Open access peer-reviewed chapter

Switchable and Reversible Superhydrophobic Surfaces: Part Two

Written By

Sabri Taleb, Thierry Darmanin and Frédéric Guittard

Reviewed: 08 December 2017 Published: 28 March 2018

DOI: 10.5772/intechopen.73020

From the Edited Volume

Interdisciplinary Expansions in Engineering and Design With the Power of Biomimicry

Edited by Gulden Kokturk and Tutku Didem Akyol Altun

Chapter metrics overview

1,521 Chapter Downloads

View Full Metrics

Abstract

In this book chapter, most of the methods used in the literature to prepare switchable and reversible superhydrophobic surfaces are described. Inspired by Nature, it is possible to induce the Cassie-Baxter-Wenzel transition using different external stimuli such as light, temperature, pH, ion exchange, voltage, magnetic field, mechanic stress, plasma, ultrasonication, solvent, gas or guest. Such properties are extremely important for various applications but especially for controllable oil/water separation membranes, oil-absorbing materials, and water harvesting systems.

Keywords

  • superhydrophobic
  • reversible
  • switchable
  • bioinspiration
  • biomimetism

1. Reversible superhydrophobic surfaces: Part two

This section provides continuation of the description of the stimuli used in the literature to induce reversible changes in surface wettability.

1.1. Ion exchange

Since 2004, the researchers have shown that the presence of charged species such as quaternary ammonium groups are sensitive to ion exchange and lead to different surface wettabilities. In 2004, Choi et al. prepared self-assembled monolayers (SAM) with imidazolium groups on smooth Au and Si/SiO2 substrates [1, 2, 3]. They studied the effect of a series of anions as shown in Figure 1 and they found that the surface hydrophobicity increases like this: Br>BF4>PF6>NO3>ClO4>TfO>Tf2N. Hence, Br ions led to the highest surface hydrophilicity and Tf2N the highest hydrophobicity.

Figure 1.

Grafting of monolayers on smooth gold substrates with imidazolium groups with sensitivity to ion exchange and surfaced hydrophobicity as a function of the ion-exchanged. Ref. [1], Copyright 2004. Reprinted with permission from American Chemical Society, USA.

The substituent has also an important influence on the surface wettability [3, 4]. For example, using a long substituent with 1-octyl-3-methylimidazolium ([omim]), θw was 68, 72, and 75° for Br, BF4 et Tf2N, respectively while with 1-benzyl-3-methylimidazolium ([bmim]) θw did not change because the anions are absent in [bmim] aggregation. Similarly, for compounds such as 1-alkyl-3-(3-triethoxysilylpropyl)imidazolium ([C1tespim]) (Figure 1),θw was 24, 30, and 42° for Cl, BF4 et PF6, respectively while with ([C4tespim]) θw did not change. Moreover, the cation nature has also an influence on the surface hydrophobicity [3].

In order to elaborate reversible superhydrophobic properties by ion exchange, gold micro and nanostructured substrates were performed by electroless etching process by immersing silicon substrates in aqueous solution of HAuCl4 and HF (Figure 2). The surface was then modified by SAM using a thiol terminated by a quaternary ammonium. Then, the surface wettability could be reversely changed from superhydrophilic to superhydrophobic after exchanging Cl ions by perfluorooctanoate (C7F15COO or PFO) ions [4].

Figure 2.

Grafting of ion-sensitive molecules on rough gold substrates. Ref. [4], Copyright 2016. Reprinted with permission from American Chemical Society, USA.

Moreover, polymers with charged species such as polyelectrolytes were also used to change the surface wettability [5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15]. For example, [PVBIm][PF6] ([1-(4-vinylbenzyle)-3-butylimidazolium hexafluorophosphate]) brushes were grafted on a silicon wafer by atom transfer radical polymerization (ATRP). The surface properties could be changed from hydrophobic (θw = 95°) to hydrophilic (θw = 41°) after exchanging PF6 ions by Cl ions. The wettability of polyelectrolyte brushes could be modified with appropriate counter-ions [12, 13, 14, 15] using poly(2-methacryl poly oxyethyl trimethylammonium chloride) (PMETAC) and poly(2-dimethylaminoethyl methacrylate chloride) (PDMAEMAC).

Superhydrophobic properties could be reached by grafting PMETAC on gold micro and nanostructured substrates by ATRP [16]. Then, reversible and switchable properties from superhydrophobic to superhydrophilic were obtained after exchanging Tf2N ions by SCN ions. PMETAC were also grafted onto the surface of multiwalled carbon nanotubes [17]. Their surface properties could be reversely changed from superhydrophobic/highly oleophobic to superhydrophilic/superoleophobic after exchanging PFO ions by SCN ions. PMETAC was also grafted on cotton fabrics leading to similar properties [18]. Zhang et al. fabricated first silicone nanofilaments and modified them by P(METAC-co-trifluoroethyl methacrylate) to the obtained reversible surface [19].

Cho et al. developed a multifunctional polyelectrolyte membranes by electrospinning of P(METAC-co-[trimethoxysilyl]propylmethacrylate) (PMETAC-co-TSPM) [20]. Here, the presence of TSPM (sol-gel precursor) was used not only to form a polymer network via intramolecular interactions but also to anchor substrates (Figure 3). The membranes could reversely change from superhydrophobic/highly oleophobic to superhydrophilic/superoleophobic after exchanging Cl ions by heptadecafluorooctanesulfonic acid (C8F17SO3 or HPS) ions. Moreover, the membranes were also highly efficient filter medium for removing multiple contaminants such as SO2 form waster gas streams. Another strategy was to deposit polyelectrolyte multilayers poly(diallyldimethylammonium chloride (PDDA) and poly(sodium 4-styrenesulfonate) (PSS) on gold micro and nanostructured substrates [21]. The authors studied the influence of the exchanged ions and the highest properties were obtained by exchanging Cl (θw < 5°) ions by PFO (θw = 164°). The authors also deposited polyelectrolyte multilayers on micro and nanostructured aluminum substrates obtained by etching in HCl and immersion in boiling water [22, 23]. Using PFO ions, the substrates were superhydrophobic and superoleophobic. When the substrates were immersed in seawater, the PFO ions were exchanged by hydrophilic Cl or SO42− making the substrates underwater superoleophobic.

Figure 3.

