a distinctive immunosuppressive profile, with the increased expression of several anti-inflammatory factors such as HGF, TGF-β1, IDO, prostaglandins, and cyclooxygenase 2 (COX-2) [108,109,110]. Prostaglandins and IDO secreted from the IFN-γ-stimulated MSCs were the main effectors in suppressing NK activation [109]. In addition, IL-2/15-activated NK cells induced less cytotoxicity to IFN-γ stimulated MSCs than nonstimulated MSCs due to their upregulation of inhibitory MHC Class I molecules, while IFN-γ-priming MSCs inhibited the proliferation of PBMCs more strongly than did the nonpriming MSCs [111], accompanied by upregulation of PD-L1 and increased secretion of COX-2-derived PGE2 [112]. The therapeutic potential of MSCs after IFN-γ pre-activation was significantly improved and demonstrated in models of CCl4-induced liver cirrhosis [113], obliterative bronchiolitis [114], and renal fibrosis [115]. This evidence illustrated that MSCs could be activated by inflammatory signalling, and sufficient to strengthen their immunoregulatory profile and therapeutic efficacy [110].

Pre-activation of MSCs with IL-1β

IL-1β is also a prevalent inflammatory cytokine in inflamed tissues produced by monocytes and macrophages [116]. It has shown that IL-1β pre-activation increases the expression of many adhesion molecules in MSCs, such as integrin LFA-1, thereby promoting adhesion to HUVECs through interaction with ICAM-1, which facilitates MSC cross-endothelium and homing [117]. In addition, Nie et al.’s study found that IL-1β pre-activated MSCs showed elevated CXCR4 expression and increased their migration towards SDF-1, leading to better therapeutic performance than naive MSCs in acute liver failure [118]. In addition, exosomes derived from IL-1β-pre-activated MSCs could induce macrophage polarization into M2 phenotype and attenuated the symptoms in the septic mice model more effectively than exosomes produced by naïve MSCs [119].

Pre-activation of MSCs with other proinflammatory cytokines and TLR ligands

In addition, some other inflammatory factors such as IL-17A and IL-25 were recently reported as alternative pre-activated means and acquired promising results. IL-17A is produced predominantly by CD4+ T helper 17 cells and plays regulatory roles in developing autoimmune and inflammatory diseases [120]. The results of a comparative study demonstrated that MSCs stimulated with IL-17A exhibited superior immunosuppressive properties than untreated MSCs [121]. IL-17A-treated MSCs showed the highest suppression of mitogen-activated CD3+ T cells compared with MSCs treated by IFN-γ, TNF-α, or IL-1β. And they also induce the generation of CD4+CD25highCD127lowFoxP3+ Tregs [121]. IL-25 is a member of the cytokine IL-17 family and has recently used to enhance MSC regulated immune response [122, 123]. Infusion of IL-25-primed MSCs significantly reduced IL-17-positive cells and increased FoxP3 positive cells, thereby alleviating intestinal inflammation in a rat model of DSS-induced colitis compared with unprimed MSC [122].

Lipopolysaccharide (LPS) is a component of the outer membrane of gram-negative bacteria that elevated in various diseases. Recent studies have shown that LPS serves as an essential mediator in the regulation of apoptosis in numerous cell types [124]. It has demonstrated that the biological effects of LPS on MSCs were closely associated with the concentration LPS used. MSCs treated with LPS at high-dose induced their apoptosis, and MSCs treated with low-dose of LPS enhanced their ability to resist oxidative stress and inhibit apoptosis, possibly depending on the upregulation of cellular FADD-like IL-1β-converting enzyme inhibitory protein. Furthermore, transplantation of low-dose LPS preactivated MSCs significantly improved MSC-mediated cardio-protection in an I/R injury model through MyD88-dependent activation of stat3 [124,125,126].

Pre-activation of MSCs with a combination of proinflammatory cytokines

The biological activities of MSCs are varied after pre-activated with various proinflammatory factors, so scientists conceived whether MSCs can be pretreatment by the combination of different inflammatory cytokines to compensate for the lower efficiency and maximize the therapeutic effect. After licensing with IFN-γ and TNF-α, MSCs retained their anti-apoptotic ability, which inhibited T cell proliferation and promoted CD14+ monocytes differentiated into anti-inflammatory CD206+ M2 macrophages more effectively than single-factor-induced MSCs. Moreover, pre-activation of ASCs with a combination of IFN-γ, TNF-α, and IL-17 dramatically enhanced their immunosuppressive effect and effectively cured concanavalin A (ConA)-induced liver injury in mice through an iNOS-dependent manner [127].

Therefore, inflammatory cytokines pre-activation can not only improve the ability of MSCs to resist oxidative stress, but also largely enhance the immunosuppressive properties of MSCs and strengthen their therapeutic efficacy. Nevertheless, some questions still need to be answered. Intensive studies need to further explore and identify the optimal concentration and action duration of inflammatory cytokine pre-activation alone or in combination. Alternatively, the possible side effects of inflammatory pre-activation, such as undesirable upregulation of class I and II HLA molecules, should also be concerned.

Pre-activation of MSCs with growth factors or regenerative cytokines

Priming MSCs with growth factors or regenerative cytokines have recently emerged and have proved to be an appealing approach. bFGF is one of potent pleiotropic cytokines and serves critical roles in regulating bio-properties of various stem cells and tissue regeneration [128]. bFGF-primed dental pulp stem cells derived from deciduous teeth (DPMSCs) show the highest angiogenic potential with the highest secretion of HGF and VEGF compared to control or hypoxic pre-activated MSCs [129]. Moreover, stimulation of canine MSCs with bFGF enabled their generation of cartilage tissue [130]. SDF-1 is also known as chemokine ligand 12 (CXCL-12), and its receptors CXCR4 and CXCR7 constitute the chemokine signalling critical for recruiting stem cells and organ repair after injury [131]. MSCs pre-activated with SDF-1 exhibited a significant anti-apoptotic capacity and proliferative potential, induced by a marked activation of Akt and ERK signalling pathways [132]. And angiogenesis also enhanced in preacitvated MSCs and was partly associated with their increased VEGF [132]. SDF-1α, the main spliced isoforms of SDF-1, has been used to priming MSCs [133]. SDF-1α pre-activation could augment the survival of MSCs in the infarcted myocardium, lessen the scar size, and enhance the cardiac systolic function [134]. In addition, infusion of MSCs pre-activated with TGF-α also obtained similar therapeutic effects in a rat model of acute myocardial I/R injury due to elevated VEGF secretion via a p38 mitogen-activated protein kinase (MAPK)-dependent mechanism [135].

Preactivating MSCs with a cocktail of growth factors revealed synergistic effects to enhance their biological function. Simultaneous pre-activation of MSCs with bFGF, IGF-1, and BMP-2 enhanced their plasticity and significantly upregulated myocardial transcription factors in the myocardial cells and MSCs co-culture model. Moreover, transplantation of these pre-activated MSCs resulted in reduced infarct size and improved cardiac function compared to transplantation of untreated MSCs [136]. However, selecting optimal cytokines to pre-activated MSCs and confer the desired biological function is a major step. Whether to choose cytokines that change in common in various diseases or to use disease-specific cytokines as a pre-activated condition is a question that needs to be thoroughly studied. Alternatively, the method of cytokine pre-activation is not economical enough because significant amounts of cytokines required for the large-scale production of pre-activated MSCs.

Pre-activation of MSCs with bioactive compounds

Bioactive compounds are a promising pre-activation method for strengthening biological properties of MSCs [28, 96, 137]. At present, the bioactive compounds used stimulation of MSCs can divided into natural (such as extracted from Traditional Chinese medicine (TCM)) and synthetic compounds according to their original; in terms of the screening principle of bioactive compounds, they either have a biological regulation effect on MSCs or have a therapeutic effect on target diseases; given their biological mechanisms for MSCs, bioactive compounds can classified as follows: promotion of the survival and migration, enhancement of the secretory activity, and reversion and reparation of disabled MSCs.

Bioactive compound for promoting the survival and migration of MSCs

Practically, exerting the inherent therapeutic properties of MSCs requires the transplanted cells to survive and function in a harsh and damaged setting [28]. A number of studies have focused on modifying MSCs to enhance their anti-apoptosis and migration capacity by using bioactive compounds. Trimetazidine (1-[2,3,4-trimethoxybenzyl]piperazine; TMZ) can lower the tissue damage caused by ischaemia and usually used to treat angina. It protected MSCs from hydrogen peroxide (H2O2)-induced oxidative stress by increasing the expression of pro-survival factors such as HIF-1α, Akt, survivin, and Bcl-2. And a significant improvement in cardiac function was observed after transplantation of TMZ pre-primed MSCs in a myocardial infarction model [97]. Tadalafil belongs to the long-acting PDE5 inhibitor group and has been applied to treat heart failure [138]. It improved ex vivo MSCs proliferation and survival via up‐regulation of miR‐21 to suppress Fas [138, 139]. It also prolonged MSC survival in vivo and promoted MSC mobilization and homing into the infarcted myocardium partly through SDF‐1α/CXCR4 cascade [139]. In addition, atorvastatin played a beneficial impact on endothelial function [140], which facilitated the survival of MSCs and promoted the therapeutic action of MSCs in infarcted hearts via eNOS/NO and SDF-1/CXCR4 pathways [140]. Vitamin E is a well-known antioxidant for its radical scavenging activity [141]. Vitamin E- pre-activated MSCs were resistant to H2O2-induced oxidative stress along with upregulation of proliferative markers (proliferating cell nuclear antigen and Ki67) and pro-regenerative markers (TGF-β and VEGF). Moreover, implantation of MSCs with Vitamin E served to repair the damaged cartilage in a rat model of osteoarthritis [142].

In recent years, TCM or its extracts has been investigated for their beneficial effects on MSCs [143]. Salvia miltiorrhiza (SM) is a widely known herb commonly found in many prescriptions of TCM for treating various diseases, including cardiovascular disease, Alzheimer's, and ischemic stroke [144,145,146]. SM effectively enhanced the viability and reduced cellular damage of MSCs under hypoxic condition. The infusion of SM modified MSCs showed the infarcted areas recovery and positive behaviour changes in the rat middle cerebral artery occlusion model [147]. Curcumin, an active component of turmeric (Curcuma longa), possesses pleiotropic effects such as antioxidant and anti-inflammatory [148, 149]. It exerts cytoprotective effects against oxidative stress-induced injury in ASCs by regulating PTEN/Akt/p53 pathway and haeme oxygenase-1 expression [150, 151]. Prior curcumin treatment significantly increased VEGF secretion in MSCs, and these pre-activated MSCs resulted in more neovascularization and functional recovery than naïve ASCs in ischemic myocardium [152]. In addition, curcumin pre-activated MSCs improved their therapeutic potential in acidic burn wounds as exhibited by improved microcirculation, pronounced granulation and hastened wound closure compared with wild type MSCs [153]. There are many other extracts of TCM, such as rosmarinic acid (RA) and gigantol, which protected the MSCs against H2O2-induced apoptosis via attenuating the expression of caspase-3, caspase-9 and Bax/Bcl-2 by regulating the PI3K/Akt and ERK1/2 signalling pathways [154,155,156]. These results indicate that they may developed as cytoprotective agent for successful MSC transplantation.

Bioactive compound for enhancing the immunomodulatory, paracrine and therapeutic potential of MSCs

In fact, the immunomodulatory and paracrine properties of MSCs are closely related to their therapeutic efficacy. Recent data demonstrate that pharmacological stimulus can boost the paracrine and immunoregulation potential of MSCs. Iron chelator deferoxamine (DFX) is a hypoxia mimetic agent with antioxidant properties. DFX used to pre-activate MSCs resulted in enhanced the secretion of anti-inflammatory (IL-4, IL-5 and COX2), pro-angiogenic factors (VEGFα and Angiopoietin-1), as well as neuroprotective factors (nerve growth factor, glial cell-derived neurotrophic factor, and neurotrophin-3) in MSCs [157]. The secretome of DFX- pre-activated MSCs could effectively reprogram LPS-induced macrophage DH82 into M2 phenotype [158]. And it also showed neuroprotective potential of dorsal root ganglion (DRG) neurons under high-dose glucose-induced injury [157]. Treprostinil, a prostacyclins analogue, was used to stimulate MSCs and produce proangiogenic effects by increasing VEGF-A production [159]. Overexpression of pro-survival, angiogenic, and pro-migration related genes, including COX‐2, HIF‐1, CXCR4, CCR2, VEGF, Ang‐2 and Ang‐4, has been found in All-trans retinoic acid (ATRA) ‐treated MSCs [160]. Moreover, wounds injected with ATRA‐treated MSCs showed significantly higher levels of vascularization, collagen deposition and re-epithelialization, resulting in accelerated wound closure compared to wounds injected with untreated MSCs [160]. Resveratrol (RSV) is a plant polyphenolic compound, which can protect MSCs from inflammation and oxidative injury [161, 162], potentiating their paracrine function, preventing their ageing and so on [163, 164]. For example, RSV pre-activation enhanced the secretion of PDGF-DD in MSCs that further activated the ERK signalling pathway in renal tubular cells, promoted angiogenesis in endothelial cells, and preferably repaired cisplatin-induced renal injury [164]. Buyang huanwu decoction (BHD) is a famous formula in TCM for supplementing Qi and activating blood and has been used to treat central nervous diseases [165]. Compared with untreated BMSCs, exosomes derived from BHD- pre-activated rat BMSCs contained more angiogenetic miRNA and elevated angiogenesis in rat brain after bilateral carotid artery ligation [166].

