Skip to main content
Log in

The BR1 Scheme is Stable for the Compressible Navier–Stokes Equations

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

A Correction to this article was published on 18 June 2018

This article has been updated

Abstract

In this work we prove that the original (Bassi and Rebay in J Comput Phys 131:267–279, 1997) scheme (BR1) for the discretization of second order viscous terms within the discontinuous Galerkin collocation spectral element method (DGSEM) with Gauss Lobatto nodes is stable. More precisely, we prove in the first part that the BR1 scheme preserves energy stability of the skew-symmetric advection term DGSEM discretization for the linearized compressible Navier–Stokes equations (NSE). In the second part, we prove that the BR1 scheme preserves the entropy stability of the recently developed entropy stable compressible Euler DGSEM discretization of Carpenter et al. (SIAM J Sci Comput 36:B835–B867, 2014) for the non-linear compressible NSE, provided that the auxiliary gradient equations use the entropy variables. Both parts are presented for fully three-dimensional, unstructured curvilinear hexahedral grids. Although the focus of this work is on the BR1 scheme, we show that the proof naturally includes the Local DG scheme of Cockburn and Shu.

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.

Similar content being viewed by others

Change history

  • 18 June 2018

    An open-source code that implements the entropy stable discontinuous Galerkin scheme with Legendere–Gauss–Lobatto collocation (DGSEM) on curved unstructured hexahedral grids for compressible Navier–Stokes equations (NSE) is available at github.com/project-fluxo.

References

  1. Abarbanel, S., Gottlieb, D.: Optimal time splitting for two-dimensional and 3-dimensional Navier–Stokes equations with mixed derivatives. J. Comput. Phys. 41(1), 1–33 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  2. Arnold, D.N., Brezzi, F., Cockburn, B., Marini, L.D.: Unified analysis of discontinuous Galerkin methods for elliptic problems. SIAM J. Numer. Anal. 39(5), 1749–1779 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  3. Arnold, D.N., Brezzi, F., Cockburn, B., Marini, D.: Discontinuous Galerkin methods for elliptic problems. In: Cockburn, B., Karniadakis, G.E., Shu, C.W. (eds.) Discontinuous Galerkin Methods. Lecture Notes in Computational Science and Engineering, vol. 11. Springer, Berlin, Heidelberg (2000)

  4. Bassi, F., Rebay, S.: A high order accurate discontinuous finite element method for the numerical solution of the compressible Navier–Stokes equations. J. Comput. Phys. 131, 267–279 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  5. Beck, A.D., Bolemann, T., Flad, D., Frank, H., Gassner, Gregor J., Hindenlang, Florian, Munz, Claus-Dieter: High-order discontinuous galerkin spectral element methods for transitional and turbulent flow simulations. Int. J. Numer. Methods Fluids 76(8), 522–548 (2014)

    Article  MathSciNet  Google Scholar 

  6. Canuto, C., Hussaini, M.Y., Quarteroni, A., Zang, T.A.: Spectral Methods: Fundamentals in Single Domains. Springer, Berlin (2006)

    MATH  Google Scholar 

  7. Carpenter, M., Fisher, T., Nielsen, E., Frankel, S.: Entropy stable spectral collocation schemes for the Navier–Stokes equations: discontinuous interfaces. SIAM J. Sci. Comput. 36(5), B835–B867 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chandrashekar, Praveen: Kinetic energy preserving and entropy stable finite volume schemes for compressible Euler and Navier–Stokes equations. Commun. Comput. Phys. 14, 1252–1286 (2013). 11

    Article  MathSciNet  MATH  Google Scholar 

  9. Cockburn, B., Shu, C.W.: The Runge–Kutta local projection \({P}^1\)-discontinuous Galerkin method for scalar conservation laws. Rairo Math. Model. Numer. Anal. Model. Math. et Anal. Numer. 25, 337–361 (1991)

    Article  MATH  Google Scholar 

  10. Cockburn, B., Shu, C.W.: Runge–kutta discontinuous Galerkin methods for convection-dominated problems. J. Sci. Comput. 16(3), 173–261 (2001)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  12. Cockburn, B., Hou, S., Shu, Chi-Wang: 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 

