Vous êtes sur la page 1sur 19

P1: GTV/MBQ

P2: GRB Final Pages

Qu: 00, 00, 00, 00

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

Rocket Motors, Hybrid


David Altman
Space Propulsion Group, Inc., and Stanford University

I.
II.
III.
IV.
V.

Introduction
Burning Rate
Motor Ballistics
Propulsion System
Propellant Combinations

GLOSSARY
Blocking parameter A measure of the reduction in heat
transfer due to blowing; represented by CH /CHO or
Cf /Cfo .
Blowing coefficient Ratio of gas flow from a surface to
the axial flow; denoted B.
Entrainment The injection of liquid droplets into a gas
stream.
Fuel mass flow The rate at which fuel is gasified from a
unit area of surface; denoted m f .
Heat of gasification Total heat required to gasify a solid
from its ambient temperature; denoted h v .
Heat transfer Rate at which heat is transferred from the
flame to a surface; denoted Q.
Mass flux Mass rate of flow of a fluid per unit area of
cross section in the port; denoted G.
Mixture ratio Ratio of oxidizer to fuel burned; denoted
O/F.
Regression rate Linear rate at which a solid fuel regresses due to gasification; denoted r .
Stanton number Coefficient that relates the heat

transfer rate to the heat of combustion; denoted


CH .

A HYBRID ROCKET uses both a liquid and a solid as


propellants. In the typical classical hybrid, the fuel is a
solid and the oxidizer is a liquid (see Fig. 1). However, it
is possible to use solid oxidizers with liquid fuels; such
combinations have been called reverse hybrids. Unless
otherwise specified, the classical hybrid with the solid fuel
grain will be assumed.
Although many components are common to liquid and
solid rockets, the operation of a hybrid is distinctly different. In the solid rocket, the oxidizer and fuel are intimately mixed in the single solid phase and combustion
occurs when the exposed surface is heated by the combustion flame to the ignition temperature. In the liquid
rocket, both oxidizer and fuel are intimately mixed in the
vicinity of the injector to form a combustible mixture. In
both cases, therefore, everywhere in the combustion chamber there is a uniform mixture of both oxidizer and fuel.
The hybrid, however, burns as a macroscopic diffusion

303

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

304

Rocket Motors, Hybrid

FIGURE 1 Schematic of a hybrid rocket.

flame where the oxidizer-to-fuel ratio (O/F) continually


decreases down the length of the chamber.

I. INTRODUCTION
A. Why Hybrid?
The fundamental difference between a hybrid and the liquid or solid rocket leads to a number of distinguishing
characteristics. The advantages are:
Safety. The fuel is inert and can be manufactured,
transported, and handled safely in accordance with standard commercial practice. The system is nonexplosive
since an intimate mixture of oxidizer and fuel is not
possible.
Simplified throttling and shutdown. The engine can
be throttled by modulating the liquid flow rate, which is
simpler than in a liquid rocket, where two flow rates must
be synchronized while being modulated. In the hybrid
rocket, the fuel flow rate, which results from vaporization from the solid surface, automatically adjusts to the
change in oxidizer flow rate. Consequently, thrust termination is simply accomplished by turning off the liquid
flow rate, a feature of significance in an abort procedure.
Grain robustness. Unlike solid rockets, fuel grain
cracks are not catastrophic because burning only occurs
down the port where it encounters the oxidizer flow.
Propellant versatility. The selection of propellants is
much greater than with either liquids or solids. In contrast to liquids, solid constituents such as dense, energetic

metals can be added to enhance both performance and


density without resorting to slurries. In contrast to solids,
the oxidizers provide much higher energy levels.
Temperature sensitivity. Since the temperature effect
on burn rate is negligible (as in liquids), no margin need
be applied to the thrust chamber weight to account for a
high ambient launch temperature, which would increase
the maximum expected operating pressure (MEOP).
Low cost. The total operational cost for hybrid systems benefits greatly from the safety features and inert
propellant. Manufacture of the fuel can be done in a commercial facility that does not require the large acreage
and many buildings for solid propellant manufacture. As
a consequence, the fuel plant can be located at or near the
launch site. Furthermore, the system can tolerate larger
design margins, resulting in a lower fabrication cost.
These advantages, however, are not enjoyed without
some disadvantages, such as:
Low regression rate. The small resulting fuel web
means that most combustion chambers over 1 ft in diameter require multiple ports. But this characteristic is desirable for long-duration applications such as target drones,
hovering vehicles, and gas generators.
Low bulk density. A consequence of the low regression rate is that a large grain surface area must be provided
to supply the required thrust. This is generally done with
multiple ports that lead to a relatively low volumetric fuel
loading and results in a moderate fuel sliver loss. However,
recent efforts to increase the regression rate have reduced
this disadvantage.

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

Rocket Motors, Hybrid

Combustion efficiency. The nature of the large diffusion flame results in a lower degree of mixing and hence
lower impulse efficiency. This loss is generally 12% more
than in either liquids or solids. In comparison with solids,
however, the delivered specific impulse Isp is greater.
O/F shift. The opening of the port during burning
causes an O/F shift with burning time, which can lower
theoretical performance. With proper design of the initial
O/F, however, this loss is usually minimal.
Slower transients. The ignition transient is slower,
as is the thrust response to throttling. In most practical
applications, where reproducibility is more important than
speed of response, this aspect is not significant.
B. History
Despite the fact that hybrid rockets have not enjoyed the
same extensive development background as solid and liquid motors, hybrid combustion involving a solid and a
fluid has been natures way of burning fuels and oxidizers. Examples are (1) a wax candle or oil lamp burning in
the presence of atmospheric oxygen, with the wick being
the flame holding device, (2) a fireplace, where the bellows serves as a means of increasing the oxidizer mass
flux and therefore the burning rate, and (3) on a grander
scale, a forest fire involving the turbulent mixing of air
and the vaporized fuel exuding from the trees. Here again,
the augmenting effect of wind velocity is well known.
Because of its nonexplosive character, safety of operation, and low cost, the hybrid rocket has been a favorite
of amateur rocketeers. The earliest work on hybrid rockets was conducted in the late 1930s in Germany at I. G.
Farben and in the United States at the California Rocket
Society. Leonid Andrussow, in conjunction with O. Lutz
and W. Noeggerath, tested a 10-kN hybrid in 1937 using coal and gaseous nitrous oxide (laughing gas). During
the same period, Oberth, in Germany, did some work on
the more energetic LOXgraphite propellant combination.
Neither of these last two efforts was successful because
the very high heat of sublimation of carbon results in a
negligible burning rate.
In the early 1940s, a more successful effort was conducted by the California Pacific Rocket Society, employing LOX and several fuels such as wood and rubber. Of
these combinations, the LOXrubber combination was the
most successful and a rocket using these propellants was
flown in June 1951 to an altitude of about 9 km. Although
the Society did not report any ballistic analyses, they did
have an accurate concept of the fundamentals of hybrid
burning as evidenced by the following statement: The
chamber pressure of a solidliquid rocket engine is proportional to the oxidizer flow and not to the internal surface
area exposed to the flame. Thus, there is no danger of ex-

