Reviews: Self-Healing Polymers

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

REVIEWS

Self-healing polymers
Siyang Wang    and Marek W. Urban    ✉
Abstract | Self-healing is the capability of a material to recover from physical damage.
Both physical and chemical approaches have been used to construct self-healing polymers.
These include diffusion and flow, shape-memory effects, heterogeneous self-healing systems,
covalent-bond reformation and reshuffling, dynamics of supramolecular chemistry or combinations
thereof. In this Review, we discuss the similarities and differences between approaches to achieve
self-healing in synthetic polymers, where possible placing this discussion in the context of
biological systems. In particular, we highlight the role of thermal transitions, network heterogeneities,
localized chemical reactions enabling the reconstruction of damage and physical reshuffling.
We also discuss energetic and length-scale considerations, as well as scientific and technological
challenges and opportunities.

Biological organisms have built-in repair mechanisms repair damage19 — and cardiovascular networks20, which
to prevent them from losing their functions. The repair use the same concept.
processes in mammals and plants occur in entirely dif- In this Review, we outline the physical, chemical and
ferent chemical and morphological environments, yet physico-chemical processes of self-healable polymers.
— in a general sense — the outcomes are similar. For We discuss how leveraging advances in synthetic mate-
example, DNA damage is a fairly common event in a rials and biological systems, while using feedback and
cell’s life that may lead to DNA mutation, uncontrollable feedforward from physico-chemical analysis and predic-
growth (cancer) or cell death. In mammals, the key com- tive computational algorithms, will lead to discoveries
ponents are pro-inflammatory cytokines, transforming and technological advances. Taking self-healing materi-
growth factors and angiogenic factors1,2. Human skin als to the next level, we discuss how exchangeable bonds
self-heals via an inflammatory response of cells below triggered by thermal, chemical or other stimuli result in
the dermis by increasing collagen production, which the development of tunable rigid or soft vitrimers.
regenerates epithelial cells and tissue. In plants, oligo-
peptides, oligosaccharides or other molecules induce Interchain diffusion
changes that signal damage and initiate a sequence Early approaches for crack healing in thermoplastic poly­
of chemical events leading to macroscopic repair3,4. mers can be broken into five stages: segmental surface
Regardless of the individual steps in any of these pro- rearrangements, surface approach, wetting, diffusion
cesses, self-healing in living systems involves a cascade and randomization5,21. During surface rearrangements,
of reactions, the exact chemistries of which are far from factors such as topography and roughness of the sur-
understood. faces, chain-end distribution and molecular-weight dis-
The main approaches to self-healable polymers tribution come into play. As two surfaces come together
involve either physical or chemical events at the molec- to enable subsequent molecular-level physical and/or
ular level, although there is overlap between the two chemical self-healing21, they form an interface and wet
approaches (Fig. 1). Examples of physical self-healing each other before diffusion occurs. Various chemical
processes are interchain diffusion5, phase-separated rebonding techniques in thermosetting and thermoplas-
morphologies6,7, shape-memory effects8 and the intro- tic self-repairing polymers have supplemented the dif-
duction of superparamagnetic nanoparticles9. By con- fusion phase22, but differentiation between the physical
Department of Materials
Science and Engineering, trast, predominantly chemical processes include the and chemical processes involved is not trivial.
Center for Optical Materials incorporation of covalent10–12, free-radical13,14 or supra- Mechanical damage creates interfacial regions. Local
Science and Engineering molecular 7,15–17 dynamic bonds. Many self-healing mobility and diffusion rates in damaged areas (especially
Technologies (COMSET), processes involve a combination of physical and chem- in interfacial regions) are important in the self-healing
Clemson University, Clemson,
SC, USA.
ical events, such as taking advantage of enhanced process23. Interfacial macromolecular interpenetration
✉e-mail: mareku@ van der Waals forces18, resulting in interdigitated copoly- was proposed in the 1960s (ref.24) and typical diffusion
clemson.edu mer morphologies — embedded, reactive, encapsulated rates in solid-state polymers are 10−5 m min−1 (ref.25).
https://doi.org/10.1038/ fluids that burst open upon damage to fill up a wound Furthermore, full recovery of mechanical strength is
s41578-020-0202-4 and trigger chemical reactions of reactive agents to approximately 0.4–0.8 times the radius of gyration26–28,

562 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a Physical approaches b Chemical approaches


Interfacial regions
Covalent rebonding
Reactive
chain ends
Free-radical rebonding

Interdiffusion Phase-separated morphologies


H-bonding
Damage Damage π–π stacking

Supra- Guest–host chemistry


molecular
chemistry

Metal–ligand coordination
γ-Fe2O3
Shape-memory recovery Melting interdiffusion
Ionic interactions
c Physico-chemical approaches

van der Waals interactions Encapsulation Cardiovascular network

Fig. 1 | Self-healing mechanisms. a | Physical processes to realize self-healing include interdiffusion of polymer chains,
the introduction of phase-separated morphologies, shape-memory effects and the introduction of active nanoparticles
into a polymer matrix. b | Chemical processes to facilitate self-healing involve either introducing reactive chain ends or
supramolecular chemistries. c | Physical and chemical processes can be combined to realize self-healing. Self-healing
is achieved by incorporating enhanced van der Waals interactions, or encapsulating nanocapsules or microcapsules
containing reactive liquids to heal a wound, or by mimicking cardiovascular architectures composed of hollow fibres filled
with reactive chemicals to heal a polymer matrix.

and the average interdiffusion depth is expressed as a owing to the excess free volume. Because newly formed
function of time X(t) by surfaces or interfaces resulting from damage exhibit
liquid-like attributes, relatively slow diffusion rates
X (t ) = X ∞(t / Tr )1/4 (1) (~10−3 m min−1) are anticipated39. Under these con-
ditions, the reptation model (that is, a polymer chain
where Tr is the reptation time, which is proportional migrating through a tube; Eq. 1) could serve to predict
to the chain molecular weight to the third power21,28,29. topological constraints imposed on a polymer back-
Accordingly, more flexible and shorter chains are more bone by the surrounding polymer chains37 in the bulk.
mobile and facilitate recovery more easily22. In contrast However, this model may not be applicable in the sur-
to the bulk, surface macromolecules exhibit high mobil- face and/or interfacial regions, where interfacial energy
ity, reflected in a lower glass-transition temperature may dominate chain mobility.
(Tg), owing to their high degrees of freedom21,30–32, thus
facilitating enhanced diffusion. This is supported by the van der Waals interactions
observation that the Tg inside a fresh cut is lower than van der Waals interactions have been of interest for
that in an undamaged surface33. over two centuries, and the presence of van der Waals
Fickian diffusion is enhanced above the Tg, owing interchain forces in polymers was established dec-
to the excess free volume, driven by entanglement ades ago. An illustrative biological system that uses
coupling and reptation34–36. However, crack healing in van der Waals forces to adhere to any surface is gecko
poly(methyl methacrylate) (pMMA) was achieved by setae 40. However, only recently have such interac-
heating above the Tg under pressure, which facilitates tions been recognized for their potential in designing
a recovery of the fracture stresses proportional to t¼, self-healing commodity copolymers18. If perturbation
where t is the heating time25,37. In general, lowering the of van der Waals forces upon mechanical damage is
Tg enhances segmental chain mobility (which can also energetically unfavourable, interdigitated alternating
be achieved through the use of plasticizing solvents25,38), or random copolymer motifs will self-heal to an ener-
thus promoting diffusion and inducing conformational getically more favourable state. Molecular dynamics
changes, but does not assure self-healing. When liquid is simulations showed that the formation of helix-like
dispersed in a solid matrix, diffusion favours the repair conformations depends on the copolymer composition
process, which will be enhanced for lower-Tg networks, and creates a viscoelastic response that energetically

NAture Reviews | MaterialS volume 5 | August 2020 | 563


Reviews

favours self-recovery upon chain separation, owing Recent studies suggest that this is governed by a combi-
to ‘key-and-lock’ associations of the neighbouring nation of hydraulic shrinking and swelling, which are
chains. In essence, van der Waals forces stabilize neigh- the main driving forces and growth-induced mechanical
bouring copolymers, which is reflected in enhanced prestresses in plant tissues47.
cohesive-energy density (CED) values. Figure 2a illus- In synthetic polymers, the shape-memory effect was
trates how induced dipole interactions for alternating discovered in the 1940s and first used in dental materials
or random poly(methyl methacrylate-alt-ran-n-butyl (methacrylic ester resin)48. In the 1960s, this discovery
acrylate) (p(MMA-alt-ran-nBA)) copolymers owing to was followed by the development of heat-shrinkable
directional van der Waals forces may enhance the CED polyethylene in films, tubing and other applications49,50.
at equilibrium (CEDeq) of entangled and side-by-side The response of shape-memory materials to external
copolymer chains. It is a challenge to measure van der stimuli was largely neglected as part of self-healing
Waals forces, and molecular-dynamics simulations are processes. However, if designed properly, polymers can
valuable for understanding and quantifying their role in ‘memorize’ a permanent shape that can be manipulated
self-healing. For that reason, various force fields for small to create a temporary form, and, under suitable condi-
molecules41,42 and polymers43,44 are useful. Although tions triggered by external or internal stimuli (for exam-
over the years they have been modified to predict ple, heat, light or deformation), transform back to the
polymer–polymer45 and polymer–inorganic interac- memorized permanent shape. Such responses manifest
tions46, there is a need for more precise dynamic modelling in conformational changes and/or chain contractions,
of damage–repair cycles in polymers. which are typically entropy-driven, resulting in mass
flow and self-healing51. One example is light-activated
Shape memory shape-memory polymers, which use one wavelength of
Shape-memory-assisted self-healing is commonly light for photocrosslinking, while a second wavelength
observed in biological systems. An example of this is cleaves the photocrosslinked bonds to enable reversible
wound closure in leaves, whereby, after a transversal switching between elastomers52. Another example is
incision, the entire leaf bends until the wound is closed. pretensioned shape-memory alloy wires53 or fibres54 in

a Distribution
of induced
dipoles
n
O O Load δ+ δ+
O O
δ–

δ+
δ–
Poly(methyl methacrylate- δ–
δ +
δ–
co-n-butyl acrylate)

δ–
δ+ C
O
H
b Crystalline
nanodomain
E–E segment of E–E segments
E-alt-AP segment
SiMe3 SiMe3

Sc+ Sc+
O

O R

Fig. 2 | Self-healing through van der Waals forces or the shape-memory effect. a | Key-and-lock configuration in
self-healing poly(methyl methacrylate-co-n-butyl acrylate) copolymer. Self-healing occurs in a narrow compositional range
for copolymer topologies that are preferentially alternating with a random component and is attributed to favourable van der
Waals forces between the polymer chains, forming key-and-lock interchain junctions. b | A self-healing and phase-separated
copolymer system comprising ethylene(E)-alt-anisyl propylene (AP). Copolymers with glass-transition temperatures below
room temperature are self-healable; however, the shape-memory effect is observed for copolymers with glass-transition
temperatures around or above room temperature. Panel b (left) adapted with permission from ref.59, ACS.

564 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a polymer matrix. Upon activation at elevated temper- clotting in a wound, but a technological challenge is
ature, self-healing occurs owing to forces pulling crack achieving high stability of the capsules during processing
surfaces towards each other. and overcoming the limitation of one-time self-repair.
Incorporating soft and hard segments into one A somewhat analogous concept to develop fibre-
copoly­mer can lead to phase-separated morphologies. reinforced composites that retard failure is based on
For example, a material that combines rubber elasticity cardiovascular networks in which hollow fibres filled
and thermoplastic stiffness will expand the applications with reactive ingredients serve as delivery systems of
of copolymers owing to enhanced mechanical properties chemicals to the wounded area. For this application,
represented by enhanced storage and loss moduli. If stiff self-healing microfibres with core–shell geometry con-
and tough polymers are combined with dynamic and taining encapsulated binary epoxies are promising mate-
flexible macromolecular assemblies facilitating mobility, rials64. These fibres can be formed by coaxial nozzles
self-healing can be achieved7. Such materials have been that encapsulate epoxy resin and crosslinker in separate
made that respond to external stimuli, including temper- cores. Upon damage, epoxy and crosslinker are released
ature, electrical or magnetic fields55, solution concentra- from the cores of damaged fibres and self-heal. This
tion56 or light57. Notably, the action of the shape-memory approach is conceptually exciting if ‘used’ self-healing
effect during self-healing restores entropic energy upon agents could be replenished after damage.
the release of the force creating damage to fill an open Polymers in a molten state — owing to their low
wound in the material. viscosity and flow — may also serve as glue to repair
The shape-memory effect can be determined quan- mechanical damage. Because of interfacial diffusion
titatively in a single dynamic mechanical analysis that in the molten state, incorporating superparamagnetic
enables determination of stored conformational entropy γ-Fe2O3 nanoparticles into a polymer matrix facilitates
in polymers that exhibit a Tg and a rubbery plateau58. repair under the application of an oscillating magnetic
Although the majority of polymers exhibit some kind field9. As a result, the polymer matrix–nanoparticle
of shape-memory effect, they do not necessarily self- interfacial regions melt, facilitating polymer flow and
heal. An attempt was made to relate shape memory the permanent repair of physically separated polymer
to self-healing by copolymerizing ethylene and anisyl- surfaces. Embedding physically responsive nanoparticles
substituted propylenes using a sterically demanding or microparticles has several technological advantages,
half-sandwich scandium catalyst59. This study showed particularly in materials where damage repair does not
nanodomain phase separation and self-healability for permit the use of electromagnetic radiation, heat or other
copolymers with Tg below room temperature, in parti­ stimuli. It may also be advantageous in fibre-reinforced
cular, in aqueous environments (Fig. 2b); however, when composites, where damage may not be easily detectable
the Tg was tuned near or above room temperature, the and accessible, such as fibre–matrix interfaces.
shape-memory effect was observed. This raises several Long-range sensing and signalling are intrig­uing
questions, including the extent to which viscoelasticity properties of biological systems, but challenging to
above the Tg contributes to self-healing; how enhanced achieve in synthetic materials. One approach to address
chain mobility near Tg affects self-healing locally and this challenge is to fabricate a Diels–Alder bond-
globally; how far from the damage area polymer chains containing polyurethane backbone with electrically con-
‘feel’ conformational changes; and the role of interfacial ductive carbon nanotubes (CNTs), where the dynamics
energy in self-healing near the damaged regions60. In of Diels–Alder bond reformation enables self-healing
attempts to address these questions, two mechanisms and the CNTs confer electrical conductivity65. In this
of self-healing were identified: first, viscoelastic shape case, when electric potential is applied to a damaged
memory driven by stored conformational entropy upon specimen, the difference in electrical resistance between
damage, which is recovered to facilitate self-healing, the damaged and undamaged areas results in temper-
and, second, surface-energy-driven or surface-tension- ature differentials reflected in infrared spectroscopy
driven processes that reduce newly generated surface thermal images, thus providing crack diagnostics. This
areas created upon damage by shallowing and widening type of composite material can also be equipped with
wounds until healed61. strain-sensing capability and has potential technological
applications in electronic strain sensors as an alternative
Heterogeneous self-healing systems to ceramic piezoelectric devices. Based on these studies,
Although interest in homogeneous self-healing polymers we envision that electrically, thermally or optically con-
via interchain diffusion dates back to the early 1980s, ductive fibres embedded into a polymer matrix may lead
heterogeneous systems were not conceptualized and to a new generation of self-healing materials. An early
developed until the 2000s. The encapsulation of reac- example is CNTs dispersed in vitrimer-based epoxy,
tive fluids that are released upon damage and, hence, fill which can facilitate self-healing using photothermal
and repair the damaged area is one of the first examples energy to activate transesterification reactions66.
of self-healing heterogeneous systems19 (Fig. 1c). Initial Multiphase heterogeneous polymeric materials
studies involved catalysed reactions of ring-opening can be produced by copolymerizations. Similar to
metathesis polymerization of dicyclopentadiene in the phase-separated polyurethane containing a combi-
presence of a ruthenium catalyst19. This approach has nation of soft and hard segments, self-healing can be
been reviewed elsewhere62,63. The presence of nanocap- achieved in copolymers by combining hard polysty-
sules and microcapsules with reactive ingredients that, rene backbones with soft polyacrylate amide (PAAm)
upon cracking, fill in the damaged area resembles blood pendant groups carrying multiple H-bonding sites7.

