You are on page 1of 10

44th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit AIAA 2008-4544

21 - 23 July 2008, Hartford, CT

A Parametric Analysis of Hybrid Rocket Motors for


Sounding Rockets

Lorenzo Casalino∗ and Dario Pastrone†


Politecnico di Torino, 10129 Torino, Italy.
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

A hybrid rocket motor is considered as the propulsion system for a sounding rocket.
The propulsion system design and the trajectory are simultaneously optimized by means
of a nested direct/indirect procedure, with the aim of maximizing the time spent in mi-
crogravity conditions given the available mass budget: Direct optimization of the param-
eters that affect the motor design is coupled with indirect optimization of the trajec-
tory. Different propellant combinations (namely, hydrogen peroxide/polyethylene, liquid
oxygen/hydroxyl-terminated polybutadiene and nitrous oxide/hydroxyl-terminated polybu-
tadiene) are considered. Nitrous oxide is a self-pressurizing oxidizer, whereas a blowdown
gas pressurized feed system is considered for hydrogen peroxide and liquid oxygen; the ben-
efit of a partially regulated feed system is also investigated when hydrogen peroxide is the
oxidizer. A parametric analysis of the influence on performance and optimal engine design
of preliminary assumptions concerning mission requirements, propellant characteristics and
other design choices, not considered in the optimization procedure, is presented.

Nomenclature

Ab = burning surface area, m2


Ap = port area, m2
Ath = nozzle throat area, m2 Superscripts
a = regression constant, m1+2n kg−n sn−1 ˙ = time derivative
CF = thrust coefficient
c∗ = characteristic velocity, m/s
D = rocket diameter, m Subscripts
F = thrust magnitude, N
J = throat area to initial port area ratio 0 = ambient
L = overall length, m 1 = combustion chamber at head-end
Lb = grain length, m a = auxiliary gas
m = mass, kg avg = average
n = mass-flux exponent BD = beginning of blowdown phase
nx = longitudinal acceleration, g c = combustion chamber at nozzle entrance
p = pressure, bar e = nozzle exit
R = port radius, m F = fuel
t = time, s f = final value
tµg = microgravity time, s i = initial value
V = volume, m3 nozzle = nozzle
Z = hydraulic resistance, (kg m)−1 O = oxidizer
α = mixture ratio p = overall propellant (oxidizer + fuel)
γ = specific heat ratio res = residual gas or oxidizer
ε = nozzle area-ratio t = oxidizer tank
ρ = density, kg/m3

∗ Associate Professor, Dipartimento di Energetica, Corso Duca degli Abruzzi, 24 Torino. AIAA Senior Member.
† Associate Professor, Dipartimento di Energetica, Corso Duca degli Abruzzi, 24 Torino. AIAA Senior Member.

1 of 10

American
Copyright © 2008 by the American Institute of Aeronautics and Institute
Astronautics, Inc. All of Aeronautics
rights reserved. and Astronautics
I. Introduction
ybrid rocket motors (HRMs) present higher density-impulse, when compared to liquid bipropellant
H rocket engines, and higher specific impulse, when compared to solid rocket motors and liquid monopro-
pellant rockets. Moreover, they are safe and reliable, can be throttled within a wide thrust range and have
shutdown-restart capability. Last but not least, they are relatively low cost and have a reduced environmental
impact. All these features make HRMs suitable for relatively-simple low-cost applications, such as sounding
rockets to be used as microgravity platforms. A large number of papers concerning numerical and exper-
imental HRM investigation can be found in literature.1–6 In most numerical investigations, a parametric
optimization is performed in order to find the best grain and nozzle geometries as well as the best operating
conditions, such as mean mixture ratio and chamber pressure. The optimization of the motor design and
operations is however strictly related to the type of mission considered. This is especially true in the present
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

