You are on page 1of 18

ARTICLE IN PRESS

Optics and Lasers in Engineering 44 (2006) 190207

Schlieren PIV for turbulent ows


Dennis R. Jonassen, Gary S. Settles, Michael D. Tronosky
Gas Dynamics Lab, Mechanical and Nuclear Engineering Department, Penn State University,
University Park, PA 16802, USA
Available online 31 May 2005

Abstract
The possibility of using commercial PIV equipment combined with schlieren optics to
measure the velocity elds of turbulent ows is explored. Given a sufciently high Reynolds
number and adequate refractive ow differences, turbulent eddies can serve as the PIV
particles in a schlieren image or shadowgram. The PIV software analyzes motion between
consecutive schlieren or shadowgraph frames to obtain velocity elds. Velocimetry examples
of an axisymmetric sonic helium jet in air and a 2D turbulent boundary layer at Mach 3 are
shown. Due to optical path integration, axisymmetric ows require the inverse Abel transform
to extract center-plane velocity data. Conditions for optimum schlieren sensitivity are
examined. In its present embodiment, schlieren PIV is not useful for laminar ows nor for
fully 3D ows. Otherwise it functions much like standard PIV under conditions where
individual particles are not resolved and velocimetry is instead based on correlation of the
motion of turbulent structures. Schlieren PIV shows signicant promise for general
refractive turbulent ow velocimetry if its integrative nature can be overcome through sharpfocusing optics.
r 2005 Elsevier Ltd. All rights reserved.
Keywords: Schlieren imaging; PIV; Velocimetry; Turbulent ows; Jets; Boundary layers

1. Introduction and literature review


Particle image velocimetry (PIV) has recently become the most important new tool
of experimental uid dynamics. Two consecutive images of a particle-laden ow are
Corresponding author. Tel.: +1 814 863 1504; fax: +1 814 865 0118.

E-mail address: gss2@psu.edu (G.S. Settles).


0143-8166/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.optlaseng.2005.04.004

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

191

interrogated for small particle displacements, from which a slice through a velocity
eld is measured. This depends upon modern digital camera equipment and desktop
computing capability previously beyond the reach of the individual investigator, but
now commercial PIV systems with convenient user interfaces are available to anyone
who can afford them.
One of the main issues in PIV is seeding the oweld with particles. These particles
should match the density of the uid in order to avoid gravitational forces, and at the
same time must not change the ow dynamics [1]. However, if the oweld is
sufciently turbulent and this turbulence involves refractive-index changes, then
turbulent eddies themselves might act as PIV particles when the ow is viewed
using either schlieren or shadowgraph methods. In such a case the oweld is selfseeding. An underlying assumption here is that the evolutionary time scale of the
eddies is much longer than the time separation of the images in a PIV pair, otherwise
no PIV correlation can occur. Likewise some ow symmetry, e.g. planar or
axisymmetric, is required in order that integrative optical methods may present an
interpretable view of a ow.
The aim of the present approach is to explore such velocimetry, not the
measurement of the refractive-index eld per se. The use of the schlieren method and
PIV software for the related purpose of refractive-index or temperature measurement is described by others, e.g. [2,3].
Schlieren velocimetry was rst tried in 1936 by Townend [4], but was not pursued in
that pre-computer era. In 1989, Papamoschou [5] used a pattern-matching algorithm
to compute the convective velocity of a supersonic shear layer from two consecutive
schlieren images. Since then, Tokumaru and Dimotakis [6] proposed an ambitious
method of image correlation velocimetry for uid ows, while Fu and Wu applied
image analysis to extract velocity elds from sequences of schlieren images [79].
Kegerise and Settles also performed schlieren velocimetry on a convective plume [10].
All the cited velocimetry examples require specic home-grown algorithms to
extract the desired velocity data from schlieren image sequences. In contrast, the
present goal is to explore the validity of using a commercially available PIV system
and software to measure velocity elds from schlieren images and shadowgrams.

2. Experimental equipment
The PIV system used here is a single-camera system made by IDT Inc. [11]. Its
software offers two different algorithms for extracting displacements from an image
pair: standard and adaptive interrogation modes. The former is the cross-correlation
approach most often used in PIV, but is prone to errors such as loss of pairing, image
truncation, and spatial averaging of velocity gradients. The adaptive interrogation
mode [12] is designed to reduce or avoid such errors. Although both algorithms
produce similar results in the present test cases, only the adaptive interrogation
mode has been used in obtaining the results presented here. Insofar as schlieren
PIV measurements are concerned, comparable equipment and software by other
manufacturers is expected to function similarly.