The development of high-throughput technologies brings us a new perspective for screening bioactive compounds to target specific genes in MSC, thus regulating the expression profile of MSCs and specifically enhancing their desired biological functions. For example, tetrandrine was selected to specifically upregulate PGE2 expression in MSCs through the NF-κB/COX-2 signalling pathway. Tetrandrine- pre-activated MSCs showed a significant reduction in TNF-α secretion after co-culture with mouse macrophages (RAW264.7) and attenuation of TNF-α level in mouse inflamed ear [167].

Bioactive compound for reversing and repairing of disabled MSCs

MSCs from perinatal tissues will undergo replicative ageing with the large-scale amplification in vitro, and MSCs from aged donors generally present premature senescence phenotype, and MSCs from patients showed the decline of MSC biological functions. These MSCs generally present downregulated cell function in proliferation, mobility, differentiation, and immunoregulation with impaired therapeutic capability [98,99,100]. Therefore, reversing the functions of these MSCs, namely the rejuvenation of MSCs, is vital for MSC-based therapy. Fortunately, recent studies have shown cellular functional decline or premature senescence can be rescued [168]. In general, approaches to rescue MSCs can broadly defined as reducing the level of intracellular oxidative stress, reprograming MSCs through adjustment of epigenetic modifications, as well as usage of senolytic drugs.

Inhibition of excessive oxidative stress

Data show that ROS, as natural by-products produced by cell metabolism, is maintained at a low level in MSCs and is essential for the proliferation and differentiation of MSCs [169, 170]. High levels of ROS-induced under chronological ageing or pathological conditions will cause severe cytotoxicity and cell damage [168]. Several studies have reported that malfunctioned MSCs can be reversed by modulation of intracellular ROS aggregation and oxidative metabolism [168]. Antioxidants have become the natural choice, and several antioxidants used for anti-ageing studies of MSCs. N-acetylcysteine (NAC), a ROS scavenger, can significantly attenuate ROS accumulation due to overactivation of Wnt/β-catenin signalling in MSCs, thereby lessening ROS induced DNA damage and downregulating the expression of senescence-associated marker p16 (INK4A), p53 and p21 [171]. Another free radical scavenger, edaravone, rescued the functions of elderly AT-MSCs by reducing ROS level and β-gal-positive cells. Moreover, it could also protect BMSCs from the intracellular accumulation of ROS caused by hypoxia and upregulation of antioxidant enzymes in UC-MSCs [172, 173]. More importantly, pre-activation with edaravone restored the elderly AT-MSCs’ in vivo therapeutic functions as decreased necrotic area in an ischemic flap mouse model [174].

Overall, ROS can act as intracellular messengers and help perform vital biological functions, so it is essential to control the optimal concentration of ROS by adjusting the amount of antioxidants. After all, high doses of antioxidants can cause DNA damage and premature senescence [175].

Bioactive compound for modifying of epigenetic dysregulation

Epigenetic regulation, an important mechanism for programming, changes the cellular phenotype by alteration of gene expression rather than DNA sequence. It is characterized by heritability and reversibility, includes DNA methylation and histone modifications, and has profound influence on MSC fate [176, 177]. The epigenetic dysregulation found in MSCs after routine culture expansion appears to be unrelated to changes in global histone acetylation level, but involves histone acetylation levels at the promoters of some genes, such as TERT, Soc2, Oct4, Runx2, and ALP, which ultimately leads to cellular senescence [178]. Moreover, a general decrease in DNA methylation has reported in MSCs derived from old compared with MSCs derived from young by using a BeadChip microarray [179]. Given the reversibility of epigenetic modification, it is a potential strategy to explore epigenetic targeted therapy for reprogramming old stem cells into youthful functional stem cells.

With current techniques, methods to reprogram ageing stem cells occur in two main ways, fully reprogramming and partial reprogramming. The former is referred to the reset of the epigenetic clock for finally obtaining the induced pluripotent stem cells (iPSCs). Functional MSCs have successfully generated from iPSCs with rejuvenated gene signature and improved cell vitality, but their immunoregulatory function for suppressing T cell proliferation is incomplete [180,181,182]. Therefore, the therapeutic efficiency of iPSCs-derived MSCs, especially the immunomodulatory functions, needs to be thoroughly assessed. In addition, due to the low efficiency, the limited number of iPSCs, and high cost, fully reprogramming is still only a means in the laboratory, and there is still a considerable distance from the bedside. Partial reprogramming involves incomplete dedifferentiation and is considered as epigenetic rejuvenation, which can achieved by regulating DNA methylation and histone modification using bioactive compounds [183].

DNA methyltransferase (DNMT) inhibitor, 5‐azacytidine (5‐AZA), is readily incorporated into DNA and inhibits methylation patterns of specific gene regions, simultaneously activating relevant genes [184]. 5‐AZA pre-activation reversed the aged phenotype of ASCs and enhanced their proliferation, shortened population doubling time, and increased extracellular vesicle secretion via reducing ROS accumulation, ameliorating superoxide dismutase activity, and increasing BCL-2/BAX ratio [185]. Moreover, RG108 is also known as a DNA methyltransferase inhibitor. RG108-educated BMSCs showed a significantly reduction in β-galactosidase-positive cells, simultaneously with up-regulation of anti-senescence genes TERT, bFGF, VEGF, and ANG and down-regulation of senescence-related genes ATM, p21, and p53 [186].

Tetramethylpyrazine (TMP), the bioactive component extracted from the rhizome of the Chinese herbal medicine Chuanxiong, can epigenetically alleviate senescent phenotype of BMSCs by regulating EZH2 (a histone-lysine N-methyltransferase enzyme)-H3k27me3 [187]. EZH2 has found to repress transcription of both p16/p14 by increasing H3K27me3 along the Ink4A locus [188]. Moreover, previous studies have revealed that TMP also possesses the capacity to significantly delay MSC senescence by suppressing NF-κB signalling and positively regulating the proliferation, lineage commitment, and anti-apoptosis [189,190,191].

Increased histone acetylation, decreased DNA methylation and hydroxymethylation, and distinct changes in H3K27me3 in the genome are prevalent in senescent cells. However, reversing stem cell senescence by altering epigenetic modifications is still in infancy. It is necessary to map out the detailed epigenetic alterations of MSCs during senescence, especially the epigenetic characteristics associated with the altered behaviour in MSC biology. In addition, the universality or uniqueness of epigenetic changes in MSC ageing from different sources needs to be confirmed.

Usage of senolytic drugs

Senotherapeutics refers to the application of senolytic drugs to selectively deplete senescent cells or delay the onset of senescence, thereby rejuvenating tissues and reducing the occurrence of age-related pathologies [192]. Several compounds have reported to hold the perspective senolytic effects, such as navitas (ABT-263), quercetin, danazol, nicotinamide riboside, dasatinib and metformin [192, 193]. A comparative study found that ABT-263, instead of quercetin, danazol, and nicotinamide riboside, suggested possessing senolytic effects in a replicative senescence model of MSCs after long-term expansion [193]. In addition, abdominal fat-derived MSCs in pregnant women with preeclampsia present a senescent phenotype with decreased cell function and viability. Treatment of them with the anti-senescence drug dasatinib was both able to selectively promote apoptosis of senescent MSCs and dramatically improve the biological activities of MSCs, including an increase in angiogenic potential, reduction in SA-β-gal positive cells, and downregulation of IL-6, IL-8, MCP-1, and p16 [194].

It is conceivable that combinations of different drugs may achieve a more pronounced senolytic effect. As shown in the study of Zhou et al., a senolytic cocktail of dasatinib and quercetin improved osteogenic potential of aged mouse-derived BMSCs in vitro or in vivo calvarial defect model, accompanied by a decrease in SA-β-gal-staining cells, and a reduction in senescence-associated and inflammation markers including p16, p21, IL-6, MCP1, and CXCL1 [195]. Notably, several findings have shown that a cocktail of dasatinib and quercetin can reduce the burden of senescent cells and meliorate the function of vital tissues such as adipose, bone, aorta, and brain [195, 196]. However, more evidence is needed to support the effectiveness of senolytic cocktail therapies in rescuing the functions of MSCs.

Generally, these findings suggest that bioactive compounds have properties to improve disabled MSCs and repair senescent MSC. Still, their dosage, combination, and suitability for MSCs from alternative sources require in-depth exploration and verification. In addition, the exploitation of more effective bioactive compounds to rescue MSCs is also the direction that scientists need to continue their efforts. Besides, the way to use bioactive compounds is also worth careful consideration. As demonstrated in several studies, using bioactive compounds as concomitant agents for MSC transplantation also has a better therapeutic effect than stem cell transplantation alone [24, 95, 197]. Still, patients with chronic diseases are also accompanied by multiple risk factors, such as age, diabetes mellitus, and cardiovascular diseases. Hence, the possible side effects and safety of concomitant drugs for stem cell transplantation should be considered. As deficiency of MSCs can reversed, this area is a potential hotbed for increasing the longevity and biological properties of in vitro expanded MSCs, and repairing patient-derived MSC for autologous transplantation [198]. However, the senescence of MSCs is a highly complex process and a thorough understanding of the underlying mechanism of senescent will help us to find more effective ways to rejuvenate ageing MSCs.

Pre-activation of MSCs with the disease-associated effector cells or patient's serum

“Individualized MSCs therapy” means that MSCs obtained by in vitro pre-activation possess customized functions and can specifically target the disease of the patient, thus achieving a better therapeutic outcomes. Therefore, as an enhancement strategy for MSC-based therapy, direct use of effector cells or their released active substances was proposed as pre-activating conditions, rather than educated with typical proinflammatory factors, cytokines, or bioactive compounds.

Mast cells (MCs) have a central role in immediate hypersensitivity and allergic reactions and are also the principal effector cells in the pathogenesis of atopic dermatitis (AD). Activated MCs release granules that contain a large number of bioactive substances, such as proinflammatory cytokines, protein mediators, lipid mediators, and growth factors, that can trigger allergic reactions [199]. Several studies have demonstrated that MSCs suppress MC activation and degranulation and induce MC apoptosis in a co-cultured system [200,201,202]. Then, pre-activated MSCs with MC granules could be a promising strategy to enhance the MSC-targeted treatment of AD. The study of Lee et al. showed that MC granule-primed UC-MSCs exhibited more immunosuppressive than non-primed cells, which are mediated by interrupting proliferation and degranulation of MCs via upregulating the COX-2/PGE2 signalling pathway [201]. In addition, in a dermatophagoides farina-induced AD model, subcutaneously infusion of MC granule-educated UC-MSCs showed a more significant decreased number of MCs and alleviated the infiltration of lymphocytes in the skin than that of naïve cells [201]. Therefore, MSCs pre-activated with effector cells or their derived active substances can accurately target the main pathogenic factors in disease development, and react and respond quickly in vivo to achieve more efficient treatment outcomes.

Alternatively, the alteration of inflammatory factors, chemokines, growth factors, cytokines, and microvesicles in blood circulation has found in a variety of diseases, including Alzheimer's disease, renal diseases, and heart disease [203,204,205]. They serve as indicators for disease diagnosis, treatment, and prognosis, along with inter-individual variations. Pre-activated expanded MSCs with patient-derived serum may allow MSCs to respond positively to the host microenvironment [203]. As Tang et al.'s study showed, compared with rats injected with control serum- pre-activated MSCs, rats injected with stroke serum- pre-activated MSCs showed a significantly improved behaviour with attenuated inflammatory cytokines, decreased brain lesion and apoptosis cells, and increased trophic growth factors in the cerebral I/R injury model [206]. In addition, strengthening the therapeutic effect of MSCs with disease-derived serum have also revealed in dextran sodium sulphate-induced colitis rat models. After MSCs pre-activated with serum derived from colitis rats or normal rats, the conditioned medium of both pre-activated cells was collected for treating colitis. The former shows more effectively impede the disease progress, better improvement in the clinical features, and much lower histological damage scores in colitis rats than the latter [207].

Therefore, disease-specific pre-activation may be a promising means to achieve "MSC customized clinical treatment." And accurate capturing of disease specificity is an essential prerequisite for this strategy, which requires comprehensive and in-depth exploration and analysis of the biological mechanism of diseases.

Perspective

In addition to pre-activating MSCs by a recreation of the physiological and pathological microenvironment, there are other means for MSC pre-activation, including photostimulation, magnetoelectric stimulation, and heat shock (HSP), etc. Low levels of lasers therapy (LLLT) is beneficial for regulating the biological functions of a variety of cells [208, 209]. For example, the biological activities of ASCs stimulated by low-level laser were enhanced, manifested by increasing survival rate, augmenting secretion, and accelerating regenerative healing compared with unstimulated ASCs [210, 211]. Moreover, pulsed electromagnetic fields have recently demonstrated to play a protective effect on BMSCs through regulating the Akt/Ras signalling pathway and upregulation of survival proteins such as Bad and Bcl-xL [212]. In addition, studies have shown that HSP can induce cytoprotective proteins and increase the ability to resist a poor external environment [213, 214]. HSP pre-activation enhanced MSCs autophagy and increased their resistance to H2O2-induced apoptosis. Besides, HSP-MSCs showed enhanced homing and survival following transplantation in a hepatic I/R injury model compared with control MSCs [215]. Furthermore, intraovarian injection of HSP-MSCs rescued the damaged ovarian structure and ameliorated endocrine function [216].

