ArticlePDF AvailableLiterature Review

Mesenchymal Stem Cells for Regenerative Medicine

Authors:

Abstract and Figures

In recent decades, the biomedical applications of mesenchymal stem cells (MSCs) have attracted increasing attention. MSCs are easily extracted from the bone marrow, fat, and synovium, and differentiate into various cell lineages according to the requirements of specific biomedical applications. As MSCs do not express significant histocompatibility complexes and immune stimulating molecules, they are not detected by immune surveillance and do not lead to graft rejection after transplantation. These properties make them competent biomedical candidates, especially in tissue engineering. We present a brief overview of MSC extraction methods and subsequent potential for differentiation, and a comprehensive overview of their preclinical and clinical applications in regenerative medicine, and discuss future challenges.
Content may be subject to copyright.
cells
Review
Mesenchymal Stem Cells for Regenerative Medicine
Yu Han 1,2,, Xuezhou Li 1 ,2 , , Yanbo Zhang 3,* , Yuping Han 4,*, Fei Chang 1,* and
Jianxun Ding 2
1Department of Orthopedics, The Second Hospital of Jilin University, 218 Ziqiang Street,
Changchun 130041, China
2Key Laboratory of Polymer Ecomaterials, Changchun Institute of Applied Chemistry,
Chinese Academy of Sciences, 5625 Renmin Street, Changchun 130022, China
3Department of Orthopedics, China-Japan Union Hospital of Jilin University, 126 Xiantai Street,
Changchun 130033, China
4Department of Urology, China-Japan Union Hospital of Jilin University, 126 Xiantai Street,
Changchun 130033, China
*Correspondence: zhangyb2012@jlu.edu.cn (Y.Z.); hyp181818@126.com (Y.H.); ccfei_cn@hotmail.com (F.C.)
These authors contributed equally to this work.
Received: 12 July 2019; Accepted: 6 August 2019; Published: 13 August 2019


Abstract:
In recent decades, the biomedical applications of mesenchymal stem cells (MSCs) have
attracted increasing attention. MSCs are easily extracted from the bone marrow, fat, and synovium,
and dierentiate into various cell lineages according to the requirements of specific biomedical
applications. As MSCs do not express significant histocompatibility complexes and immune
stimulating molecules, they are not detected by immune surveillance and do not lead to graft rejection
after transplantation. These properties make them competent biomedical candidates, especially in
tissue engineering. We present a brief overview of MSC extraction methods and subsequent potential
for dierentiation, and a comprehensive overview of their preclinical and clinical applications in
regenerative medicine, and discuss future challenges.
Keywords: mesenchymal stem cell; extraction; cell dierentiation; biomedical application
1. Introduction
Since the discovery of spindle-shaped, bone marrow-derived plastic-adherent cells in the
mid-1970s [
1
], science has come a long way, and studies have found that these cells could
dierentiate into osteoblasts and chondrocytes [
2
,
3
]. Techniques for extraction, culture, and induction
of mesenchymal stem cells (MSCs) have improved, with almost all MSC types derived from
various tissues now capable of dierentiation into osteocytes and end-stage lineages [
4
]. The rapid
development of molecular biology and transplantation techniques has benefitted MSC applications in
regenerative medicine.
MSCs are an ideal cell source for tissue regeneration, owing to the excellent properties as follows.
MSCs exist in almost all tissues, including bone marrow, adipose, and synovium [
5
], and are easily
extracted. MSCs can dierentiate into almost any end-stage lineage cells to enable their seeding
in specific scaolds (Figure 1) [
6
]. Their immunological properties, including anti-inflammatory,
immunoregulatory, and immunosuppressive capacities, contribute to their potential role as immune
tolerant agents [7,8].
Numerous studies have explored MSCs for tissue regeneration in several animal models
in vitro
;
trials have not been limited to preclinical validation. Several clinical reports verify the potential ecacy
of MSC-based cell therapy; although its eectiveness remains limited, the outcomes are inspiring.
We present a brief overview of MSC extraction methods and subsequent potential for dierentiation and
Cells 2019,8, 886; doi:10.3390/cells8080886 www.mdpi.com/journal/cells
Cells 2019,8, 886 2 of 32
provide a comprehensive overview of future applications of various MSCs in regenerative medicine,
as well as the challenges.
Cells 2019, 8, x 2 of 36
outcomes are inspiring. We present a brief overview of MSC extraction methods and subsequent
potential for differentiation and provide a comprehensive overview of future applications of
various MSCs in regenerative medicine, as well as the challenges.
Figure 1. Schematic diagram of regenerative medicine based on mesenchymal stem cells (MSCs).
The MSCs can be easily extracted from varies tissues, and the multilineage differentiation and
immunoregulatory properties of MSCs make them an ideal cell therapeutic candidate.
2. Discovery and Extraction of MSCs from Different Sources
The rich source of MSCs is the critical basis for their extensive researches and applications. It is
known that MSCs can be isolated from various tissues, such as bone marrow, adipose, and
synovium, and human umbilical cord blood, and bone marrow is one of the essential sources of
MSCs.
Figure 2. Typical extraction process of adipose-derived mesenchymal stem cells from adipose tissue
of mouse.
MSCs exist in various tissues and organs apart from bone marrow, with multilineage cells
from human umbilical cord blood, first reported in early 2000 [9]. Adipose tissue was subsequently
demonstrated as a rich source of MSCs in 2001 [10], and synovium-derived MSCs (SMSCs) were
successfully isolated [11]. MSCs from other tissues or organs were detected, and protocols were
Figure 1.
Schematic diagram of regenerative medicine based on mesenchymal stem cells (MSCs).
The MSCs can be easily extracted from varies tissues, and the multilineage dierentiation and
immunoregulatory properties of MSCs make them an ideal cell therapeutic candidate.
2. Discovery and Extraction of MSCs from Dierent Sources
The rich source of MSCs is the critical basis for their extensive researches and applications. It is
known that MSCs can be isolated from various tissues, such as bone marrow, adipose, and synovium,
and human umbilical cord blood, and bone marrow is one of the essential sources of MSCs.
MSCs exist in various tissues and organs apart from bone marrow, with multilineage cells
from human umbilical cord blood, first reported in early 2000 [
9
]. Adipose tissue was subsequently
demonstrated as a rich source of MSCs in 2001 [
10
], and synovium-derived MSCs (SMSCs) were
successfully isolated [
11
]. MSCs from other tissues or organs were detected, and protocols were
established for their extraction, identification, and culture (Figure 2and Table 1) [
12
30
]. Figure 2and
Table 1describe the general protocols used for MSC extraction. Briefly, the process involves isolation
of various tissues, digestion to obtain cells, and culture for three to five days, followed by discarding
non-adherent cells and continuous culture of adherent cells to the desired passage. The primary culture
medium for MSCs includes low-glucose Dulbecco’s modified Eagle medium (LG-DMEM) with 1%
(W/V) antibiotic/antimycotic and 10% (V/V) fetal bovine serum (FBS). Additionally, Table 1lists a
variety of markers expressed on the MSC surface. Notably, rabbit is the most frequently used animal
model for experiments, involving cartilage or bone tissue regeneration, and should receive increased
focus concerning MSC identification. Moreover, the surface markers of rabbit tissue-derived MSCs
require further verification.
Cells 2019,8, 886 3 of 32
Cells 2019, 8, x 2 of 36
outcomes are inspiring. We present a brief overview of MSC extraction methods and subsequent
potential for differentiation and provide a comprehensive overview of future applications of
various MSCs in regenerative medicine, as well as the challenges.
Figure 1. Schematic diagram of regenerative medicine based on mesenchymal stem cells (MSCs).
The MSCs can be easily extracted from varies tissues, and the multilineage differentiation and
immunoregulatory properties of MSCs make them an ideal cell therapeutic candidate.
2. Discovery and Extraction of MSCs from Different Sources
The rich source of MSCs is the critical basis for their extensive researches and applications. It is
known that MSCs can be isolated from various tissues, such as bone marrow, adipose, and
synovium, and human umbilical cord blood, and bone marrow is one of the essential sources of
MSCs.
Figure 2. Typical extraction process of adipose-derived mesenchymal stem cells from adipose tissue
of mouse.
MSCs exist in various tissues and organs apart from bone marrow, with multilineage cells
from human umbilical cord blood, first reported in early 2000 [9]. Adipose tissue was subsequently
demonstrated as a rich source of MSCs in 2001 [10], and synovium-derived MSCs (SMSCs) were
successfully isolated [11]. MSCs from other tissues or organs were detected, and protocols were
Figure 2.
Typical extraction process of adipose-derived mesenchymal stem cells from adipose tissue
of mouse.
Cells 2019,8, 886 4 of 32
Table 1. Extraction, discrimination, and culture of MSCs derived from various tissues.
MSC Type Source Extraction Approach Culture Medium Marker Reference
BMSCs
Human: tubular
bones and iliac crest
bone marrow
1. Aspirate 1 mL of bone marrow for bone canal;
2. Extraction is diluted in PBS (1:1) and
centrifuged for 30 min at 3000 rpm;
3. The obtained buy coat is isolated, washed,
and plated on culture flasks for incubation
LG-DMEM with 1% (W/V)
antibiotic/antimycotic,
10% (V/V) FBS
CD29+, CD44+, CD73+, CD90+, CD105+,
Sca-1+, CD14, CD34, CD45, CD19,
CD11b, CD31, CD86, Ia,
and HLA-DR
[1315]
Mouse, rat, and rabbit:
tubular bones, e.g.,
femurs and tibias
1. Collect femurs and tibias, cleanse the tissue
with scissors, and wash the bones with 70%
(V/V) ethanol and then PBS;
2. Cut othe proximal and distal parts of bones,
and flush out bone marrow from bone canal by a
spring to culture flasks for incubation;
3. At days 3–5, non-adherent cells are removed
Mouse: CD29+, CD44+, CD73+, CD90+,
CD105+, Sca-1+, CD14, CD34, CD45,
CD11b, CD31, Vcam-1, C-Kit,
CD135, CD11b, Ia, and CD86
[1214,16]
Rat: CD29+, CD44+, CD54+, CD73+,
CD90+, CD105+, CD106+, Sca-1+, CD14,
CD34, CD45, and CD11b
[17,31]
Rabbit: CD29+, CD44+, CD73+, CD81+,
CD90+, CD166+, CD14, CD34, CD45,
CD117, and HLD-DR
[15]
ADSCs
Human:
subcutaneous adipose
in abdomen, buttocks,
and abdominal zone
1. Separate adipose from host body, and mince it
with scissors or scalpel;
2. Digested by collagenase type I for 1 h at 37
C
gently shaking in a water bath;
3. Centrifuge the sample and discard the
superior lipid layer;
4. Filtered through 100 and 40, or 70 µm filters;
5. Washed by 10 mL PBS and centrifuged again;
6. Discard the supernatant, resuspend the cells,
and transfer them to culture flask for incubation
DMEM with 1% (W/V)
P/S, 10% (V/V) FBS
Human: CD29+, CD44+, CD73+, CD90+,
CD105+, CD146+, CD166+, MHC-I+,
CD31, CD45, and HLA-DR
[1821]
Mouse, rat, and rabbit:
subcutaneous adipose
Mouse: CD34+, CD44+, CD45+, CD90+,
MHC-I+, MHC-II+, and CD117.[30]
Rat: CD44+, CD73+, CD90+, MHC-I+,
CD31, and CD45
Rabbit: CD44+, CD105+, NG2+, CD34,
and CD45[22,23]
Cells 2019,8, 886 5 of 32
Table 1. Cont.
MSC Type Source Extraction Approach Culture Medium Marker Reference
SMSCs
Synovium, especially
in knee joints, of
human, mouse, rat,
rabbit, pig, etc.
1. Separate synovium from host knee joint,
and mince it with scissors or scalpel;
2. Digested by collagenase type II, D or P at 37
C, and filtered through a 70 µm nylon filter;
3. The released cells are washed and
resuspended in a culture medium for incubation
DMEM or
α
MEM with 1%
(W/V) P/S, 250 ng mL1
amphotericin B, and 10%
(V/V) FBS
Human: CD10+, CD13+, CD49+, CD44+,
CD73+, CD90+, CD105+, CD147+,
CD166+, CD14, CD20, CD31, CD34,
CD45
, CD62
, CD68
, CD113
, CD117
,
HLA-DR, and ALP
[24]
Mouse: CD29+, CD44+, CD90+, CD34,
CD45, and CD107[25]
Rat: CD90+, CD11b, and CD45[31]
Rabbit: CD44+, CD90+, and CD105+[26,27]
UCB-MSCs Umbilical cord blood
of human
1. Harvest of human umbilical cord blood;
2. Mononuclear cells (MNC) are isolated from
the buy coat layer;
3. Seed into 25 cm
2
flask, and non-adherent cells
are removed after 48 h
LG-DMEM, 1% P/S, 250
ng mL1amphotericin B,
and 10% (V/V) FBS
CD29+, CD44+, CD73+, CD90+, CD105+,
CD166+, CD14, CD31, CD34, CD45,
CD106, and HLA-DR
[28]
Cells 2019,8, 886 6 of 32
3. Dierentiation Potentials of MSC Types
The multi-directional dierentiation potential is one of the most critical characteristics of MSCs.
In addition, dierent tissue sources aect the dierentiation tendency and proliferation capability
of MSCs.
There is an increasing number of publications addressing the heterogeneity of MSCs [
47
].
The transcriptome, proteome, immunophenotype, and immunomodulatory activities of various MSC
types dier, implying that MSCs exhibit unique dierentiation potentials. As a critical MSC-specific
property, dierentiation potential aects MSC fate; dierent tissue-derived MSCs display distinct
tendencies to dierentiate into dierent end-stage lineage cells, such as osteoblasts and chondrocytes.
As a critical source of MSCs for tissue engineering, bone marrow-derived MSCs (BMSCs) exhibit
superior capacities for osteogenesis and chondrogenesis under standard dierentiation protocols [
48
],
and SMSCs show more significant proliferation and chondrogenic potential than adipose-derived
MSCs (ADSCs) [
49
]. Umbilical cord blood-derived MSCs (UCB-MSCs) exhibit biological advantages
relative to other adult sources, including their capability for longer culture times, larger-scale expansion,
more significant retardation of senescence, and higher anti-inflammatory eects [
50
]. Researchers must
choose the desired MSC type according to the specific purpose. Table 2summarizes fundamental and
in vivo experiments and identifies dierentiation conditions based on previous studies.
Cells 2019,8, 886 7 of 32
Table 2. Summary of dierentiation researches and application potential of MSCs.
Dierentiation
Direction *
Preferred
MSC Type
Basic Induction Medium Identify Methods Application Field Reference
Basic Medium Induce Agents Staining IHC RT-PCR Others
Osteoblast BMSCs
LG-DMEM, 10% (V/V)
FBS, 1% (W/V)
antibiotic/antimycotic
(In some studies,
the osteogenic medium
used HG-DMEM
solution)
10.0 mM β-glycerophosphate,
50.0 µg mL1ascorbic acid,
and 100 nM dexamethasone
Alizarin red
staining, Von
Kossa
Staining
Col I, OCN,
OPN
Col I, OCN, OPN, ALP,
BSP, Osterix, RUNX2
ALP activity, Calcium
assay kit
Bone regeneration
[32]
Chondrocyte
50.0 µM ascorbic acid, 100 nM
dexamethasone, 10.0 ng mL1
TGF-β1/TGF-β3
Alcian blue
staining,
Toluidine
blue staining
Col II
Col II, SOX-9, Aggrecan,
SOX-5, SOX-6, NOX 4, Col
X, Chondroitin
4-sulfotransferase
GAG assay kit Cartilage
regeneration [33,34]
Neurocyte BMSCs,
ADSCs
10.0 ng mL1EGF, 20 ng mL1
HGF, 20 ng mL1VEGF; 8 days
later, 200 µM BHA, 5.0 mM KCl,
2.0 µM valproic acid, 10 µM
forskolin, 1.0
µ
M hydrocortisone,
and 5.0 µM insulin are added to
the medium
Enolase,
Tubulin-βIII,
GFAP, S100,
MBP, MAP2,
NF
Tubulin-βIII,GFAP,
Enolase, NeuN, NCAM,
Glial cell marker, NANOG,
OCT4 and SOX-2, MAP2,
NF-M, GAP 43
Nerve
regeneration [3537]
Cardiomyocyte
ADSCs
10.0 µg L1bFGF, 10.0 µM
5-azacytidine;
one day later, the medium
maintained in the same
conditions without 5-azacytidine
for four weeks
Desmin,
M-cadherin,
MHC,
α-cardiac
actin, cTnI
Desmin, MYOD1, MYOG,
MHC, α-cardiac actin,
cTnT, MYF5/6, MEF2C,
TNNI1/2, CKM, Myosin2,
HCN2, HCN4
Heterotypic Cell Fusion
Assay
Myocardial
regeneration [3840]
Hepatocyte PDSCs
1X ITS, 108M dexamethasone,
20.0 ng mL1EGF, 20.0 ng mL1
FGF, 40.0 ng mL1OsM, 40 ng
mL1HGF;
After two weeks, the medium is
replaced with hepatic
dierentiation medium with an
increased concentration of
dexamethasone at 10
5
M and/or
1.0 µM TSA
PAS staining
ALB, AFP,
CK-18,
PanCK, CK 19,
Transthyretin
ALB, AFP,β-actin, CK-18,
HNF-4α, Transthyretin,
TDO2, and CYP7A1
LDL/CM-Dil uptake
assay; Cell morphology;
Ammonia clearance;
Albumin production;
ELISA assay
Liver
regeneration [4144]
Keratocyte
No
comparative
studies
LG-DMEM: F-12 3:1,
5% FBS, 1% (W/V)
antibiotic/antimycotic
Induction medium: without
pyruvate, 25.0 ng mL1BMP-4,
1.0 mM all-trans retinoic acid,
and 10.0 ng mL1EGF;
Dierentiation medium: 5.0 µg
mL1insulin, 2.0 nM
tri-iodothyronine, 2.0 nM
adenine, and 10.0 ng mL1EGF
H&E staining
CK3,
β1-integrin,
and E-cadherin,
p63, CK12,
CK8, CK14,
CK15
ABCG2, β1-integrin,
CEBP
δ
, CK3, and p63, Oct4,
Sox2, Nanog, Rex1, DSC1,
and DSG1
Transepithelial
electrical resistance
Corneal
regeneration [45,46]
* The culture media for another end-stage lineage cells have not been standardized except osteoblasts and chondrocytes. The bold words in identified methods mean the main identified
staining, proteins of immunohistochemistry, genes of real-time reverse transcription polymerase chain reaction (RT-PCR), and other methods.
Cells 2019,8, 886 8 of 32
4. MSC-Based Regenerative Medicine
So far, MSCs have been widely studied and applied in regenerative medicine. In this section,
we summarize reports concerning the latest preclinical and clinical trials of various MSC types for
tissue engineering. The topics mainly focus on the reconstruction of fragile tissues, including those
associated with the musculoskeletal system, nervous system, myocardium, liver, cornea, trachea,
and skin, as shown in Figure 3.
Cells 2019, 8, x 9 of 36
Figure 3. Applications of mesenchymal stem cells with multiple differentiation potential for repair
of various tissues.
3. Differentiation Potentials of MSC Types
The multi-directional differentiation potential is one of the most critical characteristics of
MSCs. In addition, different tissue sources affect the differentiation tendency and proliferation
capability of MSCs.
There is an increasing number of publications addressing the heterogeneity of MSCs [47]. The
transcriptome, proteome, immunophenotype, and immunomodulatory activities of various MSC
types differ, implying that MSCs exhibit unique differentiation potentials. As a critical MSC-specific
property, differentiation potential affects MSC fate; different tissue-derived MSCs display distinct
tendencies to differentiate into different end-stage lineage cells, such as osteoblasts and
chondrocytes. As a critical source of MSCs for tissue engineering, bone marrow-derived MSCs
(BMSCs) exhibit superior capacities for osteogenesis and chondrogenesis under standard
differentiation protocols [48], and SMSCs show more significant proliferation and chondrogenic
potential than adipose-derived MSCs (ADSCs) [49]. Umbilical cord blood-derived MSCs (UCB-
MSCs) exhibit biological advantages relative to other adult sources, including their capability for
longer culture times, larger-scale expansion, more significant retardation of senescence, and higher
anti-inflammatory effects [50]. Researchers must choose the desired MSC type according to the
Figure 3. Applications of mesenchymal stem cells with multiple dierentiation potential for repair of
various tissues.
4.1. Bone Regeneration
Bone defects frequently accompany recovery from trauma, revision arthroplasty, or tumor resection
surgeries. Autologous bone grafting represents the gold standard therapeutic strategy, despite its many
drawbacks, including (1) the limited supply of autologous bone, (2) increased operation time and blood
loss, (3) temporary disruption of bone structure in the donor site, and (4) donor site morbidity [
51
].
Allografting carries a risk of disease and/or infection [
52
]. Therefore, MSC-based bone regeneration is
considered an optimal approach [53].
