Direct Numerical Simulations of Turbulent Flows over Superhydrophobic Surfaces

Similar documents
Simulations of Turbulence over Superhydrophobic Surfaces

SECONDARY MOTION IN TURBULENT FLOWS OVER SUPERHYDROPHOBIC SURFACES

Drag Reduction in Turbulent Flows over Superhydrophobic Surfaces

Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows

arxiv: v1 [physics.flu-dyn] 2 Jun 2014

Application of wall forcing methods in a turbulent channel flow using Incompact3d

The JHU Turbulence Databases (JHTDB)

A dynamic global-coefficient subgrid-scale eddy-viscosity model for large-eddy simulation in complex geometries

Turbulence Modeling I!

The JHU Turbulence Databases (JHTDB)

Exercise 5: Exact Solutions to the Navier-Stokes Equations I

BOUNDARY LAYER ANALYSIS WITH NAVIER-STOKES EQUATION IN 2D CHANNEL FLOW

An evaluation of a conservative fourth order DNS code in turbulent channel flow

DIRECT NUMERICAL SIMULATION OF SPATIALLY DEVELOPING TURBULENT BOUNDARY LAYER FOR SKIN FRICTION DRAG REDUCTION BY WALL SURFACE-HEATING OR COOLING

Interpretation of the mechanism associated with turbulent drag reduction in terms of anisotropy invariants

Turbulent drag reduction by streamwise traveling waves

INFLUENCE OF THE BOUNDARY CONDITIONS ON A TEMPERATURE FIELD IN THE TURBULENT FLOW NEAR THE HEATED WALL

Slow viscous flow in a microchannel with similar and different superhydrophobic walls

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION

A combined application of the integral wall model and the rough wall rescaling-recycling method

arxiv: v1 [physics.flu-dyn] 24 Sep 2018

WALL RESOLUTION STUDY FOR DIRECT NUMERICAL SIMULATION OF TURBULENT CHANNEL FLOW USING A MULTIDOMAIN CHEBYSHEV GRID

Drag Reduction In Turbulent Flows Over Micropatterned Superhydrophobic Surfaces

Masters in Mechanical Engineering. Problems of incompressible viscous flow. 2µ dx y(y h)+ U h y 0 < y < h,

1. Introduction, tensors, kinematics

2.3 The Turbulent Flat Plate Boundary Layer


Experience with DNS of particulate flow using a variant of the immersed boundary method

ROLE OF THE VERTICAL PRESSURE GRADIENT IN WAVE BOUNDARY LAYERS

Large eddy simulation of turbulent flow over a backward-facing step: effect of inflow conditions

arxiv: v1 [physics.flu-dyn] 5 May 2009

Anisotropic grid-based formulas. for subgrid-scale models. By G.-H. Cottet 1 AND A. A. Wray

Fundamentals of Fluid Dynamics: Elementary Viscous Flow

Boundary layer flows The logarithmic law of the wall Mixing length model for turbulent viscosity

Numerical Heat and Mass Transfer

Preliminary Study of the Turbulence Structure in Supersonic Boundary Layers using DNS Data

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

Modeling Complex Flows! Direct Numerical Simulations! Computational Fluid Dynamics!

FLUID MECHANICS. Chapter 9 Flow over Immersed Bodies

Generation of initial fields for channel flow investigation

Simulating Drag Crisis for a Sphere Using Skin Friction Boundary Conditions

7. Basics of Turbulent Flow Figure 1.

The Effect of the DNS Data Averaging Time on the Accuracy of RANS-DNS Simulations

Sustainable Drag Reduction in Turbulent Taylor-Couette Flows by Depositing Sprayable Superhydrophobic Surfaces

Turbulent flow over anisotropic porous media

LARGE EDDY SIMULATION OF MASS TRANSFER ACROSS AN AIR-WATER INTERFACE AT HIGH SCHMIDT NUMBERS

Explicit algebraic Reynolds stress models for internal flows

Chapter 9: Differential Analysis

On the feasibility of merging LES with RANS for the near-wall region of attached turbulent flows

