You are on page 1of 15

Article

Cite This: J. Agric. Food Chem. 2019, 67, 3966−3980 pubs.acs.org/JAFC

A Dual Cross-Linked Strategy to Construct Moldable Hydrogels with


High Stretchability, Good Self-Recovery, and Self-Healing Capability
Yang Qin,†,‡,§ Jinpeng Wang,†,‡,§ Chao Qiu,†,‡,§ Xueming Xu,†,‡,§ and Zhengyu Jin*,†,‡,§

State Key Laboratory of Food Science and Technology,‡ School of Food Science and Technology, and §Synergetic Innovation
Center of Food Safety and Nutrition, Jiangnan University, Wuxi, Jiangsu 214122, People’s Republic of China
*
S Supporting Information

ABSTRACT: Most conventional synthetic hydrogels suffer from poor mechanical properties; despite recent significant
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

progress in fabricating tough hydrogels, it is still a challenge to simultaneously realize high stretchability, self-recovery, and self-
healing capability in a hydrogel. In this work, a new type of starch/PVA/borax hybrid dual cross-linked (DC) hydrogel was
Downloaded via UNIV DE CONCEPCION UDEC on June 7, 2022 at 17:37:25 (UTC).

synthesized by a one-pot method. The as-prepared DC hydrogels exhibited mechanical properties of remarkable extensibility
(ca. 2485%), excellent toughness (ca. 290.5 kJ m−3), high compression strength (ca. 547.8 kPa), rapid recoverability (81.9%
energy recovery after 30 min), and free-shapeable behavior. More impressively, the DC gels sustained approximately 300 times
their own weight and exhibited an outstanding self-healing capability at room temperature both in air and underwater.
Furthermore, the adsorption amount of methylene blue onto the anionic DC gel (144.68 mg/g) was much higher than that of
corn starch gel. Consequently, the eco-friendly, stable, and biodegradable hydrogels will have a great potential application in
removing anionic dyes from the wastewater produced by agriculture and industry.
KEYWORDS: dual cross-linked network, starch matrix, hydrogels, free-shapeable behavior, swelling-resistant, methylene blue

■ INTRODUCTION
One of the most promising soft and hydrophilic materials,
composed of two different polymer networks with asymmetric
structures, i.e., a rigid and brittle first network, which provides
hydrogels have a three-dimensional cross-linked network a skeleton structure to the hydrogel, and a soft and stretchable
structure composed of polymer chains and adjustable physical second network, which provides elasticity.26,27 Owing to the
and chemical properties.1,2 These smart materials have unique dual-network structure, DN hydrogels have exhibited
received considerable attention from academia and industry desirable mechanical properties and energy-dissipating per-
due to their unique applications in a diverse range of fields, formance. DN hydrogels are considered impressive materials
such as tissue-engineered scaffolds,3−5 coating materials,6 water because they demonstrate excellent mechanical properties.12,19
treatment,7,8 drug delivery,9−11 and biosensors.12,13 Due to the Although these DN hydrogels demonstrate improved proper-
intrinsic structural heterogeneity and/or lack of an effective ties over their corresponding single-network (SN) parent
energy dissipation mechanism, most conventional hydrogels hydrogels, the mechanical properties do not improve as
suffer from poor mechanical properties,14−16 which severely significantly as expected.
limits their application in industrial and commercial fields. Recently, dual cross-linked (DC) network hydrogels, as a
High mechanical strength, stretchability, and self-healing special DN gel, have attracted widespread interest from
capacity are highly desirable but are inversely related properties academia as they optimize and combine different properties
of gel materials. Superior mechanical strength and stretch- of each network structure, including dual physically cross-
ability avert breakage of the materials, while the self-healing linked hydrogels,2,16 dual chemically cross-linked hydrogels,28
capacity enables the materials to spontaneously recover their and hybrid cross-linked hydrogels.29 However, chemically
functionality and mesh structure after being damaged.1 The cross-linked points lack the ability to efficiently recover from
presence of diverse mechanical properties in hydrogel systems damage caused by stretching and compressing, resulting in a
has recently become a hotspot for the research focused on the loss of mechanical properties due to the permanent breaking of
field.11,17,18 However, most hydrogels possess only one of these chemical bonds.30,31 In contrast, physically cross-linked points
properties and very few possess all properties. Therefore, it is were reversible, which allowed hydrogels to recover or self-heal
still a challenging task to combine high stretchability and after a large deformation or disruption, i.e., the first network
strength with good self-healing capacity in a single material.19 unzips and dissipates energy upon deformation, and thus, the
Great efforts have been made to develop versatile hydrogels physical bonds can reform after removal of the stress.
to solve this predicament by introducing new network Afterward, many efforts have been made to endow a self-
structures and/or energy dissipation strategies to strengthen healing capacity and multifunctional properties to hydrogels.
the mechanical properties of hydrogels,2,11 such as double-
network (DN) hydrogels, nanocomposite hydrogels, ionically Received: September 20, 2018
cross-linked hydrogels, hydrogen-bonded hydrogels,20−24 etc. Revised: March 4, 2019
First proposed by Gong et al.25 in 2003, DN hydrogels are a Accepted: March 19, 2019
promising way to achieve this goal. The DN hydrogel is Published: March 19, 2019

© 2019 American Chemical Society 3966 DOI: 10.1021/acs.jafc.8b05147


J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

The research works have focused on the construction of DC Scheme 1. Schematically Fabrication of the Method and
network hydrogels by dynamic and reversible physical bonds Formation Mechanism of CS/PVA/B Double Cross-Linked
(e.g., hydrogen bond, hydrophobic interaction, π−π stacking, (DC) Hydrogel (a) and the Single and Double Cross-
or ionic bond).14−16,24,32 However, there are a number of self- Linked Network (SC and DC) Formed by Several Specific
healing hydrogels that are frequently plagued with problems Interactions among Corn Starch (CS), PVA, and Borax (B):
such as long self-healing time and stimulus dependence in Hydrogen Bond, Borate Bond (Reversible Ionic Crosslink),
application.6,9,11,14,16 Moreover, it is still worth evaluating the and Chain Entanglement (b)
bioactivity of the synthetic hydrogels owing to a number of
vinyl monomers as the building block for the DC network.
Recently, rising needs in the biochemical, medical, and
functional food fields spurred considerable interest in
developing an efficient strategy for the preparation of nontoxic,
environment-friendly, biocompatible, and biodegradable multi-
functional hydrogels as replacements for synthetic materials.
To address these concerns, as a scaffold or functional matrix,
biomacromolecules have been used to construct bioactive DC
hydrogels (e.g., gelatin,17 cellulose,33 chitosan,34 carrageenan,31
and protein2,35). Due to its great cross-linking ability in the
presence of abundant hydroxyl groups, starch is known to be a
good precursor material for preparing hydrogels.36 However,
the starch-based hydrogels exhibit poor mechanical properties
and brittleness, which thus hampers the development of
unmodified starch as a stable hydrogel.36,37 Due to the −OH
groups in each of the repeating molecular units, poly(vinyl
alcohol) (PVA) is proven to simultaneously have good gel- and
film-forming capability.38 Thus, the introduction of PVA
improves the mechanical properties and stability of starch-
based gels.39,40 Furthermore, studies have demonstrated that
introducing cross-linking agents (epichlorohydrin) markedly Within the DC gels, the borate ions further assured the
strengthens the mechanical and thermal properties of starch/ interpenetration, tangling, and cross-linking of corn starch
PVA composite films through cross-linking the chains of starch (CS) and/or PVA chains by covalent bonds (Scheme 1b).
and PVA.41 Borax (B), as a reversible cross-linker, dissolves Consequently, the hydrogen bonds and borate ions jointly
and dissociates into borate ions (B(OH)4−) in water. The PVA promoted the strong interaction between CS chains and PVA
chains and borate ions can easily form a single cross-linked chains, making the interconnection of the CS and PVA
(SC) network structure. On the basis of the above strategy, networks denser. Among the networks, CS networks provide a
Sreedhar et al. investigated the effect of B concentrations on firm platform to bear stress, and the dynamic PVA/B networks
the thermal, mechanical, and surface characterizations of the endow the hydrogels with a self-healing capacity. As a result,
soluble starch and PVA blend films.42 Moreover, boron the DC hydrogels achieved high stretchability and toughness as
complexes of PVA and starch blend films were also prepared well as self-recovery and self-healing abilities at the same time.
by Bursali et al., who investigated the effect of the cross-linking Additionally, because of the presence of numerous borate ions
agent (glutaraldehyde) on the antimicrobial activity of the (B(OH)4−) in DC gel, the anionic DC gel exhibited an
composite films.43 These studies had prepared PVA matrix efficient adsorption of MB from the aqueous environment
composite film as the starting point and mainly investigated the through electrostatic attraction. This work provides a new
effect of the addition of starch and the cross-linking agent on design strategy to prepare starch-based gels with highly
the physicochemical properties of the films, rather than mechanical, stable, self-healing, and high adsorption properties;
investigating the mechanical behaviors and the interaction consequently, we hope the renewable, biocompatible, and
mechanisms. Starch, as a poly hydroxyl macromolecule, is biodegradable DC gel will have a promising application in the
similar to PVA. Inspired by this, we wondered if we could fields of agriculture, biomedicine, and pollutant adsorbents.


