Vous êtes sur la page 1sur 13

Copyright 1996, American Institute of Aeronautics and Astronautics, Inc.

AIAA Meeting Papers on Disc, July 1996


A9637098, AIAA Paper 96-2914

Starting characteristics of supersonic inlets


D. M. Van Wie

Johns Hopkins Univ., Laurel, MD

F. T. Kwok

Johns Hopkins Univ., Laurel, MD

R. F. Walsh

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

Johns Hopkins Univ., Laurel, MD

AIAA, ASME, SAE, and ASEE, Joint Propulsion Conference and Exhibit, 32nd, Lake
Buena Vista, FL, July 1-3, 1996
The starting characteristics of a small-scale rectangular inlet with a thick ingested boundary layer were
investigated at nominal Mach 3 conditions. Parameters investigated included Reynolds number, cowl length, and
cowl height. Measurements of the maximum and restart contraction ratios were made. Depending on the test
configuration, the unstarts were classified into two broad categories as either 'hard' or 'soft'. The hard unstarts
appear to occur when the flow at the inlet throat chokes. The soft unstarts occur as large-scale separation develops
within the inlet. The ability of the classical Kantrowitz limit to predict the restart contraction ratio was assessed,
and it was shown to be applicable for the hard unstart/restart configurations. The role of fluid injection upstream
of the unstarted inlet was also assessed. The use of this injection may ultimately lead to improving the starting
characteristics of inlets. (Author)

Page 1

STARTING CHARACTERISTICS OF SUPERSONIC INLETS


D. M. Van Wie,* F. T. Kwok,** and R. F. Walsh*
The Johns Hopkins University
Applied Physics Laboratory
Laurel, Maryland

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

Abstract

Subscripts

The starting characteristics of a small-scale rectangular inlet with a thick ingested boundary layer were investigated at nominal Mach 3 conditions. Parameters investigated included Reynolds number, cowl length, and
cowl height. Measurements of the maximum and restart
contraction ratios were made. Depending on the test configuration, the unstarts were classified into two broad categories as either "hard" or "soft." The hard unstarts appear to occur when the flow at the inlet throat chokes.
The soft unstarts occur as large-scale separation develops
within the inlet. The ability of the classical Kantrowitz
limit to predict the restart contraction ratio was assessed,
and it was shown to be applicable for the hard unstart/
restart configurations. The role of fluid injection upstream
of the unstarted inlet was also assessed. The use of this
injection may ultimately lead to improving the starting
characteristics of inlets.
Nomenclature
A
hc
Lc
rh
M
M
P
Pt
x, y
y
6
6*
0
9C
p

Area
Cowl height
Cowl length
Mass flow
Mach number
Mass-averaged Mach number
Pressure
Freestream total pressure
Cartesian coordinates
Ratio of specific heats
Boundary layer thickness
Boundary layer displacement thickness
Boundary layer momentum thickness
Cowl angle
Density

* Principal staff engineer, senior member AIAA


** Senior staff engineer, member AIAA
^ Associate staff engineer, member AIAA
Copyright American Institute of Aeronautics and
Astronautics, Inc., 1996. All rights reserved.

0
2
4
cl
inj

Freestream
Entrance to cowl
Inlet throat
Wall conditions at cowl lip
Injectant
Introduction

Airbreathing engines that operate at supersonic and


hypersonic speeds require inlets to capture and compress
air for processing by the remainder of the engine. The
goal in the design of any inlet is to define a minimum
weight geometry that provides an efficient compression
process, generates minimum drag, produces nearly uniform flow entering the compressor or combustor, and provides these characteristics over a wide range of flight and
engine operating conditions. For efficient operation and
moderate induced drag, most inlets use a combination of
external and internal compression. The introduction of
internal contraction in an inlet adds complexity in the
design and analysis process in that the starting of the inlet
must be ensured.
For efficient operation, supersonic and hypersonic
inlets must operate in a started mode. The process of inlet
starting and unstarting is well understood at a conceptual
level, although significant details remain to be resolved.
Some variation exists in the very definition of a started
inlet. One convention states that a started inlet is one with
supersonic flow in the inlet throat, but it is well known
that some unstarted inlets can have complex internal
flowfields with a significant fraction of supersonic flow
in the inlet throat. In the present work, the term "started"
is used to denote operation under conditions where flow
phenomena in the internal portions of the inlet do not alter the air capture characteristics of the inlet. (Reduction
in the captured mass flow through the use of bleed holes
or bypass channels is not considered in assessing whether
an inlet is started.) An inlet can be unstarted by either
over-contracting to the point where the flow chokes at the
inlet throat or by raising the back pressure beyond the
level that can be sustained by the inlet.
Currently, a significant uncertainty exists regarding
the conditions under which an inlet will unstart or restart.
Part of this uncertainty is due to the large variety of

