Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

Journal Pre-proof

Numerical and experimental investigation of throttleable hybrid rocket motor with


aerospike nozzle

Zhu Hao, Hui Tian, Zihao Guo, Liu Hedong, Chengen Li

PII: S1270-9638(20)30665-9
DOI: https://doi.org/10.1016/j.ast.2020.105983
Reference: AESCTE 105983

To appear in: Aerospace Science and Technology

Received date: 26 February 2019


Revised date: 18 February 2020
Accepted date: 29 April 2020

Please cite this article as: Z. Hao et al., Numerical and experimental investigation of throttleable hybrid rocket motor with aerospike
nozzle, Aerosp. Sci. Technol. (2020), 0, 105983, doi: https://doi.org/10.1016/j.ast.2020.105983.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and
formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early visibility of the article. 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.

© 2020 Published by Elsevier.


Numerical and Experimental Investigation of Throttleable Hybrid

Rocket Motor with Aerospike Nozzle

Corresponding author: Zhu Hao

School of Astronautics, Beihang University, Beijing 100191, China

zhuhao@buaa.edu.cn

First author:Hui Tian

School of Astronautics, Beihang University, Beijing 100191, China

tianhui@buaa.edu.cn

Second author:Zihao Guo

School of Astronautics, Beihang University, Beijing 100191, China

guozihao@buaa.edu.cn

Third author:Liu Hedong

School of Astronautics, Beihang University, Beijing 100191, China

lhdhnd@163.com

Forth author:Chengen Li

School of Astronautics, Beihang University, Beijing 100191, China

lichengen@buaa.edu.cn

1
Abstract:

Hybrid rocket motors (HRMs) easily achieve variable thrust by changing oxidizer

mass flow rate. This paper represents numerical and experimental studies on the

throttleable HRM with aerospike nozzle, which can enhance the performances of the

throttleable HRM theoretically. By comparing the performance of the throttleable HRM

with the aerospike nozzle and that with the de Laval nozzle, the characteristics of the

aerospike nozzle applied to the throttleable HRM are explored. Both numerical and

experimental investigations are conducted. In the simulation part, characteristic

velocity, combustion efficiency, thrust coefficient and specific impulse of the HRM

with the two types of nozzles at different oxidizer mass flow rate are obtained.

Comparing with the HRM with the de Laval nozzle, the HRM with the aerospike nozzle

can improve specific impulse by about 8.9%~23.4% when the working pressure ratio

is lower than the design pressure ratio. The combustion efficiency of the throttleable

HRM with the aerospike nozzle is 3.9%~8.6% higher than that with the de Laval nozzle

because of the core structure of the aerospike nozzle. When the oxidizer mass flow rate

is 50 g/s, the thrust coefficient of the aerospike nozzle is increased by 12.3% relative to

that of the de Laval nozzle which flow separation occurs in. The corresponding firing

tests are performed by the lab-scale motor with 90% hydrogen peroxide (HP) and

2
polyethylene (PE) grain. The numerical and experimental results agree well. The

combination of the throttleable HRM and the aerospike nozzle has the advantages to

improve the performance of HRMs.

Keywords: Hybrid rocket motor; Aerospike nozzle; Numerical simulation;

Experimental study.

Nomenclature

A Arrhenius pre-exponential factor

Ar area expansion ratio

c* characteristic velocity

CF thrust coefficient

E activation energy

F thrust

h heat transfer coefficient

Is specific impulse

Ma Mach number

m (fuel) mass consumption

m mass flow rate

O/F oxidizer-to-fuel mass ratio

p pressure

R (nozzle) radius

r regression rate

3
T temperature

t combustion time

u velocity

γ specific-heat ratio

ρ density

λ thermal conductivity

η efficiency

¯ overbar, to indicate a time-averaged value

Subscripts

a atmosphere

c combustor chamber

e nozzle-exit plane

f fuel

g gas

ini initial

itf iteration

ox oxidizer

s solid

sim simulation

th Theoretical

1. Introduction

4
Hybrid rocket motor (HRM) is a rapid-developing propulsion technology that owns

many advantages over conventional liquid rocket engine and solid rocket motor[1][2],

such as low cost, safety and simplified throttling. The HRM is becoming more and more

popular, and it is considered as a promising technology[3][4]. The HRM can simply

adjust thrust by changing the oxidizer mass flow rate[5][6][7]. However, the chamber

pressure of the throttleable HRM changes when the oxidizer mass flow rate is adjusted.

The throttleable HRM with the de Laval nozzle is over-expanded at a small chamber

pressure and thrust. The performance loss of de Laval nozzle is inevitably in that

condition.

With the altitude compensation effect, the aerospike nozzle can maintain high

performance in a wide altitude range[8]. It has the advantages of compact structure and

small size. When the nozzle-exit pressure is equal to the ambient pressure, the working

pressure ratio pc/pe is the design pressure ratio. The altitude compensation performance

is related to the design pressure ratio. When the working pressure rate pc/pe is not lower

than the design pressure ratio, the thrust characteristics of the aerospike nozzle are the

same as those of the conventional de Laval nozzle; when the working pressure ratio

pc/pe is lower than the design pressure ratio, the aerospike nozzle possesses attitude

compensation characteristics due to the special structure and aerodynamic

phenomenon[9].

Combining the HRM with the aerospike nozzle is a new development field.

California State University of Technology used the oxidizer (N2O) of the HRM to cool

5
the aerospike nozzle for reusable[10]. The University of Washington proved that the

aerospike nozzle can be well integrated with the HRM through firing tests[11]. Arizona

State University conducted a firing test on the HRM with an aerospike nozzle and found

that the thrust coefficient of the HRM with an aerospike nozzle was 6.4% higher than

that of HRM with a de Laval nozzle in over-expanded state[12]. Utah State University

designed the HRM with the aerospike nozzle for small satellites; the HRM used

oxidizer (N2O) to regeneratively cool the aerospike nozzle, and thrust vector adjustment

was achieved through the secondary injection; the firing test results showed that the

HRM had thrust vector adjustment capability and the regenerative cooling system

worked normally[13].

Our previous work analyze the effects of aerospike nozzle structure on HRM

performance through simulation research [14]. Nozzles with three different expansion

ratios were selected, corresponding to design conditions of high altitude and ground.

The above study did not involve the combination of the aerospike nozzle and the

throttleable HRM. The attitude compensation performance of aerospike nozzle makes

the HRM maintain high nozzle performance when the working pressure ratio is lower

than the design pressure ratio. Therefore, this paper explores the impact of the

performance of the throttleable HRM with the aerospike nozzle.

