You are on page 1of 17

American Institute of Aeronautics and Astronautics

1
Numerical Simulation and Experiments of Jets in Cross
Flow
Z. Li., S. Murugappan, E. Gutmark, L.Vallet
Department of Aerospace Engineering and Engineering Mechanics
University of Cincinnati, Cincinnati, Ohio, 45221-0070
A numerical simulation was carried out to compute the penetration, mixing and
turbulence structures of a jet injected perpendicular into a free stream through different
circular nozzles. Six cases were studied, including 3 different jet diameters and 2 different
blowing ratios. The formation of the Counter rotating Vortex Pair (CVP) and the interaction
between the free stream and jet flow is discussed. The d, rd and r
2
d scaling parameters
associated with the centerline jet trajectory shows bifurcation into two separate branches.
The bifurcation was related to the different Reynolds numbers. Flow features related to
mixing, centerline velocity decay, size and shape of the recirculation bubble formed behind
the jet are discussed. Higher blowing ratios show higher velocity decay rate. Both time
averaged flow field and turbulence are compared with 2D Particle Image Velocimetry (PIV)
data taken along the jet center plane. Good match between the experiments and
computation was observed.
I. Nomenclature

d = hydrautic diameter of nozzle
I = turbulence intensity
k = turbulence kinetic energy
r = square of momentum flux ratio
Re
y
= Reynolds number based on the distance y to the wall
Re
DH
= Reynolds number based on the hydraulic diameter
j
u = jet velocity
'
u = fluctuating velocity
avg
u = mean velocity

u = free stream velocity


+
y = dimensionless sublayer-scaled distance
x, y, z = streamwise, transverse and spanwise coordinate directions
= density of fluid
= dissipation per unit mass
= molecular viscosity
t
= eddy viscosity


II. Introduction

ETS in cross flow (JICF) belong to classical flows that are found in many engineering applications, including:
smoke issuing from chimneys, pollutant dispersal, turbine blade cooling, V/STOL aircrafts, fuel injection in
supersonic flows, dilution zones in gas turbine combustors and reaction control jets in rockets and missiles.
J
44th AIAA Aerospace Sciences Meeting and Exhibit
9 - 12 January 2006, Reno, Nevada
AIAA 2006-307
Copyright 2006 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

American Institute of Aeronautics and Astronautics

2
Overall, there are four important vortical structures which have been well established in the studies of jets in
cross flow. These are: (a) Horse shoe vortices
1-3
which form upstream of the jet exit and wrap around the jet
column, (b) Jet shear layer vortices
1
which form at the interface between the jet and the cross flow, (c) Wake
vortices
3
which form on the lee-side of the jet and persist far downstream, (d) Counter-rotating vortex pair
4
which
forms after the jet has been turned by the cross stream.
The horseshoe vortex system is a result of the interaction between the wall boundary layer and the transverse jet
1-3
.The adverse pressure gradient formed at the injection wall forces the wall boundary layer to separate and form
the horseshoe vortex. This vortex system is then convected and stretched around the jet periphery like a necklace.
This is analogous to the vortex system formed when an approaching boundary layer interacts with a cylinder
mounted on a wall
4
. The horseshoe vortex system exhibits oscillating modes which correlate with the periodic
motions of the upright wake vortices
1-3
.
The upright wake vortices have been identified by Fric and Roshko
1
by means of smoke wire visualization. The
boundary layer of the cross flow has been observed to provide the main source of vorticity in the wake vortices.
Fric and Roshko
1
have also identified separate events where the wall boundary layer forms vortices which attach
themselves to the lee-side of the jet and eventually form the wake vortex system. This finding is called an upright
wake vortex system since one end of the vortex string is connected to the jet and follows the jet trajectory whereas
the other end stays close to the cross flow wall, positioning the vortex in an upright orientation.
The counter-rotating vortex (CVP) pair has been observed to be the dominating structure in the far field region
of the jet cross section
5
with evidence of its initiation in the near field
6,7
. Knowledge of the origin and growth of the
CVP are critical to the control of vorticity generation and evolution which are primary factors in the mixing of a
transverse jet with cross flow. It is generally accepted that the fold and rollup of the jet shear layer near the jet exit
contributes to the formation of the CVP
3
. Also, tilting and folding of the vortical structures contributes to the
vorticity in the far field which causes, on a time average basis, the formation of the counter rotating vortical
structures.
Margason
8
provided an extensive review of past work before 1993 on jet in cross flow. In many of the studies,
the main interests are the trajectories prediction, the formation, evolution and interaction of the counter-rotating-
vortex-pair (CVP) and their respective applications. Both experimental and computational efforts were conducted to
investigate the details of different flow structures of JICF. The flow field of a vertical jet in cross flow is observed
to be primarily influenced by the square root of fluid momentum ratio
2
1
2
2

=
cf cf
j j
u
u
r

(1.1)
or a simplified effective velocity ratio
cf j
u u r = for incompressible flows.

Different range of velocity ratio determines different flow regime. Typically, for r <1, jet flow is usually so weak
that it cannot break through the flat plate boundary layer and acts as a flow obstacle. Jet in cross flow within 1< r
<10 constitutes the most common flow regimes in engineering applications, while jet at r >10 behaves more like a
free jet in static flow.

