You are on page 1of 12

48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition AIAA 2010-921

4 - 7 January 2010, Orlando, Florida

A DES Procedure Applied to the Flow Over a


NACA0012 Airfoil
Radoslav Bozinoski∗ and Roger L. Davis†
University of California at Davis, Davis, California, 95616

This work describes a detached-eddy simulation (DES) for massively separated flow
over a NACA 0012 airfoil at off design conditions. Simulations were performed using the
MBFLO1, 2 simulation procedure. The Navier-Stokes equations are solved using an explicit,
multi-grid procedure for the steady-flow Reynolds-averaged (RANS) computations. A
dual time-step procedure is used for unsteady Reynolds-averaged (URANS) or DES. A
two-equation k − ω turbulence model is used for the entire turbulence field in the RANS
and URANS simulations as well as in the near-wall regions of the DES. The DES results
are obtained using the model presented by Bush and Mani3 and are compared with the
unsteady Reynolds-averaged Navier-Stokes solutions and experimental data for the NACA
0012 airfoil at an angle of attack of 60◦ . The results of the two- and three-dimensional
computations are compared to each other as well as to experimental results. Differences
between two- and three-dimensional computations are observed in terms of the coeffiients
of lift and drag as well as vorticity and entropy contours. The two-dimensional DES showed
an improvement in lift and drag predictions when compared to two- and three-dimensional
URANS, however, the three-dimensional DES lift and drag are closer to the experimental
results in the literature as expected. A method accounting for the three-dimensional DES
effects in two-dimensional DES is also explored.

Nomenclature
E total energy
e internal energy
H total enthalpy
h static enthalpy
k turbulent kinetic energy
p pressure
Pr Prandtl number
P rt turbulent Prandtl number
Sij mean strain-rate tensor
ui velocity component
V velocity magnitude
μ coefficient of viscosity
μT turbulent coefficient of viscosity
ω turbulent dissipation frequency
τ̂ij total shear stress tensor
τij laminar shear stress tensor
τij
R
Reynolds shear stress tensor
ρ density
ldis dissipation length scale
Cdes DES coefficients

∗ Graduate Student, Mechanical and Aeronautical Engineering Department, Student Member AIAA.
† Professor, Mechanical and Aeronautical Engineering Department, Associate Fellow AIAA.

1 of 12

American
Copyright © 2010 by Radoslav Bozinoski & Roger L. Davis. Published by Institute of Aeronautics
the American and Astronautics
Institute of Aeronautics and Astronautics, Inc., with permission.
I. Introduction

S imulation of “steady” and “unsteady” flow of aerodynamic bodies has matured a great deal over the past
decade. Aerodynamic performance and flow structures can be predicted with acceptable accuracy except
in the complex flow regions of mixing and at off-design conditions near stall. In these regions, complex
flow structures with multiple eddies that mingle and mix are often not predicted well due to inadequate
computational grid density and a breakdown of turbulence models.

II. Governing Equations


The unsteady, Favre-averaged governing flow-field equations for an ideal, compressible gas in the right-
handed, Cartesian coordinate system using primary variables are used in the MBFLO code. The three-
dimensional continuity, momentum, and energy equations can be written in a conservative form as follows,

∂ρ ∂ (ρuj )
+ =0 (1)
∂t ∂xj
∂ρui ∂ (ρuj ui ) ∂P ∂ τ̂ij
+ =− + (2)
∂t ∂xj ∂xi ∂j
   
∂E ∂ (ρuj H) ∂ μ μT ∂h
+ = ui τ̂ij + + (3)
∂t ∂xj ∂xj P r P rt ∂xj

τ̂ij = τij + τij


R
= (μ + μT ) Sij (4)
 
1 ∂ui ∂uj 2 ∂uk
Sij = + − δij (5)
2 ∂xj ∂xi 3 ∂xk
Since we are dealing with compressible flows, we require an equation of state to relate the energy with
pressure and enthalpy.
 
