Skip to main content
Log in

A Kinetic Energy Preserving DG Scheme Based on Gauss–Legendre Points

  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

In the context of numerical methods for conservation laws, not only the preservation of the primary conserved quantities can be of interest, but also the balance of secondary ones such kinetic energy in case of the Euler equations of gas dynamics. In this work, we construct a kinetic energy preserving discontinuous Galerkin method on Gauss–Legendre nodes based on the framework of summation-by-parts operators. For a Gauss–Legendre point distribution, boundary terms require special attention. In fact, stability problems will be demonstrated for a combination of skew-symmetric and boundary terms that disagrees with exclusively interior nodal sets. We will theoretically investigate the required form of the corresponding boundary correction terms in the skew-symmetric formulation leading to a conservative and consistent scheme. In numerical experiments, we study the order of convergence for smooth solutions, the kinetic energy balance and the behaviour of different variants of the scheme applied to an acoustic pressure wave and a viscous shock tube. Using Gauss–Legendre nodes results in a more accurate approximation in our numerical experiments for viscous compressible flow. Moreover, for two-dimensional decaying homogeneous turbulence, kinetic energy preservation yields a better representation of the energy spectrum.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

References

  1. Allaneau, Y., Jameson, A.: Kinetic energy conserving discontinuous Galerkin scheme. In: Proceedings of the 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition. AIAA-2011-198 (2011)

  2. Bassi, F., Franchina, N., Ghidoni, A., Rebay, S.: A numerical investigation of a spectral-type nodal collocation discontinuous Galerkin approximation of the Euler and Navier-Stokes equations. Int. J. Numer. Methods Fluids 71(10), 1322–1339 (2013)

    Article  MathSciNet  Google Scholar 

  3. Bassi, F., Rebay, S.: Numerical evaluation of two discontinuous Galerkin methods for the compressible Navier–Stokes equations. Int. J. Numer. Methods Fluids 40, 197–207 (2002)

    Article  MATH  Google Scholar 

  4. Carpenter, M.H., Nordström, J., Gottlieb, D.: A stable and conservative interface treatment of arbitrary spatial accuracy. J. Comput. Phys. 148, 341–365 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chandrashekar, P.: Kinetic energy preserving and entropy stable finite volume schemes for compressible Euler and Navier–Stokes equations. Commun. Comput. Phys. 14, 1252–1286 (2013)

    Article  MathSciNet  Google Scholar 

  6. Del Rey Fernández, D.C., Boom, P.D., Zingg, D.W.: A generalized framework for nodal first derivative summation-by-parts operators. J. Comput. Phys. 266, 214–239 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  7. Fisher, T.C., Carpenter, M.H., Nordström, J., Yamaleev, N.K., Swanson, C.: Discretely conservative finite-difference formulations for nonlinear conservation laws in split form: theory and boundary conditions. J. Comput. Phys. 234, 353–375 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  8. Gassner, G.J.: A skew-symmetric discontinuous Galerkin spectral element discretization and its relation to SBP-SAT finite difference methods. SIAM J. Sci. Comput. 35, A1233–A1253 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  9. Gassner, G.J.: A kinetic energy preserving nodal discontinuous Galerkin spectral element method. Int. J. Numer. Methods Fluids 76, 28–50 (2014)

    Article  MathSciNet  Google Scholar 

  10. Gassner, G.J., Beck, A.D.: On the accuracy of high-order discretizations for underresolved turbulence simulations. Theor. Comput. Fluid Dyn. 27, 221–237 (2013)

    Article  Google Scholar 

  11. Gassner, G.J., Kopriva, D.A.: A comparison of the dispersion and dissipation errors of Gauss and Gauss–Lobatto discontinuous Galerkin spectral element methods. SIAM J. Sci. Comput. 33(5), 2560–2579 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  12. Hicken, J.E., Del Rey Fernández, D.C., Zingg, D.W.: Multidimensional summation-by-parts operators: General theory and application to simplex elements. SIAM J. Sci. Comput. 38(4), A1935–A1958 (2016)

  13. Ishiko, K., Ohnishi, N., Ueno, K., Sawada, K.: Implicit large eddy simulation of two-dimensional homogeneous turbulence using weighted compact nonlinear scheme. J. Fluids Eng. 131, 061401:1–061401:14 (2009)

