AIAA_ERDC_1
AIAA_ERDC_1
AIAA_ERDC_1
High-speed turbulent boundary layers has been revitalized in recent years due to an
increasing interest in minimizing the frictional drag of hypersonic vehicles. Large eddy
simulations (LES) is a good candidate against direct numerical simulations (DNS), which
is very prohibitive from the computational cost standpoint, especially for high Reynolds
number flows. However, to make LES more affordable it is desired to model the turbulent
flow at the wall using wall models. In this paper, we test several wall-stress models into
Loci-CHEM CFD code, which is a high-order finite-volume flow solver supporting the use
of multi-species finite-rate chemistry. The test results and discussions that we include are
subsonic and supersonic turbulent channel flows.
I. Introduction
The computational cost of wall-resolved LES applied to high-Reynolds number flows is highly prohibitive,
almost the same caliber as DNS because the mesh resolution has to be high enough to resolve the smallest
scales near the wall. There are two main approaches to run LES of wall turbulence which includes either
wall-resolved or wall-modeled depending on whether these near-wall flow structures are resolved on the
computational grid or modeled, respectively. The latter makes LES suitable to high Reynolds number wall-
bounded flows because a model is used in the inner layer, while directly resolving the predominant flow
scales in the outer layer. Further, wall-modeled LES methods can be broadly split into two categories:
hybrid Reynolds averaged Navier-Stokes (RANS)/LES formulation, where RANS is employed in proximity
to the wall LES away from the wall; and formulations where the LES governing equations are solved in
the entirety of the domain and wall-stress models are used at the wall as a boundary condition. In the
former category, the separation between RANS and LES domains is either explicit, as in zonal approaches,
or implicit usually modeled using closure models, such as detached-eddy simulation (DES) (Spalart13 ).
In the wall-stress modeling approach, the mesh covers the entire flow domain, with sufficient resolution
requirements in the outer portion of the boundary layer (the outermost 80-90%) to resolve the important
flow scale, and a different set of requirements in the inner part of the boundary layer, where there is not
sufficient resolution to resolve the small scales at the wall. Therefore, since the LES mesh is not capable to
resolve the important energetic scales in the inner layer, wall-stress models come into play. These models
take the instantaneous LES velocity and temperature as the input and yield the instantaneous wall stress
and heat flux as the output to the LES.
∗ Research Assistant Professor, Center for Advanced Vehicular Systems, Mississippi State University, MS 39762; member
AIAA.
† Graduate Research Assistant, Department of Aerospace Engineering, Mississippi State University, MS 39762; member
AIAA.
‡ Professor, Department of Computer Science and Engineering, Mississippi State University, MS 39762; member AIAA.
§ Associate Professor, Department of Mechanical Engineering, Mississippi State University, MS 39762; member AIAA.
¶ Research Engineer, US Army Engineer Research and Development Center.
‖ Associate Professor, Department of Aerospace Engineering, Mississippi State University; Associate Fellow AIAA.
1 of 17
A. Governing equations
The governing equations are three-dimensional unsteady compressible Navier-Stokes equations for single-
phase and single-species, which involves mass, momentum and energy equations:
∂ρ
+ ∇ · (ρũ) = 0 (1)
∂t
∂ρũ
+ ∇ · (ρũũ) = −∇p + ∇ · (σ + τ ) (2)
∂t
∂ρE
+ ∇ · (ρũH) = κ∆T + ∇ · ũ · (σ + τ ) + ∇ · q (3)
∂t
where, ∇ and ∆ are gradient and Laplacian operators, respectively. ρ and ũ are fluid density and velocity,
respectively. κ is thermal conductivity. σ are the viscous stresses:
2
σ = µ 2D − (∇ · ũ)I (4)
3
where, I is the identity matrix and µ is dynamic viscosity. D = 12 (∇ũ + ∇ũT ) is the rate-of-strain tensor.