NAture Reviews | MaterialS volume 5 | August 2020 | 565


Reviews

a Polyurethane-radical rebonding networks. It is worth noting, however, that commercial


styrene and acrylic copolymers are non-self-healable.
H
O To achieve self-healing, the chemical make-up of the
(CH2)6 (CH2)6
DBTDL– copolymer, control of the copolymer topologies and the
NH NH
O DBTDL– processing conditions are crucial. For example, it was
O C C O
HO O HO NH2
shown for PAAm that phase separation may induce
NH O
O
UV gel–glass-like transitions using a combination of good
O O O O
O O
O O
n
O
HO n and poor solvents68. In addition, phase-separated supra-
NH
HO
O
NH molecular architectures can also facilitate self-healing,
O C O
C O as demonstrated for acrylic or styrene-based triblock
copolymers69.
O O
H-bonding can also be used to achieve self-healing
in heterogeneous systems. For example, bio-derived
b Polycarbonate-radical rebonding carboxyl cellulose nanocrystals can be used to construct
HO
HO H-bonding interactions with chitosan-decorated epoxy
OH
OH natural rubber latex, where carboxyl cellulose nano-
O crystal molecular chains comprise rich carboxylic and
O
hydroxyl groups. Moreover, chitosan molecular chains
O O
O provide abundant amino, acetamide and hydroxyl
Room
O
temperature groups, interacting with each other via H-bonding70.
+ Using condensation polymerization, elastomers com-
O
O O posed of polytetraethylene and tetraethylene gly-
O
col have been prepared that are highly stretchable,
O
O tough and capable of self-healing71. The key interac-
tion responsible for these properties is quadrupole
HO
HO H-bonding.
OH
OH

DABBF ABF Covalent-bond reforming


Free-radical recoupling can reform covalent bonds in
c Generation of reactive groups polyurethanes and polycarbonates cleaved by mechani-
O
cal damage (Fig. 3a,b). The number of free radicals formed
O
C C6H12
O in these processes may vary depending upon their stabil-
O N C C6H12
O
C
N
C6H12 H H N O ity; for example, in polyurethanes modified by oxetanes
H H R 1: O CH2CH2 OH H H
HO
H O
R2: H2N HDI H O
or oxolanes, the C–O cleavage results in fairly stable
H O HO
HO
CO2 H
H R3: Urethane N H free radicals. In such cases, kinetic factors are important
H O O H
H O H H in self-healing: if free radicals or other reactive groups
O OCH3 O O
O H OCH3 C H OCH3
Sn O C O O O NH R C remain reactive for sufficient time after bond cleavage,
N C6H12 O NH
Bu Bu
H
Sn C6H12
C6H12
the chain ends terminated with free radicals will react
Bu Bu
before they are quenched by oxidative processes. Thus,
Polyurethane
stability and reactivity are important considerations.
Fig. 3 | Reformation of covalent bonds. a,b | Free-radical rebonding13,14. a | Upon Mechanical damage resulted in cleavage of a constrained
mechanical damage of an oxetane-substituted chitosan precursor incorporated in four-membered ring oxetane generating stable free rad-
a two-component polyurethane network, four-membered oxetane rings open to create icals33, and self-healing upon mechanical damage of
two reactive ends. Upon exposure to ultraviolet (UV) light, chain scission of the chitosan a crosslinked polyurethane network can be achieved
occurs, which crosslinks with the reactive oxetane ends, thus repairing the network. by exposing the damaged area to 302-nm ultraviolet
b | Diarylbibenzofuranone (DABBF), a dimer of arylbenzofuranone (ABF) and known (UV) radiation13 (Fig. 3a). Although mechanistically
antioxidant, was used as a dynamic reversible covalent bond capable of self-healing and kinetically different, self-healing was also realized
at room temperature in the dark. c | Generation of reactive amines10 capable of self-
with oxolane–chitosan macromonomers introduced to
healing. When methyl-α-d-glucopyranoside (MGP) molecules react with hexamethylene
diisocyanate (HDI) and polyethylene glycol (PEG), crosslinked MGP–polyurethane
poly­urethane networks72. In polycarbonates produced
networks are formed, which self-repair in air. DBTDL, dibutyltin dilaurate. by heat-induced reactions and diisocyanate-terminated
poly(propylene glycol)14, mechanical damage generates
the arylbenzofuranone radicals produced upon the dis-
This approach relies on localized variations in Tg intro- sociation of tetrahydroxy-functionalized diarylbibenzo-
duced by phase-separated high-T g polystyrene and furanone (DABBF). For these radicals to facilitate the
low-T g PAAm segments, which facilitates network rebonding of polycarbonate radical (Fig. 3b), the radicals
remodelling through hydrophobic and hydrophilic must have low or no sensitivity to oxygen to avoid radi-
interactions. Taking this concept further, triblock copoly­ cal quenching via free-radical reactions with molecular
mers can be prepared, benefiting from the elastomeric oxygen73, which impede self-healing. A similar approach
properties of microphase-separated thermoplastic was developed for the reformation of poly­carbonate
block copolymers with the reversible H-bonding67. This networks by adding a base (Na 2CO 3) to introduce
is an example of how localized compositional hetero­ additional ester reactions74. Subsequent substitutions
geneities can be used in designing self-healing polymer between phenoxide and phenyl–carbonyl chain ends

566 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

along with CO2 recombination facilitates the repair of generating reactive amine functionalities that react to
polycarbonate network. reform urethane and carbonate linkages. More specif-
Adduct (exo or endo) formation between furan ically, methyl-α-d-glucopyranoside (MGP) reacts with
(diene) and maleimide (dienophile) groups can also hexamethylene diisocyanate trimer and polyethylene
facilitate self-healing, owing to covalent rebonding75,76. In glycol (PEG) to form crosslinked MGP–polyurethane
these studies, mechanical activation of a DABBF-based networks10 (Fig. 3c).
scissile using dynamic covalent mechanophores was As described earlier, thermal energy and electro-
covalently attached to the surface of cellulosic nanocrys- magnetic radiation are commonly used to initiate
tals. The mechanically activated radicals recombined self-healing. It is often difficult to determine if electro­
much more slowly than those formed by DABBF mono- magnetic radiation alone causes self-healing or if
mers and DABBF-containing polymers in the other envi- energy absorption results in temperature changes,
ronments, such as solutions, bulk and gels77. Free-radical thus leading to the same net effect — network repair.
stability and self-healing has also been explored in This energy conversion influences the reversibility
self-healable alkoxyamines prepared by click chemistry, and efficiency of self-healing reactions in reversible
in which UV irradiation cleaved polymer crosslinks, fol- crosslinking via Diels–Alder reactions (for example,
lowed by recombination of radical species78. Similarly, in epoxies12,81–83, bismaleimides84,85, polyurethanes86,87,
multifold nitroxide-exchange reactions between tri- anthracene–maleimide-based polymers88,89, caprolac-
alkoxyamines and trinitroxide monomers form cova- tones90,91, poly(ethylene oxide)92, polyesters93, crosslinked
lently crosslinked dynamic self-healable network with a poly­lactic acid94 and acrylics95,96). Retro-Diels–Alder
tunable degree of crosslinking79. The uniqueness of this reactions offer a disconnection between diene and dieno-
approach is the reversibility driven by concentration phile, but elevated temperatures reconstruct the covalent
changes of 2,2,6,6-tetramethylpiperidinyloxyl (TEMPO) bonds to repair the crack12 (Fig. 4a). The photochemical
radicals. [2+2] cycloaddition of a 1,1,1-tris-(cinnamoyloxy-methyl)
The Calvin cycle — a biological process that con- ethane (TCE) monomer can be used for self-healing
verts light energy into sugars — is a sequence of reac- reactions to form cyclobutane structures via the reversi-
tions that includes the fixation of carbon dioxide and bility of cyclobutene to C=C bond conversion97 (Fig. 4b).
various enzymatic reactions80. This sequence of reactions In this example, the cycloaddition reaction98 capitalizes
is driven by forming and reforming reactive groups. on the C–C bond cleavage of cyclobutane rings between
Achieving the fixation of carbon dioxide using synthetic TCE monomers induced by mechanical damage, result-
materials has also been realized and, similarly, enables ing in the formation of the original cinnamoyl groups;
the reformation of bonds and the renewal of materi- however, healing occurs because of the reversibility of
als10. These materials are capable of self-repair in the cyclobutane crosslinks of TCE via [2+2] photocycload-
presence of atmospheric carbon dioxide and water by dition upon UV exposure. A low-temperature reversible

a Reversible Diels–Alder reactions b [2+2] Cycloaddition reactions


O O

O O
O Damage O

Damage
hv (>280 nm)
O O
O O Heat
N O O
O O
N

c Exchange reactions d Michael addition


X X X X O
O
X X NC
X X N NC
H Damage N
Disulfide exchange: X = S H
S +
Diselenide exchange: X = Se Heat SH
Siloxane exchange: X – X=Si – O – Si

Fig. 4 | Reversible reactions enabling reformation of covalent bonds. a | Thermally reversible Diels–Alder reaction of furan
and maleimide moieties forms a highly crosslinked network that is thermally reversible by the retro-Diels–Alder reaction12.
b | Cycloaddition reactions can be used to photocrosslink cinnamate monomer, such as 1,1,1-tris-(cinnamoyloxy-methyl)ethane,
and uncrosslink to original cinnamoyl groups, thus facilitating crack healing97. c | Three different exchange reactions can
lead to self-healing. Disulfide-exchange reactions enable self-healing of n-butyl acrylate-based star polymers by attaching
bis(2-methacryloyloxyethyl) disulfide to the arm end, resulting in a crosslinked star polymer106. The presence of disulfide
functional groups enabled cleavage, owing to reduction reactions at the chain ends with thiol groups. Diselenide-exchange
reactions113 (dynamic diselenide bonds incorporated into polyurethanes facilitate visible-light-assisted self-healing materials).
Siloxane-exchange reactions116 can serve in self-healing of silicone-based polymers. Under stress, siloxane bonds in the
silicone networks are cleaved, but the addition of either acid or base will catalyse stress relaxation, resulting in self-healing,
d | Michael-addition reaction115 between a trithiol and a bisbenzylcyanoacetamide derivative forms a self-healable dynamic
network above 60 °C.

NAture Reviews | MaterialS volume 5 | August 2020 | 567


Reviews

system using acid-activated dithioesters as dienophiles disulfide-based materials and reprocessability at


in hetero-Diels–Alder reactions with cyclopentadiene99 temperatures as low as 100 °C (ref.114). Reversible thiol–
has slow self-healing kinetics in the absence of a catalyst, ene click reactions have also been used in a Michael-
but when hetero-Diels–Alder contains cyanodithioester, addition reaction, in which trithiol was reacted with
a catalyst is not required100. Using cyanodithioester and a bisbenzylcyanoacetamide derivative to generate a
cyclopentadiene building blocks, healing can be accel- self-healable dynamic polymer network above 60 °C
erated at 120 °C (ref.101)). It is also well established that (ref.115) (Fig. 4d).
Diels–Alder reactions can be catalysed by antibod- Silicone-based polymers — owing to their dynamic
ies and RNA102, which offers the possibility of using network rearrangements — are perhaps the most tech-
these reactions in chemical biology. The dynamics of nologically promising, yet least explored, self-healing
Diels–Alder bonds for self-healing in polyurethane and materials 116. For example, tetramethylammonium-
CNT composites was also used in electrically conductive silanolate-initiated ring-opening copolymerization of
networks65. octamethylcyclotetrasiloxane and bis(heptamethylcyclo-
Sulfur and selenium chemistry, which is vital in bio- tetrasiloxanyl)ethane produced a polymer that could
logical systems, also enables self-healing. Particularly, self-heal, owing to its ethylene bridges and active
the robust nature of thiol–ene chemistry has led to silanolate end groups 117. Perhaps one of the most
many synthetic opportunities103,104. For example, fea- attractive features of silicone-based materials with
tures of reversible reshuffling reactions of S–H and S–S self-healing properties are energy118,119 and biomedi-
bonds have been used in trithiocarbonates, in which cal120 applications. Of particular interest are self-healing
copolymerization of n-butyl acrylate and a trithiocar- dielectric-silicone-based elastomers that exhibit
bonate crosslinker resulted in high mobility of poly- high dielectric permittivity. These properties can be
mer segments, triggering homolysis of C–S bonds105. achieved using an interpenetrating polymer network of
The reversible nature of S–S bonds, that is, reduction to silicone elastomer and ionic silicone polymers121. The
form two thiol (S–H) groups, and oxidation to restore latter are crosslinked through proton exchange between
the disulfide (S–S) linkages, can be effectively used in amines and acids, and are able to self-heal after electrical
self-healing106 (Fig. 4c). Introducing reversible disulfide breakdown or mechanical damage. The self-healing is
(S–S) crosslinks into covalently crosslinked networks attributed to the reassembly of the ionic bonds during
can be achieved by poly(n-butyl acrylate)-grafted star damage. Self-healing in silicone-based polymers was
polymers107. In this example, the polymer networks also accomplished using thiol-functionalized silicone
originate from crosslinked cores composed of poly(eth- oils containing silver122 and magnetic Fe3O4 nanopar-
ylene glycol diacrylate) and macroinitiators for the ticles containing mussel-inspired metal-coordination
consecutive chain extension of bis(2-methacryloyl) bonds with dopamine molecules123,124.
oxyethyl disulfide. In another example, photoinduced Taking self-healing further, latexes synthesized via a
thiol–ene click-type radical addition, which generates simple colloidal process undergo not only self-healing
lightly sulfide-crosslinked polysulfide-based networks but also colour changes in the damaged area125. The syn-
with an excess of thiols, and the subsequent oxidation of thesis of these materials was accomplished by emulsion
these thiols enables the formation of dynamic disulfide copolymerization of a small fraction of photochromic
crosslinks to yield dual sulfide–disulfide crosslinked net- monomer (spirooxazine) into methyl methacrylate and
works with fast self-healing rates108. Triblock copolymers n-butyl acrylate copolymers. Upon mechanical dam-
with a centre poly(ethylene oxide) block and dithiolane age, colourless damaged areas become red and, upon
blocks crosslinked with dithiol also enable self-healing exposure to electromagnetic radiation, not only does
by reversible ring opening of the pendent 1,2-dithiolanes self-healing occur but also the red colour in the damaged
via disulfide exchange between 1,2-dithiolanes and area reverts to its original colourless appearance. During
dithiols109. Bis(4-aminophenyl) disulfide can also be repair, which is initiated by visible light, the spirooxazine
effectively used as a dynamic crosslinker in self-healing rings close; however, the neighbouring copolymer seg-
poly(urea–urethane) elastomers without the use of a ments form intermolecular H-bonding that forces the
catalyst110. Disulfide chemistry was also used in self- copolymer backbone to remain in an extended confor-
healable polyurethanes111. An advantage of disulfide- mation. This network is an example of how combining
exchange reactions is that S–S bonds are capable of covalent bonding and supramolecular chemistry may
dynamic rearrangements upon heat, UV light and redox lead to self-healing and sensing.
conditions, and, when incorporated into low-Tg gel
networks, temperature-reversible self-healing can be Dynamic covalent-bond reformation
achieved. The concept of disulfide links incorporated in The interest in covalently bonded self-healable hydro-
a rubber network resulted in the restoration of mecha­ gels is driven by its biomedical applications, which range
nical properties at moderate temperatures (~60 °C)112, from regenerative medicine to drug and protein deliv-
but the main challenge is to achieve self-healing in ery systems or tissue–material barriers. Self-healing in
higher-T g networks. Applying the same exchange hydrogels can be achieved using reversible Schiff-base
concept, a series of diselenide bond-containing poly­ linkages126. For example, such hydrogels can be formed
urethane elastomers was prepared113. Aromatic diselenides from the reaction of OH-PEG with 4-formylbenzoic
have also been incorporated into polyurethane networks acid and, subsequently, glycol chitosan127 (Fig. 5a). In
using a para-substituted amine diphenyl diselenide, this case, a large dynamic strain (300%) can change the
resulting in faster self-healing than the corresponding elastic modulus from approximately 1.5 kPa to 10 Pa and

568 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a Schiff-base bond regeneration b Acylhydrazone rebonding