simultaneously introduce the PVA chains and borate ions to
starch-based gels for constructing a novel DC hydrogel with MATERIALS AND METHODS
excellent mechanical properties, self-recovery, and self-healing Materials. Corn starch (CS, 26.9% amylose content) was
abilities at room temperature. Moreover, it is necessary to purchased from Ingredient China Ltd. (Guangdong, China). Poly-
understand whether cross-linked interactions occurred among (vinyl alcohol) (PVA, DS = 1750 ± 50, degree of deacetylation,
the starch and PVA chains and the borate ions. To the best of 99.0%; Mw = 75000−80000 g/mol) and sodium tetraborate
our knowledge, the starch-based DC hydrogels with high decahydrate (Borax, over 99.5% purity; Mw = 381.37 g/mol) were
stretchability, mechanical strength, and self-healing properties supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai,
have never been reported in the literature previously. China). Methylene blue (MB) was supplied by Aladdin and was used
In this paper, we report a new type of hybrid (physical and as an adsorbate. All reagents were of reagent-grade.
Synthesis of CS/PVA/B DC Hydrogel. CS/PVA/B hydrogels
chemical hybrid) DC starch-based hydrogel through a one-pot with corn starch, PVA, and B were prepared via one-pot tandem
method, i.e., the retrogradation technology combined with reactions as shown in Scheme 1. PVA powders were dispersed in
freezing−thawing technology (Scheme 1). The as-prepared deionized water and stirred at 95 °C for 30 min to ensure that the
DC hydrogels consist of a hydrogen-bond-associated starch PVA powders dissolved enough. The PVA solution and the
network and a dynamic borate bond cross-linked network. gelatinized starch solution were well mixed in a PVA solution with

3967 DOI: 10.1021/acs.jafc.8b05147


J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

continuous magnetic stirring (300 rpm/min) to form a mixture of hydrogel dosage (25−250 mg), initial MB concentration (50−250
homogeneous solutions of 10 wt % (starch) and 4 wt % (PVA) in mg/L), and temperature (25, 37, 50, and 55 °C) have been explored
concentration, respectively. Afterward, the mixed solution was heated to evaluate the performance of sorbents. Then, the adsorption
at 95 °C in a water bath with magnetic stirring for 1 h. Subsequently, capacity was calculated using the following equation:
B powders were slowly added to the CS/PVA mixed solution with the
concentration of 0.026 mol/L, which were stirred vigorously (300 qt = (C0 − Ct )V /m
rpm/min for 2 h) until there was a formation of homogeneous sol. In
order to avoid the evaporation of the water, the heating and stirring where qt (mg/g) is the adsorbed amount after time t, C0 and Ct (mg/
process was conducted in a hermetically sealed beaker. Moreover, the L) are the initial concentration and remaining concentration after
sol underwent ultrasonic treatment for 5 min to remove air bubbles. time t, respectively. V (mL) is the volume of MB solution, and m
Finally, the resultant uniform sol was poured into a cylindrical plastic (mg) is the weight of gels.
mold (45 mL) and first stored at 4 °C for 12 h to form the gel Mechanical Property Measurement. Compressive Tests. The
network of starch. Afterward, to further enhance the network of the compressive properties of the hydrogels were analyzed using a texture
hydrogels, the samples were frozen at −18 °C for 4 h and unfrozen at analyzer (TA-XTplus, Stable Micro Systems, Surrey, U.K.). Both
25 °C for 8 h, and the freezing−thawing process was performed one uniaxial and cyclic compressive tests were all conducted on a
time. In addition, all of the other hydrogels were prepared according cylindrical sample (25 mm in diameter and 16 mm in height).
to the conditions processed for the CS/PVA/B hydrogels. The Afterward, it was placed on the lower plate of the instrument and
control sample, starch, PVA, and starch hydrogels with PVA or B were compressed with a P36R probe at a strain rate of 0.2 mm s−1 at room
synthesized by using the same experiment process. The eventual temperature (25 °C) in air. The compressive modulus was calculated
samples were named CS, PVA, CS/B, PVA/B, and CS/PVA. All by fitting the initial slope of the stress−strain curves.44 Compression
specimens with various compositions were listed in Table 1. strain and stress data were recorded during the experiments.
Hysteresis Tests. The cylindrical hydrogel specimens were first
compressed by a loading cycle to a maximum strain of 70% with a
Table 1. Hydrogels with Various Compositions and the speed of 0.2 mm s−1 and then unloaded at the same rate. As for the
Corresponding Water Contentsa recovery experiments, the gel samples were relaxed at 25 °C for a
certain waiting time (0 min, 30 min, and 5 h) before the next loading
starch PVA borax water content
sample (wt %) (wt %) (mol/L) (%) process, during which time the specimens were coated with a silicon
oil surface and then encased in a plastic wrap to avoid and minimize
CS 10.0 0.0 0.0 89.14 ± 2.01
water evaporation. The energy dissipation was calculated from the
PVA 0.0 4.0 0.0 94.99 ± 1.39 area between the loading−unloading curves.45 The residual strain
CS/B 10.0 0.0 0.026 86.79 ± 0.95 ratio was calculated by the ratio of height change along the direction
PVA/B 0.0 4.0 0.026 93.35 ± 1.32 of stress after removal of stress with respect to the original height of
CS/PVA 10.0 4.0 0.0 83.97 ± 1.41 the specimen.46
CS/PVA/B 10.0 4.0 0.026 79.46 ± 0.79 Tensile Tests. The hydrogels were cut into a dumbbell shape with a
a
Values represent the mean ± standard deviation of triplicate tests. gauge length of 20 mm, a width of 10 mm, and a thickness of 5 mm.
Subsequently, the samples were elongated to breakage at a rate of 100
mm min−1. The tensile strain was determined as the elongation
Characterization. FTIR spectra of the hydrogels were obtained
divided by the initial length. The tensile stress was defined as the load
on a Nicolet IS10 spectrometer (Thermo Nicolet, Inc., Waltham, MA,
force divided by the original specimen cross-sectional area. The elastic
U.S.A.) equipped with an attenuated total reflection (ATR) accessory
modulus of the gels was calculated by fitting the initial linear regime of
over the wavenumber range of 4000−400 cm−1. XRD diffraction
the stress−strain curve.47 For the self-healing experiments, the strip
results for the hydrogels were obtained using an XRD diffractometer
hydrogels were cut into halves, and then the two parts were placed in
(D2 PHASER, Bruker AXS, Inc., Germany) equipped with Cu Kα
contact with each other at room temperature in air. Moreover, there
radiation in the range of 4−40° at a scan rate of 0.02°/min. All
was no other outside stimulus applied during the healing process.
samples were freeze-dried prior to analysis. The morphology of the
After self-healing, the tensile test was carried out again to calculate the
specimens was taken with a scanning electron microscope (SEM,
healing efficiency. In addition, the hardness of the hydrogel was
7500F, JEOL, Japan) at an accelerating voltage of 5 kV. The hydrogels
analyzed using a texture analyzer fitted with a P 0.5 probe. The test
were frozen in liquid nitrogen, snapped immediately, and further dried
speed was 0.5 mm s−1, and the hydrogels were compressed twice to a
for 24 h in vacuum. Afterward, the cross sections of the specimens
strain of 70%.


were coated with gold vapor, observed, and photographed. Water
content and degree of swelling of the hydrogel were determined by
the gravimetric method. The weight of the cylindrical hydrogel (25 RESULTS AND DISCUSSION
mm × 16 mm) was determined (W1). Then, the samples were dried Fabrication Process and Formation Mechanism.
at −80 °C for 48 h to obtain the dry mass (Wd). The water content is Scheme 1 schematically indicates the fabrication process and
defined as (W1 − Wd)/W1. The gels were continuously retrieved from formation mechanism of the CS/PVA/B DC hydrogel. When
the deionized water at room temperature. The swelling ratio of the
hydrogel was measured gravimetrically as a function of time (0−96 h) the gelatinized starch solution was well mixed in a PVA
after excess water was removed on the surface of the gels with dry solution completely, PVA chains transferred into the interstitial
paper. The swelling ratio of the hydrogels can be calculated as (Wt − water phases of the starch gels. Afterward, the resulting pastes
Wd)/Wd, where Wt is the weight of the swollen gel. The zeta potential were stored at 4 °C, and the amylose and/or amylopectin
of the swollen DC gel solution was determined by a dynamic light chains in starch cross-linked via numerous hydrogen bonds and
scattering instrument (Zetasizer Nano ZS, Malvern, U.K.) under formed a strong physical cross-linking framework in the
different pH conditions adjusted by using NaOH and HCl solutions. gel.48,49 Neighboring PVA chains formed crystallites through
MB Adsorption Test. The adsorption of MB from aqueous the freezing−thawing method,50 and the crystallites served as
solution onto gel adsorbent was systematically evaluated. A certain linked points for the PVA chains to fabricate another soft
amount of gel adsorbent was added into 100 mL of MB solution (150
mg L−1), and the mixture was placed in an isothermal water bath network. Borate ions further assured the interpenetration,
shaker (25 °C) at 250 rpm. After a period of time, the supernatant tangling, and cross-linking of CS and/or PVA chains by
was centrifuged and the concentration was measured with a UV−vis covalent bonds and hydrogen bonds (Scheme 1b). Due to the
spectrophotometer (UV-1800, Shimadzu Corporation, Japan) at 663 high affinity toward hydroxyl groups, the borate ions formed
nm as λmax of MB. Effects of contact time (0−48 h), pH level (2−10), complexes rapidly and reversibly in aqueous media with the
3968 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 1. (a) FTIR spectra and (b) X-ray diffraction patterns of CS, CS/B, PVA, PVA/B, CS/PVA, and CS/PVA/B hydrogels. SEM images
exhibiting the microstructure of (c) CS, (d) CS/B, (e) PVA, (f) PVA/B, (g) CS/PVA, and (h) CS/PVA/B hydrogels.

