ArticlePDF Available

Population, characteristics and kinematics of vortices in a confined rectangular jet with a co-flow

Authors:
  • Guangdong Technion--Israel Institute of Technology

Abstract and Figures

Vortex behavior and characteristics in a confined rectangular jet with a co-flow were examined using vortex swirling strength as a defining characteristic. On the left side of the jet, the positively (counterclockwise) rotating vortices are dominant, while negatively rotating vortices are dominant on the right side of the jet. The characteristics of vortices, such as population density, average size and strength, and deviation velocity, were calculated and analyzed in both the cross-stream direction and the streamwise direction. In the near-field of the jet, the population density, average size and strength of the dominant direction vortices show high values on both sides of the center stream with a small number of counter-rotating vortices produced in the small wake regions close to jet outlet. As the flow develops, the wake regions disappear, these count-rotating vortices also disappear, and the population of the dominant direction vortices increase and spread in the jet. The mean size and strength of the vortices decrease monotonically with streamwise coordinate. The signs of vortex deviation velocity indicate the vortices transfer low momentum to high-velocity region and high momentum to the low velocity region. The developing trends of these characteristics were also identified by tracing vortices using time-resolved particle image velocimetry data. Both the mean tracked vortex strength and size decrease with increasing downstream distance overall. At the locations of the left peak of turbulent kinetic energy, the two-point spatial cross-correlation of swirling strength with velocity fluctuation and concentration fluctuation were calculated. All the correlation fields contain one positively correlated region and one negatively correlated region although the orientations of the correlation fields varied, due to the flow transitioning from wake, to jet, to channel flow. Finally, linear stochastic estimation was used to calculate conditional structures. The large-scale structures in the velocity field revealed by linear stochastic estimation are spindle-shaped with a titling stream-wise major axis.
Content may be subject to copyright.
RESEARCH ARTICLE
Population, characteristics and kinematics of vortices
in a confined rectangular jet with a co-flow
B. Kong M. G. Olsen R. O. Fox
J. C. Hill
Received: 30 March 2010 / Revised: 3 September 2010 / Accepted: 11 October 2010 / Published online: 20 November 2010
ÓSpringer-Verlag 2010
Abstract Vortex behavior and characteristics in a con-
fined rectangular jet with a co-flow were examined using
vortex swirling strength as a defining characteristic. On the
left side of the jet, the positively (counterclockwise)
rotating vortices are dominant, while negatively rotating
vortices are dominant on the right side of the jet. The
characteristics of vortices, such as population density,
average size and strength, and deviation velocity, were
calculated and analyzed in both the cross-stream direction
and the streamwise direction. In the near-field of the jet, the
population density, average size and strength of the dom-
inant direction vortices show high values on both sides of
the center stream with a small number of counter-rotating
vortices produced in the small wake regions close to jet
outlet. As the flow develops, the wake regions disappear,
these count-rotating vortices also disappear, and the pop-
ulation of the dominant direction vortices increase and
spread in the jet. The mean size and strength of the vortices
decrease monotonically with streamwise coordinate. The
signs of vortex deviation velocity indicate the vortices
transfer low momentum to high-velocity region and high
momentum to the low velocity region. The developing
trends of these characteristics were also identified by
tracing vortices using time-resolved particle image veloc-
imetry data. Both the mean tracked vortex strength and size
decrease with increasing downstream distance overall. At
the locations of the left peak of turbulent kinetic energy,
the two-point spatial cross-correlation of swirling strength
with velocity fluctuation and concentration fluctuation
were calculated. All the correlation fields contain one
positively correlated region and one negatively correlated
region although the orientations of the correlation fields
varied, due to the flow transitioning from wake, to jet, to
channel flow. Finally, linear stochastic estimation was used
to calculate conditional structures. The large-scale struc-
tures in the velocity field revealed by linear stochastic
estimation are spindle-shaped with a titling stream-wise
major axis.
1 Introduction
The study of coherent structures in turbulent shear flows is
of great importance due to the structures’ significant con-
tribution to fluid entrainment and the transfer of mass,
momentum and heat. The properties of these structures,
such as population, size, circulation, and energy, can be
useful in the understanding of turbulence and property
transport and can aid in the development of more efficient
and more environmentally benign applications. Moreover,
detailed measurements of the behavior and characteristics
of large-scale structures can be used to validate the pre-
dictions of computer modeling techniques, such as large
eddy simulation, where the largest scales of the turbulence
are resolved. The objective of the work presented here is to
use a vortex identification method to analyze the experi-
mental data collected in a confined rectangular liquid jet
with a co-flow with regard to the population, size, kine-
matics and characteristics of vortices within the jet.
In the chemical process industry, liquid jets usually are
placed in a confined space, as in a mixer or a reactor, and
are commonly applied within a co-flow or cross-flow to
improve mixing and chemical reaction between two or
more fluids, like a concentric pipeline mixer or a coaxial jet
mixer (Lu et al. 1997). A typical geometry for this type of
B. Kong M. G. Olsen (&)R. O. Fox J. C. Hill
Department of Mechanical Engineering,
Iowa State University, Ames, IA 50010, USA
e-mail: mgolsen@iastate.edu
123
Exp Fluids (2011) 50:1473–1493
DOI 10.1007/s00348-010-0995-9
mixer is an axisymmetric jet with a co-flow or cross-flow.
However, cylindrical reactors pose challenges in the lab-
oratory with laser techniques such as particle image ve-
locimetry (PIV) due to the curvature and associated image
distortion in the measurements. An alternative approach is
to use a rectangular configuration (Feng et al. 2005), which
can provide much better measurement accuracy and also a
much simpler geometry to conduct CFD simulations. When
the aspect ratio of a rectangular jet nozzle (w/h, where
wand hare the long and short sides of the nozzle) is not
large enough to eliminate the 3D effect, the jet usually can
no longer be considered a planar jet, but instead a rectan-
gular jet (Deo et al. 2007a). For example, Pope (2000)
pointed out that the aspect ratio of a plane nozzle must be
significantly large, typically 50 or more, to ensure that the
flow is statistically two-dimensional and free from the
effects of sidewalls. For a liquid phase jet, especially when
designed to study the mixing process in a high Schmidt
number environment, it is not practical to build and study a
high aspect ratio planar jet mixer. The aspect ratio of the jet
in the current research is five, which indeed is a rectangular
jet. But according to Deo et al. (2007b), the sidewalls of
the confinement can significantly limit the development of
the jet in the spanwise direction, and make the 2-D region
extend a longer distance than a rectangular jet without
sidewalls. In this study, the farthest observation down-
stream location was X/d=30, and thus the measured flow
fields are more comparable to planar jets, instead of most
studies of low- aspect-ratio rectangular jets in which the
focus is mainly on a free jet, in which the 3-D vortex ring
structure can break-down faster than in axisymmetric jets
due to the presence of sharp corners (Grinstein 2001;
Gutmark and Grinstein 1999). Although the confined
rectangular jet with a co-flow is not common in turbulent
jet research, it provides a simple and well-defined geometry
suitable for both experimental measurements and numeri-
cal simulations, especially for PIV experiment and 3-D
CFD simulations.
The jet flapping phenomenon first reported by Golds-
chmidt and Bradshaw (1973) is one of the first indications
of large-scale structure in the planar turbulent jet. How-
ever, Oler and Goldschmidt (1982) suggested that corre-
lation results indicating the flapping motion can also be
explained by the presence of an antisymmetric array of
counter-rotating spanwise vortices. Antonia et al. (1983)
also performed correlation measurements, which support
this concept and showed that the apparent flapping could
indeed be explained in terms of the passage of vortical
structures past the fixed probe pair and was not associated
with bulk lateral displacement of the jet. The existence of a
large-scale structural array has also been demonstrated
later by the velocity fluctuation two-point correlation fields
obtained in different jet facilities by Mumford (1982),
Antonia et al. (1986) and Thomas and Brehob (1986). Such
large-scale structural arrays propagate at approximately
60% of the local centerline mean velocity, estimated by
Goldschmidt et al. (1981), Antonia et al. (1983) and Tho-
mas and Brehob (1986). Recently, Gordeyev and Thomas
(2000) used proper orthogonal decomposition (POD) to
analyze experimental data, also suggesting that the pres-
ence of planar structures aligned in the spanwise direction
as well as three-dimensional structures with asymmetrical
shape in the cross-stream direction and pseudo-periodically
distributed in the spanwise direction.
Most of the earlier coherent structure investigations on
planar jets were focused on the large structures, because
they were conducted using point-wise velocity measure-
ments and then analyzed with a correlation technique.
In the past two decades, particle image velocitimetry
(PIV) has become a popular experimental technique for
turbulence studies. Since PIV provides instantaneous
two-dimensional velocity field data, it is well suited for
visualizing and identifying vortical structures. Methods of
analyzing and interpreting these spatially resolved velocity
field data were discussed in Adrian (1999) and Adrian et al.
(2000a). Prime examples of using PIV to study coherent
structure are provided by Christensen and Adrian (2001,
2002a,b) in turbulent boundary layers and Agrawal and
Prasad (2002a,b,2003)in turbulent axisymmetric jets. Of
particular interest to the present study, Agrawal and Prasad
(2003) and Chhabra et al. (2006) reported results on vor-
tices present in the axial plane of a self-similar turbulent
axisymmetric jet by using a high-pass filter and vortex
extraction method based on velocity quadrant and angular
variance. These studies include measurements of the vortex
characteristics, such as the vortex population, energy,
vorticity, and rms of velocity fluctuations. Their studies
also show that the centers of a larger vortices spin faster
than centers of a smaller vortices. After normalization,
vorticity results for different vortex radii collapse upon
each other. In contrast, the similar studies of small-scale
vortices in planar jets and rectangular jets are lacking.
Besides the geometry of the jet, another new element of
current study is the presence of co-flow. Although the vast
majority of turbulent jet research has been focused on jets
discharged into ambient fluids, a jet with a co-flow is more
commonly seen in industrial applications. One of the ear-
liest investigations of jets with a co-flow is the analytical
study done by Abramovich et al. (1984). He showed the
presence of three essential regions in co-flow jets: the
initial, principal and transition areas. When the flow is
confined, the process of the co-flow driven by the jet is
modified and the mixing process depends strongly not only
on the velocity ratio but also on the interaction between the
boundary layer, the mixing layer and the main flow
(Gazzah 2010). Gutmark and Wygnanski (1976) found that
1474 Exp Fluids (2011) 50:1473–1493
123
a jet exhausting into a slow-moving co-flowing stream is
narrower than a comparable jet exhausting into quiescent
surroundings. Curtet (1958) was interested by recirculation
phenomena generated by a considerable pressure gradient,
and proposed a parameter of similarity called the parameter
of CrayaCurtet, which was formulated by Steward and
Guruz (1977). For a CrayaCurtet parameter value greater
than 0.8, recirculation can be avoided. Because of its great
practical importance, an increasing attention has been
given to the jet development and mixing in jets with a co-
flow (Bradbury and Riley 1967; Nickels and Perry 1996;
Chu et al. 1999; Benayad et al. 2001; Enjalbert et al.
