Sie sind auf Seite 1von 9

Cowl and Cavity Effects on Mixing and Combustion

in Scramjet Engines
Sang Hun Kang,

Yang Ji Lee,

and Soo Seok Yang

Korea Aerospace Research Institute, Daejeon 305-333, Republic of Korea


and
Michael K. Smart

and Milinda V. Suraweera

University of Queensland, Brisbane, Queensland 4072, Australia


DOI: 10.2514/1.48818
To investigate the supersonic combustion patterns in scramjet engines, a model scramjet engine was tested in the
T4 free-piston shock tunnel. The test model had a rectangular intake, which compressed the freestreamowthrough
a series of four shock waves upstream of the combustor entrance. A cavity ame holder was installed in the
supersonic combustor to improve ignition. The freestream test condition was xed at Mach 7.6, at an altitude of
31 km. This experimental study investigated the effects of varying fuel equivalence ratios, the inuence of the cavity
ame holder, and the effects of cowl shape. As a result, supersonic combustion was observed at equivalence ratios
between 0.11 and 0.18. Measurements indicated that the engine thermally choked at a fuel equivalence ratio of 0.40.
Furthermore, the cavity ame holder and the W-shaped cowl showed improved pressure distribution due to greater
reaction intensity. With the aid of numerical analysis, the cavity and the W-shaped cowl are shown to be effective in
fuelair mixing.
Nomenclature

h = average heat transfer coefcient,


_ m
f
= fuel mass ow rate,
P
f
= measured pressure in plenum chamber,
P
o;f
= nal pressure in fuel reservoir,
P
o;i
= initial pressure in fuel reservoir,
(P
t
)
noz
= nozzle supply pressure,
P
o
= freestream pressure,
R
f
= ideal gas constant of fuel
T
o;i
= initial temperature in fuel reservoir
t
f
= nal time
t
i
= initial time
V
o
= volume of fuel reservoir (1:66 10
3
m
3
)
VA
y
H
2
= variance of normalized fuel mass ux
= fuel calibration constant
= fuel equivalence ratio
I. Introduction
T
HE scramjet engine is one of the most promising candidates for
future transport systems because of its high specic impulse and
light weight [1]. Even though a combined-cycle rocket engine or gas
turbine would be more practical for transportation [2,3], the
realization of supersonic combustion is still a critical aspect of
hypersonic airbreathing engine development, and it has been
extensively studied at the component level [47].
Because of a strong ram effect, a scramjet engine does not require
an air compressor; therefore, its geometry is very simple. However,
inow perturbations due to the intake conguration, such as shock
waves and ow nonuniformity, are difcult to eliminate in the
absence of a compressor. Perturbations in the combustor inow can
inuence the performance of fuel injectors and ame holders.
Therefore, to develop a clear understanding of various phenomena in
a scramjet engine, a free jet test setup is most suitable [810].
In this paper, we investigate the supersonic combustion char-
acteristics of various component congurations using free jet tests of
the model scramjet engine in the T4 free-piston shock tunnel. Results
of the numerical analysis are also presented to characterize the engine
performance.
A scramjet model was designed to be tested in the T4 free-piston
shock tunnel. The design ight Mach number was 7.6, and the ight
altitude was set to 31 km. A rectangular intake with a four-shock-
wave system was employed for a high total pressure recovery and
robust combustion. The intake ramp angles were determined using
LevenbergMarqurdts optimization method and Korkegis criteria
[11,12]. With the installation of the W-shaped cowl, intake
startability was also enhanced. In the combustor, a cavity was in-
stalled for mixing enhancement and ame holding. The perfor-
mance of the supersonic combustor was predicted by the Rayleigh
line theory [13] and the perfectly stirred reactor model [14,15]. For
the design of the cavity ame holder, the ow residence time in the
cavity was determined by the Davis and Bowersox relation [14].
II. Experimental Apparatus
A. T4 Shock Tunnel
T4 is a free-piston driven, reected shock tunnel. It consists of an
annular reservoir, a free piston, a compression tube, a shock tube, a
nozzle, a test section, and a dump tank. An unscored Brightformsteel
primary diaphragm of varying thickness separates the driver gas in
the compression tube fromthe test gas in the shock tube. Asecondary
mylar diaphragm (0.1 mm thick) separates the shock tube section
from the test section. The specications of the T4 shock tunnel are
presented in Table 1 [16,17].
An axisymmetric contoured nozzle, capable of producing
Mach 7.6 ows, is used in the present study. The nozzle exit diameter
is 270 mm, and its length is 1150mm. The nozzle consists of an initial
conical section that can produce an expanded uniform source ow
and a contoured section that can straighten the ow. Based on pitot-
pressure survey results, the test core ow diameter is 200 mm at the
nozzle exit plane and reduced to 140 mmat a downstreamdistance of
500 mm from the nozzle exit plane [18]. The test model is placed
Received 6 January 2010; accepted for publication 17 June 2011.
Copyright 2011 by the American Institute of Aeronautics and Astronautics,
Inc. All rights reserved. Copies of this paper may be made for personal or
internal use, on condition that the copier pay the $10.00 per-copy fee to the
Copyright Clearance Center, Inc., 222 RosewoodDrive, Danvers, MA01923;
include the code 0748-4658/11 and $10.00 in correspondence with the CCC.

