Hostname: page-component-cb9f654ff-r5d9c Total loading time: 0 Render date: 2025-08-18T10:56:04.046Z Has data issue: false hasContentIssue false

Effects of settling on inertial particle slip velocity statistics in wall-bounded flows

Published online by Cambridge University Press:  01 August 2025

Andrew P. Grace*
Affiliation:
Department of Civil and Environmental Engineering and Earth Sciences, University of Notre Dame, Notre Dame, IN 46556, USA
Tim Berk
Affiliation:
Department of Mechanical and Aerospace Engineering, Utah State University, Logan, UT 84322, USA
Andrew D. Bragg
Affiliation:
Department of Civil and Environmental Engineering, Duke University, Durham, NC 27708, USA
David H. Richter
Affiliation:
Department of Civil and Environmental Engineering and Earth Sciences, University of Notre Dame, Notre Dame, IN 46556, USA
*
Corresponding author: Andrew P. Grace, agrace4@nd.edu

Abstract

Developing reduced-order models for the transport of solid particles in turbulence typically requires a statistical description of the particle–turbulence interactions. In this work, we utilize a statistical framework to derive continuum equations for the moments of the slip velocity of inertial, settling Lagrangian particles in a turbulent boundary layer. Using coupled Eulerian–Lagrangian direct numerical simulations, we then identify the dominant mechanisms controlling the slip velocity variance, and find that for a range of Stokes number ${S{\kern-0.5pt}t}^+$, Settling number ${S{\kern-0.5pt}v}^+$ and Reynolds number $\textit{Re}_\tau$ (based on frictional scales),the slip variance is primarily controlled by local differences between the ‘seen’ variance and the particle velocity variance, while terms appearing due to the inhomogeneity of the turbulence are subleading until ${S{\kern-0.5pt}v}^+$ becomes large. We also consider several comparative metrics to assess the relative magnitudes of the fluctuating slip velocity and the mean slip velocity, and we find that the vertical mean slip increases rapidly with ${S{\kern-0.5pt}v}^+$, rendering the variance relatively small – an effect found to be most substantial for ${S{\kern-0.5pt}v}^+\gt 1$. Finally, we compare the results with a model of the acceleration variance (Berk & Coletti 2021 J. Fluid Mech. 917, A47) based the concept of a response function described in Csanady (1963 J. Atmos. Sci. 20, 201–208), highlighting the role of the crossing trajectories mechanism. We find that while there is good agreement for low ${S{\kern-0.5pt}v}^+$, systematic errors remain, possibly due to implicit non-local effects arising from rapid particle settling and inhomogeneous turbulence. We conclude with a discussion of the implications of this work for modelling the transport of coarse dust grains in the atmospheric surface layer.

Information

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

1. Introduction

The study of the transport of inertial particles through fluids finds numerous applications in the natural sciences and in industry. A significant focus is placed on understanding how small but heavy particles respond to turbulence processes, and how to model these processes in a physically coherent way. One such example is understanding the global transport of coarse dust particles (30–100 μm) once they are emitted from the surfaces of arid regions (Rosenberg et al. Reference Rosenberg2014; Meng et al. Reference Meng, Huang, Leung, Li, Adebiyi, Ryder, Mahowald and Kok2022; Adebiyi et al. Reference Adebiyi2023; Kok et al. Reference Kok, Storelvmo, Karydis, Adebiyi, Mahowald, Evan, He and Leung2023). These particles can be lofted high into the turbulent atmosphere where they can be transported many hundreds to thousands of kilometres depending on their size (Shao Reference Shao2008; Van Der Does et al. Reference van der Does, Knippertz, Zschenderlein, Giles Harrison and Stuut2018). The interactions between the dust particles and the carrier phase must be parameterized since such interactions occur at the particle scale, and cannot be represented explicitly due to unrealistic computational requirements. Understanding the impacts of turbulence on particle transport characteristics, such as emission and deposition (Kok et al. Reference Kok, Parteli, Michaels and Karam2012), will help us to understand their overall role in global climate processes (Kok Reference Kok2011; Ryder et al. Reference Ryder, Highwood, Walser, Seibert, Philipp and Weinzierl2019; Kok et al. Reference Kok, Storelvmo, Karydis, Adebiyi, Mahowald, Evan, He and Leung2023), biogeochemical cycles (Ryder et al. Reference Ryder2018) and human health.

From a dynamical perspective, solid particles are subjected to various forces as they travel through a turbulent flow. For small (relative to the local Kolmogorov scale) and dense (relative to the carrier phase) spherical particles, the most important forces are due to gravity and hydrodynamic drag (Maxey & Riley Reference Maxey and Riley1983). Since the seminal work of Wang & Stock (Reference Wang and Stock1993), there has been a significant push to try to understand how gravity and turbulent drag couple together to affect both mean and fluctuating quantities through experiment and simulation (Aliseda et al. Reference Aliseda, Cartellier, Hainaux and Lasheras2002; Good et al. Reference Good, Ireland, Bewley, Bodenschatz, Collins and Warhaft2014; Rosa et al. Reference Rosa, Parishani, Ayala and Wang2016; Tom & Bragg Reference Tom and Bragg2019; Mora et al. Reference Mora, Obligado, Aliseda and Cartellier2021; Ferran et al. Reference Ferran, Machicoane, Aliseda and Obligado2023). Importantly, in a turbulent flow, the bias created by gravity leads to a more rapid decorrelation of the turbulence along particle trajectories, meaning that there is an implicit and nonlinear coupling between gravity and turbulent drag, resulting in a fundamental change in the forcing induced by the turbulence. This is an example of the crossing trajectories effect (Yudine Reference Yudine1959; Csanady Reference Csanady1963). Crossing trajectories occurs when Lagrangian particle trajectories cross fluid particle trajectories due to mean Lagrangian particle drift (which can occur even in the absence of turbulence). The net effect is that Lagrangian particles may experience higher frequency fluid fluctuations leading to an increase in the particle dispersion. Crossing trajectories has also been shown to increase the horizontal and vertical components of particle acceleration variance in simulations of settling Lagrangian point particles in omogeneous isotropic turbulence (HIT) (Ireland, Bragg & Collins Reference Ireland, Bragg and Collins2016b ), in numerical simulations of turbulent boundary layers (TBL) (Lavezzo et al. Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010) as well as laboratory experiments in both set-ups (Gerashchenko et al. Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008; Berk & Coletti Reference Berk and Coletti2021). The turbulent drag is often quantified via the particle slip velocity, which is the difference between the fluid velocity seen by the particle, and the particle’s velocity. Understanding the controlling mechanisms of the particle slip velocity and their magnitude within wall-bounded turbulence is key for estimating the particle Reynolds number (Balachandar Reference Balachandar2009), as well building physically coherent stochastic dispersion models for Reynolds-averaged Navier–Stokes (Arcen & Tanière Reference Arcen and Tanière2009) applications.

In this work, we are particularly focused on understanding the dynamic regimes characteristic of coarse dust particles in the Earth’s atmospheric surface layer (the lowest 100 m of the Earth’s atmosphere). Specifically, we consider how gravitational acceleration implicitly modifies the mechanisms controlling the particle slip velocity in a TBL. In a TBL, turbulence is driven by fluid shear originating at the solid lower boundary, resulting in turbulence inhomogeneity and a height dependence of the parameters governing both the turbulent flow and the particle transport. Dynamically, a TBL is characterized by a very thin laminar sublayer where viscosity plays a dominant role, followed by a smooth transitional layer, known as the buffer layer, to a layer where viscous effects become negligible, known as the logarithmic layer. A review of TBLs can be found in Smits, McKeon & Marusic (Reference Smits, McKeon and Marusic2011).

Many studies of particle transport in TBLs tend to ignore the impact of gravitational acceleration a priori in an attempt to try to decouple the effects of turbulent drag and gravity (Marchioli et al. Reference Marchioli, Soldati, Kuerten, Arcen, Tanière, Goldensoph, Squires, Cargnelutti and Portela2008; Balachandar Reference Balachandar2009; Zamansky, Vinkovic & Gorokhovski Reference Zamansky, Vinkovic and Gorokhovski2011; Johnson, Bassenne & Moin Reference Johnson, Bassenne and Moin2020). However, due to the implicit coupling between gravity and turbulent drag, we must take care when extrapolating results from studies without gravity to those with gravity (Brandt & Coletti Reference Brandt and Coletti2022). Furthermore, much of our understanding of particle-laden flows under the influence of gravitational settling comes from numerical (Bec, Homann & Ray Reference Bec, Homann and Ray2014; Good et al. Reference Good, Ireland, Bewley, Bodenschatz, Collins and Warhaft2014; Ireland et al. Reference Ireland, Bragg and Collins2016b ; Tom & Bragg Reference Tom and Bragg2019) or laboratory (Aliseda et al. Reference Aliseda, Cartellier, Hainaux and Lasheras2002; Mora et al. Reference Mora, Obligado, Aliseda and Cartellier2021; Ferran et al. Reference Ferran, Machicoane, Aliseda and Obligado2023) configurations of HIT, due to the relative simplicity of the set-up. There are a few studies aiming to understand the statistical behaviour of settling inertial particles in TBLs (Lavezzo et al. Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010; Lee & Lee Reference Lee and Lee2019; Berk & Coletti Reference Berk and Coletti2020; Bragg, Richter & Wang Reference Bragg, Richter and Wang2021a , Reference Bragg, Richter and Wangb ; Berk & Coletti Reference Berk and Coletti2023), and while they are far less numerous, they indicate the potential for gravitational settling to modify the dynamics of particle settling and two-way coupling due to the presence of the solid boundary. Indeed, Bragg et al. (Reference Bragg, Richter and Wang2021b ) showed that gravitational settling can have a strong impact on the particle transport in a TBL even for very small settling numbers for which is has been traditionally assumed that the effect of settling should be negligible. Having quantitative evidence as to when we may apply models designed under the assumptions of homogeneous turbulence to dynamics in a TBL, and when the settling is important for the particle transport, would be a useful starting point when designing a more unified theory.

In the following work, our goals are:

  1. (i) to derive continuum equations for moments of the particle slip velocity and identify the leading-order balance of the variance throughout the TBL;

  2. (ii) to determine the parametric conditions under which the slip velocity is governed by its mean component, fluctuating component or some combination of both;

  3. (iii) to compare the results from the direct numerical simulations (DNS) in a TBL with a model based on the response function in homogeneous turbulence approach outlined in Csanady (Reference Csanady1963) and Berk & Coletti (Reference Berk and Coletti2021), and identify potential discrepancies;

  4. (iv) to discuss implications for the transport of coarse particles in the atmospheric surface layer.

Section 2 provides the technical background on the carrier and particle phase equations as well as the governing parameters. We also derive the diagnostic equation for the vertical component of the slip velocity variance and discuss the model hierarchy. We choose to focus on the slip velocity variance specifically for several reasons. First, accurate estimates of the particle Reynolds number rely on the estimates of the magnitude of the slip velocity, and this quantity is not easily accessible in a laboratory or field setting directly. Second, it is related to the particle acceleration variance, which gives us a clue as to how the particles respond to turbulent structures within the flow. Section 3 presents the results of the study, while § 4 summarizes and provides a discussion on how the results relate to coarse dust transport in the atmospheric surface layer.

2. Technical background

2.1. Carrier phase

In this work, we use the NCAR Turbulence with Lagrangian Particles Model (Richter & Chamecki Reference Richter and Chamecki2018) to model one-way coupled inertial particles settling through a TBL. This code has been validated and used in multiple studies focused on inertial particle settling and transport in TBLs (Richter & Chamecki Reference Richter and Chamecki2018; Wang et al. Reference Wang, Fong, Coletti, Capecelatro and Richter2019; Bragg et al. Reference Bragg, Richter and Wang2021a ; Gao, Samtaney & Richter Reference Gao, Samtaney and Richter2023; Grace, Richter & Bragg Reference Grace, Richter and Bragg2024). For the carrier phase, we use DNS to solve the three-dimensional, incompressible Navier–Stokes equations in a turbulent open channel flow set-up,

(2.1) $\frac {D\boldsymbol{u}}{Dt} = -\frac {1}{\rho _a}\nabla p + \nu \nabla ^2\boldsymbol{u} - \frac {1}{\rho _a}\frac {{\textrm d} P}{{\textrm d}x}\hat {\boldsymbol{x}}, $
(2.2) $ \nabla \cdot \boldsymbol{u} = 0. $

A schematic of the set-up is presented in figure 1. In the above equations, ${D}/{Dt}$ represents the material derivative, $\boldsymbol{u}$ represents the three-dimensional flow velocity, $p$ represents the turbulent pressure field, $\rho _a$ is the carrier phase density and $\nu$ is the kinematic viscosity. Unconditional fluid velocities will be referred to by their components ( $u,v,w$ ) with no subscript, and fluctuating quantities will be denoted by a prime. For example, as we are primarily focused on wall normal motion in this work, we will refer to fluctuating vertical fluid velocities as $w^\prime$ . The code uses a pseudospectral method in the horizontal directions and a second-order finite difference method in the vertical. We have clustered points near the solid lower boundary using an algebraic grid stretching procedure, so that $\Delta z_1 =0.8\nu /u_\tau$ ( $u_\tau$ is the friction velocity, defined below), where $\Delta z_1$ is the first grid point and $(\Delta x_\eta ,\Delta y_\eta ,\Delta z_\eta ) = (2.8,1.4,0.9)$ at the midplane of the flow, where the subscript $\eta$ represents the grid spacing in terms of the local Kolmogorov scale, defined below.

