You are on page 1of 12

Carbohydrate Polymers 251 (2021) 117036

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Injectable hydrogel derived from chitosan with tunable mechanical


properties via hybrid-crosslinking system
Jeong Wook Seo a, Su Ryon Shin b, Min-Young Lee c, Jae Min Cha d, Kyung Hyun Min e,
Sang Cheon Lee f, Seon Young Shin g, Hojae Bae a, *
a
Department of Stem Cell and Regenerative Biotechnology, KU Convergence Science and Technology Institute, Konkuk University, Seoul, 05029, Republic of Korea
b
Division of Engineering in Medicine, Department of Medicine, Harvard Medical School, Brigham and Women’s Hospital, Cambridge, MA 02139, USA
c
Smart Healthcare Research Institute, Biomedical Engineering Research Center, Samsung Medical Center, 81, Irwon-ro, Gangnam-gu, Seoul, 06351, Republic of Korea
d
Department of Mechatronics, College of Engineering, Incheon National University, Incheon, 22012, Republic of Korea
e
Department of Pharmacy, School of Pharmacy, Jeonbuk National University, Jeonju, Jeonbuk 54896, Republic of Korea
f
Department of Maxillofacial Biomedical Engineering and Institute of Oral Biology, School of Dentistry, Kyung Hee University, Seoul, 02447, Republic of Korea
g
Department of Internal Medicine, The Catholic University of Korea, Seoul, 02841, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Keywords: Thermo-sensitive injectable hydrogels that spontaneously react to physiological temperature have been widely
Chitosan studied to be used in biomedical fields. However, several challenges on their unstable structures with large-sized
Hydroxybutyl methacrylation pores and low mechanical strength under physiological conditions must be addressed to enable their practical
Injectable hydrogel
applications. We synthesized the hydroxybutyl methacrylated chitosan (HBC-MA) hydrogel that possesses both
Hybrid-crosslinking
Microstructure
thermo-sensitive and photo-crosslinkable properties. The HBC-MA showed effective sol-gel transition under
physiological temperature as well as a sensitive photo-crosslinkable property with visible light capable of skin
penetration. The co-nonsolvency property and thermo-sensitivity of HBC-MA prevented unintended loss of the
hydrogel graft after being subcutaneously injected in mice. Subsequently applied visible light on the skin beneath
which the hydrogel was injected significantly improved the mechanical strength and stability of the graft. The
injectable HBC-MA hydrogel developed in this study can be applicable to a wide range of biomedical fields such
as drug delivery system and tissue engineering.

1. Introduction physiological conditions, limiting its potential application (Cho et al.,


2016; Potta, Chun, & Song, 2010). For example, owing to such charac­
Hydrogels have been used in various biomedical applications teristics of thermo-sensitive hydrogels, an unintended early ‘burst’
because of their advantageous properties with high water content (Wang release of drugs could be caused when applied to drug delivery systems.
et al., 2010), biocompatibility (Wei et al., 2015), and three-dimensional (Bae, Okano, & Kim, 1989; Dong et al., 2012; Mukae, Bae, Okano, &
network structure (Qi et al., 2010; Zhu, Harris, & Zhang, 2016). Kim, 1990). On the other hand, photo-crosslinkable hydrogels retain
Thermo-sensitive hydrogels can spontaneously provoke a sol-gel tran­ different aspects in the injectability features, such that they show su­
sition in response to physiological temperature without additional perior mechanical strength and stability to the once they have been
chemical reactions and have been considered a useful biomaterial for photo-polymerized. However, if the graft was injected as
drug delivery system, regenerative medicine, and tissue engineering photo-polymerized, they are less versatile in the shapes of graft, and
(Chen et al., 2016; Fu et al., 2012). The controlled sol-gel transition of when injected as a prepolymer solution before polymerization, any
the hydrogels allows minimally invasive injection via syringes or cath­ leakages of the sample cannot be avoided, which would lead to a sig­
eters even to the irregularly shaped sites (Liu et al., 2017; Shibata et al., nificant loss of graft.
2010). However, structural variation in thermo-sensitive hydrogels can Chitosan is a biomaterial that has shown the following beneficial
result in low mechanical strength and instability of the gel under abilities: inhibition of infection (Lin, Lin, & Chen, 2009),

* Corresponding author at: KU Convergence Science and Technology Institute, Department of Stem Cell and Regenerative Biotechnology, Konkuk University,
Hwayang-dong, Gwangjin-gu, Seoul, 05029, Republic of Korea.
E-mail address: hojaebae@konkuk.ac.kr (H. Bae).

https://doi.org/10.1016/j.carbpol.2020.117036
Received 19 February 2020; Received in revised form 19 August 2020; Accepted 31 August 2020
Available online 6 September 2020
0144-8617/© 2020 Elsevier Ltd. All rights reserved.
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

anti-inflammation (Azuma, Osaki, Minami, & Okamoto, 2015), mold with a diameter of 8 mm and a height of 2.5 mm was filled with
biocompatibility, and biodegradability (Rinaudo, 2006). In addition, it HBC-MA solution and then Omnicure S2000 (Lumen Dynamics; Mis­
consists of many reactive functional hydroxyl and amino/acetamido sissauga, ON, Canada) equipped with a visible light filter (450− 550 nm)
groups with ease of diverse chemical modifications. In the previous re­ was used. The luminous intensity was 10 mW/cm2, and each sample was
ports, hydroxybutylation of chitosan resulted in an efficient sol-gel exposed to visible light for 10, 20, and 30 s at room temperature (27 ◦ C).
transition under physiological temperatures by influencing hydrogen For hybrid-crosslinking, the samples were incubated at 37 ◦ C in PBS.
bonds between hydrophilic and hydrophobic groups (Bin, Changzheng,
1
Chunlin, Qisheng, & Dajun, 2010). Meanwhile, methacrylation of chi­ 2.3. H nuclear magnetic resonance (NMR) spectroscopy
tosan has been frequently used to make photo-crosslinkable hydrogels
1
with photo-initiators to enable photo-polymerization. H NMR spectroscopy was performed to determine the degree of
The aim of this study is to develop a novel injectable, stable, robust, substitution (DS) of HBC-MA. Chitosan, HBC, and HBC-MA were dis­
chitosan-based hydrogel using photo-crosslinkable and thermo-sensitive solved in 0.25 % DCl at 1% (w/v) and analyzed using the 500 MHz FT-
chemical strategies, to enable robust & conformable gel administration NMR spectrometer (Varian; Palo Alto, California, USA).
to different shaped injection sites. Our hypothesis is that HBC-MA ob­
tained by conjugating hydroxybutylation and methacrylation groups to 2.4. Fourier transform infrared spectroscopy (FTIR)
the side chains of chitosan will form a hybrid-crosslinked hydrogel
under physiological conditions in the presence of a photo-initiator. Our The chemical structures of chitosan, HBC, and HBC-MA at room
investigation uses the crosslinking strategies of thermal, photo, and temperature were analyzed using an FTIR - 4100 (Jasco; Tokyo, Japan).
hybrid-crosslinking of HBC-MA. Cytotoxicity and the physicochemical Two (2) mg of samples (chitosan, HBC, and HBC-MA) was mixed with
properties including degradation, de-swelling are investigated. Inject­ 100 mg potassium bromide, and pellets were prepared for the analysis
ability of the novel hydrogel is investigated using a mouse model. (Shigemasa, Matsuura, Sashiwa, & Saimoto, 1996).

