Hostname: page-component-745bb68f8f-hvd4g Total loading time: 0 Render date: 2025-02-06T06:32:00.385Z Has data issue: false hasContentIssue false

On drag reduction scaling and sustainability bounds of superhydrophobic surfaces in high Reynolds number turbulent flows

Published online by Cambridge University Press:  07 February 2019

Amirreza Rastegari
Affiliation:
Department of Mechanical Engineering, The University of Michigan, Ann Arbor, MI 48109-2125, USA
Rayhaneh Akhavan*
Affiliation:
Department of Mechanical Engineering, The University of Michigan, Ann Arbor, MI 48109-2125, USA
*
Email address for correspondence: raa@umich.edu

Abstract

The drag reduction characteristics and sustainability bounds of superhydrophobic (SH) surfaces in high Reynolds number turbulent flows are investigated using results from direct numerical simulation (DNS) and scaling-law analysis. The DNS studies were performed, using lattice Boltzmann methods, in turbulent channel flows at bulk Reynolds numbers of $Re_{b}=3600$ ($Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$) and $Re_{b}=7860$ ($Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$) with SH longitudinal microgrooves or SH aligned microposts on the walls. Surface microtexture geometrical parameters corresponding to microgroove widths or micropost spacings of $4\lesssim g^{+0}\lesssim 128$ in base flow wall units and solid fractions of $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ were investigated at interface protrusion angles of $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$. Analysis of the governing equations and DNS results shows that the magnitude of drag reduction is not only a function of the geometry and size of the surface microtexture in wall units, but also the Reynolds number of the base flow. A Reynolds number independent measure of drag reduction can be constructed by parameterizing the magnitude of drag reduction in terms of the friction coefficient of the base flow and the shift, $(B-B_{0})$, in the intercept of a logarithmic law representation of the mean velocity profile in the flow with SH walls compared to the base flow, where $(B-B_{0})$ is Reynolds number independent. The scaling laws for $(B-B_{0})$, in terms of the geometrical parameters of the surface microtexture in wall units, are presented for SH longitudinal microgrooves and aligned microposts. The same scaling laws are found to also apply to liquid-infused (LI) surfaces as long as the viscosity ratios are large, $N\equiv \unicode[STIX]{x1D707}_{o}/\unicode[STIX]{x1D707}_{i}\gtrsim 10$. These scaling laws, in conjunction with the parametrization of drag reduction in terms of $(B-B_{0})$, allow for a priori prediction of the magnitude of drag reduction with SH or LI surfaces in turbulent flow at any Reynolds number. For the most stable of these SH surface microtextures, namely, longitudinal microgrooves, the pressure stability bounds of the SH surface under the pressure loads of turbulent flow are investigated. It is shown that the pressure stability bounds of SH surfaces are also significantly curtailed with increasing Reynolds number of the flow. Using these scaling laws, the narrow range of SH surface geometrical parameters which can yield large drag reduction as well as sustainability in high Reynolds number turbulent flows is identified.

Type
JFM Papers
Copyright
© 2019 Cambridge University Press 

1 Introduction

Superhydrophobic (SH) surfaces have received much attention as a means of skin-friction drag reduction in wall-bounded turbulent flows in recent years (Rothstein Reference Rothstein2010). These are surfaces with apparent receding contact angle exceeding a certain value, e.g. $150^{\circ }$ (Schellenberger et al. Reference Schellenberger, Encinas, Vollmer and Butt2016), which repel liquids by trapping gas inside the nano- or micro-scale features of a textured hydrophobic surface. The entrapped gas prevents direct contact between the liquid and the wall, leading to the so-called Cassie–Baxter state, and providing a mechanism for liquid slip at the wall. This effective slip has been shown to be the primary mechanism of skin-friction drag reduction with SH surfaces in both the laminar and turbulent flow regimes (Rastegari & Akhavan Reference Rastegari and Akhavan2015, Reference Rastegari and Akhavan2018a ).

Drag reductions of up to 50 % (Daniello, Waterhouse & Rothstein Reference Daniello, Waterhouse and Rothstein2009) and up to 75 % (Park, Sun & Kim Reference Park, Sun and Kim2014) have been reported in laboratory-scale experiments in turbulent channel flows and turbulent boundary layer flows, at $135\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 250$ , using SH longitudinal microgrooves with solid fractions of $\unicode[STIX]{x1D719}_{s}=0.5$ and $\unicode[STIX]{x1D719}_{s}=0.05$ , respectively, where $Re_{\unicode[STIX]{x1D70F}_{0}}\equiv u_{\unicode[STIX]{x1D70F}_{0}}\unicode[STIX]{x1D6FF}/\unicode[STIX]{x1D708}$ is the friction Reynolds number in the base flow, $u_{\unicode[STIX]{x1D70F}_{0}}$ is the wall friction velocity in the base flow, $\unicode[STIX]{x1D6FF}$ is the boundary layer thickness or channel half-height and $\unicode[STIX]{x1D708}$ is the kinematic viscosity. It has been noted that the magnitude of drag reduction increases with increasing width of the surface microtexture indentations, $g$ , and decreasing solid fraction, $\unicode[STIX]{x1D719}_{s}$ , approaching a maximum drag reduction of $DR\sim (1-\unicode[STIX]{x1D719}_{s})$ in the limit of large $g$ (Daniello et al. Reference Daniello, Waterhouse and Rothstein2009; Rothstein Reference Rothstein2010; Park, Park & Kim Reference Park, Park and Kim2013). Scaling laws have been proposed for the slip length and the slip velocity in terms of $g$ and $\unicode[STIX]{x1D719}_{s}$ in both the laminar (Ybert et al. Reference Ybert, Barentin, Cottin-Bizonne, Joseph and Bocquet2007) and turbulent (Seo & Mani Reference Seo and Mani2016) flow regimes. However, the relation between these scaling laws and the magnitude of drag reduction has not been clarified.

These drag reduction capabilities of SH surfaces, however, are lost if the SH surface undergoes a wetting transition, from the Cassie–Baxter state to a Wenzel state, whereby the trapped gas pockets inside the surface microtexture are depleted and replaced with the working fluid. This wetting transition can occur because of (i) depinning of the contact line and/or sagging of the liquid–gas interface under high local pressures (Zheng, Yu & Zhao Reference Zheng, Yu and Zhao2005; Checco et al. Reference Checco, Ocko, Rahman, Black, Tasinkevych, Giacomello and Dietrich2014) or (ii) diffusion or entrainment of the gas layer into the working liquid under high local shear rates (Samaha, Tafreshi & Gad-el Hak Reference Samaha, Tafreshi and Gad-el Hak2012; Karatay, Tsai & Lammertink Reference Karatay, Tsai and Lammertink2013; Wexler, Jacobi & Stone Reference Wexler, Jacobi and Stone2015; Ling et al. Reference Ling, Katz, Fu and Hultmark2017) of turbulent flow.

There is also anecdotal evidence, from both direct numerical simulation (DNS) and experiments, which suggests that, for a given geometry and size of the surface microtexture in wall units, the drag reduction performance of SH surfaces degrades with increasing Reynolds number of the base flow (Rastegari & Akhavan Reference Rastegari and Akhavan2018a ). Here, and throughout this study, ‘wall units’ refers to normalization with respect to the wall friction velocity of the drag reduced flow, $u_{\unicode[STIX]{x1D70F}}$ , and the kinematic viscosity, $\unicode[STIX]{x1D708}$ , and quantities normalized in this manner are denoted by a $+$ superscript. This degradation in drag reduction has been recently demonstrated in DNS studies of turbulent channel flows with SH longitudinal microgrooves at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$ (Rastegari & Akhavan Reference Rastegari and Akhavan2018a ), and can also be observed in earlier DNS studies of turbulent channel flows with SH longitudinal microgrooves at $180\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 590$ (Park et al. Reference Park, Park and Kim2013, figure 8b), even though the authors do not comment on it. Evidence of this degradation in drag reduction can also be found in recent experimental studies in turbulent boundary layer flows with spray coated SH walls, where a given upwards shift of the logarithmic layer in wall units, corresponding to $\unicode[STIX]{x0394}U^{+}=1.1$ , was found to result in 12 % drag reduction at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863$ , but only 11 % drag reduction at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1408$ (Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016).

While other experimental studies have reported an enhancement of drag reduction with increasing Reynolds number in SH turbulent boundary layer flows (Zhang et al. Reference Zhang, Tian, Yao, Hao and Jiang2015) or in SH or liquid-infused (LI) turbulent Taylor–Couette flows (Srinivasan et al. Reference Srinivasan, Kleingartner, Gilbert, Cohen, Milne and McKinley2015; Van Buren & Smits Reference Van Buren and Smits2017), it should be noted that these experiments were not performed with a fixed size of the surface microtexture in wall units at different Reynolds numbers. Instead, they used a given surface microtexture with a fixed size in ‘physical units’ at different Reynolds numbers, which leads to an increase in the size of the surface microtexture in wall units as the Reynolds number increases. As will be discussed in detail in §§ 3 and 4, an increase in the size of the surface microtexture in wall units can affect the drag reduction in a number of conflicting and competing ways. In the absence of a wetting transition or roughness effects, an increase in the size of the surface microtexture in wall units enhances the magnitude of drag reduction. However, an increase in the size of the surface microtexture in wall units also makes the surface more susceptible to a wetting transition, at which point the surface microtexture can begin to act as surface roughness and cause a drag increase. In addition, for random surface microtextures, where the height of the surface micro-features is variable, the protrusion heights of the surface elements above the liquid/gas interface will increase in wall units as the Reynolds number increases. This can also give rise to roughness effects with increasing Reynolds number, which can degrade the drag reduction. The interplay between these competing effects determines the trends in drag reduction when a surface microtexture of a given size in physical units is employed at different Reynolds numbers. Depending on which features are at play or dominate, the magnitude of drag reduction can be enhanced with increasing Reynolds number (Srinivasan et al. Reference Srinivasan, Kleingartner, Gilbert, Cohen, Milne and McKinley2015; Zhang et al. Reference Zhang, Tian, Yao, Hao and Jiang2015; Van Buren & Smits Reference Van Buren and Smits2017), or degrade (Bidkar et al. Reference Bidkar, Leblanc, Kulkarni, Bahadur, Ceccio and Perlin2014; Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016; Reholon & Ghaemi Reference Reholon and Ghaemi2018), or show enhancement at low Reynolds numbers followed by degradation at higher Reynolds numbers (Aljallis et al. Reference Aljallis, Sarshar, Datla, Sikka, Jones and Choi2013; Bidkar et al. Reference Bidkar, Leblanc, Kulkarni, Bahadur, Ceccio and Perlin2014; Gose et al. Reference Gose, Golovin, Boban, Mabry, Tuteja, Perlin and Ceccio2018). Collectively, the available experimental data in high Reynolds number SH turbulent boundary layer flows or SH turbulent channel flows, performed with a fixed size of the surface microtexture in physical units, show a trend towards enhancement of drag reduction with increasing Reynolds number up to a maximum of ${\sim}30\,\%$ drag reduction at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1500$ (Aljallis et al. Reference Aljallis, Sarshar, Datla, Sikka, Jones and Choi2013; Zhang et al. Reference Zhang, Tian, Yao, Hao and Jiang2015; Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016), followed by degradation of drag reduction at higher Reynolds numbers, approaching negative drag reductions at $Re_{\unicode[STIX]{x1D70F}_{0}}\gtrsim 5000$ (Aljallis et al. Reference Aljallis, Sarshar, Datla, Sikka, Jones and Choi2013; Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016). These results all point to the need for an understanding of the scaling of drag reduction and the sustainability bounds of SH and LI surfaces with surface microtexture and flow parameters in turbulent flow.