O N H2N NH N NH
C + H2N C CHO + CH

R O O
R

c Oxime rebonding d Boronic-ester reformation


–H O + 2H 2O
2O
O O OH HO
R1 H2N R2 R1 N R2 B
O + + B
+H O
2 O – 2H2O
OH HO

e Boronic-ester reformation in rotaxane-based polymers

H H H H
N O N N O N
O O O O
n m n m
O O O O
O O HO OH
B
OH HO OH
B
O NH2
H2 H2
C C r C C O NH2
H x H 100-x n H2 H2
C C r C C
H x H 100-x n

Fig. 5 | The dynamic reformation of covalent bonds. a | Schiff-base rebonding reaction127. b | Acylhydrazone
rebonding108,130. c | Oxime rebonding131. d | Radical thiol–ene click chemistry is initiated by light, which enables
the reversible formation of boronic esters to yield self-healable polymers135. e | Self-healing polymer gels are formed
by crosslinked polyrotaxane and poly(acrylamide) with reversible boronate linkages142.

induce gel–sol transitions. When implanted into the cen- However, the property that makes boronic acids
tral nervous system, neurosphere-like progenitors prolif- unique in the context of self-healing is their ability to
erate in the hydrogel and differentiate into neuron-like form reversible dynamic covalent bonds with diols
cells, causing neural repair. Schiff-base networks can to form cyclic boronate esters. For example, photoini-
also be integrated with dynamic reversible acylhydra- tiated radical thiol–ene click chemistry was used to
zone via acylhydrazines and aldehyde functionalities form hydrogels capable of repeated healing under
into Diels–Alder crosslinked networks128. A hybrid ambient conditions135 (Fig. 5d). In another approach,
self-healing hydrogel system combining Schiff-base and combining reversible and dynamic boronate ester and
amine reactions with micelles exhibits rapid self-healing, disulfide chemistries facilitated the development of
extensibility and compressibility for wound-dressing pH-responsive, glucose-responsive, redox-responsive
applications129. and self-healable properties in hydrogels136. Self-healable
Hydrozenes (C=N-X)108,130 (Fig.  5b) and oximes131 hydrogels can also be prepared by covalent transester-
(Fig. 5c) are commonly used conjugates labile to hydro­ ification reactions of boronic acid with diols137. These
lysis. Keto-functional copolymers can be prepared boronate-ester-crosslinked hydrogels are capable of
by conventional radical polymerization of N,N- self-healing under neutral and acidic conditions, owing
dimethylacrylamide (DMAA) and diacetone acryl­ to the presence of an intramolecular coordinating
amide (DAA)131. The resulting water-soluble copolymers boronic-acid monomer, 2-acrylamidophenylboronic
(p(DMAA-stat-DAA)) can be chemically crosslinked acid 138 . Synthetic hydrogels functionalized with
with difunctional alkoxyamines to obtain hydrogels boronic acid can also be used as matrices for 3D cell
via oxime formation and sol–gel transitions induced culture139, and those generated by boronic-ester and
by the addition of excess monofunctional alkoxy­ disulfide-exchange chemistry can self-heal in response
amines to promote competitive oxime exchange under to several stimuli136.
acidic conditions at 25 °C. The dynamic nature of In addition to boronic acid and boronate ester,
oxime functionalities facilitates reversible self-healing. self-healing gels with dynamic covalent bonds can be
Similarly, polyurethane-like dynamic covalent polymers formed from a range of other reactive groups. For exam-
(poly(oxime-urethanes)) with self-healing at 120 °C can ple, alkoxyamines with dynamic covalent bonds can
be prepared132. facilitate self-healing, although the sensitivity of atmos-
Boronic acids form a variety of dynamic covalent pheric oxygen at elevated temperatures (90–130 °C)
bonds133. This can be achieved, for example, through leads to C–ON bond dissociations140. Another approach
dehydration of boronic acids to form boroxines by is the photoinduced [2+2] cycloaddition of cinnamoyl
reversible hydrolysis134. The boroxine–boronic acid equi- groups to reversibly form cyclobutene-derivative poly-
librium can be shifted by adjusting the temperature, add- mer gels97. This mechanism is driven by trithiocarbonate
ing a Lewis base or changing the water concentration. units that undergo reshuffling reactions; these units are

NAture Reviews | MaterialS volume 5 | August 2020 | 569


Reviews

particularly attractive as photoinitiators in reversible forming triple hydrogen bonds (Fig. 6a). When supra-
addition–fragmentation chain transfer (RAFT) poly­ molecular polyisobutylene networks are equipped with
merization105. This concept was expanded to self-healing directional associative end groups, tunable dynamic
networks using trithiocarbonate as photoresponsive behaviour, including self-healing, can be achieved151.
units to achieve remarkable self-healing properties141. Combining four hydrogen bonds offers high associa-
The latter is an excellent example of how covalently tion constants152,153. An example of this is a functional
crosslinked polymers, through light stimulation and unit of urea isopyrimidone with enhanced association
macroscopic fusion of separate pieces, synchronize strength between units when incorporated in poly-
self-healing events. Moreover, polyrotaxanes crosslinked siloxanes, polyethers and polyesters154–156. In contrast
by reversible bond reformation between ring molecules to bis-urea H-bonding, which leads to crystallization
and vinyl polymers using boronic linkages are another or clustering, resulting in brittle materials, applying
example of rapid self-recovery upon damage142 (Fig. 5e). thiourea moieties leads to zigzag H-bonded arrays,
Hydrogels with double networks were also developed which eliminate unfavourable crystallization157. The
to achieve toughness comparable to that of rubber143, dense H-bonds between thiourea appear to be ideal to
but many applications require the material to withstand crosslink low-molecular-weight polymers to achieve
repeated loading–unloading cycles in a short time, along tough, room-temperature self-healable materials. In
with self-healing upon injury. To tackle this challenge, another example, dual-amide H-bonds doped with
polyurethane-based hydrogels with enhanced dipole– poly(3,4-ethylenedioxythiophene)–poly(styrene sul-
dipole and H-bonding interactions were prepared144. fonate) (PEDOT–PSS) offer another venue for fabri-
These materials have superior mechanical properties cation of self-healing hydrogels158, whereas integrating
to conventional double-network hydrogels and are also DNA-grafted polypeptides and DNA linkers provides
capable of self-healing. reversible DNA recognition for self-healing159. Also,
sacrificial H-bonding was introduced by incorporation
Supramolecular dynamic chemistry of secondary amide side chains into olefin-containing
Supramolecular chemistry is non-covalent bonding networks to achieve self-healing160.
represented by hydrogen bonding, metal–ligand coor- As a consequence of dynamic association–dissociation,
dination, π–π, ionic, guest–host and van der Waals inter- H-bonding plays a role in tuning mechanical proper-
actions145,146. Although these interactions are relatively ties. Supramolecular polymers based on bifunctional
weak compared with covalent bonding, collectively, they ureidopyrimidinone derivatives behave like mechani-
form mechanically strong and very dynamic systems. cally stable, high-molecular-weight thermoplastic poly­
Many biological assemblies are based on the principles mers; however, their mechanical properties, such as
of supramolecular chemistry. Accordingly, supramo- Young’s modulus and tensile strength, exhibit a strong
lecular processes are typically bottom-up and involve temperature dependence, owing to H-bonding dissoci-
non-equilibrium states. ations161. Along the same lines, polysiloxane elastomers
The field of supramolecular chemistry has been containing a mixture of strong and weak H-bonding
around for several decades147,148 and is attractive for (Fig. 6b) offer tunable mechanical properties, including
development of self-healing materials because of its stretchability, toughness and autonomous self-healing
reversibility, directionality and sensitivity. In contrast ability, even under water162. Owing to the presence of
to covalent bonding, networks held together with H-bonding of different strengths, these self-healable
non-covalent bonds can be remodelled reversibly — elastomers can distinguish external signals of different
from fluid-like, low-density and high-free-volume strength, thus opening the possibilities for applications
states to solid-like, low-free-volume, elastic and plastic in human–machine interactions.
networks. Supramolecular polymers usually exhibit low Fatty diacids and triacids from renewable resources
Tg, which results in soft polymers, making them popular have been used in a two-step synthetic route (Fig. 6c)
in the development of hydrogels. Hydrogels based on to form self-healing networks by bringing together
supramolecular chemistry have been used in applica- two cut ends at room temperature, without external
tions such as injectable bioimplants, printable biological heat15. Condensation polymerization of acid groups
compounds and artificial skin. with an access diethylenetriamine and subsequent
In this section, we cover the different supramolecular reactions with urea involving amidoethyl imidazoli-
interactions and chemistries used to produce self-healing done, diamidoethyl urea and diamido tetraethyltriurea
materials, namely, those based on H-bonding, metal–ligand, groups resulted in an oligomer mixture with excessive
ionic, host–guest and π–π interactions. H-bonding. In another example, which took advantage
of variable-strength H-bonding and segmental mobil-
Designs based on H-bonding. H-bonding is typically ity of hydrogels, self-healing could be switched on or off
among the strongest of non-covalent interactions and by adjusting temperature or pH163.
its directionality and high per-volume concentra- Using the crosslinker 2-ureido-4-pyrimidone-
tion confers acceptable mechanical strength. Owing 4-hydroxybutyl acrylate (UPyHCBA), which consists
to these attributes and reversibility, it can be used in of an acrylic head, a hydrophobic alkyl spacer con-
self-healing of thermoplastic polymers149,150. For exam- nected by carbamate and a 2-ureido-4-pyrimidone
ple, high-segmental-mobility polyisobutylenes func- tail, both hydrophobic association and H-bonding can
tionalized with thymine and 2,6-diaminotriaine end be achieved. Sodium dodecyl sulfate micelles can pro-
groups assemble into strong rubber-like materials by vide a hydrophobic environment for the UPyHCBA

570 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a Triple H-bonding
HN H O NH2 O
N N
PIB N H N PIB N + H N
N N N N
HN H O PIB NH2 O PIB

b Combining strong and weak H-bonds


O O O
H H
Si O N N Si O
N N N N O N N O
H H H H n H H n
x O 1-x

Weak H-bonding (anti-cooperative)


Strong H-bonding Weak H-bonding H H O
N N N
N
O H H

O O
N N N N
H H H H

H H O
N N
N N
O H H
H H O
N N
N N
O H H

PDMS linker Strong H-bonding (cooperative)


O O
N N N N
H H H H
O O
N N N N
H H H H

c Excessive H-bonding O O
O
N
N N H N N
N
H H H 2N O H
O
O O O
(CH2)7 C6H13 N N
Amidoethyl imidazolidone N N N N
H H H H
O NH2 O NH
C8H17 (CH2)7 O O
H H
N
N N
H H HN O
O NH2 H H
N N
C6H13 (CH2)7 N
Di(amidoethyl) urea H-bonding O O
(CH2)7 C6H13 O H
O O O H
N
N N N N
C8H17 (CH2)7 N N N N O
H H H H
O NH2 O NH2
HOOC
H
COOH N Diamido tetraethyltriurea
2 NH2
+
COOH
HOOC Step 1 Step 2
COOH
Oligomeric mixture

Fig. 6 | Examples of hydrogen bonding in self-healing polymers. a | Triple H-bonds between thymine and
2,6-diaminotriaine form self-healable supramolecular networks. b | The combination of strong and weak hydrogen
bonding in polydimethylsiloxane (PDMS)–4,4′-methylenebis(phenylurea) form self-healable networks. c | A mixture of
fatty diacid and triacid reacts with diethylenetriamine to form self-healable copolymers. Subsequent reactions with urea
form oligomeric networks with strong hydrogen bonds. PIB, polyisobutylene. Panel a adapted with permission from ref.151,
RSC. Panel b adapted with permission from ref.162, Wiley-VCH. Panel c adapted from ref.15, Springer Nature Limited.

prepared by micellar polymerization of UPyHCBA and Metal–ligand interactions. When metal ions and appro-
acrylamide164. Another useful self-healing mechanism priate ligands — either at the end of a polymer chain or
is the formation of urea–water clusters in urea-based as a pendent group — are brought together, they form
polyurethanes in the presence of moisture165. a coordination complex, linking the polymer chains

NAture Reviews | MaterialS volume 5 | August 2020 | 571


Reviews

together. Metal–ligand complexes have advantages asso- endows fast self-healing under ambient conditions. In
ciated with their ability to coordinate different metal ions the second example, the coordination between Fe3+ and
and ligand substitutes, thus leading to different associa- catechol ligands resulted in a pH-induced crosslinked
tion strengths. When mechanical forces are applied, such self-healing polymer with near-covalent elastic mod-
complexes dissociate, and their reformation results in uli172 (Fig. 7c). The uniqueness of this approach is the
self-healing. Another appealing feature is minimal ability of mono-catechol–Fe3+, bis-catechol–Fe3+ or
or no side reactions during repeated damage–repair tris-catechol–Fe3+ formation at different pH values, thus
cycles. Reversibility upon mechanical damage is facili- providing control of crosslinking without Fe3+ precip-
tated by metal-ion–ligand dissociations, with reforming itation173. Incorporating Eu–iminodiacetate coordina-
triggered by exposure to electromagnetic radiation or tion with hydrophilic poly(N,N-dimethylacrylamide)174
elevated temperatures. and dynamic ionic interactions between carboxylic acid
Exemplifying a temperature-responsive system, when groups of poly(acrylic acid) (PAA) and ferric ions175 or
hydroxyl ethylene diamine triacetic acid is used to con- polyelectrolytes176 are other examples of self-healable
trol the molecular weight and crosslinker density in hydrogels177,178.
the presence of terpyridine–Ru, temperature changes Nature offers opportunities for designing self-healing
decouple Ru metal from the ligand, resulting in bonding systems. Inspired by catechol-containing biopolymers
and debonding166. In a metallosupramolecular system, that mimic sea-mussel adhesives, crosslinked hydro-
the terpyridine–Fe2+ complexation pair with poly(alkyl gels were prepared by the complexation of branched
methacrylate) accomplishes self-healing at elevated catechol-derivatized PEG with 1,3-benzenediboric
temperature167, although, in this case, the elevated tem- acid. This material exhibits tunable, covalently bonded
peratures may lead to adverse effects, such as polymer gel behaviour under alkaline pH, but dissociates into a
degradation. In a chemically different, photoresponsive viscous liquid under neutral and acidic conditions168,179.
metallosupramolecular system, polymers comprising It has been used in surface modifications of synthetic
an amorphous poly(ethylene-co-butylene) core with polyacrylate and polymethacrylate materials with
2,6-bis(1-methylbenzimidazolyl)pyridine ligands can mussel-inspired catechols, which resulted in self-healing
form metal–ion binding, which, upon mechanical dam- initiated and accelerated by H-bonding between
age, can be mended through exposure to light16 (Fig. 7a). interfacial catechol moieties180.
Because exposure to UV light causes the metal–ligand Finally, tuning the strength of metal–ligand interac-
coordination to be electronically excited via the absorp- tions has implications for the dynamic mechanical prop-
tion process, the desorption of energy may lead to heat erties of supramolecular elastomers. For example, when
dissipation and temporary cleavage of the metal–ligand 2,6-pyridinedicarboxamide ligands are incorporated
linkages. Other examples of the formation of self-healing along PDMS backbones, complexation of the polymer
complexes are N-heterocyclic carbenes and transition with Fe(III) (Fig. 7d) results in an elastic material that
metals168. However, a drawback of these examples is the self-heals at room temperature. This behaviour is attrib-
substantial external energy input required. uted to the bonding energies of Fe(III)–Npyridyl, Fe(III)–
It is important to match the electronic network prop- Namido and Fe(III)–Oamido, which may range from strong
erties and excitation sources when designing self-healing to weak. The weaker bonds are responsible for energy
polymer networks using metal–ligand coordination dissipation on stretching and on-demand self-healing,
chemistry. An example of this is polyethylenimine– whereas the metal ions maintain their location near the
copper supramolecular polymer networks with revers- ligands, thus resulting in stronger interactions and rapid
ible UV-induced self-healing by the reformation of bond reformation181.
Cu–N coordination bonds169. This system experiences
virtually no temperature change during exposure to Host–guest interactions. Guest–host chemistry is com-
electromagnetic radiation, providing high photon effi- monly used in fabricating hydrogels. For example,
ciency without side reactions. The network undergoes hydrophobic cavities in β-cyclodextrin can accommo-
square-planar-to-tetrahedral (D 4h → Td) symmetry date a variety of guest moieties182–185. If one surface con-
changes at the C2H5N–Cu coordination complex cen- tains a cyclodextrin host and the other guest molecules,
tre without side reactions facilitated by charge trans- host–guest interactions will result in bonding17. The
fer between σ(N) bonding and dx2-y2(Cu) antibonding supramolecular polymers shown in Fig. 8a are equipped
orbitals. Combining the properties of polyurethane and with multipoint molecular recognition sites achieved
polydimethylsiloxane (PDMS) networks into one supra- by various water-soluble polymer backbones modified
molecular and covalently crosslinked system catalysed with β-cyclodextrin hosts and hydrophobic adaman-
by CuCl2, this system reforms both coordination and tine as the guest at the side chain. This simple method
covalent bonds170. produces a transparent, flexible and tough hydrogel
Other pathways to self-healing in metal–ligand that self-heals in both wet and dry states186. Because of
coordination systems are shown in Fig. 7b,c. For exam- their amenability to chemical modifications and bio-
ple, a self-healing dielectric elastomer has been synthe- compatibility, supramolecular hydrogels composed of
sized using 2ʹ-bipyridine-5,5ʹ-dicarboxylic amide as modified hyaluronic acid with either adamantane or
ligand and Fe2+ and Zn2+ with various counter anions β-cyclodextrin are able to rapidly form intermolecular
in which metal–ligand coordination serve as crosslink- host–guest bonds187,188. The mechanical properties of this
ing sites in non-polar PDMS171 (Fig. 7b). The kineti- system can be tuned by changing the concentration and
cally labile coordination between Zn2+ and bipyridine ratio of the guest-to-host moieties.