    Article  Google Scholar 

  14. Ismail, F., Roe, P.L.: Affordable, entropy-consistent Euler flux functions II: entropy production at shocks. J. Comput. Phys. 228, 5410–5436 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  15. Jameson, A.: Formulation of kinetic energy preserving conservative schemes for gas dynamics and direct numerical simulation of one-dimensional viscous compressible flow in a shock tube using entropy and kinetic energy preserving schemes. J. Sci. Comput. 34, 188–208 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  16. Javadi, A., Pasandideh-Fard, M., Malek-Jafarian, M.: Modification of k-\(\epsilon \) turbulent model using kinetic energy-preserving method. Numer. Heat Transf. Part B Fundam. 68, 554–577 (2015)

    Article  Google Scholar 

  17. Kopriva, D.A., Gassner, G.J.: On the quadrature and weak form choices in collocation type discontinuous Galerkin spectral element methods. J. Sci. Comput. 44, 136–155 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  18. Kreiss, H.O., Scherer, G.: Finite element and finite difference methods for hyperbolic partial differential equations. In: Boor, C.D. (ed.) Mathematical Aspects of Finite Elements in Partial Differential Equations, pp. 195–212. Academic Press, London (1974)

    Chapter  Google Scholar 

  19. Morinishi, Y.: Skew-symmetric form of convective terms and fully conservative finite difference schemes for variable density low-mach number flows. J. Comput. Phys. 229, 276–300 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  20. Morinishi, Y., Lund, T., Vasilyev, O., Moin, P.: Fully conservative higher order finite difference schemes for incompressible flow. J. Comput. Phys. 143, 90–124 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  21. Nordström, J.: Conservative finite difference formulations, variable coefficients, energy estimates and artificial dissipation. J. Sci. Comput. 29, 375–404 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  22. Nordström, J., Forsberg, K., Adamsson, C., Eliasson, P.: Finite volume methods, unstructured meshes and strict stability for hyperbolic problems. Appl. Numer. Math. 45, 453–473 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  23. Ranocha, H., Öffner, P., Sonar, T.: Summation-by-parts operators for correction procedure via reconstruction. J. Comput. Phys. 311, 299–328 (2016)

  24. San, O., Staples, A.E.: High-order methods for decaying two-dimensional homogeneous isotropic turbulence. Comput. Fluids 63, 105–127 (2012)

    Article  MathSciNet  Google Scholar 

  25. Stelling, G.S., Duinmeijer, S.P.A.: A staggered conservative scheme for every Froude number in rapidly varied shallow water flows. Int. J. Numer. Methods Fluids 43, 1329–1354 (2003)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  27. Svärd, M., Nordström, J.: Review of summation-by-parts schemes for initial-boundary-value problems. J. Comput. Phys. 268, 17–38 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  28. Teukolsky, S.A.: Short note on the mass matrix for Gauss-Lobatto grid points. J. Comput. Phys. 283, 408–413 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  29. van Leer, B.: Flux-vector splitting for the Euler equations. In: Krause, E. (ed.) Eighth International Conference on Numerical Methods in Fluid Dynamics, Lecture Notes in Physics, vol. 170, pp. 507–512. Springer, Berlin (1982)

    Chapter  Google Scholar 

  30. van’t Hof, B., Veldman, A.E.: Mass, momentum and energy conserving (MAMEC) discretizations on general grids for the compressible Euler and shallow water equations. J. Comput. Phys. 231, 4723–4744 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  31. Verstappen, R., Veldman, A.: Symmetry-preserving discretization of turbulent flow. J. Comput. Phys. 187, 343–368 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  32. Yu, J., Yan, C., Jiang, Z.: On the use of the discontinuous galerkin method for numerical simulation of two-dimensional compressible turbulence with shocks. Sci. China Phys. Mech. Astron. 57(9), 1758–1770 (2014)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sigrun Ortleb.

Appendix

Appendix

In the following, the proofs of Lemma 2 in Sect. 3 and Lemma 3 in Sect. 5.1 are given.

1.1 Proof of Lemma 2

For the assertion of conservativity, we multiply the scheme (25) from left by \(\underline{1}^T\underline{\underline{M}}\). Thus, the rate of change of the cell mean within a cell \([x_i,x_{i+1}]\) multiplied by the cell size \(\varDelta x\) is given by

