You are on page 1of 22

RESEARCH ARTI CLE

Three-dimensional features of a Mach 2.1 shock/boundary layer


interaction
D. B. Helmer

L. M. Campo

J. K. Eaton
Received: 13 January 2012 / Revised: 23 July 2012 / Accepted: 6 August 2012 / Published online: 29 August 2012
Springer-Verlag 2012
Abstract 2D particle image velocimetry was used to study
the three-dimensionality of the shock-boundary layer inter-
action generated by a small 20 compression ramp in a low
aspect ratio continuously operated wind tunnel. High-resolu-
tion data were taken in four streamwise-wallnormal planes:
three planes located in the sidewall boundary layer and one
near the tunnel centerline. The incoming boundary layer was
found to show three-dimensionality, with signicant over-
shoot in the velocity proles observed near the sidewall. The
size of the wedge inuenced the interaction, which was
weaker than that observed in the case of a large compression
wedge. The ow turning angle was &8 near the tunnel
centerline and changed signicantly across the span. Mea-
surements behind the compression wedge in the centerline
plane showed that both velocity and turbulence properties
were nearly fully recovered &14d behind the compression
corner. The shock angle varied with spanwise position, and a
multi-shock structure was observed in the sidewall planes.
The size of the interaction decreased in the sidewall boundary
layer. Non-monotonic variations in both velocity and turbu-
lence proles across the sidewall planes suggest the presence
of signicant spanwise ows, possibly corner vortices.
1 Introduction
Due to its complexity and real-world signicance, the
shock-boundary layer interaction (SBLI) has been a topic
of interest for decades. This phenomenon has practical
relevance to a variety of problems, including re-entry
vehicle and supersonic transport design. The interaction is
highly three-dimensional with considerable unsteadiness
and a wide range of frequency content. Shock-boundary
layer systems pose signicant challenges to both experi-
mental and numerical investigations.
The improvement of existing models is a necessary step
in order to effectively use computational uid dynamics
(CFD) in the design process for SBLI systems. While direct
numerical simulation (DNS) can be used to analyze some
features of the SBLI (e.g. Wu and Martin 2008; Pirozzoli
and Grasso 2006), the computational cost effectively limits
its application to low Reynolds numbers and renders it
impractical for use in a design cycle. Reynolds Averaged
Navier-Stokes (RANS) calculations are much less expen-
sive, but are generally not adequately validated for use in
design (Knight and Degrez 1998; Dolling 2001). The
inaccuracies in these methods are not well understood or
quantied due to a lack of appropriate validation data.
Some recent studies such as Touber and Sandham (2009)
have used Large Eddy Simulation (LES) to study the SBLI,
but while such studies can provide insight into the behavior
of the SBLI, they are very difcult to apply to real systems,
particularly if a range of conditions are to be studied. In
addition, high delity simulations have generally been
performed on idealized systems, most notably neglecting
sidewall effects, so the behavior of the SBLI in such
regions has not been the subject of much numerical work
and is not yet well understood.
There is a fairly extensive experimental database on the
SBLI dating back several decades. Early experiments such
as Settles et al. (1976) and Bogdonoff and Kepler (1955)
relied on Schlieren and shadowgraph techniques, as well as
oil ows and pressure measurements. Numerous groups
D. B. Helmer L. M. Campo (&) J. K. Eaton
Department of Mechanical Engineering,
Stanford University, Stanford, CA 94305, USA
e-mail: lcampo@stanford.edu
D. B. Helmer
e-mail: dhelmer@stanford.edu
1 3
Exp Fluids (2012) 53:13471368
DOI 10.1007/s00348-012-1363-8
have acquired time-resolved pressure measurements to
study the unsteadiness of the SBLI (Selig et al. 1989;
Dolling and Murphy 1983; Beresh et al. 2002). In Selig
et al. (1989) and Smits and Muck (1987), hot-wire ane-
mometry was used to study Mach 2.9 interactions of various
strengths. Beresh et al. (1998) used double-pulse planar
laser scattering to analyze a Mach 5 SBLI. Bookey et al.
(2005) used ltered Rayleigh scatter and surface oil ow
visualizations in the same system to study the SBLI gen-
erated by a 24 compression ramp. They operated at a very
low Re
h
(&2,400) to make their results accessible to DNS,
unlike typical experiments, which operate at Re
h
[10,000.
The majority of recent experiments used particle image
velocimetry (PIV) techniques to measure the velocity eld.
Humble et al. (2007) studied a Mach 2.1 interaction using
PIV. Hou et al. (2003) used a wide eld PIV system with
multiple cameras to study development of the ow through
a Mach 2 SBLI generated by a 20 compression ramp.
Beresh et al. (2002) used a combination of PIV and time-
resolved pressure measurements to study a Mach 5 SBLI,
and they observed signicant low-frequency (\10 kHz)
motion of the shock. They also found a correlation between
the incoming boundary layer prole fullness and the
position of the shock, and theorized that the low-frequency
shock motion was caused by the presence of long structures
of high- and low velocity in the incoming boundary layer.
This mechanism has been supported by Ganapathisubra-
mani et al. (2007) and Humble et al. (2009), among others.
An alternate mechanism in which the shock motion is
controlled by the bubble dynamics has also been proposed
and has received considerable support in the community as
well (Piponniau et al. 2009).
Recently, more advanced PIV techniques have also been
applied to the study of the SBLI to acquire three-component
or time-resolved velocity data. Souverein et al. (2010) used
dual-plane PIV to study interactions of a variety of strengths
and concluded that a combination of mechanisms controlled
the shock motion, with the dominant mechanism depending
on the interaction strength. Humble et al. (2009) used
tomographic PIV to measure the three-component velocity
eld in two 7 9 4 9 1 cm
3
measurement volumes in a Mach
2.1 SBLI. They studied the instantaneous three-dimensional
structure in a nominally two-dimensional interaction.
Because of the volumetric nature of the data, more infor-
mation about the instantaneous three-dimensional structure
can be inferred from this experiment than from 2-D data,
although the method requires a relatively coarse resolution.
The three-dimensionality studied here was the inherent
three-dimensionality of the SBLI, which should be consid-
ered distinct from the focus of the present work, which is
focused on the behavior of the SBLI near channel sidewalls.
The existing experimental library covers a wide range of
Mach numbers, compression corner angles, and Reynolds
numbers. However, much of this work has been focused on
gaining fundamental understanding of the interaction and is
difcult to use in direct comparisons to CFD. Supersonic
systems present numerous challenges to experiments,
which can impact those experiments suitability for CFD
validation. The sensitivity of turbulent ows is one such
challengesmall geometric imperfections can generate
shock waves and expansions that alter the ow. If these
shocks are small, they are frequently assumed to be neg-
ligible and are not generally documented, but for detailed
CFD validation this is not ideal. Similarly, when intrusive
methods such as hot-wire anemometry or pitot probes are
used, they alter the ow by generating shock waves. While
there are corrections for these effects, they are generally
empirical and may not be properly suited to the SBLI. An
additional problem is that the large mass ow rate required
for supersonic wind tunnels often forces experimentalists to
use blowdown wind tunnels. While some experiments can
be run for several minutes, this still creates the possibility
of transient effects, particularly when one considers the
evolution of the wall temperature. Simulations do not
include these effects, thus creating a possible basis for
differences between CFD and the experimental results
other than limitations of CFD models. Seeding at high mass
ow rates can also be a challenge, and in some systems,
this is simplied by seeding downstream of the nozzle (e.g.
Beresh et al. 2002). This injection is again something that
can alter the ow and is not modeled in simulations.
Some of the largest challenges to the use of experimental
results in CFD validation come from the fact that the
experimental congurations are not designed specically
for that use. Boundary conditions, even in more modern
experiments, are frequently inadequately specied or
impossible to recreate in simulations. In some systems, the
inlet stagnation temperature and pressure conditions are
incompletely specied. In many wind tunnel experiments,
the geometry is designed to limit sidewall effects and
generate a nominally 2-D interaction, but this can affect the
system suitability for use in CFD validation. One example
of this can be seen in experiments such as Settles et al.
(1976), where aero-isolators are used to eliminate sidewall
effects. These isolators may mitigate the inuence of side-
walls, but they are prohibitively expensive to include in a
simulation. High aspect ratios can also be used to encourage
two-dimensionality, but again are very expensive to simu-
late. In some systems, a non-spanning wedge is used, which
is again very difcult to include in a simulation.
In real systems, there can be appreciable three-dimen-
sional effects, for example due to the sidewalls of the inlet
isolator of a SCRAMjet engine. However, these effects
have not been widely studied. Near the sidewalls, the lower
local velocity results in a different shock angle than in a
two-dimensional region. Also, the presence of transverse
1348 Exp Fluids (2012) 53:13471368
1 3
velocity gradients can change the physics of the ow by
introducing a different vorticity component. In most
experiments, efforts are made to either eliminate these
effects from the system or to take data in a region where
these effects are not present (e.g. Humble et al. 2007).