Figure 1. A schematic of the numerical set-up. The domain is a rectangular channel of height $H$ , streamwise length $2\pi\!{H}$ and spanwise width $\pi\!{H}$ . The flow is periodic in the horizontal and is driven by a constant pressure gradient in the streamwise direction, while the no-stress and no-slip boundary conditions are enforced at the top and bottom boundaries, respectively. Particles are injected at the upper boundary at a random horizontal location with an initial velocity equal to the fluid velocity at their location and removed when they contact the bottom boundary. They are allowed to rebound elastically off the upper boundary.

At the lower boundary, a no-slip boundary condition is enforced, while at the upper boundary, a no-stress boundary condition ( $\partial u/\partial z = \partial v/\partial z = 0$ at $z = H$ ) is enforced. The domain is periodic in the $x$ and $y$ directions. The background state of the carrier phase is established by accelerating the flow with an imposed pressure gradient, $-{\textrm d}P/{\textrm d}x\gt 0$ (note that $\hat {\boldsymbol{x}}$ is the unit vector in the streamwise direction) and allowing the flow to become turbulent. The magnitude of the pressure gradient allows us to define a friction velocity $u_\tau = \sqrt {\tau _w/\rho _a}$ , where $\tau _w$ is the stress at the lower boundary. Using the friction velocity, the height of the domain, $H$ , and viscosity of the carrier phase, we can define a friction Reynolds number of $\textit{Re}_\tau = {u_\tau H}/{\nu }$ . Friction Reynolds numbers for each simulation presented in this work can be found in table 1. This set-up is identical to that used in Grace et al. (Reference Grace, Richter and Bragg2024).

Table 1. Table of cases discussed throughout the work. Parameter definitions can be found in the main text. The case with $\textit{Re}_\tau = 1260$ was run on a $512^3$ grid, all cases with $\textit{Re}_\tau =630$ were run on a $256^3$ grid, while cases with $\textit{Re}_\tau =315$ were run on a $128^3$ grid.

We can define the local Kolmogorov time scale, velocity scale and acceleration scale,

(2.3) \begin{equation} \tau _\eta = \left (\frac {\nu }{\epsilon }\right )^{1/2},\quad v_\eta = \left (\nu \epsilon \right )^{1/4},\quad a_\eta = \left (\frac {\epsilon ^3}{\nu }\right )^{1/4}, \end{equation}

respectively, which represent the smallest relevant length scales of the turbulence. These parameters will be used to characterize the turbulent flow below. In statistically stationary HIT, the above scales are constants, but for wall-bounded turbulence they depend on height. Since the turbulence intensity decreases with height outside of the very thin viscous sublayer adjacent to the wall, so too does the kinetic energy dissipation rate resulting in a height variation of the Kolmogorov microscales. When the TBL is horizontally homogeneous, the mean dissipation rate $\epsilon$ is a function of the distance from the solid boundary. Within the logarithmic layer, it scales as

(2.4) \begin{equation} \epsilon \sim O\left (\frac {u_\tau ^3}{\kappa z}\right ), \end{equation}

where $u_\tau$ is the friction velocity, and $\kappa$ is the von Kármán constant. However, when calculating the Kolmogorov time scale, velocity scale and length scale, we use the dissipation computed in the DNS.

2.2. Dispersed phase

Our main focus is towards on-coarse dust transport in the atmospheric surface layer. Dust particles can range in size, but even coarse grains (roughly 30–100 μm) are significantly smaller than the local Kolmogorov scale, which can be in the range of several millimetres. Indeed, these particles are also significantly denser than the carrier phase, though their volume fractions can be quite low once they are above the emission layer. With these assumptions in mind, for each particle (the dispersed phase), we apply the point-particle approximation and apply the conservation of momentum for a rigid spherical particle subjected to linear hydrodynamic drag and gravity. Note that we have purposefully omitted particle lift as it has been shown to play negligible role in the dynamics discussed in the present work (Costa, Brandt & Picano Reference Costa, Brandt and Picano2020). Furthermore, as we are concerned with the dilute limit, two-way coupling and particle–particle interactions may be ignored. Since these particles are assumed to be much denser than the carrier phase, we may also ignore added mass and Basset history forces. The one-way coupled point particle approximation also has the added benefit that each particle is independent from each other particle, effectively removing the volume fraction as a governing parameter as the number of particles is increased. This allows us to increase the number of particles to ensure convergence of the statistics of interest without affecting the flow. As such, our results are assumed to be valid in the dilute limit, where particle densities are much greater than that of the fluid phase.

The equations of motion are

(2.5) \begin{equation} \frac {{\textrm d}\boldsymbol{u}_p}{{\textrm d}t} = \frac {\Psi }{\tau _p}(\boldsymbol{u}_f - \boldsymbol{u}_p ) - \boldsymbol{g}, \end{equation}
(2.6) \begin{equation} \frac {{\textrm d}\boldsymbol{x}_p}{{\textrm d}t} = \boldsymbol{u}_p. \end{equation}

Here, $\boldsymbol{u}_p = (u_p,v_p,w_p)$ is the three-dimensional velocity vector for each particle, $\boldsymbol{x}_p = (x,y,z)$ is the location of each particle in space, $\boldsymbol{g}$ is the gravitational acceleration (which only affects accelerations in the $z$ direction), $\boldsymbol{u}_f = (u_f,v_f,w_f)$ is the three-dimensional instantaneous flow velocity evaluated at the location of the particle. Much of our focus will be placed on the vertical component which we denote as $w_f$ (not bold).

Particles are injected at the upper boundary at a random horizontal location with an initial velocity equal to the fluid velocity at their location. We have performed a sensitivity test to investigate the importance of the particle initial condition by comparing results when injected at their laminar Stokes settling velocity. We have concluded that the choice of initial particle velocities at injection makes a negligible quantitative change to our results, and does not affect the results qualitatively.

Particles are removed when they contact the bottom boundary, and once removed, are reinjected at a random horizontal location at the upper boundary. This maintains a constant particle number in the domain, and a constant vertical particle flux. Particles are allowed to come to statistical equilibrium, and statistics are recorded for no less than 100 eddy turnover times after this time. Finally, particles are allowed to rebound elastically off the upper boundary, and the particle boundary conditions are periodic in the horizontal directions.

The relaxation time scale of the particles, $\tau _p$ , is defined as

(2.7) \begin{equation} \tau _p = \frac {\rho _pd_p^2}{18\rho _a\nu }, \end{equation}

where $\rho _p$ is the particle density, $d_p$ is the particle diameter and the Stokes settling velocity is defined as $v_g = \tau _p g$ . Here $\Psi = 1 + 0.15\textit{Re}_p^{0.687}$ is the Schiller–Neumann correction to the drag force, and $\textit{Re}_p$ is the particle Reynolds number. The particle Reynolds number remains small enough such that $\Psi \approx 1$ . For the theory discussed below in § 2.3, we follow Bragg et al. (Reference Bragg, Richter and Wang2021a ) and make the assumption that $\Psi = 1$ for analytical tractability.

We can now define a set of non-dimensional parameters characterizing the system,

(2.8) \begin{equation} {S{\kern-0.5pt}t}^+ = \frac {\tau _pu_\tau ^2}{\nu },\quad {S{\kern-0.5pt}v}^+ = \frac {v_g}{u_\tau },\quad {St}_\eta = \frac {\tau _p}{\tau _\eta },\quad {S{\kern-0.5pt}v}_\eta = \frac {v_g}{v_\eta }. \end{equation}

These are the friction Stokes number and the settling velocity parameter based on the viscous scales, and the Stokes number and the settling velocity parameter based on the local Kolmogorov scales. Here ${St}_\eta$ and ${S{\kern-0.5pt}v}_\eta$ are functions of the distance from the solid boundary, whereas ${S{\kern-0.5pt}t}^+$ and ${S{\kern-0.5pt}v}^+$ are constant parameters.

Other studies focused on particle settling define other parameters such as a Froude number, $\textit{Fr} = a_\eta /g$ , (Bec et al. Reference Bec, Homann and Ray2014; Berk & Coletti Reference Berk and Coletti2021), or a scaled gravity

(2.9) \begin{equation} g^+ = \frac {g\nu }{u_\tau ^3}, \end{equation}

which is simply the ratio of ${S{\kern-0.5pt}v}^+$ to ${S{\kern-0.5pt}t}^+$ . These parameters are useful as they describe the relative role of gravitational accelerations and turbulent accelerations without referring to $\tau _p$ . We will refer to $g^+$ at various points throughout the discussion when necessary. The values for all parameters considered in this work, including the associated ranges of ${St}_\eta$ and ${S{\kern-0.5pt}v}_\eta$ , can be found in table 1.

2.3. Statistics of inertial particles in a turbulent boundary layer

We adopt the particle phase space approach used in Bragg et al. (Reference Bragg, Richter and Wang2021a ), focusing only on the vertical component of the particle equations of motion. First they define the particle probability density function (PDF) in position-velocity space as

(2.10) \begin{equation} {\mathcal{P}} = \langle \delta (z_p-z)\delta (w_p-w)\rangle , \end{equation}

which describes the distribution of the vertical components of the particle position and velocity, $z_p(t)$ and $w_p(t)$ , in the phase space with coordinates $z$ , $w$ , and $\langle \cdot \rangle$ represents an ensemble average over all realizations of the system (note that $\delta$ represents the Dirac delta function). Note we make frequent use of conditional averages throughout this work, denoted by $\langle \cdot \rangle _{z,w}$ which is short hand for $\langle \cdot | z_p = z,w_p = w\rangle$ . We can form an evolution equation for the PDF,

(2.11) \begin{equation} \frac {\partial {\mathcal{P}}}{\partial t} + \frac {\partial }{\partial z}\left (w{\mathcal{P}}\right ) + \frac {\partial }{\partial w}\left (\langle a_p\rangle _{z,w}{\mathcal{P}}\right ) = 0 \end{equation}

where we have defined $\langle a_p\rangle _{z,w} = \tau _p^{-1} (\langle w_f\rangle _{z,w}-w ) - g$ as the vertical particle acceleration conditioned on $z_p = z$ and $w_p = w$ based on the vertical component of (2.5). The utility of this equation comes from the fact that we can derive evolution equations for each moment. Recall that the $n$ th moment is defined as

(2.12) \begin{equation} \langle w_p^n\rangle _z = \frac {1}{\varrho }\int _{-\infty }^{\infty }w^n{\mathcal{P}}\,{\textrm d}w, \end{equation}

where the notation $\langle \cdot \rangle _z$ represents an ensemble average conditioned on $z_p = z$ and $\varrho$ is zeroth moment, which is related to the particle number concentration. Of importance to the present work will be the first and second moments (i.e. $n=1$ and $n=2$ ). The evolution equation for the first moment is

(2.13) \begin{equation} \langle w_p\rangle _z = \langle w_f\rangle _z - v_g -\frac {\tau _p}{\varrho }\frac {{\textrm d}}{{\textrm d}z}\varrho \langle w_p^2\rangle _z. \end{equation}

The details of the derivation of this equation can be found in Bragg et al. (Reference Bragg, Richter and Wang2021a ), and the general form of this equation (for the case where $v_g=0$ ) for arbitrary moments can be found in Johnson et al. (Reference Johnson, Bassenne and Moin2020). Equation (2.13) says that the average settling velocity comes from the average fluid velocity sampled by the particles, the laminar Stokes settling velocity, and a term that depends on the derivative of the mean square particle velocity which may be important near a solid boundary. For compactness, we can rearrange (2.13) into a relationship describing how the mean slip velocity varies with height,

(2.14) \begin{equation} \langle w_s\rangle _z = v_g + \frac {\tau _p}{\varrho }\frac {{\textrm d}}{{\textrm d}z}\varrho \langle w_p^2\rangle _z, \end{equation}

where $\langle w_s\rangle _z=\langle w_f\rangle _z - \langle w_p\rangle _z$ .

Our goal is to derive a continuum equation for the slip velocity variance. To do this, we multiply (2.11) by $w^2$ and integrate over all $w$ . The full details of this operation can be found in Johnson et al. (Reference Johnson, Bassenne and Moin2020). After integrating, we are left with the following relationship:

(2.15) \begin{equation} \frac {{\textrm d}}{{\textrm d}z}\varrho \langle w_p^3\rangle _z - 2\varrho \langle a_pw_p\rangle _z = 0. \end{equation}

To expand the acceleration–velocity covariance, we expand to get $\langle a_pw_p\rangle _z = \langle w_fw_p\rangle _z - \langle w_p^2\rangle _z$ and use the fact that $\langle w_s^2\rangle _z =\langle w_f^2\rangle _z - 2\langle w_fw_p\rangle _z + \langle w_p^2\rangle _z$ to arrive at

(2.16) \begin{equation} 2\langle a_pw_p\rangle _z = \frac {1}{\tau _p}(\langle w_f^2\rangle _z - \langle w_s^2\rangle _z - \langle w_p^2\rangle _z) - 2\langle w_p\rangle _zg. \end{equation}

Putting (2.15) and (2.16) together, we get an equation for the mean squared slip velocity,

(2.17) \begin{equation} \langle w_s^2\rangle _z = \langle w_f^2\rangle _z - \langle w_p^2\rangle _z -2\langle w_p\rangle _z v_g-\frac {\tau _p}{\varrho }\frac {{\textrm d}}{{\textrm d}z}\varrho \langle w_p^3\rangle _z. \end{equation}

Equation (2.17) indicates that the mean squared slip velocity variance is a function of the mean squared sampled velocity $\langle w_f^2\rangle _z$ , the mean squared particle velocity $\langle w_p^2\rangle _z$ , a drift due to the non-zero average vertical velocity $\langle w_p\rangle _z$ and the vertical derivative of the mean cubed particle velocity $\langle w_p^3\rangle _z$ , all of which are implicit functions of particle inertia and gravity.

