30.10.2023 Views

Study on atomization and combustion characteristics of -- Fang, Xin-xin; Shen, Chi-bing -- Acta Astronautica, 136, pages 369-379, 2017 jul -- Elsevier -- 10.1016_j.actaastro.2017.03

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Acta Astronautica 136 (2017) 369–379

Contents lists available at ScienceDirect

Acta Astronautica

journal homepage: www.elsevier.com/locate/actaastro

Study on atomization and combustion characteristics of LOX/methane

pintle injectors

Xin-xin Fang a,b , Chi-bing Shen a,b, ⁎

MARK

a College of Aerospace Science and Engineering, National University of Defense Technology, Changsha, Hunan 410073, People's Republic of China

b Science and Technology on Scramjet Laboratory, National University of Defense Technology, Changsha, Hunan 410073, People's Republic of China

ARTICLE INFO

Keywords:

Liquid oxygen/methane rocket engine

Pintle injector

Atomization cone angle

Combustion characteristics

ABSTRACT

Influences of main structural parameters of the LOX/methane pintle injectors on atomization cone angles and

combustion performances were studied by experiments and numerical simulation respectively. In addition,

improvement was brought up to the structure of the pintle injectors and combustion flow fields of two different

pintle engines were obtained. The results indicate that, with increase of the gas-liquid mass flow ratio, the

atomization cone angle decreases. In the condition of the same gas-liquid mass flow ratio, as the thickness of the

LOX-injection gap grows bigger, the atomization cone angle becomes smaller. In the opposite, when the half

cone angle of the LOX-injection gap grows bigger, the atomization cone angle becomes bigger. Moreover, owing

to the viscous effects of the pintle tip, with increase of the ‘skip distance’, the atomization cone angle gets larger.

Two big recirculation zones in the combustor lead to combustion stability of the pintle engines. When the value

of the non-dimensional ‘skip distance’ is near 1, the combustion efficiency of the pintle engines is the highest.

Additionally, pintle engines with LOX injected in quadrangular slots can acquire better mixing efficiency of the

propellants and higher combustion efficiency as the gas methane can pass through the adjacent slots. However,

the annular-channel type of pintle injectors has an ‘enclosed’ area near the pintle tip which has a great negative

influence on the combustion efficiency.

1. Introduction

Toxic propellants are used frequently in modern space activities.

Through decades of continuing development, the toxic propellant

rocket engines can achieve high performance. However, the toxicology

and corrosiveness of the propellants lead to many problems, like high

costs and environmental pollution [1]. Non-toxic propellants have the

advantages of environment-friendly, high safety, good maintainability

and low cost, etc. With enhancement of people's awareness of

environmental protection and rapid development of astronautical

technology, using non-toxic propellants in space vehicles begins to be

an inexorable trend. Development of green propellants depends on two

characteristics: high performance and low cost [2]. Many researches

had been conducted to figure out the possibility of methane being used

in rocket engines. The results indicate that methane has the advantages

of safety, low cost, low corrosiveness [3], high cooling performance [4]

low carbon deposition [5]. Additionally, Masters A I et.al pointed out

that methane can overcome the weakness of low specific impulse of

kerosene and low density of hydrogen, but has both the advantages of

them [6]. Thus, rocket engines using liquid oxygen (LOX) and methane

in combination as propellants developed rapidly under that background

[7–10].

On the other hand, Goddard R H had raised the necessity of

controlling the thrust of rocket engines in early 20th century. There are

two basic ways to control the thrust of rocket engines, one of which is to

control the thrust itself, and the other is to control the duration of the

thrust. To control the thrust itself, the most important way is to control

the mass flow rate of the propellants. Although the mass flow rate

varies from different injection pressure-drop of the propellants, it is

difficult to achieve high thrust variation or high performance by this

manner, as the combustion efficiency can be significantly affected by

the injection pressure-drop. Therefore, one practicable way is to alter

the area of the injection channel of the injectors in order to maintain

constant injection pressure-drop. Thus, pintle injectors, which use a

removable component named ‘needle’ to alter the area of injection

channel in order to change the mass flow rate of both oxidizer and fuel,

were applied widely, especially in Apollo project [11,12]. Pintle

injectors, originating from experiments at the Caltech Jet Propulsion

Laboratory (JPL) in 1957, were once used to characterize the reaction

rates of candidate rocket propellants [13]. Then, after improvement of

⁎ Corresponding author at: College of Aerospace Science and Engineering, National University of Defense Technology, Changsha, Hunan 410073, People's Republic of China.

E-mail address: cbshen@nudt.edu.cn (C.-b. Shen).

http://dx.doi.org/10.1016/j.actaastro.2017.03.025

Received 17 September 2016; Received in revised form 10 March 2017; Accepted 27 March 2017

Available online 31 March 2017

0094-5765/ © 2017 IAA. Published by Elsevier Ltd. All rights reserved.


