You are on page 1of 13

Computers & Fluids 75 (2013) 22–34

Contents lists available at SciVerse ScienceDirect

Computers & Fluids


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m p fl u i d

Analysis and design optimization of solid rocket motors in viscous flows


Mine Yumusßak ⇑
ROKETSAN Missiles Industries, Ankara 06780, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: The objective of this study is to develop a design tool that can be used in viscous flows. The flow analysis
Received 28 August 2012 is based on the axisymmetric Navier–Stokes and k–e turbulence equations. These coupled equations are
Received in revised form 1 January 2013 solved using an explicit finite difference method. The accuracy of the analysis code is validated for viscous
Accepted 5 January 2013
flows in solid rocket motor combustion chamber and nozzle. The gradient-based numerical optimization
Available online 31 January 2013
model is used to maximize the thrust of solid rocket motor under a constraint of propellant weight. The
sensitivity analysis that measures the response of the flow with respect to a geometry perturbation is cal-
Keywords:
culated by finite differencing. The optimization of design study employs a commercial optimization pack-
Design optimization
Viscous flow
age. The performance of design optimization method is tested in solid rocket motor combustion chamber
Turbulence and nozzle design.
Solid rocket motor Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction pressure and friction losses of internal and external flows may in-
crease the range or payload capacity of rockets. Vehicle require-
Computational Fluid Dynamics (CFD) is extensively used in so- ment such as performance characteristics, envelope constraints
lid rocket motor (SRM) designs. In traditional design methods, and mission profile play important role in combustion chamber de-
engineers modify initial prototypes with their experience and ana- sign. Grain design and burning surface evaluation are very impor-
lyze the modified configurations with CFD solvers. In general, these tant in the design of combustion chambers. The essence is to
modifications are based on the trial and error approach and several evolve the burning surface and establish a relation between web
configurations have to be analyzed before the final decision is burnt and the burning surface. The nozzle is another important
reached. Traditional methods have some disadvantages. It is diffi- component of rockets because its efficiency significantly affects
cult to determine whether the final design is the best one, and rocket’s performance. By minimizing the losses in the nozzle, the
computational time is very long. In order to reduce the computa- thrust of a rocket can be increased by keeping the grain geometry
tional time, sometimes simpler models are used for flow analyses. fixed.
However, these models may reduce the reliability of the design re- The subject of solid rocket motor (SRM) design and optimiza-
sults. In recent years, new advancements have been made to in- tion is quite vast. Many approaches from the gradient based class
crease the reliability and efficiency of design methods. Automatic of methods, non-gradient based methods, heuristics and paramet-
design methods developed by combining the CFD and optimization ric methods have been used for design and optimization processes
codes have several advantages. The experience and information of SRM system parameters and sub-components. Kamran and Gua-
needed for design can be reduced, and innovative designs can be zhu [1] present an approach using hyper-heuristic based on meta-
made in the conditions that have never been tried before. The de- heuristics method for combustion chamber, nozzle and grain de-
sign cost can also be decreased by reducing the engineers’ work- signs. In this study, design objective is to minimize the gross mass
load and design time. Another advantage of the automatic of motor that would cater for a minimum inert mass along with
methods is that the design can be obtained as a solution to the maximum specific impulse. The grain design and burning surface
optimization problem. In this way, if there is no solution with evolution of a complex 3D solid rocket propellant grain are mod-
the specified design conditions, the design can be evaluated as a eled by CAD software using parametric modeling. Anderson and
best available solution. Burkhalter [2] used genetic algorithms to design solid rocket mo-
The performance of a rocket can be improved with aerodynam- tors as a component within an overall missile system. In this study,
ics and aero-thermodynamics design optimization. Minimizing the multiple goals, such as maximized range, minimized g-loading,
minimized takeoff weight and maximized fuel volume are used
⇑ Address: Internal Ballistics Modeling Department. Tel.: +90 312 860 55 00; fax: to test the ability of genetic algorithms to work efficiently within
+90 312 863 42 08. a multidisciplinary framework. The 0-dimensional ballistics
E-mail address: myumusak@roketsan.com.tr performance prediction code is coupled with the design tool to

0045-7930/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compfluid.2013.01.011
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 23