ARTICLE IN PRESS
192

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

Three different schlieren optical instruments were used in this study, one lensbased and two mirror-based systems [13]. The lens-based system had two 152 mm
diameter f =5:67 telescope objective lenses as eld elements. The mirror-based
systems had z-type layouts with twin f/8 parabolic mirrors of either 108 or 152 mm
diameter as eld elements.
Special schlieren light sources are required in order to provide pulsed illumination
for PIV. For low- to moderate-speed ows, inexpensive microsecond white-light
xenon strobe lamps sufce, e.g. [14]. Two such lamps are mounted perpendicularly to
one another as shown in Fig. 1. A 50/50 beamsplitter combines the beams from these
lamps and directs them, via a condenser lens and slit, along the optical axis of the
present 108 mm-aperture z-type schlieren system. Neutral density lters are required
after each bulb to balance the illumination from the two ashes. One ashbulb is
triggered by PIV trigger signal A from the controlling computer and timing box, the
second by trigger signal B. In our case, since the PIV system is designed for pulsed
lasers, proper strobe timing required an external delay circuit between the PIV timing
box and the rst strobe lamp [15].
For high-speed ows requiring time separations less than 5 ms between images A
and B of a PIV pair, strobe illumination fails and pulsed laser illumination is
required. This is provided by a New Wave Gemini 200 dual-head Nd:YAG laser.
This powerful laser (200 mJ/pulse) needs strong attenuation for use as a schlieren
illuminator, here accomplished by two beam reections from the planar sides of
fused silica plano-concave lenses. Finally, a beam expander overlls the rst mirror
of the 152 mm-diameter z-type schlieren system to yield approximately uniform laser
illumination.
Dual pulsed-laser illumination has the distinct advantage of a virtually unlimited
range of pulse-separation timing. However, coherent laser light produces schlieren
images that are inferior to those produced by non-coherent white light for reasons
given in [13] and [16]. Briey, the geometric-optical approach to schlieren system
performance breaks down and a very small focal spot occurs in the cutoff plane. The
rst issue produces coherent artifact noise and a reduced signal-to-noise ratio
compared to the usual evenly illuminated white-light schlieren image. The second

Fig. 1. Diagram of the twin strobe arrangement for schlieren PIV.

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

193

issue rules out the traditional knife-edge cutoff for laser illumination, requiring
instead one of several milder cutoff options (graded lter, partially transmitting
cutoff, etc.). Here, a simple half-plane sooted microscope slide cutoff [13] is used.
The schlieren sensitivity using this cutoff is determined by the optical transmission of
the sooted section, which is measured by transmission densitometry. Within limits,
greater optical transmission of the cutoff yields less sensitivity in a laser-illuminated
schlieren system and vice versa. By comparison, schlieren sensitivity with a
traditional extended white-light source and a knife-edge cutoff is determined by
the amount of cutoff and the width of the light source image.
All of the present schlieren instruments are easily converted to focused
shadowgraphy [13] by the removal of the cutoff, image focus adjustment, and
reduction of the source slit size to an effective point in the case of white-light
illumination. Of course, these optical methods present somewhat different views of a
oweld, but both reveal refractive turbulence, thus both are included in principle
under the general name schlieren PIV.
Two different turbulent ows, both 2D in the mean, are used here to test the
validity of the schlieren PIV concept: a planar turbulent boundary layer and an
axisymmetric turbulent jet. The boundary layer is formed on the test section oor of
the Penn State Supersonic Wind Tunnel, a cold-ow blowdown facility with a
15  16.3  60 cm test section, up to 2 min test duration, and a Mach number range
of 1.54.0. Present tests are conducted at Mach 3, boundary-layer thickness
d 25 mm, and momentum-thickness Reynolds number Rey 98000. The lens-type
schlieren system described above was used in these experiments.
A converging nozzle of exit diameter 0.787 mm, discharging helium into ambient
air, produces the axisymmetric jet tested here. The nozzle stagnation pressure of
208 kPa yields a choked helium jet with a theoretical exit velocity of 890 m/s and a
Reynolds number of about 7300. The small nozzle scale allows a large nondimensional jet length to be studied (up to about 200 nozzle diameters downstream)
within the elds-of-view of the z-type schlieren instruments employed here.

