ddb04af8c1b840713b0719d3ea80ad4c (1)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Accepted Manuscript

Numerical and experimental study of the thermochemical erosion of a graphite nozzle


in a hybrid rocket motor with a star grain

Tian Hui, Yu Ruipeng, Li Chengen, Zhao Sheng, Zhu Hao

PII: S0094-5765(18)30945-7
DOI: https://doi.org/10.1016/j.actaastro.2018.11.007
Reference: AA 7174

To appear in: Acta Astronautica

Received Date: 31 May 2018


Revised Date: 28 September 2018
Accepted Date: 7 November 2018

Please cite this article as: T. Hui, Y. Ruipeng, L. Chengen, Z. Sheng, Z. Hao, Numerical and
experimental study of the thermochemical erosion of a graphite nozzle in a hybrid rocket motor with a
star grain, Acta Astronautica (2018), doi: https://doi.org/10.1016/j.actaastro.2018.11.007.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Numerical and experimental study of the thermochemical erosion of a graphite nozzle in a


hybrid rocket motor with a star grain
Tian Hui a,b, Yu Ruipeng a,b, Li Chengen a,b, Zhao Sheng c, Zhu Hao a,b
a
School of Astronautics, Beihang University, 100191, China
b
Key Laboratory of Spacecraft Design Optimization & Dynamic Simulation Technologies, Ministry

PT
of Education, China
c
Research and Development Center, China Academy of Launch Vehicle Technology, Beijing 100076,

RI
China

*Corresponding author: tianhui@buaa.edu.cn (H. Tian).

SC
Abstract

Hybrid rocket motors is a promising propulsion system because of its intrinsic advantages over a
conventional solid rocket motor and liquid rocket engine. However, serious nozzle erosion is a key

U
problem that prevents hybrid rocket motors from being widely used, especially for propulsion systems
with long operating times. In this paper, the erosion of a graphite-based nozzle coupled with a
AN
combustion flow field is studied in a hybrid rocket motor with a star grain. As the oxidizer and fuel,
90% hydrogen peroxide and hydroxide-terminated polybutadiene are adopted, respectively. The
nozzle erosion was simulated coupled with the flow field in a typical hybrid rocket motor through
M

three-dimensional numerical simulations. The simulations are based on a pure-gas steady numerical
model considering turbulence, fuel pyrolysis, oxidizer/fuel reactions, thermal conduction and solid-gas
boundary interactions on the fuel and nozzle surfaces. The results indicate that the nozzle erosion is
D

greatly influenced by the inner flow field. The flame near the grain trough is thicker than that near the
TE

grain peak. Therefore, the maximum erosion rate (0.042 mm/s) occurs near the nozzle throat
corresponding to the grain trough. The OH and H2O contribute 49.8% and 45.5% to the erosion rate,
respectively, in this area. Furthermore, 56.6% and 31.9% contributions are made by OH and H2O,
respectively, in the area corresponding to the grain peak. The O, CO2 and O2 make much lower
EP

contributions to the total erosion. In addition, a firing test is carried out to characterize the graphite
nozzle erosion on a full-scale hybrid rocket motor with star grain. The nozzle inner profiles before and
after test show that the erosion behavior of the graphite material is strictly related to the fuel shape.
C

Key word: Hybrid rocket motor; Erosion rate; Star grain


AC

Nomenclature

Variables
A Arrhenius pre-exponential constant
c* characteristic exhaust velocity
d diameter of nozzle throat
E activation energy
H sensible enthalpy
m& mass flow rate

1
ACCEPTED MANUSCRIPT

Nc number of species in the erosion reaction,


P pressure
R universal gas constant
T temperature
r net reaction rate
rk the net rate for the surface reaction k

PT
r& regression rate
S inward coordinate normal to nozzle surface
t time

RI
v velocity
vki the stoichiometric coefficients for reactant i in reaction k
Y mass of specie

SC
y+ dimensionless wall distance
Ρ density
λ thermal conductivity

U
η outward coordinate normal to nozzle surface
ξ oxidizer-fuel ratio
AN
Subscripts
c graphite
ch chemical
M

f solid fuel
g gas phase
i reactant species
D

k reaction
ox oxidizer
TE

s nozzle surface

1 Introduction
EP

In a typical hybrid rocket motor (HRM), the oxidizer is often stored as a liquid, and the fuel is
stored as a solid. HRMs have several advantages, including propellant versatility, safety, reduced
environmental pollution and simplified throttling and shutdown compared with solid rocket motors
C

and greater simplicity and lower cost than liquid rocket motors [1,2]. However, nozzle erosion is a
challenge in the development of HRMs [3]. Typical HRMs have both nozzle and material technologies
AC

in common with solid rocket motors (SRMs). However, in HRMs, the oxidizer and fuel are stored in
different physical phases. A macroscopic diffusion flame forms during the firing process and
distributes along the fuel grain in the boundary layer [4]. Two regions are formed from the diffusion
flame in the chamber flow fields. One region is oxygen-rich, and the other is fuel-rich. The flame
structure is also affected by grain shape. Further mixing of the oxidizer and fuel occurs in the
aft-combustion chamber (ACC) [5-8]. Nozzles convert high thermal energy of chamber gases to
kinetic energy and generate thrust. However, nozzles are chemically and mechanically eroded during
motor firing due to the complex interaction between the gas flow and nozzle material. In most firing

2
ACCEPTED MANUSCRIPT

experiments, nozzle erosion may be as high as 2.5 times greater in HRMs than in SRMs [9,10]. The
more severe erosion in the HRMs is due to the greater concentration of oxidizing species in the flow
field [10-11]. Nozzle erosion during the firing process reduces the engine thrust and specific impulse,
which degrades of the motor performance.
To evaluate the erosion behavior of nozzle protection materials, numerical and experimental
works have been conducted by various groups. Bianchi and Nasuti used a computational fluid

PT
dynamics (CFD) tool with integrated gas/surface interaction modeling to describe the nozzle erosion
in hybrid rocket motors [10]. That tool considers different propellant pairs and operating conditions
under the assumption of steady-state ablation. O/F shift and combustion efficiency are found to affect

RI
the erosion rate. In a separate study, the effects of propellant combinations, O/F ratio, combustion
efficiency and other factors on the erosion rate are studied [12]. Their research reveals that erosion rate
tends to increase linearly with chamber pressure. Experimental research has been carried out to test

SC
the ablation behavior of SiC-based refractory micro-concrete nozzles [13]. The strong dependence of
the ablation phenomenon on the static temperature and the molar mass of oxidizing species at the
nozzle throat, as well as on the combustion time, is observed in the tests: an increase in these