Consequently, while modern experiments have trended
toward greater accessibility for CFD, there is still a decit
of appropriate validation data for fully 3-D simulations of
realistic systems.
The specic objective of this experiment was to inves-
tigate the behavior of the SBLI generated by a compression
wedge near sidewalls and to provide a dataset appropriate
for validation of fully 3D simulations of a non-idealized
SBLI system. The present experiment was designed and
performed in parallel with a RANS simulation effort aimed
at uncertainty quantication of modern CFD in the SBLI. A
low aspect ratio, continuously operated Mach 2.05 wind
tunnel with a 20 compression wedge was studied using
high-resolution PIV. This low aspect ratio allowed direct
replication of the full system with CFD, and the relatively
thick boundary layers ([10 % of the channel width)
emphasized 3D effects caused by sidewall boundary layers.
A short (3 mm long), fully spanning wedge was used. This
conguration is easy to replicate in CFD, and the small size
simplies measurements in the recovery region behind the
wedge, as it limits the region over which optical access is
needed and speeds the recovery process. Also, the small
height relative to the incoming boundary layer results in a
behavior in the central region which differs from inviscid
predictions, which is not the case for larger wedges. This
provides a useful metric for CFD, as the ability to predict
the shock angle and strength in the inviscid region is tied to
the performance of the simulation near the compression
corner. In addition, this setup was used in a Monte Carlo-
type sensitivity study (Helmer et al. 2011), and the small
wedge size ensured a sensitive interaction. Velocity data are
taken in a series of transverse locations, including three
locations in the sidewall boundary layer to study the three-
dimensional effects near the sidewalls. One objective of this
study was to acquire detailed velocity uctuations data and
to study how these uctuations changed near the sidewall.
Consequently, a system allowing large data records to be
collected with steady inlet conditions was used, as the
upstream boundary conditions were continuously controlled
within a narrow band with well-dened uncertainties.
2 Experimental setup and data processing
2.1 Flow facility
The experiments were performed in a Mach 2.05 contin-
uous operation wind tunnel at Stanford University. A
schematic of the wind tunnel test section is shown in
Fig. 1. Note that the geometry in the test section is two-
dimensional, so only a cross section is shown here. The
tunnel has a 45-mm high by 47.5-mm wide constant area
cross section fed from a 2D converging/diverging nozzle
designed using the method of characteristics to eliminate
ow disturbances due to inlet shocks and expansion waves.
There is a 325-mm long boundary layer development
section downstream of the nozzle followed by a short 20
compression wedge machined into the upper wall of the
wind tunnel. The wedge fully spans the wind tunnel width
and extends 3 mm in the streamwise direction, resulting in
a 1.09-mm wedge height which reduces the tunnel height
in the downstream section to &43.9 mm. An expansion fan
is generated at the end of this ramp which affects the shock
dynamicsthis effect is not shown in Fig. 1. After the end
of the compression wedge, the constant area duct extends
for an additional 250 mm prior to dumping into an exhaust
plenum and mufer.
The tunnel sidewalls and nozzle were made of black
anodized aluminum to limit reections while the top and
bottom walls were made of polished plexiglass to provide
optical access for the laser. A 15-cm long plexiglass win-
dow was inserted in the sidewall to provide optical access
for the camera in the region around the wedge. The main
air ow was supplied by an Ingersoll-Rand SSR-XF400
compressor. The ow was passed from the compressor
through an Ingersoll-Rand TM1900-KTE4 compressed air
dryer to remove water vapor. The ow then passed through
a Norgren F18-C00-A3DA air lter. A Norgren R18-C00-
RNXA pressure regulator was used to control the upstream
pressure. A series of ve 2kW resistance heaters were
available to heat the ow, and a shell-tube heat exchanger
with a cooling water supply was used for cooling. The ow
passed through a long 3
00
copper pipe until it reached the
wind tunnel, where it then fed through a series of honey-
combs and grids prior to passing through the nozzle. The
stagnation conditions upstream of the nozzle were contin-
uously monitored: stagnation temperature was monitored
with an Omega model TJ36-44004 thermistor and was kept
between 29 and 31 C; stagnation pressure was measured
M = 2.05
20 wedge
oblique
shock
Fig. 1 Wind tunnel test section. The wedge height is not drawn to
scale. It is enlarged for clarity
Exp Fluids (2012) 53:13471368 1349
1 3
with a kiel probe connected to a Setra model 204D dif-
ferential pressure transducer and was kept within 1 kPa of
154 kPa gauge at all times. In order to eliminate transients,
the tunnel was run for a period of at least 20 min prior to
data acquisition. A summary of ow conditions and tunnel
geometry is provided in Table 1.
2.2 Pressure measurements
Mean wall pressure measurements were used to study this
ow. Static pressure taps were placed at a variety of
streamwise positions along the midline of the tunnel side-
wall, including one tap placed at the inlet nozzle throat. An
array of 50 pressure taps was placed at the base of the
compression wedge. There were rows of 10 taps at 5
spanwise locations. The rst row was placed &8.4 mm
from the sidewall, and the rows were spaced 7 mm apart to
allow access for a laser sheet between them. Taps were
spaced 2.1 mm in the streamwise direction, with the rst
tap located &1.25 mm from the base of the compression
wedge. A Setra Model 204D 25 psi differential transducer
was used to measure the pressure at these locations, with a
ScanCo model SSS-48C scanivalve system used to switch
between pressure taps.
2.3 Particle image velocimetry (PIV)
Two-dimensional PIV was used to quantify the mean and
uctuating velocity elds. Data were collected in 5
streamwise-wallnormal planes located 2.5, 4, 5.5, 21, and
26.5 mm from the sidewall in order to study the three-
dimensionality of the ow. The 2.5, 4, and 5.5 mm planes
are collectively referred to as the sidewall planes in this
paper. Data were acquired for both the compression corner
and the reected shock interactions; this paper discusses
only the results at the compression corner. In addition, due
to their similarities to the 21-mm plane, the 26.5-mm plane
results are not presented here. Figure 2 shows the locations
of measurement planes in the tunnel as well as the coor-
dinate system used in the gures to follow. Table 2 sum-
marizes the PIV parameters used for data collection, which
are described in more detail in this section.
A NewWave Solo-200XT dual-pulse PIV laser with a
wavelength of 532 nm and a pulse rate of 4 Hz was used.
Images were acquired with a TSI model 630047 PIV
camera with a 1,024 9 1,280 pixel array and a Nikon AF
Micro-Nikkor 200 mm 1:4D lens. The image resolution
was &8.4 lm/pixel, resulting in a eld of view of
&10.8 9 8.6 mm per PIV tile. Due to the high resolution
of the data, each plane is the composite of multiple PIV
tiles. The position of each tile is determined by imaging at
each camera location, a precisely machined metal grid
insert. Five thousand image pairs were collected in each
tile in the 21-mm plane as well as in all tiles at the com-
pression corner for the sidewall planes. In all other regions
of the sidewall planes, at least 1500 image pairs per tile
were acquired.
A TSI model 9307-6 Laskin nozzle seeder was used to
seed the ow with olive oil. Air for the seeder was provided
by an Ingersoll-Rand model 2340 compressor connected to
Table 1 Summary of ow conditions and geometry
M 2.05
P
0
154 1.0 kPa gauge
T
0
30 1 C
d
0
(21 mm upstream of wedge) 5.4 mm
a (wedge angle) 20
Cross section (upstream) 45 mm 9 47.5 mm
Cross section (downstream) 43.9 mm 9 47.5 mm
Fig. 2 Wind tunnel test section showing locations of PIV measure-
ment planes and coordinate system. The wedge is not drawn to scale.
It is enlarged for clarity
Table 2 Summary of PIV parameters
Camera aperture f/11
Camera working distance 32 cm
Laser sheet thickness 0.7 mm
Inter-frame time 0.8 ls
Particle size &1 lm
Particle time constant &1 ls
Field of view (per tile) 10.8 mm 9 8.6 mm
Image pairs per tile (21-mm plane and near
corner in sidewall planes)
5,000
Image pairs per tile (other regions of sidewall
planes)
At least 1,500
1350 Exp Fluids (2012) 53:13471368
1 3
a Norgren B74G-4AK-AD1-RMG regulator. The seed was
injected upstream of the ow straighteners to encourage a
uniform seed distribution and to avoid disturbing the ow
downstream of the nozzle. The nominal particle size was
1 lm, which coupled with the nominal density and vis-
cosity of olive oil gives a particle time constant of
&45 ls. The true time constant was evaluated by
examining the PIV results through the primary shock, and a
signicantly smaller value of &1 ls was found. Using the
experimentally determined time constant along with the
ow timescale given by the freestream velocity of 525 m/s
and boundary layer thickness of 5.4 mm gives a nominal
Stokes number of &0.1.