case, which considers a sounding rocket that performs a sub-orbital flight. Thrust has a major influence,
since it affects both the motor design and the trajectory performance; a compromise must be sought, as
greater thrusts reduce the gravitational losses while increasing drag losses and structural mass. For this
reason, the search for the optimal design should include trajectory analysis and, possibly, its optimization.
With this purpose, a hybrid direct/indirect optimization procedure7, 8 is employed to couple the optimization
of motor design with trajectory optimization.
The choice of the propellant combination has a major influence on a rocket design. Three propellant
combinations are widely used today: Hydrogen-peroxide (HP) / polyethylene (PE), liquid oxygen (LOX) /
hydroxyl-terminated polybutadiene (HTPB) and nitrous oxide (N2O) /HTPB. Nitrous oxide can be used
either as a cryogenic liquid propellant or as a storable liquefied gas propellant at ambient temperature.
In the latter case N2O shows modest performance due to its low density and high vapor pressure, thus
requiring heavy tanks. On the other hand, nitrous oxide presents some inherent attractive advantages,
such as being non-toxic, stable and comparatively unreactive at ordinary temperatures. In contrast with
hydrogen peroxide, nitrous oxide does not require catalytic decomposition. Moreover, the aforementioned
high vapor pressure at ambient temperature determines its self-pressurizing capability. This fact and the
safety characteristics of N2O explain the wide use of nitrous oxide as an oxidizer in HRMs (as an example,
a N2O/HTPB hybrid rocket was used for a manned suborbital flight to a 100-km altitude,9 and nitrous
oxide will be probably employed for future commercial space flight10 ). Nitrous oxide has also been tested
with paraffin fuels,11 which yield superior regression rates, thus permitting the simplest single-port grain
geometries, while preserving safety and non-toxicity12 (liquefying hybrid).
A gas-pressure feed system is adopted when HP or LOX are the oxidizer, because of its simplicity and lower
cost compared to pump-fed engines. In order to assess the benefit provided by the oxidizer mass flow control
while facing complexity, the blowdown feed system option is first considered and then compared, for the
HP/PE combination, to a partially regulated pressure system. Only self pressurization is instead adopted
for N2O. The propellant tank pressurization modeling plays a key role in predicting rocket performance.
When a self-pressurizing propellant such as N2O is used, the mass flux between the two phases becomes a
fundamental control mechanism of the tank pressure change. A simple two-phase model,8 which considers
superheated liquid and saturated vapor in the oxidizer tank and does not require any evaluation of the
liquid/vapor heat transfer rate, is adopted to describe the self-pressurization of nitrous oxide.
The optimization procedure provides the optimal engine design given the mission requirements. Addi-
tional preliminary assumptions, concerning propellant characteristics and other design choices, not considered
in the optimization procedure, must be adopted. The present paper provides a discussion about the changes
in design and performance which are experienced when mission requirements and preliminary assumptions
are modified.

II. Grain Geometry and Ballistic Model


A circular-section grain with a singular circular port is adopted, while considering a uniform regression
rate along the port axis. The regression rate, i.e., the time derivative of the port radius R, is determined by
the oxidizer mass flux and grain geometry

Ṙ = a (ṁO /Ap )n ∝ ṁnO R−2n (1)

with the values for the considered propellant combinations provided in Table 1.

2 of 10

American Institute of Aeronautics and Astronautics


Table 1. Values of a and n to be employed in Eq. (1) when SI units are used.

Propellants a n Source
HP/PE 7.00 10−6 0.800 Ref.13, 14
LOX/HTPB 9.24 10−6 0.852 Ref.15
N2O/HTPB 1.87 10−4 0.347 Ref.16

No pyrolysis of the lateral ends is considered. Pressure losses inside the combustion chamber are taken
into account by relating the chamber head-end pressure p1 to the chamber nozzle-stagnation pressure pc .
An approximate relation, similar to that proposed by Barrere et al.17 for side-burning grains, is used
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

"  2 #
Ath
p1 = 1 + 0.2 pc (2)
Ap

The hydraulic resistance Z in the oxidizer flow path from the tank to the combustion chamber determines
the oxidizer flow rate. Under the assumption of incompressible turbulent flow
p
ṁO = (pt − p1 )/Z (3)

The value of Z is assumed to be constant during motor operation. When N2O is used and the liquid phase
ends, gaseous propellant flows from the propellant tank into the combustion chamber. In this case the
oxidizer mass flow through the injectors is evaluated assuming a choked flow.8
The fuel mass flow is
ṁF = ρF ṘAb ∝ ṁnO R1−2n (4)
and the propellant mixture ratio is
ṁO 1−n
α= ∝ ṁO R2n−1 (5)
ṁF
An isentropic expansion in the nozzle is assumed, and the chamber nozzle-stagnation pressure pc is deter-
mined by
(ṁO + ṁF )c∗
pc = (6)
Ath
The performance of the propellant combination is evaluated18 as a function of the mixture ratio α,
assuming pc = 10 bar. Even though the actual pressure in the combustion chamber can span over a wide
range during motor operations, the error is small for chamber pressures and mixture ratios considered in
this paper. Frozen equilibrium expansion is assumed; the exhaust gas maintains throughout the nozzle the
composition that it has in the combustion chamber. This conservative assumption of frozen equilibrium
expansion is adopted to account for the low combustion efficiency of HRMs; in addition a 0.96 c∗ -efficiency19
is introduced. Third-degree polynomial curves fitting the characteristic velocity and specific heat ratio are
embedded in the code to compute the proper values as the mixture ratio changes during motor operations.