U
parameters increases the ablated surface.
Numerical simulations have been recently carried out on nozzle erosion [4, 10]. However, neither
AN
of them did not pay much attention to the effect of three-dimensional flow fields on erosion behavior.
For the present work, the graphite nozzle erosion in a HRM with a star grain is simulated coupled with
the inner flow field. A three-dimensional numerical model is developed considering fuel pyrolysis,
M

solid-gas interface coupling and thermochemical reactions. The thermochemical reactions of H2O,
CO2, OH, O and O2 with C are all considered. In addition, a firing test of a 90% hydrogen peroxide
(HP)/hydroxyl-terminated polybutadiene (HTPB) HRM with a star grain is conducted to study the
D

nozzle erosion. The influence of the fuel type on the nozzle erosion and species distribution is
discussed based on the numerical and experimental results.
TE

2 Numerical Models

Three-dimensional numerical models are established with fluid dynamics, turbulence, solid fuel
EP

pyrolysis, thermochemical erosion of the nozzle and gas phase reactions. The numerical models and
simulation methods are described as follows. The gas-phase governing equations in the simulations
coupled the three-dimensional Navier–Stokes equations with the continuity equation, energy
C

conservation equation and species transport equations. The realizable k–ε turbulence model is used in
this paper, and it demonstrates a superior ability to capture the mean flow of complex structures
AC

compared with other measures.


2.1 Pyrolysis fuel model
1,3-Butadiene (C4H6) is treated as the main product of the HTPB pyrolysis, as indicated by
Chiaverini et al. [14]. The pyrolysis rate of the HTPB follows an Arrhenius-type law.
E
r& = A exp(− ) (1)
RT
For pure HTPB, when T>722 K, A=11.04 mm·s-1 and E=20.5 kJ·mol-1; when T<722 K, A=3965
mm· s-1 and E=55.8 kJ·mol-1 [15].

3
ACCEPTED MANUSCRIPT

Due to the non-premixed combustion in the hybrid rocket motor, a fuel-rich region exists below
the diffusion flame and an oxidizer-rich region resides above the flame. The gaseous oxidizer reacts
with the fuel pyrolysis products and a diffusion flame is formed in the boundary layer, which provides
heat to sustain further fuel pyrolysis through convection and radiation heat transfer.
As shown in [4], the mass conservation equation in the fuel surface is given by

ρ g v = ρ f r&f (2)

PT
where ρf is the gas density near the fuel surface, v is the radial velocity of the gas near the fuel surface,

and ρf and r&f are the density and the regression rate of the solid fuel, respectively.

RI
The total convective heat flux received from the gas equals the heat contributing to the
temperature rise and the phase transition of the solid fuel. The radiative heat flux can be neglected

SC
when no metal additives are added to the solid fuel. The energy conservation at the solid grain surface
is expressed as follows.

∂T

U
−λg = ρ f r&(hCTS4H6 − hHTPB
T0
) (3)
∂y +
AN
where y is the outward normal direction of the fuel surface, with the orientation pointing out the flow
T
field as the positive direction, hCS4H6 is the enthalpy of C4H6 at the fuel surface temperature TS, and
M

T0
hHTPB is the enthalpy of HTPB at the initial fuel temperature T0. The detailed derivation of Eqs. (2)

and (3) can be found in Ref. [4].


D

2.2 Chemical models of oxidizer/fuel reaction


TE

The description of the chemical reaction mechanism has a significant influence on the simulation
results, especially when different reaction products are considered. In Ref. [12], radicals (OH, H, O)
are not included in the reaction models, leading to the calculated flame temperature being higher than
EP

the result of chemical equilibrium calculations. In this paper, to simplify the calculation, it is assumed
that the liquid HP injected into the combustion chamber is fully decomposed. The evaporation and
decomposition processes are ignored. The decomposition of the HP can be expressed as
H2O2 → H2O + 0.5O2 . According to the thermodynamic calculation using the software CPROPEP1),
C

the adiabatic decomposition temperature of the 90% HP is approximately 1000 K, and the mass
AC

fraction of O2 is 0.4235.
The solid fuel in this study contains 28%Al and 10%Mg. Considering the average OF value in
the experiment is 4.6, the mass fraction of Al counted less than 5% of the total mass of propellant. The
mass and mole fractions of the products could be calculated by CPROPEP. The result is as follows:
Table 1. Mass and mole fractions of the combustion products at the inlet of nozzle
YAl2O3 YCO YCO2 YH2O YMgAl2O4 YOH YO2 Yother
Mass 0.021 0.086 0.229 0.505 0.102 0.026 0.022 0.009
Mole 0.005 0.073 0.123 0.663 0.017 0.036 0.016 0.067

4
1) Data available online at http://rocketworkbench.sourceforge.net/ [retrieved 13 October 2001].
ACCEPTED MANUSCRIPT

The thermal calculation results show that the mass and mole fractions of Al2O3 are 0.021 and
0.005, the mass and mole fractions of MgAl2O4 are 0.102 and 0.017. The mass and mole fractions of
metal species are relatively low. Therefore, the metal particle in the fuel can be ignored.
For the reaction between C4H6 and O2, there is usually a 2-step reaction mechanism considering
H2O, CO2, CO and H2 as the products [16] and a 13-step reaction mechanism considering H2O, CO2,
CO, H2, OH, H, O and O2 as products [15]. To ensure the flame temperature is within a reasonable

PT
range, the erosion caused by OH and O on the nozzle surface are considered. The CPROPEP and the
methods developed by Nicolas are used to simplify the 13-step reaction [18]. An overall reaction
model can be developed from the chemical equilibrium calculation as follows.

RI
59C4H6 + 267.5O2 → 156CO2 + 113H2O
+80CO + 47H2 + 28OH + 6H + 2O (4)

SC
The eddy dissipation model is used to calculate the reaction rate in Eq. (4) [7]. It is suitable when
the mixing time of the reactants is much longer than the reaction time, which is coincident with the
reaction mechanism in HRM.

U
2.3 Solid-gas coupling model of the nozzle surface
AN
The mass and energy conservations are considered with the coordinate system based on the
nozzle surface. η and s are, respectively, the outward and inward coordinates perpendicular to
nozzle surface.
M

According to Milos et al. [19], for non-charring thermal protection materials like C/C, the surface
mass balance of the nonequilibrium flow is

∂yi Nc
ρD = ∑ vki rk M i + ms ( yi − Ysi )
D

∂η k =1 (5)
i = 1, N c
TE

where D is the diffusion coefficient, N c is the number of species in the erosion reaction, rk is the
EP

net rate for the surface reaction k, vki are the stoichiometric coefficients for reactant i in reaction k,

Ysi is the mass of species i produced in the ablation gas per mass of the ablated graphite material
C

ablated, and Ysi is positive for erosion products and negative for atmospheric species consumed in
AC

the erosion process, with ∑Y


i
si = 1.

