You are on page 1of 6

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

Transient Characteristics of C3H8/O2 Turbulent Mixing in a Hypersonic Pulse Detonation Engine


KHALID M. SAQR*, AHMED FAIZ, HASSAN KASSEM, MOHSIN SIES AND MAZLAN A. WAHID High-Speed Reacting Flow Laboratory, Faculty of Mechanical Engineering Universiti Teknologi Malaysia 81310 Skudai, Johor Darul'Takzim MALAYSIA * Corresponding Author: khaledsaqr@gmail.com
Abstract: - We present the results of a time-dependent three-dimensional numerical simulation of the turbulent mixing characteristics in the mixing chamber of a hypersonic pulse detonation engine (PDE). Fuel (C3H8) was injected through one supersonic injector, while the oxidizer (O2) were injected through four supersonic injectors. The spatial distributions of air to fuel to ratios were traced during 240 millisecond in the mixing chamber of the PDE. The 3D velocity variations was observed as well. Key-Words: - High altitude aircrafts; Ionospheric flight; Hypersonic propulsion; Supersonic jet; Turbulent mixing; Pulse Detonation Engine

1 Introduction
The Pulse Detonation Engine (PDE) was firstly associated with hypersonic propulsion technology during the 1980s when the case of the alleged SR-91 Aurora aircraft came to publicity [1-4]. The concept of a PDE depends on creating pulsating detonation waves in the engine tube. With the high mach number of such detonation waves (up to 9 for some fuels), the PDE can provide sufficient thrust to propel a hypersonic aerial vehicle [5]. There are several texts that discuss the state of the art propulsion systems for supersonic aircrafts, such as the texts of Oates [6], Flack [7], McCloy [8]and Curran and Murthy [9]. On the other hand, the PDE fundamentals and technology has not been yet the sole topic of any text. However, a comprehensive review for the topic was presented by Roy et al [10] in 2004. This review summarizes the recent trends and challenges facing the successful development of PDEs for hypersonic aerial propulsion. On the practical side, the only recorded flight, away from the scattered and officially denied SR-91 Aurora sightings [11], was in 2008 [12]. The duration of this flight was only 10 seconds, with the purpose of PDE concept validation. However, a vast number of theoretical and experimental researches have been conducted in during the last decade to explore, design, develop and test pulse detonation engines with different operating conditions. The following section discusses some of these research in order to provide a comparative basis for the present research.

2 Literature Review
A complete thermodynamic analysis for the PDE cycle and processes was firstly presented by Wu et al [13] in 2002. The objective of their study was to identify the most crucial phases of the PDE cycle for performance optimization objectives. Similar thermodynamic cycle analyses were presented in the same time by Kentfield [14-15] and Heiser and Pratt [16]. The first published work to study the flow inside a PDE was in 1999 by Chow et al [17] using the Methods of Characteristics. Their investigation explored some of the flow properties behind the shock front of the pulsated detonations. One and two dimensional CFD-based investigations were conducted by Ebrahimi and Merkle [18] for hydrogen/oxygen PDE. A finite rate, multi-step reaction model was used to study the detonation shock speed, pressure spike behavior and CJ conditions. They have also used their model to compute the thrust and specific impulse for the PDE. In the following years, several other publications discussed the numerical simulation of the reactive flow inside PDEs using 1D, 2D and 3D transient mathematical models [19-30]. In its simplest form, a PDE receives fuel and oxidizer from high-pressure injectors that operate outside the detonation tube. Once the detonative mixture propagates and fills the detonation tube, ignition commences. The detonation wave is then generated, by either direct initiation of DDT (i.e. deflagration to detonation transition). Once the detonation wave exits the PDE tube, another cycle

ISSN: 1790-5117

325

ISBN: 978-960-474-166-3

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

begins by fuel and oxidizer injection. Since detonation is a chaotic phenomenon, it is highly dependent on initial conditions. For this reason, several researches have reported the utilization of a mixing chamber in which fuel and oxidizer injection occur. Huang et al [31]studied experimentally the effect of the geometry and disturbance device on the mixing properties of stoichiometric gasoline-air mixtures in a PDE. Such study was expanded by Jiang and Wu [32] to study the effect of injector location and alignment on the mixing quality using numerical simulation. To the best of the authors' knowledge, there is no available literature on the transient mixing of gaseous phase (i.e. fuel and oxidizer are both gaseous) reactant in a mixing chamber of a PDE. This is the concern of the present article.

4 Mixing Chamber Configuration


The configuration of the mixing chamber and injectors are shown in figure 2. The injectors are assembled in three rows on a radial direction. The injector ports are aligned on the inner surface of the mixing chamber. The length and diameter of the chamber are 142 mm and 50 mm, respectively.

