Skip to main content
Log in

An analysis of discontinuous Galerkin methods for the compressible Euler equations: entropy and \(L_2\) stability

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

The objective of this article is to characterize the entropy and \(L_2\) stability of several representative discontinuous Galerkin (DG) methods for solving the compressible Euler equations. Towards this end, three DG methods are constructed: one DG method with entropy variables as its unknowns, and two DG methods with conservative variables as their unknowns. These methods are employed in order to discretize the compressible Euler equations in space. Thereafter, the resulting semi-discrete formulations are analyzed, and the entropy and \(L_2\) stability characteristics are evaluated. It is shown that the semi-discrete formulation of the DG method with entropy variables is entropy and \(L_2\) stable. Furthermore, it is shown that the semi-discrete formulations of the DG methods with conservative variables are only guaranteed to be entropy and \(L_2\) stable under the following assumptions: the entropy projection errors vanish, or the terms containing the entropy projection errors are non-positive. Thereafter, the semi-discrete formulation with entropy variables, and one of the semi-discrete formulations with conservative variables, are discretized in time with an ‘algebraically stable’ Runge–Kutta (RK) scheme. The resulting formulations are fully-discrete and can be immediately applied to practical problems. In this article, they are employed to simulate a vortex propagating for long distances. It is shown that temporal stability is maintained by the DG method with entropy variables, but the DG method with conservative variables exhibits instability.

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

Similar content being viewed by others

References

  1. Alexander, R.: Diagonally implicit Runge-Kutta methods for stiff ODE’s. SIAM J. Numer. Anal. 14(6), 1006–1021 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  2. Balay, S., Abhyankar, S., Adams, M., Brown, J., Brune, P., Buschelman, K., Dalcin, L.D., Eijkhout, V., Gropp, W., Kaushik, D.: PETSc users manual revision 3.8. Tech. rep., Argonne National Lab.(ANL), Argonne, IL (United States) (2017)

  3. Balsara, D.S., Shu, C.W.: Monotonicity preserving weighted essentially non-oscillatory schemes with increasingly high order of accuracy. J. Comput. Phys. 160(2), 405–452 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  4. Barth, T., Charrier, P., Mansour, N.N.: Energy stable flux formulas for the discontinuous Galerkin discretization of first order nonlinear conservation laws. Tech. rep, NASA (2001)

    Google Scholar 

  5. Barth, T.J.: Numerical methods for gasdynamic systems on unstructured meshes. In: An Introduction to Recent Developments in Theory and Numerics for Conservation Laws, pp. 195–285. Springer, Berlin Heidelberg (1999)

  6. Barth, T.J.: On discontinuous Galerkin approximations of Boltzmann moment systems with Levermore closure. Comput. Methods Appl. Mech. Eng. 195(25), 3311–3330 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  7. Boyd, S., Vandenberghe, L.: Convex Optimization. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  8. Burrage, K.: Efficiently implementable algebraically stable Runge-Kutta methods. SIAM J. Numer. Anal. 19(2), 245–258 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  9. Castonguay, P., Vincent, P., Jameson, A.: Application of high-order energy stable flux reconstruction schemes to the Euler equations. In: 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition (2011)

  10. Chan, J., Demkowicz, L., Moser, R.: A DPG method for steady viscous compressible flow. Comput. Fluids 98, 69–90 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  11. Chan, J., Demkowicz, L., Moser, R., Roberts, N.: A class of discontinuous Petrov–Galerkin methods. Part V: Solution of 1D Burgers and Navier–Stokes equations. ICES Report 29 (2010)

  12. Ciarlet, P.G., Raviart, P.A.: General Lagrange and Hermite interpolation in Rn with applications to finite element methods. Arch. Ration. Mech. Anal. 46(3), 177–199 (1972)

    Article  MATH  Google Scholar 

  13. Cockburn, B.: Discontinuous Galerkin methods. ZAMM—J. Appl. Math. Mech. 83(11), 731–754 (2003)

  14. Cockburn, B., Hou, S., Shu, C.W.: The Runge–Kutta local projection discontinuous Galerkin finite element method for conservation laws. IV. The multidimensional case. Math. Comput. 54(190), 545–581 (1990)

    MathSciNet  MATH  Google Scholar 

  15. Cockburn, B., Karniadakis, G.E., Shu, C.W.: The development of discontinuous Galerkin methods. In: Discontinuous Galerkin Methods, pp. 3–50. Springer (2000)

  16. Cockburn, B., Lin, S.Y., Shu, C.W.: TVB Runge–Kutta local projection discontinuous Galerkin finite element method for conservation laws III: one-dimensional systems. J. Comput. Phys. 84(1), 90–113 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  17. Cockburn, B., Shu, C.W.: TVB Runge–Kutta local projection discontinuous Galerkin finite element method for conservation laws II. General framework. Math. Comput. 52(186), 411–435 (1989)

    MathSciNet  MATH  Google Scholar 

  18. Cockburn, B., Shu, C.W.: The Runge–Kutta local projection \(p^{1}\)-discontinuous Galerkin finite element method for scalar conservation laws. RAIRO-Modélisation mathématique et analyse numérique 25(3), 337–361 (1991)

    MathSciNet  MATH  Google Scholar 

  19. Cockburn, B., Shu, C.W.: The local discontinuous Galerkin method for time-dependent convection-diffusion systems. SIAM J. Numer. Anal. 35(6), 2440–2463 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  20. Cockburn, B., Shu, C.W.: The Runge–Kutta discontinuous Galerkin method for conservation laws V: multidimensional systems. J. Comput. Phys. 141(2), 199–224 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  21. Dafermos, C.: Hyperbolic Conservation Laws in Continuum Physics. Springer, Berlin (2005)

    Book  MATH  Google Scholar 

  22. Demkowicz, L., Gopalakrishnan, J.: A class of discontinuous Petrov–Galerkin methods. Part I: the transport equation. Comput. Methods Appl. Mech. Eng. 199(23), 1558–1572 (2010)

