You are on page 1of 11

pubs.acs.

org/acsapm Article

Tough and Biocompatible Hydrogel Tissue Adhesives Entirely Based


on Naturally Derived Ingredients
Ziyi Xu, Haihui Zhang, Yuan Huang, Hao Zhong, Peiwu Qin, Sibo Cheng, Yuenan Wang,
and Canhui Yang*
Cite This: ACS Appl. Polym. Mater. 2024, 6, 1141−1151 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Hydrogel tissue adhesives have tremendous poten-


tial applications in biological engineering. Existing hydrogel tissue
Downloaded via NEWCASTLE UNIV on February 14, 2024 at 18:40:55 (UTC).

adhesives generally do not have adequate mechanical robustness


and acceptable biocompatibility at the same time. Herein, we
report a one-step method to synthesize tough and biocompatible
hydrogel tissue adhesives entirely made of naturally derived
ingredients. We select two natural polymers, chitosan and gelatin,
to construct the backbone and a bioderived compound, genipin, as
the cross-linker. We show that, upon gelation, genipins cross-link
chitosan and gelatin to form two interpenetrated networks and
interlink them to tissue surfaces. Meanwhile, hydrogen bonds form
in the matrix to strengthen the networks and at the interface to
strengthen the adhesion between the hydrogel and tissue. Furthermore, we elaborately use high initial polymer contents to induce
topological entanglements in the polymer networks to toughen the hydrogel. The resulting chitosan−gelatin hydrogel provides a
tough matrix, and the robust covalent interlinks and hydrogen bonds provide a strong interface, achieving a tensile strength of ∼190
kPa, a fracture toughness of 205.7 J/m2, a mode I adhesion energy of 197.6 J/m2, and a mode II adhesion energy of 51.2 J/m2. We
demonstrate that the hydrogel tissue adhesive is injectable, degradable, and noncytotoxic and can be used for the controlled release
of the anticancer drug cisplatin. All-natural ingredient-based tough and biocompatible hydrogels are promising as tissue adhesives for
biomedical and related applications.
KEYWORDS: hydrogel, tissue adhesive, natural polymer, toughness, biocompatibility

■ INTRODUCTION
Adhesives that can adhere strongly to tissues have tremendous
typically on the order of 10 J/m2,16,17 which greatly limits the
applications. Highly desired is to develop new polymer-based
potential applications in biological engineering as diverse as tissue adhesive materials that are soft but can adhere strongly
to tissues.18
wound dressing,1 tissue repair,2 drug delivery,3 biomedical
A hydrogel is a three-dimensional polymer network
devices,4 and bioelectronic interfaces.5 Among the many types
infiltrated with a large amount of water. Hydrogels can be
of adhesives, polymer-based tissue adhesives are appealing
categorized as natural, synthetic, and hybrid hydrogels
alternatives to sutures in various types of wound closure, such
depending on the sources of the polymers.19 Tough hydrogel
as those in the skin,1 bone,6 vessels,7 heart,8 and organ
adhesion has two prerequisites: a strong interface and tough
transplantation,9 owing to their better hemostatic effect, less
matrix. A strong interface prevents interfacial crack prop-
equipment required, faster operation speed, and noninvasive
agation and enables efficient load transfer to elicit dissipation
nature.10,11
in the bulk, and a tough matrix dissipates energy to toughen
Over the past few decades, polymeric tissue adhesives have
the adhesion. Following this basic mechanical principle,
been extensively studied. In general, an ideal tissue adhesive is
significant progress has been made recently in forming tough
expected to possess the attributes of excellent biocompatibility,
adhesion between hydrogels and tissues using various tough
ease of use, biodegradability, and robust adhesion.12,13
Cyanoacrylate has long been the strongest commercially
available tissue adhesive, but it is cytotoxic and rigid and Received: August 19, 2023
generates enormous heat during its immediate polymerization Revised: December 8, 2023
upon exposure to water. Other commercial tissue adhesives Accepted: December 8, 2023
such as the fibrin glue14 and the polyethylene glycol−based Published: December 26, 2023
adhesives15 are soft and biocompatible, but the interfacial
adhesion is weak and the toughness of the matrix is low,

© 2023 American Chemical Society https://doi.org/10.1021/acsapm.3c01926


1141 ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 1. Tough and biocompatible hydrogel tissue adhesive with injectability and degradability. The precursor contains naturally derived
ingredients of chitosan, gelatin, and genipin and is shear-thinning to be injectable. After injection, genipin cross-links chitosan and gelatin to form
two interpenetrated networks and interlinks them to the amino group-rich tissue surface. Meanwhile, hydrogen bonds form to strengthen the
networks and the interface between the hydrogel and tissue. In addition, topological entanglements form to further toughen the networks. Upon
separation, the robust interfacial adhesion elicits enormous dissipation in the hydrogel matrix, resulting in a tough adhesion. After degradation, the
hydrogel disintegrates and the tough adhesion vanishes. The chemical structures of the ingredients, the cross-link, the hydrogen bonds, and the
reactions between genipins and amino groups are illustrated.

hydrogels.20−27 However, most of the tough hydrogels consist to tissues (Table S1).32 Moreover, the preparation process can
of synthetic polymers, e.g., polyacrylamide, that are prepared be cumbersome, so the raw natural polymers need to be
from cytotoxic raw materials, e.g., acrylamide,28,29 making modified with functional groups such as dopamine, vinyl
biocompatibility an indispensable concern. groups, adipic acid dihydrazide, or aldehyde moieties. There-
In terms of biocompatibility, natural polymers are more ideal fore, developing tough hydrogel tissue adhesives entirely based
materials of choice for tissue adhesives and have been widely on naturally derived ingredients via facile and biocompatible
used to synthesize hydrogels.30 Nevertheless, existing natural processes is significant but remains a central challenge.
polymer-based hydrogel tissue adhesives are often mechan- Herein, we report a simple one-step method to synthesize
ically weak and cannot form a tough adhesion with tissues. tough hydrogel tissue adhesives entirely based on naturally
Besides, the synthesis usually involves other cytotoxic derived ingredients. Note that we aim to propose a general
components as initiators and cross-linkers, e.g., glutaraldehyde synthetic strategy instead of new chemistries. We use the well-
and carbodiimide,31 and the synthetic procedures such as known natural polymers chitosan and gelatin to construct the
prolonged ultraviolate illumination and heating are not friendly polymer backbone and the naturally derived compound
1142 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 2. Mechanical characterizations. (a) Photographs showing a dumbbell-shaped hydrogel (φ70) being stretched to a strain of 2. (b) Nominal
uniaxial tensile stress−strain curves of hydrogels with different water contents. (c) Tensile moduli and works of fracture of various hydrogels. (d)
Photographs showing a cylindrical hydrogel (φ60, diameter: 1 cm, height: 5 mm) being compressed and springing back. (e) Nominal compressive
stress−strain curves of hydrogels with different water contents. (f) Compressive moduli and works of fracture of various hydrogels. (g) Photographs
showing the unnotched/notched hydrogels (φ70) at different deformation states. (h) Pure shear stress−strain curves of unnotched (solid curves)
and notched (dashed curves) hydrogels. (i) Fracture energies of various hydrogels.

