Sunteți pe pagina 1din 7

Proceedings

of the 37th
& 4th International
Conference
on Mechanics
Fluid Mechanics
Fluid
Power
Proceedings of the
37th National
& 4National
th International
Conference
on Fluid
andand
Fluid
Power
December 16-18, 2010, IIT Madras, Chennai, India.
December 16-18, 2010, IIT Madras, Chennai, India
FMFP10 - CR
- 06
FMFP2010
547

OPTIMIZATION OF THE FUEL MANIFOLD LOCATIONS INSIDE A JET ENGINE


AFTERBURNER
N Maheswara Reddy
Dept of Aerospace Engineering
Indian Institute of Technology
Madras
Chennai 600 036, India
mahesh_jayam@yahoo.com

E G Tulapurkara
Dept of Aerospace Engineering
Indian Institute of Technology
Madras
Chennai 600 036, India
egt@ae.iitm.ac.in

V Ganesan
Dept of Mechanical Engineering
Indian Institute of Technology
Madras
Chennai 600 036, India
vganesan@iitm.ac.in

engine producing the same maximum thrust. The


afterburner is located in a jet pipe downstream of
turbine. The real flow inside afterburner is highly
complex due to presence of struts, diffuser, fuel
manifolds, flame stabilizer, screech holes and
cooling holes (Fig. 1). The flow is turbulent in
nature and has high-pressure gradients and
recirculation. The afterburner works in two modes afterburning and non-afterburning. Analyses of
non-reacting flow and reacting flows have been
presented by Maheshwara Reddy et al. (2008,2009).
In this paper we use the computation technique to
optimize the location of the manifold in the
afterburner.
Ravichandran and Ganesan (1994) conducted
experimental and numerical investigations of threedimensional flow fields in an isothermal model
after burner. Ravichandran and Ganesan (1996)
computed the three-dimensional flow fields and
confirmed the capability of a code based on
SIMPLE algorithm to handle this flow situation.
Most of the previous investigations e.g. Rajkumar
and Ganesan (2003) are restricted to investigation
of simple geometries and ignore the effect of struts,
V-gutter brackets etc and are limited to 600 sector of
afterburner.
As regards the investigations of reacting flows,
Zhang et. al. (1987) developed a numerical method
and a computer code for modeling afterburner spray
combustion based on the Navier- Stokes equations
and finite difference method. The problem was
considered as two-dimensional, axisymmetric flow.
Modified k-! and hybrid eddy break up models
were used to account for the turbulence and

ABSTRACT
An afterburner is used in the engines of military
aircraft as a thrust-augmenting device during
operations like take off, steep climb, quick
acceleration and steep turns. The afterburner is
located in a jet pipe downstream of turbine. The
flow inside an afterburner is highly complex due to
presence of struts, diffuser, fuel manifolds, flame
stabilizer, screech holes and cooling holes. The
flow is turbulent in nature and has high-pressure
gradients and recirculation.
The locations of the fuel manifolds are crucial for
the combustion to be complete and for low levels of
CO emissions. In the present investigation, an
existing afterburner with four fuel manifolds is
considered. The locations of the two inner
manifolds are varied and analysis is carried in ten
different cases. The FLUENT software, with RNG
k-! model for turbulence and PDF model for
combustion is used. The air-fuel ratio is taken as
25:1. The results show that with suitable locations
of the inner manifolds, the concentration of CO
could be reduced from nearly 3700 ppm to 56 ppm.
Keywords : Jet engine; afterburner; reacting flow;
optimization
INTRODUCTION
An afterburner is used in the engines of military
aircraft as a thrust-augmenting device during
operations like take-off, steep climb, quick
acceleration and steep turns. The advantage of an
afterburner is that the weight of the augmented
engine is much less than the weight of turbojet

