You are on page 1of 12

AIAA 98-3315

COMPRESSIBLE LARGE EDDY SIMULATION USING


UNSTRUCTURED GRIDS: CHANNEL AND BOUNDARY
LAYER FLOWS
Nora Okong'o* and Doyle D. Knight*
Department of Mechanical and Aerospace Engineering
Rutgers, The State University of New Jersey
98 Brett Road, Piscataway, New Jersey 08854-8058

Abstract 1 Introduction
An unstructured grid algorithm for tetrahedral As the speed and memory capabilities of com-
cells has been developed for Large Eddy Simula- puters continue to increase, the prospect of per-
tion (LES). The finite volume form of the com- forming Large Eddy Simulation (LES) of engi-
pressible Navier-Stokes equations are solved for neering problems becomes more feasible. In LES,
cell-averaged conservative variables. Inviscid flux the Navier-Stokes equations are spatially filtered
computations are performed by applying a Rie- to compute the large scale flow structures, while
mann solver across each face, the values at the modeling only the small or sub-grid scale (SGS)
points on the faces being obtained by least-squares structures [4, 6]. As the small scale structures
reconstruction from the cell-averaged values. The are expected to be approximately universal in tur-
viscous fluxes and heat transfer are obtained by bulent flows, LES can potentially be applied to
application of Gauss' theorem. The numerical wider classes of problems than traditional turbu-
scheme is explicit, with second-order spatial and lence models, but at reduced computational cost
temporal accuracy. The sub-grid scale model is compared to direct numerical simulation (DNS).
of the constant coefficient Smagorinsky type, with However, whereas most of the work in the field has
van Driest damping for the viscous sublayer. been on structured meshes in the context of finite-
For channel flow, results at a Reynolds num- difference or spectral methods [7], application of
ber (based on channel height and bulk velocity) of LES to engineering problems requires methodolo-
5600 and Mach number of 0.5 are compared to in- gies suitable for complex geometries. Towards
compressible direct numerical simulations (DNS) this end, we have developed a three-dimensional
and experiments. Mean velocity and velocity fluc- Large Eddy Simulation (LES) algorithm for un-
tuations compare well with the DNS and experi- structured meshes [12, 21].
mental data. For boundary layer flow, results at a Our LES unstructured grid algorithm uses a
Reynolds number (based on the inflow boundary cell-centered storage architecture, with the cell-
layer thickness) of 20,000 and Mach 3 are com- averaged values stored at the centroid of each
pared to experimental data. Preliminary results tetrahedral cell. The inviscid fluxes are computed
for profiles of the streamwise mean velocity and by applying a Riemann solver across each face.
turbulence intensity and for the skin friction show The values of the flow variables at one or more
good agreement with experiment. points on either side of the face are obtained from
'Graduate Research Assistant, Member AIAA the cell-averaged values by function reconstruc-
okongo@jove.rutgers.edu tion. The function reconstruction is based on the
http://cronos.rutgers.edu/~okongo
'Professor, Associate Fellow AIAA the least squares method of Ollivier-Gooch [22],
knight@jove.rutgers.edu and is second-order accurate. Higher order recon-
http://www-srac.rutgers.edu/~knight struction can also be achieved using this method.
Copyright ©1998 by Nora Okong'o and Doyle D. Knight. Details of our implementation can be found in Ref.
Published by the American Institute of Aeronautics and [21]. The simulation is advanced in time using
Astronautics, Inc. with permission. second-order explicit Runge-Kutta time integra-
tion. The algorithm incorporates a constant co- features missed by their k — c model, such as recir-
efficient Smagorinsky sub-grid scale (SGS) model culation for flow in a ribbed channel. These mixed
with van Driest damping for the viscous sublayer. results illustrate the drawbacks of traditional tur-
The objective of this paper is to validate our bulence models, and hence the need for LES.
LES unstructured grid algorithm for two turbulent Several researchers have performed incompress-
shear flows. The first flow is low Mach number tur- ible LES of channel flow on structured grids us-
bulent channel flow, which is a canonical test case ing spectral methods, including Moin and Kim
for LES. We compare our results to the incom- [17], and Piomelli et al [23]. Spectral methods
pressible DNS of Kim et al [11] and experimental cannot be used in general geometries, prompting
data of Eckelmann [5] and Kreplin and Eckelmann some attempts to use finite-difference methods,
[13]. The second flow is a supersonic turbulent e.g. [1]. Most sub-grid scale (SGS) models are