  23. Demkowicz, L., Gopalakrishnan, J.: A class of discontinuous Petrov–Galerkin methods. Part II: optimal test functions. Numer. Methods Partial Differ. Equ. 27(1), 70–105 (2011)

  24. Demkowicz, L., Gopalakrishnan, J., Niemi, A.H.: A class of discontinuous Petrov–Galerkin methods. Part III: adaptivity. Appl. Numer. Math. 62(4), 396–427 (2012)

    MATH  Google Scholar 

  25. Di Pietro, D.A., Ern, A.: Mathematical Aspects of Discontinuous Galerkin Methods, vol. 69. Springer, Heidelberg (2011)

    MATH  Google Scholar 

  26. Dutt, P.: Stable boundary conditions and difference schemes for Navier–Stokes equations. SIAM J. Numer. Anal. 25(2), 245–267 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  27. Ellis, T., Demkowicz, L., Chan, J.: Locally conservative discontinuous Petrov–Galerkin finite elements for fluid problems. Comput. Math. Appl. 68(11), 1530–1549 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  28. Fjordholm, U.S., Mishra, S., Tadmor, E.: Arbitrarily high-order accurate entropy stable essentially nonoscillatory schemes for systems of conservation laws. SIAM J. Numer. Anal. 50(2), 544–573 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  29. Fjordholm, U.S., Mishra, S., Tadmor, E.: ENO reconstruction and ENO interpolation are stable. Found. Comput. Math. 13(2), 139–159 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  30. Galbraith, M.C., Allmaras, S., Darmofal, D.L.: A verification driven process for rapid development of CFD software. In: 53rd AIAA Aerospace Sciences Meeting (2015)

  31. Galbraith, M.C., Allmaras, S.R., Darmofal, D.L.: SANS RANS solutions for 3D benchmark configurations. In: 2018 AIAA Aerospace Sciences Meeting, Kissimmee, Florida (2018)

  32. Godunov, S.K.: An interesting class of quasilinear systems. In: Dokl. Akad. Nauk SSSR, pp. 521–523 (1961)

  33. Harten, A.: On the symmetric form of systems of conservation laws with entropy. J. Comput. Phys. 49(1), 151–164 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  34. Harten, A.: ENO schemes with subcell resolution. J. Comput. Phys. 83(1), 148–184 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  35. Harten, A., Engquist, B., Osher, S., Chakravarthy, S.R.: Uniformly high order accurate essentially non-oscillatory schemes: III. In: Upwind and High-Resolution Schemes, pp. 218–290. Springer (1987)

  36. Harten, A., Osher, S.: Uniformly high-order accurate nonoscillatory schemes: I. SIAM J. Numer. Anal. 24(2), 279–309 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  37. Harten, A., Osher, S., Engquist, B., Chakravarthy, S.R.: Some results on uniformly high-order accurate essentially nonoscillatory schemes. Appl. Numer. Math. 2(3–5), 347–377 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  38. Hesthaven, J.S., Warburton, T.: Nodal Discontinuous Galerkin Methods: Algorithms, Analysis, and Applications. Springer, New York (2007)

    MATH  Google Scholar 

  39. Hiltebrand, A., Mishra, S.: Entropy stable shock capturing space–time discontinuous Galerkin schemes for systems of conservation laws. Numerische Mathematik 126(1), 103–151 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  40. Hou, S., Liu, X.D.: Solutions of multi-dimensional hyperbolic systems of conservation laws by square entropy condition satisfying discontinuous Galerkin method. J. Sci. Comput. 31(1–2), 127–151 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  41. Hu, C., Shu, C.W.: Weighted essentially non-oscillatory schemes on triangular meshes. J. Comput. Phys. 150(1), 97–127 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  42. Hughes, T.J.R., Franca, L.P., Mallet, M.: A new finite element formulation for computational fluid dynamics: I. Symmetric forms of the compressible Euler and Navier–Stokes equations and the second law of thermodynamics. Comput. Methods Appl. Mech. Eng. 54(2), 223–234 (1986)

  43. Hughes, T.J.R., Mallet, M.: A new finite element formulation for computational fluid dynamics: III. The generalized streamline operator for multidimensional advective-diffusive systems. Comput. Methods Appl. Mech. Eng. 58(3), 305–328 (1986)

  44. Hughes, T.J.R., Mallet, M.: A new finite element formulation for computational fluid dynamics: IV. A discontinuity-capturing operator for multidimensional advective-diffusive systems. Comput. Methods Appl. Mech. Eng. 58(3), 329–336 (1986)

  45. Hughes, T.J.R., Mallet, M., Akira, M.: A new finite element formulation for computational fluid dynamics: II. Beyond SUPG. Comput. Methods Appl. Mech. Eng. 54(3), 341–355 (1986)

  46. Jiang, G.S., Shu, C.W.: On a cell entropy inequality for discontinuous Galerkin methods. Math. Comput. 62(206), 531–538 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  47. Kirby, R.M., Karniadakis, G.E.: De-aliasing on non-uniform grids: algorithms and applications. J. Comput. Phys. 191(1), 249–264 (2003)

    Article  MATH  Google Scholar 

  48. Lax, P.D.: Hyperbolic Systems of Conservation Laws and the Mathematical Theory of Shock Waves, vol. 11. SIAM, Philadelphia (1973)

    Book  MATH  Google Scholar 

  49. Lefloch, P.G., Mercier, J.M., Rohde, C.: Fully discrete, entropy conservative schemes of arbitrary order. SIAM J. Numer. Anal. 40(5), 1968–1992 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  50. Liu, X.D., Osher, S., Chan, T.: Weighted essentially non-oscillatory schemes. J. Comput. Phys. 115(1), 200–212 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  51. Mock, M.S.: Systems of conservation laws of mixed type. J. Differ. Equ. 37(1), 70–88 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  52. Nesterov, Y.: Introductory Lectures on Convex Optimization: A Basic Course. Springer, New York (2004)

