Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Numerical simulation of fluid-structure interaction for thermochemical ablation

ed
of solid rocket engine nozzle

Zhitan Zhou*, Dehua Cao, Ranhui Liang

School of Aircraft Engineering, Nanchang Hangkong University, Nanchang 330063, China

iew
Abstract: When a solid rocket engine is ignited, the throat lining of the nozzle is prone to chemical

ablation due to high-temperature gas erosion, resulting in thrust loss. In this paper, a coupled fluid-solid

model for thermochemical ablation on the nozzle wall is established based on the multi-component

v
Navier-Stokes equations, SST k-ω turbulence model, finite-rate chemical reaction model on the nozzle

re
wall, variable transport properties of nozzle material, and the heat conduction equation. By comparing

with experimental data, the maximum error of the calculated ablation rate is 4.37%, validating the

effectiveness of the model. Subsequently, the influence of different combustion chamber components,
er
chamber pressure, and chamber temperature on the ablation rate of C/C throat lining is studied. The

results indicate that the temperature at the nozzle throat is the highest, resulting in the maximum ablation
pe
rate. As the aluminum mass fraction at the nozzle inlet increases, with less oxidizer mass fraction, the

thermochemical ablation rate of the nozzle decreases. The inlet pressure and temperature of the nozzle

are positively correlated with the ablation rate, with temperature having a more significant impact

compared to pressure. The research findings can provide theoretical guidance for the thermal protection
ot

design of rocket engine nozzles.

Keywords: Laval nozzle; Ablation; C/C composite materials; Multiphysics coupling


tn

1. Introduction

With the rapid development of modern military technology, the specific impulse and combustion
rin

temperature of future solid rocket motors (SRMs) are expected to increase, leading to a more complex

working environment for SRMs [1-3]. Traditional graphite throat linings, widely used in the past, are

prone to mechanical performance failure during the operation of medium and large-sized engines due to
ep

insufficient strength and mediocre thermal shock resistance. They cannot meet the design requirements

for future solid rocket motor nozzles [4,5]. In 1958, the Chance Vought aerospace company in the United

States developed C/C composite materials. Upon testing, it was found that C/C composite materials
Pr

exhibit excellent high-temperature performance and good mechanical properties. Compared to graphite

throat linings, C/C material linings have advantages such as high strength, resistance to erosion, good

*Corresponding author.
E-mail address: ztzhou93@nchu.edu.cn (Z. Zhou).

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
resistance to thermal shock, resistance to solid particle erosion, and a low coefficient of thermal

ed
expansion, making them an ideal material for solid rocket motor nozzle linings [6-8].

However, during engine operation, high-temperature gases flow over the nozzle wall, forming a

turbulent boundary layer and transferring heat to the nozzle wall, causing an increase in wall temperature.

iew
After reaching a certain temperature, oxidation components in the gas undergo a heterogeneous chemical

reaction with the C/C material, resulting in thermochemical ablation and causing recession of the nozzle

wall. Especially for nozzles operating over extended periods, thermochemical ablation can lead to an

increase in throat area, resulting in a loss of engine performance [9,10]. Therefore, the study of

v
thermochemical ablation in C/C composite material nozzles is of significant practical importance. There

re
is an urgent need to establish ablation models that reflect real physical processes and accurately predict

nozzle ablation rates.

To investigate the impact of ablation on the thermal protection materials of nozzles, extensive
er
numerical simulations and experimental work have been conducted. Brian Evans et al. [11], in order to

assess the influence of different propellants on the chemical erosion rate of graphite nozzles, designed
pe
and tested solid propellant rocket engines. They found that the erosion rate of throat diameter increases

initially and reaches a relatively constant rate in the later stages of the experiments. Additionally,

scanning electron microscope (SEM) images of the recovered nozzle surface structure from non-
ot

metalized propellant tests showed a rougher surface compared to the nozzles used in metalized propellant

tests. Wang Chen [12] conducted ablation experiments on C/C composite materials using arc plasma jets.
tn

They established an ablation model for C/C materials using energy balance, mass conservation, and

chemical equilibrium. The model describes the thermal response and ablation rate of the ablative body

in high-speed flow. The theoretical model was validated through temperature measurements and online
rin

emission spectroscopy diagnostics during ablation experiments, showing good agreement between

simulation results and experimental outcomes. Daniele Bianchi et al. [13] studied the corrosion behavior

of graphite/carbon-carbon nozzles in solid rocket engines with various propellant formulations. The
ep

results indicated that, for metalized propellants, the concentration of major oxidizing species (H 2O, OH,

and CO2) significantly decreased due to the formation of aluminum oxides, leading to a substantial
Pr

reduction in the thermal-chemical erosion of these oxidizing species compared to non-metalized

propellants. Piyush Thakre et al. [14] developed a method to predict the chemical erosion of

graphite/carbon-carbon nozzles in solid propellant rocket engines. They found that H2O is the most

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
detrimental oxidizing species influencing the erosion of graphite/carbon-carbon nozzles, followed by OH

ed
and CO2.Xiaotian Zhang et al. [15] established Fluent DPM and Abaqus DEM models to investigate the

chemical ablation and mechanical erosion mechanisms of nozzle thermal protection materials. The

results revealed that, in the downstream of the converging section and upstream of the throat, wall

iew
degradation is dominated by a combination of chemical ablation and mechanical erosion, while at other

locations, chemical ablation rates prevail.

