You are on page 1of 9

Large-Eddy Simulation of a Round Jet in Crossow

Jorg Ziee
1
Ph. D. student, Institute of Fluids Dynamics
2
, ETH Zurich, Switzerland
Leonhard Kleiser
Full professor, Institute of Fluids Dynamics, ETH Zurich, Switzerland
Abstract
A numerical simulation of a compressible round turbulent jet issuing perpendicu-
larly into a laminar boundary layer (jet in crossow, JICF) is performed at a jet-to-
crossow momentum ratio of 3.3 (dened with the bulk momentum of the jet and the
free-stream momentum of the boundary layer) and a Reynolds number of 2100 (based
on the free-stream momentum, jet diameter and dynamic viscosity at the wall). To be
able to compare the results with incompressible reference data while still allowing for
an ecient time integration, a Mach number of 0.2 is chosen. The mixing behaviour
of the JICF is investigated by computing the evolution of a passive scalar at a Schmidt
number that equals the Prandtl number of the ow.
The spatial discretisation of the computational domain consists of a block-structured
multi-block grid with 58 blocks and a total of 8.9 million cells. The simulation code
NSMB uses a nite-volume discretisation with a skew-symmetric fourth-order central
scheme and a second-order Runge-Kutta method for time integration. To account
for the subgrid scales, the approximate deconvolution model (ADM) is employed in a
multiblock formulation.
Of special interest in this investigation are the mixing behaviour of the jet with the
crossow and the complex vortex systems in the mixing region. Results are compared
with incompressible LES data at similar ow parameters as mentioned above.
The evolution of a jet under the inuence of a crossow (jet in crossow,
JICF) has been subject to extensive research for more than half a century [?].
While many earlier investigations have focussed on aerospace applications such
as vertical and/or short take-o and landing (V/STOL) aircraft, steering of
rockets, lm cooling of turbine blades or fuel injection into combustors, in
more recent times ecological aspects such as plumes of smokestacks, volcanoes
or (tunnel) res have also gained attention.
The goals of our study of a jet in crossow are threefold. In the course of the
recent work, the approximate-deconvolution subgrid-scale model (ADM) [?, ?,
?] has been implemented into a semi-industrial ow solver, supporting domain
decomposition and parallel computing. After a validation study using a sim-
pler conguration [?], the complex JICF case provides a means to validate the
domain-decomposition features of the new model code. Additionally, we want
to assess the suitability of ADM for separated ows in a low-order numeri-
cal code. Finally, the experience and infrastructure developed with this ow
case is applicable for the simulation of a ow conguration of great industrial
interest, namely lm-cooling of turbine-blades, a work currently in progress.
1
Email: ziefle@ifd.mavt.ethz.ch
2
WWW: htp://www.ifd.mavt.ethz.ch
1
We selected the experimental JICF setup of Sherif and Pletcher [?], which
was also thoroughly investigated numerically by LES in the Ph. D. thesis of
Yuan [?]. In contrast to many other works, the Reynolds number is quite low
here, rendering well-resolved incompressible LES possible even with the com-
pressible simulation approach we used. The two investigated jet-to-crossow
velocity ratios of 2 and 3.3 are low enough for a rather small vertical size of
the computational domain, but still large enough for a highly complex and in
both cases diering ow structure. The specic ow case, see gure ??, con-
sists of a turbulent jet issuing perpendicularly into a laminar boundary layer.
All relevant ow parameters and geometric dimensions are listed in table ??.
x
y
z

L
z
L
x
L
upstream
L
nozzle
L
pipe
D
Figure 1: Schematic of the jet in crossow conguration. x Pipe simulation,
y nozzle and z boundary layer of main JICF simulation. / Data transfer
between pipe and main JICF simulations (see text).
The jet nozzle consists of two independent parts, marked x and y in g-
ure ??. The upper part y, which ends at the boundary layer, obtains its inow
data from the lower part x by the usual block coupling interface, symbolised
by connection . The lower part of the nozzle, however, is not coupled to the
upper one but assumes streamwise-periodic boundary conditions (connection
). In these so-called blind blocks, an independent simulation of turbulent
pipe ow is carried out concurrently with the main JICF simulation. This
one-way block-coupling mechanism in conjunction with the simultaneous in-
ow data generation is a convenient and easy-to-implement way of providing
2
Table 1: Geometric and calculation parameters. Unless otherwise noted, all
boundary-layer quantities refer to the hypothetical laminar ow state in the
centre of the jet nozzle (x = 0) without the presence of a jet.
L
x
= 16 R = ( w)
Vb
/( u)
,BL
= 3.3
L
y
= 10 Re
JICF
= ( u)
,BL
D/
wall
= 2100
L
z
= 10 Re
jet
= ( w)
Vb
D/
wall
= 6930
L
upstream
= 5.5 Re
BL
= ( u)
,BL

