Sie sind auf Seite 1von 26

Advances in Colloid and Interface Science 169 (2011) 80105

Contents lists available at SciVerse ScienceDirect

Advances in Colloid and Interface Science


journal homepage: www.elsevier.com/locate/cis

Mimicking natural superhydrophobic surfaces and grasping the wetting process: A review on recent progress in preparing superhydrophobic surfaces
Y.Y. Yan a,, N. Gao a, W. Barthlott b
a b

Energy & Sustainability Research Division, Faculty of Engineering, University of Nottingham, UK Nees-Institut fr Biodiversitt der Panzen, Rheinische Friedrich-Wilhelms-Universitt Bonn, Meckenheimer Allee 170, D-53115 Bonn, Germany

a r t i c l e

i n f o

a b s t r a c t
A typical superhydrophobic (ultrahydrophobic) surface can repel water droplets from wetting itself, and the contact angle of a water droplet resting on a superhydrophobic surface is greater than 150, which means extremely low wettability is achievable on superhydrophobic surfaces. Many superhydrophobic surfaces (both manmade and natural) normally exhibit micro- or nanosized roughness as well as hierarchical structure, which somehow can inuence the surface's water repellence. As the research into superhydrophobic surfaces goes deeper and wider, it is becoming more important to both academic elds and industrial applications. In this work, the most recent progress in preparing manmade superhydrophobic surfaces through a variety of methodologies, particularly within the past several years, and the fundamental theories of wetting phenomena related to superhydrophobic surfaces are reviewed. We also discuss the perspective of natural superhydrophobic surfaces utilized as mimicking models. The discussion focuses on how the superhydrophobic property is promoted on solid surfaces and emphasizes the effect of surface roughness and structure in particular. This review aims to enable researchers to perceive the inner principles of wetting phenomena and employ suitable methods for creation and modication of superhydrophobic surfaces. 2011 Elsevier B.V. All rights reserved.

Available online 14 September 2011 Keywords: Superhydrophobic surface Wetting Lotus effect Biomimetic Surface morphology

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . Theoretical models . . . . . . . . . . . . 2.1. Wetting models . . . . . . . . . . . 2.2. Contact angles and their hysteresis . . 2.3. Surface tension and surface free energy 2.4. Wetting transitions . . . . . . . . . 2.4.1. Critical contact angle . . . . 2.4.2. Energy equations . . . . . . Superhydrophobic surfaces . . . . . . . . . 3.1. Superhydrophobic surfaces in nature . 3.2. Articial superhydrophobic surfaces . 3.2.1. Imprinting and templating . . Lithography . . . . . . . . . . . . . . . . Templating . . . . . . . . . . . . . . . . Plasma treatment . . . . . . . . . . . . . 6.1. Chemical deposition . . . . . . . . . CVD-based surface treatment . . . . . . . . 7.1. Layer-by-layer (LBL) deposition . . . 7.2. Colloidal assembling and aggregation . 7.3. Solgel process . . . . . . . . . . . 7.4. Electrospinning and electrospraying . 7.5. Phase separation . . . . . . . . . . 7.6. TiO2 and ZnO induced wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 81 81 82 83 84 84 85 86 86 89 89 89 90 92 93 93 94 96 98 99 101 102

3.

4. 5. 6. 7.

Corresponding author. Tel.: + 44 115 951 3168; fax: + 44 115 951 3159. E-mail address: yuying.yan@nottingham.ac.uk (Y.Y. Yan). 0001-8686/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.cis.2011.08.005

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

81

8. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

102 102 103

1. Introduction The superhydrophobicity (hydrophobicity) of solid surfaces has been investigated with considerable attention over the past few years and remarkable progress has been achieved. It has been discovered that water droplets on hydrophobic surfaces can exhibit a contact angle higher than 90, and some can even be approaching approximately up to 180 [15]. In particular, the contact angles related to superhydrophobic (or ultrahydrophobic) surfaces are greater than 150. And those superhydrophobic surfaces are very likely to have phenomenal roughness with micro- or nanosized (or even smaller) protrusions coming out of the surface [6,7]. Therefore, the liquid might contact only a few bits of the superhydrophobic surface without fully wetting it. Indeed, uid interacting with superhydrophobic surfaces is one important discipline of research in the 21st century, and can essentially inuence a lot of cutting-edge topics in engineering and biotech research which involve surface structures, uid motivation, and their physical and chemical properties. Basically, the contact angle related wetting phenomena are of great interest and importance to current research progress. A considerable amount of work has been carried out to study the involved mechanisms and principles [831]. Interestingly, many methods that are used to create manmade superhydrophobic surfaces are inspired by the Lotus Effect [32]. Recent research has indicated that there are even more superhydrophobic surfaces in nature [13,3335]; and this helps to promote the applications of biomimetic ideas into practical elds [3639]. Also, it is noticed that the underlying theories interpreting the phenomena are mainly focused on Young equation, Wenzel equation, and CassieBaxter equation [40,41], while the wetting phenomena have been studied over the past decade. Based on these theories, people have understood that either surface energy or surface structure (roughness) can inuence the contact angle of liquid droplets on solid surfaces. However, such equations are not sufcient to thoroughly explain the mechanisms of wetting phenomena, although they are still necessary. This review attempts to discuss concisely the most recent research progress into the superhydrophobicity of solid surfaces with regard to the wetting process, both theoretically and experimentally. From an overall point of view, the work discussed in this paper will include reviewing the fundamental theories and their complement, the superhydrophobic surfaces that are discovered in nature, and the way manmade superhydrophobic surfaces are prepared. To make this work clearly described, we refer to wetting as wetting phenomena of liquid droplets on solid surfaces within the context. 2. Theoretical models As discussed in the Introduction, although some research work has been undertaken to study the wetting phenomena on superhydrophobic surfaces, there are still quite a few critical questions remaining unsolved. For example, it is not yet fully known how and when a stable wetting state can be achieved. An interesting observation that both homogeneous and heterogeneous wetting states could coexist on the same surface, as shown in Fig. 1, was reported by Callies and Qur in 2005 [42]. The homogeneous state (Fig. 1, right) could be obtained by pressing one liquid droplet in the heterogeneous state (Fig. 1, left) against the solid substrate. This meant the state of a droplet on a solid surface was somehow related to the process how the

droplet was delivered and formed. However, it was also suggested by other research work that the wetting state should be induced by surface roughness [10]. To clarify the complex wetting process and mechanism, this section focuses on the fundamental theories of wetting, involving contact angles, wetting models, state transitions, and the role of surface energy. 2.1. Wetting models The starting point of a wetting model is dened by the Young equation, derived for a sessile drop on an ideal rigid, homogeneous, at and inert surface. The drop contacts its substrate on a disk (radius = R) where the three phases of the system coexist, and the so-called three phase contact line is formed, as shown in Fig. 2. The liquid joins the solid at a contact angle . Each interface draws the contact line so as to minimize the corresponding surface area, balancing the surface tensions on the direction of potential motion so that there will yield a relation attributed to Young, as shown in Eq. (2.1), although it does not explicitly show up in Young's publication [42]: cos SV SL LV 2:1

where is the surface tension that indicates the energy per unit surface area of the interface. The Young equation unambiguously relates the intrinsic contact angle (Young Contact Angle) to interfacial tensions (energies). However, since real surfaces usually vary in the surface conditions, most scenarios regarding contact angles in practice cannot be fully explained by the Young equation. Wenzel then proposed an equation relating the contact angle to surface roughness and surface energies [40,43]. It can be written as: r SV SL LV cos W

2:2

where * is the apparent Wenzel contact angle, which measures the W apparent contact angle inuenced by the roughness of solid surfaces. r corresponds to the roughness factor, also referred to as the roughness area ratio of the actual surface with respect to the geometric surface. The reformed Wenzel equation is normally written as: cos W r cos :

2:3

The Wenzel equation assumes the water will penetrate into the grooves caused by the surface roughness, and therefore the Wenzel

Fig. 1. Millimetric water drops (of the same volume) deposited on a superhydrophobic substrate. The heterogeneous state (left) and homogeneous state (right) can coexist. Image reproduced from Ref. [42], with permission from Royal Society of Chemistry, Copyright 2005.

82

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

equation is related to the homogeneous wetting regime [44]. The apparent Wenzel contact angle is in contrast with the real angle , because close inspection of a rough surface would always reveal the real contact angle on any element that could be regarded as smooth [45]. The measurement of contact angles might therefore be somewhat arbitrary, for one might pursue this argument to molecular or atomic dimensions. Fortunately, for practical purposes, it is possible to have a reasonable division of surfaces into smooth and rough. The Wenzel equation states that the wettability can be improved by the surface roughness for a hydrophilic surface, but might get worse for a hydrophobic one [46]. In general, there are two types of stable wetting states, namely, homogeneous wetting state and heterogeneous wetting state, and the metastable states in-between. While the Wenzel equation is applicable to the homogeneous wetting regime, the CassieBaxter equation corresponds to the heterogeneous wetting regime. If the surface roughness is illustrated as pillars or protrusions, the 2-dimensional schematics of those wetting states can be shown in Fig. 3. Thus, a homogeneous state corresponds to the fact that the liquid drop lls up the roughness grooves [Fig. 3(a)] and the Wenzel equation is applied; a heterogeneous state refers to the fact that air bubbles are entrapped inside the grooves underneath the liquid [Fig. (3b)] and the Cassie Baxter (CB) equation is then applied. In the heterogeneous state, the liquid only contacts the solid at the top of the protrusions on a fraction denoted as S [20,41], which is the ratio of the total area of the solidliquid interface with respect to the total area of solidliquid and liquidair interfaces in a plane geometrical area of unity parallel to the rough surface. If only air was present between the solid and the liquid, the contact angle would be 180. The contact angle of a fakir drop (Fig. 1, left) is caused by both the solid and the air, and this yields [41,42]: cos 1 S cos 1 S cos S 1:

Fig. 3. The homogeneous wetting state (a) and heterogeneous wetting state (b). The liquid drop in the homogeneous wetting state follows the solid surface and penetrates into the grooves caused by the protrusions; the liquid drop in the heterogeneous wetting state only contacts the top of the protrusions, leaving air below into the grooves.

To sum up briey for this part, the Young equation builds up the basic theory for wetting phenomena on ideal smooth surfaces. The Wenzel equation is valid if the surface is homogeneously wetted, while the CB equation applies to a heterogeneous state. At this point, it seems that people have already got the theoretical support to study those involved scenes. However, the real situation is not that simple. As recently realized [44,48,49], both Wenzel equation and CB equation could be correct only if the drop is sufciently large compared with the typical roughness scale. It is still under debate that when and how a stable wetting state can be achieved. Also, there might be a transition between homogeneous state and heterogeneous state, and it is even not fully understood how this transition happens. Thus, we can only say that those mentioned theories are to some extent necessary, but not sufcient for entirely describing wetting phenomena on solid surfaces. To solve these questions, further investigation related to the wetting process is needed. In the following section, we will discuss the issues of contact angles to interpret the involved mechanisms. 2.2. Contact angles and their hysteresis Tsujii and his co-workers conducted a series of experiments to test the water repellence on fractal surfaces [3,5052]. One typical process was to prepare fractal surfaces made of alkylketene dimer (AKD) and measure the contact angles. The maximum contact angle reported in their work was 174, as shown in Fig. 4. They argued that the maximum contact angle could be approaching to 180 if there was no adsorption [3]. This was a sparking point in the long run of studying contact angles on superhydrophobic surfaces, for it opened the door of possibility for succeeding researchers to approach the extremely high contact angles in practical elds [26,46,5355]. Afterwards, there indeed were a few reported cases of high contact angles [23,53,5662]. Some researchers even reported contact angles up to 180 [63,64]. Fig. 5 shows a few cases of the extremely high contact angles of liquid droplets on superhydrophobic surfaces. However,

2:4

Besides, if we turn to the ratio of the actual wetted area to the projected area, rf, which is also referred to as the roughness ratio of the fraction, it will give rise to the modied form of the CB equation [44,47]: cos CB rf S cos S 1:

2:5

When S = 1 and rf = r, the CB equation turns into the Wenzel equation. Eq. (2.5) interprets the multilayered roughness and is more suitable for the hierarchical surface structure, which has been found much morphologically closer to the natural model of superhydrophobic surfaces [10,33]. Thus it is conceived that both surface roughness and its structure (morphology) can have impact on the value of contact angles and therefore the states of wetting. If the intrinsic contact angle of a water-repelling surface is xed for a particular material, it will be the way to prepare the surfaces that determines the surfaces' wettability.

Fig. 2. A liquid drop showing contact angle balanced by three interfaces. The letter A indicates the interfaces as well as their contact areas. SV, SL and LV correspond to the interfaces between solid, liquid and vapor respectively.

Fig. 4. A water droplet of which the contact angle is 174 placed on the AKD surface. Image reprinted from Ref. [52], with permission from American Chemical Society, Copyright 1996.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

83

Fig. 5. High contact angles of liquid droplets on articial superhydrophobic surfaces. i: 171.2 1.6, measured on poly(vinyl alcohol) nanobers [57]; ii: 175 0.7, measured on the coating of SiO2 nanoparticles [62]; iii: 180, measured on submicrometered particles of tetrauoroethylene oligomers [63]; iv: 180, measured on uorinated nanoparticles [64]; v: 175180, achieved by using DRIE patterning of controlled parameters [61]; vi: 173( 3, + 6), achieved by using silicon etching and gas phase isotropic etching [53]. Images reprinted with permission from (i) Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2002; (ii) Elsevier, Copyright 2009; (iii) American Chemical Society, Copyright 2007; (iv) Springer Berlin/Heidelberg, Copyright 2008; (v) IEEE, Copyright 2002; and (vi) American Chemical Society, Copyright 2009.

