Assessment of Turbulent Shock-Boundary Layer-Oliver
Assessment of Turbulent Shock-Boundary Layer-Oliver
Assessment of Turbulent Shock-Boundary Layer-Oliver
The performance of two popular turbulence models, the Spalart-Allmaras model and
Menter’s SST model, and one relatively new model, Olsen & Coakley’s Lag model, are eval-
uated using the OVERFLOW code. Turbulent shock-boundary layer interaction predictions
are evaluated with three different experimental datasets: a series of 2D compression ramps
at Mach 2.87, a series of 2D compression ramps at Mach 2.94, and an axisymmetric cone-
flare at Mach 11. The experimental datasets include flows with no separation, moderate
separation, and significant separation, and use several different experimental measurement
techniques (including laser doppler velocimetry (LDV), pitot-probe measurement, inclined
hot-wire probe measurement, preston tube skin friction measurement, and surface pressure
measurement). Additionally, the OVERFLOW solutions are compared to the solutions of
a second CFD code, DPLR. The predictions for weak shock-boundary layer interactions
are in reasonable agreement with the experimental data. For strong shock-boundary layer
interactions, all of the turbulence models overpredict the separation size and fail to predict
the correct skin friction recovery distribution. In most cases, surface pressure predictions
show too much upstream influence, however including the tunnel side-wall boundary layers
in the computation improves the separation predictions.
I. Introduction
A useful computational fluid dynamics (CFD) tool for the Space Shuttle and future space vehicles needs
to accurately predict the surface heat transfer over the entire vehicle, including any localized effects due to
protuberances or shock wave-boundary layer interactions (SWBLIs). The acreage flow (where the variations
in geometry are relatively gradual) is generally attached, so for these regions, the CFD code will be required
to accurately model turbulent boundary layers. Perhaps more importantly, however, the flow around protu-
berances generates shock-boundary layer interactions that can produce localized peaks in the heating rates
that may be several times larger than the acreage heating levels.
Advances in computer speeds have given rise to advances in high accuracy methods of simulating and
modeling turbulent flows. However, grid requirements typically restrict direct numerical simulations (DNS)
and even large eddy simulations (LES) to low Reynolds numbers. For complex configurations such as the
Space Shuttle, grid requirements and the need for time accurate solutions make DNS and LES impractical.
Although DNS and LES employ a more physics based representation of the fluid dynamics, the RANS
equations, along with a turbulence model, are still a valuable tool for aerodynamic analysis.
The present work continues the development of modeling high-speed compressible flows over complex
vehicles. The goal of the current project is to evaluate the performance of a set of baseline turbulence
∗ Graduate Research Assistant, Student Member AIAA
† PhD Candidate, Student Member AIAA; Aerospace Engineer, NASA/JSC, Aerosciences and CFD Branch
‡ Undergraduate Research Assistant, Student Member AIAA
§ Associate Professor, Senior Member AIAA
¶ Professor, Associate Fellow AIAA.
1 of 26
A. Flat Plates
In order to address several relevant features of a high-speed zero pressure gradient boundary layers, three
datasets have been selected for comparison. Experimental studies by Cary10 at Mach 6.0 and Mach 4.9
have been chosen for surface heating data. A DNS simulation by Gatski11, 12 provides detailed profiles of
mean and turbulent properties. Finally, experiments conducted at the Princeton Gas Dynamics Lab13, 14
at Mach 2.87 have profiles at multiple locations upstream of 8◦ and 16◦ compression corners, and will be
used for comparison of high Reynolds number experimental velocity profiles and boundary layer growth
measurements. This paper will present the results for the shock/boundary layer interaction datasets. The
zero pressure gradient results have been published previously.8, 9
B. Compression Ramps
Two supersonic compression ramp datasets with freestream Mach numbers near 2.9 have been chosen for com-
parison. The experiments of Kuntz15–17 provide two component laser doppler velocimetry (LDV) measure-
ments of turbulent boundary layers on ramps of 8◦ , 12◦ , 16◦ , 20◦ , and 24◦ , and the series of experiments13, 14
conducted at the Princeton Gas Dynamics Laboratory will be used to compare turbulence quantities down-
stream of the shock and for skin friction and surface pressure distributions on ramps of 8◦ , 16◦ , 20◦ , and
24◦ .