During PIV data collection runs, the throat pressure was
continuously monitored to provide information on any
changes in the mass ow rate. PIV images were acquired at
4 Hz, with an interframe time of 800 ns, resulting in
freestream particle displacements of approximately 50
pixels. Thousands of images were acquired at each location
to ensure sufcient yield for statistical convergence of both
mean velocity and turbulence statistics. To mitigate the
effect of blurring near the shock, images were high-pass
ltered to give sharp particle representations.
Data processing was performed using a conventional
cross-correlation PIV algorithm written at Stanford (Han
2001) which was modied to allow large streamwise dis-
placements and efcient parallel processing. Details of the
modications to the algorithm can be found in Helmer
(2011). Vectors were validated using two lters: a mini-
mum correlation value of 0.5 and a 3r lter which elimi-
nated any velocity vectors which differed by more than 3
standard deviations from the mean. The results were not
sensitive to the minimum correlation value; data were
reprocessed using values of 0.75 and 0.9 with the only
effect being a reduction in the number of valid vectors. Due
to the large displacements and dynamic range, an initial
window offset was applied based on an estimated velocity
prole. The system was tested for a variety of reasonable
initial offsets and showed no dependence on the input
prole. An iterative interrogation window scheme was
used, with a nal interrogation region size of 16 9 16
pixels, corresponding to a physical dimension of &130
lm 9 130 lm. Fifty percent overlap was used, resulting in
a vector spacing of approximately 65 lm. Due to the fact
that particle displacements ranged as high as 400 lm,
however, the interrogation region size is ner than the true
resolution of the data in the streamwise direction.
The camera magnication was chosen such that the
particle images were resolved (d
s
/ d
pix
&3.9). In addition,
a Gaussian sub-pixel estimator was used to mitigate peak-
locking effects in the PIV processing. The sheet thickness
was &700 lm, the minimum size that could be used to
generate images of sufcient particle density without
burning the plexiglass walls. Due to the high resolution of
the data, seed density near the wall was signicantly lower
than in the freestream. This resulted in peak-locking, which
reduced the accuracy of the sub-pixel peak-tting algo-
rithm. However, no evidence of seeding bias was observed,
and the data at the highest resolution did not deviate sig-
nicantly from data processed at a lower resolution. This
consistency suggests that aside from the peak-locking
error, the reduced seed density had no effect other than to
reduce the number of valid samples in a given interrogation
window, increasing uncertainty on some statistics.
2.4 Experimental uncertainty
The accuracy of the sub-pixel resolution peak-tting algo-
rithm would normally be considered to be 0.2 pixels, but
peak-locking results in a value nearer to 0.5 pixels. With
displacements of &50 pixels near the tunnel centerline, this
translated to &1 % uncertainty on instantaneous measure-
ments. The large dynamic range resulting from the high
resolution causes the effect of peak-locking on the uncer-
tainty of the velocity statistics to be negligible. Due to the
large number of samples, the statistical uncertainty in the
measurement of the mean was also\1 %. The mean velocity
measurements in all planes had a total uncertainty of\5 m/s.
In some locations, reections or glare from the wall created
false peaks in the correlationthese areas were identied in
the data processing and omitted fromthe data representation.
Similarly, oil owing along the windowblurred the image in
some regions, resulting in poor PIV qualitythese regions
were manually removed from the data representation.
The uncertainty of the turbulence measurements was
more difcult to quantify, as there are no universally
accepted standards for quantifying this uncertainty in PIV.
However, an estimate was made by examining the uctu-
ations in the turbulence levels for regions where the values
should not be changingspecically the asymptotic region
far from the top wall in the incoming boundary layer. For
the 21-mm plane, the statistical uncertainty in u
0
and v
0
was
less than 1 % due to the high yield of the data in this plane.
In the 5.5 and 4 mm planes, the yield was fairly high,
resulting in uncertainties of &0.5 % of U
1
in the u
0
and v
0
estimates, corresponding to uncertainties of a few percent.
For the 2.5-mm plane, lower seeding density resulted in
lower yields, resulting in uncertainties of nearly 1 % of U
1
in the u
0
and v
0
estimates. In addition to these uncertainties,
the uncertainty due to peak-locking is estimated as &3 %
based on a comparison to the work of Christensen (2004)
the high dynamic range again serves to mitigate this
uncertainty. The exception to this is in the wallnormal
velocity component very near (y \1 mm) the wall, where
the results display a strong negative bias. A similar near-
wall effect was observed in Souverein (2010).
Exp Fluids (2012) 53:13471368 1351
1 3
Estimates of the uncertainty of the correlation between
streamwise and wallnormal uctuations were more difcult
to make, particularly for the 2.5-mm plane, where the
potential presence of transverse structures makes it unclear
where the correlation should be constant. Based on uc-
tuations near the tunnel centerline, the uncertainty in hu
0
v
0
i
was estimated to be between 0.05 and 0.15 % of U
2
1
in the
2.5-mm plane depending on the distance from the top wall.
Note that this uncertainty is expressed as a percentage of
U
2
1
, not of the measured value. This uncertainty was
&0.05 % of U
2
1
for the 4 and 5.5 mm planes, and
&0.02 % of U
2
1
for the 21-mm plane. It is important to
note that this uncertainty is a function of both wallnormal
distance and streamwise location, as it depends strongly on
the seed density. Lower seed density near the wall results
in increased uncertainty, while increased seed density
behind the shock reduces the uncertainty. Because the
stated uncertainty bounds were derived from the upstream
boundary layer, they are conservative estimates for the
uncertainty in the interaction region. This effect is most
pronounced for the 2.5-mm plane. These estimates do not
include the effect of particle travel, which is most signi-
cant in regions of high streamwise velocities and gradients,
such as near shocks and expansion waves. The small par-
ticle time constant limits the errors induced by nite par-
ticle time response to a very small region behind the shock
(&500 lm). A detailed analysis of this is presented in
Helmer (2011).
In order to estimate the effects of peak-locking on the
computation of the Reynolds shear stress, hu
0
v
0
i, the fol-
lowing procedure was used. The Reynolds shear stress was
computed at several points in the upstream boundary layer
from the appropriate ensembles of valid pixel displace-
ments in the streamwise and wallnormal directions. Next,
the data in these ensembles of pixel displacements were
perturbed by independent samples from a zero-mean nor-
mal distribution with a standard deviation of 0.25 pixels,
corresponding to the estimated error due to peak-locking.
The Reynolds shear stresses were then recomputed using
these perturbed ensembles of pixel displacements and
compared to the original hu
0
v
0
i estimates. To simulate the
most severe case of peak-locking possible, the Reynolds
shear stresses were again computed after rounding all of
the perturbed pixel displacements to their nearest integer
values. Differences were negligible between the original
values of hu
0
v
0
i from the raw data and the estimates of
hu
0
v
0
i from the randomly perturbed data. The values of
hu
0
v
0
i computed from the worst-case scenario peak-locked
ensembles differed from the original estimates by up to
4 %, with the most pronounced differences in the sidewall
planes. The relatively low sensitivity of the velocity sta-
tistics to peak-locking is due to the large dynamic range of
pixel displacements, which is consistent with the ndings
of Christensen (2004). The peak-locking effects are neg-
ligible compared to the primary source of uncertainty on
hu
0
v
0
i which is the slow rate of convergence of this statistic.
The pressure data were not time-resolved, as all pressure
taps were connected to the scanivalve using long 1/16
00
diameter tubes which acted as low-pass lters. The Setra
204D had an RSS accuracy of 0.11 % of the full scale,
corresponding to &190 Pa for the 25 psi transducer used
to measure the tunnel pressures or &380 Pa for the 50 psi
transducer used to measure the upstream stagnation pres-
sure. The offset drift was monitored, and the largest
recorded drift was less than 5 mV, a maximum potential
bias of 170 Pa for the 25 psi transducer and 340 Pa for the
50 psi transducer. Each measurement of the upstream
stagnation pressure was the average of 1,000 readings
taken at 250 Hz to minimize noise effects. Thus, the
uncertainty in each upstream stagnation pressure mea-
surement was &500 Pa, corresponding to &0.3 % of the
measured value. During the mean pressure data acquisition
for the static pressure taps in the tunnel, 100 readings were
taken at each location. Each reading was the mean of 1,000
samples acquired at 200 Hz, meaning that each reading
was an average over 5 s and that each mean value repre-
sented the average pressure over a period of 500 s. This
long sampling eliminated any potential transient effects
and minimized statistical uncertainty, resulting in an
uncertainty of &250 Pa.
