You are on page 1of 7

Downloaded from http://iranpaper.

ir
http://www.itrans24.com/landing1.html

Heat Mass Transfer (2008) 44:579–585


DOI 10.1007/s00231-007-0280-5

ORIGINAL

Numerical studies on erosive burning in cylindrical solid


propellant grain
K. Srinivasan Æ S. Narayanan Æ O. P. Sharma

Received: 11 November 2006 / Accepted: 12 April 2007 / Published online: 16 May 2007
 Springer-Verlag 2007

Abstract This paper addresses erosive burning of a A+ damping constant in van Driest’s
cylindrical composite propellant grain. Equations govern- hypothesis (26)
ing the steady axisymmetric, chemically reacting boundary C1 fi C4,Cl, Cx constants in turbulence models (C1 =
layer are solved numerically. The turbulence is described 1, C2 = 1.3, C3 = 1.57, C4 = 2,
by the two equation (k-e) model and Spalding’s eddy break C = 0.18, Cl = 0.09)
Px
up model is employed for the gas phase reaction rate. The Cp Yk Cpk average heat capacity of
k
governing equations are transformed and solved in the reacting gases, (1.254 kJ/kg K)
normalized stream function coordinate system. The results Cpk heat capacity of kth species (kJ/kg K)
indicate that the dominant reaction zone lies within 20% of Cs heat capacity of solid propellant,
the boundary layer thickness close to the wall. The sharp (1.59 kJ/kg-K)
gradient of the temperature profile near the wall is found dAp average diameter of ammonium
responsible for bringing the maximum heat release zone per-chlorate particles
near the surface and hence enhancement in the burning D port diameter of rocket motor (m)
rate. The model reproduces the experimental observation Df diffusion coefficient in Fick’s law
that erosive burning commences only above a threshold (m2/s)
value of axial velocity. Eas activation energy in propellant
surface decomposition,
(62.7 kJ/kmole)
List of symbols Dhof,k heat of formation of kth species,
a pre exponent in strand-burning rate (233.662 kJ/kg for fuel, –3937.56 kJ/
law, (0.245 · 10–2 (m/s)/(Mpa)n) kg for oxidizer, –4753.914 kJ/kg for
A cross-sectional flow area products)
As Arrhenius frequency factor in k von Karman’s constant (0.41)
propellant surface decomposition, K u0i u0i =2; turbulent kinetic energy
(5.65 m/s) (m2/s2)
‘ mixing length (m)
K. Srinivasan (&)  S. Narayanan n exponent in strand-burning rate law
Department of Mechanical Engineering,
Indian Institute of Technology Madras,
(0.41)
Chennai 600036, India P pressure (Pa)
e-mail: ksri@iitm.ac.in Pr Cp l/k, Prandtl number based upon
molecular properties of fluid
O. P. Sharma
Department of Aerospace Engineering,
Prt Prandtl number for turbulent flow
Indian Institute of Technology Kanpur, (0.9)
Kanpur 208016, India r coordinate in radial direction (m)

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

580 Heat Mass Transfer (2008) 44:579–585

rb total burning rate of a solid propellant c constant-pressure to constant-volume