Pratte and Baines
9
first employed a length scale analysis, to conclude a general form to collapse the trajectories:
B
rd
x
A
rd
y

= (1.2)
in which a global length scale in the flow rd is adopted in the region away from the jet exit. They also obtained
A=2.05 and B=0.28 from their experimental data. Other researchers found the trajectories collapsed with different
A, B values and proposed a different length scale: d r
2
. Some other researchers also deduced different power laws
formula using numerical methods, similarity analysis and asymptotic theory.

Much of the computational effort in circular jets has been concerned with predicting the flow field characteristics
such as penetration, spreading and mixing of a passive scalar. Most computations are able to predict the gross flow
field using relatively simple turbulence models, such as k-e, as shown by Patankar et al
10
. While the gross field was
predicted using the simplest turbulence models, the details of the CVP are much more complex to discern. Reynolds
stress models have been used by Crabb et al
11
. Their results compared well in reproducing the peak vorticity and

American Institute of Aeronautics and Astronautics

3
CVP strength. Yuan and Street
12
, Yuan et al
13
were among the first to use LES simulations to compute the flow
field. Among their results is a proposed mechanism for the formation of the CVP by means of breakdown of quasi-
steady vortices that extend upwards and downstream from the lateral edges of the jet. Muppidi and Mahesh
14
did
DNS Computation of the JICF. They identified a new scaling law which depends on the inlet jet velocity profile and
the cross flow boundary layer. They proposed an analytical expression for the new length scale which is a measure
of the relative inertia of the jet and the cross flow. They observed that incorporating this length scale improves the
scaling the trajectories.
The focus of the current paper is compare simulation and experiments
15
for six different cases. The effects of
blowing ratio r on the jet trajectory scaling are studied. The flow field and mixing characteristics of the jet with the
free stream is discussed. Different parameters such as the Turbulent Kinetic Energy and centerline velocity decay
are evaluated for the six cases. The size, shape of the recirculation bubble and possible scaling with r and d is
investigated.
II. Computation Set Up

Figure 1a shows a schematic of the JICF solved numerically using FLUENT 6.0.12. The jet begins as a turbulent
pipe flow which then issues into a flat plate boundary layer through an orifice that is flushed with the plane of the
flat plate. The coordinate system used in the current simulation is depicted in figure 1. The origin x=y=z=0 is
located at center of the circular orifice. Here x, y and z represent the streamwise, vertical and spanwise directions,
respectively. Six cases are studied in the current paper. Table 1 lists the jet diameter, D is the jet diameter, bulk jet
velocity, V, free stream jet velocity, blowing ratio (r ) and the Reynolds number based on jet diameter

VD
D
= Re .


Figure 1a Flow configuration schematic with 3D coordinate setting used in the simulation


Diameter Velocity of the jet
velocity of
freestream
Blowing
ratio, r,
ReD
Case1 10m/s 8.45
Case2
Nozzle1 10.922mm 84.46m/s
17m/s 4.97
63165
Case3 13.6m/s 8.38
Case4
Nozzle2 9.398mm 114.89m/s
23m/s 4.96
73917
Case5 22m/s 8.55
Case6
Nozzle3 4.648mm 188.09m/s
35m/s 5.37
59849

Table 1. Computational Parameters



The computational mesh is mainly unstructured and consists of hexahedral elements. The use of unstructured mesh
near the jet exit and wake region provides easier handling of variation in mesh size and relatively higher
computational efficiency, provided by the finite volume method in FLUENT. The computation domain for case1 is
showed in Figure1b. It is a combination of square channel and a long cylinder pipe. Dimensions of the simulation
domain vary depending on the velocity ratio, r. For higher r simulations, size in y-direction is set larger as the jet
has a deeper penetration into the cross-flow. For lower r cases with less jet penetration vertical extent of the domain

American Institute of Aeronautics and Astronautics

4
is reduced to minimize the memory requirement and CPU time. A total of 0.9~2
6
10 control volumes are used to
discretize the domain, of which over 90% are placed within the near field at the jet exit. Coarse elements are mostly
located far away upstream and downstream from the jet exit in the free stream. The near wall regions are treated
with refined boundary layer grids as they approach the solid surface. The viscous sub-layer is not captured in the
current simulations due to the time constraints. The pipe cross-section is hybrid meshed with unstructured grid at
center surrounded by structured fine grid in order to have a high grid resolution in the region of strong shear stress.
In core region of channel, unstructured girds are of approximately the same edge length and the mesh size increases
gradually in the radial direction. The block volume mesh is created by sweeping the mesh in y-direction from flat
plate surface and pipe inlet face. The mesh size varies from 0.0006d to 1.5d over the core computational domain.