poly hydroxyl polymers,51,52 and then the borate ions acted as were enhanced. Therefore, the broad band from 3100 to 3700
an ionic cross-linker between the CS and/or PVA chains. cm−1 may be assigned to the −OH groups of starch and PVA,
Moreover, hydrogen bonds and borate ions jointly promoted as well as the complexes that formed among starch and/or
the strong interactions between CS and PVA chains, and then, PVA with borate ions. A similar result was reported by Manna
the two networks were connected more closely. The two and Patil (2009).56 After the addition of B molecules within
networks become entangled and sustain each other via the hydrogels, they displayed several distinct peaks character-
hydrogen bonds and ionic interactions, resulting in the strong istic of B and borate, i.e., 661 cm−1 (bending of B−O−B
DC hydrogel.25,27 linkages within borate networks), 833 cm−1 (B−O stretching
Structures and Morphologies of the Hydrogels. To from residual B), and 1330 and 1420 cm−1 (asymmetric
confirm whether there were ionic interactions between CS and stretching of B−O−C from the complexes).57 When borate
PVA chains in the presence of the cross-linking agent (B), ions were introduced into the networks of CS, PVA, and CS/
FTIR was carried out to gain insight into the vibration of PVA gels, separately, there were two new vibrational
molecules. As shown in Figure 1a, a large peak attributed to absorption peaks observed at 1330 and 1420 cm−1. The CS/
O−H stretching was observed and located at 3100−3700 cm−1 PVA/B DC hydrogel spectra showed the characteristic
in all spectra. The C−O−C bond of starch showed discernible absorption peaks of starch, PVA, and B. This confirmed not
absorption peaks at 1160 and 1080 cm−1.53,54 In the pure only the occurrence of complexes between borate ions and CS
spectrum of PVA, the absorption peak at 1100 cm−1 was or PVA chains but also the presence of cross-linking networks
related to the C−O in stretching mode for PVA.55 Compared among CS chains and PVA chains with borate ions within the
with the CS/PVA gel, the −OH stretching peak of the CS/ hydrogels.
PVA/B DC hydrogel shifted to a lower wavenumber (blue- The XRD profiles of SN (CS and PVA), SC (CS/B and
shift). Stronger H-bonding leads to a more significant shift in PVA/B), DN (CS/PVA), and DC (CS/PVA/B) gels are
vibrational frequency.50 Because borate ions are probably presented in Figure 1b. The retrograded CS SN gel at low
attached to the −OH groups of PVA, amylose, and/or temperature exhibited a B-type with diffraction peaks at 2θ of
amylopectin chains via H-bonds, the interactions of H-bonds 17.1° and 22.2°.58 With the addition of B, the typical peaks of
3969 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 2. Photographs showing the processes of the compressing and loosening behaviors of shapeable CS-based hydrogels: (a) CS single network
hydrogel, (b) CS/PVA double network hydrogel, (c) CS/B single cross-linked hydrogel, and (d) CS/PVA/B double cross-linked hydrogel.
Schematic illustration of the possible mechanisms of the CS-based gel network compression and loosening process.

CS crystallites disappeared, with only a blunt amorphous peak Morphology of the Hydrogels. SEM images of the inner
centered at 2θ = 22.2°. These results revealed that the strong morphologies of freeze-dried hydrogels are shown in Figure
interaction between CS chains and borate irons inhibited the 1(c−h). The pure CS gels showed a lamellar and loose
retrogradation of the amylose and/or amylopectin in starch microstructure. After borate ions were introduced, the CS/B
and led to the decrease of CS crystallites. The PVA gel showed hydrogel structure became dense in the pore wall. The fracture
two typical peaks at 2θ of 19.4° and 22.9°.59 Borate ions surface of the PVA gel was rather flat and smooth without
destroyed the organized configuration of PVA, and thus the holes, while the addition of B to the hydrogel resulted in a fine,
diffraction peaks of PVA/B gel became weaker.51 The intensity micrometer-porous network structure. The phenomenon may
of the diffraction peak at 17.1° of CS/PVA gel decreased be caused by the formation of densely cross-linked complexes
significantly, indicating that the hydrogen bonding interactions among starch (amylose and/or amylopectin) or PVA chains
with borate ions. Moreover, CS/PVA gel exhibited a
between CS and PVA chains were disadvantageous to
comparatively coarse and porous surface, which combined
improving the ordering arrangement of the amylose and/or
the characteristics of both CS and PVA. Compared with two
amylopectin chains. However, the CS/PVA/B DC gel different cross-linked gels (CS/B and PVA/B), the surface
displayed very low crystallinity, even lower than those of the image of the CS/PVA/B DC gel in Figure 1h formed an
CS/PVA gels. The result indicates that borate ions could enhanced microstructure with a large number of small, regular
weaken the self-hydrogen bonding of the DC gel to a cellular nests. From the micromorphology, the internal
considerable degree by forming new cross-linking bonds with skeleton of the DC gel was more orderly and compact,
CS and/or PVA chains and thus further restrains the which could lead to its better mechanical properties.
crystallization of CS/PVA gels. It was thus concluded that Steadiness of the Hydrogels. The cylindrical shaped
the presence of borate ions and PVA chains would inhibit the samples were kept at room temperature for 60 s, and a shape
retrogradation of amylose and amylopectin in starch, and in change was observed (Figure S1). In particular, the shape of
turn affected the starch crystal structure of DC hydrogels. the PVA/B SC hydrogel showed significant changes, proving
3970 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 3. Mechanical properties of hydrogels at room temperature: (a) compressive stress−strain curves, (b) compressive loading−unloading
curves, and (c) compression modulus and hysteresis energy of CS, CS/B, PVA/B, CS/PVA, and CS/PVA/B hydrogels. (d) Photographs revealing
the notable compression resistance of CS/PVA/B DC hydrogel.

that the bonds of borate ions and PVA hydroxyls are properties. In spite of withstanding high-level deformation
dynamic.32 The prepared pure PVA specimen was in the sol compression (over 50%), the CS/B and CS/PVA/B gels
state and exhibited good fluidity. Compared with the PVA/B exhibited only deformation behavior, while no stiff CS network
SC hydrogel, the shape of the DC hydrogel remained the same. or flexible PVA network destruction was observed. The
The reason for the higher stability of the hydrogel is the greater mechanical strength and toughness of the SC and DC gels
number of hydrogen bonds and the stronger cross-linked was an enhanced result owing to the cross-linked interactions
interactions obstructing the motion of the PVA chains.60,61 of CS and/or PVA chains with borate ions. More importantly,
Moreover, the first starch network also provided an additional upon removal of the deformation force, the compressed DC
platform to firm the structure, thus improving the stability of gels could recover their initial height and shape (after five
the second network. Although all of the CS-based hydrogels repetitions), while the CS/B SC hydrogels partly returned to
(CS, CS/B, CS/PVA, and CS/PVA/B) could remain at the their initial height (approximately 60%). CS supported many
same initial shape, significant differences in the mechanical hydroxyl groups on their surfaces that formed hydrogen bonds
strengths were exhibited. with PVA chains, forming a cross-linked network with chain
Mechanical Properties. Then, we initially compared the entanglement. The borate ions cross-linked the chain
compression behavior of CS-based hydrogels. While the entanglement of CS and PVA and further reinforced the
hydrogels were compressed, the cylindrical CS SN and CS/ networks, which enhanced the ductility of the DC gels (Figure
PVA DC gels exhibited different degrees of damage (Figure 2a, 5k, l). The cyclic compression and stress relaxation processes
b). Because the polysaccharide-based hydrogels exhibited poor revealed that SC and DC gels exhibited excellent shape-
mechanical strength and a brittle network structure,62,63 we recovery properties.30,64 To further evaluate whether CS-based
speculated that the breakage of the SN and DN hydrogels were hydrogels could withstand continuous compression stress,
related to the first starch networks formed through hydrogen optical photographs of the mechanical performance of the
bonds of amylose and/or amylopectin chains (Figure 2e, f). As hydrogels were taken, as shown in Figure S2. The CS, CS/B,
the borate ions were introduced into the hydrogels, the CS/B and CS/PVA hydrogels break down directly when the gels
SC and CS/PVA/B DC hydrogels could be compressed using were subjected to weight (∼1 kg) for 5, 60, and 15 s,
high strain without breaking (Figure 2c, d), indicating that respectively. By contrast, the DC hydrogel with a slight
introducing borate ions markedly strengthens the networks of deformation sustained a loading weight of ∼1 kg (approx-
the two hydrogels and endows extraordinary mechanical imately 300 times its own weight) for more than 5 min. Upon
3971 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 4. Self-recovery of CS/PVA/B DC hydrogels at room temperature: (a) compressive curves, (b) loading−unloading curves, and (c) time-
dependent hysteresis recovery and hysteresis energy histogram of DC hydrogels and the corresponding recovery photographs (inserted image). (d)
The consecutive compression curves with gradient increased in maximum strain on CS/PVA/B DC hydrogels; the DC hydrogel withstand the
compressive strain of 99% without breakage.