16:56

305
plosions due to cracks and fissures in the charge as with
solid propellant rockets.
In the mid-1950s, two significant hybrid efforts occurred. One was by G. Moore and K. Berman at General
Electric, involving the use of 90% hydrogen peroxide and
polyethylene in a rod and tube grain design. The combustion was very smooth, resulting in a high combustion
efficiency. The authors drew several very significant conclusions: (1) the longitudinal uniformity of burning was
remarkable, (2) grain cracks had no effect on combustion,
(3) hard starts were never observed, (4) combustion was
stable since the fuel surface acted as its own flame holder,
and (5) throttling was easily accomplished by a single
valve. The authors observed, however, that the burning
rate was low and could not be varied significantly. The
second significant effort was by William Avery at the Applied Physics Laboratory, who investigated a reverse hybrid composed of a liquid fuel (JP) and a solid oxidizer
(ammonium nitrate). The primary motivation for this propellant selection was low cost. Technically the program
was not successful because of the rough combustion and
poor performance.
During the 1960s, two European countries engaged in
hybrid studies leading to flight tests of sounding rockets. These organizations were ONERA (in conjunction
with SNECMA and SEP) in France and Volvo-Flygmotor
in Sweden. The ONERA development used a hypergolic
propellant based on nitric acid and an amine fuel. The
first flight of this vehicle occurred in April 1964, followed
by three flights in June 1965 and four flights in 1967. All
eight flights were successful, reaching altitudes of 100 km.
The Volvo-Flygmotor rocket was based on a hypergolic
combination using nitric acid and Tagaform (polybutadiene plus an aromatic amine). It was flown successfully in
1969 to an altitude of 80 km carrying a 20-kg payload.
United Technologies Center (Chemical Systems Division [CSD] of United Technologies Corp.) and Beech Aircraft developed a high-altitude supersonic target drone in
the late 1960s called the Sandpiper, using MON-25 (25%
NO, 75% N2 O4 ) and polymethyl methacrylate (PMM)Mg as the fuel. The first of six flights occurred in January
1968 and these rockets flew for over 300 sec and with a
range in excess of 100 miles. The HAST, a second version,
carried a heavier payload and was based on an IRFNAPB/PMM propellant combination. This 13-in.-diameter
motor was throttleable over a 10/1 range. A later version
of this vehicle, the Firebolt, was developed by Chemical Systems Division (CSD) and Teledyne Aircraft, using
the same propulsion configuration as the HAST. These
three programs were successfully conducted over a 15year period until the mid-1980s. These target drone rockets
were the only hybrid propulsion systems built to military
specifications.

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

306

FIGURE 2 Hybrid motor firing with FLOX oxidizer (F2 + O2 ).

An investigation of the use of high-energy hybrid propellants based on a lithium-containing fuel and FLOX
(F2 + O2 ) as the oxidizer was conducted by CSD in the
mid-1960s. This led to a hypergolic propellant system that
was throttleable and demonstrated a vacuum Isp of 380 sec
at 93% Isp efficiency. A firing of this 42-in.-diameter motor in 1970 is shown in Fig. 2, which is taken from the
cover of the January 1970 issue of Aviation Week.
The largest hybrid rockets built to date had a thrust
level of 250,000 lb and used LOX/HTPB propellant. They
were made by American Rocket Company (AMROC) and
were designed for use in a space vehicle and a high altitude
rocket (HYTOP). Another hybrid rocket at about the same
thrust level built by a consortium of Lockheed Martin,
Thiokol, and CSD and sponsored by NASA was fired in
the late 1990s. However, neither of these rockets was flight
tested.

II. BURNING RATE


A. The Combustion Model
The combustion process in the hybrid motor is distinctly
different from that in either the liquid or solid motor,
where the oxidizer and fuel are intimately mixed. In
the hybrid motor, combustion occurs as a macroscopic

Rocket Motors, Hybrid

diffusion flame, where the oxidizer is injected at the head


end and the fuel is uniformly vaporized down the port.
Although this fundamental distinction appears at first
as a potential complexity in hybrid ballistic design, it is
precisely this feature of separation of oxidizer and fuel
that accounts for the nonexplosive characteristic of the
hybrid and its consequent greater safety in manufacture,
storage, handling, and operation.
In the hybrid combustion chamber an atomized liquid
flows down the port reacting near the surface of the solid
fuel. Experimental investigations have shown that the controlling factors in the combustion process are the rate of
heat transfer to the solid surface and the heat of decomposition of the solid-phase fuel. The mass flux, which is
regulated by the rate of flow of the liquid phase, determines the rate of heat generated in the combustion zone
and hence determines both the heat transfer and the thrust
magnitude.
The combustion phenomenon is similar to that of a turbulent diffusion flame, where the flame zone is established
within the boundary layer. Experimental measurements
of gaseous oxygen reacting with Plexiglas fuel have confirmed this basic model, as is shown in the Schlieren photograph in Fig. 3. It is convenient to represent this process
by an idealized model in which the flame zone is treated
as the region where the opposing flows of oxidizer from
the port and vapor from the fuel surface meet, as shown
in Fig. 4.
In the real case of finite combustion kinetics, the flame
zone is thickened, with continuous gradients in both temperature and composition. The combustion zone is established at that point where an approximate stoichiometric
mixture ratio has been achieved. The important feature of
this model shows that the combustion zone occurs within
the turbulent boundary layer at a distance from the solid
wall yb which is less than the momentum boundary layer
thickness. The axial velocity at the flame u b is also less
than that at the outer edge of the boundary layer u e . This
is the basic model used to develop the mathematics of the
combustion process as discussed in the following sections.
B. The Regression Rate Equation
In the literature on hybrid motors, it has become common
to refer to the burning rate of the fuel as the regression
rate since, as can be seen in Fig. 4, combustion is displaced from the fuel surface and actually occurs within
the momentum boundary layer. The heat transfer from the
combustion zone to the fuel surface causes it to regress as it
vaporizes to feed the combustion layer. The common term
for regression rate, r , stems from early derivations, where
r referred to the radius of a circular port in the grain and r
to its rate of change. In solid propellants, however, early

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

307

Rocket Motors, Hybrid

FIGURE 3 Schlieren photographs of the hybrid combustion boundary layer at Go = 0.0216 lb/in.2 -sec.

burning rate theories considered flat cigarette burning


surfaces and the term r is commonly used as an abbreviation for the rate of burning.
1. Convection Model
The energy equation that expresses the balance between
the heat transfer and the heat of gasification is
Q w = m f h v ,

(1)

where Q w is the rate of heat transferred to the wall (fuel),


m f is the rate of fuel gasification, and h v is the total effective heat of gasification, i.e., the heat content of the

gaseous fuel at the surface less the heat content of the


solid fuel at the ambient temperature. In general, h v includes three terms: (1) the heat to warm the solid to the
surface temperature, (2) thermal changes prior to gasification (such as depolymerization), and (3) the heat of
vaporization. If the fuel contains a solid which does not
vaporize at the surface, then it includes the heat to simply
raise the particle temperature from ambient to the surface
temperature.
The convective heat transfer to the wall can be expressed
in terms of the Stanton number CH :
Q c = CH b u b h,

FIGURE 4 Hybrid combustion model. The combustion zone is within the boundary layer at a region where the counter
diffusion flows of oxidizer and fuel are approximately equal. The thickened combustion zone is due to turbulence and
finite kinetics.

(2)

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

308

Rocket Motors, Hybrid

where b and u b are, respectively, the density and velocity


of the gas in the flame zone within the boundary layer and
h is the specific enthalpy difference between the flame
and the wall. The evaluation of CH is similar to that in
conventional pipe flow in that it is given in terms of the
friction factor Cf , but with two major corrections. One
accounts for the fact that the heat source is located in the
combustion zone within the boundary layer, resulting in
the formula
CH = (Cf /2)

e u 2e
.
b u 2b

(3)

The other correction accounts for the fact that the heat flow
must travel upstream against a counterflow of fuel vapor.
This latter correction, known as blowing, is expressed
in a term that relates the fuel flow from the surface to the
main core flow along the axis of the port and is designated
as B:
(v)w
r
B=
,
(4)
=
(Cf /2)e u e
(Cf /2)G
where the axial flux e u e is replaced by the total mass flux
G and (v)w by r = m f , the mass flow of gas from the
surface.
Equations (1)(3) may now be combined to yield for
the regression rate


Cf
u e h

mf = r =
.
(5)
G
2
ub hv
Comparing Eq. (5) with Eq. (4), we can see that where
convection is dominant, the blowing term B is also given
by the term in the parenthesis:


u e h
B=
= Bt .
(6)
ub hv
The term Bt , also known as as the mass transfer number, can now be evaluated in terms of thermodynamic
quantities. An approximate value of u e /u b has been
found to be in the range of 1.51.7 for several systems and is less variant than h / h v . The evaluation of
the friction coefficient is made from the turbulent empirical law in the absence of blowing (Prandtl number
Pr = 1): Cfo /2 = 0.03(/Gx)0.2 , where is the viscosity
and G is the mass flux at the distance x. The remaining
parameter to be evaluated is the ratio Cf /Cfo (=CH /CHO )
reflecting the blocking effect on the heat transfer coefficient as shown in Eq. (3). The general equation for the
regression rate is now
r = (Cf /Cfo )0.03(/x) G
0.2

0.8

Bt .