genipin as the cross-linker. Chitosan, with marked biocompat- 205.7 J/m2, a mode I adhesion energy of 197.6 J/m2, and a
ibility, biodegradability, nontoxicity, and cell binding ability,33 mode II adhesion energy of 51.2 J/m2. Moreover, we show that
has been widely used in biomedical engineering.34−36 Gelatin the hydrogel tissue adhesive is injectable, biodegradable, and
produces nontoxic and biocompatible products after degrada- noncytotoxic and can be used for the controlled release of the
tion.37 Genipin is extracted from plants and exhibits excellent anticancer drug cisplatin.
biocompatibility38 and has been adopted as the cross-linking
agent for pharmaceutical applications.39 Although chitosan,
gelatin, and genipin have been used as constituent materials for
hydrogels, previously obtained hydrogels are mechanically
■ RESULTS AND DISCUSSION
Principle of the Tough and Biocompatible Hydrogel
weak and cannot afford tough tissue adhesives (Table S2). Tissue Adhesive. The principles of the tough and
It has been noted recently that the mechanical properties of biocompatible hydrogel tissue adhesive are illustrated in Figure
hydrogels are profoundly affected by the synthesis con- 1. The precursor contains long-chain natural polymers chitosan
ditions.40−42 In this work, we rationalize the synthetic and gelatin and naturally derived cross-linker genipin. Chitosan
procedures such that, upon gelation, genipin covalently is a natural cationic polysaccharide consisting of D-glucos-
cross-links chitosan and gelatin to form two interpenetrated amine and N-acetyl-D-glucosamine units.43 Gelatin is a
networks and interlinks the two networks to tissues. The mixture of peptide hydrolyzed from collagen.44 Genipin is an
resulting chitosan−gelatin hydrogel provides a tough matrix, aglycone derivative obtained from the enzymatic hydrolysis of
and the robust interlinks provide a strong interface. We use geniposide with β-glucosidase.45,46 In general, the hydrogel
high initial polymer contents such that topological entangle- precursor is viscous and non-Newtonian, featuring shear-
ments are formed during the formation of the network to thinning, i.e. the viscosity reduces under shear strain. The
further toughen the hydrogel. The chitosan−gelatin hydrogel injectability is particularly advantageous in the scenarios of
achieves a tensile strength of ∼190 kPa, a fracture toughness of tissue adhesives that the precursor can well conform to the
1143 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 3. Tissue adhesion characterizations. (a) Schematics of the 90° peel test. (b) Force/width versus displacement curves for different
hydrogels. (c) Mode I adhesion energies of different hydrogels. (d) Lap-shear test of a layer of hydrogel sandwiched between two layers of porcine
skin. (e) Nominal stress−strain curves of different hydrogels under lap-shear. (f) Shear strengths and mode II adhesion energies of various
hydrogels under lap-shear. Photographs showing the robust adhesion of the hydrogel with (g) pig heart, (h) liver, and (i) bone.

surface asperities of tissue for tight sealing. Besides, delivering Biomedical materials have excellent degradability, and it is
the hydrogel tissue adhesive through injection is much less necessary for them to be absorbed by the surrounding tissues
invasive compared to surgical operations. during applications to avoid surgical resection. Both chitosan
During the curing process, genipin cross-links chitosan and and gelatin are biodegradable. For example, the lysozyme can
gelatin to form a tough interpenetrated network. The cross- degrade chitosan into low molecular weight oligomers by
linking mechanism has three folds: two genipin molecules can cutting the β-1,4 glycosidic bonds between the D-glucosamine
cross-link two chitosan chains, two gelatin chains, or one and D-acetylglucosamine units under weakly acidic con-
chitosan chain and one gelatin chain. The underlying ditions.49 As another example, the peptide chains of gelatin
chemistry is that genipins react with the primary amino groups will be broken into low molecular weight amino acids or
to form secondary amides and heterocyclic amines.47 In polypeptides under the action of a collagenase degradation
addition to covalent cross-links, abundant hydrogen bonds accelerator.50 After degradation, the hydrogel disintegrates, and
the tough adhesion vanishes spontaneously without any
form between chitosan chains, gelatin chains, and chitosan and
cytotoxic degradation products.
gelatin chains to strengthen the matrix.
Mechanical Properties of Chitosan−Gelatin Hydro-
Since the tissue surface is generally rich in primary amino
gels. Robust mechanical properties are crucial for hydrogel
groups, during the curing process, genipin molecules interlink tissue adhesives to resist premature cohesive rupture. We
chitosan and gelatin chains to the tissue surface by forming perform uniaxial tension, compression, and pure shear tests to
covalent bonds.2 Besides, hydrogen bonds also form between assess the mechanical properties. A dumbbell-shaped sample
chitosan and gelatin chains and the tissue surface to further can withstand a tensile strain of 200% (Figure 2a). The
strengthen the interface. Upon separation, the robust synthesis conditions have prominent effects on the mechanical
interfacial adhesion enables efficient stress transmission from properties of the hydrogels synthesized from monomers.42 We
the interface to the bulk network, eliciting enormous synthesized three types of chitosan−gelatin hydrogels with
dissipation in the hydrogel matrix, such as the break of different water contents for the precursor, i.e., 60, 70, and 80%.
hydrogen bonds and the distributed chain scission caused by The hydrogel precursor becomes too viscous to manipulate
network imperfection,48 to register a tough adhesion. when the water content is below 60%. The uniaxial tensile
1144 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 4. Rheological characterizations. (a) Photographs showing the sol−gel transition of precursors in a bottle (top row) and a star-shaped mold
(bottom row). (b) Storage modulus (G’) and loss modulus (G”) vary with time. (c) Viscosity varies with time. (d) Gelation time (T1) and
saturation time (T2) for different hydrogels.

