You are on page 1of 13

Biomaterials

Science
View Article Online
PAPER View Journal | View Issue

Evaluation of absorbable hemostatic agents of


Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

Cite this: Biomater. Sci., 2018, 6,


polyelectrolyte complexes using carboxymethyl
3332 starch and chitosan oligosaccharide both in vitro
and in vivo
Xingtao Chen, a Yonggang Yan,*a Hong Li,a Xuehong Wang,b Songchao Tang,b
Quan Li,c Jie Weib and Jiacan Su *c

Absorbable hemostatic agents with a high hemostatic efficacy play an important role in surgical and
severely traumatic hemostasis. In the present study, by applying polyelectrolyte assembly, polyelectrolyte
complexes (PECs), using carboxymethyl starch (CMS) and chitosan oligosaccharide (COS), with control-
lable physicochemical properties were prepared and optimized for hemorrhage control. Particle size, zeta
potential, morphology and water absorption of the PECs with different CMS/COS ratios were systemati-
cally evaluated. The results of in vitro degradation in PBS suggested that CMS/COS PECs were degradable
and their degradation rates, which decreased with the increase of the COS content, were suitable for
absorbable hemostatic agents. The in vivo hemostatic efficacy of the PECs with 10 wt% COS content
(PEC 10), which was evaluated in a rabbit hepatic hemorrhage model, was better than CMS but decreased
with the increase of the COS content. The plasma coagulation evaluation revealed that the PECs could
significantly activate and accelerate the coagulation cascade through both the intrinsic and extrinsic path-
ways but could not directly affect the common pathway. CMS/COS PECs also showed antimicrobial
activity against S. aureus, which enhanced with the increase of the COS content, but failed against E. coli.
Received 6th June 2018, Moreover, PEC 10 displayed excellent cytocompatibility with MC3T3-L1 and good tissue compatibility in a
Accepted 2nd August 2018
rabbit liver model. These findings not only suggest that CMS/COS PECs with a suitable COS content were
DOI: 10.1039/c8bm00628h promising absorbable hemostatic agents for internal use but they are also useful to understand the under-
rsc.li/biomaterials-science lying principles for designing PEC based hemostatic agents.

1. Introduction by oxidized cellulose (e.g., Surgicel® and Oxycel®) limits its


application in sensitive tissues such as nervous and cardiac
Absorbable hemostatic agents have been widely developed for systems;4 chitosan (e.g., HemCon® and Gelox®) suffers from
surgical and severely traumatic bleeding, because in some its animal origin which may increase the risk of disease or
cases control of bleeding by pressure, ligature, and other con- viral infection;5,6 starch microspheres (e.g., Arista®) are insuffi-
ventional procedures is ineffective or impractical.1 The cient to stop severe bleeding and have no antimicrobial
approved absorbable hemostatic agents, including collagen, activity.7–9 Therefore, it is necessary to find a more effective
oxidized cellulose, chitosan and starch, have been investigated agent with better hemostatic efficacy as well as biocompatibil-
and are commercially available.2,3 However, some issues with ity, nonantigenicity, biodegradability and antimicrobial
regard to biological safety, potential infections, hemostatic activity.
efficacy and biodegradation rate limit their wide use, and none Carboxymethyl starch (CMS) is a derivative of starch in
of these agents can be considered as the ideal absorbable which some of the hydroxyl groups are modified into carboxy-
hemostatic agents for surgery. For instance, the low pH caused methyl groups. Owing to carboxymethylation, CMS is an
anionic polysaccharide and can be dissolved in aqueous solu-
tion at room temperature. It has been used as a biomaterial in
a
College of Physical Science and Technology, Sichuan University, Chengdu 610064, wound healing, regenerative medicine and drug delivery due
China. E-mail: yan_yonggang@vip.163.com to its good biocompatibility, solubility and biodegradabil-
b
Key Laboratory for Ultrafine Materials of Ministry of Education, East China
ity.10,11 Compared to starch, CMS is more biodegradable and
University of Science and Technology, 200237 Shanghai, China
c
Department of Orthopaedics Trauma, Changhai Hospital, Second Military Medical hydrophilic,12 which is favorable for application as an absorb-
University, Shanghai 200433, China. E-mail: jiacansu@126.com able hemostatic agent. In addition, previous studies suggested

3332 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper

that a negatively charged carboxyl group may directly activate degradability, hemostatic efficacy, plasma coagulation, anti-
the coagulation cascade through the intrinsic pathway leading microbial activity and cytotoxicity of CMS/COS PECs were sys-
to the accelerated formation of thrombin and fibrin clot, thus tematically evaluated and the potential for internal use as
being beneficial for hemostasis.13,14 Moreover, carboxyl in hemostatic agents was studied in a rabbit hepatic hemorrhage
CMS can be modified or incorporated with an active agent by model.
chemical bonding or electrostatic interaction to improve its
hemostatic efficacy (e.g., Ca2+)15 or endow with multifunction-
ality (e.g., Cu2+ and Ag+ for antimicrobial activity).16,17 2. Materials and methods
Chitosan is a natural cationic polysaccharide derived by
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