to the measured values. The results showed close


agreement between the model and the test results
for NOx and CO levels and lack agreement on CO2
due to poor sampling method in combustor. There
is a reduction in the average values of NOx with the
progressive fuel admission in afterburner and the
predicted levels of CO2 and CO match with the
estimated values.
Unaune and Ganesan (2004) carried out reacting
flow analysis in an afterburner using computational
fluid dynamics (CFD). A 600 sector of afterburner is
considered for the analysis. Numerical calculations
are performed using SIMPLE algorithm and RNG
k-!" #$%&'" ()" *)&%" +$," -*,.*'&/0&1" 23&" 0$#.*)-($/"
phenomenon is modeled using probability density
function (PDF) approach. The study revealed that
the pressure loss reduces for leaner mixtures
whereas the pattern factor increases with increase in
air-fuel ratio. Most of the previous investigations
e.g. Rajkumar and Ganesan (2003) and Unaune and
Ganesan (2004) are restricted to investigation of
simple geometries and ignore the effect of struts in
diffuser, V-gutter brackets, metering holes in
bypass region etc. The effect of core inlet swirl is
also neglected. Unaune and Ganesan (2004) have
not considered the effect of compressibility in
reacting flow analysis. In the present analysis all
these effects are taken into account. FLUENT
software is used as a computational platform.

combustion. SIMPLER (Semi-Implicit Method for


Pressure linked Equations Revised) method
(Patankar (1980)) to solve gas phase conservation
equations and the DROPLET procedure (Patankar
(1980)) for droplet equations were adopted. A solid
triangular bluff-body was considered as flame
holder.
Shu-Hao Chuang and Jiunn-Shean Jiang (1990)
analyzed the diffusion flame of an afterburner as a
function of the air-fuel ratio by employing the
SIMPLE-C algorithm and the turbulence k-! model.
Numerical analysis assumed the flow to be twodimensional axisymmetric, steady state, one-step
chemical reaction with infinitely fast reaction rate
and single phase. Results of the analysis conducted
for various air-fuel ratios showed that the
recirculation zone pattern behind the flame holder is
weakly dependent on air-fuel ratio. Better
combustion efficiency of an afterburner with two
V-gutter flame holders is obtained in a slightly fuellean condition.
Ramana Reddy et.al. (1994) carried out numerical
study of four different combustion models, namely
Arrhenius model, eddy break-up model, k-!-g
model and eddy break-up model with predefined
combustion zone for afterburner combustion. The
results of the flow simulation behind a V-gutter by
the code PHOENICS were in good agreement with
the available experimental data and hence the code
has been used further to simulate afterburner
combustion. The results showed that the Arrhenius
model predicts high reaction rates everywhere and
initiates combustion even before the V-gutter. The
eddy break model and k-!-g model predict
combustion near the V-gutter, the predictions were
in accordance with the expected trends. The eddy
break-up model, which does not need an extra
conservation equation for g, as in case of the k-!-g
model, yields nearly similar results, and can be used
for simulating afterburner combustion.
Haran et.al (1996) attempted to measure and predict
the emission levels during afterburning. The
existing program for constant pressure combustion
calculation in hydro-carbon-air system was used
with modifications for predicting emissions from
TF30-P-3 engine afterburner. The modifications
were based on combustor and afterburner models
and the predicted emission results were comparable

GEOMETRICAL MODELING AND GRID


GENERATION
Figure 1 shows the geometry of the afterburner with
details such as core inlet, bypass inlet, struts (8
numbers), fuel manifolds (4 numbers), flame
stabilizer, screech holes, cooling holes and C-D
nozzle. The flame holder has 18 top radial gutters
and 6 bottom radial gutters connected by a ring
gutter. The radial gutters are inclined at 150 to
vertical and the include angle of V in all gutter is
300. Five hundred and sixty seven holes of 6.4 mm
diameter are present in screech rings and two
hundred and thirty four holes of same diameter are
present in four cooling rings. Twelve metering
holes are present to maintain flow through screech
holes and cooling holes during afterburning and
non-afterburning modes. Ten block unstructured
grid has been created using GAMBIT software.

1.719 million tetrahedral cells are used over entire


domain after the grid independence study involving
1.425, 1.719 and 2.105 million cells. Fine mesh is
employed near the flame stabilizer where large
gradients in flow variables are expected. Actual
dimensions are not given owing to proprietary
nature of geometry.