(m/s) specific heat ratio (1.26)
rbo apn strand burning rate of a solid k thermal conductivity of gas
propellant (mm/s) (kJ/m s K)
R port radius of rocket motor (m) ks thermal conductivity of solid
Rh roughness height (m) propellant (kJ/m s K)
Ru universal gas constant (J/kmol k) l gas viscosity (kg/m s)
Sc l=qDi ; Schmidt number based upon leff l + lt, effective viscosity (kg/m s)
molecular properties of fluid lt turbulent viscosity (kg/m s)
Sct Schmidt number for turbulent flow (l/Pr )eff l/Pr + lt/Prt (kg/m s)
T temperature (K) (l/Sc )eff l/Sc + lt/Sct (kg/m s)
Tci initial centerline temperature (K) mk number of moles of kth species
To reference temperature, 298.14 K (1 kmol for fuel, 3.23 kmol for
Qs,ref surface heat release due to pyrolysis oxidizer, 5.9 kmol for products)
at reference temperature (J/kg) q gas density (kg/m3)
Tp propellant temperature (K) qs solid propellant density (1600 kg/m3)
Tpi propellant initial temperature, s leff ou=oy; local shear stress (N/m2)
(298 K) xk rate of production of species k due to
Tps propellant surface temperature, chemical reactions (kg/m3 s)
(800 K)
Toi initial stagnation temperature (K) Subscripts
Tps reference surface temperature of b bulk or averaged variable
propellant (K) c centerline condition
u gas velocity in x-direction (m/s) k species index representing fuel gas [F], oxidizer gas
U axial velocity outside boundary layer [O], and product gas [P].
(m/s) ¥ free stream condition
Uci initial centerline velocity (m/s) w wall (propellant surface) condition
u* qffiffiffiffiffi
sw
q1 ; friction velocity (m/s)
v gas velocity in y-direction (m/s)
 1
W P
Yk =W average molecular 1 Introduction
k
weight of gases (kg/kmol) Erosive burning is defined as the augmentation of the
Wk molecular weight of kth species, burning rate by the flow of gases across a burning surface.
kg/kmol (30 kg/kmol for fuel, On ignition, the surface of the solid propellant pyrolyses
27.9 kg/kmol for oxidizer, giving out gaseous fuel and gaseous oxidizer which mix
20.4 kg/kmol for products) with each other above the propellant surface, react and
x coordinate in axial direction (m) form predominant diffusion flames. The burning rate of a
y coordinate normal to propellant composite solid propellant is enhanced by the flow of gases
surface (m) along surface. This phenomenon is called erosive burning.
Yk mass fraction of k-th species The combustion phenomenon occurring inside the solid
YFS mass fraction of fuel in a composite rocket motor is highly complex involving heat, mass and
solid propellant (0.25) momentum transfer in a multicomponent as well as mul-
YOS mass fraction of oxidizer in a tiphase mixture.
composite solid propellant (0.75) Corner [1] attempted a theoretical model to explain the
ðÞ time-averaged quantity enhancement of burning rate of an organic substance and
ð~Þ Favre averaged quantity found that the burning rate increases with increase in
ðÞ0 fluctuating quantity velocity. This was attributed to the increase in transport
d boundary-layer thickness (m) properties by turbulence. Green [2] predicted the existence
e of a threshold velocity only above which the burning rate
lu0ij u0ij =q; turbulent dissipation
strongly depends on velocity. Lenoir and Robillard [3] based
(m2/s3) on mathematical models predicted enhancement in burning