3 Incoming ow properties
The properties of the SBLI are likely to be strongly
dependent on the upstream velocity proles, so the
incoming ow is described here for all four planes of
interest. The ow upstream of the interaction shows sig-
nicant three-dimensional effects. The proles discussed
here were taken 21 mm upstream of the wedge, beyond any
inuence of the interaction region. Figure 3 shows
streamwise velocity proles for each of the planes. Five
velocity proles in the 21-mm plane between 23 and
21 mm upstream of the compression wedge were used to
calculate the boundary layer thickness d
0
. The average
value of 5.4 mm is used to non-dimensionalize all spatial
coordinates, while all velocities are non-dimensionalized
by the centerline velocity of 525 m/s. The momentum
thickness was estimated to be h & 450 lm, resulting in a
Reynolds number based on the momentum thickness of
Re
h
& 6,500.
The most obvious feature of the velocity proles is that
the asymptotic velocity is lower for the planes in the side-
wall boundary layer, but there are two other noteworthy
1352 Exp Fluids (2012) 53:13471368
1 3
facets of the streamwise velocity proles. The rst is that
while the 21-mm plane prole shows a nearly asymptotic
approach to the centerline velocity value, the other proles
show a signicant overshoot. The severity of this overshoot
increases with the depth in the sidewall boundary layer,
with a peak velocity in the 2.5-mm plane which is more than
40 m/s higher than the asymptotic value. A second feature
of note is that the wall shear stress does not appear to change
as the sidewall is approached. This can be seen by exam-
ining the velocity gradients near the wall. The ow inside
y = 0.5d
0
is nearly identical in all four measurement
planes. Small wallnormal velocities were observed, sug-
gesting either imperfectly canceled shocks or a weak sec-
ondary ow. The 2.5 and 4 mm planes show a small
velocity component (&5 m/s) directed toward the wall,
while the 21-mm plane velocity has a small component
directed away from the wall.
After correcting for density effects, the velocity proles
can be compared to the incompressible log-law. Applying
the density calculation outlined in (Van Driest 1951) and
using a recovery factor of 0.89 to determine the wall
temperature, the density-scaled streamwise velocity prole
for the 21-mm plane was calculated. This prole is shown
in Fig. 4 along with the incompressible log-law velocity
prole for a friction velocity of 22.1 m/s (selected based on
the best t to the data). Standard log-law coefcients of
j = 0.41 and B = 5.2 were used. The velocity prole
shows good agreement with the log-law prole from the
measurement point closest to the wall to y
?
& 550. This
agreement indicates a developed turbulent boundary layer
and suggests that potential issues such as seeding bias near
the wall do not signicantly affect the measured velocities.
The Reynolds averaged turbulence properties in the
incoming boundary layer were also three-dimensional.
Turbulence statistics were calculated by averaging over at
least 1,000 velocity samples at each location. For the
analysis of the upstream region where streamwise gradients
are small, the data were averaged over a streamwise dis-
tance of &0.4 mm to reduce statistical noise.
The density-scaled streamwise velocity uctuations are
shown in Fig. 5. The turbulence prole from the top wall
out to wallnormal distances of approximately 0.6d
0
is
insensitive to the spanwise position. More dramatic dif-
ferences are seen further from the wall. While all of the
proles show some initial decay from the near-wall
0 1 2 3
0.5
0.6
0.7
0.8
0.9
1
21mm
5.5mm
4mm
2.5mm
Fig. 3 Streamwise velocity proles 21 mm upstream of the com-
pression wedge. Gaps in proles represent areas where optical effects
introduce false peaks in the cross-correlation
10
1
10
2
10
3
10
4
10
15
20
25
30
35
Fig. 4 Density-scaled streamwise velocity prole 21 mm upstream
of the compression wedge in the 21-mm plane versus log-law prole
with u
s
= 22.1 m/s
0 1 2 3
0
0.5
1
1.5
2
2.5
z = 21mm
z = 5.5mm
z = 4mm
z = 2.5mm
Fig. 5 Density-scaled streamwise velocity uctuations 21 mmupstream
of the compression wedge
Exp Fluids (2012) 53:13471368 1353
1 3
turbulence levels, this decay is arrested at different loca-
tions for the different planes. The streamwise velocity
uctuations in the 2.5-mm plane stop decaying around
0.6d
0
, then show a slight increase to a value of approxi-
mately 1.6U
s
(&7 % of U
1
) at the edge of the measure-
ment domain. In the 4-mm plane, the streamwise
turbulence intensity is minimum at y/d
0
& 1.1, then
increases to a value of &1.2U
s
(&5 % of U
1
) at y/d
0
= 3.
The 4 mm prole also shows a more gradual decay of
uctuations across the boundary layer than the 5.5 mm
prole. The streamwise velocity uctuations in the 5.5-mm
plane do not stop decaying until the edge of the boundary
layer, reaching a nal value of &0.35U
s
(&1.5 % of U
1
).
In the 21-mm plane, the streamwise turbulence intensity
appears to approach the freestream value of &0.08U
s
(&0.3 % of U
1
) almost monotonically. There are small
uctuations in the turbulence levels of the 21-mm plane far
from the wall, possibly due to an imperfectly canceled inlet
shock; slight variations in the mean velocity in this area
similarly suggest the presence of a ow disturbance. There
is a clear trend of increasing turbulence levels far from the
top wall with decreasing distance from the sidewalls. In
addition, the size of the region in which the uctuations
decay decreases as the measurement plane approaches the
sidewall.
Figure 6 shows the density-scaled wallnormal velocity
uctuations. All four proles show similar peak intensities,
and all show a broad peak at approximately y/d
0
& 0.3.
However, as with the streamwise turbulent uctuations, the
behavior changes further out in the boundary layer. The
proles in all three sidewall planes decay near the wall, then
reverse this trend and gradually increase further from the
wall. The asymptotic turbulence level decreases with
increasing distance from the sidewall, while the distance
from the wall where the decay is arrested increases with
distance from the sidewall. As in the streamwise uctua-
tions, the 4 mm prole decay rate is smaller than that of the
5.5-mm plane. The distances at which the decay ends match
the locations observed in the streamwise uctuations.
Comparing the u
0
and v
0
proles allows for some anal-
ysis of the turbulence anisotropy. Near the wall, all four
planes show a similar degree of anisotropy, with the
streamwise uctuations several times as large as the wall-
normal uctuations. The u
0
:v
0
ratio starts above 2 near the
wall, then gradually declines with increasing distance from
the top wall. All three sidewall planes have a u
0
:v
0
ratio of
&1.5 far from the top wall. The 21-mm plane does not
match this behavior, as the turbulence levels do not stop
decreasing, and a disparity between the decay rates of u
0
and v
0
results in a changing degree of anisotropy. The value
near the centerline is signicantly lower than that of the
other proles and is close to 1, reecting minimal anisot-
ropy near the tunnel centerline.
The density-corrected Reynolds shear stress proles
(Fig. 7) all collapse inside of y/d
0
& 0.4. This agrees with
Fig. 3, which indicates that the wall shear stress is the same
for all 4 proles. Using the density scaling and normalizing
by the estimated friction velocity, U
s
, the peak Reynolds
shear stress approaches a value of 1 as expected. Outside of
the boundary layer, the shear stress drops to zero for all
planes. There appear to be slight differences in the decay
from the peak stress levels, with the 2.5-mm plane showing
faster decay than the other proles.
One use of the experimental data will be for comparison
to CFD, so specication of the boundary conditions to be
applied in such a comparison is important. The wall
boundary conditions are simple since all walls of the facility
were at and polished smooth. The tunnel was operated for
0 1 2 3
0
0.2
0.4
0.6
0.8
1
z = 21mm
z = 5.5mm
z = 4mm
z = 2.5mm
Fig. 6 Density-scaledwallnormal velocityuctuations 21 mmupstream
of the compression wedge
0 1 2 3
0
0.2
0.4
0.6
0.8
1
1.2 z = 21mm
z = 5.5mm
z = 4mm
z = 2.5mm
Fig. 7 Density-scaled Reynolds shear stresses 21 mm upstream of
the compression wedge
1354 Exp Fluids (2012) 53:13471368
1 3
at least 20 min prior to taking any data, allowing the tem-
perature distribution to reach a steady state and justifying an
assumption of adiabatic tunnel walls. The data in Figs. 5
and 6 show that the freestream uctuation levels, which
include contributions of freestream turbulence and any
overall ow unsteadiness, are well below 1 % of U
1
. The
velocity proles at x = -21 mm presented in Fig. 3 clearly
show that the ow is three-dimensional. The data are con-
sistent with the presence of corner vortices as would be
expected [see e.g. Davis and Gessner (1989)], and we
believe inclusion of these vortices is likely critical to
computation of the three-dimensional features of the SBLI.
The PIV data do not include measurements of the spanwise
velocity component and are not sufciently resolved in the
corner to fully describe these secondary ows. Therefore,
calculations should begin far upstream, perhaps even at
the nozzle throat. The nozzle acceleration is so strong that
the boundary layers are extremely thin at that point, and the
Reynolds number is high enough that transition occurs
shortly downstream of the nozzle exit.
