Hostname: page-component-848d4c4894-cjp7w Total loading time: 0 Render date: 2024-06-22T10:26:37.388Z Has data issue: false hasContentIssue false

Linear and nonlinear modelling of a theoretical travelling-wave thermoacoustic heat engine

Published online by Cambridge University Press:  05 February 2015

Carlo Scalo
Affiliation:
Center for Turbulence Research, Stanford, CA 94305, USA
Sanjiva K. Lele
Affiliation:
Department of Aeronautics and Astronautics and Mechanical Engineering, Stanford, CA 94305, USA
Lambertus Hesselink
Affiliation:
Department of Aeronautics and Astronautics and Electrical Engineering, Stanford, CA 94305, USA

Abstract

We have carried out three-dimensional Navier–Stokes simulations, from quiescent conditions to the limit cycle, of a theoretical travelling-wave thermoacoustic heat engine (TAE) composed of a long variable-area resonator shrouding a smaller annular tube, which encloses the hot (HHX) and ambient (AHX) heat exchangers, and the regenerator (REG). Simulations are wall-resolved, with no-slip and adiabatic conditions enforced at all boundaries, while the heat transfer and drag due to the REG and HXs are modelled. HHX temperatures have been investigated in the range 440–500 K with the AHX temperature fixed at 300 K. The initial exponential growth of acoustic energy is due to a network of travelling waves thermoacoustically amplified by looping around the REG/HX unit in the direction of the imposed temperature gradient. A simple analytical model demonstrates that such instability is a localized Lagrangian thermodynamic process resembling a Stirling cycle. An inviscid system-wide linear stability model based on Rott’s theory is able to accurately predict the operating frequency and the growth rate, exhibiting properties consistent with a supercritical Hopf bifurcation. The limit cycle is governed by acoustic streaming – a rectified steady flow resulting from high-amplitude nonlinear acoustics. Its key features are explained with an axially symmetric incompressible model driven by the wave-induced stresses extracted from the compressible calculations. These features include Gedeon streaming, Rayleigh streaming in the resonator, and mean recirculations due to flow separation. The first drives the mean advection of hot fluid from the HHX to a secondary heat exchanger (AHX2), in the thermal buffer tube (TBT), necessary to achieve saturation of the acoustic energy growth. The direct evaluation of the nonlinear energy fluxes reveals that the efficiency of the device deteriorates with the drive ratio and that the acoustic power in the TBT is balanced primarily by the mean advection and thermoacoustic heat transport.