2. Materials and methods 2.5. Scanning electron microscopy (SEM)

2.1. Materials The HBC-MA hydrogels were prepared under three conditions
(thermal-crosslinking, photo-crosslinking, and hybrid-crosslinking). The
Chitosan (50~190 kDa; 20~300 cP, 1 wt. % in 1% acetic acid; prepared hydrogel was carefully sealed in a 50 ml conical tube, then
75~85 % deacetylated), methacrylic anhydride (MA), 1,2-epoxybutane, immersed in liquid N2 for 10 min and frozen. After unsealing, lyophili­
eosin Y (dye content~99 %), triethanolamine (TEOA), 1-vinyl-2-pyrro­ zation was performed at − 80 ◦ C for 24 h. All samples were cut with a
lidinone (NVP, contains sodium hydroxide as an inhibitor, >99 %), sharp knife to provide a clear view of the interior. Samples were then
2,4,6-trinitrobenzene sulfonic acid solution (TNBS), and glycine were secured on a stub via carbon tapes and coated with platinum. Cross-
purchased from Sigma-Aldrich (St. Louis, MO, USA). Fetal bovine serum sectional morphologies were imaged with an SU-8010 Scanning Elec­
(FBS), penicillin streptomycin (P/S), high glucose Dulbecco’s Modified tron Microscope (Hitachi; Tokyo, Japan). The pore size was confirmed
Eagle’s Medium (DMEM), Dulbecco’s Phosphate Buffered Saline by measuring all lyophilized pores in the SEM image at 300x
(DPBS), and trypsin-EDTA solutions were purchased from WelGene magnification.
(Daegu, Gyeongbuk, Korea). All other chemical agents used in this paper
were analytical grade. 2.6. Reversible characterization of hybrid-crosslinked hydrogel

2.2. Preparation of HBC-MA solution and hydrogel The 2,4,6-Trinitrobenzenesulfonic acid (TNBS) assay was applied to
measure glycine mixed in the water evacuated during thermal cross­
The HBC-MA was synthesized through two processes of hydrox­ linking (Zhang & Chu, 2004). Glycine was dissolved at a concentration
ybutylation and methacrylation of chitosan. To start, 1 g of chitosan was of 2 mg/mL in a 3 % (w/v) HBC-MA solution. Glycine-loaded hydrogel
completely dissolved in 60 ml of 0.1 M Acetic acid followed by pH was then immersed in PBS (pH 7.4) for 24 h in the 37 ◦ C incubator. The
adjustment to 6 using 5 M NaOH and addition of 20 ml of 1,2-epoxybu­ hydrogel was then carefully transferred to PBS at 20 ◦ C and 37 ◦ C for
tane (99 %). Then the mixture was reacted at 55 ◦ C for 24 h. After the intervals of 30 min. To measure evacuated glycine, 0.25 ml of 0.01 %
reaction was complete, washing was performed using a centrifuge to (w/v) TNBS solution was added to 0.5 ml of each sample solution. The
remove any remaining 1,2-epoxybutane. Finally, hydroxybutyl chitosan samples were then incubated at 37 ◦ C for 2 h. Absorbance was measured
(HBC) solution was lyophilized for 4 days (Tsukamoto, Akagi, Shima, & at 335 nm using a UV/vis spectrophotometer (Optizen POP; Seoul,
Akashi, 2017). Korea).
To make HBC-MA, 1 g of the recovered HBC was dissolved in 60 ml of
distilled water overnight followed by the addition of methacrylic an­ 2.7. Rheological analysis
hydride (94 %) at a rate of 1 ml/min until 15 % (v/v) was reached, and
was allowed to react for 4 h (pH 7). The mixture was dialyzed against The temperature-dependent rheological properties of the HBC-MA
distilled water for 1 week using 12− 14 kDa cut-off dialysis tubing to hydrogel were measured using the MCR302 Rheometer (Anton Paar
remove residual methacrylic anhydride, and finally the purified HBC- Ltd.; Graz, Austria) with parallel plate geometry (50 mm in diameter).
MA polymer was lyophilized (Yoon et al., 2016). The temperature was increased at a rate of 1 ◦ C/min between 4 and 50
For preparation of HBC-MA solution, lyophilized HBC-MA was ◦
C. Small amplitude y (0.02) and low frequency x (1.00 rad/s) were
completely dissolved in distilled water to a 3 % (w/v) concentration and applied for temperature sweep measurements of the HBC-MA solution.
then 0.01 mM Eosin Y which is a Food and Drug Administration (FDA)-
approved photoinitiator responding to visible light (450− 550 nm), 0.1 2.8. Mechanical testing
% (v/v) TEOA as a co-initiator, and 37 nM NVP as a co-monomer were
dissolved. Based on a previously reported study, we refer to a compo­ A CT3 Texture analyzer (Brookfield; Toronto, ON, Canada) with a
sition with high biocompatibility of photo-initiators (Bahney et al., 4500 g load cell (Brookfield; Toronto, ON, Canada) in compression
2011; Noshadi et al., 2017). The HBC-MA solution was then adjusted to mode was used to measure the compressive strength of photo-
pH 7.4. crosslinked HBC-MA hydrogel. Two types of hydrogel (photo-cross­
To perform photo-crosslinking, a polydimethylsiloxane (PDMS) linked and hybrid-crosslinked hydrogel) were prepared with three types

2
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

of visible light exposure time (10, 20, and 30 s) to determine the dif­ 37 ◦ C in 5 % CO2 atmosphere. After 6, 12, and 24 h incubation, 700 μl of
ference in strength. Each sample was exposed to visible light (as the staining solution was added according to the manufacturer’s pro­
described in Section 2.2) at room temperature. Photo-crosslinked tocol, and cells were imaged using fluorescence ECLIPSE Ts2 microscope
hydrogel samples were stabilized in PBS at RT for 15 min so that no (Nikon; Tokyo, Japan). The imaged cells were counted using ImageJ
further crosslinking occurred. Hybrid-crosslinked hydrogel samples 1.50a software (National Institutes of Health, Bethesda, Maryland). The
were then reacted at 37 ◦ C for 15 min for further thermal-crosslinking. A ratio of live cells to the total number of cells was used as a metric of cell
probe 12.7 mm in diameter was used to compress with a trigger load of viability.
0.05 N at a test speed of 0.05 mm/s. The compressive modulus was
determined as the slope of the linear region corresponding to 5~15 % 2.13. Cell counting kit-8 (CCK-8) assay
strain (Nichol et al., 2010).
To further investigate the cell proliferation and cytotoxicity of the
2.9. Swelling test HBC-MA, cell viability was investigated using a Cell Counting Kit-8
(CCK-8; Dojindo, Kumamoto, Japan). HBC-MA was prepared using the
The HBC-MA solution was pipetted into a PDMS mold. The sample same method (as described in the Section 2.13). Cells were seeded in 96-
was exposed to visible light (as described in the Section 2.2) at room well plates at 1000 cells per well in 100 μl of DMEM (10 % FBS, 1 %
temperature. Single group samples were prepared that were exposed to penicillin) supplemented with 0.1, 1, and 10 % (v/v) sterilized HBC-MA
visible light for 20 s. The photo-crosslinked hydrogels were stored at 37 solution, and incubated at 37 ◦ C in 5 % CO2 atmosphere. After 24, 48,

C. The swollen weight (St) was measured at the predetermined time and 72 h incubation, the CCK-8 assay was performed according to the
intervals in the PBS-containing samples, and the dry weight (Dt) was manufacturer’s protocol. After incubation at 37 ◦ C for 2 h, the absor­
measured after lyophilization. Finally, swelling ratio (%) was calculated bance of the samples was measured at 480 nm with a Microplate
as follows: Spectrophotometer (Biotek Instruments, Vernon Winston, USA).