Turbulent boundary layer

Drag Reduction via Transversal Wave Motions of Structured Surfaces

Chapter 9: Differential Analysis of Fluid Flow

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Predicting natural transition using large eddy simulation

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

Numerical simulation of wave breaking in turbulent two-phase Couette flow

CFD based design and analysis of micro-structured. surfaces with application to drag and noise reduction

ENERGY PERFORMANCE IMPROVEMENT, FLOW BEHAVIOR AND HEAT TRANSFER INVESTIGATION IN A CIRCULAR TUBE WITH V-DOWNSTREAM DISCRETE BAFFLES

Fluid Mechanics II Viscosity and shear stresses

Wall treatments and wall functions

Numerical investigation of multiple slot jets in air knife coating

Hybrid LES RANS Method Based on an Explicit Algebraic Reynolds Stress Model

Fluid contacting a solid - Slip or no slip? Wonjae Choi NNF, MIT

Direct numerical simulation of self-similar turbulent boundary layers in adverse pressure gradients

Applications of parabolized stability equation for predicting transition position in boundary layers

Chapter 7 The Time-Dependent Navier-Stokes Equations Turbulent Flows

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

Wall turbulence with arbitrary mean velocity profiles

Direct Numerical Simulation of Jet Actuators for Boundary Layer Control

Simulating Interfacial Tension of a Falling. Drop in a Moving Mesh Framework

The Role of Splatting Effect in High Schmidt Number Turbulent Mass Transfer Across an Air-Water Interface

Resolving the dependence on free-stream values for the k-omega turbulence model

Velocity Fluctuations in a Particle-Laden Turbulent Flow over a Backward-Facing Step

Microfluidics 1 Basics, Laminar flow, shear and flow profiles

NEW SCALING LAWS FOR TURBULENT POISEUILLE FLOW WITH WALL TRANSPIRATION

CHAPTER 7 NUMERICAL MODELLING OF A SPIRAL HEAT EXCHANGER USING CFD TECHNIQUE

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

Turbulence - Theory and Modelling GROUP-STUDIES:

Simulation of T-junction using LBM and VOF ENERGY 224 Final Project Yifan Wang,

Reynolds stress budgets in Couette and boundary layer flows

A Finite-Element based Navier-Stokes Solver for LES

Chapter 6: Incompressible Inviscid Flow

COMPUTATIONAL FLUID DYNAMICS ANALYSIS OF A V-RIB WITH GAP ROUGHENED SOLAR AIR HEATER

ROLE OF LARGE SCALE MOTIONS IN TURBULENT POISEUILLE AND COUETTE FLOWS

WAVE-LENGTH-DEPENDENT REARRANGEMENT OF SECONDARY VORTICES OVER SUPERHYDROPHOBIC SURFACES WITH STREAMWISE GROOVES

Direct Numerical Simulations of converging-diverging channel flow

arxiv: v1 [physics.flu-dyn] 4 Aug 2014

A multiscale framework for lubrication analysis of bearings with textured surface

Flow Characterizations and Drag Reduction on the Hydrophobic Surface with a Full Covering Gas Film

Applied Mathematics and Mechanics (English Edition)

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

Reynolds number scaling of inertial particle statistics in turbulent channel flows

LES AND RANS STUDIES OF OSCILLATING FLOWS OVER FLAT PLATE

Shear Turbulence. Fabian Waleffe. Depts. of Mathematics and Engineering Physics. Wisconsin

Estimation of Turbulent Dissipation Rate Using 2D Data in Channel Flows

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh

NEAR-WALL TURBULENCE-BUBBLES INTERACTIONS IN A CHANNEL FLOW AT Re =400: A DNS/LES INVESTIGATION

DNS, LES, and wall-modeled LES of separating flow over periodic hills

Turbulence Momentum Transport and Prediction of the Reynolds Stress in Canonical Flows