Generally, the purpose of existing pre-activating approaches is to take full advantage of the functional plasticity of MSCs and assign the desired properties to MSCs in advance so that when MSC reencounters similar environment, the cell protection mechanism initiated, the response mechanism activated rapidly, and the corresponding biological response acted quickly.

Notwithstanding, the application of the pre-activated MSCs confronted with several challenges. The first is to choose reasonable and effective MSC pre-activating methods. To date, so many complementary methods have been proposed for improving the therapeutic efficacy of MSCs, and finding the best pre-activation way is an essential pursuit in our future research. Each pretreatment targets improving a specific aspect of MSCs, and the optimal combinations of diverse strategy are conceivable to maximize the therapeutic outcome of MSCs. For example, MSCs were inoculated into an injectable gel of collagen microcarriers cross linked with bFGF or TGF-β1 to promote expansion and chondrogenic differentiation [217]. Furthermore, effector cell-based or patient serum-based pre-activation of MSCs may be relatively more targeted for disease treatment, which requires amounts of comparative studies to further determine this inference. Definitely, novel enhancement strategies to generate therapeutically effective MSCs are still demanded and will undoubtedly receive constant attention.

The second is the heterogeneity of MSCs that mainly manifested in two aspects: one is that the biological characteristics of MSCs from diverse origins appear to vary in terms of differentiation, phenotype panel, and secretion profile, which means that as a novel cellular drug, MSCs from different tissues have inconsistent therapeutic effects on the same disease. As demonstrated in Liu et al.'s study, ASCs held the most pronounced effect on promoting re-epithelialization and wound closure compared with BMSCs and amnion MSCs (AMSCs). Furthermore, ASCs had the most excellent impact on enhancing the migration of dermal fibroblasts and the expression of pro-repair cytokines such as VEGF, bFGF, and TGF-β in contrast with BMSCs and AMSCs [218]. Therefore, it is necessary to gain insight into the unique biological characteristics of each tissue-derived MSCs and to determine the disease trophic properties of each MSCs in combination with different disease models. The other is the discrepancy of MSCs from different tissue in responsiveness to the same precondition, suggesting an optimal pre-activation way for various MSCs. For example, TNF-α-educated BMSCs showed significantly higher level of IL-10 than TNF-α-primed UC-MSC [219]. Further, response mechanisms of MSCs derived from different tissues vary when faced with the same pre-activating condition. For instance, co-culture of stimulated/unstimulated UC-MSCs with phytohaematoagglutinin-activated lymphocytes resulted in early activation of a negative co-stimulatory molecule CTLA4 in UC-MSCs, whereas changed IL-12 expression in co-cultured BMSCs [219]. These discrepancies could be partly determined by the functional heterogeneity between MSCs from different tissues. This may require a deeper understanding of their developmental processes and signalling regulatory networks, and a combination of high-throughput techniques and basic research can provide possible clues to determining the optimal pre-activation methods for each type of MSCs.

The third is to develop standard platforms for evaluating the safety and therapeutic characteristics of pre-activated MSCs (Fig. 3). Extensive research evidence confirms that MSCs mainly rely on paracrine to exert biological effects. The sum of MSCs secretion products can be regarded as the secretome, which mainly contains soluble proteins and extracellular vesicles, of which the former includes cytokines, chemokines, growth factors, and the latter can divided into exosomes and microvesicles [220, 221] (Fig. 1). Therefore, it is a troublesome problem to accurately select a component both as an indicator of preactivation and therapeutic effect of educated MSCs. However, there is increasing evidence that in vivo apoptosis of transplanted MSCs is closely associated with stem cell therapeutic effects in multiple animal models, such as GvHD [222], sepsis and acute lung injury [223, 224]. The findings of Pang et al. are further illustrated impeding MSC apoptosis by ablation of BAK/BAX reduces their immunomodulatory capacity in the model of OVA-induced asthma, suggesting that the in vivo biological mechanism of MSCs is far more complicated than we thought [225]. Therefore, inhibition and resistance to stem cell apoptosis as a commonly used evaluation index of MSC pre-activation are questionable.

Fig. 3
figure 3

Cycle of naive MSCs to preconditioned MSCs towards clinical treatment. Preconditioned MSCs achieved favourable therapeutic efficacy, with enhanced injury repair and disease recovery

Finally, the essence of MSC therapy in vivo is to deliver pro-reparative regulatory factors and extracellular vesicles, etc. Therefore, researchers have begun to use MSC-derived soluble proteins and extracellular vesicles to replace stem cells for treatment [226, 227], whose release is precisely regulated and their composition changes with the pre-activated condition [228]. This method can avoid the risk of in vivo proliferation, differentiation of MSCs, and secretion of unpredictable paracrine factors in MSCs. Recent results showed that MSCs were enucleated into "cargocytes" by density-gradient centrifugation to form a bioinspired delivery method. Cargocytes retain paracrine secretion capacities, do not proliferate, or permanently engraft in the host. Application of cargocytes not only avoids the adverse events associated with the direct use of MSCs, such as pulmonary or cerebral emboli, but also improves biodistribution and enhances homing to target tissues in vivo [229]. Therefore, derivations from the MSCs may be therapeutic vehicles to deliver curative cargos with the potential to treat diseases in a controllable and effective manner.

Conclusion

In summary, despite pre-activated MSCs remaining problematic, they still hold considerable promise for the treatment of various refractory diseases due to their tremendous regenerative potential. To date, there is a growing consensus that pre-activated MSCs indeed exhibit better therapeutic benefits than naive MSCs in a variety of pathological conditions. In future, we should effectively use the "pre-activation" tool to maximize the therapeutic potential of MSCs on the one hand and on the other hand to modify them suitable for targeted disease, opening a new chapter for clinical application of MSCs.

Availability of data and materials

Not applicable.

Abbreviations

MSC:

Mesenchymal stem cell

ASCs:

Adipose-derived MSCs

T2D:

Type 2 diabetes

VEGF:

Vascular endothelial growth factor

3D:

Three-dimensional

HIF:

Hypoxia-inducible factors

HPHs:

HIF-1 prolyl-hydroxylases

HRE:

Hypoxia-response element

HGF:

Hepatocyte growth factor

BMSCs:

Bone marrow-derived mesenchymal stem cells

PrPC:

Cellular prion protein

bFGF:

Fibroblast growth factor-basic

ANG:

Angiogenin

TNF-α:

Tumour necrosis factor-alpha

PGE-2:

Prostaglandin E2

TGF-β1:

Transforming growth factor-β1

ECM:

Extracellular matrix

Hap:

Hydroxyapatite

MCP-1:

Monocyte chemoattractant protein-1

IL:

Interleukin

SCI:

Spinal cord injury

GMSCs:

Gingival tissue-derived MSC

IDO:

Indoleamine 2,3, dioxygenase

COX-2:

Cyclooxygenase 2

LPS:

Lipopolysaccharide

CXCL-12:

Chemokine ligand 12

TCM:

Traditional Chinese medicine

H2O2 :

Hydrogen peroxide

TMZ:

1-[2,3,4-Trimethoxybenzyl]piperazine

SM:

Salvia miltiorrhiz

RA:

Rosmarinic acid

DFX:

Deferoxamine

ATRA:

All-trans retinoic acid

RSV:

Resveratrol

BHD:

Buyang huanwu decoction

iPSCs:

Induced pluripotent stem cells

5‐AZA:

5‐Azacytidine

TMP:

Tetramethylpyrazine

ABT-263:

Navitas

AD:

Atopic dermatitis

MCs:

Mast cells

HSP:

Heat shock

AMSCs:

Amnion MSCs

References

  1. Fisher SA, Doree C, Mathur A, Taggart DP, Martin-Rendon E. Stem cell therapy for chronic ischaemic heart disease and congestive heart failure. Cochrane Database Syst Rev. 2016;12:CD007888.

    PubMed  Google Scholar 

  2. Radrizzani M, Lo Cicero V, Soncin S, Bolis S, Surder D, Torre T, Siclari F, Moccetti T, Vassalli G, Turchetto L. Bone marrow-derived cells for cardiovascular cell therapy: an optimized GMP method based on low-density gradient improves cell purity and function. J Transl Med. 2014;12:276.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Seo Y, Shin TH, Kim HS. Current strategies to enhance adipose stem cell function: an update. Int J Mol Sci. 2019;20(15):3827.

    Article  PubMed Central  Google Scholar 

  4. Soria-Juan B, Escacena N, Capilla-Gonzalez V, Aguilera Y, Llanos L, Tejedo JR, Bedoya FJ, JuanV, De la Cuesta A, Ruiz-Salmeron R, Andreu E, Grochowicz L, Prosper F, Sanchez-Guijo F, Lozano FS, Miralles M, Del Rio-Sola L, Castellanos G, Moraleda JM, Sackstein R, Garcia-Arranz M, Garcia-Olmo D, Martin F, Hmadcha A, Soria B, Collaborative Working Group "Noma Project T. Cost-effective, safe, and personalized cell therapy for critical limb ischemia in type 2 diabetes mellitus. Front Immunol. 2019;10:1151.

  5. Yasuhara T, Kameda M, Sasaki T, Tajiri N, Date I. Cell therapy for Parkinson’s disease. Cell Transplant. 2017;26(9):1551–9.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Hass R, Kasper C, Bohm S, Jacobs R. Different populations and sources of human mesenchymal stem cells (MSC): a comparison of adult and neonatal tissue-derived MSC. Cell Commun Signal. 2011;9:12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Kucharzewski M, Rojczyk E, Wilemska-Kucharzewska K, Wilk R, Hudecki J, Los MJ. Novel trends in application of stem cells in skin wound healing. Eur J Pharmacol. 2019;843:307–15.

    Article  CAS  PubMed  Google Scholar 

  8. Mareschi K, Ferrero I, Rustichelli D, Aschero S, Gammaitoni L, Aglietta M, Madon E, Fagioli F. Expansion of mesenchymal stem cells isolated from pediatric and adult donor bone marrow. J Cell Biochem. 2006;97(4):744–54.

    Article  CAS  PubMed  Google Scholar 

  9. Tong Z, Solanki A, Hamilos A, Levy O, Wen K, Yin X, Karp JM. Application of biomaterials to advance induced pluripotent stem cell research and therapy. EMBO J. 2015;34(8):987–1008.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Lee SH, Jin SY, Song JS, Seo KK, Cho KH. Paracrine effects of adipose-derived stem cells on keratinocytes and dermal fibroblasts. Ann Dermatol. 2012;24(2):136–43.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Schlosser S, Dennler C, Schweizer R, Eberli D, Stein JV, Enzmann V, Giovanoli P, Erni D, Plock JA. Paracrine effects of mesenchymal stem cells enhance vascular regeneration in ischemic murine skin. Microvasc Res. 2012;83(3):267–75.

    Article  CAS  PubMed  Google Scholar 

  12. Wagers AJ. The stem cell niche in regenerative medicine. Cell Stem Cell. 2012;10(4):362–9.

    Article  CAS  PubMed  Google Scholar 

  13. Klyushnenkova E, Mosca JD, Zernetkina V, Majumdar MK, Beggs KJ, Simonetti DW, Deans RJ, McIntosh KR. T cell responses to allogeneic human mesenchymal stem cells: immunogenicity, tolerance, and suppression. J Biomed Sci. 2005;12(1):47–57.

    Article  CAS  PubMed  Google Scholar 

  14. Ryan JM, Barry FP, Murphy JM, Mahon BP. Mesenchymal stem cells avoid allogeneic rejection. J Inflamm (Lond). 2005;2:8.

    Article  CAS  Google Scholar 

  15. Onate B, Vilahur G, Camino-Lopez S, Diez-Caballero A, Ballesta-Lopez C, Ybarra J, Moscatiello F, Herrero J, Badimon L. Stem cells isolated from adipose tissue of obese patients show changes in their transcriptomic profile that indicate loss in stemcellness and increased commitment to an adipocyte-like phenotype. BMC Genom. 2013;14:625.

    Article  CAS  Google Scholar 

  16. Qi Y, Ma J, Li S, Liu W. Applicability of adipose-derived mesenchymal stem cells in treatment of patients with type 2 diabetes. Stem Cell Res Ther. 2019;10(1):274.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Serena C, Keiran N, Ceperuelo-Mallafre V, Ejarque M, Fradera R, Roche K, Nunez-Roa C, Vendrell J, Fernandez-Veledo S. Obesity and type 2 diabetes alters the immune properties of human adipose derived stem cells. Stem Cells. 2016;34(10):2559–73.