3. Schlieren sensitivity to turbulence


As in traditional PIV, image quality is a key issue in schlieren PIV. The image scale
and resolution are determined by the nature of the CCD sensor in the PIV camera
and by the lens employed, and vary with the scale of the ow under study.
Aside from these traditional concerns, however, schlieren PIV has additional
image quality issues arising from the sensitivity of the optics to the refractive
disturbances in the turbulent ows under investigation. Schlieren sensitivity is here
analogous to particle size selection and seeding density in traditional PIV. For both
schlieren and shadowgraphy an optimum sensitivity range can be dened over which
the best PIV results are had.
In simple geometricoptical terms, schlieren sensitivity S is given by Eq. (1)
(adapted from [13]), where a is the width of the unobstructed light-source image in
the schlieren cutoff plane, b is the width of the entire light-source image, and f 2 is the

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

194

focal length of the second schlieren eld lens or mirror. Since f 2 and b are typically
xed for a given schlieren instrument, we may conveniently discuss sensitivity in
terms of the percent of knife-edge cutoff of the light-source image. For example,
100% cutoff refers to the complete blockage of the source image by the knife-edge (a
situation in which Eq. (1) predicts innite sensitivity, the geometricoptical
description fails, and diffraction effects take over [13]).
S

f2
f2

 .

a
b 1  %cutoff=100

(1)

For the white-light strobe-illuminated schlieren system described earlier, a razor


blade knife-edge oriented perpendicular to the streamwise direction of the ow is
used as the cutoff. Fig. 2 shows the results of the range of cutoffs used with the
present helium jet. For 10% and 20% cutoff only an underlying very-slight
shadowgraph effect is seen. These images contain insufcient information for PIV
processing. On the other hand, high-cutoff images result in over-ranging, shown for
example by the entirely white regions in the 90%-cutoff image in Fig. 2. When overranging occurs the local ow details are washed out, which is likewise detrimental to
schlieren PIV.
To avoid over-ranging, especially when high cutoff is required by a need for high
schlieren sensitivity, one can increase the overall measuring range by increasing the
light-source slit width b. Better solutions, such as replacing the knife-edge by a
graded lter [13] are also available.
The optimum case for schlieren PIV requires sufcient sensitivity to reveal the
smallest turbulent structures without signicant over-ranging. This occurs, for
example, in the 3060% cutoff range in Fig. 2. Here, unlike the case in some other
schlieren applications, it is the ne-scale turbulencePIV particlesthat must be
revealed.
As noted earlier, laser schlieren illumination requires a less-severe cutoff than the
knife-edge just described, whose optical transmission f is 0%. In Fig. 3, different halfplane-sooted microscope slides are used as cutoffs to vary the schlieren sensitivity.
Cutoffs with fp26%, being almost as opaque as a knife-edge, yield binarized images
in which some or all of the ne turbulence scales are washed out. Likewise the f
93% cutoff is almost transparent, yielding practically no cutoff and a vanishing
schlieren effect. Therefore, the present optimum sensitivity range for laser-illuminated
schlieren with partially transmitting cutoffs is roughly 50%ofo90%.
Coherent artifact noise is quite apparent in Fig. 3. It has the undesirable effect in
schlieren PIV of obscuring some of the ne-scale turbulent motion. Moreover, with
twin pulsed-laser illumination a pseudo-velocity can result from the coherent artifact
noise if the two laser beams are not aligned with perfect coincidence.
As already noted, focused shadowgram pairs can also be used for PIV analysis.
Here, since shadowgrams involve no cutoff, the product of the focus offset distance
L and the Laplacian of the refractive-index eld dictates the sensitivity [13]. Fig. 4
shows an example of laser-illuminated shadowgrams of the helium jet with L ranging
from 0 to 130 cm. As expected, the sharply focused case yields no usable information
for PIV analysis. Shadowgraph PIV becomes possible in this example when L

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

195

Fig. 2. Helium-jet white-light schlieren images with different amounts of knife-edge cutoff (b 0:86 mm
source-slit width).

Fig. 3. Helium-jet laser-schlieren results with different optical transmission values f of the half-plane
sooted microscope slide cutoff: (a) f 13%, (b) 26%, (c) 61%, (d) 72%, (e) 88%, and (f) 93%.