LIQUID FILM THICKNESS OF OSCILLATING FLOW IN A MICRO TUBE

Transcription:

Under consideration for publication in J. Fluid Mech. 1 Direct Numerical Simulations of Turbulent Flows over Superhydrophobic Surfaces M I C H A E L B. M A R T E L L 1 J. B L A I R P E R O T A N D J O N A T H A N P. R O T H S T E I N 1 Department of Mechanical and Industrial Engineering, University of Massachusetts Amherst, MA, 01003 USA (Received?? and in revised form??) Direct numerical simulations are used to investigate the drag reducing performance of superhydrophobic surfaces in turbulent channel flow. Slip velocities, wall shear stresses, and Reynolds stresses are considered for a variety of superhydrophobic surface microfeature geometry configurations at a friction Reynolds number of Re τ = 180. For the largest micro-feature spacing of 90µm an average slip velocity over 75% of the bulk velocity is obtained, and the wall shear stress reduction is nearly found to be nearly 40%. The simulation results suggest that the mean velocity profile near the superhydrophobic wall continues to scale with the wall shear stress, but is offset by a slip velocity that increases with increasing microfeature spacing. 1. Introduction Significant effort has been placed on the development of surfaces which reduce the amount of drag experienced by a fluid as it passes over the surface. Drag reduction in turbulent flows can be achieved via a number of very different mechanisms including the addition of polymers (Lumley (1969)), the addition of bubbles (Sanders et al. (2006)), riblets (Bechert et al. (1997)), compliant walls (Hahn et al. (2002)), and active blowing and suction (Kim (1999)). Laminar drag reduction is much harder to achieve. Macroscale laminar drag reduction is possible with liquids using surface or fluid electric charges (Maynes & Webb (2003)), and via surface chemistry (surface hydrophobicity) (Tretheway & Meinhart (2002)). Recent work (see Ou et al. (2004); Ou & Rothstein (2005); P. Joseph et al. (2006)) has shown that liquid laminar drag reduction is achievable in larger channels using superhydrophobic surfaces. Superhydrophobic surfaces combine hydrophobic chemistry with micron scale topological features to produce slip lengths on the order of tens of microns. In this work it will be demonstrated that superhydrophobic surfaces can also produce significant drag reduction for liquids operating in the turbulent regime. Experiments by S. Gogte et al. (2005) and early results from Daniello & Rothstein (2008) were at high enough Reynolds numbers to be turbulent and showed some drag reduction could be acheived. A theoretical analysis by Fukagata et al. (2006) (that assumes classical scaling arguments still hold when a superhydrophobic surface is present) suggests how a small alteration of the laminar sublayer can affect the entire turbulent boundary layer and alter the subsequent drag. More recently, Min & Kim (2004, 2005) performed a direct numerical simulation (DNS) in a channel at a shear stress Reynolds number, Re τ = 180 in which the turbulence was fully resolved but the superhydrophobic surface itself was modeled. The ( surface boundary condition was modeled as u wall = b du dy ) wall, where b was an assumed