There are a few common problems with these datasets that should be explicitly noted. First of all,
it should be noted that the two-dimensionality assumption is a rather large assumption, and it does not
necessarily hold for the the larger ramp angles. Additionally, both of the datasets use ramp models attached
to the floor of the wind tunnel, so the turbulent boundary layers had previously experienced a strong
distortion through the nozzle. The effect this had on the turbulence is unknown. Finally, Settles and
Dodson report in their 1994 shock-boundary layer interaction database18 that these two datasets disagree
on the magnitude of the Reynolds stresses by as much as a factor of 4. To date, the authors are not aware
of an explanation for this discrepancy, so following the suggestion of Settles and Dodson, these values are
taken to be reasonable bounds of the actual Reynolds stresses.
2 of 26
1. Princeton
Several researchers conducted a series of compression corner experiments in the Princeton 20 cm × 20 cm
high Reynolds number supersonic wind tunnel. A significant number of parameters were studied in this
series of experiments; however, only a subset of this data will be considered here. A description of these
selected experiments follows.
The ramp angles considered (8◦ , 16◦ , 20◦ , and 24◦ ) provide flows that range from unseparated to sig-
nificantly separated. The 6 inch wide compression ramp models were mounted on the wind tunnel floor
with one inch of clearance on either side to permit the passage of the sidewall boundary layers. The ramp
model station was located far enough downstream of the nozzle that the turbulence in the boundary layer
appeared to have fully recovered from distortion in the nozzle. The boundary layer was nominally 26 mm
thick (θ = 1.3 mm) upstream of the ramp corner location, although the specific values varied for each case.
The freestream Reynolds number was approximately 6.3 × 107 m−1 , which gives Reθ = 7.8 × 104 , two inches
upstream of the ramp corner. The temperature of the tunnel walls and ramp models was not controlled;
however, the temperature was observed to be approximately 1.05 times the stagnation temperature (hot-wall
conditions). A summary of the case specific conditions is given in Table 1.
Settles et al.13 analyzed the mean flowfields using several methods. Pitot and static pressure probes were
used to make measurements of mean velocity in the boundary layers upstream, inside, and downstream of the
interaction. Static pressure measurements were made on the surfaces of all models, and Preston tubes were
used to measure skin friction. Surface streak and schlieren photography were used to observe separation
behavior. Two dimensionality of the flow was verified with surface streak measurements (aerodynamic
fences were necessary for the higher ramp angles to achieve two dimensionality). Interference effects due to
the various probes were assessed with schlieren photos, surface pressure measurements, and surface streak
observations.
Smits et al.14 revisited the Settles et al.13 dataset and added to it by using hot-wires to measure
fluctuating quantities. It does not appear that Smits et al. repeated any new mean flow measurements
(likely because the flow conditions were not significantly different from those used by Settles et al.) and used
the Settles et al. measurements in their data reduction process (see note on page 64 of reference19 ). Single
normal hot-wires were used to measure the normal Reynolds stresses and longitudinal mass flux fluctuations;
inclined wires were used to measure Reynolds shear stresses. Evans et al.20 instrumented the 16◦ ramp with
thin-film gages and measured heat transfer from the surface of the ramp in the recovering boundary layer.
These runs deviate from the Settles et al. flow conditions more than the Smits et al. runs, as the wall
temperature increased by 13 K (5%). The Settles et al.13 dataset and the Smits et al.14 datasets are both
listed as acceptable experiments in the Settles and Dodson database,18, 19 and the Settles et al.13 dataset is
included in the 1980-81 AFOSR-HTTM-Stanford Conference on Complex Turbulent Flows.21
2. Kuntz
Kuntz et al.15–17 conducted a series of compression corner SWBLI experiments using a non-intrusive two-
component LDV system to measure flow velocity and turbulence properties. The range of ramp angles
were chosen to span the full range of possibilities: flows with no separation, incipient separation, and
significant separation. The ramp models were located on the floor of a 10.2 cm × 10.2 cm blow-down wind
tunnel, and spanned the entire width of the tunnel test section. The Mach number in the test section was
determined to be 2.94 based on surface static pressure measurements, and two-dimensionality was determined
via observation of surface-streak patterns. Three-dimensionality was observed in the 16◦ , 20◦ , and 24◦ cases,
but the reattachment line was reportedly straight for the inner 7 cm of the 10.2 cm test section width. The
stagnation pressure was approximately 483 kPa (70 psia), and the total temperature varied from 21◦ C to
29◦ C (the room temperature of the facilities). Given the short run times (90 seconds max) and the thick
aluminum construction, the walls likely approximated isothermal walls near the stagnation temperature.22
Incoming boundary layer properties were measured using the LDV system in the tunnel without a ramp
model in place. The no-ramp setup was also used to determine the freestream conditions. These runs
3 of 26
C. Cone-Flare
One major drawback to the two-dimensional plane compression ramp is that the flow is not exactly two
dimensional; end effects are always present and their influence on the interaction is likely very complicated.
