Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

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

Andrew P. Grace\aff1 \corresp agrace4@nd.edu    David Richter\aff1    Tim Berk\aff2    Andrew D. Bragg\aff3 \aff1Department of Civil and Environmental Engineering and Earth Sciences, University of Notre Dame, Notre Dame, Indiana 46556, U.S.A., \aff2Department of Mechanical and Aerospace Engineering, Utah State University, Logan, Utah 84322, U.S.A., \aff3Department of Civil and Environmental Engineering, Duke University, Durham, North Carolina 27708, U.S.A.
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 St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, and ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, 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 sub-leading until Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, rendering the variance relatively small — an effect found to be most substantial for Sv+>1superscriptSv1\mathrm{Sv}^{+}>1roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT > 1. Finally, we compare the results to a model of the acceleration variance (Berk & Coletti, 2021) based the concept of a response function described in Csanady (1963), highlighting the role of the crossing trajectories mechanism. We find that while there is good agreement for low Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, 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 modeling the transport of coarse dust grains in the atmospheric surface layer.

keywords:

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 μ𝜇\muitalic_μm) once they are emitted from the surfaces of arid regions (Rosenberg et al., 2014; Meng et al., 2022; Adebiyi et al., 2023; Kok et al., 2023). 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, 2008; Van Der Does et al., 2018). 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., 2012), will help us to understand their overall role in global climate processes (Kok, 2011; Ryder et al., 2019; Kok et al., 2023), biogeochemical cycles (Ryder et al., 2018), 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, 1983). Since the seminal work of Wang & Stock (1993), 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., 2002; Good et al., 2014; Rosa et al., 2016; Tom & Bragg, 2019; Mora et al., 2021; Ferran et al., 2023). Importantly, the bias created by gravity leads to a more rapid decorrelation of the turbulence along particle trajectories, meaning that there is an implicit and non-linear coupling between gravity and turbulent drag, resulting in a fundamental change in the forcing induced by turbulence. One of the effects of the implicit coupling between gravity and turbulent drag is known as crossing trajectories (Yudine, 1959; Csanady, 1963), which has been shown to increase the horizontal and vertical components of particle acceleration variance in simulations of settling Lagrangian point particles in HIT (Ireland et al., 2016b), in numerical simulations of turbulent boundary layers (Lavezzo et al., 2010), as well as laboratory experiments in both setups (Gerashchenko et al., 2008; Berk & Coletti, 2021). 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 diagnosing a particle Reynolds number (Balachandar, 2009), as well building physically coherent stochastic dispersion models for RANS (Arcen & Tanière, 2009) applications.

In this work, we are particularly focused on understanding the dynamic regimes characteristic of coarse dust particles in Earth’s atmospheric surface layer (the lowest 100 metres of Earth’s atmosphere). Specifically, we consider how gravitational acceleration implicitly modifies the mechanisms controlling the particle slip velocity in a turbulent boundary layer (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 turbulent boundary layer 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 turbulent boundary layers can be found in Smits et al. (2011).

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., 2008; Balachandar, 2009; Zamansky et al., 2011; Johnson et al., 2020). 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, 2022). Furthermore, much of our understanding of particle-laden flows under the influence of gravitational settling comes from numerical (Good et al., 2014; Bec et al., 2014; Ireland et al., 2016b; Tom & Bragg, 2019) or laboratory (Aliseda et al., 2002; Mora et al., 2021; Ferran et al., 2023) configurations of homogeneous isotropic turbulence (HIT), due to the relative simplicity of the setup. There are a few studies aiming to understand the statistical behaviour of settling inertial particles in turbulent boundary layers (Lavezzo et al., 2010; Lee & Lee, 2019; Berk & Coletti, 2020; Bragg et al., 2021a, b; Berk & Coletti, 2023), 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. (2021b) 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. 1.

    Derive continuum equations for moments of the particle slip velocity and identify the leading order balance of the variance throughout the turbulent boundary layer.

  2. 2.

    Determine the parametric conditions under which the slip velocity is governed by its mean component, fluctuating component, or some combination of both.

  3. 3.

    Compare the results from the DNS in a TBL to a model based on the response function in homogeneous turbulence approach outlined in Csanady (1963) and Berk & Coletti (2021), and identify potential discrepancies.

  4. 4.

    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 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 section 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, 2018) to model one-way coupled inertial particles settling through a turbulent boundary layer. This code has been validated and used in multiple studies focused on inertial particle settling and transport in turbulent boundary layers (Richter & Chamecki, 2018; Wang et al., 2019; Bragg et al., 2021a; Gao et al., 2023; Grace et al., 2024). For the carrier phase, we use direct numerical simulations (DNS) to solve the three-dimensional, incompressible Navier-Stokes equations in a turbulent open channel flow setup:

D𝒖Dt𝐷𝒖𝐷𝑡\displaystyle\frac{D\boldsymbol{u}}{Dt}divide start_ARG italic_D bold_italic_u end_ARG start_ARG italic_D italic_t end_ARG =1ρap+ν2𝒖1ρadPdx𝒙^,absent1subscript𝜌𝑎𝑝𝜈superscript2𝒖1subscript𝜌𝑎𝑑𝑃𝑑𝑥^𝒙\displaystyle=-\frac{1}{\rho_{a}}\nabla p+\nu\nabla^{2}\boldsymbol{u}-\frac{1}% {\rho_{a}}\frac{dP}{dx}\hat{\boldsymbol{x}},= - divide start_ARG 1 end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG ∇ italic_p + italic_ν ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_italic_u - divide start_ARG 1 end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG divide start_ARG italic_d italic_P end_ARG start_ARG italic_d italic_x end_ARG over^ start_ARG bold_italic_x end_ARG , (1)
𝒖𝒖\displaystyle\nabla\cdot\boldsymbol{u}∇ ⋅ bold_italic_u =0.absent0\displaystyle=0.= 0 . (2)

A schematic of the setup is presented in figure 1. In the above equations, DDt𝐷𝐷𝑡\frac{D}{Dt}divide start_ARG italic_D end_ARG start_ARG italic_D italic_t end_ARG represents the material derivative, 𝒖𝒖\boldsymbol{u}bold_italic_u represents the three dimensional flow velocity, p𝑝pitalic_p represents the turbulent pressure field, ρasubscript𝜌𝑎\rho_{a}italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is the carrier phase density, and ν𝜈\nuitalic_ν is the kinematic viscosity. As we are primarily focused on wall normal motion in this work, we will refer to fluctuating vertical fluid velocities as wsuperscript𝑤w^{\prime}italic_w start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

At the lower boundary, a no-slip boundary condition is enforced, while at the upper boundary, a no-stress boundary condition (du/dz=dv/dz=0𝑑𝑢𝑑𝑧𝑑𝑣𝑑𝑧0du/dz=dv/dz=0italic_d italic_u / italic_d italic_z = italic_d italic_v / italic_d italic_z = 0 at z=H𝑧𝐻z=Hitalic_z = italic_H) is enforced. The domain is periodic in the x𝑥xitalic_x and y𝑦yitalic_y directions. The background state of the carrier phase is established by accelerating the flow with an imposed pressure gradient, dP/dx>0𝑑𝑃𝑑𝑥0-dP/dx>0- italic_d italic_P / italic_d italic_x > 0 (note that 𝒙^^𝒙\hat{\boldsymbol{x}}over^ start_ARG bold_italic_x end_ARG 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 uτ=τw/ρasubscript𝑢𝜏subscript𝜏𝑤subscript𝜌𝑎u_{\tau}=\sqrt{\tau_{w}/\rho_{a}}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = square-root start_ARG italic_τ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT / italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG, where τwsubscript𝜏𝑤\tau_{w}italic_τ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT is the stress at the lower boundary. Using the friction velocity, the height of the domain, H𝐻Hitalic_H, and viscosity of the carrier phase, we can define a friction Reynolds number of Reτ=uτHνsubscriptRe𝜏subscript𝑢𝜏𝐻𝜈\mathrm{Re}_{\tau}=\frac{u_{\tau}H}{\nu}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = divide start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_H end_ARG start_ARG italic_ν end_ARG. Friction Reynolds numbers for each simulation presented in this work can be found in Table 1. This setup is identical to that used in Grace et al. (2024).

We can define the local Kolomogorov timescale, velocity scale, and acceleration scale,

τη=(νϵ)1/2,vη=(νϵ)1/4,aη=(ϵ3ν)1/4,formulae-sequencesubscript𝜏𝜂superscript𝜈italic-ϵ12formulae-sequencesubscript𝑣𝜂superscript𝜈italic-ϵ14subscript𝑎𝜂superscriptsuperscriptitalic-ϵ3𝜈14\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},italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = ( divide start_ARG italic_ν end_ARG start_ARG italic_ϵ end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , italic_v start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = ( italic_ν italic_ϵ ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT , italic_a start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = ( divide start_ARG italic_ϵ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT , (3)

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, homogeneous isotropic turbulence, 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 ϵitalic-ϵ\epsilonitalic_ϵ is a function of the distance from the solid boundary. Within the logarithmic layer, it scales as

ϵO(uτ3κz),similar-toitalic-ϵ𝑂superscriptsubscript𝑢𝜏3𝜅𝑧\epsilon\sim O\Big{(}\frac{u_{\tau}^{3}}{\kappa z}\Big{)},italic_ϵ ∼ italic_O ( divide start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_κ italic_z end_ARG ) , (4)

where uτsubscript𝑢𝜏u_{\tau}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT is the friction velocity, and κ𝜅\kappaitalic_κ is the von Karman constant. However, when calculating the Kolmogorov timescale, velocity scale, and length scale, we use the dissipation computed in the DNS.

Refer to caption
Figure 1: A schematic of the numerical setup. The domain is a rectangular channel of height H𝐻Hitalic_H, streamwise length 2πH2𝜋𝐻2\pi H2 italic_π italic_H and spanwise width πH𝜋𝐻\pi Hitalic_π italic_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.

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 μm𝜇m\mathrm{\mu m}italic_μ roman_m) 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. Furthermore, as we are concerned with the dilute limit, two-way coupling and particle-particle interactions may be ignored. Since these particles are much denser than the carrier phase, we may also ignore added mass and Basset-History forces. The one-way coupled point particle approach also has the added benefit that each particle is independent from each other particle, effectively removing the volume fraction as a governing parameter. This allows us to increase the number of particles to ensure convergence of the statistics of interest without affecting the flow.