III. Motor Design and Operation


According to the chosen ballistic model, the design of the HRM is defined by the nozzle expansion ratio
ε, and the initial values of thrust level Fi , mixture ratio αi , tank pressure (pt )i , chamber pressure (pc )i , and
the ratio J of the throat area to the initial port area.
The initial chamber pressure is assigned by imposing (pc )i = 0.4 (pt )i ; actually, the ratio pt /pc varies
during operation, but the assumed initial ratio is usually sufficient to guarantee pt /pc > 1.5 and to avoid
coupling between the hybrid motor and the oxidizer feed system. The initial port area to throat area ratio
J should be as large as possible but not exceed 0.5 to avoid excessive pressure losses and nonuniform grain
regression: J = 0.5 is therefore assumed throughout. Additional parameters are to be given, depending on
the chosen oxidizer and feed system.
A simple blowdown (BD) pressurization system (option 1) is considered for the HP/PE and LOX/HTPB
combinations. In addition, a different option is adopted for the HP/PE combination only, to highlight the

3 of 10

American Institute of Aeronautics and Astronautics


benefit of a partially regulated feed system (option 2), with a phase at constant tank pressure, maintained
by means of helium flowing from an auxiliary tank, followed by a blowdown phase. When option 1 is chosen,
the gas is assumed to expand isentropically. The value of the initial ullage volume (Vg )i (or, equivalently,
the mass of the pressurizing gas) is the only additional parameter. On the other hand, when option 2 is
adopted, the initial ullage volume is assumed to be 3% of the oxidizer volume, in order to have a stable
regulator response when the out flow starts.20 In this case two additional parameters are the auxiliary gas
tank volume Va and the initial pressurizing gas pressure pa ; large values should be adopted for the latter and
pa = 200 bar is assumed throughout. The parameter Va is conveniently replaced by the exhausted oxidizer
mass at the beginning of the blowdown phase (mO )BD . When the tank pressure is kept constant pt = (pt )i ,
whereas pt is again calculated assuming an isentropic expansion of the pressurizing gas in the tank during
the subsequent blowdown phase. Calling (Vg )BD the gas volume in the propellant tank at the beginning of
the blowdown phase, one has
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

 γ
(Vg )BD
pt = (pt )i (7)
Vg
where Vg = (Vg )i + mO /ρO . Obviously (Vg )BD = (Vg )i when the simpler blowdown pressurization is chosen
(option 1), otherwise (Vg )BD = (Vg )i + (mO )BD /ρO for option 2.
For a self-pressurizing oxidizer, liquid and vapor are in equilibrium before motor operation and the tank
pressure is only related to the temperature. An initial value of 45 bar (which corresponds to a temperature of
287.2 K) is assumed for (pt )i . Self pressurization occurs until the liquid phase ends; an isentropic blowdown
phase follows, ruled by  γ
mOt
pt = (pt )BD (8)
(mOt )BD
In order to avoid poor combustion, the thruster is turned off when the pressure in the combustion chamber
reaches 1 bar. The tank volume is determined by the indirect procedure during the trajectory optimization
to obtain the prescribed pressure at burnout; a tentative starting value must be given to begin the procedure.
The initial ullage volume is assumed to be 3% of the oxidizer tank volume.
A two-phase lumped model,8 which neglects heat transfer between the tank walls and the oxidizer, is
adopted to evaluate tank pressure during oxidizer outflow. A saturated vapor and a warmer liquid phase
are considered; a pressure/temperature relation laying between the saturation line and the liquid spinodal
line is prescribed for the liquid phase. Liquid and vapor are considered to be in equilibrium before motor
operation. During operation their temperatures are assumed to be uniform (i.e. no stratification effects
are taken into account) albeit distinct. Condensation and evaporation/boiling occur at the vapor-liquid
interface. The general form of energy conservation for a variable control volume enclosing an unsteady open
process is applied, both kinetic and potential energy being neglected. It is assumed that the condensing
vapor leaves the ullage volume as saturated liquid, and that the evaporating liquid enters the ullage volume
with saturated vapor enthalpy at the temperature of the liquid phase.
In the adopted model, condensation occurs and forces the vapor phase to follow the saturation line. No
liquid evaporation/boiling is considered until the liquid reaches a limiting pressure, which is assumed to lie
between the saturation line and the spinodal line. After reaching this limit, evaporation/boiling takes place
and the liquid is constrained to follow the limiting curve. Under these assumptions, one can derive8 a set of
differential equations that provide the time derivatives of tank pressure, liquid and vapor temperature and
mass transfer rates between the two phases, while not needing a liquid/vapor heat transfer rate evaluation.
Given the set of design parameters, the motor geometry is first determined. The relevant properties of
the combustion gases can be computed, owing to the fact that the initial values of c∗ and γ can be calculated
from αi via the aforementioned curve fittings. The thrust coefficient CF can then be evaluated by assuming
an isentropic one-dimensional expansion with constant γ, provided the ambient pressure is known. With a
0.98 correction factor introduced to modify the vacuum thrust coefficient, one has
v 
 u
u 2γ 2
   γ + 1 
  γ  