Fig. 6. Ellipse tting (a), circle tting (b), tangent searching (c), and LaplaceYoung tting (d) make static contact angles of the same water droplet different. The gures include the simulation lines of the prole of the water droplets and the horizontal baselines. Images reprinted from Ref. [65], with permission from Royal Society of Chemistry, Copyright 2008.

there was no clear interpretation for the contact angle hysteresis (CAH) in the experiment and theory related to the aforementioned AKD superhydrophobic surfaces. Also, the lack of completely consistent measurement standards might make other high contact angles uncertain to decide. Because of the contact angle hysteresis, the maximum value of a measured contact angle is not the only criterion to dene a stable superhydrophobic state. To describe a superhydrophobic state, the static angle as well as the CAH should be measured both. For an extremely stable superhydrophobic state, its static contact angle should be as high as possible, and its CAH as small as possible, otherwise the achieved wetting state might transit to a different one. Also, the tting modes that people adopted to measure the contact angles can affect their values [65]: it was reported that ellipse tting, circle tting, tangent searching, and LaplaceYoung tting could cause various values of contact angles of similarly shaped droplets as shown in Fig. 6. That's why the liquid drops in Fig. 5 just look similar but have different values. Therefore, the tting mode should be clearly mentioned to reect the real situation of wetting. On the other hand, the CAH can be explained schematically in two manners. First, as shown in Fig. 7, the liquid drop will advance at the lower side and recede at the upper side when the substrate is inclined at D, which is the sliding (tilt [66] or slip [67]) angle. In order to let the liquid droplet slide off, the substrate must be tilted at/over the sliding angle. On this occasion, the liquid wets the substrate at the advancing point and dewets it at the receding point. Thus, A is the advancing contact angle and R is the receding contact angle. They stay constant during the sliding process as long as the surrounding conditions don't change. Secondly, if one withdraws liquid from a droplet on the solid surface, the volume of the droplet as well as the contact angle will decrease, but the contact area of the droplet on the surface will not change until it begins to recede, as shown in Fig. 8(a). Similarly, if one adds water to the droplet, the volume as well as the contact angle will increase, but the contact area will not change until the droplet begins to advance, as shown in Fig. 8(b). Thus, the contact angles during receding and advancing are denoted as receding contact angle R and advancing contact angle A, respectively. The difference between advancing and receding contact angles is termed as the

CAH [18,22]. Corresponding to the description, a typical experimental process of CAH is shown in Fig. 8(c). The contact angle of a metastable droplet formed on a solid surface can be at any value between receding and advancing contact angles. Some CAH is quite low, within 5 [58], which does not affect the superhydrophobicity dramatically. However, some can be as much as 40 [68], which might even change the wetting state. From this point of view, it is not only the static contact angle but also the receding and advancing contact angles that can characterize a surface. 2.3. Surface tension and surface free energy Although the surface tension is referred to as the free energy per unit area, it can also be equally thought of as the force per unit length, which is measured in N/m [69]. The surface free energy is an important characteristic of every surface or interface. In the bulk of the body, chemical bonds exist between the molecules. To break the chemical bonds, a certain amount of energy has to be applied. Molecules that do not form bonds at the side of the surface have higher (potential) energy than those that form the bonds. This additional energy caused by the higher energy is called surface or interface free energy and measured in the energy per area units, that is, in the SI system, J/m 2 [70]. Thermodynamically, the concepts of surface tension, surface tension force, surface energy density, and surface free energy are different [71,72]. However, these concepts, in most cases studying superhydrophobic or superhydrophilic surfaces, are now numerically equivalent to each other when temperature and pressure are assumed constant [73]. To make it clear, it is necessary to assume that there is no adsorption at the interfaces as well. That's the ideal condition for the assumption that the surface tension is equivalent to the surface free energy. For a superhydrophobic surface, the dry area should have lower surface energy than the wet area. And the shape of a droplet on the superhydrophobic surface spontaneously assumes a more spherical shape to minimize its energy. Thus the liquidair surface is promoted. The surface roughness can adjust the effect of surface energy as an amplier. Hence a liquid droplet forms a

84

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 7. (a) A liquid drop theoretically sliding on a declination of D without acceleration. A is the advancing angle, and R is the receding angle. (b) The schematic of static contact angle and dynamic contact angles. Image (b) reprinted from Ref. [67], with permission from American Chemical Society, Copyright 2009.

high contact angle on the rough low-energy surface and a low contact angle on the rough high-energy surface [46].

pinning of liquid droplets. Thus, we continue to discuss the transition based on contact angles between wetting states. 2.4.1. Critical contact angle Bico et al. suggested a critical contact angle c between Wenzel state and CB state by equating Eq. (2.3) with Eq. (2.4) [21]. This critical contact angle was determined by the surface design. That the air was entrapped into the surface grooves should be favored only if was greater than c [19,55]. Moreover, the impact from rf, the roughness ratio of projected protrusions, will need to be included into the critical contact angle for further investigation. Thus, after equating Eq. (2.3) with Eq. (2.5), one can get: cos c
0

2.4. Wetting transitions It is no doubt that those that obtained high contact angles mentioned above should be due to the superhydrophobicity of the surfaces, which can be prompted in quite a few manners (see subsequent sections). However, for most cases, the wetting state has not been presented clearly when a high contact angle is shaped up. As stated previously in Section 2.1, Wenzel and CassieBaxter states might be both suitable for superhydrophobic surfaces; but on the other hand, whether or not the liquid droplet rests partially on the solid surface as well as the air respectively is still under investigation. Also, the CAH in the CB state can be normally lower than in the Wenzel state, and therefore CB state might be more stable [42]. There is an apparent difference between those two states. If the CAH is relatively big, a transition might happen, and this makes the wetting less stable or metastable [5,74]. It is noted from our aforesaid discussion that the advancing contact angle related to a superhydrophobic surface might be greater than 150 and the receding contact angle less than 150. When it comes down to a hydrophobic surface, the advancing contact angle might be greater than 90 and receding contact angle less than 90. Strictly speaking, these states are not 100% superhydrophobic or hydrophobic, as the advancing and receding contact angles can change the wetting states. It should be attributed to the incongruous and inhomogeneous distribution of structural and chemical properties on the surface. That is to say, those factors including structural and chemical properties and even adhesive force can cause the hysteresis of contact angles and

1S rf S r

2:6

' where c is the critical contact angle with the effect of rf [75]. Then the homogeneous wetting will be preferred only if the Young ' contact angle is smaller than c, otherwise the drop will exhibit a heterogeneous wetting state, theoretically, as shown in Fig. 9. However, it has been found in many cases that the CB state can exist in a scenario of moderate contact angles, which means that Wenzel and CB states may even coexist on the same particular surface [47,7686]. Our previous research indicates that how the droplet is delivered or formed also has much to do with the contact angle [87]. As shown in Fig. 10, the contact angle decreased from around 170 to around 145 when the droplet was being pressed to spread on the substrate. Meanwhile, the advancing and receding angles changed with the contact angles. From this point of view, the state of a liquid droplet on a solid surface is related to how the droplet adapts itself with the

Fig. 8. (a) The droplet is pinned at the three-phase contact line until R is reached at stage 2 and R remains constant, if external force and/or energy is not considered, during subsequent volume decreasing. (b) The droplet is pinned at the three-phase contact line until A reaches stage 6. (c) A growing and shrinking drop sequence. Images (a) and (b) reprinted from Ref. [18], with permission from American Chemical Society, Copyright 2006; Image (c) reprinted from Ref. [208], with permission from American Chemical Society, Copyright 2009.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

85

Fig. 11. Filling the liquid up to a rough surface through the solid substrate: (a) homogenous Wenzel fashion; (b) heterogeneous CB fashion. R is the radius of the circular contact area between solid and liquid.

Fig. 9. Wenzel and CB wetting models with impact from rf. Theoretically, Wenzel state: ' ' ' 90 b b c; CB state: c b b 180. However, the CB regime can also be observed for b c, shown as the dotted line after the critical point.

surface. This gives rise to the design of superhydrophobic surfaces because a transition between the two states can cause a signicant change of contact angles as mentioned earlier on. 2.4.2. Energy equations The transitions occur between the wetting states: either from CB state to Wenzel state or from Wenzel state to CB state. We herein discuss the transitions from an energetic point of view. Patankar has supposed that the shape of the droplet in equilibrium state is energetically minimized for a given value of the apparent contact angle [47,54]: G ALV LV cos ASL LV where 8 > cos SV SL eff =LV < 2 > ALV 2R 1 cos : 2 ASL R sin2 :
eff

vapor interface inside the contact area under the drop and ASL should have included the effect caused by surface geometry as well. Thus, we include the surface roughness and the fraction of the protrusions (pillars) to reform the energy equations for heterogeneous and homogeneous wetting states respectively. For a heterogeneous wetting state, the contact areas of liquid vapor and solidliquid interfaces are: ALV 2R2 1 cos 1S R2 sin2 o 0 ASL R2 rf S sin2 :
0

2:10

And thus the energy required for a liquid drop to get its heterogeneous equilibrium state is: GCB ALV LV cos ASL LV :
0 0

2:11

2:7

Subsequently Lagrange Multiplying is needed to minimize the equilibrium energy of the drop under a restricted volume V: F . h G 3 2 V R 1 cos 2 cos 3 LV 2:12

2:8

(SV SL) is assumed as the effective energy per unit area of liquidsubstrate contact (interface) with respect to the dry substrate. is the contact angle of the droplet formed on the substrate surface. G is the energy change from the initial state to the nal state. And for a given apparent contact angle , the energy of the droplet is minimized. The energy change is: G 2=3 1=3 1 cos 2 cos : 9 1=3 V 2=3 LV 2:9

8 > F > > > R 0 > < F > 0 > > . > > : V R3 1 cos 2 2 cos :
3

2:13

Eq. (2.13) can be considered as Eq. (2.12)'s boundary conditions if cos is about to be calculated. As a result, cos cos CB rf S cos S 1:

2:14

From the right side of Eq. (2.9), it's noted that the energy increases with monotonically when 0 * 180. Consequently, a higher apparent contact angle will lead to a higher energy and vice versa. To include the impact from surface roughness, we establish a theoretical setup which is rough with surface protrusions [75], as shown in Fig. 11. It can be noticed that ALV should have covered the liquid

This proves that should spontaneously and nally achieve the CB * apparent contact angle CB , as the form of Eq. (2.14) looks exactly the same to CB equation; Similarly, for a homogeneous wetting state, the contact areas of liquidvapor and solidliquid interfaces are: AW 2R2 1 cos LV AW R2 r sin2 : SL o 2:15

The energy required to get its homogeneous equilibrium state is: G ALV LV cos ASL LV and consequently we get cos cos W r cos :
W W

2:16

2:17

Fig. 10. Contact angles varying from around 170 to less than 145. Image reprinted from Ref. [87], with permission from Elsevier, Copyright 2010.

Thus, in a homogeneous wetting state, the liquid drop will spontaneously form a Wenzel contact angle to minimize the energy. We

86

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

now have theoretically proved that, from an energetic point of view, the contact angle should form either the Wenzel contact angle W * or the CB contact angle CB to minimize its energy state. Substituting * by CB and * , the energy required for a drop in heterogeneous * W equilibrium and homogeneous equilibrium should be GCB and GW respectively:
2 2 GCB R 1 cos CB 2 cos CB LV 2=3 1=3 2=3 1=3 3V 1 cos CB 2 cos CB LV 2 2 GW R 1 cos W 2 cos W LV 1=3 2=3 2=3 1=3 3V 1 cos W 2 cos W LV :

3. Superhydrophobic surfaces As explained by Wenzel and CB equations, surface roughness and its structure play a very important role in promoting extremely high contact angles. Also, thanks to the ndings related to the surface of lotus leaves, people have started the bio-inspired research to understand and design those water repelling surfaces. The effort includes exploring natural surfaces which have complex roughness and significant hydrophobicity. This type of research work is extended in not only plants but also animals at the moment (see the following sections). Also, with the knowledge and technologies being developed, it is possible to articially produce superhydrophobic surfaces based on biomimetic approaches, which refer to learning from nature [36,81,88,89]. The research work considers both the strategies of designing surface patterns and the methodologies of processing materials. In this section, we will review the recent achievement that has been made to investigate natural and manmade surfaces with superhydrophobic properties. 3.1. Superhydrophobic surfaces in nature

2:18

2:19

By combining Eqs. (2.18) and (2.19), we nally get: G


1=3

3V

2=3

1 cos

2=3

2 cos

1=3

LV :

2:20

This is consistent with previous research reports [44,54]. The surface roughness ratio and the fraction of the projected areas are included. Thus, the corresponding equations to both homogeneous and heterogeneous fashions are improved, and Eq. (2.20) can be rearranged to a dimensionless form as G 1=3 2=3 1=3 1 cos 2 cos : 3V 2=3 LV

2:21

Therefore, it is understood that the energy change from the initial state to the eventual state is a monotonic increasing quantity of * (90 b * b 180) for a given volume V. As the transition happens, the shape of the drop may change between CB and Wenzel states. The energy difference G between the CB and Wenzel states is: G jGCB GW j: 2:22

Also, because of the CAH, the energy change should consider the maximum and minimum values of contact angles. To trigger the transition between the wetting states, the energy added onto the drop must overcome this energy difference. If we dene an energy barrier GB between the wetting states, it should include many issues that might not be measurable or controllable, particularly in small scale. Although the transition between those two wetting states is not fully understood yet, this energy barrier GB should be higher than either GWorGCB. Based on the discussion, it is noted that working on the detail of surface structure is important to precisely prompt the wetting state of a liquid droplet. Therefore, the approaches that have been applied to create and modify the micro- and nanoscaled patterns for superhydrophobic surfaces will be discussed in the following section.