Cells 2019,8, 886 9 of 32
The MSC osteoblast-dierentiation capacity has been identified [
2
,
3
], with BMSCs representing
the most frequently applied cells for osteoblast dierentiation [
2
]. Comparative studies evaluating the
osteogenic ability of other MSC types yielded no definitive conclusions. By contrast, UCB-MSCs show
better angiogenic capacity, supporting more abundant blood supply during bone regeneration [
54
],
which promotes rapid tissue reconstruction. In addition to BMSCs, human dental pulp stem cells
(hDPSCs) show excellent vascular dierentiation potential while dierentiating into osteoblasts,
which subsequently support bone regeneration [
55
]. However, these hDPSCs are screened from a
stromal vascular dental pulp fraction; therefore, this represents a limited source for further research
and application. Since ADSCs can be routinely isolated from lipoaspirate with a high degree of purity
with minimal donor site morbidity or patient discomfort, ADSCs are considered to have the most
significant potential as a primary source for clinical bone tissue engineering [
48
,
56
,
57
]. Additional
comparative and screening studies are necessary to identify other cell sources with applications in
bone reconstruction.
Stimulating factors play an important role in directing MSC dierentiation into target cells
in vitro
.
The most commonly used inducing factor for osteogenesis is the bone morphogenetic protein-2 (BMP-2),
which is usually immobilized on scaolds to promote osteoblast dierentiation. BMP-2 exhibits a
strong osteogenic ability, which can be tested by the osteoblast activity and/or expression of bone
markers, such as alkaline phosphatase (ALP), osteopontin (OPN), and osteocalcin (OCN) [
58
60
].
BMP-7 activates the transforming growth factor-
β
(TGF-
β
)/SMAD signaling in CD105
+
MSCs to
enhance the expression of osteogenesis-related genes [
61
]; Wnt11 enhances the osteogenic potential of
BMP-9 [
62
]. Nano-hydroxyapatite [
58
] and strontium [
63
] are used as osteogenic regulators in tissue
engineering to promote osteogenic dierentiation of MSCs while changing the physical properties of
the scaolds.
Studies of MSC-based cell therapy for bone defects and the use of novel scaolds describe
inspiring advances
in vitro
and
in vivo
[
64
,
65
]. Clinical applications of MSCs in bone reconstruction
have been described, including those involving implantation of scaolds seeded with MSCs into bone
defect sites. Specifically, dentists have used this technique to address alveolar cleft defects, jaw defect
reconstruction, and maxillary sinus augmentation, with excellent outcomes [
66
68
]. Defects in or
non-union of human tubular bone have been tentatively treated via local implantation of MSCs with or
without scaolds [69,70].
4.2. Cartilage Repair
Cartilage defect repair is one of the significant challenges faced by orthopedic surgeons. Due to
the inherent avascular nature of cartilage and the proliferation of mature chondrocytes, cartilage
is greatly limited in its ability to repair itself. Currently, the clinically applied cartilage repair
techniques, such as bone marrow stimulation and osteochondral transplantation, have their limitations.
Fibrocartilage produced by bone marrow stimulation is not strong enough, and grafts for osteochondral
transplantation are challenging to integrate.
MSCs oer a new strategy for the repair of damaged cartilage, as they can dierentiate
into chondrocytes [
2
,
3
]. An integrated cartilage reconstruction unit comprises cells, scaolds,
and stimulatory factors, with BMSCs [
71
], ADSCs [
56
], and SMSCs [
72
], used as the primary cell
sources. Among these, BMSCs displayed better chondrogenic capacity
in vitro
and
in vivo
[
33
,
34
,
71
],
although SMSCs show better proliferation and dierentiation potential and less hypertrophy than
BMSCs and ADSCs [
72
]. Cartilage reconstruction requires a combination of multiple stimulating
factors; co-culture of chondrocytes and MSCs would achieve better results than the application of
MSCs alone [73].
Novel bioactive three-dimensional (3D) scaolds, such as hydrogels [
55
] and electrospun
scaolds [
74
], have undergone constant improvement. These scaolds provide an optimal 3D
microenvironment for cartilage regeneration. Moreover, hydrogels can regulate MSC proliferation
and dierentiation due to their high water content and biocompatibility and similar properties to the
Cells 2019,8, 886 10 of 32
extracellular matrix (ECM) [
55
]. Injectable hydrogels enable minimally invasive treatment of large
areas of cartilage defects [
75
]; thus, hydrogels loaded with MSCs and stimulating factors are highly
ecacious for the repair of cartilage damage. The development of electrospun scaolds suggests that
the arrangement of nanofibers also aects cell dierentiation and provides a dierent approach to
cartilage repair [
74
]. Stimulating factors are necessary for cartilage engineering and responsible for
inducing, accelerating, and/or enhancing cartilage formation. Common stimulating factors include
BMP-2/-4, insulin-like growth factor (IGF)-1, and TGF-
β
1/-
β
3 [
76
78
]. Moreover, physical stimuli,
such as hydrostatic pressure and dynamic compression, have been explored to induce MSC-mediated
cartilage formation [79].
The first preclinical trial of MSC application for cartilage repair occurred in 1994 [
80
]. MSCs were
seeded into a collagen (Col) gel to treat a full-thickness defect in rabbit femoral cartilage, resulting in
better outcomes than those observed in a control group. This defect model was subsequently used
as a classical cartilage defect model for cartilage regeneration; subsequently, numerous trials have
been conducted in both animal models and humans to evaluate MSC-based therapy for cartilage
damage [81].
Despite clinical trials being conducted, there are no commercially available products for MSC-based
cartilage reconstruction [
82
,
83
]. Several studies investigated the eects of expanded MSCs
in vivo
on damage to human articular cartilage. Transplantation of expanded autologous BMSCs improved
cartilage quality in patients with chronic knee osteoarthritis [
83
], although the clinical improvement was
not significant [
82
]. Other studies reported the injection of allogeneic MSCs into joints in the presence
or absence of pre-mixing with autologous chondrocytes [
81
]. All the clinical outcomes indicated the
safety of these therapeutic approaches, and their ability to relieve some symptoms, although their
ability to repair the eects of cartilage damage was not always apparent. MSC transplantation showed
better results for early lesions [81].
There remain many challenges for MSC-based cartilage regeneration, including the identification
of optimal cell sources. Additional studies are needed to enable the use of MSC-based materials as
commercial products for implantation to promote cartilage regeneration.
4.3. Regeneration of Other Musculoskeletal Tissues
Recent studies investigated the MSC-mediated regeneration of musculoskeletal tissues outside
the bone and cartilage, including the meniscus, tendons and ligaments, and intervertebral discs (IVD).
Meniscus regeneration has received increasing attention. Intra-articular administration of MSCs
to promote meniscal regeneration was first performed with favorable outcomes [
84
]. Similar to its
use for cartilage regeneration, hydrogels [
85
] and electrospun scaolds [
86
] loaded with MSCs were
used to reconstruct the meniscus. Moreover, the meniscus-derived decellularized matrix shows better
histocompatibility and is more capable of inducing MSC dierentiation as compared with natural
or synthetic polymer materials [
87
]. Scaold-free tissue-engineered constructs show promise as an
MSC-based implantation technique to repair meniscal lesions [
88
]. Tarafder et al. [
89
] proposed the
recruitment of synovial MSCs through connective tissue TGF and TGF-
β
3 to repair meniscus injury,
thereby avoiding the disadvantages of cell-based techniques. Mechanical stimulation is crucial for
meniscus growth and maintenance, with mechanical stimuli, such as dynamic compression and tensile
loads applied for meniscus repair [
90
]. Although satisfactory results were obtained in animal models,
there remains a lack of evidence in humans regarding the capability of MSCs for forming durable
tissues similar to the meniscus [91].
Tendon injury is a common problem associated with sports [
92
]. BMP-14 induces myogenic
dierentiation of BMSCs via the sirtuin-1
Janus N-terminal kinase (JNK)/SMAD1-peroxisome
proliferator-activated receptor-
γ
signaling pathway [
93
]. Studies describing tendogenic dierentiation
of MSCs were not limited to stimulating factors [
94
] and scaolds [
95
] but also referred to mechanical
stimuli that play essential roles in MSC dierentiation into tendon lineages. Uniaxial cyclic stretching
Cells 2019,8, 886 11 of 32
promoted tendogenic dierentiation of MSCs
in vitro
and
in vivo
[
96
]; however, MSCs did not repair
tendon injury but only delayed lesion progression [97].
With the increasing age of populations, IVD degeneration has become prevalent. MSCs represent
promising candidates for disc regeneration; scaolds made of Col provide readily available support for
chondrogenic dierentiation of MSCs
in vitro
, although the phenotype of the dierentiated MSCs is not
yet equivalent to that of nucleus pulposus (NP) cells [
98
]. The acellular matrices derived from NP cells
stimulated by TGF-
β
3 also enhance MSC dierentiation [
99
]. Transfection of adenoviral expression of
SOX-9 and BMP-2 in BMSCs increased Col II and aggrecan expression, and promoted IVD repair [
100
].
Varma et al. [
101
] used a hydrogel loaded with two dierent concentrations of MSCs to repair NP,
and showed that MSC inoculation at a lower density resulted in a better NP-specific matrix phenotype.
A systematic review of MSC-based cell therapy for IVD indicated the safety and eectiveness of
short-term MSC transplantation, as well as the necessity for human-based clinical trials [
102
]. In 2011,
expanded autologous BMSCs injected into patients with lumbar disc degeneration revealed several
advantages and better prognosis relative to the current gold standard treatments [
103
]. Clinical
percutaneous injection of autologous bone marrow concentrate cells into a patient with degenerative
IVD resulted in decreased lumbar discogenic pain within 13 years [
104
]. Clinical studies [
105
] indicated
that MSC transplantation represents a safe treatment option for degenerative IVD; the specific eects
need verification by additional clinical trials.
4.4. Central Nervous System Rebuilding
The adult central nervous system (CNS) lacks the ability to repair damaged neurons, so the
damage of CNS is irreversible, and there is currently no eective repair method for CNS injury in
clinical practice repair. In the area of CNS regeneration, MSC-based therapy mainly focuses on two
areas: Damage or injury of the CNS caused by severe trauma and continuous ischemia and CNS
dysfunction caused by neurologic diseases. To date, BMSCs and ADSCs are the most extensively
studied cell sources for CNS repair, with each showing similar neuronal dierentiation potential [
35
,
36
].
BMSCs can reduce scar formation around spinal cord injury (SCI) lesions and promoted axonal
regeneration [
106
]; however, ADSCs might represent a more suitable cell source owing to their easy
extraction and abundant sources. ADSCs inhibit inflammation of the nervous system and improve
the recovery of function from traumatic brain injury via neural stem cells [
107
]. UCB-MSCs can be
induced to dierentiate into neuron-like cells
in vitro
[
37
]; DPSCs can dierentiate into neurons and
express multiple factors that promote neuronal and axonal regeneration [108].
MSC expression of neuronal or astrocytic markers has been observed
in vitro
[
109
] and
in vivo
[
110
].
To promote MSC-based CNS restoration, gene-modified MSCs, such as neurotrophin-3-transferred
BMSCs, showed improved neuronal dierentiation
in vivo
[
111
]. Persistent release of specific
cytokines and growth factors, which can facilitate neurogenesis, angiogenesis, and synaptogenesis,
creates a favorable microenvironment for angiogenesis or remyelination during reconstruction [
112
].
IGF-1-transfected spinal cord-derived neural stem cells displayed higher viability and the ability to
dierentiate into oligodendrocytes [
113
]. Moreover, MSCs can induce T cell tolerance and release of
paracrine anti-inflammatory factors, such as TGF-β, that promote neuroprotective eects [114].
Animal models of traumatic and ischemic brain injury or SCI have been used to evaluate
MSC-based therapy [
115
,
116
]. A meta-analysis of 1568 rats with traumatic SCI showed that MSC
therapy provided a substantial beneficial eect on locomotor recovery [
116
]. Clinical studies indicated
MSC-based therapy as a safe and feasible technique for patients with SCI and/or traumatic brain
injury [
117
,
118
]. Migration of MSC pretreated under hypoxic conditions to the peri-cerebral injury
area of cerebral hemorrhagic stroke victimized rats resulted in the release of various growth factors to
promote neurogenesis and neurological recovery [
119
]. For neurological diseases, non-expanded or
expanded MSCs have been widely used for the treatment of multiple sclerosis [
120
], amyotrophic lateral
sclerosis [
121
], ischemic stroke [
122
], and Parkinson’s disease [
123
]. Most of the beneficial eects of MSCs
on neurological diseases are associated with their immunomodulatory and neuroprotective properties
Cells 2019,8, 886 12 of 32
exerted following local injection of non-expanded or expanded autologous MSCs, with clinical trials
assessing their ability to achieve promising outcomes and dierent degrees of remission.
4.5. Peripheral Nervous System Rebuilding
Peripheral nervous system (PNS) injury is mainly caused by severe trauma, usually accompanied
by bone fracture and vascular damage. Autologous nerve grafting (autologous nerve bridging) is
the gold standard for peripheral nerve repair; however, limited donor nerve resources and other
issues preclude the search for new therapeutic strategies. Schwann cells, neurotrophic factors,
and anti-inflammatory cells work together to promote peripheral nerve regeneration, with this process
involving axonal sprouting and fiber myelination.
There are few comparative studies concerning the eects of dierent MSC types on animal
models of peripheral nerve injury. ADSCs are more suitable cell sources for neural regeneration
in vitro
[
35
], and a nerve growth factor transcript has been identified in ADSC-secreted nanovesicles
that promotes synaptic growth
in vitro
and repair of sciatic nerve damage
in vivo
[
124
].
Sun et al.
[
125
]
proposed a new protocol called intermittent induction that alternates complete and incomplete
induction media to induce ADSC dierentiation into SLCs. Compared with traditional protocols,
SLCs obtained by intermittent induction secrete neurotrophic factors and promote axonal growth
in vitro
and more eectively repair rat sciatic nerve injury
in vivo
. Notably, SLCs seeded in acellular
nerve grafts show better functional recovery as compared with MSCs [
126
]. BMSCs [
111
], ADSCs [
127
],
and UCB-MSCs [
128
] have also been seeded onto a variety of biodegradable scaolds, with almost
all resulting in better recovery relative to controls. MSCs form a neuroblast-like sheath following
transplantation at the site of nerve injury and secrete neurotrophic factors that provide physical and
chemical barriers for the inner nerve fibers [
110
]. Compared with polymers, acellular neural matrix
hydrogels show better biocompatibility and tissue specificity and support Schwann cell proliferation
in vitro
and repair rat sciatic nerve defects
in vivo
[
129
]. Furthermore, 3D-bioprinting technology has
enabled the development of 3D scaolds with complex structures to address the challenges of nerve
tissue regeneration [130].
For nerve regeneration studies, sciatic nerve crush and nerve gap animal models were
established [
131
,
132
], and the eects of local implantation [
133
] or intravenous administration [
134
,
135
]
of neural stem cells or MSCs in peripheral nerve injury models have been investigated, resulting in
excellent outcomes relative to controls. ADSCs displayed relevant therapeutic potential not only via
their direct release of growth factors but also through the indirect modulation of neurocyte behavior in
an animal model of acute axonal injury [
134
]. Moreover, intravenously infused MSCs ameliorated
function recovery post-acute peripheral nerve injury in a sciatic nerve crush model [135].
Although preclinical studies show the feasibility of MSC-based therapy in animal models of
peripheral nerve injury, there are few reports of its clinical application [126].
4.6. Myocardium Restoration
Cardiac disease is characterized by substantial morbidity and mortality, and serious adverse
consequences. In addition to congenital heart disease, almost all cardiac diseases involve insucient
blood supply to critical regions, resulting in myocardial damage and necrosis. Although myocardium
has limited regenerative capacity, restoration of severe damage to cardiomyocytes due to catastrophic
myocardial infarction or other myocardial diseases is inadequate.
A role for MSC in attenuating myocardium damage was first reported in 2002 [
136
]; purified BMSCs
engrafted in the murine myocardium appeared to dierentiate into cardiomyocytes. Several subsequent
studies evaluated the potential of dierent MSC sources to dierentiate into cardiomyocytes [
38
40
],
finding ADSCs as the most suitable. Spraying was found to be a more cost-eective and less
invasive method for transferring ADSCs into a pig heart infarction model to promote cardiac function
recovery [
137
]. MSC-specific mechanisms associated with the repair of damaged cardiomyocytes
involve three factors: (1) myogenic and angiogenic capacities; (2) the ability to supply massive
Cells 2019,8, 886 13 of 32
amounts of angiogenic, anti-apoptotic, and mitogenic factors; and (3) the inhibition of myocardial
fibrosis (Figure 4) [
138
,
139
]. Butler et al. [
140
] demonstrated the safety of MSC therapy, and that it
improved the left ventricular ejection fraction in patients with non-ischemic cardiomyopathy via its
immunomodulatory eects. Co-culture of MSCs with cardiomyocytes promotes resistance to high
oxidative stress in heart tissue after myocardial infarction [
141
]. 5-Azacytidine is an eective factor
for inducing MSC dierentiation into cardiomyocytes [
142
]. IGF-1-transfected MSCs protected the
myocardium from fibrosis and cardiomyocyte apoptosis and reduced infarct size after myocardial
infarction in rats [
143
]. Interleukin (IL)-7 enhances the fusion of MSCs with cardiomyocytes to improve
cardiac function, with this attributed to the ability of IL-7 to promote cell proliferation and support
damaged myocardial regeneration [
144
]. In addition to their ability to dierentiate into cardiomyocytes,
MSCs promoted angiogenesis by secreting vascular endothelial growth factor (VEGF) in a critical limb
ischemia model [145], resulting in cardiac reconstruction.
MSC-based therapy is a feasible strategy to improve cardiac function, according to preclinical and
clinical findings [
146
]. Furthermore, MSC therapy has been intensively investigated as a treatment for
myocardial infarction [
147
], peripheral ischemic vascular diseases [
148
], dilated cardiomyopathy [
138
],
and pulmonary hypertension [149].
Cells 2019, 8, x 15 of 36
Figure 4. Schematic mechanisms of mesenchymal stem cells (MSCs) for cardiac regeneration.
Angiogenesis, vasculogenesis, and cardiomyocytes differentiation capacities of MSCs make them
possible for cardiac repair. Moreover, the paracrine effects of MSCs provide different kinds of
growth and anti-inflammatory factors for the immunoregulation after ischemia of heart [139].
4.7. Liver Regeneration
The liver is an essential human organ, failure of which causes fatal illnesses. Until recently, the
only effective therapy for liver failure was organ transplantation [150]; however, the availability of
transplantable livers is scarce, and many patients do not survive the wait time for transplantation.
Cell therapy provides a possible solution by either building a partial or full liver for transplantation
or addressing the damage to the liver. Although hepatocytes isolated from the livers of donors have
been studied, their therapeutic efficacy is questionable owing to their immunogenicity. Recently,
MSCs have become a therapeutic option owing to their ability to exert potent immunosuppressive
and anti-inflammatory effects [8]. Importantly, the multilineage potential of MSCs to differentiate
into different types of end-stage cells, including hepatocyte, makes them an attractive candidate for
liver-specific therapeutics [151].
The clinical use of MSCs to treat liver failure is mainly based on their immunomodulatory
capacity [152]. Preclinical and clinical studies revealed the mechanisms associated with
immunoregulation following MSC-based treatment, including transdifferentiation, fusion,
inhibition of Col deposition, paracrine effects, modulation of matrix metalloproteinase expression
and activity, neoangiogenesis, and vascular support [153–155]. A clinical phase II trial showed that
transplantation of autologous BMSCs inhibited tissue fibrosis and improved liver function in
alcoholic cirrhosis [156]; intravenous injection of allogeneic MSCs positively affected the treatment
of autoimmune cirrhosis [157]. Administration of UCB-MSCs reduces the liver inflammatory
response and hepatocyte injury, as well as the possibility of liver failure, by inhibiting T cell and B
cell proliferation and upregulating the levels of regulatory T cells [158].
Figure 4.
Schematic mechanisms of mesenchymal stem cells (MSCs) for cardiac regeneration.
Angiogenesis, vasculogenesis, and cardiomyocytes dierentiation capacities of MSCs make them
possible for cardiac repair. Moreover, the paracrine eects of MSCs provide dierent kinds of growth
and anti-inflammatory factors for the immunoregulation after ischemia of heart [139].
4.7. Liver Regeneration
The liver is an essential human organ, failure of which causes fatal illnesses. Until recently,
the only eective therapy for liver failure was organ transplantation [
150
]; however, the availability
of transplantable livers is scarce, and many patients do not survive the wait time for transplantation.
Cells 2019,8, 886 14 of 32
Cell therapy provides a possible solution by either building a partial or full liver for transplantation
or addressing the damage to the liver. Although hepatocytes isolated from the livers of donors have
been studied, their therapeutic ecacy is questionable owing to their immunogenicity. Recently,
MSCs have become a therapeutic option owing to their ability to exert potent immunosuppressive
and anti-inflammatory eects [
8
]. Importantly, the multilineage potential of MSCs to dierentiate
into dierent types of end-stage cells, including hepatocyte, makes them an attractive candidate for
liver-specific therapeutics [151].