Fig. 2 Location of Manifolds ahead of V-gutter

y
x
Fig. 1 1800 sector of computational domain with 1.7
million tetrahedral cells

Fig.3 Plot of Velocity Vectors (m/s) at mid plane

LOCATIONS OF FUEL MANIFOLDS


Fuel is injected against exhaust gas flow coming
from turbine through the small diameter orifices
located on ring manifolds such that the liquid jet
enters the gas stream in a transverse direction.
During the penetration process, the air stream tears
the jet apart and small droplets are generated with
diameters typically in the range of 10-100 microns.
Heat transfer from the hot gas stream vaporizes the
small droplets.
Four manifolds are present for supplying fuel and
these are located ahead the V-gutter (cyan color) as
shown in Fig. 2. Manifold number 1 (green color)
is the innermost manifold and is located to ensure
fuel supply to all inner radial gutters. Manifold
number 2 (purple color) is located right behind the
ring gutter to ensure the fuel supply to the wake
region of the ring gutter. Manifolds. No. 3 and No.
4 are located so as to ensure fuel supply to the outer
radial gutters and the space in between. Figure 3
shows the locations of the fuel manifolds in the
midplane of the afterburner.

BOUNDARY CONDITIONS
Core Inlet: Mass flow rate (for 1800 sector) = 43.37
kg/s, Total temperature = 1040 K,
Turbulence intensity = 10 %
Bypass inlet: Mass flow rate (for 1800 sector) = 9.2
kg/s, Total temperature = 517 K, Turbulence
intensity = 5 %
Outlet: Sea level atmospheric pressure = 1.01353
bar
Walls: Adiabatic no-slip boundary condition is
applied on struts, fuel manifolds, gutter and liner.
Interior: To simulate the flow through screech holes
and cooling holes, they are specified as interior
boundary condition (see FLUENT Users guide,
section 6.1.1).
COMPUTATIONAL DETAILS
The flow has been simulated by solving the time
averaged conservation equations for mass,
momentum and energy along with RNG k-! model
and PDF model for combustion. The equations are
omitted for the sake of brevity [see Maheswara
Reddy (2004) for details]. This set of equations has

been solved by finite volume approach using


FLUENT 6 solver.
VALIDATION
For validation purpose the experimental results in a
model afterburner (Ravichandran and Ganesan,
1994) are computed. The model afterburner has a
conical diffuser with an extension pipe of diameter
(D) 200 mm. The V- gutter is placed inside the
pipe. The bulk mean velocity U0 (9.8 m/s) is the
average axial velocity at the inlet of the test section.
Uo and D/2 are used as reference quantities.
1,59,065 tetrahedral cells are employed.
Comparison between computed and experimental
data for the axial velocity profiles in the 30" plane
at distances of x = 70 mm, 80 mm, 90 mm, 105
mm, 120 mm and 240 mm for non-reacting case are
shown in Fig.4. Reasonable agreement between
predicted and experimental data (black color
squares) can be seen. Figure 4 also shows the
results of Suresh and Ganesan (2001) who
employed RNG k-!"#$%&'1"It is seen that RNG k-!"
model (blue color line) gives closer comparison
with experiments. With this confidence the same
code has been used for computation of flow in a
real afterburner and to examine the optimum
location of the fuel manifolds.

Fuel used

: C12H23 (Kerosene)

Air fuel ratio


Sauter mean
diameter

: 25:1

Type of injection

Injection
temperature

: 1000C

: 60 m
Upstream axial 300 cone
spray

Maheswara Reddy et. al. (2008) present results for


non-reacting flow in the afterburner. Maheswara
Reddy et. al. (2009) present results for reacting flow
in the same configuration. Figures 5 and 6 show the
fuel mass flow fraction and total temperature in the
midplane of the original configuration.

Fig. 5 Plot of Fuel Mass Fraction at mid plane


Original configuration

Fig. 4 Mean axial velocity variation along the


radius of pipe at 30o plane
Fig. 6 Plot of Total Temperature (K) at mid plane
for 25:1 Air-fuel ratio Original configuration

RESULTS AND DISCUSSION


The geometry of the afterburner is described in
earlier sections. The fuel injection parameters
specified for 1800 sector of afterburner are:

Figure 7 presents the mass fraction of CO at various


stations in the afterburner. Numerical results
indicate that the mass fraction of CO at the outlet

was 3722 ppm, which was felt to be rather high.


