Hostname: page-component-848d4c4894-jbqgn Total loading time: 0 Render date: 2024-06-14T16:56:54.687Z Has data issue: false hasContentIssue false

Numerical investigation of thermochemical non-equilibrium effects in Mach 10 scramjet nozzle

Published online by Cambridge University Press:  22 May 2024

J.P. Wang
Affiliation:
School of Mechanical and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
C.F. Zhuo*
Affiliation:
School of Mechanical and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
C.L. Dai
Affiliation:
School of Mechanical and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
B. Sun
Affiliation:
School of Mechanical and Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
*
Corresponding author: C.F. Zhuo; Email: njust203zcf@126.com

Abstract

High-temperature non-equilibrium effects are prominent in scramjet nozzle flows at high Mach numbers. Hence, the thermochemical non-equilibrium gas model incorporating the vibrational relaxation process of molecules in the hydrocarbon-air reaction is developed to numerically simulate the flow of a hydrocarbon fuel scramjet nozzle at Mach 10. Besides, the results computed by the models of the thermally perfect gas, chemically non-equilibrium gas, and thermally non-equilibrium chemically frozen gas are applied for comparative studies. Results indicate that chemical non-equilibrium effects are more significant for the flow-field structure and parameters compared to thermal non-equilibrium effects. Meanwhile, vibrational relaxation and chemical reactions interact in the flow-field. The heat released from the chemical reactions in the flow-field of the thermochemical non-equilibrium gas model makes the thermal non-equilibrium effects weaker compared to the thermally non-equilibrium chemically frozen gas model; the chemical reactions in the thermochemical non-equilibrium gas model are more intense than in the chemically non-equilibrium gas model. Due to the slow relaxation of vibrational energy, the thermal non-equilibrium models predicted nozzle thrust lower than the thermal equilibrium models by approximately 1.11% to 1.33%; when considering the chemical reactions, the chemical non-equilibrium models predicted nozzle thrust higher than the chemical frozen models by approximately 7.30% to 7.54%. Hence, the structural design and performance study of the high Mach numbers scramjet nozzle must consider thermochemical non-equilibrium effects.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Royal Aeronautical Society

