Nano to Micro Structural Hierarchy Is Crucial for Stable Superhydrophobic and Water-Repellent Surfaces

Similar documents
Directional adhesion of superhydrophobic butterfly wings{

Wetting behaviours of a-c:h:si:o film coated nano-scale dual rough surface

Anti-icing surfaces based on enhanced self-propelled jumping of condensed water microdroplets

Numerical Simulation of Drops Impacting on Textured Surfaces

One-Step Preparation of Regular Micropearl Arrays for Two-Direction Controllable Anisotropic Wetting

Droplet Impact Simulation of Hydrophobic Patterned Surfaces by Computed Fluid Dynamics

Recently, the solid surface with the unusual wettability

Bioinspired surfaces for robust submerged superhydrophobicity: insights from molecular dynamics

Hierarchical roughness optimization for biomimetic superhydrophobic surfaces

Superhydrophobic surfaces. José Bico PMMH-ESPCI, Paris

Topography driven spreading. School of Biomedical & Natural Sciences, Nottingham Trent University. Clifton Lane, Nottingham NG11 8NS, UK.

Generalized Cassie-Baxter equation for wetting of a spherical droplet within a smooth and heterogeneous conical cavity

A study on wettability of the dual scale by plasma etch and nanohonycomb structure

Silicone brushes: Omniphobic Surfaces with Low Sliding Angle

Supporting Information

For rough surface,wenzel [26] proposed the equation for the effective contact angle θ e in terms of static contact angle θ s

Science and Technology, Dalian University of Technology, Dalian , P. R. China b

Electrowetting-Induced Dewetting Transitions on Superhydrophobic Surfaces

Multifunctionality and control of the crumpling and unfolding of

J. Bico, C. Tordeux and D. Quéré Laboratoire de Physique de la Matière Condensée, URA 792 du CNRS Collège de France Paris Cedex 05, France

A lichen protected by a super-hydrophobic and. breathable structure

Supporting Infromation

Groovy Drops: Effect of Groove Curvature on Spontaneous Capillary Flow

Wetting and Spreading of Drops on Rough Surfaces

Impalement of fakir drops

Modelling and analysis for contact angle hysteresis on rough surfaces

A Hydrophilic/Hydrophobic Janus Inverse-Opal

Lock-and-Key Geometry Effect of Patterned Surfaces: Wettability and Switching of Adhesive Force**

The Wilhelmy balance. How can we measure surface tension? Surface tension, contact angles and wettability. Measuring surface tension.

ANALYSIS FOR WETTING ON ROUGH SURFACES BY A THREE-DIMENSIONAL PHASE FIELD MODEL

Measurements of contact angles at subzero temperatures and implications for ice formation

Soft Matter

Enabling Self-propelled Condensate Flow During Phase-change Heat Rejection Using Surface Texturing

Superhydrophobicity and contact-line issues

Supporting Information

The Edge Termination Controlled Kinetics in Graphene. Chemical Vapor Deposition Growth

Support Information. A multi-functional oil/water separator from a selectively pre-wetted. superamphiphobic paper

Supporting Information

Fabrication of superhydrophobic AAO-Ag multilayer mimicking dragonfly wings

Fabrics with tunable oleophobicity

Direct Oil Recovery from Saturated Carbon Nanotube Sponges

The Pennsylvania State University The Graduate School Department of Chemical Engineering WETTING CHARACTERISTICS OF PHYSICALLY-PATTERNED

(Communicated by Aim Sciences)

Supporting Information

Modeling of Interfacial Debonding Induced by IC Crack for Concrete Beam-bonded with CFRP

(Communicated by Benedetto Piccoli)

Droplet Impact Dynamics on Lubricant-Infused Superhydrophobic Surfaces: The Role of Viscosity Ratio

The three-phase contact line shape and eccentricity effect of anisotropic wetting on hydrophobic surfaces

Nanosheet-Constructed Porous BiOCl with Dominant {001} Facets for Superior Photosensitized Degradation

7 Wettability of 60 kev Ar-ion irradiated rippled-si surfaces

c 2011 by Huan Li. All rights reserved.

Unusual Dual Superlyophobic Surfaces in Oil Water Systems: The Design Principles

Dr. Nikos Kehagias. Catalan Institute of Nanoscience and Nanotechnology, CSIC and The Barcelona Institute of Science and Technology, Barcelona, Spain

Supporting Information. Co 4 N Nanosheets Assembled Mesoporous Sphere as a Matrix for Ultrahigh Sulfur Content Lithium Sulfur Batteries

Projective synchronization of a complex network with different fractional order chaos nodes

A novel and inexpensive technique for creating superhydrophobic surfaces using Teflon and sandpaper

Noise Shielding Using Acoustic Metamaterials

Dewetting Transitions on Superhydrophobic Surfaces: When Are Wenzel Drops Reversible?