The pressure (p) is defined assuming fluid to be a perfect gas:
p = ρRT (5)
where, R is universal gas constant and T is the fluid temperature. Energy E and enthalpy H are:
1
E = ρcv T + ρũ · ũ (6)
2
1
H = ρcp T + ũ · ũ (7)
2
2 of 17
ũ = ρc
u/ρ̂ (8)
The turbulence stress and heat transfer are:
τ = − (ρuu
d − ρ̂ûû) (9)
q = − ρuT
d − ρ̂ûT̂ (10)
where f is used to relate the wall stresses in wall-tangential directions to the velocity ũ0 and the pressure p0
of the exterior flow at the coupling position x̃0 . ν is the kenimatic molecular viscosity. f can be an algebraic
function such as a logarithm law, or a differential equation such as the thin boundary layer equation with
a simple algebraic turbulence model. In this paper, three wall stress turbulence models are investigated,
namely Werner-Wengle model, equilibrium ODE model and Nichols-Nelson wall function, and are introduced
as follows.
with
1
A = 8.3 and B = . (13)
7
The relations
up yp uτ
u+ = and y + = (14)
uτ ν
are defined for the average velocity within the first cell. up is the magnitude of gas velcoity at the first
1
grid point parallel to the wall, and y is the distance from the wall. For y + ≤ A 1−B , the friction velocity
+ +
can be obtained from the average cell velocity magnitude up by solving u = y for the friction velocity.
For greater values of y + the average velocity in wall coordinates in the first cell needs to be determined by
averaging the piecewise function. This yields a piecewise definition for uτ (up ). The model assumed that the
instantaneous tangential velocity components at the first near-wall nodes are parallel to the wall and are
in phase with the wall shear stress. The values of the tangential velocity can be related to the wall shear
stress by integrating the velocity profile over the height of wall cells. As a result, this model can provide the
3 of 17
where ∆y is the height of the first cells near the wall. While most of the wall stress models calculate the
mean wall shear stress using the iterative solution, the wall shear stress in the Werner-Wengle model can be
explicitly expressed and applied as the boundary conditions at the wall.
y+
r
∗ τw
µT = kρy 1 − exp − + (17)
ρ A
where k = 0.41 is the von Karman constant, and the model parameter A+ = 17. The wall-model simulation
takes the LES flow variables (up , T ) and impose them at the outer boundary, which is located at a specified
wall-normal distance, hwm (in our simulations, this is the third LES mesh point from the wall):
dT
at y = 0, up = 0, T = Tw or =0 (18)
dy
at y = hwm , up = uLES , T = TLES , p = pLES (19)
Figure 1. The ODE grids and their relationship to the LES grid.
After the system of equations (16) is solved numerically, the the shear stress and the temperature or the
heat flux, depending on whether the wall is adiabatic or isothermal, respectively, are calculated from velocity
and temperature, respectively, and interpolated into the LES. We solved equations (16) numerically using
4 of 17
where
2Γu+ − β
κ
+
ywhite = exp{ √ sin−1 − φ }exp(−κB) (21)
Γ Q
The constants κ and B are generally taken as 0.4 and 5.5 respectively. The wall function boundary condition
can be tuned for such effects as surface roughness by adjusting the κ and B constants. Compressibility and
heat transfer effects can be included by replacing the regular incompressible adiabatic law-of-the-wall term
with outer velocity form of White and Christoph? given by
" √ + !#
1 Γ y
u+ = β + Qsin φ + ln (22)
2Γ κ y0+
where 1
P r 1 u2τ
Γ=
2Cp Tw
qw µw
β=
ρw Tw kw uτ
−β
(23)
φ = sin−1
Q
Q = (β 2 + 4Γ)1/2
y0+ = exp(−κB)
The non-dimensional parameter Γ models compressibility effects and the parameter β models heat transfer
effects.
C. Numerical methods
The solver uses advanced generalized grid algorithms based on cell-central finite-volume methods and high-
resolution Riemann solvers. Second-order spatial accuracy on unstructured grids is typically obtained by
using a least-squares gradient reconstruction combined with a Barth7 or Venkatakrishnan-style8 limiter for
compressible flow problems. A third-order expilict Runge-Kutta scheme is used for time integration. In
order to support LES simulations, the inviscid flux is calculated using a low-dissipation skew-symmetric
central difference scheme which is 4th-order accurate on uniform meshes. A user-selectable amount of 3rd-
order upwinding can be blended into this scheme for stability and robustness. The low-dissipation flux
significantly improves the dispersion and dissipation characteristics of the inviscid flux calculation which
is essential for LES modeling. It should be noted that, at the present time, the low-dissipation central
difference scheme functions more effectively on hexahedral or polyhedral meshes than tetrahedral meshes. A
description of this flux scheme follows in section D below.