2 γ−1  pe γ − 1  pe  p0
u
CF = 0.98 t 1− +ε −ε (9)

 γ − 1 γ + 1 p c p c
 pc
 

The mass flow rates at rocket ignition (i.e., at t = 0) are found from the initial thrust Fi
1 + αi Fi
(ṁp )i = (1 + αi )(ṁF )i = (ṁO )i = ∗ (10)
αi ci (CF )i

4 of 10

American Institute of Aeronautics and Astronautics


The throat, initial port, and initial burning areas Ath , (Ap )i and (Ab )i are then determined

(ṁp )i Ath (Ap )ni (ṁF )i


Ath = (Ap )i = = πRi2 (Ab )i = = 2πRi Lb (11)
(pc )i c∗i J aρF (ṁO )ni

thus specifying the grain geometry. The nozzle throat area Ath is considered to be constant during operation.
The head-end pressure is computed with Eq. (2) and, knowing the initial tank pressure, also the hydraulic
resistance Z can be determined by applying Eq. (3) at t = 0. Once the initial ullage volume in the propellant
tank has been assumed and the pressurization system has been specified, the motor performance can be
evaluated during operation.
The tank pressure rules motor operation; either pt = (pt )i or pt is provided by Eq. (7), when a pressurizing
gas is used. Numerical integration of Eqs. (1), (3), and (4), allows the evaluation of the fuel grain geometry,
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

the exhausted masses of oxidizer and fuel, and their mass flow rates. The equations concerning the behavior
of the self-pressurizing oxidizer8 are also integrated to determine the tank pressure when N2O is used as the
oxidizer. At each instant t, once the tank pressure pt and the motor geometry are known, the regression rate,
the propellant flow rates (and their ratio α), c∗ , pc and p1 are computed by numerically solving Eqs. (1)-(6)
while the curve fits for c∗ as a function of α are used. Then, the thrust level F = pc Ath CF is determined by
evaluating CF at the actual altitude via Eq. (9), in order to integrate the trajectory equations. At burnout
the overall propellant is finally evaluated, and an estimation of the structural masses can be obtained. The
initial ullage volume is (Vg )i for the blowdown case; when the regulated feed system or the self-pressurizing
oxidizer are employed, the initial ullage volume is proportional to the oxidizer volume. A tentative value
is assumed for the overall oxidizer mass and is corrected by the indirect method during the trajectory
optimization, in order to match the value obtained by integrating Eq. (3) up to the orbit insertion.
The optimization procedure aims at finding the motor design parameters and the corresponding trajectory
that maximize the mission performance index, which is, in this paper, the time spent above a 100-km altitude,
which is conventionally considered to correspond to microgravity conditions (it has been verified that tµg
differs by less that 5 seconds from the time during which the rocket actually experiences a state of free-fall,
with drag acceleration lower than 1 µg). A mixed optimization procedure7 is here adopted. An indirect
method21 optimizes the trajectory for each choice of the motor parameters. These are instead optimized by
means of a direct procedure.22 Both methods have been developed at the Politecnico di Torino. The details
of the optimization of a hybrid sounding rocket can be found in Ref. 8.

IV. Parametric Analysis


The nominal case is defined by the baseline problem considered in Ref. 8. The trajectory lies on the
equatorial plane. The rocket initial mass is 500 kg, with 100 kg reserved for payload and other masses, which
are not related to propulsion system characteristics (e.g., avionics). The remaining 400 kg mass budget has
to be shared between propellants, combustion chamber, nozzle, tanks, and rocket casing. The combustion
chamber has a 6-mm insulating liner (with density equal to that of the solid fuel) and an aluminum alloy
cylindrical wall. The cylindrical propellant tank has a diameter which is determined by the optimization
procedure, which seeks a compromise between drag loss reduction and minimal structural mass, for fixed
rocket aspect ratio. The wall thickness of the propellant tank is determined assuming a 1.25 safety factor.
The rocket diameter is a consequence of the propellant tank diameter; a 1-mm aluminum case is assumed.
A rocket length (except the payload) to outer diameter ratio of 12 is imposed. A 45-deg convergent and a
20-deg divergent nozzle with a phenolic silica ablative layer is considered. A uniform thickness is assumed
and is evaluated according to Ref. 24, using average values of the transport properties and heat flux. The
nozzle mass is computed accordingly; the structural mass is small compared to the ablative layer mass and
is thus neglected. A constant 0.2 drag coefficient is considered, with a reference area equal to the rocket
cross section. The most significant results of the nominal case for the considered propellant combinations
and feed systems are shown in Tables 2, 3, 4, 5, and are compared to the results which are obtained when
the parameters and data that define the mission are changed.
The mixture ratio and thrust histories for the nominal cases are shown in Fig. 1. A large thrust decrease
occurs when a blowdown pressurization system is adopted, partially mitigated by the constant pressure phase
of the regulated feed system; a more uniform thrust profile is instead obtained with the self pressurizing N2O.
In contrast, N2O exhibits a large mixture ratio decrease, whereas α is almost constant in the other cases.
This is related to the different value of the mass-flux exponent n, which makes ṁF increase with R when