When the radiation is not considered, the surface energy balance is:

 ∂T  Nc
∂yi  ∂T 
λg   ∑ ρ hi D
+ = −λs   + m& s (hg − hs ) (6)
 ∂η  g i =1 ∂η  ∂s  s
The terms on the left of Eq. (7) are the heat flux due to the conduction and diffusion of the gas,
while the terms on the right are the conduction in the solid and the change of enthalpy between the

5
ACCEPTED MANUSCRIPT

products and reactants. λg is the gas thermal conductivity, hg is the enthalpy of the gas fuel at the

pyrolysis temperature, and hs is the enthalpy of the solid fuel at the pyrolysis temperature.

2.4 Thermal conduction model in the nozzle

PT
The transient conduction of the nozzle material has a significant influence on the surface
temperature and erosion rate. The specific conditions are related to the material and structure of the
nozzle. Considering the one-dimension heat conduction in the direction of the surface depth [19,20], a

RI
more comprehensive heat conduction described by vector form becomes
∂ r
( ρs hs ) = ∇ ⋅ ( s ρ s hs ) + ∇ ⋅ ( λs∇T ) (7)
∂t

SC
where the first term on the right is the energy transfer caused from the surface regression, and the
direction of s is perpendicular to the inner contour of the nozzle and points towards the inner space.

The second term on the right-hand side of Eq. (7) is the heat conduction flux, ρs is the density of the

Ur
solid fuel, λs is the thermal conductivity, s is the erosion vector of the nozzle surface and is
AN
perpendicular to the wall surface of the nozzle and points towards the inner space. The actual direction
depends on the inner contour of the nozzle.
M

2.5 Erosion model of nozzle


In this work, a validated thermochemical erosion model [4] of the C/C and graphite nozzles is
D

used to study the erosion condition in an HRM with star grain. Thermochemical erosion caused by the
heterogeneous reactions between the oxidizing components and C is studied. In solid rocket motors,
TE

the nozzle erosion includes physical erosion and thermochemical erosion [21]. However, the mass and
mole fractions of metal species are relatively low in the flow field as introduced in section 2.2. The
sublimation of the carbon, the physical erosion from the metal oxide particles and the structural failure
caused by the heat stress are ignored [4].
EP

The overall thermochemical erosion rate of the nozzle is determined from both the diffusion rate
of the oxidizing species across the boundary layer to the wall and the kinetic rate of reactions. The
controlling mechanism depends on the type of reaction, concentration of reactants, surface
C

temperature and chamber pressure. According to Ref. [10], reactions in some hybrid motors are
AC

diffusion limited, while others are kinetic limited. In this paper, the diffusion rate and kinetic rate are
both calculated, and the net reaction rate is determined by the lesser of the two.
In the calculation of the nozzle erosion in SRMs, the main species that react with the nozzle
material are H2O, CO2 and OH [22-25]. However, in HRMs, there may be local oxygen enriched
regions in the convergent section of the nozzle. Therefore, the O and O2 are also considered in this
study.
In the C/C and graphite nozzle, the kinetic-limited erosion rate of carbon and the oxidizing
species is as follows (see Refs. [20, 26]):

6
ACCEPTED MANUSCRIPT

Nc Nc
1 1
rc ,ch =
ρs
∑ Rk =
k =1 ρs
∑c ∏ P
k =1
k
i
i
vki
(8)

where Nc is the total number of heterogeneous reactions. Rk is the erosion rate caused from reaction

k, Pi is the Dalton partial pressure of reactant i, and vki are the stoichiometric coefficients for

PT
reactant i in reaction k.

 E 

RI
ck = As ,k Ts Bk exp  − a , s , k  (9)
 RuTs 
where Ts is the wall temperature, and As,k, Bk and Ea,s,k are the pre-exponential factor, the temperature

SC
exponent, and the activation energy in the Arrhenius reaction-rate expression for the kth reaction,
respectively. The heterogeneous reaction rate constants in the calculation are listed in Table 2.
Evaluated at surface temperature 2000 K and partial pressure 0.1 MPa, the reaction rates as functions

U
of pressure and temperature are shown in Figure 1 [4].
Table 2. Heterogeneous reaction rate constants for C/C [4]
AN
Reaction-rate parameters ck (kg m-2 s-1 atm-1)
Reaction rate
Reaction Ea,s,k
k Bk As,k Rk (kg m-2 s-1)
(kcal mol-1)
M

C( s ) + OH → CO + H 1 -0.5 361kg K 0.5 m -2 s -1 atm -1 0.0 R1 =k1 POH


D

C ( s )+O → CO 2 -0.5 665.5 kg K 0.5 m −2 s -1 atm -1 0.0 R2 = k3 PO


TE

C( s ) + H 2O → CO + H 2
3 0 4.80 ×105 kg m -2 s -1 atm -0.5 68.8 R3 = k3 PH0.52O
EP

4 0 9.00 ×103 kg m -2 s -1 atm -0.5 68.1 R4 = k4 PCO


0.5
2

C ( s ) + CO2 → 2CO 5 0 2.40 ×103 kg m -2 s -1 atm -1 30.0 k5 PO 2Y


R5 = + k 7 P (1 − Y )
C

1 + k6 PO 2
6 0 21.3 atm -1 -4.1
AC


7 0 0.535 kg m -2 s -1 atm -1 15.2  k 
Y = 1+ 8 
C( s ) + 1/ 2O2 → CO  k7 PO 2 
8 0 1.81×107 kg m -2 s -1 97.0

7
ACCEPTED MANUSCRIPT

PT
(a) Rk-temperature (b) Rk-pressure

RI
Figure 1. Reaction rate Rj as a function of temperature and pressure.

3 Experimental Setup

SC
The experimental system comprises of the pressurization system, oxidizer delivery system, test
platform, measurement and control system, and motor configurations. Oxidizer is supplied by using
high-pressure nitrogen gas, and the oxidizer mass flow rate is controlled by the venturi tube. The

U
pressure of the source nitrogen gas is about 28 MPa. A pressure reducer is installed between the source
gas and oxidizer tank. The oxidizer mass flow rate is determined by measuring the pressure before and
AN
after the venturi tube. If the throat of the venturi tube is in the cavitating condition, the oxidizer mass
flow rate is only controlled by the inlet pressure.

m& o = µ A 2 ρ0 ( p1 − psat )
M

(10)

where µ is the discharge coefficient, which can be acquired by the water flow calibration experiment.
A is the area of the venturi tube throat, ρ0 is the density of H2O2, p1 is the inlet pressure of the venturi
D

tube, and psat is the saturation pressure of H2O2.


