Page 1 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

1 Genetic engineering of mesenchymal stem cells for regenerative medicine Adam Nowakowski1, Piotr Walczak2-4, Miroslaw Janowski1-3, Barbara Lukomska1 Affiliations: 1

NeuroRepair Department, Mossakowski Medical Research Centre, Polish Academy of Sciences,

Warsaw, Poland 2

Russell H. Morgan Dept. of Radiology and Radiological Science, Division of MR Research,

3

Cellular Imaging Section and Vascular Biology Program, Institute for Cell Engineering, The Johns

Hopkins University School of Medicine, Baltimore, USA 4

Dept of Radiology, Faculty of Medical Sciences, University of Warmia and Mazury, Olsztyn, Poland

Running Head: Genetic engineering of mesenchymal stem cells for regenerative medicine

Keywords: mesenchymal stem cells, transfection, gene therapy

Corresponding:

Dr. Adam Nowakowski [email protected]

Abstract Mesenchymal stem cells (MSCs), which can be obtained from various organs and easily propagated in vitro, are one of the most extensively used types of stem cells and have been shown to be efficacious in a broad set of diseases. The unique and highly desirable properties of MSCs include high migratory capacities toward injured areas, immunomodulatory features, and the natural ability to differentiate into connective tissue phenotypes. These phenotypes include bone and cartilage, and these properties predispose MSCs to be therapeutically useful. In addition, MSCs elicit their therapeutic effects by paracrine actions, in which the metabolism of target tissues is modulated. Genetic engineering methods can greatly amplify these properties and broaden the therapeutic capabilities of MSCs, including transdifferentiation toward diverse cell lineages. However, cell engineering can also affect 1

Page 2 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

2 safety and increase the cost of therapy based on MSCs; thus, the advantages and disadvantages of these procedures should be discussed. In this review, the latest applications of genetic engineering methods for MSCs with regenerative medicine purposes are presented.

Introduction Since their first description in the 1960s, mesenchymal stem cells (MSCs) have been of interest to researchers who use their therapeutic properties for the treatment of various pathological conditions. Regenerative medicine and oncology are the major fields of potential application [1-3]. Considering the wide breadth of literature on MSC engineering, we chose to limit our review to regenerative medicine. The simplicity of obtaining these cells from different organs and the subsequent ease of propagation in vitro [4], as well as the potential for autologous transplantation, makes them a tempting possibility for clinical use. However, special emphasis must be placed on the correct characterization of the obtained cells from different sources and species, to be able to properly compare and make right conclusions on the ground of multiples studies [5,6]. Independently of their original tissue location, MSCs have a natural ability to differentiate into mature mesenchymal phenotypes and form bone or cartilage, so they can be used in the treatment of injuries where these types of tissues are in need of repair [7,8]. What is more, MSCs act through paracrine effects, releasing a plethora of beneficial compounds [9]. While Prochymal (Mesoblast Ltd.), a preparation of allogeneic MSCs, has recently received conditional approval from the FDA, the engineered MSCs did not reach the level of clinical trials. However, genetically engineered neural stem cells are being tested in clinical trial sponsored by ReNeuron (NCT01151124 and NCT02117635, Clinicaltrials.gov). However, the diverse biological improvements offered by genetic engineering have the potential to greatly increase the therapeutically useful qualities of MSCs, and tailor them to specific diseases, so they are more probably to be used clinically in the future. A variety of studies have investigated the engineering of MSCs, aiming at stimulation of their direct differentiation toward endothelial cells and angiogenesis [10-16]. In addition, in in vitro (Fig. 1) culture conditions and in vivo studies (Fig. 2), MSCs can be differentiated into lineage-specific cells or lineage-specific-like cells, such as 2

Page 3 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

3 hepatocytes [17], cardiomyocytes [18], pacemaking cells [19], and neuronal cells [20]. Genetically modified MSCs can also be used to ameliorate several neurological disorders by exploiting their paracrine characteristics [21,22]. Moreover, MSCs could be modified to gain anti-fibrotic properties [23]. Finally, MSCs can be committed to overproduce anti-inflammatory cytokines that alleviate local tissue inflammatory states [24]. The abundance of genetic engineering modifications in MSCs quoted in this review was summarised in Tab.1. This review focuses on a set of examples of the various genetic modifications of MSCs in order to enhance and/or expand MSCs differentiation potential toward multiple cell types for therapeutic purposes, which have been published in recent years, revealing the growing interest directed at these cells and the diversity of strategies (Fig. 3).

Vectors, methods, and approaches to MSC engineering There are various vectors used for delivery of genetic materials to the cells, such as viruses [25], DNA plasmids [26,27], constructs of minimalistic, immunologically defined gene expression (MIDGE) [28], transposons constructs such as the Sleeping Beauty System [29], mRNA [30], miRNA [31], and siRNA [32]. While viruses directly transduce cells, non-viral methods require specific methods to cross the cell membrane, such as lipofection or electroporation. The vectors and methods for crossing the cell membrane by genetic material have been recently extensively reviewed [33]. From the biological point of view, there are also various approaches through which genetically engineered modification could be accomplished. The most basic and widespread approaches are related to the overexpression or knockdown of genes that encode extracellular exported proteins [34], specific membrane anchored [35], or cytoplasmic [36]. The more powerful approach is to induce the expression of transcription factor genes, which allows for cell reprogramming, and, in turn, leads to a complete change of phenotype [37]. Recently, the use of miRNAs is garnering growing interest, and permits the control of entire intracellular regulatory pathways, due to the relationship of one miRNA to several genes [38].

3

Page 4 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

4 Angiogenesis stimulation approaches To start with, one of the applications for MSCs in stem cell therapies is the field of induced posttraumatic angiogenesis. The problem of new blood vessel formation after injury is particularly important in ischemia, where the original arrangement of the blood vessels is destroyed, leading to a lack of oxygen and nutrient supply to the affected area. The process of angiogenesis naturally occurs after injury, but there is still a therapeutic interest in the enhancement of this physiological process. The MSC-based cell therapy may address some of these issues. One of the most popular approaches in this field is an overexpression of membrane protein: vascular endothelial growth factor (VEGF) gene engineering. VEGF properties have already been reviewed, revealing the important role of VEGF in angiogenesis during tissue injuries, such as in myocardial infarction, stroke, ischemia, and wound healing [39]. There are several examples of studies related to this kind of MSC modification, confirming that VEGF-engineered bone marrow-derived MSCs (BM-MSCs) possessed improved capabilities for tissue repair via the increased rate of angiogenesis [11-13,15,16]. An increased capillary density, a reduction in the infarct size, and elevated VEGF expression could be reached also by the use of survivin-overexpressing BM-MSCs [40]. Survivin was found to stimulate VEGF gene expression by up-regulating its promoter [41]. On the other hand, the VEGF-modified umbilical cord-derived MSCs (UC-MSCs) found an application in the tissue-engineered dermis for the treatment of skin defect wounds. In that case, the acellular dermal matrix scaffolds were used as platforms for the microencapsulated VEGF gene-transduced UC-MSCs, which effectively improved the vascularization of the graft and participated in the improved quality of wound healing [42]. The VEGF engineering can be accompanied by co-transfections with another factor that could enhance the pro-angiogenic effect of the VEGF-engineered BM-MSCs. The dual transfection of the TERT and VEGF genes was found to bring prolonged, beneficial pro-angiogenic effects in BM-MSCs derived from aged donors. An extended life span of the exogenous cells and improved angiogenic capacity were observed in those cells [43]. In another example of encapsulated BM-MSCs with angiogenic purposes, in a mouse hindlimb ischemia model, BM-MSCs were modified to overexpress the glucagon-like peptide-1 (GLP-1) 4

Page 5 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

5 gene, which is known as an anti-apoptotic, pro-angiogenic, and cardioprotective factor. Perivascular administration of the GLP-1-producing BM-MSCs improved postischemic reperfusion [44]. Angiopoietin-1 (Ang-1) is a potent growth factor that was found to be essential for new vein formation [45]. Based on this characteristic, Ang-1-engineered BM-MSCs were proposed for studies on angiogenesis. In a rat ischemic limb model, the Ang-1-modified BM-MSCs contributed to improved blood flow recovery, as well as to an increase of capillary density in the ischemic limb [46]. The cotransfection of the Ang-1 gene with the bone morphogenetic protein 2 gene (BMP-2) into BM-MSCs was found to significantly improve osteogenic and angiogenic differentiation [47]. Furthermore, the Ang-1-expressing UC-MSCs were delivered intravenously to a rat model of severe acute pancreatitis where these cells were involved in the promotion of pancreatic angiogenesis by reducing pancreatic injury and serum levels of pro-inflammatory cytokines [48]. The enhanced angiogenic potential of genetically engineered BM-MSCs was found for several additional gene modifications, such as for granulocyte macrophage-colony-stimulating factor (GMCSF) [49]; hepatocyte growth factor (HGF) [50]; human basic fibroblast growth factor (bFGF) [51], a tissue inhibitor of matrix metalloproteinase-3 (TIMP-3) [52]; the Notch homolog 1 [53]; heme oxygenase-1 (HO-1) [54], a member of the family of zinc finger transcription factors; GATA-4 [55]; HIF-1α [56]; and the LIM-homeobox transcription factor islet-1 (Islet-1) [57]. Some studies exist that relate to pro-angiogenic miRNA and engineered MSCs. miR-126 was featured in neo-angiogenesis after myocardial infarction, and with the posterior maintenance of vascular integrity [58]. In a mouse model of myocardial infarction, the transplantation of BM-MSCs expressing miR-126 resulted in increased angiogenesis and in the recovery of cardiac function in the infarcted zones of hearts [59,60]. Genetic engineering of MSCs could be applied as well in the field of endothelial cell differentiation. For instance, angiotensin converting enzyme 2 (ACE2) overexpression in BM-MSCs contributed to the prevention of endothelium-mediated inflammation and improved endothelial repair in in vitro conditions [61]. Physiologically, ACE2 is involved in the degradation of angiotensin II, which, in turn, is involved in vascular injury by contributing to the generation of pro-inflammatory, 5

Page 6 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

6 pro-thrombotic, and pro-fibrotic reactive oxygen species [62]. For this reason, ACE2, as a physiological antagonist of angiotensin II, is well suited as an anti-inflammatory factor that protects endothelial cells [63]. In addition to this, the enzymatic activity of stearoyl-CoA desaturase 1 (SCD-1) protects human arterial endothelial cells from lipotoxicity [64], while its overexpression in BM-MSCs was found to stimulate the expression of endothelial lineage marker genes [65]. Another example is the findings on the use of the angiogenesis‐related miRNA, miR-126, in BM-MSC differentiation toward endothelial cells [10]. In that case, miR-126-engineered BM-MSCs adopted the typical endothelial cobblestone-like structures in vitro, exhibited the expression of a series of endothelialspecific markers, and increased paracrine abilities.

Engineering of MSCs to increase blood production Engineered MSCs could have a role in the treatment of hemophilia B, since the encapsulated factor IX-engineered umbilical cord blood-derived MSCs (UCB-MSCs) produced sufficient amounts of that therapeutic protein in in vitro three-dimensional cultures [66]. Anemia has been also addressed through the administration of engineered MSCs. There are some examples of erythropoietin (Epo), the main inducer of red blood cell production [67], and its gene overexpression in MSCs in order to stimulate blood-cell formation in host animals after BMMSC administration. In animals that received either subcutaneously injected [68] or intraperitoneally implanted [69] human Epo-producing BM-MSCs, an increased hemoglobin level was reported. Moreover, HIF-1α-overexpressing BM-MSCs co-cultured with hematopoietic stem cells (HSCs) in in vitro conditions were shown to stimulate the growth of the latter cells [70]. A similar phenomenon was observed with the Jagged-1-producing BM-MSCs co-cultured with HSCs in vitro [71]. Apart from this, there have been attempts to convert UC-MSCs into hematopoietic cells. For instance, the miR218 overexpression in UC-MSCs facilitated differentiation toward CD34+ and CD45+ cells, if combined with additional stimulation [72].

Cartilage production applications 6

Page 7 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

7 One of the hallmarks of MSCs is their ability to differentiate toward chondrocytes under appropriate in vitro culture conditions, as confirmed for BM-MSCs [73] and the use of scaffolds was reported to enhance this process in UCB-MSCs [74], UC-MSCs and BM-MSCs [75], and dentalderived MSCs (D-MSCs) [76]. Moreover, multiple studies are underway to increase the chondrogenesis of MSCs by diverse genetic modifications. Previously, it was reported that TGF-β superfamily members were able to induce BM-MSCs to differentiate into chondrocytes under in vitro culture conditions [77]. The TGF-β1 gene could be efficiently delivered to MSCs with low toxicity, converting this biological pathway into a plausible solution for the BM-MSC modifications toward chondrogenic phenotypes [78]. The TGF-β1 gene transfection into BM-MSCs was combined with a calcium alginate gel to create three-dimensional, tissue-engineered, artificial cartilage to mimic the cell pattern growth in vivo. Type II collagen synthesis by the TGF-β1-transduced BM-MSCs was detected, indicating that, in the experimental conditions, the TGF-β1 modified BM-MSCs gradually differentiated into cartilage [79]. These results are consistent with the results of synovium-derived MSCs (S-MSCs) transfected with the TGF-β1 gene, where the modified cells were more proliferative and their chondrogenic differentiation potential was increased [80]. Another member of the TGF-β superfamily, TGF-β3 was documented as a forceful inducer of chondrogenesis in engineered BM-MSCs. The transfected BM-MSCs were transplanted into a rat model of osteoarthritis-induced cartilage injury. The administered cells differentiated toward chondrocytes and effectively secreted cartilage-relevant matrix compounds [81]. In in vitro studies, the lentiviral TGF-β3-transfected BM-MSCs were able to differentiate into chondrocytes in silk scaffolds, which was confirmed by the accumulation of collagen type II and cartilage extracellular matrix components [82]. In addition, another member of the TGF-β superfamily, the cartilage-derived morphogenetic protein 1 (CDMP1), was studied as a potential inducer of chondrogenic differentiation for BM-MSCs. The CDMP1-modified BM-MSCs were seeded on scaffolds and used to successfully repair laryngeal cartilage defects [83].

7

Page 8 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

8 The insulin-like growth factor-1 (IGF-1) is postulated to be another good candidate for the improvement of chondrogenic induction in MSCs, since it has been shown to influence MSCs chondrogenesis by increasing proliferation and inducing the expression of factors involved in the chondrocyte phenotype [84]. Adipose tissue-derived MSCs (ASCs) were co-transfected with the IGF1 and BMP-2 genes. The appropriate chondrogenenic process was appreciated in these so-transformed cells [85], despite the fact that the overexpression [86] and the treatment of BMP-2 alone is reported to facilitate osteogenesis in MSCs [87]. There is an example of the application of another BMP family member, BMP7, in which the BMP7 gene transfection of BM-MSCs resulted in favorable chondrocyte formation [88]. Interestingly, the IGF-1 overexpression was able to stimulate the expression of the Sox9 gene; however, the expression of osteogenic marker genes was also elevated, suggesting that IGF-1 gene transfections into BM-MSCs with chondrogenic purposes should be considered cautiously [89]. Moreover, the IGF-1 gene transfection, combined with different genes involved in chondrogenesis, resulted in an increase of the chondrogenic differentiation capacities of modified ASCs, especially in the case of IGF-1/fibroblast growth factor-2 double-transfected ASCs [90]. In addition, non-canonical Wnt signaling was reported to be useful for this purpose, since Wnt11modified BM-MSCs had the enhanced expression of chondrogenic genes, such as Sox9, and their chondrogenic differentiation was more profound compared to un-transfected cells [91]. Other genetic candidates for the induction of chondrogenesis for MSCs are the Sox trio genes (Sox 5, 6, and 9). They are master regulatory transcription factors known to be engaged in the activation of cartilage-specific genes [92]. The Sox9 gene, the key important gene in the early stages of chondrogenesis, has been the most studied in BM-MSCs [93-95]. In one case, in order to reinforce the chondrogenesis process, the genetic engineering strategy involved the pro-chondrogenic stimulation represented by the induction of Sox9 gene overexpression and the simultaneous inhibition of the opposite differentiation route toward osteogenic commitment represented by the application of Cbfa-1 gene silencing. The Cbfa-1 gene expression was linked to the osteogenic differentiation of BM-MSCs [96]. For this reason, at the same time, BM-MSCs were transfected with the Sox9 genecontaining vector and the small interfering RNA against the Cbfa-1 transcript, both conjugated to 8

Page 9 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

9 poly(ethyleneimine) and coated onto poly(lactide-co-glycolic acid) nanoparticles [94]. In a different study, the Sox9 gene up-regulation, and hence, increased chondrogenic differentiation of BM-MSCs, was achieved due to zinc-finger protein 145 (ZNF145) gene overexpression, which was established as a new upstream regulator of Sox9 gene expression [97]. Indeed, there are some data on the improved results when all three Sox family genes were simultaneously overexpressed in BM-MSCs [98,99] and ASCs [100].

