ArticlePDF Available

Heterogeneity of mesenchymal and pluripotent stem cell populations grown on nanogrooves and nanopillars

Authors:

Abstract and Figures

Surface nanotopographies are an important way of mimicking the stem cell niche on biomaterial surfaces. Previous studies have focused on the differentiation of stem cell into a defined lineage using...
Content may be subject to copyright.
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7927
Cite this: J. Mater. Chem. B, 2017,
5,7927
Heterogeneity of mesenchymal and pluripotent
stem cell populations grown on nanogrooves
and nanopillars
Peng-Yuan Wang, *
ab
Sheryl Ding,
a
Huseyin Sumer,
a
Raymond Ching-Bong Wong
c
and Peter Kingshott
a
Surface nanotopographies are an important way of mimicking the stem cell niche on biomaterial
surfaces. Previous studies have focused on the differentiation of stem cells into a defined lineage using
nanotopographies, but they have rarely considered the homogeneity of cell populations produced.
We examined the impact of two types of substrates (i.e. nanogrooves and nanopillars made by soft
lithography) on the surface-induced differentiation of human amniotic membrane-derived mesenchymal
stem cells (hAM-MSCs) and mouse embryonic stem cells (mESCs) without the use of additional chemical
induction medium components. Cell morphology and proliferation were analysed at day 1 and day 3.
Gene expression was analysed at day 14 for hAM-MSCs and at day 7 for mESC-derived embryoid bodies
(mEBs) using quantitative real-time polymerase chain reaction (qPCR). The substrates with nanogrooves
had a noticeable effect on cell alignment in a depth dependent manner with both cell types showing
strong alignment along the deep grooves. On the other hand, the nanopillar substrates showed
inhibition of cell spreading for both cell types. The nanogrooves showed inhibition of hAM-MSC growth
but enhanced mEB proliferation, especially on the deeper grooves. The nanopillars did not significantly
affect hAM-MSC growth, but can modulate mEB growth depending on the pillar density, indicating that
mEBs are more sensitive to nanotopographies in terms of proliferation, while hAM-MSCs are only
sensitive to specific structures and sizes. Genes associated with bone, cartilage, and fat were
investigated for hAM-MSCs, whereas genes of the endoderm, mesoderm, ectoderm, and pluripotency
were investigated for mEBs. In general, gene expression for hAM-MSCs was not enhanced significantly
by the nanotopographies compared to the flat control. On the other hand, genes of bone, cartilage,
skeletal muscle, heart, and liver were up-regulated on both nanopillars and nanogrooves, especially
OP65 (ordered pillars with 65% density) and SG40 (shallow grooves with 40 nm depth) in a feature size
dependent manner. We found that a small portion of mEBs was composed of cardiac-like beating cells
(i.e. GFP-NKX2.5 positive) and a bone cell marker (i.e. OCN) indicating a heterogeneous cell population
being generated on those types of surfaces. This work highlights the importance of nanotopographies in
stem cell differentiation and how studying multiple properties of the substrate and cells is needed as we
strive to generate homogeneous and mature cell populations using biomaterials.
Introduction
Stem cell technology is an area of great excitement, with the
potential to revolutionise regenerative medicine and biomedical
research.
1
Advanced stem cell technologies have the potential
for future development of patient-specific cell therapies, tissue
regeneration, and/or more accurate research models. Stem cells
are characterised by their prolonged proliferation and ability
to differentiate into a range of cell lineages, depending on
the potency and source of the cells. Pluripotent stem cells, such
as embryonic stem cells (ESCs) derived from the blastocyst
inner cell mass, are able to differentiate into any cell type.
a
Department of Chemistry and Biotechnology, Faculty of Science, Engineering and
Technology, Swinburne University of Technology, Hawthorn, Victoria 3122,
Australia. E-mail: pengyuanwang@swin.edu.au; Tel: +61 3 9214 5939
b
Graduate Institute of Nanomedicine and Medical Engineering, College of
Biomedical Engineering, Taipei Medical University, Taipei 110, Taiwan
c
Centre for Eye Research Australia & Ophthalmology, Department of Surgery,
The University of Melbourne, Victoria 3004, Australia
Electronic supplementary information (ESI) available. See DOI: 10.1039/
c7tb01878a
Received 12th July 2017,
Accepted 22nd August 2017
DOI: 10.1039/c7tb01878a
rsc.li/materials-b
Journal of
Materials Chemistry B
PAPER
7928 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 201 7
Multipotent stem cells, such as amniotic membrane-derived
mesenchymal stem cells (AM-MSCs), are more limited in their
differentiation potential, but have their own advantages such as
safety.
2
Although stem cell technologies including in vitro
expansion have been investigated over the last decade, there
are still major barriers to be overcome such as the maintenance
of stemness during expansion and maturation of differentiated
stem cells before stem cells can become widely used in clinical
or commercial applications.
3
One such challenge is the ability
to modulate stem cell behaviour and differentiate stem cells
into a pure and mature cell population on a large scale with
high efficiency and cost-effectiveness.
Biomaterials modified with surface nanotopographies are
showing huge potential in directing stem cell behaviour and
developing better cell culture materials along with biomechanical
stimulation.
4
Surface nanotopographies can be well-defined for
stem cell manipulation compared to traditional biochemical
methods. This biophysical approach has some advantages such
as a longer lifetime, easier characterisation, higher stability,
and better reproducibility with little batch-to-batch variation
compared to biochemical approaches.
5
Nanotopographical
features can be used to mimic the in vivo stem cell niche present
on the extracellular matrix (ECM), which plays a role in the
coordination of cell morphology, self-renewal, migration, differ-
entiation, and many other functions.
1
Studies have shown that
surface nanotopographies can have striking influences on
almost all stem cell behaviour in vitro.
6–10
However, systematic
studies are still required, owing to the vast range of possible
topographical patterns being discovered, different stem cell
sources, and combinational stimulations using both biochemical
and biophysical cues, resulting in difficulties comparing results
between studies. Also, studies have often focused on one specific
cell type generated from one type of stem cell using one type of
surface structure, and often lacks information such as the homo-
geneity of surface-induced differentiated cell populations.
For example, Yim et al. used 350 nm-width PDMS nano-
grooves to study neurogenic differentiation of human mesenchymal
stem cells.
11
A positive outcome was found using 350 nm nano-
grooves compared to 10 mm-width microgrooves and flat controls
in neural differentiation. Another independent study showed that
microgrooves can improve cardiac differentiation from either
cardiac progenitors
12
or MSCs.
13
In addition, there was other
evidence showing that aligned MSCs expressed higher markers
of anisotropic tissue lineage such as tendon.
14
The exact effect of
nanogrooves on lineage commitment is unclear. Studies have
also used diverse nanotopographies, feature sizes, materials,
surface coatings, and cell sources resulting in difficulties in
elucidating the exact effects of the nanotopography.
In this study, we examined the effect and homogeneity of
surface-induced differentiation of stem cells, i.e. human amniotic
membrane-derived mesenchymal stem cells (hAM-MSCs) and
mouse embryonic stem cells (mESCs), on nanotopographies
consisting of two types of nanogrooves and four types of nano-
pillars. These were fabricated using a combination of colloidal
self-assembly, spin coating, reactive ion etching, electron beam
lithography, and soft lithography (Fig. 1). The morphology,
proliferation, and gene expression of the two stem cell lines
were analysed on the various surface topographies without
additional chemical or growth factor stimulation. Thus, the
role of surface-only induction of stem cell differentiation can be
ascertained and this systematic study explores this concept,
which is often overlooked in biomaterials and stem cell engi-
neering research.
Materials and methods
Materials
Polystyrene (PS) was purchased from Nihon Shiyaku Industries
(Tokyo, Japan). Polydimethylsiloxane (PDMS; Sylgard 184) was
received from Dow Corning (Midland, MI, USA). 40,6-Diamidino-
2-phenylindole dihydrochloride (DAPI) and phalloidin-tetra-
methylrhodamine B isothiocyanate (phalloidin-TRITC) were
purchased from Thermo Fisher Scientific (Waltham, MA, USA).
Human mesenchymal stem cells were obtained from amniotic
membranes (hAM-MSCs) during delivery by caesarean section,
using a previously described protocol,
15
and were provided by
Prof. Richard Manasseh at Swinburne University of Technology.
hAM-MSC culture medium was composed of mesenchymal stem
cell growth medium (MSCGMt, Lonza, AU) supplemented with
MSCGMtSingleQuotst(Lonza). Murine embryonic stem cells
(cardiac-specific NKX2.5-eGFP transgenic mESCs) were generated
according to a previous study,
16
and were provided by Dr Huseyin
Sumer at Swinburne University of Technology. mESC cell culture
medium was composed of 90% (v/v) Dulbecco’s modified Eagle’s
medium (DMEM, 10569077, Thermo Fisher) supplemented with
10% ES screened fetal bovine serum (FBS; Thermo Fisher),
1mlmL
1
leukemia inhibitory factor (LIF; PMC9484, Sigma-
Aldrich, Castle Hill, NSW, AU), 1MEM NEAA (11140050,
Thermo Fisher), 1 mlmL
1
2-mercaptoethanol (21985023, Thermo
Fisher), and 1antibiotic antimycotic solution (A5955, Sigma-
Aldrich). 1st anti-vinculin (V19131), FITC-2nd-antibody (F2012),
and TRITC-2nd-antibody (T5393) were purchased from Sigma-
Aldrich. 1st anti-OCN (AB10911) was purchased from Merck, AU.