The equations of motion are

d𝒗pdt=𝑑subscript𝒗𝑝𝑑𝑡absent\displaystyle\frac{d\boldsymbol{v}_{p}}{dt}=divide start_ARG italic_d bold_italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = Ψτp(𝒖f𝒗p)𝒈,Ψsubscript𝜏𝑝subscript𝒖𝑓subscript𝒗𝑝𝒈\displaystyle\frac{\Psi}{\tau_{p}}\left(\boldsymbol{u}_{f}-\boldsymbol{v}_{p}% \right)-\boldsymbol{g},divide start_ARG roman_Ψ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( bold_italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - bold_italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) - bold_italic_g , (5)
d𝒙pdt=𝑑subscript𝒙𝑝𝑑𝑡absent\displaystyle\frac{d\boldsymbol{x}_{p}}{dt}=divide start_ARG italic_d bold_italic_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = 𝒗p.subscript𝒗𝑝\displaystyle\boldsymbol{v}_{p}.bold_italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT . (6)

Here, 𝒗psubscript𝒗𝑝\boldsymbol{v}_{p}bold_italic_v start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the three dimensional velocity vector for each particle, 𝒙psubscript𝒙𝑝\boldsymbol{x}_{p}bold_italic_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the location of each particle in space, 𝒈𝒈\boldsymbol{g}bold_italic_g is the gravitational acceleration (which only affects accelerations in the z𝑧zitalic_z direction), 𝒖fsubscript𝒖𝑓\boldsymbol{u}_{f}bold_italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT 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 ufsubscript𝑢𝑓u_{f}italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT (not bold). τpsubscript𝜏𝑝\tau_{p}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the relaxation timescale of the particle, defined as

τp=ρpdp218ρaν,subscript𝜏𝑝subscript𝜌𝑝superscriptsubscript𝑑𝑝218subscript𝜌𝑎𝜈\tau_{p}=\frac{\rho_{p}d_{p}^{2}}{18\rho_{a}\nu},italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 18 italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_ν end_ARG , (7)

where ρpsubscript𝜌𝑝\rho_{p}italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the particle density and dpsubscript𝑑𝑝d_{p}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the particle diameter, and the Stokes settling velocity is defined as vg=τpgsubscript𝑣𝑔subscript𝜏𝑝𝑔v_{g}=\tau_{p}gitalic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_g. Ψ=1+0.15Rep0.687Ψ10.15superscriptsubscriptRe𝑝0.687\Psi=1+0.15\mathrm{Re}_{p}^{0.687}roman_Ψ = 1 + 0.15 roman_Re start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.687 end_POSTSUPERSCRIPT is the Schiller-Neumann correction to the drag force, and RepsubscriptRe𝑝\mathrm{Re}_{p}roman_Re start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the particle Reynolds number. Throughout this work, we change τpsubscript𝜏𝑝\tau_{p}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT by modifying ρpsubscript𝜌𝑝\rho_{p}italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (ensuring that ρpρamuch-greater-thansubscript𝜌𝑝subscript𝜌𝑎\rho_{p}\gg\rho_{a}italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≫ italic_ρ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT) in order to maintain dpηmuch-less-thansubscript𝑑𝑝𝜂d_{p}\ll\etaitalic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≪ italic_η. This means that for each case considered, dp+superscriptsubscript𝑑𝑝d_{p}^{+}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (the particle diameter normalized by the viscous scales) is constant across all cases. As we keep the particle diameter fixed for all cases in this work, the particle Reynolds number remains small, meaning that Ψ1Ψ1\Psi\approx 1roman_Ψ ≈ 1. For the theory discussed below in section 2.3, we follow Bragg et al. (2021a) and make the assumption that Ψ=1Ψ1\Psi=1roman_Ψ = 1 for analytical tractability.

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

St+=τpuτ2ν,Sv+=vguτ,Stη=τpτη,Svη=vgvη.formulae-sequencesuperscriptStsubscript𝜏𝑝superscriptsubscript𝑢𝜏2𝜈formulae-sequencesuperscriptSvsubscript𝑣𝑔subscript𝑢𝜏formulae-sequencesubscriptSt𝜂subscript𝜏𝑝subscript𝜏𝜂subscriptSv𝜂subscript𝑣𝑔subscript𝑣𝜂\mathrm{St}^{+}=\frac{\tau_{p}u_{\tau}^{2}}{\nu},\quad\mathrm{Sv}^{+}=\frac{v_% {g}}{u_{\tau}},\quad\mathrm{St}_{\eta}=\frac{\tau_{p}}{\tau_{\eta}},\quad% \mathrm{Sv}_{\eta}=\frac{v_{g}}{v_{\eta}}.roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ν end_ARG , roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = divide start_ARG italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT end_ARG , roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG , roman_Sv start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = divide start_ARG italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG . (8)

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. StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT and SvηsubscriptSv𝜂\mathrm{Sv}_{\eta}roman_Sv start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT are functions of the distance from the solid boundary, whereas St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT are constant parameters.

Other studies focused on particle settling define other parameters such as a Froude number, Fr=aη/gFrsubscript𝑎𝜂𝑔\mathrm{Fr}=a_{\eta}/groman_Fr = italic_a start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT / italic_g, (Bec et al., 2014; Berk & Coletti, 2021), or a scaled gravity

g+=gνuτ3,superscript𝑔𝑔𝜈superscriptsubscript𝑢𝜏3g^{+}=\frac{g\nu}{u_{\tau}^{3}},italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = divide start_ARG italic_g italic_ν end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG , (9)

which is simply the ratio of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT to St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. These parameters are useful as they describe the relative role of gravitational accelerations and turbulent accelerations without referring to τpsubscript𝜏𝑝\tau_{p}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. We will refer to g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT at various points throughout the discussion when necessary. The values for all parameters considered in this work, including the associated ranges of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT and SvηsubscriptSv𝜂\mathrm{Sv}_{\eta}roman_Sv start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, can be found in Table 1.

ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT dp+superscriptsubscript𝑑𝑝d_{p}^{+}italic_d start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT StηsubscriptSt𝜂\mathrm{St_{\eta}}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT SvηsubscriptSv𝜂\mathrm{Sv}_{\eta}roman_Sv start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT
315 0.025 10 2.5×1032.5superscript1032.5\times 10^{-3}2.5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 0.236 0.31–4.25 0.038–0.14
315 0.025 50 5×1045superscript1045\times 10^{-4}5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 0.236 1.58–21.25 0.038–0.14
315 0.25 10 2.5×1022.5superscript1022.5\times 10^{-2}2.5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 0.31–4.25 0.38–1.4
315 0.25 50 5×1035superscript1035\times 10^{-3}5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 0.236 1.58–21.25 0.38–1.4
315 0.8 10 8×1028superscript1028\times 10^{-2}8 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 0.3–4.29 1.22–4.47
315 2.5 10 2.5×1012.5superscript1012.5\times 10^{-1}2.5 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT 0.236 0.31–4.25 3.8–14.0
315 2.5 50 5×1025superscript1025\times 10^{-2}5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 1.58–21.25 3.8–14.0
630 0.025 10 2.5×1032.5superscript1032.5\times 10^{-3}2.5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 0.236 0.24–4.7 0.037–0.16
630 0.025 50 5×1045superscript1045\times 10^{-4}5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 0.236 1.18–23.17 0.037–0.16
630 0.025 100 2.5×1042.5superscript1042.5\times 10^{-4}2.5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 0.236 2.38–46.33 0.037–0.16
630 0.25 10 2.5×1022.5superscript1022.5\times 10^{-2}2.5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 0.24–4.7 0.37–1.60
630 0.25 50 5×1035superscript1035\times 10^{-3}5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 0.236 1.18–23.17 0.37–1.60
630 0.25 100 2.5×1032.5superscript1032.5\times 10^{-3}2.5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 0.236 2.38–46.33 0.37–1.60
630 0.8 10 8×1028superscript1028\times 10^{-2}8 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 0.24–4.7 1.17–5.21
630 2.5 10 2.5×1012.5superscript1012.5\times 10^{-1}2.5 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT 0.236 0.24–4.7 3.66–16.03
630 2.5 50 5×1025superscript1025\times 10^{-2}5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 1.18–23.17 3.66–16.03
630 2.5 100 2.5×1022.5superscript1022.5\times 10^{-2}2.5 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 2.38–46.33 3.66–16.03
1260 0.8 10 8×1028superscript1028\times 10^{-2}8 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 0.236 0.2357–6.29 1.0–5.21
Table 1: Table of cases discussed throughout the work. Parameter definitions can be found in the main text. The case with Reτ=1260subscriptRe𝜏1260\mathrm{Re}_{\tau}=1260roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 1260 was run on a 5123superscript5123512^{3}512 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT grid, all cases with Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630 were run on a 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT grid, while cases with Reτ=315subscriptRe𝜏315\mathrm{Re}_{\tau}=315roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 315 were run on a 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT grid.

2.3 Statistics of inertial particles in a turbulent ASL

We adopt the particle phase space approach used in Bragg et al. (2021a), focusing only on the vertical component of the particle equations of motion. First they define the particle PDF in position-velocity space as

𝒫=δ(zpz)δ(wpw),𝒫delimited-⟨⟩𝛿subscript𝑧𝑝𝑧𝛿subscript𝑤𝑝𝑤\mathcal{P}=\langle\delta(z_{p}-z)\delta(w_{p}-w)\rangle,caligraphic_P = ⟨ italic_δ ( italic_z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_z ) italic_δ ( italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - italic_w ) ⟩ , (10)

which describes the distribution of the vertical components of the particle position and velocity, zp(t)subscript𝑧𝑝𝑡z_{p}(t)italic_z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ) and wp(t)subscript𝑤𝑝𝑡w_{p}(t)italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t ), in the phase space with coordinates z𝑧zitalic_z, w𝑤witalic_w, and delimited-⟨⟩\langle\cdot\rangle⟨ ⋅ ⟩ represents an ensemble average over all realizations of the system. Note we make frequent use of conditional averages throughout this work, denoted by z,wsubscriptdelimited-⟨⟩𝑧𝑤\langle\cdot\rangle_{z,w}⟨ ⋅ ⟩ start_POSTSUBSCRIPT italic_z , italic_w end_POSTSUBSCRIPT which is short hand for |zp=z,wp=w\langle\cdot|z_{p}=z,w_{p}=w\rangle⟨ ⋅ | italic_z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_z , italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_w ⟩. We can form an evolution equation for the PDF:

𝒫t+z(w𝒫)+w(apz,w𝒫)=0𝒫𝑡𝑧𝑤𝒫𝑤subscriptdelimited-⟨⟩subscript𝑎𝑝𝑧𝑤𝒫0\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)=0divide start_ARG ∂ caligraphic_P end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG ∂ end_ARG start_ARG ∂ italic_z end_ARG ( italic_w caligraphic_P ) + divide start_ARG ∂ end_ARG start_ARG ∂ italic_w end_ARG ( ⟨ italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z , italic_w end_POSTSUBSCRIPT caligraphic_P ) = 0 (11)

where we have defined apz,w=τp1(ufz,ww)gsubscriptdelimited-⟨⟩subscript𝑎𝑝𝑧𝑤superscriptsubscript𝜏𝑝1subscriptdelimited-⟨⟩subscript𝑢𝑓𝑧𝑤𝑤𝑔\langle a_{p}\rangle_{z,w}=\tau_{p}^{-1}\left(\langle u_{f}\rangle_{z,w}-w% \right)-g⟨ italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z , italic_w end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z , italic_w end_POSTSUBSCRIPT - italic_w ) - italic_g as the vertical particle acceleration conditioned on zp=zsubscript𝑧𝑝𝑧z_{p}=zitalic_z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_z and wp=wsubscript𝑤𝑝𝑤w_{p}=witalic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_w based on the vertical component of (5). The utility of this equation comes from the fact that we can derive evolution equations for each moment. Recall that the n𝑛nitalic_nth moment is defined as

wpnz=1ϱwn𝒫𝑑wsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝𝑛𝑧1italic-ϱsuperscriptsubscriptsuperscript𝑤𝑛𝒫differential-d𝑤\langle w_{p}^{n}\rangle_{z}=\frac{1}{\varrho}\int_{-\infty}^{\infty}w^{n}% \mathcal{P}dw⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_ϱ end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_w start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT caligraphic_P italic_d italic_w (12)

where the notation zsubscriptdelimited-⟨⟩𝑧\langle\cdot\rangle_{z}⟨ ⋅ ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT represents an ensemble average conditioned on zp=zsubscript𝑧𝑝𝑧z_{p}=zitalic_z start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_z. Of importance to the present work will be the first and second moments (i.e. n=1𝑛1n=1italic_n = 1 and n=2𝑛2n=2italic_n = 2). The evolution equation for the first moment is:

wpz=ufzvgτpϱddzϱwp2z.subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩subscript𝑢𝑓𝑧subscript𝑣𝑔subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle w_{p}\rangle_{z}=\langle u_{f}\rangle_{z}-v_{g}-\frac{\tau_{p}}{% \varrho}\frac{d}{dz}\varrho\langle w_{p}^{2}\rangle_{z}.⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT . (13)

The details of the derivation of this equation can be found in Bragg et al. (2021a), and the general form of this equation (for the case where vg=0subscript𝑣𝑔0v_{g}=0italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 0) for arbitrary moments can be found in Johnson et al. (2020). Eq. (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 re-arrange (13) into a relationship describing how the mean slip velocity varies with height:

usz=vg+τpϱddzϱwp2z,subscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscript𝑣𝑔subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle u_{s}\rangle_{z}=v_{g}+\frac{\tau_{p}}{\varrho}\frac{d}{dz}\varrho% \langle w_{p}^{2}\rangle_{z},⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (14)

where usz=ufzwpzsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscriptdelimited-⟨⟩subscript𝑢𝑓𝑧subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧\langle u_{s}\rangle_{z}=\langle u_{f}\rangle_{z}-\langle w_{p}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT.

