You are on page 1of 9

Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


journal homepage: www.elsevier.com/locate/colsurfb

Surface functionalization of styrenic block copolymer elastomeric


biomaterials with hyaluronic acid via a grafting to strategy
Xiaomeng Li a,b , Shifang Luan a, , Shuaishuai Yuan a,b , Lingjie Song a,b , Jie Zhao a,b ,
Jiao Ma a,b , Hengchong Shi a , Huawei Yang a , Jing Jin a , Jinghua Yin a,
a

State Key Laboratory of Polymer Physics and Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, PR
China
b
University of Chinese Academy of Sciences, Beijing 100049, PR China

a r t i c l e

i n f o

Article history:
Received 13 May 2013
Received in revised form 2 July 2013
Accepted 26 July 2013
Available online 3 August 2013
Keywords:
Styrenic block copolymer (SBC)
Hyaluronic acid (HA)
Surface functionalization
Hemocompatibility
Cytocompatibility

a b s t r a c t
As a biostable elastomer, the hydrophobicity of styrenic block copolymer (SBC) intensely limits its
biomedical applications. In order to overcome such shortcoming, the SBC lms were grafted with
hyaluronic acid (HA) using a coupling agent. The surface chemistry of the modied lms was examined by ATR-FTIR and XPS techniques, and the surface morphology was visually described by AFM. The
biological performances of the HA-modied lms were evaluated by a series of experiments, such as
protein adsorption, platelet adhesion, and in vitro cytocompatibility. It was found that the HA-modied
samples showed a low adhesiveness to broblast at the initial stage; however, it stimulated the growth
of broblast. The L929 broblast growth presented a strong dependence on the molecular weight (MW)
of HA. The samples modied with 17 kDa HA exhibited the worst wettability and platelet adhesion, while
providing the best results of supporting broblast proliferation.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Styrenic block copolymer (SBC) (Fig. S1(a)) typically consists of
two polystyrene hard segments on each end and a central polyolene soft segment. The high glass transition temperature (Tg )
styrenic regions as physical crosslinking points provide strength for
this elastomer, while the low Tg olenic regions contribute exibility and elasticity. The physical crosslinking points are reversible
upon heating, therefore imparting the multiple melt processing
properties to this bioelastomer, which makes it superior to a vulcanized rubber. As a result of its physiological inertness, good
thermal and oxidative stability, as well as low toxicity, the biostable
SBC elastomer have a wide range of biomedical applications [14],
such as disposable medical devices, microchips [5], urinary tract [6],
glaucoma shunt [7], articial heart valve [8] and stent drug delivery
coating [9,10].
For some specic biomedical applications, SBC bioelastomer
should be further subjected to surface modication. A desired
surface can be tailored by immobilizing various biocompatible substances [11,12], such as poly(ethylene glycol) (PEG) and
its derivatives [1315], poly(vinylpyrrolidone) (PNVP) [1619],
heparin [2022], zwitterionic materials [2325], peptides [26,27],

Corresponding authors. Tel.: +86 431 85262109; fax: +86 431 85262109.
E-mail addresses: suan@ciac.jl.cn, luanshifang@163.com (S. Luan),
yinjh@ciac.jl.cn (J. Yin).
0927-7765/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfb.2013.07.048

proteins [28], chitosan [29,30] and hyaluronic acid (HA) [3135].


As reviewed by Burdick and Prestwich [36], HA (Fig. S1(b))
is an attractive starting material in the construction of hydrogels for various tissue engineering applications. A huge number
of HA-based hydrogels with desired morphology, stiffness and
bioactivity have been tailored by two main strategies: addition
or condensation reactions [37], and free radical polymerization
reactions [38]. Jia et al. [39,40] have pioneered techniques for
the production of HA-based hydrogel particles (HGPs) of micron
to submicron dimensions via inverse emulsion polymerization
processes. The presence of nanoscale pores within the HGPs
makes them ideal candidates as release depots for protein-based,
biomacromolecular drugs. Up to now, just a few articles have
reported that HA could be used as biocompatible surface modier for a polymeric substrate, but they have aroused great interest
[3235]. For instance, Lees group [35] reported a novel musselinspired method for the homogeneous and robust coating of
HA onto various substrates via Michael addition or Schiff base
reactions between HA and the surface-adherent polydopamine
layer.
In this work, HA was used to modify a SBC lm using a coupling
agent. The surface densities of bioactive HA were quantied, and
the surface morphology was visually described by AFM. The biological interactions of the substrates modied with different molecular
weight HAs were evaluated by a series of experiments, such as protein adsorption, platelet adhesion, and in vitro response of L929
broblast.