For the formation and culture of embryoid bodies (EBs) from
mESCs, the cell culture medium composition was identical except
that LIF was excluded.
Fabrication of surface nanotopographies
Nanogrooved substrates were fabricated using a combination
of electron beam lithography (EBL) and soft lithography as
described in a previous study.
17
In brief, grooved substrates
with 800 nm width/400 nm depth or 800 nm width/40 nm depth
were used and named ‘‘deep nanogrooves’’ (DG400) and ‘‘shallow
nanogrooves’’ (SG40), respectively. Nanopillar substrates
were fabricated using colloidal lithography (CL) and then soft
lithography. An ordered colloidal mask was fabricated using
self-assembly of 820 nm PS colloids into close packed mono-
layers on Si wafers according to a previous protocol.
18
Random
colloidal masks were produced by spin coating of different
concentrations of colloidal solutions on Si wafers (1 :4, 1 : 8 and
1 : 12 in water/ethanol) at 2500 rpm for 1 min. Reactive ion
Paper Journal of Materials Chemistry B
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7929
etching (RIE; Oxford Instruments PLASMALAB100 ICP380,
MCN, Melbourne, Australia) was then applied on four different
colloidal masks with a gas flow of Cl
2
(50 sccm), SF
6
(3 sccm)
and O
2
(3 sccm) for 60 s, resulting in four nanopillar patterns:
ordered nanopillars (OP65), high density nanopillars (HP50),
medium density nanopillars (MP20), and low density nano-
pillars (LP10) with a 500 nm pillar height characterized using
contact model profilometry (Ambios XP 200, MCN, Melbourne,
Australia). Any residual PS colloids were removed from the
surfaces by using acetone and then ethanol under ultrasonic
agitation.
Soft lithography was then used to replicate the master patterns
according to a previous study.
19
In brief, master silicon patterns
were coated with octadecyltrichlorosilane (OTS)/toluene for
15 min, and then rinsed with toluene, dried in air, and rinsed
in 100% ethanol. The OTS-coated wafers were heated for
2 minutes at 120 1C. An approx. 5 mm layer of PDMS solution
(1-to-10: curing-to-base agent) was poured over the master
patterns, degassed in a vacuum desiccator for 1 h, and then
cured at 60 1C for 2 h. The PDMS stamps were removed from
the master patterns for making nanopatterns. Nanopatterns
were fabricated on either poly(ethylene teraphthalate) (PET) or
glass slides with a size of approx. 1.5 cm 1.5 cm. For glass
substrates, a thin layer of PS was first spin coated, and then
heated at 120 1C for 5 min. A drop of 5% PS/toluene was
imprinted by PDMS stamps on either PET or glass slides and
dried overnight in a hood. Surface nanotopographies were
examined using field emission scanning electron microscopy
(FE-SEM, ZEISS SUPRA 40 VP, Carl Zeiss, Germany) at 5 keV and
atomic force microscopy (Dimension Icon, Bruker, Germany) at
0.5 Hz (tapping model). The density and dimensions of nano-
pillars and nanogrooves were determined using SEM and AFM
images (n= 5).
Stem cell culture and analysis
Prior to cell culture, all substrates were treated with air
plasma (air flow rate of 2.0 10
2
mbar and 50 W power)
for 2 min, sterilised in 70% ethanol, and washed twice in
Fig. 1 (A) Fabrication of nanopillars and nanogrooves. Colloidal lithography and spin coating based approaches were used for nanopillars, while electron
beam lithography was used for nanogrooves. (B) Scanning electron microscopy (SEM) and atomic force microscopy (AFM) images of different
nanotopographies. AFM images were scanned over a 5 5mm area.
Journal of Materials Chemistry B Paper
7930 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 201 7
phosphate-buffered saline (PBS). All substrates were placed in
12-well plates for cell culture. Each surface was seeded with
hAM-MSCs at a density of 1 10
4
cells per cm
2
in 1 mL media.
Embryoid bodies of mESCs (mEBs) were formed by culturing
mESCsfor3daysinlow-attachment well plates (Corning, USA)
without LIF. Prior to cell culture, substrates were coated with
0.1% gelatin for 1 h at room temperature, and approx. eight
EBs were seeded on each surface.
Cell morphology was examined after 1 and 3 days using
fluorescence microscopy (Eclipse Ti-E microscope, Nikon, Japan),
confocal laser scanning microscopy (FV3000, FLUOVIEW,
Olympus, Japan) and FE-SEM. For fluorescence staining,
samples were fixed with 4% paraformaldehyde for 30 minutes,
permeabilised by 0.2% Triton X-100/PBS, and then washed with
PBS. After 24 h, focal adhesions of cells were stained using 1st
anti-vinculin (1 : 100 in PBS) and FITC-2nd antibody (1 : 200 in
PBS) for 1 h. After 1 week, immunostaining of osteocalcin
(OCN) of mEBs was done using 1st anti-OCN (1 : 100 in PBS)
and TRITC-2nd antibody (1 : 200 in PBS) for 1 h. Cell nuclei
and F-actin were stained using 3 mM DAPI and 0.165 mM
phalloidin–TRITC, respectively. Samples were imaged at excitation/
emission wavelengths of 346/442 nm for DAPI, 405/519 nm for
FITC, and 555/565 nm for TRITC.
For SEM imaging, samples were fixed with 4% paraform-
aldehyde and dehydrated by immersion in graded ethanol
at concentrations of 50%, 70%, 80%, 90%, 95%, and absolute
ethanol. Samples were dried and sputter coated with a 5 nm
gold layer for SEM imaging.
The alignment of cells was analysed using ImageJ according
to a previous study.
20
In brief, the cell or colony outline was
circled. The major and minor axes of cells or colonies were
determined automatically, so that an angle of major axis between
0 and 180 degrees was obtained. All angles were converted
between 0 and 45 degrees, where 0 degrees indicates prefect
alignment, while 45 degrees indicates a random distribution of
cells or colonies.
To determine proliferation, cell numbers or colony sizes were
determined after 1 and 3 days using fluorescence microscopy.
Growth rate is the number of cells at day 3 divided by that at day 1.
For hAM-MSCs, the cell number was counted using DAPI-stained
images. For mEBs, colony areas were measured using phalloidin-
TRITC-stained images. Five images were taken randomly from
each sample at a magnification of 10(n=5).
Quantitative real-time polymerase chain reaction (qPCR)
Gene expression of the different lineages was analysed at
specific time points. hAM-MSCs were analysed on various
nanotopographies in normal culture media after 14 days, while
mEBs were analysed on various surfaces in normal culture
media without LIF after 7 days. In normal culture medium,
surface-induced differentiation of MSCs was conducted for a
longer time (i.e. 2 weeks), while that of EBs was conducted for a
shorter time (i.e. 1 week) because spontaneous differentiation
of EBs was expected. At each time point, mRNA was extracted
using the RNeasy Plus Mini kit (Qiagen, Hilden, Germany).
qPCR was conducted using an iTaqtUniversal One-Step
RT-qPCR kit (Bio-Rad Laboratories, California, USA) and a
CFX96 TouchtReal-Time PCR Detection System thermocycler
(Bio-Rad Laboratories, California, USA). The primers of selected
genes are listed in the ESI(Table S1). Three lineage genes of
hAM-MSCs were analysed and the various genes selected
included aggrecan (ACAN), type II collagen (COL2) and SOX9
for chondrocytes, type I collagen (COL1), alkaline phosphatase
(ALP) and RUNX2 for osteoblasts; and PPARgfor adipocytes. Five
lineage genes of mEBs were analysed and the various genes
selected included COL1 and OCN for osteoblasts, COL2 for
chondrocytes, GATA6,GATA4,NKX2.5 and myosin heavy chain
(MHC) for cardiomyocytes, myogenin (MYOG) for skeletal
myoblasts, albumin (ALB) and alpha-fetoprotein (AFP) for
hepatocytes, NESTIN for neural cells, platelet-derived growth
factor receptor-A (PDGFR) for mesodermal tissues, and Kruppel-
like factor 4 (KLF4) and OCT4 for pluripotency. GAPDH was used
as the housekeeping gene. Gene expression on nanopatterns
was normalised to tissue culture polystyrene (TCPS; value = 1)
and compared to the flat control. Experiments were repeated
at least three times. Each sample was triplicated with three
technical repeats (n= 3). The DDCt method was used to
calculate and compare relative mRNA levels according to a
previous study.
21
Statistical analysis
All experiments were repeated at least three times. Data
are expressed as mean standard deviation (STDEV) unless
otherwise specified. Statistical analysis was performed for each
experimental group by one-way ANOVA and Student–Newman–
Keuls multiple comparison tests using GraphPad InStat V.3
(GraphPad Software, Inc.).
Results
Nanopatterns
The surface nanotopographies were fabricated using a combi-
nation of approaches and characterized using SEM and AFM
imaging (Fig. 1). The nanogrooves were designed to have a
width of 800 nm (equal groove-to-ridge ratio) and a depth of
either 40 nm or 400 nm. However, the obtained nanopattern
sizes were slightly different due to the fabrication process.
Nevertheless, the difference in groove depth was still clearly
visible. Two types of nanogrooves were created: shallow grooves
(SG40) and deep grooves (DG400).