(7)

This equation shows the combined effectiveness of blowing through the blocking parameter Cf /Cfo and the effective heat of gasification h v which appears in Bt . Figure 5

FIGURE 5 Plot of the blocking parameter versus the blowing


term B. As shown in the text, the simple expression C f /C fo =
B0.68 gives an excellent fit in the range 5 < B < 20 covering
most hybrid fuels. It has the virtue of resulting in the simplified
regression expression shown in Eq. (8). () Exact, (+) [ln (1 + B)/
B ]1.088 , () B0.68 .

shows a graph of Cf /Cfo versus the blowing term B. The


exact equation is
 
0.2

Cf
ln(1 + B) 0.8 1 + 1.3B + 0.364B 2
=
,
Cfo
B
(1 + B/2)2 (1 + B)
This can be approximated by Cf /Cfo = [ln(1 + B)
/B]1.088 and shows that the blocking can appreciably reduce the heat transfer. In the typical B range of 520 for
classical hybrid fuels, the Cf /Cfo ratio is empirically given
very closely by B 0.68 , resulting in the final simplified regression rate equation
r = 0.03(/x)0.2 G 0.8 Bt0.32 .

(8)

Note that the important fuel property that determines r


is h v , which is incorporated in Bt as shown in Eq. (6).
Equation (8), however, shows that this term is only raised
to the 0.32 power, which reflects the impact of blowing
on the heat transfer. Thus, the regression rate is not very
sensitive to the vaporization heat of the fuel, an observation
that holds true for a wide variety of fuels.
The practical equation used in engineering development
is a condensation of Eq (8):
r = aG n x m ,

(9)

where the constants a, n, and m are empirically determined. Equation (9) shows in principle that r varies down
the port, which could therefore result in a nonuniform axial burning. This is so because the total mass flux G is the
sum of the oxidizer flux G o and the fuel flux G f , which increases down the port. Countering this effect, however, is

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

309

Rocket Motors, Hybrid

the boundary layer growth term expressed by x m , where


m is theoretically 0.2. The combination of these two
effects tends to cancel, which results in a fairly uniform
axial regression rate. This is illustrated by conducting a
stepwise computer solution on Eq. (9) in conjunction with
the following two equations:

for the correction, which can then be useful for assigning


empirical constants from experimental data.
Assuming a reasonable estimate of Q r is obtained and
the optical absorptivity in the solid is large, we can obtain
its impact on the regression rate from the modified energy
equation as shown:

G(x) = G o + G f (x)
 x
4
G f (x) =
r (x, D) d x.
D2 0

(10a)

m f h v = r h v = Q cr + Q r ,

(10b)

where Q cr is the conductive heat transfer in the presence


of a radiation field. It is important to make this distinction because the increased m f increases the blowing and
hence decreases the convective heat flow. Clearly, if radiation is added to convection, the increased heat transfer
will increase r , but not in proportion to the increase in
Q r because of the increased blockage. The equation that
reflects this coupling effect is approximately given by

Computer simulations of port diameter versus x at different times in a circular port motor using Eq. (9) show that
except for the region close to the origin, the port diameter
remains quite uniform. This result has indeed been observed experimentally and is of considerable significance
in practical motor designs since it leads to the use of a
space-average regression rate.
If a suitable space average D can be assigned, then it
can be moved outside of the integral in Eq. (10b). Equations (9), (10a), and (10b) can now be combined and after
separation of variables and a Taylor expansion, the result
is


2na  x 1+m
r = a  G no x m 1 +
,
(11)
DG 1n
o
where a  = a/(1 + m). The advantage of this expression is
that it only contains input parameters and therefore does
not require iteration. In practice, it is further found that
the term in the parentheses does not vary much, and to a
fair approximation, the simple expression
r = aG no x m

(12)

is very useful in preliminary design. Investigators in the


field will use any of the three equations (9), (11), or (12)
based on the desire for precision or ease of use.
2. Radiation Correction
The radiation transfer from a gas at temperature Tr to a
wall at temperature Tw is given by the equation


Q r = w g Tr4 Tw4 ,
(13)
where is the StefanBoltzmann constant and w and
w are the emissivities of the wall (fuel surface) and gas,
respectively. The term Tw4 is usually neglected since it is
small relative to Tr 4 . The problem in applying this classical
equation to the hybrid motor is the nonuniform radiation
environment, where the hot combustion gas occupies a
relatively thin zone in the port. To take this feature into
account, it is customary to assign a typical average temperature in the combustion region. Although the equation
does not provide an accurate quantitative prediction of the
radiation transfer, it has a reasonable mathematical form

r /rc = exp(0.37Q r /Q c ),

(14)

(15)

where rc and Q c refer to values in the absence of radiation. As an example, if the radiation input is 50% of the
conductive contribution, the increase in r is only 20%.
3. Pressure Effects
Some hybrid propellants have evidenced a pressure dependence of the regression rate in certain regions of operation.
This dependence can occur for two reasons: (1) radiation
emissivity is pressure sensitive through the optical density
term, i.e.,
g = 1 ebpz ,

(16)

where p is the pressure, b is a constant depending on flame


temperature, composition, and percentage of metal, and z
is the optical path length from the flame to the wall; and (2)
combustion kinetics when the reaction time, which is pressure dependent, becomes comparable to or greater than the
diffusion time of reacting species in the boundary layer.
The pressure effect due to radiation can be represented
by a simple formula by combining Eqs. (12)(15) to give
r = rc + k(1 ebpz )
= rc + kbzp.

(17)

The last part of the equation is an expanded version for


low emissivity when bpz < 1. In a typical port, the optical
path z will depend on the port diameter. Thus, the radiation
correction will be influenced not only by the pressure p,
as shown, but also by the motor scale through z. This
behavior has been observed experimentally where, at very
low flux rates (where G 0), the regression rate tends
to flatten out, being dependent now on Q r , the radiation
environment. The term (bz) can vary slightly during a
run with changes in O/F and port diameter. The equation
shows the right intuitive behavior for the radiation pressure

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

310
effect, i.e., no effect for large emissivity when the radiation
is saturated, but an effect at low G when rc becomes small
relative to the radiation term.
The kinetic pressure effect arises out of the pressure
dependence of the kinetics of combustion, where the rate
depends on the concentration of reacting species, which
is directly related to pressure. Typically, in diffusion combustion, where the reactants are not premixed, the rate of
reaction is much faster than the diffusion of the reactants.
In this case, the rate of combustion is diffusion limited.
In the hybrid combustion model, the diffusion rate is governed by the mass flux G. We may expect, however, that
at high flux levels and low pressure, the kinetics can be
rate determining, in which event a pressure dependence
can exist in r .
The combination of both radiation and kinetic effects
on the regression rate is shown in Fig. 6 as a function of
the mass flux G. Since the pressure effects due to radiation
and kinetics occur at opposite ends of the mass flux, this
characteristic provides a clue to the origin of the pressure
sensitivity. The large bulk of hybrid fuels under typical
operating conditions falls in the central diffusion-limited
region. Some aluminized fuels have shown a pressure dependence that generally has been attributed to the higher
emissivity of a particle-laden gas.
4. Temperature Sensitivity
The temperature sensitivity refers to the increase in burning rate and therefore chamber pressure with ambient temperature. This quantity is especially important in solid
rockets, which are not throttled, because of its impact on
the maximum expected operating pressure (MEOP) where

Rocket Motors, Hybrid

the motor case has to be designed to sustain the highest


pressure that can be realized during operation. In typical
hybrid fuel compositions, the latent heat of vaporization
(or decomposition) of the solid is large compared to the
variation in heat content due to the expected extremes of
ambient temperature. Therefore, the initial temperature of
the solid material only slightly affects the rate of regression. From Eq. (5), we can express r as a function of the
solid ambient temperature T , where h vo is evaluated at the
reference temperature To . The result is
r [h vo c(T To )]0.32 .

(18)

The temperature sensitivity of burn rate p is defined as


p =

0.32c
1 r
=
.
h vo c(T To )
r T

(19)

Because the main influence of ambient temperature on


the flow rate of combustion gases is through the burning
rate equation, the temperature sensitivity of pressure for
hybrids (k ) is simply
k =

p
1 p
=
.
p T
1 + O/F

(20)

In solids, the term (1 + O/F) is omitted because the entire propellant including oxidizer and fuel is temperature
sensitive. In a hybrid, however, the important consideration is the effect p has on the variation in fuel consumed
due to an ambient temperature change. For a typical hybrid (c = 0.4 cal/g C and h vo = 300 cal/g at To = 273 K)
p 0.044%/K which reflects the percent variation of fuel
weight. This is well within usual weight margins.
C. Regression Rate Measurements
1. Experimental Methods

FIGURE 6 Radiation and kinetics effects on regression rate. At


low G, radiation dominates heat transfer and the optical path Pz is
influential. For large motors, z is determined mainly by D. At high
G, turbulence diffusion rates can dominate and reaction rates,
which are influenced by P, become rate determining.