1
American Institute of Aeronautics and Astronautics

geometries that have been considered in designing engines. The design of an inlet is strongly affected by vehicle considerations, and a variety of two-dimensional
planar, axisymmetric, and three-dimensional inlet designs

Oswatisch inlet1

HRE-type inlet2

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

have been investigated. The diversity found in inlet designs can be seen in the sample inlets shown in Fig. I.'"8
Preliminary estimates of the internal contraction that
will self-start can be obtained from the Kantrowitz limit.9
This limit is determined by assuming a normal shock wave
at the beginning of the internal contraction and calculating the one-dimensional, isentropic internal area ratio that
will produce sonic flow at the inlet throat. For a perfect
gas, the Kantrowitz limit can be calculated as follows:
70 sweep sidewall
30 sweep sidewall
compression inlet5 compression inlet5

A2]
/KANTROWITZ

l)MJ

Multiple inward-turning
scoop inlet6

7-1

M,
7+1

7-1

7+1

(1)

7+1

For a specific class of inlet geometry and set of


freestream conditions, an inlet will start at a certain contraction ratio. After the inlet starts, variable geometry features of the inlet, if present, can be adjusted to increase
the contraction ratio up to a maximum value, at which

point further increases would cause the inlet to unstart A


summary of published results for a wide variety of inlets
is presented in Fig. 2 for both inlet starting and maximum
contraction ratios. Data are also shown for several inlets
where the operating contraction ratio (rather than maximum contraction ratio) has been reported. These operating contraction ratios were obtained from reported tests
in which no explicit attempt was made to determine a
maximum contraction ratio. When considering the starting contraction ratio, the horizontal scale refers to M 2 .
When the maximum contraction ratio is considered, the
horizontal scale refers to the freestream Mach number,
M0. In this figure, the data are plotted in terms of the
inverse of the contraction ratio (i.e., A^/A2 and A^/AQ). Also
shown is the Kantrowitz limit (Eq. 1) and the isentropic
contraction limit, which can be calculated as follows:

Fig. 1 Examples of scramjet inlets (HRE = hypersonic


research engine).

Mach number, M2 or MQ
' /ISENTROPIC
2(7-1)

(2)

Fig. 2 Starting and maximum contraction ratio (CR)


limits.

American Institute of Aeronautics and Astronautics

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

The allowable starting and maximum contraction ratios both increase (i.e., AJAi and AJ/AQ both decrease) as
the Mach number increases, but the experimental data
show a significant scatter at higher speeds. This scatter is
due to differences in inlet geometry, Reynolds number,
and wall-to-freestream temperature ratio. The data also
suggest that the Kantrowitz limit becomes conservative
at high speeds because a single, normal shock wave is
assumed.
The maximum contraction ratio data shown in Fig. 2
show that higher contraction ratios can be generated as
the Mach number increases. Obviously, the limiting contraction ratio is lower than that for the corresponding isentropic contraction owing to the effects of shock waves
and viscous losses. An empirical fit to the data for this

limit is as follows:
M

M2,

2.5 < M0 < 10

(3)

Because this line lies below the observed maximum contraction data, this equation provides an estimate of the

limit on the achievable contraction ratio.


Several points of interest can be seen in the data presented in Fig. 2. For example, the Kantrowitz limit can be

relaxed significantly if the flexibility exists to provide


bypass duct areas or educated bleed holes in the internally contracting portion of the inlet. Educated bleed holes
have been used in the internal portions of the inlet to increase the starting contraction ratio beyond the Kantrowitz
limit.12'26'27 For example, the data provided in Ref. 12 are
plotted in Fig. 2 where it is seen that the Kantrowitz limit
was exceeded by a factor of 2. Although this approach
has been studied, the complexity of the bypass duct design and the drag associated with the bypass flow have
prevented its extensive use.
In Ref. 10, the starting characteristics of a twodimensional inlet with a thick boundary layer were
investigated. As shown in the data provided in Fig. 2, a
significant scatter in the starting contraction ratio was
observed. This scatter was caused by changes in the inlet
convergence angle, cowl lip bluntness, and Reynolds number. It was concluded10 that the Kantrowitz limit was valid
in situations where hJ5 was greater than 2.
Inlets tested in pulse facilities have been experimentally observed to start at contraction ratios significantly in
excess of the Kantrowitz limit.4'1U3-28-29 When testing in
a pulse facility, an inlet is placed in a test section with a
low initial pressure. As the facility starts, a starting shock
system sweeps through the test section and over the model.
When the conditions are suitable, the inlet will impulsively
start. Detailed information of the starting characteristics
of a simple two-dimensional inlet has been collected
using a Mach 8.3 gun tunnel. 13 By varying the initial

pressure in the test section and running the facility at a


constant Mach number and total pressure, the effect of the

pressure ratio across the starting shock could be investigated. As shown in Fig. 3 (from Ref. 13), the starting contraction ratio was found to be a strong function of inlet
shock strength and initial pressure ratio across the inlet.
In addition, pulse facility tests of an inlet are described28
where an inlet initially starts and then unstarts as the inlet

boundary layers develop.