The clinical use of MSCs to treat liver failure is mainly based on their immunomodulatory
capacity [
152
]. Preclinical and clinical studies revealed the mechanisms associated with immunoregulation
following MSC-based treatment, including transdifferentiation, fusion, inhibition of Col deposition,
paracrine effects, modulation of matrix metalloproteinase expression and activity, neoangiogenesis,
and vascular support [
153
155
]. A clinical phase II trial showed that transplantation of autologous BMSCs
inhibited tissue fibrosis and improved liver function in alcoholic cirrhosis [
156
]; intravenous injection
of allogeneic MSCs positively affected the treatment of autoimmune cirrhosis [
157
]. Administration of
UCB-MSCs reduces the liver inflammatory response and hepatocyte injury, as well as the possibility
of liver failure, by inhibiting T cell and B cell proliferation and upregulating the levels of regulatory T
cells [158].
In addition to their immunomodulatory eects, MSCs can dierentiate into hepatocytes to
promote liver regeneration. MSCs can dierentiate into several types of liver cells under certain
physiopathological conditions [
151
]. Functional hepatocyte-like cells were obtained from multipotent
adult progenitor cells [
159
]. Hepatocyte growth factor (HGF) and oncostatin M have been used
to successfully induce human BMSCs and human UCB-MSCs to dierentiate into hepatocyte-like
cells [
160
]. Hepatocyte-like cells have been derived from BMSCs [
161
], ADSCs [
41
,
42
], UCB-MSCs [
43
],
and placenta-derived MSCs (PDSCs) [
44
] via appropriate culture conditions, including co-culture
with hepatocytes on 2D or 3D scaolds [
142
], floating culture, or treatment with serum collected from
rats after partial hepatectomy [
162
]. Liver reconstruction using dierentiated MSCs has only been
demonstrated in the animal models, although recently, a phase II trial showed that transplantation of
dierentiated autologous MSCs could be used as a potential clinical treatment for liver cirrhosis [14].
There remain two obstacles to MSC-based liver regeneration: transplanted MSCs cannot eciently
dierentiate into hepatocyte-like cells, and transplanted MSCs might promote fibrogenesis [
163
].
Moreover, a previous study suggested that transplanted MSCs might contribute to the myofibroblast
pool to enhance fibrotic processes within the liver [
164
]. However, a clinical study [
165
] reported
that MSC injections were suciently safe, but their immunosuppressive eects following liver
transplantation were insucient to replace immunosuppressants, despite the severe side eects of
immunosuppressants commonly used in clinical practice.
4.8. Corneal Reconstruction
The cornea represents a transparent avascular connective tissue that provides most of the refractive
ability of the eye and acts as the primary barrier against infection and mechanical damage to internal
structures. Because the cornea is fragile and directly exposed to the external environment, a variety
of clinical disorders, such as aniridia and Stevens-Johnson syndrome, or chemical, mechanical,
and thermal injury potentially causes corneal injury. Human corneal epithelial cells can be renewed by
stem cells located in the peripheral region of the cornea, containing limbal epithelial stem cells (LESCs).
However, if the entire corneal layer is damaged, vascularization, conjunctivalization, keratinization,
corneal scarring, and opacification occur and lead to impaired vision and even blindness. Currently,
keratoplasty is the primary method used for corneal restoration, because sources of donor tissue are
restricted, and post-operation immune rejection compromises transplantation ecacy.
Transplantation of corneal epithelial stem cells by limbal allograft restored useful vision to some
patients with severe ocular surface disorders [
166
]. LESCs have been investigated as the main cell
therapeutic candidates for corneal disorders, with clinical and/or preclinical studies demonstrating
Cells 2019,8, 886 15 of 32
that LESCs cultured on an amniotic membrane eectively inhibited inflammation and reconstruct the
injured corneal surface [
167
]. However, although LESCs have been widely used in clinical practice,
visual recovery after successful transplantation remains sub-optimal [
168
]. MSCs were used for corneal
repair include BMSCs [
45
], ADSCs [
169
], and UCB-MSCs [
170
]. Allogeneic MSCs exhibited good
immunosuppressive eects in a pre-sensitized rat model of corneal transplantation [
171
]; however,
no comparative studies are evaluating the eects of dierent MSC types on cornea reconstruction.
MSC transplantation successfully reconstructed the damaged corneal surface in a rat model,
and the main therapeutic eect of MSCs was not epithelial dierentiation but instead an inhibition of
inflammation and angiogenesis following transplantation [
172
]. Transplanted amniotic fluid-derived
MSCs reduce neovascularization and promote an anti-inflammatory and anti-fibrotic environment [
167
].
Several studies have investigated the dierentiation processes associated with corneal reconstruction.
Rabbit BMSCs promoted the healing of injured corneal epithelium and could be induced to dierentiate
into corneal epithelial-like cells expressing CK3, a corneal epithelium-specific marker,
in vivo
and
ex vivo [
173
], and dierentiated MSCs cultured on an amniotic membrane expressed the CK3
marker [
46
]. MSCs cultured in the keratocyte-conditioned medium dierentiated into keratocyte-like
cells [
169
], and Yamashita et al. [
170
] induced UCB-MSC dierentiation into tissue-engineered corneal
endothelial cells
in vitro
, which subsequently displayed improved preservation of corneal thickness
and transparency. Further investigation is necessary to reveal whether MSCs play a role in corneal
reconstruction through their transdierentiation eects.
The immunoregulatory, anti-inflammatory, and immunosuppressive capacities ofMSCs play essential
roles in their therapeutic effects and have been demonstrated in studies of their transdifferentiation
effects [
172
,
174
]. The associated mechanisms include downregulation of CD45 and IL-2 expression [
172
],
downregulation of macrophage inflammatory protein-1
α
and VEGF [
175
], secretion of soluble
factors [
176
], suppression of T cell proliferation [
177
], reduction of activation and migration of chemokine
receptor-7-antigen-presenting cells [
174
], and abortion of early inflammatory responses by secretion of
tumor necrosis factor-
α
-stimulated gene/protein-6 (TSG-6) [
178
]. MSCs employed as therapeutic agents
for reconstructing corneas in animal models have resulted in mostly good prognoses [46].
Most available studies describe the strategies of MSC-based corneal reconstruction as involving
their anti-inflammatory eects associated with corneal angiogenesis and their ability to dierentiate
into keratocyte-like cells; however, their dierentiation eciency and purity need to be improved.
4.9. Tracheal Reconstruction
Tracheal stenosis and other severe diseases, such as tracheobronchial cancer, require partial
tracheal resection and reconstruction. Recently, clinical trials have been undertaken to promote tracheal
reconstruction using tracheal substitutes, including autologous grafts, homografts, and prostheses [
179
].
Based on the progress in MSC-based tissue regeneration, investigations have been conducted concerning
their ecacy for tracheal reconstruction.
For regeneration of airway epithelium, dierent culture methods, such as co-culture with epithelial
cells or in combination with inducing factors, such as VEGF, brain-derived neurotrophic factor, TGF-
β
1,
and activin A, have been explored. These four factors play a vital role in MSC dierentiation into
epithelial cells by triggering the appropriate signaling pathways [
180
]. Other induction factors also
promote epithelial-lineage dierentiation, including epidermal growth factor (EGF), keratinocyte
growth factor, HGF, and IGF-2 [
181
]. Physical culture conditions are also involved in regulating
epithelial-lineage dierentiation, with studies revealing that compartmentalized or polarized culture
conditions provide a suitable environment for MSC dierentiation into epithelial progenitor cells with
tight junction formation [
182
]. Similarly, BMSCs and porous tracheal scaolds implanted
in vivo
after
co-culture
in vitro
maintain their structural integrity and significantly reduce immune rejection [
183
].
For whole trachea regeneration, preclinical trials using animal models verified the ecacy and
safety of MSC-based tracheal reconstruction. Seeding MSCs onto decellularized trachea scaolds
represents a promising means of trachea engineering in rats [
184
]; an acellular tracheal matrix inoculated
Cells 2019,8, 886 16 of 32
with BMSCs showed excellent biocompatibility and immunogenicity [
185
]. A 3D-bioprinted artificial
scaold coated with MSCs seeded in fibrin was constructed to repair partial tracheal defects, resulting in
successful restoration in the absence of graft rejection in four rabbits [
186
]. An aortic allograft for trachea
replacement with MSC transplantation was also demonstrated as a feasible reconstruction method for
tracheal defects [
187
]; Jorge et al. [
188
] used an acellular amniotic membrane for tracheal reconstruction,
revealing that the membrane promoted cartilage regeneration, neovascularization, and epithelialization
and reduced the risk of postoperative complications, such as tracheal stenosis. However, low porosity of
the acellular tracheal matrix might lead to incomplete cartilage regeneration [
189
]; therefore, they used
laser micropore technology to alter scaold porosity, with
in vivo
results confirming that the technique
improved the cartilage matrix and the mechanical strength of the scaolds. Furthermore, artificial
biomaterial-based scaolds seeded with MSCs, such as Col-based electrospun scaolds [
190
] and
core
shell nanofibrous scaolds [
191
], represent viable options for replacing a damaged trachea,
although scaolds loaded with MSCs promote epithelium formation and angiogenesis
in vivo
but in
the absence of cartilage formation [192].
A tissue regeneration trial involving tracheal reconstruction supported the use of MSCs for
clinical tracheal tissue engineering based on their contribution to grafted tissue integration and
angiogenesis [
193
]. In 2008, the first transplantation of a tissue-engineered trachea was performed on a
30-year-old woman with end-stage left-main bronchus malacia and involved a trachea comprising
decellularized donor trachea, epithelial cells, and BMSC-derived chondrocytes [
194
]. At a five-year
follow-up, the patient exhibited normal lung function, ciliary function, cough reflex, and mucus
clearance in the absence of stem cell-related teratoma and anti-donor antibodies [
195
]. This result
demonstrated that the cell-seeded scaold was clinically safe and feasible, and was followed by another
BMSC-based tissue-engineered tracheal replacement in a 12-year-old boy, which also showed good
results at a two-year follow-up [196].
To reconstruct a functional trachea, collaborative techniques, including those associated with stem
cell biology, biomaterials, and tissue engineering, are essential, with the trachea required to assemble
around the dierentiating MSCs and matrix components.
4.10. Skin Regeneration
Skin is the first line of defense against microorganisms or physical damage, with infections
and/or trauma potentially leading to skin defects. Treatment of large-scale skin damage is sometimes
inadequate because autologous transplantation of skin is limited by skin tissue availability. Additionally,
allogeneic skin grafts always cause immunological rejection or communicable illnesses.
Wound healing is a complex process involving interactions between soluble mediators, the ECM,
and infiltrating blood cells [
197
]. MSC-based therapy combined with artificial scaolds oers a
promising strategy to promote wound healing or complete reconstruction of full-thickness skin
(Figure 5) [
198
]. Feldman et al. [
199
] used TGF-
β
3, albumin-based scaolds, and MSCs to treat pressure
ulcers, with good therapeutic results. Recent progress has also been reported in skin regeneration via
MSC-based therapy; MSCs significantly improve wound condition and angiogenesis. MSC-secreted
TSG-6 improves wound healing by limiting macrophage activation, inflammation, and fibrosis [
200
];
VEGF secreted by MSCs promotes keratinocyte-mediated wound healing [
201
]. Angiotensin II promotes
BMSC dierentiation into keratinocytes through the mitogen-activated protein kinase/JNK/Janus
kinase 2 signaling pathways [
202
]. Furthermore, gene-modified MSCs provide another possible
way to promote skin regeneration. EGF-transfected MSCs [
203
], adenovirus-transfected C-X-C motif
chemokine receptor-4-overexpressing BMSCs [
204
], VEGF-modified human UCB-MSCs [
205
], stromal
cell-derived factor-1-transfected BMSCs [
206
], and ectodysplasin-modified MSCs [
207
] promote
MSC-mediated wound healing activity in skin defects, as does physical stimuli, such as laser
therapy [
208
]. A large number of scaolds, such as those involving fibrin hydrogels [
209
], 3D-hybridized
chitosan (CS) and poly(
ε
-caprolactone) (PCL) [
210
], Col
CS [
211
], sodium carboxymethylcellulose [
212
],
and electrospun nanofibrous silk fibroin [
213
], have been developed to support MSC-based regeneration
Cells 2019,8, 886 17 of 32
of defective skin. Qi et al. [
214
] developed a photo-cross-linkable sericin hydrogel to repair full-thickness
skin damage; this hydrogen inhibited inflammation, stimulate angiogenesis, and recruit MSCs to the
site of injury to regenerate skin appendages.
Recently autologous and allogeneic MSCs were both transplanted into humans to promote the
regeneration of skin defects, and several clinical trials have been conducted involving the ecacy of MSC
transplantation alone or combined with grafts for treating severe skin burns [
215
], perianal fistula [
216
],
non-healing ulcers due to diabetes [
217
], dystrophic epidermolysis bullosa [
218
], and radiation-related
skin lesions [
219
]. Estrogen treatment repairs diabetic wounds by significantly increasing MSC
viability and proliferation and promoting neovascularization [
220
]. For appendage regeneration,
Sheng et al.
[
221
] reported the successful transplantation of MSCs to regenerate functional sweat glands
in five patients in 2009, followed by subsequent reports of this procedure was successfully performed,
with ectodysplasin-modified MSCs [
207
] and BMSCs on EGF-loaded scaolds [
222
] demonstrated to
dierentiate into the functional sweat-gland cells.
Cells 2019, 8, x 19 of 36
was successfully performed, with ectodysplasin-modified MSCs [207] and BMSCs on EGF-loaded
scaffolds [222] demonstrated to differentiate into the functional sweat-gland cells.
Figure 5. Strategies for wound healing: various mesenchymal stem cells (MSCs) are isolated and
identified, and then the cells are augmented and differentiated in a specific culture condition. To
realize the skin regeneration, MSCs secrete numerous factors to modulate inflammation and induce
angiogenesis.
4.11. Other Examples of MSC-Based Therapeutics
Systemic administration of MSCs alters the course of kidney injury via paracrine and/or
endocrine mechanisms [223]. Intratracheal injection of BMSCs was demonstrated as a feasible
strategy for managing lung diseases [30]; another study described the promotion of bladder
regeneration by MSCs seeded onto PCL/CS scaffolds [224]. MSCs loaded into bioprinted
vascularized tissue resulted in thicker tissue than that produced using current bioprinting methods
and capable of surviving for only short periods [225]. Current evidence supports the use of MSCs
for tissue regeneration in a variety of scenarios.
5. Potential Risk of Implanting MSCs
Although a large number of preclinical and clinical studies have been reported, the safety of
MSC-related therapies remains the biggest problem for clinical applications. The principal risks of
MSCs are tumorigenicity, proinflammation, and fibrosis.
Tumorigenicity is one of the most severe risk factors. On the one hand, MSCs have the ability
to develop into tumors, and some studies have shown that Ewing’s sarcoma cells are derived from
MSCs [226]. Additionally, a case of gliomas has occurred four years after stem cell transplantation
for ataxia-telangiectasia, and the tumor cells were shown to be derived from grafts [227]. On the
other hand, MSCs promote the development of tumors. Excess cytokines produced by MSCs, such
Figure 5.
Strategies for wound healing: various mesenchymal stem cells (MSCs) are isolated and
identified, and then the cells are augmented and dierentiated in a specific culture condition.
To realize the skin regeneration, MSCs secrete numerous factors to modulate inflammation and
induce angiogenesis.
4.11. Other Examples of MSC-Based Therapeutics
Systemic administration of MSCs alters the course of kidney injury via paracrine and/or endocrine
mechanisms [
223
]. Intratracheal injection of BMSCs was demonstrated as a feasible strategy for
managing lung diseases [
30
]; another study described the promotion of bladder regeneration by MSCs
seeded onto PCL/CS scaolds [
224
]. MSCs loaded into bioprinted vascularized tissue resulted in
Cells 2019,8, 886 18 of 32
thicker tissue than that produced using current bioprinting methods and capable of surviving for only
short periods [
225
]. Current evidence supports the use of MSCs for tissue regeneration in a variety
of scenarios.
5. Potential Risk of Implanting MSCs
Although a large number of preclinical and clinical studies have been reported, the safety of
MSC-related therapies remains the biggest problem for clinical applications. The principal risks of
MSCs are tumorigenicity, proinflammation, and fibrosis.
Tumorigenicity is one of the most severe risk factors. On the one hand, MSCs have the ability
to develop into tumors, and some studies have shown that Ewing’s sarcoma cells are derived from
MSCs [
226
]. Additionally, a case of gliomas has occurred four years after stem cell transplantation
for ataxia-telangiectasia, and the tumor cells were shown to be derived from grafts [
227
]. On the
other hand, MSCs promote the development of tumors. Excess cytokines produced by MSCs, such as
chemokines and growth factors, directly act on receptors on the surface of cancer cells, thereby
regulating tumor growth. The immunosuppression ability of MSCs contributes to tumor growth and
tumor cell metastasis [
228
,
229
]. Additionally, MSCs have pro-angiogenic functions in the context of
tumor development [230].
MSCs show immunosuppressive eects when exposed to suciently high levels of pro-
inflammatory cytokines. However, they promote inflammatory responses in the presence of low levels
of TNF-
α
and IFN-
γ
[
231
]. It demonstrates that MSCs need to be triggered by inflammatory cytokines
to become immunosuppressants, and the inflammatory environment is a crucial factor aecting the
immune regulation of MSCs.
In addition to tissue repair, MSCs also produce fibrotic reactions. For instance, MSCs can
dierentiate into myofibroblasts [
232
]. Additionally, the balance between repair and fibrosis of MSCs
is broken in the process of injury repair, which leads to fibrotic lung disease [233].
In order to improve the therapeutic eect of MSCs and reduce the potential risks, some measures
should be implemented, such as reducing excessive cytokines, further exploring the immunomodulatory
eects of MSCs, and establishing strict preclinical biosafety testing rules. Fortunately, major adverse
events were rare according to clinical trials evaluation of MSC therapy [
234
]. However, this study only
proves that MSCs are well tolerated and safe in the short term. As the development of cancer is a
continuous process, further longer and larger controlled clinical trials are still necessary to determine
the safety of MSCs.
6. Conclusions and Perspectives
The magic capability of regeneration of damaged parts of the body to regain lost function has long
been a dream of humanity. It has been 50 years since MSCs were first identified, and advancements in
the MSC-based tissue engineering have followed. In recent years, optimizations of extraction, culture,
and dierentiation methods have allowed MSCs to progress closer to clinical applications for disease
therapy and tissue reconstruction. Three MSC properties make them optimal for tissue regeneration:
(1) Immunoregulatory capacity beneficial to alleviate abnormal immune responses, (2) paracrine or
autocrine functions that generate growth factors, and (3) the ability to dierentiate into target cells.
Previous studies of MSC-based regenerative medicine mainly focused on musculoskeletal tissues;
however, recent progress has expanded their applications into other tissues, including the CNS, heart,
liver, cornea, and trachea. Induction factors are one of the most critical factors aecting the outcome of
MSC therapy, which sharply accelerate the repair process of MSCs on tissues. Scaolds provide the
environment for proliferation and dierentiation of MSCs, and produce some mechanical stimulation to
MSCs, which is beneficial for further applications of MSCs. Moreover, scaolds loaded with induction
factors enhance the therapeutic eects of MSCs, which is also worthy of further study. Scaolds and
induction factors remain indispensable agents in these processes; therefore, future investigation of
Cells 2019,8, 886 19 of 32
advanced materials and ecient inducing factors will promote the further applications of MSCs in
regenerative medicine.
Although MSCs have several advantages, there are still many challenges to overcome. The unique
immunomodulatory properties of MSCs are essential for their functions, but the mechanisms of MSC
immune regulation have not been elucidated. Dierent researchers have their distinct methods of
isolating and culturing MSC, although the primary medium is similar. However, dierent culture
conditions, such as FBS, supplements, cell seeding density, and oxygen, may aect cell proliferation
and dierentiation potential [
235
237
]. Therefore, a standard protocol needs to be formulated for
the
in vitro
culture of MSCs. Additionally, cryopreserved MSCs have low viability, which will aect
the further applications of cells [
238
]. Moreover, the age of the donors also aects the proliferation
and dierentiation potential of MSCs, and MSCs from young donors show lower damage and better
proliferation [
239
]. We conclude that many factors influence the therapeutic potential of MSCs, such as
induction factors, oxygen concentrations, and mechanical stimuli. Therefore, optimizing the culture
conditions of MSCs may be an eective means to improve the therapeutic potential of MSCs to achieve
tissue repair successfully.
For clinical applications, autologous and allogeneic MSCs have both been reported as sources
for tissue regeneration. Specifically, autologous MSCs represent the primary sources considered
safe for transplantation and minimization of immunological risk, despite the lack of documented
complaints regarding allogeneic MSC-based therapy. The eect of MSCs on human clinical outcomes
has not readily achieved the predominantly positive outcomes in murine. Furthermore, the oncogenic
potential of uncontrolled MSC dierentiation needs to be further investigated. The dierentiation
potential, surface markers, and transcription of various tissue-derived MSCs are challenging to be
unified, which undoubtedly become a hindrance to the clinical transformation of MSCs.
In the clinic, the eect of MSCs on human clinical outcomes has not readily achieved
the predominantly positive outcome in murine. The dierent results mainly attribute to the
immunocompatibility and fitness of MSCs. Expanding the indications of diseases and reducing
the dierences among dierent individuals are challenges in the future research of mesenchymal stem
cells. Further research is required on cell physiology about how MSCs function
in vivo
and how to
achieve accurate administration.