Intervertebral disc regeneration In the area of intervertebral disc regeneration, several modifications were reported to improve BM-MSC differentiation toward intervertebral disc-like cells [101,102]. For instance, the growth and differentiation factor 5 (GDF5) gene engineering was used to obtain GDF5-BM-MSCs. GDF5 stimulated chondrogenic BM-MSCs differentiation in vitro [103] and induced the repair of degenerated disc in vivo [104]. The GDF5-producing BM-MSCs differentiated into intervertebral disclike cells, expressing chondrogenic-specific marker genes, and, when transplanted to the intervertebral disc papain degeneration organ culture model, a partial recovery was detectable [102]. Matrix metalloproteinases (MMPs) are involved in extracellular matrix degradation during intervertebral disc degeneration [105]. For this reason, the tissue inhibitors of metalloproteinases convert into an interesting target for MSC engineering with the TIMP metallopeptidase inhibitor 1 (TIMP1) as an example. In this case, the TIMP1-producing BM-MSCs transplanted into the intervertebral injury site elicited double-positive effects by diminishing the negative effects of the MMPs activity, and, on the other hand, promoted the synthesis of additional extracellular matrix elements [106]. Worth noting is the case of Bcl-2 overexpressing BM-MSCs that became Sox9-overexpressing cells and were able to accumulate proteoglycans, supporting the notion that anti-apoptotic BM-MSC engineering could change these cells toward the cells with a nucleus pulposus-like phenotype in vitro [101].

9

Page 10 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

10 Bone formation attempts using MSCs In bone-formation approaches based on engineered MSCs, the most popular strategy, with a plethora of examples, is related to BMP-2 gene transfection. BMP-2 is a powerful positive regulator of bone formation in vivo [107], and stimulates osteoblast growth in vitro [108]. The BMP-2 overexpressing BM-MSCs have been used in multiple studies in order to produce new bone tissues in vitro and in vivo [86,109-111]. For example, the use of anti-BMP-2 antibodies encapsulated with BMMSCs to entrap BMP-2 in the vicinity of BM-MSCs enabled the proper induction of osteogenesis [112]. In many cases, the BMP-2 overexpression in MSCs was accompanied by the overexpression of other genes that strengthened the effects of BMP-2. For instance, the SDF-1β overexpression in BMMSCs potentiated osteogenic differentiation and chemotaxis in in vitro-expanded BMP-2-engineered BM-MSCs [113]. Of interest, the simultaneous pro-angiogenic and pro-osteogenic strategies applied to BM-MSC engineering were found to substantially improve bone formation potential. The HIF-1α gene overexpression was reported to have beneficial effects on new bone formation by inducing the pro-angiogenic genes in the double-HIF-1α/BMP-2-positive BM-MSCs [114]. What is interesting is that the HIF-1α gene overexpression alone enhanced the osteogenic differentiation of BM-MSCs just as well [115]. Similar results were reported in the double Ang-1/BMP-2 and VEGF/BMP-2overexpressing BM-MSCs [47,116-118]. However, BM-MSCs engineered to exclusively overexpress the pro-angiogenic VEGF gene were found to be strong inducers of new bone formation by the promotion of new vessel formation among the scaffolds used as templates [119,120]. Curiously, BMMSCs were converted into bone tissue after the overexpression of the Epo gene [121]. Epo, apart from its well-known role in hematopoiesis, promoted bone regeneration by stimulation of VEGF and BMP2 gene expression [122,123]. In a different study, the BMP-2-producing BM-MSCs were seeded together with the endothelial progenitor cells (EPCs) in porous nano-calcium, sulfate/alginate scaffolds, and EPCs contributed to the increased new vascular network growth. Meanwhile, the BMP2-producing BM-MSCs promoted bone regeneration [124]. Interestingly, other BMP family members were reported to effectively contribute to bone formation from the transduced BM-MSCs to

10

Page 11 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

11 overexpress genes, such as BMP-4, -7 [125] and BMP-6, claimed to be even more potent proosteogenic factors than BMP-2 in the BMP-2-engineered BM-MSCs [126]. Apart from the BMPs-based strategies, other genes physiologically engaged in osteogenesis and bone regeneration were considered for attempts with BM-MSCs, with various genes, such as the osteogenic growth peptide (OGP) [127]; the insulin-like growth factor I (IGF-1) [128,129]; leptin [130]; platelet-derived growth factor (PDGF) [131]; ephrinB2 [132]; the integrin α2 subunit ITGA2 [133]; and NEL-like protein (NELL-1) overexpression reported in ASCs [134]. In addition, the targeting of Wnt signaling proved useful in the bone formation studies on MSCs. BM-MSCs were transformed to either overexpress β-catenin (the pivotal element in the canonical Wnt signaling pathway) or the ROR2 gene (the pivotal molecule in the β-catenin-independent, non-canonical signaling pathway). In both cases, the alkaline phosphatase activity was promoted, as well as the expression of the osteogenic-specific genes [135]. In addition, transcription factor core binding factor alpha l (Cbfa-1) [136], and a transcriptional co-activator with a PDZ-binding motif (TAZ), were reported to stimulate osteogenesis in BM-MSCs [137]. The forkhead box protein C2 (Foxc2) overexpressing BM-MSCs were found to differentiate toward osteoblasts, mediated by the activation of the canonical Wnt/β-catenin signaling [138]. Surprisingly, despite its main commitment to the selfrenewal of embryonic stem cells, the overexpressed Nanog was reported to effectively stimulate osteogenic differentiation in modified BM-MSCs [139]. Finally, in order to ensure the high expression of pro-osteogenic genes in MSCs during the differentiation procedure, the regulatory properties of specific miRNAs could be used. During osteoblast differentiation, the miR-29b positively regulates this process by down-regulating the inhibitory factors of osteogenic signaling pathways [140]. The osteoblast differentiation was obtained in the miR-29b-overexpressing BM-MSCs [141]; however, in the case of ASCs Liao et al. reported that more pro-osteogenic potential was attributed to the miR-148b than to the miR-29b [142]. Furthermore, miR-148b-overexpressing ASCs also co-transfected with the BMP-2 gene possessed enhanced and prolonged BMP-2 gene expression. Different study reported that the miR-31 suppression obtained in the anti-miR-31-overexpressing BM-MSCs resulted in a significant increase 11

Page 12 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

12 in the expression of osteogenic lineage genes in in vitro studies, and bone formation was detected in vivo [143].

Engineered MSCs in liver regeneration In the case of liver regeneration attempts, VEGF was likewise recognized to promote BMMSC differentiation into hepatocyte-like cells under the specific condition of HGF and EGF stimulation [144]. However, most studies have been focused on attempts to induce hepatic differentiation of engineered MSCs with physiologically committed factors. To this end, a group of transcription factors involved in the physiological control of the hepatic commitment [145] were the subject of interest in the MSCs studies. One of them, the hepatocyte nuclear factor 4α (HNF4α), is a specific hepatic lineage nuclear receptor [146]. The HNF4α-producing UC-MSCs differentiated into hepatocyte-like cells under hepatic differentiation conditions and activated various hepatic-specific genes, highlighting the key role of HNF4α in the hepatic lineage commitment [147]. Another HNF family member that regulates liver development, the representative of an HNF3 subfamily, HNF3β, was also investigated in terms of its ability to induce MSC differentiation. An efficient incorporation of the modified BM-MSCs into liver grafts was attributed to the HNF3β effects on the stimulation of the expression of anti-fibrotic molecules [148]. In the case of ASCs, an interesting approach was proposed by Alizadeh et al. [17] in which the let-7 family miRNA member, let-7b, was targeted, which assumes the physiological role of an inhibitor of hepatic differentiation that silences the expression of HNF4α [149]. In that study, the Let-7b-5p inhibitor was employed for ASCs transduction. As a consequence, HNF4α gene expression was enhanced by removing the physiological inhibitor, thus allowing the engineered ASCs to differentiate toward hepatocyte-like cells [17]. What is more, the miR-122 is a key regulator in liver embryonic development, and maintenance of hepatic homeostasis, and its overexpression is sufficient for hepatic differentiation of embryonic stem cells (ESCs) [150]. In the case of human adipose tissue-derived MSCs, the overexpressed miR-122 was found to induce ASC differentiation into hepatocyte-like cells, confirmed by the increased expression of the specific hepatocyte markers and the production of urea and albumin by the differentiated ASCs 12

Page 13 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

13 [151]. On the contrary, Cui et al. found that the miR-122 was incapable of inducing hepatic differentiation alone in human UC-MSCs [152]. The hepatocyte-like cells were exclusively obtained from UC-MSCs after the concomitant transfection with six different miRNAs—miR-1290, miR-1246, miR-30a, miR-148a, miR-424, and miR-542-5p—previously thought to be overexpressed during hepatic differentiation [153]. Finally, MSCs may serve as therapeutic tools in the personalized medicine treatment of rare, inherited liver disorders. Promising results were obtained from a study on the Wilson’s disease model (WD). The hallmark of Wilson’s disease is the accumulation of elevated levels of copper in the liver due to the failure of the copper transporter, ATP7B, which is responsible for copper homeostasis [154]. The ATP7B-positive BM-MSCs were more resistant to the toxic effects exerted by the copper present in the culture medium than were control cells, indicating that this approach could be a promising solution for cell therapy of patients suffering from WD [155].

MSCs in cardiac-targeted therapies Native and genetically modified BM-MSCs have found an application in cardiac injury studies [156,157]. In that regard, the experiments have been focused on cardiomyocyte and pacemaker cell differentiation, while other strategies are focused on the improvement of the cardioprotective properties of the therapeutic cells after their administration to the injured myocardium. BM-MSCs could be efficiently modified to gain cardioprotective properties. For example, extracellular superoxide dismutase (ecSOD) - overexpressing MSCs were reported to exert cardiac protection and improve cardiac function by the anti-oxidative protection achieved by the inactivation of reactive oxygen species in mice with infarcted hearts [158]. Glucagon-like peptide-1 (GLP-1) is one of the factors that exhibits cardioprotective properties [159]. Therefore, GLP-1 was used for BM-MSC engineering with posterior encapsulation, and a prolonged supply of GLP-1 was maintained after the in vivo study [160]. The Wnt/β-catenin pathway is another subject of studies involving MSC engineering. Wnt proteins could act through two independent pathways: canonical, which includes β-catenin 13

Page 14 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

14 involvement; and β-catenin-independent non-canonical ways of signal transmission [161]. Furthermore, it has been suggested that the Wnt canonical signaling pathway promotes the cardiac commitment of the precursor cells at early developmental stages, but inhibits it later [162]. This dual Wnt/ β-catenin signaling feature was used in BM-MSC differentiation attempts, while the BM-MSCs overexpressing the interference sequence for the β-catenin gene were shown to decrease the canonical Wnt pathway in favor of the non-canonical pathway, thus contributing to the myocardial-like cell commitment [163]. However, the overexpression of the Wnt11, which participates in the noncanonical Wnt signaling, resulted in the transdifferentiation of the engineered BM-MSCs toward cells with the cardiac phenotype [164]. HO-1 is an enzyme that degrades heme into biliverdin, free iron, and carbon monoxide [165]. HO-1 also elicits beneficial results during myocardial infarction after exogenous delivery [166]. The cardioprotective properties of HO-1 were evaluated in studies conducted on the HO-1-overexpressing BM- [167] and ASCs [168] delivered to injured cardiac tissue. The differentiation of MSCs to generate new cardiomyocytes may occur by inducing the expression of genes physiologically involved in cardiomyocyte precursor differentiation signaling pathways. An example of this is myocardin, a crucial factor in heart development [169], whose overexpression was also accompanied by TERT gene overexpression, and together, promoted promyogenic gene expression and sustained the growth capacity of the modified ASCs [170]. One other protein involved in myocardial differentiation is a GATA family member, GATA-4. The representatives of the GATA family are transcription factors that bind to the GATA motifs among DNA regulatory sequences. In particular, GATA-4 is the pivotal controller of cardiac differentiation and regulates cell survival in the adult heart [171]. In the GATA-4-overexpressing BM-MSCs, an increase in other genes committed to myocardial differentiation was detected, and cell contraction was observed that might indicate that the abundant presence of GATA-4 paves the way for the BM-MSCs to transdifferentiate into and act as functional myocardial cells [172,173]. However, in the GATA-4 overexpressing BM-MSCs, the members of the miR-15 family were downregulated, which was translated into an increased resistance to the ischemic environment by the stimulation of anti-apoptotic 14

Page 15 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

15 Bcl-2 family members [174]. In addition, the GATA-4-engineered BM-MSCs gained the cardiomyocyte protection properties by the production of microvesicles containing elevated amounts of the miR-221 [175]. Previously, it was stated that the miR-221 targeted the 3’-UTR of the p53upregulated modulator of apoptosis (PUMA), which was a pro-apoptotic member of the Bcl-2 protein family. Furthermore, the inhibition of PUMA expression contributed significantly to cell survival by inhibition of the apoptosis process [176]. Cardiomyocyte differentiation of MSCs was reported after transfection with the antisense oligonucleotide, AMO-124, targeting miR-124, which, in turn, is involved in neurogenesis [177], but its depletion in BM-MSCs was reported to stimulate myogenic differentiation [178]. Cardiomyocyte commitment was also observed in the miR-133a and miR-499 - overproducing BM-MSCs [18,179]. The miR-499 function was assigned to promote cardiac muscle differentiation by provoking the repression of the histone deacetylase 4, or sex-determining region Y-box 6 genes [180], while the miR-133a was shown to be expressed in adult cardiac and skeletal muscle tissues [181]. Functional recovery of the ischemic heart was reported after the cardiac administration of the miR-210 microvesicle-producing BM-MSCs [182], which is concordant with the reported neuroprotective effects of the miR-210 inhibiting apoptosis by silencing pro-apoptotic gene expression [183]. In order to produce cardiac pacemaker cells, MSCs were engineered to overexpress the hyperpolarization-activated, cyclic, nucleotide-gated channel (HCNs) genes. HCNs underlie the production of hyperpolarization-acivated cation currents (If), the common feature of the excitable cells [184]. The HCN family consists of four members where each channel type induces a hyperpolarization-activated current with a special activation kinetic [185]. The HCN4-modified BMMSCs enabled a stable pacemaking function and an appropriate chronotropic response in dog hearts with an atrioventricular block [186]. However, in a different study, the HCN4-producing BM-MSCs were indeed able to produce certain If currents, but were poorly connected with the host cardiomyocytes, since there was a low presence of the gap-junctions [187]. Next to this, the HCN1modified BM-MSCs cultivated in vitro with neonatal rabbit ventricular myocytes were able to produce If currents and compel an increase in the spontaneous beating rate of the co-cultured myocytes [19]. 15

Page 16 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

16 There is also a piece of evidence that BM-MSCs could be utilized for engineered-tissue heart valves [188]. For this purpose, MSCs were transfected with the fibroblast inducible factor 14 (Fn14) gene, which, among other pathways, is involved in cardiomyocyte proliferation [189].

MSCs in neural differentiation and repair There has been increasing interest in MSC employment with neurological purposes. This interest stems from the fact that MSCs possess multiple immanent abilities to modulate neural cell growth and survival. It seems that MSCs act through paracrine actions rather than through direct differentiation, by attracting host neural stem cells to the sites of injury, and by enhancing their proliferation and differentiation in the damaged region [190]. In addition, native MSCs secrete a plethora of biologically active compounds, such as chemokines, anti- and pro- inflammatory cytokines, angiogenic factors, growth factors, and growth factor-binding proteins [9]. Even more significant for the neurological context, MSCs are able to express a variety of neuro-regulatory molecule genes, such as the axon guidance molecules, neural cell adhesion molecules, neuriteinducing factors, neurotransmitter receptors, and neurotrophic factors [191]. Thus, MSCs of different origin have become an attractive pursuit in the studies on animal models of stem cell-based therapies in stroke (i.e., BM-MSCs) [192,193], traumatic brain injury (i.e., BM-MSCs) [194], and in several models of neurological disorders, such as in Parkinson’s disease (PD) (i.e., UC-MSCs) [195], amyotrophic lateral sclerosis (ALS) (i.e., BM-MSCs) [196], multiple sclerosis (MS) (i.e.,ASCs) [197], Alzheimer’s disease (AD) (i.e.,ASCs) [198], and Huntington disease (HD) (i.e., UC-MSCs) [199]. MSCs could also be considered useful in studies with spinal cord injury models [200]. In addition to the experiments carried out with the administration of native MSCs, many examples exist of the use of genetic engineering methods to enhance the therapeutic effects of MSCs by granting MSCs new characteristics as a result of genetic manipulation. In this regard, in the case of traumatic brain injury, intra-ventricularly transplanted BM-MSCs overexpressing one of the neurotrophic factors, brain-derived neurotrophic factor (BDNF), were presented to attenuate neuronal injury due to the elevated concentration of BDNF present in the injured brain [201]. In that case, BDNF exerted its 16

Page 17 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

17 beneficial effects deriving from its role as a member of the neurotrophin family, whose members are involved in neuronal survival during development and injury states [202]. In peripheral nerve regeneration studies, the BDNF-producing ASCs were engaged in nerve regeneration, and differentiated into non-myelinating Schwann cells, fostering significant functional recovery of the examined animals [203]. In addition, the BDNF-releasing BM-MSCs preserved motor function in a mouse model of HD [204]. In the case of studies on stroke, the BDNF-producing BM-MSCs transplanted into ischemic stroke animals protected host cells from apoptosis and accelerated the proliferation and maturation of endogenous neural stem cells (NSCs) situated in the subventricular zone (SVZ) of the host brains [22]. It is worth noting that the CXCR4-overexpressing BM-MSCs were able to ameliorate stroke effects after femoral vein injection due to the participation in the enhanced growth of capillary vascular volume among the stroke penumbra, a depletion in the volume of the damaged infarction area, and, finally, by an improvement in neurological functions [205]. Glial cell line-derived neurotrophic factor (GDNF) is critical for the development and protection of the adult central nervous system and peripheral neurons [206]. This feature was exploited to generate the GDNF-overexpressing BM-MSCs that exhibited elevated neuroprotective effects after their transplantation into rats with intracerebral hemorrhage [207]. In another case, genetically modified BM-MSCs that overexpressed GDNF, together with neurotrophin-3 (NT-3), another member of the neutrophin family, were able to differentiate into neuronal cells and express nerve markers when induced by fetal gut culture medium. Those results indicate that these cells could be helpful in future therapies for Hirschsprung’s disease, characterized by the disrupted migration of neural crest cells during embryogenesis [208]. In stroke models, there are studies describing the application of genetic engineering methods to produce neurons from MSCs. For instance, Neurogenin2 (Ngn2), which belongs to the transcription factor family with helix-loop-helix motifs, demonstrated pro-neural activities, confirming that it is one of the key factors during corticogenesis [209]. Ngn2 gene transfection to BM-MSCs was adequate to convert these cells into neuron-like cells in vitro and Ngn2-positive BM-MSCs were able to diminish 17