Several important factors affect the maximum inlet
contraction ratio, including geometry, Mach number, wall
temperature, and throat profiles. Wall temperature has been
shown to be a major factor in inlet starting,2 but this finding has also been contradicted.11 A final factor that can
significantly affect the unstart contraction ratio is the flow
profiles in the inlet throat. The minimum throat Mach
number was determined as a function of the flow profile
in the inlet throat as shown in Fig. 4 (from Ref. 15). The

more uniform the flow was, the lower the Mach number
could be driven. This factor is one reason that the inlet

geometry strongly affects the starting and maximum contraction ratios for an inlet.
Several studies of the starting characteristics of threedimensional inlets have been undertaken.30"32 In these
studies, a modified Kantrowitz limit was derived taking
into account the three-dimensional spillage that occurs
during the starting process.
Given the large scatter in the available data and the
uncertainty that exists in the prediction of the starting

96-5965 F3

Fig. 3 Effect of initial test cabin pressure on starting


characteristics of a two-dimensional inlet13 (5C from
Ref. 13 = 0C in this paper).

American Institute of Aeronautics and Astronautics

(a)

Typical boundary

Traversable

Slideable compression surface


in position for diffusing the
supersonic boundary layer

probe

layer profiles

investigate fluid injection as a means for enhancing the


starting characteristics of inlets. In the final portion of
this paper, the effects of steady fluid injection on the starting characteristics are described.
Rig Description

Nozzle

Convergent
section

Boundary layer

section*.
-^

>

Station 0

k /..
Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

Throat

^ubsonic dftus

Open symbols; 5 model

(b)

Half-solid symbols; 7 model


Solid symbols; 10 model
Flagged symbols; low Reynolds number

1.4

1.8

2.2

A photograph and schematic of the test hardware


are shown in Figs. 5 and 6, respectively. A Mach 3
nozzle supplies flow to a test section, which is nominally
1 in. x 1 in. and 14 in. long. The inlet is simulated by a
single plate that spans the width of the test section. The

leading edge of this simulated cowl is sharp, with a 10


bevel on the top side of the plate. The maximum thickness of the cowl is O.I00 inch. The hardware was constructed so that the height of the cowl above the lower
floor could be varied. Tests were conducted with inlet
heights (hc) of 0.2, 0.3, 0.5, and 0.7 in. and cowl lengths
(Lc) of 0.75, 1.00, 1.25, 1.50 1.75, 2.00, and 2.50 inches.
The hardware is also constructed such that the cowl angle
could be changed by rotation about the leading edge while

_ 2.6

Average Mach number, M


96-5965 F4

Fig. 4 Variation of distortion index with averaged Mach


number.15 fa) Experimental setup, (b) Data obtained
at maximum model contraction ratio.

contraction ratios for inlets, an investigation of inlet starting characteristics was undertaken. The objective of this
activity was threefold. First, it was desired to use a relatively simple inlet geometry to investigate the parameters
that affect inlet starting. To this end, a small-scale rectangular facility was built in which a large number of potential parameters could be tested. In the following sections,
a description of this test facility is provided together with

the results from an initial test series. The second objective was to provide a database that would allow for an
assessment of the accuracy of computational fluid dynam-

ics codes in the prediction of inlet starting characteristics. This assessment is currently under way and will be
the focus of future reporting. The third objective was to

Fig. 5 Experimental hardware used in the current


investigation.

(15,0.836) (16.5,0.836)

(13.5,0.485

(3, 0.446)

(13,-0.483)

(3, -0.446)

(12.25,-0.479)

Fig. 6 Dimensions of experimental hardware (in inches).

American Institute of Aeronautics and Astronautics

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

the tunnel was running. This movement allowed the inlet to


be unstarted and restarted at will during tunnel operation.
The facility is equipped with a high-pressure air supply system. Tests were conducted with nominal supply
total pressures (Pt) of 50 and 100 psia. Ambient temperature air is used for the supply. A unit Reynolds number
range between 6.675 x 105/in. and 1.334 x 106/in. was
achieved.
The inlet was located in the aft portion of the test
section so that a thick boundary layer could be generated.
Calculations of the boundary layer characteristics were
made using an integral boundary layer code starting at
the throat of the supersonic nozzle; results are shown in
Fig. 7. At the location of the cowl lip, the boundary layer
thickness is approximately 0.18 inch. Given the variation
in cowl height, the ratio of hjb varied between 1.1 and
3.9.