The Programmable Logic Controller (PLC) controls the time sequence of the output signals,
TE

including ignition signal and electronic valves. Experimental data of pressures and thrust are gathered
by the National Instruments (NI) devices. The data acquisition system can provide more than 200
channels. The sampling frequency per channel in the test is set to 1000 per second. Several pressure
EP

sensors are used to get the pressures in oxidizer tank, before venturi tube, after venturi tube, oxidizer
chamber, and combustion chamber. The measurement range of the pressure sensor (CYB-20S,
BEIJING WESTZH MACHINER&ELECTRON TECHNOLOGY COR. Itd) is 10 MPa and the
C

precision is 0.25% FS. The test platform has a maximum capability of 40 kN thrust level rocket motor
experiment. The foundation of the test platform is sustained by two elastic reeds. A thrust sensor
AC

(BK-1) bought from China Academy of Aerospace Aerodynamics (CAAA) is settled before the motor.
The measurement range is 10 kN and the precision is 0.2% FS.

8
ACCEPTED MANUSCRIPT

PT
RI
Figure 2. Scheme of the head shell

SC
The igniter provides energy for the decomposition of H2O2 and pyrolysis of the solid fuel. In the
time sequence of the test, the igniter starts working 1s before the opening of the oxidizer feeding valve.
The oxidizer is divided into two parts (Main channel and auxiliary channel), flowing through the

U
venturi tube and injection panel and finally entering the combustion chamber. In the first stage, the
Main channel and auxiliary channel are both open. In the second stage, the main channel is open and
AN
the auxiliary channel is shut down. Figure 2 shows the details of the head shell.
The igniter provides energy for the decomposition of H2O2 and pyrolysis of the solid fuel. In the
time sequence of the test, the igniter starts working 1s before the opening of the oxidizer feeding valve.
M

At the end of the test, purging nitrogen gas is used to terminate the combustion of the motor. Table 3
shows the time sequence of the firing test.
Table 3. Time sequence in the firing test
D

No. Time (s) Name Action


1 0.0 Measurement trigger on
TE

2 18.0 Igniter firing signal on


3 18.2 Igniter firing signal off
4 19.0 Oxidizer feeding valve 1 on
EP

5 19.0 Oxidizer feeding valve 2 on


6 22.0 Oxidizer feeding valve 2 off
7 64.0 Oxidizer feeding valve 1 off
8 64.0 Purging nitrogen gas valve on
C

9 84.0 Purging nitrogen gas valve off


10 85.0 Measurement trigger off
AC

The experimental motor configuration is shown in Figure 3. The star grain consists of 60%HTPB,
28%Al, 10%Mg and 2%C. The diameters of Al, Mg and C are 2 (±0.5) µm, 20(±5)µm and 115(±15)
nm. There is a head shell, an injector plate, a pre-combustion chamber (PCC), an ACC and other
sections in the motor. The nozzle is composed of T705 graphite (detailed properties in Table 4), a nozzle
insert, a heat-insulating layer of high silica phenolic resin and a metal shell. The nozzle throat is a
cylindrical insert with a 5-mm length. Figure 4 shows the cross section of the nozzle. There are two
stages in the experiment, and the mass flow rates of HP in each stage are 4.2 kg/s and 2.0 kg/s. The

9
ACCEPTED MANUSCRIPT

engine is placed on a thrust stand equipped with a leaf spring deformation type load cell for measuring
thrust, shown in Figure 5.

PT
RI
Figure 3. Scheme of the motor configuration.

U SC
AN
M

Figure 4 Detailed diagram of the nozzle.


D
TE
C EP
AC

Figure 5. Image of the engine and the test platform.

10
ACCEPTED MANUSCRIPT

PT
RI
Figure 6 Thermochemical properties of C/C [4]

SC
The nozzle material properties are listed in Table 4, and the detailed parameters of the motor are
shown in Table 5.
Table 4. Nozzle material properties

U
T705 graphite Silica/phenolic Steel
-3
AN
Density, kg·m 1890 1670 7930
Specific heat, J·(kg·K)-1 Set as Figure 6 800 498
Thermal conductivity, W·(m·K)−1 Set as Figure 6 1 31
M

Table 5. Main parameters of the test motor


Name Value
D

Number of star tips 8


Nozzle throat diameter, mm 58
TE

Nozzle throat length, mm 5


Fuel grain outer diameter, mm 285
Fuel grain length, mm 1000
EP

Fuel grain thickness, mm 25


Divergence half-angle, degree 15
Convergence half-angle, degree 45
C

Nozzle expansion area ratio 5


Pre-chamber pressure in stage 1, MPa 3.2
AC

Pre-chamber pressure in stage 2, MPa 1.5


Time of the first stage, s 3
Time of the second stage, s 42

4 Results and discussion

Due to the symmetric structure of the motor, the half star tip is set as the computational domain,
and a structured hexahedral mesh is adopted. The thicknesses of the first layer grid near the fuel and
nozzle surface are 0.05 mm and 0.005 mm, respectively, which ensures that the dimensionless wall
distance parameter y+ of the first near-fuel cell center is approximately one [27]. The computational

11
ACCEPTED MANUSCRIPT

mesh is shown in Figure 7, and the grids of the cross sections in the middle of the grain and nozzle
throat are shown in Figure 8. A and B in Figure 8 are defined to represent the positions of the grain
peaks and troughs, respectively. In the same way, the A and B positions are also present in the
convergence section and nozzle throat.

PT
RI
U SC
Figure 7. Mesh of the computational domain.
AN
M
D
TE
EP

(a) Mesh of the cross sections in the middle of the grain; (b) Mesh of the cross section in the nozzle
Figure 8. Mesh of the cross section in the fuel grain and nozzle throat.
The nozzle erosion simulation is based on the ANSYS FLUENT platform with user-defined
C

functions (UDFs). The main process and iterations are performed using FLUENT solvers. The
heterogeneous reactions of the graphite nozzle erosion and fuel pyrolysis are conducted with UDFs,
AC

and the simulation process is shown in Figure 9 [4]. The governing equations are solved using a
steady-state solver with a second-order upwind spatial discretization, and the pressure-based coupled
algorithm is adopted.
To simplify the simulation process, a steady-state simulation of the two stages based on the initial
grain profile is conducted in the calculation. The oxidizer is injected uniformly from the injection plate
with mass flow rates of the oxidizer in the first and second stages of 4.2 kg/s and 2.0 kg/s, respectively.
The decomposed temperature of the HP is 1028 K in the calculation to match the firing test. The
surface of the fuel grain and the nozzle insert exposed to the flow field are set as the fluid-solid

12
ACCEPTED MANUSCRIPT

coupled walls. All other surfaces are set as adiabatic walls. The boundary conditions on the outlet are
acquired using numerical extrapolation.
Establish simulation models

Geometry and mesh


UDFs

Initialization User-defined
Initialization and boundary conditions
memories

PT
Solve mesh, momentum energy,
Solve user-defined functions
Species, turbulence equations

Update properties Update user-defined properties

RI
No Convergence?

