Axisymmetric Inlet Design For Combined-Cycle Engin
Axisymmetric Inlet Design For Combined-Cycle Engin
Axisymmetric Inlet Design For Combined-Cycle Engin
net/publication/245435655
CITATIONS READS
23 4,571
3 authors, including:
Ryan Starkey
University of Colorado Boulder
80 PUBLICATIONS 756 CITATIONS
SEE PROFILE
All content following this page was uploaded by Ryan Starkey on 19 August 2015.
are presented using the inlet of the Lockheed SR-71 as a baseline. A numerical
characteristics are quantified based upon the Kantrowitz limit. The original SR-71
inlet is analyzed throughout the designed self-starting regime, beginning at Mach 1.7
and ending with the shock-on-lip condition at Mach 3.2. The characteristics model is
considered for their ability to extend the range of the inlet into the hypersonic flight
regime. Self-starting characteristics of these new designs are also characterized;
results indicate that two new designs can maintain self-starting capability into the
Mach 6-7 range. Full external and internal flow properties of the new designs are
determined using the characteristics model. Mach number, total pressure ratio,
temperature, pressure and mass flow properties (and their levels of distortion) are
By
Jesse R. Colville
Advisory Committee:
Professor Mark J. Lewis, Chairman/Advisor
Associate Professor Kenneth Yu
Assistant Professor Christopher Cadou
© Copyright by
Jesse R. Colville
2005
DEDICATION
To my parents, Donald and Diane Colville, for their support through the first twenty-
ii
ACKNOWLEDGEMENTS
First and foremost, I have to thank my advisor Dr. Mark J. Lewis. His advice,
guidance and support have been unwavering since I first approached his office as a
committee members, Dr. Kenneth Yu and Dr. Christopher Cadou, for their time and
essentially being my mentor for the past four to five years. I honestly do not know
where I would be at in this work without your help and insight. Thank you for letting
me bug you so often without you (at least openly) getting annoyed! Maybe I’ll see
I would also like to express some gratitude to those who backed this project.
This work has been sponsored by the Space Vehicle Technology Institute, under grant
NCC3-989, one of the NASA University Institutes, with joint sponsorship from the
Harry Cikanek of the NASA Glenn Research Center, and to Dr. John Schmisseur and
iii
Now to the office. Thank you to Andy, Justin, Kerrie, Josh, Marc, Greg,
Amardip, etc. for creating a comfortable atmosphere in which to work in and for
putting up with my insanity and outbursts while finishing this work (and whenever
I also have to thank my entire family for their love and support from day one.
I know I have not been the easiest child to raise or brother to befriend, but you all
Finally, I have to thank my fiancée Michelle. You have been my crutch for
the past three years. You have put up with every one of my mood swings and dealt
with the many, many six to seven day work weeks that I had to undertake the past two
years. You are now about to venture across the country with me as I move on into the
industry. I can only hope to be able to fully repay you at some point in the future.
iv
TABLE OF CONTENTS
ABSTRACT................................................................................................................... i
DEDICATION.............................................................................................................. ii
ACKNOWLEDGEMENTS......................................................................................... iii
TABLE OF CONTENTS.............................................................................................. v
1 Introduction........................................................................................................... 1
1.1 Motivation..................................................................................................... 1
2 Model Development............................................................................................ 19
v
2.3 Solution Procedure...................................................................................... 23
4 CFD Validation................................................................................................... 55
vi
5.3 Re-extension with a Widened Shoulder...................................................... 77
vii
LIST OF FIGURES
Figure 2.4 Example of Characteristics Mesh and Reflected Shock System ............... 28
viii
Figure 2.7 Diagram of Shock Point Calculation nearest to Boundary........................ 33
Figure 2.9 Diagram of Lambda Shock and Triple Point Location ............................. 36
Figure 3.4 Full Vehicle Mach 3.2 CFD Shockwave Patterns (top view)39 ................. 44
Figure 4.1 External Grid used for CFD Analysis (1/5th the resolution)...................... 57
Figure 4.2 Internal Grid used for CFD Analysis (1/5th the resolution)....................... 57
ix
Figure 4.9 Viscous CFD Results................................................................................. 64
Figure 5.3 Re-extended Spike Inlet Total Pressure Ratio Contours ........................... 69
Figure 5.6 Re-extended Spike Inlet Mass Flow Ratio Contours ................................ 72
Figure 5.13 Widened Shoulder Centerbody Inlet Mach Number Contours ............... 80
Figure 5.14 Widened Shoulder Centerbody Inlet Total Pressure Ratio Contours...... 81
Figure 5.15 Widened Shoulder Centerbody Inlet Temperature Ratio Contours ........ 82
Figure 5.16 Widened Shoulder Centerbody Inlet Pressure Ratio Contours ............... 83
Figure 5.17 Widened Shoulder Centerbody Inlet Mass Flow Ratio Contours ........... 84
x
Figure 5.23 Self-Starting Characteristics of the Variable Cone with Re-extension ... 91
Figure 5.24 Variable Cone with Re-extension Inlet Mach Number Contours ........... 93
Figure 5.25 Variable Cone with Re-extension Inlet Total Pressure Ratio Contours .. 94
Figure 5.26 Variable Cone with Re-extension Inlet Temperature Ratio Contours..... 95
Figure 5.27 Variable Cone with Re-extension Inlet Pressure Ratio Contours ........... 96
Figure 5.28 Variable Cone with Re-extension Inlet Mass Flow Ratio Contours ....... 97
Figure 5.29 Supersonic Variable Cone with Re-extension Throat Properties ............ 99
Figure 5.30 Subsonic Variable Cone with Re-extension Throat Properties ............. 100
xi
LIST OF SYMBOLS
A = Area, ft2
C = Characteristic Line
M = Mach number
P = Pressure, lb/ft2
T = Temperature, oR
V = Velocity, ft/s
V = Non-dimensional velocity
Vl = Limiting velocity
xii
x = axial direction or component of vector
φ = potential function
ρ = density, slug/ft3
ψ = stream function
Subscripts
4 = at throat station
l = limiting condition
o, t = total condition
xiii
θ = derivative in angular direction
xiv
1 Introduction
1.1 Motivation
vehicles and high speed missiles. TBCC systems are designed to merge the low
speed attributes of turbojets with the high speed capabilities of ramjets and scramjets.
The integration of the TBCC cycle modes (gas turbine, ramjet and scramjet) into a
single system poses a considerable challenge; each requires unique flow properties in
order to operate properly. Of particular importance is that the inlet must provide
efficient compression for all three modes across a wide Mach spectrum. Recent
TBCC cycle analysis1 has demonstrated that inlet performance imposes a significant
constraint on the overall operation of the engine and can be, in fact, the limiting factor
The basic criteria for designing both supersonic and hypersonic inlets have
been well documented in the literature2-4. Some key issues are as follows:
1
• Diffusing the required amount of air needed for engine performance with
• Minimizing both shockwave and viscous losses for good total pressure
recovery
concerning combined cycle inlets and outline a roadmap for an advanced inlet
program to develop the critical inlet technologies. The particular topics addressed in
the paper are the transition from one cycle mode to another, boundary layer control,
distortion, unstart and overall performance. Their analysis, conducted with 1D, 2D
and total pressure recovery issues need to be corrected before any practical design is
realized.
Another design issue for TBCC inlets is identifying a specific design speed.
Inlets for a cruising vehicle are traditionally designed at the cruise speed since a large
amount of time is spent flying at that velocity. On the other hand, an inlet for an
discrete design points for the inlet poses a tougher task. Variable geometry is likely
2
needed attain multiple design points.
The main issues that are dealt with in the present work are total pressure
recovery, flow distortion and inlet starting. Total pressure recovery is important as it
plays a key role in engine thrust and efficiency; the lower the recovery, the lower the
thrust. In theory, the pressure losses can be augmented with an increase in fuel flow
rate, but this leads to larger tank sizes and larger vehicle sizes (and lower specific
impulse).
uniform flowfields can lead to these components performing poorly. For the purpose
this study, the distortion (or non-uniformity) is quantified as the difference between
the minimum and maximum property (for example, χ) value divided by the area
χ max − χ min ∆χ
= . (1.1)
χ avg χ avg
inlets for TBCC engines must function properly during transition from one cycle
mode to another across the entire flight regime. Hill and Peterson6 provide a very
section. For an inlet to be self-started, supersonic flow must exist at the minimum
area (or throat) of the duct. In essence, the inlet remains self-started if flow
3
disturbances (both external and internal) do not change the captured air characteristics
and the inlet is able to maintain supersonic flow at the throat. These disturbances
could be a sudden rise in back pressure downstream of the throat, a gust of wind
contraction ratios that will self-start can be obtained by calculating the Kantrowitz
beginning of the internal compression section. The internal area ratio that produces
sonic flow at the throat is then found by assuming one-dimensional, isentropic flow
through the duct. For a perfect gas, the limit can be calculated as follows:
γ +1
γ γ −1 2 ( γ −1)
1 +
1
A2 1 (γ + 1)M 22 γ −1 γ +1 γ −1
M 22
= ⋅ ⋅ 2 (1.2)
M 2 (γ − 1)M 2 + 2 2γM 2 − (γ − 1) γ +1
2 2
A4
2
Note the M2 in the Equation 2 is the mass-averaged Mach number across the cowl
Van Wie8 discusses experimental data found from a wide variety of inlets and
compares their starting characteristics to that predicted by the Kantrowitz limit. This
work shows that the Kantrowitz limit can be, under some circumstances, exceeded
with bleed holes and/or bypasses. However, for the purposes of this research, the
4
1.2.1 Previous Work
has yet to be developed. Likewise, the inlet for such an engine has not been designed
either. However, supersonic airplanes and missiles have been flying for decades.