Currently, research on the chemical ablation mechanism of C/C material throat is relatively

comprehensive, and an initial ablation model for C/C materials has been established. However, the study

v
of the thermal-fluid-solid coupling mechanism of the flow field inside the nozzle, nozzle heat transfer,

re
and wall recession is not yet complete. Additionally, existing research often focuses on the ablation

behavior of nozzles with different propellant formulations, with limited attention given to the influence

of pressure and temperature on nozzle ablation. Considering the significant differences in engine pressure
er
and temperature, investigating the ablation characteristics of engine nozzles under different total

temperature and total pressure conditions is of great importance for the thermal protection design of
pe
nozzles.

Therefore, this paper aims to establish a numerical model for the ablation heat transfer of C/C

material nozzles during the operation of solid rocket engines. Numerical simulations of throat ablation
ot

under different propellant formulations and combustion chamber pressure/temperature conditions will

be conducted. The goal is to understand the influence of propellant formulations and combustion chamber
tn

pressure/temperature on the wall ablation rate of the nozzle. This research has significant guidance

implications for the design optimization of solid rocket engines.

2. Computational models
rin

2.1 Geometry model

This paper focuses on the numerical simulation research of the internal flow field and

thermochemical ablation on the nozzle wall of the 70-1b BATES engine [16]. The geometric structure of
ep

the engine nozzle is illustrated in Figure 1, with a throat diameter of 50.8mm, a convergent half-angle of

45°, an expansion half-angle of 15°, an expansion ratio of 9.5, and a combustion time of 2 seconds. The
Pr

throat material of the nozzle is composed of C/C composite material, and its physical parameters are

represented as follows in Eq. (1):

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
c p = 6.36110−12 T 4 + 9.755 10−9 T 3 − 4.224 10−4 T 2 + 1.618T + 252.3(300 K  T  3200 K )

ed
c p = 2093(T  3200 K )
 (1)
s = 40W / (m  K )
  = 1830kg / m3
 s

Where cp, λs, ρs denote specific heat, thermal conductivity and density, respectively.

v iew
re
er
pe
Fig.1 Geometry of 70-1b BATES motor and nozzle.

2.2 Mesh model

This study investigates the thermo-fluid-solid coupling process of the internal flow field, nozzle

heat transfer, and wall recession. A two-dimensional axisymmetric model is employed during grid
ot

generation, as illustrated in Figure 2. In the figure, the purple area represents the gas fluid domain, while

the green area represents the solid domain of the nozzle wall. The inlet of the nozzle is set as a pressure
tn

inlet with combustion chamber pressure and temperature values of 6.9 MPa and 3580 K, respectively.

The outlet of the nozzle is set as a pressure outlet with a pressure of 101325 Pa. Due to the complex
rin

physical and chemical processes involved at the fluid-solid interface, including heat transfer, ablation,

and recession, the mesh in this region is refined. The entire domain is discretized using quadrilateral

structured meshes. To satisfy the conditions for the k-ω (SST) turbulence model, the first mesh point
ep

needs to be located within the turbulent viscous sublayer, i.e., y+<1 (where y+ is a dimensionless quantity

representing the distance and velocity near the wall in turbulent flow). The final thickness of the first

layer of mesh is 0.2 μm, with a maximum calculated value of y+ equal to 0.45, meeting the computational
Pr

requirements. The mesh growth rate is set to 1.1.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
The wall recession during the ablation process is implemented using dynamic mesh technology. The

ed
recession amount is calculated based on the consumption of C/C material in the chemical reaction. The

dynamic mesh is updated using the elastic smoothing method and grid reconstruction method. User-

defined functions (UDF) and the DEFINE_GRID_MOTION macro are utilized to control the movement

iew
of mesh nodes on the internal wall of the nozzle, and the displacement of nodes is determined by the

calculated ablation amount.

v
re
Fig.2 Computational grid for nozzle.

