Sie sind auf Seite 1von 10

Current Opinion in Colloid & Interface Science 11 (2006) 193 – 202

www.elsevier.com/locate/cocis

Superhydrophobic surfaces
Minglin Ma, Randal M. Hill ⁎
Institute for Soldier Nanotechnologies and Department of Chemical Engineering, Massachusetts Institute of Technology,
77 Massachusetts Avenue, Cambridge 01239, Massachusetts, USA
Dow Corning Corp., Midland, MI 48686, USA
Available online 11 October 2006

Abstract

Non-wettable surfaces with high water contact angles (WCAs) and facile sliding of drops, called superhydrophobic or ultrahydrophobic, have
received tremendous attention in recent years. New publications have appeared in the last year documenting many new ways to prepare such
surfaces—ranging from application driven work to make robust self-cleaning surfaces to careful model studies of patterned surfaces seeking to
understand the relationship between surface morphology and wettability and droplet sliding. This review summarizes this recent work and looks
ahead to future developments. The emphasis of the review is on the diverse methods that have been developed to make such surfaces.
© 2006 Published by Elsevier Ltd.

Keywords: Superhydrophobic; Water-repellent; Surface; Coating; Self-cleaning; Lotus effect

1. Introduction “superhydrophobic”. However, some natural examples (the wings


of some insects) do not exhibit two length scales and many studies
Many surfaces in nature are highly hydrophobic and self- (for example, [6••,7]) in which nanostructured surfaces were
cleaning. Examples include the wings of butterflies [1] and the prepared have found large contact angles and low sliding angles
leaves of plants such as cabbage and Indian cress [2••]. Some calling into question the necessity for a double length scale.
undesirable plants such as gorse (Ulex europeaus) introduced Since the group at Kao [8••] first demonstrated artificial
into New Zealand, and many common yard weeds have waxy superhydrophobic surfaces in the mid-1990s, a very large
leaf surfaces that make wetting them with water-based her- number of clever ways to produce rough surfaces that exhibit
bicides very difficult. The trisiloxane superwetters [3] were superhydrophobicity have been reported. Besides water repel-
developed for their remarkable ability to wet such hard-to-wet lency, other properties such as transparency and color, an-
surfaces and enhance herbicide efficacy. The best known isotropy, reversibility, flexibility and breathability have also
example of a hydrophobic self-cleaning surface is the leaves of been incorporated into superhydrophobic surfaces. The inten-
the lotus plant (Nelumbo nucifera). Electron microscopy of the tion of this article is to provide readers the current status of
surface of lotus leaves shows protruding nubs about 20–40 μm studies on superhydrophobic surfaces, concentrating mainly on
apart each covered with a smaller scale rough surface of publications appearing in the past year. Useful recent reviews
epicuticular wax crystalloids [4•]. Numerous studies have con- have been published by Nakajima et al. [5•] and Sun et al. [9•].
firmed that this combination of micrometer-scale and nanome- Quere [10••] critically discussed the surface chemistry of non-
ter-scale roughness, along with a low surface energy material sticking surfaces from the original papers by Cassie and Wenzel
leads to apparent WCAs N 150°, a low sliding angle and the self- to the present.
cleaning effect [5•]. Surfaces with these properties are called Before we go into the details, it is worth pausing to recall
that the lotus plant achieves an apparent WCA N 160° and nil
sliding angle using paraffinic wax crystals containing predom-
⁎ Corresponding author. Dow Corning Corp. Visiting Scientist MIT Institute for
inantly –CH2– groups. Nature does not require the lower
Soldier Nanotechnologies 500 Technology Square Fourth Floor Cambridge, MA
surface energy of –CH3 groups or fluorocarbons to achieve
02140, USA. Tel.: +1 617 324 6446. these effects. This plainly demonstrates that extremely low
E-mail address: rmhill@mailaps.org (R.M. Hill). surface energy is not necessary to achieve non-wetting. Rather,
1359-0294/$ - see front matter © 2006 Published by Elsevier Ltd.
doi:10.1016/j.cocis.2006.06.002
194 M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202

the ability to control the morphology of a surface on micron and hydrophilicity (dedoped or neutral state) was obtained upon
nanometer length scales is the key. This decoupling of wetting changing the applied electrochemical potential.
from simple surface energy opens up many possibilities for
engineering surfaces. 2.2. Silicones
Controlling the wetting of surfaces is an important problem
relevant to many areas of technology. The interest in self- Another well-known material with low surface energy is
cleaning surfaces is being driven by the desire to fabricate such polydimethylsiloxane (PDMS). Because of its intrinsic deform-
surfaces for satellite dishes, solar energy panels, photovoltaics, ability and hydrophobic property, PDMS can readily be
exterior architectural glass and green houses, and heat transfer made into superhydrophobic surfaces using various methods
surfaces in air conditioning equipment. Non-wettable surfaces [18–20]. For example, Khorasani et al. [18] treated PDMS
may also impart the ability to prevent frost from forming or using a CO2-pulsed laser as an excitation source. The WCA for
adhering to the surface. The fact that liquid in contact with the treated PDMS was as high as 175° which was believed to be
such a surface slides with lowered friction suggests applica- due to both the porosity and chain ordering on the PDMS
tions such as microfluidics, piping and boat hulls. Most of surface (Fig. 2a). Similarly, Jin et al. [19] used a laser etching
these applications involve solid surfaces, but the emergence of method to make a rough surface of PDMS elastomer containing
flexible membrane forms should lead to uses in garments and micro-, submicro- and nanocomposite structures. Such a surface
barrier-membranes—both Cassie and Wenzel were originally exhibited a superhydrophobicity with WCA higher than 160°
involved in work on waterproofing textiles [11,12]. The non- and sliding angle lower than 5°. Sun et al. [20] recently reported
wettable character has been claimed in biomedical applica- a nanocasting method to make superhydrophobic PDMS sur-
tions ranging from blood vessel replacement to wound face. They first made a negative PDMS template using lotus leaf
management. We assume that unexpected applications will as an original template and then used the negative template to
emerge as the technology to make non-wettable surfaces make a positive PDMS template—a replica of the original lotus
matures—nature uses this property in all known ecosystems leaf. The positive PDMS template (Fig. 2b) had the same
from polar bears to ducks to butterflies and water-walkers to surface structures and superhydrophobicity as the lotus leaf.
plant leaves.
Techniques to make superhydrophobic surfaces can be sim-
ply divided into two categories: making a rough surface from a
low surface energy material and modifying a rough surface with
a material of low surface energy.