removal of the load, the DC hydrogel rapidly recovered its decreasing trend. Previous studies reported that excessive
original shape within 15 s. These results determined that cross-linking could cause the first and second networks to
introducing PVA chains and borate ions reinforced DC become brittle and lose ductility.17 Therefore, the CS/PVA/B
hydrogels and made the DC gels able to withstand continuous DC gels prepared under optimal conditions were used in the
compression and, consequently, possess higher stiffness and subsequent experiments.
compression strength, which far exceed those of the parent gels Favorable mechanical properties are of great significance for
(CS, CS/B, and CS/PVA). hydrogels in practical applications. The pure CS SN and CS/
To better understand how the second PVA/B network PVA DN hydrogels exhibited brittle behavior (Figure 3a) and
compositions affect the mechanical properties of DC gels, we fractured at a strain of 32.3 and 29.4% (Figure S5a),
conducted a series of compressive tests to investigate the effect respectively, which meant the cross-linked networks of the
of PVA chains or borate ions on the mechanical properties of hydrogen bonds were not able to withstand large deformations.
DC hydrogels (Figures S3 and S4). When the PVA The load force changed the orientation and relative position of
concentrations were increased from 2 to 4 wt %, the DC amylose and amylopectin chains in the CS gel. Meanwhile, the
gels exhibited a monotonically enhanced compression stress, interstitial waters were squeezed out, and thus, the
compression modulus, and hysteresis energy. The result was compression stress of the CS increased with the increasing
possibly due to more CS chains cross-linked with PVA chains strain. However, the interspace for chain movement
via borate ions, thus leading to an enhancement in the determined the increase of the compression strain. Once the
mechanical properties of the DC hydrogel. However, a further compressive deformation reached its limit, a further increase in
increase in the PVA content resulted in the decrease of the load resulted in much more breakage of the gel networks.
mechanical strength in the DC gel, which might be attributed By contrast, the reinforced CS/B and PVA/B SC hydrogels did
to the inhibited formation of the first starch networks by not break at strains as high as 70%, suggesting a soft yet tough
residual PVA chains. Similar to the PVA concentration effect, behavior. Moreover, incorporation of borate ions and PVA
when the DC gels contained 0.026 mol/L of B, both chains to form the PVA/B SC gel also conquered the low
compression stress and hysteresis energy achieved maximal modulus of the PVA SN gel. It can be clearly seen that when
values of 219.6 ± 15.7 kPa and 5790 ± 450 J m−3, respectively the second PVA/B network was combined with the CS
(Figure S4bd). However, at the high B concentration (0.039 network, the mechanical strength of the DC gel increased
mol/L), the mechanical behavior of the DC gels exhibited a dramatically. For instance, the compression stress reached up
3972 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 5. CS/PVA/B DC hydrogels exhibiting an excellent free-shapeable property to form any complex shape with very fine structures, including
starfish (a), shell (b), conch (c), and cherry blossom (d, e). DC gels are highly strong and flexible to withstand stretch winding (f), convolving (g),
knotting (h), knot stretching (i), and stretching (j) without breakage. Tensile stress−strain curves of DC hydrogel: the DC gel achieved the highest
strain of 2485% (k). The image of stretched hydrogel (l) and the schematic diagram of original and stretched DC gel network structure (m).

to 209.8 ± 16.8 kPa for the DC gel, which was 1090%, 1200%, hydrogels, confirming the more efficient energy dissipation of
and 1410% higher than those of the corresponding CS SN the DC gels. Moreover, the energy dissipation of the DC gels
(19.2 ± 2.3 kPa), CS/B SC (17.4 ± 1.6 kPa), and CS/PVA reached as high as 5750 ± 450 J/m−3, which was 4.5, 8.0, and
DN (14.7 ± 1.3 kPa) hydrogels, respectively (Figure S5b). 11.5 times higher than the CS, CS/B, and CS/PVA gels,
Moreover, the compression modulus was 13.54 kPa for the DC respectively (Figure 3c). Intriguingly, the values of the
hydrogel, 8.61 kPa for the CS SN hydrogel, and 7.21 kPa for cumulative energy dissipation of all the control specimen gels
the CS/PVA DN hydrogel (Figure 3c), indicating the (CS, CS/B, CS/PVA, and PVA/B) were approximately two
mechanical strength of the DC gel exceeded those of either
times lower than that of the DC hydrogel. The results further
of its parent gels. Furthermore, the photographs of the DC
illustrated that incorporating PVA chains and borate ions
hydrogel for the presentation of mechanical properties have
been recorded as well. As shown in Figure 3d, the DC gels can improved the mechanical properties of the starch hydrogels.
partly return to their initial shapes after compression with a We speculated that the reason for the notable hysteresis loop
strain of 70%−80%, showing a much improved toughness than could stem from the much more reversible dissociation−
conventional hydrogels.65 Cyclic compressive assays were association of CS and PVA chains cross-linked by the ionic
performed to evaluate the energy dissipation capacity of the bridge of the borate ions upon compression. Meanwhile, the
hydrogels (Figure 3b, c). The CS/PVA/B DC hydrogel had a reversible dissociation−association process enabled an efficient
larger hysteresis loop than the CS-based and PVA/B SC stress transfer and energy dissipation,46 contributing more
3973 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

significant viscous dissipation and shock absorbing capacity the DC gels was due to the chain entanglement of amylose
than those of H-bonds in the SN gels. and/or amylopectin chains and PVA chains through ionic
Self-Recovery Behaviors of DC Hydrogels. In addition interaction of the borate ions (Figure 5m). The fantastic
to the reinforcement of mechanical strength, another advanta- integration of good ductility, flexibility, and free-shapeability
geous feature of the DC network structure is the ability to would broaden the scope of application for the fabricated
undergo repeated association to endow excellent self-recovery composite hydrogels.33
behaviors of gels (Figure 2d). To evaluate the self-recovery Self-Healing Capacity of DC Hydrogels. The novelty of
efficiency of DC hydrogels, the recovery performance of the this DC hydrogel is not only reflected in its excellent
gels at a strain of 70% with different interval times (0−5 h) was mechanical and free-shapeable properties but also in its
tested. The compression strength and hysteresis loop of the outstanding automatic self-healing capacity. We conducted a
DC hydrogels gradually increased with the extension of the macro self-healing experiment where the cylindrical DC gel
resting time (Figure 4a, b). Moreover, we used the hysteresis was cut into three pieces, and the pieces were kept in close
energy to calculate the recovery efficiency of the DC hydrogel, proximity, as represented in Figure 6a. After contact for 15 s, it
and the recovery efficiency of DC gel exceeded 81.9% and
93.5% after 30 min and 5 h, respectively. As the resting time
was prolonged from 0 min to 5 h, the hysteresis energy
increased from 3970 ± 190 to 5360 ± 310 J m−3. Owing to the
dynamic nature of the coordination interactions (hydrogen
bonds, borate ion cross-link interactions, etc.) within the DC
hydrogel, they can be rapidly reestablished when the external
load is removed. Therefore, the gels could recover all or part of
the original mechanical properties if the second load was
delayed for a longer time, highlighting their efficiency as an
energy dissipating mechanism. The morphology of the DC gels
after recuperating for 30 min and 5 h is shown in Figure 4c.
The consecutive compression curves with gradient increases in
the maximum strain on DC hydrogels are shown in Figure 4d.
Upon increasing the strain (70%−99%), the compressive
strength of the gels increased from 211.7 to 547.8 kPa. The
result suggested that the DC gels mentioned above could
sustain the high-level compressive deformation (strain up to
99%), and the breakage surface of the gel was not observed
after the load was removed.
Free-Shapeable Properties and Stretchability of DC
Hydrogels. Figure 5 shows that the DC hydrogels displayed a
milky white color as well as strong and free-shapeable
properties. As shown in Figure 5a−e, DC gels were highly
free-shapeable to form different fine structures, such as a Figure 6. Visual evidence of self-healing of a CS/PVA/B DC
starfish, shell, conch, and cherry blossom shapes. This is hydrogel. A cubic DC hydrogel is cut into three pieces (a1); two or
important to extend the starch-based hydrogel’s application, three parts of the hydrogels are then placed in contact with each other
especially in the bioengineering field where artificial tissues for 15 s (a2, a3), and the whole hydrogel is lifted without any of the
pieces falling. (a4) The self-healed hydrogel on a bridge support which
with specific geometrical shapes are used to achieve their can withstand 50 g of weight (approximately 5-fold its own weight).
biologically mimicking functions.13,28 Additionally, DC gels are After the pieces were in contact for 30 s, the self-healed hydrogel
highly ductile and flexible to withstand stretch winding (Figure could withstand 1000 g of weight (approximately 100-fold its own
5f), convolving (Figure 5g), knotting (Figure 5h), knot weight) without breakage. Conch shaped hydrogel: the cut sample of
stretching (Figure 5i), and stretching (Figure 5j). Moreover, the conch hydrogel (b1). The self-healed gel can bear its own weight
an intact DC gel was easily stretched to approximately 4-fold (b2) and sustain bending without breaking (b3, reverse; b4, lateral).
its original length (Figure 5j). Particularly upon removal of the Shell shaped hydrogel: strip shaped hydrogel (c1); the self-healed shell
deformation force, the convolving hydrogels could recover hydrogel also can withstand its weight (c2) and bend in a U shape
their initial shapes, which further confirmed that the gels without breaking (c3, up−down; c4, left−right). The shell hydrogel is
cut into small pieces (c5); the pieces are put in the shell- or crab-
exhibit excellent shape-recovery properties. Afterward, the
shaped mold, and are molded in air or water for 5 h. The gel pieces
recuperative DC gel was knotted again, which also bore a large heal as integral shell- and crab-shaped hydrogels (c6, c7). The points
deformation of 350% without breakage (Figure 5i). As a proof- of contact are encircled by a red circle.
of-concept, Figure 5k and l shows the tensile stress−strain
properties of the DC hydrogel. The pure starch matrix
hydrogels were too brittle to be tested for mechanical tensile should be noted that the pieces of hydrogel joined together to
properties, while the obtained DC hydrogel exhibited much form a single entity without a visible joint interface, and then
better tensile properties. The fracture strain of the DC gel the integrated gel could be lifted without falling (Figure 6a2,
reached up to 2480% with a tensile strength of ∼12500 Pa, a3), supported its own weight by placing it on a bridge (Figure
which are the top values ever reported in starch hydrogel 6a4), and withstood an ∼50 g weight (approximately 5-fold its
systems. In addition, the visual inspection of the stretching own weight). If the contact time of the pieces was delayed for
demonstrated that the DC hydrogels had excellent mechanical 30 s, we found that the self-healed DC hydrogels did not break
tensile properties (Figure 5i). The enhanced stretchability of under a static load of ∼1 kg (more than 100-fold its own
3974 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 7. Schematic illustration of the mechanism of CS/PVA/B DC hydrogel self-healing process in air at room temperature (a).The tensile
stress−strain curves of the self-healed DC hydrogels with various healing time at room temperature (b), the corresponding fracture stress and
fracture strain (c), and the elastic modulus and toughness (d) of self-healed DC hydrogels for the indicated times. Numbers denote the healing
efficiency of DC gels.