ARTICLE IN PRESS
196

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

Fig. 4. Helium-jet shadowgrams with increasing sensitivity: (a) L 0 cm, (b) 12.5 cm, (c) 38 cm, (d) 53 cm,
and (e) 130 cm.

exceeds 12 cm, but fails for L4130 cm because the increasing blur eventually
obscures the ne-scale turbulence that serves as PIV particles. These L limits will
be different, of course, for schlieren objects of different strength than the present
helium jet.
Note especially that a shadowgram is not a focused image, and that coherent
artifact noise in the case of laser illumination can pose the same troublesome issue of
pseudo-velocity as described above for schlieren PIV.
To better understand exactly how PIV algorithms perceive schlieren images,
120  120 pixel turbulence samples were taken from each image in Fig. 2 and
converted into binarized 2D intensity maps, which were then analyzed for the number
and size of the particles within a 24  24 pixel interrogation window. For simplicity,
no window overlap was used in the determination of the number of particles per
window, and the particles found at the edges of the window were included in the
particle count. As shown in Fig. 5, the number of apparent particles decreases with
increasing knife-edge cutoff, while the average particle size increases.
Keane and Adrian [17] proposed a method to quantify the minimum number of
particles required in a PIV interrogation window. It is based on the number of
particles N, the out-of-plane displacement F 0 , and the in-plane displacement F i . In
order to achieve 95% accuracy in the probability of displacement detection, they
claim that the product of these three parameters should be larger than ve. However,
Raffel et al. [1] claim that a value of three or four for this product is sufcient for
practical situations. When applied to schlieren PIV, F 0 equals unity since all
particles are always in the plane. F i can likewise be set essentially to unity when
the adaptive interrogation mode is used. Thus the number of schlieren particles in
the PIV interrogation window should be at least four.
Fig. 5 reveals that a knife-edge cutoff of 50% or less thus ensures an adequate
number of particles in a 24  24 pixel window for the schlieren equipment used
here (f 2 864 mm, b 0:86 mm). Cutoffs exceeding 50% may yield interrogation
windows without particles, thereby requiring software interpolation to compute
displacements.

ARTICLE IN PRESS

16

16

14

14

12

12

10

10

197

Number of Apparent "Particles" in a


24 x 24 pixel interrogaiton window

Mean Apparent "Particle" Diameter [pixel]

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

0
0

10

20

30

40

50

60

70

80

90

100

Percent Cutoff
Fig. 5. Particle size and concentration vs. percent schlieren cutoff (based on Fig. 2 and using a 108 mm
diameter f =8 z-type schlieren system with b 0:86 mm source-slit width).

Particle size is likewise a key parameter in PIV analysis. According to Raffel et al.
[1], the optimum particle diameter for conventional PIV is two pixels, based on a
simulation of particles traveling one pixel per frame in a light sheet. In this controlled
case, particle size is directly related to the RMS uncertainty of particle displacement.
Above a particle diameter of two pixels, however, the RMS uncertainty increases
exponentially, although it decreases as the interrogation window grows [1]. Based on
this, particle diameters should be minimized for successful schlieren PIV, especially
when small interrogation windows are needed for high spatial resolution. Fig. 5 shows
that the present particle size increases almost linearly over the range of 1060%
cutoff, above which over-ranging of the helium-jet schlieren image begins.
We therefore conclude that effective schlieren PIV calls for the minimum knifeedge cutoff compatible with particle visibility. One may anticipate extreme cases
where a combination of low cutoff and a weak turbulent refractive eld prevents the
success of schlieren PIV. Even for the present strongly refractive helium jet, a cutoff
in the ideal range of 10% produces insufcient contrast between the turbulence
and the background, i.e. a signal-to-noise ratio that is too small to be effective. Thus
the optimum schlieren sensitivity for PIV measurements in the present experiment
occurs within the 3060% range of knife-edge cutoff using white-light illumination.

4. PIV image processing


Processing schlieren image pairs with commercially available PIV software is no
different than in standard PIV. The rst consideration is spatial resolution: high

ARTICLE IN PRESS
198

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

resolution of the velocity eld requires small interrogation windows. But errors can
result due to the particle size and concentration constraints noted above. Smaller
windows also increase the probability of other errors such as loss of pairing and
particle truncation. Both errors result in velocity underestimation [12]. The adaptive
interrogation window described earlier alleviates most of these problems.
To further increase the probability of detection, a PIV image offset may sometimes
be required. This literally shifts the second image of the PIV pair with respect to the
rst by a predetermined integer pixel value. Such an offset helps the correlation
algorithm to determine sub-pixel displacements, permitting the use of smaller
interrogation windows.
Certain situations further require that the PIV eld be analyzed in zones, rather
than as a whole. For example, high velocity gradients can produce neighboring
regions of a oweld requiring different image offsets (often by just a single pixel).
This zonal evaluation technique was required in analyzing the present compressible
boundary layer test case.