After the initial mesh division, three sets of computational meshes with element numbers of 9184,
er
58774, and 147175 were respectively created to carry out grid independence verification. Numerical

simulations for the ablation of the nozzle in a solid rocket engine using an aluminum-containing

propellant were conducted. Ablation rate results at three different axial points were extracted, as shown
pe
in Table 1. Notably, the results from the first mesh set exhibit significant differences compared to the

results from the other meshes. In contrast, the results from the second and third mesh sets are in good

agreement. Considering both accuracy and computational time, the second mesh set is selected as the
ot

grid model for this study.

Table 1 The ablation rate of three grid groups.


tn

Ablation rate /(mm/s)


Number of grids
x=75mm x=88mm x=95mm
5184 0.21485 0.1989 0.0242
rin

18774 0.37175 0.3602 0.0241


47175 0.37035 0.3600 0.0241

3. Numerical method
ep

This study is based on the following assumptions: the gas components are treated as ideal gases;

chemical reactions between gas components are not considered; the impact of mechanical erosion caused

by condensing particles is disregarded; radiation heat transfer is neglected; and the influence of gravity
Pr

is ignored.

3.1 Governing Equations

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
The governing equations include the mass conservation equation, momentum conservation equation,

ed
and energy conservation equation. In a Cartesian coordinate system, the general form of the compressible

Navier-Stokes equations is expressed as follows [17]:


U F G H Fv Gv H v
+ + + = + + (2)
t x y z x y z

iew
U = [  , u,  v,  w,  e]T (3)

Where ρ,u,v,w,e, respectively, refers to the density, x velocity, y velocity, z velocity and specific

kinetic energy; F, G and H are convective term fluxes; Fv, Gv and Hv are viscous fluxes.

v
The turbulence model used is the SST k-ω model. This model exhibits good numerical stability near

the wall and provides favorable agreement between the computed results in the boundary layer and

re
experimental data [18]. The eddy viscosity, k-equation, and ω-equation for the SST k-ω model can be

expressed in the following forms:


t
( k ) +

xi
er  
x j  

(  kui ) =    + t
k
 k

 x j

 + Gk − Yk + Sk
 (4)

 
pe
     
t
(  ) + (
x j
)
u j =
x j
  + t
 
  + G − Y + D + S
 (5)
  x j 
k 1
t =
  1 SF  (6)
max   , 2 
  a1 
ot

Where Gk is the turbulent kinetic energy; Gω is the generation of ω; Yk and Yω denote the dissipation

of k and ω in the presence of turbulence; Sk and Sω are user-defined source terms; σk and σω represent the
tn

turbulent Prandtl numbers for k and ω; μt is the turbulence viscosity; S is the modulus of the mean strain-

rate tensor; and a* is the inhibition of turbulent viscosity coefficient.

The instantaneous thermal conductivity equation of solid phase materials is:


rin

Ts 1   T    T 
s Cs = s r s  +  s s  (7)
t r r  r  x  x 

Where ρs is the density of the solid material; Cs is the specific heat of solid phase material; Ts is the
ep

temperature of the solid phase material; λs is the thermal conductivity of the solid phase material.

3.2 Gas/Surface Boundary Conditions

Mass conservation equation of gas-surface can be written as:


Pr

(  g v )w = m =  s r (8)

Component conservation equation of gas-surface can be written as:

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
(  g vYi )w =   g Di,m Yr  + i (9)
w

ed
Energy conservation equation of gas-surface can be written as:

 T   Tg  i =1
−s  s  = −g   + i qi (10)
 r  w  r w N

iew
Where ρg is the density of the gas mixture; m is the material mass consumption rate; r is the solid

phase material consumption rate; Di,m is the diffusion coefficient of a gas component i; i is the rate of

formation or consumption of a gas component i; λg is the thermal conductivity of the gas mixture; qi is

v
the heat flux density of a gas component i; The subscript w represents the gas-solid interface.

The kinetics of the reaction between the solid phase material and gas on the inner surface of the

re
nozzle adheres to the Arrhenius law, as articulated in Eq. (11). Eq. (13) governs the consumption rate of

the solid phase material within the reaction process.

ki = AiTsb exp ( − Ei / ( RuTs ) )


er (11)

Wmix,s
pi ,s = psYi ,s (12)
Wi
pe
ri = ki pi ,s (13)

Where Yi,s and pi,s represent the mass fraction and partial pressure of the i component

respectively; Wmix,s, ps, Ts represent the molecular weight, surface pressure, and temperature of the
ot

gas mixture, respectively.

Given the relatively low concentration of OH in BATES engine gas, its impact on the ablation
tn

rate is deemed negligible. Consequently, the analysis focuses solely on the reaction involving C with

oxidizing components H2O and CO2. The corresponding reaction equation and its associated kinetic

parameters are detailed in Table 2 [19].


rin

Table 2 Kinetic parameters of chemical reaction.

