Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

157

Self-Healing Materials for Environmental Applications

5.1 Biomimetic Self-Healing Materials


Biological materials from plants and animals always exhibit their optimized
functional systems during the process of evolution. One of the most amazing
performances is self-healing capability of restoring health and soundness of
a system and regenerating the tissue structure and functions when damaged
by external forces. In all plants and animals, firstly, a self-sealing phase and,
secondly, a self-healing phase can be identified. The rapid self-sealing prevents
from infection by germs and gives time for the subsequent self-healing of
the injury, resulting in wound closure and the organ restoration. The scale of
biological self-healing can range from molecule level such as the restore of DNA
to macroscopic level like the union of the fractured bones. In addition, for the
biological materials, they can heal the damaged surface functional components
by re-secretion of relevant chemical substances.
Biological materials provide a great inspiration for scientists and engineers to
make bio-inspired artificial materials that can heal themselves when damaged
[1]. Recently, inspired by inherent regeneration capability of biological materials
under external damage, more and more artificial materials have been designed
by integrating self-healing properties with mechanical/electrical/chemical
functions, which are beneficial for long-term use and improving maintenance,
reliability, and durability in their practical application. The introduction of
self-healing ability can effectively cut down the replacement cost of some easily
worn materials with high price by healing the materials instead of replacing them.
Self-healing materials avoid the original degradation through the initiation of a
repair mechanism in response to the micro-damage. It is thus undisputed that
self-healing capability provides the man-made materials enormous possibilities,
in particular for applications where long-term reliability in poorly accessible
areas is important.
In such a concept, a wide range of intelligent self-healing materials with mul-
tifunctions has been developed and explored in diverse scientific areas such as
biomedical materials, smart wearable electric devices, environmental applica-
tion, etc. [1–6].

Artificially Intelligent Nanomaterials for Environmental Engineering, First Edition.


Peng Wang, Jian Chang, and Lianbin Zhang.
© 2019 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2019 by Wiley-VCH Verlag GmbH & Co. KGaA.
158 5 Self-Healing Materials for Environmental Applications

5.2 Overview of Self-Healing Materials


Early in ancient time, the Romans used a form of lime mortar that was found
to possess self-healing performance. Self-healing materials can be traced in the
1970s. A hard elastic polypropylene was observed to be capable of healing inter-
lamellar microvoids, which were formed after stretching in the perpendicular
direction [7, 8]. However, self-healing materials were widely recognized and stud-
ied in the scientific area until the twenty-first century. The initial overview of
self-healing materials gradually covered the entire spectrum of materials from
polymers to metals and ceramics.
In general, self-healing materials could be widely defined as an artificially or
synthetically created intelligent substance, which can spontaneously repair dam-
age by themselves without any external diagnosis of the problem or human inter-
vention. Hence the occurrence of damage can be also seen as a trigger to initiate
the generation of healing agents toward damage sites, leading to the directed mass
transport for filling the defect and the subsequent mending reaction by physical
interaction or chemical bonding. In the self-healing process, the healing agents
filled into damage sites and gradually became stable and immobilized to guaran-
tee perfect recover of original properties with the healed structures or chemical
composition.
Resembling biological systems, an ideal self-healing is automatically triggered
only by damage. Namely, the damage itself can be applied as a stimulus for the
self-healing process. It is mainly based on the intrinsic structure and compo-
sition of materials with the ability of detecting the damage and synchronous
self-healing. However, in most of the self-healing systems, external stimuli other
than the trigger of damage are generally applied to initiate the movement of
the healing agent toward damage sites. The typical external stimuli include
heat [9–12], pH [13–15], light [16–22], redox [23, 24], electrical field [25, 26],
magnetic field [27], etc.
In terms of self-healing materials, although inorganic materials such as metals
and ceramics have been also used in the construction industry as a healing agent
to heal physically structural cracks of materials [28], their diffusive mass transport
toward damage sites is always activated by high temperatures (>600 ∘ C for metals
and >800 ∘ C for ceramics) [1], due to the fact that their solute atoms as heal-
ing agents are relatively tiny and possess a relatively low mobility at the prevail-
ing operating temperatures, which limits their application breadth. In contrast,
polymer chains can be more readily mobile for self-healing at low temperature
(<120 ∘ C) [1]. Nowadays, self-healing polymeric materials attract much more
attention than inorganic ones due to their lightweight, wide availability, flexibility,
easy processing and manipulation, etc. [29]. This is especially true in environmen-
tal processes. Hence, in this chapter, we mainly discuss the polymeric self-healing
materials in environmental applications, as well as their chemical mechanism and
self-healing modes behind.
5.3 Extrinsic and Intrinsic Self-Healing Materials 159

5.3 Extrinsic and Intrinsic Self-Healing Materials


To be more convenient to discuss the design and fundamental mechanism of
self-healing materials, the self-healing materials are categorized into two types:
extrinsic and intrinsic self-healing materials [1, 30, 31], depending on the source
of self-healing agents.

5.3.1 Extrinsic Self-Healing Materials


Extrinsic self-healing materials do not have self-healing capability in the matrix.
They have external healing agents embedded across the entire host matrix mate-
rials in the form of micro/nanocapsules or vascular networks [1, 32, 33].
When the damage occurs, the embedded capsules or vascular networks break
to release the healing agents, then heal the matrix materials based on the healing
agents’ polymerization reactions, and prevent damage propagation.
In general, encapsulation of healing agents could be carried out by various
methods including in situ polymerization, interfacial polymerization, internal
phase separation, and solvent evaporation technique [34]. Typically, various
capsule shells could be synthesized through abovementioned techniques, such
as urea–formaldehyde (UF) resin, melamine formaldehyde (MF), polyurea
(PU), poly(methyl methacrylate) (PMMA), etc. [35–40]. Correspondingly, dicy-
clopentadiene (DCPD), polydimethylsiloxane (PDMS), glycidyl methacrylate,
epoxy resins, isocyanates, linseed oil, aliphatic amines, and so forth are typically
selected as healing agents and microencapsulated [38, 41–43]. Among them,
some kinds of isocyanates, such as isophorone diisocyanate (IPDI), hexam-
ethylene diisocyanate (HDI), polyaryl polymethylene isocyanates (PAPI), etc.
[44–48], can react with moisture and are being used as potential healing agents
to create catalyst-free self-healing materials in moist or aqueous environments.
In 2001, White et al. firstly reported extrinsically structural self-healing mate-
rials that could automatically heal cracks by incorporating a microencapsulated
healing agent (Figure 5.1a) [32]. The microencapsulated self-healing agents
(DCPD monomer) could be released after damage, and the released agents could
intrude into the crack plane through capillary action and then contact with
an embedded Ruthenium-based Grubbs’ catalyst in the matrix to trigger the
ring-opening metathesis polymerization and heal the crack faces.
Although the capsule-based healing concept has been commonly employed
owing to simple fabrication, the inherent drawback of microcapsule-based
self-healing materials restricts their practical application, stemming from the
local healing agent’s depletion after a single damage event [33].
Based on this challenge, Sottos and coworkers (Figure 5.1b) further developed
the bio-inspired vascular self-healing systems [49], where a three-dimensional
interpenetrating microvascular network was embedded into the substrate.
When physical damages occur, the healing agent inside networks could be
160 5 Self-Healing Materials for Environmental Applications

Cut

Epidermis
Capillaries

sels
Dermis d ves
r bloo
Large

(b)

(a) (c)

Figure 5.1 (a) Capsule-based healing concept. A microencapsulated healing agent is


embedded in a composite matrix containing a catalyst capable of polymerizing the healing
agent. Source: White et al. 2001 [32]. Reprinted with permission of Nature Publishing Group.
(b) Schematic illustration of a cut in the epidermis layer of the skin and the capillary network in
the dermis layer. (c) Schematic view of a vascular self-healing system, consisting of an
interpenetrating microvascular network that supplies two fluids (red and blue) to a crack
plane, where mixing occurs (purple). Source: Hansenet al. 2009 [49]. Reprinted with permission
of John Wiley and Sons.

continuously delivered to cracks via the microvascular networks and heal the
damage (Figure 5.1c). The existence of such microvascular networks guarantees
the large-volume and repeatable self-healing of mechanical damage on the
same damage region. Meanwhile, such extrinsic self-healing materials mostly
exhibit autonomic self-healing without any trigger, which the damage itself can
stimulate the process of self-healing.
Even though large-volume self-healing can be realized, some disadvantages also
exist, including relatively complicated design and fabrication process, as well as
slow and limited-time healing since the amount of healing agents is depleted in
the healed region. Moreover, the damage regions that are not lying in the path of
microcapsules would not be healed.

5.3.2 Intrinsic Self-Healing Materials


In contrast, intrinsic self-healing materials do not need an additional healing
agent, and the molecules that constitute the matrix can act as the healing agent
by themselves [1, 10].
The healing process involves polymer chain internal diffusion or/and external
migration to form reversible covalent bonds and noncovalent bonds, depending
on system design [50–53].
5.3 Extrinsic and Intrinsic Self-Healing Materials 161

Bond reformation
Network repair
Physical damage

Segmental mobility and


conformational changes bring
reactive chain ends to contact
Diffusion of low molecular
weight segments
Physical damage:
Chain cleavage
Chain slippage
Result: Reactive chain ends
Undamaged network

Evolution of damage and repair

Figure 5.2 Schematic illustration of the evolution of an ideal damage-repair process in


polymeric materials. Source: Yang and Urban 2013 [54]. Reprinted with permission of Royal
Society of Chemistry.

Upon structural damage, cleavage of polymer chains leads to the release of


reactive groups (either pendent groups or cleaved chain ends), such as free rad-
icals and/or —C—C—, —COOH, —NH2 , —OH, —Si—O, SH/S—S, or —C—O,
which facilitate re-bonding to realize self-healing [54]. As depicted in Figure 5.2,
mechanical damage could cause polymer network cleavage and/or slippage. Sub-
sequently, the resultant migration and/or diffusion of polymer chains leads their
reactive groups in contact with each other to heal the wound by bond refor-
mation and physical network repair. Additionally, for special wetting surfaces,
self-healing stems from the free-energy-driven self-migration of hydrophobic or
hydrophilic polymer chains to heal the damaged surface chemical component.
Generally, most covalent bonds are irreversible, and hence conventional
covalent polymers are permanently damaged by the formation and propaga-
tion of cracks. However, reversible covalent linkages have been exploited to
realize self-healing by reforming new bonds toward the damaged site [55].
Reversible covalent bonds are based on reversible chemical reaction, such as
Diels–Alder (DA) reaction, disulfide bonds, reversible radical reactions, and
photocycloaddition [10, 56–61]. Due to their high bonding strength, reversible
covalent polymers mainly are applicable for the majority of polymeric structural
engineering materials [62].
Additionally, reversible noncovalent bonds in polymers can also be exploited
for self-healing performance [54], consisting of hydrogen bond [63, 64],
metal–ligand bonding [65–67], ionic interaction [68], host–guest interaction
[69], π–π interaction [70], and so on. The majority of reversible covalent
cross-linking requires relatively complex treatment, like thermal or light treat-
ment, to trigger the reaction. In contrast to the covalent interaction, materials
associated by noncovalent bonds can achieve the self-healing process under more
mild conditions with a higher spatial density of bonding [1, 69, 71]. Furthermore,
since noncovalent bond interactions have relatively lower strength as compared
162 5 Self-Healing Materials for Environmental Applications

with the covalent bonds, they are highly reversible and easily accessible to be
incorporated into polymer structures to realize self-healing. Nevertheless, the
ultimate and ideal goal is to develop materials with the capability of autonomous
self-healing.
As a consequence, intrinsic self-healing mode further avoids some issues
of integration and healing-agent compatibility that arise in vascular and
capsule-based self-healing materials. Meanwhile, intrinsic self-healing materials
can achieve multi-time reversible healing through inherent reversibility of
molecular interactions of the matrix polymers. Moreover, intrinsic self-healing
systems are able to heal tiny damages even at the molecular scale. In contrast,
capsule- or vascular-based healing systems are only applicable to heal large
damage volumes.
Hence, such intrinsic self-healing materials can achieve high flexibility, fast
healing, biocompatibility, and incorporation of multifunctions. Due to these
advantages, more attention has been held on the intrinsic self-healing materials
with specific functions suitable for environmental applications than the extrinsic
self-healing materials.