Figure 1b: Computational domain of the channel and pipe


A segregated, 3D, implicit, uncompressible and steady solver is employed for all these simulations. A standard
k model is chosen for the initial iterations, and a renormalization group (RNG) k model is selected for the
successive iterations. In a standard k model, the turbulence kinetic energy, k, and its rate of dissipation, , are
obtained from the following transport equations:
K M b k
j k
t
j
i
i
S Y G G
x
k
x
ku
x
k
t
+ + +

] ) [( ) ( ) ( (2.1)
and

S
k
C G C G
k
C
x x
u
x t
b k
j
t
j
i
i
+ + +

2
2 3 1
) ( ] ) [( ) ( ) ( (2.2)
G
k
represents the generation of turbulence kinetic energy due to the mean velocity gradients. G
b
is the
generation of turbulence kinetic energy due to buoyancy. Y
M
represents the contribution of the fluctuating dilatation
in compressible turbulence to the overall dissipation rate and can be neglected here.
1
C ,
2
C , and
3
C are
constants.
k
and

are the turbulent Prandtl numbers for k and , respectively. S


k
and

S are source terms. In


RNG k model, all
k
t

+ term are changed as


eff k
. And equation (2.2) is added an additional term

R given by:
k
S
S S
R
k
k k
] ) ( 012 . 0 1 [
)
38 . 4
1 ( ) ( 0845 . 0
3
2 3

= .
This term produces negative contribution in regions of large strain rate, thus yielding a lower turbulent viscosity
than the standard k model.

American Institute of Aeronautics and Astronautics

5
X
0
0.05
0.1
0.15
0.2
0.25
Y
-0.03
-0.02
-0.01
0
0.01
0.02
Z
0
0.05
0.1
1
2
3
4
5
6
7
8
9
10
X Y
Z
The boundary conditions are set as follows. A uniform inlet velocity profile is specified at the entrance to the flat
plate. A turbulent boundary layer represented by a 1/7 power law approximation was imposed on the flat wall. It
has been observed in the LES of Yuan et al
13
and DNS of Muppidi and Mahesh
14
that the flow inside the pipe need
to be modeled in order to capture asymmetry in the jet profile at the exit and possible reverse flow inside the pipe. In
the current simulations the flow inside the pipe was modeled with a uniform inflow velocity profile. Length of the
pipe is designed to be sufficiently long for the jet fluid to fully develop and velocity profiles were checked with
experiments for validation after the computation. At the span-wise boundary and the top boundary, a symmetry
condition is set. On the downstream outflow face, a zero-gradient pressure outlet was applied. The turbulence
intensity at the entrance to the channel was set to 0.04%. No heat transfer in considered in simulation.
















Figure 2 Time-averaged contours of Y velocity and some characteristic streamlines on symmetry center plane
for case 3














(a) (b)
Figure 3. (a), Streamlines originating from the jet in the center plane (Z=0) for case 3. (b) Evolution of the
Streamlines originating from the jet flow.

III. Results

3.1Flow feature

Figure 2 shows the contours of velocity magnitude and some characteristic streamlines for case 3 on the Z = 0
symmetry center plane. Upstream of the jet exit the cross flow fluid close to the wall appears to stagnate upon
encountering the jet, streamlines upstream of the jet and above the injection wall bend around the jet. Streamlines
downstream show strong entrainment of the cross flow into the jet stream. Downstream from the jet exit and above
the injection wall exists a node with positive divergence (x=0.11rd, y=0.05rd). The streamlines from this node
indicate that the part of cross stream is entrained into the jet whereas the streamlines close to the wall follows the

American Institute of Aeronautics and Astronautics

6
streamwise direction. The existence of nodes has been observed in the past by Kelso et al
3
, Hasselbrink and
Mungal
16
and Muppidi and Mahesh
14
.
















(a) (b)












(c)




Figure 4. (a) Schematic showing the location, where the free stream fluid marker was placed along the
centerline Y=0 upstream of the jet; (b) The jet fluid (Red) and free stream fluid (Blue and Green)
represented in different colors; (c) Cross section at 4 different planes

Figure 3 a and b show the streamlines originating from the jet and their evolution at 10 different cross sectional
planes. The ten planes were sliced normal to the jet trajectory coordinate, s. The formation and growth of the counter
rotating vortex pair is clearly evident in these slices. In order to better understand the interaction of the jet flow with
the cross stream, fluid originating from the jet and cross stream were tracked. Two different regions upstream of the
jet with cross stream flow were marked with tracers. One region was closer to the wall and the other was right above
the first region. Both these regions were 3.2d upstream of the jet and had an equal width z=2.35d (refer to figure
4a). The height of the lower and upper region was 0.62d and 0.53d. In Figure 4b, the jet fluid is represented in red,
while free stream flowing through the upper region is denoted by blue, and free stream through the lower region is in
green. The shape of fluid flow is constituted by numerous streamlines that is uniformly distributed in the two
regions. The cross stream fluid originating from both the regions wraps around the jet fluid as they both come in
contact. x-z cross sections (not shown here) indicate that the blue region marked by the free stream fluid does not
penetrate the jet fluid until z=0.09. The free stream region marked in blue is convected along with the jet fluid and
follows the jet trajectory. The formation of the CVP provides a higher contact area between the free stream region
marked in blue and the jet fluid. This enhances mixing between the jet and cross steam. The free stream fluid
originating from the region closer to the wall has a different behavior. Interestingly, the free stream fluid marked in
green flows around the jet, part of it is convected up into the jet and the other part diverges like a fan following the
3.2d