2.1 Synthesis of COS


partial deacetylation of chitin. The hemostatic activity and
safety of chitosan have been confirmed by many studies and COS were prepared by oxidative degradation of chitosan using
its FDA approved products are available on the market for hydrogen peroxide according to previous studies with some
external use, but its application as an absorbable surgical modifications.32 Low molecular weight chitosan (18 g)
hemostatic agent is limited due to poor solubility at physio- (Aladdin Reagent, Shanghai) was suspended in 300 mL of a
logical pH, insufficient antihemorrhagic effect, potential risk hydrogen peroxide solution (6 wt%) under stirring for 1 h at
inducing inflammatory response as well as intestinal and per- 60 °C. After centrifugation, the supernatant containing COS
itoneal adhesion and low degradation rate in vivo.5,6,18 One fre- was collected and COS were precipitated by adding excess
quently used strategy to overcome these drawbacks is the ethanol to the supernatant. The COS precipitates were col-
chemical modification of the primary amines of chitosan. lected by centrifugation, washed twice with ethanol, and then
Water-soluble chitosan derivatives (e.g., chitosan-catechol and dried under vacuum at 40 °C. The viscosity of the COS was
carboxymethyl chitosan) with enhanced hemostatic perform- measured in a solution of NaCl (0.20 M) and CH3COOH (0.10
ance and biodegradability have been developed and used for M) at 30 °C using an Ubbelohde viscometer and the molecular
surgical hemostasis.19–23 An alternative strategy is the weight was determined according to the classic Mark–
reduction of the molecular weight of chitosan. As the degraded Houwink equation.33
products of chitosan, chitosan oligosaccharides (COS) have
been produced and their biological activities including anti- 2.2 Preparation and characterization of CMS/COS PECs
inflammatory and antimicrobial activities have been investi- Aqueous stock solutions (2% w/v) of CMS (Mw = 1.5 × 105,
gated in some research studies.24,25 In comparison to chitosan, cassava source, degree of substitution (DS) = 0.4, Aladdin
COS have improved biodegradability and solubility in aqueous Reagent, Shanghai) and COS were prepared by separately dis-
solutions due to a smaller molecular weight and showed solving CMS and COS in distilled water with magnetic stirring.
hemostatic and antimicrobial activities if the molecular weight The PEC solutions were prepared by mixing the two solutions
is controlled appropriately,26,27 thus may be being more suit- with a predetermined mass ratio (CMS to COS of 9 : 1, 4 : 1 and
able as an absorbable hemostatic agent. 7 : 3, abbreviated as PEC 10, PEC 20 and PEC 30, respectively)
Polyelectrolyte complexes (PECs) may be formed from CMS using an emulsifying machine (JRJ300-S; Shanghai Specimen
and COS and held together by ion pairing interactions Model Factory, Shanghai) at 4000 rpm for 40 min. The
between negatively charged carboxyl groups on CMS and posi- acquired PEC gels were collected after centrifugation at 4000
tively charged amino groups on COS. The formation of these rpm for 10 min and then washed three times with distilled
polyvalent interactions, which is driven by the entropic release water before lyophilization.
of counter ions and water molecules, was termed polyelectro- Freeze-dried PEC granules were characterized by Fourier
lyte assembly.28,29 It was reasonable to hypothesize that, by transform infrared spectrometry (FTIR) (Nicolet 6700; Thermo
applying polyelectrolyte assembly, PECs using negatively Fisher Scientific, Waltham, MA, USA). In order to determine
charged CMS and positively charged COS could combine the the particle size and zeta potential of the PECs in phosphate
two polysaccharides that individually have hemostatic effects buffer saline (PBS, pH = 7.4), the as-prepared PECs without
and are more suitable as novel absorbable hemostatic agents drying as well as CMS were adjusted at 7.4 by PBS and ana-
with improved hemostatic efficacy, biocompatibility, tunable lyzed with a laser particle size analyzer (JL-1177; Chengdu
physicochemical properties and antimicrobial activity. Jingxin Powder Analyzer Instruments Co., Chengdu, China) for
Furthermore, some research indicated that forming PECs particle size distribution and laser Doppler electrophoresis
could improve tissue compatibility and cytotoxicity of the poly- (Zetasizer Nano, Malvern Instruments, USA) for zeta potential.
electrolyte via affecting its surface charge and charge The span value (SP) was used to evaluate size distribution and
density.30,31 In the present study, COS was synthesized by oxi- calculated according to the formula,
dative degradation using hydrogen peroxide and PECs with
SP ¼ ðD90  D10 Þ=D50
various ratios of CMS to COS prepared by simply adding COS
solution to CMS solution with an emulsifying machine fol- where D90, D10, and D50 are volume size diameters at 90, 10,
lowed by lyophilization. Since the physicochemical properties and 50% of the cumulative volume, respectively. A smaller SP
of PECs strongly depend on parameters such as the molar indicated narrower size distribution. The morphology of PEC
ratio, the influences of the CMS/COS ratio on water absorption, granules and lyophilized gel prepared by mixing PEC granules

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3333
View Article Online

Paper Biomaterials Science

and PBS solutions until reaching the equilibrium swelling ground into powders and tested by the same method for
state was observed through scanning electron microscopy comparison.
(SEM) (JSM-6360LV, JEOL, Japan) after coating with gold in a
vacuum. 2.5 Plasma coagulation
To explore the effect of CMS/COS PECs on plasmatic coagu-
2.3 Water absorption and in vitro degradation lation, various coagulation assays were performed according to
The water absorption capacity of the PECs was evaluated using previous studies using a semi-automatic coagulation analyzer
simulated body fluid (SBF) solution with ion concentrations (Biomerieux, France) including the plasma recalcification time
approximately equal to those of human blood plasma. Briefly, (PRT), activated partial thromboplastin time (APTT), prothrom-
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