boundary layer. Preliminary results are compared based on the eddy-viscosity concept. The sim-
with the conventional one-seventh power law for plest is the constant coefficient Smagorinsky sub-
the mean velocity and the experimental data of grid scale (SGS) model [26]. Although dynamic
Yanta and Crapo [29] for streamwise turbulent in- and Lagrangian methods of determining the coef-
tensity. ficient by re-filtering the flowfield have also been
used [16, 19, 30], the additional complexity ap-
pears not to justify the small improvement in the
2 Previous Work results. This is mainly because at low Reynolds
numbers most of the turbulence energy is con-
2.1 Turbulent Channel Flow tained in the resolved scales, and thus the choice
of SGS model is not critical. In our research,
Turbulent channel flow is a classic benchmark case
we opt to use the simplest (constant-coefficient)
for validating turbulent flow simulations. It has
Smagorinsky-type model.
the advantage of geometric simplicity while admit-
There are few LES computations of turbulent
ting complex three-dimensional flow structures.
channel flow using unstructured grids. Haworth
When the flow reached statistical equilibrium, the
and Jansen [8] use an unstructured hexahedral
time-averaged flow statistics are functions of the
wall-normal direction only. This flow is charac- grid finite volume method. The results compare
terized by long streamwise vortices near the walls
favorably with the DNS of Kim et al [11] for the
[5, 13, 17, 23]. incompressible flow at Reynolds number of 5600.
The work of Haworth and Jansen is similar to ours
Data for fully-developed incompressible channel
in that we both use second-order algorithms. How-
flow at a Reynolds number of 5600 (based on bulk
ever they use variable explicit/implicit character-
velocity and channel height) is available from the
istics based advection scheme, Lagrangian SGS
experiments by Eckelmann [5] and Kreplin [13].
model and hexahedral elements. Compared to the
Turbulence statistics are also available from the
DNS, their computed centerline velocity is about
DNS of Kim et al [11]. These show that the near-
5% higher (1.22 compared to 1.16). The stream-
wall streaks have a spanwise spacing of about 100
wise fluctuations are somewhat over-predicted and
wall units and a maximum in turbulence fluctua-
the spanwise and wall-normal fluctuations are
tions at a distance of 12 or 13 wall units. The fric-
under-predicted, but show the correct behavior.
tion velocity, normalized by the bulk velocity, is
Jansen [10] uses a finite-element method with an
0.0643. DNS of supersonic channel flows by Cole-
unstructured tetrahedral grid. Results for the tur-
man et al [2] agree with the incompressible results
bulence intensities are in good agreement with the
when scaled to account for mean property varia-
DNS of Kim et al [11].
tions. Their results support Morkovin's hypothe-
sis, that relationships between statistics are unaf-
fected by compressibility if the root-mean-square 2.2 Supersonic Turbulent Bound-
density fluctuations are small. ary Layer
Computations of channel flow using the k — c
model give mixed results. Nisizima and Yoshizawa The supersonic turbulent boundary layer is an im-
[20] compute channel flow using an anisotropic portant benchmark for validating LES. Accurate
k — e model. There is good agreement for the simulation of the turbulent boundary layer is nec-
mean profiles and Reynolds stress, but poor agree- essary for providing a realistic (unsteady) inflow
ment for the turbulent kinetic energy. Ciofalo and condition for more complex flow simulations such
Collins [1] report that LES is able to capture flow as shock wave turbulent boundary layer interac-
tions. There are few LES computations of super- where /j is the applied body force
sonic turbulent boundary layer flow as a problem
in its own right, rather than as part of a more com- (7)
plex simulation. Spyropoulos and Blaisdell [28]
simulated a spatially evolving Mach 2.25 boundary
layer corresponding to the experiments of Shutts
Qk = -pcp(Tuk-fuk} (9)
et al [25] using both DNS and LES. They exam-
ined a variety of finite-difference schemes and con-
ducted grid refinement studies. They employed
a modified version of the code used by Rai et al
pe 1_
= pcvT + -pUjUi + pk (11)
[24] for DNS of the same configuration. Hunt and
Nixon [9] performed a structured grid LES of a
pk = -( (12)
supersonic compression ramp which included the
inflow turbulent flat plate boundary layer. Comte
The assumptions made in deriving the energy
and David [3] performed LES of a Mach 2.5 turbu-
equation, Equation 5, are discussed in detail in
lent boundary layer over a compression ramp us-
Ref. [12]. The standard Smagorinsky sub-grid
ing a structured grid. Results show the formation
scale model is [26]:
of Gortler vortices downstream of the compression
corner.