$$\begin{aligned} \underline{1}^T\underline{\underline{M}}\frac{\varDelta x_i}{2}\,\frac{d}{dt} \underline{u}^i= & {} - \underline{1}^T\underline{\underline{M}}\underline{\underline{D}}\underline{f}^i - \underline{1}^T\underline{\underline{M}}\left[ -\underline{\underline{D}}\underline{\underline{\alpha }}^i\underline{\beta }^i + \underline{\underline{\alpha }}^i\underline{\underline{D}}\underline{\beta }^i + \underline{\underline{\beta }}^i\underline{\underline{D}}\underline{\alpha }^i\right] \\&+\, \underline{1}^T[(f^i_{h} + \underline{\underline{\alpha }}^i\beta ^i_h + k^i_h)\underline{L}]_{-1}^{1} - \underline{1}^T[f^{*,i}\underline{L}]_{-1}^{1} -\underline{1}^T[\underline{\underline{\alpha }}^i\beta ^{*,i}\underline{L}]_{-1}^{1}\\&+\, \underline{1}^T\left( k_i^{*,+}\underline{L}(-1) - k_i^{*,-}\underline{L}(1)\right) \\= & {} - \underline{1}^T(\underline{\underline{B}}-\underline{\underline{D}}^T\underline{\underline{M}}) \underline{f}^i\\&- \,\underline{1}^T\left[ -(\underline{\underline{B}}-\underline{\underline{D}}^T\underline{\underline{M}})\underline{\underline{\alpha }}^i\underline{\beta }^i + \underline{\underline{\alpha }}^i(\underline{\underline{B}}-\underline{\underline{D}}^T\underline{\underline{M}})\underline{\beta }^i + \underline{\underline{\beta }}^i\underline{\underline{M}}\underline{\underline{D}}\underline{\alpha }^i\right] \\&+ \,\underline{1}^T[(f^i_{h} + \underline{\underline{\alpha }}\beta ^i_h + k^i_h)\underline{L}]_{-1}^{1} -\underline{1}^T[f^{*,i}\underline{L}]_{-1}^{1} -\underline{1}^T[\underline{\underline{\alpha }}^i\beta ^{*,i}\underline{L}]_{-1}^{1}\\&+ \,\underline{1}^T\left( k_i^{*,+}\underline{L}(-1) - k_i^{*,-}\underline{L}(1)\right) , \end{aligned}$$

using the SBP property \(\underline{\underline{M}}\underline{\underline{D}}= \underline{\underline{B}}- \underline{\underline{D}}^T\underline{\underline{M}}\).

Furthermore, the interpolation of boundary values and the corresponding boundary terms are encoded in the matrix \(\underline{\underline{B}}\). Thus, due to Lemma 1 we can replace the corresponding terms by \(\underline{\underline{B}}\underline{f}^i = [f^i_{h}\underline{L}]_{-1}^{1},\ \underline{\underline{B}}\underline{\underline{\alpha }}^i\underline{\beta }^i = [(\alpha \beta )^i_{h}\underline{L}]_{-1}^{1}\) and \((\underline{\alpha }^i)^T\underline{\underline{B}}\underline{\beta }^i = \underline{1}^T\underline{\underline{\alpha }}^i[\beta ^i_{h}\underline{L}]_{-1}^{1}\). Since this cancels out the terms containing \(f^i_h\) and \(\underline{\underline{\alpha }}^i\beta ^i_h\), it holds that

$$\begin{aligned} \underline{1}^T\underline{\underline{M}}\frac{\varDelta x_i}{2}\,\frac{d}{dt} \underline{u}^i= & {} (\underline{\underline{D}}\underline{1})^T\underline{\underline{M}}\underline{f}^i - \left[ (\underline{\underline{D}}\underline{1})^T\underline{\underline{M}}\underline{\underline{\alpha }}^i\underline{\beta }^i - (\underline{\alpha }^i)^T\underline{\underline{D}}^T\underline{\underline{M}}\underline{\beta }^i + (\underline{\beta }^i)^T\underline{\underline{M}}\underline{\underline{D}}\underline{\alpha }^i\right] \\&+\, \underline{1}^T[\left( (\alpha \beta )^i_{h}+k^i_h\right) \underline{L}]_{-1}^{1}\\&- \,\underline{1}^T[f^{*,i}\underline{L}]_{-1}^{1} - \underline{1}^T[\underline{\underline{\alpha }}^i\beta ^{*,i}\underline{L}]_{-1}^{1} + \underline{1}^T\left( k_i^{*,+}\underline{L}(-1) - k_i^{*,-}\underline{L}(1)\right) . \end{aligned}$$