The degradation in the drag reduction performance of microtextured surfaces with increasing Reynolds number of the base flow, for a fixed geometry and size of the surface microtexture in wall units, is a well-known feature of riblets (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997; Spalart & McLean Reference Spalart and McLean2011), where the highest drag reductions achieved in high Reynolds number turbulent flows of practical interest, with optimal ‘blade riblets’ of spacing ${\sim}15$ wall units (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997), were found to be nearly half the highest drag reductions, of ${\sim}10\,\%$ , obtained with the same geometry and size of the riblets in wall units in laboratory-scale experiments (Spalart & McLean Reference Spalart and McLean2011). It has been suggested (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997; Spalart & McLean Reference Spalart and McLean2011; García-Mayoral & Jiménez Reference García-Mayoral and Jiménez2012) that a Reynolds number independent measure of drag reduction can be constructed by parameterizing the magnitude of drag reduction in terms of the friction coefficient of the base flow, $C_{f_{0}}$ , and the shift, $(B-B_{0})$ , in the intercepts, $B$ and $B_{0}$ , of log-law representations of the normalized mean velocity profiles in the drag reduced and the base flow, respectively, where $(B-B_{0})$ is Reynolds number independent and only a function of the geometry and size of the surface microtexture in wall units. In Rastegari & Akhavan (Reference Rastegari and Akhavan2018a ), this formulation was extended to SH surfaces to obtain a parameterization of the magnitude of SH drag reduction in terms of $C_{f_{0}}$ and the shift $(B-B_{0})$ .

While such a parametrization allows the drag reduction results from low Reynolds number turbulent flows to be extrapolated to higher Reynolds number flows of practical interest (Rastegari & Akhavan Reference Rastegari and Akhavan2018a ), developing optimal designs of SH surfaces also requires an understanding of the scaling laws for $(B-B_{0})$ in terms of the geometrical parameters of the surface microtexture, as well as the sustainability bounds of SH surfaces in turbulent flow.

In this study, scaling arguments are combined with results from DNS of SH turbulent channel flows at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$ , performed with a wide range of SH longitudinal microgroove or aligned micropost surface microtexture geometrical parameters, to demonstrate the Reynolds number dependence of the drag reduction, and the Reynolds number independence of $(B-B_{0})$ , and to present the scaling laws for $(B-B_{0})$ in turbulent flow in terms of the geometrical parameters of the SH surface in wall units. The combination of these parametrizations allow for a priori prediction of the magnitude of drag reduction with SH surfaces in turbulent flow at any Reynolds number. For the most stable of such SH surfaces, namely longitudinal microgrooves, we further present the pressure stability bounds of the SH surface in turbulent flow in terms of the Weber number, the geometrical parameters of the SH surface in wall units and the friction Reynolds number of the flow. These parameterizations are used to identify the narrow range of SH surface geometrical parameters which can ensure large drag reduction as well as sustainability in high Reynolds number turbulent flows of practical interest.

2 Numerical methods and simulation parameters

The DNS studies were performed in turbulent channel flows using standard D3Q19 (three-dimensional, 19 discrete velocity), single-relaxation-time lattice Boltzmann methods (Succi Reference Succi2001) with grid embedding (Lagrava et al. Reference Lagrava, Malaspinas, Latt and Chopard2012), of grid ratio $4:1$ , in the near-wall regions to better resolve the flow features near the SH walls. The resulting grid spacings were $\unicode[STIX]{x1D6E5}_{f}^{+0}\approx 0.5$ for $z^{+0}\lesssim 30$ , and $\unicode[STIX]{x1D6E5}_{c}^{+0}\approx 2$ for $z^{+0}\gtrsim 30$ in all three directions in all the simulations, where $\unicode[STIX]{x1D6E5}_{f}$ and $\unicode[STIX]{x1D6E5}_{c}$ denote the grid spacings on the fine grid and the coarse grid, respectively, and the $+0$ superscript denotes normalization with respect to the kinematic viscosity, $\unicode[STIX]{x1D708}$ , and the wall-friction velocity, $u_{\unicode[STIX]{x1D70F}_{0}}$ , of a ‘base’ turbulent channel flow with smooth, no-slip walls at the same bulk Reynolds number as the channel flow with SH walls. To preserve the order of accuracy of the lattice Boltzmann method, and ensure smoothness and continuity of the variables at the transition between the coarse and the fine grids, the two grids were overlapped for one cell width of the coarse grid, as suggested by Lagrava et al. (Reference Lagrava, Malaspinas, Latt and Chopard2012), and the interpolations between the two grids were performed using third-order bi-cubic Hermite splines in space, and second-order Hermite polynomials in time. The details of the numerical methods and verification studies performed to ensure the accuracy of the numerical methods and adequacy of the domain sizes and grid resolutions are described in the appendix of Rastegari & Akhavan (Reference Rastegari and Akhavan2018a ).

The liquid/gas interfaces on SH walls were modelled as stationary, curved or flat shear-free boundaries, with the shape of the meniscus determined from an analytical solution of the Young–Laplace equation (Rastegari & Akhavan Reference Rastegari and Akhavan2018a ). Two interface protrusion angles of $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ were investigated, corresponding to liquid/gas interfaces which were either flat or at maximum advancing contact angle ( $\unicode[STIX]{x1D703}_{c}=\unicode[STIX]{x1D703}_{F,adv}=120^{\circ }$ ), respectively.

The simulations were performed in channels of size $5h\times 2.5h\times 2h$ in the streamwise ( $x$ ), spanwise ( $y$ ) and wall-normal ( $z$ ) directions, respectively, as shown in figure 1. A constant flow rate was maintained in the channel during the course of all simulations, corresponding to bulk Reynolds numbers of $Re_{b}\equiv q/2\unicode[STIX]{x1D708}=3600$ or $Re_{b}=7860$ , where $q$ denotes the flow rate per unit spanwise width in the channel. For reference, DNS studies were also performed in base turbulent channel flows with smooth, no-slip walls at the same $Re_{b}$ as the SH channels. The corresponding friction Reynolds numbers in the base turbulent channel flows were $Re_{\unicode[STIX]{x1D70F}_{0}}\equiv u_{\unicode[STIX]{x1D70F}_{0}}h/\unicode[STIX]{x1D708}\approx 222$ at $Re_{b}=3600$ , and $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ at $Re_{b}=7860$ .

Figure 1. Schematic of the channels with SH walls, the coordinate system and the computational grid: (a) SH longitudinal microgrooves; (b) SH aligned microposts; (c) computational grid and protrusion angle.

A total of 53 simulations were performed with SH longitudinal microgrooves, covering the range of microgroove widths $4\lesssim g^{+0}\lesssim 128$ , solid fractions of $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ , protrusion angles of $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ and bulk Reynolds numbers of $Re_{b}=3600$ ( $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ ) and $Re_{b}=7860$ ( $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ ), as shown in table 1. In addition, as shown in table 1, four simulations were also performed with SH aligned square microposts with $g^{+0}\approx 28,56$ , $\unicode[STIX]{x1D719}_{s}=1/64$ and $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ at $Re_{b}=3600$ ( $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ ) and $Re_{b}=7860$ ( $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ ).

Table 1. Summary of simulation parameters and resulting drag reductions, $DR$ , in the present study.

3 The scaling of drag reduction with surface microtexture and flow parameters

It has been suggested that in drag reduction with microtextured surfaces, be they riblets, SH surfaces or liquid-infused (LI) surfaces, the magnitude of drag reduction depends not only on the geometry and size of the surface microtexture in wall units, but also the Reynolds number of the base flow (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997; Spalart & McLean Reference Spalart and McLean2011; García-Mayoral & Jiménez Reference García-Mayoral and Jiménez2012; Rastegari & Akhavan Reference Rastegari and Akhavan2018a ). It has also been suggested (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997; Spalart & McLean Reference Spalart and McLean2011) that a Reynolds number independent measure of drag reduction can be constructed by representing the normalized mean velocity profiles in the base flow and in the flow with microtextured walls as logarithmic profiles, $\langle \overline{U}\rangle /u_{\unicode[STIX]{x1D70F}_{0}}=(1/\unicode[STIX]{x1D705})\ln z^{+0}+B_{0}$ and $\langle \overline{U}\rangle /u_{\unicode[STIX]{x1D70F}}=(1/\unicode[STIX]{x1D705})\ln z^{+}+B$ , throughout the cross-section of the flow, and expressing the magnitude of drag reduction in terms of the shift, $(B-B_{0})$ , in the intercepts of these logarithmic-law representations of the mean velocity profiles, as shown in the figure 2. Here, and throughout this study, an overbar denotes Reynolds averaging in time and any homogeneous flow directions, and brackets $\langle \,\rangle$ denote averaging in the wall-parallel directions.

Using this formulation, the magnitude of drag reduction with microtextured surfaces can be expressed as (Rastegari & Akhavan Reference Rastegari and Akhavan2018a )

(3.1) $$\begin{eqnarray}\left(1-DR\right)=\left\{1+\left[\frac{1}{2\unicode[STIX]{x1D705}}\ln (1-DR)+(B-B_{0})\right]\sqrt{\frac{C_{f_{0}}}{2}}\right\}^{-2},\end{eqnarray}$$

where $DR\equiv 1-C_{f}/C_{f_{0}}$ is the magnitude of drag reduction, $C_{f}$ and $C_{f_{0}}$ are the skin-friction coefficients in the flow with microtextured walls and in the base flow with smooth no-slip walls, defined as $C_{f}\equiv 2u_{\unicode[STIX]{x1D70F}}^{2}/U_{b}^{2}$ and $C_{f_{0}}\equiv 2u_{\unicode[STIX]{x1D70F}_{0}}^{2}/U_{b_{0}}^{2}$ in internal flow and $C_{f}\equiv 2u_{\unicode[STIX]{x1D70F}}^{2}/U_{\infty }^{2}$ and $C_{f_{0}}\equiv 2u_{\unicode[STIX]{x1D70F}_{0}}^{2}/U_{\infty }^{2}$ in external flow, respectively, $\unicode[STIX]{x1D705}$ is the von Kármán constant, and $(B-B_{0})$ is Reynolds number independent and only a function of the geometry and size of the surface microtexture in wall units. The appearance of $C_{f_{0}}$ in (3.1) leads to a degradation in the magnitude of drag reduction with increasing Reynolds number of the base flow for a given $(B-B_{0})$ .