2009). Another study of interest is the experimental
investigation of an axisymmetric jet discharging in a co-
flowing air stream by Antonia and Bilger (2006), which
shows that the far jet flow may be strongly dependent on
the nozzle injection conditions. Despite all of this previous
work, the behavior of turbulent vortices in such a flow
remains relatively unstudied. Thus, in the work that fol-
lows, the characteristics of vortices in a confined rectan-
gular turbulent jet with a co-flow have been investigated.
2 Experimental facility and methodology
The flow facility used in the experiments presented here is
shown in Fig. 1. The measurements are carried out in a
Plexiglas test section with a rectangular cross-section
measuring 60 mm by 100 mm and with an overall length of
1 m. There are three streams separated by two splitter plates,
each emitting from its own flow conditioning section con-
sisting of a packed bed and turbulence reducing screens and
a 16:1 contraction section. The slope of the surface of the
splitter plates is 3°along the side channels and 1°along the
center channel, and the thickness of the tips of the splitter
plates are less than 0.5 mm. Two different flow cases as
defined by different bulk flow Reynolds numbers of 20,000
and 50,000 (based on average velocity and hydraulic
diameter) were studied. The volumetric flow rates for the
two flow cases are listed in the Table 1.
Experimental data from simultaneous PIV/PLIF exper-
iments and high-speed PIV experiments were collected and
analyzed. The simultaneous PIV/PLIF technique has been
used to study many turbulent mixing problems, such as
Westerweel et al. (2002,2005) in a turbulent jet.
Approximately 24 g of hollow glass spheres (Spherical,
Potters Industries, Inc.) were added to the water reservoir
with total volume of 3,500 l in both the regular and high-
speed PIV measurements. The nominal diameter of the
seed particles was 11.7 lm and the density of the particles
was 1.1 g/cm
3
. The 0.5-mm-thick laser sheet used in high-
speed PIV experiments was produced by a Quantronix
Darwin-Duo dual oscillator, single head Nd:YLF CW laser,
which passes through the centerline of the test section in
the z-direction. A Photron ultima APX-RS high-speed 10-
bit CMOS camera was used to capture particle images. The
image magnification of the CMOS camera was 0.27, and
the numerical aperture was 8. The image capturing fre-
quency was 125 Hz for the Re =20K case and 250 Hz for
the Re =50K case. A multi-pass cross-correlation tech-
nique with decreasing window sizes was used to compute
the velocity field. The final interrogation spot size mea-
sured 16 pixels by 16 pixels, corresponding to 1.02 mm on
each side. With 50% overlap between interrogation win-
dows, the velocity vector spacing was 0.51 mm in both the
x- and y-directions. The time interval between two laser
pulses was 1 ms in the low flow rate case, and 600 ls in the
high flow rate case. Since the memory capacity of CMOS
camera is limited, 1,024 PIV image pairs can be collected
in each run. At each observed location, the experiment was
repeated 20 times, thus a total of 20,480 PIV realizations
were collected and analyzed at each locations in both cases.
The optical set up for the PIV/PLIF experiments is
shown in Fig. 2. In the simultaneous PIV/PLIF measure-
ments, illumination was provided by a New Wave
Research Gemini PIV laser. PIV and PLIF images were
obtained using two 12-bit LaVision Flowmaster 3S CCD
cameras. The image magnification of the two CCD cameras
Fig. 1 Photograph and schematic of the confined planar jet test
section
Table 1 Flow rates of two cases
Reynolds number Flow rate (l/s)
Center stream Outer stream
20K 0.8 0.4
50K 2.0 1.0
Exp Fluids (2011) 50:1473–1493 1475
123
was 0.12, and the numerical aperture was 8 for PIV and 5.6
for PLIF. A dichroic mirror (Q545LP, Chroma Technology
Corp.) was placed at an angle of 45°to the laser sheet to
separate the light paths and direct them to either the PIV or
the PLIF camera. The PLIF camera lens was filtered with a
long-pass optical filter (E560LP, Chroma Technology
Corp.), and the PIV camera lens was filtered with a narrow-
band-pass optical filter (Z532/10X, Chroma Technology
Corp.) The seeding method, time interval between two
laser pulses and cross-correlation technique in PIV mea-
surement in the simultaneous PIV/PLIF experiments were
the same as in high-speed PIV measurement. The fluores-
cent dye Rhodamine 6G was used as a passive scalar in
PLIF. In the center stream, the source concentration of
Rhodamine 6G was 45 lg/l, while the other two streams
were pure water. The in-plane spatial resolution of the
PLIF measurements in the present study was actually
limited by the flow area imaged per pixel, which was
approximately 56 lm. The simultaneous PIV/PLIF data at
high Reynolds number are a subset of measurements
reported by Feng et al. (2007b). Since the experimental
apparatus and procedure have been described in detail
elsewhere (Feng et al. 2005,2007b), the reader is directed
to the literature for further information. At each observed
location, 3,250 simultaneous velocity and concentration
realizations were analyzed in high Re cases and 2,500
realizations in low Re case. The simultaneous PIV/PLIF
experimental data for the low Re case are available at two
locations, X/d=4.5 and X/d=7.5, where Xis down-
stream distance and dis the initial jet width, 20 mm.
The smallest Kolmogorov scale in the flow field can be
estimated based on the exit width of the jet and the tur-
bulent kinetic energy at the tips of the two splitter plates.
Although the smallest turbulent scales of the flow cannot
be fully resolved, second-order quantities such as velocity
fluctuations and characteristics of large-scale structures can
be measured accurately. As in Prasad et al. (1992), the
random error in the PIV measurements was estimated as
one-tenth of the effective particle image diameter. The
measurement resolution and uncertainty of the two cases
are listed in the Table 2.
To assistant in the understanding of the flow configu-
ration of current investigation, Fig. 3gives the profiles of
the mean of streamwise velocity, U, for both Reynolds
number cases at 3 different downstream locations, X/d=1,
4.5 and 15. After normalization by DU, which is the
average velocity difference between the center stream and
two side streams (DU¼0:2 m/s for Re = 20K and DU¼
0:5 m/s for Re =50K), the profiles of these two flow cases
are very similar. Near the channel inlet, at X/d=1, two
small wake regions on both sides of the tips of the splitter
plates can be observed in the velocity profile. These result
from the boundary layers formed on the splitter plates. At
the further downstream location, X/d=4.5, the two wake
regions disappear and are replaced by two mixing layers
regions that quickly grow together, resulting in the poten-
tial core in the center jet disappearing for both cases. At
X/d=15, the potential cores in the outer streams also
disappear, and the flow continues its development toward
channel flow because of the confinement by the two walls
in cross-stream direction. The wake regions in the near-
field of a jet and confinement configuration of the jet are
rather uncommon in turbulent jet research Again, as
mentioned in the introduction, the design of the jet con-
figuration is determined by the practical considerations of
simulating flow conditions in an industrial reactor.
3 Swirling strength vortex-ID method
Several methods have been suggested and implemented by
researchers to identify vortices in instantaneous two-
dimensional velocity fields, such as those obtained by PIV.
Adrian et al. (2000a,b) described a number of vortex
Table 2 PIV spatial resolution and measurement uncertainty
Reynolds
number
Spatial resolution (g:
local Kolmogorov
scale)
Uncertainty
Center
stream
(%)
Outer
stream
(%)
High speed 20K 5.4g±1.2 ±2.5
PIV 50K 6.7g±0.8 ±1.6
Simultaneous 20K 4.6g±1.7 ±3.4
PIV/PLIF 50K 6.0g±1.3 ±2.7
Fig. 2 Schematic of the optical setup for the combined PIV and PLIF
experiments
1476 Exp Fluids (2011) 50:1473–1493
123
eduction techniques and compared their effectiveness in
identifying vortices. They found that swirling strength
provides a reliable means of extracting the small-scale
vortices, including those which are not visible in velocity
decompositions. Vollmers (2001) also validated and com-
pared several coherent structure eduction methods. He also
concluded that the swirling strength is the best indicator for
the presence of vortices in turbulent flow. Thus, the authors
have chosen a vortex extraction technique based on
swirling strength to analyze the turbulent vortices in a
confined rectangular turbulent jet in this paper.
The concept of swirling strength, k
ci
, was described by
Adrian (1999) as the imaginary part of the complex con-
jugate eigenvalues (k
cr
±k
ci
) of the local velocity gradient
tensor. Physically, k
ci
-1
represents the period required for a
fluid particle to swirl once about the k
cr
-axis (Piomelli et al.
1996; Adrian 1999). Thus, a non-zero k
ci
indicates a local
swirling motion, and spatially connected regions of non-
zero k
ci
represent vortices (Wu and Christensen 2006).
Tomkins and Adrian (2003) multiplied k
ci
by the sign of the
local vorticity to capture the direction of rotation at each
location in the flow field. This modified form of swirling
strength can be written as (Wu and Christensen 2006)
Kciðx;yÞ¼kci ðx;yÞxzðx;yÞ
jxzðx;yÞj ð1Þ
This method has two great advantages. First, it is frame
independent, meaning a priori information of the bulk
motion of the vortex core is not necessary. Second, it only
reveals the regions with vortices, which means the flow
regions with high vorticity but no local swirling motion,
such as shear layers, do not produce non-zero k
ci
. However,
like other vortex extraction methods based on the velocity
gradient, the method does have one important drawback
when applied to PIV data. Due to the sensitivity of velocity
gradient to noise, this type of method is often not applicable
to experimental data if no smoothing has been applied
(Vollmers 2001), since the method may falsely identify
measurement noise as vortices. However, too much velocity
smoothing may eliminate the small vortices. Thus, in
addition to carefully choosing experiment parameters to
reduce the measurement uncertainty and applying slight
smoothing, we utilized a threshold of jKci j1:5Krms
ci (Wu
and Christensen 2006) to limit the influence of experiment
noise, which only affects the edges of clusters and does not
alter the populations appreciably. Further information about
this method can be found in the references mentioned above.
4 Vortex characteristics analysis
4.1 Instantaneous structures
Two examples of the instantaneous swirling strength field
at the near-field 0 \X/d\2.5 and far-field 6.5 \X/d\9
of the jet at Re =20K are shown in Fig. 4to illustrate the
effectiveness of the swirling strength as a vortex indicator.
It is contrasted with the mean-bulk Galilean decomposed
velocity field in Fig. 4a1, a2, which reveals only the vor-
tices moving at their respective spatial convective veloci-
ties. The local Galilean decomposed velocity fields are
shown in Fig. 4b1, b2, where the local convective veloci-
ties have been subtracted from each vortex center. The
background contours are the clusters of Kci after filtering
with a universal threshold of jKci j1:5Krms
ci . Where the
clusters of filtered Kci identified vortices, all of the local
Galilean decomposed velocity vectors on Fig. 4b1, b2
display a clear swirling motion. This demonstrates the
effectiveness of using the swirling strength as a vortex
identification method in the present flow field. Due to
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.5
1
1.5
2
2.5
Mean of Stream wise velocity
Re = 20K
X/d
<U> / ΔU (ΔU = 0.2 m/s)
−1.5 −1 −0.5 0 0.5 11.5
0
0.5
1
1.5
2
2.5
Re = 50K
X/d
<U> / ΔU (ΔU = 0.5 m/s)
Fig. 3 The mean of streamwise
velocity. Filled circle:
X/d=1, open square:
X/d=4.5, asterisk:X/d=15
Exp Fluids (2011) 50:1473–1493 1477
123
spatial resolution limitations, clusters of Kci with fewer
than three grid points across their span in both the xand
ydirections are not considered.