determine the most efficient configuration of the motor. Kumar sists in the mean flow, even if the flow is highly turbulent over the
et al. [3] carried out a parametric study to examine the geometrical duct cross section.
influence on turbulent separated flows in SRMs. Five different Several numerical simulations have been performed to exam-
physical models with different port geometries are examined. Re- ine the flow-field within a rocket combustion chamber. Sabnis
sults of interests such as reattachment length, size of the recircula- et al. [14] and Tseng [15] used conventional single-point turbu-
tion zone and the axial velocity variations are reported to illustrate lence closure schemes at low injection rates and obtained a good
the influence of transition region on the flow characteristics of tur- comparison of the mean flow field with the experimental data,
bulent mixed convection downstream of a solid rocket motor with but over predicted turbulence intensity levels within the cham-
divergent port. Such detailed results are needed for an integrated ber. Apte and Yang [16] conducted two dimensional simulations
design and optimization of the high-performance solid rockets port of flows through nozzleless rocket motors at very high injection
geometry and its allied igniters with confidence. He concluded that rates to achieve a better comparison of turbulence intensity and
the narrow port and long flow development ahead of the steep axial velocity profiles compared with the second-order turbu-
divergence are shown to favor flow separation, which might lead lence closures. Apte and Yang also performed a Large Eddy Sim-
to high peak pressure, pressure-rise rate and thrust oscillations ulation (LES) of internal flow development in a three
during the starting transient period of operation of motors with dimensional rectangular rocket motor. Nicaoud et al. [17] per-
divergent ports. formed Direct Numerical Simulation (DNS) at high injection rates
The studies related to the design of rocket nozzles are not new. in an attempt to reproduce flow conditions representative of a
Optimization techniques used to design nozzle contours have been SRM in order to investigate the effect of a high blowing rate
utilized since 1950. Rao [4] developed a method which optimizes a on the wall layer.
rocket nozzle contour for a given length or expansion ratio so as to The objective of this study is to develop a design tool that can
achieve maximum thrust. Rao’s method is based on the assump- be used for rocket combustion chamber and nozzles. The reli-
tion of inviscid isentropic flow. This method has been used in many ability of design results depends on the accuracy of flow model
studies for different classes of rocket nozzle design. In a study by used in design method. In order to capture the viscous flow
Farley and Campbell [5], three Rao optimized bell contour nozzles physics, the axisymmetric Navier–Stokes and k–e turbulence
were compared to a 15° conical nozzle. The thrust produced by the equations are solved simultaneously. The classical k–e turbulence
optimized nozzles was greater than that obtained with the conical model is modified to predict the mean-flow velocity profile and
nozzle. In fact, bell contour nozzles have been used routinely for turbulence intensities accurately near the mass injecting propel-
many years in large liquid rocket engines [6]. Conical nozzles are lant surface and wall regions. A gradient-based finite difference
typically used only when fabrication and design costs outweigh numerical optimization method is used for design optimization.
performance. In recent years, more accurate design methods have In gradient based design optimizations, the derivatives of objec-
been developed by using Navier–Stokes equations [7,8]. tive function with respect to design variables are needed. In lit-
The flow physics in a SRM is quite complex. In a SRM, the erature, these derivatives are called sensitivities. The accurate
flow accelerates from the head-end, resulting in streamwise and efficient calculation of sensitivities is important for the per-
inhomogeneity, undergoing transition to turbulence in the mid- formance of design method. There are two methods to obtain
section of the motor, becoming fully turbulent further down- sensitivity derivatives. The first is the analytical approach, which
stream and finally reaching supersonic conditions at the exit of calculates sensitivities by analytically differentiating the govern-
the nozzle. The velocity profile and turbulence characteristics ing equations with respect to the design variables. The analytical
near surfaces determine the convective heating loads and there- method provides accurate and efficient sensitivity calculations.
fore affect both insulation thickness requirements and the poten- However using this method is not easy; developing a sensitivity
tial erosive burning of the propellant. Early attempts to code requires considerable amount of programming effort. The
investigate the internal flow-field of a SRM were mainly con- analytical method is more efficient if the analysis code uses an
cerned with the distribution of the mean velocity. In a SRM, near implicit numerical solution method. The other method is the fi-
the head-end, the flow is laminar and its velocity profiles can be nite difference approach, which requires the flow field solution
determined by the laminar similarity theory derived by Culick for each perturbed design variable. The main advantage of the fi-
[9] and Taylor [10]. For uniform injection at a constant mass nite difference approach is that it is much easier to implement
flow rate, and for laminar and incompressible flow, the laminar than the analytical approach. It does not require an additional
similarity velocity profile was given as: programming effort to build a dedicated sensitivity code. Even
though the finite difference approach requires flow solution eval-
"   # uations for the perturbation of each design variable, the compu-
u x p p r aþ1 tational cost can be reduced if the geometry perturbations are
¼ 1 a cos
uinj rw 2 2 rw small and the flow solution iterations for each perturbed geom-
"   # ð1Þ
v rw  a p r aþ1 etry start from an already converged base solution. However, the
¼ sin finite difference sensitivity calculations have accuracy problems.
uinj r 2 rw
The accuracy of the finite difference sensitivity derivatives de-
pends on the choice of the magnitude of the design variable per-
where the value of a is 0 for planar and 1 for axisymmetric flows. In turbations or the finite difference step size. However, the
the equation above, uinj is the injection velocity on propellant sur- accuracy of the finite difference sensitivities can be improved
face, u is the axial velocity, v is the radial velocity, x is the axial by finding the optimum step size that can minimize the norm
direction, r is the radial direction and rw is the propellant port value of total error in sensitivities [18]. In the present study,
radius. sensitivities are calculated using the finite difference method
The experiments performed by Dunlap et al. [11] and Yamada and numerical design model uses a commercially available con-
et al. [12] showed that the mean flow field is accurately repre- strained optimization package DOT [19]. It searches for a feasible
sented by Culick’s laminar velocity profile in the forward region direction and step size to minimize or maximize a specified
of a cylindrical port rocket chamber. Later, Traineau et al. [13] con- objective function under a set of constraints. The performance
ducted an experimental study on a nozzleless rocket motor at high of design optimization model is demonstrated for solid rocket
injection rates and concluded that the laminar velocity profile per- motor combustion chamber and nozzle designs.
24 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

2. Flow model Table 1


Functions in the Shima model [25].

To model viscous flows in the combustion chamber and the Functions Expressions
nozzle, the axisymmetric Navier–Stokes and k–e equations are Pii 2qsij @u i
@xj
solved simultaneously. The coupled equations in generalized coor- u1 1.5 A (Pii/2qe  1)
dinates can be written in non-dimensional form as below:

@b
F ðW b W
c Þ @ Gð cÞ @ b b v ðW
cÞ @ G
F v ðW cÞ
þ   b ¼0
þH ð2Þ
@n @g @n @g implemented in the code. The algorithm is verified for the
various inviscid and viscous flow problems in SRM [21,22].
where W c is the conservative vector, b
F; b b G
F v and G; b v are the flux vec-
A finite rate chemistry model version of the algorithm is also avail-
b is the axisymmetric source vector.
tors in n and g direction, H able [23].
 
c ¼ 1 ½ qquqv qeqkqe  Fb ¼ 1 qU quU þ nx Pqv U
To account for the near-wall flow effect, the low-turbulent Rey-
W nolds number model having the same form as the Jones and Laun-
J J ðqe þ PÞU qkU qeU
    der [24] near-wall model is used however, it is only applied on the
b¼ 1 qV q uV þ g x P q v V þ g y P b ¼1 qv quv qvv
G H solid wall. No damping function is applied on the propellant sur-
J ðqe þ PÞV qkV qeV rJ ðqe þ PÞv qkv qev face. The following damping functions (fl of lt and f2), and the con-
2 3
qU quU þ nx Pqv U þ ny P stants (D1 and D2) are introduced:
6 ðqe þ PÞU 7  
b 16 7 2:5
Fv ¼ 6 l 7
fl ¼ exp ; f 2 ¼ 1  0:3 exp½ðRet Þ2 ; D1
J64 ðl þ lt Þ @k @x
7
5 1 þ Ret =50
lt @ e
ðl þ l Þ @x pffiffiffi!2 !2
2 3 @ k @2u
qV quV þ gx Pqv V þ gy P ¼ 2l ; D2 ¼ 2llt ð6Þ
@x2 @x2
6 ðqe þ PÞV 7
6 7
bv ¼ 1 6
G l 7
J6 ð l þ t
Þ @k 7 where
4 l @r 5
lt @ e
ðl þ l Þ @r 2
Ret ¼ qk =le ð7Þ
2 3
qui uj @u @xj
i
 qe  D1 The eddy viscosity is then calculated by:
H¼4 5
 C e2 f2 qkee þ D2
~
ðC e1 þ /1 ÞC e1 ke qui uj @u@x
i
C l fl ðqkÞ2
j
lt ¼ ð8Þ
pffiffiffi!2
e
h i
~e ¼ e  2
l @ k 3:4
where C l ¼ 0:09 and f l ¼ exp 1þRe .
ð3Þ t =20
q @r Sutherland’s law is applied in computing the dynamic viscosity
l as a function of temperature (T):
In the equation above, U and V are the contravariant velocity com-
3
ponents, J is the coordinate transformation Jacobian, n and g are the lðTÞ ¼ T 2 BðT þ CÞ ð9Þ
curvilinear coordinates, nx, ny, gx and gy are the transformation met-
rics, q is the density, P is the pressure, e is the total energy per unit where C = 110.4 K and B = 1.456.10–6 kg/ms.
volume, k is the turbulence kinetic energy, e is the dissipation rate, l In Eq. (3), the values of the constant coefficients are C e1 ¼ 1:42
is the dynamic viscosity and lt is the turbulent viscosity. and C e2 ¼ 1:92.Compared to the standard k–e model, the function
An explicit finite difference algorithm is used for obtaining the u1 has been added for simulating injection induced flows by Shima
solutions to the Navier–Stokes equations. [25]. Since the value of the function can be too large in comparison
" # with the original value of C e1 , u1 is bounded by |u1| < 0.125 C e1 . This
bn bn bn bn
c n  Dt @ F þ @ G  @ F v  @ G v þ H n
c nþ1 ¼ W
W ð4Þ
has the effect of preventing the early laminarization of the flows. The
@n @g @n @g bÞ
¼Rð W function u1 depends on the second and third invariants A2 = aijaji,
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} A3 = aijajkaki, the flatness coefficient parameter A = 1  9/8
where n is the index for present time level, whereas n + 1 is the in- (A2  A3). The functions proposed by Shima [25] are listed in Table 1.