weight). Moreover, the self-healing property of the conch- and healed gels were defined as the self-healing efficiency.68 The
shell-shaped DC hydrogel was also proven by lifting it with fractured parts were put back together, in contact, for different
fingers, standing it upright on a table, and by bending the gel in lengths of time (15, 30, 60, and 600 s) in the air at room
a U shape (conch-shaped, reverse and lateral; shell-shaped, temperature. The healed hydrogel was stretched to test the
up−down and left−right) as shown in Figure 6b and c (b1−b4 tensile strength (Figure 7). In the case of 15 s healing time, the
and c1−c4) where the self-healed pieces are intact. The point of cut gels were partially self-healed; the healed gel could sustain a
self-healing can be noted from the remaining scar (red circle). tensile stress of 6300 Pa with elongation strain of 985% before
Afterward, we further challenged our DC gels to inspect their fracture. Moreover, the elastic modulus, fracture strain, and
self-healing properties. As shown in Figure 6c, the self-healed healing efficiencies significantly increased (p < 0.05) with an
shell-shaped DC hydrogels were cut into multiple fragments increased healing time. When the healing time exceeding 600 s,
again (c5) and put into a shell- and crab-shaped mold, the fracture energy of the healed gel was 199.1 ± 21.1 kJ m−3
respectively. It is noteworthy that the fragments could also be with a strain over 2000% (the fracture energy and fracture
remolded and could form an unabridged shell- and crab- strain of the initial gel were 290.5 ± 19.4 kJ m−3 and 2485%,
shaped hydrogel both in air and underwater after contact for 5 respectively). The healing efficiency was 68.5% based on stress
h (c6, c7), which suggested excellent self-healing performance or 81.2% based on strain. The tensile strain recovered to more
of the DC hydrogels. The dynamic cross-linking of CS and/or than 80%, and thus, we speculated that the CS network could
PVA chains with borate ions was possibly still carried out, with be responsible for the healing process. Therefore, we also
these chains intertwining with each other, and thus the investigated the self-healing behaviors of the parents of the DC
hydrogels could heal. With the rapid self-healing capacity both hydrogel. Although the CS SN and CS/PVA DN hydrogels did
in air and underwater, the DC gels will have huge potential not exhibit self-healing (Figure S6), the introduced borate ions
applications, especially in fields that require a fast response or cross-linked CS/B hydrogel possessed a self-healing behavior
fast working behaviors. (Figure S7). As shown in Figure S7, the two pieces of CS/B
Healing efficiency is an important parameter to characterize SC hydrogel formed an integrated gel which was lifted and
the degree of self-healing; therefore, tensile tests were could withstand its weight (b, c3). Afterward, the two broken
conducted on the original and self-healed samples.66,67 starfish-shaped SC hydrogels healed completely without a
Fracture strain and stress recovery percentages of the self- visible joint interface after 12 h at 25 °C. The healed gel could
3975 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 8. Swelling properties of hydrogels. Photos of the shapeable CS, CS/B, CS/PVA, and CS/PVA/B aerogels before (a) and after (b) swelling
in deionized water. Light-weight CS/PVA/B DC aerogel could be supported by a tender leaf (c). Swelling kinetics of CS, CS/B, PVA, PVA/B, CS/
PVA, and CS/PVA/B aerogels in the deionized water at room temperature. The PVA, CS/B, and PVA/B aerogels reach equilibrium within 9 h,
while the CS, CS/PVA DN, and CS/PVA/B DC aerogels can reach equilibrium at approximately 36, 30, and 60 h, respectively (d). Compression
stress−strain curves of the CS, CS/B, CS/PVA, and CS/PVA/B gels in the swollen state (e).

withstand gentle stretching (c4). These results further This phenomenon is possible due to the stronger double cross-
confirmed the occurrence of cross-linking networks between linked structure and more chain entanglements via hydrogen
CS chains with borate ions within the CS/B SC hydrogels. In bonds, making the network closer and hindering the water
addition, we predicted the healing mechanisms of the DC gel molecules from accessing the hydrogel cage, which results in
(Figure 7a). When the two fracture surfaces were put in lower EWC. When borate ions are introduced into the
contact at room temperature, the borate bonds within the network, we noticed the EWC of the cross-linked network gels
network were able to recross-link the broken CS and/or PVA (CS/B, PVA/B, and CS/PVA/B) was larger than the
chains by forming new complexes. The CS chains were corresponding control (CS, PVA, and CS/PVA). The borate
dragged by PVA chains and tangled with each other via ions may influence the formation of the crystallinity of the
hydrogen bonds to form a new network. Self-healing starch and PVA networks and result in an increased
capabilities were demonstrated by uniting the two pieces amorphous region of the hydrogels; thus, their water-holding
into one.69 Once they were repaired with the reversible ionic abilities increased. This observation was also noted in the X-ray
cross-link and hydrogen bond mechanism, the hydrogel results (Figure 1b). At the same time, Figure 8b exhibits the
samples could regain their mechanical properties. Overall, dried and swollen CS, CS/B, CS/PVA, and CS/PVA/B gels,
the fast self-healing of the DC gels resulted from the which nearly hold their initial shapes in a swollen state.
reestablishment of the network structure. However, the swollen PVA and PVA/B gels were collapsed and
Swelling Properties of Hydrogels. Figure 8d shows the could not retain their shapes, and thus, the PVA and PVA/B
time dependence of the water absorbency of the hydrogels. were not shown in the figure.
Due to the loose network structure of the hydrogel, the SN Then, we further inspected the compressive properties and
(CS and PVA) and SC (CS/B and PVA/B) aerogels exhibited textural hardness of the pillar-shaped hydrogels in their swollen
fast initial swelling. Compared with the SN and SC gels, the states (Figures 8c and S5). The mechanical properties of the
CS/PVA DN and CS/PVA/B DC gels were capable of swollen CS and CS/B gels were slightly decreased compared
absorbing fewer quality water molecules. Moreover, the with the original levels. This result was due to the destruction
equilibrium water content (EWC) of DC and DN was only of the network structure during the freeze-drying process,
4.52 and 3.15 g/g, respectively (Figure 8a). These results causing the SN and SC gels to exhibit a weaker mechanical
suggested that the DC hydrogel exhibited good swelling property. As listed in Table 1, the water content of the gels
resistance as compared with those of either of its parent gels. remained at a high level (70−95%) owing to the strong H-
3976 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Figure 9. Effects of contact time, pH, temperature, hydrogel dosage, and initial MB concentration on the adsorption amount of MB onto gels. (a)
Contact time (t = 0−48 h, C0 = 150 mg/L, V = 100 mL, T = 25 °C, pH = 6.45, dos. = 100 mg), (b) temperature (CS/PVA/B DC gel, C0 = 150
mg/L, V = 100 mL, t = 24 h, pH = 6.45, dos. = 100 mg), pH (CS/PVA/B DC gel, C0 = 150 mg/L, V = 100 mL, t = 24 h, T = 25 °C, dos. = 100
mg), (c) initial concentration (CS/PVA/B DC gel, t = 24 h, V = 100 mL, T = 25 °C, dos. = 100 mg), and hydrogel dosage (CS/PVA/B DC gel, C0
= 150 mg/L, t = 24 h, V = 100 mL, T = 25 °C, pH = 6.45). Here, qe represents the adsorption amount of MB onto gels at equilibrium.