5 of 17
where the density, ρ(p, T ), and internal energy, e(p, T ), are provided by the equation-of-state, ~n is the face
normal vector, and the averaged variables pavg , Tavg , ~uavg , and kavg are provided by the following averaging
relations:
1
pavg = (pl + pr )
2
1
Tavg = (Tl + Tr )
2
1 (25)
~uavg = (~ul + ~ur )
2
1
kavg = ~uavg · ~uavg − (~ul · ~ul + ~ur · ~ur )
4
The above formulation reproduces the compressible kinetic energy consistent (KEC) form of Subbareddy
et al.? due to the definition of average fluid kinetic energy, kavg . However, this scheme differs from the
Subbareddy flux form in that averages of temperature and pressure are formed and then density and internal
energy are derived from the equation-of-state (EoS), whereas the original KEC flux formed independent
averages for these variables. Other options for averaging, including density and internal energy averaging,
were also evaluated and it was found that the above averaging process generally produced errors that are
lower by a factor of up to two.
The second order skew-symmetric flux of eq. (24) can be used as a basis to construct a 4th order scheme
following the construction of Pirozzoli?, ? and given here as the form:
where β = 61 and ll denotes the value of the cell left-of-left and rr denotes the value of the cell right-of-right.
This 4th order skew-symmetric flux (obtained when β = 61 ) will be denoted as the SSF scheme. When
β = 0 this flux reduces to the KEC scheme of Subbareddy et al.. This flux will provide the central flux
scheme in our hybridized low dissipation scheme. In practice the skew symmetric fluxes (Eqn (27)) fail for
shocked flows and therefore it is blended with standard upwind fluxes to enable simulation of high speed,
compressible flows.31
6 of 17
Temperature T 300K
Pressure P 1atm
Specific heat cp 1006 J/kg − K
Gas constant R 287.05 J.kg − K
Density ρ 1.161kg/m3
Specific heat ratio γ 1.4
√
Speed of sound a = γRT 347.75 m/s
Mach number M 0.1
Centerline velocity Uc = M a 34.75 m/s
Reynolds number based on Uc Rec 12485
Half channel height H 1m
Kinematic viscosity ν = Uc H/Rec 0.002783 m2 /s
Dynamic viscosity µ = ρν 0.0032317 kg/m.s
Prandlt number Pr 0.72
Re based on friction velocity Reτ 590
Friction velocity uτ = Reτ ν/H 1.64195 m/s
Wall stress τw = ρu2τ 3.13 N/m2
∂P τw
Pressure gradient ∂x = − H 3.13N/m3
Both subsonic (M = 0.1) and supersonic (M = 3) are simulated using implicit LES with Werner-Wengle
wall model, equilibrium ODE model and Nichols-Nelson wall laws to model the turbulent near the wall.
Periodic boundary conditions are applied in the streamwise and spanwise directions. For subsonic case,
adiabatic wall boundary condition is prescribed. Channel walls for supersonic case are kept isothermal and
are cooled at a temperature of Tw = 500K. To ensure the flow a statistically stationary state that preserves
∂p
the initial bulk velocity, body force that equals the x component of momentum removed from the walls ( ∂x )
∂p
is imposed through out the volume. For isothermal case (M = 3), kinetic energy ( ∂x ux ) is also added as
a source term and the heat flux balances this energy source term when the flow reaches statistically steady
state. To generate initial turbulence, perturbation based on Fourier mode is adopted in the flow initial
conditions as follow:
ρ(x, y, z) = ρb
u(x, y, z) = ub
(28)
v(x, y, z) = −0.5ub sin(2y)cos(2z)
w(x, y, z) = 0.5ub cos(2y)sin(2z)
After certain time of initial run, online averaging is adopted to gather ensemble averaged statistics followed
by space averaging.
