Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Exp Fluids (2015) 56:17

DOI 10.1007/s00348-014-1892-4

RESEARCH ARTICLE

High‑speed particle image velocimetry for the efficient


measurement of turbulence statistics
Christian E. Willert

Received: 22 October 2013 / Revised: 19 December 2014 / Accepted: 19 December 2014 / Published online: 13 January 2015
© Springer-Verlag Berlin Heidelberg 2015

Abstract A high-frame-rate camera and a continu- Abbreviations


ous-wave laser are used to capture long particle image cf Friction coefficient
sequences exceeding 100,000 consecutive frames at fram- f Focal length
ing frequencies up to 20 kHz. The electronic shutter of f# f-Number (lens aperture)
the high-speed CMOS camera is reduced to 10 µs to pre- IW Interrogation window size
vent excessive particle image streaking. The combination M Magnification
of large image number and high frame rate is possible N Number of samples
by limiting the field of view to a narrow strip, primar- Rij Space-time correlation function of scalar
ily to capture temporally resolved profiles of velocity quantities i, j
and derived quantities, such as vorticity as well as higher Reδ Reynolds number based on U99 and δ99
order statistics. Multi-frame PIV processing algorithms Reδ ∗ Reynolds number based on U99 and δ ∗
are employed to improve the dynamic range of recovered Reθ Reynolds number based on U99 and θ
PIV data. The recovered data are temporally well resolved Reτ Reynolds number based on uτ and δ99
and provide sufficient samples for statistical convergence S Shape factor, δ ∗ /θ
of the fluctuating velocity components. The measure- t Time
ment technique is demonstrated on a spatially developing u, v, w Streamwise, wall-normal and spanwise velocity
turbulent boundary layer inside a small wind tunnel with components
Reδ = 4,800, Reτ = 240 and Reθ = 515. The chosen mag- u′ , v′ , w′ Fluctuating velocity components
nification permits a reliable estimation of the mean velocity uτ Skin friction velocity
profile down to a few wall units and yields statistical infor- Ucl Center line velocity
mation such as the Reynolds stress components and prob- U99  0.99 Ucl
ability density functions. By means of single-line correla- x, y, z Streamwise, wall-normal and spanwise
tion, it is further possible to extract the near-wall velocity coordinates
profile in the viscous sublayer, both time-averaged as well x + , y+ , z+ Distances in wall units based on uτ
as instantaneous, which permits the estimation the wall δ99 Boundary layer thickness at 99% Ucl
shear rate γ̇ and along with it the shear stress τw and fric- δ ∗ Boundary layer displacement thickness
tion velocity uτ. These data are then used for the calcula- t Time delay between two illumination pulses
tion of space-time correlation maps of wall shear stress and x, y Streamwise and wall-normal displacement in pixels
velocity.  Wavelength of light
ν Kinematic viscosity
θ Boundary layer momentum thickness
γ̇ Wall shear rate ≡ ∂u/∂yy=0
C. E. Willert (*)
τw Wall shear stress
German Aerospace Center (DLR), Institute for Propulsion
Technology, 51170 Cologne, Germany ωx , ωy , ωz Streamwise, wall-normal and spanwise vorticity
e-mail: chris.willert@dlr.de components

13
17 Page 2 of 17 Exp Fluids (2015) 56:17

1 Introduction The present article has exploratory character and intends


to bring to attention a number of previously less explored
The steady improvement of numerical methods for flow capabilities offered by high-speed imaging for turbulence
simulation relies on experimental data for their validation. research. The hardware requirements are minimal in that
In this regard, the PIV method (Adrian and Westerweel all high-speed measurements presented in the following
2010; Adrian 1991; Raffel et al. 2007) has gained increased involve only a single high-speed camera with sufficiently
popularity in the past two decades and has matured to the large memory and a strong continuous laser light source.
point of providing access to the complete velocity gradient Image data are acquired for a narrow strip. The correspond-
tensor (Elsinga et al. 2006; Scarano 2013). ing reduction in DSR is compensated through the availabil-
Inherently, the PIV technique has two important limita- ity of continuous time traces of velocity data from which
tions: on one hand, PIV relies on finite-sized imaging statistical quantities and spectra can be recovered. As the
arrays with edge lengths generally in the range of 1,000– image data is temporally over-sampled, multi-strided PIV
5,000 elements which along with the finite sampling win- processing can be utilized improving the DVR (Hain and
dow size (IW) limits the achievable dynamic spatial range Kähler 2007; Sciacchitano et al. 2012).
(DSR) to about 50. . .300 (for IW = 16 pixels). On the other The utilization of time-resolved PIV in turbulence
hand, the sampling window size defines the maximum research is by no means new as testified in the literature,
resolvable displacement to generally less than half the sam- especially with the wider availability of high-speed lasers
ple size1 and is bounded on the lower end by statistical and high-speed cameras from the mid-2000s onwards
noise. In practice, this random error is of the order of 0.05– (Dennis and Nickels 2011; LeHew et al. 2013). Even
0.1 pixel such that the achievable dynamic velocity range time-resolved tomographic PIV measurements have been
(DVR) typically is of the order of 100:1 (Adrian 1997, applied, for instance, to track and characterize vortex struc-
2005). tures within the turbulent boundary layer (Elsinga and
Another shortcoming arises due to the rather low sam- Marusic 2010; Gao et al. 2013; Schröder et al. 2008, 2011).
pling frequency of PIV with respect to the temporal time The approach presented in this work aims at bridging the
scales of the studied flow. Significant technological pro- gap between precise single-point measurements offered, for
gress has been made in recent years so that PIV image data instance, by HWA and LDA on the one end and the often
can be captured at up to 10,000 frames per second (fps) at a coarser but global/topological views offered through planar
resolution of one megapixel. While this is impressive in its PIV or tomographic PIV on the other end. By restricting
own right, the time scales in many technical flows require the measurement area to a narrow strip, both the acquisition
camera frame rates that are at least one order of magnitude frequency and sample count can be significantly increased,
higher, along with an equivalent increase in laser pulsing resulting in time-resolved velocity profile measurements.
frequencies. The first part of the following article describes a small
In the context of providing data for the validation wind tunnel facility used for the demonstration of the pro-
of numerical methods, the PIV technique is frequently posed measurement technique. Conventional (low speed)
removed of one of its most powerful capabilities, namely PIV measurements of the developing boundary layer inside
in providing spatially resolved instantaneous maps of the the test section are performed a priori to serve as reference.
investigated flow. It is common practice to first record a The second part is devoted to the high-speed PIV measure-
rather significant number of snapshots and then ensemble ment technique and provides several examples of making
average the recovered velocity data to provide maps of use of the time-resolved image data for estimating flow
the first and second moments of the flow field, that is, the statistics, spectra, probability density functions and cor-
mean and standard deviations of the velocity components relations. Furthermore, a method is presented for estimat-
along with the Reynolds stresses. Convergence of the lat- ing both the mean and unsteady wall shear stress from the
ter requires a large number of samples (N > 1,000) which acquired image data. The latter is then used for the calcula-
is not always feasible, in particular, when low-frame-rate tion of space-time correlation maps.
PIV systems are utilized on facilities with high operational While some of the experimental data presented in the
costs or short running times. Storage requirements for large following sections are compared to DNS data of a turbu-
data sets of high-resolution image data impose a further lent boundary layer (TBL), the reader should be aware of
constraint. the fact that the flow inside the small wind tunnel is not
fully compatible with the classical flat plate boundary layer
flow to which it is compared. The growing boundary layer
inside the confined geometry of the square duct results in
1
Displacements larger than 21IW are recoverable with advanced PIV an increasing centerline velocity and introduces a favorable
processing methods but impose other restrictions. pressure gradient. While agreement between measurement