At this point, an important distinction must be made. Though the above equation is valid for inertial particles settling through a TBL, it should be noted that when particles settle under gravity, the mean squared quantities are not equal to their variances in general. This arises from the non-zero average settling velocity. Thus, to derive a relationship between the variances, we must do a Reynolds decomposition of each term. The details of this process are omitted here, but can be found in Appendix A . The final relationship is

(2.18) \begin{equation} \langle {w^{\prime}_s}^2\rangle _z = \underbrace {\overbrace {\underbrace {\langle {w^{\prime}_f}^2\rangle _z - \langle {w^{\prime}_p}^2\rangle _z}_{(1)} + R_t}^{(2)} + R_g}_{(3)} \end{equation}

where $R_t$ and $R_g$ are defined as

(2.19) \begin{eqnarray} R_t &=& -\frac {\tau _p}{\varrho }\frac {{\textrm d}}{{\textrm d}z}\varrho \langle {w^{\prime}_p}^3\rangle _z ,\end{eqnarray}
(2.20) \begin{eqnarray} R_g &=& -\frac {\tau _p}{\varrho }\frac {{\textrm d}}{{\textrm d}z}\big(\varrho \langle w_p\rangle ^3_z + 3\varrho \langle w_p\rangle _z\langle {w^{\prime}_p}^2\rangle _z\big) + 2\langle w_p\rangle _z( \langle w_s\rangle _z -v_g), \end{eqnarray}

respectively. This model contains many terms and is quite complex, but it can be broken down into three parts, arranged in order of increasing problem complexity, and the level of complexity highlighted by the over and underbraces.

First, grouped under $(1)$ , are the terms that would appear for particles settling through homogeneous turbulence. In this limit, the slip velocity variance is determined exclusively by the difference between $\langle {w^{\prime}_f}^2\rangle _z$ and $\langle {w^{\prime}_p}^2\rangle _z$ . This is the case for particles both with and without gravity, since gravity and inertia implicitly modify these terms. Next, grouped under $(2)$ , are the terms that appear for particles dispersing vertically through a TBL in the absence of gravity. Note that all terms encompassed by $(1)$ are included in $(2)$ , but when considering those terms covered under $(1)$ in the context of a TBL, they gain implicit height dependence since their magnitudes vary with the distance from the boundary. Furthermore, at this level, a new term appears, denoted by $R_t$ . This term is proportional to the derivative of the product of the concentration and the particle velocity triple moment and increases with particle inertia. As ${Sv^+}\rightarrow 0$ , $\langle {w^{\prime}_p}^2\rangle _z$ approaches $\langle w_p^2\rangle _z$ , but $R_t$ remains, regardless. Finally, by incorporating gravity, the mean particle velocity is no longer zero, leading to a new term grouped under $(3)$ , denoted by $R_g$ . The quantities composing $R_g$ are explicitly dependent on both the inhomogeneity of the flow through the vertical derivative, and the non-zero particle settling velocity. For clarity of interpretation, the second term on the right-hand side of (2.20) is written in terms of $\langle w_s\rangle _z - v_g$ , which we can see from (2.14) is identical to ${\tau _p}/{\varrho }{\rm d}/{{\textrm d}z}\varrho \langle w_p^2\rangle _z$ . In summary, by considering the continuum equations for the first and second moment of the particle velocity, we have been able to derive an equation for the particle slip velocity. We have identified a hierarchy of terms that appear in homogeneous turbulence and TBLs with and without settling (grouped under $(1)$ ), those that appear in a TBL without settling (grouped under $(2)$ ) and those that appear in a TBL with settling (grouped under $(3)$ ). In the following section, we will identify the importance of these terms throughout the TBL.

3. Results

3.1. Tendencies governing the slip velocity variance

In this section, we consider vertical profiles of $\langle {w^{\prime}_s}^2\rangle _z$ , $\langle {w^{\prime}_p}^2\rangle _z$ , $\langle {w^{\prime}_f}^2\rangle _z$ and $R_g$ . Figure 2 shows the tendencies in (2.18) for several cases scaled by $u_\tau ^2$ . For the results presented in this section, $R_t$ was observed to be very small relative to the other terms in (2.18), and is omitted from figure 2. Note that all profiles of a given ${S{\kern-0.5pt}t}^+$ and ${Sv^+}$ in figure 2(b,e,h) and figure 2(c,f,i) when added together return the profiles shown in figure 2(a,d,g), as per (2.18). An example of this is shown in Appendix B . Each row corresponds to a different friction Stokes number highlighted on the left-hand side of the figure, while each curve on a given plot corresponds to a different value of ${S{\kern-0.5pt}v}^+$ . Overall, this figure highlights the dominant tendencies controlling the slip velocity variance as ${S{\kern-0.5pt}t}^+$ and ${S{\kern-0.5pt}v}^+$ are independently changed.

Figure 2. Controlling tendencies for the slip velocity variance according to (2.18) at $\textit{Re}_\tau =630$ . Panels (a), (d) and (g) show the normalized slip velocity variance for each ${S{\kern-0.5pt}v}^+$ , while (a–c), (d–f) and (g–i) are for a different value of ${S{\kern-0.5pt}t}^+$ (shown on the left-hand side of the figure). Panels (b), (e) and (h) show the (negative) velocity variance and the (positive) seen velocity variance. Panels (c), (f) and (i) show the contributions from $R_g$ . Note that $R_t$ is omitted from this figure as it is small across the entire domain relative to the other terms. All terms are normalized by $u_\tau ^2$ .

Generally, the behaviour of the slip variance as a function of the vertical coordinate is qualitatively similar between all cases considered, evident from figure 2(a,d,g). However, the magnitude of the slip variance for a given case varies throughout the domain, and becomes increasingly sensitive to ${S{\kern-0.5pt}v}^+$ when ${S{\kern-0.5pt}v}^+$ approaches unity. This is evident in 2(a), where we see a negligible changes when ${S{\kern-0.5pt}v}^+$ is varied between 0.025 and 0.25, but larger changes as it is increased to 0.8 and beyond. For particles of ${S{\kern-0.5pt}v}^+\sim {O}(1)$ , we expect that the vertical turbulent velocities in the logarithmic region (which themselves scale in magnitude with $u_\tau$ ) tend to be too weak on average to overcome the strong settling velocity, and this leads to increasingly larger slip velocities. Within the logarithmic layer, the slip variance at constant ${S{\kern-0.5pt}v}^+$ (curves of fixed colour) tends to increase rapidly between ${S{\kern-0.5pt}t}^+=10$ and ${S{\kern-0.5pt}t}^+=50$ , but more slowly between ${St^+}=50$ and ${S{\kern-0.5pt}t}^+=100$ . This occurs because the particle velocity variance rapidly decreases in magnitude towards zero, while the variance of the fluid velocity seen by the particle does not, evident by considering figure 2(b,e,h). However, very near the solid boundary (i.e. below $z/H=0.1$ ), the variance of the fluid velocity seen by the particle approaches zero, while the particle velocity variance remains finite. This is also where $R_g$ is relatively large, indicating that the primary terms that control the slip variance very near the wall are these terms and negative particle velocity variance.

Likewise, the slip variance at constant ${S{\kern-0.5pt}t}^+$ tends to increase most rapidly as ${S{\kern-0.5pt}v}^+$ surpasses unity. This is due to the fact that particles tend to settle out of locally correlated regions of turbulence faster than they would in the absence of settling, thus experiencing a higher variance in accelerations, and thus their slip velocity. Interestingly, there is some variation in the variance of the fluid velocity seen by the particle with ${S{\kern-0.5pt}v}^+$ . This dependence arises due to the preferential sampling of the fluid velocity field, and the mechanisms responsible for this are essentially the same as those responsible for $\langle w_f\rangle _z$ deviating from zero for an inertial particle (see detailed discussion in Bragg et al.(Reference Bragg, Richter and Wang2021a )).

Lastly, $R_g$ , shown in figure 2(c,f,i), are non-zero but are not leading order within the interior of the domain (note the change in the scale of the horizontal axes of these panels), though they are relatively important within the viscous sublayer. Note that above $z/H = 0.1$ , these terms are almost completely negligible aside from when ${S{\kern-0.5pt}v}^+=2.5$ . It is important to note that previous studies (Gerashchenko et al. Reference Gerashchenko, Sharp, Neuscamman and Warhaft2008; Lavezzo et al. Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010) have shown that the impact of fluid shear and wall-normal gravity on inertial particle accelerations is most pronounced in the buffer layer $z^+\lt 50$ . When normalized against the large -scale parameters of the flow (at $\textit{Re}_\tau =630)$ , this height corresponds to a height of roughly $z/H\lt 0.1$ , so it should be expected that it is $R_g$ that strongly influences the slip velocity variance in that region of the flow.

In summary, these profiles highlight the fact that within the logarithmic layer, the slip variance is primarily governed by the differences between the variance of the flow velocities sampled by the particles and the particle velocity variance, with contributions coming from $R_g$ when ${S{\kern-0.5pt}v}^+$ increases beyond unity. However, it is clear that higher-order moments of the continuum equations (encoded in $R_g$ ) may be subleading and negligible in most other cases. It is not until the viscous sublayer where the contribution from $R_g$ becomes significant in determining the slip variance. Furthermore, the subleading behaviour of $R_g$ in the logarithmic layer does not imply that the inhomogeneity of the turbulence is irrelevant. In fact, the remaining terms ( $\langle {w^{\prime}_f}^2\rangle _z$ and $\langle {w^{\prime}_p}^2\rangle _z$ ) may have implicit dependence on the inhomogeneity of the turbulence, and this will be discussed later.

Figure 3. The horizontal and vertical components of $\varphi _r^{(i)}$ (panels (a) and (b)) and $\varphi _s^{(i)}$ (panels (c) and (d)) for all cases in table 1 plotted against ${S{\kern-0.5pt}v}^+$ . The filled markers correspond to cases at $\textit{Re}_\tau =630$ , while the open-faced markers correspond to cases at $\textit{Re}_\tau = 315$ .

As the particles settle, they experience a mean vertical slip velocity according to (2.14), and they also experience a mean horizontal slip due to both the mean shear and the height dependent correlations between horizontal and vertical particle velocity fluctuations. Figure 3 illustrates the relative contribution of the slip fluctuations to the overall slip velocity by considering two metrics: the integrated relative slip, and the integrated slip variance. These metrics are defined as

(3.1) \begin{equation} \varphi _r^{(i)} = \frac {1}{D}\int _D\frac {\langle {u_{s,i}^\prime }^2\rangle _z}{\langle {u_{s,i}}\rangle _z^2+\langle {u_{s,i}^\prime }^2\rangle _z} {\textrm d}z,\quad \varphi _s^{(i)} =\frac {1}{D}\int _D\frac {\langle {u_{s,i}^\prime }^2\rangle _z}{u_\tau ^2} {\textrm d}z, \end{equation}

respectively, where $D$ is the vertical subregion of the domain between $z^+=50$ and $z/H = 0.75$ and $i = x,\,z$ (we have temporarily adopted this notation for compactness). We have chosen this subregion as it represents the logarithmic region of the flow – these bounds were chosen to eliminate edge effects from the upper boundary condition, and our results are not qualitatively affected by the choice of the bounds of integration. The integrated relative slip helps us to understand the relative importance of the slip fluctuations relative to the mean slip, while the integrated variance provides a simple metric to assess the average slip variance. The integrated relative slip is shown for the horizontal (streamwise) and vertical components of the slip variance (there is no mean slip in the spanwise, so this component is ignored for this discussion) in figures 3(a) and 3(b), while the components of the integrated slip variance are shown in figures 3(c) and 3(d). Each marker style corresponds to a different value for ${S{\kern-0.5pt}t}^+$ , and the results are plotted against ${S{\kern-0.5pt}v}^+$ to highlight the role of settling. Filled markers correspond to runs with $\textit{Re}_\tau = 630$ , empty markers correspond to runs with $\textit{Re}_\tau = 315$ , and the empty marker filled with a cross corresponds to the run with $\textit{Re}_\tau = 1260$ . Note that since this is an integrated quantity, there is necessarily no information regarding the vertical structure of the profiles. However, an analysis of the profiles themselves (not shown) would provide the same conclusions.