5.4 Self-Healing Materials in Environmental


Applications
Overall, triggered by damage or external stimuli, healing agent from self-healing
systems can be initiated and move toward damage sites via extrinsic or intrinsic
self-healing approach to achieve self-healing process (Figure 5.3). The damages
of materials generally include physical crack and surface function loss. The latter
one may be owing to the removal of surface chemically functional components
or due to surface contamination [30, 72, 73].
Accordingly, the following discussions are organized into three parts (i)
self-healing of physical cracks, (ii) self-restoring of surface chemical compo-
nents, and (iii) self-cleaning of contaminated surfaces. The following sections
present and discuss the examples of self-healing materials, which have been
preliminarily applied into the environmental field in the above categories. As the
concept of self-healing has slowly but steadily been appreciated by and applied
to environmental fields in the past decade, some exploratory works on endowing
conventional environmental processes with self-healing materials have been

Extrinsic Damage
Physical
self-healing itselt
Self- crack
materials External Damage
healing Triggers Surface
Intrinsic stimuli sites
agents self-healing (heat, light, function
...) loss
materials

Figure 5.3 Classification of self-healing materials.


5.4 Self-Healing Materials in Environmental Applications 163

conducted, including, but not limited to, water filtration [74–76], oil/water
separation [77], pollutant absorption [78], antifouling surface, etc. [79–81].

5.4.1 Self-Healing of Physical Cracks


During the installation and operation of materials and devices, accidents and
wear and tear often lead to cracked damages, which are detrimental to their
designed performances. Endowing these materials and devices with self-healing
capability would prolong their lifetime and reduce operational cost. The
self-healing of physical crack is a broad topic, and this part of the chapter focuses
only on the recent development of the application of self-healing materials to
environmental processes.
The application of self-healing materials that can heal physical cracks in
environmental processes is relatively new. In 2012, Tyagi et al. reported a porous
membrane by assembling micelles of triblock copolymer poly(styrene-co-
acrylonitrile)-b-poly(ethylene oxide)-b-poly(styrene-co-acrylonitrile) (PSAN-b-
PEO-b-PSAN) for water filtration (Figure 5.4) [74]. The evaporation of the
copolymer solvent results in the formation of a porous film through the dynamic
self-assembly of micelles with the diameter of 50 nm. The membrane material
was not cross-linked and therefore could self-heal its macroscopic structural
defects under an applied pressure of 0.8 bar by the rearrangement of micelles
and the reformation of block copolymer bridges on the damage site.
Furthermore, in 2016, Kim et al. fabricated a self-healing filtration membrane
by embedding the extrinsic microcapsules composed of PU shell and IPDI core
into PES membrane [75]. Once damaged, the healing agent of isocyanate was
released from the broken microcapsules and diffused toward the crack sites and
then reacted with the surrounding water to form PU plug to cover the dam-
age sites (Figure 5.5a). Once healed, the water flux and particle rejection perfor-
mances of the membrane were recovered to 103% and 90% of the original ones,
respectively.
Nevertheless, the microcapsule-type extrinsic self-healing agent results in a
limited number of healing cycles at a given region. Kim and coworkers further
developed a water filtration PES membrane with membrane pores filled with
poly(2-acrylamido-2-methyl-1-propane sulfonic acid) (PAMPS) hydrogel [76].
The self-healing capability of the membrane was majorly attributed to the
swelling effect of the pore-filling hydrogel at the damage site and the strong
hydrogen bonding and molecular interdiffusion of the hydrogel polymer chains
(Figure 5.5b). As a consequence, the particle rejection efficiency of the healed
membrane restored to 99%, while the efficiency of the damaged membrane
was as low as 30%. In 2017, the same group further developed an improved
in situ healing method by using branched polyethylenimine-functionalized
silica microparticles [82]. The in situ healing recovered the membrane’s particle
rejection to 99.1% of the original one without any flux decline.
In 2013, Lu and coworkers fabricated a self-healing and pollutant-adsorbing
hydrogel by using polydopamine (PDA)-modified clay as the main building
block and Fe3+ ions as the physical cross-linkers (Figure 5.6a) [78]. Upon the
removal of applied large stress, this hydrogel could be self-healed via reformation
164 5 Self-Healing Materials for Environmental Applications

COOH

O
O O
N O O
y x m O n x y mO N
EtO P CN
CN O P OEt
OEt
COOH OEt
(a)

Solvent concentration

(b) Nanoporosity

S–4800 ×50.0k
600 nm
(c)

Figure 5.4 (a) Chemical structure of the triblock copolymer. (b) Sketch of the membrane
formation: during the course of the solvent evaporation, the increase of the block copolymer
concentration triggers their self-assembly into micelles, which assemble in three dimensions
forming a dynamic interactive membrane film. (c) SEM image of the film. Source: Tyagi et al.
2012 [74]. Reprinted with permission of John Wiley and Sons.

of damaged catechol–Fe3+ complexes. Moreover, working as an adsorbent, it


effectively removed an organic pollutant Rhodamine 6G (Rh6G) from water
with an adsorption capacity of 150 mg g−1 owing to the hydrogen-bonding and
π–π stacking interactions between the aromatic moieties of PDA and Rh6G.
As demonstrated in Figure 5.6b, after dye-containing water was slowly poured
through a hydrogel-filled syringe, the Rh6G was trapped in the hydrogel, while
5.4 Self-Healing Materials in Environmental Applications 165

Polyurethane Healed site


microcapsules

Membrane

Broken
microcapsules

(a) Nonwoven fabric

Pristine membrane Damaged membrane Healed membrane

Hydrogel
expansion

Molecular
interdiffusion
O O
S OH
N
O
Pore-filling Support Hydrogen H H
n

O
hydrogel substrate bonding N
HO S
n

O O
(b)

Figure 5.5 (a) Schematic illustration of the self-healing process of microcapsule-embedded


membranes. Source: Kim et al. 2016 [75]. Reprinted with permission of American Chemical
Society. (b) Schematic illustration of self-healing pore-filled membranes with the pore-filling
hydrogel anchored on PES membrane that acted as the active layer by allowing water to pass
through and repelling unwanted particles (red spheres). Source: Getachew et al. 2017 [76].
Reprinted with permission of American Chemical Society.

D-clay hydrogel Fe3+ as cross-linker for D-clay

Water
with
Rh6G
Fe3+

Clay Rh6G
removed

A thin layer of
PDOPA coating
(a) (b)

Figure 5.6 (a) Schematic illustration of the formation mechanism for the hydrogels.
(b) Photograph showing the removal of Rh6G by the hydrogel. Source: Huang et al. 2013 [78].
Reprinted with permission of American Chemical Society.
166 5 Self-Healing Materials for Environmental Applications

the clear water could pass through the hydrogel and could be observed at the
end of the syringe, indicating the complete removal of the Rh6G dye.
Except that, in 2014, Yan et al. described a self-healing supramolecular gel that
was prepared from tetrazolyl derivatives and alkylamine via hydrogen-bonding
self-assembly [83]. This gel could selectively congeal crude oil from an oil/water
mixture and displayed remarkable self-healing properties. It is noted that hydro-
gels with self-healing properties provide a great promise in biological tissues and
environmental applications owing to their hydrated nature and biocompatibility.
In 2015, Sun and coworkers fabricated self-healing and antifouling films via
layer-by-layer (LbL) assembly of PEGylated branched poly(ethylenimine) (bPEI)
and hyaluronic acid (HA). The damaged structure and antifouling function of
the films could be healed rapidly in water due to the high mobility of polyelec-
trolytes and the resultantly rapid reformation of reversible electrostatic and
hydrogen-bonding interactions [84].
Similarly, in the same year, Zeng and coworkers reported a mussel-inspired
injectable hydrogel with self-healing and anti-biofouling capabilities, which
was based on self-assembly of triblock copolymer consisting of PNIPAM and
poly(ethylene oxide) (PEO) [85]. The obtained hydrogel remained its self-healing
ability with experiencing repeated structural damage through catechol-mediated
hydrogen-bonding interactions and aromatic interactions (Figure 5.7) and also

Hydrogen bonding
H
OH H O
O

O H HO
H H
O
OH

O HO
H
OH
HO
O
H
HO
H O
H

Aromatic interaction

HO OH OH
HO
HO OH
HO
HO
PEO

Cross-link with catechol groups (red tetragon)


trapped into the PNIPAM block (blue dot)
Figure 5.7 Schematic illustration of the chemical structure of the triblock copolymeric
hydrogel. Source: Li et al. 2015 [85]. Reprinted with permission of John Wiley and Sons.
5.4 Self-Healing Materials in Environmental Applications 167

effectively prevented nonspecific cell attachment owing to the presence of PEO


with strong antifouling property.
The membrane technology in CO2 separation and capture is significantly
beneficial to address the risk in the greenhouse gas emission, and as a novel
research direction, this technology has more advantages than traditional method
owing to its less energy consumption and satisfying ecofriendly needs. However,
in practical use, even subtle crack in the membrane would lead to entirely
losing the original functions. Hence in 2018, Cao et al. reported a self-healable
polymeric elastomer with recoverable gas-separation performance, which
was synthesized by reacting urea with amine-terminated PDMS elastomers
[86]. The obtained elastomers exhibited an extremely high elongation at break
(984–5600%) with adjustable mechanical strength, elasticity, and extensibility.
After mechanical damage, mechanical strength and gas-separation performance

Applied
scratch After
healing

(a)

5 × 103
Original
Permeability (barrier)

4 × 103
After healing
3
3 × 10

2 × 103

1 × 103

15
Selectivity/α (CO2/N2)

10

0
(b) U-PDMS0.9K-E U-PDMS3.0K-E U-PDMS5.0K-E U-PDMS30K-E

Figure 5.8 (a) Images showing the gas-separation membranes with applied scratch and after
healing at 40 ∘ C for 30 minutes. (b) Comparative CO2 permeability (up) and CO2 /N2 selectivity
(bottom) of original polymeric membranes (blue) and polymeric membranes after cut and
healing (orange). Source: Caoet al. 2018 [86]. Reprinted with permission from of John Wiley
and Sons.
168 5 Self-Healing Materials for Environmental Applications

with CO2 permeability and CO2 /N2 selectivity can be healed in 120 minutes at
ambient temperature or in 30 minutes at 40 ∘ C (Figure 5.8).
In addition, the liquid-repellent surfaces have been widely investigated in fuel
transport, antifouling, and other potential environmental applications. Mean-
while, the introduction of the self-healing capability can enhance the service
life term and stability upon physical damage [87]. In 2011, Wong et al. reported
a self-healing slippery liquid-infused porous surface (SLIPS) with exceptional
liquid and ice repellency, pressure stability, and enhanced optical transparency
[88]. This surface was capable to repel various simple and complex liquids such
as water, hydrocarbons, crude oil, and blood, with low contact angle hysteresis
(CAH) (<2.5∘ ), and could resist ice adhesion and remain its functions at high
pressures (up to about 680 atm). Serving as an intelligent coating with self-healing
performance, it rapidly and repeatedly restored the liquid-repellent function
(within 0.1–1 second) upon large-area physical damage by recurring abrasion or
impact, since the fluidic nature of the lubricating layer forced the liquid to flow
toward the damaged area by surface-energy-driven capillary action (Figure 5.9).
Following this concept, in 2013, Vogel et al. employed colloidal templating to
design transparent, slippery liquid-infused nanoporous surface structures with
self-healing capability [89]. (1H,1H,2H,2H-Tridecafluorooctyl)-trichlorosilane
was locked into the porous network and formed a stable liquid film. The
obtained film exhibited efficient liquid repellency, prevention of adsorption of
liquid-borne contaminants, reduction of ice adhesion strength, and the unique
self-healing properties arising from the liquid nature of the lubricant.
In 2018, Heng and coworkers further developed an anisotropic slippery surface
using directional, porous, and conductive polymer (PCDTPT) films and silicone
oils with different viscosities, which combined with self-healing property and
electrically driven smart control of droplet motion [90]. As a result, the critical
self-healing thickness increased with increasing silicone oil viscosity. On the
other hand, the responsive voltage for controlling the droplet sliding motion
reduced with decreasing the viscosity of silicone oil.
Earthworm could pass through sticky soil without inducing stains due to a
self-recovery thick lubricating layer on their texture skins. Recently, inspired
by earthworm, Cui and coworkers created artificially intelligent coatings with
self-replenishing lubrication for adaptive friction reduction, wear resistance,
and antifouling property in a solid environment [91]. As shown in Figure 5.10a,
lubricants were embedded as discrete droplets into a supramolecular polymer
matrix made from the copolymer of urea and PDMS. Meanwhile, the poly-
meric surface was prepared through the solution casting method under humid
conditions where solvent (THF) evaporation would induce water condensation
on the forming polymer surface and thus rough structure formed directly
(Figure 5.10b). The obtained coatings exhibited self-regulated liquid release
since the lubricant stored in the droplets could be site specifically and quickly
released under external mechanical stimulus, thus to restore oil layer and heal
the slippery functional surface (Figure 5.10c).
Ice accretion poses a severe risk for human safety, and traditional methods
of ice removal using antifreeze fluids or heating, caused high energy con-
sumption and unavoidable environmental problems. SLIPS possess icephobic
Slips
t = 0 ms 1 cm Tilting = 5° Tilting = 5° Tilting = 5°
Damage
Physical Crude
oil
damage
5 mm
t=0s t=1s t=2s
Teflon AF treated flat surface
t = 150 ms Tilting = 5° Tilting = 10° Tilting > 10°
Self-healed 1 cm