13
Exp Fluids (2015) 56:17 Page 3 of 17 17

Fig. 1  Photograph of wind tun-


nel facility

Fig. 2  Schematic diagram


of wind tunnel facility with
approximate indication of
high-speed PIV measurement
location

and DNS is quite prevalent throughout the article, its prin- Table 1  Flow channel parameters
cipal aim is to point out the capabilities and shortcomings Quantity Symbol Value Unit
of the described PIV measurement approach.
Test section area d 76 × 76 mm2
Test section length L 830 mm
2 Flow facility Nozzle contraction ratio – 10:1 –
Turbulence grid, solidity S 37.7 % –
Turbulence measurements are performed in the near-wall Turbulence grid, rod spacing rod 7.6 mm
region of a developing turbulent boundary layer inside a Turbulence grid, rod width drod 1.6 mm
square duct. The test facility, shown in Fig. 1, is a small-
scale wind tunnel designed for the investigation of mix-
ing processes within confined geometries and is used in a
“clean” configuration for the present investigations, that the present study. Details of the wind tunnel are summa-
is, without mixing port or heated wall. The wind tunnel is rized in Table 1.
operated in suction mode and has a 830 mm long test sec-
tion with a square cross section of side length d = 76 mm. 2.1 Reference PIV measurements
Quartz windows with 45◦ edge bevels provide optical
access to the entire cross section. As outlined in Fig. 2, Conventional two-component, dual-frame PIV measure-
flow conditioning is provided by a settling chamber con- ments were performed a priori to serve as a basis for com-
taining screens and straightening tubes. Reproducible tur- parison with the high-speed PIV profile measurement data
bulent flow conditions within the test section are provided described later. All PIV measurements described here,
by a turbulence grid made of perforated steel that is placed including the high-speed imaging, rely on aerosol par-
immediately upstream of the test section at the exit of the ticles consisting of 1 µm paraffin droplets created by a
contraction nozzle. Aside from introducing turbulence to Laskin atomizer. An impactor in the seeding device sepa-
the bulk flow, the grid also trips the turbulent boundary lay- rates the large size fraction prior to introducing the seeded
ers on the walls of the test section which are the subject of air to a settling chamber upstream of the facility’s flow

13
17 Page 4 of 17 Exp Fluids (2015) 56:17

straighteners and screens (see Fig. 2). The seeded chan- of 7,000 fps at a resolution of 1,024 × 1,024 pixel and a
nel flow is illuminated by a dual cavity, frequency doubled PCO dimax HS4 with a maximum frame rate of 2,000 fps
Nd:YAG laser with a rated pulse energy of 50 mJ (Brilliant at 2,016 × 2,016 pixel. The cameras have sensor resolu-
Twins, Quantel/BigSky Laser). The laser light is spread into tions of 20 µm/pixel and 11 µm/pixel, respectively. For
a light sheet of about 150 mm width and has a thickness of the present application, the cameras are operated with a
about 1 mm in the area of interest. The scattered light is reduced field of view (FOV) of 1,024 × 256 which allows
imaged by a macro lens (Zeiss Macro-Planar, 100/f# 2.0) at the frame rate to be increased up to 20,000 fps. A macro
an aperture of f# 2.8 and magnification of m = 26.9 pixel/ lens (Zeiss Macro-Planar 100 mm/f# 2.0) images the illu-
mm (M = 5.72:1) onto a scientific CMOS camera array minated particles onto the sensor at a magnification of
with a sensor resolution of 2,560 × 2,160 pixels (ILA.PIV. M ≈ 1.6:1. Additionally, a f = 200 mm macro lens (Nikon
sCMOS, ILA GmbH). With the laser pulse separation set at Micro-Nikkor 200 mm/f4.0) is used to further increase the
t = 300 µs, particle displacements of x ≈ 32 pixels are magnification to M ≈ 1:1.
reached for the tunnel centerline velocity of Ucl = 4 m/s. Three acquired image sequences from each of the cam-
The rather long pulse separation is chosen to improve the eras are the subject of this study. Their characteristics are
measurement of the relatively low turbulence levels in tun- summarized in Table 2. All sequences were acquired at a
nel’s bulk flow. fixed laser power of 5W which translates to an integrated
The reference PIV data are processed using conventional energy of 100 µJ for a 20 µs exposure time of the camera.
PIV image analysis based on PIVview-v3.5 software (PIV- At this energy level, the particles are imaged by the Pho-
TEC GmbH), featuring a coarse-to-fine processing scheme tron camera at a moderate 20–100 counts on a uniform
starting with a sample size of 192 × 192 and a final size background intensity of about 26–27 counts. The PCO
of 24 × 16 (0.906 × 0.604 mm2). The final grid spacing is camera has an increased sensitivity, providing peak par-
6 × 6 (0.227 × 0.227 mm2). Validation relies on established ticle intensities of several 1,000 counts on a background
methods such as the normalized median filter (Westerweel of about 2,000 counts. Figure 3 shows two representative
and Scarano 2005). The valid vector rate exceeds 99 % in images of the acquired sequences using inverted intensity
the bulk flow decreasing to about 90 % in the near-wall, (black indicates high brightness). In flow direction, the
high-shear regions. imaged particles are streaked proportional to their velocity
and camera exposure time. Normal to the flow direction the
particle images generally cover less than two pixels which
3 Velocity profile measurements using high‑speed PIV can be attributed to the large pixel size of the high-speed
CMOS sensors. Owing to the constant intensity of the CW
3.1 Experimental hardware laser light source, the frame-to-frame intensity variation of
the particle images is minimal (e.g., within a few percent)
The high-speed particle image sequences are made pos- and well suited for PIV processing.
sible through the use of a high-power continuous-wave
(CW) laser light source. To reduce streaking of the particle
images on the camera sensor, the camera itself is electroni-
cally shuttered during the acquisition of each image frame.
Table 2  Acquired high-speed PIV image sequences and reference
The laser (Verdi 5, Coherent) produces up to 6W single- PIV measurement
frequency green laser light at a wave length of 532 nm. As
Sequence number Seq. 1a Seq. 1b Seq. 2 Ref.
the laser fluence is constant in time, short exposure times
Camera Photron Photron Dimax-HS4 sCMOS
and high framing rates significantly reduce the amount of
light scattered by the particles for any given frame and Frame rate, f (Hz) 10,000 20,000 20,000 15
therefore typically is not suitable for PIV measurements Duration, T (s) 4.367 2.184 9.331 68.3
in air flows. However, by limiting the width of the light Samples, N 43,674 43,674 186,612 1,024
sheet to a few millimeters (4–8 mm) and less than 0.5 mm TUcl /δ 992 490 2,100 16,550
thickness, the fluence can be increased to provide suffi- Sample interval, uτ2 /(f ν) 0.275 0.134 0.134 183
cient light scattering from micrometer-sized particles to Exposure, texp (μs) 20 25 20 <0.01
adequately expose to camera sensor. Beam profiler meas- Field of view, W (mm) 8.18 5.03 7.18 96.7
urements determined the light sheet width and thickness H (mm) 32.73 20.13 28.74 81.6
(1/e2) to be 7.0 mm and 0.2 mm, respectively. At a laser Magnification, M (pixel/mm) 31.286 50.875 35.632 26.482
power of 5W, the fluence is about 3.3 W/mm2.
Magnification, M (y+/pixel) 0.430 0.265 0.373 0.502
Measurements were performed using two high-speed
Plotting symbol △ ▽  ◦
camera models: a Photron SA5 with a maximum frame rate

13
Exp Fluids (2015) 56:17 Page 5 of 17 17

Fig. 3  Representative images


from Seq. 1a (left) and Seq.
1b (right). Field of view is
32.73 × 8.18 mm2 (left) and
20.13 × 5.03 mm2 (right); glass
tunnel wall is horizontal line
near the bottom edge of the
image, mean flow is left to right