An axisymmetric cone-flare model at zero angle of attack alleviates this problem.
The dataset obtained by Holden et al.24 reports heat transfer and surface pressure measurements on two
cone-flare models in Mach 11 flow in the Calspan 96-inch shock tunnel. These models used a 9 foot (4.12 m)
long 6◦ cone to generate a turbulent boundary layer with natural transition. Flares with angles of 36◦ and
42◦ (measured from the axis of symmetry) attached to the end of the cone result in compression-corner
SWBLIs that are unseparated and separated, respectively. The freestream Reynolds number outside the
conical shock wave is Re∞ = 3.30 · 105 in−1 .
There is a fair amount of uncertainty surrounding this dataset at the time of this writing. Only one
paper has been found that describes this dataset, although the data also appears in electronic format in
the Settles and Dodson database (the database description outlines the cone-flare experiments; however,
the references it cites describe three-dimensional swept shock boundary layer interaction experiments). The
experimental boundary layer thickness is not explicitly stated in the paper (although it appears as a note
in one of the plots). The wall temperature is not stated, so room temperature was assumed. Likewise, the
exact model geometry is not provided; however, the geometry is given in the Settles and Dodson description
of the experiment. Finally, the freestream stagnation conditions are not explicitly stated and had to be
backed out from the stagnation values prior to the shock heating (the paper describes how to do this).
Despite these issues, this test case is still briefly considered in this study. The heat transfer data shows
significant heating at the reattachment point, a common and important feature of SWBLI’s that is not
addressed by the Mach 3 compression corner experiments.
4 of 26
B. Turbulence Models
OVERFLOW has numerous turbulence models coded in, but this study focuses on the two most popular
turbulence models in the OVERFLOW code: the Spalart-Allmaras (SA) model and the SST model, and also
on a promising new model: the lag model. All three of these turbulence models make use of the Boussinesq
approximation, meaning that the effects of the Reynolds stress terms are included in the Navier-Stokes
equations through an eddy viscosity. Each of the test cases is computed with all three turbulence models to
allow comparisons of the models.
C. Grid Generation
All of the 2D grids were generated using the Chimera Grid Tools (CGT) package30 and simple in-house
grid generation codes. The 3D grids used in the end-effects study were generated with CGT and Gridgen.
Double fringes were used in the overlap region interpolation stencils, and grids split simply for manual load
balancing had 5 points of perfect overlap. Table 2 summarizes the number of grid points used in each case.
5 of 26
2. Cone-Flares
The cone-flare cases use the single zone, axisymmetric grids illustrated in Figure 3. The radius of the cone
tip is assumed sharp, and an inviscid feeder block is used to permit the shock wave to develop as it may. The
outer boundary is loosely fit to the shock wave and the wall spacing is 1.0 · 10−5 in. As with the compression
corner cases, the flare ‘top’ (a cylindrical section at the end of the ramp) is included in the grid, although
this geometric feature is not discussed in the literature (and probably was not present). The ramp length
was determined based on maximum x-value shown in the data plots in reference.24
D. Boundary Conditions
The boundary conditions for the 2D compression corner grid systems are freestream (IBTYP=40) at the
inflow boundary, characteristic freestream (IBTYP=47) on the boundary opposite the viscous wall, and ex-
trapolation (IBTYP=30) at the outflow. The viscous wall is isothermal (IBTYP=7). These same conditions
were used for the 3D grids, with the tunnel floor and side walls as viscous walls and the tunnel ’ceiling’
freestream (ie: the tunnel ceiling was not modeled). The cone-flare cases use characteristic freestream (IB-
TYP=47) opposite the isothermal viscous wall (IBTYP=8). Freestream (IBTYP=40) is specified on the
inflow plane, the inviscid feeder block uses the inviscid adiabatic wall (IBTYP=1), and the outflow boundary
uses extrapolation (IBTYP=30).