Crude oil
Physical
damage Pinned
droplet Pinned
droplet
t=0s t=2s t = 17 s
(a) (b)

Figure 5.9 Self-healing and optical transparency of SLIPS. (a) Time-lapse images showing the self-healing capability of an SLIPS upon physical damage, 50 mm
wide on a timescale of the order of 100 ms. (b) Time-lapse images showing the restoration of liquid repellency of an SLIPS after physical damage, as compared
to a typical hydrophobic flat surface (coated with DuPont Teflon AF amorphous fluoropolymers) on which oil remains pinned at the damage site. Source:
Wong et al. 2011 [88]. Reprinted with permission of Springer Nature.
170 5 Self-Healing Materials for Environmental Applications

(i) Solvent evaporation in moist air (ii) Locking of liquid layer

uPDMS, silicone oil, THF Breath figure


phase separation
gelation
uPDMS: Liquid supplying regime

(iv) Rebuilding of liquid layer (iii) Consumption of surface liquid

Mechano-stimulus

Silicone oil Gel matrix


(a)

2h 5h 12 h

50 μm 50 μm 50 μm

(b)

Washed film
Pressed region

200 μm 200 μm

(c)

Figure 5.10 The fabrication and the liquid-release process of gel films. (a) Schematic
illustration of the synthetic strategy and stimuli-responsive release of the polymer coating.
(b) Top-view image of polymer coating formed at different times. (c) The oil release behavior of
a freshly water-washed sample under a local pressing. The thickness of the films in (b) and (c) is
about 400 μm. Source: Zhao et al. 2018 [91]. Reprinted with permission of John Wiley and Sons.

surface with low ice adhesion strengths below 15, 1.7, and 0.4 kPa in different
studies [92–94]. Short life spans of state-of-the-art icephobic surfaces still
need to address. In 2018, Zhang’s group integrated an interpenetrating poly-
mer network (IPN) into an autonomous self-healing elastomer, consisting of
Fe-pyridinedicarboxamide-containing PDMS (Fe-Py-PDMS) and commercial
PDMS (Sylgard 184) [95]. Fe-Py-PDMS, as a self-healing elastomer with low
modulus and surface energy, is cross-linked via metal–ligand coordination
bonds, while Sylgard 184 is cross-linked via covalent bonds (Figure 5.11). The
obtained material realized an ultralow ice adhesion strength of 6.0 ± 0.9 kPa and
remained at a low value of ∼12.2 kPa after 50 icing/deicing cycles. Meanwhile,
such material can be self-healed upon mechanical damage in 24–48 hours.
5.4 Self-Healing Materials in Environmental Applications 171

CH3 CH3
Si O x Si O
y
CH3 CH2
CH2 CH3
O Si y O Si x
CH3 CH3

Sylgard 184

O
CH3 CH3 CH3 N m
NH Si O Si O Si N O CH3 CH3 CH3
n Fe
CH3 CH3 CH3 O N Si O Si nO Si NH
m
N CH3 CH3 CH3
O
Fe-Py-PDMS

Figure 5.11 Scheme of the self-healing IPN elastomer consisting of Fe-Py-PDMS and Sylgard
184. Source: Zhuo et al. 2018 [95]. Reprinted with permission of American Chemical Society.

5.4.2 Self-Restoring of Surface Functional Components


The surface function, especially surface wettability, is usually achieved by
depositing an ultrathin functional layer (e.g. a single molecular layer) on a
substrate material, which makes the surface function being easily degraded
even without any appreciable structural cracks. Accordingly, developing durable
wetting surface materials with self-healing capability has become one of the most
intriguing topics in the area of surface science, especially involving self-healing
liquid-repellent surfaces and highly hydrophilic antifouling surfaces.

5.4.2.1 Chemical Mechanism


Particularly, the self-healing of surface wettability (e.g. hydrophobicity,
hydrophilicity) is largely based on the surface-free-energy-driven migration of
hydrophobic or hydrophilic polymer chains. In practice, surface hydrophobic
molecules can be gradually decomposed or removed by mechanical damage
or upon exposure to light irradiation and highly oxidative chemicals. With the
hydrophobic molecules gone from the surface, the surface energy would rise
in the air due to the exposed hydrophilic groups. Then the low-surface-energy
molecules underneath/near the damaged sites would migrate onto the outer
surface to reduce the surface energy and restore the original hydrophobicity
[96]. Alternatively, this process can also be explained by the fact that the air
is a very hydrophobic medium and the low-surface-energy molecules like to
make contact with the air according to the principle of “like dissolves like.”
Additionally, except the automatic process of self-healing, such motion and
172 5 Self-Healing Materials for Environmental Applications

migration of healing agent can be initiated or accelerated by external stimuli,


such as heating [97–99], moisture [100], solar irradiation [101–103], organic
solvent [104], etc.
Similarly, for the hydrophilic surfaces, after the hydrophilic polymer chains
were degraded or detached from the surfaces, the polymer chains stored inside
the matrix would spontaneously migrate to the damaged surface to form a new
hydrophilic polymer layer, which was driven by an emerging gradient in a chem-
ical potential or free energy [79]. As can be derived, for the above self-healing
mechanisms to work, a matrix to store the self-healing components is needed.

5.4.2.2 Hydrophobic Self-Healing


As a proof to this surface wettability self-healing concept, in 2010, Sun and
coworkers reported a self-healing superhydrophobic coating, which was fabri-
cated by chemical vapor deposition (CVD) of fluoroalkylsilane on the porous
LbL-assembled polyelectrolyte film [96]. As Figure 5.12 shows, the surface
superhydrophobicity, once lost, could be self-restored at room temperature by
means of self-migration of fluoroalkyl chains from the underneath polyelec-
trolyte film to the outer surface to minimize the interfacial free energy. The
moisture could accelerate the rate in self-healing since after decomposition of the
superhydrophobic layer, the surface became superhydrophilic and can absorb
water in a humid environment, which further softens the coating matrix and

1 2 3
Decompose Self-healing

H2
C CH
3
F F F F F O H2
F H2 Fluoroalkyl chain
Si O C
C C C C CH3
F C C C C in reacted POTS
H2 O CH
F F F F F F C 3

(a) H2
POTS

C O O
n

(b) SO3H

Figure 5.12 (a) Working principle of self-healing superhydrophobic coatings. (1) The porous
polymer coating with micro- and nanoscaled hierarchical structures can preserve an
abundance of healing agent units of reacted fluoroalkylsilane. (2) The top fluoroalkylsilane
layer is decomposed, and thus the coating loses its superhydrophobicity. (3) The healing
agents can migrate to the coating surface and recover superhydrophobicity. (b) Chemical
structure of sulfonated poly(ether ether ketone) (SPEEK). Source: Li et al. 2010 [96]. Reprinted
with permission of John Wiley and Sons.
5.4 Self-Healing Materials in Environmental Applications 173

drives the migration of fluoroalkyl chains. Also, such restoration process could
be repeated many times without decreasing the superhydrophobic property.
Even since, the self-healing superhydrophobic coatings and textiles have been
widely prepared from a variety of polymers, such as 1H,1H,2H,2H-perfluoro-
octyltriethoxysilane (POTS) [105], fluorinated-decyl polyhedral oligomeric
silsesquioxane (FD–POSS) [106–108], fluorinated alkyl silane (FAS) [102, 106,
109], perfluorooctanesulfonic acid (PFOS) lithium salt [105], perfluorooctyl acid
[110], poly(2,2,3,3,4,4,4-heptafluorobutylmethylsiloxane) (PMSF) [101], PDMS
[104], and octadecylamine (ODA) [100].
In 2015, Wang and coworkers further applied this superhydrophobicity
self-healing mechanism to a photothermal conversion membrane, which was
the first report on using polymeric photothermal materials, for the purpose of
solar-driven water evaporation and seawater desalination (Figure 5.13a). The
membrane was fabricated by deposition of light-to-heat conversion material
of polypyrrole onto a porous stainless steel mesh, followed by hydrophobic
fluoroalkylsilane modification [111]. As-prepared membrane can self-float at
the water/air interfaces to realize automated and uninterrupted interfacial solar
heating. By locally heating the interfacial water under light irradiation, the
PPy-based hydrophobic membrane rises the rate of water evaporation. More
importantly, it was revealed in this study that the migration and enrichment of
fluoroalkyl chains to the outer surface of the coating could be accelerated by solar
light irradiation and multiple cycles of self-healing were achieved (Figure 5.13b).
In 2016, Fang et al. demonstrated a self-healing electrospun N-perfluorooctyl-
substituted PU fibrous membrane for oil/water separation [77]. The membrane,
with O2 plasma treatment, lost its superhydrophobicity but could recover its
superhydrophobicity back with assistance by gentle heat treatment for more than
70 cycles. The oil/water separation efficiency maintained at above 98% after 20
cycles of wettability loss-and-restoration cycles since the fluorine-containing PU
could self-migrate to the outer surface of the fibers to restore the superhydropho-
bicity of the membrane. Accordingly, upon plasma-and-heating treatment, the

Point-of-use solar distillation device


160
Contact angle (°)

Solar-powered fan
120
Condensing
Plasma
treatment
irradiation

chamber 80
Light

Evaporation
chamber 40
0

PPy-coated mesh membrane


0 1 2 3 4 5
Healing cycles
(a) (b)

Figure 5.13 (a) Schematic configuration of the point-of-use device for direct and all-in-one
solar distillation. (b) Reversible water contact angle changes on the plasma-treated and
light-irradiated PPy-coated mesh. The insets in (b) were the shapes of the water droplets on
the surfaces after plasma treatment and light irradiation. Source: Zhang et al. 2015 [111].
Reprinted with permission of John Wiley and Sons.
174 5 Self-Healing Materials for Environmental Applications

6000 n-Hexane 100


Ether

Separation efficiency (%)


Permeate flux (l m–2 h–1) Petroleum ether
5000 Toluene 98 n-Hexane
Bromoethane
Petroleum ether
4000
96 Bromoethane
3000 Ether
94 Toluene
2000
1000 92

0 90
0 5 10 15 20 0 5 10 15 20
(a) Self-healing cycles (b) Self-healing cycles

Figure 5.14 Permeate flux (a) and separation efficiency (b) of the electrospun N-
perfluorooctyl-substituted PU fibrous membrane after 20 self-healing cycles. Source: Fang
et al. 2016 [77]. Reprinted with permission of John Wiley and Sons.

oil permeation flux experienced only a small decline for a variety of oils (7.8% for
n-hexane, 8.5% for petroleum ether, 8.7% for bromoethane, 8.1% for ether, and
8.2% for toluene) (Figure 5.14a,b).
In the same year, Liu and coworkers fabricated a self-healing anti-smudge
coating by modifying LbL-assembled poly(diallyldimethylammonium) (PDDA)
and poly(styrenesulfonate) (PSS) film with perfluorooctanoate (PFO) [81]. The
film can be deposited onto various commonly available substrates (e.g. Si, glass,
plastic, steel, and wood). The modified film on these substrates all showed
sliding angle <12∘ for a variety of low surface tension liquids of 10 μl and easily
self-restored its oil repellency upon lost due to the replenishing surface.
In 2017, in Guo’s group, superhydrophobic fabrics were prepared through an
in situ growth method for depositing MnO2 nanoparticles on cotton fabric sur-
face to construct a hierarchical structure and subsequent stearic acid modifica-
tion for lowing the surface energy [112]. The obtained fabrics can automatically
recover their superhydrophobicity by a short-time heat treatment, without obvi-
ous reduction on the water CA after 8 cycles of the repetitive etching–heating
cycles. Meanwhile, the self-healing superhydrophobic fabric exhibited excellent
mechanical stability, wear resistance, and high efficiency of oil/water separation.