2 M. B. Martell, J. B. Perot and J. P. Rothstein (a) (b) Figure 1. Representations of the superhydrophobic surfaces. Subfigure (a) shows ridges, arranged on a plane with periodic boundary conditions, with width d, spacing w, and channel height H. Subfigure (b) shows posts, arranged similar to the ridges shown in (a). Posts are square and spaced evenly in the streamwise (x) and spanwise (z) directions. Note these diagrams are not to scale. slip length. In the current work, both the turbulence and the surface boundary conditions are fully resolved. At the Reynolds number simulated in this work, the surface microfeatures have a size that ranges from 17 + wall units wide to 34 + wall units wide, which is below the streak width of 100 + wall units. Superhydrophobic surfaces can be made using a variety of micro-features. In this work, regular arrays of micro-ridges and micro-posts are used, so that the micro-topology is very well characterized. However, superhydrophobic surfaces are also effective with random post arrangements (plasma etched plastics and the Lotus leaf being two common examples). The physical arrangements of micro-ridge and micro-post surfaces considered in this work are illustrated in Figure 1. The SHS is only placed on the bottom wall of the channel. The spacing and size of the micro-features allows a liquid-gas interface to be supported between the individual posts or ridges. The gaps between post/ridges allow a nonzero average velocity to be present on the surface (called the slip velocity) and can result in drag reduction. The role of hydrophobicity in superhydrophobic surfaces is simply to keep the liquid out of the gaps. Normal surface chemistry would tend to pull the fluid into the gaps via a strong capillary action. While the chemical hydrophobicity of the surface (or its magnitude as measured by the static contact angle) has absolutely no affect on the drag reduction properties of the superhydrophobic surface, it does dictate the maximum pressure that the free-surface can support before the free-surface fails and the liquid is forced (by the static pressure) to move into the gap (Ou et al. (2004)). In this investigation, only the liquid is simulated. The top surface of each micro-feature is taken to be a no-slip boundary on the liquid, and the suspended liquid-gas interface between the micro-features is simulated as a flat and shear-free boundary on the liquid. While the drag is locally zero on the free-surface (between the posts/ridges), it is much higher than normal over the post/ridges because the fluid rapidly decelerates there. The net drag is the sum of these two effects, and depends on which effect dominates. While it is often stated that the drag reduction is a result of reduced wall contact with the fluid, this is an incomplete explanation of the phenomena, as the spacing between the microfeatures is also critically important. The boundary assumptions on the liquid mean that the free-surface in these simulations is assumed to experience no out-of-plane deflection. The deflection be quantified by the maximum free-surface displacement, s. When the deflection ratio s/w (where w is the gap width) is greater than roughly 20% (the exact value depends on the hydrophobic contact angle) the surface will fail. For small deflections (< 20%) the deflection is well

DNS of Turbulent Flows over Superhydrophobic Surfaces 3 approximated by s/w (w p)/(8σ), where σ is the surface tension. For our maximum gap size of 90µm and water with σ = 7 10 2, this means that pressures up to 700 Pa produce small deflections (< 10%). The work of Ybert et al. (2007) confirms that the affect of small deflections (< 20%) on laminar drag reduction is negligible. The assumption of shear-free flow on the free-surface is reasonable if the micro-posts and micro-ridges are tall enough (i.e. same order of magnitude as the spacing). If the posts/ridges are significantly shorter than the gap width, then the circulating air in the gap between the micro-features could cause some drag on the free-surface and cause such a surface to produce less drag reduction (more total drag) than these DNS calculations predict. 2. Direct Numerical Simulations A CFD code was developed to perform direct numerical simulation of turbulent channel flow over surfaces with varying boundary conditions. The governing Navier-Stokes equations are numerically solved using a second-order accurate Cartesian staggered mesh method with classical projection for the pressure solution. A second-order accurate, three step, low storage Runge Kutta scheme is used to advance the solution in time. The mesh has non-uniform spacing in the vertical direction (normal to the surfaces), in order to allow greater resolution near the channel s top and bottom walls. A Cartesian mesh is well suited to the channel and micro-feature geometry being investigated. A staggered scheme is employed because it conserves mass and momentum, as well as vorticity and kinetic energy. There is no numerical viscosity or artifical dissipation in this scheme (Perot (2000)), which is vital to accurately predicting the turbulent energy cascade (see Mittal & Moin (1998)). The code is fully parallel, using MPI libraries, and optimized for execution on supercomputers. The computational domain is a box of ratio 6:2:3 (length:height:width). x is the streamwise direction of length L x, y the vertical direction of height L y, and z the spanwise direction of width L z. For the data presented in this paper, all resolutions are 128 grid points in each direction, except the 15µm-15µm case, which has 256 grid points in each direction. The boundary conditions consist of alternating regions of no-slip and no-shear, which coorespond to the top of the micro-feature (a post or ridge) and the interface supported between micro-features, respectively. Several ridge and post configurations were examined, where the feature width d and feature spacing w were varied (see Figure 1). Eight ridges (in the spanwise direction) were present for all ridge simulations, except for the 15µm-15µm case which had 16 across the channel. Similarly, a minimum of eight posts in the spanwise direction, and sixteen in the streamwise direction, were present for all post simulations. Note that in the case of ridges, the configurations are referred to by the ratio of their spanwise width (in z) to spanwise spacing (in z). For example, a ridge which is 30µm wide and spaced at 30µm is a 30µm-30µm ridge. All micro-ridge cases investigated involve uniformly spaced ridges of equal width. Similarly, all micropost cases investigated involve square posts (whose widths d in both the streamwise and spanwise direction are equal) that are uniformly spaced (spacing w in both the streamwise and spanwise direction is equal). For clarity, the symbols used for each geometry remain consistent throughout the figures and are: regular channel flow (benchmark) r; 15µm-15µm ridges ; 30µm-30µm ridges ; 30µm-50µm ridges ; 30µm-90µm ridges ; 30µm-30µm posts ; 30µm-50µm posts ; 30µm-90µm posts. In every case, the sum of each feature width and spacing (w+d) is kept constant at 16 grid points and 67.5 + wall units. The instantaneous variables calculated in the CFD code are averaged. As such, each subdomain around a single feature produces a set of temporally averaged velocity and