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

Heat Mass Transfer (2008) 44:579–585 581

rate and attributed it to the added heat flux from the core flow To predict the erosive burning rate and its dependence on
through the flame to the surface. Klimov [4] used vorticity other parameters it is necessary to solve the boundary layer
transport equation to model erosive burning. Lengelle [5] equations, inside the hollow cylindrical propellant. The
considered the effect of pressure and velocity fluctuations on formulation and the underlying assumptions have been
the burning rate of Ammonium Perchlorate propellant using presented by Kuo [13] and Razdan and Kuo [10], and are
an analytical description of turbulent boundary layer on a repeated here for convenience: (1) the boundary layer is
flat plate with injection. He obtained a linearized response of axisymmetric, quasi-steady, chemically reacting, (2) body
the propellant to pressure and velocity fluctuations. He forces are absent, (3) radiative heat transfer is negligible,
found that the pressure response is strongly amplified when (4) Lewis number is unity, (5) Fick’s law of diffusion is
the erosive effect becomes more pronounced. Yamada and valid, (6) the chemical reaction is represented by a single
Goto [6] simulated erosive burning of solid rocket motors, step forward reaction, (7) turbulence chemistry interaction
0
using channels with porous walls. They concluded that there qu0 v0 and l0 ou
is neglected, (8) q0 u0 v0 \\ o
u
oy \\l oy ; (9) the
exists an intimate relation between turbulence intensity and decomposition of solid propellant takes place in a thin layer
rate of heat transfer and hence erosive burning rate. Beddini of negligible thickness, (10) heat conduction in the solid
[7, 8] presented a theoretical analysis of erosive burning phase is negligible in the x-direction compared to
using a reacting turbulent boundary layer approach and y-direction, and (11) the reacting gas mixture is ideal.
observed that for constant flow velocity, burning rate de-
creases along the propellant surface. 2.1 Governing equations
Vilyunov et al. [9] studied erosive burning from the
view point of the gasdynamic vibrations of the flowing gas. The equation of mass conservation is given by Eq. 1:
Razdan and Kuo [10] studied the erosive burning of
composite propellants by solving a steady, two-dimen- o o
ðr
qu~Þ þ ðr
qv~Þ ¼ 0 ð1Þ
sional, chemically reacting, and turbulent boundary layer ox oy
over the propellant surface. They found that the increase in
free stream gas velocity brings the location of peak tur- The equation of momentum conservation is given by
bulent intensity and heat release zone closer to the pro- Eq. 2:
pellant surface, and concluded that it results in increasing  
o~
u o~
u 1o o~
u op
the burning rate. Mukunda and Paul [11] found out a rel- q~
u þ q~
v ¼ rleff  ð2Þ
ox oy r oy oy ox
atively simple nondimensional relationship between the
ratio of the actual to nonerosive burn rate that matched well
with the experiments. It was concluded that the correlation
may be adopted universally for most practical propellants. (a) CYLINDRICAL
Godon et al. [12] studied the erosive burning of ammonium PROPELLANT
perchlorate inert binder propellants both experimentally
and from modeling viewpoint. A correlation law was ob-
tained for shear stress.
Y
In the present work, the reacting turbulent boundary
layer equations are solved to model the augmentation in
X
heat transfer to the propellant surface due to forced con-
vection and the enhancement in the burn rate of the pro-
pellant due to this increase. The shift in the location of an (b) Axis of symmetry

equivalent flame zone which comprises the different


flamelets in the burning of composite propellant, due to the y
change in the flow velocity over the burning surface has HOT COMBUSTION r
GASES
been considered. The composition of the propellant con-
sidered is 75% Ammonium Perchlorate and 25% polybu-
tadiene acrylic acid (PBAA)/Epoxy resin (EPON).
PROPELLANT X

2 Numerical methodology

When hot gases flow over a propellant surface, a chemi- Fig. 1 a Schematic of the cylindrical propellant, b schematic of
cally reacting boundary layer develops, as shown in Fig. 1. boundary layer flow over the propellant

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

582 Heat Mass Transfer (2008) 44:579–585

Here, the variable y is in opposite direction to r, along Y~o;1 : These have been arrived at using the mass flow
the concave normal of the propellant surface. In addition, balance at the solid–gas interface given by
the species conservation (Eq. 3) is as follows:
oY~i
  qv~Y~i Þgas  qs rb Yi;s  ð
ð qDf Þ ð9Þ
~
oYi ~
oYi 1 o  l  oYi ~ oy gas
q~
u þ q~
v ¼ r _ i
þx ð3Þ
ox oy r oy Sc eff dy
Further, the erosive burning rate is obtained from the
energy balance at the interface as follows:
The subscript ‘i’ indicates the fuel species when set to ‘F’,
oxidizer species when set to ‘O’, and product species when
oT~ oTS

set to ‘P’. k g ¼ ks jS þ qs rb ðCP  CS Þ TS  TS;ref þ QS;ref
oy oy
The conservation of energy is:
ð10Þ
 
