Hostname: page-component-745bb68f8f-g4j75 Total loading time: 0 Render date: 2025-02-11T19:44:28.522Z Has data issue: false hasContentIssue false

Particle–fluid–wall interaction of inertial spherical particles in a turbulent boundary layer

Published online by Cambridge University Press:  11 December 2020

Lucia J. Baker*
Affiliation:
Department of Aerospace Engineering and Mechanics, University of Minnesota, Minneapolis, MN55455, USA St. Anthony Falls Laboratory, University of Minnesota, Minneapolis, MN55414, USA
Filippo Coletti
Affiliation:
Department of Aerospace Engineering and Mechanics, University of Minnesota, Minneapolis, MN55455, USA St. Anthony Falls Laboratory, University of Minnesota, Minneapolis, MN55414, USA
*
Email address for correspondence: bake0616@umn.edu

Abstract

We study the dynamics of dilute, slightly negatively buoyant, millimetre-size spherical particles fully suspended in a smooth-wall open channel flow. The Reynolds number is $Re_{\tau } = 570$ and the particle Stokes number is ${{\textit {St}}}^{+} = 15$. Particle image velocimetry and tracking are used to obtain simultaneous, time-resolved flow fields and particle trajectories. Particles travel at a lower mean velocity than the fluid: in the log layer this is due to the oversampling of slow fluid regions, but closer to the wall the cause is instantaneous slip between particles and fluid. The particle Reynolds stresses exceed those of the fluid. Near the wall, the particle streamwise diffusivity is larger than the momentum diffusivity, while the opposite is true for the wall-normal component. The particle transport is strongly linked to ejections, while the role of sweeps is marginal, and there is no evidence of turbophoresis. The concentration profile follows a power law with a shallower slope than predicted by equilibrium theories that neglect particle inertia. Upward-/downward-moving particles display positive/negative mean streamwise acceleration due to the particle–fluid slip. The particles that contact the wall are faster than the local fluid both before reaching the wall and after leaving it. Therefore, they are decelerated by drag and pushed downward by shear-induced lift. The durations of wall contact follow exponential distributions with characteristic time scale close to the particle response time. Lift-offs coincide with particles meeting a fluid ejection. These observations emphasize the competing effects of inertia and gravity.

Type
JFM Papers
Copyright
© The Author(s), 2020. Published by Cambridge University Press

1. Introduction

Describing and predicting the behaviour of inertial particles in turbulent boundary layers has been a major goal in fluid dynamics since the work of Shields (Reference Shields1936), Bagnold (Reference Bagnold1936), Rouse (Reference Rouse1937) and Prandtl (Reference Prandtl1952), who first began quantifying the transport of sediment in air and water flows. The experiments of Shields (Reference Shields1936) revealed a relationship predicting the initiation of motion of the sediment from the particle inertia and the fluid shear stress on the wall. Bagnold (Reference Bagnold1936) observed the saltation of sand grains in air and developed an empirical relationship quantifying their flux as a function of the fluid flow parameters. Rouse (Reference Rouse1937) analytically derived an expected concentration profile of suspended sediment as a function of wall-normal height, under equilibrium between turbulent resuspension and gravitational settling. Under the same assumption, Prandtl (Reference Prandtl1952) used a linear diffusivity model to derive a parabolic concentration profile which is still the de facto standard. These models are effective at describing the observed bulk transport properties but do little to shed light on the underlying physical mechanisms at the particle scale.

More recently, thanks to non-intrusive measurement techniques, several researchers used detailed experiments to further understand the interaction between the inertial particles and the wall turbulence. Due to its relevance to sediment transport in water bodies, many studies in the geophysical research literature focused on the case in which the suspension is eroded from and deposited to a bed of particles. For example, Hurther & Lemmin (Reference Hurther and Lemmin2003) investigated the transport of suspended particles in a turbulent open channel flow by comparing statistics of particle mass flux with turbulent momentum flux and concluded that coherent structures are a dominant mechanism of particle transport. Lajeunesse, Malverti & Charru (Reference Lajeunesse, Malverti and Charru2010) used imaging to investigate the relation between the flow turbulence and the intermittent motions of the particles alternating rest and flight. Heyman, Bohorquez & Ancey (Reference Heyman, Bohorquez and Ancey2016) empirically determined closure equations for particle bed load transport in terms of known flow conditions, such as particle diffusivity and suspension/deposition rates.

The presence of a changing wall roughness and the mobilization of a polydisperse bed complicate the task of isolating the particle–fluid dynamics. Somewhat surprisingly, the smooth-wall case has been considered in a limited number of experimental studies. Kaftori, Hetsroni & Banerjee (Reference Kaftori, Hetsroni and Banerjee1995a) measured velocity and concentration profiles of sub-millimetre polystyrene particles in a horizontal water flume and found that particles preferentially concentrate in regions of low fluid velocity associated with near-wall streaks. Niño & Garcia (Reference Niño and Garcia1996) confirmed that particles in a smooth-wall flume arranged themselves in long streaks generated by streamwise vortices in the inner layer. Tanière, Oesterlé & Monnier (Reference Tanière, Oesterlé and Monnier1997) looked at solid particles in a wind tunnel and reported large fluctuations of the dispersed phase velocity, attributing them to saltation at the wall. Kiger & Pan (Reference Kiger and Pan2002) found evidence that particles congregate in specific structures in the turbulent boundary layer, and showed a difference between ascending and descending particles: upward-moving particles were concentrated in ejections (events with negative streamwise fluctuation and positive wall-normal fluctuation of the fluid velocity); whereas downward-moving particles showed a weaker association with sweeps (events with positive streamwise fluctuation and negative wall-normal fluctuation). This was later confirmed by detailed time-resolved measurements, e.g. van Hout (Reference van Hout2011, Reference van Hout2013) and Rabencov, Arca & van Hout (Reference Rabencov, Arca and van Hout2014). Righetti & Romano (Reference Righetti and Romano2004) studied glass particles in water at a volume fraction of $10^{-3}$ and observed significant modulation of the fluid turbulence, which they attributed to the inter-phase momentum exchange during entrainment from and deposition to the wall. Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008) measured the acceleration of inertial particles in a turbulent boundary layer and found that acceleration variance increased with particle inertia, contrary to what happens in isotropic turbulence (Bec et al. Reference Bec, Biferale, Boffetta, Celani, Cencini, Lanotte, Musacchio and Toschi2006). Ebrahimian, Sanders & Ghaemi (Reference Ebrahimian, Sanders and Ghaemi2019) investigated in detail the particle acceleration and its relation to turbulent events and showed that particles may slide along the wall for considerable time. Tee, Barros & Longmire (Reference Tee, Barros and Longmire2020) performed detailed, three-dimensional measurements on a single large spherical particle interacting with the wall in a turbulent boundary layer and found that spheres underwent minimal rotation while lifting off the wall, and that spanwise forces on the particles can be important. Berk & Coletti (Reference Berk and Coletti2020) considered microscopic glass beads in a wind tunnel. They highlighted how the particle inertia is responsible for discrepancies from the concentration profile predicted by the Rouse–Prandtl theory (Rouse Reference Rouse1937; Prandtl Reference Prandtl1952), and investigated the roots of the relative particle–fluid velocity. Due to the limited accuracy in locating the small particles, they could not measure the settling velocity, which is known to be strongly altered by turbulence (Nielsen Reference Nielsen1993; Wang & Maxey Reference Wang and Maxey1993; Sabban & van Hout Reference Sabban and van Hout2011; Petersen, Baker & Coletti Reference Petersen, Baker and Coletti2019).

The numerical studies of inertial particles in smooth-wall turbulence have been much more numerous and have allowed in-depth analysis of the problem, especially using direct numerical simulation of the fluid flow coupled with advection of point particles, e.g. Rouson & Eaton (Reference Rouson and Eaton2001), Marchioli & Soldati (Reference Marchioli and Soldati2002), Zhao, Andersson & Gillissen (Reference Zhao, Andersson and Gillissen2010), Zamansky, Vinkovic & Gorokhovski (Reference Zamansky, Vinkovic and Gorokhovski2011), Sardina et al. (Reference Sardina, Schlatter, Brandt, Picano and Casciola2012), Bernardini (Reference Bernardini2014), Richter & Sullivan (Reference Richter and Sullivan2014) and Lee & Lee (Reference Lee and Lee2015). Typically, a simplified version of the particle transport equation was used, which is only valid in the limit of vanishingly small particle Reynolds number (Maxey & Riley Reference Maxey and Riley1983). Also, in the majority of these cases (included the ones mentioned above) gravity was neglected; this isolates the effect of particle inertia but prevents the direct application of the results to practical settings. The few studies that considered wall-normal gravity (among others, Lavezzo et al. Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010; Lee & Lee Reference Lee and Lee2019) underscored its importance and interplay with the mean shear. In particular, Lee & Lee (Reference Lee and Lee2019) indicated that gravity greatly reduced turbophoresis, i.e. the tendency of the particles to drift down the gradient of Reynolds stresses (hence towards the wall) due to the interaction with streamwise vortices (Marchioli & Soldati Reference Marchioli and Soldati2002).

The limitations of the point-particle approach may be exacerbated in particle-laden turbulent boundary layers where wall-normal gravity is important: the assumption of negligible particle-size effects is particularly limiting near the wall, where the flow scales are the smallest and the number density the highest. Forces which are normally considered negligible for microscopic particles (e.g. added mass, Saffman and Magnus lift; see Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011) may be important, and advanced methods are especially needed to account for the flow distortion by the particles. While effective point-particle strategies to address the latter issue have recently been proposed (Capecelatro & Desjardins Reference Capecelatro and Desjardins2013; Gualtieri et al. Reference Gualtieri, Picano, Sardina and Casciola2015; Horwitz & Mani Reference Horwitz and Mani2016; Ireland & Desjardins Reference Ireland and Desjardins2017; Balachandar, Liu & Lakhote Reference Balachandar, Liu and Lakhote2019), numerical advances and ever-growing computational resources have enabled particle-resolved direct numerical simulations (Kidanemariam et al. Reference Kidanemariam, Chan-Braun, Doychev and Uhlmann2013; Picano, Breugem & Brandt Reference Picano, Breugem and Brandt2015; Lin et al. Reference Lin, Shao, Yu and Wang2017; Wang, Abbas & Climent Reference Wang, Abbas and Climent2017). Still, these studies are mostly limited to small Reynolds numbers and relatively large particles. Importantly, it is hard to validate these simulations against the scarce experimental studies focused on the detailed particle–wall–fluid dynamics, especially considering that most of the experimental literature is concerned with geophysical flows including particle polydispersity, bed roughness and mobile beds.

Here, we consider the fundamental case of highly dilute, mono-dispersed, spherical particles suspended by a turbulent boundary layer over a horizontal smooth wall, where particles interact with the wall but do not deposit on it. We perform laboratory experiments and leverage simultaneous time-resolved imaging of both phases to explore the details of the particle–fluid interaction across the boundary layer. The organization of the paper is as follows: the experimental methods, facility, and image processing are described in § 2; in § 3 we report on the fluid and particle velocity (§ 3.1), particle slip velocity (§ 3.2), Reynolds stresses of both phases (§ 3.3), correlation of particle and fluid fluctuations (§ 3.4), particle diffusivity (§ 3.5), concentration and flux profiles (§ 3.6), acceleration (§ 3.7) and interaction with the wall (§ 3.8); the conclusions are summarized in § 4.

2. Experimental method

2.1. Experimental facility

A recirculating open channel with water as the working fluid is used for this experiment. Complete details of the channel design and its performance to produce fully developed turbulent boundary layers can be found in Adhikari (Reference Adhikari2013) and Baker & Coletti (Reference Baker and Coletti2019) and are just summarized here. The channel walls and floor are made of transparent acrylic. The channel width is 15 cm, with the water filled to a depth $H = 15$ cm. Guide vanes are placed in each of the four corners to reduce secondary flows produced at the turns. The test section is located 1.4 m downstream of a corner, allowing the flow to reach a developed state, which was verified by comparing fluid velocity statistics from the upstream and downstream ends of the test section. A diagram of the channel is shown in figure 1. The flow is provided by a paddlewheel with 16 paddles driven by a 1/4 HP permanent magnet motor (Leeson, USA) at a constant angular speed of 10 revolutions per minute. This is used instead of a centrifugal pump to avoid damaging the particles and the pump. The resulting free-stream velocity is $0.42\ \textrm {m}\ \textrm {s}^{-1}$, which is measured to be constant in time within experimental uncertainty. Two wire screens with a grid spacing of 4 mm and one honeycomb with a cell size of 7 mm and a depth of 25 mm are placed upstream of the test section, as shown in figure 1.

Figure 1. Diagram of the water channel showing key components and dimensions. The bold arrow indicates the direction of the flow.

2.2. Particles

Spherical polystyrene (PS) particles (Composition Materials Co., USA) are used. The particles are transparent, but their index of refraction causes significant scattering of the illumination light, and they appear as shown in figure 2. Because polystyrene is hydrophobic, the particles are first mixed in a dilute solution of water and a surfactant (dish soap) before introducing them into the channel to allow them to disperse.