Figure 3 shows that fluctuations of the slip velocity are less important at larger values of ${S{\kern-0.5pt}v}^+$ . For example, for small ${S{\kern-0.5pt}v}^+$ (independent of ${S{\kern-0.5pt}t}^+$ ), the normalized variance is nearly unity, indicating that the slip variance induced by the turbulent fluctuations are the leading-order controller of the overall slip velocity for both the horizontal and vertical components. However, we can see that by increasing ${S{\kern-0.5pt}v}^+$ , there is a decrease in the relative slip variance for both components. The reason for this is clear from an examination of figures 3(c) and 3(d), which show the integrated slip variance in the horizontal and the vertical, respectively. We can see that in a bulk sense, the slip variance in both directions does not strongly vary with ${S{\kern-0.5pt}v}^+$ (except for perhaps particles with ${St^+}=10$ in the horizontal). The implication is that the strong decrease in the relative variance in figure 3(b) (the vertical component) does not come from a decrease in the magnitude of the slip variance itself, but instead a strong increase in the magnitude of the mean slip induced by gravitational settling. Moreover, the same mechanism does not occur in the horizontal slip variance, as the decrease in the normalized slip variance is not nearly as strong. Furthermore, as ${S{\kern-0.5pt}v}^+$ increases, particles with ${S{\kern-0.5pt}t}^+ =10$ have the largest change in the vertical relative slip variance due to their small slip variance values. This is indicative of the fact that these particles most faithfully follow the flow in the absence of gravity, and as a result, are most sensitive to the growing mean slip as ${S{\kern-0.5pt}v}^+$ increases.

We also briefly consider a comparison between several Reynolds numbers. There are some slight differences in these metrics as the Reynolds number is varied between 315 and 1260. Here, changes in the relative slip variance are more strongly reflected in the vertical component. As the Reynolds number is increased, the relative slip variance in the vertical, figure 3(b), decreases. Since there are only small variations in the integrated slip variance as the Reynolds number is increased, the change in the integrated relative variance come from changes in the mean slip. The explanation for this change is as follows. Appearing in the equation of the mean slip, (2.14), is the term ${\tau _p}/{\varrho }({\rm d}/{{\textrm d}z})\varrho \langle w_p^2\rangle _z$ , which represents the combined effects of turbophoresis and diffusion-like behaviour near the solid boundary. It is known from Bragg et al. (Reference Bragg, Richter and Wang2021a ) that this term is negative within the logarithmic region of the flow, representing a mechanism that decreases the slip velocity. However, this term tends to be significant only near the solid boundary for a given $\tau _p$ , so as the Reynolds number increases, the relevance of this term is diminished when integrated across the region $D$ . Therefore, in the integrated sense, $\langle w_s\rangle _z^2$ tends to increase towards $v_g^2$ when $\textit{Re}_\tau$ increases which results in a decrease in $\varphi _r^{(3)}$ . However, this conclusion is only qualitative as more data at higher Reynolds numbers are necessary to make more quantitative conclusions.

The conclusion of figure 3 is that while the overall slip magnitude in the horizontal may be controlled by the slip fluctuations, the mean may end up being the main controller of the slip in the vertical at large ${S{\kern-0.5pt}v}^+$ and small to moderate ${S{\kern-0.5pt}t}^+$ , but the relative importance of the mean may be impacted by the Reynolds number of the flow.

3.2. Relationship to the acceleration statistics

The slip velocity statistics are directly related to the acceleration statistics of the particles. By considering the acceleration statistics, we can gain an understanding of how strongly gravity implicitly modifies the drag felt by the particles as they traverse the TBL. Moreover, the acceleration variance often gives a clue regarding the turbulent structures which particles are interacting with. For example, Yeo, Kim & Lee (Reference Yeo, Kim and Lee2010) showed that the elongated tails of fluid particle acceleration PDFs within the buffer layer and viscous sublayer are due to the vortical structures impinging on the viscous sublayer. Lavezzo et al. (Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010) attributed changes in the vertical and streamwise acceleration variance of settling inertial to these same vortical structures. In a two-way coupled turbulent Couette flow, Richter & Sullivan (Reference Richter and Sullivan2013) found that particles tend to damp vertical fluctuations of the near wall dynamics, suggesting a complex feedback cycle.

In the limit of ${g}^+\rightarrow 0$ (recall that $g^+={S{\kern-0.5pt}v}^+/{S{\kern-0.5pt}t}^+)$ , or when gravitational accelerations are ignored a priori, Bec et al. (Reference Bec, Biferale, Boffetta, Celani, Cencini, Lanotte, Musacchio and Toschi2006) demonstrated that particles tend to cluster in strain dominated regions of the flow for low Stokes number (i.e. ${St_\eta }\lesssim 0.3$ ), leading to a decrease in the acceleration variance of the particles. At larger Stokes number, the acceleration variance continues to decrease, but is instead due to inertial filtering; particles can no longer respond to turbulent fluctuations with time scales greater than $\tau _p^{-1}$ . Ultimately, both processes work to reduce the acceleration variance, but for different reasons (Bragg, Ireland & Collins Reference Bragg, Ireland and Collins2015). However, by introducing gravity, particles can settle out of strain-dominated regions of the flow, which may actually contribute to an increase in their acceleration variance, and this often referred to as the crossing trajectory mechanism (Csanady Reference Csanady1963). Ireland et al. (Reference Ireland, Bragg and Collins2016) and Berk & Coletti (Reference Berk and Coletti2021a ) showed that the importance of the crossing trajectories mechanism on the acceleration statistics is due to both ${St}_\eta$ and ${S{\kern-0.5pt}v}_\eta$ (or alternatively $1/\textit{Fr} = g/a_\eta$ ). They showed that for large $g/a_\eta$ (equivalent to large $g^+$ in our context), gravitational accelerations become increasingly important to the dynamics, leading to a peak in the acceleration variance at sufficiently high $g/a_\eta$ , around ${St}_\eta ={O}(1)$ . In the following results, we highlight some similarities of the computed slip and acceleration variance to results from Berk & Coletti (Reference Berk and Coletti2021), who focused on modelling the slip and acceleration variance in homogeneous turbulence in terms of the variance of the fluid along the particle trajectory. Our goal is to compare our model results in a TBL with the predictions of their model for homogeneous turbulence (derived in Appendix C).

Figure 4. Panel (a) shows the slip variance normalized by the seen variance for all cases in table 1 at $\textit{Re}_\tau = 630$ . The horizontal dashed lines denote heights of $z^+=50$ and $z/H = 0.75$ . Panel (b) shows the normalized acceleration variance over the same range plotted against the local value of ${St}_\eta$ , while the dashed line represents the ${St}_\eta ^{-2}$ scaling. The colours of each curve in panels (a) and (b) correspond to values of ${S{\kern-0.5pt}t}^+$ , while the line styles correspond to values of ${S{\kern-0.5pt}v}^+$ . Panel (c) shows ratio of the seen variance to the unconditional variance averaged over the entire vertical extent plotted against ${S{\kern-0.5pt}v}^+$ for all cases at $\textit{Re}_\tau = 630$ .

First, we consider profiles of the relative slip variance, which we define as $\langle {w_s^\prime }^2\rangle _z/\langle {w_f^\prime }^2\rangle _z$ , shown in figure 4(a). Here we again focus on the region $D$ , which is the region between $z^+=50$ and $z/H = 0.75$ , in order to omit effects from viscous sublayer and the upper boundary condition, respectively (denoted by black dashed lines on the figure). The general trend is that by increasing ${S{\kern-0.5pt}v}^+$ at a given ${S{\kern-0.5pt}t}^+$ , the relative slip variance tends to increase, with the most dramatic increase coming as ${S{\kern-0.5pt}v}^+$ is increased beyond unity, which is probably a reflection of the crossing trajectories mechanism. Furthermore, we can see that relative change between ${S{\kern-0.5pt}v}^+=0.25$ and ${Sv^+}=2.5$ decreases as ${S{\kern-0.5pt}t}^+$ increases. The reason for this is discussed more below.

We can also consider the relative acceleration variance, $\langle {a_p^\prime }^2\rangle _z/\langle {w_f^\prime }^2\rangle _z\tau _\eta ^{-2}$ , plotted against the local value of ${St}_\eta$ , shown in figure 4(b). We can see that as the range of ${St}_\eta$ increases (by changing ${S{\kern-0.5pt}t}^+$ ), the relative acceleration variance approaches the asymptotic relationship ${St}_\eta ^{-2}$ , i.e. is a decreasing function of the particle stokes number. So, in spite of the slip variance increasing with particle inertia at fixed ${S{\kern-0.5pt}v}^+$ (evident in figure 4 a), it does so at a slower rate than ${St}_\eta ^{2}$ , meaning that the particle acceleration variance decreases with particle inertia. This agrees with the work of Lavezzo et al. (Reference Lavezzo, Soldati, Gerashchenko, Warhaft and Collins2010), who showed that the vertical particle acceleration variance decreased as ${S{\kern-0.5pt}t}^+$ was increased at fixed ${S{\kern-0.5pt}v}^+$ . Note that in their study, they considered a fixed $g^+={S{\kern-0.5pt}v}^+/{S{\kern-0.5pt}t}^+$ , but the conclusions presented here are the same. However, for moderately inertial particles, characterized by the ${S{\kern-0.5pt}t}^+=10$ cases, there is more potential for gravity to increase the relative acceleration variance. For example, due to the crossing trajectories mechanism, we can see that when ${S{\kern-0.5pt}v}^+ = 2.5$ for these particles, the acceleration variance is much larger, and tends to scale as ${St}_\eta ^{-1}$ , which is consistent with the results from Balachandar (Reference Balachandar2009) when $\tau _\eta \ll \tau _p \ll \tau _{l,p}$ and from Berk & Coletti (Reference Berk and Coletti2021) in roughly the same range of ${St}_\eta$ . From figure 4(b), we can see that the crossing trajectories mechanism is not strong for large ${St^+}$ particles, since these particles approach the asymptotic ${St}_\eta ^{-2}$ scaling across the entire TBL. The implication here is that extremely inertial particles tend not to respond to high frequency and intermittent turbulent fluctuations associated with changes in the sampled fluid environment anywhere across the TBL. However, when particles are moderately sized, such that they achieve ${St}_\eta \sim 1$ there is a region within the logarithmic layer of the TBL where they become susceptible to crossing trajectory effects, and this leads to an increase in their relative acceleration variance.

As a final point on this discussion, a potential shortcoming of analysing the relative slip variance and the relative acceleration variance is that they are written in terms of the fluid velocity variance along the particle trajectory, which is an unknown quantity a priori. We can relate this to the unconditional variance, for which there are well-known models (see Kunkel & Marusic (Reference Kunkel and Marusic2006) for example). In figure 4(c) we show the vertically integrated ratio of the vertical components of the fluid velocity variance along particle trajectories to the unconditional fluid velocity variance against ${S{\kern-0.5pt}v}^+$ . This integrated ratio approaches unity as ${S{\kern-0.5pt}v}^+$ increases implying that the seen variance approaches the unconditional variance in this limit. Moreover, this ratio is no less than roughly 0.8 for our range of parameters, implying an acceptable correspondence between these two quantities. This is significant for modelling purposes as our results show that the fluid velocity variance along the particle trajectory may be substituted for the seen variance in a TBL without incurring significant error, having implications for the predictive power the particle statistics in a TBL. Berk & Coletti (Reference Berk and Coletti2021) also arrived at this conclusion in homogeneous turbulence.

We can also consider the impact of Reynolds number on the relative acceleration variance, as well as the averaged components of (2.18). We can see in figure 5(a) the relative variance decreases as a function of ${St}_\eta$ , as we would expect based on figure 4, but it is interesting that increasing $\textit{Re}_\tau$ further decreases the relative acceleration variance. The reason for this is shown in figure 5(b). Both the the averaged fluid velocity variance seen by the particles (filled markers), and the particle velocity variance (empty markers; black outline) increase with $\textit{Re}_\tau$ , but since the particle velocity variance increases faster with $\textit{Re}_\tau$ , the net effect is a decrease in the slip variance with increasing Reynolds number (and consequently the relative acceleration variance since they are related through $\tau _p$ ). The increases in $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ with Reynolds number when averaged across the domain is probably due to the increasing size of the quasihomogeneous region of the flow (Kunkel & Marusic Reference Kunkel and Marusic2006). Models of the unconditional vertical fluid velocity variance suggest that this quantity asymptotically approaches a constant in the limit of high Reynolds number. Though our Reynolds numbers are still quite low with regards to those found in the atmospheric surface layer, this notion can still provide some guidance to interpreting our data. As both $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ are related to the unconditional fluid velocity variance in some way, we expect that they should follow this behaviour, at least qualitatively. Finally, $R_g$ and $R_t$ are non-zero within the interior of the domain, but their magnitude, discussed later, is secondary to both $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ when taking the average across the logarithmic layer (recall the $D$ does not include the viscous sublayer).

Figure 5. Panel (a) shows the normalized acceleration variance for cases with ${S{\kern-0.5pt}t}^+=10$ and ${S{\kern-0.5pt}v}^+=0.8$ at three different Reynolds numbers as a function of ${St}_\eta$ . Panel (b) shows the seen variance (filled markers), slip variance (coloured empty markers) and particle velocity variance (black markers) normalized by the unconditional variance integrated over the range $D$ as a function of $\textit{Re}_\tau$ .