3 Governing Equations x I Si, - ISkkSij (13)

The governing equations are the three-dimensional Qj = pCr-^-A^SmnSmn-^


—— (14)
filtered unsteady compressible Navier-Stokes eq-
uations. For a function /, its filtered form / is:
(15)

'=£/ V Jv
QfdV (1) The two model constants are CR, the compress-
ible Smagorinsky constant, and Prt, the turbu-
where Q is the filtering function, and its Favre- lent Prandtl number. The values used are based
averaged form / is: on previous studies [6, 18]. We use CR = 0.012
for channel flow, CR = 0.0042 for boundary layer
flow and Prt — 0.4 for both flows.
/=?
P
(2)
For wall-bounded flows the constant CR is mul-
tiplied by the van Driest damping factor, 1 —
where p is the density. From the Navier-Stokes eq-
ey /26 This expression uses wall-units: y+ =
uations for the instantaneous flow variables den-
yuT/v (y measured from the nearest wall). The
sity (p), velocity in the zth coordinate direction
friction velocity is given by UT — \Jrwj pw where
( u i ) , pressure (p) and temperature (T), averaging
TW is the average wall shear stress over both walls
and filtering yield the filtered Navier-Stokes equa-
and pw is the similarly averaged density at the
tions (here written using the Einstein summation
wall. This approach is similar to that of other
notation where repeated indices denote summa- researchers [1, 15, 17].
tion): In finite volume form
{ _Q
(3)
dt ^ / QdV + I (Fi + Gj + Hk) • ndA
at Jv Jdv
+ I (Ri + Sf+ Tit) • ndA + I BdV = 0 (16)
Jdv Jv
dpe d
where
P
r\ f\ pu
a—— (Qk + qk) Uf + -z—— (Tik + &ik) U{ - fiUi (5)
Q= pv (17)
pw
= pRT (6) pe
0 averaged dependent variables stored at the cen-
/I troid of cell i are:
B= h (18)
fa Qi = 77 Qdxdydz (28)

The value of Q at the cell-centroid is a second-


_ pu _ order accurate approximation of Q;.
2
pu + p The governing equations can be written as:
F= puv (19)
puw
(pe+p)u
^(QM) + d = Q (29)

pv where C1,- is the net flux across the cell faces:


puv
G= pv2+p (20) d = I (Fz + Gj + Hk) -ndA+ I BdV
JdV, JV,
pvw
(pS + p)v (Ri + Sj + Tk) • ndA (30)
pw
puw It is assumed that Vj, the volume of the cell, is con-
H = pvw (21) stant. The inviscid flux vector (FT+Gj+Hk)-ndA
pw2 +p is computed using an exact Riemann solver. A
(pi +p)w least-squares reconstruction scheme is used to ob-
tain the values needed as input into the Rie-
0 mann solver from the cell-averaged values. For
each cell, a linear system for the spatial deriva-
R= -Txy - (Txy (22) tives at the cell-centroid are solved using singular
value decomposition, implemented using house-
holder transformations. The flow variables at any
-Q* -qx-px
location within the cell can then be obtained us-
0 ing a Taylor series expansion. The viscous flux
— Txy ~ &xy and heat transfer vector (Ri + Sj + Tk) • ndA is
5= —— Tyy —— Oyy (23) obtained by application of Gauss' theorem to con-
-Tyz - ffyz trol volumes whose vertices are the centroids of
—Qy — qy ~ Py the cells that share each node. Details of the nu-
merical methods can be found in Ref. [21]. The
0 algorithm has been implemented for parallel com-
putation using a one-dimensional domain decom-
T= (24) position and message passing via MPI.
The overall algorithm using least squares recon-
-Qz - qz ~ Pz struction is:
• Read in the grid and the initial cell-averaged
values for each cell
(25)
• Determine the cells to be in each stencil for
the least squares reconstruction
(26)
• Using the Householder method, solve the
Pz = (TXZ + f f x z ) u + (Tyz + 0-yz]
least squares problem for the spatial deriva-
(27) tives of each conservative variable at each cell-
centroid; store the derivatives