Deciphering buried air phases on natural and bioinspired superhydrophobic surfaces using synchrotron radiation-based X-ray phase-contrast imaging

Elastic Instability and Contact Angles on Hydrophobic Surfaces with Periodic Textures

Electronic Supplementary Information

Tuning the Shell Number of Multi-Shelled Metal Oxide. Hollow Fibers for Optimized Lithium Ion Storage

Bioassay on a Robust and Stretchable Extreme Wetting. Substrate through Vacuum-Based Droplet Manipulation

Self-assembled pancake-like hexagonal tungsten oxide with ordered mesopores for supercapacitors

Highly Stretchable and Transparent Thermistor Based on Self-Healing Double. Network Hydrogel

Faceted drops on heterogeneous surfaces

Femtosecond laser controlling underwater oil-adhesion of glass surface

Self-cleaning of hydrophobic rough surfaces by coalescence-induced wetting transition

Bioinspired Surfaces with Special Wettability TAOLEI SUN, LIN FENG, XUEFENG GAO, AND LEI JIANG*,,

Superoleophobic Cotton Textiles

Femtosecond Laser Weaving Superhydrophobic Patterned PDMS Surfaces with Tunable Adhesion

Immersed Superhydrophobic Surfaces: Gas exchange, slip and drag reduction properties

Kinetically-Enhanced Polysulfide Redox Reactions by Nb2O5. Nanocrystal for High-Rate Lithium Sulfur Battery

Thermodynamic equilibrium condition and model behaviours of sessile drop

COMPARISON OF WETTABILITY AND CAPILLARY EFFECT EVALUATED BY DIFFERENT CHARACTERIZING METHODS

Electronic Supplementary Information

Revelation of the Excellent Intrinsic Activity. Evolution Reaction in Alkaline Medium

Hydrothermally Activated Graphene Fiber Fabrics for Textile. Electrodes of Supercapacitors

An Advanced Anode Material for Sodium Ion. Batteries

Flexible Asymmetrical Solid-state Supercapacitors Based on Laboratory Filter Paper

Hierarchical MoO 2 /Mo 2 C/C Hybrid Nanowires for High-Rate and. Long-Life Anodes for Lithium-Ion Batteries. Supporting Information

Keywords Galvanic exchange reactions, superhydrophobic thin films, one-step process

Supporting Information

Wetting properties on nanostructured surfaces of cicada wings

On the characterization of crystallization and ice adhesion on smooth and rough surfaces using molecular dynamics

Microfluidics 2 Surface tension, contact angle, capillary flow

EXPERIMENTAL INVESTIGATION OF HIGH TEMPERATURE THERMAL CONTACT RESISTANCE WITH INTERFACE MATERIAL

An Experimental Investigation on Surface Water Transport and Ice Accreting Process Pertinent to Wind Turbine Icing Phenomena

Supporting Information

Wetting properties on nanostructured surfaces of cicada wings

Continuum modeling of van der Waals interactions between. carbon nanotube walls

MAKING OF SUPERHYDROPHOBIC SURFACES BY THE DEPOSITION OF NANOCOATIGS FROM SUPERCRITICAL CARBON DIOXIDE

Sta$s$cal mechanics of hystere$c capillary phenomena: predic$ons of contact angle on rough surfaces and liquid reten$on in unsaturated porous media

those research efforts, the number of scientific publications, patents and review articles in the field has also shown dramatic growth.

Relaxor characteristics of ferroelectric BaZr 0.2 Ti 0.8 O 3 ceramics

Supporting Information

Key words. Wenzel equation, Cassie equation, wetting, rough surface, homogenization

bifunctional electrocatalyst for overall water splitting

Depinning of 2d and 3d droplets blocked by a hydrophobic defect

Engineering of Hollow Core-Shell Interlinked Carbon Spheres for Highly Stable Lithium-Sulfur Batteries

Transcription:

pubs.acs.org/langmuir 2010 American Chemical Society Nano to Micro Structural Hierarchy Is Crucial for Stable Superhydrophobic and Water-Repellent Surfaces Yewang Su, Baohua Ji,*,, Kai Zhang, Huajian Gao, Yonggang Huang,*,^ and Kehchih Hwang Department of Engineering Mechanics, Tsinghua Universtiy, Beijing 100084, P. R. China, Division of Engineering, Brown University, Provindence, Rhode Island 02912, and ^Department of Civil and Environmental Engineering and Department of Mechanical Engineering, Northwestern University, Evanston, Illinois 60208. Present address: Biomechanics and Biomaterials Laboratory, Department of Applied Mechanics, Beijing Institute of Technology, Beijing 100081, P.R. China Received September 27, 2009. Revised Manuscript Received November 19, 2009 Water-repellent biological systems such as lotus leaves and water strider s legs exhibit two-level hierarchical surface structures with the smallest characteristic size on the order of a few hundreds nanometers. Here we show that such nano to micro structural hierarchy is crucial for a superhydrophobic and water-repellent surface. The first level structure at the scale of a few hundred nanometers allows the surface to sustain the highest pressure found in the natural environment of plants and insects in order to maintain a stable Cassie state. The second level structure leads to dramatic reduction in contact area, hence minimizing adhesion between water and the solid surface. The two level hierarchy further stabilizes the superhydrophobic state by enlarging the energy difference between the Cassie and the Wenzel states. The stability of Cassie state at the nanostructural scale also allows the higher level structures to restore superhydrophobicity easily after the impact of a rainfall. Introduction Nature has evolved hierarchical surface-structures to achieve superior water-repellent capabilities in some plants and insects, 1-6 allowing rain drops to roll freely on Lotus (Nelumbo nucifera) leaves 1,2 and water striders (Gerris remigis) to stand effortlessly and move quickly on water. 3,4 Lotus leaves exhibit two levels of hierarchy, 1,2,5 the first level having characteristic dimension on the order of 100-500 nm and the second level on the order of 20-100 μm (Figure 1a). Water strider s legs also show a two-level hierarchy with the smallest structure size around 400 nm 6 (Figure 1b). The present paper aims to address why these biological surfaces are hierarchically structured from nanoscale up and what are the underlying mechanisms that determine the length scales and the number of hierarchies involved. *Corresponding authors. E-mail: (B.J.) bhji@bit.edu.cn; (Y.H.) y-huang@ northwestern.edu (1) Barthlott, W.; Neinhuis, C. Planta 1997, 202, 1 8. (2) Neinhuis, C.; Barthlott, W. Ann. Bot. 1997, 79, 667 677. (3) Hu, D. L.; Chan, B.; Bush, J. W. M. Nature 2003, 424, 663 666. (4) Suter, R. B.; Rosenberg, O.; Loeb, S.; Wildman, H.; Long, J. H. J. Exp. Biol. 1997, 200, 2523 2538. (5) Zhang, L.; Zhou, Z. L.; Cheng, B.; DeSimone, J. M.; Samulski, E. T. Langmuir 2006, 22, 8576 8580. (6) Gao, X. F.; Jiang, L. Nature 2004, 432, 36 36. (7) Lafuma, A.; Quere, D. Nat. Mater. 2003, 2, 457 460. (8) Mohammadi, R.; Wassink, J.; Amirfazli, A. Langmuir 2004, 20, 9657 9662. (9) Oner, D.; McCarthy, T. J. Langmuir 2000, 16, 7777 7782. (10) Yoshimitsu, Z.; Nakajima, A.; Watanabe, T.; Hashimoto, K. Langmuir 2002, 18, 5818 5822. (11) Wang, Z.; Koratkar, N.; Ci, L.; Ajayan, P. M. Appl. Phys. Lett. 2007, 90, 143117. (12) Gao, L. C.; McCarthy, T. J. Langmuir 2006, 22, 5998 6000. (13) Wenzel, R. N. Ind. Eng. Chem 1936, 28, 988 994. (14) Cassie, A. B. D.; Baxter, S. Trans. Faraday Soc. 1944, 40, 546 551. (15) Johnson, R. E.; Dettre, R. H. In Contact angle, wettability, and adhesion; Gould, R. F., Ed.; American Chemical Society: Washington, D. C., 1964; pp 112-135. (16) Marmur, A. Langmuir 2004, 20, 3517 3519. (17) Patankar, N. A. Langmuir 2004, 20, 8209 8213. (18) Zheng, Q. S.; Yu, Y.; Zhao, Z. H. Langmuir 2005, 21, 12207 12212. (19) Carbone, G.; Mangialardi, L. Eur. Phys. J. E 2005, 16, 67 76. (20) Herminghaus, S. Europhys. Lett. 2000, 52, 165 170. During the past decade, many experimental 1-12 and theoretical studies 13-22 have been dedicated to understanding the so-called Lotus effect. The connections between surface roughness and hydrophobicity were first worked out by Wenzel 13 and Cassie and Baxter. 1,2,14 In the so-called Wenzel state, 13 water forms seamless contact with a surface and the contact angle changes with the surface roughness. In contrast, the Cassie state 14 refers to water forming incomplete contact with a rough surface with air trapped between the liquid and solid. Although both Wenzel and Cassie states can lead to a high contact angle of water droplets on a rough surface, there are large differences in their adhesion properties which, together with their hydrophobic properties, determine the superhydrophobic properties of the surface. [Normally, a surface is called superhydrophobic if the contact angle of a water droplet on the surface exceeds 150. 6-8 More rigorously, however, superhydrophobicity should be defined by both a large contact angle and a low roll-off angle (or low adhesion). 16,18,19 ] Experimental and theoretical studies 1,2,7,16 have shown that the Cassie state is more reasonable for the Lotus effect in view of its incomplete contact and relatively weak adhesion with water droplets. From this point of view, it is critically important for water-repellent biological surfaces to maintain a stable Cassie state even under harsh environmental conditions. A number of previous studies 7,15-21,23-31 have attempted to investigate the stability of the Cassie state and the transition between Cassie to Wenzel states. Johnson and Dettre 15 compared (21) Liu, B.; Lange, F. F. J. Colloid Interface Sci. 2006, 298, 899 909. (22) Otten, A.; Herminghaus, S. Langmuir 2004, 20, 2405 2408. (23) Barbieri, L.; Wagner, E.; Hoffmann, P. Langmuir 2007, 23, 1723 1734. (24) Callies, M.; Quere, D. Soft Matter 2005, 1, 55 61. (25) Cao, L. L.; Hu, H. H.; Gao, D. Langmuir 2007, 23, 4310 4314. (26) He, B.; Patankar, N. A.; Lee, J. Langmuir 2003, 19, 4999 5003. (27) Nosonovsky, M. Langmuir 2007, 23, 3157 3161. (28) Patankar, N. A. Langmuir 2003, 19, 1249 1253. (29) Quere, D. Rep. Prog. Phys. 2005, 68, 2495 2532. (30) Marmur, A. Langmuir 2003, 19, 8343 8348. (31) Yang, C.; Tartaglino, U.; Persson, B. N. J. Phys. Rev. Lett. 2006, 97, 116103. 4984 DOI: 10.1021/la9036452 Published on Web 01/21/2010