Chemical reaction Ai / (kg/(m2·s)) b Ei / (J/mol)


ep

C(s)+H2O→CO+H2 1508.00 0 287859.2

C(s)+CO2→2CO 28.27 0 284930.4


Pr

3.3 Wall recession

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
During the operational phase of the nozzle, chemical ablation along the inner wall induces the

ed
depletion of solid material and consequent alterations in the nozzle profile. Notably, investigations have

demonstrated that a mere 5% augmentation in nozzle throat area can exert a considerable impact on

engine performance [20]. Thus, a comprehensive understanding of the chemical ablation process

iew
necessitates an in-depth examination coupled with the nozzle wall recession phenomenon. Under

unsteady conditions, the nozzle undergoes chemical ablation, leading to intricate interactions between

the flow field and the inner wall. Within this dynamic ablation process of the nozzle, the alterations in

the flow field and inner wall surface are integral components. In this study, a combined chemical ablative

v
model and a dynamic mesh model are employed to compute the consumption rate of solid material based

re
on the Arrhenius equation. The product of the consumption rate and the time step serves as the criterion

for grid node movement, thereby facilitating the simulation of the dynamic ablative wall recession

process of the nozzle. er


3.4 Coupling method

During nozzle operation, the carbon-carbon composite material within the solid domain of the nozzle
pe
engages in a reaction with the gas, leading to the recession of the combustion surface. In this intricate

process, a two-way fluid-structure coupling is established between the flow field and the nozzle profile

surface. As shown in Figure 3, the coupling calculation procedure is elucidated in the following steps:
ot

(1) At the initial moment, the fluid domain and the solid domain are initialized with appropriate

values.
tn

(2) The fluid domain and solid domain are solved in accordance with the specified inlet boundary

conditions. Parameters such as pressure and composition near the nozzle wall are derived from the

calculation results of the inner wall temperature and flow field.


rin

(3) The chemical reaction rate and ablation rate are computed based on the parameters obtained in

Step (2).

(4) Utilizing dynamic mesh technology, the mesh nodes on the inner surface of the nozzle are
ep

displaced according to the corresponding ablation rate.

(5) Steps (2) to (4) are iteratively repeated until the calculation reaches completion.
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
ed
v iew
re
er
Fig.3 Coupled computing flow.

4. Results and discussion


pe
4.1 Effect of Al mass fraction on ablation

In comparison to liquid propellants, solid propellants exhibit a higher specific impulse owing to the

inclusion of high-energy metal particles such as aluminum and magnesium. On the one hand, the
ot

combustion of these high-energy metal particles releases substantial heat, elevating the temperature

within the combustion chamber and promoting throat ablation reactions. On the other hand, the
tn

combustion of metal particles also consumes oxidizing components in the gas, mitigating gas oxidation

and inhibiting throat ablation reactions. Consequently, this paper investigates the influence of metal

particle mass fraction on the ablation of C/C lining.


rin

Based on real experiments [21,22] on the composition of gas after propellant combustion under

operational conditions, five working conditions with varying combustion chamber components are

proposed (Table 3). Notably, the aluminum mass fraction in the propellant ranges from 15% to 27%.
ep

With an increase in aluminum mass fraction, the gas's total temperature rises, while the total pressure

remains constant. The gas composition sees an increase in CO concentration, a decrease in H 2O and CO2
Pr

concentrations, and relatively stable concentrations of other components. The calculation time step is set

at 1ms, and the total calculation time is 2 seconds.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
Figure 4 depicts pressure, Mach number, and temperature cloud maps before and after ablation at

ed
15% aluminum mass fraction. In Figure 4(a) and 4(b), the pressure distribution in the flow field remains

consistent before and after ablation, mirroring the Mach number cloud diagram. In the nozzle contraction

section, gas pressure gradually diminishes, and gas acceleration persists. Notably, the gas in the

iew
supersonic flows no longer adheres strictly to the principle of "large flow rate at small cross section and

small flow rate at large cross section," but rather exhibits a reverse trend-larger cross-sections correspond

to faster flow rates. Consequently, in the nozzle expansion section, gas pressure continues to decline, and

acceleration persists. Figure 4(c) reveals that the temperature of the solid domain after ablation exceeds

v
that before ablation. However, the temperature of the fluid domain remains largely unchanged, with only

re
a slight decrease near the nozzle wall. This is attributed to the limited ablation of the nozzle throat lining

(less than 1mm at 2s), which is negligible compared to the throat lining radius of 50.8mm. Consequently,

the overall temperature of the fluid domain is unaffected by the changes in the nozzle profile.
er
Nevertheless, due to the heat transfer from high-temperature gas to the solid wall of the nozzle, the

temperature of the fluid domain near the nozzle experiences a slight decrease, while the temperature of
pe
the solid domain undergoes a substantial increase.