The lotus ower is considered as a symbol of purity in a few religions. Its leaves can be kept from contamination or pollution without being folded even when the lotus is around with muddy water. This phenomenon illustrates that nature can protect itself from omnipresent dirt and pathogenic organisms. And ideally, if this property is applied to functional surfaces, self-cleaning effect can be prompted in almost any materials in the open air by rain water. As mostly accepted, Barthlott and Neinhuis' early work has started the recent research of superhydrophobic surfaces inspired by nature [32,90]. Surface roughness has then been recalled to explain the surface's extreme repellence against liquid droplets, as shown in Fig. 12(a). Subsequent and further research has indicated that the plant cuticle is technically a composite material mainly built up by a network of cutin and hydrophobic waxes, where surface structuring arises at different hierarchical levels [27]. The composite or hierarchical surface structure, formed by a combination of two (or even more) layers in different sizes, is built by convex cells and a much smaller superimposed layer of hydrophobic three-dimensional wax tubules [33], as shown in Fig. 12(b and c) with different magnications. It has been argued that wetting of such surfaces is minimized, because air is enclosed in the cavities of convex cell sculptures. Along with the lotus leaves, there are other natural superhydrophobic surfaces in the plant kingdom. For example, taro (Colocasia esculenta) leaves have been used to demonstrate the self-cleaning effect in the original paper [32]. It has been observed that the elliptic protrusions with an average diameter of about 10 m form the microstructure on the taro leaf, and the nanoscaled pins form the hierarchical structure with the microstructure, which is however similar to the lotus leaf [91], as shown in Fig. 13(ac). Our SEM images of taro leaves shown in Fig. 13(b' and c') have avoided the shrinkage by using critical-point-drying techniques. They are more conformal to the real structure on the taro leaves. Similarly, India canna (Canna

Fig. 12. (a) A glycerol drop on Euphorbia myrsinites, which is a robust specimen and well suited to show the surface's repellence against the liquid droplet. Scale bar = 80 m. (b) The upper side surface of the lotus leaf without the shrinkage artifact. Scale bar = 8 m. (c) The wax tubules from the upper side of the lotus leaf. Scale bar = 1 m.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

87

Fig. 13. (a) A few water droplets on a taro leaf. (b) and (c) show the SEM images of a taro leaf with different magnications. The inset of (c) is a water droplet on a taro leaf, with a contact angle of 159 2. The scale bars of (b) and (c) are 20 m and 5 m, respectively. (b') and (c') correspond to our SEM images using critical-point-drying techniques on observing taro leaves when shrinkage is avoided. The scale bars of (b') and (c') are 10 m and 2 m, respectively. (d) Water droplets oating on an India canna leaf. (e) and (f) show the SEM images of an India canna leaf with different magnications. The inset of (f) is a water droplet on an India canna leaf, with a contact angle of 165 2. The scale bars of (e) and (f) are 50 m and 1 m, respectively. (g) A water droplet oating on a rice leaf (Oryza sativa L.). (h) and (i) show the SEM images of the rice leaf with different magnications. The inset of (i) is a water droplet on a rice leaf with a contact angle of 157 2. The scale bars of (h) and (i) are 50 m and 1 m, respectively. Images, except for (b') and (c'), reprinted from Ref. [91], with permission from Elsevier, copyright 2007.

generalis bailey) leaves and the rice leaves (whatever the kind of rice) have been reported to be superhydrophobic and have binary structures (micro- and nanostructures) on the surface [91], as shown in Fig. 13[(df) and (gi)]. These multi-layers have positive importance to make up the wax crystals. The dimensions of wax structures range from a few nanometers to several micrometers, and are responsible for the maintenance of wettability [33,92]. Lotus leaves, India canna leaves and taro leaves form tubule-like and platelet-like wax crystals, which are the most prominent types [93] of wax structures. Also, the arrangement of surface structure can have great inuence on wettability of solid materials. It is pointed out that the water droplets can roll off a rice leaf easily along the direction of the arrow, shown in Fig. 13(g), at a sliding angle of 4, but move much harder along the perpendicular direction at a sliding angle of 12 [91]. Thus, it is clear that the anisotropic arrangement of the protrusions inuences a rice leaf very much. However, due to the homogeneous arrangement of surface protrusions, the sliding angle of a lotus leaf can be down to below 2 [94]. Also, as recently realized, a colocasia leaf shows superhydrophobic characteristics but has different structure from the lotus leaf [1,95]. The surface structure of a colocasia leaf has similar bumps to lotus, but the bumps are separated by the surrounding ridges. The bump and ridge both contribute to the hydrophobic nature of colocasia since they might create air pockets between the water droplet and the surface. At this point, it is necessary to emphasize that the surface cuticle and its waxes are important to surface wettability by folding the cuticle or by forming three-dimensional wax crystals on the plant surface [33]. Thus, it has been brought into consideration that some natural surfaces with hierarchical structure and roughness can produce signicant superhydrophobicity, not only from plants but also animals. It has been revealed that the leg of a water strider [Fig. 14(a)] has numerous oriented needle-shaped setae [Fig. 14(b)] with their diameters ranging from

3 m down to several hundred nanometers [28,96]. Many elaborate nanoscaled grooves are noticeable on each microseta [Fig. 14(c)], forming a hierarchical structure, which origins the superhydrophobicity of the water strider's legs with assistance of the hydrophobic secreted wax. On the other hand, the scales on the wing surfaces of some butteries have regularly arranged edges which are overlapping like roof tiles [34,97], as shown in Fig. 14(df). The lengths and widths of each individual scales are roughly ranged 50150 m and 3570 m, respectively, while the primary distance between the middle points on the long axis of two adjacent scales is within 100 m. However, not every hierarchical surface in nature has been found with superhydrophobic behaviors. A gecko has hundreds of thousands of keratinous hairs or setae on its foot [98], as shown in Fig. 14(gi); each seta is around 30130 m long and contains hundreds of submicron spatulae, forming hierarchical morphologies [37]. This type of hierarchical structure does not necessarily have superhydrophobic properties, although it is still employable as a model for comparison. This indicates again the importance of the pattern of surface structures in wettability. But the investigation into superhydrophobicity by means of biomimetics has not been discouraged. Phenomenally, Byun et al. selected 10 orders and 24 species of insects to characterize the functions of natural structures by focusing on both lower and upper surfaces of the wings [35]. They argued that the hierarchical architectures which included micro- and nanoscaled layers on the upper surfaces of insect wings (Fig. 15) promoted hydrophobicity, thereby enabling water droplets to roll off the wings and remove the dirt particles. It is already becoming vivid from the discussions mentioned previously that the mechanisms of individual superhydrophobic surfaces are not necessarily the same with each other in every aspect. Different from the lotus effect, the so-called petal effect describes the phenomenon that a water droplet on the petal surface of a red rose (rosea Rehd) forms a spherical shape, but does not roll off even when the

88

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 14. Natural hierarchical surfaces. (a) A water strider (Gerris remigis) standing on the water surface. Scanning electron microscope (SEM) images of a strider leg showing numerous oriented spindly microsetae (b) and the ne nanogrooved structures on a seta (c). Scale bars: 20 m (b), and 200 nm (c); (d): schematic of a buttery (Pontia daplidice). SEM images show the transection (e) of the buttery wing surface and its at arranging (f); (g): A gecko (Gekko gecko) with emphasis on its toe. A toe of gecko contains hundreds of thousands of setae and each seta contains hundreds of spatulae. (h) and (i): SEM images of rows of setae at different magnications. ST: seta; BR: branch. Image (a) reprinted from Ref. [28], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2006; Images (b) and (c) from Ref. [96], with permission from Nature Publishing Group, Copyright 2004; Images (d), (e) and (f) from Ref. [88], with permission from Elsevier, Copyright 2009; Images (g), (h) and (i) from Ref. [98], with permission from Elsevier, Copyright 2005.

petal is turned upside down [13], as shown in Fig. 16. The diverse design in surface microstructure on rose petal results in different dynamic wetting from lotus leaves. When a small water drop wets the petal surface, the liquid lm impregnates the textured regime. However, the liquid only wets the grooves between the projected pillars, leaving the plateaus dry, which forms the Cassie impregnating state, as shown in Fig. 17. For the petal surfaces, the dimensions of hierarchical micro- and nanostructures both are found larger than those

related to the lotus leaf. In the Cassie impregnating wetting regime, water droplets enter into the large grooves of the petal but not into the small ones. This indicates why small water drops sealed in micropapillae would be clinched to the petal's surface, showing a high CAH even if the surface was turned upside down. Thus, the mechanisms for different natural superhydrophobic surfaces and their wetting states might seem vague to follow. However, the common characteristic of superhydrophobic surfaces rests on the

Fig. 15. SEM images of the micro/nano structures and measured contact angles on upper wing surfaces of insects. (a) Homoptera Meimuna opalifera (Walker), showing contact angle of 165; (b) Orthoptera Acrida cinerea cinerea (Thunberg), showing contact angle of 151; (c) Hymenoptera Vespa dybowskii (Andre), showing contact angle 126; (d) Diptera Tabanus chrysurus (Loew), showing contact angle of 156. Vespa dybowskii, Acrida cinerea cinerea, and Tabanus chrysurus show micro- and nanoscaled hierarchical structures. Although Homoptera Meimuna opalifera (Walker) does not have such hierarchical structures, the wing shows superhydrophobicity due to its very ne nanoscaled structure. Images reprinted from Ref. [35], with permission from Elsevier, Copyright 2009.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

89

Fig. 16. SEM images (a, b) of the surface of a red rose petal, showing a periodic array of micropapillae and nanofolds on each papillae top. (c) A water droplet on the petal's surface, indicating its superhydrophobicity with a contact angle of 152.4. (d) Prole of a water droplet on the petal surface when turned upside down. Images reprinted from Ref. [13], with permission from American Chemical Society, Copyright 2008.

congenially periodical structures that are hierarchically organized in micro- and nanoscale. In particular, the natural models that are related or similar to lotus effect have been mostly used due to their prominent features. Following this, we continue to review the recently published work that successfully manufactured superhydrophobic surfaces with typical features. 3.2. Articial superhydrophobic surfaces It is noticeable that many of those natural superhydrophobic surfaces discussed above have intense roughness and complex structure in micro- and nanoscales. Some symbolic surfaces have been utilized as the primary models for researchers to mimic superhydrophobicity. For example, a combination of poly(vinyl chloride) (PVC) and ethanol has been used to form superhydrophobic lms [99], as shown in Fig. 18. The formed PVC surfaces become rougher with the increase of ethanol content (under 50%) in the PVC solution, and more pores and nanocomposites are formed. These superhydrophobic surfaces form

lotus-leaf-like structures consisting of many nanoparticles with the sizes ranging from 100 nm to 300 nm. Also, as pointed out previously, typical superhydrophobic surfaces discovered in the nature have shown multiscaled roughness consisting of nanometer sized akes on top of micrometer sized protrusions. This type of morphology for superhydrophobic surfaces can be achieved using lithographic methods, template-based techniques, plasma treatment, self-assembly and self-organization, chemical deposition, layer-by-layer (LBL) deposition, colloidal assembly, phase separation, and electrospinning. Most of these methods are in accordance with the theoretical and natural models discussed in previous sections, and one strategy commonly used to invoke superhydrophobicity for many processes is producing rough surfaces with hierarchical structure and low surface energy. The following section will discuss those methods as well as the manmade superhydrophobic surfaces. As some of the previous research has already been discussed in Refs [26] and [100], we will be concentrating on most up-to-date articles that have been recently published over the past several years. 3.2.1. Imprinting and templating The imprinting techniques to produce superhydrophobic surfaces usually involve lithography, templating, and plasma treatment. The common feature of lithography and templating is embodied that they usually need a master piece and a replica substrate, while a plasma treatment is somehow related to surface etching. Those technologies are not necessarily independent on each other, but to the contrary, can be used as a combination to realize a superhydrophobic surface. Theoretically, a surface processed by lithography can be treated further by plasma etching to achieve complex roughness with multilayers and then used as a featured template to produce replicas. However, the main principle is to make the process as uncomplicated as possible for the creation of a superhydrophobic surface. We herein will investigate the technical principles and characteristics of those methods, respectively. 4. Lithography

Fig. 17. Schematic illustrations of a drop of water in contact with the petal of a red rose (the Cassie impregnating wetting state) and a lotus leaf (the Cassie's state). Image reprinted from Ref. [13], with permission from American Chemical Society, Copyright 2008.

The lithographic process is a well-established technique and its subtechniques used in making superhydrophobic surfaces include optical lithography (photolithography) [53,59,67,101,102], soft lithography

90

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 18. SEM images of PVC surfaces obtained with the different ethanol content in the PVC solution: (a) 0% (v/v); (b) 16.7% (v/v); (c) 37.5% (v/v); (d) 44.4% (v/v); (e) 50% (v/v); (f) is the higher magnication SEM image of (e). Insets are the water contact angles on the as-prepared PVC surfaces. Images reprinted from Ref. [99], with permission from Springer Science + Business Media, LLC, Copyright 2008.

[56,103107], nanoimprint lithography [94,108], electron beam lithography [24,109], X-ray lithography [92], and colloidal lithography [110]. On a general level, this sort of method prepares superhydrophobic surfaces by copying the information from a master, which can be either rigid or soft, and then transferring it to a replica in an opposite form. It should be noted that these branches of lithography are not rigidly or denitely distinguished. For instance, the superhydrophobic surfaces obtained by using nanolithography [94,108] also utilized the optical technique in processing the master piece. The colloidal lithography [110] and optical lithography [102] may involve nanolithography as well. And this applies to the soft lithography too [107]. Conventional lithography normally uses a at plate as the master surface. Now the lithography is often referred to as photolithography, which transfers the entire geometric pattern from an optical mask to a thin lm of photoresist sensitive to light (mostly ultraviolet), for its powerful effect and common usage. For the soft lithography, the key elements include elastomeric stamps, molds, and/or mask, which are utilized to generate patterns and structures in micro- and nanoscaled size [111,112]. And recently, this type of lithography has been advanced rapidly [113]. The whole process can also involve self-assembling as well as the replica molding. The photo reactive lm is not essential for the soft lithography to create the micro- and nanostructures, although it can still be adopted. The electron beam lithography uses focused electron beams rather than scattered light to create the surface patterns, and the mask is not necessarily used any more [114]. X-ray beam lithography also needs mask and resist, which means it is similar to photolithography in terms of working procedures, while the main difference rests on the source of radiation [115]. X-ray radiation is selectively absorbed by the mask and thus a pattern can be formed. The so-called colloidal lithography for making superhydrophobic surfaces is relatively new and still under development. One of the applications has used a monolayer of silica colloidal spheres as the mask, which is followed by reactive ion etching (RIE) to fabricate superhydrophobic surfaces [110]. The optical lithography is a frequently used technique, and it is always coupled with other lithographic methods to create surface patterns. Jeong et al. fabricated micro- and nanoscaled hierarchical structures using UV-assisted capillary force lithography [68], which was based on the sequential application of engraved polymer molds followed by a surface treatment. The fabrication procedures of the dualscale structure were to generate microstructures rst by using

microscaled poly(urethane acrylate) (PUA) or poly(dimethyl siloxane) (PDMS) mold and then nanostructures on the top by using nanoscaled PUA mold, respectively. Partial ultraviolet (UV) curing was used in the rst step to form micropillars (microposts), while complete UV curing was used in the second step to form nanopillars. To induce surface hydrophobicity, the fabricated hierarchical surface was treated with a trichloro (1H,1H,2H,2H-peruorooctyl) silane in vapor phase. Fig. 19 shows the SEM images of representative hierarchical structures. The substrates made of poly(ethylene terephthalate) (PET), silicon or glass could be equally used during the process. As a recently developed combination of nanoprocessing and lithography, the nanoimprint lithography (NIL) can create patterns more precisely down to nanoscale. Lee and Kwon's team conducted a series of experiments to fabricate superhydrophobic surfaces by using nanoimprint lithography [94,116]. They designed a set of UVNIL equipment and fabricated photopolymer replicas. The whole replication was performed in a clean environment and composed of two processing steps, making nickel mold by electroforming [Fig. 20(a)] and making replication by polymer casting [Fig. 20(b)] or UV-NIL [Fig. 20(c)]. The nickel mold was directly patterned upon a lotus leaf, and therefore the topographic feature and contact angle of produced replicas from polymer casting and UV-NIL both showed close similarity to the original lotus leaf. The nanoimprint lithography discussed above was based on a combination of nanoreplication and optical lithography. The photopolymer performed as a photoresist and there was no photo mask used during the lithographic process. This decreased the number of processing steps and still secured the superhydrophobicity caused by optics-aided NIL. It was an example that nanotechniques and lithography were involved in the same process of making superhydrophobic surfaces. In modern times, the subdivisions of lithography are not completely isolated with each other anymore. It is possible to combine the advantages of those lithographic techniques to make superhydrophobic surfaces. 5. Templating Template-based methods are another imprint-related way to prepare superhydrophobic surfaces [117]. To develop the superhydrophobic surface pattern, templating can be involved with lithography [56,107,108,110,118,119]. Thus it is plausible that the templates can

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

91

Fig. 19. Tilted SEM images of typical dual-scale hierarchical structures. (ac) For microstructures, 30 m posts of 40 m in spacing and 50 m in height were used. Magnied images in (b and c) show well-dened 400 nm dots (spacing of 800 nm, height of 500 nm) formed on the microposts. (df) For microstructures, high aspect ratio micropillars (diameter of 20 m, height of 100 m and spacing of 20 m) were used. Magnied images in (e and f) show the formation of 400 nm pillars (spacing of 400 nm, height of 2.5 m) on top of the asformed micropillars. Images reprinted from Ref. [68], with permission from Elsevier, Copyright 2009.

be fabricated by using lithography. Also, templating is used to assist other methods preparing superhydrophobic surfaces [120]. The original prototypes of the templates can be lter paper [121], insect wings [122], reptile skins [123], and plant leaves [124]. From a chemical and morphological point of view, the template can even be molecules [117] and polymers [125].