Similarly, the size of nanopillars was slightly different after
RIE and soft lithography. Colloidal masks were made from
820 nm PS particles. After fabrication, the diameter of nano-
pillars was approx. 500 nm and the height of nanopillars was
approx. 450–750 nm depending on the particle density (Fig. 1
and Table 1). A higher particle density appears to protect the Si
substrate during etching resulting in a shorter pillar height.
The density of nanopillars decreased from 65% to 10% (Fig. 1
and Table 1). Thus, four types of nanopillars were created:
ordered nanopillars (OP65), high density nanopillars (HP50), medium
density nanopillars (MP20), and low density nanopillars (LP10).
Paper Journal of Materials Chemistry B
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7931
The surface roughness of OP65 was the lowest (approx. 73 nm in
Rq), while that of MP20 was the highest (approx. 273 nm in Rq;
Table 1).
Two stem cell types, hAM-MSCs and mEBs, were studied
on the nanopatterns. The cell morphology and cell–surface
interactions were analysed after 1 and 3 days (Fig. 2 and
Fig. S1, ESI). Fluorescence staining showed that cell alignment
of hAM-MSCs was induced only by the deep nanogrooves
(i.e. DG400), and not others (Fig. 2A). The density of nanopillars
on the surfaces also had an effect on cell morphology, where
cell spreading was inhibited on lower density patterns, resulting
in more elongated cells (arrows in HP50 and LP10 in Fig. 2A).
Similarly, mEBs elongated and aligned with the deep nanogrooves
(DG400,Fig.2B),andthemEBshadmore3D-likemorphologies
on the nanopillars compared to the flat control (arrows in Fig. 2B).
That is, they extended from the tops of the pillars to the bare
substrate regions.
Morphological analysis of the two cell types showed that
the DG400 substrate caused alignment of hAM-MSCs (DG400:
angle B171, Fig. 2C) with a more elongated morphology
(DG400: elongation B5, Fig. 2D), which was significantly
different from the flat and other surfaces (po0.001 vs. flat,
Fig. 2C and D). mESC colonies showed cell alignment (DG400:
angle B191, Fig. 2D) and elongation (DG400: elongation
B2, Fig. 2D) along the DG400 substrate (Fig. 2D). With both
hAM-MSCs and mEBs, there was no significant difference in
cell alignment or elongation associated with the nanopillar
patterns (except for hAM-MSCs on LP10) compared to the flat
control.
The growth of hAM-MSCs, not mEBs, was inhibited on the
DG400 substrate compared to the flat control (Fig. 2E and
Fig. S3, ESI). On nanopillars, hAM-MSC growth was slightly
affected by the nanotopographies without significance. The
mEB size was however inhibited by OP65 and LP10 nanopillars,
but increased on the HP50 nanopillars (Fig. 2E and Fig. S3, ESI).
The details of cell–surface interactions were imaged on day 3
(Fig. 3). The filopodia or extensions of the hAM-MSCs and
mEBs ‘‘touched’’ the floor of the DG40 surfaces because the
height of the grooves was shallow. The majority of filopodia
and cell bodies of the two cell types attached to the top of ridges
of the DG400 surfaces due to the high aspect ratio and narrow
ridge spacing (Fig. 3A). This phenomenon resulted in the
alignment and elongation of the two cell types on the DG400,
not DG40, surfaces. Interestingly, hAM-MSCs and mEBs cultured
on the OP65 and HP50 surfaces attached only to the uppermost
surfaces of the nanopillars (solid arrows, Fig. 3B). On the HP
surfaces, the main body of the cells remained on the top of
nanopillars and the filopodia seemed to be guided by the pillar
arrangement (HP50, Fig. 3A and B). On the other hand, on the
MP20 and LP10 surfaces, both cell types interacted with both
the nanopillars and substrate floor because the pillar density
was too low to allow focal adhesions to form on the nanopillars
alone (empty arrows, Fig. 3A and B). Nanopillars can be seen
through the cells in the thinner parts of the cell body indicating
that nanopillars were accommodated by attached cells and
possibly semi-penetrated the cells.
Vinculin staining showed that focal adhesion was affected
by the underlying nanotopographies (Fig. 4). Both hAM-MSCs
and mESCs had clear aligned F-actin and vinculin distributions
on the DG400 surfaces. hAM-MSCs had a similar vinculin
distribution on OP65 and the flat control (Fig. 4A). mESCs
had a clear vinculin distribution at the periphery of cells, while
the distribution on the flat control was not obvious (Fig. 4B).
Gene expression of stem cells on nanotopographies
Nanotopography-induced differentiation was characterized
using qPCR. Various genes from different lineages were analysed.
Changes greater than 2-fold compared to the flat control were
considered significant and highlighted. For hAM-MSCs, ALP and
PPARgexpression was slightly enhanced on nanopatterns, while
other genes either had no change (i.e. COL1,RUNX2,andSOX9)or
decreased (i.e. ACAN) without significant changes compared to the
flat control (Fig. 5A). The COL2 gene was not expressed (i.e. most
of Ct values were 438; Fig. S2, ESI).
For mEBs, the expression of OCN,MYOG and ALB increased
on the nanogrooves and the fold was dependent on the groove
depth (Fig. 5B). On nanopillars, the expression of COL1,OCN,
COL2,MYOG, and ALB increased and the fold was dependent
on the pillar density (Fig. 5B). Interestingly, the expression of
GATA4 decreased on DG400, HP50, and LP10 significantly
(Fig. 5B). Two pluripotent genes, i.e. OCT4 and KLF4, and
the hepatocyte marker AFP were downregulated on nanotopo-
graphies compared to TCPS, but no significant difference
occurred compared to the flat control. Two neural genes, i.e.
PDGFR and NESTIN, were upregulated on nanogrooves and the
OP65 nanopillar surface. Genes including MHC,GATA6 and
NKX2.5 for mESCs were not expressed (i.e. Ct 438) as shown in
Fig. S2 (ESI).
Heterogeneity of stem cells on nanotopographies
Multiple lineage genes were up-regulated on the nanotopo-
graphies especially on OP65 (Fig. 6A). Co-expression of OCN
and NKX2.5 proteins within one colony was found on OP65 and
the flat control, which was consistent with the gene expression
Table 1 Characterisation of nanotopographies. Value = mean STDEV (n=5)
OP65 HP50 MP20 LP10 SG40 DG400
Pillar density (% area) 65.5 1.5 50.2 1.7 21.8 0.6 10.3 0.5 —
Pillar diameter (nm) 542.1 17.9 482.6 32.8 447.7 54.8 478.6 34.3 —
Height (nm) 445.6 23.0 463.7 21.6 761.3 31.7 732.9 12.8 48.5 0.1 389.8 0.6
Roughness (nm) 72.9 208 273 217 22.7 181
Groove width (nm) — — — — 802.7 66.9 924.2 23.6
Ridge width (nm) — — — — 649.9 38.7 516.7 23.3
Journal of Materials Chemistry B Paper
7932 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 20 17
Fig. 2 Cell morphology of (A) hAM-MSCs and (B) mEBs after 3 days. The arrows in G400 indicate the direction of the nanogrooves. Full data are shown
in the ESI(Fig. S1). BF: bright field; fluore: DAPI and phalloidin–TRITC staining. Nanotopography-induced morphological changes of (C) hAM-MSCs and
(D) mEBs after 3 days. All measurements were taken from fluorescence microscope images at day 3 of cell culture. The alignment (degrees) and elongation
(length-to-width) of single hAM-MSCs and mEBs were analysed. Value = mean SEM (n=74238).*,po0.05; **, po0.01; ***, po0.001 vs. flat.
(E) Growth rate of hAM-MSCs and colony size change of mEBs after 3 days. For hAM-MSCs, the cell number at day 3 was compared to that at day 1.
For mEBs, the colony size at day 3 was compared to that at day 1. Value = mean SEM (n=1012).*,po0.05; **, po0.01; ***, po0.001 vs. flat.
Paper Journal of Materials Chemistry B
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7933
data (Fig. 6B). A small portion of mEB-derived beating cells
(GFP positive cells, Video S1B and D, ESI) were found within
one colony on the surfaces (e.g. OP65 and DG400 Fig. 6C and
Video S1A–D, ESI) indicating that the cell population was
diverse. Specific nanotopographies, for example OP65, can thus
induce stem cell differentiation compared to the flat control,
but may not be strong enough to guide mESCs into one specific
lineage.
Discussion
Surface nanotopography-induced morphological changes and
differentiation of stem cells have previously been studied.
1
However, the majority of the work has focused on one type of
surface structure, one type of stem cell and/or one downstream
differentiated cell type. Various cell types, different cell culture
protocols, and the diverse types of surfaces/materials used lead
to difficulties in drawing clear conclusions about the specific
role played by nanotopography. For example, Pan et al. reported
that 350 nm-width nanogrooves enhanced neurogenesis
(i.e. increase of neuronal markers) for hiPSCs (i.e. human
induced pluripotent stem cells) compared to microgrooves and
flat controls.
22
However, analysis of other genes was missing,
although the cell morphology was neuron-like. Another study
Fig. 3 SEM images of cell–nanotopography interactions. The images
show the interaction of (A) hAM-MSCs and (B) mEBs with the nanotopo-
graphies. Double-headed arrow indicates groove direction. Arrowheads
indicate the detailed location between the cells (i.e. filopodia and/or cell
extensions) and the nanoscale features.
Fig. 4 Confocal images of stem cells on nanotopographies. (A) AM-MSCs
and (B) mESCs were cultured on DG400, OP65, and the flat control for
24 h. Cell nucleus, F-actin, and vinculin were stained using DAPI (blue),
TRITC (red), and FITC (green). White arrows indicate the clear F-actin and
vinculin overlapping.