Comparing the definition of swirling strength above
with Robinson’s definition of a vortex, ‘‘A vortex exists
when instantaneous streamlines mapped onto a plane nor-
mal to the vortex core exhibit a roughly circular or spiral
pattern, when viewed from a reference frame moving with
the center of the vortex core’’ (Robinson 1991), the edges
of swirling strength clusters do not necessarily accurately
define the very edge of a vortex. By its definition, the
swirling strength is the indicator of how strong the swirling
motion of a particular local fluid particle is, or how fast the
particle rotates. Therefore, the swirling strength clusters are
only the parts of vortices that only include fluid particles
with distinguishable strong swirling motions. For those
vortices with weak rotation at the vortex edge, the filtering
process performed to eliminate the effect of the measure-
ment noise will also eliminate the swirling strength sig-
natures of these regions. A small fluid particle with close to
zero swirling strength, with no rotation or swirling motions
itself, could be part of large-scale coherent motion.
Therefore, the swirling strength clusters identified from the
instantaneous velocity fields are only defined as vortex
cores, or the highly rotating part of a vortex. As Agrawal
and Prasad (2003) argued, vortices in turbulent flows range
in size from the integral length scale down to the Kol-
mogorov scale, and there are always small-scale vortices
embedded in large-scale structures. Large-scale vortices
account for most of the turbulent energy, while the small
Y/d
X/d
-1 0 1
6.5
7
7.5
8
8.5
9
50
40
30
20
10
0
-10
-20
-30
-40
-50
Y/d
X/d
-1 0 1
6.5
7
7.5
8
8.5
9
Y/d
X/d
-1.5 -1 -0.5 0 0.5 1 1.5
0
0.5
1
1.5
2
50
40
30
20
10
0
-10
-20
-30
-40
-50
Y/d
X/d
-1.5 -1 -0.5 0 0.5 1 1.5
0
0.5
1
1.5
2
(a1)
(b1)
(a2)
(b2)
Fig. 4 Example of vortex identification in an instantaneous PIV
velocity field in near-field 0\X/d\2.5 and far-field 6.5 \X/d\9
of the jet, at Re =20K. aThe Galilean decomposition of this
instantaneous velocity field with 0.7U
c
, where U
c
is the mean velocity
of the center stream at each Xlocation. bLocal Galilean decompo-
sition of vortices with the velocity at each vortex core. The contours
of Kci ¼1:5Krms
ci are also shown in the background of the figure aand
b
1478 Exp Fluids (2011) 50:1473–1493
123
scales carry almost the entire vorticity. These swirling
strength clusters are those small-scale vortices in their
argument.
Figure 4shows that on the left side of the jet, the pos-
itively (counterclockwise) rotating vortices are dominant,
while negatively rotating vortices are dominant on the right
side of the jet. One may notice from the instantaneous
fields of vortices that in both the near-field and the far-field
of the jet, the distribution of the vortices seems random and
does not display as the large-scale structural array in planar
jets suggested by the Mumford (1982), Antonia et al.
(1986), nor the roller and helical modes revealed by low-
pass filtering axisymmetric jets (Agrawal and Prasad
2002a). The velocity fluctuation correlation reported by
Feng et al. (2007a) actually shows, in this jet, there are the
chain of large coherent structures in the stream-wise
direction, as previous researchers suggested in planar jets.
As mentioned earlier, the large-scale coherent motions
visualized by the correlation technique are not the swirling
motions of the small-scale vortices in the individual
instantaneous velocity field. Also, notice that besides those
dominant positively rotating vortices, there are also some
negatively rotating vortices on the left side of the jet. The
counter-rotating vortices are due to the small wake regions
downstream of the splitter plates. In the far-field of the jet,
the instantaneous vortex fields bear some resemblance to
the small-scale vortex field revealed by a high-pass filter in
the far-field of an axisymmetric jet (Agrawal and Prasad
2002b,2003), where small-scale vortices spread across the
jet body. Spatial correlation calculations with the swirling
strength were also performed in the present study, like the
calculations done by Christensen and Adrian (2001), who
found a strong position and angle preference between the
vortices in wall turbulence. The results here show that the
vortices in entire body of this jet have no strong spatial
correlations with other vortices at all, confirming the high
level of randomness observed in the instantaneous fields. In
addition to the vortices in the jet region, there are also some
vortices in the two boundary layers near the walls, which
are not of interest in the current study.
4.2 Vortex population
To identify the population trends of the vortices, a similar
vortex definition as Wu and Christensen (2006) is used to
define the population density of vortices, PpðnÞ.PpðnÞis
herein defined as the local ensemble-averaged number of
detectable positively rotating or negatively rotating vorti-
ces whose centers reside at a given PIV grid node. Figure 5
shows PpðnÞcross-streamwise profiles at 3 downstream
locations for the two different Reynolds numbers
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.5
1
Y/d
Per Realization
Re = 20K X/d = 1
−1.5 −1 −0.5 0 0.5 11.5
0
0.5
1
Y/d
Per Realization
Re = 50K X/d = 1
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.5
1
Y/d
Per Realization
Re = 20K X/d = 4.5
−1.5 −1 −0.5 0 0.5 11.5
0
0.5
1Re = 50K X/d = 4.5
Y/d
Per Realization
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.5
1
Y/d
Per Realization
Re = 20K X/d = 15
−1.5 −1 −0.5 0 0.5 11.5
0
0.5
1
Y/d
Per Realization
Re = 50K X/d = 15
Fig. 5 Local vortex population
density of positively and
negatively rotating vortices at
different streamwise locations.
Open circle: positively rotating
vortices, dot: negatively rotating
vortices
Exp Fluids (2011) 50:1473–1493 1479
123
investigated. To minimize scatter in the profiles, PpðnÞis
presented as the vortex population over a half jet width
(d/2 =1 cm) long area in the streamwise direction. The
plots presented in Fig. 5demonstrate that the distributions
of the positively and negatively rotating vortices are nearly
axially symmetric at all three downstream locations in both
two cases.
These profiles confirm the observations from the
instantaneous vortices fields above. At X/d=1 in the near-
field of the jet, the population density profiles of the
dominant direction vortices and counter-rotating vortices
both show clear peaks, at the locations on either side of the
center of the wake region, according to the streamwise
velocity mean profile in Fig. 3. The population of the
counter-rotating vortices almost disappears completely at
X/d=4.5, as the wake regions decay and are replaced by
mixing layer-like region population profiles. The popula-
tion profiles of the dominant vortices also broaden with
increasing downstream distance. These vortices are ini-
tially confined to the high shear, mixing layer regions with
few vortices existing in the free streams, but as the flow
develops toward channel flow at the farthest downstream
locations, the distribution of the vortices spreads through-
out the entire width of the reactor. At X/d=15, the pop-
ulation density profiles approach the vortex distribution of
a channel flow. One may notice the remarkable difference
between the results here and the small-scale vortex distri-
bution results in far-field of a axisymmetric free jet in
Agrawal and Prasad (2002b), which shows that the prob-
ability of finding a vortex is uniform up to the edge of the
jet and a substantial number of both clockwise and counter-
clockwise rotating eddies exist on both sides of the jet axis.
In the far-field of the jet, the two walls on the spanwise
direction produce the new source of the turbulence, which
reaching into the center of the jet as the flow develops. This
causes the small-scale vortices distributed as in a channel
flow, instead of a free jet.
Also notice that the highest vortex population actually
occurs in a very small region of the wall boundary layers,
which is not a surprise because of the higher velocity
gradient in the wall region. As mentioned earlier, the
population here only includes the vortices whose core
(swirling strength cluster) diameters are larger than 13.8g
in the low Re case and 33.9gin the high Re case.
By adding the number of positively rotating vortices
together between the center of the center stream Y/d =0
and the center of the left side stream Y/d =-1 in this
1-cm-long region, the streamwise population changing
trend of detectable positively rotating vortex can be more
easily observed, as shown in Fig. 6. The number of the
vortices in the same size region per velocity realization at
different downstream locations initially increased and then
decreased. This is reasonable because right after the flow
enters the test section, the high velocity gradients produce
many vortices. At the same time, the decay and dissipation
of these vortices also occurs. As the flow progresses further
downstream, the potential cores disappear, and fewer vor-
tices are produced. The vortex dissipation process contin-
ues, though, and later in the far-field of the jet it eventually
overwhelms the jet vortex production. Also notice the
population is much higher in the high Reynolds number
case than in the low Reynolds case, specially in the far-
field of the jet, where the flow approaches a fully devel-
oped channel flow and the boundary layers on the walls
play a bigger role in the vortex production than the shear
layer caused by the jet. Figure 6also shows that the
development toward channel flow is more rapid for the low
Re case, which of course makes sense, since the entrance
length in a channel or pipe increases with increasing
Reynolds number (Sadri and Floryan 2002).
4.3 Size and strength of vortex cores
In addition to the vortex population analysis presented
earlier, other properties of the identified vortices can be
obtained from the experimental data. The results of the size
of vortex core are presented in a similar manner to the
vortex populations. Figure 7shows the mean vortex core
size profiles across the channel, and Fig. 8gives stream-
wise profiles. The values of mean vortex core size appear
to be rather small compared to the jet width. Also the
smallest size of swirling strength cluster we can identify is
d
2
/400 for both investigated Reynolds numbers.
0 5 10 15 20 25 30
6
7
8
9
10
11
12
13
X/d
Per Realization
Re = 20K
Re = 50K
Fig. 6 Mean populations of positively rotating vortex in the half jet
width (1 cm) streamwise long region between the centers of the
center of center stream (Y/d=0) and the center of left side stream
(Y/d=-1). Open circle:Re =50K, filled square:Re =20K
1480 Exp Fluids (2011) 50:1473–1493
123
The cross-stream profiles in Fig. 7show that on both
sides of the jet where the dominant direction vortex pop-
ulation is highest, the average vortex size is also the largest,
especially at X/d=1. Although the population density of
the vortices in the wall boundary layers is highest, the size
of the vortices in the boundary layer is small. Figure 8
shows that the average vortex core size between the center
of the center stream and the left stream monotonically
decreases with increasing downstream distance. This might
initially seem contrary to what was reported in Feng et al.
(2007b). In the Re = 50K case, the size of the coherent
structures was observed linearly growing almost from
X/d=1 through at least X/d=15, which would be
expected for jet flow. Glancing back at the instantaneous
vortex field in Fig. 4, the areas of the identified swirling
strength clusters in the far-field are indeed smaller than in
the near-field. In the near-field of the jet, the current vortex
areas are about two orders of magnitude smaller than in the
previous report; in the far-field of the jet, they are almost
four order of magnitude smaller than in the previous report.