99%,BL
/
wall
= 1050
L
nozzle
= 5 Re

,BL
|
inlet
= 309.28
L
pipe
= 5 Ma

= 1/

R
gas
= 0.2
D = 1
99%,BL
/D = 0.5
unsteady turbulent inow data.
The current computational grid was generated with the commercial mesh
generator ICEM CFD Hexa. Its topology consists of 58 structured blocks
with matching interfaces, of which ve blocks comprise the pipe region (blind
blocks) and another ve blocks are used for the pipe nozzle. The total number
of cells is approximately 8.9 million, of which 4.5% each belong to the pipe
simulation and the jet nozzle.
A complete simulation cycle consists of the following steps. First solely
the pipe ow simulation is run, until a statistically stationary ow state is
obtained. During this stage, no simulation takes place in the blocks of the
main JICF simulation domain (y and z). When the pipe ow simulation
reaches a stationary state, the full JICF simulation setup is turned on. After
a suciently long time, in which the jet penetrates the crossow and the char-
acteristic vortex systems have evolved, statistical sampling is started. The
simulation is continued until the statistics have converged to the desired de-
gree.
Time-averaging was performed by sampling the instantaneous ow eld ev-
ery 30 time-steps during a period of more than 250 time units L/ u
,BL
. This
averaging time is signicantly longer than in the reference computation[?],
where available computational resources limited sampling to 80 time units.
However, our statistics were still found to be not perfectly symmetric, es-
pecially in the far-eld downstream of the jet exit, where the ow is highly
unsteady and irregular.
The computation was carried out on four processors of the NEC SX/8
vector machine at HLRS, using more than 3000 CPU hours (including initial
transients). During time-averaging, 12 GB of memory were necessary. One in-
stantaneous ow eld (including mesh and restart data) consumes 1.4 GB disk
space, while the time-averaged ow eld (73 quantities) requires 5.9 GB disk
space. The total accumulated data (including a series of instantaneous ow
elds for animation and frequency analysis purposes) comprises approximately
700 GB.
3
In the following a few sample results are shown. More details are found in
the corresponding publication [?]. The jet trajectory is one of the most impor-
tant characteristics of a JICF. Naturally, its determination is to some extent
a matter of denition. Here we compute the jet trajectory in three dierent
ways that are commonly used in literature [?]. As the rst possibility, see
gure ??, the streamline of the mean ow eld through the centre of the jet
nozzle ((x, y, z) = 0) marks the jet trajectory. Additionally, the streamlines
starting at the upstream and downstream edges of the jet nozzle are shown
as an approximation of the jet boundary in the centre plane. Secondly, as
depicted in gure ??, the trajectory can be dened as the locus through all
points with maximum scalar concentration in the centre plane y = 0. Alter-
natively, the maximum velocity magnitude can be used as dening quantity,
as shown in gure ??. All trajectories exhibit generally good agreement with
the computational results of Yuan [?] (experimental results were not avail-
able for comparison), even though the sensitivity to the particular method of
determination is quite strong for the latter two methods.
To allow for a more quantitative comparison of our LES results with the
incompressible reference simulation [?] and experiment [?], vertical proles
of the mean ow eld in the symmetry plane are depicted at four dierent
streamwise positions in gure ??. The rst series of pictures, gures ??-??,
shows the average velocity magnitude | u|. The overall agreement with the
reference data is quite good. The rst local minimum ts the experimental
data better in our LES than in the one of Yuan. In the next row of gures,
??-??, the normalised rms uctuations of the velocity magnitude | u|
rms
/| u|
are plotted. Again our LES generally reproduces the comparison results well.
The isocontour of the passive-scalar concentration in gure ?? gives a good
impression of the unsteady and irregular instantaneous ow state caused by
the mixing of the two streams. The roll-up and breakdown of the jet shear
layer is also observable: the initially smooth contour surface above the jet exit
shows a regular pattern of ripples after a distance along the jet trajectory of
approximately one jet hole diameter. A few diameters farther downstream the
structure changes to an irregular form, which coincides with the breakup of
the shear-layer vortices [?].
Figure ?? depicts an isocontour of the instantaneous vortex-identication
quantity