Page 18 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

18 infarct volume and reduce the number of apoptotic cells in an in vivo stroke model. Furthermore, the functional recovery of the treated animals was appreciated as well [210]. However, BM-MSCs overexpressing another member of the transcription factors with a helix-loop-helix motif family, neurogenin-1 (Ngn1), were able to migrate with high efficiency to the central nervous system and exerted their paracrine functions on the host neural cells in an ALS mice model [211]. Another case reported data on the functional recovery after stroke obtained from animals with an BM-MSC-derived, exosome-mediated miRNA delivery to the injured area of stroke. BM-MSCs were engineered to overexpressed miR-133b, which was formerly recognized as a neural cell modulator, since the miR133b delivery to neural cells resulted in neurite outgrowth in vitro [212]. In an animal study, the miR133b-containing exosomes were effectively transmitted from the administered miR-133boverexpressing BM-MSCs to host neurons and astrocytes, and influenced gene expressions with a subsequent positive effect on axonal plasticity and neurite remodeling in the ischemic boundary zone [213]. Hippocampal neovascularization was detected after the delivery of the VEGF-overexpressing BM-MSCs to the lateral ventricle in a murine model of AD [21], with the resultant beneficial behavioural effects showing the degree of the vascular damage in neurodegenerative diseases, such as AD. In addition, GLP-1, which was previously noted to exhibit neuroprotective capabilities in in vitro studies [214], was used for BM-MSC engineering, followed by right lateral ventricle administration into an AD animal model. Encapsulated GLP-1-positive BM-MSCs were responsible for a reduction in the accumulated amyloid-beta (Abeta) peptides in a mouse model of AD [215]. Studies carried out with PD rat models show some examples of the substantial, beneficial behavioural results after engineered BM-MSC administration with either GDNF [216,217] or persephin [218], which is another neurotrophic factor that promotes survival, proliferation, and differentiation of neurons [219]. MSC modifications, with the purpose of obtaining cell-based therapies to alleviate the symptoms of MS, have generally been focused on the inhibition of the auto-immune response in which the myelin sheaths around the axons of the central nervous system are destroyed, causing 18

Page 19 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

19 demyelination and neuronal loss [220]. Vasoactive intestinal peptide (VIP)-producing ASCs [221], IL10-secreting ASCs [222], and interferon-β [223]-releasing BM-MSCs effectively migrated to the inflamed central nervous system and diminished the auto-immune response of the host by a decrease in pro-inflammatory cytokine secretion, T cell activity modulation, and the protection of the bloodbrain barrier from new injuries. In the case of spinal cord injuries, NT-3 was transduced in MSCs because of its action as a neuroprotective agent [224]. NT-3-positive BM-MSC administration to the injured spinal cords facilitated axonal regeneration, promoted neuronal survival [225], and favoured remyelination and functional recovery in the demyelinated spinal cords of the studied animals [226]. NT-3 acts through tropomyosin-related kinase receptor tyrosine kinases (Trk) [227], a feature that was used for BM-MSC transduction with Trk type C (TrkC) gene overexpression [228]. TrkC-producing MSCs possessed the ability to prolong their survival after delivery, improve differentiation toward neuron-like cells, and participated in corticospinal tract regeneration in the injured areas, mainly due to the elevated concentration of NT-3. Another neurotrophic factor that acts through the Trk-mediated signaling pathway is multineurotrophin (MNTS-1). MNTS1 is a mutated form of NT-3 that could bind efficiently all members of the Trk family [229]. After the administration of the MNTS1-producing BM-MSCs to the sites of spinal cord injuries, a noticeable reduction in the cavity volume and intense axonal growth were detected [230]. Sonic hedgehog (Shh) is a pleiotropic factor in the developing central nervous system and spinal cord [231]. Engineered BM-MSCs secreting Shh enhanced VEGF expression in the injured area, thereby causing an improvement in the local microenvironment and leading to enhanced functional recovery [232]. Finally, there is some evidence of positive outcomes after genetically engineered BM-MSC administration in animal models of epilepsy [233], autoimmune encephalomyelitis [234], and vascular dementia [235]. Another broad-range research interest is to establish efficient approaches for MSCs to force them to differentiate toward functional neurons in vitro by means of genetic engineering. For this 19

Page 20 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

20 purpose, several genes were selected to be overexpressed in MSCs. Nerve growth factor (NGF), a member of the neurotrophin family, noted to stimulate neuronal survival and neurite outgrowths, was used in MSC engineering. The NGF-producing BM-MSCs were effectively changed into cells with a neuron-like phenotype [236]. GDNF overexpression, combined with RA stimulation, resulted in the enhanced ability of transfected BM-MSCs to differentiate into neuron-like cells, which was confirmed by the expression of specific neuronal marker microtubule-associated protein 2 (MAP-2) [20]. Caveolin-1 was found to interact with the membrane receptor Notch-1, which, in turn, is involved in neuronal [237] and astroglial differentiation of NSCs [238]. In the case of BM-MSCs, the downregulation of the Caveolin-1 gene expression was achieved with siRNA engineering, and modified cells were more sensitive for neuronal differentiation, which was affirmed by the expression of several neural-specific markers [239]. Trans-retinoid acids (RA) participated in embryonic nervous system development and nerve cell differentiation by acting through the retinoic acid receptor β (RAR-β) [240]. For this reason, the RAR-β gene was overexpressed in BM-MSCs in order to improve neuronal differentiation [241]. The increased presence of the RAR-β in the nuclei of transduced BM-MSCs made these cells more susceptible to the RA signaling pathway, which, in turn, helped facilitate neuronal differentiation of the RAR-β-presenting BM-MSCs. Another protein, Nurr1, is a transcription factor that is elemental for the development and maintenance of dopaminergic neurons [242]. BM-MSCs were modified to overexpress the Nurr1 gene with simultaneous stimulation with electrical pulses. As a result, the Nurr1-positive BM-MSCs produced neurites and expressed a series of neurogenic markers, such as Tuj1, nestin, and Map 2 [243]. Yet another protein used, a basic helix-loop-helix transcription factor member, mammalian achaete-scute homologue-1 (Mash1), was reportedly an efficient inducer of BM-MSC neuronal differentiation [244]. The miRNA approaches could be valid as well. The miR-9, with its target, the zinc-finger protein 521 (Zfp521), the inhibitor of neuronal differentiation, and miR-125b, important in neuron-like 20

Page 21 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

21 cell differentiation, were implemented for BM-MSC engineering. In both cases, the modified BMMSCs were shown to become neuron-like cells [245,246].

Other applications Ultimately, there are some studies on the use of MSCs in the therapeutic strategy against bladder (i.e. fetal BM-MSCs) [247] and hepatic fibrosis by the overexpression of hepatocyte growth factor (HGF) by UC- and BM- MSCs [248-250]; differentiation of ASCs into hair cell-like cells by Atoh1 overexpression [251]; Achilles allograft replacement through transforming growth factor-1 transfection produced by BM-MSCs [252]; attenuation of intestinal injury by heparin-binding EGFlike growth factor overexpression [27]; and protection against acute lung injury by keratinocyte growth factor (KGF) expression [253]. MSCs have also been used against acute retinal injury through neurotrophin-4 production [254], and in UC-MSCs a decrease in excessive inflammatory injury after acute radiation injury by thioredoxin gene overexpression [255]. Further, there is some evidence that the DAZL gene overexpression in BM-MSCs converts these cells into putative germ cells [256]. Several attempts were focused on pancreatic lineage differentiation. The favourable MSC differentiation toward insulin-producing cells was achieved by the overexpression of several genes, such as betacellulin reported for BM-MSCs [257], and in the case of BM-MSCs and UC-MSCs are some data for pancreatic, and duodenal homeobox factor 1 (PDX-1) [258,259], combined with other lineage-specific genes [260,261], and pancreatic characteristic miRNAs, such as the miR-375 in the human placental decidua basalis-derived MSCs (PDB-MSCs) [262], or the insulin gene itself in hair follicle-derived MSCs (HF-MSCs) [263]. The VEGF and human IL-1 receptor antagonist (IL-1Ra)overexpressing BM-MSCs were shown to protect islet cells from cytokines during islet transplantation [264]. Similarly, the immunomodulatory properties of IL1Ra were also shown in the case of IL-1Ramodified ASCs [265] and BM-MSCs [266] in a cartilage repair model, where the modified MSCs protected the development of new cartilage tissue from the effects of IL-1 within pro-inflammatory surroundings.

21

Page 22 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

22 In the studies on nephritis, the glutathione S-transferase M1(GSTM1) - producing BM-MSCs ameliorated renal pathological damage due to the inhibition of the oxidative stress-induced inflammation by reactive oxygen species removal [267]. The HGF-overexpressing BM-MSCs reduced renal fibrosis after ureteral obstruction [23], and survivin producing BM-MSCs were found beneficial in renal injury in a mouse model, probably owing to the prolonged survival and increased cytokine production by the engrafted BM-MSCs [268]. In addition to its anti-fibrotic effects, in the case of the HGF-expressing ASCs, HGF was reported to stimulate thymus regeneration in a rat thymus involution model [269]. One more application of engineered MSCs is the introduction of genetic modifications to gain more prominent anti-inflammatory properties. For instance, the expression of anti-inflammatory cytokines, such as IL-10 was achieved in BM-MSCs [24] and ASCs [222,270] and IL-4 in BM-MSCs [271]. Curiously, in BM-MSCs transfected with pro-inflammatory TNF-α, an immunosuppressive activation was observed [272].

Safety and feasibility issues regarding future therapies with engineered MSCs Fortunately, therapies based on the transplantation of MSCs do not raise significant ethical and legal objections, which is in stark contrast with some other types of stem cells, especially human embryonic stem cells [273]. Due to the growing interest and hopes for therapies based on naïve and engineered MSCs, safety and feasibility aspects are extremely important and should be taken into account. These two aspects, if not properly addressed, could pose significant drawbacks to clinical use, and, for this reason, should be considered in addition to the aforementioned beneficial effects. Only a comprehensive approach will result in the realistic and practical use of the therapeutic properties of engineered MSCs. One of the fundamental problems related to genetic cell engineering, when considering clinical trials, is achieving reliable expression and low risks. Viral transduction methods, despite their reported high efficiency of transgene expression, are unlikely to be allowed in clinical trials, because of their high risk of insertional mutagenesis [274], chromosomal instability [275], and proto-oncogene 22

Page 23 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

23 activation. Even the use of self-inactivating lentiviral vectors has been shown not to be a safe solution [276]. A transposon system is a non-viral transfection method that has been used for stem cell engineering [277] including MSCs [278,279]. However, in this case there is also a risk of insertional mutagenesis [280]. For this reason, the approaches that use putative safe integration sites among target genomes are under consideration for MSCs engineering, such as the Zinc-finger nuclease-based genome-specific locus targeting used for Epo gene introduction into the MSC genome [281]. What is more, other site-specific genome engineering tools exist, such as transcription activator-like effector nucleases (TALENs) [282] and clustered regulatory interspaced short palindromic repeat (CRISPR)/Cas-based RNA-guided DNA endonuclease targeting [283], which could be potentially engaged in MSC modifications. However, despite the aforementioned solutions, methods that do not lead to genomic integration, such as DNA- and RNA- based transfections, pose a tempting solution for stem cell-based therapies [33]. Unfortunately, in MSCs, DNA-based transfections often result in rather low efficiency; thus, new approaches aimed at improving this strategy are being developed [284]. The goal is to provide an efficient, safe, and yet effective transfection method, which would result in a transduction efficiency comparable to viral methods [285]. It has been shown that viral-free transfection of MSCs is not without risk, as it may affect cellular viability and differentiation potential [286]. Therefore, there is still a need to develop new, more sophisticated and technically advanced methods, such as, for example, plasmid delivery with nanoparticles based on the polysaccharide from Angelica Sinensis [287], or by using microbubble-mediated ultrasound and a polyethylenimine compound [288]. Another attractive method for cell engineering that does not lead to genome integration and provides fairly high transfection efficiency is mRNA-based transfection [30]. This method has been shown effective in modulating MSC differentiation [289]; however, it should be always borne in mind that mRNA-based transfection is of a transient nature, so it may pose a problem in applications that require a long operating time, especially in long-term clinical trials. In addition to biological issues, the financial and legal regulations must be considered when developing putative stem cell therapies [290,291]. Despite the fact that the engineering of MSCs has 23

Page 24 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

24 proven a promising strategy in preclinical research, the application of that technology in the clinic requires a cost-benefit analysis, to assess whether the advantages of cell engineering outweigh the cost. The overall costs apply to two main categories: the cost of the engineering procedure itself, and the cost of approval by regulatory bodies. Both these can be prohibitive; however, this may differ in various countries, introducing the risk of unequal access to therapeutic procedures in the global healthcare market. The cost-effectiveness has been already evaluated in the field of tissue engineering [292], showing that the tissue engineering costs are less than those associated with obtaining material from donors for transplantation. This means that tissue engineering may be more accessible to a larger number of patients. A similar financial analysis is expected to be applied to the field of cell engineering, once such therapies prove efficacious in clinical trials. If the magnitude of improvement is significantly greater than the costs, that will stimulate the marketing of therapies based on engineered cells [293,294]. Overall, the relatively high financial expenses and complex biological aspects, as well as the wide possibilities offered by genetic engineering, make MSCs-based therapies a very interesting research area, but, due to still-unresolved issues, this field is in the early phase of development and therapeutic applications are at the preclinical stage.

Conclusion This review presents an overview of recently published data regarding the spectrum of experiments that have focused on the increased therapeutic relevance of MSCs by means of genetic engineering applicable to regenerative medicine. The presented modifications relate to many aspects of MSC physiology, improve their natural ability to differentiate into bone or cartilage tissues and introduced other multiple approaches designed to improve the therapeutic potential of MSCs, which subsequently could be applied in various types of pathological incidences, such as cardiovascular and neurological diseases. All of these aspects, as well as safety considerations, pose an interesting target for future clinical trials with engineered MSCs.

24

Page 25 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

25 Acknowledgments The study was supported by a National Centre for Research and Development grant No. 101 in the ERA-NET NEURON project: “MEMS-IRBI”. We thank Mary McAllister for editorial assistance.

Author Disclosure Statement All of the authors declare no conflict of interest.

References 1.

Chou SH, SZ Lin, WW Kuo, P Pai, JY Lin, CH Lai, CH Kuo, KH Lin, FJ Tsai and CY

Huang. (2014). Mesenchymal stem cell insights: prospects in cardiovascular therapy. Cell Transplant. 23(4-5):513-529. 2.

Dasari VR, KK Veeravalli and DH Dinh. (2014). Mesenchymal stem cells in the treatment of

spinal cord injuries: A review. World J Stem Cells. 6(2):120-133. 3.

Serakinci N, U Fahrioglu and R Christensen. (2014). Mesenchymal stem cells, cancer

challenges and new directions. Eur J Cancer. 50(8):1522-1530. 4.

Pittenger MF, AM Mackay, SC Beck, RK Jaiswal, R Douglas, JD Mosca, MA Moorman, DW

Simonetti, S Craig and DR Marshak. (1999) Multilineage potential of adult human mesenchymal stem cells. Science. 284(5411):143-7. 5.

DiMarino et al. 2013: Dimarino AM, Al. Caplan and TL Bonfield. (2013) Mesenchymal

stem cells in tissue repair. Front Immunol. 4:201. 6.

Pittenger MF. (2013) MSCs: science and trials. Nat Med. 19(7):811.

7.

Filardo G, H Madry, M Jelic, A Roffi, M Cucchiarini and E Kon. (2013). Mesenchymal stem

cells for the treatment of cartilage lesions: from preclinical findings to clinical application in orthopaedics. Knee Surg Sports Traumatol Arthrosc. 21(8):1717-1729. 8

Knight MN and KD Hankenson. (2013). Mesenchymal Stem Cells in Bone Regeneration. Adv

Wound Care (New Rochelle). 2(6):306-316.

25

Page 26 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

26 9

Skalnikova H, J Motlik, SJ Gadher and H Kovarova. (2011). Mapping of the secretome of

primary isolates of mammalian cells, stem cells and derived cell lines. Proteomics. 11(4):691-708. 10 Huang F, ZF Fang, XQ Hu, L Tang, SH Zhou and JP Huang. (2013). Overexpression of miR126 promotes the differentiation of mesenchymal stem cells toward endothelial cells via activation of PI3K/Akt and MAPK/ERK pathways and release of paracrine factors. Biol Chem. 394(9):12231233. 11 Guo JW, C Chen, Y Huang and B Li. (2012). Combinatorial effects of naomai yihao capsules and vascular endothelial growth factor gene-transfected bone marrow mesenchymal stem cells on angiogenesis in cerebral ischemic tissues in rats. J Tradit Chin Med. 32(1):87-92. 12 Helmrich U, A Marsano, L Melly, T Wolff, L Christ, M Heberer, A Scherberich, I Martin and A Banfi. (2012). Generation of human adult mesenchymal stromal/stem cells expressing defined xenogenic vascular endothelial growth factor levels by optimized transduction and flow cytometry purification. Tissue Eng Part C Methods. 18(4):283-292. 13 Lai T, M Li, L Zheng, Y Song, X Xu, Y Guo, Y Zhang, Z Zhang and Y Mei. (2012). Overexpression of VEGF in marrow stromal cells promotes angiogenesis in rats with cerebral infarction via the synergistic effects of VEGF and Ang-2. J Huazhong Univ Sci Technolog Med Sci. 32(5):724-731. 14 Wang GX, L Hu, Z Zhang and DP Liu. (2014). Construction of an adenoviral expression vector carrying FLAG and hrGFP-1 genes and its expression in bone marrow mesenchymal stem cells. Genet Mol Res. 13(1):1070-1078. 15 Wang T, T Liao, H Wang, W Deng and D Yu. (2014). Transplantation of bone marrow stromal cells overexpressing human vascular endothelial growth factor 165 enhances tissue repair in a rat model of radiation-induced injury. Chin Med J (Engl). 127(6):1093-1099. 16 Moon HH, MK Joo, H Mok, M Lee, KC Hwang, SW Kim, JH Jeong, D Choi and SH Kim. (2014). MSC-based VEGF gene therapy in rat myocardial infarction model using facial amphipathic bile acid-conjugated polyethyleneimine. Biomaterials. 35(5):1744-1754.