Generally speaking, the templating process includes preparing a featured template master, then molding the replica and nally removing the templates. One typical process was the fabrication of superhydrophobic surfaces by mimicking the hairy structure of gecko's feet [123]. The gecko-mimetic nanopillar structures of hard-PDMS were fabricated by using nanoporous anodic aluminum

Fig. 20. Schematic representation of the fabrication process of superhydrophobic surfaces composed of two steps. Nickel mold making (a), followed by replication using polymer casting (b) or UV-NIL (c). Images reprinted from Ref. [94], with permission from IOP Publishing Ltd, Copyright 2007.

92

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

oxide (AAO) membranes as templates. The aluminum plate was initially anodized to remove electrically oxidized aluminum layer, and further anodized to produce the nanoporous AAO template with cylindrical pore channels. Afterwards, h-PDMS precursors were spincoated on the fabricated AAO template. The sample was cured after conformal contact with the vinyl-terminated glass, and the molded h-PDMS lm was peeled off from the AAO template. The general process is shown in Fig. 21. The length of cylindrical pore channels of AAO templates was controlled by varying anodization time at the second anodization step, and was optimized to make the superhydrophobic h-PDMS surface, as shown in Fig. 22. Thus, it is not arduous to notice that the templating-based methods are effective in creating regular surface patterns, even though the structures need to be relatively complex. A high-quality template can be repeatedly used to make a large number of samples. However, when it comes to micro- and nanoscaled structures, the operators are required to be essentially careful to avoid damaging both the samples and the templates. Also, the templating methods might

not be t to produce irregular surface patterns with excessively complex structure, due to the peeling off procedure. 6. Plasma treatment In quite a few cases, the surface pattern has been formed by the plasma treatment after lithography or templating methods [110,126,127]. However, the plasma treatment can also be used before lithography or templating [128,129], and sometimes the plasma treatment and lithography can even alternate with each other during the surface processing [130]. Moreover, as mentioned earlier, the plasma treatment is somehow connected to etching techniques for preparing superhydrophobic surfaces. Kim et al. fabricated superhydrophobic surfaces which had sharp tips at nanoscale by using deep reactive ion etching (DRIE) [61], which was a developed form of the RIE. A typical RIE process for prompting superhydrophobicity uses chemically reactive ions from the plasma which is generated by an electromagnetic eld. It creates surface patterns by removing material from the wafer surface, while the DRIE gives deep features to the surfaces [67,130132]. Kwon et al. formed microscaled textures on a silicon surface using deep reactive etching and nanoscaled textures on the microscaled textures using XeF2 etching, as shown in Fig. 23 [53]. The XeF2 etching could provide uniform roughness at nanoscale on the surface, but the etching time should be carefully controlled to avoid over-etching. The etched surface was coated with a heptadecauoro-1,1,2,2-tetrahydrodecyl trichlorosilane (HDFS) self-assembled monolayer (SAM) to prompt large intrinsic contact angles and therefore the CB superhydrophobicity. In the droplet bouncing test, the droplet that landed on the surface with double-layered roughness would always nd the CB state energetically favorable, and it could bounce back from the surface (Fig. 24), showing great superhydrophobicity. Zhang et al. reported the fabrication of silicon cone arrays (SCAs) by using colloidal lithography and RIE to form micro- or nanoscaled surfaces [110]. Silica colloidal spheres were transferred onto PDMS lm by using soft lithography and subsequently onto the silicon substrate, which was spin-coated with a thin lm (~200 nm) of poly (vinyl alcohol) (PVA, Mw = 71,000), by using the printing technique. RIE was performed with a gas mixture of CHF3 at 30 sccm (standard cubic centimeters per minute) and SF6 at 4 sccm. Thus, the silicon microstructure arrays were obtained, as shown in Fig. 25. The maximum contact angle on the hierarchical arrays was 158.5, and the rough structure formed greater contact angles than the smooth arrays for most of the cases. This proved that the CassieBaxter's theory related to hierarchical structures to some degree. However, in some other cases, the contact angles on smooth arrays were greater than that on the hierarchical arrays, which meant a Wenzel state might be preferred for some reason. CF4 RIE was used to fabricate polymeric superhydrophobic surfaces which were three-dimensionally featured by prism holographic lithography in advance [133]. The adhesion force for water on superhydrophobic surfaces originated from gecko surfaces was normalized to be 0.01 N/cm 2, which was claimed as the highest yet reported for an articial superhydrophobic surface at that moment. A superhydrophobic hierarchical surface (Fig. 26) consisting of nanopillars superposed on microbumps was prepared by using a combination of photolithography and RIE [134]. The static contact angle of a water droplet could reach up to 160 on the articial surface. Argon plasma treatment on a superhydrophobic biodegradable polyester substrate was used to control the wettability and to produce lms with gradient of wettability [135]. A high-density microwave plasma system was used as a dry etching technique on PDMS to prepare superhydrophobic surfaces [136]. Other plasma treatments that were involved to create superhydrophobic surfaces also include plasma polymerization [126,137,138], plasma electrospray [139], plasma-enhanced chemical vapor deposition or its combination with plasma etching [140143], and plasma immersion ion implantation (PIII) [144]. From the discussion above, it can be seen that plasma treatment is applicable to the

Fig. 21. Procedure for the fabrication of gecko-mimetic h-PDMS nanopillar lms on the vinyl-terminated glass substrate. Image reprinted from Ref. [123], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2008.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

93

Fig. 22. FE-SEM images of AAO templates anodized at the second anodization for 10 min (a) and 20 min (b), and their h-PDMS replicas. The bottom shows a static water contact angle on each replica surface. The scale bar is 500 nm. The h-PDMS nanopillars stretch out from the hexagon-like arrangements of hemispherical convexes. The nanopillar structure is conrmed by the tilted view of FE-SEM. Image reprinted from Ref. [123], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2008.

production of regular superhydrophobic surfaces on a variety of substrates and materials. It can either be adopted as the only technique or coupled with other techniques as a combination to create surface patterns. The source gasses are required to react well with the materials, otherwise the surface patterns would not be realized properly. Therefore, the reaction conditions have to be controlled strictly, which means it requires a lot of physical effort to produce surface patterns with extremely complex roughness on large scaled surfaces via plasma treatment. 6.1. Chemical deposition Superhydrophobic surfaces can also be produced by chemically depositing a thin lm on the selected substrates. Typical chemical deposition methods include chemical vapor deposition (CVD), electrochemical deposition, and LBL deposition. A CVD process normally refers to exposing the selected substrate to a gaseous precursor to deposit the desired lm or powder, and chemical reactions are involved during this process. The electrochemical deposition process however can be used to deposit lms of solid metal or its oxide onto electrically conductive substrates. LBL deposition is another versatile way to prepare superhydrophobic surfaces with multilayered lms. In the following section, we will review several typical cases of preparing superhydrophobic surfaces based on chemically depositing methods.

7. CVD-based surface treatment The plasma-enhanced chemical vapor deposition (PECVD) has recently received more attention and already been used to fabricate superhydrophobic surfaces. Phenomenally, superhydrophobic surfaces with supported Ag/TiO2 core-shell nanobers were prepared at low temperature by PECVD [145]. The bers were formed by an inner nanocrystalline silver thread covered by a TiO2 overlayer. Water contact angles depended on both the width of the bers and their surface concentrations, and a maximum contact angle close to 180 was reached. As materials for reference, TiO2 was simultaneously deposited on at Si substrates. Fig. 27 shows a series of planar view SEM images corresponding to Ag/TiO2 and TiO2 surfaces, respectively. The water contact angles' varying with the width of the bers was attributed to Wenzel's theory that surface roughness would have an obvious effect on contact angles. When irradiated with UV light, the superhydrophobic surfaces became superhydrophilic. The decrease rate of contact angles depended on both the crystalline state of the titania and the size of the individual TiO2 domains. This was one example that a superhydrophobic surface could transform reversibly into a superhydrophilic one triggered by light irradiation. Multiwalled carbon nanotubes (CNTs) with considerable mechanical strength were fabricated by using catalyst-assisted chemical vapor deposition (CCVD) [107]. The multiwalled CNT composites were

94

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 23. Droplet on the fabricated nano-microroughened hierarchical surface. (ac) Nanoscaled roughness etched by XeF2 gas that conformally covers the microscale array of pillars fabricated through deep reactive etching. (de) Droplet sitting on the double roughness with the value of the pillar spacing to width ratio at 7.5, supported by only several pillars. The contact angle at this state is 173 (+6,3). Images reprinted from Ref. [53], with permission from American Chemical Society, Copyright 2009.

deposited on microstructured Si surface with pillars using soft lithography. And the at epoxy resin was deposited using a spraying method. After spraying the CNTs on the surfaces, the CNT composite structures (Fig. 28) were then annealed in order to improve the mechanical properties. For comparison's sake, hierarchical structures at nanoscale were also created by depositing self-assembled lotus waxes. After introducing CNT nanostructure on top of the micropatterned Si replica, the hierarchical structure with CNTs showed a static contact angle of 170 and a CAH of 2. On the other hand, the highest static contact angle of 173 and lowest CAH of 1 were found on hierarchical structure with lotus wax. This meant that the CNT hierarchical structure and lotus wax hierarchical structure had more or less similar superhydrophobicity. But the waterfall/jet tests showed that CNT composite structures had a good stability of wetting properties not only from long-term exposure to water but also high water pressure. The nanostructure with lotus wax, however, could be damaged by water with high pressure, resulting in the loss of superhydrophobicity.

Besides, similar approaches in the production of the surface patterns with complex roughness were reported a lot. Superhydrophobic CNT forests were obtained using PECVD [60]. Superhydrophobic properties were induced on paper surfaces via a combination of selective etching by an oxygen plasma and deposition of a uorocarbon lm from pentauoroethane (PFE) via PECVD [140]. Microwave-plasma enhanced chemical vapor deposition (MPECVD) from trimethylmethoxysilane was carried out to produce superhydrophobic micropatterns [141]. A superhydrophobic lm with high hardness and transparency was created through MPECVD [146]. PECVD of ribbonlike uorocarbon lms from modulated C2F4 was used to induce nanostructured superhydrophobic surfaces [143]. As a matter of fact, the CVD-based methods were often used to produce superhydrophobic surfaces based on their feasibility in maintaining and enhancing the mechanical strength of the formed structure. These methods are able to create complex surface structures at nanoscale, which means extreme wettability can thus be induced. However, due to the relatively extensive operation procedures, it is difcult to control precisely the detail at each individual spot of the surface. 7.1. Layer-by-layer (LBL) deposition The fundamental principle of the LBL deposition rests on changing the charge of a substrate, and this allows the construction of multilayered lms [147]. The process normally involves the assembling of layers with spontaneous adsorptions, while the size of the substrate can be in a wide range. It does not necessarily need to prepare the master for replication like imprinting methods or provide a particular environmental chamber like plasma treatment and CVD. Therefore, this type of method is relatively facile and potentially economical. To achieve the surface roughness for a superhydrophobic state, nanoparticles are often added into the deposition [148]. This, however, will put more perplexities for the precise control of surface patterns. Cohen and Rubner's team performed a series of research work using the LBL method to produce functional surfaces [149153].

Fig. 24. Droplet (liquid volume = 10 L) bouncing on the double-roughness surface. This gure was a sequential shot of the droplet bouncing on the superhydrophobic surface. A high-speed camera (MotionPro HS-4) was used to take the video image at a rate of 5000 frames per second. Image reprinted from Ref. [53], with permission from American Chemical Society, Copyright 2009.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

95

Fig. 25. Cross sectional SEM images of morphologies of silicon microstructure arrays obtained after etching for 0 s (a), 70 s (b), 140 s (c), 175 s (d), 210 s (e), and 300 s (f). Insets are the top-view SEM images of the relevant ordered structure arrays. The average diameter of silica sphere used as mask is 1115 nm. The structure height from bf is approximately 575, 1075, 1351, 1431, and 1086 nm, respectively. Scale bars are 1 m. Images reprinted from Ref. [110], with permission from American Chemical Society, Copyright 2009.

Typically, they utilized LBL processing schemes to create transparent superhydrophobic lms containing SiO2 nanoparticles with various sizes [151]. The deposition process consisted of immersing the substrates repetitively into different aqueous solutions to form three main parts of the lm, which were adhesion, body and top layers, as shown in Fig. 29. Briey, the LBL procedures were rstly dipping the substrate into the cationic solution and then dipping the substrate into the anionic solution with proper assistant operation. Poly(allylamine hydrochloride) (PAH) was used to create the cationic solution and poly(sodium 4-styrenesulfonate) (SPS) to create the anionic solution for the adhesion layers, which could enhance the binding of the polymernanoparticle system. PAH was used to create the cationic solution and two differently sized silica nanoparticles (50 and 20 nm) were respectively used to create the anionic solution for the body layers. The cationic solution used for the top layers was PAH and the anionic solution was distilled

water solution containing 20-nm silica particles. Advancing contact angles up to 160 with the CAH as small as 10 were obtained on the optimized multilayered thin lm, and the optical transmission levels were preserved above 90%. On the other hand, Zhai's group fabricated a switchable superhydrophobic/superhydrophilic surface by using similar LBL techniques [154]. Uniform and conformal multilayered lms of PAH and silica nanoparticles were built upon the surface with a precise control over lm thickness and roughness. The surface was further functionalized with poly(N-isopropylacryamide) (PNIPAAm), which was a thermoresponsive polymer exhibiting low critical solution temperature in an aqueous solution, and a low surface energy material(1H,1H,2H,2H-peruorooctyl) silane (peruorosilane). The combination of surface roughness and chemical composition created switchable superhydrophobic/superhydrophilic surfaces, which

96

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

could be potentially applied as hydrophobic valves into microuidic channels. Jindasuwan et al. reported another example of producing superhydrophobic surfaces by mimicking the structure features of the lotus leaf [62]. In their work, superhydrophobic lms were dipcoated onto glass substrates using a LBL deposition of polyelectrolyte. Afterwards, SiO2 nanoparticles were deposited to modify the surface roughness, and nally a semiuorinated silane was applied to lower down the surface energy. 7.2. Colloidal assembling and aggregation The assembling or aggregation of colloidal particles can be used to assist chemical deposition, solgel process, and lithography to make the micro- or nanostructures for superhydrophobic surfaces [120,122,151,155160]. The colloidal assembly itself is the process based on forming assemblies of monodispersed particles through chemical bonding or van der Waals forces. If the sizes of the assembling particles are well controlled and have different scales, multilayered

Fig. 26. SEM photo of a hierarchical surface made by lithography and RIE. The inset photo shows a water droplet with a static contact angle of 160 on this articial surface. Image reprinted from Ref. [134], with permission from American Institute of Physics, Copyright 2009.