stress−strain curves show that the hydrogel with 60% water owing to the toughening effects of topological entangle-
content possesses the highest mechanical strength (Figure 2b), ments,40 the fracture energy of the hydrogel with 60% water,
whereas the rupture strains are mostly the same, ∼220%. The 205.73 J/m2, is higher than that of the hydrogel with 80%
strengths are 192.16, 53.96, and 29.93 kPa, the tensile moduli water, 12.42 J/m2, by 1 order of magnitude.
are 74.02, 20.19, and 9.91 kPa, and the works of fracture are Adhesion Performances. The adhesion between the
166.27, 61.74, and 15.49 kJ/m3 for the water contents of 60, chitosan−gelatin hydrogel and tissue has not been quantified
70, and 80%, respectively (Figure 2c). before. We perform the 90° peel test to measure the mode I
The strength decreases with the water content since the adhesion energy between the hydrogel and porcine skin
density of polymer chains decreases while the modulus does (Figure 3a). The peel force versus displacement curves of the
not. For a hydrogel consist of Gaussian chains, the modulus is hydrogels with different water contents are plotted in Figure
negatively proportional to water content φ 3b. The mode I adhesion energies are 197.61 67.31, and 39.11
N/m (Figure 3c) for the hydrogels with 60, 70, and 80% water
E (1 )NkT (1)
contents, respectively.
where N is the number of polymer chains per unit volume, k is We further perform the lap-shear test to measure the mode
the Boltzmann constant, and T is the absolute temperature. II adhesion energy by sandwiching a layer of chitosan−gelatin
Taking the water contents of 60 and 80% as examples, the hydrogel between two layers of porcine skin, which are in turn
theoretical ratio of Eφ60/Eφ80 is 2, which is much lower than the
d d
glued to two PMMA plates for chasing (Figure 3d). The
experimental value, 7.47. The giant deviation between theory hydrogels with 70% water content exhibit the highest shear
and experiment manifests the effects of synthesis conditions strength (Figure 3e). The samples with 60% water content
and is affected by the topological entanglements, which undergo adhesive failure, and only half contain hydrogel
function as additional cross-links to increase the effective residues. In comparison, the samples with 70 and 80% water
value of N, during the formation of the hydrogel network.40 contents undergo cohesive failure. The hydrogels of φ60 having
Figure 2d shows that a cylindrical sample can sustain the highest modulus deform the least, while the hydrogels of
compression and return to its original state. Under the φ80 having the lowest strength and toughness dissipate the least
compression test, similarly, the hydrogel with 60% water energy. Overall, the hydrogels of φ70 have optimal perform-
content exhibits the highest compressive strength, while the ances. Specifically, the strengths are 19.51, 26.25, and 13.75
rupture strains of all hydrogels are about 80% (Figure 2e). The kPa, and the mode II adhesion energies are 41.66, 51.19, and
compression moduli are 217.01 79.55, and 31.84 kPa, and the 26.73 J/m2 for the hydrogels with 60, 70, and 80% water
works of fracture are 22.72, 14.51, and 5.76 kJ/m3 for the water contents, respectively (Figure 3f).
contents of 60, 70, and 80%, respectively (Figure 2f). Figure 2g We demonstrate the appreciable adhesion between the
shows the images of the pure-shear samples, with/without a chitosan−gelatin hydrogel and pig heart (Figure 3g) and liver
notch, at different strains, and Figure 2h plots the stress−strain (Figure 3h) by laminating and then peeling the hydrogel layer
curves of various unnotched (solid curves)/notched (dashed from the tissues. It can be seen that both tissues have been
curves) hydrogels. Figure 2i shows the histogram of the substantially deformed, indicating good adhesion. Using the
average fracture energies of the different hydrogels. Again, hydrogel as an injectable arthrosis glue (Figure 3i), we show
1145 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 5. Degradation characterizations. (a) Sequential images showing the degradation processes of different hydrogels in PBS buffer solution and
enzyme solution. (b) Percentages of degradation vary with time for different hydrogels.

that the hydrogel tissue adhesive can withstand the gravity of needs to be cured appreciably for reliable mechanical
100 g of weight. robustness before functioning. In our experiments, we define
Injectability. Injectability is often desired for tissue the time when the viscosity reaches 80% of the plateau
adhesives for the sake of high moldability, minimal invasion, viscosity as the saturation time T2. The gelling times are 117,
and easy and effective drug/cell encapsulation. The chitosan− 871, and 1174 s and the saturation times so defined are 6593,
gelatin hydrogels exhibit a highly tunable injectability. A light- 7089, and 7338 s for different water contents (Figure 4d). In
yellow chitosan−gelatin hydrogel precursor can be cured inside practice, the gelling time should be tuned according to the
a bottle or be injected into a start-shaped mold and then cured specific application scenarios while the saturation time
(Figure 4a). We conduct rheological measurements to quantify generally should be as short as possible. Note that the cross-
the injectability of the chitosan−gelatin hydrogel. The viscosity linking reaction of genipin with the amino groups is benign and
of the chitosan−gelatin solution decreases with shear rate, a thus could proceed in vivo.
feature of shear-thinning liquid (Figure S1). Figure 4b plots Degradability. As mentioned before, in situ degradation is
the variations of the storage modulus (G’) and loss modulus advantageous for hydrogel tissue adhesives to eliminate the
(G″) with time for hydrogel precursors with different water need of surgical removal, which often causes secondary trauma.
contents. When the G” is greater than the G’, the hydrogel Degradation mechanisms include hydrolysis, erosion, and
precursor behaves liquid-like; when the G’ is greater than the solubilization. In particular, chitosan is the only natural
G”, the hydrogel precursor behaves solid-like. The intersection cationic polymer that can be degraded by human enzymes
between G’ and G″ determines the gelling point. The time at due to its reactive amino groups, and gelatin can be easily
the gelling point is defined as the gelling time, denoted as T1. degraded by collagenase. We carry out degradation character-
The final storage moduli are 76, 23, and 12 kPa, respectively, izations by soaking the chitosan−gelatin hydrogels with
for 60, 70, and 80% water contents. Since the hydrogel different water contents in a PBS buffer solution and enzyme
precursor contains a cross-linker, the viscosity of the precursor solution. The sequential images in Figure 5a show the
will gradually increase as the cross-linking reaction proceeds. morphological changes of different samples at different times.
Figure 4c plots the variations of viscosity with time for the Figure 5b plots the variation of the degradability with time.
different hydrogels. As expected, throughout the test, viscosity For the samples in PBS solution, before functioning, the
increases monotonically with time and the viscosity of the active chemicals need to diffuse into the hydrogel network,
hydrogel with 60% water content is the highest. which is associated with diffusion of water molecules. The
The injection of the hydrogel precursor should be operated imbibition of water leads to the swelling of the hydrogel.
within T1 since the precursor may lose injectability after the Overall, the hydrogels with an initial water content of 70%
gelling point. To guarantee a smooth injection process requires have a balanced chemical potential of water and network
not only optimization of the rheological properties of the elasticity and exhibit the smallest degradation values. The
hydrogel precursor but also the parameters of the injection degradation values of the samples soaked in enzyme solutions
instruments such as the injection speed and the length and are all smaller than those of the samples soaked in PBS
radius of the needle, as well as the environment such as solutions since the enzyme can destroy the chemical bonds of
temperature and pH value. On the other hand, the hydrogel the backbones of chitosan and gelatin.
1146 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

Figure 6. Swelling and drug release characterizations and in vitro cytotoxicity test. (a) Swelling kinetics of different hydrogels in deionized water.
The solid curves are fittings. (b) Schematics of the swelling of different hydrogels. (c) Comparison of storage modulus (G’) and loss modulus (G″)
before and after swelling. (d) Release kinetics of cisplatin of different hydrogels in PBS solution. The solid curves are fittings. The inset shows the
chemical structure of cisplatin. (e) Saturated amount of released drug versus the saturated time for different hydrogels. (f) Schematic illustrating the
diffusion of a cisplatin molecule (green circle) within the hydrogel network of mesh size ξ under the confinement of hydrogen bonds. (g) Live cell
fluorescence imaging (calcein-positive) of HEK293T cells after the treatment with different concentrations of hydrogel extract solution. (h)
Cytotoxicity test of different hydrogel extracts after culturing for 24, 48, and 72 h.