1 2
E =ρ e+ V (6)
2
1 γ P 1 E+P
H =h+ V2 = + V2 = (7)
2 γ−1 ρ 2 ρ
Additional governing equations, as developed by Wilcox4–6 , are solved for the transport of turbulent
kinetic energy and turbulence dissipation rate in regions of the flow where the computational grid or global
time-step size cannot resolve the turbulent eddies. In regions of the flow where the larger-scale eddies can be
resolved with the computational grid, techniques borrowed from large-eddy simulation are used to represent
the viscous shear and turbulent viscosity. The large-eddy sub-grid model described by Smagorinsky7 is
modified according to the detached-eddy considerations described by Strelets8 and Bush and Mani.3

III. Numerical Techniques


The conservation of mass, momentum, and energy equations are solved using a Lax-Wendroff control-
volume, time-marching scheme as developed by Ni,9 Dannenhoffer,10 and Davis.11, 12 Numerical solutions of
unsteady flows can be performed with either the explicit9 or a dual time-step procedure.13 These techniques
are second-order accurate in time and space. A multiple-grid convergence acceleration scheme9 is used for
steady, Reynolds-averaged solutions and the inner convergence loop of the unsteady simulations using the
dual time-step scheme. The approach is called MBFLO and has two-dimensional,2 axi-symmetric14 (with
and without swirl), and three-dimensional1 versions. The two- and three-dimensional procedures, MBFLO2P
and MBFLO3P (two/three-dimensional, parallel), and results for a DES for flow over a NACA 0012 airfoil
at an angle-of-attack of 60◦ are described here.
It is recognized that two-dimensional DES lack the three-dimensional effects resulting from the interaction
of eddies in the third dimension15 and therefore, tend to under-predict diffusion. However, two-dimensional

2 of 12

American Institute of Aeronautics and Astronautics


design is an important part of the overall design process of wings, turbomachinery blades, and other similar
lifting bodies so that a robust two-dimensional simulation capability is highly desirable. The hope is to
compare the two-dimensional and three-dimensional URANS and DES and to possibly account for the three-
dimensional eddy-interaction effects through additional modeling in the future. One approach for modeling
these differences is introduced and discussed below.
The combined second- and fourth-difference dissipation model of
Jameson16 is used in the current procedure for both the mean flow 
and turbulence equations. The fourth-difference dissipation is scaled  

 


by the inverse of the absolute value of the mean strain rate squared.  




 
This function decays the numerical dissipation in all viscous flow 



regions, including boundary layers, wakes, large eddies, secondary 
flows, etc. 
In the MBFLO suite of codes, parallelization is performed using 
the Message Passing Interface (MPI) library.17 Figure 1 and Table 1 
show the typical speed-up and associated efficiencies as functions of 
         
the number of processors for the MBFLO3P code. The configuration 
 




used to generate this data was similar to that shown below in the
Figure 1: Typical Speed-Up Factor
results section where the computational grid consisted of 1.80 million
as a function of Number of Processes
grid points. The data was generated on a Linux cluster consisting
of 3.6GHz Intel Xeon processors. Figure 1 shows that a speed-up
factor of 18.06 is realized with 20 processes yielding a 90% parallel efficiency. In Table 1 we can see that
efficiencies of 90% and higher can be obtained if no less than 100, 000 grid points per process are used.

Table 1: Typical Speed-Up and Efficiencies


Processes Speed Up Efficiency Pts/Process
1 1.00 100% 1,780,440
2 2.00 100% 892,440
4 4.00 100% 448,440
8 7.79 97% 226,440
16 14.60 91% 115,440
20 18.06 90% 93,240
40 31.39 78% 48,840

IV. Results and Discussion


To determine the solution capability of the URANS and DES procedures for massively separated flows,
the flow over a NACA 0012 airfoil was performed. These simulations were run at an angle-of-attack of 60◦ ,
a freestream Mach number of 0.2, and a Reynolds number of 100, 000 with a freestream total temperature
set to 463.68◦ R. Both two- and three-dimensional simulations were performed to assess the importance of
three-dimensional flow structures over two-dimensional geometries. The dual time-step procedure for the
unsteady URANS and DES was utilized with a global time-step of 1.0×10−4 seconds. For the DES procedure,
a CDES 1 coefficient of 0.65 was used. Figure 2 shows the computational grids used in these simulations.
The two-dimensional grid consisted of 74, 185 points with the far field approximately 50 chords away. For
the three-dimensional simulations the two-dimensional grid was used and extruded one half chord into the
span-wise direction resulting in a 1, 854, 625 point grid. Using Table 1, it was determined that the three-
dimensional grid could be decomposed into 20 blocks to operate MBFLO3P at 90% parallel computational
efficiency.
The temporal periodicity and unsteady behavior was initially studied using the two-dimensional URANS
procedure. The information obtained was then used to determine the number of time steps necessary to
resolve a minimum of 10 periodic cycles for the subsequent two- and three-dimensional URANS simulation
and DES at the given global time-step. Figures 3(a) and 3(c) show the signal history for the instantaneous
lift, CL , and drag, CD , coefficients as well as the Power Spectral Density (PSD) as a function of the number