    Book  MATH  Google Scholar 

  53. Pugh, C.C.: Real Mathematical Analysis. Springer, New York (2002)

    Book  MATH  Google Scholar 

  54. Quarteroni, A., Sacco, R., Saleri, F.: Numerical Mathematics, vol. 37. Springer, New York (2010)

    MATH  Google Scholar 

  55. Saad, Y., Schultz, M.H.: GMRES: a generalized minimal residual algorithm for solving nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 7(3), 856–869 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  56. Shakib, F., Hughes, T.J.R., Johan, Z.: A new finite element formulation for computational fluid dynamics: X. The compressible Euler and Navier–Stokes equations. In: Computer Methods in Applied Mechanics and Engineering, pp. 141–219 (1991)

  57. Shu, C.W.: Essentially non-oscillatory and weighted essentially non-oscillatory schemes for hyperbolic conservation laws. In: Advanced Numerical Approximation of Nonlinear Hyperbolic Equations, pp. 325–432. Springer (1998)

  58. Shu, C.W., Osher, S.: Efficient implementation of essentially non-oscillatory shock-capturing schemes. J. Comput. Phys. 77(2), 439–471 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  59. Shu, C.W., Osher, S.: Efficient implementation of essentially non-oscillatory shock-capturing schemes: II. In: Upwind and High-Resolution Schemes, pp. 328–374. Springer (1989)

  60. Spiegel, S.C., Huynh, H.T., DeBonis, J.R.: A survey of the isentropic Euler vortex problem using high-order methods. In: 22nd AIAA Computational Fluid Dynamics Conference (2015)

  61. Strang, G.: Introduction to Linear Algebra. Wellesley-Cambridge Press Wellesley, Wellesley (2016)

    MATH  Google Scholar 

  62. Svärd, M.: Weak solutions and convergent numerical schemes of modified compressible Navier-Stokes equations. J. Comput. Phys. 288, 19–51 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  63. Svärd, M., Özcan, H.: Entropy-stable schemes for the Euler equations with far-field and wall boundary conditions. J. Sci. Comput. 58(1), 61–89 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  64. Tadmor, E.: The numerical viscosity of entropy stable schemes for systems of conservation laws. I. Math. Comput. 49(179), 91–103 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  65. Tadmor, E.: Entropy stability theory for difference approximations of nonlinear conservation laws and related time-dependent problems. Acta Numerica 12, 451–512 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  66. Toro, E.F.: Riemann Solvers and Numerical Methods for Fluid Dynamics: A Practical Introduction. Springer, Berlin (2013)

    Google Scholar 

  67. Wang, Z.J.: Adaptive High-Order Methods in Computational Fluid Dynamics, vol. 2. World Scientific, Singapore (2011)

    Book  MATH  Google Scholar 

  68. Wang, Z.J., Liu, Y., May, G., Jameson, A.: Spectral difference method for unstructured grids II: extension to the Euler equations. J. Sci. Comput. 32(1), 45–71 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  69. Williams, D.: An entropy stable, hybridizable discontinuous Galerkin method for the compressible Navier-Stokes equations. Math. Comput. 87(309), 95–121 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  70. Zitelli, J., Muga, I., Demkowicz, L., Gopalakrishnan, J., Pardo, D., Calo, V.M.: A class of discontinuous Petrov–Galerkin methods. Part IV: the optimal test norm and time-harmonic wave propagation in 1D. J. Comput. Phys. 230(7), 2406–2432 (2011)

Download references

Acknowledgements

The author would like to thank Dr. Siddhartha Mishra (ETH Zürich), Dr. Sigal Gottlieb (University of Massachusetts, Dartmouth), Dr. Dmitry Kamenetskiy (Boeing Research & Technology), Dr. Krzysztof Fidkowski (University of Michigan), Dr. Jesse Chan (Rice University), and Dr. Jason Hicken (Rensselaer Polytechnic Institute) for their participation in conversations that helped shape this work. The author would also like to thank Dr. David Darmofal, Dr. Marshall Galbraith, and Dr. Steven Allmaras (Massachusetts Institute of Technology) for developing the Solution Adaptive Numerical Simulator (SANS) code [30, 31] that was utilized for the numerical experiments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David M. Williams.

Additional information

Publisher's Note

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

Appendices

Appendix A: Notational conventions

The principal results in this paper are expressed via a mixture of index notation and vector notation. This combination of notation is best explained with an example. Consider the term \( \left( \varvec{w}^{h}_{,\varvec{x}_i} \right) ^T \varvec{f}^{i} \left( \varvec{u}^h \right) \). This can be expanded as follows when \(d =2\)

$$\begin{aligned} \left( \varvec{w}^{h}_{,\varvec{x}_i} \right) ^T \varvec{f}^{i} \left( \varvec{u}^h \right) = \left( \frac{\partial \varvec{w}^{h}}{\partial \varvec{x}_1} \right) ^T \varvec{f}^{1} \left( \varvec{u}^h \right) + \left( \frac{\partial \varvec{w}^{h}}{\partial \varvec{x}_2} \right) ^T \varvec{f}^{2} \left( \varvec{u}^h \right) , \end{aligned}$$

where the repeated index i facilitates the standard Einstein summation over d dimensions, and the transpose facilitates the standard dot product between a pair of m-vectors [for instance \(\varvec{w}^{h}_{,\varvec{x}_1}\) and \(\varvec{f}^{1} \left( \varvec{u}^h \right) \)].

Appendix B: Supporting lemmas

Lemma B.1

Suppose that the pressure p and density \(\rho \) are bounded in the following fashion

$$\begin{aligned} 0< p \le M, \quad 0 < \rho \le N, \end{aligned}$$
(B.1)

where M and N are positive numbers. Then, the symmetric Jacobian matrices \({\widetilde{A}}_0\) and \({\widetilde{A}}_{0}^{-1}\ (\)defined in Eq. (3.7)) are guaranteed to be positive definite (PD).

