You are on page 1of 42

Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

BURNING MECHANISM OF AMMONIUM PERCHLORATE PROPELLANTS1

by
Martin Summer-field 2, G. S,. Sutherland3 ,
M. J. Webb4, H. J. Taback5, K. P. H a l l 6
Guggenheim Jet Propulsion Center
Princeton University
Princeton, New Jersey

ABSTRACT
The purpose of this research is to clarify the burning
mechanism of heterogeneous solid propel I ants, with the u l t i -
mate objective of providing a rational basis for prediction
and/or control of the steady state burning rates of propel I ants

Presented at the ARS 13th Annual Meeting, New York, Nov.


17, 1958.

1. This work was supported at various intervals by funds pro-


vided by the Aerojet General Corporation, the A i r Reduction
Company, and by the Office of Naval Research through Project
Squid. The work is continuing under direct contract with
the Office of Naval Research. The authors are grateful for
the support of these organizations. The experimental re-
sults are described in greater detail in the Ph. D. thesis
of Sutherland and in the M.S.E. theses of Taback and Webb.

2. Professor of Jet Propulsion, Department of Aeronautical


Engineering.

3. Former Ph. D. Candidate; now at Boeing Airplane Company.

4. Former M.S.E. Candidate; now Research Associate, Department


of Aeronautical Engineering.

5. Former M.S.E, Candidate; now with Aerojet General Corporatbn

6. Research Associate, Department of Aeronautical Engineering.

141
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

of various compositions. Inasmuch as the burning rate must


depend directly on the physical and chemical structure of the
flame zone, the experiments were directed toward the determi-
nation of temperature, composition, fiow characteristics, re-
action intensity, and physical conditions inside the very
thin reaction zone. The f i n d i n g s of these experiments a l l
combine to indicate that for ammonium perchiorate propel I ants
the reaction zone is mostly, if not entirely, In the gas phase,
that it is less than O.I mm thick at pressures over 10 atm, that
its thickness becomes smaller as the pressure is raised, and
that diffuslonal m i x i n g occurs simultaneously with chemical
reaction. A theory of steady state burning is developed on
the basis of these f i n d i n g s which leads to a novel two-
parameter burning rate law:
I a b
" "

The burning rates measured in a strand-burner with a polysty-


rene ammonium perchiorate propel lant show excellent agreement
with the pressure effect predicted by this theory. The para-
meters "a" and "b" are obtainable from the experiments, and
the magnitudes thus obtained are shown to be consistent with
their physical meaning in the theory. The effects of oxidizer
particle size, fuel-oxidant ratio, and copper chromite catalyst
on the burning rate curve are shown to be e x p l i c a b l e in terms
of specific effects on "a" and "b". The conclusion is that
the burning of a certain class of heterogeneous propel lants is
controlled by the rate of reaction in a so-called granular
diffusion flame zone of the type deduced from these experiments.

I. Structure of the Burning Zone of an Ammonium Perchiorate


Propellant

Introduction

In the design of s o l i d propellant rockets, knowledge of


the b u r n i n g rate of the propellant under a w i d e range of oper-
ating conditions is of primary importance. The p r i n c i p a l
variables that can affect the burning rate of a specified type
of heterogeneous propellant are the fuel-oxidant ratio, the
chamber pressure, the temperature of the solid, and the particle
size distribution of the ground oxidizer. Other variables that
the charge designer must take into account are the gas velocity
in the grain port (erosive b u r n i n g effects), size and shape
of grain (scale effects), and processing conditions (curing
time, for example). Certain typesof additives in s m a l l per-
centages are shown to modify strongly the burning rate behavior
of various propellants. At present, the information on burning
142
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

rates and the effects of various operating conditions is


purely empirical, and frequently incomplete, even for so-called
standard propeMants. There is a great need, therefore, for a
theory, or at least a set of principles, to unify the mass of
empirical data, and more important, to point the way toward
propel I ant improvement. Such theory w i l l undoubtedly have
several aspects: the explanation of the differences between
oxidizers w i l l probably be on thermodynamic and chemical
kinetic grounds; the explanation of the differences between
fuel binders w i l l surely envolve h i g h polymer chemistry; and
the explanation of the effects of pressure, fuel-oxidant ratio,
and particle size w i l l probably lie in the domain of flame
science. It is the latter aspect with which this paper is con-
cerned .
The objective, therefore, is to develop a theory of burn-
ing that w i l l predict at least the effects of pressure, particle
size, and fuel-oxidant ratio. For this purpose it is necessary
to have a physico-chemical model of the flame zone describing
the temperature distribution, rate of reaction, d i f f u s i o n a l
processes, gas flow conditions, etc. Although a few attempts
have been made before to develop a burning theory for heter-
ogeneous solid propellants, they rested on scanty information
about the structure of the flame, and partly because of this,
were totally incorrect in their results. Part I of this paper
describes the structure of ammonium perchlorate propellant
flames as determined by a series of different types of experi-
ments. Part II presents a novel theory of burning that y i e l d s
certain predictions as to the effects of pressure, fuel-oxidant
ratio, particle size, and burning rate catalyst content. Part
III describes b u r n i n g rate tests for a series of ammonium
perchlorate propellants, and it is shown that the predictions
of the theory are confirmed.

Conditions at the Burning Surface

The propellant that was chosen as the subject of this


research is a s o l i d two-phase mixture of f i n e l y ground ammonium
perchlorate mixed w i t h a styrene-base resin that enters the
process as a l i q u i d and is f i n a l l y polymerized to a s o l i d . The
resin is prepared and marketed by the Rohm and Haas Company
under the formula designation P-I3. Ammonium perchlorate was
chosen as the oxidizer because of its broad technological im-
portance in present-day rocketry, and it appears that it w i l l
continue to be important for years to come. The resin P-13
was chosen as one of many submitted by various companies
because of the s i m p l i c i t y of processing , desirable physical
properties, absence of personnel hazards, and because it is

143
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

representative of a w i d e class of s o l i d hydrocarbon fuels.

It is first necessary to establish the conditions at the


surface during burning. Is the surface wet (that is, molten)
or dry? The question is whether the oxidizer gasifies by first
m e l t i n g or by dry sublimation, and whether the fuel does like-
wise. This has a bearing on the theory of the flame, since the
fuel and oxidant gases might issue from a wet surface and enter
the gaseous reaction zone p a r t i a l l y pre-mixed, as where the
gases from a dry surface must enter the reaction zone i n i t i a l l y
unmixed. The next and closely related question is whether the
gases entering the flame zone proceed to mix by turbulent
diffusion, or whether the flow is non-turbulent so that m i x i n g
must occur by molecular d i f f u s i o n . Photography was used to
provide experimental answers to these questions. (Figs. 1-5)

Fig. I shows two photomicrographs of the propellant


matrix. Thin slices of a so-called coarse-ground propel I ant
and a fine-ground propellant were photographed under a
microscope with transmitted l i g h t after the oxidizer crystals
were leached out with water. If the burning surface of the
propellant remains dry d u r i n g burning, as is shown in Fig. 4,
then these illustrations show the type of surface from which
the oxidant and fuel gases issue.