2. Roughening a low surface energy material

2.1. Fluorocarbons

Fluorinated polymers are of particular interest due to their


extremely low surface energies. Roughening these polymers in
certain ways leads to superhydrophobicity directly [13–15].
For example, Zhang et al. [13] reported a simple and effective
way to achieve a superhydrophobic film by stretching a poly
(tetrafluoroethylene) (Teflon®) film. The extended film con-
sisted of fibrous crystals with a large fraction of void space in
the surface which was believed responsible for the super-
hydrophobicity. Shiu et al. [14] treated a Teflon® film with
oxygen plasma and obtained a rough surface with a WCA of
168°. Due to their limited solubility, many fluorinated
materials have not been used directly but linked [16•] or
blended [17••] with other materials (which are often easy to
roughen) to make superhydrophobic surfaces. Yabu and
Shimomura [16•] prepared a porous superhydrophobic mem-
brane by casting a fluorinated block polymer solution under
humid environment (Fig. 1a). The membrane was also
transparent due to the small pore size. Xu et al. [17••] Fig. 1. SEM images of superhydrophobic surfaces made by roughening
fabricated a double-roughened perfluorooctanesulfonate fluorinated materials. (a) The honeycomb-patterned film (top image and cross
(PFOS) doped conducting polypyrrole (PPy) film by a section) cast from a solution of the polymer shown in the inset under humid
conditions [16•] (reproduced by permission of the American Chemical Society).
combination of electropolymerization and chemical polymer- (b) The PPy porous film made by electro- and chemical polymerization and
ization (Fig. 1b). Interestingly, a reversible switching between transition of hydrophobic states (inset) [17••] (reproduced by permission of
superhydrophobicity (doped or oxidized state) and super- Wiley-VCH).
M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202 195

Fig. 2. SEM images of superhydrophobic surfaces made by roughening silicone-based materials. (a) PDMS surface treated by CO2-pulsed laser [18] (reproduced by
permission of Elsevier); (b) lotus leaf-like PDMS surface by nanocasting [20]; (c) PS-PDMS/PS electrospun fiber mat and the droplets on it [21•]; (d) PS-PDMS surface
cast from a 5 mg/ml solution in dimethylformamide (DMF) in humid air [22]. (Fig. 2b, c and d are reproduced by permission of the American Chemical Society.)

Given the difference in composition and consequent surface method to produce a highly porous superhydrophobic surface of
energy between the lotus leaf (paraffinic wax crystals, –CH2–, polyethylene (PE) by controlling its crystallization behavior. WCA
30–32 mN/m) and the PDMS replica (–CH3, 20 mN/m), the up to 173° was obtained by adding nonsolvent (cyclohexanone) to
similarity of the hydrophobicity obtained is surprising. the PE/xylene solution to form nanostructured floral-like crystal
Another way to exploit the low surface energy of PDMS is to structures (Fig. 3a). Jiang et al. [24••] showed that by electrostatic
use a block copolymer such as poly(styrene-b-dimethylsilox- spinning and spraying a PS solution in dimethylformamide (DMF)
ane) (PS-PDMS). For instance, Ma et al. [21•] made a super- they obtained a superhydrophobic film composed of porous
hydrophobic membrane in the form of a nonwoven fiber mat by microparticles and nanofibers as shown in Fig. 3b. Lee et al. [25]
electrospinning a PS-PDMS block copolymer blended with PS produced vertically aligned PS nanofibers by using nanoporous
homopolymer (Fig. 2c). The superhydrophobicity with WCA of anodic aluminum oxide as a replication template in a heat- and
163° was attributed to the combination of enrichment of PDMS pressure-driven nanoimprint pattern transfer process. As the aspect
component on fiber surfaces and the surface roughness due to ratio of the PS nanofibers increased, the nanofibers could not stand
small fiber diameters (150 nm to 400 nm). The flexibility, upright but formed twisted bundles resulting in a three-dimension-
breathability and free-standing feature of the membrane are of ally rough surface (Fig. 3c) with advancing and receding WCA of
particular interest in areas such as textile and biomedical appli- 155.8° and 147.6°, respectively.
cations. More recently, Zhao et al. [22] prepared a super- Other organic materials such as polyamide [26], polycar-
hydrophobic surface by casting a micellar solution of PS-PDMS bonate [27] and alkylketene dimer [28] have also recently been
in humid air based on the cooperation of vapor-induced phase made into superhydrophobic surfaces. Yan et al. [29•] syn-
separation and surface enrichment of PDMS block (Fig. 2d). thesized a poly(alkylpyrrole) film by electrochemical polymer-
ization—the needle-like poly(alkylpyrrole) structures grown
2.3. Organic materials perpendicularly to the surface of the electrode yielded an
environmentally stable superhydrophobicity (Fig. 3d).
Although fluorocarbons and silicones are known as hydro-
phobic materials, nature achieves non-wetting and self-cleaning 2.4. Inorganic materials
using paraffinic hydrocarbons. Recently, several groups have
demonstrated superhydrophobic surfaces made from organic Certain inorganic materials have also been made into super-
materials. Lu et al. [23•] proposed a simple and inexpensive hydrophobic surfaces. For example, superhydrophobic surfaces
196 M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202