While the engines operating on aircraft may not be (or even similar to) combined
cycles, their inlet characteristics are very analogous to that needed on a TBCC inlet.
The following sections will discuss a sample of supersonic aircraft, along with their
engines and inlets, to gain insight into the design strategies of existing supersonic
inlets. It should be noted that some of the vehicles reported on were (or are still)
scarce. For the most part, this work has relied upon open-literature NASA reports.
The Lockheed SR-71 (Figure 1.1) is still the world’s fastest manned air-
breathing aircraft. The airplane was designed to serve as a long range reconnaissance
vehicle and could fly at speeds upwards of Mach 3.2 with a ceiling of about 85,000
ft9. The SR-71 was powered by two Pratt & Whitney J-58D engines (Figure 1.2).
Each engine was an axial flow turbojet using a nine stage compressor, two stage
turbine with bypass and produced 34,000 lbs of thrust with the afterburner. It was the
first jet engine designed to operate over long periods of time using its afterburner. At
high Mach numbers, the turbo-machinery was heavily bypassed thereby creating a
Interestingly, a study10,11 was completed in the mid-90’s looking at the SR-71 as the
5
first stage of a small payload launch vehicle.
compression system. Ben Rich, program manager for the propulsion system of the
SR-71, called the design of the inlets the “single most complex and vexing
engineering problem of the entire project12.” He also went on to add that “developing
this [system] was the most exhausting, difficult, and nerve-racking work of my
professional life.” An intensive review of the inlet operation is given in the SR-71
Flight Manual13. The external system utilized a translating conical spike that
retracted 1 ⅝ inches per 0.1 Mach number starting near Mach 1.6 and ending at Mach
3.2 with shock-on-lip condition. The internal system consisted of a series of reflected
tripping the flow; the inlet self-starts when the spike begins to translate. Originally
controlled by the pilot, the inlet system was outfitted with a Digital Automated Flight
and Inlet Control System (DAFICS)14 in the early 1980’s to allow for faster response
to unstart.
and bypasses, is to control the amount of flow entering the engine and to hold the
terminal shock downstream of the throat (when M∞ > 1.6) to avoid unstart. At low
Mach numbers, the forward air inlet bypass doors open to exhaust air that is not
needed by the engine. As the spike translates, the forward bypass doors begin to
close in order to position the shock downstream of the throat. Shock trap and
centerbody bleeds, along with aft bypass doors, also aid in the process of keeping the
6
shockwave/boundary layer interactions. The spike translation increases the captured
stream tube area from 8.7 ft2 to 18.5 ft2. In the event of unstart of one of the engines,
the DAFICS would push both engine spikes forward and control the bleeds and
bypasses such that normal operation and normal shock position could be reobtained
within seconds.
7
Figure 1.3 SR-71 Inlet9
One of the most secret aircraft ever developed, the D-21 Drone was an
unmanned, expendable reconnaissance vehicle that flew in excess of Mach 3.5 above
90,000 ft.9 Shown in Figure 1.4, The 42 ft. long vehicle was originally launched from
an M-21 (a cousin of the SR-71); however, after a fatal crash during separation, the
drone was fixed with a carrier rocket and launched from a modified B-52 instead.
The spy vehicle was designed to eject the reconnaissance data prior to crashing. The
drone was outfitted with a Marquardt XRJ-43-MA20S-4 ramjet capable of run times
upwards of 2 hours12.
The D-21 inlet is shown in Figure 1.5. Due to the intense secrecy of the
program, there are no open-literature sources detailing the exact performance and
geometry of the D-21 inlet. However, the inlet used on the Bomarc B Missile
Interceptor used a similar ramjet (XRJ-43-MA20) and its inlet was a fixed geometry,
Mach 2.35 axisymmetric isentropic inlet19. In all likelihood, since the D-21 was able
to fly past Mach 3.5, its design Mach number was not Mach 2.35. However, the inlet
8
does appear to have an isentropic spike. The original spike was probably fixed, as
well, since the flight envelope of the D-21 was not very wide. Interestingly, in 1999
outfitting the D-21 with a Rocket Based Combined Cycle (RBCC) engine. The
cycle Operation). The modification to the inlet system required the replacement of
the original fixed geometry axisymmetric inlet with a translating axisymmetric inlet
9
Figure 1.5 D-21 Inlet9
hypersonic research ramjet/scramjet and to flight test the concept from Mach numbers
3 through 8—this project was known as the Hypersonic Research Engine (HRE)22.
Shown in Figure 1.6, the HRE was intended to be the first project to integrate years of
research in high speed inlets and nozzles, and the first to use a direct-connect scramjet
however, because of the cancellation of the X-15 program, no flight tests were ever
performed.
A schematic of the inlet used on the HRE is show in Figure 1.7. It was a
mixed compression system that utilized an axisymmetric translating spike. The spike
consisted of a 10o initial half angled cone that isentropically turned the flow up to a
22o surface angle. The spike was fixed from Mach 4 to 6 with the spike tip shock
falling outside of the cowl lip. Starting at Mach 6, the spike tip shock impinged on
the cowl leading edge—this condition was maintained through Mach 8 by pushing the
centerbody forward to maintain the shock on lip criteria. The engine was operated in
10
a ramjet mode (i.e. subsonic combustion) from Mach 4 to 6. However, starting
between Mach 5 and 6, the engine transitioned from ramjet to scramjet mode. The
bomber with a gross weight of about 500,000 pounds and a cruising speed of Mach
325. The schematic in Figure 1.8 shows the design features. The design highlighted a
65.6o swept back leading edge delta wing with folding wing tips (creating a quasi-
11
waverider design). The vehicle propulsion system resided underneath the aircraft and
consisted of two, two-dimensional inlets mounted side-by-side. Each inlet fed three
1.9. The inlet started at Mach 2.0 (i.e. the terminal shock laid across the inlet cowl
face). Each inlet was composed of variable ramps and bypass doors that could be
positioned to maximize performance through the flight profile. The boundary layer
was controlled through a porous wall bleed system in the throat region. Each intake
had six bypass doors installed just forward of the engine face that, used in
conjunction with the variable throat, would position the normal shock and match
engine airflow requirements. Control of the inlet system was both manual and
semiautomatic in the first prototype XB-70; however, the second was fully automatic.
The purpose of the variable throat and bypass system, as with the SR-71, was to
position the terminal normal shock slightly downstream of the aerodynamic throat to
12
Figure 1.8 Boeing XB-70 Supersonic Bomber26
1.2.1.5 Concorde
For over 25 years, up until 2003, the Concorde (Figure 1.10) was the only
operational civil supersonic transport. It was designed for a maximum cruise speed of
Mach 2.227. The major design features of the vehicle area modified delta wing, no
13
Each engine (OL 593) was capable of generating over 38,000 pounds of thrust at
single-stage low pressure turbine, both with cooled stator and rotor blades. An
afterburner was also used to provide adequate thrust at takeoff and for transonic
acceleration.