To determine whether the flow is turbulent or non-


turbulent, two types of photographs were taken. Fig. 2 shows
the flame as it appears with its own emitted l i g h t . From these
snapshots, there is no evidence of turbulence in the thin
reaction layer less than I mm from the surface. There is an
indication of turbulent flow in the more distant regions of the
flame, but this may be largely due to jet m i x i n g of the issuing
combustion gas with the surrounding inert nitrogen in the
chamber, and so it would be an edge effect far removed from
the important reaction zone. This interpretation is confirmed
by the spark schlieren photographs in Fig. 3. These photographic
results are not to be taken to i m p l y that the flame is micro-
scopically steady adjacent to the surface. It is rather that
the unsteadiness observed in h i g h speed color motion pictures
and inferred from the unsteadiness of the temperature traces of
Fig. 9 below is that of v a r i a b l e unmixedness in the reaction
layer. In the theory of Part II, it w i l l be postulated that
the i n i t i a l unmixedness occurs in random non-steady fashion but
that the stream is non-turbulent.

The question of whether the burning surface is dry or


wet is answered by the silhouttte type cinephotomicrographs of
Figs. 4 and 5. The first figure shows the burning surface of a
thin slice of ammonium perchlorate propellant in the open
144
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

atmosphere. The pertinent feature of this series of frames


is the microscopically jagged profile, indicating that it is
a dry unmelted surface. This interpretation is more convinc-
ing after inspection of the second of these figures, which
shows the b u r n i n g surface of potassium perch I orate. When the
perch I orate decomposes it is known that the KCL residue must
melt before it can vaporize, and so it is not surprising to
f i n d a smooth p r o f i l e and evidence of b u b b l i n g action. (The
growth and collapse of a bubble can be seen in this particular
sequence.) The temperatures pertinent to this discussion are:
KCI melting point, I044K; KG 164 decomposition temperature,
about 800 K, that is, below the m e l t i n g point, 883 K; probable
decomposition temperature of NH4.CI04, about 1200 K; probable
pyrolysis temperature of P-13 resin, about 1000 K. (cf. Ref.I,
p. 325).

Thus, it is established that the burning surface is dry


and hence the gases liberated at the surface of the ammonium
perchlorate propellant are i n i t i a l l y unmixed as they enter the
gaseous flame zone. The next question relates to the rate of
reaction in the gaseous flame, or moresimply, the thickness
of the reacting layer. At one time it was thought that the
height of the v i s i b l e flame might be indicative of the reaction
thickness, and so photographic measurements were made by
several investigators (cf. Ref. 2) of the variation of flame
height with pressure. The measured heights varied from 5 mm
at pressures about 200 psi down to 2 mm at 2000 psi . However,
the sequence of Fig. 4 provides by chance some evidence that
the adiabatic flame temperature is achieved in a much smaller
distance. In connection with the brightness-emissivity
temperature determinations described in the section below,
a small amount of NaCI was added to the propellant to provide
the yellow D l i n e s of sodium. In Fig. 4 an NaCI particle is
shown glowing at a distance of about 50 microns from the surface
showing that its b o i l i n g point of about I700K has been reached
at this distance. This is evidence of an extremely thin
reaction zone.

Temperature Profile of the Reaction Zone (Thermocouple Method)

More detailed information about the temperature d i s t r i -


bution is obtainable by running a fine thermocouple through
the reaction zone. The method of K l e i n et al (Ref. 3) was used
for this purpose. Thermocouples of 1/2 m i l platinum-rhodium
were imbedded in a strand of propellant, and the output was
recorded on an oscillograph w h i l e the flame burned past it.
The voltage-time record was then transformed to a temperature-
distance record, examples of which are shown in Fig. 6.
Although the traces cannot indicate the actual level of the
145
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

temperature of the surface, it Is clear that the gaseous reac-


tion zone is less than 100 microns in effective thickness at
pressures above 200 psi, if it is assumed that the surface
temperature is in the neighborhood of 700*-800C.

Temperature Profi|e of Gaseous Reaction Layer (Optical Method)

Another approach to the temperature distribution is by


spectrophotometric determination of the brightness and
emissivity of the flame in a spectral interval where the
emission is strong enough to be measured and where the parti-
cular emitter is in thermal e q u i l i b r i u m with the hot gases.
Inasmuch as the sodium D l i n e s have been shown in various
researches to satisfy these requirements adequately, even in
the active reaction zone of a gaseous flame, a small amount
of NaC1 was added to the ammonium perchlorate propellant, and
the double monochromator was set for 5893 AU with a s l i t width
just sufficient to pass both D lines. The apparatus is shown
diagrammatically in Fig. 7.

The details of the method have been described by others


who have used it for the measurement of flame temperature, for
example, Mi Mar et a| (Ref. 4). The p r i n c i p l e is to measure
the emissive power (brightness) of the flame and at the same
time its emissivity, then to compare the brightness of the
flame with that of a radiation source whose brightness has
been calibrated as a function of temperature. In the arrange-
ment shown in Fig. 7, the radiation beam that enters the photo-
m u l t i p l i e r tube produces a deflection on the oscilloscope
proportional to the intensity of the beam. The deflections
for three conditions are of importance: (I) the beam from
the comparison lamp alone; (2) the beam from the flame alone;
(3) the combination of the beam from the flame and that from
the lamp after passing through the flame. The lamp itself
is calibrated by means of a standard optical pyromter.

Condition (1) produces the deflection Dj shown on the


upper trace of Fig. 8, before the propellant strand is i n s t a l l e d
on the axis. The signal is chopped in order to get a r e l i a b l e
zero determination. Condition (2) produces the deflection D2
In the lower trace, w h i c h is recorded from left to right whi le
the strand burns past the beam axis. Condition (3) produces
the deflection D2+ 03. The zero l i n e is recorded at intervals
In order to compensate for zero drift. The ratio of 03 to Di
determines the spectral absorptivity of the flame, and on the
assumption that Kirchoff's law is v a l i d , the e m i s s i v i t y can
be equated to it. The ratio of D2 to Dj determines the bright-
ness temperature of the flame, and by correcting for the
emissivity just determined, the true temperature can be
146
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

computed. By chopping the signal 250 times per second, a


temperature determination is possible every 4 milliseconds.
For a burning rate of 5mm/sec, this corresponds to a temperature
point every 20 microns in the fjame zone. Such spatial resolu-
tion would be v a l i d only i f the l i g h t beam could be accurately
focussed in the flame and confined to an area of that size.
Actually, the results reported here are believed to have a
spatial resolution of about 30 microns. Details of the focus-
sing procedures and of the beam confinement methods are given
in detail in the cited Ph. D. thesis.