SC
Yes
Export result

Figure 9. Flow chart of the simulation process [4].

U
4.1 Flowfield analysis
AN
The first and second stages have similar flow field characteristics. Therefore, only the flow field
of the second stage is discussed here. The three-dimensional fuel regression distribution contours of
the fuel surfaces are shown in Figure 10. The temperature contours can be seen in Figure 11. Due to
M

the special motor structure, a U-shape flame forms near the grain surface, and the flame vortex with
three-dimensional characteristics forms in the ACC. The distributions of the species in the flow fields
show apparent three-dimensional features, as shown in Figures 12-16. At position B, the grain surface
D

is concave so that the oxidizer has more space and time to react with the fuel. The flame at position B
is thicker than that at position A and the temperature in the regions corresponding to position B in the
TE

nozzle throat is higher than that in the regions corresponding to position A. Meanwhile, in comparison
to mass fractions of H2O and CO2 at position A, those at position B are bigger. Those results mean that
the O/F ratio at position B is closer to the stoichiometric value. The flow fields above the fuel surface
EP

is divided by the flame into the oxidizer-rich zone near the axis of motor and the fuel-rich zone near
the grain surface. The O2 and C4H6 are restricted by the flame on the oxidizer-rich and fuel-rich sides,
respectively. The H2O is the main product of the HP decomposition and C4H6/O2 reaction. As a result,
C

the mass fraction of H2O is larger in the oxidizer-rich zone and flame zone, and reduces gradually on
the fuel-rich side. The CO2, OH, O and other products are produced from the combustion reaction, and
AC

their mass fractions are large in the flame zone. Then radial diffusion of flame and combustion
products (H2O, CO2, OH, O) occurs in the ACC, which can be seen in Figures 12-15. However, the
mass fractions of oxidizer species (OH, H2O) are still smaller at position A than those at position B
because the diffusion rate is lower, shown in Figures 12 and 14. OH and H2O are the main species that
cause nozzle erosion, which is inferred in Figure 28. As a result, the nozzle erosion rate at position B
is higher that would be discussed in the following. The temperature distribution inside the high silica
phenolic resin of the nozzle convergence section is shown in Figure 17. The temperature at position B
of the convergence section is approximately 100 K higher than that at position A due to the flame
structure.

13
ACCEPTED MANUSCRIPT

PT
RI
Figure 10 Regression contour of the fuel surface.

U SC
AN
M

Figure 11. Temperature contours in the inner flow field.


D
TE
EP
C

Figure 12. Mass fraction of H2O in the inner flow field.


AC

14
ACCEPTED MANUSCRIPT

PT
RI
Figure 13. Mass fraction of CO2 in the inner flow field.

U SC
AN
M

Figure 14. Mass fraction of OH in the inner flow field.


D
TE
EP
C

Figure 15. Mass fraction of O in the inner flow field.


AC

15
ACCEPTED MANUSCRIPT

PT
RI
Figure 16. Mass fraction of O2 in the inner flow field.

U SC
AN
M

Figure 17. Temperature contour of the nozzle surface in the nozzle convergence section.
D

4.2 Thermochemical erosion of the graphite nozzle


Figure 18 presents the pre-chamber pressure (Pc) and the thrust (F) curves from the test. The
TE

pressure of the first stage is approximately 3.2 MPa. When the oxidizer mass flow rate decreases to 2
kg/s, the pressure and thrust of the second stage reduces with time. The pressure remains
approximately 1.5 MPa but has a small downward trend. The combustion chamber pressure peak in
EP

the tailoff of the motor is due to the higher purging gas pressure.
C
AC

Figure 18. Experimental curves of the test motor.


The temperature contours and erosion rates in the nozzle throat insert are shown in Figures 19-20.
The flame near position B makes the local erosion more severe than that of position A in the nozzle

16
ACCEPTED MANUSCRIPT

throat insert. Thus, a high-temperature gradient forms between position A and position B, which is the
main factor influencing the erosion distribution.

PT
RI
SC
Figure 19. Temperature contour in the nozzle insert.

U
AN
M
D

Figure 20. Erosion contour in the nozzle insert.


The species distributions at the cross section of the nozzle entrance in the S2 are shown in Figure
TE

21. H2O and O2 have higher mass fractions near the axis. This is consistent with the conclusion in
Figures 12-16. At position A, the species mass fractions change more rapidly to stable values in radial
direction when compare to those at position B.
C EP
AC

(a) Species mass fraction at position A (b) Species mass fraction at position B
Figure 21. Species mass fractions at the nozzle entrance at position A and B.

17
ACCEPTED MANUSCRIPT

The temperature and pressure distributions in the first stage (S1) and second stage (S2) are shown
in Figures 22-23. A high thermal gradient forms between the nozzle insert surface and the
heat-insulating surface due to different settings for the boundary conditions, as shown in Figure 22. The
nozzle insert surface is set as a coupled surface, and the heat-insulating surface of high silica phenolic
resin is set as an adiabatic wall. Therefore, the surface temperature of the nozzle insert is lower than that
of the heat-insulating surface of the high-silica phenolic resin due to the thermal conduction inside the

PT
graphite insert. In the axial direction of the nozzle, the surface temperature of the nozzle insert increases
in the convergence section while remaining stable in the nozzle throat. The temperature at position A is
lower than that at position B, and the pressure distribution along position A is similar to that along

RI
position B, as shown in Figure 23.
The nozzle erosion rates along positions A and B in the two stages are shown in Figure 24. The
maximum erosion rate occurs in the convergence section of the nozzle. In S1, the erosion rates of the

SC
nozzle throat along A and B are in the range of 0.053 mm/s~0.051 mm/s and 0.096 mm/s~0.091 mm/s,
respectively. Accordingly, in S2, the erosion rates of nozzle throat along A and B are in the range of
0.023 mm/s~0.020 mm/s and 0.042 mm/s~0.039 mm/s, respectively.