Now, discrete differentiation yields \(\underline{\underline{D}}\underline{1} = \underline{0}\) and obviously, \(-\underline{\alpha }^T\underline{\underline{D}}^T\underline{\underline{M}}\underline{\beta }+ \underline{\beta }^T\underline{\underline{M}}\underline{\underline{D}}\underline{\alpha }= \underline{0}\) holds. This reduces the rate of change of mass contained in the specific cell to

$$\begin{aligned} \underline{1}^T\underline{\underline{M}}\frac{\varDelta x_i}{2}\,\frac{d}{dt} \underline{u}^i= & {} \underline{1}^T[\left( (\alpha \beta )^i_{h} + k^i_h\right) \underline{L}]_{-1}^{1} -\underline{1}^T[f^{*,i}\underline{L}]_{-1}^{1} -\underline{1}^T[\underline{\underline{\alpha }}^i\beta ^{*,i}\underline{L}]_{-1}^{1}\\&+\, \underline{1}^T\left( k_i^{*,+}\underline{L}(-1) - k_i^{*,-}\underline{L}(1)\right) \end{aligned}$$

For conservativity, the remaining terms may only contain interface contributions. Hence, the balance of flux contributions at cell interfaces has to be investigated. We consider the interface with index \((i-1,i)\) between two cells \([x_{i-1},x_i]\) and \([x_i,x_{i+1}]\). As the interpolation property yields \(\underline{1}^T\underline{L}(\xi ) = 1\), we obtain the following fluxes over cell interfaces. Denoting by \(C^{*,i-1}_{i-1,i}\) the corresponding flux from left to right based on the values in the left cell, neglecting the term \(f_{i-1,i}^*\), we have

$$\begin{aligned} C^{*,i-1}_{i-1,i} = -(\alpha \beta )^{i-1}_{h}(1) - k^{i-1}_h(1) + \alpha ^{i-1}_h(1)\beta _{i-1,i}^*+ k_{i-1}^{*,-}, \end{aligned}$$

whereas the analogous quantity \(C^{*,i}_{i-1,i}\) based on the right cell is given by

$$\begin{aligned} C^{*,i}_{i-1,i} = -(\alpha \beta )^i_{h}(-1) - k^i_h(-1) + \alpha _h^i(-1)\beta _{i-1,i}^*+ k_{i}^{*,+} . \end{aligned}$$

Hence, conservativity precisely requires

$$\begin{aligned} C^{*,i-1}_{i-1,i} = C^{*,i}_{i-1,i} =: C^{*}_{i-1,i}, \end{aligned}$$

yielding the first assertion.

As in the second assertion, we now assume that \(k^i_h\) only depends on a combination of the interior values \(\underline{\alpha }^i,\underline{\beta }^i\) and that \(k_i^{*,+}\) only depends on \(\beta _{i-1,i}^*\) and \(\alpha ^i_h(-1)\). If we then modify any of the input values other than \(\underline{\alpha }^i,\underline{\beta }^i,\beta ^*_{i-1,i}\), the value of \(C^{*,i}_{i-1,i}\) does not change and thus \(C^{*}_{i-1,i}\) remains constant as well. If we furthermore assume that \(k_i^{*,-}\) only depends on \(\beta _{i,i+1}^*\) and \(\alpha ^i_h(1)\), modifying \(\underline{\alpha }^i\) or \(\underline{\beta }^{i}\) does not change \(C^{*,i-1}_{i-1,i}\) and hence does not change \(C^{*}_{i-1,i}\). Therefore, \(C^{*}_{i-1,i}\) may only depend on \(\beta ^*_{i-1,i}\) which proves assertion 2.

Now, consistency in the finite volume sense refers to a consistent numerical flux. With \(k^i_h\) and \(k_i^{*,\pm }\) set as in the third assertion, the scheme (25) for \(N=0\) reduces to

$$\begin{aligned} \varDelta x_i\frac{d}{dt} u^i= & {} [f^{*,i}]_{-1}^{1} -[\alpha ^i\beta ^{*,i}]_{-1}^{1} + \left( k_i^{*,+} - k_i^{*,-}\right) = [f^{*,i}]_{-1}^{1} - \left( C_{i,i+1}^{*} - C_{i-i,i}^{*}\right) . \end{aligned}$$

Since \(f^*\) is consistent to f with \(f_{i-1,i}^*(u,u) = f(u)\) and \(\beta ^*\) is consistent to \(\beta \), this above finite volume scheme is consistent if and only if

$$\begin{aligned} f^*_{i-1,i}(u,u) + C^*_{i-1,i}(\beta _{i-1,i}^{*}(u,u)) = f(u) + C^*_{i-1,i}(\beta (u)) = f(u), \end{aligned}$$

for any interface index \((i-1,i)\) and any value of the conserved variable u. Hence, we have the requirement \(C_{i-1,i}(\beta (u)) = 0\), which yields the last assertion.