To conclude this section, we comment on the applicability of the model proposed by Berk & Coletti (Reference Berk and Coletti2021) for $\langle {w_s^\prime }^2\rangle _z$ (which we will refer to as BC2021 within the text). A sketch of the derivation of their model is presented in Appendix C . In short, they invoke an argument from Csanady (Reference Csanady1963) to relate the particle velocity variance to the seen fluid energy spectrum. They then use the fact that the fluid energy spectrum along the particle trajectory is the Fourier transform of the autocorrelation of the fluid velocity along the particle trajectory, which they represent using a two time scale model derived by Sawford (Reference Sawford1991). The autocorrelation function involves the decorrelation time scale of the turbulence along the particle trajectory, $\tau _{l,p}$ , for which they use the model derived in Csanady (Reference Csanady1963). This model assumes that the turbulence is homogeneous and isotropic, and thus particles experience no spatial change in the statistics of the turbulence along their trajectory. Our goal is to compare how the computed relative slip variance within the TBL compares with the modelled slip variance in BC2021. As BC2021 was developed under the assumption of homogeneous turbulence, we can extend it to a TBL by making a locally homogeneous approximation, meaning that any change the slip variance with height is occurs due to local changes in the turbulent dissipation. For brevity, we will refer to the slip variance modelled by BC2021 as $BC$ . We first consider (2.18) normalized by the seen variance,

(3.2) \begin{equation} \frac {\langle {w_s^\prime }^2\rangle _z}{\langle {w_f^\prime }^2\rangle _z} = \overbrace {\underbrace {1 - \frac {\langle {w_p^\prime }^2\rangle _z}{\langle {w_f^\prime }^2\rangle _z}}_{\text{$PV$}} + \frac {R_t}{\langle {w_f^\prime }^2\rangle _z} + \frac {R_g}{\langle {w_f^\prime }^2\rangle _z}}^{\text{$M$}}. \end{equation}

We have highlighted two subsets of these terms. First, we denote the first two terms on the right-hand side of (3.2) as $PV$ (for ‘particle velocity’). Our focus on this subset of terms is motivated by the BC2021 model (derived in Appendix C). The model BC2021 was developed for settling particles in homogeneous turbulence, meaning that there are no spatial variations in the turbulent statistics. This implies that $R_g$ and $R_t$ play no role in the dynamics, which leaves behind only the $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ terms. We also denote the full right-hand side of (3.2) as $M$ (for ‘model’). As we have discussed previously, our analysis will be limited to the logarithmic region of the flow. We also comment (but do not show) that the variance predicted by BC2021 is smaller in magnitude than the slip variance computed by the DNS throughout this region.

First, we consider how BC2021 compares with the full right-hand side of (3.2), (denoted by $M$ ) which includes $R_t$ and $R_g$ (recall that $R_t$ and $R_g$ cannot appear in BC2021 a priori due to their assumption of homogeneous turbulence). Specifically, we consider the metric

(3.3) \begin{equation} \xi _1 = \frac {1}{D}\int _D\|M - BC\| {\textrm d}z, \end{equation}

which is plotted in figure 6(a). Here, $\xi _1$ represents the difference between the DNS data (which includes $R_t$ and $R_g$ ) and the BC2021 model averaged over the region $D$ . The main point here is that the differences between the BC2021 and $M$ are relatively small. This should be the case since the buffer layer and viscous sublayer where $R_t$ and $R_g$ are expected to be most important has been omitted in $D$ .

Figure 6. Shown in panels (a)–(d) are $\xi _1$ , $\xi _2$ and the root mean squares (r.m.s.) values of $R_t$ and $R_g$ normalized by the seen variance for all cases in table 1, respectively. Open-faced markers represent cases with $\textit{Re}_\tau = 315$ , filled markers represent cases with $\textit{Re}_\tau = 630$ and the marker with an $\otimes$ ${Re_\tau }=1260$ .

This motivates our second comparison between $PV$ and the BC2021 model ( $BC$ ); mathematically, we consider

(3.4) \begin{equation} \xi _2 = \frac {1}{D}\int _D\|PV - BC\| {\textrm d}z, \end{equation}

which is shown in figure 6(b). Here, $\xi _1$ represents the difference between the DNS data omitting $R_t$ and $R_g$ and the BC2021 model averaged over the region $D$ . This metric highlights a disparity between the DNS data and BC2021, but the differences amount to no more than roughly $0.3$ . This discrepancy may come from several sources; one may be due to the fact that the underlying statistics of the turbulent flow change along the particle’s trajectory due to the presence of the wall. This behaviour is reflected in the general increase of $\xi _2$ as ${S{\kern-0.5pt}v}^+$ increases, and will be discussed more in § 4. It is also interesting to note that there are differences associated with ${S{\kern-0.5pt}t}^+$ , and these differences tend to plateau at large ${S{\kern-0.5pt}t}^+$ for fixed ${S{\kern-0.5pt}v}^+$ .

Interestingly, from figures 6(a) and 6(b), it appears that the inclusion of $R_t$ and $R_g$ (which did not occur in BC2021) actually work to decrease the associated differences between the DNS and BC2021. However, as we established figure 2(c,f,i), $R_g$ is primarily negative, causing a reduction of the relative slip variance. As we noted previously, $M$ underestimates the computed relative slip variance within the logarithmic region of the turbulence, and the result is a decrease in the differences between the BC2021 and the DNS data. Moreover, by considering the r.m.s. of $R_t$ and $R_g$ normalized by $\langle {w^{\prime}_f}^2\rangle _z$ , shown in figures 6(c) and 6(d), we can see that $R_g$ more important tendency, rather than $R_t$ . Again, the importance of this term appears when both ${S{\kern-0.5pt}t}^+$ and ${S{\kern-0.5pt}v}^+$ are large, but decreases significantly as $\textit{Re}_\tau$ increases.

Moreover, there are likely to be differences between the DNS data and BC2021 associated with the relatively small Reynolds numbers considered in this study. Within BC2021, there are several semiempirical models required to calculate characteristic parameters (i.e. $C_0$ and $a_0$ ; see Appendix C) of the turbulence, and there may be an associated error incurred when the Reynolds number is low (for example, see the discussion in Lien & D’Asaro (Reference Lien and D’Asaro2002)). However, we can see that by increasing $\textit{Re}_\tau$ , the differences between the DNS and BC2021 tend to decrease suggesting a correspondence at large enough Reynolds number.

In summary, BC2021 gives a reasonable estimate for the DNS data in the logarithmic region of the flow. However, at large ${St^+}$ and ${S{\kern-0.5pt}v}^+$ , the size of the differences increases, and this is due to combination of the changing statistics of the turbulence along the particle trajectories, as well as the low Reynolds numbers in the DNS, and how these are represented within BC2021. However, an important conclusion is that at large Reynolds number, we expect correspondence between the DNS data and BC2021, evidenced by the fact that the differences between BC2021 and the DNS decrease in this limit. Moreover, outside of the viscous sublayer, the importance of $R_t$ and $R_g$ will also be reduced as $\textit{Re}_\tau$ increases, meaning correspondence between predictions by the BC2021 model and the variance measured in a TBL will become stronger in this limit. These results suggest that when ${S{\kern-0.5pt}t}^+$ and ${S{\kern-0.5pt}v}^+$ are not too large, we can estimate the slip variance in the logarithmic region of the flow by using BC2021 without incurring significant error.

4. Summary and discussion

4.1. Summary

Motivated by coarse particle transport in the atmospheric surface layer, we used coupled Eulerian–Lagrangian simulations to simulate the dynamics of ensembles of inertial particles in boundary layer turbulence. We examined the impact of particle inertia and settling on the mean and fluctuating particle slip velocity. We adapted a mathematical model discussed in Bragg et al. (Reference Bragg, Richter and Wang2021a ) and Johnson et al. (Reference Johnson, Bassenne and Moin2020) for the slip velocity variance for settling inertial particles in a TBL and highlighted the controlling factors throughout the domain. We showed that to leading order, the slip variance above of the viscous sublayer was determined by the difference between the seen variance, $\langle {w_f^\prime }^2\rangle _z$ , and the particle velocity variance $\langle {w_p^\prime }^2\rangle _z$ , except for the largest value of ${S{\kern-0.5pt}v}^+$ , where all terms in (2.18) became comparable. Consequently, as changes in the seen variance were relatively small, changes in the slip variance within the logarithmic layer were primarily governed by a decrease of the particle velocity variance, which was implicitly a function of particle inertia and the particle settling velocity (more on this below). Within the viscous sublayer, the balance became more complicated. In all cases, $\langle {w_f^\prime }^2\rangle _z$ tended towards zero to adhere to the no-slip condition enforced at the bottom boundary. However, the slip variance remained finite as the particles tended towards $z^+ = 0$ as $\langle {w_p^\prime }^2\rangle _z$ , $R_t$ and $R_g$ remained finite. The higher-order terms tended to peak within this layer, and the magnitude of the peak tended to increase with ${S{\kern-0.5pt}v}^+$ . However, by using domain averages, we demonstrated that the relative magnitude of the higher-moment terms tended to decrease as the Reynolds number increased, reflecting the fact that at higher Reynolds number, the viscous sublayer becomes much thinner, leading to a smaller contribution when averaged across the domain. As discussed above, these terms may still be important within the viscous sublayer depending on particle parameters, though.

We also showed that the fluid velocity variance along the particle trajectories exhibited only small changes with ${S{\kern-0.5pt}t}^+$ , ${S{\kern-0.5pt}v}^+$ and $\textit{Re}_\tau$ , and was approximately 80 % of the unconditional fluid velocity variance when averaged across the domain. The differences between the seen and unconditional variances are largest at the smallest ${S{\kern-0.5pt}v}^+$ considered in this work, though the differences are still relatively small. These differences are likely a result of the relatively large spectrum of turbulent motions at high Reynolds number, and the impact of crossing trajectories, which works to implicitly affect the seen variance (see the discussion in Csanady (Reference Csanady1963), for example). This conclusion is quantitatively consistent to the conclusions presented in Berk & Coletti (Reference Berk and Coletti2021) for their laboratory experiments in homogeneous turbulence. This correspondence is significant as the seen variance is not known a priori. There exist approaches to modelling this quantity (Pozorski & Minier Reference Pozorski and Minier1998; Minier & Peirano Reference Minier and Peirano2001), but the results of our work suggest that even in a TBL where there is spatial dependence in the variance of the turbulent quantities, we can approximate the fluid variance along the particle trajectories with the unconditional fluid variance without introducing significant errors. While this reduces the overall accuracy of the slip variance estimate, it increases the predictive power at the field scale, as models for the unconditional variance, such as those discussed in Kunkel & Marusic (Reference Kunkel and Marusic2006), can be employed.

To examine the relative importance of the fluctuating and mean slip, we considered the ratio of the slip variance to the total mean squared slip velocity, $\langle {w_s}^2\rangle _z$ (i.e. the square of the mean plus the fluctuation). We found that relative to the mean, the vertical slip variance decreased much faster than the horizontal as ${S{\kern-0.5pt}v}^+$ was varied. However, for both components, we showed that the overall magnitudes in the average sense did not change significantly, indicating that at relatively large ${S{\kern-0.5pt}v}^+,$ the mean slip was the determining factor in the vertical, while the fluctuating slip was the determining factor in the horizontal. This effect was also accentuated for the smallest particles considered, due to their relatively small slip variance. To further complicate the behaviour, the relative size of the slip variance tended to decrease as the Reynolds number was increased, though due to computational restrictions, we can only provide limited guidance on this issue.

We also compared the slip variance computed by the DNS with a model derived for HIT by Berk & Coletti (Reference Berk and Coletti2021) (as no such model currently exists for a TBL and is the focus of future work). The main conclusion is that the globally averaged differences (in the absolute sense) were relatively small, but the higher moment terms act as a confounding factor to reduce differences between the model and the DNS data. Thus, care should be taken when extrapolating results from low Reynolds number DNS in TBLs to higher Reynolds number experiments in homogeneous turbulence. However, as we know the size of higher moment terms tends to decrease as Reynolds number increase when integrated across the domain, we expect that DNS at higher Reynolds number should tend towards the results derived in Berk & Coletti (Reference Berk and Coletti2021) outside of the thin viscous sublayer.