Fig. 3. SEM images of superhydrophobic surfaces by roughening organic materials. (a) Floral-like crystal structures of PE [23•]; (b) PS surface made by electrostatic
spinning and spraying [24••]; (c) aligned PS nanofibers replicated from nanoporous anodic aluminum oxide [25] (reproduced by permission of the American Chemical
Society); (d) double-roughened poly(alkylpyrrole) film made by electrochemical polymerization. Scare bar: 15 μm [29•]. (Fig. 3a, b and d are reproduced by
permission of Wiley-VCH).

have been produced from ZnO [30••,31] and TiO2 [32]. Fig. 4a
shows SEM image of the ZnO nanorods Feng et al. [30••]
synthesized via a two-step solution method. The ZnO nanorod
films were superhydrophobic due to the surface roughness and the
low surface energy of the (001) plane of the nanorods on the surface
of the film as confirmed by X-ray diffraction (see the inset of Fig.
4a). Interestingly, the superhydrophobicity was transformed to
superhydrophilicity upon UV irradiation which generated electron-
hole pairs and resulted in hydroxyl adsorption in the ZnO surface.
Dark storage of the UV irradiated film for 7 days made it
superhydrophobic again. Similarly, they also obtained TiO2
nanorod films with reversibly switchable wettability (Fig. 4b).

3. Making a rough substrate and modifying it with low


surface energy materials

Methods to make superhydrophobic surfaces by roughening


low surface energy materials are mostly one-step processes and
have the advantage of simplicity. But they are always limited to a
small set of materials. Making superhydrophobic surfaces by a

Fig. 4. SEM of superhydrophobic surfaces by roughening inorganic materials.


(a) Aligned ZnO nanorods prepared by a two-step solution approach. The insets
show the XRD pattern and hydrophobicity transition of the nanorod film [30••]
(reproduced by permission of the American Chemical Society). (b) TiO2
nanorod film and a single papilla at high magnification [32] (reproduced by
permission of Wiley-VCH).
M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202 197

totally different strategy, i.e., making a rough substrate first and [14,33], laser etching [19,34] and chemical etching [35•,36]
then modifying it with a low surface energy material, decouples have all been used in the past year to fabricate superhydrophobic
the surface wettability from the bulk properties of the material and surfaces. For example, Teshima et al. [33] obtained a transparent
enlarges potential applications of superhydrophobic surfaces. superhydrophobic surface from a poly(ethylene terephthalate)
There are many ways to make rough surfaces, including (PET) substrate via selective oxygen plasma etching followed by
those mentioned above such as mechanical stretching, laser/ plasma-enhanced chemical vapor deposition using tetramethyl-
plasma/chemical etching, lithography, sol–gel processing and silane (TMS) as the precursor (Fig. 5a). Qian and Shen [35•]
solution casting, layer-by-layer and colloidal assembling, described a simple surface roughening method by dislocation-
electrical/chemical reaction and deposition, electrospinning selective chemical etching on polycrystalline metals such as
and chemical vapor deposition. There are also several methods aluminum (Fig. 5b). After treatment with fluoroalkylsilane, the
commonly used to modify the chemistry of a surface. For etched metallic surfaces exhibited superhydrophobicity.
example, covalent bonds can be formed between gold and alkyl Lithography (e.g. photolithography, electron beam lithogra-
thiols. Silanes are often used to decrease the surface energy. phy, X-ray lithography, soft lithography, nanosphere lithogra-
Physical binding, adsorption and coating can also change the phy and so on) is a well-established technique for creating large-
surface chemistry. In this section, we will focus on the important area periodic micro-/nanopatterns [37••,38]. Abdelsalam et al.
techniques reported in the past two years to make rough surfaces [39] conducted a systematic study of the wetting of structured
(not necessarily from low surface energy materials) and gold surfaces formed by electrodeposition through a template of
subsequent modifications of the surface chemistry. Although submicrometer spheres and discussed the role of the pore size
many of these studies are directed towards making practical and shape in controlling wetting (Fig. 5c). Martines et al. [6••]
surfaces, some important model studies are included here also. fabricated ordered arrays of nanopits and nanopillars by using
electron beam lithography and plasma etching. They obtained a
3.1. Etching and lithography superhydrophobicity with WCA of 164° and hysteresis of 1° for
a surface consisting of tall pillars with cusped tops (see Fig. 5d
Etching is a straightforward and effective way to make rough for the SEM image) after a hydrophobization with octadecyl-
surfaces. Different etching methods including plasma etching tricholorosilane. Composite surfaces with superhydrophobic