Actual osci Ilographic traces are shown in Fig. 9 for two


propellants, and typical computed temperature profiles are
shown in Fig. 10. Several points deserve comment. One is that
this method, l i k e any method based on v i s i b l e radiation, is im-
practical for recording temperatures below about I800K because
the brightness of a thermal source becomes too weak to measure.
Another is that excess temperatures by as much as 200K are
possible as a result of n o n - e q u i l i b r i u m excitation of the sodium
atoms. A third uncertainty is the nature of the averaging
process, since the flame is r e a l l y inhomogeneous both micro-
scopically in its interior and macroscopica1 Iy by reason of the
cool periphery. F i n a l l y , the temperature p r o f i l e in the first
100 microns from the surface may be i n v a l i d in some cases
because of t i l t i n g of the b u r n i n g surface.

Despite these uncertainties, two conclusions emerge. One


is that the adiabatic flame temperature is reached at a distance
of 100 microns or less from the surface, confirming the thermo-
couple results. Another is that the flame zone is microscopic-
a l l y inhomogeneous with inhomogeneities of the order of at least
ten microns, a situation that can be inferred from the r i p p l i n g
of the s i g n a l when the beam of the flame is being recorded.
These conclusions lend support to the t h i n gaseous d i f f u s i o n
flame model that is described in Part II.

Spectrographic Scanning of the Flame Zone

An important method for investigating the structure of a


flame zone is to survey the spectral emissions of the flame
s p a t i a l l y . A H i l g e r Intermediate Quartz Spectrograph was
focussed on the flame in two arrangements, as shown in Fig. II.
In Position I, the spectrum was first recorded by a l l o w i n g
the propellant strand to burn past the optic axis of the spec-
trograph w h i l e the plate shutter was continuously open. This
produced the spectrum shown at the bottom of Fig. 12. With
such long-duration exposures, the p r i n c i p a l emitters in the
flame were i d e n t i f i e d : OH, C2, CH, CN and NH. The b u r n i n g
took place i n the open a i r .
147
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

Spectrograms are shown In Fig. 12 that were snapped (1/25


to 1/5 of a second) w h i l e the propellant flame was momentarily
in the proper location to produce a streamwise profile of the
emitted radiation. It is shown that the significant radical
emissions are situated strongly in a very thin layer adjacent
to the propellant surface. The spatial resolution of this
method is not as f i n e as that of the opticaI temperature traverse:
it is possible only to say that the active reaction zone l i e s
w i t h i n I/2mm of the surface.

Another indication of the spatial extent of the active


reaction zone is provided by the spectrograms of Position 2,
with the burning surface facing the si it. It so happens that
the flame gas issuing from the burning surface of a strand at
atmospheric conditions fans outward to form a jet wider than
the strand itself. Thus, in the spectrograms shown in Fig. 13,
the emitters that l o g i c a l l y radiate in the w i d e flame brush,
l i k e OH, Co, Cu, Na, and incandescent carbon e x h i b i t lines that
are longer than the strand dimension, w h i l e l i k e CN, NH, NO, C2>
and CH, which are generated exclusively in the perchlorate
decomposition and oxidizing zones, exhibit lines that are only
as long as the strand dimension. (The cobalt l i n e s come from
the cobalt napthenate accelerator used in the polymerization
process.) These results confirm thespatial distribution of the
chemical reaction zone revealed in Fig. 12. It should be ob-
served that the emissions in the thin reaction zone are s i m i l a r
to those of a premixed flame of a hydrocarbon mixed with oxides
of nitrogen. (Ref. 5). Evidently, at atmospheric pressure,
the diffusional m i x i n g process is sufficiently rapid in compar-
ison with the oxidizing reaction rates that a good deal of
premixing can occur before the onset of oxidation. This concept,
that at low pressures the chemical reactions are the rate-
control I ing process, is employed i n the theory of Part II.

II. Theory of Burning of a Composite Solid Propellant

Physico-chemical Model of the Flame

The object of the f o l l o w i n g analysis is to develop a


theoretical burning rate law for heterogeneous s o l i d propel I ants
of the ammonium perchlorate type that w i l l predict the depend-
ence on pressure, oxidant-fuel ratio and particle size, and
possibly the effects of burning rate catalysts. The theory
must start with a physico-chemical model of the reaction zone
that is consistent with the f i n d i n g s of the experiments of
Part I.

The following one-dimensional flame model is postulated.

148
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

A qua si-steady gaseous flame exists adjacent to the burning


surface, as indicated in Fig. 14. The surface is dry, that is,
the oxidant and fuel gases are liberated directly from the
solid phase by sublimation or pyrolysis. Mixing of the oxidant
and fuel occurs only in the gas phase. No chemical reactions
of significance for flame theory occur in the interior of the
solid, and so no heat release occurs until the actual surface
is reached. The sublimation or pyrolysis reactions on the
surface may be either exothermic or endothermic. Heat of
oxidation is released in the thin reaction zone in the gas
phase.
The vapors of fuel or oxidizer, or both, are released in
the form of pockets of a certain mass content, and then these
pockets proceed to burn in the surrounding medium of the
opposite reactant. It is assumed that the average mass content
of each pocket is very much smaller than that of an average
oxidizer crystal, but that its mass is somehow related to the
mass of the crystal. The bigger the crystal, the bigger the
pockets. In the theory, the mass per pocket is taken as in-
dependent of pressure. The pockets are gradually consumed at
a rate controlled by diffusional mixing and chemical reaction,
as they pass through the flame zone, and the heat of combustion
is released at a corresponding rate. This type of flame zone
can be called a "granular diffusion flame".
The physical reason for the occurrence of such pockets,
and particularly their independence of pressure, is s t i l l un-
known, but it is the object of a current research project.
The most p l a u s i b l e assumption is that the pockets contain fuel
vapor, and that these fuel pockets are carried along i n the
surrounding stream of oxidizer gas w h i l e they burn. At the
present moment, the only defense of this assumption is that it
works. If it is true, it is significant for the possible con-
trol of burning rates by methods heretofore unsuspected.