The schematic of the four inlets is shown in Figure 1.11. Each two-
dimensional inlet consists of an initial, fixed ramp, followed by a second hinged ramp
that utilizes isentropic compression28. A terminal, strong oblique shock emanates off
of the cowl lip. Below Mach 1.3, both ramps are fully open and the bleed door is
closed. Above Mach 1.3, the ramps and dump door are moved per the engine
requirements. The boundary layer diverter and ramp bleed slot are employed to
reduce the negative effects of boundary layer growth. The design combines a high
measure of internal recovery and flow quality with low external drag. Control of the
14
Figure 1.10 Concorde Supersonic Transport28
The preceding five sections detailed the inlet design and operation for five
distinct supersonic systems. The designs showed that both two-dimensional (XB-70
and Concorde) and axisymmetric (SR-71, D-21 and HRE) schemes are viable options
for the design of a supersonic inlet. With the exception of the original D-21 inlet,
15
spike, hinged doors or a variable throat. Variable geometry was needed, of course,
because the range of flight speeds for the vehicle (or engine) required operation under
1.3 Objective
in conjunction with TBCC engines. Instead of designing an inlet from scratch, the
previous TBCC-like engines. The process will examine the effects of overspeeding a
known inlet, and then incorporate design changes from lessons learned in an attempt
to increase its operational range into the hypersonic flight regime. To this end, the
limit of previous high-speed inlet designs will be explored, to both determine their
For this present work, the benchmark for a TBCC-like engine was chosen to
be the Pratt & Whitney J-58D used on the Lockheed SR-71 Blackbird previously
discussed in Section 1.2.1.1. As stated above, the inlet will be analyzed for its
traditional use to fully understand its normal operational characteristics and will then
be oversped, employing design changes in order to improve its high speed capabilities
and to conceptualize the design needs for future TBCC inlets. Obviously, the J-58D
could not work much past its intended design speed; however, for the purposes of this
thought experiment, the modified inlet(s) will be decoupled from the engine and
16
the inlet. The advantage this type of analysis has over higher order computational
methods, such as computational fluid dynamics (CFD), is that MOC solutions are
much faster and the results are more readily available. However, CFD will be used to
validate the characteristics model developed for this project and to preview the effects
of viscosity on the system. Additionally, the effects of drag and added installation
system within a duct. The functions developed within the program are
performed. The inlet geometry was reverse engineered from NASA YF-
flow properties are found for flight speeds of Mach 1.7 to 3.2. Final flow
• Various methods to modify the SR-71-type inlet are surmised with the
17
supersonic and subsonic conditions are resolved at the exit plane.
18
2 Model Development
All of the inlets considered in this report consisted of a centerbody spike acting
as the lower surface and a cowl acting as the upper boundary. A method of
begins by calculating the supersonic flow over a cone, with the condition that the
external tip shockwave lay outside of or on the cowl leading edge. From that point
on, the remainder of the flow is analyzed by calculating the reflected internal
shockwaves until the point where a normal shockwave was prescribed – the
downstream regions of subsonic flow will not be considered here. The ratio of
specific heats (γ) was held constant throughout the procedure and level, steady flight
was assumed at zero angle of attack. All cases were run on a Dell Latitude D600 with
a 1.6 GHz Pentium M processor and 1 GB of RAM. The initial data line always
begins with 402 points—typical run times were on the order of 15-20 seconds.
19
2.2 Method of Characteristics
reviewed. The equations governing steady, inviscid, supersonic flow are hyperbolic
equations for axially symmetric flow. The characteristic lines for two-dimensional or
axially symmetric flow are Mach lines as shown in Figure 2.1. Solutions can be
found only along these lines, i.e. information can only propagate downstream in
supersonic flow. For the present work, solutions to the equations governing
supersonic, axially symmetric flow will be used. The equations, as shown by Ferri30,
are briefly introduced below. The following sections are meant only as a brief
C+
m
V
A
q
m
C-
20
2.2.1 Supersonic Potential Flow with Axial Symmetry
In a velocity field, if the curl of the velocity is equal to zero at every point
within the flow, the flow is called irrotational. This implies that each fluid element
has no angular velocity. This type of flow is also called potential flow. The velocity
potential applies to one, two and three-dimensional irrotational flows only. The
partial differential equation governing supersonic potential flow with axial symmetry
is
φ x2 ∂ 2φ φ r2 ∂ 2φ 2φ xφ r ∂ 2φ v
1 − 2 2 + 1 − 2 2 − 2 + =0 (2.1)
a ∂x a ∂r a ∂r∂x r
dr
= tan(θ + µ ) (2.2)
dx
and the – (minus) characteristic (also known as the second family) is defined by
dr
= tan(θ − µ ) . (2.3)
dx
2
Vl 1 2
= 2 = 1+ sin 2 ( µ ) (2.4)
V W γ −1
21
dW dx sin( µ ) sin(θ ) tan( µ )
+ dθ tan( µ ) − = 0. (2.6)
W r cos(θ − µ )
Equations (2.5) and (2.6) can be numerically solved using first order differences for
the derivatives.
uniform flow is turned into itself, the resulting shockwave is curved. When a
the shock is no longer isentropic. Therefore, the flow must be considered rotational.
ψ r = ru (1 − W 2 )γ −1
1
(2.7)
ψ x = −rv(1 − W )
1
2 γ −1
and
vx − ur
f (ψ ) = − . (2.8)
r (1 − W )
γ
2 γ −1
The following equation can be obtained from the equations of state, continuity and
22
u 2Vl 2 2uvVl 2 v 2Vl 2 ψ
1 − 2 ψ xx − ψ xr + 1 − 2 ψ rr − r −
a a2 a r
(2.9)
W 2Vl 2
r 2 (1 − W 2 )γ −1
γ +1
− 1 f (ψ ) = 0
a
2
After a considerable derivation, the characteristic lines become the same as Eqns.
(2.2) and (2.3) for the C+ and C- characteristics, respectively, and the compatibility
equations become
Again, just as with the potential flow equations, Eqns. (2.10) and (2.11) are solved
using first order differences for the derivatives. However, the entropy derivative must
be handled in a particular manner and is outlined in Ref. 30. Comparing Eq.’s 2.5
and 2.10 along with Eq.’s 2.6 and 2.11, the rotational flow equations are the potential
flow equations with an entropy correction term. This is expected since (for two-
flowfield within the inlet including flowfield points, boundary points and shockwave
points.
23
2.3.1 Initial Data Line
initial characteristic line in which all the flow properties are known) must be found.
For all of the cases analyzed in this report, the initial data line was found by solving
for the flow over a cone. The solution to supersonic flow over a cone for any given
freestream Mach number and cone angle is given by the Taylor-Maccoll equation31
γ −1
2
[V 2
max
]
2 dV
− Vr2 − Vθ 2Vr + Vθ cot (θ ) + θ −
dθ
(2.12)
dV dVθ
Vθ VrVθ + r =0
dθ dθ
where Vr is the radial component of the velocity, θ is the angle from the centerline of
the cone to the ray of the solution, and Vmax is the maximum theoretical velocity
where ho is the freestream total enthalpy. The normal component of the velocity can
be found by
dVr
Vθ = Vr' = . (2.14)
dθ
V
V = (2.15)
Vmax
24
γ −1
2
dV dVr d 2Vr
1 − Vr − r 2Vr + cot (θ ) + −
2
2 dθ dθ dθ 2
. (2.16)
dVr dVr dVr d 2Vr
Vr + = 0
dθ dθ dθ dθ
2
Equation (2.16) can be rearranged and solved for the second derivative of the
ray angle, the normalized radial velocity and the first derivative of the normalized
radial velocity. The equation can be integrated numerically, for instance using the
Anderson33, the freestream Mach number and cone angle are prescribed and the
Now that the shockwave angle is known, the initial data line can be
calculated. From the Taylor-Maccoll equation, the flow properties (Mach number
and flow angle) can be found as a function of ray angle. To start the solution, these
flow properties need to be transformed into physical (x, r) coordinates, i.e. the
creation of an initial data line. Conical flow is non-uniform, meaning the flow angle
relative to the axis decreases (and Mach number increases) from the surface to the
shock. Therefore, the initial characteristic line is curved. The angle from the tip of
the centerbody to the leading edge of the cowl is known as the capture angle as shown
in Figure 2.2. This angle is always less than or equal to (in the case of shock on lip)
ray angle increments from the shockwave until the cowl. Two methods to generate
25
the initial data line from the cowl have been investigated. The first is the simplest but
has some drawbacks. Using the geometry as shown in Figure 2.3, the initial data line
is generated by marching in equal ray angle increments from the cowl (the first point
i) and intersecting the C+ characteristic with the ray angle at the next point (i+1). The
angle of the C+ characteristic line is the average of the flow angles plus the average of
the Mach angles of the i and i+1 points. The process is repeated until the centerbody
surface is reached. The drawback of using this method is that the mesh tends to be
highly concentrated towards the surface, creating an unequally spaced mesh—this can
tend to incur errors when marching downstream using the method of characteristics.
Instead of marching in equal ray angle increments, the second method divides the
initial characteristic into equal changes in x and solves for the subsequent changes in
ray angle at each point—again, the process marches from the cowl to the centerbody.
This produces an initial characteristic that is equally spaced in the x-direction from
Cowl
b First Characteristic
Centerbody
qcap
26
i
.5(mi+mi+1)+.5(qi+qi+1)
i+1
qray,i+1
From the initial data line, the solution proceeds to calculate the next
characteristic line and the internal reflected shock system, an example of which is
shown in Figure 2.4. As the characteristics net is formed, the shock waves are
propagated through the system. In the first zone, the area between the initial data line
and the first reflected cowl shock, the flow is isentropic so the potential flow
compatibility relations are used. In the remaining zones, since all of the reflected
shockwaves will be curved, the rotational flow compatibility relations are used [In
reality, the compatibility relations found from the stream function can be used for the
entire flowfield]. The remaining subsections discuss how each point in the flow is
handled.
