Hostname: page-component-54dcc4c588-r5qjk Total loading time: 0 Render date: 2025-09-27T09:02:46.385Z Has data issue: false hasContentIssue false

Analysis of flow-wall deformation coupling in high Reynolds number compliant wall boundary layers

Published online by Cambridge University Press:  23 September 2025

Yuhui Lu
Affiliation:
Department of Mechanical Engineering, Johns Hopkins University, Baltimore, MD 21218, USA
Tianrui Xiang
Affiliation:
Department of Mechanical Engineering, Johns Hopkins University, Baltimore, MD 21218, USA
Tamer A. Zaki
Affiliation:
Department of Mechanical Engineering, Johns Hopkins University, Baltimore, MD 21218, USA
Joseph Katz*
Affiliation:
Department of Mechanical Engineering, Johns Hopkins University, Baltimore, MD 21218, USA
*
Corresponding author: Joseph Katz, katz@jhu.edu

Abstract

Interactions of turbulent boundary layers with a compliant surface are investigated experimentally at Reτ = 3300–8900. Integrating tomographic particle tracking with Mach–Zehnder interferometry enables simultaneous mapping of the compliant wall deformation and the three-dimensional velocity and pressure fields. Our initial study (J. Fluid. Mech. vol. 980, R2) shows that the flow–deformation correlations decrease with increasing Reτ, despite an order of magnitude increase in deformation amplitude. To elucidate the mechanisms involved, the same velocity, pressure and kinetic energy fields are decomposed to ‘wave-coherent’ and ‘stochastic’ parts using a Hilbert projection method. The phase dependent coherent variables, especially the pressure, are highly correlated with the wave, but decrease with increasing Reτ. While the coherent energy is 6 %–10 % of the stochastic level, the pressure root mean square is comparable near the wall. The energy flux between the coherent and stochastic parts and the pressure diffusion reverse sign at the critical layer. To explain the Reτ dependence, the characteristic deformation wavelength (three times the thickness) is compared with the scales of the energy-containing eddies in the boundary layer represented by the k−1 range in the energy spectrum. When the deformation wavelength is matched with the kxEuu peak at the present lowest Reτ, the flow–deformation correlations and coherent pressure become strong, even for submicron deformations. In this case, the flow and wall motion become phase locked, suggesting resonant behaviours. As Reτ increases, the wall wavelengths and spectral range of attached eddies are no longer matched, resulting in reduced correlations and lower coherent energy and pressure, despite larger deformation.

Information

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

The interactions of compliant surfaces with laminar or turbulent boundary layers have been the subject of theoretical, experimental and numerical research over the past decades. The main research interests have included changes to skin friction, flow–deformation interactions as well as noise suppression. In the present study, we focus on the flow–deformation interactions of a fully developed boundary layer with a viscoelastic compliant wall. The early attempts to study the compliant wall boundary layer date back to Kramer’s (Reference Kramer1957, Reference Kramer1962) effort to mimic the dolphin’s skin aimed at potential drag reduction. Despite the numerous studies that have followed, there is no consensus on whether one can reduce drag using a compliant wall. The past studies can be divided to those focusing on compliance effects on laminar boundary layer transition to turbulence, and those investigating turbulent flows. Early studies (Benjamin Reference Benjamin1960; Landahl Reference Landahl1962; Lee, Fisher & Schwarz Reference Lee, Fisher and Schwarz1995; Wang, Yeo & Khoo Reference Wang, Yeo and Khoo2006) in favour of drag reduction fall in the first category, and suggest that the wall compliance could suppress the Tollmien–Schlichting waves. For turbulent boundary layers, results have been mixed – some report drag reduction (e.g. Fisher & Blick Reference Fisher and Blick1966; Choi et al. Reference Choi, Yang, Clayton, Glover, Atlar, Semenov and Kulik1997; Fukagata et al. Reference Fukagata, Kern, Chatelain, Koumoutsakos and Kasagi2008), some find negligible effects (e.g. Lissaman & Harris Reference Lissaman and Harris1969; Mcmichael, Klebanoff & Mease Reference Mcmichael, Klebanoff and Mease1980; Endo & Himeno Reference Endo and Himeno2002) and others observe drag increase (Boggs & Hahn Reference Boggs and Hahn1962). More recent experimental studies using two-dimensional (2-D) particle image velocimetry (PIV) (Wang, Koley & Katz Reference Wang, Koley and Katz2020; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022) and three-dimensional (3-D) tomographic particle tracking (Lu et al. Reference Lu, Xiang, Zaki and Katz2024) show an increasing downward shift in the mean velocity profile, indicating drag increase, with increasing deformation magnitude. Recent direct numerical simulation (DNS) by Rosti & Brandt (Reference Rosti and Brandt2017) and Esteghamatian, Katz & Zaki (Reference Esteghamatian, Katz and Zaki2022) have also observed the momentum deficit, as well as an increase in near-wall turbulence with decreasing material stiffness, consistent with the experiments.

Several types of wave propagation have been observed on the compliant surface. When the free stream velocity (U 0) is larger than the shear speed of the material ( $c_{t}=\sqrt{G/\rho _{s}}$ , where G and $\rho _{s}$ are the shear modulus and density), e.g. U 0/ $c_{t}$ > 2.8, Gad-El-Hak, Blackwelder & Riley (Reference Gad-El-Hak, Blackwelder and Riley1984) and Duncan (Reference Duncan1986) show that the deformation wave contains a slowly propagating (∼0.05U 0), high-amplitude, ‘static divergence wave’. This wave is also observed in Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) for U 0/ $c_{t}$ > 3.5. With increasing material stiffness, the deformation wave consists primarily of ‘travelling-wave flutter’, which travels at a fraction of the free stream velocity (Gad-el-Hak Reference Gad-El-Hak1986). Data compiled in the review by Carpenter, Davies & Lucey (Reference Carpenter, Davies and Lucey2000) indicates that the phase speed of this wave varies between 0.4U 0 to 0.8U 0, consistent with recent experiments (Zhang et al. Reference Zhang, Wang, Blake and Katz2017; Wang et al. Reference Wang, Koley and Katz2020; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022; Lu et al. Reference Lu, Xiang, Zaki and Katz2024). This phase speed does not show a consistent trend with either hydrodynamic or material parameters, a topic that has not been resolved yet. In addition, Rosti & Brandt (Reference Rosti and Brandt2017) and Wang et al. (Reference Wang, Koley and Katz2020) observe long streamwise deformations resembling boundary layer superstructures (Hutchins & Marusic Reference Hutchins and Marusic2007). Essentially all the recent experimental and computational studies (Rosti & Brandt Reference Rosti and Brandt2017; Wang et al. Reference Wang, Koley and Katz2020; Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022) report that the wall deformations enhance the near-wall turbulence and Reynolds shear stress, although the observed profiles vary. Wang et al. (Reference Wang, Koley and Katz2020) also observe that the deformations decrease the correlation length scales of streamwise velocity fluctuations, which they attribute to ‘scrambling’ of the eddies by the surface motions.

Studying the dynamic interaction between flow and deformation requires time-resolved measurements of both. Early experiments were limited to single-point flow data and separate deformation mapping (e.g. Gad-El-Hak et al. Reference Gad-El-Hak, Blackwelder and Riley1984; Lee, Fisher & Schwarz Reference Lee, Fisher and Schwarz1993a ). Over recent years, techniques such as holography (Lee, Fisher & Schwarz Reference Lee, Fisher and Schwarz1993b ), laser Doppler vibrometer (Castellini, Martarelli & Tomasini Reference Castellini, Martarelli and Tomasini2006), Mach–Zehnder interferometry (Zhang, Miorini & Katz Reference Zhang, Miorini and Katz2015), background oriented schlieren (Charruault, Greidanus & Westerweel Reference Charruault, Greidanus, Breugem and Westerweel2018) and digital image correlation (Huynh & Mckeon Reference Huynh and Mckeon2020) have been used to measure wall deformation. Zhang et al. (Reference Zhang, Miorini and Katz2015) and Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) have performed the first simultaneous measurement, combining tomographic PIV and Mach–Zehnder interferometry (MZI). The pressure field is measured by spatial integration of the material acceleration and correlated with the spatiotemporal distribution of deformation. In their studies, the wall is quite rigid (c t = 6.8U 0), resulting in submicron deformations with heights (d) that are 2–3 orders of magnitude smaller than the wall unit (δ ν = ν/u τ , where ν is the kinematic viscosity and u τ = (τ $_{w}$ )1/2 is the friction velocity, τ $_{w}$ is the wall shear stress and ρ is the fluid density). Hence, the interaction mostly involves one-way coupling, i.e. the flow causes wall deformation, but the deformation is too small to have a significant effect on the flow. Conditional sampling shows that surface bumps preferentially reside under pressure minima located between the legs of hairpin vortices, and dimples are associated with pressure maxima at the transition between sweeps and ejections. Subsequent experiments by Wang et al. (Reference Wang, Koley and Katz2020) consist of MZI applications and separate 2-D PIV measurements under conditions of two-way coupling, involving a softer material with d being of the same order as δ ν . In these cases, the velocity deficit in the inner part of the boundary layers starts to appear even for deformations of the order of 0.1δ ν . For the same compliant wall, Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) apply tomographic particle tracking to measure the 3-D flow and pressure, and simultaneous MZI for mapping the 2-D wall deformation. This approach is also used for the present study.

Theoretical predictions of linear viscoelastic material response to harmonic excitation have been introduced by Chase (Reference Chase1991) and Benschop et al. (Reference Benschop, Greidanus, Delfos, Westerweel and Breugem2019). Applications of these models by Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) and Wang et al. (Reference Wang, Koley and Katz2020) lead to the conclusion that the characteristic wavelength of the peak material response is three times the coating thickness (l 0). The experiments by Zhang et al. (Reference Zhang, Wang, Blake and Katz2017), Wang et al. (Reference Wang, Koley and Katz2020) and Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) have confirmed these predictions. For a wide range of material properties (U 0 < 3.4C t ), the deformation root mean square (r.m.s.) scaled by l 0 shows a linear relationship with the liquid pressure r.m.s., scaled by the shear modulus (Benschop et al. Reference Benschop, Greidanus, Delfos, Westerweel and Breugem2019; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022; Lu et al. Reference Lu, Xiang, Zaki and Katz2024), indicating a linear normal stress–strain relationship.

In numerical simulations, a proper model for the compliant material is essential. Recent DNS (Rosti & Brandt Reference Rosti and Brandt2017; Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022) model the compliant wall as a hyperelastic material and simulate the flow/motion in both media. While Rosti & Brandt (Reference Rosti and Brandt2017) focus on the mean flow profiles and turbulence statistics, Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022) provide detailed accounts of the flow structure. For example, they show the contributions of surface acceleration and pressure gradient on the generation of vorticity. For large deformation, d ∼ 20δ ν , the near-wall shear layer detaches at the deformation peak up to a height where the local advection speed is equal to the Rayleigh wave speed. The low momentum fluid on the leeward side is subsequently lifted up by the upward wall motion, and then entrained into the high-speed flow at higher elevations. For small deformation, d ≤ δ ν , the experiments of Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) and Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) show that the pressure-deformation correlation also peaks at an elevation where the local mean velocity is equal to the surface wave speed. Adopting terminology introduced by Miles (Reference Miles1957), and widely used in the air–sea interaction community, Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) refer to this elevation as a ‘critical layer’, and show that it plays an important role in the flow–deformation interactions. For example, the near-wall turbulence is highly correlated and phase-locked with the deformation up to the critical height (y c ), i.e. it has the same advection speed as the deformation wave. At higher elevations, the turbulence is advected with the local mean flow. In these experiments, the critical height increases from y c + = 60–190 as Re τ increases from 3300 to 8900. Here, Re τ = u τ δ/ν and y c + = y c ν , where δ is the boundary layer thickness. It should be noted that the phase lock persists at the lowest Re τ when the characteristic deformation height is of the order of 0.1δ ν . In contrast to roughness effects, where such small surface perturbations are expected to have a minimal effect on the flow, the turbulence-wave interactions over the compliant wall have significant impact on the near wall dynamics. A plausible explanation for the mechanisms involved is one of the topics of the present paper.

Owing to its significance to the present study, this paragraph provides a brief background on the critical layer and its effect on wind–wave interactions in oceanography. Early efforts to model air–sea interactions date back to Jeffreys’ (Reference Jeffreys1925) sheltering hypothesis, which assumes that the symmetric distribution of surface pressure is broken by flow separation on the leeward side, resulting in pressure work on the wave. Phillips (Reference Phillips1957) attributes the wave growth to resonant forcing of the surface waves by turbulent pressure fluctuations, while Miles (Reference Miles1957) proposes that an inviscid instability involving a resonance of waves with the airflow occurs at the critical height. Subsequently, Lighthill (Reference Lighthill1962) further elucidates Miles’s theory based on vortex induced forces, leading to the conclusion that the wind–wave energy transfer is concentrated at the critical layer. While Miles’s theory has triggered some criticism over the years (e.g. Krasitskii & Zaslavskii Reference Krasitskii and Zaslavskii1978; Riley, Donelan & Hui Reference Riley, Donelan and Hui1982), more recent open-ocean data (e.g. Hristov, Miller & Friehe Reference Hristov, Miller and Friehe2003; Grare, Lenain & Melville Reference Grare, Lenain and Melville2013) and laboratory experiments (Carpenter, Buckley & Veron Reference Carpenter, Buckley and Veron2022) support the important role of the critical layer. As noted above, in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024), we show that for a compliant wall wave, the wave–turbulence correlations peak near the critical height.

In studies of flow-wave interactions, instead of the typical Reynolds decomposition, Hussain & Reynolds (Reference Hussain and Reynolds1970) introduce a triple decomposition that separates the flow variables to mean, wave-coherent and background turbulence components. The corresponding decomposed conservation equations for kinetic energy are derived in Reynolds & Hussain (Reference Reynolds and Hussain1972), and modified equations for a curvilinear coordinate system have been presented recently by Yousefi & Veron (Reference Yousefi and Veron2020). Separating the wave-coherent motions from the turbulence in experimental data has been a challenge, especially for cases with high background turbulence. Several methods have been utilized, including phase averaging (e.g. Einaudi & Finnigan Reference Einaudi and Finnigan1993; Yang & Shen Reference Yang and Shen2010), Hilbert projection (e.g. Hristov, Friehe & Miller Reference Hristov, Friehe and Miller1998; Hristov & Plancarte Reference Hristov and Ruiz-Plancarte2014) and spectral filtering (e.g. Grare et al. Reference Grare, Lenain and Melville2013). Phase averaging is particularly effective when the wave is monochromatic or when only one dominant component is considered, and the latter two methods have been employed when the wave field is broadband, e.g. in field ocean data. In the present study, we use the Hilbert projection to characterize the flow components that are coherent with the wall deformation. In the oceanic boundary layer, the production of wave-coherent kinetic energy is smaller than that of the ‘stochastic’ background turbulent kinetic energy (Rutgersson & Sullivan Reference Rutgersson and Sullivan2005; Yousefi, Veron & Buckley Reference Yousefi, Veron and Buckley2021) but not negligible. Several studies have shown an energy shift from the wave-coherent motions to the background turbulence (Hsu, Hsu & Street Reference Hsu, Hsu and Street1981; Makin & Kudryavtsev Reference Makin and Kudryavtsev1999; Yousefi et al. Reference Yousefi, Veron and Buckley2021), while others (Rutgersson & Sullivan Reference Rutgersson and Sullivan2005; Hara & Sullivan Reference Hara and Sullivan2015) have observed an opposite trend. The present study examines these energy fluxes in the compliant wall boundary layer.

This paper builds upon the introductory analysis in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024), in which we demonstrate the above-mentioned linear pressure-deformation scaling, and the changes to the advection speed of turbulent fluctuations across the critical layer, from the wave speed to the local velocity. Here, we utilize the same three laboratory datasets at Re τ = 3300–8900, to investigate the effect of the critical layer on the kinetic energy budget, and to decompose the unsteady kinetic energy to deformation-wave-coherent and stochastic motions in order to elucidate their interactions.

Furthermore, in an attempt to understand the phase locking between the turbulence and the submicron-scale wall motions, we compare the streamwise scales of wall deformations with those of the ‘attached eddies’ in the boundary layer (Townsend Reference Townsend1976; Perry & Chong Reference Perry and Chong1982; Nickels et al. Reference Nickels, Marusic, Hafez and Chong2005). A brief relevant background on this topic is provided while presenting the data. The experiments involve simultaneous measurements of the time-resolved volumetric flow fields and spatial distribution of wall deformation. The rest of the paper is organized as follows. Descriptions of the flow configuration, experimental set-up and data analysis procedures are summarized in § 2. Results of the wave-coherent and stochastic flow variables (velocity, pressure, vorticity, stresses, pressure r.m.s., kinetic energy and their budget terms) are presented in § 3. A discussion comparing the deformation wavelength with the length scales of energy-containing eddies in the boundary layer, which is aimed at explaining the Reynolds number scaling of flow–deformation correlations, is provided in § 4, followed by conclusions in § 5.