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

2. Experimental

147

AFM images. The reported data were the mean values of triplicate
specimens.

2.1. Materials and reagents


2.5. Surface amine density of the SBC substrate
Poly(styrene-b-(ethylene-co-butylene)-b-styrene) copolymer
(Kraton G 1652) was purchased from Shell Chemicals. Diethylenetriamine (DETA) was provided by Shanghai Aladdin Chemicals.
HA (Molecular weight (MW) = 17, 47 and 310 kDa, MW was
tested by light scattering) was purchased from Liuzhou Chemicals. 1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC)
hydrochloride and N-hydroxysuccinimide (NHS) were provided
by Alfa Aesar. Phosphate-buffered saline (PBS, pH 7.4), bovine
serum albumin (BSA), bovine serum brinogen (BFg), MES buffer
(2-(N-morpholino) ethanesulfonic acid, pH 5.4) and sodium
dodecyl sulfate (SDS) were provided by Dingguo Biotechnology.
Micro BCATM protein assay reagent kit (AR1110) was provided by
Boster Biotechnology. Ninhydrin was provided by Shanghai Huishi
Biochemical Reagent. Human hyaluronic acid binding protein
(HABP) ELISA Kit was supplied by Shanghai kanu biotechnology.
Dulbeccos modied Eagles medium (DMEM) and 0.25 wt.% trypsin
were purchased from Beijing Solarbio Science & Technology. Sterile
ltered fetal bovine serum (FBS) was supplied by Beijing Yuanhengjinma Biotechnology. Fluorescein Isothiocyanate Labeled
Phalloidin (FITC-Phalloidin) and 4 ,6-diamidino-2-phenylindole
(DAPI) dihydrochloride were provided by Sigma. Other reagents
were AR grade and used without further purication.
2.2. Surface functionalization of SBC elastomer
The HA-functionalized substrate was prepared as follows
(Fig. 1). The SBC lms (1.5 cm 1.5 cm for each lm) were washed
with deionized water and ethanol, and dried for use. Then they
were subjected to an approximate 10 Pa oxygen atmosphere in a
DT-03 plasma apparatus (Suzhou Omega Technology, China) with
a plasma power of 30 W for 30 s. The pre-treated SBC lms were
soaked in an acetonitrile solution containing DETA (0.04, 0.08, 0.12,
0.16 and 0.20 mol/L) at 75 C for 16 h. The lms were respectively
washed with deionized water and ethanol, and nally dried under
vacuum for 24 h to obtain the DETA-modied lm (denoted as SBCDETA). 1 mL solution of HA in 0.05 mol/L MES buffer (0.5, 1.0, 5.0 and
1.0 mg/mL) was prepared, and followed by adding 1 mL 0.1 mol/L
EDC and 1 mL 0.05 mol/L NHS solution for each lm. The DETAmodied lms were soaked in the as-prepared HA solution at room
temperature for 16 h for the HA immobilization, then rinsed with
deionized water, and dried in a vacuum oven. The samples modied by HA with a MW of 17 kDa, 47 kDa and 310 kDa were denoted
as HA17, HA47, and HA310.
2.3. Surface chemistry characterization of the SBC lms
ATR-FTIR spectra were obtained from a Fourier transform
infrared spectrometer (FTIR, BRUKER Vertex 70). A total of 32
scans were accumulated with a resolution of 4 cm1 for each spectrum. Surface elemental compositions were examined by an X-ray
photoelectron spectroscopy (XPS, VG Scientic ESCA MK II Thermo
Avantage V 3.20 analyzer) with Al/K (h = 1486.6 eV) anode monoX-ray source at the detection angle of 90 . Scans between 1200 and
0 eV were performed.
2.4. Surface morphology of the SBC lms
Surface morphology was examined in an atomic force
microscopy (AFM) with contact mode (SPA300HV with a
SPI 3800 controller, Seiko Instruments Industry, Japan). The
root-mean-square (RMS) roughness was evaluated directly from

Surface amine density was tested via a ninhydrin method.