27
Cowl
Initial
Data
Line Shockwaves
Centerbody
Figure 2.5. The nearest point to the boundary on the working characteristic (the C+
characteristic in the figure) is used to find the next boundary point. The C-
characteristic from point 1 is intersected with the boundary (at point 2’) as shown by
the dashed line. At this new location, the flow angle is set equal to the angle of the
surface (since all streamlines must be parallel to the surface) and the correct
compatibility equation is solved at this point for the remaining variables (Mach
number, entropy, Mach angle, etc.)—if the flow is irrotational (i.e. in the region
upstream of the first reflected shock), Eqn. (2.6) is used; otherwise, if the flow is
rotational (i.e. downstream of the first reflected shockwave), Eqn. (2.11) is used.
Note that if the working characteristic is the C- curve, then the C+ characteristic is
intersected with the boundary and the respective compatibility equation is solved.
approximating the characteristic curve with a straight line. A correction can be made
28
in which the C- characteristic line slope from point 1 is set to the average of the flow
angles and Mach angles at both points 1 and 2’ and the new flow properties and the
new point 2’ are found. This process is repeated until the location of point 2’ does not
instances the solution tended to diverge. The equations used to calculate the
C+
C-
1
2 2
Figure 2.6. The C+ characteristic from point 2 (which could be the first boundary
point or another flowfield point) is intersected with the C- characteristic from point 1
which is located on the previous data line. Point 3’ is the intersection of these lines.
At this point, no flow properties are yet known. A coupled solution of Eqns. (2.5)
and (2.6) (for potential flow) or Eqns. (2.10) and (2.11) (for rotational flow) is found.
From the solution, the flow properties (flow angle, Mach number, entropy, etc.) at
29
point 3’ are found. Again, using the updated properties at point 3’, a correction is
found setting the slopes of the characteristic lines to be the average of the upstream
and downstream Mach lines. The process is iterated until the location of point 3’ does
not change within a margin of error at successive iterations. As mentioned earlier, the
problems. This procedure is repeated throughout the remainder of the flowfield along
the working characteristic until point 1 is the final point on the previous data line.
One of the drawbacks of the method of characteristics, is that at each successive step,
the new data line is one point shorter in length; however, with a sufficiently fine
starting grid, the resulting decrease in numerical accuracy because of the lost data
point is not an issue. The process is simply reversed when the working characteristic
section.
30
C+
C-
shock also is generated when two characteristic lines of the same family cross each
other. In the present model, the shockwaves are fit to the system, i.e. the shock is
treated as a discontinuity in the flowfield. The following sections discuss how the
shocks are propagated through the flowfield in the present work. All of the internal
The location of the first reflected cowl shock is always known a priori. It is
the leading edge of the cowl. At the leading edge, the upstream flow properties, as
well as the angle of the cowl surface, are known values. Since the flow downstream
31
of the shock must be parallel to the surface, the shock deflection angle is set equal to
the difference between the upstream flow angle and surface angle. The shockwave
angle is determined by solving the θ−β−M equation which is presented, along with
the shockwave thermodynamic properties, in Section 2.4. With the origin of the
The procedure to calculate the second shock point (the point nearest to the
boundary) is shown in Figure 2.7. The shockwave segment from the origin is
intersected with the next characteristic line (in the case of the figure, the next C-
upstream of the shock are found by a linear interpolation of the data at the two points
before and after point 1. The deflection angle is assumed to be the same as the at the
previous shock point. The properties behind the shock are calculated using the
oblique shock relations. Next, using the new downstream shock properties, points 2
and 3 in the figure are found by sending the C+ and C- characteristics, respectively, to
intersect the surface and solving each respective compatibility relation to find the
flow properties at both points. The characteristics at points 2 and 3 are then sent back
and intersected to find updated downstream properties at point 1. The new flow angle
is then used to calculate a new deflection angle at point 1. The process of calculating
the shock properties and finding points 2 and 3 is repeated using the new deflection
angle. This procedure is iterated until the downstream flow angle and flow angle
found from the intersection of the characteristics from points 2 and 3 agree. For
further correction, the location of point 1 is then updated by averaging the slope of the
shock at the origin and the new shock angle at point 1. The entire process is repeated
32
until the location of point 1 and the properties of point 1 do not change (within a
starting point of the next downstream C+ characteristic and is used to calculate the
next shock point in the flow (as is discussed in the next section). The procedure is
simply mirrored if the shock reflection is off of the upper surface instead of the lower
1
Impinging Reflected
shock shock
C-
C+ Surface
3
2
The procedure to calculate the shock points within the flowfield is very
similar to that of the previous section and is shown in Figure 2.8. Points 0 (the
previous shock point) and 2 (the previous downstream flowfield point) can be thought
intersecting the shock with the next working characteristic line. Again, the deflection
angle is assumed to be the same as that of the previous shock point. The properties
behind the shockwave are found using the shock relations. Point 3 is found by
intersecting the C- characteristic with the C+ characteristic from point 2 and solving
the coupled compatibility relations to find the flow properties at point 3. Point 2’ is
found by intersecting the C+ characteristic from point 1 with the C- characteristic that
33
runs from point 0 to point 2. The properties at point 2’ are found by a linear
interpolation of the data at points 0 and 2. Then, as in the previous section, the
characteristics from point 3 and 2’ are sent back to point 1 and a new flow angle at
point 1 downstream of the shock is found. Again, the entire process is repeated until
the flow properties and location of point 1 do not change with successive iterations.
This process is repeated until the shock propagates to the other boundary, in
which case the location of point 1 is found by intersecting point 0 with the surface.
(points 2 and 3 in the below figure) will inevitably reflect off of the shock boundary.
In the event that point 3 crosses the shock boundary, the method of characteristics (as
a new characteristic line. The procedure to find points 1 and 3 is then repeated. The
more curved the shock, the more times a new downstream characteristic line will
need to be found. Once the shock is transmitted through the flowfield to the other
boundary, the downstream characteristic line that was found when calculating the
shock propagation becomes the new working characteristic line that is used to
1
C-
3
Shockwave
C+
0 C+
2
2
34
2.3.3.3 Lambda Shock and Triple Point Calculation
(in the case of this report) the flow is tripped by a normal shockwave as is shown in
Figure 2.9. At point A, it is assumed there is a step (as shown in the figure) in the
cowl which would trigger the normal shock. From point A, a normal shockwave
(shock angle of 90 degrees) is propagated until it intersects with the last reflected
shockwave at point T. Both types of intersection are shown in the figure. This is an
the trip point and would position itself according to the back pressure in the inlet.
However, for the purpose of this research, assuming the constant 90 degree normal
shock is adequate.
generated and a slip line (shown by the dashed line at point T) is formed. Across a
slip line, flow angle and pressure must equal but temperature and velocity are not
these criteria is impossible without a higher order solution. It was chosen to match
flow angle instead of pressure. This decision was made because the slip line could, in
theory, adjust quickly downstream of the triple point to match the pressure since the
flow is subsonic and information is able to pass across the slip line. The alternative
situation of having the flow angle be different at a single point is more non-physical.
As a result, the reflected shock angle is determined by matching the flow angle at the
triple point. This shock angle (another approximation) is held constant for the
remainder of the reflected shock propagation until it intersects the centerbody from
35
point T. At this stage, the flow is subsonic and the program is complete.
Trip Location
Cowl A Cowl A
T
T
Centerbody Centerbody
surface is designed such that small increases in surface angle compress the flow
without causing the characteristic lines to cross (causing shock waves to form).
Ideally, this is done so that all working characteristic lines intersect at the cowl
leading edge as shown in the figure. From the initial data line, as in Section 2.3.2.1,
the nearest point to the boundary is sent to intersect the surface. However, the second
surface point is an unknown at this point. To start the solution, two guesses are made,
a change in surface angle that is large and one that is very small. The remaining
flowfield points are calculated for both starting guess. Then, using the false position
method32, the change in surface angle is iterated until the next characteristic line
intersects at the cowl leading edge. This process is repeated until the desired surface
angle and/or surface Mach number is reached. This procedure was validated by
successfully testing the program with the design cases used in Reference 34.