Senior Researcher, Aero Propulsion SystemDepartment. Member AIAA.

Head of Department, Aero Propulsion System Department.

Professor, Centre for Hypersonics, Division of Mechanical Engineering.


Senior Member AIAA.

Postdoctoral Research Fellow, Centre for Hypersonics, Division of


Mechanical Engineering; currently Research Engineer, Gexcon. Member
AIAA.
JOURNAL OF PROPULSION AND POWER
Vol. 27, No. 6, NovemberDecember 2011
1169
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

within the core ow by being inserted 275 mm into the nozzle. The
core ow diameter at the cowl capture plane, which is 270 mm away
from the nozzle exit, is 190 mm.
B. Model Installation
Figure 1 shows the conguration of the test model. The ow
turning angles due to the intake ramps and cowls are 12, 15, 12, and
15

. As a result, the freestream Mach 7.6 ow is compressed and


slowed down to a Mach 2.02.3 owat the entrance of the supersonic
combustor. In the combustor, a 3-mm-deep and 9-mm-long cavity is
placed 75.5 mm downstream of the cowls leading edge. Gaseous
hydrogen is injected at an angle of 45

to the local owthrough a row


of four sonic injectors. The injector holes are 2 mm in diameter and
are spaced laterally at 25 mmintervals. For the intake startability, the
W-shaped cowl is used. By cutting out the cowl in a W shape, the
internal contraction ratio of the intake stays within the Kantrowitz
limit [19]. The details of this design are explained in Kang et al. [20].
Figure 2 shows the pressure transducers installed in the test model
within the test section. In this picture, the Mach 7.6 nozzle is tempo-
rarily unequipped to provide a clearer view of the model. In the test
model, static pressures are measured at 32 stations. Kulite
TM
and
PCB
TM
piezoelectric pressure transducers are used to measure the
pressure levels. Kulite XTEL-190M piezoelectric pressure trans-
ducers have an excitation voltage of 10 V and pressure ranges of
070, 0170, and 0700 kPa. High pressure levels, such as pitot
pressures and plenum chamber pressures, are measured by PCB
type 111A26 piezoelectric pressure transducers. The transducers
sensing faces are thermally protected from the ow by 25 m
cellophane disks covering them. Using measurement uncertainty
analysis, the total systematic uncertainties in the pressure measured
by the Kulite and PCB transducers were estimated to be 2:5 and
3:8%, respectively.
Gaseous hydrogen is injected into the scramjet engine combustor
through a rowof four holes. Figure 3 shows the layout and schematic
of the fuel supplysystemwithin the test model. Fuel is injected froma
room-temperature reservoir through a fast-acting solenoid valve. The
reservoir is a coiled Ludwieg tube that keeps the temperature of the
fuel constant at approximately 300 K during injection. The injection
ow is initiated at least 8 ms before the test ow arrival.
The fuel system was calibrated before testing, in order to deter-
mine the mass ow rate of hydrogen as a function of the reservoir
pressure. The calibration procedure for the shock tunnel fuel system
is described in Robinson et al. [17]. The instantaneous mass owrate
of the fuel is given by
_ m
f
=
1

P
(1)=2
o;i
P
(1)=2
f
(1)
where is the experimentally determined fuel calibration constant,
which in turn is given by
Table 1 Specications of T4 free-piston shock tunnel
Description Quantity
Piston mass 92 kg
Compression tube 229 mm inside diameter 26 m long
Shock tube 76 mm ID 10 m long
Nozzles Machs 4, 6, 7, 7.6, 8, and 10
Enthalpy range 2:515 MJ=kg
Pressure range 1050 MPa
Fig. 1 Model scramjet engine conguration.
Fig. 2 Test model installation of the test section.
1170 KANG ETAL.
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