Firstly, the lms (1.5 cm 1.5 cm) were placed in 3 mL test tubes
with 0.5 mL 0.1 mol/L sodium citrate (pH 5) and 1 mL ninhydrin
reagent, and heated to 100 C for 15 min. Finally, the solution
was cooled to room temperature and the colorimetric response
was measured at 490 nm by a microplate reader (TECAN SUNRISE, Swiss). A standard curve was prepared from a series of DETA
standard solutions. The reported data were the mean values of
triplicate specimens.
2.6. Surface HA density of the SBC substrate
Surface HA density was measured by human hyaluronic acid
binding protein (HABP) ELISA kit. Briey, the HA-grafted lms
reacted with 50 L hyaluronic acid binding protein conjugated with
horseradish peroxidase (HRP) at 37 C for 30 min to create HA complexes on the substrate. After completely rinsing with PBS solution,
50 L HRP-conjugate reagent was added, and incubated at 37 C for
30 min, and followed by adding 50 L chromogen solution A and
50 L chromogen solution B, nally the reaction was terminated
by adding 50 L stop solution. The absorbance values of the solutions were tested by a microplate reader (TECAN SUNRISE, Swiss)
at a wavelength of 450 nm. A standard curve was calculated from a
series of HA standard solutions. Each result is an average of at least
ve parallel experiments.
2.7. Wettability
Water contact angle (WCA) of the SBC lms was measured with
a drop shape analysis instrument (DSA, KRSS GMBH, Germany) at
room temperature. Each result is an average of at least ve parallel
experiments.
2.8. Protein adsorption
After soaking in PBS solution at room temperature for 12 h, the
SBC lms were dipped into PBS solution containing BSA (0.1, 0.5
and 1.0 mg/mL) or BFg (0.1, 0.5 and 1.0 mg/mL) at 37 C for 2 h.
Each lm was sequentially rinsed ve times with fresh PBS, soaked
in an aqueous solution of 1.0 wt.% SDS, and oscillated at 37 C for
1.0 h to remove the adsorbed proteins from the substrate. Based on
bicinchoninic acid (BCA) protein assay kit, the absorbance values
of the SDS solution containing proteins was measured at 570 nm
with a microplate reader (TECAN SUNRISE, Swiss), and amount of
the adsorbed proteins was calculated. Each result is an average of
at least three parallel experiments.
2.9. Platelet adhesion
The SBC lms were soaked in PBS solution at room temperature for 2 h. Fresh platelet-rich plasma (RRP) was obtained from a
healthy rabbit by centrifugation at 800 rpm for 10 min. Then 20 L
RRP was dropped on the lms and incubated at 37 C for 1 h. After
washing three times with PBS solution, the platelets adhered on the
lms were xed by 2.5 wt.% glutaraldehyde at 4 C for at least 10 h.
Finally, the lms were washed three times with PBS, and sequentially dehydrated with a series of ethanol aqueous solution (10, 30,
50, 70, 90, 100 vol%). The samples were observed with eld emitted
scanning electron microscopy (SEM, XL 30 ESEM FEG, FEI Company,
USA). The number of platelet adhesion on the SBC substrates was

148

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

Fig. 1. The preparation procedure of the HA-functionalized surface.

analyzed using a one-way ANOVA method (*p < 0.05, **p < 0.01).
Each result is an average of at least three parallel experiments.
2.10. Cytocompatibility
Cytocompatibility was studied using murine broblast cell
line L929. Cells were grown in DMEM supplemented with
10 vol% FBS, 4.5 g/L glucose, 100 units/mL penicillin, 5958 mL/L N2-Hydroxyethylpiperazine-N -2-ethanesulfonic acid (HEPES) and
100 g/mL streptomycin and maintained in a humidied 5%
CO2 /95% air incubator at 37 C.
The SBC lms with a diameter of 16 mm were pretreated with
75 vol% ethanol aqueous solution for 1 h, and followed by extensive
rinsing in PBS. Those lms were then soaked in cell culture medium
and kept in cell culture incubator. After reaching 95% conuency,
cells were detached by 0.25 wt.% trypsin and then suspended in culture medium at a density of 1.0 105 cells/mL. Then 1 mL medium
and 0.5 mL cell suspension were added to each sample. Cell viability and cell numbers were measured after cultured for 1, 2 and 4
days. The cell culture medium was removed and followed by extensive rinsing in PBS, and the cell adhered on the lms were xed
by 4.0 wt.% paraformaldehyde at 4 C for 30 min to obtained the
cell-immobilized samples. For observing the uorescence images of
cells in different stages, the cell-immobilized samples were stained
via adding 0.2 mL uorescence-staining solution at room temperature for 30 min. Finally, they were washed three times with PBS
and once with deionized water, and vacuum freeze dehydration.
Fluorescence images of FITC and DAPI-stained cells were obtained
by an inverted TE-2000-U digital uorescence microscope (Nikon)
attached with a digital camera (DXM1200F). For observing the SEM
images of cells, the cell-immobilized samples were washed three
times with PBS and once with deionized water, and vacuum freeze
dehydration. Images from scanning electron microscopy (SEM, XL
30 ESEM FEG, FEI Company, USA) were used to calculate cell spreading area and cell density by using Image J software. Cell density and
cell surface area on the SBC substrates were analyzed using a oneway ANOVA method (*p < 0.05, **p < 0.01, ***p < 0.001). Each result
is an average of at least three parallel experiments.
3. Results and discussion
The most widely used hydrophilic surface modiers are based
on PEG and its derivatives [4145]. We have found that the introduction of PEG could impart good hemocompatibity to a SBC