evolution toward a stable Cassie state under harsh environmental conditions. We will also show that the higher level structures in the hierarchy will further stabilize the Cassie state and also lead to dramatic reduction in contact area, thereby largely removing adhesion between water and the surface. Results and Discussion Nanostructure is Crucial for a Stable Cassie State. Let us first address the effect of structure size on the wetting state of a surface. Consider a two-dimensional model of water droplet in contact with a surface profile with periodic sinusoidal protrusions,15,19 8 < -h cosðkðx þ dþþ -d -λ=2 e x < -d ð1þ hs ðxþ ¼ -h -d e x < d : -h cosðkðx -dþþ d e x e d þ λ=2 Figure 1. Biological surfaces and theoretical model of the Cassie wetting state of an N-level hierarchical surface structure. (a) Twolevel hierarchical structure of Lotus leaves (N. nucifera)1,2 consisting of epidermal cells at microscale (papillae; bar = 20 μm) and a superimposed layer of hydrophobic wax crystals at nanoscale (see inset; bar = 1 μm);5 (b) Two-level hierarchical structure of water strider s leg (G. remigis)6 exhibiting microsetae (bars = 20 μm) with fine nanoscale grooves (see inset; bar = 200 nm). (c) A liquid drop in contact with an N-level hierarchical wavy surface, where air is trapped between the liquid and structure at each level. The lowest panel is for an illustration of the contact angle at the triple-line and the physical quantities involved in the problem. the free energies associated with the Cassie and Wenzel states of a sinusoidal surface and found that the Cassie state is preferred at large wave heights. The criterion of the stability of a Cassie state was discussed from the point of view of design of hydrophobic surfaces.25-28 Qu er e and co-workers7,24,29 studied the transition between the Cassie and Wenzel states by applying a pressure on a liquid drop on a microtextured surface and observed a transition in wetting state as the pressure is increased. They found that there is a critical pressure that induces an irreversible transition from Cassie to Wenzel states. Extending the work of Johnson and Dettre,15 Carbone and Mangialardi19 found a critical height of a sinusoidal surface structure beyond which the Cassie state becomes more stable than the Wenzel state, and determined the critical pressure for the Cassie-to-Wenzel (CW) transition. Zheng et al.18 performed more rigorous studies on the stability, metastability and instability of the Cassie and Wenzel states, as well as their transitions under pressure acting on the surface of a periodic array of micropillars. Recently, the effect of the characteristic size of the microstructure on the stability of the Cassie state has been explored.21,32,33 The roles of the hierarchy of surface structures were studied by Li and Amirfazli34 and Yu et al.35 However, few studies have been dedicated to the quantitative understanding of the questions why the biological surfaces are hierarchically structured from nanoscale up and what are the underlying mechanisms that determine the length scales and the number of hierarchies involved. The point of view adopted in the present study is that the nanometer length scale should be vital to the water-repellent properties of biological surfaces and this size selection might stem from natural (32) Reyssat, M.; Pepin, A.; Marty, F.; Chen, Y.; Quere, D. Europhys. Lett. 2006, 74, 306 312. (33) Tuteja, A.; Choi, W.; Ma, M. L.; Mabry, J. M.; Mazzella, S. A.; Rutledge, G. C.; McKinley, G. H.; Cohen, R. E. Science 2007, 318, 1618 1622. (34) Li, W.; Amirfazli, A. Soft Matter 2008, 4, 462 466. (35) Yu, Y.; Zhao, Z. H.; Zheng, Q. S. Langmuir 2007, 23, 8212 8216. where h is the height of the surface protrusions, λ is the surface wavelength, k = 2π/λ is the wavenumber, d is the half-distance between the bases of two adjacent protrusions (Figure 1c), and F = 2h/λ is the slenderness ratio of the structure. If the surface is in the Cassie state, the water will rest on top of the surface structure with air trapped between the liquid and solid under the support of surface tension. As long as the contact angle is smaller than 180, there will be a finite liquid-solid contact area near the top of the surface, as shown in Figure 1c. The liquid imposes a pressure P on the substrate as a result of an externally applied load or impact of a droplet on the surface (as in rainfall). Consideration of equilibrium at the triple-junction yields the following normalized relationship between the water pressure and properties (geometrical and chemical) of the sinusoidal protrusion (see the Supporting Information), P ¼ -h sin a cos θ0 -sin θ0 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 β 1 þ h sin2 a ð2þ Here, P = P/(kγLA) is the normalized pressure, γla is the liquid-air interfacial energy (similarly, γls and γsa denote the liquid-solid and solid-air interfacial energies, respectively); β = a þ d, β = kβ, a = ka, h = kh, and θ0 is the contact angle of the water droplet. This solution is valid only for the Cassie state with air trapped between the liquid and solid as shown in Figure 1c. To focus on the effect of structure size λ on the Cassie-to-Wenzel (CTW) transition, we set d = 0 in the following discussions without loss of generality. The projected fraction of contact area is then defined as f ¼ 1-2a=λ ¼ 1 -a=π ð3þ From eq 2, we can determine the evolution of the normalized pressure P as a function of the projected fraction of contact area f. Figure 2a shows five curves of such evolution for the sinusoidallike surface with different values of the slenderness ratio F. In each case, P increases as f increases at the initial stage, as expected. At a critical value fc, the pressure reaches a maximum value defined as the critical nondimensional pressure Pc. Further increase of pressure beyond this critical value induces an instability that causes the liquid bridge to collapse onto the solid surface, resulting in complete contact with the solid with fc equal to 1. This is the essence of the CTW transition. Note that we have defined the surface to be in the Cassie state when f < fc. The Wenzel state corresponds to f = 1. There exists an intermediate state for fc <f < 1, which is not stable according to our analysis. DOI: 10.1021/la9036452 4985