Through numerical simulations and test researches, this paper compares the

performance of the throttleable HRM with aerospike nozzle and de Laval nozzle. The

aerospike nozzle and the de Laval nozzle are designed with the same throat area and

6
area expansion ratio. Numerical simulations of the throttleable HRMs with the

aerospike nozzle and the de Laval nozzle are carried. The parameters such as

combustion efficiency, thrust coefficient and specific impulse are obtained. On this

basis, firing tests are carried out. The laboratory scale motors are tested under different

oxidizer mass flow rate with 90% hydrogen peroxide (HP) and tubular polyethylene

(PE) grain.

2. Numerical simulations

2.1. Geometry models and numerical models

The geometry model of the lab-scale motor with aerospike nozzle and the geometric

parameters is shown in Fig. 1, and the unit is mm. The geometry model and the

geometric parameters of the HRM with de Laval nozzle are the same as that of the HRM

with aerospike nozzle except the nozzle. The throat area and area expansion ratio of the

de Laval nozzle are the same as that of the aerospike nozzle. The throttleable HRM

adjusts thrust by changing the oxidizer mass flow rate. The variation range of the

oxidizer mass flow rate is between 50 g/s and 500 g/s. Table 1 shows the simulation

cases of the oxidizer mass flow rate. The letter A represents the HRM with the aerospike

nozzle and L represents the HRM with the de Laval nozzle. The stable working pressure

of the HRM increases with oxidizer mass flow rate. When the oxidizer mass flow rate

is about 250 g/s, the chamber pressure is about 1.8 MPa according to the previous

research[5]. The two types of nozzles are expected to be fully expanded on the ground

7
under this chamber pressure. The nozzle-exit Mach number and area expansion ratio of

the aerospike nozzle are acquired by

γ −1
2  pc γ 
Ma = ( ) ( ) − 1 (1)
γ − 1  pa 
γ +1
1  1 γ − 1 2  γ −1
Ar = ( )(1 + Ma )  (2)
Ma  γ + 1 2 
where γ = 1.192, pa = 0.1 MPa, pc = 1.8 MPa, Ma = 2.46 and Ar = 3.48. The plug

expansion section of the aerospike nozzle is designed according to the simplified design

method of Gianfranco Angelino[15]. The length and relative angle of the characteristic

line of the expanded wave cluster are obtained through the method. Therefore, the shape

of plug expansion section is determined. Fig. 2 shows simplified diagram of the

aerospike nozzle. R1 and R2 are the radial of the drop-shaped plug and the ring throat

at the throat section, respectively. θ is the angle between the throat section and the

centerline, which is equal to 40.5°. The radius of the base is 1.56 mm. The outer

diameter and the inner diameter of the plug straight section are 12 mm and 25mm,

respectively. The throat area of the aerospike nozzle is designed to be equal to that of

the de Laval nozzle. The area expansion ratio of the aerospike nozzle and the de Laval

nozzle are both 3.48. Fig. 3 shows the 3D structure of the aerospike nozzle which is

used in the firing test. The outer wall and the plug of the nozzle are connected by four

ribs. The ribs are made of stainless steel with a width of 8 mm. An adiabatic plate is

placed in front of the plug and the ribs.

8
Fig. 1. Cutaway sketch of a lab-scale motor.

Fig. 2. Partial structure diagram of aerospike nozzle.

Fig. 3. Structure of aerospike nozzle.

9
Table 1 The simulation cases setting.

No. m ox (g/s) No. m ox (g/s)

A1 50 L1 50
A2 100 L2 100
A3 150 L3 150
A4 200 L4 200
A5 250 L5 250
A6 300 L6 300
A7 350 L7 350
A8 400 L8 400
A9 450 L9 450
A10 500 L 10 500

The two-dimensional (2D) axisymmetric mesh model is shown in Fig. 4, which is

similar to the model in ref.[14]. The mesh of the HRM with the aerospike nozzle ignores

the plug connection structure. The three-dimensional (3D) mesh model of the HRM

with the aerospike nozzle is designed, which verifies the error of calculation results of

the simplified 2D model. The size of external flow field of the 3D mesh model is the

same as the 2D axisymmetric mesh model, whose length is 600mm and radius is

300mm. In order to make the mesh have good orthogonalization, the computational

domain grid uses structured meshes. The meshes are clustered near the fuel surface and

walls to meet the requirement of the turbulence model for the numerical simulation.

The 3D flow field computational domain grid of the HRM with the aerospike nozzle is

demonstrated in Fig. 5. It contains 1141967 cells.

10
External Flow Field

Oxidizer Nozzle
Inlet Fuel Surface

Internal Flow Field Axis

(a) Mesh of the HRM with aerospike nozzle.

External Flow Field

Oxidizer
Inlet Fuel Surface Nozzle

Internal Flow Field Axis

(b) Mesh of the HRM with de Laval nozzle.

Fig. 4. The two-dimensional (2D) axisymmetric mesh model

(a) Mesh of integrated flow field.

11
(b) Mesh detail of nozzle flow field

(c) Symmetrical surface 1


Y

Z X

(d) Symmetrical surface 2.

Fig. 5. Mesh of the HRM with aerospike nozzle.

12
2.2. Numerical models

Steady-state numerical simulations are conducted through ANSYS Fluent platform

with the User defined functions (UDFs). The 2D numerical models are the same as the

previous paper[14], which is omit here. The 3D numerical models are introduced below.

2.2.1. Governing equation

The gas-phase governing equations in the simulations couple the three-dimensional

Navier-Stokes equations with transport equations and turbulence model equations. The

realizable k-ε turbulence model is employed in the equations. The gas governing

equations can be described as:

∂ρ ∂
+ ( ρ ui ) = H ρ (3)
∂t ∂xi

∂ ∂ ∂
( ρ ui ) + ( ρ ui u j + pt δ ij ) = τ ij + H ui (4)
∂t ∂xi ∂xi

∂ ∂ ∂  ∂T 
e+ [(e + pt )ui ] =  uiτ ij + λ  + He (5)
∂t ∂xi ∂xi  ∂xi 

∂ ∂ ∂  ∂Y 
( ρYm ) + ( ρ uiYm ) =  ρ Dm i  + H Ym (6)
∂t ∂xi ∂x  ∂xi 

where H contains the source terms combustion, turbulence and user defined sources, pt

is the effective pressure expressed as pt = p + (2/3) ρk, Ym is the mass fraction of the

mth species and τij is the viscous tensor.

