You are on page 1of 19

nature reviews materials https://doi.org/10.

1038/s41578-024-00672-3

Review article Check for updates

Design principles for strong


and tough hydrogels
Xueyu Li 1
& Jian Ping Gong 1,2

Abstract Sections

Hydrogels are crosslinked polymer networks swollen with water. Introduction

Owing to their soft and water-containing nature, hydrogels are Model of basic hydrogel
promising materials for applications in many fields, such as biomedical mechanics

engineering, soft robotics and environmental studies. One of the Design strategies
main obstacles to the practical application of hydrogels is their Conclusions and outlook
low mechanical strength and toughness. Since the 2000s, many
breakthroughs in the development of mechanically strong and tough
hydrogels have led to enormous advances in the study of soft materials
and our understanding of their failure mechanisms. Research has also
been conducted on long-term mechanical stability — that is, the cyclic
fatigue resistance and self-strengthening properties of hydrogels — to
enable their application as load-bearing materials. This Review provides
a comprehensive overview of the design principles for tough hydrogels.
Strategies to obtain self-growing and reinforced hydrogels that can
adapt to their surrounding mechanical environment are also presented.

Laboratory of Soft and Wet Matter, Faculty of Advanced Life Science, Hokkaido University, Sapporo, Japan.
1

Institute for Chemical Reaction Design and Discovery (WPI-ICReDD), Hokkaido University, Sapporo, Japan.
2

e-mail: lixueyu@sci.hokudai.ac.jp; gong@sci.hokudai.ac.jp

Nature Reviews Materials


Review article

Introduction been proposed. For elastic gels, these strategies involve modulations
As a combination of crosslinked solid and liquid components, hydro- at the molecular level, such as lengthening the polymer chain 49,50,
gels are similar to soft tissues in the human body and are promising increasing entanglement26,27,51,52, and unfolding or degrading the
materials for uses such as scaffolds in tissue engineering1, medical crosslinker53–57. For viscoelastic gels, mesoscale modifications have
implants2 or wound dressings3. Many applications require hydrogels to been proposed, such as microphase separation42,58–60, microcrystalliza-
bear mechanical loads and to resist failure under static or cyclic load- tion and fibrils18,47,61–65, and nanocomposites66–68. The fatigue threshold
ing conditions. The macroscopic failure of materials originates from is greatly enhanced for gels with hierarchical structures69, resolving the
the growth of small defects. The ability to resist defect growth — the conflict between the modulus and the fatigue threshold.
toughness — is correlated to the fracture energy Γ of materials, which After about two decades of effort, nowadays the fracture energy
is the energy required for crack growth per unit area4. Typically, load- and fatigue threshold of hydrogels can reach ~200 kJ m−2 (refs. 16,44) and
bearing biological tissues have a high toughness. For instance, typical ~10 kJ m−2 (ref. 18), respectively. Elastic modulus from submegapascal
fracture energies are ~2.5 kJ m−2 for skeletal muscle, ~1 kJ m−2 for carti- to hundreds of megapascals, strength from submegapascal to tens of
lage and 20–30 kJ m−2 for tendons5. However, conventional synthetic megapascals, and stretching ratio from several to hundreds70,71 can be
hydrogels, which are composed of a single non-uniform network of achieved. Moreover, taking inspiration from biological systems, efforts
hydrophilic polymers, are in general very weak, with low fracture energy have been made to develop self-healing hydrogels and anisotropic hydro-
(<10 J m−2)6–8. This poor mechanical performance greatly limits the gels. Unlike other solid materials, hydrogels are permeable to small
application of such hydrogels. In addition to the toughness, elastic molecules. Thus, hydrogels could be used as an open system to develop
modulus E (stiffness), fracture strength σf (strength), fracture stretch mechanically triggered new network growth. Self-growing and strength-
ratio λf (deformability) and fatigue threshold Γ0 are important mechani- ening hydrogels based on mechanochemistry mechanisms have also
cal parameters for applications. For a simple polymer network material, attracted attention for the purpose of developing materials that adapt
these mechanical properties are correlated and conflicting. Increases to their surrounding environment, resembling biological systems72,73.
in elastic modulus and strength result in decreases in deformability, The advancements in hydrogel design strategies, focusing on
fracture energy and fatigue threshold. Moreover, as hydrogels are achieving comprehensive mechanical performance, have been summa-
polymer networks swollen in water (solvent), these mechanical param- rized in several outstanding review papers, particularly those empha-
eters are strongly related to the swelling ratio λs of the hydrogels. At the sizing the structure of polymer networks5,74 and the nonlinear elastic
equilibrium swelling, the swelling ratio is determined by the balance fracture mechanics75. Given the importance of swelling and deswelling
of osmotic pressure and network elasticity9,10. Therefore, designing characteristics in hydrogels, and the practical need for hydrogels to
and developing hydrogels with high toughness without sacrificing the be in an equilibrium state of swelling in a liquid medium, this Review
modulus and strength is a challenge. explores design strategies to attain high mechanical performance with
The poor mechanical performance of conventional hydrogels is a focus on the perspective of swelling and deswelling. We initially use a
attributed to several intrinsic features. One is the network inhomo- basic hydrogel model to clarify the influences of the molecular structure
geneity in polymer density distribution and polymer strand length and swelling or deswelling on the mechanical properties, encompass-
between the crosslinking points. Thus, hydrogels are susceptible ing the elastic modulus, extensibility, strength, toughness and fatigue
to stress concentration at loading, initiating cracks. Another is the resistance. We then discuss the main design strategies to obtain tough
rubber-like elasticity caused by a lack of energy dissipation mecha- hydrogels, followed by advances in developing fatigue-resistant hydro-
nisms, resulting in low resistance against crack propagation. In addi- gels and in understanding their underlying mechanisms. We highlight
tion, conventional hydrogels usually contain abundant water. The recently developed strategies for developing self-reinforcement hydro-
low amount of load-bearing solid phase further results in weak and gels with tissue-like self-growing properties. To close, we emphasize
fragile mechanical properties. Since the 2000s, the development of challenges and trends in developing inelastic fracture mechanics theory
mechanically strong and tough hydrogels11–18 has led to important to capture the large deformation behaviour and the next generation of
advances in the study of soft materials and our understanding of their tough hydrogels for practical biological applications.
failure mechanisms. The main strategies to improve the mechanical
properties of hydrogels through their structural components can Model of basic hydrogel mechanics
be classified into three categories: the design of topological struc- Various theoretical models, including the classic Flory–Rehner statis-
tures, such as slide-ring gels14,15,19–21, homogeneous four-arm gels22–25 tical model9, Arruda–Boyce eight-chain constitutive model76 and Gent
and highly entangled gels26–28, to distribute stress in single-network continuum mechanics model77, along with their combinations10,78, have
systems; the introduction of energy dissipation mechanisms by sac- been explored to explain the swelling and large deformation of polymer
rificial bonds, such as in double-network (DN) hydrogels11,29–33 and networks. These models are discussed in reviews elsewhere79–82. Here,
dual-crosslinked hydrogels13,34–39; and the introduction of high-order we analyse a simple affine network to represent the structure of a hydro-
structures, such as microphase separations16,40–43, microcrystals, and gel, and we discuss its rubber elasticity and fracture at different swell-
fibrils or fabrics15,44–48. From the perspective of mechanical dynamics, ing states. The hydrogel features a uniform polymer network
they can be classified into elastic and viscoelastic hydrogels that are comprising strands characterized by the number of Kuhn monomers
strain-rate independent and dependent, respectively, in the common in each strand (Nx) and the length of each monomer (b) (Fig. 1a). In the
observation window. reference state, the end-to-end distance of each strand (R0) follows an
Fatigue resistance — that is, long-term stability under cyclic loads — ideal Gaussian chain, approximately given by R0 ≃ bNx 1/2, and the net-
is vital for some applications such as artificial cartilage. Since the ini- work has a strand density denoted as νx,0 (refs. 76,83). On contact with
tial work in 2017 (ref. 6), the study of fatigue-resistant mechanisms water, the hydrogel undergoes a size change by a factor of λs in length
and the development of fatigue-resistant hydrogels have attracted (λ s > 1 indicates swelling, and λ s < 1 indicates deswelling). Operating as
considerable interest. Numerous antifatigue design strategies have an affine network, where the deformations of the bulk hydrogel and

Nature Reviews Materials


Review article

individual network strands are identical83,84, the end-to-end distance the maximum stretch ratio of a single strand, which goes from λ sR0 to
of a strand becomes R = λ sR0. Consequently, the strand density trans- its physical limit or contour length (Rmax = bNx), is
−3
forms to νx,0λ s (Fig. 1a). Subsequently, we investigate the influence
of λs on the elastic modulus, stretchability and fracture characteristics of Rmax Nx 1/2
λ max = ≃ (3)
the hydrogel in the absence of viscoelastic effect, unless specified. λ s R0 λs

Elastic modulus The maximum engineering stress, which is the product of a single
The Young’s modulus (E) of a hydrogel is determined by the product strand’s rupture force ( fb) and the areal density of strands crossing the
of the elasticity (k) or stiffness per strand and the density of strands plane perpendicular to the loading direction in the undeformed state
−3 −3 3
(νx,0λ s ), expressed as E = kνx,0λ s . As the elasticity of a strand increases (Σ ≃ λ sR0νx ,0 / λ s ), is given by
with swelling (λs > 1) and strand number per volume decreases with
swelling, E exhibits a non-monotonic dependence on λs. In this context, fb bC0
σmax ≃ 1/2 (4)
we use the freely jointed chain model to represent the elastic energy of 2
λ s Nx
a strand, considering the finite extensibility effect (see Supplementary
Information)85,86. By taking the second derivative of the elastic energy Here, the monomer concentration in the as-synthesized state is
of a strand with respect to the stretching ratio, we derive the elasticity of C0 = νx ,0Nx . From equations (3) and (4), the true maximum stress
the strand, expressed as: (σmax T = σmaxλ max ) of the simple hydrogel is proportional to the bond
3
2 3 + α4 rupture force and bond density of the network (C0 /λ s ), independent
k = kBTλ s (1) of Nx.
(1 − α 2) 2
f bC0
where α = R
λ R
s 0 λ
≃ 1/2
s
. Consequently, the Young’s modulus of the hydrogel σmax T ≃ b 3 (5)
max Nx λs
is given by:
ve,0kBT 3 + α 4 Thus, the stretchability and strength of a hydrogel decrease by
E= (2) swelling.
λ s (1 − α 2) 2
λ sR 0 R
In Fig. 1a, we depict E/E0 as a function of α = R max for Rmax ≃ Nx 1/2 = Stress concentration effect
0
20 as a representative example. Here E0 = 3ve,0kBT is the Young’s modu- For real hydrogels, as for other brittle materials, the theoretical σmax T
lus of the hydrogel at the reference state (λ s = 1, α ≃ 0). The curve in has never been observed macroscopically because of stress concentra-
Fig. 1a illustrates the non-monotonic evolution of the elastic modulus tion at defects or crack tips. For hydrogels for which nonlinear elastic
with swelling ratio. For small or modest λ s, where λ s R0 is much smaller fracture mechanics apply, two characteristic lengths, the nonlinear
than the contour length of the strand Rmax = bNx, polymer strands can elastic length (ℓ, denoted as the distance from the crack tip below which
be approximated as Gaussian chains, and the elasticity is proportional the deformation is dominated by elastic nonlinearity at the onset of
2
to λ s . Therefore, E decreases as λ s increases because the strand density crack initiation)75 and defect-insensitive length (ξ, denoted as the
−3
is proportional to λ s . For large λ s where λ s R0 approaches Rmax, the critical crack length for the transition from defect-insensitive to defect-
stiffness of strands increases rapidly owing to the limited extensibility sensitive rupture)89 or dissipative length, are important75. The true
of the strand85. The increase of strand stiffness with λ s is much stronger tensile stress σ T around the crack tip inversely scales with the distance
2
than λ s , surpassing the effect from the decrease of strand density and from the crack tip (r), provided that the length of a defect or crack (c)
resulting in an increase of E with λ s. exceeds a certain nonlinear elastic length ℓ , and r is in the range of
For a deswelling hydrogel (λ s < 1), when there are physical interac- ξ < r < ℓ < c (refs. 75,84) (Fig. 1b).
tions between polymer strands, the elastic modulus comprises two
Γ
̇ where Epr is from the primary network and increases
terms, E = Epr + E(ε), σ T (r ) ≃ (when ξ < r < ℓ < c) (6)
r
with deswelling (Fig. 1a, regime C), and E(ε)̇ is from dynamic crosslink-
ing of interstrand physical bonds and changes with the observation Here, the fracture energy Γ represents the energy required to cre-
time (or strain rate ε)̇ relative to the characteristic relaxation time (τ) ate a unit surface by propagating a pre-existing crack. The true stress
of the viscoelastic hydrogel. At low strain rate (ετ̇ ≪ 1), the dynamic converges to the external stress applied to the material when the dis-
bonds have little contribution (E(ε)̇ ≃ 0) to the modulus, and the hydro- tance from the crack is much larger than the crack size (r ≫ c). There-
gel behaves as a soft solid with elastic modulus Epr, like the elastic fore, even if the applied stress is still far below the theoretical maximum
hydrogel. At high strain rate (ετ̇ ≫ 1), the dynamic bonds play a similar stress σmax T, the σ T(r ) near a defect could already reach σmax T, resulting
role to the permanent crosslinking, and the hydrogel behaves as a hard in material failure. Owing to this stress concentration effect at the crack
̇ In the intermediate strain rate, the modu-
elastic solid with a large E(ε). tip, the elastic energy per unit volume that a soft material can store
lus increases as the strain rate increases (viscoelastic regime). The mod- before undergoing rupture (Wb) decreases inversely proportional to c,
ulus of viscoelastic hydrogels typically follows the time–temperature provided that c exceeds ξ (ref. 75) (Fig. 1c).
superposition principle, as the bond association time is strongly
Γ
temperature dependent and influences τ (refs. 83,87,88). Wb ≃ (when c > ξ ) (7)
c
Maximum stress and stretch ratio For an elastic hydrogel, Wb corresponds to the area under the
The theoretical maximum stretching ratio ( λ max) and engineering stress stress–strain curve of the material. Thus, Wb is not an intrinsic material
(σmax) strongly depend on the swelling behaviour. Here we use a simple parameter and depends on the pre-existing crack length c. Equation (7)
1D model to see the effect of λ s on λ max and σmax. Under deformation, explains why in real hydrogels, in which defects inevitably exist, the