5.4.2.3 Hydrophilic Self-Healing


Bio-inspired by fish scales, artificial underwater surface materials with
hydrophilicity enables to repel oils or organic liquids only when submerged in
water. This kind of functional surface is commonly utilized as antifouling mate-
rial operating in seawater or resisting the undesirable adhesion of protein, cells,
and microorganisms [113–116]. Undoubtedly, physical rubbing or damage in an
underwater environment could cause the loss of superoleophobic or antifouling
properties of the functional surface, accordingly recalling the introduction of
self-healing design for long-lasting use and operation.
In 2013, Minko and his coworkers grafted PEO both on the surface and inside
of P2VP polymer 3D network [79]. When the surface was damaged underwater,
the grafted PEO polymers residing in a polymer network underwent a spon-
taneous rearrangement and migration to replenish the lost PEO chains at the
5.4 Self-Healing Materials in Environmental Applications 175

Surface grafting
PEO brush
Fouling
Network
Degradation

Substrate

(a)

Self–healing
antifouling surface

PEO brush 3D grafting

Degradation

Network
with the grafted PEO

Replacement
(b) of the detached chains

Figure 5.15 Schematic illustration of the PEO-grafted P2VP network films. (a) The grafting of
PEO to the surface of a P2VP film and (b) 3D polymer grafting on the surface and inside of a
P2VP film. The self-healing aspect of the antifouling property is due to the rearrangement of
internally grafted polymers to the interface (marked as dark blue chains). Source: Kuroki et al.
2013 [79]. Reprinted with permission of John Wiley and Sons.

water–network interface (Figure 5.15). As a result, the antifouling performance of


the material showed a fourfold increase as compared with the traditional antifoul-
ing materials in physiological conditions (pH 7.4 and 37 ∘ C). Furthermore, owing
to the existence of the P2VP block, PEO-grafted P2VP films also demonstrated
the high stability of pH-responsive performance during the healing process.
Based on the same concept, in 2015, Wu and coworkers fabricated a similar
material by grafting self-assembled hydrophilic copolymeric chains on the
hierarchical microgel spheres. The prepared material achieved the capability of
self-restoring antifouling properties once its surface is mechanically damaged
(Figure 5.16a) [80]. Due to the swelling of the microgel spheres and movement of
176 5 Self-Healing Materials for Environmental Applications

Water
: MHMS : Thermoset acrylate resins and t-HDI Oil

Underwater
Abrading or superoleophobic

crushing Self-healing Protein

Water
Anti-biofouling
(a)
Surface
Press Retract
160
OCA(°)

(c) Water
155
Surface
150 Press Retract
Oil repellent
Oil adhesive
0 1 2 3 4 5 6
(b) (d) Water
Crushing-repairing cycles

Figure 5.16 (a) Schematic illustration of the structure and self-healing process of underwater
superoleophobic and anti-biofouling coating. (b) Changes of OCAs of hexadecane on the
MHMS-based coating as a function of the cycle number of the crushing and self-repairing. The
images show that the oil droplet of hexadecane is adhesive on the pressed surface (c) or
nonadhesive on the surface of the immersed coating at 20 ∘ C for 25 minutes (d). Source: Chen
et al. 2016 [80]. Reprinted with permission of American Chemical Society.

hydrophilic polymeric chains in the coating, this crushed surface can recover its
original underwater superoleophobicity (Figure 5.16b) and acquire oil-repellent
property again after cycles of press and immersion in water at 20 ∘ C for
25 minutes (Figure 5.16c,d).

5.4.3 Self-Cleaning of Contaminated Surfaces


Self-cleaning is a desired performance for human beings and applicable in a wide
domain of promising practical applications, such as industry, agriculture, and
the military [117, 118]. However, the surfaces of environmental materials dur-
ing their applications inevitably suffer from the contaminant, such as attach-
ment and absorption of dirt, bacteria, oil, and protein, resulting in weakened and
impaired surface structures and performances [119]. Therefore, materials with
self-cleaning surfaces are highly desirable in many environmental processes and
are necessary before ideally antifouling surfaces, if any, can be developed.
With the great progress on the different synthesis strategies and designs of
self-cleaning surfaces, recently, a wide range of self-cleaning surfaces have also
been commercialized [120, 121]. According to the mechanism of self-cleaning,
the majority of self-cleaning surfaces can be placed into three categories: super-
hydrophobicity, superhydrophilicity, and photocatalysis.

5.4.3.1 Superhydrophobicity-Induced Self-Cleaning


The self-cleaning effect of lotus leaves was observed 2000 years ago in Asia
as a miracle. However, the actual mechanism behind the mysterious miracle
5.4 Self-Healing Materials in Environmental Applications 177

was not scientifically revealed until the 1960s when the modern scanning
electron microscope (SEM) was developed [122, 123]. Lotus leaf surfaces
possess superhydrophobic and low-adhesion characteristics with water CA of
160∘ and SA of 2∘ , due to the combination of micro/nanoscaled hierarchical
structures and hydrophobic epicuticular waxes. Inspired by lotus leaves, a
multitude of self-cleaning materials with superhydrophobicity and low-adhesion
characteristics have been fabricated, which have very high water contact angles
(CAs) (>150∘ ), small water sliding angles, and small CAH (<10∘ ) in air (Cassie
state) [122, 124, 125]. Water on such self-cleaning superhydrophobic surface
tends to form spherical droplets, which can easily roll off the surface and
meanwhile carry dust and dirt with them to achieve the self-cleaning property
[122, 125–128]. In recent years, the self-cleaning superhydrophobic materials
have been widely applied in the fields of antifogging [129, 130], anti-icing [131],
corrosion resistance [132–134], solar water evaporation [135], light energy
harvesting by solar cell [136, 137], water drop energy harvesting [137, 138], etc.
For example, in 2015, Jiang and coworkers designed a carbon-black-based
superhydrophobic gauze that floats on the water surface and locally heats at
the air/water interface under solar irradiation, resulting in the enhancement
of evaporation rate [135]. Furthermore, the superhydrophobic black gauze was
endowed with self-cleaning ability, arising from its high water repellency and low
water adhesive force. As shown in Figure 5.17, a moving water droplet effectively
picked up the surface contamination of the NaCl crystals without any residue in
the superhydrophobic surface, revealing a lotus-leaf-like self-cleaning ability.
In the same year, Choi and coworkers fabricated a transparent hybrid cell
with the self-cleaning property that was able to complementarily harvest
raindrop and solar energy with immunity to dust contamination [137]. Such
hybrid cell consists of transparent triboelectric nanogenerator (TENG) and
conventional solar cell, and the transparent TENG was fabricated with PDMS
and indium tin oxide/polyethylene naphthalate (ITO/PEN) substrate. Due to

Contact and move

NaCl crystal

(a) (b) (c)

“Pick up” NaCl Self-cleaning

(d) (e) (f)

Figure 5.17 Demonstration of the self-cleaning ability of the superhydrophobic black gauze.
A moving droplet can effectively remove the surface contamination of NaCl crystals. Source:
Liu et al. 2015 [135]. Reprinted with permission of American Chemical Society.
178 5 Self-Healing Materials for Environmental Applications

the superhydrophobicity of the PDMS with the microbowl array, the solar cell
efficiency of the hybrid cell could recover to 84% after a cleaning process.

5.4.3.2 Superhydrophilicity-Induced Self-Cleaning



In case of superhydrophilic surfaces, the surfaces have water CA as small as 0
and allow water to spread out quickly to form a thin water film between fouling
debris and the underneath surface, which leads to separation of the debris from
the surface [125, 139–141].
On the other hand, hydrophilic polymers have been widely employed to
make self-cleaning and antifouling coatings [142]. Commonly, these antifoul-
ing surfaces have been reported through surface modification of hydrophilic
polymers onto substrates, especially involving PEO, PNIPAM, zwitterionic
polyelectrolytes, and copolymers with hydrophilic domains [139–141, 143–147].
PEO is commonly used in antifouling application since PEO molecule can
tightly bind with water by hydrogen-bonding interaction and form a powerful
barrier to resist fouling adsorption. Furthermore, PNIPAM-based materials
can perfectly release fouling only by low-temperature water rinsing due to the
hydrophilicity transition below LCST. In terms of zwitterionic polyelectrolytes,
zwitterions integrating both positively and negatively charged units in one group
can superiorly bind water molecules via electrostatically induced hydration. As a
result, a tightly bound water layer could be produced at the surfaces contacting
with water and form a barrier for self-cleaning and biofouling [140, 148–150].
For example, in 2015, Liu and coworkers grafted poly(2-methacryloyloxylethyl
phosphorylcholine) (PMPC), a typical zwitterionic polyelectrolyte, onto steel
meshes [141]. The PMPC modified meshes not only exhibited complete oil
repellency in a water-wetted state but also realized effective cleaning of oil
contamination on dry surfaces with simply water rinsing due to the high water
affinity of zwitterionic phosphorylcholine groups. As shown in Figure 5.18a, the
Nile red-labeled oil that fouled on the dry PMPC mesh underwent complete
dewetting and detachment in only 1.12 seconds after immersion into water,
which enabled a thorough cleaning of the oil-fouled surface by water washing
(Figure 5.18b).
In 2016, Jin and coworkers reported a sodium polyacrylate-grafted
poly(vinylidene fluoride) (PAAS-g-PVDF) hydrogel membrane via an
alkaline-induced phase inversion process, which possessed underwater
anti-crude-oil-adhesion property [151]. PAAS-g-PVDF exhibited underwater
ultralow-oil-adhesion and self-cleaning properties to crude oils with underwater
oil CAs all above 165∘ and nearly zero underwater oil adhesion force due to
the strong hydration capability of PAAS (Figure 5.19). A PAAS-g-PVDF-coated
copper mesh can effectively separate crude oil/water mixtures with high flux
up to 50 000 l m−2 h−1 and high separation efficiency of 97.1% and is able to be
recycled for long-term use.
Similarly, Zhang and coworkers reported a hydrogel-embedded tight ultrafil-
tration membrane consisting of sodium polyacrylate-modified polyacrylonitrile
(PAAS-m-PAN) via alkaline-assisted phase inversion process [152]. Driven by
an externally applied pressure of only 1 bar, this membrane can effectively sep-
arate dyes from both monovalent and divalent salts solutions with more than
140 l m−2 h−1 bar−1 water permeance and up to 94% dye rejection. Meanwhile,
5.4 Self-Healing Materials in Environmental Applications 179

t = 1.12 s after immersion


into water

Water

(a)

Under visible light Under UV light

Water washing

Under visible light Under UV light

(b)

Figure 5.18 (a) Photos of stainless steel meshes coated without (left sample) and with (right
sample) PMPC coatings, fouled by 30 μl Nile red-labeled canola oil in air, before (left panel) and
after (right panel) immersion into water. (b) Photos, taken under visible (left panels) and UV
(right panels) light, of the stainless steel meshes coated without (left sample) and with (right
sample) PMPC coatings, fouled by Nile red-labeled canola oil in air, before (upper panels) and
after (lower panels) being washed by water alone. It can be observed that the oil fouled on the
PMPC-coated mesh can be thoroughly cleaned by water washing, as evidenced by the
absence of luminescence of Nile red under UV light. Source: He et al. 2015 [141]. Reprinted
with permission of American Chemical Society.