Dynamics of Swelling and Drug Release. Hydrogels are be L2/D ∼ 105 s, which is in the same order of magnitude with
ideal depots of therapeutic agents. Loading drugs into the the experiments.
hydrogel matrix endows a hydrogel tissue adhesive with Experimental results show that the higher the initial water
additional functionalities such as accelerated wound healing content, the higher the saturated swelling ratio, which is
and antibacterial properties. Upon deployment, the hydrogel consistent with the literature.42 A higher initial polymer
tissue adhesives often undergo swelling, which enhances the content (i.e., a lower initial water content) will lead to more
release of loaded drugs. Figure 6a plots the variations of the topological entanglements between polymer chains, which act
swelling ratios of different hydrogels with time. The swelling as additional cross-links to strengthen the polymer network to
ratio is defined as Δm/m, where Δm is the change of mass and resist swelling (Figure 6b). As expected, both the storage
m is the original mass. Assuming a free diffusion process, we modulus (G’) and loss modulus (G”) decrease after swelling
can fit the experimental data to the following formula: (Figure 6c). Nevertheless, due to the topological entangle-
(x / ) ments, the equilibrium storage modulus decreases with the
y = (1 e ) (2)
increase in the initial water content. The storage moduli after
where x represents the swelling time, y represents Δm/m, α is a swelling equilibrium are 829.8, 360.4, and 172.9 Pa for 60, 70,
dimensionless fitting parameter, and β is a fitting parameter and 80% water contents, respectively.
with a unit of time. Specifically, β represents the characteristic A piece of therapeutic agent-loaded hydrogel tissue adhesive
swelling time. The characteristic times are 38.5, 37.1, and 18.4 is a delivery platform for local administration. We loaded the
h for 60, 70, and 80% water contents, respectively. For the chitosan−gelatin hydrogel with cisplatin and characterized the
samples of size L ∼ 4 mm and the diffusion coefficient of water drug release. Cisplatin is a chemotherapy medication widely
D ∼ 10−10 m2/s, we estimate the characteristic swelling time to used in clinics to treat a number of cancers. The molecular
1147 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

structure of cisplatin is the square planar coordination complex functional groups, introduction of nanocrystalline domains, or
cis-[Pt(NH3) 2Cl 2 ]. Because the release of cisplatin is constructing networks with enormous entanglements.41 A
accompanied by swelling, it can be seen from Figure 6d that mechanical property-biodegradability conflict exists in which
the release kinetics of cisplatin is similar to the swelling the mechanical properties will deteriorate when the hydrogel
kinetics. The highest water content (φ80) with the highest network undergoes biodegradation. Nevertheless, the conflict
swelling ratio gives the fastest release rate and the highest can be circumvented by tailoring the function times. For
saturated drug release ratio. Fitting the drug release data to the instance, the characteristic time needed for appreciable
above exponential formula, with x representing the drug release biodegradation can be rationally tuned so that the hydrogel
time and y representing the drug release, gives the character- can maintain its mechanical robustness within a certain time
istic times of 15.4, 12.9, and 9.0 h for 60, 70, and 80% water window. The proposed strategy, in principle, applies to other
contents, respectively. biopolymers rich in amino groups. For example, carboxymethyl
The saturated drug release ratios are 47.1, 38.7, and 28.2 for chitosan can be used and the obtained genipin cross-linked
80, 70, and 60% water contents, respectively (Figure 6e). The carboxymethyl chitosan−gelatin hydrogel exhibits decent
amino groups of cisplatin form hydrogen bonds with the mechanical properties and good adhesion to the liver (Figure
functional groups on the chitosan−gelatin network. When the S2).
initial water content is high, the mesh size of the hydrogel
network is large and the confinement of cisplatin is small due
to the larger distance. The hydrogel with a higher initial water
■ CONCLUSIONS
In conclusion, we have reported tough and biocompatible
content has faster swelling kinetics and a larger saturated
hydrogel tissue adhesives completely based on naturally
swelling ratio, which further facilitates the release of cisplatin
derived ingredients. The hydrogel contains a tough matrix
(Figure 6f). The controlled drug release and the good adhesion
consisting of two interpenetrating networks of chitosan and
of the tough and biocompatible chitosan−gelatin hydrogel to
gelatin cross-linked by genipin, which also covalently interlinks
native tissue are expected to be potentially used for the
the hydrogel matrix to tissue surfaces. Topological entangle-
targeted elimination of tumors. Besides, the chitosan−gelatin
ments are formed to toughen the chitosan−gelatin hydrogel
hydrogel can be loaded with therapeutic agents on demand for
matrix and thus the adhesion by increasing the initial polymer
specific applications.
contents. We have shown that the chitosan−gelatin hydrogel is
An ideal biomaterial should be nontoxic. We investigated the
injectable and biodegradable with tailorable injection and
cytotoxicity of the chitosan−gelatin hydrogels by evaluating
biodegradation dynamics. Moreover, we have demonstrated
the in vitro cell viability of the extracted media leached from
the drug loading and release performances as well as excellent
the chitosan−gelatin hydrogels with different water contents.
biocompatibility of the chitosan−gelatin hydrogel. This work
We used fluorescence imaging and lactate dehydrogenase
puts a step toward tough and biocompatible hydrogel tissue
(LDH) assays and observed the living cells through confocal adhesives entirely based on naturally derived ingredients for
laser scanning microscopy (CLSM). The CLSM images injectable drug-loaded hydrogel materials for biomedicine and
demonstrate that cell viability is not significantly affected engineering.
after culturing with different concentrations of extracted
hydrogel media compared to the control group (Figure 6g).
We further analyze the cell viability by LDH release when
treating cells with different concentrations of hydrogel extracts
■ EXPERIMENTAL SECTION
Materials. We procured chitosan (CS, H1928001) and acetic acid
for 24, 48, and 72 h. Cell viabilities of all hydrogel precursor- (G2121128) from Aladdin, gelatin (Gel, CLP813) from Shanghai
treated media are higher than 80% (Figure 6h), which are Biad Pharmaceutical Technology Co., Ltd., and genipin (E100770)
comparable to that of the untreated controls. Moreover, we from Anergy Co., Ltd. All chemicals were obtained without further
purification.
conduct a one-way variance analysis using IBM SPSS Statistics Preparation of Hydrogels. First, gelatin, chitosan, and 1 mL of
and find no significant difference (P > 0.05) in cell viability deionized water were mixed in a 4 mL centrifuge tube at 65 °C to
between the samples with different concentrations of hydrogel dissolve gelatin. Then, a genipin solution was added at a
extracts and the control group. These results confirm that the concentration of 0.035 g/mL. Finally, 20 μL of glacial acetic acid
chitosan−gelatin hydrogel possesses excellent biocompatibility. was added to dissolve chitosan and catalyze the chemical reaction
The mechanical properties of the chitosan−gelatin hydrogel between genipin and the amino groups on the polymer chains. The
should be decent to meet many biomedical applications, pH value of the precursor was ∼7. After curing, the once light-yellow
although they are still relatively weak compared to those of liquid-state precursor became a dark-green solid-state hydrogel. The
masses of gelatin Ma1 = 0.83 g, Ma2 = 0.42 g, and Ma3 = 0.27 g,
tough synthetic tough hydrogels. The main reason is that the
chitosan Mb1 = 0.067 g, Mb2 = 0.034 g, and Mb3 = 0.022 g, and genipin
hydrogel network of our hydrogel is amorphous, entirely made solution Vc1 = 25.6 μL, Vc2= 12.8 μL, and Vc3 = 8.5 μL were used to
of natural polymers, and covalently cross-linked without any prepare the hydrogels with water contents of 60, 70, and 80%,
further treatment, for the sake of biocompatibility and ease of respectively. The water content was quantified by weight percentage.
synthesis. The bulky side chains and the strong intra/ Rheological Tests. Rheological measurements were conducted to
interchain interactions restrict the flexibility of the polymer evaluate the gelling processes of the precursors and the viscoelastic
chains such that the configuration of the polymer chain is more properties of the cured hydrogels by using a rheometer (TA
difficult to change to adapt to the applied force. Besides, the Instruments HR30) equipped with a Peltier plate for temperature
natural polymer-based network is formed by preformed long- control. A parallel steel geometry (diameter: 20 mm) was used for all
experiments. Oscillatory measurements were performed at a constant
chain polymers so that the network topology is much less frequency of 1 Hz and a small strain of 0.1%. The precursor was
flexible. Subject to an applied force, the natural polymer-based added to the loading platform of the rheometer, and then the gelation
network more easily concentrates stress and causes the nearby process was monitored by measuring the evolutions of storage and
chains to break. The mechanical properties of natural polymer- loss moduli (G’ and G”, respectively) as a function of time at 37 °C.
based hydrogels could be improved via, e.g., modification of To prevent the hydrogel from dehydration during the test, a