Despite the current challenges, MSC-based tissue engineering represents a promising clinical
strategy in the field of regenerative medicine. Moreover, improving the cultural environment of MSCs
and selecting appropriate scaolds and induction factors are essential for MSC therapy.
Author Contributions:
Conceptualization, Y.Z., F.C., and J.D.; Writing—original draft preparation, Y.H. and X.L.;
Writing—review and editing, Y.Z., Y.P.H., and J.D.; Funding acquisition, F.C. and J.D.
Funding:
This work was finically supported by the National Natural Science Foundation of China (Grant Nos.
51873207, 51803006, 81701811, 51673190, and 81671804), the Development and Reform Commission of Jilin
Province (Grant No. 2018C052-4), the Health and Family Planning Commission of Jilin Province (Grant No.
20165061), the Science and Technology Development Project of Jilin Province (Grant No. 20190201068JC), the Youth
Talents Promotion Project of Jilin Province (Grant No. 181909), the Youth Innovation Promotion Association of
Chinese Academy of Sciences (Grant No. 2019005), and the Natural Science Foundation of Shandong Province
(Grant No. ZR2018BH008).
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Friedenstein, A.J.; Gorskaja, U.F.; Kulagina, N.N. Fibroblast precursors in normal and irradiated mouse
hematopoietic organs. Exp. Hematol. 1976,4, 267–274. [PubMed]
2.
Ashton, B.A.; Allen, T.D.; Howlett, C.; Eaglesom, C.; Hattori, A.; Owen, M. Formation of bone and cartilage by
marrow stromal cells in diusion chambers
in vivo
.Clin. Orthop. Relat. Res.
1980
,151, 294–307. [CrossRef]
3.
Owen, M.; Friedenstein, A. Stromal stem cells: Marrow-derived osteogenic precursors. Ciba. Found Symp.
1988,136, 42–60. [PubMed]
Cells 2019,8, 886 20 of 32
4.
Salgado, A.J.; Oliveira, J.M.; Martins, A.; Teixeira, F.G.; Silva, N.A.; Neves, N.M.; Sousa, N.; Reis, R.L. Tissue
engineering and regenerative medicine: Past, present, and future. Int. Rev. Neurobiol.
2013
,108, 1–33.
[PubMed]
5.
Campagnoli, C.; Roberts, I.A.G.; Kumar, S.; Bennett, P.R.; Bellantuono, I.; Fisk, N.M. Identification of
mesenchymal stem/progenitor cells in human first-trimester fetal blood, liver, and bone marrow. Blood
2001
,
98, 2396–2402. [CrossRef] [PubMed]
6.
Chen, Y.; Shao, J.Z.; Xiang, L.X.; Dong, X.J.; Zhang, G.R. Mesenchymal stem cells: A promising candidate in
regenerative medicine. Int. J. Biochem. Cell Biol. 2008,40, 815–820. [CrossRef]
7.
Gao, F.; Chiu, S.M.; Motan, D.A.L.; Zhang, Z.; Chen, L.; Ji, H.L.; Tse, H.F.; Fu, Q.L.; Lian, Q. Mesenchymal
stem cells and immunomodulation: Current status and future prospects. Cell Death Dis.
2016
,7, e2062.
[CrossRef]
8.
Aggarwal, S.; Pittenger, M.F. Human mesenchymal stem cells modulate allogeneic immune cell responses.
Blood 2005,105, 1815–1822. [CrossRef]
9.
Erices, A.; Conget, P.; Minguell, J.J. Mesenchymal progenitor cells in human umbilical cord blood. Br. J.
Haematol. 2000,109, 235–242. [CrossRef]
10. Zuk, P.A.; Zhu, M.; Mizuno, H.; Huang, J.; Futrell, J.W.; Katz, A.J.; Benhaim, P.; Lorenz, H.P.; Hedrick, M.H.
Multilineage cells from human adipose tissue: Implications for cell-based therapies. Tissue Eng.
2001
,7,
211–228. [CrossRef]
11.
De Bari, C.; Dell’Accio, F.; Tylzanowski, P.; Luyten, F.P. Multipotent mesenchymal stem cells from adult
human synovial membrane. Arthritis Rheum. 2001,44, 1928–1942. [CrossRef]
12.
Soleimani, M.; Nadri, S. A protocol for isolation and culture of mesenchymal stem cells from mouse bone
marrow. Nat. Protoc. 2009,4, 102–106. [CrossRef]
13.
Zhu, H.; Guo, Z.K.; Jiang, X.X.; Li, H.; Wang, X.Y.; Yao, H.Y.; Zhang, Y.; Mao, N. A protocol for isolation and
culture of mesenchymal stem cells from mouse compact bone. Nat. Protoc. 2010,5, 550–560. [CrossRef]
14.
El-Ansary, M.; Abdel-Aziz, I.; Mogawer, S.; Abdel-Hamid, S.; Hammam, O.; Teaema, S.; Wahdan, M. Phase II
trial: Undierentiated versus dierentiated autologous mesenchymal stem cells transplantation in egyptian
patients with HCV induced liver cirrhosis. Stem Cell Rev. Rep. 2012,8, 972–981. [CrossRef]
15.
Mao, P.; Joshi, K.; Li, J.F.; Kim, S.H.; Li, P.P.; Santana-Santos, L.; Luthra, S.; Chandran, U.R.; Benos, P.V.;
Smith, L.; et al. Mesenchymal glioma stem cells are maintained by activated glycolytic metabolism involving
aldehyde dehydrogenase 1A3. Proc. Natl. Acad. Sci. USA 2013,110, 8644–8649. [CrossRef]
16.
Caniglia, C.J.; Schramme, M.C.; Smith, R.K. The eect of intralesional injection of bone marrow derived
mesenchymal stem cells and bone marrow supernatant on collagen fibril size in a surgical model of equine
superficial digital flexor tendonitis. Equine Vet. J. 2012,44, 587–593. [CrossRef]
17.
Davies, L.C.; Lonnies, H.; Locke, M.; Sundberg, B.; Rosendahl, K.; Gotherstrom, C.; Le Blanc, K.; Stephens, P.
Oral mucosal progenitor cells are potently immunosuppressive in a dose-independent manner. Stem Cells
Dev. 2012,21, 1478–1487. [CrossRef]
18.
Herrmann, R.; Sturm, M.; Shaw, K.; Purtill, D.; Cooney, J.; Wright, M.; Phillips, M.; Cannell, P. Mesenchymal
stromal cell therapy for steroid-refractory acute and chronic graft versus host disease: A phase 1 study. Int. J.
Hematol. 2012,95, 182–188. [CrossRef]
19.
Zuk, P.A.; Zhu, M.; Ashjian, P.; De Ugarte, D.A.; Huang, J.I.; Mizuno, H.; Alfonso, Z.C.; Fraser, J.K.;
Benhaim, P.; Hedrick, M.H. Human adipose tissue is a source of multipotent stem cells. Mol. Biol. Cell
2002
,
13, 4279–4295. [CrossRef]
20.
Smadja, D.M.; d’Audigier, C.; Guerin, C.L.; Mauge, L.; Dizier, B.; Silvestre, J.S.; Dal Cortivo, L.; Gaussem, P.;
Emmerich, J. Angiogenic potential of BM MSCs derived from patients with critical leg ischemia. Bone Marrow
Transplant. 2012,47, 997–1000. [CrossRef]
21.
Connick, P.; Kolappan, M.; Crawley, C.; Webber, D.J.; Patani, R.; Michell, A.W.; Du, M.Q.; Luan, S.L.;
Altmann, D.R.; Thompson, A.J.; et al. Autologous mesenchymal stem cells for the treatment of secondary
progressive multiple sclerosis: An open-label phase 2a proof-of-concept study. Lancet Neurol.
2012
,11,
150–156. [CrossRef]
22.
Doorn, J.; Siddappa, R.; van Blitterswijk, C.A.; de Boer, J. Forskolin enhances
in vivo
bone formation by
human mesenchymal stromal cells. Tissue Eng. Part A. 2012,18, 558–567. [CrossRef]
Cells 2019,8, 886 21 of 32
23.
Higuera, G.A.; Schop, D.; Spitters, T.W.G.M.; van Dijkhuizen-Radersma, R.; Bracke, M.; de Bruijn, J.D.;
Martens, D.; Karperien, M.; van Boxtel, A.; van Blitterswijk, C.A. Patterns of amino acid metabolism by
proliferating human mesenchymal stem cells. Tissue Eng. Part A. 2012,18, 654–664. [CrossRef]
24.
Zhao, D.W.; Cui, D.P.; Wang, B.J.; Tian, F.D.; Guo, L.; Yang, L.; Liu, B.Y.; Yu, X.B. Treatment of early stage
osteonecrosis of the femoral head with autologous implantation of bone marrow-derived and cultured
mesenchymal stem cells. Bone 2012,50, 325–330. [CrossRef]
25.
Futami, I.; Ishijima, M.; Kaneko, H.; Tsuji, K.; Ichikawa-Tomikawa, N.; Sadatsuki, R.; Muneta, T.;
Arikawa-Hirasawa, E.; Sekiya, I.; Kaneko, K. Isolation and characterization of multipotential mesenchymal
cells from the mouse synovium. PloS ONE 2012,7, e45517. [CrossRef]
26.
Kim, M.J.; Shin, K.S.; Jeon, J.H.; Lee, D.R.; Shim, S.H.; Kim, J.K.; Cha, D.H.; Yoon, T.K.; Kim, G.J. Human
chorionic-plate-derived mesenchymal stem cells and Wharton’s jelly-derived mesenchymal stem cells:
A comparative analysis of their potential as placenta-derived stem cells. Cell Tissue Res.
2011
,346, 53–64.
[CrossRef]
27.
Marx, R.E.; Harrell, D.B. Translational research: The CD34
+
cell is crucial for large-volume bone regeneration
from the milieu of bone marrow progenitor cells in craniomandibular reconstruction. Int. J. Oral. Maxillofac.
Implants 2014,24, e201–209. [CrossRef]
28.
Sibov, T.T.; Severino, P.; Marti, L.C.; Pavon, L.F.; Oliveira, D.M.; Tobo, P.R.; Campos, A.H.; Paes, A.T.;
Amaro, E.; Gamarra, L.F.; et al. Mesenchymal stem cells from umbilical cord blood: Parameters for isolation,
characterization and adipogenic dierentiation. Cytotechnology 2012,64, 511–521. [CrossRef]
29.
Gao, L.R.; Pei, X.T.; Ding, Q.A.; Chen, Y.; Zhang, N.K.; Chen, H.Y.; Wang, Z.G.; Wang, Y.F.; Zhu, Z.M.;
Li, T.C.; et al. A critical challenge: Dosage-related ecacy and acute complication intracoronary injection of
autologous bone marrow mesenchymal stem cells in acute myocardial infarction. Int. J. Cardiol.
2013
,168,
3191–3199. [CrossRef]
30.
Dai, R.R.; Yu, Y.C.; Yan, G.F.; Hou, X.X.; Ni, Y.M.; Shi, G.C. Intratracheal administration of adipose derived
mesenchymal stem cells alleviates chronic asthma in a mouse model. BMC Pulm. Med.
2018
,18, 131.
[CrossRef]
31.
Yoshimura, H.; Muneta, T.; Nimura, A.; Yokoyama, A.; Koga, H.; Sekiya, I. Comparison of rat mesenchymal
stem cells derived from bone marrow, synovium, periosteum, adipose tissue, and muscle. Cell Tissue Res.
2007,327, 449–462. [CrossRef]
32.
Choudhery, M.S.; Badowski, M.; Muise, A.; Harris, D.T. Comparison of human mesenchymal stem cells
derived from adipose and cord tissue. Cytotherapy 2013,15, 330–343. [CrossRef]
33.
Park, J.S.; Shim, M.S.; Shim, S.H.; Yang, H.N.; Jeon, S.Y.; Woo, D.G.; Lee, D.R.; Yoon, T.K.; Park, K.H.
Chondrogenic potential of stem cells derived from amniotic fluid, adipose tissue, or bone marrow encapsulated
in fibrin gels containing TGF-β3. Biomaterials 2011,32, 8139–8149. [CrossRef]
34.
Pievani, A.; Scagliotti, V.; Russo, F.M.; Azario, I.; Rambaldi, B.; Sacchetti, B.; Marzorati, S.; Erba, E.; Giudici, G.;
Riminucci, M.; et al. Comparative analysis of multilineage properties of mesenchymal stromal cells derived
from fetal sources shows an advantage of mesenchymal stromal cells isolated from cord blood in chondrogenic
dierentiation potential. Cytotherapy 2014,16, 893–905. [CrossRef]
35.
Chung, C.S.; Fujita, N.; Kawahara, N.; Yui, S.; Nam, E.; Nishimura, R. A comparison of neurosphere
dierentiation potential of canine bone marrow-derived mesenchymal stem cells and adipose-derived
mesenchymal stem cells. J. Vet. Med. Sci. 2013,75, 879–886. [CrossRef]
36.
Bae, K.S.; Park, J.B.; Kim, H.S.; Kim, D.S.; Park, D.J.; Kang, S.J. Neuron-like dierentiation of bone
marrow-derived mesenchymal stem cells. Yonsei. Med. J. 2011,52, 401–412. [CrossRef]
37.
Hou, L.L.; Cao, H.; Wang, D.M.; Wei, G.R.; Bai, C.X.; Zhang, Y.; Pei, X.T. Induction of umbilical cord blood
mesenchymal stem cells into neuron-like cells in vitro. Int. J. Hematol. 2003,78, 256–261. [CrossRef]
38.
Lei, H.; Yu, B.; Huang, Z.; Yang, X.; Liu, Z.; Mao, X.; Tian, G.; He, J.; Han, G.; Chen, H.; et al. Comparative
analysis of mesenchymal stem cells from adult mouse adipose, muscle, and fetal muscle. Mol. Biol. Rep.
2013,40, 885–892. [CrossRef]
39.
Xu, W.; Zhang, X.; Qian, H.; Zhu, W.; Sun, X.; Hu, J.; Zhou, H.; Chen, Y. Mesenchymal stem cells from
adult human bone marrow dierentiate into a cardiomyocyte phenotype
in vitro
.Exp. Biol. Med.
2004
,229,
623–631. [CrossRef]
Cells 2019,8, 886 22 of 32
40.
Yang, J.; Song, T.; Wu, P.; Chen, Y.; Fan, X.; Chen, H.; Zhang, J.; Huang, C. Dierentiation potential of human
mesenchymal stem cells derived from adipose tissue and bone marrow to sinus node-like cells. Mol. Med.
Rep. 2012,5, 108–113.
41.
Yin, L.B.; Zhu, Y.H.; Yang, J.G.; Ni, Y.J.; Zhou, Z.; Chen, Y.; Wen, L.X. Adipose tissue-derived mesenchymal
stem cells dierentiated into hepatocyte-like cells
in vivo
and
in vitro
.Mol. Med. Rep.
2015
,11, 1722–1732.
[CrossRef]
42.
Banas, A.; Teratani, T.; Yamamoto, Y.; Tokuhara, M.; Takeshita, F.; Quinn, G.; Okochi, H.; Ochiya, T. Adipose
tissue-derived mesenchymal stem cells as a source of human hepatocytes. Hepatology
2007
,46, 219–228.
[CrossRef]
43.
Zhou, R.P.; Li, Z.K.; He, C.Y.; Li, R.L.; Xia, H.B.; Li, C.Y.; Xiao, J.; Chen, Z.Y. Human umbilical cord
mesenchymal stem cells and derived Hepatocyte-like cells exhibit similar therapeutic eects on an acute
liver failure mouse model. PloS ONE 2014,9, e104392. [CrossRef]
44.
Lee, H.J.; Jung, J.; Cho, K.J.; Lee, C.K.; Hwang, S.G.; Kim, G.J. Comparison of
in vitro
hepatogenic
dierentiation potential between various placenta-derived stem cells and other adult stem cells as an
alternative source of functional hepatocytes. Dierentiation 2012,84, 223–231. [CrossRef]
45.
Katikireddy, K.R.; Dana, R.; Jurkunas, U.V. Dierentiation potential of limbal fibroblasts and bone marrow
mesenchymal stem cells to corneal epithelial cells. Stem Cells 2014,32, 717–729. [CrossRef]
46.
Rohaina, C.M.; Then, K.Y.; Ng, A.M.H.; Halim, W.H.W.A.; Zahidin, A.Z.M.; Saim, A.; Idrus, R.B.H.
Reconstruction of limbal stem cell deficient corneal surface with induced human bone marrow mesenchymal
stem cells on amniotic membrane. Transl. Res. 2014,163, 200–210. [CrossRef]
47.
Andrzejewska, A.; Lukomska, B.; Janowski, M. Concise review: Mesenchymal stem cells: from roots to boost.
Stem Cells. 2019,37, 855–864. [CrossRef]
48.
Lin, H.; Sohn, J.; Shen, H.; Langhans, M.T.; Tuan, R.S. Bone marrow mesenchymal stem cells: Aging and
tissue engineering applications to enhance bone healing. Biomaterials 2018,203, 96–110. [CrossRef]
49.
Mochizuki, T.; Muneta, T.; Sakaguchi, Y.; Nimura, A.; Yokoyama, A.; Koga, H.; Sekiya, I. Higher chondrogenic
potential of fibrous synovium- and adipose synovium-derived cells compared with subcutaneous fat-derived
cells: Distinguishing properties of mesenchymal stem cells in humans. Arthritis Rheum.
2006
,54, 843–853.
[CrossRef]
50.
Jin, H.J.; Bae, Y.K.; Kim, M.; Kwon, S.J.; Jeon, H.B.; Choi, S.J.; Kim, S.W.; Yang, Y.S.; Oh, W.; Chang, J.W.
Comparative analysis of human mesenchymal stem cells from bone marrow, adipose tissue, and umbilical
cord blood as sources of cell therapy. Int. J. Mol. Sci. 2013,14, 17986–18001. [CrossRef]
51.
Younger, E.M.; Chapman, M.W. Morbidity at bone graft donor sites. J. Orthop. Trauma.
1989
,3, 192–195.
[CrossRef]
52.
Lozano-Calderon, S.A.; Swaim, S.O.; Federico, A.; Anderson, M.E.; Gebhardt, M.C. Predictors of soft-tissue
complications and deep infection in allograft reconstruction of the proximal tibia. J. Surg. Oncol.
2016
,113,
811–817. [CrossRef]
53.
Jiang, Y.H.; Jahagirdar, B.N.; Reinhardt, R.L.; Schwartz, R.E.; Keene, C.D.; Ortiz-Gonzalez, X.R.; Reyes, M.;
Lenvik, T.; Lund, T.; Blackstad, M.; et al. Pluripotency of mesenchymal stem cells derived from adult marrow.
Nature 2002,418, 41–49. [CrossRef]
54.
Kargozar, S.; Mozafari, M.; Hashemian, S.J.; Milan, P.B.; Hamzehlou, S.; Soleimani, M.; Joghataei, M.T.;
Gholipourmalekabadi, M.; Korourian, A.; Mousavizadeh, K.; et al. Osteogenic potential of stem cells-seeded
bioactive nanocomposite scaolds: A comparative study between human mesenchymal stem cells derived
from bone, umbilical cord Wharton’s jelly, and adipose tissue. J. Biomed. Mater. Res. B
2018
,106, 61–72.
[CrossRef]
55.
Yang, J.Z.; Zhang, Y.S.; Yue, K.; Khademhosseini, A. Cell-laden hydrogels for osteochondral and cartilage
tissue engineering. Acta Biomater. 2017,57, 1–25. [CrossRef]
56.
Bougioukli, S.; Sugiyama, O.; Pannell, W.; Ortega, B.; Tan, M.H.; Tang, A.H.; Yoho, R.; Oakes, D.A.;
Lieberman, J.R. Genetherapy for bone repair using human cells: Superior osteogenic potential of bone
morphogenetic protein 2-transduced mesenchymal stem cells derived from adipose tissue compared to bone
marrow. Hum. Gene Ther. 2018,29, 507–519. [CrossRef]
57.
Burrow, K.L.; Hoyland, J.A.; Richardson, S.M. Human adipose-derived stem cells exhibit enhanced
proliferative capacity and retain multipotency longer than donor-matched bone marrow mesenchymal stem
cells during expansion in vitro. Stem Cells Int. 2017,15, 2541275. [CrossRef]
Cells 2019,8, 886 23 of 32
58.
Ye, K.Q.; Liu, D.H.; Kuang, H.Z.; Cai, J.Y.; Chen, W.M.; Sun, B.B.; Xia, L.G.; Fang, B.; Morsi, Y.; Mo, X.M.
Three-dimensional electrospun nanofibrous scaolds displaying bone morphogenetic protein-2-derived
peptides for the promotion of osteogenic dierentiation of stem cells and bone regeneration. J. Colloid Interface
Sci. 2019,534, 625–636. [CrossRef]
59.
Aragon, J.; Salerno, S.; De Bartolo, L.; Irusta, S.; Mendoza, G. Polymeric electrospun scaolds for bone
morphogenetic protein 2 delivery in bone tissue engineering. J. Colloid Interface Sci.
2018
,531, 126–137.
[CrossRef]
60.
Decambron, A.; Devriendt, N.; Larochette, N.; Manassero, M.; Bourguignon, M.; El-Hafci, H.; Petite, H.;
Viateau, V.; Logeart-avramoglou, D. Eect of the bone morphogenetic protein-2 doses on the osteogenic
potential of human multipotent stromal cells- containing tissue engineered constructs. Tissue Eng. Part A
2019,25, 642–651. [CrossRef]
61.