=
R
f
T
o;i
V
o
(P
o;f
P
o;i
)
P
(1)=2
o;i
Z
f
i
P
(1)=2
f
(2)
Figure 4 shows a typical fuel mass owrate change during testing.
As demonstrated in the gure, the fuel mass ow rate remains
constant within 3% of the maximum value during the test interval.
C. Test Conditions
The freestream conditions are xed at a Mach number of 7.6 and
an altitude of 31 km. A detailed list of the ow conditions of the
nominal freestreamis shown in Table 2. With these parameters xed,
the effects of fuel equivalence ratios and varying component cong-
urations were investigated. The tests are summarized in Table 3.
III. Experimental Results
A. Characteristics of Nonreacting Flows
In Fig. 5, the normalized pressure distributions within the model
for nonreacting ow cases are shown. Here, pressure is normalized
by the freestream value P
o
. When fuel is absent, pressure uc-
tuations are observed in the combustor due to shock and expansion
wave reections. However, the mean pressure level is largely
unchanged within the combustor. At the intake, there is a region of
increased pressure near the deection point at x =420 mm. This
pressure rise is indicative of a separation bubble. During the design
phase, intake compression angles were determined using Korkegis
criteria for two-dimensional (2-D) at-plate ramp cases [12].
Nevertheless, according to Korkegi, if three-dimensional (3-D)
effects from the sidewalls are strong enough, separation can occur
even in stable regions [12]. Therefore, effects from the test model
sidewalls are believed to be signicant in this case. However, the size
of the separation bubble is too small to affect the ow along the
second ramp and the combustor. Moreover, at the end of the second
ramp, at x =564 mm, the measured pressure level is the same as the
theoretical prediction. This indicates that the performance of the
intake is acceptable, even if 3-D effects from the sidewalls are
distorting the inow.
As is also evident from Fig. 5, the measured pressure levels in the
case of fuel injection into a nitrogen freestream are similar in
magnitude and trend to the case with no fuel, demonstrating that fuel
injection alone has little inuence on the ow at an equivalence ratio
of 0.12.
B. Characteristics of Reacting Flows
In Fig. 6, pressure distributions within the model for reacting ow
cases are presented. When =0:11 and 0.18, the measured pressure
levels of the fuel-into-air test start to rise above those for the fuel-into-
nitrogen test at approximately 700 mm from the leading edge,
indicating a combustion phenomenon. The increase in measured
pressure levels due to combustion is greater for =0:18, with a
maximum normalized pressure level of P=P
o
~226; 785 mm
downstream of the leading edge. For the case in which =0:40,
pressure levels start to rise upstream of the fuel injection point. The
pressure distribution suggests that the boundary layer has separated
and that there are subsonic regions within the combustion chamber. If
the combustion efciency of the model is assumed to be 0.7,
theoretical analysis based on the Rayleigh line ow[13] predicts that
thermal choking will occur when >0:374.
To further investigate choking, Fig. 7 shows time histories of the
measurements by the rst pressure sensor (Cb1), located just
Fig. 3 Schematic of the fuel delivery system.
Time (mms)
P
r
e
s
s
u
r
e
(
M
P
a
)
M
a
s
s
f
l
o
w
r
a
t
e
(
g
/
s
)
0 1 2 3 4
6
8
10
12
14
16
18
20
0
0.5
1
1.5
2
2.5
3
(P
t
)
noz
Fuel flow
Test time
Fig. 4 Typical fuel mass ow rate change during testing.
Table 2 Detailed conditions of
nominal freestream ow
Parameter Value
P
s
, MPa 11.8
T
s
, K 2595
H
s
, MJ=kg 3.0
P
pitot
, kPa 85.4
P
e
, kPa 1.1
T
e
, K 243

e
, kg=m
3
0.016
U
e
, m=s 2360
M
e
7.56
1.39
Unit Reynolds number 2:4 10
6
Table 3 Test run summary
Shot number Test gas Amount of fuel in
g=s (equivalence ratio, )
Cavity Cowl
1 9481 Air 0.0 (0.0) Y W cowl
2 9486 Air 1.551 (0.11) Y W cowl
3 9487 Air 2.464 (0.18) Y W cowl
4 9483 Air 5.620 (0.40) Y W cowl
5 9489 N
2
1.603 (0.12) Y W cowl
6 9493 Air 0.0 (0.0) N W cowl
7 9494 Air 1.518 (0.11) N W cowl
8 9508 Air 0.0 (0.0) Y Flat cowl
9 9509 Air 1.523 (0.11) Y Flat cowl
Distance (mm)
P
/
P