American Institute of Aeronautics and Astronautics

7
0
2
4
6
8
10
12
14
16
0 1 2 3 4 5 6
X/d
Y
/
d
Case1 (r=8.45,
d=10.92mm, Re=63000)
Case2 (r=4.97,
d=10.92mm, Re=63000)
Case3 (r=8.38,
d=9.398mm, Re=74000)
Case4 (r=4.96,
d=9.398mm, Re=74000)
Case5 (r=8.55,
d=4.648mm, Re=60000,
Fr=00)
Case6 (r=5.37,
d=4.648mm, Re=60000)
Yuan & Street
(d=13.84mm, r=2,
Re=2100, Fr=00)
Yuan & Street
(d=13.84mm, r=2,
Re=2100, Fr=10)
Yuan & Street
(d=13.84mm, r=3.3
Re=1050, Fr=00)
Yuan & Street
(d=13.84mm, r=3.3,
Re=2100, Fr=00)
Su & Mugal (r=5.7,
d=4.53mm, mean
turbulent, Re=5000)
streamwise direction behind the jet wake. Figure 4c shows the interaction of the jet and free stream at four constant-
x slices. The observations made in the figure 4b are supported by the y-z planes. The free stream region marked by
green shows one mechanism of mixing encountered in JICF where the free stream flow in the jet wake is convected
up into the jet fluid through wake-upright vortices. The other mixing mechanism involves the free stream fluid
which is carried along by the jet stream and follows the jet trajectory. As the CVP grow in size there is higher
contact area between the jet and the free stream that enables molecular mixing between the two streams.


3.2 Trajectories

Different definitions have been used to track the jet trajectory in experiments. Depending on the flow diagnostic
technique either the local velocity maxima (Kamotani and Greber
17
) or the local scalar concentration maxima (Smith
and Mungal
6
). The latter definition is usually showed to have about 5~10% deeper penetration into the flow than
the former one (Haniu and Ramaprian
18
). Other definitions such as streamline originating from the center of the jet
exit on the symmetry plane have been used by Yuan and Street
12
and Muppidi and Mahesh
14
. They justify using this
definition, since both the maximum velocity and scalar concentration has multiple maxima at the jet exit. In the
current work, the jet trajectory is defined as locus of maximum velocity in the center plane. In Figure 5 a-c,
trajectories of all 6 cases are plotted with three different normalizations: d, rd and r
2
d. Trajectories from experiments
and simulations of other researchers are also compared. Figure 5a shows the jet trajectory normalized by d. The d-
scaled plot shows that the jet penetrates deeper into the cross stream for cases 1, 3 and 5 where r ~8.5 when
compared to case 2, 4 and 6 with r~5. The data from Yuan and street
13
simulations at r=2 and 3.3, Su and Mungal
19

at r=5.7 and Pratte and Baines
9
experiments are included. Two issues become clear from the d scaled plots: 1) the
data does not collapse for the different r and higher r is associated with larger jet penetration into the cross stream.
Figure 5b shows the jet trajectory normalized by rd. The data curves seem to bifurcate. The difference between the
bifurcating branches arises from Re in both the present data and other previously published JICF trajectories. The
higher Reynolds numbers (Re>2500) show a larger jet penetration and follow on the top branch, while the lower set
of curves relate to Re<2500. Included in the rd scaling is the empirical relation (eq 1.2) proposed by Smith and
Mungal
6
. They observed that coefficient A in equation 1.2 was found to vary between 1.4 and 1.8. The lower set of
curves seems to collapse with an empirical constant of A=1.4, while the high Re cases fall on the curve that has the
empirical constant A=1.8. The r
2
d scaling of the jet trajectories is presented in figure 5c. The data still shows
bifurcation similar to the d scaling where the higher r fall on the lower set of curves and smaller r fall on the higher
set of curves. In summary, from the limited data plotted in figure 5, r and d alone could not be used to scale the data.
The effect of Reynolds number could be an additional factor that should be considered for scaling the data.









(a)














American Institute of Aeronautics and Astronautics

8
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0 0.5 1 1.5 2 2.5 3
x/rd
y
/
r
d
Su & Mungal PIV (r=5.7,
d=4.53mm,mean turbulent, Re=5000)
Case1 (r=8.45, d=10.92mm,
Re=63000)
Case3 (r=8.38, d=9.398mm,
Re=74000)
Case2 (r=4.97, d=10.92mm,
Re=63000)
Case5 (r=8.55, d=4.648mm,
Re=60000)
Case4 (r=8.38, d=9.398mm,
Re=74000)
Case6 (r=5.37, d=4.648mm,
Re=60000)
Roth et al (r=6)
Roth et al (r=4)
Smith & Mungal (scaled
coefficient=1.4)
Smith & Mungal (scaled
coefficient=1.8)
Yuan & Street (d=13.84mm, r=2,
Re=2100, Fr=00)
Yuan & Street (d=13.84mm, r=2,
Re=2100, Fr=10)
Yuan & Street (d=13.84mm, r=3.3,
Re=1050, Fr=00)
Yuan & Street (d=13.84mm, r=3.3,
Re=2100, Fr=00)
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
0 0.1 0.2 0.3
x/r
2
d
y
/
r
2
d
Su & Mungal PIV (r=5.7,
d=4.53mm, mean turbulent,
Re=5000)
Case1 (r=8.45, d=10.92mm,
Re=63000)
Case2 (r=4.97, d=10.92mm,
Re=63000)
Case3 (r=8.38, d=9.398mm,
Re=74000)
Case4 (r=4.96, d=9.398mm,
Re=74000)
Case5 (r=8.55, d=4.648mm,
Re=60000)
Case6 (r=5.37, d=4.648mm,
Re=60000)
Yuan & Street (r=2,
d=13.84mm, Re=2100,
Fr=00)
Yuan & Street (r=2,
d=13.84mm, Re=2100,
Fr=10)
Yuan & Street (r=3.3,
d=13.84mm, Re=1050,
Fr=00)
Yuan & Street (r=3.3,
d=13.84mm, Re=2100,
Fr=00)










(b)
















(c)










Figure 5: Jet Trajectories with different scaling (a) d, (b) rd and (c) r
2
d.