elastomeric substrate [46,47]; however, these modiers simultaneously suppressed cell adhesion and proliferation. The covalent
attachment of acrylated HA to this substrate via a photografting polymerization could obtain both hemocompatiblity and
cytocompatiblity [48]. In view of the typical characteristics of
photo-initiated free radical polymerization and the inhomogeneous distribution of photoinitiator physically adsorbed on the
surface, a large polydispersity index and inhomogeneous distribution of the grafted chains were inevitable. The morphology
of the modied substrate was certainly heterogeneous on the
micro- and macro-scales. This rendered the dependence of biological response on the MW of the immobilized HA ambiguous.
In this study, HA with different MW were respectively immobilized onto a SBC substrate under mild condition via a grafting
to strategy (Fig. 1). Both the biocompatibility and the MW
effect of HA were systematically investigated for the HA-modied
systems.
3.1. Surface chemistry
3.1.1. ATR-FTIR
As ATR-FTIR spectra shown in Fig. 2a, two new characteristic peaks at 3387 cm1 (N H stretching) and 1265 cm1 (C N
stretching) appeared after surface modication of the SBC substrate with diethylenetriamine. As for the HA-modied substrates,
the ATR-FTIR spectra further changed. Specically, characteristic
peaks at 3387 cm1 (N H stretching) and 1265 cm1 (C N stretching) strengthened, and new peaks at 1665 cm1 (C O stretching),
1099 cm1 (C OH stretching) and 812 cm1 (C O C stretching)
appeared.
3.1.2. XPS
Surface composition of the SBC lms was detected by XPS
(Fig. 2b). After sequential surface modication with DETA and HA,
the intensity of N1s peak nearby 399 eV immensely strengthened,
and the ratios of N/C increased from 0.85% up to 2.85% (Table
S1). Due to oxygen contamination, the high-resolution C1s spectra of the virgin SBC lm were divided into two peaks using a
Gaussian peak tting algorithm (Fig. S2(a)). After covalent immobilization with DETA, a new peak at 285.6 eV attributing to C N
group appeared (Fig. S2(b)). As for the HA-modied lm, the highresolution C1s spectra were decomposed into ve peaks: a C H
(C C) peak at 284.6 eV, a C N peak at 285.6 eV, a C O C peak at
286.2 eV, a N C O peak at 288.2 eV and a O C O peaks at 289.0 eV,

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

149

Fig. 2. ATR-FTIR (a), and XPS (b) spectra of the SBC lms. (I) Virgin SBC, (II) SBC-DETA, (III) HA17, (IV) HA47, and (V) HA310.

respectively (Fig. S2(c)(e)). The composition ratios of N C O were


larger than that of O C O, suggesting that activated by EDC/NHS
partner, carboxyl groups in HA chains reacted with amine groups
on the substrate (Table S2).

3.1.3. Surface amine density of the SBC lms


As quantied using a ninhydrin assay, the surface amine density gradually rose with the increment of DETA concentration, and
reached a maximum of 60 ng/cm2 at a DETA concentration of
0.20 mol/L (Fig. 3a). After the HA immobilization (10 mg/mL HA
(17 kDa) in the loading solution), the exposed amines on the substrates decreased, conrming that DETA participated in the HA
grafting reactions. It should be noted that the secondary amines,
which already existed in HA chains and newly created by the linking reaction between HA and the grafted DETA, also reacted with
the ninhydrin reagent, so amines were still detected for the HAmodied lms.

3.1.4. Surface HA density of the SBC substrate


Surface HA density was determined by a quantitative assay
based on the binding of labeled HA-binding protein (HABP) to the
grafted HA [49]. As shown in Fig. 3b, the amount of the grafted
HA generally increased as both HA concentration and MW rose.
A maximum density of the grafted HA reached 53 ng/cm2 at a
concentration of 10.0 mg/mL HA with a 310 kDa MW. Using the
solutions of 10.0 mg/mL HA with a 17 kDa MW, 5.0 mg/mL HA with
a 47 kDa MW, and 1.0 mg/mL HA with a 310 kDa MW, a similar HA
density of 40 ng/cm2 could be tailored and used in the subsequent
experiments.
For the HA binding and recognition, HABP required HA having a
continuous segments of 12002000 Da, that is, a 35 disaccharide
repeating units, so the grafted amount of HA was underestimated
by this assay kit [33]. This underestimation was supported by the
ATR-FTIR and XPS results. Similarly, the biological functions of HA
to other proteins also required a minimum HA repeating sequence,
and the biological interaction between HA and cells was expressed

150

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

Fig. 3. Surface density of amine groups (a) (n = 3) and HA (b) (n = 5) on the SBC lms.