13
17 Page 6 of 17 Exp Fluids (2015) 56:17

3.2 PIV processing

Except for the shear rate estimation presented in Sect. 3.7,


the image data are processed using conventional dual-
frame PIV image analysis based on PIVview-v3.5 soft-
ware (PIVTEC GmbH). The processing scheme outlined
in Fig. 4 makes use of the fact that the data are tempo-
rally over-sampled to allow an image stride greater than
the individual frame separation. This results in a reduced
uncertainty in the displacement estimate due to the pro- Fig. 4  PIV processing scheme for the analysis of temporally over-
portionally increased displacement magnitude, but ignores sampled PIV recordings for an image stride of 2
higher order effects such as particle path curvature or accel-
eration. For the present application, the image strides varies
between n = 1 and n = 2, while the sampling window size required friction velocity uτ is determined using the meth-
is varied between 32 × 12 and 64 × 6 pixels. This allows odology described in Sect. 3.7.
improving the spatial resolution in the vicinity of the wall In Fig. 6 the unsteady behavior of the boundary layer
while maintaining sufficient particle image counts per sam- flow is shown in the form of space-time diagrams of the
ple. The recovered PIV data are then filtered in time with a streamwise and wall-normal velocity components u+ , v+
moving average of Gaussian weight distribution to reduce as well as the spanwise vorticity profile ωz+. In these plots
high-frequency (spurious) noise. In the present application, each column of pixels represents a single column of PIV
the filter base length is ±1 to ±3 time steps. data from a given data set in the sequence, while the hori-
Histograms compiled from the complete data sets of the zontal time axis covers 1,024 time steps (0.0512 s). Thus,
processed sequences Seq. 1b and Seq. 2 are presented in a given row represents a time trace at a fixed wall distance.
Fig. 5. Both the horizontal (x) and the vertical displace- These time traces will be used later for the calculation of
ment component (y) are provided in unscaled (pixel) spectra and cross-correlation functions.
units to highlight the possible presence of pixel locking
in the data. Pixel locking refers to the frequently observed 3.3 Mean velocity profile and higher order statistics
tendency of PIV data toward integral pixel displacement
values and is typically rooted in under-sampled image An overview of several characteristic quantities for the tur-
data. Of the four histograms in Fig. 5, only the vertical bulent boundary layer as calculated from the high-speed
displacement component for Seq. 1b obtained with the PIV image sequences are provided in Table 3. Many of
Photron camera (Fig. 5, top left) exhibits significant pixel these values have been obtained through integration of the
locking. This effect is not observed for the PCO camera mean velocity profile which will be described in the fol-
which can be explained by its roughly halved pixel size of lowing. The mean streamwise velocity profile is shown in
11 µm compared with 20 µm for the Photron camera. The Fig. 7 for all three sequences, all of which were processed
horizontal displacement component x of neither camera with a sample size of 64 × 6 pixels, corresponding to ver-
exhibits any sign of pixel locking which is a consequence tical sizes of 2.6y+ (Seq. 1a), 1.6y+ (Seq. 1b) and 2.2y+
of preferentially horizontal image streaking on the sen- (Seq. 2). Symbols are only provided for the near-wall region
sor due to the finite exposure of continuously illuminated (<10y+) to improve the visibility of differences between the
particles. This results in particle images wider than the respective profiles. The reference PIV data with a sample
minimum 2.0 pixel size required for PIV data being free of height of 8.0y+ is indicated as orange-filled circles up into
pixel locking (Adrian and Westerweel 2010; Raffel et al. the buffer layer and orange line above. In addition the result
2007). The effect of pixel locking on the turbulence statis- of direct numerical simulation (DNS) of an evolving turbu-
tics, although an interesting subject in itself, is not investi- lent boundary layer with zero pressure gradient (ZPGTBL)
gated within the scope of the present study. Surprising (or at Reτ = 255, Reθ = 590 (Schlatter et al. 2009) is plotted in
maybe not) is the fact that none of the figures involving the solid red. The solid blue line indicates the linear dependence
vertical velocity component v presented in the following inside the viscous sublayer where u+ = y+. The boundary
seems to exhibit artifacts that could be attributed to pixel layer thickness δ99 % is reached at ≈ 240y+ which corre-
locking. sponds to 17.8 mm.
The majority of the following figures are scaled to Overall the measured velocity profiles do not exhibit
wall units (inner scaling) which is indicated by the super- a clear logarithmic portion which is to be expected at
script ‘+’ and represents the viscous scaling of length this rather low Reynolds number (Reτ = 240). The data
y+ = yuτ /ν, velocity u+ = u/uτ and time t + = tuτ2 /ν. The obtained from the sequences were compared favorably with

13
Exp Fluids (2015) 56:17 Page 7 of 17 17

Fig. 5  Histograms for complete


data sets of horizontal (left) and
vertical displacement compo-
nent (right) for Seq. 1b (top)

Rel. Frequency
Rel. Frequency
and Seq. 2 (bottom); bin width
0.05 pixel

0 5 10 15 20 -2 -1 0 1 2
dx [pixel] dy [pixel]

Rel. Frequency
Rel. Frequency

0 5 10 15 -2 -1 0 1 2
dx [pixel] dy [pixel]

Fig. 6  Temporal evolution


of the normalized streamwise
component u+ (top), the wall-
normal component v+ (middle)
and spanwise vorticity ωz+ (bot-
tom) for a portion of Seq. 1b

the ZPGTBL DNS data at Reτ = 255. A small vertical off- The reference PIV data with its roughly four times larger
set of 0.5U + toward higher values at 20 < y+ < 150 with a vertical sample size 8y+ are not able to reliably recover
less pronounced excursion in the outer layer (>150y+) may this peak and deviates strongly for smaller wall distances.
be related to the fact that the evolving turbulent boundary Deviations between the DNS and measured profiles for
layer inside the duct is bounded from both sides and the the streamwise and wall-normal fluctuations are more
top and has a small (favorable) streamwise pressure gradi- pronounced in the region above 50y+ which again is most
ent (∂p/∂x < 0). likely attributed to the finite streamwise pressure gradient
The wall-normal profiles of the normalized fluctuat- in the duct.
ing velocity components �ui′ uj′ �+ are provided in Fig. 8 Figure 9 shows profiles of the skewness S and flatness
for all three sequences alongside with the reference PIV or kurtosis K of the streamwise velocity. Only the data for
measurements (orange-filled circles) and ZPGTBL DNS the longest sequence Seq. 2, processed with a larger sam-
data (red line). The peak turbulence production is located ple window of 64 × 12 pixels, are presented. Data for the
near 13.5y+ and is in good agreement with the DNS data. shorter sequences Seq. 1a and Seq. 1b is not shown because

13
17 Page 8 of 17 Exp Fluids (2015) 56:17

Table 3  Characteristic Sequence number Seq. 1a Seq. 1b Seq. 2 Ref.


quantities calculated from high-
speed PIV data Camera Photron Photron Dimax-HS4 sCMOS

2,770 2,700 2,685 –



−1
Wall shear rate, ∂u
∂y y=0 (s )


Friction velocity, uτ (m/s) 0.2045 0.2019 0.2008 –


Wall unit, ν/uτ (μ/s) 74.3 75.3 75.7 –
Max. mean velocity, Umax (m/s) 4.106 4.049 4.026 4.034
Max. mean displacement, xmax (pixel) 12.84 10.30 7.17 32.05
BL thickness, δ (mm) 31.21 31.26 27.60 29.36
δ99 % (mm) 18.08 18.02 17.83 17.12
Displacement thickness, δ∗ (mm) 2.953 2.979 2.917 2.887
Momentum thickness, θ (mm) 1.937 1.933 1.894 1.922
Shape factor, S (–) 1.525 1.542 1.540 1.502
Reynolds number, Reδ99 (–) 4,915 4,833 4,755 4,575
Reθ (–) 527 518 505 513
Reτ (–) 245 241 239 –
Friction coeff., cf (–) 0.00496 0.00506 0.00503 0.00501

20 8
〈 u’u’ 〉+

6
15
+
〈 u’i u’j 〉

4
U+

10

2
〈 v’v’ 〉+
5
0
〈 u’v’ 〉+

0 100 101 102


100 101 102 +
+
y
y

Fig. 7  Mean streamwise velocity profiles U + of turbulent boundary 8


〈 u’u’ 〉+
layer inside a square duct; DNS data for Reτ = 255 from Schlatter
et al. (2009) (red line). Symbols only shown for y+ < 10 for clarity:
6
triangle Seq. 1a, inverter triangle Seq. 1b, square Seq. 2, circle con-
ventional 2-C PIV
〈 u’i u’j 〉+

of lack of convergence. In comparison, the reference PIV 2


data cover a longer sampling period in time and therefore 〈 v’v’ 〉+
are able to follow the same trends, even though not fully 0
converged. Toward the upper edge of the boundary layer the 〈 u’v’ 〉+
skewness tends toward zero and kurtosis toward 3, which 0 100 200 300
is indicative of the expected normal Gaussian distribution +
y
of the weakly turbulent outer bulk flow. A good match to
the ZPGTBL DNS data at Reτ = 252 (Schlatter and Örlü
Fig. 8  Averaged, normalized velocity fluctuations of the turbulent
2010) can be observed even for values down into the vis-
channel boundary layer, top profiles: < u′ u′ >; middle< v′ v′ >; bot-
cous sublayer. The peaks in Su and Ku for the measured tom < u′ v′ >; red lines DNS data for Reτ = 255 from Schlatter and
data in the outer region at ≈ 200y+ are not as pronounced Örlü (2010)