2. Experimental set-up and procedures

Throughout the paper, x, y and z are the streamwise, wall-normal and spanwise directions, respectively, and the corresponding velocity components are u, ${v}$ and ${w}$ . The free stream velocity, material thickness and critical height are denoted as U 0, l 0 and y c , respectively. Here, U 0 is the velocity outside of the boundary layer at the location of measurements, and as demonstrated for a rigid wall in the same facility (Wang et al. Reference Wang, Koley and Katz2020), the grooves generate a fully developed turbulent boundary layer in the sample area with the typical velocity profile, consisting of viscous, buffer, log and outer layers. Accordingly, U 0 is not expected to change significantly over the sample area. The friction velocity is $u_{\tau }=\sqrt{\tau _{w}/\rho }$ , where $\tau _{w}$ and $\rho$ are the wall shear stress and the fluid density, the viscous wall unit is $\delta _{\nu }=\nu /u_{\tau }$ , where $\nu$ is the kinematic viscosity and the friction Reynolds number is $\textit{Re}_{\tau }=\delta /\delta _{\nu }$ , where $\delta$ is the boundary layer thickness. Superscript ‘+’ denotes normalization with inner variables, namely by $\delta _{\nu }$ for length scales, $u_{\tau }$ for velocity, ${u_{\tau }}^{2}$ for Reynolds stresses and $\rho {u_{\tau }}^{2}$ for pressure.

The experiments have been performed in the recently constructed Johns Hopkins University refractive index-matched water tunnel, which is sketched in figure 1. This facility is filled with a 62 % by weight aqueous sodium iodide (NaI) solution, which has the same refractive index as acrylic (1.487), but not that of the compliant wall, a kinematic viscosity of 1.1 × 10−6 m2 s−1, and a density of 1850 kg m−3 (Bai & Katz Reference Bai and Katz2014). The flow is driven by a 60 HP axial pump whose speed is controlled by an inverter. The top part of this tunnel has been used before in Wang et al. (Reference Wang, Koley and Katz2020) as an extension to another facility, but the pump and the 30.5 cm diameter vertical and lower piping have been constructed recently. The upper section includes a 279.4 cm long, mild inlet diffusor expanding from a 30.5 cm diameter pipe to a rectangular 34.8 × 40.6 cm2 settling chamber containing flow straighteners, a nozzle with an area ratio of 4.6 : 1, a test section and an outlet diffusor. The test section, with 50.8 mm thick acrylic windows on all sides (figure 1 b), is 83.8 cm long, and has a cross-section of 15.2 × 20.3 cm2. A half-filled tank above the test section is connected to sources of compressed gas and a vacuum pump to control the mean pressure in the loop. The tunnel is also equipped with a 0.92 m diameter and 1.52 m high cyclone bubble separator (figure 1 c) that can either be connected inline or bypassed using gate valves. The flow enters the cyclone circumferentially at the bottom of the tank and leaves circumferentially at the top of it. The inlet jet velocity is adjustable by restricting the inflow opening using a nozzle. The internal swirl causes bubble migration to the centre of the cyclone, where they are led to the pressure control tank. The cyclone is effective for degassing the water. The current set-up has been shown to remove free stream bubbles larger than 60 µm when the velocities at the entrance to the cyclone is 4.5 m s−1 (Lu et al. Reference Lu, Ram, Jose, Agarwal and Katz2021).

Figure 1. Schematics of (a) the refractive index-matched water tunnel, (b) the test section and (c) the cyclone separator.

Figure 2(a) is a sketch of the compliant wall coating, showing the location of the sample volume. Following Wang et al. (Reference Wang, Koley and Katz2020), the l 0 = 5 mm thick viscoelastic material is manufactured by mixing polydimethylsiloxane (PDMS, Dow Corning Sylgard 184) with a silicone gel softener (Sylgard 527) at a ratio of 1 : 7.5 by weight. The bottom acrylic tunnel window serves as the moulding base for the coating. Except for a narrow peripheral rim used for securing the coating, the entire bottom window is coated. A chemical primer (Dow Corning Dowsil PR-1204) is applied over the entire area except for the immediate vicinity of the sample area to avoid potential optical distortions. A fresh coating has been used for the present measurements. The present compliant material has a storage modulus (E) of 158 kPa and loss tangent of 0.01. Details on the coating properties and manufacturing procedures are available in Wang et al. (Reference Wang, Koley and Katz2020). Boundary layer trips, consisting of six, 0.5 mm high, triangular tripping grooves are installed immediately downstream of the nozzle to trigger boundary layer transition at the entrance to the test section. The flow and deformation measurements with the compliant wall are carried out 48 cm downstream of the tripping grooves. As demonstrated for a rigid wall (Wang et al. Reference Wang, Koley and Katz2020), these grooves generate a fully developed turbulent boundary layer at the sample area with the typical velocity profile, consisting of viscous, buffer, log and outer layers. The compliant wall modifies the velocity profiles, causing momentum loss that increases with increasing E/ρU 0 2 , as discussed in the introduction.

Figure 2. (a) The compliant coating and location of the sample volume, and (b,c) the optical set-up of the integrated TPTV-MZI system shown in (b) front view and (c) top view.

Figures 2(b) and 2(c) illustrate the optical set-up for simultaneous measurements of the time-resolved 3-D flow field using tomographic particle tracking velocimetry (TPTV), and 2-D wall deformation using MZI. This set-up replicates the approach introduced by Zhang et al. (Reference Zhang, Miorini and Katz2015), but the velocity measurements are based on tomographic particle tracking instead of tomographic PIV. The experiments have been conducted at three Re τ ranging from 3300 to 8900. The values of τ $_{w}$ and δ are determined from a fit to the mean velocity profile in the log layer based on TPTV data and separate stereo-PIV measurements (details follow) that cover the entire boundary layer, respectively. The experimental parameters, such as the sample volume size, field of view, resolution, frame rate and database size (all varying with the Reynolds numbers) are summarized in table 1. The corresponding flow parameters, including E/ρU 0 2 , deformation r.m.s. and critical heights, all taken from Lu et al. (Reference Lu, Xiang, Zaki and Katz2024), are also included. Since the optical set-up has been described before in great detail (Zhang et al. Reference Zhang, Miorini and Katz2015), it is only summarized here. The 10 mm thick light sheet is generated using a high-speed Nd:YLF laser (Photonics DM60–527), and the flow is seeded with 13 µm diameter silver-coated hollow glass spheres with a specific gravity of 1.6. The corresponding Stokes number is 4.6 × 10−2, hence these particles are expected to follow the flow (Raffel et al. Reference Raffel, Willert, Scarano, Kähler, Wereley and Kompenhans2018). The test section is located between two backside-polished mirrors (denoted as M1 and M2), which reflect 99.9 % of the light to illuminate the flow field, and the remaining 0.1 % is used for the MZI measurements. The tomographic particle tracking data are recorded by four high-speed cameras (PCO Dimax S4), denoted as Cam 1–4. As the (0.1 %) light transmitted through mirror M2 passes through the compliant wall, its phase is modulated by the surface deformation. The modulated beam is then combined with the reference beam transmitted through M1, and the resulting interference pattern is captured by a fifth high-speed camera (Phantom V2640), marked as Cam 5 in figure 2(b,c). Their optical path lengths are matched by adjusting the reference beam (mirrors M4–M7) until the fringes appear.

Table 1. The experimental conditions and scales of data acquisition.

Next, we summarize the data processing procedure for the combined TPTV-MZI measurements. In TPTV, the flow field is measured using the ‘shake-the-box’ Lagrangian particle tracking (Schanz, Gesemann & Schröder Reference Schanz, Gesemann and Schröder2016) available in the Davis 10.2 software. The calibration follows the typical two-step procedure involving coarse calibration using a moving target, followed by fine calibrations using the seeded flow field. For the present data, the mean tomographic disparity is approximately 0.05 pixel, with a spatial standard deviation of 0.14 pixel. Data analysis involves tomographic reconstruction, followed by particle tracking, which typically gives approximately 10 000 instantaneous particle tracks. The unstructured velocity and material acceleration are evaluated directly from the particle trajectories, and then interpolated onto a structured grid using constrained cost minimization procedure introduced and analysed by Agarwal et al. (Reference Agarwal, Ram, Wang, Lu and Katz2021) and implemented in Agarwal et al. (Reference Agarwal, Ram, Lu and Katz2023). This iterative method minimizes the difference between the interpolated values and the measured one, while constraining the differences between iterations. It also forces the interpolated velocity field to be divergence-free, and the acceleration field, curl-free, while accounting for viscous effects near the wall. The pressure distribution is calculated by spatially integrating the interpolated material acceleration using a 3-D, parallel line, omnidirectional integration method introduced in Wang, Zhang & Katz (Reference Wang, Zhang and Katz2019), and utilized in Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) and Agarwal et al. (Reference Agarwal, Ram, Lu and Katz2023). To account for viscous effect near the wall, we also include the viscous diffusion terms in the pressure integration as well. The integration procedure also detects paths with significant acceleration errors, based on the local curl, and reduces the weight of these paths. In Agarwal et al. (Reference Agarwal, Ram, Wang, Lu and Katz2021), the uncertainty of constrained cost minimization and pressure integration are evaluated using synthetic tracks reproduced from the DNS of turbulent channel flow at Re τ = 1000 obtained from Johns Hopkins Turbulence database (Graham et al. Reference Graham2016). It shows that the r.m.s. velocity and pressure errors are approximately 3 % and 10 % at y + > 80, respectively. The velocity error increases to 6 % at lower elevations while the pressure error remains at around 10 %. As discussed in Wang et al. (Reference Wang, Zhang and Katz2019) and Agarwal et al. (Reference Agarwal, Ram, Wang, Lu and Katz2021), the primary parameter affecting the uncertainty in pressure are the seed particle concentration (scaled by the wall unit), and image acquisition frequency. While the simultaneous flow and wall deformation measurement techniques utilize the same light source, the data acquisition and analysis procedures are totally independent, and do not bias each other (Zhang et al. Reference Zhang, Miorini and Katz2015, Reference Zhang, Wang, Blake and Katz2017). Further details about the uncertainty in pressure integration using omnidirectional techniques can be found in Liu & Katz (Reference Liu and Katz2006) and Wang et al. (Reference Wang, Zhang and Katz2019). To obtain the mean velocity profile from the TPTV data, the calculated particle tracks are projected onto a 2-D plane, and then analysed using a 2-D PIV-based sum-of-correlation method (Meinhart, Wereley & Santiago Reference Meinhart, Wereley and Santiago2000), using 4 × 4 pixels (118 × 118 μm2) interrogation windows. Due to the resolution limit of the TPTV data, the present measurements do not resolve the viscous sublayer and only part of the buffer layer. The first data point for the three Reynolds numbers are located at y + = 30, 59 and 79.

Figure 3. The PDFs of the compliant wall deformations at the indicated Reynolds numbers.

The MZI data analysis involves reconstruction of the instantaneous surface shape from the phase distribution of the fringes. Following Zhang et al. (Reference Zhang, Miorini and Katz2015) and using in-house developed software, the fringes are temporally normalized, and then enhanced using a correlation-based spatial filtering method that generates fringes with uniform amplitudes. The phase distributions are subsequently evaluated from the arc cosine of the enhanced fringes, followed by temporal and spatial unwrapping, as well as detrending. The resulting phase distribution is proportional to the fluctuations in the wall thickness, and to the difference between the refractive index of the NaI solution and that of the compliant material. Calibration tests described in Zhang et al. (Reference Zhang, Miorini and Katz2015) demonstrate that the deformation height resolution is of the order of 20 nm, ranging between 0.1 % to 0.7 % of the deformation amplitude. The planar spatial resolution of the present deformation measurement, which is defined by the fringe spacing, is 200 μm. Probability density functions (PDFs) of the compliant wall deformations normalized by their r.m.s. values (available in table 1) for the three Reynolds numbers are presented in figure 3. They demonstrate that in addition to the increase of d rms with increasing Reynolds number, the PDFs deviate further from a Gaussian distribution in the range of infrequent extreme events (d/d rms > 3). The comparison with the Gaussian distribution also highlights that the distributions are asymmetric, with more large negative deformations than positive ones. Table 1 also provides information on the skewness and kurtosis of the wall deformation, parameters that have been used for characterizing the effect of rigid roughness on boundary layers (e.g. Flack & Schultz Reference Flack and Schultz2014; Busse & Jelly Reference Busse and Jelly2023). Consistent with the PDFs in figure 3, the magnitude of wall skewness and the kurtosis increase with Reynolds number. However, they remain rather small, with the skewness varying between −0.07 and −0.19, and the kurtosis, between 3.5 and 4.1, i.e. they both do not deviate substantially from those of a Gaussian distribution. These variations are too small for determining the impact of high-order deformation statistics on the wall-deformation coupling.

Figure 4. Mean velocity profiles based on the present TPTV (reproduced from Lu et al. Reference Lu, Xiang, Zaki and Katz2024), stereo PIV and the 2-D PIV data of Wang et al. (Reference Wang, Koley and Katz2020). Dashed lines indicate the boundary layer heights, and dash–dotted lines, the critical heights.

Since the TPTV field of view (FOV) does not cover the entire boundary layer, to measure its thickness, we have also performed complementing stereo-PIV measurements with a sample area of 30 × 48 cm2 at the same location as the TPTV experiments. The images are recorded using two of the high-speed cameras (PCO Dimax S4), fitted with Nikon 105 mm lens, one at each side of the tunnel. The image size is 1100 × 1870 pixels, and the spatial resolution is 38 μm pixel−1. While the resolution of these measurements is of the same order as that of the TPTV data, the vertical extent of the FOV covers the entire boundary layers, allowing us to determine δ. The mean velocity profiles are obtained using 4 × 4 pixels sum-of-correlation available in the Davis 10.2 package, which determines the displacement from the ensemble-averaged cross-correlation. Figure 4 includes the mean velocity profiles from the stereo PIV measurements scaled with inner variables, showing that they resolve the outer part of the log layer and the outer layer. Determining δ from the point where U/U 0 = 0.99, and u τ from a fit to the velocity profile in the log layer, gives the values listed in table 1, along with the three Re τ of the present experiments, namely δ + = 3300, 6700 and 8900, as indicated also in figure 4. Table 1 also provides the momentum thickness, θ, and the corresponding $Re_{\theta }$ , showing that θ is around 1/10 of δ, and increases slightly with Reynolds number. Figure 4 also contains the mean streamwise velocity profiles obtained from the present TPTV measurements, and the 2-D PIV data from Wang et al. (Reference Wang, Koley and Katz2020), both reproduced from Lu et al. (Reference Lu, Xiang, Zaki and Katz2024). As is evident, with decreasing E/ρU 0 2, there is a growing deficit in the mean velocity in the log and buffer layers. The 2-D PIV data at E/ρU 0 2 = 2.4, which is obtained for the same compliant wall, and in the same test section, indicate that the velocity deficit extends to the viscous sublayer as well. Similar general behaviour, but with varying extents, are also reported in other experiments (Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022) and DNS (Rosti & Brandt Reference Rosti and Brandt2017; Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022).

One should note that the drag increase (i.e. the downward shift of mean velocity profiles in figure 4) rests on the determination of u τ based on a log-law fit. This method is widely used in experimental studies (e.g. Fernholz & Finley Reference Fernholz and Finley1996; De Graaff & Eaton Reference De Graaff and Eaton2000), and its limitation has also been discussed extensively (e.g. George & Castillo Reference George and Castillo1997; Wei, Schmidt & Mcmurtry Reference Wei, Schmidt and Mcmurtry2005; Kumar & Mahesh Reference Kumar and Mahesh2022). For example, the bias in log-law fit estimated from channel flow DNS data and pipe flow experimental data are less than 4.3 % and 8 %, respectively (Wei et al. Reference Wei, Schmidt and Mcmurtry2005). For a smooth rigid wall in the same facility as the present experiments, the high-resolution 2-D PIV measurements by Wang et al. (Reference Wang, Koley and Katz2020) demonstrate a good agreement with the values of u τ obtained from a log fit and velocity gradients in the viscous sublayer, with discrepancies decreasing from 5.4 % to 1.8 % with increasing Reynolds number. For a compliant wall, in addition to the velocity gradients, one has to account for the non-zero Reynolds shear stress and unsteady form drag (e.g. Rosti & Brandt Reference Rosti and Brandt2017; Wang et al. Reference Wang, Koley and Katz2020; Esteghamatian, Zaki & Katz Reference Esteghamatian, Katz and Zaki2022; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022). Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) compare the wall shear stress in compliant wall boundary layers estimated from a log fit with calculations based on force measurement on a large plate containing the compliant surface. They show a discrepancy of less than 4.7 % between the two methods, confirming that the log fit is a reliable approach when other techniques are not available.

Figure 5. An instantaneous sample snapshot at Re τ = 3300 of the velocity vectors and pressure contours in two planes, along with the wall deformations presented in exaggerated scales. The 3-D blobs are isosurfaces of ${\lambda _{2}}^{+}$ = −1.4 × 10−3.