Journal of Materials Chemistry B Paper
7934 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 201 7
showed that 250 nm-diameter nanopillars enhanced chondro-
genic differentiation of hMSCs. Again, the study did not analyse
other mesoderm genes such as adipocytes.
23
To address this, in
this systematic study we investigated the effects of two types of
surface nanostructures made by a combination of approaches
and examined expression of genes associated with various
lineages of the two stem cell types. The aim was to understand
whether a specific nanotopography can stimulate stem cells into
one specific cell type in early stages of growth without chemical
induction.
Colloidal lithography is a cost-effective approach which uses
particle monolayers as masks to generate nanostructures after
etching the substrate.
18,24
Surprisingly, very few studies have
used colloidal lithography to fabricate nanostructures for cell
culture. One possible reason is that the ability of controlling
colloidal self-assembly for the fabrication of either close packed
or randomly distributed particle masks has been lacking.
In this study, we used an exquisite approach to assemble
submicron polystyrene particles into a monolayer at the air–
water interface, and then deposited the monolayer on a silicon
wafer. Other masks with decreasing particle densities were
prepared using spin coating. Then, reactive ion etching and
soft lithography were used for the production of surface nano-
topographies. To the best of our knowledge, this is the first
time that various nanopillar densities generated in a large area
(2 2cm
2
) for stem cell culture have been made. This
combination of approaches and the types of nanostructures
produced are very useful in biomedical engineering as well as
other fields such as photonics and solar cells.
25
We previously demonstrated that cell alignment on nano-
grooves was mainly depth dependent (i.e. C2C12 myoblasts,
19,26,27
rat BM-MSCs,
17
mouse cardiomyocytes,
8
and porcine anterior
cruciate ligament cells
20
). In this study, the same phenomenon
was observed again on hAM-MSCs and mEBs. The morphology of
Fig. 5 Gene expression of different lineage markers of (A) hAM-MSCs and (B) mEBs on the different nanotopographies. hAM-MSCs were cultured on
nanopatterns for 14 days, while mEBs were cultured for 7 days in normal culture medium. All gene expression was normalized to TCPS (value = 1). The
gene expression of each gene on different surfaces was compared to that on the flat control, and the fold changes are indicated using red boxes. A gene
expression greater than 2-fold vs. flat was considered significant. *, 42-fold; **, 43-fold; ***, 44-fold. Value = mean STDEV (n= 3). Genes from bone
(COL1,ALP,RUNX2RUNX2), cartilage (SOX9,ACAN), and fat (PPARg) were selected for hAM-MSCs. Genes from bone (COL1,OCN), cartilage (COL2),
skeletal muscle (MYOG), heart (GATA4), nerve (NESTIN), mesodermal (PDGFR), liver (ALB,AFP) and pluripotency (KLF4,OCT4) were selected for mEBs.
Paper Journal of Materials Chemistry B
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7935
hAM-MSCs and spheroidal mEBs was elongated on DG400
(800 nm width/400 nm depth) indicating that the submicron
nanogrooves of this specific size are an effective modulator of
contact guidance. On the other hand, nanopillars do not provide
any specific guidance on cell direction. hAM-MSCs on nanopillars
have smaller cell sizes, especially on the HP50 and LP10 surfaces,
while mEBs on nanopillars seem to have more 3D structure,
especially on OP65 and LP10, indicating that nanopillars inhibit
the spreading of both types of stem cells. This phenomenon was
reported previously on different materials with nanopillar
28
and
nanoprotrusion
18,24
structures.
This inhibition changes subsequent cell proliferation, espe-
cially for mEBs. The elongation and alignment of hAM-MSCs
on DG400 decreases the growth rate after 3 days. On the
nanopillars, the growth rate of hAM-MSCs did not significantly
change compared to the flat control. On the other hand, the
growth rate, or more specifically the colony size, of mEBs,
increased on DG400 and HP50, but decreased on the OP65
and LP10 surfaces. There is also a correlation between cell
morphology and growth rate. On the DG400 and HP50 surfaces,
colonies appear flat, while on the OP65 and LP10 surfaces they
take on a more 3D-like structure. Decreasing the pillar density
from HP to LP seems to inhibit cell growth. However, the cell
growth rate on the OP65 surface is low and it seems that this is
another factor that interferes with cell growth such as the pillar
size. The pillar size between surfaces is different and the
symmetry of pillars could be another factor causing growth
inhibition.
10,29
Since we do not have a nanopattern having the
same pillar density but different ordering (due to fabrication
difficulties), the effect of symmetry needs to be studied in the
future if these types of patterns can be made. Nevertheless,
we prove here that nanotopography can modulate stem cell
morphology, which in turn changes the cell proliferation in a
cell type- and pattern-dependent manner.
Our gene expression results are interesting and show that
these nanotopographies do not improve hAM-MSC differentiation.
Although ALP,RUNX2,andPPARgare slightly higher and SOX9 and
ACAN are slightly lower on some patterns compared to the flat
control, none of them exhibit more than 2-fold changes compared
to the flat control. Previous studies reported that nanogrooves
suppress osteogenesis of osteoprogenitors.
30
Similarly, specific
pillar structures enhance osteogenesis of human mesenchymal
dental pulp-derived stem cells.
31
Compared to these previous
studies, we did not find the same outcome in this study. This is
possible due to the different cell types and pattern sizes used in
this study compared to previous ones. Overall, the nanogrooves
and nanopillars used in this study did not improve the differentia-
tion of hAM-MSCs toward any of the cell types of the mesenchymal
lineage.
On the other hand, mEBs exhibited higher sensitivity to
the different nanotopographies. Nanogrooves enhanced OCN,
NESTIN,PDGFR,MYOG and ALB expression, which are at least
2-fold greater compared to the flat control. Nanopillars, especially
OP65, enhanced COL1,OCN,MYOG,COL2,NESTIN,PDGFR,and
ALB expression, but inhibited GATA4 expression, compared to flat
controls in a surface dependent manner. Interestingly, three
cardiac genes, i.e. NKX2.5,GATA6,andMHC have no expression
on the nanopatterns, indicating that the selected nanotopo-
graphies were unable to stimulate cardiac differentiation of
mESCs. It is important to note that the OP65 surface can
stimulate mEBs to express multiple lineage markers without
biochemical stimulation. This outcome implies that biophysical
stimulation alone may not be strong enough for driving stem cell
fate specifically although the fate of stem cells was modulated.
Pluripotent stem cells such as mESCs routinely display hetero-
geneous gene expression in in vitro culture.
32
Aspects of
this heterogeneity are linked to differences in the propensity
of individual cells to either self-renew or commit towards
differentiation. On our nanotopographies, especially the OP65
and SG40 surfaces, the expressions of various lineage genes
were upregulated. This indicates that these two nanotopo-
graphies may induce stem cell differentiation but may not be
Fig. 6 Correlation of gene and protein expressions of mEBs. (A) Gene expression on OP65 normalized to TCPS and the flat control. (B) Immunostaining
of OCN (red) on OP65 and the flat control. Cell nuclei were stained using DAPI (blue) and NKX2.5 positive cells expressed GFP (green). White arrows
indicate strong GFP positive cells. (C) Video images show that a small portion of cells were cardiac-like beating cells, while other cells were non-beating
cells. Double-headed arrow indicates groove direction. Bright-field and GFP videos are available in the ESI(Video S1A–D).
Journal of Materials Chemistry B Paper
7936 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 2017
able to direct the differentiation into a specific lineage. It is
possible that either each mEB contains a heterogeneous cell
population or diverse mEBs are formed on the surface. To
confirm this, further studies using single colony or single cell
analysis such as fluorescence activated cell sorting (FACS) and
other emerging advanced technologies
33
are needed.
Previous studies have shown that extracellular matrix (ECM)
proteins can guide human bone marrow MSCs into different
lineages.
34
For example, collagen participates in the formation of
fibrils guiding the differentiation of cardiomyocytes, adipocytes,
and osteoblasts, while fibronectin does not significantly affect
stem cell differentiation.
34
Cells or stem cells can sense
both biochemical and biophysical signals surrounding them.
In the same way collagen can form fibrils to induce stem cell
differentiation into different cell types, nanogrooves and nano-
pillars are also present in the microenvironment of various
tissues. Thus, it is possible that one type of nanopattern
can induce multiple lineage differentiation of stem cells even
without the need for additional growth factors.
We propose a mechanism by which nanotopography can
affect stem cell function most likely indirectly through the
changes in the cytoskeleton and nucleus shape. In general,
biophysical cues are not as efficient as biochemical cues.
Biochemical cues such as small molecules or growth factors
can either bind to cell surface receptors or permeabilize the
cellular membrane to directly activate intracellular signalling
pathways. On the other hand, biophysical cues normally pre- or
post-exert stress and/or stretching on cells, which leads to
changes in focal adhesions, the cytoskeleton (i.e. cell morphology)
and/or the shape of the cell nucleus. These changes in turn can
affect intracellular molecular assembly and then cell function.