1.2 Proof of Lemma 3

The columns of the matrices \(\underline{\underline{B_\xi }}= \underline{\underline{B}}\otimes \underline{\underline{M}}_{1D}\) and \(\underline{\underline{B_\eta }}= \underline{\underline{M}}_{1D}\otimes \underline{\underline{B}}\) can be related to the Lagrange polynomials as follows. Let \(\nu (i,j)=(i-1)(N+1)+j\). From the proof of Lemma 1 in Sect. 2, we recall that \(\underline{\underline{B}}\) has the entries \(B_{jk}=[L_jL_k]^1_{-1}\) and \(\underline{\underline{M}}_{1D}\) is diagonal with entries \(\omega _j\). Therefore, the \(\nu \)th columns of the above matrices \(\underline{\underline{B_\xi }}\) and \(\underline{\underline{B_\eta }}\) are given by \(\left. \underline{B_\xi }\right. _{\nu (i,j)} = [\underline{L}(\xi )L_i(\xi )]^1_{-1}\otimes \omega _j\underline{\epsilon }_j\) and \(\left. \underline{B_\eta }\right. _{\nu (i,j)} = \omega _i\underline{\epsilon }_i\otimes [\underline{L}(\xi )L_j(\xi )]^1_{-1}\), respectively, where \(\underline{\epsilon }_j\) denotes the jth unit vector.

Therefore, we obtain

$$\begin{aligned} \underline{\underline{B_\xi }}\underline{g}^\xi= & {} \sum _{i=1}^{N+1}\sum _{j=1}^{N+1}\left. \underline{B_\xi }\right. _{\nu (i,j)} g^\xi _{\nu (i,j)} = \sum _{i=1}^{N+1}\sum _{j=1}^{N+1}\left( [\underline{L}(\xi )L_i(\xi )]^1_{-1}\otimes \omega _j\underline{\epsilon }_j\right) g^\xi _{\nu (i,j)}\\= & {} \left[ \sum _{j=1}^{N+1}\omega _j\left( \sum _{i=1}^{N+1}L_i(\xi ) g^\xi _{\nu (i,j)}\right) \underline{L}(\xi )\otimes \underline{\epsilon }_j\right] ^1_{-1} = \left[ \sum _{j=1}^{N+1}\omega _jg^\xi _h(\xi ,\eta _j)\underline{L}(\xi )\otimes \underline{L}(\eta _j)\right] ^1_{-1}. \end{aligned}$$

Since for the normal vectors and Gauss nodes on edges \(e=2\) and \(e=4\) we have \(n^2_\xi =1, n^4_\xi =-1\) and \(\xi ^2_j = 1, \xi ^4_j = -1, j=1,\ldots ,N+1\), it holds that

$$\begin{aligned} \underline{\underline{B_\xi }}\underline{g}^\xi= & {} \sum _{e=2,4}\sum _{j=1}^{N+1}n^e_\xi \omega _jg^\xi _h(\xi ^e_j,\eta ^e_j)\underline{L}(\xi ^e_j)\otimes \underline{L}(\eta ^e_j) \end{aligned}$$

With analogous arguments we obtain

$$\begin{aligned} \underline{\underline{B_\eta }}\underline{g}^\eta= & {} \left[ \sum _{i=1}^{N+1}\omega _i\underline{\epsilon }_i\otimes \underline{L}(\eta )\sum _{j=1}^{N+1}L_j(\eta ) g^\eta _{\nu (i,j)}\right] ^1_{-1}\\= & {} \sum _{e=1,3}\sum _{i=1}^{N+1}n^e_\eta \omega _ig^\eta _h(\xi ^e_i,\eta ^e_i)\underline{L}(\xi _i^e)\otimes \underline{L}(\eta ^e_i). \end{aligned}$$

Summing up \(\underline{\underline{B_\xi }}\underline{g}^\xi + \underline{\underline{B_\eta }}\underline{g}^\eta \) and considering that \(n^1_\xi =n^2_\eta =n^3_\xi =n^4_\eta =0\) we obtain the assertion.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ortleb, S. A Kinetic Energy Preserving DG Scheme Based on Gauss–Legendre Points. J Sci Comput 71, 1135–1168 (2017). https://doi.org/10.1007/s10915-016-0334-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-016-0334-2

Keywords

Mathematics Subject Classification

Navigation