Our goal is to derive a continuum equation for the slip velocity variance. To do this, we multiply (11) by w2superscript𝑤2w^{2}italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and integrate over all w𝑤witalic_w. The full details of this operation can be found in Johnson et al. (2020). After integrating, we are left with the following relationship:

ddzϱwp3z2ϱapwpz=0.𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧2italic-ϱsubscriptdelimited-⟨⟩subscript𝑎𝑝subscript𝑤𝑝𝑧0\frac{d}{dz}\varrho\langle w_{p}^{3}\rangle_{z}-2\varrho\langle a_{p}w_{p}% \rangle_{z}=0.divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - 2 italic_ϱ ⟨ italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 . (15)

To expand the acceleration-velocity covariance, we expand to get awpz=ufwpzwp2zsubscriptdelimited-⟨⟩𝑎subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩subscript𝑢𝑓subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle aw_{p}\rangle_{z}=\langle u_{f}w_{p}\rangle_{z}-\langle w_{p}^{2}% \rangle_{z}⟨ italic_a italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and use the fact that us2z=uf2z2ufwpz+wp2zsubscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑓2𝑧2subscriptdelimited-⟨⟩subscript𝑢𝑓subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle u_{s}^{2}\rangle_{z}=\langle u_{f}^{2}\rangle_{z}-2\langle u_{f}w_{p}% \rangle_{z}+\langle w_{p}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - 2 ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT to arrive at

2awpz=1τp(uf2zus2zwp2z)2wpzg2subscriptdelimited-⟨⟩𝑎subscript𝑤𝑝𝑧1subscript𝜏𝑝subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧2subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧𝑔2\langle aw_{p}\rangle_{z}=\frac{1}{\tau_{p}}\left(\langle u_{f}^{2}\rangle_{z% }-\langle u_{s}^{2}\rangle_{z}-\langle w_{p}^{2}\rangle_{z}\right)-2\langle w_% {p}\rangle_{z}g2 ⟨ italic_a italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) - 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_g (16)

Putting (15) and (16) together, we get an equation for the mean squared slip velocity:

us2z=uf2zwp2z2wpzvgτpϱddzϱwp3z.subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧2subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscript𝑣𝑔subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧\langle u_{s}^{2}\rangle_{z}=\langle u_{f}^{2}\rangle_{z}-\langle w_{p}^{2}% \rangle_{z}-2\langle w_{p}\rangle_{z}v_{g}-\frac{\tau_{p}}{\varrho}\frac{d}{dz% }\varrho\langle w_{p}^{3}\rangle_{z}.⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT . (17)

Eq. (17) indicates that the mean squared slip velocity variance is a function of the mean squared sampled velocity uf2zsubscriptdelimited-⟨⟩superscriptsubscript𝑢𝑓2𝑧\langle u_{f}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, the mean squared particle velocity wp2zsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle w_{p}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, a drift due to the non-zero average vertical velocity wpzsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧\langle w_{p}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, and the vertical derivative of the mean cubed particle velocity wp3zsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧\langle w_{p}^{3}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, 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 turbulent boundary layer, 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

us2z=uf2zwp2z(1)+Rt(2)+Rg(3)subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptsuperscriptsubscriptsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧1subscript𝑅𝑡2subscript𝑅𝑔3\langle{u_{s}^{\prime}}^{2}\rangle_{z}=\underbrace{\overbrace{\underbrace{% \langle{u_{f}^{\prime}}^{2}\rangle_{z}-\langle{w_{p}^{\prime}}^{2}\rangle_{z}}% _{(1)}+R_{t}}^{(2)}+R_{g}}_{(3)}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = under⏟ start_ARG over⏞ start_ARG under⏟ start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_POSTSUBSCRIPT ( 1 ) end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT + italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_POSTSUBSCRIPT ( 3 ) end_POSTSUBSCRIPT (18)

where Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are defined as

Rt=subscript𝑅𝑡absent\displaystyle R_{t}=italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = τpϱddzϱwp3zsubscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝3𝑧\displaystyle-\frac{\tau_{p}}{\varrho}\frac{d}{dz}\varrho\langle{w_{p}^{\prime% }}^{3}\rangle_{z}- divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (19)
Rg=subscript𝑅𝑔absent\displaystyle R_{g}=italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = τpϱddz(ϱwpz3+3ϱwpzwp2z)+2wpz(uszvg),subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptsuperscriptdelimited-⟨⟩subscript𝑤𝑝3𝑧3italic-ϱsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧2subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscript𝑣𝑔\displaystyle-\frac{\tau_{p}}{\varrho}\frac{d}{dz}\left(\varrho\langle w_{p}% \rangle^{3}_{z}+3\varrho\langle w_{p}\rangle_{z}\langle{w_{p}^{\prime}}^{2}% \rangle_{z}\right)+2\langle w_{p}\rangle_{z}\left(\langle u_{s}\rangle_{z}-v_{% g}\right),- divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG ( italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + 3 italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) + 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) , (20)

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)1(1)( 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 uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. This is the case for particles both with and without gravity, since gravity and inertia implicitly modify these terms. Next, grouped under (2)2(2)( 2 ), are the terms that appear for particles dispersing vertically through a turbulent boundary layer in the absence of gravity. Note that all terms encompassed by (1)1(1)( 1 ) are included in (2)2(2)( 2 ), but when considering those terms covered under (1)1(1)( 1 ) in the context of a turbulent boundary layer, 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 Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. 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+0superscriptSv0\mathrm{Sv^{+}}\rightarrow 0roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT → 0, wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT approaches wp2zsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\langle w_{p}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, but Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT remains, regardless. Finally, by incorporating gravity, the mean particle velocity is no longer zero, leading to a new term grouped under (3)3(3)( 3 ), denoted by Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. The quantities composing Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT 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 (20) is written in terms of uszvgsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscript𝑣𝑔\langle u_{s}\rangle_{z}-v_{g}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, which we can see from (14) is identical to τpϱddzϱwp2zsubscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\frac{\tau_{p}}{\varrho}\frac{d}{dz}\varrho\langle w_{p}^{2}\rangle_{z}divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. 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)1(1)( 1 )), those that appear in a TBL without settling (grouped under (2)2(2)( 2 )), and those that appear in a TBL with settling (grouped under (3)3(3)( 3 )). In the following section, we will identify the importance of these terms throughout the turbulent boundary layer.

3 Results

3.1 Tendencies governing the slip velocity variance

Refer to caption
Figure 2: Controlling tendencies for the slip velocity variance according to equation (18) at Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630. The left column shows the normalized slip velocity variance for each Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, while each row is for a different value of St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (shown on the left side of the figure). The centre column shows the (negative) velocity variance and the (positive) seen velocity variance. The right column shows the contributions from Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. All terms are normalized by uτ2superscriptsubscript𝑢𝜏2u_{\tau}^{2}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

In this section, we consider vertical profiles of us2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧\langle{u_{s}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. Figure 2 shows the tendencies in (18) for several cases scaled by uτ2superscriptsubscript𝑢𝜏2u_{\tau}^{2}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Note that all profiles of a given St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv^{+}}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT in columns two and three of figure 2 when added together return the profiles shown in column one, as per (18). An example of this is shown in Appendix C. Each row corresponds to a different friction Stokes number highlighted on the left side of the figure, while each curve in the leftmost and center columns corresponds to a different value of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. Within the third column, the solid curves correspond to Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT while the dashed curves correspond to Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. Overall, this figure highlights the dominant tendencies controlling the slip velocity variance as St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT are independently changed.

Generally, the behaviour of the slip variance as a function of the vertical coordinate is qualitatively similar between all cases considered, evident from Figures 2(a), (d), and (g). However, the magnitude of the slip variance for a given case varies throughout the domain, and becomes sensitive to Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT when Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT becomes larger than unity. Within the logarithmic layer, the slip variance at constant Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (curves of fixed color) tends to increase rapidly between St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 10 and St+=50superscriptSt50\mathrm{St}^{+}=50roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 50, but more slowly between St+=50superscriptSt50\mathrm{St^{+}}=50roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 50 and St+=100superscriptSt100\mathrm{St}^{+}=100roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 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 figures 2(b), (e), and (h). However, very near the solid boundary (i.e. below z/H=0.1𝑧𝐻0.1z/H=0.1italic_z / italic_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 Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are 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 St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT tends to increase most rapidly as Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. 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 ufzsubscriptdelimited-⟨⟩subscript𝑢𝑓𝑧\langle u_{f}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT deviating from zero for an inertial particle (see detailed discussion in Bragg et al. (2021a)).

Lastly, the higher order terms (Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT), shown in figures 2(c), (f), and (i), are non-zero but are not leading order within the interior of the domain (note the change in the horizontal scale of these panels), though they are relatively important within the viscous sublayer. Note that above z/H=0.1𝑧𝐻0.1z/H=0.1italic_z / italic_H = 0.1, these terms are almost completely negligible aside from when Sv+=2.5superscriptSv2.5\mathrm{Sv}^{+}=2.5roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 2.5.

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 Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT when Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT increases beyond unity. However, it is clear that higher order moments of the continuum equations (Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT) may be sub-leading and negligible in most other cases. It is not until the viscous sublayer where contributions from Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT become significant in determining the slip variance. Furthermore, the sub-leading behaviour of Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT in the logarithmic layer does not imply that the inhomogeneity of the turbulence is irrelevant. In fact, the remaining terms (uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT) may have implicit dependence on the inhomogeneity of the turbulence, and this will be discussed later.

Refer to caption
Figure 3: The horizontal and vertical components of φr(i)superscriptsubscript𝜑𝑟𝑖\varphi_{r}^{(i)}italic_φ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT (panels (a) and (b)) and φs(i)superscriptsubscript𝜑𝑠𝑖\varphi_{s}^{(i)}italic_φ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT (panels (c) and (d)) for all cases in table 1 plotted against Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. The filled markers correspond to cases at Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630, while the open faced markers markers correspond to cases at Reτ=315subscriptRe𝜏315\mathrm{Re}_{\tau}=315roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 315.