2.2.2. Turbulence model

The model transport equations for the turbulence kinetic energy k and its dissipation

rate ε in the realizable k-ε turbulence model[16][17] can be expressed as follows.

13
∂ ∂
(ρk ) + ( ρ kui )
∂t ∂xi
(7)
∂  μt ∂k 
= (μ + )  + Gk + Gb − ρε − Ym + S k
∂x j  σ k ∂x j 
∂ ∂
( ρε ) + ( ρε u j )
∂t ∂x j
(8)
∂  μt ∂ε  ε2 ε
=  ( μ + )  + ρ C S ε − ρ C + C1ε C3ε Gb + Sε
∂x j  σ ε ∂x j  1 2
k + vε k

where μt is the turbulence viscosity, C1 = max [0.43, ], η = S , S= 2𝑆 𝑆 .

2.2.3. Chemical reaction model

The oxidizer is assumed to be completely decomposed to GOX and water vapor when

it passes through the catalyst bed. The standard molar formation enthalpy of PE is

-41.2 KJ/mol. The gaseous ethylene (C2H4) is considered to be the pyrolysis product of

PE pyrolysis when the fuel is heated[18][19][20]. The gaseous ethylene is injected

from the fuel surface[19]. The chemical reaction equation between C2H4 and O2 is

presented as follows

38C 2 H 4 +101O 2 → 36CO+40CO 2 +58H 2 O+20OH+6H 2 +8O+4H (9)

The combustion in HRM is diffusion combustion. The mixing time of reactions is

much longer than the reaction time. Eddy dissipation model is used to calculate the

reaction rate.

2.2.4. solid-gas coupling model

The pyrolysis of the solid fuel grain is a complex fluid-solid coupling process which

satisfies the law of mass conservation and the law of energy conservation[21]. The mass

conservation equation at the surface of the solid fuel grain can be described as

14
ρ g v = − ρs r (10)

where ρg is the density of the fuel pyrolysis gas, v is the velocity of the fuel pyrolysis

gas and ρs is the density of the PE fuel.

The energy balance equation at the fuel grain surface can be described as

Q conv,in + Q rad,in = Q conv,out + Q pyr,chg + Q rad,out (11)

Considering that there is no metal contained in the fuel in this study, the radiation

could be safely ignored[21][22]. Then

∂T ∂T
-λg = -λf + ( - ρg vhg − ρ f rh
 f,itf ) (12)
∂y g.itf
∂y f.itf

The heat transferred into the solid fuel grain can be determined as

∂T
λf = ρf cp r (Tf,itf − Tf,ini ) (13)
∂y

The energy balance equation[23] can be described as

∂T
-λ g = ρ s r ( hC 2 H 4 ,Ts − hs,ini ) (14)
∂y g.itf

where λg is the thermal conductivity of the gas product, hC 2H4 ,Ts


is the enthalpy of

the C 2 H 4 at the grain surface temperature and hs,in i is the enthalpy of the fuel at the

reference temperature (300 K).

The regression rate of PE fuel is related to the temperature of the grain surface, which

can be described by following Arrhenius equation


E

r = A e RTs
(15)

where A is the pre-exponential factor, E is the activation energy, R is the universal gas

constant and Ts is the surface temperature of the fuel grain. The value of A is 2678.1

15
m/s and the value of E is 125604.0 J/mol[24][25]. The fuel grain surface temperature

and the fuel regression rate can be calculated by Eq. (14) and Eq. (15) in the UDFs.

2.2.5. boundary conditions

The 90% HP is assumed to have completely decomposed. Therefore, the oxidizer

inlet is defined as the mass flow inlet of GOX and water vapor at the temperature of

1028 K. The atmospheric boundary condition of the external flow field of the pressure

and the temperature is 0.1 MPa and 300 K, respectively. The surface of the solid fuel

grain is defined as solid-gas interface, which is governed by the solid-gas coupling

model. Other solid surfaces including the end face of the grain are defined as adiabatic

and no-slip walls.

2.3. Simulations and discussions

Both 2D and 3D simulations of the HRM with the aerospike nozzle are carried out.

The connection structure of four ribs is omitted in the 2D simulations. Subsequent

discussions indicate that the difference of results between the simplified 2D simulations

and the 3D simulations is very small. In order to facilitate the analysis and comparison

of flow field differences, 2D simulations of the HRM with the aerospike nozzle are

selected for comparison with the simulation of the HRM with the de Laval nozzle. Case

A1, A5, A9, L1, L5 and L9 are chosen to study the flow field characteristics. The

performance of the throttleable HRM with the aerospike nozzle and the de Laval nozzle

is discussed.

2.3.1 Flow field characteristics

16
Fig. 6 displays the temperature contours of the two motors. The upper half of every

figure belong to the HRM with the aerospike nozzle, and the lower part of the figures

belong to the HRM with the de Laval nozzle. The temperature distribution between two

motors in the pre-chamber and the grain gas passage portion is basically similar. The

fuel pyrolysis products mix and react with the oxidizer stream near the grain wall. A

thin layer near the grain wall is the center of the combustion reaction, whose

temperature is significantly higher than other parts. Because the aerospike nozzle

structure is special, the temperature distribution in the post-chamber and the nozzle is

different.

Temperature: 400 700 1000 1300 1600 1900 2200

A1

L1

(a) Temperature contours of A1 and L1

17
(b) Temperature contours of A5 and L5

(c) Temperature contours of A9 and L9

Fig. 6 Temperature contours of the two motors.

18
Fig. 7 presents the flow field characteristic of the post-chamber and the nozzle more

clearly. Since the motor temperature contour and the stream lines of different oxidizer

mass flow are basically similar, case A1 and L1 are chosen to study the flow field

characteristics of two motors in post-chamber and nozzle. The combustion gas reflux

forms vortex in post-chamber. The vortex improves the mixing degree and the residence

time of the gas in the chamber. The average temperature of the post-chamber is higher

than that of the central flow from Fig. 7.