4 Numerical Algorithm • Use the derivatives to reconstruct each vari-


able on both sides of all cell face centroids;
An unstructured grid of tetrahedra is employed, obtain the inviscid fluxes by solving the Rie-
with a cell-centered storage architecture. The cell- mann problem using these variables
• Compute the viscous fluxes [13] and direct numerical simulation on structured
grids published by Kim et al [11] will be used to
• Compute body-force if needed compare the results. In the presentation of re-
• Update the cell-averaged variables using the sults, statistics are accumulated at all grid points.
computed fluxes Planar averages are obtained by averaging at all
points on planes parallel to the walls, thus mak-
• Repeat the last five steps as required for ing the statistics functions of the wall-normal (y)
the temporal integration; for explicit multi- direction only.
stage Runge-Kutta integration, the deriva-
tives need to be recomputed at each stage
5.2 Results
The results for the friction velocity UT and cen-
5 Channel Flow terline velocity Uc, both normalized by the bulk
velocity C/j,, are summarized in Table 1. Figure 1
5.1 Computational Configuration shows the history of the friction velocity, which
Results have been obtained for low Mach number by the end of the computation is 0.0606 and has
turbulent channel flow. The mean flow is in the in- an average value of 0.0594. This is within 8%
direction. A body-force is applied in this direction of the DNS and experimental values of 0.0643.
and is adjusted at each timestep to keep the bulk Figure 2 shows the planar averages of the time-
velocity constant. Using the global x-momentum averaged streamwise velocity. The centerline ve-
conservation for fully developed flow, /i = —2u^. locity is 1.155 which is within 0.6% of the experi-
The bulk velocity is also the reference velocity mental value of 1.162. Figure 3 shows the planar
and so is set equal to 1. The boundary condi- averages of the time-averaged velocity, normalized
tions are periodic in the streamwise (x) and span- by the friction velocity, compared to the veloc-
wise (2) directions with no-slip boundaries in the ity predicted by the linear law and the log law.
wall-normal (?/) direction, with isothermal walls. The LES is close to the DNS in the near-wall re-
The Reynolds number based on the channel height gion, but overshoots it slightly in the core of the
and bulk velocity is 5600. The Mach number channel, consistent with the lower shear stress of
is 0.5; however, the static temperature variation the LES. Notably, the LES has excellent agree-
across the channel is small (less than 4%) so the ment with the experimental data of Eckelmann
flow is effectively incompressible. Computations [5], whose log-law of u+ — 2.651nj/+ +5.9 is based
have been performed using the constant-coefficient on data at Rec of 5600 and 8200. The LES Rec
Smagorinsky sub-grid scale model (CR = 0.012, of about 6470 is between these values. The exper-
Prt= 0.4). The coefficient CR is multiplied by the imental data points shown in Figure 3 are for Rec
van Driest damping factor 1 — ey /26 to integrate of 5600 [5].
to the walls. The length scale A for the SGS model
is the nodal spacing in the ^-direction. Table 1: Friction and Centerline Velocities for
The grid size is 2?r x 1 x 27T/3, where the lengths Channel Flow Computations
have been non-dimensionalized by the channel
height. The grid has 65 x 65 x 65 nodes (274,625 LES DNS [11]
nodes), 1,310,720 cells, and 2,646,016 faces. The L/x X jLy X Liz 2n x 1 x 2?r/3 2;r x 1 x TT
grid is stretched in the y-direction, with mini- NxxNyx Nz 65 x 65 x 65 192 x 129 x 160
mum spacing (at the walls) of 0.00278 and a max- UT 0.0594 0.0643
imum spacing (in the channel center) of 0.0457. uc 1.155 1.162
The initial condition is obtained from interpolat-
ing the fully-developed solution on a grid with Figures 4, 5 and 6 compare the root-mean-
33 x 65 x 33 nodes with the same wall-normal (y) square velocity fluctuations from the LES with
spacing but double the streamwise and spanwise the DNS results. Also shown are the experimen-
spacing. The flow has been simulated for about tal results from Kreplin and Eckelmann [13]. The
twenty-two flow-through times, with statistics ac- experiments have Rec of 7700 and ReT of 388,