Burning occurs as a result of energy fed back from the


flame to the exposed surface. The mode of energy transfer is
taken to be mainly thermal conduction. Radiation is ignored
in the present analysis, but a paper w i l l be offered in the
near future on the modification of the theory to take this into
account because there is some experimental evidence that radia-
tive transfer is important in some cases. (See the paper by
Blair, Bastress, Hermance, H a l l , and Summerfield in this
VoIume.)
The two transport processes, thermal conduction and gas
diffusion, are taken to be molecular in nature and not turbulent,

149
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

on the basis of the findings in Part I. Further support for


this assumption is that the Reynolds Number based on the flame
thickness is so low as to preclude turbulence, in the light of
experiments on f l u i d flow *n packed beds, a closely analogous
situation. (Ref. 9)
Ana I ys is

The following analysis is based on the quasi -one-d imen-


sional model of F i g . 14. Although this analysis is relatively
elementary in the l i g h t of the present state of the theory of
pre-mixed gaseous flames, it is w e l l known that s i m p l e formula-
tions can give the correct pressure effects although they do
not y i e l d the absolute burning rates. S i m i l a r l y , this type of
analysis w i l l not y i e l d absolute burning rates.

Jlfi = mass burning rate =


C& = specific heat of sol id (average of fuel and
ox id i zer )
»S = femperature at surface of solid (average)
\gs= avera9e thermal conductivity of flame gas
3t surface
(AT/ \
V %Wg£ = temperature gradient in flame gas at surface
y. = heat of combustion of the propel lant
net heat release (positive) for gasification
of the propel lant
Qs is probably negative for the fuel component, and is
p l u s 256 cal/gm for ammonium perch I orate. The average is
probably at least p l u s 100 cal/gm. This term is therefore
smaller than the cs (Ts~To) term, but not n e g l i g i b l e .

The meaning of Ts, the surface temperature, deserves ex


planation. It is taken here to be an average of the tempera-
tures on the oxidizer and fuel surfaces, which are probably
different. A series of researches by Chaiken, Schultz, et al
at Aerojet (Ref. 6) on the linear pyro lysis rates of separate
fuel and oxidizer components show that the rates can be corre
lated by Arrhenius expressions for r^ and rQ of this form:

= bo «xp (Eo/ RTos)

150
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

The values of B and E are different for the two compo-


nents. It is postulated that the pyrolysis processes proceed
independently, but that the rates ro and rf are equal to the
observed burning rate r for the total propellant. For this to
be true, the two surface temperatures have to be different.
In our theory, the two-temperature postulate is accepted as
v a l i d , but we deal only with Ts, the average of the two tem-
peratures, for s i m p l i c i t y . (See the Note Added in Proof at
the end of this paper.)

The temperature Ts is taken for the present as a quantity


independent of pressure, particle size, or mixture ratio, but
it may be expected to depend on the amount of copper chromite,
a known decomposition catalyst for perchlorate. The approxima-
tion that Ts is independent of pressure is a l l o w a b l e only if
the pyrolysis is governed by an activation energy E that is
large compared with RTS. For ammonium perchlorate and poly-
styrene, the activation energies are about 20,000 cal/mole,
which justifies the approximation. Otherwise the Arrhenius
equation would have to be coupled to the energy equation for
the flame, and the two equations would be solved simultaneously
for r and Ts as functions of pressure.

In the above equation, an expression is needed for the


flame thickness L as a function of pressure, that is, the rate
of burning in the gaseous flame as a function of pressure.
Two extreme cases are possible. First, at very low pressure,
the rate of molecular diffusion is very much faster than the
oxidation reaction, so that the latter occurs essentially in a
premixed gas. W e l l known theory is a v a i l a b l e for this case.
Second, at very high pressure, the chemical reaction rate i s
so fast that the rate of burning is controlled entirely by the
rate of inter-diffusion. A separate theory is needed here.
The joining of the two extremes is sr d i f f i c u l t mathematical
problem which we w i l l treat in an approximate manner.

The low pressure pre-mixed flame can be regarded s i m p l y


as a stream of velocity /fl/^^in which a second order reaction
is taking place.

(Average values of £ , ^ M-fe J >J°*± > e + c »> are


assumed.)

151
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

If this is inserted in the energy equation, it i s pos-


s i b l e to solve for <yp , the mass burning rate.

<rn -T> )A
CTs-TV) -
Thus, the burning rate is proportional to the pressure,
with an exponent of unity. The flame thickness is seen to be
inversely proportional to pressure.

Now, the other extreme case of a "granular diffusion


flame 11 is considered. In accordance with the assumption de-
scribed above, pockets of fuel vapor of mass ^cand of dimen-
sion d_ are generated at the surface, and these masses are
unaffected by pressure but dependent on crystal size.

The lifetime of a pocket is determined by the rate of


gas diffusion, inside and outside, to the surrounding flame
surface. D is the molecular diffusivity, the same for both
gases, and averaged over the flame zone. Then, the flame
thickness and the mass burning rate can be computed as before.

152
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

It can be seen that the mass burning rate is proportional


to the one-third power of the pressure, which is in the right
range for ammonium perchlorate propellants. The flame thick-
ness is inversely proportional to the one-third power of p.

The two extreme cases at low and high pressure are thus
solved. The general intermediate case is not simple. We make
the assumption, therefore, that the flame thickness varies
with pressure partly as if it were reaction rate controlled
(L|) and partly as if it were diffusion controlled

This expression for the flame thickness can be introduced


into the energy equation to get the burning rate.

The pressure dependence then emerges in the following


s i m p l e two-parameter formula:

Possible Tests of the Theory


*-f+>
It would be d i f f i c u l t to compute accurate theoretical
values for the coefficients a^ and j^ for testing the theory
against experiment, but a number of significant tests are
sti I I possible:

(I) The theoretical burning rate law involves only two adjust-
able parameters. It is proposed to compare the theory against
measured burning rate curves over a very wide range of pressure.
The usual empirical two-parameter equations are known not to

153
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

fit the data, except over limited ranges. Thus, the proposed
law should be better than:

r = apn
and r = a + bp

(2) The parameters a_ and J^ as obtained from the f i t of the


experimental data should be tested for their sensitivity, re-
spectively, to oxidant-fuel ratio and particle size. It can
be seen from the above analysis that a^ should be very sensitive
to flame temperature and b_ should be very sensitive to average
particle size. For convenience, we can refer to a^ as the re-
action time parameter and b_ as the diffusion time parameter.
(3) A decomposition catalyst such as copper chromite (Ref. 7)
should raise the burning rate by depressing the value of Ts
and, therefore, a propel lant of this type should show a_ and _b^
reduced by s i m i l a r ratios.
These three test objectives formed the program for the
burning rate studies reported i n Part III.
Other Attempts at a Theory of Burning
There have been other attempts at a theory for hetero-
geneous propellant burning. 0. K. Rice proposed in 1945 that
the flame zone consists of streamers of fuel and oxidizer gas
that burn according to the well-known theory for laminar diffu-
sion flames in the form of ax{symmetric jets, with the idea
that the internal oxidation processes are very fast in compari-
son with the diffusion process. This jet diffusion model leads
to the conclusion that the burning rate is unaffected by pres-
sure; which of course is incorrect. (Ref. I, p 322 for a
summary.) Recently, Nachbar and Penner have formulated the
Rice model in a more accurate way mathematically, and have
broadened the physical assumptions to a l l o w for slow as w e l l
as fast oxidation reactions. As can be seen from the Summer-
f i e l d theory, the latter aspect of the Nachbar-Penner approach
is a step in the right direction, but there s t i l l remains the
great d i f f i c u l t y of obtaining useful solutions to the system
of equations. Moreover, there is good reason to b e l i e v e that
the more exact analysis w i l l lead at h i g h pressure to the same
incorrect zero pressure effect (although with greater effort)
since the physical model is essentially the same (Ref. 6).