4 Mean pressure data
Pressure measurements were taken for 3 tunnel operating
conditions. All pressure data were normalizedbythe upstream
stagnation pressure to enable direct comparisons of the pro-
les. No dependence on the mass owrate was observed. The
mean pressure data for the 5 spanwise rows of pressure taps
taken at the conditions used for PIV data acquisition are
shown in Fig. 8. The proles all exhibit the same basic
behavior; a consistent upstream pressure followed by a sharp
rise in pressure near the wedge due to the oblique shock.
Because of the breakdown of the shock in the subsonic region
of the boundary layer, the pressure rise occurs over several
millimeters as opposed to the width of a shock.
The upstream pressure varies by approximately 1 kPa
across the measurement region, possibly due to imperfect
shock cancelation in the nozzle. The largest difference in
the pressure proles is observed 5.5 mm upstream of the
wedge, near the start of the interaction. This difference is
less than 2 kPa, and given the large gradients in the pres-
sure prole in the interaction is likely indicative of very
small changes in shock position across the tunnel. The
pressure measurements suggest minimal spanwise changes
in the ow outside of the sidewall boundary layers.
5 Features of the SBLI
This section outlines the streamwise evolution of the
velocity statistics through the shock-boundary layer inter-
action. The data are not corrected for density because only
the wall pressure is known in the interaction region, and
accurate density corrections cannot be inferred from the
velocity proles as they were in the simpler upstream eld.
Figures 9, 10, 11, 12 show contour plots of the
streamwise velocity for each of the measured planes. Each
of these plots is a composite of 12 PIV tiles. White regions
on the plots indicate areas where optical effects prevented
reliable data acquisition. The origin of the streamwise axis
is placed at the base of the compression corner.
All four planes show both an oblique shock wave
upstream of the compression wedge and an expansion fan
behind the wedge. The 4 and 2.5 mm planes show a lower
velocity far from the wall. A large low-velocity region
surrounds the wedge. The size of this region varies across
the span as is most easily observed by considering the
green contour representing velocities below &300 m/s.
This contour extends far below the corner for the 21 and
5.5 mm planes before turning back toward the wall after
the end of the wedge. In the 4-mm plane, this contour still
extends beyond the wedge, but has noticeably begun to
atten out. In the 2.5-mm plane, this contour barely
extends beyond the wedge, indicating a weakening of the
interaction in this plane.
There is much less three-dimensionality behind the
interaction zone. The velocity contours pinch rapidly
toward the wall after the corner due to the presence of the
expansion fan at the edge of the wedge. In the 21-mm
plane, this is followed by further thinning of the boundary
layer and then a region of boundary layer thickening. In the
three sidewall planes, there is no further thinning of the
boundary layer after the pinching caused by the expansion.
Instead, these planes all show a rapid initial thickening of
the boundary layer after the interaction zone ends, followed
by a region of slower development. The recovery region of
20 15 10 5 0
0.13
0.14
0.15
0.16
0.17
0.18
0.19
0.2
0.21
(mm)
z=8.4mm
z=15.4mm
z=22.4mm
z=29.4mm
z=36.4mm
Fig. 8 Mean pressure measurements upstream of the compression
wedge. Pressures are normalized by upstream stagnation pressure
Exp Fluids (2012) 53:13471368 1355
1 3
the 21-mm plane shows rapid boundary layer thickening at
the end of the measurement domain due to the reection of
the shock from the bottom wall of the tunnel.
The three-dimensional nature of the interaction is more
clearly seen in the wallnormal velocity contours shown in
Figs. 13, 14, 15, 16. The strength of the shock decreases
Fig. 9 Streamwise velocity
contours in the 21-mm plane.
White regions indicate areas
where data could not be
acquired. This includes all
locations where the yield was
below 10 % and regions where
optical effects were known to
prevent reliable data acquisition
Fig. 10 Streamwise velocity
contours in the 5.5-mm plane
Fig. 11 Streamwise velocity
contours in the 4-mm plane
1356 Exp Fluids (2012) 53:13471368
1 3
near the sidewall, as can be seen by the reduction in the
downward velocity behind the shock for the near-wall
planes. The behavior behind the expansion fan changes
with spanwise position as well. In the 5.5-mm plane, there
is a clear region of positive wallnormal velocity, meaning
the ow is driven toward the wall. This feature is weaker in
the 21 and 4 mm planes and is not observed in the 2.5-mm
plane. This change through the sidewall boundary layer is
Fig. 12 Streamwise velocity
contours in the 2.5-mm plane
Fig. 13 Wallnormal velocity
contours in the 21-mm plane
Fig. 14 Wallnormal velocity
contours in the 5.5-mm plane
Exp Fluids (2012) 53:13471368 1357
1 3
consistent with the structure observed in the streamwise
velocity contoursas the extent of the region of low
streamwise velocity decreases, there is less of a need to
drive the low-velocity region toward the wall.
Examining the upstream edge of the interaction zone
reveals signicant spanwise variation. Because the cen-
terline Mach number decreases near the sidewall, the main
shock angle increases approaching the sidewall. As a
result, in the area far from the top wall, the shock is seen
signicantly further upstream in the 2.5 and 4 mm planes.
In the 21-mm plane, a single shock appears at the
leading edge of the interaction. A weak shock (indicated by
the purple region upstream of the main shock in Fig. 13),
most likely an imperfectly canceled nozzle shock, is seen
impinging on the shock generated by the compression
wedge. The leading edge is no longer a distinct single
downturn in the measurement planes near the sidewall.
Instead, the ow turns strongly downward, then back
toward horizontal, and then turns downward again. This is
evidence of a pair of shocks separated by an expansion.
Because this structure is more dened nearer the wall and
is not evident in the 21-mm plane, this suggests that the
rst shock and expansion start in the corner and interact
before reaching the centerline. The nozzle shock observed
in the 21-mm plane is not seen in any of the sidewall
planes.
Another feature revealed by the wallnormal velocity is
that the inuence of the shock is rst seen several milli-
meters away from the wall. Not only do the velocity con-
tours fail to penetrate all the way to the wall, the rounding
of the contours means that their furthest upstream inuence
is not the location of deepest penetration in the boundary
layer.
5.1 Detailed examination of the 21-mm plane
Figure 17 shows streamwise and wallnormal velocity
proles upstream of the wedge for the 21-mm plane. The
streamwise location is dened relative to the compression
wedge location. Note that the wallnormal coordinate is
Fig. 15 Wallnormal velocity
contours in the 4-mm plane
Fig. 16 Wallnormal velocity
contours in the 2.5-mm plane
1358 Exp Fluids (2012) 53:13471368
1 3
dened as positive in the downward direction (away from
the top wall). For clarity, only every fourth data point is
plotted. The rst changes in the velocity prole are seen
approximately 2.5d
0
upstream of the corner, as the
streamwise velocities in the boundary layer are slightly
smaller than in the incoming boundary layer. The wall-
normal velocity component at the edge of the boundary
layer, which is positive in the incoming ow, becomes
slightly negative at the start of the interaction region. As
the ow approaches the wedge, the streamwise velocity
prole decreases further, and a positive wallnormal
velocity component appears due to the presence of the
oblique shock.
The rst signs of the peak in wallnormal velocity are
observed approximately 1.2d
0
upstream of the compression
corner, signicantly downstream of where the streamwise
velocity decrease begins. The wallnormal velocity proles
show a dramatic increase as the ow approaches the
wedge, as expected. The slightly negative wallnormal
velocities observed for y/d
0
[1.5 can be attributed to the
presence of an imperfectly canceled nozzle shock. No ow
reversal is observed to a distance of 250 lm from the top
wall. This is likely due to the small size of the wedge, as
comparable investigations in the literature (e.g. Gana-
pathisubramani et al. 2007) observed ow reversal and a
larger interaction region at the compression corner.
The evolution of the velocity proles downstream of the
wedge can be seen in Fig. 18. The rst three proles are
taken along the wedge, the fourth is 1.1d
0
downstream of
the end of the wedge, and the nal prole is just upstream
of the return of the reected shock. Proceeding along the
wedge, the streamwise velocity near the wall increases,
possibly due to the inuence of the expansion fan or
streamline curvature. Further from the wall, the trend is
reversed, with the more downstream locations along the
wedge showing a lower streamwise velocity.