0 200 400 600 800 1000


0
50
100
150
200
Fuel off, Air
Fuel-N
2
: =0.12
Theoretical Prediction
Separation
Fig. 5 Normalized pressure distribution for nonreacting ows.
KANG ETAL. 1171
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

upstream of the fuel injection point in the combustor. The pressure
distribution for the case in which =0:40 shows severe uctuations
that are not seen at lower fuel equivalence ratios. This suggests the
occurrence of the inlet buzz phenomenon, in which the inlet starts
and unstarts rapidly. This is evidenced by a considerable transient
pressure rise emanating in the aft section of the combustor, moving
upstreamtoward the combustor entrance during testing. The pressure
increase is immediately followed by a rapid reduction in pressure
within the combustor, indicating ow spillage.
C. Effects of Component Variations
For the investigation of cavity and cowl shape effects, the model
conguration was changed, as shown in Fig. 8. The normalized
pressure distributions for different model congurations are shown
in Fig. 9. It can be seen that the conguration with the cavity and the
W-shaped cowl shows higher pressures than the congurations
without them. Figure 10 compares the increases in pressure due to
combustion for various congurations. In the case involving fuel
injection and the cavity, the pressure levels start to surpass those
without fuel injection at the position of pressure sensor Cb4
(x =705 mm). However, in the case without the cavity, pressure
levels starts to rise at approximately 725 mm from the leading edge.
Ben-Yaker and Hanson conrmed that the cavity generates ow
oscillations that enhance mixing in the supersonic shear layer,
creating a recirculation zone that acts as a continuous ignition
source [4].
In the case with the at cowl, the starting point of the pressure rise
due to combustion is almost the same as that in the case with the W-
shaped cowl. However, with the at cowl, the slope of the pressure
rise is smaller. Figure 10 shows that the case with the at cowl
exhibits a lower combustion pressure. As a result, the maximum
normalized pressure is 162, whereas it is 187 with the W-shaped
Distance (mm)
P
/
P

0 200 400 600 800 1000


0
100
200
300
400
Fuel Off
Fuel-N
2
: =0.12
Fuel-Air : =0.11
Fuel-Air : =0.18
Fuel-Air : =0.40
Thermal Choking
Supersonic
Combustion
Fuel Injection
Fig. 6 Normalized pressure distribution for reacting ows.
Time (ms)
P
r
e
s
s
u
r
e
(
k
P
a
)
0 5000 10000 15000
0
100
200
300
400
500
600
Fuel-Air : =0.11
Fuel-Air : =0.18
Fuel-Air : =0.40
Fig. 7 Time history of the rst pressure sensor measurement in the
combustor.
a) Cavity and W cowl
b) No cavity and W cowl
c) Cavity and flat cowl
Fig. 8 Variations of the test model conguration.
Distance (mm)
P
/
P

0 200 400 600 800 1000


0
50
100
150
200
Cavity and W cowl
No Cavity and W cowl
Cavity and Flat cowl
Fig. 9 Effects of component variations on the normalized pressure
distribution for reacting ows ( 0:11).
1172 KANG ETAL.
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

cowl. Therefore, it can be established that the cavity and the
W-shaped cowl have better combustion enhancing effects.
IV. Numerical Analysis
A. Numerical Method
Numerical simulations were carried out to analyze the detailed
ow and combustion characteristics of the model. Computations
were performed using the commercial code Fluent
TM
, along with the
coupled-implicit solver. For turbulent characteristics, the shear-stress
transport (SST) k-! model was used [21]. The SST model is known
for its good prediction of mixing layers and jet ows [22,23].
The wall boundary conditions of scramjet engine combustors are
hard to describe using numerical analysis. Computations by Mitani
and Kouchi [24], done with simple adiabatic or isothermal wall
conditions, showed combustion characteristics that differed from
experimental data. Complicated wall boundary conditions had to be
used to describe the experimental conditions. Star et al. tried various
other wall boundary conditions [25]. In their study, it was concluded
that simple isothermal wall conditions are not suited for the accurate
prediction of the wall pressure distribution. On the other hand, esti-
mation of the heat ux into the wall agreed better with experimental
data.
In the present study, a steady-state convective wall boundary
condition is used for the combustor wall. Because of the high-speed
owin the combustor, the oweld can be expected to reach a steady
state in the short test times of a shock tunnel [8]. For the convective
wall boundary condition, the local Nusselt number for the turbulent
ow is calculated as follows [26]:
Nu
x
=St Re
x
Pr =0:0296Re
4=5
x
Pr
1=3
; 0:6 < Pr <60 (3)
For the given Reynolds number, Prandtl number, and geometry,
the calculated average heat transfer coefcient