26

Page 27 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

27 17 Alizadeh E, A Akbarzadeh, MB Eslaminejad, A Barzegar, S Hashemzadeh, K Nejati-Koshki and N Zarghami. (2014). Up Regulation of Liver-enriched Transcription Factors HNF4a and HNF6 and Liver-Specific MicroRNA (miR-122) by Inhibition of Let-7b in Mesenchymal Stem Cells. Chem Biol Drug Des. 2014 Jul 24. 18 Lee SY, O Ham, MJ Cha, BW Song, E Choi, IK Kim, W Chang, S Lim, CY Lee, JH Park, J Lee, Y Bae, HH Seo, E Choi, Y Jang and KC Hwang. (2013). The promotion of cardiogenic differentiation of hMSCs by targeting epidermal growth factor receptor using microRNA-133a. Biomaterials. 34(1):92-99. 19 Zhou YF, XJ Yang, HX Li, LH Han and WP Jiang. (2013). Genetically-engineered mesenchymal stem cells transfected with human HCN1 gene to create cardiac pacemaker cells. J Int Med Res. 41(5):1570-1576. 20 Du J, X Gao, L Deng, N Chang, H Xiong and Y Zheng. (2014). Transfection of the glial cell line-derived neurotrophic factor gene promotes neuronal differentiation. Neural Regen Res. 9(1):33-40. 21 Garcia KO, FL Ornellas, PK Martin, CL Patti, LE Mello, R Frussa-Filho, SW Han and BM Longo. (2014). Therapeutic effects of the transplantation of VEGF overexpressing bone marrow mesenchymal stem cells in the hippocampus of murine model of Alzheimer's disease. Front Aging Neurosci. 6:30. 22 Jeong CH, SM Kim, JY Lim, CH Ryu, JA Jun and SS Jeun. (2014). Mesenchymal stem cells expressing brain-derived neurotrophic factor enhance endogenous neurogenesis in an ischemic stroke model. Biomed Res Int. 2014:129145. 23 Liu X, W Shen, Y Yang and G Liu. (2011). Therapeutic implications of mesenchymal stem cells transfected with hepatocyte growth factor transplanted in rat kidney with unilateral ureteral obstruction. J Pediatr Surg. 46(3):537-545. 24 Levy O, W Zhao, LJ Mortensen, S Leblanc, K Tsang, M Fu, JA Phillips, V Sagar, P Anandakumaran, J Ngai, CH Cui, P Eimon, M Angel, CP Lin, MF Yanik and JM Karp. (2013).

27

Page 28 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

28 mRNA-engineered mesenchymal stem cells for targeted delivery of interleukin-10 to sites of inflammation. Blood. 122(14):e23-32. 25 Boura JS, M Vance, W Yin, C Madeira, C Lobato da Silva, CD Porada and G AlmeidaPorada. (2014) Evaluation of gene delivery strategies to efficiently overexpress functional HLA-G on human bone marrow stromal cells. Mol Ther Methods Clin Dev. 2014(1). 26 Marquez-Curtis LA, H Gul-Uludag, P Xu, J Chen and A Janowska-Wieczorek. (2013) CXCR4 transfection of cord blood mesenchymal stromal cells with the use of cationic liposome enhances their migration toward stromal cell-derived factor-1. Cytotherapy. 15(7):840-9. 27 Watkins DJ, Y Zhou, MA Matthews, L Chen and GE Besner. (2014). HB-EGF augments the ability of mesenchymal stem cells to attenuate intestinal injury. J Pediatr Surg. 49(6):938-944; discussion 944. 28 Mok PL, SK Cheong, CF Leong, KH Chua and O Ainoon. (2012) Extended and stable gene expression via nucleofection of MIDGE construct into adult human marrow mesenchymal stromal cells. Cytotechnology. 64(2):203-16. 29 Martin PK, RS Stilhano, VY Samoto, CM Takiya, GB Peres, YM da Silva Michelacci, FH da Silva, VG Pereira, V D'Almeida, FL Marques, AH Otake, R Chammas and SW Han. (2014) Mesenchymal stem cells do not prevent antibody responses against human α-L-iduronidase when used to treat mucopolysaccharidosis type I. PLoS One. 9(3):e92420. 30 Rejman J, G Tavernier, N Bavarsad, J Demeester and SC De Smedt. (2010). mRNA transfection of cervical carcinoma and mesenchymal stem cells mediated by cationic carriers. J Control Release. 147(3):385-391. 31 Meng YB, X Li, ZY Li, J Zhao, XB Yuan, Y Ren, ZD Cui, YD Liu and XJ Yang. (2015) microRNA-21 promotes osteogenic differentiation of mesenchymal stem cells by the PI3K/βcatenin pathway. J Orthop Res. 33(7):957-64. 32 Ríos CN, RJ Skoracki and AB Mathur. (2012) GNAS1 and PHD2 short-interfering RNA support bone regeneration in vitro and in an in vivo sheep model. Clin Orthop Relat Res. 470(9):2541-53. 28

Page 29 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

29 33 Nowakowski A, A Andrzejewska, M Janowski, P Walczak and B Lukomska. (2013). Genetic engineering of stem cells for enhanced therapy. Acta Neurobiol Exp (Wars). 73(1):1-18. 34 Li J, CQ Zheng, Y Li, C Yang, H Lin and HG Duan. (2015) Hepatocyte Growth Factor GeneModified Mesenchymal Stem Cells Augment Sinonasal Wound Healing. Stem Cells Dev. 2015 Apr 30. [Epub ahead of print] 35 Yang JX, N Zhang, HW Wang, P Gao, QP Yang and QP Wen. (2015) CXCR4 receptor overexpression in mesenchymal stem cells facilitates treatment of acute lung injury in rats. J Biol Chem. 290(4):1994-2006. 36 Eun LY, BW Song, MJ Cha, H Song, IK Kim, E Choi, W Chang, S Lim, EJ Choi, O Ham, SY Lee, KH Byun, Y Jang and KC Hwang. (2010) Overexpression of phosphoinositide-3-kinase class II alpha enhances mesenchymal stem cell survival in infarcted myocardium. Biochem Biophys Res Commun. 402(2):272-9. 37 Xu L, S Huang, Y Hou, Y Liu, M Ni, F Meng, K Wang, Y Rui, X Jiang and G Li. (2015) Sox11-modified mesenchymal stem cells (MSCs) accelerate bone fracture healing: Sox11 regulates differentiation and migration of MSCs. FASEB J. 29(4):1143-52. 38 Mariner PD, E Johannesen and KS Anseth. (2012) Manipulation of miRNA activity accelerates osteogenic differentiation of hMSCs in engineered 3D scaffolds. J Tissue Eng Regen Med. 6(4):314-24. 39 Crafts TD, AR Jensen, EC Blocher-Smith and TA Markel. (2014). Vascular endothelial growth factor: Therapeutic possibilities and challenges for the treatment of ischemia. Cytokine. pii:S1043-4666(14)00523-7. 40 Fan L, C Lin, S Zhuo, L Chen, N Liu, Y Luo, J Fang, Z Huang, Y Lin and J Chen. (2009). Transplantation with survivin-engineered mesenchymal stem cells results in better prognosis in a rat model of myocardial infarction. Eur J Heart Fail. 11(11):1023-1030. 41 Abbate A, S Scarpa, D Santini, J Palleiro, F Vasaturo, J Miller, C Morales, GW Vetrovec and A Baldi. (2006). Myocardial expression of survivin, an apoptosis inhibitor, in aging and heart failure. An experimental study in the spontaneously hypertensive rat. Int J Cardiol. 111(3):371-376. 29

Page 30 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

30 42 Han Y, R Tao, Y Han, T Sun, J Chai, G Xu and J Liu. (2014). Microencapsulated VEGF genemodified umbilical cord mesenchymal stromal cells promote the vascularization of tissueengineered dermis: an experimental study. Cytotherapy. 16(2):160-169. 43 Tang H, Y Xiang, X Jiang, Y Ke, Z Xiao, Y Guo, Q Wang, M Du, L Qin, Y Zou, Y Cai, Z Chen and R Xu. (2013). Dual expression of hTERT and VEGF prolongs life span and enhances angiogenic ability of aged BMSCs. Biochem Biophys Res Commun. 440(4):502-508. 44 Katare R, F Riu, J Rowlinson, A Lewis, R Holden, M Meloni, C Reni, C Wallrapp, C Emanueli and P Madeddu. (2013). Perivascular delivery of encapsulated mesenchymal stem cells improves postischemic angiogenesis via paracrine activation of VEGF-A. Arterioscler Thromb Vasc Biol. 33(8):1872-1880. 45 Arita Y, Y Nakaoka, T Matsunaga, H Kidoya, K Yamamizu, Y Arima, T Kataoka-Hashimoto, K Ikeoka, T Yasui, T Masaki, K Yamamoto, K Higuchi, JS Park, M Shirai, K Nishiyama, H Yamagishi, K Otsu, H Kurihara, T Minami, K Yamauchi-Takihara, GY Koh, N Mochizuki, N Takakura, Y Sakata, JK Yamashita and I Komuro. (2014). Myocardium-derived angiopoietin-1 is essential for coronary vein formation in the developing heart. Nat Commun. 5:4552. 46 Piao W, H Wang, M Inoue, M Hasegawa, H Hamada and J Huang. (2010). Transplantation of Sendai viral angiopoietin-1-modified mesenchymal stem cells for ischemic limb disease. Angiogenesis. 13(3):203-210. 47 Liu X, B Zeng and C Zhang. (2011). Osteogenic and angiogenic effects of mesenchymal stromal cells with co-transfected human Ang-1 gene and BMP2 gene. Biotechnol Lett. 33(10):1933-1938. 48 Hua J, ZG He, DH Qian, SP Lin, J Gong, HB Meng, TS Yang, W Sun, B Xu, B Zhou and ZS Song. (2014). Angiopoietin-1 gene-modified human mesenchymal stem cells promote angiogenesis and reduce acute pancreatitis in rats. Int J Clin Exp Pathol. 7(7):3580-3595. 49 da Cunha FF, L Martins, PL Martin, RS Stilhano, EJ Paredes Gamero and SW Han. (2013). Comparison of treatments of peripheral arterial disease with mesenchymal stromal cells and

30

Page 31 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

31 mesenchymal stromal cells modified with granulocyte and macrophage colony-stimulating factor. Cytotherapy. 15(7):820-829. 50 Su GH, YF Sun, YX Lu, XX Shuai, YH Liao, QY Liu, J Han and P Luo. (2013). Hepatocyte growth factor gene-modified bone marrow-derived mesenchymal stem cells transplantation promotes angiogenesis in a rat model of hindlimb ischemia. J Huazhong Univ Sci Technolog Med Sci. 33(4):511-519. 51 Zhang JC, GF Zheng, L Wu, LY Ou Yang and WX Li. (2014). Bone marrow mesenchymal stem cells overexpressing human basic fibroblast growth factor increase vasculogenesis in ischemic rats. Braz J Med Biol Res. 47(10):886-894. 52 Yao J, SL Jiang, W Liu, C Liu, W Chen, L Sun, KY Liu, ZB Jia, RK Li and H Tian. (2012). Tissue inhibitor of matrix metalloproteinase-3 or vascular endothelial growth factor transfection of aged human mesenchymal stem cells enhances cell therapy after myocardial infarction. Rejuvenation Res. 15(5):495-506. 53 Dao M, CC Tate, M McGrogan and CC Case. (2013). Comparing the angiogenic potency of naïve marrow stromal cells and Notch-transfected marrow stromal cells. J Transl Med. 11:81. 54 Zeng B, G Lin, X Ren, Y Zhang and H Chen. (2010). Over-expression of HO-1 on mesenchymal stem cells promotes angiogenesis and improves myocardial function in infarcted myocardium. J Biomed Sci. 17:80. 55 Li H, S Zuo, Z He, Y Yang, Z Pasha, Y Wang and M Xu. (2010). Paracrine factors released by GATA-4 overexpressed mesenchymal stem cells increase angiogenesis and cell survival. Am J Physiol Heart Circ Physiol. 299(6):H1772-1781. 56 Razban V, AS Lotfi, M Soleimani, H Ahmadi, M Massumi, S Khajeh, M Ghaedi, S Arjmand, S Najavand and A Khoshdel. (2012). HIF-1α Overexpression Induces Angiogenesis in Mesenchymal Stem Cells. Biores Open Access. 1(4):174-183. 57 Liu J, W Li, Y Wang, W Fan, P Li, W Lin, D Yang, R Fang, M Feng, C Hu, Z Du, G Wu and AP Xiang. (2014). Islet-1 overexpression in human mesenchymal stem cells promotes vascularization through monocyte chemoattractant protein-3. Stem Cells. 32(7):1843-1854. 31

Page 32 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

32 58 Fish JE, MM Santoro, SU Morton, S Yu, RF Yeh, JD Wythe, KN Ivey, BG Bruneau, DY Stainier and D Srivastava. (2008). miR-126 regulates angiogenic signaling and vascular integrity. Dev Cell. 15(2):272-284. 59 Chen JJ and SH Zhou. (2011). Mesenchymal stem cells overexpressing MiR-126 enhance ischemic angiogenesis via the AKT/ERK-related pathway. Cardiol J. 18(6):675-681. 60 Huang F, X Zhu, XQ Hu, ZF Fang, L Tang, XL Lu and SH Zhou. (2013). Mesenchymal stem cells modified with miR-126 release angiogenic factors and activate Notch ligand Delta-like-4, enhancing ischemic angiogenesis and cell survival. Int J Mol Med. 31(2):484-492. 61 He HL, L Liu, QH Chen, SX Cai, JB Han, SL Hu, P Chun, Y Yang, FM Guo, YZ Huang and HB Qiu. (2014). MSCs modified with ACE2 restore endothelial function following LPS challenge by inhibiting the activation of RAS. J Cell Physiol. 230(3):691-701. 62 Montezano AC, A Nguyen Dinh Cat, FJ Rios and RM Touyz. (2014). Angiotensin II and vascular injury. Curr Hypertens Rep. 16(6):431. 63 Ferreira AJ, TM Murça, RA Fraga-Silva, CH Castro, MK Raizada and RA Santos. (2012). New cardiovascular and pulmonary therapeutic strategies based on the Angiotensin-converting enzyme 2/angiotensin-(1-7)/mas receptor axis. Int J Hypertens. 2012:147825. 64 Peter A, C Weigert, H Staiger, K Rittig, A Cegan, P Lutz, F Machicao, HU Häring and E Schleicher. (2008). Induction of stearoyl-CoA desaturase protects human arterial endothelial cells against lipotoxicity. Am J Physiol Endocrinol Metab. 295(2):E339-349. 65 Lu Y, Z Zhou, J Tao, B Dou, M Gao and Y Liu. (2014). Overexpression of stearoyl-CoA desaturase 1 in bone marrow mesenchymal stem cells enhance the expression of induced endothelial cells. Lipids Health Dis. 13:53. 66 Sayyar B, M Dodd, J Wen, S Ma, L Marquez-Curtis, A Janowska-Wieczorek and G Hortelano. (2012). Encapsulation of factor IX-engineered mesenchymal stem cells in fibrinogenalginate microcapsules enhances their viability and transgene secretion. J Tissue Eng. 3(1):2041731412462018.