Because the subsonic portion of the boundary layer
allows for gradual curvature of the sonic line, the ow does
not show a single discrete jump in wallnormal velocity
(even after accounting for particle travel effects). A more
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
0 0.5 1 1.5 2
0.02
0
0.02
0.04
0.06
0.08
0.1
x = 3.9
0
x = 2.5
0
x = 1.9
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 2.5
0
x = 1.9
0
x = 0.7
0
x = +0.0
0
Fig. 17 Streamwise
development of streamwise
(left) and wallnormal (right)
velocity proles in the 21-mm
plane upstream of the
compression wedge
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
0 0.5 1 1.5 2 2.5 3
0.02
0
0.02
0.04
0.06
0.08
0.1
x = +0.0
0
x = +0.3
0
x = +0.5
0
x = +1.7
0
x =+13.7
0
x = +0.0
0
x = +0.3
0
x = +0.5
0
x = +1.7
0
x =+13.7
0
Fig. 18 Streamwise
development of streamwise
(left) and wallnormal (right)
velocity proles in the 21-mm
plane downstream of the
compression wedge
Exp Fluids (2012) 53:13471368 1359
1 3
gradual ow turning is observed, caused by a series of
compression waves which eventually coalesce further from
the wall into a single shock.
At x/d
0
= 0.5, the presence of a local wallnormal
velocity minimum at y/d
0
& 0.2 shows the two distinct
mechanisms acting on the wallnormal velocity. The wall
angle forces the ow to turn, resulting in high wallnormal
velocities near the wall, and the shock creates positive
wallnormal velocities further from the wall. The peak
closest to the wall should control the dynamics of the
expansion fan after the compression corner. A mismatch
between the ow angle near the wall and the ow angle
behind the shock may explain why the ow is not perfectly
horizontal after the expansion. The proles at x/d
0
= 1.7
show the rapid recovery of the velocity prole for
y/d
0
\0.7 after the expansion fan as well as the inuence
of the shock wave and expansion fan far from the wall. A
small wallnormal velocity component is observed which
indicates that the expansion fan and shock wave strengths
are not perfectly matched. Finally, at x/d
0
= 13.7, the
velocity prole has nearly recovered to a fully developed
turbulent boundary layer prole near the wall with
u
s
& 21.2 m/s, slightly lower than the friction velocity in
the incoming boundary layer. The boundary layer thickness
has increased by a factor of approximately 1.4.
The length of the compression wedge has a signicant
effect on the ow. The maximum ow angle observed
behind the shock is &8, signicantly smaller than the
wedge angle of 20. More evidence of the effect of the
wedge length can be seen when considering the shock
angle, which was calculated by comparing the location of
the peak in the wallnormal velocity uctuations at mul-
tiple distances away from the top wall. Locations where
the shock was far from the wall were used so that effects
of PIV spatial averaging and particle tracking were con-
stant. A shock angle of &37 was measured, signicantly
smaller than the shock angle of &52 predicted by
oblique shock relations for a 20 wedge. However, this
angle agrees well with the 36 predicted by oblique shock
relations for an 8 wedge. The upstream inuence of the
interaction allows for smaller streamline curvature than
the inviscid case allows, resulting in a weaker shock
structure.
Figures 19 and 20 show the unscaled velocity uctua-
tions upstream and downstream of the compression corner.
Only locations with a minimum of 500 valid velocity
measurements are represented, and every other data point is
omitted for clarity. At x/d
0
= -2.5 and x/d
0
= -1.9, the
turbulence and Reynolds shear stress proles show little
change from the unperturbed prole. The turbulence levels
appear slightly elevated near the edge of the boundary
layer. As the ow approaches the wedge, the peak
streamwise turbulence intensity does not appear to increase
signicantly, but the peak moves a short distance from the
wall. The wallnormal peak turbulence level increases sig-
nicantly in the interaction zone, and the peak induced by
the interaction is located signicantly further from the wall
than the peak in the streamwise turbulence. This reects
the fact that the wallnormal uctuation peak tracks the
shock, while the main increase in streamwise uctuations is
the result of the uctuations in the boundary layer induced
by the shock.
As the ow proceeds along the wedge, the turbulence
levels decrease, and the peaks move back toward the wall.
At x/d
0
= 1.7, the turbulence levels near the wall are still
elevated, but are rapidly decreasing due to the inuence of
the expansion fan. Two peaks are observed far from the
wallthis indicates the presence of two shocks. The
weaker of these shocks is induced by the mismatch
between the oblique shock and expansion fan discussed
earlier; the ow is driven toward the wall by the strength of
the expansion fan, causing the formation of a second shock.
0 0.5 1 1.5 2
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
0.06
x = 3.9
0
x = 2.5
0
x = 1.9
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 2.5
0
x = 1.9
0
x = 0.7
0
x = +0.0
0
Fig. 19 Streamwise
development of unscaled
streamwise (left) and
wallnormal (right) velocity
uctuations in the 21-mm plane
upstream of the compression
wedge
1360 Exp Fluids (2012) 53:13471368
1 3
In order to consider the relaxation of the turbulence
proles, Fig. 21 scales the downstream prole by the local
boundary layer thickness instead of d
0
. The wallnormal
turbulence at x/d
0
= 13.7 has relaxed completely and
matches the upstream prole throughout the measurement
domain. Small differences are observed in the streamwise
turbulencethe turbulence level very near the wall appears
to be lower at x/d
0
= 13.7, and the decay rate at the edge of
the boundary layer appears to be slightly larger for the
more downstream prole. However, the relaxation appears
to be nearly complete.
Figure 22 shows the unscaled Reynolds shear stress
throughout the measurement domain. No differences are
seen in the most upstream proles. The prole at x/d
0
=
-0.7 shows an increase in the Reynolds shear stress near
the wall. This peak broadens and moves away from the
wall as the ow approaches the wedge. As with the
streamwise and wallnormal turbulence, the Reynolds shear
stresses for y/d
0
\1 decrease to levels similar to those seen
in the unperturbed region downstream of the wedge. At
x/d
0
= 1.7, the inuence of the weaker shock on the
Reynolds shear stress is much smaller than the effect
observed on the normal stress levels.
5.2 Detailed examination of the sidewall planes
Figures 23, 24, 25 show the streamwise evolution of the
velocity proles for the three planes in the sidewall
boundary layer. Only locations with a minimum of 100
valid samples are included, and the plots display every
fourth point for clarity.
The most prominent feature of the sidewall proles at
the furthest upstream location is the non-monotonic
behavior of the streamwise velocity across the boundary
layer. The strength of this effect increases as the sidewall is
approached and decreases in the streamwise direction as
0 0.5 1 1.5 2 2.5 3
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5 2 2.5 3
0
0.01
0.02
0.03
0.04
0.05
0.06
x = +0.0
0
x = +0.3
0
x = +0.5
0
x = +1.7
0
x =+13.7
0
x = +0.0
0
x = +0.3
0
x = +0.5
0
x = +1.7
0
x =+13.7
0
Fig. 20 Streamwise
development of unscaled
streamwise (left) and
wallnormal (right) velocity
uctuations in the 21-mm plane
downstream of the compression
wedge
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
x = 3.9
0
x =+13.7
0
x = 3.9
0
x =+13.7
0
Fig. 21 Comparison of
unscaled velocity uctuations in
the 21-mm plane upstream and
downstream of the interaction
Exp Fluids (2012) 53:13471368 1361
1 3
0 0.5 1 1.5 2
0
0.5
1
1.5
2
2.5
x 10
3
x = 3.9
0
x = 2.5
0
x = 1.9
0
x = 0.7
0
x = +0.0
0
0 0.5 1 1.5 2 2.5 3
0
0.5
1
1.5
2
2.5
x 10
3
x = +0.0
0
x = +0.3
0
x = +0.5
0
x = +1.7
0
x =+13.7
0
Fig. 22 Streamwise
development of unscaled
Reynolds shear stress in the
21-mm plane upstream (left)
and downstream (right) of the
compression wedge
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
0 0.5 1 1.5 2 2.5 3
0.02
0
0.02
0.04
0.06
0.08
0.1
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 23 Streamwise
development of streamwise
(left) and wallnormal (right)
velocity proles in the 5.5-mm
plane
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
0 0.5 1 1.5 2 2.5 3
0.02
0
0.02
0.04
0.06
0.08
0.1
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 24 Streamwise
development of streamwise
(left) and wallnormal (right)
velocity proles in the 4-mm
plane
1362 Exp Fluids (2012) 53:13471368
1 3
the boundary layer begins to feel the inuence of the
wedge. The trend of this overshoot in the sidewall proles
is consistent with the secondary corner vortices observed
by Davis and Gessner (1989) in a supersonic turbulent duct
ow. Davis and Gessner describe two counter-rotating
vortical structures that are symmetric about the corner
bisector of the spanwise-wallnormal plane and act to force
high momentum uid from the freestream into the corner.
The wallnormal velocity proles at the upstream edge of
the interaction (x/d
0
= -1.7) all show a small negative
wallnormal velocity. This feature was also observed in the
21-mm plane data, and it indicates that the ow is pinched
toward the wall before pushing away from the wall as it
approaches the wedge. The streamwise velocity proles in
the sidewall planes retain their non-monotonic shape from
far upstream until approximately x = 0, where the proles
no longer exhibit an overshoot. In addition, the shock-
boundary layer interaction region has a shorter streamwise
extent in the sidewall boundary layer than near the tunnel
centerline. This is seen most clearly by examining the
differences between the unperturbed boundary layer
(x/d
0
= -3.9d
0
) and the prole at x/d
0
= -1.7 in Figs. 17
and 23, 24, 25. The differences between these proles are
signicantly smaller in the sidewall planes than they are in
the 21-mm plane.