From the stream lines of the post-chamber and the nozzle in Fig. 7, there is a vortex

in the plug straight section. The plug straight section increases the residence time of the

combustion gas and provides a place for further reaction. For the case of the HRM with

the de Laval nozzle, the core oxidizer directly flows out of the nozzle. This is

determined by the diffusion combustion of the HRM. The average temperature in the

post-chamber and nozzle inlet of the HRM with the aerospike nozzle is always higher

than that with the de Laval nozzle. The core structure of the aerospike acts as a spoiler

to improve the reaction degree of the gas and increase the average temperature of the

gas in post-chamber. In Fig. 7a, gas flow separation emerges in the de Laval nozzle at

oxidizer mass flow rate of 50 g/s. In case L1, the chamber pressure is low, and the

nozzle expansion ratio is relatively too large. The nozzle is serious over-expansion,

which causes gas flow separation. The shock wave generates inside the nozzle

expansion section. After the shock wave, the temperature of the gas flow increases

significantly. At the same time, outside air enters the nozzle. Therefore, the

19
phenomenon of temperature variation roughly from 2200 K to 300 K inside the de Laval

nozzle appears in Fig. 7.

(a) case A1 and L1

(b) case A5 and L5

(c) case A9 and L9

Fig. 7 Flow field temperature contours and the stream lines of the post-chamber and

nozzle.

20
2.3.2. Combustion efficiency

Details of simulation results are summarized in Table 2 and Table 3. η c * is

combustion efficiency of the HRM which is acquired by

*
csim
ηc* = (16)
cth*

where c th* is the theoretical characteristic velocity, which is obtained by thermal

calculation software RPA1) (Rocket Propulsion Analysis). csim


*
is acquired by

pc ⋅ At
*
csim = (17)
m

where m is the sum of oxidizer mass flow rate and fuel mass flow rate. Fig. 8a shows

the relationship between the characteristic velocity and oxidizer mass flow rate of the

2D simulations. Fig. 8b shows the relationship between the combustion efficiency and

oxidizer mass flow rate of the 2D simulations. When the oxidizer mass flow rate

increases, the O/F ratio is further away the optimal O/F ratio which value is about 6.8,

so the characteristic velocity of the motor is reduced. The aerospike nozzle does not

affect the combustion in the grain passage, so the O/F ratio of the two motors are equal.

At the same oxidizer mass flow rate, the cth* of the two motors are almost the same.

The average chamber pressure of the HRM with the aerospike nozzle is always higher

than that with the de Laval nozzle, so the csim


*
of the HRM with the aerospike nozzle

is always greater than that of the HRM with the de Laval nozzle.

1) data available online at http://www.propulsion-analysis.com/index.htm.

21
Table 2 Summary of two-dimensional simulation results.

NO. m ox pc O /F F c th* *
csim Is CF η c*

(g/s) (MPa) (N) (m/s) (m/s) (m/s)

A1 50 0.477 8.33 82.88 1575.52 1505.23 1480.02 0.9832 0.955


A2 100 0.877 10.52 179.66 1517.44 1415.33 1640.69 1.1592 0.932
A3 150 1.259 12.0 279.22 1476.97 1369.13 1718.31 1.2550 0.927
A4 200 1.641 12.90 381.91 1453.97 1345.66 1772.20 1.3170 0.925
A5 250 2.012 13.81 481.95 1432.33 1326.19 1797.64 1.3555 0.925
A6 300 2.378 14.49 581.43 1417.18 1310.35 1813.01 1.3836 0.924
A7 350 2.739 15.15 680.04 1403.32 1297.30 1822.68 1.4050 0.924
A8 400 3.097 15.68 777.78 1392.64 1286.22 1827.91 1.4212 0.923
A9 450 3.456 16.18 874.90 1383.06 1278.21 1831.09 1.4325 0.924
A10 500 3.813 16.61 972.10 1375.25 1271.11 1833.80 1.4427 0.924
L1 50 0.434 8.33 67.16 1574.61 1369.54 1199.36 0.8757 0.869
L2 100 0.799 10.52 154.35 1517.06 1289.45 1409.55 1.0931 0.850
L3 150 1.161 12.00 251.82 1476.81 1262.56 1549.67 1.2274 0.854
L4 200 1.526 12.90 350.70 1453.88 1251.36 1627.37 1.3005 0.860
L5 250 1.885 13.81 448.16 1432.28 1242.48 1671.61 1.3454 0.867
L6 300 2.243 14.49 545.24 1417.15 1235.96 1700.16 1.3756 0.872
L7 350 2.598 15.15 641.83 1403.29 1230.52 1720.27 1.3980 0.876
L8 400 2.951 15.68 737.86 1392.63 1225.58 1734.10 1.4149 0.880
L9 450 3.302 16.18 833.41 1383.04 1221.25 1744.27 1.4283 0.883
L10 500 3.654 16.61 928.92 1375.24 1218.10 1752.35 1.4386 0.885

Table 3 Summary of three-dimensional simulation results of HRM with aerospike

nozzle.

NO. m ox pc O /F F c th* *
csim Is CF η c*

(g/s) (MPa) (N) (m/s) (m/s) (m/s)

A1-3d 50 0.479 8.33 82.44 1575.56 1511.54 1472.15 0.9739 0.959


A2-3d 100 0.882 10.53 179.20 1517.46 1423.40 1636.54 1.1497 0.938
A3-3d 150 1.265 12.00 278.48 1476.98 1375.66 1713.74 1.2458 0.931
A4-3d 200 1.650 12.90 381.19 1453.98 1353.04 1768.87 1.3073 0.931
A5-3d 250 2.021 13.81 481.23 1432.33 1332.12 1794.96 1.3474 0.9300
A6-3d 300 2.389 14.49 580.73 1415.66 1316.82 1811.38 1.3756 0.9302
A7-3d 350 2.757 15.15 680.14 1403.32 1305.82 1822.66 1.3960 0.9304
A8-3d 400 3.118 15.69 777.51 1392.65 1294.94 1827.29 1.4111 0.9298

22
A9-3d 450 3.480 16.19 874.85 1383.06 1287.08 1831.00 1.4226 0.9306
A10-3d 500 3.838 16.61 971.43 1375.25 1279.44 1832.54 1.4323 0.9303

(a) (b)

Fig. 8 Characteristic velocity (a) and Combustion efficiency (b)

of two-dimensional simulation.

The combustion of oxidizer and fuel are diffusion combustion in the HRM.

Compared to premixed combustion, diffusion combustion requires more time for

oxidizer and fuel to fully react. When the mass flow rate is relatively small, the oxidizer

flow velocity is slow. There is more residence time for propellant in the chamber. The

plug of the aerospike nozzle has spoiler effect, which is favorable for the reaction to

proceed sufficiently. The combustion efficiency of case A1 and A2 is obviously

improved. When the oxidizer mass flow rate is greater than 150 g/s, the O/F ratio is

much larger than the optimum O/F ratio. The combustion efficiency tends to be stable.