3 HiREF Pulse Detonation Engine


The high-Speed Reacting Flow Laboratory (HiREF) of Universiti Teknologi Malaysia is currently developing a PDE prototype under direct funding from the Malaysian government. The ongoing project includes the design and development of the PDE fuel admission, detonation tube, control, and data acquisition systems. The HiREF PDE is operating on propane and oxygen and has a single detonation tube. The PDE is designed to operate on frequencies up to 100 Hz with a cycle that consists of four stages. The PDE cycle can be depicted in figure 1.
1 2 3 4

(a)

t in

t dt

tpj

1- Injection 2- Ignition 3- Purge

tin injection time tdt detonation time tpj Purge time

(b) Figure 2. (a) 3D Mixing chamber and injectors configuration (b) Dimensions in mm

Figure 1. HiREF PDE operating cycle The total cycle time is the sum of tin, tdt and tpj. The present study focus on the flow structure and mixing properties in the mixing chamber of the PDE during the injection phase; tin. The geometry of the mixing chamber is illustrated in figure 2 with the injectors in place. The mixing chamber has a diameter and length of 50 mm and 142 mm. The oxidizer; O2 is injected through four injectors aligned as indicated in figure 2. Fuel; CH4 is injected through one injector. The operating pressure of the injectors is 3 bar and the operating pressure of the chamber is atmospheric.

4 Mathematical Formulation
The flow inside the mixing chamber, which is resulting from the O2 and C3H8 injection, was represented by the three dimensional Navier Stokes equations, and the turbulence flow field was resolved using the Spalart-Allmaras [33] single equation turbulence model. The conservation equations in tensor notations are given as follows:

4.1 Continuity equation + u j = 0 t x j 4.2 Momentum equation (ui ) + ui u j + p ij ji = 0 t x j

[ ]

(1)

(2)

ISSN: 1790-5117

326

ISBN: 978-960-474-166-3

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

The details of the governing equations can be found in several texts such as [34] and [35]. For the sake of brevity, the equations of the SpalartAllmaras turbulence model are not given in the present article. Readers who wish to have a detailed discussion of the single-equation turbulence model can refer to its original paper [33] or to the text of Plaek [34].

4 Numerical Model
4.1 Flow domain and discretization
The flow domain inside the mixing chamber was represented as a cylinder with five inlets, which represent the nozzles of the injectors. Each nozzle had a diameter of 0.70 mm. The flow domain was spatially descretized to 480737 grid cells. The grid cells were hybrid, incorporating both hexahedra and tetrahedral cells. The regions where the nozzles intersect with the mixing chamber boundary were descretized using tetrahedral cells, while the rest of the domain was descretized using hexahedral cells. The discretized flow domain is shown in figure 3 with a focus on once of the locations where the injector nozzles intersect with the flow boundary. The hybrid meshing can be depicted form such figure.

the physical boundary conditions are given in table 1. Table 1. Summary of the physical boundary conditions Boundary condition Value O2 mass flow rate 2.12 10-4 kg/s C3H8 mass flow rate 2.24 10-4 kg/s Initial chamber pressure 1.013 kpa Initial chamber temperature 290 K Exit pressure of the chamber 1.013 kpa

5 Results and Discussion


5.1 Simulation strategy
Fuel and oxygen were injected into the mixing chamber during 240 ms, and the development of spatial species concentration as well as velocity and pressure fields were recorded every 30 ms. The results of the present study details the spatiotemporal depiction of the C3H8/O2 ratio in the mixing chamber.

5.2 Flow structure


The flow in the mixing chamber is quite complex and there are several phenomena to investigate. This is due to the interacting jets of fuel and oxidizer. To the scope of the present study, the effect of jet flow on the mixing is the most important. To develop an indication on such phenomenon, the 3D velocity profiles are plotted in figure 4 on regions where the C3H8 and O2 meet the stoichiometric ratio. The temporal development of mixing, as affected by the intersection between the five jets is rather evident from such figure. Although the volumetric flammability limit of propane ranges from 2.15% to 9.6% in oxygen, the stoichiometric regions give a significant indication on the mixture flammability in the mixing chamber.

Figure 3. Discretized flow domain showing the hybrid meshing

4.2 Solver details and boundary conditions


The problem was solved using a coupled CFD solver that solves the pressure and velocity terms simultaneously with a first order spatially implicit finite volume scheme. The temporal discretization was undertaken using a first order implicit time formulation.