Some recent experiments by Friedman on the seIf-decompo-


sition of pressed pure ammonium perchlorate sticks show burning
rates of the same order as those of propellants incorporating

154
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

fuels, that is, about 5 mm/sec. This has raised the idea that
the burning of actual ammonium perchlorate propel I ant is con-
trolled by the decomposition rate of the perchlorate crystals.
Chaiken has a p p l i e d this idea to a theory of burning in which
the linear regression rate of the crystal surface surrounded
only by its own monopropellant flame determines the burning
rate of the propellant, w h i l e the regression of the fuel sur-
face somehow takes care of itself. (See his paper in this
Volume.) The theory makes no use whatever of the oxidation
reaction in the gas phase: it admits its existence but it
does not use either the flame temperature or the reaction rate.
Of this theory, several remarks can be made: (I) The physical
model is hardly plausible. (2) The analysis of the linear re-
gression rate of a crystal is not really acceptable even i f
the physical model is accepted. (3) Even i f the analysis is
accepted, the results as to pressure dependence, mixture ratio
dependence, and particle size effects cannot be reconciled
with experimental burning rates. Yet, It must be conceded
that Friedman's experimental burning rates of pressed perchlo-
rate may have a bearing on the burning of a mixed propellant,
which future research may disclose. (See the Note Added in
Proof at the end of this paper.)

In Part III of this paper, the Summerfield theory is


tested against measured burning rates. It is shown therein
that the theory explains in many respects the experimental
facts.

Ill. Comparison of Burning Rates with Theory


Experimental Method

The object of these experiments was to measure the burn-


ing rates of a series of ammonium perchlorate propellants over
a wide range of pressure, that is, from 15 to 1500 psia, to
test the v a l i d i t y of the Summerfield burning rate equation.
It was of particular interest to have propellants of different
oxidant-fuel ratios, different average particIe sizes, differ-
ent amounts of copper chromite catalyst, and different distri-
butions of particle size.
The propellant was processed in the laboratory at
Princeton by starting with commercial ammonium perchlorate,
g r i n d i n g it in a hammermi I I grinder to a certain size distribu-
tion, m i x i n g it thoroughly in 600 gram batches with P -13
resin (a styrene polyester copolymer), curing the mixture in a
rectangular mold at elevated temperature until polymerization
to a s o l i d had occurred, and then sawing the s o l i d into the

155
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

desired burning-rate strands. The strands were typically


about 6 to 7 inches long and 3/16 to 1/4 inch square, restrict-
ed on the long sides with a quick-drying lacquer.
The burning rates were measured under a controlled pres-
sure of nitrogen from 15 to 1500 psi in strand-burner s i m i l a r
to that o r i g i n a l l y used by Crawford et a I (Ref. 10). The
strand was threaded with four wires, from the top end to the
bottom, which were connected to external circuits through
sealed electrical leads. The top wire was the ignition wire;
next was a f u s i b l e wire that started two timers when the flame
melted it; next was a fusible wire at the midpoint that stopped
one timer; the last at the bottom was a f u s i b l e wire that
stopped the second timer. The burning rate is s i m p l y the dis-
tance between successive f u s i b l e wires d i v i d e d by the time of
traverse of the flame.
Two determinations were made on each strand, and the
probable error of the method was determined by a statistical
study of the differences between such pairs of rates. It was
concluded that the probable error was about plus or minus 3#,
and that very l i t t l e of this was due to the apparatus itself.
Photographic study of the process of burning of typical strands
showed that random t i l t i n g of the burning surface during the
progress of a run could account for the f u l l range of the ob-
served error. The t i l t i n g was probably due to inhomogeneities
in the strand resulting from some imperfection in the process-
ing, and to imperfect restriction of the sides of the strand.
Many details of the experimental method require attention
in order to insure successful measurements. The reader inter-
ested in conducting such experiments should consult the theses
of Sutherland, Taback, and Webb (Refs. I I , 12, 13).

Effects of Particle Size and Fuel-Oxidant Ratio

The experimental burning rate versus pressure data re-


ported in Sutherland's thesis are shown in Fig. 15. To deter-
mine whether these curves follow the Summerfield theory, the
burning rate equation is rewritten in the following form:

(JL)=
The experimental values of r and p are then plotted as
(p/r) versus (p^/-^). jf -j-^e cjata conform to the equation, it
should be possible to draw a straight l i n e through the points.
The intercept on the (p/r) axis y i e l d s the value of a^, which

156
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

i n c i d e n t a l l y must come out p o s i t i v e if the theory is v a l i d .


The slope y i e l d s the value of JD_.

It can be seen in F i g . 16 that Sutherland's data a l l con-


form to the proposed equation. The fit i s remarkably good. It
can be pointed out that no data points were rejected, a l l were
used.

Four propellants were tested, corresponding to a coarse


grind (average 120 microns) and a fine grind (average 16
microns), and a cool mixture (15% oxidizer) and a hot mixture
(80$ oxidizer). The values of a^ and ^determined from the
lines of Fig. 16 are as follows:
Propel I ant a. J^
(1) Cool - coarse 365 39.0
(2) Cool - fine 400 19.8
(3) Hot - coarse 245 27.0
(4) Hot - fine 160 17.3

It can be seen that the values of £ and j^ show the antici-


pated response to mixture ratio and particle size. A hot pro-
pel lant produces a flame with faster kinetics than a cool one;
compare (3) with (I) and (4) with (2). A fine oxidizer results
in a shorter diffusional time than a coarse one; compare (2)
with (I) and (4) with (3).
It can be said that these results agree with the predic-
tions of the theory. However, because of a number of debatable
questions about the data on burning rates in Fig. 15, particu-
larly in the low pressure region, Webb undertook to re-determine
the effects of particle size and fuel-oxidant ratio. His re-
sults are shown in Fig. 17 and the re-plot is shown in Fig. 18.
Again, the data fit the theoretical law rather w e l l over
the entire range from about 5 psia to 2000 p s i a . The values of
bj the d i f f u s i o n time parameter, seem to conform to the ex-
pected e f f e c t of p a r t i c l e size, but the cross-infIuence of mix-
ture ratio is stronger than in Sutherland's data. The system-
atic influence of oxidant-fuel ratio on a^ found in Sutherland's
data and expected on the basis of the theory is curiously lack-
ing in Webb's data, and his values are much smaller than
Sutherland's. For this we have no good explanation, except
that the intercept a^ depends markedly on the low pressure burn-
ing rates, and low pressure rates are hard to measure exactly.
Webb reported some evidence of incomplete combustion at low
pressures, which may have something to do w i t h the d i f f i c u l t y .