7 of 17
Figures 2 and 3 show the comparison between the simulation results from the wall models on the fine grid
(643 ) and DNS results. Figures 2(a) and 3(a) are profiles of non-dimensional mean tangential velocity along
the non-dimensional height of the channel given by y + = yρw uτ /µw . For subsonic case, the commonly
p used
form of non-dimensional mean tangential velocity is adopted, defined as u+ = u/uτ with uτ = τw /ρw .
In supersonic case, since the flow is compressible and the fluid properties such as density and viscosity
experience large changes caused by viscous heating, Van Driest transformation28, 30 taking into account for
the compressibility effect is applied to compute the u+ expressed as
ˆ u+ p
u+
VD = ρ/ρw du (29)
0
Figures 2 and 3 demonstrate that simulation solutions of mean velocity and fluctuation field from the three
wall models on fine grids compare well with DNS results in general. However, for the incompressible case
(Fig. 2), the equilibrium ODE model predicts velocity fluctuation well (Fig 2(b)(c)(d)(e)) while the the
Werner-Wengle and Nichols-Nelson wall models appear to over-predict and under-predict the fluctuation
respectively. The good prediction by the wall models are expected since the first grid point is within the
viscous sublayer
Next, we focus on the evaluation of the wall models in LES simulations on the coarse grid, which is
the purpose of the models to decrease the computational cost. As mentioned in the description of the
equilibrium wall model, the wall model simulation takes the LES flow variables (up , T ) and impose them
as the top boundary conditions in the ODE solver. The location where the LES flow solutions are taken
is called exchange location, and this is not normally the first cell point from the wall due to the boundary
effect. Instead, the exchange location is generally chosen to be further away from the wall, and takes as the
third cell off the wall in this study. We apply this exchange location for all the three turbulent wall models
for the channel flow LES simulations on the coarse grid.
Figures 4 and 5 show the comparison between the simulation results from the wall models on the coarse
grid (62 × 31 × 64) and DNS solutions. There are good agreement of mean and fluctuated velocity field
between LES simulation and DNS results for subsonic case (Fig. 4) although the Werner-Wengle wall model
underpredicts the mean velocity. For supersonic case (Fig. 5), the equilibrium model and Nichols-Nelson wall
law predict the profile of mean velocity and mach number well, while the Werner-Wengle model significantly
8 of 17
(c) (d)
(e)
Figure 2. Comparison of simulation results from three wall models for M = 0.1 subsonic channel flow on fine
grids (643 ): (a) profile of mean velocity and (b) profile of x component of velocity fluctuation (variance in u) (c)
profile of y component of velocity fluctuation (variance in v) (d) profile of z component of velocity fluctuation
(variance in w) (e) profile of Reynolds shear stress (co-variance in uv)
9 of 17
(c) (d)
Figure 3. Comparison of simulation results from three wall models for M = 3 subsonic channel flow on fine
grids (643 ): (a) profile of mean velocity and (b) profile of Mach number (c) profile of streamwise stress (d)
profile of Reynolds shear stress
10 of 17
p p
where wall friction velocity is uτ = τw /ρw and local friction velocity is u∗τ = τw /ρ. SL stands for
semi-local. It should be noted that since damping function itself is empirical and adjusted to yield the
correct near-wall eddy-viscosity behavior, it appears that the choice of a suitable damping function scaling
for compressible flows must also determined empirically.
We first performed a priori analysis where the equilibrium wall model equation (Eqn 16) are solved using
time-averaged data from DNS database at exchange location from the wall and return the wall shear stress
and heat flux. A second-order
finite
volume scheme is used to solve the couple ODE. The relative error
−QDN S
defined as EQ (%) = Qwm QDN S × 100% between the simulation results and DNS are then computed
where Q is either shear stress or heat flux at the wall. Figure 8 shows the profile of error for the three
different scaling method for M ach = 3 compressible channel flow. It shows that the local scaling works the
best, while traditional wall scaling has the worst performance. The error peaks at the exchange location
about y/H ≈ 0.1.