4 M. B. Martell, J. B. Perot and J. P. Rothstein pressure data, denoted ū t and p t. These temporal averages are then ensemble averaged (denoted by angle brackets) over all the post/ridges in the simulation to produce ū t and p t. The Reynolds stresses are calculated as R(x, y, z) = (u i ū t ) (u j ū t ) t. For graphical presentation, we calculate velocity averages over xz planes to remove the local micro-feature variations, U(y) = ū t xz = 1 Lx Lz L x L z ū t dxdz. Similarly, the planar 0 0 averaged Reynolds stresses are R ij (y) = R xz = 1 Lx Lz L x L z R dxdz. Finally, the bulk 0 0 velocity, U bulk = 1 + Ly 2 L y Ly 2 U(y)dy, is used as a velocity scale, noting the center of the channel is located at y = 0. The slip velocity, U slip = U( L y /2) is the xz planar average of the streamwise velocity at the lower wall. This slip velocity is a macroscopic value, as the actual velocities at the superhydrophobic surface are either zero (above the noslip feature) or some non-zero value (above the interface). The average wall shear stress ( ) τ w = (τw T +τw B )/2 is calculated by averaging the top wall shear stress τw T du = µ dy ( ) y=l y/2 and the bottom wall shear stress τw B = µ where µ is the viscosity. The du dy y= L y /2 top and bottom wall shear stresses are also macroscopic quantities. At statistical steadystate, the average wall shear stress τ w is directly related to the average channel pressure gradient dp dx by τ w = (2/L y ) ( ) dp dx. Note the pressure gradient and not the mass flux is prescribed in these calculations. 3. Validation Previous DNS of turbulent channel flow performed by Moser et al. (1998) was employed as a means of benchmarking the CFD code. In Figure 2 the planar averaged streamwise velocity profile and the four non-zero components of the planar averaged Reynolds stress tensor are compared to the results of Moser et al. (1998) with quantitative agreement at a friction Reynolds number Re τ = (u τ L y ) /2ν = 180. The friction velocity is given by u τ = (τ w /ρ) 1 2. The results are plotted against the non-dimensional channel height y + = u τ y/ν, where y is the vertical position in the channel. The left end of the axis in Figure 2 cooresponds to the bottom wall of the channel (at y = L y /2), and the right end corresponds to the center of the channel (at y = 0). For the standard channel flow, the velocity and Reynolds stress profiles are symmetric about the channel s vertical center plane. This assumption of symmetry does not apply to the subsequent simulations of channels with an superhydrophobic lower surface and standard upper surface. 4. Results Velocity profiles and Reynolds stresses were computed for the four different ridge cases. Figure 3(a) shows the mean velocity profiles normalized by the average friction velocity u τ. The channel height (y) is normalized by the channel half-height δ = L y /2. The standard channel flow data is provided (with a dashed line) as a reference. Results from the ridge cases show slip velocities and peak mean velocities that increase with increased width to spacing ratio d/w. In addition, the location of the peak velocity moves closer to the superhydrophobic surface. In Figure 3(b), the velocity profiles are shown close to the superhydrophobic surface and plotted against y +. In this figure both axis are scaled by the bottom wall shear stress τ B w. Note that the 15µm-15µm ridge case was performed with 256 grid points in the vertical (y) direction, which is why it has a data point closer to the superhydrophobic surface. This resolution increase was necessary in order to properly resolve the smaller feature sizes and spacings.