1148 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

humidifier was used to pump humid moisture during the test. Each c
II =t ( )d
test was repeated at least three times. 0 (6)
Mechanical Tests. For compression, a cylindrical acrylate mold
with dimensions of 10 mm in diameter and 5 mm in height was where t is the thickness of the sandwiched hydrogel layer at the
prepared by using a laser cutter (CMA0604-B-A), and the precursor undeformed state, γc is the nominal rupture shear strain, and τ is the
was poured into the mold to prepare cylindrical samples. The samples nominal shear stress.
were loaded to a mechanical testing machine (Instron 5966) with a Swelling Test. The initial masses of the hydrogel samples with
100 N load cell and compressed at a loading velocity of 30 mm/min different water contents were measured. Then, the samples were
at 25 °C. The force−displacement curves were recorded to calculate soaked in phosphate buffered saline (PBS) buffer solutions. After
the stress−strain curves. Unless otherwise specified, nominal stress some time, the weights of the hydrogels were measured again. The
and nominal strain were used throughout the text. selected time intervals were 0.5, 1, 2, 3, 6, 9, 24, 30, 48, 56, 68, 72, 84,
For uniaxial tension, a large 2 mm-thick sheet was prepared,
96, 120, 150, 180, 200, and 250 h. Each measurement was repeated
incubated for 30 min at 65 °C, and aged for 20 min at 25 °C. Then,
three times.
dumbbell-shaped samples with a gauge length of 12 mm and a width
Degradation Test. Cubic samples (1 × 1 × 1 cm3) were
of 2 mm were cut from the sheet. The samples were loaded to a
mechanical testing machine (Instron 5966) with a 100 N load cell and immersed in ethanol for 24 h to remove the unreacted contents.
elongated at a loading velocity of 30 mm/min at 25 °C. The force− Then, the samples were freeze-dried for 48 h and weighed.
displacement curves were recorded to calculate the stress−strain Subsequently, the samples were immersed in pure PBS solution
curves. The work of fracture is defined as the energy per unit volume (control group) and PBS solution containing lysozyme (1 mg/mL)
needed to break the hydrogel and calculated as the area underneath and collagenase (0.16 mg/mL) (test group).51,52 The samples were
the stress−strain curve cultured at 37 °C, and the solutions were renewed every 2 days.
Finally, the samples were taken out at different times, washed with
c
Wf = ( )d ethanol for 3 h, and freeze-dried for 24 h. The percentage of weight
0 (3) loss was calculated using the following formula:
where εc is the rupture strain. Mf
To measure the fracture toughness, two sets of samples (width × remaining weight = × 100%
Mi (7)
length × thickness: 1 × 7 × 0.2 cm3) of identical dimensions were
prepared, with one being precut with a single-edge notch along the
middle line of the sample and the other being intact. For the notched where “Mi” is the initial weight of the freeze-dried hydrogel before
samples, we made a precrack of 20 mm using a razor blade. After the degradation and “Mf” is the weight of the freeze-dried hydrogel after
sample was loaded to the testing machine, the effective width (along degradation. Each test was repeated at least five times.
the loading direction) of the sample is 10 mm. The size of the In Vitro Drug Release. Cisplatin was selected as the model drug
precrack was much larger than the width while much smaller than the in this study. To prepare the drug-loaded hydrogels, 4 μL of cisplatin
length of the sample. The unnotched samples are stretched along the solution (2.5 mg/mL) was added to the hydrogel precursors. The
direction of length (may or may not up to rupture), and their nominal concentration of cisplatin in the resulting hydrogels is 10 μg/mL.
stress−strain curves were measured. The notched samples were Hydrogel samples (0.15 g) were immersed in 50 mL of PBS solution
stretched until the notch started to propagate catastrophically at a (pH 6.8) at 37 °C. Then, 5 mL of the solution was taken out at
critical strain εc. The fracture toughness of the hydrogel is calculated different times, namely, 0.5, 1, 2, 3, 6,12, 24, 30, 36, 48, 60, 72, 84, and
by 96 h; meanwhile, 5 mL of fresh PBS was added. The abstracted
solution was added with 500 μL of concentrated nitric acid, sealed,
c
=H ( )d and placed in an oven at 65 °C overnight. Finally, the platinum (Pt)
0 (4) concentration in the sample was detected by inductively coupled
plasma-mass spectrometry (ICP-MS) to obtain the drug release curve.
where H is the width of the sample along the loading direction at the
undeformed state and σ is the nominal stress. In Vitro Cytotoxicity Test. The cytotoxicity detection kit
The 90° peel test was conducted to measure the mode I adhesion (Roche) was used to measure the in vitro cytotoxicity of hydrogels
energy. A piece of chitosan−gelatin hydrogel sample (width × length according to the supplier’s protocol. The principle of measurement is
× thickness: 2 × 10 × 0.1 cm3) was in situ synthesized on pigskin, based on lactate dehydrogenase (LDH) release from damaged cells to
which was fixed on an acrylic plate. A sheet of nonwoven fabric (width media. The cell viability was calculated by the percentage of treated
× length: 2.5 × 20 cm2) was used as the soft but inextensible backing. cells with respect to untreated cells. The hydrogel (5 g) was first
The sample was loaded onto an Instron 5966 machine with a 100 N washed with PBS (pH 7.4) three times. Next, the hydrogel was
load cell and pulled at a loading speed of 30 mm/min. The peel incubated with Dulbecco’s Modified Eagle Medium (Gibco) with
force−displacement curves were recorded to calculate the mode I 10% fetal bovine serum, 100 U/mL penicillin, and 0.1 mg/mL
adhesion energy. As the peel displacement increases, the peel force streptomycin (Thermo Fisher Scientific) for 30 min at 37 °C. To
increases and then plateaus at a steady state value. At the steady state, obtain the hydrogel extract stock solution, different concentrations of
the mode I adhesion energy is calculated as hydrogel were immersed with 20 mL of complete growth media for 24
h at 37 °C and then passed through a 0.22 μm filter. NIH-3T3 cells
Fss
I = (ATCC) were plated in a 96-well plate (Corning) at a density of 5000
W (5) cells per well and incubated for 24 h at 37 °C and 5% CO2 in a
humidified incubator. Cells were treated with 100 μL of different
where Fss is the steady-state peel force and W is the width of the
sample. concentrations of hydrogel extract solutions and cultured for 24, 48,
The lap-shear test was conducted to measure the mode II adhesion and 72 h. Each hydrogel extract (50 μL) was transferred to a new 96-
energy. A layer of 0.5 mm-thick chitosan−gelatin hydrogel was well plate and mixed with 50 μL of reaction mixture for 30 min.
prepared in situ between two pieces of porcine skin. Two acrylic plates Absorbance was measured at 490 nm by using a plate reader (Tecan
were glued to the porcine skin pieces and then clamped to the Instron Microplate Reader Spark). For fluorescence imaging, the cells were
5966 machine with a 100 N load cell and pulled at a loading speed of transfected with TagBFP fluorescent proteins and cultured with
30 mm/min. The lap-shear force−displacement curves were recorded different concentrations of hydrogel extract solutions (0, 60, and 80%)
to calculate the mode II adhesion energy. The mode II adhesion for 48 h and then imaged by a confocal microscope (Nikon A1R SI
energy is calculated as Confocal).