Swelling ratio (%) = St / Dt × 100


2.14. Subcutaneous injection study of HBC-MA hydrogels in mice

All experimental procedures were approved by the Institutional


2.10. Hydrogel degradation Animal Care and Use Committee (IACUC) of Konkuk University (KU-
20035) and these mice were handled in accordance with the guidelines
Samples for degradation test were prepared by the method described approved by the IACUC of Konkuk University. The hydroxybutyl
in Section 2.2. Degradation tests were performed using lysozyme methacrylated chitosan (HBC-MA) solution was injected into mice
(40,000 U/mg, Sigma, USA). The samples were placed in PBS containing without incision, and experiments were conducted to confirm that
lysozyme at concentrations of 0.5 mg/mL. Control samples were placed thermal- and photo-crosslinked stable hydrogels could be formed.
in PBS without lysozyme. The hydrogels were then carefully stored at 20 Additionally, it was intended to determine the biodegradation tendency

C and 37 ◦ C for a defined time course. To keep the enzyme active, the of hydrogels according to the control of photo-crosslinking in vivo. HBC-
PBS containing lysozyme was changed daily. After 1, 3, 5, 7, 14, and 21 MA solution was sterilized using a PES syringe filter (0.22 μm pore size;
d, the samples were thoroughly washed with deionized water to remove Millipore; Bedford, Massachusetts, USA). Athymic nude mice (20~30 g,
any remaining PBS and lysozyme, and the water remaining on the sur­ 8 weeks old, male) were purchased from Orient Bio (Seongnam,
face was removed by blotting with a kimwipe. Then, each hydrogel Kyunggido, Korea). Four mice per group (0 s, 10 s, 20 s, and 30 s) were
sample was lyophilized and the dry weight (Dt) was recorded. Four used and then the implanted constructs were retrived at 1, 3, and 5 days.
samples were measured for freeze-dry weight immediately after prepa­ HBC-MA solution without visible-light exposure (0 s) was used as con­
ration and the mean weight (D0) was recorded. Gel remaining (%) was trol. The dorsal skin of each mice was divided into 4 types according to
calculated as follows: the light exposure time. Total of twelve mice were weighed and intra­
peritoneally injected using alfaxan 24 mg/kg and xylazine 10 mg/kg.
Mass loss (%) = (D0 − Dt) / D0 × 100 After sterilizing the injection site with 70 % (v/v) ethanol (99 %), HBC-
MA solution was injected into the subcutaneous dorsal through the
needle. All injection procedures were performed identically. To main­
tain a constant injection volume, a syringe pump (Legato 101; KD Sci­
2.11. In vitro hydrogel crosslinking test
entific, Holliston, MA, USA) was run for 10 s at an injection rate of 0.6
ml/min for all the mice to inject 100 μl of 3 % (w/v) HBC-MA solution.
The HBC-MA solution at 20 ◦ C was added into deionized water at 37
Photo-crosslinking was induced by exposing the injection site to visible

C in a petri dish using a 20 G needle syringe. Compression force of the
light using the Omnicure S2000 (Lumen Dynamics, Canada), which is
syringe was adjusted manually (Fig. S1A and B, Supporting Informa­
equipped with a visible light filter (450~550 nm), for a defined time
tion). The sample was exposed to visible light (as described in Section
period (0, 10, 20, and 30 s). Other than at the injection site, the light was
2.2). The distance between the sample and the light source was main­
completely blocked. The light intensity was 10 mW/cm2. For safety
tained at 2 cm (Fig. S1C, Supporting Information).
reasons, the light guide was always at least 3 cm away from the mice.
After the visible-light exposure step, the mice were placed in the mouse
2.12. Live/Dead fluorescence assay cage. After 1, 3, 5 d, mice were intraperitoneally injected using alfaxan
and xylazine, and then the dorsal subcutaneous skin was resected using
To investigate the cytotoxic effect of the HBC-MA, cell viability was surgical instruments. The obtained dorsal skin samples were fixed in 4%
investigated using a calcein AM/ethidium homodimer Live/Dead assay
formaldehyde and then hydrogel samples were carefully isolated. The
kit (Invitrogen, Carlsbad, CA). NIH3T3 fibroblast cells were grown and separated hydrogel was weighed immediately. Then each hydrogel
maintained in Dulbecco’s Modified Eagle Medium (DMEM; Welgene,
sample was lyophilized and the dry weight (Dt) was recorded. Four
Daegu, Korea) containing 10 % fetal bovine serum (FBS). The cells were hydrogel samples made with 100 μl HBC-MA solution were measured for
cultured in an incubator supplied with 5 % CO2 at 37 ◦ C. The culture
freeze-dry weight immediately after preparation and the mean weight
medium was changed daily. Cells were seeded in 48-well plates at 103 (D0) was recorded. Mass loss (%) was calculated as follows:
cells per well in 700 μl DMEM (10 % FBS, 1 % penicillin) supplemented
with 0.1, 1, and 10 % (v/v) sterilized HBC-MA solution, and incubated at Mass loss (%) = (D0 − Dt) / D0 × 100

3
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 1. (A) Chemical structure of HBC-MA. Hydroxybutylation substituted for pristine chitosan is shown in blue, and methacrylation is shown in red. (B) 1H NMR of
pristine chitosan (black) versus HBC (blue), HBC-MA (red) to confirm chemical modification. (C) FTIR spectra of chitosan, HBC, and HBC-MA. (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article).

methacrylic anhydride bound to HBC-MA were labeled at 1.8 ppm (peak


4). This demonstrated that HBC with the addition of hydroxybutyl
groups, a modification of chitosan, resulted in the successful addition of
2.15. Statistical analysis methacrylamide to a proton of -NH2. The peaks from 3.5 to 4.0 ppm
(peaks 5~9, Fig. 1A and B) corresponded to the non-anomeric protons
To evaluate statistical significance, an ordinary one- and two-way (Yang et al., 2012). The degree of methacrylation was calculated based
analysis of variance (ANOVA) followed by Tukey test was performed. on the ratio of the integrated area of the peaks of chitosan moiety be­
Data are mean ± standard deviation (SD) and means were compared tween 3.5 and 4.0 ppm (peaks 5~9, Fig. 1A and B) to that of the CH2
using unpaired Student’s t-tests. A p-value<0.05 was considered to peaks of the methacrylate group at 5.7 and 6.11 ppm (peak 3, Fig. 1A
indicate statistical significance. All analyses were performed using and B). Since methacrylic anhydride bonds mainly to reactive amine
GraphPad Prism 8.0.2 (GraphPad Software; La Jolla, CA). groups on the chitosan backbone, the degree of methacrylation of chi­
tosan was found to be 24.8 % by 1H NMR using the above method
3. Results and discussion (Valmikinathan et al., 2012).