dex for next time level, Dt is the computational time step and Rð W
is the residual vector. 3. Design optimization model
For a cell vertex centered finite difference method, residual vec-
tor can then be written as: The present numerical design model employs a commercial
optimization package DOT [19]. It searches for a feasible direction
c Þ ¼ ðb
Rð W F iþ1;j  b b i;jþ1  G
F i1;j Þ þ ð G b i;j1 Þ  ð b
F v iþ1;j  b
F v i1;j Þ and step size to minimize or maximize a specified objective func-
 ðG b v i;j1 Þ þ H
b v i;jþ1  G b i;j ð5Þ tion under a set of constraints. The general optimization problem
can be written mathematically as follows:
In this study, time integration is performed using an explicit, multi-
stage Runga–Kutta solution scheme based on the work of Jameson Minimize : SðXÞ
et al. [20]. In this scheme, a different number of stages can be se- Subject to : g j ðXÞ 6 0 J ¼ 1; J
lected, but a four-stage scheme is used because of its efficiency. A X Lm 6 X m 6 X Um m ¼ 1; M ð10Þ
spatially-varying time step and implicit residual smoothing is used
to maintain stability and thus increase the convergence rate. The where S(X) is the objective function and gj(X) is the constraint
algorithm uses a second-order central difference method and re- function. The vector X is the set of design variables Xm. J and M
quires artificial dissipation terms. In order to improve shock captur- are the total numbers of design constraints and design variables,
ing and prevent odd–even decoupling, second and fourth order respectively. Equality constraints can be included in the formation
differences are used in artificial dissipation terms. Mass injection, but are not considered in the present design, because in practice,
supersonic exit, symmetry and wall boundary conditions are the favorable direction of each constraint can be pre-determined.
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 25

From nonlinear programming, it is useful to form the sible direction is the one that satisfies the constraints. Hence, a
Lagrangian: search direction Z should satisfy the usability requirement

X
J rSðXÞ  Z 6 0 ð15Þ
LðX; kÞ ¼ SðXÞ þ kj g j ðXÞ ð11Þ
j¼1 and the feasibility requirement

where kj are the Lagrangian multipliers. The Kuhn–Tucker condition rg j ðXÞ  Z 6 0 ð16Þ
for optimality requires that the Lagrangian, LðX; kÞ, must have a van-
ishing gradient at the optimum values of the design variables de- When no constraints are active, the search direction is iteratively
noted by X. That is obtained using the Fletcher–Reeves conjugate direction method
by imposing the orthogonality condition between search directions.
X
J An active constraint is the one whose value becomes zero within
rSðX  Þ þ kj rg j ðX  Þ ¼ 0 ð12Þ some numerical tolerance. When some constraints are active or vio-
j¼1 lated, a sub-optimization process is used to find the search direc-
tion. The sub optimization process minimizes the left hand side of
The gradient operator, r, is the sensitivity of a function with re-
Eq. (15) subject to the active constraints given by Eq. (16).
spect to design variables and is calculated by using finite differ-
ences. For example, the mth element rS is defined by
4. Results and discussion
@S SðX þ dX m Þ  SðXÞ
¼ ð13Þ
@X m dX m 4.1. Flow analysis validation
The optimization process starts with an initial guess of the design
The quality of a design depends on the ability to accurately pre-
variables. The design is then updated using an iterative procedure
dict the flow field. Therefore, the reliability of the Navier–Stokes
given by
code is first evaluated by comparing its analysis result with exper-
X nþ1 ¼ X n þ bZ n ð14Þ imental data. For this reason, a cylindrical port rocket chamber is
used. Dunlap et al. [26] designed an experimental apparatus in or-
where the superscript n is the iteration number, the vector Z is the der to provide a realistic simulation of rocket chamber turbulent
search direction, and the scalar b is the step size to move in the flow-field and measured the turbulence properties at various loca-
direction Z. The optimum step size is found using a one-dimen- tions. He simulated the cylindrical port rocket chambers by putting
sional search via interpolation. The process is iterated until it con- together separate segments to produce overall L/D ratios of 9.5 or
verges to an optimum. 14.3, at injection Mach numbers and Reynolds numbers typical
The search direction must be usable and feasible. A suitable of rocket motor values. Individual segment construction is illus-
direction is the one that reduces the objective function, and a fea- trated in Figs. 1 and 2. Compressed ambient temperature nitrogen

Fig. 1. Sketch of Dunlap’s cylindrical port segment [26].

r
x

rw D/2
Nozzle
Injecting wall
Fig. 2. Boundary conditions of Dunlap’s cylindrical port rocket.
26 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