Fig. 4. Amount of protein adsorption on the SBC surfaces. (a) BSA, and (b) BFg. (The
error bars: standard deviations, n = 3.)

3.3. Biocompatibility of the SBC lms


through cell surface proteins [50]. This HA measurement appeared
inappropriate, but accurately presented the amount of bioactive HA
which had biological response to protein and cell.

3.2. Surface morphology


Chemical and physical signals from a substrate, such as surface energy, topography, electrostatic charge and wettability, play
a key role in stimulating cell adhesion and inuencing cell growth
behavior. As indicated by AFM three-dimensional images, there
were several big holes and a few large granules on the SBC controls (Fig. S3(a)), and the RMS value was 27 nm (Fig. S3(e)). Once
HA was grafted to the substrate, the RMS value could increase up
to 65 nm. The constrained mobility stemming from the solid substrate strongly prevented the short chains of the low MW HA from
spreading. The high MW HA chains had a larger steric hindrance,
but the long HA chains had good mobility, and their grafting yields
was much larger, so fully spread out on the substrates as a homogeneous and continuous HA layer. Thus, the RMS value had an
inverse correlation to the MW of the grafted HA, i.e., 65 nm for
HA17, 59 nm for HA47 and 39 nm for HA310, respectively (Fig.
S3(b)(e)).

Protein adsorption is the rst event in bloodmaterial interactions and some proteins in blood, which plays an important
role in material-associated clotting. It is widely recognized that a
hydrophilic surface possesses good anti-protein adsorption properties, a moderate hydrophilicity is generally optimal for a polymer.
The high amounts of protein adsorption were observed on the
hydrophobic SBC samples (a WCA of 103 ) (Table S3). Correspondingly, a large number of disk-shaped (a typical stage of early
activation), and spreading (a totally activated status) platelets were
observed on the virgin SBC surface (Fig. 5a). The serious thrombosis would be initiated by platelets adhesion and activation and
breakage of platelets. In contrast, as for the HA-modied samples (a WCA of 6269 ) (Table S3), HA hydration layer prevented
the hydrophobic interaction between protein and substrate, so
the amount of BSA and BFg adsorbed on the HA-modied substrates was smaller than that of the hydrophobic SBC controls
(Fig. 4). Depending on protein concentration, although the amounts
of the proteins adsorbed on the HA-modied substrates had a
slight variation, the adsorbed amount of BSA was generally reduced
by 2050%, while that of BFg was even up to 91% relative to
the SBC controls. The amount of protein adsorption was mainly
dependent on the concentration and species of protein, rather than
the MW of HA (Fig. 4). BFg could bind to the platelet GP IIb/IIIa

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

151

Fig. 5. Adhesion of platelets on the SBC lms. (a) virgin SBC, (b) SBC-DETA, (c) HA17, (d) HA47, (e) HA310, and (f) the statistical number of adherent platelets. (The error bars:
standard deviations, n = 3; data analyzed using a one-way ANOVA, *p < 0.05, **p < 0.01.)

receptor, and formed bridges between platelets, resulting in


platelet activation and aggregation, while BSA was generally nonadhesive for platelet. In addition, the ratio of brinogen to albumin
adsorbed on the substrate decreased, suggesting that the HAimmobilized surfaces has good hemocompatibility [51].In the case
of platelet adhesion, the hydrophilic HA chains quickly hydrated
when contacting with platelet-rich plasma, which weakened the
interaction force between platelets and the substrate because
of the steric hindrance effect, bringing about the less adhered
platelets (Fig. 5ce). The values of the adhered platelets were
1.5 104 cells/mm2 (HA17 and HA47) and 0.9 104 cells/mm2
(HA310) relative to 2.5 104 cells/mm2 of SBC controls (Fig. 5f).
The difference between the SBC controls and the HA-modied
groups was statistically signicant, i.e., p < 0.05 for HA17 and HA47,
p < 0.01 for HA310 (Fig. 5f). In addition, HA17 and HA47 groups
had statistically signicant differences in comparison with HA310
group. As for the SBC controls, the ratio of activated platelets was