Again, this can be explained based on the difference
between the two measurement techniques. In Fig. 6, from
X/d=1toX/d=7.5, the populations of vortex cores
increase by at least 40%, while the mean sizes only
decrease by 15%. Therefore, the overall swirling area in the
jet grows by approximately 20%. However, as the jet
progresses downstream, there are more small structures
with increasing downstream distance, as previously men-
tioned. The increase in the number of small structures
brings the mean size values down. When the flow becomes
fully developed channel flow, the mean structure size will
remain constant beyond this point. This explains why the
vortex core size becomes smaller with increasing down-
stream distance and eventually stabilizes as the flow fully
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/ d2
Re = 20K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/ d2
Re = 50K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/ d2
Re = 20K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/ d2
Re = 50K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/ d2
Re = 20K X/d = 15
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.005
0.01
0.015
Y/d
Size/d2
Re = 50K X/d = 15
Fig. 7 Profiles of the size of
vortex cores. Open circle:
positively rotating vortices, dot:
negatively rotating vortices
0 5 10 15 20 25 30
0.0085
0.009
0.0095
0.01
0.0105
0.011
0.0115
0.012
X/d
Size/ d2
Re = 20K
Re = 50K
Fig. 8 Mean vortex size in the half jet width (1 cm) streamwise long
region between the centers of the center of center stream (Y/d=0)
and the center of left side stream (Y/d=-1). Open circle:
Re =50K, filled square:Re =20K
Exp Fluids (2011) 50:1473–1493 1481
123
develops. This is to be expected as the confined jet flow
transitions from a jet-like flow to a channel-like flow,
which would have the vortex size independent of down-
stream location.
The strength of the vortex cores also can be determined.
Note that inside the swirling strength clusters, the strength
value is not evenly distributed. The highest swirling
strength value of the vortex core is used to characterize the
vortex, since it gives the maximum rotation speed of the
given vortex, according to the swirling strength definition.
Figure 9shows cross-stream profiles of the average peak
value of the vortex core and Fig. 10 gives streamwise
profiles. All these profiles were normalized using DU=d.In
Fig. 9, it can be observed that the cross-stream profiles of
the maximum strength of vortex core show very similar
characteristics as the plot of the size of vortex core. The
center of the shear region of the jet has the maximum
vortex population density, the largest vortex size, and the
strongest swirling motion. Figure 10 shows that as the flow
develops, the strength of the vortices decreases, then
eventually stabilizes, just as the vortex size. Comparing the
two Reynolds number cases, the values for the Re =20K
case are higher than the Re =50K case after normaliza-
tion, especially at X/d=1. However, the difference
between the two Reynolds numbers becomes smaller and
smaller as the flow progresses downstream.
The cross-stream and streamwise profiles of both vortex
size and vortex strength all indicate that the high Reynolds
number case develops more slowly than the low Reynolds
number case, just as the profiles of vortex population
density show.
−1.5 −1 −0.5 0 0.5 1 1.5
−10
−5
0
5
10
Y/d
Λcid /ΔU
Re = 20K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
−5
0
5
Y/d
Λcid /ΔU
Re = 50K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
−4
−2
0
2
4
Y/d
Λcid /ΔU
Re = 20K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
−4
−2
0
2
4
Y/d
Λcid /ΔU
Re = 50K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
−4
−2
0
2
4
Y/d
Λcid /ΔU
Re = 20K X/d = 15
−1.5 −1 −0.5 0 0.5 1 1.5
−4
−2
0
2
4
Y/d
Λcid /ΔU
Re = 50K X/d = 15
Fig. 9 Profiles of the peak
swirling strength value of vortex
cores. Open circle: positively
rotating vortices, dot: negatively
rotating vortices
0 5 10 15 20 25 30
2
2.5
3
3.5
4
4.5
5
5.5
6
X/d
Λcid /ΔU
Re = 20K
Re = 50K
Fig. 10 Mean the peak swirling strength value of the vortex cores in
a half jet width (1 cm) streamwise long region between the centers of
the center of center stream (Y/d=0) and the center of left side stream
(Y/d=-1). Open circle:Re =50K, filled square:Re =20K
1482 Exp Fluids (2011) 50:1473–1493
123
4.4 Deviation velocity of vortex cores
The instantaneous velocity field data can also lend insight
into the motion of the vortices. The velocity of vortex cores
can be mapped back to the velocity field once the vortex is
identified. By subtracting the mean velocity from the
velocity values of the vortex cores, what we define as
the deviation velocity of the vortex can be obtained, i.e., the
deviation of the instantaneous vortex convection velocity
from the local mean velocity. Figures 11 and 12 show
profiles of the average cross-stream (V) and streamwise
(U) deviation velocities cross the channel at three down-
stream locations, which all are normalized by DU.
Figure 11 shows that the mean deviation velocity Vis
zero at the location of the population density peak.
Considering the left side of the jet, on the right side of
population density peak, vortices tend to move toward the
center of the center stream, and on the left side, the
vortices tend to move toward the center of the left side
(or outer) stream, which corresponds to the spreading of
the shear layers at the center regions of the jet. Particu-
larly, at X/d=1, the deviation velocity Vis maximum at
the very edge of the potential core of the free streams.
Since there are virtually no vortices in the potential core,
the average deviation velocity there is zero. In addition,
the normalized values of vortex deviation velocity Vare
higher in the low Reynolds number case than in the high
Reynolds number case.
The mean deviation velocity Uis also zero at the
location of the population density peak as shown in Fig. 12.
At X/d=1, toward the jet center, the values are negative,
and toward the side streams, the values are positive, which
corresponds to one of the roles of vortices in the flow:
transferring low momentum fluid to the high-velocity
region and transferring high momentum fluid to the low
velocity region. However, at further downstream locations,
X/d=4.5 and X/d=15, the profiles of mean deviation
velocity Uof the counter-rotating vortices overlap with the
profiles of dominant rotating direction vortices. Once
again, glancing back at the instantaneous vortex field in the
far-field of the jet in Fig. 4, these small number of counter-
rotating vortices are blending with the dominant direction
vortices.
The probability density function of vortex deviation
velocity was also calculated at three downstream locations,
X/d=1, X/d=4.5 and X/d=15 for the Re =20K case,
and compared to the PDF of the velocity fluctuations of the
overall flow, which are shown in Figs. 13,14 and 15.At
each downstream location, three observation points were
used, one is at the location of the peaks of positively rotating
vortex population density (Y/d =-0.45 at X/d=1,
Y/d=-0.42 at X/d=4.5), one is on the left side and one is
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
0
0.1
Y/d
Vλ/ΔU
Re = 20K X/d =1
−1.5 −1 −0.5 0 0.5 11.5
−0.1
−0.05
0
0.05
0.1
Y/d
Vλ/ΔU
Re = 50K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
−0.05
0
0.05
0.1
Y/d
Vλ/ΔU
Re = 20K X/d = 7.5
−1.5 −1 −0.5 0 0.5 11.5
−0.1
−0.05
0
0.05
0.1
Y/d
Vλ/ΔU
Re = 50K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
−0.05
0
0.05
Y/d
Vλ/ΔU
Re = 20K X/d = 15
−1.5 −1 −0.5 0 0.5 11.5
−0.05
0
0.05
Y/d
Vλ/ΔU
Re = 50K X/d = 15
Fig. 11 Vortex deviation
velocity in the cross-stream
direction. Open circle:
positively rotating vortices, dot:
negatively rotating vortices
Exp Fluids (2011) 50:1473–1493 1483
123
on the right side where the middle points of the population
density peak declining slopes are. In each case, all the
distributions of the deviation velocity appear narrower than
the velocity fluctuation distribution. One reason to explain
this is the swirling strength vortex identification method
reveals the core of the vortex. It is actually the part of the
−1.5 −1 −0.5 0 0.5 1 1.5
−0.2
−0.1
0
0.1
0.2
Y/d
Uλ/ΔU
Re = 20K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
−0.05
0
0.05
0.1
Y/d
Uλ/ΔU
Re = 50K X/d =1
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
0
0.1
Y/d
Uλ/ΔU
Re = 20K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
0
0.1
Y/d
Uλ/ΔU
Re = 50K X/d = 7.5
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
−0.05
0
0.05
0.1
Y/d
Uλ/ΔU
Re = 20K X/d = 15
−1.5 −1 −0.5 0 0.5 1 1.5
−0.1
−0.05
0
0.05
0.1
Y/d
Uλ/ΔU
Re = 50K X/d = 15
Fig. 12 Vortex deviation
velocity in the streamwise
direction. Open circle:
positively rotating vortices, dot:
negatively rotating vortices
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.57
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.45
−0.5 00.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.31
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.57
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.45
−0.5 00.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.31
Fig. 13 Vortex deviation
velocity distribution, at X/
d=1, Re =20K. Times:
velocity fluctuation distribution,
open circle: vortex deviation
velocity distribution
1484 Exp Fluids (2011) 50:1473–1493
123
vortex close to the vortex edge that generates high velocity
fluctuations, where the swirling motion could be very weak.
Another possible reason is the vortex identification method
used here does have spatial limitations, and vortices that are
too small to be detected could play some role in the flow
unsteadiness. It is also possible that some of the fluctuations
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
VΔU, Y/d = −0.72
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.42
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.16
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.72
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.42
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
U/ΔU, Y/d = −0.16
Fig. 14 Vortex deviation
velocity distribution, at X/
d=4.5, Re =20K. Times:
velocity fluctuation distribution,
open circle: vortex deviation
velocity distribution
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.82
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.52
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.21
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.82
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.52
−0.5 0 0.5
0
0.02
0.04
0.06
0.08
Value
Distribution
V/ΔU, Y/d = −0.21
Fig. 15 Vortex deviation
velocity distribution, at X/
d= 15, Re =20K. Times:
velocity fluctuation distribution,
open circle: vortex deviation
velocity distribution
Exp Fluids (2011) 50:1473–1493 1485
123
are due to other unsteadiness or large-scale coherent
motions in the jet flow, such as stream-wise jet pulsing or
cross-stream flapping of the jet.
Similar to the mean vortex deviation velocity profiles, at
the location of the population density peak, the mean of
both the vortex deviation velocities are zero. However, the
mean of velocity fluctuations are also close to zero at this
location. Interestingly, on both sides of the peak location,
the distribution peaks of vortex deviation velocity appear
on the opposite side of zero as the distribution peaks of
velocity fluctuation. It makes sense that the velocity fluc-
tuation close to the jet center is negative, because low
momentum fluid is transferring into the jet center (where
the mean velocity is high); the velocity fluctuation close to
the outer stream center (where the mean velocity is low) is
positive, because high momentum fluid is transferring in.
Notice also that compared to the distributions at X/d=1,
as downstream distance increases, the vortex deviation
velocity distributions become closer to the distribution of
the velocity fluctuation (an exception to this is at X/d=4.5
and Y/d =-0.16, where the streamwise velocity fluctua-
tion distribution has a much higher peak, which is simply
because the location is very close to the jet center). As the
flow progresses further downstream, the vortex deviation
velocity should become more similar to the distribution
shape of the velocity fluctuations, as the mean vortex
deviation velocity profiles indicate.
5 Vortex tracking analysis
From the high speed PIV data, the trajectories of individual
vortices can also be determined. After identifying a par-
ticular vortex in a pair of velocity realizations, the velocity
of the swirling strength cluster can be found from the
change in the vortex position between the two frames.