h is calculated to be
1368:5 W=m
2
K.
For the numerical analysis of the combustor, 3-D calculations are
carried out. A nine-species 18-step detailed chemical kinetics model
is employed for the simulation of hydrogen combustion. Because of
owsymmetry, only half the combustor along the symmetry plane is
solved for. In the grid resolution study, results with 1,760,000 grid
points, 980,000 grid points, and 520,000 grid points all predicted the
same pressure level. Thus, for reliable computational uid dynamics
(CFD) results, 980,000 grid points are chosen for the combustor
analysis. To provide accurate inow conditions, the computational
domain of the combustor includes a part of the second ramp of the
intake. At the inlet of the computational domain, only two oblique
shock waves fromthe leading edge and the second ramp inuence the
inow. Acomplicated shock and expansion wave reection region is
included in the computational domain. By the 3-D calculation of the
intake compression ramps, the boundary-layer thickness is con-
rmed tobe smaller than 0.3 mmat the computational domain inlet of
the combustor.
For the intake analysis, a 2-D calculation is performed due to the
simplicity of the geometry. Sixty-thousand grid points are distributed
in the computational domain. Here, an adiabatic wall boundary
condition is used. The computational domains of the intake and the
combustor are shown in Fig. 11.
B. Results of Numerical Analysis
1. Comparison of Computational Fluid Dynamics Results
with Experimental Data
In Fig. 12, numerical results for unfueled and fueled cases, with air
as the test gas, are compared with corresponding experimental results
for the setup with no cavity and the W-shaped cowl. As evident in the
gure, there is good overall agreement between the numerical and
experimental test results. The slight pressure differences near the
combustor exit might be due to accumulated error in the estimations
of the pressure and the boundary layer from the inlet to the exit. In
particular, the growing boundary layer could affect the overall
pressure level by changing the owpath area. In the combustor, shock
waves generated by the cowl are continuously reected by the
combustor walls and propagated downstream. In this process,
complex shock wave/boundary-layer interactions occur along the
owpath. However, Knight et al. have pointed out that Reynolds-
Distance (mm)
P
/
P

0 200 400 600 800 1000


0
50
100
150
200
Fuel Off
Fuel-Air : =0.11
Sensor Cb4
a) Cavity and W cowl
Distance (mm)
Distance (mm)
P
/
P

0 200 400 600 800 1000


0
50
100
150
200
Fuel Off
Fuel-Air : =0.11
Sensor Cb4
b) No cavity and W cowl
P
/
P

0 200 400 600 800 1000


0
50
100
150
200
Fuel Off
Fuel-Air : =0.11
Sensor Cb4
c) Cavity and flat cowl
Fig. 10 Normalized pressure distribution changes due to combustion
for different congurations.
Fig. 11 Computational domain.
KANG ETAL. 1173
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

averaged NavierStokes turbulence models are not accurate enough
to predict shock wave/boundary-layer interactions precisely [27].
Nevertheless, overall ow patterns and combustion characteristics
can be observed with such turbulence models [22].
2. Unfueled Air Run
For the unfueled air run, the static temperature contour plots of the
combustor are shown in Fig. 13. As can be seen here, in the no-cavity
case, the high-temperature regions are conned to the local thin
boundary layer at the combustor wall. However, when a cavity is
present, a high static temperature region is formed inside the cavity.
Since this high-temperature region is located 6 mmdownstreamfrom
the fuel injector, the cavity can play the role of a ame holder.
In Fig. 14, static pressure contour plots of the combustor with a
nonreacting oware shown. As seen in the gure, the static pressures
in the conguration with the W-shaped cowl show transverse
directional uctuations. Such pressure distributions could cause
enhancement of mixing and reactions.
Fig. 13 Static temperature contours of the combustor for nonreacting
ow cases for different congurations.
Fig. 14 Static pressure contours of the combustor for nonreacting ow
cases for different congurations.
Fig. 15 Static temperature contours of the combustor for reacting ow
cases with different congurations.
Distance (mm)
P
r
e
s
s
u
r
e
(
k
P
a
)
0 200 400 600 800
-50
0
50
100
150
200
250
CFD, intake (2-D)
CFD, combustor (3-D)
Experiment
a) Unfueled run
Distance (mm)
P
r
e
s
s
u
r
e
(
k
P
a
)
0 200 400 600 800
-50
0
50
100
150
200
250
CFD, intake (2-D)
CFD, combustor (3-D)
Experiment
b) Fueled run
Fig. 12 Numerical analysis and experimental results comparison for
the conguration with no cavity, W cowl.
1174 KANG ETAL.
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