  13. 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 

  14. Fisher, T., 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 

  15. Frank, Hannes M., Munz, Claus-Dieter: Direct aeroacoustic simulation of acoustic feedback phenomena on a side-view mirror. J. Sound Vib. 371, 132–149 (2016)

    Article  Google Scholar 

  16. Gassner, G.: A skew-symmetric discontinuous Galerkin spectral element discretization and its relation to SBP-SAT finite difference methods. SIAM J. Sci. Comput. 35(3), A1233–A1253 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  17. Gassner, G., 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, 2560–2579 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  18. Gassner, Gregor J.: A skew-symmetric discontinuous Galerkin spectral element discretization and its relation to SBP-SAT finite difference methods. SIAM J. Sci. Comput. 35(3), A1233–A1253 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  19. Gassner, G.J., Beck, Andrea D.: On the accuracy of high-order discretizations for underresolved turbulence simulations. Theor. Comput. Fluid Dyn. 27(3–4), 221–237 (2013)

    Article  Google Scholar 

  20. Gassner, G.J., Winters, A.R., Kopriva, David A.: Split form nodal discontinuous Galerkin schemes with summation-by-parts property for the compressible Euler equations. J. Comput. Phys. 327, 39–66 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  21. Hindenlang, F., Gassner, G.J., Altmann, C., Beck, A., Staudenmaier, Marc, Munz, Claus-Dieter: Explicit discontinuous Galerkin methods for unsteady problems. Comput. Fluids 61, 86–93 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  22. Ismail, F., Roe, P.L.: Affordable, entropy-consistent Euler flux functions II: entropy production at shocks. J. Comput. Phys. 228(15), 5410–5436 (2009)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  24. Kirby, R.M., Karniadakis, G.E.: Selecting the numerical flux in discontinuous galerkin methods for diffusion problems. J. Sci. Comput. 22(1), 385–411 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  25. Kopriva, D.A.: Metric identities and the discontinuous spectral element method on curvilinear meshes. J. Sci. Comput. 26(3), 301–327 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kopriva, D.A., Gassner, G.: On the quadrature and weak form choices in collocation type discontinuous Galerkin spectral element methods. J. Sci. Comput. 44(2), 136–155 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  27. Kopriva, D.A., Gassner, G.J.: An energy stable discontinuous Galerkin spectral element discretization for variable coefficient advection problems. SIAM J. Sci. Comput. 36(4), A2076–A2099 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  28. Kopriva, David A.: Implementing Spectral Methods for Partial Differential Equations. Scientific Computation. Springer, Berlin (2009)

    Book  MATH  Google Scholar 

  29. Kopriva, D.A.: A polynomial spectral calculus for analysis of DG spectral element methods. arXiv:1704.00709 (2017)

  30. Kopriva, D.A., Gassner, G.J.: An energy stable discontinuous Galerkin spectral element discretization for variable coefficient advection problems. SIAM J. Sci. Comput. 34(4), A2076–A2099 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  31. Kopriva, D.A., Gassner, G.J.: Geometry effects in nodal discontinuous Galerkin methods on curved elements that are provably stable. Appl. Math. Comput. 272(Part 2), 274–290 (2016)

    MathSciNet  Google Scholar 

  32. Mengaldo, G., De Grazia, D., Moxey, D., Vincent, P.E., Sherwin, S.J.: Dealiasing techniques for high-order spectral element methods on regular and irregular grids. J. Comput. Phys. 299, 56–81 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  33. Merriam, Marshal L.: An entropy-based approach to nonlinear stability. NASA Tech. Memo. 101086(64), 1–154 (1989)

    MathSciNet  Google Scholar 

  34. Nelson, D.A., Kopriva, D.A., Jacobs, G.B.: High-order curved boundary representation with discontinuous-Galerkin methods. Theor. Comput. Fluid Dyn. 30(4), 363–385 (2016)