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Nomenclature

π Pi

LOX Liquid Oxygen

m ĊH4 Mass flow rate of gaseous methane, kg/s

ṁLOX Mass flow rate of liquid oxygen, kg/s

L s Length of needle in combustor or ‘skip distance’, mm

D fo Outside diameter of methane channel, mm

D p Outside diameter of pintle tip, mm

Ls/

Dp Non-dimensional ‘skip distance’

D pi Inside diameter of needle, mm

h o Thickness of LOX-injection gap, mm

α o Half cone angle of LOX-injection gap, °

L c Cylindrical section length of combustor, mm

Diameter of combustor, mm

D c

D t Diameter of nozzle throat, mm

L* Characteristic length of combustor, m

N Number of quadrangular slots

δ o Circumferential size of quadrangular slots, mm

L o Axial size of quadrangular slots, mm

BF Blockage Factor, BF =( Nδo)/( πDp)

η Combustion efficiency

C* Characteristic velocity, m/s

C th Theoretical characteristic velocity, m/s

P cs , Stagnation pressure in the combustor, Pa

A t Throat area, mm 2

SMD Sauter Mean Diameter, μm

TMR Total Momentum Ratio

R-R distribution Rosin-Rammler distribution

the structure, pintle injectors have the characteristics of deep throttling,

fast response and ‘face shut off’ [14]. Over 60 different pintle

engines have been developed at least to the point of hot fire

characterization testing in TRW (Thompson-Ramo-Wooldridge Inc.)

and over 130 bipropellant engines using pintle injectors have flown

successfully [15]. In addition, pintle injectors were used in gel

propellants tactical missiles [16] and cryogenic propellants rocket

engines [17–23].

However, fundamental research about pintle injectors is little. The

conical liquid sheet of the pintle injectors is more stable if the inner

face of the injection channel is longer than the outer face [24].

Numerical simulation results about inner flow process of the pintle

injectors, which have been demonstrated by experiments [25], indicate

that manufacturing process tolerance has a great influence on flow

state of the conical liquid sheet in the exit of the injectors [26].

Moreover, effects of environmental pressure, gas-liquid momentum

ratio and Weber number on atomization characteristics of the pintle

injectors were studied experimentally [27–29]. The combustion flow

fields of the pintle injectors which are important to understand the

combustion characteristics of the pintle engines were rarely studied.

In the present study, the influences of main structural parameters

of the pintle injectors on atomization cone angles are studied experimentally,

and combustion flow fields inside the pintle engines are

obtained through numerical simulation. Finally, a comparison of

combustion flow fields of two different pintle engines is given to come

up with the improved structure of the pintle injectors. The results

provide a reference for the structural optimization of the pintle engines.

distribution of the pintle injectors, the Malvern measurement system is

used in the experiments. The SMD distribution results provide

references to the inflow parameters of the numerical simulation.

2.2. Pintle injector

Fig. 2(a) presents the pintle injector used in the experiments. The

pintle injector is mainly composed of installation fixture, pintle, needle

and base. The needle can be positioned along the axial direction. It is

achieved by replacing the gaskets which have different thickness. There

are two pressure sensors in the pintle injectors to measure the injection

pressure of the water and air respectively. The pintle and needle form

the injection channel of the water, while the base and needle form the

injection channel of the air. The physical configuration of the pintle

injector is shown in Fig. 2(b).

In the experiments, the total momentum ratio (TMR) is the same

between the simulants and the real propellants. The TMR is defined by

Eq. (1) in which ρ represents density, v represents velocity and A

represents the area of injection channel. The subscript index “g” and “l”

2. Experimental facilities

2.1. Experimental platform

The atomization experimental platform used in this research is

shown in Fig. 1. The nitrogen resource works as pressurization bottle.

Due to bad atomization characteristics of liquid/liquid pintle injectors,

the gaseous methane rather than liquid methane was chosen as fuel,

mainly because the gaseous methane can improve the atomization

characteristics of the pintle injectors. A combination of water and air is

selected as simulant of the LOX and gaseous methane respectively. The

water and air are marked as simulant No. 1 and No. 2.

The pressure of the simulants is controlled by a gas pressure

regulator which is installed in the gas distribution board. The injected

water is collected by collecting system and pumped out by the air draft

system. The measurement system is independent from the experimental

apparatus. A Tamron high speed camera with a lens of Nikon

80–200 mm is used to obtain atomization images in the experiments.

The frame frequency in this research is 10,000 f/sand the exposure

time is 1/40,000 s. To measure the Sauter Mean Diameter (SMD)

Fig. 1. Schematic configuration for experimental platform.

370


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

2.3. Image processing method

To obtain the atomization cone angles, the atomization images

should be processed following the process shown in Fig. 3. Fig. 3(a) is a

background image, in which x and y represent the axial direction and

radial direction respectively. An original atomization image is shown in

Fig. 3(b), in which the partial image in red box whose height is 30 mm

is extracted. After that, the partial image is converted to grayscale

image firstly, and then to binary image. The boundary between black

and white is atomization gas-liquid boundary as shown in Fig. 3(c). To

obtain the atomization cone angles, the atomization gas-liquid boundary

was fitted by ordinary least squares techniques as shown in

Fig. 3(d), in which the red curve represents the atomization gas-liquid

boundary, the yellow lines represent the fitting curve and the blue lines

represent the fitting error. The angle between the two yellow lines is

defined as the atomization cone angle. The scale between the atomization

images and the physical dimensions in Fig. 3 is 76.5 mm per 648

pixels.