6 of 26
V. Results
Results from the various test cases computed by OVERFLOW are presented in this section along with
discussion of the results. Where possible, each case uses the same non-dimensionalizing reference values used
in the literature describing the dataset. These values are tabulated in Table 3.
A. 2D Surface Pressure
Figures 5 and 6 show the surface pressure predicted by OVERFLOW for the 2D Princeton and Kuntz
compression ramps. The Kuntz dataset shows a greater tendency to separate, a likely effect of the lower
Reynolds number. In general, for the 8◦ and 16◦ ramps, only small differences exist between the three
turbulence models. Note that the apparent shift in the experimental pressures on the Kuntz 8◦ ramp is
present in the experimental data. This offset is present for all ramp angles and no effort was made to shift
this data; it was used as reported (it is discussed by Kuntz in reference17 ). Differences between the models
start to become apparent at 16◦ , and continue to grow as the ramp angle increases to 20◦ and 24◦ . The SST
model consistently shows the largest upstream influence, and the SA model shows the least. The SA model
consistently predicts the most rapid pressure recovery of the three models, and the SST model shows the
slowest. The lag model shows a tendency to rapidly recover once it reattaches, however the late reattachment
appears to result in pressure predictions bounded by the SA and SST solutions. It is interesting to note
that as the ramp angle increases, no one model does the best job of representing the profile relative to the
experimental data; for each ramp angle, the ‘best’ model changes.
The surface pressure distributions predicted on the cone-flare models are shown in Figure 8. As with the
weak compression ramp interactions, the unseparated case (36◦ flare) shows only small differences between
the three turbulence models. For the separated case, however, the SST model shows significantly larger
upstream influence than the other two models, and appears to predict the upstream influence reasonably
well for this ramp angle. The SST model also shows a significant overprediction of the peak heating on the
flare. It is really interesting that the lag model mimics the SA model so closely for both geometries of this
test case; the reason for this behavior is unknown.
Wilcox29 claims that RANS models typically show too little upstream influence and too large a wall
pressure in the separated region. This observation appears to hold for the OVERFLOW solutions of the
shallow ramp angles; however, it tends to predict too much upstream influence for the steeper ramps. It also
7 of 26
C. Comparison to DPLR
Figure 11 shows the comparison of the OVERFLOW solutions and the DPLR solutions for the Princeton
8◦ and 24◦ ramps. The DPLR and OVERFLOW implementations of the SA and SST turbulence model
shows slight differences in the prediction of the undisturbed boundary layer. They show identical trends in
8 of 26
D. 3D End Effects
Figure 12 shows the effect of adding the side-wall boundary layers to the compression corner shock boundary
layer interaction. The pseudo-surface streak lines show slightly curved separation and reattachment lines,
and indicate a very complex flow pattern near the walls. The red streamlines near the centerline show an
escape path for the flow trapped in the recirculating separation bubble. This would seem to imply that
the separation bubble could be smaller in this case than for a strictly 2D flow. Figure 13 shows the skin
friction on the centerline of the 3D runs compared to the 2D computations for each turbulence model. The
end-effects do indeed cause the separation point to move closer to the corner for all three turbulence models,
improving the prediction for each model. The reattachment prediction improves for the lag model, remains
the same for the SST model, and actually gets worse for the SA model. The surface pressures shown in
Figure 14 show similar improvements, with a rather remarkable improvement in the prediction by the lag
model. Clearly for this test case, the three-dimensional end-effects cannot be neglected. Present efforts are
directed at performing 3D computations for the Princeton dataset.
