1. Introduction
Solid particles in strong shear flows, such as rocket motor plumes and boundary layers, experience lift forces. Lift forces can be caused by asymmetric pressure and shear stress distributions across the particle surface, rotation induced by asymmetric viscous forces (Magnus effect) and asymmetric compression if near a wall (wall effect), as shown in figure 1. In this work, we focus on the lifting effects due to shear only. We develop a lift model for a solid spherical particle in an unbounded laminar compressible shear flow.

Figure 1. Lifting forces in a shear flow.
Rotation-induced lift is driven by asymmetric pressure distributions across the particle surface. On the advancing side of the sphere, fluid is drawn forward by the rotation, which effectively induces a higher fluid velocity relative to the free stream. The higher relative velocity induces a higher pressure on the advancing side of the sphere. The opposite occurs on the retreating side of the sphere, and the resulting pressure differential creates lift in the direction from the advancing side towards the retreating side. This effect has been studied in detail for incompressible flow since the 1950s (Howarth Reference Howarth1951; Rubinow & Keller Reference Rubinow and Keller1961). Nagata et al. (Reference Nagata, Nonomura, Takahashi, Mizuno and Fukuda2018b ) (among others) extended the study of rotation effects to compressible flow. Particles in shear flows tend to begin rotating in a direction that equalises asymmetric shear stresses across the particle surface. This creates rotation-induced lift in the direction aligned with the gradient. For example, the shear flow in figure 1 would induce rotation in the clockwise direction due to stronger shear stresses on the top half of the sphere. The clockwise rotation would then create a lifting force in the upwards direction (Nagata et al. Reference Nagata, Nonomura, Takahashi, Mizuno and Fukuda2018b ).
Wall-effect lift is driven by competing pressure and viscous effects (Cox & Hsu Reference Cox and Hsu1977; Vasseur & Cox Reference Vasseur and Cox1977; Zeng, Balachandar & Fischer Reference Zeng, Balachandar and Fischer2005). In a subsonic flow, as flow enters the restricted gap between a near-wall body and the body itself, the flow accelerates. This causes the pressure between the body and the wall to decrease compared with the pressure on the opposite side of the body, which creates a lifting force towards the wall. Conversely, as vorticity is generated at the body surface, it is advected and diffused downstream. The presence of the wall on one side of the body disrupts the symmetry of the resultant vorticity field, thereby disrupting the symmetry of the wake. For subsonic flows, this tends to create a lifting force away from the wall. The second of these two effects tends to dominate for subsonic flow, resulting in a net lift away from the wall (Takemura & Magnaudet Reference Takemura and Magnaudet2003). However, in a supersonic flow, the position of the particle relative to the sonic line and the size of the particle relative to the boundary layer thickness could change the direction of wall-effect lift, as shock reflections and other strongly compressible phenomena interact with the body and surrounding flowfield.
This work focuses on the lifting effects due to shear. In a compressible flow, both density and velocity gradients can be present in the free stream. The asymmetric free stream flow results in asymmetric pressure and shear stress distributions on the upstream portion of the sphere, which produces lift. The asymmetric free stream flow also means that asymmetric vorticity is created at the sphere surface, then asymmetrically stretched and tilted as it advects downstream, which disrupts the symmetry of the wake. This results in asymmetric forces on the downstream portion of the sphere, which also produces lift. This effect has been studied extensively for velocity gradients in incompressible flows, especially at low Reynolds number (Saffman Reference Saffman1965; McLaughlin Reference McLaughlin1991; Patankar et al. Reference Patankar, Huang, Ko and Joseph2001; Zeng et al. Reference Zeng, Najjar, Balachandar and Fischer2009). A review of many other works is provided by Shi & Rzehak (Reference Shi and Rzehak2020). At Reynolds numbers less than 10, both pressure and viscous forces tend to lift the particle in the direction aligned with the velocity gradient. However, as Reynolds number increases, the viscous and pressure contributions change sign, resulting in lift in the direction opposite the velocity gradient at Reynolds numbers greater than 100 (Shi & Rzehak Reference Shi and Rzehak2019). Eames & Hunt (Reference Eames and Hunt1997) studied the lifting effects of weak density gradients from an inviscid perspective in incompressible flow and found that a particle would tend to be lifted in the direction of the gradient. For the finite Reynolds number flows in this work, we find that density gradients create lift in the direction opposite the gradient. This discrepancy is likely due to the impacts of viscosity and compressibility not considered by Eames & Hunt (Reference Eames and Hunt1997) on the pressure, velocity and vorticity fields. Despite this significant literature on modelling lift for incompressible shear flows, there exists no general lift model for compressible shear flows.
Liu et al. (Reference Liu, Wu, Su and Kang2017) developed a general method to calculate aerodynamic forces on a body in viscous compressible flows, but their method assumed a constant free stream density and velocity, and thus did not include free stream shear. Furthermore, their method assumed that flow around the body could be described as a potential flow, and since mapping a real compressible flow back to potential flow is difficult at best, the method remains generally untested. Parmar, Haselbacher & Balachandar (Reference Parmar, Haselbacher and Balachandar2012) developed equations of motion for a sphere in inhomogeneous unsteady flow. However, their solution uses the linearised Navier–Stokes equations and therefore requires that the scales of particle motion be small relative to the fluid scales. This is usually not the case for particles moving in strong shear flows.
Thus, there is currently no general compressible shear-induced lift model for a sphere. We use direct numerical simulation and simple scaling arguments to develop a model for shear-induced lift on a sphere in an unbounded laminar compressible linear shear flow. The model can be incorporated into particle tracking simulations to model the motion of a simulated point-mass sphere through a fluid. To develop the lift model, this work will do the following.
-
(i) Show that lift coefficient scales with non-dimensional dynamic pressure gradient for steady flows via direct numerical simulation and scaling arguments (§ 3.2).
-
(ii) Show how changing wake structure affects lift coefficient (§ 3.4).
-
(iii) Develop a predictive delineation between wake structures (§ 4).
-
(iv) Formulate a model for lift force direction and lift coefficient magnitude (§ 5).
2. Computational methodology
The compressible Navier–Stokes equations (NSE) for modelling continuum fluid flow are solved via direct numerical simulation (DNS) using US3D (Nompelis, Drayna & Candler Reference Nompelis, Drayna and Candler2004; Candler et al. Reference Candler, Johnson, Nompelis, Subbareddy, Drayna and Gidzak2015), an unstructured finite-volume CFD code developed at the University of Minnesota. US3D has been previously validated against many test cases.
2.1. Governing equations
The NSE for a perfect gas are written in conservative form as
where the solution vector is
\begin{equation} U = \left [ \begin{matrix} \rho \\ \rho \boldsymbol{u} \\ E \end{matrix} \right ] \end{equation}
and the convective and viscous fluxes respectively are
\begin{equation} F^c = \left [ \begin{matrix} \rho \boldsymbol{u} \\ \rho \boldsymbol{u} \otimes \boldsymbol{u} + \boldsymbol{I}\! P \\ (E+P) \boldsymbol{u} \end{matrix} \right ]\!, \end{equation}
\begin{equation} F^v = \left [ \begin{matrix} 0 \\ \boldsymbol{\sigma } \\ \boldsymbol{u} \boldsymbol{\cdot }\boldsymbol{\sigma } - \boldsymbol{q} \end{matrix} \right ]\!, \end{equation}
where
$\rho$
is density,
$ \boldsymbol{u}$
is the velocity vector,
$P$
is pressure,
$\boldsymbol{I}$
is the identity matrix,
$E=\rho c_{v} T + \rho ( { \boldsymbol{u} \boldsymbol{\cdot }\boldsymbol{u}}/{2})$
is the total energy, where
$T$
is temperature and
$c_v$
is the specific heat at constant volume, and
$ \boldsymbol{\sigma }=2 \mu \boldsymbol{S} - ({2}/{3}) \mu \boldsymbol{I}\! \boldsymbol{S}$
is the viscous stress tensor, where
$\mu$
is the dynamic viscosity of the fluid,
$ \boldsymbol{S}$
is the strain tensor and
$ \boldsymbol{q}=\kappa ({\partial T}/{\partial \boldsymbol{x}})$
is the heat flux vector with thermal conductivity
$\kappa$
. Pressure is defined by the ideal gas equation of state
$P=\rho \textit{RT}$
, where the specific gas constant for air is
$R=287.051 \rm \: J \,kg^{-1} \,K^{-1}$
. Dynamic viscosity is calculated using Sutherland’s law (Sutherland Reference Sutherland1893) as
$\mu =1.456 \times 10^{-6} ({T^{3/2}}/{T+110.3})$
. Thermal conductivity is calculated as
$\kappa = {\mu c_{\textit{p}}}/\textit{Pr}$
, where
$\textit{Pr}=0.72$
is the Prandtl number for air and the ratio of specific heats is
$\gamma ={c_{\textit{p}}}/{c_v}=1.4$
.
2.2. Numerical scheme
For cases with steady flow, convective fluxes are calculated using the Modified-Steger–Warming (MSW) upwinding flux scheme (MacCormack & Candler Reference MacCormack and Candler1989). Viscous fluxes are calculated using weighted least-squares-error gradients with a deferred-correction approach (Kim, Makarov & Caraeni Reference Kim, Makarov and Caraeni2003). Time is advanced using the data parallel line relaxation (DPLR) time stepping scheme near walls (Wright, Candler & Bose Reference Wright, Candler and Bose1998) with the full matrix data parallel (FMDP) method (Wright, Candler & Prampolini Reference Wright, Candler and Prampolini1996) in regions of the flowfield not near a wall.
Cases with unsteady behaviour are simulated using a low-dissipation numerical method. The details of the method are beyond the scope of this work, so we provide only a brief summary here. The reader is encouraged to refer to the cited works for details. Time stepping is performed using an implicit second-order backward difference (BDF2) scheme. The convective flux
$F^c$
is calculated from a fourth-order central kinetic-energy-consistent (KEC) flux
$F^k$
(Subbareddy & Candler Reference Subbareddy and Candler2009) and a dissipative MSW upwind flux
$F^d$
(MacCormack & Candler Reference MacCormack and Candler1989) as
where the switch
$\alpha$
determines the level of dissipation in the flux calculation.
For supersonic flows,
$\alpha$
is set using a shock sensor. The sensor takes the form of a Larsson switch (Larsson & Lele Reference Larsson and Lele2009):
where
$\alpha$
is the sensor value,
$\epsilon$
is the tuning parameter,
$u_s$
is the free stream velocity on the stagnation streamline,
$\delta$
is a normalisation parameter defined as the approximate boundary layer thickness and
$\alpha _m$
is the minimum value for the sensor to maintain numerical stability (set to 0.05 in this work). Here,
$\epsilon$
is tuned so that
$\alpha$
is highest inside shocks and low everywhere else.
For subsonic flows, a numerical sponge layer is created by setting
$\alpha$
to be high near inflow/outflow boundaries and low everywhere else using the logistic function as
where
$L$
is the distance from face centroid to the nearest inflow/outflow,
$k_h$
is the sharpening parameter which determines how quickly
$\alpha$
increases, and
$\epsilon$
is the distance from the boundary below which
$\alpha$
increases to
$1$
. As for supersonic cases,
$\alpha _m$
is set to
$0.05$
.
3. Direct numerical simulation of a spherical particle in a dynamic pressure gradient
In this section, we study the effects of dynamic pressure gradient, Mach number, Reynolds number and wall temperature ratio on the lift coefficient for a spherical particle via DNS. We isolate lifting effects due to asymmetric pressure and shear stress by studying unbounded laminar compressible shear flow as shown in figure 2. The particle is held fixed (i.e. does not rotate). All flow quantities are in the particle reference frame. The particle is oriented so that free stream flow is parallel to the
$x$
-axis. Free stream velocity and density vary only along the
$y$
-axis. It is assumed that the free stream flow contains no static pressure gradient normal to the
$x$
-axis, as the presence of a static pressure gradient would cause the flow to turn and therefore be no longer aligned with the
$x$
-axis. This work uses a linear shear approximation, applicable to flows where dynamic pressure gradient is nearly constant across the particle diameter.