3.1. Synthesis and characterization of HBC-MA conjugate 3.1.2. Fourier transform infrared spectroscopy (FTIR)
The synthesis of HBC-MA was also confirmed through FTIR com­
3.1.1. 1H Nuclear magnetic resonance spectroscopy parison. In the FTIR of chitosan, a broad band in the region of
Chitosan was reacted with epoxybutane to obtain HBC through a 3247~3398 cm− 1 corresponded to the overlapping of the vibrations of
substitution reaction of epoxybutane to hydroxy and amine group under − OH and -NH. The bands at 1024 and 1067 cm− 1 presented the char­
neutral condition. Then, HBC-MA was obtained through reaction be­ acteristic band of 6− OH in chitosan. A peak of 1153 cm− 1 is typically
tween unsubstituted amine group of HBC and methacrylic anhydride. attributed to asymmetric C–O–C bridge. The absorption bands at 2911
New signals at 0.92 and 1.5 ppm (peak 1~2, Fig. 1A and B) were and 2873 cm− 1 can be presented to C–H symmetric and asymmetric
observed due to the substitution of methyl group of the hydroxybutyl. stretching (Fernandes Queiroz, Melo, Sabry, Sassaki, & Rocha, 2015). In
The degree of hydroxybutylation was calculated based on the ratio of the the FTIR of HBC, the band at 1020~1060 cm− 1, which mainly represents
integrated area of the peaks of chitosan moiety between 3.5 and 4.0 ppm 6− OH, is found to be less than that of the FTIR of chitosan spectra. The
(peaks 5~9, Fig. 1A and B) to that of the CH2 peaks of the hydroxybutyl new peaks at 2944~2892 cm− 1 and 1585 cm− 1 indicate bending and
group at 0.92 and 1.5 ppm. The degree of hydroxybutylation of chitosan C–H stretching of the newly formed CH3 groups due to the introduction
was found to be 73.64 % by 1H NMR using the above method. Peaks at of 1,2-epoxybutane. It was confirmed that the hydroxybutyl group was
5.7 and 6.11 ppm (peak 3, Fig. 1A and B) are methylene peaks that occur mainly substituted with 6− OH (Wang et al., 2013). In the FTIR of
as covalent bonds in methacryloyl groups are replaced by meth­ HBC-MA, a peak rise for HBC-MA at 1614 cm− 1 associated with the NH2
acrylamide. The methyl protons of the methacryloyl group in

4
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 2. Reversible physical properties of HBC-MA hydrogel. (A) Schema to describe the network formation of thermal-crosslinked HBC-MA hydrogel. Interactions
between hydrophobic chains occur with deswelling of H2O at physiological temperatures. (B) Schema to describe a hydrogel network; the methacrylic group of HBC-
MA was crosslinked using visible light irradiation in the presence of a photo-initiator at 20 ◦ C. (C) Schema to describe Hybrid-crosslinked HBC-MA hydrogel network
with both thermal-crosslinking and photo-crosslinking at 37 ◦ C. This was accompanied by a deswelling of H2O and showed reversible characteristics with respect to
temperature. (D~F) SEM images of micropore structure according to different crosslinking methods: thermal, photo, and hybrid-crosslinking. The magnifications of
SEM images are x100 and x300. (G~I) Pore distribution frequency obtained from SEM images of hydrogels prepared by different crosslinking methods: thermal,
photo, and hybrid-crosslinking. (J) Average pore size according to cross-linking method: thermal, photo, and hybrid-crosslinking. * p < 0.05, *** p < 0.001 vs.
hybrid-crosslinking group. Data are represented in means ± SD, n = 3. (K) Absorbance graph measured using TNBSA for glycine deswelling by reversible hybrid-
crosslinking. Data are represented in means ± SD, n = 3. A photograph of the water deswelling is attached.

band was observed. Newly formed amide 3, 2, and 1 band absorbances 3.2. Preparation and characterization of HBC-MA hydrogel
corresponded at 1306, 1542, and 1710 cm− 1. This confirms that the
additional substitution from HBC to HBC-MA was successfully conju­ 3.2.1. Morphological analysis
gated (Fig. 1B). The effect of cross-linking methods on the morphology of HBC-MA
hydrogels was investigated by SEM. Scanning electron microscopy

5
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 3. Rheological analysis and mechanical characterization. (A) Temperature dependent storage modulus (G’) and loss modulus (G’’) of HBC-MA aqueous solution.
In this process, temperature is increased from 4 to 50 ◦ C (heating rate: 1 ◦ C; frequency: 1 rad/s). (B) The temperature-gelation point at which G’ intersects G’’ is
shown. (C) Mechanical properties of photo-crosslinked and hybrid-crosslinking HBC-MA with various visible light exposure times. Compressive modulus was
compared with the time of exposure to visible light. *** p < 0.001 vs. photo-crosslinking 10 s; ### p < 0.001 vs. hybrid-crosslinking; +++ p < 0.001 between the
indicated group. Data are represented in means ± SD, n = 8. (D) A representative stress-strain curve of photo-crosslinking and hybrid-crosslinking HBC-MA hydrogel
with visible light exposure time.

micrographs clearly show the difference in pore size of the lyophilized 3.2.2. Reversible characteristics of HBC-MA hydrogel
hydrogel according to the crosslinking mechanism. HBC-MA solution The temperature reversible characteristics of hybrid-crosslinking
forms hydrophobic interaction domains by binding hydrophobic chains hydrogels were investigated by applying the internal water molecule
surrounded by water molecules at 37 ◦ C. This process is accompanied by evacuation phenomenon (Zhang, Xu, Cheng, & Zhuo, 2008). The reac­
an evacuation phenomenon in which water molecules are deswelled tion of HBC-MA with temperature occurs through
(Zhang, Xu, Cheng, & Zhuo, 2008). The thermal crosslinked hydrogel hydrophobic-interaction. As the temperature rises, crosslinking occurs
forms a porous matrix, but the micropores are relatively larger in between hydrophobic chains. This process is accompanied by an evac­
diameter (Fig. 2D). When the HBC-MA solution is exposed to visible uation phenomenon in which water molecules are released. The
light, photo-initiators bind between methacrylamide chains to form photo-crosslinked HBC-MA hydrogel gradually became opaque due to
crosslinking networks (Kuwajima, Yoshida, & Hayashi, 1976). When the the hydrophobic-interaction that occurred during hybrid-crosslinking at
degree of methacrylation of chitosan was 24.8 % and exposed to short 37 ◦ C, and gradually decreased in volume as water releasing occurred
time visible light for 10 s, the photo-crosslinking bridge was shorter than (Videos 1, Supporting Information). Conversely, the hybrid-crosslinked
the chitosan backbone, resulting in an elongated micropore (Fig. 2E). If hydrogel showed reversible characteristics in response to low temper­
more methacrylamide is introduced or exposed to light for a longer time, ature (20 ◦ C). When the temperature is lowered, crosslinking between
it is expected to form a denser matrix. Hydrogels formed by the hydrophobic chains is released. The opaque HBC-MA hydrogel
photo-crosslinking are responsive to temperature and can form tighter became transparent enough to see the text on the back and the volume
and more consistent micropores in response to temperature (37 ◦ C) increased again by swelling the surrounding water quickly (Videos 2,
(Fig. 2F). There was a difference in distribution of pore size depending Supporting Information). As a representative material of hydrogel,
on the crosslinking method used. The micropore size of the PEG-DMA, which has low reactivity to temperature, was used as a
thermal-crosslinked samples showed an irregular distribution (Fig. 2G), control. If the photo-crosslinked HBC-MA hydrogel can form additional
whereas the micropore size of the photo-crosslinked sample was uni­ crosslinking according to temperature, it is assumed that the evacuation
formly distributed (Fig. 2H). A smaller pore size and more regular phenomenon will occur, and the hydrogel will release the
structure than the pore structure of thermal-crosslinking hydrogel was glycine-dissolved water. In this case, it is possible to know whether the
observed. In the hybrid-crosslinked sample, the pore distribution was thermal crosslinking of HBC-MA occurred when the glycine concentra­
more constant than the photo-crosslinked sample (Fig. 2I). This more tion of PBS in which the hydrogel remained when the crosslinking
compact and regular pore structure provides additional mechanical occurred, and the crosslinking was released by temperature was
strength stability. The average pore sizes for lyophilized temperature, measured. When glycine-loaded photo-crosslinked HBC-MA hydrogel
photo, and hybrid-crosslinking hydrogel were 69.5 ± 5.5, 27.9 ± 3.5, was transferred to PBS at 20 ◦ C and 37 ◦ C at 30-minute intervals, the
and 18.0 ± 1.9 μm, respectively (Fig. 2J). absorbance was measured after reacting the released glycine with