Chen, F.; Bi, D.; Cheng, C.; Ma, S.; Liu, Y.; Cheng, K. Bone morphogenetic protein 7 enhances the osteogenic
dierentiation of human dermal-derived CD105
+
fibroblast cells through the Smad and MAPK pathways.
Int. J. Mol. Med. 2019,43, 37–46. [CrossRef]
62.
Zhu, J.H.; Liao, Y.P.; Li, F.S.; Hu, Y.; Li, Q.; Ma, Y.; Wang, H.; Zhou, Y.; He, B.C.; Su, Y.X. Wnt11 promotes
BMP9-induced osteogenic dierentiation through BMPs/Smads and p38 MAPK in mesenchymal stem cells.
J. Cell Biochem. 2018,119, 9462–9473. [CrossRef]
63.
Bizelli-Silveira, C.; Pullisaar, H.; Abildtrup, L.A.; Andersen, O.Z.; Spin-Neto, R.; Foss, M.; Kraft, D.C.E.
Strontium enhances proliferation and osteogenic behavior of periodontal ligament cells
in vitro
.J. Periodont.
Res. 2018,53, 1020–1028. [CrossRef]
64.
Tang, D.; Tare, R.S.; Yang, L.Y.; Williams, D.F.; Ou, K.L.; Oreo, R.O.C. Biofabrication of bone tissue:
Approaches, challenges and translation for bone regeneration. Biomaterials 2016,83, 363–382. [CrossRef]
65.
Westhauser, F.; Senger, A.S.; Reible, B.; Moghaddam, A.
In vivo
models for the evaluation of the osteogenic
potency of bone substitutes seeded with mesenchymal stem cells of human origin: A concise review. Tissue
Eng. Part C 2017,23, 881–888. [CrossRef]
66.
Khojasteh, A.; Fahimipour, F.; Jafarian, M.; Sharifi, D.; Jahangir, S.; Khayyatan, F.; Eslaminejad, M.B. Bone
engineering in dog mandible: Coculturing mesenchymal stem cells with endothelial progenitor cells in
a composite scaold containing vascular endothelial growth factor. J. Biomed. Mater. Res. B
2017
,105,
1767–1777. [CrossRef]
67.
Katagiri, W.; Watanabe, J.; Toyama, N.; Osugi, M.; Sakaguchi, K.; Hibi, H. Clinical study of bone regeneration
by conditioned medium from mesenchymal stem cells after maxillary sinus floor elevation. Implant. Dent.
2017,26, 607–612. [CrossRef]
68.
Gjerde, C.; Mustafa, K.; Hellem, S.; Rojewski, M.; Gjengedal, H.; Yassin, M.A.; Feng, X.; Skaale, S.; Berge, T.;
Rosen, A.; et al. Cell therapy induced regeneration of severely atrophied mandibular bone in a clinical trial.
Stem Cell Res. Ther. 2018,9, 213. [CrossRef]
69.
Liebergall, M.; Schroeder, J.; Moshei, R.; Gazit, Z.; Yoram, Z.; Rasooly, L.; Daskal, A.; Khoury, A.; Weil, Y.;
Beyth, S. Stem cell-based therapy for prevention of delayed fracture union: A randomized and prospective
preliminary study. Mol. Ther. 2013,21, 1631–1638. [CrossRef]
70.
Giannotti, S.; Trombi, L.; Bottai, V.; Ghilardi, M.; D’Alessandro, D.; Danti, S.; Dell’Osso, G.; Guido, G.;
Petrini, M. Use of autologous human mesenchymal stromal cell/fibrin clot constructs in upper limb non-unions:
Long-term assessment. PloS ONE 2013,8, e73893. [CrossRef]
71.
Honarpardaz, A.; Irani, S.; Pezeshki-Modaress, M.; Zandi, M.; Sadeghi, A. Enhanced chondrogenic
dierentiation of bone marrow mesenchymal stem cells on gelatin/glycosaminoglycan electrospun nanofibers
with dierent amount of glycosaminoglycan. J. Biomed. Mater. Res. Part A 2019,107, 38–48. [CrossRef]
72.
Fernandes, T.L.; Kimura, H.A.; Pinheiro, C.C.G.; Shimomura, K.; Nakamura, N.; Ferreira, J.R.M.; Gomoll, A.H.;
Hernandez, A.J.; Bueno, D.F. Human synovial mesenchymal stem cells good manufacturing practices for
articular cartilage regeneration. Tissue Eng. Part C 2018,24, 709–716. [CrossRef]
73.
Zou, J.Y.; Bai, B.; Yao, Y.C. Progress of co-culture systems in cartilage regeneration. Expert. Opin. Biol. Ther.
2018,18, 1151–1158. [CrossRef]
74.
Girao, A.F.; Semitela, A.; Ramalho, G.; Completo, A.; Marques, P. Mimicking nature: Fabrication of 3D
anisotropic electrospun polycaprolactone scaolds for cartilage tissue engineering applications. Composites,
Part B 2018,154, 99–107. [CrossRef]
Cells 2019,8, 886 24 of 32
75.
Liu, M.; Zeng, X.; Ma, C.; Yi, H.; Ali, Z.; Mou, X.B.; Li, S.; Deng, Y.; He, N.Y. Injectable hydrogels for cartilage
and bone tissue engineering. Bone Res. 2017,5, 20. [CrossRef]
76.
Legendre, F.; Ollitrault, D.; Gomez-Leduc, T.; Bouyoucef, M.; Hervieu, M.; Gruchy, N.; Mallein-Gerin, F.;
Leclercq, S.; Demoor, M.; Galera, P. Enhanced chondrogenesis of bone marrow-derived stem cells by using
a combinatory cell therapy strategy with BMP-2/TGF-
β
1, hypoxia, and COL1A1/HtrA1 siRNAs. Sci. Rep.
2017,7, 3406. [CrossRef]
77.
Crecente-Campo, J.; Borrajo, E.; Vidal, A.; Garcia-Fuentes, M. New scaolds encapsulating TGF-
β
3/BMP-7
combinations driving strong chondrogenic dierentiation. Eur. J. Pharm. Biopharm.
2017
,114, 69–78.
[CrossRef]
78.
Gugjoo, M.B.; Amarpal; Abdelbaset-Ismail, A.; Aithal, H.P.; Kinjavdekar, P.; Pawde, A.M.; Kumar, G.S.;
Sharma, G.T. Mesenchymal stem cells with IGF-1 and TGF-
β
1 in laminin gel for osteochondral defects in
rabbits. Biomed. Pharmacother. 2017,93, 1165–1174. [CrossRef]
79.
Fahy, N.; Alini, M.; Stoddart, M.J. Mechanical stimulation of mesenchymal stem cells: Implications for
cartilage tissue engineering. J. Orthop. Res. 2018,36, 52–63. [CrossRef]
80.
Wakitani, S.; Goto, T.; Pineda, S.J.; Young, R.G.; Mansour, J.M.; Caplan, A.I.; Goldberg, V.M. Mesenchymal
cell-based repair of large, full-thickness defects of articular cartilage. J. Bone Joint Surg. Am.
1994
,76, 579–592.
[CrossRef]
81.
de Windt, T.S.; Vonk, L.A.; Slaper-Cortenbach, I.C.; van den Broek, M.P.; Nizak, R.; van Rijen, M.H.; de
Weger, R.A.; Dhert, W.J.; Saris, D.B. Allogeneic mesenchymal stem cells stimulate cartilage regeneration and
are safe for single-stage cartilage repair in humans upon mixture with recycled autologous chondrons. Stem
Cells 2017,35, 256–264. [CrossRef]
82.
Wakitani, S.; Imoto, K.; Yamamoto, T.; Saito, M.; Murata, N.; Yoneda, M. Human autologous culture
expanded bone marrow mesenchymal cell transplantation for repair of cartilage defects in osteoarthritic
knees. Osteoarthr. Cartilage 2002,10, 199–206. [CrossRef]
83.
Orozco, L.; Munar, A.; Soler, R.; Alberca, M.; Soler, F.; Huguet, M.; Sentis, J.; Sanchez, A.; Garcia-Sancho, J.
Treatment of knee osteoarthritis with autologous mesenchymal stem cells: A pilot study. Transplantation
2013,95, 1535–1541. [CrossRef]
84.
Shapiro, S.A.; Kazmerchak, S.E.; Heckman, M.G.; Zubair, A.C.; O’Connor, M.I. A prospective, single-blind,
placebo-controlled trial of bone marrow aspirate concentrate for knee osteoarthritis. Am. J. Sports Med.
2017
,
45, 82–90. [CrossRef]
85.
Sasaki, H.; Rothrau, B.B.; Alexander, P.G.; Lin, H.; Gottardi, R.; Fu, F.H.; Tuan, R.S.
In vitro
repair of
meniscal radial tear with hydrogels seeded with adipose stem cells and TGF-
β
3. Am. J. Sports Med.
2018
,46,
2402–2413. [CrossRef]
86.
Shimomura, K.; Rothrau, B.B.; Hart, D.A.; Hamamoto, S.; Kobayashi, M.; Yoshikawa, H.; Tuan, R.S.;
Nakamura, N. Enhanced repair of meniscal hoop structure injuries using an aligned electrospun nanofibrous
scaold combined with a mesenchymal stem cell-derived tissue engineered construct. Biomaterials
2018
,192,
346–354. [CrossRef]
87.
Liang, Y.; Idrees, E.; Szojka, A.R.A.; Andrews, S.H.J.; Kunze, M.; Mulet-Sierra, A.; Jomha, N.M.; Adesida, A.B.
Chondrogenic dierentiation of synovial fluid mesenchymal stem cells on human meniscus-derived
decellularized matrix requires exogenous growth factors. Acta Biomater. 2018,80, 131–143. [CrossRef]
88.
Toratani, T.; Nakase, J.; Numata, H.; Oshima, T.; Takata, Y.; Nakayama, K.; Tsuchiya, H. Scaold-free
tissue-engineered allogenic adipose-derived stem cells promote meniscus healing. Arthroscopy
2017
,33,
346–354. [CrossRef]
89.
Tarafder, S.; Gulko, J.; Sim, K.H.; Yang, J.; Cook, J.L.; Lee, C.H. Engineered healing of avascular meniscus
tears by stem cell recruitment. Sci. Rep. 2018,8, 8150. [CrossRef]
90.
Chen, M.X.; Guo, W.M.; Gao, S.; Hao, C.X.; Shen, S.; Zhang, Z.Z.; Wang, Z.H.; Li, X.; Jing, X.G.; Zhang, X.L.;
et al. Biomechanical stimulus based strategies for meniscus tissue engineering and regeneration. Tissue Eng.
Part B 2018,24, 392–402. [CrossRef]
91.
Chew, E.; Prakash, R.; Khan, W. Mesenchymal stem cells in human meniscal regeneration: A systematic
review. Ann. Med. Surg. 2017,24, 3–7. [CrossRef]
92.
Clanton, T.O.; Coupe, K.J. Hamstring strains in athletes: Diagnosis and treatment. J. Am. Acad. Orthop. Surg.
1998,6, 237–248. [CrossRef]
Cells 2019,8, 886 25 of 32
93.
Wang, D.; Jiang, X.H.; Lu, A.Q.; Tu, M.; Huang, W.; Huang, P. BMP14 induces tenogenic dierentiation of
bone marrow mesenchymal stem cells in vitro. Exp. Ther. Med. 2018,16, 1165–1174. [CrossRef]
94.
Aktas, E.; Chamberlain, C.S.; Saether, E.E.; Duenwald-Kuehl, S.E.; Kondratko-Mittnacht, J.; Stitgen, M.;
Lee, J.S.; Clements, A.E.; Murphy, W.L.; Vanderby, R. Immune modulation with primed mesenchymal stem
cells delivered via biodegradable scaold to repair an Achilles tendon segmental defect. J. Orthop. Res.
2017
,
35, 269–280. [CrossRef]
95.
Park, S.H.; Choi, Y.J.; Moon, S.W.; Lee, B.H.; Shim, J.H.; Cho, D.W.; Wang, J.H. Three-dimensional bio-printed
scaold sleeves with mesenchymal stem cells for enhancement of tendon-to-bone healing in anterior cruciate
ligament reconstruction using soft-tissue tendon graft. Arthroscopy 2018,34, 166–179. [CrossRef]
96.
Wang, T.; Thien, C.; Wang, C.; Ni, M.; Gao, J.J.; Wang, A.; Jiang, Q.; Tuan, R.S.; Zheng, Q.J.; Zheng, M.H. 3D
uniaxial mechanical stimulation induces tenogenic dierentiation of tendon-derived stem cells through a
PI3K/AKT signaling pathway. FASEB J. 2018,32, 4804–4814. [CrossRef]
97.
Gulecyuz, M.F.; Macha, K.; Pietschmann, M.F.; Ficklscherer, A.; Sievers, B.; Rossbach, B.P.; Jansson, V.;
Muller, P.E. Allogenic myocytes and mesenchymal stem cells partially improve fatty rotator cudegeneration
in a rat model. Stem Cell Rev. Rep. 2018,14, 847–859. [CrossRef]
98.
Bertolo, A.; Mehr, M.; Aebli, N.; Baur, M.; Ferguson, S.J.; Stoyanov, J.V. Influence of dierent commercial
scaolds on the
in vitro
dierentiation of human mesenchymal stem cells to nucleus pulposus-like cells. Eur.
Spine J. 2012,21 Suppl 6, S826–838. [CrossRef]
99.
Zhou, X.; Wang, J.; Huang, X.; Fang, W.; Tao, Y.; Zhao, T.; Liang, C.; Hua, J.; Chen, Q.; Li, F. Injectable
decellularized nucleus pulposus-based cell delivery system for dierentiation of adipose-derived stem cells
and nucleus pulposus regeneration. Acta Biomater. 2018,81, 115–128. [CrossRef]
100.
Hou, Y.; Liang, L.; Shi, G.D.; Shi, J.G.; Sun, J.C.; Xu, G.H.; Guo, Y.F.; Yuan, W.
In vivo
evaluation of
the adenovirus-mediated Sox9 and BMP2 double gene transduction on treatment of intervertebral disc
degeneration in rabbit annular puncture model. J. Biomater. Tissue Eng. 2018,8, 1091–1099. [CrossRef]
101.
Varma, D.M.; DiNicolas, M.S.; Nicoll, S.B. Injectable, redox-polymerized carboxymethylcellulose hydrogels
promote nucleus pulposus-like extracellular matrix elaboration by human MSCs in a cell density-dependent
manner. J. Biomater. Appl. 2018,33, 576–589. [CrossRef]
102.
Kraus, P.; Lufkin, T. Implications for a stem cell regenerative medicine based approach to human intervertebral
disk degeneration. Front. Cell Dev. Biol. 2017,5, 17. [CrossRef]
103.
Orozco, L.; Soler, R.; Morera, C.; Alberca, M.; Sanchez, A.; Garcia-Sancho, J. Intervertebral disc repair by
autologous mesenchymal bone marrow cells: A pilot study. Transplantation 2011,92, 822–828. [CrossRef]
104.
Pettine, K.A.; Suzuki, R.K.; Sand, T.T.; Murphy, M.B. Autologous bone marrow concentrate intradiscal
injection for the treatment of degenerative disc disease with three-year follow-up. Int. Orthop.
2017
,41,
2097–2103. [CrossRef]
105.
Kumar, H.; Ha, D.H.; Lee, E.J.; Park, J.H.; Shim, J.H.; Ahn, T.K.; Kim, K.T.; Ropper, A.E.; Sohn, S.;
Kim, C.H.; et al. Safety and tolerability of intradiscal implantation of combined autologous adipose-derived
mesenchymal stem cells and hyaluronic acid in patients with chronic discogenic low back pain: 1-year
follow-up of a phase I study. Stem Cell Res. Ther. 2017,8, 262. [CrossRef]
106.
Kim, M.; Kim, K.H.; Song, S.U.; Yi, T.G.; Yoon, S.H.; Park, S.R.; Choi, B.H. Transplantation of human bone
marrow-derived clonal mesenchymal stem cells reduces fibrotic scar formation in a rat spinal cord injury
model. J. Tissue Eng. Regen. Med. 2018,12, E1034–E1045. [CrossRef]
107.
Jahan-Abad, A.J.; Negah, S.S.; Ravandi, H.H.; Ghasemi, S.; Borhani-Haghighi, M.; Stummer, W.; Gorji, A.;
Ghadiri, M.K. Human neural stem/progenitor cells derived from epileptic human brain in a self-assembling
peptide nanoscaold improve traumatic brain injury in rats. Mol. Neurobiol.
2018
,55, 9122–9138. [CrossRef]
108.
Luo, L.H.; He, Y.; Wang, X.Y.; Key, B.; Lee, B.H.; Li, H.Q.; Ye, Q.S. Potential roles of dental pulp stem cells in
neural regeneration and repair. Stem Cells Int. 2018,2018, 1731289. [CrossRef]
109.
Fesharaki, M.; Razavi, S.; Ghasemi-Mobarakeh, L.; Behjati, M.; Yarahmadian, R.; Kazemi, M.; Hejazi, H.
Dierentiation of human scalp adipose-derived mesenchymal stem cells into mature neural cells on
electrospun nanofibrous scaolds for nerve tissue engineering applications. Cell J. 2018,20, 168–176.
110.
Ma, Y.H.; Zeng, X.; Qiu, X.C.; Wei, Q.S.; Che, M.T.; Ding, Y.; Liu, Z.; Wu, G.H.; Sun, J.H.; Pang, M.; et al.
Perineurium-like sheath derived from long-term surviving mesenchymal stem cells confers nerve protection
to the injured spinal cord. Biomaterials 2018,160, 37–55. [CrossRef]
Cells 2019,8, 886 26 of 32
111.
Comar, M.; Delbue, S.; Zanotta, N.; Valencic, E.; Piscianz, E.; Del Savio, R.; Tesser, A.; Tommasini, A.;
Ferrante, P.
In vivo
detection of polyomaviruses JCV and SV40 in mesenchymal stem cells from human
umbilical cords. Pediatr. Blood Cancer 2014,61, 1347–1349. [CrossRef]
112.
van Velthoven, C.T.J.; Kavelaars, A.; Heijnen, C.J. Mesenchymal stem cells as a treatment for neonatal
ischemic brain damage. Pediatr. Res. 2012,71, 474–481. [CrossRef]
113.
Shi, B.; Ding, J.X.; Liu, Y.; Zhuang, X.M.; Zhuang, X.L.; Chen, X.S.; Fu, C.F. ERK1/2 pathway-mediated
dierentiation of IGF-1-transfected spinal cord-derived neural stem cells into oligodendrocytes. PloS ONE
2014,9, e106038. [CrossRef]
114.
de Araujo Farias, V.; Carrillo-Galvez, A.B.; Martin, F.; Anderson, P. TGF-
β
and mesenchymal stromal cells in
regenerative medicine, autoimmunity and cancer. Cytokine Growth Factor Rev. 2018,43, 25–37. [CrossRef]
115.
Zhang, Y.L.; Chopp, M.; Zhang, Z.G.; Katakowski, M.; Xin, H.Q.; Qu, C.S.; Ali, M.; Mahmood, A.; Xiong, Y.
Systemic administration of cell-free exosomes generated by human bone marrow derived mesenchymal
stem cells cultured under 2D and 3D conditions improves functional recovery in rats after traumatic brain
injury. Neurochem. Int. 2017,111, 69–81. [CrossRef]
116.
Oliveri, R.S.; Bello, S.; Biering-Sorensen, F. Mesenchymal stem cells improve locomotor recovery in traumatic
spinal cord injury: Systematic review with meta-analyses of rat models. Neurobiol. Dis.
2014
,62, 338–353.
[CrossRef]
117.
Oh, S.K.; Choi, K.H.; Yoo, J.Y.; Kim, D.Y.; Kim, S.J.; Jeon, S.R. A phase III clinical trial showing limited
ecacy of autologous mesenchymal stem cell therapy for spinal cord injury. Neurosurgery
2016
,78, 436–447.
[CrossRef]
118.
Wang, S.; Cheng, H.B.; Dai, G.H.; Wang, X.D.; Hua, R.R.; Liu, X.B.; Wang, P.S.; Chen, G.M.; Yue, W.; An, Y.H.
Umbilical cord mesenchymal stem cell transplantation significantly improves neurological function in
patients with sequelae of traumatic brain injury. Brain Res. 2013,1532, 76–84. [CrossRef]
119.
Hu, C.X.; Li, L.J. Preconditioning influences mesenchymal stem cell properties
in vitro
and
in vivo
.J. Cell
Mol. Med. 2018,22, 1428–1442. [CrossRef]
120.
Harris, V.K.; Stark, J.; Vyshkina, T.; Blackshear, L.; Joo, G.; Stefanova, V.; Sara, G.; Sadiq, S.A. Phase I trial of
intrathecal mesenchymal stem cell-derived neural progenitors in progressive multiple sclerosis. EBioMedicine
2018,29, 23–30. [CrossRef]
121.