To reduce errors during image processing, atomization cone angles

of 1000 images were obtained firstly, and then an average was

computed. Fig. 4 shows average value of the atomization cone angles

for different number of images processed. With increase of images

processed, the average value fluctuates. But when the number of the

images is larger than 300, the atomization cone angles tend to be

stabilizing. Thus, it is reasonable to suppose the stable value as the

atomization cone angle in certain operating condition. In this paper,

1000 images were selected to obtain the atomization cone angles.

3. Numerical simulation set-up

3.1. Numerical simulation conditions

represent gaseous and liquid simulants respectively. In our experiments,

the mass flow rate of the simulants is smaller than that of the

real propellants. The values of TMR for different h o are given in Table 1.

TMR =( ρ v A )/( ρ v A )

Fig. 2. Pintle injector.

g g 2 g l l 2 l

(1)

During to the limits of the experimental platform, the effects of

ambient pressure on the atomization characteristics of the pintle

injectors are not taken into account yet. In addition, the influences of

the dissimilar of fluid properties between the simulants and the real

propellants on the experimental results were supposed keep the same

for each condition. So the conclusion of the experiments is reasonable.

Numerical simulation on LOX/methane pintle engines was conducted.

The influences of the main structural parameters of the pintle

injector and combustor on combustion performance were studied. The

numerical simulation is performed on the ANSYS Fluent platform. A

second-order, double-precision solver was utilized to conduct the

simulations.

The schematic configuration of the pintle engines is shown in Fig. 5.

Because the structure is axisymmetric, the numerical simulation was

conducted in two-dimensional mainly. The fuel and oxidizer are

gaseous methane and LOX respectively. The numerical simulation

conditions are listed in Table 2. The throat diameter and the contraction

ratio of the pintle engines are 47.17 mm and 4.49 respectively. The

boundary condition and the mesh of a close-up view of the injection

area in the two-dimensional numerical simulation are shown in Fig. 6.

The heat transfer was not taken into account in the numerical

simulation. And a “No Slip” boundary condition is used on the wall

in the numerical simulations. Demonstration of the grids could be

found in Section 3.2.

The Rosin-Rammler (R-R) distribution was used in the Discret

Phase Model (DPM). The R-R distribution is a cumulative mass

fraction distribution:

f ( d) = exp[−( d/ d) n

]

(2)

In Eq. (2) f ( d) represents the cumulative mass fraction of particles

Table 1

Gas-liquid momentum ratio for different h o.

h o (mm)

0.06 3.02

0.08 4.03

0.10 5.03

0.12 6.04

Momentum ratio

371


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 3. Acquisition process of atomization cone angles.

Table 2

Numerical simulation conditions of pintle engines.

No. ṁLOX m ĊH 4 D p Ls/

Dp h o D fo Dc / DP L c L*

(kg/s) (kg/s) (mm) (mm) (mm) (mm) (m)

1 2.056 0.793 30 1.00 0.10 31.2 3.3 220 1.0

2 2.056 0.793 30 0.75 0.10 31.2 3.3 220 1.0

3 2.056 0.793 30 0.25 0.10 31.2 3.3 220 1.0

4 2.056 0.793 30 1.50 0.10 31.2 3.3 220 1.0

5 2.056 0.793 30 1.00 0.08 31.2 3.3 220 1.0

6 2.056 0.793 30 1.00 0.06 31.2 3.3 220 1.0

7 2.056 0.793 30 1.00 0.12 31.2 3.3 220 1.0

8 2.056 0.793 30 1.00 0.10 31.2 3.3 180 0.8

9 2.056 0.793 30 1.00 0.10 31.2 3.3 260 1.2

10 2.056 0.793 30 1.00 0.10 31.2 3.3 300 1.4

Fig. 4. Atomization cone angles for different number of pictures processed.

Fig. 5. Schematic configuration for pintle engines.

whose diameters are bigger than d. In addition, d and n represent the

mean diameter and spread parameter respectively. The mean diameter

is discussed in Section 4.

The eddy-dissipation model was used as the chemical reaction

model, as the combustion process in rockets is a non-premixed

diffusion process. The model and its applicability in the numerical

simulation of rockets had been demonstrated by other researchers

[30,31].

In the present simulations, a single-step global reaction (Eq. ()) is

used as other researchers [32,33]. It is reasonable as the global reaction

allows the coupling of pressure and unsteady heat release to be

captured and reduces the amount of computation [34].

CH4 +2O2 → CO2 +2H2 O

(3)

Researches have demonstrated that the non-equilibrium evaporation

model is more accurate than the quasi-equilibrium one, especially

for the droplet with small radius [35–37]. So the non-equilibrium

evaporation model is adopted in the present simulation. The details

about the model can be found in reference [38].

The standard k–ε turbulence model is applied for modeling the

turbulence. In order to obtain a high-quality solution of the flow in the

turbulent boundary layer, the standard wall functions are used for the

near-wall treatment [32].