572 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a Photoresponsive system

N N N N
N N N N
UV
O N Zn2+ N O O N N O

Repair N
N N N
N N N N

2NTf2– Zn2++ 2NTf2–

b Self-healing dielectric elastomer

H 2N O O NH2
Si m Metal salts
crosslinker : N M2+ = Zn2+, Fe2+
Cl N Cl
FeCl2, Fe(BF4)2 N N
Zn(CF3SO3)2, M2+ = Si O
O N O
ZnCl2, Zn(ClO4)2 O n
N N
H H
N N N O N
Si O =
n
m N N N
O N O N

N N
N ClO4–
N N N
CF3SO3– N Zn CF3SO3– N N
N Cl – N N Zn
Cl Zn N N CF3SO3
Zn N N
Cl CF3SO3
N N N
N N
N N N

c pH-induced self-healing d Tuning metal–ligand interactions


+ +
H H
H N N
O N N
+Force N Force
HO OH N O O O
Fe
Fe
O O Fe3+ O N
–Force O O
O O OH HO N
Fe3+ N O N
H N N
O O OH HO H H

Strong Medium Weak


coordination coordination coordination

Fig. 7 | Self-healing using metal–ligand coordination chemistry. a | Coordination of Zn(NTf)2 and 2,6-bis
(1-methylbenzimidazolyl)pyridine ligands responsible for self-healing in poly(ethylene-co-butylene)16. b | Synthetic route
to prepare Zn2+- or Fe2+-containing crosslinked polydimethylsiloxane networks and a schematic illustration of the dynamic
interactions within the Zn2+–ligand complex, counter anions polymer system under mechanical stress171. c | Reversible
tris-catechol–Fe3+ complex172. d | Reversible coordination complexes with 2,6-pyridinedicarboxamide ligands and Fe(III)
centre as crosslinker for self-healable polydimethylsiloxane network181.

Another well-known host is cucurbit[8]uril (Fig. 8b), conductive and transparent networks can be stretched
which has high molecular weight and sufficient chain to 100 times their original length and hold objects
entanglement for physical crosslinking, and can accept 2,000 times their own weight192. Although this example
two guests, naphthyl and viologen189. Some seminal and those above are remotely related to self-healing,
studies using self-assembled peptides have demonstrated if applied in the host–guest environments, potential
the potential of these interactions in biomedical appli- applications in self-healing of materials are feasible.
cations190,191. At low concentration (2.5 mol%), dynamic
cucurbit[8]uril-mediated non-covalent crosslinking Ionic interactions. Ions have a key role in many bio-
yields extremely stretchable and tough supramolecular logical processes. For example, molecular motors are
polymer networks, exhibiting remarkable self-healing powered by chemical processes that result in the swell-
capability at room temperature. These ionically ing and shrinkage of macromolecular segments, thus

NAture Reviews | MaterialS volume 5 | August 2020 | 573


Reviews

causing motion. These processes are often attributed As a consequence of electrostatic interactions, mac-
to the imbalance between osmotic and entropic forces romolecules carrying opposite charges form neutrally
resulting from stored energy between electrostatic charged polyelectrolyte complexes. Stiff and self-healing
repulsions and negatively charged filaments. Ionic gels can be fabricated with the polyelectrolyte compl-
interactions also influence polymer networks and are exation of polyamines with phosphate-bearing multi­
primarily manifested by the formation of ionomers193; valent anions196 (Fig. 9b). Self-healing is also exhibited in
however, only some ionomers self-heal. For example, poly­electrolyte complexes synthesized from (PAA)/poly
poly(ethylene-co-methacrylic acid) (pEMAA) and (allylamine hydrochloride) (PAH) pairs and NaCl, with
polyethylene-g-poly(hexylmethacrylate)194 (pEHMA) the self-healing efficiency increasing with NaCl con-
self-repair (Fig. 9a) under ambient conditions and at ele- centration197 (Fig. 9c). In this case, self-healing originates
vated temperatures upon projectile puncture testing195. from the disturbance of ionic interactions by adding salt,
Moreover, a ballistic puncture in low-density polyethy­ resulting in enhanced chain mobility.
lene does not heal, whereas a puncture in pEMAA Although ionic bonds may be more difficult to break
does194. This is a two-stage process whereby projectile than covalent bonds, they can be disrupted by introduc-
impact disrupts the ionomeric network and the heat ing strong electrostatic interactions (for example, disso-
generated by friction during damage is transferred to ciations upon addition of water or adding strong polar
polymer matrix surroundings, resulting in a localized salts or solvent). Therefore, the addition of water, salt
melt state. The molten polymer surfaces fuse via interdif- or polar solvents will influence self-healing. In the pres-
fusion to seal the puncture, followed by rearrangement ence of water, polyelectrolyte multilayered assemblies
of the ionic clustered regions and long-term network also exhibit self-healing upon mechanical damage. An
relaxation. example of this is polyampholyte self-healing hydrogels

a OH
CH2CH r CH2CH r CH2CH
O O OH C O C O R
NH x NH y 100-x-y
O
O
OH
HO
HO

HO O HO n
OH O

OH
HO O O O
OH
O β-Cyclodextrin R=
O NH2 N N
HO H
OH
pAAm pDMAAm pNIPAAm
O OH O O
O
HO OH
HO N OH O
HO H
HO
OH pHMAAm pHEA
O
O
OH
O O

HO

+ O
N N
≡ – ≡ N N CH2
Br
H H
Cucurbit[8]uril N N CH2

Fig. 8 | Host–guest chemistry in self-healing systems. a | Water-soluble polymer backbones modified with β-cyclodextrin
as host and hydrophobic adamantine as guest at the side chain. Polymers selected to form the water-soluble backbone
of the system include poly(acrylamide) (pAAm), poly(N,N-dimethylacrylamide) (pDMAAm), poly(N-isopropylacrylamide)
(pNIPAAm), poly(hydroxymethylacrylamide) (pHMAAm) and poly(hydroxyethylacrylate) (pHEA)186. b | Stepwise formation
of ternary host–guest supramolecular complexation between cucurbit[8]uril and guest molecule with macrocyclic host
cucurbit[8]uril and polymerizable guest molecules (1-benzyl-3-vinylimidazolium) and acrylamide192. Panel b adapted with
permission from ref.192, Wiley-VCH.

574 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a Self-healing b Polyelectrolyte complexes c Polyelectrolyte form reversible crosslinked networks when triggered by
ionomer with phosphate anions complexes with NaCl Ca ions199. Mechanically and electrically self-healing
pEMAA PAH/tripolyphosphate PAA PAH hydrogels can be obtained using dynamic ionic inter-
O OH O
actions between carboxylic groups of PAA and ferric
O O
P P +
+
ions178. Moreover, stretchable, transparent, self-healable
+ P Na
+
H 3N

O – O
O–
O
O
– O– –
O O H 3N
+
ionic conductors can be formed based on ion–dipole
CH3 O –
interactions200 (Fig. 9e). In these systems, the polar poly­
x y 5
Cl – +
Cl
Na
mer network is usually composed of poly-(vinylidene
d Multilayered assemblies
fluoride-co-hexafluoropropylene) (PVDF-co-HFP).
Cationic monomer Anionic monomer Cationic monomer This copolymer, which contains two building units,
DMAEA-Q NaSS MPTC VDF (crystalline and less polar) and HFP (amorphous
Cl–
O
O and highly polar), is capable of room-temperature
Cl–
+
N
O
+Na–O S
+ self-healing without external stimulus.
N N
O O
H Moving towards commodity self-healable polymers,
ionic interactions may be important in the optimization
and modification of commercial rubber. Commercially
P(NaSS-co-DMAEA-Q) P(NaSS-co-MPTC) popular and inexpensive butyl rubber can be modified
into a self-healing product by ionic modifications with-
e Ion–dipole interactions
out the addition of conventional curatives or vulcanizing
F F agents. For example, the mechanical properties of bro-
+
mobutyl rubber (which is conventionally sulfur-cured)
N N
F CF3 0.45
Me Et can be improved by transforming its bromine function-
F F 0.55
alities into ionic imidazolium bromide groups (Fig. 9f),
PVDF-co-HFP-5545 1-Ethyl-3-methylimidazolium resulting in reversible ionic associations that exhibit
physical crosslinking201. In summary, ionic interactions
Et
N
F F in polymers may provide a unique opportunity for the
F F
+
F F development self-healable commodity materials.
N CF3 F
CF3 F F
Me F
F F
CF3 F F
F F π–π interactions. π–π interactions are often viewed as
F F F F
F F F F
F
an extension of coordination chemistry. As the name
CF3 F
F F Et implies, these interactions are facilitated by π orbitals,
F N CF3
Me N +
F
CF3 which are strongly dependent on chemical structure and
stereochemistry. π–π interactions are well-documented
f Self-healing in a commodity polymer in many biological studies of peptides, interactions of
aromatic side groups in proteins, and nucleic acids or
Br
Heat
N N DNA. Using a combination of π–π, metal-coordination
+
chemistry and/or H-bonding, several self-healing elas-
Br
Br –
N tomers have been developed. For example, combining
N Pt···Pt and π–π interactions between a cyclometalated
platinum(II) complex and a PDMS backbone enables
high stretchability and self-healing202. Composites with
Fig. 9 | Examples of ionic interactions applied in self-healing. a | An early example metal nanoparticles can also be commonly applied. For
of a self-healing ionomer, namely, poly(ethylene-co-methacrylic acid) (pEMAA)194. example, the blend of pyrene-functionalized poly­amide
b | Polyelectrolyte complexes with phosphate anions, for example, poly(allylamine (π-electron donor), polydiimide (π-electron accep-
hydrochloride) (PAH)/tripolyphosphate ionic gel196. c | Polyelectrolyte complexes tor) and pyrene-functionalized gold nanoparticles can
can be formed by ultracentrifugation with NaCl, for example, poly(acrylic acid) produce thermally induced π–π stacking interactions
(PAA)/PAH197. d | Chemical structures of anionic monomer sodium 4-styrenesulfonate
between functionalized gold nanoparticles and the poly­
(NaSS), cationic monomers acryloyloxyethyltrimethyl ammonium chloride (DMAEA-Q)
and [3-(methacryloylamino)propyl]trimethylammonium chloride (MPTC) used for mer matrix, resulting in self-healing203. Combining π–π
synthesizing polyampholyte hydrogels, p(NaSS-co-DMAEA-Q) and p(NaSS-co-MPTC)198. and H-bonding interactions can lead to thermally trig-
e | Ionic conductor fabrication using ion–dipole interactions between poly-(vinylidene gered, self-healable, tweezer-shaped structures consist-
fluoride-co-hexafluoropropylene) (PVDF-co-HFP) and 1-ethyl-3-methylimidazolium, ing of bis-pyrenyl end groups and naphthalene-diimide
where PVDF-co-HFP is the polar polymer network comprising VDF (which is crystalline chains, along with a non-tweezer combination of
and less polar) and HFP (which is amorphous and highly polar)200. f | Reversible ionic naphthalene-diimide end groups and mono-pyrenyl
transformation between the bromine functionalities of bromobutyl rubber and ionic groups204. Incorporating an ionic moiety into the same
imidazolium bromide groups results in self-healing rubber material201. π–π stacking system may provide technological oppor-
tunities for the development of conductive self-healing
(90% healing efficiency) and tougher hydrogels (0.2 MPa polymers, such as a cathode in lithium–sulfur batteries205.
stress at break)198 (Fig. 9d). In this system, high density of Regardless of the chemical reactions involved in the
weak bonds facilitates bond reforming, but the softness self-healing of polymers, the main challenge is the ability
enhances contact across the interface, thereby increas- of the network to rearrange upon mechanical damage.
ing the self-healing efficiency. In another example, With a few exceptions — such as some heterogeneous
star-shaped PEG chains functionalized with alendronate networks206 — recent studies have focused on low-Tg

NAture Reviews | MaterialS volume 5 | August 2020 | 575


Reviews

polymers. Another challenge is to achieve self-healing properties. The network rearrangements impart high
in higher-Tg polymers, in which the limited free volume mechanical recyclability and the ability to undergo ther-
diminishes the segmental mobility of macromolecular mal healing, as well as rapid stress relaxations, which are
chains and diffusion is unfavourable. Thus, the pres- not found in conventional crosslinked thermoset poly-
ence of localized low-Tg and high-Tg components may mers. These networks can also exhibit shape-memory
be necessary for achieving self-healing in functional properties, allowing for autonomous arrangement
materials. into complex shapes through dynamic crosslinks by
the introduction of exchangeable chemical bonds.
Vitrimers A unique property of epoxy-based vitrimers is that their
Vitrimers are a relatively new class of synthetic material Tg can be controlled by tuning the rates of transester-
that resemble enzymatic cleavage of linkers in biologi- ification210 (Fig. 10b). An alternative to transesterifica-
cal systems. Their characteristic feature is exchangeable tion is catalyst-free transamidation exchange reactions.
covalent bonds, which, upon cleavage, can reshuffle207,208. Because amide groups are thermodynamically more
The exchange may be triggered by thermal, chemical or stable than esters, the resulting vitrimer networks are
other stimuli, and rigidity and plasticity are tunable209. less susceptible to hydrolysis208 (Fig. 10c).
Vitrimers form covalent networks capable of changing Another attractive class of vitrimers is based on
their topology through thermoactivated bond-exchange bis(cyclic carbonate)s, which can react with triamines or
reactions (Fig. 10a). At high temperatures, they flow and hydroxyl groups211 (Fig. 10d). Using transamination reac-
behave like viscoelastic liquids; however, at low temper- tions (Fig. 10e), vinylogous urethane networks can be pro-
atures, slow exchange reactions result in thermoset-like duced via the condensation reaction between acetoacetates

a Concept b Transesterification exchange reactions


O Heat O
R1 + R2OH R2 + R1OH
O O
Catalyst
c Transamidation exchange reactions
O Heat O
R1 + R2NH2 R2 + R1NH2
O N
Catalyst H

d Transcarbamoylation exchange reactions


H Heat H
N O N O
R1 + R OH R2 + R1OH
2
O Catalyst O

e Transamination of vinylogous amides or urethanes


R1 Heat R2
O HN O HN
+ R2NH2 + R1NH2
X Catalyst X

X = CH2: Vinylogous amide


X = O: Vinylogous urethane
f Transcarbonation exchange reactions
O O

R1 O O R1 + R2 OH
R1 O O R2 + R1 OH
Catalyst
g Dioxaborolane metathesis
O O
B + B
O O

O O O O
B + B + B + B
O O O O

Fig. 10 | Vitrimer systems. a | General concept of topological rearrangements through exchange reactions in vitrimers.
b | Transesterification of hydroxy-ester networks207. c | Transamidation exchange reactions in polyhydroxyurethane
networks208. d | Transcarbamoylation reactions leading to polyhydroxyurethanes211. e | Transamination of vinylogous
amides or urethanes208. f | Transcarbonation reactions leading to polycarbonates212. g | Dioxaborolane metathesis
reactions213. Panel a reprinted with permission from ref.207, AAAS.