The combustion efficiency of the HRM with the aerospike nozzle is significantly higher

than that with de Laval nozzle and always higher than 0.92, which is attributed to the

core structure of the aerospike nozzle.

23
With the increase of the oxidizer mass flow rate, the combustion efficiency of the

HRM with the de Laval nozzle decreases first and then increases. When the oxidizer

mass flow rate is about 100 g/s, the combustion efficiency of the HRM with the de

Laval nozzle is at the minimum value. When the oxidizer mass flow rate is small, the

diffusion of the fuel pyrolysis product into the oxidizer mainstream has a great influence

on the combustion efficiency. As the oxidizer mass flow rate increases, the diffusion

effect decreases, so the combustion efficiency decreases. When the oxidizer mass flow

rate is larger, temperature of the combustion gas in the post-chamber gets higher. The

higher temperature indicates that the blending degree between the fuel and the oxidizer

is improved. Therefore, the combustion efficiency increases correspondingly.

Under the same oxidizer mass flow rate, the combustion efficiency of the HRM with

the aerospike nozzle is 3.9%~8.6% higher than that of the HRM with the de Laval

nozzle because of the core structure of the aerospike nozzle.

Fig. 9 shows 2D and 3D simulation temperature flow field contours for oxidizer mass

flow rate of 50 g/s. The temperature contours show that the average temperature of the

post-chamber of the 3D simulation is higher than that of the 2D simulation. The 2D

simulation ignores the connection structure of four ribs, which can increase the degree

of gas mixing in the post-chamber.

Fig 10 shows the results of the characteristic velocity and the combustion efficiency

of the 3D simulation and 2D axisymmetric flow field simulation of the HRM with the

aerospike nozzle. The simulation characteristic velocity and combustion efficiency of

24
the 3D flow field are slightly larger than that of the 2D axisymmetric flow field. The

plug connection structure of the aerospike nozzle can slightly improve the characteristic

velocity and combustion efficiency of the HRM.

Fig. 9 Flow field temperature contours of 2D and 3D simulation

(a) (b)
Fig 10. Characteristic velocity (a) and Combustion efficiency (b) of simulation results
of the HRM with aerospike nozzle

2.3.3. Thrust coefficient

F is acquired by

F = ( pe − pa ) Ae +m ⋅ ue (18)

where pa is the ambient pressure and Ae indicates the nozzle-exit area. CF is acquired
by

25
F
CF = (19)
pc ⋅ At

Fig. 11 shows the relationship between the thrust coefficient of the HRM with the

two types of nozzles and oxidizer mass flow rate. As can be seen in Fig. 7a, the flow

separation occurs in the de Laval nozzle of case L1. The difference between the thrust

coefficient of case A1 and L1 is as large as 12.3% of the thrust coefficient of the HRM

with the de Laval nozzle. The flow separation reduces performance of de Laval nozzle.

Due to the altitude compensation characteristic, the aerospike nozzle can theoretically

be considered in the fully-expanded state. Therefore, the aerospike nozzle is superior to

the de Laval nozzle in these cases.

When the oxidizer mass flow rate is less than 250 g/s, the de Laval nozzle is in over-

expansion state. The thrust coefficient of the aerospike nozzle is much larger than that

of the de Laval nozzle especially when the oxidizer mass flow rate is 100 g/s and

50 g/s. When the oxidizer mass flow rate is greater than 250 g/s, the two types of nozzles

are both in the incomplete-expanded state. Since the chamber pressure of the HRM with

aerospike nozzle is stronger than that of the HRM with de Laval nozzle at the same

oxidizer mass flow rate, the working pressure ratio of the HRM with the aerospike

nozzle is greater than that of the HRM with the de Laval nozzle. Therefore, the thrust

coefficient of the HRM with the aerospike nozzle is slightly larger.

26
Fig. 11 Thrust coefficient of two-dimensional simulation.

Fig. 12 shows the results of the thrust coefficient of the 3D simulation and 2D

axisymmetric flow field simulation of the HRM with the aerospike nozzle. The

connection structure of four ribs in the aerospike nozzle reduces the velocity of gas

entering the nozzle. As shown in the Fig. 12, the thrust coefficient of the 3D simulation

is slightly smaller than that of the 2D simulation.

Fig. 12 Thrust coefficient of simulation results of the HRM with aerospike nozzle

27
2.3.4. Specific impulse

In Table 2 and Table 3, Is is acquired by


I s =csi* m C F (20)

Fig. 13 shows the relationship between the specific impulse of the HRM with the two

types of nozzles and oxidizer mass flow rate. The specific impulse of the HRM with

the two types of nozzles increases with the increase of the oxidizer mass flow rate, and

the growth rate gradually decreases. When the oxidizer mass flow rate is the same, the

specific impulse of the HRM with the aerospike nozzle is always greater than that of

the HRM with the de Laval nozzle. The specific impulse of the HRM with the aerospike

nozzle is 8.9%~23.4% higher than that of the HRM with the de Laval nozzle when the

working pressure ratio is lower than the design pressure ratio.

Fig. 13 Specific impulse of two-dimensional simulation.

3. Experimental research

3.1. Experimental setup

28
Fig. 14 shows the test platform (a), the lab-scale motor (b) and the firing test (c). The

lab-scale motor and the aerospike nozzle have been introduced in 2.1 geometry model.

The de Laval nozzles are made by C-C composite material, whose area expansion ratio

is 3 and throat diameter is 15 mm. The throat area of the aerospike nozzle and the de

Laval nozzle are the same. The area expansion ratio of the aerospike nozzle is 3.48. The

drop-shaped plug and the ring throat are made of copper infiltrated tungsten. The outer

wall of the nozzle and the plug is connected by four ribs. The width of the four ribs is

8 mm. An insulation plate is placed in front of the plug and the ribs. 90% HP is catalyzed

by the catalytic bed. The motor is ignited by catalytic products of 90% HP. The oxidizer

mass flow rate is controlled by a venturi tube during the firing test. The fuel grain is PE

with a density of 950 kg/m3. The nitrogen purge valve is opened 0.5 s after closing the

oxidizer supply valve, and the residual 90% HP in the pipeline is blown into the

combustion chamber. The combustion chamber and the nozzle are cooled to facilitate

the next test.

(a) (b)

29
(c)

Fig. 14 Test platform (a), assembled lab-scale motor (b) and firing test (c).