3.2. Mesh study

The grid size has a great influence on the simulation results,

especially the mesh near the wall. The first grid height is 0.05 mm in

the boundary and the standard grid has a total of 147,196 cells. In

addition, the y plus near the wall is larger than 10 by checking the

numerical results to make sure the validation of the standard wall

functions. There are both 25 cells for the inlet of gaseous methane and

372


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 6. Boundary condition and mesh of numerical simulation.

Table 3

Pressure and temperature of monitor points of different grids in pre-simulation.

Monitor points a b c d

Grids

Parameters

72,538 Pressure (MPa) 2.930 2.923 2.908 2.898

Relative deviation (%) 1.52 1.39 0.972 1.68

Temperature (×10 3 K) 2.938 2.886 2.863 2.815

Relative deviation (%) 16.17 15.07 12.50 10.31

147,196 Pressure (MPa) 2.895 2.893 2.876 2.863

Relative deviation (%) – – – –

Temperature (×10 3 K) 2.489 2.532 2.561 2.535

Relative deviation (%) – – – –

218,550 Pressure (MPa) 2.886 2.883 2.880 2.850

Relative deviation (%) −0.31 −0.35 0.14 −0.46

Temperature (×10 3 K) 2.529 2.508 2.545 2.552

Relative deviation (%) 1.58 −0.96 −0.63 0.67

Fig. 8. Atomization cone angles for different α o.

Fig. 7. Atomization cone angles for different gas-liquid mass flow ratio.

Fig. 9. Atomization cone angles for different Ls/

Dp.

liquid oxygen. In order to demonstrate the grid convergence, presimulation

was conducted. Two grids which have smaller and larger

quantity of cells were used to compare with the standard grid. The two

grids have a total of 72,538 and 218,550 cells respectively. Four points,

named a, b, c and d (see in Fig. 6) were set to compare their pressure

and temperature measured in the three grids. The positions of the four

points (a, b, c and d) are (0.03 m, 0), (0.03 m, 0.03 m), (0.12 m, 0) and

(0.12 m, 0.03 m) respectively. The results are shown in Table 3. We can

see that, compared with the results of the standard grid, the relative

deviation of the larger quantity grid is negligible, while that of the

smaller quantity grid is large. So the standard grid is suitable for the

present simulation.

4. Experimental results

In the experiments, the mass flow of the gaseous simulant were kept

unchanged, while the mass flow of the liquid simulant changed

according to different gas-liquid mass flow ratio. Fig. 7 shows the

variation curve of the atomization cone angles along with the gas-liquid

mass flow ratio in the condition of different h o (see in Fig. 5). With

increase of the gas-liquid mass flow ratio, the atomization cone angles

decrease, and the trend becomes flat. In addition, as h o grows bigger,

the atomization cone angles become smaller under the condition of

same gas-liquid mass flow ratio. It is because when h o becomes bigger,

373


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 10. SMD and spread parameter along axial direction.

the injected liquid sheet slows down and the gas-liquid momentum

ratio (see in Table 1) becomes higher.

Fig. 8 shows the atomization cone angles for different α o (see in

Fig. 5). In the condition of same gas-liquid mass flow ratio, the bigger

α o is, the larger the atomization cone angle becomes. It is because with

the increase of α o , the radial momentum of injected liquid sheet

becomes greater.

Fig. 9 shows the variation curves of the atomization cone angles for

different Ls/

Dp (see in Fig. 5). In the condition of same gas-liquid mass

flow ratio, the atomization cone angles become larger along with

increase of Ls/

Dp, although the difference between the case of 0.25

and 0.50 is not significant. It is because the larger Ls/

Dp is, the longer

the gas propellant travels before colliding with the liquid sheet. And in

the effect of the viscous force of the needle's outer wall, velocity of the

gas propellant decreases. Thus, the gas propellant becomes slower and

the gas-liquid momentum ratio becomes lower in the case of high

Ls/

Dp.

The SMD and spread parameters of R-R distribution of the particle

diameters in six different positions were measured. The positions

locate at x=2 cm, 4 cm, 6 cm, 7 cm, 8 cm, 10 cm, and y=0 for all the

points (the coordinate axes can be seen in Fig. 3). The results can be

seen in Fig. 10. It can be seen that the SMD along the central

atomization zone of the pintle injectors is about 70( ± 10) μm. The

spread parameters are in the range of 1.5 and 1.8. The relation between

the average diameter and SMD is given by Lefebvre [39]. The SMD and

spread parameter are chosen as 70 µm and 1.65 respectively. As a

result, the average diameter is 50.7 µm. So in the numerical simulation

of the pintle engines the mean diameter of the R-R distribution was

chosen 50.7 µm.

Fig. 12. Combustor pressure for different L s .

5. Numerical simulation results

5.1. Influence of Ls/

Dp on combustion performance

To study influences of Ls/

Dp on combustion performance of the

pintle engines, different Ls/

Dp like 0.25, 0.50, 1.00 and 1.50 are

selected. As D p =30 mm (see in Table 2), L s are 7.5 mm, 15 mm,

30 mm and 45 mm respectively. The minimize mesh size near the wall

and the inlet of the propellants keep unchanged and the construction

method of the grid refinement is the same for different pintle lengths.