Type
Papers
Copyright
© 2015 Cambridge University Press 

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

Backhaus, S. & Swift, G. W. 1999 A thermoacoustic Stirling heat engine. Nature 399 (6734), 335338.CrossRefGoogle Scholar
Backhaus, S. & Swift, G. W. 2000 A thermoacoustic-Stirling heat engine: detailed study. J. Acoust. Soc. Am. 107, 31483166.Google Scholar
Bauwens, L. 1996 Oscillating flow of a heat-conducting fluid in a narrow tube. J. Fluid Mech. 324, 135161.CrossRefGoogle Scholar
Bejan, A. 2004 Convective Heat Transfer. John Wiley & Sons Inc.Google Scholar
de Blok, C. M.1998 Thermoacoustic system, 1998. Patent. International application number PCT/NL98/00514. Issued 2001.Google Scholar
Boluriaan, S. & Morris, P. J. 2003 Suppression of traveling wave streaming using a jet pump. In Proceedings of the 41st Aerospace Science Meeting and Exhibit. AIAA.Google Scholar
Boluriaan, S. & Morris, P. J. 2009 Acoustic streaming: from Rayleigh to today. Intl J. Aeroacoust. 11 (3–4), 255292.Google Scholar
Ceperley, P. H. 1979 A pistonless Stirling engine – the traveling wave heat engine. J. Acoust. Soc. Am. 66, 12391244.Google Scholar
Gardner, D. L. & Swift, G. W. 2003 A cascade thermoacoustic engine. J. Acoust. Soc. Am. 114 (4), 19051919.Google Scholar
Garrett, S. L. 2004 Resource letter: TA-1: thermoacoustic engines and refrigerators. Am. J. Phys. 72, 1117.Google Scholar
Gary, J., O’Gallagher, A. & Radebaugh, R.1994 A numerical model for regenerator performance, Tech. Rep. NIST-Boulder.Google Scholar
Gedeon, D. 1995 Sage: object-oriented software for cryocooler design. In Cryocoolers 8 (ed. Ross, R. G. Jr), pp. 281292. Springer.Google Scholar
Gedeon, D. 1997 DC gas flows in Stirling and pulse-tube cryocoolers. Cryocoolers 9, 385392.CrossRefGoogle Scholar
Ham, F., Mattsson, K., Iaccarino, G. & Moin, P. 2007 Towards Time-Stable and Accurate LES on Unstructured Grids, Lecture Notes in Computational Science and Engineering, vol. 56. pp. 235249. Springer.Google Scholar
Hamilton, M. F., Ilinksii, Yu. A. & Zabolotskaya, E. A. 2002 Nonlinear two-dimensional model for thermoacoustic engines. J. Acoust. Soc. Am. 111 (5), 20762086.Google Scholar
Hireche, O., Weisman, C., Baltean-Carles, D., Quere, P. L., Francois, M. & Bauwens, L. 2010 Numerical model of a thermoacoustic engine. C. R. Méc. 338 (1), 1823.Google Scholar
In ’T Panhuis, P. H. M. W., Rienstra, S. W., Molenaar, J. & Slot, J. J. M. 2009 Weakly nonlinear thermoacoustics for stacks with slowly varying pore cross-sections. J. Fluid Mech. 618, 4170.Google Scholar
Jensen, B. L., Sumer, B. M. & Fredsøe, J. 1989 Turbulent oscillatory boundary layers at high Reynolds numbers. J. Fluid Mech. 206, 265297.Google Scholar
Karypis, G. & Kumar, V. 1998 Multilevel algorithms for multi-constraint graph partitioning. In Proceedings of Supercomputing ’98.Google Scholar
Kirchhoff, G. 1868 Über den Einfluss der Wärmeleitung in einem Gase auf die Schallbewegung. Pogg. Ann. 134, 177193.Google Scholar
Kramers, H. A. 1949 Vibrations of a gas column. Physica 15 (11–12), 971984.Google Scholar
Lele, S. K. 1994 Compressibility effecs on turbulence. Annu. Rev. Fluid Mech 26, 211254.CrossRefGoogle Scholar
Lighthill, J. 1978 Acoustic streaming. J. Sound Vib. 61 (3), 391418.Google Scholar
Mariappan, S. & Sujith, R. I. 2011 Modelling nonlinear thermoacoustic instability in an electrically heated Rijke tube. J. Fluid Mech. 680, 511533.Google Scholar
Müller, U. A. & Rott, N. 1983 Thermally driven acoustic oscillations, part VI: excitation and power. Z. Angew. Math. Phys. 34 (5), 609626.Google Scholar
Lycklama à Nijeholt, J. A., Tijani, M. E. H. & Spoelstra, S. 2005 Simulation of a traveling-wave thermoacoustic engine using computational fluid dynamics. J. Acoust. Soc. Am. 118 (4), 22652270.CrossRefGoogle Scholar
Olson, J. R. & Swift, G. W. 1997 Acoustic streaming in pulse tube refrigerators: tapered pulse tubes. Cryogenics 37, 769776.Google Scholar
Organ, A. J. 1992 Thermodynamics and Gas Dynamics of the Stirling Cycle Machine. Cambridge University Press.Google Scholar
Penelet, G., Guedra, M., Gusev, V. & Devaux, T. 2012 Simplified account of Rayleigh streaming for the description of nonlinear processes leading to steady state sound in thermoacoustic engines. Intl J. Heat Mass Transfer 55, 60426053.Google Scholar
Penelet, G., Gusev, V., Lotton, P. & Bruneau, M. 2005a Experimental and theoretical study of processes leading to steady-state sound in annular thermoacoustic engines. Phys. Rev. E 72, 016625.CrossRefGoogle ScholarPubMed