    Article  CAS  PubMed  Google Scholar 

  18. Scruggs BA, Semon JA, Zhang X, Zhang S, Bowles AC, Pandey AC, Imhof KM, Kalueff AV, Gimble JM, Bunnell BA. Age of the donor reduces the ability of human adipose-derived stem cells to alleviate symptoms in the experimental autoimmune encephalomyelitis mouse model. Stem Cells Transl Med. 2013;2(10):797–807.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wang H, Liang X, Xu ZP, Crawford DH, Liu X, Roberts MS. A physiologically based kinetic model for elucidating the in vivo distribution of administered mesenchymal stem cells. Sci Rep. 2016;6:22293.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Eggenhofer E, Benseler V, Kroemer A, Popp FC, Geissler EK, Schlitt HJ, Baan CC, Dahlke MH, Hoogduijn MJ. Mesenchymal stem cells are short-lived and do not migrate beyond the lungs after intravenous infusion. Front Immunol. 2012;3:297.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Lee RH, Pulin AA, Seo MJ, Kota DJ, Ylostalo J, Larson BL, Semprun-Prieto L, Delafontaine P, Prockop DJ. Intravenous hMSCs improve myocardial infarction in mice because cells embolized in lung are activated to secrete the anti-inflammatory protein TSG-6. Cell Stem Cell. 2009;5(1):54–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Matsuura K, Honda A, Nagai T, Fukushima N, Iwanaga K, Tokunaga M, Shimizu T, Okano T, Kasanuki H, Hagiwara N, Komuro I. Transplantation of cardiac progenitor cells ameliorates cardiac dysfunction after myocardial infarction in mice. J Clin Invest. 2009;119(8):2204–17.

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Kale VP. Application of “primed” mesenchymal stromal cells in hematopoietic stem cell transplantation: current status and future prospects. Stem Cells Dev. 2019;28(22):1473–9.

    Article  PubMed  Google Scholar 

  24. Li H, Zhu H, Ge T, Wang Z, Zhang C. Mesenchymal stem cell-based therapy for diabetes mellitus: enhancement strategies and future perspectives. Stem Cell Rev Rep. 2021;17(5):1552–69.

    Article  CAS  PubMed  Google Scholar 

  25. Asahara T, Kalka C, Isner JM. Stem cell therapy and gene transfer for regeneration. Gene Ther. 2000;7(6):451–7.

    Article  CAS  PubMed  Google Scholar 

  26. Mitrecic D, Nicaise C, Klimaschewski L, Gajovic S, Bohl D, Pochet R. Genetically modified stem cells for the treatment of neurological diseases. Front Biosci (Elite Ed). 2012;4:1170–81.

    Article  Google Scholar 

  27. Song SH, Lee MO, Lee JS, Jeong HC, Kim HG, Kim WS, Hur M, Cha HJ. Genetic modification of human adipose-derived stem cells for promoting wound healing. J Dermatol Sci. 2012;66(2):98–107.

    Article  CAS  PubMed  Google Scholar 

  28. Bui TVA, Hwang JW, Lee JH, Park HJ, Ban K. Challenges and limitations of strategies to promote therapeutic potential of human mesenchymal stem cells for cell-based cardiac repair. Korean Circ J. 2021;51(2):97–113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Nayerossadat N, Maedeh T, Ali PA. Viral and nonviral delivery systems for gene delivery. Adv Biomed Res. 2012;1:27.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Bernardo ME, Fibbe WE. Mesenchymal stromal cells: sensors and switchers of inflammation. Cell Stem Cell. 2013;13(4):392–402.

    Article  CAS  PubMed  Google Scholar 

  31. Abdul-Al M, Kyeremeh GK, Saeinasab M, Heidari Keshel S, Sefat F. Stem cell niche microenvironment: review. Bioengineering (Basel). 2021;8(8):108.

    Article  CAS  Google Scholar 

  32. Augustine R, Dan P, Hasan A, Khalaf IM, Prasad P, Ghosal K, Gentile C, McClements L, Maureira P. Stem cell-based approaches in cardiac tissue engineering: controlling the microenvironment for autologous cells. Biomed Pharmacother. 2021;138:111425.

    Article  CAS  PubMed  Google Scholar 

  33. Vunjak-Novakovic G, Scadden DT. Biomimetic platforms for human stem cell research. Cell Stem Cell. 2011;8(3):252–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Lee JH, Yoon YM, Lee SH. Hypoxic preconditioning promotes the bioactivities of mesenchymal stem cells via the HIF-1alpha-GRP78-Akt axis. Int J Mol Sci. 2017;18(6):1320.

    Article  PubMed Central  CAS  Google Scholar 

  35. Goossens GH, Blaak EE. Adipose tissue oxygen tension: implications for chronic metabolic and inflammatory diseases. Curr Opin Clin Nutr Metab Care. 2012;15(6):539–46.

    Article  CAS  PubMed  Google Scholar 

  36. Tsai CC, Yew TL, Yang DC, Huang WH, Hung SC. Benefits of hypoxic culture on bone marrow multipotent stromal cells. Am J Blood Res. 2012;2(3):148–59.

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Estrada JC, Albo C, Benguria A, Dopazo A, Lopez-Romero P, Carrera-Quintanar L, Roche E, Clemente EP, Enriquez JA, Bernad A, Samper E. Culture of human mesenchymal stem cells at low oxygen tension improves growth and genetic stability by activating glycolysis. Cell Death Differ. 2012;19(5):743–55.

    Article  CAS  PubMed  Google Scholar 

  38. Fehrer C, Brunauer R, Laschober G, Unterluggauer H, Reitinger S, Kloss F, Gully C, Gassner R, Lepperdinger G. Reduced oxygen tension attenuates differentiation capacity of human mesenchymal stem cells and prolongs their lifespan. Aging Cell. 2007;6(6):745–57.

    Article  CAS  PubMed  Google Scholar 

  39. Han KH, Kim AK, Kim MH, Kim DH, Go HN, Kim DI. Enhancement of angiogenic effects by hypoxia-preconditioned human umbilical cord-derived mesenchymal stem cells in a mouse model of hindlimb ischemia. Cell Biol Int. 2016;40(1):27–35.

    Article  CAS  PubMed  Google Scholar 

  40. Schive SW, Mirlashari MR, Hasvold G, Wang M, Josefsen D, Gullestad HP, Korsgren O, Foss A, Kvalheim G, Scholz H. Human adipose-derived mesenchymal stem cells respond to short-term hypoxia by secreting factors beneficial for human islets in vitro and potentiate antidiabetic effect in vivo. Cell Med. 2017;9(3):103–16.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Stamati K, Mudera V, Cheema U. Evolution of oxygen utilization in multicellular organisms and implications for cell signalling in tissue engineering. J Tissue Eng. 2011;2(1):2041731411432365.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Brahimi-Horn MC, Pouyssegur J. Oxygen, a source of life and stress. FEBS Lett. 2007;581(19):3582–91.

    Article  CAS  PubMed  Google Scholar 

  43. Jaakkola P, Mole DR, Tian YM, Wilson MI, Gielbert J, Gaskell SJ, von Kriegsheim A, Hebestreit HF, Mukherji M, Schofield CJ, Maxwell PH, Pugh CW, Ratcliffe PJ. Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science. 2001;292(5516):468–72.

    Article  CAS  PubMed  Google Scholar 

  44. Weidemann A, Johnson RS. Biology of HIF-1alpha. Cell Death Differ. 2008;15(4):621–7.

    Article  CAS  PubMed  Google Scholar 

  45. Choi JR, Pingguan-Murphy B, Wan Abas WA, Noor Azmi MA, Omar SZ, Chua KH, Wan Safwani WK. Impact of low oxygen tension on stemness, proliferation and differentiation potential of human adipose-derived stem cells. Biochem Biophys Res Commun. 2014;448(2):218–24.

    Article  CAS  PubMed  Google Scholar 

  46. Yamamoto Y, Fujita M, Tanaka Y, Kojima I, Kanatani Y, Ishihara M, Tachibana S. Low oxygen tension enhances proliferation and maintains stemness of adipose tissue-derived stromal cells. Biores Open Access. 2013;2(3):199–205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Feng Y, Zhu M, Dangelmajer S, Lee YM, Wijesekera O, Castellanos CX, Denduluri A, Chaichana KL, Li Q, Zhang H, Levchenko A, Guerrero-Cazares H, Quinones-Hinojosa A. Hypoxia-cultured human adipose-derived mesenchymal stem cells are non-oncogenic and have enhanced viability, motility, and tropism to brain cancer. Cell Death Dis. 2015;6:e1797.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kakudo N, Morimoto N, Ogawa T, Taketani S, Kusumoto K. Hypoxia enhances proliferation of human adipose-derived stem cells via HIF-1a activation. PLoS ONE. 2015;10(10):e0139890.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Kim JH, Park SH, Park SG, Choi JS, Xia Y, Sung JH. The pivotal role of reactive oxygen species generation in the hypoxia-induced stimulation of adipose-derived stem cells. Stem Cells Dev. 2011;20(10):1753–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hu X, Wei L, Taylor TM, Wei J, Zhou X, Wang JA, Yu SP. Hypoxic preconditioning enhances bone marrow mesenchymal stem cell migration via Kv2.1 channel and FAK activation. Am J Physiol Cell Physiol. 2011;301(2):C362–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rosova I, Dao M, Capoccia B, Link D, Nolta JA. Hypoxic preconditioning results in increased motility and improved therapeutic potential of human mesenchymal stem cells. Stem Cells. 2008;26(8):2173–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Chacko SM, Ahmed S, Selvendiran K, Kuppusamy ML, Khan M, Kuppusamy P. Hypoxic preconditioning induces the expression of prosurvival and proangiogenic markers in mesenchymal stem cells. Am J Physiol Cell Physiol. 2010;299(6):C1562–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Han YS, Lee JH, Yoon YM, Yun CW, Noh H, Lee SH. Hypoxia-induced expression of cellular prion protein improves the therapeutic potential of mesenchymal stem cells. Cell Death Dis. 2016;7(10):e2395.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Engela AU, Baan CC, Peeters AM, Weimar W, Hoogduijn MJ. Interaction between adipose tissue-derived mesenchymal stem cells and regulatory T-cells. Cell Transplant. 2013;22(1):41–54.

    Article  PubMed  Google Scholar 

  55. Xiang C, Xie QP. Protection of mouse pancreatic islet function by coculture with hypoxia pretreated mesenchymal stromal cells. Mol Med Rep. 2018;18(3):2589–98.

    CAS  PubMed  Google Scholar 

  56. Roemeling-vanRhijn M, Mensah FK, Korevaar SS, Leijs MJ, van Osch GJ, Ijzermans JN, Betjes MG, Baan CC, Weimar W, Hoogduijn MJ. Effects of hypoxia on the immunomodulatory properties of adipose tissue-derived mesenchymal stem cells. Front Immunol. 2013;4:203.

    Article  CAS  Google Scholar 

  57. Saparov A, Ogay V, Nurgozhin T, Jumabay M, Chen WC. Preconditioning of human mesenchymal stem cells to enhance their regulation of the immune response. Stem Cells Int. 2016;2016:3924858.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Silva LHA, Antunes MA, Dos Santos CC, Weiss DJ, Cruz FF, Rocco PRM. Strategies to improve the therapeutic effects of mesenchymal stromal cells in respiratory diseases. Stem Cell Res Ther. 2018;9(1):45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Hu X, Yu SP, Fraser JL, Lu Z, Ogle ME, Wang JA, Wei L. Transplantation of hypoxia-preconditioned mesenchymal stem cells improves infarcted heart function via enhanced survival of implanted cells and angiogenesis. J Thorac Cardiovasc Surg. 2008;135(4):799–808.

    Article  CAS  PubMed  Google Scholar 

  60. Follin B, Juhl M, Cohen S, Pedersen AE, Kastrup J, Ekblond A. Increased paracrine immunomodulatory potential of mesenchymal stromal cells in three-dimensional culture. Tissue Eng Part B Rev. 2016;22(4):322–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Madl CM, Heilshorn SC. Engineering hydrogel microenvironments to recapitulate the stem cell niche. Annu Rev Biomed Eng. 2018;20:21–47.

    Article  CAS  PubMed  Google Scholar 

  62. Lech W, Sarnowska A, Kuczynska Z, Dabrowski F, Figiel-Dabrowska A, Domanska-Janik K, Buzanska L, Zychowicz M. Biomimetic microenvironmental preconditioning enhance neuroprotective properties of human mesenchymal stem cells derived from Wharton’s Jelly (WJ-MSCs). Sci Rep. 2020;10(1):16946.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Guo L, Zhou Y, Wang S, Wu Y. Epigenetic changes of mesenchymal stem cells in three-dimensional (3D) spheroids. J Cell Mol Med. 2014;18(10):2009–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Cheng NC, Wang S, Young TH. The influence of spheroid formation of human adipose-derived stem cells on chitosan films on stemness and differentiation capabilities. Biomaterials. 2012;33(6):1748–58.

    Article  CAS  PubMed  Google Scholar 

  65. Zhang S, Liu P, Chen L, Wang Y, Wang Z, Zhang B. The effects of spheroid formation of adipose-derived stem cells in a microgravity bioreactor on stemness properties and therapeutic potential. Biomaterials. 2015;41:15–25.