6
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 4. Swelling and degradation characteristics. (A) Swelling of photo-crosslinked HBC-MA hydrogel from 0 to 36 h at 37 ◦ C in PBS. Photographs were taken before
and after the deswelling. Data are represented in means ± SD, n = 3. (B) Degradation of hybrid-crosslinked HBC-MA hydrogel from 0 to 21 d at 37 ◦ C. Two groups
(PBS and 0.5 mg/mL lysozyme) with three visible light exposure conditions (10, 20, and 30 s) were tested according to the lysozyme concentration. Data are
represented in means ± SD, n = 3. (C) Degradation of photo-crosslinked and hybrid-crosslinked HBC-MA hydrogel from 0 to 21 d. Two groups (20 and 30 ◦ C) with
three visible light exposure conditions (10, 20, and 30 s) were tested according to the temperature condition. Data are represented in means ± SD, n = 3.

TNBSA. Hydrogel reswelled the water at 20 ◦ C and released the water experiment, the gelation temperature of 3 % (w/v) HBC-MA solution
containing glycine at 37 ◦ C. The results showed that glycine was was 27.3 ◦ C. It was confirmed that thermal-crosslinking could be
released even after repeated reversible temperature changes. This con­ exhibited at physiological temperature.
firms that hybrid-crosslinking and photo-crosslinking can be reversibly
switched on and off by changing the temperature. In addition, the 3.2.4. Mechanical analysis
shorter photo-crosslinking time (10 s) resulted in unexpected glycine The reactivity of the HBC-MA solution was considerably high, so that
loss, even at 20 ◦ C. Perhaps the loss of glycine occurs as the pore, where photo-crosslinking occurred in a short time. When exposed to visible
the hybrid-crosslinking hydrogel condenses, is released again (Fig. 2K). light, photo-crosslinking between the methacrylamide group and the
It was confirmed that the temperature-reactive crosslinking was a photo-initiator resulted in the formation of a tighter pores as the visible
reversible reaction and was accompanied by a swelling and deswelling light exposure time increased. When exposed to visible light for 10 s,
phenomenon. Through this data, the photo-crosslinked hydrogel can be compressive modulus of the hydrogel sample was 15.1 ± 1.7 kPa. It was
a hybrid-crosslinked hydrogel at physiological temperature, and the confirmed once again that hydrogels could be formed even after expo­
hybrid-crosslinked hydrogel can be converted into photo-crosslinked sure to visible light for a short time (10 s). The 20 and 30 s groups
hydrogel in response to low temperature (20 ◦ C). This characteristic showed a statistically high difference compared to the 10 s group. The
can be expected to be applied to induce stimuli-drug release by inten­ 20 and 30 s groups were 34.1 ± 2.9 kPa (*** p < 0.001) and 42.0 ± 4.1
tionally creating low temperature environment in the hydrogel during kPa (*** p < 0.001), respectively. In the case of hybrid-crosslinking, the
drug release through continuous deswelling via physiological tempera­ compressive modulus of 10 s (hybrid) was 25.7 ± 2.7 kPa (+++ p <
ture in the body. 0.001), which is 41.5 % higher than 10 s exposure only to visible light.
This was a statistically significant difference. Due to thermal cross­
3.2.3. Rheological analysis linking, additional pore formation affected compressive modulus. The
Rheology analysis was performed to characterize the sol-gel transi­ 20 and 30 s (hybrid) groups did not show statistically significant dif­
tion of HBC-MA solution as a function of temperature. As shown in ferences when compared to the 20 and 30 s groups. The 20 and 30 s
Fig. 3A and B, the loss modulus (G”) is higher than the storage modulus (hybrid) groups were 37.4 ± 2.2 kPa and 45.4 ± 2.8 kPa, respectively.
(G’) at the beginning of the measurement at 4 ◦ C, but both show very Increasing the degree of photo-crosslinking by increasing the visible
low coefficients. As the temperature rises, the sol-gel transition takes light exposure time decreases the mobility of the polymer, thus limiting
place, and the storage modulus also rises slightly (Fig. 3A). However, the hydrophilic-hydrophobic interactions that occurred as the temper­
depending on the increase in temperature, storage modulus (G’) and loss ature rises, which led to no significant difference in modulus (Fig. 3C).
modulus (G”) showed different rates of rising. The gelation point was Differences between photo-crosslinking and hybrid-crosslinking groups
defined as the temperature at which the storage modulus (G’) and loss were also observed in the strain-stress curve. As the visible light expo­
modulus (G’’) intersect. The intersection of the storage modulus and loss sure time increased, the initial strain (~ 40 %) showed high stiffness for
modulus of the HBC-MA solution was 27.3 ◦ C, confirming that this point each group. In strain later point, yield points appeared at lower strains
is the temperature-gelation point (Fig. 3B). As the temperature rises past and yield stresses tended to decrease. As a result, both groups had
the gelation point, the storage modulus rises at a faster rate. In this stronger intensities with increased exposure to visible light and had

7
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 5. Cell viability analysis by Live/Dead assay. (A) Image of NIH3T3 fibroblasts stained for 6 to 24 h by calcein-AM (green) / ethidium homodimer (red) Live/
Dead assay in media containing various concentrations of HBC-MA prepolymer solution concentration (0, 0.1, 1, and 10 %). The scale bars represent 100 μm and
magnification is x100. (B) NIH3T3 viability (%) according to HBC-MA concentration from Live/Dead image analysis. There was no statistically significant difference.
Data are represented in means ± SD, n = 3. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article).

lower elastic regions. But when comparing the two groups at the same interaction could occur due to the temperature (37 ◦ C) even after
visible light exposure time, a unique difference was observed. At 10 s, photo-crosslinking. The interaction of hydrophobic residues with hy­
hybrid-crosslinking samples with additional thermal crosslinking drophilic residues results in condensation and release of water (Zhang &
compared to photo-crosslinking only had a lower yield stress value that Chu, 2004). During formation of the polymer network by temperature,
differed by more than 40 kPa at lower or similar strain points. At 20 s, the ratio between the swollen hydrogel and the dry hydrogel was
the strain point of hybrid-crosslinking was reduced compared to photo- reduced. When the occurrence of additional crosslinking due to tem­
crosslinking, but the difference gap was not significant compared to 10 s. perature reached saturation, the swelling ratio of the hydrogel showed
However, the addition of thermal crosslinking increased the deforma­ the lowest ratio. The lowest swelling ratio was observed at 4–6 h (2 h).
tion rate. At 30 s, the strain of hybrid-crosslinking increased by about 10 Hydrogel converted to hybrid-crosslinking after saturated thermal
% compared to photo-crosslinking. But in contrast to the previous re­ crosslinking showed an increase in the swelling curve from 6 to 12 h.
sults, it was found that the fracture stress showed only minor change After 12 h, the hybrid-crosslinked hydrogel a maintained equilibrium
(Fig. 3D). The 10 s (hybrid) sample with the largest change in thermal swelling ratio (Fig. 4A). This data showed that temperature-sensitive
properties showed a decreased fracture stress as the compressive crosslinking is possible even after photopolymerization once again,
modulus increased due to thermal crosslinking. On the other hand, the and deswelling can be continuously maintained in a physiological
30 s (hybrid) sample, which had no significant difference in compressive temperature environment such as 37 ◦ C.
modulus due to the change of thermal properties, showed a large change
in deformation as strain increased to the highest value. 3.2.6. Degradation test
Chitosan is known to degrade glycoside bonds by lysozyme (Pan­
3.2.5. Swelling test gburn, Trescony, & Heller, 1982). The lysozyme mainly contributes to
The swelling of the hydrogel network is important in applications as the initial degradation rate because it binds to the hexamer bond site
it affects mechanical properties and solute diffusion (Peppas, Hilt, with 3–4 acetylated units of chitosan (Pangburn et al., 1982). To confirm
Khademhosseini, & Langer, 2006). As has already been reported, in the difference in the biodegradation pattern of HBC-MA hydrogel ac­
conventional polymer scaffolds, the gel expands due to the interaction cording to the light exposure time, experiments were conducted under
between the polymer and solvent depending on the pore size (Chen, three light exposure time conditions (10, 20, and 30 s). First, to further
Park, & Park, 1999). But in the case of HBC-MA polymer, hydrophobic investigate the degradation patterns of HBC-MA, lysozyme (40,000