576 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

a b
–1.5 3.5
Non-equilibrated
A - Free chain ends Equilibrated

Cohesive-energy density (× 105 kJ m–3)


B - Tethered chain ends
–1
3

–0.5 ΔG < 0
ΔG (102 kJ mol–1)

spontaneous Healing curve


2.5
0 A
–RTInN
ΔG = 0
0.5 B
2 Self-healable
A (N, KT) = (20, 0.63) range
1
80 B (N, KF) = (20, 0.56)
65 ΔG > 0 1.5
N 50 requires Eext 0 20 40 60 80 100
un - rep
its ea 35 0.75 0.8
tin 0.6 0.65 0.7 Molar % of A in p(A/B) copolymer
g 20
0.45 0.5 0.55
0.4
f - Flexibility parameter

Fig. 11 | Energy considerations of self-healing. a | A plot of the change in Gibbs free energy (ΔG = −TΔS) as a function of
the number of segments (N) in a polymer chain and the flexibility parameter (f) for free and tethered chain ends (surfaces A
and B, respectively). At equilibrium, ΔG = 0. The space above this plane (ΔG < 0) represents spontaneous repair, whereas
below it (ΔG > 0), self healing will not occur. The red dots mark the critical f values for free (KF) and tethered (KT) chain
ends for N = 20. The solid red curve shows how ΔG is influenced by N. Eext is the external energy input. b | Cohesive-
energy density (CED) as a function of molar ratio for non-equilibrated and equilibrated copolymers. Panel a adapted with
permission from ref.22, RSC.

and amines208. The bisacetoacetate monomers can be transition (Tv) (viscoelastic solid to viscoelastic liquid),
prepared from readily available diol monomers and their both controlled by crosslink density, monomer rigidity,
combination with commercially available polyamine the kinetics of exchange reactions and the concentra-
monomers allows tunability of the mechanical proper- tion of exchangeable bonds217. The presence of the two
ties of the poly(vinylogous urethane) polymers. Another transitions is an opportunity for developing self-healable
example of a vitrimer system is the reaction of bis(cyclic polymers using the exchange-reactions concept. One can
carbonate) and diols212 (Fig. 10f). The problem, however, envision that, if localized damage generates a sufficient
with the aforementioned examples is their need for a amount of heat to reach a Tv, thus triggering exchange
catalyst. By contrast, polyhydroxyurethane vitrimers can reactions, self-healing may occur for high-Tg polymers,
be formed without a catalyst211, as can those formed via as long as local damage generates sufficient segmental
metathesis reactions of dioxaborolanes213 (Fig. 10g). chain mobility to facilitate, for example, wound closure.
Vitrimers can either be introduced to other poly­mer
networks or have additives added to them to improve Energetic considerations
their properties. For example, vitrimer-based liquid- During the damage–repair cycle, a polymer’s response
crystalline elastomers exhibit processability owing to is in non-equilibrium state. One example is phase-
monodomain alignments, thus allowing robust con- separated morphologies, where each phase responds
struction of actuators with complex 3D structures; at different rates to damage, thus resulting in gradients
however, the stability of these elastomers needs to be of local volume expansion and different rates of filling
improved214. Introducing oligoaniline into vitrimer the empty space. To understand these processes, knowl-
networks resulted in a covalently crosslinked material edge of the rates of response for each phase is required.
that can respond to heat, light, pH, voltage, metal ions Assuming that mechanical damage causes chain cleavage
and redox chemicals215. This material exhibited shape or slippage near newly created surfaces, loose chain ends
memory, self-healing, recyclability, electrochromism and may or may not reform. A Gibbs free energy (∆G) < 0
adsorption of metal ions. As mentioned earlier, CNTs will favour recombination if the entropy (∆S) > 0.
embedded in common vitrimers are able to remotely Assuming that the damage–repair cycle is a transi-
trigger localized transesterification66 and, more inter- tion from a non-equilibrium to an equilibrium state
estingly, induce fast exchange reactions in crosslinked achieved by a series of infinitely small equilibrated steps
liquid-crystalline elastomers, thus allowing spatial con- described by the recoupling lattice model22,218, chain
trol of the alignment and fabrication of dynamic 3D flexibility — macroscopically reflected by the Tg — will
constructs in a short time216. influence the number of chain configurations. To pre-
The main challenge is designing self-healable poly- dict self-healing, a recoupling lattice model (Fig. 11a) can
mers with high thermal resistance. These materials typ- be used. This relates Flory’s flexibility parameter to the
ically exhibit high Tgs. However, in the design of vitrimer number of repeating units needed for a given ∆G (ref.219).
polymers, two transitions should be taken into account: Self-repair is favourable when the enthalpy (∆H) < 0,
T g (glassy to rubbery state) and topology-freezing implying that exothermic processes are dominated

NAture Reviews | MaterialS volume 5 | August 2020 | 577


Reviews

by macromolecular miscibility and favourable inter- high molecular weight) will self-heal via shape-memory
molecular and intramolecular chains (defined by the effects owing to their ability to store entropic energy, but
Flory–Huggins parameter218). that low-junction-density polymers will self-heal as a
In copolymers, the composition determines the CED consequence of surface-tension effects. For example, in
— the minimal amount of energy required to remove thermoresponsive polyurethanes formed via the Diels–
(or add) a macromolecular chain from a surrounding Alder reaction between furan and maleimide moieties,
network. CED values plotted as a function of the copoly­ the shape-memory effect facilitated self-healing with-
mer composition predict a linear dependence in the out the need for an external force (in contrast to other
non-equilibrated state (Fig. 11b). These predictions show self-healing polymers)86. Expanding the application tem-
that when the molar fraction of A in the p(A/B) copoly- perature range to 130 °C and thermal stability to 250 °C
mer (A with higher Tg and B with lower Tg) increases, the (compared with 60 °C for more common self-healing
CED values gradually increase, owing to the larger con- polymers based on polycaprolactone222) makes these
tent of B units in the chain. However, when the system is materials particularly attractive.
equilibrated, taking into account composition-dependent Self-healing is a local phenomenon with macroscopic
conformations, a maximum CED value is reached in the consequences. If polymer components exhibit local-
vicinity of the self-healing composition18. The maxi- ized, endothermic, stimuli-responsive transitions223,224
mum molar ratio corresponds to copolymer conforma- below the Tg at or near the damaged surface, macro-
tions that preferentially exhibit helical configurations molecular flow will occur225,226 and repair may occur.
(self-healing is typically not observed for other copoly­ In the presence of localized heterogeneities, such as in
mer compositions or for homopolymer blends of the elastomers, concurrent H-bonding and covalent-like
same compositions). One can envision a copolymer bonded polyvalent clusters introduce balance between
backbone with a helical topology as a spring, which, the steady state and non-equilibrium states15. The term
upon compression, is distorted. When an external force is autonomous self-healing is often used to emphasize
released, the ‘spring’ decompresses, bringing the cleaved self-healing under ambient conditions, which implies
ends together to self-heal. If low Tg was the only driv- a bulk phenomenon. However, the local dynamics of
ing force for self-healing, compositions with the higher macromolecular segments, reactivity of reactive compo-
B content would also exhibit self-healing characteris- nents (such as catalysts or reactive groups) and/or stere-
tics; however, they do not and, instead, they may form ospecificity are responsible for temperature-dependent
sticky, flowing films. These materials are commercially entropic and enthalpic contributions to the Gibbs free
known as pressure-sensitive adhesives. energy, leading to self-healing.
Polymer-chain separation or rupture — an outcome
of physical damage — creates isolated chain ends, From molecular to microrepair
which can be in a form of loose chain ends, free radicals Self-healing in biological systems occurs across all length
(depending on the polymer) or other reactive groups. scales. At the molecular level, biological systems use met-
The repair process begins longitudinally along the bot- abolic processes, such as in autophagy, to self-heal227.
tom of a scratch owing to energetically favourable inter- Self-healing materials can have similar metabolic charac-
actions, manifested by lower near-surface Tg. During ter if ‘outdated’ or degradation products are replaced by
this process, a percolation transition (that is, the min- new components, eliminating undesirable by-products.
imum threshold of reconnected ends or recombined A problem is that side reactions may lead to undesirable
chains) occurs after a given number of macromolecules products and are not easily controllable. However, com-
are connected longitudinally. The driving force is high bining the attributes of two or more self-healing motives in
interfacial energy in the smaller-curvature areas at the synthetic materials may lead to autonomous self-healing
bottom of a scratch220, which leads to close physical with a metabolic character. The future of self-healing poly­
proximity, high segmental mobility and entropic energy mer networks may comprise controllable supramolecular
storage during damage. If the interfacial energy and networks or van der Waals interactions, which do not
entropy can be recovered during repair, the material will generate side reactions or involve covalent rebonding.
autonomously self-heal. At the atomic scale of the covalent bond length in
If polymer chains are in extended conformations, synthetic and natural materials, cleavage of aliphatic
mechanical damage may lead to their compression. chains leads to the formation of CH or CH2 free radi-
Upon release of an applied force, decompression driven cals, or slippage of polymer chains. For chain cleavage,
by stored entropic and interfacial energies of spring-like simple geometrical considerations indicate that, relative
copolymer segments will restore their initial state. The to terminal −CH3 represented by sp3 hybridization, CH
spring model of the shape-memory effect resulting and CH2 free radicals adopt more directional sp and
from entropy storage during mechanical damage may sp2 hybridizations. The distances between the free rad-
also explain the driving force for self-healing behav- icals to initiate coupling are roughly 2.35–3.44 Å, and
iour; however, validation requires quantitative analy­ an energy of ~20 kcal mol−1 is typically required. When
sis. In fact, the maximum storable strain (and, thus, radical coupling occurs, the bonding distances decrease
shape-memory storage capacity) can be accurately pre- to 0.95–1.60 Å and the energy required decreases to
dicted using the junction density (that is, chemical or ~10 kcal mol−1. For radical recoupling to occur, inter-
physical crosslinks) and shape-memory factor221. chain diffusion should occur, which is often driven
We anticipate that high-junction-density polymers by the access of interfacial energy in the damaged area
(that is, those with a high number of entanglements and and the Tg-dependent kinetics of interdiffusion (Eq. 1).

578 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

At the nanometre scale of skeletal muscle fibres, For example, the rolling motion of microcapsules
mesogenic precursors are activated in the site of an was modelled using Brownian dynamics on an
injured fibre, leading to the proliferation of myoblasts, adhesive-coated substrate to design particle-filled
which repair damage by differentiating into multi­ microcapsules that heal cracks in the substrate236. To
nucleated myocytes2. By analogy with supramolecular model fracture and self-healing, most models adopt
self-healing chemistry, satellite and myelomonocytic continuum-damage approaches in which cracking and
cells also contribute to repair processes, in which dys- healing are considered as the degradation and recov-
ferlin acts as a Ca2+ sensor of the muscle membrane dam- ery of materials properties. By contrast, cohesive-zone
age, triggering vesicle fusion and directing Ca2+ along approaches treat damage as a discrete event, allowing
the membrane to seal the lesion by sending messages to explicit modelling of crack evolution237, including simu-
neutrophils for repair3,4. lation of property recovery over repeated healing events.
An example of cascading processes at the micrometre Looking forwards, there are opportunities for using and
and millimetre scales are skin-injury wounds, which developing computational models that could tackle 3D
penetrate into the dermis layer of skin, to which the red self-healing in heterogeneous systems. Collectively, com-
blood cells transfer as a result of rupture of blood ves- puter simulations and modelling are useful for answer-
sels. Platelets and inflammatory cells or cytokines rush ing many of the puzzling questions that are not easily
to the site of the injury and receptors signal the event by experimentally measured, particularly for predicting
activating fibroblasts and other connective-tissue cells dynamic materials properties238–240.
to deposit collagen. As a result, new tissue at the injury
site and wound healing occurs5. The cascading processes Outlook
leading to such self-healing (not just in mammals but The past couple of decades have brought remarkable
also in plants) brings a different perspective to designing advances in the controllable synthesis of self-healing
self-healing in synthetic materials. However, it is neces- polymers and the development of ‘living-like’ program-
sary to answer fundamental questions of how localized mable polymeric materials. The synthetic capabilities
individual reactions are capable of generating cascading afforded by continuously improving polymerizations
microscale responses leading to macroscale self-healing. provide access to new, controlled-architecture macro-
This requires the measurement of the molecular-level molecules with precision placement of the functional-
events — usually inside a scratch — responsible for ity needed for self-healing. New strategies for installing
self-healing of materials. self-healable moieties into existing commodity polymers
have further expanded the preparation toolbox. Further
Tools for studying self-healing manipulation of molecular weight and molecular-weight
Self-healing is typically measured by bulk mechanical distribution, molecular architecture, functional-group
analysis, in which properties before and after damage placement and precise copolymer compositions will
are assessed (ideally, these should match). Although use- yield self-healing polymers with new properties.
ful, this analysis does not reveal the molecular processes However, the technological success of self-healing poly­
responsible for self-healing. To identify mass transport mers will depend on how commodity (co)polymers can
during the damage–repair cycle, atomic force micros- be cost-effectively converted into their self-healable
copy may be useful if acquisition times are fast enough counterparts via precisely controllable and affordable
to capture self-healing. The same is applicable to optical polymerization processes, and how the structural under-
microscopy. standing can be translated into specific functionalities
Many self-healing mechanisms are not trivial to study and applications.
experimentally. For detecting local molecular events, In addition to one-time and repeated-cycle self-
the most sensitive spectroscopic tools are infrared228 healing, there are opportunities and challenges in the
and Raman spectroscopy229, which provide satisfactory development of sequential and parallel reactions that
spatial resolution. By contrast, electron paramagnetic may lead to physical remodelling at macroscopic scales.
resonance230 and NMR spectroscopy231,232, in particular, Localized heterogeneous macromolecular networks
nuclear Overhauser effect spectroscopy (NOESY) and capable of transient communication and signalling of
correlated spectroscopy (COSY) 2D 1H NMR exhibit damage that enable synchronous macroscopic rear-
decent sensitivity but insufficient spatial resolution. A rangements will become increasingly important in the
promising approach for monitoring self-healing events development of self-healing materials. Localized Tg
in polymers is dynamic nuclear polarization (DNP) gradients and associated local stimuli-responsive tran-
NMR spectroscopy, if it were applied to assess confor- sitions241,242 will become crucial for achieving reasonable
mational changes of dangling chain ends at the interfa- self-healing kinetics. It is interesting to note that some
cial regions generated during damage. In fact, the DNP of the observations and questions from the late 1970s
approach has the potential to improve the sensitivity of and early 1980s remain unanswered. For example, for
1
H NMR spectroscopy up to ~660-fold. Solid-state DNP pMMA slightly above Tg, a diffusional interpenetra-
13
C spectroscopy has been used in microporous organic tion of chain segments occurs, but after short pene-
polymers 233 but requires an external bis-nitroxide tration times (~5 min), fracture toughness is regained
radical-polarizing agent. Other intrinsic radicals as in short-term experiments, but long-term properties
polarizing agents can also be used234,235. cannot be restored243.
Computational research is increasingly important The ultimate challenge will be to produce polymer
to predict or explain observed self-healing behaviour. networks with metabolic and anabolic characteristics,

