Sie sind auf Seite 1von 37

Prog. Aerospace Sci. Vol. 34, pp.

219—256, 1998
( 1998 Published by Elsevier Science Ltd
Printed in Great Britain. All rights reserved
0376—0421/98 $34.00
PII: S0376–0421(98)00005–0

SOLID FUEL RAMJET COMBUSTOR DESIGN


S. KRISHNAN and PHILMON GEORGE
Department of Aerospace Engineering, Indian Institute of Technology Madras, Chennai — 600 036, India

Abstract—Combustion aspects of solid fuel ramjet (SFRJ) are reviewed. On the point of view of the ability
of an SFRJ to operate satisfactorily at all off-design conditions the areas of concern to propulsion system
designer are (1) selection of a fuel type, (2) flame holding requirements that limit maximum fuel loading,
(3) understanding the fuel regression rate behaviour as a function of flight speed and altitude, (4) diffusion-
controlled combustion process and its efficiency enhancement, and (5) inlet/combustor matching. Consider-
ing these areas, the following aspects are reviewed from the information available in open literature:
(1) different experimental set-up conditions adopted in combustor research, (2) various suitable fuel types,
(3) flammability limits, (4) fuel regression rate behaviour, (5) methods of achieving high efficiency in
metallized fuel, and (6) various modelling efforts. Detailed discussion is presented on two different types of
regression rate mechanism in SFRJ: one that is controlled by the heat transfer processes downstream of the
reattachment region and the other by that in the region itself. With a view to demonstrate the use of the
information collected through this review, a preliminary design procedure is presented for an SFRJ-assisted
gun launched projectile of pseudo-vacuum trajectory. ( 1998. Published by Elsevier Science Ltd.

CONTENTS

NOTATION 219
1. INTRODUCTION 220
2. EXPERIMENTAL RESULTS 224
2.1. Experimental set-ups 224
2.2. Fuel types 228
2.3. Flammability limits 229
2.4. Regression rate 230
2.5. Combustion efficiency 236
3. MODELLING STUDIES 238
4. DESIGN OF AN SFRJ ASSISTED PROJECTILE 242
4.1. Analysis 244
4.1.1. Launch velocity and trajectory 244
4.1.2. Mass of projectile 244
4.1.3. Drag 245
4.1.4. Inlet area and nozzle throat area 246
4.1.5. Air inlet area of rearward step and port area 247
4.2. Procedure 247
4.3. Results 249
5. CONCLUDING REMARKS 251
ACKNOWLEDGEMENTS 253
REFERENCES 253

NOTATION

A—constant in the regression rate equation c* —theoretical characteristic velocity


5)
A —area of a fin Dg—total drag on the projectile
&
A —projectile inlet area Dg —body skin friction drag
* B
A —rearward step area Dg —fin skin friction drag
*/ &
A —port area Dg —interference drag
1 I
A —nozzle throat area Dg —wave drag
5 8
b, bo—pressure index in regression rate equation D —rearward step diameter
*/
C —wave drag coefficient D —fuel port diameter
D8 1
C —skin-friction coefficient corrected for com- f—fuel air ratio
&
pressibility F—thrust
C —skin-friction coefficient for turbulent incom- g—acceleration due to gravity
&5
pressible flow G —air mass flux
!
C —specific heat h—rearward step height
p
c* —experimental characteristic velocity h —convective heat transfer coefficient
%9 #

219
220 S. Krishnan and P. George

h —latent heat of decomposition or gasification of p—combustion chamber pressure


7
condensed fuel p —atmospheric pressure at any altitude
!
H—maximum altitude attained by projectile p —nozzle exit static pressure
%
ID —inner diameter of combustion chamber p —stagnation pressure of air at inlet
CH 0!
housing p —stagnation pressure at station x, Fig. 16
09
ID —initial inner diameter of fuel grain Pr—Prandtl number
F
ID —inner diameter of liner q —convective heat transfer rate
L #0/
Isp —air-specific impulse rR —regression rate
!
Isp —fuel-specific impulse R—range of projectile
F
k—thermal conductivity Re—Reynolds number
K , K —drag coefficient multiplication factors for Re —Reynolds number in the rearward step
B F */
protuberance and junctions Ru—universal gas constant
l/d—slenderness ratio of the ogival nose t—time
¸ —aft mixing chamber length t —total time of flight
!&5 505
¸ —average length of fin ¹ —temperature of inlet air in test facilities; also
& !
¸ —length of fuel grain stagnation temperature of ambient air corre-
F
¸ —length of projectile sponding to M
P
¸ —recirculation zone length ¹ —flame temperature
3 #
m, mo—temperature index in regression rate equa- ¹ —static temperature at station number three
3
tion ¹ —stagnation temperature at the exit of the aft
06
mR —mass flow rate of air mixing chamber
!
mR —mass flow rate of fuel ¹@ —adiabatic flame temperature
F 06
m —mass of projectile ¹ —fuel wall temperature
P 8
m —reference mass u—flow velocity
0
m —mass of combustion chamber º—projectile velocity
C
m —mass of combustion chamber housing º —exit velocity under adapted condition
CH %
m —total fuel mass º —launch velocity of reference projectile
F 0
m —mass of liner c—specific heat ratio of combustion products
L
M—flight Mach number c —specific heat ratio of air
!
M —Mach number at launch *t—time interval
1
M —Mach number at rearward step *¸ —combustion chamber wall thickness
2 CH
M —Mach number at port entry *¸ —combustion chamber liner thickness
3 L
Mo —molecular mass of combustion products g —combustion efficiency
1 "
n, no—mass flux index in regression rate equation g —boron combustion efficiency
""
n —number of fins k—viscosity
&
Nu—Nusselt number o—density
Nu —average Nusselt number o —density of air
!7% !
Nu —Nusselt number of fully developed flow o —density of combustion chamber housing
&$ CH
Nu —maximum Nusselt number material
.!9
OD —outer diameter of combustion chamber o —density of fuel
CH F
OD —outer diameter of fuel grain o —density of liner
F L
OD —outer diameter of liner h—projectile launch angle
L
OD —outer diameter of projectile
P

1. INTRODUCTION

A ramjet with subsonic combustion is the simplest and hence the most reliable and the
least expensive air-breathing propulsion system for supersonic flight. With the adoption of
supersonic combustion it is the only available air-breathing propulsion system for sustained
hypersonic flight. The ramjet then is known as supersonic combustion ramjet or scramjet.
However, the major disadvantage of a ramjet is that it cannot produce thrust under static
condition. Moreover, sufficiently high supersonic flight is necessary for an acceptable flight
economy. Therefore, the ramjet is required to be launched at a supersonic speed. Basically,
there are two types of ramjets, namely, liquid-fuel ramjet (LFRJ) and solid-fuel ramjet
(SFRJ). Hybridisation is always found in engineering systems and ramrocket or ducted
rocket is the result of such hybridisation between ramjet and rocket. Similarly there can be
turbo ramjet which is the combination of turbojet and ramjet. Now, of the two basic types
of ramjet, SFRJ (Fig. 1) represents a design of utmost simplicity due to the absence of any
moving part in its fundamental configuration. The configuration consists of an air intake
system, an exhaust nozzle, and a combustor. The combustor comprises of a flame holder, an
igniter, and the combustion chamber containing solid fuel.
The liquid fuel used in LFRJ and the solid fuel used in SFRJ are primarily of hydrocar-
bon type. Therefore, the fuel heating values per unit mass are essentially the same (approx-
imately around 44 MJ/kg). But the density of solid fuel used is generally 15—20% higher.
This can further be enhanced by the inclusion of metallic fuel powders such as aluminium or
Solid fuel ramjet combustor design 221

Fig. 1. Solid fuel ramjet.

boron. This addition will result in not only higher fuel density but also higher heating value
per unit mass. Therefore, the density fuel specific impulse (fuel density ] fuel specific
impulse) of SFRJ is expected to be about 15% higher than that of LFRJ. Approximately,
the density fuel specific impulse is inversely proportional to system size. Furthermore, the
SFRJ avoids fuel-control, fuel-storage, and the feed system comprising pumps and valves.
Therefore, SFRJ exhibits a better space utilisation resulting in a less complex and possibly
lighter system. Another benefit of SFRJ is the high degree of combustion stability experi-
enced with a wide variety of combustor types and sizes. The basic diffusion-controlled SFRJ
combustion processes result in a distributed energy release throughout the combustor. This
uniform energy release is believed to be the basic reason why no combustion instability
problems have been encountered in over 2500 combustion test firings with many different
combustion configurations at United Technologies Chemical Systems, U.S.A.(50)
The potential for a highly reliable and storable SFRJ propulsion system exists at a cost
only slightly higher than a solid rocket motor. In a way compensating even this higher cost,
the SFRJ can provide specific impulse in the 6000—7000 N s/kg range, resulting in typically
a 200—300% range-increase over a comparable size and weight of a solid rocket motor.(50)
When an SFRJ flies at a lower altitude as the air there is dense it ingests large mass flow
rate of air with high values of air mass flux, pressure, and temperature in the combustion
chamber. The requirement of correspondingly high fuel flow rate for this large mass flow
rate of air can be met since the regression rate of fuel is proportional to air mass flux,
pressure, and temperature. At higher altitudes as the air there is thin, the SFRJ ingests low
mass flow rate of air with reduced values of air mass flux, pressure, and temperature in the
combustion chamber. The requirement of reduced fuel flow rate at this condition can again
be met because of the above regression rate dependency. However, this ‘‘self-throttling’’
characteristic of SFRJ is not complete in itself. To achieve the desired trajectory or the
optimum fuel-air ratio over wide operating limits one has to additionally adopt suitable
methods of control for engine mass flow rate. Presently, two methods are possible, one the
bypass control of inlet air and the other the fuel regression rate control. The former is an old
concept and is found adopted also in other operating systems; for example, propulsion
systems of YF-12 and Concord use bypass control of inlet air.(41,65) But the fuel regression
rate control is new and specific to SFRJ.(28)
SFRJ has been a propulsion system of research-interest at least for the last 30 years.
Based on open literature, the countries which are taking keen interest in SFRJ application
in missile systems are China (Taiwan), Germany, Israel, Netherlands, Russia, Sweden, and
USA.(24,76,79,80,82,84) Considerable information on the related experimental research is
available from Germany, Israel, Netherlands and USA; the Sweden have reported to be
working on missiles using SFRJ. The profiles of the four types of SFRJ powered mis-
siles/projectiles reported from USA are shown in Figs 2 and 3.(50,76) The 229 mm (9 in)
air-to-air, air-to-surface, and surface-to-air missile shown in Fig. 2b has an SFRJ with solid
222 S. Krishnan and P. George

Fig. 2. SFRJ-powered missiles(50).

Fig. 3. SFRJ-assisted gun launched projectiles(76).

rocket motor booster. The 75 mm SFRJ propelled gun launched projectile (Fig. 3a),
designed by US Army Ballistic Research Laboratory, is reported to be of two versions,
namely, spin-stabilised (267.7 mm length) and fin-stabilised ('267.7 mm length).(59,60) The
concept adopts the very high pointing accuracy of a gun system. The projectile uses
a tubular unit into which is cast the solid fuel that generates sufficient thrust after gun
launch to sustain the projectile at its launch velocity. This results in a significant enhance-
ment in range. The projectile does not have an igniter and the fuel-autoignition capability
with air under the gun-launch condition was demonstrated as early as 1980. Mermagen and
Yalamanchili conducted in 1984 free-flight tests of the fin-stabilised version (Fig. 3a) with
hydroxyl-terminated polybutadiene (HTPB) solid fuel.(59) They measured the velocity and
Solid fuel ramjet combustor design 223

drag versus range for these projectiles with different internal configurations and solid fuel
compositions of HTPB. The SFRJs generated about 1100 N of thrust during the 1.6 s
burning time. Nordan systems of USA reported their studies on the SFRJ projectile known
as ‘‘cannon launched advanced indirect fire system (AIFS)’’ that was to be launched using
the M110A-2 cannon.(14,71,72) The projectile is of 203 mm diameter and 2548 mm (100 in)
length (Fig. 3b). It approximately weighs 114 kg and has a range greater than 60 km. By the
control of air mass flow rate through the use of a sensitive accelerometer, this projectile is
designed for pseudo-vacuum (thrust equals drag) trajectory.(71) A fire-and-forget version of
this projectile has a mix of submunitions as payload. Reference(80) presents the development
of SFRJ assisted gun launched projectile and air-to-air missile by Dutch (Figs 4 and 5).
Prins Maurits Laboratory and the Delft University of Technology in the Netherlands have
conducted studies on gun launched SFRJ assisted ‘‘tank-to-tank’’ projectile known as
‘‘kinetic energy penetrator’’ (M"4 and range"2500 m at sea level; 75 mm/90 mm dia-
meter).(80) An AGARD publication indicates the flight testing of an SFRJ projectile prior to
1992.(1) National Defence Research Establishment of Sweden has reported the development
of a spin-stabilised SFRJ assisted anti-aircraft projectile (M"4.3 and burn time"2 to 3 s;
40 mm diameter and 200 mm length).(24,82) A publication from the Naval Postgraduate
School of U.S.A. explains the possible use of a dual mode SFRJ for hypersonic tactical
missiles operating at design point of M"6 and altitude of 25 km (Fig. 6).(79)
On the point of view of ability of SFRJ to operate satisfactorily at all off-design
conditions the areas of concern to propulsion system designer are (1) selection of a fuel type,
(2) flame holding requirements that limit maximum fuel loading, (3) understanding the fuel
regression rate behaviour as a function of flight speed and altitude, (4) diffusion controlled
combustion process and its efficiency enhancement that requires special mixing section, and
(5) inlet/combustor matching. The objective of the present paper is to review the available
information on the combustor design with respect to the first four points and identify grey

Fig. 4. Geometry of the SFRJ-assisted antitank missile(80).