As the particles settle, they experience a mean vertical slip velocity according to (14), and they also experience a mean horizontal slip due to the background shear. 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:

φr(i)=1DDus,i2zus,iz2+us,i2z𝑑z,φs(i)=1DDus,i2zuτ2𝑑z,formulae-sequencesuperscriptsubscript𝜑𝑟𝑖1𝐷subscript𝐷subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠𝑖2𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑖𝑧2subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠𝑖2𝑧differential-d𝑧superscriptsubscript𝜑𝑠𝑖1𝐷subscript𝐷subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠𝑖2𝑧superscriptsubscript𝑢𝜏2differential-d𝑧\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}}dz,\quad\varphi_{s}^{(i)}=\frac{1}{D}\int_{D}\frac{\langle{u_{s,i}% ^{\prime}}^{2}\rangle_{z}}{u_{\tau}^{2}}dz,italic_φ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_D end_ARG ∫ start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT divide start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_s , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_s , italic_i end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⟨ italic_u start_POSTSUBSCRIPT italic_s , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG italic_d italic_z , italic_φ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_D end_ARG ∫ start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT divide start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_s , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_d italic_z , (21)

respectively, where D𝐷Ditalic_D is the vertical sub-region of the domain between z+=50superscript𝑧50z^{+}=50italic_z start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 50 and z/H=0.75𝑧𝐻0.75z/H=0.75italic_z / italic_H = 0.75, and i=x,z𝑖𝑥𝑧i=x,\,zitalic_i = italic_x , italic_z. These bounds were chosen to eliminate edge effects from the upper boundary condition, though the results are not significantly affected by the choice of the lower bound of the 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 (b), while the components of the integrated slip variance are shown in 3(c) and (d). Each marker style corresponds to a different value for St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, and the results are plotted against Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT to highlight the role of settling. Filled markers correspond to runs with Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630, empty markers correspond to runs with Reτ=315subscriptRe𝜏315\mathrm{Re}_{\tau}=315roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 315, and the empty marker filled with an x corresponds to the run with Reτ=1260subscriptRe𝜏1260\mathrm{Re}_{\tau}=1260roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. For example, for small Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (independent of St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, 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 (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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT (except for perhaps particles with St+=10superscriptSt10\mathrm{St^{+}}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT increases, particles with St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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, figures 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 comes from the fact that the the magnitude of τpϱddzϱwp2zsubscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧\frac{\tau_{p}}{\varrho}\frac{d}{dz}\varrho\langle w_{p}^{2}\rangle_{z}divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT decreases within the region D𝐷Ditalic_D. It is known from Bragg et al. (2021a) that this term is negative within the logarithmic region of the flow, so as it decreases, there is an associated increase in the squared mean slip, usz2superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧2\langle u_{s}\rangle_{z}^{2}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. As the interior of the domain becomes decoupled from the solid boundary as the Reynolds number increases, the relevance of this term becomes diminished, contributing to an overall decrease in the relative slip variance. 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and small to moderate St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, 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 et al. (2010) showed that the elongated tails of fluid particle accelerations within the buffer layer and viscous sublayer are due to the vortical structures impinging on the viscous sublayer. Lavezzo et al. (2010) attributed particles settling through these same vortical structures to the increase in spanwise and streamwise acceleration variance. In a two-way coupled turbulent Couette flow, Richter & Sullivan (2013) found that particles tend to damp vertical fluctuations of the near wall dynamics, suggesting a complex feedback cycle.

In the limit of g+0superscriptg0\mathrm{g}^{+}\rightarrow 0roman_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT → 0 (recall that g+=Sv+/St+)g^{+}=\mathrm{Sv}^{+}/\mathrm{St}^{+})italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT / roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ), or when gravitational accelerations are ignored a priori, Bec et al. (2006) demonstrated that particles tend to cluster in strain dominated regions of the flow for low Stokes number (i.e. Stη0.3less-than-or-similar-tosubscriptSt𝜂0.3\mathrm{St_{\eta}}\lesssim 0.3roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT ≲ 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 timescales greater than τp1superscriptsubscript𝜏𝑝1\tau_{p}^{-1}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Ultimately, both processes work to reduce the acceleration variance, but for different reasons (Bragg et al., 2015). 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, 1963). Berk & Coletti (2021) and Ireland et al. (2016a) showed that the importance of the crossing trajectories mechanism on the acceleration statistics is due to both StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT and SvηsubscriptSv𝜂\mathrm{Sv}_{\eta}roman_Sv start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT (or alternatively 1/Fr=g/aη1Fr𝑔subscript𝑎𝜂1/\mathrm{Fr}=g/a_{\eta}1 / roman_Fr = italic_g / italic_a start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT). They showed that for large g/aη𝑔subscript𝑎𝜂g/a_{\eta}italic_g / italic_a start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT (equivalent to large g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT in our context), gravitational accelerations become increasingly important to the dynamics, leading to a peak in the acceleration variance at sufficiently high g/aη𝑔subscript𝑎𝜂g/a_{\eta}italic_g / italic_a start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, around Stη=𝒪(1)subscriptSt𝜂𝒪1\mathrm{St}_{\eta}=\mathcal{O}(1)roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT = caligraphic_O ( 1 ). In the following results, we highlight some similarities of the computed slip and acceleration variance to results from Berk & Coletti (2021), who focused on modeling 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 turbulent boundary layer to the predictions of their model for homogeneous turbulence (derived in Appendix B).

Refer to caption
Figure 4: Panel (a) shows the slip variance normalized by the seen variance for all cases in table 1 at Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630. The horizontal dashed lines denote heights of z+=50superscript𝑧50z^{+}=50italic_z start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 50 and z/H=0.75𝑧𝐻0.75z/H=0.75italic_z / italic_H = 0.75. Panel (b) shows the normalized acceleration variance over the same range plotted against the local value of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, while the dashed line represents the Stη2superscriptsubscriptSt𝜂2\mathrm{St}_{\eta}^{-2}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT scaling. The colors of each curve in panels (a) and (b) correspond to values of St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, while the line styles correspond to values of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. Panel (c) shows ratio of the seen variance to the unconditional variance averaged over the entire vertical extent plotted against Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT for all cases at Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630.

First, we consider profiles of the relative slip variance, which we define as us2z/uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{s}^{\prime}}^{2}\rangle_{z}/\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, shown in figure 4(a). Here we again focus on the region D𝐷Ditalic_D, which is the region between z+=50superscript𝑧50z^{+}=50italic_z start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 50 and z/H=0.75𝑧𝐻0.75z/H=0.75italic_z / italic_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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT at a given St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, the relative slip variance tends to increase, with the most dramatic increase coming as Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT is increased beyond unity, which is probably a reflection of the crossing trajectories mechanism. Furthermore, we can see that relative change between Sv+=0.25superscriptSv0.25\mathrm{Sv}^{+}=0.25roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 0.25 and Sv+=2.5superscriptSv2.5\mathrm{Sv^{+}}=2.5roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 2.5 decreases as St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT increases. The reason for this is discussed more below.

We can also consider the relative acceleration variance, ap2z/uf2zτη2subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑎𝑝2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧superscriptsubscript𝜏𝜂2\langle{a_{p}^{\prime}}^{2}\rangle_{z}/\langle{u_{f}^{\prime}}^{2}\rangle_{z}% \tau_{\eta}^{-2}⟨ italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, plotted against the local value of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, shown in figure 4(b). We can see that as the range of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT increases (by changing St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), the relative acceleration variance approaches the asymptotic relationship Stη2superscriptsubscriptSt𝜂2\mathrm{St}_{\eta}^{-2}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. However, for moderately inertial particles, characterized by the St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 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 Sv+=2.5superscriptSv2.5\mathrm{Sv}^{+}=2.5roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 2.5 for these particles, the acceleration variance is much larger, and tends to scale as Stη1superscriptsubscriptSt𝜂1\mathrm{St}_{\eta}^{-1}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, which is consistent with the results from Balachandar (2009) when τητpτl,pmuch-less-thansubscript𝜏𝜂subscript𝜏𝑝much-less-thansubscript𝜏𝑙𝑝\tau_{\eta}\ll\tau_{p}\ll\tau_{l,p}italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT ≪ italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≪ italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT and from Berk & Coletti (2021) in roughly the same range of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT. From figure 4(b), we can see that the crossing trajectories mechanism is not strong for large St+superscriptSt\mathrm{St^{+}}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT particles, since these particles approach the asymptotic Stη2superscriptsubscriptSt𝜂2\mathrm{St}_{\eta}^{-2}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT 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 Stη1similar-tosubscriptSt𝜂1\mathrm{St}_{\eta}\sim 1roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT ∼ 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 analyzing 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 (2006) 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. This integrated ratio approaches unity as Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 modeling purposes as our results show that the fluid velocity variance along the particle trajectory may be substituted for the seen variance in a turbulent boundary layer without incurring significant error, having implications for the predictive power the particle statistics in a TBL. Berk & Coletti (2021) 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 (18). We can see in figure 5(a) the relative variance decreases as a function of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, as we would expect based on figure 4, but it is interesting that increasing ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT 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 ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, but since the particle velocity variance increases faster with ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, 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 τpsubscript𝜏𝑝\tau_{p}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT). The increases in uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT with Reynolds number when averaged across the domain is probably due to the increasing size of the quasi-homogeneous region of the flow (Kunkel & Marusic, 2006). 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 uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are related to the unconditional fluid velocity variance in some way, we expect that they should follow this behaviour, at least qualitatively. Finally, Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT and Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT are non-zero within the interior of the domain, but their magnitude, discussed later, is secondary to both uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT when taking the average across the logarithmic layer (recall the D𝐷Ditalic_D does not include the viscous sublayer).

Refer to caption
Figure 5: Panel (a) shows the normalized acceleration variance for cases with St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 10 and Sv+=0.8superscriptSv0.8\mathrm{Sv}^{+}=0.8roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 0.8 at three different Reynolds numbers as a function of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT. Panel(b) shows the seen variance (filled markers), slip variance (colored empty markers), and particle velocity variance (black markers) normalized by the unconditional variance integrated over the range D𝐷Ditalic_D as a function of ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT.
Refer to caption
Figure 6: Shown in panels (a)–(d) are D1subscript𝐷1D_{1}italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, D2subscript𝐷2D_{2}italic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and the RMS values of Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT normalized by the seen variance for all cases in Table 1, respectively. Open faced markers represent cases with Reτ=315subscriptRe𝜏315\mathrm{Re}_{\tau}=315roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 315, while filled markers represent cases with Reτ=630subscriptRe𝜏630\mathrm{Re}_{\tau}=630roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = 630.

To conclude this section, we comment on the applicability of the model proposed by Berk & Coletti (2021) for us2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧\langle{u_{s}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (which we will refer to as BC2021 within the text). A sketch of the derivation of their model is presented in Appendix B. In short, they invoke an argument from Csanady (1963) 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 auto-correlation of the fluid velocity along the particle trajectory, which they represent using a two timescale model derived by Sawford (1991). The auto-correlation function involves the decorrelation timescale of the turbulence along the particle trajectory, τl,psubscript𝜏𝑙𝑝\tau_{l,p}italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT, for which they use the model derived in Csanady (1963). 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 modeled 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. We first consider (18) normalized by the seen variance:

us2zuf2z=1wp2zuf2zH+Rtuf2z+Rguf2zS.subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧superscriptsubscript1subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧𝐻subscript𝑅𝑡subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧subscript𝑅𝑔subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧𝑆\frac{\langle{u_{s}^{\prime}}^{2}\rangle_{z}}{\langle{u_{f}^{\prime}}^{2}% \rangle_{z}}=\overbrace{\underbrace{1-\frac{\langle{w_{p}^{\prime}}^{2}\rangle% _{z}}{\langle{u_{f}^{\prime}}^{2}\rangle_{z}}}_{\text{$H$}}+\frac{R_{t}}{% \langle{u_{f}^{\prime}}^{2}\rangle_{z}}+\frac{R_{g}}{\langle{u_{f}^{\prime}}^{% 2}\rangle_{z}}}^{\text{$S$}}.divide start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG = over⏞ start_ARG under⏟ start_ARG 1 - divide start_ARG ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG end_ARG start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT + divide start_ARG italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG end_ARG start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT . (22)

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 B2021 is smaller in magnitude than the slip variance computed by the DNS throughout the this region.

As the model developed in Berk & Coletti (2021) was for settling particles in homogeneous turbulence, the slip variance in that case was only due to the difference between uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (see Appendix B). We denote this subset of terms in (22) by H𝐻Hitalic_H after normalizing by uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Therefore, our first comparison is between H𝐻Hitalic_H and the BC2021 model (which we denote by M𝑀Mitalic_M for brevity); mathematically, we consider

D1=1DDHM𝑑z,subscript𝐷11𝐷subscript𝐷norm𝐻𝑀differential-d𝑧D_{1}=\frac{1}{D}\int_{D}\|H-M\|dz,italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_D end_ARG ∫ start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ∥ italic_H - italic_M ∥ italic_d italic_z , (23)

which is shown in figure 6(a).