4.3 Boundary conditions


The boundary conditions of the problem were set such that constant oxygen and propane mass flow rates are injected in the chamber. The summary of

ISSN: 1790-5117

327

ISBN: 978-960-474-166-3

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

1.8 1.5

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 20 40 60 80 100 120 140 30 ms

Axial Distance (mm)


1.8 1.5

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0


1.8

60 ms

20 40 60 80 100 120 140

Axial Distance (mm)


1.5

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 20 40 60 80 100 120 140 90 ms

Figure 4. Velocity contours of the fuel and oxidizer in regions where they meet stoichiometric condition It is obvious that the jet interaction in the mixing chamber provide a flammable mixture in almost all the mixing chamber. In addition, the velocity values inside the mixing chamber are relatively low. After 240 ms, the average mixture velocity ranges between 0.0 m/s to 9.0 m/s. This is important to ensure that the onset of detonation is not affected by large spatial fluctuations of velocity.

Axial Distance (mm)


1.8

1.5

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0


1.8 1.5

120 ms

20 40 60 80 100 120 140

Axial Distance (mm)

5.3 Concentration of C3H8/O2


The ration of the C3H8 mass fraction to the O2 mass fraction is plotted on the chamber centerline at different times during the injection as in figure 4. It is clear that the mixture becomes flammable only near to the chamber outlet section. Before the middle of the chamber (x=0:70 mm) there is approximately no fuel concentration. The fuel concentration reaches to a maximum value very near to the fuel injector location (x=100:120).

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 20 40 60 80 100 120 140 150 ms

Axial Distance (mm)

ISSN: 1790-5117

328

ISBN: 978-960-474-166-3

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

1.8 1.5

References:
[1] [2] Sweetman, B., Aurora - is Mach 5 a reality?, in Interavia Aerospace Review. 1990. p. 1009. Rogers, J., RAF Radar Tracked 'Aurora' Over Scotland at Speeds From Mach 3 to Mach 6, in Inside the Air Force. 1992. p. 10-11. Sweetman, B., Mystery contact may be Aurora, in Jane's Defense Weekly. 1992. p. 333. Group, J.s.I., Jane's international defense review: IDR. 2006. 39(7-12). Hutchins, T.E. and M. Metghalchi. Energy and exergy analyses of the pulse detonation engine. in American Society of Mechanical Engineers, Advanced Energy Systems Division (Publication) AES. 2001. New York, NY. Oates, G.C., Aircraft propulsion systems technology and design. 1989, Michigan: AIAA. Flack, R.D., Fundamentals of jet propulsion with applications. 2005: Cambridge University Press. McCloy, R.W., The fundamentals of supersonic propulsion. 1967: Boeing Co., Supersonic Propulsion Test Group. Curran, E.T. and S.N.B. Murthy, Scramjet Propulsion. 2000: AIAA. Roy, G.D., et al., Pulse detonation propulsion: Challenges, current status, and future perspective. Progress in Energy and Combustion Science, 2004. 30(6): p. 545-672. Gregory, T. and P.M. Pope, America's New Secret Aircraft, in Popular Mechanics. 1991, Hearst Magazines. p. 32-36. Norris, G., Pulse Power: Pulse Detonation Enginepowered Flight Demonstration Marks Milestone in Mojave, in Aviation Week & Space Technology. 2008. p. 60. Wu, Y., F. Ma, and V. Yang, System performance and thermodynamic cycle analysis of air-breathing pulse detonation engines. Zhongguo Hangkong Taikong Xuehui Huikan/Transactions of the Aeronautical and Astronautical Society of the Republic of China, 2002. 34(1): p. 1-11. Kentfield, J.A.C., Thermodynamics of airbreathing pulse-detonation engines. Journal of Propulsion and Power, 2002. 18(6): p. 1170-1175. Kentfield, J.A.C., Fundamentals of idealized airbreathing pulse-detonation engines. Journal of Propulsion and Power, 2002. 18(1): p. 77-83. Heiser, W.H. and D.T. Pratt, Thermodynamic cycle analysis of pulse detonation engines. Journal of Propulsion and Power, 2002. 18(1): p. 68-76. Chow, W.L., Y. Yong, and C. Yiannas, Combustion and gas dynamics as related to pulsedetonation engine. Physics and Modern Topics in Mechanical and Electrical Engineering, 1999: p. 112-123. Ebrahimi, H.B. and C.L. Merkle, Numerical simulation of a pulse detonation engine with hydrogen fuels. Journal of Propulsion and Power, 2002. 18(5): p. 1042-1048.