Figure 5 shows a sample instantaneous snapshot of the integrated TPTV/MZI data at Re τ = 3300. The wavy surface at the bottom shows the instantaneous wall shape with its submicron amplitude variations exaggerated for clarity. Based on statistical analysis of the surface shape (Lu et al. Reference Lu, Xiang, Zaki and Katz2024), the dominant wavelength of the surface wave is 2.8l 0, close to the prediction of the Chase (Reference Chase1991) linear model of the surface response to harmonic excitation (3l 0). The vectors in the two selected (x,y) and (y,z) planes represent the velocity fluctuations (u , ${v}^{\prime}$ , ${w}^{\prime}$ ). The in-plane colour contours indicate the distribution of the pressure fluctuations, and the 3-D blobs representing vortical structures show isosurfaces of λ 2 + = −1.4 × 10−3 (Jeong & Hussain Reference Jeong and Hussain1995). In this snapshot, the deformation peak appears to be aligned with a pressure minimum, and the trough, with a maximum, consistent with results of conditional analysis described in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024). They also show that the pressure-deformation correlations peak near the critical heights whose elevations are listed in table 1, and indicated as dash–dotted lines in figure 4. Supplemental movies in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) demonstrate that the large-scale pressure features appear to be phase locked with the deformation and travel with it. The flow and pressure phenomena that appear to travel with the deformation extend to several hundred wall units despite the less than 1 μm (∼0.1δ ν ) deformation height. In the present sample, the pressure minimum above the surface bump appears to be associated with a vortical structure, also consistent with the results of conditional sampling.

Figure 6. Wall normal profiles of Reynolds stresses: (a) $\overline{u'u'}^{+}$ , (b) $\overline{v'v'}^{+}$ and (c) $-\overline{u'v'}^{+}$ non-dimensionalized using inner scaling and compared with results of DNS for rigid smooth walls (solid lines).

3. Results

3.1. Reynolds stresses

Figure 6 presents the wall normal profiles of Reynolds stresses, $\overline{u'u'}$ , $\overline{v'v'}$ and $-\overline{u'v'}$ , and compares them with DNS data obtained for smooth-wall rigid wall channel flow at Re τ = 1000 (Graham et al. Reference Graham2016) and Re τ = 5200 (Lee & Moser Reference Lee and Moser2015), both of them obtained from the Johns Hopkins Turbulence Database (Li et al. Reference Li, Perlman, Wan, Yang, Meneveau, Burns, Chen, Szalay and Eyink2008). Here, the prime indicates the fluctuation from mean value and the overbar refers to ensemble averaging. The present Reynolds stresses are calculated based on spatial ‘binning’ of the particle tracking data, using tools available in the Davis 10.2 software, with 8 × 8 × 8 voxel windows (0.24 × 0.24 × 0.24 mm3). The corresponding window size in terms of wall units are presented in table 1. This method first calculates the mean velocity using all the tracks in the interrogation volume, and then subtract the mean from each track, followed by ensemble averaging to obtain the Reynolds stresses. The contribution of each track is also weighted by its distance from the interrogation volume centre using a Gaussian weighting function. Consequently, the stresses are under-resolved, especially close to the wall. In the outer part of the boundary layer, the present distribution of $\overline{u'u'}^{+}$ and $\overline{v'v'}^{+}$ increase with Reynolds number in the outer layer, as expected, and mostly fall between the two DNS datasets. The profiles of both components collapse as the wall is approached, also consistent with the smooth wall data. However, the present increase in $\overline{u'u'}^{+}$ , and the decrease in $\overline{v'v'}^{+}$ with decreasing y + appear to be faster than those of the smooth wall. In figure 6(c), the three profiles of $-\overline{u'v'}^{+}$ peak in the log layer, but their magnitudes are considerably lower than those of the rigid wall DNS. As discussed in Wang et al. (Reference Wang, Koley and Katz2020) based on comparisons of 2-D PIV data obtained for rigid and complaint walls, the spatial extent of uu and ${v}^{\prime}{v}^{\prime}$ , two-point correlations over the compliant wall are substantially lower than those of the smooth wall. This finding implies that the interaction with the compliant material ‘scrambles’ the turbulence, reducing the correlation between the two velocity components, hence the Reynolds shear stress. Yet, figure 6(c) shows that the locations of shear stress peaks are consistent with the empirical relationship obtained for smooth wall boundary layers, y m + = 2.17 $\sqrt{\delta ^{+}}$ , by Morill-Winter, Philip & Klewicki (Reference Morrill-Winter, Philip and Klewicki2017). Interestingly, y m + is quite close to the critical height, y c + , at Re τ = 6700 and 8900, but is nearly twice of y c + at Re τ = 3300.

Figure 7 demonstrates that trends are different when the same results are plotted using mixed scaling, following Schultz & Flack (Reference Schultz and Flack2013), i.e. the Reynolds stresses are scaled with u τ 2, and y with δ. Replotting the data with y scaled with y c (not shown) is not significantly different since y c remains nearly a constant in our measurements (table 1, Lu et al. Reference Lu, Xiang, Zaki and Katz2024). To assess the effect of the limited spatial resolution, these plots also compare the results obtained for binning volumes of 8 × 8 × 8 voxels with those obtained using 6 × 6 × 6 voxels. As is evident, trends with elevations remain similar, and that increasing resolution increases $\overline{u^{\prime}u^{\prime}}^{+}$ and $\overline{v^{\prime}v^{\prime}}^{+}$ by 10 %– 20 % at low elevations and overlap in the outer layer, while $-\overline{u^{\prime}v^{\prime}}^{+}$ is affected by less than 4 %. Both normal stresses nearly collapse in the outer layer, and deviate below the log layer, where $\overline{u^{\prime}u^{\prime}}^{+}$ increases and $\overline{v^{\prime}v^{\prime}}^{+}$ decreases with decreasing Re τ . The mixed scaling seems to nearly collapse the $-\overline{u^{\prime}v^{\prime}}^{+}$ profiles (figure 7 c). These trends, but not the magnitudes, are consistent with those of smooth wall data (DeGraaf & Eaton Reference De Graaff and Eaton2000; Hultmark et al. Reference Hultmark, Vallikivi, Bailey and Smits2012; Schultz & Flack Reference Schultz and Flack2013).

Figure 7. Wall normal profiles of Reynolds stresses plotted using mixed scaling: (a) $\overline{u'u'}^{+}$ , (b) $\overline{v^{\prime}v^{\prime}}^{+}$ , (c) $-\overline{u^{\prime}v^{\prime}}^{+}$ . Circles, 8 × 8 × 8 voxels binning; lines, 6 × 6 × 6 voxel binning.

We have also tried to compare the present profiles with those presented in Wang et al. (Reference Wang, Koley and Katz2020) and Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) for compliant wall boundary layers. In both cases, results do not agree (not shown here). The Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) data corresponds to at least an order of magnitude lower E/ρU 0 2, but Re τ on the 5800–8000 range, and their spatial resolution is lower than the present levels. The Wang et al. (Reference Wang, Koley and Katz2020) stresses are calculated based on standard 2-D PIV measurements at a similar spatial resolution, speeds and wall properties. In mixed scale plots (not shown), trends of $\overline{u'u'}^{+}$ are similar to the present data, but the present magnitudes are slightly higher. The trends of $\overline{v'v'}^{+}$ and $-\overline{u'v'}^{+}$ with increasing Re τ do not agree, especially their observed sharp increase in stresses near the wall at high Re τ .

3.2. Decomposition of the wave-coherent motion and stochastic turbulence

The velocity and pressure fluctuations in a boundary layer over travelling waves originate either from the shear-driven turbulence or from wave-correlated motions. The wavenumber-frequency spectral analyses in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) have shown that turbulence is advected with the deformation wave below the critical layer, but follow the mean flow at higher elevations. Furthermore, at all three Reynolds numbers, the flow–deformation coherence peaks at the critical height. To understand the flow–deformation interactions, we would like to examine the characteristic turbulent flow phenomena that propagate with the deformation. Hence, the unsteady flow is decomposed into a wave-coherent part that presumably travels with the wave, and a ‘stochastic’ turbulent part that is not correlated with the wave. To this end, we adopt the ‘triple decomposition’ approach introduced by Hussain & Reynolds (Reference Hussain and Reynolds1970) and the Hilbert projection method introduced by Hristov et al. (Reference Hristov, Friehe and Miller1998), as described briefly here. Each flow variable $f$ is decomposed into its mean $\overline{\!f}$ (ensemble averaged value), and fluctuating/unsteady part, $f'$ . The latter is further decomposed into a wave-coherent, $\tilde{f}$ , and stochastic (incoherent), $f''$ , parts, i.e.

(3.1) \begin{equation} f=\overline{\!f}+f'=\overline{\!f}+\tilde{f}+f''. \end{equation}

Figure 8. (a) Power spectral densities of the wall deformation, and (b) sample time segments of the instantaneous wall deformation at the indicated times, z + = 0, and Re τ = 3300.

The Hilbert projection method used for determining $\tilde{f}$ , is particularly suitable for data that does not have a dominant periodic frequency or wavelength. A detailed summary of the procedure is provided in the Appendix. The reason that one cannot rely on simple phase averaging is discussed next. Figure 8(a) presents the power spectral densities of the wall deformation, E dd , at the present three Reynolds numbers. They are calculated using fast Fourier transform with a Hanning window available in MATLAB for every point on the surface, and then spatially averaged. The angular frequency ω in rad−1 is normalized by U 0 /l 0, and E dd by l 0 d rms 2 /U 0, where d rms is the r.m.s. of the wall deformation, presented in table 1. Clearly, the three deformation waves have broad ranges of frequencies, all peaking at a frequency corresponding to a wavelength of 3l 0 advected at 0.53U 0, the present wave speed. Results for the two higher Reynolds numbers nearly collapse, but the peak at the lowest Reynolds number is lower, and has a larger low frequency content. Furthermore, figure 8(b) provides sample instantaneous one-dimensional deformation waves at three instants, t 1, t 2 and t 3, that are separated by a time gap of 50ν/u τ 2, all at Re τ = 3300. As is evident, the wave amplitude varies spatially and in time, making phase averaging extremely challenging. In contrast, Hristov & Plancarte (Reference Hristov and Ruiz-Plancarte2014) and Hristov et al. (Reference Hristov, Friehe and Miller1998) show for field ocean waves that a Hilbert projection method can be used for estimating the wave-coherent component of flow variables in the atmospheric boundary layer. Further details about this procedure, including a demonstration that it is effective in extracting the wave-coherent part of a flow variable, are provided in the Appendix. The procedure involves the following steps. First, the wall deformation d(x,z,t) is divided into multiple narrowband signals, d k (x,z,t), each with a frequency bandwidth of Δω = 0.15U 0 /l 0. The division is performed in the Fourier domain, and each bandpass filtered signal is converted to the time domain. Second, the unsteady flow variable $f^{\prime}$ is projected onto each d k using

(3.2) \begin{equation} \tilde{f}=\sum _{k}\left\{\frac{\langle f',{d}_{k}\rangle }{\left\| {d}_{k}\right\| ^{2}}{d}_{k}+\frac{\langle f',\mathcal{D}_{k}\rangle }{\left\| \mathcal{D}_{k}\right\| ^{2}}\mathcal{D}_{k}\right\},\end{equation}

where $\mathcal{D}_{k}$ is the Hilbert transform of d k , $\langle$ · $\rangle$ indicates an inner product and ||·||2 is the squared norm. The Appendix shows that $\tilde{d}$ has a 99.9 % correlation with d, i.e. they are nearly identical, and that the maximum d– $\tilde{p}$ and d– $\tilde{v}$ correlations exceed 96 % and 75 %, respectively. The remaining incoherent part of the signal, $f''$ , is obtained by subtracting the $\tilde{f}$ from $f'$ .

To verify that the wave coherent and stochastic turbulence are orthogonal, we have calculated the vertical profiles of the cross term, $\overline{\tilde{f}f^{\prime\prime}}$ , and compared them with the corresponding coherent values for the streamwise and vertical velocity components as well as for the pressure. In all cases, the mean cross terms normalized by the corresponding coherent terms (not shown) remain below 10–7, indicating that the wave-coherent motion and the stochastic turbulence are essentially orthogonal. It appears that the Hilbert projection method is effective even for the present non-stationary wave condition. Considering the small magnitudes of cross terms, they are neglected in subsequent discussions.

The next discussion compares the two-point correlation between the wall deformation and the wave-coherent flow variables with that between the deformation and the stochastic turbulence. The conditional correlation is conditioned on (i) a deformation magnitude exceeding d rms , the local temporal r.m.s. deformation averaged over the entire area (table 1), and (ii) a surface peak (bump) or a minimum (dimple) at $\Delta$ x = 0. The $\tilde{f}$ –d correlation for a bump is defined as

(3.3) \begin{align} & \left.C_{\tilde{f}-d}\left({\unicode[Arial]{x0394}} x,y,{\unicode[Arial]{x0394}} z\right)\right| _{d\left(x_{0},z_{0},t\right)\gt {d_{\textit{rms}}}} \nonumber \\ & =\dfrac{\left.\langle \tilde{f}\left(x_{0}+{\unicode[Arial]{x0394}} x,y,z_{0}+{\unicode[Arial]{x0394}} z,t\right)d\left(x_{0},z_{0},t\right)\rangle \right| _{d\left(x_{0},z_{0},t\right)\gt {d_{\textit{rms}}}}}{\left.\left(\langle \tilde{f}\left(x_{0}+{\unicode[Arial]{x0394}} x,y,z_{0}+{\unicode[Arial]{x0394}} z,t\right)^{2}\rangle \left\langle d\left(x_{0},z_{0},t\right)^{2}\right\rangle \right)^{\frac{1}{2}}\right| _{d\left(x_{0},z_{0},t\right)\gt {d_{\textit{rms}}}}}.\end{align}

Figure 9. Conditional correlations of the deformation with the indicated flow variables based on (a) bumps, and (b) dimples, both at Re τ = 3300. In each set, subpanels (i), (iii), (v), (vii) and (ix) show the correlations with wave-coherent components ( $C_{\tilde{f}-d}$ ), subpanels (ii), (iv), (vi), (viii) and (x) the correlation with the ‘stochastic’ turbulence ( $C_{f''-d}$ ). The subpanels from (i) to (vi) show the distributions of $C_{u-d}$ , $C_{v-d}$ and $C_{p-d}$ . Subpanels (vii) and (viii) present the conditionally averaged wall shape, and (ix) and (x), the variations of peak magnitudes of conditional correlations with Reynolds number.

A corresponding expression is used for a dimple. For the stochastic signal, $\tilde{f}$ is replaced by $f''$ . The results for both a bump and a dimple are shown in figure 9(a,b). In each set, (i,iii,v,vii,ix) and (ii,iv,vi,viii,x) correspond to the $\tilde{f}$ d and $f^{\prime\prime}$ d correlations, respectively. Only results for Re τ = 3300 are presented here since those of the higher Reynolds numbers look very similar. The conditional wall shapes are plotted in (vii,viii), where the bump is bounded by dimples on both sides, and the dimple by bumps. The distance between dimples (or bumps), representing the characteristic wavelength of the deformation is 2.8l 0, very close to the predicted theoretical value based on a linear analysis of the compliant wall response to harmonic excitations. The critical height is marked by a yellow dashed line. The conditional correlations of the coherent velocity and pressure based on bumps and dimples are essentially identical (figure 9 a). Both are highly correlated with the deformation, with $C_{\tilde{u}-d}$ and $C_{\tilde{v}-d}$ having peaks of around ±0.7 on both sides of the bump/dimple. The domains of high correlation extend well above the critical layer, with the high $C_{\tilde{v}-d}$ peaks extending to further than those of $C_{\tilde{u}-d}$ . Near the wall, the transitions between negative and positive $C_{\tilde{u}-d}$ are offset slightly downstream of the bump. Away from the wall, at y + > 400, the patterns switch phase for reasons discussed in the next section. In contrast, the transitions between negative and positive $C_{\tilde{v}-d}$ occur above the bump, and there are no phase shifts in the correlations at higher elevations. As expected, the $\tilde{p}$ –d correlations are negative above the bump, with values lower than –0.9 extending deep into the log layer and remaining below –0.5 at y + = 900. The minimum correlations, –0.98, are measured at y + = 90, close to the critical height. The negative correlation peaks are bounded on both sides by positive peaks that also extend to y + > 800, with peak magnitudes of 0.6 aligned with the deformation troughs in the data conditioned on bumps, and bumps in the analysis conditioned on dimples. The implications and flow phenomena associated with these patterns are discussed in the next section. In contrast, the incoherent velocities and pressures are poorly correlated with the deformation, confirming that the $f^{\prime\prime}$ variables are incoherent with the deformation. The correlations with the incoherent velocity components (u′′ and $v^{\prime\prime}$ ), conditioned on bumps and dimples, have opposite signs with peak magnitudes remaining below 0.2. They indicate a broad sweeping flow above the bump, and a diminishing sweeping flow above the dimple, with $C_{v''-d}$ > 0 to the left of the dimple and nearly zero to the right of it. There is also non-zero but very low correlation with the pressure, indicating a weak minimum above the bump, and a weak maximum above the dimple. It should be noted, as discussion follows, that except for the pressure near the wall, for most of the flow field, the magnitudes of u′′ , $v^{\prime\prime}$ and p′′ are substantially higher than the corresponding coherent variables. Furthermore, as shown in subpanels (ix) and (x), the peak correlation magnitudes involving coherent variables, namely the highest values of $C_{\tilde{u}-d}$ , $C_{\tilde{v}-d}$ and $C_{\tilde{p}-d}$ , decrease with increasing Reynolds number. This decrease occurs in spite of the increase in the scaled deformation height (table 1). Finally, the maxima of $C_{u''-d}$ , $C_{v''-d}$ and $C_{p''-d}$ magnitudes, which are also presented in subpanels (ix) and (x), remain low and decrease with increasing Reynolds number.