Figure 2. Asymmetric pressure and viscous forces in a linear shear flow.
Lift and drag coefficients
$C\!_L$
and
$C_D$
are defined as
where
$A_{\textit{ref}}=\pi ( {D^2}/{4})$
,
$F_L$
and
$F_D$
are the lift and drag force, and subscript
$s$
refers to quantities in the free stream on the stagnation streamline as shown in figure 2. In this work, drag is always in the
$x$
-direction, whereas the lift coefficient can have
$y$
- and
$z$
-components
$C\!_{L,y}$
and
$C\!_{L,z}$
. Dynamic pressure
$Q$
is defined as
Because we vary free stream velocity and density only along the
$y$
-axis, the free stream dynamic pressure gradient is one-dimensional along the
$y$
-axis and is defined as
This means that there are infinite combinations of density and velocity gradients which create a given dynamic pressure gradient. We list the particular combinations used in this work in tables and appendices when results are presented. Note that since we assume no static pressure gradient in the free stream, the ideal gas law
$P=\rho \textit{RT}$
implies that density gradients in the free stream are accompanied by opposing temperature gradients, i.e.
$ {\mathrm{d} T_\infty }/{\mathrm{d} y} = - ( {P_\infty }/{R \rho _s^2}) ({\mathrm{d} \rho _\infty }/{\mathrm{d} y})$
.
We define a non-dimensional dynamic pressure gradient to characterise the strength of the gradient relative to the particle diameter and the free stream dynamic pressure. The non-dimensional dynamic pressure gradient is
We will show that lift coefficient scales with the non-dimensional dynamic pressure gradient. Shear flows can also be characterised by the normalised velocity shear rate
$\widetilde {\boldsymbol{\nabla }u} = D( {||\boldsymbol{\nabla }u_{\infty }||}/{u_s})$
and the normalised density gradient
$\widetilde {\boldsymbol{\nabla \!}\rho } = D ({\boldsymbol{\nabla }||\rho _{\infty }||}/{\rho _s})$
. The non-dimensional dynamic pressure gradient is related to these parameters by
$\widetilde {\boldsymbol{\nabla\! }Q} = \widetilde {\boldsymbol{\nabla \!}\rho } + 2 \widetilde {\boldsymbol{\nabla }u}$
.
Reynolds number characterises how strongly viscous effects influence flow behaviour. The particle Reynolds number is
Mach number characterises how strongly compressibility effects influence flow behaviour. Mach number on the stagnation line in the free stream is
Flows have been split along typical delineations (Henderson Reference Henderson1976) into three Mach number regimes:
\begin{align} \textrm {Regime} = \left \{ \begin{array}{ll} \textrm {Subsonic} & M_s\lt 1.0, \\ \textrm {Low Supersonic} & 1.0 \leqslant M_s \leqslant 1.75, \\ \textrm {High Supersonic} & M_s\gt 1.75. \end{array} \right . \end{align}
Compressible flows can also be characterised by wall temperature ratio
$ {T_w}/{T_{aw}}$
, where
$T_w$
is the particle wall temperature and
Table 1. Grid sizes for shear-induced lift simulations.

is an estimate of the particle wall temperature if the particle wall was adiabatic. The recovery factor
$r$
for laminar flow is approximated using the Prandtl number
$\textit{Pr}$
for air as
$r \simeq \sqrt {\textit{Pr}} \simeq 0.85$
.
The combination of
$M_s$
,
$\textit{Re}_{\textit{p}}$
,
$ {T_w}/{T_{aw}}$
,
$D$
and
$T_w$
provide a complete description of each case. For contextualising the non-dimensional numbers discussed herein, we note that the diameter of the particle under study is
$D=100 \rm \, \mu m$
and that an isothermal wall is used for all cases with
$T_w = 1000 \,\rm K$
. Dimensional free stream flow conditions are then determined using the following method:
-
(i)
$T_s$
is computed using specified
$M_s$
and
$ {T_w}/{T_{aw}}$
; -
(ii)
$u_s$
is computed using
$M_s$
and the free stream speed of sound, which depends on
$T_s$
; -
(iii)
$\rho _s$
is computed from
$\textit{Re}_{\textit{p}}$
, which includes free stream viscosity
$\mu _s$
computed from
$T_s$
.
Table 2. Convergence test set for steady-flow lifting effects.