5. Schlieren PIV results


5.1. Helium jet
The helium jet studied here has distinct near- and far-eld zones. The near-eld
extends to about 15 nozzle diameters downstream of the nozzle exit. Its dening
feature is its laminar core, containing no turbulent structures (particles) for schlieren
PIV. The effect is clearly seen in Fig. 6: a double peak appears in the integrated
velocity prole at about y 8 mm from the nozzle exit. The reason for this is the
path-integration of the schlieren beam through this axisymmetric ow. A light ray
through the edge of the laminar jet core incurs more refraction due to turbulence
than one traversing the jet centerline. The former ray thus yields a greater apparent
particle displacement and a higher velocity is found.
No schlieren PIV is possible within y 1 mm of the nozzle exit, where the
accompanying shadowgram reveals an initially laminar jet. Also, due to optical
path integration and turbulent mixing, the measured velocities in the 40 m/s range in
Fig. 6 are low compared to the theoretical jet nozzle-exit speed of 890 m/s.
The far-eld jet region begins around 15 diameters (12 mm) from the nozzle exit
and is fully turbulent. Fig. 7 shows laser- and white-light-illuminated schlieren
images and a laser shadowgram of this region. The white-light schlieren image,
Fig. 7b, has a more severe knife-edge cutoff than the laser schlieren image, and both
laser frames have coherent artifact noise that can lead to velocity errors.
Nonetheless, each of these three varied depictions of the helium jet does in fact
yield, with good accuracy, the same convective velocity eld upon schlieren PIV
analysis (150 image pairs at a time delay of 5 ms between paired images). The
resulting mean velocity prole normal to the jet axis is given in Fig. 8.
For comparison, traditional laser-sheet-illuminated PIV measurements were made
using micron-sized oil droplets that were entrained into the helium jet by way of a

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

199

Fig. 6. Laser shadowgram of the helium jet near eld (left, L20 cm, nozzle exit diameter 0.787 mm)
and the corresponding schlieren PIV velocity contour plot (right). Velocities are given in m/s.

Fig. 7. (a) Laser schlieren, (b) white-light schlieren, and (c) laser shadowgram of helium jet showing
far-eld.

slow coaxial seeded jet. Only far-eld measurements were possible in this manner.
Example oweld images and PIV velocity results are compared in Fig. 9. It is clear
from this comparison that traditional PIV yields higher velocities than schlieren PIV
in this helium jet case, as expected. The difference is attributed to the pathintegration of the schlieren optics. Since the jet is axisymmetric, a proper comparison

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

200

Axial Velocity, path-averaged (m/s)

30
Schlieren, white-light illumination
Schlieren, laser illumination
Shadowgraph, laser illumination

25

20

15

10

0
-15

-10

-5

0
X (mm)

10

15

Fig. 8. Mean helium-jet velocity proles from schlieren PIV analysis of the cases of Fig. 7 (y 42 mm
downstream of nozzle exit). Only half-proles are shown for the laser schlieren and shadowgraph cases.

can be made by applying the Abel transform to the traditional PIV results, Eq. (2)
(see, e.g. [18]).
iP
edge

V pa i; j

V pl i; j

iio

N io !iedge

(2)

where V pa i; j is the path-averaged velocity, V pl i; j is the planar velocity, N io !iedge is


the number of discrete points between io and iedge , and the subscripts o and edge
represent the location of V pa i; j within the jet and the jet edge, respectively. The
result is shown in Fig. 10, where the Abel transform applied to the traditional-PIV jet
center-plane data recovers approximately the same velocity prole as that measured
directly by schlieren PIV.
Alternatively, the inverse Abel transform can be applied to the schlieren PIV
results to yield equivalent jet center-plane data. In most cases this is the preferable
procedure, since the equivalent center-plane data are the most useful form of the
experimental results. The opposite procedure has been adopted here merely as a
matter of convenience, since deconvolution of the data makes the inverse Abel
transform approach somewhat more involved. Nonetheless modern digital computing power is more than sufcient for either of these procedures.
To further test the schlieren PIV results, one may compare the jet centerline
velocity distribution vs. y with traditional PIV results and with the extensive jet data