Additionally, due to the inhomogeneous nature of the turbulence, non-local effects implicit to $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ may occur at large ${S{\kern-0.5pt}v}^+$ and ${S{\kern-0.5pt}t}^+$ . Isolating the importance of non-local effects in a TBL is the focus on ongoing research (for a recent example for a model of $\langle {w_p^\prime }^2\rangle _z$ in a TBL, see Zhang, Bragg & Wang (Reference Zhang, Bragg and Wang2023) and references therein), but incorporating them in a model is beyond the scope of the current article. These effects may arise due to the fact that the statistics of the turbulence may change significantly as the particle travels vertically. For example, consider the distance a settling particle travels over one relaxation time: $\delta \sim |\tau _p\langle w_p\rangle _z|$ , where $\langle w_p\rangle _z$ is the average particle settling velocity conditioned on a height $z$ given by (2.13). In order for the particle trajectory to be altered, turbulent fluctuations must be correlated over this distance. However, if this distance is comparable to the distance over which the characteristics of the turbulence change, then we expect that the particle feels the inhomogeneous nature of the flow. To formalize this quantitatively, consider the local turbulent kinetic energy at a height $z$ , $k$ . Taylor expanding about this point and truncating after the second term, we have

(4.1) \begin{equation} k(z-\delta ) = k(z)- \left .\delta \frac {{\textrm d}k}{{\textrm d}z}\right |_z. \end{equation}

To make a locally homogeneous approximation about the turbulence, we must have that

(4.2) \begin{equation} k(z)\gg \left .\delta \frac {{\textrm d}k}{{\textrm d}z}\right |_z, \end{equation}

i.e. the kinetic energy in a small neighbourhood about $z$ (defined by the distance $\delta$ ) is primarily defined by the kinetic energy measured at a height $z$ . Using (4.1), we can estimate under what conditions a locally homogeneous approximation would be appropriate by looking for cases where the second term is small compared with the first. By assuming that the gradient of the turbulent kinetic energy scales as $u_\tau ^2z^{-1}$ (also implying $k$ can be scaled by $u_\tau ^2$ ) (Smits et al. Reference Smits, McKeon and Marusic2011) we have that $z\gg \delta$ . By normalizing both sides of this inequality by the r.m.s. turbulent velocity, $u^\prime = k^{1/2}$ , we can write $\delta$ in terms of the sum of the settling enhancement, $E = ({\langle w_p\rangle _z + v_g})/{u^\prime }$ and $v_g$ , (Good et al. Reference Good, Ireland, Bewley, Bodenschatz, Collins and Warhaft2014; Loth Reference Loth2023) as

(4.3) \begin{equation} \tau _p\left |E + \frac {v_g}{w^\prime }\right | \ll \frac {z}{u^\prime }. \end{equation}

Now normalizing by $\tau _{\eta }$ , and observing that $u^\prime \sim u_\tau$ , we can simplify both side of this inequality to reveal that

(4.4) \begin{eqnarray} {St}_\eta \left |E + {S{\kern-0.5pt}v}_\ell \right | \ll \left (\frac {zu_\tau }{\nu }\right )^{1/2}, \end{eqnarray}

where we have used the dissipation scaling in (2.4) to relate $\tau _\eta$ to the vertical coordinate, $z$ . This relationship indicates that both particle inertia and gravity have an explicit role, and an implicit role (through $E$ ) to play in potential non-local effects.

We know from Good et al. (Reference Good, Ireland, Bewley, Bodenschatz, Collins and Warhaft2014) and Loth (Reference Loth2023) that in small scale laboratory experiments, $E\sim 0.2$ as ${St}_\eta$ and ${S{\kern-0.5pt}v}_\ell$ approach unity, but as both of these parameters increase, $E$ tends back towards zero, and may even become negative (Ferran et al. Reference Ferran, Machicoane, Aliseda and Obligado2023). Therefore, for the coarse particles we are concerned with in this work, we can make a locally homogeneous approximation when ${St_\eta Sv_\ell }\ll ({zu_\tau }/{\nu } )^{1/2}$ . This may not particularly restrictive for the atmospheric surface layer as the Reynolds numbers are ${O}(10^6)$ , but for laboratory experiments, the integral scales tend to scale with the size of the experimental domain (i.e. $z\sim h$ where $h$ could be the half-height of a channel). This could present a problem making a locally homogeneous approximation for coarse particles in a wind tunnel set-up.

For the DNS presented in this work, the above relationship shows that non-local effects are likely only important for cases when both ${St^+}$ and ${S{\kern-0.5pt}v}^+$ are large, as ${S{\kern-0.5pt}v}_\ell$ tends to scale with ${S{\kern-0.5pt}v}^+$ since $u^\prime \sim u_\tau$ . For example, if we consider cases with ${St^+}=100$ and ${S{\kern-0.5pt}v}^+ = 2.5$ (and assume $E\approx 0$ ), we can see immediately that (4.4) is not satisfied. This may explain the differences between the DNS and the model in Berk & Coletti (Reference Berk and Coletti2021) in figures 6(a) and 6(b) for these particles. One of the main conclusions of this work is that for the governing continuum equation for the particle slip velocity in a TBL, there are tendencies that arise due to the inhomogeneities in the turbulence associated with the presence of the wall and the fact that the particle settling velocity is non-zero. However, as we have shown, for moderately sized particles (characterized by ${S{\kern-0.5pt}t}^+$ or ${St}_\eta$ ), and ${S{\kern-0.5pt}v}^+\lt 1$ , these terms are subleading outside of the viscous sublayer. Moreover, the magnitudes of these terms in the logarithmic layer tend to diminish as $\textit{Re}_\tau$ increases. Thus, outside of the viscous sublayer, and at moderate local ${St}_\eta$ and ${S{\kern-0.5pt}v}^+$ , we can extend models designed for homogeneous turbulence (like that described in Berk & Coletti (Reference Berk and Coletti2021)) to a TBL, where we must interpret the model as local to a height $z$ . However, outside of this regime (i.e. for very large and strongly settling particles), there are implicit non-local effects that appear as particles tend to settle through the flow due to the vertical variation of the turbulent statistics along the particle trajectories.

4.2. Implications for modelling coarse particle transport in the atmospheric surface layer

Interpreting DNS results in terms of the laboratory or field scales must be done with care, as the Reynolds numbers in DNS numbers are much smaller than those found at these scales. However, by scaling up the results in this work, we can gain valuable qualitative insights into the drag on inertial settling dust particles. For example, using estimates of turbulent dissipation for an atmospheric surface layer of roughly $10^{-3}$ $\mathrm{m^2\, s^{-3}}$ , we can define a rough Kolmogorov time scale as $10^{-1}$ s. Thus, for quartz dust particles ( $\rho _p = 2650$ $\mathrm{kg\,m^{-3}}$ ) that range between 30–100 μm, we should expect a value of ${St}_\eta$ to range between 0.1–10. We can see that the ranges in our DNS are in the correct neighbourhood to model these same coarse dust particles. Moreover, we can use the values of $g^+$ from table 1 (recall $g^+ = {Sv^+/St^+}$ ) to estimate an equivalent friction velocity, $u_\tau ^*$ (note that since we are rescaling, $u_\tau ^*$ is necessarily different than the value of $u_\tau$ used in this work), which effectively gives us a qualitative estimate of the intensity of the turbulence in an atmospheric surface layer. The effective friction velocity is given by

(4.5) \begin{equation} u_\tau ^* = \left (\frac {g\nu }{g^+}\right )^{1/3}. \end{equation}

Assuming $g = 9.81\mathrm{\,m\,s^{-2}}$ and $\nu = 1.57 \times 10^{-5}\mathrm{\,m^2\, s^{-1}}$ , and some relationship between 10 m wind velocity and the friction velocity (see Kantha & Clayson (Reference Kantha and Clayson2000) for example), this gives us a proxy for wind speed at the field scale. For the values of $g^+$ in this manuscript (see table 1), the effective friction velocities vary between 0.09 $\mathrm{m\,s^{-1}}$ and 0.84 $\mathrm{m\,s^{-1}}$ , which covers a wide range of friction velocities on the Earth (Vickers, Mahrt & Andreas Reference Vickers, Mahrt and Andreas2015). Since $g^+$ is proportional to ${S{\kern-0.5pt}v}^+$ , we can see the effective wind speed increases as ${S{\kern-0.5pt}v}^+$ decreases.

Therefore, the insight we can gain is that the slip velocity in high wind conditions (small ${S{\kern-0.5pt}v}^+)$ should be primarily governed by the fluctuations associated with the turbulence, as opposed to the mean induced by gravitational settling and the presence of the solid boundary. Conversely, at lower wind speeds, the drag induced by turbulent fluctuations is much smaller relative to the mean slip. Thus, the magnitude of the slip velocity should instead be controlled by the average, which itself is controlled by the Stokes settling velocity and the turbophoretic term. Likewise, we expect the higher moment terms governing the slip variance (i.e. $R_t$ and $R_g$ ) to be more important to the dynamics farther away from the surface in this limit, relatively speaking.

This is significant when applying models like BC2021 to particle transport in field scale systems. For example, as we have described previously, BC2021 can be used (in conjunction with a model for the unconditional fluid velocity variance) to predict the slip velocity variance for inertial settling particles in homogeneous turbulence. Under low ${S{\kern-0.5pt}v}^+$ conditions, our results show that the magnitude of the slip velocity is primarily governed by its fluctuating component, which is in turn associated with the interactions with the turbulence. Moreover, as we have discussed, a locally homogeneous approximation may be used when ${S{\kern-0.5pt}v}^+$ is small enough (see (4.4)), our work suggests that BC2021 can also be applied to inhomogeneous turbulence, like that of the atmospheric surface layer, provided we are not concerned with dynamics too close to the ground and the wind conditions are strong enough. Another interesting related application is towards modelling the particle Reynolds number, which is known to affect the associated drag on the particles (Balachandar Reference Balachandar2009; Berk & Coletti Reference Berk and Coletti2024). For example, it is known that loitering effects are typically associated with large particle Reynolds number (Rosa et al. Reference Rosa, Parishani, Ayala and Wang2016), and these loitering effects work to reduce the average particle settling velocity (Good et al. Reference Good, Ireland, Bewley, Bodenschatz, Collins and Warhaft2014). Accurate modelling of loitering effects could explain discrepancies between numerical simulations and laboratory experiments with respect to the measurement of settling velocities (Ferran et al. Reference Ferran, Machicoane, Aliseda and Obligado2023). Moreover, our results may gain some insights into further than expected horizontal transport of giant dust particles off of the West African coast (Van Der Does et al. Reference van der Does, Knippertz, Zschenderlein, Giles Harrison and Stuut2018), which could be linked to loitering effects.

Funding.

The authors would like to acknowledge grant no. W911NF2220222 from the U.S. Army Research Office. The authors would also like to thank the Center for Research Computing at the University of Notre Dame, and the three anonymous reviewers whose comments have increased the overall quality of the work.

Declaration of interests.

The authors report no conflict of interest.

Appendix A. Mathematical details of slip velocity model hierarchy

By including gravitational settling in the particle equation of motion, there is now a mean settling velocity. Due to preferential sweeping, the average particle settling velocity can be increased (or decreased in some cases) beyond the laminar settling velocity, $v_g$ , leading to there is a non-zero average slip velocity. Since we know that the average settling velocity of the particles will be non-zero due to the presence of gravity, the average of the square is not equivalent to the variance, i.e. $\langle F^2\rangle \neq \langle {F'}^2\rangle$ , ( $F$ is some arbitrary quantity, and a prime indicates a fluctuation about the mean of $F$ ) meaning we must be careful to discern between the variance and squared means,

(A1) \begin{align} \langle w_p^2\rangle _z = \langle w_p\rangle _z^2 + \langle {w^{\prime}_p}^2\rangle _z, \end{align}
(A2) \begin{align} \langle w_s^2\rangle _z = \langle w_s\rangle _z^2 + \langle {w^{\prime}_s}^2\rangle _z,\quad \langle w_s\rangle _z = \langle w_f\rangle _z - \langle w_p\rangle _z, \end{align}
(A3) \begin{align} \langle w_p^3\rangle _z = \langle w_p\rangle _z^3 + \langle w_p\rangle _z\langle {w^{\prime}_p}^2\rangle _z + \langle {w^{\prime}_p}^3\rangle _z. \end{align}

Equation (2.17), derived in Johnson et al. (Reference Johnson, Bassenne and Moin2020), assumed that particles did not settle under the action of gravity, meaning that $\langle w_p\rangle _z = 0$ . However, by substituting in the above Reynolds decompositions, it can be readily extended to settling particles. Doing this, we arrive at

(A4) \begin{equation} \def\lumina\hspace*{-10pt} \langle {w^{\prime}_s}^2\rangle _z = \langle {w^{\prime}_f}^2\rangle _z {-} \langle {w^{\prime}_p}^2\rangle _z{-}\frac {\tau _p}{\varrho }\frac {d}{{\textrm d}z}\varrho \langle w_p^3\rangle _z + \langle w_f\rangle _z^2 -\langle w_s\rangle _z^2 - \langle w_p\rangle _z^2 - 2\langle w_p\rangle _zv_g, \end{equation}

where we have not expanded $\langle w_p^3\rangle _z$ in terms of its variance and mean components yet. We can see from this equation that the slip velocity variance is due to the variance of seen velocities, the variance of the particle velocity, and several other terms. These terms are difficult to interpret in their current form, so we simplify them next.