Figure 3. Computational grid for subsonic flows (heavily coarsened for visualisation purposes).

Figure 4. Lift and drag coefficients for various grid resolutions.
3.1. Computational grid
An overview of the subsonic computational grid is shown in figure 3. The grid extends 10 diameters upstream, 8 diameters in the free stream-normal direction and 15 diameters downstream from the particle centre. The supersonic grid takes the same form as the subsonic grid, but extends only 2 diameters upstream. The number of elements in each grid is listed in table 1.
Grid convergence is assessed for the subset of cases shown in table 2 using the lift and drag coefficients. Figure 4 plots the magnitude of the computed lift and drag coefficient for subsonic and supersonic flows. For subsonic flows, the medium and fine grids produce lift and drag coefficients that agree within 6 %, indicating that the medium grid is converged for subsonic flows. For supersonic flows, the coarse and medium grids produce lift and drag coefficients that agree within 2 %, indicating that the coarse grid is converged for supersonic flows. Furthermore, the measured drag coefficient agrees with at least one known drag model within 6 % in all cases except for a low supersonic case where
$M_s=1.2$
, where the agreement is still within 20 %. Some deviation in this regime is not unexpected, as drag models in this regime are less accurate as compared with the subsonic and high supersonic flow regimes (Nagata et al. Reference Nagata, Nonomura, Takahashi and Fukuda2020).
The size of the computational domain can also influence the computed lift coefficient (Shi & Rzehak Reference Shi and Rzehak2019), especially for low-Mach-number flows. To ensure sufficient domain size, run A.23 from table 2 was re-run with a domain wherein the distance between the sphere and the domain boundaries was doubled. This changed the lift coefficient by only
$3\,\%$
, indicating that the domain size is sufficient.
3.2. Lift coefficient scaling with non-dimensional dynamic pressure gradient
In this section, we study the direction of lift force and the magnitude of the lift coefficient relative to the dynamic pressure gradient. Drag force is always in the direction of the free stream flow, which by convention is in the
$x$
-direction. Lift force is normal to the drag force. This means that the direction of lift force is not predetermined, since any vector with zero
$x$
component is normal to the drag force. Therefore, we must determine the lift force direction ourselves.
The directional lift coefficient for all steady-flow cases in Appendix A.1 is shown in figure 5. Lift coefficient is plotted in the
$y$
–
$z$
plane; normal to the
$x$
-direction free stream flow. The
$y$
-axis is aligned with the direction of the dynamic pressure gradient. The
$z$
-axis is in the cross-gradient direction. As shown in figure 5, the lift coefficient is only in the negative
$y$
-direction. There is negligible lift in the
$z$
-direction. This shows that shear-induced lift force is in the direction opposite the dynamic pressure gradient, with negligible cross-gradient lift. This means that particles tend to be lifted from regions of higher dynamic pressure to regions of lower dynamic pressure.

Figure 5. Lift polar for a spherical particle in a dynamic pressure gradient.
Next, we show that the lift coefficient is a function of the non-dimensional dynamic pressure gradient using the subset of cases in Appendix A.2. Figure 6 plots the lift coefficient for a non-dimensional dynamic pressure gradient
$\widetilde {\boldsymbol{\nabla\! }Q} = 10^{-3}$
created from only a density gradient (
$\boldsymbol{\nabla}\kern-1pt \rho _{\infty }$
), only a velocity gradient (
$\boldsymbol{\nabla }u_{\infty }$
), and both density and velocity gradients (
$\boldsymbol{\nabla}\kern-1pt \rho _{\infty } + \boldsymbol{\nabla }u_{\infty }$
). We have shown already that the lift coefficient is only in the negative
$y$
-direction, so we plot the negative of the
$y$
-direction lift coefficient. The total lift coefficient
$C_{L,y}$
is plotted, as well as its pressure and viscous components
$C_{L,y,P}$
and
$C_{L,y,\tau }$
. We see that pressure lift dominates viscous lift at
$\textit{Re}_{\textit{p}} \geqslant 200$
and pressure lift is similar in magnitude to viscous lift at
$\textit{Re}_{\textit{p}}=100$
. This is consistent with the observations of Shi & Rzehak (Reference Shi and Rzehak2019). When the dynamic pressure gradient is created with only a velocity gradient, the measured lift coefficient at low Mach number (
$M_s=0.3,{} \textit{Re}_{\textit{p}}=100$
) is within the variation of the DNS used to develop the shear lift correlation for incompressible flow proposed by Shi & Rzehak (Reference Shi and Rzehak2019).
As shown in figure 6, at
$M_s \geqslant 0.8$
,
$\textit{Re}_{\textit{p}} \geqslant 200$
, both the total lift coefficient and its pressure and viscous components do not change regardless of whether the dynamic pressure gradient was created from a density and/or velocity gradient. At
$M_s \leqslant 0.6$
,
$\textit{Re}_{\textit{p}} \leqslant 100$
, the lift coefficient starts to vary based on how the dynamic pressure gradient was created, though the variation is small. This shows that lift is dependent on the dynamic pressure gradient, not on the density or velocity gradient.
The physical mechanism behind the change in behaviour at lower Mach and Reynolds number is nuanced and requires careful examination. As Reynolds number decreases, viscous forces in the flowfield increase in strength relative to inertial (pressure) forces. This affects the behaviour of the entire flow, not just the boundary layer – so the changing Reynolds number affects both pressure and viscous lift at the particle surface. Recall that in this study, flows with a density gradient come with an opposing temperature gradient such that zero static pressure gradient is maintained in the free stream. A temperature gradient in the free stream indicates a viscosity gradient in the free stream (Sutherland Reference Sutherland1893). This means that for runs with a density gradient, there is an additional viscosity gradient affecting the behaviour of the flow. As Reynolds number decreases and viscous forces increase in strength relative to inertial forces, this ‘second’ gradient due to viscosity contributes further to asymmetries in the flowfield, as compared with runs with only a velocity gradient. This results in stronger lifting effects for particles in density gradients. In figure 6 (for which all runs have the same non-dimensional dynamic pressure gradient), this effect is reflected in stronger lift at
$\textit{Re}_{\textit{p}}=100$
when the dynamic pressure gradient is created with a density gradient (red) than when the dynamic pressure gradient is created with a velocity gradient (blue); wherein runs with both density and velocity gradients (green) fall between runs with only a density or velocity gradient.
Now that we have shown lift coefficient depends on the dynamic pressure gradient, we must determine the scaling of lift coefficient magnitude with dynamic pressure gradient. We use the subset of cases in Appendix A.2. In figure 7, we plot lift coefficient versus non-dimensional dynamic pressure gradient; wherein the lift coefficient is normalised by its value at
$\widetilde {\boldsymbol{\nabla\! }Q} = 10^{-3}$
. As shown in figure 7, lift coefficient scales linearly with non-dimensional dynamic pressure gradient. For all Mach and Reynolds numbers except at
$M_s=0.8$
,
$\textit{Re}_{\textit{p}}=200$
, if non-dimensional dynamic pressure gradient doubles, lift coefficient doubles. This relationship between
$\widetilde {\boldsymbol{\nabla\! }Q}$
and
$C\!_L$
indicates that as
$\widetilde {\boldsymbol{\nabla\! }Q}$
approaches zero,
$C\!_L$
also approaches zero. This makes sense, because if
$\widetilde {\boldsymbol{\nabla\! }Q}=0$
, there is no shear in the flow, so we would expect there to be no shear-induced lift.

Figure 6. Lift coefficient in density and/or velocity gradients at
$\widetilde {\boldsymbol{\nabla\! }Q} = 10^{-3}$
.