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 20 40 60 80 100 120 140 180 ms


[3] [4] [5]

Axial Distance (mm)


1.8 1.5

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 1.8 1.5 20 40 60 80 100 120 140
[8] [6]

210 ms

[7]

Axial Distance (mm)

[9] [10]

C3H8/O2

1.2 0.9 0.6 0.3 0.0 0 20 40 60 80 100 120 140


[12] [11]

240 ms

Axial Distance (mm)


Figure 4. Spatiotemporal development of the ration between C3H8/O2 mass fractions
[13]

6 Conclusion
The injection of C3H8 and O2, and the flow field inside the mixing chamber of the HiREF experimental PDE mixing chamber has been simulated using a 3D RANS CFD model. The characteristics of mixture velocity and mixing quality have been briefly investigated in the present study. The results show that the mean flow field has low average velocity during several time intervals of the injection. It was also shown that the mixture is flammable in the mixing chamber. The quantitative results of the mixing quality on the centerline of the chamber show that the highly flammable mixture exits near the chamber exit section, which is mostly the demand to permit such mixture of flowing inside the main detonation tube.
[14]

[15]

[16]

[17]

[18]

ISSN: 1790-5117

329

ISBN: 978-960-474-166-3

Proceedings of the 9th WSEAS International Conference on APPLICATIONS of COMPUTER ENGINEERING

[19] Austin, J.M. and J.E. Shepherd, Detonations in hydrocarbon fuel blends. Combustion and Flame, 2003. 132(1-2): p. 73-90. [20] He, X. and A.R. Karagozian, Numerical Simulation of Pulse Detonation Engine Phenomena. Journal of Scientific Computing, 2003. 19(1-3): p. 201-224. [21] Hutchins, T.E. and M. Metghalchi, Energy and exergy analyses of the pulse detonation engine. Journal of Engineering for Gas Turbines and Power, 2003. 125(4): p. 1075-1080. [22] Kailasanath, K., Recent developments in the research on pulse detonation engines. AIAA Journal, 2003. 41(2): p. 145-159. [23] Kawai, S. and T. Fujiwara, Numerical Analysis of First and Second Cycles of Oxyhydrogen Pulse Detonation Engine. AIAA Journal, 2003. 41(10): p. 2013-2019. [24] Murty, V.D. A numerical simulation of heat pipe application to the cooling of pulse detonation engines. in Proceedings of the ASME/JSME Joint Fluids Engineering Conference. 2003. Honolulu, HI. [25] Wang, C., M. Situ, and Z.Y. Han, Numerical investigations on cold flowfields of shock focussing for ignition of pulse detonation. Tuijin Jishu/Journal of Propulsion Technology, 2003. 24(2): p. 156-159. [26] Lee, S.Y., et al., Deflagration to detonation transition processes by turbulence-generating obstacles in pulse detonation engines. Journal of Propulsion and Power, 2004. 20(6): p. 1026-1036. [27] Li, H.H., J.M. Yang, and L.G. Xu, Numerical simulation on the nozzle flow of pulse detonation engine. Tuijin Jishu/Journal of Propulsion Technology, 2004. 25(6): p. 553-556. [28] Wang, J.P., Y.F. Liu, and T.W. Li, Numerical simulations of pre-detonation ignition of pulse detonation engines. Kongqi Donglixue Xuebao/Acta Aerodynamica Sinica, 2004. 22(4). [29] Li, M., et al., Numerical simulation of and experimental study on effect of ring obstacles on detonation initiation and propagation. Xibei Gongye Daxue Xuebao/Journal of Northwestern Polytechnical University, 2006. 24(3): p. 299-303. [30] Li, T.W., J.P. Wang, and Z.H. Ye, Numerical simulation of one-dimensional detonation with detailed chemical reaction model. Kongqi Donglixue Xuebao/Acta Aerodynamica Sinica, 2007. 25(2): p. 199-204. [31] Huang, X.Q., et al., Experimental study of the mixing process of pulse detonation combustion. Ranshao Kexue Yu Jishu/Journal of Combustion Science and Technology, 2005. 11(3): p. 282-285. [32] Jiang, R.H. and X.S. Wu, Numerical simulation of injection and mixing on two-phase for pulse detonation engine. Xitong Fangzhen Xuebao / Journal of System Simulation, 2009. 21(15): p. 4912-4915. [33] Spalart, P. and S. Allmaras., A one-equation turbulence model for aerodynamic flows. , in

Technical Report AIAA-92-0439, merican Institute of Aeronautics and Astronautics. 1992. [34] Blaek, J., Computational fluid dynamics: principles and applications. 2001: Elsevier. [35] Yeoh, G.H. and K.K. Yuen, Computational Fluid Dynamics in Fire Engineering: Theory, Modelling and Practice. 2009: Butterworth-Heinemann.

ISSN: 1790-5117

330

ISBN: 978-960-474-166-3

You might also like