The instrumentation provides measurements of the


supply total pressure, the static pressure distribution on
the centerline of the lower wall of the test section, the
cowl angle, and Schlieren and shadowgraph flow visualization. The tunnel total pressure was measured with 200psia transducers, and the static pressures were measured
with 25-psia transducers. The accuracy of each transducer
is 0.25% of full-scale. The cowl angle measurement was
made using a rotary potentiometer that had an accuracy
of 0.1. All data were recorded at a rate of 1 sample per
second.
In the operation of the tests, the flow was initialized
and stabilized at the desired test condition with the cowl
angle near 0. That angle would then be manually increased and decreased to induce the unstart and restart.
At least two unstart/restart cycles were obtained for every configuration.

Basic Unstart/Restart Characteristics


In conducting the unstart/restart tests, two general
types of unstart behavior were observed that can be classified as "hard" or "soft." Examples of hard unstarts are
shown in Figs. 8 and 9 for Pt = 100 psia, Le = 2.5 in., and
hc - 0.7 inch. In Fig. 8, the ratio of the pressure immediately below the cowl lip to the supply total pressure is
plotted as a function of cowl angle. Here, two sweeps of

the cowl are shown, with unstart occurring near 7.44 and
restart occurring near 6.63. Note the region of hysteresis
between the unstart and restart conditions. This type of
unstart occurs when the flow at the aft end of the inlet
chokes, creating a disturbance that propagates forward at
nearly the local speed of sound. For the conditions of this
test, the change in operation from a started to an unstarted

state occurs in approximately 0.25 ms.


Note that the point of unstart in Fig. 8 is slightly different between the two sweeps of the cowl angle. This
effect is caused by a combination of the natural spread in

0.03
-8.0 -7.8 -7.6 -7.4 -7.2 -7.0 -6.8 -6.6 -6.4 -6.2 -6.0

0CC (deg)

96-5965 F8

Fig. 8 Effect of cowl angle on entrance pressure showing a typical hard unstart (Pt = 100 psia, Lc = 2.5 in.,
Ac = 0.7 in.).
0.25

0.12

0.20
= 0.15

0.10

0.05

10

15

11.0

11.5 12.0

12.5

13.0

13.5

14.0

14.5 15.0
96-S96S F9

Fig. 7 Boundary layer characteristics at Mach 3 and Pt


= 100 psia as calculated with an integral boundary
layer code.

Fig. 9 Effect of cowl angle on wall pressure distribution showing a typical hard unstart (Pt = 100 psia, Lc
= 2.5 in., hc = 0.7 in.).

American Institute of Aeronautics and Astronautics

the unstart and restart conditions due to unsteady disturbances in the flow as well as quantization errors in the
recording system. In a hard unstart, the unstart angle can
be defined within two measurement samples, so the rate
at which the cowl angle is adjusted affects the measurement resolution for a fixed sampling frequency. Since the
cowl angle is manually adjusted, some variation exists in
this quantization error with respect to the unstart and restart measurements.

-- 0.83

-- 1.64
-*- 2.17

-*- 2.70
-*- 3.31
-*- 3.79

-t- 4.01

- 4.36
- 5.30

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

The effect of a hard unstart on the lower wall static


pressure distribution is shown in Fig. 9. For cowl angles
below 7.44, the inlet is started and the variation in the
cowl angle affects only the aft-most pressures. At cowl
angles above 7.44, the inlet unstarts with the unstart disturbance moving out forward of the inlet. Note that both
started and unstarted pressure distributions are shown for
a cowl angle of 7.44. In this case, the inlet transitioned
from started to unstarted operation while the cowl was
stationary. This behavior was observed when the cowl
angle was set at a fixed position near the maximum contraction point. Small disturbances in the freestream could
drive the inlet from the started to unstarted condition.
These disturbances were not extensively investigated, so
some uncertainty in the measured unstart and restart is
present owing to the unsteady operating characteristics
of the test rig. For most conditions, this factor is believed
to be small.
The characteristics of a soft unstart are illustrated in

Figs. 10 and 11 for a test with Pt = 100 psia, Lc = 2.5 in.,


and hc = 0.3 inch. In Fig. 10, the pressure on the test section wall immediately below the cowl lip is shown as a
function of cowl angle. Note that no hysteresis is visible
as the cowl is swept through a range of angles. This type
of unstart is seen as a continuous process rather than the
discrete transition that occurs with the hard unstart. The
effect of cowl angle on wall static pressure distributions
is shown in Fig. 11 for this case. A gradual transition in

0.03
-6.0

-5.6

Fig. 10 Effect of cowl angle on entrance pressure showing a typical soft unstart (P, = 100 psia, Lc = 2.5 in.,
/ic = 0.3in.).

11.0

11.5

12.0

12.5

13.0

13.5

14.0

14.5 15.0
96-5965 F11

Fig. 11 Effect of cowl angle on wall pressure distribution showing a typical soft unstart (Pt = 100 psia, Lc
= 2.5 in., hc = 0.3 in.).

the pressure distribution is observed as the cowl angle is


increased.
Figures 12 and 13 are Schlieren photographs of the
inlet flow for the two types of unstart. In Fig. 12, the
sequence of photographs shows a hard unstart for Pt =
100 psia, Lc = 1.0 in., and hc = 0.5 inch. As the cowl angle
is increased from 0 to 7, the cowl shock strength and
angle increase. Between 7 and 8, the inlet unstarts. As
the cowl angle is increased above 8, the unstart separation region moves forward. Near a cowl angle of 17, the
entire test section unstarts. In Fig. 13, a soft unstart is
shown for Pt = 100 psia, Lc = 2.0 in., and hc = 0.3 inch.
The more gradual nature of the soft unstart is evident.