The mass flow rates in Eqs. (10a) and (10b) show that r is
a function of the local axial position and mass flux. However, as mentioned earlier, the reasonably uniform axial
burning contour justifies the use of a space-average regression rate that simplifies engineering design at a small
compromise in precision. As a consequence, the large bulk
of experimental data reported in the literature has been
obtained as space-average values. However, investigators
whose prime interest is in studying basic mechanisms have
pursued local methods of regression rate measurement.
The typical local measuring techniques have involved
the use of ultrasonic methods and embedded thermocouples that locally display the regression rate at a given
location.
A simple method for measuring the space-average regression rate is to make a weight measurement of the fuel
grain before and after the test. Although the change in

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

311

Rocket Motors, Hybrid

web thickness or grain diameter for a circular port can be


measured directly, a more accurate evaluation is done by
a weight loss measurement of the fuel. This is shown as
follows, where m is the weight loss of the grain, D2 and
D1 are the final and initial values of the diameter, respectively, r is the average rate, is the fuel density, and L is
its length:

 2
m =
D2 D12 L
4

= (D2 + D1 ) (D2 D1 ) Lt
(21a)
4
D2 D1
4m
r =
=
.
(21b)
t
(D2 + D1 ) Lt
Since the measurement represents an average value over
the time interval t, it is desirable to have the runs fairly
short so that the average values are close to instantaneous.
The measurements are made over a range of average oxidizer mass flux G o and a logarithmic plot of r versus G o
gives the values of a and n in the regression rate equation (12). The problem with this approach, which is especially pronounced in small motors, is that if the burned
web is small, measurement errors can be appreciable.
On the other hand, if thicker webs are used to reduce the
weight measurement error, an uncertainty occurs in the
proper definition of G o . There are three averages for

the mass flux, which is defined as m/port


area. The first
approach selects a simple average D = (D1 + D2 )/2 for
calculating the port area. The second employs an average
D 2 = (D12 + D22 )/2, and the third averages the initial and
final values of G o , i.e. G o = (1/2)(G 01 + G 02 ). These definitions are shown in the following equations, where m o
is the constant oxidizer flow rate:
G o1 =

16m o
(D1 + D2 )2

(22a)

G o2 =

8m o

D12 + D22

(22b)

G o1 + G o2
2m o
G o3 =
=
2

1
1
+ 2 .
D12
D2

(22c)

It can be shown by integrating the regression rate Eq. (12)


for a circular port that the minimum error incurred is one
in which a linear average is used in Eq. (22a). As an example, if the ratio of final to initial port diameter is 2, the error
incurred is about 2% for a linear average, about 7% for
the average area, and about 20% for the average mass flux.
Figure 7 shows a typical calculation with the three averages compared with the accurate equation. The significant
conclusion is that with G o1 , relatively large diameter ratios
(i.e., large burned webs) can be used to incur only small
errors. This conclusion, which may seem counterintuitive,

FIGURE 7 Regression rate constant, based on Go average.


The simple linear average for initial and final port diameters results in minimum error. This somewhat surprising conclusion is
based on the mass flux power law. () Go1 = 4m o / D 2 , (+)
Go2 = 4m o /D 2 , () Go3 = 12 (Go1 +Go2 ).

assumes that the conventional equation (12) is valid. It also


applies to Eq. (9), which employs total G = G o + G f . In
the latter case, m o in Eqs. (22a)(22c) should be replaced
by m o + m f /t as a close approximation when m f < m o , a
condition satisfied by most common hybrids.
Some investigators have attempted to obtain the regression rate parameters in a single motor test with a circular
port. As the burning proceeds and the port diameter opens
up, the mass flux decreases as well as the pressure. The
rates of change of these parameters are related to the regression rate constants. By tabulating the pressure at various times during the run and knowing theoretical c as a
function of O/F, one can in principle obtain regression rate
values during the course of the run. The auxiliary equation
needed is
CD Pc At
c =
,
(23)
m o + m f
where is the combustion efficiency, CD the discharge
coefficient, and m f = D L r . A stepwise integration of
m f gives the values of m f burned at a time t, and the
port diameter D is then obtained from the mass balance:
m f = (D 2 D12 ) L/4. The fuel mass flow m f is integrated over the course of the run and is then compared
with the measured fuel burned. Care must be taken during
the integration to record the change in c with O/F. The
unknown parameter is the combustion efficiency , which
is then adjusted by trial and error until the calculated and
measured values of fuel burned are equal. Although this
approach is mathematically neat, experience has shown
that it is fraught with uncertainty due to (1) the combustion
efficiency adjustment, (2) the possibility of the change in

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

312

Rocket Motors, Hybrid

nozzle throat area At during the course of the run, and (3)
calculation uncertainties in a stepwise integration.
2. Summary of Working Relations
The regression rate equations that have been used are summarized as follows:
r = aG 0.8 L 0.2

(24a)

r = aG L

(24b)

r = aG no L m

r = aG no L m 1 +

2an L 1+m
DG 1n
o

(24c)

(24d)

r = aG n L m [1 exp(D /Do )]

(24e)

r = aG no L m D d p p .

(24f)

Equation (24a) is the classical theoretical expression based


on the model shown in Fig. 4. The values of the exponents are obtained from the boundary layer theory. Equation (24b) allows these exponents to be assigned empirically based on modifications to the model which normally
occur. It also has been shown to be an excellent technical
representation of experimental data, but it has the drawback of requiring an iterative calculation on the computer.
For preliminary design estimates, it is convenient to use
Eq. (24c), which only involves input data. Because of its
convenience, this equation has received the most extensive use in the experimental literature. As discussed earlier, Eq. (24d) is a modification of (24c) to incorporate
the effects of the fuel mass flux on G as derived earlier in
Section II.B.1. Mathematically it approximates Eq. (24b)
with the virtue of containing only input data. It has been
found in many small-motor experiments that the port diameter has an impact on the regression rate, which may
involve a refinement of the theory to account for both
boundary layer interaction and radiation through the optical path. To take these effects into account, Eq. (24e)

has sometimes been employed, where Do is an empirical constant. This value is usually fairly small, so that in
scaling to larger motors, the exponential approaches zero.
Finally, the completely empirical equation (24f ) has been
proposed because it follows a classical chemical engineering treatment. It is useful in describing variations within
a mass of data where only interpolations are required.
However, because of the arbitrary relation between the
variables, extrapolations are hazardous. In summary, the
equations most extensively employed are (24b) and (24c).
Table I shows a summary of experimental data on a
LOXHTPB system covering 35 data points with a wide
range of port diameters from 2 to 11 in. The following
observations are of interest: (1) All expressions give a
satisfactory representation of the data based on the average error, (2) the average error is decreased with Eq. (24e),
although the improvement is small, and (3) the simple theoretical equation (24a) with assigned n and m exponents
provides an acceptable small average error and is useful
in initially estimating regression rates with sparse data.
D. High-Regression-Rate Fuels
The conventional hybrid fuels are characterized by low
regression rates as compared to solid propellants. As a
consequence, relatively thin webs are required in most typical applications, leading to poor volumetric fuel loadings.
The various efforts to increase the regression rate generally fall into two categories. The first approach involves
attempts to increase the heat transfer by generating turbulence through swirl motion or roughening the surface.
This technique, however, has had limited effectiveness because the increased gasification rate acts to decrease the
heat transfer. The second approach involves incorporating additives in the fuel which create an exothermic reaction at the surface, thus effectively reducing the heat
of vaporization. Typical additives include solid oxidizers
such as NH4 ClO4 , NN4 NO3 , and nitro compounds such
as RDX and HMX. In principle, these additives are added

TABLE I Regression Rates for LOXHTPBa


Eq.