U
Circumferentially uneven ablation occurs in both the simulation and experiment. The simulated
erosion rates in the front of the nozzle throat in S1 and S2 are shown in Figure 25 and were used to
AN
calculate the final simulation throat diameter. In S1, the erosion rates in positions A and position B are
set to 0.053 mm/s and 0.096 mm/s, respectively. In S2, the erosion rates are set to 0.023 mm/s and
0.042 mm/s in positions A and B, respectively. In this case, considering that the first stage lasts 3 s and
M

the second stage lasts 42 s, the average throat diameter will be 60.28 mm and 62.27 mm in positions A
and B, respectively. The calculated shape of the throat after burning can be seen in Figure 26.
In the experiment, the erosion rate is determined by
D

d 2 − d1
r&s = (11)
TE

2t
where d1 and d2 are the initial inner diameter and final inner diameter of the nozzle throat, respectively,
and t is the working time of the motor.
In the experiment, the average diameter of the nozzle throat increased from 58 mm to the final
EP

diameter in the range of 59.4 mm~63.7 mm. Because the working time is 45 s, the mean erosion rate
is 0.016 mm/s~0.063 mm/s. The comparison of the experimental and simulation results is also shown
in Figure 27. The erosion condition is more scattered in the experiment. At positions 1 and 2 in Figure
C

27, the erosion conditions for the firing test are more severe than those for the simulation. At position
3 in Figure 27, the situation is the opposite. Metal ablation, asymmetric fuel injection and combustion
AC

could lead to an asymmetric inner flow field and asymmetric nozzle erosion in an experiment.
Figure 28 shows the erosion rates caused by different species in the S2, with OH and H2O being
the main oxidant species contributing to the nozzle erosion. At position B, the OH and H2O contribute
49.8% and 45.5% to the total erosion rate, respectively. Additionally, 56.6% and 31.9% contributions
are made by the OH and H2O at position A. O, CO2 and O2 make much lower contributions to the total
erosion. The concentrations of H2O, OH, and CO2 are high near the nozzle throat. However, the
reaction rate Rj of CO2 is smaller than that of OH [7] when the pressure is below 1.5 MPa and the
temperature is below 2100 K. Therefore, CO2 has less influence on the erosion.

18
ACCEPTED MANUSCRIPT

PT
RI
Figure 22. Temperature distribution in the nozzle insert.

U SC
AN
M

Figure 23. Pressure distribution in the nozzle insert.


D
TE
EP
C

Figure 24 Erosion rate distribution in the nozzle insert.


AC

19
ACCEPTED MANUSCRIPT

PT
RI
(a) Erosion rate distribution in S1 (b) erosion rate distribution in S2
Figure 25 Erosion rate in the nozzle.

U SC
AN
M

Figure 26 Throat radius before and after the erosion.


D
TE
EP
C
AC

Figure 27 Comparison between the experiment and simulation results.

20
ACCEPTED MANUSCRIPT

PT
RI
Figure 28 Erosion rate distribution by species.

SC
4.3 Oxidizer-Fuel ratio and combustion efficiency
The O/F ratio are also calculated by a method titled “Nozzle-Throat Reconstruction Technique”
(NTRT) [28]. The equations used in the method are as follows:

U
The thrust F equation is:

F =λue m& + ( Pe − Pa ) Ae (12)


AN
in which λ, ue, m& , Pe, Pa, and Ae are the thrust correction factor, nozzle-exit velocity, propellant
mass-flow rate, exit pressure, atmospheric pressure, and nozzle-exit cross-sectional area, respectively.
M

Theoretical characteristic exhaust velocity c* and specific-heat ratio γ are solved by CPROPEP.

c* = f ( Pc , ξ ) (13)
D

γ =g ( Pc , ξ ) (14)
TE

& is calculated for a given oxidizer mass-flow rate m& ox by


The propellant mass-flow rate m

1
m& = m& ox (1 + )
EP

(15)
ξ
The nozzle throat area is calculated for a given c* and efficiency η∗ by
C

m&η * c *
At = (16)
Pc
AC

The value for theoretical exit pressure Pe is determined implicitly from Eq. (17):

 Pe  γ +1  Pe  Ae γ + 1 1 (1−γ )
  1 − ( ) (γ − 1) γ = ( ) (17)
 Pc  γ −1  Pc  At 2
And theoretical exit velocity ue is calculated explicitly by
(γ +1) ( γ −1)
2γ 2  2   Pe (γ −1)/ γ 
ue = c * 1 − ( ) 
γ − 1  γ + 1 
(18)
 Pc 

21
ACCEPTED MANUSCRIPT

Lastly, overall fuel-mass consumption Mf is calculated by

m& ox
Mf =∫ dt (19)
ξ

PT
RI
SC
Figure 29 Oxidizer mass flow rate and OF in the experiment

U
Figure 29 shows the oxidizer mass flow rate and oxidizer-fuel ratio in the experiment. The
oxidizer mass flow rate between the two stages has a 5-seconds transition section. O/F ratio increases
AN
slowly in two stages. However, there is a step change of O/F ratio in the transition of experiment
stages. At the end of the experiment, the O/F value reaches its maximum, 11.0. The average O/F value
of the test is 4.6 and the O/F ratio of the steady state simulation is 4.55. The combustion efficiency is
M

0.88 in the calculation. The erosion rate reduces with the decrease of O/F value [10]. Due to the
varying O/F in the experiment, the calculated erosion rate in the experiment is quite different from that
in the simulation. This is an important reason causing the erosion rate difference between the
D

experiment and simulation. The simulation model would be extended to transient state and consider
the metal particle. Figure 31 shows the calculation process of the NTRT method [28].
TE
C EP
AC

Figure 30 NTRT solutions for nozzle-throat-erosion history

22
ACCEPTED MANUSCRIPT

Upload Experimental Data


m& ox (t ), F (t ), Pc (t ), M f , Rt, f

Loop A
test η∗

Loop B
test ξ(t)

PT
& ox (t ), ξ (t ), Pc (t )
Input: m
1111111111111
Eqs. (13)-(15)

RI
m& (t ), c *(t ), γ (t )
Output: 222222222222

m& (t ), c *(t ), Pc (t ), η *
Input: 11111111111111111

SC
Eq. (16)
Output: At(t)

Input: Pc(t),γ(t), At(t)

U
Eq. (17)
AN
Output: Pe(t)

Input: c*(t), Pc(t),γ(t), Pe(t)


Eq. (18)
M

Output: ue(t)

& (t ), ue (t ), Pe (t )
Input: m
222222222222
D

Eq. (12)
Output: Fcal(t)
TE

NO

Fcal (t ) − F (t ) < F ×10−5


EP

YES
NO
t =t
C

f
t = t + ∆t
YES
AC

NO
Rt (tf ) − R < R ×10−4
t, f t, f

YES

Solutions: ξ(t), η*

Figure 31 Flowchart of the NTRT method [28].