The measure system adopts pressure and thrust sensors to acquire the pressure and

thrust dates. The percentage errors of pressure and thrust sensors are 0.2% and 0.1%

respectively. The oxidizer mass flow rate is measured by a flowmeter which percentage

error is 0.1%. The weight of the fuel grain is measured by an electronic weight scale

which error is 1 g. The programmable logic controller (PLC) is selected to control the

valves in the test.

3.2. Results analysis

3.2.1 Experimental phenomena

The aerospike nozzle used in case A1~A3 have no obvious ablation while the

aerospike nozzle used in case A4~A6 has a certain degree of ablation. The de Laval

nozzles have no obvious ablation. The typical experimental traces of the firing test (A4,

L4) are presented in Fig. 15. According to the chamber pressure curve, the working

process of the HRM is divided into five sections, including valve response section,

30
catalytic bed startup section, combustion pressure building section, chamber pressure

stability section and combustion after-effect section, represented by s0, s1, s2, s3 and s4

respectively.

Fig. 15 Experimental traces of firing test.

It is assumed that the pyrolysis of the fuel grain and the combustion of the propellant

only exist on s2, s3 and s4, so the analysis is based on these sections. In section s4, after

the nitrogen purge valve opens, the residual HP in the pipeline is blown into the

combustion chamber which reacts with the fuel. Table 4 shows the test conditions and

results, where Δmf is the fuel grain mass loss after the firing test, Δmf_a is the fuel grain

mass consumption in s4, m f_w is the average fuel mass flow rate and O/F is the average

mass flow ratio of oxidizer to fuel in s2 and s3. If Δmf_a is estimated, the fuel grain

consumption in the working section (s2, s3) can be obtained. The pressure of the

nitrogen supply per test is in the range of 3.0~3.8 MPa. The amount of residual HP in

each firing test is assumed the same, so the fuel grain mass consumed in s4 is

substantially the same. Δmf_a is acquired by

t1

Δmf_a =

tw
pc dt ⋅ Δmf
(21)
t1

t0
pc dt

31
The two fuel grains of case A1 and case A5 are tested three times and twice

respectively. For case A1, the time set in the first two tests is not long enough to ignite

the fuel grain, so the third test is carried out. For case A5, there is a sensor measuring

point abnormality, so the tests are carried out twice. In case A1 and case A5, the value

fuel grain mass consumption in s4 of each test needs to be divided by the number of

test times. It can be seen from Table 4 that the fuel grain mass consumption in s4 of

each test is between 15.7 g and 20.2 g. Therefore, the estimation of Δmf_a is considered

reasonable. m f_w is acquired by

Δmf - Δmf_a
m f_w = (22)
t w - t0

The fuel grains of the same oxidizer mass flow rate are compared after the test. The

fuel grains consumption at the end face of the fuel grain with the aerospike nozzle is

larger than that of the HRM with the de Laval nozzle at the same oxidizer mass flow

rate. Table 4 shows that the fuel mass flow rate of the HRM with the aerospike nozzle

is larger than that with the de Laval nozzle.

Table 4 Summary of firing test results.

NO. pc m o_w Δmf Δ m f_a m f_w O /F c* cth* η c*

(MPa) (g/s) (g) (g) (g/s) (m/s) (m/s)

A1 0.58 53.7 101 54.7 6.9 7.77 1532.48 1590.07 0.964


A2 1.09 99.0 142 21.0 9.7 10.20 1459.07 1528.02 0.955
A3 1.76 156.0 148 19.9 15.9 9.80 1565.04 1542.71 1.014
A4 1.90 180.8 125 17.9 17.1 10.57 1483.20 1518.99 0.976
A5 2.23 211.0 265 38.3 20.5 10.75 1444.23 1513.75 0.954
A6 2.56 255.6 176 16.6 26.9 9.88 1467.49 1541.39 0.952
L1 0.48 53.7 126 20.2 5.5 9.76 1128.92 1536.55 0.735

32
L2 0.87 100.0 95 16.2 8.2 12.21 1137.84 1470.59 0.774
L3 1.38 153.4 110 16.9 13.2 11.60 1220.24 1488.21 0.820
L4 1.67 186.9 126 17.9 17.1 10.95 1297.10 1507.27 0.861
L5 1.99 222.1 134 16.8 20.4 10.86 1317.92 1510.20 0.873
L6 2.34 234.1 168 15.7 24.9 11.30 1294.84 1494.53 0.866

According to Table 4 , the chamber pressure of the HRM with the aerospike nozzle

is significantly higher than that of the HRM with the de Laval nozzle under the same

working condition. The total mass flow rate of the HRM with the aerospike nozzle and

the de Laval nozzle is basically equal in the contrast cases. The smallest gap between

the drop-shaped plug and the ring throat is the annular throat of the aerospike nozzle,

which is only about 2.1 mm. The throat area of the aerospike nozzle is assumed to be

smaller than the design area during the test, which causes the chamber pressure of the

HRM with the aerospike nozzle to become larger. The throat area of the aerospike

nozzle is acquired by

π ( R22 − R12 )
At = (23)
sin θ

where R1 and R2 are marked in Fig. 2, θ is the angle between the nozzle throat and the

centerline, which equals 40.5°. The machining uncertainty of At is acquired by

ΔAt 2( R2 ΔR2 +R1ΔR1 )


= (24)
At R22 − R12

where the machining accuracy of ΔR1 and ΔR2 is 0.05 mm and 0.02 mm, respectively.

The calculated machining uncertainty of At is 4.98%. The left side of Fig. 16 is a top

view of the aerospike nozzle after firing tests and the right side of the figure is the ring

throat. There is a layer of blue object covering the surface of the ring throat, which is

considered to be the product of adiabatic ablation. The liquid copper film forms on the

33
surface during firing test. These two effects on the ventilating area of the nozzle are

ignored. The machining uncertainty of At of the de Laval nozzle is acquired by

ΔAt 2ΔR
= (25)
At R

where the machining accuracy of ΔR is 0.02 mm and throat radius R equals 15 mm, so

the machining uncertainty of At is 0.27%.

Fig. 16 The aerospike nozzle after firing test.