Fig. 11 shows streamline for different L s . It can be seen that there are

two big recirculation zones in the combustor.

The combustion efficiency is defined by Eq. (4) below [40]. The η

represents the combustion efficiency, while C* and C th represents the

characteristic velocity of numerical simulation and theoretical value.

The P cs , represents the stagnation pressure in the combustor while the

A t represents the throat area.

C

η = *

Cth

Pcs

, At

C* =

m ̇ CH4 + m ̇ LOX

(4)

The combustion efficiency when Ls/

Dp is 0.25, 0.5, 1.0 and 1.5 is

0.884, 0.907, 0.958 and 0.927 respectively (see in Fig. 12). Heister S D

points out that the pintle engines have demonstrated high combustion

efficiencies in the 96–99% range for most of the engines which have

been developed for flight programs [15,41]. The low combustion

efficiency of the pintle engines in the present study is due to the bad

mixing efficiency of the propellants. So the structure of the pintle

Fig. 11. Streamline for different L s .

374


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 13. Temperature fields for different L s .

Fig. 14. Contours of mass fraction of O 2 and particle traces of LOX drops when Ls is

7.5 mm.

engines needs to be improved. Actually, the gaseous methane has two

aspects of influences on the injected LOX drops. Firstly, the acceleration

effect of the gaseous methane on the injected LOX drops is higher

when Ls/

Dp is smaller, so the LOX drops is faster in this case. As a

result, the residence time of the LOX drops is shorter and this is the

main reason why the combustion efficiency is lower when Ls/

Dp is

smaller. Secondly, the gaseous methane makes the injected LOX drops

in the condition of smaller Ls/

Dp break into smaller drops more quickly,

and this has a positive influence on the combustion efficiency. Thus, the

variation of the combustion efficiency along with Ls/

Dp is a tradeoff

between that two factors. In addition, when Ls/

Dp is too big, the

effective characteristic length of the combustor in which the flame

exists is smaller as there is no flame near the pintle tip (see in Fig. 13).

Thus, the combustion efficiency when Ls/

Dp is bigger than 1.0

decreases.

The combustor pressure (area average total pressure) has a similar

varying trend as the combustion efficiency for different Ls/

Dp (see in

Fig. 12). The combustor pressure increases firstly and then decreases

along with enlarging of L s . In considering of the combustion efficiency,

L s should be chosen 30 mm. In other words, when Ls/

Dp is around 1.0,

the pintle engines could acquire the best combustion efficiency. This

similar conclusion was got by Heister S D who analyzed the liquid/

liquid propellants pintle engines [41]. He pointed out that the typical

value of the “skip distance” (Ls/

Dp) is around 1.0.

The temperature fields for different L s are shown in Fig. 13. There

are two low temperature zones (big temperature gradient) in the

combustor marked as A and B in Fig. 13. Contours of the mass fraction

of O 2 and particle traces of LOX drops when Ls is 7.5 mm is shown in

Fig. 14. We can see that there exists much O 2 near the pintle tip and in

the center of the combustor which are the positions of zone A and B in

Fig. 13. From the particle traces in Fig. 14 we can see that there are

some LOX drops rebound from the wall and move toward the center of

the combustor. And the evaporation of LOX drops and mass fraction of

O 2 is much high in the region A. In addition, there are some particles

reach the center of the combustor in the position of region B. Thus it is

the evaporation of LOX drops which causes the low temperature in

region A and B in the combustor. In addition, the gas temperature near

the first half of the combustor is low, which is because there is no flame

exists in this region, and it provides protection to the wall of the

combustor.

5.2. Influence of h o on combustion performance

Different values of h o were selected like 0.06 mm, 0.08 mm,

0.10 mm and 0.12 mm to study influences of h o on combustion

performance of the pintle engines. Different h o leads to different gasliquid

momentum ratio (see in Table 1).

Fig. 15 shows the streamline for different h o . The size of the

recirculation zone D varies little, while the size of the recirculation

zone C decreases along with increase of h o . The bigger h o is, the smaller

the velocity of the injected LOX becomes and the bigger gas-liquid

momentum ratio becomes. As a result, the size of the recirculation zone

C decreases. Both the recirculation zone C and D have positive

influences on the combustion stability of the pintle engines. For

recirculation zone D, its effects are little, as the recirculating flow is

formed by unreacted propellants [42]. But the low-temperature

recirculating flow has a cooling effect on the front part of the combustor

375


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 15. Streamline for different h o.

brought into the center of the combustor. Finally, the high-temperature

combustion products lead to more rapid spray mixing.

Fig. 16 shows the combustor pressure for different h o . The

combustor pressure decreases and the trend becomes flat along with

increase of h o , and decrease of the size of the recirculation zone C is

responsible for that. Thus, when designing pintle injectors, the velocity

of LOX could not be chosen too small or the gas-liquid momentum

ratio could not be chosen too big.