Membranes sensitivity to ion exchange prepared by electrospinning of PMETAC-co-TSPM. The resulting membranes could switch from superhydrophobic/highly oleophobic to superhydrophilic/superoleophobic by changing the counter-ions. Ref. [20], copyright 2012. Reprinted with permission from Royal Society of Chemistry, United Kingdom.

1.2. Magnetic field

Environment protection against oil leakage during oil tankers sinking is a major global problem. Finding new materials to separate oil/water mixtures is hence extremely important [24, 25, 26]. Athanassiou et al. reported the formation of a novel composite material based on polyurethane (PU) foams functionalized with colloidal superparamagnetic iron oxide (spinel-cubic γ-Fe2O3/Fe3O4) nanoparticles and submicrometer PTFE particles [27]. The resulting foams could efficiently separate oil from water. The combination of the functionalization of the PTFE-treated foam surfaces with colloidal iron oxide nanoparticles significantly increased the speed of oil absorption. The foams were also magnetically responsive because they could be magnetically actuated. Durable and magnetic PU sponges were also reported by CVD of tetraethoxysilane (TEOS) to bind Fe3O4 nanoparticles [28]. The sponges exhibited fast magnetic responsivity with a saturation magnetization of 22.73 emu/g and can be easily manipulated with a magnet. The sponges were also superhydrophobic and superoleophilic, quickly absorbed floating oils on the water surface and also displayed high mechanical properties.

Other materials were also used in the literature [29, 30, 31, 32, 33, 34, 35]. Inspired by marine mussel adhesive, Jiang et al. used dopamine to link Fe3O4 nanoparticles on electrospun PVDF [29]. After fluorination, the materials displayed oil adsorption properties and could move towards a magnet. Using magnetically responsive mesh substrates, micro-robots were fabricated (Figure 4) [30]. The meshes could float and move on air/water and oil/water interfaces and could be guided by a magnetic field.

Figure 4.

Fabrication of micro-robots using magnetically responsive mesh substrates and Fe3O4 nanoparticles. Ref. [30], Copyright 2015. Reprinted with permission from Royal Society of Chemistry, United Kingdom.

Zhang et al. reported superamphiphobic elastic and magnetic silicone sponges with excellent thermal stability [31]. Their sponges were prepared by hydrolytic condensation of methyltrimethoxysilane, dimethoxydimethylsilane in the presence of Fe3O4@SiO2 nanoparticles and finally fluorinated. Superhydrophobic and magnetic quartz fibers were also reported by loading in cobalt and modification with PDMS [32].

Another potential application of magnetically responsive materials is the control of liquid moving using a magnet [36]. Randomly oriented hierarchical arrays with control geometries (diameter, height, and density) could be prepared by the mouldless self-assembly of solutions comprising procured polymers and magnetic particles under a magnetic field. With their actuating and superhydrophobic properties, these flexible films enabled active, fast, precise, and reversible manipulation of droplets with the use of a magnet.

Moreover, using superparamagnetic droplets on a magnetic superhydrophobic film, it was also shown to switch from superhydrophobic (low adhesion) to parahydrophobic (high adhesion) properties after magnetization/demagnetization [37]. It is also possible to control the speed, the shape, and the self-assembly of magnetic droplets on the superhydrophobic surface by modulating the magnetic field (Figure 5) [38, 39].

Figure 5.

Modification of the speed, the shape and the self-assembly of magnetic droplets on the superhydrophobic surface by modulating the magnetic field. Ref. [38], Copyright 2013. Reprinted with permission from The American Association for the Advancement of Science, United Kingdom.

1.3. Mechanical stress

Most of the time, reversible switchable wettability is governed by the change in the surface chemistry (surface energy). However, the use of mechanical strains is a very interesting way to control the surface wettability by modifying the surface roughness. In 2004, Zhang et al. showed that the hydrophobicity of PTFE can reversely change from 108 to 165° as the material is extended to 190% [40]. This was attributed to an increase in the distance between the fibrous PTFE crystals. They also described a novel method to reversibly change the wettability (from superhydrophilic to superhydrophobic) of a polyamide film with a triangular net-like structure upon biaxial extension and unloading [41]. The average side-length of the triangular structures was around 200 μm before extension and the surface was superhydrophobic, and 450 μm and superhydrophilic after extension. The phenomena could be repeated around 20 times because the film had a good elasticity.

Using a flexible PDMS with micropillar arrays, it was reported the feasibility to drastic change the surface wettability by simple modification of the substrate curvature (Figure 6) [42, 43]. When the substrate was not incurved it was parahydrophobic (high water adhesion) while it became superhydrophobic after curvature become the curvature could induce air injection into the pillar arrays. This easy and reversible process could be used in microfluidic devices [42, 43, 44, 45].

Figure 6.

Change in the water adhesion by inducing flexion on a flexible micro-patterned substrate. Ref. [43], Copyright 2011. Reprinted with permission from AIP Publishing LLC, USA.

To obtain micro and nanostructures on PDMS, Singh et al. deposited Ag nanorods arrays using oblique angle deposition on prestretched PDMS (Figure 7) [46]. The substrates displayed both microbuckles/wrinkles and nanorods. Superhydrophobic properties with θw = 154.8° under 30% prestretching, which is due to optimal amplitude and periodicity of the wrinkles. The substrates also displayed anisotropic wetting and water droplets could move only along the direction parallel to the wrinkles. Reversible contact angles from 154.8 to 126.2° were also reported by simple stretching/relaxation cycles because the stretching changes the dimensions of the microstructures. Yang et al. also reported the properties of PDMS elastomers with microscale ripples and SiO2 nanoparticles, allowing to reach anisotropic superhydrophobic properties [47]. The surface properties were dependent on the ripple amplitude and periodicity and also on the surface chemistry. Here, the sliding angle could also reversibly be tuned with external strains and with fast response.

Figure 7.

Preparation of microwrinkles of PDMS by pre-stretching following by oblique angle deposition of Ag nanorod. Ref. [46], Copyright 2015. Reprinted with permission from American Chemical Society, USA.

Self-healing superhydrophobic textiles with mechanical responsivity were also reported [48]. To induce this property, polydopamine nanocapsules with trapped hydrophobic agents were coated on the textiles. The nanocapsules could be released using different mechanical stresses, such as stretching, compression, friction, and even mechanical washing, and lead to self-healing properties.