PEC granules were vacuum dried at 50 °C overnight to elimin- bin time (PT) and thrombin time (TT).7,9,17,35 The platelet poor
ate the residual water completely and then weighed (recorded plasma (PPP) and platelet rich plasma (PRP) were prepared by
as W1) prior to immersing. Dried samples were dipped into a the centrifugation of citrated rabbit blood at 3000 rpm for
centrifuge tube with 30 mL SBF for 10 min. Afterwards, the 10 min and at 1000 rpm for 5 min, respectively. For the PRT
samples were collected by air pump filtration until no fluid test, the sample was mixed and incubated with 100 µL of PRP
dropped into the flask and then the weight was measured for at 37 °C for 2 min. Subsequently, the mixture was recalcified
each sample (recorded as W2). The water absorption of each with 100 µL of pre-incubated CaCl2 solution (25 mM) and the
sample was calculated according to the formula, PRT was recorded as the clotting time. Prior to APTT, PT and
TT tests, the samples were mixed and incubated with PPP at
A% ¼ ½ðW 2  W 1 Þ=W 1   100%:
37 °C for 2 min. The influence of hemostatic agents on APTT
In vitro degradation behaviors were investigated by immer- was determined by mixing 100 µL of the mixture with 100 µL
sing the samples into PBS at a weight to volume ratio of of the APTT reagent, incubating for 4 min, and adding 100 µL
0.4 g : 5 mL in plastic tubes, followed by incubating at 37 °C in of CaCl2 solution (25 mM), in sequence. PT or TT under the
a shaker at 100 rpm for up to 21 days, with the solution being influence of samples was tested by mixing 100 µL of the
refreshed every three days. At predetermined time intervals mixture with 200 µL of the PT or TT reagent, respectively. All
(3, 7, 14, 21 days), the sample was collected after being centri- the reagents were preincubated at 37 °C for 2 min prior to
fuged at 4000 rpm for 5 min, gently rinsed with deionized using. The same plasma without the addition of the hemo-
water, and vacuum dried at 50 °C until a constant weight. The static agent was tested by the same method as the control.
weight of each sample before (W1) and after (W2) immersion
was accurately measured and the weight loss of each sample 2.6 Antimicrobial activity
was calculated according to the equation, The antimicrobial activity of the PECs against Escherichia coli
(E. coli, Gram-negative, ATCC 8739) and Staphylococcus aureus
D% ¼ ½ðW 1  W 2 Þ=W 1   100%:
(S. aureus, Gram-positive ATCC 6538) was investigated by a
The pH values of the supernatants were tested with a pH modified colony counting method.17,36 Sterilized samples were
meter (PS-25; Shanghai Leici Chuangyi Apparatus and incubated with the bacteria suspension, which was adjusted to
Instrument Co., Shanghai, China). the turbidity of the 0.5 McFarland standard followed by
dilution with sterile normal saline solution (0.85%, w/v) result-
2.4 Whole blood clotting ing in a concentration of about 5log CFU mL−1, at 37 °C up to
The whole blood clotting experiments were performed to 24 h in the vials with constant shaking. The suspension
measure the blood-clotting efficiency of PECs according to pre- (0.1 ml) obtained after a fixed incubating time (8, 16, 24 h) was
vious studies with some modifications.3,5,34 Fresh blood was diluted, and then seeded on Luria–Bertani agar. After incu-
obtained from New Zealand white rabbits and mixed with a bation at 37 °C for additional 24 h, the resulting bacterial colo-
one-tenth volume of 3.8% (w/v) sodium citrate. The citrated nies on the plates were counted. The growth inhibition of bac-
blood (300 µL) was dropped into a 15 mL tube containing 0.3 g teria was expressed using the difference between the number
of hemostatic agent followed by vortex mixing the powders of colonies from bacteria with samples and that in the vial as
with the blood well. Subsequently, CaCl2 solution (30 µL, 0.2 the control.
M) was added to the tube in order to initiate blood clotting,
and then incubated at 37 °C for various times (1, 2, 3, 4, 2.7 Cytotoxicity
5 min). At each time point, deionized water (10 mL) was care- Cytotoxicity was assayed using MC3T3-L1 fibroblasts by the
fully added to each tube without disturbing the clot so as to Cell Counting KIT-8 (CCK-8). The cells were cultured in a
hemolyze the red blood cells (RBCs) which were not trapped in 24-well plate for the cell viability test and a confocal dish for
the clot. The absorbance of hemoglobin in uncoagulated the observation of cell morphology, both in a 5% CO2 incuba-
blood was measured at 540 nm with a microplate reader. The tor at 37 °C. The culture medium was Dulbecco’s Modified
blood clot formed after 3 min incubation was collected, fixed Eagle’s Medium (DMEM) containing the samples and sup-
in 2.5% glutaraldehyde solution, dehydrated in a graded series plemented with 10% fetal bovine serum and 2% antibiotics
of ethanol and dried in a fume hood at room temperature for (200 mg mL−1 penicillin and 200 mg mL−1 streptomycin).
SEM analysis. Commercial gelatin sponge (GS) was gently After 1, 4 and 7 days of incubation, the medium was replaced

3334 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper

by a culture medium containing 10% CCK-8 solution and the post hoc test, where p < 0.05 was considered statistically
cells were incubated under the same conditions for 3 h. The significant.
absorbance of the culture medium was measured at 450 nm
and the cell viability was presented as the viability of cells rela-
tive to negative control (without the samples) at each time 3. Results
point. At the same time point, the confocal dishes were
3.1 Characterization of CMS/COS PECs
washed three times with PBS and the cultured cells were fixed
in 4% paraformaldehyde for 20 min for cell morphology obser- In order to identify the interactions involved in CMS/COS
vation. Prior to confocal laser scanning microscopy (CLSM) PECs, FTIR analysis was performed and the spectra are shown
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

(Nikon A1R, Japan) analysis, the fixed cells were stained with in Fig. 1. In the spectrum of CMS, two strong peaks at 1606
rhodamine phalloidin (Cytoskeleton Inc., America) for 40 min and 1427 cm−1 corresponded to the asymmetric and sym-
and 4′,6-diamidino-2-phenylindole (DAPI) (Beyotime Institute metric stretching vibration of –COO− groups and a broad band
of Biotechnology, China) for 10 min, sequentially. at 3447 cm−1 was ascribable to the stretching vibration of –OH
groups. COS exhibited a broad peak at 1629 cm−1 attributed to
2.8 Evaluation of hemostatic efficacy and tissue compatibility N–H bending vibration, a sharp peak at 1384 cm−1 assigned to
in vivo the stretching vibration of C–N, and a broad band at
3409 cm−1 corresponding to O–H and N–H stretching
The hemostatic efficacy and tissue compatibility were evalu- vibrations. Compared with the spectrum of CMS and COS, the
ated in a model of rabbit liver injury. With the approval of the peaks corresponding to the asymmetric stretching vibration of
ethical committee of the Shanghai University of Traditional –COO− groups and bending vibration of N–H of the PECs
Chinese Medicine, all animal experiments complied with the coupled and blue-shifted to 1603 cm−1 in PEC 10, 1600 cm−1
ARRIVE guidelines37 and were carried out strictly in accord- in PEC 20 and 1596 cm−1 in PEC 30, together with blue-shifts
ance with the “Guide for the Care and Use of Laboratory of symmetric stretching vibration of –COO− groups to
Animals” by the 2003 National Research Council. Thirty New 1424 cm−1 in PEC 10, 1422 cm−1 in PEC 20 and 1421 cm−1 in
Zealand white rabbits (weight: 3–4 kg) were randomly assigned PEC 30, implying that ionic interaction occurred between CMS
to five groups (CMS, PEC 10, PEC 20, PEC 30 and GS group, and COS and enhanced with the increase of the COS
n = 6 per group) and were fasted 24 h before experiments. All content.38,39 In addition, the band corresponding to O–H and
samples were vacuum dried at 50 °C overnight and then N–H stretching vibrations blue-shifted to 3419 cm−1 in PEC 10,
gamma-ray sterilized (Heming Co. Ltd, Shanghai) for 2 h 3416 cm−1 in PEC 20 and 3403 cm−1 in PEC 30, suggesting
before use. that both ionic interaction and hydrogen bonding might be
Rabbits were anesthetized and cleaned. The abdomen was involved in the PECs.
opened to expose the liver, and then an “X” shaped incision According to the viscosity of COS along with the Mark–
(approximately 15 × 15 mm with 10 mm depth) was created on Houwink equation ([μ] = 1.8 × 10−3 M0.93), the molecular
the left lateral hepatic lobe by a scalpel. Free bleeding was weight of COS was ∼5 kDa. Table 1 depicts that the mean par-
allowed for 10 s, the hemostatic sample (1 g) was applied to ticle size and zeta potential ( pH = 7.4) of CMS was 11.2 μm
the bleeding site subsequently with manual compression for and −22.7 mV, respectively, which was similar to that of PEC
10 s, and no additional hemostatic technique was used. The 10 (12.4 μm and −20.6 mV, respectively). However, when the
hemostatic time was recorded when no blood flow was
observed and hemostasis was achieved (only the hemostasis
that could maintain 10 s was recorded), and then the abdomi-
nal cavity was sutured. Postoperative rabbits of CMS, PEC 10
and PEC 30 treatment groups were used for tissue compatibil-
ity evaluation and were individually housed, allowing free
access to water and diet, and other rabbits were euthanized. At
1 and 4 weeks after experiments, rabbits for tissue compatibil-
ity evaluation were euthanized and the injured sites of liver
were photographed and sampled, fixed in 10% buffered forma-
lin, dehydrated in grading ethanol, embedded in paraffin,
then serial sectioned at 5 mm and finally stained with hema-
toxylin and eosin (H&E) according to the protocol for micro-
scopic observation.