Table 3 Nozzle inlet conditions under different Al mass fraction conditions.

No. Al/% pc/MPa Tc/K yco yco2 yHCl yH2 yH2O yN2 yAl2O3
A1 15 6.9 3580 0.175 0.040 0.240 0.02 0.145 0.10 0.28
ot

A2 18 6.9 3655 0.180 0.025 0.230 0.02 0.105 0.10 0.34


A3 21 6.9 3715 0.200 0.015 0.195 0.02 0.070 0.10 0.40
tn

A4 24 6.9 3750 0.200 0.005 0.190 0.02 0.045 0.10 0.44


A5 27 6.9 3745 0.200 0.005 0.190 0.02 0.025 0.10 0.46
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
ed
v iew
re
er
pe
Fig.4 Comparison of pressure (a), Mach number (b), and temperature contours (c) before and after

ablation.

Table 4 presents the calculated ablation rate at the throat of the nozzle (x=88mm) and the ablation
ot

rate obtained from Geisler's experiments [21,23]. The average error is 3.056%, and the maximum error

is 4.37%. The comparison between the calculated results and experimental data demonstrates the
tn

accuracy and effectiveness of the ablative heat transfer coupling model proposed in this paper.

Table 4 Comparison of throat ablation rate and test value under different Al mass fraction conditions.

Experimental Calculated Relative


No. Al%
rin

value/(mm/s) value/(mm/s) error/%


A1 15 0.3531 0.3602 2.01
A2 18 0.2845 0.29055 2.13
A3 21 0.2000 0.20485 2.43
ep

A4 24 0.1245 0.1299 4.34


A5 27 0.0686 0.0656 -4.37

To investigate the impact of ablation on the nozzle contour, the nozzle contour map of the throat

region at the 2-second mark under A1 condition was extracted, as shown in Figure 5. It can be observed
Pr

that there is a noticeable recession in the vicinity of the throat, with the maximum recession at the throat

reaching 0.7204mm. In contrast, the recession of the wall surface away from the throat is relatively small,

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
and the expansion section wall surface shows little change, consistent with the actual ablation scenario

ed
in rocket engines.

v iew
re
Fig.5 Nozzle profile before and after ablation.

To verify the correctness of the numerical simulation results regarding the representation of physical
er
and chemical processes in thermochemical ablation, we conducted an analysis of the temperature

distribution and ablation rate along the axial direction of the nozzle wall. Figure 6(a) illustrates the wall
pe
temperature distribution using a dual-axis coordinate system, where the x-axis represents the axial

distance of the nozzle, the left y-axis denotes the static temperature at the nozzle wall, and the right y-

axis represents the radial distance from the nozzle wall. It can be observed from the figure that the wall
ot

temperature rises before the nozzle throat, reaching its maximum value just before the throat, followed

by a rapid decrease. With an increase in the mass fraction of aluminum particles, the temperature field in
tn

the gas flow raises, leading to a gradual increase in the peak temperature at the nozzle wall. Figure 6(b)

displays the distribution of ablation rates along the nozzle, with the x-axis indicating the axial distance

of the nozzle, the left y-axis representing the ablation rate at the nozzle, and the right y-axis representing
rin

the radial distance from the nozzle wall. In the flow of the nozzle, gas temperature and pressure decrease

gradually along the flow direction, while gas velocity increases. It can be observed from Figure 6(b) that

as the aluminum mass fraction increases, the ablation rate at the nozzle throat significantly decreases.
ep

This is attributed to the increased aluminum mass fraction, resulting in an increase in Al2O3 mass

fraction in the combustion products, a decrease in CO2 and H2O mass fraction, and a weakening of the
Pr

thermal ablation reaction with the C/C throat lining, ultimately leading to a reduction in the wall ablation

rate. Considering the results from Figure 6(a) and Figure 6(b), it is evident that although the wall

temperature of the nozzle increases with higher aluminum mass fraction, the ablation rate decreases. This

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
suggests that when the temperature rises to a certain value, the ablation rate of the nozzle is not solely

ed
determined by the temperature. In this scenario, the ablation rate primarily depends on the diffusion rate

of oxidizing components towards the wall. A higher aluminum mass fraction in the metallic propellant

results in a lower mass fraction of oxidizing components, leading to a smaller diffusion rate and,

iew
consequently, a reduced ablation rate.