E. SWBLI Discussion
Knight et al.33 states that RANS computations for compression corner shock boundary layer interactions
are accurate for weak interactions, but significant discrepancies existed for strong interactions. While it is
difficult to define the dividing line between ‘weak’ and ‘strong’ interactions, the present results seem to agree
with this assessment. The 8◦ ramp results tend to be reasonably accurate, but the accuracy decreases as the
interactions are strengthened.
Knight et al.33 places a strong emphasis on the fact that the RANS computations fail to predict the low-
frequency shock unsteadiness;34 the OVERFLOW solutions likewise do not predict any unsteadiness. This
should be expected, however, as recent findings35 suggest that the low frequency oscillations are the result
of large streamwise turbulent structures in the upstream boundary layer. Since the RANS models average
out all of the turbulent fluctuations, including these large streamwise features, there is no physical source
for this type of shock unsteadiness. In preliminary computations, the shock waves did occasionally show an
oscillatory motion of the shock foot location that prevented convergence to a steady solution; however, efforts
to run these solutions time-accurate failed to produce consistent and repeatable results. This ‘unsteadiness’
was considered to be a numerical problem and was fixed by reducing the grid aspect ratio near separation
and by increasing the CFL number.
One more thing is interesting to note about the presented results. On the Princeton 16◦ ramp, Figure 5(b)
shows a rather accurate prediction of the surface pressure, but Figure 9 shows a very poor heat transfer
prediction. This shows that the code can accurately predict the inviscid flow field (and hence, gets the
inviscid pressure correct), but still get the viscous boundary layer wrong. Assessing the accuracy of a CFD
simulation using pressure only does not indicate anything about the accuracy of skin friction or heating
prediction.
Finally, it has been shown that the 2D assumption is indeed a questionable one for plane compression
ramp flows. It is helpful to note, however, that knowledge of the exact geometry run can help reconcile the
difference as far as model validation is concerned.
Acknowledgments
This work is supported by the NASA Johnson Space Center under Grant No. NNJ04HI12G. We would
also like to thank Tom Gatski for providing information on his DNS simulations, Dave Kuntz and Lex Smits
for providing us with documents and information on each of their experiments.
References
1 Jespersen, D. C., Pulliam, T. H., and Buning, P. G., “Recent Enhancements to OVERFLOW,” AIAA Paper No. 97-0644,
January 1997.
9 of 26
January 2005.
5 Spalart, P. R. and Allmaras, S. R., “A One-Equation Turbulence Model for Aerodynamic Flows,” AIAA Paper No.
and Shock-Boundary Layer Interaction Computations With the OVERFLOW Code,” AIAA Paper No. 2006-0894, January
2006.
9 Oliver, A. B., Evaluation of Two-Dimensional High-Speed Turbulent Boundary Layer and Shock-Boundary Layer Inter-
action Computations With the OVERFLOW Code, MS Thesis, Purdue University, West Lafayette, IN, August 2006.
10 Cary, A. M., Turbulent Boundary Layer Heat Transfer and Transition Measurements With Extreme Surface Cooling in
Tech. Rep. NASA TM 211934, NASA Langley Research Center, Hampton, VA, September 2002.
12 Pirozzoli, S., Grasso, F., and Gatski, T. B., “Direct Numerical Simulation and Analysis of a Spatially Evolving Supersonic
Turbulent Boundary Layer at M=2.25,” Physics of Fluids, Vol. 16, No. 3, March 2004, pp. 530–545.
13 Settles, G. S., Fitzpatrick, T. J., and Bogdonoff, S. M., “Detailed Study of Attached and Separated Compression Corner
Flowfields in High Reynolds Number Supersonic Flow,” AIAA Journal, Vol. 17, No. 6, June 1979, pp. 579–585.
14 Smits, A. and Muck, K., “Experimental Study of Three Shock Wave/Turbulent Boundary Layer Interactions,” Journal
Wave/Boundary-Layer Interaction,” AIAA Journal, Vol. 25, May 1987, pp. 668–675.
16 Kuntz, D., Amatucci, V. A., and Addy, A. L., “The Turbulent Boundary Layer Properties Downstream of the Shock-
Tech. Rep. 177638, Pennsylvania State University, University Park, PA, April 1994, NASA Contractor Report.