In this study, we found that the population of differentiated mESCs
was heterogeneous even with nanotopographical modulation. This
could be due to the fact that, in EBs, cell–cell contact has a
stronger effect on cell function than cell–surface interactions,
resulting in only a small portion of cells being influenced by
surface nanotopographies. Another possibility is that the dimen-
sions of selected nanostructures were not optimized. Since the
substrate solely affects the cell–substrate interface to indirectly
alter cell function, this effect may be minimized when cells do
not exist as a monolayer culture. Nevertheless, our results show
that the OP65 nanotopography stimulates early differentiation of
mESCs especially into the mesodermal lineage (i.e. bone and
muscle) compared to the flat control. Future studies that assess
the long-term effect of nanotopographies on stem cells would
allow us to further understand the effect of early stimulation of
gene and protein expression on the function of stem cells.
Conclusion
Substrate nanotopographies have a tremendous influence on the
morphology, migration, and proliferation of stem cells. However,
downstream events such as nanotopography-induced differen-
tiation remain unclear. Although some specific nanotopographies
such as nanogrooves have been suggested to be able to trigger
stem cell differentiation into, for example, neurons, it is unclear
to what extent nanotopographies contribute and whether or not
the cell population is homogenous. In this systematic study, we
demonstrated that nanogrooves and nanopillars can affect the
growth and proliferation of mESCs and hAM-MSCs with different
levels of cellular response occurring depending on the specific
nanogrooves and nanopillars. In terms of the cell populations,
nanotopography induced or inhibited mEB differentiation was
found to be dependent on the structure and size of the features.
We observed an increase of various lineage genes and markers
on nanotopographies, especially on OP65, indicating that the
cell population is heterogeneous. This study uses accessible
approaches for the fabrication of commonly nanostructures
found in the native extracellular matrix, and demonstrates that
hAM-MSCs are not sensitive to these nanotopographies, while
mEBs can be affected by nanotopographies but most likely
other factors will be needed such as biochemical factors to
precisely control the stem cell fate and homogeneity.
Conflicts of interest
The authors declare no competing financial interest.
Acknowledgements
This work was supported by the Biomedical Research Victoria
for the Undergraduate Research Opportunities Program (UROP).
The Australian Research Council (ARC) is acknowledged for
providing the Discovery Early Career Researcher Award (DECRA:
DE150101755) to P.-Y. Wang. Taipei Medical University and MOST,
Taiwan are also acknowledged for providing funding support
to P.-Y. Wang (TMU105-AE1-B13 and 105-2314-B-038-088-MY2).
R. Wong was supported by the Medical Advances Without Animal
Fellowship and the Louisa Jean de Bretteville Bequest. We espe-
cially thank Prof. Wei-Bor Tsai for providing the nanogrooved
patterns. This work was performed at both the Biointerface
Engineering Hub at Swinburne University and MCN as part of
the Victorian Node of the Australian National Fabrication Facility,
a company established under the National Collaborative Research
Infrastructure Strategy to provide nano- and microfabrication
facilities to Australia’s researchers. The Centre for Eye Research
Australia receives operational infrastructure support from the
Victorian Government.
References
1 P.-Y. Wang, H. Thissen and P. Kingshott, Modulation
of Human Multipotent and Pluripotent Stem Cells Using
Surface Nanotopographies and Surface-Immobilised Bioactive
Signals: A Review, Acta Biomater., 2016, 45, 31–59.
2 P. O. Favaron, R. C. Carvalho, J. Borghesi, A. R. Anunciacao
and M. A. Miglino, The Amniotic Membrane: Development
and Potential Applications – a Review, Reprod. Domest. Anim.,
2015, 50, 881–892.
Paper Journal of Materials Chemistry B
This journal is ©The Royal Society of Chemistry 201 7 J. Mater. Chem. B, 2017, 5, 7927--7938 | 7937
3 B. Lo and L. Parham, Ethical Issues in Stem Cell Research,
Endocr. Rev., 2009, 30, 204–213.
4 S. Ding, P. Kingshott, H. Thissen, M. Pera and P.-Y. Wang,
Modulation of Human Mesenchymal and Pluripotent Stem
Cell Behavior Using Biophysical and Biochemical Cues:
A Review, Biotechnol. Bioeng., 2017, 114, 260–280.
5 L. G. Villa-Diaz, A. M. Ross, J. Lahann and P. H. Krebsbach,
Concise Review: The Evolution of Human Pluripotent Stem
Cell Culture: From Feeder Cells to Synthetic Coatings, Stem
Cells, 2013, 31, 1–7.
6 P.-Y. Wang, S. S.-C. Hung, H. Thissen, P. Kingshott and
R. C.-B. Wong, Binary Colloidal Crystals (BCCs) as a Feeder-
Free System to Generate Human Induced Pluripotent Stem
Cells (hiPSCs), Sci. Rep., 2016, 6, 36845.
7 P.-Y. Wang, L. R. Clements, H. Thissen, A. Jane, W.-B. Tsai
and N. H. Voelcker, Screening Mesenchymal Stem Cell
Attachment and Differentiation on Porous Silicon Gradients,
Adv. Funct. Mater., 2012, 22, 3414–3423.
8 P.-Y. Wang, J. Yu, J.-H. Lin and W.-B. Tsai, Modulation of
Alignment, Elongation and Contraction of Cardiomyocytes
through a Combination of Nanotopography and Rigidity of
Substrates, Acta Biomater., 2011, 7, 3285–3293.
9 M.J.Dalby,D.McCloy,M.Robertson,H.Agheli,D.Sutherland,
S. Affrossman and R. O. Oreffo, Osteoprogenitor Response to
Semi-Ordered and Random Nanotopographies, Biomaterials,
2006, 27, 2980–2987.
10 M. J. Dalby, N. Gadegaard, R. Tare, A. Andar, M. O. Riehle,
P. Herzyk, C. D. W. Wilkinson and R. O. C. Oreffo, The
Control of Human Mesenchymal Cell Differentiation Using
Nanoscale Symmetry and Disorder, Nat. Mater., 2007, 6,
997–1003.
11 E. K. F. Yim, S. W. Pang and K. W. Leong, Synthetic
Nanostructures Inducing Differentiation of Human
Mesenchymal Stem Cells into Neuronal Lineage, Exp. Cell
Res., 2007, 313, 1820–1829.
12 C. Morez, M. Noseda, M. A. Paiva, E. Belian, M. D. Schneider
and M. M. Stevens, Enhanced Efficiency of Genetic Program-
ming toward Cardiomyocyte Creation through Topographical
Cues, Biomaterials, 2015, 70, 94–104.
13 S. R. Gu, Y. G. Kang, J. W. Shin and J. W. Shin, Simultaneous
Engagement of Mechanical Stretching and Surface
Pattern Promotes Cardiomyogenic Differentiation of Human
Mesenchymal Stem Cells, J. Biosci. Bioeng., 2017, 123, 252–258.
14 A. Islam, M. Younesi, T. Mbimba and O. Akkus, Collagen
Substrate Stiffness Anisotropy Affects Cellular Elongation,
Nuclear Shape, and Stem Cell Fate toward Anisotropic
Tissue Lineage, Adv. Healthcare Mater., 2016, 5, 2237–2247.
15 F. Marongiu, R. Gramignoli, Q. Sun, V. Tahan, T. Miki,
K. Dorko, E. Ellis and S. C. Strom, Isolation of Amniotic
Mesenchymal Stem Cells, Curr. Protoc. Stem Cell Biol., 2010,
ch. 1, Unit 1E 5.
16 S. M. Wu, Y. Fujiwara, S. M. Cibulsky, D. E. Clapham,
C. L. Lien, T. M. Schultheiss and S. H. Orkin, Developmental
Origin of a Bipotential Myocardial and Smooth Muscle Cell
Precursor in the Mammalian Heart, Cell, 2006, 127,
1137–1150.
17 P.-Y. Wang, W.-T. Li, J. Yu and W.-B. Tsai, Modulation of
Osteogenic, Adipogenic and Myogenic Differentiation of
Mesenchymal Stem Cells by Submicron Grooved Topography,
J. Mater. Sci.: Mater. Med., 2012, 23, 3015–3028.
18 P.-Y. Wang, D. T. Bennetsen, M. Foss, T. Ameringer,
H. Thissen and P. Kingshott, Modulation of Human
Mesenchymal Stem Cell Behavior on Ordered Tantalum
Nanotopographies Fabricated Using Colloidal Lithography
and Glancing Angle Deposition, ACS Appl. Mater. Interfaces,
2015, 7, 4979–4989.
19 P.-Y. Wang, H. Thissen and W.-B. Tsai, The Roles of RGD
and Grooved Topography in the Adhesion, Morphology, and
Differentiation of C2C12 Skeletal Myoblasts, Biotechnol.
Bioeng., 2012, 109, 2104–2115.
20 P.-Y. Wang, T.-H. Wu, P.-H. Chao, W.-H. Kuo, M.-J. Wang,
C.-C. Hsu and W.-B. Tsai, Modulation of Cell Attachment
and Collagen Production of Anterior Cruciate Ligament
Cells Via Submicron Grooves/Ridges Structures with Different
Cell Affinity, Biotechnol. Bioeng., 2013, 110, 327–337.
21 P.-Y. Wang, H. Thissen and P. Kingshott, Stimulation of
Early Osteochondral Differentiation of Human Mesenchymal
Stem Cells Using Binary Colloidal Crystals (BCCs), ACS Appl.
Mater. Interfaces, 2016, 8, 4477–4488.
22 F. Pan, M. Zhang, G. Wu, Y. Lai, B. Greber, H. R. Scholer and
L. Chi, Topographic Effect on Human Induced Pluripotent
Stem Cells Differentiation Towards Neuronal Lineage,
Biomaterials, 2013, 34, 8131–8139.