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

201

Fig. 9. (Above left) white-light helium-jet schlieren image, (above right) particle-seeded image of the same
ow with laser sheet illumination through the jet center-plane. Below the images are shown average PIV
velocity contour maps for comparison (schlieren PIV left, traditional PIV right).

correlation by Kleinstein and Witze [19]:


!
1
uc 1  exp
,
 0:5
kx re
 Xc

(3)

where k 0:074, x x=rj , re re =rj (e and j refer to the ambient and jet exit
conditions, respectively), and X c 0:7. The centerline velocity is normalized by the
jet nozzle-exit velocity, 890 m/s.

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

202
50

Axial Velocity (m/s)

40

Particle PIV
Particle PIV, Abel transformed
Schlieren PIV

30

20

10

0
0

10

X (mm)

Fig. 10. Velocity half-proles of the far-eld helium jet at y 65 mm, comparing schlieren PIV with
traditional particle PIV results with and without the application of the Abel transform to the latter.

The results show that the present helium-jet centerline velocity data, obtained by
traditional PIV in two separate experiments, agree well with the KleinsteinWitze
correlation as shown in Fig. 11. Also, applying the Abel transform to the traditional
PIV data yields results comparable to the schlieren PIV data, demonstrating once
again that schlieren PIV gives a path-integrated representation of the axisymmetric velocity eld. The noise level of the traditional PIV data in Fig. 11 makes it
difcult to determine the exact jet edge location, which may account for some error
incurred during the Abel transformation.
5.2. Compressible turbulent boundary layer
The above helium-jet schlieren PIV results reveal the dominant effect of optical
path integration in measuring an axisymmetric oweld. No such effect is expected
in a true 2D ow, however, where schlieren PIV and traditional PIV results should
be directly comparable even though the former still integrates across the ow.
To test this, the compressible turbulent boundary layer on the test-section oor of
the Penn State Supersonic Wind Tunnel was measured by schlieren PIV. Given a
span of 6d and end effects of less than one d, approximately 2D ow is expected.
Note, however, that it is not possible with the current schlieren optics to eliminate
end effects by focusing on the center-plane of this ow.
Double-pulsed laser illumination through the lens-type schlieren system described
earlier is the only approach that allows a sufciently small time delay between PIV
frames (0.55 ms) for this Mach 3 ow. To provide adequate schlieren sensitivity a

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

203

Fig. 11. Helium jet centerline velocity decay as measured by traditional PIV, predicted by the KleinsteinWitze correlation, and measured by schlieren PIV with path integration.

half-plane sooted microscope slide with 72% light transmission was required. Given
the 15 cm extent of the boundary layer along the optical axis, converting to
shadowgraphy required only the removal of the knife-edge, without any focus
adjustment. (In other words, the shadowgraph image is always present and is
combined with the schlieren image when a schlieren cutoff is used.)
An example of schlieren image and shadowgram are given in Fig. 12. In both cases
ne-scale turbulence is observed, though it does not end as expected at the
boundary-layer edge, but rather appears to continue into the freestream. This is, in
fact, an end effect due to the wind-tunnel sidewall boundary layers on the glass
windows. Slight over-ranging is also seen in the schlieren image, but this is not
serious enough to affect the overall PIV results.
One hundred and fty schlieren PIV image pairs were processed, as before, using
the IDT softwares adaptive mode with a 20  20 pixel interrogation window.
However, since the boundary-layer velocity prole extends from 0 at the wall to
613 m/s in the freestream, the required image offset varies across the height of the
boundary layer as discussed earlier, starting at 6 pixels near the wall and ending at
8 pixels near the freestream. Since three separate image-offset zones are needed to
analyze this ow, three different meshes are also required. The zones with their mesh
locations and corresponding image offsets are as follows:
1. 0.0py/dp0.17 6-pixel offset,
2. 0.14py/dp0.30 7-pixel offset,
3. 0.27py/dp1.01 8-pixel offset.

ARTICLE IN PRESS
204

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

Fig. 12. Example schlieren image (left) and shadowgram (right) of the Mach 3 turbulent boundary layer
on the test section oor of the Penn State Supersonic Wind Tunnel.