Figure 7. Scaling of lift coefficient with
$\widetilde {\boldsymbol{\nabla\! }Q}$
.
Special behaviour is observed for the case at
$M_s=0.8$
,
$\textit{Re}_{\textit{p}}=200$
. This is because this combination of Mach and Reynolds number results in a planar-symmetric wake, whereas the other cases in figure 7 have axisymmetric wakes. For planar-symmetric wakes, some lift is induced along the symmetry plane (Nagata et al. Reference Nagata, Nonomura, Takahashi and Fukuda2020) due to counter-rotating vortices in the wake, even when
$\widetilde {\boldsymbol{\nabla\! }Q}=0$
. The symmetry plane would tend to be randomly oriented when
$\widetilde {\boldsymbol{\nabla\! }Q}=0$
. However, in the range of
$\widetilde {\boldsymbol{\nabla\! }Q}$
considered in this work, the symmetry plane tends to align with the gradient, amplifying shear-induced lift (see § 3.4). This means that for planar-symmetric wakes, as
$\widetilde {\boldsymbol{\nabla\! }Q} \rightarrow 0$
, nonlinear lifting effects are induced as the orientation of the symmetry plane changes from gradient-aligned to randomly oriented. However, figure 7 shows that even for planar-symmetric wakes,
$C\!_L$
is linear with
$\widetilde {\boldsymbol{\nabla\! }Q}$
in the range of
$\widetilde {\boldsymbol{\nabla\! }Q}$
studied here.
This assessment of lift direction and magnitude is consistent with simple scaling arguments. Lift force is by definition normal to the free stream flow, so we ignore forces in the
$\hat {i}$
direction as they would be included in drag force. We align the dynamic pressure gradient with the
$y$
-axis, so in steady flow, there are no asymmetric forces with a component in the
$z$
-direction. Therefore, lift force is one-dimensional along the
$y$
-axis:
where
$P_w$
is the pressure force at the wall and
$\tau _{w,y}$
is shear stress at the wall in the
$y$
-direction. The angles
$\theta$
and
$\phi$
are defined in figure 2.
Consider first the lift force due to pressure. There exists no analytic model for surface pressure in a general compressible flow. However, Newtonian aerodynamics (Heybey Reference Heybey1964) provides a crude estimate of surface pressure at high Mach number. Newtonian aerodynamics neglects pressure forces on the downstream half of the sphere as small and approximates pressure coefficient on the upstream half of the sphere as
Using the definition of pressure coefficient and the definition of dynamic pressure, we can approximate the wall pressure as
Recall we have aligned our axes such that the free stream dynamic pressure gradient is parallel to and in the direction of the
$y$
-axis, so
$\boldsymbol{\nabla \!}Q_{\infty }$
only has one dimension and can be represented by the scalar
$||\boldsymbol{\nabla \!}Q_{\infty }||$
. Therefore, the free stream dynamic pressure at a given
$y$
-coordinate is
where
$D$
is the diameter of the spherical particle. Recall also we assume no static pressure gradient in the free stream, so
$P_{\infty } = P_s$
everywhere in the free stream. Thus, wall pressure scales as
By substituting into the first term in (3.10) and evaluating the integral, we see that
The negative sign on the right-hand side in (3.15) means that lifting force is in the direction opposite the dynamic pressure gradient – i.e. particles will tend to be lifted down the gradient, from regions of higher dynamic pressure to regions of lower dynamic pressure.
Applying the definition of lift coefficient, using the definition of non-dimensional dynamic pressure gradient from (3.5), taking the reference area for a sphere to be
$A_{\textit{ref}}=\pi R^2=\pi ( {D^2}/{4})$
and removing constant factors, we see that
which shows that lift due to pressure forces scales with non-dimensional dynamic pressure gradient and is in the direction opposite the gradient. This is consistent with the DNS results.
While Newtonian aerodynamics can provide some limited insight into pressure forces on the upstream portion of the sphere, it does not provide any insight into the viscous forces on the upstream portion of the sphere, nor does it provide insight into the evolution of the vorticity field or forces on the downstream portion of the sphere. We explore in § 3.4 how wake structure influences lift coefficient, but no analytic map back to the vorticity field or viscous stresses exists for a general compressible flow.
3.3. Effects of Mach Number, Reynolds number and wall temperature ratio
In this section, we study effects of changing Reynolds number, Mach number and wall temperature ratio by varying each of these parameters independently. We consider a subsonic baseline case at
$\textit{Re}_{\textit{p}}=100, \: M_s=0.3$
and a supersonic baseline case at
$\textit{Re}_{\textit{p}}=1000, \: M_s=2.0$
. Both baseline cases use
$ {T_w}/{T_{aw}}=0.5$
. Case details are listed in Appendix A.3. Figure 8 shows how increasing each parameter by a factor of 2 affects both pressure and viscous lift by plotting the lift coefficient against each parameter, normalised by their value for the baseline case. As shown in figure 8, Reynolds number has the largest effect on both pressure and viscous lift, especially at low Mach number. This is because changing the Reynolds number results in a change in wake structure (Nagata et al. Reference Nagata, Nonomura, Takahashi and Fukuda2020). Discussion on this matter can be found in § 3.4. Mach number has a smaller but non-negligible effect, primarily on the pressure lift coefficient at lower Reynolds number. Effects of wall temperature ratio are even smaller, especially for pressure lift, which drives the overall lift coefficient in an unbounded compressible flow.
These conclusions are consistent with parameters that drive drag forces. Drag coefficient on a spherical particle in continuum flow is a function of Mach number and Reynolds number, with wall temperature effects becoming important only for rarefied flows (Singh et al. Reference Singh, Kroells, Li, Ching, Ihme, Hogan and Schwartzentruber2020; Loth et al. Reference Loth, Despit, Jeong, Nagata and Nonomura2021). Thus, we would also expect Mach and Reynolds number to drive lift coefficient for the continuum flow studied in this work.

Figure 8. Effects of
$\textit{Re}_{\textit{p}}$
,
$M_s$
and
$ {T_w}/{T_{aw}}$
on pressure lift (top row) and viscous lift (bottom row).
3.4. Effects of wake structure on lift coefficient
In this section, we examine the effect of wake structure on lift coefficient. Both steady and unsteady wake structures significantly affect lift.
Unsteady wakes can develop cross-gradient asymmetries. When the cross-gradient asymmetric behaviour is stronger than the dynamic pressure gradient, cross-gradient lifting forces in directions not readily predictable are induced. While this behaviour is not intuitive, it is well documented. Williams & Smits (Reference Williams and Smits2024) compiled dozens of observations regarding asymmetric behaviour in nominally symmetric flows. They concluded that asymmetric behaviours can and do exist, mainly because they evolve on time scales much longer than are currently possible to observe in most simulations or experiments.

Figure 9. Wakes at
$M_s=0.3$
and
$\widetilde {\boldsymbol{\nabla\! }Q}=10^{-3}$
: top, steady axisymmetric (SA); middle, steady planar-symmetric (SP); bottom, unsteady hairpin with azimuthal oscillations (HaWAO).