is injected uniformly along the walls of 10.2 cm diameter, porous- Comparisons with the experimental data, Figs. 4 and 5, are
tube chambers connected to a choked sonic nozzle. based on the velocity and turbulent intensity profiles at several
The following boundary condition is specified on the injecting axial positions. The development of the axial velocity profile
wall: along the length of the port is shown in Fig. 4. For comparison,
the analytical solution for an incompressible laminar flow derived
by Culick, Eq. (1), is also included. At x/D = 1.8 and 3.04, the
€ ¼ 3:05 kg=m2 s; k ¼ 0:01 m2 =s2 ;
T 0 ¼ 303 K; m e ¼ 0:18 m2 =s3 velocity agrees quite well with cosine distribution. The axial
The no-slip and adiabatic wall condition is imposed on the non- velocity profiles are in good agreement with the laminar
injecting wall, where the velocity and normal pressure gradients incompressible similarity theory over most of the first 5
are zero. The turbulent quantities k and e are set to zero on the diameters (x/D 6 5.46). Further downstream, in response to mod-
non-injecting wall. For the outflow, the supersonic exit boundary erate turbulent shear, the velocity profile transits into the
condition is used. shape of fully developed turbulent flow. There is a gradual round-
The grid-independence study is performed by checking the ax- ing of the profile in the mid region and some steeping near the
ial mean-velocity and turbulence intensity at specified locations as wall.
shown in Fig. 3. Three sets of grids, 121  21 (coarse grid), Turbulence intensity distribution is shown in Fig. 5. The turbu-
223  42 (medium grid) and 362  84 (fine grid) are examined. lence intensity is small at the head end of the motor (x/D 6 5.46).
The smallest grid spacing is 1 lm near the injecting wall based The peak in the turbulence intensity increases rapidly toward the
on the fine grid. For this grid set, the centers of the cells adjacent downstream. Because the mass influx is kept constant in the pres-
to the injecting wall are located at y+  0.08 at x/D = 0 and ent study, the density decreases and injection velocity increases as
y+  1.05 near the nozzle exit region to accurately capture the the flow accelerates toward the exit. The increasing injection
near-wall phenomena. Reynolds number based on the injection velocity tends to reduce the wall-damping effect, and therefore, lit-
velocity (uinj) and the port radius, Reinj = uinjrwq/l, is approximately tle change in the vertical location of turbulence intensity peak oc-
18,000. The normalized axial velocity profiles of Fig. 3 clearly indi- curs beyond x/D = 5.46. The higher experimental levels observed in
cates that the coarse grid does not provide enough spatial resolu- the initial part of the chamber are probably due to the ‘‘pseudotur-
tion to capture the velocity and turbulence intensity profiles; bulence’’ of the injected flow; as reported by Traineau et al. [13]. As
both medium and fine grids give much better results even though the flow develops further, a strong turbulent flow regime is ob-
some small discrepancy is observed. Therefore, all of the results served. The distribution has undergone a significant change be-
presented here after have been obtained with the fine grid tween x/D = 5.46 and 6.46, where the transition to turbulent flow
(362  84). The run time for fine grid simulations is around 8 has occurred and the flow becomes fully turbulent beyond this
CPU minutes on a HP Z600 workstation machine. location.

(a) (b)
x/D = 3.04 x/D = 9.06
0.0 0.0

Grid size:
Grid size: 185x45
121x21 Grid
Grid size: 180x45
121x21
-0.2 Grid size:
Grid size: 310x90
223x42 -0.2 Grid size:
Grid size: 310x90
223x42
Grid size:
Grid size: 560x180
362x84 Grid size: 560x180
Grid 362x84
ExperimentalData
Experimental Data ExperimentalData
Experimental Data
-0.4 -0.4
r/rw
r/rw

-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
u/u1C u/u1C

x/D = 6.46 x/D = 10.3


1.0
(c) 1.0
(d)
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2 Grid
Gridsize:
size: 121x21
180x45
Grid size: 185x45
Grid 121x21
Grid
Gridsize:
size: 223 x42
r/rw

310x90
r/rw

0.0 Grid size: 310x90 0.0


Grid 223x42 Grid size: 362 x84
Grid size: 560x180 Grid size: 560x180
-0.2 Grid 362x84 -0.2 Experimental
ExperimentalData
Data ExperimentalData
Data
Experimental
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1.0 -1.0
0.00 0.03 0.06 0.09 0.12 0.15 0.00 0.03 0.06 0.09 0.12 0.15
I I

Fig. 3. Grid independence study for axial mean-velocity and turbulence intensity.
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 27

x/D = 1.80 Experimental Data


Experimental
Experimental data
Data x/D = 3.04 Experimental
ExperimentalData
Experimental Data
Data
1.0 k ε− model
k- εExperimental
ROKETSAN Solver,
Turbulence k- ε
DataModel 1.0 kk-− εmodel
εExperimental
ROKETSAN Solver,
Turbulence k- ε
Data Model
Culick’s
Laminar
Culick's solution
Solution
Solution
ROKETSAN Solver Laminar
Culick's
Culick’sSolution
Solution
Solution
ROKETSAN Solver
0.8 Laminar Soln 0.8 Laminar Soln

0.6 0.6

0.4 0.4

0.2 0.2

r/rw
0.0 0.0
r/rw

-0.2 -0.2

-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.00 0.20 0.40 0.60 0.80 1.00 0.00 0.20 0.40 0.60 0.80 1.00
u1 / u1C u1 / u1C
Experimental
ExperimentalData
Experimental Data Experimental
Experimental Data
Experimental Data
x/D = 5.46 x/D = 6.46 k ε− model
Data
ε Turbulence
1.0
k ε− model
k- ε Turbulence
ROKETSAN Solver, k- ε
Model
Experimental Data
1.0 ROKETSAN
k- Solver,
Experimental k- ε
DataModel
Laminar
Culick’s
Culick'sSolution
Solution Solver
Solution
ROKETSAN Laminar
Culick's
Culick’s Solution
Solution
Solution
ROKETSAN Solver
0.8 0.8 Laminar Soln
Laminar Soln

0.6 0.6

0.4 0.4

0.2 0.2

r/rw
0.0 0.0
r/rw

-0.2 -0.2

-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.00 0.20 0.40 0.60 0.80 1.00 0.00 0.20 0.40 0.60 0.80 1.00
u1 / u1C u1 / u1C
Experimental
Experimental Data
Experimental Data
Data Experimental
Experimental Data
Experimental Data
Data
x/D = 7.88 k ε− model
ε Turbulence
ROKETSAN
k- Solver, k- ε
Model
x/D = 9.06 kk-ε−Experimental
ROKETSANε Turbulence
model Solver, k- ε
Model
1.0 Experimental Data 1.0 Data
Laminar
Culick’s
Culick's Solution
Solution
Solution
ROKETSAN Solver Laminar
Culick’s
Culick's Solution
Solution
ROKETSAN Solver
0.8 Laminar Soln 0.8 Laminar Soln

0.6 0.6

0.4 0.4

0.2 0.2
r/rw

0.0
r/rw

0.0
-0.2 -0.2

-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.00 0.20 0.40 0.60 0.80 1.00 0.00 0.20 0.40 0.60 0.80 1.00
u1 / u1C u1 / u1C
Experimental
Experimental Data
Data Experimental
Experimental Data
Experimental Data
x/D = 10.3 Experimental Data
kε−model
ε Turbulence
x/D = 12.72 k ε− Experimental
ε Turbulence
k- ε Data k- ε
ROKETSAN Solver, k- model
ROKETSAN Solver,Model
1.0 k- Experimental DataModel 1.0
Laminar
Culick's
Culick’s Solution
Solution
Solution
ROKETSAN Solver Culick’s
Laminar
Culick's Solution
Solution
Solution
ROKETSAN Solver
0.8 Laminar Soln 0.8 Laminar Soln

0.6 0.6
0.4 0.4
0.2 0.2
r/rw

r/rw

0.0 0.0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1.0 -1.0
0.00 0.20 0.40 0.60 0.80 1.00 0.00 0.20 0.40 0.60 0.80 1.00
u1 / u1C u1 / u1C

Fig. 4. Axial mean-velocity in vertical direction at various axial locations.