36
q
3
q
2
q
c
properties for a calorically perfect gas from point-to-point. The equations were found
by relating the adiabatic total properties between two points 1 and 2 and then relating
γ −1 γ −1 2
T2 T1 Pt 2 γ 1+ M1
= 2 (2.18)
T∞ T∞ Pt1 1+ γ −1 M 2
2
2
37
1
γ −1 2 γ −1
γ 1+
1
ρ 2 ρ1 t 2
P M1
= 2 (2.19)
ρ ∞ ρ ∞ Pt1 1 + γ − 1 M 2
2
2
Of course, when the flow is isentropic between two points the total pressure ratio is
equal to 1. The mass flow ratio (defined as the ratio of the local density times
velocity to the freestream density times velocity) is found from the conservation of
mass as
A∞ ρu ρ Mi Ti ρi M i Pi
= i i = i = (2.21)
Ai ρ ∞u∞ ρ ∞ M ∞ T∞ ρ∞ M ∞ P∞
The well known θ−β−M relation that describes the relationship between
upstream Mach number, deflection angle and the corresponding shockwave angle for
approach. However, closed form solutions have been developed. The solution given
by Wellmann35 provides a very fast and accurate solution to the equation36 and is used
38
through a shockwave are given by33
P 2γ
P2
= 1 1 + (M 12 sin 2 (β ) − 1) (2.23)
P∞ P∞ γ + 1
ρ2 ρ (γ + 1)M 12 sin 2 (β )
= 1 (2.24)
ρ ∞ ρ ∞ (γ − 1)M 12 sin 2 (β ) + 2
(γ + 1)M 12 sin 2 (β )
−1
T2 T1 2γ
= 1+ (M 1 sin (β ) − 1)
2 2
(2.25)
T∞ T∞ γ + 1 (γ − 1)M 1 sin (β ) + 2
2 2
−1 γ
and
2
M 12 sin 2 (β ) +
1 γ −1
M2 = (2.27)
sin( β − θ ) 2γ 2
γ − 1 M 1 sin (β ) − 1
2
where the subscript 1 corresponds to the condition before the shockwave and the
39
3 SR-71-Type Inlet Analysis & Results
representative geometry had to be inferred from various YF-1237,38 reports and the
SR-71 Flight Manual. Figure 3.1 is a detailed schematic of the YF-12 inlet. The
dimensions given in the figure are in terms of x/R (length normalized by the radius of
the cowl). The radius of the cowl was determined by assuming that the inlet had zero
spillage at the design Mach number. The captured stream tube area at the cruise
Mach number was given as 18.5 ft2 in Section 1.2.1.1. From that number, a cowl
radius of about 2.43 ft. was assumed—this allowed the true dimensions in Figure 3.1
to be found. From Ref. 37 and the newly determined dimensions of the figure, the
angle of the conical spike was found to be 13o. The shoulder of the inlet (the
transition from the linear spike to the linear rear section of the centerbody) was
approximated by a cubic spline. The inner cowl was assumed to be a circular arc.
After several iterations of the method of characteristics solution, the initial cowl angle
was found to be 3o—at this angle the reflected shock waves remain attached at Mach
1.7 (the starting Mach number) but not at Mach 1.6. The resultant geometry is shown
40
in Figure 3.2, where the solid line is the position of the spike in the most forward
position and the dashed line is the position of the spike in the most aft (at the cruise
speed) position. [Note the step in the figure at about x = 3.5 is the location of the
throat section where the computer code assumes the start of a lambda shock
structure].
41
3.2 Self-Starting Characteristics
Since the geometry is now known, the area across the cowl face and at the
throat can be calculated. Using those two areas and the mass-averaged Mach number
entering the cowl (found from the solution to the Taylor-Maccoll equation), the self-
starting characteristics are calculated using the Kantrowitz limit, Eq. (1.2). Figure 3.3
is the resulting plot. As the figure shows, the inlet does indeed self-start at Mach 1.7
(i.e. the area ratio is below the Kantrowitz limit) and then proceeds to move away
from the Kantrowitz limit, in parallel with the isentropic limit. Note this figure does
not take into account any of the bleed or bypass losses that would invariably change
the amount of mass flow in the internal contraction section and change the properties
of the figure.
It should also be noted that the properties of the SR-71-type inlet (and the
proposed modifications) were calculated under the assumption that the incoming air
is uniform and traveling at the flight Mach number. In reality, the flow will be
perturbed depending upon the placement of the inlet on the aircraft. In the real case
of the SR-71, the flow going into the inlet is disturbed by the conical shock coming of
off of the nose (~18o cone) of the vehicle. CFD analysis39 performed on the SR-71
showed that the nose shock fell outside of the inlet at the design speed of Mach 3.2 as
shown in Figure 3.4. Johnson and Montoya37 state that on average the local Mach
number at the spike tip of the YF-12 (from flight data) was never more than 3 percent
less than the freestream Mach number. This perturbation is less than the theoretical
Mach number behind a conical shock off of an 18o cone; however, in looking closely
42
at Figure 3.4, it does appear that the shockwave has weakened before it draws near to
the inlet, most likely due to the expansion waves generated on the lateral surfaces of
the vehicle. Nevertheless, as previously stated, the effects of the nose shock are
43
Figure 3.4 Full Vehicle Mach 3.2 CFD Shockwave Patterns (top view)39
Figures 3.5 through 3.9. The figures show contours of Mach number, total pressure
ratio, temperature ratio, pressure ratio and mass flow ratio for freestream Mach
numbers of 1.7, 2.0, 2.5, 3.0 and the final shock-on-lip design point of Mach 3.2.
Note the line emanating off of the tip of the centerbody is the spike tip shock and the
first zone of each contour plot represents the first captured, left running characteristic,
shockwaves. Also note that the contours only show the supersonic conditions prior to
the lambda shock system (plotting a sliver of subsonic flow would distort the contour
levels).
Starting at Mach 1.7, the contours indicate that the initial reflected shock off
of the cowl leading edge intersects the centerbody on the shoulder (i.e. the curved
44
segment of the centerbody), causing the flow to re-expand to a higher-than-freestream
Mach number (locally). This increase in local Mach number allows the second
reflected shock to remain attached (Note the previous discussion of determining the
angle of the cowl leading edge). This also produces a near-sonic (choked flow)
condition just downstream of the second reflection system. Because the shoulder
continues to turn, the flow is able to expand, and subsequent shocks can be formed.
The remainder of the shock train progresses (the solutions shows about 6-7 total
reflections) and the effect of the expansion waves propagating through the system is
readily apparent, especially in the temperature plots. Due to the presence of the
expansion waves and shock train, temperature and pressure end up being near to or
less than the freestream values at the throat entrance at the lower Mach numbers.
As the freestream Mach number increases and the spike retracts, the plots
show the progression of the shock train through the duct. The initial reflection off the
cowl leading edge gradually moves off of the shoulder to the conical segment of the
centerbody stopping any expansion prior to the first reflection. The expansion wave
angles also begin to decrease as is shown by the movement of the expansion system
down the duct—this is to be expected with higher speeds. By Mach 2.5, the initial
cowl shock has moved onto the conical section of the centerbody and the total
reflections occur. Local regions of high temperature and pressure on the cowl are
also apparent at Mach 3.2 but by the time the flow nears the throat, this is effect is
45
Figure 3.5 SR-71 Inlet Mach Number Contours
46
Figure 3.6 SR-71 Inlet Total Pressure Ratio Contours
47
Figure 3.7 SR-71 Inlet Temperature Ratio Contours
48
Figure 3.8 SR-71 Inlet Pressure Ratio Contours
49
Figure 3.9 SR-71 Inlet Mass Flow Ratio Contours
50
3.4 Throat Conditions
The area-averaged supersonic and subsonic properties at the throat are shown
in Figures 3.10 and 3.11, respectively. The levels of distortion are quantified as well.
The supersonic plot shows a gradual rise in pressure, temperature and mass flow with
freestream Mach number. The supersonic throat Mach number initially decreases and
then increases (with a final throat entrance Mach number of about 1.8) whereas the
total pressure ratio initially increases and then decreases as the freestream speed is
increased. The total pressure ratio distortion is small and always stays below three
percent. Distortion levels for Mach number and temperature are moderate, remaining
at or below 20 percent. Pressure and mass flow distortion levels are large, rising to
~50 percent and ~30 percent, respectively. However, this is to be expected as the
area averaging for the supersonic throat conditions takes place across a shockwave
The subsonic plots agree show that Mach number initially increases and then
decreases (with a final throat exit Mach number of about 0.61) whereas all of the
other flow properties have the same general trend as the supersonic plot. However,
the levels of the distortion of the total pressure ratio increased significantly while the
levels of distortion of Mach number, temperature and pressure decreased from the
supersonic case (both mass flow plots are identical, as they should be). Any errors in
the subsonic properties are a direct result of the assumed lambda shock structure
As shown, the subsonic temperature and pressure ratios at Mach 3.2 are about
51
2.8 and 30, respectively. The pressure ratio is lower than the value of 40 reported by
Kelly Johnson (former head of the Skunk Works)18. The temperature ratio is below
the maximum compressor inlet temperature ratio of 3.23 (assuming flight at 90,000
ft.)13. However, real gas and viscous effects have been ignored. Additionally, there
will be changes to the flowfield as the air moves through the subsonic portion of the
inlet prior to entering the compressor. If the flow (at a static pressure ratio of 30)
were isentropically compressed to zero velocity, the pressure ratio would rise to 39,
52
Figure 3.10 Supersonic SR-71 Throat Properties
53
Figure 3.11 Subsonic SR-71 Throat Properties
54
4 CFD Validation
The accompanying grid generator is called OVERGRID41. Its many features include
a hyperbolic grid generator and the ability to create separate, overlapping grids.