Figure 2. Instantaneous realization of the particle-laden flow. Both the mm-sized spherical PS particles and the microscopic silver-coated glass tracer particles are visible.

The physical properties of the particles are listed in table 1. The diameter $D_p$ is measured by imaging about 230 particles placed on a tray in a single layer. Their detection and sizing are performed via a circle-finding function based on the Hough transform. The probability density function (PDF) of the particle diameters is plotted in figure 3. For completeness, we also report the value of the Galileo number ${\textit {Ga}} = [(\rho _p/\rho _f - 1)gD_p^{3}/\nu ^{2}]^{1/2}$ and the Shields number ${\textit {Sh}} = u_\tau ^{2}/[(\rho _p/\rho _f - 1)gD_p]$, where $\rho_p$ is the particle density, $\rho_f$ is the fluid density, $g$ is gravitational acceleration, $\nu$ is the fluid kinematic viscosity and $u_\tau$ is the friction velocity (defined in § 3.1). These indicate that the effect of gravity is significant (as also indicated by the ratio $V_t/u_\tau$, of order one), although the particles are in a continuous transport (full suspension) regime.

Table 1. Properties of the PS particles; $D_p$ is the mean particle diameter, $D_p^{+}$ is the diameter normalized by the viscous length scale of the flow, $\rho _p$ is the density, $V_t$ is the terminal velocity in still water, $Re_{p,V_t}$ is the Reynolds number based on $V_t$, $\tau _p$ is the particle response time, ${{\textit {St}}}^{+}$ is the particle Stokes number based on the viscous time scale of the flow, Ga is the Galileo number, Sh is the Shields number and $\varPhi _v$ is the particle volume fraction.

Figure 3. Probability density function of PS particle diameters. The standard deviation is 12 % of the mean value.

The terminal velocity $V_t$ is measured by dropping individual particles from rest in a large tank of quiescent water and recording 60 frame-per-second videos. Particles are tracked using the same method used for the particle-laden flow measurements, which will be described in the next section. The tank is deep enough (0.3 m) for the particles to reach a steady-state velocity before touching the bottom. The nominal particle Reynolds number is computed based on the terminal velocity, $Re_{p,V_t} = \rho _fV_tD_p/\mu$, where $\rho _f$ and $\mu = \rho _f\nu$ are the water density and dynamic viscosity, respectively. This Reynolds number is used to correct the Stokes drag coefficient according to the Schiller & Naumann correction (Clift, Grace & Weber Reference Clift, Grace and Weber2005), which in turn is used to estimate the particle density $\rho _p$ from the measured terminal velocity.

To quantify particle inertia, we refer to the Stokes number, i.e. the ratio between the particle response time and a relevant fluid time scale. As for the particle response time, we consider the characteristic time scale with which the particle exponentially approaches the steady-state velocity of the surrounding fluid, $\tau _p=\rho _pD_p^{2}/(18\mu )$. We favour this definition over the other commonly used response time, $(\rho _p-\rho _f)D_p^{2}/(18\mu )$, which describes the exponential approach to terminal velocity of a particle settling in a still fluid. As for the fluid, both the viscous time scale $\tau ^{+}$ and the Kolmogorov time scale $\tau _\eta$ are relevant; $\tau ^{+}$ is based on the friction velocity $u_\tau$, estimated from fitting the log law to the measured velocity profile (see § 3.1), from which we define ${{\textit {St}}}^{+}=\tau _p/\tau ^{+}$ (the superscript ‘$+$’ denoting, here and in the following, normalization by wall units). The value of $\tau _\eta$ varies with the wall-normal distance and is estimated from the production–dissipation balance in the turbulent boundary layer (Pope Reference Pope2000). This gives a range for the Stokes number ${{\textit {St}}}_\eta =\tau _p/\tau _\eta$ and for the ratio of particle diameter to Kolmogorov length $D_p/\eta$, both reported in figure 4.

Figure 4. Wall-normal profiles of (a) the particle Stokes number based on the Kolmogorov scale and (b) the particle diameter normalized by the Kolmogorov scale.

The volume fraction of the particles in the system, $\varPhi _v$, is approximately $10^{-4}$. Thus, at the present particle-to-fluid density ratio, the momentum two-way coupling effects are expected to be localized and have a minimal impact on the fluid statistics. This is verified in § 3.1 by showing that the unladen and laden fluid velocity profiles overlap within experimental uncertainty.

2.3. Fluid velocity measurements

Time-resolved planar particle image velocimetry (PIV) is used to measure the velocity of the fluid. The water is seeded with 13 micron silver-coated glass bubbles (Potters Industries) to act as tracers. A 300 W near-infrared pulsed laser with a wavelength of 808 nm (Oxford Lasers, Firefly 300W) is used for illumination. The laser is positioned above the channel and emits a 1 mm light sheet perpendicular to the floor and parallel to the streamwise direction, illuminating the channel symmetry plane. A 15 cm square acrylic plate is fixed at the water surface to avoid distortion of the laser sheet. This results a in shear layer below the plate less than 1 cm deep, which does not affect our region of interest. Images are captured with a high-speed, 4-megapixel CMOS camera (Phantom VEO 640L) viewing through one of the sidewalls. The camera mounts a 105 mm lens, capturing the bottom 6 cm of the channel. For optimal tracking, the frame rate is chosen to obtain typical displacements of about one particle diameter (approximately 20 pixels). The recording time amounts to approximately 1900 boundary-layer turnover times.

The image processing routine is similar to what is described in Petersen et al. (Reference Petersen, Baker and Coletti2019). First the PS particles are identified (using the method described in § 2.4) and substituted with Gaussian noise having the same mean and standard deviation as the background image. The resulting tracer-only images are used for PIV processing performed with a custom-written software. A minimum-intensity background subtraction is then performed which removes consistent bright spots caused by reflections and glare off the wall. Multi-pass cross-correlation with an overlap of 75 % between interrogation windows is used to compute fluid displacement fields. Initial, intermediate and final interrogation window sizes of $128^{2}$, $64^{2}$ and $32^{2}$ pixels are used, respectively. A signal-to-noise ratio criterion and a universal outlier detection (Westerweel & Scarano Reference Westerweel and Scarano2005) are used to reject spurious velocity vectors. The imaging and PIV processing parameters are summarized in table 2.

Table 2. Imaging and PIV processing parameters: $f_s$ is the imaging frequency; $N$ is the number of images; $w$ and $h$ are the field of view width and height, respectively; $w_i$ is the final-pass PIV interrogation window size; and $\delta x$ is the PIV vector spacing.

2.4. Particle detection and tracking

To locate the particles, a convolution method using a particle template image is used, similar to van Hout, Sabban & Cohen (Reference van Hout, Sabban and Cohen2013) (figure 5). First, a low-pass median filter with a width of nine pixels is applied to the original images (figure 5a) to remove the tracers (figure 5b). Then, images are convolved with a particle template image (figure 5c). The particle centroids are then identified as convolution peaks which surpass a specified threshold (figure 5d), whose exact value is verified to have negligible impact on the results.

Figure 5. Convolution method for particle detection: (a) original image, (b) median filtered image, (c) particle template image to be convolved with the filtered image and (d) convolution peak, the red cross indicating the detected particle centroid.

The centroids are tracked between successive image pairs using a PIV-based predictor: a first-guess displacement is estimated from the mean fluid velocity profile interpolated at the wall-normal location of each particle centroid and subtracted from the second frame in the pair. Then, a nearest-neighbour search with a search radius of one particle diameter is used to match particle centroids in the first frame with the shifted centroids in the second one. As the inter-frame particle displacement is about one particle diameter, there is no ambiguity in matching particle images. Approximately 2400 particles were tracked within the image set. In the data analysis, we will consider the fluid velocity at the particle location, $u_{f|p}$. This is obtained by interpolating the PIV vectors onto the instantaneous particle centroid location using an inverse-distance-weighted average of the fluid velocity in the $4\times 4$ vector neighbourhood surrounding each particle (a $2\times 2$ neighbourhood is used for $y^{+} < 25$ to account for the greater shear). As the particles have finite size, this definition does not accurately represent an undisturbed fluid velocity at the particle location (as used in the correct definition of the drag force, Horwitz & Mani Reference Horwitz and Mani2016), but it will serve the purpose of investigating the fluid flow events experienced by the particles.

To obtain particle velocities and accelerations, the particle trajectories are convolved with the first and second derivative of a Gaussian kernel, respectively. This method, introduced for fluid tracers (Voth et al. Reference Voth, La Porta, Crawford, Alexander and Bodenschatz2002; Mordant, Crawford & Bodenschatz Reference Mordant, Crawford and Bodenschatz2004), has been used in several studies of inertial particles in turbulence (Gerashchenko et al. Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008; Nemes et al. Reference Nemes, Dasari, Hong, Guala and Coletti2017; Ebrahimian et al. Reference Ebrahimian, Sanders and Ghaemi2019). The optimal width of the kernel $t_k$ is determined from the variance of the particle acceleration magnitude in the data set: the latter is calculated for a range of kernel widths, and the smallest value for which the variance start decaying exponentially is adopted (figure 6). This corresponds to a duration of 17 successive snapshots, or about $12\tau ^{+}$, where $\tau ^{+} = \nu /u_\tau$ is the time scale based on wall units.

Figure 6. Streamwise acceleration variance as a function of smoothing kernel width. The solid line indicates the fit of the acceleration variance over its exponential range, and the optimal kernel width is denoted by the open circle.

2.5. Measurement uncertainty

Uncertainty in the particle statistics is estimated by considering both random uncertainty (due to the finite sample size) and bias uncertainty (due to imperfect centroid locations). The random uncertainty is estimated by computing 90 % confidence intervals on the statistics (Bendat & Piersol Reference Bendat and Piersol2011). The bias uncertainty is estimated by performing detection, tracking and smoothing on a set of synthetic particle images and comparing the position, velocity and acceleration of the smoothed tracks to the known values. The associated uncertainty on the particle location, defined as the root-mean-square (r.m.s.) difference between the actual and calculated values, is found to be approximately 0.04 mm (1 pixel), and therefore random error is the dominant source of uncertainty for particle statistics. To evaluate random uncertainty, we assume a number of independent realizations equal to the number of recorded trajectories. When statistics are computed within wall-normal bins, we assume a number of independent realizations equal to the number of trajectories in each bin.

Uncertainty in the PIV is also dominated by random error. The number of independent samples in the PIV statistics is estimated as the number of temporally independent realizations (i.e. the number of boundary-layer turnover times in the recording) multiplied by the number of spatially independent samples in each realization (i.e. $w/\delta _{99}$, where $\delta _{99}$ is the boundary-layer thickness).

For the fluid velocity evaluated at the particle location, the interpolation also contributes to the uncertainty. This uncertainty is estimated by applying an artificial particle mask to images where the actual velocity vectors are known, performing PIV analysis on the masked images, then interpolating the resulting fluid velocity at the location of the artificial particles. The actual fluid velocity is then compared with the artificially interpolated values. The resulting interpolation error on the fluid velocity, again defined as the r.m.s. difference between the actual and calculated values, is approximately $1\ \textrm {mm}\ \textrm {s}^{-1}$, significantly smaller than the random error. The particles closest to (and in contact with) the wall are detected at $y^{+} \approx D_p^{+}/2 = 8$, requiring an extrapolation from the nearest PIV vectors at $y^{+} = 12$. While this may lead to larger uncertainties, these are not believed to overshadow any of the reported trends.

3. Results and discussion

3.1. Fluid and particle velocity

Here and in the following, the streamwise and wall-normal coordinates are indicated by $x$ and $y$, respectively, and $u$ and $v$ indicate the respective velocity components. These are Reynolds decomposed as $u = \langle u\rangle + u'$ and $v = \langle v\rangle + v'$, where angle brackets denote the time average and the prime denotes the fluctuating part. Subscripts f and p denote quantities referring to fluid and particles, respectively, and the subscript $f|p$ denotes fluid quantities interpolated at the particle location. In figures 7 and 8, the particle-laden and unladen fluid velocity statistics are compared to turbulent boundary-layer measurements by De Graaff & Eaton (Reference De Graaff and Eaton2000) at a similar Reynolds number ($Re_\theta = 1430$). The particle-laden and unladen fluid velocity profiles overlap within experimental uncertainty, indicating that significant two-way momentum coupling between the particles and fluid is not present. From the mean velocity profile, the friction velocity is determined by iterative fitting, with the von Kármán constant $\kappa = 0.41$ and the additive constant $B = 5.5$. The Reynolds stress profiles show some discrepancies in the magnitudes of the peak stresses due to the non-canonical features of the channel design (unconventional forcing, limited channel width) and limited spatial resolution. Physical parameters of the water channel and the boundary-layer properties are reported in table 3.

Figure 7. Mean streamwise velocity profile of the particle-laden (black dots) and unladen (red dots) fluid, in (a) outer units and (b) wall units. The dashed line in (b) indicates the logarithmic-law fit. The profile is compared with De Graaff & Eaton (Reference De Graaff and Eaton2000) in (b), shown in blue open circles.