19 Settles, G. S. and Dodson, L. J., “Hypersonic Shock/Boundary-Layer Interaction Database,” Tech. Rep. 177577, Penn-
sylvania State University, University Park, PA, April 1991, NASA Contractor Report.
20 Evans, T. T. and Smits, A. J., “Measurements of the Mean Heat Transfer in a Shock Wave - Turbulent Boundary Layer
Interaction,” Experimental Thermal and Fluid Science, , No. 12, 1996, pp. 87–97.
21 Kline, S. J., editor, The 1980-81 AFOSR-HTTM-Stanford conference on complex turbulent flows: Comparison of com-
turbulence model for application in hypersonic flows,” AIAA Paper No. 1997-2023, June 1997.
28 Olsen, M. E., Lillard, R. P., and Coakley, T. J., “The Lag Model Applied to High Speed Flows,” AIAA Paper No.
Boundary Layer Interactions,” Progress in Aerospace Sciences, Vol. 39, 2003, pp. 121–184.
34 Dolling, D. S., “50 Years of Shock Wave/Boundary Layer Interaction - What Next?” AIAA Paper No. 2000-2596, June
2000.
10 of 26
structure of upstream turbulent boundary layer on shock induced separation,” AIAA Paper No. 2006-0324, January 2006.
Table 1. Freestream and inflow boundary layer data for compression ramp cases
11 of 26
12 of 26
Princeton 8◦ 123x167 182x144 181x144 358x28 250x39 120x24 95.5E+3
Princeton 16◦ 123x167 180x144 180x144 355x28 240x39 120x24 94.6E+3
Princeton 20◦ 153x208 177x144 130x144 302x35 221x39 45x19 96.1E+3
Princeton 24◦ 153x208 176x144 155x144 326x35 221x39 - 99.5E+3
Cone-Flare 36◦ 1452x363 - - - - - 527.1E+3
Cone-Flare 42◦ 1451x363 - - - - - 526.7E+3
Case δRef [m] MRef VRef [m/s] PRef [Pa] ρRef [kg/m3 ]
Princeton Ramps:
8◦ 0.026 2.87 591.5 22856 -
◦
16 0.026 2.85 577.2 23561 -
20◦ 0.025 2.85 566.3 23561 -
◦
24 0.023 2.84 569.9 23922 -
Kuntz Ramps:
13 of 26
8◦ 0.00827 2.94 611.8 14391 -
16◦ 0.00827 2.94 621.0 14391 -
◦
20 0.00827 2.94 621.0 14391 -
24◦ 0.00827 2.94 620.0 14391 -
Cone-Flares:
36◦ 0.01778 10.96 1805 - 0.0327
(b)
Figure 1. General grid topology for the 2D compression ramp cases. (a) Entire domain, (b) Detail of corner
region.
14 of 26
15 of 26
(b)
(c)
Figure 3. General grid topology for the cone-flare cases. (a) Entire domain, (b) Detail of tip region, (c) Detail
of corner region, colored by static pressure (color scale adjusted to show shock locations).
16 of 26
17 of 26
18 of 26
Figure 5. Surface pressure distributions for Princeton series. (a) 8◦ ramp, (b) 16◦ ramp, (c) 20◦ ramp, (d) 24◦ ramp.
(a) (b)
19 of 26
Figure 6. Surface pressure distributions for Kuntz series. (a) 8◦ ramp, (b) 16◦ ramp, (c) 20◦ ramp, (d) 24◦ ramp.
(a) (b)
20 of 26
Figure 7. Skin friction distribution for Princeton series. (a) 8◦ ramp, (b) 16◦ ramp, (c) 20◦ ramp, (d) 24◦ ramp.
(a)
(b)
Figure 8. Surface pressure distribution on cone-flares. (a) 36◦ flare, (b) 42◦ flare.
21 of 26
22 of 26
(b)
Figure 10. Heat transfer distribution on cone-flares. (a) 36◦ flare, (b) 42◦ flare.
23 of 26
(b)
Figure 11. Skin friction predicted by DPLR of (a) Princeton 8◦ ramp, (b) Princeton 24◦ ramp.
24 of 26
25 of 26
Figure 14. Surface pressure distribution along centerline of Kuntz 24◦ ramp.
26 of 26