1.4. Plasma

In a plasma chamber, highly ionized species are created by applying an electric field between two electrodes. When the plasma interacts with a substrate, different effects can be produced such as the surface cleaning, the formation of chemical groups, and/or the formation of surface structures. These effects are highly dependent on the plasma parameters such as the used gas, the pressure or the power and also on the substrate nature. For example, if a hydrophobic monolayer is used to prepare a superhydrophobic surface, the plasma treatment can remove the monolayer and leads to superhydrophobic properties [49, 50, 51, 52]. However, to obtain reversible properties, it is often necessary to have again the hydrophobic monolayer and/or to storage in the dark.

Otherwise, following the used gas, the plasma treatment can also change the surface chemistry. For example, meshes substrates were coated by CVD with the nanocrystalline diamond film [53]. After H2 plasma, the surface diamond termination changed with hydrogen atom leading to superhydrophobic properties. Moreover, the surface properties were reversible by annealing in air at 500°C.

1.5. Ultrasound

Very recently, full reversibility between the Cassie-Baxter and Wenzel states were reported through acoustic pressure [54]. Ultrasonication is used both for the nucleation of bubbles directly on superhydrophobic surfaces. Oppositely, the collapse of an entrapped air film was promoted to reversibly manipulate the material surface properties and the interaction with the environment.

1.6. Guest

It is known that specific interactions between “guest” and “host” molecules can be extremely strong. This is the case, for example, between crown ethers and ions. Li et al. used this strategy to develop superhydrophobic surfaces with specific sensitivity [55]. They grafted by click chemistry calix[4]azacrown (C4AC) on micro/nanostructured silicon substrates. Different organic ion pairs were selected (Figure 8): 1-butyl-3-methylimidazolium (C4mim) with different counter ions (Cl, Br and PF6). Reversible surfaces from superhydrophobic to superhydrophilic were observed but only in the presence of [C4mim]Cl because [C4mim]Cl interacts with C4AC cavities and cone resulting in 1,3-alternate conversion. Calix[4]arene lipoic acid (C4LA) was also grafted on the gold substrate [56]. The substrate had specific interaction with guest molecules, resulting in a high decrease of surface hydrophobicity. The interaction was specific with methomyl (a carbamate pesticide) but not with four other tested carbamate pesticides. With the aim to remove metal ions from the environment, responsive mesh was prepared by coating with poly(acrylic acid) hydrogel [57]. The resulting mesh was superhydrophilic and underwater superoleophobic. The mesh could complex Hg+ ions by complexation with the COO groups of poly(acrylic acid) leading to an increase in water contact angle and a decrease in underwater oil contact angle.

Figure 8.

Preparation of guest-responsive surface using calix[4]azacrown derivatives. Ref. [55], Copyright 2012. Reprinted with permission from American Chemical Society, USA.

Metal-organic frameworks (MOFs), also known as porous coordination polymers, were also used in the literature. In order to prepare hydrophobic MOFs, Planas et al. used an ortho-carborane functionalized with pyridylmethyl alcohol groups at the C-positions as a hydrophobic linker to Zn-1,4-benzenedicarboxylate [58]. Moreover, the substrate could change from highly hydrophobic to superhydrophilic by immersion in NaOH/DMF. In order to obtain fluorescent and hydrophobic MOFS, pyrene was incorporated in the structure [59]. The resulting materials displayed super-absorbency and could remove oils from oil/water mixture. They could also C60 as guest offering great promise for application in sensors.

In order to develop biosensors with switchable wettability, different strategies were used in order to bind biological molecules. For that, a tricomponent copolymer containing phenylthiourea and phenylboronic acid was designed [60]. As a strong hydrogen-bonding donor, phenylthiourea was used to combine with the phosphate units and phenylboronic acid to combine with the pentose rings. The copolymer was grafted on structured silicon substrates. The substrates could switch from superhydrophobic to superhydrophilic after immersion in adenosine diphosphate (ADP) aqueous solution. Using copolymers with phenylboronic acid units, biosensors that can bind sugars such as glucose were also reported in the literature [61].

1.7. Solvent and gas

When structured polymer materials are immersed into a solvent such as water, the interactions between the polymer chains and in contact with the solvent can highly varied leading to possible changes in surface morphology and wettability. Mixed polymer brushes with hydrophobic and hydrophilic blocks were used and fixed to a substrate. After exposure to different solvents, the organization of these chains can be highly affected as shown in Figure 9 [62, 63]. As a consequence, the surface energy and as a consequence θw vary [64, 65].

Figure 9.

Changing in the organization of mixed polymer brushes with hydrophobic and hydrophilic blocks after exposure to different solvents. Ref. [63], Copyright 2002. Reprinted with permission from American Chemical Society, USA.

To verify this hypothesis, molecules in Y-shape composed of hydrophobic polystyrene PS and hydrophilic poly(acrylic acid) (PAA) were grafted on a substrate [66, 67]. When the substrate was immersed in toluene, a good solvent for PS, the layer at the extreme surface was mainly composed of PS brushes and reversely the orientation of the polymers chains could highly vary as a function of the affinity with the solvent. PAA-block-PS was also grafted on multiwall carbon nanotubes. The surface could reversely switch from superhydrophobic to parahydrophobic by immersion in water and heating [68].

In order to reach superhydrophobic properties, Minko et al. grafted carboxyl-terminated poly(styrene-co-2,3,4,5,6-pentafluorostyrene) (PSF-COOH) and carboxyl-terminated poly(2-vinylpyridine) (PVP-COOH) on flat and rough (with needle-like structures) PTFE substrates functionalized with hydroxyl and amino groups (Figure 10) [69, 70]. After exposure to toluene, the top layer was mainly composed of hydrophobic PSF while after exposure to water the surface topography changed: PSF formed round domains inside the hydrophilic PVP matrix. As a consequence, on flat substrates, θw varies from 118 to 25° while on rough substrates θw varies from 160 to 0°.

Figure 10.

Change in the surface topography and organization of needle-like PTFE nanostructures grafted with PSF-COOH and PVP-COOH after exposure to different solvents. Ref. [70], Copyright 2003. Reprinted with permission from American Chemical Society, USA.