Figure 8. Profiles of particle-laden (black dots) and unladen (red dots) streamwise turbulent normal stress (a), wall-normal normal stress (b) and shear stress (c) of the fluid in wall units. The profiles are compared with De Graaff & Eaton (Reference De Graaff and Eaton2000) shown in blue open circles.

Table 3. Physical parameters of the water channel and boundary-layer properties. $U_\infty$ is the free-stream velocity, H is the water depth, W is the channel width, $\delta _{99}$ is the boundary-layer thickness and $u_\tau$ is the shear velocity. The boundary thickness is defined such that $u(\delta _{99}) = 0.99U_\infty$; $Re = U_\infty H/\nu$, $Re_\tau = u_\tau \delta _{99}/\nu$ and $Re_\theta = U_\infty \theta /\nu$ are the free-stream, friction and momentum thickness Reynolds numbers, respectively. Standard water properties at $22\,^{\circ }\textrm {C}$ are used in the calculations.

Profiles of particle velocity are obtained by defining wall-normal layers (bins) and taking the mean of particle velocities within each. Particles are more numerous near the wall and sparser in the outer region (approximately following a power law, see § 3.6), thus the bins are logarithmically spaced to equalize the numbers of particles in each, as well as to capture the high shear in the near-wall region. The mean streamwise and wall-normal particle velocity profiles are shown in figure 9. In the free stream, the particle streamwise velocity is very similar to the fluid's (figure 9a), as expected since there the particles are in equilibrium with a steady flow having negligible fluctuations. Closer to the wall ($y^{+} \lesssim 200$) the particles generally lag the fluid, due to their inertia in responding to turbulence. Past experiments found that mean velocity of inertial particles exceeded that of the fluid in the viscous sublayer; see Kaftori et al. (Reference Kaftori, Hetsroni and Banerjee1995a), Righetti & Romano (Reference Righetti and Romano2004) and Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019). The present PIV resolution does not allow reliable measurements at such small heights, but the canonical shape of the boundary-layer profile suggests that the lag is vanishing approaching the wall. We remark that those previous studies considered smaller particles (in wall units) whose centroid could reach closer to the wall.

Figure 9. Wall-normal profiles of streamwise (a) and wall-normal (b) mean particle velocity (red crosses) compared with the mean fluid velocity (black dots).

The vertical velocity profile (figure 9b) shows that in the free stream the particles settle through the fluid at a speed close to the still-fluid terminal velocity. This is again consistent with the fact that particles at those heights fall through a quasi-laminar flow. For $y^{+} \lesssim 200$ (the same range for which particles lag the fluid streamwise velocity), the vertical velocity decays in magnitude, but remains negative. We note that a downward mean particle velocity is expected under equilibrium conditions, i.e. when the wall-normal turbulent flux balances the gravitational settling (Rouse Reference Rouse1937; Prandtl Reference Prandtl1952). Using the concentration measurements (see § 3.6), we estimate the total vertical flux to be several orders of magnitude smaller than the streamwise flux, confirming approximate equilibrium conditions. The Rouse–Prandtl theory, however, assumes a constant settling velocity (generally taken to be equal to the still-fluid terminal velocity) throughout the boundary layer, while here it shrinks to vanishingly small values approaching the wall. We investigate the roots of this effect, as well as the streamwise velocity lag, in the following.

3.2. Slip velocity

The reduced streamwise velocity of the particles has been often attributed to the preferential sampling of slow fluid regions (Kaftori et al. Reference Kaftori, Hetsroni and Banerjee1995a; Kiger & Pan Reference Kiger and Pan2002). We investigate this issue first by separating the mean slip velocity ($\langle u_p\rangle - \langle u_f\rangle$) in two separate contributions: the ‘particle-conditioned’ slip velocity, i.e. the mean slip velocity at the particle location, $\langle u_p - u_{f|p}\rangle$, and the ‘apparent’ slip velocity due to the oversampling of fluid regions faster or slower than the average, $\langle u_{f|p}\rangle - \langle u_f\rangle$ (Kiger & Pan Reference Kiger and Pan2002)

(3.1)\begin{equation} \langle u_p\rangle - \langle u_f\rangle = \langle u_p - u_{f|p}\rangle + \langle u_{f|p}\rangle - \langle u_f\rangle . \end{equation}

Figure 10(a) displays both contributions. (We only report the slip velocity down to the location of the PIV vector closest to the wall, to avoid extrapolation.) It indicates that, for $y^{+} > 20$, the particles do oversample fluid regions with negative streamwise fluctuations ($\langle u_{f|p}\rangle - \langle u_f\rangle < 0$). Closer to the wall, however, the particle-conditioned slip plays a dominant role in determining the particle lag from the fluid: the term $\langle u_p - u_{f|p}\rangle$ becomes larger in magnitude than $\langle u_{f|p}\rangle - \langle u_f\rangle$. This is consistent with the recent findings of Berk & Coletti (Reference Berk and Coletti2020) who considered solid particles in air at much larger $Re_\tau$ and a broad range of ${{\textit {St}}}^{+}$.

Figure 10. Wall-normal profiles of mean streamwise (a) and wall-normal (b) particle slip velocity, separated into the particle-conditioned mean slip (blue circles) and the apparent mean slip (red crosses).

A similar decomposition can be carried out for the vertical velocity component, noting that $\langle v_f\rangle = 0$:

(3.2)\begin{equation} \langle v_p\rangle = \langle v_p - v_{f|p}\rangle + \langle v_{f|p}\rangle, \end{equation}

which highlights the separate contributions of the particle-conditioned slip and the vertical fluid velocity at the particle location. Figure 10(b) shows that the first term on the right-hand side dominates in the free stream and decreases approaching the wall. The second term, representing the preferential sampling of upward/downward fluid fluctuations, is negligible in the free stream and it becomes comparable to the first term as the wall is approached. In particular, particles near the wall oversample upward fluid motions ($\langle v_{f|p}\rangle > 0$). This is consistent with their tendency of favouring negative streamwise fluctuations, which are correlated with upward fluctuations in a turbulent shear flow. This point will be further discussed in § 3.4. In general, throughout the boundary layer, the effect of the turbulence on the particle settling is opposite to what one would expect from homogeneous turbulence studies, where the predominant effect is the enhancement of settling speed by preferential sweeping (Wang & Maxey Reference Wang and Maxey1993; Petersen et al. Reference Petersen, Baker and Coletti2019).

The mean particle-conditioned slip velocity can be used to define profiles of particle Reynolds numbers, $Re_{p,u_{slip}}=\langle u_p-u_{f|p}\rangle D_p/\nu$ and $Re_{p,v_{slip}}=\langle v_p-v_{f|p}\rangle D_p/\nu$, shown in figure 11. Given the observed ranges, the particle wakes are likely to extend for less than one diameter (Rimon & Cheng Reference Rimon and Cheng1969). At the present volume fraction, particles are hardly ever found so close to each other, indicating that the momentum coupling between particles can be assumed to be negligible. However, the Reynolds numbers are well into the nonlinear drag regime, especially in the near-wall region. While drag corrections are available (Clift et al. Reference Clift, Grace and Weber2005), those are developed for steady/quiescent flows. In a turbulent flow laden with particles at finite $Re_p$, the nonlinearities undermine the use of superposition (assumed in the derivation of the particle equation of motion, Maxey & Riley Reference Maxey and Riley1983), making the particle–fluid dynamics challenging to capture with point-particle simulations (Wang et al. Reference Wang, Fong, Coletti, Capecelatro and Richter2019).

Figure 11. Profiles of mean instantaneous particle Reynolds number based on the particle-conditioned streamwise (blue circles) and wall-normal (red crosses) slip velocity.

The instantaneous slip experienced by the particles with respect to the surrounding fluid is related to their ability to retain memory of the flow events experienced at previous times. Thus, one might infer that the gravitational drift plays a major role in determining the streamwise slip velocity, as the settling particles cross flow trajectories and attain large relative velocities with respect to the fluid. This view is certainly valid in homogeneous turbulence (e.g. Csanady Reference Csanady1963; Elghobashi & Truesdell Reference Elghobashi and Truesdell1992). In a turbulent boundary layer, however, the tendency of the particles to sample slow flow regions, combined with the mean wall-normal velocity gradient, leads to a different outcome. Let us consider separately the mean streamwise velocity profile for ascending and descending particles; that is, particles with positive and negative instantaneous vertical velocity, respectively. Figure 12(a) shows that ascending particles, on average, move slower than the fluid and account for most of the mean slip velocity reported above; while descending particles roughly match the mean fluid velocity in the outer layer. This trend, in agreement with Kiger & Pan (Reference Kiger and Pan2002) and van Hout (Reference van Hout2011), is explained by the fact that ascending particles come from slower-moving regions of the flow nearer to the wall, and therefore have lower streamwise velocities. Therefore, unlike in homogeneous flows, it is the ascending particles suspended by the turbulence that determine the large slip velocity, rather than the descending ones that settle due to gravity. Remarkably, the vertical velocity of the ascending particles is comparable in magnitude to that of the descending ones, both being the order of the friction velocity. As the descending particles near the wall are more numerous, their contributions dominate the statistics and the mean vertical velocity of the dispersed phase is negative.

Figure 12. Profiles of mean streamwise (a) and wall-normal (b) particle velocity conditioned on ascending (red crosses) and descending (blue circles) particles compared to the fluid velocity (black dots).

3.3. Reynolds stresses

Profiles of particle and fluid Reynolds stresses are compared in figure 13. The streamwise normal stress of the particles are comparable to that of the fluid, while the particle wall-normal normal stress and shear stress exceed that of the fluid in the range $20 \lesssim y^{+} \lesssim 200$. Qualitatively similar results were reported for particles in water (Kaftori, Hetsroni & Banerjee Reference Kaftori, Hetsroni and Banerjee1995b; van Hout Reference van Hout2011) and in air (e.g. Tanière et al. Reference Tanière, Oesterlé and Monnier1997; Fong, Amili & Coletti Reference Fong, Amili and Coletti2019). The relatively large particle velocity fluctuations are interpreted as a consequence of the spread in momentum of particles with different pathways, retaining memory of their interactions with disparate flow structures. This in turn results in particles often being surrounded by fluid with velocity different from their own, contributing to the instantaneous slip reported above.

Figure 13. Wall-normal profiles of the Reynolds stresses of the particles (red crosses) and fluid (black dots) normalized by the squared shear velocity.

To explore the particle–turbulence interaction, we again condition the fluid statistics on the location of ascending and descending particles. The Reynolds shear stresses are of special interest: they feature in the turbulence production and represent the statistical signature of the instantaneous sweep and ejection events that are believed to play a crucial role in the particle transport (Marchioli & Soldati Reference Marchioli and Soldati2002). Figure 14(a) displays the fluid Reynolds shear stress evaluated at the locations of ascending and descending particles, while figure 14(b) shows the ‘particle Reynolds shear stress’ associated with ascending and descending particles. The ascending particles appear to sample regions with much larger fluid shear stress magnitudes compared to the descending particles, while the shear stress magnitudes of ascending and descending particles themselves are nearly the same. That suggests that the motion of the former is strongly driven by turbulent ejection events, while the latter have a weaker relation to sweep events.

Figure 14. Wall-normal profiles of the Reynolds shear stress of (a) the fluid velocity at particle locations and (b) the particle velocity, conditioned on ascending and descending particles (red crosses and blue circles, respectively). The profiles are compared to the overall fluid shear stress profile (black dots).

3.4. Quadrant analysis

Previous work (e.g. Niño & Garcia Reference Niño and Garcia1996; Kiger & Pan Reference Kiger and Pan2002; Marchioli & Soldati Reference Marchioli and Soldati2002; van Hout Reference van Hout2011) found evidence that inertial particles are strongly affected by sweep and ejection events in the turbulent boundary layer. Sweeps are identified as simultaneous $u_f' > 0$ and $v_f' < 0$ events (fourth quadrant of the $u_f'$$v_f'$ plane, or Q4) and ejections as simultaneous $u_f' < 0$ and $v_f' > 0$ (second quadrant, or Q2). They are often associated with coherent structures in wall turbulence, such as streamwise rollers and hairpin vortices (Robinson Reference Robinson1991). We first plot in figure 15 the joint PDF of streamwise and wall-normal fluid velocity fluctuations: for the near-wall region ($y^{+} < 100$, figure 15a) and farther away from the wall ($y^{+} > 100$, figure 15b). As expected, sweeps and ejections are dominant near the wall, whereas away from the wall the fluctuations are weaker and there are no dominant quadrants. We then consider the quadrant events at the location of ascending and descending particles, in the near-wall and outer regions. Near the wall, ascending particles are found to strongly oversample ejection events (figure 15c), supporting the view that ejections are a major mechanism driving particle resuspension. Even in the outer region, although the ejections themselves are weaker, ascending particles still occupy the second quadrant almost exclusively (figure 15d). In contrast, descending particles are found to oversample sweep events near the wall, but the preference is weaker (figure 15e), and in the outer region, descending particles do not preferentially sample any quadrant (figure 15f). This indicates that sweeps weakly influence the descent of particles towards the wall, while ejections are a key factor in lifting the particles away from the wall.