32

Page 33 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

33 67 Lodish H, J Flygare and S Chou. (2010). From stem cell to erythroblast: regulation of red cell production at multiple levels by multiple hormones. IUBMB Life. 62(7):492-496. 68 Mok PL, SK Cheong, CF Leong, KH Chua and O Ainoon. (2012). Human mesenchymal stromal cells could deliver erythropoietin and migrate to the basal layer of hair shaft when subcutaneously implanted in a murine model. Tissue Cell. 44(4):249-256. 69 Scheibe F, N Gladow, P Mergenthaler, AH Tucker, A Meisel, DJ Prockop and J Priller. (2012). Nonviral gene delivery of erythropoietin by mesenchymal stromal cells. Gene Ther. 19(5):550-560. 70 Kiani AA, J Abdi, R Halabian, MH Roudkenar, N Amirizadeh, M Soleiman Soltanpour and A Kazemi. (2014). Over expression of HIF-1α in human mesenchymal stem cells increases their supportive functions for hematopoietic stem cells in an experimental co-culture model. Hematology. 19(2):85-98. 71 Duryagina R, S Thieme, K Anastassiadis, C Werner, S Schneider, M Wobus, S Brenner and M Bornhäuser. (2013). Overexpression of Jagged-1 and its intracellular domain in human mesenchymal stromal cells differentially affect the interaction with hematopoietic stem and progenitor cells. Stem Cells Dev. 22(20):2736-2750. 72 Hu K, C Xu, H Ni, Z Xu, Y Wang, S Xu, K Ji, J Xiong and H Liu. (2014). Mir-218 contributes to the transformation of 5-Aza/GF induced umbilical cord mesenchymal stem cells into hematopoietic cells through the MITF pathway. Mol Biol Rep. 41(7):4803-4816. 73 Johnstone B, TM Hering, AI Caplan, VM Goldberg and JU Yoo. (1998). In vitro chondrogenesis of bone marrow-derived mesenchymal progenitor cells. Exp Cell Res. 238(1):265272. 74 Chen X, F Zhang, X He, Y Xu, Z Yang, L Chen, S Zhou, Y Yang, Z Zhou, W Sheng and Y Zeng. (2013). Chondrogenic differentiation of umbilical cord-derived mesenchymal stem cells in type I collagen-hydrogel for cartilage engineering. Injury. 44(4):540-549. 75 Montoya F, F Martínez, M García-Robles, C Balmaceda-Aguilera, X Koch, F Rodríguez, C Silva-Álvarez, K Salazar, V Ulloa and F Nualart. (2013). Clinical and experimental approaches to 33

Page 34 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

34 knee cartilage lesion repair and mesenchymal stem cell chondrocyte differentiation. Biol Res. 46(4):441-451. 76 Moshaverinia A, X Xu, C Chen, K Akiyama, ML Snead and S Shi. (2013). Dental mesenchymal stem cells encapsulated in an alginate hydrogel co-delivery microencapsulation system for cartilage regeneration. Acta Biomater. 9(12):9343-9350. 77 Tuli R, S Tuli, S Nandi, X Huang, PA Manner, WJ Hozack, KG Danielson, DJ Hall and RS Tuan. (2003). Transforming growth factor-beta-mediated chondrogenesis of human mesenchymal progenitor cells involves N-cadherin and mitogen-activated protein kinase and Wnt signaling crosstalk. J Biol Chem. 278(42):41227-41236. 78 Deng WW, X Cao, M Wang, R Qu, WY Su, Y Yang, YW Wei, XM Xu and JN Yu. (2012). Delivery of a transforming growth factor β-1 plasmid to mesenchymal stem cells via cationized Pleurotus eryngii polysaccharide nanoparticles. Int J Nanomedicine. 7:1297-1311. 79 Zhu S, T Zhang, C Sun, A Yu, B Qi and H Cheng. (2013). Bone marrow mesenchymal stem cells combined with calcium alginate gel modified by hTGF-β1 for the construction of tissueengineered cartilage in three-dimensional conditions. Exp Ther Med. 5(1):95-101. 80 Kim YI, JS Ryu, JE Yeo, YJ Choi, YS Kim, K Ko and YG Koh. (2014). Overexpression of TGF-β1 enhances chondrogenic differentiation and proliferation of human synovium-derived stem cells. Biochem Biophys Res Commun. 450(4):1593-1599. 81 Wu W, Y Dan, SH Yang, C Yang, ZW Shao, WH Xu, J Li, XZ Liu and D Zheng. (2013). Promotion of chondrogenesis of marrow stromal stem cells by TGF-β3 fusion protein in vitro. J Huazhong Univ Sci Technolog Med Sci. 33(5):692-699. 82 Sun L, H Li, L Qu, R Zhu, X Fan, Y Xue, Z Xie and H Fan. (2014). Immobilized lentivirus vector on chondroitin sulfate-hyaluronate acid-silk fibroin hybrid scaffold for tissue-engineered ligament-bone junction. Biomed Res Int. 2014:816979. 83 Wu G, Y Cui, YT Wang, M Yao, J Hu, JX Li, Y Wang and B Zhang. (2014). Repair of cartilage defects in BMSCs via CDMP1 gene transfection. Genet Mol Res. 13(1):291-301.

34

Page 35 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

35 84 Danišovič L, I Varga and S Polák. (2012). Growth factors and chondrogenic differentiation of mesenchymal stem cells. Tissue Cell. 44(2):69-73. 85 An C, Y Cheng, Q Yuan and J Li. (2010). IGF-1 and BMP-2 induces differentiation of adipose-derived mesenchymal stem cells into chondrocytes-like cells. Ann Biomed Eng. 38(4):1647-1654. 86 Wang GX, L Hu, HX Hu, Z Zhang and DP Liu. (2014). In vivo osteogenic activity of bone marrow stromal stem cells transfected with Ad-GFP-hBMP-2. Genet Mol Res. 13(2):4456-4465. 87 Knippenberg M, MN Helder, B Zandieh Doulabi, PI Wuisman and J Klein-Nulend. (2006). Osteogenesis versus chondrogenesis by BMP-2 and BMP-7 in adipose stem cells. Biochem Biophys Res Commun. 342(3):902-908. 88 Bai X, G Li, C Zhao, H Duan and F Qu. (2011). BMP7 induces the differentiation of bone marrow-derived mesenchymal cells into chondrocytes. Med Biol Eng Comput. 49(6):687-692. 89 Frisch J, JK Venkatesan, A Rey-Rico, G Schmitt, H Madry and M Cucchiarini. (2014). Influence of insulin-like growth factor I overexpression via recombinant adeno-associated vector gene transfer upon the biological activities and differentiation potential of human bone marrowderived mesenchymal stem cells. Stem Cell Res Ther. 5(4):103. 90 Garza-Veloz I, VJ Romero-Diaz, ML Martinez-Fierro, IA Marino-Martinez, M GonzalezRodriguez, HG Martinez-Rodriguez, MA Espinoza-Juarez, DA Bernal-Garza, R Ortiz-Lopez and A Rojas-Martinez. (2013). Analyses of chondrogenic induction of adipose mesenchymal stem cells by combined co-stimulation mediated by adenoviral gene transfer. Arthritis Res Ther. 15(4):R80. 91 Liu S, E Zhang, M Yang and L Lu. (2014). Overexpression of Wnt11 promotes chondrogenic differentiation of bone marrow-derived mesenchymal stem cells in synergism with TGF-β. Mol Cell Biochem. 390(1-2):123-131. 92 Akiyama H, MC Chaboissier, JF Martin, A Schedl and B de Crombrugghe. (2002). The transcription factor Sox9 has essential roles in successive steps of the chondrocyte differentiation pathway and is required for expression of Sox5 and Sox6. Genes Dev. 16(21):2813-2828.

35

Page 36 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

36 93 Kim JH, JS Park, HN Yang, DG Woo, SY Jeon, HJ Do, HY Lim, JM Kim and KH Park. (2011). The use of biodegradable PLGA nanoparticles to mediate SOX9 gene delivery in human mesenchymal stem cells (hMSCs) and induce chondrogenesis. Biomaterials. 32(1):268-278. 94 Jeon SY, JS Park, HN Yang, DG Woo and KH Park. (2012). Co-delivery of SOX9 genes and anti-Cbfa-1 siRNA coated onto PLGA nanoparticles for chondrogenesis of human MSCs. Biomaterials. 33(17):4413-4423. 95 Venkatesan JK, M Ekici, H Madry, G Schmitt, D Kohn and M Cucchiarini. (2012). SOX9 gene transfer via safe, stable, replication-defective recombinant adeno-associated virus vectors as a novel, powerful tool to enhance the chondrogenic potential of human mesenchymal stem cells. Stem Cell Res Ther. 3(3):22. 96 Frith J and P Genever. (2008) Transcriptional control of mesenchymal stem cell differentiation. Transfus Med Hemother. 35(3):216-27. 97 Liu TM, XM Guo, HS Tan, JH Hui, B Lim and EH Lee. (2011). Zinc-finger protein 145, acting as an upstream regulator of SOX9, improves the differentiation potential of human mesenchymal stem cells for cartilage regeneration and repair. Arthritis Rheum. 63(9):2711-2720. 98 Kim HJ and GI Im. (2011). Electroporation-mediated transfer of SOX trio genes (SOX-5, SOX-6, and SOX-9) to enhance the chondrogenesis of mesenchymal stem cells. Stem Cells Dev. 20(12):2103-2114. 99 Park JS, HN Yang, DG Woo, SY Jeon, HJ Do, HY Lim, JH Kim and KH Park. (2011). Chondrogenesis of human mesenchymal stem cells mediated by the combination of SOX trio SOX5, 6, and 9 genes complexed with PEI-modified PLGA nanoparticles. Biomaterials. 32(14):3679-3688. 100 Yang HN, JS Park, DG Woo, SY Jeon, HJ Do, HY Lim, SW Kim, JH Kim and KH Park. (2011). Chondrogenesis of mesenchymal stem cells and dedifferentiated chondrocytes by transfection with SOX Trio genes. Biomaterials. 32(30):7695-7704.

36

Page 37 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

37 101 Fang Z, Q Yang, W Luo, GH Li, J Xiao, F Li and W Xiong. (2013). Differentiation of GFPBcl-2-engineered mesenchymal stem cells towards a nucleus pulposus-like phenotype under hypoxia in vitro. Biochem Biophys Res Commun. 432(3):444-450. 102 Bucher C, A Gazdhar, LM Benneker, T Geiser and B Gantenbein-Ritter. (2013). Nonviral Gene Delivery of Growth and Differentiation Factor 5 to Human Mesenchymal Stem Cells Injected into a 3D Bovine Intervertebral Disc Organ Culture System. Stem Cells Int. 2013:326828. 103 Coleman CM, EE Vaughan, DC Browe, E Mooney, L Howard and F Barry. (2013). Growth differentiation factor-5 enhances in vitro mesenchymal stromal cell chondrogenesis and hypertrophy. Stem Cells Dev. 22(13):1968-1976. 104 Chujo T, HS An, K Akeda, K Miyamoto, C Muehleman, M Attawia, G Andersson and K Masuda. (2006). Effects of growth differentiation factor-5 on the intervertebral disc--in vitro bovine study and in vivo rabbit disc degeneration model study. Spine (Phila Pa 1976). 31(25):29092917. 105 Vo NV, RA Hartman, T Yurube, LJ Jacobs, GA Sowa and JD Kang. (2013). Expression and regulation of metalloproteinases and their inhibitors in intervertebral disc aging and degeneration. Spine J. 13(3):331-341. 106 Yi Z, T Guanjun, C Lin and P Zifeng. (2014). Effects of Transplantation of hTIMP1Expressing Bone Marrow Mesenchymal Stem Cells on the Extracellular Matrix of Degenerative Intervertebral Discs in an in vivo Rabbit Model. Spine (Phila Pa 1976). 2014 Apr 8. 107 Holloway JL, H Ma, R Rai and JA Burdick. (2014). Modulating hydrogel crosslink density and degradation to control bone morphogenetic protein delivery and in vivo bone formation. J Control Release. 191:63-70. 108 Wutzl A, M Rauner, R Seemann, W Millesi, P Krepler, P Pietschmann and R Ewers. (2010). Bone morphogenetic proteins 2, 5, and 6 in combination stimulate osteoblasts but not osteoclasts in vitro. J Orthop Res. 28(11):1431-1439.

37

Page 38 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

38 109 Zhang Z, G Wang, C Li and D Liu. (2013). Construction and characterization of a recombinant human adenovirus vector expressing bone morphogenetic protein 2. Exp Ther Med. 6(2):329-334. 110 Jin H, K Zhang, C Qiao, A Yuan, D Li, L Zhao, C Shi, X Xu, S Ni, C Zheng, X Liu, B Yang and H Sun. (2014). Efficiently engineered cell sheet using a complex of polyethylenimine-alginate nanocomposites plus bone morphogenetic protein 2 gene to promote new bone formation. Int J Nanomedicine. 9:2179-2190. 111 Wei L, GH Lei, HW Yi and PY Sheng. (2014). Bone formation in rabbit's leg muscle after autologous transplantation of bone marrow-derived mesenchymal stem cells expressing human bone morphogenic protein-2. Indian J Orthop. 48(4):347-353. 112 Moshaverinia A, S Ansari, C Chen, X Xu, K Akiyama, ML Snead, HH Zadeh and S Shi. (2013). Co-encapsulation of anti-BMP2 monoclonal antibody and mesenchymal stem cells in alginate microspheres for bone tissue engineering. Biomaterials. 34(28):6572-6579. 113 Herberg S, S Fulzele, N Yang, X Shi, M Hess, S Periyasamy-Thandavan, MW Hamrick, CM Isales and WD Hill. (2013). Stromal cell-derived factor-1β potentiates bone morphogenetic protein2-stimulated osteoinduction of genetically engineered bone marrow-derived mesenchymal stem cells in vitro. Tissue Eng Part A. 19(1-2):1-13. 114 Liu D, L Hu, Z Zhang, QY Li and G Wang. (2013). Construction of human BMP2-IRESHIF1αmu adenovirus expression vector and its expression in mesenchymal stem cells. Mol Med Rep. 7(2):659-663. 115 Zou D, W Han, S You, D Ye, L Wang, S Wang, J Zhao, W Zhang, X Jiang, X Zhang and Y Huang. (2011). In vitro study of enhanced osteogenesis induced by HIF-1α-transduced bone marrow stem cells. Cell Prolif. 44(3):234-243. 116 Xiao C, H Zhou, G Liu, P Zhang, Y Fu, P Gu, H Hou, T Tang and X Fan. (2011). Bone marrow stromal cells with a combined expression of BMP-2 and VEGF-165 enhanced bone regeneration. Biomed Mater. 6(1):015013.

38

Page 39 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

39 117 Lin Z, JS Wang, L Lin, J Zhang, Y Liu, M Shuai and Q Li. (2014). Effects of BMP2 and VEGF165 on the osteogenic differentiation of rat bone marrow-derived mesenchymal stem cells. Exp Ther Med. 7(3):625-629. 118 Zhang C, HM Liu, QW Li, GW Chen, X Liang and CY Meng. (2014). Construction of recombinant adenovirus vector containing hBMP2 and hVEGF165 genes and its expression in rabbit Bone marrow mesenchymal stem cells. Tissue Cell. 46(5):311-317. 119 Liu B, X Li, G Liang and X Liu. (2011). VEGF expression in mesenchymal stem cells promotes bone formation of tissue-engineered bones. Mol Med Rep. 4(6):1121-1126. 120 Kasten P, M Beverungen, H Lorenz, J Wieland, M Fehr and F Geiger. (2012). Comparison of platelet-rich plasma and VEGF-transfected mesenchymal stem cells on vascularization and bone formation in a critical-size bone defect. Cells Tissues Organs. 196(6):523-533. 121 Li C, J Ding, L Jiang, C Shi, S Ni, H Jin, D Li and H Sun. (2014). Potential of mesenchymal stem cells by adenovirus-mediated erythropoietin gene therapy approaches for bone defect. Cell Biochem Biophys. 70(2):1199-1204. 122 Holstein JH, M Orth, C Scheuer, A Tami, SC Becker, P Garcia, T Histing, P Mörsdorf, M Klein, T Pohlemann and MD Menger. (2011). Erythropoietin stimulates bone formation, cell proliferation, and angiogenesis in a femoral segmental defect model in mice. Bone. 49(5):10371045. 123 Shiozawa Y, Y Jung, AM Ziegler, EA Pedersen, J Wang, Z Wang, J Song, J Wang, CH Lee, S Sud, KJ Pienta, PH Krebsbach and RS Taichman. (2010). Erythropoietin couples hematopoiesis with bone formation. PLoS One. 5(5):e10853. 124 He X, R Dziak, X Yuan, K Mao, R Genco, M Swihart, D Sarkar, C Li, C Wang, L Lu, S Andreadis and S Yang. (2013). BMP2 genetically engineered MSCs and EPCs promote vascularized bone regeneration in rat critical-sized calvarial bone defects. PLoS One. 8(4):e60473. 125 Yuan SH and Z Bi. (2012). Effect of recombinant adeno-associated BMP-4/7 fusion gene on the biology of BMSCs. Mol Med Rep. 6(6):1413-1417.

