Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-27T06:21:16.969Z Has data issue: false hasContentIssue false

Shock train response to pulse backpressure forcing

Published online by Cambridge University Press:  17 January 2024

X. Ma (马晓敏)*
Affiliation:
School of Power and Energy, Northwestern Polytechnical University, Xi’an, China
Y. Zhang (张永辉)
Affiliation:
School of Power and Energy, Northwestern Polytechnical University, Xi’an, China
J. Yuan (袁菁涛)
Affiliation:
School of Power and Energy, Northwestern Polytechnical University, Xi’an, China
W. Fan (范玮)
Affiliation:
School of Power and Energy, Northwestern Polytechnical University, Xi’an, China
*
Corresponding author: X. Ma; Email: xiaominma@mail.nwpu.edu.cn

Abstract

Transient numerical simulations were conducted to investigate the influence of large amplitude and fast impact backpressure on a shock train. The fundamental problem consists of a shock train within a constant-area channel with a Ma=1.61 inflow and a pulse backpressure applied to the outlet. The pressure disturbance in the isolator has an intense forcing-response lag. From the moment of the backpressure peak appearance, it takes 36 times the backpressure duration for the pressure disturbance to reach the upstream end. It moves upstream with time in the form of a normal shock wave. As time progresses, the normal shock degenerates into a $\lambda $ shock and a compression wave behind due to the action of viscous dissipation in the boundary layer. Eventually, a multi-stage shock train is formed. The maximum backpropagation distance is a quadratic function of both the pulse backpressure peak and duration, and the relationship between these variables was determined by fitting. When the integral value of backpressure to time is fixed, reducing the backpressure peak while increasing the duration will reduce the backpressure pulsation at the isolator outlet, which will be more conducive to shortening the maximum backpropagation distance than reducing the duration and increasing the backpressure peak. The values of backpressure peak and duration are obtained from the detonation combustion case, which ensures the authenticity of backpressure characteristics. The relevant research conclusions can provide a reference for the design of the isolator of pulse detonation ramjet.