Figure 15. Joint PDFs of streamwise and wall-normal fluctuating fluid velocities for $y^{+} < 100$ (a) and $y^{+} > 100$ (b). Joint PDFs of streamwise and wall-normal fluctuating fluid velocities at particle locations for $y^{+} < 100$ (c,e) and $y^{+} > 100$ (d,f) conditioned on whether the particle is ascending (c,d) or descending (e,f).

The prevalence of fluid ejection over sweeps in influencing particle transport was reported by previous studies focused on heavy particles suspended in horizontal wall-bounded flows over a wide range of physical parameters (Kiger & Pan Reference Kiger and Pan2002; van Hout Reference van Hout2011; Li et al. Reference Li, Wang, Liu, Chen and Zheng2012; Zhu et al. Reference Zhu, Pan, Wang, Liang and Ji2019; Berk & Coletti Reference Berk and Coletti2020). This is in contrast with configurations in which gravity does not participate to the wall-normal transport: in no-gravity simulations (e.g. Marchioli & Soldati Reference Marchioli and Soldati2002) and in vertical channel flow experiments (e.g. Fong et al. Reference Fong, Amili and Coletti2019) sweep events crucially contribute to the turbophoretic drift that produces a multi-fold increase in near-wall concentration. In horizontal particle-laden flows, by contrast, the near-wall concentration has been found to be smaller than what predicted by the Rouse–Prandtl equilibrium theory (Kiger & Pan Reference Kiger and Pan2002; Zhu et al. Reference Zhu, Pan, Wang, Liang and Ji2019; Berk & Coletti Reference Berk and Coletti2020). In § 3.6 we will show this to be the case also in the present configuration. Because the Rouse–Prandtl theory does not account for turbophoresis, we deduce the latter is not playing a significant role in the particle transport for the present conditions, despite ${{\textit {St}}}^{+}$ being in the turbophoretic regime according to point-particle simulations without gravity (see, e.g. Bernardini Reference Bernardini2014). This could be partly due to the relatively large size of our particles, which influences the ability of inner-layer streamwise vortices to accumulate them at the wall. However, considering the findings of previous studies with much smaller particles (e.g., Berk & Coletti Reference Berk and Coletti2020), the more likely reason is that gravitational drift disrupts the particle interaction with coherent turbulent motions.

3.5. Particle diffusion

The question of dispersion is central in particle-laden flows. A large body of experimental and numerical work in homogeneous turbulence has established that heavy particles disperse differently from tracers due to two distinct and competing effects. Particle inertia increases the integral time scale of their Lagrangian velocity autocorrelation, and hence their diffusivity, due to the finite response time (Squires & Eaton Reference Squires and Eaton1991; Wang & Stock Reference Wang and Stock1993; Jung, Yeo & Lee Reference Jung, Yeo and Lee2008). Meanwhile, particle drift due to gravity or other body forces causes them to cross fluid trajectories with consequent decorrelation of motion and reduction of diffusivity compared to tracers (Csanady Reference Csanady1963; Squires & Eaton Reference Squires and Eaton1991; Elghobashi & Truesdell Reference Elghobashi and Truesdell1992; Wang & Stock Reference Wang and Stock1993). In wall-bounded flows, Lagrangian stochastic models have been proposed in order to predict dispersion of inertial particles (Tanière & Arcen Reference Tanière and Arcen2016; Marchioli Reference Marchioli2017) but have been mostly tested against point-particle numerical simulations, usually without gravity.

To address the issue of dispersion, we first consider the temporal coherence of particle motion by computing Lagrangian autocorrelations of the particle velocity (in these definitions we only refer to $u_p$ for brevity, but all definitions apply to $v_p$ as well), given by

(3.3)\begin{equation} \rho_{u_p}({\rm \Delta} t, y_0) = \frac{\langle u_p'(t_0, y_0) u_p'(t_0+{\rm \Delta} t, y_0) \rangle}{\langle u_p'(t_0, y_0) u_p'(t_0, y_0)\rangle} . \end{equation}

The subscript ‘0’ denotes the origin of a trajectory, so that $t_0$ and $y_0$ are the initial time and wall-normal location of each trajectory, respectively. Here, the fluctuating velocities, $u_p'$, are determined by subtracting the Lagrangian mean velocity from each trajectory, $\langle u_p({\rm \Delta} t, y_0)\rangle _L$, as follows:

(3.4)\begin{equation} u_p'(t_0+{\rm \Delta} t, y_0) = u_p(t_0+{\rm \Delta} t, y_0) - \langle u_p({\rm \Delta} t, y_0)\rangle_L . \end{equation}

The Lagrangian autocorrelation is computed within five logarithmically spaced wall-normal bins, such that each contains a comparable number of samples. The autocorrelations of streamwise and wall-normal particle velocities are plotted in figures 16(a) and 16(b), respectively. They both drop off more steeply near the wall, and the streamwise particle velocity remains correlated over a longer length of time than the wall-normal velocity. These trends are consistent with results for fluid tracers in channel flow simulations (Choi, Yeo & Lee Reference Choi, Yeo and Lee2004). They are attributed to the smaller flow scales affecting the particle motion near the wall, and the streamwise-elongated structures that characterize the boundary layer, contributing to the turbulence anisotropy.

Figure 16. Lagrangian autocorrelations of streamwise (a) and wall-normal (b) particle velocity for five wall-normal bins (dots) shown with their respective exponential fits (dashed lines). The wall-normal locations listed in the legend correspond to the centre of each bin.

The integral time scale of the particle motions in streamwise and wall-normal directions can be defined as $\tau _{L,x} = \int _0^{\infty } \rho _{u_p}({\rm \Delta} t)\,\mathrm {d}t$ and $\tau _{L,y} = \int _0^{\infty } \rho _{v_p}({\rm \Delta} t)\,\mathrm {d}t$, respectively. Since in practice the integral can only extend to finite values, and recognizing that $\rho _{u_p}$ and $\rho _{v_p}$ approximately follow an exponential decay, we fit an exponential function to the autocorrelations and consider the time lag that results in an e-fold drop of the exponential fit. Applying the theory of Taylor (Reference Taylor1921) on the Lagrangian statistics of particle displacements, we evaluate the long-time particle diffusivities in both directions, $\varepsilon _{p,x} = \tau _{L,x}\langle u_p^{'2}\rangle$ and $\varepsilon _{p,y} = \tau _{L,y}\langle v_p^{'2}\rangle$. In figure 17 this is shown for the five wall-normal bins and for both velocity components, computed using the variance of the particle velocity within the respective bins. For comparison, we also plot the classic estimate for the fluid momentum diffusivity in the log-law region, $\varepsilon _f = \kappa yu_\tau$ (Prandtl Reference Prandtl1952), and that for the defect layer ($y > 0.2\delta _{99}$), estimated as $\varepsilon _f = 0.09\delta _{99}u_\tau$ (Pope Reference Pope2000). In the log-law region, the turbulence causes the particles to disperse much faster in the streamwise direction, greatly exceeding the momentum diffusivity. This indicates that the effect of particle inertia (which increases particle dispersion) is dominating over the effect of gravitational drift (which reduces it), at least in what pertains streamwise dispersion. This agrees with theoretical arguments of Reeks (Reference Reeks1977) which predicted particles to disperse faster than tracers when the settling velocity $V_s < \langle u_f^{'2}\rangle ^{1/2}$, as it is the case here. Laboratory observations had confirmed this in homogeneous turbulence (Wells & Stock Reference Wells and Stock1983; Sabban & van Hout Reference Sabban and van Hout2011), and to our knowledge the present results are the first experimental observation of this effect in wall turbulence. On the other hand, $\varepsilon _{p,y}$ is equal to or smaller than the momentum diffusivity across the boundary layer, indicating that, in the vertical direction, the effect of gravity in decorrelating the particle motion slightly dominates.

Figure 17. Wall-normal profiles of streamwise and wall-normal diffusivity (blue circles and red crosses, respectively) compared to the theoretical profile of fluid momentum diffusivity (black dashed line).

3.6. Particle concentration and flux

Mean particle relative concentration as a function of wall-normal distance is plotted in figure 18. This is obtained by counting particles within logarithmically spaced wall-normal bins and normalizing by the mean concentration in the lowest bin, $C_0$. The observed power-law behaviour prompts a comparison with the concentration profile predicted by the theory of Rouse (Reference Rouse1937) and Prandtl (Reference Prandtl1952). This follows from the balance between gravitational settling and wall-normal turbulent flux

(3.5)\begin{equation} \langle C\rangle V_s - \varepsilon \frac{\partial\langle C\rangle}{\partial y} = \varPhi, \end{equation}

where $V_s$ is the particle settling velocity and $\varPhi$ is the net wall-normal flux of particles. Assuming equilibrium conditions ($\varPhi = 0$), the particles falling at $V_s = V_t$ and having the same diffusivity as the momentum in the turbulent boundary layer ($\varepsilon = \kappa yu_\tau$), leads to the well-known concentration profile (Prandtl Reference Prandtl1952)

(3.6)\begin{equation} \frac{\langle C\rangle}{\langle C\rangle_{ref}} = \left(\frac{y}{y_{ref}} \right)^{-{{\textit{Ro}}}} , \end{equation}

where the subscript denotes an arbitrary reference height and the corresponding concentration. Here, ${{\textit {Ro}}}=-V_t/(\kappa u_\tau )$ is the Rouse number, which quantifies the relative strength of gravitational settling and turbulent resuspension of the particles. Equation (3.6) is also plotted in figure 18 for comparison, which shows a much steeper drop in concentration with height than the measurements. A departure from Rouse–Prandtl theory is expected, notably because the latter does not account for particle inertia. In particular, Berk & Coletti (Reference Berk and Coletti2020) recently carried out a wind tunnel study of particle transport in turbulent boundary layers, and also reported a reduced slope of the concentration profile compared to the Rouse–Prandtl theory for a wide range of Stokes numbers. They hypothesized this to be due to a near-wall settling rate below the terminal velocity but could not accurately measure the particle vertical velocity. The present measurements corroborate their hypothesis (see figure 9b).

Figure 18. Wall-normal profile of mean particle concentration normalized by the concentration at the lowest wall-normal bin (black crosses). The power-law profile predicted by Rouse–Prandtl theory (red dashed line) is calculated from (3.6), where the arbitrary reference height is taken at $y_r^{+} = 90$.

The particle streamwise mass flux $Q_x$ is often of interest, especially in geophysical flows. Assuming advection dominates on the turbulent transport, the mean flux can be approximated from the mean concentration and mean velocity profiles, i.e. $\langle Q_x\rangle \approx \langle C\rangle \langle u_p\rangle$. Here, we compute the flux directly by counting particles crossing wall-normal planes, and verify that it does not vary with streamwise location within the imaging window, and that it is indistinguishable from the mean advective flux $\langle C\rangle \langle u_p\rangle$. The profile in figure 19, albeit with experimental scatter, suggests a power-law behaviour. In sediment transport and aeolian transport studies, the flux is often observed to decay exponentially with wall-normal height, thus identifying a characteristic length scale (e.g. Bagnold Reference Bagnold1941; Nishimura & Hunt Reference Nishimura and Hunt2000; Guala et al. Reference Guala, Manes, Clifton and Lehning2008; Kok et al. Reference Kok, Parteli, Michaels and Karam2012). However, those processes are inherently different from the present one: they are characterized by beds of particles mobilized by the impact of other particles, with their transport largely concentrated in a ‘saltation layer’. The present case instead is governed by suspension, and as such it does not possess a specific length scale beyond those associated with the fluid turbulence.

Figure 19. Wall-normal profile of particle streamwise mass flux normalized by the mass flux at the lowest wall-normal bin (black crosses), compared with a power-law fit (red dashed line).

3.7. Particle acceleration

The statistics of the particle acceleration provide insight not only on the kinematics but also on the forces at play. In figure 20(a) we present the profiles of mean and r.m.s. acceleration for both in-plane components. The mean acceleration profiles are in close agreement with the recent experiments of Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019), who considered particles with ${{\textit {St}}}^{+} = 3.9$ and $D_p^{+} = 6.8$ in a channel flow at $Re_\tau = 410$. The mean acceleration is negligible in the free stream, as expected in a fully developed flow. The streamwise acceleration increases to a positive maximum at $y^{+} \sim 50$, and then rapidly decays to become negative near the wall. When expressed in wall units, the negative mean streamwise acceleration we observe near the wall is significantly smaller in magnitude compared with the experiments of Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008). These authors considered microscopic water droplets, and therefore the discrepancy is likely rooted in the larger size of our particles compared to the viscous length scale, and the consequent effect of forces other than drag and gravity (Maxey & Riley Reference Maxey and Riley1983). The difference could also be related to particle interactions with the wall (addressed in § 3.8), which were likely inconsequential for the statistics of Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008).

Figure 20. (a) Profiles of mean streamwise and wall-normal particle acceleration. (b) Profiles of mean streamwise and wall-normal particle acceleration conditioned on whether particles are ascending (filled symbols) or descending (open symbols).

The mean wall-normal acceleration shows a qualitatively opposite trend compared to the streamwise component. At first, this would appear consistent with sweep and ejection being at the root of the acceleration profiles: upward vertical fluid fluctuations lifting the particles away from the wall also tend to decelerate them in streamwise direction, and vice versa for downward fluctuations. This view, however, does not consider the fact that the mean velocity of the particles is downward: thus, negative wall-normal accelerations are mostly associated with a decrease in magnitude of the negative velocity, i.e. hindering of the settling. We shall come back to this point shortly, when considering the fluid events associated with these particle statistics.

In Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008) and Lavezzo et al. (Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010) it was argued that the gravitational settling of particles through the shear flow plays a crucial role in determining the streamwise acceleration. To explore this issue, we condition again the statistics on ascending and descending particles (figure 20b). It becomes apparent that the regions of positive and negative streamwise acceleration can be attributed to the ascending and descending particles, respectively. This is in line with arguments presented by Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008) and Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019): particles that fall vertically find themselves surrounded by slower fluid, which decelerates them; and vice versa for particles moving vertically upward. The mean vertical acceleration is instead unaffected by conditioning on ascending/descending particles.

We present profiles of r.m.s. particle acceleration in figure 21. The streamwise component is quantitatively close to what is reported by Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008), Zamansky et al. (Reference Zamansky, Vinkovic and Gorokhovski2011) and Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019). The wall-normal component instead reaches a peak 30 %–60 % larger than in those studies, being comparable to the streamwise component. We believe this to be mainly caused by gravitational effects. Already Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008) found that the particle r.m.s. acceleration of inertial particles was larger than that of tracer particles, in contrast with known trends in homogeneous turbulence (Bec et al. Reference Bec, Biferale, Boffetta, Celani, Cencini, Lanotte, Musacchio and Toschi2006; Ayyalasomayajula et al. Reference Ayyalasomayajula, Gylfason, Collins, Bodenschatz and Warhaft2006). Our particles are more inertial than those considered by Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008) (who had maximum ${{\textit {St}}}^{+} = 5.3$) and have larger settling velocity ($V_t/u_\tau = 0.75$, versus $V_t/u_\tau = 0.38$ in that study). In laboratory experiments it is not possible to conclusively disentangle inertia from gravity to determine what drives the higher vertical r.m.s. acceleration, but we can find hints in previous numerical studies where gravity could be suppressed. Lavezzo et al. (Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010) demonstrated that gravity was the cause of increased r.m.s. acceleration in streamwise direction. This appears to be the case also for the wall-normal component, enhanced because the falling particles encounter rapidly changing flow structures. In further support of this view, the zero-gravity simulations of Zamansky et al. (Reference Zamansky, Vinkovic and Gorokhovski2011) found a decrease of both components of r.m.s. acceleration with increasing ${{\textit {St}}}^{+}$, while the recent simulations of Lee & Lee (Reference Lee and Lee2019) that included gravity showed the opposite trend, in agreement with Gerashchenko et al. (Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008). Taken together, the above indicates that it is gravitational drift that causes heavy particles to have large r.m.s. accelerations, in both streamwise and wall-normal directions.

Figure 21. Wall-normal profiles of streamwise (blue circles) and wall-normal (red crosses) particle r.m.s. acceleration.

The acceleration of fluid tracers and inertial particles in turbulence are known to be intermittent, with super-Gaussian probabilities of high-acceleration events that strongly depend on the particle inertia. When gravity is absent or has negligible influence, it is established that increasing Stokes number leads to reduced intermittency (Bec et al. Reference Bec, Biferale, Boffetta, Celani, Cencini, Lanotte, Musacchio and Toschi2006; Gerashchenko et al. Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008; Zamansky et al. Reference Zamansky, Vinkovic and Gorokhovski2011). We present the PDF of streamwise and wall-normal accelerations in figure 22. We plot separately the distributions for particles below and above $y^{+} = 100$, which helps in better understanding the observed behaviour. The stretched exponential tails indicate significant intermittency, with relatively large probability of extreme events. This is, however, much stronger for the particles far from the wall than for those close to it: the flatness of the distributions for $y^{+} < 100$ and $y^{+} > 100$ is 5.5 and 20.5 for the streamwise component, and 6.4 and 10.4 for the wall-normal component, respectively. This is because the particles closer to the wall interact with turbulent eddies of smaller time scales, and therefore have an effectively larger Stokes number (see figure 4a), which in turn causes them to ‘filter out’ intense fluid flow events (Bec et al. Reference Bec, Biferale, Boffetta, Celani, Cencini, Lanotte, Musacchio and Toschi2006). The crossing-trajectory effect, on the other hand, is weaker for the near-wall particles, which have small mean vertical velocity (figure 9b). Combined with the results for the r.m.s. accelerations, these results underscore the competing influence of gravity and inertia: the first enhances the variance and intermittency of the particle accelerations, while the second damps them, the net result depending on the relative importance of both effects.

Figure 22. PDFs of streamwise (blue filled symbols) and wall-normal (red open symbols) particle acceleration conditioned on $y^{+} < 100$ (circles) and $y^{+} > 100$ (triangles). The PDFs for $y^{+} > 100$ are shifted upward by a factor of $10^{3}$ for clarity.

The particle trajectories are influenced not only by the instantaneous values of their accelerations, but also by their temporal coherence. This is explored by calculating the Lagrangian autocorrelation of streamwise particle accelerations, defined analogously to the velocity autocorrelations and presented in figure 23 for various wall distances. The wall-normal acceleration (not shown) follows similar trends, but less clearly so due to experimental uncertainty on measuring quantities of smaller magnitude. Close to the wall ($y^{+} \lesssim 100$), the acceleration has significant temporal coherence, with the autocorrelation showing e-folding time of the order of the response time of the particles, before decaying monotonically to zero. This suggests that near-wall particles move in and out of streamwise-coherent turbulent structures and behave as if responding to step changes in the surrounding fluid velocity. We will show in § 3.8 that this picture is consistent with the behaviour of the particles that directly interact with the wall. At larger heights, the temporal coherence is greatly reduced and the oscillations around zero suggest a quick alternation of positive and negative accelerations. This is likely caused by settling with significant speed (close to the particle terminal velocity for $y^{+} \gtrsim 100$, see figure 9b), which causes the particles to move in a less coherent fashion. This is in line with the above-mentioned drop in streamwise diffusivity with increasing distance from the wall (figure 17).

Figure 23. Lagrangian autocorrelation of streamwise particle acceleration for five wall-normal bins. The wall-normal locations listed in the legend correspond to the centre of each bin.

3.8. Particle–wall interactions

In order to analyse the direct interactions with the wall, we consider the trajectories of particles that come in contact with it. An example of one such trajectory is shown in figure 24. Actual physical contact cannot be ascertained by the present imaging, and a lubrication layer is possibly maintained during those events. Still, we will use the word ‘contact’ to indicate the instances with no measurable wall–particle separation. More precisely, considering the variance of the particle diameters and the uncertainty in locating both the particles and the wall, we record contact when particle centroids are within 1.3 mean particle radii from the wall, i.e. when their wall-normal height is $y^{+} \leqslant 10.4$. The results below are robust to small modifications of this threshold. We define ‘touch-down’ and ‘lift-off’ events when this threshold is crossed by particles approaching and leaving the wall, respectively. In figure 24, touch-down occurs at about $\tau ^{+} = 15$ and lift-off at about $\tau ^{+} = 40$; in between, the particle is considered to be in continued contact with the wall. Below we present results from averaging 241 trajectories leading to a touch-down and 151 trajectories following a lift-off.

Figure 24. An example of an interaction between a particle and the wall. The blue curve shows the wall-normal distance of the particle centroid vs. time, with error bars representing the uncertainty in particle position. The dashed line represents a height of one particle radius above the wall, with the grey shaded region indicating the uncertainty in the particle radius and the wall location.

First, particle–wall interactions are characterized in terms of duration of contact, $t_{wall}$. Figure 25 presents the PDF of $t_{wall}$ in viscous units, which is well described by an exponential distribution. This is reminiscent of the distribution of ‘waiting times’ between touch-down and lift-off of the particles mobilized by the flow over a sediment bed (Einstein Reference Einstein1950; Ancey et al. Reference Ancey, Böhm, Jodeau and Frey2006; Fan et al. Reference Fan, Singh, Guala, Foufoula-Georgiou and Wu2016). In the case of bedload transport, an exponential distribution follows from assuming that touch-down and lift-off are time- and space-invariant Markovian processes (Ancey et al. Reference Ancey, Böhm, Jodeau and Frey2006), i.e. independent events that are influenced only by the present and local state. Accordingly, waiting times have often been modelled as induced by a random (Poissonian) distribution of turbulent fluctuations having sufficient strength to mobilize the particles (e.g. Papanicolaou et al. Reference Papanicolaou, Diplas, Evaggelopoulos and Fotopoulos2002). There are obvious differences between bedload transport and the present case: most notably, our particles are never at rest, as the wall shear stress exerted by the flow is much larger than the critical value for mobilizing them (Shields Reference Shields1936). Still, our distribution of contact times suggests these can be modelled by a similar stochastic process driven by randomly occurring fluid flow events above a certain threshold. We note that Cameron, Nikora & Witz (Reference Cameron, Nikora and Witz2020) recently showed the importance of very-large-scale motions (non-random, spatially correlated turbulent events) for sediment entrainment. While this points to the limitations of considering turbulent fluctuations to be random, their study was conducted at much higher Reynolds numbers for which very-large-scale motions are more prominent than in our case (Smits, McKeon & Marusic Reference Smits, McKeon and Marusic2011).

Figure 25. PDF of duration of particle–wall interactions, defined as the length of time particles spend continuously within 1.3 mean particle radii of the wall.

The exponential distribution of $t_{wall}$ suggests a characteristic time $\tau _{wall}$, such that its duration probability can be approximated as $\mathrm {PDF}(t_{wall}) \propto \exp ({-t_{wall}}/{\tau _{wall}} )$. A least-squares fit returns $\tau _{wall} = 13.5\tau ^{+}$, which is remarkably close to the particle response time $\tau _p/\tau ^{+} = {{\textit {St}}}^{+} = 15$. This suggests that both particle inertia and the fluid fluctuations play a role in determining the duration of the wall contact, and one may hypothesize a scaling $\tau _{wall}^{+} \sim {{\textit {St}}}^{+}$. This is indeed consistent with the results reported by Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019), whose particles had ${{\textit {St}}}^{+} = 3.9$ and a mean wall contact time around $4\tau ^{+}$. Because that is the only previous study reporting contact times for suspended particles, further research is warranted to corroborate this ansatz. We note that, in sediment transport, the waiting time was originally suggested by Einstein (Reference Einstein1950) to be inversely proportional to the particle settling velocity, but it was later recognized that an inverse proportionality with fluid velocity showed better agreement with observations, and with the view that the particle entrainment rate depends on the strength of the turbulent fluctuations (Ancey et al. Reference Ancey, Böhm, Jodeau and Frey2006, Reference Ancey, Davison, Böhm, Jodeau and Frey2008).

As the particles are never at rest, one may consider the possibility that the short contact times are associated with a rebound process. However, even assuming an incident velocity $V_i$ equal to the maximum vertical velocity in the $20\tau ^{+}$ preceding touch-down (see below), the impact Stokes number $\rho _pV_iD_p/(9\mu )$ is of order unity – well below the threshold for a non-zero restitution coefficient (Joseph et al. Reference Joseph, Zenit, Hunt and Rosenwinkel2001; Gondret, Lance & Petit Reference Gondret, Lance and Petit2002). Thus, in the present regime, the kinetic energy of the particle associated with its wall-normal motion is expected to be dissipated at contact, with no effective rebound.

The particle–wall interactions can also be characterized by the angle $\theta _p$ of near-wall particle trajectories as they approach or recede from the wall. The trajectory angle is defined as $\theta _p = \mathrm {atan}(v_p/u_p)$, such that $\theta _p$ is negative for particles approaching the wall and positive for particles receding from it. The set of particles within one particle diameter of the wall ($y^{+} < 16$) is considered. The PDF of $\theta _p$ is shown in figure 26; it displays stretched tails (with a flatness of 6.7), and measurable preference for positive values (with a skewness of 0.65). The distribution of angles is narrower than what reported by Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019). This is attributed to the larger inertia of the present particles (${{\textit {St}}}^{+} = 15$ versus 3.9), resulting in a slower response time to fluid velocity fluctuations and thus shallower near-wall trajectory angles. The positive skewness indicates that, despite gravity, relatively sharp lift-offs may occur.

Figure 26. PDF of the angle of trajectories for near-wall ($y^{+} < D_p^{+}$) particles.

The mechanisms behind particle touch-down and lift-off can be further understood by considering the Lagrangian averages along the particle trajectories. To this end, we first identify the moments of touch-down and lift-off (within the temporal resolution of $1/500\ \textrm {s} = 0.7\tau ^{+}$), and then average over all trajectories leading to and following those events, respectively. Figure 27 displays the average particle wall-normal distance, streamwise velocity and vertical velocity during the $20\tau ^{+}$ before touch-down and after lift-off. We do not report Lagrangian statistics during particle–wall contact, as the range of contact durations thwarts a consistent averaging process. However, we note that the wall-normal positions, velocities and accelerations at touch-down are very close to those at lift-off.

Figure 27. Lagrangian means of particle wall-normal coordinate (a,b), streamwise velocity (c,d) and wall-normal velocity (e,f) averaged over all identified touch-down events (a,c,e) and lift-off events (b,d,f). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.