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Curran, E.T. Scramjet engines: the first forty years. J. Propul. Power, 2001, 17, (6), pp 11381148. doi: 10.2514/2.5875 CrossRefGoogle Scholar
Laurence, S.J., Schramm, J.M., Karl, S. and Hannemann, K. An experimental investigation of steady and unsteady combustion phenomena in the HyShot II combustor. In 17th AIAA International Space Planes and Hypersonic Systems and Technologies Conferences, AIAA Paper, 2011. doi: 10.2514/6.2011-2310 CrossRefGoogle Scholar
Karl, S. and Hannemann, K. CFD Analysis of the HyShot II scramjet experiments in the HEG shock tunnel. In 15th AIAA International Space Planes and Hypersonic Systems and Technologies Conferences, AIAA Paper, 2008. doi: 10.2514/6.2008-2548 CrossRefGoogle Scholar
Kindler, M., Lempke, M., Blacha, T., Gerlinger, P. and Aigner, M. Numerical investigation of the HyShot supersonic combustion configuration. In 44th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, AIAA Paper, 2008. doi: 10.2514/6.2008-5167 CrossRefGoogle Scholar
Kumar, S., Pandey, K.M. and Sharma, K.K. Recent developments in technological innovations in scramjet engines: a review. Mater. Today Proc., 2021, 45, pp 68746881. doi: 10.1016/j.matpr.2020.12.1086 CrossRefGoogle Scholar
Dai, C.L., Sun, B., Zhou, C.S., Zhuo, C.F., Du, L. and Zhou, S.B. Numerical investigation of real-gas effect of inward-turning inlet at Mach 12. Aerosp. Sci. Technol., 2021, 115, p 106786. doi: 10.1016/j.ast.2021.106786 CrossRefGoogle Scholar
Dai, C.L., Sun, B., Zhou, S.B., Zhuo, C.F., Zhou, C.S. and Yue, L.J. Influence of high temperature non-equilibrium effects on Mach 12 scramjet inlet. Acta Astronaut., 2022, 193, pp 237254. doi: 10.1016/j.actaastro.2022.01.013 CrossRefGoogle Scholar
Zuo, F.Y., Mölder, S. and Hu, S.L. Thermochemical non-equilibrium effects on hypersonic wavecatcher intake at Mach 12. Acta Astronaut., 2022, 198, pp 5668. doi: 10.1016/j.actaastro.2022.05.030 CrossRefGoogle Scholar
Fiévet, R. and Raman, V. Effect of vibrational nonequilibrium on isolator shock structure. J. Propul. Power., 2018, 34, (5), pp 13341344. doi: 10.2514/1.B37108 CrossRefGoogle Scholar
Ao, Y., Wu, K., Lu, H.B., Ji, F. and Fan, X.J. Combustion dynamics of high Mach number scramjet under different inflow thermal nonequilibrium conditions. Acta Astronaut., 2023, 208, pp 281295. doi: 10.1016/j.actaastro.2023.04.020 CrossRefGoogle Scholar
Shi, L.S., Shen, H., Zhang, P., Zhang, D.L. and Wen, C. Assessment of vibrational non-equilibrium effect on detonation cell size. Combust. Sci. Technol., 2017, 189, (5), pp 841853. doi: 10.1080/00102202.2016.1260561 CrossRefGoogle Scholar
Voelkel, S., Raman, V. and Varghese, P.L. Effect of thermal nonequilibrium on reactions in hydrogen combustion. Shock Waves, 2016, 26, pp 539549. doi: 10.1007/s00193-016-0645-0 CrossRefGoogle Scholar
Koo, H., Raman, V. and Varghese, P.L. Direct numerical simulation of supersonic combustion with thermal nonequilibrium. Proc. Combust. Inst., 2015, 35, (2), pp 21452153. doi: 10.1016/j.proci.2014.08.005 CrossRefGoogle Scholar
Kaneko, M., Men’Shov, I. and Nakamura, Y. Computation of nozzle starting process with thermal and chemical nonequilibrium in high-enthalpy shock tunnel. In 40th AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper, 2002. doi: 10.2514/6.2002-142 CrossRefGoogle Scholar
Schramm, J.M., Karl, S., Hannemann, K. and Steelant, J. Ground testing of the HyShot II scramjet configuration in HEG. In 15th AIAA International Space Planes and Hypersonic Systems and Technologies Conference, AIAA Paper, 2008. doi: 10.2514/6.2008-2547 CrossRefGoogle Scholar
Kaneko, M., Men’Shov, I. and Nakamura, Y. Numerical simulation of nonequilibrium flow in high-enthalpy shock tunnel with EIH scheme. In 15th AIAA Computational Fluid Dynamics Conference, AIAA Paper, 2001. doi: 10.2514/6.2001-2860 CrossRefGoogle Scholar
Sagnier, P. and Marraffa, L. Parametric study of thermal and chemical nonequilibrium nozzle flow, AIAA J., 1991, 29, (3), pp 334343. doi: 10.2514/3.59921 CrossRefGoogle Scholar
Zeitoun, D., Boccaccio, E., Druguet, M.C. and Imbert, M. Reactive and viscous flow in hypersonic nozzles. AIAA J. 1994, 32, (2), pp 333340. doi: 10.2514/3.11989 CrossRefGoogle Scholar
Teixeira, O. and Páscoa, J.Catalytic wall effects for hypersonic nozzle flow in thermochemical non-equilibrium. Acta Astronaut., 2023, 203, pp 4859. doi: 10.1016/j.actaastro.2022.11.031 CrossRefGoogle Scholar
Zidane, A., Haoui, R., Sellam, M. and Bouyahiaoui, Z. Numerical study of a nonequilibrium H2−O2 rocket nozzle flow. Int. J. Hydrogen Energy, 2019, 44, pp 43614373. doi: 10.1016/j.ijhydene.2018.12.149 CrossRefGoogle Scholar
Yu, K.K., Chen, Y.L., Huang, S. and Xu, J.L. Inverse design method on scramjet nozzles based on maximum thrust theory. Acta Astronaut., 2020, 166, pp 162171. doi: 10.1016/j.actaastro.2019.10.024 CrossRefGoogle Scholar
Park, C. Assessment of a two-temperature kinetic model for dissociating and weakly ionizing nitrogen. J. Thermophys. Heat Transf., 1988, 2, (1), pp 816. doi: https://doi.org/10.2514.3/55 CrossRefGoogle Scholar
Park, C. Assessment of two-temperature kinetic model for ionizing air. J. Thermophys. Heat Transf., 1989, 3, (3), pp 233244. https://arc.aiaa.org/doi/10.2514/3.28771 CrossRefGoogle Scholar
Goldberg, U., Peroomian, O. and Chakravarthy, S. A wall-distance-free k- $\varepsilon$ model with enhanced near-wall treatment. J. Fluids Eng., 1998, 120, pp 457462. doi: 10.1115/1.2820684 CrossRefGoogle Scholar
Harten, A. High-resolution schemes for hyperbolic conservation laws. J. Comput. Phys., 1983, 49, (3), pp 357393. doi: 10.1016/0021-9991(83)90136-5 CrossRefGoogle Scholar
Gupta, R.N., Yos, J.M. and Thompson, R.A. A review of reaction rates and thermodynamic and transport properties for the 11-species air model for chemical and thermal nonequilibrium calculations to 30000 K, Technical Report No. NASA/TP1990-1323.Google Scholar
Edwards, J.R. A diagonal implicit/nonlinear multigrid algorithm for computing hypersonic, chemically-reacting viscous flows. In 32nd Aerospace Sciences Meeting and Exhibit, AIAA Paper, 1994. doi: 10.2514/6.1994-762 CrossRefGoogle Scholar
Nietubicz, C. and Gibeling, H. Navier-Stokes computations for a reacting, M864 base bleed projectile. In 31st Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, 1993. doi: 10.2514/6.1993-504 CrossRefGoogle Scholar
Weisgerber, H., Fischer, M., Magens, E., Winandy, A., Foerster, W. and Beversdorff, M. Experimental analysis of the flow of exhaust gas in a hypersonic nozzle. In 8th AIAA International Space Planes and Hypersonic Systems and Technologies Conference, AIAA Paper, 1998. doi: 10.2514/6.1998-1600 CrossRefGoogle Scholar
Link, T. and Koschel, W.W. Computation of a nonequilibrium expansion flow in a single expansion ramp nozzle. J. Propul. Power., 2001, 17, (6), pp 13531360. doi: 10.2514/2.5887 CrossRefGoogle Scholar