This metric highlights a disparity between the DNS and BC2021, but the differences amount to no more than roughly 0.3uf2z0.3subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧0.3\langle{u_{f}^{\prime}}^{2}\rangle_{z}0.3 ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (note both H𝐻Hitalic_H and M𝑀Mitalic_M have a factor of uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT in their denominator). 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 D1subscript𝐷1D_{1}italic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT increases, and will be discussed more in section 4. It is also interesting to note that there are differences associated with St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, and these differences tend to plateau at large St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT.

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 sub-models required to calculate characteristic parameters (i.e. C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT; see Appendix B) 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 (2002)). However, we can see that by increasing ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, the differences between the DNS and BC2021 tend to decrease suggesting a correspondence at large enough Reynolds number.

Now, we consider how BC2021 compares to the full right hand side of 22, which includes Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (note that Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT cannot appear in BC2021 a priori due to their assumption of homogeneous turbulence). We will denote this as S𝑆Sitalic_S and consider the metric

D2=1DDSM𝑑z,subscript𝐷21𝐷subscript𝐷norm𝑆𝑀differential-d𝑧D_{2}=\frac{1}{D}\int_{D}\|S-M\|dz,italic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_D end_ARG ∫ start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ∥ italic_S - italic_M ∥ italic_d italic_z , (24)

which is plotted in figure 6(b).

It is apparent that the inclusion of Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT actually works to decrease differences between the DNS data and BC2021. However, as we established 2(c), (f), and (i), Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is negligible within the logarithmic region, and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is primarily negative, causing a reduction of the relative slip variance. As we noted previously, M𝑀Mitalic_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 root-mean-squared of the normalized Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, shown in figures 6(c) and (d), we can see that Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is much more important than Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. Again, the importance of this term appears when both St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT are large, but decreases significantly as ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT increases.

In summary, BC2021 gives a reasonable estimate for the DNS data in the logarithmic region of the flow. However, at large St+superscriptSt\mathrm{St^{+}}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, 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 Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT will also be reduced as ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT 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 St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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. (2021a) and Johnson et al. (2020) for the slip velocity variance for settling inertial particles in a turbulent boundary layer 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, uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, and the particle velocity variance wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, except for the largest value of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, where all terms in (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, uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT 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+=0superscript𝑧0z^{+}=0italic_z start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 0 as wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT remained finite. The higher order terms tended to peak within this layer, and the magnitude of the peak tended to increase with Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. 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 St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 (1963), for example). This conclusion is quantitatively consistent to the conclusions presented in Berk & Coletti (2021) for their laboratory experiments in homogeneous turbulence. This correspondence is significant as the seen variance is not known a priori. There exist approaches to modeling this quantity (Pozorski & Minier, 1998; Minier & Peirano, 2001), but the results of our work suggest that even in a turbulent boundary layer where there is spatial dependence in the 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 (2006), 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, us2zsubscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠2𝑧\langle{u_{s}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 Sv+,superscriptSv\mathrm{Sv}^{+},roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT , 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 to a model derived for homogeneous isotropic turbulence by Berk & Coletti (2021) (as no such model currently exists for a turbulent boundary layer 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 turbulent boundary layers 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 (2021) outside of the thin viscous sublayer.

Additionally, due to the inhomogeneous nature of the turbulence, non-local effects implicit to uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT may occur at large Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT. Isolating the importance of non-local effects in a TBL is the focus on ongoing research (for a recent example for a model of wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT in a TBL, see Zhang et al. (2023) 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: δ|τpwpz|similar-to𝛿subscript𝜏𝑝subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧\delta\sim|\tau_{p}\langle w_{p}\rangle_{z}|italic_δ ∼ | italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT |, where wpzsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧\langle w_{p}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the average particle settling velocity conditioned on a height z𝑧zitalic_z given by (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 z𝑧zitalic_z, k𝑘kitalic_k. Taylor expanding about this point and truncating after the second term, we have

k(zδ)=k(z)δdkdz|z.𝑘𝑧𝛿𝑘𝑧evaluated-at𝛿𝑑𝑘𝑑𝑧𝑧k(z-\delta)=k(z)-\left.\delta\frac{dk}{dz}\right|_{z}.italic_k ( italic_z - italic_δ ) = italic_k ( italic_z ) - italic_δ divide start_ARG italic_d italic_k end_ARG start_ARG italic_d italic_z end_ARG | start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT . (25)

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

k(z)δdkdz|z,much-greater-than𝑘𝑧evaluated-at𝛿𝑑𝑘𝑑𝑧𝑧k(z)\gg\left.\delta\frac{dk}{dz}\right|_{z},italic_k ( italic_z ) ≫ italic_δ divide start_ARG italic_d italic_k end_ARG start_ARG italic_d italic_z end_ARG | start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (26)

i.e. the kinetic energy in a small neighborhood about z𝑧zitalic_z (defined by the distance δ𝛿\deltaitalic_δ) is primarily defined by the kinetic energy measured at a height z𝑧zitalic_z. Using (25), we can estimate under what conditions a locally homogeneous approximation would be appropriate by looking for cases where the second term is small compared to the first. By assuming that the gradient of the turbulent kinetic energy scales as uτ2z1superscriptsubscript𝑢𝜏2superscript𝑧1u_{\tau}^{2}z^{-1}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (also implying k𝑘kitalic_k can be scaled by uτ2superscriptsubscript𝑢𝜏2u_{\tau}^{2}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) (Smits et al., 2011) we have that zδmuch-greater-than𝑧𝛿z\gg\deltaitalic_z ≫ italic_δ. By normalizing both sides of this inequality by the root-mean-squared turbulent velocity, u=k1/2superscript𝑢superscript𝑘12u^{\prime}=k^{1/2}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_k start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT, we can write δ𝛿\deltaitalic_δ in terms of the sum of the settling enhancement, E=wpz+vgu𝐸subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscript𝑣𝑔superscript𝑢E=\frac{\langle w_{p}\rangle_{z}+v_{g}}{u^{\prime}}italic_E = divide start_ARG ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG and vgsubscript𝑣𝑔v_{g}italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, (Good et al., 2014; Loth, 2023) as

τp|E+vgw|zu.much-less-thansubscript𝜏𝑝𝐸subscript𝑣𝑔superscript𝑤𝑧superscript𝑢\tau_{p}\left|E+\frac{v_{g}}{w^{\prime}}\right|\ll\frac{z}{u^{\prime}}.italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT | italic_E + divide start_ARG italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_w start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG | ≪ divide start_ARG italic_z end_ARG start_ARG italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG . (27)

Now normalizing by τηsubscript𝜏𝜂\tau_{\eta}italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT, and observing that uuτsimilar-tosuperscript𝑢subscript𝑢𝜏u^{\prime}\sim u_{\tau}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∼ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT, we can simplify both side of this inequality to reveal that

Stη|E+Sv|(zuτν)1/2,much-less-thansubscriptSt𝜂𝐸subscriptSvsuperscript𝑧subscript𝑢𝜏𝜈12\displaystyle\mathrm{St}_{\eta}\left|E+\mathrm{Sv}_{\ell}\right|\ll\left(\frac% {zu_{\tau}}{\nu}\right)^{1/2},roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT | italic_E + roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT | ≪ ( divide start_ARG italic_z italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT end_ARG start_ARG italic_ν end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , (28)

where we have used the dissipation scaling in 4 to relate τηsubscript𝜏𝜂\tau_{\eta}italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT to the vertical coordinate, z𝑧zitalic_z. This relationship indicates that both particle inertia and gravity have an explicit role, and an implicit role (through E𝐸Eitalic_E) to play in potential non-local effects.

We know from Good et al. (2014) and Loth (2023) that in small scale laboratory experiments, E0.2similar-to𝐸0.2E\sim 0.2italic_E ∼ 0.2 as StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT and SvsubscriptSv\mathrm{Sv}_{\ell}roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT approach 1, but as both of these parameters increase, E𝐸Eitalic_E tends back towards zero, and may even become negative (Ferran et al., 2023). Therefore, for the coarse particles we are concerned with in this work, we can make a locally homogeneous approximation when StηSv(zuτν)1/2much-less-thansubscriptSt𝜂subscriptSvsuperscript𝑧subscript𝑢𝜏𝜈12\mathrm{St_{\eta}Sv_{\ell}}\ll\left(\frac{zu_{\tau}}{\nu}\right)^{1/2}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ≪ ( divide start_ARG italic_z italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT end_ARG start_ARG italic_ν end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. This may not particularly restrictive for the atmospheric surface layer as the Reynolds numbers are 𝒪(106)𝒪superscript106\mathcal{O}(10^{6})caligraphic_O ( 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ), but for laboratory experiments, the integral scales tend to scale with the size of the experimental domain (i.e. zhsimilar-to𝑧z\sim hitalic_z ∼ italic_h where hhitalic_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 setup.

For the DNS presented in this work, the above relationship shows that non-local effects are likely only important for cases when both St+superscriptSt\mathrm{St^{+}}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT are large, as SvsubscriptSv\mathrm{Sv}_{\ell}roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT tends to scale with Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT since uuτsimilar-tosuperscript𝑢subscript𝑢𝜏u^{\prime}\sim u_{\tau}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∼ italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT. For example, if we consider cases with St+=100superscriptSt100\mathrm{St^{+}}=100roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 100 and Sv+=2.5superscriptSv2.5\mathrm{Sv}^{+}=2.5roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 2.5 (and assume E0𝐸0E\approx 0italic_E ≈ 0), we can see immediately that 28 is not satisfied. This may explain the differences between the DNS and the model in Berk & Coletti (2021) in figure 6(a) and (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 turbulent boundary layer, 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 St+superscriptSt\mathrm{St}^{+}roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT or StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT), and Sv+<1superscriptSv1\mathrm{Sv}^{+}<1roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT < 1, these terms are subleading outside of the viscous sublayer. Moreover, the magnitudes of these terms in the logarithmic layer tend to diminish as ReτsubscriptRe𝜏\mathrm{Re}_{\tau}roman_Re start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT increases. Thus, outside of the viscous sublayer, and at moderate local StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT and Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, we can extend models designed for homogeneous turbulence (like that described in Berk & Coletti (2021)) to a TBL, where we must interpret the model as local to a height z𝑧zitalic_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 modeling 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 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT m2/s3superscriptm2superscripts3\mathrm{m^{2}/s^{3}}roman_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_s start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, we can define a rough Kolmogorov timescale as 101superscript10110^{-1}10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT s. Thus, for quartz dust particles (ρp=2650subscript𝜌𝑝2650\rho_{p}=2650italic_ρ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 2650 kg/m3kgsuperscriptm3\mathrm{kg/m^{3}}roman_kg / roman_m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) that range between 30-100 μm𝜇m\mathrm{\mu m}italic_μ roman_m, we should expect a values of StηsubscriptSt𝜂\mathrm{St}_{\eta}roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT to range between 0.1–10. We can see that the ranges in our DNS are in the correct neighborhood to model these same coarse dust particles. Moreover, we can use the values of g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT from table 1 (recall g+=Sv+/St+superscript𝑔superscriptSvsuperscriptStg^{+}=\mathrm{Sv^{+}/St^{+}}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT / roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT) to estimate an equivalent friction velocity, uτsuperscriptsubscript𝑢𝜏u_{\tau}^{*}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (note that since we are re-scaling, uτu_{\tau}*italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT ∗ is necessarily different than the value of uτsubscript𝑢𝜏u_{\tau}italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT 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

uτ=(gνg+)1/3.superscriptsubscript𝑢𝜏superscript𝑔𝜈superscript𝑔13u_{\tau}^{*}=\left(\frac{g\nu}{g^{+}}\right)^{1/3}.italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = ( divide start_ARG italic_g italic_ν end_ARG start_ARG italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT . (29)

Assuming g=9.81m/s2𝑔9.81msuperscripts2g=9.81\mathrm{\,m/s^{2}}italic_g = 9.81 roman_m / roman_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and ν=1.57×105m2/s𝜈1.57superscript105superscriptm2s\nu=1.57\times 10^{-5}\mathrm{\,m^{2}/s}italic_ν = 1.57 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_s, and some relationship between 10 metre wind velocity and the friction velocity (see Kantha & Clayson (2000) for example), this gives us a proxy for wind speed at the field scale. For the values of g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT in this manuscript (see table 1), the effective friction velocities vary between 0.09 m/sms\mathrm{m/s}roman_m / roman_s and 0.84 m/sms\mathrm{m/s}roman_m / roman_s, which covers a wide range of friction velocities on Earth (Vickers et al., 2015). Since g+superscript𝑔g^{+}italic_g start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT is proportional to Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, we can see the effective wind speed increases as Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT decreases.