oT oT 1 o  l  oT 
q~
u þ q~
v ¼ r where, ks oT
oy js ¼ TS  TS;i qS rb CS and rb is obtained from
S
ox oy r oy Pr eff oy
the surface temperature using Arrhenius law of surface
   2    
l ou u dp 1 X o  E
pyrolysis rb ¼ As exp  Rua;sTs
þ þ  Dhf ;k xk ð4Þ
Cp eff oy Cp dx Cp
The boundary conditions for turbulent kinetic energy
and dissipation are given as follows: Neumann boundary
For turbulence closure, K–e equations are used which
conditions are applicable at the free stream, i.e.,
are given in Eqs. 5 and 6. oK oe
oy ¼ oy ¼ 0 at y=d. Near the wall the boundary conditions
     2 are prescribed as per Razdan and Kuo [10]
oK oK 1 o lt oK o~
u
q~
u þ q~
v ¼ r lþ þ lt  qe
ox oy r oy C1 oy oy  2
½kDv ðy þ DyÞ2 o~
u
ð5Þ K¼ pffiffiffiffiffiffi ð11Þ
Cl oy
   
oe oe 1 o l oe 3
q~
u þ q~
v ¼ r lþ t o~
u
ox oy r oy C2 oy e ¼ ½kDv ðy þ DyÞ2 ð12Þ
 2 oy
o~
u e e2 h i hpffiffiffiffiffiffi
þ C3 lt  C4 q ð6Þ
oy K K where, Dv ¼ 1  exp  ðyþDyþ
A l
Þ
qu  s
sw ; Dy ¼ 0:9 q
l
u Rþ h

2
 Ke and Cl is a constant.
where lt ¼ Cl q Rþ
h expð 6 Þ;
h
and Rþ
h ¼ ðq
u Rh Þ=l:
Further, the equation of state is used The rate of production of fuel species is given by
pffiffiffiffi Y~
 K oor
Lockwood [14] as x_ F ¼ Cx q F
; where cx is con-
qRu T~
P¼ ð7Þ stant, based on Spalding’s Eddy-Break-Up (EBU) model.
W
The coefficients of momentum, mass and energy are given
l l lt
The governing equation in the freestream (inviscid) by the relations leff ¼ l þ lt ; Sc ¼ þ and
  eff Sc Sct
region is the Euler’s equation: l l lt
Pr eff ¼ Pr þ Prt : The molecular viscosity l is given by
pffiffiffiffiffi 0:65
dUc d
p 8
the relation l ¼ 8:7  10 WT [10]. Prandtl number
qc Uc ¼ ð8Þ
dX dx c
Pr is calculated from the formula Pr ¼ ð1:77c0:45 Þ [10].
where the
pressure gradient
 is obtained from the equations
A dp
¼ 2pRs þ d
q AU 2 d
and dx ðqb AUb Þ ¼ 2pRqs rb : 2.3 Solution procedure
dx w dx b b
The isentropic assumption in the core flow enables the
calculation of
centerline stagnation In order to overcome the typical difficulties of solving the
temperature of the core equations in the axisymmetric (x–y) coordinates, the
flow To;c ¼ Tc 1 þ 12 ðc  1ÞMc2 :
equations are first transformed to von-Mises coordinates
2.2 Boundary conditions (x–w) (where w is stream function) and then to Patankar–
Spalding coordinates (x – x ), where x ¼ www I
Subse-
E wI
The boundary conditions include (1) At the wall, quently, the equations are solved in this coordinate system,
ðy ¼ 0Þ; u~ ¼ 0; v~ ¼ qqs rb ; T~ ¼ Ts (2) At the boundary layer using Patankar and Spalding’s space marching technique
g
edge ðy ¼ dÞ; u~ ¼ U1 ; T~ ¼ T1 ; Y~F ¼ Y~F;1 ; and Y~o ¼ [15]. The overall solution algorithm involves prescription

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

Heat Mass Transfer (2008) 44:579–585 583

of starting profiles, computation of transport coefficients,


source terms, boundary values using the wall and core flow
regions, and obtaining the solutions of u, Yi¢s, T, K, and e
for a particular station. If the values of the variables at a
particular station do not simultaneously satisfy the equa-
tions, modifications are effected and iterated until conver-
gence.