3.2.2. Combustion efficiency

In Table 4, η c is the combustion efficiency, which is acquired by


*

c*
ηc = * (26)
cth*
tw

c =
*  t0
pc dt ⋅ At
(27)
(m f_w + m o_w ) ⋅ (tw - t0 )
Δc* is the uncertainty of c*, which is acquired by
tw

Δc ∗
=
t0
Δpc dt
+
ΔAt Δm
+ (28)
c∗ tw
At m
t0
pc dt

34
The percentage uncertainty of c* of the HRM with the aerospike nozzle and the de

Laval nozzle is 5.3% and 0.57%, respectively. From Fig. 17, the characteristic velocity

and combustion efficiency of the HRM with the aerospike nozzle are significantly

greater than that of the HRM with the de Laval nozzle. The combustion efficiency of

the HRM with the aerospike nozzle is about 8%~23% higher than that of the HRM with

the de Laval nozzle. The chamber pressure of the HRM with the aerospike nozzle is

higher than that of the HRM with the de Laval nozzle. One reason is that the fuel mass

flow rate and characteristic velocity of the HRM with aerospike nozzle are greater than

that of the HRM with the de Laval nozzle. The other reason is that the actual throat area

of the aerospike nozzle becomes smaller than the designed value. The smaller throat

area leads to large calculated values of test characteristic velocity and combustion

efficiency. Considering the influence of the smaller throat area, the combustion

efficiency of the HRM with aerospike nozzles is not such a large increase than that of

the HRM with the de Laval nozzle.

(a) (b)

Fig. 17 Characteristic velocity (a) and Combustion efficiency (b) of testing results

3.2.3. thrust coefficient and specific impulse

35
In Table 5, CF is acquired by

F
CF = (29)
At ⋅ pc

where F and pc is the average thrust and the chamber pressure of the test stable

section. Since the thrust sensor of some cases is not working normally, the thrust

coefficient of those cases in Table 5 are obtained. ΔCF is the uncertainty of CF, which

is acquired by

ΔF ⋅ pc F ⋅ Δpc F ⋅ pc ⋅ ΔAt
ΔCF = + + (30)
At At At2

ΔCF ΔF Δpc ΔAt


= + + (31)
CF F pc At

The percentage uncertainty of CF of the aerospike nozzle and the de Laval nozzle is

5.3% and 0.57%, respectively.

Table 5 Test results of thrust coefficient and specific impulse.

NO. F (N) CF I s p (m/s) O/F

A1 100.2 0.9768 1599.8 7.77


A2 —— —— —— ——
A3 —— —— —— ——
A4 429.1 1.2748 1855.4 10.57
A5 506.8 1.2870 1833.4 10.75
A6 613.1 1.3573 1953.9 9.88
L1 —— —— —— ——
L2 —— —— —— ——
L3 309.2 1.2663 1497.2 11.60
L4 396.2 1.3438 1700.9 10.95
L5 469.6 1.3375 1720.1 10.86
L6 —— —— —— ——

36
Fig. 18 Thrust coefficient with the variation of combustion chamber pressure.

Fig. 18 shows the variation of thrust coefficient with the oxidizer mass flow rate. The

simulation thrust coefficient is also shown in the figure as a comparison. The test thrust

coefficient of the aerospike nozzle is consistent with the simulation results. Since the

simulation does not consider the end face burning of the grain, the test thrust coefficient

should be higher than the simulation result. The test thrust coefficient of the de Laval

nozzle is higher than the simulation results. The uncertainty of the thrust coefficient of

the aerospike nozzles is calculated to be 5.3%. The main uncertainty source of the thrust

coefficient is the machining uncertainty of the throat area. The actual throat area of the

aerospike nozzle is assumed to be the upper limit of the machining error range. When

the oxidizer mass flow rate is about 50 g/s, the thrust coefficient of the aerospike nozzle

is consistent with the simulation results. When the oxidizer is higher than 150 g/s, the

test thrust coefficient of the aerospike nozzle has little advantage over the test thrust

coefficient of the de Laval nozzle.

37
In Table 5, I sp is acquired by
tw

Isp =
t0
Fdt
(32)
(m f_w + m o_w ) ⋅ (tw − t0 )

Fig. 19 shows the variation of average specific impulse with the oxidizer mass flow

rate. The simulation specific impulse is also shown in the figure as a comparison. The

test average specific impulse of two HRMs is consistent with the trend of the simulation

specific impulse. Since the simulation does not consider the end face burning of the

fuel grain, the O/F ratio of the test cases is smaller than the simulation cases. The O/F

ratio of the test cases are closer to the optimal O/F ratio than that of the simulation cases,

so the test average specific impulse is generally higher than the simulation results as

shown in Fig. 19. The test average specific impulse of case L3 is lower than the

simulation results. O/F ratio of case L3 is slightly larger, which causes the test average

specific impulse of case L3 to be small. In general, the test specific impulse of the HRM

with the aerospike nozzle is significantly higher than that of the HRM with the de Laval

nozzle, especially when the oxidizer flow rate is small.

Fig. 19 Specific impulse with the variation of combustion chamber pressure.

38
4. Conclusion

This paper focuses on performance comparison of the throttleable HRM with the

aerospike nozzle and that with the de Laval nozzle, including comparison of

characteristic velocity, combustion efficiency, thrust coefficient and specific impulse.

Conclusions can be drawn as follows:

(1) The simulation results show that the combustion efficiency of the throttleable HRM

with the aerospike nozzle is 3.9%~8.6% higher than that with the de Laval nozzle.

The test combustion efficiency of the HRM with the aerospike nozzle is about

8%~23% higher than that of the HRM with the de Laval nozzle. because of the core

structure of the aerospike nozzle. The aerospike nozzle with the core plug structure

can improve the combustion efficiency of HRMs.

(2) For the simulation results, when the oxidizer mass flow rate is less than 250 g/s, the

thrust coefficient of the aerospike nozzle is increased by 2%~12% compared with

that of the de Laval nozzle; when the oxidizer mass flow rate is greater than 250 g/s,

the thrust coefficient of the aerospike nozzle is slightly larger. The test thrust

coefficient agree with the simulation.

(3) The simulation results show that the specific impulse of throttleable HRM with the

aerospike nozzle is 8.9%~23.4% higher than that with the de Laval nozzle when the

working pressure ratio is lower than the design pressure ratio. The test specific

impulse of the HRM with the aerospike nozzle is significantly higher than that of

the HRM with the de Laval nozzle, especially when the oxidizer flow rate is small.

39
References

[1] C. Dunn, G. Gustafson, J. Edwards, T. Dunbrack, C. Johansen, Spatially and

temporally resolved regression rate measurements for the combustion of paraffin

wax for hybrid rocket motor applications, Aerosp. Sci. Technol. 72 (2018) 371–

379. doi:10.1016/j.ast.2017.11.024.