Following a priori test, we then conduct a posteriori analysis for the same channel case using the three
different scaling functions expressed by Eqn (30) in the definition of eddy viscosity (Eqn (17)). Figure 9
compares the profile of mean velocity from various damping function scaling, and it shows the semi-local
scaling produces the best agreement with DNS, different from a priori result. The shear stress and heat flux
are decoupled in a priori test, while they are tightly coupled in a posteriori test. Therefore, the results for
the best scaling selection can be inconsistent for the two tests, which complicates the conculsion of scaling
analysis. Note that Iyer et. al performed the similar a priori and a posteriori studies on supersonic channel
and boundary layer flows to document the effect of different damping function scaling in equilibrium wall
model,32 and they included more scaling definitions by mixing the scalings defined in Eqn 30. Their results
on scaling predictivity are not consistent with different flow conditions, and do not agree completely with
our findings. Furthermore, the local scaling approach may not retain a monotonic increasing function for
+ +
ySL and ylocal which is at the very least non-physical. The ad-hoc nature of these blending approaches to
the damping function are unlikely to produce a universal modeling solution. Instead a more foundational
approach to damping that accounts for the physical effects of the variation of density near the wall is needed.
11 of 17
(c) (d)
(e)
Figure 4. Comparison of simulation results from three wall models for M = 0.1 subsonic channel flow on coarse
grids (64 × 31 × 64): (a) profile of mean velocity and (b) profile of x component of velocity fluctuation (variance
in u) (c) profile of y component of velocity fluctuation (variance in v) (d) profile of z component of velocity
fluctuation (variance in w) (e) profile of Reynolds shear stress (co-variance in uv)
12 of 17
(c) (d)
Figure 5. Comparison of simulation results from three wall models for M = 3 subsonic channel flow on coarse
grids (64 × 31 × 64): (a) profile of mean velocity and (b) profile of Mach number (c) profile of streamwise stress
(d) profile of Reynolds shear stress
13 of 17
(c)
Figure 6. Instantaneous velocity field for M ach = 3 turbulent channel flow (a) Equilibrium ODE wall model
(b) Werner-Wengle wall model (c) Nichols-Nelson wall function
14 of 17
Figure 7. Comparison of wall models with different exchange location (a) Equilibrium ODE wall model (b)
Nichols-Nelson wall function
(a) (b)
Figure 8. Effect of damping function scalings for the equilibrium wall model (a) profile of relative error on
shear stress (b) profile of relative error on heat flux
15 of 17
IV. Conclusion
Implicit LES simulations with three wall-stress models (Werner-Wengle, equilibrium ODE and Nichols-
Nelson) are carried out on adiabatic subsonic and isothermal supersonic channel flows. In general, the
equilibrium ODE and Nichols-Nelson model perform well compared with the DNS results, while the Werner-
Wengle wall model underpredicts the mean velocity for the supersonic case probably due to the use of the
laminar model for heat transfer. The exchange points where the LES results are taken as the input to the
wall models are important since the LES solutions next to wall could be deteriorated due to the impact by
the wall boundary. The scaling analysis on damping function in the equilibrium ODE wall model shows the
semi-local scaling produces the best results for the supersonic channel flow, however, the blending approach
is ad-hoc and a physics based approach would likely provide a more universal modeling approach.
Acknowledgments
This work by Mississippi State University was financially supported by the U.S. Department of Defense
(DoD) High Performance Computing Modernization Program, through the US Army Engineer Research and
Development Center (ERDC) Contract #W912HZ21C0011. The views and conclusions contained herein
are those of the authors and should not be interpreted as necessarily representing the official policies or
endorsements, either expressed or implied, of the U.S. Army ERDC or the U.S. DoD.
References
1 Boiko, A.V., Ivanov, A.V., Kachanov, Y.S. and Mischenko, D.A. (2010) Steady and unsteady GöTell boundary-layer
instability on concave wall, European Journal of Mechanics B/Fluids, Vol. 29, pp. 61-83.
2 Choudhari, M. and Fischer, P. (2005) Roughness induced Transient Growth, AIAA Paper 2005-4765.
3 Depmsey, L.D., Hall, P. & Deguchiu, K. (2017) The excitation of Gortler vortices by free stream coherent structures, J.