5 of 10

American Institute of Aeronautics and Astronautics


Table 2. Optimal design and performance of the N2O/HTPB rocket.

Case Fi αi αavg ε (pt )i mp mres tµg D L Ri Rf


kN bar kg kg s m m m m
Nominal 16.2 10.92 6.72 4.14 45.0 340 8.7 177 0.40 4.75 0.065 0.131
mi = 1000 kg 33.6 10.68 6.80 4.17 45.0 686 17.9 209 0.51 6.10 0.094 0.162
mu /mi = 0.1 15.7 11.14 6.76 4.65 45.0 383 9.7 366 0.41 4.91 0.065 0.138
L/D = 9 14.9 11.12 6.77 4.13 45.0 340 8.6 148 0.45 4.02 0.063 0.132
CD = 0.4 13.4 11.08 6.62 4.12 45.0 341 8.5 53 0.39 4.69 0.060 0.134
a = 2.06 10−4 16.5 10.91 6.68 4.14 45.0 340 8.6 181 0.39 4.71 0.066 0.136
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

Table 3. Optimal design and performance of the LOX/HTPB rocket with blowdown pressurization.

Case Fi αi αavg ε (pt )i mp mres tµg D L Ri Rf


kN bar kg kg s m m m m
Nominal 23.5 3.20 3.52 4.53 59.9 328 0.9 219 0.46 5.52 0.067 0.111
mi = 1000 kg 45.2 3.32 3.53 4.60 57.9 664 2.0 253 0.60 7.19 0.094 0.144
mu /mi = 0.1 21.5 3.26 3.56 4.45 48.2 376 0.9 411 0.48 5.75 0.072 0.118
L/D = 9 21.8 3.24 3.62 4.45 60.2 327 0.9 169 0.54 4.83 0.064 0.111
CD = 0.4 21.4 3.15 3.55 4.52 62.6 328 0.9 124 0.45 5.42 0.064 0.110
a = 10.16 10−6 24.1 3.12 3.45 4.47 58.8 329 0.9 231 0.45 5.43 0.068 0.115

n < 0.5 (N2O/HTPB) and decrease with R when n > 0.5 (LOX/HTPB and HP/PE), as shown by Eq. (4).
In the latter case, the effect of the radius increase adds to the effect produced on the fuel flow rate by the
decrease of ṁO , whereas it contrasts it when N2O is employed. The opposite hold for the mixture ratio,
ruled by Eq. (5).
First, a larger rocket initial mass is considered, while the payload fraction is kept constant at 0.2, in
order to highlight the effects of the rocket scale. In particular, an initial mass of 1000 kg (and a payload
of 200 kg) is assumed. The larger rocket exhibits better performance, with a microgravity time increase of
approximately 30 s, for all the propellant combinations. A liquid propellant rocket could be scaled in a simple
way: If the initial mass is doubled, the rocket would provide the exactly same performance if the propellant
masses, mass flow rates, throat and exit areas are also doubled, and only the rocket aspect ratio L/D would

14 25
N2O/HTPB
12 LOX/HTPB
HP/PE blowdown 20
HP/PE regulated
10
Mixture ratio α

Thrust F, kN

15
8

6 10

4 5
2
0
0 20 40 60 80 0 20 40 60 80

Time t, s Time t, s
(a) Mixture ratio. (b) Thrust.

Figure 1. Mixture ratio and thrust histories (nominal cases).

6 of 10

American Institute of Aeronautics and Astronautics


Table 4. Optimal design and performance of the HP/PE rocket with blowdown pressurization.