bonds and sufficient interaction between the hydroxyl groups gradually slowed until equilibrium was reached (Figure 9a).
of adjacent polymer strands.70 Therefore, we speculated that This may be ascribed to the higher MB concentration, faster
the reason for the decreased strain was more water molecules diffusion, and more adsorptive sites at the initial adsorption
were introduced into the CS network from the swelling, phase.71 With increasing adsorption time, the vacant adsorptive
thereby leading to decreased interactions between the CS sites of the DC gel decreased, which lowered the probability of
strands. The freeze-drying process enhanced the crystalline and interactions between MB and the adsorptive site, and
cross-linking density of PVA within the hydrogel, and consequently, the adsorption gradually slowed and eventually
consequently, the mechanical strength of the CS/PVA gel tended to balance. When the adsorption of the hydrogels
increased. It is noteworthy that although the strength of the achieved equilibrium, the MB adsorption capacities of CS/B,
swollen DC hydrogels was lower than that of the initial DC PVA/B, and CS/PVA/B respectively reached 80.32, 65.26, and
gels upon addition of borate ions, the swollen DC hydrogels 144.68 mg/g, which were higher than those of the
still exhibited excellent mechanical properties. Figure 8c shows corresponding CS (16.43 mg/g), PVA (38.29 mg/g), and
the compression stress−strain curves of the swollen hydrogels, CS/PVA (26.41 mg/g) gels. Considering the anionic
and one can see the DC hydrogel demonstrated high characteristic of the hydrogels, we hypothesized that a possible
mechanical performance with 166.2 kPa compression stress, reason for the increase of the adsorbed ionic dyes might be
70% compressive strain, and 1610.7 g of hardness (Figure S5). attributed to the electrostatic interactions.71 Furthermore, the
The mechanical strength recovered to a great extent, i.e., the porous cross-linked structure would be beneficial for the
recovery efficiency of swollen DC gel reached as high as 79.2% adsorption process since it increased surface area and
and 100% based on the compression stress and strain, facilitated diffusion of dyes into the adsorbent.72 Thus, a
respectively. In addition, the CS/PVA/B DC aerogel was significant increase in the adsorption capacity of CS/PVA/B
lightweight and could be supported by a tender leaf (Figure was observed. Moreover, the color of the solution with CS/
8c). PVA/B DC gel turned significantly transparent (Figure S8),
Absorption of MB. Contact time is an important factor especially in comparison with CS, indicating the beneficial
that indicates whether the adsorbent removes the target effects of borate ions and PVA on MB adsorption.
contaminant to reach the equilibrium. It is apparent that the The effects of temperature and pH on the removal amount
adsorption amount rapidly increased initially and then are important factors to evaluate dye adsorption in agricultural
3977 DOI: 10.1021/acs.jafc.8b05147
J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

and industrial wastewater at different temperatures and pH break. What’s more, because of its unique reversible network
values. When the temperature increased, the adsorption structures, the gel system is able to form any complex shape.
capacity of MB on DC gel decreased; however, it was still The combination of good mechanical properties, self-recovery
higher than that of the gels (CS, CS/B, PVA, PVA/B, and CS/ of strength and toughness, and free-shapeable ability, along
PVA) at 25 °C. As the solution pH was changed from acidic to with an easy synthesis method, allow the materials to have
basic, the zeta potential of the swollen DC gel solution many potential applications in the fields of biomaterials and
decreased from −0.24 to −25.60 mV (Figure S9). It can be load-bearing artificial soft tissues. In addition, the introduction
speculated that the isoelectric point (pI) for the DC gel was of borate ions and PVA play an important role in the
less than pH 1.0. When the solution pH > pI, the DC gel is adsorption of MB by DC hydrogel, whose maximum removal
negatively charged, which was favorable for adsorbing MB amount reaches 144.68 mg/g, which is much higher than that
owing to the electrostatic attraction. It can be seen that the MB of CS gel. Consequently, the prepared DC gels could serve as
adsorption of DC hydrogels was highly pH dependent. Close eco-friendly, stable, and efficient adsorbents for cationic dyes
to neutrality or under alkaline conditions, an obvious increase from agricultural and industrial effluents.
in the adsorption amount was observed. Moreover, the
adsorption amount of MB on DC gel markedly increased
from 56.33 to 144.29 mg/g as the solution pH increase from

*
ASSOCIATED CONTENT
S Supporting Information
2.0 to 8.0 (Figure 9b). Under acidic conditions (pH of 1.0− The Supporting Information is available free of charge on the
5.0), the protonation of boric acid (B(OH)3) groups was ACS Publications website at DOI: 10.1021/acs.jafc.8b05147.
unfavorable for the adsorption process due to the fewer
Compressive stress−strain curves, compression stress
hydrogen bonds between the DC gel and MB. At basic
and modulus, compressive loading−unloading curves,
conditions, the deprotonation of B(OH)4− resulted in the
and corresponding hysteresis energy of hydrogels with
destruction of hydrogen bonding interactions and the “charge
different PVA and B concentrations; hydrogel steadiness
screening effect” of excess Na+ ions also prevented effective images; photographs of the mechanical performance of
electrostatic interactions between the borate ions of gel and CS-based hydrogels; self-healing properties of CS, CS/
MB. Under neutral conditions, the moderate residual B(OH)4− PVA, and CS/B hydrogels; compressive strain, stress,
and B(OH)3 groups may be conducive to the formation of and textural hardness of hydrogels in the initial and
electrostatic interactions and hydrogen bonds between DC gel swollen states; and images of the MB dye solution
and MB, leading to the optimal adsorption capacity of DC gel adsorption on different hydrogels (PDF)


for MB. As depicted in Figure 9c, because the initial MB
concentration is fixed, when the gel dosage increased, the total
adsorption amount of MB on the hydrogel increased, but the AUTHOR INFORMATION
adsorption amount per unit mass of DC gel was decreased and Corresponding Author
the utilization of adsorbent would be greatly reduced. In *E-mail: fpcenter@jiangnan.edu.cn.
addition, the adsorption amount was significantly affected by ORCID
the initial concentrations of dye solution. With increasing the Zhengyu Jin: 0000-0001-6880-9955
MB concentrations, the adsorption amount exhibited an
Funding
initially increasing and then an equilibrium trend (Figure
This work was supported by the National First-Class Discipline
9c). The removal amount was 143.66 mg/g at 150 mg/L MB
Program of Food Science and Technology
dye initial concentration, which increased to 170.29 mg/g at
(JUFSTR20180203).
200 mg/L initial concentration.
In summary, we have designed and fabricated a one-pot Notes
The authors declare no competing financial interest.


method, which is a simple, time-saving, and easily controlled
process to prepare hybrid DC hydrogels. Within the DC gel,
hydrogen bonds cross-linked to CS chains created the rigid and REFERENCES
brittle first network, and dynamic borate bonds cross-linked to (1) Li, S.; Yi, J.; Yu, X.; Shi, H.; Zhu, J.; Wang, L. Preparation and
PVA chains formed the second network. The first network Characterization of Acid Resistant Double Cross-Linked Hydrogel for
provides reversible sacrificial bonds and dissipates energy, Potential Biomedical Applications. ACS Biomater. Sci. Eng. 2018, 4,
whereas the second network provides elasticity to the DC 872−883.
(2) Xu, R.; Ma, S.; Lin, P.; Yu, B.; Zhou, F.; Liu, W. High Strength
hydrogel. Because of the distinct hybrid physically and Astringent Hydrogels using Protein as the Building Block for
chemically cross-linked structure, the DC hydrogels showed Physically Cross-Linked Multi-Network. ACS Appl. Mater. Interfaces
notably improved mechanical properties over the single 2018, 10, 7593−7601.
physically cross-linked starch-based hydrogels. The tensile (3) Wang, H.; Heilshorn, S. C. Adaptable Hydrogel Networks with
fracture strains and toughness of the DC hydrogels could be as Reversible Linkages for Tissue Engineering. Adv. Mater. 2015, 27,
high as 2485% and 290.5 kJ m−3, respectively. Moreover, the 3717−3736.
compression strength of the DC gels reached 547.8 kPa at a (4) Webber, M. J.; Appel, E. A.; Meijer, E. W.; Langer, R.
compression strain of 99%, which was significantly higher than Supramolecular Biomaterials. Nat. Mater. 2016, 15, 13−26.
those of the starch matrix hydrogels previously reported in the (5) Luo, K.; Yang, Y.; Shao, Z. Physically Crosslinked Biocompatible
Silk-Fibroin-Based Hydrogels with High Mechanical Performance.
literature. They retained their original shapes after continuous
Adv. Funct. Mater. 2016, 26, 872−880.
loading−unloading compressive cyclic tests, demonstrating (6) Loh, X. J.; Peh, P.; Liao, S.; Sng, C.; Li, J. Controlled Drug
their superior self-recovery ability. Furthermore, the DC gels Release from Biodegradable Thermoresponsive Physical Hydrogel
exhibited remarkable self-healing properties both in air and Nanofibers. J. Controlled Release 2010, 143, 175−182.
underwater at room temperature. Especially, the healed DC (7) Zhu, P.; Hu, M.; Deng, Y.; Wang, C. One-Pot Fabrication of a
hydrogel specimens still showed ∼20 times elongation at their Novel Agar-Polyacrylamide/Graphene Oxide Nanocomposite Double