DNS of Turbulent Flows over Superhydrophobic Surfaces 5 (a) (b) Figure 2. Velocity profiles and Reynolds stresses calculated with the present code (r) for turbulent channel flow at Re τ = 180 compared with results of Moser et al. (1998) (symbols). (a) Near wall velocity profiles, scaled by u τ. (b) Planar averaged Reynolds stresses scaled by u 2 τ. R 11( ), R 22( ), R 33( ), and R 12( ). (a) (b) Figure 3. Ridge cases: Planar averaged velocity profiles. 15µm 15µm( ), 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ), with standard channel flow (r). (a) Velocity profiles scaled by the average friction velocity. (b) Velocity profiles near the superhydrophobic surface (left), scaled by the bottom wall (superhydrophobic surface) friction velocity. Symbols are employed to differentiate lines, and do not reflect actual data points. The Reynolds stresses R ij (y) are shown in Figure 4 and are scaled by the average friction velocity squared, u 2 τ. Since the mean pressure gradient remains constant in these simulations, a net reduction in drag or shear on the lower (superhydrophobic) boundary leads to a corresponding increase in the net shear on the top (regular) wall. With reduced shear, turbulence levels decrease near the superhydrophobic boundary. The turbulence levels increase on the regular boundary because of the higher shear now present near the upper wall. In the middle of the channel, the slope of the shear stress (R 12 ) must balance the fixed pressure gradient, and this is why these curves are all parallel in this region. Note that, due to the slip velocities present in the x and z directions, R 11 and R 33 become non-zero at the superhydrophobic surface. Figures 5 and 6 show U(y) and R ij (y) for posts. Similar to the results found with ridges, the slip velocities and peak mean velocities for the post cases increase with increased d/w. For posts, the slip velocities tend to be larger than those from equivalently sized

6 M. B. Martell, J. B. Perot and J. P. Rothstein (a) R 11 (b) R 22 (c) R 33 (d) R 12 Figure 4. Ridge cases: Reynolds stresses scaled by u 2 τ. 15µm 15µm( ), 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ), with standard channel flow (r). Symbols are employed to differentiate lines, and do not reflect actual data points. (a) (b) Figure 5. Post cases: Planar averaged velocity profiles. 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ), with standard channel flow (r). (a) Velocity profiles scaled by the average friction velocity. (b) Velocity profiles near the superhydrophobic surface (left), scaled by the bottom wall (superhydrophobic surface) friction velocity. Symbols are employed to differentiate lines, and do not reflect actual data points.