Fig. 17 shows the temperature field for different h o . The gas

temperature near the pintle tip decreases along with increase of h o .

When h o is 0.06 mm there is a strong temperature gradient along the

axis of the combustor. The reasons are the same as that in Section 5.1.

The recirculation zones C in Fig. 15 has great influences on the

combustion stability of the pintle engines as discussed before. The size

of the recirculation zone C is bigger and more high-temperature gas is

entrained into it when h o is small.

Fig. 16. Combustor pressure for different h o.

(see in Fig. 13). As for recirculation zone C, it has great positive

influences on the combustion stability of the pintle engines. There are

three reasons for this [42]. Firstly, it provides a continuous heat source

for ignition purpose. Secondly, it enables the combustion zone to be

5.3. Influence of L* on combustion performance

Four different parameters were selected, which are 0.8 m, 1.0 m,

1.2 m and 1.4 m to study influences of L*. The diameter of the

combustor keeps constant, while the lengths are 180 mm, 220 mm,

260 mm and 300 mm respectively (see in Table 2). For different L*, the

size of the grids keeps unchanged. Thus, bigger L* has greater amount

Fig. 17. Temperature field for different h o.

376


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

Fig. 18. Streamline for different L*.

Fig. 21. Schematic configuration for two different pintle engines in three-dimensional

numerical simulation.

Fig. 19. Combustor pressure for different L*.

Fig. 22. Temperature (K) field for two different pintle engines.

Fig. 20. Schematic configuration for improved pintle engines.

of cells. Fig. 18 shows the streamline for different L*. The size of the

two recirculation zones changes little.

Fig. 19 shows the combustor pressure for different L*. It can be seen

that the larger L* is, the higher the combustion pressure of the pintle

engines is. But increase of the combustor pressure is little while L* is

larger than 1.2 m. It has been analyzed in Section 5.1 that the bigger

Ls/

Dp is, the smaller the effective characteristic length of the combustor

is. When the characteristic length is small, the time for mixing and

reaction is less. The biggest particle residence time is 2.33 ms, 3.05 ms,

3.27 ms and 3.36 ms when the characteristic lengths are 0.8 m, 1.0 m,

1.2 m and 1.4 m respectively. Bigger particle residence time means the

Fig. 23. Mole fraction distribution of gaseous methane.

oxygen can react with the methane more sufficiently. Thus bigger

characteristic length means higher combustion efficiency and combustor

pressure (see in Fig. 19). On the other hand, larger values of L*

mean heavier pintle engines. So the value of L* is a tradeoff between the

combustion efficiency, weight and cooling of the wall of the pinle

engines.

5.4. Improvement of pintle structure

From the two-dimensional numerical simulation results, the combustion

efficiency of the pintle engines is low (around 0.96). It is

377


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

because the LOX is injected as conical sheet and as a consequence it

forms an ‘enclosed’ area near the pintle tip which leads to bad mixing of

the propellants. Thus, improved structure of the pintle injectors is

brought up [41] (see in Fig. 20). Fig. 20(b) is an enlarged schematic

configuration of the new pintle tip. The LOX injection channel is

changed from the circularity (see in Fig. 5) to the quadrangular slots

(see in Fig. 20). The most important advantage of the slots is that the

gas methane can pass through between the adjacent slots, so mixing of

the propellants is much more efficient.

The combustion field of the pintle engines with different LOXinjection

channels was studied by three-dimensional numerical simulation.

The methane is injected through the outer annular channel,

while the LOX is injected through the circularity in Fig. 21(a) (scheme

1) and quadrangular slots in Fig. 21(b) (scheme 2). There are 16

quadrangular slots on the pintle tip in scheme 2. The circumferential

and axial sizes of the quadrangular slots are 2 mm and 2.95 mm

respectively. Thus, BF of the scheme 2 with quadrangular slots is 0.5.

Other sizes of the pintle injectors and combustor are same as condition

No. 1 in Table 2.

Fig. 22 shows the temperature field of the two different pintle

engines. The temperature of the gas near the combustor wall is lower in

scheme 1, while the high-temperature area is larger in scheme 2. The

cooling effect of the low-temperature methane on the wall of the

combustor is better for scheme 1. However, the mixing efficiency of the

propellants and the combustion efficiency (0.981) are better for scheme

2. So in consideration of the combustion efficiency of the pintle

engines, scheme 2 is better than scheme 1. In addition, the combustor

pressure is higher for scheme 2.

For the pintle engines with quadrangular slots, the temperature

field is more uniform in the last half than the first half of the

combustor. The mole fraction distribution of the gaseous methane is

shown in Fig. 23. It can be seen that for scheme 1 the methane

distributes mainly near the combustor wall, while the methane is

mainly in the center of the combustor for scheme 2. So the mixing

efficiency of scheme 2 is much higher than scheme 1. All in all,

considering the mixing efficiency and combustion efficiency of the

pintle engines, LOX injected through the quadrangular slots is better

than the circularity.