In addition, while the magnitudes of the u′′–d correlations for a bump and a dimple are low, their inclination and elongation resemble those of u–u correlations in a typical rigid wall boundary layer (e.g. Ganapathisubramani et al. Reference Ganapathisubramani, Hutchins, Hambleton, Longmire and Marusic2005; Sillero, Jiménez & Moser Reference Sillero, Jiménez and Moser2014). This trend likely arises from the fact that conditional correlations are restricted to deformations exceeding the r.m.s. value. Such a restriction might cause a weak bias towards sweeping motions, where the near wall velocity is elevated, and the flow is subjected to adverse pressure gradients. Furthermore, the inclination angle of the u′′–d correlation is ∼15°, consistent with that of the u–u correlation in a solid wall boundary layer, a trend that has been associated with coherent structures (Ganapathisubramani et al. Reference Ganapathisubramani, Hutchins, Hambleton, Longmire and Marusic2005; Adrian Reference Adrian2007; Jiménez Reference Jiménez2018). This trend suggests that high wall deformations might be weakly correlated with ‘naturally occurring’ coherent structures in the boundary layer.

Figure 10 presents the distributions of phase-averaged coherent flow field, $\widehat{\tilde{f}}$ , as a function of deformation phase, at the three Reynolds numbers. The phase averaging is performed by dividing a wave cycle into 20 phase bins, i.e. each with a width of π/10, and averaging all the data (> 8.8 × 105 samples) inside each bin. The wall-normal axis is scaled both as y/y c and y/δ, with the results for Re τ = 3300, 6700 and 8900 plotted in figures 10(a,b), 10(c,d) and 10(e,f), respectively. The phase-averaged wall deformation is presented in figure 10(g) and repeated three times for convenience. Note that the phase-averaged dimples are slightly deeper than the bumps, consistent with the PDFs of deformation in figure 3, e.g. at Re τ = 8900, d max ≈ −0.93d min . For all the three cases, when the vertical axis is scaled with outer variables, they display similar distributions of flow quantities penetrating deep into the log layer in spite of substantial differences in the heights of the deformation. The velocity contours have alternate signs, which are not in phase with the windward (–π < φ< 0) or leeward (0 < φ <π) sides. The $\widehat{\tilde{u}}$ contours (figure 10 a,c,e) are inclined upstream at y/δ≤ 0.15, and shift significantly at y/δ ∼ 0.19. Near the wall, the velocity vectors indicate Q2 events (ejections, $\widehat{\tilde{u}}$ < 0, $\widehat{\tilde{v}}$ > 0) generally on the windward side and near the crest, and Q4 events (sweeps, $\widehat{\tilde{u}}$ > 0, $\widehat{\tilde{v}}$ < 0) on the leeward side and the trough. The distribution of $\widehat{\tilde{v}}$ appear to be periodic at all elevation, with positive values above the windward side and negative on the leeward side, without the phase shift observed in the horizontal velocity.

Figure 10. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,c,e) $\widehat{\tilde{u}}$ +, (b,d,f) $\widehat{\tilde{v}}$ + and (g) $\hat{d}/\delta$ with deformation phase; (a,b) Re τ = 3300, (c,d) Re τ = 6700 and (e,f) Re τ = 8900. The arrows in (a,c,e) show the velocity vectors, and the yellow dashed lines indicate the critical heights. Panels (h) and (i) display the variations of peak (h) $\widehat{\tilde{u}}$ + and (i) $\widehat{\tilde{v}}$ + with Reynolds number.

Due to the Q2 and Q4 events, the horizontal velocity maxima are located slightly above the critical layer, whereas the coherent vertical velocity maxima extend to around three times the critical height. Overall, the coherent velocity peaks decrease with increasing Reynolds numbers, in contrast with the increase in deformation height. Further understanding of the flow structure can be obtained by examining the corresponding distributions of $\widehat{\widetilde{\omega _{z}}}$ and $\widehat{\tilde{p}}$ presented in figure 11. Like the velocity, the vorticity and pressure magnitudes decrease with increasing Reynolds number. Below the critical height, $\widehat{\widetilde{\omega _{z}}}$ (figure 11 a,c,e) is negative on the leeward side and positive on the windward side. This pattern shifts at higher elevations, where the $\widehat{\widetilde{\omega _{z}}}$ < 0 region is aligned with the top of a bump, and $\widehat{\widetilde{\omega _{z}}}$ > 0 with the trough. At y > y c , the transitions between positive and negative areas remain largely vertical. The negative regions in the distributions of pressure (figure 11 b,d,f) appear to be aligned with the wave crests, with the peaks around π/10 ahead of the summits. The positive peaks are aligned with the troughs, with their maxima also located approximately π/10 ahead of the valleys. The ∼π/10 phase shift is consistent with the trends of p–d correlations reported in Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) for an order of magnitude stiffer compliant wall, as inferred from the offset of the peak correlation and the deformation wavelength. They attribute this delay primarily to the flow structure in the boundary layer, and to a lesser extent, to the material response. Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022) also observe a similar phase lag, but in their case, the delay decreases with the deformation magnitude, e.g. from ∼π/4 at d rms + = 5.6 to ∼π/8 at d rms + = 0.55. Finally, figures 10(h,i) and 11(h,i) show that for all the flow variables presented in figures 10 and 11, the magnitudes of the coherent parts decrease with increasing Reynolds number in spite of the significant increase in d rms + . A plausible cause for these contradicting trends is discussed later in the paper.

Figure 11. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,c,e) $\widehat{\widetilde{\omega _{z}}}$ +, (b,d,f) $\widehat{\tilde{p}}$ +,and (g) $\hat{d}/\delta$ with deformation phase; (a,b) Re τ = 3300, (c,d) Re τ = 6700 and (e,f) Re τ = 8900. The yellow dashed lines indicate the critical heights. Panels (h) and (i) display the variations of peak (h) $\widehat{\widetilde{\omega _{z}}}$ + and (i) $\widehat{\tilde{p}}$ + with Reynolds number.

The locations of the negative and positive vorticity peaks above the crest and trough, respectively, are consistent with results of conditional averaging of the flow structures above bumps and dimples, without decomposition to coherent and stochastic turbulence, as mentioned briefly (but not shown) in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024). Here, the conditional averaging also uses d > d rms for a surface bump and d < d rms for a dimple. The results of this conditional sampling, which is summarized in figure 12, show that bumps preferentially form under a negative spanwise vortex, owing to the pressure minimum that this vortex generates (figure 12 a,c). A sweeping flow induced downstream of this vortex impinges on the surface, generating a pressure maximum and a dimple at the sweep to ejection transition (figure 12 b,d). The present coherent part of the flow contains a negative vorticity peak above the bump, where the pressure is minimum, and a sweeping-ejection transition above the dimple, where the pressure is maximum. In attempts to explain the curious shift in the vorticity distribution below the critical height, one possibility is generation of counter rotating vorticity as the vortical structure above interacts with the boundary, generating shear in the opposite direction. Another possibility might involve viscous vorticity flux from the wall associated either with the pressure gradients (Lighthill Reference Lighthill and Rosenhead1963) or the surface-parallel material acceleration (Morton Reference Morton1984). These two contributors are compared based on DNS data by Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022). Results show these two contributors have opposite effects, with a net impact that varies with the compliant material stiffness and amplitude of the deformation. Unfortunately, the present data does not have sufficient spatial resolution to determine the near wall vorticity or the surface acceleration, so we cannot comment on the viscous vorticity diffusion. Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022) also show for relatively large deformation (d rms + = 5.6), thickening of the boundary layer and flow separation close to the wave crest inject negative vorticity to the flow above the trough. However, the present deformations are significantly smaller, and the vorticity distribution shift occurs also for d rms + << 1 (figure 11 a,c,e). Hence flow separation is unlikely to be a significant contributor. This discussion suggests that the first option, namely formation of counter-rotating vorticity as the flow induced by a vortex interacts with the wall, seems to be the most viable option.

Figure 12. Conditionally averaged flow variables and deformation at Δz + = 0 for (a,c) a surface bump, and (b,d) a dimple, both at Re τ = 3300: (a,b) pressure contours and in-plane velocity vectors, and (c,d) compliant wall shape.

Before proceeding, it should be noted that the present distributions of $\widehat{\tilde{u}}$ and $\widehat{\tilde{v}}$ are not consistent with results obtained for wind-wave interactions (e.g. Buckley & Veron Reference Buckley and Veron2019; Cao & Shen Reference Cao and Shen2021; Do, Wang & Chang Reference Do, Wang and Chang2024), and the simulated flow over a very soft compliant material (Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022). Discrepancies exist also among results obtained for ocean waves owing to differences in wave amplitude, slope and ‘age’ (U $_{\textit{s}w}$ /u τ ). There is better consistency among the distributions of pressure, where the minima are persistently centred in the vicinity of the wave crest, and the maxima around the trough. The differences from Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022) might be related to the more than an order of magnitude higher wave amplitude owing to the much softer material in their simulations. Consequently, their wave crest is more prone to flow separation. Another possibility is related to the relationship between wave speed and flow velocity, which might affect the compliant wall–flow interactions. In the simulations, the magnitude of U $_{\textit{s}w}$ is close to the shear wave speed, corresponding to the advection speed of Rayleigh waves in elastic material (∼0.95C t , Freund Reference Freund1998). In contrast, in the present experiments as well as the data presented in Carpenter et al. (Reference Carpenter, Davies and Lucey2000), Zhang et al. (Reference Zhang, Wang, Blake and Katz2017), Wang et al. (Reference Wang, Koley and Katz2020), Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) and Lu et al. (Reference Lu, Xiang, Zaki and Katz2024), U $_{\textit{s}w}$ varies between 40 %–80 % of U 0 and does not scale with C t .

Questions remain whether it is possible to relate the unsteady wave-coherent motion to the flow induced by a steady wall roughness. Nakato et al. (Reference Nakato, Onogi, Himeno, Tanaka and Suzuki1985) suggest that a rough wall boundary layer should be considered as flow over a wavy surface when the roughness slope is less than 6°. In this case, the momentum deficit in the log layer is strongly affected by the slope of the roughness element, increasing from ∼1 to ∼9.5 when the slope increases from 1° to 6° (e.g. Napoli, Armenio & Marchis Reference Napoli, Armenio and De Marchis2008; Schultz & Flack Reference Schultz and Flack2009). In the present study, the slope changes, 0.01° to 0.07°, are two orders of magnitude smaller, yet the downward shifts in velocity profiles increase from 1.1 to 2.6, i.e. it is of the same order as that of the rough wall. However, in the present measurements, the deformation amplitude also increases with the Reynolds number, which is also expected to affect the momentum deficit. Furthermore, the flow over stationary wavy surfaces with low slopes has also been modelled theoretically by Hunt, Leibovich & Richards (Reference Hunt, Leibovich and Richards1988). This model assumes that the wave amplitude and the roughness length are much smaller than wavelength, conditions that are satisfied in the present experiments. They divide the flow field into an inner region, where surface-induced shear stress is significant, and an outer region, where the flow perturbations are inviscid. The present measurements do not resolve the inner region. In the outer region, the present coherent vertical velocity component (figure 10 b,d,f) peaks at the elevation as the theoretical prediction. However, the present coherent horizontal velocity peaks in the outer layer, in contrast to the model prediction that places this peak in the inner layer. Therefore, some of the trends of the compliant wall boundary layer appear to be consistent with those of the flow over a stationary surface undulation, while others do not.

Figure 13. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,d,g) $\widehat{\tilde{u}\tilde{u}}$ +, (b,e,h) $\widehat{\tilde{v}\tilde{v}}$ +, (c,f,i) $-\widehat{\tilde{u}\tilde{v}}$ + and (j) $\hat{d}/\delta$ with deformation phase: (a–c) Re τ = 3300, (d–f) Re τ = 6700 and (g–i) Re τ = 8900. Yellow dashed lines indicate the critical heights. Panels (k–m) display the variations of peak (k) $\widehat{\tilde{u}\tilde{u}}$ +, (l) $\widehat{\tilde{v}\tilde{v}}$ + and (m) $-\widehat{\tilde{u}\tilde{v}}$ + with the Reynolds number.

The next discussion compares the magnitudes of $\widetilde{u_{{i}}}\widetilde{u_{{j}}}$ with those of $u_{i}^{\prime\prime}u_{j}^{\prime\prime}$ , and $\tilde{p}_{\textit{rms}}$ with $p^{\prime\prime} _{\textit{rms}}$ . As figure 13 demonstrates, all the coherent stresses are concentrated near the wall, decaying rapidly with elevation. The values of $\widehat{\tilde{u}\tilde{u}}$ peak at or just above the critical layer, and those of the other stresses, at higher elevations, all consistent with the peak locations of their mean values. All the coherent stresses have maxima on the windward and leeward sides, which are asymmetric both in phase and in magnitude. The phase lags are consistent with the delay between the coherent velocity and the deformation, as depicted in figure 10. All the stresses are stronger on the leeward side, and decrease with increasing Reynolds number, trends that are evident in figure 13(k–m). The difference between the windward and leeward values appears to be associated with infrequent, high amplitude sweep-ejection events above the troughs. The formation of stress maxima on the windward and leeward sides is consistent with the DNS data in Esteghamatian et al. (Reference Esteghamatian, Katz and Zaki2022), with some phase differences. The simulations also show strong negative shear stress peaks between the positive maxima very near the wall (y + ≤ 4), which the experimental data cannot resolve.

Figure 14. Variations of the stochastic, spatially and temporally phase-averaged: (a,d,g) $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ +, (b,e,h) $\widehat{v^{\prime\prime}v^{\prime\prime}}$ +, (c,f,i) $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ + and (j) $\hat{d}/\delta$ with deformation phase; (a–c) Re τ = 3300, (d–f) Re τ = 6700 and (g–i) Re τ = 8900. Panels (k–m) display the variations of peak (k) $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ +, (l) $\widehat{v^{\prime\prime}v^{\prime\prime}}$ + and (m) $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ + with the Reynolds number.

Figure 14 presents the phase-averaged distributions of $u_{i}^{\prime\prime}u_{j}^{\prime\prime}$ for the three Reynolds numbers. As is evident, the magnitudes of the incoherent stresses are higher than the coherent ones by nearly an order of magnitude everywhere, indicating that most of the unsteady motion is incoherent, and that the trends with elevation are consistent with those of the total Reynolds stresses. In particular the maximum in $-\widehat{\textit{u}^{\prime\prime}{v}^{\prime\prime}}$ at the two higher Reynolds number occurs at the critical height. While the variations of $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ and $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ with deformation phase are milder than those of the coherent motions, the ‘stochastic’ stresses are still not distributed uniformly. Both $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ and $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ peak on the downwind sides of the bump, with the maxima in $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ occurring upstream of those of $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ . Their variations with phase seem to decrease with increasing Reynold number. Since the leeward side experience an adverse pressure gradient, the turbulence level is expected to increase. Interestingly, the peaks in $\widehat{\tilde{u}\tilde{u}}$ do not occur in the same phase as those of $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$ , but the maxima in - $\widehat{\tilde{u}\tilde{v}}$ and $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$ do coincide. The distributions of $\widehat{v^{\prime\prime}v^{\prime\prime}}$ are more uniform, and do not display consistent or clear trends with Reynolds number. As discussed later, these phase variations will affect the energy exchange between the coherent and incoherent parts of the kinetic energy.

Figure 15. Variations of the spatially and temporally phase-averaged: (a,c,e) $\widehat{\tilde{p}}_{\textit{rms}}$ +, (b,d,f) $\widehat{p''}_{\textit{rms}}$ + and (g) $\hat{d}/\delta$ with deformation phase: (a,b) Re τ = 3300, (c,d) Re τ = 6700 and (e,f) Re τ = 8900. Panels (h,i) display the variations of peak (h) $\tilde{p}_{\textit{rms}}$ +, and (i) $p''_{\textit{rms}}$ + with Reynolds number.

The phase-averaged distributions of coherent and stochastic r.m.s. pressure are compared in figure 15. The coherent r.m.s. pressure peaks are concentrated near the wall, peaking slightly beyond the crest and the trough of the deformation (figure 15 a,c,e), and their magnitudes decrease with increasing Reynolds number (figure 15 h). The stochastic r.m.s. pressure also peaks at the wall but does not vary significantly with phase (figure 15 b,d,f). In contrast to the coherent values, the stochastic r.m.s. pressure increases with Reynolds number (figure 15 i). Owing to these opposite trends, while the coherent and stochastic maxima have similar magnitudes at Re τ = 3300, the stochastic r.m.s. is more than twice higher than that at Re τ = 8900.