Error (%)

(24a)
(24b)

0.21

0.80

0.20

5.6

(24c)
(24d)

aG n L m
aG n L m
aG no L m
aG no L m (1 + 2 an L 1 + m /DG 1o n )

0.16
0.22

0.75
0.75

0.13
0.14

4.0
5.5

0.19

0.76

0.14

4.1

(24e)

aG n L m [1 exp(D/Do )]

0.28

0.76

0.24

2.9

(24f)

aG no L m D 0.06 p 0.05

0.17

0.72

0.05

3.7

a D is an empirical constant, about 1 in. for this propellant. The data are based on 35
o
tests in which the diameter ranged from 2 to 11 in. The 24 small motors have average port
diameters of 2 in. most of which are circular, whereas the 11 large motors have triangular
ports with Dh values from about 7 to 11 in .

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

313

Rocket Motors, Hybrid

in amounts just short of maintaining sustained burning.


In that sense, they resemble low-energy solid propellants.
However, in common with solid propellants, their burning
rates show a pressure dependence. These propellant systems have also been referred to as liquid-augmented solids
where the incentive was to both impart throttleability and
increase the Isp with a more energetic oxidizer. Although
these additives have been effective, the amounts added to
appreciably increase the regression rate elevate the shipping classification to explosive, and requires a more expensive manufacturing facility, similar to solid propellants. It
should also be noted that since such partially oxidized fuel
grains border on sustained burning, they can compromise
the shutdown capability, especially in larger motors.
A more recent concept, which has proven to be much
more effective, is based on utilizing fuels that develop
a melt layer of low viscosity which can entrain liquid
droplets under the influence of the shear force of the axial
gas flow in the port. It has been shown that if the melt layer
is sufficiently fluid, i.e., with low viscosity and surface tension, then under a high gas flux environment (G), the shear
stress can create an instability in the liquid layer, resulting in the formation of liquid droplets that are expelled
from the surface. This behavior is typical of cryogenic
compounds, which crystallize on solidification and exhibit high fluidity on melting. Also included are a class of
nonpolymerizable fuels, such as the alkanes, whose melt
layers are quite fluid. When liquid entrainment occurs, the
resulting propellant flow from the surface is composed of
a mixture of vapor and liquid droplets. This occurrence
can have a profound effect on the regression rate, increasing it by many factors. There are two reasons for this
effectiveness:
1. The droplets that are formed do not require a heat of
vaporization, but only a heat of fusion, which is
generally several times less than vaporization.
2. Since the gas flow from the surface is reduced, the
blowing term B is diminished, resulting in a reduced
blocking of the convective heat transfer.
The effect of blowing on heat transfer (Fig. 5), shows
that with classical polymeric hybrid fuels (5 < B < 20),
the heat transfer is reduced by a factor of about three to
six, whereas with entrainment, the reduction is typically
less than a factor of two. The impact of this reduction is
shown in Eq. (7). This effect exemplifies the effectiveness
of porous wall cooling in combustion chambers where
blocking is desired. By the same token, it explains why,
when gas emanates from a body moving in a fluid, the
drag is reduced.
The cryogens that have exhibited this behavior in
small lab-scale motors are the solid forms of CO, O2 , H2 ,

C2 H2 , CH4 , and pentane (C5 H12 ). The higher alkanes, including many of the paraffins, can also form liquid layers of sufficiently low viscosity to exhibit entrainment of
liquid droplets.
For a hybrid with a melt layer, a certain degree of
entrainment will occur depending on the fluidity of the
melt and the mass flux. The quantitative calculation of the
amount of entrainment is quite complex theoretically, but
some empirical relations have been developed. However,
we can formulate the energy equation and the form of the
regression rate in terms of the mass fraction of droplets
in the gaseous mixture emanating from the surface. If re
represents the regression rate due to entrainment and rv
that due to vaporization, then r = rv + re . If we define the
fraction entrained as = re /r , then the total heat of gasification of the two-phase mixture is
h = (1 )h v + (h v L v ) = h v L v ,

(25)

where L v is the reduction in heat required due to the


formation of the liquid droplets. With entrainment, therefore, the term h v in the mass transfer number Bt in Eq. (6)
is replaced by h and the blowing term B in Eq. (4) is
multiplied by (1 ) to reflect the fact that it is the gas
emanating from the surface that accounts for the blocking.
By incorporating these corrections into Eq. (5), we can
now express the ratio of rates for entrainment relative to
no entrainment when = 0, i.e., R() = r ()/r (0). With
these corrections to both B and Bt , we can now calculate the effect of the entrainment ratio on the regression
rate ratio R(). The Cf /Cfo ratio in Eq. (7) (equivalent to
CH /CHO ) is shown in Fig. 5 as a function of the blowing
parameter B. This curve, which is based both on theory
and experiment, can be represented very closely over the
entire range by the expression [ln (1 + B)/B]1.088 . Table II
shows the values of the amplified regression rates as a
function of the entrainment ratio and at two values of
L v / h v . Note that for the reference B value of 10, regression rate increases up to factors of 1015 are theoretically
possible. Magnifications in this range were indeed suggested in the deep cryogenic solids such as solid methane
and oxygen, whereas for the more moderate cryogen solid
pentane, the increase is of the order of 3. This suggests an
entrainment ratio of 6070% for pentane and over 90%
for CH4 .
Although this method of augmenting regression rates
appears very promising, there are several drawbacks. The
use of cryogens can be expensive and incurs significant
weight penalties because of the insulation required. For
the noncryogens that have fluid melt layers, such as certain paraffins, the class of fuels available is limited since
a low-viscosity melt is required at the surface temperature to allow entrainment. One further caution is that this
system can evidence sloughing if the solid conductivity is

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

314

Rocket Motors, Hybrid

TABLE II Regression Rate Ratio R() versus Fraction Entrained for Two values of Lv /hv a
R()

Cf /Cfo

Lv / hv = 0.5

Lv / hv = 0.7

0.00
0.20
0.40
0.60
0.70
0.75
0.80
0.85
0.90
0.95
0.98
1.00

10.00
8.00
6.00
4.00
3.00
2.50
2.00
1.50
1.00
0.50
0.20
0.00

0.211
0.245
0.294
0.371
0.432
0.472
0.521
0.585
0.671
0.796
0.904
0.999

1.00
1.29
1.74
2.51
3.14
3.57
4.11
4.81
5.77
7.17
8.38
9.45

1.00
1.35
1.93
3.03
4.00
4.69
5.60
6.83
8.58
11.24
13.62
15.75

is the fraction entrained, B is the blowing = m(g)/(C


f /2) G ,
r = const Cf /C f o /(1 L v / h v ), and R( ) = r ( )/r ( = 0).

FIGURE 8 Plot of O/F shift versus port diameter ratio at several


n values.

too high. Nevertheless, laboratory and small-scale experiments appear promising.

III. MOTOR BALLISTICS


In contrast to the solid and liquid motors, hybrid combustion occurs as a diffusion flame that extends the length of
the grain. As a consequence, many of the operational parameters, such as the oxygen-fuel ratio (O/F), thrust, and
fuel flow rate, will depend on the grain dimensions and
therefore change with time as the fuel port opens up. One
must understand these variations in order to minimize the
performance decrements that might result from the associated Isp variations due to the O/F change. In discussing this
behavior, the equation in common use, r = aG no L m , will
be used for simplicity. In the case of a noncircular port, a
hydraulic diameter DH = 4A /P will be used, where A is
the cross section of the port and P is its perimeter.

n

K m no L 1+m
,
D 2n 1
(26)
where K = 4n a 1n , a constant depending only on material properties. For a noncircular port, m f = r P L =
a(m o /A)n P L 1+m . The O/F for a circular port is given by

O/F = m 1n
D 2n1 K L 1+m
(27a)
o
m f = r DL = a

m o
D 2 /4

L m DL =

and the corresponding formula for a noncircular port is



O/F = m 1+n
An a P L 1+m .
(27b)
o
From Eq. (27a) we can see that O/F will increase during
burning according to D 2n1 . Typical O/F shifts in a circular port are shown in Fig. 8 for diameter ratios from 1.2
to 2.0 corresponding to volumetric loading efficiencies of
3075%. When n = 0.5, there is no shift and for a typical n value of 0.7 at a volumetric efficiency of 69%, the
O/F shift is 27% at constant oxidizer flow. As further seen
from Eqs. (27a) or (27b) this shift can be reduced with a regressive oxidizer flow. This fortunately is a typical design
requirement to limit vehicle acceleration during flight.