Fig. 5. Superhydrophobic surfaces produced by etching and lithography. (a) AFM image of the PET surfaces coated with TMS layer after the oxygen plasma treatment
[33] (reproduced by permission of Elsevier). (b) SEM image of the aluminum surfaces etched with a Beck's dislocation etchant for 15 s at room temperature and the
shape of a droplet on the surface after fluoroalkylsilane coating (inset) [35•]. (c) SEM image of a gold film electrodeposited through a submicrometer sphere template
[39]. (d) SEM image of the nanopillars after hydrophobization the base diameter of the pillars is about 120 nm [6••]. (Fig. 5b, c and d are reproduced by permission of
the American Chemical Society.)
198 M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202

3.3. Layer-by-layer (LBL) and colloidal assembly

LBL self-assembly is a rich process to fabricate conformal


thin film coatings with molecular level control over film thick-
ness and chemistry using electrostatic interaction and hydrogen
bonding. Recently, LBL process has been used by several groups
to make rough surfaces for superhydrophobicity [45–48]. For
example, Zhai et al. [49••] used an LBL technique to create a
pH-sensitive poly(allylamine hydrochloride)/poly(acrylic acid)
(PAH/PAA) multilayer which formed a honeycomb-like struc-
ture on the surface after an appropriate combination of acidic
treatments. After cross-linking the structure, they deposited
silica nanoparticles on the surface via alternating dipping of the
substrates into an aqueous suspension of the negatively charged
nanoparticles and an aqueous PAH solution, followed by a final
dipping into the nanoparticle suspension. Fig. 7a shows the SEM
image of the resultant hierarchically roughened structures.
Superhydrophobicity was obtained after the surface was modified
by a chemical vapor deposition of (tridecafluoro-1,1,2,2-
tetrahydrooctyl)-1-trichlorosilane followed by a 2 h thermal
annealing. LBL self-assembly could also be combined with
electrochemical deposition to prepare superhydrophobic surfaces
as shown by Zhang et al. [45,46,50•] during the past 2 years.
Assembly from colloidal systems has also been demonstrat-
ed to provide right surface roughness for superhydrophobicity
[51•,52••,53]. Ming et al. [52••] prepared a double roughened
Fig. 6. Superhydrophobic surfaces prepared by sol–gel process. (a) SEM image
of the methyltriethoxysilane (MTEOS) sol–gel foam. The insets: phenolphtha-
lein in water on the sol–gel foam heated to 390 °C (left) and 400 °C (right)
[41••] (reproduced by permission of the Royal Society of Chemistry). (b) AFM
image of sol–gel film containing 30 wt.% colloidal silica. Image area: 5 × 5 μm2;
the inset shows surface condensation of water vapor on this film [42]
(reproduced by permission of the American Chemical Society).

and superhydrophilic patterns have also been reported by Notsu


et al. [40] using photolithography.

3.2. Sol–gel processing

Sol–gel processes have been used to fabricate superhydro-


phobic surfaces from a variety of materials [30••,32,41••,42–44].
In most of these investigations, no post-process hydrophobization
was used for the achievement of superhydrophobicity since the
low surface energy materials were already included in the sol–gel
process. For example, Shirtcliffe et al. [41••] prepared porous
sol–gel foams from organo-triethoxysilanes which exhibited
binary switching between superhydrophobicity and superhydro-
philicity when exposed to different temperatures (Fig. 6a). Hikita
et al. [42] used colloidal silica particles and fluoroalkylsilane as
the starting materials and prepared a sol–gel film with super-
liquid-repellency by hydrolysis and condensation of alkoxysilane
compounds (Fig. 6b). Instead of blending low surface energy
materials in the sols, Shang et al. [43] described a procedure to
make transparent superhydrophobic surface by modifying silica-
based gel films with a fluorinated silane. On a similar note, Wu
et al. [44] made a microstructured ZnO-based surface via a wet- Fig. 7. Double-roughened superhydrophobic surfaces made by (a) layer-by-layer
chemical process and obtained the superhydrophobicity after assembly [49••] and (b) colloidal assembly [52••] (reproduced by permission of
coating the surface with long-chain alkanoic acids. the American Chemical Society).
M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202 199

Fig. 8. SEM images of the superhydrophobic surfaces made by electrochemical reaction and deposition or chemical bath deposition. (a) Dendritic gold cluster formed
on an ITO electrode modified with a polyelectrolyte multilayer [50•]. (b) Copper surface with dual-scale roughness and the droplet on it (inset) [54•] (reproduced by
permission of Wiley-VCH). (c) The copper surface after electrochemical reaction with sulfur gas [55] (reproduced by permission of the Royal Society of Chemistry).
(d) The BCH-LA nanopin films and the contact angle (inset) [56••]. (Fig. 8a and d are reproduced by permission of the American Chemical Society.)