3.3. Kinetic energy budgets for wave-coherent flow and stochastic turbulence

Following the established framework, the total turbulent kinetic energy can also be separated into a coherent ‘wave kinetic energy’ (WKE), and an incoherent ‘stochastic kinetic energy’ (SKE),

(3.4) \begin{equation} \begin{array}{c} \frac{1}{2}\overline{u_{{i}}'u_{{i}}'}=\frac{1}{2}\overline{\widetilde{u_{{i}}}\widetilde{u_{{i}}}}+\frac{1}{2}\overline{u_{{i}}^{\prime\prime}u_{{i}}^{\prime\prime}}.\end{array} \end{equation}

As verified before, the cross terms are negligible since $\widetilde{u_{{i}}}$ and $u_{{i}}^{\prime\prime}$ are uncorrelated. Following Reynolds & Hussain (Reference Reynolds and Hussain1972), the turbulent kinetic energy budget equation is also decomposed. The WKE budget is given by

(3.5) \begin{equation} \begin{array}{c} 0=\left(-\overline{\widetilde{u_{{i}}}\widetilde{u_{{j}}}}\right)\partial _{j}\overline{u_{{i}}}+\overline{\left(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}}\right)\partial _{{j}}\widetilde{u_{{i}}}}-\partial _{j}\left(\overline{\widetilde{u_{{j}}}\tilde{p}}\right)-\overline{u_{{j}}}\partial _{j}\left(0.5\overline{\widetilde{u_{{i}}}\widetilde{u_{{i}}}}\right)-\partial _{j}\left[\overline{\widetilde{u_{{j}}}\left(0.5\widetilde{u_{{i}}}\widetilde{u_{{i}}}\right)}\right]\\\quad -\partial _{j}\left[\overline{\widetilde{u_{{i}}} \left(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}}\right)}\right] +\nu \partial _{j}\left[\overline{\widetilde{u_{{i}}}\left(\partial _{{j}}\widetilde{u_{{i}}}+\partial _{{i}}\widetilde{u_{{j}}}\right)}\right]-2\nu \overline{\left(\partial _{{j}}\widetilde{u_{{i}}}+\partial _{{i}}\widetilde{u_{{j}}}\right)\left(\partial _{{j}}\widetilde{u_{{i}}}+\partial _{{i}}\widetilde{u_{{j}}}\right)}, \end{array} \end{equation}

where $(-\overline{\widetilde{u_{{i}}}\widetilde{u_{{j}}}})\partial _{j}\overline{u_{{i}}}$ is the WKE production by mean flow; $\overline{(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}})\partial _{{j}}\widetilde{u_{{i}}}}$ is the WKE production by the stochastic turbulence; $-\partial _{j}(\overline{\widetilde{u_{{j}}}\tilde{p}})$ is the coherent pressure diffusion; $-\overline{u_{{j}}}\partial _{j}(0.5\overline{\widetilde{u_{{i}}}\widetilde{u_{{i}}}})$ is the advection by mean flow; and the fifth to seventh terms represent the transport of WKE by the wave-coherent stresses, stochastic turbulent stresses and viscous stresses, respectively. The last term is the viscous dissipation. Accordingly, the budget equation for the SKE is

(3.6) \begin{equation} \begin{array}{c} 0=\left(\!-\overline{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}}\right)\partial _{j}\overline{u_{{i}}}-\overline{\left(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}}\right)\partial _{{j}}\widetilde{u_{{i}}}}-\partial _{j}\left(\overline{u_{{j}}^{\prime\prime}\textit{p}^{\prime\prime}}\right)-\overline{u_{{j}}}\partial _{j}\left(0.5\overline{u_{{i}}^{\prime\prime}u_{{i}}^{\prime\prime}}\right)-\partial _{j}\!\left[\overline{u_{{j}}^{\prime\prime}\left(0.5u_{{i}}^{\prime\prime}u_{{i}}^{\prime\prime}\right)}\right]\\ -\overline{\widetilde{u_{{j}}}\partial _{{j}}\left(0.5\widetilde{u_{{i}}^{\prime\prime}u_{{i}}^{\prime\prime}}\right)}+\nu \partial _{j}\left[\overline{u_{{i}}^{\prime\prime}\left(\partial _{{j}}u_{{i}}^{\prime\prime}+\partial _{{i}}u_{{j}}^{\prime\prime}\right)}\right]-2\nu \overline{\left(\partial _{{j}}u_{{i}}^{\prime\prime}+\partial _{{i}}u_{{j}}^{\prime\prime}\right)\left(\partial _{{j}}u_{{i}}^{\prime\prime}+\partial _{{i}}u_{{j}}^{\prime\prime}\right)},\end{array} \end{equation}

where $(-\overline{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}})\partial _{j}\overline{u_{{i}}}$ and $-\overline{(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}})\partial _{{j}}\widetilde{u_{{i}}}}$ are production of SKE by the mean flow and the wave-coherent motions, respectively. The other terms follow the same order as those associated with WKE. One term, $\overline{(\widetilde{u_{{i}}^{\prime\prime}u_{{j}}^{\prime\prime}})\partial _{{j}}\widetilde{u_{{i}}}}$ , appears in both equations with the opposite sign, representing transfer of energy from the stochastic turbulence to the wave-coherent motions when positive.

The profiles of WKE and SKE budget terms at Re τ = 3300 are compared in figure 16. As is evident, near the wall, the production by mean flow is larger than the rest of the terms by more than an order of magnitude, with shear productions, namely $(-\overline{\tilde{u}\tilde{v}})\partial \overline{u}/\partial y$ and $(-\overline{u''v''})\partial \overline{u}/\partial y$ , being the dominant contributors. Near the wall, the WKE production is only 6 %–10 % of the SKE counterpart, decreasing to 3 % in the log layer. Figure 17(a,b) shows the variations of shear production terms with Reynolds number. When plotted using inner variables, the profiles of both production terms collapse far away from the interface, but they deviate near the wall, consistent with the reduction of the stochastic and coherent shear stresses with increasing Reynolds number. Trends do not collapse when plotted using mixed variables either (insert in figure 17 a). While the stresses nearly collapse, the mean velocity gradients do not. Both production terms still increase with decreasing elevation at the lowest point that can be resolved by the present 3-D measurements, implying that both terms peak below the critical height. For the resolved range, the profiles of SKE production (figure 17 b) are consistent with those obtained in DNS of a smooth rigid wall channel flow at Re τ = 5200 (Lee & Moser Reference Lee and Moser2015), especially in the outer layer. However, compliant wall boundary layer DNS (Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022) shows that the Reynolds stress becomes negative at y + < 4, implying that reverse energy transfer, from the wave to the mean flow, occurs very near the wall. Finally, figures 17(c) and 17(d) compare the variation of shear production terms with wave phase. In accordance with the distributions of shear stresses, the WKE production peaks are located slightly upstream of the crest and the trough. The variation in SKE production with phase is milder, with higher values on the leeward side, in the same area as the phase-averaged transition from sweep to ejection.

Figure 16. Wall-normal profiles of the ensemble-averaged kinetic energy budget terms for: (a) WKE and (b) SKE, both at Re τ = 3300.

Figure 17. Shear production rates: (a,b) ensemble-averaged profiles of (a) WKE, and (b) SKE at the indicated Reynolds numbers; (c,d) distributions of (c) $(-\widehat{\tilde{u}\tilde{v}}\partial \overline{u}/\partial y)^{+}$ and (d) $(-\widehat{u''v''}\partial \overline{u}/\partial y)^{+}$ at Re τ = 3300.

Figure 18. The axial contributor to wave–turbulence energy exchange: (a,b,c) temporally and spatially phase-averaged distributions at (a) Re τ = 3300, (b) Re τ = 6700 and (c) Re τ = 8900. (d) Variations of the peak values with Reynolds number.

The next discussion shifts to the energy exchange between WKE and SKEs. Consistent with trends of wind–wave interaction (Zhang, Wang & Liu Reference Zhang, Wang and Liu2024), the axial extension/contraction term, i.e. $\overline{(\widetilde{u''u''})\partial \tilde{u}/\partial x}$ , is the dominant contributor. Plots of the phase-averaged axial contraction (figure 18 a–c) indicate that this term is negative on the windward side, i.e. energy is transferred from WKE to SKE owing to the wave-induced streamwise contraction, and positive on the leeward side, owing to axial extension. The net energy flux is therefore the difference between the contributions of contraction and extension. At all three Reynolds numbers, the axial contraction and extension peaks are aligned along the critical height but decrease in magnitude with increasing Re τ (figure 18 d). Profiles of the net energy fluxes (figure 19) are presented scaled using inner variables (figure 19 a), and mixed variables, u τ 3/λ (figure 19 b), where the wall unit is replaced by the surface wavelength (λ = 3l 0). Both have been used for normalizing data in studies of wind–wave interaction (e.g. Yousefi et al. Reference Yousefi, Veron and Buckley2021; Zhang et al. Reference Zhang, Wang and Liu2024). With either scaling, the magnitudes of energy flux decrease with increasing Reynolds number, but the differences between them are significantly smaller under mixed scaling. Finally, the energy fluxes at all the three Reynolds numbers change sign across the critical layer. At y> y c , kinetic energy is transferred from the stochastic turbulence to the wave-coherent field. Conversely, at y< y c , kinetic energy flows from the coherent to the stochastic turbulence. A sign change across the critical layer is also observed in the coherent pressure diffusion, in which the wall-normal gradient of the $\tilde{p}$ $\tilde{v}$ correlation is the main contributor. Profiles of the ensemble averaged values (figure 20 a,b) indicate that mixed scaling leads to better collapse of results than the inner variables. In contrast to the wave–turbulence exchange, pressure diffusion adds energy to the wave-coherent field at y< y c and depletes it at y> y c . A comparison with the profiles of WKE, shown in figure 21, indicates that the peak negative pressure diffusion is located at nearly the same elevation as the maximum WKE, and the sign change indicates transport towards the wall, across the critical layer.

Figure 19. Wall-normal profiles of the ensemble-averaged axial wave–turbulence energy exchange term for the three Reynolds numbers, normalized using (a) inner and (b) mixed parameters.

Figure 20. Wall-normal profiles of the ensemble-averaged coherent pressure diffusion term for the three Reynolds numbers, normalized using (a) inner and (b) mixed parameters.

Figure 21. Wall-normal profiles of the WKE at the indicated Reynolds numbers.

4. Discussion

In the previous sections we have noticed that in contrast to the substantial increase in surface wave amplitude with increasing Reynolds number, the coherent phase-averaged velocity, vorticity and pressure, as well as the coherent stresses, r.m.s. pressure, kinetic energy, WKE production rate and WKE–SKE exchange decrease. Furthermore, both the p–d correlation and coherence, presented in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024) and reproduced as figure 22 decay with increasing Reynolds number at all elevations. This section attempts to discuss possible reasons for these apparently contradicting trends, namely that all the parameters associated with flow–deformation interaction are strongest while the deformation height is the smallest at the lowest Reynolds number. Furthermore, figure 22(a) also shows the p–d correlation of the one-way coupling case in Zhang et al. (Reference Zhang, Wang, Blake and Katz2017), which as noted before, involves a thicker and stiffer compliant material (l 0 = 16 mm, E = 930 Kpa). Clearly, the latter is considerably lower than those of the present experiments, despite the agreement in the magnitude of d rms + (∼0.03), and the similarity in Re τ to the present slowest case. The analysis aimed at explaining these disagreements is based on a comparison between the wavelength of energy containing eddies in the boundary layer and the characteristic wavelength of the deformation.

Figure 22. Wall-normal profiles of the (a) p–d correlation conditioned on a bump for the present and the Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) data, and (b) p–d coherence for the present data, both at the indicated Reynolds numbers. The present results are reproduced from Lu et al. (Reference Lu, Xiang, Zaki and Katz2024).

In the compliant wall turbulent boundary layer, the interaction between flow and deformation is not restricted to a single wavelength. Some of the theoretical approaches examine the deformation as surface responses to harmonic flow excitations at different frequencies and wavenumbers (e.g. Chase Reference Chase1991; Benschop et al. Reference Benschop, Greidanus, Delfos, Westerweel and Breugem2019). Therefore, it is necessary to compare the length scales of the turbulent structures in the boundary layer with the wavelength and frequency dependent responses of the compliant surface deformation. The former has been studied extensively in smooth-wall boundary layers, both experimentally (e.g. Mathis, Hutchins & Marusic Reference Mathis, Hutchins and Marusic2009) and numerically (e.g. Jimenez et al. Reference Jiménez, Hoyas, Simens and Mizuno2010). The boundary layer scales, denoted as λ x in the following discussion, vary from the viscous length scales to several times the boundary layer thickness. The energetic coherent structures, characterized by the spectral domain with a slope of $-1$ in the streamwise velocity spectra (Perry & Chong Reference Perry and Chong1982; Nickels et al. Reference Nickels, Marusic, Hafez and Chong2005; Calaf et al. Reference Calaf, Hultmark, Oldroyd, Simeonov and Parlange2013), are consistent with the so-called ‘attached eddies’ introduced by Townsend (Reference Townsend1976). As demonstrated by Mathis et al. (Reference Mathis, Hutchins and Marusic2009), these eddies have a dominant peak in the buffer layer at λ x + ≈1000 in the premultiplied streamwise velocity spectra, k x E uu . The peak response of the compliant wall to harmonic excitation, based on a solution to the Helmholtz equation (Chase Reference Chase1991; Zhang et al. Reference Zhang, Wang, Blake and Katz2017; Benschop et al. Reference Benschop, Greidanus, Delfos, Westerweel and Breugem2019; Wang et al. Reference Wang, Koley and Katz2020), for a broad range of frequencies, is approximately three times the compliant layer thickness (3l 0), as long as the flow velocity and wave speed are lower than C t , as demonstrated in figure 23(a). This preferred wavelength has been confirmed experimentally for several material properties and wall thicknesses by Zhang et al. (Reference Zhang, Wang, Blake and Katz2017), Wang et al. (Reference Wang, Koley and Katz2020), Greidanus et al. (Reference Greidanus, Delfos, Picken and Westerweel2022) and Lu et al. (Reference Lu, Xiang, Zaki and Katz2024). It should be noted that since the wave propagates at a speed scaled with U 0, with coefficients varying between 0.4–0.8 (Carpenter et al. Reference Carpenter, Davies and Lucey2000), each frequency involves a different wavelength. Focusing on the frequency of peak material response, figure 23(b) provides sample experimental wavenumber spectra extracted from the Wang et al. (Reference Wang, Koley and Katz2020) results demonstrating that the peak wavelength is 3l 0.

Figure 23. Wavelength of the compliant surface response at the indicated frequencies, based on (a) a solution to the Chase (Reference Chase1991) model, and (b) the large field of view experimental data of Wang et al. (Reference Wang, Koley and Katz2020). The solid grey lines mark the location of λ = 3l 0.

Figure 24. Premultiplied energy spectra, $k_{x}{E_{u'u'}}^{+}$ , for the full turbulence at (a) Re τ = 3300, (b) Re τ = 6700 and (c) Re τ = 8900. The horizontal dashed lines correspond to λ = 3l 0, the characteristic wavelength of the compliant wall deformation.

Figure 24 compares the energy spectra of the streamwise velocity fluctuations at the three Reynolds numbers premultiplied by k x to emphasize the k −1 region, which represent the energetic attached eddies. Since the wavenumber range of the TPTV field of view is limited, these spectra are calculated in the frequency domain and converted to wavelength using Taylor’s frozen turbulence hypothesis. The peaks in all the premultiplied energy spectra are centred around λ x + ≈ 1000, consistent with the findings in typical boundary layers over a wide range of Reynolds numbers (Mathis et al. Reference Mathis, Hutchins and Marusic2009). However, the present data does not show their secondary peak at the wavelength of λ x ≈ 10δ for large-scale motions, possibly since the present recording time is only approximately 130δ/U 0. The black dashed lines mark the preferred deformation wavelength, 3l 0. As is evident, at Re τ = 3300, the dominant deformation wavelength is well aligned with the peak of the premultiplied energy spectrum, indicating that the wall waves excite the flow at the preferred wavelength of the attached eddies. In contrast, the preferred surface wavelength falls at the edge of the k 1 range at the two higher Reynolds numbers, suggesting that it would cause weaker response and lower correlations with the flow. These observations are consistent with the decrease in all the coherent flow and stress parameters, as well as the decrease in flow–deformation correlations and coherence at the higher Reynolds numbers in spite of the substantially larger deformations. Additional evidence is provided by the comparison with the Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) results, for which the wall thickness, hence the preferred wavelength, is 3.2 times larger. In that case, the p–d correlations are low (figure 22 a) despite the near agreement in the magnitudes of d rms + at Re τ = 2300 with those of the present values. The corresponding premultiplied spectrum calculated from that data, which is presented in figure 25, shows that the larger deformation wavelength deviates significantly from that of the energy-containing eddies. Hence, one should not be surprised by the low pressure-deformation correlations. Interestingly, in the Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) data, the mean flow hardly responds to the wall deformation, resulting in one-way flow–deformation coupling. In contrast, there are clear (but small) changes to the boundary layer mean velocity profile in the Wang et al. (Reference Wang, Koley and Katz2020) data at a similar Re τ and d rms + , but with a wall thickness matching the present values, suggesting that the boundary layer is more ‘sensitive’ to wall deformation.