3.3 Comparison between Experimental and Computational Results

2D time averaged flow field computed from PIV measurements on the Z=0 center plane are compared with the
computational results for all six cases. The data shows a good match in predicting the entire flow field. Details for
Case 3 alone are shown in this manuscript. Side views on symmetrical center plane of the velocity magnitude
contours and streamlines are shown in figure 6 a and b. The simulations predict the entrainment of the free stream
into the jet flow, the existence of nodes and the stagnation region on the leeward and windward side close to the
injection wall. The potential core in the simulation was found to be 10% smaller in experiments when compared to
the simulations. This indicates that the jet centerline decays faster with experiments than simulations.


American Institute of Aeronautics and Astronautics

9
-0.2
0
0.2
0.4
0.6
0.8
1
-0.12 -0.02 0.08 0.18 0.28 0.38
X/rd
V
/
V
m
a
x
computation
experiment
- 0. 2
0
0. 2
0. 4
0. 6
0. 8
1
1. 2
-0.12 -0.02 0.08 0.18 0.28 0.38
X/rd
V
/
V
m
a
x
computation
experiment

(a) (b)
Figure. 6 Side view of the velocity magnitude contour with streamlines on symmetrical center plane. a)
experimental; b) Numerical
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
-0.2 -0.1 -0.1 0.0 0.1 0.1 0.2 0.2
X/rd
V
/
V
m
a
x
experiment
computation

(a)
















(b) (c)

American Institute of Aeronautics and Astronautics

10
-0.1
0.1
0.3
0.5
0.7
0.9
-0.12 -0.02 0.08 0.18 0.28 0.38
X/rd
V
/
V
m
a
x
experiment
computation
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
-0.12 -0.02 0.08 0.18 0.28 0.38
X/rd
V
/
V
m
a
x
computation
experiment
0
0.5
1
1.5
2
2.5
3
0 0.5 1 1.5 2 2.5 3
X/rd
Y
/
r
d
Case1 experiment
Case2 experiment
Case3 experiment
Case1 simulation
Case2 simulation
Case3 simulation













(d) (e)
Figure. 7: The Y velocity profile at jet exit vicinity region: a) Y=0, b) Y=0.13rd, b) Y=0.25rd, c) Y=0.38rd, d)
Y=0.51rd
















Figure. 8: Comparison of Jet Trajectory from Experiments and Simulations

In order to validate the simulations, the flow field near the jet exit region is compared quantitatively between the
computations and the experiments. Figure 7 shows a series of y velocities profiles at y=0, 0.13, 0.25, 0.38 and
0.51rd. The y velocity profiles are normalized by rd. The results show a good agreement with each other in the near
field. Further downstream, the peak velocity shifts to the right indicating that the jet begins to bend. Also evident in
these figures is the broadening on the leeward side of the jet along the vertical direction, since the jet begins to mix
with the cross stream. The jet trajectory is also compared for cases1-3. The simulations agree fairly well with the
experiments. Figure 9 shows the windward and leeward boundaries from the simulations. The boundaries of the jet
are identified as 40% of the maximum jet velocity. Case 1 and 3 shows very similar jet upper and lower boundary
trajectories, possibly due to the small variation between actual r and d between these cases. The trajectory in case 2
does not match with case 1 and 3. In all cases, the jet width becomes narrow until the end of the potential core
(y<0.6rd). Past 0.6rd in the y direction, the jet boundary begin to spreads.

4.4 Entrainment and mixing

Figure 11 a depicts the centerline velocity magnitude decay along the jet trajectory. The velocity is normalized by
the velocity magnitude at the exit. They are compared with the mean centerline velocity decay of a circular free jet
and an elliptic free jet with similar Reynolds number. For the circular free jet, the velocity decay is inversely
proportional to the streamwise distance downstream from the potential core. The decay equation is mainly
controlled by velocity profile at the exit and the virtual jet origin distance (Malmstrom, 1997). The elliptic jet had a
more rapid decay when compared to the circular jet. Velocity magnitude of the jets in cross flow begins to drop
immediately after leaving the nozzle (S/D=0.5~1). In contrast, a circular free jet has a potential core ending at about
S/D=5, and an elliptic jet at about S/D=5, at similar Reynolds numbers. Figure 11b shows the centerline velocity

American Institute of Aeronautics and Astronautics

11
magnitude, Vmmax/Vm, as a function of the normalized distance. Two different regimes are noticed; the higher r
cases (case 1, 3 and 5) collapse on to one set and lower r (cases 2, 4 and 6) collapse onto the lower set. The slopes
were computed for both the curves in the regions (1.2<s/rd<2.2). The lower set of curves had a slope =1.4 and
higher set of curves show a slope =2.5. The larger the slope, the higher is the velocity decay. Clearly, the larger r
cases show a 44% increase when compared to the lower r. In a review by Holdeman (1993), one of the important
observations he made was that smaller momentum flux ratios required a greater downstream distance for equivalent
mixing. Since, centerline decay is one of the parameter that could be used to evaluate large scale mixing, it
supports the observation that larger r have higher decay rate at the same distance downstream when compared to
lower r. Farther downstream (s>3.25rd) , all the plots tend to asymptote to the free stream velocity.


