Skip to main content

Advertisement

Log in

The exergy concept and compressible turbulence

  • Original Article
  • Published:
Theoretical and Computational Fluid Dynamics Aims and scope Submit manuscript

Abstract

Turbulence models facilitated by Kolmogorov’s theory play an important role for compressible flows. Typically the basis of these models is the power spectrum of the velocity \({\mathbf {u}}\) or of the density-weighted velocity \({\mathbf {w}}\equiv \rho ^{1/3}{\mathbf {u}}\). While for incompressible flow the quantity turbulent kinetic energy characterises turbulent motions, from the thermodynamic point of view, due to fluctuations of the density and the temperature other kinds of energies play a role at the different scales in compressible turbulence. We generalise the power spectrum of the velocity \({\mathbf {u}}\) from incompressible flows to compressible flows by introducing the exergy spectrum as an application of the exergy concept. Furthermore, we discuss the application of the concept of turbulent exergy to turbulence modelling and demonstrate this approach with a direct numerical simulation and a Large-Eddy-Simulation of homogeneous isotropic turbulence. The advantage of turbulence modelling based on turbulent exergy is shown on the example of the Approximate Deconvolution Model (ADM) where, at smallest scales for its newly introduced entropy formulation, more available energy is extracted from the flow, and this occurs in a more physical way than for the classical equation set of the model using the total energy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. The literature is not uniform here, the quantity should not be mistaken for Gibbs energy or Helmholtz energy.

  2. An alternative naming of turbulent exergy could be “turbulent available energy”. We found it more clear for the present contribution to use the term turbulent exergy for consistency with the term turbulent kinetic energy.

  3. Note that the terms used in this reference do not necessarily correspond to the commonly used ones in thermodynamics.

References

  1. Bataille, F., Zhou, Y., Bertoglio, J.P.: Energy transfer and triadic interactions in compressible turbulence (NASA/CR-97-206249, ICASE Report No. 97-62) (1997)

  2. Bejan, A.: Entropy Generation Through Heat and Fluid Flow. Wiley, New York (1982)

    Google Scholar 

  3. Bejan, A.: Advanced Engineering Thermodynamics, 4th edn. Wiley, New York (2016)

    Book  Google Scholar 

  4. Bosse, T., Kleiser, L., Meiburg, E.: Small particles in homogeneous turbulence: settling velocity enhancement by two-way coupling. Phys. Fluids 18(2), 027102 (2006)

    Article  Google Scholar 

  5. Chernyshov, A.A., Karelsky, K.V., Petrosyan, A.S.: Development of large eddy simulation for modeling of decaying compressible magnetohydrodynamic turbulence. Phys. Fluids 19(5), 055106 (2007)

    Article  MATH  Google Scholar 

  6. Cho, J., Lazarian, A.: Compressible sub-Alfvénic MHD turbulence in low-\(\beta \) plasmas. Phys. Rev. Lett. 88(24), 245001 (2002)

    Article  Google Scholar 

  7. Eswaran, V., Pope, S.: An examination of forcing in direct numerical simulations of turbulence. Comput. Fluids 16(3), 257–278 (1988)

    Article  MATH  Google Scholar 

  8. Federrath, C.: On the universality of supersonic turbulence. Mon. Not. R. Astron. Soc. 436(2), 1245–1257 (2013)

    Article  Google Scholar 

  9. Fureby, C., Tabor, G., Weller, H., Gosman, A.: A comparative study of subgrid scale models in homogeneous isotropic turbulence. Phys. Fluids 9(5), 1416–1429 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  10. Garnier, E., Adams, N., Sagaut, P.: Large Eddy Simulation for Compressible Flows. Springer, Berlin (2009)

    Book  MATH  Google Scholar 

  11. Gatski, T.B.: Modeling compressibility effects on turbulence. In: Métais, O. (ed.) New Tools in Turbulence Modelling, pp. 73–104. Springer, Berlin (1997)