1149 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

■ ASSOCIATED CONTENT
* Supporting Information

hen Municipality (ZDSYS20210623092005017), the National
Natural Science Foundation of China (no. 12302212), the
The Supporting Information is available free of charge at Natural Science Foundation of Guangdong Province
https://pubs.acs.org/doi/10.1021/acsapm.3c01926. (2022A1515010601), and the Stable Support Plan Program
of Shenzhen Natural Science Fund Grant
Variations of viscosity with shear rate, performances of
(20200925174603001). The authors acknowledge the assis-
the carboxymethyl chitosan−gelatin hydrogel, prepara-
tion and properties of natural polymer-based hydrogels tance of SUSTech Core Research Facilities.
and hydrogels based on chitosan, gelatin, and genipin,
fitting parameters for swelling and drug release (PDF) ■ REFERENCES

■ AUTHOR INFORMATION
Corresponding Author
(1) Li, J.; Celiz, A. D.; Yang, J.; Yang, Q.; Wamala, I.; Whyte, W.;
Seo, B. R.; Vasilyev, N. V.; Vlassak, J. J.; Suo, Z.; Mooney, D. J. Tough
adhesives for diverse wet surfaces. Science 2017, 357 (6349), 378−
381.
Canhui Yang − Shenzhen Key Laboratory of Soft Mechanics (2) Yuk, H.; Varela, C. E.; Nabzdyk, C. S.; Mao, X.; Padera, R. F.;
and Smart Manufacturing, Department of Mechanics and Roche, E. T.; Zhao, X. Dry double-sided tape for adhesion of wet
Aerospace Engineering, Southern University of Science and tissues and devices. Nature 2019, 575 (7781), 169−174.
Technology, Shenzhen 518055, China; orcid.org/0000- (3) Li, J.; Mooney, D. J. Designing hydrogels for controlled drug
0001-5674-834X; Email: yangch@sustech.edu.cn delivery. Nat. Rev. Mater. 2016, 1 (12), 16071.
(4) Feiner, R.; Engel, L.; Fleischer, S.; Malki, M.; Gal, I.; Shapira, A.;
Authors Shacham-Diamand, Y.; Dvir, T. Engineered hybrid cardiac patches
Ziyi Xu − Shenzhen Key Laboratory of Soft Mechanics and with multifunctional electronics for online monitoring and regulation
Smart Manufacturing, Department of Mechanics and of tissue function. Nat. Mater. 2016, 15 (6), 679−685.
Aerospace Engineering, Southern University of Science and (5) Yuk, H.; Wu, J. J.; Zhao, X. H. Hydrogel interfaces for merging
Technology, Shenzhen 518055, China humans and machines. Nat. Rev. Mater. 2022, 7 (12), 935−952.
Haihui Zhang − Institute of Biopharmaceutical and Health (6) Hoffmann, B.; Volkmer, E.; Kokott, A.; Augat, P.; Ohnmacht,
Engineering, Tsinghua Shenzhen International Graduate M.; Sedlmayr, N.; Schieker, M.; Claes, L.; Mutschler, W.; Ziegler, G.
School, Tsinghua University, Shenzhen 518055, China Characterisation of a new bioadhesive system based on poly-
Yuan Huang − Shenzhen Key Laboratory of Soft Mechanics saccharides with the potential to be used as bone glue. Journal of
and Smart Manufacturing, Department of Mechanics and Materials Science-Materials in Medicine 2009, 20 (10), 2001−2009.
(7) Giraudo, M. V.; Di Francesco, D.; Catoira, M. C.; Cotella, D.;
Aerospace Engineering, Southern University of Science and
Fusaro, L.; Boccafoschi, F. Angiogenic Potential in Biological
Technology, Shenzhen 518055, China Hydrogels. Biomedicines 2020, 8 (10), 436.
Hao Zhong − Shenzhen Key Laboratory of Soft Mechanics (8) Hong, Y.; Zhou, F.; Hua, Y.; Zhang, X.; Ni, C.; Pan, D.; Zhang,
and Smart Manufacturing, Department of Mechanics and Y.; Jiang, D.; Yang, L.; Lin, Q.; Zou, Y.; Yu, D.; Arnot, D. E.; Zou, X.;
Aerospace Engineering, Southern University of Science and Zhu, L.; Zhang, S.; Ouyang, H. A strongly adhesive hemostatic
Technology, Shenzhen 518055, China hydrogel for the repair of arterial and heart bleeds. Nat. Commun.
Peiwu Qin − Institute of Biopharmaceutical and Health 2019, 10, 2060.
Engineering, Tsinghua Shenzhen International Graduate (9) Liu, K.; Yang, H.; Huang, G. B.; Shi, A. H.; Lu, Q.; Wang, S. P.;
School, Tsinghua University, Shenzhen 518055, China; Qiao, W.; Wang, H. H.; Ke, M. Y.; Ding, H. F.; Li, T.; Zhang, Y. C.;
orcid.org/0000-0002-7829-8973 Yu, J. W.; Ren, B. Y.; Wang, R. F.; Wang, K. L.; Feng, H.; Suo, Z. G.;
Sibo Cheng − Suzhou Soft Intelligent Materials Co. Ltd., Tang, J. D.; Lv, Y. Adhesive anastomosis for organ transplantation.
Suzhou 215123, China Bioact. Mater. 2022, 13, 260−268.
(10) Ghobril, C.; Grinstaff, M. W. The chemistry and engineering of
Yuenan Wang − Department of Therapeutic Radiology, Yale polymeric hydrogel adhesives for wound closure: a tutorial. Chem. Soc.
University, New Haven, Connecticut 06511, United States Rev. 2015, 44 (7), 1820−1835.
Complete contact information is available at: (11) Ma, Y.; Yao, J.; Liu, Q.; Han, T.; Zhao, J.; Ma, X.; Tong, Y.; Jin,
https://pubs.acs.org/10.1021/acsapm.3c01926 G.; Qu, K.; Li, B.; Xu, F. Liquid Bandage Harvests Robust Adhesive,
Hemostatic, and Antibacterial Performances as a First-Aid Tissue
Author Contributions Adhesive. Adv. Funct. Mater. 2020, 30 (39), 2001820.
C.Y. and Z.X. conceived the research ideas and designed the (12) Ahmed, E. M. Hydrogel: Preparation, characterization, and
research. Z.X. designed, synthesized, and characterized the applications: A review. J. Adv. Res. 2015, 6 (2), 105−121.
(13) Zhou, L.; Dai, C.; Fan, L.; Jiang, Y.; Liu, C.; Zhou, Z.; Guan, P.;
mechanical properties and drug release performances. H.Z. Tian, Y.; Xing, J.; Li, X.; Luo, Y.; Yu, P.; Ning, C.; Tan, G. Injectable
conducted the cytotoxicity tests. Y.H. synthesized and Self-Healing Natural Biopolymer-Based Hydrogel Adhesive with
characterized the carboxymethyl chitosan-based hydrogels. Thermoresponsive Reversible Adhesion for Minimally Invasive
Z.X. analyzed the experimental results with inputs from other Surgery. Adv. Funct. Mater. 2021, 31 (14), 2007457.
authors. Z.X. drafted the manuscript. C.Y. revised the (14) Sierra, D. H. Fibrin Sealant Adhesive Systems: A Review of
manuscript with inputs from other authors. C.Y. supervised Their Chemistry, Material Properties and Clinical Applications. J.
the study. Biomater. Appl. 1993, 7 (4), 309.
Notes (15) Wallace, D. G.; Cruise, G. M.; Rhee, W. M.; Schroeder, J. A.;
The authors declare no competing financial interest. Prior, J. J.; Ju, J.; Maroney, M.; Duronio, J.; Ngo, M. H.; Estridge, T.;
Coker, G. C. A tissue sealant based on reactive multifunctional