Proof

In what follows, we present the proof for the 2D case. The proof for the 3D case and for higher dimensions, is very similar. In 2D, one may begin by examining the matrix \({\widetilde{A}}_0\), which can be expressed as follows

$$\begin{aligned} {\widetilde{A}}_0 = \begin{bmatrix} \rho&\quad \rho u&\quad \rho v&\quad \rho H-p \\ \rho u&\quad \rho u^2 + p&\quad \rho u v&\quad \rho u H \\ \rho v&\quad \rho u v&\quad \rho v^2 + p&\quad \rho v H \\ \rho H - p&\quad \rho u H&\quad \rho v H&\quad \rho H^2 - \gamma e p \end{bmatrix}, \end{aligned}$$

where

$$\begin{aligned} H = \gamma e + \frac{1}{2} \left( u^2 + v^2 \right) . \end{aligned}$$

The characteristic polynomial of \({\widetilde{A}}_0\) takes the following form

$$\begin{aligned} \frac{1}{4} \left( \lambda - p \right) \left( a \lambda ^3 + b \lambda ^2 + c \lambda +d \right) = 0, \end{aligned}$$
(B.2)

where

$$\begin{aligned} a&= 4, \nonumber \\ b&= - \rho \left( \left( 2 + u^2 + v^2 \right) ^2 + 4 e \left( \gamma \left( 1 + u^2 + v^2 + e \right) - 1 \right) \right) , \nonumber \\ c&= \rho p \left( \left( 2 + u^2 + v^2 \right) ^2 + 4e \left( \gamma e + 1 \right) \right) , \nonumber \\ d&= -4 \rho e p^2. \end{aligned}$$
(B.3)

It immediately follows from Eq. (B.2), that the pressure p is an eigenvalue. Evidently, this eigenvalue is positive since Eq. (B.1) holds.

One may determine the signs of the remaining eigenvalues of \({\widetilde{A}}_0\) by first observing that the matrix is symmetric and will have all real entries if Eq. (B.1) is satisfied. As a result, the eigenvalues of \({\widetilde{A}}_0\) will be real, i.e. the roots of its characteristic polynomial will be real. In accordance with Decartes’ rule of signs, a cubic polynomial with real roots, and with coefficients \(a > 0\), \(b < 0\), \(c > 0\), and \(d < 0\) is guaranteed to have three positive roots. By inspection of Eq. (B.3), it immediately follows that the coefficients a, b, c, and d will have the required signs if Eq. (B.1) holds, if \(\gamma > 1\) (which is true for virtually all gases), and if one notes that \(e = p /(\rho \left( \gamma -1 \right) ) > 0\). This completes the proof that \({\widetilde{A}}_0\) has positive eigenvalues.

The eigenvalues of \({\widetilde{A}}_{0}^{-1}\) are obtained by taking the reciprocals of the eigenvalues of \({\widetilde{A}}_{0}\). The eigenvalues are positive as long as Eq. (B.1) holds, as under these circumstances, the eigenvalues of \({\widetilde{A}}_0\) remain positive, and the reciprocals of positive numbers are positive. This completes the proof that \({\widetilde{A}}_{0}^{-1}\) has positive eigenvalues. \(\square \)

Lemma B.2

The jump in the entropy functional \(F^i \left( \varvec{v}\right) \) across an interface is governed by the following

$$\begin{aligned}&\left[ \!\left[ F^i \left( \varvec{u}^h \right) \right] \! \right] ^{-}_{+} + \left( \left\{ \!\left\{ \varvec{v}\left( \varvec{u}^h \right) \right\} \! \right\} ^{-}_{+}\right) ^T \left[ \!\left[ \varvec{f}^i \left( \varvec{u}^h \right) \right] \! \right] ^{+}_{-} \nonumber \\&\quad = \frac{1}{2} \int _{0}^{1} \left( 1 - \theta \right) \left( \left[ \!\left[ \varvec{v}\left( \varvec{u}^h \right) \right] \! \right] ^{+}_{-}\right) ^T \left( {\widetilde{A}}_i \left( \overline{\overline{\varvec{v}}} \left( \theta \right) \right) - {\widetilde{A}}_i \left( \overline{\varvec{v}} \left( \theta \right) \right) \right) \left[ \!\left[ \varvec{v}\left( \varvec{u}^h \right) \right] \! \right] ^{+}_{-}d \theta .\nonumber \\ \end{aligned}$$
(B.4)

Proof

The proof is virtually identical to the proofs of similar statements in [69] and [5]. It is repeated here for the sake of completeness.

Recall that \({\mathcal {F}}^{i} = {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}\right) \right) \). On utilizing this fact in conjunction with Taylor’s theorem, one can obtain the following

$$\begin{aligned}&{\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{+} \right) \right) - {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}_{-}^h \right) \right) - {\mathcal {F}}^{i}_{, \varvec{v}} \left( \varvec{v}\left( \varvec{u}^h_{+} \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}_{-}^h \right) \right) \nonumber \\&\quad + \int _{0}^{1} \left( 1- \theta \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}_{-}^h \right) \right) ^T {\mathcal {F}}^{i}_{,\varvec{v}, \varvec{v}} \left( \overline{\varvec{v}} \left( \theta \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}_{-}^h \right) \right) d \theta = 0,\nonumber \\ \end{aligned}$$
(B.5)
$$\begin{aligned}&{\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{+} \right) \right) - {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{-} \right) \right) - {\mathcal {F}}^{i}_{, \varvec{v}} \left( \varvec{v}\left( \varvec{u}^h_{-} \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) -\varvec{v}\left( \varvec{u}^h_{-} \right) \right) \nonumber \\&\quad - \int _{0}^{1} \left( 1- \theta \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) ^T {\mathcal {F}}^{i}_{,\varvec{v}, \varvec{v}} \left( \overline{\overline{\varvec{v}}} \left( \theta \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) d \theta = 0. \nonumber \\ \end{aligned}$$
(B.6)