surface consisting of raspberry-like particles which were made electrochemical reaction of Cu or Cu–Sn alloy plated on steel
by covalently grafting amine-functionalized silica particles of sheets with sulfur gas, and subsequent perfluorosilane treatment
70 nm to epoxy-functionalized silica particles of 700 nm via the (Fig. 8c). It is worth mentioning that chemical bath deposition
reaction between epoxy and amine groups (Fig. 7b). The surface (CBD) has also been used to make nanostructured surfaces. For
became superhydrophobic after being modified with PDMS. example, Hosono et al. [56••] used this technique to fabricate a
nanopin film of brucite-type cobalt hydroxide (BCH) and
3.4. Electrochemical reaction and deposition obtained a WCA as high as 178° after further modification of
lauric acid (LA) (Fig. 8d).
Electrochemical reaction and deposition has been extensively
used to prepare superhydrophobic surfaces. For instance, Zhang 3.5. Other methods
et al. [50•] demonstrated that the surface covered with dendritic
gold clusters, which was formed by electrochemical deposition Electrospinning is powerful technique to make ultrafine fibers
onto indium tin oxide (ITO) electrode modified with a poly- and has been found by several groups to provide sufficient surface
electrolyte multilayer, showed superhydrophobic properties roughnesses for superhydrophobicity [15,21•,24••,57–59••,60].
after further deposition of a n-dodecanethiol monolayer. (See Electrospinning a hydrophobic material led to superhydrophobi-
Fig. 8a for typical structures of the gold clusters.) Shirtcliffe et al. city in one step. Ma et al. [59••] showed that an even higher
[54• ] prepared a double-roughened copper surface which superhydrophobicity with WCAs up to 175° could be obtained
resembled “chocolate chip cookies” by electrodeposition and by applying a thin layer of conformal coating of a fluorinated
patterning technique. Further hydrophobization with fluorocar- polymer to electrospun mats by initiated chemical vapor de-
bons yielded a superhydrophobicity with WCA of 160° position (iCVD). This paper contained also a useful discussion of
(Fig. 8b). Cho's group [55] recently described the fabrication the impact of fiber morphology on superhydrophobicity (see
of lotus leaf-like superhydrophobic metal surfaces by using Fig. 9a).
200 M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202

without contact angle hysteresis or sliding angle values. Details of


methods used to measure these quantities are often not mentioned.
Though central to this topic, the concept of “roughness” is
difficult to reduce to a model that includes all the diverse forms
found in nature or made in the laboratory. While the simple Cassie
and Wenzel models provide a useful framework to understand
high contact angles, a more detailed model will be necessary to
relate hysteresis to roughness. A recent attempt to broaden the
scope of “roughness” has been made by Nosonovsky and
Bhushan [64]. Contact angles N170° have been reported, for
which the Cassie model calculates about 1% contact between the
water and the solid surface [65]. This seems problematic and
demands direct experimental validation. It is worth noticing that
an alternative view of superhydrophobicity or hysteresis from the
perspective of contact line rather than contact area has been
extensively reported in particular by Oner and McCarthy [66••].
A fundamental question about superhydrophobic surfaces
especially for potential applications is the robustness of the
effect. Theoretical modeling of the transition between the het-
erogeneous wetting (Cassie) state and the homogeneous wetting
(Wenzel) state [67,68], hydraulic pressure experiments [69,70]
and water condensation experiments [71] have been reported in
the past year to provide new insight into this problem. Both
Cheng et al. [71] and Jin et al. [72] found high contact angles
and sticky drops, which challenges the perception that rough
surfaces lead to Cassie states.

Fig. 9. (a) Superhydrophobic surfaces by modifying electrospun fiber mats with Acknowledgments
a fluorinated polymer. The plot shows the effect of fiber morphology on the
contact angles [59••]; (b) SEM image of ZnO-coated CNTs. The insets are a
TEM of a single coated CNT and the contact angle on the surface [63]
This research was supported in part by the U.S. Army
(reproduced by permission of the American Chemical Society). through the Institute for Soldier Nanotechnologies, under
Contract DAAD-19-02-D-0002 with the U.S. Army Research
Chemical (or physical) vapor deposition (CVD or PVD) has Office. The content does not necessarily reflect the position of
been widely used for the modification of surface chemistry as well the Government, and no official endorsement should be
as the synthesis of nanostructured surfaces [61–63]. Huang et al. inferred. The authors would like to acknowledge the support
[63] reported a stable superhydrophobic surface composed of of Dow Corning Corporation and many useful discussions with
aligned carbon nanotubes (CNTs) synthesized by chemical vapor Gregory Rutledge.
deposition on an Fe–N catalyst layer. The superhydrophobicity
was achieved after a thin layer coating of ZnO (see Fig. 9b). References and recommended reading •,••

4. Conclusions and perspectives [1] Wagner T, Neinhuis C, Barthlott W. Wettability and contaminability of
insect wings as a function of their surface sculptures. Acta Zool
1996;77:213–25.
This is an active research field with many publications [2••] Neinhuis C, Barthlott W. Characterization and distribution of water-
appearing each month dealing with methods to make super- repellent, self-cleaning plant surfaces. Ann Bot 1997;79:667–77.This
hydrophobic surfaces and theoretical understanding of the re- paper characterized and discussed a lot of self-cleaning and water-
lationship between surface morphology and wettability/sliding. repellent plant surfaces.
[3] Hill RM. Superspreading. Curr Opin Colloid Interface Sci 1998;3:247–54.
Theoretical understanding is maturing with models appearing
[4•] Barthlott W, Neinhuis C. Purity of the sacred lotus, or escape from
capable of dealing with more complex forms of surface mor- contamination in biological surfaces. Planta 1997;202:1–8.This paper
phology. Techniques to make non-wettable solid surfaces using gave a nice demonstration of self-cleaning effect in Nature.
polymers or sol–gel chemistry have been widely documented. [5•] Nakajima A, Hashimoto K, Watanabe T. Recent studies on super-
Other types of surfaces, including flexible membrane forms are hydrophobic films. Monatsh Chem 2001;132:31–41.This paper
receiving more attention. Making such surfaces in a microfluidic reviewed most work through 2001.
[6••] Martines E, Seunarine K, Morgan H, Gadegaard N, Wilkinson CDW,
device will probably be demonstrated shortly. Studies that present Riehle MO. Superhydrophobicity and superhydrophilicity of regular
results for other aspects such as drag reduction are just beginning
to appear.
Many studies in this area present only superficial results for •
Of special interest.
••
wettability and slip—frequently only a contact angle is cited Of outstanding interest.
M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202 201