Penelet, G., Job, S., Gusev, V., Lotton, P. & Bruneau, M. 2005b Dependence of sounds amplification on temperature distribution in annular thermacoustic engines. Acta Acoust. 91, 567577.Google Scholar
Penelet, G., Gusev, V., Lotton, P. & Bruneau, M. 2006 Nontrivial influence of acoustic streaming on the efficiency of annular thermoacoustic prime movers. Phys. Lett. A 351, 268273.Google Scholar
Poinsot, T. & Veynante, D. 2011 Theoretical and Numerical Combustion, 3rd edn. R. T. Edwards, Inc.Google Scholar
Rott, N. 1969 Damped and thermally driven acoustic oscillations in wide and narrow tubes. Z. Angew. Math. Phys. 20, 230243.CrossRefGoogle Scholar
Rott, N. 1973 Thermally driven acoustic oscillations, part II: stability limit for helium. Z. Angew. Math. Phys. 24, 5472.CrossRefGoogle Scholar
Rott, N. 1974 The influence of heat conduction on acoustic streaming. Z. Angew. Math. Phys. 25, 417421.CrossRefGoogle Scholar
Rott, N. 1975 Thermally driven acoustic oscillations, part III: second-order heat flux. Z. Angew. Math. Phys. 26, 4349.Google Scholar
Rott, N. 1976a Thermally driven acoustic oscillations, part IV: tubes with variable cross-section. Z. Angew. Math. Phys. 27, 197224.CrossRefGoogle Scholar
Rott, N. 1976b Ein ‘Rudimentarer’ Stirlingmotor. Neue Zurecher Ztg. 197 (210).Google Scholar
Rott, N. 1980 Thermoacoustics. Adv. Appl. Mech. 20, 135175.Google Scholar
Rott, N. 1984 Thermoacoustic heating at the closed end of an oscillating gas column. J. Fluid Mech. 145, 19.Google Scholar
Rudenko, O. V. & Soluyan, S. I. 1977 Theoretical Foundations of Non-Linear Acoustics. Acoustic Streaming. Consultants Bureau, Plenum.Google Scholar
Schloegel, K., Karypis, G. & Kumar, V. 2000 Parallel multilevel algorithms for multi-constraint graph partitioning. In Proceedings of the EuroPar-2000 – Parallel Processing, Lecture Notes in Computer Science, vol. 1900, pp. 296310. Springer.Google Scholar
Shang, D. Y. & Wang, B. X. 1990 Effect of variable thermophysical properties on laminar free convection of gas. Intl J. Heat Mass Transfer 33 (7), 13871395.Google Scholar
Swift, G. W. 1988 Thermoacoustic engines. J. Acoust. Soc. Am. 84 (4), 11451181.Google Scholar
Swift, G. W. 1992 Analysis and performance of a large thermoacoustic engine. J. Acoust. Soc. Am. 92 (3), 15511563.Google Scholar
Swift, G. W. & Ward, W. C. 1996 Simple harmonic analysis of regenerators. J. Thermophys. Heat Transfer 10 (4).Google Scholar
Thomas, B. & Pittman, D. 2000 Update on the evaluation of different correlations for the flow friction factor and heat transfer of Stirling engine regenerators. In Energy Conversion Engineering Conference and Exhibit, 2000 (IECEC) 35th Intersociety, vol. 1, pp. 7684.Google Scholar
Thompson, M. W., Atchley, A. A. & Maccarone, M. J. 2004 Influences of a temperature gradient and fluid inertia on acoustic streaming in a standing wave. J. Acoust. Soc. Am. 117 (4 Pt 2), 18391849.Google Scholar
Tijani, M. E. H. & Spoelstra, S. 2011 A high performance thermoacoustic engine. J. Appl. Phys. 110, 093519.Google Scholar
de Waele, A. T. A. M. 2009 Basic treatment of onset conditions and transient effects in thermoacoustic Stirling engines. J. Sound Vib. 325, 974988.Google Scholar
Ward, W. C. & Swift, G. W. 1994 Design environment for low-amplitude thermoacoustic engines. J. Acoust. Soc. Am. 95 (6), 36713672.Google Scholar
Wu, X. & Moin, P. 2008 A direct numerical simulation study on the mean velocity characteristics in turbulent pipe flow. J. Fluid Mech. 608, 81112.Google Scholar
Yazaki, T., Iwata, A., Maekawa, T. & Tominaga, A. 1998 Traveling wave thermoacoustic engine in a looped tube. Phys. Rev. Lett. 81 (15), 31283131.Google Scholar
Zouzoulas, G. & Rott, N. 1976 Thermally driven acoustic oscillations, part V: Gas–liquid oscillations. Z. Angew. Math. Phys. 27 (3), 325334.Google Scholar

Scalo et al. supplementary movie

Instantaneous visualizations of temperature contours (see colorbar) showing streaming of hot fluid in the thermal buffer tube and vorticity magnitude (white) showing intense vortex shedding and transitional turbulence. Data is extracted from case for $T_h=500$ at the limit cycle.

Download Scalo et al. supplementary movie(Video)
Video 7.1 MB

Scalo et al. supplementary movie

Instantaneous visualizations of temperature contours (see colorbar) showing streaming of hot fluid in the thermal buffer tube and vorticity magnitude (white) showing intense vortex shedding and transitional turbulence. Data is extracted from case for $T_h=500$ at the limit cycle.

Download Scalo et al. supplementary movie(Video)
Video 10.6 MB