3 Results and discussions

The starting profiles were as follows: The velocity profile


was prescribed as per Cole’s Law of the wall/Law of the
wake, given by
  p y 
u 1 us d 2p
¼ ln þ B þ sin2 ð13Þ
us k m1 k 2d

where k is Karman’s constant. The temperature profile was Fig. 3 Variation of centerline static pressure with axial distance
prescribed as a piecewise linear profile, varying from sur-
face temperature (800 K) to 2016 K at x = 0.2, and then flow inside the rocket motor is accelerated by the strong
to the starting free-stream value of 2235 K at x = 1. The favorable pressure gradient.
species profiles were also assumed linear. The k and e The species profile is shown in Fig. 4. The species
profiles were prescribed as per Chambers and Wilcox [16]. profile of the products can be inferred from the relation
The dimensionless velocity (U/Uci) variation within the YP = 1 – (YF + YO). The fuel and oxidizer species steeply
boundary layer is shown in Fig. 2. The comparison between decrease near the boundary indicating that the dominant
the present study and that of Razdan and Kuo [10] is made reaction zone lies within 20% of the boundary layer
at (X/D = 15). Similar trend is seen. The flow acceleration thickness close to the wall. The use of eddy break up model
is remarkably reflected in the profile, strongly influenced by brings in the role of turbulent kinetic energy in determining
the axially nonlinear pressure gradient shown in Fig. 3. The the reaction zone. It is also noted that although the initial
turbulent nature of the flow is evidenced by the steepness of profiles of fuel and oxidizer species prescribed were linear,
the velocity gradient at the wall. The velocity gradient near the resultant profiles downstream are predicted realisti-
the wall increases with the downstream distance, since the cally, showing a sharp gradient near the wall and a gradual

Fig. 4 Mass–fraction profiles of fuel and oxidizer in the boundary


Fig. 2 Velocity Profile in the reacting boundary layer layer

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

584 Heat Mass Transfer (2008) 44:579–585

decrease near the edge of the boundary layer. The profile of


product species can be easily inferred from the plots of
oxidizer and fuel species, and reveals a sharp increase in
the reaction zone.
The dimensionless temperature (T/Tc) profile is shown
in Fig. 5. The temperature shoots up in the region near to
the wall, in conformance with the reaction zone indicated
by the species profile shown in Fig. 4. Beyond the domi-
nant reaction zone (0.2 < x < 1), the temperature rise is
gradual, as it attains the free stream value. The sharp gra-
dient of the temperature profile near the wall is responsible
for bringing the maximum heat release zone near the sur-
face and hence enhancement in the burning rate. It may be
expected that the increase in turbulence near the surface of
the propellant plays an important role in the erosive
burning mechanism.
The erosive burning variation with velocity is shown in Fig. 6 Erosive burning rate relation with centerline velocity
Fig. 6. The burning rate enhancement is seen by comparing
the erosive burning rate to the burning rate that occurs at
the same local pressure for no cross flow (rb/rb0). Using the flow depends not only on temperature in accordance with
strand burning law with values given in the nomenclature, the Arrhenius law but also on its fluctuations’’. Based on
the quantity rb0 is 5.6 mm/s. At the initial locations, neg- this argument, they showed that the most favorable con-
ative erosion is observed. This may be due to the ditions for negative erosion develop at the laminar-turbu-
quenching effect due to developing turbulent structures. lent transition. In the present study, velocity increases with
However, this effect quickly diminishes with increasing increase in X/D, and when it exceeds 230 m/s positive
axial distance, until positive erosion is reached. Vilyunov erosion is observed. This is due to the increase in peak
and Dvoryashin [17] also observed negative erosion and value of heat generation rate and its location getting closer
reasoned that ‘‘the average reaction rate in the turbulent to the propellant surface, corroborated by the steep slope in
the temperature profile. This is responsible for the in-
creased burning rate and is in conformance with the
observations of Vilyunov and Dvoryashin [17]. When the
initial centerline velocity was 200 m/s, pressure 7.5 MPa,
stagnation temperature 2250 K, erosive burning starts be-
yond 230 m/s.