Case Fi αi αavg ε (pt )i mp mres tµg D L Ri Rf


kN bar kg kg s m m m m
Nominal 24.8 8.58 8.30 4.99 60.1 339 0.9 299 0.42 5.01 0.068 0.101
mi = 1000 kg 48.2 8.76 8.30 5.04 58.4 684 1.9 326 0.54 6.52 0.097 0.132
mu /mi = 0.1 22.6 8.81 8.48 4.82 47.6 388 0.8 512 0.43 5.21 0.075 0.108
L/D = 9 23.1 8.82 8.69 4.93 60.0 339 0.9 269 0.48 4.31 0.066 0.100
CD = 0.4 22.5 8.58 8.42 5.02 62.2 339 0.9 231 0.41 4.92 0.064 0.099
a = 7.70 10−6 25.4 8.39 8.16 4.92 59.3 340 0.9 307 0.41 4.95 0.070 0.104
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

Table 5. Optimal design and performance of the HP/PE rocket with regulated pressurization.

Case Fi αi αavg ε (pt )i mp mres tµg D L Ri Rf


kN bar kg kg s m m m m
Nominal 16.4 8.02 8.24 4.66 38.3 348 1.1 328 0.41 4.91 0.072 0.104
mi = 1000 kg 31.4 8.12 8.21 4.67 36.8 703 2.2 358 0.53 6.38 0.103 0.136
mu /mi = 0.1 14.7 8.21 8.41 4.51 30.7 398 0.9 558 0.42 5.08 0.079 0.111
L/D = 9 15.0 8.26 8.61 4.64 38.0 350 1.1 303 0.48 4.30 0.070 0.103
pa = 250 bar 16.5 7.93 8.21 4.84 39.9 348 1.1 336 0.42 4.93 0.071 0.103
CD = 0.4 14.3 8.03 8.42 4.67 38.6 349 1.0 260 0.40 4.82 0.067 0.102
a = 7.70 10−6 16.7 7.85 8.12 4.62 15.1 348 1.0 336 0.40 4.84 0.074 0.106


be reduced by a 2 factor. If the aspect ratio is held constant, the performance benefits from the reduced
drag-to-mass ratio. This simple scaling law cannot be adopted in a hybrid rocket, because the port area (here
assumed to be proportional to the throat area) influences the fuel flow rate. The relative increase of port
radius during operation is lower for the larger rocket, and the mixture ratio shifting results to be enhanced for
the LOX/HTPB and HP/PE combinations (see Fig. 2 for HP/PE with a blowdown pressurization system)
whereas it is slightly reduced for the N2O/HTPB combination (Fig. 3). These phenomena are taken into
account by the optimal design, and the initial mixture ratio is increased for the propellant combinations
with enhanced shifting and decreased for the N2O/HTPB combination. On the other hand, the thrust

9.5 8
Longitudinal acceleration nx , g

nominal
mi = 1000 kg
mu / mi = 0.1
L/D = 9
6
9.0
Mixture ratio α

CD = 0.4
-6
a = 7.70 ⋅ 10

4
8.5

8.0
0 30 60 90 0 30 60 90

Time t, s Time t, s
(a) Mixture ratio. (b) Thrust.

Figure 2. Mixture ratio and longitudinal acceleration histories (HP/PE blowdown).

7 of 10

American Institute of Aeronautics and Astronautics


13 8
nominal

Longitudinal acceleration nx , g
mi = 1000 kg
11
mu / mi = 0.1
6
L/D = 9
9
Mixture ratio α

CD = 0.4
-4
a = 2.06 ⋅ 10
7 4

5
2
3
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

1 0
0 30 60 90 0 30 60 90

Time t, s Time t, s
(a) Mixture ratio. (b) Thrust.

Figure 3. Mixture ratio and longitudinal acceleration histories (N2O/HTPB).