To further investigate the influence of combustion chamber components on the ablation rate of the

C/C throat lining, we extracted ablation data for the nozzle throat (x=88mm), as shown in Figure 7. It

can be observed that the ablation rate at the nozzle throat significantly decreases with an increase in

v
aluminum mass fraction. This is due to the increased aluminum mass fraction, resulting in an increase in

re
Al2O3 mass fraction in the combustion products, a decrease in CO2 and H2O mass fraction, and a

weakening of the thermal ablation reaction with the C/C throat lining, ultimately leading to a reduction

in the wall ablation rate. er


pe
ot
tn

Fig.6 Effect of Al mass fraction variation on (a) wall temperature and (b) ablation rate.
rin
ep
Pr

Fig.7 Ablation rate of nozzle throat under different Al mass fraction.

4.2 Effect of combustion chamber pressure on ablation rate

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
According to the Eq. (13), it is evident that pressure and temperature are crucial factors influencing

ed
the ablation process. The reaction of H2O, CO2, and C in the throat lining leads to a C/C throat lining

ablation rate proportional to the square root of the pressure. Therefore, with an increase in pressure, the

ablation rate of C/C throat should also increase. Additionally, as combustion chamber pressure increases,

iew
combustion temperature also increases, and temperature is also a significant factor affecting the rate of

ablation of the C/C throat. With an increase in temperature, the ablation rate of the C/C throat also

increases, as reflected in the ablation rate Eq. (11). To investigate the influence of pressure on the ablation

of the C/C throat, numerical simulations are conducted for different combustion chamber pressures in

v
the range of 4.9 MPa to 9.9 MPa.

re
The combustion products of the engine propellant are consistent with the conditions outlined in

Section 4.1, Case A1. Table 5 provides the mass fractions of various components under different

pressures. Figure 8(a) illustrates the temperature distribution at various locations on the nozzle wall. It
er
can be observed that as the inlet pressure increases, the wall temperature of the nozzle increases to

varying degrees. The temperature rise is smaller at the nozzle throat and more pronounced at the inlet.
pe
This is due to the increase in total pressure at the inlet, but with a relatively small change in velocity,

resulting in an increase in gas temperature near the inlet and a more significant temperature change on

the wall surface near the inlet. Figure 8(b) shows the ablation rate at various locations on the nozzle wall.
ot

It can be observed that the ablation rate at the nozzle inlet is small, and the ablation rate in the expansion

section tends to be zero. From the nozzle inlet to the nozzle outlet, the wall ablation rate first increases
tn

and then decreases, ultimately reaching a steady state. Notably, the ablation rate is highest near the nozzle

throat, where the recession distance is also the largest.

By varying the pressure at the nozzle inlet, it is observed that as the inlet pressure increases, the wall
rin

temperature increases, and the throat ablation rate increases. The maximum ablation rate at the throat

linearly increases with pressure. As the combustion chamber pressure increases, the heat transfer from

the gas to the nozzle wall and the diffusion of oxidative components to the wall enhance, leading to an
ep

increase in the heterogeneous chemical reaction rate on the wall. Consequently, the ablation rate of the

nozzle increases, and it approximately follows a proportional relationship with pressure.


Pr

Figure 9 illustrates the relationship between the ablation rate of the nozzle throat and combustion

chamber pressure. With an increase in pressure, the throat ablation rate exhibits a nearly linear growth

trend. The variation in ablation rate with pressure can be approximated by the following equation:

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
r = 0.1059P - 0.3736 (14)

ed
Table 5 Nozzle inlet conditions under different pressure conditions.

NO. pc/MPa Tc/K Al/% yco yco2 yHCl yH2 yH2O yN2 yAl2O3
B1 4.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28
B2 5.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28

iew
B3 6.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28
B4 7.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28
B5 8.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28
B6 9.9 3580 15 0.175 0.040 0.240 0.02 0.145 0.10 0.28

v
re
er
pe
Fig.8 Effect of combustion chamber pressure on (a) wall temperature and (b) ablation rate .
ot
tn
rin

Fig.9 Ablation rate of nozzle throat under different pressure.

4.3 Effect of combustion chamber temperature on ablation


ep

To investigate the influence of temperature on the ablation of the C/C throat, numerical simulations

are conducted for different combustion chamber temperatures in the range of 2580 K to 3580 K. The

combustion products of the engine propellant are consistent with the conditions described in Section 4.1,
Pr

and Table 6 provides the mass fractions of different components at various temperatures. Figure 10(a)

illustrates the temperature distribution at various locations on the nozzle wall. It can be observed that as

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
the inlet temperature increases, the wall temperature of the nozzle increases to varying degrees. Unlike

ed
the increase in inlet pressure, the temperature rise at the nozzle throat is essentially consistent with that

at the inlet. Additionally, with a linear increase in the inlet temperature, the temperature rise on the wall

surface near the throat becomes smaller. Figure 10(b) shows the ablation rate at various locations on the

iew
nozzle wall. The wall ablation rate increases and then decreases along the nozzle axis (x-direction),

tending to zero in the expansion section. By varying the inlet temperature of the nozzle, it is observed

that with an increase in the inlet temperature, the overall wall ablation rate increases, and the maximum

ablation rate at the throat increases significantly with temperature. Compared to an increase in

v
combustion chamber pressure, an increase in combustion chamber temperature more significantly causes

re
an increase in throat ablation rate.