8
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 6. Cell viability analysis by cell counting kit-8 (CCK-8) assay. (A) NIH3T3 fibroblasts in media containing various concentrations of HBC-MA (0, 0.1, 1 and 10 %)
from 12 to 72 h by CCK-8 assay. It is a relative graph comparing the measured absorbance with that of the same time course (h) based on the 0 % group. ** p < 0.01,
*** p < 0.001 vs. Same time course (h) in 0% group. Data are represented in means ± SD, n = 7. (B) CCK-8 absorbance graph from 12 to 72 h of 10 % (v/v) HBC-MA
concentration in culture media group. ** p < 0.01, *** p < 0.001 vs. 12 h. Data are represented in means ± SD, n = 7. (C) Image by concentration at 48 h. The scale
bars represent 100 μm and magnification is x100.

U/mg, Sigma-aldrich, USA) was added to PBS at a concentration of 0.5 successfully crosslinked by the co-nonsolvency effect but was not diluted
mg/mL and compared at 37 ◦ C. In the 10 s condition where the most upon injection in deionized water. Subsequent thermal-crosslinking by
biodegradation occurred, 47.9 ± 2.4 % degraded, showing the fastest the temperature took place and was indicated by the alteration of
biodegradation rate in the early stage. In comparison, the biodegrada­ hydrogel color (opaque pink) (Fig. S1B, Supporting Information).
tion rate was relatively slow at 32.2 ± 2.7 % at 30 s condition. At 14 d, Sensitivity to temperature was confirmed to operate as expected. In
most of 10 s condition degraded, and most the 20 s condition degraded addition, it was exposed to visible light to confirm its sensitivity to light
at 21 d. However, in the 30 s condition, only 76.1 ± 3.2 % was bio­ (Fig. S1C, Supporting Information). Eosin Y, which was used as a
degraded even at 21 d, and the remaining hydrogel was present. On the photo-initiator, is shown as yellow and colorless when exposed to visible
other hand, in the PBS group without lysozyme, the biodegradation light (Noshadi et al., 2017). The area cross-linked by visible light turned
tendency was relatively slow for the initial 7 d. The mass loss ratio did yellow or colorless, providing visual confirmation of the sensitivity to
not exceed 40 % in 10, 20, and 30 s conditions for 21 d. Through this, the light (Fig. S1D, Supporting Information). HBC-MA was not diluted in
modified chitosan HBC-MA did not inhibit the activity of lysozyme and PBS as well as deionized water, and it was confirmed that the temper­
was confirmed to be affected by lysozyme (Fig. 4B). Next, the biode­ ature and photo-crosslinking functioned as behavior of HBC-MA. In
gradability of HBC-MA hydrogel at low temperature (20 ◦ C) was addition, it was confirmed that crosslinking with visible light is possible
confirmed. Unlike the biodegradation tendency shown at 37 ◦ C, no after temperature sensing (Videos 3, Supporting Information).
significant change occurred after 7 d under all conditions. In addition,
no mass loss took place more than 20 % (Fig. 4C). Through this, it was
3.3. In vitro cytotoxicity and cell viability test
confirmed that HBC-MA showed tendency to biodegrade at low tem­
perature. It can be anticipated that HBC-MA hydrogel or solution can be
3.3.1. Live/dead fluorescence assay
stored for a relatively long time when stored at low temperatures. These
To evaluate the effect of HBC-MA and photoinitiator on cell viability,
data suggested biodegradability of HBC-MA can be controlled according
NIH3T3 cells were cultured in media containing different concentrations
to temperature and light exposure time and has the potential to be
of HBC-MA solution. To compare the cytotoxicity of HBC-MA solution,
optimized and applied according to the specific application.
we used the NIH3T3 grown in cell-cultured media as a control group. At
the 6, 12 and 24 h time point, cell viability of all groups was higher than
3.2.7. Visual confirmation of co-nonsolvency effect and crosslinking in vitro
90 %, and therefore statistically significant cytotoxicity compared to
Hydroxybutyl methacrylamide chitosan (HBC-MA) is a polymer in
NIH3T3 cells grown without HBC-MA solution was not shown (Fig. 5A
which both hydrophilic and hydrophobic groups coexist, and when
and B).
HBC-MA solution is added to photo-initiator, the HBC-MA solution
shows a co-nonsolvency effect (Wolf & Willms, 1978). We hypothesized
3.3.2. Cell counting kit-8 (CCK-8) assay
that if a co-nonsolvency effect can be exhibited in an aqueous environ­
To further evaluate the effect of HBC-MA and photoinitiator on cell
ment, crosslinking would be possible. Deionized water was prepared to
viability and proliferation, NIH3T3 cells were cultured with media
minimize the effects of ions and HBC-MA solution (20 ◦ C) was carefully
containing different concentrations of HBC-MA solution and subject to
injected into the 37 ◦ C deionized water using a syringe (Fig. S1A, Sup­
CCK-8 assay at 12, 24, 48, and 72 h. To confirm the cytotoxicity of HBC-
porting Information). As expected, the HBC-MA solution was
MA, relative cell viability was compared based on the NIH3T3 control

9
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

Fig. 7. Subcutaneous injection study of HBC-MA hydrogels. (A) Using a syringe pump, HBC-MA solution was injected into mice at 100 μl. (B) Visible light was
irradiated on the back of respiratory anesthesia mice without incision. The light source was at least 3 cm away from the mice. The light intensity is 10 mW/cm2. The
0 s type was excluded from this process. (C) Confirmation of hydrogel formation after HBC-MA injection 1, 3, and 5 d later. (D) Representative photograph showing
samples after 1, 3, and 5 d for each visible light exposure time after injection. (E) Hydrogel samples were measured for each visible exposure time after injection and a
defined day. Data are represented in means ± SD, n = 4. (F) Degradation of injected HBC-MA hydrogel from 1 to 5 d in mice. Four types (0, 10, 20, and 30 s) were
tested according to visible light exposure time. Data are represented in means ± SD, n = 4.

group grown on cell-cultured media. In the 0.1 % and 1% HBC-MA which showed a decrease in cell viability, the absorbance was checked
prepolymer solution groups, except for 0.1 % of 24 h, there was no again to see if proliferation was limited in the cell. Only the 10 % group
statistically significant difference in cell viability and proliferation from was rechecked to see if cell viability was reduced due to limited cell
the control group. However, at 48 and 72 h, the 10 % group showed a proliferation (Fig. 6B). Cell proliferation was occurring at all time
statistically significant decrease of 70.4 ± 0.7 % (*** p < 0.001) and points, including the 48 and 72 h samples, which showed a decrease in
64.2 ± 2.4 % (*** p < 0.001), respectively (Fig. 6A). In the same group, cell viability. As a result, the absorbance (480 nm) showed a statistically