71% (Table S4), while that of the HA-modied surfaces was respectively reduced to 54%, 52% and 38%, conrming that the hydrophilic
HA graft chains suppressed platelet spread and activation.
Usually, the interaction force between the vinculin on the Cell
Membrane and the substrate was also weakened because of the
hydrated layer effect, bringing about the less adhered cells. For
example, anti-fouling PEG and its derivatives inevitably inhibited
cell adhesion and proliferation. As shown in Figs. 6, 8 and S4, the
HA-modied lms suppressed L929 cell adhesion (Fig. 8a), which
was consistent with the previous reports that the low adhesiveness
of the HA surfaces to cells [52]. However, the HA-modied lms
signicantly promoted cell proliferation, especially for HA17, as
compared with the SBC controls. As for the HA17 lms, the number
of L929 cell cultured on their surfaces increased in time dependent
manner, their increasing rates are much larger than that of HA47
and HA310 groups, presenting the best results of supporting broblast proliferation. With respect to the 1020 m size of cells, the

152

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

Fig. 6. Fluorescent photomicrographs showing L929 cell numbers at different time points on various surfaces. (I) virgin SBC, (II) HA17, (III) HA47, and (IV) HA310, after
incubation for 1 day (A), 2 days (B) and 4 days (C). (Size of the scale bars: 100 m.)

Fig. 7. SEM images of L929 cell at different time points on various surfaces: (I) virgin SBC, (II) HA17, (III) HA47, and (IV) HA310, after incubation for 1 day (A), 2 days (B) and
4 days (C). (Size of the scale bars: 200 m.)

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

153

de-adhesion and recovery of these cells. Thus, the cell-spreading


area of the HA-modied samples was actually much larger than
that analyzed from SEM images. The testing results of cell adhesion, proliferation, morphology and spreading area suggested that
HA imparted good cytocompatibility to the SBC substrates.
As we know, a single molecule chain of HA could bind to multiple
cell surface receptors, the HA binding domains to different cell surface proteins were different, and new cell surface recognizing HA
proteins was continuously discovered, so the biological response
of HA to cell was still incompletely elucidated even for the HA in
aqueous solution. The constrained mobility of the grafted HA made
the interaction between HA and cell more complex.
4. Conclusions
The SBC elastomeric biomaterials modied with hyaluronic acid
(HA) signicantly inhibited protein adsorption and platelet adhesion relative to the SBC controls. The HA-modied samples had
a low adhesiveness to broblast at the initial stage; however, it
stimulated the growth of broblast. The L929 broblast growth
showed a strong dependence on the MW of HA. The samples
modied with the 17 kDa HA presented the worst wettability and
platelet adhesion, while producing the best results of supporting
broblast proliferation. Although the complex interaction between
the immobilized HA and cell remained to be incompletely elucidated, the HA modication imparted excellent hemocompatiblity
and cytocompatibility to the SBC biomaterials.
Acknowledgements
The authors acknowledge the nancial support of National
Science Foundation of China (Project number: 21274150 and
51273200), Chinese Academy of Sciences-Wego Group High-tech
Research & Development Program (20112013), and Scientic Development Program of Jilin Province (Project number:
20130102064JC).
Fig. 8. Cell density (a) and cell surface area (b) on the SBC substrates for culturing
1, 2 and 4 days. (The error bars: standard deviations, n = 3; data analyzed using a
one-way ANOVA, *p < 0.05, **p < 0.01, ***p < 0.001.)

HA layer on the SBC surface could be thought as homogeneous and


continuous. The HA17 and HA47 samples had the similar roughness, while the broblast growth of HA17 samples was statistically
signicant better than that of HA47 samples. Therefore, it could be
concluded that the L929 broblast growth had a strong dependence
on the MW of the HA immobilized on the SBC substrate, similar to
that of the HA in other forms [53].
Fine topographies of L929 cells were obtained by SEM (Fig. 7),
and the cell spreading area as well as morphology were summarized in Fig. 8b and Table S5, respectively. After 1 day culturing, the
SBC controls were mostly covered with round-shaped cells, while
the HA-modied substrates were mainly adhered by the spreading
cells, i.e., spindle-shaped, triangle-shaped and stellate-shaped cells
(Fig. 7A, Table S5). The cell-spreading area of HA17 was larger than
those of the SBC controls, HA47 and HA310 groups, and it had statistically signicant differences as compared with the SBC controls
and HA310 group (Fig. 8b). As culture time increased, cells rapidly
proliferated and fully spread out on the HA-modied substrates
(Fig. 7). Comparing with the SBC controls, the HA-modied lms
had more spreading cells (Table S5), as well as larger cell-spreading
area (Fig. 8b), especially for HA17. It should be noticed that uorescent photographs (Fig. 6) and SEM images (Fig. 7) presented
obvious differences as for cell morphology, particularly for the HAmodied samples at initial stage of cell culturing. It was probably
due to that the cells were not tightly bound to the HA-modied
substrates, so the treatment for SEM observation resulted in the