The Lagrangian averages of the wall-normal distance indicate that the particles descending to the wall follow a somewhat steeper trajectory than those ascending from it (figures 27a and 27b). This is confirmed by the plots of the vertical velocities (figures 27e and 27f) and is due to a combination of gravity and Saffman lift (which is directed downward, as demonstrated below). The streamwise velocity declines before touch-down (figure 27c), because the particles approaching the wall are surrounded by slower and slower fluid, which applies on them drag in the negative $x$-direction. After lift-off, although the particles ascend at a similar pace as they descended, they regain streamwise velocity much more slowly (figure 27d). This fits the view of particles being lifted from the wall by ejections, i.e. events with smaller streamwise fluid momentum compared to the local mean.

The Lagrangian average of the fluid velocity experienced by the particles along their trajectory helps in the interpretation of the above trends. Figures 28(a) and 28(b) show that, before and after wall contact, the streamwise velocity of the fluid surrounding the particles is significantly lower than the local particle velocity (see figures 27c and 27d). (This is not captured by the unconditioned mean profiles presented §§ 3.1 and 3.2. We remark that the particles coming in contact with the wall are a small fraction of the total in the corresponding bin.) The direction of the streamwise slip velocity implies that, close to wall contact, drag acts against the direction of motion and Saffman lift is directed downward (in consideration of the slip velocity and the fluid velocity gradient, Saffman Reference Saffman1965). The streamwise fluid velocity along the trajectory is also significantly lower than the local mean fluid velocity (assuming a canonical boundary-layer profile, see figure 7); therefore, the particles are in regions of negative streamwise fluctuations. Because the vertical fluid velocity is negative before touch-down (figure 28c) and positive after lift-off (figure 28d), particles approaching and leaving the wall experience Q3 events and Q2 events, respectively. This confirms once more that ejections (in Q2) are critical to lift particles from the wall, while the role of sweeps (in Q4) is not dominant near wall contact. Importantly, figure 28(d) indicates that the vertical fluid velocity becomes positive just as the particles leave the wall: that is, lift-off happens when the particle starts to experience a fluid ejection. This is in line with the view that the duration of the contact time is dictated by the occurring of turbulent events.

Figure 28. Lagrangian means of streamwise (a,b) and wall-normal (c,d) fluid velocity at particle location averaged over all identified touch-down events (a,c) and lift-off events (b,d). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.

A material frame of reference attached to the particles is the most appropriate to investigate the forces acting on them, and thus we now consider the accelerations tangential and normal to the local trajectory, $a_{t,p}$ and $a_{n,p}$, respectively

(3.7)\begin{gather} a_{t,p} = a_{x,p}\cos (\theta_p) + a_{y,p}\sin (\theta_p) \end{gather}
(3.8)\begin{gather}a_{n,p} = a_{y,p}\cos (\theta_p) - a_{x,p}\sin (\theta_p). \end{gather}

Note that since $\theta _p$ is relatively small, the tangential and normal accelerations are very similar in magnitude to the horizontal and vertical components, respectively. We can make inferences on the role and respective importance of drag, gravity and Saffman lift. Other forces are expected to be important as well, such as added mass, fluid acceleration, Basset history and Faxén corrections and rotation-induced (Magnus) lift (Maxey & Riley Reference Maxey and Riley1983; Crowe et al. Reference Crowe, Schwarzkopf, Sommerfeld and Tsuji2011; Mathai et al. Reference Mathai, Loeffen, Chan and Wildeman2019); however, it is not trivial to infer their behaviour from experimental data, so our qualitative considerations will not focus on these forces.

The Lagrangian averages of the accelerations before and after wall contact are plotted in figure 29. As the particles approach the wall before touch-down, they experience increasingly negative tangential acceleration (figure 29a), reflecting the drag force opposing their motion. Notice that, while we are outside the assumptions of viscous steady flow made to derive analytical results, we expect the wall proximity to enhance drag for a particle moving parallel to a solid boundary (Brenner Reference Brenner1965). The negative streamwise acceleration peaks around $5\tau ^{+}$ before touch-down and then decreases in magnitude. This is possibly due to lubrication effects, as the average particle is then only a fraction of its diameter away from the wall. Higher resolution measurements are needed to draw firm conclusions in this regard.

Figure 29. Lagrangian means of particle tangential acceleration (a,b) and normal acceleration (c,d) averaged over all identified touch-down events (a,c) and lift-off events (b,d). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.

As the particles leave the wall and reach faster fluid layers, they at first experience negative streamwise acceleration, but this shrinks in magnitude and eventually reaches a small positive plateau (figure 29b). This follows from the particles being entrained into faster fluid strata, which reduces the negative drag. However, because the particles are still faster than the surrounding fluid (see figures 27d and 28b), the drag is expected to remain negative. Thus, the positive streamwise acceleration must be a consequence of other forces. Added mass and fluid acceleration force are prime candidates, because the acceleration of the fluid along the particle trajectory is positive and larger than that of the particle (compare again figures 27d and 28b). Further studies to quantify these forces are warranted.

The normal acceleration becomes positive before touch-down (figure 29c) since the downward particle velocity is decreasing in magnitude as the wall is approached. This is qualitatively consistent with the increase in drag predicted in viscous flows when a particle travels towards a wall (Brenner Reference Brenner1965). The interaction with a lubrication layer may also contribute to slowing down the descent. The downward-directed Saffman lift increases the magnitude of the downward particle velocity before touch-down and decreases the magnitude of the upward velocity after lift-off. Thus, Saffman lift (along with gravity) is the likely cause of the drop in positive normal acceleration after the particles has left the wall (figure 29d). Simulations of Lee & Balachandar (Reference Lee and Balachandar2010) suggested the Magnus force to be relatively small, but the estimates of Ebrahimian et al. (Reference Ebrahimian, Sanders and Ghaemi2019) suggested it could be sufficient to overcome gravity. Here, the magnitude of the Magnus force cannot be estimated without knowing whether the particles are actually in contact with the wall, as this would dictate their rolling motion.

4. Conclusions

We have investigated the transport of inertial particles suspended in a turbulent boundary layer through simultaneous fluid and particle velocity measurements. While the considered regime is especially relevant to geophysical water flows, we have taken a fundamental viewpoint and focused on the seemingly simple case of monodispersed spherical particles not depositing over the smooth wall. Leveraging the time-resolved, two-phase nature of the measurements, we documented in detail the particle–fluid dynamics and delved in several central aspects of this archetypical configuration. Given the richness of the flow physics, we chose to limit our attention to one specific regime, leaving parametric studies for the future.

We confirm that the particles travel at a lower mean velocity than the fluid, at least above the viscous sublayer. In the past, this has largely been attributed to the oversampling of slow velocity regions. While this is found to be the dominant factor in the logarithmic layer, closer to the wall the main cause for the mean velocity defect is the instantaneous particle slip from the surrounding fluid. Indeed, the expectation that particles would have a small instantaneous slip velocity is only tenable for small particle inertia and/or slow turbulence fluctuation, i.e. for small Stokes numbers. In the present regime (${{\textit {St}}}^{+} = 15$), a significant portion of the fluid fluctuations experienced by the near-wall particles have shorter time scales than $\tau _p$, and the inability of the particles to respond to them naturally translates in a sizeable slip velocity. By inspecting the profiles conditioned on ascending and descending particles, it appears that the mean velocity lag is due to the ascending ones, thus it is not directly related to gravity. The mean vertical velocity of the particles is close to the quiescent-fluid terminal velocity in the free stream, but it becomes vanishingly small approaching the wall. This is an effect of oversampling upward turbulent fluctuations, which is opposite to the behaviour in homogeneous turbulence.

The near-wall particle diffusivity in the streamwise direction is larger than the momentum eddy diffusivity, while the opposite is true in the wall-normal direction. This is one of the several instances in which the distinct (and competing) effects of gravity and inertia are on display. The dichotomy is highlighted here both by conditioning the statistics on ascending and descending particles, and by comparing outer and inner layers. This helps reconcile contrasting results in the literature and underscores the importance of both effects in realistic situations.

The near-wall particle transport is strongly linked to ejection events that promote resuspension, while the link with sweep events is weak. This suggests that turbophoresis, which is often considered a dominant feature according to studies where gravity is either absent or acting parallel to the wall, is at best a second-order effect when gravity is wall normal. Under equilibrium between settling and turbulent suspension, the concentration profile follows a power law with a much more gradual decrease from the wall than predicted by the Rouse–Prandtl theory (which neglects particle inertia), consistent with the observed reduction of settling velocity. While the particles in the inner layer do interact with the wall, the absence of bedload is reflected in the lack of a saltation layer in the mass flux profile.

Separating ascending and descending particles also helps us understand the root of the mean acceleration profile. Regions of positive and negative streamwise acceleration can be attributed to ascending and descending particles, respectively. Ascending particles find themselves surrounded by slower fluid which decelerates them (and more so because they are strongly associated with ejections); while descending particles are in the opposite condition, although this is due to gravity and not sweeps. By comparing with previous studies, we confirm that acceleration variance is larger for heavier particles, in stark contrast with the well-known trend in homogeneous turbulence. This is strictly a consequence of the crossing trajectories effect caused by gravitational drift, rather than by inertia. Indeed, near the wall, where the gravitational drift is smaller and the local ${{\textit {St}}}_\eta$ is larger, the accelerations are much less intermittent than further away from the wall.

Contrasting the effects of gravity and inertia is important to estimate the trends in different regimes. If the same particles are suspended in a turbulent boundary layer at higher Reynolds numbers, for example, the near-wall turbulence time scales will become smaller. Thus, the Stokes number (encapsulating the effect of inertia) will grow; while the friction velocity will increase, and so the Rouse number (representing the relative importance of gravitational drift) will shrink. Our conclusions then provide indications on the expected trends in such a case: the particles would lag the fluid more markedly; the settling rate would decrease; the particle Reynolds stresses would be even larger than the fluid's; the particle diffusivity would increase (the streamwise component becoming increasingly larger than the eddy diffusivity, and the vertical component getting closer to it); the concentration profile would get flatter; and the acceleration variance and intermittency would decrease. Future parametric studies are warranted to verify these predictions.

The duration of the wall–particle contact follows an exponential distribution, suggesting an analogy with the waiting times between deposition and entrainment in sediment transport: after touch-down, the particles slide along the wall (possibly over a lubrication layer we cannot resolve) until fluid fluctuations of sufficient strength lift them back into the stream. When expressed in wall units, the time scale of wall contact (i.e. the inverse of the entrainment rate) is close to ${{\textit {St}}}^{+}$, in line with our interpretation of the process: a balance between particle inertia and fluid turbulence. The Lagrangian average along the trajectories before and after wall contact indicate that the particles descend more steeply towards the wall than they recede from it, due to a combination of gravity and Saffman lift. The latter is directed downward, because the particles in close proximity to the wall are faster than the fluid surrounding them. This also implies that the drag force opposes the particle streamwise motion. The impact velocity is well below the threshold for significant energy restitution, therefore there is effectively no rebound. The lift-off coincides with the particles meeting a fluid ejection, reaffirming the importance of these events in the particle transport. The particles accelerate after leaving the wall, even while they are still faster than the surrounding fluid; this is likely an effect of added mass and fluid acceleration forces. Future experiments focused on the near-wall region with higher resolution and the ability to detect particle rotation are needed to determine the fluid velocity gradients surrounding the particles and assess the role of lubrication forces and rotation-induced lift.

Different configurations (e.g. particles of different size and density) will display different behaviours. In particular, the particle size is expected to have an important impact on the near-wall behaviour, where the velocity gradients are sharper. The present results, however, paint a coherent picture of the multi-faceted problem, shed some new light on a series of long-standing issues, and isolate new questions that require an even deeper investigation. Moreover, inasmuch as we reported novel statistics for several previously undocumented quantities, the results should be useful for developing and validating numerical models.

Acknowledgements

This work was funded by the Legislative-Citizen Commission on Minnesota Resources (LCCMR). L.J.B. was supported by the Department of Defense (DoD) through the National Defense Science & Engineering Graduate Fellowship (NDSEGF) Program.

Declaration of interests

The authors report no conflict of interest.

Footnotes

Present address: Department of Mechanical and Process Engineering, ETH Zurich, Switzerland.