39

Page 40 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

40 126 Mizrahi O, D Sheyn, W Tawackoli, I Kallai, A Oh, S Su, X Da, P Zarrini, G Cook-Wiens, D Gazit and Z Gazit. (2013). BMP-6 is more efficient in bone formation than BMP-2 when overexpressed in mesenchymal stem cells. Gene Ther. 20(4):370-377. 127 An G, Z Xue, B Zhang, QK Deng, YS Wang and SC Lv. (2014). Expressing osteogenic growth peptide in the rabbit bone mesenchymal stem cells increased alkaline phosphatase activity and enhanced the collagen accumulation. Eur Rev Med Pharmacol Sci. 18(11):1618-1624. 128 Granero-Moltó F, TJ Myers, JA Weis, L Longobardi, T Li, Y Yan, N Case, J Rubin and A Spagnoli. (2011). Mesenchymal stem cells expressing insulin-like growth factor-I (MSCIGF) promote fracture healing and restore new bone formation in Irs1 knockout mice: analyses of MSCIGF autocrine and paracrine regenerative effects. Stem Cells. 29(10):1537-1548. 129 Sheng MH, KH Lau and DJ Baylink. (2014). Role of Osteocyte-derived Insulin-Like Growth Factor I in Developmental Growth, Modeling, Remodeling, and Regeneration of the Bone. J Bone Metab. 21(1):41-54. 130 Han G, Y Jing, Y Zhang, Z Yue, X Hu, L Wang, J Liang and J Liu. (2010). Osteogenic differentiation of bone marrow mesenchymal stem cells by adenovirus-mediated expression of leptin. Regul Pept. 163(1-3):107-112. 131 Elangovan S, SR D'Mello, L Hong, RD Ross, C Allamargot, DV Dawson, CM Stanford, GK Johnson, DR Sumner and AK Salem. (2014). The enhancement of bone regeneration by gene activated matrix encoding for platelet derived growth factor. Biomaterials. 35(2):737-747. 132 Tierney EG, K McSorley, CL Hastings, SA Cryan, T O'Brien, MJ Murphy, FP Barry, FJ O'Brien and GP Duffy. (2013). High levels of ephrinB2 over-expression increases the osteogenic differentiation of human mesenchymal stem cells and promotes enhanced cell mediated mineralisation in a polyethyleneimine-ephrinB2 gene-activated matrix. J Control Release. 165(3):173-182. 133 Hu HM, L Yang, Z Wang, YW Liu, JZ Fan, J Fan, J Liu and ZJ Luo. (2013). Overexpression of integrin a2 promotes osteogenic differentiation of hBMSCs from senile osteoporosis through the ERK pathway. Int J Clin Exp Pathol. 6(5):841-852. 40

Page 41 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

41 134 Liu Y, C Chen, H He, D Wang, L E, Z Liu and H Liu. (2012). Lentiviral-mediated gene transfer into human adipose-derived stem cells: role of NELL1 versus BMP2 in osteogenesis and adipogenesis in vitro. Acta Biochim Biophys Sin (Shanghai). 44(10):856-865. 135 Cai SX, AR Liu, HL He, QH Chen, Y Yang, FM Guo, YZ Huang, L Liu and HB Qiu. (2014). Stable genetic alterations of β-catenin and ROR2 regulate the Wnt pathway, affect the fate of MSCs. J Cell Physiol. 229(6):791-800. 136 Li ZH, W Liao, Q Zhao, T Huan, P Feng, X Wei, Y Yi and NS Shao. (2014). Effect of Cbfa1 on osteogenic differentiation of mesenchymal stem cells under hypoxia condition. Int J Clin Exp Med. 7(3):540-548. 137 Suh JS, JY Lee, YJ Choi, HK You, SD Hong, CP Chung and YJ Park. (2014). Intracellular delivery of cell-penetrating peptide-transcriptional factor fusion protein and its role in selective osteogenesis. Int J Nanomedicine. 9:1153-1166. 138 You W, L Fan, D Duan, L Tian, X Dang, C Wang and K Wang. (2014). Foxc2 overexpression in bone marrow mesenchymal stem cells stimulates osteogenic differentiation and inhibits adipogenic differentiation. Mol Cell Biochem. 386(1-2):125-134. 139 Ogasawara T, S Ohba, F Yano, H Kawaguchi, UI Chung, T Saito, Y Yonehara, T Nakatsuka, Y Mori, T Takato and K Hoshi. (2013). Nanog promotes osteogenic differentiation of the mouse mesenchymal cell line C3H10T1/2 by modulating bone morphogenetic protein (BMP) signaling. J Cell Physiol. 228(1):163-171. 140 Li Z, MQ Hassan, M Jafferji, RI Aqeilan, R Garzon, CM Croce, AJ van Wijnen, JL Stein, GS Stein and JB Lian. (2009). Biological functions of miR-29b contribute to positive regulation of osteoblast differentiation. J Biol Chem. 284(23):15676-15684. 141 Suh JS, JY Lee, YS Choi, CP Chung and YJ Park. (2013). Peptide-mediated intracellular delivery of miRNA-29b for osteogenic stem cell differentiation. Biomaterials. 34(17):4347-4359. 142 Liao YH, YH Chang, LY Sung, KC Li, CL Yeh, TC Yen, SM Hwang, KJ Lin and YC Hu. (2014). Osteogenic differentiation of adipose-derived stem cells and calvarial defect repair using baculovirus-mediated co-expression of BMP-2 and miR-148b. Biomaterials. 35(18):4901-4910. 41

Page 42 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

42 143 Deng Y, X Bi, H Zhou, Z You, Y Wang, P Gu and X Fan. (2014). Repair of critical-sized bone defects with anti-miR-31-expressing bone marrow stromal stem cells and poly(glycerol sebacate) scaffolds. Eur Cell Mater. 27:13-24; discussion 24-25. 144 Tan Y, EH Xiao, LZ Xiao, YH Yuan, C Ma, QL Shang, DJ Bian, YH Li, Z Chen and Q Chang. (2012). VEGF(165) expressing bone marrow mesenchymal stem cells differentiate into hepatocytes under HGF and EGF induction in vitro. Cytotechnology. 64(6):635-647. 145 Cereghini S. (1996). Liver-enriched transcription factors and hepatocyte differentiation. FASEB J. 10(2):267-282. 146 Zamule SM, DM Coslo, F Chen and CJ Omiecinski. (2011). Differentiation of human embryonic stem cells along a hepatic lineage. Chem Biol Interact. 190(1):62-72. 147 Hang H, Y Yu, N Wu, Q Huang, Q Xia and J Bian. (2014). Induction of highly functional hepatocytes from human umbilical cord mesenchymal stem cells by HNF4α transduction. PLoS One. 9(8):e104133. 148 Cho JW, CY Lee and Y Ko. (2012). Therapeutic potential of mesenchymal stem cells overexpressing human forkhead box A2 gene in the regeneration of damaged liver tissues. J Gastroenterol Hepatol. 27(8):1362-1370. 149 Davoodian N, AS Lotfi, M Soleimani, SJ Mola and S Arjmand. (2014). Let-7f microRNA negatively regulates hepatic differentiation of human adipose tissue-derived stem cells. J Physiol Biochem. 70(3):781-789. 150 Deng XG, RL Qiu, YH Wu, ZX Li, P Xie, J Zhang, JJ Zhou, LX Zeng, J Tang, A Maharjan and JM Deng. (2014). Overexpression of miR-122 promotes the hepatic differentiation and maturation of mouse ESCs through a miR-122/FoxA1/HNF4a-positive feedback loop. Liver Int. 34(2):281-295. 151 Davoodian N, AS Lotfi, M Soleimani and SJ Mowla. (2014). MicroRNA-122 overexpression promotes hepatic differentiation of human adipose tissue-derived stem cells. J Cell Biochem. 115(9):1582-1593.

42

Page 43 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

43 152 Cui L, Y Shi, X Zhou, X Wang, J Wang, Y Lan, M Wang, L Zheng, H Li, Q Wu, J Zhang, D Fan and Y Han. (2013). A set of microRNAs mediate direct conversion of human umbilical cord lining-derived mesenchymal stem cells into hepatocytes. Cell Death Dis. 4:e918. 153 Cui L, X Zhou, J Li, L Wang, J Wang, Q Li, J Chu, L Zheng, Q Wu, Z Han, Y Shi, Y Han and D Fan. (2012). Dynamic microRNA profiles of hepatic differentiated human umbilical cord liningderived mesenchymal stem cells. PLoS One. 7(9):e44737. 154 Polishchuk EV, M Concilli, S Iacobacci, G Chesi, N Pastore, P Piccolo, S Paladino, D Baldantoni, SC van IJzendoorn, J Chan, CJ Chang, A Amoresano, F Pane, P Pucci, A Tarallo, G Parenti, N Brunetti-Pierri, C Settembre, A Ballabio and RS Polishchuk. (2014). Wilson disease protein ATP7B utilizes lysosomal exocytosis to maintain copper homeostasis. Dev Cell. 29(6):686700. 155 Sauer V, R Siaj, T Todorov, A Zibert and HH Schmidt. (2010). Overexpressed ATP7B protects mesenchymal stem cells from toxic copper. Biochem Biophys Res Commun. 395(3):307311. 156 Zhao Y, Li T, Wei X, Bianchi G, Hu J, Sanchez PG, Xu K, Zhang P, Pittenger MF, Wu ZJ and BP Griffith. (2012) Mesenchymal stem cell transplantation improves regional cardiac remodeling following ovine infarction. Stem Cells Transl Med. 1(9):685-95. 157 Li J, Zhu K, Wang Y, Zheng J, Guo C, Lai H and C Wang. (2015) Combination of IGF‑1 gene manipulation and 5‑AZA treatment promotes differentiation of mesenchymal stem cells into cardiomyocyte‑like cells. Mol Med Rep. 11(2):815-20. 158 Pan Q, X Qin, S Ma, H Wang, K Cheng, X Song, H Gao, Q Wang, R Tao, Y Wang, X Li, L Xiong and F Cao. (2014). Myocardial protective effect of extracellular superoxide dismutase gene modified bone marrow mesenchymal stromal cells on infarcted mice hearts. Theranostics. 4(5):475-486. 159 Wang D, P Luo, Y Wang, W Li, C Wang, D Sun, R Zhang, T Su, X Ma, C Zeng, H Wang, J Ren and F Cao. (2013). Glucagon-like peptide-1 protects against cardiac microvascular injury in diabetes via a cAMP/PKA/Rho-dependent mechanism. Diabetes. 62(5):1697-1708. 43

Page 44 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

44 160 Wright EJ, KA Farrell, N Malik, M Kassem, AL Lewis, C Wallrapp and CM Holt. (2012). Encapsulated glucagon-like peptide-1-producing mesenchymal stem cells have a beneficial effect on failing pig hearts. Stem Cells Transl Med. 1(10):759-769. 161 Ling L, V Nurcombe and SM Cool. (2009). Wnt signaling controls the fate of mesenchymal stem cells. Gene. 433(1-2):1-7. 162 Ueno S, G Weidinger, T Osugi, AD Kohn, JL Golob, L Pabon, H Reinecke, RT Moon and CE Murry. (2007). Biphasic role for Wnt/beta-catenin signaling in cardiac specification in zebrafish and embryonic stem cells. Proc Natl Acad Sci U S A. 104(23):9685-9690. 163 Gao Q, X Hu, X Jiang, M Guo, H Ji, Y Wang and Y Fan. (2014). Cardiomyocyte-like cells differentiation from non β-catenin expression mesenchymal stem cells. Cytotechnology. 66(4):575584. 164 He Z, H Li, S Zuo, Z Pasha, Y Wang, Y Yang, W Jiang, M Ashraf and M Xu. (2011). Transduction of Wnt11 promotes mesenchymal stem cell transdifferentiation into cardiac phenotypes. Stem Cells Dev. 20(10):1771-1778. 165 Tenhunen R, HS Marver and R Schmid. (1968). The enzymatic conversion of heme to bilirubin by microsomal heme oxygenase. Proc Natl Acad Sci U S A. 61(2):748-755. 166 Bilbija D, JA Gravning, F Haugen, H Attramadal and G Valen. (2012). Protecting the heart through delivering DNA encoding for heme oxygenase-1 into skeletal muscle. Life Sci. 91(1718):828-836. 167 Tsubokawa T, K Yagi, C Nakanishi, M Zuka, A Nohara, H Ino, N Fujino, T Konno, MA Kawashiri, H Ishibashi-Ueda, N Nagaya and M Yamagishi. (2010). Impact of anti-apoptotic and anti-oxidative effects of bone marrow mesenchymal stem cells with transient overexpression of heme oxygenase-1 on myocardial ischemia. Am J Physiol Heart Circ Physiol. 2298(5):H13201329. 168 Preda MB, T Rønningen, A Burlacu, M Simionescu, JØ Moskaug and G Valen. (2014). Remote transplantation of mesenchymal stem cells protects the heart against ischemia-reperfusion injury. Stem Cells. 32(8):2123-2134. 44

Page 45 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

45 169 Huang J, J Elicker, N Bowens, X Liu, L Cheng, TP Cappola, X Zhu and MS Parmacek. (2012). Myocardin regulates BMP10 expression and is required for heart development. J Clin Invest. 122(10):3678-3691. 170 Madonna R, DA Taylor, YJ Geng, R De Caterina, H Shelat, EC Perin and JT Willerson. (2013). Transplantation of mesenchymal cells rejuvenated by the overexpression of telomerase and myocardin promotes revascularization and tissue repair in a murine model of hindlimb ischemia. Circ Res. 113(7):902-914. 171 Peterkin T, A Gibson, M Loose and R Patient. (2005). The roles of GATA-4, -5 and -6 in vertebrate heart development. Semin Cell Dev Biol. 16(1):83-94. 172 Li H, S Zuo, Z Pasha, B Yu, Z He, Y Wang, X Yang, M Ashraf and M Xu. (2011). GATA-4 promotes myocardial transdifferentiation of mesenchymal stromal cells via up-regulating IGFBP-4. Cytotherapy. 13(9):1057-1065. 173 Li HX, YF Zhou, B Jiang, X Zhao, TB Jiang, X Li, XJ Yang and WP Jiang. (2014). GATA-4 induces changes in electrophysiological properties of rat mesenchymal stem cells. Biochim Biophys Acta. 1840(6):2060-2069. 174 Yu B, M Gong, Z He, YG Wang, RW Millard, M Ashraf and M Xu. (2013). Enhanced mesenchymal stem cell survival induced by GATA-4 overexpression is partially mediated by regulation of the miR-15 family. Int J Biochem Cell Biol. 45(12):2724-2735. 175 Yu B, M Gong, Y Wang, RW Millard, Z Pasha, Y Yang, M Ashraf and M Xu. (2013). Cardiomyocyte protection by GATA-4 gene engineered mesenchymal stem cells is partially mediated by translocation of miR-221 in microvesicles. PLoS One. 8(8):e73304. 176 Zhang CZ, JX Zhang, AL Zhang, ZD Shi, L Han, ZF Jia, WD Yang, GX Wang, T Jiang, YP You, PY Pu, JQ Cheng and CS Kang. (2010). MiR-221 and miR-222 target PUMA to induce cell survival in glioblastoma. Mol Cancer. 9:229. 177 Kapsimali M, WP Kloosterman, E de Bruijn, F Rosa, RH Plasterk and SW Wilson. (2007). MicroRNAs show a wide diversity of expression profiles in the developing and mature central nervous system. Genome Biol. 8(8):R173. 45

Page 46 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

46 178 Cai B, J Li, J Wang, X Luo, J Ai, Y Liu, N Wang, H Liang, M Zhang, N Chen, G Wang, S Xing, X Zhou, B Yang, X Wang and Y Lu. (2012). microRNA-124 regulates cardiomyocyte differentiation of bone marrow-derived mesenchymal stem cells via targeting STAT3 signaling. Stem Cells. 30(8):1746-1755. 179 Zhang LL, JJ Liu, F Liu, WH Liu, YS Wang, B Zhu and B Yu. (2012). MiR-499 induces cardiac differentiation of rat mesenchymal stem cells through wnt/β-catenin signaling pathway. Biochem Biophys Res Commun. 420(4):875-881. 180 Sluijter JP, A van Mil, P van Vliet, CH Metz, J Liu, PA Doevendans and MJ Goumans. (2010). MicroRNA-1 and -499 regulate differentiation and proliferation in human-derived cardiomyocyte progenitor cells. Arterioscler Thromb Vasc Biol. 30(4):859-868. 181 Chen JF, EM Mandel, JM Thomson, Q Wu, TE Callis, SM Hammond, FL Conlon and DZ Wang. (2006). The role of microRNA-1 and microRNA-133 in skeletal muscle proliferation and differentiation. Nat Genet. 38(2):228-233. 182 Kim HW, S Jiang, M Ashraf and KH Haider. (2012). Stem cell-based delivery of Hypoxamir210 to the infarcted heart: implications on stem cell survival and preservation of infarcted heart function. J Mol Med (Berl). 90(9):997-1010. 183 Qiu J, XY Zhou, XG Zhou, R Cheng, HY Liu and Y Li. (2013). Neuroprotective effects of microRNA-210 against oxygen-glucose deprivation through inhibition of apoptosis in PC12 cells. Mol Med Rep. 7(6):1955-1959. 184 Biel M, C Wahl-Schott, S Michalakis and X Zong. (2009). Hyperpolarization-activated cation channels: from genes to function. Physiol Rev. 89(3):847-885. 185 Moosmang S, J Stieber, X Zong, M Biel, F Hofmann and A Ludwig. (2001). Cellular expression and functional characterization of four hyperpolarization-activated pacemaker channels in cardiac and neuronal tissues. Eur J Biochem. 268(6):1646-1652. 186 Lu W, N Yaoming, R Boli, C Jun, Z Changhai, Z Yang and S Zhiyuan. (2013). mHCN4 genetically modified canine mesenchymal stem cells provide biological pacemaking function in complete dogs with atrioventricular block. Pacing Clin Electrophysiol. 36(9):1138-1149. 46

Page 47 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

47 187 Nong Y, C Zhang, L Wei, Z Zhang, J Cheng, L Wen and Z Song. (2013). In situ investigation of allografted mouse HCN4 gene-transfected rat bone marrow mesenchymal stromal cells with the use of patch-clamp recording of ventricular slices. Cytotherapy. 15(8):905-919. 188 Huang W, DZ Xiao, Y Wang, ZX Shan, XY Liu, QX Lin, M Yang, J Zhuang, Y Li and XY Yu. (2014). Fn14 promotes differentiation of human mesenchymal stem cells into heart valvular interstitial cells by phenotypic characterization. J Cell Physiol. 229(5):580-587. 189 Ren MY and SJ Sui. (2012). The role of TWEAK/Fn14 in cardiac remodeling. Mol Biol Rep. 39(11):9971-9977. 190 Lin YT, Y Chern, CK Shen, HL Wen, YC Chang, H Li, TH Cheng and HM Hsieh-Li. (2011). Human mesenchymal stem cells prolong survival and ameliorate motor deficit through trophic support in Huntington's disease mouse models. PLoS One. 6(8):e22924. 191 Crigler L, RC Robey, A Asawachaicharn, D Gaupp and DG Phinney. (2006). Human mesenchymal stem cell subpopulations express a variety of neuro-regulatory molecules and promote neuronal cell survival and neuritogenesis. Exp Neurol. 198(1):54-64. 192 Gu N, C Rao, Y Tian, Z Di, Z Liu, M Chang and H Lei. (2014). Anti-inflammatory and Antiapoptotic Effects of Mesenchymal Stem Cells Transplantation in Rat Brain with Cerebral Ischemia. J Stroke Cerebrovasc Dis. 23(10):2598-2606. 193 Huang P, N Gebhart, E Richelson, TG Brott, JF Meschia and AC Zubair. (2014). Mechanism of mesenchymal stem cell-induced neuron recovery and anti-inflammation. Cytotherapy. pii:S1465-3249(14)00605-7. 194 Zhang R, Y Liu, K Yan, L Chen, XR Chen, P Li, FF Chen and XD Jiang. (2013). Antiinflammatory and immunomodulatory mechanisms of mesenchymal stem cell transplantation in experimental traumatic brain injury. J Neuroinflammation. 10(1):106. 195 Shetty P, AM Thakur and C Viswanathan. (2013). Dopaminergic cells, derived from a high efficiency differentiation protocol from umbilical cord derived mesenchymal stem cells, alleviate symptoms in a Parkinson's disease rodent model. Cell Biol Int. 37(2):167-180.