Sykova, E.; Rychmach, P.; Drahoradova, I.; Konradova, S.; Ruzickova, K.; Vorisek, I.; Forostyak, S.; Homola, A.;
Bojar, M. Transplantation of mesenchymal stromal cells in patients with amyotrophic lateral sclerosis: Results
of phase I/IIa clinical trial. Cell Transplant. 2017,26, 647–658. [CrossRef]
122.
Shichinohe, H.; Kawabori, M.; Iijima, H.; Teramoto, T.; Abumiya, T.; Nakayama, N.; Kazumata, K.; Terasaka,S.;
Arato, T.; Houkin, K. Research on advanced intervention using novel bone marrOW stem cell (RAINBOW):
A study protocol for a phase I, open-label, uncontrolled, dose-response trial of autologous bone marrow
stromal cell transplantation in patients with acute ischemic stroke. BMC Neurol. 2017,17, 179. [CrossRef]
123.
Venkatesh, K.; Sen, D. Mesenchymal stem cells as a source of dopaminergic neurons: A potential cell based
therapy for Parkinson’s disease. Curr. Stem Cell Res. T. 2017,12, 326–347. [CrossRef]
124.
Bucan, V.; Vaslaitis, D.; Peck, C.T.; StrauSs, S.; Vogt, P.M.; Radtke, C. Eect of exosomes from rat
adipose-derived mesenchymal stem cells on neurite outgrowth and sciatic nerve regeneration after crush
injury. Mol. Neurobiol. 2019,56, 1812–1824. [CrossRef]
125.
Sun, X.; Zhu, Y.; Yin, H.Y.; Guo, Z.Y.; Xu, F.; Xiao, B.; Jiang, W.L.; Guo, W.M.; Meng, H.Y.; Lu, S.B.; et al.
Dierentiation of adipose-derived stem cells into Schwann cell-like cells through intermittent induction:
Potential advantage of cellular transient memory function. Stem Cell Res. Ther. 2018,9, 133. [CrossRef]
126.
Fan, L.H.; Yu, Z.F.; Li, J.; Dang, X.Q.; Wang, K.Z. Schwann-like cells seeded in acellular nerve grafts improve
nerve regeneration. BMC Musculoskel. Dis. 2014,15, 165. [CrossRef]
127.
Hu, F.H.; Zhang, X.F.; Liu, H.X.; Xu, P.; Doulathunnisa; Teng, G.J.; Xiao, Z.D. Neuronally dierentiated
adipose-derived stem cells and aligned PHBV nanofiber nerve scaolds promote sciatic nerve regeneration.
Biochem. Biophys. Res. Commun. 2017,489, 171–178. [CrossRef]
128.
Papa, S.; Vismara, I.; Mariani, A.; Barilani, M.; Rimondo, S.; De Paola, M.; Panini, N.; Erba, E.; Mauri, E.;
Rossi, F.; et al. Mesenchymal stem cells encapsulated into biomimetic hydrogel scaold gradually release
CCL2 chemokine in situ preserving cytoarchitecture and promoting functional recovery in spinal cord injury.
J. Controlled Release 2018,278, 49–56. [CrossRef]
Cells 2019,8, 886 27 of 32
129.
Lin, T.; Liu, S.; Chen, S.H.; Qiu, S.; Rao, Z.L.; Liu, J.H.; Zhu, S.; Yan, L.W.; Mao, H.Q.; Zhu, Q.T.; et al. Hydrogel
derived from porcine decellularized nerve tissue as a promising biomaterial for repairing peripheral nerve
defects. Acta Biomater. 2018,73, 326–338. [CrossRef]
130.
Lee, S.J.; Esworthy, T.; Stake, S.; Miao, S.; Zuo, Y.Y.; Harris, B.T.; Zhang, L.G. Advances in 3D bioprinting for
neural tissue engineering. Adv. Biosyst. 2018,2, 18. [CrossRef]
131.
Decosterd, I.; Woolf, C.J. Spared nerve injury: An animal model of persistent peripheral neuropathic pain.
Pain 2000,87, 149–158. [CrossRef]
132.
Angius, D.; Wang, H.; Spinner, R.J.; Gutierrez-Cotto, Y.; Yaszemski, M.J.; Windebank, A.J. A systematic
review of animal models used to study nerve regeneration in tissue-engineered scaolds. Biomaterials
2012
,
33, 8034–8039. [CrossRef]
133.
O’Rourke, C.; Day, A.G.E.; Murray-Dunning, C.; Thanabalasundaram, L.; Cowan, J.; Stevanato, L.; Grace, N.;
Cameron, G.; Drake, R.A.L.; Sinden, J.; et al. An allogeneic ‘othe shelf ’ therapeutic strategy for peripheral
nerve tissue engineering using clinical grade human neural stem cells. Sci. Rep. 2018,8, 11. [CrossRef]
134.
Thomson, A.A.; Cunha, G.R. Prostatic growth and development are regulated by FGF10. Development
1999
,
126, 3693–3701.
135.
Matthes, S.M.; Reimers, K.; Janssen, I.; Liebsch, C.; Kocsis, J.D.; Vogt, P.M.; Radtke, C. Intravenous
transplantation of mesenchymal stromal cells to enhance peripheral nerve regeneration. Biomed. Res. Int.
2013,2013, 573169. [CrossRef]
136.
Toma, C.; Pittenger, M.F.; Cahill, K.S.; Byrne, B.J.; Kessler, P.D. Human mesenchymal stem cells dierentiate
to a cardiomyocyte phenotype in the adult murine heart. Circulation 2002,105, 93–98. [CrossRef]
137.
Mori, D.; Miyagawa, S.; Yajima, S.; Saito, S.; Fukushima, S.; Ueno, T.; Toda, K.; Kawai, K.; Kurata, H.;
Nishida, H.; et al. Cell spray transplantation of adipose-derived mesenchymal stem cell recovers ischemic
cardiomyopathy in a porcine model. Transplantation 2018,102, 2012–2024. [CrossRef]
138.
Nagaya, N.; Kangawa, K.; Itoh, T.; Iwase, T.; Murakami, S.; Miyahara, Y.; Fujii, T.; Uematsu, M.; Ohgushi, H.;
Yamagishi, M.; et al. Transplantation of mesenchymal stem cells improves cardiac function in a rat model of
dilated cardiomyopathy. Circulation 2005,112, 1128–1135. [CrossRef]
139.
Gnecchi, M.; Danieli, P.; Cervio, E. Mesenchymal stem cell therapy for heart disease. Vascul. Pharmacol.
2012
,
57, 48–55. [CrossRef]
140.
Butler, J.; Epstein, S.E.; Greene, S.J.; Quyyumi, A.A.; Sikora, S.; Kim, R.J.; Anderson, A.S.; Wilcox, J.E.;
Tankovich, N.I.; Lipinski, M.J.; et al. Intravenous allogeneic mesenchymal stem cells for nonischemic
cardiomyopathy safety and ecacy results of a phase II-a randomized trial. Circ. Res.
2017
,120, 332–340.
[CrossRef]
141.
Choe, G.; Park, J.; Jo, H.; Kim, Y.S.; Ahn, Y.; Lee, J.Y. Studies on the eects of microencapsulated human
mesenchymal stem cells in RGD-modified alginate on cardiomyocytes under oxidative stress conditions
using in vitro biomimetic co-culture system. Int. J. Biol. Macromol. 2018,123, 512–520. [CrossRef]
142.
Jiang, S.; Zhang, S. Dierentiation of cardiomyocytes from amniotic fluid-derived mesenchymal stem cells
by combined induction with transforming growth factor
β
1 and 5-azacytidine. Mol. Med. Rep.
2017
,16,
5887–5893. [CrossRef]
143.
Jung, S.; Kim, J.H.; Yim, C.; Lee, M.; Kang, H.J.; Choi, D. Therapeutic eects of a mesenchymal stem cell-based
insulin-like growth factor-1/enhanced green fluorescent protein dual gene sorting system in a myocardial
infarction rat model. Mol. Med. Rep. 2018,18, 5563–5571. [CrossRef]
144.
Haneef, K.; Ali, A.; Khan, I.; Naeem, N.; Jamall, S.; Salim, A. Role of IL-7 in fusion of rat bone marrow
mesenchymal stem cells with cardiomyocytes
in vitro
and improvement of cardiac function
in vivo
.
Cardiovasc. Ther. 2018,36, e12479. [CrossRef]
145.
Fujita, Y.; Kawamoto, A. Stem cell-based peripheral vascular regeneration. Adv. Drug Deliv. Rev.
2017
,120,
25–40. [CrossRef]
146.
Shafei, A.E.; Ali, M.A.; Ghanem, H.G.; Shehata, A.I.; Abdelgawad, A.A.; Handal, H.R.; Talaat, K.A.;
Ashaal, A.E.; El-Shal, A.S. Mesenchymal stem cell therapy: A promising cell-based therapy for treatment of
myocardial infarction. J. Gene. Med. 2017,19, e2995. [CrossRef]
147.
Tse, H.F.; Kwong, Y.L.; Chan, J.K.; Lo, G.; Ho, C.L.; Lau, C.P. Angiogenesis in ischaemic myocardium by
intramyocardial autologous bone marrow mononuclear cell implantation. Lancet
2003
,361, 47–49. [CrossRef]
Cells 2019,8, 886 28 of 32
148.
Eng, G.; Lee, B.W.; Parsa, H.; Chin, C.D.; Schneider, J.; Linkov, G.; Sia, S.K.; Vunjak-Novakovic, G. Assembly
of complex cell microenvironments using geometrically docked hydrogel shapes. Proc. Natl. Acad. Sci. USA
2013,110, 4551–4556. [CrossRef]
149.
Cheng, G.S.; Wang, X.Y.; Li, Y.X.; He, L. Let-7a-transfected mesenchymal stem cells ameliorate
monocrotaline-induced pulmonary hypertension by suppressing pulmonary artery smooth muscle cell
growth through STAT3-BMPR2 signaling. Stem Cell Res. Ther. 2017,8, 11. [CrossRef]
150. Bernal, W.; Wendon, J. Acute liver failure. N. Engl. J. Med. 2013,369, 2525–2534. [CrossRef]
151.
Petersen, B.E.; Bowen, W.C.; Patrene, K.D.; Mars, W.M.; Sullivan, A.K.; Murase, N.; Boggs, S.S.;
Greenberger, J.S.; Go, J.P. Bone marrow as a potential source of hepatic oval cells. Science
1999
,284,
1168–1170. [CrossRef]
152.
Shi, D.Y.; Zhang, J.M.; Zhou, Q.; Xin, J.J.; Jiang, J.; Jiang, L.Y.; Wu, T.Z.; Li, J.; Ding, W.C.; Li, J.; et al.
Quantitative evaluation of human bone mesenchymal stem cells rescuing fulminant hepatic failure in pigs.
Gut 2017,66, 955–964. [CrossRef]
153.
Higashiyama, R.; Inagaki, Y.; Hong, Y.Y.; Kushida, M.; Nakao, S.; Niioka, M.; Watanabe, T.; Okano, H.;
Matsuzaki, Y.; Shiota, G.; et al. Bone marrow-derived cells express matrix metalloproteinases and contribute
to regression of liver fibrosis in mice. Hepatology 2007,45, 213–222. [CrossRef]
154.
Liu, W.H.; Song, F.Q.; Ren, L.N.; Guo, W.Q.; Wang, T.; Feng, Y.X.; Tang, L.J.; Li, K. The multiple functional
roles of mesenchymal stem cells in participating in treating liver diseases. J. Cell Mol. Med.
2015
,19, 511–520.
[CrossRef]
155.
Lou, G.H.; Yang, Y.; Liu, F.F.; Ye, B.J.; Chen, Z.; Zheng, M.; Liu, Y.N. MiR-122 modification enhances the
therapeutic ecacy of adipose tissue-derived mesenchymal stem cells against liver fibrosis. J. Cell Mol. Med.
2017,21, 2963–2973. [CrossRef]
156.
Suk, K.T.; Yoon, J.H.; Kim, M.Y.; Kim, C.W.; Kim, J.K.; Park, H.; Hwang, S.G.; Kim, D.J.; Lee, B.S.; Lee, S.H.;
et al. Transplantation with autologous bone marrow-derived mesenchymal stem cells for alcoholic cirrhosis:
Phase 2 trial. Hepatology 2016,64, 2185–2197. [CrossRef]
157.
Liang, J.; Zhang, H.; Zhao, C.; Wang, D.; Ma, X.; Zhao, S.; Wang, S.; Niu, L.; Sun, L. Eects of allogeneic
mesenchymal stem cell transplantation in the treatment of liver cirrhosis caused by autoimmune diseases.
Int. J. Rheum. Dis. 2017,20, 1219–1226. [CrossRef]
158.
Fang, X.Q.; Zhang, J.F.; Song, H.Y.; Chen, Z.L.; Dong, J.; Chen, X.; Pan, J.J.; Liu, B.; Chen, C.X. Eect of
umbilical cord mesenchymal stem cell transplantation on immune function and prognosis of patients with
decompensated hepatitis B cirrhosis. Zhonghua Gan Zang Bing Za Zhi. 2016,24, 907–910.
159.
Schwartz, R.E.; Reyes, M.; Koodie, L.; Jiang, Y.H.; Blackstad, M.; Lund, T.; Lenvik, T.; Johnson, S.;
Hu, W.S.; Verfaillie, C.M. Multipotent adult progenitor cells from bone marrow dierentiate into functional
hepatocyte-like cells. J. Clin. Invest. 2002,109, 1291–1302. [CrossRef]
160.
Lee, K.D.; Kuo, T.K.C.; Whang-Peng, J.; Chung, Y.F.; Lin, C.T.; Chou, S.H.; Chen, J.R.; Chen, Y.P.; Lee, O.K.S.
In vitro
hepatic dierentiation of human mesenchymal stem cells. Hepatology
2004
,40, 1275–1284. [CrossRef]
161.
Haga, H.; Yan, I.K.; Takahashi, K.; Matsuda, A.; Patel, T. Extracellular vesicles from bone marrow-derived
mesenchymal stem cells improve survival from lethal hepatic failure in mice. Stem Cell Transl. Med.
2017
,6,
1262–1272. [CrossRef]
162.
Papanikolaou, I.G.; Katselis, C.; Apostolou, K.; Feretis, T.; Lymperi, M.; Konstadoulakis, M.M.; Papalois, A.E.;
Zografos, G.C. Mesenchymal stem cells transplantation following partial hepatectomy: A new concept to
promote liver regeneration-systematic review of the literature focused on experimental studies in rodent
models. Stem Cells Int. 2017,22, 7567958. [CrossRef]
163.
Di Bonzo, L.V.; Ferrero, I.; Cravanzola, C.; Mareschi, K.; Rustichell, D.; Novo, E.; Sanavio, F.; Cannito, S.;
Zamara, E.; Bertero, M.; et al. Human mesenchymal stem cells as a two-edged sword in hepatic regenerative
medicine: Engraftment and hepatocyte dierentiation versus profibrogenic potential. Gut
2008
,57, 223–231.
[CrossRef]
164.
Decker, M.; Martinez-Morentin, L.; Wang, G.N.; Lee, Y.J.; Liu, Q.X.; Leslie, J.; Ding, L. Leptin-receptor-
expressing bone marrow stromal cells are myofibroblasts in primary myelofibrosis. Nat. Cell Biol.
2017
,19,
677–688. [CrossRef]
165.
Detry, O.; Vandermeulen, M.; Delbouille, M.H.; Somja, J.; Bletard, N.; Briquet, A.; Lechanteur, C.; Giet, O.;
Baudoux, E.; Hannon, M.; et al. Infusion of mesenchymal stromal cells after deceased liver transplantation:
A phase I-II, open-label, clinical study. J. Hepatol. 2017,67, 47–55. [CrossRef]
Cells 2019,8, 886 29 of 32
166.
Tsubota, K.; Satake, Y.; Kaido, M.; Shinozaki, N.; Shimmura, S.; Bissen-Miyajima, H.; Shimazaki, J. Treatment
of severe ocular-surface disorders with corneal epithelial stem-cell transplantation. N. Engl. J. Med.
1999
,
340, 1697–1703. [CrossRef]
167.
Navas, A.; Magana-Guerrero, F.S.; Dominguez-Lopez, A.; Chavez-Garcia, C.; Partido, G.;
Graue-Hernandez, E.O.; Sanchez-Garcia, F.J.; Garfias, Y. Anti-inflammatory and anti-fibrotic eects of
human amniotic membrane mesenchymal stem cells and their potential in corneal repair. Stem Cell Transl
Med. 2018,7, 906–917. [CrossRef]
168.
Haagdorens, M.; Van Acker, S.I.; Van Gerwen, V.; Ni Dhubhghaill, S.; Koppen, C.; Tassignon, M.J.; Zakaria, N.
Limbal stem cell deficiency: Current treatment options and emerging therapies. Stem Cells Int.
2016
,2016,
9798374. [CrossRef]
169.
Shen, T.; Shen, J.; Zheng, Q.Q.; Li, Q.S.; Zhao, H.L.; Cui, L.; Hong, C.Y. Cell viability and extracellular matrix
synthesis in a co-culture system of corneal stromal cells and adipose-derived mesenchymal stem cells. Int. J.
Ophthalmol. 2017,10, 670–678.
170.
Yamashita, K.; Inagaki, E.; Hatou, S.; Higa, K.; Ogawa, A.; Miyashita, H.; Tsubota, K.; Shimmura, S. Corneal
endothelial regeneration using mesenchymal stem cell derived from human umbilical cord. Stem Cells Dev.
2018,27, 1097–1108. [CrossRef]
171.
Lohan, P.; Murphy, N.; Treacy, O.; Lynch, K.; Morcos, M.; Chen, B.L.; Ryan, A.E.; Grin, M.D.; Ritter, T.
Third-party allogeneic mesenchymal stromal cells prevent rejection in a pre-sensitized high-risk model of
corneal transplantation. Front Immunol. 2018,9, 2666. [CrossRef]
172.
Ma, Y.L.; Xu, Y.S.; Xiao, Z.F.; Yang, W.; Zhang, C.; Song, E.; Du, Y.Q.; Li, L.S. Reconstruction of chemically
burned rat corneal surface by bone marrow-derived human mesenchymal stem cells. Stem Cells
2006
,24,
315–321. [CrossRef]
173.
Gu, S.F.; Xing, C.Z.; Han, J.Y.; Tso, M.O.M.; Hong, J. Dierentiation of rabbit bone marrow mesenchymal
stem cells into corneal epithelial cells in vivo and ex vivo.Mol. Vis. 2009,15, 99–107.
174.
Lee, H.J.; Ko, J.H.; Ko, A.Y.; Kim, M.K.; Wee, W.R.; Oh, J.Y. Intravenous infusion of mesenchymal stem/stromal
cells decreased CCR7
+
antigen presenting cells in mice with corneal allotransplantation. Curr. Eye. Res.
2014
,
39, 780–789. [CrossRef]
175.
Yao, L.; Li, Z.R.; Su, W.R.; Li, Y.P.; Lin, M.L.; Zhang, W.X.; Liu, Y.; Wan, Q.; Liang, D. Role of mesenchymal
stem cells on cornea wound healing induced by acute alkali burn. PloS ONE 2012,7, e30842. [CrossRef]
176.
Oh, J.Y.; Kim, M.K.; Shin, M.S.; Lee, H.J.; Ko, J.H.; Wee, W.R.; Lee, J.H. The anti-inflammatory and
anti-angiogenic role of mesenchymal stem cells in corneal wound healing following chemical injury.
Stem Cells 2008,26, 1047–1055. [CrossRef]
177.
Bray, L.J.; Heazlewood, C.F.; Munster, D.J.; Hutmacher, D.W.; Atkinson, K.; Harkin, D.G. Immunosuppressive
properties of mesenchymal stromal cell cultures derived from the limbus of human and rabbit corneas.
Cytotherapy 2014,16, 64–73. [CrossRef]
178.
Oh, J.Y.; Lee, R.H.; Yu, J.M.; Ko, J.H.; Lee, H.J.; Ko, A.Y.; Roddy, G.W.; Prockop, D.J. Intravenous mesenchymal
stem cells prevented rejection of allogeneic corneal transplants by aborting the early inflammatory response.
Mol. Ther. 2012,20, 2143–2152. [CrossRef]
179.
Den Hondt, M.; Vranckx, J.J. Reconstruction of defects of the trachea. J. Mater. Sci.: Mater. Med.
2017
,28, 11.
[CrossRef]
180.
Sun, Z.; Wang, Y.; Gong, X.; Su, H.; Han, X. Secretion of rat tracheal epithelial cells induces mesenchymal
stem cells to dierentiate into epithelial cells. Cell Biol. Int. 2012,36, 169–175. [CrossRef]
181.
Paunescu, V.; Deak, E.; Herman, D.; Siska, I.R.; Tanasie, G.; Bunu, C.; Anghel, S.; Tatu, C.A.; Oprea, T.I.;
Henschler, R.; et al.
In vitro
dierentiation of human mesenchymal stem cells to epithelial lineage. J. Cell
Mol. Med. 2007,11, 502–508. [CrossRef]
182.
Shu, C.; Li, T.Y.; Tsang, L.L.; Fok, K.L.; Lo, P.S.; Zhu, J.X.; Ho, L.S.; Chung, Y.W.; Chan, H.C. Differentiation of adult
rat bone marrow stem cells into epithelial progenitor cells in culture. Cell Biol. Int. 2006,30, 823–828. [CrossRef]
183.