Ji et al. developed an asymmetric free-standing film layer-by-layer (LbL) assembly using polyethyleneimine (PEI) and PAA [71]. After coating with Teflon on one side, the free-standing film was superhydrophobic on one side and hydrophilic on the other side. As a function of the humidity, the free-standing film could be in a control manner bended and unbended (Figure 11). In a similar manner, Sun et al. developed films that could reversely induce wrinkles in the presence of humidity by depositing hydrophobic SiO2 nanoparticles on multilayer assembly of poly(allylamine hydrochloride) (PAH) and PAA [72]. Aizenberg et al. also reported a very innovative strategy by embedding nanoarrays in a hydrogel [73]. When the material was exposed to water, the orientation of the nanostructures changes leading to different surface hydrophobicity.

Figure 11.

Control in the bending of free-standing films prepared by LbL of PEI and PAA as a function of humidity. Ref. [71], Copyright 2010. Reprinted with permission from American Chemical Society, USA.

Tang et al. reported that 1,2,3,4,5-hexaphenylsilole is a unique solvent-sensitive material with switchable luminescence properties (Figure 12) [74]. When this material is spin-coated on a substrate, the resulting material is smooth, slightly hydrophobic (θw = 97.0°) and displayed a green light luminescence. After exposure to ethanol vapor, the substrate became nanostructured, highly hydrophobic (θw = 136.3°) and displayed a blue light luminescence. Moreover, the surface properties could be reversely switched using toluene vapor.

Figure 12.

Reversible control in the surface structuration, hydrophobicity, and luminescence properties of 1,2,3,4,5-hexaphenylsilole films after exposure to ethanol vapor and toluene vapor, respectively. Ref. [74], Copyright 2008. Reprinted with permission from American Chemical Society, USA.

Inorganic materials can also be used to obtain solvent-sensitive but, here, it is often possible when interactions of the substrates with the solvent induce changes in the surface chemistry [75, 76, 77]. This is the case of titanate nanostructures in water. The authors reported a change in the surface hydrophobicity due to physically adsorbed water molecules [75]. Moreover, the surface was reversible by simple heating. Otherwise, the immersion of inorganic materials can also lead to the removal of the hydrophobic treatment and as a consequence highly decreases the surface hydrophobicity [78, 79].

Different materials sensitive to gases were also reported in the literature [80, 81, 82, 83, 84, 85, 86]. 19-25 Jiang et al. developed superhydrophobic indium hydroxide (InOH)3) with microcubes and nanorods using a hydrothermal process in the presence of InCl3 and urea (Figure 13) [80]. The surface could switch from superhydrophobic to superhydrophilic in the presence of ammonia (NH3·H2O). Indeed, In(OH)3 being acidic, the ammonia molecules would anchor to the surface and form an ammonia layer on the surface of In(OH)3. Moreover, the bonds created between In(OH)3 and NH3.H2O being weak, they can be easily broken by heating leading again to superhydrophobic properties. In(OH)3-PDMS sponges with NH3 sensitivity were also prepared in the presence of polydopamine [81]. In a similar manner, CO2 is an acid gas; it can react at the surface of a material if amine groups are present. This strategy was used by Yuan et al. [83]. They electrospun PMMA-co-poly(N,N-dimethylaminoethyl methacrylate), a copolymer with amine groups. The resulting materials could switch from highly hydrophobic and underwater oleophobic to highly hydrophilic and underwater oleophilic in the presence of CO2. Moreover, the properties were reversible in the presence of N2. A very interesting work was reported by Wang et al. using the allochroic material (a material that can changes color) crystal violet lactone (CVL) [84]. To obtain superhydrophobic properties, palygorskite@polysiloxane was modified with CVL. The resulting materials could change from blue (CVL+: the carboxylate group of CVL) to discolored (CVL) in the presence of different gas vapor such as acetone.

Figure 13.

Influence of NH3·H2O on the surface hydrophobicity of In(OH)3 with microcubes and nanorods. Ref. [80], Copyright 2008. Reprinted with permission from American Chemical Society, USA.

A palladium-based superhydrophobic substrate was reported by Pd coating on vertically aligned Si nanowires [85]. The resulting substrates could reversely switch from superhydrophobic to parahydrophobic (high adhesion) properties in the presence of H2 or air, respectively. Here, the changes are due to a change in the surface energy due to the formation of β-phase Pd hydride (PdHx). Moreover, the substrates could potentiate the therapeutic efficiency of 3D stem cell spheroids. ZnO nanowires could also be reversely switched from superhydrophobic to superhydrophilic in the presence of O2 and H2, respectively, but at high temperature (300°C) [86]. Using a model based on density functional theory, the authors showed that oxygen-related defects are responsible for the wettability switching.

1.8. Chemical reactions

Different grafting strategies can also be envisaged to prepare the surface with reversible wettability [87, 88, 89]. For example, nanofibers of poly(3,4-ethylenedioxythiophene) (PEDOT) functionalized with azido groups (N3) were prepared by electropolymerization [87]. Then, dithiolane groups were introduced by reaction with lipoic acid also called thioctic acid. Then, various thiols were introduced to modify the surface properties. Both the surface morphology and the surface hydrophobicity were affected by the post-treatment with the used thiol. Interestingly, the use of fluorinated thiols highly changed the surface morphology and porosity leading to superhydrophobic properties and highly oleophobic properties. Moreover, the surface hydrophobicity and oleophobicity were reversible by reaction with dithiothreitol (DTT) to form the dithiolane groups again (Figure 14).

Figure 14.

Reversible wetting properties by reaction of hydrophobic thiols on dithiolane and DTT [87].

The boronic ester chemistry was also used to reversely change the surface properties [88]. This time, nanostructured PEDOT films functionalized with protected 1,2-diol were prepared by electropolymerization. After deprotection, the surface could easily react with different boronic acids containing different aromatic groups as shown in Figure 15. Similarly, both the surface morphology and the surface hydrophobicity were affected by the use of boronic acid post-treatment. The highest properties were obtained with the pyrene group for which the surface was parahydrophobic (extremely high water adhesion) with θw = 135°.

Figure 15.

Reversible wetting properties by reaction of hydrophobic boronic acids on 1,2-diols and HCl [88].