    Chapter  Google Scholar 

  12. Grete, P., Vlaykov, D.G., Schmidt, W., Schleicher, D.R.G.: A nonlinear structural subgrid-scale closure for compressible MHD. II. A priori comparison on turbulence simulation data. Phys. Plasmas 23(6), 062317 (2016)

    Article  Google Scholar 

  13. Grete, P., Vlaykov, D.G., Schmidt, W., Schleicher, D.R.G.: Comparative statistics of selected subgrid-scale models in large-eddy simulations of decaying, supersonic magnetohydrodynamic turbulence. Phys. Rev. E 95, 033206 (2017)

    Article  Google Scholar 

  14. Hanifi, A., Schmid, P.J., Henningson, D.S.: Transient growth in compressible boundary layer flow. Phys. Fluids 8(3), 826–837 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hickel, S., Adams, N.A., Domaradzki, J.A.: An adaptive local deconvolution method for implicit LES. J. Comput. Phys. 213(1), 413–436 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  16. Honein, A.E., Moin, P.: Higher entropy conservation and numerical stability of compressible turbulence simulations. J. Comput. Phys. 201(2), 531–545 (2004)

    Article  MATH  Google Scholar 

  17. Ishihara, T., Gotoh, T., Kaneda, Y.: Study of high-Reynolds number isotropic turbulence by direct numerical simulation. Annu. Rev. Fluid Mech. 41, 165–180 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  18. Jagannathan, S., Donzis, D.A.: Reynolds and Mach number scaling in solenoidally-forced compressible turbulence using high-resolution direct numerical simulations. J. Fluid Mech. 789, 669–707 (2016)

    Article  MathSciNet  Google Scholar 

  19. Jocksch, A.: Direct numerical simulation of turbulent spots in high-speed boundary layers. Ph.D. thesis, ETH Zurich, Zurich (2009)

  20. Jocksch, A., Kleiser, L.: Exergetic aspects of high-speed boundary layers. Proc. Appl. Math. Mech. 6(1), 529–530 (2006)

    Article  Google Scholar 

  21. Jocksch, A., Kraushaar, M., Daverio, D.: Optimized all-to-all communication on multicore architectures applied to FFTs with pencil decomposition. Pract. Exper. Concurr. Comput. 31, e4964 (2018)

    Google Scholar 

  22. Karlsson, S.: Energy, entropy and exergy in the atmosphere. Ph.D. thesis, Chalmers University of Technology, Göteborg (1990)

  23. Kolmogorov, A.N.: The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. Proc. R. Soc. Lond. A 434, 9–13 (1991). Reprint from Dokl. Akad. Nauk SSSR 30, 301 (1941)

  24. Kotas, T.: The Exergy Method of Thermal Analysis. Butterworths, London (1985)

    Google Scholar 

  25. Kraichnan, R.H.: The structure of isotropic turbulence at very high Reynolds numbers. J. Fluid Mech. 5(4), 497–543 (1959)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kritsuk, A.G., Norman, M.L., Padoan, P., Wagner, R.: The statistics of supersonic isothermal turbulence. Astrophys. J. 665(1), 416–431 (2007)

    Article  Google Scholar 

  27. Lele, S.K.: Compressibility effects on turbulence. Annu. Rev. Fluid Mech. 26(1), 211–254 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  28. LeVeque, R.J.: Finite Volume Methods for Hyperbolic Problems. Cambridge Texts in Applied Mathematics. Cambridge University Press, Cambridge (2002). https://doi.org/10.1017/CBO9780511791253

    Book  MATH  Google Scholar 

  29. Lior, N., Sarmiento-Darkin, W., Al-Sharqawi, H.S.: The exergy fields in transport processes: their calculation and use. Energy 31(5), 553–578 (2006)

    Article  Google Scholar 

  30. Mathew, J., Lechner, R., Foysi, H., Sesterhenn, J., Friedrich, R.: An explicit filtering method for large eddy simulation of compressible flows. Phys. Fluids 15(8), 2279–2289 (2003)