Therefore, the insight we can gain is that the slip velocity in high wind conditions (small Sv+)\mathrm{Sv}^{+})roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) 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. Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT) to be more important to the dynamics further 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 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 Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT is small enough (see (28)), 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 modeling the particle Reynolds number, which is known to affect the associated drag on the particles (Balachandar (2009), Berk & Coletti 2024 (under review)). For example, it is known that loitering effects are typically associated with large particle Reynolds number (Rosa et al., 2016), and these loitering effects work to reduce the average particle settling velocity Good et al. (2014). Accurate modeling of loitering effects could explain discrepancies between numerical simulations and laboratory experiments with respect to the measurement of settling velocities (Ferran et al., 2023). 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., 2018), which could be linked to loitering effects.

5 Acknowledgements

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.

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, vgsubscript𝑣𝑔v_{g}italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, 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 squared mean is not equivalent to the average, i.e. F2F2delimited-⟨⟩superscript𝐹2delimited-⟨⟩superscriptsuperscript𝐹2\langle F^{2}\rangle\neq\langle{F^{\prime}}^{2}\rangle⟨ italic_F start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ≠ ⟨ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩, (F𝐹Fitalic_F is some arbitrary quantity, and a prime indicates a fluctuation about the mean of F𝐹Fitalic_F) meaning we must be careful to discern between the variance and squared means:

wp2z=wpz2+wp2z,subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝2𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧2subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\displaystyle\langle w_{p}^{2}\rangle_{z}=\langle w_{p}\rangle_{z}^{2}+\langle% {w_{p}^{\prime}}^{2}\rangle_{z},⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (30)
us2z=usz2+us2z,usz=ufzwpz,formulae-sequencesubscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠2𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧2subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscriptdelimited-⟨⟩subscript𝑢𝑓𝑧subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧\displaystyle\langle u_{s}^{2}\rangle_{z}=\langle u_{s}\rangle_{z}^{2}+\langle% {u_{s}^{\prime}}^{2}\rangle_{z},\quad\langle u_{s}\rangle_{z}=\langle u_{f}% \rangle_{z}-\langle w_{p}\rangle_{z},⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (31)
wp3z=wpz3+wpzwp2z+wp3z.subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧3subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝3𝑧\displaystyle\langle w_{p}^{3}\rangle_{z}=\langle w_{p}\rangle_{z}^{3}+\langle w% _{p}\rangle_{z}\langle{w_{p}^{\prime}}^{2}\rangle_{z}+\langle{w_{p}^{\prime}}^% {3}\rangle_{z}.⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT . (32)

Eq. (17), derived in Johnson et al. (2020), assumed that particles did not settle under the action of gravity, meaning that wpz=0subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧0\langle w_{p}\rangle_{z}=0⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0. However, by substituting in the above Reynolds decompositions, it can be readily extended to settling particles. Doing this, we arrive at

us2z=uf2zwp2zτpϱddzϱwp3z+ufz2usz2wpz22wpzvg,subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑓𝑧2superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧2superscriptsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧22subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscript𝑣𝑔\langle{u_{s}^{\prime}}^{2}\rangle_{z}=\langle{u_{f}^{\prime}}^{2}\rangle_{z}-% \langle{w_{p}^{\prime}}^{2}\rangle_{z}-\frac{\tau_{p}}{\varrho}\frac{d}{dz}% \varrho\langle w_{p}^{3}\rangle_{z}+\langle u_{f}\rangle_{z}^{2}-\langle u_{s}% \rangle_{z}^{2}-\langle w_{p}\rangle_{z}^{2}-2\langle w_{p}\rangle_{z}v_{g},⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , (33)

where we have not expanded wp3zsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧\langle w_{p}^{3}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT 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, usz2superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧2\langle u_{s}\rangle_{z}^{2}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT can be expanded as

usz2=ufz22ufzwpz+wpz2.superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑠𝑧2superscriptsubscriptdelimited-⟨⟩subscript𝑢𝑓𝑧22subscriptdelimited-⟨⟩subscript𝑢𝑓𝑧subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧superscriptsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧2\langle u_{s}\rangle_{z}^{2}=\langle u_{f}\rangle_{z}^{2}-2\langle u_{f}% \rangle_{z}\langle w_{p}\rangle_{z}+\langle w_{p}\rangle_{z}^{2}.⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (34)

Upon substitution of the above into (33) and simplifying, we can express (33) as

us2z=uf2zwp2zτpϱddzϱwp3z+2wpz(uszvg)subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝3𝑧2subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscript𝑣𝑔\langle{u_{s}^{\prime}}^{2}\rangle_{z}=\langle{u_{f}^{\prime}}^{2}\rangle_{z}-% \langle{w_{p}^{\prime}}^{2}\rangle_{z}-\frac{\tau_{p}}{\varrho}\frac{d}{dz}% \varrho\langle w_{p}^{3}\rangle_{z}+2\langle w_{p}\rangle_{z}\left(\langle u_{% s}\rangle_{z}-v_{g}\right)⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) (35)

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.

us2z=uf2zwp2z(1)τpϱddzϱwp3z(2)τpϱddz(ϱwpz3+3ϱwpzwp2z)+2wpz(uszvg)(3)subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧superscriptsubscriptsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧1subscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝3𝑧2subscriptsubscript𝜏𝑝italic-ϱ𝑑𝑑𝑧italic-ϱsubscriptsuperscriptdelimited-⟨⟩subscript𝑤𝑝3𝑧3italic-ϱsubscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧2subscriptdelimited-⟨⟩subscript𝑤𝑝𝑧subscriptdelimited-⟨⟩subscript𝑢𝑠𝑧subscript𝑣𝑔3\langle{u_{s}^{\prime}}^{2}\rangle_{z}=\overbrace{\underbrace{\langle{u_{f}^{% \prime}}^{2}\rangle_{z}-\langle{w_{p}^{\prime}}^{2}\rangle_{z}}_{(1)}-\frac{% \tau_{p}}{\varrho}\frac{d}{dz}\varrho\langle{w_{p}^{\prime}}^{3}\rangle_{z}}^{% (2)}\\ \underbrace{-\frac{\tau_{p}}{\varrho}\frac{d}{dz}\left(\varrho\langle w_{p}% \rangle^{3}_{z}+3\varrho\langle w_{p}\rangle_{z}\langle{w_{p}^{\prime}}^{2}% \rangle_{z}\right)+2\langle w_{p}\rangle_{z}\left(\langle u_{s}\rangle_{z}-v_{% g}\right)}_{(3)}start_ROW start_CELL ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = over⏞ start_ARG under⏟ start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_POSTSUBSCRIPT ( 1 ) end_POSTSUBSCRIPT - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL under⏟ start_ARG - divide start_ARG italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_ϱ end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_z end_ARG ( italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + 3 italic_ϱ ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) + 2 ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) end_ARG start_POSTSUBSCRIPT ( 3 ) end_POSTSUBSCRIPT end_CELL end_ROW (36)

Appendix B – Response function model of acceleration variance

In this section, we sketch the derivation of the semi-analytical model proposed by Berk & Coletti (2021) for the slip velocity variance. The following is based on the concept of a response function, described in Csanady (1963), 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

us2=0(ωτp)2Ep𝑑ω,wp2=0Ep𝑑ω.formulae-sequencedelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2superscriptsubscript0superscript𝜔subscript𝜏𝑝2subscript𝐸𝑝differential-d𝜔delimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2superscriptsubscript0subscript𝐸𝑝differential-d𝜔\langle{u_{s}^{\prime}}^{2}\rangle=\int_{0}^{\infty}(\omega\tau_{p})^{2}E_{p}d% \omega,\quad\langle{w_{p}^{\prime}}^{2}\rangle=\int_{0}^{\infty}E_{p}d\omega.⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( italic_ω italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_d italic_ω , ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_d italic_ω . (37)

In this equation Epsubscript𝐸𝑝E_{p}italic_E start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the kinetic energy spectrum of the particles. As discussed in Csanady (1963), Epsubscript𝐸𝑝E_{p}italic_E start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is related to the kinetic energy spectrum of the fluid motion sampled along the particle’s trajectory, denoted by Ef,psubscript𝐸𝑓𝑝E_{f,p}italic_E start_POSTSUBSCRIPT italic_f , italic_p end_POSTSUBSCRIPT, through a response function

H(ω)=11+(ωτp)2,𝐻𝜔11superscript𝜔subscript𝜏𝑝2H(\omega)=\frac{1}{1+\left(\omega\tau_{p}\right)^{2}},italic_H ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG 1 + ( italic_ω italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (38)

meaning that

us2=0(ωτp)2H(ω)Ef,p𝑑ω,wp2=0H(ω)Ef,p𝑑ω.formulae-sequencedelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2superscriptsubscript0superscript𝜔subscript𝜏𝑝2𝐻𝜔subscript𝐸𝑓𝑝differential-d𝜔delimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2superscriptsubscript0𝐻𝜔subscript𝐸𝑓𝑝differential-d𝜔\langle{u_{s}^{\prime}}^{2}\rangle=\int_{0}^{\infty}(\omega\tau_{p})^{2}H(% \omega)E_{f,p}d\omega,\quad\langle{w_{p}^{\prime}}^{2}\rangle=\int_{0}^{\infty% }H(\omega)E_{f,p}d\omega.⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( italic_ω italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_H ( italic_ω ) italic_E start_POSTSUBSCRIPT italic_f , italic_p end_POSTSUBSCRIPT italic_d italic_ω , ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_H ( italic_ω ) italic_E start_POSTSUBSCRIPT italic_f , italic_p end_POSTSUBSCRIPT italic_d italic_ω . (39)

Using the stochastic model for the particle velocity auto-correlation outlined in Sawford (1991), Berk & Coletti (2021) used the fact that the auto-correlation and the spectra are fourier transform pairs. The particle velocity auto-correlation described in Sawford (1991) is

Rf,p(t)=uf2τl,pτ2(τl,pet/τl,p+τ2et/τ2),subscript𝑅𝑓𝑝𝑡delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2subscript𝜏𝑙𝑝subscript𝜏2subscript𝜏𝑙𝑝superscript𝑒𝑡subscript𝜏𝑙𝑝subscript𝜏2superscript𝑒𝑡subscript𝜏2R_{f,p}(t)=\frac{\langle{u_{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),italic_R start_POSTSUBSCRIPT italic_f , italic_p end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT - italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ( italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) , (40)

where τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is proportional to the fluid acceleration variance and appears due to the finite Reynolds number. For this work, we use

τ2=C02a0τη,subscript𝜏2subscript𝐶02subscript𝑎0subscript𝜏𝜂\tau_{2}=\frac{C_{0}}{2a_{0}}\tau_{\eta},italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT , (41)

where C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are universal constants modeled by