3 of 12

American Institute of Aeronautics and Astronautics


(a) Stream-wise view (b) Span-wise view

Figure 2: NACA0012 computational mesh (every fourth point shown).

of time-steps. The PSD of lift and drag was plotted here as a function of the number of time-steps to more
easily see time-periodicity in the flow. In Fig. 3(c), we see that the dominant signal is repeated approximately
every 260 time-steps, which corresponds to a frequency of 39 Hz. This frequency was used to determine the
period over which the solution was time-averaged. Once the period was determined, this case was run for a
minimum of 20 periodic cycles and time-averaged. Using the information obtained from the two-dimensional
URANS simulation, a two-dimensional DES was performed. Figure 3(b) shows the signal histories for the
two-dimensional DES. Here, we notice that the CL and CD signals contain signal power at more frequencies
than their URANS counterparts, which is also reflected in Figs. 3(d) and 3(f). However, the PSD shows
a dominant signal repeating every 300 time-steps. Using this information the two-dimensional DES was
time-averaged for 16, 000 time-steps, which is well over 20 periodic cycles. A dominant PSD frequency of 38
Hz was noticed for the two-dimensional DES.
Figure 4 below shows the corresponding three-dimensional CL and CD signal histories as well as their
PSD’s for URANS and DES. In Fig. 4(a), we can see the periodic lift and drag signals for the URANS
case. Due to the high dissipation of the turbulence model, there was no significant three dimensionality
observed and the solution had identical streamwise contours in the spanwise direction. It was reported3
that no three-dimensionality would be noticed for the NACA0012 airfoil case unless the spanwise domain
is extended past two chord lengths. Figure 4(c) shows that the dominant power signal for the URANS was
repeated every 300 time-steps corresponding to a frequency of 35 Hz. The lift and drag signals for the two-
and three-dimensional URANS simulations produced very similar PSD plots with dominant frequencies at
approximately 30 and 60 Hz shown in Figs. 3(e) and 4(e).
The DES lift and drag signals, shown in Fig. 4(b), for the three-dimensional simulation showed signal
power at more frequencies and reflected the three dimensionality of the flowfield. In Fig. 4(d), we can
see that there are three dominant signals repeating every 100, 230, and 420 time-steps, and these can be
attributed to the three-dimensional flowfield in the separation region on the suction side of the airfoil. It
should be noted that a Hanning window function was used to ensure no artifical frequencies are present due
to signal clipping. When comparing the PSDs in Figs. 3(f) and 4(f), we can see that both the two- and
three-dimensional DES have a larger range of frequencies present in the lift and drag signals compared to
URANS. An interesting observation when comparing the PSD for the URANS and DES is that the dominant
power signals for the lift and drag are at approximately the same frequency. The two-dimensional DES has
a dominant lift and drag power signal at 38 Hz and the three-dimensional DES at 42 Hz.
Based on the results found by analyzing the lift and drag signals for the two- and three-dimensional
URANS and DES, the total number of unsteady time-steps for each case was chosen to allow for the time-
averaging of a minimum of 10 periodic cycles. The global time-step was also set so that the dominant
frequencies being resolved would be within the 35 to 45 Hz range. Table 2 shows a comparison of the time
averaged CL and CD coefficients against experimentally measures values. Here we can see that the two- and
three-dimensional URANS as well as the two-dimensional DES predicted similar results and over predicted
both CL and CD . The two-dimensional URANS predicted CL and CD values of 1.33 and 2.25, respectively.
The two-dimensional DES predicted CL and CD values of 1.25 and 2.12, respectively, and did show a slight
improvement over the URANS procedure at the given time-step. The time-averaged three-dimensional CL

