Vous êtes sur la page 1sur 27

Flow, Turbulence and Combustion 65: 177203, 2000.

2001 Kluwer Academic Publishers. Printed in the Netherlands.


177
LES of Jets in Cross Flow and Its Application
to a Gas Turbine Burner
J.U. SCHLTER and T. SCHNFELD
CERFACS, 42, Av. Gaspard Coriolis, 31057 Toulouse Cedex 1, France
Received 14 December 1999; accepted in revised form 12 April 2000
Abstract. LES computations of jets in cross ow (JICF) were performed. Experimental investiga-
tions reported in literature are reproduced with good agreement concerning the momentum eld and
the mixing of a passive scalar. The results validate the ability of the present LES approach to compute
fuel injection of the type JICF. LES computations of fuel injection in an industrial gas turbine burner
are presented.
Key words: LES, jet in cross ow, mixing, gas turbine burner.
1. Introduction
The increasing demand in reducing pollutants in gas turbines [31] forces gas tur-
bine manufacturers to investigate possibilities to mix fuel and oxydizer in the best
way possible before their ignition. It is well known that fuel rich combustion
leads locally to higher temperatures and higher pollutant formations. Although
Reynolds-averaged mean values might suggest a perfect mixedness, the unsteady
nature of turbulent combustion may result in a temporally surplus of fuel in the
combustion chamber. Gas turbine manufacturers seek to avoid these conditions in
their burners as from the design stage on.
The motivation for this project originates in a demand from the Siemens Power
Generation (KWU) company to undertake a Large Eddy Simulation (LES) analysis
of a gas turbine burner conguration and to explore possibilities for using LES as a
design tool. Figure 1 shows such a gas turbine burner. It is of swirl type, where the
swirl is induced by several circumferential vanes. The fuel (natural gas) is injected
on the surfaces of the vanes. In Figure 2 a close-up on the vane is shown, where air
is coming from the left and the fuel is injected perpendicular to the airow.
The objectives of this LES analysis are to determine the mixing quality of the
burner and to track vortex developments around the vane, which might disturb later
on in the combustion chamber and potentially lead to combustion instabilities. The
problem met herein is that measurements inside the burner are extremely difcult to
achieve under operating conditions because of the limited access of measurement
probes and optical devices to the inner geometry of the burner. Measurements on
178 J.U. SCHLTER AND T. SCHNFELD
Figure 1. A Siemens gas turbine burner currently under investigation at CERFACS.
Figure 2. Close-up on the vane of the burner. The fuel is injected on the surface of the vanes.
The air (from the left) and the fuel form a JICF.
the inlet and outlet of the burner have been made, but they are too far away from
the fuel injection to give signicant insight to the ow around the vane.
For this reason, a direct comparison between LES and measurement is nearly
impossible to validate the LES results. Hence, before applying the LES approach
to the gas turbine burner, a validation of the LES technique and the underlying
ow solver is done on simpler test cases. The ability of the underlying ow solver
to reproduce foil ows has been shown previously [21], so that the description of
the vane itself did not pose a problem. However, the ability of the ow solver to
reproduce the fuel injection still had to be done. As a simplied test case the jet in
cross ow (JICF) has been chosen to investigate the fuel injection. Here, the vane
is replaced by a plane wall.
The computation of the burner ow requires a high number of mesh points
to describe the principle ow features: the inlet ow, the vane ow, the injector
ow and the swirling ow at the outlet. The variety and number of different ow
LES OF JETS IN CROSS FLOW 179
problems encountered here will need each a high number of mesh points to describe
the ow properly. It is desirable to know the minimum number of mesh points to
describe the different problems in order to keep low the total number of mesh
points in the full burner computation. Hence, the present investigations of the JICF
are limited to computations on relatively coarse meshes.
2. Flow Physics of the Jet in Cross Flow
2.1. INTRODUCTION
The JICF has attracted attention in uid mechanics research: turbine blade cooling,
V/STOL aircrafts, roll control of missiles or chimney ows are examples of the
wide eld of applications, but most of the interest in research has been focused on
the application to combustors.
The JICF is a very pleasant ow conguration with regard to mixing. It is one
of the most effective way to mix two uids in a limited space, which is superior to
other ow constellations like the mixing layer or the jet in coow [8].
Investigations on the JICF have started in the 1930s [28] with the mixing of
chimney plumes. Since, there have been numerous investigations on the JICF lead-
ing to the perception, that the JICF, in contrast to other ow congurations like
the jet and the mixing layer, cannot be generalized in terms of self similarity and
Reynolds dependence, due to strong nonlinear effects. The systematic analysis of
the JICF started in the 1970s with the discovery and acceptance of coherent struc-
tures [9, 10]. A clear denition of coherent structures cannot be given. Hussain
[15, 16] tries to dene them as a connected, large scale uid mass with a phase
correlated vorticity over his spatial extend, but still this denition is incomplete.
However, coherent structures are able to explain various nonlinear effects in the
JICF. Up to now, at least four different types of coherent structures are determined
in the JICF. They are shown in Figure 3. The most dominant vortex system is the
counterrotating vortex pair. The three other vortices, the jet shear layer vortices,
the wake vortices and the horseshoe vortex are often called secondary vortices, as
they play a minor, although not neglectable role in the far eld of the jet.
In the following some recent developments in JICF research are summarized.
Note, that there exists an extensive review from Margason [19] from 1993.
2.2. THE MOMENTUM RATIO
The most dominant quantity to characterize a JICF is the momentum ratio r dened
as:
r =
_