the PAAS-m-PAN membrane exhibited an unprecedented antifouling property,


without detecting observed dye adhesion on the membrane surfaces.
In 2016, Zhu and coworkers prepared a strong underwater superoleophobic
PNIPAM–clay nanocomposite hydrogel, which was developed as an oil/water
separation filtration membrane [153]. As the hybrid hydrogel was immersed into
crude oil, the adhered oil could be easily cleaned up by water washing, thus indi-
cating ultralow affinity to crude oil.
In 2017, a Janus membrane was designed by integrating an omniphobic sub-
strate and a hydrophilic and underwater superoleophobic surface layer [154]. The
electrospun membrane of cetyltrimethylammonium bromide/poly(vinylidene
fluoride-co-hexafluoropropylene) (CTAB/PVDF-HFP) was decorated with
fluorinated silicon nanoparticles (SiNPs) to serve as the omniphobic substrate
in air. Then the surface layer consisting of SiNPs, chitosan (CTS), and PFO was
applied onto the omniphobic substrate using spray coating (Figure 5.20a). The
Janus membranes maintained stable MD performance with both their fluxes
180 5 Self-Healing Materials for Environmental Applications

PAA-g-PVDF Copper mesh Nickel foam Nonwoven PP plastic

5 cm 5 cm 5 cm 5 cm 5 cm

Crude oil Crude oil Crude oil Crude oil Crude oil

(a) CA: ∼165° CA: ∼165° CA: ∼172° CA: ∼167° CA: ∼166°

Nonwoven

(b)

Steel mesh

(c)

Steel mesh

(d)

Latex glove

(e) Coated Uncoated

Figure 5.19 Underwater anti-oil-adhesion and self-cleaning property of PAAS-g-PVDF


coatings. (a) Optical images and underwater crude oil CAs of PAAS-g-PVDF-coated substrates
including copper mesh, nickel foam, nonwoven, and polypropylene (PP) plastic plate. (b–e) A
series of implementing processes to demonstrate the anti-viscous-oil-adhesion and
self-cleaning performance of PAAS-g-PVDF-coated nonwoven (b), stainless steel mesh
(c and d), and latex glove (e). The oil used in (b), (c), and (e) is crude oil. The oil used in (d) is
soybean oil. Optical images in the red frame show the oil fouling of the substrates without
PAAS-g-PVDF coating. Source: Gao et al. 2016 [151]. Reprinted with permission of John Wiley
and Sons.

and salt rejections over 10 hours of operation (Figure 5.20b) and exhibited low
adhesion of oil foulants onto the surface (Figure 5.20c). As-prepared composite
membrane overcame the existing limits in conventional MD membranes,
amphiphilic surfactant-induced wetting and fouling by hydrophobic contami-
nants, and thus enabled an effective desalination of hypersaline wastewater with
complex compositions driven by low-grade thermal energy.
5.4 Self-Healing Materials in Environmental Applications 181

Fibrous network: Second-level reentrant structure


First-level reentrant structure
Electrospinning

Janus(o) top

Florinated SiNPs CTS/PFO-SiNP


CTAB/PVDF-HFP Janus(o) bottom
(positively charged)
composite Janus(o)
(omniphobic)
(a)
Janus(o)
1.2
Normalized flux (J/J0)

100
Before

Salt rejection (%)


1.0
80
Janus(h)
Janus(o)

0.8
0.6
60
0.4 Janus(h) After
Janus(o)
0.2 40
0.0
0 1 2 8 9 10 Non-fouled
(b) Time (h) (c)

Figure 5.20 (a) Fabrication procedure of the Janus membrane. The first step involves
electrospinning a fibrous substrate of CTAB/PVDF-HFP; the second step involves adsorption of
SiNPs followed by surface fluorination; in the last step, a CTS/PFO-SiNP (nanoparticle–polymer
composite) coating was applied onto the omniphobic substrate. (b) Normalized water flux
(blue) and salt rejection (red) for MD wetting experiments. (c) Photographic images of the
membranes before and after the wetting experiments. Source: Huang et al. 2017 [154].
Reprinted with permission of American Chemical Society.

5.4.3.3 Photocatalytic Self-Cleaning


Unlike wetting surfaces that rely solely on water rinsing to clean the surface,
on a photocatalytic self-cleaning surface, the contaminant is washed away due
to the synergistic effect of photocatalysis and photoinduced superhydrophilicity
[118, 128]. Photocatalysis is basically inspired from the photosynthesis process
of the green leaves, which absorb sunlight to drive the chemistry. The artificial
photocatalyst materials commonly stem from semiconductor materials such as
TiO2 , WO3 , ZrO2 , ZnO, CdS, and so forth, which can decompose organic con-
tamination by photocatalyst reaction [118].
TiO2 , as an extensively investigated photocatalyst material, not only possesses
excellent photocatalytic degradation efficiency of organic pollutants in sun-
light but also shows chemical stability under light, low cost, and nontoxicity.
As demonstrated in Figure 5.21, TiO2 can absorb the light energy that is
equal to or higher than its bandgap energy, leading to the generation of the
excited charge carriers, photoelectrons and holes. Therefore, photogenerated
holes can decompose adsorbed organic molecules. Meanwhile, electrons can
integrate with atmospheric oxygen to generate the superoxide radical, which
rapidly attacks surrounding organic molecules. The process of photocatalytic
decomposition exhibits potential self-cleaning ability upon the UV irradiation
[118, 128, 155–157].
182 5 Self-Healing Materials for Environmental Applications



– Conduction
band

(a) + Valence
–++ band

– +
+
(a) +– – Photoexcited electron
+ Photoexcited hole

– + +
(b) (d)
A– + D+
– (c)
TiO2

A D

Figure 5.21 Upon irradiation of ultra-bandgap light, the semiconductor undergoes


photoexcitation. The electron and the hole that result can follow one of several pathways:
(a) electron–hole recombination on the surface, (b) electron–hole recombination in the bulk
reaction of the semiconductor, (c) electron acceptor (A) reduced by photogenerated electrons,
and (d) electron donor (D) oxidized by photogenerated holes. Source: Parkin and Palgrave
2005 [118]. Reprinted with permission of Royal Society of Chemistry.

The first photocatalytic self-cleaning surface emerged in 1995 when Paz et al.
fabricated a transparent TiO2 film coating on glass [158]. The photocatalytic
self-cleaning concept later extended into the fields of oil/water separation and
wastewater treatment [159–162]. For example, in 2013, Zhang et al. fabricated
oil/water separation materials with underwater superoleophobicity through
LbL assembly of sodium silicate and TiO2 nanoparticles on a stainless steel
mesh (Figure 5.22a) [163]. Under UV irradiation, the mesh effectively removed
and decomposed fouling oil contaminants, leading to a facile recovery of its
wettability and oil/water separation ability (Figure 5.22b).
Similarly, in 2016, Wang et al. prepared TiO2 -doped PVDF electrospun
nanofibrous membrane for on-demand oil/water separation [160]. The obtained
membrane possesses antifouling and self-cleaning performance, arising from
the photocatalytic property of TiO2 , which has practical significance in saving
solvents and recycling materials.
Since membrane fouling severely restricts membrane separation efficiency
for wastewater treatment, in 2017, Xu and coworkers fabricated a self-cleaning
wastewater treatment film. Correspondingly, a PDA/polyethyleneimine (PEI)
intermediate layer is fabricated on an ultrafiltration substrate, followed by
mineralization of a photocatalytic layer consisting of photocatalytic β-FeOOH
nanorods [162]. As shown in Figure 5.23, the PDA–PEI layer not only acted
as nanofiltration selective layer but also served as the intermediate layer for
5.5 Conclusion 183

LbL assembly of sodium


silicate and TiO2
Stainless steel mesh

Oil/water separation UV light

Self-cleaning
(a)

120

100
Contact angle (°)

80

60

40

20
0 1 2 3 4 5
(b) Cycle numbers

Figure 5.22 (a) Preparation of a self-cleaning underwater superoleophobic mesh for oil/water
separation by LbL assembly of sodium silicate and TiO2 on a steel mesh. (b) The water contact
angle of the coated mesh could be repeatedly recovered by UV illumination after it was
contaminated by oleic acid. (▴) (silicate/TiO2 )*20 coated mesh; (•) after oleic acid
contamination; (⬧) after UV illumination recovery. Source: Zhang et al. 2013 [163]. Reprinted
with permission of Springer Nature.

anchoring β-FeOOH nanorods through the strong coordination interaction


between Fe3+ and catechol groups. Under visible light, the β-FeOOH nanorod
modified membrane exhibited efficient photocatalytic activity for degrading
organic contaminant through the photo-Fenton reaction in the presence of
hydrogen peroxide, thus showing the self-cleaning property. Moreover, the
modified membrane exhibited good stability and maintained high nanofiltration
performance (≈97.3%) during five photocatalytic filtration cycles.

5.5 Conclusion
Overall, endowing materials with self-healing ability would improve materials’
service life and stability and also would minimize external intervention, includ-
ing monitoring, maintenance, and repairing, during their lifetime of operation.
The field is still at its birth stage, and the applications of the self-healing
184 5 Self-Healing Materials for Environmental Applications

UF substrate Visible light


e– H2O2

OH
PDA–PEI codeposition
h+ Dye solution
OH
GA crosslinking OH
Dyes Degraded
products
Photocatalytic

nanofiltration

Fe3+
Hydrolysis

PDA–PEI NFM Mineralized NFM Colorless filtrate

Figure 5.23 Schematic representation of preparation process for the β-FeOOH-mineralized


nanofiltration membrane and its application in photocatalytic nanofiltration. Source: Lv et al.
2017 [162]. Reprinted with permission of John Wiley and Sons.

materials in the environmental domain are limited to membrane filtration,


oil/water separation, and antifouling surface. Thus the application perspectives
of self-healing materials ought to be further explored and expanded to other
environmental processes.
At the state of the art, the following points are noteworthy.
(1) Although some promising advances have been made, there is still a long way
to go to implement self-healing environmental materials in real-world con-
ditions. Moreover, the methods of integrating self-healing agents into base
materials are typically nontrivial, expensive, and time consuming. The envi-
ronmental safety and toxicity of many self-healing agents/materials used in
the environmental processes have not been assessed. The future self-healing
materials that are suitable to practical applications should offer fast healing,
environmental compatibility, and cost effectiveness.
(2) Polymeric self-healing agents are commonly used due to their inherent
flexibility and softness, which, to some extent, discourage their applications
to rigid materials, and the design of high-strength self-healing materials
would break this bottleneck [1]. In 2018, Sun and coworkers have realized
mechanically robust polymer composites with self-healing ability by com-
pounding nitrogen-coordinated boroxine-containing poly(propylene glycol)
with poly(acrylic acid) (PAA) though complexation and hydrogen-bonding
interactions between them [164]. However, this concept is still not applied
in the environmental field, recently.
(3) Computer modeling can provide a deeper and more thorough insight into the
self-healing behavior of polymers and thereby aid the formulation of effective
guidelines for optimizing both the synthesis of these polymers and the design
of the system [165].
(4) Design of self-cleaning materials with the capabilities of self-healing of physi-
cal cracks and/or self-restoration of chemical components would be another
key future research since the self-cleaning components on the surface are
References 185

always fragile and easily damaged when operating in extreme environments


[81, 88].
(5) The recent research about environmental materials with self-healing ability
still remains in its early state. Eventually, it will exist in large volumes
to develop fast, repeatable, and highly effective self-healing materials in
some remarkable field in advanced environmental materials, involving the
self-healable protection of oil pipeline [166], the surface recovery of fog
harvesting [167–169], self-stored photothermal materials [170–173], etc.

The future development direction for the self-healing materials is to further


emulate nature. Bio-inspired self-healing materials have recently developed
as a major branch of intelligent materials, which were designed to recover
mostly mechanical damage or/and surface functions at ambient conditions
without using external stimuli or energy input [1, 8]. With better understanding
the healing phenomenon at the nano- to microscale, more novel self-healing
materials and scalable fabrication strategies could be created. The devices made
from self-healing materials can eventually be functional in severe working
conditions, like deep sea or even outer space. The maintenance and replacement
costs of these devices could also be reduced.