4 of 12

American Institute of Aeronautics and Astronautics


Avg. CL = 1.33, CD = 2.25 Avg. CL = 1.25, CD = 2.12
2.0 2.5
1.8 2.0
1.6
1.5
CL

CL
1.4
1.0
1.2
1.0 0.5
0.8 0.0
3.0 4.0
2.8 3.5
2.6 3.0
2.4 2.5
CD

CD
2.0
2.2 1.5
2.0 1.0
1.8 0.5
1.6 0.0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
Seconds Seconds

(a) URANS Signals (b) DES Signals

×106 ×106
2.0 1.2
1.0
1.5
0.8
P SDCL

P SDCL

1.0 0.6
0.4
0.5
0.2
0.0 0.0
×106 ×105
2.5 8
7
2.0 6
P SDCD

P SDCD

1.5 5
4
1.0 3
0.5 2
1
0.0 0
0 100 200 300 400 500 0 100 200 300 400 500
Period [iterations] Period [iterations]

(c) URANS PSD (d) DES PSD

×106 ×106
2.0 1.2
1.0
1.5
0.8
P SDCL

P SDCL

1.0 0.6
0.4
0.5
0.2
0.0 0.0
×106 ×105
2.5 8
7
2.0 6
P SDCD

P SDCD

1.5 5
4
1.0 3
0.5 2
1
0.0 0
0 20 40 60 80 100 0 20 40 60 80 100
Frequency [Hz] Frequency [Hz]

(e) URANS PSD (Hz) (f) DES PSD (Hz)

Figure 3: NACA0012 lift and drag history for two-dimensional simulations.

5 of 12

American Institute of Aeronautics and Astronautics


Avg. CL = 1.40, CD = 2.37 Avg. CL = 1.09, CD = 1.89
1.8
1.6
2.0
1.4
1.2
CL

CL
1.5
1.0
0.8
1.0
0.6
0.4
4.0
3.5 2.5
3.0
2.0
CD

CD
2.5
2.0 1.5
1.5
1.0 1.0
0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Seconds Seconds
(a) URANS Signals (b) DES Signals

×105 ×104
3.0 2.5
2.5 2.0
2.0
P SDCL

P SDCL
1.5
1.5
1.0
1.0
0.5 0.5
0.0 0.0
×105 ×104
4.0 3.0
3.5 2.5
3.0
2.0
P SDCD

P SDCD

2.5
2.0 1.5
1.5 1.0
1.0
0.5 0.5
0.0 0.0
0 100 200 300 400 500 0 100 200 300 400 500
Period [iterations] Period [iterations]

(c) URANS PSD (d) DES PSD

×105 ×104
3.0 2.5
2.5 2.0
2.0
P SDCL

P SDCL

1.5
1.5
1.0
1.0
0.5 0.5
0.0 0.0
×105 ×104
4.0 3.0
3.5 2.5
3.0
2.0
P SDCD

P SDCD

2.5
2.0 1.5
1.5 1.0
1.0
0.5 0.5
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
Frequency [Hz] Frequency [Hz]

(e) URANS PSD (Hz) (f) DES PSD (Hz)

Figure 4: NACA0012 lift and drag history for three-dimensional simulations.

6 of 12

American Institute of Aeronautics and Astronautics


and CD predictions for the URANS procedure were higher than the two-dimensional URANS with values at
1.40 and 2.37, respectively. Similar observation were reported by Bush et al.3 and this may be due to the
relatively coarse spanwise spatial resolution. The three-dimensional time averaged DES predicted a CL of
1.09 and CD of 1.89 and matched the experimental CL and CD values much more closely.

Table 2: NACA0012 CL and CD comparisons

CL CD
2D URANS 1.33 2.25
2D DES 1.25 2.12
3D URANS 1.40 2.37
3D DES 1.09 1.89
Experiment 0.90 1.60

A comparison of the instantaneous two- and three-dimensional vorticity contours for the URANS and
DES can be seen in Figs. 5 and 6. The three-dimensional plots for both URANS and DES have been
spatially averaged over the span. In Figs. 5(a) and 6(a), we can see the larger vortical structures with a
greater level of dissipation when compared to Figs. 5(b) and 6(b). We can also see a larger range of scales
in the DES vorticity contours when compared to the URANS simulations. Similar results can be seen in the
entropy contours in Figs. 7 and 8.