13
Exp Fluids (2015) 56:17 Page 9 of 17 17

10
6
4 8
5
6

4
2 4

Ku-3
Su

Ξ
2 3

0 0
2

-2
1
-2 -4

100 101 102 0


+
100 101 102
y y +

Fig. 9  Skew (solid line) and kurtosis (dashed line) of streamwise +


velocity component u for Seq. 2 processed with a sampling window Fig. 10  Profiles of the log-law indicator functions Ξ = y+ ∂U ∂y+ ;
of 64 × 12 pixels (24.0x + × 4.5y+). Conventional PIV results circle inverter triangle Seq. 1b, circle conventional 2-C PIV, straight line
obtained with sample size of 24 × 16 pixels (12.0x + × 8y+), red lines DNS data for Reτ = 255 from Schlatter and Örlü (2010)
DNS data for Reτ = 255 from Schlatter and Örlü (2010)

as in the DNS, for which there is no immediate explana- 0.5

tion. Other TBL experiments at higher Reynolds numbers


report similar peak values of Su ≈ −2 and Ku ≈ 10 (Vin- 0.4
centi et al. 2013). The different positions of the outer peaks
between the current experiment and the DNS are related, in 0.3
+
u’ /U

part, to the difference in Reynolds numbers.


+

0.2
3.4 Diagnostic plots and log‑law indicator function
0.1
To better assess the quality of the measurements with
regard to their suitability to providing reliable turbulence
statistics, three different methods of plotting the experi- 0
100 101
2
10
mental data are reported in the literature and shall be pre- y
+

sented in the following.


Plots of the so called log-law indicator function
+ Fig. 11  Ratio of rms streamwise velocity to local mean velocity
Ξ = y+ ∂U ∂y+ for sequence Seq. 1b, the reference PIV meas-

( u′2 /u) for sampling window heights of 1.6y+ (Seq. 1b, inverter tri-
urements and the DNS are shown in Fig. 10. The local min- angle) and 4.5y+ (Seq. 2, square), 8.0y+ (Ref., circle), red line DNS
imum at y+ ≈ 50 provides an estimate for the inverse of data for Reτ = 255 from Schlatter and Örlü (2010)
the von Kármán constant 1/κ appearing in the logarithmic
wall function:
are a consequence of the considerably larger number of
1
U+ = ln y+ + B (1) samples provided by the sequences. 
κ Figure 11 shows the ratio of u′2 to the streamwise
For fully developed turbulent boundary layers this constant mean velocity u alongside the ZPGTBL DNS data set for
reaches κ = 0.41 with B ≈ 5. The ZPGTBL DNS data of Reτ = 255. As pointed out by de Graaff and Eaton (2000)
Schlatter et al. (2009), Schlatter and Örlü (2010) predict a this quantity provides an estimate of the quality of the near-
value of κ = 0.38 at 48y+ which agrees well to the experi- wall measurements as both u′2 and u approach zero near
mental data shown in Fig. 10. Using the value of κ = 0.38 the wall and hence are susceptible to measurement errors.
the offset constant B is estimated to be B ≈ 4.5 and The DNS data (as well as other experiments) predict a
B ≈ 4.66 for the experimental data and DNS data, respec- value of u′2 /u ≈ 0.4 near the wall for distances y+ < 4.
tively. Compared with the reference PIV measurements, the For the present measurement data, the smallest interroga-
data obtained from the sequence show less scatter which tion window (IW = 64 × 6 pixels) achieved with Seq. 1b

13
17 Page 10 of 17 Exp Fluids (2015) 56:17

0.1

0.4
0.2 0.05

0 0.3
0 0.1 0.2
u’/Ucl

+
ωz
0.1
0.2

0.1

0
0 0.2 0.4 0.6 0.8 1
0 0
101
2
u/Ucl 10 10
+
y

Fig. 12  Normalized rms streamwise velocity u′2 /Ucl with respect
Fig. 13  Root mean square of spanwise vorticity ωz normalized
to u/Ucl for sampling window heights of 2.58y+ (Seq. 1a, triangle),
by mean wall shear rate γ̇ . Red line DNS data for Reτ = 255 from
1.6y+ (Seq. 1b, inverter triangle), 4.5y+ (Seq. 2, square), 8.0y+ (Ref.,
Schlatter and Örlü (2010), orange symbols circle and line reference
circle), red line DNS data for Reτ = 255 from Schlatter and Örlü
PIV data
(2010)

measurements are trustworthy (e.g., free of wall interfer-


retrieves this value remarkably well down to ≈ 1. As y+ ence effects).
expected for larger interrogation windows, the deviation
commences at correspondingly larger wall distances, for 3.5 Spanwise vorticity distribution
Seq. 2a in Fig. 11 at y+ < 4. In the outer region, the meas-
ured value asymptotes to a finite level (≈ 0.02) which is The availability of spatially resolved velocity data allows
attributed to the turbulence in the bulk flow induced by the the estimation of differential quantities such as vorticity
grid located upstream. from each PIV recording. This data can then be analyzed
The third approach to assess the data quality was sug- statistically analogous to the velocity data. As an exam-
gested by Alfredsson and Örlü (2010). The so called ple, the root-mean-square distribution of the wall-parallel
diagnostic plot shows the root-mean-squared streamwise vorticity component ωz is presented in Fig. 13. Vorticity
velocity as a function of the streamwise mean velocity is computed from the velocity field data using a circula-
both normalized by the outer mean flow U∞ which is tion approach applied on the eight nodes surrounding a
plotted in Fig. 12 for the present data sets with U∞ = Ucl. given data point [see e.g., (Raffel et al. 2007), page 194].
This representation collapses the data into the range These data are then normalized by the mean wall shear
0 < u/Ucl < 1 with data taken at the same Reynolds rate γ̇ whose estimation is presented in Sect. 3.7. The vor-
number falling on top of each other, if measured prop- ticity data provided in Fig. 13 were determined for the
erly (Alfredsson and Örlü 2010). In the present case, three high-speed sequences using common sample sizes of
the sequence data match closely with some deviations 32 × 12 pixels. In the outer region of the turbulent bound-
and outliers near the wall where both the mean veloc- ary layer, the wall-parallel vorticity has a nearly constant
ity u and the fluctuations u′ decrease. For the refer- value of ωz+ ≈ 0.02 − 0.03, related to the free stream tur-
ence PIV data, this deviation is much more severe for bulence in the wake of the grid, and increases to a maxi-
values u/Ucl < 0.4 which correspond to a wall distance mum of ωz+ ≈ 0.4 in the immediate vicinity of the wall
of about 15y+. (Alfredsson and Örlü 2010) also point which agrees with ZPGTBL DNS data provided by Schlat-
out that the slope of the linear portion near the wall is ter and Örlü (2009, 2010). The reference PIV data under-
related to the rms level of wall shear stress which was estimate the rms vorticity which can be attributed to both
determined experimentally to be τ ′ /τw = 0.39 Alfreds- spatial and temporal filtering of the velocity estimates due
son et al. (1988) and 0.403 to 0.429 for DNS (Schlatter to finite window size and sampling interval t. The refer-
et al. 2009; Schlatter and Örlü 2010). The dashed blue ence data have a vertical sampling window height of 8.0y+
line in Fig. 12 has a slope of 0.4 with which the current and a sampling interval of 0.80 t + (300 µs) compared
experimental data line up, indicating that the near-wall with 3.2 − 5.2y+ and 0.27 t + (100 µs) for the high-speed

13
Exp Fluids (2015) 56:17 Page 11 of 17 17

Fig. 14  Power spectra of u 10


-1
10
-1

(straight line) and v (dotted 1815 Hz 1116 Hz


-2 -2 -5/3
line) for Seq. 1b at 100y+ (left) 10 10
-5/3
and 15y+ (right). Vertical lines