Upon multiplying Eqs. (B.5) and (B.6) by (1/2) and summing the results, one obtains

$$\begin{aligned}&{\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{+} \right) \right) - {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{-} \right) \right) - \frac{1}{2} \left( \varvec{f}^i \left( \varvec{u}^h_{+} \right) + \varvec{f}^i \left( \varvec{u}^h_{-} \right) \right) ^T \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) \nonumber \\&\quad = \frac{1}{2} \int _{0}^{1} \left( 1- \theta \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) ^T \left( {\widetilde{A}}_i \left( \overline{\overline{\varvec{v}}} \left( \theta \right) \right) - {\widetilde{A}}_i \left( \overline{\varvec{v}} \left( \theta \right) \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) d \theta ,\nonumber \\ \end{aligned}$$
(B.7)

where the fact that \(\left( \varvec{f}^i \right) ^T = {\mathcal {F}}^{i}_{, \varvec{v}}\) and \({\widetilde{A}}_{i} = {\mathcal {F}}^{i}_{,\varvec{v}, \varvec{v}}\) has been used (cf. Eqs. (3.3) and (3.7)). Setting Eq. (B.7) aside for the moment, consider the following jump identity that derives from Eq. (3.5)

$$\begin{aligned}&F^i \left( \varvec{u}^h_{+} \right) - F^i \left( \varvec{u}^h_{-} \right) + {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{+} \right) \right) - {\mathcal {F}}^i \left( \varvec{v}\left( \varvec{u}^h_{-} \right) \right) \nonumber \\&\quad = \frac{1}{2} \left( \varvec{v}\left( \varvec{u}^h_{+} \right) + \varvec{v}\left( \varvec{u}^h_{-} \right) \right) ^T \left( \varvec{f}^i \left( \varvec{u}^h_{+} \right) - \varvec{f}^i \left( \varvec{u}^h_{-} \right) \right) \nonumber \\&\qquad + \frac{1}{2} \left( \varvec{f}^i \left( \varvec{u}^h_{+} \right) + \varvec{f}^i \left( \varvec{u}^h_{-} \right) \right) ^T \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) . \end{aligned}$$
(B.8)

Substituting Eq. (B.7) into Eq. (B.8) yields

$$\begin{aligned}&F^i \left( \varvec{u}^h_{-} \right) - F^i \left( \varvec{u}^h_{+} \right) + \frac{1}{2} \left( \varvec{v}\left( \varvec{u}^h_{-} \right) + \varvec{v}\left( \varvec{u}^h_{+} \right) \right) ^T \left( \varvec{f}^i \left( \varvec{u}^h_{+} \right) - \varvec{f}^i \left( \varvec{u}^h_{-} \right) \right) \nonumber \\&\quad = \frac{1}{2} \int _{0}^{1} \left( 1- \theta \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) ^T \left( {\widetilde{A}}_i \left( \overline{\overline{\varvec{v}}} \left( \theta \right) \right) - {\widetilde{A}}_i \left( \overline{\varvec{v}} \left( \theta \right) \right) \right) \left( \varvec{v}\left( \varvec{u}^h_{+} \right) - \varvec{v}\left( \varvec{u}^h_{-} \right) \right) d \theta .\nonumber \\ \end{aligned}$$
(B.9)

Upon substituting the spatial jump [Eq. (5.1)] and average [Eq. (5.2)] operators into Eq. (B.9) one obtains Eq. (B.4). \(\square \)

Lemma B.3

Suppose that the initial condition \(\varvec{u}\left( t_0 \right) \in \varvec{L}_2 \left( \varOmega \right) \), and the pressure p and density \(\rho \) are positive and bounded for all convex combinations of states \(\varvec{u}\left( t_0 \right) \) and \(\varvec{u}^{*}\) defined by \(\widehat{\widehat{\varvec{u}}} \left( \theta \right) = \varvec{u}^{*} + \theta \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) \), where \(0 \le \theta \le 1\). Under these circumstances, the following statements hold: (i) the \(L_2\) norm of \(\varvec{u}\left( t_0 \right) - \varvec{u}^{*}\) is well-defined; (ii) the eigenvalues of \({\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \) are real and positive; and (iii) the broken integral of \(H \left( \varvec{u}\left( t_0 \right) , \varvec{u}^{*} \right) \) over the domain \({\mathcal {T}}_h\) is bounded in the following fashion

$$\begin{aligned} \sum _{T_k \in {\mathcal {T}}_h} \int _{T_k} H \left( \varvec{u}\left( t_0 \right) , \varvec{u}^{*} \right) dx \le \frac{\lambda _{1,_{{\mathcal {T}}_h}}}{2} \Vert \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \Vert _{L_2, {\mathcal {T}}_h}^2, \end{aligned}$$
(B.10)

where \(\lambda _{1,_{{\mathcal {T}}_h}}\) is the maximum eigenvalue of \({\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \) over all elements in the domain.

Note: throughout this lemma we implicitly assume that \(\varvec{u}\left( t_0 \right) = \varvec{u}\left( \varvec{v}^h \left( t_0 \right) \right) \).

Proof

The proof of part (i) follows from inspection, and the proof of part (ii) is given in Lemma B.1. Therefore, it remains to prove part (iii). One may begin the proof by utilizing Taylor’s Theorem in order to express \(U \left( \varvec{u}\left( t_0 \right) \right) \) in terms of \(U \left( \varvec{u}^{*} \right) \) as follows

$$\begin{aligned} U \left( \varvec{u}\left( t_0 \right) \right)&= U \left( \varvec{u}^{*} \right) + \left( U_{,\varvec{u}} \left( \varvec{u}^{*} \right) \right) ^T \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) \nonumber \\&\quad + \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) ^T U_{,\varvec{u},\varvec{u}} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) d \theta . \end{aligned}$$
(B.11)