28 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

x/D = 1.80 x/D = 3.04


1.0 Experimental Data 1.0
Experimental Data
k − ε Turbulence Model
Experimental Data
0.8 0.8 k − Experimental Data
ε Turbulence Model
k-ε model
ROKETSAN Solver k-ε model
ROKETSAN Solver
0.6 0.6
0.4 0.4
0.2 0.2

r/rw
r/rw

0.0 0.0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1.0 -1.0
0.00 0.03 0.06 0.09 0.12 0.15 0.00 0.03 0.06 0.09 0.12 0.15
I I
x/D = 5.46 x/D = 6.46
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
Experimental Data Data
Experimental Experimental Data Data
Experimental

r/rw
r/rw

0.0 k − εk- ε model


ROKETSAN
Turbulence Solver
Model 0.0 k − εROKETSAN
ε model Model
k-Turbulence Solver
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.00 0.03 0.06 0.09 0.12 0.15 0.00 0.03 0.06 0.09 0.12 0.15
I I
x/D = 7.88 x/D = 9.06
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
r/rw
r/rw

Experimental
ExperimentalData
Data Experimental Data Data
Experimental
0.0 k −ROKETSAN
k-εε Turbulence
model Solver
Model 0.0 k − εROKETSAN
ε model Solver
k-Turbulence Model
-0.2 -0.2

-0.4 -0.4

-0.6 -0.6
-0.8
-0.8
-1.0
-1.0 0.00 0.03 0.06 0.09 0.12 0.15
0.00 0.03 0.06 0.09 0.12 0.15
I I
x/D = 10.3 x/D = 12.72
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2 Experimental Data Data


Experimental Data Data
Experimental Experimental
k − εk-
r/rw
r/rw

0.0 k − ε ROKETSAN
Turbulence 0.0 ε model Model
Turbulence
ROKETSAN Solver
k- ε model Model
Solver
-0.2 -0.2

-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1.0 -1.0
0.00 0.03 0.06 0.09 0.12 0.15 0.00 0.03 0.06 0.09 0.12 0.15
I I
Fig. 5. Turbulence intensity in vertical direction at various axial locations.
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 29

4.2. Design optimization Design optimization process of ONERA C1xb subscale motor is
split into three separate parts. Firstly, only nozzle geometry is opti-
The goal of the present design is to generate a rocket motor that mized while the combustion chamber’s geometry has initial base-
produces maximum thrust. Hence, the objective function is defined line configuration. Then, combustion chamber is optimized while
as: the nozzle has baseline conical geometry. Finally, both nozzle
Z and combustion chamber are optimized simultaneously.
Thrust ¼ ðqu2 þ PÞdA ð17Þ
Aexit
4.2.1. Design optimization of nozzle
The objective function is normalized with respect to the initial of The total geometry change of nozzle wall, Dr, perpendicular to
thrust value and this function is tried to be maximized during de- the axial axis is defined as the linear combination of base functions
sign optimization processes. In Eq. (18), the Thrust represents the fi:
thrust of rocket motor, the sub-0 denotes the initial baseline thrust
value.
X
I
DrðxÞ ¼ di fi ðxÞ ð20Þ
Thrust i¼1
Objective function : S ¼ ð18Þ
Thrust0 where x is the normalized supersonic part position on the nozzle,
A constraint is formulated to control the weight of propellant. This and I stands for the number of base functions to be used. The
constraint is formulated in Eq. (19) where mpropellant represents the weighting coefficients, di, in the equation are the design variables,
weight, the sub-0 denotes the initial baseline weight value. Xi in Eq. (10), to be determined through the optimization process.
The performance of the design process is strongly influenced by
mpropellant
Constraint function : g ¼ 1  <0 ð19Þ the choice of the shape functions. The shape functions can influ-
mpropellant0 ence the convergence rate of the optimization cycle and may gen-
The constraint is formulated such that it is satisfied with the nega- erate a wavy surface in the design result. The present study uses
tive value per the guidelines of DOT. Grid generation code coupled three different types of base functions.
with optimization tool automatically updates the mesh for each
function calls of design process. 4.2.1.1. Hicks–Henne functions. Hicks–Henne functions are one of
The performance of design code is tested for another rocket mo- the frequently used shape functions in airfoil optimization [30].
tor. For this reason, a small rocket motor called as ‘‘whistling mo- The first one and the next seven and ten sinusoidal functions are
tor’’ or ‘‘C1xb’’, designed by ONERA is used. This rocket motor was given as follows:
constructed by ONERA for the validation of turbulence models,
f1 ðxÞ ¼ x0:25 ð1  xÞe20x for i ¼ 1
propellant response and two-phase flow effects [27–29]. This rock- ð21Þ
3
et motor has a cylindrical chamber of 700 mm long with a 85 mm fi ðxÞ ¼ sin ½pxeðiÞ  for i > 1
inner diameter. It contains an axisymmetric propellant grain, lo-
where eðiÞ ¼ lnð0:5Þ
lnðxi Þ
.
cated in the head-end half of the chamber, with a conical diverging
downstream end. The grain thickness is 18 mm. The nozzle geom- xi ¼ 0:15; 0:30; 0:45; 0:60; 0:75; 0:90 for i ¼ 7
etry has conical shape. Figs. 6 and 7 present the motor geometry xi ¼ 0:15; 0:25; 0:35; 0:45; 0:55; 0:65; 0:75; 0:85; 0:90 for i ¼ 10
and boundary conditions. This geometry is referred as baseline
geometry for design optimization process.
The following boundary condition is specified on the injecting 4.2.1.2. Wagner functions. Wagner functions were first applied to
wall: airfoil optimization by Ramamoorthy and Padmavathi [31]. They
€ ¼ 13:05 kg=m2 s; k ¼ 0:4 m2 =s2 ;
T 0 ¼ 3375 K; m e ¼ 0:14 m2 =s3 provide fairly large variations. However, high harmonics may cause
waviness in the design geometry.

Fig. 6. Sketch of ONERA C1xb subscale motor.

r
x

rw
D/2
Nozzle
Injecting wall
Non-injecting

Fig. 7. Boundary conditions of ONERA C1xb subscale motor.