Pan, S.; Zhong, Y.; Shan, Y.; Liu, X.; Xiao, Y.; Shi, H. Selection of the optimum 3D-printed pore and the surface
modification techniques for tissue engineering tracheal scaffold
in vivo
reconstruction. J. Biomed. Mater. Res. A
2019,107, 360–370. [CrossRef]
184.
Ghorbani, F.; Moradi, L.; Shadmehr, M.B.; Bonakdar, S.; Droodinia, A.; Safshekan, F. In-vivo characterization
of a 3D hybrid scaold based on PCL/decellularized aorta for tracheal tissue engineering. Mater. Sci. Eng. C
2017,81, 74–83. [CrossRef]
Cells 2019,8, 886 30 of 32
185.
Pan, S.; Liu, X.; Shi, H. The biocompatibility and immunogenicity study of decellularized tracheal matrix.
Zhongguo Xiu Fu Chong Jian Wai Ke Za Zhi 2018,32, 441–447.
186.
Chang, J.W.; Park, S.A.; Park, J.K.; Choi, J.W.; Kim, Y.S.; Shin, Y.S.; Kim, C.H. Tissue-engineered tracheal
reconstruction using three-dimensionally printed artificial tracheal graft: Preliminary report. Artif. Organs
2014,38, E95–E105. [CrossRef]
187.
Siddiqi, S.; de Wit, R.; van der Heide, S.; Oosterwijk, E.; Verhagen, A. Aortic allografts: Final destination?-a
summary of clinical tracheal substitutes. J. Thorac. Dis. 2018,10, 5149–5153. [CrossRef]
188.
Jorge, L.F.; Francisco, J.C.; Bergonse, N.; Baena, C.; Carvalho, K.A.T.; Abdelwahid, E.; Neto, J.R.F.;
Moreira, L.F.P.; Guarita-Souza, L.C. Tracheal repair with acellular human amniotic membrane in a rabbit
model. J. Tissue Eng. Regen. Med. 2018,12, E1525–E1530. [CrossRef]
189.
Xu, Y.; Li, D.; Yin, Z.Q.; He, A.J.; Lin, M.M.; Jiang, G.N.; Song, X.; Hu, X.F.; Liu, Y.; Wang, J.P.; et al.
Tissue-engineered trachea regeneration using decellularized trachea matrix treated with laser micropore
technique. Acta Biomater. 2017,58, 113–121. [CrossRef]
190.
Jing, H.; Gao, B.; Gao, M.; Yin, H.; Mo, X.; Zhang, X.; Luo, K.; Feng, B.; Fu, W.; Wang, J.; et al. Restoring tracheal
defects in a rabbit model with tissue engineered patches based on TGF-
β
3-encapsulating electrospun poly(L-lactic
acid-co-ε-caprolactone)/collagen scaffolds. Artif. Cells, Nanomed., Biotechnol. 2018,46, 985–995. [CrossRef]
191.
Wang, J.; Sun, B.; Tian, L.; He, X.; Gao, Q.; Wu, T.; Ramakrishna, S.; Zheng, J.; Mo, X. Evaluation of the potential of
rhTGF-
β
3 encapsulated P(LLA-CL)/collagen nanofibers for tracheal cartilage regeneration using mesenchymal
stems cells derived from Wharton’s jelly of human umbilical cord. Mater. Sci. Eng. C
2017
,70, 637–645. [CrossRef]
192.
Bae, S.W.; Lee, K.W.; Park, J.H.; Lee, J.; Jung, C.R.; Yu, J.; Kim, H.Y.; Kim, D.H. 3D bioprinted artificial trachea
with epithelial cells and chondrogenic-dierentiated bone marrow-derived mesenchymal stem cells. Int. J.
Mol. Sci. 2018,19, 14. [CrossRef]
193.
Maughan, E.F.; Hynds, R.E.; Proctor, T.J.; Janes, S.M.; Elliott, M.; Birchall, M.A.; Lowdell, M.W.; De Coppi, P.
Autologous cell seeding in tracheal tissue engineering. Curr. Stem Cell Rep. 2017,3, 279–289. [CrossRef]
194.
Macchiarini, P.; Jungebluth, P.; Go, T.; Asnaghi, M.A.; Rees, L.E.; Cogan, T.A.; Dodson, A.; Martorell, J.;
Bellini, S.; Parnigotto, P.P.; et al. Clinical transplantation of a tissue-engineered airway. Lancet
2008
,372,
2023–2030. [CrossRef]
195.
Gonfiotti, A.; Jaus, M.O.; Barale, D.; Baiguera, S.; Comin, C.; Lavorini, F.; Fontana, G.; Sibila, O.; Rombola, G.;
Jungebluth, P.; et al. The first tissue-engineered airway transplantation: 5-year follow-up results. Lancet
2014
,
383, 238–244. [CrossRef]
196.
Elliott, M.J.; De Coppi, P.; Speggiorin, S.; Roebuck, D.; Butler, C.R.; Samuel, E.; Crowley, C.; McLaren, C.;
Fierens, A.; Vondrys, D.; et al. Stem-cell-based, tissue engineered tracheal replacement in a child: A 2-year
follow-up study. Lancet 2012,380, 994–1000. [CrossRef]
197.
Wang, P.H.; Huang, B.S.; Horng, H.C.; Yeh, C.C.; Chen, Y.J. Wound healing. J. Chin. Med. Assoc.
2018
,81,
94–101. [CrossRef]
198.
Laverdet, B.; Micallef, L.; Lebreton, C.; Mollard, J.; Lataillade, J.J.; Coulomb, B.; Desmouliere, A. Use of
mesenchymal stem cells for cutaneous repair and skin substitute elaboration. Pathol. Biol.
2014
,62, 108–117.
[CrossRef]
199.
Feldman, D.S.; McCauley, J.F. Mesenchymal stem cells and transforming growth factor-
β3
(TGF-
β3
) to
enhance the regenerative ability of an albumin scaold in full thickness wound healing. J. Funct. Biomater.
2018,9, 65. [CrossRef]
200.
Qi, Y.; Jiang, D.S.; Sindrilaru, A.; Stegemann, A.; Schatz, S.; Treiber, N.; Rojewski, M.; Schrezenmeier, H.;
Beken, S.V.; Wlaschek, M.; et al. TSG-6 released from intradermally injected mesenchymal stem cells
accelerates wound healing and reduces tissue fibrosis in murine full-thickness skin wounds. J. Invest.
Dermatol. 2014,134, 526–537. [CrossRef]
201.
Ong, H.T.; Dilley, R.J. Novel non-angiogenic role for mesenchymal stem cell-derived vascular endothelial
growth factor on keratinocytes during wound healing. Cytokine Growth Factor Rev.
2018
,44, 69–79. [CrossRef]
202.
Jiang, X.; Wu, F.; Xu, Y.; Yan, J.X.; Wu, Y.D.; Li, S.H.; Liao, X.; Liang, J.X.; Li, Z.H.; Liu, H.W. A novel role
of angiotensin II in epidermal cell lineage determination: Angiotensin II promotes the dierentiation of
mesenchymal stem cells into keratinocytes through the p38 MAPK, JNK and JAK2 signalling pathways.
Exp. Dermatol. 2019,28, 59–65. [CrossRef]
Cells 2019,8, 886 31 of 32
203.
Kolakshyapati, P.; Li, X.Y.; Chen, C.Y.; Zhang, M.X.; Tan, W.Q.; Ma, L.E.; Gao, C.Y. Gene-activated matrix/bone
marrow-derived mesenchymal stem cells constructs regenerate sweat glands-like structure
in vivo
.Sci. Rep.
2017,7, 13. [CrossRef]
204.
Yang, D.; Sun, S.; Wang, Z.; Zhu, P.; Yang, Z.; Zhang, B. Stromal cell-derived factor-1 receptor
CXCR4-overexpressing bone marrow mesenchymal stem cells accelerate wound healing by migrating
into skin injury areas. Cell Reprogram. 2013,15, 206–215. [CrossRef]
205.
Kisseleva, T.; Uchinami, H.; Feirt, N.; Quintana-Bustamante, O.; Segovia, J.C.; Schwabe, R.F.; Brenner, D.A. Bone
marrow-derived fibrocytes participate in pathogenesis of liver fibrosis. J. Hepatol. 2006,45, 429–438. [CrossRef]
206.
Nakamura, Y.; Ishikawa, H.; Kawai, K.; Tabata, Y.; Suzuki, S. Enhanced wound healing by topical
administration of mesenchymal stem cells transfected with stromal cell-derived factor-1. Biomaterials
2013,34, 9393–9400. [CrossRef]
207.
Cai, S.; Pan, Y.; Han, B.; Sun, T.Z.; Sheng, Z.Y.; Fu, X.B. Transplantation of human bone marrow-derived
mesenchymal stem cells transfected with ectodysplasin for regeneration of sweat glands. Chin. Med. J.
2011
,
124, 2260–2268.
208.
Solmaz, H.; Ulgen, Y.; Gulsoy, M. Photobiomodulation of wound healing via visible and infrared laser
irradiation. Lasers Med. Sci. 2017,32, 903–910. [CrossRef]
209.
Murphy, K.C.; Whitehead, J.; Zhou, D.J.; Ho, S.S.; Leach, J.K. Engineering fibrin hydrogels to promote the
wound healing potential of mesenchymal stem cell spheroids. Acta Biomater. 2017,64, 176–186. [CrossRef]
210.
Tavakol, D.N.; Jeffries, L.A.; Schwager, S.C.; Kelly-Goss, M.R.; Peirce, S.M. Design of a 3D-bioprinted mesenchymal
stem-cell laden construct for accelerating angiogenesis in diabetic wounds. FASEB J. 2017,31, 2.
211.
Shokrgozar, M.A.; Fattahi, M.; Bonakdar, S.; Ragerdi Kashani, I.; Majidi, M.; Haghighipour, N.; Bayati, V.;
Sanati, H.; Saeedi, S.N. Healing potential of mesenchymal stem cells cultured on a collagen-based scaold
for skin regeneration. Iran. Biomed. J. 2012,16, 68–76.
212.
Rodrigues, C.; de Assis, A.M.; Moura, D.J.; Halmenschlager, G.; Sa, J.; Xavier, L.L.; Fernandes Mda, C.;
Wink, M.R. New therapy of skin repair combining adipose-derived mesenchymal stem cells with sodium
carboxymethylcellulose scaold in a pre-clinical rat model. PLoS ONE 2014,9, e96241. [CrossRef]
213.
Navone, S.E.; Pascucci, L.; Dossena, M.; Ferri, A.; Invernici, G.; Acerbi, F.; Cristini, S.; Bedini, G.; Tosetti, V.;
Ceserani, V.; et al. Decellularized silk fibroin scaold primed with adipose mesenchymal stromal cells
improves wound healing in diabetic mice. Stem Cell Res. Ther. 2014,5, 7. [CrossRef]
214.
Qi, C.; Xu, L.M.; Deng, Y.; Wang, G.B.; Wang, Z.; Wang, L. Sericin hydrogels promote skin wound healing
with eective regeneration of hair follicles and sebaceous glands after complete loss of epidermis and dermis.
Biomater. Sci. 2018,6, 13. [CrossRef]
215.
Pelizzo, G.; Avanzini, M.A.; Mantelli, M.; Croce, S.; Maltese, A.; Vestri, E.; De Silvestri, A.; Percivalle, E.;
Calcaterra, V. Granulation tissue-derived mesenchymal stromal cells: A potential application for burn wound
healing in pediatric patients. J. Stem Cells Regen. Med. 2018,14, 53–58.
216.
Lightner, A.L.; Wang, Z.; Zubair, A.C.; Dozois, E.J. A systematic review and meta-analysis of mesenchymal
stem cell injections for the treatment of perianal Crohn’s disease: Progress made and future directions.
Dis. Colon Rectum. 2018,61, 629–640. [CrossRef]
217.
Wu, Q.N.; Lei, X.T.; Chen, L.; Zheng, Y.L.; Huang, H.M.; Qian, C.; Liang, Z.W. Autologous platelet-rich gel
combined with
in vitro
amplification of bone marrow mesenchymal stem cell transplantation to treat the
diabetic foot ulcer: A case report. Ann. Transl. Med. 2018,6, 7. [CrossRef]
218.
Kuhl, T.; Mezger, M.; Hausser, I.; Handgretinger, R.; Bruckner-Tuderman, L.; Nystrom, A. High local
concentrations of intradermal MSCs restore skin integrity and facilitate wound healing in dystrophic
epidermolysis bullosa. Mol. Ther. 2015,23, 1368–1379. [CrossRef]
219.
Portas, M.; Mansilla, E.; Drago, H.; Dubner, D.; Radl, A.; Coppola, A.; Di Giorgio, M. Use of human cadaveric
mesenchymal stem cells for cell therapy of a chronic radiation-induced skin lesion: A case report. Radiat.
Prot. Dosim. 2016,171, 99–106. [CrossRef]
220.
Zhuge, Y.; Regueiro, M.M.; Tian, R.; Li, Y.; Xia, X.; Vazquez-Padron, R.; Elliot, S.; Thaller, S.R.; Liu, Z.J.;
Velazquez, O.C. The eect of estrogen on diabetic wound healing is mediated through increasing the function
of various bone marrow-derived progenitor cells. J. Vasc. Surg. 2018,68, 127S–135S. [CrossRef]
221.
Sheng, Z.; Fu, X.; Cai, S.; Lei, Y.; Sun, T.; Bai, X.; Chen, M. Regeneration of functional sweat gland-like
structures by transplanted dierentiated bone marrow mesenchymal stem cells. Wound Repair Regen.
2009
,
17, 427–435. [CrossRef]
Cells 2019,8, 886 32 of 32
222.
Xia, Y.; You, X.E.; Chen, H.; Yan, Y.J.; He, Y.C.; Ding, S.Z. Epidermal growth factor promotes mesenchymal
stem cell-mediated wound healing and hair follicle regeneration. Int. J. Clin. Exp. Pathol.
2017
,10, 7390–7400.
223.
Jiang, M.H.; Li, G.L.; Liu, J.F.; Liu, L.S.; Wu, B.Y.; Huang, W.J.; He, W.; Deng, C.H.; Wang, D.; Li, C.L.;
et al. Nestin
+
kidney resident mesenchymal stem cells for the treatment of acute kidney ischemia injury.
Biomaterials 2015,50, 56–66. [CrossRef]
224.
Zhou, Z.; Yan, H.; Liu, Y.D.; Xiao, D.D.; Li, W.; Wang, Q.; Zhao, Y.; Sun, K.; Zhang,M.; Lu, M.J. Adipose-derived
stem-cell-implanted poly(
ε
-caprolactone)/chitosan scaold improves bladder regeneration in a rat model.
Regen. Med. 2018,13, 331–342. [CrossRef]
225.
Kolesky, D.B.; Homan, K.A.; Skylar-Scott, M.A.; Lewis, J.A. Three-dimensional bioprinting of thick
vascularized tissues. Proc. Natl. Acad. Sci. USA 2016,113, 3179–3184. [CrossRef]
226.
Lin, P.P.; Wang, Y.; Lozano, G. Mesenchymal stem cells and the origin of Ewing’s sarcoma. Sarcoma
2011
,
2011, 276463. [CrossRef]
227.
Amariglio, N.; Hirshberg, A.; Scheithauer, B.W.; Cohen, Y.; Loewenthal, R.; Trakhtenbrot, L.; Paz, N.;
Koren-Michowitz, M.; Waldman, D.; Leider-Trejo, L.; et al. Donor-derived brain tumor following neural
stem cell transplantation in an ataxia telangiectasia patient. PLoS Med. 2009,6, e1000029. [CrossRef]
228.
Uccelli, A.; Moretta, L.; Pistoia, V. Mesenchymal stem cells in health and disease. Nat. Rev. Immunol.
2008
,8,
726–736. [CrossRef]
229.
Karnoub, A.E.; Dash, A.B.; Vo, A.P.; Sullivan, A.; Brooks, M.W.; Bell, G.W.; Richardson, A.L.; Polyak, K.;
Tubo, R.; Weinberg, R.A. Mesenchymal stem cells within tumour stroma promote breast cancer metastasis.
Nature 2007,449, 557–563. [CrossRef]
230.
Weis, S.M.; Cheresh, D.A. Tumor angiogenesis: Molecular pathways and therapeutic targets. Nat. Med.
2011
,
17, 1359–1370. [CrossRef]
231.
Li, W.; Ren, G.; Huang, Y.; Su, J.; Han, Y.; Li, J.; Chen, X.; Cao, K.; Chen, Q.; Shou, P.; et al. Mesenchymal stem
cells: A double-edged sword in regulating immune responses. Cell Death Differ. 2012,19, 1505–1513. [CrossRef]
232.
Popova, A.P.; Bozyk, P.D.; Goldsmith, A.M.; Linn, M.J.; Lei, J.; Bentley, J.K.; Hershenson, M.B. Autocrine
production of TGF-
β
1 promotes myofibroblastic dierentiation of neonatal lung mesenchymal stem cells.
Am. J. Physiol.: Lung Cell. Mol. Physiol. 2010,298, L735–743. [CrossRef]
233.
Collins, J.J.; Thebaud, B. Lung mesenchymal stromal cells in development and disease: To serve and protect?
Antioxid. Redox. Signal. 2014,21, 1849–1862. [CrossRef]
234.
Lalu, M.M.; McIntyre, L.; Pugliese, C.; Fergusson, D.; Winston, B.W.; Marshall, J.C.; Granton, J.; Stewart, D.J.
Safety of cell therapy with mesenchymal stromal cells (SafeCell): A systematic review and meta-analysis of
clinical trials. PLoS ONE 2012,7, e47559. [CrossRef]
235.
Tsutsumi, S.; Shimazu, A.; Miyazaki, K.; Pan, H.; Koike, C.; Yoshida, E.; Takagishi, K.; Kato, Y. Retention of
multilineage dierentiation potential of mesenchymal cells during proliferation in response to FGF. Biochem.
Biophys. Res. Commun. 2001,288, 413–419. [CrossRef]
236.
Shahdadfar, A.; Fronsdal, K.; Haug, T.; Reinholt, F.P.; Brinchmann, J.E.
In vitro
expansion of human
mesenchymal stem cells: Choice of serum is a determinant of cell proliferation, dierentiation, gene
expression, and transcriptome stability. Stem Cells 2005,23, 1357–1366. [CrossRef]
237.
Berniakovich, I.; Giorgio, M. Low oxygen tension maintains multipotency, whereas normoxia increases
dierentiation of mouse bone marrow stromal cells. Int. J. Mol. Sci. 2013,14, 2119–2134. [CrossRef]
238.
Holubova, M.; Lysak, D.; Vlas, T.; Vannucci, L.; Jindra, P. Expanded cryopreserved mesenchymal stromal
cells as an optimal source for graft-versus-host disease treatment. Biologicals. 2014,42, 139–144. [CrossRef]
239.
Stolzing, A.; Jones, E.; McGonagle, D.; Scutt, A. Age-related changes in human bone marrow-derived
mesenchymal stem cells: Consequences for cell therapies. Mech. Ageing. Dev.
2008
,129, 163–173. [CrossRef]
©
2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
Article
Microfluidic cell encapsulation has provided a platform for studying the behavior of individual cells and has become a turning point in single-cell analysis during the last decade. The engineered microenvironment, along with protecting the immune response, has led to increasingly presenting the results of practical and pre-clinical studies with the goals of disease treatment, tissue engineering, intelligent control of stem cell differentiation, and regenerative medicine. However, the significance of cell-substrate interaction versus cell-cell communications in the microgel is still unclear. In this study, monodisperse alginate microgels were generated using a flow-focusing microfluidic device to determine how the cell microenvironment can control human bone marrow-derived mesenchymal stem cells (hBMSCs) viability, proliferation, and biomechanical features in single-cell droplets versus multi-cell droplets. Collected results show insufficient cell proliferation (234 % and 329 %) in both single- and multi-cell alginate microgels. Alginate hydrogels supplemented with poly-L-lysine (PLL) showed a better proliferation rate (514 % and 780 %) in a comparison of free alginate hydrogels. Cell stiffness data illustrate that hBMSCs cultured in alginate hydrogels have higher membrane flexibility and migration potency (Young's modulus equal to 1.06 kPa), whereas PLL introduces more binding sites for cell attachment and causes lower flexibility and migration potency (Young's modulus equal to 1.83 kPa). Considering that cell adhesion is the most important parameter in tissue engineering, in which cells do not run away from a 3D substrate, PLL enhances cell stiffness and guarantees cell attachments. In conclusion, cell attachment to PLL-mediated alginate hydrogels is crucial for cell viability and proliferation. It suggests that cell-cell signaling is good enough for stem cell viability, but cell-PLL attachment alongside cell-cell signaling is crucial for stem cell proliferation and selfrenewal.