157
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

Effect of Particle Size Distribution

In searching to e x p l a i n the effects of particle size, the


question was raised as to whether a broad range of sizes or a
mixture of two normal distributions (bimodal distribution)
would e x h i b i t anomalous behavior. Taback tested several d i f -
ferent oxidizer grinds, as shown in Fig. 19. The size distri-
butions were measured with a Sharpies Company Micromeragraph,
which c l a s s i f i e s particles according to Stokes1 Law by their
rate of f a l l in an air column. These distributions were used
to obtain the burning rate data shown in F i g . 20. A l l four
propellants were made with the same fuel-oxtdant ratio, 11.5%
perchlorate. It should be observed that the large amount of
scatter in the very f i n e ground propellant was probably due to
the d i f f i c u l t y of casting this type of propellant mix without
flaws or bubbles.

The re-plot of these burning rate points is shown in


Fig. 21. The data follow quite w e l l the expected straight
lines. The values of a_, although not very precise, f a i l in the
same range as Sutherland's values, corresponding to h i s 80$
oxidizer propellant.
The interesting thing is that the .b. va I ues range from
10.I to 36.0, exactly where they would be expected, and they
l i n e up in the order that would be expected on the basis of
weight-mean particle diameter. Numerically, these values of JD_
lead to a mass content for the average pocket that corresponds
roughly to the mass of the interstitial fuel between the packed
crystals in the solid matrix. This point is of interest since,
on the basis of the present flame theory, it is the weight-
mean particle size that would determine the average temperature
gradient and hence the burning rate. Other investigators have
previously proposed a surface-mean particle size as the sig-
nificant average; this question should be tested in the future
more systematically.
In summary, the study of particle size effects reveals
agreement with the theory.
Effect of Copper Chromlte Catalyst
F i n a l l y , it was decided to test the theory with propel-
lants in which small percentages of copper chromite powder
have been added. Friedman's experiments had shown that this
is a catalyst for the thermal decomposition of ammonium per-
chlorate. If so, it should affect the theory by depressing
Ts, the mean surface temperature during burning. This would

158
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

have the effect of reducing both a^ and J^ by s i m i l a r fractions.


Taback's determinations of burning rates are shown in F i g . 22
and the re-plot in F i g . 23.

It is evident that these data, too, fit the present


theory, and the expected reductions in a_ and b^ showed up ap-
proximately as expected. It should be r e c a l l e d a g a i n that
precise values of a^ are d i f f i c u l t to obtain since they depend
e n t i r e l y on low pressure data.

Cone I us ions

The new burning rate equation for ammonium perch I orate


propellants correlates the measured burning rates with a h i g h
degree of success. The effects of average particle size,
fuel-oxidant ratio, combustion pressure, and copper chromite
catalyst content enter l o g i c a l l y into the equation. The suc-
cess of the s i m p l e concept that chemical reaction is rate-
control I ing at low pressure w h i l e diffusion is rate-controlling
at h i g h pressure is demonstrated by the experimental fact that
particle size effects tend to vanish at low pressure w h i l e they
are all-important at h i g h pressure. Application of this con-
cept in the present theory w i l l a l l o w one to predict the maxi-
mum burning rate that can be achieved at each pressure by
g r i n d i n g to the finest possible size. Other practical pre-
dictions w i l l become apparent.

Research on this problem in the future w i l l be concerned


with the role of radiative heat transfer, which was ignored in
this theory. Also, the basic reason for the appearance i n the
flame of pockets of gas with average mass content independent
of pressure w i l l be investigated. F i n a l l y , it is now appro-
priate to extend the research to the phenomenon of plateaus,
which probably involve a mechanism not included in the present
theory.

Note Added in Proof

In the time since this paper was written in 1958, further


work has been done on the effects of pressure and particle
size, particularly with narrow size range oxidizer (so-called
unimodal distributions), and with p o l y s u l f i d e and epoxy fuels
as w e l l as P-13. Along with the determinations of burning
rates, observations with a microscope have been made of ex-
tinguished burning surfaces to see how closely the condition
of the surface corresponds to that c a l l e d for in the theory of
this paper. The higher pressure range, above 1000 psi, in the
case of a p o l y s u l f i d e propellant, reveals an unexpected

159
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

behavior: The particles of perchlorate seem to pyrolyze much


more rapidly than the surrounding fuel, leaving tiny bore-
holes in the surface where they have burned out. This is
interpreted as evidence that the monopropeliant-type burning
of the perchlorate dominates the assumed two-gas diffusion
flame, in this high pressure range. At the same time, in this
pressure range, these unimodal propellants burn at rates that
agree f a i r l y closely with the rate of burning of pure per-
chlorate sticks (Ref. 7). On the other hand, at lower pres-
sures, say below 500 psi, the condition of the surface seems
to correspond to the assumed model of the granular diffusion
flame theory, and the burning rate data tend to agree with
thi s theory.
These results indicate that the nature of the surface
may not always agree with the demands of the theory, depending
on how w i d e l y different the two separate pyrolysis rates are.
Experimentally, it is found that these "bore-hole11 effects
and these departures from the granular diffusion flame theory
tend to be masked and disappear when bimodal distributions
are used and when more h i g h l y oxidized mixtures are used.
These results and the hypotheses to explain them are the
subject of the forthcoming Ph.D. Thesis of E. K. Bastress.
(See h i s paper in this Volume.)
References
1. Geckler, Richard D., The Mechanism of Combustion of Solid
PropeHants Selected Combustion Problems,
pp 289-339, Butterworth Scientific Publications,
1954.
2. Lawrence, R. W. and Dekker, A. 0., Structure of Flames
from Ammonium Perchtorate PropelJant Jet Propulsion,
Vol 25, p 81, February 1955.

3. K l e i n , R. Mentser, M., von Elbe, G., and Lewis, B., Peter-


mi nation of the Thermal Structure of a Combustion
Wave by Five Thermocouples Jour. Phys and C o l l o i d
Chem. Vol 54, pp 877-884, June 1950.

4. Mi Ilar, G. H., Winans, J. G., Uyehara, 0. A., and


Myers, P. S., A Fast Electro-Optical Hot-Gas
Pyrometer. Jour. Opt. Soc. Am. Vol. 43, p 609,
1954.