47

Page 48 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

48 196 Boido M, A Piras, V Valsecchi, G Spigolon, K Mareschi, I Ferrero, A Vizzini, S Temi, L Mazzini, F Fagioli and A Vercelli. (2014). Human mesenchymal stromal cell transplantation modulates neuroinflammatory milieu in a mouse model of amyotrophic lateral sclerosis. Cytotherapy. 16(8):1059-1072. 197 Hedayatpour A, I Ragerdi, P Pasbakhsh, L Kafami, N Atlasi, V Pirhajati Mahabadi, S Ghasemi and M Reza. (2013). Promotion of remyelination by adipose mesenchymal stem cell transplantation in a cuprizone model of multiple sclerosis. Cell J. 15(2):142-151. 198 Yan Y, T Ma, K Gong, Q Ao, X Zhang and Y Gong. (2014). Adipose-derived mesenchymal stem cell transplantation promotes adult neurogenesis in the brains of Alzheimer's disease mice. Neural Regen Res. 9(8):798-805. 199 Fink KD, J Rossignol, AT Crane, KK Davis, MC Bombard, AM Bavar, S Clerc, SA Lowrance, C Song, L Lescaudron and GL Dunbar. (2013). Transplantation of umbilical cordderived mesenchymal stem cells into the striata of R6/2 mice: behavioral and neuropathological analysis. Stem Cell Res Ther. 4(5):130. 200 Oliveri RS, S Bello and F Biering-Sørensen. (2014). Mesenchymal stem cells improve locomotor recovery in traumatic spinal cord injury: systematic review with meta-analyses of rat models. Neurobiol Dis. 62:338-353. 201 Wang Z, W Yao, Q Deng, X Zhang and J Zhang. (2013). Protective effects of BDNF overexpression bone marrow stromal cell transplantation in rat models of traumatic brain injury. J Mol Neurosci. 49(2):409-416. 202 Lipsky RH and AM Marini. (2007). Brain-derived neurotrophic factor in neuronal survival and behavior-related plasticity. Ann N Y Acad Sci. 1122:130-143. 203 Tseng TC and SH Hsu. (2014). Substrate-mediated nanoparticle/gene delivery to MSC spheroids and their applications in peripheral nerve regeneration. Biomaterials. 35(9):2630-2641. 204 Dey ND, MC Bombard, BP Roland, S Davidson, M Lu, J Rossignol, MI Sandstrom, RL Skeel, L Lescaudron and GL Dunbar. (2010). Genetically engineered mesenchymal stem cells

48

Page 49 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

49 reduce behavioral deficits in the YAC 128 mouse model of Huntington's disease. Behav Brain Res. 214(2):193-200. 205 Yu X, D Chen, Y Zhang, X Wu, Z Huang, H Zhou, Y Zhang and Z Zhang. (2012). Overexpression of CXCR4 in mesenchymal stem cells promotes migration, neuroprotection and angiogenesis in a rat model of stroke. J Neurol Sci. 316(1-2):141-149. 206 Airaksinen MS and M Saarma. (2002). The GDNF family: signalling, biological functions and therapeutic value. Nat Rev Neurosci. 3(5):383-394. 207 Yang C, L Zhou, X Gao, B Chen, J Tu, H Sun, X Liu, J He, J Liu and Q Yuan. (2011). Neuroprotective effects of bone marrow stem cells overexpressing glial cell line-derived neurotrophic factor on rats with intracerebral hemorrhage and neurons exposed to hypoxia/reoxygenation. Neurosurgery. 68(3):691-704. 208 Gao H, M Wei, Y Wang, X Wu and T Zhu. (2012). Differentiation of GDNF and NT-3 dual gene-modified rat bone marrow mesenchymal stem cells into enteric neuron-like cells. J Huazhong Univ Sci Technolog Med Sci. 32(1):87-91. 209 Ross SE, ME Greenberg and CD Stiles. (2003). Basic helix-loop-helix factors in cortical development. Neuron. 39(1):13-25. 210 Cheng F, XC Lu, HY Hao, XL Dai, TD Qian, BS Huang, LJ Tang, W Yu and LX Li. (2014). Neurogenin 2 converts mesenchymal stem cells into a neural precursor fate and improves functional recovery after experimental stroke. Cell Physiol Biochem. 33(3):847-858. 211 Chan-Il C, L Young-Don, K Heejaung, SH Kim, H Suh-Kim and SS Kim. (2013). Neural induction with neurogenin 1 enhances the therapeutic potential of mesenchymal stem cells in an amyotrophic lateral sclerosis mouse model. Cell Transplant. 22(5):855-870. 212 Xin H, Y Li, B Buller, M Katakowski, Y Zhang, X Wang, X Shang, ZG Zhang and M Chopp. (2012). Exosome-mediated transfer of miR-133b from multipotent mesenchymal stromal cells to neural cells contributes to neurite outgrowth. Stem Cells. 30(7):1556-1564. 213 Xin H, Y Li, Z Liu, X Wang, X Shang, Y Cui, ZG Zhang and M Chopp. (2013). MiR-133b promotes neural plasticity and functional recovery after treatment of stroke with multipotent 49

Page 50 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

50 mesenchymal stromal cells in rats via transfer of exosome-enriched extracellular particles. Stem Cells. 31(12):2737-2746. 214 Perry T, NJ Haughey, MP Mattson, JM Egan and NH Greig. (2002). Protection and reversal of excitotoxic neuronal damage by glucagon-like peptide-1 and exendin-4. J Pharmacol Exp Ther. 302(3):881-888. 215 Klinge PM, K Harmening, MC Miller, A Heile, C Wallrapp, P Geigle and T Brinker. (2011). Encapsulated native and glucagon-like peptide-1 transfected human mesenchymal stem cells in a transgenic mouse model of Alzheimer's disease. Neurosci Lett. 497(1):6-10. 216 Glavaski-Joksimovic A, T Virag, TA Mangatu, M McGrogan, XS Wang and MC Bohn. (2010). Glial cell line-derived neurotrophic factor-secreting genetically modified human bone marrow-derived mesenchymal stem cells promote recovery in a rat model of Parkinson's disease. J Neurosci Res. 88(12):2669-2681. 217 Shi D, G Chen, L Lv, L Li, D Wei, P Gu, J Gao, Y Miao and W Hu. (2011). The effect of lentivirus-mediated TH and GDNF genetic engineering mesenchymal stem cells on Parkinson's disease rat model. Neurol Sci. 32(1):41-51. 218 Yin X, H Xu, Y Jiang, W Deng, Z Wu, H Xiang, P Sun and J Xie. (2014). The effect of lentivirus-mediated PSPN genetic engineering bone marrow mesenchymal stem cells on Parkinson's disease rat model. PLoS One. 9(8):e105118. 219 Sidorova YA, K Mätlik, M Paveliev, M Lindahl, E Piranen, J Milbrandt, U Arumäe, M Saarma and MM Bespalov. (2010). Persephin signaling through GFRalpha1: the potential for the treatment of Parkinson's disease. Mol Cell Neurosci. 44(3):223-232. 220 Høglund RA and AA Maghazachi. (2014). Multiple sclerosis and the role of immune cells. World J Exp Med. 4(3):27-37. 221 Cobo M, P Anderson, K Benabdellah, MG Toscano, P Muñoz, A García-Pérez, I Gutierrez, M Delgado and F Martin. (2013). Mesenchymal stem cells expressing vasoactive intestinal peptide ameliorate symptoms in a model of chronic multiple sclerosis. Cell Transplant. 22(5):839-854.

50

Page 51 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

51 222 Payne NL, G Sun, C McDonald, L Moussa, A Emerson-Webber, S Loisel-Meyer, JA Medin, C Siatskas and CC Bernard. (2013). Human adipose-derived mesenchymal stem cells engineered to secrete IL-10 inhibit APC function and limit CNS autoimmunity. Brain Behav Immun. 30:103-114. 223 Ryu CH, KY Park, Y Hou, CH Jeong, SM Kim and SS Jeun. (2013). Gene therapy of multiple sclerosis using interferon β-secreting human bone marrow mesenchymal stem cells. Biomed Res Int. 2013:696738. 224 Beutner C, V Lepperhof, A Dann, B Linnartz-Gerlach, S Litwak, I Napoli, M Prinz and H Neumann. (2013). Engineered stem cell-derived microglia as therapeutic vehicle for experimental autoimmune encephalomyelitis. Gene Ther. 20(8):797-806. 225 Zhang W, Q Yan, YS Zeng, XB Zhang, Y Xiong, JM Wang, SJ Chen, Y Li, IC Bruce and W Wu. (2010). Implantation of adult bone marrow-derived mesenchymal stem cells transfected with the neurotrophin-3 gene and pretreated with retinoic acid in completely transected spinal cord. Brain Res. 1359:256-271. 226 Zhang YJ, W Zhang, CG Lin, Y Ding, SF Huang, JL Wu, Y Li, H Dong and YS Zeng. (2012). Neurotrophin-3 gene modified mesenchymal stem cells promote remyelination and functional recovery in the demyelinated spinal cord of rats. J Neurol Sci. 313(1-2):64-74. 227 Uren RT and AM Turnley. (2014). Regulation of neurotrophin receptor (Trk) signaling: suppressor of cytokine signaling 2 (SOCS2) is a new player. Front Mol Neurosci. 7:39. 228 Ding Y, Q Yan, JW Ruan, YQ Zhang, WJ Li, X Zeng, SF Huang, YJ Zhang, JL Wu, D Fisher, H Dong and YS Zeng. (2013). Electroacupuncture promotes the differentiation of transplanted bone marrow mesenchymal stem cells overexpressing TrkC into neuron-like cells in transected spinal cord of rats. Cell Transplant. 22(1):65-86. 229 Urfer R, P Tsoulfas, D Soppet, E Escandón, LF Parada and LG Presta. (1994). The binding epitopes of neurotrophin-3 to its receptors trkC and gp75 and the design of a multifunctional human neurotrophin. EMBO J. 13(24):5896-5909.

51

Page 52 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

52 230 Kumagai G, P Tsoulfas, S Toh, I McNiece, HM Bramlett and WD Dietrich. (2013). Genetically modified mesenchymal stem cells (MSCs) promote axonal regeneration and prevent hypersensitivity after spinal cord injury. Exp Neurol. 248:369-380. 231 Fuccillo M, AL Joyner and G Fishell. (2006). Morphogen to mitogen: the multiple roles of hedgehog signalling in vertebrate neural development. Nat Rev Neurosci. 7(10):772-783. 232 Jia Y, D Wu, R Zhang, W Shuang, J Sun, H Hao, Q An and Q Liu. (2014). Bone marrowderived mesenchymal stem cells expressing the Shh transgene promotes functional recovery after spinal cord injury in rats. Neurosci Lett. 573:46-51. 233 Long Q, B Qiu, K Wang, J Yang, C Jia, W Xin, P Wang, R Han, Z Fei and W Liu. (2013). Genetically engineered bone marrow mesenchymal stem cells improve functional outcome in a rat model of epilepsy. Brain Res. 1532:1-13. 234 Lu Z, X Hu, C Zhu, D Wang, X Zheng and Q Liu. (2009). Overexpression of CNTF in Mesenchymal Stem Cells reduces demyelination and induces clinical recovery in experimental autoimmune encephalomyelitis mice. J Neuroimmunol. 206(1-2):58-69. 235 Wang F, G Chang and X Geng. (2014). NGF and TERT co-transfected BMSCs improve the restoration of cognitive impairment in vascular dementia rats. PLoS One. 9(6):e98774. 236 Yuan J, G Huang, Z Xiao, L Lin and T Han. (2013). Overexpression of β-NGF promotes differentiation of bone marrow mesenchymal stem cells into neurons through regulation of AKT and MAPK pathway. Mol Cell Biochem. 383(1-2):201-211. 237 Campos LS, L Decker, V Taylor and W Skarnes. (2006). Notch, epidermal growth factor receptor, and beta1-integrin pathways are coordinated in neural stem cells. J Biol Chem. 281(8):5300-5309. 238 Li Y, WM Lau, KF So, Y Tong and J Shen. (2011). Caveolin-1 promote astroglial differentiation of neural stem/progenitor cells through modulating Notch1/NICD and Hes1 expressions. Biochem Biophys Res Commun. 407(3):517-524.

52

Page 53 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

53 239 Wang S, Q Kan, Y Sun, R Han, G Zhang, T Peng and Y Jia. (2013). Caveolin-1 regulates neural differentiation of rat bone mesenchymal stem cells into neurons by modulating Notch signaling. Int J Dev Neurosci. 31(1):30-35. 240 Siegenthaler JA, AM Ashique, K Zarbalis, KP Patterson, JH Hecht, MA Kane, AE Folias, Y Choe, SR May, T Kume, JL Napoli, AS Peterson and SJ Pleasure. (2009). Retinoic acid from the meninges regulates cortical neuron generation. Cell. 139(3):597-609. 241 Bi Y, M Gong, Y He, X Wei, J Chen and T Li. (2013). Adenovirus-mediated RAR-β overexpression enhances ATRA-induced neuronal differentiation of rat mesenchymal stem cells. Arch Med Sci. 9(2):314-322. 242 Kadkhodaei B, A Alvarsson, N Schintu, D Ramsköld, N Volakakis, E Joodmardi, T Yoshitake, J Kehr, M Decressac, A Björklund, R Sandberg, P Svenningsson and T Perlmann. (2013). Transcription factor Nurr1 maintains fiber integrity and nuclear-encoded mitochondrial gene expression in dopamine neurons. Proc Natl Acad Sci U S A. 110(6):2360-2365. 243 Park JS, HN Yang, DG Woo, SY Jeon, HJ Do, SH Huh, NH Kim, JH Kim and KH Park. (2012). Exogenous Nurr1 gene expression in electrically-stimulated human MSCs and the induction of neurogenesis. Biomaterials. 33(29):7300-7308. 244 Wang K, Q Long, C Jia, Y Liu, X Yi, H Yang, Z Fei and W Liu. (2013). Over-expression of Mash1 improves the GABAergic differentiation of bone marrow mesenchymal stem cells in vitro. Brain Res Bull. 99:84-94. 245 Han R, Q Kan, Y Sun, S Wang, G Zhang, T Peng and Y Jia. (2012). MiR-9 promotes the neural differentiation of mouse bone marrow mesenchymal stem cells via targeting zinc finger protein 521. Neurosci Lett. 515(2):147-152. 246 Wu R, N Wang, M Li, W Zang and Y Xu. (2013). Experimental study on the facilitative effects of miR-125b on the differentiation of rat bone marrow mesenchymal stem cells into neuronlike cells. Cell Biol Int. 37(8):812-819.

53

Page 54 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

54 247 Song YS, HJ Lee, SH Doo, SJ Lee, I Lim, KT Chang and SU Kim. (2012). Mesenchymal stem cells overexpressing hepatocyte growth factor (HGF) inhibit collagen deposit and improve bladder function in rat model of bladder outlet obstruction. Cell Transplant. 21(8):1641-1650. 248 Yu Y, L Lu, X Qian, N Chen, A Yao, L Pu, F Zhang, X Li, L Kong, B Sun and X Wang. (2010). Antifibrotic effect of hepatocyte growth factor-expressing mesenchymal stem cells in small-for-size liver transplant rats. Stem Cells Dev. 19(6):903-914. 249 Seo KW, SY Sohn, DH Bhang, MJ Nam, HW Lee and HY Youn. (2014). Therapeutic effects of hepatocyte growth factor-overexpressing human umbilical cord blood-derived mesenchymal stem cells on liver fibrosis in rats. Cell Biol Int. 38(1):106-116. 250 Kim MD, SS Kim, HY Cha, SH Jang, DY Chang, W Kim, H Suh-Kim and JH Lee. (2014). Therapeutic effect of hepatocyte growth factor-secreting mesenchymal stem cells in a rat model of liver fibrosis. Exp Mol Med. 46:e110. 251 Lin Z, P Perez, Z Sun, JJ Liu, JH Shin, KL Hyrc, D Samways, T Egan, MC Holley and J Bao. (2012). Reprogramming of single-cell-derived mesenchymal stem cells into hair cell-like cells. Otol Neurotol. 33(9):1648-1655. 252 Wei X, Z Mao, Y Hou, L Lin, T Xue, L Chen, H Wang and C Yu. (2011). Local administration of TGFβ-1/VEGF165 gene-transduced bone mesenchymal stem cells for Achilles allograft replacement of the anterior cruciate ligament in rabbits. Biochem Biophys Res Commun. 406(2):204-210. 253 Chen J, C Li, X Gao, C Li, Z Liang, L Yu, Y Li, X Xiao and L Chen. (2013). Keratinocyte growth factor gene delivery via mesenchymal stem cells protects against lipopolysaccharideinduced acute lung injury in mice. PLoS One. 8(12):e83303. 254 Machalińska A, M Kawa, E Pius-Sadowska, J Stępniewski, W Nowak, D Rogińska, K Kaczyńska, B Baumert, B Wiszniewska, A Józkowicz, J Dulak and B Machaliński. (2013). Longterm neuroprotective effects of NT-4-engineered mesenchymal stem cells injected intravitreally in a mouse model of acute retinal injury. Invest Ophthalmol Vis Sci. 54(13):8292-8305.