Figure 2. Schematic representation of $(B-B_{0})$ and $U_{s}^{+}$ in turbulent flow with microtextured walls: —— (black), –  $\cdot$   $\cdot$  – (turquoise), mean velocity profiles in the base turbulent flow and the flow with microtextured walls, respectively; – – – (black), –  $\cdot$  – (turquoise), logarithmic representations of the mean velocity profiles in the base turbulent flow and the flow with microtextured walls, respectively.

It was brought to our attention by one of the anonymous referees that (3.1) can be recast into an explicit expression for $DR$ , given by

(3.2) $$\begin{eqnarray}DR=1-\unicode[STIX]{x1D705}^{2}\left(\frac{2}{C_{f_{0}}}\right)\left\{W\left(\unicode[STIX]{x1D705}\sqrt{\frac{2}{C_{f_{0}}}}\exp \left(\unicode[STIX]{x1D705}\left[(B-B_{0})+\sqrt{\frac{2}{C_{f_{0}}}}\right]\right)\right)\right\}^{-2},\end{eqnarray}$$

where $W$ denotes the Lambert $W$ function.

Figure 3. The magnitude of drag reduction, $DR$ , as a function of $(B-B_{0})$ and the friction Reynolds number of the base flow, $Re_{\unicode[STIX]{x1D70F}_{0}}$ . Present DNS studies in SH turbulent channel flow: 

,
 (navy), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/2$ , $1/8$ , $1/16$ , $1/32$ , $1/64$
 (navy), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ , $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/2$ , $1/8$ , $1/16$ , $1/32$ , $1/64$
,  
 (blue), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/8$ , $1/16$ , $1/32$ , $1/64$
 (blue), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/8$ , $1/16$ , $1/32$ , $1/64$
(navy), SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$
 (blue), SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$
 (lavender), DNS of turbulent channel flow with LI longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 180$ , $\unicode[STIX]{x1D703}_{p}=0^{o}$ , $\unicode[STIX]{x1D719}_{s}=1/2$ , $N=0.1,1,2.5,10,20,100$ (Fu et al. Reference Fu, Arenas, Leonardi and Hultmark2017); ▪ (red), experiments in turbulent boundary layer flow with SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 460$ and $560$ (Zhang et al. Reference Zhang, Tian, Yao, Hao and Jiang2015); ♦ (green), experiments in turbulent Taylor–Couette flow with SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1193$ and $3810$ (Srinivasan et al. Reference Srinivasan, Kleingartner, Gilbert, Cohen, Milne and McKinley2015); ● (orange), experiments in turbulent boundary layer flow with SH longitudinal microgrooves or SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863,1408$ and $4287$ (Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016). The black lines show the predictions of (3.1) in turbulent channel flow for: ——, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ ( $Re_{b}=3600$ ); — —, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ ( $Re_{b}=7860$ ); – – –, $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{3}$ ( $Re_{b}\approx 2\times 10^{4}$ ); –  $\cdot$  –,  $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{4}$ ( $Re_{b}\approx 2.6\times 10^{5}$ ); –  $\cdot$   $\cdot$  –,  $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{5}$ ( $Re_{b}\approx 3.2\times 10^{6}$ );  $\cdots \cdots$ , $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{6}$ ( $Re_{b}\approx 3.8\times 10^{7}$ ).

Figure 3 shows the predictions of (3.1) in turbulent channel flow for $222\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 10^{6}$ compared to results from present DNS studies in turbulent channel flows with SH longitudinal microgrooves or SH aligned microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$ , as summarized in table 1. For $Re_{\unicode[STIX]{x1D70F}_{0}}\geqslant 1000$ , the values of $C_{f_{0}}$ in (3.1) were obtained from the logarithmic skin-friction correlation suggested by Zanoun, Nagib & Durst (Reference Zanoun, Nagib and Durst2009), which has been shown to give improved agreement with experimental data at high Reynolds numbers compared to the classical Dean’s correlation (Dean Reference Dean1978). In agreement with the predictions of (3.1), for a given value of $(B-B_{0})$ , the DNS data show small degradations of ${\approx}1.0{-}2.5\,\%$ in drag reduction as the friction Reynolds number of the base flow increases from $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ to $442$ . Also shown in figure 3 are the experimental data of Zhang et al. (Reference Zhang, Tian, Yao, Hao and Jiang2015), obtained in turbulent boundary layer flow with SH random microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 460$ and $560$ , the experimental data of Ling et al. (Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016), obtained in turbulent boundary layer flow with SH longitudinal microgrooves at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863$ or with SH random microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863$ , $1408$ and $4287$ , and the experimental data of Srinivasan et al. (Reference Srinivasan, Kleingartner, Gilbert, Cohen, Milne and McKinley2015), obtained in turbulent Taylor–Couette flow with SH random microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1193$ and $3810$ . In plotting these experimental data, the values of $(B-B_{0})$ were approximated by the values of $\unicode[STIX]{x0394}U^{+}$ reported in the experiments. Good agreement can be seen between the predictions of (3.1) and the experimental data, including the case where the SH surface gives a negative drag reduction of $-10\,\%$ in turbulent boundary layer flow at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 4287$ (Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016). Evidence of the degradation of drag reduction with increasing Reynolds number of the base flow, for a given value of $(B-B_{0})$ , can also be seen in these experimental data. For example, in turbulent boundary layer flow with SH random microposts, a $(B-B_{0})=1.1$ is reported to give rise to $12\,\%$ drag reduction at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863$ , but only $11\,\%$ drag reduction at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1408$ (Ling et al. Reference Ling, Srinivasan, Golovin, McKinley, Tuteja and Katz2016). Similarly, a $(B-B_{0})=3.2$ is found to give rise to $24.7\,\%$ drag reduction in turbulent boundary layer flow with SH random microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 560$ (Zhang et al. Reference Zhang, Tian, Yao, Hao and Jiang2015), but only $22\,\%$ drag reduction in turbulent Taylor–Couette flow with SH random microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 3810$ (Srinivasan et al. Reference Srinivasan, Kleingartner, Gilbert, Cohen, Milne and McKinley2015). While these degradations in drag reduction are small at moderate Reynolds number, they become far more significant at higher Reynolds numbers, requiring ever higher values of $(B-B_{0})$ to achieve a given level of drag reduction, as shown in figure 3. Nevertheless, with a judicious choice of $(B-B_{0})$ , drag reductions of up to ${\sim}50\,\%$ should still be feasible at $Re_{\unicode[STIX]{x1D70F}_{0}}\sim 10^{5}{-}10^{6}$ of practical interest. Finally, it should be noted that the drag reduction relation given by (3.1) is equally valid for all microtexture surfaces, including LI surfaces. This can be seen in figure 3 by a comparison of the predictions of (3.1) with the DNS results of Fu et al. (Reference Fu, Arenas, Leonardi and Hultmark2017), obtained in turbulent channel flows with LI longitudinal microgrooves at viscosity ratios of $0.1\leqslant N\equiv \unicode[STIX]{x1D707}_{o}/\unicode[STIX]{x1D707}_{i}\leqslant 100$ with $\unicode[STIX]{x1D719}_{s}=1/2$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 180$ .

Designing optimal SH surfaces for drag reduction in high Reynolds number turbulent flows requires, in addition to (3.1), an understanding of the scaling of $(B-B_{0})$ with the geometrical parameters of the surface microtexture. Progress can be made by noting that the shift, $(B-B_{0})$ , arises primarily because of the presence of an effective streamwise slip velocity, $U_{s}^{+}$ , at the wall, as shown in figure 2 (Min & Kim Reference Min and Kim2004; Seo & Mani Reference Seo and Mani2016; Rastegari & Akhavan Reference Rastegari and Akhavan2018a ). However, the magnitude of $(B-B_{0})$ is always smaller than $U_{s}^{+}$ due to concurrent presence of spanwise slip. As such, $(B-B_{0})$ and $U_{s}^{+}$ can be expected to have similar scalings with surface microtexture geometrical parameters, but with a smaller multiplicative coefficient in the scaling law for $(B-B_{0})$ compared to $U_{s}^{+}$ .

Figure 4. Contours of instantaneous streamwise velocity, $U^{+}(x,y,z)$ , at $z=0$ for: (a) SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ ; (b) SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ ; (c) SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 56$ .

Figure 5. Production and dissipation of turbulence kinetic energy in turbulence channel flow with SH longitudinal microgrooves at (a,c,d,e) $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ and (b,f,g,h) $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ : (a,b) ——, — —, $\cdots \cdots$ , –  $\cdot$   $\cdot$  – (turquoise), total production $\langle {\mathcal{P}}\rangle ^{+}=-\langle \overline{u_{i}u_{j}}\overline{S}_{ij}\rangle ^{+}$ , production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$ , production arising from $-\langle \overline{uv}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\rangle ^{+}$ , dissipation $\langle \unicode[STIX]{x1D700}\rangle ^{+}=\langle 2\unicode[STIX]{x1D708}\overline{s_{ij}s_{ij}}\rangle ^{+}$ , in turbulent channel flow with SH walls; – – –, –  $\cdot$  – (black), production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$ and dissipation, $\langle \unicode[STIX]{x1D700}\rangle ^{+}$ , in base turbulent channel flow with smooth, no-slip walls; (c,d,e; f,g,h) spanwise variation of the mean streamwise velocity, $\bar{U}^{+}$ , total production, ${\mathcal{P}}^{+}$ , and dissipation, $\unicode[STIX]{x1D700}^{+}$ , at different $z^{+}$ in turbulent channel flow with SH longitudinal microgrooves at $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , respectively. The blue bars on the lower horizontal axes of (c–h) denote the locations of the no-slip surfaces. $L=g+w$ denotes the pitch of the microgrooves.

Figure 6. Production and dissipation of turbulence kinetic energy in turbulent channel flow with SH aligned microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 56$ : (a) total production $\langle {\mathcal{P}}\rangle ^{+}=-\langle \overline{u_{i}u_{j}}\overline{S}_{ij}\rangle ^{+}$ , production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$ , production arising from $-\langle \overline{uv}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\rangle ^{+}$ , dissipation $\langle \unicode[STIX]{x1D700}\rangle ^{+}=\langle 2\unicode[STIX]{x1D708}\overline{s_{ij}s_{ij}}\rangle ^{+}$ , in turbulent channel flow with SH walls, compared to base turbulent channel flow with smooth, no-slip walls; (b) schematic representation of sections A-A and B-B where turbulence statistics are shown in (c–e) and (f–h); (c–e; f–h) spanwise variation of the mean streamwise velocity, $\bar{U}^{+}$ , total production, ${\mathcal{P}}^{+}$ , dissipation, $\unicode[STIX]{x1D700}^{+}$ , at different $z^{+}$ of sections A-A and B-B, respectively. Line types in (a) as in figure 5. The blue bars on the lower horizontal axes of (c–h) denote the locations of the no-slip surfaces. $L=g+w$ denotes the pitch of the microposts.