Fig. 5. Geometry of the high speed air-to-air missile(80).


224 S. Krishnan and P. George

Fig. 6. Schematic diagram of a solid fuel, dual-mode ramjet(79).

areas for further investigation. At the end, with a view to demonstrate the use of the
information collected through this review, a preliminary design procedure is presented for
an SFRJ-assisted gun launched projectile of 155 mm diameter.

2. EXPERIMENTAL RESULTS

2.1. EXPERIMENTAL SET-UPS

A schematic diagram of an SFRJ combustion and nozzle flow region is shown in


Fig. 7.(16,67) The combustion chamber is basically a hollow cylinder in which a cylindrical
fuel grain, usually with a circular perforation, is placed. Incoming-air flows through the fuel
port. An often used combustor geometry consists of a number of different regions and
features: (1) the head end with the air inlet and rearward step, (2) the main combustor
section where the solid fuel grain is placed, (3) the aft mixing chamber often with a mixer
plate at its front, and (4) the exit nozzle. The combustion in the solid fuel grain is mostly
through boundary layer diffusion flame and hence slow and not very efficient. For enhance-
ment in the overall combustion efficiency the aft mixing chamber is necessary. In this the
reaction between fuel and air is completed due to better mixing. Sometimes the aft mixing
chamber is fitted with a bypass air injection. In the case of certain metallized fuels being
used, introducing swirl to inlet air flow and/or injecting bypassed air into the aft mixing
chamber are found necessary to have acceptable combustion efficiency.
The details of experimental set-ups and test conditions adopted by Israel Institute of
Technology (Haifa, Israel), Institute for Chemical Propulsion and Engineering (DLR,
Hardthausen-Lampoldshausen, Germany), Naval Postgraduate School (California, U.S.A.),
and Delft University of Technology and Prins Maurits Laboratory TNO (Rijswijk, The
Netherlands) are given in Table 1. A typical experimental set-up adopted for SFRJ
combustor development is shown in Fig. 8.(85)
In Israel, (3,17,22,52~54,58,85,86) the experimental set-up used is small and it is for an SFRJ
of initial fuel port diameter 7.5—25 mm. To simulate the flight conditions compressed air was
heated electrically up to 800 K. A 30 kW electric heater and water-cooled three-way valve
provided the necessary controllable hot, high-pressure pure air flow rates up to 40 g/s.
A nitrogen purge system was used to extinguish the flame and cool the combustor when
required. Figure 9 gives the details of the combustor test section.(58) The assembly of air
inlet injector permitted testing with different air inlets and rearward steps. It consisted of
a 20 mm diameter plenum chamber, and a converging section (half-angle 30°) followed by
a 20—26 mm long cylindrical air inlet. The combustor attached to the air injector included
the fuel grain and an aft-mixing chamber with a mixer plate at its front. The ignition was
achieved by a short time injection of hot solid propellant combustion gases into the
chamber through an annular slot at the wake of rearward step. Continuous measurements
Solid fuel ramjet combustor design 225

Fig. 7. SFRJ combustor flow field(16,67).

of inlet air pressure and temperature, air flow rate, chamber pressure, and thrust were made
and recorded on an optoelectrical recorder. The exhaust flow composition was analysed by
an Orsat device. The tests were automatically operated and controlled from the control
room. The transparency of the PMMA fuel grain enabled continuous observation and
determination of the spatial and temporal variations of the fuel regressing surface. A video
camera with a zoom lens and a graded calibrated screen was used for this. The video
measurements enabled the determination of the instantaneous internal contour with an
uncertainty of less than $0.05 mm. Distortion of the internal contour image was elimi-
nated by using fuel grains with rectangular external cross section.
In Germany,(9,38,45,61,67~69,81) the experimental set-up used is considerably larger and it
is for an SFRJ of initial fuel port diameter, 60—120 mm. The air was heated by hydrogen-
oxygen combustion with oxygen make up. The supply systems of high-pressure air (200 bar),
hydrogen, and oxygen provided the controllable air flow rates up to 5.5 kg/s and a max-
imum temperature of 900 K. The hydrogen/oxygen air heater chamber consisted of ten
burners distributed on its front end cover plate. By choosing the number of burners, a wide
range of temperature could be realised. The maximum amount of water vapour in the air on
account of combustion of hydrogen and oxygen was about 6% of the total mass flow.
Through oxygen makeup the amount of oxygen in the vitiated air was kept constant at 23%
by mass for all inlet air temperatures.
In the experimental ramjet motor, the length of fuel grain could be varied up to 1 m. With
different rearward step inserts and nozzles, the ratios of grain port to air inlet area between
1 and 5 and the ratios of grain port to nozzle throat area between 1 and 4 could be
investigated. Either a pyrotechnic (mixture of aluminium and Teflon powder) or an H /O
2 2
spark torch igniter could be used for ignition. At the air inlet several fixed vane swirlers were
installed to induce swirling flow into the ramjet combustor, so that the effect of swirl on
solid fuel combustion could be studied. The whole assembly of the combustor and the air
heater was mounted on a thrust stand and connected through flexible joints to the supply
systems. The experimental set-up had a provision to measure temperature and species
concentration profiles throughout the combustor. The temperature measurements were
made using platinum/rhodium 70/30 platinum/rhodium 94/6 (EI 18) thermocouples of
0.2 mm diameter.(69) With the help of a motor drive, the thermocouple could be moved
radially at the chosen location of the combustor. The positioning accuracy of the ther-
mocouple bead was $0.1 mm. The concentration measurements were undertaken by
a gas-chromatographic technique. The gas samples were taken from the combustor by
a water-cooled probe of an internal diameter of 0.6 mm and an external diameter of 3 mm.
The probe was inserted into a hole drilled through the fuel block. As the concentration of
species in a ramjet combustion chamber is dependent on time and the location with respect
to the regressing fuel surface, the gas sampling was done within 2—3 s. The gases were
226
Table 1. Experimental set-ups and test conditions of SFRJ combustor adopted by different investigators

Test conditions Israel Germany U.S.A. The Netherlands

Inlet air temperature (K) 290—800 electrically heated air 288—900 oxygen/hydrogen burners 288—600 methane/air burners 288—1000 methane/oxygen burning
and condition heat up heat up; oxygen make up with oxygen makeup
Initial port diameter 7.5—25 (circular port) 60—120 (circular port) 37—45 (circular port) and 40—60 (circular port)
(mm) 63.4 * 25.4 (rectangular port)
Chamber pressure (bar) 3—10 4—12 4—10 3—6
Air flow rate (kg/s) Up to 0.04 Up to 5.5 Up to 0.15 Up to 4.0
Inlet air flow condition Straight flow with single stream Straight flow and swirl flow/ Straight flow and swirl flow, Straight flow
single stream single stream and bypas
Testing duration (s) Up to 8 Up to 40 — Up to 35
Rearward step height, h 0—0.35 0.10—0.30 0.25 0.28—0.40
by initial port
diameter, D (h/D )
S. Krishnan and P. George

1 1
Fuels used PB#PS (3 : 1), PE, PMMA, PP HTPB, HTPB#B, HTPB#B C, HTPB#B, HTPB#B C, HTPB, PE
4 4
HTPB#Mg, HTPB#PE, HTPB#Mg, PMMA
HTPB#PMMA, PE
Ignition Pyrotechnic (solid propellant) (1) Pyrotechnic (mixture of Al and Ethylene—oxygen ignition torch H /O spark igniter
2 2
Teflon powders)
(2) H /O spark torch igniter
2 2
HTPB: Hydroxyl-terminated polybutadiene; PB: Polybutadiene; PE: Polyethylene; PMMA: Polymethylmethacrylate; PP: Polypropylene; PS: Polystyrene.
Solid fuel ramjet combustor design 227

Fig. 8. A typical SFRJ test set-up(85).

Fig. 9. SFRJ combustor test section(58).

analysed for N , O , CO, CO and H and were separated by a Porapak T and a molecular
2 2 2 2
sieve 5 As column.
In USA,(2,7,23,26,37,40,51,55,70,83) the experimental set-up used is smaller than that of
Germany. Here the air was heated by methane/air or hydrogen/air combustion with oxygen
make up. In order to study the effect of adding bypass air in the aft mixing chamber special
provisions existed in the set-up. Air flowed from high pressure (200 bar) tanks through
a pressure regulator and choked nozzles to two air heaters for the main and bypass air
supplies. Hydrogen and methane were used as fuels for the main and the bypass air heaters,
respectively, and oxygen was injected downstream of the heaters to ensure that the vitiated
air contained 23% oxygen by mass as that of ambient air. Both air heaters were acoustically
isolated from the ramjet combustor with sonically choked nozzles. The set-up had facility to
conduct experiments with two-dimensional as well as axisymmetric combustors. The
two-dimensional combustor incorporated a variable rearward step height, driven by a vari-
able-speed, reversible, high-torque AC motor which could position the step during SFRJ
228 S. Krishnan and P. George

operation. The combustor was 63.5 mm wide with a 25.4 mm height between the fuel slabs,
which were bonded to the top and bottom surfaces of the combustor. The fuel slabs were
406 mm long and 6.4 mm thick. Two air-purged viewing windows (12.7 mm diameter) were
used, one within the recirculation region (38 mm aft of the inlet) and the other within the
boundary layer combustion region near the aft end of the fuel slab (343 mm aft of the inlet).
Two rotating prism high-speed cameras were used to observe the combustion process. Most
of the recordings were taken at 6000 frames/s. The axisymmetric combustor configuration
was of about 45 mm initial port diameter. This facility was adopted mainly to study the
combustion characteristics of highly metallised (up to 50%) solid fuel grains with
boron/boron carbide and magnesium. This facility had a provision to measure the particle
size distribution in the combustor. The particle size measurements were taken with a Mal-
vern Particle Sizer. A 2 mW He—Ne laser produced a collimated, 9 mm diameter, monochro-
matic beam of light that illuminated the particles. The incident light was diffracted by
the particles to give a stationary diffraction pattern independent of particle position
and velocity. The measurement was based on far-field, near-forward Fraunhofer light
diffraction.
In the Netherlands,(8,11~13,19,31,32) the experimental set-up used is smaller than that of
Germany and it is for an SFRJ of initial fuel port diameter, 40 mm. Here the air was heated
by methane/oxygen combustion with oxygen make up. This facility had a computer-
operated gas supply system which separately supplied air, oxygen, hydrogen, nitrogen, and
methane. Mass flows were controlled by choked nozzles with an accuracy of about 1.5%. In
the air heater, methane was burned with oxygen to raise the inlet air temperature up to
1000 K, and the oxygen-content in the vitiated air was kept constant at 23% by mass. An
H /O spark torch igniter was used at the wake of the rearward step for the ignition of solid
2 2
fuel. The assembly of air heater and combustor was mounted on a thrust stand. Fuel grains
could be varied up to 190 mm outer diameter and 1000 mm length. To determine the
instantaneous local regression rate, an ultrasonic pulse echo technique was employed.
A sound pulse emitted by an ultrasonoscope was reflected at the solid—gas interface of the
fuel. The ultrasonic analyser determined the time elapsed between the emitted and the
received reflected pulse. By knowing the velocity of sound, the momentary wall thickness
and, hence, the regression rate could be determined. Reference 8 gives further details of the
adopted ultrasonic pulse echo technique.

2.2. FUEL TYPES

There is a wide variety of solid fuel binder materials that have been investigated to date.
Depending on the mission requirements, the best gravimetric or volumetric heat-releasing
fuel binder can be selected. SFRJ test combustion chambers employ hydrocarbon (HC)
fuels, usually polymers such as PB, PE, PMMA, PP, and PS. One of the suitable solid fuels
developed to date consists of a blend of PB and PS, providing high gravimetric heat release,
good mechanical properties, good regression characteristics, and high combustion efficien-
cy over a wide range of conditions.(50)
It is possible to increase both the gravimetric heat release and the volumetric heat release
of solid fuels by use of metal additives. The boron and boron carbide (B C) additives offer
4
the highest potential increase in both gravimetric and volumetric heat release. Boron
carbide is capable of providing almost same energy of boron; in addition it costs much less.
However, extracting the energetic potential from boron or boron carbide has been a difficult
task due to very complicated ignition and combustion processes. The boron or boron
carbide particles in the fuel matrix are covered with a thin boron oxide (B O ) layer that
2 3
serves as a barrier-crest for the oxygen to react with the material beneath. Ignition of the
particle is possible only when the oxide layer is removed. This can be done either by the
evaporation of the oxide layer or by the reaction of the oxide with water vapour that is
found in the HC—O reaction products.(18,29,30) The rate of oxide evaporation depends
2
mainly on the particle temperature. Thus the particle heating process, by way of convection,
radiation, or reaction with oxygen, seems to be the dominant factor that controls particle
Solid fuel ramjet combustor design 229

ignition. In order to achieve ignition the particles need an environmental temperature of


approximately 1900 K and an adequate residence time in this environment. Aluminium and
magnalium (alloy of magnesium and aluminium) are the other attractive metal additives for
solid fuels; between the two, with less difficulty the fuel with magnalium can be ignited and
made to perform with higher combustion efficiency.