10
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

significant increase (*** p < 0.001). These results occurred when HBC- hydrogels. Despite chemically photo-crosslinking, the hybrid-
MA solution reacted with 37 ◦ C of the incubator to form a tiny hydrogel. crosslinked hydrogel was biodegradable in the presence of lysozyme.
Tiny hydrogels in the 10 % group were clouded over the cell, partially In addition, it was confirmed that degradation optimization according to
interfering with its nutrient transport. For additional reasons, this also temperature and light exposure degree is possible. Slow biodegrad­
affected the lack of physical space for cell proliferation in cell-culture ability at 20 ◦ C showed the storage potential of HBC-MA solution or
dishes (Fig. 6C). HBC-MA was not toxic to NIH3T3 cells but also did hydrogel at low temperature. Cytotoxicity and cell viability test using
not provide a space for cells to grow physically due to a lack of cell NIH3T3 cell confirmed the biocompatibility of HBC-MA conjugate. The
adhesion to NIH3T3 while forming hydrogels. subcutaneous injection of the HBC-MA solution into the mice allowed
the formation of thermogel successfully without incision. Further visible
3.4. Subcutaneous injection study of HBC-MA hydrogels in mice light irradiation from outside the skin greatly increased the stability of
the hydrogel in vivo. The results of the in vivo experiment showed a
Subcutaneous injection study was performed without an incision to similar tendency as the in vitro experiment. It has been demonstrated
identify the potential for application as a minimally invasive injectable that the photo-crosslink density can be adjusted in vivo with visible light
hydrogel and biodegradability of the hydrogel in vivo. For the subcu­ exposure time, and various elements (degradation and mechanical
taneous injection study, 100 μl of HBC-MA solution was injected using a properties) can be optimized accordingly. Taken together, the biocom­
syringe pump (Fig. 7A). As expected, the HBC-MA solution retained a patible and hybrid-crosslinkable HBC-MA can be considered as an
convex shape throughout the experiment at the injection site without injectable hydrogel material with high potential in a variety of appli­
any manipulation due to the viscosity and co-nonsolvency effects of cations, including drug delivery system and tissue engineering.
HBC-MA. There was no change in the convex morphology even after
exposure of visible light (0, 10, 20, and 30 s) at different times for each Author contributions
type in the dorsal skin of mice without incision (Fig. 7B). After a defined
time course, the mice were sacrificed, and the injected hydrogels were Hojae Bae and Jeong Wook Seo designed the experiments; Jeong
carefully separated using surgical instruments (Fig. 7C and D). Photo­ Wook Seo and Seon Young Shin performed the experiments. Hojae Bae
polymerization of the HBC-MA solution was successful and temperature and Jeong Wook Seo arranged the data and wrote the manuscript; Hojae
sensitivity was also confirmed by the 0 s types in mice. All samples Bae, Jeong Wook Seo, Su Ryon Shin, Min Young Lee, Jae Min Cha,
obtained from mice were weighed and compared. After day 1, the Kyung Hyun Min, and Sang Cheon Lee revised the manuscript. All au­
weight of the hydrogel obtained from mice was heavier with longer thors significantly contributed to the study and to the interpretation of
visible light exposure time to visible light. The 1 d group in the 0 s type the data.
with the lowest weight was 31.4 ± 5.2 mg, and the 1 d group in the 30 s
type with the most weight was 13.4 ± 2.3 mg, which showed significant Funding
difference. The tendency for weight to decrease over time also showed a
significant difference (Fig. 7E). As can be seen in previous experiments, This article was supported by a 2016 grant from the Konkuk
increasing the degree of photo-crosslinking tended to limit thermal University.
crosslinking. Conversely, a sample with a low degree of photo-
crosslinking formation initially showed more water release due to Declaration of Competing Interest
thermal crosslinking, thus showing a low hydrogel weight. However, it
was judged that the degree of water content was not the only factor All the authors declare no financial or commercial conflict of interest
causing the change in hydrogel weight. After the hydrogel obtained from that are directly related to the content of this article.
mice was lyophilized to remove all moisture, degradation ratio was
analyzed. Through the in vitro degradation experiment, it was confirmed Acknowledgements
that there was a difference in the tendency of degradation of HBC-MA
hydrogel according to the crosslinking density. Samples from early in None
vivo time points showed faster biodegradability. The 0 s type with
temperature crosslinking only showed high biodegradability of 61.3 ± Appendix A. Supplementary data
16 % in 1 d group. At 5 d group, most hydrogels were degraded during
the experiment, with mass loss of 92.5 ± 4.1 %. On the other hand, the Supplementary material related to this article can be found, in the
30 s type with the longest visible light exposure and additional thermal- online version, at doi:https://doi.org/10.1016/j.carbpol.2020.117036.
crosslinking due to physiological temperature was 25.2 ± 12.3 % at 1
d group, and 36.5 ± 3.3 % at 5 d group. It showed similar results to the References
degradation experiment using lysozyme in vitro. It was confirmed that
there was a difference in degradation of HBC-MA hydrogel according to Azuma, K., Osaki, T., Minami, S., & Okamoto, Y. (2015). Anticancer and anti-
inflammatory properties of chitin and chitosan oligosaccharides. Journal of
the crosslinking density. Unlike in vitro experiments, the degree of Functional Biomaterials, 6(1), 33–49.
degradation with temperature could not be tested. Bae, Y. H., Okano, T., & Kim, S. W. (1989). Insulin permeation through thermo-sensitive
hydrogels. Journal of Controlled Release, 9(3), 271–279.
Bahney, C. S., Lujan, T. J., Hsu, C. W., Bottlang, M., West, J. L., & Johnstone, B. (2011).
4. Conclusion Visible light photoinitiation of mesenchymal stem cell-laden bioresponsive
hydrogels. European Cells & Materials, 22, 43–55.
In this study, we synthesized hydroxybutyl methacrylated chitosan Bin, Z., Changzheng, W., Chunlin, H., Qisheng, G., & Dajun, C. (2010). Preparation and
characterization of hydroxybutyl chitosan. e-Polymers, 10(1), 146.
conjugate and investigated its thermo-sensitive and photo-crosslinkable Chen, J., Park, H., & Park, K. (1999). Synthesis of superporous hydrogels: Hydrogels with
characteristics for injectable hydrogel system. The HBC-MA solution fast swelling and superabsorbent properties. Journal of Biomedical Materials Research,
exhibits co-nonsolvency characteristics at neutral pH at 37 ◦ C, so ther­ 44(1), 53–62.
Chen, Y., Luan, J., Shen, W., Lei, K., Yu, L., & Ding, J. (2016). Injectable and
mal crosslinking can be achieved even under physiological conditions.
thermosensitive hydrogel containing liraglutide as a long-acting antidiabetic system.
The photo-crosslinked HBC-MA hydrogel formed smaller micropore ACS Applied Materials & Interfaces, 8(45), 30703–30713.
sizes than the thermo-crosslinked gel, and hybrid-crosslinked HBC-MA Cho, I. S., Cho, M. O., Li, Z., Nurunnabi, M., Park, S. Y., Kang, S.-W., & Huh, K. M. (2016).
hydrogel significantly improved mechanical strength and structural Synthesis and characterization of a new photo-crosslinkable glycol chitosan
thermogel for biomedical applications. Carbohydrate Polymers, 144, 59–67.
stability under physiological conditions. The improved stability is ex­ Dong, Y., Hassan, W., Zheng, Y., Saeed, A. O., Cao, H., Tai, H., … Wang, W. (2012).
pected to broaden the application range of thermal-crosslinkable Thermoresponsive hyperbranched copolymer with multi acrylate functionality for in