6. Conclusions

In this paper, atomization characteristics of the pintle injectors

were studied experimentally. With increase of the gas-liquid mass flow

ratio, the atomization cone angle decreases. In the condition of the

same gas-liquid mass flow ratio, as h o grows bigger, the atomization

cone angle becomes smaller. In the opposite, when the half cone angle

of the LOX-injection gap grows bigger, the atomization cone angle

becomes bigger. Owing to the viscous effects of the pintle tip, with

increase of Ls/

Dp, the atomization cone angle becomes larger. The SMD

along the central atomization zone of the pintle injectors is about 70( ±

10) μm. And the spread parameters are in the range of 1.5 and 1.8.

Then, the influences of the main structural parameters of the pintle

injectors and combustor on the combustion performances of the LOX/

methane pintle engines were studied by numerical simulation. Two big

recirculation zones exist in the combustor, which lead to combustion

stability of the pintle engines. When L s is too small, there exist two low

temperature zones in the combustor, and that will lead to decrease of

the combustion efficiency. When L s is too big, the effective characteristic

length of the combustor reduces. In consideration of the combustion

efficiency, Ls/

Dp should be chosen around 1.0. Additionally, the

combustion efficiency decreases along with increase of h o , and the

decrease of the size of the recirculation zones is responsible for that.

Finally, improvement was brought up to the structure of the pintle

injectors and a comparison of the combustion flow fields of the two

different pintle engines is given by three-dimensional numerical

simulation. Pintle engines with LOX injected in the quadrangular slots

can acquire better mixing efficiency of propellants and higher combustion

efficiency as the gas methane can pass through the adjacent slots

while the annular-channel type has an ‘enclosed’ area near the pintle

tip which has a negative influence on the combustion efficiency.

Acknowledgements

The authors would like to express their gratitude for the financial

support provided by the Fund of Innovation, Graduate School of NUDT

(No. S150105). The authors are also grateful to the reviewers for their

extremely constructive comments.

References

[1] D. Haeseler, A. Götz, A. Fröhlich, et al., Non-Toxic Propellents for Future Advanced

Launcher Propulsion Systems. AIAA2000-3687.

[2] John D. DeSain, Green Propulsion: Trends and Perspectives. Crosslink, 〈http://

www.Aero.org/publications/crosslink/summer2011/04.html〉.

[3] S.D. Rosenberg, M.L. Gage Compatibility of Hydrocarbon Fuels with Booster

Engine Combustion Chamber Liners. AIAA88-3215.

[4] P. Pempi, T. Frohlich, H. Vernin, LOX/Methane and LOX/Kerosene High Thrust

ngine trade-off. AIAA 2001–3542.

[5] A.J. Giovanetti, L.J. Spadaccini, E.J. Szetela, Deposit Formation and Heat Transfer

in Hydrocarbon Rocket Fuels. NASA CR-168277.

[6] A.I. Masters, J.E. Colbert, A.W. Brooke, FLOX/Methane Pump-Fed Engine

Systems. AIAA 69-510.

[7] K. Breisacher, K. Ajmani, LOX/Methane Main Engine Igniter Tests and Modeling.

AIAA 2008-4757.

[8] C. Daniele, P. Mario, R. Daniele, et al., Numerical Simulation of a LO 2 -CH 4 Rocket

Engine Demonstrator, 65th International Astronautical Congress, Toronto, 2014.

[9] J.B. Michael, J.M. Eric, E.A. William, Student Design/Build/Test of a Throttlable

LOX/LCH 4 Thrust Chamber, 65th International Astronautical Congress, Toronto,

2014.

[10] W.M. Marshall, J.E. Kleinhenz, Summary of Altitude Pulse Testing of a 100-lbf

LO 2 /LCH 4 Reaction Control Engine, JANNAF 8th MSS/6th LPS/5th LPS Joint

Subcommittee Meeting, Huntsville, 2011.

[11] G. Elverum, P. Staudhammer, J. Miller, et al., The Descent Engine for the Lunar

Module. AIAA 67-521.

[12] R. Gilroy, R. Sackheim, The Lunar Module Descent Engine - A Historical

Perspective. AIAA 89-2385.

[13] Personal Conversation Between Dr. Pete Staudhammer and Grodon Dressler at

TRW, Inc., Redondo Beach, CA, 2000.

[14] B. Siegel, Research of Low-Thrust Bipropellant Engines, Independent Research

Program Annual Progress Report, Redondo Beach, USA, 1961, pp. 59–73.

[15] A.D. Gordon, J.M. Bauer, TRW Pintle Engine Heritage and Performance

Characteristics. AIAA 2000-3871.

[16] K.F. Hodge, T.A. Crofoot, S. Nelson, Gelled Propellants for Tactical Missile

Applications. AIAA 99-2976.

[17] G. Dressier, F. Stoddard, K. Gavitt, et al., Test Results from a Simple, Low-cost,

Pressure-fed Liquid Hydrogen/Liquid Oxygen Rocket Combustor, 1993 JANNAF

Propulsion Meeting, Monterey, 1993.

[18] J.P. Henneberry, F.J. Stoddard, A.L. Gu, et al., Low-cost Expendable Launch

Vehicles, in: Proceedings of the 28th Joint Propulsion Conference and Exhibit,

AIAA/SAE/ASME/ASEE, Nashville, 1992.