2.3. FLAMMABILITY LIMITS

The nature of the flow field in the SFRJ combustor is also shown in Fig. 7. The SFRJ uses
a rearward step to produce the recirculation zone that ends at the reattachment point from
where the stabilised boundary layer diffusion flame propagates downstream. Although no
true reattachment point can be defined because of mass blowing from solid fuel walls, it is
possible to define a region where effective reattachment does occur. At the beginning,
a minimum (or critical) ratio of rearward step height to port diameter, (h/D ) * is required
1
for ignition and sustained combustion in an SFRJ and this ultimately limits the maximum
pressure recovery and the maximum fuel grain loading, thus the range of missile. This limit
can be determined from ‘‘go/no-go’’ type tests. Tests that exhibit ignition and sustained
combustion are considered within the flammability, whereas those resulting in no ignition
are considered out of flammability limits.
Israel performed several hundreds of such tests with three different fuel types, PMMA,
PE and PB/PS, for various A /A and h/D ratios.(58) The initial port diameter of all the
1 5 1
grains was 10 mm. Air inlet temperatures of 800, 520, and 290 K (room temperature) were
adopted to simulate flight Mach numbers of 3, 2, and a low subsonic value respectively. At
the highest temperature of 800 K, all the three fuels exhibited very good flammability—low
(h/D )*. The (h/D )* increased at 520 K and became very high at the room temperature. The
1 1
studies with various A /A ratios demonstrated that the poorest flammability is obtained for
1 5
the lowest A /A (A /A *1). With increase in A /A the flammability improves—(h/D )*
1 5 1 5 1 5 1
decreases. Above certain A /A , the (h/D )* is essentially constant as schematically shown in
1 5 1
Fig. 10a. The PB/PS fuel type exhibited better flammability limits. This latter observation
was argued to be due the highest regression rate of PB/PS fuel that could give at a lower
(h/D )* the sufficient fuel/air mixtures in the recirculation region to sustain combustion in
1
the rest of the combustor. DLR group of Germany also conducted flammability experi-
ments at three different air temperatures, 673, 473, and 288 K.(67) PE and HTPB fuel grains

Fig. 10. Flammability limits.


230 S. Krishnan and P. George

of two different port diameters, 60 and 120 mm were used in their study. They found similar
flammability results on the effects of air temperature and A /A . In addition, they showed
1 5
that a larger grain port diameter had a better flammability; the grains with 120 mm port
diameter required lower (h/D )* than the ones with 60 mm port diameter.
1
A windowed two-dimensional combustor incorporated with a variable rearward step
height was used by U.S.A. to study separately the ignition and sustained combustion
characteristics of several types of fuel compositions.(83) This was in significant departure
from the earlier go/no-go type experiments. The test conditions were G " 13.7 to
!
14.3 g/cm2 s, ¹ "510 to 533 K and p"8.7—9.5 bar. Wooldridge and Netzer(83) showed
!
that (h/D )* for ignition was higher than that required for sustained combustion. This
1
demonstrates that the region of ignitability is narrower than that of sustained combus-
tion—a typical characteristic of any open combustion chamber.
All the studies show that a lower (h/D )* can be provided if, (1) A /A is higher (the
1 1 5
combustion chamber velocity is lower), (2) the inlet air temperature is higher, (3) the port
diameter is larger, and (4) the fuel regression rate is faster. These trends are qualitatively
presented in Fig. 10.
Netzer and Gany(58) compared their results with the studies of United Technologies
Chemical Systems, U.S.A.(76) and of Schulte(67) to give flammability limits for SFRJs using
different fuels, port diameters, and air inlet stagnation temperatures. These can be sum-
marised as given in Tables 2 and 3.
Another point of interest for the recirculation flow field is the variation of recirculation
zone length (¸ ) which changes with the rearward step height. With increasing rearward step
3
height the recirculation zone length increases. The ratio of recirculation zone length to port
diameter, ¸ /D increases with h/D ; and this ratio does not have pronounced dependence
3 1 1
on port flow Reynolds number and port diameter. For these, depending upon test-
conditions, empirical fits as given below are obtained: (58,67,13)
¸ /D "!0.64#9.24h/D
3 1 1
¸ /D "!0.73#11.93h/D
3 1 1
¸ /D "!1.53#11.77h/D
3 1 1

2.4. REGRESSION RATE

In SFRJ, the solid fuel regression pattern along the grain passage is complex. The
regression is at a minimum just downstream of the rearward step. It increases with axial
distance and attains a maximum at the flow reattachment region, which itself is shifting

Table 2. Flammability limits of polyethylene SFRJ

¹ "290 K ¹ "470 K ¹ "670 K


! ! !
D (mm) (h/D )* (A /A ) (h/D )* (A /A ) (h/D )* (A /A )
1 1 1 5 .*/ 1 1 5 .*/ 1 1 5 .*/
10 0.29 3.0 0.23 1.8 0.09 1.8
60 0.22 2.8 0.21 1.7 — —
120 0.18 2.2 0.17 1.5 0.09 1.2

Table 3. Flammability limits of polybutadiene SFRJ

¹ "290 K ¹ "520 K ¹ "800 K


! ! !
D (mm) (h/D )* (A /A ) (h/D )* (A /A ) (h/D )* (A /A )
1 1 1 5 .*/ 1 1 5 .*/ 1 1 5 .*/
10 0.28 2.8 0.20 1.2 0.0 1.3
'50 0.26 1.9 0.14 1.4 — —
Solid fuel ramjet combustor design 231

Fig. 11. Solid fuel regression patterns vs axial distance for different burning times.

downstream from time to time as the rearward step height is increasing. Further
downstream of the reattachment region the regression rate decreases gradually from its
maximum value. This typical behaviour is shown in Fig. 11, that has been re-plotted from
Ref. (85). ¸ocal and instantaneous regression rates can be measured by (1) the high-speed
photography of the combustion front through transparent windows; (2) the ultrasonic pulse
echo technique;(8,33) or (3) a suitable flame quenching procedure. In many studies, as an
alternative to this regression rate measurement, average regression rate is calculated by the
time averaging procedure—weighing the fuel grain before and after each run and calculat-
ing the averaged web burn rate. In this procedure the calculated regression rate is evidently
a space- and time-interval-dependent. The effects of various fuel types on regression rate are
summarised in Table 4. Various investigators have obtained the empirical relations for
average fuel regression rate and these are given in Table 5. Examining these relations we
find significant differences in the values of indices and coefficients, and Table 6 summarises
the different indices obtained by various studies. Since the fuel regression rate is an
important parameter in the design of SFRJ, an attempt is made to explain the reasons for
such differences in the following paragraphs.
In relation to hybrid rocket combustion, a large number of studies on combustion
chamber configurations similar to that of SFRJ were conducted in the 60s and 70s and also
in recent years.(5,6,42—44,63,73—75,78) Although the combustion characteristics in hybrid
rockets are similar to those of SFRJ, there are two major differences. First, the hybrid rocket
employs an oxidiser other than air; and second, it does not necessarily feature sudden
expansion of gaseous inlet flow. Due to this sudden expansion three distinct flow fields exist
in SFRJ: (1) recirculation zone, (2) reattachment region, and (3) zone downstream of
reattachment. Following the intense mixing of fuel vapour, air, and hot combustion
products, and the pre-flame reactions in the recirculation zone, the maximum temperature
occurs closest to the surface at the reattachment region. The flame stabilised in this region
propagates throughout the combustor as the mixing limited diffusion flame, established
within the boundary layer. The local fuel-regression rate is differently influenced by these
three fields. However, depending upon the combustion chamber operating conditions the
spatially averaged regression rate can be taken to be largely influenced either by the
convective and radiative heat transfer from the high temperature gas-phase reacting-flow-
field in the reattachment region or by that in the zone downstream of it. In such a situation,
for the present, it can be assumed that the convective heat transfer from the high temper-
ature flow-field is the only energy source at the fuel wall for its endothermic gasification
processes. The use of this assumption yields

q
rR " #0/ (1)
o h
F 7
232

Table 4. Experimental results obtained using various fuels on regression rates and flammability limits

Results reported from


Results reported from Results reported from U.S.A.(2,7,23,26,37,40,51,55,70,83) and the
Fuel types tested Israel(3,17,22,52~54,58,85,86) Germany(9,38,45,61,67~69,81) Netherlands(8,11—13,19,31,32) Remarks

HTPB Among PB#PS, PE, PMMA, and PP Among HTPB, HTPB#PE, U.S.A: PMMA’s heat of combustion is low
HTPB#B (1) PE and PP have the lowest HTPB#PMMA, and PE HTPB#B, HTPB#B C, whereas HTPB, PE, and PS have
4
HTPB#B C regression rate while PB#PS has (1) PE has the lowest regression rate HTPB#Mg, and PMMA high and essentially equal heats of
4
HTPB#Mg the highest regression rate, and while HTPB has the highest are the fuels reported to combustion (*Q '*Q '*Q )
PE HTPB PS
HTPB#PE (2) PMMA has the narrowest regression rate, have been studied.
HTPB#PMMA flammability limits whereas PB#PS (2) Cast fuel regresses slower than
The Netherlands:
HTPB#PS has the widest limits the compressed ones,
PE (3) Addition of Mg in HTPB HTPB, PE, PMMA, PS are the fuels
reported to have been studied.
PMMA increases regression rate, and
S. Krishnan and P. George

PP (4) HTPB#B ()40%) and


PS HTPB#B C ()40%) can be
4
combusted efficiently with swirl flow.
Solid fuel ramjet combustor design 233

Table 5. Empirical relations for average regression rate of solid fuel

Empirical power law Initial port diameter


Investigator (mm/s) Parameter D (mm) Fuels
1

G
Israel rR JG0.9~1.0 10 PMMA
! —
[Ref. (58)] *¹0.95 PE, & PB
!
*(1!2h/D )~1
1
rR JD~0.4 5—60
1
Germany rR "0.008p0.28¹0.5 G "25 g/cm2 s 60  PE
! !
[Ref. (67)] rR "0.013p0.26¹0.42 G "36 g/cm2 s 120 
! ! 
rR "0.066G0.26p0.39 ¹ "673 K 120 
! !
rR JD~0.23 G "25—36 g/cm2 s 60&120 
1 !

G G
USA rR "0.047G0.38p0.29 G "2.82 g/cm2s 38 PMMA
! !
[Ref. (40)] rR JG0.41p0.51¹0.34 — — PMMA
! !
[Ref. (76)] rR JG0.6D~0.4¹0.3 — 100—260 PB
! 1 !
Netherlands rR "0.0072mR 0.61 ¹ "300 K
! !
[Ref. (13)] rR "0.0117mR 0.58 ¹ "500 K  
! !  
rR "0.0083mR 0.56 h/D "0.30  
! 1
rR "0.0072mR 0.61 h/D "0.33  
! 1  40  PE
rR "0.0071¹0.55
!
h/D "0.33
1  
rR "0.0051¹0.58 h/D "0.30  
! 1  
rR "0.0072¹0.60 mR "250 g/s  
! !
D "(mm); G "(g/cm2s); mR "(g/s); p"(bar); rR "(mm/s); ¹ "(K).
1 ! ! !

Table 6. Experimental correlations for average regression rate

Indices for G D 1!(2h/D ) ¹ p Fuels


! 1 1 !