(a) 2D URANS (b) 2D DES

Figure 5: NACA0012 instantaneous vorticity contours.

Instantaneous vorticity contour comparison between three-dimensional URANS and DES is shown in Fig.
9. Here, we see span-wise vorticity contour slices of the flow domain at y/c = 0.15, 0.35, and 0.55 normal to
the surface of the NACA0012 airfoil. As expected, the 3D URANS procedure produced a two-dimensional
flow structure with no noticeable variation in the spanwise direction. The 3D DES procedure, however, starts
to show three-dimensional vortical structures close to the surface at y/c = 0.15 that continue throughout
the wake. To further highlight the three-dimensionality in the flow domain of the DES, Fig. 10 shows a
iso-surface contour plot of the instantaneous vorticity color by pressure.
As previously mentioned, two-dimensional simulations are considered a valuable tool during design. How-
ever, these 2D simulations lack the three-dimensional flow effects as previously discussed. A technique to
include local 3D effects in 2D simulations would be very useful for improving two-dimensional simulation
accuracy. There are various ways to model three-dimensional effects in two-dimensional calculations. Crude
techniques include modeling aerodynamic geometry with quasi-two-dimensional extensions using streamtube
height or area variation. Another approach could be the use of source terms similar to deterministic stress
models18, 19 used previously to model unsteady effects in steady simulations. For instance, let T (u) be the
time-averaged solution from a URANS or DES simulation and S(u) be the steady solution obtained from a

7 of 12

American Institute of Aeronautics and Astronautics


(a) 3D URANS (b) 3D DES

Figure 6: NACA0012 instantaneous vorticity contours.

(a) 2D URANS (b) 2D DES

Figure 7: NACA0012 instantaneous entropy contours.

(a) 3D URANS (b) 3D DES

Figure 8: NACA0012 instantaneous entropy contours.

8 of 12

American Institute of Aeronautics and Astronautics


(a) y/c = 0.15 (URANS) (b) y/c = 0.35 (URANS) (c) y/c = 0.55 (URANS)

(d) y/c = 0.15 (DES) (e) y/c = 0.35 (DES) (f) y/c = 0.55 (DES)

Figure 9: NACA0012 spanewise vorticity contours

Figure 10: NACA0012 airfoil non-dimensional vorticity isosurface (colored by pressure) for instantaneous
DES.

9 of 12

American Institute of Aeronautics and Astronautics


RANS simulation that satisfies the conservation equations Ds (u) = 0. A discrete deterministic stress model
for unsteady effects in a steady simulation can be generated by initializing the flow field with T (u), running
one iteration in the steady RANS simulation, Ds (T (u)), and capturing the residual, Rs (u). The residual
can be thought of as a deterministic source term that is required to drive the steady procedure to produce
the time-averaged solution.
Now consider using a similar approach to drive a two-dimensional, unsteady procedure to produce a
three-dimensional, unsteady solution. Complete temporal similitude would require a discrete source term
for every time-step in the simulation. This is a considerable task and outside the scope of what is considered
an engineering model. However, it would be feasible to model the three-dimensional, unsteady effects in a
two-dimensional procedure with time-invariant source terms determined from the time-averaged solutions.
A discrete model for the three-dimensional effects could be computed using the following steps:

1. Initialize the flow field of the unsteady, two-dimensional simulation with the three-dimensional locally,
spanwise-averaged, time-averaged flow field, T3D (u). Note that if the three-dimensional flow field is
periodic in the spanwise direction, then the spanwise-average, time-averaged flow field should not have
any spanwise velocity component. This is the case for the simulations shown here. However, in general,
it would not be the case and the spanwise velocities would need to be neglected.