    Article  MATH  Google Scholar 

  31. Müller, S.: Numerical investigations of compressible turbulent swirling jet flows. Ph.D. thesis, ETH Zurich, Zurich (2007)

  32. Myers, M.K.: Transport of energy by disturbances in arbitrary steady flows. J. Fluid Mech. 226, 383–400 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  33. Petersen, M.R., Livescu, D.: Forcing for statistically stationary compressible isotropic turbulence. Phys. Fluids 22(11), 116101 (2010)

    Article  Google Scholar 

  34. Pope, S.B.: Turbulent Flows. Cambridge University Press, Cambridge (2000)

    Book  MATH  Google Scholar 

  35. Sagaut, P., Cambon, C.: Homogeneous Turbulence Dynamics, 2nd edn. Springer, Berlin (2018). https://doi.org/10.1007/978-3-319-73162-9

    Book  MATH  Google Scholar 

  36. Schlichting, H., Gersten, K.: Boundary-Layer Theory, 9th edn. Springer, Berlin (2017)

    Book  MATH  Google Scholar 

  37. Sciubba, E., Zullo, F.: Exergy dynamics of systems in thermal or concentration non-equilibrium. Entropy 19(6), 263 (2017)

    Article  Google Scholar 

  38. Sciubba, E., Zullo, F.: On the quantification of non-equilibrium exergy for thermodynamic systems evolving according to Cattaneo’s equation. Int. J. Thermodyn. 22(1), 19–24 (2019)

    Article  Google Scholar 

  39. Stolz, S., Adams, N.A., Kleiser, L.: The approximate deconvolution model for large-eddy simulations of compressible flows and its application to shock-turbulent-boundary-layer interaction. Phys. Fluids 13(10), 2985–3001 (2001)

    Article  MATH  Google Scholar 

  40. Subbareddy, P.K., Candler, G.V.: A fully discrete, kinetic energy consistent finite-volume scheme for compressible flows. J. Comput. Phys. 228(5), 1347–1364 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  41. Szargut, J.: Exergy Method: Technical and Ecological Applications. WIT Press, Southampton (2005)

    Google Scholar 

  42. Tailleux, R.: Available potential energy and exergy in stratified fluids. Annu. Rev. Fluid Mech. 45(1), 35–58 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  43. Visbal, M.R., Gaitonde, D.V.: High-order-accurate methods for complex unsteady subsonic flows. AIAA J. 37(10), 1231–1239 (1999)

    Article  Google Scholar 

  44. Xie, C., Wang, J., Li, H., Wan, M., Chen, S.: A modified optimal LES model for highly compressible isotropic turbulence. Phys. Fluids 30(6), 065108 (2018)

    Article  Google Scholar 

Download references

Acknowledgements

The author would like to thank Leonhard Kleiser (Institute of Fluid Dynamics, ETH Zurich), William Sawyer (CSCS), Stephan Brunner, Laurent Villard and Claudio Gheller (Swiss Plasma Center, EPFL) for helpful discussions. Furthermore he would like to thank the anonymous reviewers for their constructive remarks.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andreas Jocksch.

Ethics declarations

Conflict of interest:

The author declares that he has no conflict of interest.

Additional information

Communicated by Sergio Pirozzoli.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Relations of exergy expressions

From Eq. (1), the averaged exergy per unit volume is expressed as

$$\begin{aligned} {\bar{\rho }}{\tilde{\epsilon }}={\bar{\rho }}{\tilde{e}}-{\bar{\rho }}e_r +p_r\left( 1-\frac{{\bar{\rho }}}{\rho _r}\right) -T_r({\bar{\rho }}{\tilde{s}} -{\bar{\rho }}s_r)+\frac{1}{2}{\bar{\rho }} {\widetilde{(u_j-u_{j,r})(u_j-u_{j,r})}}. \end{aligned}$$
(31)