2.9 Statistical analysis


All samples were run in triplicate and the data points were
expressed as mean ± standard deviation (SD). Statistical ana-
lysis was performed using one-way ANOVA with the Bonferroni Fig. 1 FTIR spectra of COS, CMS and CMS/COS PECs.

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3335
View Article Online

Paper Biomaterials Science

Table 1 Size distribution, zeta potential and water absorption of CMS and CMS/COS PECs

Samples Mean particle size (μm) Span value Zeta potential (mV) Water absorption (%)

CMS 11.2 ± 1.5 1.55 ± 0.09 −22.7 ± 1.1 —


PEC 10 12.4 ± 1.1 1.54 ± 0.08 −20.6 ± 1.5 660 ± 34
PEC 20 25.1 ± 2.2 1.59 ± 0.15 −11.2 ± 1.3 471 ± 24
PEC 30 50.1 ± 3.8 1.65 ± 0.11 +4.8 ± 0.3 272 ± 16
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

COS content was increased from 10% to 30%, the mean par- CMS was miscible with SBF at any proportion, while the
ticle size and the zeta potential increased significantly (from absorption ratio of the PECs dropped remarkably with the
12.4 to 50.1 μm and from −20.6 to +4.8 mV, respectively), increase in the COS content. More specifically, the absorption
suggesting that negatively charged CMS and positively charged ratio of PEC 10 (660%) was much higher than that of PEC 20
COS were aggregated into polyelectrolyte complexes with (471%) and that of PEC 30 (272%). The SEM images of freeze-
varied ratios. The span of each sample was relatively large dried PEC granules and gels are shown in Fig. 2. Some pores
(1.54–1.65) and no significance was found between them, indi- and porous structures were found in the granules of PEC 10
cating a large and similar size distribution width of the and PEC 20, whereas those of PEC 30 showed fewer pores and
materials. The water absorption (Table 1) of CMS/COS PECs in more solid and compacted morphology. The micrographs of
SBF was also strongly dependent on the ratio of CMS to COS. the lyophilized gels (Fig. 2B, D and F) depicted that all the PEC

Fig. 2 SEM micrographs of PEC 10 (A, B), PEC 20 (C, D) and PEC 30 (E, F) in the form of xerogel granules (A, C, E) and lyophilized gels prepared by
mixing PEC granules and PBS solutions (B, D, F) (scale bar: 10 μm for (A, C, E) and 50 μm for (B, D, F)).

3336 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper

granules could coalesce and form integrated porous gel recovered slowly to 7.4 after 21 days. No significant difference
matrixes when equilibrium absorbed PBS and, with an in pH values was observed among the PECs with different
increase of the COS content, smaller pores and less intercon- ratios of CMS to COS. The results suggested that the degra-
nected porous structures were found, which was in agreement dation rate of the PECs could be regulated by changing the
with the results of water absorption. ratio of CMS to COS without affecting the pH values.

3.2 Degradation in PBS 3.3 Whole blood clotting


The degradation profiles of CMS and CMS/COS PECs in PBS, The blood clotting experiments were performed to evaluate
which were expressed as the weight loss ratio as a function of blood clotting efficiency of the PECs and study how they inter-
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

the incubation time, are presented in Fig. 3(A). CMS powders acted with red blood cells (RBCs). As illustrated in Fig. 4(E),
were completely dissolved in PBS within one day and forming the absorbance at 540 nm corresponded to the amount of red
PECs with COS significantly reduced their degradation rate. All blood cells that were not trapped in the clot at each time
the PECs showed a rapid weight loss rate during the first week, point, which means the samples with lower absorbance have
followed by gradually reduced degradation rate until the end higher efficiency to form clots and trap RBCs. All the absor-
of incubation time. The degradation rates were highly depen- bance profiles showed a dramatic reduction within 3 min
dent on the ratio of CMS to COS and reduced significantly because of proceeding of blood clotting, and then became flat
with a higher COS content in the PECs. After 21 days of incu- when relatively stable blood clots were formed. The highest
bation, the weight loss of PEC 10 was 92%, while that of PEC blood clotting efficiency (lowest absorbance at 540 nm) was
20 and PEC 30 was 77% and 65%, respectively. The changes of found with both CMS and PEC 10, indicating that forming
pH values in the PBS are shown in Fig. 3(B). The pH values of PECs with a small amount of COS would not affect the blood
all the PECs dropped to approximately 7.0 within 3 days and clotting capacity. However, when the COS content in the PECs

Fig. 3 Weight loss (A) and pH change (B) of PBS after immersion of CMS and CMS/COS PECs.

Fig. 4 SEM observations of the stable clot formed after 3 min treatment with CMS (A), PEC 10 (B), PEC 20 (C) and PEC 30 (D) (scale bar: 10 μm); (E)
whole blood clotting efficiency.