NAture Reviews | MaterialS volume 5 | August 2020 | 579


Reviews

whereby synchronized events between sequential or mechanisms leading to macroscopic reconfiguration of


concurrent self-healing paths occur. An approach in this the materials. The challenge is how to implement inter-
direction is combining phase-separated supramolecular nal directional actuation mechanisms while retaining
and covalent bonding with hydrophilic and hydrophobic physical and chemical stability. If answered, technolog-
interactions. For example, learning from nature, when a ical advances may enable new strategies for materials
protein composed of a hydrophobic packed interior and developments. It is not hard to envision that hip or other
hydrophilic exterior is placed in an aqueous environment body part replacements may have extended lifetimes
and exposed to hydrostatic pressure (or cutting), water when self-healing materials are used. Biological sens-
molecules are forced into the protein interior by filling ing and signalling pathways are mediated by interfacial
cavities and breaking up the hydrophobic structure244,245. processes linked to biological growth and development.
When the forces are removed, hydrophobic inter­actions One can envision that self-healing will be crucial for
prevail and the protein regains its structural and func- interactions between synthetic and biological systems.
tional features. There are opportunities for further To achieve this, we need to better understand the molec-
advances when hydrophobic and hydrophilic interactions ular processes that govern biosynthetic signalling and
in synthetic polymers compete during the damage–repair decision-making processes in synthetic materials. Using
cycle. Similar to biological systems, the key to controlling the analogy of metabolic processes in cells, we envision
these macromolecular interactions is precise molecular that, if repeating units of a copolymer represent a her-
DNA-like sequencing. From a macroscopic view, the for- itable genetic identity derived during its growth, differ-
mation of self-knotted polymer-chain topologies promises ent interfacial interactions (the ‘genotype’) will lead to
the ability of self-healing of fibres in composites. different materials ‘phenotypes’.
Applications of self-healing polymers are likely to Overall, the outcomes will be less energy used and less
be found in agriculture, the food industry, medicine, waste generated. Self-healable polymers with more tun-
transportation, recycling and upcycling. All these sec- able time-sensitive properties are needed; a grocery bag
tors face challenges to meet new requirements and reg- does not have to last more than a day or two, but paint
ulations, and polymers with upcycling and self-healing on a bridge or vehicle needs to last much longer. An ulti-
attributes may become a new standard for many indus- mate goal of future studies should lead to the develop-
tries. Despite the urgency of agricultural needs, our ment of organism-like materials with encoded molecular
understanding of how plants and other species trans- features that dictate their growth and structural assembly
port nutrients to their destinations in vivo is still largely in response to the environment. Organism-like materials
unanswered. Similarly, there are many unanswered or will consist of elements responsible for autonomous and
unasked questions in self-healing polymers. One such adaptive properties enabling the creation of synthetic
question is how the shape and symmetry of a struc- integrated systems and devices.
ture will determine the arrangements of the building
blocks enabling self-healing through internal actuation Published online 5 June 2020

1. Diegelmann, R. F. & Evans, M. C. Wound healing: 14. Imato, K. et al. Self-healing of chemical gels cross-linked 27. Wool, R. P. Polymer Interfaces: Structure and Strength
an overview of acute, fibrotic and delayed healing. by diarylbibenzofuranone-based trigger-free dynamic (Hanser Publishers, 1995).
Front. Biosci. 9, 283–289 (2004). covalent bonds at room temperature. Angew. Chem. 28. Sperling, L. H. in Introduction to Physical Polymer
2. Han, R. & Campbell, K. P. Dysferlin and muscle Int. Ed. 51, 1138–1142 (2012). Science Ch. 4.4 (John Wiley & Sons, 2005).
membrane repair. Curr. Opin. Cell Biol. 19, 409–416 15. Cordier, P., Tournilhac, F., Soulié-Ziakovic, C. 29. Welp, K. A. et al. Direct observation of polymer
(2007). & Leibler, L. Self-healing and thermoreversible rubber dynamics: mobility comparison between central and
3. París, R., Lamattina, L. & Casalongué, C. A. Nitric from supramolecular assembly. Nature 451, 977–980 end section chain segments. Macromolecules 32,
oxide promotes the wound-healing response of potato (2008). 5127–5138 (1999).
leaflets. Plant Physiol. Biochem. 45, 80–86 (2007). 16. Burnworth, M. et al. Optically healable supramolecular 30. Ellison, C. J. & Torkelson, J. M. The distribution of
4. Biggs, A. Suberized boundary zones and the chronology polymers. Nature 472, 334–337 (2011). glass-transition temperatures in nanoscopically
of wound response in tree bark. Phytopathology 75, 17. Nakahata, M., Takashima, Y., Yamaguchi, H. & confined glass formers. Nat. Mater. 2, 695–700
1191–1195 (1985). Harada, A. Redox-responsive self-healing materials (2003).
5. Wool, R. P. & O’Connor, K. M. A theory crack healing formed from host–guest polymers. Nat. Commun. 2, 31. Bodiguel, H. & Fretigny, C. Reduced viscosity in thin
in polymers. J. Appl. Phys. 52, 5953–5963 (1981). 511 (2011). polymer films. Phys. Rev. Lett. 97, 266105 (2006).
6. Yang, Y., Davydovich, D., Hornat, C. C., Liu, X. & 18. Urban, M. W. et al. Key-and-lock commodity self-healing 32. Fakhraai, Z. & Forrest, J. A. Measuring the surface
Urban, M. W. Leaf-inspired self-healing polymers. copolymers. Science 362, 220–225 (2018). dynamics of glassy polymers. Science 319, 600–604
Chem 4, 1928–1936 (2018). 19. White, S. R. et al. Autonomic healing of polymer (2008).
7. Chen, Y., Kushner, A. M., Williams, G. A. & Guan, Z. composites. Nature 409, 794–797 (2001). 33. Ghosh, B., Chellappan, K. V. & Urban, M. W.
Multiphase design of autonomic self-healing 20. Kessler, M. R., Sottos, N. R. & White, S. R. Self-healing Self-healing inside a scratch of oxetane-substituted
thermoplastic elastomers. Nat. Chem. 4, 467–472 structural composite materials. Compos. Part A Appl. chitosan-polyurethane (OXE-CHI-PUR) networks.
(2012). Sci. Manuf. 34, 743–753 (2003). J. Mater. Chem. 21, 14473–14486 (2011).
8. Nji, J. & Li, G. A biomimic shape memory polymer 21. Wool, R. P. Self-healing materials: a review. Soft Matter 34. de Gennes, P.-G. Reptation of a polymer chain in the
based self-healing particulate composite. Polymer 51, 4, 400–418 (2008). presence of fixed obstacles. J. Chem. Phys. 55,
6021–6029 (2010). 22. Yang, Y. & Urban, M. W. Self-healing polymeric 572–579 (1971).
9. Corten, C. C. & Urban, M. W. Repairing polymers materials. Chem. Soc. Rev. 42, 7446–7467 (2013). 35. Klein, J. Evidence for reptation in an entangled
using oscillating magnetic field. Adv. Mater. 21, 23. Prager, S. & Tirrell, M. The healing process at polymer melt. Nature 271, 143–145 (1978).
5011–5015 (2009). polymer–polymer interfaces. J. Chem. Phys. 75, 36. Roland, C. M. & Ngai, K. L. Segmental relaxation and
10. Yang, Y. & Urban, M. W. Self-repairable polyurethane 5194–5198 (1981). the correlation of time and temperature dependencies
networks by atmospheric carbon dioxide and water. 24. Voyutskii, S. S. Autohesion and Adhesion of High in poly(vinyl methyl ether)/polystyrene mixtures.
Angew. Chem. Int. Ed. 53, 12142–12147 (2014). Polymers (Interscience Publishers, 1963). Macromolecules 25, 363–367 (1992).
11. Ying, H., Zhang, Y. & Cheng, J. Dynamic urea bond for 25. Grinsted, R. A., Clark, L. & Koenig, J. L. Study of cyclic 37. Kim, Y. H. & Wool, R. P. A theory of healing at a
the design of reversible and self-healing polymers. sorption-desorption into poly(methyl methacrylate) polymer-polymer interface. Macromolecules 16,
Nat. Commun. 5, 3218 (2014). rods using NMR imaging. Macromolecules 25, 1115–1120 (1983).
12. Chen, X. et al. A thermally re-mendable cross-linked 1235–1241 (1992). 38. Lin, C., Lee, S. & Liu, K. Methanol-induced crack
polymeric material. Science 295, 1698–1702 26. Kim, K. D., Sperling, L. H., Klein, A. & Hammouda, B. healing in poly(methyl methacrylate). Polym. Eng. Sci.
(2002). Reptation time, temperature, and cosurfactant effects 30, 1399–1406 (1990).
13. Ghosh, B. & Urban, M. W. Self-repairing oxetane- on the molecular interdiffusion rate during polystyrene 39. Jud, K., Kausch, H. H. & Williams, J. G. Fracture
substituted chitosan polyurethane networks. Science latex film formation. Macromolecules 27, 6841–6850 mechanics studies of crack healing and welding of
323, 1458–1460 (2009). (1994). polymers. J. Mater. Sci. 16, 204–210 (1981).

580 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

40. Autumn, K. et al. Evidence for van der Waals adhesion 68. Sato, K. et al. Phase-separation-induced anomalous with one-way and two-way shape-memory properties
in gecko setae. Proc. Natl Acad. Sci. USA 99, stiffening, toughening, and self-healing of through Diels–Alder reactions. Chem. Eur. J. 17,
12252–12256 (2002). polyacrylamide gels. Adv. Mater. 27, 6990–6998 10135–10143 (2011).
41. Buckingham, A., Fowler, P. & Hutson, J. M. Theoretical (2015). 92. Sedaghat-Herati, R., Chacon, A., Hansen, M. E. &
studies of van der Waals molecules and intermolecular 69. Chen, S., Mahmood, N., Beiner, M. & Binder, W. H. Yalaoui, S. New poly(oxyethylene) derivatives from
forces. Chem. Rev. 88, 963–988 (1988). Self-healing materials from V- and H-shaped Diels–Alder reactions of 3-[methoxypoly(oxyethylene)]
42. Brunauer, S., Deming, L. S., Deming, W. E. & Teller, E. supramolecular architectures. Angew. Chem. Int. Ed. methylene furan. Macromol. Chem. Phys. 206,
On a theory of the van der Waals adsorption of gases. 54, 10188–10192 (2015). 1981–1987 (2005).
J. Am. Chem. Soc. 62, 1723–1732 (1940). 70. Cao, J. et al. Multiple hydrogen bonding enables 93. Watanabe, M. & Yoshie, N. Synthesis and properties
43. Dzyaloshinskii, I. E., Lifshitz, E. M., Pitaevskii, L. P. & the self-healing of sensors for human–machine of readily recyclable polymers from bisfuranic
Priestley, M. G. in Perspectives in Theoretical Physics interactions. Angew. Chem. Int. Ed. 56, 8795–8800 terminated poly(ethylene adipate) and multi-
(ed. Pitaevskii, L. P., translated from Russian by Sykes, (2017). maleimide linkers. Polymer 47, 4946–4952 (2006).
J. B. & ter Haar, D.) 443–492 (Elsevier, 1992). 71. Yan, X. et al. Quadruple H-bonding cross-linked 94. Yamashiro, M., Inoue, K. & Iji, M. Recyclable
44. Sun, H. COMPASS: an ab initio force-field optimized supramolecular polymeric materials as substrates shape-memory and mechanical strength of poly(lactic
for condensed-phase applications overview with for stretchable, antitearing, and self-healable thin acid) compounds cross-linked by thermo-reversible
details on alkane and benzene compounds. J. Phys. film electrodes. J. Am. Chem. Soc. 140, 5280–5289 Diels-Alder reaction. Polym. J. 40, 657–662 (2008).
Chem. B 102, 7338–7364 (1998). (2018). 95. Kavitha, A. A. & Singha, N. K. “Click chemistry” in
45. Bharadwaj, R. K., Berry, R. J. & Farmer, B. L. 72. Ghosh, B., Chellappan, K. V. & Urban, M. W. UV-initiated tailor-made polymethacrylates bearing reactive
Molecular dynamics simulation study of norbornene– self-healing of oxolane–chitosan–polyurethane furfuryl functionality: a new class of self-healing
POSS polymers. Polymer 41, 7209–7221 (2000). (OXO–CHI–PUR) networks. J. Mater. Chem. 22, polymeric material. ACS Appl. Mater. Interfaces 1,
46. Prathab, B., Subramanian, V. & Aminabhavi, T. 16104–16113 (2012). 1427–1436 (2009).
Molecular dynamics simulations to investigate 73. Korth, H. G. Carbon radicals of low reactivity against 96. Kavitha, A. A. & Singha, N. K. Smart “all acrylate”
polymer–polymer and polymer–metal oxide oxygen: radically different antioxidants. Angew. Chem. ABA triblock copolymer bearing reactive functionality
interactions. Polymer 48, 409–416 (2007). Int. Ed. 46, 5274–5276 (2007). via atom transfer radical polymerization (ATRP):
47. Speck, O., Schlechtendahl, M., Borm, F., Kampowski, T. 74. Takeda, K., Unno, H. & Zhang, M. Polymer reaction in demonstration of a “click reaction” in thermoreversible
& Speck, T. Humidity-dependent wound sealing in polycarbonate with Na2CO3. J. Appl. Polym. Sci. 93, property. Macromolecules 43, 3193–3205 (2010).
succulent leaves of Delosperma cooperi–An adaptation 920–926 (2004). 97. Chung, C.-M., Roh, Y.-S., Cho, S.-Y. & Kim, J.-G. Crack
to seasonal drought stress. Beilstein J. Nanotechnol. 9, 75. Stevens, M. P. & Jenkins, A. D. Crosslinking of healing in polymeric materials via photochemical
175–186 (2018). polystyrene via pendant maleimide groups. J. Polym. [2+2] cycloaddition. Chem. Mater. 16, 3982–3984
48. Vernon, L. B. & Vernon, H. M. Process of Sci. Polym. Chem. Ed. 17, 3675–3685 (1979). (2004).
manufacturing articles of thermoplastic synthetic 76. Liu, Y. L. & Chen, Y. W. Thermally reversible 98. Egerton, P. L. et al. Photocycloaddition in liquid
resins. US Patent 2234993 (1941). cross-linked polyamides with high toughness ethyl cinnamate and in ethyl cinnamate glasses. The
49. Rainer, W. C., Redding, E. M., Hitov, J. J., Sloan, A. W. and self-repairing ability from maleimide- and photoreaction as a probe into the micromorphology
& Stewart, W. D. Heat-shrinkable polyethylene. US furan-functionalized aromatic polyamides. Macromol. of the solid. J. Am. Chem. Soc. 103, 3859–3863
Patent 3144398 (1964). Chem. Phys. 208, 224–232 (2007). (1981).
50. Perrone, R. J. Silicone-rubber, polyethylene 77. Imato, K. et al. Dynamic covalent 99. Guimard, N. K. et al. Harnessing entropy to direct
composition; heat shrinkable articles made therefrom diarylbibenzofuranone-modified nanocellulose: the bonding/debonding of polymer systems based
and process therefor. US Patent 3326869 (1967). Mechanochromic behaviour and application in on reversible chemistry. Chem. Sci. 4, 2752–2759
51. Cussler, E. L. Diffusion: Mass Transfer in Fluid Systems self-healing polymer composites. Polym. Chem. 8, (2013).
3rd edn Ch. 5 (Cambridge Univ. Press, 2009). 2115–2122 (2017). 100. Oehlenschlaeger, K. K. et al. Fast and catalyst-free
52. Habault, D., Zhang, H. & Zhao, Y. Light-triggered self- 78. Telitel, S. et al. Introduction of self-healing hetero-Diels–Alder chemistry for on demand cyclable
healing and shape-memory polymers. Chem. Soc. Rev. properties into covalent polymer networks via bonding/debonding materials. Polym. Chem. 4,
42, 7244–7256 (2013). the photodissociation of alkoxyamine junctions. 4348–4355 (2013).
53. Kirkby, E. L. et al. Embedded shape-memory Polym. Chem. 5, 921–930 (2014). 101. Oehlenschlaeger, K. K. et al. Adaptable hetero
alloy wires for improved performance of self-healing 79. An, Q. et al. Recycling and self-healing of dynamic Diels–Alder networks for fast self-healing under
polymers. Adv. Funct. Mater. 18, 2253–2260 (2008). covalent polymer networks with a precisely tuneable mild conditions. Adv. Mater. 26, 3561–3566
54. Li, G. & Shojaei, A. A viscoplastic theory of shape crosslinking degree. Polym. Chem. 10, 672–678 (2014).
memory polymer fibres with application to self-healing (2019). 102. Stocking, E. M. & Williams, R. M. Chemistry and
materials. Proc. R. Soc. A 468, 2319–2346 (2012). 80. Raines, C. A. The Calvin cycle revisited. Photosynth. biology of biosynthetic Diels–Alder reactions. Angew.
55. Mohr, R. et al. Initiation of shape-memory effect by Res. 75, 1–10 (2003). Chem. Int. Ed. 42, 3078–3115 (2003).
inductive heating of magnetic nanoparticles in 81. Bai, N., Saito, K. & Simon, G. P. Synthesis of a diamine 103. Hoyle, C. E., Lee, T. Y. & Roper, T. Thiol–enes:
thermoplastic polymers. Proc. Natl Acad. Sci. USA cross-linker containing Diels–Alder adducts to produce chemistry of the past with promise for the future.
103, 3540–3545 (2006). self-healing thermosetting epoxy polymer from a J. Polym. Sci. Part A Polym. Chem. 42, 5301–5338
56. Huang, W. M., Yang, B., An, L., Li, C. & Chan, Y. S. widely used epoxy monomer. Polym. Chem. 4, (2004).
Water-driven programmable polyurethane shape 724–730 (2013). 104. Kade, M. J., Burke, D. J. & Hawker, C. J. The power of
memory polymer: demonstration and mechanism. 82. Peterson, A. M., Jensen, R. E. & Palmese, G. R. thiol-ene chemistry. J. Polym. Sci. Part A Polym. Chem.
Appl. Phys. Lett. 86, 114105 (2005). Reversibly cross-linked polymer gels as healing 48, 743–750 (2010).
57. Lendlein, A., Jiang, H., Jünger, O. & Langer, R. agents for epoxy–amine thermosets. ACS Appl. Mater. 105. Nicolay, R., Kamada, J., Van Wassen, A. &
Light-induced shape-memory polymers. Nature 434, Interfaces 1, 992–995 (2009). Matyjaszewski, K. Responsive gels based on a dynamic
879–882 (2005). 83. Tian, Q., Yuan, Y. C., Rong, M. Z. & Zhang, M. Q. covalent trithiocarbonate cross-linker. Macromolecules
58. Hornat, C. C., Yang, Y. & Urban, M. W. Quantitative A thermally remendable epoxy resin. J. Mater. Chem. 43, 4355–4361 (2010).
predictions of shape-memory effects in polymers. 19, 1289–1296 (2009). 106. Kamada, J. et al. Redox responsive behavior of thiol/
Adv. Mater. 29, 1603334 (2017). 84. Chen, X., Wudl, F., Mal, A. K., Shen, H. & Nutt, S. R. disulfide-functionalized star polymers synthesized via
59. Wang, H. B. et al. Synthesis of self-healing polymers New thermally remendable highly cross-linked atom transfer radical polymerization. Macromolecules
by scandium-catalyzed copolymerization of ethylene polymeric materials. Macromolecules 36, 1802–1807 43, 4133–4139 (2010).
and anisylpropylenes. J. Am. Chem. Soc. 141, (2003). 107. Yoon, J. A. et al. Self-healing polymer films based on
3249–3257 (2019). 85. Billiet, S., Van Camp, W., Hillewaere, X. K. D., Rahier, H. thiol–disulfide exchange reactions and self-healing
60. Hornat, C. C. & Urban, M. W. Shape memory effects in & Du Prez, F. E. Development of optimized autonomous kinetics measured using atomic force microscopy.
self-healing polymers. Prog. Polym. Sci. 102, 101208 self-healing systems for epoxy materials based on Macromolecules 45, 142–149 (2011).
(2020). maleimide chemistry. Polymer 53, 2320–2326 108. Kuhl, N. et al. Acylhydrazones as reversible covalent
61. Hornat, C. C. & Urban, M. W. Entropy and interfacial (2012). crosslinkers for self-healing polymers. Adv. Funct. Mater.
energy driven self-healable polymers. Nat. Commun. 86. Heo, Y. & Sodano, H. A. Self-healing polyurethanes 25, 3295–3301 (2015).
11, 1028 (2020). with shape recovery. Adv. Funct. Mater. 24, 109. Barcan, G. A., Zhang, X. Y. & Waymouth, R. M.
62. Murphy, E. B. & Wudl, F. The world of smart healable 5261–5268 (2014). Structurally dynamic hydrogels derived from
materials. Prog. Polym. Sci. 35, 223–251 (2010). 87. Du, P. et al. Synthesis and characterization of 1,2-dithiolanes. J. Am. Chem. Soc. 137, 5650–5653
63. Yang, Y., Ding, X. & Urban, M. W. Chemical and physical linear self-healing polyurethane based on thermally (2015).
aspects of self-healing materials. Prog. Polym. Sci. reversible Diels–Alder reaction. RSC Adv. 3, 110. Rekondo, A. et al. Catalyst-free room-temperature
49–50, 34–59 (2015). 15475–15482 (2013). self-healing elastomers based on aromatic disulfide
64. Lee, M. W., Yoon, S. S. & Yarin, A. L. Solution-blown 88. Syrett, J. A., Mantovani, G., Barton, W. R., Price, D. & metathesis. Mater. Horiz. 1, 237–240 (2014).
core–shell self-healing nano- and microfibers. Haddleton, D. M. Self-healing polymers prepared 111. Xu, Y. & Chen, D. A novel self-healing polyurethane
ACS Appl. Mater. Interfaces 8, 4955–4962 (2016). via living radical polymerisation. Polym. Chem. 1, based on disulfide bonds. Macromol. Chem. Phys.
65. Pu, W. et al. Realizing crack diagnosing and 102–106 (2010). 217, 1191–1196 (2016).
self-healing by electricity with a dynamic crosslinked 89. Yoshie, N., Saito, S. & Oya, N. A thermally-stable 112. Canadell, J., Goossens, H. & Klumperman, B.
flexible polyurethane composite. Adv. Sci. 5, 1800101 self-mending polymer networked by Diels–Alder Self-healing materials based on disulfide links.
(2018). cycloaddition. Polymer 52, 6074–6079 (2011). Macromolecules 44, 2536–2541 (2011).
66. Yang, Y. et al. Carbon nanotube–vitrimer composite 90. Sugane, K., Yoshioka, Y., Shimasaki, T., Teramoto, N. 113. Ji, S., Cao, W., Yu, Y. & Xu, H. Visible-light-induced
for facile and efficient photo-welding of epoxy. & Shibata, M. Self-healing 8-armed star-shaped self-healing diselenide-containing polyurethane
Chem. Sci. 5, 3486–3492 (2014). ε-caprolactone oligomers dually crosslinked by the elastomer. Adv. Mater. 27, 7740–7745 (2015).
67. Chen, Y. & Guan, Z. Multivalent hydrogen bonding Diels-Alder and urethanization reactions. Polymer 114. An, X. et al. Aromatic diselenide crosslinkers to
block copolymers self-assemble into strong and 144, 92–102 (2018). enhance the reprocessability and self-healing of
tough self-healing materials. Chem. Commun. 50, 91. Raquez, J. M. et al. Design of cross-linked polyurethane thermosets. Polym. Chem. 8,
10868–10870 (2014). semicrystalline poly (ε-caprolactone)-based networks 3641–3646 (2017).