Figure 25. Contours of $k_{x}{E_{u'u'}}^{+}$ calculated from the Zhang et al. (Reference Zhang, Wang, Blake and Katz2017) data at Re τ = 2300. The horizontal dashed line corresponds to λ = 3l 0.

It should be noted that in addition to the main peaks, the premultiplied spectra in figure 24 contain several lobes with a wide range of scales. Similar lobes have been seen in other experiments involving smooth rigid wall boundary layers and pipe flow (e.g. Guala, Hommema & Adrian Reference Guala, Hommema and Adrian2006; Mathis et al. Reference Mathis, Hutchins and Marusic2009), but have not received significant attention. In order to determine whether the present spectral undulations are associated with the deformation, figure 26(a–c) displays distribution of k x E uu after subtracting the wave-coherent motions, leaving only stochastic part of the turbulence. While still visible, some of the lobes, e.g. those near λ x = 3l 0 in all the three spectra, and those at λ x +> 10 000 at the two higher Reynolds numbers, have significantly lower magnitude, indicating that at least part of the energy in the lobes is associated with the wave-coherent motions, as is also demonstrated in the distribution of $k_{x}E_{\tilde{u}\tilde{u}}$ in figure 26(d–f). While the coherent energy at the lowest Reynolds number is broad and extends to the entire range of energy-containing eddies, the results at the higher Reynolds numbers are patchy and extend well beyond the critical height, consistent with the distributions of the horizontal velocity fluctuation.

Figure 26. Contours of (a,b,c) $k_{x}{E_{u''u''}}^{+}$ , and (d,e,f) $k_{x}{E_{\tilde{u}\tilde{u}}}^{+}$ at (a,d) Re τ = 3300, (b,e) Re τ = 6700 and (c,f) Re τ = 8900.

Figure 27. A comparison of the temporal spectra of the present total, wave-coherent and stochastic pressures at y + = 30 and Re τ = 3300 with the DNS channel flow pressure spectra at the same height and Re τ = 1000. The yellow background marks the k –1 region in the kinetic energy spectra, and the dotted line shows the experimental Nyquist frequency.

The next discussion examines the spectral content of the coherent and stochastic pressures. Figure 27 compares the total pressure spectrum at y + = 30 for Re τ = 3300, with those of the coherent and stochastic pressures. Also included is the pressure spectrum at the same elevation obtained from the Johns Hopkins Turbulence Database (Li et al. Reference Li, Perlman, Wan, Yang, Meneveau, Burns, Chen, Szalay and Eyink2008; Graham et al. Reference Graham2016) DNS for a rigid wall channel flow at Re τ = 1000. The Nyquist frequency of the experimental pressure spectra at ω + = 1.4 is also marked. As is evident from figure 27, the present stochastic pressure spectra collapse well with the DNS spectra at ω + < 0.4 but deviate at higher frequencies. As shown in Agarwal et al. (Reference Agarwal, Ram, Wang, Lu and Katz2021), where the present particle tracking based pressure calculation method is introduced and calibrated, the calculated pressure deviates from the DNS data at frequencies exceeding 20 % of the Nyquist frequency, consistent with the trends depicted by the present results. The area highlighted by a yellow background corresponds to frequencies where the value of the k 1 premultiplied spectrum in figure 24(a) is higher than 75 % of its maximum value. The choice to use 75 % of the peak value as a threshold for defining the frequency bandwidth of the peak has been selected arbitrarily for illustration purpose, and is not used in any other analysis. Within this range, the present stochastic and DNS pressure spectra collapse, but the total pressure amplitude is clearly higher, especially in the vicinity of 3l 0. The difference is associated with the peak in the coherent pressure in the same wavenumber range. In fact, around the peak, the coherent pressure amplitude is nearly equal to that of the stochastic level. Considering that the coherent pressure spectral peak fall in the energetic attached eddy range suggests that the flow and wall deformation are ‘resonating’. Figure 28 is a comparison between trends of the pressure spectra at the three Reynolds numbers, this time in a linear amplitude scale to highlight differences. These pressure spectra are sampled at y/δ = 0.009, i.e. at the same distance from the wall as in figure 27. While the peak $E_{\tilde{p}\tilde{p}}$ at Re τ = 3300 is comparable to $E_{p''p''}$ in the energetic eddy range, $E_{p''p''}$ has a maximum at a significantly lower wavelength, demonstrating that the overall peak is attributable to $E_{\tilde{p}\tilde{p}}$ . In contrast, the relative contributions of $E_{\tilde{p}\tilde{p}}$ at the higher Reynolds numbers are clearly much lower than the stochastic levels. We believe that these findings explain how submicron deformations at a frequency matching that of the energetic eddies in the boundary layer can cause more flow–deformation coupling than orders of magnitude higher deformations with mismatched frequencies.

Figure 28. Temporal spectra of the total, coherent and stochastic pressures at (a) Re τ = 3300, (b) Re τ = 6700 and (c) Re τ = 8900.

5. Summary and conclusions

The interactions between high Reynolds number boundary layers with a compliant wall are investigated experimentally using simultaneous time-resolved TPTV to measure the 3-D flow and pressure field, and MZI to map the 2-D distribution of surface deformation. The present analysis extends the preliminary results described in Lu et al. (Reference Lu, Xiang, Zaki and Katz2024), in which we demonstrate the important role of the critical layer, where the local mean flow speed is equal to the surface wave speed. For the present compliant wall, the critical height increases from 60δ ν at Re τ = 3300, to 190δ ν at Re τ = 8900. Below the critical height, the turbulence is preferentially advected at the wave speed, while above it, the advection speed is equal to the local mean velocity. Furthermore, the coherence and correlation of flow parameters with the wall deformation peak at or near the critical height. However, both the flow–deformation correlations and the coherence decrease with increasing Reynolds number in spite of the order of magnitude increase in deformation height. To explain this puzzling trend, in the present study, a Hilbert projection method is used for decomposing the velocity and pressure fluctuations to wave-coherent motions that are highly correlated with the surface waves, and stochastic turbulence that has a low correlation with the wall motion. The structure and trends of the two unsteady fields are then investigated separately, including the phase-averaged flows and pressures, Reynolds stresses, r.m.s. pressures, vorticity distributions as well as kinetic energies and their production rates.

The spatial distributions of phase-averaged coherent and stochastic flow variables reveal that all the wave-coherent flow variables have wave-phase dependent distributions. While all the coherent variables have high correlations with the surface wave, the correlation with the pressure is particularly high, peaking at 98 %. A negative spanwise vortex generates the pressure minimum at the deformation crest and a sweep-ejection transition is preferentially associated with pressure maximum at the trough. The negative vorticity peak aligned with the crest and the positive vorticity aligned with the trough are centred above the critical height. However, interaction of these vortices with the wall generates counter rotating vorticity below the critical layer. The coherent Reynolds stresses and pressure r.m.s. are an order of magnitude smaller than their stochastic counterparts, except for the pressure r.m.s. very near the wall, where the wave-coherent and stochastic values are comparable. Decomposition of the kinetic energy budget shows that both the wave-coherent and stochastic energy production rates are focused near the wall, well below the critical height, and that the former is 6 %–10 % of the latter, i.e. the mean flow transfer energy predominantly to the stochastic turbulence. The exchange between coherent and stochastic kinetic energies changes sign across the critical layer, flowing from the stochastic to the wave-coherent energy above the critical layer, and from the wave-coherent to the stochastic energy below the critical height. The coherent pressure diffusion term is negative around the peak of coherent energy above the critical layer, and positive below it. Finally, all the coherent flow variables and their correlations with the wave, from the phase-averaged velocity and pressure, to the Reynolds stresses and pressure r.m.s. and to the kinetic energy and transport terms, decrease with increasing Reynolds number, in contrast to the trends of the deformation amplitude.

To explain the strong flow–deformation interactions at the lowest Reynolds number, despite the submicron deformation height, we compare the characteristic deformation wavelength, three times the wall thickness, with the scales of the energy-containing attached eddies dominating the k 1 range of the turbulent energy spectrum. For this case, the deformation wavelength falls in the middle of the premultiplied streamwise energy spectrum (k x E uu ), suggesting that the compliant wall waves resonate with the attached eddies. Indeed, k x E uu contains lobes, some of them centred around the preferred surface wavelength, which contain significant wave-coherent energy. At higher Reynolds numbers, the deformation wavelength falls at the edge of the attached eddy range, as the deformation wavelength remains the same while these eddies decrease in size. Hence, the interactions become weaker even for considerably higher deformation amplitudes. Furthermore, when the scales overlap at the lowest Reynolds number, the near-wall coherent pressure spectrum at a wavelength corresponding to the surface wave is comparable in magnitude to the stochastic pressure, causing a significant increase in the overall pressure fluctuations. At higher Reynolds number, the coherent pressure is significantly smaller than the stochastic pressure spectra, which peak at other frequencies. These findings provide a plausible explanation on how turbulent eddies in a boundary layer become phase locked with submicron compliant wall deformations. They could guide the choice of compliant wall thickness that would maximize (or minimize) the wall-flow interactions in future efforts aimed at developing flow control strategies. Future experiments should also focus on the flow–deformation interactions below the critical layer, hence require a higher resolution, and include means to determine the horizontal and vertical velocity fields of the compliant material.

One should keep in mind that the present deformation r.m.s. values are still smaller than δ ν . The flow–deformation interactions are expected to change significantly with further decrease in E/ρU 0 2 and the resulting increase in deformation amplitude (Wang et al. Reference Wang, Koley and Katz2020; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022). For example, flow separation might occur at the crest of the deformation (Esteghamatian et al. Reference Esteghamatian, Katz and Zaki2022), and surface instabilities, e.g. static-divergence waves, might develop when the flow speed exceeds the material shear speed (Gad-El-Hak et al. Reference Gad-El-Hak, Blackwelder and Riley1984; Duncan Reference Duncan1986; Greidanus et al. Reference Greidanus, Delfos, Picken and Westerweel2022).

Funding

This project is funded by the Office of Naval Research under grant numbers N00014-23-1-2681 and N00014-20-1-2778. Greg Orris, Christine Sanders, and Meredith Hutchinson are the programme officers. The authors would also like to thank the late Y. Ronzhes for designing the water tunnel.

Declaration of interests

The authors report no conflict of interest.

Appendix A. The Hilbert projection method

For a real-valued deformation, d(x,z,t), the Hilbert transform (see comprehensive tutorials in Bendat & Piersol Reference Bendat and Piersol1986), denoted as $\mathcal{D}$ (t), is given by

(A1) \begin{equation} \mathcal{D}(t)=\mathcal{H}\left\{{d}(t)\right\}=\dfrac{1}{\unicode{x03C0} } p.v.\int _{-\infty }^{+\infty }\dfrac{{d}(\tau)}{t-\tau }{\rm d}\tau , \end{equation}

where $\mathcal{H}$ indicates the Hilbert transform operator, p.v. represents the Cauchy principal value of the integral and τ is a time shift. Equivalently, $\mathcal{D}$ (t) is the convolution (denoted by *) of d(t) with (1/πt),

(A2) \begin{equation} \begin{array}{c} \mathcal{D}(t)=d(t)*\left(1/\unicode{x03C0} t\right).\end{array} \end{equation}

The Fourier transform of $\mathcal{D}$ (t) is given by

(A3) \begin{equation} \begin{array}{c} \mathcal{F}\left\{\mathcal{D}(t)\right\}=\mathcal{F}\left\{d(t)\right\}\boldsymbol{\cdot }\mathcal{F}\left\{\dfrac{1}{\unicode{x03C0} t}\right\}=\hat{d}\left(\omega \right)\boldsymbol{\cdot }\left\{-i \textit{sgn}\left(\omega \right)\right\}=\begin{cases} i\hat{d}\left(\omega \right)\quad\,\textrm{for}\,\,\, \omega \lt 0,\\ 0 \quad\quad\,\,\,\,\,\,\,\textrm{for}\, \,\,\omega =0,\\ -i\hat{d}\left(\omega \right)\,\,\textrm{for}\,\,\, \omega \gt 0, \end{cases}\end{array} \end{equation}

where $\hat{d}(\omega )$ is the Fourier transform of d(t). Hence, $\mathcal{D}$ (t) can be readily calculated from the inverse Fourier transform of (A3). The functions d(t) and $\mathcal{D}$ (t) form a quadrature pair, i.e. they have a phase difference of π/2, defining the following complex analytic form:

(A4) \begin{equation} \begin{array}{c} \psi (t)=d(t)+i\mathcal{D}(t)=A(t)e^{i\phi (t)}, \end{array} \end{equation}

where $A(t)=\sqrt{d(t)^{2}+\mathcal{D}(t)^{2}}$ and $\phi (t)=\tan ^{-1} \{\mathcal{D}(t)/d(t)\}$ are the amplitude and phase of $\psi (t)$ , respectively.

Following Hristov et al. (Reference Hristov, Friehe and Miller1998), the Hilbert projection method consists of the following steps. First, the original d(t) signal is divided into a series of narrow bandpass filtered signals denoted as d k (t) in the frequency domain, each with a bandwidth of Δω = 0.15U 0 /l 0. Using the Fourier modes of d(t),

(A5) \begin{equation} d_{k}(t)=\sum _{\omega _{k}-{\unicode[Arial]{x0394}} \omega /2\lt \omega _{n}\lt \omega _{k}+{\unicode[Arial]{x0394}} \omega /2}C_{n}e^{i{(\omega _{n}}t)}, \end{equation}

where the Fourier coefficients, $ C_{n}$ , are obtained using fast Fourier transform in MATLAB. For each d k (t), we construct its analytic signal $\psi _{k}(t)$ , similar to equation (A4). The wave coherent part of a signal, $\tilde{f}$ , is obtained by projecting $f'$ onto $\psi _{k}(t)$ , namely

(A6) \begin{equation} \tilde{f}=Real\left\{\sum _{k}\frac{\langle f',\psi _{k}\rangle }{\left\| \psi _{k}\right\| ^{2}} \psi _{k}\right\}\!, \end{equation}

where <·> indicates an inner product, and ||·||2 is the squared norm. In Hristov & Plancarte (Reference Hristov and Ruiz-Plancarte2014) and Wu, Hristov & Rutgersson (Reference Wu, Hristov and Rutgersson2018), (A6) is further clarified to indicate that

(A7) \begin{equation} \begin{array}{c} \tilde{f}=\sum _{k}\left\{\dfrac{\langle f',d_{k}\rangle }{\left\| d_{k}\right\| ^{2}}d_{k}+\dfrac{\langle f',\mathcal{D}_{k}\rangle }{\left\| \mathcal{D}_{k}\right\| ^{2}}\mathcal{D}_{k}\right\} \end{array}\! \!.\end{equation}

Figure 29. Sample instantaneous snapshots of the original, phase-averaged and Hilbert-projected deformations. The presented region is for x + = −200 ∼ 750, z + = −200 ∼ 0 at Re τ = 3300.

Since $d_{k}$ and $\mathcal{D}_{k}$ have a phase difference of π/2, the ratio of these projections gives the phase lag between $\tilde{f}$ and $d_{k}$ .

To demonstrate the advantage of the Hilbert projection method introduced by Hristov et al. (Reference Hristov, Friehe and Miller1998), we first examine its ability to reproduce the original wave, i.e. the wave-coherent deformation should be nearly equal to the deformation itself, $\tilde{d}$ (x,z,t)∼d(x,z,t). Subsequently, correlations are used for confirming the high coherence between the deformation and projected flow variables, such as pressure and velocity. As a baseline for comparison, the projection results are compared with the phase-averaged deformation and flow variables, denoted as $\hat{d}$ (x,z,t) and $\hat{f}$ (x,z,t). The latter are evaluated using the following steps. For each snapshot, an instantaneous phase is assigned to each (x,z) point based on the Hilbert Transform of d(x,z,t). Based on the definitions above, phases 0 and ±π are assigned to the wave crest and trough, respectively. The interval between peaks is divided to 20 bins, each with a phase width of π/10, and record the location and flow variables corresponding to each bin. Results obtained for the entire data set are then phase-averaged to obtain $\hat{d}$ and $\hat{f}$ . In addition, figure 29 compares a sample snapshot of d(x,z,t) at Re τ = 3300, with the instantaneous $\tilde{d}$ (x,z,t) and $\hat{d}$ (φ), where the instantaneous phase of the peak in d(x,z,t) is matched with $\phi$ = 0 of the phase-averaged waveform. As is evident, $\tilde{d}$ (x,z,t) reproduces d(x,z,t), and in fact, the temporal correlation between them for the entire data is 99.9 %. In contrast, the temporal correlation of $\hat{d}$ with the original signal is 78 %. Figure 30 compares the coherent pressure and vertical velocity component with the phase-averaged distributions. The maximum temporal $\tilde{p}$ –d and $\tilde{v}$ –d correlations, obtained after shifting the time series relative to each other and finding the maximum values, are 96 % and 75 %, respectively. The corresponding maximum $\hat{p}$ –d and $\hat{v}$ –d correlations are only 76 % and 58 %. Clearly, the Hilbert projection is an effective tool for finding the coherent part of the signal. The remaining incoherent part of the pressure or flow variables, which are referred to as ‘stochastic’ are obtained by subtracting the coherent signal from the original data.