    Article  PubMed  CAS  Google Scholar 

  66. Cheng NC, Chen SY, Li JR, Young TH. Short-term spheroid formation enhances the regenerative capacity of adipose-derived stem cells by promoting stemness, angiogenesis, and chemotaxis. Stem Cells Transl Med. 2013;2(8):584–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Potapova IA, Brink PR, Cohen IS, Doronin SV. Culturing of human mesenchymal stem cells as three-dimensional aggregates induces functional expression of CXCR4 that regulates adhesion to endothelial cells. J Biol Chem. 2008;283(19):13100–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Burand AJ Jr, Di L, Boland LK, Boyt DT, Schrodt MV, Santillan DA, Ankrum JA. Aggregation of human mesenchymal stromal cells eliminates their ability to suppress human T cells. Front Immunol. 2020;11:143.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Bartosh TJ, Ylostalo JH, Bazhanov N, Kuhlman J, Prockop DJ. Dynamic compaction of human mesenchymal stem/precursor cells into spheres self-activates caspase-dependent IL1 signaling to enhance secretion of modulators of inflammation and immunity (PGE2, TSG6, and STC1). Stem Cells. 2013;31(11):2443–56.

    Article  CAS  PubMed  Google Scholar 

  70. Bartosh TJ, Ylostalo JH, Mohammadipoor A, Bazhanov N, Coble K, Claypool K, Lee RH, Choi H, Prockop DJ. Aggregation of human mesenchymal stromal cells (MSCs) into 3D spheroids enhances their antiinflammatory properties. Proc Natl Acad Sci USA. 2010;107(31):13724–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Ylostalo JH, Bartosh TJ, Tiblow A, Prockop DJ. Unique characteristics of human mesenchymal stromal/progenitor cells pre-activated in 3-dimensional cultures under different conditions. Cytotherapy. 2014;16(11):1486–500.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Zimmermann JA, McDevitt TC. Pre-conditioning mesenchymal stromal cell spheroids for immunomodulatory paracrine factor secretion. Cytotherapy. 2014;16(3):331–45.

    Article  CAS  PubMed  Google Scholar 

  73. Potapova IA, Gaudette GR, Brink PR, Robinson RB, Rosen MR, Cohen IS, Doronin SV. Mesenchymal stem cells support migration, extracellular matrix invasion, proliferation, and survival of endothelial cells in vitro. Stem Cells. 2007;25(7):1761–8.

    Article  CAS  PubMed  Google Scholar 

  74. Bhang SH, Lee S, Shin JY, Lee TJ, Kim BS. Transplantation of cord blood mesenchymal stem cells as spheroids enhances vascularization. Tissue Eng Part A. 2012;18(19–20):2138–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Kouroupis D, Correa D. Increased mesenchymal stem cell functionalization in three-dimensional manufacturing settings for enhanced therapeutic applications. Front Bioeng Biotechnol. 2021;9:621748.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Barczyk M, Carracedo S, Gullberg D. Integrins. Cell Tissue Res. 2010;339(1):269–80.

    Article  CAS  PubMed  Google Scholar 

  77. Kwon SG, Kwon YW, Lee TW, Park GT, Kim JH. Recent advances in stem cell therapeutics and tissue engineering strategies. Biomater Res. 2018;22:36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Li J, Liu Y, Zhang Y, Yao B, Li Z, Song W, Wang Y, Duan X, Yuan X, Fu X, Huang S. Biophysical and biochemical cues of biomaterials guide mesenchymal stem cell behaviors. Front Cell Dev Biol. 2021;9:640388.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Caliari SR, Harley BA. Collagen-GAG scaffold biophysical properties bias MSC lineage choice in the presence of mixed soluble signals. Tissue Eng Part A. 2014;20(17–18):2463–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Lee EJ, Kasper FK, Mikos AG. Biomaterials for tissue engineering. Ann Biomed Eng. 2014;42(2):323–37.

    Article  PubMed  Google Scholar 

  81. Shao Z, Zhang X, Pi Y, Wang X, Jia Z, Zhu J, Dai L, Chen W, Yin L, Chen H, Zhou C, Ao Y. Polycaprolactone electrospun mesh conjugated with an MSC affinity peptide for MSC homing in vivo. Biomaterials. 2012;33(12):3375–87.

    Article  CAS  PubMed  Google Scholar 

  82. Fauzi MB, Lokanathan Y, Aminuddin BS, Ruszymah BHI, Chowdhury SR. Ovine tendon collagen: extraction, characterisation and fabrication of thin films for tissue engineering applications. Mater Sci Eng C Mater Biol Appl. 2016;68:163–71.

    Article  CAS  PubMed  Google Scholar 

  83. Gurumurthy B, Bierdeman PC, Janorkar AV. Composition of elastin like polypeptide-collagen composite scaffold influences in vitro osteogenic activity of human adipose derived stem cells. Dent Mater. 2016;32(10):1270–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Offeddu GS, Ashworth JC, Cameron RE, Oyen ML. Structural determinants of hydration, mechanics and fluid flow in freeze-dried collagen scaffolds. Acta Biomater. 2016;41:193–203.

    Article  CAS  PubMed  Google Scholar 

  85. Ewa-Choy YW, Pingguan-Murphy B, Abdul-Ghani NA, Jahendran J, Chua KH. Effect of alginate concentration on chondrogenesis of co-cultured human adipose-derived stem cells and nasal chondrocytes: a biological study. Biomater Res. 2017;21:19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Blair HC, Teitelbaum SL, Ghiselli R, Gluck S. Osteoclastic bone resorption by a polarized vacuolar proton pump. Science. 1989;245(4920):855–7.

    Article  CAS  PubMed  Google Scholar 

  87. Teitelbaum SL. Bone resorption by osteoclasts. Science. 2000;289(5484):1504–8.

    Article  CAS  PubMed  Google Scholar 

  88. Kook YJ, Lee DH, Song JE, Tripathy N, Jeon YS, Jeon HY, Oliveira JM, Reis RL, Khang G. Osteogenesis evaluation of duck’s feet-derived collagen/hydroxyapatite sponges immersed in dexamethasone. Biomater Res. 2017;21:2.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Valles G, Bensiamar F, Crespo L, Arruebo M, Vilaboa N, Saldana L. Topographical cues regulate the crosstalk between MSCs and macrophages. Biomaterials. 2015;37:124–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Stucky EC, Schloss RS, Yarmush ML, Shreiber DI. Alginate micro-encapsulation of mesenchymal stromal cells enhances modulation of the neuro-inflammatory response. Cytotherapy. 2015;17(10):1353–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Follin B, Juhl M, Cohen S, Pedersen AE, Gad M, Kastrup J, Ekblond A. Human adipose-derived stromal cells in a clinically applicable injectable alginate hydrogel: phenotypic and immunomodulatory evaluation. Cytotherapy. 2015;17(8):1104–18.

    Article  CAS  PubMed  Google Scholar 

  92. Barminko J, Kim JH, Otsuka S, Gray A, Schloss R, Grumet M, Yarmush ML. Encapsulated mesenchymal stromal cells for in vivo transplantation. Biotechnol Bioeng. 2011;108(11):2747–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Li LM, Han M, Jiang XC, Yin XZ, Chen F, Zhang TY, Ren H, Zhang JW, Hou TJ, Chen Z, Ou-Yang HW, Tabata Y, Shen YQ, Gao JQ. Peptide-tethered hydrogel scaffold promotes recovery from spinal cord transection via synergism with mesenchymal stem cells. ACS Appl Mater Interfaces. 2017;9(4):3330–42.

    Article  CAS  PubMed  Google Scholar 

  94. Jiao D, Zheng A, Liu Y, Zhang X, Wang X, Wu J, She W, Lv K, Cao L, Jiang X. Bidirectional differentiation of BMSCs induced by a biomimetic procallus based on a gelatin-reduced graphene oxide reinforced hydrogel for rapid bone regeneration. Bioact Mater. 2021;6(7):2011–28.

    Article  CAS  PubMed  Google Scholar 

  95. Lee BC, Kang KS. Functional enhancement strategies for immunomodulation of mesenchymal stem cells and their therapeutic application. Stem Cell Res Ther. 2020;11(1):397.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Linares GR, Chiu CT, Scheuing L, Leng Y, Liao HM, Maric D, Chuang DM. Preconditioning mesenchymal stem cells with the mood stabilizers lithium and valproic acid enhances therapeutic efficacy in a mouse model of Huntington’s disease. Exp Neurol. 2016;281:81–92.

    Article  CAS  PubMed  Google Scholar 

  97. Wisel S, Khan M, Kuppusamy ML, Mohan IK, Chacko SM, Rivera BK, Sun BC, Hideg K, Kuppusamy P. Pharmacological preconditioning of mesenchymal stem cells with trimetazidine (1-[2,3,4-trimethoxybenzyl]piperazine) protects hypoxic cells against oxidative stress and enhances recovery of myocardial function in infarcted heart through Bcl-2 expression. J Pharmacol Exp Ther. 2009;329(2):543–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Lee BC, Yu KR. Impact of mesenchymal stem cell senescence on inflammaging. BMB Rep. 2020;53(2):65–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Turinetto V, Vitale E, Giachino C. Senescence in human mesenchymal stem cells: functional changes and implications in stem cell-based therapy. Int J Mol Sci. 2016;17(7):1164.

    Article  PubMed Central  CAS  Google Scholar 

  100. Yu KR, Lee JY, Kim HS, Hong IS, Choi SW, Seo Y, Kang I, Kim JJ, Lee BC, Lee S, Kurtz A, Seo KW, Kang KS. A p38 MAPK-mediated alteration of COX-2/PGE2 regulates immunomodulatory properties in human mesenchymal stem cell aging. PLoS ONE. 2014;9(8):e102426.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  101. Katsuda T, Kosaka N, Takeshita F, Ochiya T. The therapeutic potential of mesenchymal stem cell-derived extracellular vesicles. Proteomics. 2013;13(10–11):1637–53.

    Article  CAS  PubMed  Google Scholar 

  102. Locksley RM, Killeen N, Lenardo MJ. The TNF and TNF receptor superfamilies: integrating mammalian biology. Cell. 2001;104(4):487–501.

    Article  CAS  PubMed  Google Scholar 

  103. Nakao Y, Fukuda T, Zhang Q, Sanui T, Shinjo T, Kou X, Chen C, Liu D, Watanabe Y, Hayashi C, Yamato H, Yotsumoto K, Tanaka U, Taketomi T, Uchiumi T, Le AD, Shi S, Nishimura F. Exosomes from TNF-alpha-treated human gingiva-derived MSCs enhance M2 macrophage polarization and inhibit periodontal bone loss. Acta Biomater. 2021;122:306–24.

    Article  CAS  PubMed  Google Scholar 

  104. Bai X, Xi J, Bi Y, Zhao X, Bing W, Meng X, Liu Y, Zhu Z, Song G. TNF-alpha promotes survival and migration of MSCs under oxidative stress via NF-kappaB pathway to attenuate intimal hyperplasia in vein grafts. J Cell Mol Med. 2017;21(9):2077–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Heo SC, Jeon ES, Lee IH, Kim HS, Kim MB, Kim JH. Tumor necrosis factor-alpha-activated human adipose tissue-derived mesenchymal stem cells accelerate cutaneous wound healing through paracrine mechanisms. J Invest Dermatol. 2011;131(7):1559–67.

    Article  CAS  PubMed  Google Scholar 

  106. Kwon YW, Heo SC, Jeong GO, Yoon JW, Mo WM, Lee MJ, Jang IH, Kwon SM, Lee JS, Kim JH. Tumor necrosis factor-alpha-activated mesenchymal stem cells promote endothelial progenitor cell homing and angiogenesis. Biochim Biophys Acta. 2013;1832(12):2136–44.

    Article  CAS  PubMed  Google Scholar 

  107. Krampera M, Cosmi L, Angeli R, Pasini A, Liotta F, Andreini A, Santarlasci V, Mazzinghi B, Pizzolo G, Vinante F, Romagnani P, Maggi E, Romagnani S, Annunziato F. Role for interferon-gamma in the immunomodulatory activity of human bone marrow mesenchymal stem cells. Stem Cells. 2006;24(2):386–98.

    Article  CAS  PubMed  Google Scholar 

  108. Croitoru-Lamoury J, Lamoury FM, Caristo M, Suzuki K, Walker D, Takikawa O, Taylor R, Brew BJ. Interferon-gamma regulates the proliferation and differentiation of mesenchymal stem cells via activation of indoleamine 2,3 dioxygenase (IDO). PLoS ONE. 2011;6(2):e14698.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Noone C, Kihm A, English K, O’Dea S, Mahon BP. IFN-gamma stimulated human umbilical-tissue-derived cells potently suppress NK activation and resist NK-mediated cytotoxicity in vitro. Stem Cells Dev. 2013;22(22):3003–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Ryan JM, Barry F, Murphy JM, Mahon BP. Interferon-gamma does not break, but promotes the immunosuppressive capacity of adult human mesenchymal stem cells. Clin Exp Immunol. 2007;149(2):353–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Lee HJ, Kim HD, Jo CH, Bok EY, Kim SB, Lee SL, Jang M, Bae SG, Yun SH, Kim SJ, Rho GJ, Lee WJ. IFN-gamma licensing does not enhance the reduced immunomodulatory potential and migratory ability of differentiation-induced porcine bone marrow-derived mesenchymal stem cells in an in vitro xenogeneic application. Biomed Res Int. 2021;2021:4604856.

    PubMed  PubMed Central  Google Scholar 

  112. Kronsteiner B, Wolbank S, Peterbauer A, Hackl C, Redl H, van Griensven M, Gabriel C. Human mesenchymal stem cells from adipose tissue and amnion influence T-cells depending on stimulation method and presence of other immune cells. Stem Cells Dev. 2011;20(12):2115–26.