Nature Reviews Materials


Review article

a Hydrogel swelling behaviour and its effect on elastic modulus b True stress as a function of distance r to the crack tip
Г
~
r Crack
Crosslink σT

1
ln(σT)

R0
Swell (λs > 1) Deswell (λs < 1) Applied σT

Elastic Viscoelastic

ln(ξ) ln(l) ln(c)


ln(r)
Dynamic
bonds
c Critical strain-energy density at crack initiation as a function
of crack length c
b
λsR0 Г/c
σ
Nx monomers W*
λsR0
c
λs < 1 λs > 1
3 Wb
C
B

1
E/E0

A Flaw Flaw
insensitive sensitive
0.3 ξ

Crack length c
Nx = 400
0.1
0.03 0.1 0.3 1
λs/Nx½

d Fracture energy of hydrogels Polymer chain lying across the crack propagation plane
Г0: Lake–Thomas model Nx monomer units Crack
σ

λcH0 ГD Crack growth λsR0


Plane of crack
propagation

Crosslink
Г = Г0 + ГD

λ
~10 nm ~0.1–1 mm ~1–10 mm ~1 mm ~10 mm
ГD ∝ dissipation lengths Simple network Double network Dual-crosslinked Strain-induced Fibre reinforcement
viscoelastic network crystallization

Nature Reviews Materials


Review article

Fig. 1 | Hydrogel swelling, deswelling and fracture behaviours. a, Schematic of crack (defect) length c. W* is the critical strain-energy density for rupturing a
depiction of hydrogel swelling behaviour and its influence on the elastic material in the absence of large cracks of c > ξ. d, Fracture energy Γ of hydrogels.
modulus. The top panel shows an affine network hydrogel in its reference state. H0 is the initial height. λc is the critical stretch ratio where the pre-existing crack
In the middle panels, the hydrogel undergoes swelling or deswelling by a length starts propagation. In general, Γ = ΓD + Γ0, where ΓD is correlated to the energy
factor of λs. The bottom panel shows the elastic modulus (E) of the hydrogel as a dissipation in the processing zone and increases with dissipation length ξ, and Γ0
function of λs, where E is normalized by its value at the reference state (E0), and λs is intrinsic fracture energy, correlated to the energy required to break polymer
is normalized by the maximum allowable value ( Nx1/2). In practice, hydrogels chains lying across the crack plane by a unit area. For a simple network, Γ = Γ0 and
exhibit distinct behaviours: typically, a neutral hydrogel swells modestly, and λs is Γ0 can be described by the Lake–Thomas model. The typical processing zone sizes
in the regime where E/E0 decreases (regime A); a strong polyelectrolyte hydrogel (or dissipation length) are ~10 nm, ~0.1–1 mm, ~1–10 mm, ~1 mm and ~10 mm for
swells substantially, and λs can access the regime where E/E0 increases (regime B); simple network hydrogels, double-network hydrogels, dual-crosslinked
and a hydrogel containing dynamic bonds deswells in water, and E/E0 is greater viscoelastic hydrogels, strain-induced crystallization hydrogels and fibre-
than 1 (regime C). b, True stress distribution along the principal axis as a function reinforced composites, respectively75. Panels b,c and d adapted with permission
of distance r to the crack tip for soft materials, based on nonlinear elastic fracture from ref. 75, Annual Reviews. Panel d adapted with permission from ref. 69, Royal
mechanics theory. ξ and l are the dissipative length and nonlinear elastic length, Society of London.
respectively. c, Critical strain-energy density Wb at crack initiation as a function

fracture stretch ratio ( λ f ) and engineering stress σf vary from sample pre-existing crack will extend when the rate of strain-energy release
to sample and are much smaller than λ max and σmax predicted from their from the stress field around the crack is at least equal to the rate at
average structure. which it is absorbed by crack extension. Commonly used experimental
setups for measuring the fracture energy for the crack initiation (the
Fracture energy onset of crack propagation) can be found in the literature4,12,92,93.
The fracture energy Γ is an intrinsic material parameter for charac-
terizing a material’s toughness. For the swelling elastic hydrogel, the Design strategies
Lake–Thomas model90 provides a description for Γ as the energy needed In practice, hydrogels typically undergo swelling when immersed in
to break a single layer of strands intersecting the crack plane: water after synthesis. The reference state of the model hydrogel
described in the theoretical section may correspond to the as-synthe-
2
Γ ≃ NxUbΣ ≃ bUbC0Nx 1/2/λ s (8) sized hydrogel. The swelling ratio λ s of a hydrogel in the equilibrium
state is governed by the equilibrium between the osmotic pressure
Here, Ub represents the activation energy required to break a arising from the polymer–solvent (water) mixing free energy and the
chemical bond in the backbone of strands. The Lake–Thomas model network elasticity of the hydrogels, which can be described by
postulates that the energy needed to break an individual strand is dir­ the Flory−Rehner theory9.
ectly proportional to the monomer number of that particular strand Generally, hydrogels made from neutral polymers display mod-
(Nx). This proportionality arises because all the bonds in the strand back- est swelling, leading to a decrease in the elastic modulus on swelling
bone are arranged in series, necessitating the stretching of each bond (Fig. 1a, regime A). By contrast, polyelectrolyte hydrogels, where ionic
to the same energy state (Ub) to induce bond rupture (Fig. 1d). By insert- osmotic pressure from Donnan equilibrium contributes to excessive
ing equations (5) and (8) into equation (6), the defect-insensitivity swelling in pure water, result in an increase in E with swelling94,95 (Fig. 1a,
length or energy dissipative length ξ for the model elastic hydrogel is regime B). When physical bonds can be formed between polymer
found to be of the order of the end-to-end distance of the polymer strands, hydrogels undergo deswelling after synthesis39,96,97.
U
strands (the network mesh size), ξ ≃ bfb λ sR0 ≃ R. For a swelling gel (λ s > 1 ), the equilibrium λ s increases as Nx
b
To bolster a material’s strength, reducing the stress concentration increases83,94,98, whereas for a deswelling gel (λ s < 1), the equilibrium λ s
is crucial, as underscored by equations (6) and (7). Increasing ξ, reduc- is barely dependent on Nx (refs. 41,42). Nx decreases as chemical
ing the size of defects or cracks (c), and increasing the energy required crosslinker density increases. Swelling hydrogels typically exhibit
for crack propagation (Γ) are potential strategies. For improving Г, the elastic behaviour with negligible interstrand interaction. On the other
central idea is to incorporate additional energy-dissipating mecha- hand, deswelling hydrogels usually exhibit viscoelastic behaviour due
nism from a processing zone at the crack tip that extends beyond to physical bonds among polymer strands (Fig. 1a). It is important to
the network mesh size. The total fracture energy can be expressed note that there are exceptions to these patterns99,100. This Review
as the sum of two components: Г = Г0 + ГD, where Γ0 represents the focuses on discussing these two typical patterns.
dissipation directly related to breaking the polymer strands bridging
the crack (the intrinsic fracture energy) and can be described by the Elastic hydrogels (typically λs > 1)
Lake–Thomas model (Fig. 1d), and ΓD accounts for the contribution In real hydrogels, there are inherent variations in both the distribution
of mechanical energy dissipation from the process zone. The process of polymer density and the length of a polymer strand (Fig. 2a). There-
zone also increases the defect-insensitivity length ξ. Different molecu- fore, hydrogels naturally contain numerous defects. A single, relatively
lar designs can substantially increase the crack-tip processing zone large defect produces remarkable stress concentration and triggers
(Fig. 1d). For a deswelling hydrogel that is dual crosslinked, the vis- crack propagation, resulting in the failure of hydrogels at macroscopic
coelasticity resulting from the dynamic bonds brings intrinsic energy stress levels that are much lower than theoretical averages would sug-
dissipation mechanisms. gest (equations (6) and (7)). Because elastic hydrogels have no energy
The Γ of elastic gels can be experimentally measured based on dissipation mechanism during deformation, real swelling hydrogels are
Griffith’s theory for a brittle fracture91. The theory postulates that a typically very weak, and the failure behaviour is hard to predict, owing

Nature Reviews Materials


Review article

a Conventional hydrogel b Topological design for tough hydrogel


Force-triggered
Ideal network Slide-ring Dense entanglement cycloreversion

c Double-network hydrogel
First network: rigid and brittle Second network: flexible Double network: tough and robust
and stretchable

No interaction between first network and second network Interaction between first network and second network

Chemical crosslinking: swollen (elastic) Deswelling in water (viscoelastic)


Load
First network Second network
Unload

First network Load


Reversible crosslinking: usually used
at as-prepared state (viscoelastic)
Load Unload

Dynamic bonds
Unload
Fig. 2 | Swelling (elastic) hydrogels. a, Structure of conventional hydrogel. b, Topological structure hydrogels. c, Double-network hydrogels. Panel b adapted with
permission from ref. 26, AAAS, and with permission from ref. 108, AAAS. Panel c adapted from ref. 200, Springer Nature Limited.

to their heightened vulnerability to defects. Enhancing the toughness between these quantities. Hydrogel swelling (λs) reduces Γ, σmax and
of elastic hydrogels is a definite challenge. λmax. Hence, simultaneously optimizing these properties in purely
elastic hydrogels necessitates molecular designs that go beyond the
Topological structure. Reducing spatial heterogeneities through the constraints of a basic elastic network. For example, slide-ring hydro-
realization of an ideally homogeneous network, movable crosslinkers gels14 can equalize the tension of polymer chains. In these hydrogels,
or scission mechanisms can improve stress homogenization (Fig. 2b). the polymer chains are topologically interlocked by figure-of-eight
Hydrogels made from well-defined symmetrical tetrahedron-like poly- polyrotaxane crosslinks, which can pass along the polymer chains
ethylene glycol (PEG) macromonomers22–25 have fewer spatial heteroge- freely in a manner similar to pulleys. As a result, the strand length
neities101–103 than hydrogels obtained from free radical polymerization between two crosslinked junctions (Nx) increases as the figure-of-
and thus show improved failure strength and extensibility due to a eight crosslinks slide away from each other and redistribute loads19,90,
lower amount of defects22,104. However, the fracture energy of such gels resulting in a higher extensibility and stronger crack growth resistance
lies in the range estimated from the Lake–Thomas model (<100 J m−2)105, compared with fixed crosslink gels106. Another strategy to transmit
because the network homogeneity improves the strength but not tension around the notch consists in applying dense entanglements
the fracture energy, consistent with the theorem for simple elastic as slip links to fabricate single-network polyacrylamide (PAAm)
network fracture. hydrogels — with polymer volume fraction around 10 vol% — that are
For a purely elastic hydrogel, a large Nx (which can be achieved stiff (E = 0.1 MPa) and stretchable (fracture stretch ratio λf = 4.5) and
by reducing the chemical crosslinking density) results in increases have superior fracture energy compared with conventional PAAm
in stretchability λmax and fracture energy Γ, but decreases in elas- hydrogels of similar modulus (Г up to 2 kJ m−2 compared with 10−2
tic modulus E and strength σmax. Therefore, there exists a tradeoff to 10−1 kJ m−2)8,26,107. The large number of trapped entanglements