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3337
View Article Online

Paper Biomaterials Science

increased to 20% and 30%, the absorbance increased signifi- 3.5 Antimicrobial activity
cantly meaning reduced blood clotting capacity. SEM obser- Antimicrobial activities of CMS and CMS/COS PECs against
vations of the clot formed after 3 min treatment with CMS and S. aureus are shown in Fig. 6. CMS showed no antimicrobial
the PECs are shown in Fig. 4(A–D). The disc-shaped RBCs activity, while the PECs exhibited different degrees of bacter-
adhered to all the materials. More RBCs were found with iostasis depending on the ratio of CMS to COS. With an
regard to CMS and PEC 10 compared with other PECs, which increase in the COS content in the PECs, the antibacterial
was in agreement with the results of the OD value measure- rate increased correspondingly at each time point. In
ments. It was interesting to note that the RBCs adhered to the addition, the antibacterial rate of PEC 20 and PEC 30
PECs seemed to expand and lose their typical biconcave mor-
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

increased with the incubation time reaching 51% and 83% at


phology, which may be attributed to the PEC gels tightly 24 h, respectively, which could be due to the sustained
adhered to the surface of the RBSs. Particularly, in the PEC release of COS from the PECs, while that of PEC 10 increased
30 group the RBSs were completely enwrapped by PEC 30 gels within 16 h and exhibited a slight reduction at 24 h which
and became a part of the gel matrix. could be attributed to insufficient COS. The digital images of
the resulting S. aureus colonies after incubation with CMS
3.4 PRT, APTT, PT, and TT and the PECs (Fig. 6A–D) are also presented, which were con-
Effects of the PECs on plasma coagulation were investigated by sistent with the results of Fig. 6E. Neither CMS nor CMS/COS
PRT, APTT, PT and TT measurements. As shown in Fig. 5(A, B PECs showed significant antimicrobial activity against E. coli
and C), the PRT, APTT and PT values were significantly (data not shown).
reduced by CMS and CMS/COS PECs compared with the
control (100%, without any agent). CMS and PEC 10 got the 3.6 Cytotoxicity
lowest values without a significant difference, while increasing Fig. 7(A) shows the cell viabilities of MC3T3-L1 fibroblasts
the COS content in the PECs (to 20% and 30%) led to pro- incubated with CMS or CMS/COS PECs, which were deter-
longed PRT, APTT and PT, indicating impaired effects on accel- mined by CCK-8 assays and normalized with the negative
erating plasma coagulation through the intrinsic pathway (PRT control (without any agent). PEC 30 exerted mild effects on
and APTT) and extrinsic pathway (PT). As for TT shown in MC3T3-L1 fibroblasts and the cell viability dropped from 95%
Fig. 5(D), no significant difference between these materials to 86% when the culture time increased from 1 to 7 days. The
and the control was observed, which indicated that CMS and relative cell viabilities retained slightly higher than 100% for
the PECs did not directly affect the common pathway. CMS, PEC 10 and PEC 20 at each time point and no evident

Fig. 5 Effects of CMS and CMS/COS PECs on plasma recalcification time (PRT) (A), activated partial thromboplastin time (APTT) (B), prothrombin
time (PT) (C) and thrombin time (TT) (D). *p < 0.05 compared with control. **p < 0.05 compared with PEC 10.

3338 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper


Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

Fig. 6 Antimicrobial activity of CMS and CMS/COS PECs against S. aureus. The resulting bacterial colonies incubated with CMS (A), PEC 10 (B), PEC
20 (C) and PEC 30 (D) for 24 h. (E) Antibacterial ratio against S. aureus at 8, 16 and 24 h.

Fig. 7 The viability (A) of MC3T3-L1 fibroblasts incubated with CMS and CMS/COS PECs and CLSM images of the cells for PEC 10 (B), PEC 20 (C),
PEC 30 (D) after 7 days of cultivation (scale bar: 50 μm).

cytotoxic effect was observed, indicating good cytocompatibil- blast characteristics with a bipolar structure were found for
ity. The cell morphology for CMS/COS PECs at day 7 was PEC 10 and PEC 20 groups, while the cell morphology for the
observed by CLSM and is shown in Fig. 7(B–D). Typical fibro- PEC 30 group was more spherical and not very well spread

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3339
View Article Online

Paper Biomaterials Science

which could be attributed to the mild effect of PEC 30 on the inflammatory response was found with some residuals of PEC
cells. 10 in the 1st week (Fig. 9a and g) and disappeared with its
complete degradation in the 4th week (Fig. 9d and j), indicat-
3.7 Hemostatic efficacy and tissue compatibility in a rabbit ing good tissue compatibility. The incision was surrounded by
hepatic hemorrhage model a small amount of fibrous tissues in the 1st week (Fig. 9a), and
Fig. 8(A–D) illustrate the hemostasis model of rabbit hepatic after 4 weeks, no fibrous tissue was observed around the
hemorrhage. After applying CMS/COS PECs on the injured incision (Fig. 9d). The CMS group (Fig. 9c, i, f and l) showed a
liver (Fig. 8A), a sticky gel matrix (Fig. 8C) which further trans- similar tissue response compared to the PEC 10 group with
formed into hemostasis clots (Fig. 8D) was observed and slightly worse wound healing performance, while acute inflam-
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

hemostasis was achieved rapidly. The hemostatic time mation was induced and adhesion between the two liver lobes
(Fig. 8E) of CMS and CMS/COS PECs was significantly shorter mediated by plenty of fibrous tissues occurred for the PEC
than that of commercial GS (121 s). More specifically, the 30 group during the 1st week (Fig. 9b and h) and then sub-
hemostatic time of PEC 10 (48 s) was significantly shorter than sided in the 4th week (Fig. 9e and k). For the PEC 30 group, it
CMS (61 s), PEC 20 (67 s) and PEC 30 (82 s). is worth noting that owing to the relatively low degradation
Fig. 9 shows the typical gross views and H&E staining of rate and persistent release of COS with PEC disassembly,
liver defects treated with PEC 10, PEC 30 and CMS after 1 and inflammatory cells were still observed and the tissue response
4 weeks, respectively. For the PEC 10 group, extremely mild remained after 4 weeks (Fig. 9k).

Fig. 8 Hemostasis in a rabbit hepatic hemorrhage model. (A) Incision and profuse bleeding of the left lobe of the liver. (B) Treatment with PEC 10
on the bleeding site. (C) Sticky gel matrix and (D) subsequent hemostasis clots created by PEC 10. (E) Corresponding hemostatic time. *p < 0.05
compared with GS. **p < 0.05 compared with CMS, PEC 20 or PEC 30.

3340 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper


Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

Fig. 9 Gross views (a–f ) and H&E staining (g–l, ×100, scale bar: 100 μm) of liver defects treated with PEC 10 (a, d, g, j), PEC 30 (b, e, h, k) and CMS
(c, f, i, l) after 1 (a–c, g–i) and 4 (d–f, j–l) weeks.