Figure 10. Effect of wake structure on lift coefficient for seven cases at
$\widetilde {\boldsymbol{\nabla\! }Q}=10^{-3}$
,
$\textit{Re}_{\textit{p}}$
from
$100$
to
$2000$
,
$M_s$
from
$0.3$
to
$2.0$
.
Steady-wake structure also influences lift. We observe that the symmetry plane for planar-symmetric wakes tends to align with the direction of the dynamic pressure gradient. This creates gradient-aligned viscous forces on the downstream half of the sphere which would otherwise be axisymmetrically balanced. Therefore, at low Mach number when viscous forces have a significant influence on the lift coefficient, planar-symmetric wakes result in higher-magnitude lift coefficients than for axisymmetric wakes.
For example, consider the wake structures shown in figure 9. At
$\textit{Re}_{\textit{p}}=100$
, the wake is steady axisymmetric (SA). (Wake classification nomenclature from Nagata et al. Reference Nagata, Nonomura, Takahashi and Fukuda2020.) At
$\textit{Re}_{\textit{p}}=200$
, the wake is steady planar-symmetric (SP) across the
$x$
–
$y$
plane, which is aligned with the dynamic pressure gradient. At
$\textit{Re}_{\textit{p}}=600$
, the wake becomes an unsteady hairpin wake with azimuthal oscillations (HaWAO). All of these structures match those predicted by Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020).
In figure 10, we plot the lift polar for three sets of cases. The first set consists of the same three cases shown in figure 9 at
$M_s=0.3$
. The other sets are pairs of cases at
$M_s=0.6$
and
$M_s=2.0$
. One member of each pair is an SA wake and the other is an SP wake. Case details are listed in Appendix A.4. All six steady-flow cases show lift in the expected negative
$y$
-direction with negligible cross-gradient (
$z$
-direction) lift. However, at
$M_s=0.3$
and
$M_s=0.6$
, the SP case has a lift coefficient nearly an order of magnitude higher than the SA case. This trend holds at
$M_s=2$
, but is much less pronounced, since lift coefficient at high Mach number is dominated by pressure forces on the upstream half of the sphere and is therefore less affected by wake behaviour. The unsteady case demonstrates significant non-zero-average cross-gradient lift, which is an asymmetric behaviour across the
$x$
–
$y$
plane in a flow that is nominally symmetric across that plane. The unsteady case also demonstrates lift in the positive
$y$
-direction – opposite the direction predicted in § 3.2. This shows that unsteady wakes drive lift force in directions not readily predicted. The changing lift coefficient with changing wake structure means that predicting lift requires predicting wake structure.
4. Delineation between wake types
We show in § 3.4 that unsteady effects induce lift force in unexpected directions. We also show that even for steady flows, lift is influenced by the type of wake. Therefore, predicting lift coefficient requires the ability to predict the wake structure.
Nagata et al. (Reference Nagata, Nonomura, Takahashi, Mizuno and Fukuda2016, Reference Nagata, Nonomura, Takahashi and Fukuda2020), Riahi et al. (Reference Riahi, Medli, Favier, Serre and Goncalves2018) and Sansica et al. (Reference Sansica, Robinet, Alizard and Goncalves2018) built on the work of Taneda (Reference Taneda1956), Magarvey & Biship (Reference Magarvey and Biship1961), Sakamoto & Haniu (Reference Sakamoto and Haniu1990), Johnson & Patel (Reference Johnson and Patel1999) and others to characterise wake behaviour of spheres at various Reynolds numbers. Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020) in particular extended the characterisation to compressible flows, demonstrated the dependence of wake behaviour on both Mach and Reynolds number, and showed delineations between various types of steady behaviours (including fully attached, steady axisymmetric and steady planar-symmetric wakes) and unsteady behaviours (including various types of hairpin and helical wakes) at Reynolds numbers up to 1000.
In this section, we extend the prior work by developing an approximate delineation between steady axisymmetric, steady planar-symmetric and unsteady wakes using a set of direct numerical simulations at Reynolds numbers up to 20 000. The complete list of cases can be found in Appendix B.
4.1. Computational grid modifications
For subsonic and low-supersonic flows, the computational grid take the same form as in § 3.1. However, for high-supersonic flows, the use of the low-dissipation numerical scheme with a shock sensor described in § 2.2 requires grid-shock alignment to minimise flickering of the shock between cells, which introduces numerical noise into the simulation (Knutson, Thome & Candler Reference Knutson, Thome and Candler2021). An overview of the grid is shown in figure 11. As shown in figure 12, the grid is approximately aligned with the shock using shock shape equations from Bedarev, Fedorov & Fomin (Reference Bedarev, Fedorov and Fomin2012):
\begin{equation} \begin{array}{ll} R = \dfrac {D}{2} \left ( 1 + \dfrac {1}{M-1} \right )\!, \\[2pt]\\ a = \textit{RM} ( M-1 ),\\[2pt] \\ b = R \sqrt {M ( M-1)}, \\[2pt]\\ \varDelta = \dfrac {D}{2} \left ( 1 + \dfrac {M}{2 ( M^2 -1) } \right )\!. \end{array} \end{equation}

Figure 11. Unsteady supersonic grid overview,
$M_s=2.4$
(heavily coarsened for visualisation purposes).

Figure 12. Sample shock-aligned grid with contours of the shock sensor value
$\alpha$
,
$M_s=2.4$
.
The number of elements in each grid is listed in table 3.
Table 3. Grid sizes for unsteady/steady simulations.

Grid convergence for the subset of cases shown in table 4 is assessed using the drag coefficient over a window of more than five unsteady periods. Mean drag coefficient is compared with existing models in figure 13. The coarse, medium and fine grids produce drag coefficients which match each other within 2 % and match at least one known drag model within 7 %. Frequency content of the drag coefficient time-history, generated using a fast Fourier transform (FFT), is shown in figure 14. The medium grid and fine grid produce frequency spectra with a maximum deviation of less than 5 %. The close agreement in both mean drag and unsteady drag frequency content indicates that the medium grid is converged.
Table 4. Convergence test set for delineating steady versus unsteady behaviour.


Figure 13. Mean drag force versus existing drag models.

Figure 14. Drag coefficient frequency spectrum in the Strouhal number domain,
$St=f( {D}/{u_s})$
.
Furthermore, wake structures observed in this work at
$\textit{Re}_{\textit{p}} \leqslant 1000$
match those predicted by Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020). Figure 15 shows wake structures over a range of Mach and Reynolds numbers. At
$\textit{Re}_{\textit{p}}=500$
and
$M_s=0.8$
, the wake demonstrates hairpin structures with azimuthal oscillation (HaWAO). At
$\textit{Re}_{\textit{p}}=1000$
and
$M_s=0.8$
, the wake demonstrates helical structures (HeW). At
$\textit{Re}_{\textit{p}}=1000$
and
$M_s=1.2$
, we observe one of the types of pure hairpin wakes (HaW2). The wake is steady and axisymmetric (SA) at
$\textit{Re}_{\textit{p}}=1000$
and
$M_s=2.0$
.

Figure 15. Sample wake structures in steady and unsteady flows.
4.2. Results
We use data from the prior work and the DNS to develop approximate delineations among steady axisymmetric, steady planar-symmetric and unsteady wakes. Delineations among wake types are represented using critical Mach numbers, each of which is a function of Reynolds number. A critical Mach number
$M_{\textit{ts}}$
which delineates steady and unsteady behaviours is

where if
$M_s \gt M_{\textit{ts}}$
, the wake will likely assume a steady behaviour and will likely be unsteady otherwise. Here,
$\textit{Re}_{\textit{ts}}=1395$
is the Reynolds number at which the delineation switches from the linear Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020) curves to the proposed logarithmic curve; found by value- and slope-matching the logarithmic curve to the linear curve to eliminate discontinuity (within rounding error) between the two. The delineation is shown graphically in figure 16, along with the continuum flow limit defined by the Knudsen number
$\textit{Kn} \leqslant 0.01$
and related to the Mach and Reynolds numbers by
\begin{equation} M_s \leqslant 0.01 \textit{Re}_{\textit{p}} \sqrt {\frac {2}{\pi \gamma }}. \end{equation}
Note that for consistency with the prior work, we develop this delineation with
$\widetilde {\boldsymbol{\nabla\! }Q} = 0$
. However, as shown by the cases at
$M_s=2, \: \textit{Re}_{\textit{p}}=1000$
,
$M_s=2, \: \textit{Re}_{\textit{p}}=2000$
and
$M_s=4,{} \: \textit{Re}_{\textit{p}}=10\,000$
, the wake structure does not change when a gradient is introduced.