A summary of the unstart and restart data is provided


in Figs. 14 and 15. In general, the cowl angle that causes
unstart increases as the cowl length decreases and the cowl
height increases. The unstart angles are also slightly higher
for operation at the higher total pressure. Note that for a
Mach 3 flow, boundary layer separation can be expected
with an impinging and reflecting shock when the cowl
angle is greater than approximately 8.5. This angle is
exceeded for the shortest cowls at the higher cowl heights.
The specification of an unstart as soft or hard is somewhat subjective in that it was not always clear which type
of unstart occurred. In the present work, the unstart angle
is defined as the cowl angle at which the pressure on the
tunnel wall is first affected by increasing the cowl angle.
A crude boundary between the soft and hard unstarts exists in the parameter space of cowl length and cowl height
(Fig. 16). For shorter cowls and higher cowl heights, hard
unstarts are prevalent, whereas the softer unstarts occur
for the longer cowl lengths and lower cowl heights.
In assessing the ability of the Kantrowitz limit to predict the restart contraction ratio, an estimate for the Mach
number at the cowl entrance is required. The local value

American Institute of Aeronautics and Astronautics

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

, = 1.3

= 0.3

= 3.3"

e, = 2.4

. = 6.7

. = 4.5

= 8.7

. = 5.5

96-5965 F13

fl = 13.2

Fig. 13 Sequence of Schlieren photographs showing soft


unstart for Pl = 100 psia, Lc = 2.0 in., and hc = 0.3
inch.
Mach number was determined at the entrance to the inlet
as a function of cowl height.

6 =17.8

Fig. 12 Sequence of Schlieren photographs showing


hard unstart for Pt = 100 psia, Lc = 1.0 in., and hc =

0.5 inch.
of static pressure was combined with the tunnel total
pressure to obtain the inviscid core Mach number.
A l/7th-power velocity profile and modified Crocco temperature distribution were combined with the calculated

boundary layer thickness (shown in Fig. 7) to estimate


the flow parameters in the boundary layers on the tunnel
floor and sidewalls. Using these profiles, a mass-averaged

In Fig. 17, the internal area ratio at which the inlets


restarted is shown as a function of the average Mach number at the cowl lip. The Kantrowitz limit is also shown.
The restart measurements indicate a tremendous amount
of spread relative to the Kantrowitz limit. In analyzing
these results it was discovered that much of the spread
was due to the conditions where the inlet unstart and restart were classified as soft. Some additional portion of
the spread occurred for the cowl height of 0.7 inch. In
reviewing the Schlieren photographs, it was discovered
that the expansion off the upper surface of the facility likely
enters the inlet when it is positioned at a height of 0.7
inch. When this happens the procedure used to estimate
the Mach number at the inlet entrance is no longer valid.
In Fig. 18, the restart contraction ratio data were replotted with all soft restart data or data obtained at cowl
heights of 0.7 in. removed. Under these conditions, the

American Institute of Aeronautics and Astronautics

7.5
*o>

12.5

10.0

*s
<D

7,5

5.0

10.0

5.0

Unstart

h = 0.2

13

2.5

0.0

Unstart

2.5

Restart
i
1.0

i
0.5

nn

OS

i
1.5

i
2.0

Restart
i
1.0

nn
0.0

0.5

i
1.5

i
2.0

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

12.5
10.0

10.0

7.5

^
oi

5.0

7.5

0>

Unstart

Unstart

a
"7>

2.5

5.0

Restart

2.5
0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

2.0

2.5

3.0

Uin.)

Min.)
12.5

10,0

10.0

/)=0.5

7.5

-,
O)

7.5

2.
o

5.0

2.5

2.5

0.0

0.0
0.0

0,5

1.0

2.0

1.5

2.5

3.0

12.5

12.5

10.0

10.0

7.50

_-

0)

i.
=&

5.0
Restart

Restart

0.0

0.5

1.0

1.5

= 0.7

7.5

Restart

5,00
2.5

2.50

0.0

00
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

1,0

1.5

2.0

2.5

3.0

98-6965 F15

96-5965 FI4

Fig. 14 Effect of cowl length and height on unstart and


restart cowl angles at Pt = 50 psia.

Fig. 15 Effect of cowl length and height on unstart and


restart cowl angles at P, = 100 psia.