1.9. Dual sensitivity

Temperature and pH: Switchable surfaces by temperature and pH were prepared using thermosensitive polymers functionalized by pH-sensitive groups such as carboxylic acids (COOH) or amines. For example, a copolymer of poly(N-isopropylacrylamide) (PNIPAAm: sensitive to temperature) and polyacrylic acid (PAA: sensitive to pH) was grafted on micro and nanostructured silicon substrates (Figure 16) [90, 91]. The resulting surface could switch from superhydrophobic (low adhesion) at high temperature (45°C) and/or low pH (2) to parahydrophobic (high adhesion) at low temperature (20°C) and/or high pH (11).

Figure 16.

Temperature and pH-switchable surfaces using PNIPAAm and PAA copolymers. Ref. [90], Copyright 2012. Reprinted with permission from Royal Society of Chemistry, United Kingdom.

Different polymers with amino or pyridinium groups were also used to induce pH sensitivity [92, 93, 94, 95, 96]. Hybrid responsive nanoparticles were prepared by grafting on a SiO2 core mixed block copolymers of PS and poly(4-vinylpyridine) (P4VP) [92, 93]. PS is a thermosensitive polymer with a Tg of 100°C while P4VP is sensitive to pH. The surface was superhydrophobic at high pH (6) and at a temperature (> 100°C) while θw decreased with the temperature and pH. The authors also observed that the surface roughness changed with the temperature and pH. Wang et al. modified stainless steel meshes with a hydrogel copolymer of 2-(dimethylamino)ethyl methacrylate and methacrylic acid (two monomers sensitive to pH) [95]. Using oil/water mixtures, the meshes could let water pass through with a separation efficiency of 98.35% but only at both high temperature (55°C) and low pH (7), or both low temperature (25°C) and high pH (13). Poly-L-lysine, as both pH and the thermosensitive peptide, was also tested on a micro and nanostructured silicon substrate [96]. The surface was superhydrophobic at both high temperature (60°C) and high pH (11.5), and became superhydrophilic as the temperature and/or the pH decrease.

Solvent and pH: To prepare polymer particles responsive to solvent and pH, copolymers were prepared using a hydrophilic moiety (PAA) and hydrophobic ones (poly(2,3,4,5,6-pentafluorostyrene) [97]. The polymer films produced by annealing in the air led to superhydrophobic properties while by annealing in water led to highly hydrophilic properties (Figure 17). This was due to the reorientation of the PAA groups in water. Moreover, due to the presence of COOH groups, the pH could also modify the surface charge between negative and neutral, which can also modify the surface hydrophobicity.

Figure 17.

Solvent and pH-switchable surfaces using poly(2,3,4,5,6-pentafluorostyrene) and PAA copolymers. Ref. [97], Copyright 2010. Reprinted with permission from American Chemical Society, USA.

Voltage and pH: Switchable surfaces by voltage and pH were prepared using conducting polymers for their sensitivity to voltage. First, polyaniline was used for its sensitivity also to pH [98, 99]. To form superhydrophobic micro and nanostructured polyaniline with urchin-like and core-shell structures, polystyrene microspheres were used as seed (hard template) for the growth of polyaniline nanofibers [98]. The surface could switch from superhydrophobic to superhydrophilic depending on the voltage and the pH. Similar properties were also reported on polyaniline-polyacrylonitrile coaxial nanofibers prepared by electrospinning [99].

Switchable surface from superhydrophobic properties to highly hydrophilic were obtained by electro-copolymerization of PEDOT monomers with both pH-sensitive groups (COOH) and fluorinated chains [100]. The authors studied the influence of the percentage of each monomer, the doping state and also the pH on both the surface hydrophobicity (Figure 18). The surface morphology changed with the percentage of each monomer. The highest wettability changes were observed for a mol% of EDOT-COOH between 12.5 and 25%.

Figure 18.

Voltage and pH-switchable surfaces using PEDOT-COOH and PEDOT-O-H12 copolymers [101].

Voltage and Ion Exchange: Switchable surfaces by voltage and ion exchange were prepared using conducting polymers for their sensitivity to voltage while the sensitivity to ion exchange could be obtained by introducing functional groups such as ammonium, imidazolium or sulfonate groups. For example, multiresponsive surfaces were obtained by grafting an imidazolium substituent on PEDOT polymers [101, 102]. Smooth polymer films were prepared by spin-coating and observed sensitivity to ion exchange, oxidative doping, temperature, and pH. Their surface hydrophobicity could be modified from 40 to 70–72° by exchanging the counter-anion of the imidazolium moiety with fluorinated bis(trifluoromethane)sulfonamide or nonafluoro-1-butanesulfonate anions. The surface properties could be also enhanced from 24 to 107° by depositing the polymer on ZnO nanowire arrays. The surface properties of PEDOT:PSS were also studied by the authors [103].

In order to reach superhydrophobic properties, PEDOT copolymers with both ion exchange functional groups and fluorinated chains were prepared by electro-copolymerization (Figure 19) [104, 106]. The authors studied the influence of the percentage of each monomer, the doping state and also the counter-ions of ion exchange functional groups on both the surface hydrophobicity (Figure 19) [104]. Surprisingly, using 25 mol% of EDOT-Py+, the surface could switch from superhydrophobic to hydrophilic by reduction (dedoping using a different voltage) and again superhydrophobic properties by ions exchange with hydrophobic bis(trifluoromethanesulfonyl)imide (Tf2N) ions while the superoleophobic properties remained unchanged. Hence, it was possible to obtain both superoleophobic and hydrophilic properties, which is extremely rare in the literature [105]. Moreover, superoleophobic properties were also prepared with poly(3,4-ethylenedioxypyrrole) PEDOP copolymers with ion exchange functional groups and fluorinated chains [106].

Figure 19.

Voltage and ion exchange-switchable surfaces using PEDOT-Py+ and PEDOT-F8 copolymers [104].

Light and ion exchange: Switchable surfaces by light and ion exchange were reported by SAM of an imidazolium moiety (sensitive to ion exchange) terminated with a fluorinated chain on nanostructured ZnO (sensitive to light) films [107]. The authors studied the influence of ZnO morphology, the counter-ions (I, BF4, PF6, Tf2N) of imidazolium groups and the light on the surface hydrophobicity. Superhydrophobic properties were obtained with ZnO nanoparticles and hydrophobic PF6 or Tf2N ions. Moreover, the surface could reversely switch from superhydrophobic to hydrophilic by UV irradiation and dark storage.