The mean slip velocity squared, $\langle w_s\rangle _z^2$ can be expanded as

(A5) \begin{equation} \langle w_s\rangle _z^2 =\langle w_f\rangle _z^2 -2\langle w_f\rangle _z\langle w_p\rangle _z +\langle w_p\rangle _z^2. \end{equation}

Upon substitution of the above into (A4) and simplifying, we can express (A4) as

(A6) \begin{equation} \langle {w^{\prime}_s}^2\rangle _z = \langle {w^{\prime}_f}^2\rangle _z -\langle {w^{\prime}_p}^2\rangle _z -\frac {\tau _p}{\varrho }\frac {d}{{\textrm d}z}\varrho \langle w_p^3\rangle _z + 2\langle w_p\rangle _z\left ( \langle w_s\rangle _z -v_g\right ). \end{equation}

We can now expand the third term on the right-hand side of the above in order to express the slip velocity variance in terms of only means and variances of other quantities,

(A7) \begin{eqnarray} \def\lumina\hspace*{-2.5pc} \langle {w^{\prime}_s}^2\rangle _z &=& \overbrace {\underbrace {\langle {w^{\prime}_f}^2\rangle _z -\langle {w^{\prime}_p}^2\rangle _z}_{(1)} -\frac {\tau _p}{\varrho }\frac {d}{{\textrm d}z}\varrho \langle {w^{\prime}_p}^3\rangle _z}^{(2)}\nonumber \\[8pt]&& \def\lumina\hspace*{-1.5pc} \underbrace {-\frac {\tau _p}{\varrho }\frac {d}{{\textrm d}z}(\varrho \langle w_p\rangle ^3_z + 3\varrho \langle w_p\rangle _z\langle {w^{\prime}_p}^2\rangle _z ) + 2\langle w_p\rangle _z\left ( \langle w_s\rangle _z -v_g\right )}_{(3)}. \end{eqnarray}

Figure 7. Panel (a) shows the slip velocity variance computed by the DNS for three values of ${S{\kern-0.5pt}v}^+$ at fixed ${S{\kern-0.5pt}t}^+=10$ (black curves), and the slip variance computed by (2.18) (dashed coloured curves). Panel(b) shows $R_t + R_g$ computed via a residual of $\langle {w_s^\prime }\rangle _z - (\langle {w_f^\prime }\rangle _z - \langle {w_p^\prime }\rangle _z)$ (black curves), and $R_t+R_g$ computed directly (dashed coloured curves) for the same cases.

Appendix B. Comparison of computed and modelled slip variance

Figure 7(a) shows a comparison between computed values of $\langle {w_s^\prime }^2\rangle _z$ and that computed using the right-hand side of (2.18). The slip variance computed directly from the DNS for three different values of ${S{\kern-0.5pt}v}^+$ at ${S{\kern-0.5pt}t}^+=10$ are shown by black curves, while the right-hand side of (2.18) are shown by coloured dashed lines. We can see that the right-hand side contains significant noise, but otherwise models the computed slip variance well. This noise is a result of the routines used to estimate the derivatives in $R_t$ and $R_g$ , and not in the computation of $\langle {w_f^\prime }^2\rangle _z$ and $\langle {w_p^\prime }^2\rangle _z$ (as evidenced in figure 2). Figure 7(b) shows a comparison between the computed values of $R_t+R_g$ (dashed curves) and that computed by a residual of (2.18) (black solid curves). We can see that by plotting $R_t$ and $R_g$ as a residual, the noise is significantly reduced.

Appendix C. Response function model of acceleration variance

In this section, we sketch the derivation of the semianalytical model proposed by Berk & Coletti (Reference Berk and Coletti2021) for the slip velocity variance. The following is based on the concept of a response function, described in Csanady (Reference Csanady1963), which is meant to quantify the fact that inertial particles require a finite amount of time to respond to turbulent fluctuations. By Fourier-transforming the vertical component of the fluctuating particle velocity and the sampled velocity through the use of standard manipulations, we can write the slip velocity variance and the particle velocity variance as

(C1) \begin{equation} \langle {w_s^\prime }^2\rangle = \int _0^{\infty }(\omega \tau _p)^2 E_p\,{\textrm d}\omega , \quad \langle {w_p^\prime }^2\rangle = \int _0^{\infty }E_p\,{\textrm d}\omega . \end{equation}

In this equation, $E_p$ is the kinetic energy spectrum of the particles. As discussed in Csanady (Reference Csanady1963), $E_p$ is related to the kinetic energy spectrum of the fluid motion sampled along the particle’s trajectory, denoted by $E_{f,p}$ , through a response function

(C2) \begin{equation} H(\omega ) = \frac {1}{1 + (\omega \tau _p)^2}, \end{equation}

meaning that

(C3) \begin{equation} \langle {w_s^\prime }^2\rangle = \int _0^{\infty }(\omega \tau _p)^2H(\omega ) E_{f,p}\,{\textrm d}\omega ,\quad \langle {w_p^\prime }^2\rangle = \int _0^{\infty }H(\omega ) E_{f,p}\,{\textrm d}\omega . \end{equation}

Using the stochastic model for the particle velocity autocorrelation outlined in Sawford (Reference Sawford1991), Berk & Coletti (Reference Berk and Coletti2021) used the fact that the autocorrelation and the spectra are Fourier transform pairs. The particle velocity autocorrelation described in Sawford (Reference Sawford1991) is

(C4) \begin{equation} R_{f,p}(t) = \frac {\langle {w_f^\prime }^2\rangle }{\tau _{l,p} - \tau _2}\left (\tau _{l,p}e^{-t/\tau _{l,p}}+\tau _{2}e^{-t/\tau _{2}}\right ), \end{equation}

where $\tau _2$ is proportional to the fluid acceleration variance and appears due to the finite Reynolds number. For this work, we use

(C5) \begin{equation} \tau _2 = \frac {C_0}{2a_0}\tau _\eta , \end{equation}

where $C_0$ and $a_0$ are universal constants modelled by

(C6) \begin{equation} C_0 = C_\infty (1 - (0.1{Re_\lambda }^{-1/2})), \quad a_0 = \left (\frac {5}{1 + 110\textit{Re}_\lambda ^{-1}}\right ), \end{equation}

defined in Lien & D’Asaro (Reference Lien and D’Asaro2002) and Sawford et al. (Reference Sawford, Yeung, Borgas, Vedula, La Porta, Crawford and Bodenschatz2003), respectively. Here, $\textit{Re}_\lambda = \sqrt {15}\langle w^2\rangle /v_\eta ^2$ is the Taylor–Reynolds number evaluated at a height $z$ . Thus, $C_0$ and $a_0$ are functions of the vertical coordinate. Note that the results in this work are not meaningfully dependent on the exact choice of model for $C_0$ and $a_0$ .

Here $\tau _{l,p}$ is the Lagrangian correlation time scale of the turbulence along the particle trajectory. Here $\tau _{l,p}$ is a function of three parameters: ratio of the laminar settling velocity to the integral velocity scale; the Lagrangian correlation time scale of the turbulence; the Eulerian correlation time scale of the turbulence which are defined as

(C7) \begin{equation} {S{\kern-0.5pt}v}_\ell =\frac {v_g}{w^\prime },\quad \tau _E = \frac {\langle w^2\rangle }{\epsilon }, \quad \tau = -\frac {\kappa z}{u_\tau }\frac {\langle uw\rangle }{\langle w^2\rangle }, \end{equation}

respectively. Note that the definition of $\tau$ can be found in Oesterlé & Zaichik (Reference Oesterlé and Zaichik2004). Here $\tau _{l,p}$ is meant to encapsulate the fact that as an inertial particle settles through a local neighbourhood of correlated motion, the turbulence it experiences decorrelates faster along its trajectory than it would if it was not settling. Berk & Coletti (Reference Berk and Coletti2021) derived a semiempirical model for the correlation along the particle trajectory using the idea of the crossing trajectories mechanism introduced by Csanady (Reference Csanady1963) as

(C8) \begin{equation} \tau _{l,p} = \tau \frac {1}{\left (1 + \left (\frac {\tau }{\tau _E}\right )^2{S{\kern-0.5pt}v}_\ell ^2\right )^{1/2}}. \end{equation}

By Fourier-transforming $R_{f,p}$ and substituting into the integral relations for the slip variance and velocity variance, we arrive at the following for the velocity variance:

(C9) \begin{equation} \langle {w_p^\prime }^2\rangle = \langle {w_f^\prime }^2\rangle \left (1 - \frac {{St}^2_\eta }{\left ({St}_\eta + \frac {\tau _{l,p}}{\tau _\eta }\right )\left ({St}_\eta + \frac {\tau _{2}}{\tau _\eta }\right )}\right ), \end{equation}

and the slip variance

(C10) \begin{equation} \langle {w_s^\prime }^2\rangle = \langle {w_f^\prime }^2\rangle \frac {{St}^2_\eta }{\left ({St}_\eta + \frac {\tau _{l,p}}{\tau _\eta }\right )\left ({St}_\eta + \frac {\tau _{2}}{\tau _\eta }\right )}. \end{equation}

If follows from the above that the particle acceleration variance is

(C11) \begin{equation} \langle {a_p^\prime }^2\rangle = \langle {w_f^\prime }^2\rangle \frac {1}{\left ({St}_\eta + \frac {\tau _{l,p}}{\tau _\eta }\right )\left ({St}_\eta + \frac {\tau _{2}}{\tau _\eta }\right )}. \end{equation}

Since the term in the brackets in (C9) is simply the model for $1 - \langle {w_s^\prime }^2\rangle /\langle {w_f^\prime }^2\rangle$ , the implied relationship between the slip variance and the particle velocity variance is

(C12) \begin{equation} \langle {w_s^\prime }^2\rangle = \langle {w_f^\prime }^2\rangle - \langle {w_p^\prime }^2\rangle , \end{equation}

which is almost identical to (2.18), except for the fact that $R_t$ and $R_g$ are not accounted for in this model.