    Article  CAS  PubMed  Google Scholar 

  113. Takeuchi S, Tsuchiya A, Iwasawa T, Nojiri S, Watanabe T, Ogawa M, Yoshida T, Fujiki K, Koui Y, Kido T, Yoshioka Y, Fujita M, Kikuta J, Itoh T, Takamura M, Shirahige K, Ishii M, Ochiya T, Miyajima A, Terai S. Small extracellular vesicles derived from interferon-gamma pre-conditioned mesenchymal stromal cells effectively treat liver fibrosis. NPJ Regen Med. 2021;6(1):19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Zheng G, Qiu G, Ge M, He J, Huang L, Chen P, Wang W, Xu Q, Hu Y, Shu Q, Xu J. Human adipose-derived mesenchymal stem cells alleviate obliterative bronchiolitis in a murine model via IDO. Respir Res. 2017;18(1):119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  115. Kanai R, Nakashima A, Doi S, Kimura T, Yoshida K, Maeda S, Ishiuchi N, Yamada Y, Ike T, Doi T, Kato Y, Masaki T. Interferon-gamma enhances the therapeutic effect of mesenchymal stem cells on experimental renal fibrosis. Sci Rep. 2021;11(1):850.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Dinarello CA. Interleukin-1 in the pathogenesis and treatment of inflammatory diseases. Blood. 2011;117(14):3720–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Lin FS, Lin CC, Chien CS, Luo SF, Yang CM. Involvement of p42/p44 MAPK, JNK, and NF-kappaB in IL-1beta-induced ICAM-1 expression in human pulmonary epithelial cells. J Cell Physiol. 2005;202(2):464–73.

    Article  CAS  PubMed  Google Scholar 

  118. Nie H, An F, Mei J, Yang C, Zhan Q, Zhang Q. IL-1beta pretreatment improves the efficacy of mesenchymal stem cells on acute liver failure by enhancing CXCR4 expression. Stem Cells Int. 2020;2020:1498315.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  119. Yao M, Cui B, Zhang W, Ma W, Zhao G, Xing L. Exosomal miR-21 secreted by IL-1beta-primed-mesenchymal stem cells induces macrophage M2 polarization and ameliorates sepsis. Life Sci. 2021;264:118658.

    Article  CAS  PubMed  Google Scholar 

  120. Jiang Y, Yang M, Zhang Y, Huang Y, Wu J, Xie Y, Wei Q, Liao Z, Gu J. Dynamics of adaptive immune cell and NK cell subsets in patients with ankylosing spondylitis after IL-17A inhibition by secukinumab. Front Pharmacol. 2021;12:738316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Sivanathan KN, Rojas-Canales DM, Hope CM, Krishnan R, Carroll RP, Gronthos S, Grey ST, Coates PT. Interleukin-17A-induced human mesenchymal stem cells are superior modulators of immunological function. Stem Cells. 2015;33(9):2850–63.

    Article  CAS  PubMed  Google Scholar 

  122. Cheng W, Su J, Hu Y, Huang Q, Shi H, Wang L, Ren J. Interleukin-25 primed mesenchymal stem cells achieve better therapeutic effects on dextran sulfate sodium-induced colitis via inhibiting Th17 immune response and inducing T regulatory cell phenotype. Am J Transl Res. 2017;9(9):4149–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Kempuraj D, Frydas S, Conti P, Kandere-Grzybowska K, Boucher W, Letourneau R, Madhappan B, Huang SH, Sugimoto K, Papadopoulou NG, Christodoulou S, Theoharides TC. Interleukin-25 (or IL-17E): a new IL-17 family member with growth factor/inflammatory actions. Int J Immunopathol Pharmacol. 2003;16(3):185–8.

    Article  CAS  PubMed  Google Scholar 

  124. Hou YS, Liu LY, Chai JK, Yu YH, Duan HJ, Hu Q, Yin HN, Wang YH, Zhuang SB, Fan J, Chu WL, Ma L. Lipopolysaccharide pretreatment inhibits LPS-induced human umbilical cord mesenchymal stem cell apoptosis via upregulating the expression of cellular FLICE-inhibitory protein. Mol Med Rep. 2015;12(2):2521–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Chu X, Xu B, Gao H, Li BY, Liu Y, Reiter JL, Wang Y. Lipopolysaccharides improve mesenchymal stem cell-mediated cardioprotection by MyD88 and stat3 signaling in a mouse model of cardiac ischemia/reperfusion injury. Stem Cells Dev. 2019;28(9):620–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Yao Y, Zhang F, Wang L, Zhang G, Wang Z, Chen J, Gao X. Lipopolysaccharide preconditioning enhances the efficacy of mesenchymal stem cells transplantation in a rat model of acute myocardial infarction. J Biomed Sci. 2009;16:74.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  127. Han X, Yang Q, Lin L, Xu C, Zheng C, Chen X, Han Y, Li M, Cao W, Cao K, Chen Q, Xu G, Zhang Y, Zhang J, Schneider RJ, Qian Y, Wang Y, Brewer G, Shi Y. Interleukin-17 enhances immunosuppression by mesenchymal stem cells. Cell Death Differ. 2014;21(11):1758–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Kurogoushi R, Hasegawa T, Akazawa Y, Iwata K, Sugimoto A, Yamaguchi-Ueda K, Miyazaki A, Narwidina A, Kawarabayashi K, Kitamura T, Nakagawa H, Iwasaki T, Iwamoto T. Fibroblast growth factor 2 suppresses the expression of C-C motif chemokine 11 through the c-Jun N-terminal kinase pathway in human dental pulp-derived mesenchymal stem cells. Exp Ther Med. 2021;22(6):1356.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Gorin C, Rochefort GY, Bascetin R, Ying H, Lesieur J, Sadoine J, Beckouche N, Berndt S, Novais A, Lesage M, Hosten B, Vercellino L, Merlet P, Le-Denmat D, Marchiol C, Letourneur D, Nicoletti A, Vital SO, Poliard A, Salmon B, Muller L, Chaussain C, Germain S. Priming dental pulp stem cells with fibroblast growth factor-2 increases angiogenesis of implanted tissue-engineered constructs through hepatocyte growth factor and vascular endothelial growth factor secretion. Stem Cells Transl Med. 2016;5(3):392–404.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Endo K, Fujita N, Nakagawa T, Nishimura R. Effect of fibroblast growth factor-2 and serum on canine mesenchymal stem cell chondrogenesis. Tissue Eng Part A. 2019;25(11–12):901–10.

    Article  CAS  PubMed  Google Scholar 

  131. Ratajczak MZ, Zuba-Surma E, Kucia M, Reca R, Wojakowski W, Ratajczak J. The pleiotropic effects of the SDF-1-CXCR4 axis in organogenesis, regeneration and tumorigenesis. Leukemia. 2006;20(11):1915–24.

    Article  CAS  PubMed  Google Scholar 

  132. Liu X, Duan B, Cheng Z, Jia X, Mao L, Fu H, Che Y, Ou L, Liu L, Kong D. SDF-1/CXCR4 axis modulates bone marrow mesenchymal stem cell apoptosis, migration and cytokine secretion. Protein Cell. 2011;2(10):845–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Marquez-Curtis LA, Janowska-Wieczorek A. Enhancing the migration ability of mesenchymal stromal cells by targeting the SDF-1/CXCR4 axis. Biomed Res Int. 2013;2013:561098.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Esmaeili R, Darbandi-Azar A, Sadeghpour A, Majidzadeh AK, Eini L, Jafarbeik-Iravani N, Hoseinpour P, Vajhi A, Oghabi Bakhshaiesh T, Masoudkabir F, Sadeghizadeh M. Mesenchymal stem cells pretreatment with stromal-derived factor-1 alpha augments cardiac function and angiogenesis in infarcted myocardium. Am J Med Sci. 2021;361(6):765–75.

    Article  PubMed  Google Scholar 

  135. Herrmann JL, Wang Y, Abarbanell AM, Weil BR, Tan J, Meldrum DR. Preconditioning mesenchymal stem cells with transforming growth factor-alpha improves mesenchymal stem cell-mediated cardioprotection. Shock. 2010;33(1):24–30.

    Article  CAS  PubMed  Google Scholar 

  136. Hahn JY, Cho HJ, Kang HJ, Kim TS, Kim MH, Chung JH, Bae JW, Oh BH, Park YB, Kim HS. Pre-treatment of mesenchymal stem cells with a combination of growth factors enhances gap junction formation, cytoprotective effect on cardiomyocytes, and therapeutic efficacy for myocardial infarction. J Am Coll Cardiol. 2008;51(9):933–43.

    Article  CAS  PubMed  Google Scholar 

  137. Wei N, Yu SP, Gu X, Taylor TM, Song D, Liu XF, Wei L. Delayed intranasal delivery of hypoxic-preconditioned bone marrow mesenchymal stem cells enhanced cell homing and therapeutic benefits after ischemic stroke in mice. Cell Transplant. 2013;22(6):977–91.

    Article  PubMed  Google Scholar 

  138. Haider H, Lee YJ, Jiang S, Ahmed RP, Ryon M, Ashraf M. Phosphodiesterase inhibition with tadalafil provides longer and sustained protection of stem cells. Am J Physiol Heart Circ Physiol. 2010;299(5):H1395–404.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Elmadbouh I, Ashraf M. Tadalafil, a long acting phosphodiesterase inhibitor, promotes bone marrow stem cell survival and their homing into ischemic myocardium for cardiac repair. Physiol Rep. 2017;5(21):e13480.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Song L, Yang YJ, Dong QT, Qian HY, Gao RL, Qiao SB, Shen R, He ZX, Lu MJ, Zhao SH, Geng YJ, Gersh BJ. Atorvastatin enhance efficacy of mesenchymal stem cells treatment for swine myocardial infarction via activation of nitric oxide synthase. PLoS ONE. 2013;8(5):e65702.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Niki E, Traber MG. A history of vitamin E. Ann Nutr Metab. 2012;61(3):207–12.

    Article  CAS  PubMed  Google Scholar 

  142. Bhatti FU, Mehmood A, Latief N, Zahra S, Cho H, Khan SN, Riazuddin S. Vitamin E protects rat mesenchymal stem cells against hydrogen peroxide-induced oxidative stress in vitro and improves their therapeutic potential in surgically-induced rat model of osteoarthritis. Osteoarthritis Cartilage. 2017;25(2):321–31.

    Article  CAS  PubMed  Google Scholar 

  143. Ocansey DKW, Pei B, Yan Y, Qian H, Zhang X, Xu W, Mao F. Improved therapeutics of modified mesenchymal stem cells: an update. J Transl Med. 2020;18(1):42.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Chen F, Li L, Tian DD. Salvia miltiorrhiza roots against cardiovascular disease: consideration of herb–drug interactions. Biomed Res Int. 2017;2017:9868694.

    PubMed  PubMed Central  Google Scholar 

  145. Lam BY, Lo AC, Sun X, Luo HW, Chung SK, Sucher NJ. Neuroprotective effects of tanshinones in transient focal cerebral ischemia in mice. Phytomedicine. 2003;10(4):286–91.

    Article  CAS  PubMed  Google Scholar 

  146. Vakili A, Kataoka H, Plesnila N. Role of arginine vasopressin V1 and V2 receptors for brain damage after transient focal cerebral ischemia. J Cereb Blood Flow Metab. 2005;25(8):1012–9.

    Article  CAS  PubMed  Google Scholar 

  147. Kim R, Lee S, Lee CY, Yun H, Lee H, Lee MY, Kim J, Jeong JY, Baek K, Chang W. Salvia miltiorrhiza enhances the survival of mesenchymal stem cells under ischemic conditions. J Pharm Pharmacol. 2018;70(9):1228–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Ke S, Zhang Y, Lan Z, Li S, Zhu W, Liu L. Curcumin protects murine lung mesenchymal stem cells from H2O2 by modulating the Akt/Nrf2/HO-1 pathway. J Int Med Res. 2020;48(4):300060520910665.

    Article  CAS  PubMed  Google Scholar 

  149. Pirmoradi S, Fathi E, Farahzadi R, Pilehvar-Soltanahmadi Y, Zarghami N. Curcumin affects adipose tissue-derived mesenchymal stem cell aging through TERT gene expression. Drug Res (Stuttg). 2018;68(4):213–21.

    Article  CAS  Google Scholar 

  150. Cremers NA, Lundvig DM, van Dalen SC, Schelbergen RF, van Lent PL, Szarek WA, Regan RF, Carels CE, Wagener FA. Curcumin-induced heme oxygenase-1 expression prevents H2O2-induced cell death in wild type and heme oxygenase-2 knockout adipose-derived mesenchymal stem cells. Int J Mol Sci. 2014;15(10):17974–99.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  151. Xiao Y, Xia J, Wu S, Lv Z, Huang S, Huang H, Su X, Cheng J, Ke Y. Curcumin inhibits acute vascular inflammation through the activation of heme oxygenase-1. Oxid Med Cell Longev. 2018;2018:3295807.

    PubMed  PubMed Central  Google Scholar 

  152. Liu J, Zhu P, Song P, Xiong W, Chen H, Peng W, Wang S, Li S, Fu Z, Wang Y, Wang H. Pretreatment of adipose derived stem cells with curcumin facilitates myocardial recovery via antiapoptosis and angiogenesis. Stem Cells Int. 2015;2015:638153.