Together with the special reference state a (Eq. 4), we obtain the turbulent exergy per unit volume

$$\begin{aligned} {\bar{\rho }}\epsilon _t={\bar{\rho }}{\tilde{\epsilon }}(r=a) ={\bar{\rho }}{\tilde{e}}-{\bar{\rho }}e_a+\frac{1}{2}{\bar{\rho }} {\widetilde{(u_j-u_{j,a}) (u_j-u_{j,a})}}. \end{aligned}$$
(32)

If we subtract Eq. (32) from Eq. (31)

$$\begin{aligned}&{\bar{\rho }}{\tilde{\epsilon }}-{\bar{\rho }}\epsilon _t =\underbrace{{\bar{\rho }}{\tilde{e}}-{\bar{\rho }}e_r+p_r \left( 1-\frac{{\bar{\rho }}}{\rho _r}\right) -T_r({\bar{\rho }}{\tilde{s}} -{\bar{\rho }}s_r)+\frac{1}{2}{\bar{\rho }} {\widetilde{(u_j-u_{j,r})(u_j-u_{j,r})}}}_{{\bar{\rho }} {\tilde{\epsilon }}}\nonumber \\&\quad -\underbrace{\left[ {\bar{\rho }}{\tilde{e}} -{\bar{\rho }}e_a+\frac{1}{2}{\bar{\rho }} {\widetilde{(u_j-u_{j,a})(u_j-u_{j,a})}}\right] }_{{\bar{\rho }} \epsilon _t}~, \end{aligned}$$
(33)

we obtain the exergy per unit volume for the special reference state a

$$\begin{aligned} ={\bar{\rho }}e_a-{\bar{\rho }}e_r+p_r\frac{{\bar{\rho }}}{\rho _a} \left( 1-\frac{\rho _a}{\rho _r}\right) -T_r({\bar{\rho }}s_a-{\bar{\rho }}s_r) +\frac{1}{2}{\bar{\rho }}(u_{j,a}-u_{j,r})(u_{j,a}-u_{j,r})~, \end{aligned}$$
(34)

which shows that the exergy is the sum of the turbulent exergy and the exergy of the special reference state a (Eq. 7).

Acoustic energy density

The exergy of compressible flow is related to the acoustic energy density as summarised subsequently: For small disturbances and a perfect gas, the equation for the exergy (Eq. 1) multiplied with \(\rho \) and expanded with a Taylor series around the reference state r becomes after Reynolds averaging the equation of the disturbance energy per unit volume

$$\begin{aligned} \epsilon _d=\frac{1}{2}\left( \frac{T_r}{\rho _r\gamma M^2}\overline{\rho '^2}+\frac{\rho _r}{\gamma (\gamma -1)T_rM^2}\overline{T'^2} +\rho _r(\overline{u_i'u_i'})\right) ~, \end{aligned}$$
(35)

introduced in Ref. [14] as an integral formulation, where \(f'=f-f_r\) represents the deviation from the average reference or base state \(f_r={\bar{f}}\). This expression is equivalent to

$$\begin{aligned} \epsilon _d=\frac{1}{2}\left( \frac{1}{\gamma p_r}\overline{p'^2}+(\gamma -1)\gamma M^4p_r\overline{s'^2}+\rho _r(\overline{u_i'u_i'})\right) ~, \end{aligned}$$
(36)

(see Ref. [35] and references therein) and is as the exergy, positive definite. For isentropic fluctuations (\(s'=0\)), Eq. (36) reduces to the well-known formula for the acoustic energy density with the contributions of potential and kinetic energy. A more general expression of disturbance energy which is closely related to the exergy has been derived in Ref. [32].