Figure 2. Critical pressure for water penetration into a sinusoidal surface as a function of the characteristic structural size. (a) Normalized pressure versus the contact area fraction as liquid bridges penetrate into the surface at different slenderness ratios of the surface protrusions. (b) Scaling laws of the penetrating pressure versus the structural size predicted by 2D theoretical model and 2D and 3D numerical simulations. The prediction of 2D numerical simulations with and without the gravity effects agrees well with the theoretical results. The penetrating pressure predicted from the 3D model is lower than that from the 2D model. According to eq 2, the normalized pressure depends on the contact angle θ 0, as well as the normalized parameters h (or the slenderness ratio F) and a (or f). However, the critical nondimensional pressure P c only depends on θ 0 and F (as shown in Figure 2a). According to the normalization P = P/(kγ LA ), the critical pressure can be expressed as P c ¼ P c kγ LA ¼ 2πP cðθ 0, FÞγ LA ð4þ λ which is inversely proportional to the surface wavelength. Therefore, the critical pressure for the CTW transition is inversely proportional to the characteristic size of the surface roughness when all other chemical (e.g., contact angle) and geometrical (e.g., the slenderness ratio) parameters of the surface are fixed. This size effect is consistent with the experimental studies by Reyssat et al., 32 and eq 4 considers more chemical and geometrical parameters than previous studies. For a given contact angle θ 0 (e.g., θ 0 =106 on a smooth wax surface) and slenderness ratio F, the smaller the structure size, the higher the critical pressure for the CTW transition and the more stable the Cassie state is, as shown in Figure 2b. This result suggests that reduction in size of the microstructure can significantly increase the critical pressure for the CTW transition, hence stabilizing the Cassie state. For Lotus leaves, the highest pressure found in nature comes from the impact by water droplets in a rainfall, which can be as large as 0.1 MPa. 36 The critical nondimensional pressure is defined once the chemical and normalized geometrical parameters of the surface, e.g., the contact angle and the slenderness ratio F, are given. The slenderness ratio of the nanostructure on Lotus leaves is generally 3-5 (Figure 1a), while that on water strider s legs is 1-3 (Figure 1b). For F = 3, the critical nondimensional pressure is found to be P c = 0.16. According to eq 4, under the highest impact pressure P = 0.1 MPa of rain drops, the critical structure size is calculated to be 740 nm (taking γ LA = 0.073J/m 2 ). Therefore, our analysis shows that the structure size λ must be smaller than λ c = 740 nm in order to maintain the Cassie state under the highest impact pressure from rain drops. To validate our theoretical model, we have numerically calculated the relationship between the critical size of the surface structure and pressure, with or without the effects of gravity, for both 2D (with 1D distribution of wavy protrusions, Figure 1c) and 3D surface configurations (with 2D distribution of revolving sinusoidal protrusions, as shown by the inset in Figure 2b) using the software package Surface Evolver. 37 The numerical results are compared with the theoretical predictions in Figure 2b. It is seen that the critical size calculated based on the 2D model is the same as our analytical result. The corresponding result based on the 3D model is smaller than the 2D value (the difference between 2D and 3D models can be rigorously obtained for a prismatic pillar structure). At the raindrop pressure of 0.1 MPa, the 3D model predicts a critical structure size of 400 nm. This result is in agreement with the observed smallest structure on Lotus leaves and suggests that the nanometer length scale may have indeed been evolved to maintain a stable Cassie state under the most severe pressure from raindrops. Similarly, nanostructures allow water striders to maintain their legs in a stable water-repellent state to be able to stand and maneuver on water. The Roles of Hierarchical Surface Structures. Recently, a number of studies 11,12,38,39 have shown that surfaces with a single level structure do not necessarily guarantee low adhesion even in the Cassie state, and some systems exhibit strongly adhesive surfaces no matter what the size of the surface structure is. However, the adhesion force dramatically decreases as soon as the second level structure is introduced. What are the underlying mechanisms of adhesion reduction via structural hierarchy? To answer this question, let us consider a self-similar hierarchical structure in the Cassie state shown in Figure 1c. The contact area fraction (the ratio of the contact area to the total surface area of the hierarchical structure) of a self-similar N-level structure is (for derivation see the Supporting Information) where S N ¼ YN s n n ¼1 Z x2 s n ¼ φða n, πþ, φðx 1, x 2 Þ¼ ð1þh 2 sin 2 xþ 1=2 dx=π ð6þ x 1 and a n denotes the position of the triple line at the nth level structure. Figure 3a shows that the contact area fraction exponentially (36) Erpul, G.; Norton, L. D.; Gabriels, D. Catena 2002, 47, 227 243. (37) Brakke, K. Surface Evolver. (38) Jin, M. H.; Feng, X. J.; Feng, L.; Sun, T. L.; Zhai, J.; Li, T. J.; Jiang, L. Adv. Mater. 2005, 17, 1977 1981. (39) Yeh, K. Y.; Chen, L. J.; Chang, J. Y. Langmuir 2008, 24, 245 251. ð5þ 4986 DOI: 10.1021/la9036452