Kantrowitz limit appears to predict the restart contraction


ratio with an accuracy of approximately 10%. This result
leads one to suspect that some of the scatter in the available

starting data in the literature may be caused by


misclassification of the starting process or uncertainties
in the approach Mach number for the tested inlets.

American Institute of Aeronautics and Astronautics

1.0

0.75

0.9

Hard

0.50 -

unstarts
N

0.8

Kantrowitz
limit

0.7
H

SXN

055

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

H/N

0.5

1.0

2.0

1.5
L

c(

0.6

ln

2.5

->

0.5

2.0

2.2

96-5965 F16

_
Mz

Fig. 16 Approximate boundary between hard and soft

unstarts (H = hysteresis present, S = small degree of


hysteresis, N = no hysteresis present).

2.4
2.4

2.6
96-5965 F18

Fig. 18 Measured restart area ratios for hard restarts and


hc < 0.7 inch.

1.0

Steady started
flowfield

0.9

"Steady" unstarted
flowfield

0.8

Recirculation
zone,-

Kantrowitz

limit

0.7

Nonequilibrium

unstarted flowfield
0.6

Fluid
injection

Recirculation
zone

0.5

2.0

96-5965 F19

2.2

2.4

2.6
96-5965 F17

Fig. 19 Effect of fluid injection on unstarted inlet


flowfield.

Fig. 17 Measured restart area ratio as a function of the

mass-averaged Mach number at the cowl entrance.

Effect of Injectant on Unstart/Restart


One purpose of the current investigation is to examine the effect of fluid injection on the restart characteristics of the inlet. The motivation for this portion of the
experiment can be explained with the aid of Fig. 19. In
the upper portion of this figure, the started inlet flowfield
is shown. As the inlet contraction rado is increased, the
inlet eventually unstarts, and a large separation zone forms
in front of the inlet. This flowfield is naturally unsteady
to a certain degree, but it oscillates around a "steady" condition. As noted previously, a significant reduction in the

inlet contraction may be required to restart the inlet. Fluid


injection from the lower surface has been proposed as a

means for altering the unstarted flowfield and influencing the restart characteristics by drawing the separated
zone forward from its "steady" position. It is believed that
if the fluid is then turned off quickly, the aftward movement of the separation will behave somewhat like the pulse

facility starting process such as that found in Refs. 4, 11,


13, 28, and 29. In this section, the effect of the steady
fluid injection on the inlet unstart and restart operation is

presented. The influence of quickly turning the flow off


will be the subject of future reporting.
In the test rig, a row of twenty-five 0.020-in.-dia.
normal injection holes was placed 1.75 in. upstream of

American Institute of Aeronautics and Astronautics

the cowl leading edge. Air was supplied to these holes


through a common plenum, and injectant pressure ratios
(i.e., Piaj/Po) of 1 to 5 were investigated. This produced
an additional momentum thickness equal to thin/ Po^O of
up to 0.046 inch.
In Fig. 20, the effects of fluid injection on the unstart

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

and restart cowl angles are shown for Lc = 2.0 and 2.5 in.
and injectant pressure ratios between 1 and 5. These results are presented for a cowl height of 0.5 inch. The effect of the steady fluid injection is to decrease the cowl
angle at which both unstart and restart occur owing to the
additional mass flow of the injection. Note that these conditions correspond to soft unstart configurations in the

data previously shown. When the test apparatus was modified for the fluid insertion investigation (i.e., holes were
added in the lower wall), a small amount of hysteresis
between unstart and restart was observed. The cause of
this different behavior is under investigation.
8.0
7.5

~L = 2.0 in.

Unstart

7.0

jection for a case with Lc = 1.0 in. and hc = 0.5 inch. Both
started and unstarted pressure distributions are provided.

Note that when the fluid injection was added to the test
hardware, several pressure taps in the vicinity of the injection were not available, so determination of the effect
of injection on the pressure distribution is difficult. On
the basis of results shown in Fig. 21, it appears that the
injectant has little effect on the unstarted separation zone.
For this case, the fluid injection may be too far upstream
of the inlet to affect the separation zone.
The effect of the injectant on the pressure distributions is shown in Fig. 22 for a second case where Lc = 2.5
in. and /zc = 0.5 inch. Here, a noticeable change in the
unstart pressure distribution is seen as the injectant pressure ratio is changed. For this case it appears that the injectant does modify the unstart separation. On the basis
on these initial results, the fluid injection may be useful
in modifying the restart characteristics of the unstarted
inlets. A more detailed mapping of the effects of fluid
injection on these characteristics is required and will be
the subject of future reporting.