Appendix A. Supplementary data


Supplementary data associated with this article can be found,
in the online version, at http://dx.doi.org/10.1016/j.colsurfb.2013.
07.048.
References
[1] L. Pinchuk, G.J. Wilson, J.J. Barry, R.T. Schoephoerster, J.M. Parel, J.P. Kennedy,
Biomaterials 29 (2008) 448.
[2] J.E. Puskas, Y. Chen, Biomacromolecules 5 (2004) 1141.
[3] J.E. Puskas, Y. Chen, Y. Dahman, D. Padavan, J. Polym. Sci. Polym. Chem. 42
(2004) 3091.
[4] C. Gotz, G.T. Lim, J.E. Puskas, V. Altstadt, J. Mech. Behav. Biomed. Mater. 10
(2012) 206.
[5] E. Roy, J.C. Galas, T. Veres, Lab. Chip 11 (2011) 3193.
[6] P. Cadieux, J.D. Watterson, J. Denstedt, R.R. Harbottle, J. Puskas, J. Howard, B.S.
Gan, G. Reid, Colloids Surf. B Biointerfaces 28 (2003) 95.
[7] M. Orozco, A.C. Acosta, E.M. Espana, L. Pinchuk, B. Weber, S. Davis, E. Arrieta, S.
Dubovy, F. Fantes, M. Aly, Y.H. Zhou, J.M. Parel, Ophthalmic Technol. XVI 6138
(2006) U1380.
[8] T.E. Claiborne, G. Girdhar, S. Gallocher-Lowe, J. Sheriff, Y.P. Kato, L. Pinchuk, R.T.
Schoephoerster, J. Jesty, D. Bluestein, ASAIO J. 57 (2011) 26.
[9] J.Z. Zhu, X.W. Xiong, R. Du, Y.J. Jing, Y. Ying, X.M. Fan, T.Q. Zhu, R.Y. Zhang,
Biomaterials 33 (2012) 8204.
[10] S.V. Ranade, R.E. Richard, M.N. Helmus, Acta Biomater. 1 (2005) 137.
[11] F.J. Xu, K.G. Neoh, E.T. Kang, Prog. Polym. Sci. 34 (2009) 719.
[12] Q.A. Yu, Y.X. Zhang, H.W. Wang, J.L. Brash, H. Chen, Acta Biomater. 7 (2011)
1550.
[13] S. Zanini, C. Riccardi, M. Orlandi, C. Colombo, F. Croccolo, Polym. Degrad. Stab.
93 (2008) 1158.
[14] K.M. Xiu, Q. Cai, J.S. Li, X.P. Yang, W.T. Yang, F.J. Xu, Colloids Surf. B Biointerfaces
90 (2012) 177.
[15] S. Zanini, M. Orlandi, C. Colombo, E. Grimoldi, C. Riccardi, Eur. Phys. J. D 54
(2009) 159.

154

X. Li et al. / Colloids and Surfaces B: Biointerfaces 112 (2013) 146154

[16] U. Edlund, M. Kllrot, A.-C. Albertsson, J. Am. Chem. Soc. 127 (2005) 8865.
[17] S.F. Luan, J. Zhao, H.W. Yang, H.C. Shi, J. Jin, X.M. Li, J.C. Liu, J.W. Wang, J.H. Yin,
P. Stagnaro, Colloids Surf. B Biointerfaces 93 (2012) 127.
[18] Z.M. Liu, Z.K. Xu, L.S. Wan, J. Wu, M. Ulbricht, J. Membr. Sci. 249 (2005) 21.
[19] H.Y. Yu, Z.K. Xu, Y.J. Xie, Z.M. Liu, S.Y. Wang, J. Membr. Sci. 279 (2006) 148.
[20] U. Edlund, S. Danmark, A.C. Albertsson, Biomacromolecules 9 (2008) 901.
[21] F.J. Xu, Y.L. Li, E.T. Kang, K.G. Neoh, Biomacromolecules 6 (2005) 1759.
[22] Y.-C. Kuo, Y.-T. Tsai, Colloids Surf. B Biointerfaces 82 (2011) 616.
[23] J. Yuan, L. Chen, X.F. Jiang, J. Shen, S.C. Lin, Colloids Surf. B Biointerfaces 39
(2004) 87.
[24] J. Zhao, Q. Shi, L.G. Yin, S.F. Luan, H.C. Shi, L.J. Song, J.H. Yin, P. Stagnaro, Appl.
Surf. Sci. 256 (2010) 7071.
[25] Y.-F. Yang, Y. Li, Q.-L. Li, L.-S. Wan, Z.-K. Xu, J. Membr. Sci. 362 (2010) 255.
[26] D. Li, H. Chen, W.G. McClung, J.L. Brash, Acta Biomater. 5 (2009) 1864.
[27] D. Li, H. Chen, S.S. Wang, Z.Q. Wu, J.L. Brash, Acta Biomater. 7 (2011) 954.
[28] L.P. Zhu, J.H. Jiang, B.K. Zhu, Y.Y. Xu, Colloids Surf. B Biointerfaces 86 (2011)
111.
[29] O. Wiarachai, N. Thongchul, S. Kiatkamjornwong, V.P. Hoven, Colloids Surf. B
Biointerfaces 92 (2012) 121.
[30] Y.F. Wang, Q.F. Hong, Y.J. Chen, X.X. Lian, Y.F. Xiong, Colloids Surf. B Biointerfaces 100 (2012) 77.
[31] M. Kallrot, U. Edlund, A.C. Albertsson, Macromol. Biosci. 8 (2008) 645.
[32] J.G. Alauzun, S. Young, R. DSouza, L. Liu, M.A. Brook, H.D. Sheardown, Biomaterials 31 (2010) 3471.
[33] M. Morra, Biomacromolecules 6 (2005) 1205.
[34] T.W. Chuang, K.S. Masters, Biomaterials 30 (2009) 5341.
[35] H. Lee, S.M. Dellatore, W.M. Miller, P.B. Messersmith, Science 318 (2007) 426.
[36] J.A. Burdick, G.D. Prestwich, Adv. Mater. 23 (2011) H41.