Figure 16. Approximate delineation between wake behaviours in continuum flow.
A critical Mach number
$M_{\textit{tp}}$
which delineates steady axisymmetric (SA) and steady planar-symmetric (SP) wakes is
\begin{align} M_{\textit{tp}} \simeq \left \{ \begin{array}{ll} 0, & \textit{Re}_{\textit{p}} \lt 175, \\ 0.0115\textit{Re}_{\textit{p}}-1.42, & 175 \leqslant \textit{Re}_{\textit{p}} \lt 210, \\ \big ( 0.72\textit{Re}_{\textit{p}}+846\big ) \times 10^{-3}, & 210 \leqslant \textit{Re}_{\textit{p}} \lt \textit{Re}_{\textit{tp}},\\ 0.78\ln {\textit{Re}_{\textit{p}}}-3.8243, & \textit{Re}_{\textit{p}} \geqslant \textit{Re}_{\textit{tp}}, \end{array} \right . \end{align}
where if
$M_s \leqslant M_{\textit{tp}}$
and
$M_s \gt M_{\textit{ts}}$
, the wake will likely be steady planar-symmetric, and steady axisymmetric if
$M_s \gt M_{\textit{tp}}$
. Here,
$\textit{Re}_{\textit{tp}}=1083$
is calculated by value- and slope-matching the linear curve to the logarithmic curve. This delineation is also shown in figure 16. Note that we observe a slight increase in
$M_{\textit{tp}}$
at low Reynolds number as compared with the delineation drawn by Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020). This is not unexpected, as some variance around this delineation also occurs in the many works cited by Nagata et al. (Reference Nagata, Nonomura, Takahashi and Fukuda2020).
It is important to acknowledge that the proposed delineations should be treated as approximations only. Various complex factors including noise in the free stream (for experimental studies), properties of the numerical scheme (for computational studies), and other considerations can have non-negligible impacts on the Mach and Reynolds number at which wake behaviour changes. This is reflected in inconsistencies around delineations between wake types in the prior work. Note also that increasing wall temperature ratio tends to delay the transition from steady to unsteady wake behaviour as Reynolds number increases for a given Mach number (Nagata et al. Reference Nagata, Nonomura, Takahashi, Mizuno and Fukuda2018a
). Since this work uses a relatively low wall temperature ratio (
${T_w}/{T_{aw}}=0.5$
), the delineations developed herein would be considered conservative for the purposes of a steady-flow lift model, because wakes are more likely to be steady for a given Mach and Reynolds number for higher wall temperature ratios.
5. Proposed lift model
In this section, we develop a lift model for a spherical particle in compressible shear flow which can be used in particle tracking simulations to model particle motion in a general flow. We show in § 3.2 that the lift force is in the direction opposite the dynamic pressure gradient and that the lift coefficient scales with non-dimensional dynamic pressure gradient. We show in § 3.3 that Mach number and Reynolds number affect the lift coefficient magnitude. We show in § 3.4 that the lift coefficient is affected by wake structure and develop a predictive delineation between wake types in § 4. We now make use of this information to formulate our lift model. Formulating the model requires three steps.
-
(i) Estimate if the flow is steady or unsteady.
-
(ii) Compute the lift force direction.
-
(iii) Compute the lift coefficient magnitude.
In addition, the lift model was tested against five blind validation cases.
First, we must assess whether or not the flow is steady. This is accomplished using (4.2). If the flow is unsteady, lift force direction is not readily predictable as shown in § 3.4. While there may be some method to identify bias in the unsteady lift force direction, we do not pursue those details here. As such, for unsteady flows, we assume the lift force is zero. If the flow is steady, we proceed by computing lift force direction and lift coefficient magnitude.
Next, we compute the lift force direction. We show in § 3.2 that lift force is in the direction opposite the dynamic pressure gradient. In this work, the dynamic pressure gradient is always aligned to the
$y$
-axis and is normal to the free stream flow, so the lift force direction is implied. However, when using this lift model in a general flow, the direction of the dynamic pressure gradient is not predetermined. Thus, to compute the lift force direction, we must extract the component of the dynamic pressure gradient normal to the free stream flow as
where
$u_s$
is the fluid velocity on the stagnation streamline in the particle reference frame. Note that the component of the dynamic pressure gradient aligned with the free stream flow is neglected as this would contribute only to drag force. Therefore, the effective non-dimensional dynamic pressure gradient is
We show in § 3.2 that for steady flows, lift force is in the direction opposite the dynamic pressure gradient. Therefore, a unit vector in the lift force direction is
Finally, we compute the lift coefficient. Note that lift coefficients in this section are relative to the lift force direction in (5.3), so lift coefficient is always positive. We split the lift coefficient into pressure and viscous components. Each component is estimated separately. Then, the total lift coefficient is computed by summing the two components. Pressure and viscous lift coefficients are fit to the DNS data using the following process.
-
(i) A limit for steady axisymmetric wakes at high Mach number is developed using scaling arguments.
-
(ii) A correction to the high-Mach-number limit is developed using the DNS data.
-
(iii) Well-behaved functions are selected to represent subsonic, low supersonic and steady planar-symmetric wake lifting effects relative to the high-Mach axisymmetric wake limit. Well-behaved functions limit to appropriate values as their independent variable(s) reach zero or infinity, and represent the DNS data in between. This process is similar to that used to develop modern drag models (Singh et al. Reference Singh, Kroells, Li, Ching, Ihme, Hogan and Schwartzentruber2020; Loth et al. Reference Loth, Despit, Jeong, Nagata and Nonomura2021).
-
(iv) Constant parameters in those well-behaved functions are fit to the DNS data.
We fit parameters in the lift model using a sweep of direct numerical simulations across
$\widetilde {\boldsymbol{\nabla \!}Q_{\perp }} = [ 5,50 ] \times 10^{-3}$
,
$M_s= [ 0.3, \: 6 ]$
and
$\textit{Re}_{\textit{p}}= [ 100, \: 20\,000 ]$
. The complete list of cases can be found in Appendix A.1.
5.1. Pressure lift
First, we develop a high-Mach limit for pressure lift coefficient. The scaling arguments in § 3.2 show that in the high-Mach limit of Newtonian aerodynamics, pressure lift coefficient scales with non-dimensional dynamic pressure gradient. This would indicate that
However, in supersonic flows, a bow shock will form in front of the particle. The shock will change the effective dynamic pressure gradient seen by the particle. We must therefore account for the presence of a shock in our dynamic pressure gradient calculation. Dynamic pressure can be written as
For a perfect gas with constant
$\gamma$
and
$R$
, the change in dynamic pressure across a shock is then
where subscript
$1$
indicates pre-shock conditions and subscript
$2$
indicates post-shock conditions. Approximating this change using the Rankine–Hugoniot normal shock relations for a perfect gas results in the dynamic pressure jump across the shock:
Conveniently, this is the same relation as the inverse density ratio
$\epsilon ={\rho _{s1}}/{\rho _{s2}}$
, which must be the case because momentum is conserved across a shock. Recall that we consider only dynamic pressure gradients normal to the free stream flow. Since a normal shock is also normal to the free stream flow, the gradient will change only in magnitude (not in direction) across the shock. So, we can take the free stream-normal gradient of
$Q_1$
and
$Q_2$
. Recall also that we aligned the
$y$
-axis of our particle to the free stream-normal dynamic pressure gradient, so
$\boldsymbol{\nabla \!}Q_{\perp }$
only has one dimension and can be represented by the scalar
$||\boldsymbol{\nabla \!}Q_{\perp }||$
. Therefore,
Taking
$M_1$
to be the free stream Mach number on the stagnation line
$M_s$
, we can define a parameter
$\sigma$
which characterises the change in dynamic pressure gradient across the shock:
\begin{equation} \sigma = \left \{ \begin{array}{ll} \dfrac {\left ( \gamma - 1 \right ) M_{s}^2 + 2}{\left ( \gamma + 1 \right ) M_{s}^2}, & M \geqslant 1.0, \\ 1.0, & \mathrm{otherwise}. \end{array} \right . \end{equation}
A high-Mach limit for pressure lift coefficient is then
Next, a high-Mach correction parameter
$C_{\textit{MH}}$
is fit to all steady axisymmetric cases where
$M_s=6$
, so that the lift coefficient due to pressure forces is
The high-Mach correction parameter is found to be
which is consistent with the high-Mach drag coefficient for a sphere
$C_D \sim 0.9$
(Hornung, Schramm & Hannemann Reference Hornung, Schramm and Hannemann2019), despite the theoretical Newtonian limit being
$C_D=1.0$
.
Next, we find well-behaved functions that represent subsonic, low-supersonic and planar-symmetric wake lifting effects relative to the high-Mach axisymmetric limit. A subsonic representation should asymptote to a fixed limit as
$M_s \rightarrow 0$
and approach 1 as
$M_s \rightarrow \infty$
. A function that fits this form and the DNS is
This leaves us with a need for a low supersonic representation that limits to 1 at both high and low Mach number, but provides a suitable adjustment for lift coefficient near
$1 \lt M_s \lt 1.75$
. A function that fits this form and the DNS is
which is effectively an inverted Gaussian centred at
$M_s=1.25$
.
Finally, we need a representation for planar-symmetric wakes. We observe in the DNS data that this correction must be stronger at lower
$\widetilde {\boldsymbol{\nabla \!}Q_\perp }$
and weaker at higher
$M_s$
. Furthermore, as
$\widetilde {\boldsymbol{\nabla \!}Q_\perp }$
decreases, this function must grow more slowly than
$\widetilde {\boldsymbol{\nabla \!}Q_\perp }$
decreases, so lift coefficient limits to a fixed value when combined with the high-Mach scaling term. This function must also limit to fixed values as
$M_s \rightarrow 0$
and
$M_s \rightarrow \infty$
. A function that fits this form and the DNS is
We must also account for the step change in lift coefficient when the wake behaviour switches from steady axisymmetric to steady planar-symmetric. We can use a combination of hyperbolic tangent functions and the delineation between wake types
$M_{\textit{tp}}$
from (4.4) to define
which can be used to smoothly switch between
$C_{\textit{WP}}$
when
$M_s \lt M_{\textit{tp}}$
and 1 when
$M_s \gt M_{\textit{tp}}$
. This leaves us with the steady planar-symmetric wake term
The final lift coefficient due to pressure is then
The model is then fitted to the DNS data. Fit parameters are listed in table 5. The pressure lift coefficient model is shown in figure 17 and generally matches the DNS data. Small discrepancies exist for certain cases because lifting effects near changing wake behaviour are often erratic and therefore difficult to quantify with simple well-behaved analytic expressions.
Table 5. Coefficients for pressure lift model.