Future Plans

' 6.5
>

Future investigations will focus on obtaining unsteady


flow measurements in the experimental rig, obtaining data
at higher Mach numbers, and investigating the effects of
pulsed fluid injection on the restart characteristics of the

6.0

5.5

inlet. High-response pressure transducers will be added

5.0

6
96-5965 F20

Fig. 20

Effect of steady injectant on unstart and restart

cowl angles for Lc = 2.0 and 2.5 in. (P, = 100 psia).

to the hardware, and the impact of unsteady effects within


the freestream on the unstart and restart process will be
examined. Tests will also be conducted at Mach 4 conditions in an attempt to generate conditions with a greater
hysteresis in the inlet operation. It is in this hysteresis
0.16

0.200

e - 3.4. P^/P, = 1.0

0.1750.150-

0.14

Unstarted

c 14.3'. P^P,. 1.2


P

'. 4.6'. P.,' =

0.12

18

, - 14.7. PnIP,- 2.1

0.10

Cow! lip

Injection

0.125-

0.08

Unstarted

JJT 0.1000.075

0.050-

Started

0.06
0.04
il'l Started

0.02

0.025
0.000

The objective of the fluid injection is to affect the


steady unstart separation zone. In Fig. 21, the lower wall
pressure distribution is shown with and without fluid in-

10

11

12

13

14

15

10

11

12

13

14

15

16

17

x(m.)
96-5965 F22

Fig. 2 1 Effect of fluid injection on started and unstarted


pressure distributions for Lc = 1 .0 in. and hc = 0.5 in.
(/>, = 100 psia).

Fig. 22 Effect of fluid injection on started and unstarted


pressure distributions for Pt = 100 psia, Lc = 2.5 in.,

and hc = 0.5 inch.

10
American Institute of Aeronautics and Astronautics

zone that the pulse fluid injection may have an effect on


the restart characteristics. The ability of computational
fluid dynamics codes to predict the unstart and restart
characteristics will also be assessed.

Conclusions
A small-scale supersonic inlet experiment has been
conducted to better understand the major factors that influence the unstart and restart characteristics of simple
inlets. The results of the experiment have shown that both
Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

hard and soft unstarts and restarts can occur, depending


principally on the inlet geometry and the amount of inviscid flow entering the inlet. Unsteady effects observed
within the test hardware were also shown to affect the
unstart characteristics of the inlets. For the inlets investigated, the Kantrowitz limit was successful in predicting
the restart contraction ratio with an accuracy of approximately 10% if only the hard unstarts were considered and
the Kantrowitz limit was based on the mass-averaged

Mach number at the cowl lip.


The role of steady fluid injection on the starting characteristics of the simulated inlet was also investigated. In
all cases, fluid injection degraded the unstart and restart
abilities of the inlet. Future investigations will focus on
the use of pulsed fluid injection for affecting the restart
characteristics of the inlet.

Acknowledgment
This work was supported by the Independent Research and Development Program at The Johns Hopkins
University Applied Physics Laboratory. The idea for the
use of the fluid injection for modifying the restart characteristics was first proposed by Mr. George McLafferty.

Scramjet Inlets in Tetraflouromethane, AIAA 90-0530


(Jan 1990).
6
White, M. E., Stevens, J. R., Van Wie, D. M., Mattes,
L. A., and Keirsey, J. L., "Investigation of Cowl Vent
Slots for Supercritical Stability Enhancement of DualMode Ramjet Inlets," /. Prop. Power 26(3), 225-226
(May-Jun 1990).
7

Waltrup, J. P., Anderson, G. Y., and Stull, F. D., "Supersonic Combustion, Ramjet (SCRAMJET) Engine Development in the United States," in Proc. Third International Symp. on Air Breathing Engines, Munich,
Germany (Mar 1970).
8
Billig, F. S., SCRAM-A Supersonic Combustion Ramjet Missile, AIAA 93-2329 (Jun 1993).
9
Kantrowitz, A., and Donaldson, C., Preliminary Investigation of Supersonic Diffuser, NACA WRL-713

(1945).
Goldberg, T. J., and Hefner, J. N., Starting Phenomena
for Hypersonic Inlets with Thick Boundary Layers at
Mach 6, NASA TN-D 6280 (Aug 1971).
1:
Gurylev, V. G., and Mamet'yev, Yu. A., "Effect of Cooling of the Central Body on the Start-Up Separation of
the Flow at the Intake and the Throttling Characteristics of Air Scoops at Supersonic and Hypersonic Velocities," Fluid Mech. Sov. Res. 7(3) (May-Jun 1978).
12
Mahoney, J. J., Inlets for Supersonic Missiles, AIAA
10

13

Education Series (1993).


McGregor, R. J., Molder, S., and Paisley, T. W., "Hypersonic Inlet Flow Starting in the Ryerson/University
of Toronto Gun Tunnel," in Investigations in the Fluid
Dynamics of Scramjet Inlets, Ryerson Polytechnical
University and the University of Toronto, Canada, pp.

Molder, S., McGregor., R. I, and Paisley, T. W., "A


Comparison of Three Hypersonic Air Inlets," in Investigations in the Fluid Dynamics of Scramjet Inlets,
Ryerson Polytechnical University and University of
Toronto, Canada, pp. 6.1-6.94 (Jul 1992).
2
Andrews, E. H., McClinton, C. R., and Pinckney, S. Z.,
Flowfield and Starting Characteristics of an

4.1-4.50 (Jul 1992).