A. O/F Shift
The O/F in rocket combustion is defined as m o /m f , which
is the reciprocal of the fuelair ratio common in airbreathing engines. In liquid rockets, this parameter is determined solely by the rates of oxidizer and fuel injection,
which are input variables. In solid rockets, the oxidizer and
fuel are intimately premixed and so the O/F is unchanged
regardless of operating conditions. In hybrids, the O/F is
determined by the fuel flow rates, which depend on the
oxidizer flux.
The mass flow of fuel per unit area of surface in a
circular port is given by

B. Stoichiometric Length
The stoichiometric length is that position in the grain
where the integrated fuel burned satisfies the required O/F.
In chemical terms the stoichiometric point (O/F)st occurs
at complete oxidation, when all carbons, hydrogens, and
metals react. In a high-energy system such as rockets,
however, the optimum specific impulse usually occurs on
the fuel-rich side. This is true for hybrid propellants employing liquid oxygen as the oxidizer and especially when
the fuel contains aluminum. In these cases, L st refers to
that length where O/F provides maximum Isp . The relation

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

315

Rocket Motors, Hybrid

between the stoichiometric length and O/F is obtained in


terms of G o :

L st =

G 1n
D
o
4a(O/F)st

1/(1+m)
,

(28)

where DH substitutes for D in the noncircular case. This


quantity is significant because in the motor design, L st
specifies the length of a port, which determines the motor envelope. Since m is usually a small negative number
(m = 0.2 theoretically), a first good guess for (L /D)st is
obtained from the relation (ao = a L m )

(L /D)st = G 1n
4ao (O/F)st .
o

(29)

For typical polymeric fuels, L /D varies between 20 and


30 for maximum Isp . In cases where the envelope is restricted in length, multiple ports are required to shorten the
overall grain length accompanied by an increase in motor diameter. For the high-regression-rate fuels discussed
in Section II.D, thicker webs are possible, resulting in
(L /D)st < 10.

C. Pressure/Thrust Variation
The pressure and consequently the thrust will vary during
a run at constant m o mainly because of the m f variation
shown in Eq. (26). Figure 9 shows how the pressure varies
during a run depending on the n exponent. The equation
governing the pressure is
/At
P = mc

(30a)

Isp ,
F = mg

(30b)

where m = m o + m f , At is the nozzle throat area, c is the


characteristic velocity, g is the gravitational constant, and
Isp is the specific impulse. For a system operating near stoichiometric ratio, the dominant term in the above equations
causing the pressure variation is m f , with c providing a
secondary effect. The curves in Fig. 9 were obtained by
allowing both m f and c to vary with time, using c values for a typical HTPB fuel. For each n value, the initial
O/F was selected on the fuel-rich side of stoichiometric to
minimize the c variation during the run. The thrust variation will closely follow the pressure variation since the
Isp change will closely parallel the c variation.
A summary of formulas for the various parameters useful in preliminary design is given in Table III. The formulas are based on the simple expression r = aG no L m and are
given for circular and noncircular ports.

FIGURE 9 Pressure versus time for various n values; circular port


with D2 /D1 = 2. In this example, the oxidizer flow rate m o and the
initial O/F are fixed, but the O/F varies with time in accordance
with Eq. (27a). The c values were permitted to vary with O/F.

IV. PROPULSION SYSTEM


The propulsion system consists of two main parts, the liquid tank with its feed system, and the combustion chamber
that houses the fuel. The liquid tank with the feed system
components is essentially identical to that of the liquid
system. The feed system typically consists of any of the
following: (1) inert pressurized gas such as N2 or helium,
(2) hot gas derived from a separate gas generator, typically H2 O2 diluted with He, (3) pump run by a gas generator, and (4) a low-boiling-point liquid such as nitrous
oxide (N2 O), which has a vapor pressure in the vicinity of
700 psi at ambient temperature. Although N2 O is not an
energetic oxidizer, it has the virtue of simplicity and low
cost, and therefore has been a favorite of amateurs and
those developing low-cost sounding rockets.
A. Combustion Chamber
The combustion or thrust chamber can be considered as
divided into five major components: the injector assembly,
TABLE III Summary of Ballistic Parametersa
Parameter

Circular

Noncircular

D 2n1
m 1n
o
K L 1+m

1/(1+m)
D
G 1n
o
4a(O/F)st

An
m 1n
o
a P L 1+m

1/(1+m)
DH
G 1n
o
4a(O/F)st

L /D (approx.)

G 1n
o
4ao (O/F)st

G 1n
o
4ao (O/F)st

m f

K m no L 1+m /D 2n 1

a(m o /A)n P L 1+m

O/F
L st

aK

= 4n a 1n , a0 = a L m .

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

316
the grain compartment, the aft mixing chamber, the nozzle, and the thrust vector control, which typically, but not
necessarily, is part of the nozzle. For flights in the atmosphere, it is usually more efficient to simply use fins or
wings with actuators.
B. Injector Assembly
The types of injectors used may include all of those used
in liquid propellant rockets plus several that are unique
to hybrids. There are two basic injector design philosophies: one entails direct injection of the oxidizer down the
port and the second involves injection into a precombustion chamber where the oxidizer is largely gasified and
heated prior to flowing down the port. The direct injection
technique was historically used because the early work involved small motors with single circular ports and direct
injection appeared obvious. With large-diameter motors,
however, where multiple ports are necessary, a precombustion chamber to vaporize the oxidizer and provide a
uniform entrance condition to all the ports is desirable.
In the early work with multiple ports, individual injectors were used, as shown in Fig. 11. In that case, it
was necessary to verify that all injectors had identical
flow characteristics to guarantee uniform burning down
the ports. A further disadvantage of this arrangement is
that if a hypergolic fuel is required for ignition, each individual injector will need its own means of injecting the
hypergolic fuel. Despite these issues, the use of individual
injectors did work well, giving relatively high combustion
efficiency (9193%) with N2 O4 as the storable oxidizer.
When a precombustor is used, one of the more conventional liquid injectors is used which provides a uniform
atomized spray. The hypergolic liquid is injected in much
the same fashion as the fuel in a liquid rocket, except, of
course, in much smaller quantities. The injector types that
can be used include the showerhead, impinging jets (doublets, etc.), use of splash plates, and swirl sprays involving
both hollow and full cone patterns. A detailed discussion
of some of these injectors can be found in texts on liquid
rockets. In the early French and Swedish programs, where
nitric acid was used as the oxidizer, the main solid fuel was
selected to be hypergolic with the oxidizer and therefore
no separate ignition fuel was needed. Hypergolicity was
imparted to the fuel by incorporating an amine compound.
When a precombustor is used, its configuration is selected
so that there is sufficient residence time to vaporize the
oxidizer.
C. Grain Design
The simplest grain configuration is the one with a single
circular port. In this case, the web thickness is typically

Rocket Motors, Hybrid

in the range of one-fourth the diameter, leading to a volumetric loading of 75% in the fuel section of the case. For
larger motors, however, the web must increase to maintain
a reasonable loading density. With typical regression rates,
this would lead to unacceptably long burning times. The
solution to this dilemma is to select a multiport design that
provides an increased burning surface in a shorter length
but larger diameter motor. Each port satisfies the L /D requirement as shown in Table III. As the motor diameter
is increased, the number of ports is increased as dictated
by the web thickness, which in turn is determined by the
regression rate and the burning time.
Figure 10 shows six typical grain designs. The single
cylinder has the advantage of being highly efficient and
desirable for applications requiring long burning times.
An example is the case of the target drone, described earlier, which required burning times of up to 5 min with a
low thrust sufficient to overcome atmospheric drag. If a
high-regression-rate fuel is employed (as discussed in Section II.D), then a thicker web can be used for high-thrust
boost applications. The seven-cylinder cluster represents
a means of retaining the simplicity of the single port in
an optimum packaging configuration. This approach also
simplifies manufacturing and motor loading since standard identical units can be employed. The design tradeoff here is a more efficient packaging envelope at the expense of additional free volume between cylinders, which
amounts to about 22%.
The wagon wheel design is very popular and has been
tested in configurations of up to 15 ports (AMROC 250klb-thrust motor). The double wagon wheel has not been
tested but is a conceptual design suitable for very large
motors. Figure 11 shows a 12-port wagon wheel design
before and after the test. An important consideration of
the multiport design is the grain support toward the end
of burn. To avoid shredding of the webs when they become thin toward the end of a burn, it has been customary
to employ a low-density web support that is fairly rigid.
A disadvantage of the multiport designs is that the sharp
corners in the ports lead to residuals at the end of burning,
which increase the inert rate. This penalty can be reduced
by using low-density sliver savers or by allowing a lowthrust tailoff, which is allowable in certain applications
such as space motors. Another consideration of the wagon
wheel design is whether to allow burning to occur in the
center circular port. It is obviously more efficient volumetrically if this occurs, but care must be taken to match
the hydraulic diameters of the circular port with the triangular or quadrilateral ports of the wagon wheel. If this
is not done, the gas flow will prefer the larger hydraulic
diameter, thereby starving the other ports.
Grain loading in the motor case is usually accomplished
in two ways. The first, which is common in solid rockets, is