Mechanic stress and magnetic field: Liu et al. developed flexible conical arrays coated with magnetic nanoparticles for fog harvesting systems [108]. Under an external magnetic field, static fog water could be spontaneously and continuously captured and directionally transported from the tip to the base of the spine through periodic vibration of the flexible conical spines driven by the magnetic field and the Laplace pressure difference arising from the conical shape of the flexible spines (Figure 20). Magnetically sensitive superomniphobic surfaces were also reported by fabricating flexible micronail caps [109]. The micronail caps could reversely bend using an external magnetic field, which changed the surface properties from superomniphobic to superomniphilic.

Figure 20.

Switchable surfaces with mechanic stress and magnetic field by creating flexible magnetic conical arrays [108].

Multiresponsivity: Switchable surfaces with responsivity to both light, heat, and pH was prepared by modifying TiO2 (sensitive to light) nanoparticles with a copolymer of poly(N-isopropylacrylamide) (PNIPAAm: sensitive to temperature) and polyacrylic acid (PAA: sensitive to pH) [110]. The resulting polymer surfaces displayed reversible wettability from superhydrophobic to superhydrophilic by UV and heat treatment at 150°C or immersion in solution of pH 12 and 2. Multiresponsive surfaces to heat, pH, and sugars was also reported by using a copolymer of PNIPAAm (sensitive to temperature) and poly(acrylamidophenylboronic acid (sensitive to both sugars and pH) [111]. The surface could switch from superhydrophobic to superhydrophilic by cooling at T = 20°C, exposure to glucose or immersion at high pH (10.1).

Advertisement

2. Conclusion

In this book chapter, most of the articles dedicated to switchable and reversible superhydrophobic surfaces were reviewed. If superhydrophobic properties are highly present in Nature, the preparation of reversible superhydrophobic properties has become one of the most studied domains. Indeed, if robust superhydrophobic surfaces can be obtained if the surface structures are able to stabilize the Cassie-Baxter state, it is possible to induce the Cassie-Baxter-Wenzel transition. In the literature, various techniques were developed to control the surface wettability using extern stimuli such as light, temperature, pH, ion exchange, voltage, magnetic field, mechanic stress, plasma, ultrasonication, solvent, gas or guest. Such properties are extremely important for various applications but especially for controllable oil/water separation membranes, oil-absorbing materials, and water harvesting systems.