23 Y. N. Wu, J. B. Law, A. Y. He, H. Y. Low, J. H. Hui, C. T. Lim,
Z. Yang and E. H. Lee, Substrate Topography Determines
the Fate of Chondrogenesis from Human Mesenchymal
Stem Cells Resulting in Specific Cartilage Phenotype Formation,
Nanomedicine, 2014, 10, 1507–1516.
24 P.-Y. Wang, D. T. Bennetsen, M. Foss, H. Thissen and
P. Kingshott, Response of Mg63 Osteoblast-Like Cells to
Ordered Nanotopographies Fabricated Using Colloidal Self-
Assembly and Glancing Angle Deposition, Biointerphases,
2015, 10, 04A306.
25 I. J. Kramer, D. Zhitomirsky, J. D. Bass, P. M. Rice,
T. Topuria, L. Krupp, S. M. Thon, A. H. Ip, R. Debnath,
H. C. Kim and E. H. Sargent, Ordered Nanopillar Struc-
tured Electrodes for Depleted Bulk Heterojunction
Colloidal Quantum Dot Solar Cells, Adv. Mater., 2012, 24,
2315–2319.
26 P.-Y. Wang, T.-H. Wu, W.-B. Tsai, W.-H. Kuo and M.-J.
Wang, Grooved PLGA Films Incorporated with RGD/YIGSR
Peptides for Potential Application on Skeletal Muscle Tissue
Engineering, Colloids Surf., B, 2013, 110C, 88–95.
27 P.-Y. Wang, H.-T. Yu and W.-B. Tsai, Modulation of Alignment
and Differentiation of Skeletal Myoblasts by Submicron
Ridges/Grooves Surface Structure, Biotechnol. Bioeng., 2010,
106, 285–294.
28 D. Bae, S. H. Moon, B. G. Park, S. J. Park, T. Jung, J. S. Kim,
K. B. Lee and H. M. Chung, Nanotopographical Control for
Maintaining Undifferentiated Human Embryonic Stem Cell
Colonies in Feeder Free Conditions, Biomaterials, 2014, 35,
916–928.
Journal of Materials Chemistry B Paper
7938 |J. Mater. Chem. B, 2017, 5, 7927--7938 This journal is ©The Royal Society of Chemistry 201 7
29 K. A. Kilian, B. Bugarija, B. T. Lahn and M. Mrksich, Geo-
metric Cues for Directing the Differentiation of Mesenchymal
Stem Cells, Proc.Natl.Acad.Sci.U.S.A., 2010, 107, 4872–4877.
30 J. W. Cassidy, J. N. Roberts, C. A. Smith, M. Robertson,
K. White, M. J. Biggs, R. O. Oreffo and M. J. Dalby, Osteo-
genic Lineage Restriction by Osteoprogenitors Cultured on
Nanometric Grooved Surfaces: The Role of Focal Adhesion
Maturation, Acta Biomater., 2014, 10, 651–660.
31 K. Kolind, D. Kraft, T. Boggild, M. Duch, J. Lovmand,
F. S. Pedersen, D. A. Bindslev, C. E. Bunger, M. Foss and
F. Besenbacher, Control of Proliferation and Osteogenic
Differentiation of Human Dental-Pulp-Derived Stem Cells
by Distinct Surface Structures, Acta Biomater., 2014, 10,
641–650.
32 M. E. Torres-Padilla and I. Chambers, Transcription Factor
Heterogeneity in Pluripotent Stem Cells: A Stochastic
Advantage, Development, 2014, 141, 2173–2181.
33 J. R. Heath, A. Ribas and P. S. Mischel, Single-Cell Analysis
Tools for Drug Discovery and Development, Nat. Rev. Drug
Discovery, 2016, 15, 204–216.
34 J. A. Santiago, R. Pogemiller and B. M. Ogle, Heterogeneous
Differentiation of Human Mesenchymal Stem Cells in
Response to Extended Culture in Extracellular Matrices,
Tissue Eng., Part A, 2009, 15, 3911–3922.
Paper Journal of Materials Chemistry B
... Nanostructures on acrylic film have stimulating effects on the physical and biophysical behavior of the cells that adhere to the surface [1]. Various designs of structures on acrylic films, such as nanopillars [2][3][4], nanogrooves [5][6][7], and nanopores [8][9][10] on a nanometer scale have been presented in a wide variety of biomedical devices. Nanostructures on the surfaces of some materials can effectively regulate the adhesion [11,12], migration [13], proliferation [14], and other behaviors of bacteria and viruses. ...
Article
Full-text available
The fabrication of nanostructures is of great importance in producing biomedical devices. Significantly, the nanostructure of the polymeric film has a significant impact on the physical and biophysical behavior of the biomolecules. This study presents an efficient nanofabrication method of nanogroove structures on an acrylic film by the micro-embossing process. In this method, a master mold was made from a thermos oxide silicon substrate using photolithography and etching techniques. An isotropic optical polymethyl methacrylate (PMMA) film is used in the experiment. The acrylic film is known for its excellent optical properties in products such as optical lenses, medical devices, and various general purpose engineering plastics. Then, the micro-embossing process was realized to fabricate nanogroove patterns on an acrylic film by using a micro-embossing machine. However, the morphology of the nanopatterns on an acrylic film was characterized by using an atomic force microscope to measure the dimensions of the nanogroove patterns. The impact of embossing temperature on the morphology of nanogroove patterns on acrylic film is experimentally investigated. The results show that when the embossing temperature is too small, the pattern is not fully formed, and slipping occurs in nanopatterns on the acrylic film. On the other hand, the effect of increasing the embossing temperature on the morphology of nanogrooves agrees with the master mold, and the crests between the nanogrooves form straight edges. It should be noted that the micro-embossing temperature also strongly influences the transferability of nanopatterns on an acrylic film. The technique has great potential for rapidly fabricating nanostructure patterns on acrylic film.
... The nanopillar OP65 and HP50 surfaces showed differentiation markers for multiple stem cell fates without biochemicals. The distribution of F-actin filaments and vinculin was seen in confocal microscopy by hAM-MSCs and mESCs on OP65 nanopillar topography for focal adhesion, as shown in Figure 3b and c. 54 Furthermore, the nanopillars of titania with an average diameter of 20−30 nm were effective for promoting the osteogenic differentiation of stem cells. 55−57 The Si nanopillars of larger diameter (100 nm) were more effective in determining the osteogenic fate of hMSCs than those of smaller diameter (50 nm). ...
Article
Full-text available
Engineered nanomaterials (ENMs) with different topographies provide effective nano−bio interfaces for controlling the differentiation of stem cells. The interaction of stem cells with nanoscale topographies and chemical cues in their microenvironment at the nano−bio interface can guide their fate. The use of nanotopographical cues, in particular nanorods, nanopillars, nanogrooves, nanofibers, and nanopits, as well as biochemical forces mediated factors, including growth factors, cytokines, and extracellular matrix proteins, can significantly impact stem cell differentiation. These factors were seen as very effective in determining the proliferation and spreading of stem cells. The specific outgrowth of stem cells can be decided with size variation of topographic nanomaterial along with variation in matrix stiffness and surface structure like a special arrangement. The precision chemistry enabled controlled design, synthesis, and chemical composition of ENMs can regulate stem cell behaviors. The parameters of size such as aspect ratio, diameter, and pore size of nanotopographic structures are the main factors for specific termination of stem cells. Protein corona nanoparticles (NPs) have shown a powerful facet in stem cell therapy, where combining specific proteins could facilitate a certain stem cell differentiation and cellular proliferation. Nano−bio reactions implicate the interaction between biological entities and nanoparticles, which can be used to tailor the stem cells' culmination. The ion release can also be a parameter to enhance cellular proliferation and to commit the early differentiation of stem cells. Further research is needed to fully understand the mechanisms underlying the interactions between engineered nano−bio interfaces and stem cells and to develop optimized regenerative medicine and tissue engineering designs.
... Chemical surface modifications, signaling, and growth factors may be needed to supplement the nanotopography to bring about similar results. 68 The surface stiffness has also been observed to provide the requisite mechanical stimulus to encourage the differentiation of stem cells to form a specific cell type. It was possible to culture polymeric cell types by utilizing the differential mechanical properties of different polymers and polymer composites to fabricate the support platform. ...
Article
Full-text available
Membrane technology is playing a crucial role in cutting-edge innovations in the biomedical field. One such innovation is the surface engineering of a membrane for enhanced longevity, efficient separation, and better throughput. Hence, surface engineering is widely used while developing membranes for its use in bioartificial organ development, separation processes, extracorporeal devices, etc. Chemical-based surface modifications are usually performed by functional group/biomolecule grafting, surface moiety modification, and altercation of hydrophilic and hydrophobic properties. Further, creation of micro/nanogrooves, pillars, channel networks, and other topologies is achieved to modify physio-mechanical processes. These surface modifications facilitate improved cellular attachment, directional migration, and communication among the neighboring cells and enhanced diffusional transport of nutrients, gases, and waste across the membrane. These modifications, apart from improving functional efficiency, also help in overcoming fouling issues, biofilm formation, and infection incidences. Multiple strategies are adopted, like lysozyme enzymatic action, topographical modifications, nanomaterial coating, and antibiotic/antibacterial agent doping in the membrane to counter the challenges of biofilm formation, fouling challenges, and microbial invasion. Therefore, in the current review, we have comprehensibly discussed different types of membranes, their fabrication and surface modifications, antifouling/antibacterial strategies, and their applications in bioengineering. Thus, this review would benefit bioengineers and membrane scientists who aim to improve membranes for applications in tissue engineering, bioseparation, extra corporeal membrane devices, wound healing, and others.