3978 DOI: 10.1021/acs.jafc.8b05147


J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Network Hydrogel with High Mechanical Properties. Adv. Eng. Mater. Network Hydrogels Using Thermoreversible Sol-Gel Polysaccharide.
2016, 18 (10), 1799−1807. Adv. Mater. 2013, 25, 4171−4176.
(8) Jia, H.; Li, Z.; Wang, X.; Zheng, Z. Facile Functionalization of a (28) Chen, H.; Chen, Q.; Hu, R.; Wang, H.; Newby, B. M. Z.;
Tetrahedron-Like PEG Macromonomer-Based Fluorescent Hydrogel Chang, Y.; Zheng, J. Mechanically Strong Hybrid Double Network
with High Strength and Its Heavy Metal Ion Detection. J. Mater. Hydrogels with Antifouling Property. J. Mater. Chem. B 2015, 3 (27),
Chem. A 2015, 3, 1158−1163. 5426−5435.
(9) Ha, W.; Yu, J.; Song, X. Y.; Chen, J.; Shi, Y. P. Tunable (29) Chen, Q.; Wei, D.; Chen, H.; Zhu, L.; Jiao, C.; Liu, G.; Huang,
Temperature-Responsive Supramolecular Hydrogels Formed by L.; Yang, J.; Wang, L.; Zheng, J. Simultaneous Enhancement of
Prodrugs as a Co-delivery System. ACS Appl. Mater. Interfaces 2014, Stiffness and Toughness in Hybrid Double-Network Hydrogels via
6, 10623−10630. the First, Physically Linked Network. Macromolecules 2015, 48, 8003−
(10) Seliktar, D. Designing Cell-Compatible Hydrogels for 8010.
Biomedical Applications. Science 2012, 336, 1124−1128. (30) Wang, Y. X.; Niu, J. Y.; Hou, J.; Wang, Z. C.; Wu, J. N.; Meng,
(11) Taylor, D. L.; in het Panhuis, M. Self-Healing Hydrogels. Adv. G. H.; Liu, Z. Y.; Guo, X. A Novel Design Strategy for Triple-Network
Mater. 2016, 28, 9060−9093. Structure Hydrogels with High-Strength, Tough and Self-Healing
(12) Ikeda, M.; Tanida, T.; Yoshii, T.; Kurotani, K.; Onogi, S.; Properties. Polymer 2018, 135, 16−24.
Urayama, K.; Hamachi, I. Installing Logic-Gate Responses to a (31) Liu, S.; Li, L. Recoverable and Self-Healing Double-Network
Variety of Biological Substances in Supramolecular Hydrogel-Enzyme Hydrogel Based on κ-Carrageenan. ACS Appl. Mater. Interfaces 2016,
Hybrids. Nat. Chem. 2014, 6, 511−518. 8, 29749−29758.
(13) Yin, M. J.; Yao, M.; Gao, S.; Zhang, A. P.; Tam, H. Y.; Wai, P. (32) Chen, W. P.; Hao, D. Z.; Hao, W. J.; Guo, X. L.; Jiang, L.
A. Rapid 3D Patterning of Poly(acrylic acid) Ionic Hydrogel for Hydrogel with Ultrafast Self-Healing Property Both in Air and
Miniature pH Sensors. Adv. Mater. 2016, 28 (7), 1394−1399. Underwater. ACS Appl. Mater. Interfaces 2018, 10, 1258−1265.
(14) Sun, J. Y.; Zhao, X.; Illeperuma, W. R.; Chaudhuri, O.; Oh, K. (33) De France, K. J.; Hoare, T.; Cranston, E. D. A Review of
H.; Mooney, D. J.; Vlassak, J. J.; Suo, Z. Highly Stretchable and Hydrogels and Aerogels Containing Nanocellulose. Chem. Mater.
Tough Hydrogels. Nature 2012, 489, 133−136. 2017, 29, 4609−4631.
(15) Deng, Z. X.; Guo, Y.; Zhao, X.; Ma, P. X.; Guo, B. L. (34) Azevedo, S.; Costa, A. M. S.; Andersen, A.; Choi, I. S.; Birkedal,
Multifunctional Stimuli-Responsive Hydrogels with Self-Healing, H.; Mano, J. F. (2017). Bioinspired Ultratough Hydrogel with Fast
High Conductivity, and Rapid Recovery through Host−Guest Recovery, Self-Healing, Injectability and Cytocompatibility. Adv.
Interactions. Chem. Mater. 2018, 30 (5), 1729−1742. Mater. 2017, 29, 1700759.
(16) Yuan, N.; Xu, L.; Wang, H.; Fu, Y.; Zhang, Z.; Liu, L.; Wang, C. (35) Upadhyay, A.; Kandi, R.; Rao, C. P. An Injectable, Self-Healing
L.; Zhao, J. H.; Rong, J. H. Dual Physically Cross-Linked Double and Stress Sustainable Hydrogel of BSA as a Functional
Network Hydrogels with High Mechanical Strength, Fatigue Biocompatible Material for Controlled Drug Delivery in Cancer
Resistance, Notch-Insensitivity, and Self-Healing Properties. ACS Cells. ACS Sustainable Chem. Eng. 2018, 6, 3321−3330.
Appl. Mater. Interfaces 2016, 8, 34034−34044. (36) Lima-Tenório, M. K.; Tenório-Neto, E. T.; Guilherme, M. R.;
(17) Yan, X. Q.; Chen, Q.; Zhu, L.; Chen, H.; Wei, D. D.; Chen, F.; Garcia, F. P.; Nakamura, C. V.; Pineda, E. A. G.; Rubira, A. F. Water
Tang, Z. Q.; Yang, J.; Zheng, J. High Strength and Self-Healable Transport Properties Through Starch-Based Hydrogel Nanocompo-
Gelatin/Polyacrylamide Double Network Hydrogels. J. Mater. Chem. sites Responding to Both pH and a Remote Magnetic Field. Chem.
B 2017, 5, 7683−7691. Eng. J. 2015, 259, 620−629.
(18) Yang, Y.; Urban, M. W. Self-Healing Polymeric Materials. (37) Dragan, E. S. Design and Applications of Interpenetrating
Chem. Soc. Rev. 2013, 42, 7446−7467. Polymer Network Hydrogels. A Review. Chem. Eng. J. 2014, 243 (5),
(19) Wang, W.; Zhang, Y.; Liu, W. Bioinspired Fabrication of High 572−590.
Strength Hydrogels from Non-Covalent Interactions. Prog. Polym. Sci. (38) Lin, H. L.; Liu, Y. F.; Yu, T. L.; Liu, W. H.; Rwei, S. P. Light
2017, 71, 1−25. Scattering and Viscoelasticity Study of Poly(vinyl alcohol)−Borax
(20) Rauner, N.; Meuris, M.; Zoric, M.; Tiller, J. C. Enzymatic Aqueous Solutions and Gels. Polymer 2005, 46, 5541−5549.
Mineralization Generates Ultrastiff and Tough Hydrogels with (39) Bagri, L. P.; Bajpai, J.; Bajpai, A. K. Cryogenic Designing of
Tunable Mechanics. Nature 2017, 543, 407−410. Biocompatible Blends of Polyvinyl Alcohol and Starch with
(21) Zhang, Y. Y.; Li, Y. M.; Liu, W. G. Dipole-Dipole and H- Macroporous Architecture. J. Macromol. Sci., Part A: Pure Appl.Chem.
Bonding Interactions Significantly Enhance the Multifaceted 2009, 46 (11), 1060−1068.
Mechanical Properties of Thermoresponsive Shape Memory Hydro- (40) Follain, N.; Joly, C.; Dole, P.; Bliard, C. Properties of Starch
gels. Adv. Funct. Mater. 2015, 25, 471−480. Based Blends. Part 2. Influence of Polyvinyl Alcohol Addition and
(22) Zhang, J.; Zhu, Y.; Song, J.; Yang, J.; Pan, C.; Xu, T.; Zhang, L. Photocrosslinking on Starch Based Materials Mechanical Properties.
A Novel Balanced Charged Alginate/PEI Polyelectrolytes Hydrogel Carbohydr. Polym. 2005, 60 (2), 185−192.
that Resists Foreign-Body Reaction. ACS Appl. Mater. Interfaces 2018, (41) Sreedhar, B.; Chattopadhyay, D. K.; Karunakar, M. S. H.;
10, 6879−6886. Sastry, A. R. K. Thermal and Surface Characterization of Plasticized
(23) González-domínguez, J. M.; Martín, C.; Durá, O. J.; Merino, S.; Starch Polyvinyl Alcohol Blends Crosslinked with Epichlorohydrin. J.
Vázquez, E. Smart Hybrid Graphene Hydrogels: A Study of the Appl. Polym. Sci. 2006, 101, 25−34.
Different Responses to Mechanical Stretching Stimulus. ACS Appl. (42) Sreedhar, B.; Sairam, M.; Chattopadhyay, D. K.; Rathnam, P. A.
Mater. Interfaces 2018, 10, 1987−1995. S.; Rao, D. V. M. Thermal, Mechanical, and Surface Characterization
(24) Jeon, O.; Shin, J. Y.; Marks, R.; Hopkins, M.; Kim, T. H.; Park, of Starch−Poly(vinyl alcohol) Blends and Borax-Crosslinked Films. J.
H. H.; Alsberg, E. Highly Elastic and Tough Interpenetrating polymer Appl. Polym. Sci. 2005, 96, 1313−1322.
Network-Structured Hybrid Hydrogels for Cyclic Mechanical (43) Bursali, E. A.; Coskun, S.; Kizil, M.; Yurdakoc, M. Synthesis,
Loading-Enhanced Tissue Engineering. Chem. Mater. 2017, 29 (19), Characterization and In Vitro, Antimicrobial Activities of Boron/
8425−8432. Starch/Polyvinyl Alcohol Hydrogels. Carbohydr. Polym. 2011, 83 (3),
(25) Gong, J. P.; Katsuyama, Y.; Kurokawa, T.; Osada, Y. Double- 1377−1383.
Network Hydrogels with Extremely High Mechanical Strength. Adv. (44) Zhang, T. T.; Zuo, T.; Hu, D. N.; Chang, C. Y. Dual physically
Mater. 2003, 15, 1155−1158. Crosslinked Nanocomposite Hydrogels Reinforced by Tunicate
(26) Gong, J. P. Materials both Tough and Soft. Science 2014, 344, Cellulose Nanocrystals with High Toughness and Good Self-
161−162. Recoverability. ACS Appl. Mater. Interfaces 2017, 9, 24230−24237.
(27) Chen, Q.; Zhu, L.; Zhao, C.; Wang, Q. M.; Zheng, J. A Robust, (45) Yu, H. C.; Zhang, H.; Ren, K. F.; Ying, Z. M.; Zhu, F. B.; Qian,
One-Pot Synthesis of Highly Mechanical and Recoverable Double J.; Ji, J.; Wu, Z. L.; Zheng, Q. Ultrathin κ-Carrageenan/Chitosan