Upon substituting the expression for \(H \left( \varvec{u}\left( t_0 \right) , \varvec{u}^{*} \right) \) (as given by definition 7.3) into Eq. (B.11), one obtains

$$\begin{aligned} H \left( \varvec{u}\left( t_0 \right) , \varvec{u}^{*} \right)&= \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) ^T U_{,\varvec{u},\varvec{u}} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) d \theta \nonumber \\&= \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) ^T {\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) d \theta , \end{aligned}$$
(B.12)

where the definition of the inverse Jacobian matrix \({\widetilde{A}}_0^{-1} = U_{,\varvec{u}, \varvec{u}}\) has been used in the last line (see Eqs. (3.4) and (3.7)). The matrix \({\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \) has a maximum eigenvalue denoted by \(\lambda _1 \left( \theta \right) = \lambda _1 \left( {\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \right) \). Therefore, the last three terms that appear under the integral sign in Eq. (B.12) can be bounded from above as follows

$$\begin{aligned}&\left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) ^T {\widetilde{A}}_0^{-1} \left( \widehat{\widehat{\varvec{u}}} \left( \theta \right) \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) \nonumber \\&\quad \le \lambda _1 \left( \theta \right) \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) ^T \left( \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right) , \nonumber \\&\quad \le \lambda _1 \left( \theta \right) \left\| \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right\| _{L_2}^2, \end{aligned}$$
(B.13)

where several standard results from Linear Algebra (cf. for instance [61]) have been used. On substituting Eq. (B.13) into Eq. (B.12) and integrating both sides over the entire domain, one obtains

$$\begin{aligned} \sum _{T_k \in {\mathcal {T}}_h} \int _{T_k} H \left( \varvec{u}\left( t_0 \right) , \varvec{u}^{*} \right) dx&\le \sum _{T_k \in {\mathcal {T}}_h} \int _{T_k} \int _{0}^{1} \left( \left( 1 - \theta \right) \lambda _1 \left( \theta \right) \left\| \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right\| _{L_2}^2 \right) d \theta \, dx, \nonumber \\&\le \sum _{T_k \in {\mathcal {T}}_h} \left( \int _{T_k} \lambda _1 \left\| \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right\| _{L_2}^2 dx \right) \int _{0}^{1} \left( 1 - \theta \right) d \theta , \nonumber \\&\le \frac{1}{2} \lambda _{1,_{{\mathcal {T}}_h}} \left\| \varvec{u}\left( t_0 \right) - \varvec{u}^{*} \right\| _{L_2, {\mathcal {T}}_h}^2, \end{aligned}$$
(B.14)

where we have defined \(\lambda _1\) to be the maximum of \(\lambda _1 \left( \theta \right) \) over all \(0 \le \theta \le 1\), and \(\lambda _{1,_{{\mathcal {T}}_h}}\) to be the maximum of \(\lambda _1\) over all elements in the domain. Finally, the proof is completed by comparing Eq. (B.14) with Eq. (B.10). \(\square \)

Lemma B.4

Suppose that the initial condition \(\varvec{v}^h \left( t_0 \right) \in \varvec{L}_2 \left( \varOmega \right) \), the initial condition \(\varvec{u}\left( t_0 \right) \in \varvec{L}_1 \left( \varOmega \right) \), each component of \(\varvec{v}\left( \varvec{u}^{*} \right) \) is bounded, and the pressure p and density \(\rho \) are positive and bounded for all convex combinations of states \(\varvec{v}^{h} \left( t_0 \right) \) and \(\varvec{v}\left( \varvec{u}^{*} \right) \) defined by \(\widehat{\widehat{\varvec{v}}} \left( \theta \right) = \varvec{v}^{h} \left( t_0 \right) + \theta \left( \varvec{v}\left( \varvec{u}^{*} \right) - \varvec{v}^{h} \left( t_0 \right) \right) \), where \(0 \le \theta \le 1\). Under these circumstances, the following statements hold: (i) the \(L_2\) norm of \(\varvec{v}^h \left( t_0 \right) - \varvec{v}\left( \varvec{u}^{*} \right) \) is well-defined; (ii) the eigenvalues of \({\widetilde{A}}_0 \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \) are real and positive; and (iii) the broken integral of \({\mathcal {H}} \left( \varvec{v}\left( \varvec{u}^* \right) , \varvec{v}^h \left( t_0 \right) \right) \) over the domain \({\mathcal {T}}_h\) is bounded in the following fashion

$$\begin{aligned} \sum _{T_k \in {\mathcal {T}}_h} \int _{T_k} {\mathcal {H}} \left( \varvec{v}\left( \varvec{u}^* \right) , \varvec{v}^h \left( t_0 \right) \right) dx \le \frac{\lambda _{1,_{{\mathcal {T}}_h}}}{2} \Vert \varvec{v}^h \left( t_0 \right) - \varvec{v}\left( \varvec{u}^{*} \right) \Vert _{L_2, {\mathcal {T}}_h}^2, \end{aligned}$$
(B.15)

where \(\lambda _{1,_{{\mathcal {T}}_h}}\) is the maximum eigenvalue of \({\widetilde{A}}_0 \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \), over all elements in the domain.

Note: throughout this lemma we implicitly assume that \(\varvec{u}\left( t_0 \right) = \varvec{u}\left( \varvec{v}^h \left( t_0 \right) \right) \).

Proof

The proof of this lemma is very similar to the proof of Lemma B.3. In particular, only part (iii) requires proof. One may begin the proof by using Taylor’s Theorem in order to express \({\mathcal {U}} \left( \varvec{v}\left( \varvec{u}^* \right) \right) \) in terms of \({\mathcal {U}} \left( \varvec{v}^h \left( t_0 \right) \right) \) as follows

$$\begin{aligned} {\mathcal {U}} \left( \varvec{v}\left( \varvec{u}^* \right) \right)&= {\mathcal {U}} \left( \varvec{v}^h \left( t_0 \right) \right) + \left( {\mathcal {U}}_{,\varvec{v}} \left( \varvec{v}^h \left( t_0 \right) \right) \right) ^T \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) \nonumber \\&\quad + \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) ^T {\mathcal {U}}_{,\varvec{v},\varvec{v}} \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) d \theta . \end{aligned}$$
(B.16)