jet
v
2
jet

crossow
v
2
crossow
. (1)
180 J.U. SCHLTER AND T. SCHNFELD
Figure 3. Vortex system for a jet in crossow (from Fric and Roshko [13]).
In most cases the jet uid and the crossow uid consist of the same species and
have the same temperature, hence
jet
=
crossow
. The momentum ratio simplies
then to the velocity ratio:
r =
v
jet
v
crossow
. (2)
Different ow regimes can be determined based on the velocity ratio r. JICF
with r < 0.5 play a special role. The jet ow is weaker than the crossow. It is not
able to break through the wall boundary layer of the crossow and plays more the
role of an obstacle for the crossow. The far eld of the jet is primarily governed
by the oncoming boundary layer. This fact introduces a large number of additional
parameters to the investigations. Dependent on the wall boundary layer thickness
this ow regime can occur up to a velocity ratio of r = 1. This ow conguration
is especially important for turbine blade coolings [7].
Velocity ratios r between 1 and 10 are common ow regimes for combustion
applications. The jet is then able to push through the boundary layer, which plays
a minor role. The JICF is now determined by free turbulence characteristics and
is easier reproducible. Andreopoulos [5] measured velocity proles for JICF with
velocity ratios from 0.5 to 2, where the described transition from wall boundary
layer to free turbulence can be seen.
JICF with velocity ratios higher than 10 have additional effects as they behave
more and more like free jets with increasing velocity ratio.
LES OF JETS IN CROSS FLOW 181
Figure 4. Left: jet shear layer vortices and trajectory of the horseshoe vortex, right: horseshoe
vortex and recirculation bubble at the upstream edge of the jet orice (from Kelso [18]).
2.3. VORTEX SYSTEMS
The mechanism for the formation of the counterrotating vortex pair (CVP) is not
fully understood. It can be taken as certain, that the vorticity of the CVP has its
origin at the sidewalls of the jet. Haven [14] investigated different nozzle geome-
tries for the jet. Rectangular jets with a low aspect ratio (the edges at the sides of
the jet are longer than the upstream and downstream edge) amplied the CVP. The
longer sidewalls produce more vorticity, which can be found later in the CVP. On
the other side, jets with a high aspect ratio have a weaker CVP.
Toy [30] investigated two closely spaced jets issuing into a crossow. The jets
were either side-by-side or in-line. Hot-wire measurements on the centerline of
the jets were made. As a major feature found, only one CVP, instead of two,
was produced, when the jets were in a side-by-side setup. The data obtained, was
not sufcient to give a reason for this behaviour. To the authors knowledge, a
mechanism which leads to a single CVP is not reported in literature.
The jet shear layer vortices are produced directly at the jet orice (Figure 4).
The two streams (jet stream and crossow) form a mixing layer with a Kelvin
Helmholtz instability, which causes a roll-up near the edges of the jet.
The horseshoe vortex forms upstream of the jet at the crossow wall. The
adverse pressure gradient at the crossow wall forces the wall boundary layer
to separate and to form a vortex. It is then convected and stretched by the ow
and wraps around the jet nozzle like a necklace. The same kind of vortices can
be observed for ows where a boundary layer hits on an obstacle, e.g. a cylinder
mounted on a wall [6]. The vortex is in interaction with the upstream edge of the
jet orice, causing a separation bubble inside the jet (see also [17, 20]). Figure 4
shows a sketch of the horseshoe vortex and the separation bubble.
The existence of the separation bubble was rst proposed by Andreopoulos [2].
He measured the velocities in the interior of the jet-pipe and found non-symmetric
proles. In this investigation, the crossow affects the jet-pipe ow up to 2D up-
stream of the nozzle.
182 J.U. SCHLTER AND T. SCHNFELD
The wake vortices were rst believed to be the consequence of a shedding
process behind the jet, with the jet acting like a solid cylinder and the wake vortices
behaving like a von Krmn vortex street. However, Perry [20] pointed out, that in
an incompressible uid no vorticity can be produced inside the ow. The vorticity
transport equation for incompressible uids:
D
Dt
= u +
2
(3)
shows a convection term on the left-hand side, a term which describes the reori-
entation and stretching of the vorticity due to the velocity gradients and a viscous
term on the right-hand side, but no explicit production term. That means, in incom-
pressible ows vorticity can only enter a ow by initial conditions and imposed
wall boundaries, but not be generated in the interior, which would be the case, if
the wake vortices would be shed from the jet like from a solid cylinder. Fric and
Roshko [13] connected the wake vortices to a separation event at the cross ow
wall at the sides of the jet nozzle. The end of the vortex string, which is close to
the nozzle is convected by the jet and follows the jet trajectory, while the other end
stays close to the wall, bringing the vortex in an upright position.
2.4. PREVIOUS NUMERICAL INVESTIGATIONS OF THE JICF
The rst numerical investigation of a JICF has been made by Sykes [29]. He
simplied the calculation by using a slip wall as the crossow wall. Hence, the
crossow boundary layer is neglected and the horseshoe vortex and the wake vor-
tices cannot be calculated. His results agree qualitatively with the measurements of
Andreopoulos [3, 5], besides a large error near the wall.
A numerical investigation from Chiu [11] tested the applicability of different
algebraic turbulence models with the result, that the turbulence models do not
improve the calculations compared with a laminar calculation. It showed the limits
of an eddy viscosity model applied to free turbulence.
The investigation of Alvarez [1] used the k model and a direct closure. The
direct closure improved the calculations, but the error was still high.
The strong unsteady behaviour of the JICF leads to the conclusion, that an
unsteady LES approach might be useful for reproducing and predicting the JICF.
Yuan [32, 33] made an LES calculation of a JICF. He calculated a JICF with a
small Reynolds number (Re = 2100) on meshes with 1.3 million mesh points.
The results agree quite well with measurements, although no direct comparison
is possible , because the Reynolds number is lower and the velocity ratio is not
exactly the same as in the experimental references.
He pointed out, that it is necessary to mesh the pipe which supplies the jet, in his
case on a length of one diameter D upstream. But, as Andreopoulos [2] mentioned,
there is an interaction between the crossow and the jet-ow. In the examined case
(r = 2) the crossow affects the pipe-ow around two diameters upstream the
LES OF JETS IN CROSS FLOW 183
pipe, so that his extension was probably not sufcient. The high number of mesh
points and the low Reynolds number make it difcult to apply this computation to
industrial congurations.
3. Mathematical Formulation of LES
3.1. GOVERNING EQUATIONS
The basic idea of LES is to resolve the larger scales of motion of the turbulence
while approximating the smaller ones. To achieve this, a lter is applied to the
continuity equation and the transport equations of momentum, energy and species.
For reacting ows, often Favre ltering is used, which is dened as:


Q = Q =
+
_

Q(x, t )G(x x

) dx

, (4)
leading to the following equations for momentum u
i
and species Y
i
:
momentum (j = 1, 2, 3)
u
i
t
+
u
i
u
j
x
i
+
p
x
i
=

ij
x
j
+
T
ij
x
j
; (5)
species mass fraction (k = 1, . . . , N)


Y
k
t
+
u
i

Y
k
x
i
=

x
i
_
D
k


Y
k
x
i
_
+


k
+

kj
x
j
. (6)
3.2. SUBGRID SCALE MODELS
The terms T
ij
and
ik
result from the convective terms u
i
u
j
and

Y
k
u
i
, which are
split into a resolved part on the left-hand side of the equation, directly delivered by
the LES calculation, and an unresolved part on the right-hand side, which needs to
be modeled.
We used an eddy viscosity approach for the subgrid scales:
T
ij
= 2
t

S
ij
+
1
3
T
ll

ij
(7)
with

S
ij
=
1
2
_
u
i
x
j
+
u
j
x
i
_
. (8)
184 J.U. SCHLTER AND T. SCHNFELD
Although the eddy viscosity approach is not valid for free turbulence, its simplic-
ity allows faster computations and by this a higher spatial discretization and an
increase of the resolved part of the spectrum.
Subgrid mixing is modeled by an eddy diffusity approach with a turbulent dif-
fusity based on the turbulent viscosity
t
of the subgrid stress model and a constant
Schmidt number Sc:

k
j =

t
Sc


Y
k
x
j
. (9)
3.3. PRESENT IMPLEMENTATION
For our LES calculations we used the AVBP parallel solver developed at CERFACS
and the Oxford University [21], based on the parallel library COUPL [22]. The
program handles structured and unstructured meshes and is second-order accurate
in space and time.
Two models were used to determine the eddy viscosity
t
. The rst one is the
Standard Smagorinsky Model [23]:

t
= (C1
x
)
2
_
2

S
ij

S
ij
(10)
with C
1
= 0.18, which has the advantage of simplicity and speed.
The second model is the Filtered Smagorinsky model [12] dened on a high-
pass lter HP:

t
= (C2
x
)
2
_
2HP(

S
ij
)HP(

S
ij
) (11)
and a constant C
2
= 0.37. This model offers a better behaviour in transitional ows
and was optimized to work in wall boundary layers.
4. Grid Resolution
The spatial discretization of the ow is based on structured meshes. Although the
AVBP solver allows unstructured meshes, the structured mesh approach provides
a better control of the point distribution in the ow. Figure 5 shows an exploded
view of the mesh. At the bottom is the plenum chamber of the jet (A) passing over
into a pipe (B). The jet nozzle is at the upper end of the pipe. An 0-grid is put in
the jet trajectory and the vicinity of the nozzle (C). A block behind the nozzle (D)
describes the ow downstream of the nozzle and several coarse blocks (E) are put
around the jet trajectory to mesh the nearly undisturbed outer ow.
Meshing the jet pipe ow is important. As already pointed out, there is an inu-
ence of the crossow to the jetow in the pipe. The existence of a recirculation zone
at the edge of the jet nozzle (see Figure 4) makes this area sensible to mesh point
distribution. The mesh has to be ne enough to capture this recirculation zone.
LES OF JETS IN CROSS FLOW 185
Figure 5. Exploded view of grid blocks.
In the LES computation of Yuan [33] the inuence of an extension of the mesh
into the jet pipe was examined. He found out, that the ow behaviour is much better
reproduced with such a mesh extension.
We found out, that a simple extension might not be sufcient. Our rst calcula-
tions have been carried out with a 3D long pipe leading to the jet nozzle and the
velocity prole u was imposed at the entry of the pipe. This led to strong pressure
oscillations in the pipe. As a numerical artefact, the pipe acts as a Helmholtz res-
onator, because the inlet below the wall forms a velocity node. The frequency of
the oscillations are determined by the length of the pipe. The jet shear layer roll-up
locks into the oscillations and the jet acts like a forced jet. Kelso [17] found, that
the jet trajectory is affected by the forcing. In our computations the trajectory is
higher than in the case where the oscillations are suppressed.
In order to avoid pressure oscillations we use a combination of two counter-
measures. The rst one is to impose the mass ux u instead of the velocity u
as a boundary condition to change the acoustic wave reections at the inlet. The
second is to extend the jet pipe mesh into the plenum chamber in front of the jet
pipe (Block A in Figure 5). The sudden change in diameter between jet pipe and
plenum chamber makes it more difcult for the system pipe/plenum chamber to act
186 J.U. SCHLTER AND T. SCHNFELD
Figure 6. Smoke visualization of the test case of Andreopoulos and Rodi, r = 2, Re = 81000.
as a resonator. This approach is more expensive, but offers as a by-product more
certainty on the jet velocity prole.
Furthermore a renement of the mesh in the low pressure region downstream
of the jet nozzle is necessary. It inuences the jet trajectory and higher trajectories
were obtained with a low resolution mesh. This mesh is automatically ne enough
to capture the wake vortices. But, additionally the wall region, where the wake
vortices have their origin, has to be well resolved as well.
However, it turned out that the number of mesh points was still too high, when
applied to a real conguration of the burner. To assess the accuracy of even coarser
meshes, additional computations were performed on an even simpler mesh. Here,
the jet pipe is meshed by a 6 6 H-mesh. The crossow is meshed with one single
block of H-type topology. The plenum chamber stays part of the computed domain
and is meshed by an H-block as well.
5. LES Validation Test Cases
In order to obtain meaningful statements on the validity of the LES computations
of the injectors of the real burner geometry, test cases are needed which are close
to the real problem with respect to the Reynolds number and the velocity ratio r.
In the real problem the Reynolds number is Re 8000, based on the exit velocity
of the injecting fuel and the injector diameter, and the velocity ratio r 2 at the
nozzle. Three cases have been chosen:
LES OF JETS IN CROSS FLOW 187
1. The rst one is a series of experiments carried out by Andreopoulos and Rodi
[25]. They provide detailed hot wire measurements of the mean velocity
components, the turbulent kinetic energy, the Reynolds stresses and measure-
ments on the turbulent kinetic energy budget. Furthermore they made measure-
ments on a slightly heated jet to obtain statements on the mixing behaviour by
measuring the temperature eld.
In the experiments the velocity ratio varies from r = 0.5 (Re = 20500) to
r = 2 (Re = 82000). Since the behaviour of the jet changes dramatically for
velocity ratios lower than r = 1 the experiments with a velocity ratio r = 2
were chosen to be reproduced despite the high Reynolds number.
2. Because the work on the gas turbine burner focuses on mixing, we have chosen
as a second test case the experiments by Smith and Mungal [2427]. They
seeded the jet air with acetone and made LIF measurements of the mixing
behaviour. Here, we have chosen to reproduce their measurement of a JICF
with a Re = 16400 and a velocity ratio r = 5.
3. An additional problem in the gas turbine concerns the interaction between
adjacent jets. Hence, the third test case is a measurement of a twin jet from
Toy et al. [30]. He measured velocity proles in the far eld of the jets with
Re = 31800 and a velocity ratio r = 6. Here the establishment of a single
CVP in the far eld of the jets shall be reproduced.
Despite the requirements on the mesh, a dexterous point distribution in the mesh
limits the number of points to 90.000 for the test cases of Andreopoulos and Rodi
and Smith and Mungal and 200.000 for the case of Toy. Tests with coarser meshes
have been done, but the results were poor, so that it can be assumed, that this is a
lower limit in terms of mesh points. The rst cell on the surface of the crossow
wall has a thickness of y
+
= 90 in the case of Andreopoulos and Rodi and Toy and
a thickness of y
+
= 50 in the case of Smith and Mungal.
6. Computational Results
6.1. REPRODUCTION OF GENERAL FLOW CHARACTERISTICS
All ow visualizations in this section have been made from the Reynolds-averaged
ow eld of the unsteady computation of the test case of Andreopoulos and Rodi.
The features shown can be found in all other computations as well.
Figure 7 shows the streamlines computed by LES in the initial region of the
CVP. It can be seen, that the CVP starts to develop very early. Haven [14] showed,
that the vorticity of the CVP has its origin at the side walls of the jet pipe. It seems
from the ow visualization of the LES computation, that this can be conrmed.
The vortices start very early, right behind the orice.
Figure 8 juxtaposes the streamline pattern proposed by Perry and Kelso [20]
with the computed streamlines. The resolution of the mesh upstream is not ne
enough to resolve all vortices in this region. Baker [6] showed that the two sec-
188 J.U. SCHLTER AND T. SCHNFELD
Figure 7. Streamlines at the jet orice, viewer is downstream of the orice looking up-
stream. The developing counterrotating vortex pair can be seen. Streamlines created from
Reynolds-averaged ow eld from the unsteady LES computation from the case of An-
dreopoulos and Rodi.
Figure 8. Side view on the computed streamlines at the jet orice fromthe case of Andreopou-