16 of 17
Jan. 1993.
9 Roe, P. L., Approximate Riemann solvers, parameter vectors, and difference schemes," Journal of Computational Physics,
for Simulation of Complex Turbulent Flows, ASME International Mechanical Engineering Congress and Exposition, Volume
7A: Fluids Engineering Systems and Technologies, 2015.
11 Thornber, B., Drikakis, D., Williams, R., and Youngs, D., On entropy generation and dissipation of kinetic energy in
high-resolution shock-capturing schemes," Journal of Computational Physics, Vol. 227, No. 10, 2008, pp. 4853-4872.
12 Thornber, B., Mosedale, A., Drikakis, D., Youngs, D., and Williams, R., An improved reconstruction method for com-
pressible flows with low Mach number features," Journal of Computational Physics, Vol. 227, No. 10, 2008, pp. 4873-4894.
13 Spalart P.R. (2009) Detached-eddy simulation. Annu. Rev. Fluid Mech., Vol. 41, pp. 181-202.
14 Werner, H., and Wengler, H. (1993) Large-eddy simulation of turbulent flow over and around a cube in a plate channel.
and Higher-Order Moments in Turbulent Wall-Bounded Flow, J. of Fluid Mech, Vol. 757, Oct. 2014, pp. 888-907.
doi:10.1017/jfm.2014.510
18 Park, G. I., and Moin, P., Space-Time Characteristics ofWall-Pressure and Wall Shear-Stress Fluctuations in Wall-
Modeled Large Eddy Simulation, Physical Review Fluids,Vol. 1, No. 2, 2016, Paper 024404. doi:10.1103/PhysRevFluids.1.024404
19 Yang, X. I. A., On the Mean Flow Behaviour in the Presence of Regional-Scale Surface Roughness Heterogeneity, Boundary
Insights for Aeolian Processes, Boundary Layer Meteorology, Vol. 162, No. 1, 2016, pp. 1-21.
21 Bou-Zeid, E., Overney, J., Rogers, B. D., and Parlange, M. B., The Effects of Building Representation and Clustering
in Large-Eddy Simulations of Flows in Urban Canopies, Boundary Layer Meteorology, Vol. 132, No. 3, 2009, pp. 415-436.
doi:10.1007/s10546-009-9410-6
22 Yang, X. I., Sadique, J., Mittal, R., and Meneveau, C., Exponential Roughness Layer and Analytical Model for Tur-
bulent Boundary Layer Flow over Rectangular-Prism Roughness Elements, J. Fluid Mech, Vol. 789, Feb. 2016, pp. 127-165.
doi:10.1017/jfm.2015.687
23 Sayadi, T., and Moin, P., Large Eddy Simulation of Controlled Transition to Turbulence, Physics of Fluids, Vol. 24, No.
Wave/Turbulent Boundary Layer Interactions Through Wall-Modelled Large-Eddy Simulations, J. Fluid Mech, Vol. 758, Nov.
2014, pp. 5-62. doi:10.1017/jfm.2014.505
25 Vane, Z., Bermejo-Moreno, I., and Lele, S. K., Wall-Modeled Large- Eddy Simulations of a Supersonic Turbulent Flow
in a Square Duct, 44th AIAA Fluid Dynamics Conference, AIAA Paper 2014-2209, 2014.
26 Larsson, J., Laurence, S., Bermejo-Moreno, I., Bodart, J., Karl, S., and Vicquelin, R., Incipient Thermal Choking
and Stable Shock-Train Formation in the Heat-Release Region of a Scramjet Combustor. Part 2: Large Eddy Simulations,
Combustion and Flame, Vol. 162, No. 4, 2015, pp. 907-920.
27 Stolz, S. & Adams, N. A. 2003 Large-eddy simulation of high-Reynolds-number supersonic boundary layers using the
approximate deconvolution model and a rescaling and recycling technique. Phys. Fluids 15, 2398-2412.
28 Foysi, H., Sarkar, S., and Friedrich, R., (2004) Compressibility Effects and Turbulence Scalings in Supersonic Channel
17 of 17