Hence, attempts are made in this investigation to
reduce it.

of injections from the no.2 manifold clearly. Sixth


position is found to be the optimum among all the
selected locations. Some details are available in
Maheswara Reddy (2004).

Fig. 7 Plot of CO Mass Fraction at various planes


along length of Afterburner for 25:1 Air-fuel ratio
Original configuration
It is seen from Fig.5 that more fuel is trapped near
the axis. Figure 3 shows that fuel manifolds 1 and 2
are located in a region where the flow turns towards
the axis. To gain more insight, the particle trances
from various manifolds are examined. Figure 8
shows the particle traces for manifolds 1 and 2 and
Fig.9 for those from 3 and 4. It is seen that the fuel
injected from manifolds 1 and 2 drift towards
afterburner axis as the flow turns towards the axis
due to the curved shape of the diffuser. The fuel
injected from manifolds 3 and 4 moves parallel to
the afterburner axis (see also Fig.3). These
observations suggest that manifolds 1 and 2 need to
be shifted from their original positions.
Six different locations of No.2 manifold from the
original position were selected for the analysis,
which include (i) 5 mm radial shift i.e. mean
diameter is increased by 5 mm, (ii) 10 mm radial
shift i.e. mean diameter is increased by 10 mm, (iii)
10 mm axial shift, (iv) 20 mm axial shift, (v) 20
mm axial and 5 mm radial shift i.e. mean diameter
is increased by 5 mm from the 10 mm axial shift
and (vi) 20 mm axial and 10 mm radial shift i.e.
mean diameter is increased by 10 mm from the 20
mm axial shift. These locations are shown in
Fig.10.
The fuel distribution for all the above cases is
studied. Fuel is not injected from the no.1 manifold
during the above analysis in order to see the effect

Fig.8 Particle Traces from manifolds 1 and 2

Fig. 9 Particle Traces from manifolds 3 and 4

Fig.10 Six different locations of manifold No.2


Four different locations of No.1 manifold were
selected for the analysis which include (i) 10 mm
axial shift, (ii) 20 mm axial shift, (iii) 10 mm axial

Table 1Mass weighted average values of various


parameters at different sections along the length of
afterburner

and 10 mm radial shift i.e. mean diameter is


increased by 10 mm from the 10 mm axial shift and
(iv) 20 mm axial and 10 mm radial shift i.e. mean
diameter is increased by 10 mm from the 20 mm
axial shift. These locations are shown in Fig. 11.

Total
pressure Original
loss (%)
configuration
In core region
25.59
In bypass region
15.7
Velocity (m/s) at
916
nozzle exit
Mach number at
1.316
nozzle exit
Total temperature (K)
At core inlet
1061
Just next to struts
1061
After
the
1819
manifolds
After the gutter
1954
End of screech
1923
holes
Just at entry to
1890
nozzle
At outlet
1810
CO mass fraction (ppm)
At core inlet
0.06
Just next to struts
0.0325
After
the
21138
manifolds
After the gutter
11734
End of screech
8699
holes
Just at entry to
6062
nozzle
At outlet
3722

Fig. 6.11 Four different locations of manifold No.1


The fuel distribution for all the four cases has been
studied. The optimized location of No.2 manifold is
taken into account in this study. Among four
locations the fourth position is found to be the
optimum location.
Detailed computations were carried out with
manifolds 1 and 2 in their optimum positions.
Figures 12, 13 and 14 show the fuel mass fraction,
total temperature and CO mass fraction in mid
plane respectively. The mass weighted average
values of various parameters at different sections
along the length of afterburner and the nozzle are
presented in Table 1. It is observed that the velocity
and total temperature at the nozzle exit are slightly
higher for the optimized configuration as compared
to the original one. The CO mass fraction has
dramatically decreased to 56 ppm from 3722 ppm
for the original.