3. Fueled Air Run
For the fueled air run, the static temperature contour plots of the
combustor are shown in Fig. 15. In the conguration with the cavity
and the W-shaped cowl, the overall temperature within the combustor
is higher than in congurations without these two elements. Further-
more, a high static temperature region appears close to the fuel
injectors in this case. Therefore, the ignition length is demonstrably
shorter in the case with the cavity and the W-shaped cowl. In Fig. 16,
static pressure contour plots of the combustor with the reacting ow
are shown. Here, pressure uctuations in the transverse direction are
still valid, even with fuel injection and the reacting ow. Such
pressure uctuations can increase the transverse directional velocity
of the inow and enhance the fuelair mixing. Figure 17 shows
contour slices of the Mach number in the combustor for different
congurations. In all these cases, there are no sonic lines across the
combustor, indicating that combustion occurs in the supersonic
regime. Figure 18 shows the OH mass fraction contour slices in the
combustor. As depicted in the plots, the reactions are initiated just
downstream of the fuel injectors and, even with high axial ow
velocity, spread in the radial direction. Furthermore, the OH mass
fractions in the combustor with the cavity and the W-shaped cowl
show much higher values than those without them. The OH mass
fraction is at its highest value inside the cavity itself, indicating that
the cavity has enhanced the reaction of hydrogen.
4. Numerical Experiment for Mixing
In the previous sections, numerical results indicated the existence
of a high-temperature region within the cavity, which can act as an
igniter. Obviously, this enhances combustion and reduces the
ignition delay time. Another role for the cavity could be as a mixer of
the fuel and the oxidizer. Mixing could also be enhanced by the
transverse directional pressure nonuniformity due to the W-shaped
cowl. Experimental results notwithstanding, it is difcult to deter-
mine how much mixing has been enhanced by the presence of the
cavity and the W-shaped cowl. For this purpose, an additional
numerical experiment was performed. Using the same inow and
fuel conditions as before, the cavity and W-shaped cowl effects on
mixing were reproduced. To observe the mixing phenomenon alone,
combustion is articially turned off. Mixing is monitored by the
variance VA
y
H
2
, dened as
VA
y
H
2
=
1
(u)
2
A
Z
(uy
H
2
uy
H
2
)
2
dA
=
1
(u)
2
A
Z
uy
H
2

1
A
Z
uy
H
2
dA

2
dA (4)
Zero variance denotes a perfectly uniform mixture, and as the
variance value increases, the mixture becomes less uniform. In
Fig. 19, variance values for different congurations are displayed.
When the rst conguration (no cavity, at cowl) is compared with
Fig. 16 Static pressure contours of the combustor for reacting ow
cases with different congurations.
Fig. 17 Mach number contours of the combustor for reacting ow
cases with different congurations.
Fig. 18 OH mass fraction contours of the combustor for reacting ow
cases with different congurations.
KANG ETAL. 1175
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