Nature Reviews Materials


Review article

contribute to the modulus and strength, whereas the slip of entan- ruptured first network causes large softening and stress–strain hys-
glements at the crack tip lengthens the strands, resulting in large teresis, contributing to a substantial ГD (refs. 115–118) (Fig. 1d). This
fracture energy simultaneously. Weak side-crosslinker scission is internal fracture phenomenon has been confirmed by luminescent
another strategy, in which the side crosslinkers are preferentially probes31,119–122 and real-time birefringence123,124. Furthermore, the bro-
cleaved under force, resulting in lengthening of the strands. This ken fragments of the first network could serve as sliding crosslinks
strategy can enhance the fracture energy ninefold without sacrificing to further delocalize the stress concentration around the crack tip
the modulus108. To demonstrate the toughening effect of hydrogels and prevent chain scissions125. Therefore, the combination of the two
fabricated through different strategies, we summarize the mechani- networks enhances resistance to crack propagation, resulting in high
cal performances and corresponding water content of various types fracture energy comparable to cartilage and tendon126,127.
of hydrogels (Table 1). We can see that the tradeoff between modulus
and fracture energy can be resolved by movable or weak crosslinker Viscoelastic hydrogels (typically λs < 1)
mechanisms (Table 1). Hydrogels with topological designs have the Viscoelastic hydrogels have an intrinsic energy dissipation mecha-
advantage of negligible hysteresis, which can avoid stress soften- nism, and they typically exhibit a large, self-recoverable mechanical
ing and relaxation during practical applications. Because they have hysteresis because the physical bonds dynamically break and reform.
no built-in energy dissipation mechanisms, the fracture energy of The mechanical properties strongly depend on the structure (dynamic
such systems, in principle, can be described by a modified Lake– bond strength and density) and the observation condition (strain rate
Thomas model in which multiple network layers near the crack plane and temperature). The fracture energy strongly depends on the crack
participate in dissipating energy109,110. propagation velocity (vc) and temperature (T), and can be expressed as
Γ(vc,T) = Γ0(1 + f(vc,T)) (ref. 87). When vc approaches zero, the fracture
Double-network structure. Swelling hydrogels are elastic and energy reduces to Γ0, which is determined by the primary network
thus do not dissipate energy during deformation. The invention of structure (equation (8)).
the DN structure with built-in sacrificial bonds and energy dissipa-
tion mechanisms was a breakthrough for that class of hydrogels11. Dual-crosslinked hydrogels. Because deswelling hydrogels typically
DN hydrogels consist of two interpenetrating elastic networks with have two crosslinking mechanisms (covalent and non-covalent), they
contrasting properties (Fig. 2c) obtained by a two-step sequential are also known as dual-crosslinked hydrogels. The non-covalent bonds
polymerization process. The first network, which is rigid and brit- can dissociate and reassociate dynamically, dissipating energy and pro-
tle, acts as a sacrificial bond network that effectively dissipates tecting the primary network crosslinked by chemical bonds from stress
energy. Meanwhile, the second network is soft and ductile, ensur- overshoot. Thus, the dynamic bonds function as reversible sacrificial
ing the hydrogel’s integrity during deformation11,32. Despite having bonds to dissipate energy. Various kinds of non-covalent bonds, includ-
a high water content (~90 wt%), DN hydrogels are stiff, strong and ing borate/di-diol complexation, electrostatic interaction, hydrogen
tough; the fracture energy is even comparable to industrial rub- bonds, metal–ligand coordinate bonds, hydrophobic interaction, π–π
bers and natural cartilage30. Achieving exceptionally high elastic interaction, cation–π interaction and host–guest interaction, have
modulus, strength and fracture energy in DN hydrogels revolves been used (Fig. 3a). Synthesis of this type of hydrogel is simple and
around two key aspects. The first network should be more brittle usually involves a one-step reaction to form the chemically crosslinked
and weaker than the second network, that is, it should have a smaller primary network. Non-covalent interactions are induced by post-
fracture stretch ratio (λf ) and fracture stress (σf ). Conventional DN treatments, such as dialysis to remove counterions for polyampholy-
gels usually use highly crosslinked polyelectrolyte networks (such tes, or immersion in a multivalent ion solution to form coordination
as PAMPS (poly(2-acylamido-2-methylpropanesulfonic acid)) and complexes. During post-treatment, intrachain and interchain dynamic
PAAc (poly(acrylic acid))) as the first network, and slightly crosslinked bonds are formed and reorganized to form aggregations, causing the
neutral polymers (such as PAAm, PDMAAm (poly(N,N-dimethylacryla- polymer strands to collapse into globule conformation, analogous to
mide)) and PHEMA (poly(2-hydroxyethyle methacrylate))) as the sec- the folded structure of proteins. Therefore, these hydrogels deswell in
ond network. The highly crosslinked polyelectrolyte network, which water (λs < 1) after post-treatment, resulting in a high polymer volume
over-swells in water owing to high osmotic pressure from counterions fraction (around 50 vol%).
(Fig. 1a, regime B), is both brittle (small λf ) and weak (small σf ) while Here we take the hydrogen bond as an example, but examples
exhibiting a high Young’s modulus29. In addition, the monomeric molar of hydrogels containing other non-covalent bonds can be found in
ratio between the sparsely crosslinked second neutral network and comprehensive reviews5,128,129. Hydrogen bonds are ubiquitous in living
the first network should range from several to a few tens, resulting systems, contributing to the formation of secondary structures in pro-
in larger λf and σf. The mechanical contrast between the two elastic teins. When designing tough hydrogels based on hydrogen bonds, the
networks results in inelastic behaviour, the DN hydrogel exhibiting presence of water molecules in hydrogels can disrupt the bonds, poten-
yielding and necking during tensile loading111 and substantial hyster- tially limiting the stability of dynamic bonds34. Therefore, to stabilize
esis during cyclic loading112. Because the second network suppresses the hydrogel, formation of multiple hydrogen bonds between mono-
the stress concentration in the defects of the first network, the yield- mer pairs is required34,130, or the hydrogen bonds can be further stabi-
ing stress is much higher in comparison to the fracture stress of the lized by hydrophobic interactions131,132. For instance, the chemically
first network alone. Above the yielding point, the short-strand first crosslinked PAAm hydrogel exhibits swelling behaviour and no self-
network is considered to rupture into fragments, dissipating energy, healing property because its hydrogen bonding is attacked by water. By
whereas the stretchable second network maintains structural integrity contrast, after the PAAm hydrogel is synthesized in the presence of tan-
and sustains further elongation113,114. In fracture tests, covalent bond nic acid, which can form multiple hydrogen bonds between the PAAm
rupture in the first network ahead of the propagating crack forms a polymer and tannic acid, the gel deswells in water with self-healing
damage zone with a thickness of several hundred micrometres, and the ability97. The deswelling behaviour of PAAm–tannic-acid hydrogel

Nature Reviews Materials


Review article

Table 1 | Mechanical properties of representative swelling and deswelling hydrogels

Category Main Water Young’s Fracture Fracture Work of Fracture Fatigue threshold Comments Refs.
components content modulus stretch stress σf extension energy Γ
(wt%) (MPa) ratio λf (MPa) Wb (MJ m−3) (kJ m−2) Γ0 (J m )
−2
Observation
resolution

Elastic hydrogels
Homogeneous Tetra-PEG ~90 0.001–0.01 2–8 0.025–0.2 0.015–0.85 0.02–0.07a 12.2 ~20 nm per – 23,181,
network cycle 194
Slide-ring gel PEG with ~90 0.009–0.08 12–15 0.02–0.12 NR 0.002–0.25b NR NR Amorphous 14,106,
polyrotaxanes 195
Dense PAAm 70–99 ~0.1 Up to 5 0.39 NR 0.1–2b 200 ~1 nm per – 26
entanglement cycle
Polyprotein PAAm ~95 0.012 11 0.08 NR 0.9b 126 ~10 nm per – 53
crosslinkers crosslinked by cycle
polyprotein
crosslinkers
DN gels PAMPS–PAAm ~90 0.1–1.0 10–32 1–10 11 1–4.5a 41–400 ~1 nm per – 32,49,
cycle 50,127,
196
Viscoelastic hydrogels
DN gels with Ca2+-alginate– ~86 0.01–0.11 2–23 0.03–0.25 NR 0.025–9b 35–50 ~1 nm per Not water 12,133,
physical PAAm cycle equilibrium 197
interactions
Dual-crosslinked P(NaSS-co- ~50 0.08–0.4 5–33 0.19–0.55 0.93–2.8 1.5–4a,b 10–100 ~1 nm per Water 40,42,
(ionic bond) DMAEAQ) cycle equilibrium 58,140
Supra­molecular PNAGA 70–86 0.05–0.15 7–16 0.16–1.1 NR 0.2–1.2a NR NR No chemical 34
hydrogel crosslinker
(hydrogen bond)
Dual-crosslinked Zr4+-P(AAm-co- 33–74 0.4–28.5 3.2–13.5 2.1–11.6 NR 3.7–24.2a NR NR – 198
(coordination AMPS)
complex)
Dual-crosslinked PAAm–DVB 86 0.027–0.9 > 102 0.4–1.5 7.3–18.8 2.5–26b 2,500 ~1 μm per – 71
(hydrophobic cycle
interaction)
Hydrogels with high-order structure
Bilayer lamellar PDGI–PAAm 93–98 0.005–0.12 8–24 0.05–0.6 Up to 5 5b NR NR – 154–157
structure
Strain-induced Slide-ring PEG 60–80 ~0.2 13–15 1–5.5 6.6–22 2.9–3.6c NR NR – 15
crystallization
Phase separation PAAc–calcium 39–74 3.4–131 1.6–7 0.19–17.9 0.16–9.6 NR NR NR – 16
(ionic bonds) acetate
Microcrystal and PVA 58–90 0.03–10 Up to 30 0.3–23.5 Up to 210 0.1–170b,c 150– ~1 nm per – 5,18,64,
nanofibre 10,000 cycle 145,149,
150
Microfibre DCC-alginate ~60 59–342.4 1.35–3.7 19.8–81.5 7.04–38 NR NR NR – 48,199
and cellulose
Fibre-woven PA–GF 38 606 NR 380 N mm−1 d NR 100–250a NR NR – 44,179
fabrics
Composite 2D PVA–GO 80 10 3.3 8 Ca. 10 NR 1,500 ~1 nm per – 67
composite cycle
hydrogel
DCC, drying under confined conditions; DN, double network; DVB, divinylbenzene; GO, graphene oxide; NR, not reported; PAAc, poly(acrylic acid); PAAm, polyacrylamide; PA–GF, woven
glass fibre fabric–polyampholyte hydrogel composite; PAMPS, poly(2-acylamido-2-methylpropanesulfonic acid); PDGI, poly(dodecyl glyceryl itaconate); PEG, polyethylene glycol; PNAGA,
poly(N-acryloyl glycinamide); P(NaSS-co-DMAEAQ), sodium styrenesulfonate and acryloyloxethyltrimethylammonium chloride copolymer; PVA, polyvinyl alcohol. aFrom tearing test; bfrom
pure shear tests; cfrom simple extension; dtearing strength.

is conducive to high extensibility (λf up to 16) and tensile strength crack propagation, imparting a high fracture energy to the hydrogel
(σf up to 2.5 MPa). In addition, the multiple hydrogen bonds between (Г up to 3.1 kJ m−2), nearly 30 times higher than that of PAAm gels solely
the polymer and tannic acid effectively dissipate energy and suppress crosslinked through chemical crosslinking.

Nature Reviews Materials


Review article

Double-network gels with physical interactions. Traditional DN gels nanocomposites, are effective approaches for toughening hydrogels
are composed of two chemically crosslinked elastic networks, but DN (Fig. 3b). Interplays between hierarchical structures from molecular
gels can also be formed by combining a physically crosslinked and a to mesoscopic scales endow strong crack resistance. Particularly, the
chemically crosslinked network. For example, DN hydrogels from Ca2+ processing zone is enlarged because of the increase in the load-transfer
crosslinked alginate network and chemically crosslinked PAAm network length by submicrometre-scale to centimetre-scale stiff constituents,
exhibit partial self-recovery and a high fracture energy of ~9 kJ m−2 in resulting in a large ГD. Meanwhile, the Г0 can be enhanced by two to
the as-prepared state, highlighting the beneficial effect of the dynamic four orders of magnitude. Consequently, incorporating reinforcement
bonds10. However, owing to osmotic pressure, this DN hydrogel swells in components is one of the most effective strategies to substantially
water and becomes weaker than in its as-prepared state12,133,134. DN gels enhance both fracture energy and fatigue threshold simultaneously,
containing no chemically crosslinked network have also been devel- to a level even better than bio-tissues (Fig. 4).
oped. For example, combining an amphiphilic triblock copolymer as
the first network and a linear PAAm as the second network, a deswollen Microphase separation. Deswelling hydrogels are prone to microphase
water-equilibrated block copolymer PAAm DN hydrogel was synthe- separations, forming polymer-dense phases and polymer-sparse phases
sized (water content ~50 wt%)135,136. The amphiphilic triblock copolymer at scales much larger than the dual-crosslinked structure (Fig. 3b).
PBMA-b-PMAA-b-PBMA (poly(butyl methacrylate)-b-poly(methacrylic The size of the microphases is governed by the competition between
acid)-b-poly(butyl methacrylate)) forms a hyperconnective physical enthalpic gain of the physical interactions and entropic penalty of the
network through strong hydrophobic associations of the end-block primary network. The interplay of the molecular and mesoscale hier-
PBMA135. Simultaneously, the midblock PMAA forms hydrogen bonds archical structures dissipates a substantial amount of energy through
with the linear PAAm chains to allow energy dissipation (Fig. 2c). The a large load-transfer length. For example, polyampholyte hydrogels
hydrogels show abnormally large non-softening, and quasilinear but (PA gels)13,137 copolymerized from cationic and anionic monomers
inelastic deformation, and they have high elastic modulus (1.7 MPa), at high monomer concentrations around the charge balance point
fracture energy (Г ≈ 3 kJ m−2), strength (σf ≈ 10 MPa), extensibility (λf ≈ 7) form a bicontinuous microphase separation, in which the hard micro-
and quick self-recovery (85% recovery within 5 min). phase network and soft microphase network are interpenetrated41,138.
The PA gels shrink to approximately 40% in volume when immersed
Hydrogels with high-order structure in water to dialyse counterions. During dialysis, the conformation of
Introducing high-order structures, such as microphase separa- the polyampholyte strands changes from coil to aggregated globule.
tion structures, microcrystals and fibrils, fibre-woven fabrics and Water-equilibrated PA gels exhibit a polymer volume fraction ~50 vol%,

a Dual-crosslinked hydrogel

Deswelling in water (viscoelastic) – +

Physical crosslinking/ Chemical crosslinking Dual crosslinked PVA–borate ions Ionic bond
supramolecular network Tough and self-healing
H O
N HO M
R
Physical crosslinking/ Hydrogen bond Coordination complex
+
supramolecular network

Hydrophobic interaction π–π interaction

Cation–π interaction Host–guest interaction

b Hydrogel containing load-bearing high-order structure


Deswelling in water (viscoelastic) Partly deswelling in water (viscoelastic) Usually used in the
as-prepared state
Phase separation Fibre Crystal Bilayer lamellar structure Nanocomposite

d1
d2

d1 ≈ 4.7 nm
d2 ≈ hundreds nm

Fig. 3 | Deswelling (viscoelastic) hydrogels and hydrogels with high-order from physical crosslinking. b, Hydrogel containing high-order structure. PVA,
structure. a, Dual-crosslinked hydrogel composed of a primary network from polyvinyl alcohol. Panel b adapted with permission from ref. 40, APS, ref. 177,
chemical crosslinking and/or trapped entanglement and dynamic networks Wiley, and ref. 156, ACS.