Figure 11 illustrates the relationship between the ablation rate of the nozzle throat and combustion

chamber temperature. The two variables show a positive correlation. With an increase in temperature,
er
the ablation rate of the throat increases slowly in the early stage but experiences a significant increase in

the later stage. The variation in ablation rate with temperature can be approximated by the following
pe
equation:

r = 4e-7T 2 - 0.0024T + 3.1924 (15)

Table 6 Nozzle inlet conditions under different temperature conditions.


ot

NO. Tc/K Al/% pc/MPa yco yco2 yHCl yH2 yH2O yN2 yAl2O3
C1 2580 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28
C2 2780 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28
tn

C3 2980 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28


C4 3180 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28
C5 3380 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28
C6 3580 15 6.9 0.175 0.040 0.240 0.02 0.145 0.10 0.28
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
ed
v iew
Fig.10 Effect of combustion chamber temperature on (a) wall temperature and (b) ablation rate.

re
er
pe

Fig.11 Ablation rate of nozzle throat under different temperature.

5. Conclusions
ot

The paper focuses on the ablation process of the solid rocket Laval nozzle throat. A coupled thermal-

chemical ablation model for the nozzle wall is established based on the multi-species Navier-Stokes
tn

equations, the SST k-ω turbulence model, a finite-rate chemical reaction model for the nozzle wall,

variable transport properties of the nozzle material, and the heat conduction equation. The effectiveness

of the model is validated by comparing with experimental data. Subsequently, numerical simulations of
rin

C/C throat ablation under different conditions are conducted to investigate the ablation rate of the nozzle

throat and the distribution of wall temperatures. The study aims to understand the influence of

combustion chamber composition, pressure, and temperature on the ablation rate of the C/C throat. The
ep

main conclusions are as follows:

(1) Under the same operating conditions, the wall temperature is highest near the throat,
Pr

corresponding to the maximum ablation rate, while the ablation rate in the expansion section tends to

zero.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
(2) With an increase in the aluminum mass fraction in the combustion chamber, a decrease in H2O

ed
and CO2 mass fraction leads to a smaller diffusion rate and a significant reduction in the throat ablation

rate.

(3) Both combustion chamber pressure and temperature are positively correlated with ablation rate

iew
of the nozzle throat. However, an increase in combustion chamber temperature has a more significant

effect on the throat ablation rate compared to total pressure.

Acknowledgements

This work is supported by the Opening Foundation of Science and Technology on Combustion,

v
Internal Flow and Thermo-Structure Laboratory, the Natural Science Foundation of Jiangxi Province of

re
China (Grant No. 20232BAB201031, Grant No. 20224BAB211010).

References

[1] G. Zhu, J. Li, K. Li, S. Cheng, Z. He, Numerical study of the impact and aggregation characteristics of alumina
er
droplets on a wall in the solid rocket motor. Aerospace Science and Technology, 2023, 137: 108242.

https://doi.org/10.1016/j.ast.2023.108242.
pe
[2] X. Meng, H. Tian, R. Yu, Y. Lu, X. Gu, G. Tan, G. Cai, Three-dimensional numerical simulation of hybrid

rocket motor based on dynamic mesh technology. Aerospace Science and Technology, 2023, 141: 108573.

https://doi.org/10.1016/j.ast.2023.108573.
ot

[3] D. Trache, T. Klapotke, L. Maiz, M. Abd-Elghany, T. DeLuca, Recent advances in new oxidizers for solid

rocket propulsion. Green Chemistry, 2017, 19(20): 4711-4736.


tn

https://doi.org/10.1039/C7GC01928A.

[4] A. Mahjub, N. Mazlan, M. Abdullah, Q. Azam, Design optimization of solid rocket propulsion: a survey of

recent advancements. Journal of Spacecraft and Rockets, 2020, 57(1): 3-11.


rin

https://doi.org/10.2514/1.A34594.

[5] G. Feng, H. Li, X. Yao, B. Ren, Y. Jia, J. Sun, Evaluation of ablation resistant ZrC-SiC multilayer coating for

SiC coated carbon/carbon composites under oxyacetylene and laser conditions. Corrosion Science, 2022, 205:
ep

110427.

https://doi.org/10.1016/j.corsci.2022.110427.
Pr

[6] T. DeLuca, Overview of Al-based nanoenergetic ingredients for solid rocket propulsion. Defence Technology,

2018, 14(5): 357-365.

https://doi.org/10.1016/j.dt.2018.06.005.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
[7] X. Wang, P. Jiang, Y. Tang, W. Zhang, S. Shi, Microstructure-based thermochemical ablation model of carbon

ed
/carbon-fiber composites. Materials, 2022, 15(16): 5695.

https://doi.org/10.3390/ma15165695.