160
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

5. Wolfhard, H. G., and Parker, W. G., Spectra and Combustion


Mechanism of Flames Supported by the Oxides of
Nitrogen. Fifth Combustion Symposium, pp 224-230,
Reinhold P u b l i s h i n g Corp., 1955.
6. Schultz, R., Green, L. and Penner, S. S., Studies of the
Decomposition Mechanism, Erosive Burning, Sonance
and Resonance for Solid Composite Propellant
Combustion and Propulsion Colloquium, AGARD,
pp 367-421, Pergamon Press, 1958.
7. Friedman, R., Nugent, R. G., Pumbel, K. E., and
Surlock, A. C., Deflagration of Ammonium Per-
chlorate Sixth Combustion Symposium, pp 612-618,
Reinhold P u b l i s h i n g Corp., 1956.

8. Chaiken, R. F., A Thermal Layer Mechanism of Combustion of


Solid Composite Propellant with Application to
Ammonium Nitrate Propellant Aerojet-General Corp.
Tech Memo 290 Azusa (Calif.) 1957. (Also in
this Volume.)
9. Bernard, R. A., and W i l h e l m , R. H., Turbulent Pi ffusion
in Fine Beds of Packed Solids. Chem. Eng.
Progress, Vol 46, pp 233-244, May 1950.

10. Crawford, B. L., Huggett, C., Daniels, F., Wilfong, R. E.,


Direct Determination of Burning Rates of Pro-
pel lant Powders., Jour. Analyt. Chem. Vol 19,
p 630, 1947.

11. Sutherland, G. S., Mechanism of Combustion of an Ammonium


Perchlorate-Polyester Resin Composite Solid
Propellant. Ph.D. Thesis, Aeronautical Engineer-
ing, Princeton, 1956. (Copies obtainable from
University Microfilm, Inc., Ann Arbor, Mich.)
12. Webb, M. J., The Dependence of Linear Burning Rate upon
Pressure for Ammonium Perch I orate - Polyester
Resin Composite Solid Propellant. M. S. E.
Thesis, Aeronautical Engineering, Princeton,
1958. (Copies obtainable from the Librarian,
Forrestal Research Center, Princeton, N. J.)
13. Taback, H. J., The Effects of Several Composition Factors
on the Burning Rate of an Ammonium Perch I orate
Sol id Propellant. M. S. E. Thesis, Aeronautical
Engineering, Princeton, 1958. (Copies obtainable

161
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

from the Librarian, ForrestaI Research Center,


Princeton, N. J . )

PHOTOMICROGRAPH OF 80:20 FINE PROPELLANT


15 MICRON SLICE LEACHED IN h O

PHOTOMICROGRAPH OF 80:20 COARSE PROPELLANT


50 MICRON SLICE LEAChED IN H20

Microscopic examination of the distribution of


ammonium perch I orate cyrstals in a composite (hetero-
geneous) propellant.
162
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

Purchased from American Institute of Aeronautics and Astronautics

en
O
6
O
-o
m

l
§
^

V i s i b l e flame of NH4CI04- PI3 propel lant b u r n i n g in


nitrogen atmosphere. Strands are 1/8 inch square.
Left to right: 300, 900, 1500 psi.
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

Purchased from American Institute of Aeronautics and Astronautics

I
o
a) 10 psi -o
yo
a) 200 psi O
•o

£
xD

b) 50 psi b) 400 psi

F i i _3. Spark schlieren photographs of flame zone of 1/2 mm


strands of burning propellant. Four pressures: 10,
50, 200, 400 psi. Reveals non-turbulent flow at
surface.
SOLID PROPELLANT ROCKET RESEARCH
CD
C
M_ . "L.
0 4- ZJ
c •o
in (0
JZ 0
a. —. U
(D 0 fO
L. a_
CO O L.
0 ZJ
CL tn
O
0 s
0 CO L.
!_ L.
0 -
a. jz tn
0 u—
c L_ (0
.— 0 0
o Q. >
0
Purchased from American Institute of Aeronautics and Astronautics

x-x E cr
0 ZJ
0 ,—.
tn c •
0 § o
E E
(0 (0 r~
L. -H
o tn
0 c
o 0 0
in 0
(a U
~ •4-
L_ E
-o ZJ
0 cn o
0 o

Q. CD —. CD
cn c c
,—. 0 .—
j— c u c
CD L.
.— ZJ .—» ZJ
X JD CO jQ
165
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.
SOLID PROPELLANT ROCKET RESEARCH
4- 0>
ex. ;s
0 r
o
X CO
CD ——
(O
- CD CD
CD ~
• o: c
en L.
U_ • JD
-h-
O C CD
•K <a c:
03 CD Z3
• ii. O ~Q
E 2-H
—— CL ——
Purchased from American Institute of Aeronautics and Astronautics

CO 03
CD CO
CO H-
OL. L. CD
L_ —— —
CD _£= O
O O £
0 CD -C
1 s 'i
-K Z5
O — <D
^I CO O
Q. CO (O
CD (D H-
--
— O
O <=*-
O
LO
CT
166
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

I8OO

1600

1400

1200

1000

800

600

400

2OO

100 200 300 400 500 6OO


Distance, microns

_6. Temperature profiles of flame zone obtained with im-


bedded thermocouples, showing very thin reaction layer,
f i n e ground ammonium perchlorate.

167
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

168
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

(A)
innnnnn BASE LINE

^-CH 0 PPERI

(B)
finiinr in
*sf*\ D2 CHOPPER 2
BASE LINE

60 CPS TIMING MARK

Idealized intensity-time trace given by apparatus of


Fig. 7 as propel I ant flame moves past axis of optical
system. (A) Tungsten lamp calibration; (B) Flame plus
lamp.

169
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

Purchased from American Institute of Aeronautics and Astronautics

co
O

-o
TO
a) 75:25 COARSE, 50 PS I O

TO
O
o
q
TO
m
c/>
TO
n

b) 80:20 FINE, 400 PS I

Actual intensity-time trace of flame plus lamp for two


tests. (A) 75:25 coarse ground at 50 psi . (B) 80:20
f i n e ground at 400 psi. Irregularities due m a i n l y
to non-steady unmixedness in flame zone.
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

2800

2
°°°0 400 800
DISTANCE FROM SURFACE, MICRONS
Fig.10, Three temperature profiles of propellant flame zone
determined by brightness-emissivity method, showing
t h i n reaction layer less than 1/10 mm.

Hilger intermediate
quartz spectrograph

Propellent strand
(Two positions!

Propellent strand with


burning surface
facing slit
"Slit jaw

Propellent strand position


Front view-time exposure
when shutter is tripped
(Position 2)

Front view-with shutter (Position I)


\ \. Two arrangements of spectrographic scanning system for
determining spatial distribution of emitters in reac-
tion layer. Position I -Streamwise distribution;
Position 2 - Lateral distribution.