References
1 Hager, M.D., Greil, P., Leyens, C. et al. (2010). Self-healing materials.
Advanced Materials 22 (47): 5424–5430.
2 Brochu, A.B., Craig, S.L., and Reichert, W.M. (2011). Self-healing biomateri-
als. Journal of Biomedical Materials Research Part A 96 (2): 492–506.
3 Huynh, T.P., Sonar, P., and Haick, H. (2017). Advanced materials for use in
soft self-healing devices. Advanced Materials 29 (19): 1604973.
4 Chang, J., Zhang, L., and Wang, P. (2018). Intelligent environmental nanoma-
terials. Environmental Science: Nano 5 (4): 811–836.
5 Murphy, E.B. and Wudl, F. (2010). The world of smart healable materials.
Progress in Polymer Science 35 (1-2): 223–251.
6 Tee, B.C., Wang, C., Allen, R., and Bao, Z. (2012). An electrically and
mechanically self-healing composite with pressure- and flexion-sensitive
properties for electronic skin applications. Nature Nanotechnology 7 (12):
825–832.
7 Wool, R.P. (1978). Material response and reversible cracks in viscoelastic
polymers. Polymer Engineering & Science 18 (14): 1057–1061.
8 Guimard, N.K., Oehlenschlaeger, K.K., Zhou, J. et al. (2012). Current trends
in the field of self-healing materials. Macromolecular Chemistry and Physics
213 (2): 131–143.
9 Zhang, Y., Broekhuis, A.A., and Picchioni, F. (2009). Thermally self-healing
polymeric materials: the next step to recycling thermoset polymers? Macro-
molecules 42 (6): 1906–1912.
10 Chen, X., Dam, M.A., Ono, K. et al. (2002). A thermally re-mendable
cross-linked polymeric material. Science 295 (5560): 1698–1702.
186 5 Self-Healing Materials for Environmental Applications

11 Inglis, A.J., Nebhani, L., Altintas, O. et al. (2010). Rapid bonding/debonding


on demand: reversibly cross-linked functional polymers via Diels-Alder
chemistry. Macromolecules 43 (13): 5515–5520.
12 Tian, Q., Yuan, Y.C., Rong, M.Z., and Zhang, M.Q. (2009). A thermally
remendable epoxy resin. Journal of Materials Chemistry 19 (9): 1289–1296.
13 Ono, T., Nobori, T., and Lehn, J.-M. (2005). Dynamic polymer
blends-component recombination between neat dynamic covalent polymers
at room temperature. Chemical Communications (12): 1522–1524.
14 Schultz, R.K. and Myers, R.R. (1969). The chemorheology of poly(vinyl
alcohol)-borate gels. Macromolecules 2 (3): 281–285.
15 Deng, G., Tang, C., Li, F. et al. (2010). Covalent cross-linked polymer gels
with reversible sol-gel transition and self-healing properties. Macromolecules
43 (3): 1191–1194.
16 Amamoto, Y., Otsuka, H., Takahara, A., and Matyjaszewski, K. (2012).
Self-healing of covalently cross-linked polymers by reshuffling thiuram
disulfide moieties in air under visible light. Advanced Materials 24 (29):
3975–3980.
17 Burnworth, M., Tang, L., Kumpfer, J.R. et al. (2011). Optically healable
supramolecular polymers. Nature 472 (7343): 334–337.
18 Coulibaly, S., Roulin, A., Balog, S. et al. (2013). Reinforcement of optically
healable supramolecular polymers with cellulose nanocrystals. Macro-
molecules 47 (1): 152–160.
19 Scott, T.F., Schneider, A.D., Cook, W.D., and Bowman, C.N. (2005). Photoin-
duced plasticity in cross-linked polymers. Science 308 (5728): 1615–1617.
20 Froimowicz, P., Frey, H., and Landfester, K. (2011). Towards the generation
of self-healing materials by means of a reversible photo-induced approach.
Macromolecular Rapid Communications 32 (5): 468–473.
21 Amamoto, Y., Kamada, J., Otsuka, H. et al. (2011). Repeatable photoin-
duced self-healing of covalently cross-linked polymers through reshuffling
of trithiocarbonate units. Angewandte Chemie International Edition 50 (7):
1660–1663.
22 Ghosh, B. and Urban, M.W. (2009). Self-repairing oxetane-substituted chi-
tosan polyurethane networks. Science 323 (5920): 1458–1460.
23 Vogt, A.P. and Sumerlin, B.S. (2009). Temperature and redox responsive
hydrogels from ABA triblock copolymers prepared by RAFT polymerization.
Soft Matter 5 (12): 2347–2351.
24 Nakahata, M., Takashima, Y., Yamaguchi, H., and Harada, A. (2011).
Redox-responsive self-healing materials formed from host-guest polymers.
Nature Communications 2: 511.
25 Kowalski, D., Ueda, M., and Ohtsuka, T. (2010). Self-healing
ion-permselective conducting polymer coating. Journal of Materials Chem-
istry 20 (36): 7630–7633.
26 Williams, K.A., Boydston, A.J., and Bielawski, C.W. (2007). Towards electri-
cally conductive, self-healing materials. Journal of the Royal Society Interface
4 (13): 359–362.
27 Corten, C.C. and Urban, M.W. (2009). Repairing polymers using oscillating
magnetic field. Advanced Materials 21 (48): 5011–5015.
References 187

28 van Dijk, N. and van der Zwaag, S. (2018). Self-healing phenomena in met-
als. Advanced Materials Interfaces 5 (17): 1800226.
29 Thakur, V.K. and Kessler, M.R. (2015). Self-healing polymer nanocomposite
materials: A review. Polymer 69: 369–383.
30 Wu, D.Y., Meure, S., and Solomon, D. (2008). Self-healing polymeric mate-
rials: a review of recent developments. Progress in Polymer Science 33 (5):
479–522.
31 Tan, Y.J., Wu, J., Li, H., and Tee, B.C. (2018). Self-healing electronic materi-
als for a smart and sustainable future. ACS Applied Materials & Interfaces
10 (18): 15331–15345.
32 White, S.R., Sottos, N., Geubelle, P. et al. (2001). Autonomic healing of poly-
mer composites. Nature 409 (6822): 794–797.
33 Blaiszik, B.J., Kramer, S.L., Olugebefola, S.C. et al. (2010). Self-healing poly-
mers and composites. Annual Review of Materials Research 40: 179–211.
34 Guo, M., Li, W., Han, N. et al. (2018). Novel dual-component microen-
capsulated hydrophobic amine and microencapsulated isocyanate used for
self-healing anti-corrosion coating. Polymers 10 (3): 319.
35 Gao, L., He, J., Hu, J., and Wang, C. (2015). Photoresponsive self-healing
polymer composite with photoabsorbing hybrid microcapsules. ACS Applied
Materials & Interfaces 7 (45): 25546–25552.
36 Brown, E.N., Kessler, M.R., Sottos, N.R., and White, S.R. (2003). In situ poly
(urea-formaldehyde) microencapsulation of dicyclopentadiene. Journal of
Microencapsulation 20 (6): 719–730.
37 Blaiszik, B., Caruso, M., McIlroy, D. et al. (2009). Microcapsules filled with
reactive solutions for self-healing materials. Polymer 50 (4): 990–997.
38 Keller, M.W., White, S.R., and Sottos, N.R. (2007). A self-healing
poly(dimethyl siloxane) elastomer. Advanced Functional Materials 17 (14):
2399–2404.
39 Liu, X., Sheng, X., Lee, J.K., and Kessler, M.R. (2009). Synthesis and char-
acterization of melamine-urea-formaldehyde microcapsules containing
ENB-based self-healing agents. Macromolecular Materials and Engineering
294 (6-7): 389–395.
40 Cho, S.H., Andersson, H.M., White, S.R. et al. (2006).
Polydimethylsiloxane-based self-healing materials. Advanced Materials 18
(8): 997–1000.
41 Yang, J., Keller, M.W., Moore, J.S. et al. (2008). Microencapsulation of iso-
cyanates for self-healing polymers. Macromolecules 41 (24): 9650–9655.
42 Suryanarayana, C., Rao, K.C., and Kumar, D. (2008). Preparation and charac-
terization of microcapsules containing linseed oil and its use in self-healing
coatings. Progress in Organic Coatings 63 (1): 72–78.
43 Jin, H., Mangun, C.L., Stradley, D.S. et al. (2012). Self-healing thermoset
using encapsulated epoxy-amine healing chemistry. Polymer 53 (2): 581–587.
44 He, Z., Jiang, S., Li, Q. et al. (2017). Facile and cost-effective synthesis of
isocyanate microcapsules via polyvinyl alcohol-mediated interfacial polymer-
ization and their application in self-healing materials. Composites Science
and Technology 138: 15–23.
188 5 Self-Healing Materials for Environmental Applications

45 Wang, W., Xu, L., Liu, F. et al. (2013). Synthesis of isocyanate microcapsules
and micromechanical behavior improvement of microcapsule shells by oxy-
gen plasma treated carbon nanotubes. Journal of Materials Chemistry A 1
(3): 776–782.
46 Yi, H., Yang, Y., Gu, X. et al. (2015). Multilayer composite microcapsules
synthesized by Pickering emulsion templates and their application in
self-healing coating. Journal of Materials Chemistry A 3 (26): 13749–13757.
47 Wu, G., An, J., Tang, X.Z. et al. (2014). A versatile approach towards multi-
functional robust microcapsules with tunable, restorable, and solvent-proof
superhydrophobicity for self-healing and self-cleaning coatings. Advanced
Functional Materials 24 (43): 6751–6761.
48 Ma, Y., Jiang, Y., Tan, H. et al. (2017). A rapid and efficient route to prepara-
tion of isocyanate microcapsules. Polymers 9 (7): 274.
49 Hansen, C.J., Wu, W., Toohey, K.S. et al. (2009). Self-healing materials
with interpenetrating microvascular networks. Advanced Materials 21 (41):
4143–4147.
50 Kim, Y.H. and Wool, R.P. (1983). A theory of healing at a polymer-polymer
interface. Macromolecules 16 (7): 1115–1120.
51 Jud, K. and Kausch, H. (1979). Load transfer through chain molecules after
interpenetration at interfaces. Polymer Bulletin 1 (10): 697–707.
52 Jud, K., Kausch, H., and Williams, J. (1981). Fracture mechanics studies of
crack healing and welding of polymers. Journal of Materials Science 16 (1):
204–210.
53 Lin, C., Lee, S., and Liu, K. (1990). Methanol-induced crack healing in poly
(methyl methacrylate). Polymer Engineering & Science 30 (21): 1399–1406.
54 Yang, Y. and Urban, M.W. (2013). Self-healing polymeric materials. Chemical
Society Reviews 42 (17): 7446–7467.
55 Wei, Z., Yang, J.H., Zhou, J. et al. (2014). Self-healing gels based on constitu-
tional dynamic chemistry and their potential applications. Chemical Society
Reviews 43 (23): 8114–8131.
56 Liu, Y.L. and Chen, Y.W. (2007). Thermally reversible cross-linked
polyamides with high toughness and self-repairing ability from maleimide-
and furan-functionalized aromatic polyamides. Macromolecular Chemistry
and Physics 208 (2): 224–232.
57 Chen, X., Wudl, F., Mal, A.K. et al. (2003). New thermally remendable highly
cross-linked polymeric materials. Macromolecules 36 (6): 1802–1807.
58 Liu, Y.L. and Hsieh, C.Y. (2006). Crosslinked epoxy materials exhibiting ther-
mal remendablility and removability from multifunctional maleimide and
furan compounds. Journal of Polymer Science Part A: Polymer Chemistry 44
(2): 905–913.
59 Murphy, E.B., Bolanos, E., Schaffner-Hamann, C. et al. (2008). Synthesis
and characterization of a single-component thermally remendable polymer
network: Staudinger and Stille revisited. Macromolecules 41 (14): 5203–5209.
60 Wouters, M., Craenmehr, E., Tempelaars, K. et al. (2009). Preparation and
properties of a novel remendable coating concept. Progress in Organic
Coatings 64 (2): 156–162.
References 189