The scaling of $U_{s}^{+}$ with surface microtexture parameters can be clarified by considering the dynamics of turbulence kinetic energy (TKE) within the ‘surface layer’. This is a layer of thickness ${\sim}g$ above the microtextured walls, in which the flow transitions from the slip/no-slip patterns at the tip of the wall microtexture, at $z=0$ , to a homogeneous turbulent flow in the wall-parallel directions at $z\sim g$ (Rastegari & Akhavan Reference Rastegari and Akhavan2015). For surface microtextures where the flow can find an unobstructed passageway through the surface micro-features, such as with longitudinal microgrooves or aligned microposts, sharp spanwise gradients of the mean streamwise velocity develop between the slip and no-slip regions of the boundary, giving rise to strong shear layers within the surface layer above these regions, as shown in figures 4(a–c), 5(c,f) and 6(c,f). These shear layers are the strongest with longitudinal microgrooves, but are also present with aligned microposts, where entire stripes aligned with the row of microposts can act as low-slip regions, as shown in figures 4(c) and 6(c,f). The presence of interface curvature somewhat mitigates the strength of these shear layers, but their mechanism is still at work, as can be seen by a comparison of figures 4(a) and 4(b) or figures 5(c) and 5(f).

These shear layers are the source of additional production of TKE within the surface layer through the $-\overline{uv}~\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y$ term, as shown in figures 5(a,b) and 6(a). This additional production of TKE, which is above and beyond the normal production of TKE in wall-bounded flows through the $-\overline{uw}~\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z$ term, dominates the overall production of TKE within the surface layer, as shown in figures 5(a,b,d,g) and 6(a,d,g). The additional production of TKE through these shear layers is dissipated by the turbulent eddies above the no-slip regions, as shown in figures 5(e,h) and 6(e,h), such that

(3.3) $$\begin{eqnarray}\{-\overline{uv}~\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\}_{shear\,layers}\sim \{\unicode[STIX]{x1D708}\overline{s_{ij}s_{ij}}\}_{no\text{-}slip}.\end{eqnarray}$$

An order of magnitude estimate of the terms in (3.3) leads to the scaling law for $U_{s}^{+}$ . Specifically, within the surface layer, the spanwise gradients of the mean streamwise velocity can be estimated as $\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\sim U_{ss}/g$ , where $U_{ss}=U_{s}/(1-\unicode[STIX]{x1D719}_{s})$ is the average slip velocity on the slip portions of the microtextured walls. Furthermore, the fluctuating strain rate above the no-slip regions can be estimated as $\{{s_{ij}\}}_{no-slip}\sim \sqrt{u_{\unicode[STIX]{x1D70F}_{n}}u_{\unicode[STIX]{x1D70F}}}/(\unicode[STIX]{x1D708}/\sqrt{u_{\unicode[STIX]{x1D70F}_{n}}u_{\unicode[STIX]{x1D70F}}})$ , where $\sqrt{u_{\unicode[STIX]{x1D70F}_{n}}u_{\unicode[STIX]{x1D70F}}}$ is the characteristic velocity of the turbulent eddies above the no-slip regions within the surface layer, estimated as the geometric average of the wall friction velocities $u_{\unicode[STIX]{x1D70F}_{n}}=u_{\unicode[STIX]{x1D70F}}/\sqrt{\unicode[STIX]{x1D719}_{s}}$ and $u_{\unicode[STIX]{x1D70F}}$ above the no-slip regions at $z=0$ and $z\sim g$ , respectively, and $\unicode[STIX]{x1D708}/\sqrt{u_{\unicode[STIX]{x1D70F}_{n}}u_{\unicode[STIX]{x1D70F}}}$ is the associated inner length scale. Finally, the Reynolds shear stress $-\overline{uv}$ can be estimated using a mixing-length model as $-\overline{uv}\sim \unicode[STIX]{x1D708}_{t}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\sim u^{\prime }\ell ~\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y$ , where $u^{\prime }\sim \sqrt{u_{\unicode[STIX]{x1D70F}_{n}}u_{\unicode[STIX]{x1D70F}}}$ is the characteristic velocity of the largest eddies within the surface layer, and $\ell \sim \sqrt{g\unicode[STIX]{x1D708}/u_{\unicode[STIX]{x1D70F}}}$ is their characteristic size, approximated as the geometric average of $g$ and $\unicode[STIX]{x1D708}/u_{\unicode[STIX]{x1D70F}}$ , representing the size of the largest eddies at $z\sim g$ and $z\sim 0$ , respectively. Substitution of these expressions into (3.3) gives the scaling law for $U_{s}^{+}$ as

(3.4) $$\begin{eqnarray}\frac{\unicode[STIX]{x1D719}_{s}^{3/8}}{(1-\unicode[STIX]{x1D719}_{s})}\,U_{s}^{+}\sim \{{g^{+}\}}^{3/4}.\end{eqnarray}$$

Given that the dominant contribution to $(B-B_{0})$ arises from $U_{s}^{+}$ , a similar scaling can also be assumed for $(B-B_{0})$ :

(3.5) $$\begin{eqnarray}\frac{\unicode[STIX]{x1D719}_{s}^{3/8}}{(1-\unicode[STIX]{x1D719}_{s})}\,(B-B_{0})\sim \{{g^{+}\}}^{3/4},\end{eqnarray}$$

with the understanding that the multiplicative coefficient in the expression for $(B-B_{0})$ would be smaller than that for $U_{s}^{+}$ .

Figure 7. The scaling of $U_{s}^{+}$ and $(B-B_{0})$ with surface microtexture parameters in turbulent flow with SH or LI walls: (a–d) present DNS data in turbulent channel flow with SH longitudinal microgrooves or SH aligned microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$ , $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $-30^{\circ }$ ; (e,f) DNS data from other investigators in turbulent channel flow with SH longitudinal microgrooves, LI longitudinal microgrooves or SH aligned microposts at $180\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 590$ , $1/36\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ . Symbols as in figure 3. Other symbols in (e,f):

 (brown), $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 180$ , $\unicode[STIX]{x1D719}_{s}=1/8$ , $1/4$ , $1/2$
 (green),  $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 395$ , $\unicode[STIX]{x1D719}_{s}=1/16$ , $1/8$ , $1/4$ , $1/2$
 (red),  $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 590$ , $\unicode[STIX]{x1D719}_{s}=1/16$ , $1/8$ , $1/4$ , $1/2$ , SH longitudinal microgrooves, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ (Park et al. Reference Park, Park and Kim2013); 
 (brown),  $Re_{\unicode[STIX]{x1D70F}}\approx 195$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $1/9$
 (green),  $Re_{\unicode[STIX]{x1D70F}}\approx 400$ , $\unicode[STIX]{x1D719}_{s}=1/36$ , $1/16$ , $1/9$ , SH aligned microposts, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ (Seo & Mani Reference Seo and Mani2016).

Figure 7(a–d) shows $U_{s}^{+}$ and $(B-B_{0})$ from the DNS studies summarized in table 1, compared to (3.4) and (3.5). At both $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$ , and for both $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , $U_{s}^{+}$ and $(B-B_{0})$ scale as $\{{g^{+}\}}^{3/4}$ , while normalization with $\unicode[STIX]{x1D719}_{s}^{3/8}/(1-\unicode[STIX]{x1D719}_{s})$ collapses the data from different $\unicode[STIX]{x1D719}_{s}$ . Both $U_{s}^{+}$ and $(B-B_{0})$ are Reynolds number independent and only functions of the geometrical parameters of the surface microtexture in wall units, in agreement with the predictions of (3.4) and (3.5).

A best fit to the DNS data gives the multiplicative coefficients in the expressions for $U_{s}^{+}$ and $(B-B_{0})$ as

(3.6) $$\begin{eqnarray}\displaystyle U_{s}^{+} & = & \displaystyle 0.52\{(1-\unicode[STIX]{x1D719}_{s})\unicode[STIX]{x1D719}_{s}^{-3/8}\}\{{g^{+}\}}^{3/4},\end{eqnarray}$$
(3.7) $$\begin{eqnarray}\displaystyle (B-B_{0}) & = & \displaystyle 0.41\{(1-\unicode[STIX]{x1D719}_{s})\unicode[STIX]{x1D719}_{s}^{-3/8}\}\{{g^{+}\}}^{3/4},\end{eqnarray}$$

with $p$ -values of 0.98, obtained from the Kolmogorov–Smirnov goodness of fit test.

Figure 7(e,f) shows the comparison of (3.6) and (3.7) with available DNS data from other investigators, obtained in turbulent channel flows using SH longitudinal microgrooves (Park et al. Reference Park, Park and Kim2013), SH aligned microposts (Seo & Mani Reference Seo and Mani2016) or LI longitudinal microgrooves at viscosity ratios of $N\equiv \unicode[STIX]{x1D707}_{o}/\unicode[STIX]{x1D707}_{i}=0.1,1,10,20,100$ (Fu et al. Reference Fu, Arenas, Leonardi and Hultmark2017). Good agreement can be seen between the predictions of (3.6) and (3.7) and available DNS data from other investigators for both SH longitudinal microgrooves and SH aligned microposts as long as $g^{+}<100$ . For $g^{+}>100$ , both $U_{s}^{+}$ and $(B-B_{0})$ saturate and branch off from the predictions of (3.6) and (3.7). This saturation has been attributed to the width of the microgrooves, $g^{+}$ , approaching the mean streak spacing of $\unicode[STIX]{x1D706}^{+}\sim 100$ in turbulent wall flows (Park et al. Reference Park, Park and Kim2013). Presumably, for $g^{+}\gtrsim 100$ , the dynamics of turbulence above the slip regions becomes independent of the dynamics of turbulence above the no-slip regions, and both $U_{s}^{+}$ and $(B-B_{0})$ approach their limiting values.

Good agreement can also be seen in figure 7(e,f) between the predictions of (3.6) and (3.7) and DNS results in turbulent channel flows with LI longitudinal microgrooves (Fu et al. Reference Fu, Arenas, Leonardi and Hultmark2017) when the viscosity ratio, $N$ , is greater than ${\sim}10$ . These results suggest that for high enough viscosity ratios, of $N\gtrsim 10$ , the drag reduction of LI surfaces begins to approach that of SH surfaces. These findings are consistent with analytical solutions in laminar shear flow with LI surfaces obtained by Schönecker, Baier & Hardt (Reference Schönecker, Baier and Hardt2014), and reproduced in figure 8, which show that for LI longitudinal microgrooves of depth $d=g$ , $\unicode[STIX]{x1D719}_{s}=0.5$ and viscosity ratio of $N=10$ , the effective slip length is ${\approx}86\,\%$ of the effective slip length predicted by Philip’s solution (Philip Reference Philip1972) for the same surface microtexture with ‘idealized’, shear-free, SH interfaces. The LI results shown in figure 7(e,f) are also consistent with experimental measurements in turbulent Taylor–Couette flow with LI longitudinal microgrooves of depth $d=g$ , $\unicode[STIX]{x1D719}_{s}=0.5$ and $0.7\leqslant N\leqslant 3.0$ (Van Buren & Smits Reference Van Buren and Smits2017), which show the drag reduction of the LI surface at $N\approx 2.0{-}3.0$ to be ${\approx}50{-}60\,\%$ of the drag reduction of a SH surface with the same surface microtexture.