3979 DOI: 10.1021/acs.jafc.8b05147


J. Agric. Food Chem. 2019, 67, 3966−3980
Journal of Agricultural and Food Chemistry Article

Hydrogel Films with High Toughness and Antiadhesion Property. (64) Feng, W.; Zhou, W.; Dai, Z.; Yasin, A.; Yang, H. Tough
ACS Appl. Mater. Interfaces 2018, 10, 9002−9009. Polypseudorotaxane Supramolecular Hydrogels with Dual-Responsive
(46) Yang, J.; Xu, F.; Han, C. R. Metal Ion Mediated Cellulose Shape Memory Properties. J. Mater. Chem. B 2016, 4 (11), 1924−
Nanofibrils Transient Network in Covalently Cross-linked Hydrogels: 1931.
Mechanistic Insight into Morphology and Dynamics. Biomacromole- (65) Oyen, M. L. Mechanical Characterisation of Hydrogel
cules 2017, 18 (3), 1019−1028. Materials. Int. Mater. Rev. 2014, 59 (1), 44−59.
(47) Chen, H.; Liu, Y.; Ren, B.; Zhang, Y.; Ma, J.; Xu, L. J.; Chen, (66) Wang, J.; Liu, F.; Tao, F.; Pan, Q. Rationally Designed Self-
Q.; Zheng, J. Super Bulk and Interfacial Toughness of Physically Healing Hydrogel Electrolyte toward a Smart and Sustainable
Crosslinked Double-Network Hydrogels. Adv. Funct. Mater. 2017, 27, Supercapacitor. ACS Appl. Mater. Interfaces 2017, 9, 27745−27753.
1703086. (67) Shao, C.; Chang, H.; Wang, M.; Xu, F.; Yang, J. High-Strength,
(48) Ji, N.; Qin, Y.; Li, M.; Xiong, L.; Qiu, L. Z.; Bian, X. L.; Sun, Q. Tough, and Self-Healing Nanocomposite Physical Hydrogels Based
J. Fabrication and Characterization of Starch Nanohydrogels via on the Synergistic Effects of Dynamic Hydrogen Bond and Dual
Reverse Emulsification and Internal Gelation. J. Agric. Food Chem. Coordination Bonds. ACS Appl. Mater. Interfaces 2017, 9, 28305−
2018, 66 (35), 9326−9334. 28318.
(49) Torres, O.; Tena, N. M.; Murray, B.; Sarkar, A. Novel Starch (68) Sabzi, M.; Samadi, N.; Abbasi, F.; Mahdavinia, G. R.;
Based Emulsion Gels and Emulsion Microgel Particles: Design, Babaahmadi, M. Bioinspired fully physically Cross-Linked Double
Structure and Rheology. Carbohydr. Polym. 2017, 178, 86−94. Network Hydrogels with a Robust, Tough and Self-Healing Structure.
(50) Chen, Y. N.; Peng, L. F.; Liu, T. Q.; Wang, Y. X.; Shi, S. J.; Mater. Sci. Eng., C 2017, 74, 374−381.
Wang, H. L. Poly (vinyl alcohol)-Tannic Acid Hydrogels with (69) Li, X. F.; Zhao, Y.; Li, D.; Zhang, G.; Long, S.; Wang, H.
Excellent Mechanical Properties and Shape Memory Behaviors. ACS Hybrid Dual Crosslinked Polyacrylic Acid Hydrogels with Ultrahigh
Appl. Mater. Interfaces 2016, 8, 27199−27206. Mechanical Strength, Toughness and Self-Healing Properties via
(51) Lu, B.; Lin, F.; Jiang, X.; Cheng, J.; Lu, Q.; Song, J.; Chen, C.; Soaking Salt Solution. Polymer 2017, 121, 55−63.
Huang, B. One-Pot Assembly of Microfibrillated Cellulose Reinforced (70) Bian, H.; Wei, L.; Lin, C.; Ma, Q.; Dai, H.; Zhu, J. Y. Lignin
PVA-Borax Hydrogels with Self-Healing and pH-Responsive Proper- Containing Cellulose Nanofibril-Reinforced Polyvinyl Alcohol Hydro-
ties. ACS Sustainable Chem. Eng. 2017, 5, 948−956. gels. ACS Sustainable Chem. Eng. 2018, 6 (4), 4821−4828.
(52) Narita, T.; Indei, T. Microrheological Study of Physical (71) Hu, T.; Liu, Q. Z.; Gao, T. T.; Dong, K. J.; Wei, G.; Yao, J. S.
Gelation in Living Polymeric Networks. Macromolecules 2016, 49 Facile Preparation of Tannic Acid−Poly(vinyl alcohol)/Sodium
(12), 4634−4646. Alginate Hydrogel Beads for Methylene Blue Removal from Simulated
(53) Sun, Q. J.; Li, G. H.; Dai, L.; Ji, N.; Xiong, L. Green Preparation Solution. ACS Omega 2018, 3, 7523−7531.
and Characterisation of Waxy Maize Starch Nanoparticles Through (72) Dai, H.; Huang, Y.; Huang, H. Eco-Friendly Polyvinyl Alcohol/
Enzymolysis and Recrystallisation. Food Chem. 2014, 162 (11), 223− Carboxymethyl Cellulose Hydrogels Reinforced with Graphene Oxide
228. and Bentonite for Enhanced Adsorption of Methylene Blue.
(54) Qin, Y.; Liu, C. Z.; Jiang, S. S.; Xiong, L.; Sun, Q. J. Carbohydr. Polym. 2018, 185, 1−11.
Characterization of Starch Nanoparticles Prepared by Nanoprecipi-
tation: Influence of Amylose Content and Starch Type. Ind. Crops
Prod. 2016, 87, 182−190.
(55) Taleb, M. F. A.; El-Mohdy, H. A.; El-Rehim, H. A. Radiation
Preparation of PVA/CMC Copolymers and Their Application in
Removal of Dyes. J. Hazard. Mater. 2009, 168, 68−75.
(56) Manna, U.; Patil, S. Borax Mediated Layer-by-Layer Self-
Assembly of Neutral Poly(vinyl alcohol) and Chitosan. J. Phys. Chem.
B 2009, 113, 9137−9142.
(57) Han, J. Q.; Yue, Y. Y.; Wu, Q. L.; Huang, C. B.; Pan, H.; Zhan,
X. X.; Mei, C. T.; Xu, X. W. Effects of Nanocellulose on the Structure
and Properties of Poly(vinyl alcohol)-Borax Hybrid Foams. Cellulose
2017, 24, 4433−4448.
(58) Liu, G. D.; Gu, Z. B.; Hong, Y.; Cheng, L.; Li, C. M. Structure,
Functionality and Applications of Debranched Starch: A Review.
Trends Food Sci. Technol. 2017, 63, 70−79.
(59) Peng, M.; Xiao, G.; Tang, X.; Zhou, Y. Hydrogen-Bonding
Assembly of Rigid-Rod Poly(p-sulfophenylene terephthalamide) and
Flexible-Chain Poly(vinyl alcohol) for Transparent, Strong, and
Tough Molecular Composites. Macromolecules 2014, 47, 8411−8419.
(60) Zhao, Y.; Nakajima, T.; Yang, J. J.; Kurokawa, T.; Liu, J.; Lu, J.;
Mizumoto, S.; Sugahara, K.; Kitamura, N.; Yasuda, K.; Daniels, A. U.
D.; Gong, J. P. Proteoglycans and Glycosaminoglycans Improve
Toughness of Biocompatible Double Network Hydrogels. Adv. Mater.
2014, 26, 436−442.
(61) Zhao, D.; Huang, J.; Zhong, Y.; Li, K.; Zhang, L.; Cai, J. High
Strength and High-Toughness Double-Cross-Linked Cellulose
Hydrogels: A New Strategy Using Sequential Chemical and Physical
Cross-Linking. Adv. Funct. Mater. 2016, 26, 6279−6287.
(62) Hu, X. Y.; Feng, L. D.; Xie, A. M.; Wei, W.; Wang, S. M.;
Zhang, J. F.; Dong, W. Synthesis and Characterization of a Novel
Hydrogel: Salecan/Polyacrylamide Semi-IPN Hydrogel with a
Desirable Pore Structure. J. Mater. Chem. B 2014, 2, 3646−3658.
(63) Zhu, B. D.; Ma, D. Z.; Wang, J.; Zhang, S. Structure and
Properties of Semi-Interpenetrating Network Hydrogel Based on
Starch. Carbohydr. Polym. 2015, 133, 448−455.

3980 DOI: 10.1021/acs.jafc.8b05147


J. Agric. Food Chem. 2019, 67, 3966−3980

You might also like