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

317

Rocket Motors, Hybrid

FIGURE 10 Examples of grain shapes.

by bonding the grain to the case wall. This is done by filling


the empty motor case with the uncured liquid fuel and appropriate mandrels to create the port holes. This approach
requires the fuel to have adequate physical properties to
withstand the ignition loads due to case expansion and acceleration loads in flight. The second approach involves
loading the fuel in a cartridge that is then inserted in the
motor. By providing the proper support for the grain and
allowing a small clearance between the cartridge and case
wall, the grain can be disengaged from the expansion of
the case wall during pressurization. As a consequence, the
elongation property requirement on the grain is considerably reduced.
Both approaches require adequate properties to avoid
grain slumping during storage and potential shear stress

FIGURE 11 Multiport grain before and after firing. [Courtesy of


the Chemical Systems Division of United Technologies.]

failure during accelerated flight. In the case of cartridge


loading, aft end support of the grain must be considered.
These considerations are generally of little importance for
small motors, but can be critical for large motors in flight.
D. Aft Combustor/Nozzle
1. Aft Chamber
Because of the turbulent diffusion flame, the combustion
gas exiting the grain will generally exhibit striated beams
of both high and low O/F values. Flow in the center of
the port will be more oxidized than near the fuel surface,
which contains a fuel-rich layer. Aside from the turbulence in the boundary layer, additional mixing is desired.
Several approaches that have been found effective include
the following:
(a) Provide an aft combustion chamber with an L/D
ratio sufficient to give increased mixing residence
time.
(b) Use mechanical mixers (turbulent generators) at the
aft end of the grain to stimulate turbulent mixing.
(c) Use a submerged mozzle to generate aft end mixing
from the protruding forward end of the nozzle.
(d) Provide for aft end injection of part of the oxidizer to
stimulate mixing. The additional benefit here is that
oxidizer makeup can be provided during low-O/F
operation, usually during the early portion of the

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

318

Rocket Motors, Hybrid

flight. The objection to this approach, however, is the


additional plumbing complexity.
(e) Use a multiport design to decrease the average
diffusion length between high- and low-O/F regions.
The simplest approach that has been found to be effective is the use of a submerged nozzle at the end of the
mixing chamber, usually with a multiported grain. The
typical combustion efficiency c is about 95%, which is
23% below that of the premixed liquid and solid motors.
2. Nozzle
The nozzle design generally follows the same ablative approach as the solid motor but with one caveat. An aluminized solid propellant usually operates in a low-O/F
regime where Isp is maximized. In this region, where the
combustion gases contain a low concentration of oxidizing
species (such as O2 , OH, H2 O, CO2 ), graphite nozzles perform well, showing low erosion. However, most hybrids,
unless aluminized, operate in a region where the oxidizing species are in a higher concentration. Consequently,
except for short-duration runs or for those designed to
run fuel-rich, a ceramic-containing nozzle insert or one
containing silica fortification has been found to provide
more resistance to oxidation than graphite. Typical values
for erosion rates for aluminized solids are in the range of
37 mils/sec, whereas typical nonaluminized hybrids fall
in the range of 1025 mils/sec depending on the nozzle
material selected and the range of O/F and pressure operation. In general, nozzle erosion is governed by the same
laws as hybrid burning and erosive burning in solid propellants. One can expect Eq. (7) to be applicable to the

behavior of an ablative nozzle in a combustion gas, with


the G o term now referring to the mass flux of oxidizing
species. By Reynolds analogy, the diffusion of reactive
oxidizing species would follow the same law as heat conduction. As a consequence, one would expect the erosion
rate to depend both on the fraction of oxidizing species
and the pressure that determines the mass flux G. These
effects have indeed been observed in practice. Table IV
shows the mole fraction of oxidizing species for some
typical propellants. Of interest is the fact that the addition
of aluminum reduces the oxidizing species since those
propellants optimize performance on the fuel-rich side.
These considerations assume that the combustion temperature is sufficiently high that the rate-determining step
in nozzle erosion is governed by diffusion of reactive
species. For cool propellants, this is not the case and
erosion effects are minimal. Similarly, during the early
transient of motor operation before the nozzle has heated
up to its steady temperature, erosion is low. For this reason,
nozzle erosion is of lesser concern in short-duration runs.

V. PROPELLANT COMBINATIONS
Because the hybrid employs both a liquid and a solid,
it enjoys the largest selection of candidate propellants.
These range from storable to cryogenic to metallized
combinations.
A. Storable Propellants
Where storability is required, any of the oxidizers that have
been employed by liquid propellants is suitable. These
include nitrogen tetroxide (N2 O4 ), nitric acid (HNO3 ),

TABLE IV Performance of Hybrid Propellantsa


Specific gravity

mf of
oxidizing
species

Tc (K)

Propellant

O/F

Fuel

Oxidizer

Average

I sp (std.)
500 psi

HTPBLOX
HTPB/AlLOX
PELOX
HTPBN2 O4

2.25
1.21
2.50

0.98
1.25
0.92

1.14
1.14
1.14

1.09
1.19
107

277
275
281

0.51
0.12
0.51

3635
3795
3530

3.30

0.98

1.43

1.29

257

0.42

3420

1.70
1.80
7.33

1.33
1.35
0.98

1.43
1.55
0.77

1.39
1.47
0.79

259
252
247

0.10
0.20
0.30

3770
3500
3370

2.80

1.56

0.77

0.89

253

0.04

3930

7.70

0.92

1.38

1.31

247

0.85

2600

0.80
2.60

1.57
0.62

1.14
1.44

0.32
1.05

408
350

<0.01
<0.02

2690
4970

HTPB/AlN2 O4
HTPB/AlRFNA
HTPBN2 O
HTPB/AlN2 O
PEH2 O2 (90%)
HTPB/BeLH2 LOX
PE/LiFLOX

a Aluminized propellants contain 15% Al in the total propellant. FLOX (with PE/Li) contains 85% F . The Be in the tribrid
2
is 80% of the solid fuel and 25% of the total propellant. The hydrogen acts essentally as a working fluid because of its low
molecular weight. The term m f is the mole fraction of oxidizing species, which is defined as 2 O2 + O + H2 O + CO2 (other
oxidizing species are usually negligible). PE is polyethylene with the formula (CH2 )x .