The grid used in the validation is shown in Figures 4.1 (external mesh) and 4.2
(enhanced internal mesh) with every fifth grid point removed for clarity. It consists
of 701 points in the x direction and 101 points in the z. The grid is stretched at both
ends of vertical plane to accurately capture the boundary layer (when solving the
viscous equations). The grid is also stretched at the x = 4 (on the centerbody surface)
and x = 10 (on the cowl surface) to improve the resolution of the shockwaves. For
the inviscid analysis, the cowl and the centerbody boundary conditions were set to be
inviscid adiabatic walls. The left plane was set to freestream conditions (Mach 3.2
flow at 90,000 ft. altitude) and the right exit plane was set to be an outflow (with
55
pressure extrapolation). The top horizontal plane between the left plane and the cowl
leading edge was set to be a simple supersonic inflow/outflow. The solution used
central differencing (fourth order) and the ARC3D 3-factor diagonal scheme. The
CFL number was determined locally. The ratio of specific heats was also held
constant at 1.4. As with the MOC cases, the code was run on a Dell Latitude D600
The results are shown in Figures 4.3 through 4.5 and compared to the contour
plots presented earlier in Chapter 3 of Mach number, temperature ratio and pressure
ratio. Note the MOC contours are scaled differently (to match the CFD contour
levels) from the plots in Section 3.3 because the CFD includes the freestream
properties whereas the MOC contours do not. The two solution methods agree quite
temperature rise and a minor under-prediction (.4 %) of the maximum pressure rise,
the MOC solution obtained almost the same results as the CFD. Note in the CFD
results, the maximum contour level of Mach number is higher than the freestream
value because the CFD had some difficulty resolving the shock at the spike tip.
Similarly, the pressure and temperature minimum level is less than freestream. Also
included are the prediction of the surface Mach number, temperature and pressure in
Figures 4.6 through 4.8. Again, excellent agreement between the CFD and MOC
solutions is obtained. The shock locations are accurately predicted by the MOC code
(shown by the discontinuities). As previously mentioned, the run times for most
cases of the MOC code (depending on the number of shock reflections) was on the
order of 15-20 seconds (with a grid 4 times as fine) whereas the CFD (at 500
56
iterations) results presented here took a little over 3 minutes.
Figure 4.1 External Grid used for CFD Analysis (1/5th the resolution)
Figure 4.2 Internal Grid used for CFD Analysis (1/5th the resolution)
57
Figure 4.3 Mach Number Contour Comparison (CFD above)
58
Figure 4.4 Temperature Ratio Contour Comparison (CFD above)
59
Figure 4.5 Pressure Ratio Contour Comparison (CFD above)
60
Figure 4.6 Surface Mach Number Comparison
61
Figure 4.8 Surface Pressure Ratio Comparison
Full viscous analysis of the SR-71 inlet was performed using OVERFLOW2.
The turbulence model used was the one-equation Spalart-Allmaras RT model (fully
turbulent). The results are presented in Figure 4.9. [Note these results are displayed
for qualitative purposes only. Each contour shown does not represent a converged
solution, i.e. this is not time accurate CFD.] At 500 iterations, the boundary layers
along both the cowl and centerbody are developing. At the impingement location of
the initial reflected cowl shock, the boundary layer separates and then reattaches—
similar results are seen in the shock train downstream. However, at 1000 iterations,
the boundary layer along the centerbody appears to have separated even further, but
62
an oblique shock train still exists. At 1500 iterations, the centerbody boundary layer
has almost fully separated, and the shock train downstream has transitioned into a
normal shock train. Finally, at 2000 iterations, the inlet has completely unstarted as
evidence by the subsonic flow present throughout the internal duct. The final contour
plot shows a second oblique shock is generated on the centerbody surface (outside of
the cowl face) followed by a very strong oblique shock that causes the entire inlet
flowfield to be subsonic. These results show why the original SR-71 inlet used a
bleed system on the spike slightly downstream of where the initial reflected cowl
shock impinged on the centerbody. Additionally, the results demonstrate why the
designers selected a design such that the impingement location remained relatively
63
Figure 4.9 Viscous CFD Results
64
5 Modification Analysis & Results
The next six sections detail a variety of means to modify the SR-71 inlet to
extend its operation into the hypersonic flight regime. With no modification to the
inlet, the centerbody spike shock will move inside of the cowl (past Mach 3.2) and
the resulting inlet flowfield would likely degrade. The effect on the overall system
(especially mass and installation factors) due to the design changes are not be taken
into account at this time. Just as with the original SR-71 inlet, the self-starting
characteristics of each scheme are analyzed and, for those inlets that are able to start
well past Mach 3.2, the full flow properties will be determined. The constraints on
these changes are that the slope of the duct after the throat remains fixed and the new
designs have to fit within the original SR-71 cowl. The throat is located at the same
axial position as was described in the original SR-71 inlet. Accordingly, the flow is
that the SR-71 flew upwards of Mach 3.5 and that the Skunk Works performed
studies to see how fast the SR-71 could actually fly. The websites cited the limiting
factors to be the interaction of the nose conical shock with the inlet spike and the heat
65
load on the airframe past Mach 3.5. However, these published reports were never
located and the modifications of the inlet (if any) are not known to the author.
Additionally, declassified CIA reports indicate that the maximum speed obtained by
the SR-71 was actually Mach 3.2944. Note that previously mentioned NASA reports
said that the inlet Mach number was generally 3 percent less than the flight Mach
most obvious way to overspeed the inlet (and the least expensive) would be to re-
the higher Mach numbers. Assuming the spike could be extended to its original low
speed location (and fixing the conical half angle at 13o), the inlet could be oversped
up to a freestream Mach number of about 6.1. Beyond this point, the conical
shockwave generated off of the spike would move inside of the cowl.
The self-starting characteristics (i.e. the ratio of throat area to cowl area
compared to the Kantrowitz limit) of the first redesign scheme are shown in Figure
5.1. As is shown, the inlet would violate the Kantrowitz limit near a freestream Mach
number of 4.8. The violation occurs because as the spike is moved forward the throat
area decreases while the cowl area increases. This causes a rapid increase in the area
66
Figure 5.1 Self-Starting Characteristics of the Re-extended Spike
Contours of the flowfield of the re-extended spike are shown in Figures 5.2
through 5.6 for a range of freestream Mach numbers within the self-starting regime.
The contours show that as freestream Mach number increases, the impingement
location of the first reflected cowl shock advances onto the shoulder. This effect,
again, causes expansion waves to be generated prior to the first reflection off of the
centerbody. As the spike retracts and the initial reflection moves up and around the
centerbody shoulder, local regions of high temperature and pressure are present along
the cowl; however, these effects are eventually overcome by the presence of the
expansion waves. The expanding of the duct also restricts the number of
reflections—as is shown by Mach 4.0, two total reflections exist prior to the throat
67
section. By Mach 4.5, the leading edge shock has moved around the shoulder and by
Mach 4.8, it is on the opposite side of the shoulder and is highly curved. Large
expansion regions exist prior to the first reflection and large gradients in the flow
68
Figure 5.3 Re-extended Spike Inlet Total Pressure Ratio Contours
69
Figure 5.4 Re-extended Spike Inlet Temperature Ratio Contours
70
Figure 5.5 Re-extended Spike Inlet Pressure Ratio Contours
71
Figure 5.6 Re-extended Spike Inlet Mass Flow Ratio Contours
Gas properties at the throat for the re-extended spike are shown in Figures 5.7
and 5.8. Because of the limited number of shock reflections, the effective entry Mach
number remains one less than the flight Mach number for the entire flight regime.
72
Entry total pressure losses are small (the total pressure ratio remains above 0.86 at
Mach 4.8), mostly because of the small number of shock reflections. Entry
temperature, pressure and mass flow ratios decrease to about 1.6, 4 and 2.3,
highly distorted (the pressure and mass flow distortion levels are at 140 and 75
percent at Mach 4.8, respectively). This is expected because of the large curvature in
The Mach number at the throat exit is rather low, beginning below Mach 0.6
and ending at ~0.45 (which is good for ramjet operation) but because the entry Mach
numbers are high, the total pressure losses are substantial (the total pressure ratio is
~0.7 at Mach 3.3 and falls to ~0.2 at Mach 4.8). Exit pressure and temperature ratios
are also large because of the high entry Mach numbers (the temperature and pressure
ratios are ~5.7 and ~55 at Mach 4.8, respectively). The distortion levels for Mach
number and temperature ratio are acceptable (remaining below 10 percent for the
entire regime), but the levels of distortion for total pressure, static pressure and mass
flow ratio are all still very high (above 50 percent for most of the regime) at the
higher speeds.