cumulated over the last twelve flow-through times. which are higher that the DNS Rec of 6600 and
(One flow-through time is the domain length in the ReT of 360 and the LES Rec of 6470 and ReT of
x-direction divided by the bulk velocity, i.e., 2?r.) 330. The velocity fluctuations from the LES, DNS
Experimental results of Eckelmann [5] and Kreplin and experiment show generally good agreement.
The streamwise fluctuations, Figure 4, of the LES tions are periodic in the spanwise (z) direction,
compare favorably with the experiment, but are zero-gradient extrapolation at the outflow, no-slip
slightly higher than the DNS. Figure 5 shows the adiabatic at the wall, and a characteristic bound-
LES wall-normal fluctuations to be lower than ary condition on the upper surface. The inflow
those of the experiment and DNS; however, the boundary condition is obtained by rescaling the
DNS although higher than the LES is still signif- flow at a downstream station, in a compressible
icantly lower than the experiment. For the span- extension of the method used by Lund et al [14].
wise fluctuations, shown in Figure 6, there is com- The Reynolds number based on the inflow bound-
parable agreement with experiment for the LES ary layer thickness 8 is 20,000. Computations
and DNS. For the LES, as in the DNS and exper- have been performed using the constant-coefficient
iment, the spanwise fluctuations are larger than Smagorinsky sub-grid scale model (CR = 0.0042,
the wall-normal fluctuations. Figure 7 shows the Prt = 0.4). The length scale A is multiplied by
Reynolds stress, which also compares well with the the van Driest damping factor 1 — ey /26 to inte-
DNS. Figure 8 shows the time-average of the y- grate to the walls. The length scale for the SGS
component of the z-momentum flux. Its linear model is the cube root of the average of two vol-
behavior indicates that statistical equilibrium has umes sharing a face.
been reached. The grid size is 14.8 x 3.4 x 4.4 where the
It should be noted that the comparison between lengths have been non-dimensionalized by the in-
the DNS and the LES is not precise because the coming boundary layer thickness S. The grid has
DNS has not been filtered onto the LES grid. 51 x 65 x 51 nodes (169,065 nodes), 800,000 cells
However, such filtering is not obvious anyway for and 1,617,800 faces. The grid is stretched in the
an unstructured grid. Additionally, for this case, y-direction, with a minimum spacing (at the wall)
most of the energy is in the resolved scales, as of 0.0048 and a maximum spacing of 0.2. The
shown by Figure 9, which shows that the SGS eddy initial condition for the streamwise velocity is the
viscosity is at most 9% of the laminar viscosity. Law of the Wall and Wake plus a random fluctu-
The higher-order velocity statistics based on the ation whose maximum was 10% of the freestream
resolved velocities are shown in Figures 10 through value. The initial conditions for normal and span-
15. The skewness for the streamwise velocity is wise velocities are independent random fluctua-
S(u') = u'u'u'/u'u' and likewise for v' and tions with maximum of 10% of the freestream
w', where the bar denotes the time-average and streamwise velocity. The initial condition for the
the prime denotes fluctuation relative to the time- static temperature assumes a Crocco integral for
average. Figure 10 shows the streamwise skew- the mean temperature plus a random perturba-
ness from the LES to have greater magnitude than tion with zero total temperature fluctuation. The
the DNS, particularly in the center of the chan- flow statistics have been accumulated for about
nel. The reasons for the disagreement with DNS eight flow-through times. (One flow-through time
are under evaluation. However, the agreement for is the domain length in the z-direction divided by
the wall-normal and spanwise skewness, Figures 11 the freestream velocity, i.e., 14.8.) Experimental
and 12 respectively, is excellent. The flatness for results of Yanta and Crapo [29] will be used to
_______ ___o
the streamwise velocity is F(u') = u'u'u'u/u'u' compare the results.
and likewise for v1 and w'. Once again the stream-
wise statistics of the LES deviate from the DNS,
as seen in Figure 13. The wall-normal and span- 6.2 Results
wise flatness, shown in Figures 14 and 15 agree