11
J.W. Seo et al. Carbohydrate Polymers 251 (2021) 117036

situ cross-linkable hyaluronic acid composite semi-IPN hydrogel. Journal of Materials monitoring. Proceedings of the National Academy of Sciences United States of America,
Science Materials in Medicine, 23(1), 25–35. 107(42), 17894.
Fernandes Queiroz, M., Melo, K. R., Sabry, D. A., Sassaki, G. L., & Rocha, H. A. (2015). Shigemasa, Y., Matsuura, H., Sashiwa, H., & Saimoto, H. (1996). Evaluation of different
Does the use of chitosan contribute to oxalate kidney stone formation? Marine Drugs, absorbance ratios from infrared spectroscopy for analyzing the degree of
13(1). deacetylation in chitin. International Journal of Biological Macromolecules, 18(3),
Fu, S., Ni, P., Wang, B., Chu, B., Zheng, L., Luo, F., … Qian, Z. (2012). Injectable and 237–242.
thermo-sensitive PEG-PCL-PEG copolymer/collagen/n-HA hydrogel composite for Tsukamoto, Y., Akagi, T., Shima, F., & Akashi, M. (2017). Fabrication of orientation-
guided bone regeneration. Biomaterials, 33(19), 4801–4809. controlled 3D tissues using a layer-by-layer technique and 3D printed a
Kuwajima, T., Yoshida, H., & Hayashi, K. (1976). Graft polymerization of methyl thermoresponsive gel frame. Tissue Engineering Part C, Methods, 23(6), 357–366.
methacrylate onto gelatin. Journal of Applied Polymer Science, 20(4), 967–974. Valmikinathan, C. M., Mukhatyar, V. J., Jain, A., Karumbaiah, L., Dasari, M., &
Lin, S.-B., Lin, Y.-C., & Chen, H.-H. (2009). Low molecular weight chitosan prepared with Bellamkonda, R. V. (2012). Photocrosslinkable chitosan based hydrogels for neural
the aid of cellulase, lysozyme and chitinase: Characterisation and antibacterial tissue engineering. Soft Matter, 8(6), 1964–1976.
activity. Food Chemistry, 116(1), 47–53. Wang, Q., Mynar, J. L., Yoshida, M., Lee, E., Lee, M., Okuro, K., … Aida, T. (2010). High-
Liu, M., Zeng, X., Ma, C., Yi, H., Ali, Z., Mou, X., … He, N. (2017). Injectable hydrogels water-content mouldable hydrogels by mixing clay and a dendritic molecular binder.
for cartilage and bone tissue engineering. Bone Research, 5(1), 17014. Nature, 463(7279), 339–343.
Mukae, K., Bae, Y. H., Okano, T., & Kim, S. W. (1990). A Thermo-sensitive hydrogel: Poly Wang, Q. Q., Kong, M., An, Y., Liu, Y., Li, J. J., Zhou, X., … Chen, X. G. (2013).
(ethylene oxide-dimethyl siloxane-ethylene oxide)/poly(N-isopropyl acrylamide) Hydroxybutyl chitosan thermo-sensitive hydrogel: A potential drug delivery system.
interpenetrating polymer networks II. On-off regulation of solute release from Journal of Materials Science, 48(16), 5614–5623.
thermo-sensitive hydrogel. Polymer Journal, 22(3), 250–265. Wei, Z., Yang, J. H., Liu, Z. Q., Xu, F., Zhou, J. X., Zrínyi, M., … Chen, Y. M. (2015).
Nichol, J. W., Koshy, S. T., Bae, H., Hwang, C. M., Yamanlar, S., & Khademhosseini, A. Novel biocompatible polysaccharide-based self-healing hydrogel. Advanced
(2010). Cell-laden microengineered gelatin methacrylate hydrogels. Biomaterials, 31 Functional Materials, 25(9), 1352–1359.
(21), 5536–5544. Wolf, B. A., & Willms, M. M. (1978). Measured and calculated solubility of polymers in
Noshadi, I., Hong, S., Sullivan, K. E., Shirzaei Sani, E., Portillo-Lara, R., Tamayol, A., … mixed solvents: Co-nonsolvency. Die Makromolekulare Chemie, 179(9), 2265–2277.
Annabi, N. (2017). In vitro and in vivo analysis of visible light crosslinkable gelatin Yang, K. K., Kong, M., Wei, Y. N., Liu, Y., Cheng, X. J., Li, J., … Chen, X. G. (2012).
methacryloyl (GelMA) hydrogels. Biomaterials Science, 5(10), 2093–2105. Folate-modified–chitosan-coated liposomes for tumor-targeted drug delivery.
Pangburn, S. H., Trescony, P. V., & Heller, J. (1982). Lysozyme degradation of partially Journal of Materials Science, 48(4), 1717–1728.
deacetylated chitin, its films and hydrogels. Biomaterials, 3(2), 105–108. Yoon, H. J., Shin, S. R., Cha, J. M., Lee, S.-H., Kim, J.-H., Do, J. T., … Bae, H. (2016).
Peppas, N. A., Hilt, J. Z., Khademhosseini, A., & Langer, R. (2006). Hydrogels in biology Cold water fish gelatin methacryloyl hydrogel for tissue engineering application.
and medicine: From molecular principles to bionanotechnology. Advanced Materials, PloS One, 11(10). e0163902.
18(11), 1345–1360. Zhang, X.-Z., & Chu, C.-C. (2004). Preparation of thermosensitive PNIPAAm hydrogels
Potta, T., Chun, C., & Song, S.-C. (2010). Dual cross-linking systems of functionally with superfast response. Chemical Communications, (3), 350–351.
photo-cross-Linkable and thermoresponsive polyphosphazene hydrogels for Zhang, X.-Z., Xu, X.-D., Cheng, S.-X., & Zhuo, R.-X. (2008a). Strategies to improve the
biomedical applications. Biomacromolecules, 11(7), 1741–1753. response rate of thermosensitive PNIPAAm hydrogels. Soft Matter, 4(3).
Qi, H., Du, Y., Wang, L., Kaji, H., Bae, H., & Khademhosseini, A. (2010). Patterned Zhang, X.-Z., Xu, X.-D., Cheng, S.-X., & Zhuo, R.-X. (2008b). Strategies to improve the
differentiation of individual embryoid bodies in spatially organized 3D hybrid response rate of thermosensitive PNIPAAm hydrogels. Soft Matter, 4(3), 385–391.
microgels. Advanced Materials, 22(46), 5276–5281. Zhu, W., Harris, B. T., & Zhang, L. G. (2016). Gelatin methacrylamide hydrogel with
Rinaudo, M. (2006). Chitin and chitosan: Properties and applications. Progress in Polymer graphene nanoplatelets for neural cell-laden 3D bioprinting. 2016 38th Annual
Science, 31(7), 603–632. International Conference of the IEEE Engineering in Medicine and Biology Society
Shibata, H., Heo, Y. J., Okitsu, T., Matsunaga, Y., Kawanishi, T., & Takeuchi, S. (2010). (EMBC), 4185–4188.
Injectable hydrogel microbeads for fluorescence-based in vivo continuous glucose

12

You might also like