nanopatterns. Nano Lett 2005;5:2097–103.This paper studied system- [28] Mohammadi R, Wassink J, Amirfazli A. Effect of surfactants on wetting
atically the hydrophobicity, hydrophilicity and sliding behavior of water of super-hydrophobic surfaces. Langmuir 2004;20:9657–62.
droplets on nanoasperities of controlled dimensions. [29•] Yan H, Kurogi K, Mayama H, Tsujii K. Environmentally stable super
[7] Feng L, Li S, Li S, Zhai J, Song Y, Jiang L, et al. Super-hydrophobic water-repellent poly(alkylpyrrole) films. Angew Chem Int Ed
surface of aligned polyacrylonitrile nanofibers. Angew Chem Int Ed 2005;44:3453–6. This paper reported super water-repellent films
2002;41:1221–3. which also showed excellent environmental stability to both temper-
[8••] Onda T, Shibuichi S, Satoh N, Tsujii K. Super-water-repellent fractal ature and organic solvents and oils.
surfaces. Langmuir 1996;12:2125–7.This paper reported first experi- [30••] Feng XJ, Feng L, Jin MH, Zhai J, Jiang L, Zhu DB. Reversible super-
mental demonstration of artificial superhydrophobic surfaces at Kao. hydrophobicity to super-hydrophilicity transition of aligned ZnO
[9•] Sun TL, Feng L, Gao XF, Jiang L. Bioinspired surfaces with special nanorod films. J Am Chem Soc 2004;126:62–3.This was the first
wettability. Acc Chem Res 2005;38:644–52. This paper provided a report of reversible switching between superhydrophobicity and
summation of many of their nice data on superhydrophobicity. superhydrophilicity made from inorganic materials.
[10••] Quere D. Non-sticking drops. Rep Prog Phys 2005;68:2495–532.This [31] Yang YH, Li ZY, Wang B, Wang CX, Chen DH, Yang GW. Self-
paper was a comprehensive review especially on the fundamental studies assembled ZnO agave-like nanowires and anomalous superhydropho-
of wetting. bicity. J Phys, Condens Matter 2005;17:5441–6.
[11] Wenzel RN. The evaluation of textile waterproofing agents. Am Dyest [32] Feng XJ, Zhai J, Jiang L. The fabrication and switchable superhydropho-
Report 1936;25:505–14. bicity of TiO2 nanorod films. Angew Chem Int Ed 2005;44:5115–8.
[12] Baxter S, Cassie ABD. The water repellency of fabrics and a new water- [33] Teshima K, Sugimura H, Inoue Y, Takai O, Takano A. Transparent ultra
repellency test. J Text Inst 1945;36:T67–90. water-repellent poly(ethylene terephthalate) substrates fabricated by
[13] Zhang JL, Li JA, Han YC. Superhydrophobic PTFE surfaces by oxygen plasma treatment and subsequent hydrophobic coating. Appl
extension. Macromol Rapid Commun 2004;25:1105–8. Surf Sci 2005;244:619–22.
[14] Shiu J-Y, Kuo C-W, Chen P. Fabrication of tunable superhydrophobic [34] Song XY, Zhai J, Wang YL, Jiang L. Fabrication of superhydrophobic
surfaces. Proceedings of SPIE-The International Society for Optical surfaces by self-assembly and their water-adhesion properties. J Phys
Engineering, vol. 5648; 2005. p. 325–32. Chem, B 2005;109:4048–52.
[15] Singh A, Steely L, Allcock HR. Poly[bis(2,2,2-trifluoroethoxy)phos- [35•] Qian BT, Shen ZQ. Fabrication of superhydrophobic surfaces by
phazene] superhydrophobic nanofibers. Langmuir 2005;21:11604–7. dislocation-selective chemical etching on aluminum, copper, and zinc
[16•] Yabu H, Shimomura M. Single-step fabrication of transparent super- substrates. Langmuir 2005;21:9007–9.This paper described a simple
hydrophobic porous polymer films. Chem Mater 2005;17:5231–4.This method to make superhydrophobic metal surfaces.
paper described a simple method to make transparent superhydrophobic [36] Guo ZG, Zhou F, Hao JC, Liu WM. Stable biomimetic super-hydrophobic
surfaces. engineering materials. J Am Chem Soc 2005;127:15670–1.
[17••] Xu L, Chen W, Mulchandani A, Yan Y. Reversible conversion of [37••] Furstner R, Barthlott W, Neinhuis C, Walzel P. Wetting and self-
conducting polymer films from superhydrophobic to superhydrophilic. cleaning properties of artificial superhydrophobic surfaces. Langmuir
Angew Chem Int Ed 2005;44:6009–12. This paper reported the first 2005;21:956–61.This paper investigated the wetting and self-cleaning
synthesis of superhydrophobic conducting polymers and the reversible properties of various artificial superhydrophobic surfaces.