the second (cavity, at cowl), it can be seen that both variances start at
the same value and decrease. However, the case with the cavityshows
a more severe decrease in the vicinity of the cavity than the case
without the cavity. Consequently, the cavity is shown to be effective
in fuelair mixing. Comparing the rst conguration (no cavity, at
cowl) with the third (no cavity, W-shaped cowl), VA
y
H
2
of the third
conguration shows a lower value than the rst from the very
beginning to the end of the combustor. The cowl is located upstream
of the fuel injection point. This allows the inow, perturbed by the W-
shaped cowl, to enhance the fuelair mixing from the upstream
region. Because of the combined effects of the cavity and the W-
shaped cowl, the fourth conguration (cavity, W-shaped cowl) shows
the lowest variance value in the gure.
V. Conclusions
The effects of different cowl and cavity congurations on fuelair
mixing and combustion in scramjet engines were investigated using
experimental and numerical approaches. The model scramjet engine
was tested with the T4 free-piston shock tunnel. The test model
consisted of a four-shock-wave systemintake, a W-shaped cowl, and
a supersonic combustor with a cavity ame holder. The test was
conducted under Mach 7.6 conditions at a 31 km altitude.
The experimental results showed active combustion for low
( =0:11) and middle ( =0:18) equivalence ratio test cases.
However, in cases of high fuel equivalence ratios, thermal choking
and inlet unstarts were observed. Furthermore, the presence of the
cavity and the W-shaped cowl resulted in greater combustion-
induced pressure increases.
Numerically, the cavity in the combustor was predicted to generate
a hot static temperature region that acted as an ignition source,
improving the mixing characteristics. With the W-shaped cowl, the
static pressure showed transverse directional uctuations and
resulted in improved mixing. Via the combined effects of the cavity
and the W-shaped cowl, earlier ignition and more active combustion
were observed.
On the whole, the cavity and the W-shaped cowl generated ow
perturbations to enhance fuelair mixing and combustion effectively.
However, in real-world applications, ow perturbations result in an
increased drag in most cases [6,28]. Sometimes, the advantages of
enhanced combustioncan be smaller than the drawbacks of increased
drag. Therefore, additional research should be conducted before the
implementation of such devices in real systems.
References
[1] Fry, R. S., A Century of Ramjet Propulsion Technology Evolution,
Journal of Propulsion and Power, Vol. 20, No. 1, 2004, pp. 2758.
doi:10.2514/1.9178
[2] Kanda, T., Tani, K., and Kudo, K., Conceptual Study of a Rocket-
Ramjet Combined-Cycle Engine for and Aerospace Plane, Journal of
Propulsion and Power, Vol. 23, No. 2, 2007, pp. 301309.
doi:10.2514/1.22899
[3] Quinn, J. E., ISTAR: Project Status and Ground Test Engine Design,
AIAA Paper 2003-5235, July 2003.
[4] Ben-Yaker, A., and Hanson, R. K., Cavity Flameholders for Ignition
and Flame Stabilization in Scramjets: Reviewand Experimental Study,
AIAA Paper 1998-3122, July 1998.
[5] Micka, D. J., and Driscoll, J. F., Combustion Characteristics of a Dual-
Mode Scramjet Combustor with Cavity Flameholder, Proceedings of
the Combustion Institute, Vol. 32, 2009, pp. 23972404.
doi:10.1016/j.proci.2008.06.192
[6] Doster, J. C., King, P. I., Gruber, M. R., Carter, C. D., Michael, D. R.,
and Hsu, K. Y., In-StreamHypermixer Fueling Pylons in Supersonic,
Journal of Propulsion and Power, Vol. 25, No. 4, 2009, pp. 885901.
doi:10.2514/1.40179
[7] Smart, M. K., Experimental Testing of a Hypersonic Inlet with
Rectangular-to-Elliptical Shape Transition, Journal of Propulsion and
Power, Vol. 17, No. 2, 2001, pp. 276283.
doi:10.2514/2.5774
[8] Stalker, R. J., Paull, A., Mee, D. J., Morgan, R. G., and Jacobs, P. A.,
Scramjets and Shock Tunnels: The Queensland Experience, Progress
in Aerospace Sciences, Vol. 41, 2005, pp. 471513.
doi:10.1016/j.paerosci.2005.08.002