C0=C(1(0.1Reλ1/2)),a0=(51+100Reλ1),formulae-sequencesubscript𝐶0subscript𝐶10.1superscriptsubscriptRe𝜆12subscript𝑎051100Rsuperscriptsubscripte𝜆1C_{0}=C_{\infty}(1-(0.1\mathrm{Re_{\lambda}}^{-1/2})),\quad a_{0}=\left(\frac{% 5}{1+100\mathrm{Re}_{\lambda}^{-1}}\right),italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ( 1 - ( 0.1 roman_Re start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ) ) , italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( divide start_ARG 5 end_ARG start_ARG 1 + 100 roman_R roman_e start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG ) , (42)

defined in Lien & D’Asaro (2002) and Sawford et al. (2003) respectively. Here, Reλ=15w2/vη2subscriptRe𝜆15delimited-⟨⟩superscript𝑤2superscriptsubscript𝑣𝜂2\mathrm{Re}_{\lambda}=\sqrt{15}\langle w^{2}\rangle/v_{\eta}^{2}roman_Re start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = square-root start_ARG 15 end_ARG ⟨ italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / italic_v start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the Taylor Reynolds number evaluated at a height z𝑧zitalic_z. Thus, C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are functions of the vertical coordinate. Note that the results in this work are not meaningfully dependent on the exact choice of model for C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

τl,psubscript𝜏𝑙𝑝\tau_{l,p}italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT is the lagrangian correlation timescale of the turbulence along the particle trajectory. τl,psubscript𝜏𝑙𝑝\tau_{l,p}italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT is a function of three parameters: ratio of the laminar settling velocity to the integral velocity scale, the lagrangian correlation timescale of the turbulence, and the Eulerian correlation timescale of the turbulence which are defined as

Sv=vgw,τE=w2ϵ,τ=κzuτuww2,formulae-sequencesubscriptSvsubscript𝑣𝑔superscript𝑤formulae-sequencesubscript𝜏𝐸delimited-⟨⟩superscript𝑤2italic-ϵ𝜏𝜅𝑧subscript𝑢𝜏delimited-⟨⟩𝑢𝑤delimited-⟨⟩superscript𝑤2\mathrm{Sv}_{\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},roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT = divide start_ARG italic_v start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_w start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG , italic_τ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = divide start_ARG ⟨ italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG start_ARG italic_ϵ end_ARG , italic_τ = - divide start_ARG italic_κ italic_z end_ARG start_ARG italic_u start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT end_ARG divide start_ARG ⟨ italic_u italic_w ⟩ end_ARG start_ARG ⟨ italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG , (43)

respectively. Note that the definition of τ𝜏\tauitalic_τ can be found in Oesterlé & Zaichik (2004). τl,psubscript𝜏𝑙𝑝\tau_{l,p}italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT is meant to encapsulate the fact that as an inertial particle settles through a local neighborhood of correlated motion, the turbulence it experiences de-correlates faster along its trajectory than it would if it was not settling. Berk & Coletti (2021) derived a semi-empirical model for the correlation along the particle trajectory using the idea of the crossing trajectories mechanism introduced by Csanady (1963) as:

τl,p=τ1(1+(ττE)2Sv2)1/2subscript𝜏𝑙𝑝𝜏1superscript1superscript𝜏subscript𝜏𝐸2superscriptsubscriptSv212\tau_{l,p}=\tau\frac{1}{\left(1+\left(\frac{\tau}{\tau_{E}}\right)^{2}\mathrm{% Sv}_{\ell}^{2}\right)^{1/2}}italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT = italic_τ divide start_ARG 1 end_ARG start_ARG ( 1 + ( divide start_ARG italic_τ end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Sv start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG (44)

By Fourier Transforming Rf,psubscript𝑅𝑓𝑝R_{f,p}italic_R start_POSTSUBSCRIPT italic_f , italic_p end_POSTSUBSCRIPT and substituting into the integral relations for the slip variance and velocity variance, we arrive at the following for the velocity variance

wp2=uf2(1Stη2(Stη+τl,pτη)(Stη+τ2τη)).delimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓21subscriptsuperscriptSt2𝜂subscriptSt𝜂subscript𝜏𝑙𝑝subscript𝜏𝜂subscriptSt𝜂subscript𝜏2subscript𝜏𝜂\langle{w_{p}^{\prime}}^{2}\rangle=\langle{u_{f}^{\prime}}^{2}\rangle\left(1-% \frac{\mathrm{St}^{2}_{\eta}}{\left(\mathrm{St}_{\eta}+\frac{\tau_{l,p}}{\tau_% {\eta}}\right)\left(\mathrm{St}_{\eta}+\frac{\tau_{2}}{\tau_{\eta}}\right)}% \right).⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ( 1 - divide start_ARG roman_St start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG start_ARG ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) end_ARG ) . (45)

and the slip variance

us2=uf2Stη2(Stη+τl,pτη)(Stη+τ2τη),delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2subscriptsuperscriptSt2𝜂subscriptSt𝜂subscript𝜏𝑙𝑝subscript𝜏𝜂subscriptSt𝜂subscript𝜏2subscript𝜏𝜂\langle{u_{s}^{\prime}}^{2}\rangle=\langle{u_{f}^{\prime}}^{2}\rangle\frac{% \mathrm{St}^{2}_{\eta}}{\left(\mathrm{St}_{\eta}+\frac{\tau_{l,p}}{\tau_{\eta}% }\right)\left(\mathrm{St}_{\eta}+\frac{\tau_{2}}{\tau_{\eta}}\right)},⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ divide start_ARG roman_St start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG start_ARG ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) end_ARG , (46)

If follows from the above that the particle acceleration variance is

ap2=uf21(Stη+τl,pτη)(Stη+τ2τη),delimited-⟨⟩superscriptsuperscriptsubscript𝑎𝑝2delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓21subscriptSt𝜂subscript𝜏𝑙𝑝subscript𝜏𝜂subscriptSt𝜂subscript𝜏2subscript𝜏𝜂\langle{a_{p}^{\prime}}^{2}\rangle=\langle{u_{f}^{\prime}}^{2}\rangle\frac{1}{% \left(\mathrm{St}_{\eta}+\frac{\tau_{l,p}}{\tau_{\eta}}\right)\left(\mathrm{St% }_{\eta}+\frac{\tau_{2}}{\tau_{\eta}}\right)},⟨ italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ divide start_ARG 1 end_ARG start_ARG ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT italic_l , italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) ( roman_St start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT + divide start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_η end_POSTSUBSCRIPT end_ARG ) end_ARG , (47)

Since the term in the brackets in (45) is simply the model for 1us2/uf21delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓21-\langle{u_{s}^{\prime}}^{2}\rangle/\langle{u_{f}^{\prime}}^{2}\rangle1 - ⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩, the implied relationship between the slip variance and the particle velocity variance is

us2=uf2wp2,delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2delimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2delimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2\langle{u_{s}^{\prime}}^{2}\rangle=\langle{u_{f}^{\prime}}^{2}\rangle-\langle{% w_{p}^{\prime}}^{2}\rangle,⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ , (48)

which is almost identical to (18), except for the fact that Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are not accounted for in this model.

Appendix C – Comparison of computed and modeled slip variance

Figure 7(a) shows a comparison between computed values of us2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑠2𝑧\langle{u_{s}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and that computed using the right hand side of (18). The slip variance computed directly from the DNS for three different values of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT at St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 10 are shown by black curves, while the right hand side of (18) are shown by colored 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 Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, and not in the computation of uf2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑢𝑓2𝑧\langle{u_{f}^{\prime}}^{2}\rangle_{z}⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and wp2zsubscriptdelimited-⟨⟩superscriptsuperscriptsubscript𝑤𝑝2𝑧\langle{w_{p}^{\prime}}^{2}\rangle_{z}⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT (as evidenced in figure 2). Figure 7(b) shows a comparison between the computed values of Rt+Rgsubscript𝑅𝑡subscript𝑅𝑔R_{t}+R_{g}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (dashed curves) and that computed by a residual of (18)(black solid curves). We can see that by plotting Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and Rgsubscript𝑅𝑔R_{g}italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT as a residual, the noise is significantly reduced.

Refer to caption
Figure 7: Panel (a) shows the slip velocity variance computed by the DNS for three values of Sv+superscriptSv\mathrm{Sv}^{+}roman_Sv start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT at fixed St+=10superscriptSt10\mathrm{St}^{+}=10roman_St start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = 10 (black curves), and the slip variance computed by (18) (dashed colored curves). Panel(b) shows Rt+Rgsubscript𝑅𝑡subscript𝑅𝑔R_{t}+R_{g}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT computed via a residual of usz(ufzwpz)subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑠𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑢𝑓𝑧subscriptdelimited-⟨⟩superscriptsubscript𝑤𝑝𝑧\langle{u_{s}^{\prime}}\rangle_{z}-\left(\langle{u_{f}^{\prime}}\rangle_{z}-% \langle{w_{p}^{\prime}}\rangle_{z}\right)⟨ italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ( ⟨ italic_u start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - ⟨ italic_w start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) (black curves), and Rt+Rgsubscript𝑅𝑡subscript𝑅𝑔R_{t}+R_{g}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT computed directly (dashed colored curves) for the same cases.