control of the wettability. [38] Callies M, Chen Y, Marty F, Pepin A, Quere D. Microfabricated
[18] Khorasani MT, Mirzadeh H, Kermani Z. Wettability of porous textured surfaces for super-hydrophobicity investigations. Microelec-
polydimethylsiloxane surface: morphology study. Appl Surf Sci tron Eng 2005;78–79:100–5.
2005;242:339–45. [39] Abdelsalam ME, Bartlett PN, Kelf T, Baumberg J. Wetting of regularly
[19] Jin MH, Feng XJ, Xi JM, Zhai J, Cho KW, Feng L, et al. Super- structured gold surfaces. Langmuir 2005;21:1753–7.
hydrophobic PDMS surface with ultra-low adhesive force. Macromol [40] Notsu H, Kubo W, Shitanda I, Tatsuma T. Super-hydrophobic/super-
Rapid Commun 2005;26:1805–9. hydrophilic patterning of gold surfaces by photocatalytic lithography.
[20] Sun MH, Luo CX, Xu LP, Ji H, Qi OY, Yu DP, et al. Artificial lotus leaf J Mater Chem 2005;15:1523–7.
by nanocasting. Langmuir 2005;21:8978–81. [41••] Shirtcliffe NJ, McHale G, Newton MI, Perry CC, Roach P. Porous
[21•] Ma M, Hill RM, Lowery JL, Fridrikh SV, Rutledge GC. Electrospun materials show superhydrophobic to superhydrophilic switching. Chem
poly(styrene-block-dimethylsiloxane) block copolymer fibers exhibit- Commun 2005:3135–7.This paper demonstrated porous materials
ing superhydrophobicity. Langmuir 2005;21:5549–54.This paper was which were either completely filled with liquid or not wetted at all
the first report of superhydrophobic surfaces in the form of nanofiber when exposed to different temperatures.
mats. [42] Hikita M, Tanaka K, Nakamura T, Kajiyama T, Takahara A. Super-
[22] Zhao N, Xie QD, Weng LH, Wang SQ, Zhang XY, Xu J. Super- liquid-repellent surfaces prepared by colloidal silica nanoparticles
hydrophobic surface from vapor-induced phase separation of copolymer covered with fluoroalkyl groups. Langmuir 2005;21:7299–302.
micellar solution. Macromolecules 2005;38:8996–9. [43] Shang HM, Wang Y, Limmer SJ, Chou TP, Takahashi K, Cao GZ.
[23•] Lu XY, Zhang CC, Han YC. Low-density polyethylene superhydro- Optically transparent superhydrophobic silica-based films. Thin Solid
phobic surface by control of its crystallization behavior. Macromol Films 2005;472:37–43.
Rapid Commun 2004;25:1606–10.This paper described a simple [44] Wu X, Zheng L, Wu D. Fabrication of superhydrophobic surfaces from
method to make superhydrophobic surfaces from a simple polymer. microstructured ZnO-based surfaces via a wet-chemical route. Lang-
[24••] Jiang L, Zhao Y, Zhai J. A lotus-leaf-like superhydrophobic surface: a muir 2005;21:2665–7.
porous microsphere/nanofiber composite film prepared by electrohydro- [45] Shi F, Wang ZQ, Zhang X. Combining a layer-by-layer assembling
dynamics. Angew Chem Int Ed 2004;43:4338–41.This was the first report technique with electrochemical deposition of gold aggregates to mimic
of superhydrophobic surface by electrostatic spinning and spraying. the legs of water striders. Adv Mater 2005;17:1005–9.
[25] Lee W, Jin M-K, Yoo W-C, Lee J-K. Nanostructuring of a polymeric [46] Zhao N, Shi F, Wang ZQ, Zhang X. Combining layer-by-layer assembly
substrate with well-defined nanometer-scale topography and tailored with electrodeposition of silver aggregates for fabricating super-
surface wettability. Langmuir 2004;20:7665–9. hydrophobic surfaces. Langmuir 2005;21:4713–6.
[26] Zhang J, Lu X, Huang W, Han Y. Reversible superhydrophobicity to [47] Jisr RM, Rmaile HH, Schlenoff JB. Hydrophobic and ultrahydrophobic
superhydrophilicity transition by extending and unloading an elastic multilayer thin films from perfluorinated polyelectrolytes. Angew
polyamide film. Macromol Rapid Commun 2005;26:477–80. Chem Int Ed 2005;44:782–5.
[27] Zhao N, Xu J, Xie QD, Weng LH, Guo XL, Zhang XL, et al. Fabrication [48] Han JT, Zheng Y, Cho JH, Xu X, Cho K. Stable superhydrophobic
of biomimetic superhydrophobic coating with a micro-nano-binary organic–inorganic hybrid films by electrostatic self-assembly. J Phys
structure. Macromol Rapid Commun 2005;26:1075–80. Chem, B 2005;109:20773–8.
202 M. Ma, R.M. Hill / Current Opinion in Colloid & Interface Science 11 (2006) 193–202