4. Discussion philicity of CMS and COS, CMS/COS PECs, especially PEC 10,
were also highly hydrophilic and had excellent water absorp-
Starch based hemostatic agents have been investigated by tion capacity which were essential for hemostasis application.
many studies, and their hemostatic efficacy mainly results Generally, the structure and properties of PECs strongly
from their function as a molecular filter by separating serum depend on parameters such as molar ratio, pH and ionic
from the cellular constituents such as platelets and erythro- strength. Highly hydrophilic PEC microgels with small particle
cytes. They can absorb much of the water portion of the blood sizes, high water content and large negative or positive charge
rapidly leading to concentrated blood solids, which then form values are obtained with an excess of polyanionic or polycatio-
a gel matrix. The gel matrix slows blood flow and serves to nic charges, respectively. In contrast, insoluble PEC precipi-
enhance blood clotting.7,8,40 Therefore, water absorption tates with large particle sizes, low water content and near zero
capacity is the key factor responsible for the hemostatic total charges are derived from equal charge molar ratios of
efficacy of starch based hemostatic agent. In the present study, polyelectrolytes.41–43 In our study, the water absorption, par-
both CMS and COS could be completely dissolved in water, ticle size, zeta potential and degradation rate of CMS/COS
while CMS/COS PECs were insoluble due to electrostatic inter- PECs were highly dependent on the ratio of CMS to COS,
actions between CMS and COS which were supported by the which can be explained by enhanced electrostatic interactions
results of FTIR analysis (Fig. 1). Owing to the excellent hydro- with the increase of the COS content (supported by FTIR

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3341
View Article Online

Paper Biomaterials Science

results). Therefore, the performance of CMS/COS PECs on clot. So the efforts to accelerate any of the above phases can be
hemostasis could also be strongly influenced by the CMS/COS helpful in achieving a high hemostatic efficacy.35,46 The hemo-
ratio as confirmed by the results of hemostatic efficacy. In static efficacy of CMS and CMS/COS PECs was evaluated
addition, the results of in vitro degradation in PBS suggested in vitro by whole blood clotting assay (Fig. 4) and in vivo in a
that CMS/COS PECs were degradable and their degradation rabbit hepatic hemorrhage model (Fig. 8) in consideration of
rates decreased with the increase of the COS content (Fig. 3). the fact that the liver is the most representative organ due to
Furthermore, their degradation rates were suitable when used its abundant blood supply and sensitivity to traumatic hemor-
as absorbable hemostatic agents, while rapidly dissolved CMS rhaging. The highest hemostatic efficacy both in vitro and
may easily enter a blood vessel leading to general or systemic in vivo was achieved by PEC 10, which was consistent with the
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

emboli and thrombosis.13 results of water absorption (highest), particle size (smallest)
A previous study suggested that material-induced blood and zeta potential (most negatively charged). Interestingly, no
plasma coagulation is a surface-mediated event, which is significant difference was found between CMS and PEC 10 in
related to both the surface area for immobilizing participants the whole blood clotting assay, whereas the hemostatic time
of surface-dependent clotting reactions and the surface pro- for PEC 10 in vivo was significantly shorter than CMS. Our
perties for activating the intrinsic pathway of the blood clotting hypothesis is that CMS and PEC 10 showed similar surface
cascade of hemostatic agents.13,44 More recently, material- properties (Table 1, Fig. 4A and B) for material-induced blood
induced rapid blood coagulation was also achieved by chito- plasma coagulation leading to similar results of blood clotting
san–catechol via interacting with plasma proteins.23 For hemo- assay (Fig. 4E);1,13,14 in the case of in vivo hemostasis, highly
static application, PEC 10 particles with a smaller particle size hydrophilic but insoluble PEC 10 could form a sticky gel
(12.4 μm) and larger negative charge value (−20.6 mV) were matrix to adhere to the injured site and occlude the blood
more propitious to contact activation of blood leading to accel- vessel rapidly (Fig. 8), while soluble CMS dissolved in blood
erated plasma coagulation demonstrated by the results of and couldn’t occlude the blood vessel until fibrin clots were
whole blood clotting (Fig. 4), with respect to PEC 20 and PEC formed.19,23
30. In summary, the hemostatic process of CMS/COS PECs
Since blood plasma coagulation is the key phase during could be depicted as follows. CMS/COS PECs could absorb
hemostasis, effects of CMS and CMS/COS PECs on plasma much of the water portion of the blood rapidly leading to a
coagulation including PRT, APTT, PT and TT were determined sticky gel matrix which could tightly adhere and occlude the
(Fig. 5) in order to further understand the hemostatic mecha- injured site to stop bleeding efficiently. Subsequently, the gel
nisms. PRT and APTT are both evaluation for the intrinsic matrix could accelerate the coagulation process by providing
coagulation pathway, while the influence of platelets is negatively charged and hydrophilic surfaces which are
involved for the PRT and not involved for the APTT.45 Both the regarded as key factors to activate the intrinsic pathway of the
PRT and APTT were significantly shortened for CMS and the coagulation cascade. Ultimately, stable fibrin clots were
PECs compared with the control, and the increase of the COS formed to reinforce the primary occlusion and hemostasis was
content in the PECs resulted in a prolonged PRT and APTT, achieved.
which along with the results of particle size and zeta potential Since infection is often induced by trauma and bleeding,
were in agreement with the principles of material-induced hemostatic agents with antimicrobial activity can be
contact activation. Interestingly, although previous research helpful.5,17,47 The antimicrobial activity of COS has been con-
reported that the hemostatic activity of chitosan and COS is firmed by previous studies, and can be explained by altered
related to platelet activation, our study suggested that the permeability of the microbial cell membrane when positively
plasma coagulation process was mainly activated and acceler- charged COS bind to the bacterial cell wall.24,48 CMS/COS
ated by negatively charged CMS rather than positively charged PECs also showed antimicrobial activity against S. aureus
COS since both the PRT and APTT of PEC 10 were significantly (Fig. 6) but failed against E. coli. Unsurprisingly, the anti-
lower than those of PEC 30. A similar trend was observed for microbial activity against S. aureus enhanced with the increase
PT, which is evaluation for the extrinsic coagulation pathway, of the COS content, suggesting that the antimicrobial activity
indicating that CMS could also accelerate the extrinsic coagu- derived from COS and the contribution of CMS were
lation pathway. As for TT, which is evaluation for the common unapparent.
coagulation pathway, all the samples did not show significant In addition to hemostatic efficacy and antimicrobial
impact on TT. Based on this evidence, it could be concluded activity, another important issue was biological safety for an
that CMS could accelerate both the intrinsic and extrinsic absorbable hemostatic agent. The cytotoxicities of CMS and
coagulation pathways and not directly affect the common CMS/COS PECs were evaluated with MC3T3-L1 fibroblasts
pathway, while the effect of COS on plasma coagulation was using CCK-8. The obtained data revealed that only PEC 30
not that dramatic. showed a mild effect on the proliferation and morphology of
Typical hemostasis refers to the process of halting the MC3T3-L1 (Fig. 7), which might be due to the excessive
bleeding from damaged blood vessels and proceeds in several amount of COS released to the cell culture medium resulting
sequential phases, including constriction of blood vessels, in too high cation concentration for cells to grow. Besides, no
activation of a coagulation cascade and formation of a blood evident cytotoxic effect was observed for CMS, PEC 10 and