G
Israel 0.9—1.0 !0.4 PMMA
!1 0.95 —
(PMMA only) PE, & PB
Germany 0.26 !0.23 — 0.42—0.50 0.26—0.39 PE
U.S.A.—I 0.38—0.41 — — 0.34 0.29—0.51 PMMA
U.S.A.—II 0.6 !0.4 — 0.30 — PB
Netherlands 0.56—0.61 — — 0.55—0.60 0.0 PE

where h is the latent heat of decomposition or gasification of condensed fuel. The


7
convective heat transfer can be expressed by using the heat transfer coefficient h ,
#
q "h (¹ !¹ ) (2)
#0/ # # 8
Assumption of constant gas temperature ¹ and constant wall temperature ¹ yields
# 8
q Jh (3)
#0/ #
hence
rR J h (4)
#
As indicated above, depending upon the combustion chamber operating conditions the
spatially averaged regression rate can be taken to be largely influenced either by the heat
transfer mechanism in the reattachment region or by that in the zone downstream of it.
Let us first consider the heat transfer mechanism in the zone downstream of reattach-
ment. The fuel regression rate in this zone can be explained by a simple analysis of heat
transfer in a turbulent pipe flow. The flow field is characterised by a fully developed
turbulent boundary layer and a constant temperature gas-phase diffusion flame within.
Applying the Reynolds analogy between heat and momentum transfers in boundary
234 S. Krishnan and P. George

layers, we get
NuJRen Pr0.33 (5)
or
h D /kJ(ouD /k)n(kC /k)0.33 (6)
# 1 1 1
Hence,
h JGnDn~1k0.33~nC0.33k0.67 (7)
# ! 1 1
SFRJ combustion chambers are designed to operate around the stoichiometric fuel—air
ratios and hence for typical solid fuel hydrocarbons these ratios are around 0.1. Further-
more, the flow fluid through the fuel grain passage changes from pure air at entry to
incompletely burned combustion products at exit. However, for the simplicity of analysis it
is assumed that the mass flux through the grain passage is only of air. Using Sutherland’s
equations for k and k for air,(4)
k"1.458]10~6¹1.5/(¹ #110.4) (kg/m s) (8)
! !
k"1.993]10~3¹1.5/(¹ #112) (W/(cm K)) (9)
! !
and the linear variation of C with temperature as
1
C "a¹ #b (10)
1 !
where a and b are constants, we get
h JGnDn~1[¹1.5/(¹ #110.4)]0.33~n(a¹ #b) 0.33[¹1.5/(¹ #112)]0.67 (11)
# ! 1 ! ! ! ! !
The above equation can be put into a simpler form as
h JGnDn~1¹m (12)
# ! 1 !
or as per Equation (4)
rR JGnDn~1¹m (13)
! 1 !
This result indicates that in a non-radiative system with zero blowing the convective heat
transfer and hence the regression rate downstream of reattachment region should depend
on the air mass flux, the port diameter, and the inlet air temperature. As has been indicated,
this sample analysis is applicable to the condition in which the regression rate is controlled
only by convective heat transfer in a turbulent pipe flow. However, at low pressures
kinetically controlled processes also become important—particularly near the flame leading
edge. For example, in PMMA-oxygen hybrid system the pressure sensitivity was found
below 17 bar.(43) Such low pressures—even to sub-atmospheric values—exist in SFRJ
combustion systems. Therefore, including the pressure effect in the above regression rate
equation we get
rR JGnDn~1¹mpb (14)
! 1 !
Under this simple analysis of heat transfer in a pipe flow, for a fully developed turbulent
boundary layer n"0.8, and for the typical SFRJ operating temperature range of
300—800 K, m"0.14. Therefore
rR JAG0.8D~0.2¹0.14pb (15)
! 1 !
This regression rate equation corresponds to the flow field downstream of reattachment and
is applicable to the SFRJ grains of large port-length to diameter ratios and high Reynolds
numbers.
Let us now turn to the heat transfer mechanism in the reattachment region. In this region
the heat transfer and hence Nusselt number reach their maxima. In a fixed geometry similar
to that of an SFRJ, Krall and Sparrow(34) experimentally obtained with a nonreacting flow
the heat transfer pattern in the reattachment region. They compared this pattern with that
in the fully developed boundary layer far downstream of reattachment. In a subsequent
study by Zvuloni et al.(86) SFRJ fuel regression pattern demonstrated a similar variation
with the heat transfer pattern obtained by krall and Sparrow. The study of Krall and
Sparrow brings out three important results.
Solid fuel ramjet combustor design 235

First, the value of Nu at the reattachment region, Nu can be directly correlated to the
.!9
inlet Reynolds number Re based on the air inlet diameter (D "D !2h), according to
*/ */ 1
Nu JRe/0 (16)
.!9 */
Equation (16) reveals a heat transfer regine that is fundamentally different from the one that
exists in a fully developed boundary layer. While the Nu in a fully developed boundary layer
is related to the grain-passage Reynolds number [based on D , Equation (5)] regardless of
1
the value of D , the Nu in the reattachment region is solely dependent on the Re .
*/ */
The second result is that the ratio between (1) the maximum heat transfer obtained at the
flow reattachment (represented by Nu ) and (2) the value of heat transfer from the fully
.!9
developed turbulent boundary layer far downstream (represented by Nu ) increases with
&$
decreasing grain-passage Reynolds number. This effect can be correlated using
Nu /Nu JRe~0.2 (17)
.!9 &$
This equation reveals that for the given air mass flux range G , since Re"G D /k, the
! ! 1
smaller the port diameter the greater the effect of reattachment region. In the case of SFRJ
having a low grain-passage Reynolds number the maximum heat transfer at the reattach-
ment region dominates the overall heat transfer behaviour in the entire port; this domi-
nance, for a fixed Re gets further accentuated for small ¸ /D ratio.
F 1
The third and final important result of the study is that the larger the h/D ratio the larger
1
the ratio of Nu /Nu . For a fixed Reynolds number, it can be shown that
.!9 &$
Nu /Nu J(1!2h/D )~2@3 (18)
.!9 &$ 1
Therefore, combining their second and third results, in situations such as low Re, low ¸ /D ,
F 1
or high h/D ,
1
Nu JNu (19)
!7% .!9
For the time-averaged values of D ,G , Re, etc.,
1 !
Re "ReD /D "G D2/(kD ) (20)
*/ 1 */ ! 1 */
Using Equations (16) and (19) which characterise the heat transfer for reattachment region,
and Equation (20) we get
Nu JG/0D2/0/(kD )/0 (21)
!7% ! 1 */
As rR Jh , the above equation becomes
#
rR JG/0D2/0~1D~/0k~/0k (22)
! 1 */
Considering as before the effects of temperature and low pressure the above equation in
a general form is
rR JG/0D2/0~1D~/0¹.0p"0 (23)
! 1 */ !
Typically, no"2/3; for the SFRJ operating-temperature range of 300—800 K, mo"0.23.
Therefore,
rR " AG0.67D0.33D~0.67¹0.23p"0 (24)
! 1 */ !
Thus we see that if the spatially averaged regression rate is influenced by the heat transfer
mechanism downstream of reattachment region it can be estimated through the equation of
the form, Equation (14); and on the other hand if it is influenced by the mechanism at the
reattachment region the equation of the form Equation (23) applies.
At this juncture it is important to recall that the forms of the Equations (14) and (23) are
derived for a non-radiative convective system with zero blowing. However, for a non-
radiative system with the blowing due to fuel regression there is a partial blockage to
convective heat transfer. A suitable correction factor for blowing can be incorporated
noting that it is a thermodynamic one determined by the air—fuel combination selected.(44)
Furthermore, in case the radiative heat transfer is significant the consequent increase in fuel
regression due to radiative component (hence increased blowing) tends to decrease the
236 S. Krishnan and P. George

convective heat transfer to the wall. On accounting for this coupling, for small ratios of
radiative heat transfer to convective heat transfer, say (0.1, it can be shown that in a given
system operating under fixed flow conditions the total heat transfer (convective and
radiative) with related-blowing is not much different from the estimated convective heat
transfer with blowing.(44)
Experimental investigations on SFRJ combustion often seek a regression rate correlation
as
rR "AGn (25)
!
and the value of n is determined through time averaging procedure. There is an interesting
discussion by Zvuloni et al.(86) indicating that in such time-averaging procedure the ‘‘appar-
ent’’ indices (typically given in Tables 5 and 6) can be significantly different from the indices
values corresponding to the fundamental heat transfer mechanism, Equations (15) and (24).
For example, the apparent n value will be less than 0.8 if the regression rate is influenced by
the heat transfer mechanism downstream of the reattachment region; but it will be greater
than 2/3 if the rate is influenced by the mechanism in the reattachment region. This
behaviour has been confirmed by Korting et al.(33) through their regression rate measure-
ments by ultrasonic pulse echo technique as well as time-averaging procedure. Further-
more, for two motors of identical initial-operating conditions the apparent indices will be
different if the time-averaging intervals are different. Finally it should also be noted that
the various investigators adopted different methods for heating the air in their studies
(Table 1).
We may summarise the above discussion on regression rate as follows. The missile
designer on the point of view of fuel regression rate law is confronted with many forms of
empirical correlations. The regression rate is usually affected by one of the two dominant
mechanisms of heat transfer to the fuel surface: (1) heat transfer mechanism downstream of
the reattachment region (typical of long fuel grains and large grain port diameter of high
Reynolds number) and (2) heat transfer mechanism in the region itself (typical of short fuel
grain and small grain port diameter of low Reynolds number). In the time-averaging
procedure, in either of these two mechanisms, the form of fuel regression rate law obtained is
further affected by the duration of the experiment. Under this procedure a sufficiently small
duration of the experiment is expected to give the forms of regression rate laws that may
closely correspond to the fundamental heat transfer mechanism, of course, accepting the
other realities. An important ‘‘other-reality’’ is that it is impossible to obtain a heat transfer
to fuel surface that is purely downstream of reattachment region or at the reattachment
region itself. Therefore, the values of indices found in various reported correlations only
demonstrate the general influence of various parameters and the application of these values
in actual design should be done with caution. The regression rate law obtained through the
procedure of local and instantaneous regression rate measurements is expected to be more
accurate, but the sophistication involved in the measurements is evidently much higher.

2.5. COMBUSTION EFFICIENCY

As indicated earlier, an aft mixing chamber is found necessary to improve combustion


efficiency. The parameters controlling combustion efficiency are: (1) aft mixing chamber
length to initial port-diameter ratio, ¸ /ID , (2) equivalence ratio, and (3) h/D . The
!&5 F 1
increase in ¸ /ID provides longer residence time for the completion of combustion
!&5 F
through mixing and burning—calculations based on turbulent boundary layer theory show
that only about 50% of the fuel gets mixed and burned at the grain-passage exit plane.(50)
Equivalence ratio is important because it determines at the exit the relative quantity of fuel
which must be subsequently mixed and burned. The h/D controls the turbulence level
1
introduced in the recirculation zone, thereby influencing mixing throughout the combustor.
The aft mixing chamber is usually fitted with a mixer plate at its front for additional
mixing and consequent enhancement in combustion efficiency (Figs 1 and 7). The combus-
tion efficiency is defined differently by various investigators. Investigators from Israel(86)
Solid fuel ramjet combustor design 237

defined it as
(¹ !¹ )
(g ) " 06 ! (26)
"1 (¹@ !¹ )
06 !
where ¹@ was the adiabatic flame temperature obtained from thermo-chemical computer
06
code such as CEC71.(20) ¹ was the experimental combustion chamber temperature
06
determined from the experimental c* value given by
p A
c* " 06 t (27)
%9 (mR #mR )
! F
The c* was calculated from the measured chamber stagnation pressure, p , nozzle throat
%9 06
area, and the overall exhaust flow rate (mR #mR ). While the air flow rate mR was a measured
! F !
quantity, mR was estimated from the average fuel consumption rate. By using a mixer plate
F
and an aft mixing chamber diameter"1.4 * ID and length"4 * ID , the typical combus-
F F
tion efficiency attained by Gany et al.(58,85,86) for unmetallised solid fuels (PMMA, PE,
PB/PS) of small initial port diameters of 10 mm was above 90%. Investigators from
Netherlands(32) defined the combustion efficiency as the ratio between experimental and
theoretical characteristic velocities,

S
c* p A ¹
(g ) " %9" 06 5 " 06 (28)
" 2 c* (mR #mR ) c* ¹@
5) ! F 5) 06
Here, the mR was determined with the ultrasonic technique and c* was calculated using
F 5)
CEC71. The value of (g ) attained by them(32) for unmetallised PMMA fuel grain was
"2
around 75% for the same ¸ /ID and h/D used by Israel. This value of 75% works out to
!&5 F 1
be around 50% for (g ) . The main difference between these two combustor configurations
"1
was the absence of mixer plate in the combustor used by the latter investigators. This
observation points out the importance of introducing a mixer plate in enhancing the
combustion efficiency. This is in agreement with the qualitative correlation found in
Ref. (76). As the experiments of the connected pipe facility for SFRJ may sometimes be
run for short durations of one second or less, care should be taken not to measure the
combustion efficiency before the test facility reaches the steady-state condition. In the study
by Dijkstra et al.,(8) this was demonstrated by showing that the combustion efficiency at the
initial stage of experiment was lower (because of severe heat loss to the experimental facility)
than that at a later stage.
Generally, the combustion efficiency attainable is lower in metallised solid fuels parti-
cularly with boron or boron carbide. For metallised fuels, the diffusion flame is weaker
because there is less hydrocarbon. In these fuels, the metal particles tend to accumulate and
agglomerate on the regressing fuel surface before they are ejected into the gaseous flow. In
addition, the surface may produce large flakes that are ejected into the flow. These large
metal-particle agglomerates and flakes are difficult to ignite and also require a longer
residence time in the combustor to complete their burning.(15,17,26) In such situations in
order to improve combustion efficiency, as indicated earlier, introducing swirl to inlet air
flow and/or injecting bypassed air into the aft mixing chamber are found necessary. The
effect of swirl on boron combustion was investigated by the DLR group in Germany.(9,61)
They varied the percentage of boron and boron carbide in HTPB fuel from 10 to 40. They
defined a boron combustion efficiency by
reacted boron reacted boron
g " " (29)
"" total boron (reacted boron#unreacted boron)
Without swirl g decreased from 40% for fuel with 10% boron to 30% for fuel with 40%
""
boron. But with swirl flow (swirl No."0.54) the g improved by 18—20%. In fuels with
""
10% boron carbide the DLR group achieved boron combustion efficiency of 80% by
introducing swirl in the inlet air.(9,61) The effect of injecting bypass air into the aft mixing
chamber was studied by Natan and Netzer.(55) The bypass air injected into the aft mixing
238 S. Krishnan and P. George