    Article  Google Scholar 

  35. Nordström, J.: A roadmap to well posed and stable problems in computational physics. J. Sci. Comput. 71(1), 365–385 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  36. Nordström, J., Svard, M.: Well-posed boundary conditions for the Navier–Stokes equations. SIAM J. Numer. Anal. 43(3), 1231–1255 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  37. Reed, W.H., Hill, T.R.: Triangular mesh methods for the neutron transport equation. Technical Report LA-UR-73-479, Los Alamos National Laboratory (1973)

  38. Tadmor, Eitan: Entropy stability theory for difference approximations of nonlinear conservation laws and related time-dependent problems. Acta Numer. 12, 451–512 (2003). 5

    Article  MathSciNet  MATH  Google Scholar 

  39. Tadmor, E., Zhong, W.: Entropy stable approximations of Navier–Stokes equations with no artificial numerical viscosity. J. Hyperb. Differ. Equ. 3(3), 529–559 (2006)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This work was supported by a grant from the Simons Foundation (#426393, David Kopriva). G. G has been supported by the European Research Council (ERC) under the European Union’s Eights Framework Program Horizon 2020 with the research project Extreme, ERC Grant Agreement No. 714487.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gregor J. Gassner.

Appendices

An Entropy Conserving Euler Flux

We explicitly present the Kinetic Energy Preserving and Entropy Conservative (KEPEC) numerical flux function for the compressible Euler equations. The KEPEC flux was first derived by Chandrashekar [8]

$$\begin{aligned} \mathbf {F}_1^{*,\mathrm{ec}} = {\small \begin{bmatrix} \rho ^{\ln }\{\!\{v_1\}\!\} \\ \rho ^{\ln }\{\!\{v_1\}\!\}^2 + \{\!\{p\}\!\} \\ \rho ^{\ln }\{\!\{v_1\}\!\}\{\!\{v_2\}\!\} \\ \rho ^{\ln }\{\!\{v_1\}\!\}\{\!\{v_3\}\!\} \\ \frac{p^{\ln }\{\!\{v_1\}\!\}}{(\gamma -1)} +\{\!\{p\}\!\}\{\!\{v_1\}\!\} + \frac{1}{2}\rho ^{\ln }\{\!\{v_1\}\!\}\overline{||\mathbf {v} ||^2}\end{bmatrix}}, \end{aligned}$$
(A.1)

where

$$\begin{aligned} \{\!\{p\}\!\}= & {} \frac{\{\!\{\rho \}\!\}}{2\{\!\{\beta \}\!\}},\qquad {p}^{\ln } = \frac{{\rho }^{\ln }}{2{\beta }^{\ln }},\nonumber \\ \overline{||\mathbf {v} ||^2}= & {} 2\left( \{\!\{v_1\}\!\}^2 + \{\!\{v_2\}\!\}^2 + \{\!\{v_3\}\!\}^2\right) \nonumber \\&-\left( \{\!\{v_1^2\}\!\} + \{\!\{v_2^2\}\!\} + \{\!\{v_3^2\}\!\}\right) , \end{aligned}$$
(A.2)

and \((\cdot )^{\ln }\) is the logarithmic mean

$$\begin{aligned} (\cdot )^{\mathrm{ln}} = \frac{\llbracket \cdot \rrbracket }{\llbracket \ln (\cdot )\rrbracket }\,. \end{aligned}$$
(A.3)

A numerically stable method to compute the logarithmic mean when \((\cdot )_{\mathrm{R}} \approx (\cdot )_{\mathrm{L}}\) is given in [22, App. B].