References

REFERENCES

Adhikari, D. 2013 Volumetric velocity measurement of aquatic predator-prey interactions. PhD thesis, University of Minnesota.Google Scholar
Ancey, C., Böhm, T., Jodeau, M. & Frey, P. 2006 Statistical description of sediment transport experiments. Phys. Rev. E 74 (1), 011302.CrossRefGoogle ScholarPubMed
Ancey, C., Davison, A. C., Böhm, T., Jodeau, M. & Frey, P. 2008 Entrainment and motion of coarse particles in a shallow water stream down a steep slope. J. Fluid Mech. 595, 83114.CrossRefGoogle Scholar
Ayyalasomayajula, S., Gylfason, A., Collins, L. R., Bodenschatz, E. & Warhaft, Z. 2006 Lagrangian measurements of inertial particle accelerations in grid generated wind tunnel turbulence. Phys. Rev. Lett. 97 (14), 144507.CrossRefGoogle ScholarPubMed
Bagnold, R. A. 1936 The movement of desert sand. Proc. R. Soc. Lond. A 157 (892), 594620.Google Scholar
Bagnold, R. A. 1941 The Physics of Blown Sand and Desert Dunes. Methuen.Google Scholar
Baker, L. J. & Coletti, F. 2019 Experimental study of negatively buoyant finite-size particles in a turbulent boundary layer up to dense regimes. J. Fluid Mech. 866, 598629.CrossRefGoogle Scholar
Balachandar, S., Liu, K. & Lakhote, M. 2019 Self-induced velocity correction for improved drag estimation in Euler–Lagrange point-particle simulations. J. Comput. Phys. 376, 160185.CrossRefGoogle Scholar
Bec, J., Biferale, L., Boffetta, G., Celani, A., Cencini, M., Lanotte, A., Musacchio, S. & Toschi, F. 2006 Acceleration statistics of heavy particles in turbulence. J. Fluid Mech. 550, 349358.CrossRefGoogle Scholar
Bendat, J. S. & Piersol, A. G. 2011 Random Data: Analysis and Measurement Procedures, vol. 729. John Wiley & Sons.Google Scholar
Berk, T. & Coletti, F. 2020 Transport of inertial particles in high-Reynolds-number turbulent boundary layers. J. Fluid Mech. 903, A18.CrossRefGoogle Scholar
Bernardini, M. 2014 Reynolds number scaling of inertial particle statistics in turbulent channel flows. J. Fluid Mech. 758, R1.CrossRefGoogle Scholar
Brenner, H. 1965 Effect of finite boundaries on the Stokes resistance of an arbitrary particle. J. Fluid Mech. 12 (1), 3548.CrossRefGoogle Scholar
Cameron, S. M., Nikora, V. I. & Witz, M. J. 2020 Entrainment of sediment particles by very large-scale motions. J. Fluid Mech. 888, A7.CrossRefGoogle Scholar
Capecelatro, J. & Desjardins, O. 2013 An Euler–Lagrange strategy for simulating particle-laden flows. J. Comput. Phys. 238, 131.CrossRefGoogle Scholar
Choi, J.-I., Yeo, K. & Lee, C. 2004 Lagrangian statistics in turbulent channel flow. Phys. Fluids 16 (3), 779793.CrossRefGoogle Scholar
Clift, R., Grace, J. R. & Weber, M. E. 2005 Bubbles, Drops, and Particles. Dover Publications.Google Scholar
Crowe, C. T., Schwarzkopf, J. D., Sommerfeld, M. & Tsuji, Y. 2011 Multiphase Flows with Droplets and Particles. CRC Press.CrossRefGoogle Scholar
Csanady, G. T. 1963 Turbulent diffusion of heavy particles in the atmosphere. J. Atmos. Sci. 20 (3), 201208.2.0.CO;2>CrossRefGoogle Scholar
De Graaff, D. B. & Eaton, J. K. 2000 Reynolds-number scaling of the flat-plate turbulent boundary layer. J. Fluid Mech. 422, 319346.CrossRefGoogle Scholar
Ebrahimian, M., Sanders, R. S. & Ghaemi, S. 2019 Dynamics and wall collision of inertial particles in a solid–liquid turbulent channel flow. J. Fluid Mech. 881, 872905.CrossRefGoogle Scholar
Einstein, H. A. 1950 The Bed-Load Function for Sediment Transportation in Open Channel Flows. US Government Printing Office.Google Scholar
Elghobashi, S. & Truesdell, G. C. 1992 Direct simulation of particle dispersion in a decaying isotropic turbulence. J. Fluid Mech. 242, 655700.CrossRefGoogle Scholar
Fan, N., Singh, A., Guala, M., Foufoula-Georgiou, E. & Wu, B. 2016 Exploring a semimechanistic episodic Langevin model for bed load transport: emergence of normal and anomalous advection and diffusion regimes. Water Resour. Res. 52 (4), 27892801.CrossRefGoogle Scholar
Fong, K. O., Amili, O. & Coletti, F. 2019 Velocity and spatial distribution of inertial particles in a turbulent channel flow. J. Fluid Mech. 872, 367406.CrossRefGoogle Scholar
Gerashchenko, S., Sharp, N. S., Neuscamman, S. & Warhaft, Z. 2008 Lagrangian measurements of inertial particle accelerations in a turbulent boundary layer. J. Fluid Mech. 617, 255281.CrossRefGoogle Scholar
Gondret, P., Lance, M. & Petit, L. 2002 Bouncing motion of spherical particles in fluids. Phys. Fluids 14 (2), 643652.CrossRefGoogle Scholar
Guala, M., Manes, C., Clifton, A. & Lehning, M. 2008 On the saltation of fresh snow in a wind tunnel: profile characterization and single particle statistics. J. Geophys. Res.: Earth 113, F03024.CrossRefGoogle Scholar
Gualtieri, P., Picano, F., Sardina, G. & Casciola, C. M. 2015 Exact regularized point particle method for multiphase flows in the two-way coupling regime. J. Fluid Mech. 773, 520561.CrossRefGoogle Scholar
Heyman, J., Bohorquez, P. & Ancey, C. 2016 Entrainment, motion, and deposition of coarse particles transported by water over a sloping mobile bed. J. Geophys. Res.: Earth 121 (10), 19311952.CrossRefGoogle Scholar
Horwitz, J. A. K. & Mani, A. 2016 Accurate calculation of Stokes drag for point–particle tracking in two-way coupled flows. J. Comput. Phys. 318, 85109.CrossRefGoogle Scholar
van Hout, R. 2011 Time-resolved PIV measurements of the interaction of polystyrene beads with near-wall-coherent structures in a turbulent channel flow. Intl J. Multiphase Flow 37 (4), 346357.CrossRefGoogle Scholar
van Hout, R. 2013 Spatially and temporally resolved measurements of bead resuspension and saltation in a turbulent water channel flow. J. Fluid Mech. 715, 389.CrossRefGoogle Scholar
van Hout, R., Sabban, L. & Cohen, A. 2013 The use of high-speed PIV and holographic cinematography in the study of fiber suspension flows. Acta Mechanica 224 (10), 22632280.CrossRefGoogle Scholar
Hurther, D. & Lemmin, U. 2003 Turbulent particle flux and momentum flux statistics in suspension flow. Water Resour. Res. 39 (5), 1139.CrossRefGoogle Scholar
Ireland, P. J. & Desjardins, O. 2017 Improving particle drag predictions in Euler–Lagrange simulations with two-way coupling. J. Comput. Phys. 338, 405430.CrossRefGoogle Scholar
Joseph, G. G., Zenit, R., Hunt, M. L. & Rosenwinkel, A. M. 2001 Particle–wall collisions in a viscous fluid. J. Fluid Mech. 433, 329346.CrossRefGoogle Scholar
Jung, J., Yeo, K. & Lee, C. 2008 Behavior of heavy particles in isotropic turbulence. Phys. Rev. E 77 (1), 016307.CrossRefGoogle ScholarPubMed
Kaftori, D., Hetsroni, G. & Banerjee, S. 1995 a Particle behavior in the turbulent boundary layer. I. Motion, deposition, and entrainment. Phys. Fluids 7 (5), 10951106.CrossRefGoogle Scholar
Kaftori, D., Hetsroni, G. & Banerjee, S. 1995 b Particle behavior in the turbulent boundary layer. II. Velocity and distribution profiles. Phys. Fluids 7 (5), 11071121.CrossRefGoogle Scholar
Kidanemariam, A. G., Chan-Braun, C., Doychev, T. & Uhlmann, M. 2013 Direct numerical simulation of horizontal open channel flow with finite-size, heavy particles at low solid volume fraction. New J. Phys. 15 (2), 025031.CrossRefGoogle Scholar
Kiger, K. T. & Pan, C. 2002 Suspension and turbulence modification effects of solid particulates on a horizontal turbulent channel flow. J. Turbul. 3 (19), 117.CrossRefGoogle Scholar
Kok, J. F., Parteli, E. J. R., Michaels, T. I. & Karam, D. B. 2012 The physics of wind-blown sand and dust. Rep. Prog. Phys. 75 (10), 106901.CrossRefGoogle Scholar
Lajeunesse, E., Malverti, L. & Charru, F. 2010 Bed load transport in turbulent flow at the grain scale: experiments and modeling. J. Geophys. Res.: Earth 115, F04001.CrossRefGoogle Scholar
Lavezzo, V., Soldati, A., Gerashchenko, S., Warhaft, Z. & Collins, L. R. 2010 On the role of gravity and shear on inertial particle accelerations in near-wall turbulence. J. Fluid Mech. 658, 229246.CrossRefGoogle Scholar
Lee, H. & Balachandar, S. 2010 Drag and lift forces on a spherical particle moving on a wall in a shear flow at finite Re. J. Fluid Mech. 657, 89125.CrossRefGoogle Scholar
Lee, J. & Lee, C. 2015 Modification of particle-laden near-wall turbulence: effect of Stokes number. Phys. Fluids 27 (2), 023303.CrossRefGoogle Scholar
Lee, J. & Lee, C. 2019 The effect of wall-normal gravity on particle-laden near-wall turbulence. J. Fluid Mech. 873, 475507.CrossRefGoogle Scholar
Li, J., Wang, H., Liu, Z., Chen, S. & Zheng, C. 2012 An experimental study on turbulence modification in the near-wall boundary layer of a dilute gas-particle channel flow. Exp. Fluids 53 (5), 13851403.CrossRefGoogle Scholar
Lin, Z.-W., Shao, X.-M., Yu, Z.-S. & Wang, L.-P. 2017 Effects of finite-size heavy particles on the turbulent flows in a square duct. J. Hydrodyn. 29 (2), 272282.CrossRefGoogle Scholar
Marchioli, C. 2017 Large-eddy simulation of turbulent dispersed flows: a review of modelling approaches. Acta Mechanica 228 (3), 741771.CrossRefGoogle Scholar
Marchioli, C. & Soldati, A. 2002 Mechanisms for particle transfer and segregation in a turbulent boundary layer. J. Fluid Mech. 468, 283315.CrossRefGoogle Scholar
Mathai, V., Loeffen, L. A. W. M., Chan, T. T. K. & Wildeman, S. 2019 Dynamics of heavy and buoyant underwater pendulums. J. Fluid Mech. 862, 348363.CrossRefGoogle Scholar
Maxey, M. R. & Riley, J. J. 1983 Equation of motion for a small rigid sphere in a nonuniform flow. Phys. Fluids 26 (4), 883889.CrossRefGoogle Scholar
Mordant, N., Crawford, A. M. & Bodenschatz, E. 2004 Experimental Lagrangian acceleration probability density function measurement. Physica D 193 (1-4), 245251.CrossRefGoogle Scholar
Nemes, A., Dasari, T., Hong, J., Guala, M. & Coletti, F. 2017 Snowflakes in the atmospheric surface layer: observation of particle–turbulence dynamics. J. Fluid Mech. 814, 592613.CrossRefGoogle Scholar
Nielsen, P. 1993 Turbulence effects on the settling of suspended particles. J. Sedim. Res. 63 (5), 835838.Google Scholar
Niño, Y. & Garcia, M. H. 1996 Experiments on particle–turbulence interactions in the near-wall region of an open channel flow: implications for sediment transport. J. Fluid Mech. 326, 285319.CrossRefGoogle Scholar
Nishimura, K. & Hunt, J. C. R. 2000 Saltation and incipient suspension above a flat particle bed below a turbulent boundary layer. J. Fluid Mech. 417, 77102.CrossRefGoogle Scholar
Papanicolaou, A. N., Diplas, P., Evaggelopoulos, N. & Fotopoulos, S. 2002 Stochastic incipient motion criterion for spheres under various bed packing conditions. J. Hydraul. Engng ASCE 128 (4), 369380.CrossRefGoogle Scholar
Petersen, A. J., Baker, L. & Coletti, F. 2019 Experimental study of inertial particles clustering and settling in homogeneous turbulence. J. Fluid Mech. 864, 925970.CrossRefGoogle Scholar
Picano, F., Breugem, W.-P. & Brandt, L. 2015 Turbulent channel flow of dense suspensions of neutrally buoyant spheres. J. Fluid Mech. 764, 463487.CrossRefGoogle Scholar
Pope, S. B. 2000 Turbulent Flows. Cambridge University Press.CrossRefGoogle Scholar
Prandtl, L. 1952 Essentials of Fluid Dynamics: with Applications to Hydraulics, Aeronautics, Meteorology and Other Subjects. Hafner Publishing Company.Google Scholar
Rabencov, B., Arca, J. & van Hout, R. 2014 Measurement of polystyrene beads suspended in a turbulent square channel flow: spatial distributions of velocity and number density. Intl J. Multiphase Flow 62, 110122.CrossRefGoogle Scholar
Reeks, M. W. 1977 On the dispersion of small particles suspended in an isotropic turbulent fluid. J. Fluid Mech. 83 (3), 529546.CrossRefGoogle Scholar
Richter, D. H. & Sullivan, P. P. 2014 Modification of near-wall coherent structures by inertial particles. Phys. Fluids 26 (10), 103304.CrossRefGoogle Scholar
Righetti, M. & Romano, G. P. 2004 Particle–fluid interactions in a plane near-wall turbulent flow. J. Fluid Mech. 505, 93121.CrossRefGoogle Scholar
Rimon, Y. & Cheng, S. I. 1969 Numerical solution of a uniform flow over a sphere at intermediate Reynolds numbers. Phys. Fluids 12 (5), 949959.CrossRefGoogle Scholar
Robinson, S. K. 1991 Coherent motions in the turbulent boundary layer. Annu. Rev. Fluid Mech. 23 (1), 601639.CrossRefGoogle Scholar
Rouse, H. 1937 Modern conceptions of the mechanics of fluid turbulence. Trans. Am. Soc. Civ. Engng 102, 463505.Google Scholar
Rouson, D. W. I. & Eaton, J. K. 2001 On the preferential concentration of solid particles in turbulent channel flow. J. Fluid Mech. 428, 149169.CrossRefGoogle Scholar
Sabban, L. & van Hout, R. 2011 Measurements of pollen grain dispersal in still air and stationary, near homogeneous, isotropic turbulence. J. Aerosol Sci. 42 (12), 867882.CrossRefGoogle Scholar
Saffman, P. G. T. 1965 The lift on a small sphere in a slow shear flow. J. Fluid Mech. 22 (2), 385400.CrossRefGoogle Scholar
Sardina, G., Schlatter, P., Brandt, L., Picano, F. & Casciola, C. M. 2012 Wall accumulation and spatial localization in particle-laden wall flows. J. Fluid Mech. 699, 5078.CrossRefGoogle Scholar
Shields, A. 1936 Anwendung der Aehnlichkeitsmechanik und der Turbulenzforschung auf die Geschiebebewegung. PhD thesis, Technical University Berlin.Google Scholar
Smits, A. J., McKeon, B. J. & Marusic, I. 2011 High–Reynolds number wall turbulence. Annu. Rev. Fluid Mech. 43, 353375.CrossRefGoogle Scholar
Squires, K. D. & Eaton, J. K. 1991 Measurements of particle dispersion obtained from direct numerical simulations of isotropic turbulence. J. Fluid Mech. 226, 135.CrossRefGoogle Scholar
Tanière, Anne & Arcen, Boris 2016 Overview of existing Langevin models formalism for heavy particle dispersion in a turbulent channel flow. Intl J. Multiphase Flow 82, 106118.CrossRefGoogle Scholar
Tanière, A., Oesterlé, B. & Monnier, J. C. 1997 On the behaviour of solid particles in a horizontal boundary layer with turbulence and saltation effects. Exp. Fluids 23 (6), 463471.Google Scholar
Taylor, G. I. 1921 Diffusion by continuous movements. Proc. Lond. Math. Soc. s2-20 (1), 196212.Google Scholar
Tee, Y. H., Barros, D. & Longmire, E. K. 2020 Motion of finite-size spheres released in a turbulent boundary layer. Intl J. Multiphase Flow 133, 103462.CrossRefGoogle Scholar
Voth, G. A., La Porta, A., Crawford, A. M., Alexander, J. & Bodenschatz, E. 2002 Measurement of particle accelerations in fully developed turbulence. J. Fluid Mech. 469, 121160.CrossRefGoogle Scholar
Wang, G., Abbas, M. & Climent, É. 2017 Modulation of large-scale structures by neutrally buoyant and inertial finite-size particles in turbulent Couette flow. Phys. Rev. Fluids 2 (8), 084302.CrossRefGoogle Scholar
Wang, G., Fong, K. O., Coletti, F., Capecelatro, J. & Richter, D. H. 2019 Inertial particle velocity and distribution in vertical turbulent channel flow: a numerical and experimental comparison. Intl J. Multiphase Flow 120, 103105.CrossRefGoogle Scholar
Wang, L.-P. & Maxey, M. R. 1993 Settling velocity and concentration distribution of heavy particles in homogeneous isotropic turbulence. J. Fluid Mech. 256, 2768.CrossRefGoogle Scholar
Wang, L.-P. & Stock, D. E. 1993 Dispersion of heavy particles by turbulent motion. J. Atmos. Sci. 50 (13), 18971913.2.0.CO;2>CrossRefGoogle Scholar
Wells, M. R. & Stock, D. E. 1983 The effects of crossing trajectories on the dispersion of particles in a turbulent flow. J. Fluid Mech. 136, 3162.CrossRefGoogle Scholar
Westerweel, J. & Scarano, F. 2005 Universal outlier detection for PIV data. Exp. Fluids 39, 10961100.CrossRefGoogle Scholar
Zamansky, R., Vinkovic, I. & Gorokhovski, M. 2011 Acceleration statistics of solid particles in turbulent channel flow. Phys. Fluids 23 (11), 113304.CrossRefGoogle Scholar
Zhao, L. H., Andersson, H. I. & Gillissen, J. J. J. 2010 Turbulence modulation and drag reduction by spherical particles. Phys. Fluids 22 (8), 081702.CrossRefGoogle Scholar
Zhu, H.-Y., Pan, C., Wang, J.-J., Liang, Y.-R. & Ji, X.-C. 2019 Sand-turbulence interaction in a high-Reynolds-number turbulent boundary layer under net sedimentation conditions. Intl J. Multiphase Flow 119, 5671.CrossRefGoogle Scholar
Figure 0