chamber improves combustion efficiency in metallised fuels because of the following two
beneficial effects. First, it promotes the ignition of metal particles in the flow through the
grain passage because of the lower mass flux there; this produces a thicker and hotter flame
zone as well as increases the residence time of the particles. Second, the use of bypass air
enhances the combustion of the already ignited particles due to better mixing of the
particles with the bypass air in the aft mixing chamber. The combustion efficiency, (g ) in
"1
a boron carbide-fuelled SFRJ (B C : Mg : HTPB"50 : 5 : 45) increased from 36% at 0%
4
bypass ratio, to 85% at 42% bypass ratio.(55) In these tests the total mass flow rate of air
(\0.59 kg/s), combustion chamber pressure (\6 bar), inlet air temperature (\585 K), initial
port diameter (44 mm), and aft mixing chamber length (700 mm) were maintained constant.
To keep the bulk equivalence ratio constant, the fuel grain length was varied from 216 mm
for 0% bypass to 330 mm for 42% bypass.
The introduction of the swirl or the use of bypass air may however give an enhanced total
pressure loss in the combustor resulting in specific impulse reduction and thus it requires
a suitable compromise. In addition to the internal flow losses expected in the use of bypass
air, there is every possibility for enhancement in the external drag due to increase in frontal
and wetted areas. Our study using HTPB : magnalium : ammonium perchlorate
(AP)"30 : 40 : 30 fuel grain with a mixer plate and an aft mixing chamber diameter and
length of 90 and 135 mm, respectively indicated a combustion efficiency, (g ) , around 95%;
"2
in this study we used an initial port diameter of 60 mm and A /A of 2.6.(35) This high
1 5
combustion efficiency for the metallised fuel without the introduction of swirl or the use of
bypass air is due to the addition of AP and also possibly due to magnesium, a metal of low
melting point.

3. MODELLING STUDIES

The investigations in Israel and Germany on SFRJ have been mainly experimental
whereas in U.S.A. both experimental and modelling studies have been
undertaken.(46,48,56,57,76,77) In the Netherlands both experimental and modelling studies
have been reported.(10,12,13)
The modelling of SFRJ combustion should account for the following features. The SFRJ
consists of a solid fuel grain that provides the walls for the combustion chamber (Fig. 12).(48)
The sudden expansion at air inlet (either axial or side dump) can be used to provide flame
stabilisation. As already discussed, a turbulent boundary layer develops downstream of the
flow reattachment and includes a diffusion-controlled flame between the fuel-rich zone near
the wall and the oxygen rich central core. On account of this diffusion flame heat is
transferred by convection and radiation into the solid surface causing vaporisation of fuel.
Computer model simulation of SFRJ combustion process started from stream function
vorticity formulation (t—u model).(56) In this formulation Spalding’s PISTEP II model(21)
on heat and mass transfer in recirculating flows was adapted. Steady, recirculating, two-
dimensional, subsonic flow was the basic assumption. The turbulent Lewis number was
taken to be unity. A modified Jones—Launder two equation turbulence model was used to

Fig. 12. SFRJ combustion chamber geometries(48).


Solid fuel ramjet combustor design 239

Fig. 13. Reattachment location for axisymmetric flows(48,56).

calculate the viscosity throughout the flow field.(25,36) The combustion was considered to be
mixing limited with a single-step chemical reaction and the formulation neglected radiation.
For the purpose of comparison with other models, these assumptions are listed in Table 7.
The t—u model qualitatively predicted the effects of combustion geometry and test
environment on reattachment location, fuel regression rate, flame pattern, and combustion
efficiency. In a subsequent work,(57) the predictions of the model were compared with the
experimental data obtained by other investigators (Table 8).
Several problems were inherent to the t—u model. The most important ones are: first, the
predicted pressure distributions were normally inaccurate, and second, the boundary
conditions were difficult to specify. In view of the above problems, primitive variable
(velocity—pressure 2D) model developed by Pun and Spalding(64) (CHAMPION 2/E/FIX)
was subsequently adapted(77) to the SFRJ geometry. The basic assumptions in this model
were same as in the previous one. This model calculated also the aft mixing region fluid
flow. Subsequently, the same model was extended to consider the effect of radiation from
carbon particles within the flame zone; however, the gas-phase radiation was neglected.(46)
The effect of this radiation-consideration increased the regression rate downstream of
reattachment point resulting in a better agreement with experiment. This velocity—pressure
model was later improved considering three-dimensional effect.(48) In this study an exten-
sion of the CHAMPION code namely, ‘‘GARRAT 3D’’ (developed by GARRAT Turbine
Engine Co, U.S.A.)(49) was adapted. The model included the finite chemical kinetics through
a four step process as well as the turbulent mixing process. The code was made to take
automatically the slower of the two processes to be rate controlling at each solution point
within the grid flow field. This study neglected the radiation effect. The assumptions
involved in the velocity—pressure models are also given in Table 7. Comparisons between
the predictions of the velocity—pressure 2D model and the experimental data are detailed in
Table 8. The velocity—pressure 3D model(48) predicted the location of the maximum
regression rate point in good agreement with the Krall and Sparrow(34) maximum heat
transfer point but with the experimental maximum regression rate still downstream. At
a combustion chamber pressure of 4 bar, the regression rate profile predicted by the 3D
model, that automatically takes into account the slower of the two processes (chemical
kinetics and turbulent mixing) to be rate controlling, is very identical to the 2D model, that
takes into account only the turbulent mixing process. This indicates that most of the process
is mixing limited at 4 bar. However, at high altitude conditions that result in situations of
low combustion chamber pressures, the chemical kinetics may become rate controlling.
In view of all the above, the SFRJ modelling effort at its present state is unable to give
quantitative results on the characteristics of fuel regression rate. However, it does qualitat-
ively predict the effects of the fuel compositions, the combustor geometry, and the inlet fluid
dynamic properties on the combustor flow field and the fuel regression rate pattern. The
modelling of other combustion characteristics such as flammability and combustion effi-
ciency are yet to be included in the effort.
240

Table 7. Assumptions involved in the t—u and velocity—pressure models

Assumptions t—u model Velocity—Pressure 2 D model Velocity—Pressure 3 D model

Heat and mass transfer Gosman—Spalding PISTEP II model Pun-Spalding CHAMPION 2/E/FIX code ‘‘GARRAT 3D’’ code
for heat and mass transfer
Flow field Steady, elliptic, two dimensional, Steady, elliptic, two dimensional, subsonic Steady, elliptic, three dimensional, subsonic
subsonic
Lewis number Lewis No."1 Lewis No."1 Lewis No."1
assumption

Viscosity Modified Jones—Launder two Modified Jones—Launder two equations turbulence model Modified Jones—Launder two equations turbulence
equations turbulence model used used to calculate viscosity model used to calculate viscosity
to calculate viscosity
Reaction and mixing Mixing limited; one step simple Mixing limited; one step simple chemical reaction Finite rate chemical kinetics; four step chemical
chemical reaction reaction and turbulent mixing—slower of the
S. Krishnan and P. George

two automatically taken as the rate controlling


Radiation Radiation neglected Initially radiation neglected, but later considered; only the Radiation neglected
dominant effect of radiation from carbon particles within the
flame zone are considered; gas phase radiation neglected
Table 8. Results of the t—u and velocity—pressure models

Results t—u model Velocity—pressure 2 D model

Reattachment location (Fig. 13)(48,56) Predicted location is downstream of Krall-Sparrow(34) maximum heat transfer Same as in t—u model
point in a non-reacting high enthalpy flow and upstream of
Phaneuf-Netzer(62) experimental reattachment location for non-reacting air
flow with 10% blowing; reattachment location is insensitive to inlet velocity
Maximum regression rate point and Predicted maximum regression rate point is upstream of the experimental value Similar as t—u model, slightly better agreement with experi-
regression rate profile (Fig. 14)(77) and slightly upstream of Krall-Sparrow maximum heat transfer point; no mental values.
information is available on the relative location of experimental maximum
regression rate point with respect to reattachment points (predicted as well as rR JG0.31—0.34
13%$*#5%$ !
rR JG0.38—0.41
experimental); regression rate is under-predicted down-stream of maximum %91%3*.%/5!- !
regression rate point and over-predicted upstream
Centerline turbulence intensity Experimental data are available only for cold flow; maximum turbulence Maximum centerline turbulence intensity underpredicted;
(Fig. 15)(77) intensity over-predicted; turbulent intensity under-predicted near the rearward step better agreement of turbulence intensity at farther points
but over-predicted at farther points
Axial pressure distribution Normally inaccurate prediction No comparison with experimental values; predicted distribution
compares well with the expected variation
Solid fuel ramjet combustor design

Radial temperature distribution 30% arbitrary increase in the regression rate in a way to compensate neglected Predictions are presented without experimental comparisons; as
radiation effect has been introduced in temperature prediction; predicted air mass flow increases the centerline temperature decreases and
temperature profiles are in better agreement with experiment down-stream of maximum temperature point moves closer to the fuel surface but
reattachment point; other predictions are presented without experimental this movement is less sensitive to mass flow than that in t—u
comparisons; the centerline temperature increase with axial distance is reasonably model; temperature profiles and flame sheet locations in the aft
predicted; as air mass flux increases the centerline temperature decreases and the mixing chamber are also predicted
maximum temperature point moves closer to the fuel surface
241
242 S. Krishnan and P. George

Fig. 14. Polymethylmethacrylate regression rate(77).

Fig. 15. Centreline turbulence intensity(77).

4. DESIGN OF AN SFRJ-ASSISTED PROJECTILE

With a view to demonstrate the use of the information collected through the present
review, a preliminary design analysis is presented in this section for an SFRJ-assisted gun
launched projectile of pseudo-vacuum trajectory. In line with the subject matter of this
review only information related to the SFRJ combustor design is elaborated here. In any
propulsive system design the ‘‘rubber engine’’ analysis is the first step. By carrying out such
analysis for different operating conditions one can estimate the engine-configuration to be
used for the projectile. Starting with this configuration, one can iteratively arrive at the final
configuration by conducting detailed flight analysis. The ‘‘rubber engine’’ analysis presented
here assumes that the inlet area, fuel grain length, and nozzle throat area are infinitely
variable; and that the nozzle is always operating under adapted condition (p "p ). The
% !
projectile envelope and its performance, and the properties of suitable fuel are the two
major aspects discussed here.
A typical construction of an SFRJ assisted gun launched projectile is as given in Fig. 16. It
is of two parts approximately of equal lengths; the front part, of diameter a little less than
the gun barrel diameter, houses a payload. At the nose of the front part is the inlet closed by
a frangible diaphragm (diaphragm not shown in the figure). The rear part is the actual
engine in which the fuel grain is stored. The outer diameter of the rear part is less than that
Solid fuel ramjet combustor design 243

Fig. 16. Solid fuel ramjet assisted projectile and its station numbers.

of the front part and this provides an annular passage between the gun barrel and the
engine-exterior. At the middle, the obturator on the periphery, together with the one-way
valve (not shown in the figure) inside the projectile separating the front and rear part, act as
a piston.
The operating principle is as follows. On firing the gun, high pressure and high temper-
ature gases are evolved by the gun propellant. These gases fill the annular passage between
the gun barrel and the engine-exterior, and the space within the engine; and, forcing the
piston, push the projectile. The fuel grain in SFRJ is thus exposed to high temperature
('3000 K) and very high pressure (few thousand bars!) combustion gases during the
projectile’s travel within the gun barrel. This travel time is very short, a few milliseconds.
The projectile is then ejected into the atmosphere at a supersonic Mach number around or
greater than 2. From this instant, the sudden depressurization within the SFRJ, the opening
of intake by the release of the frangible diaphragm and the one-way valve, and the gushing
of air into the SFRJ take place in very quick successions in a matter of a few milliseconds.
Air flows into the SFRJ with a stagnation temperature of 540 K or more, and through
external and internal deceleration processes its Mach number is decreased to a value
around 0.3—0.4 at the air inlet of the rearward step. Having been exposed to high temper-
ature and very high pressure gases within the gun barrel and now on being exposed to high
temperature air, the fuel grain may get automatically ignited. This is definitely possible at
very high ejection Mach numbers, say 4 or more for which the stagnation temperature is
well in excess of 1200 K—ejection Mach number around 4 is possible for very small
projectiles. However, a point not in support of easy automatic ignition is the cooling caused
by the sudden depressurization within the SFRJ. Nevertheless, in order to ensure ignition
within practical time-limits, one or both of the following may be adopted: (1) provision of
a pyrotechnique igniter (schematically shown in Fig. 1) and (2) giving a thin coat of an easily
ignitable propellant on the fuel grain port. Though the grain composition can be all-fuel,
giving maximum specific impulse, flight tested solid fuels to date contain certain oxidiser
fractions in order to realise easy ignition and desirable burning rate(59)—of course, at the
expense of fuel specific impulse. The nature of the flow field in the combustor has already
been explained. The hot combustion products thus formed are accelerated through the
nozzle with exit momentum greater than the inlet value, thereby producing thrust. The
range of projectile, therefore, can be considerably enhanced with the assistance of an SFRJ.
A pseudo-vacuum trajectory of a projectile in air is the one in which the drag experienced
by the projectile is always balanced by the thrust produced by its propulsive unit. Adoption
of this trajectory has two principal advantages.(71) The first one is the easy and accurate
predictability of the trajectory. The second is the insensitiveness of the trajectory to the
external disturbances such as winds. Due to this, a circular error probable of even one order
of magnitude less than that from an equivalent conventional-trajectory is feasible. Unlike in
the fire-and-forget type, in a manoeuvrable version the major part of the flight path of
244 S. Krishnan and P. George

projectile is psuedo-vacuum one; with the change in the launch angle the range can be
altered. At a pre-set time corresponding to an appropriate altitude/range the terminal
manoeuvre is initiated after (1) blowing off the combustion (say, by shutting of the inlet),
(2) deploying the canards (see Fig. 16), and (3) reducing the flight path decent angle. A seeker
control device placed at the nose of the projectile scans through its field of view. When
a target is located canard control signals are developed to command a manoeuvre that will
impact the projectile against the target. Since the fuel grain in the SFRJ has to cater for the
complete trajectory till touch-down and our aim is to design mainly the combustor the
above terminal manoeuvre need not be considered on our design.