los and Rodi juxtaposed with the proposed streamline pattern of Perry and Kelso [20]. The
horseshoe vortex system is not fully resolved by the LES computation and thus simpler. The
recirculation zone at the upstream edge of the jet pipe can be seen. Streamlines created from
Reynolds-averaged ow eld from the unsteady LES computation.
ondary vortices of a horseshoe vortex are weak. The weakness and the small size
of these vortices make it difcult for a numerical investigation to capture these
structures. Nevertheless a simplied horseshoe vortex develops upstream of the
orice.
Figure 9 juxtaposes the streamline pattern in the xy plane on the centerline of
the jet found by Kelso [17] and the streamline pattern of the LES computation of
the case of Andreopoulos and Rodi. The comparison can be done only qualitatively,
because of the different experimental setups. In both ow elds a velocity node in
the wake, close to the wall can be found.
In Figure 10 the side view and top view of the streamline patterns are compared.
The streamline patterns resemble, except for one important difference: the vortex
in the xz plane directly behind the jet nozzle turns in opposite direction. Kelso
LES OF JETS IN CROSS FLOW 189
Figure 9. Qualitative comparison of streamline patterns in the xy plane on the centerline
of the jet, left Kelso [17] with a v
r
= 2.2 and the LES computation of Smith and Mungal
with a v
r
= 5. Streamlines created from Reynolds-averaged ow eld from the unsteady LES
computation.
Figure 10. Qualitative comparison of streamline patterns in the xy plane on the centerline
of the jet, left Kelso [17] with v
r
= 2.2 and the LES computation of Andreopoulos and Rodi
with v
r
= 2, the half-circle denotes the position of the jet nozzle. Streamlines created from
Reynolds averaged ow eld from the unsteady LES computation.
admitted in his investigation, that the resolution of his measurement grid is quite
coarse in this region. Because of the higher resolution in the LES computation we
believe, that the orientation of this vortex has to be clockwise. The direction of the
vortex can be conrmed from Figure 7. The vortex pair developing directly behind
the nozzle is in a nearly upright position at the nozzle itself and bends over with
the jet. That means, the vortex seen behind the nozzle in Figure 10 must have the
same orientation as the CVP. This is only the case, when the vortex behind the
nozzle spirals out clockwise. The origin of this vortex is unclear. Neither the jet
nor the wall boundary layer of the crossow carries the vorticity of this sign. A
computation, where the plane wall from the crossow has been replaced by a slip
wall, still showed this vortex. This indicates, that the origin of this vortex is located
in the pipe wall.
190 J.U. SCHLTER AND T. SCHNFELD
Figure 11. Comparison of the momentum elds on the centerline of the jet, circles: mea-
surements, solid line: LES Filtered Smag. Model, dashed line: LES Standard Smag. Model,
averaging time 0.25 s.
In the literature this vortex is unmentioned, with the exception from the investi-
gation of Kelso [17]. It seems, further investigation on the origin and further history
of this vortex is necessary, especially to shed light into the development of the CVP.
6.2. TEST CASE OF ANDREOPOULOS AND RODI
6.2.1. Momentum Field
Comparisons between the hot wire data obtained from Andreopoulos and Rodi and
our LES computations were made to quantify the reproducibility. In order to show
the ability of LES to simulate the experiment on low-resolved meshes, all LES
computations are made on meshes with 90000 mesh points. The u component of
velocity is compared in Figure 11 for different positions downstream on the jet
centerline. Regarding the prole at x/D = 2 the measurements and the LES com-
putation with the Filtered Smagorinsky model agree well. The LES computation
using the Standard Smagorinsky model shows a wrong trajectory, the location of
the velocity maximum is too high by 0.4D, but the right order of magnitude. As
already mentioned, the Filtered Smagorinsky model was optimized for boundary
layers. Hence, the oncoming wall boundary layer is better described and the mo-
mentum ratio close to the wall is better predicted. This has an inuence on the jet
trajectory. The u velocity proles downstream show, that the Filtered Smagorinsky
LES OF JETS IN CROSS FLOW 191
Figure 12. Comparison of the turbulent kinetic energy elds on the centerline of the jet. The
graphs concerning LES contain only the resolved part of the turbulent motion. Circles: mea-
surements, solid line: LES Filtered Smag. Model, dashed line: LES Standard Smag. Model,
averaging time 0.25 s.
model has advantages over the Standard Smagorinsky model, although the trajec-
tory is slightly too high. The measured velocity decit below the jet trajectory is
leveled out in the LES computation too early.
The turbulent kinetic energy (TKE) k
2
= u
2
+v
2
+w
2
was chosen to compare
dynamic variables. The TKE of the LES calculations presented here represent only
the TKE of the resolved spectrum of turbulence. This means, that the subgrid
turbulence does not appear in the graphs. In regions, where the level of subgrid
turbulence is high, the TKE of the LES computations is underestimated. This is
especially the case in wall boundary layers, where the high shear stress at the wall
implies a production of small scale structures, which are not captured by the mesh
resolution. In free turbulence far off the wall, the LES mesh is ne enough to allow
comparisons with measurements.
Figure 12 shows a comparison of TKE. The prole at x/D = 4 shows a
good agreement of the LES calculation with the Filtered Smagorinsky model.
The Standard Smagorinsky model shows a too high trajectory (which conrms
the observation from Figure 11) and overestimates the TKE. Obviously the Stan-
dard Smagorinsky model is not dissipating sufciently the turbulent energy. The
constant C
1
of the Standard Smagorinsky model could be increased to adapt the
model to the ow, but this would question the universal validity of the model. The
downstream proles show a quite good agreement far off the wall, but close to the
192 J.U. SCHLTER AND T. SCHNFELD
Figure 13. Comparison of the velocity proles at the jet orice. Circles: measurements, solid
line: LES, averaging time 0.25 s.
wall the TKE of the LES computations is well below the measured TKE. Here, as
mentioned above, the level of subgrid turbulence is not neglectable.
A look at the measured velocity prole at the jet orice (Figure 13) from
Andreopoulos and Rodi shows a major discrepancy in the recirculation bubble
between the hot wire measurements and the LES computation, which can be ex-
plained as follows: although Andreopoulos already proposed the possibility of a
temporal recirculation zone at this location for very short time spans [2], he did
not expect a steady recirculation zone. But the major limitation of hot wire mea-
surements is, that they cannot measure the direction of the ow and thus always
pretend to measure a positive velocity. This explains, that in the region of the recir-
culation bubble Andreopoulos and Rodi still measure a positive velocity. The LES
computations in contrast, show the recirculation zone and conrm the observations
of Perry and Kelso [20] (see also Figure 8).
6.2.2. Scalar Field
Due to the lack of reliable measurement techniques for mixing in the 1980s,
Andreopoulos heated the jet slightly by 4