state and useless for the hydrophobicity of biological surfaces). Generalizing eq 4, we have the critical size for higher level structure, λ N ¼ 2πP cðθ N -1, FÞγ LA P c ð8þ According to eqs s10 and s11 (Supporting Information), θ N-1 can be calculated as cos θ N -1 ¼ γ SA, N -1 -γ LS, N -1 γ LA ¼ - F CB N -1 γ LA ð9þ Figure 3. Effects of structural hierarchy on superhydrophobicity. (a) Contact area fraction as a function of the number of hierarchical levels under liquid pressures P = 0 and 146 Pa. (b) Energy difference ΔF N between the Wenzel and Cassie states as a function of the number of hierarchical levels under liquid pressures at P =0 and 146 Pa. The open squares in black stand for P = 0 and the solid triangles in red are stand for P = 146 Pa. θ 0 =106,andF =3. decreases as the number of hierarchical levels increases (We chose P=0 and 146 Pa in the calculation, where 146 Pa is the static pressure induced by the surface tension of a liquid drop of 2 mm in diameter). For example, the contact area fraction is S 1 =0.255 for the single level structure but becomes S 2 = 0.003 for the twolevel structure. Since a small contact area between liquid and solid is an important factor for adhesion reduction, 8,39 the structural hierarchy keeps the adhesion force sufficiently low for free rolling of water droplets on lotus leaves and walking of water striders on water. In addition, the hierarchical structure can significantly enhance the stability of the superhydrophobic state of a surface. Considering different wetting states of the self-similar N-level hierarchical structure shown in Figure 1c, we can calculate the difference in Helmholtz free energy per unit area (unit length in the 2D model) between the Cassie and the Wenzel states of the surface as (for derivation see the Supporting Information), ΔF N ¼ F W N ð7þ In this equation, F W N and F CB N are determined by the recursive equations F W N = F W N-1 φ(0, π) andf CB N = F CB N-1 φ (a N, π)þγ LA ψ(a N,1/P N ), where F W 0 = F CB 0 = γ LS -γ SA, ψ(x 1, x 2 )=x 2 arcsin (x 1 /x 2 )/π, a N denotes the position of the triple line at the Nth level structure, and P N denotes the normalized pressure of the Nth level. Figure 3b shows the relationship between this energy difference and the number of hierarchies in the cases of P =0 and 146 Pa. It is seen that the energy difference between the Cassie and the Wenzel states increases exponentially as the number of hierarchies increases. This suggests that the structural hierarchy also tends to enlarge the energy difference between the Wenzel and the Cassie states of the surface, thereby stabilizing the Cassie state. Why do biological surfaces only have two levels of hierarchies? To address this question, we assume that the higher level surface structures can sustain at least the static pressure of liquid droplet or the pressure (on the order of 100 Pa) induced by the weight of a water strider (10 dyn) 3 (otherwise they will be in the Wenzel -F CB N For the static pressure of 146 Pa, the critical size of the third level structure is calculated to be 2.6 mm. This size is larger than the size of a typical rain droplet of intermediate size (e.g., 2 mm), and therefore useless for enhancing superhydrophobicity. For the water strider, this critical size is much larger than the dimension of its leg (on the order of 100 μm). If one reduces the size of the third level structure by 1 order of magnitude, i.e., 260 μm, it will be close to the upper limit of the second level structure, which is 20-100 μm, as well as the dimension of the water strider s leg. The third level structure has no physical meaning either. This might be the reason why the surface structures in nature generally has only two levels of hierarchy. We therefore suggest that the static pressure of liquid determines the maximum number of hierarchies of the biological surfaces. The Reversibility of the Cassie State of Higher Level Structures. How do the higher-level structures, e.g., the second level structure, survive the impact from rain drops and maintain the Cassie state? This question can be understood by studying the Gibbs free energy of the system as the wetting state of the surface evolves under external pressure. The Gibbs free energy of the system loaded by a pressure P is written as G = F - PV, where V = hλ - V 0 is the volume of liquid penetrating into the structure, and V 0 is the volume of air trapped under the liquid (see Figure 1c and the Supporting Information). In the calculation of free energy, we choose P=146Paand0.1MPa,i.e.,thestatic pressure of a 2 mm water droplet and the raindrop pressure, respectively. Two different structure sizes λ = 200 nm and 20 μm are selected in the calculation. The upper panel in Figure 4 plots the Gibbs free energy per unit area of surface G = G/λ of a single level structure as a function of the area fraction of liquid-solid contact for the two different structure sizes. At the structure size of λ = 200 nm, the wetting state stays as a stable Cassie state with minimum free energy at f = f c under both values of the external pressure. This suggests that the external load is not able to overcome the forward energy barrier for the CTW transition at this sufficiently small structure size. However, for the large structure size of, λ = 20 μm, although it can maintain a stable Cassie state at the static pressure P =146 Pa (due to the forward energy barrier ΔG 1 f ), the wetting state would evolve from the Cassie state at f = f c to the Wenzel state at f = 1 under the raindrop pressure. Let us examine the wetting behavior of the surface with larger structure size λ =20μm when the surface is released from the previously loaded raindrop pressure back to the static pressure of the liquid. As the pressure is reduced from 0.1 MPa to 146 Pa, the wetting state immediately transfers from state A to state B as shown in the upper panel of Figure 4, which is still a Wenzel state. In order for the surface to restore its original Cassie state, i.e., from B to C, a reverse energy barrier from the Wenzel to the Cassie states, ΔG 1 r,mustbeovercome. DOI: 10.1021/la9036452 4987