3342 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018
View Article Online

Biomaterials Science Paper

PEC 20, indicating good cell compatibility. In accordance Conflicts of interest


with the cytotoxicity results of CMS/COS PECs, PEC 10
showed good tissue compatibility with extremely mild inflam- There are no conflicts to declare.
matory response which weakened over time, while the inten-
sity of the inflammatory response increased remarkably with
the increase of the COS content (PEC 30). The differences Acknowledgements
between PEC 10 and PEC 30 can be explained by different
contents of COS, which can cause inflammatory tissue This work was supported by the National Natural Science
responses and tissue adhesion,3,18 and different degradation Foundation of China (51772194, 51502340 and 81771990) and
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

rates. PEC 10 with a low COS content and fast degradation Key Medical Program of Science and Technology Development
rate resulted in a slight and fast release of COS leading to of Shanghai (17441900600 and 17441902000).
extremely mild inflammation weakened and eventually dis-
appeared, whereas PEC 30 with a high COS content and low
degradation rate resulted in a plentiful and long-term release References
of COS leading to acute inflammation and adhesion which
were still observed after 4 weeks. Interestingly, in consider- 1 T. A. Ostomel, Q. H. Shi and G. D. Stucky, J. Am. Chem.
ation of the good cytocompatibility and tissue compatibility Soc., 2006, 128, 8384–8385.
of CMS confirmed by this research, these in vitro and in vivo 2 M. Gabay, Am. J. Health-Syst. Pharm., 2006, 63, 1244–1253.
results also suggested that a large amount of water-soluble 3 S. Pourshahrestani, E. Zeimaran, N. A. Kadri, N. Gargiulo,
COS could cause mild cytotoxicity and induce intense inflam- H. M. Jindal, S. V. Naveen, S. D. Sekaran, T. Kamarul and
mation and adhesion despite of the accepted good biocom- M. R. Towler, ACS Appl. Mater. Interfaces, 2017, 9, 31381–
patibility of chitosan. 31392.
4 S. Ohta, T. Nishiyama, M. Sakoda, K. Machioka, M. Fuke,
S. Ichimura, F. Inagaki, A. Shimizu, K. Hasegawa,
5. Conclusion N. Kokudo, M. Kaneko, Y. Yatomi and T. Ito, J. Biosci.
Bioeng., 2015, 119, 718–723.
Three CMS/COS PECs with different ratios of CMS to COS 5 S. Y. Ong, J. Wu, S. M. Moochhala, M. H. Tan and J. Lu,
(9/1, 4/1, 7/3) have been developed as absorbable hemostatic Biomaterials, 2008, 29, 4323–4332.
agents for surgery and their hemostatic efficacy, hemostatic 6 R. L. Gu, W. Z. Sun, H. Zhou, Z. N. Wu, Z. Y. Meng,
mechanism and biocompatibility were evaluated. The results X. X. Zhu, Q. Tang, J. Dong and G. F. Dou, Biomaterials,
demonstrated that PEC 10 displayed superior hemostatic 2010, 31, 1270–1277.
efficacy but reduced with the increase of the COS content. 7 F. P. Chen, X. Y. Cao, X. L. Chen, J. Wei and C. S. Liu,
The results of whole blood clotting, plasma coagulation and J. Mater. Chem. B, 2015, 3, 4017–4026.
hemostasis in a rabbit hepatic hemorrhage model revealed 8 K. Bjorses, L. Faxalv, C. Montan, K. Wildt-Persson, P. Fyhr,
that the possible hemostasis mechanism of CMS/COS PECs J. Holst and T. L. Lindahl, Acta Biomater., 2011, 7, 2558–
was attributed to (1) physical absorption of the fluid com- 2565.
ponent of blood, (2) occlusion of the bleeding sites by 9 Y. Hou, Y. Xia, Y. K. Pan, S. C. Tang, X. F. Sun, Y. Xie,
forming a sticky gel matrix, and (3) activation and accelera- H. Guo and J. Wei, Mater. Sci. Eng., C, 2017, 76, 340–349.
tion of the coagulation cascade by negatively charged and 10 E. Assaad and M. A. Mateescu, Int. J. Pharm., 2010, 394, 75–84.
hydrophilic surfaces. To our knowledge, these findings 11 N. Masina, Y. E. Choonara, P. Kumar, L. C. du Toit,
provide the first evaluation and optimization of CMS/COS M. Govender, S. Indermun and V. Pillay, Carbohydr. Polym.,
PECs for hemostatic applications and are useful to under- 2017, 157, 1226–1236.
stand the hemostasis mechanism and underlying principles 12 L. P. Massicotte, W. E. Baille and M. A. Mateescu,
for designing PEC based hemostatic agents. The degradation Int. J. Pharm., 2008, 356, 212–223.
rates, which decreased with the increase of the COS content, 13 X. Yang, W. Liu, N. Li, M. Wang, B. Liang, I. Ullah,
were also suitable for absorbable hemostatic agents. In A. L. Neve, Y. Feng, H. Chen and C. C. Shi, Biomater. Sci.,
addition, PEC 10 displayed excellent cytocompatibility with 2017, 5, 2357–2368.
MC3T3-L1 and good tissue compatibility with a rabbit liver. 14 T. A. Ostomel, Q. H. Shi, P. K. Stoimenov and G. D. Stucky,
CMS/COS PECs also showed antibacterial activities against Langmuir, 2007, 23, 11233–11238.
S. aureus, which were enhanced with the increase of the COS 15 Y. X. Chen and G. Y. Wang, Colloids Surf., A, 2006, 289, 75–
content. Thus, it is reasonable to conclude that CMS/COS 83.
PEC with a suitable COS content (e.g., PEC 10) has a great 16 A. Bari, N. Bloise, S. Fiorilli, G. Novajra, M. Vallet-Regí,
potential for internal use as an absorbable hemostatic agent G. Bruni, A. Torres-Pardo, J. M. González-Calbet, L. Visai
and the above findings will open a new door for designing a and C. Vitale-Brovarone, Acta Biomater., 2017, 55, 493–504.
host of PEC based hemostatic agents with great potential in 17 C. L. Dai, Y. Yuan, C. S. Liu, J. Wei, H. Hong, X. S. Li and
clinical applications. X. H. Pan, Biomaterials, 2009, 30, 5364–5375.