changes significantly during the operation of HP/PE and LOX/HTPB rockets, but, for the larger rockets,
the thrust decrease is reduced by the lower decrease of the fuel mass flow rate. The optimal initial thrust
of the larger rocket is less than twice that of the smaller rocket, but the average acceleration results to be
larger, with a reduction of the gravitational losses. The initial thrust must instead be more than doubled
when the N2O/HTPB combination is employed to guarantee the increased average acceleration, because of
the more pronounced reduction of the fuel mass flow rate. Also, the initial tank pressure can be reduced for
the gas-pressurized propellant combinations, in order to diminish the structural mass, which also benefits
from the less-than-proportional increase of liner and rocket casing masses (whose thicknesses are fixed) in
advantage of the propellant mass. Finally, the rocket cross section and the related drag are less than doubled,
with additional benefit on the performance. Also, the web thickness Rf − Ri and the burning times do not
exhibit relevant changes.
As expected, a payload reduction improves the microgravity time, since more propellant can be loaded.
The more relevant change in the design parameters occurs for the initial tank pressure. The lower pressure
allows to store a larger amount of oxidizer in a tank which has roughly the same mass of the nominal case,
even though a larger shifting in mixture ratio and thrust occurs during operation. As a matter of fact, it
results to be more convenient the containment of the structural mass. Minor adjustments affect the other
design parameters, mainly to assure the correct average values of mixture ratio and acceleration, which are
slightly larger compared to the nominal case; the initial values are therefore farther from the average value,
also to account for the larger shifting, as more propellant is consumed. Note that the rocket dimensions
grow, to accommodate the larger propellant mass (the payload is here excluded in the compute of the rocket
length). The peak acceleration is quite large due to the low final mass. This fact may require the introduction
of a constraint on its maximum value and/or on the peak thermal flux and dynamic pressure during ascent.
The discussion of these issues is left for future work.
The optimal value of the rocket aspect ratio is dictated by the opposite requirements of low drag against
low structural mass and uniform thrust and mixture ratio; large values of L/D are required for optimal
performance and the value is constrained to 12 in the nominal case. A decrease of the rocket aspect ratio
penalizes the rocket performance, as the influence of the larger drag prevails. Lower values of initial thrust
are adopted to reduce the rocket acceleration and the drag losses, while accepting larger gravitational losses.
The tank pressure is slightly increased to reduce the tank volume and limit the rocket cross-section. One
should note that the blowdown cases experience larger penalties, in terms of microgravity time, compared
to the self-pressurized and partially regulated cases. Also, larger values of mixture ratio (in particular for
the HP/PE combination) are employed.
An increase of the initial pressure pa in the auxiliary tank, which is required for the pressurizing gas in
the case of the regulated feed system, slightly improves the performance. The pressurizing gas is now stored

8 of 10

American Institute of Aeronautics and Astronautics


in a more compact tank, and the oxidizer tank dimension and weight are reduced. This allows for a shorter
tank, diminishing the case mass and allowing the use of a nozzle with a larger expansion area ratio, thus
improving the propellant specific impulse (the nozzle is under-expanded for most part of the trajectory) and
the microgravity time.
Besides the design choices, the rocket performance depend on some parameters that may be known
with uncertainty, at least during preliminary analysis, such as the drag coefficient or the coefficient of the
combustion ballistic. The influence if these parameter has also been investigated. The increase of the drag
coefficient penalizes the performance; the consequences on the optimal design are similar to those produced
by a reduction of the rocket aspect ratio, with lower values of initial thrust and larger values of initial tank
pressure to be preferred. The regression rate is also difficult to estimate, and the results for regression
constant values increased by 10% with respect to the nominal cases have also been obtained. The rocket
perrormance is improved by average values of the mixture ratio closer to the values which correspond to the
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

optimal specific impulse. The increased regression constant allows for an improved grain design, and a larger
fuel mass, while also reduces the grain and the rocket length.

V. Conclusions
An optimization procedure, which couples direct and indirect optimization methods has been applied to
the optimization of the design of a hybrid sounding rocket. The optimization of the parameters that affect
the design of the motor is coupled with the trajectory optimization. The coupling is fundamental, as there
is a mutual dependence of the mission requirements and engine optimal characteristics. The optimization
method shows to be fast and reliable and provide motor design and the corresponding trajectory to maximize
the microgravity time of the sounding rocket. Few minutes are sufficient on a standard PC.
A sensitivity analysis concerning the influence on design and performance of mission requirements and
preliminary assumptions not involved in the optimization process shows that the optimization method is fast
and robust and can be applied to different cases. The results show that, as far as propellant and feed system
comparison are concerned, the influence of the preliminary assumptions is quite limited and does not modify
in depth the conclusions that can be derived. On the other hand, the design changes are significant in terms
of performance optimization, in order to adjust the rocket features to the particular mission requirements.

References
1 Dyer, J., Doran, E., Dunn, Z., Lohner, K., Bayart, C., Sadhwani, A., Zilliac, G., Cantwel, B., and Karabeyoglu, M.A.,

“Design and Development of a 100km Nitrous Oxide/Paraffin Hybrid Rocket Vehicle,” Paper AIAA 2007-5362, July 2007.
2 Kwon, S.T., Park, B.K., Lee, C., and Lee, J.,“Optimal Design of Hybrid Motor for the First Stage of Air Launch Vehicle,”

Paper AIAA 2003-4749, July 2003.


3 Sackheim, R., Ryan, R., and Threet, E., “Survey of Advanced Booster Option for Potential Shuttle-Derivative Vehicles,”

Paper AIAA 2001-3414, July 2001.