Fig. 27. SEM micrographs of the Ag/TiO2 composites (left) and TiO2 (right) surfaces prepared by PECVD on respectively a silver membrane and a at Si substrate. (a) and (b), deposition at 403 K, bers are now formed. (c) and (d), deposition at 523 K. (e) and (f), deposition at 298 K, no bers are formed. The insets show the images of the water droplets that formed on the surfaces of different samples during water contact angle measurements. Images reprinted from Ref. [145], with permission from American Chemical Society, Copyright 2008.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

97

Fig. 28. SEM micrographs taken at 45 tilt angle, show three magnications of nano- and hierarchical structures fabricated with CNTs after 3 h at 120 C. Images reprinted from Ref. [107], with permission from American Chemical Society, Copyright 2009.

roughness can be obtained [161]. The multilayer is normally formed by immersing the substrate into the solution of particles or spin-coating the solution onto the substrate [162,163]. For instance, Min et al. synthesized monodispersed silica particles (~70 nm) which could be spin-coated over a large area [164]. Fig. 30(a) shows a typical transmission electron microscope (TEM) image of the as-made nanoparticles. The diameter of the spheres was calculated to be 72 6 nm using the image analysis software (Scion Image). Fig. 30(b) shows a topview SEM image of the multilayered crystal structure prepared by spin-coating at 10,000 rpm for 10 min. Fig. 30(c) shows the spin-coated colloidal crystal which was non-close-packed. The spheres of the top layer lled in the triangularly arranged crevices made by the spheres of the second layer. On this occasion, the spheres were separately distributed in the same single layer. The thickness of the colloidal crystals could be controlled by adjusting speed and time of spin-coating. The case of achieving superhydrophobic states discussed right above actually had much to do with multilayered colloidal crystals based on the self-assembly techniques. Directed by internal interactions or forces among the colloidal particles, a self-assembly method forms the patterns on superhydrophobic surfaces [148]. The use of self-assembled monolayers (SAMs) requires that a stable monolayer lm, which has specic and favorable interaction with the selected solid substrate, should be formed [165]. As a result, the formed lm must remain attached rmly onto the substrate because of the

Fig. 29. Simplied schematic of the multilayer lm showing the three main assembly blocks. Image reprinted from Ref. [151], with permission from American Chemical Society, Copyright 2007.

interaction, even after the substrate is taken off from the solution which originates the SAMs. And the multilayered structure can be created by repeatedly conducting the SAMs [119,131,166]. Zhu et al. reported a self-assembly process to prepare superhydrophobic dandelion-like 3D microstructure, which was self-assembled of 1D nanobers of polyaniline (PANI) in the presence of peruorosebacic acid (PFSEA) [167]. It was proposed that the self-assembling of PANI 1D nanobers, driven by the combined interaction of hydrogen bonding, stacking and hydrophobic interactions, led to the formation of the 3D microstructures. Shen et al. utilized supramolecular chemistry to fabricate structural templates by using hierarchical selfassembling of a fullerene derivative bearing three long aliphatic chains [166]. As shown in Fig. 31, the assembled fullerene derivative provided surface roughness, while the SAMs of thiols imparted superhydrophobic or superhydrophilic characteristics. Also, the SAMs of low surface energy material introduce superhydrophobic properties to micro- or nanostructure through chemical modication. The combination of the conventional silicon wet-etching technique (for microstructures) and a solution method for ZnO nanorod formation (for nanostructures) was reported to fabricate dual-scaled surfaces [58]. The maximum difference of contact angles between the surfaces with and without SAMs could be up to 167. Another example reported superhydrophobic and transparent glass surfaces that were treated with SAMs of tridecauoro-1,1,2,2-tetrahydrooctyltrichlorosilane (FOTS) [131]. The combination of colloidal lithography and plasma etching was performed to produce tower-shaped nanostructures, which were then treated with uoroalkylsilane SAMs to obtain the superhydrophobic property. Motornov et al. reported the fabrication of superhydrophobic surfaces by depositing (casting) nanoparticles aggregates from waterborn solutions [168]. The particle aggregates were hierarchical colloids, consisting of three levels, as shown in Fig. 32(ac). The grafted copolymer layer was considered to be a mixed brush formed by P4VP and poly(styrene-block-4-vinylpyridine) P(S-b-4VP) chains with a broad distribution of length [Fig. 32(d)]. Initially, the formed coating was hydrophilic, but heating above the glass transition temperature of polystyrene (PS) switched the particle shell into a hydrophobic state and even resulted in the superhydrophobicity. Changing the pH of the particle dispersions could result in different surface morphologies and thus different wetting properties, as shown in Fig. 33.

98

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 30. (a) TEM image of synthesized silica nanoparticles. (b) SEM image of a multilayer colloidal crystal prepared by spin-coating. (c) Higher-magnication SEM image of the sample in (b). Images reprinted from Ref. [164], with permission from IOP Publishing Limited, Copyright 2008.

Xue et al. prepared superhydrophobic surfaces on cotton textiles with a dual-sized hierarchical structure based on a complex coating of silica nanoparticles [169]. The nanoparticles were prepared by polymerization of tetraethylorthosilicate (TEOS), followed by aminofunctionalization and epoxy-functionalization respectively. The cotton textiles, which had been epoxy-functionalized by a mixture of epichlorohydrin and alkaline solution in advance, were dipped into the ethanol solution of amino-functionalized silica nanoparticles and then the ethanol solution of epoxy-functionalized silica nanoparticles with proper nipping. The silica particles were covalently bonded to the cotton bers during the process, and the outer surface of the ber was full of epoxy groups available for further surface grafting. To hydrophobize the cotton textiles, stearic acid, 1H,1H,2H,2H-peruorodecyltrichlorosilane (PFTDS), and/or their combination could be nally grafted onto the rough surface. The methodology is schematically shown in Fig. 34. From those discussed examples, it can be seen that the process based on colloidal assembly and aggregation is very effective in modifying surface energy and promoting surface roughness (structure) to create extreme surface wettability. The chemical reactions which are often accompanied with the process can produce fairly considerable connection between the colloidal aggregates and the substrates, and thus maintain the durability of the formed structure. It does not necessarily need complicated production equipment or operation procedures.

Therefore, the colloidal assembling and aggregation is relatively more economical and efcient than traditional imprinting methods and even some CVD methods. It has the potential to produce both sufcient surface coverage and precise order of structure. Nonetheless, the primary underlying task to widely apply the process is resting on grafting the desired properties into the colloidal solutions as well as keeping the system stable and uniform.

7.3. Solgel process The solgel methods have been utilized to create superhydrophobic surfaces in accordance with Wenzel or CassieBaxter's theories since the very early stage of mimicking lotus leaves' surface structure [4,158,170175]. It involves a chemical solution deposition, during which the chemical solution or sol is utilized as a precursor on the selected substrate to form a gel-like network. Material of low surface energy and micro- or nanoparticles can be added into the network to create superhydrophobic surfaces. For its compatibility with glass, this type of method is particularly favored in creating transparent and superhydrophobic lms on glass surfaces [157,176181]. To provide generally regular surface pattern to the network, extra control over the process like spin-coating is always needed.

Fig. 31. SEM images of globular assemblies of the fullerene derivative (a), Au nanoakes transcribed from the globular objects of the fullerene derivative (b), top-view (c) and 34.7 tilted-view (d) of an Au nanoake. Images reprinted from Ref. [166], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2009.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

99

Fig. 32. Different morphologies of synthesized hybrid nanoparticles. (a) Primary aggregates coated by the collapsed P4VP (red) and stretched PS (blue) chains in toluene, a good solvent for PS, and a poor solvent for P4VP; (b) primary aggregates coated by the collapsed PS and stretched P4VP in water at pH b 4, a good solvent for P4VP and a poor solvent for PS; (c) secondary aggregate in water at pH N 5; (d) schematic representation of the grafting block copolymer brushes on the surface of primary aggregates of silica particles. Images reprinted from Ref. [168], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2008.

One of the recently reported examples was based on the use of a eutectic liquid in a solgel process to form silica lms with optical transparency and superhydrophobicity [178]. The involved eutectic liquid (composite of urea and choline chloride) had an extremely low vapor pressure relative to a common solgel solvent (for example of ethanol). The eutectic liquid remained in the lm throughout the process under ambient conditions, which contributed to the control over lm thickness and surface roughness. Porous silica thin lms with asperities were formed after gelation in the presence of a base catalyst, which was composed of tetraethoxysilane, ethanol, HCl and H2O, and subsequent extraction of the eutectic liquid. Proper uoroalkylsilane (PFOS) treatment imparted the superhydrophobicity with a contact angle of 170 and a hysteresis of 23. To further promote transparency of the superhydrophobic lms, the sol solution was diluted with ethanol and then spin-coated on substrates. Fig. 35 shows SEM images of surfaces spin-coated with sol dilutions with the ratios of ethanol against sol being 1:1, 3:1, 5:1 and 7:1, respectively. The surface roughness and lm thickness of spin-coated lms

decreased continuously with dilution, while the superhydrophobicity of the surface was maintained at a certain level. As a result, the transparency of the lms was improved. 7.4. Electrospinning and electrospraying Strictly speaking, electrospraying and electrospinning are two similar but different techniques which can both be used to fabricate micro- or nanostructures. But when it comes down to fabricating superhydrophobic surfaces, those two methods are very easy to be garbled. Electrospinning is a simple and versatile method to produce continuous polymer bers in micro- and nanoscale [182184]. It is an effective way to create superhydrophobic surfaces by prompting the roughness [38]. To form uniform bers, the molecular weight of the polymer and the concentration of the solution should be properly prepared [185]. On the other hand, electrospraying is not necessarily restricted in bers only. The polymer lms deposited by electrospraying can range from spheres to bers [186]. But on a general level, it is

Fig. 33. 3D AFM images obtained for the coatings on silicon wafers prepared from aqueous dispersions (a) at pH 2 and (b) at pH 6. Images reprinted from Ref. [168], with permission from Wiley-VCH Verlag GmbH & Co. KGaA, Copyright 2008.

100

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

Fig. 34. Schematic illustration of preparation procedure of superhydrophobic surfaces on a cotton substrate. Image reprinted from Ref. [169], with permission from Elsevier, Copyright 2009.

considered that bers are produced during electrospinning process, and beads are produced during electrospraying process [187]. Ding et al. reported superhydrophobic surfaces with ZnO nanostructures obtained via electrospinning [188]. Composite poly(vinyl alcohol) (PVA)/ZnO nanobrous lms, formed by electrospinning the PVA/zinc acetate solution, were calcined in air to get ZnO brous lms. Then the as-prepared rough surface was modied by uoroalkylsilane (FAS) coating to induce low surface energy for superhydrophobicity. The eld emission scanning electron microscopy (FESEM) images in Fig. 36 (a and a`) show the inorganic brous ZnO lms. The bers formed the base structure and their average diameter was around 100 nm. Grain-like ZnO particles with an average diameter of 69 nm were formed on the ber surface and they composed of the secondary structure. Before FAS coating, the brous ZnO lms were superhydrophilic, showing a water contact angle of 0. After FAS coating, however, the brous FAS coated ZnO (FZnO) lms

showed apparent superhydrophobicity with a water contact angle up to 165 and a sliding angle down to 5. The FE-SEM images in Fig. 36 (b and b`) show that the brous FZnO lms still maintained the brous structure and grain-like ZnO particles on the ber surface. Han et al. recently investigated coaxial electrospinning to produce core-sheath-structured nano/microbers that combined different properties from individual core and sheath materials [189]. A normally nonelectrospinnable material such as Teon AF (2,2-bis-triuoromethyl4,5-diuoro-1,3-dioxole with tetrauoroethylene) was successfully electrospun when combined with an electrospinnable core material. As shown in Fig. 37(a), a coaxial nozzle consisted of a central tube surrounded by a concentric annular tube. Two polymer solutions for the core and sheath materials were separately fed into the coaxial nozzle from which they were ejected simultaneously. A compound pendant droplet emerged from the coaxial nozzle in Fig. 37(b). Upon a sufcient voltage, a charged liquid jet, referred to as the Taylor cone, was ejected

Fig. 35. SEM surface images of dilute-sol-coated glass slides. The ratios of ethanol:sol are 1:1 (a), 3:1 (b), 5:1 (c), and 7:1 (d). Images reprinted from Ref. [178], with permission from Elsevier, Copyright 2009.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

101

Fig. 36. FE-SEM images of ZnO (a) and FZnO (b) brous lms. The insets are the corresponding proles of a water droplet on the lms. (a`) and (b`) are the magnication corresponding to (a) and (b), respectively. Images reprinted from Ref. [188], with permission from Elsevier, Copyright 2008.

from the tip of a distorted droplet [Fig. 37(c)] that consisted of the core material enveloped by the sheath material. Superhydrophobic and oleophobic membranes were fabricated using the coaxial electrospinning method, while the mechanical strength of the core material was preserved. Similarly, Li et al. reported a large-scale superhydrophobic mat surface fabricated via a four-jet electrospinning technique [190]. The electrospun surface was imparted with an enhanced mechanical property by blending PS and polyamide 6 (PA6) bers. There were four syringes xed in the system to electrospin the bers. The mechanical properties of PS mats were regulated by tuning the ratio of jets of PS/ PA6 in the electrospinning process. Burkarter et al. employed the electrospray technique to deposit polytetrauoroethylene (PTFE) lms on uorine doped tin oxide (FTO) coated glass slides [191]. A superhydrophobic PTFE surface was achieved, with the water contact angle being up to 160 and the sliding angle down to ~2. As a matter of fact, the electrospray technique used in this case was quite similar to electrospinning. Because no apparent

evidence of PTFE ber formation was found, the method used was considered as electrospraying. Fig. 38 shows a characteristic SEM image of the electrosprayed PTFE coating with peculiar micro- and nanosized features, which provided typical rough structure required for the superhydrophobic surfaces. 7.5. Phase separation Colloidal aggregation methods can also involve phase separation, which separates the solid phase from a metastable mixture of substances by changing surrounding conditions, for example of temperature and pressure, to fabricate patterned surface structures. The surface structure and roughness formed through phase separation can be macroscopic, microscopic, and even nanoscopic [99,192]. Due to its working mechanism, this method is often related to the sol

Fig. 37. Coaxial electrospinning operation. (a) Diagram of the coaxial nozzle; (b) coresheath droplet without bias; (c) Taylor cone and coaxial jet formation at 12.5 kV. Images reprinted from Ref. [189], with permission from American Chemical Society, Copyright 2009.