The ow turning angle also changes with spanwise
position; this angle shrinks to approximately 6 at the 5.5-
mm plane and decreases to approximately 5 in the 4 and
2.5 mm planes. This suggests a stronger shock near the
centerline than at the corners, which could drive ow
toward the sidewalls. The angle in the 2.5-mm plane
appears slightly larger than that observed in the 4-mm
plane. This difference may result in a slightly higher pres-
sure nearer the corner than at the 4mm plane, which could
result in a local minimum in pressure near the 4-mm plane.
In order to emphasize the differences between the cen-
terline and sidewall ow features, Figs. 26 and 27 show the
mean streamwise and wallnormal velocity proles at
x = -0.7d
0
and x = 0 plotted across the span. While the
streamwise velocity proles at x = -0.7d
0
show similar
0 0.5 1 1.5 2 2.5 3
0
0.2
0.4
0.6
0.8
1
0 0.5 1 1.5 2 2.5 3
0.02
0
0.02
0.04
0.06
0.08
0.1
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 25 Streamwise
development of streamwise
(left) and wallnormal (right)
velocity proles in the 2.5-mm
plane
0 1 2 3
0
0.2
0.4
0.6
0.8
1
0 1 2 3
0.02
0.01
0
0.01
0.02
0.03
0.04
0.05
z=21mm
z=5.5mm
z=4mm
z=2.5mm
z=21mm
z=5.5mm
z=4mm
z=2.5mm
Fig. 26 Comparison of
streamwise (left) and
wallnormal (right) velocity
proles across the span of the
tunnel at x = -0.7d
0
Exp Fluids (2012) 53:13471368 1363
1 3
trends to the upstream (unperturbed) proles shown in
Fig. 3, the wallnormal velocity plot shows no clear trend
across the sidewall boundary layer. The peak wallnormal
velocity is signicantly higher for the 4mm plane than
either the 5.5 or 2.5 mm planes, while the 2.5-mm plane
shows a larger region of positive velocity than the other
sidewall planes. This may be evidence of a corner vortex or
upwash produced by the spanwise pressure gradients gen-
erated by the changing shock strength and shock angle
across the sidewall boundary layer.
Further downstream at x = 0, there are clear trends in
the wallnormal velocity proles across the sidewall
boundary layer. The peak in the velocity prole is broader
near the sidewall, and the peak velocity decreases with
decreasing distance from the sidewall. In addition, the
streamwise velocity prole in the 2.5-mm plane has a fuller
shape inside y = 0.5d
0
than the other three proles. This
suggests that in this near-corner region, there is a signi-
cant spanwise component of velocity pushing high
momentum uid toward the sidewall. Unlike the upstream
sidewall proles, there is no overshoot in the sidewall
proles at x = 0, which suggests that outside of y/d
0
= 0.5,
there may be a signicant spanwise component of velocity
away from the wall carrying low momentum uid toward
the center of the channel.
As the ow approaches the wedge, it is posited that the
symmetry of the corner vortices that assumed to be present in
the upstream unperturbed ow is broken. This is due to the
inuence of the streamwise adverse pressure gradient and
shock wave which cause the owto turn downward as well as
the spanwise pressure gradient due to the varying shock
strength. The lack of a clear trend in wallnormal velocity
proles at x/d
0
= -0.7 could be due to the corner vortex near
the top wall being forced toward the sidewall. The center of
such a vortex is likely between the 4 and 2.5 mm planes,
causing stronger downward velocity in the 4-mm plane and
weaker downward velocity in the 2.5-mm plane. By the time
the ow reaches the wedge, it is likely that the secondary
corner ow evolves into a single elongated vortical structure
along the sidewall which washes high momentum toward the
sidewall near the top wall and pushes low momentum uid
from the sidewall boundary layer into the freestream further
away fromthe top wall. This feature could explain the lack of
overshoot observed in the sidewall proles at x = 0 and is
consistent with the trends in wallnormal velocity at this
location. A series of qualitative sketches (Fig. 28) show the
0 1 2 3
0
0.2
0.4
0.6
0.8
1
0 1 2 3
0.02
0
0.02
0.04
0.06
0.08
0.1
z=21mm
z=5.5mm
z=4mm
z=2.5mm
z=21mm
z=5.5mm
z=4mm
z=2.5mm
Fig. 27 Comparison of
streamwise (left) and
wallnormal (right) velocity
proles across the span of the
tunnel at x = 0
Fig. 28 Qualitative sketches of a zoomed in view of the upper left
corner of a spanwise-wallnormal plane showing the proposed corner
vortex evolution. a shows a plane in the upstream region where the
vortices are symmetrical. bd show the presumed evolution of the
structures in successive downstream planes as the wedge is
approached. In the last panel, the downward velocity vectors are
due to the inuence of the shock wave
1364 Exp Fluids (2012) 53:13471368
1 3
proposed evolution of these features as the wedge is approa-
ched. A direct numerical simulation of this ow may be
required in order to validate these conclusions since mea-
surement in the cross-plane would be very difcult.
Figures 29, 30, 31, 32, 33, 34 show the unscaled turbu-
lent uctuations and Reynolds shear stresses. Only loca-
tions with a minimum of 400 valid samples are included.
Without density scaling, only very general observations
0 0.5 1 1.5 2
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
0.06
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 29 Streamwise
development of streamwise
(left) and wallnormal (right)
unscaled velocity uctuations in
the 5.5-mm plane
0 0.5 1 1.5 2
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
0.06
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 30 Streamwise
development of streamwise
(left) and wallnormal (right)
unscaled velocity uctuations in
the 4-mm plane
0 0.5 1 1.5
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0 0.5 1 1.5
0
0.01
0.02
0.03
0.04
0.05
0.06
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 31 Streamwise
development of streamwise
(left) and wallnormal (right)
unscaled velocity uctuations in
the 2.5-mm plane
Exp Fluids (2012) 53:13471368 1365
1 3
about the turbulence structure can be made. As was seen in
the 21-mm plane, all three sidewall planes show the
development of a peak in the streamwise turbulence level at
the compression corner which has moved a short distance
away from the wall. The turbulence levels increase signif-
icantly in the interaction zone, although the degree of
increase cannot be quantied without density data. The
wallnormal turbulence levels become more uniform closer
to the sidewall, with no clear peak observed in the 2.5-mm
plane. The behavior of the shear stress also changes sig-
nicantly across the sidewall boundary layer. In the 5.5-mm
plane, there is a clear jump within the interaction, with the
peak moving out from the wall as the ow approaches the
corner. This behavior changes in the 4mm plane, where the
proles at x/d
0
= -0.7 and at the corner are very similar,
with the only differences appearing in the decay near the
edge of the boundary layer. In the 2.5-mm plane, a small
increase in the Reynolds shear stress is again observed in
0 0.5 1 1.5
2
0
0.5
1
1.5
2
2.5
x 10
3
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 32 Streamwise development of unscaled Reynolds shear stress
in the 5.5-mm plane
0 0.5 1 1.5 2
0
0.5
1
1.5
2
2.5
x 10
3
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 33 Streamwise development of unscaled Reynolds shear stress
in the 4-mm plane
0 0.5 1 1.5
0
0.5
1
1.5
2
2.5
x 10
3
x = 3.9
0
x = 1.7
0
x = 0.7
0
x = +0.0
0
Fig. 34 Streamwise development of unscaled Reynolds shear stress
in the 2.5-mm plane
0 0.5 1 1.5 2
0.6
0.65
0.7
0.75
0.8
0.85
0.9
0.95
1
x = 3.9
0
x =+10.0
0
Fig. 35 Comparison of velocity proles upstream and downstream of
the interaction in the 2.5-mm plane
1366 Exp Fluids (2012) 53:13471368
1 3
the interaction. This lack of a clear spanwise trend may be
the result of a corner vortex or other spanwise ow gener-
ated by the non-uniform spanwise pressure distribution.
The nal region considered along the upper wall is the
recovery region for the 2.5-mm plane. It was observed in the
21-mm plane that at x/d
0
& 13.7, the velocity proles
recovered to standard turbulent boundary layer proles, but
the turbulence had not completely recovered. At this
streamwise location in the 2.5-mm plane, the inuence of
the shock reecting off the bottom wall can already be seen.
Therefore, it does not make sense to compare the 2.5-mm
plane prole at x/d
0
= 13.7 to the incoming prole to
determine whether the boundary layer has relaxed. Instead,
a prole at a less downstream location of x/d
0
= 10 where
the effect of the reected shock is not yet felt in the 2.5-mm
plane was used. A comparison between this prole and the
incoming prole is shown in Figs. 35 and 36.