References

Adebiyi, A. et al. 2023 A review of coarse mineral dust in the Earth system. Aeolian Res. 60, 100849.CrossRefGoogle Scholar
Aliseda, A., Cartellier, A., Hainaux, F. & Lasheras, J.C. 2002 Effect of preferential concentration on the settling velocity of heavy particles in homogeneous isotropic turbulence. J. Fluid Mech. 468, 77105.Google Scholar
Arcen, B. & Tanière, A. 2009 Simulation of a particle-laden turbulent channel flow using an improved stochastic Lagrangian model. Phys. Fluids 21 (4), 043303.10.1063/1.3115056CrossRefGoogle Scholar
Balachandar, S. 2009 A scaling analysis for point–particle approaches to turbulent multiphase flows. Intl J. Multiphase Flow 35 (9), 801810.10.1016/j.ijmultiphaseflow.2009.02.013CrossRefGoogle 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, 349.Google Scholar
Bec, J., Homann, H. & Ray, S.S. 2014 Gravity-driven enhancement of heavy particle clustering in turbulent flow. Phys. Rev. Lett. 112 (18), 184501.CrossRefGoogle ScholarPubMed
Berk, T. & Coletti, F. 2020 Transport of inertial particles in high-Reynolds-number turbulent boundary layers. J. Fluid Mech. 903, A18.10.1017/jfm.2020.597CrossRefGoogle Scholar
Berk, T. & Coletti, F. 2021 Dynamics of small heavy particles in homogeneous turbulence: a Lagrangian experimental study. J. Fluid Mech. 917, A47.CrossRefGoogle Scholar
Berk, T. & Coletti, F. 2023 Dynamics and scaling of particle streaks in high-Reynolds-number turbulent boundary layers. J. Fluid Mech. 975, A47.10.1017/jfm.2023.885CrossRefGoogle Scholar
Berk, T. & Coletti, F. 2024 An analytical model for the slip velocity of particles in turbulence. J. Fluid Mech. 996, A1.10.1017/jfm.2024.620CrossRefGoogle Scholar
Bragg, A.D., Ireland, P.J. & Collins, L.R. 2015 On the relationship between the non-local clustering mechanism and preferential concentration. J. Fluid Mech. 780, 327343.CrossRefGoogle Scholar
Bragg, A.D., Richter, D.H. & Wang, G. 2021 a Mechanisms governing the settling velocities and spatial distributions of inertial particles in wall-bounded turbulence. Phys. Rev. Fluids 6 (6), 064302.10.1103/PhysRevFluids.6.064302CrossRefGoogle Scholar
Bragg, A.D., Richter, D.H. & Wang, G. 2021 b Settling strongly modifies particle concentrations in wall-bounded turbulent flows even when the settling parameter is asymptotically small. Phys. Rev. Fluids 6 (12), 124301.10.1103/PhysRevFluids.6.124301CrossRefGoogle Scholar
Brandt, L. & Coletti, F. 2022 Particle-laden turbulence: progress and perspectives. Annu. Rev. Fluid Mech. 54 (1), 159189.10.1146/annurev-fluid-030121-021103CrossRefGoogle Scholar
Costa, P., Brandt, L. & Picano, F. 2020 Interface-resolved simulations of small inertial particles in turbulent channel flow – CORRIGENDUM. J. Fluid Mech. 891, E2.10.1017/jfm.2020.199CrossRefGoogle Scholar
van der Does, M., Knippertz, P., Zschenderlein, P., Giles Harrison, R. & Stuut, J.-B.W. 2018 The mysterious long-range transport of giant mineral dust particles. Sci. Adv. 4 (12), eaau2768.10.1126/sciadv.aau2768CrossRefGoogle ScholarPubMed
Csanady, G.T. 1963 Turbulent diffusion of heavy particles in the atmosphere. J. Atmos. Sci. 20 (3), 201208.2.0.CO;2>CrossRefGoogle Scholar
Ferran, A., Machicoane, N., Aliseda, A. & Obligado, M. 2023 An experimental study on the settling velocity of inertial particles in different homogeneous isotropic turbulent flows. J. Fluid Mech. 970, A23.10.1017/jfm.2023.579CrossRefGoogle Scholar
Gao, W., Samtaney, R. & Richter, D.H. 2023 Direct numerical simulation of particle-laden flow in an open channel at. J. Fluid Mech. 957, A3.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
Good, G.H., Ireland, P.J., Bewley, G.P., Bodenschatz, E., Collins, L.R. & Warhaft, Z. 2014 Settling regimes of inertial particles in isotropic turbulence. J. Fluid Mech. 759, R3.10.1017/jfm.2014.602CrossRefGoogle Scholar
Grace, A.P., Richter, D.H. & Bragg, A.D. 2024 A reinterpretation of phenomenological modeling approaches for lagrangian particles settling in a turbulent boundary layer. Boundary-Layer Meteorol. 190 (4), 15.10.1007/s10546-024-00858-wCrossRefGoogle Scholar
Ireland, P.J., Bragg, A.D. & Collins, L.R. 2016 a The effect of Reynolds number on inertial particle dynamics in isotropic turbulence. Part 1. Simulations without gravitational effects. J. Fluid Mech. 796, 617658.CrossRefGoogle Scholar
Ireland, P.J., Bragg, A.D. & Collins, L.R. 2016 b The effect of Reynolds number on inertial particle dynamics in isotropic turbulence. Part 2. Simulations with gravitational effects. J. Fluid Mech. 796, 659711.10.1017/jfm.2016.227CrossRefGoogle Scholar
Johnson, P.L., Bassenne, M. & Moin, P. 2020 Turbophoresis of small inertial particles: theoretical considerations and application to wall-modelled large-eddy simulations. J. Fluid Mech. 883, A27.CrossRefGoogle Scholar
Kantha, L.H. & Clayson, C.A. 2000 Small Scale Processes in Geophysical Fluid Flows. Elsevier.Google Scholar
Kok, J.F. 2011 A scaling theory for the size distribution of emitted dust aerosols suggests climate models underestimate the size of the global dust cycle. Proc. Natl Acad. Sci. USA 108 (3), 10161021.CrossRefGoogle ScholarPubMed
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), 106901.10.1088/0034-4885/75/10/106901CrossRefGoogle ScholarPubMed
Kok, J.F., Storelvmo, T., Karydis, V.A., Adebiyi, A.A., Mahowald, N.M., Evan, A.T., He, C. & Leung, D.M. 2023 Mineral dust aerosol impacts on global climate and climate change. Nat. Rev. Earth Environ. 4 (2), 7186.10.1038/s43017-022-00379-5CrossRefGoogle Scholar
Kunkel, G.J. & Marusic, I. 2006 Study of the near-wall-turbulent region of the high Reynolds-number boundary layer using an atmospheric flow. J. Fluid Mech. 548(-1), 375402.10.1017/S0022112005007780CrossRefGoogle 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, 229246.10.1017/S0022112010001655CrossRefGoogle 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
Lien, R.-C. & D’Asaro, E.A. 2002 The Kolmogorov constant for the Lagrangian velocity spectrum and structure function. Phys. Fluids 14 (12), 44564459.10.1063/1.1518695CrossRefGoogle Scholar
Loth, E. 2023 Particles in a turbulent gas: diffusion, bias, modulation and collisions. Prog. Energy Combust. 97, 101094.CrossRefGoogle Scholar
Marchioli, C., Soldati, A., Kuerten, J.G.M., Arcen, B., Tanière, A., Goldensoph, G., Squires, K.D., Cargnelutti, M.F. & Portela, L.M. 2008 Statistics of particle dispersion in direct numerical simulations of wall-bounded turbulence: results of an international collaborative benchmark test. Intl J. Multiphase Flow 34 (9), 879893.10.1016/j.ijmultiphaseflow.2008.01.009CrossRefGoogle 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.Google Scholar
Meng, J., Huang, Y., Leung, D.M., Li, L., Adebiyi, A.A., Ryder, C.L., Mahowald, N.M. & Kok, J.F. 2022 Improved parameterization for the size distribution of emitted dust aerosols reduces model underestimation of super coarse dust. Geophys. Res. Lett. 49 (8), e2021GL097287.10.1029/2021GL097287CrossRefGoogle ScholarPubMed
Minier, J.-P. & Peirano, E. 2001 The Pdf Approach to Turbulent Polydispersed Two-Phase Ows. Physics Reports.Google Scholar
Mora, D.O., Obligado, M., Aliseda, A. & Cartellier, A. 2021 Effect of Re 03BB; and Rouse numbers on the settling of inertial droplets in homogeneous isotropic turbulence. Phys. Rev. Fluids 6 (4), 044305.10.1103/PhysRevFluids.6.044305CrossRefGoogle Scholar
Oesterlé, B. & Zaichik, L.I. 2004 On Lagrangian time scales and particle dispersion modeling in equilibrium turbulent shear flows. Phys. Fluids 16 (9), 33743384.10.1063/1.1773844CrossRefGoogle Scholar
Pozorski, J. & Minier, J.-P. 1998 On the Lagrangian turbulent dispersion models based on the Langevin equation. Intl J. Multiphase Flow 24 (6), 913945.10.1016/S0301-9322(98)00016-0CrossRefGoogle Scholar
Richter, D. & Chamecki, M. 2018 Inertial effects on the vertical transport of suspended particles in a turbulent boundary layer. Boundary-Layer Meteorol. 167 (2), 235256.CrossRefGoogle Scholar
Richter, D.H. & Sullivan, P.P. 2013 Momentum transfer in a turbulent, particle-laden Couette flow. Phys. Fluids 25 (5), 053304.10.1063/1.4804391CrossRefGoogle Scholar
Rosa, B., Parishani, H., Ayala, O. & Wang, L.-P. 2016 Settling velocity of small inertial particles in homogeneous isotropic turbulence from high-resolution DNS. Intl J. Multiphase Flow 83, 217231.10.1016/j.ijmultiphaseflow.2016.04.005CrossRefGoogle Scholar
Rosenberg, P.D. et al. 2014 Quantifying particle size and turbulent scale dependence of dust flux in the sahara using aircraft measurements: AIRCRAFT MEASUREMENTS OF DUST FLUX. J. Geophys. Res.: Atmos. 119 (12), 75777598.Google Scholar
Ryder, C.L., Highwood, E.J., Walser, A., Seibert, P., Philipp, A. & Weinzierl, B. 2019 Coarse and giant particles are ubiquitous in Saharan dust export regions and are radiatively significant over the Sahara. Atmos. Chem. Phys. 19 (24), 1535315376.10.5194/acp-19-15353-2019CrossRefGoogle Scholar
Ryder, C.L. et al. 2018 Coarse-mode mineral dust size distributions, composition and optical properties from AER-d aircraft measurements over the tropical eastern Atlantic. Atmos. Chem. Phys. 18 (23), 1722517257.CrossRefGoogle Scholar
Sawford, B.L. 1991 Reynolds number effects in Lagrangian stochastic models of turbulent dispersion. Phys. Fluids A: Fluid Dyn. 3 (6), 15771586.CrossRefGoogle Scholar
Sawford, B.L., Yeung, P.K., Borgas, M.S., Vedula, P., La Porta, A., Crawford, A.M. & Bodenschatz, E. 2003 Conditional and unconditional acceleration statistics in turbulence. Phys. Fluids 15 (11), 34783489.10.1063/1.1613647CrossRefGoogle Scholar
Shao, Y. 2008 Dust transport and deposition. In Physics and Modelling of Wind Erosion, pp. 247301, Springer Netherlands.Google Scholar
Smits, A.J., McKeon, B.J. & Marusic, I. 2011 High–Reynolds number wall turbulence. Annu. Rev. Fluid Mech. 43 (1), 353375.10.1146/annurev-fluid-122109-160753CrossRefGoogle Scholar
Tom, J. & Bragg, A.D. 2019 Multiscale preferential sweeping of particles settling in turbulence. J. Fluid Mech. 871, 244270.CrossRefGoogle Scholar
Vickers, D., Mahrt, L. & Andreas, E.L. 2015 Formulation of the sea surface friction velocity in terms of the mean wind and bulk stability. J. Appl. Meteorol. Clim. 54 (3), 691703.10.1175/JAMC-D-14-0099.1CrossRefGoogle 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, 103105.10.1016/j.ijmultiphaseflow.2019.103105CrossRefGoogle Scholar
Wang, L.-P. & Stock, D.E. 1993 Dispersion of heavy particles by turbulent motion. J. Atmos. Sci. 50 (13), 18971913.10.1175/1520-0469(1993)050<1897:DOHPBT>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Yeo, K., Kim, B.-G. & Lee, C. 2010 On the near-wall characteristics of acceleration in turbulence. J. Fluid Mech. 659, 405419.10.1017/S0022112010002557CrossRefGoogle Scholar
Yudine, M.I. 1959 Physical considerations on heavy-particle diffusion. Adv. Geophys. 6, 185191.10.1016/S0065-2687(08)60106-5CrossRefGoogle Scholar
Zamansky, R., Vinkovic, I. & Gorokhovski, M. 2011 Acceleration statistics of solid particles in turbulent channel flow. Phys. Fluids 23 (11), 113304.10.1063/1.3662006CrossRefGoogle Scholar
Zhang, Y., Bragg, A.D. & Wang, G. 2023 Asymptotic closure model for inertial particle transport in turbulent boundary layers. Phys. Rev. Fluids 8 (1), 014301.10.1103/PhysRevFluids.8.014301CrossRefGoogle Scholar
Figure 0

Figure 1. A schematic of the numerical set-up. The domain is a rectangular channel of height $H$, streamwise length $2\pi\!{H}$ and spanwise width $\pi\!{H}$. The flow is periodic in the horizontal and is driven by a constant pressure gradient in the streamwise direction, while the no-stress and no-slip boundary conditions are enforced at the top and bottom boundaries, respectively. Particles are injected at the upper boundary at a random horizontal location with an initial velocity equal to the fluid velocity at their location and removed when they contact the bottom boundary. They are allowed to rebound elastically off the upper boundary.

Figure 1

Table 1. Table of cases discussed throughout the work. Parameter definitions can be found in the main text. The case with $\textit{Re}_\tau = 1260$ was run on a $512^3$ grid, all cases with $\textit{Re}_\tau =630$ were run on a $256^3$ grid, while cases with $\textit{Re}_\tau =315$ were run on a $128^3$ grid.

Figure 2

Figure 2. Controlling tendencies for the slip velocity variance according to (2.18) at $\textit{Re}_\tau =630$. Panels (a), (d) and (g) show the normalized slip velocity variance for each ${S{\kern-0.5pt}v}^+$, while (a–c), (d–f) and (g–i) are for a different value of ${S{\kern-0.5pt}t}^+$ (shown on the left-hand side of the figure). Panels (b), (e) and (h) show the (negative) velocity variance and the (positive) seen velocity variance. Panels (c), (f) and (i) show the contributions from $R_g$. Note that $R_t$ is omitted from this figure as it is small across the entire domain relative to the other terms. All terms are normalized by $u_\tau ^2$.

Figure 3

Figure 3. The horizontal and vertical components of $\varphi _r^{(i)}$ (panels (a) and (b)) and $\varphi _s^{(i)}$ (panels (c) and (d)) for all cases in table 1 plotted against ${S{\kern-0.5pt}v}^+$. The filled markers correspond to cases at $\textit{Re}_\tau =630$, while the open-faced markers correspond to cases at $\textit{Re}_\tau = 315$.

Figure 4

Figure 4. Panel (a) shows the slip variance normalized by the seen variance for all cases in table 1 at $\textit{Re}_\tau = 630$. The horizontal dashed lines denote heights of $z^+=50$ and $z/H = 0.75$. Panel (b) shows the normalized acceleration variance over the same range plotted against the local value of ${St}_\eta$, while the dashed line represents the ${St}_\eta ^{-2}$ scaling. The colours of each curve in panels (a) and (b) correspond to values of ${S{\kern-0.5pt}t}^+$, while the line styles correspond to values of ${S{\kern-0.5pt}v}^+$. Panel (c) shows ratio of the seen variance to the unconditional variance averaged over the entire vertical extent plotted against ${S{\kern-0.5pt}v}^+$ for all cases at $\textit{Re}_\tau = 630$.

Figure 5

Figure 5. Panel (a) shows the normalized acceleration variance for cases with ${S{\kern-0.5pt}t}^+=10$ and ${S{\kern-0.5pt}v}^+=0.8$ at three different Reynolds numbers as a function of ${St}_\eta$. Panel (b) shows the seen variance (filled markers), slip variance (coloured empty markers) and particle velocity variance (black markers) normalized by the unconditional variance integrated over the range $D$ as a function of $\textit{Re}_\tau$.

Figure 6

Figure 6. Shown in panels (a)–(d) are $\xi _1$, $\xi _2$ and the root mean squares (r.m.s.) values of $R_t$ and $R_g$ normalized by the seen variance for all cases in table 1, respectively. Open-faced markers represent cases with $\textit{Re}_\tau = 315$, filled markers represent cases with $\textit{Re}_\tau = 630$ and the marker with an $\otimes$${Re_\tau }=1260$.

Figure 7

Figure 7. Panel (a) shows the slip velocity variance computed by the DNS for three values of ${S{\kern-0.5pt}v}^+$ at fixed ${S{\kern-0.5pt}t}^+=10$ (black curves), and the slip variance computed by (2.18) (dashed coloured curves). Panel(b) shows $R_t + R_g$ computed via a residual of $\langle {w_s^\prime }\rangle _z - (\langle {w_f^\prime }\rangle _z - \langle {w_p^\prime }\rangle _z)$ (black curves), and $R_t+R_g$ computed directly (dashed coloured curves) for the same cases.