23
ACCEPTED MANUSCRIPT

It is hard to measure the final “area” in the experiment. The nozzle throat radius is measured by
an inside micrometer at 13 different radial angles (every thirty in 0-360 degrees). The average value
(62.67 mm) of the measured data is set as the final “radius” to be the convergence criterion of NTRT
method. The nozzle throat radius history calculated by NTRT method is illustrated in Figure 30. To
clearly show the changing process of the nozzle throat radius, the NTRT result is smoothed by the
Adjacent-Averaging method. As the figure shows, the process of nozzle erosion consists of three

PT
periods. Of the 3 periods, period (1) is the first stage of the test and period (2) and (3) belongs to the
second stage. The rapid rise (0.249 mm/s) of the nozzle throat radius happens in the period (1) and (2),
which is caused by the high oxidizer mass flow rate. Meanwhile, the average nozzle erosion rate of

RI
period (3) is low (0.061 mm/s) which results from the small oxidizer mass flow rate. From 30s to 45s,
the nozzle throat radius rarely rises, and after 45s in the solution, the nozzle throat radius increases at a
constant erosion rate. The increase of erosion rate in period (3) is owing to the upward trend of O/F

SC
value. A vibration appears in the nozzle throat radius history at the end of the test due to the similar
oscillations of pressure and O/F values.
4.4 Uncertainty Analysis

U
The uncertainty in the NTRT method and the calculation of nozzle erosion rate is determined by
AN
analyzing the bias introduced from experimental measurements. The overall uncertainty Uy is
calculated from the equation as follows:
∂f
δ y 2 = ∑( δ xi )2
M

∂xi
(20)
∂f ∂f ∂f ∂f
= ( δ x1 ) 2 + ( δ x2 ) 2 + ( δ x3 )2 + ... + ( δ xn ) 2
∂x1 ∂x2 ∂x3 ∂xn
D

in which xi represents an experimental measurement and the δ is the uncertainty in the measurement.
TE

The uncertainties in pressure and thrust measurements are determined as follows:

δ y 2 = b2 + υ 2 (21)
EP

in which b represents the precision limit of the pressure sensor or load cell, and υ is a measure of
variance between the filtered and unfiltered data calculated by Eq. (22):
N
1
υ2= ∑ (x − x )2
C

i fil ,i (22)
N i =1
AC

in which xi represents the unfiltered pressure or thrust data, and xfil,i represents the value of the filtered
data at the corresponding time. Thus, the uncertainty in oxidizer mass-flow rate can be calculated
according to Eq. (23):
∂ m& ox ∂ m& ox
δ m2& = ( δ p )2 + ( δ p )2 =
ox
∂p1 1
∂psat sat

(23)
∂ m& ox ∂ m& ox
( bp21 + υ p21 ) 2 + ( bp2sat + υ p2sat )2
∂p1 ∂psat

24
ACCEPTED MANUSCRIPT

The pressure sensor precision is 0.25%. Then the uncertainty of oxidizer mass-flow rate would be
0.91% g. The uncertainty in fuel-mass consumption is the precision limit of the digital scale which is
0.02kg. The nozzle radius is measured by an inside micrometer and repeated 15 times, and the
precision of measuring tool is 0.01mm. The uncertainty of erosion rate can be calculated by Eq. (24):

∂r&s ∂r&
δ r&2 = ( δ d1 ) 2 +( s δ d2 )2 (24)
s
∂d1 ∂d 2

PT
The δd1 and δd2 are both 0.00258mm, and the uncertainty of erosion rate is 0.000157 mm/s.

RI
5 Conclusions

Both a numerical simulation and firing test are conducted to study the thermochemical erosion of
the graphite nozzle in a 90% HP/HTPB hybrid rocket motor. The three-dimensional geometry of the

SC
model is based on a typical hybrid rocket motor that includes a pre-combustion chamber, an
aft-combustion chamber and a nozzle. The firing test employed two stages with a high flow rate of 4.2
kg/s and a low flow rate of 2.0 kg/s. A steady state simulation based on the initial fuel type is

U
conducted in the simulation process. The heterogeneous reaction of H2O/CO2/OH/O/O2 with C are
considered to study the nozzle erosion. The erosion behavior of the nozzle material is presented based
AN
on the simulation results. The following main conclusions have been drawn.
1) Under the conditions of the fuel type and the oxidizer mass flow rate in this study, the
oxidizing radicals, especially OH, cannot be ignored in the simulation of nozzle erosion. At position B,
M

the OH and H2O contribute 49.8% and 45.5% to the total erosion rate, respectively. Additionally,
56.6% and 31.9% contributions are made by the OH and H2O at position A. Although the mass
fraction of CO2 is large near the nozzle throat, the CO2/C reaction rate is relatively slow due to the
D

special reaction mechanism.


2) The flame at position B of the grain surface is thicker and the temperature in the regions
TE

corresponding to position B in the nozzle throat is higher than those in the regions corresponding to
position A. Meanwhile, in comparison to mass fractions of H2O and CO2 at position A, those at
position B are bigger. Those results are because that the O/F ratio at position B is closer to the
EP

stoichiometric value. The mass fractions of oxidizer species (OH, H2O) are smaller at position A than
those at position B because the radial diffusion rate at aft-combustion chamber is lower. Therefore, the
nozzle erosion rate at position B is higher than that at position A.
C

3) Both the experimental and numerical results show that the flow fields and nozzle erosion
present obvious correlations with the star grain. The flame that gathers in the area facing the grain
AC

peaks is thicker than that in the area facing the grain troughs, which can be seen in the simulation
results. Accordingly, the surface temperature of the regions near the nozzle insert corresponding to the
grain peaks is approximately 100 K higher than that facing the grain troughs in the simulation results.
As a consequence, the erosion rate along position B in the nozzle insert is greater than that along
position A.
4) The erosion rate in the nozzle is more scattered in the firing test than in the simulation results.
The maximum erosion rate in the experiment is higher than that in the simulation, and the minimum

25
ACCEPTED MANUSCRIPT

erosion rate is smaller in the experiment. This could have been the result of the asymmetric fuel
injection, asymmetric combustion or metal particle ablation in the experiment.
5) The NTRT solutions demonstrate that O/F value increases slowly in two stages of the
experiment. Meanwhile, there is a step change of O/F ratio in the transition of experimental stages.
Due to the varying O/F in the experiment, the calculated erosion rate in the experiment is quite
different from that in the simulation. Simultaneously, in the NTRT result, it is obvious that the nozzle

PT
erosion is more serious in the stage of lager oxidizer mass flow rate. In the case of small mass flow
rate, O/F value would greatly influence the nozzle erosion rate.

RI
References

1 D. ALTMAN AND A. HOLZMAN. Overview and history of hybrid rocket propulsion. In: KUO
K K, CHIAVERINI M. Fundamentals of Hybrid Rocket Combustion and Propulsion. Virginia;

SC
AIAA, 2007:1-36.
2 N.A. Davydenko, R.G. Gollender, A.M. Gubertov. et al. Hybrid rocket engines: the benefits and
prospects. Aerospace Science and Technology, 2007, 11: 55-60.