Figure 4. Gibbs free energy per unit area (length in the 2D model) of the liquid-air-solid system (θ 0 =106 and F =3)asliquid penetrates into the surface. The Cassie state, Wenzel state and energy barrier between them can be clearly identified. The upper panel for a single level structure shows that the energy barrier can not be overcome when the structure size λ is 200 nm at the external pressure conditions P = 146 Pa and 0.1 MPa. However, the CTW transition spontaneously occurs at P = 0.1 MPa when the structure size is λ =20μm. The reverse energy barrier for Wenzel-to- Cassie transition G 1 r is about 1/20 of the surface tension of water; The lower panel is for a two-level hierarchical surface structure with λ 1 = 200 nm and λ 2 =20μm. The reverse energy barrier G 2 r of the second level structure is about 3 orders of magnitude smaller than G 1 r due to the presence of the first level nanostructure, which significantly enhances the reversibility of the Cassie state of the surface. The reverse energy barrier ΔG r for the Nth level structure can be calculated as, ΔG r N CB ¼ -F N -1 φð0, ac N Þþγ LA ψðac N,1=P NÞþγ LA P N V c N =ð2πþ ð10þ where a c N is the critical position of the triple line of the Nth level at which the structure reaches the maximum value of the c Gibbs free energy, and V 0 N is the corresponding value of V 0 at a = a c N. According to eq 10, we find that the reverse energy barrier decreases exponentially with increasing number of hierarchical levels (see Figure 5), a behavior similar to that of the contact area fraction. For the single-level structure of size 20 μm, the reverse energy barrier ΔG r 1 is about 1/20 of surface energy (see Figure 5). However, for the second level structure of hierarchical surface of the same size, the reverse energy barrier ΔG r 2 is 3 orders of magnitude smaller than ΔG r 1 (at the same time the forward energy barrier ΔG f 2 is much larger than ΔG f 1 ) (see the lower panel in Figure 4 and Figure 5). For example, the energy input needed for the transition at the second level structure is only 8 10-12 J. In nature, a disturbance by a wind of velocity 0.1 m/s is on the order of 4 10-8 J, which is big enough for inducing the transition. Therefore, the structural hierarchy allows the Cassie state to be restored after the impact of a rainfall, i.e., from B 0 to C 0. A very recent study by Boreyko and Chen 40 has demonstrated how a disturbance can restore superhydrophobicity from the Wenzel state. (40) Boreyko, J. B.; Chen, C.-H. Phys. Rev. Lett. 2009, 103, 174502. Figure 5. Relationship between the reverse energy barrier G N r and the structure level N in a self-similar hierarchical surface at P = 0 and 146 Pa. The open squares in black are results for P = 0, while the solid triangles in red are for P =146Pa. Conclusions In summary, the present study indicates that the nano- to microstructural hierarchy is essential for the superhydrophobic surfaces seen in water-repellent plants and insects. The nanostructure allows the surface to sustain the highest pressure in nature in order to maintain a secure Cassie state, while the microstructure significantly reduce the contact area, thereby largely removing adhesion between solid and fluid at the macroscopic level. The two level hierarchy further stabilizes the superhydrophobic state by enlarging the energy difference between the Cassie and the Wenzel states. The stability of Cassie state at the nanostructural scale also allows the higher level structures to restore the Cassie state easily after the impact of a rainfall. This study shows that the characteristic size of the smallest structure of biological surfaces is determined by dynamic loading (rainfall pressure), while the number of hierarchical levels seems to be determined by the static loading (static pressure of the liquid) in the natural environment of relevant plants and insects. The principle of such superhydrophobic surfaces can provide useful guidelines for engineering new materials for industrial applications. Compared to previous energetic studies of fractal surfaces, 15-20,23-28,30,31 the main contribution of our present study is that we have considered the roles of an external force (from either the impact pressure of rain drops or static pressure of a water droplet due to surface tension) in the stability of the Cassie state and the critical size of each level structure of the hierarchy. Our model provides feasible explanations for why the nanostructure of the biological surface can sustain the impact by rain drops and remains in a stable Cassie state. Our method has also been used to study the energy barrier between the Cassie state and the Wenzel state under the external force, and the effects of the higher level structure on energy barrier for the inverse state transition from the Wenzel to the Cassie, and to discuss the reversibility of the Cassie state (from the Wenzel state) of the surface at higher level structures after impact by the rain drops. Previous studies 41-44 have shown that the nanometer structure size and structural hierarchy play essential roles in mechanical properties of biological systems. There are interesting analogueies (41) Gao, H.; Ji, B.; Jager, I. L.; Arzt, E.; Fratzl, P. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 5597 5600. (42) Ji, B.; Gao, H. J. Mech. Phys. Solids 2004, 52, 1963 1990. (43) Gao, H.; Yao, H. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 7851 7856. (44) Yao, H.; Gao, H. J. Mech. Phys. Solids 2006, 54, 1120 1146. 4988 DOI: 10.1021/la9036452