2. Run one iteration of the unsteady, two-dimensional simulation, Dus (T3D (u)), and capture the residual,
R̄us−3D (u). This residual represents the discrete deterministic stresses necessary to drive the two-
dimensional unsteady solution to the three-dimensional time-averaged solution. The residuals are
essentially the time-average of the instantaneous deterministic stresses:


R̄us−3D (u) = Rus−3D (u)/T (8)

3. We are interested in determining the time-invariant deterministic stresses that when applied to the
unsteady, two-dimensional simulation and then time-averaged, produces the three-dimensional time-
averaged solution instead of the two-dimensional time-averaged solution. In addition, the instantaneous
two-dimensional simulation with these additional modeled 3D effects would also produce a solution
similar to the locally spanwise averaged three-dimensional instantaneous solution. An approximation
to these discrete unsteady source terms can be found by assuming that the three-dimensional effects per
time-step are time-invariant and equal to the time-averaged deterministic stress, R̄us−3D (u). Numerical

results, however, have shown that the instantaneous deterministic stress, Rus−3D (u), applied to each
time-step should be a fraction of the time-averaged value, R̄us−3D (u). As shown below, a value of

Rus−3D (u) = Δt/T R̄us−3D (u) works reasonably well.

Figure 11 shows the discrete time-averaged three-dimensional deterministic stresses that arise using this
technique with the previously described detached-eddy simulations. A Δt/T fraction of these discrete 3D
effects are subtracted as source terms from the left-hand side of the unsteady, two-dimensional Navier-Stokes
equations. Ultimately, the development of an analytical model for these effects, either in the form of source
terms for each equation or as additional stresses in the momentum and energy equations, is the goal of this
research. The results shown below demonstrate that the use of this type of modeling approach for 3D effects
in two-dimensional simulations is viable.

(a) Continuity (b) Energy (c) X-Momentum (d) Y-Momentum

Figure 11: Discrete three-dimensional effects.

10 of 12

American Institute of Aeronautics and Astronautics


Figures 12 and 13 show a comparison between the two-dimensional RANS, two-dimensional time-averaged
DES, and three-dimensional time- and space-averaged DES for entropy and vorticity. The 3D time-averaged
DES entropy and vorticity contours are more similar to the 2D steady RANS contours than the 2D time-
averaged DES contours. The next step is to determine if the differences between the 3D time/spanwise-
averaged DES and 2D time-averaged DES can be modeled using the deterministic stress approach described
above.

(a) 2D RANS (b) 2D TA-DES (c) 3D TA-DES

Figure 12: Time-averaged and RANS entropy contours.

(a) 2D RANS (b) 2D TA-DES (c) 3D TA-DES

Figure 13: Time-averaged and RANS vorticity contours.

Figure 14 shows the results of applying the discrete 3D effect source terms, shown in Fig. 11, on the
instantaneous and time-average 2D DES. We can see that by adding the source terms, both the time-
averaged vorticity and entropy contours more closely resemble the time-averaged 3D DES. Further research
and numerical experimentation is required, however, to determine an appropriate analytical model for local
3D effects based on the discrete source terms resulting from many simulations.

(a) Instantaneous entropy (b) Time-averaged entropy (c) Instantaneous vorticity (d) Time-averaged vorticity

Figure 14: NACA0012 airfoil 2D DES contours with modeled 3D effects.

11 of 12

American Institute of Aeronautics and Astronautics


V. Conclusion
A general Reynolds-averaged/detached-eddy simulation procedure was used to predict the flow structure
over a NACA0012 airfoil at 60◦ . As expected, the three-dimensional DES procedure produced the best
results when compared to the experimental values for CL and CD . With the use of the model described
by Bush et al.,3 the detached-eddy simulation showed significant improvement over the Reynold’s-averaged
Navier-Stokes simulation. This can be attributed to the DES model’s ability to better capture the flows
inherent three-dimensionality and to resolve the smaller vortical structures. The two- and three-dimensional
URANS, as well as the two-dimensional DES, over-predicted both the lift and drag with the two-dimensional
DES showing no improvement when compared to the URANS procedures. The two-dimensional DES did,
however, capture the smaller vortical structures and showed less dissipation in the wake of the NACA0012
airfoil. Finally, a new approach for modeling the effects of local three-dimensional flow in two-dimensional
simulations has also been introduced and discussed.

Acknowledgments
The authors would like to thank Dr. John Clark and the managers of the turbine branch at the Wright-
Patterson Air Force Research Laboratory in Dayton, Ohio for their support of this effort under contract
09-S590-0009-20-C1. We would also like to thank the managers of Pratt & Whitney Rocketdyne for donating
the computer resources used to conduct these simulations.