■ ACKNOWLEDGMENTS
This work is supported by the National Key Research and
polyethylene glycol. J. Biomed. Mater. Res. 2001, 58 (5), 545−555.
(16) Thistlethwaite, P. A.; Luketich, J. D.; Ferson, P. F.; Keenan, R.
J.; Jamieson, S. W. Ablation of persistent air leaks after thoracic
Development Program of China (grant 2023YFB3812500), the procedures with fibrin sealant. Annals of Thoracic Surgery 1999, 67
Science, Technology, and Innovation Commission of Shenz- (2), 575−577.

1150 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151
ACS Applied Polymer Materials pubs.acs.org/acsapm Article

(17) Dastjerdi, A. K.; Pagano, M.; Kaartinen, M. T.; McKee, M. D.; (36) Pok, S.; Vitale, F.; Eichmann, S. L.; Benavides, O. M.; Pasquali,
Barthelat, F. Cohesive behavior of soft biological ″glues″. Experiments M.; Jacot, J. G. Biocompatible Carbon Nanotube-Chitosan Scaffold
and modeling. 2012, 8 (9), 3349−3359. Matching the Electrical Conductivity of the Heart. ACS Nano 2014, 8
(18) Giri, T. K.; Thakur, D.; Alexander, A.; Ajazuddin; Badwaik, H.; (10), 9822−9832.
Tripathi, D. K. Alginate based Hydrogel as a Potential Biopolymeric (37) Dong, Y.; Sigen, A.; Rodrigues, M.; Li, X.; Kwon, S. H.; Kosaric,
Carrier for Drug Delivery and Cell Delivery Systems: Present Status N.; Khong, S.; Gao, Y.; Wang, W.; Gurtner, G. C. Injectable and
and Applications. Curr. Drug Delivery 2012, 9 (6), 539−555. Tunable Gelatin Hydrogels Enhance Stem Cell Retention and
(19) Buwalda, S. J.; Boere, K. W. M.; Dijkstra, P. J.; Feijen, J.; Improve Cutaneous Wound Healing. Adv. Funct. Mater. 2017, 27
Vermonden, T.; Hennink, W. E. Hydrogels in a historical perspective: (24), 1606619.
From simple networks to smart materials. J. Controlled Release 2014, (38) Yu, Y. B.; Xu, S.; Li, S. M.; Pan, H. Genipin-cross-linked
190, 254−273. hydrogels based on biomaterials for drug delivery: a review. Biomater.
(20) Yang, J.; Bai, R.; Chen, B.; Suo, Z. Hydrogel Adhesion: A Sci. 2021, 9 (5), 1583−1597.
Supramolecular Synergy of Chemistry, Topology, and Mechanics. (39) Chang, W. H.; Chang, Y.; Chen, Y. C.; Sung, H. W.
Adv. Funct. Mater. 2020, 30 (2), 1901693. Hemoglobin polymerized with a naturally occurring crosslinking
(21) Yang, C. H.; Suo, Z. G. Hydrogel ionotronics. Nat. Rev. Mater. agent as a blood substitute: In vitro and in vivo studies. Artificial Cells
2018, 3 (6), 125−142. Blood Substitutes and Biotechnology 2004, 32 (2), 243−262.
(22) Yuk, H.; Zhang, T.; Lin, S. T.; Parada, G. A.; Zhao, X. H. (40) Kim, J.; Zhang, G. G.; Shi, M. X.; Suo, Z. G. Fracture, fatigue,
Tough bonding of hydrogels to diverse non-porous surfaces. Nat. and friction of polymers in which entanglements greatly outnumber
Mater. 2016, 15 (2), 190−196. cross-links. Science 2021, 374 (6564), 212−216.
(23) Gao, Y.; Han, X.; Chen, J.; Pan, Y.; Yang, M.; Lu, L.; Yang, J.; (41) Nian, G.; Kim, J.; Bao, X.; Suo, Z. Making Highly Elastic and
Tough Hydrogels from Doughs. Adv. Mater. 2022, 34 (50), 2206577.
Suo, Z.; Lu, T. Hydrogel-mesh composite for wound closure. Proc.
(42) Sun, X. J.; Rao, P.; He, X. T.; Yang, C. H.; Hong, W.
Natl. Acad. Sci. U.S.A. 2021, 118 (28), No. e2103457118.
Chemically identical gels I-under-crosslinked networks. J. Mech. Phys.
(24) Freedman, B. R.; Kuttler, A.; Beckmann, N.; Nam, S.; Kent, D.;
Solids 2023, 175, No. 105278.
Schuleit, M.; Ramazani, F.; Accart, N.; Rock, A.; Li, J. Y.; Kurz, M.;
(43) Suh, J. K. F.; Matthew, H. W. T. Application of chitosan-based
Fisch, A.; Ullrich, T.; Hast, M. W.; Tinguely, Y.; Weber, E.; Mooney, polysaccharide biomaterials in cartilage tissue engineering: a review.
D. J. Enhanced tendon healing by a tough hydrogel with an adhesive Biomaterials 2000, 21 (24), 2589−2598.
side and high drug-loading capacity. Nat. Biomed. Eng. 2022, 6 (10), (44) Wang, S.; Li, K.; Zhou, Q. High strength and low swelling
1167−1179. composite hydrogels from gelatin and delignified wood. Sci. Rep.
(25) Liu, J.; Qu, S.; Suo, Z.; Yang, W. Functional hydrogel coatings. 2020, 10 (1), 17842.