Repeating this frame by frame over the high speed PIV
data set, the location of the vortex in each frame can be
estimated. The vortex on the next frame which has the
maximum overlap with the estimated location is considered
the same vortex as the selected vortex from the previous
frame. Figure 16 shows two trajectories of a positively
rotating vortex on the left side of the jet and a negatively
rotating vortex on the right side of the jet to demonstrate
the effectiveness of this tracking method. Only vortices that
newly appear at the bottom of the velocity fields were
traced to maximize tracking time because the length of
observation window is limited. Once a vortex has been
identified in a series of velocity realizations at different
time steps, how the size and strength of the vortex changes
with time can be determined. Because the vortex property
development of one given vortex could be rather noisy, to
obtain the main characteristics of vortex development,
averaging was performed for certain vortices which trav-
eled through the current watching window, t0:2 s for
Re =20K and t0:08 s for Re =50K. Those vortices
which disappeared during this time frame are not included.
The mean vortex size and strength are shown in Figs. 17
and 18, using observation windows beginning at X/d=0,
X/d=3.5 and X/d=7.5. Since the experimental data at
these three locations were collected at different times, the
vortices that were tracked in one window definitely could
not appear in the other windows. The mean tracked vortex
maximum strength decreases with increasing downstream
distance after the flow enters the channel, the same
development as we observed in the previous section.
However, the mean tracked vortex size increases slightly
after the vortices enter the reactor, and then decreases. The
vortices that were tracked close to the jet outlet must have
already existed when the flow entered the test section. They
are likely those well organized vortices reported in wall
turbulence (Adrian et al. 2000b), and continue to develop
for a while before the changing of the flow field affects
them. When these vortices meet the counter-rotating vor-
tices on the other side of the splitter plates, and when
vortices begin to be produced by the mean shear in the jet,
the properties of the vortex core will change. They can be
weakened or pushed aside by the counter-rotating vortices.
Also, the swirling strength values in the swirling strength
clusters could be redistributed, as some parts of the rotating
fluid particles can spin off from the cluster, a phenomenon
observed in some instantaneous vortices fields from the
high speed PIV data.
Also compared to the mean vortex size and strength
profiles, at X/d=1 and X/d=4.5, presented in Figs. 7and
Y/d
X/d
Re = 20K
−1.5 −1 −0.5 0 0.5 1 1.5
0
0.5
1
1.5
2
2.5
−60
−40
−20
0
20
40
60
Fig. 16 The trajectories of a positively rotating vortex and a
negatively rotating vortex, at the jet outlet, Re =20K. The
background contours are the plots of the swirling strength clusters
at several different time steps
1486 Exp Fluids (2011) 50:1473–1493
123
9, all the values calculated by tracing individual vortices
using high speed PIV data are higher. This is because the
averaging here is actually conditional averaging as
mentioned earlier. Those vortices chosen here have rela-
tively long life spans, a probable consequence of their
being larger and stronger than the ‘‘average’’ vortex.
0.05 0.1 0.15
2
3
4
5
6
7
Λcid /ΔU
Starting at X/d = 0
0.05 0.1 0.15
2
3
4
5
6
7
X/d = 3.5
0.05 0.1 0.15
2
3
4
5
6
7
X/d = 15
0.04 0.06
2
3
4
5
6
7
time (sec)
Λcid /ΔU
Starting at X/d = 0
0.04 0.06
2
3
4
5
6
7
time (sec)
X/d = 3.5
0.04 0.06
2
3
4
5
6
7
time (sec)
X/d = 15
Re = 50K
Re = 20K
Fig. 17 The mean tracked
vortex maximum strength
starting at three different
downstream locations
0.05 0.1 0.15
0.015
0.02
0.025
0.03
Size / d2
Starting at X/d = 0
0.05 0.1 0.15
0.015
0.02
0.025
0.03 X/d = 3.5
0.05 0.1 0.15
0.015
0.02
0.025
0.03 X/d = 15
0.04 0.06
0.02
0.025
0.03
time (sec)
Size / d2
Starting at X/d = 0
0.04 0.06
0.015
0.02
0.025
0.03
time (sec)
X/d = 3.5
0.04 0.06
0.015
0.02
0.025
0.03
time (sec)
X/d = 15
Re = 50K
Re = 20K
Fig. 18 The mean tracked
vortex size starting at three
different downstream locations
Exp Fluids (2011) 50:1473–1493 1487
123
6 Spatial correlation functions and linear stochastic
estimation
6.1 Two-point spatial correlations
Normalized by the rms of swirling strength and some flow
property, r, the two point spatial cross correlation of
swirling strength and the fluctuation of rcan be defined as
Rkr0ðX;Y;x;yÞ¼ hkciðX;YÞr0ðx;yÞi
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hk2
ciðX;YÞihr02ðx;yÞi
qð2Þ
where (X,Y) and (x,y) are the coordinates of the basis point
(i.e. the point about which the correlation is calculated) and
an arbitrary point in the flow field, respectively. In the
present study, locations along the left peak of turbulent
kinetic energy at different downstream locations were
chosen as basis points. This location also corresponds to
the peak of the population density of positively rotating
vortices. The spatial correlations of Rku0and Rkv0for basis
points located at four downstream locations for both Re
cases are presented in Figs. 19 and 20. The streamwise
flow direction is from the bottom to the top of the figures.
The cross correlation fields of swirling strength and
fluctuations of the two velocity components exhibit ‘‘but-
terfly’’ like shapes. The right wing of Rku0contains positive
values and the left wing contains negative values. The
symmetry line of this ‘‘butterfly’’ shape in Rku0, called ‘‘the
axis’’ of these correlation fields hereafter, is oriented ver-
tically in the streamwise direction, which the contour line
of Rku0=0 overlaps in the region of the basis point. The
symmetry line of Rkv0is oriented horizontally with negative
values downstream of the basis point and positive values
upstream of the basis point. Notice at X/d=1, the axis of
Rku0is not oriented vertically as it is at the other down-
stream locations. Instead, the axis tilts about 30°away from
the jet center, which is caused by high velocity gradients in
this narrow wake region. Once the wake region disappears,
the tilting of the axis becomes less significant due to
smaller and smaller velocity gradients. There is a second
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
Fig. 19 Two-point cross correlation Rkciu0(first row), Rkci v0(second row). At X/d=1, X/d=4.5, X/d=7.5 and X/d=15, Re =50K. The black
dot indicates the location of swirling strength basis point
1488 Exp Fluids (2011) 50:1473–1493
123
positively correlated region on the left side of the negative
peaks, and a second negatively correlated region on the
right side of the positive peaks in the Rku0contours, espe-
cially in the low Re case, at X/d=1. Also, there is a
weakly correlated negative region upstream of the strong
positive region close to the basis point and a weak positive
region downstream of the strong negative region in Rkv0.
Assuming the vortex at the basis point is accompanied by
vortices upstream and downstream of the basis point
location, these two correlation fields indicate locations of
accompanying vortices. Also the correlation in the region
far from the basis point is actually very weak (that is close
to zero), which makes the location of the Rku0=0 contour
somewhat random. This agrees with the observation from
the instantaneous vortex fields, in Fig. 4. Also, as the flow
progresses downstream, the correlation areas of both
Rku0=0 and Rkv0=0 grow larger. Notice the peak values
of these correlations in Fig. 22. The peak values of Rku0are
approximately 0.2 for both cases close to the jet inlet. The
peak values of Rkv0begin at approximately 0.4 near jet inlet
in both cases and decrease with downstream distance,
approaching a value of 0.25 for both Reynolds number
cases after the flow passes beyond X/d=12.
Figure 21 shows Rkci/0, the cross-corrlation of swirling
strength with the concentration fluctuation, for Re =50K
and Re =20K, respectively. Note that simultaneous PIV/
PLIF data were only collected for two downstream loca-
tions for Re =20K, compared to six locations for
Re =50K. There are two correlated areas in the contours;
one positively correlated region downstream of the basis
point and one negatively correlated region upstream. This
shape suggests that the positively rotating (counterclock-
wise) vortices bring high concentration fluid (i.e., positive
concentration fluctuations) from the center stream to the
side stream downstream of the vortex core and bring the
low concentration fluid from the side stream to the center
stream upstream of the core. These two correlation areas
are not only comparable in size but also in absolute peak
values. As mentioned before, the locations chosen here also
are the locations of the peak of positive vortex population
density, which means there are predominately positively
rotating vortices centered at this location. The contour lines
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
Fig. 20 As Fig. 19 but at Re =20K
Exp Fluids (2011) 50:1473–1493 1489
123
of Rkci/0¼0:1 and Rkci /0¼0:1 are nearly symmetrical
about the basis point, except at the location X/d=1. As
previously stated, because the axis of positively rotating
vortices close to the jet outlet is obliquely oriented, positive
Rkci/0is located in the second quadrant instead of the first
quadrant. Considering these two regions together provides
an explanation for the elliptic shape of the R/0/0with the
major axis inclined at 45°, reported in a previous paper
(Feng et al. 2007a). Also, considering the location of the
correlation peaks with respect to the basis point, the results
are consistent with the location of the vortex core in linear
stochastic estimates of the velocity field given the event of
/0(x
o
)=?2/0
rms
(x
o
) (Feng et al. 2007a). Notice also that
with increasing downstream distance, the size of the cor-
relation area becomes steadily larger. However, the peak
values of Rkci/0, shown in Fig. 22, also first increase, then
decrease in the high Re case. Compared to the high Re
case, the correlation area in the low Re case is considerably
larger and the peak values are lower. Also, unlike the high
Re case, the negative region not only occupies most of the
third quadrant but also the fourth quadrant, with the
contour line of Rkci/0¼0 inclined at approximately 30°
with respect to the xdirection.
6.2 Linear stochastic estimation
One useful tool to interpret spatial correlation data is
linear stochastic estimation, in which conditional aver-
ages are calculated from measured correlation fields
(Adrian 1994; Olsen and Dutton 2002,2003) Although
conditional averages can be calculated directly from an
experiment dataset, this requires either a very large
ensemble size or averaging over a large flow region
instead of a particular location in the flow, due to the
low population density of the vortices. Derived directly
from the two point spatial correlation, LSE can give the
typical underlying flow structures more precisely with a
smaller ensemble size.
Letting k
ci
(X,Y) be the swirling strength value at loca-
tion (X, Y), the linear stochastic estimate of the velocity
fluctuation u
i
0(x,y) over the entire flied given the condition
k
ci
(X,Y) is,
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.8
7
7.2
7.4
7.6
7.8
8
8.2
Fig. 21 Two-point cross-correlation Rkci/0.AtX/d=1, X/d=4.5, X/d=7.5 and X/d=15, for Re =50K (first row) and at X/d=4.5, X/
d=7.5, for Re =20K (second row)
1490 Exp Fluids (2011) 50:1473–1493
123
hu0
iðx;yÞjkciðX;YÞi  Likci ðX;YÞð3Þ
The linear coefficient, L
i
, can be determined in the
following way,
Li¼hu0
iðx;yÞkciðX;YÞi
hkciðX;YÞkci ðX;YÞi ð4Þ
where hu0
iðx;yÞkciðX;YÞi is the (unnormalized) spatial
correlation of u
i
0and k
ci
.