E11, E22 [m2 s-2 Hz-1]

E11, E22 [m2 s-2 Hz ]


-1
Nyquist cutoff frequencies 10-3 10-3
determined from mean velocity -4 -4
10 10
and sample size
-5 -5
10 10

-6 -6
10 10

-7 -7
10 10

10-8 10-8
101 102 103 101 102 103
Freq. [Hz] Freq. [Hz]

sequences. At the same time the high-magnification data similar Reynolds number [e.g., Figs. 16, 17 by Wu and
set (Seq. 1b) overestimates the rms vorticity which could Moin (2009)].
be related to pixel locking effects present in the wall-nor-
mal velocity component v (see Fig. 5, top right). 3.7 Estimation of the wall shear rate

3.6 Velocity spectra estimation The normalization of boundary layer data is performed


using inner scaling based on the friction velocity uτ which
The acquired sequences are both sufficiently long and tem- is linearly related to the shear stress τw:
porally well resolved to allow the estimation of frequency
∂u 
spectra of the velocity components. Figure 14 shows the τw = µ  (3)
∂y y=0
power spectrum for u at y+ = 15 and y+ = 100 obtained
for Seq. 1b by averaging multiple Hanning window spectra 
of 2,048 point subsamples at a stride of 512. Up to frequen-

τw ∂u 
cies of about 1 kHz the spectra follow a nearly linear decay
uτ = = ν  (4)
ρ ∂y y=0
in energy. Beyond this frequency, the spectra are affected
by the finite size sampling window IW. The 20-kHz For the estimation of shear velocity uτ several approaches
sequence data presented in Fig. 14 were processed with a are commonly used, such as the Clauser method (Clauser
sampling window of 48 pixel width (IWx = 0.94 mm) and 1954) which estimates the shear stress through a nonlinear
an image separation of 2 frames. With a mean velocity of fit of the data from the logarithmic outer layer to Prandtl’s
u ≈ 2.1 m/s at 15y+ this results in an effective sample time law of the wall. Other approaches rely on the properties of
of: the laminar sublayer in which the velocity increases line-
arly with increasing distance from the wall, that is, the non-
IWx
tIW = ≈ 4.48 · 10−4 s (2) dimensional wall-normal coordinate and the velocity are
u linearly related u+ = y+. Measurements of the shear stress
The inverse 1/tIW provides an estimate of the highest are performed using protruding sensors, such as hot wires
resolvable frequencies: 2,233 Hz at 15y+ and 3,630 Hz at and fences (Fernholz et al. 1996; Ruedi et al. 2003), laser
100y+. These values match the relative difference in the Doppler velocimetry (Naqwi and Reynolds 1987; Czar-
Nyquist cutoff frequencies shown in Fig. 14. ske et al. 2002; Shirai et al. 2006), laser gradient meters
Due to the comparatively low turbulence of the channel (Obi et al. 1996), oil film interferometry (Fernholz et al.
flow (Reτ = 240) the typically observed k −5/3 decay in 1996; Ruedi et al. 2003) or even MEMS devices employ-
the power spectrum is practically nonexistent indicating ing hotwires (Löfdahl et al. 2003) or hot films (Jiang et al.
the absence of an inertial subrange. Rather, the presenta- 1996). Other shear stress measurement devices are based
tion of these spectra intends to demonstrate the feasibil- on micro-pillars (Brücker et al. 2005; Gnanamanickam
ity of spectral estimation using temporally resolved PIV et al. 2013, shear sensitive films (Amili and Soria 2011)
data. Estimation of spatial power spectra from high-reso- and micro-optical devices (Ayaz et al. 2011; Miyagi et al.
lution PIV recordings has been demonstrated for instance 2000).
by Herpin et al. (2008). Overall, the spectra match those The wall shear measurement technique presented here
reported in the literature for turbulent boundary layer at is a velocity gradient-based approach and follows the

13
17 Page 12 of 17 Exp Fluids (2015) 56:17

streamwise velocity for the given wall distance.


+
Figure
u (2.08
+
y) 16
+ +
u (1.55 y )
3 + +
u (1.02 y )
+ +
u (0.50 y )

+
u
1

0
0.05 0.06 0.07 0.08 0.09 0.1
Time [ s ]

Fig. 16  Temporal evolution of the streamwise velocity in the viscous


sublayer for 1,024 time steps (0.0512 s) matching the time frame of
Fig. 15

3 1400
Fig. 15  Visualization of temporal particle movement for Seq. 1b
along single wall-parallel line of pixels at difference wall distances.
1360
Time axis is along horizontal, spanning 1,024 time steps (0.0512 s)
2
1320
approach proposed by Kähler et al.Kähler et al. (2006) by

y [ µm ]
evaluating imaged particle trajectories in the immediate
+

1280
y

vicinity of the wall. Figure 15 shows the particle movement Linear fit:
1 y0 = 1181.33 µm
very close to the tunnel wall at distances of y ≈ 0.5y+ , 1y+ = 60.128 pixel 1240
m = 369.78 µm / (m/s)
and 1.5y+. The visualizations are created by extracting sin-
-1
gle lines of pixels at a fixed wall distance from each of the dU/dy = 2704.31 s
1200
images of the sequence and placing these lines side-by-side 0
such that the horizontal axis corresponds to a time axis. 1160
0 0.1 0.2 0.3 0.4 0.5
Therefore, each horizontal pixel equals a time step of 50 µs.
Within this space-time map, the slope of the streak caused U [ m/s ]
by a given particle directly corresponds to the particle’s
velocity. With increasing wall distance, the slope increases Fig. 17  Mean streamwise velocity u obtained by single-line correla-
tion from Seq. 1b with respect to vertical distance (1 pixel ≡ 19.656 µ
as expected. Temporal changes in the slope indicate veloc-
m ≈ 0.265y+)
ity changes that can be observed to occur simultaneously
for all wall distances shown in Fig. 15. The length of the provides a short segment (1,024 time steps) of the time-
streaks of up to a hundred and more time steps indicates resolved, near-wall velocity at four wall distances. A ver-
a sufficient light sheet thickness in this area to account for tical distance of 2 pixels between the estimates is chosen
the spanwise (out-of-plane) motion of the particles. to guarantee minimal particle image-induced correlation
While the near-wall particle streak visualizations are between adjacent velocity estimates as the imaged particles
of qualitative nature, it is also possible to retrieve actual typically cover less than 2 pixels vertically. The velocity
velocity data from the image sequences using a rather data in Fig. 16 were calculated using an image separation
straightforward approach that has been utilized previously of 4 time steps (200 µs) to improve the DVR and subse-
by Nguyen et al. (2010) to estimate the mean wall shear quently temporally filtered with a moving average of ±2
rate along curved interfaces. The method relies on the one- time steps.
dimensional cross-correlation of single, streamwise rows of Estimates of the mean streamwise velocity for each
pixels that are taken at the same wall-normal distance but wall distance are obtained by averaging the instantaneous
separated by a few time steps (here 2–5 time steps). A one- velocity estimates over the entire length of the sequences.
dimensional Gaussian peak fit at the location of maximum A least squares fitting to the linear portion of this data pro-
correlation then provides sub-pixel accurate displacement vides an estimate of the average wall shear rate γ̇ (Fig. 17)
information, which, in conjunction with magnification that is directly proportional to the wall shear stress τw. The
factor M and time difference t, yields an estimate of the line fit also provides a sub-pixel estimate of the actual wall

13
Exp Fluids (2015) 56:17 Page 13 of 17 17

position where u = 0—assuming a nonslip condition at


the wall. At this point it should be pointed out that in com-
parison with most of the existing shear stress measurement 0.4

techniques, the described approach does not require cali-


bration, since the spatial derivative used in the estimation is
invariant with the magnification factor: 0.3

∂u 1 M (�x2 − �x1 )
γ̇ = ≈ (5)