between the roles of the hierarchical structures in wetting behaviors and those in mechanical properties of bone and shells as well as in adhesion ability of Gecko. For example, the nanometer scale renders the critical structure components in bone and shells insensitive to crack-like flaws, thereby allowing these materials to fail near their theoretical strength, 41,42 and hierarchical structures further enhance the fracture toughness of bone and shells. Also, the nanoscale fibrillar structures on the feet of Gecko can achieve theoretical strength of adhesion and structure hierarchy enlarges the adhesion energy so that Gecko can adhere and move along vertical walls and ceilings with strong adhesion ability. 43,44 In the superhydrophobicity of surface of plants and insects, nanostructure ensures the stability of the Cassie state and the hierarchical structure reduces the contact area and at the same time provides the robustness of Cassie state at higher scales. Therefore, it seems that nature uses similar strategies, i.e., with hierarchical design from nanoscale, to optimize or control different material properties. Acknowledgment. This research was supported from the National Natural Science Foundation of China through Grant No. 10502031, 10628205, 10732050, 10872115, and National Basic Research Program of China through Grant No. 2004CB619304 and 2007CB936803. The work of HG is partially supported by the A*Star Visiting Investigator Program Size Effects in Small Scale Materials hosted at the Institute of High Performance Computing in Singapore. Supporting Information Available: Text giving the mathematical derivations. This material is available free of charge via the Internet at http://pubs.acs.org. DOI: 10.1021/la9036452 4989