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

Zvegintsev, V.I., Ivanov, V.S., Frolov, S.M., Shamshin, I.O. and Zangiev, A.E. Testing of hydrogen-fueled detonation ramjet in aerodynamic wind tunnel at Mach 1.5 and 2.0, AIP Conf. Proc., 2021, 2351, p 030056.CrossRefGoogle Scholar
Ivanov, V.S., Frolov, S.M., Zangiev, A.E., Zvegintsev, V.I. and Shamshin, I.O. Updated conceptual design of hydrogen/ethylene fueled detonation ramjet: Test fires at Mach 1.5, 2.0, and 2.5, Aerospace Sci. Technol., 2022, 126, (107602), pp 110.CrossRefGoogle Scholar
Frolov, S.M., Shamshin, I.O., Kazachenko, M.V., Aksenov, V.S., Bilera, I.V., Ivanov, V.S. and Zvegintsev, V.I. Polyethylene pyrolysis products: Their detonability in air and applicability to solid-fuel detonation ramjets, Energies, 2021, 14, (4), p 820.CrossRefGoogle Scholar
Lu, J., Zhang, H., Tan, W. and Zheng, L. Isolation of back pressure waves in a valveless pulse detonation combustor with inlet air of high total temperature, Fuel, 2023, 331, (2), p 125866.Google Scholar
Qiu, H., Xiong, C. and Li, J. A theoretical and 1-D numerical investigation on a valve/valveless air-breathing pulse detonation engine, Chin. J. Aeronaut., 2021, 34, (1), pp 6878.CrossRefGoogle Scholar
Gerdroodbary, M.B. Aerodynamic Heating in Supersonic and Hypersonic Flows: Advanced Techniques for Drag and Aero-Heating Reduction, Elsevier, Oxford, UK, 2022.Google Scholar
Gerdroodbary, M.B. Scramjets: Fuel Mixing and Injection Systems, Elsevier, Oxford, UK, 2020.Google Scholar
Matsuo, K., Miyazato, Y. and Kim, H. Shock train and pseudo-shock phenomena in internal gas flows, Progr. Aeronaut. Sci., 1999, 35, (1), pp 33100.CrossRefGoogle Scholar
Chyu, W.J., Kawamura, T. and Bencze, D.P. Navier-Stokes solutions for mixed compression axisymmetric inlet flow with terminal shock, J. Propul. Power, 2015, 5, (1), pp 45.CrossRefGoogle Scholar
Hamed, A. and Shang, J.S. Survey of validation data base for shockwave boundary-layer interactions in supersonic inlets, J. Propul. Power, 1991, 7, (4), pp 617625.CrossRefGoogle Scholar
Sajben, M., Donovan, J.F. and Morris, M.J. Experimental investigation of terminal shock sensors for mixed-compression inlets, J. Propul. Power, 1992, 8, (1), pp 168174.CrossRefGoogle Scholar
Matsuo, K., Sasaguchi, K., Mochizuki, H. and Takechi, N. Investigation of the starting process of a supersonic wind tunnel, Bull. JSME, 2008, 23, (186), pp 19751981.CrossRefGoogle Scholar
Keanini, R.G., Tkacik, P.T., Srivastava, N., Thorsett-Hill, K. and Tomsyck, J. Millisecond-scale shock-train evolution in high-pressure ratio nozzles: Schlieren imaging and qualitative analysis of shock–boundary layer interaction, Proc. Inst. Mech. Eng. Part G, 2013, 228, (7), pp 10761082.CrossRefGoogle Scholar
Mousavi, S.M. and Roohi, E. Large eddy simulation of shock train in a convergent-divergent nozzle, Int. J. Mod. Phys. C, 2014, 25, (4), p 1450003.CrossRefGoogle Scholar
Islam, S., Abir, S.K.S.H. and Hasan, A.B.M.T. Shock train structure in variable pressure condition, AIP Conf. Proc., 2017, 1851, (1), p 020109.CrossRefGoogle Scholar
Waltrup, P.J. and Billing, F.S. Prediction of precombustion wall pressure distributions in scramjet engines, J. Spacecraft Rockets, 1973, 10, (9), pp 620622.CrossRefGoogle Scholar
Heiser, W., Pratt, D., Daley, D. and Mehta, U. Hypersonic Airbreathing Propulsion, AIAA, 1994, Washington, D.C.CrossRefGoogle Scholar
Billing, F.S. Inlet-combustor interface problems in scramjet engines, First International Symposium on Air-Breathing Engine, 1972.Google Scholar
Chang, J., Li, N., Xu, K., Bao, W. and Yu, D. Recent research progress on unstart mechanism, detection and control of hypersonic inlet, Prog. Aeronaut. Sci., 2017, 89, pp 122.CrossRefGoogle Scholar
Sekar, K.R., Karthick, S.K., Jegadheeswaran, S. and Kannan, R. On the unsteady throttling dynamics and scaling analysis in a typical hypersonic inlet–isolator flow, Phys. Fluids, 2020, 32, (12), p 126104.CrossRefGoogle Scholar
Cheng, C., Wang, C. and Cheng, K. Response of an oblique shock train to downstream periodic pressure perturbations, Proc. Inst. Mech. Eng. Part G, 2019, 233, (1), pp 5770.CrossRefGoogle Scholar
Wang, Z., Chang, J., Li, Y., Chen, R., Hou, W., Guo, J. and Yue, L. Oscillation of the shock train under synchronous variation of incoming Mach number and backpressure, Phys. Fluids, 2022, 34, (4), p 046104.CrossRefGoogle Scholar
Gillespie, A. and Sandham, N.D. Shock train response to high-frequency backpressure forcing, AIAA J., 2022, 60, (6), pp 37363748.CrossRefGoogle Scholar
Peng, C., Fan, W., Zheng, L., Wang, Z. and Yuan, C. Experimental investigation on valveless air-breathing dual-tube pulse detonation engines, Appl. Therm. Eng., 2013, 51, (1-2), pp 11161123.CrossRefGoogle Scholar
Sethuraman, V.R.P., Kim, T.H. and Kim, H.D. Effects of back pressure perturbation on shock train oscillations in a rectangular duct, Acta Astronaut., 2021, 179, (6), pp 525535.CrossRefGoogle Scholar
Matsuoka, K., Muto, K., Kasahara, J., Watanabe, H., Matsuo, A. and Endo, T. Development of high-frequency pulse detonation combustor without purging material, J. Propul. Power, 2016, 33, (1), pp 4350.CrossRefGoogle Scholar
Gerdroodbary, M.B., Moradi, R. and Babazadeh, H. Computational investigation of multi hydrogen jets at inclined supersonic flow, Int. J. Energy Res., 2021, 45, (2), pp 16611672.CrossRefGoogle Scholar
Du, S., Al-Rashed, A.A., Gerdroodbary, M.B., Moradi, R., Shahsavar, A. and Talebizadehsardari, P. Effect of fuel jet arrangement on the mixing rate inside trapezoidal cavity flame holder at supersonic flow, Int. J. Hydrogen Energy, 2019, 44, (39), pp 2223122239.CrossRefGoogle Scholar
Sun, C., Gerdroodbary, M.B., Abazari, A.M., Hosseini, S. and Li, Z. Mixing efficiency of hydrogen multijet through backward-facing steps at supersonic flow, Int. J. Hydrogen Energy, 2021, 46, (29), pp 1607516085.CrossRefGoogle Scholar
Liu, X., Gerdroodbary, M.B., Sheikholeslami, M., Moradi, R., Shafee, A. and Li, Z. Effect of strut angle on performance of hydrogen multi-jets inside the cavity at combustion chamber, Int. J. Hydrogen Energy, 2020, 45, (55), pp 3117931187.CrossRefGoogle Scholar
Carroll, B.F. and Dutton, J.C. Characteristics of multiple shock wave/turbulent boundary-layer interactions in rectangular ducts, J. Propul. Power, 1988, 6, (2), pp 186193.CrossRefGoogle Scholar
Kim, Y., Xie, Z.T. and Castro, I.P. A forward stepwise method of inflow generation for LES, AIP Conf. Proc., 2011, pp 134136.CrossRefGoogle Scholar
Kim, Y. Wind Turbine Aerodynamics in Freestream Turbulence, Doctoral thesis, University of Southampton, 2013.Google Scholar
Ma, X., Li, Q., Zhang, Y. and Fan, W. Study on coupling characteristics of supersonic inlet and pulse detonation combustor, J. Northwest. Polytech. Univ., 2023, 41 (2) pp 354362.CrossRefGoogle Scholar