Nature Reviews Materials


Review article

a Fracture energy and fatigue threshold versus stiffness

III III
105 104

II II IV

Fatigue threshold Γ0 (J m–2)


Fracture energy Γ (J m–2)

104 103
IV

103 I
I 102

102
101

101

100
100
10–1 100 101 102 103 104 105 106 107 10–1 100 101 102 103 104 105 106 107
Modulus (kPa) Modulus (kPa)

b Fracture energy and fatigue threshold versus strength


106
104

10 5
II
Fatigue threshold Γ0 (J m–2)
IV III
Fracture energy Γ (J m–2)

IV III 103
104
II

103 102
I
I
10 2

101
101

100
100

10–1 100 101 102 103 104 105 106 10–1 100 101 102 103 104 105
Strength (kPa) Strength (kPa)

c Fatigue threshold versus fracture energy

I: Elastic hydrogels II: Viscoelastic hydrogels


104
Simple network PAAm Physical network–PAAm DN
II
Highly entangled PAAm PA
Fatigue threshold Γ0 (J m–2)

III
103 Tetra-PEG Coordination interaction
IV Protein/polypeptide crosslinked PAAm Dynamic covalent bond
PAMPS–PAAm DN Hydrophobic interaction
I
102 Hydrogen bonding
III: High-order structure
Phase separation PA IV: Bio-tissues
101
Phase separation PMMA–PAAc
Bio-tissues
Crystal PVA
Crystal PEG
100
Salt-out PVA
100 101 102 103 104 105 106
2D PVA–GO
Fibre-reinforced hydrogel
Fracture energy Γ (J m–2)

Fig. 4 | Fracture energy and fatigue threshold versus elastic modulus and topological network26,51–53,181,202,203 and double network (DN)49,50 (I). Viscoelastic
strength of hydrogels fabricated by various strategies. a, Fracture energy hydrogels with dynamic bonds42,54,55,58,71,133,140,180,197,204–206 (II). Hydrogels containing
and fatigue threshold versus elastic modulus. b, Fracture energy and fatigue high-order structure18,27,42,47,58,59,62,64,65,145,150,166,182,207,208 (III). Load-bearing bio-
threshold versus strength (engineering fracture stress). c, Fatigue threshold tissues5,209 (IV). GO, graphene oxide; PA, polyampholyte; PAAm, polyacrylamide;
versus fracture energy. Elastic hydrogels including simple network6,201, PEG, polyethylene glycol; PVA, polyvinyl alcohol.

independent of chemical crosslinking density but slightly dependent scales of nanometres, tens of nanometres and hundreds of nanometres,
on the chemical structure of the monomeric units41,42. The hierarchical respectively, play a crucial role in achieving exceptional mechanical
structures, including the transient network due to dynamic bonds, the properties (λf up to ~35, modulus up to ~2 MPa, strength up to ~4 MPa,
primary network, and the bicontinuous phase networks, which exist at self-healing capability up to 90%, and fracture energy up to ~4 kJ m−2)

Nature Reviews Materials


Review article

through multistep energy dissipation40. The hierarchical structures in bilayers (PDGI) inside the polymer matrix (PAAm)154,155 (Fig. 3b). Owing
PA gels also contribute to excellent mechanical adaptation to cyclic train- to the water-impermeable nature of the hydrophobic bilayers, these
ing139. The value of λf is very close to the theoretical maximum stretch hydrogels show 1D swelling, anisotropic molecular permeation and
ratio λmax predicted from the average network structure, suggesting that diffusion, and substantial mechanical anisotropy along the thickness
such PA hydrogels have a defect-insensitive length ξ much larger than and in-plane directions. The rigid PDGI lamellar layers, having a lipid-
the mesh size of the primary network. The small-strain moduli, large like mobile nature, act as reversible sacrificial bonds that disassociate
deformation energy dissipation, and fracture energy of PA gels not only during deformation, thereby contributing to substantial energy dis-
obey the time–temperature superposition principle87, but also obey the sipation. Furthermore, at large deformation, the single-domain PDGI
time–salt superposition principle, where salt ions screen the ionic bond- lamellar bilayers transform into a hierarchical fibrous structure con-
ing between polyampholytes, accelerating the dynamic processes like sisting of micrometre-thick fibre bundles made from nanometre-thick
the temperature, and the dynamic mechanical behaviour for different fibrils, which impedes crack growth and results in crack blunting156,157.
salt concentrations can be shifted onto a single master curve using salt- Therefore, the hydrogels exhibit large hysteresis, high strength and
concentration-dependent shift factors for the frequency140–142. Through extensibility, and extraordinary toughness. Lamellar hydrogels are
time–salt superposition, one can access a wide range of timescales that often called photonic hydrogels because they usually exhibit structural
are difficult to access at room temperature. colours due to Bragg’s reflection on the multilayer planes. The colours
can be dynamically tuned over a broad spectral range, spanning from
Microcrystals and microfibres. Hydrogels containing microcrys- the ultraviolet–visible region to the near-infrared region, by tuning the
tals and microfibres typically have coexisting swelling (amorphous) hydrogel layer thickness through the application and release of stress
and deswelling (crystalline or fibre) domains (Fig. 3b). Because water or strain. Therefore, these tough gels could be used as stress or strain
weakens polymer–polymer interactions, special strategies are required sensors and deformation-based colour displays. Various chemical
to introduce microcrystalline domains or fibrils into hydrogels. Com- structures have been incorporated into the soft layers to achieve differ-
mon strategies include freeze–thaw143, mechanical stretching144, sol- ent functionalities. For instance, PDGI–h-PAAm (partially hydrolysed
vent exchange145, thermal drawing and rapid quenching146, and salting PAAm) hydrogels have ultrafast colour response over the whole visible
out18,147,148. The formation of microcrystals can suppress crack propa- wavelength158, and PDGI–PAAcNa (sodium polyacrylate) hydrogels
gation because the crystalline domains have high chain density, and show colour tunability under perpendicular or parallel electric fields
the energy for rupturing a crystalline domain is higher than that for in the directions of the hydrogel layers159.
rupturing the amorphous polymers5. Meanwhile, under loading, micro-
crystals convert to fibrils149, which are highly anisotropic and prone to Organic–inorganic nanocomposite. Polymer networks that are physi-
crack deflection. For example, freeze–thawing and air-drying resulted cally adsorbed on hard inorganic nanoparticles160–162, nanotubes163,164 or
in polyvinyl alcohol (PVA) hydrogels with high crystallization, which nanosheets17,67,165,166 form one type of organic–inorganic nanocompos-
showed superior crack resistance and therefore superior mechanical ite. The inorganic components act as a high-energy phase to toughen
properties150. Combining freeze-casting and salting out led to the PVA the hydrogel167–169 (Fig. 3b). To achieve superior mechanical perfor-
hydrogels with a highly anisotropic hierarchical structure, consisting mances in the water-equilibrated state, one should choose a hydrogel
of micrometre-scale honeycomb-like pore walls and interconnected showing deswelling behaviour or containing crystalline structure as
nanofibril meshes18. This anisotropic hierarchical structure contributes polymer matrix to enhance the interfacial affinity166, or the polymer
to the suppression of crack propagation, leading to superior mechanical matrix should be post-crosslinked after the formation of a strong inter-
performances (σf = 23.5 MPa, λf = 30, Г = 170 kJ m−2 and Г0 = 10.5 kJ m−2). face between the polymer and the inorganic components170. Otherwise,
the swelling mismatch between the polymer network and inorganic
Strain-induced crystallization. Similar to natural rubber151–153, water- components weakens the load-transfer interface. Such nanocomposite
containing hydrogels reinforced by strain-induced crystallization gels with a swelling mismatch in organic–inorganic components have
have also been developed. Strain-induced crystallization can serve superior performance in the as-synthesized state (usually σf > 0.1 MPa,
as a damage-free reinforcement strategy for slide-ring PEG hydro- λf > 10 and Г > 2 kJ m−2 for an appropriate amount of polymer and
gels with reduced slidable crosslinker hydroxypropyl-α-cyclodextrin inorganic components)17,171–173, whereas their mechanical properties
rings15. Because the crosslinks can slide along the PEG chains, the become weak at the swollen equilibrium state174.
initially amorphous polymer strands between the crosslinks can
become long and uniformly stretched under large deformation, Fibres and fibre-woven fabric composite. Incorporating macroscale
resulting in highly oriented PEG chains. Microcrystals form and melt stiff fibres or fibre-woven fabrics into hydrogels is a powerful strategy
with elongation and retraction, respectively. Consequently, the crack for fabricating strong and tough composites (Fig. 3b). This strategy
resistance during loading is substantially enhanced, resulting in a relies on two key principles62,175: strong interface adhesion between the
fracture energy Г of up to 3.6 kJ m−2, much higher than gels with fixed fibres or fabrics and the soft matrix to effectively disperse stress at the
crosslinker or non-strain-induced crystallization systems (Г from about crack tip; and a large modulus contrast between the stiff fibres or
0.05 to 0.24 kJ m−2). fabrics and the soft hydrogel. The load-transfer length lT of unidirec-
tional fibre-reinforced soft composites correlates with the modulus
Bilayer lamellar structure. The unidirectionally aligned hydrophobic contrast between the fibre and matrix, lT ≈ EfAf /μm , where Ef is the
lamellar bilayers embedded in simple hydrophilic networks lead to Young’s modulus of the fibre, μm is the shear modulus of the matrix,
hydrogels with coexisting swelling and deswelling layers. A unique and Af is the cross-sectional area of a fibre176,177. Along this line, polymer
example is anisotropic photonic poly(dodecyl glyceryl itaconate)/ fibres have been introduced into swelling hydrogels63,178. To achieve
polyacrylamide (PDGI/PAAm) hydrogels, which consist of macroscopic, good adhesion and load transfer between the matrix and polymer
single-domain, periodical stacking of integrated microscopic lamellar fibres, covalent bonds were used for interlinking62. Such composite

Nature Reviews Materials


Review article

a Cyclic fatigue measurement

L0
Elastic gel
C0 Cyclic to
H0 steady state

∆c/∆N
Gel with
Undeformed state
hierarchical
structure
1st cycle
N cycles

Stress (MPa)
N=1 2 … 0
λm c G0 Gtran G
… Nth cycle

1 New surface
Time W(λm) λm

λ G0: Fatigue threshold Γ0 = G0


G = W(λm) H0 Gtran: Slow-to-fast crack growth transition

b Fatigue resistance of elastic hydrogels c Fatigue resistance of viscoelastic hydrogels

Chemical crosslink Reversible crosslink

Cyclic Cyclic
loads loads
Chain Chain
scission scission

d Fatigue crack resistance by hierarchical structure


Phase separation Fibre

Polymer-dense phase Fibre Matrix

Cyclic Cyclic
loads loads

Polymer-sparse phase

Fig. 5 | Fatigue fracture measurement and the mechanisms proposed to the fatigue threshold G0 below which the crack does not grow. b, Chain scission
enhance the fatigue resistance of hydrogels. a, Pure shear test for cyclic mechanism for elastic hydrogels. c, Fatigue resistance of viscoelastic hydrogels.
fatigue test. From the fatigue resistance curve (Δc/ΔN versus G), one can obtain d, Fatigue resistance enhanced by high-order structures.

hydrogels showed a certain reinforcement in the as-synthesized state, Fatigue resistance of hydrogels
but were rarely used in water equilibrium state, since swelling resulted Long-term mechanical stability under prolonged cyclic loads is extremely
in weak mechanical performance owing to stress mismatch between important for tough hydrogels to emulate load-bearing biological tis-
the hydrogel matrix and the fibres. Using deswelling hydrogels to sues, such as muscles and cartilage, that are constantly undergoing
form the fibre-reinforced hydrogels is a more successful strategy. For reciprocal cycles in daily life. An early study on the fatigue of hydrogels
example, composites obtained from glass fibre-woven fabric and poly- was reported in 2017 (ref. 6). Fatigue damage occurs when a sample is
ampholyte hydrogels show exceptional mechanical performance with subjected to irreversible changes in mechanical properties under cyclic
a superior tearing strength of 380 N mm−1, a tensile modulus of 606 MPa loading92. Fatigue is usually characterized by two aspects: the critical
and a fracture energy of 250 kJ m−2, which are several orders of magni- energy release rate G0 for a pre-existing crack to grow, and the crack
tude greater than individual neat materials44,179. Such excellent mechan- growth rate per cycle (dc/dN) versus the energy release rate G of cyclic
ical properties stem from the strong physical adhesion of the deswelling loading180 (Fig. 5a). Typically, a sample with a pure shear or single-edge
polyampholyte matrix to the negatively charged glass fibre surface in notch geometry is used for cyclic fatigue tests. In an example of pure
water as well as from the high viscoelastic energy dissipation density shear geometry (Fig. 5a), a pre-notch perpendicular to the loading direc-
of the polyampholyte matrix. tion is made at the middle edge of the sample. At a certain loading

Nature Reviews Materials


Review article

amplitude (λm) and cycle period (or speed), the crack length c grows The observation resolution of several representative hydrogels is listed
above a certain energy release rate G, and c increases gradually with in Table 1.
number of cycles N. The energy release rate can be obtained from
the stress–stretch curve of the corresponding unnotched sample at the Fatigue-resistant behaviour of elastic hydrogels. Currently, the
λ
same loading condition, by G = W(λm)H0, where W (λ m) = ∫ m σ dλ is fatigue mechanisms in hydrogels are mainly explained by rubber elastic
1
the strain energy when the cyclic loading curve reaches steady state, theory92. The fatigue threshold Г0 of a single-network hydrogel fol-
and H0 is the initial sample height along loading direction. From the lows the chain scission mechanism (Fig. 5b), that is, the Lake–Thomas
crack growth rate per cycle versus the energy release rate, one obtains model (equation (8)), which was also confirmed in tetra-PEG gels with
the fatigue-resistant curve and the fatigue threshold Г0 = G0 (dc/dN → 0) a relatively homogeneous network structure181. Therefore, increasing
below which the crack does not grow (infinite lifetime). A comparison Nx in simple network hydrogels is an effective strategy to improve Г0.
of Г0 between synthetic hydrogels (including elastic, viscoelastic hydro- However, a large Nx results in small elastic modulus E (ref. 92). To solve
gels and hydrogel containing high-order structures) and bio-tissues is the issue, fatigue-resistant hydrogels that use tandem-repeat pro-
presented in Fig. 4. When comparing Г0 among different materials, atten- teins (polyproteins) as crosslinkers and coiled PAAm for percolating
tion should be paid to the observation resolution of the fatigue thresh- networks were developed53. During cyclic fatigue tests, the folda-
old. Г0 is denoted as the threshold below which crack growth is ble polyprotein crosslinkers around the crack tip unfold, yielding a
undetectable at the camera resolution under certain fatigue cycles. fatigue threshold (Г0 = 126 J m−2) that is substantially superior to that of

a Force-induced mechanical activation of dibromocyclopropane mechanophores

Br
O O
8

Br O O O
Br Stress activation
TBA SA
8 O
Crosslinking
Br Br
O O

Closed Open
Br

Stress activation and


chain scission Crosslinking

b Force-induced radicals in DN gel c Force-induced cyanofluorene radicals to initiate the polymerization of side chains

Physical
C C crosslinking
or
N N
Force
Force

R
R

Fig. 6 | Strategies proposed for mechanochemical strengthening and induce polymerizing monomers to reconstruct and strengthen the network.
self-growing hydrogels. a, Self-strengthening through ring-opening reaction Compared with the first network crosslinked by strong C–C bonds, the weak
of gem-dibromocyclopropanes and subsequent crosslinking by nucleophilic azoalkane crosslinked first network can produce more mechanoradicals in DN
displacement reactions with TBA SA (ditetrabutylammonium salt of sebacic gels. c, Force-induced cyanofluorene radicals to initiate the polymerization of
acid). b, Force-induced mechanoradicals in double-network (DN) gel, which can side chains. Panel a adapted from ref. 185, Springer Nature Limited.