61 Zhang, M.Q. and Rong, M.Z. (2013). Intrinsic self-healing of covalent poly-
mers through bond reconnection towards strength restoration. Polymer
Chemistry 4 (18): 4878–4884.
62 Wojtecki, R.J., Meador, M.A., and Rowan, S.J. (2011). Using the dynamic
bond to access macroscopically responsive structurally dynamic polymers.
Nature Materials 10 (1): 14–27.
63 Sijbesma, R.P., Beijer, F.H., Brunsveld, L. et al. (1997). Reversible polymers
formed from self-complementary monomers using quadruple hydrogen
bonding. Science 278 (5343): 1601–1604.
64 Scherman, O.A., Ligthart, G.B.W.L., Sijbesma, R.P., and Meijer, E.W. (2006).
A selectivity-driven supramolecular polymerization of an AB monomer.
Angewandte Chemie International Edition 45 (13): 2072–2076.
65 Hofmeier, H. and Schubert, U.S. (2003). Supramolecular branching and
crosslinking of terpyridine-modified copolymers: complexation and decom-
plexation studies in diluted solution. Macromolecular Chemistry and Physics
204 (11): 1391–1397.
66 Schubert, U.S. and Eschbaumer, C. (2002). Macromolecules containing
bipyridine and terpyridine metal complexes: towards metallosupramolecular
polymers. Angewandte Chemie International Edition 41 (16): 2892–2926.
67 Fustin, C.A., Guillet, P., Schubert, U.S., and Gohy, J.F. (2007).
Metallo-supramolecular block copolymers. Advanced Materials 19 (13):
1665–1673.
68 Varley, R.J. and van der Zwaag, S. (2008). Towards an understanding of
thermally activated self-healing of an ionomer system during ballistic pene-
tration. Acta Materialia 56 (19): 5737–5750.
69 Kakuta, T., Takashima, Y., Nakahata, M. et al. (2013). Preorganized hydrogel:
self-healing properties of supramolecular hydrogels formed by polymeriza-
tion of host-guest-monomers that contain cyclodextrins and hydrophobic
guest groups. Advanced Materials 25 (20): 2849–2853.
70 Burattini, S., Colquhoun, H.M., Fox, J.D. et al. (2009). A self-repairing,
supramolecular polymer system: healability as a consequence of
donor-acceptor π–π stacking interactions. Chemical Communications 44:
6717–6719.
71 Zhang, M., Xu, D., Yan, X. et al. (2012). Self-healing supramolecular gels
formed by crown ether based host-guest interactions. Angewandte Chemie
International Edition 51 (28): 7011–7015.
72 Bergman, S.D. and Wudl, F. (2008). Mendable polymers. Journal of Materials
Chemistry 18 (1): 41–62.
73 Youngblood, J.P. and Sottos, N.R. (2008). Bioinspired materials for
self-cleaning and self-healing. MRS Bulletin 33 (8): 732–741.
74 Tyagi, P., Deratani, A., Bouyer, D. et al. (2012). Dynamic interactive mem-
branes with pressure-driven tunable porosity and self-healing ability.
Angewandte Chemie International Edition 51 (29): 7166–7170.
75 Kim, S.-R., Getachew, B.A., Park, S.-J. et al. (2016). Toward
microcapsule-embedded self-healing membranes. Environmental Science
& Technology Letters 3 (5): 216–221.
190 5 Self-Healing Materials for Environmental Applications

76 Getachew, B.A., Kim, S.-R., and Kim, J.-H. (2017). Self-healing hydrogel
pore-filled water filtration membranes. Environmental Science & Technology
51 (2): 905–913.
77 Fang, W., Liu, L., Li, T. et al. (2016). Electrospun N-substituted polyurethane
membranes with self-healing ability for self-cleaning and oil/water separa-
tion. Chemistry--A European Journal 22 (3): 878–883.
78 Huang, S., Yang, L., Liu, M. et al. (2013). Complexes of
polydopamine-modified clay and ferric ions as the framework for
pollutant-absorbing supramolecular hydrogels. Langmuir 29 (4): 1238–1244.
79 Kuroki, H., Tokarev, I., Nykypanchuk, D. et al. (2013). Stimuli-responsive
materials with self-healing antifouling surface via 3D polymer grafting.
Advanced Functional Materials 23 (36): 4593–4600.
80 Chen, K., Zhou, S., and Wu, L. (2016). Self-healing underwater superoleo-
phobic and antibiofouling coatings based on the assembly of hierarchical
microgel spheres. ACS Nano 10 (1): 1386–1394.
81 Yu, L., Chen, G.Y., Xu, H., and Liu, X. (2016). Substrate-independent, trans-
parent oil-repellent coatings with self-healing and persistent easy-sliding oil
repellency. ACS Nano 10 (1): 1076–1085.
82 Kim, S.-R., Getachew, B.A., and Kim, J.-H. (2017). In situ healing of compro-
mised membranes via polyethylenimine-functionalized silica microparticles.
Environmental Science & Technology 51 (21): 12630–12637.
83 Yan, L., Li, G., Ye, Z. et al. (2014). Dual-responsive two-component
supramolecular gels for self-healing materials and oil spill recovery. Chemical
Communications 50 (94): 14839–14842.
84 Chen, D., Wu, M., Li, B. et al. (2015). Layer-by-layer-assembled healable
antifouling films. Advanced Materials 27 (39): 5882–5888.
85 Li, L., Yan, B., Yang, J. et al. (2015). Novel mussel-inspired injectable
self-healing hydrogel with anti-biofouling property. Advanced Materials
27 (7): 1294–1299.
86 Cao, P.F., Li, B., Hong, T. et al. (2018). Superstretchable, self-healing poly-
meric elastomers with tunable properties. Advanced Functional Materials 28
(22): 1800741.
87 Quéré, D. (2008). Wetting and roughness. Annual Review of Materials
Research 38: 71–99.
88 Wong, T.-S., Kang, S.H., Tang, S.K. et al. (2011). Bioinspired self-repairing
slippery surfaces with pressure-stable omniphobicity. Nature 477 (7365):
443–447.
89 Vogel, N., Belisle, R.A., Hatton, B. et al. (2013). Transparency and damage
tolerance of patternable omniphobic lubricated surfaces based on inverse
colloidal monolayers. Nature Communications 4: 2176.
90 Wang, Z., Heng, L., and Jiang, L. (2018). Effect of lubricant viscosity on
the self-healing properties and electrically driven sliding of droplets on
anisotropic slippery surfaces. Journal of Materials Chemistry A 6 (8):
3414–3421.
91 Zhao, H., Sun, Q., Deng, X., and Cui, J. (2018). Earthworm-inspired
rough polymer coatings with self-replenishing lubrication for adaptive
References 191

friction-reduction and antifouling surfaces. Advanced Materials 30 (29):


1802141.
92 Kim, P., Wong, T.-S., Alvarenga, J. et al. (2012). Liquid-infused nanostruc-
tured surfaces with extreme anti-ice and anti-frost performance. ACS Nano
6 (8): 6569–6577.
93 Wang, Y., Yao, X., Chen, J. et al. (2015). Organogel as durable anti-icing
coatings. Science China Materials 58 (7): 559–565.
94 Urata, C., Dunderdale, G.J., England, M.W., and Hozumi, A. (2015).
Self-lubricating organogels (SLUGs) with exceptional syneresis-induced
anti-sticking properties against viscous emulsions and ices. Journal of Mate-
rials Chemistry A 3 (24): 12626–12630.
95 Zhuo, Y., Håkonsen, V., He, Z. et al. (2018). Enhancing mechanical durability
of icephobic surfaces by introducing autonomous self-healing function. ACS
Applied Materials & Interfaces 10 (14): 11972–11978.
96 Li, Y., Li, L., and Sun, J. (2010). Bioinspired self-healing superhydrophobic
coatings. Angewandte Chemie International Edition 49 (35): 6129–6133.
97 Dikić, T., Ming, W., van Benthem, R.A. et al. (2012). Self-replenishing sur-
faces. Advanced Materials 24 (27): 3701–3704.
98 Puretskiy, N., Stoychev, G., Synytska, A., and Ionov, L. (2012). Surfaces with
self-repairable ultrahydrophobicity based on self-organizing freely floating
colloidal particles. Langmuir 28 (8): 3679–3682.
99 Esteves, A., Luo, Y., van de Put, M.W. et al. (2014). Self-replenishing dual
structured superhydrophobic coatings prepared by drop-casting of an
all-in-one dispersion. Advanced Functional Materials 24 (7): 986–992.
100 Liu, Q., Wang, X., Yu, B. et al. (2012). Self-healing surface hydrophobicity by
consecutive release of hydrophobic molecules from mesoporous silica. Lang-
muir 28 (13): 5845–5849.
101 Chen, K., Zhou, S., and Wu, L. (2014). Facile fabrication of self-repairing
superhydrophobic coatings. Chemical Communications 50 (80):
11891–11894.
102 Chen, K., Zhou, S., Yang, S., and Wu, L. (2015). Fabrication of
all-water-based self-repairing superhydrophobic coatings based on
UV-responsive microcapsules. Advanced Functional Materials 25 (7):
1035–1041.
103 Liu, Y., Pei, X., Liu, Z. et al. (2015). Accelerating the healing of superhy-
drophobicity through photothermogenesis. Journal of Materials Chemistry A
3 (33): 17074–17079.
104 Xue, C.-H., Zhang, Z.-D., Zhang, J., and Jia, S.-T. (2014). Lasting and
self-healing superhydrophobic surfaces by coating of polystyrene/SiO2
nanoparticles and polydimethylsiloxane. Journal of Materials Chemistry
A 2 (36): 15001–15007.
105 Li, Y., Chen, S., Wu, M., and Sun, J. (2014). All spraying processes for the
fabrication of robust, self-healing, superhydrophobic coatings. Advanced
Materials 26 (20): 3344–3348.
106 Wang, H., Xue, Y., Ding, J. et al. (2011). Durable, self-healing superhy-
drophobic and superoleophobic surfaces from fluorinated-decyl polyhedral
192 5 Self-Healing Materials for Environmental Applications

oligomeric silsesquioxane and hydrolyzed fluorinated alkyl silane. Ange-


wandte Chemie International Edition 50 (48): 11433–11436.
107 Chen, S., Li, X., Li, Y., and Sun, J. (2015). Intumescent flame-retardant and
self-healing superhydrophobic coatings on cotton fabric. ACS Nano 9 (4):
4070–4076.
108 Wu, M., Ma, B., Pan, T. et al. (2016). Silver-nanoparticle-colored cotton
fabrics with tunable colors and durable antibacterial and self-healing super-
hydrophobic properties. Advanced Functional Materials 26 (4): 569–576.
109 Zhou, H., Wang, H., Niu, H. et al. (2013). Robust, self-healing superam-
phiphobic fabrics prepared by two-step coating of fluoro-containing polymer,
fluoroalkyl silane, and modified silica nanoparticles. Advanced Functional
Materials 23 (13): 1664–1670.
110 Wang, X., Liu, X., Zhou, F., and Liu, W. (2011). Self-healing superamphipho-
bicity. Chemical Communications 47 (8): 2324–2326.
111 Zhang, L., Tang, B., Wu, J. et al. (2015). Hydrophobic light-to-heat con-
version membranes with self-healing ability for interfacial solar heating.
Advanced Materials 27 (33): 4889–4894.
112 Li, D. and Guo, Z. (2017). Stable and self-healing superhydrophobic
MnO2 @fabrics: applications in self-cleaning, oil/water separation and wear
resistance. Journal of Colloid and Interface Science 503: 124–130.
113 Liu, M., Wang, S., Wei, Z. et al. (2009). Bioinspired design of a superoleo-
phobic and low adhesive water/solid interface. Advanced Materials 21 (6):
665–669.
114 Callow, J.A. and Callow, M.E. (2011). Trends in the development of environ-
mentally friendly fouling-resistant marine coatings. Nature Communications
2: 244.
115 Cai, Y., Lin, L., Xue, Z. et al. (2014). Filefish-inspired surface design for
anisotropic underwater oleophobicity. Advanced Functional Materials 24 (6):
809–816.
116 Kota, A.K., Kwon, G., Choi, W. et al. (2012). Hygro-responsive membranes
for effective oil-water separation. Nature Communications 3: 1025.
117 Liu, K., Yao, X., and Jiang, L. (2010). Recent developments in bio-inspired
special wettability. Chemical Society Reviews 39 (8): 3240–3255.
118 Parkin, I.P. and Palgrave, R.G. (2005). Self-cleaning coatings. Journal of
Materials Chemistry 15 (17): 1689–1695.
119 Van der Bruggen, B., Mänttäri, M., and Nyström, M. (2008). Drawbacks of
applying nanofiltration and how to avoid them: a review. Separation and
Purification Technology 63 (2): 251–263.
120 Liu, K. and Jiang, L. (2011). Bio-inspired design of multiscale structures for
function integration. Nano Today 6 (2): 155–175.
121 Yao, X., Song, Y., and Jiang, L. (2011). Applications of bio-inspired special
wettable surfaces. Advanced Materials 23 (6): 719–734.
122 Ganesh, V.A., Raut, H.K., Nair, A.S., and Ramakrishna, S. (2011). A
review on self-cleaning coatings. Journal of Materials Chemistry 21 (41):
16304–16322.
References 193