On substituting the expression for \({\mathcal {H}} \left( \varvec{v}\left( \varvec{u}^* \right) , \varvec{v}^h \left( t_0 \right) \right) \) (as given by Definition 7.6) into Eq. (B.16), one obtains

$$\begin{aligned}&{\mathcal {H}} \left( \varvec{v}\left( \varvec{u}^* \right) , \varvec{v}^h \left( t_0 \right) \right) \nonumber \\&\quad = \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) ^T {\mathcal {U}}_{,\varvec{v},\varvec{v}} \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) d \theta \nonumber \\&\quad = \int _{0}^{1} \left( 1 - \theta \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) ^T {\widetilde{A}}_0 \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \left( \varvec{v}\left( \varvec{u}^* \right) - \varvec{v}^h \left( t_0 \right) \right) d \theta , \end{aligned}$$
(B.17)

where the definition of the Jacobian matrix \({\widetilde{A}}_0 = {\mathcal {U}}_{,\varvec{v}, \varvec{v}}\) has been used in the last line [see Eqs. (3.2) and (3.7)]. The remainder of the proof requires a simple eigenvalue analysis in order to find the maximum eigenvalue \(\lambda _1 \left( \theta \right) = \lambda _1 \left( {\widetilde{A}}_0 \left( \widehat{\widehat{\varvec{v}}} \left( \theta \right) \right) \right) \), and thereafter, the construction of an upper bound with the maximum eigenvalue and the \(L_2\) norm

$$\begin{aligned} \Vert \varvec{v}^h \left( t_0 \right) - \varvec{v}\left( \varvec{u}^{*} \right) \Vert _{L_2}, \end{aligned}$$

in accordance with the proof of Lemma B.3. \(\square \)

Lemma B.5

The functional \(\; U \left( \varvec{u}\right) \) is strongly convex on the set \({{\mathscr {C}} \subset \varvec{\mathrm {dom}} \; U \left( \varvec{u}\right) }\), under the assumptions that \({\mathscr {C}}\) is a convex set, and that the eigenvalues of the matrix \({\widetilde{A}}_{0}^{-1} \left( \varvec{u}\right) \) are bounded away from zero for all \(\varvec{u}\in {\mathscr {C}}\), where \({\widetilde{A}}_{0}^{-1}\) is defined in Eq. (3.7).

Proof

In order to prove the strong convexity of U, one must obtain a particular inequality governing the Hessian, \(U_{, \varvec{u}, \varvec{u}}\), for all \(\varvec{u}\in {\mathscr {C}}\). Towards this end, one may first note that the following identity holds in accordance with Eqs. (3.4) and (3.7)

$$\begin{aligned} U_{, \varvec{u}, \varvec{u}} = \varvec{v}_{,\varvec{u}} = {\widetilde{A}}_{0}^{-1}. \end{aligned}$$

Here, the inverse Jacobian (\({\widetilde{A}}_{0}^{-1}\)) is SPD because the Jacobian itself (\({\widetilde{A}}_{0}\)) is SPD under the assumptions of Lemma B.1. As a result, the eigenvalues of \(U_{, \varvec{u}, \varvec{u}} = {\widetilde{A}}_{0}^{-1}\) are greater than zero, and are guaranteed to satisfy the following inequality

$$\begin{aligned} \lambda _{1} \ge \cdots \ge \lambda _{m} > 0. \end{aligned}$$

Furthermore, it is common practice (cf. [6, 62]) to assume that the minimum eigenvalue of \({\widetilde{A}}_{0}^{-1}\), for all \(\varvec{u}\in {\mathscr {C}}\), is bounded away from zero

$$\begin{aligned} \lambda _{m} \ge 2 C, \end{aligned}$$

where \(C > 0\) is a constant independent of \(\varvec{u}\). Setting this result aside for the moment, one may utilize the Hessian \(U_{, \varvec{u}, \varvec{u}}\) in order to construct a quadratic form

$$\begin{aligned} \varvec{u}_{b}^T \left( U_{, \varvec{u}, \varvec{u}} \left( \varvec{u}_{a} \right) \right) \varvec{u}_{b}, \end{aligned}$$
(B.18)

where \(\varvec{u}_{a}\) and \(\varvec{u}_{b}\) are arbitrary values of \(\varvec{u}\in {\mathscr {C}}\). In accordance with several standard results from Linear Algebra (cf. for instance [61]), the quadratic form in Eq. (B.18) satisfies the following inequalities

$$\begin{aligned} \varvec{u}_{b}^T \left( U_{, \varvec{u}, \varvec{u}} \left( \varvec{u}_{a} \right) \right) \varvec{u}_{b}&\ge \lambda _m \varvec{u}_{b}^T \varvec{u}_{b}, \nonumber \\&= \lambda _m \left\| \varvec{u}_{b} \right\| _{L_2}^2, \nonumber \\&\ge 2 C \left\| \varvec{u}_{b} \right\| _{L_2}^2. \end{aligned}$$
(B.19)

From Eq. (B.19) and the strong convexity criterion in [7], p. 459, it immediately follows that \(U \left( \varvec{u}\right) \) is strongly convex. \(\square \)

Lemma B.6

The functional \(\; {\mathcal {U}} \left( \varvec{v}\right) \) is strongly convex on the set \({{\mathscr {C}} \subset \varvec{\mathrm {dom}} \; {\mathcal {U}} \left( \varvec{v}\right) }\), under the assumptions that \({\mathscr {C}}\) is a convex set, and that the eigenvalues of the matrix \({\widetilde{A}}_{0} \left( \varvec{v}\right) \) are bounded away from zero for all \(\varvec{v}\in {\mathscr {C}}\), where \({\widetilde{A}}_0\) is defined in Eq. (3.7).

Proof

The proof of this lemma is very similar to the proof of Lemma B.5. The proof begins with the observation that \({\widetilde{A}}_{0}\) is SPD under the assumptions of Lemma B.1, and that

$$\begin{aligned} {\mathcal {U}}_{, \varvec{v}, \varvec{v}} = \varvec{u}_{,\varvec{v}} = {\widetilde{A}}_{0}, \end{aligned}$$

in accordance with Eqs. (3.2) and (3.7). The rest of the proof requires a simple eigenvalue analysis, and the construction of inequalities involving the quadratic form,

$$\begin{aligned} \varvec{v}_{b}^T \left( {\mathcal {U}}_{, \varvec{v}, \varvec{v}} \left( \varvec{v}_{a} \right) \right) \varvec{v}_{b}, \end{aligned}$$

in accordance with the proof of Lemma B.5. \(\square \)

Appendix C: Algebraically stable SDIRK methods

In this section, the Butcher tables for some well-known algebraically stable SDIRK methods are presented.