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

16:56

319

Rocket Motors, Hybrid

red fuming nitric acid (HNO3 + N2 O4 ), hydrogen peroxide (H2 O2 ), and certain fluorine compounds such as ClF3 .
Combinations of some of the oxidizers have also been
used, such as NO dissolved in N2 O4 (to lower the freezing
point) and an aqueous solution of hydroxylamine nitrate
(N2 H4 O4 ). The former was used in the CSD drone flight
program and the latter has only been tested to date in the
laboratory.
Also included in this class are the reverse hybrids,
where the solid is the oxidizer and the liquid is the
fuel. Typical oxidizers include ammonium perchlorate
(NH4 ClO4 ), ammonium nitrate (NH4 NO3 ), and the nitrates of sodium and potassium. Higher energy oxidizers, such as nitronium perchlorate (NOClO4 ), hydrazine
nitrate (N2 H5 NO3 ), and hexanitroethane [C2 (NO2 )6 ],
have also been considered but not developed because of
their unstable character. The typical fuels used with AN
and AP have been kerosene (or JP) and the hydrazines
(N2 H4 and UDMH).
B. Cryogenic Propellants
The most extensively employed oxidizer in this class is liquid oxygen, or LOX. It delivers high performance since
it is not diluted with other inert elements, and is suitable
with a wide variety of fuels, principally the polybutadienes. The higher energy form of oxygen, namely ozone,
has been considered but not employed in practice because
of its cost and instability.
A higher energy oxidizer in this class is fluorine. This
oxidizer is typically used as a solution in liquid oxygen
(FLOX, containing 6070% F2 ) for three reasons. One is
that with a typical hydrocarbon (like HTPB), the maximum energy output is obtained when the carbon is oxidized to CO and the hydrogen to HF. The energy level can
be further improved by incorporating lithium in the fuel
to deliver a vacuum Isp in excess of 380 sec. The second
advantage in using FLOX is that its corrosivity is reduced
relative to pure F2 . The third reason is the high density of
liquid fluorine (1.5 g/cm3 ), which results in an improved
mass fraction in a vehicle. Despite these advantages, although this oxidizer has been tested at thrust levels of
40,000 lb, it has not been seriously considered for operational use because of its toxicity. It remains, however, as
a serious option for certain space engines.
Finally, there is the class of cryogenic solids discussed
in Section II.D. These include solidified O2 , CO, OH4 , H2 ,
and C5 H12 and have only been tested on a laboratory scale.
Their advantage of high Isp and high regression rate makes
them suitable in special circumstances. However, for practical applications, this advantage is overshadowed by the
expense, insulating weight penalty, and inconvenience of
handling these cryogens in the solid state.

C. Other Propellants
Because the hybrid fuel can sustain modest cracks and
voids with no disastrous effects, the physical property requirements, compared to solids, are greatly reduced. This
expands the field of eligible candidates. Indeed, early testing included fuels such as wood, coal, lucite, and even
compacted garbage fuels held together with 510%
binder. An example of this latter fuel that has been tested
is shown in Fig. 12. Because of its useful performance, as
shown in the figure, it has been considered for auxiliary
power in space stations.
The highest performing propellant not involving esoteric ingredients is a tribrid involving the bipropellant
H2 O2 with beryllium incorporated in a polymeric fuel
binder. The H2 O2 is burned in the precombustor and
the hot gas flows down the port, burning the berylliumcontaining fuel. This propellant combination provides a
calculated vacuum impulse in excess of 500 sec. Figure 13
shows a graph of the performance of this hybrid as compared to H2 O2 under the same operating conditions. Were
it not for the toxic nature of BeO, this propellant would
be a serious candidate for space applications.

D. Summary
A compilation of hybrid propellant combinations is shown
in Table IV. Performance data are displayed both under
standard conditions and in space for the higher performing propellants. The most popular propellant has been the
LOXPB (such as HTPB) combination because of its high
performance and reasonable cost. For storable combinations, the most practical have been those using N2 O4 and
H2 O2 as oxidizers. The target drone programs that enjoyed 15 years of successful flight experience used both
N2 O4 - and HNO3 -based oxidizers. These oxidizers have
high density and are hypergolic with amine fuels, such as
the hydrazines, which provides a convenient means of ignition. The highest density propellant combination in this
list is aluminized HTPB with RFNA as oxidizer. Highdensity propellants are generally desirable because they
result in higher mass fraction propulsion systems. The
next to the last column lists the mole fraction of oxidizing
species, which provides an indication of the corrosiveness
of the combustion gases on carbon nozzle throat inserts.
Note that all the metallized combinations have low concentrations of oxidizing species. Indeed, experience has
shown that the aluminized solid propellants show lower
erosive behavior to carbon or graphite nozzles than the
nonaluminized combinations. This is true despite the fact
that the metallized propellants have high flame temperatures because of their energy. To resolve this problem, the
nonmetallized propellants have sometimes used ceramic

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

320

EN014G-835

July 31, 2001

16:56

Rocket Motors, Hybrid

FIGURE 12 The trash rocket. Left: Grains formulated with 2% and 4% HTPB binder. Right: Waste vacuum specific
impulse potential as a function of oxidizer composition. NAR, North American Rockwell. [Courtesy of the Chemical
Systems Division of United Technologies.]

P1: GTV/MBQ

P2: GRB Final Pages

Encyclopedia of Physical Science and Technology

EN014G-835

July 31, 2001

Rocket Motors, Hybrid

FIGURE 13 Performance of Be/H2 /HTPBO2 tripropellant compared to that of H2 O2 under the same operating conditions;
p c = 500 psi, area ratio = 50.

or silica-filled plastics instead of carbon-based materials.


It may be surprising to observe that the highest flame temperature propellant, FLOX plus a lithium-containing fuel,
ran for over 50 sec in a relatively large motor without undue erosion. This may be attributed to the exceedingly low
oxidizer content since the gases were principally CO, HF,
Li, H2 , and H.

SEE ALSO THE FOLLOWING ARTICLES


COMBUSTION CRYOGENICS ELECTRIC PROPULSION
FUELS HEAT TRANSFER ROCKET MOTORS, LIQUID ROCKET MOTORS, SOLID SPACECRAFT CHEMICAL
PROPULSION SPACE NUCLEAR PROPULSION

BIBLIOGRAPHY
Altman, D. (1991). Hybrid rocket development history, Presented at
the 27th Joint Propulsion Conference, AIAA.

16:56

321
Altman, D., and Humble, R. (1995). Hybrid rocket propulsion systems,
In Space Propulsion Analysis and Design (R. Humble, G. Henry,
and W. Larson, eds.), pp. 365441, McGraw-Hill, New York.
Barrere, M., and Moutet, A. (1963). La Propulsion par Fusees
Hybrides, International Astronautical Congress.
Boardman, T. A., and Carpenter, R. L. (1997). A comparative study of
the effects of liquid-versus gaseous-oxygen injection on combustion
stability in 11-inch-diameter hybrid motors, Presented at the 33rd
Joint Propulsion Conference, AIAA, Seattle, WA.
Calabro, M. (1991). European hybrid propulsion history, Presented at
the AIAA Hybrid Propulsion Lecture Series, 29th Aerospace Sciences
Meeting, Reno, NV.
Green, L., Jr. (1964). Progress in astronautics and aeronautics. In
Heterogeneous Combustion, Vol. 15, p. 451, Academic Press, New
York.
Karabeyoglu, M. A., and Altman, D. (1997). Transient behavior
in hybrid Rockets, Presented at the 33rd Joint Propulsion Conference, AIAA, Seattle, WA.
Karabeyoglu, M. A., Cantwell, B. J., and Altman, D. (2001). Development and Testing of Paraffin-Based Hybrid Rocket Fuels, AIAA
Paper No. 2001-4503, Joint Propulsion Conference, July 811, Salt
Lake City, Utah.
Larsen, C. W., Pfeil, K. L., De Rose, M. E., and Carrick, P. G. (1996).
High pressure combustion of cryogenic solid fuels for hybrid rockets, Presented at the 32nd Joint Propulsion Conference, AIAA.
Marxman, G. A., and Gilbert, M. (1992). Turbulent boundary combustion in the hybrid rocket. In Ninth Symposium (International) on
Combustion, pp. 371383, Academic Press, New York.
Netzer, D. W. (1972). Hybrid rocket internal ballistics, CPIA Publication 222.
Rocketdyne. (1962). Research on hybrid combustion, Summary
Report, January 30, 1962, Contract Nonr 3016(00).
Smoot, L. D., and Price, C. F. (1965). Regression rates of metallized
hybrid fuel systems, Presented at the 6th Solid Propellant Rocket
Conference, AIAA, Washington, DC.
St. Clair, C., Rice, E., Knuth, W., and Gramer, D. (1998). Advanced cryogenic solid hybrid rocket engine developments: Concept
and testing, Presented at the 34th Joint Propulsion Conference,
AIAA.
Wooldridge, C. E., and Muzzy, R. J. (1964). Measurements in a turbulent
boundary with porous wall injection and combustion. In Tenth Symposium (International) on Combustion, pp. 13511362, The Combustion Institute.
Wooldridge, C. E., and Muzzy R. M. (1966). Internal ballistics considerations in hybrid rocket design, AIAA Paper No. 66628, Colorado
Springs, CO.

Vous aimerez peut-être aussi