References

  1. 1. Lee BS, Chi YS, Lee JK, Choi IS, Song CE, Namgoong SK, Lee S. Journal of the American Chemical Society. 2004;126:480-481
  2. 2. Chi YS, Lee JK, Lee SG, Choi IS. Langmuir. 2004;20:3024-3027
  3. 3. Zhao Y, Li M, Lu Q. Langmuir. 2008;24:3937-3943
  4. 4. Osicka J, Ilcikova M, Popelka A, Filip J, Bertok T, Tkac J, Kasak P. Langmuir. 2016;32:5491-5499
  5. 5. Zhang Y, Shen Y, Yuan J, Han D, Wang Z, Zhang Q, Niu L. Angewandte Chemie International Edition. 2006;45:5867-5870
  6. 6. Park MJ, Lee JK, Lee BS, Lee YW, Choi IS, Lee SG. Chemistry of Materials. 2006;18:1546-1551
  7. 7. Dobbelin M, Marcilla R, Tollan C, Pomposo JA, Sarasua JR, Mecerreyes D. Journal of Materials Chemistry. 2008;18:5354-5358
  8. 8. Shen Y, Zhang Y, Zhang Q, Niu L, You T, Ivaska A. Chemical Communications. 2005:4193-4195
  9. 9. He X, Yang W, Pei X. Macromolecules. 2008;41:4615-4621
  10. 10. Law G, Watson PP. Langmuir. 2001;17:6138-6141
  11. 11. Carvalho PJ, Freire MG, Marrucho IM, Queimada AJ, Coutinho JAP. Journal of Chemical and Engineering Data. 2008;53:1346-1350
  12. 12. Azzaroni O, Brown AA, Huck WTS. Advanced Materials. 2007;19:151-154
  13. 13. Wang L, Lin Y, Peng B, Su Z. Chemical Communications. 2008:5972-5974
  14. 14. Azzaroni O, Moya S, Farhan T, Brown AA, Huck WTS. Macromolecules. 2005;38:10192-10199
  15. 15. Zhou F, HY H, Yu B, Osborne VL, Huck WTS, Liu WM. Analytical Chemistry. 2007;79:176-182
  16. 16. Lim HS, Lee SG, Lee DH, Lee DY, Lee S, Cho K. Advanced Materials. 2008;20:4438-4441
  17. 17. Jiang C, Wang Q, Wang T. New Journal of Chemistry. 2013;37:810-814
  18. 18. Jiang C, Wang Q, Wang T. New Journal of Chemistry. 2012;36:1641-1645
  19. 19. Hua Z, Yang J, Wang T, Liu G, Zhang G. Langmuir. 2013;29:10307-10312
  20. 20. Lee CH, Kang SK, Lim JA, Kwark YJ, Lim HS, Kim J, Cho JH. Journal of Materials Chemistry. 2012;22:14656-14660
  21. 21. Wang L, Peng B, Su Z. Langmuir. 2010;26:12203-12208
  22. 22. Zhang G, Zhang X, Huang Y, Su Z. ACS Applied Materials and Interfaces. 2013;5:6400-6403
  23. 23. Zhang G, Zhang X, Li M, Su Z. ACS Applied Materials and Interfaces. 2014;6:1729-1733
  24. 24. Yang F, Dong Y, Guo Z. Colloids and surfaces, A. 2014;463:101-109
  25. 25. Ali N, Zhang B, Zhang H, Zaman W, Li X. Colloids and surfaces, A. 2015;472:38-49
  26. 26. Zhang J, Li Y, Bao M, Yang X, Wang Z. Environmental Science and Technology. 2016;50:8809-8816
  27. 27. Calcagnile P, Fragouli D, Bayer IS, Anyfantis GC, Martiradonna L, Cozzoli PD, Cingolani R, Athanassiou A. ACS Nano. 2012;6:5413-5419
  28. 28. Wu L, Li L, Li B, Zhang J, Wang A. ACS Applied Materials and Interfaces. 2015;7:4936-4946
  29. 29. Wu J, Wang N, Zhao Y, Jiang L. Nanoscale. 2015;7:2625-2632
  30. 30. Zhang J, Feng H, Zao W, Zhao Y, Zhang H, Liu Y. RSC Advances. 2015;5:47892-47899
  31. 31. Li L, Zhang J. Advanced Materials Interfaces. 2016;3:1600517/1-1600517/7
  32. 32. Du R, Zhao Q, Li P, Ren H, Gao X, Zhang J. ACS Applied Materials and Interfaces. 2016;8:1025-1032
  33. 33. Su C, Yang H, Song S, Lu B, Chen R. Chemical Engineering Journal. 2017;309:3636-3373
  34. 34. Peng H, Wang H, Wu J, Meng G, Wang Y, Shi Y, Liu Z, Guo X. Industrial and Engineering Chemistry Research. 2016;55:832-838
  35. 35. Cheng M, Ju G, Jiang C, Zhang Y, Shi F. Journal of Materials Chemistry A. 2013;1:13411-13416
  36. 36. Kim JH, Kang SM, Lee BJ, Ko H, Bae WG, Suh KY, Kwak MK, Jeong HE. Scientific Reports. 2015;5:17843
  37. 37. Cheng Z, Feng L, Jiang L. Advanced Materials. 2008;18:3219-3225
  38. 38. Timonen JVI, Latikka M, Leibler L, Ras RHA, Ikkala O. Science. 2013;341:253-257
  39. 39. Hong X, Gao X, Jiang L. Journal of the American Chemical Society. 2007;129:1478-1479
  40. 40. Zhang JL, Li J, Han YC. Macromolecular Rapid Communications. 2004;25:1105-1108
  41. 41. Zhang JL, XY L, Huang WH, Han YC. Macromolecular Rapid Communications. 2005;26:477-480
  42. 42. Wu D, SZ W, Chen QD, Zhang YL, Yao J, Yao X, Niu LG, Wang JN, Jiang L, Sun HB. Advanced Materials. 2011;23:545-549
  43. 43. Wu SZ, Wang JN, Niu LG, Yao J, Wu D, Li AW. Applied Physics Letters. 2011;98:081902/1-081902/3
  44. 44. Vasileiou T, Gerber J, Prautzsch J, Schutzius TM, Poulikakos D. Proceedings of the National Academy of Sciences U.S.A. 2016;98:081902/1-081902/3
  45. 45. Liu Y, Andrew M, Li J, Yeomans JM, Wang Z. Nature Communications. 2015;6 10034
  46. 46. Goel P, Kumar S, Sarkar J, Singh JP. ASC Applied MaterialsInterfaces. 2015;7:8419-8426
  47. 47. Lin PC, Yang S. Soft Matter. 2009;5:1011-1018
  48. 48. Liu Y, Liu Y, Hu H, Liu Z, Pei X, Yu B, Yan P, Zhou F. Journal of Physical Chemistry. C. 2015;119:7109-7114
  49. 49. Zhu X, Zhang Z, Wang K, Yang J, Xu X, Men X, Zhou X. New Journal of Chemistry. 2012;36:1280-1284
  50. 50. Zhu X, Zhang Z, Men X, Yang J, Xu X, Zhou X. Applied Surface Science. 2011;257:3753-3757
  51. 51. Liu J, Chen F, Zheng H, Liu S, Sun J, Huang S, Song J, Jin Z, Liu X. RSC Advances. 2016;6:79437-79447
  52. 52. Wu J, Bai HJ, Zhang XB, JJ X, Chen HY. Langmuir. 2010;26:1191-1198
  53. 53. Yang Y, Li H, Cheng S, Zou G, Wang C, Lin Q. Chemical Communications. 2014;50:2900-2903
  54. 54. Pinchasik BE, Wang H, Möhwald H, Asanuma H. Advanced Materials Interfaces. 2016;3:1600722/1-1600722/6
  55. 55. Feng N, Zhao H, Zhan J, Tian D, Li H. Organic Letters. 2012;14:1958-1961
  56. 56. Zhang GF, Zhan JY, Li HB. Organic Letters. 2011;13:3392-3395
  57. 57. Xu L, Liu N, Cao Y, Lu F, Chen Y, Zhang X, Feng L, Wei Y. ACS Applied Materials and Interfaces. 2014;6:13324-13329