As Christensen and Adrian (2001) argued, since the
event of the conditional average is a single scalar value, the
character of the conditionally averaged results remains the
same for all values, just the magnitudes are simply
amplified and attenuated. Thus, the thresholding of the k
ci
is not necessary for just examining the fluctuation field. But
to see the real typical structure in the flow field, the esti-
mated fluctuation fields need to be added with the mean
field. Thus, the thresholding of the event will only change
the size of the structure. Here we chose the event of kci ¼
2krms
ci in order to get the features of typical structure in the
flow.
Adding the mean field and subtracting U(x
o
) from each
vector results in Fig. 23, which shows the typical velocity
field of one vortex at different downstream locations. All
the roller structures revealed by the linear scholastic esti-
mation are spindle-shaped with a stream-wise major axis.
The tilting of the major axis becomes much more visible
here than in the spatial correlation fields. The high
momentum of the jet pushes the upper part of the roller
structure outwards, and the low momentum fluid was
entrained inwards to the center of the jet by the lower part
of the structure. In the near-field of the jet, there are two
vortices seen adjacent to the vortex at the base point in both
upstream and downstream direction. As the flow progresses
downstream, the roller structures grow larger. Although the
sizes of the vortex cores decrease, the large-scale coherent
motions of the flow, which the swirling motions at the base
point are part of, are indeed growing. Comparing the two
Reynolds number case, the roller structures for Re =20K
are larger than those in the Re =50K case.
7 Conclusions
A vortex identification method based on swirling strength
was employed to analyze the properties of vortices in a
confined rectangular jet. Experimental data from simulta-
neous PIV/PLIF experiments and high speed PIV experi-
ments were used in this analysis. Swirling strength fields
were computed from velocity fields, and then filtered with
a universal threshold of jKcij1:5Krms
ci . By identifying
clusters of filtered Kci, vortex structures were identified.
Instantaneous swirling strength field data indicate that
positively (counterclockwise) rotating vortices are domi-
nant on the left side of the jet and negatively (clockwise)
rotating vortices are dominant on the right side. The pop-
ulation density, average size and strength, deviation
velocity of vortices were calculated and analyzed, in both
the cross-stream direction and the streamwise direction. In
the region close to the channel inlet, the population density,
average size and strength all show high values on both
sides of the center stream. There are some counter-rotating
vortices next to the dominant direction vortices that are
indicative of a wake region formed downstream of the
splitter plate tips by the boundary layers that form on both
sides of the splitter plates. At the further downstream
location, X/d=3.5, the wakes disappear, as do most of the
counter-rotating vortices. As the flow develops toward
channel flow at the farthest downstream locations, the
distribution of the vortices spreads throughout the entire
reactor. The mean size and strength of the vortices decrease
continuously downstream from the channel inlet. The mean
vortex deviation velocity in both the Xand Ydirections are
zero at the location of the population density peak. The
signs of vortex deviation velocity Vindicate the vortices
move from the high vortex population region to the low
vortex population region. The signs of mean deviation
velocity Uare negative on the side near the jet center and
positive on the side of near center of the outer stream,
which indicate vortices transfer low momentum fluid to
high-velocity region and transfer high momentum fluid to
the low velocity region.
The development trends of vortex size and strength were
also identified by tracking vortices using high speed PIV
experimental data. Both the average tracked vortex
strength and size decrease with increasing downstream
Fig. 22 Average of positive and negative peak values of cross
correlations. (open square:Rku0;open triangle:Rkv0;times:Rk/0;solid
line:Re ¼20K;dotted line:Re ¼50K)
Exp Fluids (2011) 50:1473–1493 1491
123
distance overall. However the average tracked vortex size
increases before it starts to decrease in the area close to the
jet inlet.
Two point spatial cross-correlations of swirling strength
with velocity fluctuations and concentration fluctuations
were calculated at the location of the left peak of turbulent
kinetic energy. The cross-correlation fields of swirling
strength and fluctuations of the two velocity components
exhibit a ‘‘butterfly’’ like shape. The right wing of Rku0
contains positive values and the left wing contains negative
values. The axis of Rku0, the contour line of Rku0=0, is
oriented vertically in the streamwise direction, slightly
tilted toward the outer stream. With increasing downstream
distance, the angle of orientation of the Rku0axis becomes
smaller. Also at the X/d=1.0 downstream locations for
both Reynolds number cases, there are a weak negatively
correlated regions upstream of the strong positively cor-
related region close to the basis point and a weak posi-
tively correlated region downstream of the strong
negatively correlated region in Rkv0. This indicates that a
vortex at the basis point is usually accompanied by at least
one counter-rotating vortex. The axis of Rkv0is aligned
with the cross-stream direction with negative values
downstream of the basis point and positive values
upstream of the basis point. There are also two correlated
areas in the Rk/0correlation field; one positively correlated
region downstream of the basis point and one negative
upstream of the basis point, indicating that the positively
rotating vortices bring high concentration field from the
center stream to the side stream downstream of the vortex
core and bring the low concentration field from the side
stream to the center stream upstream of the core. Finally,
linear stochastic estimation was used to calculate condi-
tional structures. The estimation was based on the swirling
strength values at chosen locations in the flow. The large-
scale structures in the velocity field revealed by linear
stochastic estimation are spindle-shaped with a titling
stream-wise major axis.
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.5
1
1.5
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
6.6
6.8
7
7.2
7.4
7.6
7.8
8
8.2
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
15.8
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
0.5
1
1.5
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2 0
3.8
4
4.2
4.4
4.6
4.8
5
5.2
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2
6.8
7
7.2
7.4
7.6
7.8
8
8.2
0.5 m/s
Y/d
X/d
-1 -0.8 -0.6 -0.4 -0.2
14.2
14.4
14.6
14.8
15
15.2
15.4
15.6
0.5 m/s
Fig. 23 Linear stochastic estimation of velocity field with U(x
o
) subtracted from each vector. At X/d=1, X/d=4.5, X/d=7.5 and X/d=15,
Re =50K (first row), Re =20K (second row)
1492 Exp Fluids (2011) 50:1473–1493
123
Acknowledgments This work was supported by the National Sci-
ence Foundation through Grants CTS-9985678 and CTS-0336435.
References
Abramovich G, Girshovich T, Krasheninnikov S, Sekundov A,
Smirnova I (1984) The theory of turbulent jets. Moscow Izdatel
Nauka
Adrian RJ (1994) Stochastic estimation of conditional structure: a
review. Appl Sci Res 53:291–303
Adrian RJ (1999) Uses, analysis and interpretation of piv data. 30th
AIAA fluid dynamics conference (c)
Adrian RJ, Christensen KT, Liu ZC (2000a) Analysis and interpre-
tation of instantaneous turbulent velocity fields. Exp Fluids
29:275–290
Adrian RJ, Meinhart CD, Tomkins CD (2000b) Vortex organization
in the outer region of the turbulent boundary layer. J Fluid Mech
422:1–54
Agrawal A, Prasad AK (2002a) Organizational modes of large-scale
vortices in an axisymmetric turbulent jet. Flow Turbul Combust
68(4):359–377
Agrawal A, Prasad AK (2002b) Properties of vortices in the self-
similar turbulent jet. Exp Fluids 33:565–577
Agrawal A, Prasad AK (2003) Measurements within vortex cores in a
turbulent jet. J Fluids Eng 125(3):561
Antonia R, Bilger R (2006) An experimental investigation of an
axisymmetric jet in a co-flowing air stream. J Fluid Mech
61(04):805–822
Antonia RA, Browne LWB, Rajagopalan S, Chambers AJ (1983) On
the organized motion of a turbulent plane jet. J Fluid Mech
134:49–66
Antonia RA, Chambers AJ, Britz D, Browne LWB (1986) Organized
structures in a turbulent plane jet: topology and contribution to
momentum and heat transport. J Fluid Mech 172:211–229
Benayad S, Salem A, Legrand J (2001) Turbulent mixing in a
coflowing liquid jet. Ind Eng Chem Res 40(3):927–932
Bradbury L, Riley J (1967) The spread of a turbulent plane jet issuing
into a parallel moving airstream. J Fluid Mech 27(02):381–394
Chhabra S, Huq P, Prasad A (2006) Characteristics of small vortices
in a turbulent axisymmetric jet. J Fluids Eng 128:439
Christensen KT, Adrian RJ (2001) Statistical evidence of hairpin
vortex packets in wall turbulence. J Fluid Mech 431:433–443
Christensen KT, Adrian RJ (2002a) Measurement of instantaneous
eulerian acceleration fields by particle image accelerometry:
method and accuracy. Exp Fluids 33(6):759–769
Christensen KT, Adrian RJ (2002b) The velocity and acceleration
signatures of small-scale vortices in turbulent channel flow.
J Turbul 3:1–28
Chu PCK, Lee JHW, Chu VH (1999) Spreading of turbulent round jet
in coflow. J Hydraul Eng 125(February):193
Curtet R (1958) Confined jets and recirculation phenomena with cold
air. Combust Flame 2(4):383–411
Deo R, Mi J, Nathan G (2007a) The influence of nozzle aspect ratio
on plane jets. Exp Therm Fluid Sci 31(8):825–838
Deo R, Nathan G, Mi J (2007b) Comparison of turbulent jets issuing
from rectangular nozzles with and without sidewalls. Exp Therm
Fluid Sci 32(2):596–606
Enjalbert N, Galley D, Pierrot L (2009) An entrainment model for the
turbulent jet in a coflow. C R Me
´c 337(9–10):639–644
Feng H, Olsen MG, Liu Y, Fox RO, Hill JC (2005) Investigation of
turbulent mixing in a confined planar-jet reactor. AIChE J
51(10):2649–2664
Feng H, Olsen MG, Fox RO, Hill JC (2007a) Conditional statistics for
passive-scalar mixing in a confined rectangular turbulent jet.
Phys Fluids 19, art. no. 055104
Feng H, Olsen MG, Hill JC, Fox RO (2007b) Simultaneous velocity
and concentration field measurements of passive-scalar mixing
in a confined rectangular jet. Exp Fluids 42:847–862
Gazzah MH (2010) The dynamic field in turbulent round jet
discharging into a co-flowing stream. Nat Sci 02(06):635–640
Goldschmidt V, Bradshaw P (1973) Flapping of a plane jet. Phys
Fluids 16:354
Goldschmidt V, Young M, Ott E (1981) Turbulent convective
velocities (broadband and wavenumber dependent) in a plane jet.
J Fluid Mech 105:327–345
Gordeyev S, Thomas F (2000) Coherent structure in the turbulent
planar jet. Part 1. Extraction of proper orthogonal decomposition
eigenmodes and their self-similarity. J Fluid Mech 414:145–194
Grinstein F (2001) Vortex dynamics and entrainment in rectangular
free jets. J Fluid Mech 437:69–101
Gutmark E, Wygnanski I (1976) The planar turbulent jet. J Fluid
Mech 73(03):465–495
Gutmark EJ, Grinstein FF (1999) Flow control with noncircular jets.