Fourier filter

In Sect. 3, the filtering procedure was introduced based on cancelling out Fourier modes in spectral space. One can obtain the same result with a real space Fourier filter which might be more illustrative, when compared to our incompressible flow analogy (Fig. 2). Note that in Sect. 3 the Fourier modes were in the range \(0\le k\le N/2\), and the complex conjugate Fourier coefficients were not stored. The zero mode and odd-ball mode N/2 coefficients were real. Here we use the notation that the range of Fourier modes is \(0\le k\le N-1\) (conjugate complex coefficients are represented in contrast to the definition of Sect. 3 and Fourier modes are transformed to positive k for simplicity).

For the one-dimensional case, the Fourier filter is constructed from the discrete Fourier transform forward and backward matrices, where the entries of the forward transformation matrix \({\mathbf {W}}\) are

$$\begin{aligned} w_{jl}=\frac{e^{(-2\pi i/N)jl}}{\sqrt{N}}\qquad 0\le j,l<N-1~, \end{aligned}$$
(37)

and the backward transformation matrix \({\mathbf {W}}^{-1}\) is

$$\begin{aligned} {\mathbf {W}}^{-1}={\mathbf {W}}^*~, \end{aligned}$$
(38)

where here \(()^*\) denotes conjugate complex numbers. Together with the matrix \({\mathbf {G}}_k\), which cancels out Fourier modes k and \(N+1-k\) in spectral space

$$\begin{aligned} g_{jl,k}= {\left\{ \begin{array}{ll} 1 &{}\quad j=l\ne k \wedge j=l\ne N+1-k~,\\ 0 &{}\quad \text {otherwise}~, \end{array}\right. } \end{aligned}$$
(39)

the filter matrix is

$$\begin{aligned} {\mathbf {F}}_k={\mathbf {W}}^{-1}{\mathbf {G}}_k{\mathbf {W}}~. \end{aligned}$$
(40)

We filter the vector

$$\begin{aligned} {\varvec{\varTheta }}= \begin{bmatrix} {\varvec{\sigma }}_0\\ {\varvec{\sigma }}_1\\ \vdots \\ {\varvec{\sigma }}_{N-1} \end{bmatrix}~, \end{aligned}$$
(41)

of the control volumes \(0,\ldots ,N-1\) of the state vector \({\varvec{\sigma }}\) mode by mode. The stepwise filtering procedure can be expressed recursively as

$$\begin{aligned} {\varvec{\varPhi }}_k= {\left\{ \begin{array}{ll} {\varvec{\varTheta }} &{}\quad k=k_{\max }~,\\ {\mathbf {F}}_{k+1}{\varvec{\varPhi }}_{k+1} &{}\quad 0\le k<k_{\max }~, \end{array}\right. } k_{\max }=N/2~, \end{aligned}$$
(42)

for the filter steps \(k,k-1,\ldots ,\)0. The computation of the exergy of mode k with respect to mode \(k-1\)

$$\begin{aligned} \begin{aligned} \epsilon _k=\sum _{l=0}^{N-1}\sum _{j=0}^{N-1}\epsilon \Big (&{\varvec{\sigma }} =\begin{bmatrix}\phi _{j1}&\phi _{j2}&\cdots&\phi _{j5}\end{bmatrix}_k, \\&{\varvec{\sigma }}_r=\begin{bmatrix}\phi _{j1}&\phi _{j2}&\cdots&\phi _{j5}\end{bmatrix}_{k-1}\Big )f_{jl, k}~,~k>0~, \end{aligned} \end{aligned}$$
(43)

is done with the filter coefficients \(f_{jl, k}\). In order to exploit the favourable computational complexity of the FFT algorithm scaling with \(N\log N\), one would typically compute \(\epsilon _k\) as the difference between the exergy of the unfiltered and filtered solution with respect to one set of modes using FFTs for the filtering step (Sect. 3). The extension of the filtering procedure to multiple dimensions is straightforward.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jocksch, A. The exergy concept and compressible turbulence. Theor. Comput. Fluid Dyn. 34, 271–286 (2020). https://doi.org/10.1007/s00162-020-00533-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-020-00533-z

Keywords

Navigation