4 Conclusions

The problem of erosive burning in a cylindrical solid


propellant grain is investigated numerically in this work by
solving the axisymmetric reacting turbulent boundary layer
inside the propellant grain. The velocity gradient near the
wall steepens with downstream distance, since the flow
inside the rocket motor is accelerated by the strong
favorable pressure gradient. The species profiles indicate
that the dominant reaction zone lies within 20% of the
boundary layer thickness close to the wall. The sharp
gradient of the temperature profile near the wall is
responsible for bringing the maximum heat release zone
near the surface and hence enhancement in the burning
rate. At initial locations, negative erosion occurs, and po-
Fig. 5 Temperature profile in the boundary layer at various axial sitive erosive burning commences only when the flow at-
distances tains a threshold velocity.

123
Downloaded from http://iranpaper.ir
http://www.itrans24.com/landing1.html

Heat Mass Transfer (2008) 44:579–585 585

References 10. Razdan MK and Kuo KK (1982) Turbulent flow analysis of


erosive burning of cylindrical composite solid propellants. AIAA
1. Corner J (1947) The effect of turbulence on heterogeneous J 20:122–128
reaction rates. Trans Faraday Soc 43:635 11. Mukunda HS and Paul PJ (1997) Universal behavior in erosive
2. Green L Jr (1954) Erosive burning of some composite solid burning of solid propellants. Combust Flame 109(1–2):224–236
propellants. Jet Propulsion 24:9 12. Godon JC, Duterque J, Lengelle G (1993) Erosive burning in
3. Lenoir JM, Robillard G (1957) A mathematical method to predict solid propellant motors. J Propulsion Power 9(6):806–811
the effects of erosive burning in solid propellant rocket, sixth 13. Kuo KK (1986) Principles of combustion. Wiley, New York
international symposium on combustion, pp 667–683 14. Lockwood FC (1977) The modeling of turbulent premixed and
4. Klimov AM (1975) Erosive burning of propellants. Combust diffusion combustion in the computation of engineering flows.
Explosion Shock Waves 11(5):678–681 Combust Flame 29:111
5. Lengelle G (1975) Model describing the erosive combustion and 15. Spalding DB (1977) GENMIX: a general computer program
velocity response of composite propellants. AIAA J 13:315 for two—dimensional parabolic phenomena. Pergamon Press,
6. Yamada K, Goto M (1976) Simulative study on the erosive New York
burning of solid rocket motors. AIAA J 14:1170 16. Chambers TL and Wilcox WC (1977) Critical examination of
7. Beddini RA (1978) Reacting turbulent boundary layer approach two-equation turbulence closure models for boundary layers.
to solid propellant erosive burning. AIAA J 16:898 AIAA J 15:821
8. Beddini RA (1980) Aerothermochemical analysis of erosive 17. Vilyunov VN, Dvoryashin AA (1971) An experimental investi-
burning in a laboratory solid rocket motor. AIAA J 18:1346 gation of the erosive burning effect. Combust Explosion Shock
9. Vilyunov VN, Isaev Yu M, Revyagin LN (1981) Erosive burning Waves 7(1):38–42
of condensed materials in an acoustic field. Combust Explosion
Shock Waves 17(4):383–386

123

You might also like