Article
Mesenchymal stem cells (MSCs) have tantalized regenerative medicine with their therapeutic potential, yet a cloud of controversies looms over their clinical transplantation. This comprehensive review navigates the intricate landscape of MSC controversies, drawing upon 15 years of clinical experience and research. We delve into the fundamental properties of MSCs, exploring their unique immunomodulatory capabilities and surface markers. The heart of our inquiry lies in the controversial applications of MSC transplantation, including the perennial debate between autologous and allogeneic sources, concerns about efficacy, and lingering safety apprehensions. Moreover, we unravel the enigmatic mechanisms surrounding MSC transplantation, such as homing, integration, and the delicate balance between differentiation and paracrine effects. We also assess the current status of clinical trials and the ever-evolving regulatory landscape. As we peer into the future, we examine emerging trends, envisioning personalized medicine and innovative delivery methods. Our review provides a balanced and informed perspective on the controversies, offering readers a clear understanding of the complexities, challenges, and potential solutions in MSC transplantation.
Article
Full-text available
Glioblastoma (GBM) is the most common and aggressive malignant brain tumor. Standard treatments including surgical resection, radiotherapy, and chemotherapy, have failed to significantly improve the prognosis of glioblastoma patients. Currently, immunotherapeutic approaches based on vaccines, chimeric antigen-receptor T-cells, checkpoint inhibitors, and oncolytic virotherapy are showing promising results in clinical trials. The combination of different immunotherapeutic approaches is proving satisfactory and promising. In view of the challenges of immunotherapy and the resistance of glioblastomas, the treatment of these tumors requires further efforts. In this review, we explore the obstacles that potentially influence the efficacy of the response to immunotherapy and that should be taken into account in clinical trials. This article provides a comprehensive review of vaccine therapy for glioblastoma. In addition, we identify the main biomarkers, including isocitrate dehydrogenase, epidermal growth factor receptor, and telomerase reverse transcriptase, known as potential immunotherapeutic targets in glioblastoma, as well as the current status of clinical trials. This paper also lists proposed solutions to overcome the obstacles facing immunotherapy in glioblastomas.
Article
Full-text available
Mesenchymal stem cells (MSCs) of placental origin hold great promise in tissue engineering and regenerative medicine for diseases affecting cartilage and bone. However, their utility has been limited by their tendency to undergo premature senescence and phenotypic drift into adipocytes. This study aimed to explore the potential involvement of a specific subset of aging and antiaging genes by measuring their expression prior to and following in vitro-induced differentiation of placental MSCs into chondrocytes and osteoblasts as opposed to adipocytes. The targeted genes of interest included the various LMNA/C transcript variants (lamin A, lamin C, and lamin A∆10), sirtuin 7 (SIRT7), and SM22α, along with the classic aging markers plasminogen activator inhibitor 1 (PAI-1), p53, and p16INK4a. MSCs were isolated from the decidua basalis of human term placentas, expanded, and then analyzed for phenotypic properties by flow cytometry and evaluated for colony-forming efficiency. The cells were then induced to differentiate in vitro into chondrocytes, osteocytes, and adipocytes following established protocols. The mRNA expression of the targeted genes was measured by RT-qPCR in the undifferentiated cells and those fully differentiated into the three cellular lineages. Compared to undifferentiated cells, the differentiated chondrocytes demonstrated decreased expression of SIRT7, along with decreased PAI-1, lamin A, and SM22α expression, but the expression of p16INK4a and p53 increased, suggesting their tendency to undergo premature senescence. Interestingly, the cells maintained the expression of lamin C, which indicates that it is the primary lamin variant influencing the mechanoelastic properties of the differentiated cells. Notably, the expression of all targeted genes did not differ from the undifferentiated cells following osteogenic differentiation. On the other hand, the differentiation of the cells into adipocytes was associated with decreased expression of lamin A and PAI-1. The distinct patterns of expression of aging and antiaging genes following in vitro-induced differentiation of MSCs into chondrocytes, osteocytes, and adipocytes potentially reflect specific roles for these genes during and following differentiation in the fully functional cells. Understanding these roles and the network of signaling molecules involved can open opportunities to improve the handling and utility of MSCs as cellular precursors for the treatment of cartilage and bone diseases.
Article
Full-text available
Background Microvesicles are membraned particles produced by different types of cells recently investigated for anticancer purposes. The current study aimed to investigate the effects of human bone marrow mesenchymal stem cell-derived microvesicles (BMSC-MVs) on the multiple myeloma cell line U266. BMSC-MVs were isolated from BMSCs via ultracentrifugation and characterized using transmission electron microscopy (TEM) and dynamic light scattering (DLS). U266 cells were treated with 15, 30, 60, and 120 µg/mL BMSC-MVs for three and seven days and the effects of treatment in terms of viability, cytotoxicity, and DNA damage were investigated via the MTT assay, lactate dehydrogenase (LDH) assay, and 8‑hydroxy-2’-deoxyguanosine (8‑OHdG) measurement, respectively. Moreover, the apoptosis rate of the U266 cells treated with 60 µg/mL BMSC-MVs was also assessed seven days following treatment via flow cytometry. Ultimately, the expression level of BCL2, BAX, and CCND1 by the U266 cells was examined seven days following treatment with 60 µg/mL BMSC-MVs using qRT-PCR. Results BMSC-MVs had an average size of ~ 410 nm. According to the MTT and LDH assays, BMSC-MV treatment reduced the U266 cell viability and mediated cytotoxic effects against them, respectively. Moreover, elevated 8‑OHdG levels following BMSC-MV treatment demonstrated a dose-dependent increase of DNA damage in the treated cells. BMSC-MV-treated U266 cells also exhibited an increased apoptosis rate after seven days of treatment. The expression level of BCL2 and CCND1 decreased in the treated cells whereas the BAX expression demonstrated an incremental pattern. Conclusions Our findings accentuate the therapeutic benefit of BMSC-MVs against the multiple myeloma cell line U266 and demonstrate how microvesicles could be of therapeutic advantage. Future in vivo studies could further corroborate these findings.
Chapter
Hard-to-heal wounds are an important public health issue worldwide, with a significant impact on the quality of life of patients. It is estimated that approximately 1–2% of the global population suffers from difficult wounds, which can be caused by a variety of factors such as trauma, infections, chronic diseases like diabetes or obesity, or poor health conditions. Hard-to-heal wounds are often characterized by a slow and complicated healing process, which can lead to serious complications such as infections, pressure ulcers, scar tissue formation, and even amputations. These complications can have a significant impact on the mobility, autonomy, and quality of life of patients, leading to an increase in healthcare and social costs associated with wound care. The preparation of the wound bed is a key concept in the management of hard-to-heal wounds, with the aim of promoting an optimal environment for healing. The TIME (Tissue, Infection/Inflammation, Moisture, Edge) model is a systematic approach used to assess and manage wounds in a targeted and personalized way. The concept of TIMER, expanding the TIME model, further focuses on regenerative processes, paying particular attention to promoting tissue regeneration and wound healing in a more effective and comprehensive way. The new element introduced in the TIMER model is “Regeneration”, which highlights the importance of activating and supporting tissue regeneration processes to promote complete and lasting wound healing. Regenerative therapies can include a wide range of approaches, including cellular therapies, growth factors, bioactive biomaterials, stem cell therapies, and growth factor therapies. These therapies aim to promote the formation of new healthy tissues, reduce inflammation, improve vascularization, and stimulate cellular proliferation to accelerate wound closure and prevent complications. Thanks to continuous progress in research and development of regenerative therapies, more and more patients suffering from difficult wounds can benefit from innovative and promising solutions to promote faster and more effective healing, improve quality of life, and reduce the risk of long-term complications.
Article
Background Magnetic intramedullary lengthening nailing has demonstrated benefits over external fixation devices for femoral bone lengthening. These include avoiding uncomfortable external fixation and associated pin site infections, scarring, and inhibition of muscle or joint function. Despite this, little has changed in the field of biologically enhanced bone regeneration. Venting the femoral intramedullary canal at the osteotomy site before reaming creates egress for bone marrow during reaming. The reamings that are extruded from vent holes may function as a prepositioned bone graft at the distraction gap. The relationship between venting and the consolidation of regenerating bone remains unclear. Questions/purposes (1) Do bone marrow reamings extruded through venting holes enhance the quality of bone regeneration and improve healing indices and consolidation times? (2) Is venting associated with a higher proportion of complications than nonventing? Methods We performed a retrospective study of femoral lengthening performed at one hospital from December 2012 to February 2022 using a magnetic intramedullary lengthening nail with or without venting at the osteotomy site before reaming. This was a generally sequential series, in which the study groups were assembled as follows: Venting was performed between July 2012 and August 2016 and again from November 2021 onward. Nonventing was used between October 2016 and October 2021 because the senior author opted to create drill holes after the reaming procedure to avoid commitment to the osteotomy level before completing the reaming procedure. Outcomes were evaluated based on bone healing time, time to achieve full weightbearing, and complications. Sixty-one femoral lengthening procedures were studied (in 33 male and 28 female patients); two patients were excluded because of implant breakage. The mean age was 17 ± 5 years. The mean amount of lengthening was 55 ± 13 mm in the venting group and 48 ± 16 mm in the nonventing group (mean difference 7 ± 21 [95% CI 2 to 12]; p = 0.07). The healing index was defined as the time (in days) required for three cortices to bridge with new bone formation divided by the length (in cm) lengthened during the clinical protocol. This index signifies the bone formation rate achieved under the specific conditions of the protocol. Full weightbearing was allowed upon bridging the regenerated gap on three sides. Consolidation time was defined as the total number of days from the completion of the lengthening phase until adequate bone union (all three cortices healed) was achieved and full weightbearing was permitted. This time frame represents the entire healing process after the lengthening is complete divided by the amount of lengthening achieved (in cm). Patient follow-up was conducted meticulously at our institution, and we adhered to a precise schedule, occurring every 2 weeks during the distraction phase and every 4 weeks during the consolidation phase. There were no instances of loss to follow-up. Every patient completed the treatment successfully, reaching the specified milestones of weightbearing and achieving three cortexes of bone bridging. Results The mean healing index time in the venting group was faster than that in the nonventing group (21 ± 6 days/cm versus 31 ± 22 days/cm, mean difference 10 ± 23 [95% CI 4 to 16]; p = 0.02). The mean consolidation time was faster in the venting group than the nonventing group (10 ± 6 days/cm versus 20 ± 22 days/cm; mean difference 10 ± 23 [95% CI 4 to 15]; p = 0.02). No medical complications such as deep vein thrombosis or fat or pulmonary embolism were seen. Two patients had lengthy delays in regenerate union, both of whom were in the nonventing group (healing indexes were 74 and 62 days/cm; consolidation time was 52 and 40 days/cm). Conclusion Femoral lengthening with a magnetic intramedullary lengthening nail healed more quickly with prereaming venting than with nonventing, and it allowed earlier full weightbearing without any major associated complications. Future studies should evaluate whether there is a correlation between the number of venting holes and improvement in the healing index and consolidation time. Level of Evidence Level III, therapeutic study.
Article
Full-text available
The influences of pore sizes and surface modifications on biomechanical properties and biocompatibility (BC) of porous tracheal scaffolds (PTSs) fabricated by polycaprolactone (PCL) using 3D printing technology. The porous grafts were surface‐modified through hydrolysis, amination, and nanocrystallization treatment. The surface properties of the modified grafts were characterized by energy dispersive spectroscopy (EDS) and scanning electron microscopy (SEM). The materials were cocultured with bone marrow mesenchymal stem cells (BMSCs). The effect of different pore sizes and surface modifications on the cell proliferation behavior was evaluated by the cell counting kit‐8 (CCK‐8). Compared to native tracheas, the PTS has good biomechanical properties. A pore diameter of 200 μm is the optimum for cell adhesion, and the surface modifications successfully improved the cytotropism of the PTS. Allogeneic implantation confirmed that it largely retains its structural integrity in the host, and the immune rejection reaction of the PTS decreased significantly after the acute phase. Nano‐silicon dioxide (NSD)‐modified PTS is a promising material for tissue engineering tracheal reconstruction. © 2018 Wiley Periodicals, Inc. J Biomed Mater Res Part A:, 2018.
Article
Full-text available
High-risk cornea transplant recipients represent a patient population with significant un-met medical need for more effective therapies to prevent immunological graft rejection due to heightened anti-donor immune response. In this study, a rat model of pre-existing anti-donor immunity was developed in which corneal allografts were rejected earlier than in non-pre-sensitized recipients. In this model, third-party (non-donor, non-recipient strain) allogeneic mesenchymal stromal cells (allo-MSC) were administered intravenously 7 and 1 days prior to transplantation. Rejection-free graft survival to 30 days post-transplant improved from 0 to 63.6% in MSC-treated compared to vehicle-treated control animals (p = < 0.0001). Pre-sensitized animals that received third-party allo-MSC prior to transplantation had significantly higher proportions of CD45⁺CD11b⁺ B220⁺ monocytes in the lungs 24 h after the second MSC injection and significantly higher proportions of CD4⁺ FoxP3⁺ regulatory T cells in the graft-draining lymph nodes at the average day of rejection of control animals. In in vitro experiments, third-party allo-MSC polarized primary lung-derived CD11b/c⁺ myeloid cells to a more anti-inflammatory phenotype, as determined by cytokine profile and conferred them with the capacity to suppress T cell activation via prostaglandin E2 and TGFβ1. In experiments designed to further validate the clinical potential of the protocol, thawed cryopreserved, third-party allo-MSC were shown to be similarly potent at prolonging rejection-free corneal allograft survival as their freshly-cultured counterparts in the pre-sensitized high-risk model. Furthermore, thawed cryopreserved third-party allo-MSC could be co-administered with mycophenolate mofetil without adversely affecting their immunomodulatory function. In conclusion, a clinically-relevant protocol consisting of two intravenous infusions of third-party allo-MSC during the week prior to transplantation, exerts a potent anti-rejection effect in a pre-sensitized rat model of high-risk corneal allo-transplantation. This immune regulatory effect is likely to be mediated in the immediate post-transplant period through the promotion, by allo-MSC, of alternatively-activated macrophages in the lung and, later, by enhanced regulatory T-cell numbers.
Article
Full-text available
Aims Mesenchymal stem cells (MSCs) hold significant promise as potential therapeutic candidates following cardiac injury. However, to ensure survival of transplanted cells in ischemic environment, it is beneficial to precondition them with growth factors that play important role in cell survival and proliferation. Aim of this study is to use interleukin‐7 (IL‐7), a cell survival growth factor, to enhance the potential of rat bone marrow MSCs in terms of cell fusion in vitro and cardiac function in vivo. Methods MSCs were transfected with IL‐7 gene through retroviral vector. Normal and transfected MSCs were co‐cultured with neonatal cardiomyoctes (CMs) and cell fusion was analyzed by flow cytometry and fluorescence microscopy. These MSCs were also transplanted in rat model of myocardial infarction (MI) and changes at tissue level and cardiac function were assessed by histological analysis and echocardiography, respectively. Results Co‐culture of IL‐7 transfected MSCs and CMs showed significantly higher (p < 0.01) number of fused cells as compared to normal MSCs. Histological analysis of hearts transplanted with IL‐7 transfected MSCs showed significant reduction (p < 0.001) in infarct size and better preservation (p < 0.001) of left ventricular wall thickness as compared to normal MSCs. Presence of cardiac specific proteins, α‐actinin and troponin‐T showed that the transplanted MSCs were differentiated into cardiomyocytes. Echocardiographic recordings of the experimental group transplanted with transfected MSCs showed significant increase in the ejection fraction and fractional shortening (p < 0.01), and decrease in diastolic and systolic left ventricular internal diameters (p < 0.001) and end systolic and diastolic volumes (p < 0.01 and p < 0.001, respectively). Conclusion IL‐7 is able to enhance the fusogenic properties of MSCs and improve cardiac function. This improvement may be attributed to the supportive action of IL‐7 on cell proliferation and cell survival contributing to the regeneration of damaged myocardium. This article is protected by copyright. All rights reserved.
Article
We have examined the role of Fibroblast Growth Factor 10 (FGF10) during the growth and development of the rat ventral prostate (VP) and seminal vesicle (SV). FGF10 transcripts were abundant at the earliest stages of organ formation and during neonatal organ growth, but were low or absent in growth-quiescent adult organs. In both the VP and SV, FGF10 transcripts were expressed only in a subset of mesenchymal cells and in a pattern consistent with a role as a paracrine epithelial regulator. In the neonatal VP, FGF10 mRNA was expressed initially in mesenchymal cells peripheral to the peri-urethral mesenchyme and distal to the elongating prostatic epithelial buds. At later stages, mesenchymal cells surrounding the epithelial buds also expressed FGF10 transcripts. During induction of the SV, FGF10 mRNA was present in mesenchyme surrounding the lower Wolffian ducts and, at later stages, FGF10 transcripts became restricted to mesenchymal cells subadjacent to the serosa. We investigated whether the FGF10 gene might be regulated by androgens by analysing the levels of FGF10 transcripts in SV and VP organs grown in serum-free organ culture. While FGF10 transcript levels increased after treatment with testosterone in the SV (but not VP), these changes were not sensitive to anti-androgen treatment, and thus it is likely that FGF10 mRNA was not directly regulated by testosterone. Also, FGF10 mRNA was observed in the embryonic female reproductive tract in a position analogous to that of the ventral prostate in males suggesting that FGF10 is not regulated by androgens in vivo. Recombinant FGF10 protein specifically stimulated growth of Dunning epithelial and BPH1 prostatic epithelial cell lines, but had no effect on growth of Dunning stromal cells or primary SV mesenchyme. Furthermore, FGF10 protein stimulated the development of ventral prostate and seminal vesicle organ rudiments in serum-free organ culture. When both FGF10 and testosterone were added to organs in vitro, there was no synergistic induction of development. Additionally, development induced by FGF10 was not inhibited by the addition of the anti-androgen Cyproterone Acetate demonstrating that the effects of FGF10 were not mediated by the androgen receptor. Taken together, our experiments suggest that FGF10 functions as a mesenchymal paracrine regulator of epithelial growth in the prostate and seminal vesicle and that the FGF10 gene is not regulated by androgens
Article
Background and aim: Bone marrow mesenchymal stem cells (MSC) are receiving increasing attention for skin wound repair. However, the specific mechanisms underlying MSC-mediated improvement in wound healing have not been fully elucidated. This study aims at testing whether epidermal growth factor (EGF) can promote MSC-mediated wound healing and hair follicle regeneration. Methods: Excisional wounds in rats were transplanted with different collagen-chitosan scaffolds: control, MSC, and MSC + EGF. Regenerated tissues were harvested 1, 3, or 5 weeks following transplantation, stained with hematoxylin and eosin and evaluated microscopically. The formation of sebaceous glands was examined by Oil red staining and the regeneration of hair follicles by immunohistochemical staining and Western blot to test the expression of hair follicle-specific factors. Results: Gross observations showed that the wounds were much smaller and the hairs grew faster in the MSC + EGF group. Histological analysis demonstrated that there were more hair follicles, sebaceous glands, and newly formed blood vessels in the MSC + EGF group compared with that in the MSC group. In addition, oil red staining showed that MSCs + EGF induced sebaceous gland regeneration. Finally, immunohistochemistry and western blot revealed that MSCs + EGF increased the expression of hair follicle-specific factors. Conclusion: MSCs alone cannot achieve the regeneration of hair follicles and EGF can promote MSC-mediated wound healing and hair follicle regeneration.
Article
It was shown as long as half a century ago that bone marrow is a source of not only hematopoietic stem cells, but also stem cells of mesenchymal tissues. Then the term of “mesenchymal stem cells” (MSCs) has been coined in early 1990s and over a decade later, the criteria for defining MSCs have been released by International Society for Cellular Therapy. The easy derivation from a variety of fetal and adult tissues and not demanding cell culture conditions made MSCs an attractive research object. It was followed by the avalanche of reports from preclinical studies on potentially therapeutic properties of MSCs, such as immunomodulation, trophic support and capability for a spontaneous differentiation into connective tissue cells, and differentiation into majority of cell types upon specific inductive conditions. Although ontogenesis, niche, and heterogeneity of MSCs are still under investigation, there is a rapid boost of attempts in clinical applications of MSCs, especially for a flood of civilization‐driven conditions in so quickly aging societies in not only developed countries, but also very populous developing world. The fields of regenerative medicine and oncology are particularly extensively addressed by MSC applications, in part due to paucity of traditional therapeutic options for these highly demanding and costly conditions. There are currently almost 1,000 clinical trials from entire world registered at clinicalTrials.gov and it seems that we are starting to witness the snowball effect with MSCs becoming a powerful global industry; however, spectacular effects of MSCs in clinic still need to be shown. Stem Cells 2019 Mesenchymal stem cells: from isolation to transplantation.
Article
With chronic wounds remaining a substantial healthcare issue, new therapies are sought to improve patient outcomes. Various studies have explored the benefits of promoting angiogenesis in wounds by targeting proangiogenic factors such as Vascular Endothelial Growth Factor (VEGF) family members to improve wound healing. Along similar lines, Mesenchymal Stem Cell (MSC) secretions, usually containing VEGF, have been used to improve angiogenesis in wound healing via a paracrine mechanism. Recent evidence for keratinocyte VEGF receptor expression, as well as proliferative and chemotactic responses by keratinocytes to exogenous VEGFA in vitro implies distinct non-angiogenic actions for VEGF during wound healing. In this review, we discuss the expression of VEGF family members and their receptors in keratinocytes in relation to the potential for wound healing treatments. We also explore recent findings of MSC secreted paracrine wound healing activity on keratinocytes. We report here the concept of keratinocyte wound healing responses driven by MSC-derived VEGF that is supported in the literature, providing a new mechanism for cell-free therapy of chronic wounds.