[2] Y. Wu, X. Yu, X. Lin, S. Li, X. Wei, C. Zhu, L. Wu, Experimental investigation

of fuel composition and mix-enhancer effects on the performance of paraffin-

based hybrid rocket motors, 83 (2018) 620–627. doi:10.1016/j.ast.2018.09.026.

[3] B. Marciniak, A. Okninski, B. Bartkowiak, M. Pakosz, K. Sobczak, W. Florczuk,

D. Kaniewski, J. Matyszewski, P. Nowakowski, D. Cieslinski, G. Rarata, P.

Surmacz, D. Kublik, D. Rysak, J. Smetek, P. Wolanski, Development of the ILR-

33 “ Amber ” sounding rocket for microgravity experimentation, Aerosp. Sci.

Technol. 73 (2018) 19–31. doi:10.1016/j.ast.2017.11.034.

[4] A. Okninski, Acta Astronautica On use of hybrid rocket propulsion for suborbital

vehicles, Acta Astronaut. 145 (2018) 1–10. doi:10.1016/j.actaastro.2018.01.027.

[5] S. Zhao, G. Cai, H. Tian, N. Yu, P. Zeng, Experimental Tests of Throttleable H

2 O 2 /PE Hybrid Rocket Motors, 51st AIAA/SAE/ASEE Jt. Propuls. Conf. (2015)

1–8. doi:10.2514/6.2015-3831.

40
[6] A. Ruffin, F. Barato, E. Paccagnella, D. Pavarin, Development of a Flow Control

Valve for a Throttleable Hybrid Rocket Motor and Throttling Fire Tests, 2018

Jt. Propuls. Conf. (2018). doi:10.2514/6.2018-4664.

[7] B. Zhao, N. Yu, Y. Liu, P. Zeng, J. Wang, Unsteady simulation and experimental

study of hydrogen peroxide throttleable catalyst hybrid rocket motor, Aerosp.

Sci. Technol. 76 (2018) 27–36. doi:10.1016/j.ast.2018.02.008.

[8] G. Hagemann, H. Immich, T. Van Nguyen, G.E. Dumnov, Advanced Rocket

Nozzles, J. Propuls. Power. 14 (1998) 620–634. doi:10.2514/2.5354.

[9] J. Ruf, P. McConaughey, J. Ruf, P. McConaughey, The plume physics behind

aerospike nozzle altitude compensation and slipstream effect, 33rd Jt. Propuls.

Conf. Exhib. (1997). doi:10.2514/6.1997-3218.

[10] P. Lemieux, W.R. Murray, Nitrous Oxide Cooled , Reusable Hybrid Aerospike

Rocket Motor : Experimental Results, (2012) 1–18.

[11] J.R. Stoffel, Experimental and Theoretical Investigation of Aerospike Nozzles

in a Hybrid Rocket Propulsion System, New Horizons. (2009) 1–11.

[12] J. Dennis, S. Shark, F. Hernandez, J.K. Villarreal, Design of a N2O/HTPB

Hybrid Rocket Motor Utilizing a Toroidal Aerospike Nozzle, 48th

AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib. 4 - 7 January 2010, Orlando,

Florida. AIAA paper (2010). doi:10.2514/6.2010-183.

[13] S.D. Eilers, S.A. Whitmorej, Z.W. Peterson, Development and testing of the

regeneratively cooled multiple use plug hybrid (for) Nanosats (MUPHyN) motor,

41
48th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib. 2012. (2012) 1–26.

doi:10.2514/6.2012-4199.

[14] H.D. Liu, H. Tian, Effects of Aerospike Nozzle Structure on Hybrid Rocket

Motor Performance, Tuijin Jishu/Journal Propuls. Technol. 39 (2018) 792–801.

doi:10.13675/j.cnki.tjjs.2018.04.009.

[15] G. ANGELINO, Approximate method for plug nozzle design, AIAA J. 2 (1964)

1834–1835. doi:10.2514/3.2682.

[16] H. Tian, Y. Li, C. Li, X. Sun, Regression rate characteristics of hybrid rocket

motor with helical grain, Aerosp. Sci. Technol. 68 (2017) 90–103.

doi:10.1016/j.ast.2017.05.006.

[17] S. Zhao, H. Tian, P.F. Wang, N.J. Yu, G.B. Cai, Steady-state coupled analysis

of flowfields and thermochemical erosion of C/C nozzles in hybrid rocket motors,

Sci. China Technol. Sci. 58 (2015) 574–586. doi:10.1007/s11431-015-5775-6.

[18] K. Das, K. Vogiatzis, G.Z. Angeli, B.C. Bigelow, W. Burgett, Computational

fluid dynamics modeling of GMT, (2018) 28. doi:10.1117/12.2312858.

[19] G.D. Di Martino, C. Carmicino, R. Savino, Transient computational

thermofluid-dynamic simulation of hybrid rocket internal ballistics, J. Propuls.

Power. 33 (2017) 1395–1409. doi:10.2514/1.B36425.

[20] M. Lazzarin, F. Barato, A. Bettella, D. Pavarin, Computational fluid dynamics

simulation of regression rate in hybrid rockets, J. Propuls. Power. 29 (2013)

1445–1452. doi:10.2514/1.B34910.

42
[21] X. Sun, H. Tian, G. Cai, Diameter and position effect determination of

diaphragm on hybrid rocket motor, Acta Astronaut. 126 (2016) 325–333.

doi:10.1016/j.actaastro.2016.04.029.

[22] M. SALITA, Comparison of four boundary layer solutions for fuel regression

ratein classical hybrid rocket motors, (1991). doi:10.2514/6.1991-2520.

[23] G. Cai, C. Li, S. Zhao, H. Tian, Transient analysis on ignition process of catalytic

hybrid rocket motor, Aerosp. Sci. Technol. 67 (2017) 366–377.

doi:10.1016/J.AST.2017.03.041.

[24] G. LENGELLE, B. FOUREST, J. GODON, C. GUIN, Condensed-phase

behavior and ablation rate of fuels for hybrid propulsion, (1993).

doi:10.2514/6.1993-2413.

[25] C. Li, G. Cai, H. Tian, Numerical analysis of combustion characteristics of

hybrid rocket motor with multi-section swirl injection, Acta Astronaut. 123

(2016) 26–36. doi:10.1016/j.actaastro.2016.02.023.

43
Conflict of interest statement

No conflicts of interest to declare.

You might also like