4.1. ANALYSIS

4.1.1. Launch velocity and trajectory

The launch velocity, º , delivered by a gun to a reference projectile mass, m is known


0 0
from the gun designer. The launch velocity, º , imparted to a projectile of a different mass,
l
m , can be obtained using the law of conservation of energy,
P
1 m º2"1 m º2 (30)
2 0 0 2 P l
However, there is an approximation in this. Generally, in the case of firing an unpowered
projectile the initial free volume to be occupied by gun propellant gases will be less than that
in the case of firing an SFRJ assisted projectile of the type shown in Fig. 16. Because of this
enhanced free volume, the gun pressure will be lower for the given mass of gun-propellant.
Therefore, the projectile launch velocity will be slightly lower than that from the Equation
(30) for the given gun propellant charge.
The trajectory equations for a pseudo-vacuum flight are
º2 sin 2h
R" l (31)
g
º2 sin2 h
H" l (32)
2g

2º sin h
t " l (33)
505 g

4.1.2. Mass of projectile

The different parts of the projectile that contribute to its total mass are: (1) intake and
struts, (2) central-body housing seeker, controls, payload, and other electronics, (3) combus-
tion-chamber housing fuel grain and aft chamber, (4) fuel grain, and (5) nozzle. Combustion
chamber mass and fuel grain mass are dependent on the fuel grain length. The combustion
chamber mass consists of the mass of the insulation liner between fuel and chamber, and
mass of the chamber. The mass of liner,
n
m " (OD2!ID2) (¸ #¸ )o (34)
L 4 L L F !&5 L

and the chamber mass,


n
m " (OD2 !ID2 ) (¸ #¸ )o (35)
CH 4 CH CH F !&5 CH

Finally, the total mass of the combustion chamber,


m "m #m (36)
C CH L
Solid fuel ramjet combustor design 245

The fuel-grain length ¸ necessary for maintaining the required fuel flow rate, mR , is
F F
mR
¸ " F (37)
F nD rR o
1 F
The total fuel mass is calculated as
n
m " (OD2!ID2) ¸ o (38)
F 4 F F F F

For the fuel found suitable for the SFRJ the regression rate equation can be of the form
rR "AG/D2¹.p" (14a)
! 1 !
where
4mR
G" ! (14b)
! nD2
1
Here, through experiments, the values of the constants n, q, m, and b [Eq. (14a)] are known
for the basic fuel being developed. Also depending upon the actual regression rate,
a ‘‘fine-tuning’’ of the regression rate equation is possible by adding certain catalysts or by
slightly changing the fuel composition, in order to arrive at a suitable value for A.

4.1.3. Drag

The fore body or nose of the projectile may have many shapes such as conical, ogival, and
hemispherical. At supersonic speeds the wave drag may be several times that due to friction
for a fore body. Therefore, a careful selection of the fore body is mandatory and usually an
ogival shape is preferred. Using the semi-empirical relations derived by Miles(47) for ogival
configuration, the wave drag coefficient,

C D
*p 2[196(l/d)2!16]
C " 1! (39)
D8 q 28(M#18) (l/d)2
where

C DA B
*p 0.096 p 1.69
" 0.083# 0 (40)
q M2 10
and

A B
1
p "2 tan~1 in degrees (41)
0 2(l/d)
The wave drag is then given by
Dg "(n/4) (1/2)o º2 C (OD )2 (42)
8 ! D8 P
The frictional drag, the second major drag component, is composed of body skin friction
drag and fin skin friction drag. The skin friction coefficient in incompressible flow,
C depends primarily on the type of flow, laminar or turbulent, and Reynolds number. The
&5
Reynolds number
o º¸
Re" ! 1 (43)
k
is used for the calculations of body skin friction drag (Dg ), whereas the Reynolds number
B
o º¸
Re" ! & (44)
k
is used for the calculations of fin skin friction drag (Dg ). The Karman—Shoenherr(39)
&
equation for incompressible turbulent flow is

JC log (C Re)"0.242 (45)


&5 10 &5
246 S. Krishnan and P. George

The expression for the compressibility effect(66) is

C 1
&" (46)

C A B D
C c !1 0.467
&5 1# ! M2
2

Therefore, the drags due to skin friction are

Dg "1 K o º2C p(OD ) ¸ (47)


B 2 B ! & P P
and
Dg "1K o º2C A n (48)
& 2 F ! & & &
where K and K are multiplication factors for protuberances and junctions. Including the
B F
interference drag (Dg ), the total projectile drag,
I
Dg " Dg #Dg #Dg #Dg (49)
8 B & I
The interference drag, Dg , as a first approximation, can be taken to be a small fraction of
I
the sum of the other three drags and 0.05 is appropriate. The variations of pressure,
temperature, and density with respect to altitude are known from appropriate standard
tables or fitted equations.
Further analysis of the SFRJ projectile requires an insight into the components’ figures of
merit (Fig. 16). Intake stagnation pressure recovery ratio for critical operation, p /p , is
02 0!
dependent on flight Mach number around the design value and can be obtained from wind
tunnel tests. For preliminary design the effect of angle of attack on p /p may be neglected.
02 0!
The stagnation pressure loss factor across the rearward step, p /p can be calculated from
03 02
the following empirical relation:(27)

C A BD
p c A 2
03"1! ! M2 1! */ (50)
p 2 2 A
02 1
The other figures of merit, viz., p /p of the combustor and p /p of the nozzle are
06 03 08 06
known from experience and can be taken to be constants.

4.1.4. Inlet area and nozzle throat area

The next step in the analysis involves the determination of mR and mR , the mass flow rates
! F
of air and fuel, respectively. To do this one must know the specific impulse values for the air
as well as the fuel. These are obtained from a knowledge of the properties of the products of
combustion and the velocity of the projectile,

F
"Isp "(1#f ) º !º (51)
mR ! %
!

S C D
2 Ru c p ((c~1)@c)
º" ¹ 1! ! (52)
% Mo (c!1) 06 p
1 08
F
"Isp "Isp / f (53)
mR F !
F
Here the value of f is taken as the stoichiometric value. Mo c, and adiabatic flame
1,
temperature ¹@ can be obtained from a standard program such as the NASA CEC71.(20)
06
A suitable value for combustion efficiency g can be assumed to calculate ¹ "
" 06
g (¹@ !¹ )#¹ ,
" 06 ! !
mR "F/Isp (54)
! !
mR "F/Isp (55)
F F
Solid fuel ramjet combustor design 247

The nozzle throat area required at each instant,

S
Ru¹
mR (1#f ) 06
! Mo
A" 1 (56)
5
A B
2 (c`1)@(2(c~1))
p Jc
07 c#1
A ’s for various instants are calculated following the ‘‘rubber engine’’ concept and a value
5
slightly above the maximum of these A ’s can be adopted for the SFRJ as a preliminary
5
design value. The inlet area, A is given by
*
mR
A" ! (57)
* ºo
!
and has its maximum value at ground level. A suitable value is to be selected for the inlet
area. As discussed in Section 1 the required control of engine mass flow rate (mR #mR ) can
! F
be brought about by either of the two methods, one the bypass control of inlet air and the
other the fuel regression rate control such as ‘‘tube in hole’’.(28) Again as a first approxima-
tion an inlet area value A slightly above this ground-level-maximum can be adopted.
*

4.1.5. Air inlet area of rearward step and port area

The Mach number at air-inlet of rearward step M should be set at a minimum possible
2
value in order to achieve better flammability limits (Section 2.3). For a gun launched SFRJ
of fixed A the maximum M occurs at launch and a value of around 0.4 appears
*/ 2
appropriate. A can be calculated for the chosen values of A p /p , launch Mach number
*/ *, 02 0!
M , and M
1 2
c !1

C D
1# ! M (c!`1)@(2(c!~1))
AM 2 2
A " * 1 (58)
*/ p c !1
02 M 1# ! M
p 2 2 1
0!
As per Tables 2 and 3, (h/D )* can be selected based on the type of fuel, the air stagnation
1
temperature at launch, ¹ and the approximate A /A .
! 1 5

4.2. PROCEDURE

Here we present a preliminary design procedure to arrive at the combustion chamber


dimensions of a 155 mm gun launched SFRJ projectile. By a separate study the following
fixed masses were approximately arrived at:
(1) Nozzle (mass proportional to total impulse)"2 kg.
(2) Payload (specified)"7 kg.
(3) Seeker controls and other electronics"3 kg.
(4) Centre body of the intake housing items (2) and (3)"20.5 kg.
(5) Intake outer shell and the struts connecting centre body and outer shell"17 kg.
(6) Foldable aft fins"1.5 kg.
The above add-up to 51 kg. The other masses to be determined are those of fuel grain,
liner, and combustion chamber. These depend upon the design parameters such as launch
angle, launch velocity, grain outer diameter, and ogival nose configuration. Based on an
independent study the following dimensions have been approximately fixed and maintained
as constants throughout the present combusion chamber design-calculations:
Length of the front part"810 mm.
Outer diameter of the front part (specified)"155 mm.
Aft mixing chamber length"250 mm.
Nozzle length"70 mm.
Aft fin dimensions"253 * 135 * 50 mm (trapezoidal).
248 S. Krishnan and P. George

For an inlet required to operate around a specified Mach number, P /P can be


02 0!
expressed in an empirical form fitted from experimental results. For the present class of
155 mm-projectile the launch Mach number is around 2. A typical fit for flight Mach
numbers around 2 is
P
02"!0.625M2#2.127M!0.936 (59)
P
0!
As the Mach number at air-inlet of rearward step, M and the average initial Mach number
2
at grain port entry are to be low (M ) 0.4 and M ) 0.3) D and ID are fixed as 60 and
2 3 */ F
90 mm, respectively (Sections 2.3 and 4.1.5). Now the dimension to be arrived at is the fuel
grain length. This is calculated as per the procedure given below. The procedure assumes
a quasi-steady-state condition during a selected time interval. One-second time interval is
found to be sufficiently accurate for the present projectile having a total flight time of
80—120 s.
To begin with, fix the following parameters: launch angle (h), ogival nose slenderness ratio
(l/d), outer diameter of the projectile-rear-part (OD ), reference mass (m ) and reference
CH 0
velocity (º ). Based on the maximum pressure and temperature experienced by the
0
chamber wall, calculate the wall thickness, *L . For the fixed OD , ID "
CH CH CH
OD !2*¸ . Noting OD "ID and choosing a suitable liner thickness *¸ arrive at
CH CH L CH L
the inner diameter of liner, ID that is equal to the outer diameter of fuel grain, OD . For
L F
the present calculations *¸ and *¸ are selected as 2.5 mm each and treated as constants
CH L
because their variations with different operating conditions are minimal. Hence OD
F
(instead of OD ) becomes the fixed parameter.
CH
Step I: Choose a trial average-value of fuel grain length, ¸ .
F5
Step II: Using Equations (34), (35), and (38) calculate the liner mass (m ), the combustion
L
chamber mass (m ), and fuel grain mass (m ), respectively. Thus, the total trial mass of the
CH F
projectile, m "(51#m #m #m ) kg.
P5 L CH F
Step III: Using equation (30) calculate the launch velocity º for m .
l P5
Step IV: Using ¸ and the previously fixed dimensions, calculate the trial total length of
F5
projectile, ¸ .
P5
Step V: Using a trial value of regression rate rR for the launch values of p, G ¹ , and
! !
D "90 mm, evaluate the constant, A in Equation (14a) (Section 4.1.2)—for the present
p
calculations n"0.4, q"!0.25,m"0.4, and b"0.4. Here ¹ is the temperature of inlet
!
air with reference to experimental test facility. At port entry, Mach number is ) 0.3.
Therefore, for a projectile in flight, ¹ can be taken as the stagnation temperature of the
!
ambient air corresponding to M.
Step VI: At each instant evaluate the following parameters:
(a) Regression rate, using Equations (14a) and (14b)—for launch condition (t"0) this is
already chosen as a trial value
(b) Drag to be overcome, using Equations (39)—(49)
(c) Mass flow rates of air and fuel required, using Equations (51)—(55)
(d) Fuel grain length required, using Equation (37)
(e) Inlet area, using Equation (57)
(f) Nozzle throat area, using Equation (56).
Step VII: Under quasi-steady-state assumption, the new port dimension after the time
interval
D "D #2rR *t (60)
p,i`1 p,i i
Step VIII: Once the total time of flight t (Equation (33)) is covered by incrementing the
505
time intervals, check the final value of the grain port diameter with OD . If they are not
F
equal (within an error limit of 0.1%), using Regular Falsi method alter the trial value of
regression rate rR , and go to Step V. This is done till the convergence on OD is reached. In
F
all these steps the value of the projectile length (which is used in drag calculations) is kept
constant, although fuel length and hence the projectile length change every time Step »I is
executed.
Solid fuel ramjet combustor design 249