C and measured the temperature eld


instead. The goal was to obtain information on the mixing behaviour of a species
injected by the jet with a Schmidt number equal to the Prandtl number. The tem-
perature difference had to be kept small in order to preserve nearly equal densities.
This helped to leave the jet velocity unharmed while maintaining the impulse ratio
r and to avoid additional effects by buoyancy in the far eld of the jet. But accuracy
problems arise when measuring the temperature with thermocouples, especially
in highly turbulent regions. An additional effect occurs close to the wall, which
LES OF JETS IN CROSS FLOW 193
Figure 14. Comparison of the mixture fraction of a passive scalar distribution (LES) with
the non-dimensionalized temperature distribution on the centerline of the jet. Circles: mea-
surements, solid line: LES Filtered Smag. Model, dashed line: LES Standard Smag. Model,
averaging time 0.25 s.
potentially heats up. For the LES computations the eld of a passive scalar with a
Schmidt number of Sc = 0.72 equal to the Prandtl number was computed.
With the knowledge of these problems, the comparison of the measured temper-
ature eld with the computed mixture fraction of a passive scalar on Figure 14 has
to be looked at carefully. At x/D = 2 the agreement is quite well, but downstream
the proles increasingly disagree, especially close to the wall. This is explained
with heating up of the wall in the experiment. However, we do not feel, that this
bad agreement is of major importance, since the following test case from Smith
and Mungal shows better agreement, so that the disagreement in the case from
Andreopoulos and Rodi can be explained by different boundary conditions.
6.3. TEST CASE OF SMITH AND MUNGAL
6.3.1. Scalar Field
The doubts in the comparability of the measured temperature eld with the com-
puted mixing of a scalar led to further investigations of the JICF with regard to
mixing. In the recent years LIF measurements became a reliable tool to predict
mixing behaviour in turbulent ows. Acetone seeded air is used to track the history
of uid particles in the ow.
194 J.U. SCHLTER AND T. SCHNFELD
Figure 15. Comparison of the mixture fraction of a passive scalar distribution of the measured
non-dimensionalized mixture fraction of acetone seeded air with the LES computation on the
centerline of the jet. Circles: measurements, solid line: LES Filtered Smag. Model, dashed
line: LES Standard Smag. Model, averaging time 0.25 s, mesh size: 90000 points.
Our LES computations that aim to reproduce the case of Smith and Mungal
show basically the same behaviour as before: a CVP is developing, the horseshoe
vortex, wake vortices, jet shear layer vortices and the recirculation zone at the
leading edge of the jet nozzle appear as seen previously.
The eld of a passive scalar with a Schmidt number Sc = 1.0 was computed
to compare it with the mixture fraction of acetone from the experiment. Figure 15
shows the comparisons between experiment and LES computations, which show
an excellent agreement. The computation using the Standard Smagorinsky model
shows a slightly too high trajectory. The better agreement compared to the previous
case can be explained by two facts. First, the Reynolds number is lower than in the
case of Andreopoulos and Rodi. This fact speaks for a sufcient accuracy for the
LES computations of an injector on the vane in the real geometry, because the
Reynolds number is even lower than in the case of Smith and Mungal. Second,
because of the higher velocity ratio, the boundary layer plays a minor role. The
origin of the boundary layer on the vane is only slightly upstream of an injector,
at the leading edge of the vane, and is accelerated then. This means, the boundary
layer thickness is supposed to be very small at the height of the injectors and has
only little effect on the jet ow.
As model for the description of the subgrid mixing the eddy diffusity model
was chosen. The results show, that even this simple model is able to reproduce the
LES OF JETS IN CROSS FLOW 195
Figure 16. Inuence of Schmidt number to mean mixture fraction distribution, LES computa-
tion of two passive scalars with different Schmidt numbers. Solid line: Sc = 0.72, dotted line:
Sc = 1.0.
mixing behaviour of a JICF. Because of its simplicity and fast execution it was
chosen for all other LES computations.
6.3.2. Inuence of the Schmidt Number
The LIF measurements with acetone seeded air and a Schmidt number of Sc 1.0
are performed to simulate the mixing behaviour of methane (Sc 0.72). To de-
termine the inuence of the Schmidt number on the mixing behaviour, an LES
computation was performed with two passive scalars. One of the passive scalars
possesses the Schmidt number Sc = 1.0 and the other Sc = 0.72. Only one com-
putation needed to be performed, because the passive scalars have no feedback on
the momentum eld. Here, the advantage of a numerical investigation is clear: both
scalars use exactly the same momentum eld. Figure 16 shows the comparison of
the distribution of the scalars. The curves are very close together and show nearly
no deviation. As expected, the distribution of the scalar with a Sc = 0.72 diffuses
slightly more. It can be concluded, that the LIF measurements simulate the mixing
behaviour of methane accurately.
6.3.3. Inuence of Mesh Coarsening
In order to assess the inuence of grid resolution, tests on simpler, less problem
adapted meshes were performed. Here, the mesh cell size and point distribution is
comparable to meshes used in the gas turbine burner. The meshes have approxi-
196 J.U. SCHLTER AND T. SCHNFELD
Figure 17. Inuence of mesh coarsening to mixing, solid line: LES computation, mesh size:
30000 points, dashed line: LES computation, mesh size 90000 points, circles: measurements.
mately 30000 mesh points. Although it cannot be assumed that the LES computa-
tion reproduces exactly quantitatively the measurements, we aim to reproduce the
measurement qualitatively.
The comparison of the scalar elds (Figure 17) shows some disagreements. The
results are worse than in previous computations, but surprisingly good given the
fact, that they are performed on insufcient meshes where not all requirements of
LES are met. The trajectory is too high, which is probably due to the bad resolution
of the boundary layer. Furthermore the scalar quantity is not sufciently diffused.
This effect results from the ltering of the small scale structures by the coarse mesh
and thus a worse description of the turbulent transport. This could be improved by
a more sophisticated subgrid mixing model, but this idea was dropped in favour of
fast computation times.