This journal is © The Royal Society of Chemistry 2018 Biomater. Sci., 2018, 6, 3332–3344 | 3343
View Article Online

Paper Biomaterials Science

18 X. F. Huang, Y. F. Sun, J. Y. Nie, W. T. Lu, L. Yang, 34 B. K. Gu, S. J. Park, M. S. Kim, C. M. Kang, J. I. Kim and
Z. L. Zhang, H. P. Yin, Z. K. Wang and Q. L. Hu, Int. J. Biol. C. H. Kim, Carbohydr. Polym., 2013, 97, 65–73.
Macromol., 2015, 75, 322–329. 35 C. L. Dai, C. S. Liu, J. Wei, H. Hong and Q. H. Zhao,
19 J. H. Ryu, Y. Lee, W. H. Kong, T. G. Kim, T. G. Park and Biomaterials, 2010, 31, 7620–7630.
H. Lee, Biomacromolecules, 2011, 12, 2653–2659. 36 H. Wang, M. Li, J. M. Hu, C. H. Wang, S. S. Xu and
20 K. Kim, J. H. Ryu, D. Y. Lee and H. Lee, Biomater. Sci., C. C. Han, Biomacromolecules, 2013, 14, 954–961.
2013, 1, 783–790. 37 C. Kilkenny, W. J. Browne, I. C. Cuthill, M. Emerson and
21 J. M. Lee, J. H. Ryu, E. A. Kim, S. Jo, B.-S. Kim, H. Lee and D. G. Altman, Osteoarthritis Cartilage, 2012, 20,
G.-I. Im, Biomaterials, 2015, 39, 173–181. 256–260.
Published on 18 September 2018. Downloaded by Kings College London on 1/21/2019 2:26:23 AM.

22 Q. Xia, Z. Liu, C. Wang, Z. Zhang, S. Xu and C. C. Han, 38 Q. Zhao, J. Qian, Q. An, C. Gao, Z. Gui and H. Jin, J. Membr.
Biomacromolecules, 2015, 16, 3083–3092. Sci., 2009, 333, 68–78.
23 M. Shin, S. G. Park, B. C. Oh, K. Kim, S. JO, M. S. Lee, 39 M. S. Belluzo, L. F. Medina, A. M. Cortizo and
S. S. Oh, S. H. Hong, E. C. Shin, K. S. Kim, S. W. Kang and M. S. Cortizoa, Ultrason. Sonochem., 2016, 30, 1–8.
H. Lee, Nat. Mater., 2017, 16, 147–152. 40 P. Rajagopal and N. Hakim, Transplant. Proc., 2011, 43,
24 P. Zou, X. Yang, J. Wang, Y. F. Li, H. L. Yu, Y. X. Zhang and 424–426.
G. Y. Liu, Food Chem., 2016, 190, 1174–1181. 41 H. V. Saether, H. K. Holme, G. Maurstad, O. Smidsrød and
25 H. K. No, N. Y. Park, S. H. Lee and S. P. Meyers, Int. J. Food B. T. Stokke, Carbohydr. Polym., 2008, 74, 813–821.
Microbiol., 2002, 74, 65–72. 42 D. L. Cerf, A. S. Pepin, P. M. Niang, M. Cristea,
26 J. Y. Kim, J. K. Lee, T. S. Lee and W. H. Park, Int. J. Biol. C. Karakasyan-Dia and L. Picton, Carbohydr. Polym., 2014,
Macromol., 2003, 32, 23–27. 113, 217–224.
27 B. C. Lee, M. S. Kim, S. H. Choi, K. Y. Kim and T. S. Kim, 43 E. Spruijt, A. H. Westphalm, J. W. Borst, M. A. C. Stuart and
Int. J. Mol. Med., 2009, 24, 327–333. J. Gucht, Macromolecules, 2010, 43, 6476–6484.
28 C. B. Bucur, Z. Sui and J. B. Schlenoff, J. Am. Chem. Soc., 44 Z. Guo, K. M. Bussard, K. Chatterjee, R. Miller,
2006, 128, 13690–13691. E. A. Vogler and C. A. Siedlecki, Biomaterials, 2006, 27, 796–
29 P. Schaaf and J. B. Schlenoff, Adv. Mater., 2015, 27, 2420– 806.
2432. 45 C. Q. Xu, Y. J. Zeng and H. Gregersen, Med. Eng. Phys.,
30 M. Nagahata, R. Nakaoka, A. Teramoto, K. Abe and 2002, 24, 587–593.
T. Tsuchiya, Biomaterials, 2005, 26, 5138–5144. 46 T. A. Ostomel, Q. Shi, C. K. Tsung, H. J. Liang and
31 J. S. Martinez, T. C. S. Keller and J. B. Schlenoff, G. D. Stucky, Small, 2006, 2, 1261–1265.
Biomacromolecules, 2011, 12, 4063–4070. 47 S. Pourshahrestani, E. Zeimaran, N. A. Kadri, N. Gargiulo,
32 Y. Lu, D. L. Slomberg and M. H. Schoenfisch, Biomaterials, S. Samuel, S. V. Naveen, T. Kamarul and M. R. Towler,
2014, 35, 1716–1724. J. Mater. Chem. B, 2016, 4, 71–86.
33 G. G. Maghami and G. A. F. Roberts, Makromol. Chem., 48 C. Qin, Y. Zhang, W. Liu, L. Xu, Y. Yang and Z. Zhou, Fish
1988, 189, 195–200. Shellfish Immunol., 2014, 40, 267–274.

3344 | Biomater. Sci., 2018, 6, 3332–3344 This journal is © The Royal Society of Chemistry 2018

You might also like