Figure 30. Sample instantaneous distributions of (a) $\tilde{p}$ and $\hat{p}$ , (b) $\tilde{v}$ and $\hat{v}$ , (c) d at z + = $-100$ and Re τ = 3300.

References

Adrian, R.J. 2007 Hairpin vortex organization in wall turbulence. Phys. Fluids 19 (4), 041301.10.1063/1.2717527CrossRefGoogle Scholar
Agarwal, K., Ram, O., Lu, Y. & Katz, J. 2023 On the pressure field, nuclei dynamics and their relation to cavitation inception in a turbulent shear layer. J. Fluid Mech. 966, A31.10.1017/jfm.2023.368CrossRefGoogle Scholar
Agarwal, K., Ram, O., Wang, J., Lu, Y. & Katz, J. 2021 Reconstructing velocity and pressure from noisy sparse particle tracks using constrained cost minimization. Exp. Fluids 62, 120.10.1007/s00348-021-03172-0CrossRefGoogle Scholar
Bai, K. & Katz, J. 2014 On the refractive index of sodium iodide solutions for index matching in PIV. Exp. Fluids 55, 16.10.1007/s00348-014-1704-xCrossRefGoogle Scholar
Bendat, J.S. & Piersol, A.G. 1986 Random Data: Analysis and Measurement Procedure. 2nd edn. Wiley InterScience.Google Scholar
Benjamin, T.B. 1960 Effects of a flexible boundary on hydrodynamic stability. J. Fluid Mech. 9, 513532.10.1017/S0022112060001286CrossRefGoogle Scholar
Benschop, H.O.G., Greidanus, A.J., Delfos, R., Westerweel, J. & Breugem, W.-P. 2019 Deformation of a linear viscoelastic compliant coating in a turbulent flow. J. Fluid Mech. 859, 613658.10.1017/jfm.2018.813CrossRefGoogle Scholar
Boggs, F.W. & Hahn, E.R. 1962 Performance of compliant skins in contact with high velocity flow in water. In Proc. 7th Joint Army-Navy-Air Force Conf. on Elastomer Research and Development, vol. 2, pp.443. Office of Naval Research.Google Scholar
Buckley, M.P. & Veron, F. 2019 The turbulent airflow over wind generated surface waves. Eur. J. Mech. (B/Fluids) 73, 132143.10.1016/j.euromechflu.2018.04.003CrossRefGoogle Scholar
Busse, A. & Jelly, T.O. 2023 Effect of high skewness and kurtosis on turbulent channel flow over irregular rough walls. J. Turbul. 24 (1–2), 5781.10.1080/14685248.2023.2173761CrossRefGoogle Scholar
Calaf, M., Hultmark, M., Oldroyd, H.J., Simeonov, V. & Parlange, M.B. 2013 Coherent structures and the k 1 spectral behaviour. Phys. Fluids 25 (12), 125107.10.1063/1.4834436CrossRefGoogle Scholar
Cao, T. & Shen, L. 2021 A numerical and theoretical study of wind over fast-propagating water waves. J. Fluid Mech. 919, A38.10.1017/jfm.2021.416CrossRefGoogle Scholar
Carpenter, J.R., Buckley, M.P. & Veron, F. 2022 Evidence of the critical layer mechanism in growing wind waves. J. Fluid Mech. 948, A26.10.1017/jfm.2022.714CrossRefGoogle Scholar
Carpenter, P.W., Davies, C. & Lucey, A.D. 2000 Hydrodynamics and compliant walls: does the dolphin have a secret? J. Curr. Sci. 79, 758765.Google Scholar
Castellini, P., Martarelli, M. & Tomasini, E.P. 2006 Laser doppler vibrometry: development of advanced solutions answering to technology’s needs. Mech. Syst. Signal Process. 20, 12651285.10.1016/j.ymssp.2005.11.015CrossRefGoogle Scholar
Charruault, F., Greidanus, A.J., Breugem, W.-P. & Westerweel, J. 2018 A dot tracking algorithm to measure free surface deformations. In Proceedings 18th International Symposium on Flow Visualization. Eth Zurich.Google Scholar
Chase, D.M. 1991 Generation of fluctuating normal stress in a viscoelastic layer by surface shear stress and pressure as in turbulent boundary-layer flow. J. Acoust. Soc. Am. 89 (6), 25892596.10.1121/1.400698CrossRefGoogle Scholar
Choi, K.-S., Yang, X., Clayton, B.R., Glover, E.J., Atlar, M., Semenov, B.N. & Kulik, V.M. 1997 Turbulent drag reduction using compliant surfaces. Proc. R. Soc. Lond. A. 453 (1965), 22292240.10.1098/rspa.1997.0119CrossRefGoogle Scholar
De Graaff, D.B. & Eaton, J.K. 2000 Reynolds-number scaling of the flat-plate turbulent boundary layer. J. Fluid Mech. 422, 319346.10.1017/S0022112000001713CrossRefGoogle Scholar
Do, J., Wang, B. & Chang, K.A. 2024 Turbulence over young wind waves dominated by capillaries and micro-breakers. J. Fluid Mech. 985, A22.10.1017/jfm.2024.308CrossRefGoogle Scholar
Duncan, J.H. 1986 The response of an incompressible viscoelastic coating to pressure fluctuations in a turbulent boundary layer. J. Fluid Mech. 171, 339363.10.1017/S0022112086001477CrossRefGoogle Scholar
Einaudi, F. & Finnigan, J. 1993 Wave-turbulence dynamics in the stably stratified boundary layer. J. Atmos. Sci. 50 (13), 18411864.10.1175/1520-0469(1993)050<1841:WTDITS>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Endo, T. & Himeno, R. 2002 Direct numerical simulation of turbulent flow over a compliant surface. J. Turbul. 3 (1), 7.10.1088/1468-5248/3/1/007CrossRefGoogle Scholar
Esteghamatian, A., Katz, J. & Zaki, T.A. 2022 Spatiotemporal characterization of turbulent channel flow with a hyperelastic compliant wall. J. Fluid Mech. 942, A35.10.1017/jfm.2022.354CrossRefGoogle Scholar
Fernholz, H.H. & Finley, P.J. 1996 The incompressible zero-pressure-gradient turbulent boundary layer – an assessment of the data. Prog. Aero. Sci. 32, 245311.10.1016/0376-0421(95)00007-0CrossRefGoogle Scholar
Fisher, D.H. & Blick, E.F. 1966 Turbulent damping by flabby skins. J. Aircraft. 3 (2), 163164.10.2514/3.43722CrossRefGoogle Scholar
Flack, K.A. & Schultz, M.P. 2014 Roughness effects on wall-bounded turbulent flows. Phys. Fluids 26 (10), 101305.10.1063/1.4896280CrossRefGoogle Scholar
Freund, L.B. 1998 Dynamic Fracture Mechanics. Cambridge University Press.Google Scholar
Fukagata, K., Kern, S., Chatelain, P., Koumoutsakos, P. & Kasagi, N. 2008 Evolutionary optimization of an anisotropic compliant surface for turbulent friction drag reduction. J. Turbul. 9, N35.10.1080/14685240802441126CrossRefGoogle Scholar
Gad-El-Hak, M. 1986 The response of elastic and viscoelastic surfaces to a turbulent boundary layer. Trans. Asme E: J. Appl. Mech. 53, 206212.10.1115/1.3171714CrossRefGoogle Scholar
Gad-El-Hak, M., Blackwelder, R.F. & Riley, J.J. 1984 On the interaction of compliant coatings with boundary-layer flows. J. Fluid Mech. 140, 257280.10.1017/S0022112084000598CrossRefGoogle Scholar
Ganapathisubramani, B., Hutchins, N., Hambleton, W.T., Longmire, E.K. & Marusic, I. 2005 Investigation of large-scale coherence in a turbulent boundary layer using two-point correlations. J. Fluid Mech. 524, 5780.10.1017/S0022112004002277CrossRefGoogle Scholar
George, W.K. & Castillo, L. 1997 Zero-pressure-gradient turbulent boundary layer. Appl. Mech. Rev. 50, 689729.10.1115/1.3101858CrossRefGoogle Scholar
Graham, J., et al. 2016 A Web services accessible database of turbulent channel flow and its use for testing a new integral wall model for LES. J. Turbul. 17 (2), 181215.10.1080/14685248.2015.1088656CrossRefGoogle Scholar
Grare, L., Lenain, L. & Melville, W.K. 2013 Wave-coherent airflow and critical layers over ocean waves. J. Phys. Oceanogr. 43 (10), 21562172.10.1175/JPO-D-13-056.1CrossRefGoogle Scholar
Greidanus, A.J., Delfos, R., Picken, S.J. & Westerweel, J. 2022 Response regimes in the fluid–structure interaction of wall turbulence over a compliant coating. J. Fluid Mech. 952, A1.10.1017/jfm.2022.774CrossRefGoogle Scholar
Guala, M., Hommema, S.E. & Adrian, R.J. 2006 Large-scale and very-large-scale motions in turbulent pipe flow. J. Fluid Mech. 554, 521542.10.1017/S0022112006008871CrossRefGoogle Scholar
Hara, T. & Sullivan, P.P. 2015 Wave boundary layer turbulence over surface waves in a strongly forced condition. J. Phys. Oceanogr. 45 (3), 868883.10.1175/JPO-D-14-0116.1CrossRefGoogle Scholar
Hristov, T.S. & Ruiz-Plancarte, J. 2014 Dynamic balances in a wavy boundary layer. J. Phys. Oceanogr. 44 (12), 31853194.10.1175/JPO-D-13-0209.1CrossRefGoogle Scholar
Hristov, T.S., Friehe, C. & Miller, S. 1998 Wave-coherent fields in air flow over ocean waves: identification of cooperative behavior buried in turbulence. Phys. Rev. Lett. 81 (23), 5245.10.1103/PhysRevLett.81.5245CrossRefGoogle Scholar
Hristov, T.S., Miller, S. & Friehe, C. 2003 Dynamical coupling of wind and ocean waves through wave-induced air flow. Nature. 422 (6927), 5558.10.1038/nature01382CrossRefGoogle ScholarPubMed
Hsu, C.T., Hsu, E.Y. & Street, R.L. 1981 On the structure of turbulent flow over a progressive water wave: theory and experiment in a transformed, wave-following co-ordinate system. J. Fluid Mech. 105, 87117.10.1017/S0022112081003121CrossRefGoogle Scholar
Hultmark, M., Vallikivi, M., Bailey, S.C.C. & Smits, A.J. 2012 Turbulent pipe flow at extreme Reynolds numbers. Phys. Rev. Lett. 108 (9), 094501.10.1103/PhysRevLett.108.094501CrossRefGoogle ScholarPubMed
Hunt, J.C.R., Leibovich, S. & Richards, K.J. 1988 Turbulent shear flows over low hills. Q. J. R. Meteorol. Soc. 114 (484), 14351470.10.1002/qj.49711448405CrossRefGoogle Scholar
Hussain, A.K.M.F. & Reynolds, W.C. 1970 The mechanics of an organized wave in turbulent shear flow. J. Fluid Mech. 41 (2), 241258.10.1017/S0022112070000605CrossRefGoogle Scholar
Hutchins, N. & Marusic, I. 2007 Evidence of very long meandering features in the logarithmic region of turbulent boundary layers. J. Fluid Mech. 579, 128.10.1017/S0022112006003946CrossRefGoogle Scholar
Huynh, D. & Mckeon, B. 2020 Measurements of a turbulent boundary layer-compliant surface system in response to targeted, dynamic roughness forcing. Exp. Fluids 61, 115.10.1007/s00348-020-2933-9CrossRefGoogle Scholar
Jeffreys, H. 1925 On the formation of water waves by wind. Proc. R. Soc. Lond. A. 107, 189206.Google Scholar
Jeong, J. & Hussain, F. 1995 On the identification of a vortex. J. Fluid Mech. 285, 6994.10.1017/S0022112095000462CrossRefGoogle Scholar
Jiménez, J. 2018 Coherent structures in wall-bounded turbulence. J. Fluid Mech. 842, P1.10.1017/jfm.2018.144CrossRefGoogle Scholar
Jiménez, J., Hoyas, S., Simens, M.P. & Mizuno, Y. 2010 Turbulent boundary layers and channels at moderate Reynolds numbers. J. Fluid Mech. 657, 335360.10.1017/S0022112010001370CrossRefGoogle Scholar
Kramer, M.O. 1957 Boundary-layer stabilization by distributed damping. J. Aero. Sci. 24, 459460.Google Scholar
Kramer, M.O. 1962 Boundary-layer stabilization by distributed damping. Naval Engrs J. 74 (2), 341348.10.1111/j.1559-3584.1962.tb05568.xCrossRefGoogle Scholar
Krasitskii, V.P. & Zaslavskii, M.M. 1978 Comments on the Phillips’-Miles’ theory of wind wave generation. Boundary-Layer Meteorol. 14 (2), 199215.10.1007/BF00122619CrossRefGoogle Scholar
Kumar, P. & Mahesh, K. 2022 A method to determine wall shear stress from mean profiles in turbulent boundary layers. Exp. Fluids 63 (1), 6.10.1007/s00348-021-03352-yCrossRefGoogle Scholar
Landahl, M.T. 1962 On the stability of a laminar incompressible boundary layer over a flexible surface. J. Fluid Mech. 13, 609632.10.1017/S002211206200097XCrossRefGoogle Scholar
Lee, M. & Moser, R.D. 2015 Direct numerical simulation of turbulent channel flow up to re τ ≈ 5200. J. Fluid Mech. 774, 395415.10.1017/jfm.2015.268CrossRefGoogle Scholar
Lee, T., Fisher, M. & Schwarz, W.H. 1993 a Investigation of the stable interaction of a passive compliant surface with a turbulent boundary layer. J. Fluid Mech. 257, 373401.10.1017/S002211209300312XCrossRefGoogle Scholar
Lee, T., Fisher, M. & Schwarz, W.H. 1993 b The measurement of flow-induced surface displacement on a compliant surface by optical holographic interferometry. Exp. Fluids 14, 159168.10.1007/BF00189506CrossRefGoogle Scholar
Lee, T., Fisher, M. & Schwarz, W.H. 1995 Investigation of the effects of a compliant surface on boundary-layer stability. J. Fluid Mech. 288, 3758.10.1017/S0022112095001054CrossRefGoogle Scholar
Li, Y., Perlman, E., Wan, M., Yang, Y., Meneveau, C., Burns, R., Chen, S., Szalay, A. & Eyink, G. 2008 A public turbulence database cluster and applications to study lagrangian evolution of velocity increments in turbulence. J. Turbul. 9, N31.10.1080/14685240802376389CrossRefGoogle Scholar
Lighthill, M.J. 1963 Boundary layer theory. In Laminar Boundary Layers,(Rosenhead, L.), pp. 46103. Oxford University Press.Google Scholar
Lighthill, M.J. 1962 Physical interpretation of the mathematical theory of wave generation by wind. J. Fluid Mech. 14 (3), 385398.10.1017/S0022112062001305CrossRefGoogle Scholar
Lissaman, P.B.S. & Harris, G.L. 1969 Turbulent skin friction on compliant surfaces. AIAA J. 7 (8), 16251627.10.2514/3.5452CrossRefGoogle Scholar
Liu, X. & Katz, J. 2006 Instantaneous pressure and material acceleration measurements using a four-exposure PIV system. Exp. Fluids 41, 227240.10.1007/s00348-006-0152-7CrossRefGoogle Scholar
Lu, Y., Ram, O., Jose, J., Agarwal, K. & Katz, J. 2021 A water tunnel with inline cyclone separator for removing freestream bubble. In Proceedings of the 11th International Symposium on Cavitation. Available at: http://cav2021.org/wp-content/uploads/2021/07/P00125_optimize-3.pdf.Google Scholar
Lu, Y., Xiang, T., Zaki, T.A. & Katz, J. 2024 On the scaling and critical layer in a turbulent boundary layer over a compliant surface. J. Fluid Mech. 980, R2.10.1017/jfm.2024.11CrossRefGoogle Scholar
Makin, V.K. & Kudryavtsev, V.N. 1999 Coupled sea surface–atmosphere model: 1. Wind over waves coupling. J. Geophys. Res: Oceans 104 (C4), 76137623.10.1029/1999JC900006CrossRefGoogle Scholar
Mathis, R., Hutchins, N. & Marusic, I. 2009 Large-scale amplitude modulation of the small-scale structures in turbulent boundary layers. J. Fluid Mech. 628, 311337.10.1017/S0022112009006946CrossRefGoogle Scholar
Mcmichael, J.M., Klebanoff, P.S. & Mease, N.E. 1980 Experimental investigation of drag on a compliant surface. Prog. Astronaut. Aeronaut. 72, 410438.Google Scholar
Meinhart, C.D., Wereley, S.T. & Santiago, J.G. 2000 A piv algorithm for estimating time-averaged velocity fields. J. Fluids Engng 122 (2), 285289.10.1115/1.483256CrossRefGoogle Scholar
Miles, J.W. 1957 On the generation of surface waves by shear flows. J. Fluid Mech. 3 (2), 185204.10.1017/S0022112057000567CrossRefGoogle Scholar
Morrill-Winter, C., Philip, J. & Klewicki, J. 2017 An invariant representation of mean inertia: theoretical basis for a log law in turbulent boundary layers. J. Fluid Mech. 813, 594617.10.1017/jfm.2016.875CrossRefGoogle Scholar
Morton, B.R. 1984 The generation and decay of vorticity. Geophys. Astrophys. Fluid Dyn. 28 (3–4), 277308.10.1080/03091928408230368CrossRefGoogle Scholar
Nakato, M., Onogi, H., Himeno, Y., Tanaka, I. & Suzuki, T. 1985 Resistance due to surface roughness. In Proceedings of the 15th Symposium on Naval Hydrodynamics, pp. 553568.Google Scholar
Napoli, E., Armenio, V. & De Marchis, M. 2008 The effect of the slope of irregularly distributed roughness elements on turbulent wall-bounded flows. J. Fluid Mech. 613, 385394.10.1017/S0022112008003571CrossRefGoogle Scholar
Nickels, T.B., Marusic, I., Hafez, S. & Chong, M.S. 2005 Evidence of the k 1 -1 law in high-reynolds number turbulent boundary layer. Phys. Rev. Lett. 95, 074501.10.1103/PhysRevLett.95.074501CrossRefGoogle Scholar
Perry, A.E. & Chong, M.S. 1982 On the mechanism of wall turbulence. J. Fluid Mech. 119, 173217.10.1017/S0022112082001311CrossRefGoogle Scholar
Phillips, O.M. 1957 On the generation of waves by turbulent wind. J. Fluid Mech. 2 (5), 417445.10.1017/S0022112057000233CrossRefGoogle Scholar
Pope, S.B. 2000 Turbulent Flows. Cambridge University Press.Google Scholar
Raffel, M., Willert, C.E., Scarano, F., Kähler, C.J., Wereley, S.T. & Kompenhans, J. 2018 Particle Image Velocimetry: A Practical Guide. 3rd edn. Springer.10.1007/978-3-319-68852-7CrossRefGoogle Scholar
Reynolds, W.C. & Hussain, A.K.M.F. 1972 The mechanics of an organized wave in turbulent shear flow. Part 3. Theoretical models and comparisons with experiments. J. Fluid Mech. 54 (2), 263288.10.1017/S0022112072000679CrossRefGoogle Scholar
Riley, D.S., Donelan, M.A. & Hui, W.H. 1982 An extended Miles’ theory for wave generation by wind. Boundary-Layer Meteorol. 22, 209225.10.1007/BF00118254CrossRefGoogle Scholar
Rosti, M.E. & Brandt, L. 2017 Numerical simulation of turbulent channel flow over a viscous hyper-elastic wall. J. Fluid Mech. 830, 708735.10.1017/jfm.2017.617CrossRefGoogle Scholar
Rutgersson, A. & Sullivan, P.P. 2005 The effect of idealized water waves on the turbulence structure and kinetic energy budgets in the overlying airflow. Dyn. Atmos. Oceans. 38 (3–4), 147171.10.1016/j.dynatmoce.2004.11.001CrossRefGoogle Scholar
Schanz, D., Gesemann, S. & Schröder, A. 2016 Shake-The-Box: Lagrangian particle tracking at high particle image densities. Exp. Fluids 57, 127.10.1007/s00348-016-2157-1CrossRefGoogle Scholar
Schultz, M.P. & Flack, K.A. 2009 Turbulent boundary layers on a systematically varied rough wall. Phys. Fluids 21 (1), 015104.10.1063/1.3059630CrossRefGoogle Scholar
Schultz, M.P. & Flack, K.A. 2013 Reynolds-number scaling of turbulent channel flow. Phys. Fluids 25 (2), 025104.10.1063/1.4791606CrossRefGoogle Scholar
Sillero, J.A., Jiménez, J. & Moser, R.D. 2014 Two-point statistics for turbulent boundary layers and channels at Reynolds numbers up to δ + ≈ 2000. Phys. Fluids 26 (10), 105109.10.1063/1.4899259CrossRefGoogle Scholar
Townsend, A.A. 1976 the Structure of Turbulent Shear Flow. 2nd edn. Cambridge University Press.Google Scholar
Wang, J., Koley, S.S. & Katz, J. 2020 On the interaction of a compliant wall with a turbulent boundary layer. J. Fluid Mech. 899, A20.10.1017/jfm.2020.446CrossRefGoogle Scholar
Wang, J., Zhang, C. & Katz, J. 2019 GPU-based, parallel-line, omni-directional integration of measured pressure gradient field to obtain the 3D pressure distribution. Exp. Fluids 60, 124.10.1007/s00348-019-2700-yCrossRefGoogle Scholar
Wang, Z., Yeo, K.S. & Khoo, B.C. 2006 On two-dimensional linear waves in Blasius boundary layer over viscoelastic layers. Eur. J. Mech. (B/Fluids) 25 (1), 3358.10.1016/j.euromechflu.2005.04.006CrossRefGoogle Scholar
Wei, T., Schmidt, R. & Mcmurtry, P. 2005 Comment on the Clauser chart method for determining the friction velocity. Exp. Fluids 38, 695699.10.1007/s00348-005-0934-3CrossRefGoogle Scholar
Wu, L., Hristov, T. & Rutgersson, A. 2018 Vertical profiles of wave-coherent momentum flux and velocity variances in the marine atmospheric boundary layer. J. Phys. Oceanogr. 48 (3), 625641.10.1175/JPO-D-17-0052.1CrossRefGoogle Scholar
Yang, D. & Shen, L. 2010 Direct-simulation-based study of turbulent flow over various waving boundaries. J. Fluid Mech. 650, 131180.10.1017/S0022112009993557CrossRefGoogle Scholar
Yousefi, K. & Veron, F. 2020 Boundary layer formulations in orthogonal curvilinear coordinates for flow over wind-generated surface waves. J. Fluid Mech. 888, A11.10.1017/jfm.2020.32CrossRefGoogle Scholar
Yousefi, K., Veron, F. & Buckley, M.P. 2021 Turbulent and wave kinetic energy budgets in the airflow over wind generated surface waves. J. Fluid Mech. 920, A33.10.1017/jfm.2021.377CrossRefGoogle Scholar
Zhang, C., Miorini, R. & Katz, J. 2015 Integrating Mach–Zehnder interferometry with TPIV to measure the time-resolved deformation of a compliant wall along with the 3D velocity field in a turbulent channel flow. Exp. Fluids 56 (11), 203.10.1007/s00348-015-2072-xCrossRefGoogle Scholar
Zhang, C., Wang, J., Blake, W. & Katz, J. 2017 Deformation of a compliant wall in a turbulent channel flow. J. Fluid Mech. 823, 345390.10.1017/jfm.2017.299CrossRefGoogle Scholar
Zhang, E., Wang, Z. & Liu, Q. 2024 A numerical investigation of momentum flux and kinetic energy transfers between turbulent wind and propagating waves. Flow 4, E14.10.1017/flo.2024.22CrossRefGoogle Scholar
Figure 0