Optimized
configuration
25.3
15.4
924
1.314
1061
1061
1838
1995
1959
1929
1843
0.054
0.029
18938
9088
4831
1600
56

Fig.13 Plot of Total Temperature (K) at mid plane


with optimized locations of manifolds No.1 and
No.2

Fig. 12 Plot of Fuel Mass Fraction at mid plane


with optimized locations of manifolds No.1 and
No.2

Maheswara Reddy, 2004. Analysis of non-reacting


and reacting flows in an aero gas turbine engine
afterburner. MS Thesis, IIT Madras.
Maheswara Reddy, N., Tulapurkara, E. G. and
Ganesan, V., 2008. Aerodynamic analysis of flow
inside an aero engine afterburner. Proc of 35th Nat.
Conf on Fluid Mechanics and Fluid Power, Dec.1113, 2008, Bangalore, pp.733-738
Maheswara Reddy, N., Tulapurkara, E. G. and
Ganesan, V., 2009. Analysis of the reacting flow
inside a jet engine afterburner. Proc of 36th Nat.
Conf on Fluid Mechanics and Fluid Power, Dec.1719, 2009, Pune, 84-ANT-pp.17.1-17.9.
Patankar, S. V., 1980. Numerical heat transfer and
fluid flow. Hemisphere Publishers, Washington D.
C.
Rajkumar, M. and Ganesan, V., 2003. Analysis of
non-reacting flow in an aircraft gas turbine
afterburner using finite volume method. Indian
Journal of Engineering & Material Science, 10,
341-352.
Ramana Reddy, M.V., Shembarkar, T. R. and J. J.
Isaac, 1994. Numerical simulation of combustion in
an afterburner. NCABE94, 40-50
Ravichandran M and Ganesan V., 1994.
Aerodynamic flow investigations in an isothermal
model of an afterburner. Experiments in Fluids 17,
59-67
Ravichandran M and Ganesan V., 1996. Isothermal
flow field modelling of gas turbine afterburners. IE
(I) Journal- MC, 77, 67-75.
Shu-Hao Chuang and Jiunn-Shean Jiang, 1990.
Diffusion flame analysis of an afterburner as a
function of the air-fuel ratio, International Journal
for Numerical Methods in Fluids, 11, 303-316.
Suresh D and Ganesan V., 2001. Modeling of
isothermal and reacting flows in an afterburner. 2nd
international SAE India Mobility Conference, IIT
Madras, Chennai, 173-178.
Unaune, S. V. and V. Ganesan., 2004, Flow field
studies in an afterburner. IE (I) Journal-MC, 84,
165-170.
Zhang, X. and H. Chiu., 1987. Numerical modeling
of afterburner combustion. International Journal of
Turbo and Jet Engines, 4, 251-262

Fig.14 Plot of CO Mass Fraction at mid plane with


optimized locations of manifolds No.1 and No.2
Comparing Figs.5 & 12, 6 &13 and 7 & 14, the
following can be observed as the result of changed
location of manifolds 1 and 2. The influence of the
curved walls of the diffuser is reduced and the fuel
injected from the manifolds 1 and 2 also moves
nearly parallel to the afterburner axis. Hence the
fuel mass fraction is almost uniform over most of
afterburner. The combustion is almost complete and
mass fraction of CO reduced to minimum.
CONCLUSIONS
The reacting flow inside a jet engine afterburner is
analyzed using FLUENT software with RNG k!#model of turbulence and PDF model for
combustion. With given locations of the fuel
manifolds, the CO mass fraction was found to be
high. It was noticed that the curved wall of the
diffuser was influencing the trajectory of the fuel
injected from the inner manifolds. When the
locations of the inner two manifolds were shifted
so that the fuel mass fraction is almost uniform
along the afterburner, the combustion is almost
complete and CO mass fractions reduces to a
minimum.
REFERENCES
FLUENT 6, Users Guide Volumes, Fluent
Incorporated.
Haran, A. P., Parminder Singh, Antonio Davis, and
V. Suresh. 1996. Aerodynamic design tool for
afterburner. NCABE 96.

S-ar putea să vă placă și