123 Solga, A., Cerman, Z., Striffler, B.F. et al. (2007). The dream of staying
clean: lotus and biomimetic surfaces. Bioinspiration & Biomimetics 2 (4):
S126–S134.
124 Li, W. and Amirfazli, A. (2007). Microtextured superhydrophobic surfaces: a
thermodynamic analysis. Advances in Colloid and Interface Science 132 (2):
51–68.
125 Nishimoto, S. and Bhushan, B. (2013). Bioinspired self-cleaning surfaces
with superhydrophobicity, superoleophobicity, and superhydrophilicity. RSC
Advances 3 (3): 671–690.
126 Marmur, A. (2004). The lotus effect: superhydrophobicity and metastability.
Langmuir 20 (9): 3517–3519.
127 Extrand, C.W. (2002). Model for contact angles and hysteresis on rough and
ultraphobic surfaces. Langmuir 18 (21): 7991–7999.
128 Liu, K. and Jiang, L. (2012). Bio-inspired self-cleaning surfaces. Annual
Review of Materials Research 42: 231–263.
129 Gao, X., Yan, X., Yao, X. et al. (2007). The dry-style antifogging proper-
ties of mosquito compound eyes and artificial analogues prepared by soft
lithography. Advanced Materials 19 (17): 2213–2217.
130 Cao, M., Guo, D., Yu, C. et al. (2015). Water-repellent properties of super-
hydrophobic and lubricant-infused “slippery” surfaces: a brief study on
the functions and applications. ACS Applied Materials & Interfaces 8 (6):
3615–3623.
131 Cao, L., Jones, A.K., Sikka, V.K. et al. (2009). Anti-icing superhydrophobic
coatings. Langmuir 25 (21): 12444–12448.
132 Liu, K., Li, Z., Wang, W., and Jiang, L. (2011). Facile creation of bio-inspired
superhydrophobic Ce-based metallic glass surfaces. Applied Physics Letters
99 (26): 261905.
133 Liu, K., Zhang, M., Zhai, J. et al. (2008). Bioinspired construction of Mg-Li
alloys surfaces with stable superhydrophobicity and improved corrosion
resistance. Applied Physics Letters 92 (18): 183103.
134 Ishizaki, T., Masuda, Y., and Sakamoto, M. (2011). Corrosion resistance and
durability of superhydrophobic surface formed on magnesium alloy coated
with nanostructured cerium oxide film and fluoroalkylsilane molecules in
corrosive NaCl aqueous solution. Langmuir 27 (8): 4780–4788.
135 Liu, Y., Chen, J., Guo, D. et al. (2015). Floatable, self-cleaning, and
carbon-black-based superhydrophobic gauze for the solar evaporation
enhancement at the air-water interface. ACS Applied Materials & Interfaces
7 (24): 13645–13652.
136 Zhu, J., Hsu, C.-M., Yu, Z. et al. (2009). Nanodome solar cells with efficient
light management and self-cleaning. Nano Letters 10 (6): 1979–1984.
137 Jeon, S.-B., Kim, D., Yoon, G.-W. et al. (2015). Self-cleaning hybrid energy
harvester to generate power from raindrop and sunlight. Nano Energy 12:
636–645.
138 Lin, Z.H., Cheng, G., Lee, S. et al. (2014). Harvesting water drop energy
by a sequential contact-electrification and electrostatic-induction process.
Advanced Materials 26 (27): 4690–4696.
194 5 Self-Healing Materials for Environmental Applications

139 Yang, C., Ding, X., Ono, R.J. et al. (2014). Brush-like polycarbonates contain-
ing dopamine, cations, and PEG providing a broad-spectrum, antibacterial,
and antifouling surface via one-step coating. Advanced Materials 26 (43):
7346–7351.
140 Jiang, S. and Cao, Z. (2010). Ultralow-fouling, functionalizable, and hydrolyz-
able zwitterionic materials and their derivatives for biological applications.
Advanced Materials 22 (9): 920–932.
141 He, K., Duan, H., Chen, G.Y. et al. (2015). Cleaning of oil fouling with water
enabled by zwitterionic polyelectrolyte coatings: overcoming the imperative
challenge of oil-water separation membranes. ACS Nano 9 (9): 9188–9198.
142 Banerjee, I., Pangule, R.C., and Kane, R.S. (2011). Antifouling coatings:
recent developments in the design of surfaces that prevent fouling by pro-
teins, bacteria, and marine organisms. Advanced Materials 23 (6): 690–718.
143 Ngang, H., Ahmad, A., Low, S., and Ooi, B. (2017). Preparation of ther-
moresponsive PVDF/SiO2 -PNIPAM mixed matrix membrane for saline oil
emulsion separation and its cleaning efficiency. Desalination 408: 1–12.
144 Wu, Y., Yan, M., Lu, J. et al. (2017). Facile bio-functionalized design of
thermally responsive molecularly imprinted composite membrane for
temperature-dependent recognition and separation applications. Chemical
Engineering Journal 309: 98–107.
145 Manso-Silván, M., Valsesia, A., Gilliland, D. et al. (2004). Ion-beam treat-
ment of PEO; towards a physically stabilized anti-fouling film. Surface and
Interface Analysis 36 (8): 733–736.
146 Bose, R.K., Nejati, S., Stufflet, D.R., and Lau, K.K. (2012). Graft polymer-
ization of anti-fouling PEO surfaces by liquid-free initiated chemical vapor
deposition. Macromolecules 45 (17): 6915–6922.
147 Rana, D. and Matsuura, T. (2010). Surface modifications for antifouling
membranes. Chemical Reviews 110 (4): 2448–2471.
148 Chen, S., Zheng, J., Li, L., and Jiang, S. (2005). Strong resistance of phos-
phorylcholine self-assembled monolayers to protein adsorption: insights
into nonfouling properties of zwitterionic materials. Journal of the American
Chemical Society 127 (41): 14473–14478.
149 He, Y., Hower, J., Chen, S. et al. (2008). Molecular simulation studies of
protein interactions with zwitterionic phosphorylcholine self-assembled
monolayers in the presence of water. Langmuir 24 (18): 10358–10364.
150 Schlenoff, J.B. (2014). Zwitteration: coating surfaces with zwitterionic func-
tionality to reduce nonspecific adsorption. Langmuir 30 (32): 9625–9636.
151 Gao, S., Sun, J., Liu, P. et al. (2016). A robust polyionized hydrogel with
an unprecedented underwater anti-crude-oil-adhesion property. Advanced
Materials 28 (26): 5307–5314.
152 Jiang, G., Zhang, S., Zhu, Y. et al. (2018). Hydrogel-embedded tight ultrafil-
tration membrane with superior anti-dye-fouling property for low-pressure
driven molecule separation. Journal of Materials Chemistry A 6 (7):
2927–2934.
153 Teng, C., Xie, D., Wang, J. et al. (2016). A strong, underwater superoleopho-
bic PNIPAM-clay nanocomposite hydrogel. Journal of Materials Chemistry A
4 (33): 12884–12888.
References 195

154 Huang, Y.-X., Wang, Z., Jin, J., and Lin, S. (2017). A novel Janus membrane
for membrane distillation with simultaneous fouling and wetting resistance.
Environmental Science & Technology 51 (22): 13304–13310.
155 Fujishima, A. and Honda, K. (1972). Electrochemical photolysis of water at a
semiconductor electrode. Nature 238 (5358): 37–38.
156 Hoffmann, M.R., Martin, S.T., Choi, W., and Bahnemann, D.W. (1995). Envi-
ronmental applications of semiconductor photocatalysis. Chemical Reviews
95 (1): 69–96.
157 Kamat, P.V. (2007). Meeting the clean energy demand: nanostructure archi-
tectures for solar energy conversion. The Journal of Physical Chemistry C
111 (7): 2834–2860.
158 Paz, Y., Luo, Z., Rabenberg, L., and Heller, A. (1995). Photooxidative
self-cleaning transparent titanium dioxide films on glass. Journal of Materials
Research 10 (11): 2842–2848.
159 Xu, C., Xu, Y., and Zhu, J. (2014). Photocatalytic antifouling graphene
oxide-mediated hierarchical filtration membranes with potential applica-
tions on water purification. ACS Applied Materials & Interfaces 6 (18):
16117–16123.
160 Wang, Y., Lai, C., Wang, X. et al. (2016). Beads-on-string structured
nanofibers for smart and reversible oil/water separation with outstand-
ing antifouling property. ACS Applied Materials & Interfaces 8 (38):
25612–25620.
161 Lee, J.A., Krogman, K.C., Ma, M. et al. (2009). Highly reactive
multilayer-assembled TiO2 coating on electrospun polymer nanofibers.
Advanced Materials 21 (12): 1252–1256.
162 Lv, Y., Zhang, C., He, A. et al. (2017). Photocatalytic nanofiltration mem-
branes with self-cleaning property for wastewater treatment. Advanced
Functional Materials 27 (27): 1700251.
163 Zhang, L., Zhong, Y., Cha, D., and Wang, P. (2013). A self-cleaning underwa-
ter superoleophobic mesh for oil-water separation. Scientific Reports 3: 2326.
164 Bao, C., Jiang, Y.J., Zhang, H. et al. (2018). Room-temperature self-healing
and recyclable tough polymer composites using nitrogen-coordinated borox-
ines. Advanced Functional Materials 28 (23): 1800560.
165 Patrick, J.F., Robb, M.J., Sottos, N.R. et al. (2016). Polymers with autonomous
life-cycle control. Nature 540 (7633): 363–370.
166 Azevedo, C.R. (2007). Failure analysis of a crude oil pipeline. Engineering
Failure Analysis 14 (6): 978–994.
167 Wu, J., Zhang, L., Wang, Y., and Wang, P. (2017). Efficient and anisotropic
fog harvesting on a hybrid and directional surface. Advanced Materials
Interfaces 4 (2): 1600801.
168 Wang, Y., Zhang, L., Wu, J. et al. (2015). A facile strategy for the fabrica-
tion of a bioinspired hydrophilic-superhydrophobic patterned surface for
highly efficient fog-harvesting. Journal of Materials Chemistry A 3 (37):
18963–18969.
169 Zhang, L., Wu, J., Hedhili, M.N. et al. (2015). Inkjet printing for direct
micropatterning of a superhydrophobic surface: toward biomimetic fog
harvesting surfaces. Journal of Materials Chemistry A 3 (6): 2844–2852.
196 5 Self-Healing Materials for Environmental Applications

170 Li, R., Zhang, L., Shi, L., and Wang, P. (2017). MXene Ti3 C2 : an effective 2D
light-to-heat conversion material. ACS Nano 11 (4): 3752–3759.
171 Shi, L., Wang, Y., Zhang, L., and Wang, P. (2017). Rational design of a
bi-layered reduced graphene oxide film on polystyrene foam for solar-driven
interfacial water evaporation. Journal of Materials Chemistry A 5 (31):
16212–16219.
172 Zhang, L., Li, R., Tang, B., and Wang, P. (2016). Solar-thermal conversion
and thermal energy storage of graphene foam-based composites. Nanoscale 8
(30): 14600–14607.
173 Wang, Y., Zhang, L., and Wang, P. (2016). Self-floating carbon nanotube
membrane on macroporous silica substrate for highly efficient solar-driven
interfacial water evaporation. ACS Sustainable Chemistry & Engineering 4
(3): 1223–1230.

You might also like