Figure 1. Schematics of (a) the refractive index-matched water tunnel, (b) the test section and (c) the cyclone separator.

Figure 1

Figure 2. (a) The compliant coating and location of the sample volume, and (b,c) the optical set-up of the integrated TPTV-MZI system shown in (b) front view and (c) top view.

Figure 2

Table 1. The experimental conditions and scales of data acquisition.

Figure 3

Figure 3. The PDFs of the compliant wall deformations at the indicated Reynolds numbers.

Figure 4

Figure 4. Mean velocity profiles based on the present TPTV (reproduced from Lu et al.2024), stereo PIV and the 2-D PIV data of Wang et al. (2020). Dashed lines indicate the boundary layer heights, and dash–dotted lines, the critical heights.

Figure 5

Figure 5. An instantaneous sample snapshot at Reτ = 3300 of the velocity vectors and pressure contours in two planes, along with the wall deformations presented in exaggerated scales. The 3-D blobs are isosurfaces of ${\lambda _{2}}^{+}$ = −1.4 × 10−3.

Figure 6

Figure 6. Wall normal profiles of Reynolds stresses: (a) $\overline{u'u'}^{+}$, (b) $\overline{v'v'}^{+}$ and (c) $-\overline{u'v'}^{+}$ non-dimensionalized using inner scaling and compared with results of DNS for rigid smooth walls (solid lines).

Figure 7

Figure 7. Wall normal profiles of Reynolds stresses plotted using mixed scaling: (a) $\overline{u'u'}^{+}$, (b) $\overline{v^{\prime}v^{\prime}}^{+}$, (c) $-\overline{u^{\prime}v^{\prime}}^{+}$. Circles, 8 × 8 × 8 voxels binning; lines, 6 × 6 × 6 voxel binning.

Figure 8

Figure 8. (a) Power spectral densities of the wall deformation, and (b) sample time segments of the instantaneous wall deformation at the indicated times, z+ = 0, and Reτ = 3300.

Figure 9

Figure 9. Conditional correlations of the deformation with the indicated flow variables based on (a) bumps, and (b) dimples, both at Reτ = 3300. In each set, subpanels (i), (iii), (v), (vii) and (ix) show the correlations with wave-coherent components ($C_{\tilde{f}-d}$), subpanels (ii), (iv), (vi), (viii) and (x) the correlation with the ‘stochastic’ turbulence ($C_{f''-d}$). The subpanels from (i) to (vi) show the distributions of $C_{u-d}$, $C_{v-d}$ and $C_{p-d}$. Subpanels (vii) and (viii) present the conditionally averaged wall shape, and (ix) and (x), the variations of peak magnitudes of conditional correlations with Reynolds number.

Figure 10

Figure 10. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,c,e) $\widehat{\tilde{u}}$+, (b,d,f) $\widehat{\tilde{v}}$+ and (g) $\hat{d}/\delta$ with deformation phase; (a,b) Reτ = 3300, (c,d) Reτ = 6700 and (e,f) Reτ = 8900. The arrows in (a,c,e) show the velocity vectors, and the yellow dashed lines indicate the critical heights. Panels (h) and (i) display the variations of peak (h) $\widehat{\tilde{u}}$+ and (i) $\widehat{\tilde{v}}$+ with Reynolds number.

Figure 11

Figure 11. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,c,e) $\widehat{\widetilde{\omega _{z}}}$+, (b,d,f) $\widehat{\tilde{p}}$+,and (g) $\hat{d}/\delta$ with deformation phase; (a,b) Reτ = 3300, (c,d) Reτ = 6700 and (e,f) Reτ = 8900. The yellow dashed lines indicate the critical heights. Panels (h) and (i) display the variations of peak (h) $\widehat{\widetilde{\omega _{z}}}$+ and (i) $\widehat{\tilde{p}}$+ with Reynolds number.

Figure 12

Figure 12. Conditionally averaged flow variables and deformation at Δz+ = 0 for (a,c) a surface bump, and (b,d) a dimple, both at Reτ = 3300: (a,b) pressure contours and in-plane velocity vectors, and (c,d) compliant wall shape.

Figure 13

Figure 13. Variations of the wave-coherent, spatially and temporally phase-averaged: (a,d,g) $\widehat{\tilde{u}\tilde{u}}$+, (b,e,h) $\widehat{\tilde{v}\tilde{v}}$+, (c,f,i) $-\widehat{\tilde{u}\tilde{v}}$+ and (j) $\hat{d}/\delta$ with deformation phase: (a–c) Reτ = 3300, (d–f) Reτ = 6700 and (g–i) Reτ = 8900. Yellow dashed lines indicate the critical heights. Panels (k–m) display the variations of peak (k) $\widehat{\tilde{u}\tilde{u}}$+, (l) $\widehat{\tilde{v}\tilde{v}}$+ and (m) $-\widehat{\tilde{u}\tilde{v}}$+ with the Reynolds number.

Figure 14

Figure 14. Variations of the stochastic, spatially and temporally phase-averaged: (a,d,g) $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$+, (b,e,h) $\widehat{v^{\prime\prime}v^{\prime\prime}}$+, (c,f,i) $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$+ and (j) $\hat{d}/\delta$ with deformation phase; (a–c) Reτ = 3300, (d–f) Reτ = 6700 and (g–i) Reτ = 8900. Panels (k–m) display the variations of peak (k) $\widehat{\textit{u}^{\prime\prime}\textit{u}^{\prime\prime}}$+, (l) $\widehat{v^{\prime\prime}v^{\prime\prime}}$+ and (m) $-\widehat{\textit{u}^{\prime\prime}v^{\prime\prime}}$+ with the Reynolds number.

Figure 15

Figure 15. Variations of the spatially and temporally phase-averaged: (a,c,e) $\widehat{\tilde{p}}_{\textit{rms}}$+, (b,d,f) $\widehat{p''}_{\textit{rms}}$+ and (g) $\hat{d}/\delta$ with deformation phase: (a,b) Reτ = 3300, (c,d) Reτ = 6700 and (e,f) Reτ = 8900. Panels (h,i) display the variations of peak (h) $\tilde{p}_{\textit{rms}}$+, and (i) $p''_{\textit{rms}}$+ with Reynolds number.

Figure 16

Figure 16. Wall-normal profiles of the ensemble-averaged kinetic energy budget terms for: (a) WKE and (b) SKE, both at Reτ = 3300.

Figure 17

Figure 17. Shear production rates: (a,b) ensemble-averaged profiles of (a) WKE, and (b) SKE at the indicated Reynolds numbers; (c,d) distributions of (c) $(-\widehat{\tilde{u}\tilde{v}}\partial \overline{u}/\partial y)^{+}$ and (d) $(-\widehat{u''v''}\partial \overline{u}/\partial y)^{+}$ at Reτ = 3300.

Figure 18

Figure 18. The axial contributor to wave–turbulence energy exchange: (a,b,c) temporally and spatially phase-averaged distributions at (a) Reτ = 3300, (b) Reτ = 6700 and (c) Reτ = 8900. (d) Variations of the peak values with Reynolds number.

Figure 19

Figure 19. Wall-normal profiles of the ensemble-averaged axial wave–turbulence energy exchange term for the three Reynolds numbers, normalized using (a) inner and (b) mixed parameters.

Figure 20

Figure 20. Wall-normal profiles of the ensemble-averaged coherent pressure diffusion term for the three Reynolds numbers, normalized using (a) inner and (b) mixed parameters.

Figure 21

Figure 21. Wall-normal profiles of the WKE at the indicated Reynolds numbers.

Figure 22

Figure 22. Wall-normal profiles of the (a) p–d correlation conditioned on a bump for the present and the Zhang et al. (2017) data, and (b) p–d coherence for the present data, both at the indicated Reynolds numbers. The present results are reproduced from Lu et al. (2024).

Figure 23

Figure 23. Wavelength of the compliant surface response at the indicated frequencies, based on (a) a solution to the Chase (1991) model, and (b) the large field of view experimental data of Wang et al. (2020). The solid grey lines mark the location of λ = 3l0.

Figure 24

Figure 24. Premultiplied energy spectra, $k_{x}{E_{u'u'}}^{+}$, for the full turbulence at (a) Reτ = 3300, (b) Reτ = 6700 and (c) Reτ = 8900. The horizontal dashed lines correspond to λ = 3l0, the characteristic wavelength of the compliant wall deformation.

Figure 25

Figure 25. Contours of $k_{x}{E_{u'u'}}^{+}$ calculated from the Zhang et al. (2017) data at Reτ = 2300. The horizontal dashed line corresponds to λ = 3l0.

Figure 26

Figure 26. Contours of (a,b,c) $k_{x}{E_{u''u''}}^{+}$, and (d,e,f) $k_{x}{E_{\tilde{u}\tilde{u}}}^{+}$ at (a,d) Reτ = 3300, (b,e) Reτ = 6700 and (c,f) Reτ = 8900.

Figure 27

Figure 27. A comparison of the temporal spectra of the present total, wave-coherent and stochastic pressures at y+ = 30 and Reτ = 3300 with the DNS channel flow pressure spectra at the same height and Reτ = 1000. The yellow background marks the k–1 region in the kinetic energy spectra, and the dotted line shows the experimental Nyquist frequency.

Figure 28

Figure 28. Temporal spectra of the total, coherent and stochastic pressures at (a) Reτ = 3300, (b) Reτ = 6700 and (c) Reτ = 8900.

Figure 29

Figure 29. Sample instantaneous snapshots of the original, phase-averaged and Hilbert-projected deformations. The presented region is for x+ = −200 ∼ 750, z+ = −200 ∼ 0 at Reτ = 3300.

Figure 30

Figure 30. Sample instantaneous distributions of (a) $\tilde{p}$ and $\hat{p}$, (b) $\tilde{v}$ and $\hat{v}$, (c) d at z+ = $-100$ and Reτ = 3300.