Fig. 38. SEM image of the electrosprayed PTFE superhydrophobic surface. Image reprinted from Ref. [191], with permission from IOP Publishing Limited, Copyright 2007.

102

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

gel process for creating controllable surface patterns [171,193]. However, phase separation, is not necessarily utilized by solgel methods only, but also can assist other techniques, including plasma treatment [135], electrospinning [182], and self-aggregation [194], to make superhydrophobic surfaces. Note that the phase separation processes involved in these various methods are more or less connected to colloidal polymerization. 7.6. TiO2 and ZnO induced wettability TiO2 is known as an efcient and stable photocatalyst [195]. TiO2 surfaces with proper treatment can be light (UV irradiation) induced into an amphiphilic (both hydrophilic and oleophilic) state, and the water contact angle on the induced surfaces can be close to 0 [196198]. However, it is also found that, for surfaces coated with thin lms which mainly consist of TiO2 nanoparticles, the water contact angle goes up and gets restored in a dark place [199,200]. Moreover, the effect of adding other nanoparticles into the TiO2 thin lms can change the surface wettability. For instance, SiO2 nanoparticles are found constructively helpful to maintain the superhydrophilicity relatively long after UV irradiation [150]. Also, adding TiO2 nanoparticles into other lms can do the job as well [122]. It has been argued that TiO2 coated surfaces can well lead to selfcleaning effect [201203], but the mechanism is quite different from the self-cleaning effect corresponding to the lotus effect. Upon UV illumination on TiO2, electron-hole pairs are generated and they produce hydroxyl and superoxide radicals which can oxidize and decompose organic compounds absorbed on the TiO2 coated surfaces [198,204,205]. For those surfaces mainly functioned by light-induced TiO2, the self-cleaning effect can be enhanced by sunlight and rainfall, as shown in Fig. 39. The light-induced characteristic of TiO2 provides a novel way to prepare surfaces with superhydrophilic and superhydrophobic properties both [118]. It should be noted that the superhydrophilicity of TiO2 coated surfaces rests not only on the intrinsic property of TiO2, of which the mechanism has been proposed to be the reconstruction of surface hydroxyl groups under UV illumination, but also on the surface morphology [197,205]. As stated previously, surface roughness can be utilized to increase the contact angle on hydrophobic surfaces, but decrease the contact angle on hydrophilic surfaces. It has also been found that, from the comparative studies, the conversion reactions of photo-induced wettability on the ZnO thin lms follow a similar mechanism with TiO2 [206], which means superhydrophilicity can also be light-induced on surfaces mainly functionalized by ZnO. Thus it is feasible to fabricate superhydrophobic/superhydrophilic surfaces by taking advantage of either ZnO or TiO2 [119]. Actually the light-induced mechanism corresponding to the wettability of ZnO coatings lies in the electronhole pairs as well. The electron and hole generated by ZnO upon UV illumination react with the lattice oxygen and

then form oxygen vacancies [207]. Water molecules on the ZnO surfaces may coordinate into the oxygen vacancy sites, leading to adsorption of the water molecules on the surface; the induced superhydrophilicity can go back to superhydrophobicity if the surface is placed in the dark [206]. Therefore, the wettability is convertible from being superhydrophilic to being superhydrophobic on ZnO surfaces, and this point is useful to control wetting states of surfaces.

8. Conclusions and outlook Although superhydrophobicity is only a recently developed concept, it has already become important to a lot of research and will be potentially important to people's life. As stated in previous sections, a great amount of effort has been put into the research to understand the mechanisms that are related to superhydrophobicity on solid surfaces. This nature-inspired theory is an interdisciplinary subject which involves physics, chemistry, material science, and even biology. In the cutting edge research, nanotechnologies are effectively used to explore the intrinsic essence of the subject. Hence, with more developed equipment and further investigation, we are able to see the real structures of the superhydrophobic models that are found in nature and fabricate manmade superhydrophobic surfaces which have signicant water-repellence and durability. The achievement of superhydrophobicity-related study is phenomenal and this tendency will maintain and continue. However, the needed research will be a long job. Young equation, Wenzel equation, and CassieBaxter equation are the fundamental theories to describe the wetting phenomena on superhydrophobic surfaces. They are very necessary and useful, but not sufcient to t all situations. When a liquid drop wets a superhydrophobic surface, its state depends on the particular surface property and the wetting process. A transition between the wetting states might take place if the energy balance of the liquid drop is upset. The question for theoretical research in the future lies on how to work out governing equations that are adequate to illustrate the stable wetting regimes as well as the transitions on superhydrophobic surfaces. With further research into the theories, tremendous work has been made to understand the wetting phenomena on superhydrophobic surfaces. We thus review the superhydrophobic surfaces in nature and the most recent progress in fabricating manmade superhydrophobic surfaces from a biomimetic point of view. The achievement people have made is phenomenal and it is possible to apply this technology into self-cleaning, drag reduction, and many other elds [65]. As discussed previously, creating a superhydrophobic surface may involve several procedures, during which a number of techniques are used together. But in some cases, it is also possible to create superhydrophobic surfaces using relatively simple and prompt methods. No matter how the superhydrophobic surfaces are made, it is always important to keep the process under control and obtain the required effect precisely. Even greater progress is in prospect due to increasing attention and developing technique. The mostly expected breakthroughs in the future research will be concentrated on precisely controlling surface roughness and structure in both small and large scales, understanding more details of the wetting process, especially the wetting transitions, and turning up the application in an even broader range of materials.

Acknowledgment We thank H. J. Ensikat from University of Bonn, Germany, for his assistance to this work. We are grateful to UK Royal Society, China Scholarship Council and the Open Project of Key Laboratory of Bionic Engineering (Jilin University), Ministry of Education, China, 2010 (project no. K201008).

Fig. 39. Schematic diagram of the decontamination process occurred on the superhydrophilic self-cleaning surface. Image reprinted from Ref. [205], with permission from Elsevier, Copyright 2006.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105

103

References
[1] Burton Z, Bhushan B. Surface characterization and adhesion and friction properties of hydrophobic leaf surfaces. Ultramicroscopy 2006;106:70919. [2] Poynor A, Hong L, Robinson IK, Granick S, Zhang Z, Fenter PA. How water meets a hydrophobic surface. Phys Rev Lett 2006;97:4. [3] Onda T, Shibuichi S, Satoh N, Tsujii K. Super-water-repellent fractal surfaces. Langmuir 1996;12:21257. [4] Bico J, Marzolin C, Quere D. Pearl drops. Europhys Lett 1999;47:2206. [5] Chen W, Fadeev AY, Hsieh MC, Oner D, Youngblood J, McCarthy TJ. Ultrahydrophobic and ultralyophobic surfaces: some comments and examples. Langmuir 1999;15:33959. [6] Yoshimitsu Z, Nakajima A, Watanabe T, Hashimoto K. Effects of surface structure on the hydrophobicity and sliding behavior of water droplets. Langmuir 2002;18:581822. [7] Kijlstra J, Reihs K, Klamt A. Roughness and topology of ultra-hydrophobic surfaces. Colloid Surf A-Physicochem Eng Asp 2002;206:5219. [8] Patankar NA. Mimicking the lotus effect: inuence of double roughness structures and slender pillars. Langmuir 2004;20:820913. [9] Seemann R, Monch W, Herminghaus S. Liquid ow in wetting layers on rough substrates. Europhys Lett 2001;55:698704. [10] Herminghaus S. Roughness-induced non-wetting. Europhys Lett 2000;52:16570. [11] Gau H, Herminghaus S, Lenz P, Lipowsky R. Liquid morphologies on structured surfaces: from microchannels to microchips. Science 1999;283:469. [12] Rosario R, Gust D, Garcia AA, Hayes M, Taraci JL, Clement T, et al. Lotus effect amplies light-induced contact angle switching. J Phys Chem B 2004;108:126402. [13] Feng L, Zhang YA, Xi JM, Zhu Y, Wang N, Xia F, et al. Petal effect: a superhydrophobic state with high adhesive force. Langmuir 2008;24:41149. [14] Nosonovsky M, Bhushan B. Roughness optimization for biomimetic superhydrophobic surfaces. -Inf Storage Process Syst 2005;11:53549. [15] Gu ZZ, Uetsuka H, Takahashi K, Nakajima R, Onishi H, Fujishima A, et al. Structural color and the lotus effect. Angew Chem Int Ed 2003;42 894-+. [16] Ma ML, Hill RM. Superhydrophobic surfaces. Curr Opin Colloid Interface Sci 2006;11:193202. [17] Marmur A. The lotus effect: superhydrophobicity and metastability. Langmuir 2004;20:35179. [18] Gao LC, McCarthy TJ. Contact angle hysteresis explained. Langmuir 2006;22: 62347. [19] Lafuma A, Quere D. Superhydrophobic states. Nat Mater 2003;2:45760. [20] Quere D. Non-sticking drops. Rep Prog Phys 2005;68:2495532. [21] Bico J, Thiele U, Quere D. Wetting of textured surfaces. International workshop on nanocapillarity: wetting of heterogenous surface and porous solid. Princeton, New Jersey: Elsevier Science Bv; 2001. p. 416. [22] Extrand CW. Model for contact angles and hysteresis on rough and ultraphobic surfaces. Langmuir 2002;18:79919. [23] Choi CH, Kim CJ. Large slip of aqueous liquid ow over a nanoengineered superhydrophobic surface. Phys Rev Lett 2006;96:4. [24] Wong TS, Huang APH, Ho CM. Wetting behaviors of individual nanostructures. Langmuir 2009;25:6599603. [25] Priest C, Albrecht TWJ, Sedev R, Ralston J. Asymmetric wetting hysteresis on hydrophobic microstructured surfaces. Langmuir 2009;25:565560. [26] Roach P, Shirtcliffe NJ, Newton MI. Progess in superhydrophobic surface development. Soft Matter 2008;4:22440. [27] Koch K, Bhushan B, Barthlott W. Multifunctional surface structures of plants: an inspiration for biomimetics. Prog Mater Sci 2009;54:13778. [28] Feng XJ, Jiang L. Design and creation of superwetting/antiwetting surfaces. Adv Mater 2006;18:306378. [29] Li W, Amirfazli A. A thermodynamic approach for determining the contact angle hysteresis for superhydrophobic surfaces. J Colloid Interface Sci 2005;292:195201. [30] Li W, Amirfazli A. Microtextured superhydrophobic surfaces: a thennodynarnic analysis. Adv Colloid Interface Sci 2007;132:5168. [31] Li W, Amirfazli A. Hierarchical structures for natural superhydrophobic surfaces. Soft Matter 2008;4:4626. [32] Barthlott W, Neinhuis C. Purity of the sacred lotus, or escape from contamination in biological surfaces. Planta 1997;202:18. [33] Koch K, Bhushan B, Barthlott W. Diversity of structure, morphology and wetting of plant surfaces. Soft Matter 2008;4:194363. [34] Fang Y, Sun G, Cong Q, Chen GH, Ren LQ. Effects of methanol on wettability of the non-smooth surface on buttery wing. J Bionic Eng 2008;5:12733. [35] Byun D, Hong J, Ko JH, Saputra, Lee YJ, Park HC, et al. Wetting characteristics of insect wing surfaces. J Bionic Eng 2009;6:6370. [36] Barnes WJP. Biomimetic solutions to sticky problems. Science 2007;318:2034. [37] Nosonovsky M, Bhushan B. Multiscale friction mechanisms and hierarchical surfaces in nano- and bio-tribology. Mater Sci Eng R-Rep 2007;58:16293. [38] Tuteja A, Choi W, Ma ML, Mabry JM, Mazzella SA, Rutledge GC, et al. Designing superoleophobic surfaces. Science 2007;318:161822. [39] Gao N, Yan YY, Chen XY, Mee DJ. Superhydrophobic surfaces with hierarchical structure. Mater Lett 2011;65:29025. [40] Wenzel RN. Resistance of solid surfaces to wetting by water. Ind Eng Chem 1936;28:98894. [41] Cassie ABD, Baxter S. Wettability of porous surfaces. Trans Faraday Soc 1944;40: 054650. [42] Callies M, Quere D. On water repellency. Soft Matter 2005;1:5561. [43] Wenzel RN. Surface roughness and contact angle. J Phys Colloid Chem 1949;53: 14667.