very well with the DNS. The computed skin friction coefficient at the inflow
boundary is ur/Ux = 0.0506, which is within 7%
of the theoretical value uT/Uoo = 0.0544 based on
6 Supersonic Boundary friction law obtained from the combined Law of
the Wall and Wake
Layer
6.1 Computational Configuration =^ In (31)
Preliminary results have been obtained for super-
sonic boundary layer flow at Mach 3. The mean where K = 0.4 is von Karman's constant, B' =
flow is in the z-direction. The boundary condi- 5.1 and II = 0.55. The van Driest transformed
freestream velocity s experiment for the skin friction, the streamwise
mean velocity and streamwise turbulence inten-
I sity. We intend to now use this algorithm to com-
pute more complex flows, such as turbulent super-
sonic flow into a compression corner. Anticipated
sin (32) algorithmic improvements include the incorpora-
tion of a dynamic sub-grid scale model.
where
1/2
(7-1) (33) 8 Acknowledgments
This work has been supported by AFOSR un-
B = - 1 (34) der Grant Number F49620-96-1-0389, monitored
by Dr. Len Sakell, and by the National Center
where Prtm = 0.89 is the mean turbulent Prandtl for Supercomputing Applications (NCSA), Uni-
number, Taw is the adiabatic wall temperature and versity of Illinois at Urbana-Champaign, under
Tw is the wall temperature. Grant Numbers ENG980001N and CTS980021N.
The computed adiabatic wall TW/TX = 2.68 Computations were performed on the NCSA Cray
which is within 3% of the theoretical value of 2.602 Origin2000, the NAVO OCEANO DOD Cray Ori-
computed from gin2000 and an SGI Power Onyx at Rutgers Uni-
versity. The results were analyzed at the Rutgers
= l+fciW M (35) University Supercomputer Remote Access Center.
We would like to thank Robert Murray for his as-
The mean velocity profile at x = 126 down- sistance with the flow visualizations. We would
stream of the inflow boundary is shown in Fig- also like to thank Gary Coleman, Dan Haworth,
ure 16 and compared with the one-seventh profile. Ken Jansen and Thomas Lund for their sugges-
Reasonable agreement is observed. tions and comments.
The mean streamwise resolved turbulence inten-
sity, normalized using the local mean density p
and wall shear stress TW, is shown in Figure 17,
References
together with the experimental data of Yanta and [I] M. Ciofalo and M. Collins. Large-Eddy Sim-
Crapo [29]. As discussed in Smits and Dussauge ulation of Turbulent Flow and Heat Transfer
[27], the scaling pu'2/Tw provides an approximate in Plane and Rib-Roughened Channels. In-
self-similar correlation of experimental data for su- ternational Journal for Numerical Methods in
personic flat plate boundary layers, although the Fluids, 15:453-489, 1992.
measurements near the wall are subject to consid-
erable uncertainty. The data of Yanta and Crapo [2] G.N. Coleman, J. Kim, and R.D. Moser.
at Mach 3 is representative of this data. With the A Numerical Study of Turbulent Supersonic
exception of the near wall region, the computed Isothermal-Wall Channel Flow. Journal of
results show good agreement with experiment. Fluid Mechanics, 305:159-183, 1995.
[3] P. Comte and E. David. Large-eddy Simu-
7 Conclusions lation of Gortlet Vortices in a Curved Com-
pression Ramp. In J.A. Desideri, B.N.
A three-dimensional unstructured grid algorithm Chetverushkin, Y.A. Kuznetsov, J. Periaux,
for accurate simulation of unsteady flows has been and B. Stoufflet, editors, Experimentation,
used for Large Eddy Simulations. The algorithm is Modelling and Computation in Flow, Turbu-
second-order in space and time and uses the con- lence and Combustion, volume 1, chapter 4.
stant coefficient Smagorinsky SGS model. Com- Wiley, Chichester, 1996.
putations of low Mach number channel flow at
a Reynolds number of 5600 agree well with in- [4] F.M. Denaro. Towards a New Model-Free
compressible DNS and experimental results for the Simulation of High-Reynolds-Flows: Local
mean velocity and turbulence fluctuations. Com- Average Direct Numerical Simulation. In-
putations of supersonic boundary layer at Mach 3 ternational Journal for Numerical Methods in
and Reynolds number of 20,000 agree well with Fluids, 23:125-142, 1996.
[5] H. Eckelmann. The Structure of the Viscous [15] P.J. Mason and N.S. Callen. On the Mag-
Sublayer and the Adjacent Wall Region in a nitude of the Subgrid-Scale Eddy Coeffi-