[8] H. Wang, R. Ji, G. Xiao, Z. Qu, Pore scale visualization of thermal-fluid-structural evolution in the ablation

iew
of carbon/carbon composites. Aerospace Science and Technology, 2022, 130: 107924.

https://doi.org/10.1016/j.ast.2022.107924.

[9] J. Lachaud, Y. Aspa, G. Vignoles, Analytical modeling of the steady state ablation of a 3D C/C composite.

International Journal of Heat and Mass Transfer, 2008, 51(9-10): 2614-2627.

v
https://doi.org/10.1016/j.ijheatmasstransfer.2008.01.008.

re
[10] J. Lachaud, Y. Aspa, G. Vignoles, Analytical modeling of the transient ablation of a 3D C/C composite.

International Journal of Heat and Mass Transfer, 2017, 115: 1150-1165.

http://doi.org/10.1016/j.ijheatmasstransfer.2017.06.130.
er
[11] B. Evans, K. Kuo, P. Ferrara, J. Moore, P. Kutzler, E. Boyd, Nozzle throat erosion characterization study using

a solid-propellant rocket motor simulator. 43rd AIAA/ASME/SAE/ASEE Joint Propulsion Conference &
pe
Exhibit. 2007:5776.

https://doi.org/10.2514/6.2007-5776.

[12] W. Chen, Numerical analyses of ablative behavior of C/C composite materials. International Journal of Heat
ot

and Mass Transfer, 2016, 95: 720-726.

https://doi.org/10.1016/j.ijheatmasstransfer.2015.12.031.
tn

[13] D. Bianchi, A. Neri, Numerical simulation of chemical erosion in vega solid-rocket-motor nozzles. Journal of

Propulsion and Power, 2018, 34(2): 482-498.

https://doi.org/10.2514/1.B36388.
rin

[14] P. Thakre, V. Yang, Chemical erosion of carbon-carbon/graphite nozzles in solid-propellant rocket motors.

Journal of Propulsion and Power, 2008, 24(4): 822-833.

https://doi.org/10.2514/1.34946.
ep

[15] X. Zhang, Z. Wang, R. Wang, C. Lu, R. Yu, H. Tian, Numerical simulation of chemical ablation and mechanical

erosion in hybrid rocket nozzle. Acta Astronautica, 2022, 192: 82-96.


Pr

https://doi.org/10.1016/j.actaastro.2021.12.012.

[16] R. Geisler, C. Beckman, The history of the BATES motors at the Air Force Rocket Propulsion Laboratory.

34th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit. 1998: 3981.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551
https://doi.org/10.2514/6.1998-3981.

ed
[17] Z. Zhou, X. Wang, C. Lu, G. Le, Numerical analysis on thermal environment of liquid rocket with afterburning

under different altitudes. Applied Thermal Engineering, 2020, 178: 115584.

https://doi.org/10.1016/j.applthermaleng.2020.115584.

iew
[18] S. Evans, S. Lardeau, Validation of a turbulence methodology using the SST k-ω model for adjoint calculation.

54th AIAA Aerospace Sciences Meeting. 2016: 0585.

https://doi.org/10.2514/6.2016-0585.

[19] K. Chelliah, A. Makino, I. Kato, N. Araki, K. Law, Modeling of graphite oxidation in a stagnation-point flow

v
field using detailed homogeneous and semiglobal heterogeneous mechanisms with comparisons to

re
experiments. Combustion and Flame, 1996, 104(4): 469-480.

https://doi.org/10.1016/0010-2180(95)00151-4.

[20] P. Sutton, Rocket propulsion elements 7th edition solution manual. Services Marketing People Technology
er
Strategy, 2010, 8(2): 475-477.

[21] L. Geisler, The relationship between solid propellant formulation variables and nozzle recession rates,
pe
JANNAF Rocket Nozzle Technology Subcommittee Workshop. 1978.

[22] D. Bianchi, F. Nasuti, E. Martelli, Coupled analysis of flow and surface ablation in carbon-carbon rocket

nozzles. Journal of Spacecraft and Rockets, 2009, 46(3): 492-500.


ot

https://doi.org/10.2514/1.40197.

[23] D. Bianchi, F. Nasuti, M. Onofri, E. Martelli, Thermochemical erosion analysis for graphite/carbon-carbon
tn

rocket nozzles. Journal of Propulsion and Power, 2011, 27(1): 197-205.

https://doi.org/10.2514/1.47754.
rin
ep
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4740551

You might also like