Step IX: On OD convergence the fuel grain lengths and the nozzle throat diameters are
F
averaged over the entire time of flight. Use this averaged fuel grain length, ¸ , to calculate
F
the total projectile length, ¸ . Check this ¸ with ¸ . If they are not equal (within an error
P P P5
limit of 0.1%), set ¸ equal to ¸ and go to Step IV.
F5 F
Step X: On ¸ convergence the projectile mass m is calculated as per procedure detailed
P P
in Step II. Check this m with m . If they are not equal (within an error limit of 0.1%), set
P P5
m equal to m and go to Step III.
P5 P
Step XI: On m convergence the projectile performance is recorded.
P

4.3. RESULTS

The above-mentioned procedure was carried out with a reference mass of 43.88 kg for all
possible combinations of the following parameters.

Reference velocities (m/s)"830 and 897.


Launch angles (deg)"30 to 45.
Ogival nose slenderness ratios (l/d),"2.0, 2.15, 2.3, and 2.5.
Outer diameter of the projectile-rear-part, OD (mm)"135, 138.3, 141.6, and 145.
CH
The typical variations of A , ¸ and A that occur for launch angle of 35°, l/d ratio of 2.5,
* F 5
and OD of 145 mm are shown in Figs 17, 18 and 19, respectively. Referring to Figs 17—19,
CH
we note the following for the infinitely varying quantities: (1) the inlet area is maximum at
launch/touch-down, (2) the fuel length variation is within a range of 6%, and (3) the nozzle
throat area is maximum at peak altitude. This indicates to us that in order that the projectile
performs satisfactorily at all conditions of operation with necessary control (bypass control
of inlet air or fuel regression rate control) the chosen fixed inlet area and throat area should
be above their maxima and the chosen fixed fuel length within its range of variation. These
dimensions can be finalised only through a detailed iterative flight analyses which is outside
the scope of the present paper. But the message is, the rubber engine analysis helps in very
closely arriving at the fixed dimensions of the practical SFRJ within 10% of the values
indicated.
The reduction in launch angle from 45 to 30° reduces the maximum height by 52% (from
11.5 to 5.6 km), and the flight time by 30% (from 96.8 to 67.5 s) but the range comes down
only by 16% (from 45.6 to 38.4 km) (Fig. 20). Also the corresponding variations in thrust
and Mach number during the flight are less. The reduced launch angle requires enhanced

Fig. 17. Variation of inlet area with time. Launch angle"35°, reference velocity"830 m/s, nose slenderness
ratio"2.5, outer diameter of the projectile-rear-part"145 mm.
250 S. Krishnan and P. George

Fig. 18. Variation of fuel grain length with time. Launch angle"35°, reference velocity"830 m/s, nose slender-
ness ratio"2.5, outer diameter of the projectile-rear-part"145 mm.

Fig. 19. Variation of nozzle throat area with time. Launch angle"35°, reference velocity"830 m/s, nose
slenderness ratio"2.5, outer diameter of the projectile-rear-part"145 mm.

projectile-length and -mass; however, these are only marginal—9 and 5%, respectively.
Thus, we see that with the reduction in launch angle the projectile and hence the SFRJ,
without much loss in range, operates with less variation in environmental operating
conditions.
The major drag contribution is due to wave drag. The increase in nose-slenderness ratio,
leading to a consequent reduction in drag, significantly improved projectile performance in
many ways: (1) enhanced range (Fig. 20), (2) higher fuel regression rate, and (3) shorter fuel
grain length (therefore shorter and lighter projectile) (Fig. 21). Low fuel-regression-rate
required for low nose-slenderness-ratio may be very difficult to realise in the SFRJ-fuel
development. A significant reduction in fuel grain length (about 200—300 mm) because of
higher nose-slenderness-ratio results in increased acceleration time in the gun barrel (higher
launch velocity) as well as better flight performance with not-too-high a projectile-slender-
ness-ratio.
Internal external pressure equalisation for the projectile during its travel in the gun-barrel
depends on the annular gap between gun-barrel and engine chamber. An increase in this
Solid fuel ramjet combustor design 251

Fig. 20. Maximum altitude and range vs launch angle for different nose slenderness ratio.

Fig. 21. Initial regression rate and projectile length vs launch angle for different nose slenderness ratio.

gap (from 5 to 10 mm) demands lower fuel regression rate and longer fuel grain (about
300—400 mm) and hence a highly slender projectile (Fig. 22).

5. CONCLUDING REMARKS

The functional simplicity of SFRJ combined with high performance makes this engine
quite attractive for powering missiles and gun-launched projectiles. Many combustion
252 S. Krishnan and P. George

Fig. 22. Initial regression rate and projectile length vs launch angle for different combustion chamber outer
diameter.

studies have been carried out on connected-pipe test facilities. Selection of fuel type,
flammability limits, solid fuel regression rate, and combustion efficiency are the different
aspects considered in these studies.
HTPB with or without polystyrene appears to be a better-suited fuel grain material.
Addition of metal powders will further increase gravimetric and volumetric heat release but
may pose problems of ignition and combustion inefficiency, particularly in the cases of
boron and aluminium additions; addition of magnalium seems to be a better compromise in
this regard. Though the grain composition can be all-fuel, giving maximum specific impulse,
flight tested solid fuels to date contain certain oxidiser fractions in order to realise easy
ignition and desirable burning rate—of course, at the expense of fuel specific impulse.
At the beginning of operation, a minimum (or critical) ratio of rearward step height to
port diameter, (h/D )* is required for combustion to occur in an SFRJ and this ultimately
1
limits the maximum pressure recovery and the maximum fuel grain loading, thus the range
of missile. Studies show that a lower (h/D )* can be provided if, (1) A /A is higher (the
1 1 5
combustion chamber velocity is lower), (2) the inlet air temperature is higher, (3) the port
diameter is larger, and (4) the fuel regression rate is faster.
In SFRJ, the solid fuel regression pattern along the grain passage is complex. The
regression is minimum just downstream of the rearward step. It increases with axial distance
and attains a maximum at the flow reattachment region, which itself is shifting downstream
from time to time as the rearward step height is increasing. Both the local and instantaneous
regression rate and the time averaged regression rate can be adopted. For the latter, various
empirical relations are available. In these relations the values of indices and coefficients are
significantly different. An attempt is made to explain the reasons for such differences.
Depending upon the combustion chamber operating conditions, the average regression rate
can be taken to be largely influenced either by the heat transfer mechanism downstream of
the reattachment region or by that in the region itself. In the first case, using the analogy of
heat transfer in a pipe-flow for a fully developed turbulent boundary layer, the regression
rate for the typical SFRJ operating temperature range of 300—800 K,

rR "AG0.8D~0.2¹0.14pb
! 1 !
Solid fuel ramjet combustor design 253

In the second case it is


rR "AG0.67D0.33D~0.67¹0.23p"0
! 1 */ !
However in the time-averaged regression rate procedure the ‘‘apparent’’ indices can be
significantly different from the indices values corresponding to the fundamental heat
transfer mechanism. Therefore, the values of indices found in various correlations only
demonstrate the influence pattern of various parameters and the application of these values
in actual design should be done with caution.
An aft mixing chamber is found necessary to improve combustion efficiency. It is usually
fitted with a mixer plate at its front for additional mixing and consequent enhancement in
combustion efficiency. In this the reaction between fuel and air is completed due to better
mixing. The parameters controlling combustion efficiency are: (1) aft mixing chamber length
to initial port-diameter ratio, ¸ /ID , (2) equivalence ratio, and (3) h/D . The typical
!&5 F 1
combustion efficiency attained is above 90% for unmetallised solid hydrocarbon fuel.
However such an efficiency is high to be achieved in metallised solid fuels particularly with
boron or boron carbide. In the case of metallised fuel being used, introducing swirl to inlet
air flow and/or injecting bypassed air into the aft mixing chamber are found necessary to
have an acceptable combustion efficiency but at the expense of total pressure.
Detailed information is not available on (1) the ignition and combustion in SFRJ at low
pressures (0.5—3 bar), (2) the rigorous correlation for fuel regression rate behaviour along
the fuel grain passage, and (3) the effect of swirl and bypass ratio towards total pressure
loss.
Fluid dynamic computations of simplified SFRJ-combustion-processes are able to dem-
onstrate the experimental trends related to reattachment location, maximum regression rate
point, and regression rate profile.
With a view to demonstrate the use of the information collected through this review,
a preliminary design procedure is presented for an SFRJ assisted gun launched projectile of
pseudo-vaccum trajectory. A parametric study conducted using this procedure reveals the
following: (1) The reduction in launch angle from 45 to 30° reduces the maximum height by
52%, and the flight time by 30% but the range comes down only by 16%. (2) Increase in
nose-slenderness-ratio with a consequent reduction in drag, significantly improves projec-
tile performance in many ways: enhanced range, higher fuel regression rate, and shorter fuel
grain length. (3) An increase in annular gap between gun-barrel and projectile demands
lower fuel regression rate and longer fuel grain and hence a highly slender projectile.

ACKNOWLEDGEMENTS

S. Sathyan and S. Sanjay (B. Tech students of the Department of Mechanical Engineering
of this Institute, class of 1997) developed the necessary computer code and obtained the
results presented here.

REFERENCES

1. AGARD (1992) Airbreathing propulsion for missiles and projectiles, AGARD Conf. Proc., Vol. 526, p. T5.
2. Angus, W. J., Witt, M. A., Laredo, D. and Netzer, D. W. (1993) Solid fuel supersonic combustion, Recherche
Aerospatiale 6, 1—8.
3. Ben-Arosh, R. and Gany, A. (1992) Similarity and scale effects in solid-fuel ramjet combustors, J. Propulsion
Power 8, 615—623.
4. Bertin, J. J. (1994) Hypersonic Aerothermodynamics. AIAA Education Series, Washington DC.
5. Boardman, T. A., Carpenter, R. L., Goldberg, B. E. and Shaeffer, C. W. (1993) Development and testing of
11- and 24-inch hybrid motors under the joint government/industry IR&D program, AIAA Paper No.
93-2552.
6. Chiraveni, M. J., Harting, G. C., Lu, Y., Kuo, K. K., Serin, N. and Johnson, D. K. (1995) Fuel decomposition
and boundary-layer combustion processes of hybrid rocket motor, AIAA Paper No. 95-2686.
7. Comphell Jr. W. H., Ko, B. N., Lowe, S. R. and Netzer, D. W. (1992) Solid-fuel ramjet fuel regression
rate/thrust modulation, J. Propulsion Power 8, 624—629.
254 S. Krishnan and P. George