... For instance, human MSCs cultured on 350-nm gratings showed decreased expression of integrin subunits α2, αV,β2, β3 and β4 and exhibited an aligned actin cytoskeleton, compared to unpatterned controls [142]. Human amniotic membrane-derived mesenchymal stem cells (hAM-MSCs) and mouse embryonic stem cells (mESCs) showed strong alignment on deep grooves and inhibition of spreading on nanopillars [143]. Topographical cues modulated focal adhesion kinase (FAK) activity and thereby the RhoA phosphorylation level. ...
Article
Full-text available
Background Whilst traditional strategies to increase transfection efficiency of non-viral systems aimed at modifying the vector or the polyplexes/lipoplexes, biomaterial-mediated gene delivery has recently sparked increased interest. This review aims at discussing biomaterial properties and unravelling underlying mechanisms of action, for biomaterial-mediated gene delivery. DNA internalisation and cytoplasmic transport are initially discussed. DNA immobilisation, encapsulation and surface-mediated gene delivery (SMD), the role of extracellular matrix (ECM) and topographical cues, biomaterial stiffness and mechanical stimulation are finally outlined. Main text Endocytic pathways and mechanisms to escape the lysosomal network are highly variable. They depend on cell and DNA complex types but can be diverted using appropriate biomaterials. 3D scaffolds are generally fabricated via DNA immobilisation or encapsulation. Degradation rate and interaction with the vector affect temporal patterns of DNA release and transgene expression. In SMD, DNA is instead coated on 2D surfaces. SMD allows the incorporation of topographical cues, which, by inducing cytoskeletal re-arrangements, modulate DNA endocytosis. Incorporation of ECM mimetics allows cell type-specific transfection, whereas in spite of discordances in terms of optimal loading regimens, it is recognised that mechanical loading facilitates gene transfection. Finally, stiffer 2D substrates enhance DNA internalisation, whereas in 3D scaffolds, the role of stiffness is still dubious. Conclusion Although it is recognised that biomaterials allow the creation of tailored non-viral gene delivery systems, there still are many outstanding questions. A better characterisation of endocytic pathways would allow the diversion of cell adhesion processes and cytoskeletal dynamics, in order to increase cellular transfection. Further research on optimal biomaterial mechanical properties, cell ligand density and loading regimens is limited by the fact that such parameters influence a plethora of other different processes (e.g. cellular adhesion, spreading, migration, infiltration, and proliferation, DNA diffusion and release) which may in turn modulate gene delivery. Only a better understanding of these processes may allow the creation of novel robust engineered systems, potentially opening up a whole new area of biomaterial-guided gene delivery for non-viral systems.
Article
Full-text available
Despite the advent of tissue engineering (TE) for the remodeling, restoring, and replacing damaged cardiovascular tissues, the progress is hindered by the optimal mechanical and chemical properties required to induce cardiac tissue-specific cellular behaviors including migration, adhesion, proliferation, and differentiation. Cardiac extracellular matrix (ECM) consists of numerous structural and functional molecules and tissue-specific cells, therefore it plays an important role in stimulating cell proliferation and differentiation, guiding cell migration, and activating regulatory signaling pathways. With the improvement and modification of cell removal methods, decellularized ECM (dECM) preserves biochemical complexity, and bio-inductive properties of the native matrix and improves the process of generating functional tissue. In this review, we first provide an overview of the latest advancements in the utilization of dECM in in vitro model systems for disease and tissue modeling, as well as drug screening. Then, we explore the role of dECM-based biomaterials in cardiovascular regenerative medicine (RM), including both invasive and non-invasive methods. In the next step, we elucidate the engineering and material considerations in the preparation of dECM-based biomaterials, namely various decellularization techniques, dECM sources, modulation, characterizations, and fabrication approaches. Finally, we discuss the limitations and future directions in fabrication of dECM-based biomaterials for cardiovascular modeling, RM, and clinical translation.
Article
In the human body, almost all cells interact with extracellular matrices (ECMs), which have tissue and organ-specific compositions and architectures. These ECMs not only function as cellular scaffolds, providing structural support, but also play a crucial role in dynamically regulating various cellular functions. This comprehensive review delves into the examination of biofabrication strategies used to develop bioactive materials that accurately mimic one or more biophysical and biochemical properties of ECMs. We discuss the potential integration of these ECM-mimics into a range of physiological and pathological in vitro models, enhancing our understanding of cellular behavior and tissue organization. Lastly, we propose future research directions for ECM-mimics in the context of tissue engineering and organ-on-a-chip applications, offering potential advancements in therapeutic approaches and improved patient outcomes.
Article
Mesenchymal stem cells (MSCs) can maintain immune homeostasis and many preclinical trials with MSCs have been carried out around the world. In vitro culture of MSCs has been found to result in the decline of immunomodulatory capacity, migration and proliferation. To address these problems, simulating the extracellular environment for preconditioning of MSCs is a promising and inexpensive method. Biophysical cues in the external environment that MSCs are exposed to have been shown to affect MSC migration, residency, differentiation, secretion, etc. We review the main ways in which MSCs exert their immunomodulatory ability, and summarize recent advances in mechanical preconditioning of MSCs to enhance immunomodulatory capacity and related mechanical signal sensing and transduction mechanisms.
Article
Full-text available
3D cardiac engineered constructs have yielded not only the next generation of cardiac regenerative medicine but also have allowed for more accurate modeling of both healthy and diseased cardiac tissues. This is critical as current cardiac treatments are rudimentary and often default to eventual heart transplants. This review serves to highlight the various cell types found in cardiac tissues and how they correspond with current advanced fabrication methods for creating cardiac engineered constructs capable of shedding light on various pathologies and providing the therapeutic potential for damaged myocardium. In addition, insight is given towards the future direction of the field with an emphasis on the creation of specialized and personalized constructs that model the region-specific microtopography and function of native cardiac tissues.
Article
Full-text available
Over the last few years, organic bioelectronics has experienced an exponential growth with applications encompassing platforms for tissue engineering, drug delivery systems, implantable, and wearable sensors. Although reducing the physical and mechanical mismatch with the human tissues allows to increase the coupling efficiency, several challenges are still open in terms of matching biological curvature, size, and interface stiffness. In this context, the replacement of bulky with more flexible and conformable devices is required, implying the transition from inorganic conventional electronics to organic electronics. Indeed, the advent of organic materials in bioelectronics, due to the indisputable benefits related to biocompatibility, flexibility, and electrical properties, has granted superior coupling properties with human tissues increasing the performances of both sensing and stimulation platforms. In this review the ease of functionalization and patterning of conductive polymers (CPs) will be analyzed as a strategy that enables the fabrication of platforms with high structural flexibility ranging from the macro to the micro/nano‐scales, leading to the increase of devices sensitivity. Drawing from the concept of biomimicry, the human body tissues interfaces will be explored through an ideal journey starting from organic platforms for epidermal sensing and stimulation. Then, devices capable of establishing a dynamic coupling with the heart will be reviewed and finally, following the circulatory system and crossing the blood‐brain barrier, the brain will be reached and novel sensing and computing implants advances that pave the way to the possibility to emulate as well as to interact with the neural functions will be analyzed.
Article
Full-text available
Human induced pluripotent stem cells (hiPSCs) are capable of differentiating into any cell type and provide significant advances to cell therapy and regenerative medicine. However, the current protocol for hiPSC generation is relatively inefficient and often results in many partially reprogrammed colonies, which increases the cost and reduces the applicability of hiPSCs. Biophysical stimulation, in particular from tuning cell-surface interactions, can trigger specific cellular responses that could in turn promote the reprogramming process. In this study, human fibroblasts were reprogrammed into hiPSCs using a feeder-free system and episomal vectors using novel substrates based on binary colloidal crystals (BCCs). BCCs are made from two different spherical particle materials (Si and PMMA) ranging in size from nanometers to micrometers that self-assemble into hexagonal close-packed arrays. Our results show that the BCCs, particularly those made from a crystal of 2 μm Si and 0.11 μm PMMA particles (2SiPM) facilitate the reprogramming process and increase the proportion of fully reprogrammed hiPSC colonies, even without a vitronectin coating. Subsequent isolation of clonal hiPSC lines demonstrates that they express pluripotent markers (OCT4 and TRA-1-60). This proof-of-concept study demonstrates that cell reprogramming can be improved on substrates where surface properties are tailored to the application.
Article
Full-text available
Statement of significance: This review focuses on the effect of surface nanotopographies and surface-bound biosignals on human stem cells. Recently, stem cell research attracts much attention especially the induced pluripotent stem cells (iPSCs) and direct lineage reprogramming. The fast advance of stem cell research benefits disease treatment and cell therapy. On the other hand, surface property of cell adhered materials has been demonstrated very important for in vitro cell culture and regenerative medicine. Modulation of cell behavior using surfaces is costeffective and more defined. Thus, we summarize the recent progress of modulation of human stem cells using surface science. We believe that this review will capture a broad audience interested in topographical and chemical patterning aimed at understanding complex cellular responses to biomaterials.