DNS of Turbulent Flows over Superhydrophobic Surfaces 7 (a) R 11 (b) R 22 (c) R 33 (d) R 12 Figure 6. Post cases: Reynolds stresses scaled by u 2 τ. 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ), with standard channel flow (r). Symbols are employed to differentiate lines, and do not reflect actual data points. or spaced ridges. Similar to results found with ridges, R 11 and R 33 become non-zero at the superhydrophobic surface, and peak values for R 11, R 22, and R 33 decrease near the superhydrophobic surface. The asymmetry present in R 11, R 22, and R 33 is more pronounced that that found in the ridged case. In addition, the peak in R 11 for both the 30µm 50µm and 30µm 90µm post cases moves so that the maximum occurs on the superhydrophobic surface itself. 5. Discussion The presence of superhydrophobic surface micro-features has a significant affect on the behavior of turbulent channel flow. Marked changes in the velocity profiles, Reynolds stresses, and wall shear stress are observed for a variety of micro-ridge and micro-post geometries. These results are consistent with the recent experimental work of Daniello & Rothstein (2008), who have investigated turbulent flow over superhydrophobic surfaces with similar micro-feature geometries. Figures 3(b) and 5(b) suggest that the essential scaling properties of turbulent boundary layers remain intact even when an superhydrophobic surface is present and significant drag reduction is occurring for that boundary layer. In those figures, u + = C + y + to

8 M. B. Martell, J. B. Perot and J. P. Rothstein Figure 7. R 12, scaled by the square of the bottom wall friction velocity, `u 2, B τ at the superhydrophobic surface, for streamwise ridges 15µm 15µm( ), 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ) and square uniformly spaced posts 30µm 30µm( ), 30µm 50µm( ), and 30µm 90µm( ), with standard channel flow R 12 (r). Symbols are employed to differentiate lines, and do not reflect actual data points. within 10% for all the cases considered (where C is the normalized slip velocity). For this result it was important to scale with the local (bottom) boundary shear stress. Figure 7 investigates if this scaling also holds for the turbulence as well as the mean flow. This figure shows R 12, profiles for all seven cases investigated, scaled by the bottom wall shear stress. The profiles all collapse onto the standard channel flow profile (r) near the bottom wall (superhydrophobic surface), suggesting that the structure of the near-wall turbulence has not fundamentally changed. Less surprisingly (and not shown), the same behavior is also exhibited at the top wall if the R 12 profiles are scaled by the top wall shear stress. The slip and drag reduction properties of the superhydrophobic surface are summarized in Figure 8. Drag reduction performance increases with increased feature spacing w. This trend is consistent for both ridges and posts, as the slip velocity attains nearly 65% of the bulk velocity for ridges, and over 75% of the bulk velocity for posts, with a width to spacing ratio just above 0.3. For equal ridge width to spacing ratios d/w, more slip is achieved with larger micro-features. This is clearly seen in Figure 8(a), where the slip velocity at 15µm 15µm ridges is nearly 40% lower than the slip velocity for the 30µm 30µm ridges. A similar trend for the drag reduction is found in Figure 8(b), where the bottom wall shear stress reduction for the 15µm 15µm ridges is nearly 30% lower that that for the 30µm 30µm ridges. For a width to spacing ratio d/w of unity, the 30µm 30µm ridges consistently outperform the 15µm 15µm ridges. This indicates that the actual size of the features, and not simply the ratio of width to spacing (or the percentage surface area contacted), plays an important role in the surface s dragreduction performance. Smaller features lead to diminished drag-reduction performance. When comparing posts and ridges, it is clear that for a given ratio of micro-feature size to spacing, d/w, posts yield higher slip velocities and larger shear stress reductions when compared to ridges. The performance advantage of posts over ridges appears to increase with increased feature spacing, w. The authors would like to thank the Office of Naval Research for support of this research under grant N00014-06-1-0497.