[19] R.L. Sackheim, F.J. Stoddard, J.A. Hardgrove, Propulsion Advancements to Lower

the Cost of Satellite Based Communications Systems. AIAA 94-0930.

[20] K. Gavittetc, Testing of the 650 K1bf LOX/LH2 Low Cost Pintle Engine. AIAA

2001-3987.

[21] K. Gavittetc, TRW LCPE 650 K1bf LOX/LH2 Test Results. AIAA 2000-3853.

[22] G. Kathy, et al., TRW's Ultra Low Cost LOX/LH2 Booster Liquid Rocket Engine,

33th Joint Propulsion Conference and Exhibit, Seattle, 1997.

[23] G. Dressler, F. Stoddard, K. Gavitt, Test Results from a Simple, Low-Cost, Pressure-

Fed Liquid Hydrogen/Liquid Oxygen Rocket Combustor. 1993 JANNAF Propulsion

Meeting, 1993.

[24] A. Marchi, J. Nouri, Y. Yan, et al., Spary stability of outwards opening Pintle

Injectors for stratified direct injection spark ignition engine operation, Int. J.

Engine Res. 11 (6) (2010) 413–437.

[25] J.M. Nouri, E. Abo-Serie, A. Marchi, et al., Internal and near nozzle flow

characteristics in an enlarged model of an outwards opening Pintle-type gasoline

injector, J. Phys.: Conf. Ser. 85 (1) (2007) 256–267.

[26] M. Gavaises, S. Tonini, A. Marchi, Modelling of internal and Near-nozzle flow of a

Pintle-type outwards-opening gasoline Piezo-injector, Int. J. Engine Res. (7) (2006)

381–397.

[27] J.M. Nouri, M.A. Hamid, Y. Yan, et al., Atomization characterization of a Piezo

Pintle-type injector for gasoline direct injection engines, J. Phys.: Conf. Ser. 85 (1)

(2007) 281–292.

[28] D. Quan, I. Tsuneaki, K. Hisanobu, A study on the atomization characteristics of a

Piezo Pintle-type injector for DI gasoline engines, J. Mech. Sci. Technol. 27 (7)

(2013) 1981–1993.

[29] S. Min, Y. Kijeong, K. Jaye, et al., Effects of momentum ratio and Weber number on

atomization half angles of liquid controlled Pintle injector, J. Therm. Sci. 24 (1)

378


X.-x. Fang, C.-b. Shen Acta Astronautica 136 (2017) 369–379

(2015) 37–43.

[30] X. Sun, H. Tian, Y. Li, et al., Regression rate behaviors of HTPB-based propellant

combinations for hybrid rocket motor, Acta Astronaut. (119) (2016) 137–146.

[31] G.B. Cai, C.G. Li, H. Tian, Numerical and experimental analysis of heat transfer in

injector plate of hydrogen peroxide hybrid rocket motor, Acta Astronaut. (128)

(2016) 286–294.

[32] J. Song, B. Sun, in LOX/methaneRocket Engines, Appl. Therm. Eng. (106) (2016)

762–773.

[33] B. Yang, K. Seshadri, Asymptotic Analysis of the Structure of Nonpremixed

Methane Air Flames Using Reduced Chemistry, Combust. Sci. Technol. 88 (1–2)

(1993) 115–132.

[34] K.J. Shipley, C. Morgan, W.E. Anderson, Computational and Experimental

Investigation of Transverse Combustion Instabilities, in: Proceedings of the 49th

AIAA/ASME/SAE/ASEE Joint Propulsion Conference, AIAA 2013-3992, San Jose,

CA, 2013.

[35] V.V. Tyurenkova, Non-equilibrium Diffusion Combustion of a Fuel Droplet, Acta

Astronaut. 75 (6) (2012) 78–84.

[36] V.R. Dushin, A.V. Kulchitskiy, V.A. Nerchenko, et al., Mathematical Simulation for

Non-equilibrium Droplet Evaporation, Acta Astronaut. 63 (11–12) (2008)

1360–1371.

[37] N.N. Smirnov, A.V. Kulchitski, Unsteady State Evaporation in Weightlessness, Acta

Astronaut. 39 (8) (1996) 561–568.

[38] G. Cai, C. Li, H. Tian, Numerical and Experimental Analysis of Heat Transfer in

Injector Plate of Hydrogen Peroxide Hybrid Rocket Motor, Acta Astronaut. (128)

(2016) 286–294.

[39] A.H. Lefebvre, Atomization and Sprays, Hemisphere Publishing Corporation,

Bristol, 1989.

[40] D.K. Huzel, D.H. Huang, Modern Engineering for Design of Liquid Propellant

Rocket Engines, American Institute of Aeronautics and Astronautics, USA, 1992.

[41] S.D. Heister, Handbook of Atomization and Sprays, Springer Science & Business

Media, Berlin, 2011.

[42] D.T. Harrje, F.H. Reardon (Eds.), Liquid Propellant Rocket Combustion Instability,

NASA SP-194, 1972.

379

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!