4 Schoonover, P.L., Crossley, W.A., and Heister, S.D., “ Application of a Genetic Algorithm to the Optimization of Hybrid

Rockets,” Journal of Spacecraft and Rockets, Vol. 37, No. 5, 2000, pp. 622-629.
5 Vonderwell, D.J., Murray, I.F., and Heister, S.D., “Optimization of Hybrid-Rocket-Booster Fuel-Grain Design,” Journal

of Spacecraft and Rockets, Vol. 32, No. 6, 1995, pp. 964-969.


6 Jansen, D.P.F.L., and Kletzkine, Ph., “Preliminary Design for a 3kN Hybrid Propellant Engine,” ESA Journal, Vol. 12,

No. 4, 1988, pp. 421-439.


7 Casalino, L., and Pastrone, D., “Optimal Design and Control of Hybrid Rockets for Access to Space,” Paper AIAA

2005-3547, July 2005.


8 Casalino, L., and Pastrone, D.,“Optimal Design of Hybrid Rocket Motors for Microgravity Platform,”Journal of Propul-

sion and Power, Vol. 24, No. 3, 2008, pp. 491-498.


9 Dornheim, M.A., “Reaching 100 km,” Aviation Week & Space Technology, Vol. 161, No. 6, Aug. 2004.
10 Benson, J., “Safe and Affordable Human Access to LEO,” Paper AIAA 2005-6758, Sep. 2005.
11 Van Pelt, D., Hopkins, J., Skinner, M., Buchanan, A., Gulman, R., Chan, H., Karabeyoglu, M.A., and Cantwell, B.,

“Overview of a 4-Inch OD Paraffin-Based Hybrid Sounding Rocket Program,” Paper AIAA-2004-3822, July 2004.
12 Karabeyoglu, M.A., Altman, D., and Cantwell, B. J., “Combustion of Liquefying Hybrid Propellants: Part 1, General

Theory,” Journal of Propulsion and Power, Vol. 18, No. 3, 2002, pp. 610-620.
13 Maisonneuve, Y., Godon, J.C., Lecourt, R., Lengelle, G., and Pillet, N., “Hybrid Propulsion for Small Satellites: Design

Logic and Test,” Combustion of Energetic Materials, Begell House, New York, 2002, pp. 90-100.
14 Wernimont, E. H., and Heister, S. D., “Combustion Experiments in Hydrogen Peroxide/Polyethylene Hybrid with

Catalitic Ignition,” Journal of Propulsion and Power, Vol. 16, No. 2, 2000, pp. 218-326.
15 Rocker, M.,“Modeling of Nonacoustic Combustion Instability in Simulations of Hybrid Motor Tests,”NASA/TP-000-

209905, Feb. 2000.

9 of 10

American Institute of Aeronautics and Astronautics


16 Doran, E., Dyer, J., Lohner, K., Dunn, Z., Cantwell, B., and Zilliac, G.,“Nitrous Oxide Hybrid Rocket Motor Fuel

Regression Rate Characterization,”Paper AIAA 2007-5352, July 2007.


17 Barrere, M., Jaumotte, A., De Veubeke, B. F., and Vandenkerckhove, J., Rocket Propulsion, Elsevier Publishing Company,

1960, pp. 251-256.


18 Mc Bride, B.J., Reno, M.A., and Gordon, S., “CET93 and CETPC: An Interim Updated Version of the NASA Lewis

Computer Program for Calculating Complex Chemical Equilibria With Applications,” NASA TM-4557, March 1994.
19 Sutton, G. P., and Biblarz, O., Rocket Propulsion Elements, John Wiley & Sons, 7th edition, 2001, pp. 64.
20 Brown, C. D., Spacecraft Propulsion, AIAA Education Series, AIAA, 1996, pag. 82.
21 Casalino, L., Colasurdo, G., and Pastrone, D., “Optimal Low-Thrust Escape Trajectories Using Gravity Assist,” Journal

of Guidance, Control, and Dynamics, Vol. 22, No. 5, 1999, pp. 637-642.
22 Casalino, L., and Pastrone, D.,“Oxidizer Control and Optimal Design of Hybrid Rockets for Small Satellites,”Journal of

Propulsion and Power, Vol. 21, No. 2, 2005, pp. 230-238.


23 Colasurdo, G., and Pastrone, D., “Indirect Optimization Method for Impulsive Transfer,” Paper AIAA 94-3762, Aug.

1994.
Downloaded by Vikram Sarbhai Space Centre LIbrary Thiruvanathapuram on February 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2008-4544

24 Barker, D.H., Kording, J.W., Belnap, R.D., and Hall, A.F., “A Simplified Method of Predicting Char Formation in

Ablating Rocket Exit Cones,” AIChE Chemical Engineering Progress Symposium Series, Vol. 61, No. 59, 1965, pp. 108-114.

10 of 10

American Institute of Aeronautics and Astronautics

You might also like