NAture Reviews | MaterialS volume 5 | August 2020 | 581


Reviews

115. Kuhl, N. et al. Self-healing polymer networks based 140. Yuan, C., Rong, M. Z., Zhang, M. Q., Zhang, Z. P. 167. Bode, S. et al. Self-healing polymer coatings based
on reversible Michael addition reactions. Macromol. & Yuan, Y. C. Self-healing of polymers via synchronous on crosslinked metallosupramolecular copolymers.
Chem. Phys. 217, 2541–2550 (2016). covalent bond fission/radical recombination. Chem. Adv. Mater. 25, 1634–1638 (2013).
116. Kantor, S. W., Grubb, W. T. & Osthoff, R. C. The Mater. 23, 5076–5081 (2011). 168. Williams, K. A., Boydston, A. J. & Bielawski, C. W.
mechanism of the acid- and base-catalyzed 141. Amamoto, Y., Kamada, J., Otsuka, H., Takahara, A. Towards electrically conductive, self-healing materials.
equilibration of siloxanes. J. Am. Chem. Soc. 76, & Matyjaszewski, K. Repeatable photoinduced J. R. Soc. Interface 4, 359–362 (2007).
5190–5197 (1954). self-healing of covalently cross-linked polymers 169. Wang, Z. & Urban, M. W. Facile UV-healable
117. Zheng, P. & McCarthy, T. J. A surprise from 1954: through reshuffling of trithiocarbonate units. Angew. polyethylenimine–copper (C2H5N–Cu) supramolecular
siloxane equilibration is a simple, robust, and obvious Chem. Int. Ed. 50, 1660–1663 (2011). polymer networks. Polym. Chem. 4, 4897–4901
polymer self-healing mechanism. J. Am. Chem. Soc. 142. Nakahata, M., Mori, S., Takashima, Y., Yamaguchi, H. (2013).
134, 2024–2027 (2012). & Harada, A. Self-healing materials formed by 170. Wang, Z. H., Yang, Y., Burtovyy, R., Luzinov, I.
118. Wang, C. et al. Self-healing chemistry enables the cross-linked polyrotaxanes with reversible bonds. & Urban, M. W. UV-induced self-repairing
stable operation of silicon microparticle anodes for Chem 1, 766–775 (2016). polydimethylsiloxane–polyurethane (PDMS–PUR)
high-energy lithium-ion batteries. Nat. Chem. 5, 143. Gong, J. P. Why are double network hydrogels so and polyethylene glycol–polyurethane (PEG–PUR)
1042–1048 (2013). tough? Soft Matter 6, 2583–2590 (2010). Cu-catalyzed networks. J. Mater. Chem. A 2,
119. Xu, Z. et al. Silicon microparticle anodes with 144. Jia, H. et al. Unconventional tough double-network 15527–15534 (2014).
self-healing multiple network binder. Joule 2, hydrogels with rapid mechanical recovery, self-healing, 171. Rao, Y. L. et al. Stretchable self-healing polymeric
950–961 (2018). and self-gluing properties. ACS Appl. Mater. Interfaces dielectrics cross-linked through metal–ligand
120. Brochu, A. B. W., Craig, S. L. & Reichert, W. M. 8, 31339–31347 (2016). coordination. J. Am. Chem. Soc. 138, 6020–6027
Self-healing biomaterials. J. Biomed. Mater. Res. Part A 145. Webber, M. J., Appel, E. A., Meijer, E. & Langer, R. (2016).
96, 492–506 (2011). Supramolecular biomaterials. Nat. Mater. 15, 13–26 172. Ceylan, H. et al. Mussel inspired dynamic cross-linking
121. Madsen, F. B., Yu, L. & Skov, A. L. Self-healing, (2016). of self-healing peptide nanofiber network. Adv. Funct.
high-permittivity silicone dielectric elastomer. 146. Herbst, F., Döhler, D., Michael, P. & Binder, W. H. Mater. 23, 2081–2090 (2013).
ACS Macro Lett. 5, 1196–1200 (2016). Self-healing polymers via supramolecular forces. 173. Zeng, H., Hwang, D. S., Israelachvili, J. N. & Waite, J. H.
122. Martín, R. et al. Room temperature self-healing power Macromol. Rapid Commun. 34, 203–220 (2013). Strong reversible Fe3+-mediated bridging between
of silicone elastomers having silver nanoparticles 147. Pedersen, C. J. Cyclic polyethers and their complexes dopa-containing protein films in water. Proc. Natl Acad.
as crosslinkers. Chem. Commun. 48, 8255–8257 with metal salts. J. Am. Chem. Soc. 89, 7017–7036 Sci. USA 107, 12850–12853 (2010).
(2012). (1967). 174. Weng, G. S., Thanneeru, S. & He, J. Dynamic
123. Jin, B., Liu, M., Zhang, Q., Zhan, X. & Chen, F. Silicone 148. Kyba, E. P., Siegel, M. G., Sousa, L. R., Sogah, G. D. coordination of Eu–iminodiacetate to control
oil swelling slippery surfaces based on mussel-inspired & Cram, D. J. Chiral, hinged, and functionalized fluorochromic response of polymer hydrogels to
magnetic nanoparticles with multiple self-healing multiheteromacrocycles. J. Am. Chem. Soc. 95, multistimuli. Adv. Mater. 30, 1706526 (2018).
mechanisms. Langmuir 33, 10340–10350 (2017). 2691–2692 (1973). 175. Liu, S. L., Oderinde, O., Hussain, I., Yao, F. & Fu, G. D.
124. Ogliani, E., Yu, L., Javakhishvili, I. & Skov, A. L. 149. Brunsveld, L., Folmer, B., Meijer, E. W. & Sijbesma, R. Dual ionic cross-linked double network hydrogel with
A thermo-reversible silicone elastomer with remotely Supramolecular polymers. Chem. Rev. 101, self-healing, conductive, and force sensitive properties.
controlled self-healing. RSC Adv. 8, 8285–8291 (2018). 4071–4098 (2001). Polymer 144, 111–120 (2018).
125. Ramachandran, D., Liu, F. & Urban, M. W. 150. Fyfe, M. C. & Stoddart, J. F. Synthetic supramolecular 176. Luo, F. et al. Oppositely charged polyelectrolytes
Self-repairable copolymers that change color. RSC Adv. chemistry. Acc. Chem. Res. 30, 393–401 (1997). form tough, self-healing, and rebuildable hydrogels.
2, 135–143 (2012). 151. Herbst, F., Seiffert, S. & Binder, W. H. Dynamic Adv. Mater. 27, 2722–2727 (2015).
126. Zhao, X. et al. Antibacterial anti-oxidant electroactive supramolecular poly(isobutylene)s for self-healing 177. Zhong, M., Liu, Y. T. & Xie, X. M. Self-healable, super
injectable hydrogel as self-healing wound dressing with materials. Polym. Chem. 3, 3084–3092 (2012). tough graphene oxide–poly(acrylic acid) nanocomposite
hemostasis and adhesiveness for cutaneous wound 152. Sijbesma, R. P. et al. Reversible polymers formed hydrogels facilitated by dual cross-linking effects through
healing. Biomaterials 122, 34–47 (2017). from self-complementary monomers using quadruple dynamic ionic interactions. J. Mater. Chem. B 3,
127. Tseng, T. C. et al. An injectable, self-healing hydrogel hydrogen bonding. Science 278, 1601–1604 (1997). 4001–4008 (2015).
to repair the central nervous system. Adv. Mater. 27, 153. Aida, T., Meijer, E. & Stupp, S. Functional 178. Darabi, M. A. et al. Skin-inspired multifunctional
3518–3524 (2015). supramolecular polymers. Science 335, 813–817 autonomic-intrinsic conductive self-healing hydrogels
128. Yu, F., Cao, X. D., Du, J., Wang, G. & Chen, X. F. (2012). with pressure sensitivity, stretchability, and 3D
Multifunctional hydrogel with good structure integrity, 154. Hirschberg, J. K. et al. Supramolecular polymers from printability. Adv. Mater. 29, 1700533 (2017).
self-healing, and tissue-adhesive property formed by linear telechelic siloxanes with quadruple-hydrogen- 179. He, L., Fullenkamp, D. E., Rivera, J. G. &
combining Diels–Alder click reaction and acylhydrazone bonded units. Macromolecules 32, 2696–2705 Messersmith, P. B. pH responsive self-healing hydrogels
bond. ACS Appl. Mater. Interfaces 7, 24023–24031 (1999). formed by boronate–catechol complexation. Chem.
(2015). 155. Folmer, B. J. B., Sijbesma, R. P., Versteegen, R. M., Commun. 47, 7497–7499 (2011).
129. Qu, J. et al. Antibacterial adhesive injectable van der Rijt, J. A. J. & Meijer, E. W. Supramolecular 180. Ahn, B. K., Lee, D. W., Israelachvili, J. N. & Waite, J. H.
hydrogels with rapid self-healing, extensibility and polymer materials: Chain extension of telechelic Surface-initiated self-healing of polymers in aqueous
compressibility as wound dressing for joints skin polymers using a reactive hydrogen-bonding synthon. media. Nat. Mater. 13, 867–872 (2014).
wound healing. Biomaterials 183, 185–199 (2018). Adv. Mater. 12, 874–878 (2000). 181. Li, C.-H. et al. A highly stretchable autonomous
130. Ono, T., Nobori, T. & Lehn, J.-M. Dynamic polymer 156. Bosman, A. W., Sijbesma, R. P. & Meijer, E. W. self-healing elastomer. Nat. Chem. 8, 618–624
blends — component recombination between neat Supramolecular polymers at work. Mater. Today 7, (2016).
dynamic covalent polymers at room temperature. 34–39 (2004). 182. Li, Z. Q., Wang, G. N., Wang, Y. G. & Li, H. R.
Chem. Commun. 1522-1524 (2005). 157. Yanagisawa, Y., Nan, Y. L., Okuro, K. & Aida, T. Reversible phase transition of robust luminescent
131. Mukherjee, S., Hill, M. R. & Sumerlin, B. S. Mechanically robust, readily repairable polymers hybrid hydrogels. Angew. Chem. Int. Ed. 57,
Self-healing hydrogels containing reversible oxime via tailored noncovalent cross-linking. Science 359, 2194–2198 (2018).
crosslinks. Soft Matter 11, 6152–6161 (2015). 72–76 (2018). 183. Rodell, C. B., Dusaj, N. N., Highley, C. B.
132. Liu, W.-X. et al. Oxime-based and catalyst-free 158. Wu, Q. et al. A robust, highly stretchable & Burdick, J. A. Injectable and cytocompatible
dynamic covalent polyurethanes. J. Am. Chem. Soc. supramolecular polymer conductive hydrogel with tough double-network hydrogels through tandem
139, 8678–8684 (2017). self-healability and thermo-processability. Sci. Rep. 7, supramolecular and covalent crosslinking. Adv. Mater.
133. Niu, W., Smith, M. D. & Lavigne, J. J. Self-assembling 41566 (2017). 28, 8419–8424 (2016).
poly(dioxaborole)s as blue-emissive materials. J. Am. 159. Li, C. et al. A writable polypeptide–DNA hydrogel with 184. Loebel, C., Rodell, C. B., Chen, M. H. & Burdick, J. A.
Chem. Soc. 128, 16466–16467 (2006). rationally designed multi-modification sites. Small 11, Shear-thinning and self-healing hydrogels as injectable
134. De, P., Gondi, S. R., Roy, D. & Sumerlin, B. S. Boronic 1138–1143 (2015). therapeutics and for 3D-printing. Nat. Protoc. 12,
acid-terminated polymers: synthesis by RAFT and 160. Neal, J. A., Mozhdehi, D. & Guan, Z. Enhancing 1521–1541 (2017).
subsequent supramolecular and dynamic covalent mechanical performance of a covalent self-healing 185. Chen, H., Ma, X., Wu, S. F. & Tian, H. A rapidly
self-assembly. Macromolecules 42, 5614–5621 material by sacrificial noncovalent bonds. J. Am. self-healing supramolecular polymer hydrogel with
(2009). Chem. Soc. 137, 4846–4850 (2015). photostimulated room-temperature phosphorescence
135. Cash, J. J., Kubo, T., Bapat, A. P. & Sumerlin, B. S. 161. Feldman, K. E. et al. Polymers with multiple responsiveness. Angew. Chem. Int. Ed. 53,
Room-temperature self-healing polymers based on hydrogen-bonded end groups and their blends. 14149–14152 (2014).
dynamic-covalent boronic esters. Macromolecules 48, Macromolecules 41, 4694–4700 (2008). 186. Nakahata, M., Takashima, Y. & Harada, A. Highly
2098–2106 (2015). 162. Kang, J. H. et al. Tough and water-insensitive flexible, tough, and self-healing supramolecular
136. Guo, R. et al. Facile access to multisensitive and self- self-healing elastomer for robust electronic skin. polymeric materials using host–guest interaction.
healing hydrogels with reversible and dynamic boronic Adv. Mater. 30, 1706846 (2018). Macromol. Rapid Commun. 37, 86–92 (2016).
ester and disulfide linkages. Biomacromolecules 18, 163. Phadke, A. et al. Rapid self-healing hydrogels. 187. Burdick, J. A. & Prestwich, G. D. Hyaluronic acid
1356–1364 (2017). Proc. Natl Acad. Sci. USA 109, 4383–4388 (2012). hydrogels for biomedical applications. Adv. Mater. 23,
137. Cromwell, O. R., Chung, J. & Guan, Z. Malleable 164. Jeon, I., Cui, J. X., Illeperuma, W. R. K., Aizenberg, J. & H41–H56 (2011).
and self-healing covalent polymer networks through Vlassak, J. J. Extremely stretchable and fast self-healing 188. Highley, C. B., Rodell, C. B. & Burdick, J. A.
tunable dynamic boronic ester bonds. J. Am. Chem. hydrogels. Adv. Mater. 28, 4678–4683 (2016). Direct 3D printing of shear-thinning hydrogels into
Soc. 137, 6492–6495 (2015). 165. Willocq, B. et al. Mechanistic insights on spontaneous self-healing hydrogels. Adv. Mater. 27, 5075–5079
138. Deng, C. C., Brooks, W. L. A., Abboud, K. A. & moisture-driven healing of urea-based polyurethanes. (2015).
Sumerlin, B. S. Boronic acid-based hydrogels undergo ACS Appl. Mater. Interfaces 11, 46176–46182 189. Janeček, E. R. et al. Hybrid supramolecular and
self-healing at neutral and acidic pH. ACS Macro Lett. (2019). colloidal hydrogels that bridge multiple length scales.
4, 220–224 (2015). 166. Heller, M. & Schubert, U. S. Polystyrene with Angew. Chem. Int. Ed. 54, 5383–5388 (2015).
139. Smithmyer, M. E. et al. Self-healing boronic acid-based pendant mixed functional ruthenium(II)-terpyridine 190. Matson, J. B. & Stupp, S. I. Self-assembling peptide
hydrogels for 3D co-cultures. ACS Macro Lett. 7, complexes. Macromol. Rapid Commun. 23, 411–415 scaffolds for regenerative medicine. Chem. Commun.
1105–1110 (2018). (2002). 48, 26–33 (2012).