In the derivation of (3.6) and (3.7), the multiplicative coefficients, of 0.52 and 0.41, were obtained based on DNS data which assumed ‘idealized’, shear-free boundaries at the air/water interfaces and, as such, did not consider the flow of air in the SH microgrooves. This raises a concern that the magnitudes of $U_{s}^{+}$ and $(B-B_{0})$ would be over-predicted by (3.6) and (3.7). An estimate of the degree of this over-prediction can be obtained by comparing analytical solutions for the effective slip length, $L_{s,N}$ , obtained with a finite viscosity ratio, $N$ , in laminar shear flow with LI longitudinal microgrooves (Schönecker et al. Reference Schönecker, Baier and Hardt2014) to the effective slip length, $L_{s,\infty }$ , which would be predicted in the same flow based on the assumption of shear-free liquid/air interfaces (Philip Reference Philip1972). Figure 8 shows $L_{s,N}/L_{s,\infty }$ as a function of $N$ for longitudinal microgrooves of depth $d=g$ and the range of solid fractions, $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ , investigated in the present study. It can be seen that for $N\approx 56$ , corresponding to air/water interfaces, the ratio $L_{s,N}/L_{s,\infty }$ ranges from $0.92\leqslant L_{s,N}/L_{s,\infty }\leqslant 0.97$ for $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ . These results suggest that for the range of $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ , considered in the present study, the errors in (3.6) and (3.7) due to the assumption of shear-free interfaces are no more than 10 %. These errors can become larger for solid fractions of $\unicode[STIX]{x1D719}_{s}\lesssim 1/64$ or for shallow grooves of $d\lesssim g/2$ (Schönecker et al. Reference Schönecker, Baier and Hardt2014). However, such configurations of microgrooves are impractical to build, either because the microgroove blades become too thin to be structurally sustainable, or because the depth of the grooves becomes so shallow that the air/water interfaces can hit the bottom of the grooves during their deformation, thus expediting their collapse.

Figure 8. Ratio of the effective slip length, $L_{s,N}$ , obtained with a finite viscosity ratio, $N$ , based on the analytical solution of Schönecker et al. (Reference Schönecker, Baier and Hardt2014), to the effective slip length, $L_{s,\infty }$ , obtained with the assumption of shear-free liquid/air interfaces (Philip Reference Philip1972), in laminar shear flow over LI longitudinal microgrooves for groove depth $d=g$ and $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$ : $\cdots \cdots$ , $\unicode[STIX]{x1D719}_{s}=1/2$ ; — ⋅ ⋅ —, $\unicode[STIX]{x1D719}_{s}=1/8$ ; — ⋅ —, $\unicode[STIX]{x1D719}_{s}=1/16$ ; – – –, $\unicode[STIX]{x1D719}_{s}=1/32$ ; ——, $\unicode[STIX]{x1D719}_{s}=1/64$ .

Another concern is the effect of dynamic deformation of the interface on (3.6) and (3.7), as the multiplicative coefficients in (3.6) and (3.7) were obtained from DNS studies which were performed with ‘stationary’, flat or curved, shear-free interfaces. The effect of dynamic deformation of the interface can be assessed by examining available DNS results obtained with deformable interfaces, but pinned contact lines, in turbulent channel flows with LI longitudinal microgrooves at a viscosity ratio of $N=2.5$ for $We_{\unicode[STIX]{x1D70F}}=0$ and $We_{\unicode[STIX]{x1D70F}}=0.024$ (Arenas-Navarro Reference Arenas-Navarro2017), where $We_{\unicode[STIX]{x1D70F}}\equiv \unicode[STIX]{x1D70C}_{o}u_{\unicode[STIX]{x1D70F}}\unicode[STIX]{x1D708}_{o}/\unicode[STIX]{x1D70E}$ is the Weber number, $\unicode[STIX]{x1D70C}_{o}$ and $\unicode[STIX]{x1D708}_{o}$ are the density and kinematic viscosity of the outer fluid, respectively, and $\unicode[STIX]{x1D70E}$ is the surface tension.

Figure 9 shows $\{\unicode[STIX]{x1D719}_{s}^{3/8}/(1-\unicode[STIX]{x1D719}_{s})\}\,U_{s}^{+}$ and $\{\unicode[STIX]{x1D719}_{s}^{3/8}/(1-\unicode[STIX]{x1D719}_{s})\}\,(B-B_{0})$ computed from these DNS databases. Both $\{\unicode[STIX]{x1D719}_{s}^{3/8}/(1-\unicode[STIX]{x1D719}_{s})\}\,U_{s}^{+}$ and $\{\unicode[STIX]{x1D719}_{s}^{3/8}/(1-\unicode[STIX]{x1D719}_{s})\}\,(B-B_{0})$ can be seen to scale as ${\sim}{g^{+}\{}^{3/4}\}$ , in agreement with (3.6) and (3.7). However, the multiplicative coefficients in the best fit lines to the LI data are smaller than those in (3.6) and (3.7) by a factor of $0.6$ for $U_{s}^{+}$ and a factor of $0.4$ for $(B-B_{0})$ . These factors are comparable to the ratio $L_{s,N}/L_{s,\infty }=0.6$ observed for $N=2.5$ and $\unicode[STIX]{x1D719}_{s}=1/2$ in figure 8.

The dynamic deformation of the interface is seen to have little effect on the slip velocity, with the values for $U_{s}^{+}$ at the two Weber numbers being nearly identical, as seen in figure 8(a). However, the surface ripples formed by the dynamic deformation of the interface can act as ‘surface roughness’ and reduce the magnitude of $(B-B_{0})$ by $2.5-25.5\,\%$ at $We_{\unicode[STIX]{x1D70F}}=0.024$ compared to $We_{\unicode[STIX]{x1D70F}}=0$ , as seen in figure 8(b). The results shown in figure 9 were obtained with LI longitudinal microgrooves at $N=2.5$ and pinned contact lines. As will be shown in § 4 (figure 12), for SH longitudinal microgrooves, a $We_{\unicode[STIX]{x1D70F}}=0.024$ is very large and will lead to interface instability for $g^{+}>6.5$ at $Re_{\unicode[STIX]{x1D70F}}\approx 200$ . Stable SH longitudinal microgrooves require much smaller values of $We_{\unicode[STIX]{x1D70F}}$ , for which the effect of dynamic deformation of the interface on $(B-B_{0})$ is much smaller than the results shown in figure 8(b). Nevertheless, these effects can be accounted for by constructing a $\unicode[STIX]{x0394}B_{k}$ associated with the ‘roughness’ of the liquid/gas interface, and subtracting this $\unicode[STIX]{x0394}B_{k}$ from the $(B-B_{0})$ given by (3.7) to get the drag reduction. More recent DNS studies (Rastegari & Akhavan Reference Rastegari and Akhavan2018b ) suggest that the motion of the contact line may be a more important source of roughness effects. A similar procedure can also be used to account for the roughness effects arising from the motion of the contact line on SH or LI longitudinal microgrooves or aligned microposts, assuming the interface has not fully collapsed.

Figure 9. Effect of dynamic deformation of the interface on (a) $U_{s}^{+}$ and (b) $(B-B_{0})$ , based on DNS results with deformable interfaces, but pinned contact lines, in turbulent channel flows with LI longitudinal microgrooves at $Re_{\unicode[STIX]{x1D70F}}\approx 180$ , $\unicode[STIX]{x1D719}_{s}=1/2$ , $N=2.5$ (Arenas-Navarro Reference Arenas-Navarro2017): 

 (lavender),  $We_{\unicode[STIX]{x1D70F}}=0$
 (dark violet), $We_{\unicode[STIX]{x1D70F}}=0.024$ .

Finally, it should be noted that, given that the derivations of (3.6) and (3.7) require an unobstructed passageway through the surface microtexture, these scalings are not expected to apply to staggered microposts or transverse microgrooves.

4 Sustainability bounds of superhydrophobic surfaces

Equations (3.1) and (3.7) provide a tool for a priori prediction of the magnitude of drag reduction with any pattern of SH longitudinal microgrooves or aligned microposts in turbulent flow at any Reynolds number. However, design of optimal SH surfaces also requires consideration of the sustainability bounds of such surfaces, as all SH surfaces are susceptible to a wetting transition from the Cassie–Baxter state to a Wenzel state. This wetting transition can occur because of (i) depinning of the contact line and/or sagging of the liquid–gas interface under high local pressures (Zheng et al. Reference Zheng, Yu and Zhao2005; Checco et al. Reference Checco, Ocko, Rahman, Black, Tasinkevych, Giacomello and Dietrich2014) or (ii) diffusion or entrainment of the gas layer into the working liquid under high local shear rates (Samaha et al. Reference Samaha, Tafreshi and Gad-el Hak2012; Karatay et al. Reference Karatay, Tsai and Lammertink2013; Wexler et al. Reference Wexler, Jacobi and Stone2015; Ling et al. Reference Ling, Katz, Fu and Hultmark2017) of turbulent flow. Of the two mechanisms, the wetting transition due to high local pressures is generally the first mode of failure, and of the different SH surface microtextures, longitudinal microgrooves are the most stable under this mode of failure, as impingement of the flow on the back of microposts creates large stagnation pressures which can expedite the wetting transition (Seo, García-Mayoral & Mani Reference Seo, García-Mayoral and Mani2015). Consequently, in this section we will focus only on the pressure stability bounds of SH longitudinal microgrooves.

For SH longitudinal microgrooves with stable interfaces, the contact line is pinned and the shape of the interface is governed by the Young–Laplace equation

(4.1) $$\begin{eqnarray}\unicode[STIX]{x0394}P=-\frac{\unicode[STIX]{x1D70E}}{R}=-\frac{2\unicode[STIX]{x1D70E}\cos (\unicode[STIX]{x1D703}_{c})}{g},\end{eqnarray}$$

where $\unicode[STIX]{x0394}P$ is the pressure difference across the interface, $\unicode[STIX]{x1D70E}$ is the surface tension, $R$ is the radius of curvature of the interface, $\unicode[STIX]{x1D703}_{c}$ is the contact angle and $g$ is the microgroove width. With increasing $\unicode[STIX]{x0394}P$ , the contact angle increases until it reaches the maximum advancing contact angle, $\unicode[STIX]{x1D703}_{F,adv}$ , at which point the contact line depins (Checco et al. Reference Checco, Ocko, Rahman, Black, Tasinkevych, Giacomello and Dietrich2014). Assuming a maximum advancing contact angle of $\unicode[STIX]{x1D703}_{F,adv}=120^{\circ }$ (Nishino et al. Reference Nishino, Meguro, Nakamae, Matsushita and Ueda1999), the criterion for stability of SH interfaces becomes $\unicode[STIX]{x0394}Pg/\unicode[STIX]{x1D70E}\leqslant 1$ , or

(4.2) $$\begin{eqnarray}(\unicode[STIX]{x0394}P^{+})(g^{+})(We_{\unicode[STIX]{x1D70F}})\leqslant 1,\end{eqnarray}$$

where $\unicode[STIX]{x0394}P^{+}\equiv \unicode[STIX]{x0394}P/(\unicode[STIX]{x1D70C}u_{\unicode[STIX]{x1D70F}}^{2})$ is the non-dimensional pressure, and $We_{\unicode[STIX]{x1D70F}}\equiv \unicode[STIX]{x1D70C}u_{\unicode[STIX]{x1D70F}}\unicode[STIX]{x1D708}/\unicode[STIX]{x1D70E}$ is the Weber number.