2
. Similar to the ow around a wall-mounted cylinder, a horseshoe
vortex wraps around the jet exit hole. The mentioned shedding of shear-layer
vortices from the upstream edge of the jet nozzle is directly visible. After a
few jet diameters along the trajectory they loose their distinct appearance.
This marks the location of their breakup which was already observed in the
isocontour of the passive-scalar concentration in gure ??.
In gure ?? we display three sets of streamlines: two originating from the
jet nozzle and one from the crossow. The red streamlines start from within
the boundary layer upstream of the jet exit at 10% boundary-layer thickness.
The green and blue streamlines are seeded across the jet nozzle throughout
the two spanwise diameters aligned with the x and y axes.
Following the red boundary-layer streamlines, we observe their deection
around the jet exit, as the jet obstructs their downstream path, and the cre-
ation of a horseshoe vortex. We can clearly recognise the bundling of the outer
4
streamlines to form the two legs of the horseshoe vortex. The streamlines close
to the symmetry plane pass under the other streamlines. After a wide circle
around the jet exit, they nally join the horseshoe leg or move downstream
parallel to it.
At the downstream edge of the jet exit, a spanwise vortex caused by the
roll-up of the jet shear layer induces negative streamwise velocities near the
wall. Therefore, some of the boundary-layer streamlines within the horseshoe
legs are forced to move upstream toward the jet exit. There they fall into the
upward inuence of the emerging jet and its counter-rotating vortex pair.
The path of the green and blue streamlines emanating from the jet nozzle
is more regular. The inner ones follow more or less the jet trajectory, while
the blue ones at the spanwise edges of the jet are subject to the inuence of a
stationary inclined vortex pair above the lateral sides of the jet exit. Farther
downstream, those hanging vortices form the characteristic counter-rotating
vortex pair (CVP).
Acknowledgements
Calculations have been performed at the High Performance Computing Cen-
ter Stuttgart (HLRS). This work was carried out under the HPC-EUROPA
project (RII3-CT-2003-506079), with the support of the European Community
Research Infrastructure Action under the FP6 Structuring the European
Research Area Programme. We are grateful to Prof. U. Rist and his group
for their hospitality.
References
[1] Margason, R. J., Fifty years of jet in cross ow research, Computational
and Experimental Assessment of Jets in Cross Flow, AGARD report CP-
534, Nov. 1993.
[2] Stolz, S. and Adams, N. A., An approximate deconvolution procedure for
large-eddy simulation, Phys. Fluids, Vol. 11, No. 7, 1999, pp. 16991701.
[3] Stolz, S., Adams, N. A., and Kleiser, L., An approximate deconvolution
model for large-eddy simulation with application to incompressible wall-
bounded ows, Phys. Fluids, Vol. 13, No. 4, 2001, pp. 9971015.
[4] Stolz, S., Adams, N. A., and Kleiser, L., The approximate deconvolution
model for large-eddy simulations of compressible ows and its applica-
tion to shock-turbulent-boundary-layer interaction, Phys. Fluids, Vol. 13,
No. 10, 2001, pp. 29853001.
[5] Ziee, J., Stolz, S., and Kleiser, L., Large-Eddy Simulation of Separated
Flow in a Channel with Streamwise-Periodic Constrictions, 17th AIAA
Computational Fluid Dynamics Conference, Toronto, Canada, June 69,
2005, AIAA Paper 2005-5353.
5
[6] Sherif, S. A. and Pletcher, R. H., Measurements of the Flow and Turbu-
lence Characteristics of Round Jets in Crossow, J. Fluids Eng., Vol. 111,
1989, pp. 165171.
[7] Yuan, L. L., Large Eddy Simulations of a Jet in Crossow, Ph. D. thesis,
Stanford University, 1997.
[8] Ziee, J. and Kleiser, L., Large-Eddy Simulation of a Round Jet in Cross-
ow, 36th AIAA Fluid Dynamics Conference, San Francisco, USA, June
58 2006, 2006, AIAA Paper 2006-3370.
Biography
Jorg Ziee is a Ph. D. student at Institute of Fluid Dynamics, ETH Zurich,
Switzerland. His work is concerned with the large-eddy simulation of sepa-
rated ows using domain decomposition and a semi-industrial nite-volume
ow solver. He holds a Master of Science in Aerospace Engineering from
Georgia Institute of Technology and a Diploma (Dipl.-Ing.) from University
of Stuttgart in the same discipline.
6
PSfrag replacements
x
z
(a)
PSfrag replacements
x
z
(b)
PSfrag replacements
x
z
(c)
Figure 2: Jet trajectory determined in dierent ways: a) Streamline trajectory,
b) maximum passive-scalar concentration trajectory and c) maximum velocity
magnitude trajectory. Present LES, LES of Yuan [?], bounding
streamlines in present LES, bounding streamlines in LES of Yuan.
7
PSfrag replacements
z
| u|
(a) x = 0
PSfrag replacements
z
| u|
(b) x = 1.84
PSfrag replacements
z
| u|
(c) x = 3.67
PSfrag replacements
z
| u|
(d) x = 5.54
PSfrag replacements
z
| u|
rms
/| u|
(e) x = 0
PSfrag replacements
z
| u|
rms
/| u|
(f) x = 1.84
PSfrag replacements
z
| u|
rms
/| u|
(g) x = 3.67
PSfrag replacements
z
| u|
rms
/| u|
(h) x = 5.54
Figure 3: Vertical proles of time-averaged quantities at four streamwise po-
sitions in the symmetry plane y = 0. a)-d) Mean velocity magnitude | u|,
e)-h) normalised rms uctuations of the velocity magnitude | u|
rms
/| u|.
Present LES, LES of Yuan [?], experiments of Sherif and Pletcher [?].
8
(a) (b)
(c)
Figure 4: a) Smoke visualisation of the instantaneous ow eld. b) Visuali-
sation of the instantaneous ow eld using the vortex identication criterion

2
. c) Streamlines of the mean velocity eld. Red: streamlines originating
from the boundary layer upstream of the jet exit hole at 10% boundary-layer
height. Green and blue: streamlines seeded from the two diameters aligned
with the x and y axis across the jet nozzle.
9

You might also like