73
Figure 5.7 Supersonic Re-extended Spike Throat Properties
74
Figure 5.8 Subsonic Re-extended Spike Throat Properties
75
5.2 Variable Cowl Leading Edge
The second proposed redesign scheme involves allowing the radial portion of
the cowl leading edge to vary while keeping the spike in its aft most position. This
would enable the leading edge of the cowl to be aligned with the conical shock (as is
shown in Figure 5.9) at every oversped Mach number. The dashed lines shown
(starting from top to bottom) illustrate the cowl position at Mach numbers of 3.3, 3.6,
figure shows, the self-starting performance is poorer than the previous redesign (the
inlet would unstart at about Mach 4). This occurs because of the rapid change in the
76
area ratio, caused by the decrease in the cowl area. This design would also introduce
a large penalty in added cowl drag. Because of the poor self-starting characteristics,
The major problem with both of the previous redesign schemes is the high rate
at which the cowl/throat area ratio increases as the inlet is oversped. One way to
alleviate this problem is to fix the throat area while re-extending the centerbody. In
order to perform this operation, the shoulder on the centerbody spike would need to
have the ability to widen as shown in Figure 5.11. The dashed lines (going from right
to left) in the figure represent the centerbody geometry at Mach numbers of 4, 5 and
6.1, respectively.
77
Figure 5.11 Diagram of the Re-extension with a Widened Shoulder
in Figure 5.12. The figure shows a marked improvement in the starting performance.
The curve still shows a sharp discontinuity but it also demonstrates that the inlet has
the ability to maintain starting capability up until the maximum oversped Mach
number of 6.1 discussed in Section 5.1. The inlet is mechanically limited by the re-
extension constraint imposed on the problem. If the inlet were allowed to re-extend
beyond the original low speed location of the SR-71 inlet, it would be able to remain
78
Figure 5.12 Self-Starting Characteristics of the Widened Shoulder Centerbody
Property contours are presented in Figures 5.13 through 5.17 for the widened
shoulder centerbody. The general trend is similar to that of the re-extended spike
except that the widened shoulder slows the progression of the initial reflected cowl
shock onto the shoulder. Unlike the re-extended spike, the reflected shock never
moves completely over the shoulder. However, both systems result in only a 2 shock
internal reflection system. Local regions of high temperature and pressure (the
temperature and pressure ratios rise to 2.4 and 15.24, respectively) are again apparent
on the cowl prior to being affected by the expansion waves. Since the initial reflected
cowl shock is not as curved for the widened shoulder centerbody, the gradients in the
flow properties are not as severe as they are with the simple re-extended spike.
79
Figure 5.13 Widened Shoulder Centerbody Inlet Mach Number Contours
80
Figure 5.14 Widened Shoulder Centerbody Inlet Total Pressure Ratio Contours
81
Figure 5.15 Widened Shoulder Centerbody Inlet Temperature Ratio Contours
82
Figure 5.16 Widened Shoulder Centerbody Inlet Pressure Ratio Contours
83
Figure 5.17 Widened Shoulder Centerbody Inlet Mass Flow Ratio Contours
84
5.3.3 Throat Conditions
The throat entrance and exit properties shown in Figures 5.18 and 5.19 show a
marked improvement in the flow quality over the re-extended spike. The supersonic
plots reveal that at similar speeds, the Mach number entering the throat is lower, the
temperature and pressure are higher, and the distortion levels are much lower for four
of the flow properties. The total pressure ratio is about the same entering the throat at
similar speeds. The mass flow ratio similarly decreases from 3.37 at Mach 3.3 to
3.32 at Mach 6.1, but the change is sufficiently small such that the mass flow ratio is
essentially constant. The plots also show that the temperature ratio starts at about 1.8,
decreases to about 1.75 near Mach 4, and then increases with a final ratio of about 2
at Mach 6.1. Similarly, the pressure ratio starts at about 7.7, and then decreases to
about 6.7 near Mach 4.5 and then increases with a final ratio of about 7 near Mach
6.1.
The subsonic plots also show that performance is better overall. Total
pressure ratio (starting at about 0.75 and ending below 0.2), Mach number (starting at
about 0.59 and ending near 0.43) and pressure ratio (starting at 30 and ending at about
140), for any given Mach number, are higher than the re-extended spike with slightly
lower distortion levels. The temperature ratio (starting at about 3 and ending at about
8), interestingly, is essentially equal, albeit with differing levels of distortion. In all
likelihood, transition to scramjet would occur somewhere along the flight path
because of the large increases in pressure ratio and the considerable losses in total
85
Figure 5.18 Supersonic Widened Shoulder Centerbody Throat Properties
86
Figure 5.19 Subsonic Widened Shoulder Centerbody Throat Properties
87
5.4 Variable Cone Centerbody
Yet another option would be to install a variable cone on the centerbody spike.
In this scenario, the spike would remain fixed its axial location at Mach 3.2 and then
increase its conical half angle to maintain the shock-on-lip condition. However, this
design has major limitations (beyond the obvious mechanical complexity) as shown
in Figure 5.20. The dashed lines in the figure are the change in the centerbody
of 3.3, 3.4, 3.5, 3.6 and 3.7. As the figure illustrates, the variable centerbody can only
of 3.7) before the throat would be completely closed. [Note this design still assumes
88
5.4.1 Self-Starting Characteristics
Besides the obvious physical problems with this inlet modification, Figure
5.21 shows that the self-starting characteristics of this inlet are very poor. The inlet
would violate the maximum isentropic contraction ratio limit (due to the decrease in
throat area) before it penetrates the outer cowl. As with the variable cowl leading
edge design, this inlet was not fully analyzed due to its poor starting performance.
The inlet shown in Figure 5.22 can be constructed by combining the variable
cone and re-extension aspects. In this situation, as the spike is pushed forward the
In the current design, the throat area remained fixed as well. Theoretically, this inlet
89
could maintain the shock-on-lip condition for any freestream Mach number (the
maximum cone angle needed would only be about 16o) before reaching the re-
extension limit. The dashed lines correspond to the centerbody geometry (starting
from right to left) at Mach numbers of about 3.5, 4, 5.25 and 8.25, respectively.
While the variable cone with re-extension inlet could be oversped to any
freestream Mach number and remain within the geometrical constraints, it would
violate the Kantrowitz limit at about Mach 7.2 (Figure 5.23). Again, this occurs
because of the change in throat/cowl area ratio. However, while it will not be
considered in the current report, the starting performance would likely improve if the
90
Figure 5.23 Self-Starting Characteristics of the Variable Cone with Re-extension
Contours for the variable cone with re-extension modification scheme are
shown in Figures 5.24 through 5.28. The new conical half angle is indicated on each
figure. Again, the basic structures of the flowfields are very similar to that of the
widened shoulder and the re-extended spike schemes. However, as the plots
shock is slower than that of the widened shoulder thus delaying the pre-reflection
expansion. Accordingly, the impingement locations of the both the initial cowl
reflected shock and the centerbody reflection appear to be constant from about Mach
5.5 through Mach 7.2. The temperature and pressure ratio plots show a rather
complex flow structure as the expansion waves propagate through the system creating
91
multiple regions of high and low temperature and pressure. The major difference
between this scheme and the two previously analyzed is that this design maintains the
92
Figure 5.24 Variable Cone with Re-extension Inlet Mach Number Contours
93
Figure 5.25 Variable Cone with Re-extension Inlet Total Pressure Ratio Contours
94
Figure 5.26 Variable Cone with Re-extension Inlet Temperature Ratio Contours
95
Figure 5.27 Variable Cone with Re-extension Inlet Pressure Ratio Contours
96
Figure 5.28 Variable Cone with Re-extension Inlet Mass Flow Ratio Contours
97
5.5.3 Throat Conditions
The throat entrance and exit conditions are shown in Figures 5.29 and 5.30,
respectively. For similar flight speeds, the entrance Mach number for the variable
cone with re-extension scheme is essentially the same as that of the widened shoulder.
Likewise, the temperature and pressure ratio entrance properties are similar, but the
total pressure ratio is lower. This makes sense because in this case the flow passes
through two stronger shockwaves and an additional reflected shock. Overall, the
distortion levels are slightly lower compared to the widened shoulder centerbody.