P
∂y �t M (y2 − y1 )
0.2
where xi is the measured displacement in pixels and yi the
wall-normal distance in pixels. It does, however, require a
sufficient number of data points (pixels) within the viscous
0.1
sublayer.
Least squares fitting may also be applied to the instan-
taneous near-wall, streamwise velocity estimates, such as
those shown in Fig. 16, in order to estimate the instanta- 0
-3 -2 -1 0 1 2 3 4 5
neous wall shear rate γ̇ (t). A sample of 1,024 time steps (τ - τavg)/τ’
(0.051 s) of the temporal evolution of the wall shear rate is
shown in Fig. 18. Prior to plotting and averaging, the shear
Fig. 19  Probability density function of the wall shear rate γ̇ obtained
rate data have been median filtered in time using a kernel of from Seq. 1b
±3 time steps. The resulting data provide an estimate of the
instantaneous wall shear stress τw, albeit averaged across
the cross section of the light sheet (≈ 5 × 0.22mm2). 10
-1

Table 4 summarizes the averaged shear rate for two


acquired sequences obtained by either averaging the
unfiltered and filtered instantaneous wall shear rates or -3
-2
10
by least squares fitting to the averaged near-wall velocity
distribution, as determined with the single-line correlation
approach.
PSD [ Hz-1]

-3
10

4000 1.5 -4
10
du/dy [ s ]
-1

Normalized

3000
1
2000
0.5
1000

0 0 10-5
0.05 0.06 0.07 0.08 0.09 0.1 101 102 103
Time [ s ]
Freq. [Hz]

Fig. 18  Temporal evolution of the wall shear rate estimated through Fig. 20  Spectra of wall shear rate obtained from Seq. 1b
least squares fitting to data in Fig. 16 for 1,024 time steps (0.0512 s).
Line at 2,700 s −1 indicates mean shear rate obtained from 43,500
time steps
The normalized probability density function of the
Table 4  Estimates mean shear strain for acquired sequences, filtering measured wall shear rate γ̇ is provided in Fig. 19. The
by mean of median filtering fluctuation, skew and kurtosis of the wall shear rate have
Seq. Duration Number of (Unfiltered) γ̇ Average shear From
respective values of σγ̇ = 0.41, S = 0.80 and K = 3.41.
num. (s) samples N (s −1) rate (filtered) γ̇ Ū(y) γ̇ These values are in good agreement both with previously
(–) (s −1) (s −1) published experimental results [see e.g., (Alfredsson et al.
1988; Colella and Keith 2003; Silva et al. 2014; Keirsbulck
1a 4.367 43,674 2,797 2,755 2,714
et al. 2012; Mathis et al. 2013)] and with reported DNS
1b 2.184 43,674 2,784 2,708 2,704
results (Hu et al. 2006; Örlü and Schlatter 2011).

13
17 Page 14 of 17 Exp Fluids (2015) 56:17

is mostly affected by the motions in the viscous sublayer


with a rapidly decreasing influence within the buffer layer.
Motions in the upper layers (>50y+) show no correlation
with the wall shear. Close to the wall (<10y+), the shear
rate shows a symmetric response to the streamwise com-
ponent u and skews toward delayed response with increas-
ing wall distances (Fig. 21, top). Since the wall-normal
velocity component v drops to zero near the wall, this area
shows no correlation with the wall shear (Fig. 21, middle).
The peak correlation of Rτ ,v = 0.58 located at position
(−3.8t + , 14.7y+) coincides with the peak streamwise fluc-
tuation umax
′+ (c.f. Fig. 8). For wall bounded turbulence, the

peak turbulence production is also located at approximately


this wall distance.
Compared with the previous velocity correlations, the
correlation map for the spanwise vorticity Rτ ,ω (Fig. 21,
bottom) exhibits a more symmetric pattern in time with
the maximum correlation approaching unity at the wall as
to be expected (vorticity at the wall is identical to the wall
shear rate). A negative correlation peak of Rτ ,v = −0.40
is located at (−4.6t + , 17.8y+) near the peak for Rτ ,v. This
negative correlation represents the response of the wall
shear stress to opposite signed vorticity in the turbulence
producing layer.
Fig. 21  Cross-correlation between the wall shear rate γ̇ and the
velocity components u (top), v (middle) and spanwise vorticity ωz
(bottom)
4 Discussion

The temporally resolved wall shear can be further used The large quantity of temporally well resolved velocity
to determine the corresponding frequency spectrum as pre- samples made available with the described PIV implemen-
sented in Fig. 20. The spectrum was obtained by averaging tation suggests it as a possible alternative to single-point
multiple Hanning window spectra of 2,048 point subsam- measurement techniques such as laser Doppler anemom-
ples at a stride distance of 512. The wall shear spectral den- etry (LDA) and hotwire anemometry (HWA). Compared
sity varies with f −3 as reported previously by Keirsbulck with single-point measurement techniques, the presented
et al. (2012) for both experimental and DNS results for PIV implementation additionally provides simultaneous
channels flows with Reτ in a similar range as the present data across the (narrow) field of view and allows the calcu-
measurements. At about 300 Hz (0.1f +) the power spec- lation of, for instance, spatial correlations and differential
trum begins to depart from the f −3 trend due to measure- quantities such as the spanwise vorticity component. How-
ment noise. ever, the finite size sampling window (here: 8.5x + × 3.2y+
With the additional availability of the correspond- or 17.0x + × 1.6y+) imposes a certain amount of spatial
ing temporally resolved wall-normal velocity profile, the averaging in the data. In comparison, the probe volume
wall shear rate can be used to calculate space-time cor- of an LDA typically has a diameter of 30–100 µm which
relations such as presented in Fig. 6. In these plots each corresponds to y+ ≈ 0.7 − 1.5 in the present application.
horizontal line at position y+ represents the normalized Employing optimized LDA configurations, spatial resolu-
cross-correlation of the shear rate γ̇ (t) with the time trace tions better than 10 µm have been reported in the literature
of a given scalar variable for a fixed wall distance y+ [i.e., (Czarske et al. 2002; Shirai et al. 2006).
u(t, y+ ), v(t, y+ ) or ωz (y+ , t)]. The width of the plots covers Hotwires for wall bounded turbulence measurements
512 time steps in each direction (±69t +). have reported diameters of d = 0.5–5 µm and lengths of
The interpretation of the space-time correlation maps about 0.5–1 mm in spanwise direction (e.g., Hutchins et al.
is as follows: a high correlation to the left of t + = 0 indi- 2009; Stanislas et al. 2008), which would, respectively,
cates a delayed response of the wall shear γ̇ to the respec- correspond to 0.007 − 0.07d + and 7 − 13z+ in the present
tive scalar variable u, v or ωz. Bearing this in mind, sev- application. In comparison, the light sheet for the reported
eral interesting features can be observed: The wall shear PIV measurements has a thickness of about 200 µm