In conclusion, the LES computation delivers reasonable qualitative solutions
even on coarse meshes and statements about mixedness can be derived from these
computations. But qualitative comparisons have to be regarded carefully.
6.4. TEST CASE OF TOY
To obtain information about the reproducibility of the merging mechanism of two
adjacent jets the experiments of Toy [30] were computed. Here, additional prob-
lems occur, because little is known about the interactions of the jet. To the authors
knowledge, no mechanism is reported to explain the behaviour of the two jets
LES OF JETS IN CROSS FLOW 197
Figure 18. Test case of Toy [30], comparison of the momentum eld on the centerline of the
two jets. Circles: measurements, solid line: LES, averaging time 0.235 s.
to form one single CVP instead of two. The aim of the LES investigation is to
reproduce this merging mechanism and to shed light into this phenomenon.
The quantitative comparison of the momentum eld on the centerline between
the two jets is shown in Figure 18. The results are not as good as in the previous
LES computations. It has to be mentioned, that the interactions of the two jets leads
to longer durations until a stationary ow pattern is established. That means, the
computed time span of t = 0.235 s might not be sufcient to obtain reliable mean
values on the centerline of the jet.
Nevertheless, some major ow characteristics can be reproduced. The two jets
merge and form a single CVP (Figure 19) in the far eld of the jets. This agrees
with the measurements of Toy [30].
The knowledge of the whole velocity eld (and hence the entire vorticity eld)
makes it possible to have a closer look at the mechanism which leads to the de-
velopment of a single CVP and the fate of the two other vortices. Fig 19 shows
a series of streamwise vorticity distributions downstream of the two jets. From
Figures 19a and 19b it can be seen, that the two jets produce two CVPs, that means,
four vortices. Further downstream the two jets attract each other and start to merge.
This attraction can be explained with the Coanda effect. The Coanda effect
normally occurs when a jet is close to a parallel wall. The entrainment of the jet
is disturbed by the wall and, due to the impossibility to entrain the xed wall, the
jet bends towards the wall. With two parallel jets a similar behaviour can be seen.
198 J.U. SCHLTER AND T. SCHNFELD
Figure 19. Test case of Toy [30], vorticity distribution at different downstream positions:
x = (a) 0 mm, (b) 20 mm, (c) 40 mm, (d) 60 mm, (e) 80 mm, (f) 100 mm, (g) 150 mm,
(h) 200 mm, (i) 300 mm. Initially two counterrotating vortex pairs develop. Due to the Coanda
effect both jets attach each other. The two middle vortices are quenched and parted in four
smaller vortices, one pair above the main CVP, one below, close to the wall.
Here, in the zone between the two jets the entrainment is disturbed and the jets
bend towards the other one as each jet tries to entrain the other.
In the case of the twin jet in crossow this causes a problem for the two inner
vortices. The two vortices are quenched (Figure 19c) and divided in two vortex
pairs (Figure 19d), one pair (2a and 3a) above the remaining two outer vortices
(1 and 4) and one pair below, close to the wall (2b and 3b). The four small
vortices spin fast to keep the torque of the initial two (inner) vortices constant. The
strong gradients in these small, fast spinning vortices lead to higher dissipation
rates and a shorter life span of the vortices. In the far eld (Figure 19i) the two
originally outer vortices survive and form one single CVP. There remains a pair of
small vortices close to the wall, though.
LES OF JETS IN CROSS FLOW 199
Figure 20. Fuel injection on a part of the vane, isosurface: mixture fraction, isocontours:
pressure, black arrows: velocity vectors. Inside the vane is the fuel supply leading to the upper
and lower injector.
7. LES of Fuel Injection in a Gas Turbine
The lack of measurements in a real burner geometry to determine the accuracy
were compensated by the investigation of the JICF. The similar ow conditions
of the JICF and the fuel injectors on the vane in the industrial application gave
condence, that the computations of the burner geometry predicts the ow behav-
iour with sufcient precision, that allows to make statements about mixedness and
vortex formations inside the burner. A full analysis of our LES computation of the
Siemens gas turbine burner is beyond the scope of this article, and we explain only
the main steps of the analysis.
Our rst approach in examining the burner was to have a look at a pair of
injectors on the vane in a homogeneous outer air ow (Figure 20). This gave us
the possibility to determine fuel mass ux rates through the injectors and to in-
vestigate vortex formations around the vane, which interact with the fuel injecting
jets. The computation of a full vane in homogeneous surrounding ow (Figure 21)
gave insight about interactions between adjacent jets, especially to the merging
mechanism of adjacent jets, which lead to unmixedness.
The nal step, the computation of a segment of the burner has been done re-
cently but for reasons of condentiality we are not able to present results of this
computation in this article.
200 J.U. SCHLTER AND T. SCHNFELD
Figure 21. Visualization of fuel injection on the vane.
Figure 22 shows a probability density function (PDF) of mixture fraction down-
stream of an injector. There is a peak at 0 (pure air) and another at 0.075. This
means, there are pockets of air alternating with pockets of fuel. The mean value
does not give any useful information, because the probability that the uid is in a
state of the mean value is quite improbable. Furthermore the mean value suggests,
that the uid at this position is perfectly mixed, which is certainly not the case at
every instant. This sort of PDF in the combustion chamber causes problems due
to the fuel-rich combustion at certain instances leading to higher NOx production.
Visualizations of the LES computation gives insight into the ow to correlate the
unmixed pockets to vortex formations around the vane. The knowledge of the rea-
sons for unmixedness in the mixing section of the burner gives the possibility for
designing engineers to improve the mixedness in the gas turbine burner.
8. Conclusions
The present LES computations have proven the ability of reproducing the main
features of jets in cross ow on reasonably coarse grids.
The necessity of including the jet pipe and the plenum chamber to the com-
puted domain became evident. The proper description of the wall boundary layer
is required to obtain the correct jet trajectory. The low pressure region downstream