Mass flow decreases as well from 3.39 at Mach 3.3 to about 3.2 at Mach 7.2, but as
The subsonic features of the variable cone with re-extension inlet show
equivalent trends with the widened shoulder centerbody. The exit Mach number
starts at about Mach 0.6 and finishes at about Mach 0.43. The total pressure ratio is
effectively the same, starting at about 0.7 and ending below 0.1. The exit temperature
and pressure ratios are fairly consistent; however, the distortion levels are moderately
98
Figure 5.29 Supersonic Variable Cone with Re-extension Throat Properties
99
Figure 5.30 Subsonic Variable Cone with Re-extension Throat Properties
100
6 Summary and Conclusions
This thesis developed a new strategy for designing inlets for turbine-based
combined-cycle engines. Rather than starting the design process with a clean slate,
the performance of a specific supersonic inlet was analyzed and certain aspects of the
inlet were redesigned to see if the flight envelope could be pushed into the hypersonic
flight regime. A numerical model was developed using the axisymmetric method of
axisymmetric inlets. This results from the model were used to quantify important
properties of the flowfield (Mach number, total pressure ratio, temperature ratio,
pressure ratio and mass flow ratio), and their levels of distortion (spatial non-
uniformity), at the throat entrance. A simple lambda shock structure was then used to
approximate the flow properties leaving the throat. The Kantrowitz limit was
Several candidate inlets were examined initially, including those used on the
Boeing XB-70, Lockheed D-21 Drone, and the Concorde. However, this work
focused on Lockheed SR-71 inlet. The J-58D engines that powered the SR-71
effectively operated as a ramjet at the higher speeds since a good deal of the airflow
101
bypassed the combustor and turbine and was dumped straight into the afterburner.
The SR-71 inlet was chosen because of the TBCC-like qualities of the J-58D engine.
The inlet of the SR-71 is a mixed compression inlet that starts at Mach 1.7 and can be
run up to Mach 3.2. The inlet utilizes a translating spike and a series of bleeds and
bypasses to control the position of the terminating normal shock and to regulate the
Investigation of the SR-71 inlet proved that the inlet indeed satisfies the
Kantrowitz limit at Mach 1.7 and maintains the self-starting capability through Mach
3.2. The characteristics model showed that at the low Mach numbers several
reflected shock waves are present within the duct. As the inlet speeds up, the shock
train propagates through the duct so that at the design Mach number of 3.2, only two
internal reflections are present. The inlet performed rather well, losing less than three
percent of total pressure (with very low levels of distortion) prior to entering the
throat. Downstream of the throat, the total pressure losses stayed above about 23
percent and the Mach number remained nearly constant, varying from 0.66 to 0.62 at
flight speeds between Mach 1.7 and 3.2. Pressure, temperature, and mass flow ratios
gradually rose at the exit plane as the Mach number was increased. A final pressure
ratio of 30 was found at the shock-on-lip Mach number of 3.2. The Mach 3.2 inviscid
solution was validated using NASA’s OVERFLOW2 CFD tool. In addition, the full
viscous solution was performed using OVERFLOW2 and the results demonstrated
why the original SR-71 employed a boundary layer bleed on the centerbody—the
interaction of the initial reflected shockwave with the boundary layer caused the
102
Five modifications to the inlet were then proposed and their high speed
performance (> Mach 3.2) was evaluated. Two schemes, the variable cowl leading
edge and the variable cone centerbody, performed poorly—the former would unstart
at Mach 4.0, the latter at Mach 3.5. The simplest (and least expensive) redesign
method, the re-extended spike, demonstrated the ability to remain self-started until
about Mach 4.8. However, the flowfield properties were highly distorted because the
initial reflected cowl shock moved all the way around the shoulder of the centerbody
The remaining two schemes, the widened shoulder centerbody and the
variable cone with re-extension, showed some promise. The widened shoulder
centerbody was able to remain self-started (with a healthy buffer from the Kantrowitz
limit) up to the maximum constrained Mach number of 6.1, while the variable cone
with re-extension remained self started through Mach 7.2. Both inlets displayed
comparable flowfield properties at the exit plane as shown in Figures 6.1 and 6.2.
The Mach number entering the throat started around Mach 2.0 and increased at the
same rate for both inlets. Similarly, for both inlets the temperature ratio entering the
throat started at about 1.8, decreased, and then increased at about the same rate.
Likewise, the pressure ratio entering the throat started at about 7.7, decreased, and
then increased at similar rates. The main difference in the throat entrance properties
between the two schemes is that the total pressure ratio is consistently higher for the
widened shoulder centerbody and that the distortion levels for the variable cone with
re-extension scheme are steadily lower. Figure 6.2 illustrates that the area-averaged
subsonic flow properties are fundamentally the same for both schemes. The final
103
figure also shows that both of these inlets exhibited very large levels of static
temperature and pressure at the throat exit at higher speeds suggesting that transition
probably needed in order to design an inlet that is operable over a span of flight
speeds that ranges from the low supersonic regime into the hypersonic corridor. The
mechanical complexity associated with the complex variable geometry is the main
obstacle that would need to be overcome to realize either of the two promising
redesign schemes developed in this work. This point underscores the difficulty in
constructing a single common inlet that would properly feed the three cycles of a
TBCC engine. More detailed flow analysis, including viscous and boundary layer
characteristics of both concepts. Finally, coupling this numerical tool to gas turbine,
104
Figure 6.1 Comparison of Area-Averaged Supersonic Throat Properties
105
Figure 6.2 Comparison of Area-Averaged Subsonic Throat Properties
106
7 Future Work
venture, there are several aspects of the process that could use some improvement.
Recommendations for improvement to this research project are listed in the following
sections.
• The ratio of specific heats (γ) was held constant throughout the entire
assumed trajectory (or integrating the trajectory into the code) should be
107
included.
distorted flow and does not allow for a clear understanding of the design
and below the triple point to go along with the MOC solution is a must.
• Viscous effects (as seen in Chapter 4) play a major role in the design of
• Incorporate the calculation of drag (both internal and external) into the
solution.
108
• The current report uses the SR-71 inlet as a baseline and suggests only
five re-design schemes to make the inlet a true TBCC inlet. Obviously,
there are many other concepts that could be considered (such as isentropic
modification to the present work would be to allow for the throat area and
• Incorporate an engine model into the program so that the inlet can match
the engine needs. This allows for the coupling of the design of both the
• CFD analysis was not done until the very end of the project. In hindsight,
much earlier in the research process as the run times were not as slow as
originally perceived. This would allow for research into bleed and/or
bypass schemes for flow control to prevent the problems associated with
perform time accurate CFD analysis with a moving grid for the entire
range of SR-71 Mach number to allow the inlet to start at the low Mach
109
Additionally, it would an interesting project to somehow reverse engineer
the suction schedule of the original bleed system on the SR-71 inlet.
initial reflected cowl shock moving up and around the shoulder of the
110
Bibliography
4. Mahoney, J.J., Inlets for Supersonic Missiles, AIAA Education Series, 1990.
5. Fernandez, R., Reddy, D. R., Benson, T. J., Iek, C., Biesiadny, T. J., and
111
in Astronautics and Aeronautics, Vol. 189, Ed. Curran, E.T., and Murthy,
S.N.B., 2000.
10. Anderson, E. C., and Lopata, J. B. “Using a Modified SR-71 Aircraft and Air-
11. Anderson, E. C., Lopata, J. B., Hoar, P. R., Nelson, D., “Performance
Analysis and Optimized Design of an SR-71 Based Air Launch System for
Paper 75-1180.
15. Syberg, J, and Hickcox, T. E., “Design of a Bleed System for a Mach 3.5
TM-X-71779, 1975.
112
Shock-Position Sensor for Mixed Compression Inlets Evaluated in Wind
18. Miller, J. Lockheed Martin’s Skunk Works. Rev. Ed. Midland Publishing
8-21 Cycle II: DRACO Flowpath Hypersonic Inlet Design.” NASA TR-
113
Propulsion and Power. Vol. 20, No. 1, 2004.
28. Rettie, I. H. and Lewis, W.G. “Design and Development of an Air Intake for
and Jet Propulsion. Vol. VI. General Theory of High Speed Aerodynamics.
Sec. G. W.R. Sears, ed., Princeton University Press, 1954, pp. 583-668.
31. Taylor, G. I., and Maccoll, J. W., “The Air Pressure on a Cone Moving at
High Speed,” Proc. Roy. Soc., Vol. 139, 1933, pp. 278-311.
32. Canale, R. P., and Chapra, S. C., Numerical Methods for Engineers. 3rd Ed.
33. Anderson, J. D., Modern Compressible Flow with Historical Perspective. 3rd
34. Connors, J. F., and Meyer, R. C. “Design Criteria for Axisymmetric and Two-
36. vonEggers Rudd, J., and Lewis, M. J., “Comparison of Shock Calculation
1973.
114
38. Cole, G. L., Cwynar, D. S., Geyser, L. C. “Wind-Tunnel Evaluation of the
40. Buning, P. G., Jespersen, D. C., Pulliam, T. H., Chan, W. M., Slotnick, J. P.,
Krist, S. E., and Renze, K. J., “Overflow User’s Manual.” Version 1.8aa.
41. Chan, W. M., Chiu, I.-T., and Buning, P. G., “User’s Manual for the HYP-
GEN Hyperbolic Grid Generator and the HGUI Graphical User Interface,”
works/v02.n056
44. McIninch, T. P., “The Oxcart Story,” Central Intelligence Agency website
http://www.odci.gov/csi/kent_csi/docs/v15i1a01p_0023.htm
115