References

  • Adebiyi et al. (2023) Adebiyi, Adeyemi, Kok, Jasper F., Murray, Benjamin J., Ryder, Claire L., Stuut, Jan-Berend W., Kahn, Ralph A., Knippertz, Peter, Formenti, Paola, Mahowald, Natalie M., Pérez García-Pando, Carlos, Klose, Martina, Ansmann, Albert, Samset, Bjørn H., Ito, Akinori, Balkanski, Yves, Di Biagio, Claudia, Romanias, Manolis N., Huang, Yue & Meng, Jun 2023 A review of coarse mineral dust in the Earth system. Aeolian Research 60, 100849.
  • Aliseda et al. (2002) 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. Journal of Fluid Mechanics 468, 77–105.
  • Arcen & Tanière (2009) Arcen, Boris & Tanière, Anne 2009 Simulation of a particle-laden turbulent channel flow using an improved stochastic Lagrangian model. Physics of Fluids 21 (4), 043303.
  • Balachandar (2009) Balachandar, S. 2009 A scaling analysis for point–particle approaches to turbulent multiphase flows. International Journal of Multiphase Flow 35 (9), 801–810.
  • Bec et al. (2006) Bec, J., Biferale, L., Boffetta, G., Celani, A., Cencini, M., Lanotte, A., Musacchio, S. & Toschi, F. 2006 Acceleration statistics of heavy particles in turbulence. Journal of Fluid Mechanics 550 (-1), 349.
  • Bec et al. (2014) Bec, Jérémie, Homann, Holger & Ray, Samriddhi Sankar 2014 Gravity-Driven Enhancement of Heavy Particle Clustering in Turbulent Flow. Physical Review Letters 112 (18), 184501.
  • Berk & Coletti (2020) Berk, Tim & Coletti, Filippo 2020 Transport of inertial particles in high-Reynolds-number turbulent boundary layers. Journal of Fluid Mechanics 903, A18.
  • Berk & Coletti (2021) Berk, Tim & Coletti, Filippo 2021 Dynamics of small heavy particles in homogeneous turbulence: a Lagrangian experimental study. Journal of Fluid Mechanics 917, A47.
  • Berk & Coletti (2023) Berk, Tim & Coletti, Filippo 2023 Dynamics and scaling of particle streaks in high-Reynolds-number turbulent boundary layers. Journal of Fluid Mechanics 975, A47.
  • Bragg et al. (2015) Bragg, Andrew D., Ireland, Peter J. & Collins, Lance R. 2015 On the relationship between the non-local clustering mechanism and preferential concentration. Journal of Fluid Mechanics 780, 327–343.
  • Bragg et al. (2021a) Bragg, A. D., Richter, D. H. & Wang, G. 2021a Mechanisms governing the settling velocities and spatial distributions of inertial particles in wall-bounded turbulence. Physical Review Fluids 6 (6), 064302.
  • Bragg et al. (2021b) Bragg, A. D., Richter, D. H. & Wang, G. 2021b Settling strongly modifies particle concentrations in wall-bounded turbulent flows even when the settling parameter is asymptotically small. Physical Review Fluids 6 (12), 124301, publisher: American Physical Society.
  • Brandt & Coletti (2022) Brandt, Luca & Coletti, Filippo 2022 Particle-Laden Turbulence: Progress and Perspectives. Annual Review of Fluid Mechanics 54 (1), 159–189.
  • Csanady (1963) Csanady, G. T. 1963 Turbulent Diffusion of Heavy Particles in the Atmosphere. Journal of the Atmospheric Sciences 20 (3), 201–208.
  • Ferran et al. (2023) Ferran, Amélie, Machicoane, Nathanaël, Aliseda, Alberto & Obligado, Martín 2023 An experimental study on the settling velocity of inertial particles in different homogeneous isotropic turbulent flows. Journal of Fluid Mechanics 970, A23.
  • Gao et al. (2023) Gao, Wei, Samtaney, Ravi & Richter, David H. 2023 Direct numerical simulation of particle-laden flow in an open channel at. Journal of Fluid Mechanics 957, A3.
  • Gerashchenko et al. (2008) Gerashchenko, S., Sharp, N. S., Neuscamman, S. & Warhaft, Z. 2008 Lagrangian measurements of inertial particle accelerations in a turbulent boundary layer. Journal of Fluid Mechanics 617, 255–281.
  • Good et al. (2014) 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. Journal of Fluid Mechanics 759, R3.
  • Grace et al. (2024) Grace, Andrew P., Richter, David H. & Bragg, Andrew D. 2024 A Reinterpretation of Phenomenological Modeling Approaches for Lagrangian Particles Settling in a Turbulent Boundary Layer. Boundary-Layer Meteorology 190 (4), 15.
  • Ireland et al. (2016a) Ireland, Peter J., Bragg, Andrew D. & Collins, Lance R. 2016a The effect of Reynolds number on inertial particle dynamics in isotropic turbulence. Part 1. Simulations without gravitational effects. Journal of Fluid Mechanics 796, 617–658.
  • Ireland et al. (2016b) Ireland, Peter J., Bragg, Andrew D. & Collins, Lance R. 2016b The effect of Reynolds number on inertial particle dynamics in isotropic turbulence. Part 2. Simulations with gravitational effects. Journal of Fluid Mechanics 796, 659–711.
  • Johnson et al. (2020) Johnson, Perry L., Bassenne, Maxime & Moin, Parviz 2020 Turbophoresis of small inertial particles: theoretical considerations and application to wall-modelled large-eddy simulations. Journal of Fluid Mechanics 883, A27.
  • Kantha & Clayson (2000) Kantha, Lakshmi H & Clayson, Carol Anne 2000 Small scale processes in geophysical fluid flows. Elsevier.
  • Kok (2011) Kok, Jasper F. 2011 A scaling theory for the size distribution of emitted dust aerosols suggests climate models underestimate the size of the global dust cycle. Proceedings of the National Academy of Sciences 108 (3), 1016–1021.
  • Kok et al. (2012) Kok, Jasper F, Parteli, Eric J R, Michaels, Timothy I & Karam, Diana Bou 2012 The physics of wind-blown sand and dust. Reports on Progress in Physics 75 (10), 106901.
  • Kok et al. (2023) Kok, Jasper F., Storelvmo, Trude, Karydis, Vlassis A., Adebiyi, Adeyemi A., Mahowald, Natalie M., Evan, Amato T., He, Cenlin & Leung, Danny M. 2023 Mineral dust aerosol impacts on global climate and climate change. Nature Reviews Earth & Environment 4 (2), 71–86.
  • Kunkel & Marusic (2006) Kunkel, Gary J. & Marusic, Ivan 2006 Study of the near-wall-turbulent region of the high-Reynolds-number boundary layer using an atmospheric flow. Journal of Fluid Mechanics 548 (-1), 375.
  • Lavezzo et al. (2010) 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. Journal of Fluid Mechanics 658, 229–246.
  • Lee & Lee (2019) Lee, Junghoon & Lee, Changhoon 2019 The effect of wall-normal gravity on particle-laden near-wall turbulence. Journal of Fluid Mechanics 873, 475–507.
  • Lien & D’Asaro (2002) Lien, Ren-Chieh & D’Asaro, Eric A. 2002 The Kolmogorov constant for the Lagrangian velocity spectrum and structure function. Physics of Fluids 14 (12), 4456–4459.
  • Loth (2023) Loth, Eric 2023 Particles in a turbulent gas: Diffusion, bias, modulation and collisions. Progress in Energy and Combustion Science 97, 101094.
  • Marchioli et al. (2008) 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. International Journal of Multiphase Flow 34 (9), 879–893.
  • Maxey & Riley (1983) Maxey, Martin R. & Riley, James J. 1983 Equation of motion for a small rigid sphere in a nonuniform flow. The Physics of Fluids 26 (4), 883–889.
  • Meng et al. (2022) Meng, Jun, Huang, Yue, Leung, Danny M., Li, Longlei, Adebiyi, Adeyemi A., Ryder, Claire L., Mahowald, Natalie M. & Kok, Jasper F. 2022 Improved Parameterization for the Size Distribution of Emitted Dust Aerosols Reduces Model Underestimation of Super Coarse Dust. Geophysical Research Letters 49 (8), e2021GL097287.
  • Minier & Peirano (2001) Minier, Jean-Pierre & Peirano, Eric 2001 The pdf approach to turbulent polydispersed two-phase ows. Physics Reports .
  • Mora et al. (2021) Mora, Daniel Odens, Obligado, Martin, Aliseda, Alberto & Cartellier, Alain 2021 Effect of Re $\lambda$ and Rouse numbers on the settling of inertial droplets in homogeneous isotropic turbulence. Physical Review Fluids 6 (4), 044305.
  • Oesterlé & Zaichik (2004) Oesterlé, Benoit & Zaichik, Leonid I. 2004 On Lagrangian time scales and particle dispersion modeling in equilibrium turbulent shear flows. Physics of Fluids 16 (9), 3374–3384.
  • Pozorski & Minier (1998) Pozorski, Jacek & Minier, Jean-Pierre 1998 On the Lagrangian turbulent dispersion models based on the Langevin equation. International Journal of Multiphase Flow 24 (6), 913–945.
  • Richter & Chamecki (2018) Richter, David & Chamecki, Marcelo 2018 Inertial Effects on the Vertical Transport of Suspended Particles in a Turbulent Boundary Layer. Boundary-Layer Meteorology 167 (2), 235–256.
  • Richter & Sullivan (2013) Richter, David H. & Sullivan, Peter P. 2013 Momentum transfer in a turbulent, particle-laden Couette flow. Physics of Fluids 25 (5), 053304.
  • Rosa et al. (2016) Rosa, Bogdan, Parishani, Hossein, Ayala, Orlando & Wang, Lian-Ping 2016 Settling velocity of small inertial particles in homogeneous isotropic turbulence from high-resolution DNS. International Journal of Multiphase Flow 83, 217–231.
  • Rosenberg et al. (2014) Rosenberg, Philip D., Parker, Douglas J., Ryder, Claire L., Marsham, John H., Garcia-Carreras, Luis, Dorsey, James R., Brooks, Ian M., Dean, Angela R., Crosier, Jonathon, McQuaid, James B. & Washington, Richard 2014 Quantifying particle size and turbulent scale dependence of dust flux in the Sahara using aircraft measurements: AIRCRAFT MEASUREMENTS OF DUST FLUX. Journal of Geophysical Research: Atmospheres 119 (12), 7577–7598.
  • Ryder et al. (2019) Ryder, Claire L., Highwood, Eleanor J., Walser, Adrian, Seibert, Petra, Philipp, Anne & Weinzierl, Bernadett 2019 Coarse and giant particles are ubiquitous in Saharan dust export regions and are radiatively significant over the Sahara. Atmospheric Chemistry and Physics 19 (24), 15353–15376.
  • Ryder et al. (2018) Ryder, Claire L., Marenco, Franco, Brooke, Jennifer K., Estelles, Victor, Cotton, Richard, Formenti, Paola, McQuaid, James B., Price, Hannah C., Liu, Dantong, Ausset, Patrick, Rosenberg, Phil D., Taylor, Jonathan W., Choularton, Tom, Bower, Keith, Coe, Hugh, Gallagher, Martin, Crosier, Jonathan, Lloyd, Gary, Highwood, Eleanor J. & Murray, Benjamin J. 2018 Coarse-mode mineral dust size distributions, composition and optical properties from AER-D aircraft measurements over the tropical eastern Atlantic. Atmospheric Chemistry and Physics 18 (23), 17225–17257.
  • Sawford (1991) Sawford, B. L. 1991 Reynolds number effects in Lagrangian stochastic models of turbulent dispersion. Physics of Fluids A: Fluid Dynamics 3 (6), 1577–1586.
  • Sawford et al. (2003) 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. Physics of Fluids 15 (11), 3478–3489.
  • Shao (2008) Shao, Yaping 2008 Dust Transport and Deposition. In Physics and Modelling of Wind Erosion, pp. 247–301. Dordrecht: Springer Netherlands.
  • Smits et al. (2011) Smits, Alexander J., McKeon, Beverley J. & Marusic, Ivan 2011 High–Reynolds Number Wall Turbulence. Annual Review of Fluid Mechanics 43 (1), 353–375.
  • Tom & Bragg (2019) Tom, Josin & Bragg, Andrew D. 2019 Multiscale preferential sweeping of particles settling in turbulence. Journal of Fluid Mechanics 871, 244–270.
  • Van Der Does et al. (2018) Van Der Does, Michèlle, Knippertz, Peter, Zschenderlein, Philipp, Giles Harrison, R. & Stuut, Jan-Berend W. 2018 The mysterious long-range transport of giant mineral dust particles. Science Advances 4 (12), eaau2768.
  • Vickers et al. (2015) Vickers, Dean, Mahrt, Larry & Andreas, Edgar L 2015 Formulation of the Sea Surface Friction Velocity in Terms of the Mean Wind and Bulk Stability. Journal of Applied Meteorology and Climatology 54 (3), 691–703.
  • Wang et al. (2019) Wang, Guiquan, Fong, Kee Onn, Coletti, Filippo, Capecelatro, Jesse & Richter, David H. 2019 Inertial particle velocity and distribution in vertical turbulent channel flow: A numerical and experimental comparison. International Journal of Multiphase Flow 120, 103105.
  • Wang & Stock (1993) Wang, Lian-Ping & Stock, Davd E. 1993 Dispersion of Heavy Particles by Turbulent Motion. Journal of the Atmospheric Sciences 50 (13), 1897–1913.
  • Yeo et al. (2010) Yeo, K., Kim, B.-G. & Lee, C. 2010 On the near-wall characteristics of acceleration in turbulence. Journal of Fluid Mechanics 659, 405–419.
  • Yudine (1959) Yudine, M.I. 1959 Physical considerations on heavy-particle diffusion. Advances in geophysics 6, 185–191.
  • Zamansky et al. (2011) Zamansky, R., Vinkovic, I. & Gorokhovski, M. 2011 Acceleration statistics of solid particles in turbulent channel flow. Physics of Fluids 23 (11), 113304.
  • Zhang et al. (2023) Zhang, Y., Bragg, A. D. & Wang, G. 2023 Asymptotic closure model for inertial particle transport in turbulent boundary layers. Physical Review Fluids 8 (1), 014301.