    PubMed  PubMed Central  Google Scholar 

  153. Azam M, Ghufran H, Butt H, Mehmood A, Ashfaq R, Ilyas AM, Ahmad MR, Riazuddin S. Curcumin preconditioning enhances the efficacy of adipose-derived mesenchymal stem cells to accelerate healing of burn wounds. Burns Trauma. 2021;9:tkab021.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Chen H, Huang Y, Huang D, Wu Z, Li Y, Zhou C, Wei G. Protective effect of gigantol against hydrogen peroxideinduced apoptosis in rat bone marrow mesenchymal stem cells through the PI3K/Akt pathway. Mol Med Rep. 2018;17(2):3267–73.

    CAS  PubMed  Google Scholar 

  155. Fang Y, Chu L, Li L, Wang J, Yang Y, Gu J, Zhang J. Tetramethylpyrazine protects bone marrow-derived mesenchymal stem cells against hydrogen peroxide-induced apoptosis through PI3K/Akt and ERK1/2 pathways. Biol Pharm Bull. 2017;40(12):2146–52.

    Article  CAS  PubMed  Google Scholar 

  156. Lin LZ, Chen HH, Lei ZX, Li YR, Zhou CH, Huang YC, Wei G. Rosmarinic acid protects on rat bone marrow mesenchymal stem cells from hydrogen peroxide-induced apoptosis. J Asian Nat Prod Res. 2018;20(6):570–80.

    Article  CAS  PubMed  Google Scholar 

  157. Oses C, Olivares B, Ezquer M, Acosta C, Bosch P, Donoso M, Leniz P, Ezquer F. Preconditioning of adipose tissue-derived mesenchymal stem cells with deferoxamine increases the production of pro-angiogenic, neuroprotective and anti-inflammatory factors: potential application in the treatment of diabetic neuropathy. PLoS ONE. 2017;12(5):e0178011.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  158. Park SM, An JH, Lee JH, Kim KB, Chae HK, Oh YI, Song WJ, Youn HY. Extracellular vesicles derived from DFO-preconditioned canine AT-MSCs reprogram macrophages into M2 phase. PLoS ONE. 2021;16(7):e0254657.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Smadja DM, Levy M, Huang L, Rossi E, Blandinieres A, Israel-Biet D, Gaussem P, Bischoff J. Treprostinil indirectly regulates endothelial colony forming cell angiogenic properties by increasing VEGF-A produced by mesenchymal stem cells. Thromb Haemost. 2015;114(4):735–47.

    PubMed  PubMed Central  Google Scholar 

  160. Pourjafar M, Saidijam M, Mansouri K, Ghasemibasir H, KarimiDermani F, Najafi R. All-trans retinoic acid preconditioning enhances proliferation, angiogenesis and migration of mesenchymal stem cell in vitro and enhances wound repair in vivo. Cell Prolif. 2017;50(1):e12315.

    Article  CAS  Google Scholar 

  161. Fu Y, Wang Y, Du L, Xu C, Cao J, Fan T, Liu J, Su X, Fan S, Liu Q, Fan F. Resveratrol inhibits ionising irradiation-induced inflammation in MSCs by activating SIRT1 and limiting NLRP-3 inflammasome activation. Int J Mol Sci. 2013;14(7):14105–18.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  162. Zhang A, Zhang X, Tan X, Cai B, Ge W, Dai G, Cai J. Resveratrol rescued the TNF-alpha-induced impairments of osteogenesis of bone-marrow derived mesenchymal stem cells and inhibited the TNF-alpha-activated NF-small ka CyrillicB signaling pathway. Int Immunopharmacol. 2015;26(2):409–15.

    Article  PubMed  CAS  Google Scholar 

  163. Jiang Q, Huang K, Lu F, Deng S, Yang Z, Hu S. Modifying strategies for SDF-1/CXCR4 interaction during mesenchymal stem cell transplantation. Gen Thorac Cardiovasc Surg. 2021.

  164. Zhang R, Yin L, Zhang B, Shi H, Sun Y, Ji C, Chen J, Wu P, Zhang L, Xu W, Qian H. Resveratrol improves human umbilical cord-derived mesenchymal stem cells repair for cisplatin-induced acute kidney injury. Cell Death Dis. 2018;9(10):965.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  165. Yang P, Chen A, Qin Y, Yin J, Cai X, Fan YJ, Li L, Huang HY. Buyang huanwu decoction combined with BMSCs transplantation promotes recovery after spinal cord injury by rescuing axotomized red nucleus neurons. J Ethnopharmacol. 2019;228:123–31.

    Article  PubMed  Google Scholar 

  166. Yang J, Gao F, Zhang Y, Liu Y, Zhang D. Buyang huanwu decoction (BYHWD) enhances angiogenic effect of mesenchymal stem cell by upregulating VEGF expression after focal cerebral ischemia. J Mol Neurosci. 2015;56(4):898–906.

    Article  CAS  PubMed  Google Scholar 

  167. Yang Z, Concannon J, Ng KS, Seyb K, Mortensen LJ, Ranganath S, Gu F, Levy O, Tong Z, Martyn K, Zhao W, Lin CP, Glicksman MA, Karp JM. Tetrandrine identified in a small molecule screen to activate mesenchymal stem cells for enhanced immunomodulation. Sci Rep. 2016;6:30263.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Zhou X, Hong Y, Zhang H, Li X. Mesenchymal stem cell senescence and rejuvenation: current status and challenges. Front Cell Dev Biol. 2020;8:364.

    Article  PubMed  PubMed Central  Google Scholar 

  169. Atashi F, Modarressi A, Pepper MS. The role of reactive oxygen species in mesenchymal stem cell adipogenic and osteogenic differentiation: a review. Stem Cells Dev. 2015;24(10):1150–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Chaudhari P, Ye Z, Jang YY. Roles of reactive oxygen species in the fate of stem cells. Antioxid Redox Signal. 2014;20(12):1881–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Zhang DY, Pan Y, Zhang C, Yan BX, Yu SS, Wu DL, Shi MM, Shi K, Cai XX, Zhou SS, Wang JB, Pan JP, Zhang LH. Wnt/beta-catenin signaling induces the aging of mesenchymal stem cells through promoting the ROS production. Mol Cell Biochem. 2013;374(1–2):13–20.

    Article  CAS  PubMed  Google Scholar 

  172. Zeng W, Xiao J, Zheng G, Xing F, Tipoe GL, Wang X, He C, Chen ZY, Liu Y. Antioxidant treatment enhances human mesenchymal stem cell anti-stress ability and therapeutic efficacy in an acute liver failure model. Sci Rep. 2015;5:11100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Zhang GW, Gu TX, Sun XJ, Wang C, Qi X, Wang XB, Li-Ling J. Edaravone promotes activation of resident cardiac stem cells by transplanted mesenchymal stem cells in a rat myocardial infarction model. J Thorac Cardiovasc Surg. 2016;152(2):570–82.

    Article  CAS  PubMed  Google Scholar 

  174. Khanh VC, Yamashita T, Ohneda K, Tokunaga C, Kato H, Osaka M, Hiramatsu Y, Ohneda O. Rejuvenation of mesenchymal stem cells by extracellular vesicles inhibits the elevation of reactive oxygen species. Sci Rep. 2020;10(1):17315.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Kornienko JS, Smirnova IS, Pugovkina NA, Ivanova JS, Shilina MA, Grinchuk TM, Shatrova AN, Aksenov ND, Zenin VV, Nikolsky NN, Lyublinskaya OG. High doses of synthetic antioxidants induce premature senescence in cultivated mesenchymal stem cells. Sci Rep. 2019;9(1):1296.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Avgustinova A, Benitah SA. Epigenetic control of adult stem cell function. Nat Rev Mol Cell Biol. 2016;17(10):643–58.

    Article  CAS  PubMed  Google Scholar 

  177. Cakouros D, Gronthos S. Epigenetic regulators of mesenchymal stem/stromal cell lineage determination. Curr Osteoporos Rep. 2020;18(5):597–605.

    Article  PubMed  PubMed Central  Google Scholar 

  178. Li Z, Liu C, Xie Z, Song P, Zhao RC, Guo L, Liu Z, Wu Y. Epigenetic dysregulation in mesenchymal stem cell aging and spontaneous differentiation. PLoS ONE. 2011;6(6):e20526.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Fernandez AF, Bayon GF, Urdinguio RG, Torano EG, Garcia MG, Carella A, Petrus-Reurer S, Ferrero C, Martinez-Camblor P, Cubillo I, Garcia-Castro J, Delgado-Calle J, Perez-Campo FM, Riancho JA, Bueno C, Menendez P, Mentink A, Mareschi K, Claire F, Fagnani C, Medda E, Toccaceli V, Brescianini S, Moran S, Esteller M, Stolzing A, de Boer J, Nistico L, Stazi MA, Fraga MF. H3K4me1 marks DNA regions hypomethylated during aging in human stem and differentiated cells. Genome Res. 2015;25(1):27–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Frobel J, Hemeda H, Lenz M, Abagnale G, Joussen S, Denecke B, Saric T, Zenke M, Wagner W. Epigenetic rejuvenation of mesenchymal stromal cells derived from induced pluripotent stem cells. Stem Cell Rep. 2014;3(3):414–22.

    Article  CAS  Google Scholar 

  181. Hynes K, Menicanin D, Han J, Marino V, Mrozik K, Gronthos S, Bartold PM. Mesenchymal stem cells from iPS cells facilitate periodontal regeneration. J Dent Res. 2013;92(9):833–9.

    Article  CAS  PubMed  Google Scholar 

  182. Spitzhorn LS, Megges M, Wruck W, Rahman MS, Otte J, Degistirici O, Meisel R, Sorg RV, Oreffo ROC, Adjaye J. Human iPSC-derived MSCs (iMSCs) from aged individuals acquire a rejuvenation signature. Stem Cell Res Ther. 2019;10(1):100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Ocampo A, Reddy P, Martinez-Redondo P, Platero-Luengo A, Hatanaka F, Hishida T, Li M, Lam D, Kurita M, Beyret E, Araoka T, Vazquez-Ferrer E, Donoso D, Roman JL, Xu J, Rodriguez Esteban C, Nunez G, Nunez Delicado E, Campistol JM, Guillen I, Guillen P, Izpisua Belmonte JC. In vivo amelioration of age-associated hallmarks by partial reprogramming. Cell. 2016;167(7):1719-1733 e12.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  184. Moore LD, Le T, Fan G. DNA methylation and its basic function. Neuropsychopharmacology. 2013;38(1):23–38.

    Article  CAS  PubMed  Google Scholar 

  185. Kornicka K, Marycz K, Maredziak M, Tomaszewski KA, Nicpon J. The effects of the DNA methyltranfserases inhibitor 5-Azacitidine on ageing, oxidative stress and DNA methylation of adipose derived stem cells. J Cell Mol Med. 2017;21(2):387–401.

    Article  CAS  PubMed  Google Scholar 

  186. Oh YS, Jeong SG, Cho GW. Anti-senescence effects of DNA methyltransferase inhibitor RG108 in human bone marrow mesenchymal stromal cells. Biotechnol Appl Biochem. 2015;62(5):583–90.

    Article  CAS  PubMed  Google Scholar 

  187. Gao B, Lin X, Jing H, Fan J, Ji C, Jie Q, Zheng C, Wang D, Xu X, Hu Y, Lu W, Luo Z, Yang L. Local delivery of tetramethylpyrazine eliminates the senescent phenotype of bone marrow mesenchymal stromal cells and creates an anti-inflammatory and angiogenic environment in aging mice. Aging Cell. 2018;17(3):e12741.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  188. Cakouros D, Isenmann S, Cooper L, Zannettino A, Anderson P, Glackin C, Gronthos S. Twist-1 induces Ezh2 recruitment regulating histone methylation along the Ink4A/Arf locus in mesenchymal stem cells. Mol Cell Biol. 2012;32(8):1433–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Li L, Chu L, Ren C, Wang J, Sun S, Li T, Yin Y. Enhanced migration of bone marrow-derived mesenchymal stem cells with tetramethylpyrazine and its synergistic effect on angiogenesis and neurogenesis after cerebral ischemia in rats. Stem Cells Dev. 2019;28(13):871–81.

    Article  CAS  PubMed  Google Scholar 

  190. Song X, Dai J, Li H, Li Y, Hao W, Zhang Y, Zhang Y, Su L, Wei H. Anti-aging effects exerted by Tetramethylpyrazine enhances self-renewal and neuronal differentiation of rat bMSCs by suppressing NF-kB signaling. Biosci Rep. 2019;39(6):BSR20190761.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Zhang L, Wang X, Lu X, Ma Y, Xin X, Xu X, Wang S, Hou Y. Tetramethylpyrazine enhanced the therapeutic effects of human umbilical cord mesenchymal stem cells in experimental autoimmune encephalomyelitis mice through Nrf2/HO-1 signaling pathway. Stem Cell Res Ther. 2020;11(1):186.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  192. Acar MB, Ayaz-Guner S, Gunaydin Z, Karakukcu M, Peluso G, Di Bernardo G, Ozcan S, Galderisi U. Proteomic and biological analysis of the effects of metformin senomorphics on the mesenchymal stromal cells. Front Bioeng Biotechnol. 2021;9:730813.