Figure 10. Probability density function and cumulative distribution function of the wall pressure fluctuations in turbulent channel flow with SH longitudinal microgrooves, compared to turbulent channel flow with smooth, no-slip walls: (a) probability density function, $f_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2})$ ; (b) cumulative distribution function, $F_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\}^{1/2})$ ; – – –, –  $\cdot$  –, SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ , $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $-30^{\circ }$ , respectively; $\cdots \cdots$ , base turbulent channel flow with smooth, no-slip walls, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ ; —— (orange), Gaussian distribution.

Assessing the stability bounds of SH surfaces in turbulent flow from (4.2) requires a knowledge of $\unicode[STIX]{x0394}P^{+}$ . In turbulent flow, $\unicode[STIX]{x0394}P^{+}$ arises from the instantaneous pressure fluctuations on the SH walls. The range of values assumed by the instantaneous pressure fluctuations on the SH walls can be assessed by examining the probability density function (p.d.f.) of the wall pressure fluctuations, $f_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2})$ . Figure 10(a) shows the p.d.f. of wall pressure fluctuations in turbulent channel flow with SH longitudinal microgrooves at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ , $\unicode[STIX]{x1D719}_{s}=1/64$ , $g^{+0}\approx 63$ and $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ , compared to turbulent channel flow with smooth, no-slip walls. The p.d.f. of instantaneous wall pressure fluctuations in turbulent channel flow with SH walls is found to be similar to that for smooth, no-slip walls. Both p.d.f.s follow a non-Gaussian distribution with exponential tails. Similar non-Gaussian distributions have also been reported experimentally in high Reynolds number turbulent boundary layers (Tsuji et al. Reference Tsuji, Fransson, Alfredsson and Johansson2007). The probability that the instantaneous pressure fluctuations fall within a given range of values, $a\leqslant p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2}<b$ , can be estimated from $F_{p}(b)-F_{p}(a)=\int _{a}^{b}f_{p}(\unicode[STIX]{x1D702})\,\text{d}\unicode[STIX]{x1D702}$ , where $F_{p}$ is the cumulative distribution function. Figure 10(b) shows the cumulative distribution functions, $F_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2})$ , for the cases shown in figure 10(a). Examination of these cumulative distribution functions shows that the probability of the instantaneous wall pressure fluctuations falling within $\pm 4\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2}$ is $99.75\,\%$ . These results are consistent with experimental measurements in turbulent boundary layer flows (Schewe Reference Schewe1983), which show a probability of $99.7\,\%$ for the wall pressure fluctuations never exceeding $4\{\overline{p^{2}}_{w}^{+}\!\}^{1/2}$ . Based on these results, an upper estimate for the magnitude of $\unicode[STIX]{x0394}P^{+}$ can be obtained as $\unicode[STIX]{x0394}P^{+}\approx 4\{\langle \overline{p^{2}}\rangle _{w,SH}^{+}\!\}^{1/2}$ , with $99.75\,\%$ confidence that the instantaneous wall pressure fluctuations will not exceed this value.

Figure 11. Mean-squared turbulent wall pressure fluctuations in turbulent channel flow with SH longitudinal microgrooves: (a) $\langle \bar{p^{2}}\rangle _{w,SH}^{+}$ as a function of $Re_{\unicode[STIX]{x1D70F}}$ ; (b) $\langle \bar{p^{2}}\rangle _{w,SH}^{+}-\unicode[STIX]{x1D6E4}(g^{+})$ as a function of $Re_{\unicode[STIX]{x1D70F}}$ , where $\unicode[STIX]{x1D6E4}(g^{+})=2.32\ln (\{g^{+}\}^{3/4})-4.5$ . Symbols as in figure 3. $\cdots \cdots$ , the fit $\langle \bar{p^{2}}\rangle _{w,smooth}^{+}=2.31\ln (Re_{\unicode[STIX]{x1D70F}})-9.5$ , suggested by Sillero, Jiménez & Moser (Reference Sillero, Jiménez and Moser2013) for turbulent channel flows with smooth, no-slip walls.

Figure 11(a) shows $\langle \overline{p^{2}}\rangle _{w,SH}^{+}$ as a function of $Re_{\unicode[STIX]{x1D70F}}$ from present DNS studies of turbulent channel flow with SH longitudinal microgrooves at fixed protrusion angle of $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$ ( $\unicode[STIX]{x1D703}_{c}=120^{\circ }$ ). For reference, the fit $\langle \overline{p^{2}}\rangle _{w,smooth}^{+}=2.31\ln (Re_{\unicode[STIX]{x1D70F}})-9.5$ , suggested by Sillero et al. (Reference Sillero, Jiménez and Moser2013), for turbulent channel flow with smooth no-slip walls is also shown. Similar to turbulent flow with smooth no-slip walls, the mean-squared wall pressure fluctuations in turbulent flow with SH walls is strongly Reynolds number dependent. However, due to the presence of spanwise slip, the magnitude of $\langle \bar{p^{2}}\rangle _{w,SH}^{+}$ on SH walls is higher than that on smooth no-slip walls at comparable $Re_{\unicode[STIX]{x1D70F}}$ . Analysis of DNS data suggests that the effect of this spanwise slip can be captured in a function $\unicode[STIX]{x1D6E4}(g^{+})$ , such that

(4.3) $$\begin{eqnarray}\langle \overline{p^{2}}\rangle _{w,SH}^{+}\approx \unicode[STIX]{x1D6E4}(g^{+})+\langle \overline{p^{2}}\rangle _{w,smooth}^{+}\approx \unicode[STIX]{x1D6E4}(g^{+})+2.31\ln (Re_{\unicode[STIX]{x1D70F}})-9.5.\end{eqnarray}$$

A best fit to the DNS data suggests the correlation

(4.4) $$\begin{eqnarray}\unicode[STIX]{x1D6E4}(g^{+})\approx \left\{\begin{array}{@{}ll@{}}2.32\ln (\{g^{+}\}^{3/4})-4.5,\quad & g^{+}\gtrsim 5\\ 0,\quad & g^{+}\lesssim 5.\end{array}\right.\end{eqnarray}$$

For $g^{+}\lesssim 5$ , the microgrooves become too narrow to generate any significant spanwise slip (Rastegari & Akhavan Reference Rastegari and Akhavan2018a ) and $\unicode[STIX]{x1D6E4}(g^{+})\approx 0$ . The correlations (4.3)–(4.4) fit the DNS data with a $p$ -value of $0.97$ , based on the Kolmogorov–Smirnov goodness of fit test. Figure 11(b) shows the comparison of these correlations to the DNS data.

Figure 12. The maximum permissible $We^{+}$ and $g^{+}$ at a given $Re_{\unicode[STIX]{x1D70F}}$ to maintain the stability of SH longitudinal microgrooves in turbulent channel flow. Lines show the predictions of (4.2) for: ——, $Re_{\unicode[STIX]{x1D70F}}\approx 200$ ; — —, $Re_{\unicode[STIX]{x1D70F}}\approx 400$ ; – – –, $Re_{\unicode[STIX]{x1D70F}}\approx 10^{3}$ ; — ⋅ —, $Re_{\unicode[STIX]{x1D70F}}\approx 10^{4}$ ; –  $\cdot$   $\cdot$  –,  $Re_{\unicode[STIX]{x1D70F}}\approx 10^{5}$ ; $\cdots \cdots$ $Re_{\unicode[STIX]{x1D70F}}\approx 10^{6}$ .

Figure 12 shows the pressure stability bounds of SH longitudinal microgrooves obtained from (4.2), with $\unicode[STIX]{x0394}P^{+}$ approximated as $\unicode[STIX]{x0394}P^{+}\approx 4\{\langle \overline{p^{2}}\rangle _{w,SH}^{+}\!\}^{1/2}$ and $\langle \overline{p^{2}}\rangle _{w,SH}^{+}$ computed from (4.3) to (4.4). Due to the strong Reynolds number dependence of $\langle \overline{p^{2}}\rangle _{w,SH}^{+}$ , the pressure stability bounds of SH surfaces become significantly curtailed with increasing Reynolds number of the flow. Specifically, for the range of $10^{4}\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 10^{5}$ and $We_{\unicode[STIX]{x1D70F}_{0}}\approx 5\times 10^{-3}$ , typical of practical applications, figure 3 shows that achieving ${\sim}50\,\%$ drag reduction requires a $(B-B_{0})\approx 12{-}14$ . At the same time, figure 12 shows that the microgroove widths should be kept to $g^{+}\lesssim 17{-}20$ to maintain the stability of the interface. The only way such large values of $(B-B_{0})$ can be achieved with such small values of $g^{+}$ is by using ‘blade’ or ‘scalloped’ longitudinal microgrooves with $\unicode[STIX]{x1D719}_{s}\lesssim 0.02{-}0.05$ , per (3.7), where the nomenclature for ‘blade’ or ‘scalloped’ longitudinal microgrooves is borrowed from the riblet literature (Bechert et al. Reference Bechert, Bruse, Hage, Van Der Hoeven and Hoppe1997).

The combination of (3.1), (3.7), (4.2) and (4.3)–(4.4) provides a set of design tools for a priori prediction of the magnitude of drag reduction and the sustainability bounds of SH surfaces in high Reynolds number turbulent flows. Given the Reynolds number, $Re_{\unicode[STIX]{x1D70F}_{0}}$ , and Weber number, $We_{\unicode[STIX]{x1D70F}_{0}}$ , of the base flow and a desired level of drag reduction, (3.1) provides an estimate of the required $(B-B_{0})$ , while (4.2) and (4.3)–(4.4) provide an estimate of the maximum groove width $g^{+}$ for stability. These $(B-B_{0})$ and $g^{+}$ values can then be used in (3.7) to determine the maximum allowable solid fraction, $\unicode[STIX]{x1D719}_{s}$ .