Nature Reviews Materials


Review article

a Force-induced self-growth and reinforcement

Mechanical stress Rest First network Open system


1st cycle
(rigid and brittle)
4 2nd
Second network 3rd
(soft and stretchable) 3 4th

F (N)
Covalent bond 2
Destruction Reconstruction
R• •R scission
1
R• Mechanoradical

Newly formed 0
network 15 30 45 60 75 90
Monomers L (mm)

b Sustainable mechanochemical growth supported by vascular-like perfusion through the inner channel

Blood vessel

Circulatory system Channel

Nutrients Pump

Chemicals

F F
Jig
Destruction
Luer lock 1st cycle 60 min
R· ·R 5 min 120 min
20 min 360 min
Reconstruction Repeat
5
Tensile deformation

4
Complement

3
F (N)

Flow
2

PNaAMPS network PAAm network 1


R · Mechanoradical Chain fracture
Monomer 0
Diffusion New network
0 15 30 45 60 75 90 105
Dp (mm)

c Force-triggered rapid microstructure growth on surface


d
Indenter

Monomer solution
Force Microstructure H Vertical pattern Horizontal pattern Smooth surface
stimuli growth
Gel surface
F Lmax D
TDN
Gel bulk
g
1 mm 1 mm 1 mm
θrec = 15° θrec = 12° θrec = 20°
F F g

Double network Chains scission in process zone New polymer chain growth θadv = 42° θadv = 50°
θadv = 30°
500 µm 500 µm 500 µm

First network Second network Monomer Mechanoradical Newly formed


(brittle) (stretchable) polymer chain

Nature Reviews Materials


Review article

Fig. 7 | Force-induced self-growing and strengthening in DN hydrogels. and water for avoiding sample dry. c, Microstructure growth on DN
a, Self-growing and strengthening of double network (DN) hydrogel hydrogel surface via force-induced self-growing strategy. θadv, advancing
through rupturing the first network to produce radicals and subsequent angle; θrec, receding angle; Dp, displacement; PAAm, polyacrylamide; PNaAMPS,
polymerization to reconstruct network. F is force and L is sample length. poly(2-acrylamido-2-methyl-1-propanesulfonic acid sodium salt). Panel a
b, Sustainable mechanochemical growth of channel-containing DN gel reprinted with permission from ref. 72, AAAS. Panel b reprinted with permission
under cyclic training. The channel can supply monomers for polymerization from ref. 187, RSC. Panel c reprinted from ref. 188, CC BY 4.0.

PAAm hydrogel crosslinked by bisacrylamide (Г0 = 7.5 J m−2), which has Rather than the nanometre-scale polymer chains, the large-scale stiff
about half the modulus of the polyprotein crosslinked PAAm. Induc- constituents are broken to advance the crack during cyclic loading.
ing highly entangled polymers into single-network elastic hydrogels Therefore, the intrinsic energy required to advance the crack per unit
is another strategy to improve Г0. For instance, PAAm hydrogel with area equals the covalent energy of a layer of the reinforcement con-
dense entanglements fabricated through a high concentration of PAAm stituents per unit area. The Lake–Thomas model can be generalized as
and sparse crosslinks has high modulus (0.1 MPa) and Г0 (200 J m−2)26. Г0 = LNU (refs. 183,184), where L is the size of reinforcement constituents,
The sparse chemical crosslinks impeding the disentanglement of N is the number of the reinforcement constituents per unit volume, and
polymer strands contribute to the high modulus, and the dense entan- U is the energy required to break the reinforcement constituents. Com-
glements allowing the transmission of tension in the polymer strands pared with typical hydrogels without reinforcement components, L can
at the crack tip to other strands contribute to the fatigue threshold. be increased by two to eight orders of magnitude, and U can be increased
As a special type of elastic material, chemically crosslinked DN by fourteen orders of magnitude184. Meanwhile, N can be adjusted within
hydrogels can resolve the conflict between Г0 and E (refs. 49,50). The a broad range. As a consequence, Г0 can be enhanced by two to four
high elastic modulus and robust fatigue resistance are ascribed to orders of magnitude. These hydrogels show fracture energy and fatigue
the densely crosslinked first network and sparse second network, thresholds comparable to or even better than bio-tissues (Fig. 4).
respectively. During cyclic fatigue tests, the first network fractures Hydrogels with hierarchical structures may exhibit multilevel
into small fragments, effectively serving as mobile crosslinks for the fatigue resistance compared with elastic hydrogels, which is derived
second network, which features long strands (large Nx). Consequently, from the minor slope of the curve of dc/dN versus G (Fig. 5a). For exam-
DN hydrogels with loosely crosslinked second networks can achieve ple, hierarchical PA gels that include ~0.1-nm ionic bonds, ~10-nm
a fatigue threshold of up to 400 J m−2 with a high Young’s modulus of primary polymer network, and ~100-nm bicontinuous phase network
0.3 MPa while still containing around 90 wt% water. structures (Fig. 5d) exhibit multilevel fatigue-resistant behaviour42,58,140.
The fatigue threshold Г0 is mainly correlated with the mesh size of the
Fatigue resistance of viscoelastic hydrogels containing dynamic primary polymer network bridged by both chemical crosslinking and
bonds or weak crosslinking. Although dynamic bonds substantially trapped entanglement. Above Г0, a transition point Gtran marking the
enhance the toughness of viscoelastic hydrogels, they contribute little jump from slow to fast crack growth occurs in PA gels with strong phase
to the fatigue threshold. The dynamic bonds are gradually destroyed separation. Gtran is correlated to whether the damage occurs in the hard
under cyclic loading, and the fatigue threshold of such tough hydrogels phase network. Fatigue fracture is delayed to a larger Gtran when there is
also follows the chain scission mechanism described by the Lake– a bigger phase contrast, whereas gels without phase separation exhibit
Thomas model (Fig. 5c). For example, Ca2+-alginate/PAAm hydrogels only a fast crack growth mode (without Gtran).
with interchain ionic bonds have a high fracture energy of several kilo- Microcrystalline domains are comprised of densely folded chains.
joules per square metre but a low Г0 of 35 J m−2, which is close to the Pulling out the polymer chain from a microcrystalline domain needs
corresponding Na+-alginate/PAAm without interchain ionic bonds133. energy multiple times what is needed to fracture a single polymer
Nevertheless, the presence of dynamic bonds does reduce the crack chain. In addition, mechanically rupturing the microcrystalline domain
growth rate above Г0. Notably, whether dynamic bonds contribute to demands energy several times that required to rupture the correspond-
gel Г0 depends on the observation timescale relative to the relaxation ing amorphous polymers61. Therefore, microcrystalline domains act as
time of the hydrogel. By extending the observation timescale of PA gels intrinsic high-energy phases. For example, the introduction of micro-
by the time–salt superposition principle140,141, Г0 was observed to be rate crystalline domains into amorphous PVA hydrogels using a freeze–
independent and matched the predicted value from the Lake–Thomas thawing and air-drying approach enhanced the fatigue threshold from
model only when the fatigue test was performed at a strain rate in the 10 to 1,000 J m−2 (ref. 150). Meanwhile, the elastic modulus of such PVA
elastic regime. By contrast, Г0 increased with increasing the strain rate hydrogels can reach 10 MPa, effectively resolving the tradeoff between
ε ̇ (Г0 ~ εν̇ with scaling parameter 0 < ν < 0.27) when the fatigue test was modulus and fatigue threshold. Similar to microcrystals, fibres62–64,182
performed in the viscoelastic regime140. Thus, the fatigue threshold can can act as strong barriers to fatigue crack extension, resulting in a high
be improved by tuning the sacrificial bond dynamics of the viscoelastic fatigue threshold (Г0 > 1 kJ m−2), especially when the fibres are aligned
hydrogels. Similar to dynamic bonds, weak side-chain crosslinks are perpendicular to the crack extension direction (Fig. 5d).
also gradually destroyed under cyclic loading. Triggering the rupture
of weak side-chain crosslinks using force can lengthen the bridging Mechanochemical strengthening and self-growing hydrogels
strands around the crack tip, improving the fatigue threshold108. Living soft tissues possess the remarkable ability to autonomously grow,
remodel and strengthen themselves in direct response to mechanical
Fatigue resistance improved by high-order structure. Introducing forces. For instance, skeletal muscles can increase in mass and strength
high-order structures, such as microphase separations42,58–60, micro- when subjected to cyclic training. There has thus been increasing interest
crystals145,150, fibrils18,47,61,62,64,182,183 and nanocomposites66,67 into polymer in developing self-growing and self-strengthening polymer networks
networks can substantially improve the fatigue resistance of hydrogels. based on force-triggered mechanoradical generation and subsequently

Nature Reviews Materials


Review article

induced chemical reactions. Mechanical forces typically cause bond biomedical fields. However, challenges in fundamental science and
scission in hydrogels, resulting in material failure. These mechano- engineering still limit their real-life applications. From the fundamental
radicals, in turn, initiate processes such as crosslinking or polymeriza- point of view, cutting-edge fracture mechanics theories were developed
tion propagation, effectively converting mechanical energy into an for nonlinear elastic materials, which assume no strain-rate depend-
increase in material strength (Fig. 6). Force-induced crosslinking was ency. However, hydrogels with dynamic bonds are strongly strain-rate
pioneered by the application of shear forces to trigger the ring opening dependent, like most biological tissues. Moreover, hydrogels with
of gem-dibromocyclopropanes to generate allylic bromides, which were relatively strong dynamic bonds show loading-history dependence,
subsequently crosslinked through nucleophilic substitution reactions resulting in richer and more complex mechanical behaviours. New
with a bifunctional carboxylate185 (Fig. 6a). The mechanically triggered theories and experimental approaches are required to understand
crosslinking outcompetes the destructive shearing forces, resulting in the toughening and antifatigue mechanisms related to such nonlinear
an increase in the material’s modulus by several orders of magnitude. viscoelastic effects. Resulting progress in understanding the dynamic
In 2019, a ‘self-growing’ strategy was introduced that relies on force- aspect of the materials and the molecular mechanisms should assist in
induced mechanoradical generation and subsequent polymerization the design of new structures in material development.
within DN hydrogels72. The densely crosslinked initial brittle network From an engineering point of view, developing hydrogels with
undergoes ruptures when subjected to loads, concurrently generating comprehensive mechanical properties that match those of specific
mechanoradicals (Fig. 6b). These mechanoradicals then initiate the biological tissues is indispensable. Currently, loading-induced sof-
polymerization of monomers, forming a new polymer network. This tening, substantial stress relaxation, and large hysteresis loops are
process, similar to open and dynamic biological systems186, leads to the main mechanical characteristics of tough hydrogels. Introducing
an increase in the mass of DN gels under repeated mechanical loading protein-like intrachain folded structures into hydrogels may be a future
through structural destruction and reconstruction with a sustained strategy to create hydrogels with fast hysteresis recovery. Moreover,
monomer supply, substantially strengthening the DN gels72 (Fig. 7a). By one needs to address issues related to biocompatibility and biochemi-
constructing a vascular-like circulatory system to supply monomers to cal stability in vivo, ensuring that implanted hydrogel artificial tissues
channel-containing DN gels, sustainable self-growth of hydrogels in air can maintain their structure and functionality. Because hydrogels are
was achieved187 (Fig. 7b). The force-induced self-growing strategy within an open system, when hydrogels are implanted into the body they may
the DN gel system allows for local remodelling and for the creation of interact with various minerals and biomolecules during prolonged use.
microstructures on the hydrogel surface on demand188, which provides a Using biopolymers as constituents or designing copolymers possessing
facile way to imbue the hydrogel surface with specific properties, such as monomer sequences that feature amino acids from functional proteins
wettability and cell adhesion (Fig. 7c). In such systems, the active mecha- could be possible strategies to solve this issue. For the latter, molecular
noradical concentration plays a crucial role on the reconstruction of the design, notably through machine learning from protein structures, and
ruptured network. DN hydrogels are unusual mechanochemical materi- precise control of the designed monomer sequences are key challenges.
als, as numerous mechanoradicals are generated by the internal fracture Tough hydrogels bring new opportunities beyond their load-
of the brittle first network. The efficiency of the mechanoradical genera- bearing use. Because they are open systems and can be permeated by
tion can be further increased by using weak azoalkane crosslinkers in small molecules, tough hydrogels are promising as force-catalysing
the first network, as the C–N=N bond (azo group connected to saturated materials that use mechanical energy to drive chemical reactions. As a
carbon atoms) in azoalkane has a lower bond dissociation energy than proof of concept, tough DN hydrogels could enable the development
the saturated C–C and C–N bonds in traditional crosslinker189 (Fig. 6b). of mechanochemical-based metabolic-like active materials. These
In common soft materials comprising a single network, the new materials could achieve functions such as self-strengthening,
breaking of covalent bonds often leads to the failure of the entire self-growth or morphogenesis by supplying specific monomers dur-
material. Therefore, weak moieties (for which bond dissociation ing mechanical loading. To develop such adaptive materials — whose
energy is lower than that of the C–C bond), such as difluorenylsuc- structure undergoes repeated destruction and reconstruction under
cinonitrile190, diselenide191 and disulfide192,193, have been introduced load — challenges include the removal of the destroyed residuals and
into soft materials to achieve mechanoradical-induced self-growing the maintaining of the internal fracture characteristics of the DN that
and strengthening (Fig. 6). In polyurethanes, for example, the dis- are essential for material regeneration after many loading cycles. For
sociation of weak difluorenylsuccinonitrile moieties in the main the former, using a reversible chemical reaction pathway could be a
chains190 (Fig. 6c) resulted in the production of pink cyanofluorene solution, whereas for the latter, a fundamental understanding of the
(CF) radicals. These CF radicals initiated the polymerization of side mechanisms of internal destruction and reformation is key. Exploring
chains with methacrylate groups, leading to new chemical crosslink- this new field of soft materials will require interdisciplinary research in
ing in the network. Therefore, the modulus increased markedly with fracture mechanics, polymer physics and polymer chemistry.
increasing mechanical stimuli cycles. Additionally, the generated CF
radicals displayed notable resilience to oxygen exposure, making this Published online: xx xx xxxx
approach well suited for applications in ambient air conditions. This
References
strategy simultaneously incorporates damage reporting (a pink colour 1. El-Sherbiny, I. M. & Yacoub, M. H. Hydrogel scaffolds for tissue engineering: progress and
appears, owing to the generation of CF radicals) and strengthening challenges. Glob. Cardiol. Sci. Pract. 2013, 38 (2013).
2. Nonoyama, T. & Gong, J. P. Tough double network hydrogel and its biomedical
capabilities.
applications. Annu. Rev. Chem. Biomol. Eng. 12, 393–410 (2021).
3. Yuk, H., Wu, J. & Zhao, X. Hydrogel interfaces for merging humans and machines.
Conclusions and outlook Nat. Rev. Mater. 7, 935–952 (2022).
4. Long, R. & Hui, C.-Y. Fracture toughness of hydrogels: measurement and interpretation.
Since the 2000s, various successful designs for tough hydrogels have
Soft Matter 12, 8069–8086 (2016).
been proposed, strengthening the position of hydrogels as promis- 5. Zhao, X. et al. Soft materials by design: unconventional polymer networks give extreme
ing soft and wet materials for diverse applications, especially in properties. Chem. Rev. 121, 4309–4372 (2021).