Annu Rev Fluid Mech 31(1):239–272
Lu Y, Dixon AG, Moser WR, Ma YH (1997) Analysis and
optimization of cross-flow reactors with staged feed policies
isothermal operation with parallel-series, irreversible reaction
systems. Chem Eng Sci 52:1349–1363
Mumford JC (1982) The structure of the large eddies in fully
developed turbulent shear flows. Part 1. The plane jet. J Fluid
Mech 118:241–268
Nickels T, Perry A (1996) An experimental and theoretical study of
the turbulent coflowing jet. J Fluid Mech 309:157–182
Oler J, Goldschmidt V (1982) A vortex-street model of the flow in the
similarity region of a two-dimensional free turbulent jet. J Fluid
Mech 123:523–535
Olsen MG, Dutton JC (2002) Stochastic estimation of large structures
in an incompressible mixing layer. AIAA J 40(12):2431–2438
Olsen MG, Dutton JC (2003) Planar velocity measurements in a
weakly compressible mixing layer. J Fluid Mech 486:51–77
Piomelli U, Yu Y, Adrian RJ (1996) Subgrid-scale energy transfer
and near-wall turbulence structure. Phys Fluids 8:215–224
Pope S (2000) Turbulent flows. Cambridge University Press,
Cambridge
Prasad AK, Adrian RJ, Landreth CC, Offutt PW (1992) Effect of
resolution on the speed and accuracy of particle image veloc-
imetry interrogation. Exp Fluids 13:105–116
Robinson SK (1991) Coherent motions in the turbulent boundary
layer. Annu Rev Fluid Mech 23
Sadri RM, Floryan JM (2002) Accurate evaluation of the loss
coefficient and the entrance length of the inlet region of a
channel. J Fluids Eng 124(3):685–694
Steward F, Guruz A (1977) Aerodynamics of a confined jet with
variable density. Combust Sci Technol 16(1):29–45
Thomas F, Brehob E (1986) An investigation of large-scale structure
in the similarity region of a two-dimensional turbulent jet. Phys
Fluids 29:1788
Tomkins CD, Adrian RJ (2003) Spanwise structure and scale growth
in turbulent boundary layers. J Fluid Mech 490:37–74
Vollmers H (2001) Detection of vortices and quantitative evaluation
of their main parameters from experiment velocity data. Meas
Sci Tech 12:1199–1207
Westerweel J, Hofmann T, Fukushima C, Hunt J (2002) The
turbulent/non-turbulent interface at the outer boundary of a
self-similar turbulent jet. Exp Fluids 33(6):873–878
Westerweel J, Fukushima C, Pedersen J, Hunt J (2005) Mechanics of
the turbulent-nonturbulent interface of a jet. Phys Rev Lett
95(17):174,501
Wu Y, Christensen KT (2006) Population trends of spanwise vortices
in wall turbulence. J Fluid Mech 568:55–76
Exp Fluids (2011) 50:1473–1493 1493
123
... This technique deduces conditionally velocity fields based on defined conditional events. This event, or signal, can be selected from any number of the flow parameters such as velocity or concentration fluctuation (Feng et al., 2007;Kong et al., 2011), the deformation tensor (Olsen and Dutton, 2002;, and pressure fluctuation (Hall and Ewing, 2010). ...
... First introduced by Adrian (1975;1994), Moin (1988), andAdrain et al. (1989), linear stochastic estimation has since been adopted by numerous researchers to analyze a wide range of fluid flows. For instance, Feng et al. (2007) and Kong et al. (2011) combined the PIV-PLIF and linear stochastic estimation (LSE) with the concentration fluctuation as the conditional event to characterize large-scale structures produced by the mixing of confined rectangular jets. By using the deformation tensor as a signal, Olsen and Dutton (2002) identified structural behaviors in incompressible mixing layers. ...
Article
Improving the efficiency of the production of uniformly sized functional nanoparticles for pharmaceutical and agricultural applications has been a problem of great interest. The macroscale multi-inlet vortex Reactor (macro-MIVR) could potentially be used for this purpose due to its ability to achieve the rapid mixing necessary for the flash nanoprecipitation nanoparticle fabrication technique. In the presented work, the coherent structures, a key contributor to the turbulent mixing, were investigated for the turbulent swirling flow within the MIVR. The two-point spatial correlations of velocity and concentration fluctuations at various basepoints were measured from instantaneous velocity and concentration fields obtained using simultaneous stereoscopic particle image velocimetry and planar laser-induced fluorescence. The basepoint locations were chosen as the middle and at the edge of the partially mixed concentration spiral arms. The correlations were found to be elliptical in shape, inclined, and peaked at the basepoints. A region near the basepoint was positively correlated and was surrounded by negatively correlated regions. Autocorrelations of concentration were also elliptical and curved toward the center of the reactor. The linear stochastic estimation was used to interpret the coherent structure features that would result in the observed spatial correlations. The linear stochastic estimates of the velocity fields were computed directly from the cross correlations of the tangential velocity fluctuations with the concentration fluctuations. The estimated conditional velocity fields revealed obliquely oriented counter-rotating vortical structures that stir the fluid from high-concentration regions to low-concentration regions, and the orientation of these vortical structures depended on the local concentration gradient.
... Accordingly, numerous flow geometries and methods that aim to enhance mixing have been extensively studied. Examples of mixingenhancing geometries include confined co-flow (Liu et al. 2008;Feng et al. 2007; Kong et al. 2011) and cross flow Page 3 of 13 28 jets (Wegner et al. 2004;Smith and Mungal 1998), rotating geometries such as rotary kilns (Larsson et al. 2015;Finnie et al. 2005) and mixing tabs (Gretta and Smith 1993;Chua et al. 2003;Bohl and Foss 1996) among many others. ...
... Several researchers have used PIV and PLIF in other turbulent mixing flows to simultaneously measure in-plane velocity and scalar concentration fields. Examples of the canonical flows that have been extensively studied using this technique are turbulent jet mixing and combustion in turbulent jets (Larsson et al. 2015;Kong et al. 2011;Feng et al. 2007;Filatyev et al. 2007;Yimer et al. 2002). However, there is a lack of similar investigations in complex swirling geometries such as the MIVR. ...
Article
Full-text available
Simultaneous measurements of velocity and concentration using stereoscopic particle image velocimetry (stereo-PIV) and planar laser-induced fluorescence (PLIF) were used to investigate the mixing performance of a scaled-up multi-inlet vortex reactor (MIVR). Data were collected in three measurement planes located at different heights from the reactor bottom (¼, ½, and ¾ of the reactor height) for Reynolds numbers of 3250 and 8125 based on the reactor inlet velocity and hydraulic diameter. The collected data were analyzed to determine turbulent flow statistics such as turbulent viscosity, turbulent diffusivity, and turbulent Schmidt number. When analyzed across 16 different azimuth angles and radial positions (r) normalized by the reactor radius (Ro), the turbulent viscosity was found to be nearly axisymmetric. In the free-vortex region (r/Ro > 0.2), the turbulent viscosity results were nearly constant. Near the center of the reactor in the forced-vortex region (r/Ro < 0.1), the turbulent viscosity significantly increased, with peak values occurring near the center. The turbulent viscosity and Reynolds shear stress were highest near the reactor exit at the ¾ plane. The dominance of high turbulent fluxes and low concentration gradients near the reactor center led to high turbulent diffusivity. Away from the center, the turbulent diffusivity was reduced because of large concentration gradients and low turbulence intensity in the spiral arm region. The turbulent Schmidt numbers were also found to correlate with concentration gradients. The turbulent Schmidt number values were found to vary from 0.1 to 1.2. The highest spatial variation in \(S{c_t}\) was observed in the spiral arms region, where the concentration gradients are also the highest. This spatial variation in Schmidt number contrasts with the common assumption of constant \(S{c_t}\) in Reynolds-averaged CFD models. Graphical abstract A typical stereo-PIV/PLIF simultaneous instantaneous measurement for Re = 8125. The color and vectors represent the instantaneous mixture fraction and in-plane velocity field, respectively. (a) ¼ Plane. (b) ½ Plane. (c) ¾ Plane. Open image in new window
... Some researchers have carried out the vortex tracking studies. For example, Kong et al. [29] have obtained the individual vortex trajectories by identifying the vortex positions between two successive frames. Berson et al. [30] have successfully tracked the vortex core locations by using two Galilean invariant functions. ...
... Actually, we have compared vortex identification by the Q and λ ci criterions, which yield qualitatively identical and quantitatively similar results comparing with the Subsequently, based on obtained λ 2 -identified vortex distributions, we develop a novel vortex tracking method. This method is inspired by PIV principles and Kong et al. [29], in that the discrete vortices (the connected regions with λ 2 < 0) are treated as individual "particles". Specifically, the λ 2 distribution is plotted using a 2-color image where vortices are shown in white like the seeding particles and the remaining non-rotating regions (λ 2 ≥ 0) are shown in black. ...
... Jet flapping and jet precession (defined as rotation of the entire jet with respect to the nozzle axis) have been identified as the primary patterns of oscillation in confined jets. The former usually occurs in planar jets under rectangular confinement [15][16][17][18]. The latter has been encountered in cases of round jets expanding into a concentric cylindrical chamber, which also induces a swirling flow within the confinement [19][20][21]. ...
Article
Full-text available
Self-sustained jet flapping is observed in a confined, premixed and preheated methane-air turbulent flame, generated in a single-nozzle jet-stabilized gas turbine model combustor designed based on the FLOXⓇ concept. The flapping frequency and its complex motion within the confinement of the combustor are characterized in detail using proper orthogonal decomposition (POD) of the flow fields measured by particle imaging velocimetry (PIV). The influence of jet flapping on combustion stability is examined using simultaneous PIV/OH chemiluminescence imaging and PIV/planar laser-induced fluorescence of OH radicals (OH PLIF) at 5 kHz repetition rate. By influencing the size and location of the recirculation zones, jet flapping modifies the flame shape and flame lift-off height. It also controls the amount of hot gas entrainment into the recirculation zones. In extreme cases, jet flapping is found to cause temporary local extinction of the flame, due to jet impingement on the combustor wall and partial blockage of burned gas entrainment. The flame is only able to recover after the jet detaches from the wall and reopens the back flow channel. The results suggest that jet flapping could play a key role in the stabilization mechanisms in similar jet-stabilized combustors.
... Note that the edges of the swirling strength contours do not necessarily accurately define the very edge of a vortex, as defined by Robinson [31]. Therefore, the swirling strength clusters are only the parts of vortices that include fluid particles with distinguishable strong swirling motions in the measurement plane [30]. Figure 4 clearly shows that the majority of the identified vortex clusters are located in the plane of the jet and a less number in the ambient nonturbulent flow. ...