Figure 17. Fitted pressure lift coefficient (
$\widetilde {\textit{Re}_{\textit{p}}}$
defined in Appendix C).
5.2. Viscous lift
Similar to pressure lift, we first develop a high-Mach limit for viscous lift coefficient. Since viscous lift and pressure lift are both functions of non-dimensional dynamic pressure gradient as shown in § 3.2, we start with the high-Mach term from (5.11). However, we need a correction term to relate the high-Mach pressure lift limit to the high-Mach viscous lift limit. We observe in the DNS data that viscous lift coefficient scales inversely with
$\widetilde {\textit{Re}_{\textit{p}}}$
and directly with
$M_s$
at high Mach number. A function that meets these requirements and represents the DNS data is
where
$\widetilde {\textit{Re}_{\textit{p}}}$
is a transformed compressible Reynolds number derived by Singh et al. (Reference Singh, Kroells, Li, Ching, Ihme, Hogan and Schwartzentruber2020). The transformed Reynolds number is based on the Howarth (Reference Howarth1948) and Stewartson (Reference Stewartson1949) compressible boundary layer transformation and accounts for the changing relationship between compressibility and viscous effects as the Mach number increases. A summary of the transformed Reynolds number can be found in Appendix C. The addition of
$0.1$
inside the
$\log (M_s)$
term limits the function as
$M_s \rightarrow 0$
. The high-Mach limit for viscous lift coefficient is then
Next, we incorporate a subsonic representation. The term should asymptote to a fixed value as
$M_s \rightarrow 0$
and approach 1 as
$M_s \rightarrow \infty$
. A function that fits this form and the DNS is
We observe in the DNS data that no additional low supersonic term is needed for viscous lift. Finally, we incorporate the planar-symmetric wake representation, which takes the same form as for pressure lift,
and is then switched on or off using (5.16) to give
Combining these terms leaves us with the viscous lift coefficient
The model is then fitted to the DNS data. Fit parameters are listed in table 6. The viscous lift coefficient model is shown in figure 18 and generally matches the DNS data.
5.3. Total lift
Finally, we compute the total lift coefficient as
Table 6. Coefficients for viscous lift model.


Figure 18. Fitted viscous lift coefficient (
$\widetilde {\textit{Re}_{\textit{p}}}$
defined in Appendix C).
The fitted model was tested against five blind validation cases excluded from the fitting process. Case details are listed in Appendix A.5. As shown in figure 19, the fitted model matches the DNS within 11 % for all blind validation cases except at
$M_s=0.6$
, where the fitted model still matches within 20 %. Note that the case at
$M_s=0.8$
has a steady planar-symmetric wake, while other cases have a steady axisymmetric wake, indicating that the model does predict effects of changing wake behaviour.

Figure 19. Blind validation of total lift coefficient,
$\widetilde {\boldsymbol{\nabla\! }Q}=15\times 10^{-3}$
.
5.4. Model limitations
The present work uses the Navier–Stokes equations to model fluid dynamics, which means that this study is limited to continuum flow. There exists little to no data in the published literature regarding particles in compressible shear flows, either in continuum or rarefied flows. Recall also that the present work only studies linear shear flow. Therefore, the proposed model comes with several limitations.
-
(i) Lifting effects at very low Reynolds or Mach number are not studied. As a result, there are no model terms that collapse to incompressible and/or Stokes flow limits.
-
(ii) The proposed model includes no term(s) for rarefaction effects.
-
(iii) The proposed model does not address lifting effects induced by unsteady wakes.
-
(iv) The proposed model applies only to linear or near-linear shear flows, i.e. flows where
$\parallel\! ( {D^2}/{Q_{s}}){\nabla} ^2 Q_{\infty} \! \parallel \ll 1$
.
It is expected that as more data become available to address these limitations, the proposed lift model will and should evolve, just as spherical drag models have evolved with data availability.
6. Conclusion
This work studies the lifting effects for a spherical particle in unbounded compressible laminar shear flows. We show in § 3.2 via direct numerical simulation that lift force for steady flows is in the direction opposite the dynamic pressure gradient and that lift coefficient scales with non-dimensional dynamic pressure gradient. This conclusion is consistent with simple scaling arguments. This means that particles tend to be lifted from regions of higher dynamic pressure to regions of lower dynamic pressure. For example, particles originating from a viscous wall would tend to be trapped inside the boundary layer, and particles near the edges of a rocket motor plume would tend to be lifted outside the plume entirely.
We show in § 3.4 that lift coefficient is affected by the wake structure. Steady planar-symmetric wakes tend to be symmetric about a plane aligned with the dynamic pressure gradient. As a result, steady planar-symmetric wakes demonstrate larger lift coefficients than steady axisymmetric wakes. Furthermore, unsteady wakes produce lift in directions not readily predictable. This includes strong cross-gradient lift despite there being no cross-gradient asymmetry in the free stream flow. This means that predicting lift requires the ability to predict wake structure. Therefore, we develop a predictive delineation between wake types in § 4, which can be used to estimate whether the wake will be steady axisymmetric, steady planar-symmetric or unsteady.
Finally, in § 5, we develop a model for lift force direction and lift coefficient magnitude for a sphere in a laminar compressible linear shear flow. The lift model includes terms for non-dimensional dynamic pressure gradient, Mach number, Reynolds number and wake structure. We predict pressure and viscous lift independently, then sum the two components to compute total lift coefficient. The model was tested against five blind validation cases and was shown to match the DNS data. The lift model can be used in conjunction with a drag model to simulate particle motion in compressible shear flow.
Acknowledgements
The authors thank Dr Pramod Subbareddy at the University of Minnesota for contributing expertise on low-dissipation numerical methods and means to characterise flowfields; Carter Vu at the University of Minnesota for his knowledge of curve fits and unsteady flows; and Dr Tyler Hendrickson at the University of Minnesota for his thoughtful critique and perspective.
Funding
E.F. acknowledges funding from Department of Defense (DoD) through the National Defense Science & Engineering Graduate (NDSEG) Fellowship Program. G.C. acknowledges funding from ONR MURI grant N00014-20-1-2682.
Declaration of interests
The authors report no conflict of interest.
Appendix A. Runsets with dynamic pressure gradients
A.1. Runset to estimate lift coefficient