0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
-0.2 0 0.2 0.4 0.6 0.8 1
X/rd
Y
/
r
d
Case1 leeward and rearward boundary
Case2 leeward and rearward boundary
Case3 leeward and rearward boundary

Figure. 9: Comparison of Jet Trajectory from Experiments and Simulations




0
1
2
3
4
5
6
7
8
9
10
0 1 2 3 4 5 6 7 8
S/rd
V
m
m
a
x
/
V
m
Case1 r=8.45, d=10.92mm, Re=63000
Case2, r=4.97, d=10.92mm, Re=63000
Case3, r=8.38, d=9.398mm, Re=74000
Case4, r=4.96, d=9.398mm, Re=74000
Case5, r=8.55, d=4.648mm, Re=60000
Case6, r=5.37, d=4.648mm, Re=60000


(a) (b)

Figure. 11 Velocity decay along trajectory under different normalization: (a)d; (b)rd slopes of case1, 3, 5 are
about 2.5 and slopes of case 2, 4, 6 are about 1.4 for the specified range.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 5 10 15 20 25 30
S/D
V
m
/
V
m
m
a
x
Case1, r=8.45, d=10.92mm, Re=63000
Case2, r=4.97, d=10.92mm, Re=63000
Case3, r=8.38, d=9.398mm, Re=74000
Case4, r=4.96, d=9.398mm, Re=74000
Case5, r=8.55, d=4.648mm, Re=60000
Case6, r=5.37, d=4.648mm, Re=60000
Elliptic free jet (Ho and Gutmark[20], Re=78000)
Circular free jet (Lee et al.[21], Re=50000)

American Institute of Aeronautics and Astronautics

12


Figure 12 displays the width of the jet boundary for case 3. The boundary was defined as 40% of the maximum
velocity on the jet trajectory, s, on the center plane. The width of the boundary is measured on the perpendicular to
the jet trajectory. The variation of the width of the boundary displays a converging-diverging tendency. The
boundary was found to converge until the end of the potential core; past the potential core the jet tends to spread.
The plot of turbulent kinetic energy (TKE) contour on the center plane is showed in Figure 14 for cases 2 and 3.
The TKE is scaled by the square of the jet exit velocity. The turbulence is most strong at the windward jet shear
layer at the nozzle exit in both the cases. The high turbulent intensity (TKE>0.025) apparently follows the
streamlines that pass at the vicinity of the jet exit edge. The high TKE produced on the windward shear layer arises
from the shear layer eddies created by the Kelvin-Helmholtz instability, and the impingement of the cross flow
stream on the jet flow at the windward side. A plot of the turbulent kinetic energy along the jet trajectory is shown in
Figure14 for all 6 cases. The TKE is normalized by the square of the jet velocity and the distance, s, is normalized
by rd. The TKE reaches its peak at about 1 rd, and it rapidly decreases to very low magnitude after a distance of 2
rd. This is mainly because the trajectory (locus of maximum velocity) crosses over the highest TKE about Z=1rd
and then continues to remain be above the higher TKE trajectory.

0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 0.5 1 1.5 2
S/rd
W
i
d
t
h
/
r
dCase1
Case2
Case3

Figure 12. The width between leeward and rearward boundary variation along the trajectory, normalized by
rd for case3.



















(a) (b)

Figure 13. Turbulent kinetic energy contour on the center plane(Y=0) with two streamlines passing the edges
of jet exit. (a) Case 2; (b) Case 3

American Institute of Aeronautics and Astronautics

13


0
0.005
0.01
0.015
0.02
0.025
0.03
0 1 2 3 4 5 6
S/rd
T
K
E
/
U
j
U
j
Case1, r=8.45, d=10.92mm, Re=63000
Case2, r=4.97, d=10.92mm, Re=63000
Case3, r=8.38, d=9.398mm, Re=74000
Case4, r=4.96, d=9.398mm, Re=74000
Case5, r=4.96, d=9.398mm, Re=60000
Case6, r=5.37, d=4.648mm, Re=60000


Figure 14. Turbulent kinetic energy variations along the trajectory, non-dimensionalized by rd


4.5 Recirculation bubble

Figure 15a-f describes the characteristics of the recirculation bubble that forms behind the jet in case 3. Figure 15a
shows the contours of static pressure on the center plane (Y=0). A high pressure region exists in front of the jet and
acts as a driving force to bend the jet . It also has a region of negative pressure behind the jet exit, generating a
reverse flow in the jet wake. This negative pressure region stretches into the jet nozzle when the blowing ratio is
small, extending the reversed flow into the nozzle. LES of Yuan et al
13
and DNS of Muppidi and Mahesh
14
also
observed reverse flow inside the pipe at low blowing ratio (r<2), due to the adverse pressure gradient set up in the
vertical direction on the windward side of the jet boundary.