[44] Marmur A. Wetting on hydrophobic rough surfaces: to be heterogeneous or not to be? Langmuir 2003;19:83438. [45] Cassie ABD. Contact angles. Discuss Faraday Soc 1948;3:116. [46] Blossey R. Self-cleaning surfaces virtual realities. Nat Mater 2003;2:3016. [47] Patankar NA. Transition between superhydrophobic states on rough surfaces. Langmuir 2004;20:7097102. [48] Wolansky G, Marmur A. Apparent contact angles on rough surfaces: the Wenzel equation revisited. Colloid Surf A-Physicochem Eng Asp 1999;156:3818. [49] Brandon S, Haimovich N, Yeger E, Marmur A. Partial wetting of chemically patterned surfaces: the effect of drop size. J Colloid Interface Sci 2003;263: 23743. [50] Tsujii K, Yamamoto T, Onda T, Shibuichi S. Super oil-repellent surfaces. Angew Chem Int Ed Engl 1997;36:10112. [51] Shibuichi S, Yamamoto T, Onda T, Tsujii K. Super water- and oil-repellent surfaces resulting from fractal structure. J Colloid Interface Sci 1998;208:28794. [52] Shibuichi S, Onda T, Satoh N, Tsujii K. Super water-repellent surfaces resulting from fractal structure. J Phys Chem 1996;100:195127. [53] Kwon Y, Patankar N, Choi J, Lee J. Design of surface hierarchy for extreme hydrophobicity. Langmuir 2009;25:612936. [54] Patankar NA. On the modeling of hydrophobic contact angles on rough surfaces. Langmuir 2003;19:124953. [55] Quere D, Lafuma A, Bico J. Slippy and sticky microtextured solids. Nanotechnology 2003;14:110912. [56] Jung YC, Bhushan B. Dynamic effects induced transition of droplets on biomimetic superhydrophobic surfaces. Langmuir 2009;25:920818. [57] Feng L, Song YL, Zhai J, Liu BQ, Xu J, Jiang L, et al. Creation of a superhydrophobic surface from an amphiphilic polymer. Angew Chem Int Ed 2003;42:8002. [58] Kim H, Kim MH, Kim J. Wettability of dual-scaled surfaces fabricated by the combination of a conventional silicon wet-etching and a ZnO solution method. J Micromech Microeng 2009;19:7. [59] Chen MH, Hsu TH, Chuang YJ, Tseng FG. Dual hierarchical biomimic superhydrophobic surface with three energy states. Appl Phys Lett 2009;95:3. [60] Journet C, Moulinet S, Ybert C, Purcell ST, Bocquet L. Contact angle measurements on superhydrophobic carbon nanotube forests: effect of uid pressure. Europhys Lett 2005;71:1049. [61] Kim JW, Kim CJ. Nanostructured surfaces for dramatic reduction of ow resistance in droplet-based microuidics. Fifteenth IEEE International Conference on Micro Electro Mechanical Systems. Tech Digest 2002:47982. [62] Jindasuwan S, Nimittrakoolchai O, Sujaridworakun P, Jinawath S, Supothina S. Surface characteristics of water-repellent polyelectrolyte multilayer lms containing various silica contents. Thin Solid Films 2009;517:50015. [63] Gao LC, McCarthy TJ. A commercially available perfectly hydrophobic material (theta(A)/theta(R) = 180 degrees/180 degrees). Langmuir 2007;23:91257. [64] Sawada H, Suzuki T, Takashima H, Takishita K. Preparation and properties of uoroalkyl end-capped vinyltrimethoxysilane oligomeric nanoparticles a new approach to facile creation of a completely superhydrophobic coating surface with these nanoparticles. Colloid Polym Sci 2008;286:156974. [65] Zhang X, Shi F, Niu J, Jiang YG, Wang ZQ. Superhydrophobic surfaces: from structural control to functional application. J Mater Chem 2008;18:62133. [66] Oner D, McCarthy TJ. Ultrahydrophobic surfaces. Effects of topography length scales on wettability. Langmuir 2000;16:777782. [67] Long CJ, Schumacher JF, Brennan AB. Potential for tunable static and dynamic contact angle anisotropy on gradient microscale patterned topographies. Langmuir 2009;25:129829. [68] Jeong HE, Kwak MK, Park CI, Suh KY. Wettability of nanoengineered dual-roughness surfaces fabricated by UV-assisted capillary force lithography. J Colloid Interface Sci 2009;339:2027. [69] Adamson AW, Gast AP. Physical chemistry of surfaces. 6th ed. New York: John Wiley & Sons; 1997. [70] Nosonovsky M, Bhushan B. Multiscale dissipative mechanisms and hierarchical surfaces: friction, superhydrophobicity, and biomimetics. Berlin: Springer; 2008. [71] Orowan E. Surface energy and surface tension in solids and liquids. Proceedings of the Royal Society of London Series a Mathematical and Physical Sciences, 316; 1970. 473-&. [72] Kumikov VK, Khokonov KB. On the measurement of surface free-energy and surface-tension of solid metals. J Appl Phys 1983;54:134650. [73] Ip SW, Toguri JM. The equivalency of surface-tension, surface-energy and surface free-energy. J Mater Sci 1994;29:68892. [74] Wang JD, Chen HS, Sui T, Li A, Chen DR. Investigation on hydrophobicity of lotus leaf: experiment and theory. Plant Sci 2009;176:68795. [75] Gao N, Yan YY. Modeling superhydrophobic contact angles and wetting transition. J Bionic Eng 2009;6:33540. [76] Koishi T, Yasuoka K, Fujikawa S, Ebisuzaki T, Zeng XC. Coexistence and transition between Cassie and Wenzel state on pillared hydrophobic surface. Proc Natl Acad Sci USA 2009;106:843540. [77] Jung YC, Bhushan B. Wetting transition of water droplets on superhydrophobic patterned surfaces. Scr Mater 2007;57:105760. [78] Ishino C, Okumura K. Wetting transitions on textured hydrophilic surfaces. Eur Phys J E 2008;25:41524. [79] Ran CB, Ding GQ, Liu WC, Deng Y, Hou WT. Wetting on nanoporous alumina surface: transition between Wenzel and Cassie states controlled by surface structure. Langmuir 2008;24:99525. [80] Kusumaatmaja H, Blow ML, Dupuis A, Yeomans JM. The collapse transition on superhydrophobic surfaces. Epl, 81; 2008. [81] Nosonovsky M, Bhushan B. Biomimetic superhydrophobic surfaces: multiscale approach. Nano Lett 2007;7:26337.

104

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105 [121] Hou WX, Wang QH. Stable polytetrauoroethylene superhydrophobic surface with lotus-leaf structure. J Colloid Interface Sci 2009;333:4003. [122] Sato O, Kubo S, Gu ZZ. Structural color lms with lotus effects, superhydrophilicity, and tunable stop-bands. Acc Chem Res 2009;42:110. [123] Cho WK, Choi IS. Fabrication of hairy polymeric lms inspired by geckos: wetting and high adhesion properties. Adv Funct Mater 2008;18:108996. [124] Yuan ZQ, Chen H, Tang JX, Gong HF, Liu YJ, Wang ZX, et al. A novel preparation of polystyrene lm with a superhydrophobic surface using a template method. J Phys D-Appl Phys 2007;40:34859. [125] Bormashenko E, Stein T, Whyman G, Bormashenko Y, Pogreb R. Wetting properties of the multiscaled nanostructured polymer and metallic superhydrophobic surfaces. Langmuir 2006;22:99825. [126] Berendsen CWJ, Skeren M, Najdek D, Cerny F. Superhydrophobic surface structures in thermoplastic polymers by interference lithography and thermal imprinting. Appl Surf Sci 2009;255:930510. [127] Winkleman A, Gotesman G, Yoffe A, Naaman R. Immobilizing a drop of water: fabricating highly hydrophobic surfaces that pin water droplets. Nano Lett 2008;8:12415. [128] Manca M, Cortese B, Viola I, Arico AS, Cingolani R, Gigli G. Inuence of chemistry and topology effects on superhydrophobic CF4-plasma-treated poly(dimethylsiloxane) (PDMS). Langmuir 2008;24:183343. [129] Tsougeni K, Papageorgiou D, Tserepi A, Gogolides E. Smart polymeric microuidics fabricated by plasma processing: controlled wetting, capillary lling and hydrophobic valving. Lab Chip.10:462-9. [130] Lee SM, Jung ID, Ko JS. The effect of the surface wettability of nanoprotrusions formed on network-type microstructures. J Micromech Microeng 2008;18. [131] Lim H, Jung DH, Noh JH, Choi GR, Kim WD. Simple nanofabrication of a superhydrophobic and transparent biomimetic surface. Chin Sci Bull 2009;54:36136. [132] Choi CH, Ulmanella U, Kim J, Ho CM, Kim CJ. Effective slip and friction reduction in nanograted superhydrophobic microchannels. Phys Fluids 2006;18:8. [133] Park SG, Moon HH, Lee SK, Shim J, Yang SM. Bioinspired holographically featured superhydrophobic and supersticky nanostructured materials. Langmuir 2010;26:146872. [134] Chen MH, Hsu TH, Chuang YJ, Tseng FG. Dual hierarchical biomimic superhydrophobic surface with three energy states. Appl Phys Lett 2009;95. [135] Song WL, Veiga DD, Custodio CA, Mano JF. Bioinspired degradable substrates with extreme wettability properties. Adv Mater 2009;21 1830-+. [136] Hwang SJ, Oh DJ, Jung PG, Lee SM, Go JS, Kim JH, et al. Dry etching of polydimethylsiloxane using microwave plasma. J Micromech Microeng 2009;19. [137] Ji YY, Kim SS, Kwon OP, Lee SH. Easy fabrication of large-size superhydrophobic surfaces by atmospheric pressure plasma polymerization with non-polar aromatic hydrocarbon in an in-line process. Appl Surf Sci 2009;255:45758. [138] Lee SM, Oh DJ, Jung ID, Bae KM, Jung PG, Chung KH, et al. Fabrication of nickel micromesh sheets and evaluation of their water-repellent and water-proof abilities. Int J Precis Eng Manuf 2009;10:1616. [139] Byun D, Lee Y, Tran SBQ, Nugyen VD, Kim S, Park B, et al. Electrospray on superhydrophobic nozzles treated with argon and oxygen plasma. Appl Phys Lett 2008;92. [140] Balu B, Breedveld V, Hess DW. Fabrication of roll-off and sticky superhydrophobic cellulose surfaces via plasma processing. Langmuir 2008;24:478590. [141] Miyahara Y, Mitamura K, Saito N, Takai O. Fabrication of microtemplates for the control of bacterial immobilization. J Vac Sci Technol, A 2009;27:11837. [142] Intranuovo F, Sardella E, Rossini P, d'Agostino R, Favia P. PECVD of uorocarbon coatings from hexauoropropylene oxide: glow vs. afterglow. Chem Vap Deposition 2009;15:95100. [143] Milella A, Di Mundo R, Palumbo F, Favia P, Fracassi F, d'Agostino R. Plasma nanostructuring of polymers: different routes to superhydrophobicity. Plasma Processes Polym 2009;6:4606. [144] Han ZJ, Tay BK, Shakerzadeh M, Ostrikov K. Superhydrophobic amorphous carbon/carbon nanotube nanocomposites. Appl Phys Lett 2009;94. [145] Borras A, Barranco A, Gonzalez-Elipe AR. Reversible superhydrophobic to superhydrophilic conversion of Ag@TiO2 composite nanober surfaces. Langmuir 2008;24:80216. [146] Ishizaki T, Saito N, Inoue Y, Bekke M, Takai O. Fabrication and characterization of ultra-water-repellent alumina-silica composite lms. J Phys D-Appl Phys 2007;40:1927. [147] Chen W, McCarthy TJ. Layer-by-layer deposition: a tool for polymer surface modication. Macromolecules 1997;30:7886. [148] Soeno T, Inokuchi K, Shiratori S. Ultra-water-repellent surface: fabrication of complicated structure of SiO2 nanoparticles by electrostatic self-assembled lms. Appl Surf Sci 2004;237:5437. [149] Wu Z, Lee D, Rubner MF, Cohen RE. Structural color in porous, superhydrophilic, and self-cleaning SiO2/TiO2 Bragg stacks. Small 2007;3:144551. [150] Lee D, Rubner MF, Cohen RE. All-nanoparticle thin-lm coatings. Nano Lett 2006;6:230512. [151] Bravo J, Zhai L, Wu ZZ, Cohen RE, Rubner MF. Transparent superhydrophobic lms based on silica nanoparticles. Langmuir 2007;23:72938. [152] Cebeci FC, Wu ZZ, Zhai L, Cohen RE, Rubner MF. Nanoporosity-driven superhydrophilicity: a means to create multifunctional antifogging coatings. Langmuir 2006;22:285662. [153] Zhai L, Cebeci FC, Cohen RE, Rubner MF. Stable superhydrophobic coatings from polyelectrolyte multilayers. Nano Lett 2004;4:134953. [154] Chunder A, Etcheverry K, Londe G, Cho HJ, Zhai L. Conformal switchable superhydrophobic/hydrophilic surfaces for microscale ow control. Colloid Surf APhysicochem Eng Asp 2009;333:18793. [155] Lvov Y, Ariga K, Onda M, Ichinose I, Kunitake T. Alternate assembly of ordered multilayers of SiO2 and other nanoparticles and polyions. Langmuir 1997;13:6195203.