30 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

 
h þ sinðhÞ 2 h example, optimization with ten Patched polynomials required
f1 ðxÞ ¼  sin for i ¼ 1
p 2
ð22Þ
twice as many function calls compared to the other shape func-
sinðihÞ sin½ði  1Þh tions. The increase in the number of shape functions improved
fi ðxÞ ¼ þ for i > 1 the design efficiency with the Hicks–Henne and Wagner functions,
ip p
1 pffiffiffi
but not with the Patched polynomials. With Patched polynomials,
where h ¼ 2 sin ð xÞ. a smaller reduction in objective function is achieved. This may be
related to the locations of the maximum height of the Patched
4.2.1.3. Patched polynomials. The following patched polynomials polynomials, xi. Changing these locations may improve the conver-
join smoothly at the position xi where the perturbation is maxi- gence of objective function. The results of thrust with ten shape
mum. A parabola on one side of xi is patched with a cubic on the functions are shown in Table 2. The maximum 3.09% thrust in-
other side. This provides continuities up to second order deriva- crease is achieved in the optimized nozzle with Hicks–Henne func-
tives without oscillations. tions. In designs with other shape functions, the increase in thrust
 2  ! is smaller.
xi  x A x Fig. 10 shows the distribution of Mach number, turbulent vis-
f1 ðxÞ ¼ 1  1þ for 0 < x 6 xi
xi ð1  xi Þ2 xi cosity ratio and turbulent kinetic energy in the baseline and opti-
 2  ! ð23Þ
mized nozzles. As expected, in both nozzles, Mach number
x  xi B 1x
fi ðxÞ ¼ 1  1þ for xi < x 6 1 increases in axial direction. However, the distribution of this prop-
1  xi ðxi Þ2 1  xi erty shows some differences in the baseline and optimized nozzles.
where Mach numbers at the exit of the optimized nozzle reach to higher
values. The same figure also shows the distribution of the turbulent
A ¼ maxð0; 1  2xi Þ viscosity ratio (lt/l) which is defined as;
B ¼ maxð0; 2xi  1Þ
lt C l fl ðqkÞ2
¼ ð24Þ
xi ¼ 0:10; 0:20; 0:30; 0:50; 0:70; 0:80; 0:90 for i ¼ 7 l e
xi ¼ 0:05; 0:15; 0:25; 0:35; 0:45; 0:55; 0:65; 0:75; 0:85; 0:90 for i ¼ 10
Cl and fl are defined in Eq. (8). The maximum of the turbulent vis-
cosity is roughly 6200 times the molecular viscosity for both nozzle
The result of the nozzle optimization with baseline nozzle is
geometries. The differences in distributions of turbulent viscosity
shown in Fig. 8. As seen from the figure, the design optimization
ratio and turbulent kinetic energy are not significant as can be seen
produces a bell shaped nozzle. Effects of the number and type of
from the figure.
shape functions on design performance are shown in Fig. 9. The re-
In order to validate the accuracy of the design results, the base-
sults indicate that an improper choice of shape functions may de-
line and optimized geometries are analyzed with a commercial
crease the efficiency and increase the computational cost. For
CFD code. The validation is performed using the FLUENT code with
the second- and third-order upwind methods and standard k–e tur-
bulence model. The optimized geometry is obtained using Hicks–

Table 2
Design optimization for nozzle (with 10 shape functions).

Shape functions Thrust of baseline Nozzle % Increase


3.8744
Thrust of optimized nozzle
Hicks–Henne shape functions 3.9945 3.09
Patched polynomials 3.9617 2.25
Fig. 8. Design optimization for nozzle (Hicks–Henne shape function) (–– baseline Wagner functions 3.9015 0.71
geometry, –– optimized geometry).

1.035 1.035
Thrust Thrust
Obj = Obj =
1.030 Thrust 0 Thrust 0
1.030

1.025 1.025
Objective

Objective

1.020 1.020

1.015 1.015

1.010 1.010

1.005 Hicks-Henne Func. 1.005 Hicks-Henne Func.


Patched Poly. Patched Poly.
Wagner Func. Wagner Func.
1.000 1.000
0 20 40 60 80 100 0 20 40 60 80 100
Design cycle Design cycle
(a) With seven shape functions (b) With ten shape functions
Fig. 9. Nozzle design convergence history.
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 31

code, fluxes are calculated using Roe’s upwind Flux-Difference


Splitting Scheme. Several interpolation schemes are available in
Mach:
0.1 0.50.1
0.90.5
1.3 0.9
1.7 1.3
2.1 1.7
2.5 2.1
2.9 2.5 2.9 FLUENT: second-order upwind and third-order upwind. The sec-
ond-order upwind scheme provides stability for supersonic flows.
Compared to the second-order upwind scheme, the third-order
scheme has a potential to improve spatial accuracy for all types
Optimized geometry
Optimized geometry of meshes by reducing numerical diffusion, most significantly for
complex three-dimensional flows, and it is available for all
transport equations [32]. The thrust value calculated with the
Baseline geometry third-order scheme for optimized nozzle is approximately 0.72%
Baseline geometry
higher than the one calculated with the present code. Another rea-
son for having different thrust values from the present and FLUENT
codes may be due to the extra term, such as the function u1, is used
in the present k–e turbulence model. The standard k–e turbulence
model used in FLUENT code does not include this term.
μt
: 200
200 1400
1400 2600
2600 38003800 5000
5000 6200 6200
μ
4.2.2. Design optimization of combustion chamber
The second part of design optimization process includes the
combustion chamber optimization. In this part, nozzle geometry
Optimized geometry
Optimized geometry is kept fixed as baseline configuration. A total of three design vari-
ables are used and they are shown in Fig. 11. The design variables
Baseline geometry are chosen as propellant divergence location (Xs), port radius (rw)
Baseline geometry and divergent angle (a). Initial baseline geometry is selected as
the combustion chamber geometry of ONERA C1xb subscale motor.
Design optimization code is tested by re-designing this combustion
chamber. Table 4 lists the values of design variables for the initial
geometry, the upper and lower bounds of design variables (UB and
LB). Design optimization is subjected to a constraint satisfying that
k: 00 242448 4871 71
95 119
95 143
119166 190166 190
143 the propellant weight of optimized geometry is not less than the
propellant weight of baseline geometry.
The baseline combustion chamber and resulted optimized
geometries are shown in Fig. 12. Table 5 lists the final design var-
Optimized geometry
Optimized geometry iable values for the optimized combustion chamber geometry.

Baseline
Baseline geometry
geometry
rw D/2
α
Xs

Fig. 11. Design variables for combustion chamber optimization.

Fig. 10. Distributions of Mach number, turbulent viscosity ratio (lt/l) and
turbulence kinetic energy (k) for baseline and optimized nozzle geometry.
Table 4
Design variable configuration for combustion chamber.

Table 3 Variable Name initial value LB UB


Analyses of baseline and optimized geometries with different flow solvers.
Xs/D 3.850 2.140 5.560
Geometry name Present code FLUENT FLUENT rw/D 0.414 0.243 0.456
(second-order (third-order a 18.08o 16.08o 60.08o
upwind) upwind)
Baseline geometry 3.8744 3.9007 3.9019
Optimized geometry 3.9945 4.0196 4.0239

Fig. 12. Design optimization for combustion chamber (–– baseline geometry, ––
Henne functions. In nozzle design, the thrust values calculated optimized geometry).
with the present and FLUENT codes are compared in Table 3. If
the second order methods are used in both codes, the thrust values
calculated with the FLUENT code are larger in both the baseline
Table 5
and the optimized geometries. The thrust values calculated with Design variable values for optimized combustion chamber.
the FLUENT code are approximately 0.68% larger than the ones pre-
Variable Name value
dicted with the present code. This may be due to using different
flux calculations schemes in the present and the FLUENT codes. Xs/D 5.560
rw/D 0.414
In the present code, fluxes are calculated with central difference
a 16.08o
scheme and artificial dissipation terms are added. In the FLUENT
32 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