The downstream velocity prole, when scaled for an
increased boundary layer thickness, was found to collapse
with the incoming prole inside y/d = 0.5. However, a
1.6:1 scaling was observed, indicating more rapid boundary
layer thickening near the sidewall, especially given that the
comparison location was further upstream in the 2.5-mm
plane. Outside of y/d = 0.5, the (rescaled) downstream
velocity prole deviates from the incoming prole. This
deviation is due to the relative strength of the primary
shock wave and expansion fan at varying wallnormal dis-
tances in the 2.5-mm plane. Outside of y/d = 0.5, the
shock is stronger than the expansion fan. Therefore, the
expansion fan does not turn the ow completely horizontal,
and it does not accelerate the ow back to the same free-
stream value as the incoming prole. This mismatch
between shock and expansion fan strengths can be seen
clearly in the cyan contour behind the shock/expansion
structures in Fig. 16.
The velocity uctuation proles also collapse inside
y/d
0
= 0.5 and disagree further out in the boundary layer.
The velocity uctuation proles never relax to the original
proles before encountering the returning reected shock.
The apparent rapid recovery may be due to the fact that the
perturbations to the turbulence quantities appeared to be
highly concentrated in the corner, and thus, there was less
of a deviation to correct after the wedge ended.
6 Conclusion
PIV and pressure measurements were taken in a Mach 2.05
continuously operated wind tunnel with a small 20 com-
pression wedge. Pressure measurements showed no sig-
nicant three-dimensionality outside of the sidewall
boundary layers of the wind tunnel. PIV measurements
showed strong three-dimensionality in the velocity and
turbulence levels in the incoming boundary layer, although
the Reynolds shear stress showed minimal variation across
the sidewall boundary layer. The velocity prole near the
centerline of the tunnel was shown to be a developed tur-
bulent boundary layer prole after density variations were
accounted for.
The shock structure depended strongly on spanwise
position. Near the tunnel sidewall, the primary shock
generated by the compression wedge gave rise to two
distinct shocks at the leading edge of the interaction. The
shock strength was decreased near the sidewall, as seen by
the decrease in wallnormal velocity. Near the tunnel cen-
terline, the shock inuence extended roughly 2.5d
0
upstream of the compression corner. Both the shock angle
and ow angle were smaller than the inviscid predictions
for a 20 compression wedge in a Mach 2.05 ow, as the
small wedge size resulted in an effective wedge angle of
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0 0.5 1 1.5 2
0
0.01
0.02
0.03
0.04
0.05
x = 3.9
0
x =+10.0
0
x = 3.9
0
x =+10.0
0
Fig. 36 Comparison of
turbulence proles upstream
and downstream of the
interaction in the 2.5-mm plane
Exp Fluids (2012) 53:13471368 1367
1 3
approximately 8. This resulted in a mismatch between
the strengths of the shock wave and the expansion fan
behind the wedge. Reynolds normal and shear stress
levels increased signicantly in the interaction zone, but
rapidly returned to their unperturbed levels behind the
expansion fan. Nearly complete relaxation was observed
at x/d
0
= 13.7.
The extent of the interaction region was smaller near
the tunnel sidewalls. The peak wallnormal velocity rose
more rapidly in the upstream portion of the interaction in
the 4-mm plane than in any other measurement plane.
Reynolds shear stress and velocity uctuation levels
varied signicantly across the boundary layer, with the
4-mm plane showing less change in the interaction zone
than either 2.5 or 5.5 mm planes. In the absence of
density data, it is difcult to compare turbulence quan-
tities to the unperturbed state. However, it is clear that
signicant three-dimensional effects exist. The strength
of the interaction changes, and no clear trends are
observed in the behavior of quantities such as the Rey-
nolds shear stress across the sidewall boundary layer.
This suggests the inuence of spanwise ow, possibly
due to multiple competing effects such as the changing
inviscid shock angle and ow turning angle across the
interaction zone.
The presence of counter-rotating corner vortices is
consistent with the velocity proles observed in the side-
wall planes upstream of the interaction. It is proposed that
these corner vortices evolve into a single elongated vortical
structure along the sidewall as the ow reaches the foot of
the compression wedge. The mechanisms proposed to
cause these changes in the corner ows are the spanwise
pressure gradient due to the varying shock strength across
the span, as well as the streamwise adverse pressure gra-
dient and shock wave which cause the ow to turn
downward.
Acknowledgments This material is based upon work supported by
the Department of Energy [National Nuclear Security Administration]
under Award Number NA28614. Support was also received from the
Graduate Research Fellowship Program of the National Science
Foundation, the National Defense Science and Engineering Graduate
Fellowship program of the Department of Defense, and the Stanford
Graduate Fellowship Program.
References
Beresh S, Comninos M, Clemens N, Dolling D (1998) The effects of
the incoming turbulent boundary layer structure on a shock-
induced separated ow. AIAA Paper 98-0629
Beresh S, Clemens N, Dolling D (2002) Relationship between
upstream turbulent boundary-layer velocity uctuations and
separation shock unsteadiness. AIAA J 40:24122422
Bogdonoff S, Kepler C (1955) Separation of a supersonic turbulent
boundary layer. J Aeronaut Sci 22:414424
Bookey P, Wyckham C, Smits A, Martin M (2005) New experimental
data of stbli at dns/les accessible reynolds numbers. AIAA Paper
2005-309
Christensen K (2004) The inuence of peak-locking errors on
turbulence statistics computed from PIV ensembles. Exp Fluids
36:484497
Davis D, Gessner F (1989) Further experiments on supersonic
turbulent ow development in a square duct. AIAA J 27:1023
1030
Dolling D (2001) Fifty years of shock-wave/boundary-layer interac-
tion research: what next? AIAA J 39:15171531
Dolling D, Murphy M (1983) Unsteadiness of the separation shock
wave structure in a supersonic compression ramp oweld.
AIAA J 21:16281634
Ganapathisubramani B, Clemens N, Dolling D (2007) Effects of
upstream boundary layer on the unsteadiness of shock-induced
separation. J Fluid Mech 585:369394
Han D (2001) Study of turbulent nonpremixed jet ames using
simultaneous measurements of velocity and CH distribution.
PhD Thesis, Stanford University
Helmer D (2011) Measurements of a three-dimensional shock-
boundary layer interaction. PhD Thesis, Stanford University
Helmer D, Campo L, Eaton J (2011) Sensitivity of a shock-boundary
layer interaction to geometric perturbations. TSFP7 Conference
Paper 7B1P
Hou Y, Clemens N, Dolling D (2003) Wide-eld PIV study of shock-
induced turbulent boundary layer separation. AIAA Paper
2003-0441
Humble R, Scarano F, van Oudheusden B (2007) Particle image
velocimetry measurements of a shock wave/turbulent boundary
layer interaction. Exp Fluids 43:173183
Humble R, Elsinga G, Scarano F, van Oudheusden B (2009) Three-
dimensional instantaneous structure of a shock wave/turbulent
boundary layer interaction. J Fluid Mech 622:3362
Knight D, Degrez G (1998) Shock wave boundary layer interactions
in high mach number ows. A critical survey of current
numerical prediction capabilities. Advisory report 319. AGARD
2:135
Piponniau S, Dussauge J, Debieve J, Dupont P (2009) A simple model
for low-frequency unsteadiness in shock-induced separation.
J Fluid Mech 629:87108
Pirozzoli S, Grasso F (2006) Direct numerical simulation of
impinging shock wave/turbulent boundary layer interaction at
M = 2.25. Phys Fluids 18:065113-1065113-17
Selig M, Andreopoulos J, Muck K, Dussauge J, Smits A (1989)
Turbulence structure in a shock wave/turbulent boundary-layer
interaction. AIAA J 27:862869
Settles G, Vas I, Bogdonoff S (1976) Details of a shock-separated
turbulent boundary layer at a compression corner. AIAA J
14:17091715
Smits A, Muck K (1987) Experimental study of three shock wave/
turbulent boundary layer interactions. J Fluid Mech 182:291314
Souverein L (2010) On the scaling and unsteadiness of shock induced
separation. Doctoral Thesis, LUniversite de Provence
Souverein L, Dupont P, Debieve J, Dussauge J, van Oudheusden B,
Scarano F (2010) Effect of interaction strength on unsteadiness
in turbulent shock-wave-induced separations. AIAA J 48:1480
1493
Touber E, Sandham N (2009) Comparison of three large-eddy
simulations of shock-induced turbulent separation bubbles.
Shock Waves 19:469478
Van Driest E (1951) Turbulent boundary layer in compressible uids.
J Aeronaut Sci 18:145160
Wu M, Martin M (2008) Analysis of shock motion in shockwave and
turbulent boundary layer interaction using direct numerical
simulation data. J Fluid Mech 594:7183
1368 Exp Fluids (2012) 53:13471368
1 3

You might also like