Article
Full-text available
In vitro manipulation of human stem cells is a critical process in regenerative medicine and cellular therapies. Strategies and methods to maintain stem cells and direct them into specific lineages are ongoing challenges in these fields. To date, a number of studies have reported that besides biochemical stimulation, biophysical cues in the form of surface patterning and external stimulation also influence stem cell attachment, proliferation and differentiation, and can be used in cell reprogramming and the maintenance of pluripotency. While biochemical cues generally are effective and easy to deliver, biophysical cues have many other advantages for scalability as they are cost-efficient, have a longer lifetime, and can be easily well-defined. However, different protocols and cell sources utilized in a variety of studies have led to difficulties in obtaining clear conclusions about the effects of the biophysical environment on stem cells. In addition, the examination of different types of external stimulation is time consuming and limited by available fabrication techniques, resulting in a delay in commercialisation and clinical applications. In this review, we aim to summarise the most important biophysical cues and methods for the culture of human stem cells, including mesenchymal and pluripotent stem cells, to facilitate their adoption in stem cell biology. The standard classical protocols of using biochemical cues will also be discussed for comparison. We believe that combining biochemical and biophysical stimulation has the greatest potential to generate functionally mature cells at a scalable and inexpensive rate for diverse applications in regenerative medicine and cell therapy. This article is protected by copyright. All rights reserved
Article
Full-text available
Rigidity of substrates plays an important role in stem cell fate. Studies are commonly carried out on isotropically stiff substrate or substrates with unidirectional stiffness gradients. However, many native tissues are anisotropically stiff and it is unknown whether controlled presentation of stiff and compliant material axes on the same substrate governs cytoskeletal and nuclear morphology, as well as stem cell differentiation. In this study, electrocompacted collagen sheets are stretched to varying degrees to tune the stiffness anisotropy (SA) in the range of 1 to 8, resulting in stiff and compliant material axes orthogonal to each other. The cytoskeletal aspect ratio increased with increasing SA by about fourfold. Such elongation was absent on cellulose acetate replicas of aligned collagen surfaces indicating that the elongation was not driven by surface topography. Mesenchymal stem cells (MSCs) seeded on varying anisotropy sheets displayed a dose-dependent upregulation of tendon-related markers such as Mohawk and Scleraxis. After 21 d of culture, highly anisotropic sheets induced greater levels of production of type-I, type-III collagen, and thrombospondin-4. Therefore, SA has direct effects on MSC differentiation. These findings may also have ramifications of stem cell fate on other anisotropically stiff tissues, such as skeletal/cardiac muscles, ligaments, and bone.
Article
Full-text available
A new surface based on self-assembly of two colloids into well-defined nanostructures so called binary colloidal crystals (BCCs) for stem cell culture was fabricated. The facile fabrication process described is able to cover large surface areas (> 3 cm-diameter, i.e. > 7 cm2) with ordered surface nanotopographies that is often a challenge particularly in biomaterials science. From our library, four different combinations of BCCs were selected using mixtures of silica, polystyrene and poly(methyl methacrylate) particles with sizes in the range from 100 nm to 5 µm. Cell spreading, proliferation, and surface-induced lineage commitment of human adipose-derived stem cells (hADSCs) was studied using quantitative polymerase chain reaction (qPCR) and immunostaining. The results showed that BCCs induced osteo- and chondro- but not adipo-gene expression in the absence of induction medium suggesting that the osteochondral lineage can be stimulated by the BCCs. When applying induction media, higher osteo- and chondro-gene expression on BCCs was found compared with tissue culture polystyrene (TCPS) and flat silica (Si) controls, respectively. Colony forming of chondrogenic hADSCs was found on BCCs and TCPS but not Si controls suggesting that the differentiation of stem cells is surface-dependent. BCCs provide access to complex nanotopographies and chemistries, which can find applications in cell culture and regenerative medicine.
Article
Full-text available
Ordered surface nanostructures have attracted much attention in different fields including biomedical engineering because of their potential to study the size effect on cellular response and modulation of cell fate. However, the ability to fabricate large-area ordered nanostructures is typically limited due to high costs and low speed of fabrication. Herein, highly ordered nanostructures with large surface areas (>1.5 × 1.5 cm2) were fabricated using a combination of facile techniques including colloidal self-assembly, colloidal lithography, and glancing angle deposition (GLAD). An ordered tantalum (Ta) pattern with 60-nm-height was generated using colloidal lithography. A monolayer of colloidal crystal, i.e., hexagonal close packed 720 nm polystyrene particles, was self-assembled and used as a mask. Ta patterns were subsequently generated by evaporation of Ta through the mask. The feature size was further increased by 100 or 200 nm using GLAD, resulting in the fabrication of four different surfaces (FLAT, Ta60, GLAD100, and GLAD200). Cell adhesion, proliferation, and mineralization of MG63 osteoblast-like cells were investigated on these ordered nanostructures over a 1 week period. Our results showed that cell adhesion, spreading, focal adhesion formation, and filopodia formation of the MG63 osteoblast-like cells were inhibited on the GLAD surfaces, especially the initial (24 h) attachment, resulting in a lower cell density on the GLAD surfaces. After 1 week culture, alkaline phosphatase activity and the amount of Ca was higher on the GLAD surfaces compared with Ta60 and FLAT controls, suggesting that the GLAD surfaces facilitate differentiation of osteoblasts. This study demonstrates that ordered Ta nanotopographies synthesized by combining colloidal lithography with GLAD can improve the mineralization of osteoblast-like cells providing a new platform for biomaterials and bone tissue engineering.
Article
Full-text available
Contents Foetal membranes are essential tissues for embryonic development, playing important roles related to protection, breathing, nutrition and excretion. The amnion is the innermost extraembryonic membrane, which surrounds the foetus, forming an amniotic sac that contains the amniotic fluid ( AF ). In recent years, the amniotic membrane has emerged as a potential tool for clinical applications and has been primarily used in medicine in order to stimulate the healing of skin and corneal diseases. It has also been used in vaginal reconstructive surgery, repair of abdominal hernia, prevention of surgical adhesions and pericardium closure. More recently, it has been used in regenerative medicine because the amniotic‐derived stem cells as well as AF ‐derived cells exhibit cellular plasticity, angiogenic, cytoprotective, immunosuppressive properties, antitumoural potential and the ability to generate induced pluripotent stem cells. These features make them a promising source of stem cells for cell therapy and tissue engineering. In this review, we discussed the development of the amnion, AF and amniotic cavity in different species, as well as the applicability of stem cells from the amnion and AF in cellular therapy.
Article
Full-text available
Generation of de novo cardiomyocytes through viral over-expression of key transcription factors represents a highly promising strategy for cardiac muscle tissue regeneration. Although the feasibility of cell reprogramming has been proven possible both in vitro and in vivo, the efficiency of the process remains extremely low. Here, we report a chemical-free technique in which topographical cues, more specifically parallel microgrooves, enhance the directed differentiation of cardiac progenitors into cardiomyocyte-like cells. Using a lentivirus-mediated direct reprogramming strategy for expression of Myocardin, Tbx5, and Mef2c, we showed that the microgrooved substrate provokes an increase in histone H3 acetylation (AcH3), known to be a permissive environment for reprogramming by "stemness" factors, as well as stimulation of myocardin sumoylation, a post-translational modification essential to the transcriptional function of this key co-activator. These biochemical effects mimicked those of a pharmacological histone deacetylase inhibitor, valproic acid (VPA), and like VPA markedly augmented the expression of cardiomyocyte-specific proteins by the genetically engineered cells. No instructive effect was seen in cells unresponsive to VPA. In addition, the anisotropy resulting from parallel microgrooves induced cellular alignment, mimicking the native ventricular myocardium and augmenting sarcomere organization. Copyright © 2015 The Authors. Published by Elsevier Ltd.. All rights reserved.
Article
It has been widely recognized and proved that biophysical factors for mimicking in vivo conditions should be also considered to have stem cells differentiated into desired cell type in vitro along with biochemical factors. Biophysical factors include substrate and biomechanical conditions. This study focused on the effect of biomimetic mechanical stretching along with changes in substrate topography to influence on cardiomyogenic differentiation of human mesenchymal stem cells (hMSCs). Elastic micropatterned substrates were made to mimic the geometric conditions surrounding cells in vivo. To mimic biomechanical conditions due to beating of the heart, mechanical stretching was applied parallel to the direction of the pattern (10% elongation, 0.5 Hz, 4 h/day). Suberoylanilide hydroxamic acid (SAHA) was used as a biochemical factor. The micropatterned substrate was found more effective in the alignment of cytoskeleton and cardiomyogenic differentiation compared with flat substrate. Significantly higher expression levels of related markers [GATA binding protein 4 (GATA4), troponin I, troponin T, natriuretic peptide A (NPPA)] were observed when mechanical stretching was engaged on micropatterned substrate. In addition, 4 days of mechanical stretching was associated with higher levels of expression than 2 days of stretching. These results indicate that simultaneous engagement of biomimetic environment such as substrate pattern and mechanical stimuli effectively promotes the cardiomyogenic differentiation of hMSCs in vitro. The suggested method which tried to mimic in vivo microenvironment would provide systematic investigation to control cardiomyogenic differentiation of hMSCs.
Article
The genetic, functional or compositional heterogeneity of healthy and diseased tissues presents major challenges in drug discovery and development. Such heterogeneity hinders the design of accurate disease models and can confound the interpretation of biomarker levels and of patient responses to specific therapies. The complex nature of virtually all tissues has motivated the development of tools for single-cell genomic, transcriptomic and multiplex proteomic analyses. Here, we review these tools and assess their advantages and limitations. Emerging applications of single cell analysis tools in drug discovery and development, particularly in the field of oncology, are discussed.