A.2. Runset to study lift coefficient scaling with non-dimensional dynamic pressure gradient
| Run ID |
|
|
|
|
|
|
|
|
|
|
| (
|
(
|
(
|
(
|
(
|
(
|
(
|
||||
| A.22 | 10 | 100 | 0.3 | 0.5 | 22.96 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.23 | 10 | 100 | 0.3 | 0.5 | 11.48 | 6673.11 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.24 | 10 | 100 | 0.3 | 0.5 | 0.00 | 13 346.22 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.26 | 10 | 100 | 0.6 | 0.5 | 11.45 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.27 | 10 | 100 | 0.6 | 0.5 | 5.73 | 13 054.88 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.28 | 10 | 100 | 0.6 | 0.5 | 0.00 | 26 109.76 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.31 | 10 | 200 | 0.8 | 0.5 | 17.13 | 0.00 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.32 | 10 | 200 | 0.8 | 0.5 | 8.57 | 17 029.38 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.33 | 10 | 200 | 0.8 | 0.5 | 0.00 | 34 058.76 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.34 | 10 | 250 | 1.2 | 0.5 | 14.18 | 0.00 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.35 | 10 | 250 | 1.2 | 0.5 | 7.09 | 24 110.38 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.36 | 10 | 250 | 1.2 | 0.5 | 0.00 | 48 220.75 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.41 | 10 | 1000 | 2 | 0.5 | 33.28 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.42 | 10 | 1000 | 2 | 0.5 | 16.64 | 34 596.02 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.43 | 10 | 1000 | 2 | 0.5 | 0.00 | 69 192.04 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.55 | 15 | 100 | 0.3 | 0.5 | 34.44 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.56 | 15 | 100 | 0.6 | 0.5 | 17.18 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.57 | 15 | 200 | 0.8 | 0.5 | 25.70 | 0.00 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.58 | 15 | 250 | 1.2 | 0.5 | 21.27 | 0.00 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.59 | 15 | 1000 | 2 | 0.5 | 49.93 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.60 | 20 | 100 | 0.3 | 0.5 | 45.92 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.61 | 20 | 100 | 0.6 | 0.5 | 22.90 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.64 | 20 | 200 | 0.8 | 0.5 | 34.27 | 0.00 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.65 | 20 | 250 | 1.2 | 0.5 | 28.36 | 0.00 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.70 | 20 | 1000 | 2 | 0.5 | 66.57 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |

A.3. Runset to study effects of varying Mach number, Reynolds number and wall temperature ratio
| Run ID |
|
|
|
|
|
|
|
|
|
|
| (
|
(
|
(
|
(
|
(
|
(
|
(
|
||||
| A.22 | 10 | 100 | 0.3 | 0.5 | 22.96 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.25 | 10 | 100 | 0.3 | 1 | 21.80 | 0.00 | 1000 | 0.2180 | 984.96 | 188.74 |
| A.26 | 10 | 100 | 0.6 | 0.5 | 11.45 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.29 | 10 | 200 | 0.3 | 0.5 | 45.92 | 0.00 | 1000 | 0.4592 | 1969.91 | 266.92 |
| A.41 | 10 | 1000 | 2 | 0.5 | 33.28 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.44 | 10 | 1000 | 2 | 1 | 30.68 | 0.00 | 1000 | 0.3068 | 595.66 | 978.52 |
| A.45 | 10 | 1000 | 4 | 0.5 | 15.09 | 0.00 | 1000 | 0.1509 | 538.32 | 1860.47 |
| A.47 | 10 | 2000 | 2 | 0.5 | 66.57 | 0.00 | 1000 | 0.6657 | 1191.31 | 1383.84 |

A.4. Runset to study effects of wake structure on lift coefficient
| Run ID |
|
|
|
|
|
|
|
|
|
|
| (
|
(
|
(
|
(
|
(
|
(
|
(
|
||||
| A.22 | 10 | 100 | 0.3 | 0.5 | 22.96 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.26 | 10 | 100 | 0.6 | 0.5 | 11.45 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.29 | 10 | 200 | 0.3 | 0.5 | 45.92 | 0.00 | 1000 | 0.4592 | 1969.91 | 266.92 |
| A.30 | 10 | 200 | 0.6 | 0.5 | 22.90 | 0.00 | 1000 | 0.2290 | 1884.85 | 522.20 |
| A.41 | 10 | 1000 | 2 | 0.5 | 33.28 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |
| A.47 | 10 | 2000 | 2 | 0.5 | 66.57 | 0.00 | 1000 | 0.6657 | 1191.31 | 1383.84 |
| A.102 | 10 | 600 | 0.3 | 0.5 | 68.87 | 6673.11 | 1000 | 1.3775 | 1969.91 | 266.92 |

A.5. Blind validation cases
| Run ID |
|
|
|
|
|
|
|
|
|
|
| (
|
(
|
(
|
(
|
(
|
(
|
(
|
||||
| A.55 | 15 | 100 | 0.3 | 0.5 | 34.44 | 0.00 | 1000 | 0.2296 | 1969.91 | 266.92 |
| A.56 | 15 | 100 | 0.6 | 0.5 | 17.18 | 0.00 | 1000 | 0.1145 | 1884.85 | 522.20 |
| A.57 | 15 | 200 | 0.8 | 0.5 | 25.70 | 0.00 | 1000 | 0.1713 | 1804.06 | 681.18 |
| A.58 | 15 | 250 | 1.2 | 0.5 | 21.27 | 0.00 | 1000 | 0.1418 | 1607.23 | 964.42 |
| A.59 | 15 | 1000 | 2 | 0.5 | 49.93 | 0.00 | 1000 | 0.3328 | 1191.31 | 1383.84 |

Appendix B. Runset to estimate delineation between steady and unsteady flows

Appendix C. Transformed compressible Reynolds number
The transformed compressible Reynolds number
$\widetilde {\textit{Re}_{\textit{p}}}$
derived by Singh et al. (Reference Singh, Kroells, Li, Ching, Ihme, Hogan and Schwartzentruber2020) accounts for the changing relationship between compressibility and viscous effects as Mach number increases. The transformed compressible Reynolds number is
where
$T_s$
is the free stream temperature on the stagnation streamline,
$T_{ps}$
is the post-shock temperature, and
$ {T_{s}}/{T_{ps}}$
is the inverse Rankine–Hugoniot ratio of temperatures across a normal shock if
$M_s\gt 1$
and 1 otherwise. Here,
$\gamma$
is the ratio of specific heats,
$\textit{Re}_{\textit{p}}$
is the Reynolds number evaluated at free stream conditions on the stagnation streamline,
$\omega$
is a viscosity power term approximated as
$\omega = 1$
. The temperature ratio
$\theta _{ps}$
is
\begin{equation} \theta _{ps} = \left ( \frac {\widetilde {T}}{T_{ps}} \right ) ^ {\frac {\gamma }{\gamma -1}}, \end{equation}
where
$T_{ps}$
is the Rankine–Hugoniot post-shock temperature if
$M_s\gt 1$
and
$T_{ps}=T_s$
otherwise, and
$\: \widetilde {T}$
is the temperature from the Howarth (Reference Howarth1948) and Stewartson (Reference Stewartson1949) compressible boundary layer transformation given by
Then,
$\alpha$
is defined as
where
$\alpha _0=0.356$
. The reader is directed to Singh et al. (Reference Singh, Kroells, Li, Ching, Ihme, Hogan and Schwartzentruber2020) for details.




