Watson, E. C, An Experimental Investigation of Mach
Numbers from 2.1 to 3.0 of Circular Internal Compression Inlets Having Translatable Center Bodies and Provisions for Boundary Layers Removal, NASA TM-X156 (Jan 1960).
15
Cnossen, J. W, and O'Brien, R. L., "Investigation of
the Diffusion Characteristics of Supersonic Streams
Composed Mainly of Boundary Layers," J. Aircr. 2(6)
(Nov-Dec 1965).
16
Hypersonic Ramjet Program for the Period I March to

Axisymmetric Mixed Compression Inlet, NASA TMX2072 (Jan 1971).


3
Karanian, A. J., and Kepler, C. B., Experimental Hypersonic Met Investigation with Application to Dual
Mode Scramjet, AIAA 65-088 (Jun 1965).

31 December 1964, Vol. 1 18-in. Engine Program,


AFAPL-TR-65-36, Vol. 1 (May 1965).
17
McFarlin, D. J., and Kepler, C. A., Mach 5 Test of Hydrogen-Fueled Variable Geometry Scramjet, AFAPLTR-68-247 (1954).

Gunther, F., Development of a Two-Dimensional Adjustable Supersonic Inlet, CIT Report 20-247, Jet
Propulsion Laboratory (1954).
19
Vahl, W. A., and Oehman, W. L, Internal Flow Characteristics of a Fixed-Geometry Induction System

References
1

Van Wie, D. M, and Ault, D. A., Internal Flowfield


Characteristics of a Two-Dimensional Scramjet Inlet
at Mach 10, AIAA 94-0584 (Jan 1994).
5
Holland, S., and Perkins, J., Mach 6 Test of Two Generic Three-Dimensional Sidewall Compression

14

18

11
American Institute of Aeronautics and Astronautics

Having Axial Symmetry at Mach Numbers from 3.8 to


4.2, NASA TM-X-759 (Jan 1963).
20
Sokata, K., Yanagi, R., Murakami, A., Shindo, S.,
Honami, S., Shizaua, T., Sakomoto, K., Shiraisha, K.,

26

and Omi, J., An Experimental Study of the Supersonic


Air Intake with 5-Shock System of Mach 3, AIAA 932305 (Jim 1993).

27

21

McLafferty, G. H., A Study of Perforation Configurations for Supersonic Diffusers, United Aircraft Corporation Research Department Report R-53372-7 (Dec
1950).
28
Williams, R. L., "Application of Pulse Facilities to Inlet Testing," /. Aircr. 1(10), 236-241 (Oct 1964).
29
Lashkov, A. I., and Niko'skii, A. A., "Shock Starting of
a Supersonic Diffuser," Inzh. 7h. 2(1), 11-16 (1962).
30
Trexler, C. A., Inlet Starting Predictions for SidewallCompression Scramjet Inlets, AIAA 88-3257 (Jul
1988).
31
Murakami, A., Yanagi, R., Shindo, S., Sakato, K.,

Downloaded by ROKETSAN MISSLES INC. on October 22, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.1996-2914

Demarest, P. E., High Temperature Tests of an External


Plus Internal Compression Inlet at Mach 6.5, ASD-TDR
62-280 (Sep 1962).
22
Karanian, A. J., and Debois, R. L., Investigation of

23

McLafferty, G. H., Tests of Perforated ConvergentDivergent Diffusers for Multi-Unit Ramjet Application,
United Aircraft Corporation Research Department Report R-53133-19 (Jun 1950).

Precompression Devices for Downward Turning Inlets,


United Airlines Research Laboratory Report M91135610 (Apr 1973).

Heins, A. E., Reed, G. J., and Woodgrift, K. E., Hydrogen


Scramjet Feasibility Program Port III Freejet Engine
Design and Performance, AFAPL-TR-70-74 (Jan 1971).
24
Ragsdale, W. C, An Investigation of Leeward Aft-

Honami, S., Tanaka, A., and Shiraish, K., Mach 3 Wind


Tunnel Tests of Mixed Compression Supersonic Inlet,

Entry Ramjet Inlet Performance, TR-73-25, Naval


Ordnance Laboratory (Feb 1973).
25
Trexler, C. A., and Souders, S. W., Design and Performance at a Local Mach Number of 6 of an Inlet for Integrated Scramjet Concept, NASATN-D-7944 (Aug 1975).

32

AIAA 92-3625 (Jul 1992).


Tani, K., Kanda,T., Kudou, K., Murakami, A., Komuro,
T., and Itoh, K., Aerodynamic Performance of Scramjet
Inlet Models with a Single Strut, AIAA 93-0741 (Jan
1993).

12

American Institute of Aeronautics and Astronautics

Vous aimerez peut-être aussi