8. Dijkstra, F., Korting, P. A. O. G. and Berg, R. van der (1990) Ultrasonic regression rate measurement in solid
fuel ramjets, AIAA Paper No. 90-1963.
9. Duesterhaus, D. A. and Hogl, A. (1988) Measurements in a solid fuel ramjet combustion with swirl, AIAA
Paper No 88-3045.
10. Elands, P. J. M. (1987) The prediction of the flow and combustion in a solid fuel combustion chamber by
means of two combustion models based on the diffusion flame concept. AIAA Paper No. 87-1702.
11. Elands, P. J. M., Dijkstra, F. and Zandbergen, B. T. C. (1990) Experimental and computational flammability
limits in a solid fuel ramjet, AIAA Paper No. 90-1964.
12. Elands, P. J. M., Dijkstra, F. and Zandbergen, B. T. C. (1991) Validation of flow and combustion process in
a solid fuel ramjet, AIAA Paper No. 91-1869.
13. Elands, P. J. M., Korting, P. A. O. G., Wijchers, T. and Kijkstra, F. (1990) Comparison of combustion
experiments and theory in polyethylene solid fuel ramjets, J. Propulsion Power 6, 732—739.
14. Fink, M. R. (1982) Aerodynamic properties of an advanced indirect fire system (AIFS) projectiles,
J. Spacecrafts Rockets 19, 36—40.
15. Gany, A. (1993) Combustion of boron-containing fuels in solid fuel ramjets. In: Combustion of Boron Based
Solid Propellants and Solid Fuels (eds K. K. Kuo and R. Pein), Begell House and CRC Press, Boca Raton, FL,
pp. 91—112.
16. Gany, A. and Netzer, D. W. (1985) Fuel performance evaluation for the solid fueled ramjet, Int. J. ¹urbo Jet
Engines 2, 157—168.
17. Gany, A. and Netzer, D. W. (1986) Combustion studies of metallized fuels for solid-fuel ramjets, J. Propulsion
Power 2, 423—427.
18. Gany, A. and Timnat, Y. M. (1993) Advantages and drawbacks of boron-fueled propulsion, Acta Astronautica
29, 181—187.
19. Geld, C. W. M. Van der, Korting, P. A. O. G. and Wijchers, T. (1990) Combustion of PMMA, PE, and PS in
a ramjet, Combust. Flame 79, 299—306.
20. Gordon, S. and McBride, J. B. (1971) Computer Program for Calculation of Complex Chemical Equilibrium
Compositions, Rocket Performance, Incident and Reflecting Shocks, and Chapman-Jouguet Detonations.
NASA SP-273, National Aeronautics and Space Administration, Washington DC.
21. Gosman, A. D., Pun, W. M., Runchal, A. K., Spalding, D. B. and Wolfshtein, M. (1969) Heat and Mass ¹ransfer
in Recirculating Flows, Academic Press, New York.
22. Hadar, I. and Gany, A. (1992) Fuel regression rate mechanism in a solid fuel ramjet, Propellants, Explosives,
Pyrotechn. 17, 70—76.
23. Hewett, M. E and Netzer, D. W. (1981) Light transmission measurements in solid fuel ramjet combustors,
J. Spacecraft Rockets 18, 127—132.
24. Ivarsson, U. (1994) Entirely up to speed, Int. Defence Review. 3, 63.
25. Jones, W. P. and Launder, B. E. (1972) The prediction of laminarization with a two-equation model of
turbulence, Int. J. Heat Mass ¹ransfer 15, 301—304.
26. Karadimitris, A., Scott II, C., Netzer, D. W. and Gany, A. (1991) Regression and combustion characteristics of
boron containing fuels for solid fuel ramjets, J. Propulsion Power 7, 341—347.
27. Kays, W. M. (1950) Loss coefficients for abrupt changes in flow cross section with low Reynolds number flow
in single and multiple tube systems, ¹rans. Am. Soc. Mech. Engrs, Paper 50-S-7, 72, pp. 1067—1074.
28. Keirsey, J. L. (1986) Solid Fuel Ramjet Flow Control Device. United States Patent No. 628,688.
29. King, M. K. (1974) Boron particle ignition in hot gas streams. Combust. Sci. ¹ech. 8, 255—273.
30. King, M. K. (1982) Ignition and combustion of boron particles and clouds, J. Spacecraft Rockets 19,
294—306.
31. Korting, P. A. O. G., Geld, C. W. M. Van der, Vos, J. B., Wijchers, T., Schoyer, H. F. R. and Nina, M. N. R.
(1986) Combustion behaviour of PMMA in a solid fuel ramjet, AIAA Paper No. 86-1401.
32. Korting, P. A. O. G., Geld, C. W. M. Van der, Wijchers, T. and Schoyer, H. F. R. (1990) Combustion of
polymethylmethacrylate in a solid fuel ramjet, J. Propulsion Power 6, 263—270.
33. Korting, P. A. O. G., Schoyer, H. F. R. and Timnat, Y. M. (1987) Advanced hybrid rocket motor experiments,
Acta Astronautica 15, 91—104.
34. Krall, A. M. and Sparrow, E. M. (1966) Turbulent heat transfer in the separated, reattached, and redevelop-
ment regions of a circular tube, J. Heat ¹ransfer 8, 131—136.
35. Krishnan, S., Philmon George and Shrotri, P. G. (1996) Solid fuel ramjet combustion chamber studies, In: 47th
Annual General Meeting, ¹he Aeronautical Society of India, IIT, Madras, India.
36. Launder, B. E. and Spalding, D. B. (1972) ¸ectures in Mathematical Models of ¹urbulence, Academic Press,
New York.
37. Lee, T. H. and Netzer, D. W. (1992) Temperature effect on solid fuel ramjet fuel properties and combustion,
J. Propulsion Power 8, 721—723.
38. Lips, H. R., Schmucker, R. H. and Witgracht, I. L. (1978) Experimental investigation of a solid fuel ramjet.
DFVLR-FB 78-27.
39. Locke, F. A. S. Jr. (1952) Recommended Definition of Turbulent Friction in Incompressible Fluids. Bureau of
Aeronautics Research Division, Rept. No. 1415.
40. Mady, C. J., Hickey, P. J. and Netzer, D. W. (1978) Combustion behaviors of solid-fuel ramjets, J. Spacecraft
Rockets 15, 131—132.
41. Mahoney, J. J. (1991) Inlets for Supersonic Missiles. AIAA Education Series, Washington, DC, pp. 185—187.
42. Marxman, G. A. and Gilbert, M. (1963) Turbulent boundary layer combustion in the hybrid rocket, In: Proc.
9th Int. Symp. on Combustion, The Combustion Institute, Pittsburg, PA, pp. 371—383.
43. Marxman, G. A. and Wooldridge, C. E. (1968) Research on the combustion mechanism of hybrid rockets,
In: Advances in ¹actical Rocket Propulsion. AGARD Conf. Proc. Vol. 1. (ed. S. S. Penner), pp. 485—522.
44. Marxman, G. A., Wooldridge, C. E. and Muzzy, R. J. (1964) Fundamentals of hybrid boundary layer
combustion, In: Heterogeneous Combustion. pp. 485—522, AIAA Progress in Aeronautics and Astronautics,
Vol. 15 (eds H. G. Wolfhard, I. Glassman and Green, Jr. L). Academic Press, New York.
Solid fuel ramjet combustor design 255

45. Meinkohn, D. and Bergmann, J. W. (1981) Experimental investigation of a hydrocarbon solid fuel ramjet.
In: Ramjets and Ramrockets for Military Applications, AGARD, CP-307, pp. 21.1—21.11.
46. Metochianakis, M. E. and Netzer, D. W. (1983) Modeling solid-fuel ramjet combustion including radiation to
the fuel surface, J. Spacecraft Rockets 20, 405—406.
47. Miles, E. R. C. (1958) Semi Empirical Formulae for Ogives. CM-505, Applied Physics Laboratory, Johns-
Hopkins University.
48. Milshtein, T. and Netzer, D. W. (1986) Three-dimensional, primitive-variable model for solid-fuel ramjet
combustion, J. Spacecraft Rockets 23, 113—117.
49. Mongia, H. C., Reynolds, R. S. and Srinivasan, R. (1984) Multidimensional turbulent combustion analysis,
applications and limitations. AIAA Paper No. 84-0477.
50. Myer, T. D. (1984) Special problems of ramjet with solid fuel. In Ramjet and Ramrocket Propulsion Systems for
Missiles. AGARD Lecture Series 136.
51. Nabity, J. A., Lee, T. H., Natan, B. and Netzer, D. W. (1993) Combustion behavior of boron carbide fuel in
solid fuel ramjets, In: Combustion of Boron Based Solid Propellants and Solid Fuels (eds K. K. Kuo and R. Pein),
Begell House and CRC Press, Boca Raton, FL, pp. 287—302.
52. Natan, B. and Gany, A. (1991) Ignition and combustion of boron particles in the flow field of a solid fuel
ramjet, J. Propulsion Power 7, 37—43.
53. Natan, B. and Gany, A. (1993) Effects of bypass air on boron combustion in solid fuel ramjets, J. Propulsion
Power 9, 155—157.
54. Natan, B. and Gany, A. (1993) Combustion characteristics of boron-fueled solid fuel ramjet with aft-burner,
J. Propulsion Power 9, 695—701.
55. Natan, B. and Netzer, D. W. (1993) Experimental investigation of the effect of bypass air on boron combustion
in a solid fuel ramjet, In: Combustion of Boron Based Solid Propellants Solid Fuels (eds K. K. Kuo and
R. Pein), Begell House and CRC Press, Boca Raton, FL, pp. 427—437.
56. Netzer, D. W. (1977) Modeling solid fuel ramjet combustion. J. Spacecraft Rockets 14, 762—766.
57. Netzer, D. W. (1978) Model applications to solid-fuel ramjet combustion, J. Spacecraft Rockets 15, 263—264.
58. Netzer, A. and Gany, A. (1991) Burning and flameholding characteristics of a miniature solid fuel ramjet
combustor, J. Propulsion Power 7, 357—363.
59. Nusca, M. J. (1990) Steady flow combustion model for solid fuel ramjet projectile, J. Propulsion Power 6,
348—352.
60. Nusca, M. J., Chakravarthy, S. R. and Goldberg, U. C. (1990) Computational fluid dynamics capability for
solid fuel ramjet projectile, J. Propulsion Power 6, 256—262.
61. Pein, R. and Vinnemeier, F. (1992) Swirl and fuel composition effects on boron combustion in solid-fuel
ramjets, J. Propulsion Power 8, 609—614.
62. Phaneuf, J. T. Jr. and Netzer, D. W. (1974) Flow characteristics in solid fuel ramjets. Naval Postgraduate
School, Monterery, California, Rept. NPS-57Nt 74081.
63. Philmon George, Krishnan, S., Lalitha, R., Varkey, P. M. and Raveendran, M. (1996) Regression rate study in
HTPB/GOX hybrid rocket motors, Defence Sci. J. 46, 337—345.
64. Pun, W. M. and Spalding, D. B. (1977) A General Computer Programme for Two-Dimensional Elliptic Flows.
Imperial College of Science and Technology. Rept. No. HTS/76/2.
65. Refinement on Inlet Design Continues. (1971) Aviation ¼eek and Space ¹echnology, 94, 62—64.
66. Rubesin, M. W., Maydew, R. C. and Varga, S. A. (1951) An Analytical and Experimental Investigation of the
Skin Friction of the Turbulent Boundary Layer on a Flat Plate at Supersonic speeds. NACA Tech. Note 2305.
67. Schulte, G. (1986) Fuel regression and flame stabilisation studies of solid-fuel ramjets, J. Propulsion Power 2,
301—304.
68. Schulte, G. and Pein, R. (1986) Regression rate study for a solid-fuel ramjet. In: Proc. 15th Congr. of the Int.
Council of the Aeronautical Sciences, pp. 1331—1336, London, UK.
69. Schulte, G., Pein, R. and Hogl, A. (1987) Temperature and concentration measurements in a solid fuel ramjet
combustion chamber, J. Propulsion Power 3, 114—120.
70. Scott II, C. K. (1986) An experimental investigation of various metallic/polymer fuels in a two-dimensional
solid fuel ramjet, Master’s thesis, Naval Postgraduate School, Monterey, California.
71. Simpson, J. A. (1983) Fight dynamics of the advanced fire system (AIFS)—cannon launched ramjet, In: Proc.
7th Int. Symp. on Ballistics. The Hague, The Netherlands.
72. Simpson, J. A., Krier, H. and Butler, P. B. (1983) Interior ballistics for launch dynamics for advanced indirect
fire system, In: Proc. 7th Int. Symp. on Ballistics. The Hague, The Netherlands.
73. Smoot, L. D. and Price, C. F. (1965) Regression rates of non-metallized hybrid fuel systems, AIAA J. 3,
1408—1413.
74. Smoot, L. D. and Price, C. F. (1966) Regression rates of metallized hybrid fuel systems, AIAA J. 4, 910—915.
75. Smoot, L. D. and Price, C. F. (1967) The pressure dependence of hybrid fuel regression rates, AIAA J. 5,
102—106.
76. Solid Fuel Ramjet. (1980) United Technologies Chemical Systems, San Jose, CA.
77. Stevenson, C. A. and Netzer, D. W. (1981) Primitive-variable model applications to solid-fuel ramjet combus-
tion, J. Spacecraft Rockets 18, 89—94.
78. Strand, L. D., Ray, R. L., Anderson, F. A. and Cohen, N. S. (1992) Hybrid fuel combustion and regression rate
study, AIAA Paper No. 92-3302.
79. Vaught, C., Witt, M., Netzer, D. W. and Gany, A. (1992) Investigation of solid fuel, dual-mode combustion
ramjets. J. Propulsion Power 8, 1004—1011.
80. Veraar, R. G. (1991) Ramjet applications of the solid fuel combustion chamber. In: Proc. Symp. on ¹he Solid
Fuel Combustion Chamber and Beyond, Rijswijk, The Netherlands, pp. 129—146.
81. Vinnemier, F. M. and de Wilde, J. P. (1990) Heat transfer in a solid fuel ramjet combustor, AIAA Paper No.
90-1783.
82. Wimmerstrom, P., Nilsson, Y. and Gunners, N. E. (1993) Initial study of a 40 mm SFRJ projectile. In: Proc.
14th Int. Symp. on Ballistics, Quebec, Canada.

Das könnte Ihnen auch gefallen