13
Exp Fluids (2015) 56:17 Page 15 of 17 17

translating to 2.7z+. As outlined in Table 2, the PIV fram- for the continuous-wave laser source utilized in the pre-
ing frequencies of 10 and 20 kHz correspond to sampling sented setup.
intervals of 0.274t + and 0.137t +. However, the finite width A possible alternative may also be offered through
of the samples (e.g., IWx = 64 pixels) imposes a tempo- pulsed light-emitting diodes (Willert et al. 2010). At pre-
ral averaging of the sampled flow of t + = �x + /u+ ≈ 3t + sent, the fluence achievable with overdriven high-power
at 15y+ reducing to ≈ 1.6t + at 200y+. In comparison, diodes approaches 2 W /mm2 which is of similar magni-
reported HWA measurements indicate sampling intervals tude as the fluence for the 5W CW laser used in the pre-
of 0.5 − 1 t + (Hutchins et al. 2009). sent experiments. However, the primary challenge of LED-
In terms of sampling duration, typical HWA applica- based illumination is the collimation of the emitted light
tions sample the flow at least an order of magnitude longer into thin light sheets which is hampered by the wide angu-
than for the presented PIV data sequences which only lar emission characteristics of LEDs (e.g., high spreading
cover between 490 and 2,100 boundary layer thicknesses angle).
(see Table 2). This duration may be insufficient to properly The high-speed PIV measurement approach described
sample the turbulent flow as exemplified by the improved herein can further benefit from improved sensor tech-
convergence of Seq. 2 (180,000 samples at 20 kHz) in com- nology, in particular with regard to light sensitivity and
parison with Seq. 1b (43,000 samples at 20 kHz). The sam- reduced pixel size. In this regard the PCO dimax HS4
pling duration can be further increased through a reduction (PCO AG., Germany) used for the current measurements
in image width since the described measurement approach exhibits a roughly four times higher sensitivity in compari-
only needs to cover a single interrogation window (e.g., son with the Photron SA-5. Therefore, the combination of
64 pixel). more powerful laser and more sensitive camera should—
The presented high-speed PIV measurement technique in principle—allow measurements of flows in the 100 m/s
was applied to a duct flow of moderate turbulence with range.
a moderate free stream velocity of 4 m/s. When apply- The technique can be readily extended to retrieve all
ing the presented measurement technique to faster flows, three velocity components through the addition of a sec-
two aspects must be taken into consideration. First of all, ond camera in a classical stereoscopic PIV configuration
the sampling frequency must be increased to keep parti- (Prasad 2000; Willert 1997). Given that the narrow field of
cle image displacements to magnitudes of roughly 10–20 view has a high aspect ratio (W ≪ H ) the use of oblique
pixels in image space. Secondly, the camera exposure has view correcting Scheimpflug mounts is not essential which
to be further reduced to prevent excessive particle image can further simplify the imaging setup.
streaking, typically to less than 25 % of the exposure With regard to the presented wall shear estimation
time. Both factors, increased frame rate and reduced expo- method, it should be kept in mind that the viscous sub-
sure time, require an equivalent increase of laser power to layer has to be sufficiently resolved in order to apply the
achieve a similar exposure of the image detector. With the single-line correlation technique. In the present case, the
utilized imaging and illumination equipment, sensible PIV height of the viscous sublayer was imaged with 10-20
recordings are feasible down to exposure times of 5 µs with pixels on the sensor, corresponding to about 5 wall units.
corresponding frame rates of 50 kHz. This translates to a Depending on the application, this resolution is not always
fivefold increase of applicability of the measurement tech- achievable and requires high-magnification imaging (e.g.,
nique to flow velocities in the 20 m/s range at a comparable long-distance microscope). However, with increased mag-
geometric imaging configuration (i.e., same magnification nification the effective seeding density typically decreases
factor). (Kähler et al. 2006, 2012) which in turn affects the reliable
An extension of the measurement technique for use at recovery of unsteady shear rate estimates using the single-
higher flow velocities beyond 10 m/s requires either a cam- line correlation approach. Furthermore, image aberration
era of increased sensitivity or a more powerful light source generally increases with increased magnification leading to
or, preferably, a combination of both. Continuously oper- blurred particle images and further compounds the problem
ating lasers, comparable to the one utilized in the present (Kähler et al. 2012).
application, are available with output powers in excess
of 20 Watts. Pulsed light sources based on diode pumped
solid state lasers are also capable of providing sufficient 5 Summary
pulse energy (≥ 100 µJ) even at frequencies beyond
100 kHz. Their use has to be balanced against their higher The primary objective of presented study was to demon-
cost, more complex handling and, in particular, reduced strate the capability of nowadays readily available high-
image quality stemming from nonuniform intensity distri- speed imaging in combination with large image counts for
butions and pulse-to-pulse variations that are not present the retrieval of statistically converged flow quantities in

13
17 Page 16 of 17 Exp Fluids (2015) 56:17

low-speed turbulent air flows. Contrary to conventional dou- Amili O, Soria J (2011) A film-based wall shear stress sensor for wall-
ble-pulsed PIV, the setup uses a continuous-wave (CW) laser bounded turbulent flows. Exp Fluids 51(1):137–147. doi:10.1007/
s00348-010-1035-5
and does not require a laser timing generator as the particle Ayaz UK, Ioppolo T, Ötügen MV (2011) Wall shear stress sensor
image exposure is directly controlled by limiting the expo- based on the optical resonances of dielectric microspheres. Meas
sure of the camera sensor (e.g., electronic shutter). This in Sci Technol 22(7):075,203. doi:10.1088/0957-0233/22/7/075203
itself reduces the complexity of the setup. Although applied Brücker C, Spatz J, Schröder W (2005) Feasability study of wall
shear stress imaging using microstructured surfaces with flex-
to an air flow, the technique is equally suitable for measure- ible micropillars. Exp Fluids 39(2):464–474. doi:10.1007/
ment of water flows and can be extended to provide three- s00348-005-1003-7
component data by means of a multi-camera imaging setup. Clauser FH (1954) Turbulent boundary layers in adverse pressure gra-
In the present application, the technique permitted dients. J Aeronaut Sci 21(2):91–108. doi:10.2514/8.2938
Colella K, Keith W (2003) Measurements and scaling of wall shear
the efficient estimation of wall-normal profiles of mean stress fluctuations. Exp Fluids 34(2):253–260. doi:10.1007/
and fluctuating velocity components such as the Reyn- s00348-002-0552-2
olds stresses. Measurements down into the viscous sub- Czarske J, Büttner L, Razik T, Müller H (2002) Boundary layer
layer of the turbulent boundary layer inside a square duct velocity measurements by a laser Doppler profile sensor with
micrometre spatial resolution. Meas Sci Technol 13(12):1979.
were possible using rather high magnifications of 1:2 and doi:10.1088/0957-0233/13/12/324
1:1 and sufficient seeding density in this region. Using a de Graaff DB, Eaton JK (2000) Reynolds-number scaling of the
wall-parallel, one-dimensional cross-correlation method, flat-plate turbulent boundary layer. J Fluid Mech 422:319–346.
it was shown that reliable velocity estimates are feasible doi:10.1017/S0022112000001713
de Silva CM, Gnanamanickam EP, Atkinson C, Buchmann NA,
in the single wall unit range. These data can then be used Hutchins N, Soria J, Marusic I (2014) High spatial range
to directly estimate the instantaneous wall shear rate γ̇ velocity measurements in a high reynolds number turbulent
and with it the shear stress τw and corresponding friction boundary layer. Phys Fluids (1994–present) 26(2):025117.
velocity uτ. Finally, the large sequence lengths in excess doi:10.1063/1.4866458
Dennis DJC, Nickels TB (2011) Experimental measurement of large-
of 100,000 samples permit the estimation of spectra and scale three-dimensional structures in a turbulent boundary layer.
space-time correlation maps from the recovered data. Part 1. Vortex packets. J Fluid Mech 673:180–217. doi:10.1017/
​The described measurement approach provides a breadth S0022112010006324
of flow-specific information that qualifies its use in a com- Elsinga G, Scarano F, Wieneke B, Oudheusden B (2006) Tomo-
graphic particle image velocimetry. Exp Fluids 41(6):933–947.
plementary fashion to established methods such as HWA, doi:10.1007/s00348-006-0212-z
LDA and “classical” PIV. Elsinga GE, Marusic I (2010) Evolution and lifetimes of flow topol-
ogy in a turbulent boundary layer. Phys Fluids (1994–present)
Acknowledgments The author would like to thank his colleagues 22(1):015102. doi:10.1063/1.3291070
J. Klinner, M. Schroll and M. Beversdorff for their assistance in the Fernholz HH, Janke G, Schober M, Wagner PM, Warnack D (1996)
wind tunnel setup and PIV measurements. Further acknowledgment New developments and applications of skin-friction measuring
goes to N. Buchmann for his valuable comments while preparing the techniques. Meas Sci Technol 7(10):1396
manuscript. Finally, the support of PCO GmbH is gratefully acknowl- Gao Q, Ortiz-Dueas C, Longmire EK (2013) Evolution of coher-
edged who provided the Dimax-HS4 high-speed camera for evalua- ent structures in turbulent boundary layers based on mov-
tion purposes. The recommendations by the anonymous reviewers ing tomographic PIV. Exp Fluids 54(12):1625. doi:10.1007/
significantly improved the overall quality of the manuscript. s00348-013-1625-0
Gnanamanickam EP, Nottebrock B, Große S, Sullivan J, Schröder W
(2013) Measurement of turbulent wall shear-stress using micro-pillars.
Meas Sci Technol 24:124,002. doi:10.1088/0957-0233/24/12/124002
References Hain R, Kähler C (2007) Fundamentals of multiframe particle image
velocimetry (PIV). Exp Fluids 42(4):575–587. doi:10.1007/
Adrian R (2005) Twenty years of particle image velocimetry. Exp s00348-007-0266-6
Fluids 39(2):159–169. doi:10.1007/s00348-005-0991-7 Herpin S, Wong C, Stanislas M, Soria J (2008) Stereoscopic PIV
Adrian R, Westerweel J (2010) Particle image velocimetry. Cam- measurements of a turbulent boundary layer with a large spa-
bridge Aerospace Series. Cambridge University Press, New York. tial dynamic range. Exp Fluids 45(4):745–763. doi:10.1007/
ISBN 978-0521440080 s00348-008-0533-1
Adrian RJ (1991) Particle-imaging techniques for experimental fluid Hu Z, Morfey CL, Sandham ND (2006) Wall pressure and shear
mechanics. Annu Rev Fluid Mech 23(1):261–304. doi:10.1146/ stress spectra from direct simulations of channel flow. AIAA J
annurev.fl.23.010191.001401 44(7):1541–1549
Adrian RJ (1997) Dynamic ranges of velocity and spatial resolution Hutchins N, Nickels TB, Marusic I, Chong MS (2009) Hot-wire spa-
of particle image velocimetry. Meas Sci Technol 8(12):1393 tial resolution issues in wall-bounded turbulence. J Fluid Mech
Alfredsson PH, Johansson AV, Haritonidis JH, Eckelmann H (1988) 635:103–136. doi:10.1017/S0022112009007721
The fluctuating wall-shear stress and the velocity field in the Jiang F, Tai YC, Gupta B, Goodman R, Tung S, Huang JB, Ho CM
viscous sublayer. Phys Fluids (1958–1988) 31(5):1026–1033. (1996) A surface-micromachined shear stress imager. In: Pro-
doi:10.1063/1.866783 ceedings of micro electro mechanical systems, 1996, MEMS ’96.
Alfredsson PH, Örlü R (2010) The diagnostic plot a litmus test for An investigation of micro structures, sensors, actuators, machines
wall bounded turbulence data. Eur J Mech B Fluids 29(6):403– and systems. IEEE, The ninth annual international workshop, pp
406. doi:10.1016/j.euromechflu.2010.07.006 110–115 . doi:10.1109/MEMSYS.1996.493838