Natl. Sci. Rev. 2021, 8 (2), nwaa254. (45) Paik, Y. S.; Lee, C. M.; Cho, M. H.; Hahn, T. R. Physical
(26) Zhang, D.; Yang, F.; He, J.; Xu, L.; Wang, T.; Feng, Z.-Q.; stability of the blue pigments formed from geniposide of gardenia
Chang, Y.; Gong, X.; Zhang, G.; Zheng, J. Multiple physical bonds to fruits: Effects of pH, temperature, and light. J. Agric. Food. Chem.
realize highly tough and self-adhesive double-network hydrogels. ACS 2001, 49 (1), 430−432.
Applied Polymer Materials 2020, 2 (3), 1031−1042. (46) Butler, M. F.; Ng, Y. F.; Pudney, P. D. A. Mechanism and
(27) Zhang, Y.; Ren, B.; Xie, S.; Cai, Y.; Wang, T.; Feng, Z.; Tang, J.; kinetics of the crosslinking reaction between biopolymers containing
Chen, Q.; Xu, J.; Xu, L. Multiple physical cross-linker strategy to primary amine groups and genipin. J. Polym. Sci., Part A: Polym. Chem.
achieve mechanically tough and reversible properties of double- 2003, 41 (24), 3941−3953.
network hydrogels in bulk and on surfaces. ACS Appl. Polym. Mater. (47) Cui, L.; Jia, J. F.; Guo, Y.; Liu, Y.; Zhu, P. Preparation and
2019, 1 (4), 701−713. characterization of IPN hydrogels composed of chitosan and gelatin
(28) Chen, Q.; Zhu, L.; Zhao, C.; Wang, Q.; Zheng, J. A robust, one- cross-linked by genipin. Carbohydr. Polym. 2014, 99, 31−38.
pot synthesis of highly mechanical and recoverable double network (48) Yang, C. H.; Yin, T. H.; Suo, Z. G. Polyacrylamide hydrogels. I.
hydrogels using thermoreversible sol-gel polysaccharide. Advanced Network imperfection. J. Mech. Phys. Solids 2019, 131, 43−55.
materials 2013, 25 (30), 4171−4176. (49) Nordtveit, R. J.; Varum, K. M.; Smidsrod, O. Degradation of
(29) Chen, Q.; Chen, H.; Zhu, L.; Zheng, J. Fundamentals of double partially N-acetylated chitosans with hen egg white and human
network hydrogels. J. Mater. Chem. B 2015, 3 (18), 3654−3676. lysozyme. Carbohydr. Polym. 1996, 29 (2), 163−167.
(30) Yuan, L.; Wu, Y.; Fang, J.; Wei, X. J.; Gu, Q. S.; El-Hamshary, (50) Pang, J. H.; Wischke, C.; Lendlein, A. In vitro Degradation
H.; Al-Deyab, S. S.; Morsi, Y.; Mo, X. M. Modified alginate and Analysis of 3D-architectured Gelatin-based Hydrogels. MRS Adv.
gelatin cross-linked hydrogels for soft tissue adhesive. Artificial Cells 2020, 5 (12−13), 633−642.
Nanomedicine and Biotechnology 2017, 45 (1), 76−83. (51) Tanuma, H.; Kiuchi, H.; Kai, W. H.; Yazawa, K.; Inoue, Y.
(31) Zhu, W.; Li, Y. L.; Liu, L. X.; Chen, Y. M.; Wang, C.; Xi, F. Characterization and Enzymatic Degradation of PEG-Cross-Linked
Supramolecular Hydrogels from Cisplatin-Loaded Block Copolymer Chitosan Hydrogel Films. J. Appl. Polym. Sci. 2009, 114 (3), 1902−
Nanoparticles and alpha-Cyclodextrins with a Stepwise Delivery 1907.
Property. Biomacromolecules 2010, 11 (11), 3086−3092. (52) Gorgieva, S.; Kokol, V. Preparation, characterization, and in
(32) Josa-Cullere, L.; Llebaria, A. In the Search for Photocages vitro enzymatic degradation of chitosan-gelatine hydrogel scaffolds as
Cleavable with Visible Light: An Overview of Recent Advances and potential biomaterials. J. Biomed. Mater. Res. Part A 2012, 100A (7),
Chemical Strategies. Chemphotochem 2021, 5 (4), 298−316. 1655−1667.
(33) Tang, W.; Wang, J.; Hou, H. W.; Li, Y.; Wang, J.; Fu, J. A.; Lu,
L.; Gao, D. D.; Liu, Z. M.; Zhao, F. Y.; Gao, X. Q.; Ling, P. X.; Wang,
F. S.; Sun, F.; Tan, H. N. Review: Application of chitosan and its
derivatives in medical materials. Int. J. Biol. Macromol. 2023, 240,
No. 124398.
(34) Huang, W.; Cheng, S.; Wang, X.; Zhang, Y.; Chen, L.; Zhang,
L. Noncompressible Hemostasis and Bone Regeneration Induced by
an Absorbable Bioadhesive Self-Healing Hydrogel. Adv. Funct. Mater.
2021, 31 (22), 2009189.
(35) Feng, Y.; Gao, H. L.; Wu, D.; Weng, Y. T.; Wang, Z. Y.; Yu, S.
H.; Wang, Z. Biomimetic Lamellar Chitosan Scaffold for Soft Gingival
Tissue Regeneration. Adv. Funct. Mater. 2021, 31 (43), 2105348.

1151 https://doi.org/10.1021/acsapm.3c01926
ACS Appl. Polym. Mater. 2024, 6, 1141−1151

You might also like