[9] Kanda, T., Tani, K., Kobayashi, T., Saito, T., and Sunami, T., Mach 8
Testing of a Scramjet Engine with Ramp Compression, Journal of
Propulsion and Power, Vol. 18, No. 2, 2002, pp. 417423.
doi:10.2514/2.5950
[10] Tomioka, S., Hiraiwa, T., Kobayashi, K., and Izumikawa, M.,
Vitiation Effects on Scramjet Engine Performance in Mach 6 Flight
Conditions, Journal of Propulsion and Power, Vol. 23, No. 4, 2007,
pp. 789796.
doi:10.2514/1.28149
[11] Ozisik, M. N., Inverse Heat Transfer, 1st ed., Taylor and Francis,
New York, 2000, pp. 3758.
[12] Korkegi, R. H., Comparison of Shock-Induced Two- and Three-
Dimensional Incipient Turbulent Separation, AIAA Journal, Vol. 13,
1975, pp. 534535.
doi:10.2514/3.49750
[13] John, J. E., Gas Dynamics, 2nd ed., Allyn and Bacon, Boston, MA,
1984, pp. 204217.
[14] Davis, D. L., and Bowersox, R. W., Stirred Reactor Analysis of Cavity
Flame Holders for Scramjets, AIAA Paper 1997-3274, July 1997.
[15] Davis, D. L., and Bowersox, R. W., Computational Fluid Dynamics
Analysis of Cavity Flame Holders for Scramjets, AIAA Paper 1997-
3270, July 1997.
[16] Portwood, T. W., Enhancement of Hydrocarbon Supersonic
Combustion by Radical Farming and Oxygen Enrichment, M.S.
Thesis, Univ. of Queensland, Brisbane, QLD, Australia, 2006.
[17] Robinson, M. J., Rowan, S. A., and Odam, J., T4 Free Piston Shock
Tunnel Operators Manual, Univ. of Queensland, Department of
Mechanical Engineering Rept. 2003-1, Brisbane, QLD, Australia,
2003.
[18] Suraweera, M. V., Mach 7.6 Nozzle Pitot Survey in T4 Shock Tunnel,
Univ. of Queensland, Dept. of Mechanical Engineering Rept. 2006-1,
Brisbane, QLD, Australia, 2006.
[19] Kantrowitz, A., and Donaldson, C., Preliminary Investigation of
Supersonic Diffusers, NACAWRL-713, 1945.
[20] Kang, S. H., Lee, Y. J., and Yang, S. S., Design of Model Scramjet
Engine for Shock Tunnel Test, 9th Asian International Conference on
Fluid Machinery, KFMA Paper 9-106, Seoul, Oct. 2007.
[21] Menter, F. R., Two-Equation Eddy-Viscosity Turbulence Models for
Engineering Applications, AIAA Journal, Vol. 32, No. 8, 1994,
pp. 15981605.
doi:10.2514/3.12149
[22] Choi, J. Y., Ma, F., and Yang, V., Combustion Oscillations in a
Scramjet Engine Combustor with Transverse Fuel Injection,
Proceedings of the Combustion Institute, Vol. 30, 2005, pp. 28512858.
doi:10.1016/j.proci.2004.08.250
[23] Bardina, J. E., Huang, P. G., and Coakly, T. J., Turbulence Modeling
Validation, AIAA Paper 1997-2121, June 1997.
[24] Mitani, T., and Kouchi, T., Flame Structures and Combustion
Efciency Computed for a Mach 6 Scramjet Engine, Combustion and
Flame, Vol. 142, No. 3, 2005, pp. 187196.
doi:10.1016/j.combustame.2004.10.004
[25] Star, J. B., Edwards, J. R., Smart, M. K., and Baurle, R. A., Numerical
Simulation of Scramjet Combustion in a Shock Tunnel, AIAA
Paper 2005-0428, Jan. 2005.
[26] Incropera, F. P., and DeWitt, D. P., Fundamentals of Heat and Mass
Transfer, 4th ed., Wiley, New York, 1996, p. 355.
Distance (mm)
V
A
y
H
2
x
1
0
4
650 675 700 725 750
-1
0
1
2
3
4
5
6
7
8
No Cavity, Flat Cowl
Cavity, Flat Cowl
Nocavity, W Cowl
Cavity, W Cowl
Fuel
Injection
Cavity
Fig. 19 Variance of hydrogen mixing for different congurations.
1176 KANG ETAL.
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

[27] Knight, D., Yan, H., Panaras, A., and Zheltovodov, A., Advances in
CFD Prediction of Shock Wave Turbulent Boundary Layer
Interactions, Progress in Aerospace Sciences, Vol. 39, 2003,
pp. 121184.
doi:10.1016/S0376-0421(02)00069-6
[28] Sunami, T., Itoh, K., Komuro, T., and Sato, K., Effects of Streamwise
Vortices on Scramjet Combustion at Mach 8-15 Flight EnthalpiesAn
Experimental Study in HIEST, 17th International Symposium on Air
Breathing Engines, AIAA Paper 2005-1028, Sept. 2005.
J. Seitzman
Associate Editor
KANG ETAL. 1177
D
o
w
n
l
o
a
d
e
d

b
y

I
N
D
I
A
N

I
N
S
T
I
T
U
T
E

O
F

T
E
C
H
N
O
L
O
G
Y

-

K
A
N
P
U
R

o
n

M
a
y

2
6
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
1
.
4
8
8
1
8

Das könnte Ihnen auch gefallen