13
Exp Fluids (2015) 56:17 Page 17 of 17 17

Kähler C, Scholz U, Ortmanns J (2006) Wall-shear-stress and near- Scarano F (2013) Tomographic PIV: principles and practice. Meas Sci
wall turbulence measurements up to single pixel resolution by Technol 24(1):012,001
means of long-distance micro-PIV. Exp Fluids 41(2):327–341. Schlatter P, Örlü R (2010) Assessment of direct numerical simulation
doi:10.1007/s00348-006-0167-0 data of turbulent boundary layers. J Fluid Mech 659:116–126.
Kähler CJ, Scharnowski S, Cierpka C (2012) On the resolution limit doi:10.1017/S0022112010003113
of digital particle image velocimetry. Exp Fluids 52(6):1629– Schlatter P, Örlü R, Li Q, Brethouwer G, Fransson JHM, Johansson
1639. doi:10.1007/s00348-012-1280-x AV, Alfredsson PH, Henningson DS (2009) Turbulent boundary
Keirsbulck L, Labraga L, el Hak MG (2012) Statistical properties of layers up to Reθ = 2500 studied through simulation and experi-
wall shear stress fluctuations in turbulent channel flows. Int J Heat ment. Phys Fluids 21(5):051702. doi:10.1063/1.3139294
Fluid Flow 37:1–8. doi:10.1016/j.ijheatfluidflow.2012.04.004 Schröder A, Geisler R, Elsinga GE, Scarano F, Dierksheide U (2008)
LeHew J, Guala M, McKeon B (2013) Time-resolved measurements Investigation of a turbulent spot and a tripped turbulent bound-
of coherent structures in the turbulent boundary layer. Exp Fluids ary layer flow using time-resolved tomographic PIV. Exp Fluids
54(4):1508. doi:10.1007/s00348-013-1508-4 44(2):305–316. doi:10.1007/s00348-007-0403-2
Löfdahl L, Chernoray V, Haasl S, Stemme G, Sen M (2003) Charac- Schröder A, Geisler R, Staack K, Elsinga G, Scarano F, Wieneke B,
teristics of a hot-wire microsensor for time-dependent wall shear Henning A, Poelma C, Westerweel J (2011) Eulerian and lagran-
stress measurements. Exp Fluids 35(3):240–251. doi:10.1007/ gian views of a turbulent boundary layer flow using time-resolved
s00348-003-0624-y tomographic PIV. Exp Fluids 50(4):1071–1091. doi:10.1007/
Mathis R, Marusic I, Chernyshenko SI, Hutchins N (2013) Estimating s00348-010-1014-x
wall-shear-stress fluctuations given an outer region input. J Fluid Sciacchitano A, Scarano F, Wieneke B (2012) Multi-frame pyramid
Mech 715:163–180. doi:10.1017/jfm.2012.508 correlation for time-resolved PIV. Exp Fluids 53(4):1087–1105.
Miyagi N, Kimura M, Shoji H, Saima A, Ho CM, Tung S, Tai YC doi:10.1007/s00348-012-1345-x
(2000) Statistical analysis on wall shear stress of turbulent bound- Shirai K, Pfister T, Büttner L, Czarske J, Müller H, Becker S, Lien-
ary layer in a channel flow using micro-shear stress imager. Int J hart H, Durst F (2006) Highly spatially resolved velocity meas-
Heat Fluid Flow 21:576–581 urements of a turbulent channel flow by a fiber-optic heterodyne
Naqwi A, Reynolds W (1987) Dual cylindrical wave laser Doppler laser-doppler velocity-profile sensor. Exp Fluids 40(3):473–481.
method for measurement of skin friction in fluid flow. Technical doi:10.1007/s00348-005-0088-3
report, Report No. TF-28, Stanford University Stanislas M, Perret L, Foucaut JM (2008) Vortical structures in the tur-
Nguyen TD, Wells JC, Nguyen CV (2010) Wall shear stress meas- bulent boundary layer: a possible route to a universal representa-
urement of near-wall flow over inclined and curved boundaries tion. J Fluid Mech 602:327–382. doi:10.1017/S0022112008000803
by stereo interfacial particle image velocimetry. Int J Heat Fluid Vincenti P, Klewicki J, Morrill-Winter C, White C, Wosnik M
Flow 31(3):442–449. doi:10.1016/j.ijheatfluidflow.2009.12.002 (2013) Streamwise velocity statistics in turbulent boundary lay-
(Sixth International Symposium on Turbulence and Shear Flow ers that spatially develop to high reynolds number. Exp Fluids
Phenomena) 54(12):1629. doi:10.1007/s00348-013-1629-9
Obi S, Inoue K, Furukawa T, Masuda S (1996) Experimental study on Westerweel J, Scarano F (2005) Universal outlier detection
the statistica of wall shear stress in turbulent channel flows. Int J for PIV data. Exp Fluids 39(6):1096–1100. doi:10.1007/
Heat Fluid Flow 17:187–192 s00348-005-0016-6
Örlü R, Schlatter P (2011) On the fluctuating wall-shear stress in zero Willert C (1997) Stereoscopic digital particle image velocimetry for
pressure-gradient turbulent boundary layer flows. Phys Fluids application in wind-tunnel flows. Meas Sci Technol 8:1465–1479
(1994–present) 23(2):021704. doi:10.1063/1.3555191 Willert C, Stasicki B, Klinner J, Moessner S (2010) Pulsed operation of
Prasad AK (2000) Stereoscopic particle image velocimetry. Exp Flu- high-power light emitting diodes for imaging flow velocimetry. Meas
ids 29:103–116 Sci Technol 21(7):075402. doi:10.1088/0957-0233/21/7/075402
Raffel M, Willert C, Wereley S, Kompenhans J (2007) Particle image Wu X, Moin P (2009) Direct numerical simulation of turbulence in
velocimetry: a practical guide. Experimental Fluid Mechanics. a nominally zero-pressure-gradient flat-plate boundary layer. J
Springer, Berlin Fluid Mech 630:5–41. doi:10.1017/S0022112009006624
Ruedi J, Nagib H, Österlund J, Monkewitz P (2003) Evaluation
of three techniques for wall-shear measurements in three-
dimensional flows. Exp Fluids 35(5):389–396. doi:10.1007/
s00348-003-0650-9

13

You might also like