Table 6 Reinj ¼ uinj rw q=l


Thrust of initial baseline and optimized geometries. ð25Þ
Rem ¼ um Lq=l
Geometry name Thrust % Increase
Baseline geometry 3.8744 44.7 where the uinj is the injection velocity at injecting wall, um is the
Optimized geometry 5.6062 mean axial velocity and L is the axial length. The transition to tur-
bulence would occur when the Rem reaches a critical value that is
dependent on the Reinj in accordance with linear stability theory
Table 7 [33]. Apte and Yang [34] stated that for high injection Reynolds
Constraint values of initial baseline and optimized geometries. numbers, the flow is mainly driven by the pressure gradient arising
Constraint name Initial value Optimized value from the mass injection; the viscous shear stress plays a much less
Propellant weight 0 0.00165
important role in determining the flow development and laminar-
to-turbulence transition point shifts to the downstream of motor.
The injection Reynolds numbers at the propellant divergence loca-
tion (Xs) is about 9110 for the baseline geometry and 9384 for the
As shown in Table 6, 44.7% thrust increase is achieved in the optimized geometry. So, the mean-flow transition in baseline geom-
optimized combustion chamber. This is due to the increase in pro- etry is attained earlier compared to the optimized geometry as seen
pellant burn area. Since the mass flow rate from injecting wall is from Fig. 13.
greater in the optimized geometry, approximately 45% thrust in- In combustion chamber design, the thrust values calculated
crease is achieved. The initial and optimized values of constraint with the present and FLUENT codes are compared in Table 8. As
variable given in Eq. (19) are listed in Table 7. Propellant weight in the nozzle design, FLUENT code uses the second- and third-order
is approximately same as with the baseline geometry. upwind schemes as opposed to central difference scheme used in
In Fig. 13, the distributions of Mach number, turbulent viscosity the present code. Results show the thrust values predicted with
ratio and turbulence kinetic energy are shown in the optimized the FLUENT code are higher for both baseline and optimized geom-
and baseline combustion chambers. As can be seen from the fig- etries. The thrust calculated by second-order upwind scheme is
ures, in the optimized combustion chamber, turbulent viscosity ra- approximately 0.69% greater than the one evaluated with the pres-
tio and turbulence kinetic energy values are smaller.
Although the entire flow field is driven by the mass injection Table 8
through the injecting wall, the characteristics of each flow regime Analyses of baseline and optimized geometries with different flow solvers.

are dominated by different physical mechanisms. Two key param- Geometry name Present code FLUENT FLUENT
eters characterizing the flow development are defined here: the (second-order (third-order
injection Reynolds number, Reinj, and the mean-flow Reynolds upwind) upwind)

number, Rem. Baseline 3.8744 3.9007 3.9019


Optimized geometry 5.6062 5.6426 5.6472

Mach : 0.01 0.08 0.13 0.30 0.60 0.90 1.20 1.50 1.80 2.10 2.40 2.70

Optimized geometry

Baseline geometry

μt
: 200 800 1400 2000 2600 3200 3800 4400 5000 5600 6200
μ

Optimized geometry

Baseline geometry

k : 0.0 3.0 22.2 200.0 700.0 1200.0 1700.0

Optimized geometry

Baseline geometry

Fig. 13. Distributions of Mach number, turbulent viscosity ratio (lt/l) and turbulence kinetic energy (k) for baseline and optimized combustion chamber geometry.
M. Yumusßak / Computers & Fluids 75 (2013) 22–34 33

Table 11
Analyses of baseline and optimized geometries with different flow solvers.

Geometry name Present code FLUENT FLUENT


Fig. 14. Design optimization for combustion chamber and nozzle (–– baseline (second-order (third-order
geometry, –– optimized geometry). upwind) upwind)
Baseline geometry 3.8744 3.9007 3.9019
Optimized geometry 5.7302 5.7680 5.7708
Table 9
Thrust of initial baseline and optimized geometries.

Geometry name Thrust % Increase than the thrust obtained by seperate optimization approach
Baseline geometry 3.8744 47.9 (3.09% for nozzle +44.7% for combustion chamber 47.7%) with a
Optimized geometry 5.7302 small amount. The major improvement in thrust is again achieved
by the combustion chamber design. The initial and optimized val-
ues of constraint variable given in Eq. (19) are listed in Table 10.
The contours of Mach number, turbulent viscosity ratio and
Table 10
Constraint values of initial baseline and optimized geometries.
turbulence kinetic energy are shown in Fig. 15. These figures indi-
cate that the simultaneous optimization reduced the turbulent vis-
Constraint name Initial value Optimized value
cosity ratio and kinetic energy in the nozzle and combustion
Propellant weight 0 0.00166 chamber region compared to the separate optimization.
Similar validation study is also performed for the combustion
chamber and nozzle design. The thrust values calculated with pres-
ent scheme. In third-order scheme, the difference becomes larger, ent and FLUENT codes are compared. Results are summarized in
it is approximately 0.73%. Table 11. Again, higher thrust values are achieved with second-or-
der upwind and third-order schemes. The thrust calculated with
the second-order upwind scheme is approximately 0.66% greater
4.2.3. Design optimization of combustion chamber and nozzle
than the one obtained with the present code. In third-order
The final part of design optimization process includes both noz-
scheme, this value reaches to approximately 0.71%.
zle and combustion chamber optimizations. They are optimized
simultaneously. Ten Hicks–Henne shape functions for the nozzle
and three design variables for the combustion chamber are used. 5. Conclusion
Fig. 14 shows initial baseline and optimized geometries.
As shown in Table 9, 47.9% thrust increase is evaluated in the The presented design optimization method improves the per-
optimized nozzle and combustion chamber. This value is greater formance of a solid rocket motor in viscous flows. The Navier–

Mach : 0.01 0.08 0.13 0.30 0.60 0.90 1.20 1.50 1.80 2.10 2.40 2.70

Optimized geometry

Baseline geometry

μt
: 200 800 1400 2000 2600 3200 3800 4400 5000 5600 6200
μ

Optimized geometry

Baseline geometry

k: 0.0 3.0 22.2 200.0 700.0 1200.0 1700.0

Optimized geometry

Baseline geometry

Fig. 15. Distributions of Mach number, turbulent viscosity ratio (lt/l) and turbulence kinetic energy (k) for baseline and optimized nozzle and combustion chamber.
34 M. Yumusßak / Computers & Fluids 75 (2013) 22–34