171
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

Purchased from American Institute of Aeronautics and Astronautics


3360A 3883A

(a) Snapshot at 1/25 sec.


Slit = 0.5mm Max. NH intensity
Propellent surface Max. CN intensity
Carbon
continuum
Top of
(b)Snapshot at 1/25 sec.
sllt
SI it =0.5 mm
Propel I ant surface o
NH \CN Carbon at surface TJ

O
-o
3360A 3590A 3883A 4315A 4730A 5I65A

(c) Snapshot at 1/5 sec. >


Z
Slit = 1.5 mm
Propellant surface
(Slit jaw opened toward red.)

Scale - Angstrom units x 10"

70
3064A 3360A 3590A 3883A 43I5A 4730A 5I65A n
421 A
f \ . . . . . . . . I. . .
(d) Strand burned by
with shutter open

Fig. |2. Spectrograms outlined with Position I of Fig. I I


shows CN and NH emission strongest in thin zone near
surface, i m p l y i n g thin reaction layer. (62 and CH
emission are obliterated by carbon continuum).
SOLID PROPELLANT ROCKET RESEARCH
n surface

0
O)
C
£
1^ en (0 D in
E c c D
u_ L.
ID p ^ en 4- o 0
O «O
"o L_ in
0 O O
6 0 § 0
L.
!!
(N (D
t c c U
*
c. o O s- U
0 .£) 0
<b _ in L. 0
£ ? £ +- in ~
o
^ c ^
O 0
in 'I __
3
5> •Q
2j O in -V
^ CL . c
Purchased from American Institute of Aeronautics and Astronautics

X 0 > (D
11I .c
-H
o 0
(0 c ^
o ~3 L.
T3 in D
0 13 u
C C L- a
— (0 (0 O
to 0
+- o c. in
a -a Z
O
•s
0
c
O
in x O in
E ^ N in
(0
L. ^
cn z
o
oo
o 4- in
u 5
0 O O *
CL JZ L. u
co in 4- o
173
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

SOLID PHASE GRANULAR HOT PRODUCT GASES


(NO REACTION) DIFFUSION (NO MORE HEAT
FLAME RELEASE) T,

/ x-———
:RATURE

Ul

MASS BURNING fc
RATE =>
Q_ /
Ul
Ts
Z
UJ _^x^0 ^L ——

DISTANCE FROM BURNING SURFACE

Fig. 14. Theoretical model of "steady-state" flame showing


profile of time-average temperature.

1.50

1.00
0.80
80=20 FINE
0.60 A 75=25 FINE -
x 80=20 COARSE
O 75:25 CQARSF

20 40 60 80 100 200 400 600 800 1000 2000


PRESSURE,PS IA

F i g . 15. BurnI ng-rate-pressure curves for four NH4CI04 - PI3


propel lants. (Sutherland's experiments.)

174
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

5000

4000

3000

2000

1000

80 120

Fi g. 16. Comparison of Sutherland's b u r n i n g rate data with


Summer-field theory. (Theoretical equation appears
as straight l i n e on this type of plot.)

175
Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

I I
EXPERIMENTAL CURVE
THEORETICAL CURVE

co
O

o
-o
Purchased from American Institute of Aeronautics and Astronautics

70
O
O

x 75-25 COARSE
A 75-25 FINE
70
D 80-20 COARSE n
O 80-20 FINE

400 600 1000 2000

GAS PRESSURE PS.I.A.

Fi q. 17. Burning-rate-pressure curves for four propellants.


( W e b b ' s experiments.).
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

7000

6500 L
6000

5500

5000

4500
<£•>
-^

4000

L 3500

3000

2500

2000

1500

1000

500

0 20 40 60 80 2 / JOO 120 140 160 180


P
Fiq. |8. Comparison of Taback 1 s particle size data w i t h theory.
(Theory gives straight lines.)

177
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

C
13
o
L_
cr>

c
O

o
or (/) CD
o — N
*0 —
L.
CD 0)
N >

a: L.
g fO
Q- (O
E

CL (0
ro CD
L- v^
en
O •
CD <•

(\J O CO CD ^ Od O

~nVAH31NI~NOdDllN 3NO d3d 1H9I3M 1V101 dO !N3Dd3d

178
Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

O
6

>
Purchased from American Institute of Aeronautics and Astronautics

O
n

n
FINE-50%
VlERY COAR5E
(149-29,

200 400 500 1500 2000


PRESSURE (PSIA)

Fig. 20. Effects of particle size on burning-rate-pressure


curves, for three bimodal distributions. (Taback's
experi ment.)
Purchased from American Institute of Aeronautics and Astronautics

SOLID PROPELLANT ROCKET RESEARCH


Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

//.
o/ 1
xv IOO°X . VERY FINE
35
™ (' • 7 / * )
50% VERY FINE;
o—50% COARSE
V
(53- 104 M)
A_._25% VERY FINE / /7 /°
\j\j ^ 75% COARSE / /' /
'2 50%VERY
x a—• FINE -50% /
/ H~ SLOPE =b= 26.6
NTERCEPT =a=200
o VER^ ' COARSE
^ (I49M •297M) / f SLOPE =b= 24.2
25 o^ / / -
tf yU. X^-SLO
•/ SJTERCEPT =a=!95

//; » /-
D
z E = b=36.(
>N f> INTEF ?CEPT = a = l
o
//
<

20 fi—————— /

//
xx
x /
/
/
/
/
^ , /

/
15
1——/
I A
? ///
/^ // x
- SLOPE =b = 10.1
INTERCEP1 =a=300

10 /
PRESSURE

/« -
/« EFFECT OF 1'ARTICLE <SIZE ON
TlHE CONST/S NTVa"!>" IN THE
x
\

Bl RNING RA TE LAW
5 _L=
**k 1
f FC R 77%rN H 4 CI04- 2;>% POLY-
E$ TER RESIh1 SOLID Pf ROPELL-
M JT
\ \ i i
40 80 120 . 160 200
PRESSURES (PSIA)2/3

Fig. 2L Comparison of Taback's particle size data with theory.


(Theory gives straight lines.)

180
PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.

Purchased from American Institute of Aeronautics and Astronautics

EFFECT 0
ON
NH 4 CI04
RF^IN

o
-o
m

o
n

15 30 60 120 250 500 1000 2000


PRESSURE (PSIA)
Fig. 22. Effect of copper chromite combustion catalyst on
burning rate with catalyst. (Theory gives straight.
Iines.)
Downloaded by PENNSYLVANIA STATE UNIVERSITY on November 22, 2015 | http://arc.aiaa.org | DOI: 10.2514/5.9781600864766.0141.0182 | Book DOI: 10.2514/4.864766

PRESSURE / BURNING RATE (PSIA/IN

00
Purchased from American Institute of Aeronautics and Astronautics

CO

You might also like