Turbulent Channel Flow. Journal of Fluid cient in Large-Eddy Simulations of Turbulent
Mechanics, 65:439-459, 1974. Channel Flow. Journal of Fluid Mechanics,
162:439-462, 1986.
[6] G. Erlebacher, M.Y. Hussaini, C.G. Speziale,
and T.A. Zang. Toward the Large-Eddy [16] C. Meneveau, T. Lund, and W. Cabot. A
Simulation of Compressible Turbulent Flows. Lagrangian Dynamic Subgrid-Scale Model of
Journal of Fluid Mechanics, 238:155-185, Turbulence. Journal of Fluid Mechanics,
1992. 319:353-385, 1996.
[7] B. Galperin and S.A. Orszag, editors. Large [17] P. Moin and J. Kim. Numerical Investigation
Eddy Simulation of Complex Engineering and of Turbulent Channel Flow. Journal of Fluid
Geophysical Flows. Cambridge University Mechanics, 118:341-377, 1982.
Press, 1993.
[18] P. Moin, K. Squires, W. Cabot, and S. Lee.
[8] D.C. Haworth and K. Jansen. Large- A Dynamic Subgrid-Scale Model for Com-
Eddy Simulation on Unstructured Deform- pressible Turbulence and Scalar Transport.
ing Meshes: Towards Reciprocating 1C En- Physics of Fluids, 3:2746-2757, 1991.
gines. Center for Turbulence Research, Sum-
[19] Y. Morinishi and T. Kobayashi. Large Eddy
mer Program 1996. Submitted to Computers
Simulation of Complex Flow Fields. Comput-
& Fluids.
ers & Fluids, 19:335-346, 1991.
[9] D. Hunt and D. Nixon. A Very Large [20] S. Nisizima and A. Yoshizawa. Turbu-
Eddy Simulation of an Unsteady Shock lent Channel and Couette Flows Using an
Wave/Turbulent Boundary Layer Interac- Anisotropic K — e Model. AIAA Journal,
tion. AIAA 95-2212, 1995. 25(3):414-420, 1987.
[10] K. Jansen. Large Eddy Simulation Using Un- [21] N. Okong'o and D.D. Knight. Accurate
structured Grids. In C. Liu and L. Sakell, Three-Dimensional Unsteady Flow Simula-
editors, First AFOSR International Confer- tion Using Unstructured Grids. AIAA 98-
ence on DNS/LES, pages 117-128. Greyden 0787, 1997. 36th Aerospace Sciences Meeting,
Press, Columbus, OH, Louisiana Tech Uni- January 12-15, 1998, Reno, Nevada.
versity, Ruston, LA, 1997.
[22] C.F. Ollivier-Gooch. High-Order ENO
[11] J. Kim, P. Moin, and R. Moser. Turbulence Schemes for Unstructured Meshes Based on
Statistics in Fully Developed Channel Flow Least Squares Reconstruction. AIAA 97-
at Low Reynolds Number. Journal of Fluid 0540, 1997. 35th Aerospace Sciences Meet-
Mechanics, 177:133-166, 1987. ing and Exhibit, January 6-10, 1997, Reno,
Nevada.
[12] D.D. Knight, G. Zhou, N. Okong'o, and
V. Shukla. Compressible Large Eddy Sim- [23] U. Piomelli, L. Ong, C. Wallace, and D. Lad-
ulation Using Unstructured Grids. AIAA 98- hari. Reynolds Stress and Vorticity in Turbu-
0535, 1998. 36th Aerospace Sciences Meet- lent Wall Flows. Applied Scientific Research,
ing and Exhibit, January 12-15, 1998, Reno, 51:365-370, 1993.
Nevada.
[24] M. Rai, T. Gatski, and G. Erlebacher. Di-
[13] H. Kreplin and H. Eckelmann. Behavior of rect Simulation of Spatially Evolving Com-
the Three Fluctuating Velocity Components pressible Turbulent Boundary Layers. AIAA
in the Wall Region of a Turbulent Channel 95-0583, 1995.
Flow. Physics of Fluids, 22(7):1233-1239,
1979. [25] W.H Shutts, W.H. Hartwig, and J.E.
Weiler. Final Report of Turbulent Bound-
[14] T. Lund, X. Wu, and K. Squires. Genera- ary Layer and Skin Friction Measurements
tion of Turbulent Inflow Data for Spatially- on a Smooth, Thermally Insulated Flat Plate
Developing Boundary Layer Simulations. at Supersonic Speeds. Technical Report 364,
Journal of Computational Physics. DRL, 1955.
[26] J. Smagorinsky. Some Historical Remarks
on the Use of Nonlinear Viscosities. In
B. Galperin and S.A. Orszag, editors, Large
Eddy Simulation of Complex Engineering and
Geophysical Flows, chapter 1, pages 3-36.
Cambridge University Press, 1993.
[27] A. Smits and J.-P. Dussauge. Turbulent Shear
Layers in Supersonic Flow. American Insti-
tute of Physics, New York, 1996.
[28] E. Spyropoulos and G. Blaisdell. Large-Eddy
Simulation of a Spatially Evolving Compress-
ible Boundary Layer Flow. AIAA 97-0429,
1997.
[29] W. Yanta and B. Crapo. Applications of the
Laser Doppler Velocimeter Measure Subsonic
and Supersonic Flows. Technical Report CP-
193, AGARD, 1976.
[30] H. Zhao and P. R. Yoke. A Dyn-
amic Subgrid-Scale Model for Low-Reynolds-
Number Channel Flow. International Journal
for Numerical Methods m Fluids, 23:19-27,
1996.
0.03