[49••] Zhai L, Cebeci FC, Cohen RE, Rubner MF. Stable superhydrophobic Macromolecules 2005;38:9742–8.This paper provided a useful insight
coatings from polyelectrolyte multilayers. Nano Lett 2004;4:1349–53. into the role of nanofiber morphology in the wettability of nonwoven mats.
This paper demonstrated a relatively simple process for creating highly [60] Shang HM, Wang Y, Takahashi K, Cao GZ, Li D, Xia YN. Nanostructured
stable superhydrophobic coatings with structures that conceptually superhydrophobic surfaces. J Mater Sci 2005;40:3587–91.
mimicked the lotus leaf. [61] Liu H, Feng L, Zhai J, Jiang L, Zhu DB. Reversible wettability of a
[50•] Zhang X, Shi F, Yu X, Liu H, Fu Y, Wang ZQ, et al. Polyelectrolyte chemical vapor deposition prepared ZnO film between superhydropho-
multilayer as matrix for electrochemical deposition of gold clusters: bicity and superhydrophilicity. Langmuir 2004;20:5659–61.
toward super-hydrophobic surface. J Am Chem Soc 2004;126:3064–5. [62] Zhu LB, Xiu YH, Xu JW, Tamirisa PA, Hess DW, Wong CP.
This paper proposed a new way for the fabrication of superhydrophobic Superhydrophobicity on two-tier rough surfaces fabricated by con-
surfaces by combining LBL technique and electrochemical deposition. trolled growth of aligned carbon nanotube arrays coated with
[51•] Shiu JY, Kuo CW, Chen PL, Mou CY. Fabrication of tunable su- fluorocarbon. Langmuir 2005;21:11208–12.
perhydrophobic surfaces by nanosphere lithography. Chem Mater [63] Huang L, Lau SP, Yang HY, Leong ESP, Yu SF, Prawer S. Stable
2004;16:561–4.This paper described a fabrication method for creating superhydrophobic surface via carbon nanotubes coated with a ZnO thin
well-ordered nanostructured surfaces whose hydrophobicity could be flm. J Phys Chem, B 2005;109:7746–8.
modeled and tuned. [64] Nosonovsky M, Bhushan B. Roughness optimization for biomimetic
[52••] Ming W, Wu D, van Benthem R, de With G. Superhydrophobic films superhydrophobic surfaces. Microsyst Technol 2005;11:535–49.
from raspberry-like particles. Nano Lett 2005;5:2298–301.This paper [65] Cassie ABD. Contact angles. Discuss Faraday Soc 1948;3:11–6.
reported a procedure for preparing robust superhydrophobic surfaces [66••] Oner D, McCarthy TJ. Ultrahydrophobic surfaces. Effects of topogra-
with hierarchical roughness. phy length scales on wettability. Langmuir 2000;16:7777–82.This
[53] Zhang G, Wang D, Gu ZZ, Mohwald H. Fabrication of superhydrophobic paper contains an important discussion on three-phase contact lines.
surfaces from binary colloidal assembly. Langmuir 2005;21:9143–8. [67] Li W, Amirfazli A. A thermodynamic approach for determining the
[54•] Shirtcliffe NJ, McHale G, Newton MI, Chabrol G, Perry CC. Dual-scale contact angle hysteresis for superhydrophobic surfaces. J Colloid
roughness produces unusually water-repellent surfaces. Adv Mater Interface Sci 2005;292:195–201.
2004;16:1929–32.This paper had a useful discussion on the effect of [68] Krasovitski B, Marmur A. Drops down the hill: theoretical study of
dual-scale roughness on superhydrophobicty. limiting contact angles and the hysteresis range on a tilted plate.
[55] Han JT, Jang Y, Lee DY, Park JH, Song SH, Ban DY, et al. Fabrication Langmuir 2005;21:3881–5.
of a bionic superhydrophobic metal surface by sulfur-induced [69] Zheng QS, Yu Y, Zhao ZH. Effects of hydraulic pressure on the stability
morphological development. J Mater Chem 2005;15:3089–92. and transition of wetting modes of superhydrophobic surfaces.
[56••] Hosono E, Fujihara S, Honma I, Zhou HS. Superhydrophobic Langmuir 2005;21:12207–12.
perpendicular nanopin film by the bottom-up process. J Am Chem Soc [70] Journet C, Moulinet S, Ybert C, Purcell ST, Bocquet L. Contact angle
2005;127:13458–9.This paper reported the highest contact angle so far measurements on superhydrophobic carbon nanotube forests: effect of
documented. fluid pressure. Europhys Lett 2005;71:104–9.
[57] Acatay K, Simsek E, Ow-Yang C, Menceloglu YZ. Tunable, super- [71] Cheng YT, Rodak DE, Angelopoulos A, Gacek T. Microscopic
hydrophobically stable polymeric surfaces by electrospinning. Angew observations of condensation of water on lotus leaves. Appl Phys Lett
Chem Int Ed 2004;43:5210–3. 2005;87:194112.
[58] Gu Z-Z, Wei H-M, Zhang R-Q, Han G-Z, Pan C, Zhang H, et al. [72] Jin MH, Feng XJ, Feng L, Sun TL, Zhai J, Li TJ, et al. Super-
Artificial silver ragwort surface. Appl Phys Lett 2005;86:201915/1–3. hydrophobic aligned polystyrene nanotube films with high adhesive
[59••] Ma M, Mao Y, Gupta M, Gleason KK, Rutledge GC. Superhydrophobic force. Adv Mater 2005;17:1977–81.
fabrics produced by electrospinning and chemical vapor deposition.

Das könnte Ihnen auch gefallen