The 1-stage 1st-order ‘Backward Euler’ method has the following Butcher table

$$\begin{aligned} \begin{array}{c|c} 1 &{} 1 \\ \hline &{} 1 \end{array} \end{aligned}$$

The 2-stage 2nd-order method due to [1] and [8] has the following Butcher table

$$\begin{aligned} \begin{array}{c|ccc} \zeta &{} \zeta &{} 0\\ c_2 &{} a_{21} &{} \zeta \\ \hline &{}b_1 &{} b_2 \end{array} \end{aligned}$$

where

$$\begin{aligned} \zeta&= \frac{\left( 2 \pm \sqrt{2} \right) }{2}, \\ a_{21}&= 1 - 2 \zeta , \qquad b_1 = \frac{1}{2}, \qquad b_2 = \frac{1}{2}, \qquad c_2 = 1 - \zeta . \end{aligned}$$

The 3-stage 3rd-order method due to [1] and [8] has the following Butcher table

$$\begin{aligned} \begin{array}{c|ccc} \zeta &{} \zeta &{} 0 &{} 0\\ c_2 &{} a_{21} &{} \zeta &{} 0 \\ c_3 &{} a_{31} &{} a_{32} &{} \zeta \\ \hline &{}b_1 &{} b_2 &{} b_3 \end{array} \end{aligned}$$

where

$$\begin{aligned} \zeta&= 0.43586652150845899941601945, \\ | d_1 |&\ge 1.774294247072785073096332, \qquad d_2 = \left( \frac{1}{2} - \zeta - d_1 \right) \frac{b_2}{b_3}, \\ a_{21}&= \frac{1}{2} - \zeta +d_1, \qquad a_{31} = 1 - 2 \zeta - d_2, \qquad a_{32} = d_2, \\ b_1&= \frac{d_1 \left( 1 -2 \zeta \right) + \frac{1}{6}}{\left( 2 \zeta - 1\right) \left( 2 \zeta - 1 -2 d_1 \right) }, \qquad b_2 = \frac{\zeta ^2 - \zeta + \frac{1}{6}}{\left( \zeta - \frac{1}{2} \right) ^2 - d_1^2}, \\&\qquad b_3 = \frac{d_1 \left( 2 \zeta - 1 \right) + \frac{1}{6}}{\left( 2 \zeta -1 \right) \left( 2 \zeta -1 + 2 d_1 \right) }, \\ c_2&= \frac{1}{2} + d_1, \qquad c_3 = 1 - \zeta . \end{aligned}$$

Note that the expression for \(d_{2}\) given here is different than the corresponding expression in [8]. The definition given here is the correct definition, as the original definition contains a typographical error that prevents the method from being algebraically stable.

Finally, the 4-stage 4th-order method due to [1, 8] has the following Butcher table

$$\begin{aligned} \begin{array}{c|cccc} \zeta &{} \zeta &{} 0 &{} 0 &{} 0\\ c_2 &{} a_{21} &{} \zeta &{} 0 &{} 0\\ c_3 &{} a_{31} &{} a_{32} &{} \zeta &{} 0 \\ c_4 &{} a_{41} &{} a_{42} &{} a_{43} &{} \zeta \\ \hline &{}b_1 &{} b_2 &{} b_3 &{} b_4 \end{array} \end{aligned}$$

where

$$\begin{aligned} \zeta&= 0.5728160624821348554080014, \\ d_1&= \zeta ^2 - \frac{\zeta }{2} + \frac{1}{12}, \qquad d_2 = -\frac{\zeta \left( \zeta -\frac{1}{3} \right) }{\left( 2 \zeta ^2 - \zeta + \frac{1}{6} \right) }, \qquad d_3 =\zeta ^3 -\frac{3 \zeta ^2}{2} + \frac{\zeta }{2} - \frac{1}{24}, \\ a_{21}&= \frac{\zeta \left( \frac{1}{3} - \zeta \right) }{2 d_1}, \qquad a_{31} = \frac{1}{2} - \zeta - \frac{d_3}{d_1} + \frac{8 d_3 \left( \zeta - \frac{1}{4} \right) }{\zeta \left( \zeta - \frac{1}{3} \right) } , \qquad a_{32} = -\frac{8 d_3 \left( \zeta - \frac{1}{4} \right) }{\zeta \left( \zeta - \frac{1}{3} \right) }, \\ a_{41}&= 1 - 2 \zeta - a_{42} - \frac{d_1}{\left( 4 \zeta -1 \right) \left( b_2 - \frac{1}{2} \right) }, \qquad a_{42} = \frac{ \left( \zeta ^2 - \zeta + \frac{1}{6} \right) /2 + d_3/d_2}{d_2 \left( \frac{1}{2} - b_2 \right) }, \\ a_{43}&= \frac{d_1}{\left( 4 \zeta -1 \right) \left( b_2 - \frac{1}{2} \right) }, \qquad b_1 = \frac{1}{2} - b_2, \qquad b_2 = \frac{d_1^2}{2 \zeta \left( \zeta - \frac{1}{3} \right) \left( \zeta - \frac{1}{4} \right) }, \qquad b_3 = b_2, \\ b_4&= \frac{1}{2} - b_2, \qquad c_2 = \frac{ \zeta \left( \zeta - \frac{1}{2} \right) ^2}{d_1}, \qquad c_3 = \frac{1}{2} - \frac{d_3}{d_1}, \qquad c_4 = 1- \zeta . \end{aligned}$$

Note that the expression for \(a_{31}\) given here is different than the corresponding expression in [8]. The definition given here is the correct definition, as the original definition contains a typographical error that prevents the method from being algebraically stable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Williams, D.M. An analysis of discontinuous Galerkin methods for the compressible Euler equations: entropy and \(L_2\) stability. Numer. Math. 141, 1079–1120 (2019). https://doi.org/10.1007/s00211-019-01027-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-019-01027-9

Mathematics Subject Classification

Navigation