Figure 15b demonstrates the relative position between the recirculation bubble (pink) and the jet fluid (green).
The recirculation region is defined as the region of reversed flow. The recirculation bubble is wrapped by the jet
fluid on its lee side, with the lower end partly extending into the jet nozzle and partly attaching to the wall. A 3-D
shape of the recirculation bubble is shown in Figure 15c for case 3. It is created by extracting the iso-surface of zero
x velocity in the flow field. The recirculation bubble appears to resemble a tilted cylinder with a slight broadening in
the diameter along the transverse y directions. All 6 cases have a shape that is similar to that shown in Figure 15c.

Table 2 compares the 6 cases in terms of the width, length and volume of the recirculation bubble. The length
and width of the bubble are defined as the maximum projected distance along the orientation of the bubble in the x-y
plane (refer to figure 15d). The length and width of the recirculation bubble have also non-dimensionalized by rd
and shown in table 2. The length of all cases is in the range of 1.2-1.3 rd and the width is 0.3-0.4 rd. The higher r
cases (cases 1 and 3) have a smaller width indicating that the bubble becomes narrower when compared to case 2
and 4. Case 1-4 shows small variation in projected length of the recirculation bubble, the lower r shows a smaller
width and length. Interestingly, for cases 5 and 6 which belong to the lowest diameter cases, the trend switches,
lower r shows smaller length and width of the recirculation bubble . Further investigation is necessary to explain this
feature. The volume of the recirculation bubble for Y>0 is calculated by integrating the cell volumes. The results
indicate that, lower blowing ratio or higher free stream velocity generates smaller recirculation region. There is
almost an order of magnitude increase in recirculation bubble volume when the jet diameter is increased by a factor
of 2. This could be observed when we compare case 3and 5 at the higher blowing ratio or case 4 and 6 at the lower
blowing ratio.


Figure 15f, g gives a comparison of recirculation bubble in 2D for case 3 between experiment and numerical
simulation. In terms of X and Y range, this region is smaller in experiment possibly due to minor differences in the
blowing ratio and upstream flow conditions between computations and experiments.

American Institute of Aeronautics and Astronautics

14





Table 2
r d(mm) Length/rd
Length
(m)
Width/rd
Width
(m)
Volume(mm
3
)
Case1 8.45 1.349 0.1245 0.331 0.03059 3.214E-05
Case2 4.97
10.922
1.393 0.07561 0.4 0.02172 1.249E-05
Case3 8.38 1.328 0.1046 0.327 0.02574 2.168E-05
Case4 4.96
9.398
1.352 0.06303 0.406 0.01891 9.280E-06
Case5 8.55 1.318 .05221 0.412 0.01636 2.760E-06
Case6 5.37
4.648
1.212 0.03024 0.385 0.0096 8.350E-07




(a) (b)

















(c) (d)







American Institute of Aeronautics and Astronautics

15


















(e) (f )

Figure. 15a-f , Structure of recirculation Bubble for case 3. (a) Static pressure contour on the center plane;
(b) The relative position of 3D recirculation bubble in red color with in contrast with jet fluid in green color;
(c) 3D close view of the recirculation bubble; (d)The dimension to compare on the side view. 2D recirculation
bubble for case 3: (e) experimental result;(f) numerical result.






V. Conclusion

Numerical simulations of a turbulent circular jet exhausting into cross flow, were performed using commercial
FLUENT Software. The simulations were conducted for a range of jet nozzle diameters and blowing ratio. The
computational results were validated by a quantitative comparison with corresponding 2D Particle Image
Velocimetry data.
The time averaged flow field was analyzed using 3-D stream shape embodied by streamline of finite thickness.
The evolution of jet fluid and its interaction with the cross flow was clearly identified in these simulations. The free
stream fluid was found to follow two different trajectories depending on where it originates. The cross stream fluid
closer to the wall gets entrained from the bottom up into the CVP. The fluid stream originating above the wall
boundary layer follows the jet trajectory and wraps around the jet fluid. The fanning motion of the free stream flow
originating close to the wall was also observed as noticed by many experimental observations (Fric and Roshko
1
,
Smith and Mungal
6
). The cross sectional slices along the jet trajectory indicate that the CVP convolute and grow in
size thereby increasing the interfacial contact between the jet and cross stream fluid which enable mixing between
the two streams.
The jet trajectory was normalized by d, rd and r
2
d. They are also compared with other published experimental
and computational data. All the three scaling laws causes bifurcation and the effect of Reynolds number seems to be
an important factor that causes the split in scaling. An examination and categorization of the magnitude of Reynolds
number is therefore necessary to obtain better scaling law.
Velocity decay, jet width and turbulent kinetic energy along the trajectories were also studied. Under the
existence of cross flow, circular jet decays much faster than a free jet. The slope of the decay was found to be 44%
higher with larger r (~8.5) as compared to smaller r (~5). The TKE along the shear layer indicates that the windward
turbulence is higher than the leeward side.
The shape, size of the recirculation bubble behind the jet was also presented. Reversible flow upstream of the
jet and in the jet wake was identified. Negative flow velocities were also detected inside the pipe. The projected
X (rd)
Y
(
r
d
)
0 0.2 0.4 0.6
0
0.2
0.4
0.6
0.8
U
-0.505806
-1.68742
-2.86903
-4.05065
-5.23226
-6.41387
-7.59548
-8.7771
-9.95871
-11.1403
-12.3219
-13.5035
-14.6852
-15.8668
-17.0484
-18.23
X (rd)
Y
(
r
d
)
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
1.2
U
-2
-4
-6
-8
-10
-12
-14
-16
-18
-20
-22
-24
-26