References
1 Bozinoski, R. and Davis, R. L., “General Three-Dimensional, Multi-Block, Parallel Turbulent Navier-Stokes Procedure,”

AIAA Paper No. 2008-756 , January 2008, 46th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV.
2 Davis, R. L. and Dannenhoffer, J. F., “A Detached-Eddy Simulation Procedure Targeted for Design,” AIAA Paper No.

2008-534 , January 2008, 46th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV.
3 Bush, R. H. and Mani, M., “A Two-Equation Large Eddy Stress Model for High Sub-Grid Shear,” AIAA Paper No.

2001-2561 , June 2001, 15th AIAA Computational Fluid Dynamics Conference, Anaheim, CA.
4 Wilcox, D. C., Turbulence Modeling for CFD, DCW Industries, Inc., 3rd ed., Nov. 2006.
5 Wilcox, D. C., “Formulation of the k-w Turbulence Model Revisited,” AIAA Journal, Vol. 46, No. 11, November 2008,

pp. 2823–2838.
6 Wilcox, D. C., “Reassessment of the Scale-Determining Equation for Advanced Turbulence Models,” AIAA Journal,

Vol. 26, No. 11, Nov. 1988, pp. 1299–1310.


7 Smagorinsky, J., “General circulation experiments with the primitive equations,” Mon. Weather Review , Vol. 91, No. 3,

March 1963, pp. 99–164.


8 Strelets, M., “Detached Eddy Simulation of Massively Separated Flows,” AIAA Paper No. 2001-879 , Jan. 2001, 39th

Aerospace Sciences Meeting and Exhibit, Reno, NV.


9 Ni, R., “A Multiple-Grid Scheme for Solving the Euler Equations,” AIAA, Vol. 20, No. 11, Nov. 1982, pp. 1565–1571.
10 Dannenhoffer, J. F., Grid Adaptation for Complex Two-Dimensional Transonic Flows, Ph.D. thesis, Massachusetts

Institute of Technology, Aug. 1987.


11 Davis, R. L., “Cascade Viscous Flow Analysis Using The Navier-Stokes Equations,” AIAA Journal of Propulsion and

Power , Vol. 3, No. 5, Sep.-Oct. 1987, pp. 406–414.


12 Davis, R. L., Hobbs, D. E., and Weingold, H. D., “Prediction of Compressor Cascade Performace Using a Navier-Stokes

Technique,” ASME Journal of Turbomachinery, Vol. 110, No. 4, June 1988, pp. 520–531.
13 Jameson, A., “Time Dependent Calculations Using Multigrid with Applications to Unsteady Flows Past Airfoils with

Wings,” AIAA Paper No. 1991-1596 , June 1991, pp. 14, 10th Computational Fluid Dynamics Conference, Honolulu, HI.
14 Andrade, A. J., Davis, R. L., and Havstad, M. A., “A RANS/DES Numerical Procedure for Axisymmetric Flows with

and without Strong Rotation,” AIAA Paper No. 2008-702 , January 2008, 46th AIAA Aerospace Sciences Meeting and Exhibit,
Reno, NV.
15 Spalart, P., “Young Person’s Guide to Detached-Eddy Simulation Grids,” Tech. Rep. CR-2001,211032, NASA, 2001.
16 Jameson, A., Schmidt, W., and Turkel, E., “Numerical solution of the Euler equations by finite volume methods using

Runge Kutta time stepping schemes,” AIAA Paper No. 1981-1259 , June 1981, 14th Fluid and Plasma Dynamics Conference,
Palo Alto, CA,.
17 Gropp, W., Lusk, E., and Skjellum, A., MPI: A Message-Passing Interface Standard, Scientific and Engineering Com-

putation Series, The MIT Press, 1994.


18 Busby, J., Sondak, D. L., Staubach, B., and Davis, R. L., “Deterministic Stress Modeling of Hot Gas Segregation in a

Turbine,” ASME , Vol. 122, No. 1, January 2000, pp. 62–67.


19 Sondak, D. L., Dorney, D. J., and Davis, R. L., “Modeling Turbomachinery Unsteadiness with Lumped Deterministic

Stresses,” AIAA Paper No. 1996-2570 , 1996, 32nd ASME, SAE, and ASEE, Joint Propulsion Conference and Exhibit, Lake
Buena Vista, FL.

12 of 12

American Institute of Aeronautics and Astronautics

You might also like