Nature Reviews Materials


Review article

6. Tang, J., Li, J., Vlassak, J. J. & Suo, Z. Fatigue fracture of hydrogels. Extreme Mech. Lett. 43. Guo, H., Sanson, N., Hourdet, D. & Marcellan, A. Thermoresponsive toughening with
10, 24–31 (2017). crack bifurcation in phase‐separated hydrogels under isochoric conditions. Adv. Mater.
7. Bonn, D., Kellay, H., Prochnow, M., Ben-Djemiaa, K. & Meunier, J. Delayed fracture of an 28, 5857–5864 (2016).
inhomogeneous soft solid. Science 280, 265–267 (1998). 44. Huang, Y. et al. Energy-dissipative matrices enable synergistic toughening in fiber
8. Tanaka, Y., Fukao, K. & Miyamoto, Y. Fracture energy of gels. Eur. Phys. J. E 3, 395–401 reinforced soft composites. Adv. Funct. Mater. 27, 1605350 (2017).
(2000). 45. Li, J., Suo, Z. & Vlassak, J. J. Stiff, strong, and tough hydrogels with good chemical
9. Flory, P. J. & Rehner, J. Jr. Statistical mechanics of cross-linked polymer networks II. stability. J. Mater. Chem. B 2, 6708–6713 (2014).
swelling. J. Chem. Phys. 11, 521–526 (2004). 46. Holloway, J. L., Lowman, A. M. & Palmese, G. R. The role of crystallization and phase
10. Hong, W., Zhao, X., Zhou, J. & Suo, Z. A theory of coupled diffusion and large deformation separation in the formation of physically cross-linked PVA hydrogels. Soft Matter 9,
in polymeric gels. J. Mech. Phys. Solids 56, 1779–1793 (2008). 826–833 (2013).
11. Gong, J. P., Katsuyama, Y., Kurokawa, T. & Osada, Y. Double-network hydrogels with 47. Liang, X. et al. Anisotropically fatigue-resistant hydrogels. Adv. Mater. 33, e2102011
extremely high mechanical strength. Adv. Mater. 15, 1155–1158 (2003). (2021).
12. Sun, J. Y. et al. Highly stretchable and tough hydrogels. Nature 489, 133–136 (2012). 48. Mredha, M. T. I. et al. A facile method to fabricate anisotropic hydrogels with perfectly
13. Sun, T. L. et al. Physical hydrogels composed of polyampholytes demonstrate high aligned hierarchical fibrous structures. Adv. Mater. 30, 1704937 (2018).
toughness and viscoelasticity. Nat. Mater. 12, 932–937 (2013). 49. Zhou, Y. et al. The stiffness-threshold conflict in polymer networks and a resolution.
14. Okumura, Y. & Ito, K. The polyrotaxane gel: a topological gel by figure-of-eight cross-links. J. Appl. Mech. 87, 031002 (2020).
Adv. Mater. 13, 485–487 (2001). 50. Zhang, W. et al. Fatigue of double-network hydrogels. Eng. Fract. Mech. 187, 74–93
15. Liu, C. et al. Tough hydrogels with rapid self-reinforcement. Science 372, 1078–1081 (2018).
(2021). 51. Zhang, W., Hu, J., Yang, H., Suo, Z. & Lu, T. Fatigue-resistant adhesion II: swell tolerance.
16. Nonoyama, T. et al. Instant thermal switching from soft hydrogel to rigid plastics inspired Extreme Mech. Lett. 43, 101182 (2021).
by thermophile proteins. Adv. Mater. 32, e1905878 (2020). 52. Liu, P., Zhang, Y., Guan, Y. & Zhang, Y. Peptide-crosslinked, highly entangled hydrogels
17. Haraguchi, K. & Takehisa, T. Nanocomposite hydrogels: a unique organic–inorganic with excellent mechanical properties but ultra-low solid content. Adv. Mater. 35, 2210021
network structure with extraordinary mechanical, optical, and swelling/de-swelling (2023).
properties. Adv. Mater. 14, 1120 (2002). 53. Lei, H. et al. Stretchable hydrogels with low hysteresis and anti-fatigue fracture based on
18. Hua, M. et al. Strong tough hydrogels via the synergy of freeze-casting and salting out. polyprotein cross-linkers. Nat. Commun. 11, 4032 (2020).
Nature 590, 594–599 (2021). 54. Shi, X. et al. Double hydrogen‐bonding reinforced high‐performance supramolecular
19. Karino, T., Shibayama, M. & Ito, K. Slide-ring gel: topological gel with freely movable hydrogel thermocell for self‐powered sensing remote‐controlled by light. Adv. Funct.
cross-links. Phys. B Condens. Matter 385–386, 692–696 (2006). Mater. 33, 2211720 (2023).
20. Bin Imran, A. et al. Extremely stretchable thermosensitive hydrogels by introducing 55. Xiao, Y. et al. Fatigue of amorphous hydrogels with dynamic covalent bonds. Extreme
slide-ring polyrotaxane cross-linkers and ionic groups into the polymer network. Mech. Lett. 53, 101679 (2022).
Nat. Commun. 5, 5124 (2014). 56. Yang, H., Chen, X., Sun, B., Tang, J. & Vlassak, J. J. Fracture tolerance induced by dynamic
21. Jiang, L. et al. Highly stretchable and instantly recoverable slide-ring gels consisting bonds in hydrogels. J. Mech. Phys. Solids 169, 105083 (2022).
of enzymatically synthesized polyrotaxane with low host coverage. Chem. Mater. 30, 57. Liu, B. et al. Tough and fatigue-resistant polymer networks by crack tip softening.
5013–5019 (2018). Proc. Natl Acad. Sci. USA 120, e2217781120 (2023).
22. Sakai, T. et al. Design and fabrication of a high-strength hydrogel with ideally 58. Li, X. et al. Mesoscale bicontinuous networks in self-healing hydrogels delay fatigue
homogeneous network structure from tetrahedron-like macromonomers. fracture. Proc. Natl. Acad. Sci. U.S.A. 117, 7606–7612 (2020).
Macromolecules 41, 5379–5384 (2008). 59. Zhang, G., Kim, J., Hassan, S. & Suo, Z. Self-assembled nanocomposites of high water
23. Kamata, H., Akagi, Y., Kayasuga-Kariya, Y., Chung, U.-I. & Sakai, T. ‘Nonswellable’ hydrogel content and load-bearing capacity. Proc. Natl Acad. Sci. USA 119, e2203962119 (2022).
without mechanical hysteresis. Science 343, 873–875 (2014). 60. Zheng, Y. et al. Nanophase separation in immiscible double network elastomers induces
24. Ohira, M. et al. Star-polymer-DNA gels showing highly predictable and tunable synergetic strengthening, toughening, and fatigue resistance. Chem. Mater. 33,
mechanical responses. Adv. Mater. 34, e2108818 (2022). 3321–3334 (2021).
25. Shibayama, M., Li, X. & Sakai, T. Precision polymer network science with tetra-PEG 61. Liu, J. et al. Fatigue-resistant adhesion of hydrogels. Nat. Commun. 11, 1071 (2020).
gels — a decade history and future. Colloid Polym. Sci. 297, 1–12 (2018). 62. Xiang, C. et al. Stretchable and fatigue-resistant materials. Mater. Today 34, 7–16 (2020).
26. Kim, J., Zhang, G., Shi, M. & Suo, Z. Fracture, fatigue, and friction of polymers in which 63. Sun, D. et al. Enhance fracture toughness and fatigue resistance of hydrogels by
entanglements greatly outnumber cross-links. Science 374, 212–216 (2021). reversible alignment of nanofibers. ACS Appl. Mater. Interfaces 14, 49389–49397 (2022).
27. Nian, G., Kim, J., Bao, X. & Suo, Z. Making highly elastic and tough hydrogels from 64. Lin, S., Liu, J., Liu, X. & Zhao, X. Muscle-like fatigue-resistant hydrogels by mechanical
doughs. Adv. Mater. 34, e2206577 (2022). training. Proc. Natl Acad. Sci. USA 116, 10244–10249 (2019).
28. Kamiyama, Y. et al. Highly stretchable and self-healable polymer gels from physical 65. Wang, M., Sun, S., Dong, G., Long, F. & Butcher, J. T. Soft, strong, tough, and durable
entanglements of ultrahigh–molecular weight polymers. Sci. Adv. 8, eadd0226 (2022). protein-based fiber hydrogels. Proc. Natl Acad. Sci. USA 120, e2213030120 (2023).
29. Nakajima, T. et al. Tough double-network gels and elastomers from the nonprestretched 66. Su, G. et al. Human-tissue-inspired anti-fatigue-fracture hydrogel for a sensitive
first network. ACS Macro Lett. 8, 1407–1412 (2019). wide-range human–machine interface. J. Mater. Chem. A 8, 2074–2082 (2020).
30. Nakajima, T. et al. A universal molecular stent method to toughen any hydrogels based 67. Liang, X. et al. Bioinspired 2D isotropically fatigue-resistant hydrogels. Adv. Mater. 34,
on double network concept. Adv. Funct. Mater. 22, 4426–4432 (2012). e2107106 (2022).
31. Zheng, Y. et al. In situ and real-time visualization of mechanochemical damage in 68. Jia, L., Wu, S., Yuan, R., Xiang, T. & Zhou, S. Biomimetic microstructured antifatigue
double-network hydrogels by prefluorescent probe via oxygen-relayed radical trapping. fracture hydrogel sensor for human motion detection with enhanced sensing sensitivity.
J. Am. Chem. Soc. 145, 7376–7389 (2023). ACS Appl. Mater. Interfaces 14, 27371–27382 (2022).
32. Gong, J. P. Why are double network hydrogels so tough? Soft Matter 6, 2583–2590 69. Lake, G. & Thomas, A. The strength of highly elastic materials. Proc. R. Soc. Lond. A 300,
(2010). 108–119 (1967).
33. Gong, J. P. Materials both tough and soft. Science 344, 161–162 (2014). 70. Jeon, I., Cui, J., Illeperuma, W. R., Aizenberg, J. & Vlassak, J. J. Extremely stretchable
34. Dai, X. et al. A mechanically strong, highly stable, thermoplastic, and self‐healable and fast self-healing hydrogels. Adv. Mater. 28, 4678–4683 (2016).
supramolecular polymer hydrogel. Adv. Mater. 27, 3566–3571 (2015). 71. Tan, S., Wang, C., Yang, B., Luo, J. & Wu, Y. Precision polymer network science with
35. Hu, X., Vatankhah-Varnoosfaderani, M., Zhou, J., Li, Q. & Sheiko, S. S. Weak hydrogen tetra-PEG gels — a decade history and future. Colloid Polym. Sci. 34, e2206904 (2022).
bonding enables hard, strong, tough, and elastic hydrogels. Adv. Mater. 27, 6899–6905 72. Matsuda, T., Kawakami, R., Namba, R., Nakajima, T. & Gong, J. P. Mechanoresponsive
(2015). self-growing hydrogels inspired by muscle training. Science 363, 504–508 (2019).
36. Wang, Y. J. et al. Ultrastiff and tough supramolecular hydrogels with a dense and robust 73. Xiong, X., Wang, H., Xue, L. & Cui, J. Self-growing organic materials. Angew. Chem. Int. Ed.
hydrogen bond network. Chem. Mater. 31, 1430–1440 (2019). 62, e202306565 (2023).
37. Han, Z. et al. A versatile hydrogel network-repairing strategy achieved by the covalent- 74. Ji, D. & Kim, J. Recent strategies for strengthening and stiffening tough hydrogels.
like hydrogen bond interaction. Sci. Adv. 8, eabl5066 (2022). Adv. Nanobiomed. Res. 1, 2100026 (2021).
38. Lin, P., Ma, S., Wang, X. & Zhou, F. Molecularly engineered dual-crosslinked hydrogel 75. Long, R., Hui, C.-Y., Gong, J. P. & Bouchbinder, E. The fracture of highly deformable
with ultrahigh mechanical strength, toughness, and good self-recovery. Adv. Mater. 27, soft materials: a tale of two length scales. Annu. Rev. Condens. Matter Phys. 12, 71–94
2054–2059 (2015). (2020).
39. Luo, F. et al. Oppositely charged polyelectrolytes form tough, self-healing, and 76. Arruda, E. M. & Boyce, M. C. A three-dimensional constitutive model for the large stretch
rebuildable hydrogels. Adv. Mater. 27, 2722–2727 (2015). behavior of rubber elastic materials. J. Mech. Phys. Solids 41, 389–412 (1993).
40. Cui, K. et al. Multiscale energy dissipation mechanism in tough and self-healing hydrogels. 77. Gent, A. N. A new constitutive relation for rubber. Rubber Chem. Technol. 69, 59–61 (1996).
Phys. Rev. Lett. 121, 185501 (2018). 78. Zhao, X. A theory for large deformation and damage of interpenetrating polymer
41. Cui, K. et al. Phase separation behavior in tough and self-healing polyampholyte networks. J. Mech. Phys. Solids 60, 319–332 (2012).
hydrogels. Macromolecules 53, 5116–5126 (2020). 79. Richbourg, N. R. & Peppas, N. A. The swollen polymer network hypothesis: quantitative
42. Li, X. et al. Effect of mesoscale phase contrast on fatigue-delaying behavior of self-healing models of hydrogel swelling, stiffness, and solute transport. Prog. Polym. Sci. 105,
hydrogels. Sci. Adv. 7, eabe8210 (2021). 101243 (2020).