54

Page 55 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

55 255 Hu J, Z Yang, J Wang, Y Tang, H Liu, B Zhang and H Chen. (2013). Infusion of Trx-1overexpressing hucMSC prolongs the survival of acutely irradiated NOD/SCID mice by decreasing excessive inflammatory injury. PLoS One. 8(11):e78227. 256 Yan G, Y Fan, P Li, Y Zhang and F Wang. (2015). Ectopic expression of DAZL gene in goat bone marrow-derived mesenchymal stem cells enhances the trans-differentiation to putative germ cells compared to the exogenous treatment of retinoic acid or bone morphogenetic protein 4 signalling molecules. Cell Biol Int. 39(1):74-83. 257 Paz AH, GD Salton, A Ayala-Lugo, C Gomes, P Terraciano, R Scalco, CC Laurino, EP Passos, MR Schneider, L Meurer and E Cirne-Lima. (2011). Betacellulin overexpression in mesenchymal stem cells induces insulin secretion in vitro and ameliorates streptozotocin-induced hyperglycemia in rats. Stem Cells Dev. 20(2):223-232. 258 Yuan H, H Liu, R Tian, J Li and Z Zhao. (2012). Regulation of mesenchymal stem cell differentiation and insulin secretion by differential expression of Pdx-1. Mol Biol Rep. 39(7):77777783. 259 Van Pham P, P Thi-My Nguyen, A Thai-Quynh Nguyen, V Minh Pham, A Nguyen-Tu Bui, L Thi-Tung Dang, K Gia Nguyen and N Kim Phan. (2014). Improved differentiation of umbilical cord blood-derived mesenchymal stem cells into insulin-producing cells by PDX-1 mRNA transfection. Differentiation. 87(5):200-208. 260 Guo QS, MY Zhu, L Wang, XJ Fan, YH Lu, ZW Wang, SJ Zhu, Y Wang and Y Huang. (2012). Combined transfection of the three transcriptional factors, PDX-1, NeuroD1, and MafA, causes differentiation of bone marrow mesenchymal stem cells into insulin-producing cells. Exp Diabetes Res. 2012:672013. 261 Wang H, Y Yang, G Ho, X Lin, W Wu, W Li, L Lin, X Feng, X Huo, J Jiang, X Liu, T Huang, C Wei and L Ma. (2013). Programming of human umbilical cord mesenchymal stem cells in vitro to promote pancreatic gene expression. Mol Med Rep. 8(3):769-774.

55

Page 56 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

56 262 Shaer A, N Azarpira, A Vahdati, MH Karimi and M Shariati. (2014). miR-375 induces human decidua basalis-derived stromal cells to become insulin-producing cells. Cell Mol Biol Lett. 19(3):483-499. 263 Wu C, F Liu, P Li, G Zhao, S Lan, W Jiang, X Meng, L Tian, G Li, Y Li and JY Liu. (2014). Engineered Hair Follicle Mesenchymal Stem Cells Overexpressing Controlled-Release Insulin Reverse Hyperglycemia in Mice with Type l Diabetes. Cell Transplant. 2014 May 15. 264 Mundra V, H Wu and RI Mahato. (2013). Genetically modified human bone marrow derived mesenchymal stem cells for improving the outcome of human islet transplantation. PLoS One. 8(10):e77591. 265 Roux C, DF Pisani, HB Yahia, M Djedaini, GE Beranger, JC Chambard, D Ambrosetti, JF Michiels, V Breuil, G Ailhaud, L Euller-Ziegler and EZ Amri. (2013). Chondrogenic potential of stem cells derived from adipose tissue: a powerful pharmacological tool. Biochem Biophys Res Commun. 440(4):786-791. 266 Glass KA, JM Link, JM Brunger, FT Moutos, CA Gersbach and F Guilak. (2014). Tissueengineered cartilage with inducible and tunable immunomodulatory properties. Biomaterials. 35(22):5921-5931. 267 Li Y, M Yan, J Yang, I Raman, Y Du, S Min, X Fang, C Mohan and QZ Li. (2014). Glutathione S-transferase Mu 2-transduced mesenchymal stem cells ameliorated anti-glomerular basement

membrane

antibody-induced

glomerulonephritis

by

inhibiting

oxidation

and

inflammation. Stem Cell Res Ther. 5(1):19. 268 Yuzeng Q, H Weiyang, G Xin, Z Qingson, K Youlin and R Ke. (2014). Effects of transplantation with marrow-derived mesenchymal stem cells modified with survivin on renal ischemia-reperfusion injury in mice. Yonsei Med J. 55(4):1130-1137. 269 Jung WS, SM Han, SM Kim, ME Kim, JS Lee, KW Seo, HY Youn and HW Lee. (2014). Stimulatory effect of HGF-overexpressing adipose tissue-derived mesenchymal stem cells on thymus regeneration in a rat thymus involution model. Cell Biol Int. 38(10):1106-1117.

56

Page 57 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

57 270 Niu J, W Yue, Y Song, Y Zhang, X Qi, Z Wang, B Liu, H Shen and X Hu. (2014). Prevention of acute liver allograft rejection by IL-10-engineered mesenchymal stem cells. Clin Exp Immunol. 176(3):473-484. 271 Tan CQ, X Gao, L Guo and H Huang. (2014). Exogenous IL-4-expressing bone marrow mesenchymal stem cells for the treatment of autoimmune sensorineural hearing loss in a guinea pig model. Biomed Res Int. 2014:856019. 272 Miettinen JA, M Pietilä, RJ Salonen, S Ohlmeier, K Ylitalo, HV Huikuri and P Lehenkari. (2011). Tumor necrosis factor alpha promotes the expression of immunosuppressive proteins and enhances the cell growth in a human bone marrow-derived stem cell culture. Exp Cell Res. 317(6):791-801. 273 Salter B and C Salter. (2013). Bioethical ambition, political opportunity and the European governance of patenting: the case of human embryonic stem cell science. Soc Sci Med. 98:286292. 274 Bokhoven M, SL Stephen, S Knight, EF Gevers, IC Robinson, Y Takeuchi and MK Collins. (2009). Insertional gene activation by lentiviral and gammaretroviral vectors. J Virol. 83(1):283294. 275 Takeuchi M, K Takeuchi, A Kohara, M Satoh, S Shioda, Y Ozawa, A Ohtani, K Morita, T Hirano, M Terai, A Umezawa and H Mizusawa. (2007). Chromosomal instability in human mesenchymal stem cells immortalized with human papilloma virus E6, E7, and hTERT genes. In Vitro Cell Dev Biol Anim. 43(3-4):129-138. 276 Modlich U, S Navarro, D Zychlinski, T Maetzig, S Knoess, MH Brugman, A Schambach, S Charrier, A Galy, AJ Thrasher, J Bueren and C Baum. (2009). Insertional transformation of hematopoietic cells by self-inactivating lentiviral and gammaretroviral vectors. Mol Ther. 17(11):1919-1928. 277 Aronovich EL, RS McIvor and PB Hackett. (2011). The Sleeping Beauty transposon system: a non-viral vector for gene therapy. Hum Mol Genet. 20(R1):R14-20.

57

Page 58 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

58 278 Tolar J, AJ Nauta, MJ Osborn, A Panoskaltsis Mortari, RT McElmurry, S Bell, L Xia, N Zhou, M Riddle, TM Schroeder, JJ Westendorf, RS McIvor, PC Hogendoorn, K Szuhai, L Oseth, B Hirsch, SR Yant, MA Kay, A Peister, DJ Prockop, WE Fibbe and BR Blazar. (2007). Sarcoma derived from cultured mesenchymal stem cells. Stem Cells. 25(2):371-379. 279 Ma K, DD Wang, Y Lin, J Wang, V Petrenko and C Mao. (2013). Synergetic Targeted Delivery of Sleeping-Beauty Transposon System to Mesenchymal Stem Cells Using LPD Nanoparticles Modified with a Phage-Displayed Targeting Peptide. Adv Funct Mater. 23(9):11721181. 280 de Jong J, W Akhtar, J Badhai, AG Rust, R Rad, J Hilkens, A Berns, M van Lohuizen, LF Wessels and J de Ridder. (2014). Chromatin landscapes of retroviral and transposon integration profiles. PLoS Genet. 10(4):e1004250. 281 Benabdallah BF, E Allard, S Yao, G Friedman, PD Gregory, N Eliopoulos, J Fradette, JL Spees, E Haddad, MC Holmes and CM Beauséjour. (2010) Targeted gene addition to human mesenchymal stromal cells as a cell-based plasma-soluble protein delivery platform. Cytotherapy. 12(3):394-9. 282 Valton J, JP Cabaniols, R Galetto, F Delacote, M Duhamel, S Paris, DA Blanchard, C Lebuhotel, S Thomas, S Moriceau, R Demirdjian, G Letort, A Jacquet, A Gariboldi, S Rolland, F Daboussi, A Juillerat, C Bertonati, A Duclert and P Duchateau. (2014) Efficient strategies for TALEN-mediated genome editing in mammalian cell lines. Methods. 69(2):151-70. 283 Liang X, J Potter, S Kumar, Y Zou, R Quintanilla, M Sridharan, J Carte, W Chen, N Roark, S Ranganathan, N Ravinder and JD Chesnut. (2015) Rapid and highly efficient mammalian cell engineering via Cas9 protein transfection. J Biotechnol. pii: S0168-1656(15)00200-X. [Epub ahead of print] 284 Madeira C, SC Ribeiro, IS Pinheiro, SA Martins, PZ Andrade, CL da Silva and JM Cabral. (2011). Gene delivery to human bone marrow mesenchymal stem cells by microporation. J Biotechnol. 151(1):130-136.

58

Page 59 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

59 285 Abdul Halim NS, KS Fakiruddin, SA Ali and BH Yahaya. (2014). A comparative study of non-viral gene delivery techniques to human adipose-derived mesenchymal stem cell. Int J Mol Sci. 15(9):15044-15060. 286 King WJ, NA Kouris, S Choi, BM Ogle and WL Murphy. (2012). Environmental parameters influence non-viral transfection of human mesenchymal stem cells for tissue engineering applications. Cell Tissue Res. 347(3):689-699. 287 Deng W, M Fu, Y Cao, X Cao, M Wang, Y Yang, R Qu, J Li, X Xu and J Yu. (2013). Angelica sinensis polysaccharide nanoparticles as novel non-viral carriers for gene delivery to mesenchymal stem cells. Nanomedicine. 9(8):1181-1191. 288 Li P, Y Gao, Z Liu, K Tan, Z Zuo, H Xia, D Yang, Y Zhang and D Lu. (2013). DNA transfection of bone marrow stromal cells using microbubble-mediated ultrasound and polyethylenimine: an in vitro study. Cell Biochem Biophys. 66(3):775-786. 289 Wang XL, P Hu, XR Guo, D Yan, Y Yuan, SR Yan and DS Li. (2014). Reprogramming human umbilical cord mesenchymal stromal cells to islet-like cells with the use of in vitrosynthesized pancreatic-duodenal homebox 1 messenger RNA. Cytotherapy. 16(11):1519-1527. 290 Ancans J. (2012). Cell therapy medicinal product regulatory framework in Europe and its application for MSC-based therapy development. Front Immunol. 3:253. 291 Blommestein HM, SG Verelst, PC Huijgens, NM Blijlevens, JJ Cornelissen and CA Uyl-de Groot. (2012). Real-world costs of autologous and allogeneic stem cell transplantations for haematological diseases: a multicentre study. Ann Hematol. 91(12):1945-1952. 292 Tan TE, GS Peh, BL George, HY Cajucom-Uy, D Dong, EA Finkelstein and JS Mehta. (2014). A cost-minimization analysis of tissue-engineered constructs for corneal endothelial transplantation. PLoS One. 9(6):e100563. 293 Sherman W, C Mazouz, R Deans and AN Patel. (2011). Commercialization of trials for peripheral artery disease. Cytotherapy. 13(10):1157-1161.

59

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

Page 60 of 68

60

294 Matthes-Martin S, U Pötschger, R Barr, M Martin, H Boztug, T Klingebiel, A Attarbaschi, W

Eibler and G Mann. (2012). Costs and cost-effectiveness of allogeneic stem cell transplantation in

children are predictable. Biol Blood Marrow Transplant. 18(10):1533-1539.

Figure legends:

60

Page 61 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

61

Fig. 1 Differentiation potential of engineered MSCs in vitro. Although MSCs have the natural ability to differentiate into osteocytes and chondrocytes, genetic engineering methods may intensify this natural behaviour of MSCs. In addition, genetic engineering techniques have been used in MSC differentiation in vitro into many cell types, such as hepatocytes, cardiomyocytes, pacemaking cells, endothelial cells, pancreatic cells, and neurons. 61

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

Page 62 of 68

62

62

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof. Page 63 of 68

63

Fig. 2 Potential therapeutic applications of engineered MSCs in vivo.

63

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

Page 64 of 68

64

The genetic engineering of MSCs converts MSCs into a multifunctional therapeutic tool that may be

used in numerous applications in vivo, such as in different organ repair, anti-fibrotic actions and anti-

inflammatory purposes.

64

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof. Page 65 of 68

65

Fig. 3 Diversity of transgene delivery routes and operation modes of modified MSCs with therapeutic

purposes.

65

Page 66 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

66 A – Transgenes can be delivered to MSCs in multiple ways, such as by viral transduction, plasmid DNA transfection, or direct delivery of mRNA. B – The therapeutic properties of MSCs can be accessed in several ways. Foremost, transgene expression is induced in order to obtain therapeutic protein synthesis, which, in turn, can be exported from MSCs to interfere with target cells in a paracrine and / or a microvesicle manner. Therapeutic proteins can also operate inside engineered MSCs by acting on inner metabolic pathways or affecting native gene expression, which may result in the differentiation of MSCs. MSCs can be modified to overexpress specific miRNAs, which modulate native gene expression, leading to therapeutic changes in MSCs, or, when exported in microvesicles, regulate gene expression in target cells. In addition, MSCs can be used as carriers for therapeutic substances that are endocytosed and then secreted to act on target cells.

Tab.1 Summary of genetic engineering approaches of MSCs depending upon genetic modification and field of application.

APPLICATIONS

ANGIOGENESIS STIMULATION APPROACHES

ENDOTHELIAL CELL DIFFERENTIATION

GENETIC ENGINEERING APPROACHES SECRETED PROTEINS

MEMBRANE BOUND PROTEINS

CYTOPLASM ACTING PROTEINS

NU PR

VEGF [11-13][15,16][42,43] GLP-1 [44] Ang-1 [46][48] Ang-1/BMP-2 [47] GM-CSF [49] HGF [50] bFGF [51] TIMP-3 [52]

Notch homolog 1 [53]

HO-1 [54]

GA HI Isl

ACE2 [61]

SCD-1 [65]

66

Page 67 of 68

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

67

BLOOD PRODUCTION

CARTILAGE PRODUCTION

INTERVERTEBRAL DISC REGENERATION

BONE FORMATION

Factor IX [66] Epo [68,69]

Jagged-1 [71]

HI

TGF-β1 [78,79,80] TGF-β3 [81,82] CDMP1 [83] IGF-1/BMP-2 [85] IGF-1/FGF-2 [90] BMP-7 [88]

Wnt11 [91]

So So ZN

NELL1 [134] β-catenin [135]

HI Cb TA Fo Na

GDF-5 [102-104] TIMP-1 [106] Bcl-2 [101]

BMP-2 [86][109-111] BMP-2/SDF-1β [113] BMP-2/HIF-1α [114] BMP-2/Ang-1 [47] BMP-2/VEGF [116-118] BMP-2/miR-148b [142] BMP-4/-6/-7 [125,126] VEGF [119, 120] Epo [121] IGF-I [128,129] OGP [127] Leptin [130] PDGF [131]

ephrinB2 [132] ITGA2 [133]

LIVER REGENERATION

VEGF [144]

ATP7B [155]

CARDIAC-TARGETED THERAPIES

ecSOD [158] GLP-1 [160]

HCN1 [19] HCN4 [186,187] Fn14 [189]

Wnt11 [164] HO-1 [167,168]

My GA

BDNF [22][201][203,204] GDNF [20][207][216,217] NGF [236] NT-3 [225,226] MNTS-1 [230] Shh [232] VEGF [21] GLP-1 [215] VIP [221] IL-10 [222] interferon-β [223]

CXCR4 [205]

TrkC [228]

Ng Ng RA Nu Ma

HGF [23]

GSTM1 [267] Survivin [268]

NEURAL DIFFERENTIATION AND REPAIR

KIDNEY REPAIR

ANTI-INFLAMMATORY PROPERTIES

HN HN

IL-4 [271] IL-10 [24][222][270]

67

Stem Cells and Development Genetic engineering of mesenchymal stem cells for regenerative medicine (doi: 10.1089/scd.2015.0062) This article has been peer-reviewed and accepted for publication, but has yet to undergo copyediting and proof correction. The final published version may differ from this proof.

Page 68 of 68

68

68

Genetic Engineering of Mesenchymal Stem Cells for Regenerative Medicine.

Mesenchymal stem cells (MSCs), which can be obtained from various organs and easily propagated in vitro, are one of the most extensively used types of...
1MB Sizes 1 Downloads 13 Views