of the nozzle inuences the jet trajectory and needs a high spatial discretization as
well.
The general ow characteristics, like the development of the counterrotating
vortex pair, the horseshoe vortex, the jet shear layer vortices and the merging of
LES OF JETS IN CROSS FLOW 201
Figure 22. Example of a probability density function of mixture fraction downstream of an
injector.
two adjacent jets are reproduced. Furthermore some light could be shed into the
merging mechanism of two adjacent jets.
The subgrid turbulence model inuences the jet trajectory. For ows, that are
strongly affected by the wall boundary layer, the Filtered Smagorinsky model
showed better results. The nature of the JICF, to be determined by large scale
motions, makes it possible for the LES approach to obtain results with a good
agreement, even with simple subgrid models.
The LES approach has been applied to a gas turbine burner to obtain informa-
tion about momentum eld and mixing. The computations gave valuable insight in
the events in the burner. This shows the capability of the LES approach to be used
as research tool in the design process of an industrial gas turbine burner.
Acknowledgement
This work has been carried out with the support of the gas turbine department of
Siemens Power Generation KWU (Combustion).
References
1. Alvarez, J., Jones, W.P. and Seoud, R., Prediction of momentum and scalar elds in a jet in cross
ow using rst and second order turbulence closures. In: Computational and Experimental
Assessment of Jets in Cross Flow, No. CP-534. AGARD (1993).
202 J.U. SCHLTER AND T. SCHNFELD
2. Andreopoulos, J., Measurements in a jet-pipe ow issuing perpendicularly into a cross stream.
Journal of Fluids Engineering 104 (1982) 493499.
3. Andreopoulos, J., Heat transfer measurements in a heated jet-pipe ow issuing into a cold cross
stream. Physics of Fluids 26(11) (1983) 32013210.
4. Andreopoulos, J., On the structure of jets in a crossow. Journal of Fluid Mechanics 157 (1985)
163197.
5. Andreopoulos, J. and Rodi, W., Experimental investigation of jets in a crossow. Journal of
Fluid Mechanics 138 (1984) 93127.
6. Baker, C.J., The laminar horseshoe vortex. Journal of Fluid Mechanics 95(2) (1979) 347367.
7. Benz, E., Beeck, A., Fottner, L. and Wittig, S., Analysis of cooling jets near the leading edge of
turbine blades. In: Computational and Experimental Assessment of Jets in Cross Flow, No. CP-
534. AGARD (1993).
8. Broadwell, J.E. and Breidenthal, R.E., Structure and mixing of a transverse jet in incompress-
ible ow. Journal of Fluid Mechanics 148 (1984) 405412.
9. Brown, G.L., Structures of turbulent shear ows: A new look. AIAA Journal 14 (1976) 1349
1357.
10. Brown, G.L. and Roshko, A., On density effect and large structure in turbulent mixing layers.
Journal of Fluid Mechanics 64 (1974) 775816.
11. Chiu, S.H., Roth, K.R., Margason, R.J. and Tso, J., A numerical investigation of a subsonic
jet in a cross ow. In: Computational and Experimental Assessment of Jets in Cross Flow,
No. CP-534. AGARD (1993).
12. Ducros, F., Nicoud, F. and Schnfeld, T., Large Eddy Simulation of compressible ows on
hybrid meshes. In: Proceedings of the Eleventh Symposium on Turbulent Shear Flows (1997)
pp. 28-128-6.
13. Fric, T.F. and Roshko, A., Vortical structure in the wake of a transverse jet. Journal of Fluid
Mechanics 279 (1994) 147.
14. Haven, B.A. and Kurosaka, M., Kidney and anti-kidney vortices in crossow jets. Journal of
Fluid Mechanics 352 (1997) 2764.
15. Hussain, A.K.M.F., Coherent structures Reality and myth. Physics of Fluids 26(10) (1983)
28162849.
16. Hussain, A.K.M.F., Coherent structures and turbulence. Journal of Fluid Mechanics 173 (1986)
303356.
17. Kelso, R.M., Lim, T.T. and Perry, A.E., An experimental study of round jets in crossow.
Journal of Fluid Mechanics 306 (1996) 111144.
18. Kelso, R.M. and Smits, A.J., Horseshoe vortex systems resulting from the interaction between
a laminar boundary layer and a transverse jet. Physics of Fluids 7(1) (1995) 153158.
19. Margason, R.J., Fifty years of jet in cross ow research. In: Computational and Experimental
Assessment of Jets in Cross Flow, No. CP-534. AGARD (1993).
20. Perry, A.E., Kelso, R.M. and Lim, T.T., Topological structure of a jet in cross ow. In:
Computational and Experimental Assessment of Jets in Cross Flow, No. CP-534. AGARD
(1993).
21. Schnfeld, T. and Rudgyard, M.A., Steady and unsteady ows simulations using the hybrid
ow solver AVBP. AIAA Journal 37(11) (1999) 13781385.
22. Schnfeld, T. and Rudgyard, M.A., COUPL and its use within hybrid mesh CFD applica-
tions. In: Ecer, A., Emerson, D., Periaux, J. and Satofuka, N. (eds), Proceedings of the 10th
International Conference on Parallel CFD 98, Hsinchu, Taiwan, May 1998. Elsevier Science
Publishers, Amsterdam (1998) pp. 433440.
23. Smagorinsky, J., General circulation experiments with the primitive equations, I. The basic
experiment. Monthly Weather Review 91(3) 99152.
LES OF JETS IN CROSS FLOW 203
24. Smith, S.H., Hasselbrink, E.F., Mungal, M.G. and Hanson, R.K., The scalar concentration eld
of the axisymmetric jet in crossow. In: 34th Aerospace Sciences Meeting and Exhibit, Reno,
Nevada. AIAA Paper No. AIAA 96-0198 (1996).
25. Smith, S.H., Lozano, A., Mungal, M.G. and Hanson, R.K., Scalar mixing in the subsonic jet in
cross ow. In: Computational and Experimental Assessment of Jets in Cross Flow, No. CP-534.
AGARD (1993).
26. Smith, S.H. and Mungal, M.G., Mixing, structure and scaling of the jet in crossow. Journal of
Fluid Mechanics 357 (1998) 83122.
27. Smith, S.H., The scalar concentration eld of the axisymmetric jet in cross ow. Ph.D. Thesis,
Stanford University, High Temperature Gas Dynamics Laboratory, HTGL Report No. T-328
(1996).
28. Sutton, O.G., A theory of eddy diffusion in the atmosphere. Proceedings of the Royal Society
of London, Ser. A CXXXV (1932) 143165.
29. Sykes, R.I., Lewellen, W.S. and Parker, S.F., On the vorticity dynamics of a turbulent jet in a
crossow. Journal of Fluid Mechanics 168 (1986) 393413.
30. Toy, N., Savory, E. and McCusker, S., The interaction region associated with twin jets and
a normal cross ow. In: Computational and Experimental Assessment of Jets in Cross Flow,
No. CP-534. AGARD (1993).
31. Wulff, A. and Hourmouziadis, J., Technology review of aeroengine pollutant emissions.
Aerospace and Science 8 (1997) 557572.
32. Yuan, L.L., Street, R.L. and Ferziger, J.H., Large eddy simulations of a round jet in crossow.
Journal of Fluid Mechanics 379 (1999) 71104.
33. Yuan, L.L., Large eddy simulations of a jet in crossow. Ph.D. Thesis, Department of
Mechanical Engineering, Stanford University (1997).

Vous aimerez peut-être aussi