Nature Reviews Materials


Review article

80. Melly, S. K., Liu, L., Liu, Y. & Leng, J. A review on material models for isotropic 117. Tanaka, Y. A local damage model for anomalous high toughness of double-network gels.
hyperelasticity. Int. J. Mech. Syst. Dyn. 1, 71–88 (2021). Europhys. Lett. 78, 56005 (2007).
81. Boyce, M. C. & Arruda, E. M. Constitutive models of rubber elasticity: a review. Rubber 118. Matsuda, T., Kawakami, R., Nakajima, T. & Gong, J. P. Crack tip field of a double-network
Chem. Technol. 73, 504–523 (2000). gel: visualization of covalent bond scission through mechanoradical polymerization.
82. Marckmann, G. & Verron, E. Comparison of hyperelastic models for rubber-like materials. Macromolecules 53, 8787–8795 (2020).
Rubber Chem. Technol. 79, 835–858 (2006). 119. Ducrot, E., Chen, Y., Bulters, M., Sijbesma, R. P. & Creton, C. Toughening elastomers with
83. Rubinstein, M. & Colby, R. H. Polymer Physics (Oxford Univ. Press, 2003). sacrificial bonds and watching them break. Science 344, 186–189 (2014).
84. Zheng, Y., Nakajima, T., Cui, W., Hui, C.-Y. & Gong, J. P. Swelling effect on the yielding, 120. Chen, Y., Yeh, C. J., Qi, Y., Long, R. & Creton, C. From force-responsive molecules to
elasticity, and fracture of double-network hydrogels with an inhomogeneous first quantifying and mapping stresses in soft materials. Sci. Adv. 6, eaaz5093 (2020).
network. Macromolecules 56, 3962–3972 (2023). 121. Slootman, J. et al. Quantifying rate- and temperature-dependent molecular damage in
85. Hoshino, K. I., Nakajima, T., Matsuda, T., Sakai, T. & Gong, J. P. Network elasticity of a elastomer fracture. Phys. Rev. X 10, 041045 (2020).
model hydrogel as a function of swelling ratio: from shrinking to extreme swelling states. 122. Slootman, J., Yeh, C. J., Millereau, P., Comtet, J. & Creton, C. A molecular interpretation of the
Soft Matter 14, 9693–9701 (2018). toughness of multiple network elastomers at high temperature. Proc. Natl Acad. Sci. USA
86. Matsuda, T., Kawakami, R., Nakajima, T., Hane, Y. & Gong, J. P. Revisiting the origins of the 119, e2116127119 (2022).
fracture energy of tough double-network hydrogels with quantitative mechanochemical 123. Zhao, X. EML webinar overview: extreme mechanics of soft materials for merging
characterization of the damage zone. Macromolecules 54, 10331–10339 (2021). human-machine intelligence. Extreme Mech. Lett. 39, 100784 (2020).
87. Sun, T. L. et al. Bulk energy dissipation mechanism for the fracture of tough and 124. Zheng, Y. et al. How chain dynamics affects crack initiation in double-network gels.
self-healing hydrogels. Macromolecules 50, 2923–2931 (2017). Proc. Natl Acad. Sci. USA 118, e2111880118 (2021).
88. de Gennes, P.-G. Soft adhesives. Langmuir 12, 4497–4500 (1996). 125. Zheng, Y., Wang, Y., Nakajima, T. & Gong, J. P. Effect of predamage on the fracture energy
89. Chen, C., Wang, Z. & Suo, Z. Flaw sensitivity of highly stretchable materials. Extreme of double-network hydrogels. ACS Macro Lett. 13, 130–137 (2024).
Mech. Lett. 10, 50–57 (2017). 126. Tanaka, Y. et al. Determination of fracture energy of high strength double network
90. Mayumi, K. & Ito, K. Structure and dynamics of polyrotaxane and slide-ring materials. hydrogels. J. Phys. Chem. B 109, 11559–11562 (2005).
Polymer 51, 959–967 (2010). 127. Nakajima, T. et al. True chemical structure of double network hydrogels. Macromolecules
91. Griffith, A. A. V. I. The phenomena of rupture and flow in solids. Phil. Trans. R. Soc. Lond. A 42, 2184–2189 (2009).
221, 163–198 (1921). 128. Wang, W., Narain, R. & Zeng, H. Rational design of self-healing tough hydrogels: a mini
92. Bai, R., Yang, J. & Suo, Z. Fatigue of hydrogels. Eur. J. Mech. A 74, 337–370 (2019). review. Front. Chem. 6, 497 (2018).
93. Creton, C. & Ciccotti, M. Fracture and adhesion of soft materials: a review. Rep. Prog. Phys. 129. Wang, S. & Urban, M. W. Self-healing polymers. Nat. Rev. Mater. 5, 562–583 (2020).
79, 046601 (2016). 130. Cui, J. & del Campo, A. Multivalent H-bonds for self-healing hydrogels. Chem. Commun.
94. Brannon-Peppas, L. & Peppas, N. A. Equilibrium swelling behavior of pH-sensitive 48, 9302–9304 (2012).
hydrogels. Chem. Eng. Sci. 46, 715–722 (1991). 131. Beijer, F. H., Sijbesma, R. P., Kooijman, H., Spek, A. L. & Meijer, E. W. Strong dimerization of
95. Tang, J. et al. Swelling behaviors of hydrogels with alternating neutral/highly charged ureidopyrimidones via quadruple hydrogen bonding. J. Am. Chem. Soc. 120, 6761–6769
sequences. Macromolecules 53, 8244–8254 (2020). (1998).
96. Sun, T. L. et al. Molecular structure of self-healing polyampholyte hydrogels analyzed 132. Zhang, X. N. et al. A tough and stiff hydrogel with tunable water content and mechanical
from tensile behaviors. Soft matter 11, 9355–9366 (2015). properties based on the synergistic effect of hydrogen bonding and hydrophobic
97. Fan, H., Wang, J. & Jin, Z. Tough, swelling-resistant, self-healing, and adhesive dual- interaction. Macromolecules 51, 8136–8146 (2018).
cross-linked hydrogels based on polymer–tannic acid multiple hydrogen bonds. 133. Zhang, W. et al. Fracture toughness and fatigue threshold of tough hydrogels.
Macromolecules 51, 1696–1705 (2018). ACS Macro Lett. 8, 17–23 (2018).
98. Canal, T. & Peppas, N. A. Correlation between mesh size and equilibrium degree of 134. Zhou, X. et al. Shape morphing of anisotropy-encoded tough hydrogels enabled by
swelling of polymeric networks. J. Biomed. Mater. Res. 23, 1183–1193 (2004). asymmetrically-induced swelling and site-specific mechanical strengthening. J. Mater.
99. Mussault, C., Guo, H., Sanson, N., Hourdet, D. & Marcellan, A. Effect of responsive graft Chem. B 6, 4731–4737 (2018).
length on mechanical toughening and transparency in microphase-separated hydrogels. 135. Ye, Y. N. et al. Molecular mechanism of abnormally large nonsoftening deformation in a
Soft Matter 15, 8653–8666 (2019). tough hydrogel. Proc. Natl Acad. Sci. USA 118, e2014694118 (2021).
100. Hartquist, C. M. et al. An elastomer with ultrahigh strain-induced crystallization. Sci. Adv. 136. Zhang, H. J. et al. Tough physical double-network hydrogels based on amphiphilic
9, eadj0411 (2023). triblock copolymers. Adv. Mater. 28, 4884–4890 (2016).
101. Matsunaga, T., Sakai, T., Akagi, Y., Chung, U.-I. & Shibayama, M. Structure characterization 137. Ihsan, A. B. et al. Self-healing behaviors of tough polyampholyte hydrogels.
of tetra-PEG gel by small-angle neutron scattering. Macromolecules 42, 1344–1351 Macromolecules 49, 4245–4252 (2016).
(2009). 138. Cui, K. et al. Effect of structure heterogeneity on mechanical performance of physical
102. Li, X., Nakagawa, S., Tsuji, Y., Watanabe, N. & Shibayama, M. Polymer gel with a flexible polyampholytes hydrogels. Macromolecules 52, 7369–7378 (2019).
and highly ordered three-dimensional network synthesized via bond percolation. 139. Li, X. et al. Role of hierarchy structure on the mechanical adaptation of self-healing
Sci. Adv. 5, eaax8647 (2019). hydrogels under cyclic stretching. Sci. Adv. 9, eadj6856 (2023).
103. Li, X. A benchmark for gel structures: bond percolation enables the fabrication of 140. Li, X. & Gong, J. P. Role of dynamic bonds on fatigue threshold of tough hydrogels.
extremely homogeneous gels. Polym. J. 53, 765–777 (2021). Proc. Natl Acad. Sci. USA 119, e2200678119 (2022).
104. Sugimura, A. et al. Mechanical properties of a polymer network of tetra-PEG gel. Polym. J. 141. Li, X. et al. Effect of salt on dynamic mechanical behaviors of polyampholyte hydrogels.
45, 300–306 (2012). Macromolecules 56, 535–544 (2023).
105. Akagi, Y., Sakurai, H., Gong, J. P., Chung, U.-I. & Sakai, T. Fracture energy of polymer gels 142. Spruijt, E., Sprakel, J., Lemmers, M., Stuart, M. A. & van der Gucht, J. Relaxation dynamics
with controlled network structures. J. Chem. Phys. 139, 144905 (2013). at different time scales in electrostatic complexes: time-salt superposition. Phys. Rev.
106. Mayumi, K., Liu, C., Yasuda, Y. & Ito, K. Softness, elasticity, and toughness of polymer Lett. 105, 208301 (2010).
networks with slide-ring cross-links. Gels 7, 91 (2021). 143. Hassan, C. M. & Peppas, N. A. in Biopolymers: PVA Hydrogels, Anionic Polymerisation
107. Norioka, C., Inamoto, Y., Hajime, C., Kawamura, A. & Miyata, T. A universal method to Nanocomposites (eds Abe, A. et al.) 37–65 (Springer, 2000).
easily design tough and stretchable hydrogels. NPG Asia Mater. 13, 34 (2021). 144. Zhang, R. et al. Stretch-induced complexation reaction between poly (vinyl alcohol)
108. Wang, S. et al. Facile mechanochemical cycloreversion of polymer cross-linkers and iodine: an in situ synchrotron radiation small- nd wide-angle X-ray scattering study.
enhances tear resistance. Science 380, 1248–1252 (2023). Soft Matter 14, 2535–2546 (2018).
109. Wang, S., Panyukov, S., Craig, S. L. & Rubinstein, M. Contribution of unbroken strands 145. Wu, Y. et al. Solvent-exchange assisted wet-annealing: a new strategy for super-strong,
to the fracture of polymer networks. Macromolecules 56, 2309–2318 (2023). tough, stretchable and anti-fatigue hydrogels. Adv. Mater. 35, e2210624 (2023).
110. Lin, S. & Zhao, X. Fracture of polymer networks with diverse topological defects. 146. Lin, S. & Zhao, X. Nanostructured artificial-muscle fibres. Nat. Nanotechnol. 17, 677–678
Phys. Rev. E 102, 052503 (2020). (2022).
111. Na, Y.-H. et al. Necking phenomenon of double-network gels. Macromolecules 39, 147. Wu, S. et al. Poly(vinyl alcohol) hydrogels with broad-range tunable mechanical
4641–4645 (2006). properties via the hofmeister effect. Adv. Mater. 33, e2007829 (2021).
112. Webber, R. E., Creton, C., Brown, H. R. & Gong, J. P. Large strain hysteresis and Mullins 148. Cui, W. et al. Strong tough conductive hydrogels via the synergy of ion‐induced
effect of tough double-network hydrogels. Macromolecules 40, 2919–2927 (2007). cross‐linking and salting‐out. Adv. Funct. Mater. 32, 2204823 (2022).
113. Fukao, K. et al. Effect of relative strength of two networks on the internal fracture 149. Zhang, Q. et al. Stretch-induced structural evolution of poly (vinyl alcohol) film in water
process of double network hydrogels as revealed by in situ small-angle X-ray scattering. at different temperatures: an in-situ synchrotron radiation small-and wide-angle X-ray
Macromolecules 53, 1154–1163 (2020). scattering study. Polymer 142, 233–243 (2018).
114. Matsuda, T. et al. Yielding criteria of double network hydrogels. Macromolecules 49, 150. Lin, S. et al. Anti-fatigue-fracture hydrogels. Sci. Adv. 5, eaau8528 (2019).
1865–1872 (2016). 151. Zhou, W. et al. Toughening mystery of natural rubber deciphered by double network
115. Yu, Q. M., Tanaka, Y., Furukawa, H., Kurokawa, T. & Gong, J. P. Direct observation of damage incorporating hierarchical structures. Sci. Rep. 4, 7502 (2014).
zone around crack tips in double-network gels. Macromolecules 42, 3852–3855 (2009). 152. Katz, J. Röntgenspektrographische Untersuchungen am gedehnten Kautschuk und ihre
116. Brown, H. R. A model of the fracture of double network gels. Macromolecules 40, mögliche Bedeutung für das Problem der Dehnungseigenschaften dieser Substanz.
3815–3818 (2007). Naturwissenschaften 13, 410–416 (1925).