[82] Lee C, Choi CH, Kim CJ. Structured surfaces for a giant liquid slip. Phys Rev Lett 2008;101. [83] Marmur A. From hygrophilic to superhygrophobic: theoretical conditions for making high-contact-angle surfaces from low-contact-angle materials. Langmuir 2008;24:75739. [84] Bormashenko E, Pogreb R, Whyman G, Erlich M. Resonance CassieWenzel wetting transition for horizontally vibrated drops deposited on a rough surface. Langmuir 2007;23:1221721. [85] Marmur A, Bittoun E. When Wenzel and Cassie are right: reconciling local and global considerations. Langmuir 2009;25:127781. [86] He B, Patankar NA, Lee J. Multiple equilibrium droplet shapes and design criterion for rough hydrophobic surfaces. Langmuir 2003;19:49995003. [87] Gao N, Yan YY, Chen XY, Zheng XF. Superhydrophobic composite lms based on THS and nanoparticles. J Bionic Eng 2010;7:S5966. [88] Sun G, Fang Y, Cong Q, Ren LQ. Anisotropism of the non-smooth surface of buttery wing. J Bionic Eng 2009;6:716. [89] Gorb SN. Biological attachment devices: exploring nature's diversity for biomimetics. Philos Trans R Soc A-Math Phys Eng Sci 2008;366:155774. [90] Neinhuis C, Barthlott W. Characterization and distribution of water-repellent, self-cleaning plant surfaces. Ann Bot 1997;79:66777. [91] Guo ZG, Liu WM. Biomimic from the superhydrophobic plant leaves in nature: binary structure and unitary structure. Plant Sci 2007;172:110312. [92] Furstner R, Barthlott W, Neinhuis C, Walzel P. Wetting and self-cleaning properties of articial superhydrophobic surfaces. Langmuir 2005;21:95661. [93] Barthlott W, Neinhuis C, Cutler D, Ditsch F, Meusel I, Theisen I, et al. Classication and terminology of plant epicuticular waxes. Bot J Linn Soc 1998;126:23760. [94] Lee SM, Kwon TH. Effects of intrinsic hydrophobicity on wettability of polymer replicas of a superhydrophobic lotus leaf. J Micromech Microeng 2007;17:68792. [95] Huger E, Rothe H, Frant M, Grohmann S, Hildebrand G, Liefeith K. Atomic force microscopy and thermodynamics on taro, a self-cleaning plant leaf. Appl Phys Lett 2009;95. [96] Gao XF, Jiang L. Water-repellent legs of water striders. Nature 2004;432 36-. [97] Fang Y, Sun G, Wang TQ, Cong Q, Ren LQ. Hydrophobicity mechanism of nonsmooth pattern on surface of buttery wing. Chin Sci Bull 2007;52:7116. [98] Gao HJ, Wang X, Yao HM, Gorb S, Arzt E. Mechanics of hierarchical adhesion structures of geckos. Mech Mater 2005;37:27585. [99] Chen H, Yuan ZQ, Zhang JD, Liu YJ, Li K, Zhao DJ, et al. Preparation, characterization and wettability of porous superhydrophobic poly(vinyl chloride) surface. J Porous Mat 2009;16:44751. [100] Li XM, Reinhoudt D, Crego-Calama M. What do we need for a superhydrophobic surface? A review on the recent progress in the preparation of superhydrophobic surfaces. Chem Soc Rev 2007;36:135068. [101] Steinberger A, Cottin-Bizonne C, Kleimann P, Charlaix E. High friction on a bubble mattress. Nat Mater 2007;6:6658. [102] Kim TI, Baek CH, Suh KY, Seo SM, Lee HH. Optical lithography with printed metal mask and a simple superhydrophobic surface. Small 2008;4:1825. [103] Reyssat M, Quere D. Contact angle hysteresis generated by strong dilute defects. J Phys Chem B 2009;113:39069. [104] He B, Lee J, Patankar NA. Contact angle hysteresis on rough hydrophobic surfaces. Colloid Surf A-Physicochem Eng Asp 2004;248:1014. [105] Joseph P, Tabeling P. Direct measurement of the apparent slip length. Phys Rev E 2005;71:4. [106] Ou J, Moss GR, Rothstein JP. Enhanced mixing in laminar ows using ultrahydrophobic surfaces. Phys Rev E 2007;76:10. [107] Jung YC, Bhushan B. Mechanically durable carbon nanotube-composite hierarchical structures with superhydrophobicity, self-cleaning, and low-drag. ACS Nano 2009;3:415563. [108] Choi YW, Han JE, Lee S, Sohn D. Preparation of a superhydrophobic lm with UV imprinting technology. Macromol Res 2009;17:8214. [109] Wang JZ, Zheng ZH, Li HW, Huck WTS, Sirringhaus H. Dewetting of conducting polymer inkjet droplets on patterned surfaces. Nat Mater 2004;3:1716. [110] Zhang XM, Zhang JH, Ren ZY, Li X, Zhang X, Zhu DF, et al. Morphology and wettability control of silicon cone arrays using colloidal lithography. Langmuir 2009;25:737582. [111] Xia YN, Whitesides GM. Soft lithography. Angew Chem Int Ed 1998;37:55175. [112] Xia YN, Whitesides GM. Soft lithography. Annu Rev Mater Sci 1998;28:15384. [113] Rogers JA, Nuzzo RG. Recent progress in soft lithography. Mater Today 2005;8:506. [114] Groves TR, Pickard D, Rafferty B, Crosland N, Adam D, Schubert G. Maskless electron beam lithography: prospects, progress, and challenges. Microelectron Eng 2002;612:28593. [115] Cerrina F. X-ray imaging: applications to patterning and lithography. J Phys DAppl Phys 2000;33:R10316. [116] Lee SM, Kwon TH. Mass-producible replication of highly hydrophobic surfaces from plant leaves. Nanotechnology 2006;17:318996. [117] Shi F, Liu Z, Wu GL, Zhang M, Chen H, Wang ZQ, et al. Surface imprinting in layerby-layer nanostructured lms. Adv Funct Mater 2007;17:18217. [118] Lai YK, Lin CJ, Wang H, Huang HY, Zhuang HF, Sun L. Superhydrophilicsuperhydrophobic micropattern on TiO2 nanotube lms by photocatalytic lithography. Electrochem Commun 2008;10:38791. [119] Lai YK, Lin ZQ, Huang JY, Sun L, Chen Z, Lin CJ. Controllable construction of ZnO/ TiO2 patterning nanostructures by superhydrophilic/superhydrophobic templates. New J Chem 2010;34:4451. [120] Lai YK, Huang YX, Wang H, Huang JY, Chen Z, Lin CJ. Selective formation of ordered arrays of octacalcium phosphate ribbons on TiO2 nanotube surface by template-assisted electrodeposition. Colloids Surf B Biointerfaces 2010;76: 11722.

Y.Y. Yan et al. / Advances in Colloid and Interface Science 169 (2011) 80105 [156] Ling XY, Phang IY, Vancso GJ, Huskens J, Reinhoudt DN. Stable and transparent superhydrophobic nanoparticle lms. Langmuir 2009;25:32603. [157] Manca M, Cannavale A, De Marco L, Arico AS, Cingolani R, Gigli G. Durable superhydrophobic and antireective surfaces by trimethylsilanized silica nanoparticles-based solgel processing. Langmuir 2009;25:635762. [158] Liu YY, Chen XQ, Xin JH. Super-hydrophobic surfaces from a simple coating method: a bionic nanoengineering approach. Nanotechnology 2006;17:325963. [159] Lai YK, Lin ZQ, Huang JY, Sun L, Chen Z, Lin CJ. Controllable construction of ZnO/ TiO2 patterning nanostructures by superhydrophilic/superhydrophobic templates. New J Chem 2010;34:4451. [160] Amigoni S, de Givenchy ET, Dufay M, Guittard F. Covalent layer-by-layer assembled superhydrophobic organicinorganic hybrid lms. Langmuir 2009;25:110737. [161] Sun C, Ge LQ, Gu ZZ. Fabrication of super-hydrophobic lm with dual-size roughness by silica sphere assembly. Thin Solid Films 2007;515:468690. [162] Zhao Y, Li M, Lu QH, Shi ZY. Superhydrophobic polyimide lms with a hierarchical topography: combined replica molding and layer-by-layer assembly. Langmuir 2008;24:126517. [163] Hsieh CT, Chen WY, Wu FL, Shen YS. Fabrication and superhydrophobic behavior of uorinated silica nanosphere arrays. J Adhes Sci Technol 2008;22:26575. [164] Min WL, Jiang P, Jiang B. Large-scale assembly of colloidal nanoparticles and fabrication of periodic subwavelength structures. Nanotechnology 2008;19. [165] Schwartz DK. Mechanisms and kinetics of self-assembled monolayer formation. Annu Rev Phys Chem 2001;52:10737. [166] Shen YF, Wang JB, Kuhlmann U, Hildebrandt P, Ariga K, Mohwald H, et al. Supramolecular templates for nanoake-metal surfaces. Chem-Eur J 2009;15:27637. [167] Zhu Y, Li JM, Wan MX, Jiang L. Superhydrophobic 3D microstructures assembled from 1D nanobers of polyaniline. Macromol Rapid Commun 2008;29:23943. [168] Motornov M, Sheparovych R, Lupitskyy R, MacWilliams E, Minko S. Superhydrophobic surfaces generated from water-borne dispersions of hierarchically assembled nanoparticles coated with a reversibly switchable shell. Adv Mater 2008;20 200-+. [169] Xue CH, Jia ST, Zhang J, Tian LQ. Superhydrophobic surfaces on cotton textiles by complex coating of silica nanoparticles and hydrophobization. Thin Solid Films 2009;517:45938. [170] Mahltig B, Bottcher H. Modied silica sol coatings for water-repellent textiles. Sci Technol 2003;27:4352. [171] Shirtcliffe NJ, McHale G, Newton MI, Perry CC. Intrinsically superhydrophobic organosilica solgel foams. Langmuir 2003;19:562631. [172] Pilotek S, Schmidt HK. Wettability of microstructured hydrophobic solgel coatings. Sci Technol 2003;26:78992. [173] Yu M, Gu GT, Meng WD, Qing FL. Superhydrophobic cotton fabric coating based on a complex layer of silica nanoparticles and peruorooctylated quaternary ammonium silane coupling agent. Appl Surf Sci 2007;253:366973. [174] Sheen YC, Chang WH, Chen WC, Chang YH, Huang YC, Chang FC. Non-uorinated superamphiphobic surfaces through solgel processing of methyltriethoxysilane and tetraethoxysilane. Mater Chem Phys 2009;114:638. [175] Jitianu A, Doyle J, Amatucci G, Klein LC. Methyl modied siloxane melting gels for hydrophobic lms. Sci Technol 2010;53:2729. [176] Hikita M, Tanaka K, Nakamura T, Kajiyama T, Takahara A. Super-liquid-repellent surfaces prepared by colloidal silica nanoparticles covered with uoroalkyl groups. Langmuir 2005;21:7299302. [177] Shang HM, Wang Y, Limmer SJ, Chou TP, Takahashi K, Cao GZ. Optically transparent superhydrophobic silica-based lms. Thin Solid Films 2005;472:3743. [178] Xiu YH, Xiao F, Hess DW, Wong CP. Superhydrophobic optically transparent silica lms formed with a eutectic liquid. Thin Solid Films 2009;517:16105. [179] Tadanaga K, Kitamuro K, Matsuda A, Minami T. Formation of superhydrophobic alumina coating lms with high transparency on polymer substrates by the sol gel method. Sci Technol 2003;26:7058. [180] Rao AV, Latthe SS, Nadargi DY, Hirashima H, Ganesan V. Preparation of MTMS based transparent superhydrophobic silica lms by solgel method. J Colloid Interface Sci 2009;332:48490. [181] Tadanaga K, Morinaga J, Minami T. Formation of superhydrophobicsuperhydrophilic pattern on owerlike alumina thin lm by the solgel method. Sci Technol 2000;19:2114. [182] Ma ML, Hill RM, Lowery JL, Fridrikh SV, Rutledge GC. Electrospun poly(styreneblock-dimethylsiloxane) block copolymer bers exhibiting superhydrophobicity. Langmuir 2005;21:554954.

105

[183] Zhu MF, Zuo WW, Yu H, Yang W, Chen YM. Superhydrophobic surface directly created by electrospinning based on hydrophilic material. J Mater Sci 2006;41: 37937. [184] Wang N, Zhao Y, Jiang L. Low-cost, thermoresponsive wettability of surfaces: poly(N-isopropylacrylamide)/polystyrene composite lms prepared by electrospinning. Macromol Rapid Commun 2008;29:4859. [185] Ma ML, Mao Y, Gupta M, Gleason KK, Rutledge GC. Superhydrophobic fabrics produced by electrospinning and chemical vapor deposition. Macromolecules 2005;38:97428. [186] Mizukoshi T, Matsumoto H, Minagawa M, Tanioka A. Control over wettability of textured surfaces by electrospray deposition. J Appl Polym Sci 2007;103:38117. [187] Zheng JF, He AH, Li JX, Xu JA, Han CC. Studies on the controlled morphology and wettability of polystyrene surfaces by electrospinning or electrospraying. Polymer 2006;47:7095102. [188] Ding B, Ogawa T, Kim J, Fujimoto K, Shiratori S. Fabrication of a superhydrophobic nanobrous zinc oxide lm surface by electrospinning. Thin Solid Films 2008;516:2495501. [189] Han DW, Steckl AJ. Superhydrophobic and oleophobic bers by coaxial electrospinning. Langmuir 2009;25:945462. [190] Li XH, Ding B, Lin JY, Yu JY, Sun G. Enhanced mechanical properties of superhydrophobic microbrous polystyrene mats via polyamide 6 nanobers. J Phys Chem C 2009;113:204527. [191] Burkarter E, Saul CK, Thomazi F, Cruz NC, Zanata SM, Roman LS, et al. Electrosprayed superhydrophobic PTFE: a non-contaminating surface. Phys 2007;40: 777881. [192] Li XH, Chen GM, Ma YM, Feng L, Zhao HZ, Jiang L, et al. Preparation of a superhydrophobic poly(vinyl chloride) surface via solventnonsolvent coating. Polymer 2006;47:5069. [193] Nakajima A, Abe K, Hashimoto K, Watanabe T. Preparation of hard superhydrophobic lms with visible light transmission. Thin Solid Films 2000;376:1403. [194] Xie QD, Xu J, Feng L, Jiang L, Tang WH, Luo XD, et al. Facile creation of a superamphiphobic coating surface with bionic microstructure. Adv Mater 2004;16 302-+. [195] Miyauchi M, Nakajima A, Hashimoto K, Watanabe T. A highly hydrophilic thin lm under 1 mu W/cm(2) UV illumination. Adv Mater 2000;12 1923-+. [196] Wang R, Hashimoto K, Fujishima A, Chikuni M, Kojima E, Kitamura A, et al. Lightinduced amphiphilic surfaces. Nature 1997;388:4312. [197] Wang R, Hashimoto K, Fujishima A, Chikuni M, Kojima E, Kitamura A, et al. Photogeneration of highly amphiphilic TiO2 surfaces. Adv Mater 1998;10 135-+. [198] Nakajima A, Koizumi S, Watanabe T, Hashimoto K. Photoinduced amphiphilic surface on polycrystalline anatase TiO2 thin lms. Langmuir 2000;16:704850. [199] Machida M, Norimoto K, Watanabe T, Hashimoto K, Fujishima A. The effect of SiO2 addition in super-hydrophilic property of TiO2 photocatalyst. J Mater Sci 1999;34:256974. [200] Hattori A, Kawahara T, Uemoto T, Suzuki F, Tada H, Ito S. Ultrathin SiOx lm coating effect on the wettability change of TiO2 surfaces in the presence and absence of UV light illumination. J Colloid Interface Sci 2000;232:4103. [201] Nakajima A, Hashimoto K, Watanabe T, Takai K, Yamauchi G, Fujishima A. Transparent superhydrophobic thin lms with self-cleaning properties. Langmuir 2000;16:70447. [202] Euvananont C, Junin C, Inpor K, Limthongkul P, Thanachayanont C. TiO2 optical coating layers for self-cleaning applications. Ceram Int 2008;34:106771. [203] Chen CC, Lin JS, Diau EWG, Liu TF. Self-cleaning characteristics on a thin-lm surface with nanotube arrays of anodic titanium oxide. Appl Phys A -Mater Sci Process 2008;92:61520. [204] Fujishima A, Zhang XT, Tryk DA. TiO2 photocatalysis and related surface phenomena. Surf Sci Rep 2008;63:51582. [205] Fujishima A, Zhang XT. Titanium dioxide photocatalysis: present situation and future approaches. C R Chim 2006;9:75060. [206] Sun RD, Nakajima A, Fujishima A, Watanabe T, Hashimoto K. Photoinduced surface wettability conversion of ZnO and TiO2 thin lms. J Phys Chem B 2001;105: 198490. [207] Liu H, Feng L, Zhai J, Jiang L, Zhu DB. Reversible wettability of a chemical vapor deposition prepared ZnO lm between superhydrophobicity and superhydrophilicity. Langmuir 2004;20:565961. [208] Kietzig AM, Hatzikiriakos SG, Englezos P. Patterned superhydrophobic metallic surfaces. Langmuir 2009;25:48217.

Das könnte Ihnen auch gefallen