582 | August 2020 | volume 5 www.nature.com/natrevmats


Reviews

191. Webber, M. J., Kessler, J. & Stupp, S. I. Emerging 211. Fortman, D. J., Brutman, J. P., Cramer, C. J., 232. Bovey, F. A. & Mirau, P. A. NMR of Polymers
peptide nanomedicine to regenerate tissues and Hillmyer, M. A. & Dichtel, W. R. Mechanically (Academic, 1996).
organs. J. Intern. Med. 267, 71–88 (2010). activated, catalyst-free polyhydroxyurethane vitrimers. 233. Blanc, F. et al. Dynamic nuclear polarization NMR
192. Liu, J. et al. Tough supramolecular polymer networks J. Am. Chem. Soc. 137, 14019–14022 (2015). spectroscopy allows high-throughput characterization
with extreme stretchability and fast room-temperature 212. Snyder, R. L., Fortman, D. J., De Hoe, G. X., of microporous organic polymers. J. Am. Chem. Soc.
self-healing. Adv. Mater. 29, 1605325 (2017). Hillmyer, M. A. & Dichtel, W. R. Reprocessable 135, 15290–15293 (2013).
193. Eisenberg, A. (ed.) Ions in Polymers (American acid-degradable polycarbonate vitrimers. 234. Casabianca, L. B., Shames, A. I., Panich, A. M.,
Chemical Society, 1980). Macromolecules 51, 389–397 (2018). Shenderova, O. & Frydman, L. Factors affecting DNP
194. Kalista, S. J. Jr & Ward, T. C. Thermal characteristics 213. Röttger, M. et al. High-performance vitrimers from NMR in polycrystalline diamond samples. J. Phys.
of the self-healing response in poly(ethylene-co- commodity thermoplastics through dioxaborolane Chem. C 115, 19041–19048 (2011).
methacrylic acid) copolymers. J. R. Soc. Interface 4, metathesis. Science 356, 62–65 (2017). 235. Cassidy, M. C., Ramanathan, C., Cory, D. G., Ager, J. W.
405–411 (2007). 214. Chen, Q. et al. Durable liquid-crystalline vitrimer & Marcus, C. M. Radical-free dynamic nuclear
195. Kalista, S. J. Jr, Ward, T. C. & Oyetunji, Z. Self-healing actuators. Chem. Sci. 10, 3025–3030 (2019). polarization using electronic defects in silicon. Phys.
of poly(ethylene-co-methacrylic acid) copolymers 215. Chen, Q. et al. Multi-stimuli responsive and Rev. B 87, 161306 (2013).
following projectile puncture. Mech. Adv. Mater. multi-functional oligoaniline-modified vitrimers. 236. Verberg, R., Dale, A. T., Kumar, P., Alexeev, A.
Struct. 14, 391–397 (2007). Chem. Sci. 8, 724–733 (2017). & Balazs, A. C. Healing substrates with mobile,
196. Huang, Y., Lawrence, P. G. & Lapitsky, Y. Self-assembly 216. Yang, Y., Pei, Z., Li, Z., Wei, Y. & Ji, Y. Making particle-filled microcapsules: designing a ‘repair and
of stiff, adhesive and self-healing gels from common and remaking dynamic 3D structures by shining go’ system. J. R. Soc. Interface 4, 349–357 (2006).
polyelectrolytes. Langmuir 30, 7771–7777 (2014). light on flat liquid crystalline vitrimer films without 237. Ponnusami, S. A., Krishnasamy, J., Turteltaub, S. &
197. Reisch, A. et al. On the benefits of rubbing salt a mold. J. Am. Chem. Soc. 138, 2118–2121 van der Zwaag, S. A cohesive-zone crack healing
in the cut: Self-healing of saloplastic PAA/PAH (2016). model for self-healing materials. Int. J. Solids Struct.
compact polyelectrolyte complexes. Adv. Mater. 26, 217. Denissen, W., Winne, J. M. & Du Prez, F. E. Vitrimers: 134, 249–263 (2018).
2547–2551 (2014). permanent organic networks with glass-like fluidity. 238. Tiwary, P. & Parrinello, M. From metadynamics to
198. Bin Ihsan, A. et al. Self-healing behaviors of tough Chem. Sci. 7, 30–38 (2016). dynamics. Phys. Rev. Lett. 111, 230602 (2013).
polyampholyte hydrogels. Macromolecules 49, 218. Yang, Y. & Urban, M. W. in Healable Polymer Systems 239. Valsson, O., Tiwary, P. & Parrinello, M. Enhancing
4245–4252 (2016). (eds Hayes, W. & Greenland, B. W.) 126–148 (Royal important fluctuations: rare events and metadynamics
199. Lopez-Perez, P. M. et al. Self-healing hydrogels Society of Chemistry, 2013). from a conceptual viewpoint. Annu. Rev. Phys. Chem.
formed by complexation between calcium ions and 219. Flory, P.-J. Statistical thermodynamics of semi-flexible 67, 159–184 (2016).
bisphosphonate-functionalized star-shaped polymers. chain molecules. Proc. R. Soc. A 234, 60–73 (1956). 240. Bochicchio, D. & Pavan, G. M. Molecular modelling of
Macromolecules 50, 8698–8706 (2017). 220. Adamson, A. W. & Gast, A. P. Physical chemistry of supramolecular polymers. Adv. Phys. X 3, 1436408
200. Cao, Y. et al. A transparent, self-healing, highly surfaces Vol. 15 (Interscience, 1967). (2018).
stretchable ionic conductor. Adv. Mater. 29, 1605099 221. Hornat, C. C. et al. Quantitative predictions of 241. Lu, C. & Urban, M. W. Stimuli-responsive polymer
(2017). maximum strain storage in shape memory polymers nano-science: shape anisotropy, responsiveness,
201. Das, A. et al. Ionic modification turns commercial (SMP). Polymer 186, 122006 (2020). applications. Prog. Polym. Sci. 78, 24–46 (2018).
rubber into a self-healing material. ACS Appl. Mater. 222. Rodriguez, E. D., Luo, X. & Mather, P. T. Linear/ 242. Liu, F. & Urban, M. W. New thermal transitions in
Interfaces 7, 20623–20630 (2015). network poly(ε-caprolactone) blends exhibiting shape stimuli-responsive copolymer films. Macromolecules
202. Mei, J.-F. et al. A highly stretchable and autonomous memory assisted self-healing (SMASH). ACS Appl. 42, 2161–2167 (2009).
self-healing polymer based on combination of Mater. Interfaces 3, 152–161 (2011). 243. Jud, K. & Kausch, H. H. Load transfer through chain
Pt···Pt and π–π interactions. Macromol. Rapid 223. Liu, F., Jarrett, W. L. & Urban, M. W. Glass (Tg) molecules after interpenetration at interfaces.
Commun. 37, 1667–1675 (2016). and stimuli-responsive (TSR) transitions in random Polym. Bull. 1, 697–707 (1979).
203. Vaiyapuri, R., Greenland, B. W., Colquhoun, H. M., copolymers. Macromolecules 43, 5330–5337 (2010). 244. Gross, M. & Jaenicke, R. Proteins under pressure:
Elliott, J. M. & Hayes, W. Molecular recognition 224. Liu, F., Jarrett, W. L. & Urban, M. W. Synergistic the influence of high hydrostatic pressure on structure,
between functionalized gold nanoparticles and temperature and pH effects on glass (Tg) and function and assembly of proteins and protein
healable, supramolecular polymer blends-a route to stimuli-responsive (TSR) transitions in poly(N-acryloyl- complexes. Eur. J. Biochem. 221, 617–630 (1994).
property enhancement. Polym. Chem. 4, 4902–4909 N′-propylpiperazine-co-2-ethoxyethyl methacrylate) 245. Hummer, G., Garde, S., García, A. E., Paulaitis, M. E. &
(2013). copolymers. Polym. Chem. 2, 963–969 (2011). Pratt, L. R. The pressure dependence of hydrophobic
204. Burattini, S. et al. A supramolecular polymer based 225. Priestley, R. D., Ellison, C. J., Broadbelt, L. J. & interactions is consistent with the observed pressure
on tweezer-type π–π stacking interactions: molecular Torkelson, J. M. Structural relaxation of polymer denaturation of proteins. Proc. Natl Acad. Sci. USA
design for healability and enhanced toughness. glasses at surfaces, interfaces, and in between. 95, 1552–1555 (1998).
Chem. Mater. 23, 6–8 (2011). Science 309, 456–459 (2005).
205. Qin, J. et al. Tuning self-healing properties of stiff, 226. O’Connell, P. A. & McKenna, G. B. Rheological Acknowledgements
ion-conductive polymers. J. Mater. Chem. A 7, measurements of the thermoviscoelastic response of This work was supported by the National Science Foundation
6773–6783 (2019). ultrathin polymer films. Science 307, 1760–1763 under awards DMR 1744306 and partial OIA-1655740. The
206. Hentschel, J., Kushner, A. M., Ziller, J. & Guan, Z. (2005). J.E. Sirrine Foundation Endowment at Clemson University is
Self-healing supramolecular block copolymers. 227. Rabinowitz, J. D. & White, E. Autophagy and also acknowledged for partial support of this work.
Angew. Chem. Int. Ed. 51, 10561–10565 (2012). metabolism. Science 330, 1344–1348 (2010).
207. Montarnal, D., Capelot, M., Tournilhac, F. & Leibler, L. 228. Otts, D. B., Zhang, P. & Urban, M. W. High fidelity Author contributions
Silica-like malleable materials from permanent organic surface chemical imaging at 1000 nm levels: internal S.W. wrote and edited the article. M.W.U. conceptualized,
networks. Science 334, 965–968 (2011). reflection IR imaging (IRIRI) approach. Langmuir 18, wrote and edited the article.
208. Denissen, W. et al. Vinylogous urethane vitrimers. 6473–6477 (2002).
Adv. Funct. Mater. 25, 2451–2457 (2015). 229. Urban, M. W. Vibrational Spectroscopy of Molecules Competing interests
209. Denissen, W. et al. Chemical control of the viscoelastic and Macromolecules on Surfaces (Wiley, 1993). The authors declare no competing interests.
properties of vinylogous urethane vitrimers. 230. Hinderberger, D. in EPR Spectroscopy: Applications in
Nat. Commun. 8, 14857 (2017). Chemistry and Biology (eds. Drescher, M. & Jeschke, G.) Publisher’s note
210. Demongeot, A., Mougnier, S. J., Okada, S., 67–89 (Springer, 2011). Springer Nature remains neutral with regard to jurisdictional
Soulié-Ziakovic, C. & Tournilhac, F. Coordination 231. Schmidt-Rohr, K. & Spiess, H. W. Multidimensional claims in published maps and institutional affiliations.
and catalysis of Zn2+ in epoxy-based vitrimers. Solid-State NMR and Polymers Chs 3–5 (Academic,
Polym. Chem. 7, 4486–4493 (2016). 2012). © Springer Nature Limited 2020

NAture Reviews | MaterialS volume 5 | August 2020 | 583

You might also like