0.02

75
TIME

Figure 1: Friction Velocity Figure 4: Planar Averages of Root-Mean-Square


Streamwise Velocity Fluctuations

Figure 2: Planar Averages of Time Averaged Ve-


Figure 5: Planar Averages of Root-Mean-Square
locity
Wall-Normal Velocity Fluctuations

DNS
LES
u'.y*
u'« 2.5 In y'-f 5.5 (DNS)
u'« 2.65 kiy't 5.8 (Exp)
Eck.hi.im (1874)

Figure 3: Planar Averages of Time Averaged Ve- Figure 6: Planar Averages of Root-Mean-Square
locity in Wall Units Spanwise Velocity Fluctuations

10
0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
0.5
Y

Figure 7: Planar Averages of Time-Averaged Figure 10: Planar Averages of Skewness of


Reynolds Shear Stress Streamwise Velocity Fluctuations

0.6

0.4

I 0.2

I 0

-0.4

-0.6

-0.8

0.5 0.5
Y Y

Figure 8: Planar Averages of Wall-Normal Com- Figure 11: Planar Averages of Skewness of Wall-
ponent of Streamwise Momentum Flux Normal Velocity Fluctuations

10"

10*

10''

0.5
Y

Figure 9: Planar Averages of SGS Eddy Viscosity, Figure 12: Planar Averages of Skewness of Span-
Normalized with Laminar Viscosity wise Velocity Fluctuations

11
7.5r

0.5
Y

Figure 13: Planar Averages of Flatness of Stream-


wise Velocity Fluctuations

0 0.25 0.5 0.75

Figure 16: Mean Velocity Profile

- >*^ _ _ _ _ - ^Sj,

'• , , , , 1 , . , , 1 , • . i i . • , , i
3 0-25 0.5 0.75 1
Y

Figure 14: Planar Averages of Flatness of Wall-


Normal Velocity Fluctuations
Computed
Yanta and Crapo (1976)

0.5 Figure 17: Streamwise Velocity Fluctuations


Y

Figure 15: Planar Averages of Flatness of Span-


wise Velocity Fluctuations

12

You might also like