DNS of Turbulent Flows over Superhydrophobic Surfaces 9 (a) (b) Figure 8. Drag reduction performance of the superhydrophobic surface with ridges and posts. See Figure 1 for definitions of d and w. (a) Slip velocity at the superhydrophobic surface as a percentage of the bulk velocity for varying feature width and spacing. (b) Wall shear stress reduction at the superhydrophobic surface, normalized by the average wall shear stress, for varying feature width and spacing. Symbols are connected by lines to illustrate trends. REFERENCES Bechert, D.W., Bruse, M., Hage, W., G., J.G.T.VanderHoeven & Hoppe 1997 Experiments on drag-reducing surfaces and their optimization with an adjustable geometry. Journal of Fluids Mechanics 338, 59 87. Daniello, R. & Rothstein, J. P. 2008 Turbulent drag reduction using ultrahydrophobic surfaces. Submitted to Physical Review Letters. Fukagata, K., Kasagi, N. & Koumoutsakos, P. 2006 A theoretical prediction of friction drag in turbulent flow by superhydrophobic surfaces. Physics of Fluids 18, 051703:1 051703:4. Hahn, S., Je, J. & Choi, H. 2002 Direct numerical simulation of turbulent channel flow with permeable walls. Journal of Fluid Mechanics 450, 259 285. Kim, J. 1999 Active control of turbulent boundary layers for drag reduction. Lecture Notes in Physics 529, 142 152. Lumley, J.L. 1969 Drag reduction by additives. Annual Review of Fluid Mechanics 1, 367. Maynes, D. & Webb, B. W. 2003 Fully developed electro-osmotic heat transfer in microchannels. International Journal of Heat and Mass Transfer 46 (8). Min, T. & Kim, J. 2004 Effects of hydrophobic surface on skin-friction drag. Physics of Fluids 16 (7), L55 L58. Min, T. & Kim, J. 2005 Effects of hydrophobic surface on stability and transition. Physics of Fluids 17, 108106:1 108106:4. Mittal, R. & Moin, P. 1998 Suitability of upwind-biased finite difference schemes for largeeddy simulation of turbulent flows. American Institute of Aeronautics and Astronautics Jorunal 35 (8). Moser, R., Kim, J. & Mansour, N. 1998 Direct numerical simulation of turbulent channel flow up to Re τ =590. Physics of Fluids 11 (4), 943 945. Ou, J., Perot, J.B. & Rothstein, J. 2004 Laminar drag reduction in microchannels using ultrahydrophobic surfaces. Physics of Fluids 16 (12), 4635 4643. Ou, J. & Rothstein, J. 2005 Direct velocity measurements of the flow past drag-reducing ultrahydrophobic surfaces. Physics of Fluids 17 (10), 13606:2 13606:10. P. Joseph, 1 C. Cottin-Bizonne, Benot, J.-M., Ybert, C., Journet, C., Tabeling, P., & Bocquet, L. 2006 Slippage ofwater past superhydrophobic carbon nanotube forests in microchannels. Physical Review Letters 97 (15), 156104 1:156104 4.

10 M. B. Martell, J. B. Perot and J. P. Rothstein Perot, J. B. 2000 Conservation properties of unstructured staggered mesh schemes. Journal of Compututational Physics 159, 58 89. S. Gogte, P. Vorobieff, Truesdell, R., Mammoli, A., van Swol, F., Shah, P. & Brinker, C. J. 2005 Effective slip on textured superhydrophobic surfaces. Physics of Fluids 17, 051701:1 051701:4. Sanders, W.C., Winkel, E.S., Dowling, D.R., Perlin, M. & Ceccio, S.L. 2006 Bubble friction drag reduction in a high-reynolds-number flat-plate turbulent boundary layer. Journal of Fluid Mechanics 552, 353 380. Tretheway, D.C. & Meinhart, C.D. 2002 Apparent fluid slip at hydrophobic microchannel walls. Physics of Fluids 14 (3). Ybert, C., Barentin, C. & Cottin-Bizonne, C. 2007 Acheiving large slip with superhydrophobic surfaces:scaling laws for generic geometries. Physics of Fluids 19, 123601:1 123601:10.