Unfortunately, traditional particle PIV is not currently possible in this boundary


layer due to particle seeding difculties. Instead, the schlieren PIV results are
evaluated by comparison with mean velocity proles derived from pitot-pressure
surveys by Garg and Settles [20], and with the established wake-wall similarity law
for turbulent boundary layer proles. The Van Driest transformation was used to
convert measured compressible-ow velocity values, u, obtained from both the
optical PIV and the pitot surveys, to equivalent incompressible values, u* (see, e.g.
[21]). The results are plotted in Fig. 13 in traditional u+ vs. y+ coordinates along
with the incompressible wall-wake law of Coles [22].
Fig. 13 reveals that the schlieren PIV data are in substantial agreement with the
pitot-survey results for this boundary layer, even though the differences in the two
measurement methods are striking. Schlieren PIV measures, in principle, the
broadband convective velocity of turbulent structures directly. Pitot pressure surveys
are converted, with the assumption of constant stagnation temperature, to mean
Mach number proles from which the mean velocity is extracted by way of an
assumed static temperature prole. Agreement between the two indicates that
schlieren PIV actually measures, by way of refractive eddy motion, the mean velocity
prole of the boundary layer.
Fig. 13 also reveals that this boundary layer has an unusually high wake-strength
component, P 1:9 vs. the usual P0:55 for a at-plate boundary layer [22]. This
doubtless results from the pressure-gradient history along the wall of the wind
tunnels long asymmetric variable-Mach-number sliding-block nozzle.
There are two boundary-layer regions in which schlieren PIV is unable to
measure the velocity satisfactorily: very near the wall and near the boundary-layer
edge. Failure near the wall is due to poor spatial resolution caused by wind-tunnel
vibrations and by laser diffraction that obscures the true location of the wall. Even
so, schlieren PIVbeing non-intrusivemeasures closer to the wall than does the

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

205

40
Pitot Pressure Survey
Wall Law
Shadowgraph PIV
Schlieren PIV
Wall-Wake Law

U+ = u*/u

35

30

M=3
Tw = 260 K
= 25 mm
cf = 8.98 x 10-4
u = 18.5 m/s
w = 2.52 x 10-5 m2/s
Re = 98,000
= 1.9

25

20

15
102

103

104

105

Y+ = y(u/w)

Fig. 13. Mean velocity proles of the Mach 3 turbulent boundary layer in wall-wake coordinates.

pitot survey method. Near the boundary-layer edge, however, schlieren PIV fails
by yielding velocities that are too low. The outer 20% of the turbulent boundary
layer is quite intermittent, so the eddies serving as PIV particles are fewer there.
Eddies in the sidewall boundary layers contaminate the integrated schlieren
measurement with incorrect end-effect velocities in this region, as shown in Fig.
12. By way of focusing schlieren optics [20,23], it should be possible to eliminate
these end effects and make an accurate measurement up to the boundary-layer edge,
but that is beyond the present scope.

6. Conclusion
Schlieren PIV is shown to yield valid velocimetry data, within certain limits, for
a 2D compressible turbulent boundary layer and an axisymmetric turbulent helium
jet in air. Effective schlieren PIV calls for the minimum knife-edge cutoff
compatible with the visibility of ne-scale turbulence. Due to optical path
integration, axisymmetric ows require the inverse Abel transform to extract
center-plane velocity data. Path integration also causes velocimetry errors due to end
effects in the boundary-layer experiment. In both ows studied here the evolutionary
time scale of turbulent eddies was much longer than the proper time separation
between the images in a PIV pair, so eddy evolution did not limit the schlieren PIV
measurements.
Despite its limitations, the method shows promise as a new optical velocimetry
tool. Schlieren PIV combines commercially available PIV equipment and software

ARTICLE IN PRESS
206

D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

with a standard schlieren optical instrument. It requires no actual particle seeding of


the ow, since it uses ne-scale turbulence as PIV particles. Otherwise it functions
much the way standard PIV does under conditions where individual particles are not
resolved and velocimetry is instead based on correlation of the motion of turbulent
structures. In its present embodiment, schlieren PIV is not useful for laminar ows
or for fully 3D ows, but it shows signicant promise for general refractive turbulent
ow velocimetry if its integrative nature can be overcome through sharp-focusing
schlieren optics [20,23]. A study of that possibility is recommended for future work,
as is the potential use of simple means to generate non-coherent schlieren
illumination directly from coherent PIV laser pulses [24]. Additional information
on the present study is given in [25].