Nature Reviews Materials


Review article

153. Trabelsi, S., Albouy, P. A. & Rault, J. Stress-induced crystallization around a crack tip in 187. Wei, G. et al. Sustainable mechanochemical growth of double-network hydrogels
natural rubber. Macromolecules 35, 10054–10061 (2002). supported by vascular-like perfusion. Mater. Horiz. 10, 4882–4891 (2023).
154. Haque, M. A., Kamita, G., Kurokawa, T., Tsujii, K. & Gong, J. P. Unidirectional alignment of 188. Mu, Q. et al. Force-triggered rapid microstructure growth on hydrogel surface for
lamellar bilayer in hydrogel: one-dimensional swelling, anisotropic modulus, and stress/ on-demand functions. Nat. Commun. 13, 6213 (2022).
strain tunable structural color. Adv. Mater. 22, 5110–5114 (2010). 189. Wang, Z. J. et al. Azo-crosslinked double-network hydrogels enabling highly efficient
155. Haque, M. A., Kurokawa, T., Kamita, G., Yue, Y. & Gong, J. P. Rapid and reversible tuning of mechanoradical generation. J. Am. Chem. Soc. 144, 3154–3161 (2022).
structural color of a hydrogel over the entire visible spectrum by mechanical stimulation. 190. Seshimo, K. et al. Segmented polyurethane elastomers with mechanochromic and
Chem. Mater. 23, 5200–5207 (2011). self-strengthening functions. Angew. Chem. Int. Ed. 60, 8406–8409 (2021).
156. Haque, M. A. et al. Lamellar bilayer to fibril structure transformation of tough photonic 191. Li, X. et al. Diselenide as a dual functional mechanophore capable of stress self-reporting
hydrogel under elongation. Macromolecules 53, 4711–4721 (2020). and self-strengthening in polyurethane elastomers. CCS Chem. 5, 925–933 (2023).
157. Haque, M. A., Kurokawa, T., Kamita, G. & Gong, J. P. Lamellar bilayers as reversible 192. Kida, J., Aoki, D. & Otsuka, H. Self-strengthening of cross-linked elastomers via the use
sacrificial bonds to toughen hydrogel: hysteresis, self-recovery, fatigue resistance, and of dynamic covalent macrocyclic mechanophores. ACS Macro Lett. 10, 558–563 (2021).
crack blunting. Macromolecules 44, 8916–8924 (2011). 193. Yang, Y. et al. Self-strengthening, self-welding, shape memory, and recyclable
158. Yue, Y. et al. Mechano-actuated ultrafast full-colour switching in layered photonic polybutadiene-based material driven by dual-dynamic units. ACS Appl. Mater. Interfaces
hydrogels. Nat. Commun. 5, 4659 (2014). 14, 3344–3355 (2022).
159. Yue, Y., Norikane, Y. & Gong, J. P. Ultrahigh‐water‐content photonic hydrogels with large 194. Akagi, Y., Gong, J. P., Chung, U.-I. & Sakai, T. Transition between phantom and
electro‐optic responses in visible to near‐infrared region. Adv. Optical Mater. 9, 2002198 affine network model observed in polymer gels with controlled network structure.
(2021). Macromolecules 46, 1035–1040 (2013).
160. Chen, W., Zhang, Z. & Kouwer, P. H. J. Magnetically driven hierarchical alignment in 195. Ito, K. Slide-ring materials using topological supramolecular architecture. Curr. Opin.
biomimetic fibrous hydrogels. Small 18, e2203033 (2022). Solid State Mater. Sci. 14, 28–34 (2010).
161. Li, H.-J., Jiang, H. & Haraguchi, K. Ultrastiff, thermoresponsive nanocomposite hydrogels 196. Nakajima, T., Kurokawa, T., Furukawa, H. & Gong, J. P. Effect of the constituent networks
composed of ternary polymer–clay–silica networks. Macromolecules 51, 529–539 (2018). of double-network gels on their mechanical properties and energy dissipation process.
162. Kamio, E., Yasui, T., Iida, Y., Gong, J. P. & Matsuyama, H. Inorganic/organic double- Soft Matter 16, 8618–8627 (2020).
network gels containing ionic liquids. Adv. Mater. 29, 1704118 (2017). 197. Bai, R. et al. Fatigue fracture of tough hydrogels. Extreme Mech. Lett. 15, 91–96 (2017).
163. Pan, C., Wang, J., Ji, X. & Liu, L. Stretchable, compressible, self-healable carbon 198. Yu, H. C. et al. Reversibly transforming a highly swollen polyelectrolyte hydrogel to an
nanotube mechanically enhanced composite hydrogels with high strain sensitivity. extremely tough one and its application as a tubular grasper. Adv. Mater. 32, e2005171
J. Mater. Chem. C 8, 1933–1942 (2020). (2020).
164. Li, S.-N. et al. Constructing dual ionically cross-linked poly(acrylamide-co-acrylic acid)/ 199. Guo, Y. Z. et al. Facile preparation of cellulose hydrogel with Achilles tendon-like super
chitosan hydrogel materials embedded with chitosan decorated halloysite nanotubes strength through aligning hierarchical fibrous structure. Chem. Eng. J. 428, 132040
for exceptional mechanical performance. Compos. B Eng. 194, 108046 (2020). (2022).
165. Zhu, X., Zhang, W., Lu, G., Zhao, H. & Wang, L. Ultrahigh mechanical strength and robust 200. Nonoyama, T. Robust hydrogel–bioceramics composite and its osteoconductive properties.
room-temperature self-healing properties of a polyurethane-graphene oxide network Polym. J. 52, 709–716 (2020).
resulting from multiple dynamic bonds. ACS Nano 16, 16724–16735 (2022). 201. Zhang, E., Bai, R., Morelle, X. P. & Suo, Z. Fatigue fracture of nearly elastic hydrogels.
166. Liang, X. et al. Impact-resistant hydrogels by harnessing 2D hierarchical structures. Soft Matter 14, 3563–3571 (2018).
Adv. Mater. 35, e2207587 (2023). 202. Zhang, W., Gao, Y., Yang, H., Suo, Z. & Lu, T. Fatigue-resistant adhesion I. Long-chain
167. Haraguchi, K. Synthesis and properties of soft nanocomposite materials with novel polymers as elastic dissipaters. Extreme Mech. Lett. 39, 100813 (2020).
organic/inorganic network structures. Polym. J. 43, 223–241 (2011). 203. Sakai, T. Gelation mechanism and mechanical properties of tetra-PEG gel. React. Funct.
168. Nepal, D. et al. Hierarchically structured bioinspired nanocomposites. Nat. Mater. 22, Polym. 73, 898–903 (2013).
18–35 (2023). 204. Lin, S., Londono, C. D., Zheng, D. & Zhao, X. An extreme toughening mechanism for soft
169. Sano, K., Ishida, Y. & Aida, T. Synthesis of anisotropic hydrogels and their applications. materials. Soft Matter 18, 5742–5749 (2022).
Angew. Chem. Int. Ed. 57, 2532–2543 (2018). 205. Xue, B. et al. Strong, tough, rapid-recovery, and fatigue-resistant hydrogels made of
170. Ji, D., Nguyen, T. L. & Kim, J. Bioinspired structural composite hydrogels with a picot peptide fibres. Nat. Commun. 14, 2583 (2023).
combination of high strength, stiffness, and toughness. Adv. Funct. Mater. 31, 2101095 206. Guo, X. et al. Stretchable hydrogels with low hysteresis and high fracture toughness for
(2021). flexible electronics. Macromol. Rapid Commun. 43, e2100716 (2022).
171. Ning, J., Li, G. & Haraguchi, K. Synthesis of highly stretchable, mechanically tough, 207. Liu, X. et al. Fatigue-resistant hydrogel optical fibers enable peripheral nerve
zwitterionic sulfobetaine nanocomposite gels with controlled thermosensitivities. optogenetics during locomotion. Nat. Methods 20, 1802–1809 (2023).
Macromolecules 46, 5317–5328 (2013). 208. Ju, Y.-X. et al. Strong silk fibroin/PVA/chitosan hydrogels with high water content inspired
172. Klein, A., Whitten, P. G., Resch, K. & Pinter, G. Nanocomposite hydrogels: fracture by straw rammed earth brick structures. ACS Sustain. Chem. Eng. 10, 13070–13080
toughness and energy dissipation mechanisms. J. Polym. Sci. B 53, 1763–1773 (2015). (2022).
173. Haraguchi, K., Farnworth, R., Ohbayashi, A. & Takehisa, T. Compositional effects 209. Liangsong, Z. et al. Flaw-insensitive fatigue resistance of chemically fixed collagenous
on mechanical properties of nanocomposite hydrogels composed of poly soft tissues. Sci. Adv. 9, eade7375 (2023).
(N,N-dimethylacrylamide) and clay. Macromolecules 36, 5732–5741 (2003).
174. Wang, T. et al. Large deformation behavior and effective network chain density of swollen Acknowledgements
poly(N-isopropylacrylamide)–laponite nanocomposite hydrogels. Soft Matter 8, 774–783 This work was supported by JSPS KAKENHI (grant nos. JP22H04968, JP22K21342, JP22K20521
(2012). and JP23K13796).
175. Wang, Z. et al. Stretchable materials of high toughness and low hysteresis. Proc. Natl
Acad. Sci. USA 116, 5967–5972 (2019). Author contributions
176. Hui, C.-Y., Liu, Z. & Phoenix, S. L. Size effect on elastic stress concentrations in Both authors contributed to the discussion, writing, reviewing and editing of the manuscript.
unidirectional fiber reinforced soft composites. Extreme Mech. Lett. 33, 100573 (2019).
177. Cui, W. et al. Fiber-reinforced viscoelastomers show extraordinary crack resistance that Competing interests
exceeds metals. Adv. Mater. 32, e1907180 (2020). The authors declare no competing interests.
178. Agrawal, A., Rahbar, N. & Calvert, P. D. Strong fiber-reinforced hydrogel. Acta Biomater. 9,
5313–5318 (2013). Additional information
179. King, D. R. et al. Extremely tough composites from fabric reinforced polyampholyte Supplementary information The online version contains supplementary material available at
hydrogels. Mater. Horiz. 2, 584–591 (2015). https://doi.org/10.1038/s41578-024-00672-3.
180. Bai, R., Yang, J., Morelle, X. P., Yang, C. & Suo, Z. Fatigue fracture of self-recovery
hydrogels. ACS Macro Lett. 7, 312–317 (2018). Peer review information Nature Reviews Materials thanks Kozho Ito; Shaoting Lin, who
181. Lin, S., Ni, J., Zheng, D. & Zhao, X. Fracture and fatigue of ideal polymer networks. co-reviewed with Zhaohan Yu; and the other, anonymous, reviewer(s) for their contribution
Extreme Mech. Lett. 48, 101399 (2021). to the peer review of this work.
182. Lei, Z., Gao, W., Zhu, W. & Wu, P. Anti‐fatigue and highly conductive thermocells for
continuous electricity generation. Adv. Funct. Mater. 32, 2201021 (2022). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
183. Ni, J. et al. Strong fatigue-resistant nanofibrous hydrogels inspired by lobster underbelly. published maps and institutional affiliations.
Matter 4, 1919–1934 (2021).
184. Li, C., Yang, H., Suo, Z. & Tang, J. Fatigue-resistant elastomers. J. Mech. Phys. Solids 134, Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
103751 (2020). article under a publishing agreement with the author(s) or other rightsholder(s); author self-
185. Ramirez, A. L. et al. Mechanochemical strengthening of a synthetic polymer in response archiving of the accepted manuscript version of this article is solely governed by the terms of
to typically destructive shear forces. Nat. Chem. 5, 757–761 (2013). such publishing agreement and applicable law.
186. Schoenfeld, B. J. The mechanisms of muscle hypertrophy and their application to
resistance training. J. Strength. Cond. Res. 24, 2857–2872 (2010). © Springer Nature Limited 2024

Nature Reviews Materials

You might also like