U
3 K. K. Kuo, C. Martin. Challenges of hybrid rocket propulsion in the 21st century. In: KUO K K,
CHIAVERINI M, eds. Fundamentals of Hybrid Rocket Combustion and Propulsion. Virginia;
AN
AIAA, 2007: 593-638.
4 S. Zhao, H. Tian, P. Wang, et al. Steady state coupled analysis of flowfields and thermochemical
erosion of C/C nozzles in hybrid rocket motors. Science China Technological Sciences, 2015,
M

58(3): 574-86.
5 G. A. Marxman, C. E. Wooldridge, R. J. Muzzy. Fundamentals of hybrid boundary layer
combustion. Heterogeneous Combustion Conference. American Institute of Aeronautics and
D

Astronautics, 1963.
6 H. Tian, X. Li, N. Yu, et al. Numerical and experimental investigation on the effects of aft mixing
TE

chamber diaphragm in hybrid rocket motor. Science China Technological Sciences, 2013, 56(11):
2721-31.
7 X. Li, H. Tian, G. Cai. Numerical analysis of fuel regression rate distribution characteristics in
EP

hybrid rocket motors with different fuel types. Science China Technological Sciences, 2013,
56(7): 1807-17.
8 G. Cai, P. Zeng, X. Li, et al. Scale effect of fuel regression rate in hybrid rocket motor. Aerospace
C

Science and Technology, 2013, 24(1): 141-6.


9 R. C. Bunker, A. Prince. Hybrid rocket motor nozzle material predictions and results. 28th
AC

AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit. Nashville, TN; AIAA, 1992.
10 D. Bianchi, F. Nasuti. Numerical analysis of nozzle material thermochemical erosion in hybrid
rocket engines. Journal of Propulsion and Power, 2013, 29(3): 547-58.
11 E. J. Wernimont, S. D. Heister. Characterization of fuel regression in hybrid rockets utilizing
hydrogen peroxide oxidizer. 31st Joint Propulsion Conference and Exhibit. San Diego, CA;
AIAA, 1995.

26
ACCEPTED MANUSCRIPT

12 D. Bianchi, F. Nasuti. CFD analysis of hybrid rocket flowfields including fuel pyrolysis and
nozzle erosion. 49th AIAA/ASME/SAE/ASEE Joint Propulsion Conference. San Jose, CA;
AIAA, 2013.
13 R. D'Elia, G. Bernhart, J. Hijlkema, T. Cutard. Experimental analysis of SiC-based refractory
concrete in hybrid rocket nozzles. Acta Astronautica, 2016,126:168-77.
14 M. J. Chiaverini, G. C. Harting, et al. Pyrolysis behavior of hybrid-rocket solid fuels under rapid

PT
heating conditions. Journal of Propulsion and Power, 1999, 15(6): 888-95.
15 G. Lengelle. Solid-Fuel Pyrolysis Phenomena and Regression Rate, Part 1: Mechanisms.
Fundamentals of Hybrid Rocket Combustion and Propulsion. American Institute of Aeronautics

RI
and Astronautics, 2007: 127-66.
16 S. Venkateswaran, C. Merkle. Size scale up in hybrid rocket motors. 34th Aerospace Sciences
Meeting and Exhibit. Reno, NV; American Institute of Aeronautics and Astronautics. 1996.

SC
17 G. C. Cheng, R. C. Farmer, H. S. Jones, et al. Numerical simulation of the internal ballistics of a
hybrid rocket motor. 32nd Aerospace Sciences Meeting and Exhibit. Reno, NV; American
Institute of Aeronautics and Astronautics, 1994.

U
18 N. Bellomo, M. Lazzarin, F. Barato, et al. Numerical investigation of the effect of a diaphragm
on the performance of a hybrid rocket motor. 46th AIAA/ASME/SAE/ASEE Joint Propulsion
AN
Conference & Exhibit. Nashville, TN; American Institute of Aeronautics and Astronautics, 2010.
19 F. S. Milos, D. J. Rasky. Review of numerical procedures for computational surface
thermochemistry. Journal of Thermophysics and Heat Transfer, 1994, 8(1): 24-34.
M

20 R. Acharya, K. K. Kuo. Effect of pressure and propellant composition on erosion rate of graphite
rocket nozzle. Journal of Propulsion and Power, 2007, 23(6): 1242-54.
21 Y. M. Wang, X. Xiong, et al. Thermal Shock and Ablation Behavior of Tungsten Nozzle
D

Produced by Plasma Spray Forming and Hot Isostatic Pressing, 2015, 24 (6):1026-1037.
22 P. Thakre, V. Yang. Chemical erosion of carbon-carbon/graphite nozzles in solid-propellant
TE

rocket motors. Journal of Propulsion and Power, 2008, 24(4):822-33.


23 D. Bianchi, A. Turchi, F. Nasuti, et al. Coupled CFD analysis of thermochemical erosion and
unsteady heat conduction in solid rocket nozzles. 48th AIAA/ASME/SAE/ASEE Joint
EP

Propulsion Conference & Exhibit. Atlanta, Georgia; American Institute of Aeronautics and
Astronautics, 2012.
24 D. Bianchi, F. Nasuti, E. Martelli. Coupled analysis of flow and surface ablation in
C

carbon-carbon rocket nozzles. Journal of Spacecraft & Rockets, 2009, 46(3): 492-500.
25 D. Bianchi, F. Nasuti, M. Onofri, et al. Thermochemical erosion analysis for
AC

graphite/carbon-carbon rocket nozzles. Journal of Propulsion and Power, 2011, 27(1): 197-205.
26 H. K. Chelliah, A. Makino, I. Kato, et al. Modeling of graphite oxidation in a stagnation-point
flow field using detailed homogeneous and semiglobal heterogeneous mechanisms with
comparisons to experiments. Combust Flame, 1996, 104(4): 469-80.
27 Z. X. Gao, C. W. Jiang, C. H. Lee. Improvement and application of wall function boundary
condition for high-speed compressible flows. Science China Technological Sciences, 2013, 56:
2501–2515.

27
ACCEPTED MANUSCRIPT

28 K. Landon, S. Yuji, K. Ryosuke, et al. Method for Determining Nozzle-Throat-Erosion History in


Hybrid Rockets. Journal of Propulsion and Power, 2017, 33(6): 1369-1377.

PT
RI
U SC
AN
M
D
TE
C EP
AC

28
ACCEPTED MANUSCRIPT
Coupled simulations of flow field and nozzle erosion are conducted.

The effects of grain structure on nozzle erosion are studied with star-shape grain.

Heterogeneous reactions are considered to study the graphite nozzle erosion.

A numerical model for nozzle erosion in hybrid rocket motor is analyzed and tested.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like