Stokes and modified k–e turbulence model flow equations are [8] Guobiao C, Jie F, Xu X, Minghao L. Performance prediction and optimization for
liquid rocket engine nozzle. Sci Technol 2007;11:155–62.
solved simultaneously by using finite difference explicit method.
[9] Culick FEC. Rotational axisymmetric mean flow and damping waves in solid
The results have been compared with the experimental data and propellant rocket motors. AIAA J 1966;4(8):1462–4.
it is seen that modified k–e turbulence model provides good accu- [10] Taylor GL. Fluid flow regions bounded by porous surfaces. Proc Roy Soc Lond
racy. The sensitivity derivatives for design optimization model are Ser 1956;234A(11199):456–75.
[11] Dunlap R, Willoughby PG, Hermsen RW. Flow field in the combustion chamber
obtained by finite difference evaluation. Solid rocket motor nozzle of a solid propellant rocket motor. AIAA J 1974;12(10):1440–2.
and combustion chamber are optimized separately and simulta- [12] Yamada K, Goto M, Ishikawa N. Simulative study on the erosive burning of
neously. Maximum improvement in thrust is achieved by simulta- solid rocket motor. AIAA J 1976;14(9):1170–6.
[13] Traineau JC, Hervat P, Kuentzmann P. Cold-flow simulation of a two-
neous design optimization. In order to validate the accuracy of dimensional nozleless solid rocket motor. AIAA paper 86-1447; 1986.
design results, the gain in objective function is evaluated using a [14] Sabnis JS, Madabhushi R, Gibeling HJ, MacDonald H. On the use of k–e
commercial code, FLUENT. Results show that even though there turbulence model for computation of solid rocket internal flows. AIAA paper
89-2558; 1989.
are some discrepancies between the thrust values calculated with [15] Tseng IS. Numerical simulation of velocity-coupled combustion response of
the present and FLUENT codes, similar gains are achieved with solid rocket propellants. Ph.D dissertation. University Park (PA): Dept. of
both codes. The parametric study that is conducted indicates that Mechanical Engineering, Pennsylvania State Univ.; 1992.
[16] Apte SV, Yang V. Effects of acoustic oscillation on flow development in a
the design efficiency can be enhanced significantly through the simulated nozzleless rocket motor. In: Yang V, Brill TB, Ren W, editors. Solid
correct choice of design variables. It also shows that the design cost propellant chemistry, combustion and motor interior ballistics. Progress in
can be controlled by the number and type of base functions, while aeronautics and astronautics. AIAA Reston, VA, vol. 185; 2000. p. 791–822.
[17] Nicoud F, Poinsot TJ, Minh MH. Direct numerical simulation of a turbulent flow
achieving a comparable level of improvement in the thrust perfor-
with massive uniform injection. In: 10th Symposium on turbulent shear flows,
mance. Utilizing analytical calculation of sensitivity derivatives vol. 3, no. 29; 1995. p. 13–8.
may provide benefit in improving the accuracy of finite difference [18] Eyi S. Finite-difference sensitivity calculation in iteratively solved problems.
method. However, in order to accomplish this improvement, Jaco- AIAA paper 3071-379; 2011.
[19] Vanderplaats GN, Hansen SR. DOT user manual. Goleta (CA): VMA
bian matrices should be derived by taking the derivatives of Na- Engineering; 1989.
vier–Stokes and turbulence equations with respect to the flow [20] Jameson A, Schmidt W, Turkel E. Numerical solutions of the Euler equations by
and design variables. Increasing the thrust may not be an easy task finite volume methods using Runga Kutta time stepping schemes. AIAA paper
81-1259; 1981.
in rocket motor design. In the present study, choosing the objective [21] Yumusßak M, Vuillot F, Tinaztepe T. Viscous flows in solid propellant rocket
function as thrust produces a gain at only a small amount. Larger motors. AIAA paper 5116-138; 2006.
improvement in objective function may be achieved by changing [22] Tinaztepe T, Yumusak M, Vuillot F. Validation of internal flow prediction codes
for solid propellant rocket motors. RTO meeting on advances in rocket
the objective function. For example, choosing the nozzle length propellant performance. Life and disposal for improved system performance
as the objective function may be more appropriate as far as the and reduced cost, Aalborg, Denmark; 2002.
reduction in objective function is concerned. [23] Yumusak M, Eyi S. Design optimization of rocket nozzles in chemically
reacting flows. Int J Comput Fluids 2012;65:25–34.
[24] P Jones W, Launder BE. The prediction of laminarization with a two-equation
Acknowledgments model of turbulence. Int J Heat Mass Transfer 1972;15:301–14.
[25] Shima N. Prediction of turbulent boundary layers with a second-moment
closure: Part I – effects of periodic pressure gradient, wall transpiration and
This research is supported by NATO RTO-AVT Support Project
free-stream turbulence. Part II – effects of streamline curvature and spanwise
Committee and The Scientific and Technological Research Council rotation. J Fluid Eng 1993;115:56–69.
of Turkey (TUBITAK). [26] Dunlap R, Blackner AM, Waugh RC, Brown RS, Willoughby PG. Internal flow
field studies in simulated cylindrical port rocket chamber. J Propulsion Power
1990;6:690–703.
References [27] Lupoglazoff N, Vuillot F. Numerical simulation of vortex shedding
phenomenon in 2D test case solid rocket motors. AIAA paper 92-0776; 1992.
[1] Kamran A, Guozhu L. An integrated approach for optimization of solid rocket [28] Lupoglazoff N, Vuillot F. Comparison between firing tests and numerical
motor. Aerosp Sci Technol 2012;17:50–64. simulation of vortex shedding in a 2D test solid rocket motors. AIAA paper 93-
[2] Anderson M, Burkhalter J. Multi-disciplinary intelligent systems approach to 3066; 1993.
solid rocket motor design, Part I, II: single and dual goal optimization. AIAA [29] Vuillot F. Vortex shedding phenomena in solid rocket motors. J Propulsion
2001:3559–600. Power 1995;11(4):626–39.
[3] Kumar VR, Kim HD, Raghunandan BN, Setoguchi T. Numerical studies on [30] Hicks R, Henne P. Wing design by numerical optimization. J Aircraft
turbulent separated flows in high-velocity transient motors. In: 15th 1978;15:407–12.
Australasian fluid mechanics conference. Sydney (Australia): The University [31] Ramamoorthy P, Padmavathi K. Airfoil design by optimization. J Aircraft
of Sydney; 2004. 1977;14:219–21.
[4] Rao GVR. Exhaust nozzle contour for optimum thrust. Jet Propulsion [32] ANSYS fluent theory guide. Release 14.5; 2012.
1958;2:377–82. [33] Beddini RA. Injection induced flows in porous-walled ducts. AIAA J
[5] Farley JM, Campbell CE. Performance of several method of characteristics 1986;24(11):1773–6.
exhaust nozzles. NASA TN D293; October, 1960. [34] Apte S, Yang V. Unsteady flow evaluation in porous chamber with surface mass
[6] Liquid rocket engine nozzles. NASA SP-8120; July, 1976. injection, Part 1: free oscillation. AIAA J 2001;39(8).
[7] Xing XQ, Damodaran M. Design of three-dimensional nozzle shapes using
hybrid optimization techniques. AIAA paper-2004-26. In: 42nd AIAA aerospace
sciences meeting and exhibit, Reno, Nevada; January 5–8, 2004.

You might also like