American Institute of Aeronautics and Astronautics

16
length and width of the recirculation bubble was found to scale well with rd. The volume of the recirculation bubble
was found to increase with increasing jet diameter. The recirculation bubble was found be smaller with lower
blowing ratios or higher free stream velocity.

Acknowledgements
The authors would like to greatly appreciate R. DiMicco and R. Ogden for their support in laboratory; Herve
Mazieres from Fluent France for their help in CFD; Irene Ibrahim for sharing her experimental data.
References
1
Fric, T.F., and Roshko, A., Vortical Structure in the wake of a transverse jet, Journal of Fluid Mechanics, Vol.
279, pp. 1-47, 1994.

2
Kelso, R.M., and Smits, A.J., Horseshoe vortex systems resulting from the interaction between the laminar
boundary layer and a transverse jet, Physics of Fluids, Vol. 7, pp. 153-158, 1995.

3
Kelso, R.M., Lim, T.T., and Perry, A.E., An experimental study of round jets in cross-flow, Journal of fluid
mechanics, Vol. 306, pp. 111-144, 1996.

4
Baker, C.J., The laminar horseshoe vortex, Journal of fluid mechanics, Vol. 95(2), pp. 347-367, 1979.

5
Kamotani, Y., and Greber, I., Experiments on a turbulent jet in cross flow, AIAA Journal, Vol. 10, pp. 1425-
1429, 1972.

6
Smith, S.H., and Mungal, M.G., Mixing, Structure and Scaling of the jet in cross flow, Journal of fluid
mechanics, Vol. 357, pp. 83-122, 1998.

7
Cortelezzi, L., and Karagozian, A.R., On the formation of counter-rotating vortex pair in transverse jets,
Journal of fluid mechanics, Vol. 446, pp. 347-373, 2001.


8
Margason, R. J. 1993 Fifty years of jet in crossflow research. In AGARD Symp. on a Jet in Cross Flow, Winchester, UK,
AGARD CP-534

9
B.P. Pratte and W.D. Baines, Profiles of the round turbulent jet in a cross flow, J. hydraul. Div., proc of the American
Society of Civil Engineers, Vol.92, No. HY6, November, 1967.


10
Patankar, S.V., Basu, D.K., Alpay, S.A., Prediction of the Three Dimensional Velocity Field of a Deflected Turbulent Jet,
Transaction of ASME I: Journal of Fluids Engineering, Vol. 99, pp. 758-762, 1977.


11
Crabb, D., Duarao, D.F.G., and Whitelaw, J.H., A Round Jet Normal to the Cross Flow, Transaction of ASME I: Journal
of Fluids Engineering, Vol. 103, pp. 142-153, 1981.


12
Yuan, L.L. and Street, R. L. 1998 Trajectory and entrainment of a round jet in crossflow. Phys. Fluids 10, 2323-2335.


13
Yuan, L.L., Street, R. L. and Ferziger, J. H. 1999 Large-eddy simulations of a round jet in crossflow. J. Fluid Mech. 379,
71-104


14
Muppidi, S. and Mahesh, K. 2005 Study of trajectories of jets in crossflow using direct numerical simulations. J. Fluid
Mech. 530, 81-100.



15
I.. Ibrahim, S, Murugappan and E. Gutmark, Penetration, Mixing and Turbulent Structures of Circular and Non-Circular
jets in Crossflow, AIAA 2005-0300


16
Hasselbrink, E.F. and Mungal, M. G. 2001b Transverse jets and jet flames. Part 2.Velocity and OH field imaging. J. fluid
Mech. 443, 27-68


17
Y. Kamotani and I. Greber, Experiments on a turbulent jet in a cross flow AIAA J.10. 1425 (1992)

American Institute of Aeronautics and Astronautics

17


18
Haniu, H. and Ramaprian, B. R. 1989 Studies on two-dimensional curved nonbuoyant jets in cross flow. Trans. ASME I: J.
Fluids Engng. 111, 78-86.

19
Su, L.K., and Mungal, M. G. 2004 Simultaneous Measurements of Scalar and Velocity Field Evolution in Turbulent Cross
Flowing Jets. J. fluid Mech. 513, 1-45.

20
T.G.Malmstrom, KirkPatrick, A.T., Christensen, B., Knappmiller, K.D., Centerline Velocity Decay Measurements in Low
Velocity Axi-symmetric Jets, Journal of Fluid Mechanics, Vol. 346, pp. 363-377, 1997.

21
C. Ho and E. Gutmark, 1987 Vortex induction and mass entrainment in a small-aspect-ratio elliptic jet. J. Fluid Mech. 179,
383-405.

You might also like