  58. 58. Rodriguez-Hermida S, Tsang MY, Vignatti C, Stylianou KC, Guillerm V, Perez-Carvajal J, Teixidor F, Vinas C, Choquesillo-Lazarte D, Verdugo-Escamilla C, Peral I, Juanhuix J, Verdaguer A, Imaz I, Maspoch D, Planas JG. Angewandte Chemie International Edition. 2016;55:1-6
  59. 59. Rao KV, Mohapatra S, Maji TK, George SJ. Chemistry - A European Journal. 2012;18:4505-4509
  60. 60. Qing G, Wang X, Fuchs H, Sun T. Journal of American Chemical Society. 2009;131:8370-8371
  61. 61. Qing G, Wang X, Jiang L, Fuchs H, Sun T. Soft Matter. 2009;5:2759-2765
  62. 62. Boyes SG, Brittain WJ, Wang X, Cheng SZD. Macromolecules. 2002;35:4960-4967
  63. 63. Jennings GK, Brantley EL. Advanced Materials. 2004;16:1983-1994
  64. 64. Zhao B, Brittain WJ. Macromolecules. 2000;33:8813-8820
  65. 65. Tan H, Sun TL, Li JH, Guo M, Xie XY, Zhong YP, Fu Q, Jiang L. Macromolecular Rapid Communications. 2005;26:1418-1422
  66. 66. Julthongpiput D, Lin YH, Teng J, Zubarev ER, Tsukruk VV. Journal of the American Chemical Society. 2003;125:912-915
  67. 67. Julthongpiput D, Lin Y, Teng J, Zubarev ER, Tsukruk VV. Langmuir. 2003;19:7832-7836
  68. 68. Yang J, Zhang Z, Men X, Xu X, Zhu X. Journal of Colloid and Interface Science. 2010;346:241-247
  69. 69. Minko S, Muller M, Motornov M, Nitschke M, Grundke K, Stamm M. Journal of the American Chemical Society. 2003;125:3896-3900
  70. 70. Sidorenko A, Minko S, Schenk-Meuser K, Duschner H, Stamm M. Langmuir. 1999;15:8349-8355
  71. 71. Shen L, Fu J, Fu K, Picart C, Ji J. Langmuir. 2010;26:16634-16637
  72. 72. Li Y, Li X, Guo W, Wu M, Sun J. Science China Chemistry. 2016;59:1568-1573
  73. 73. Sidorenko A, Krupenkin T, Aizenberg J. Journal of Materials Chemistry. 2008;18:38418-33846
  74. 74. Heng L, Dong Y, Zhai J, Tang B, Jiang L. Langmuir. 2008;24:2157-2161
  75. 75. Tokudome Y, Okada K, Nakahira A, Takahashi M. Journal of Materials Chemistry A. 2014;2:58-61
  76. 76. Lian Z, Xu J, Wang Z, Weng Z, Xu Z, Yu H. Applied Surface Science. 2016;390:244-247
  77. 77. Lv LB, Cui TL, Zhang B, Wang HH, Li XH, Chen JS. Angewandte Chemistry International Edition. 2015;54:15165-15169
  78. 78. Liu N, Cao Y, Lin X, Chen Y, Feng L, Wei Y. ACS Applied Materials and Interfaces. 2014;6:12821-12826
  79. 79. Basu BJ, Manasa J. Applied Physics A. 2011;103:343-348
  80. 80. Zhu W, Zhai J, Sun Z, Jiang L. Journal of Physical Chemical C. 2008;112:8338-8342
  81. 81. Wang J, Guo J, Si P, Cai W, Wang Y, Wu G. RSC Advances. 2016;6:4329-4334
  82. 82. Si Y, Yang F, Guo Z. Journal of Colloid and Interface Science. 2016;484:173-182
  83. 83. Che H, Huo M, Peng L, Fang T, Liu N, Feng L, Wei Y, Yuan J. Angewandte Chemistry, International Edition. 2015;54:8934-8938
  84. 84. Zhang Y, Dong J, Sun H, Yu B, Zhu Z, Zhang J, Wang A. ACS Applied Materials and Interfaces. 2016;8:27346-27353
  85. 85. Seo J, Lee JS, Lee K, Kim D, Yang K, Shin S, Mahata C, Jung HB, Lee W, Cho SW, Lee T. Advanced Materials. 2014;26:7043-7050
  86. 86. Yadav K, Mehta BR, Bhattacharya S, Singh JP. Scientific Reports. 2016;6:35073
  87. 87. Godeau G, Darmanin T, Guittard F. Reactive and Functional Polymers. 2015;96:44-49
  88. 88. Taleb S, Noyer E, Godeau G, Darmanin T, Guittard F. Chemphyschem. 2016;17:305-309
  89. 89. Godeau G, Darmanin T, Guittard F. Journal of Applied Polymer Science. 2016;133 43130
  90. 90. Cheng Z, Lai H, Du M, Zhu S, Zhang N, Sun K. Soft Matter. 2012;8:9635-9641
  91. 91. Xia F, Feng L, Wang S, Sun T, Song W, Jiang W, Jiang L. Advanced Materials. 2006;18:432-436
  92. 92. Motornov M, Sheparovych R, Lupitskyy R, MacWilliams E, Minko S. Advanced Materials. 2008;20:200-205
  93. 93. Motornov M, Sheparovych R, Lupitskyy R, MacWilliams E, Minko S. Journal of Colloid and Interface Science. 2007;310:481-488
  94. 94. JD W, Zhang C, Jiang DJ, Zhao SF, Jiang YL, Cai GQ, Wang JP. RSC Advances. 2016;6:24076-24082
  95. 95. Li T, Shen J, Zhang Z, Wang S, Wei D. RSC Advances. 2016;6:40656-60663
  96. 96. Guo Y, Xia F, Xu L, Li J, Yang W, Jiang L. Langmuir. 2010;26:1024-1208
  97. 97. Rodriguez-Hernandez J, Munoz-Bonilla A, Bousquet A, Ibarboure E, Papon E. Langmuir. 2010;26:18617-18620
  98. 98. Tan L, Cao L, Yang M, Wang G, Sun D. Polymer. 2011;52:4770-4776
  99. 99. Zhu Y, Feng L, Xia F, Zhai J, Wan M, Jiang L. Macromolecular Rapid Communications. 2007;28:1135-1141
  100. 100. Darmanin T, Guittard F. Chemphyschem. 2013;14:2529-2533
  101. 101. Doebbelin M, Tena-Zaera R, Marcilla R, Iturri J, Moya S, Pomposo JA, Mecerreyes D. Advanced Functional Materials. 2009;19:3326-3333
  102. 102. Doebbelin M, Pozo-Gonzalo C, Marcilla R, Blanco R, Segura JL, Pomposo JA, Mecerreyes D. Journal of Polymer Science, Part A: Polymer Chemistry. 2009;47:3010-3021
  103. 103. Doebbelin M, Marcilla R, Tollan C, Pomposo JA, Sarasua JR, Mecerreyes D. Journal of Materials Chemistry. 2008;18:5354-5358
  104. 104. Taleb S, Darmanin T, Guittard F. ACS Applied Materials and Interfaces. 2014;6:7953-7960
  105. 105. Taleb S, Darmanin T, Guittard F. RSC Advances. 2014;4:3550-3555
  106. 106. Taleb S, Darmanin T, Guittard F. Materials Today Communications. 2016;6:1-8
  107. 107. Li HL, Xin BW, Feng L, Hao JC. Science China Chemistry. 2014;57:1002-1009
  108. 108. Peng Y, He Y, Yang S, Ben S, Cao M, Li K, Liu K, Jiang L. Advanced Functional Materials. 2015;25:5967-5971
  109. 109. Grigoryev A, Tokarev I, Kornev KG, Luzinov I, Minko S. Journal of the American Chemical Society. 2012;134:12916-12919
  110. 110. Sun W, Zhou S, You B, Wu L. Journal of Materials Chemistry A. 2013;1:3146-3154
  111. 111. Xia F, Ge H, Hou Y, Sun T, Chen L, Zhang G, Jiang L. Advanced Materials. 2007;19:2520-2524

Written By

Sabri Taleb, Thierry Darmanin and Frédéric Guittard

Reviewed: 08 December 2017 Published: 28 March 2018