Article
The developing region of a turbulent square jet is investigated using high-resolution particle image velocimetry (PIV). The mean velocity and turbulence stresses are presented in various horizontal planes, along the jet centerline covering the initial region of the jet as well as the transition to the self-similar region. To study the flow structure away from the central plane, velocity measurements in two additional horizontal planes, one located halfway from the jet central plane toward the edge and the other at the edge of the square jet, are also examined. Analysis of the instantaneous velocity fields reveal the presence of an arrow-like feature in the square jet due to the higher instability generated in the jet shear layer compared with a round jet. To elucidate the imprints of the vortex structures present in the jets, a swirling strength-based vortex identification methodology is applied on a large ensemble of instantaneous velocity fields. Statistical analysis of the number of vortex cores, and their size and rotational strength in the measurement plane is undertaken. Vortex population at the edge was found to be very different compared with that in the central plane.
Article
Various clinical observations reported that swirling blood flow is a normal physiological flow pattern in various vasculatures. The swirling flow has beneficial effects on blood circulation through the blood vessels. It enhances oxygen transfer and reduces low-density lipoprotein concentration in the blood vessel by enhancing cross-plane mixing of the blood. However, the fluid-dynamic roles of the swirling flow are not yet fully understood. In this study, inhibition of material deposition at the post-stenosis region by the swirling flow was observed. To reveal the underlying fluid-dynamic characteristics, pathline flow visualization and time-resolved particle image velocimetry measurements were conducted. Results showed that the swirling inlet flow increased the development of vortices at near wall region of the post-stenosis, which can suppress further development of stenosis by enhancing transport and mixing of the blood flow. The fluid-dynamic characteristics obtained in this study would be useful for improving hemodynamic characteristics of vascular grafts and stents in which the stenosis frequently occurred. Moreover, the time-resolved particle image velocimetry measurement technique and vortex identification method employed in this study would be useful for investigating the fluid-dynamic effects of the swirling flow on various vascular environments. © IMechE 2015.
Article
The effects of swirling flow on the flow field in 45° end-to-side anastomosis are experimentally investigated using a particle image velocimetry technique to reveal fluid dynamic advantages of swirling flow in the vascular graft. Non-swirling Poiseuille inlet flow unnecessarily induces pathological hemodynamic features, such as high wall shear stress (WSS) at the ‘bed’ side and large flow separation at the ‘toe’ side. The introduction of swirling flow is found to equalize the asymmetric WSS distribution and reduces the peak magnitude of WSS. In particular, the intermediate swirling intensity of S = 0.45 induces the most uniform axial velocity and WSS distributions compared with weaker or stronger swirling flows, which addresses the importance of proper selection of swirling intensity in the vascular graft to obtain optimum flow fields at the host vessel. In addition, swirling flow reduces the size of flow separation because it disturbs the formation of Dean-type vortices in secondary flow and inhibits secondary flow collision. The beneficial fluid dynamic features of swirling flow obtained in this study are helpful for designing better vascular graft suppressing pathological hemodynamic features in the recipient host vessel.
Article
Full-text available
Time-resolved particle-image velocimetry measurements are made in the streamwise-wall-normal plane of turbulent channel flow at Re τ = 550 and 1747. Temporal and convective derivatives of velocity are computed from this data in order to evaluate the small-scale behaviour of these quantities as well as of the velocity itself. Instantaneous velocity fields indicate that the flow is dominated by small-scale vortex cores believed to be associated with hairpin/hairpin-like vortices. These vortices have been observed in realizations of the random velocity in other wall turbulence studies. In this work, a deterministic ‘vortex signature’ is determined by conditional averaging techniques. This average signature is consistent with the hairpin vortex signature defined by Adrian and co-workers: circular streamlines with a strong ejection of low-speed fluid away from the wall (a Q 2 event) just upstream of the vortex head. In addition, the spatial extent of these small-scale vortices appears to remain relatively constant within the Reynolds-number range studied herein.
Data
Full-text available
The effects of a co-flow on a spreading and en-trainment rate of turbulent round jets have been studied numerically. The first and second order closure models are used and have been comp-ared with existing experimental data. The influ-ence of theses models on the dynamic fields is examined. The results of the models in general agree well with the trends observed experiment-tally. The co-flowing imposed noticeable restri-ctions on the spreading and the turbulent mix-ing. Finally, an entrainment hypothesis has been introduced to describe the development of tur-bulent jets issuing into a stagnant or co-flowing air. It relates the mass flow rate of the surround-ing fluid entrained into the jet to the character-istic velocity difference between the jet and the co-flow. It is obvious that the co-flow decreases considerably the entrainment of air.
Article
Full-text available
High-vector-densityplanar velocity fields were obtained for a weakly compressible mixing layer using particle image velocimetry (PIV). The velocity ratio of the mixing layer was 0.53, the density ratio was 0.67, and the convective Mach number was 0.38. At the location where the PIV images were obtained, Rex=3.7×10^6 and Reδω=1.8×10^5. The instantaneous planar velocity fields fall into three regimes characterized by the size and number of large-scale structures present. The large-scale rollers are either circular or elliptical, with the elliptical rollers having, in general,horizontal major axes. The transverse velocity fluctuations and Reynolds shear stress are suppressed for the weakly compressible mixing-layer as compared tothe incompressible case. The spatial correlations of velocity fluctuations also occupy a smaller fraction of the mixing-layer thickness than for an incompressible mixing layer. The linear stochastic estimate of a roller structure is elliptical with the major axis oriented in the streamwise direction and with an eccentricity greater than for the incompressible case. The linear stochastic estimate of a braid suggests that the braids are vertically oriented, as opposed to the oblique orientation seen in incompressible mixing layers. In addition, the braids in the weakly compressible case have a vertically oriented stagnation line, as opposed to the braids in the incompressible mixing layer where stagnation occurs at a point.
Article
Full-text available
This paper reports a systematic investigation of the effect of nozzle aspect ratio (AR) on plane jets. The aspect ratio AR (≡ w/h, where h and w are the nozzle height and width) was varied from 15 to 72. The present velocity measurements were performed using single hot-wire anemometry, over a downstream distance of up to 85h and at a nozzle-height-based Reynolds number of Reh = 1.80 × 104. Results obtained reveal that both the extent and character of statistical two-dimensionality of a plane jet depend significantly on AR. Most aspects of the near field flow exhibit an asymptotic-like dependence on AR, but do not become independent of AR within the range of AR investigated. A region of statistically two-dimensional (2-D) mean velocity field is achieved only for AR ⩾ 20, and its axial extent increases with AR. However, the centerline turbulence intensity in the far field displays an asymptotic-like convergence only for AR ⩾ 30. In the self-similar region, both the mean decay and spreading rates of the jet increase as AR increases and do not reach an asymptotic value, even at AR = 72. The aspect ratio of the local jet (w/local velocity half-width) at the end of the 2-D region becomes asymptotically independent of nozzle aspect ratio, for approximately AR ⩾ 30. That is, the plane jet ceases to be statistically 2-D at a fixed value of local jet aspect ratio for nozzle aspect ratios greater than 30. The skewness and flatness factors also depend on AR. These results imply that independence of AR, even in the near field, will require very much larger aspect ratios than have been investigated previously.
Article
Full-text available
Experimental results for the conditional statistics, such as the velocity conditioned on a conserved scalar and the scalar conditioned on velocity, in a confined liquid-phase rectangular jet are presented and analyzed for a data set collected using simultaneous particle image velocimetry and planar laser-induced fluorescence. The joint velocity-scalar probability density function (PDF) is not joint Gaussian in this flow, as the PDF of the conserved scalar is accurately described by a beta-PDF. The conditional mean velocity is found to agree with a linear model when the scalar is close to its local mean value. A gradient PDF model is found to give poor predictions for the streamwise conditional velocity. However, the improved gradient PDF model predicts both the streamwise and transverse conditional velocities well. A linear model for the scalar fluctuation conditioned on velocity is also tested against the experimental data, showing that this model only obtains good approximations when the joint velocity-scalar PDF approaches a joint Gaussian at farther downstream locations in this flow.
Article
Preface; Nomenclature; Part I. Fundamentals: 1. Introduction; 2. The equations of fluid motion; 3. Statistical description of turbulence; 4. Mean flow equations; 5. Free shear flows; 6. The scales of turbulent motion; 7. Wall flows; Part II. Modelling and Simulation: 8. Modelling and simulation; 9. Direct numerical simulation; 10. Turbulent viscosity models; 11. Reynolds-stress and related models; 12. PDF models; 13. Large-eddy simulation; Part III. Appendices; Bibliography.
Article
An experimental investigation of large-scale structural patterns in the similarity region of a two-dimensional turbulent jet was performed. Two-point correlation and coherence-based measurements obtained from both longitudinal and lateral component velocity fluctuations are, in general, supportive of the existence of an antisymmetric array of counterrotating vortices in the similarity region. Particularly graphic evidence is obtained from detailed isocorrelation contour maps that are presented. A novel method of determining structural convective velocity is developed and indicates the structural array propagates at 60&percnt; of the local centerline mean velocity. Very limited structural two dimensionality is exhibited in the direction of mean flow homogeneity and it is suggested that this is likely to be related to the existence of multiple coherent modes in the flow.
Article
Conditional averages of the velocity field, subgrid-scale (SGS) stresses and SGS dissipation are calculated using the velocity fields obtained from the DNS of plane channel flow. The detection criteria isolate the coherent turbulent structures that contribute most strongly to the energy transfer between the large, resolved scales and the subgrid, unresolved, ones. Separate averages are computed for forward and backward scatter. The interscale energy transfer is found to be strongly correlated with the presence of the turbulent structures typical of wall-bounded flows: quasi-streamwise and hairpin vortices, sweeps and ejections. In the buffer layer, strong SGS dissipation is observed near lifted shear layers; the forward scatter is associated with ejections, the backscatter with sweeps. Both backward and forward scatter occur in close proximity to longitudinal vortices that form a very shallow angle to the wall. Further away from the solid boundary, in the logarithmic region and beyond, both forward and backward energy transfer are associated prevalently with ejections. Eddy viscosity models do not predict the three-dimensional structure of these events adequately, while scale-similar models reproduce the correlation between the large-scale coherent structures and the SGS events more accurately.
Article
High-vector density planar velocity fields were obtained for an incompressible mixing layer using particle image velocimetry (PIV) for the purpose of determining spatial correlations of velocity fluctuations and linear stochastic estimates of the large-scale structures. The linear stochastic estimates were calculated based on the deformation tensor. The velocity ratio of the mixing layer was 0.575, and the density ratio was unity. At the location where the PIV images were obtained, Re-x = 1.8 x 10(5), Re-deltaomega = 1.1 X 10(4), and the pairing parameter was Rx/lambda = 8. Preliminary hot-film measurements showed the mixing-layer mean velocity and turbulence profiles to be self-similar at this location. The mixing layer was found to be largely two-dimensional with well-organized Brown-Roshko roller structures and braids (Brown, G. L., and Roshko, A., "On Density Effects and Large Structures in Turbulent Mixing Layers," Journal of Fluid Mechanics, Vol. 64, 1974, pp. 775-814). Measured velocity fluctuations and Reynolds shear stress data agree well with previous experimental results. The R-u'u' correlation is a horizontally oriented ellipse, with a slight inclination from horizontal. The R-v'v' correlation is a vertically oriented ellipse. The linear stochastic estimate of a roller suggests that these structures are very slightly elliptical, with a horizontal major axis. The linear stochastic estimate of a braid suggests that the braids are obliquely oriented with stagnation occurring at a point.