Figure 1. Diagram of the water channel showing key components and dimensions. The bold arrow indicates the direction of the flow.

Figure 1

Figure 2. Instantaneous realization of the particle-laden flow. Both the mm-sized spherical PS particles and the microscopic silver-coated glass tracer particles are visible.

Figure 2

Table 1. Properties of the PS particles; $D_p$ is the mean particle diameter, $D_p^{+}$ is the diameter normalized by the viscous length scale of the flow, $\rho _p$ is the density, $V_t$ is the terminal velocity in still water, $Re_{p,V_t}$ is the Reynolds number based on $V_t$, $\tau _p$ is the particle response time, ${{\textit {St}}}^{+}$ is the particle Stokes number based on the viscous time scale of the flow, Ga is the Galileo number, Sh is the Shields number and $\varPhi _v$ is the particle volume fraction.

Figure 3

Figure 3. Probability density function of PS particle diameters. The standard deviation is 12 % of the mean value.

Figure 4

Figure 4. Wall-normal profiles of (a) the particle Stokes number based on the Kolmogorov scale and (b) the particle diameter normalized by the Kolmogorov scale.

Figure 5

Table 2. Imaging and PIV processing parameters: $f_s$ is the imaging frequency; $N$ is the number of images; $w$ and $h$ are the field of view width and height, respectively; $w_i$ is the final-pass PIV interrogation window size; and $\delta x$ is the PIV vector spacing.

Figure 6

Figure 5. Convolution method for particle detection: (a) original image, (b) median filtered image, (c) particle template image to be convolved with the filtered image and (d) convolution peak, the red cross indicating the detected particle centroid.

Figure 7

Figure 6. Streamwise acceleration variance as a function of smoothing kernel width. The solid line indicates the fit of the acceleration variance over its exponential range, and the optimal kernel width is denoted by the open circle.

Figure 8

Figure 7. Mean streamwise velocity profile of the particle-laden (black dots) and unladen (red dots) fluid, in (a) outer units and (b) wall units. The dashed line in (b) indicates the logarithmic-law fit. The profile is compared with De Graaff & Eaton (2000) in (b), shown in blue open circles.

Figure 9

Figure 8. Profiles of particle-laden (black dots) and unladen (red dots) streamwise turbulent normal stress (a), wall-normal normal stress (b) and shear stress (c) of the fluid in wall units. The profiles are compared with De Graaff & Eaton (2000) shown in blue open circles.

Figure 10

Table 3. Physical parameters of the water channel and boundary-layer properties. $U_\infty$ is the free-stream velocity, H is the water depth, W is the channel width, $\delta _{99}$ is the boundary-layer thickness and $u_\tau$ is the shear velocity. The boundary thickness is defined such that $u(\delta _{99}) = 0.99U_\infty$; $Re = U_\infty H/\nu$, $Re_\tau = u_\tau \delta _{99}/\nu$ and $Re_\theta = U_\infty \theta /\nu$ are the free-stream, friction and momentum thickness Reynolds numbers, respectively. Standard water properties at $22\,^{\circ }\textrm {C}$ are used in the calculations.

Figure 11

Figure 9. Wall-normal profiles of streamwise (a) and wall-normal (b) mean particle velocity (red crosses) compared with the mean fluid velocity (black dots).

Figure 12

Figure 10. Wall-normal profiles of mean streamwise (a) and wall-normal (b) particle slip velocity, separated into the particle-conditioned mean slip (blue circles) and the apparent mean slip (red crosses).

Figure 13

Figure 11. Profiles of mean instantaneous particle Reynolds number based on the particle-conditioned streamwise (blue circles) and wall-normal (red crosses) slip velocity.

Figure 14

Figure 12. Profiles of mean streamwise (a) and wall-normal (b) particle velocity conditioned on ascending (red crosses) and descending (blue circles) particles compared to the fluid velocity (black dots).

Figure 15

Figure 13. Wall-normal profiles of the Reynolds stresses of the particles (red crosses) and fluid (black dots) normalized by the squared shear velocity.

Figure 16

Figure 14. Wall-normal profiles of the Reynolds shear stress of (a) the fluid velocity at particle locations and (b) the particle velocity, conditioned on ascending and descending particles (red crosses and blue circles, respectively). The profiles are compared to the overall fluid shear stress profile (black dots).

Figure 17

Figure 15. Joint PDFs of streamwise and wall-normal fluctuating fluid velocities for $y^{+} < 100$ (a) and $y^{+} > 100$ (b). Joint PDFs of streamwise and wall-normal fluctuating fluid velocities at particle locations for $y^{+} < 100$ (c,e) and $y^{+} > 100$ (d,f) conditioned on whether the particle is ascending (c,d) or descending (e,f).

Figure 18

Figure 16. Lagrangian autocorrelations of streamwise (a) and wall-normal (b) particle velocity for five wall-normal bins (dots) shown with their respective exponential fits (dashed lines). The wall-normal locations listed in the legend correspond to the centre of each bin.

Figure 19

Figure 17. Wall-normal profiles of streamwise and wall-normal diffusivity (blue circles and red crosses, respectively) compared to the theoretical profile of fluid momentum diffusivity (black dashed line).

Figure 20

Figure 18. Wall-normal profile of mean particle concentration normalized by the concentration at the lowest wall-normal bin (black crosses). The power-law profile predicted by Rouse–Prandtl theory (red dashed line) is calculated from (3.6), where the arbitrary reference height is taken at $y_r^{+} = 90$.

Figure 21

Figure 19. Wall-normal profile of particle streamwise mass flux normalized by the mass flux at the lowest wall-normal bin (black crosses), compared with a power-law fit (red dashed line).

Figure 22

Figure 20. (a) Profiles of mean streamwise and wall-normal particle acceleration. (b) Profiles of mean streamwise and wall-normal particle acceleration conditioned on whether particles are ascending (filled symbols) or descending (open symbols).

Figure 23

Figure 21. Wall-normal profiles of streamwise (blue circles) and wall-normal (red crosses) particle r.m.s. acceleration.

Figure 24

Figure 22. PDFs of streamwise (blue filled symbols) and wall-normal (red open symbols) particle acceleration conditioned on $y^{+} < 100$ (circles) and $y^{+} > 100$ (triangles). The PDFs for $y^{+} > 100$ are shifted upward by a factor of $10^{3}$ for clarity.

Figure 25

Figure 23. Lagrangian autocorrelation of streamwise particle acceleration for five wall-normal bins. The wall-normal locations listed in the legend correspond to the centre of each bin.

Figure 26

Figure 24. An example of an interaction between a particle and the wall. The blue curve shows the wall-normal distance of the particle centroid vs. time, with error bars representing the uncertainty in particle position. The dashed line represents a height of one particle radius above the wall, with the grey shaded region indicating the uncertainty in the particle radius and the wall location.

Figure 27

Figure 25. PDF of duration of particle–wall interactions, defined as the length of time particles spend continuously within 1.3 mean particle radii of the wall.

Figure 28

Figure 26. PDF of the angle of trajectories for near-wall ($y^{+} < D_p^{+}$) particles.

Figure 29

Figure 27. Lagrangian means of particle wall-normal coordinate (a,b), streamwise velocity (c,d) and wall-normal velocity (e,f) averaged over all identified touch-down events (a,c,e) and lift-off events (b,d,f). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.

Figure 30

Figure 28. Lagrangian means of streamwise (a,b) and wall-normal (c,d) fluid velocity at particle location averaged over all identified touch-down events (a,c) and lift-off events (b,d). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.

Figure 31

Figure 29. Lagrangian means of particle tangential acceleration (a,b) and normal acceleration (c,d) averaged over all identified touch-down events (a,c) and lift-off events (b,d). The time axis shows the time interval before or after lift-off or touch-down, respectively, and ${\rm \Delta} t^{+} = 0$ represents the moment of lift-off or touch-down.