5 Summary and conclusions

Using scaling arguments and analysis of results from DNS, we present scaling laws which allow for a priori prediction of the magnitude of skin-friction drag reduction with SH or LI surfaces and the pressure stability bounds of SH surfaces in turbulent flow. It is shown that the magnitude of drag reduction can be parameterized in terms of the friction coefficient of the base flow and the shift, $(B-B_{0})$ , in the intercept of logarithmic-law representations of the mean velocity profiles in the flow with SH or LI walls compared to the base flow, where $(B-B_{0})$ is Reynolds number independent. The parameterization of $(B-B_{0})$ in terms of the geometrical parameters of the SH or LI surface, and the pressure stability bounds of SH longitudinal microgrooves in turbulent flow are presented. It is shown that, for a given geometry and size of the surface microtexture in wall units, both the magnitude of drag reduction and the sustainability bounds of SH surfaces degrade with increasing Reynolds number of the base flow. Using these scalings, the narrow range of surface microtexture geometrical parameters which can yield large drag reduction while maintaining the stability of SH interfaces in high Reynolds number turbulent flows have been identified.

Acknowledgement

This work was supported by an unrestricted grant from the M. R. Prince Foundation, and NSF XSEDE Allocation TG-CTS070067N. This support is gratefully acknowledged.

References

Aljallis, E., Sarshar, M. A., Datla, R., Sikka, V., Jones, A. & Choi, C. H. 2013 Experimental study of skin friction drag reduction on superhydrophobic flat plates in high Reynolds number boundary layer flow. Phys. Fluids 25 (2), 025103.Google Scholar
Arenas-Navarro, I.2017 Numerical simulations for turbulent drag reduction using liquid infused surfaces. PhD thesis, The University of Texas at Dallas.Google Scholar
Bechert, D. W., Bruse, M., Hage, W., Van Der Hoeven, J. G. T. & Hoppe, G. 1997 Experiments on drag-reducing surfaces and their optimization with an adjustable geometry. J. Fluid Mech. 338, 5987.10.1017/S0022112096004673Google Scholar
Bidkar, R. A., Leblanc, L., Kulkarni, A. J., Bahadur, V., Ceccio, S. L. & Perlin, M. 2014 Skin-friction drag reduction in the turbulent regime using random-textured hydrophobic surfaces. Phys. Fluids 26 (8), 085108.10.1063/1.4892902Google Scholar
Checco, A., Ocko, B. M., Rahman, A., Black, C. T., Tasinkevych, M., Giacomello, A. & Dietrich, S. 2014 Collapse and reversibility of the superhydrophobic state on nanotextured surfaces. Phys. Rev. Lett. 112 (21), 216101.Google Scholar
Daniello, R. J., Waterhouse, N. E. & Rothstein, J. P. 2009 Drag reduction in turbulent flows over superhydrophobic surfaces. Phys. Fluids 21 (8), 085103.Google Scholar
Dean, R. B. 1978 Reynolds number dependence of skin friction and other bulk flow variables in two-dimensional rectangular duct flow. Trans. ASME J. Fluids Engng 100 (2), 215223.10.1115/1.3448633Google Scholar
Fu, M. K., Arenas, I., Leonardi, S. & Hultmark, M. 2017 Liquid-infused surfaces as a passive method of turbulent drag reduction. J. Fluid Mech. 824, 688700.10.1017/jfm.2017.360Google Scholar
García-Mayoral, R. & Jiménez, J. 2012 Scaling of turbulent structures in riblet channels up to Re 𝜏 ≈ 550. Phys. Fluids 24 (10), 105101.Google Scholar
Gose, J., Golovin, K., Boban, M., Mabry, J., Tuteja, A., Perlin, M. & Ceccio, S. 2018 Characterization of superhydrophobic surfaces for drag reduction in turbulent flow. J. Fluid Mech. 845, 560580.10.1017/jfm.2018.210Google Scholar
Karatay, E., Tsai, P. A. & Lammertink, R. G. H. 2013 Rate of gas absorption on a slippery bubble mattress. Soft Matt. 9 (46), 1109811106.10.1039/c3sm51928gGoogle Scholar
Lagrava, D., Malaspinas, O., Latt, J. & Chopard, B. 2012 Advances in multi-domain lattice Boltzmann grid refinement. J. Comput. Phys. 231 (14), 48084822.Google Scholar
Ling, H., Katz, J., Fu, M. & Hultmark, M. 2017 Effect of Reynolds number and saturation level on gas diffusion in and out of a superhydrophobic surface. Phys. Rev. Fluids 2 (12), 124005.10.1103/PhysRevFluids.2.124005Google Scholar
Ling, H., Srinivasan, S., Golovin, K., McKinley, G. H., Tuteja, A. & Katz, J. 2016 High-resolution velocity measurement in the inner part of turbulent boundary layers over super-hydrophobic surfaces. J. Fluid Mech. 801, 670703.10.1017/jfm.2016.450Google Scholar
Min, T. & Kim, J. 2004 Effects of hydrophobic surface on skin-friction drag. Phys. Fluids 16 (7), L55.Google Scholar
Nishino, T., Meguro, T., Nakamae, K., Matsushita, M. & Ueda, Y. 1999 The lowest surface free energy based on -CF3 alignment. Langmuir 15, 43214323.10.1021/la981727sGoogle Scholar
Park, H., Park, H. & Kim, J. 2013 A numerical study of the effects of superhydrophobic surface on skin-friction drag in turbulent channel flow. Phys. Fluids 25 (11), 110815.Google Scholar
Park, H., Sun, G. & Kim, C.-J. 2014 Superhydrophobic turbulent drag reduction as a function of surface grating parameters. J. Fluid Mech. 747, 722734.10.1017/jfm.2014.151Google Scholar
Philip, J. R. 1972 Flows satisfying mixed no-slip and no-shear conditions. Z. Angew. Math. Phys. 23 (3), 353372.Google Scholar
Rastegari, A. & Akhavan, R. 2015 On the mechanism of turbulent drag reduction with super-hydrophobic surfaces. J. Fluid Mech. 773, R4.Google Scholar
Rastegari, A. & Akhavan, R. 2018a The common mechanism of turbulent skin-friction drag reduction with super-hydrophobic longitudinal microgrooves and riblets. J. Fluid Mech. 838, 68104.10.1017/jfm.2017.865Google Scholar
Rastegari, A. & Akhavan, R. 2018b Effect of interface dynamics on drag reduction and sustainability of superhydrophobic and liquid-infused surfaces in turbulent flow. Bull. Am. Phys. Soc. 63 (13), 143.Google Scholar
Reholon, D. & Ghaemi, S. 2018 Plastron morphology and drag of a superhydrophobic surface in turbulent regime. Phys. Rev. Fluids 3, 104003.Google Scholar
Rothstein, J. P. 2010 Slip on superhydrophobic surfaces. Annu. Rev. Fluid Mech. 42, 89109.10.1146/annurev-fluid-121108-145558Google Scholar
Samaha, M. A., Tafreshi, H. V. & Gad-el Hak, M. 2012 Influence of flow on longevity of superhydrophobic coatings. Langmuir 28 (25), 97599766.10.1021/la301299eGoogle Scholar
Schellenberger, F., Encinas, N., Vollmer, D. & Butt, H.-J. 2016 How water advances on superhydrophobic surfaces. Phys. Rev. Lett. 116 (9), 096101.Google Scholar
Schewe, G. 1983 On the structure and resolution of wall-pressure fluctuations associated with turbulent boundary-layer flow. J. Fluid Mech. 134, 311328.10.1017/S0022112083003389Google Scholar
Schönecker, C., Baier, T. & Hardt, S. 2014 Influence of the enclosed fluid on the flow over a microstructured surface in the Cassie state. J. Fluid Mech. 740, 168195.10.1017/jfm.2013.647Google Scholar
Seo, J., García-Mayoral, R. & Mani, A. 2015 Pressure fluctuations and interfacial robustness in turbulent flows over superhydrophobic surfaces. J. Fluid Mech. 783, 448473.10.1017/jfm.2015.573Google Scholar
Seo, J. & Mani, A. 2016 On the scaling of the slip velocity in turbulent flows over superhydrophobic surfaces. Phys. Fluids 28, 025110.10.1063/1.4941769Google Scholar
Sillero, J. A., Jiménez, J. & Moser, R. D. 2013 One-point statistics for turbulent wall-bounded flows at Reynolds numbers up to 𝛿+ ≈ 2000. Phys. Fluids 25 (10), 105102.Google Scholar
Spalart, P. R. & McLean, D. 2011 Drag reduction: enticing turbulence, and then an industry. Phil. Trans. R. Soc. Lond. A 369, 15561569.Google Scholar
Srinivasan, S., Kleingartner, J. A., Gilbert, J. B., Cohen, R. E., Milne, A. J. B. & McKinley, G. H. 2015 Sustainable drag reduction in turbulent Taylor–Couette flows by depositing sprayable superhydrophobic surfaces. Phys. Rev. Lett. 114 (1), 014501.10.1103/PhysRevLett.114.014501Google Scholar
Succi, S. 2001 The Lattice Boltzmann Equation: For Fluid Dynamics and Beyond. Oxford University Press.Google Scholar
Tsuji, Y., Fransson, J. H. M., Alfredsson, P. H. & Johansson, A. V. 2007 Pressure statistics and their scaling in high-Reynolds-number turbulent boundary layers. J. Fluid Mech. 585, 140.Google Scholar
Van Buren, T. & Smits, A. J. 2017 Substantial drag reduction in turbulent flow using liquid-infused surfaces. J. Fluid Mech. 827, 448456.10.1017/jfm.2017.503Google Scholar
Wexler, J. S., Jacobi, I. & Stone, H. A. 2015 Shear-driven failure of liquid-infused surfaces. Phys. Rev. Lett. 114 (16), 168301.10.1103/PhysRevLett.114.168301Google Scholar
Ybert, C., Barentin, C., Cottin-Bizonne, C., Joseph, P. & Bocquet, L. 2007 Achieving large slip with superhydrophobic surfaces: scaling laws for generic geometries. Phys. Fluids 19 (12), 123601.10.1063/1.2815730Google Scholar
Zanoun, E.-S., Nagib, H. & Durst, F. 2009 Refined c f relation for turbulent channels and consequences for high-re experiments. Fluid Dyn. Res. 41 (2), 021405.Google Scholar
Zhang, J., Tian, H., Yao, Z., Hao, P. & Jiang, N. 2015 Mechanisms of drag reduction of superhydrophobic surfaces in a turbulent boundary layer flow. Exp. Fluids 56 (9), 179.Google Scholar
Zheng, Q.-S., Yu, Y. & Zhao, Z.-H. 2005 Effects of hydraulic pressure on the stability and transition of wetting modes of superhydrophobic surfaces. Langmuir 21 (26), 1220712212.10.1021/la052054yGoogle Scholar
Figure 0

Figure 1. Schematic of the channels with SH walls, the coordinate system and the computational grid: (a) SH longitudinal microgrooves; (b) SH aligned microposts; (c) computational grid and protrusion angle.

Figure 1

Table 1. Summary of simulation parameters and resulting drag reductions, $DR$, in the present study.

Figure 2

Figure 2. Schematic representation of $(B-B_{0})$ and $U_{s}^{+}$ in turbulent flow with microtextured walls: —— (black), – $\cdot$ $\cdot$ – (turquoise), mean velocity profiles in the base turbulent flow and the flow with microtextured walls, respectively; – – – (black), – $\cdot$ – (turquoise), logarithmic representations of the mean velocity profiles in the base turbulent flow and the flow with microtextured walls, respectively.