    Article  PubMed  PubMed Central  Google Scholar 

  193. Grezella C, Fernandez-Rebollo E, Franzen J, Ventura Ferreira MS, Beier F, Wagner W. Effects of senolytic drugs on human mesenchymal stromal cells. Stem Cell Res Ther. 2018;9(1):108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Suvakov S, Cubro H, White WM, Butler Tobah YS, Weissgerber TL, Jordan KL, Zhu XY, Woollard JR, Chebib FT, Milic NM, Grande JP, Xu M, Tchkonia T, Kirkland JL, Lerman LO, Garovic VD. Targeting senescence improves angiogenic potential of adipose-derived mesenchymal stem cells in patients with preeclampsia. Biol Sex Differ. 2019;10(1):49.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  195. Zhou Y, Xin X, Wang L, Wang B, Chen L, Liu O, Rowe DW, Xu M. Senolytics improve bone forming potential of bone marrow mesenchymal stem cells from aged mice. NPJ Regen Med. 2021;6(1):34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Hickson LJ, Langhi Prata LGP, Bobart SA, Evans TK, Giorgadze N, Hashmi SK, Herrmann SM, Jensen MD, Jia Q, Jordan KL, Kellogg TA, Khosla S, Koerber DM, Lagnado AB, Lawson DK, LeBrasseur NK, Lerman LO, McDonald KM, McKenzie TJ, Passos JF, Pignolo RJ, Pirtskhalava T, Saadiq IM, Schaefer KK, Textor SC, Victorelli SG, Volkman TL, Xue A, Wentworth MA, Wissler Gerdes EO, Zhu Y, Tchkonia T, Kirkland JL. Senolytics decrease senescent cells in humans: preliminary report from a clinical trial of Dasatinib plus Quercetin in individuals with diabetic kidney disease. EBioMedicine. 2019;47:446–56.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Luo R, Lu Y, Liu J, Cheng J, Chen Y. Enhancement of the efficacy of mesenchymal stem cells in the treatment of ischemic diseases. Biomed Pharmacother. 2019;109:2022–34.

    Article  CAS  PubMed  Google Scholar 

  198. Cakouros D, Gronthos S. Epigenetic regulation of bone marrow stem cell aging: revealing epigenetic signatures associated with hematopoietic and mesenchymal stem cell aging. Aging Dis. 2019;10(1):174–89.

    Article  PubMed  PubMed Central  Google Scholar 

  199. Kawakami T, Ando T, Kimura M, Wilson BS, Kawakami Y. Mast cells in atopic dermatitis. Curr Opin Immunol. 2009;21(6):666–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Cuerquis J, Romieu-Mourez R, Francois M, Routy JP, Young YK, Zhao J, Eliopoulos N. Human mesenchymal stromal cells transiently increase cytokine production by activated T cells before suppressing T-cell proliferation: effect of interferon-gamma and tumor necrosis factor-alpha stimulation. Cytotherapy. 2014;16(2):191–202.

    Article  CAS  PubMed  Google Scholar 

  201. Lee BC, Kim JJ, Lee JY, Kang I, Shin N, Lee SE, Choi SW, Cho JY, Kim HS, Kang KS. Disease-specific primed human adult stem cells effectively ameliorate experimental atopic dermatitis in mice. Theranostics. 2019;9(12):3608–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Su WR, Zhang QZ, Shi SH, Nguyen AL, Le AD. Human gingiva-derived mesenchymal stromal cells attenuate contact hypersensitivity via prostaglandin E2-dependent mechanisms. Stem Cells. 2011;29(11):1849–60.

    Article  CAS  PubMed  Google Scholar 

  203. Haque N, Kasim NH, Rahman MT. Optimization of pre-transplantation conditions to enhance the efficacy of mesenchymal stem cells. Int J Biol Sci. 2015;11(3):324–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Kim HJ, Choi W, San Lee J, Choi J, Choi N, Hwang KS. Clinical application of serological Alzheimer’s disease diagnosis using a highly sensitive biosensor with hydrogel-enhanced dielectrophoretic force. Biosens Bioelectron. 2022;195:113668.

    Article  CAS  PubMed  Google Scholar 

  205. Yuan D, Kuan T, Ling H, Wang H, Feng L, Zhao Q, Li J, Ran J. Serum metabolomics of end-stage renal disease patients with depression: potential biomarkers for diagnosis. Ren Fail. 2021;43(1):1479–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Tang W, Lv X, Huang J, Wang B, Lin L, Shen Y, Yao Y. Neuroprotective effect of stroke pretreated MSCs against cerebral ischemia/reperfusion injury in rats. World Neurosurg. 2021.

  207. Qi LL, Fan ZY, Mao HG, Wang JB. The therapeutic efficacy of adipose tissue-derived mesenchymal stem cell conditioned medium on experimental colitis was improved by the serum from colitis rats. Front Bioeng Biotechnol. 2021;9:694908.

    Article  PubMed  PubMed Central  Google Scholar 

  208. AlGhamdi KM, Kumar A, Moussa NA. Low-level laser therapy: a useful technique for enhancing the proliferation of various cultured cells. Lasers Med Sci. 2012;27(1):237–49.

    Article  PubMed  Google Scholar 

  209. Hawkins D, Abrahamse H. Effect of multiple exposures of low-level laser therapy on the cellular responses of wounded human skin fibroblasts. Photomed Laser Surg. 2006;24(6):705–14.

    Article  CAS  PubMed  Google Scholar 

  210. Kim H, Choi K, Kweon OK, Kim WH. Enhanced wound healing effect of canine adipose-derived mesenchymal stem cells with low-level laser therapy in athymic mice. J Dermatol Sci. 2012;68(3):149–56.

    Article  CAS  PubMed  Google Scholar 

  211. Yin K, Zhu R, Wang S, Zhao RC. Low-level laser effect on proliferation, migration, and antiapoptosis of mesenchymal stem cells. Stem Cells Dev. 2017;26(10):762–75.

    Article  CAS  PubMed  Google Scholar 

  212. Urnukhsaikhan E, Cho H, Mishig-Ochir T, Seo YK, Park JK. Pulsed electromagnetic fields promote survival and neuronal differentiation of human BM-MSCs. Life Sci. 2016;151:130–8.

    Article  CAS  PubMed  Google Scholar 

  213. Jiang B, Xiao W, Shi Y, Liu M, Xiao X. Heat shock pretreatment inhibited the release of Smac/DIABLO from mitochondria and apoptosis induced by hydrogen peroxide in cardiomyocytes and C2C12 myogenic cells. Cell Stress Chaperones. 2005;10(3):252–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Venkatakrishnan CD, Tewari AK, Moldovan L, Cardounel AJ, Zweier JL, Kuppusamy P, Ilangovan G. Heat shock protects cardiac cells from doxorubicin-induced toxicity by activating p38 MAPK and phosphorylation of small heat shock protein 27. Am J Physiol Heart Circ Physiol. 2006;291(6):H2680–91.

    Article  CAS  PubMed  Google Scholar 

  215. Qiao PF, Yao L, Zhang XC, Li GD, Wu DQ. Heat shock pretreatment improves stem cell repair following ischemia-reperfusion injury via autophagy. World J Gastroenterol. 2015;21(45):12822–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Chen X, Wang Q, Li X, Wang Q, Xie J, Fu X. Heat shock pretreatment of mesenchymal stem cells for inhibiting the apoptosis of ovarian granulosa cells enhanced the repair effect on chemotherapy-induced premature ovarian failure. Stem Cell Res Ther. 2018;9(1):240.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Bertolo A, Arcolino F, Capossela S, Taddei AR, Baur M, Potzel T, Stoyanov J. Growth factors cross-linked to collagen microcarriers promote expansion and chondrogenic differentiation of human mesenchymal stem cells. Tissue Eng Part A. 2015;21(19–20):2618–28.

    Article  CAS  PubMed  Google Scholar 

  218. Liu X, Wang Z, Wang R, Zhao F, Shi P, Jiang Y, Pang X. Direct comparison of the potency of human mesenchymal stem cells derived from amnion tissue, bone marrow and adipose tissue at inducing dermal fibroblast responses to cutaneous wounds. Int J Mol Med. 2013;31(2):407–15.

    Article  PubMed  CAS  Google Scholar 

  219. Prasanna SJ, Gopalakrishnan D, Shankar SR, Vasandan AB. Pro-inflammatory cytokines, IFNgamma and TNFalpha, influence immune properties of human bone marrow and Wharton jelly mesenchymal stem cells differentially. PLoS ONE. 2010;5(2):e9016.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  220. Driscoll J, Patel T. The mesenchymal stem cell secretome as an acellular regenerative therapy for liver disease. J Gastroenterol. 2019;54(9):763–73.

    Article  PubMed  PubMed Central  Google Scholar 

  221. Tsuji K, Kitamura S, Wada J. Secretomes from mesenchymal stem cells against acute kidney injury: possible heterogeneity. Stem Cells Int. 2018;2018:8693137.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  222. Galleu A, Riffo-Vasquez Y, Trento C, Lomas C, Dolcetti L, Cheung TS, von Bonin M, Barbieri L, Halai K, Ward S, Weng L, Chakraverty R, Lombardi G, Watt FM, Orchard K, Marks DI, Apperley J, Bornhauser M, Walczak H, Bennett C, Dazzi F. Apoptosis in mesenchymal stromal cells induces in vivo recipient-mediated immunomodulation. Sci Transl Med. 2017;9(416):eaam7828.

    Article  PubMed  CAS  Google Scholar 

  223. Liu FB, Lin Q, Liu ZW. A study on the role of apoptotic human umbilical cord mesenchymal stem cells in bleomycin-induced acute lung injury in rat models. Eur Rev Med Pharmacol Sci. 2016;20(5):969–82.

    PubMed  Google Scholar 

  224. Sung PH, Chang CL, Tsai TH, Chang LT, Leu S, Chen YL, Yang CC, Chua S, Yeh KH, Chai HT, Chang HW, Chen HH, Yip HK. Apoptotic adipose-derived mesenchymal stem cell therapy protects against lung and kidney injury in sepsis syndrome caused by cecal ligation puncture in rats. Stem Cell Res Ther. 2013;4(6):155.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  225. Pang SHM, D’Rozario J, Mendonca S, Bhuvan T, Payne NL, Zheng D, Hisana A, Wallis G, Barugahare A, Powell D, Rautela J, Huntington ND, Dewson G, Huang DCS, Gray DHD, Heng TSP. Mesenchymal stromal cell apoptosis is required for their therapeutic function. Nat Commun. 2021;12(1):6495.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  226. Hu C, Zhao L, Zhang L, Bao Q, Li L. Mesenchymal stem cell-based cell-free strategies: safe and effective treatments for liver injury. Stem Cell Res Ther. 2020;11(1):377.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Phinney DG, Pittenger MF. Concise review: MSC-derived exosomes for cell-free therapy. Stem Cells. 2017;35(4):851–8.

    Article  CAS  PubMed  Google Scholar 

  228. Garcia-Martin R, Wang G, Brandao BB, Zanotto TM, Shah S, Kumar Patel S, Schilling B, Kahn CR. MicroRNA sequence codes for small extracellular vesicle release and cellular retention. Nature. 2021;601(7893):446–51.

    Article  PubMed  CAS  Google Scholar 

  229. Wang H, Alarcon CN, Liu B, Watson F, Searles S, Lee CK, Keys J, Pi W, Allen D, Lammerding J, Bui JD, Klemke RL. Genetically engineered and enucleated human mesenchymal stromal cells for the targeted delivery of therapeutics to diseased tissue. Nat Biomed Eng. 2021.

Download references

Acknowledgements

We are grateful to the National Nature Science Foundation of China (81971841, 81830064, 81721092, 81941021, 81901973), the National Key Research and Development Plan (2017YFC1104701, 2017YFC1103304, 2018YFC2000400), the CAMS Innovation Fund for Medical Sciences (CIFMS, 2019-I2M-5-059) and the Military Medical Research and Development Projects (AWS17J005, 2019-126).

Funding

This study was supported in part by the National Nature Science Foundation of China (81971841, 81830064, 81721092, 81941021, 81901973), the National Key Research and Development Plan (2017YFC1104701, 2017YFC1103304, 2018YFC2000400), the CAMS Innovation Fund for Medical Sciences (CIFMS, 2019-I2M-5-059) and the Military Medical Research and Development Projects (AWS17J005, 2019-126). The National S&T Resource Sharing service platform Project of China (YCZYPT[2018]07), and the General Hospital of PLA Medical Big Data R&D Project (MBD2018030).

Author information

Authors and Affiliations

Authors

Contributions

LMR and JYF involved in writing—reviewing and editing, HQ involved in literature investigation and collation, ZYL involved in literature investigation and collation, ZLZ involved in pictures in the article, and FXB took part in conceptualization. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Meirong Li or Xiaobing Fu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no conflicts of interest.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, M., Jiang, Y., Hou, Q. et al. Potential pre-activation strategies for improving therapeutic efficacy of mesenchymal stem cells: current status and future prospects. Stem Cell Res Ther 13, 146 (2022). https://doi.org/10.1186/s13287-022-02822-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-022-02822-2

Keywords