References
[1] Raffel M, Willert C, Kompenhans J. Particle image velocimetry. Berlin: Springer; 1998.
[2] Raffel M, Richard H, Meier GEA. On the applicability of background oriented optical tomography
for large scale aerodynamic investigations. Exp.Fluids 2000;28:47781.
[3] Papadopoulos G. Novel shadow image velocimetry technique for inferring temperature.
J Thermophys Heat Transfer 2000;14(4):593600.
[4] Townend HCH. A method of airow cinematography capable of quantitative analysis. J Aero Sci
1936;3(10):34352.
[5] Papamoschou DI. A two-spark schlieren system for very-high velocity measurement. Exp Fluids
1989;7(5):3546.
[6] Tokumaru PT, Dimotakis PE. Image correlation velocimetry. Exp Fluids 1995;19:115.
[7] Fu S, Wu Y. Detection of velocity distribution of a ow eld using sequences of Schlieren images. Opt
Eng 2001;40(8):16616.
[8] Fu S, Wu Y. Quantitative analysis of velocity distribution from schlieren images. In: Carlomagno
GM, editor. Proceedings of the eighth international symposium on ow visualization, Sorrento.
Paper 233, 1998.
[9] Fu S, Wu Y, Kothari RD, Xing H. Flow visualization using the negative-positive grid schlieren
system and its image analysis. In: Grant I, Carlomagno GM, editors. Proceedings of the ninth
international symposium on ow visualization, Edinburgh. Paper 324, 2000.
[10] Kegerise MA, Settles GS. Schlieren image-correlation velocimetry and its application to freeconvection ows. In: Grant I, Carlomagno GM, editors. Proceedings of the ninth international
symposium on ow visualization, Edinburgh. Paper 380, 2000.
[11] IDT Inc., Tallahassee, FL, USA, http://www.idtpiv.com.
[12] Lourenco LM, Krothapalli A. TRUE resolution PIV: a mesh-free second order accurate algorithm.
In: Proceedings of the international conference on applications of laser uid mechanics, Lisbon, 2000.
[13] Settles GS. Schlieren and shadowgraph techniques, 1st ed. Berlin: Springer; 2001.
[14] Model 1531-AB Genrad Strobotac, IET Labs, Inc., (516) 334-5959, sales@ietlabs.com, http://
www.ietlabs.com/.
[15] Tronosky MD. Schlieren PIV. MS thesis, Department of Mechanical and Nuclear Engineering, Penn
State University, 2003.
[16] Oppenheim AK, Urtiew PA, Weinberg FJ. On the use of laser light sources in schliereninterferometer systems. Proc Roy Soc A 1966;291:27990.
[17] Keane RD, Adrian RJ. Optimization of particle image velocimeters. Part I: double pulsed systems.
Meas Sci Technol 1992;1:96374.
[18] Watt DW, Donker Duyvis FJ, van Oudeusden BW, Bannink WJ. Calibrated schlieren and
incomplete Abel inversion for the study of axisymmetric wind tunnel ows. In: Grant I, Carlomagno

ARTICLE IN PRESS
D.R. Jonassen et al. / Optics and Lasers in Engineering 44 (2006) 190207

[19]
[20]
[21]
[22]
[23]
[24]

[25]

207

GM, editors. Proceedings of the ninth international symposium on ow visualization, Edinburgh.


Paper 363, 2000.
Witze PO. Centerline velocity decay of compressible free jets. AIAA J 1974;12(4):4178.
Garg S, Settles GS. Measurements of a supersonic turbulent boundary layer by focusing schlieren
deectometry. Exp Fluids 1998;25:25464.
White FM. Viscous uid ow, 2nd ed. Boston: McGraw-Hill; 1997.
Coles D. The young persons guide to the data. In: Proceedings of AFOSR-IFP-Stanford Conference
Computation Turb BLs, vol. 2. Stanford University Press; 1968. p. 145.
Alvi FS, Settles GS, Weinstein LM. A sharp-focusing schlieren optical deectometer. AIAA paper
93-0629.
Murphy MJ, Adrian RJ, Elliott GS, Thomas KA, Kennedy JE. Visualization in extreme
environments. In: Mueller TJ, Grant I, editors. Proceedings of the 11th international symposium
on ow visualization, Notre Dame University paper 155, 2000.
Jonassen DR. Schlieren PIV and its application to two-dimensional and axisymmetric turbulent
ows. MS thesis, Department of Mechanical and Nuclear Engineering, Penn State University, 2004.

You might also like