[37] J.X. Zhang, A. Skardal, G.D. Prestwich, Biomaterials 29 (2008) 4521.


[38] S.K. Seidlits, Z.Z. Khaing, R.R. Petersen, J.D. Nickels, J.E. Vanscoy, J.B. Shear, C.E.
Schmidt, Biomaterials 31 (2010) 3930.
[39] A.K. Jha, W.D. Yang, C.B. Kirn-Safran, M.C. Farach-Carson, X.Q. Jia, Biomaterials
30 (2009) 6964.
[40] X. Xu, A.K. Jha, R.L. Duncan, X.Q. Jia, Acta Biomater. 7 (2011) 3050.
[41] S. Zanini, C. Riccardi, E. Grimoldi, C. Colombo, A.M. Villa, A. Natalello, P. GattiLafranconi, M. Lotti, S.M. Doglia, J. Colloid Interface Sci. 341 (2010) 53.
[42] Z.Q. Wu, H. Chen, X.L. Liu, J.L. Brash, Macromol. Biosci. 12 (2012) 126.
[43] S. Alibeik, S. Zhu, J.L. Brash, Colloids Surf. B Biointerfaces 81 (2010) 389.
[44] H. Chen, L. Wang, Y.X. Zhang, D. Li, W.G. McClung, M.A. Brook, H. Sheardown,
J.L. Brash, Macromol. Biosci. 8 (2008) 863.
[45] H. Chen, Y.X. Zhang, D. Li, X.Y. Hu, L. Wang, W.G. McClung, J.L. Brash, J. Biomed.
Mater. Res. A 90A (2009) 940.
[46] X.M. Li, S.F. Luan, H.W. Yang, H.C. Shi, J. Zhao, J. Jin, J.H. Yin, P. Stagnaro, Appl.
Surf. Sci. 258 (2012) 2344.
[47] H.W. Yang, S.F. Luan, J. Zhao, H.C. Shi, X.M. Li, L.J. Song, J. Jin, Q. Shi, J.H. Yin, D.A.
Shi, P. Stagnaro, Polymer 53 (2012) 1675.
[48] X.M. Li, S.F. Luan, H.C. Shi, H.W. Yang, L.J. Song, J. Jin, J.H. Yin, P. Stagnaro, Colloids
Surf. B Biointerfaces 102 (2013) 210.
[49] A. Suzuki, P. Angulo, J. Lymp, D. Li, S. Satomura, K. Lindor, Liver Int. 25 (2005)
779.
[50] J. Lesley, I. Gal, D.J. Mahoney, M.R. Cordell, M.S. Rugg, R. Hyman, A.J. Day, K.
Mikecz, J. Biol. Chem. 279 (2004) 25745.
[51] M. Holmberg, X.L. Hou, Colloids Surf. B Biointerfaces 84 (2011) 71.
[52] A. Khademhosseini, K.Y. Suh, J.M. Yang, G. Eng, J. Yeh, S. Levenberg, R. Langer,
Biomaterials 25 (2004) 3583.
[53] S. Ibrahim, B. Joddar, M. Craps, A. Ramamurthi, Biomaterials 28 (2007) 825.

You might also like