Figure 3

Figure 3. The magnitude of drag reduction, $DR$, as a function of $(B-B_{0})$ and the friction Reynolds number of the base flow, $Re_{\unicode[STIX]{x1D70F}_{0}}$. Present DNS studies in SH turbulent channel flow: , , , ,  (navy), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/2$, $1/8$, $1/16$, $1/32$, $1/64$; , , , ,  (navy), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$, $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/2$, $1/8$, $1/16$, $1/32$, $1/64$; , ,  ,  (blue), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/8$, $1/16$, $1/32$, $1/64$; , , ,  (blue), SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/8$, $1/16$, $1/32$, $1/64$;  (navy), SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$;  (blue), SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$;  (lavender), DNS of turbulent channel flow with LI longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 180$, $\unicode[STIX]{x1D703}_{p}=0^{o}$, $\unicode[STIX]{x1D719}_{s}=1/2$, $N=0.1,1,2.5,10,20,100$ (Fu et al.2017); ▪ (red), experiments in turbulent boundary layer flow with SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 460$ and $560$ (Zhang et al.2015); ♦ (green), experiments in turbulent Taylor–Couette flow with SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 1193$ and $3810$ (Srinivasan et al.2015); ● (orange), experiments in turbulent boundary layer flow with SH longitudinal microgrooves or SH random microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 863,1408$ and $4287$ (Ling et al.2016). The black lines show the predictions of (3.1) in turbulent channel flow for: ——, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ ($Re_{b}=3600$); — —, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$ ($Re_{b}=7860$); – – –, $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{3}$ ($Re_{b}\approx 2\times 10^{4}$); – $\cdot$ –, $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{4}$ ($Re_{b}\approx 2.6\times 10^{5}$); – $\cdot$ $\cdot$ –, $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{5}$ ($Re_{b}\approx 3.2\times 10^{6}$); $\cdots \cdots$, $Re_{\unicode[STIX]{x1D70F}_{0}}=10^{6}$ ($Re_{b}\approx 3.8\times 10^{7}$).

Figure 4

Figure 4. Contours of instantaneous streamwise velocity, $U^{+}(x,y,z)$, at $z=0$ for: (a) SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 63$; (b) SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 63$; (c) SH aligned microposts, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 56$.

Figure 5

Figure 5. Production and dissipation of turbulence kinetic energy in turbulence channel flow with SH longitudinal microgrooves at (a,c,d,e) $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 63$ and (b,f,g,h) $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 63$: (a,b) ——, — —, $\cdots \cdots$, – $\cdot$ $\cdot$ – (turquoise), total production $\langle {\mathcal{P}}\rangle ^{+}=-\langle \overline{u_{i}u_{j}}\overline{S}_{ij}\rangle ^{+}$, production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$, production arising from $-\langle \overline{uv}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\rangle ^{+}$, dissipation $\langle \unicode[STIX]{x1D700}\rangle ^{+}=\langle 2\unicode[STIX]{x1D708}\overline{s_{ij}s_{ij}}\rangle ^{+}$, in turbulent channel flow with SH walls; – – –, – $\cdot$ – (black), production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$ and dissipation, $\langle \unicode[STIX]{x1D700}\rangle ^{+}$, in base turbulent channel flow with smooth, no-slip walls; (c,d,e; f,g,h) spanwise variation of the mean streamwise velocity, $\bar{U}^{+}$, total production, ${\mathcal{P}}^{+}$, and dissipation, $\unicode[STIX]{x1D700}^{+}$, at different $z^{+}$ in turbulent channel flow with SH longitudinal microgrooves at $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $\unicode[STIX]{x1D703}_{p}=-30^{\circ }$, respectively. The blue bars on the lower horizontal axes of (c–h) denote the locations of the no-slip surfaces. $L=g+w$ denotes the pitch of the microgrooves.

Figure 6

Figure 6. Production and dissipation of turbulence kinetic energy in turbulent channel flow with SH aligned microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 56$: (a) total production $\langle {\mathcal{P}}\rangle ^{+}=-\langle \overline{u_{i}u_{j}}\overline{S}_{ij}\rangle ^{+}$, production arising from $-\langle \overline{uw}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}z\rangle ^{+}$, production arising from $-\langle \overline{uv}\unicode[STIX]{x2202}\bar{U}/\unicode[STIX]{x2202}y\rangle ^{+}$, dissipation $\langle \unicode[STIX]{x1D700}\rangle ^{+}=\langle 2\unicode[STIX]{x1D708}\overline{s_{ij}s_{ij}}\rangle ^{+}$, in turbulent channel flow with SH walls, compared to base turbulent channel flow with smooth, no-slip walls; (b) schematic representation of sections A-A and B-B where turbulence statistics are shown in (c–e) and (f–h); (c–e; f–h) spanwise variation of the mean streamwise velocity, $\bar{U}^{+}$, total production, ${\mathcal{P}}^{+}$, dissipation, $\unicode[STIX]{x1D700}^{+}$, at different $z^{+}$ of sections A-A and B-B, respectively. Line types in (a) as in figure 5. The blue bars on the lower horizontal axes of (c–h) denote the locations of the no-slip surfaces. $L=g+w$ denotes the pitch of the microposts.

Figure 7

Figure 7. The scaling of $U_{s}^{+}$ and $(B-B_{0})$ with surface microtexture parameters in turbulent flow with SH or LI walls: (a–d) present DNS data in turbulent channel flow with SH longitudinal microgrooves or SH aligned microposts at $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 222$ and $442$, $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $-30^{\circ }$; (e,f) DNS data from other investigators in turbulent channel flow with SH longitudinal microgrooves, LI longitudinal microgrooves or SH aligned microposts at $180\lesssim Re_{\unicode[STIX]{x1D70F}_{0}}\lesssim 590$, $1/36\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$. Symbols as in figure 3. Other symbols in (e,f): , ,  (brown), $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 180$, $\unicode[STIX]{x1D719}_{s}=1/8$, $1/4$, $1/2$, , , ,  (green), $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 395$, $\unicode[STIX]{x1D719}_{s}=1/16$, $1/8$, $1/4$, $1/2$, , , ,  (red), $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 590$, $\unicode[STIX]{x1D719}_{s}=1/16$, $1/8$, $1/4$, $1/2$, SH longitudinal microgrooves, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ (Park et al.2013); ,  (brown), $Re_{\unicode[STIX]{x1D70F}}\approx 195$, $\unicode[STIX]{x1D719}_{s}=1/64$, $1/9$, , ,  (green), $Re_{\unicode[STIX]{x1D70F}}\approx 400$, $\unicode[STIX]{x1D719}_{s}=1/36$, $1/16$, $1/9$, SH aligned microposts, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ (Seo & Mani 2016).

Figure 8

Figure 8. Ratio of the effective slip length, $L_{s,N}$, obtained with a finite viscosity ratio, $N$, based on the analytical solution of Schönecker et al. (2014), to the effective slip length, $L_{s,\infty }$, obtained with the assumption of shear-free liquid/air interfaces (Philip 1972), in laminar shear flow over LI longitudinal microgrooves for groove depth $d=g$ and $1/64\leqslant \unicode[STIX]{x1D719}_{s}\leqslant 1/2$: $\cdots \cdots$, $\unicode[STIX]{x1D719}_{s}=1/2$; — ⋅ ⋅ —, $\unicode[STIX]{x1D719}_{s}=1/8$; — ⋅ —, $\unicode[STIX]{x1D719}_{s}=1/16$; – – –, $\unicode[STIX]{x1D719}_{s}=1/32$; ——, $\unicode[STIX]{x1D719}_{s}=1/64$.

Figure 9

Figure 9. Effect of dynamic deformation of the interface on (a) $U_{s}^{+}$ and (b) $(B-B_{0})$, based on DNS results with deformable interfaces, but pinned contact lines, in turbulent channel flows with LI longitudinal microgrooves at $Re_{\unicode[STIX]{x1D70F}}\approx 180$, $\unicode[STIX]{x1D719}_{s}=1/2$, $N=2.5$ (Arenas-Navarro 2017):  (lavender), $We_{\unicode[STIX]{x1D70F}}=0$;  (dark violet), $We_{\unicode[STIX]{x1D70F}}=0.024$.

Figure 10

Figure 10. Probability density function and cumulative distribution function of the wall pressure fluctuations in turbulent channel flow with SH longitudinal microgrooves, compared to turbulent channel flow with smooth, no-slip walls: (a) probability density function, $f_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\!\}^{1/2})$; (b) cumulative distribution function, $F_{p}(p_{w}^{+}/\{\langle \overline{p^{2}}\rangle _{w}^{+}\}^{1/2})$; – – –, – $\cdot$ –, SH longitudinal microgrooves, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$, $\unicode[STIX]{x1D719}_{s}=1/64$, $g^{+0}\approx 63$, $\unicode[STIX]{x1D703}_{p}=0^{\circ }$ and $-30^{\circ }$, respectively; $\cdots \cdots$, base turbulent channel flow with smooth, no-slip walls, $Re_{\unicode[STIX]{x1D70F}_{0}}\approx 442$; —— (orange), Gaussian distribution.

Figure 11

Figure 11. Mean-squared turbulent wall pressure fluctuations in turbulent channel flow with SH longitudinal microgrooves: (a) $\langle \bar{p^{2}}\rangle _{w,SH}^{+}$ as a function of $Re_{\unicode[STIX]{x1D70F}}$; (b) $\langle \bar{p^{2}}\rangle _{w,SH}^{+}-\unicode[STIX]{x1D6E4}(g^{+})$ as a function of $Re_{\unicode[STIX]{x1D70F}}$, where $\unicode[STIX]{x1D6E4}(g^{+})=2.32\ln (\{g^{+}\}^{3/4})-4.5$. Symbols as in figure 3. $\cdots \cdots$, the fit $\langle \bar{p^{2}}\rangle _{w,smooth}^{+}=2.31\ln (Re_{\unicode[STIX]{x1D70F}})-9.5$, suggested by Sillero, Jiménez & Moser (2013) for turbulent channel flows with smooth, no-slip walls.

Figure 12

Figure 12. The maximum permissible $We^{+}$ and $g^{+}$ at a given $Re_{\unicode[STIX]{x1D70F}}$ to maintain the stability of SH longitudinal microgrooves in turbulent channel flow. Lines show the predictions of (4.2) for: ——, $Re_{\unicode[STIX]{x1D70F}}\approx 200$; — —, $Re_{\unicode[STIX]{x1D70F}}\approx 400$; – – –, $Re_{\unicode[STIX]{x1D70F}}\approx 10^{3}$; — ⋅ —, $Re_{\unicode[STIX]{x1D70F}}\approx 10^{4}$; – $\cdot$ $\cdot$ –, $Re_{\unicode[STIX]{x1D70F}}\approx 10^{5}$; $\cdots \cdots$$Re_{\unicode[STIX]{x1D70F}}\approx 10^{6}$.