It is also possible to add additional dissipation to the entropy conservative scheme and create an entropy stable (ES) approximation. One example is with a matrix dissipation entropy stable flux for the compressible Euler equations of the form

$$\begin{aligned} \mathbf {F}_1^{*,\mathrm{ec}} = \mathbf {F}_1^{*,\mathrm{ec}} - \frac{1}{2}{{\hat{\varvec{\mathcal {R}}_1}}}|\varvec{\hat{\varLambda }}|\hat{\varvec{\mathcal {T}}}{{\hat{\varvec{\mathcal {R}}_1}}}^T\llbracket \mathbf {w}\rrbracket . \end{aligned}$$
(A.4)

The average components of the dissipation term are given by

$$\begin{aligned} \begin{aligned} {{\hat{\varvec{\mathcal {R}}_1}}}&= {\begin{bmatrix} 1&1&0&0&1 \\ \{\!\{v_1\}\!\} - \bar{a}&\{\!\{v_1\}\!\}&0&0&\{\!\{v_1\}\!\} + \bar{a} \\ \{\!\{v_2\}\!\}&\{\!\{v_2\}\!\}&1&0&\{\!\{v_2\}\!\} \\ \{\!\{v_3\}\!\}&\{\!\{v_3\}\!\}&0&1&\{\!\{v_3\}\!\} \\ \bar{h} - \{\!\{v_1\}\!\}\bar{a}&\frac{1}{2}\overline{\Vert \mathbf v \Vert ^2}&\{\!\{v_2\}\!\}&\{\!\{v_3\}\!\}&\bar{h} + \{\!\{v_1\}\!\}\bar{a} \\ \end{bmatrix}},\\ \hat{\varvec{\varLambda }}&= \text {diag}\left( \{\!\{v_1\}\!\} - \bar{a},\{\!\{v_1\}\!\},\{\!\{v_1\}\!\},\{\!\{v_1\}\!\},\{\!\{v_1\}\!\} + \bar{a}\right) ,\\ \hat{\varvec{\mathcal {T}}}&= \text {diag}\left( \frac{\rho ^{\ln }}{2\gamma },\frac{\rho ^{\ln }(\gamma -1)}{\gamma },\{\!\{p\}\!\},\{\!\{p\}\!\},\frac{\rho ^{\ln }}{2\gamma }\right) , \end{aligned} \end{aligned}$$
(A.5)

where

$$\begin{aligned} \bar{a} = \sqrt{\frac{\gamma \{\!\{p\}\!\}}{\rho ^{\ln }}},\qquad \bar{h} = \frac{\gamma }{2\beta ^{\ln }(\gamma -1)} + \frac{1}{2}\overline{||\mathbf {v} ||^2}\,. \end{aligned}$$
(A.6)

We note that the selection of the discrete dissipation operator (A.5) creates a scheme that is able to exactly resolve stationary contact discontinuities. The proof of this property follows the same structure as that presented by Chandrashekar [8].

For completeness, we provide the entropy conserving (and stable) fluxes for the other Cartesian directions

$$\begin{aligned} \mathbf {F}_2^{*,{\mathrm{ec}}}= & {} {\begin{bmatrix} \rho ^{\ln }\{\!\{v_2\}\!\} \\ \rho ^{\ln }\{\!\{v_1\}\!\}\{\!\{v_2\}\!\} \\ \rho ^{\ln }\{\!\{v_2\}\!\}^2 + \{\!\{p\}\!\} \\ \rho ^{\ln }\{\!\{v_2\}\!\}\{\!\{v_3\}\!\} \\ \frac{p^{\ln }\{\!\{v_2\}\!\}}{(\gamma -1)} +\{\!\{p\}\!\}\{\!\{v_2\}\!\} + \frac{1}{2}\rho ^{\ln }\{\!\{v_2\}\!\}\overline{||\mathbf {v} ||^2}\end{bmatrix}},\nonumber \\ \mathbf {F}_3^{*,\mathrm{ec}}= & {} {\begin{bmatrix} \rho ^{\ln }\{\!\{v_3\}\!\} \\ \rho ^{\ln }\{\!\{v_1\}\!\}\{\!\{v_3\}\!\} \\ \rho ^{\ln }\{\!\{v_2\}\!\}\{\!\{v_3\}\!\} \\ \rho ^{\ln }\{\!\{v_3\}\!\}^2 + \{\!\{p\}\!\} \\ \frac{p^{\ln }\{\!\{v_3\}\!\}}{(\gamma -1)} +\{\!\{p\}\!\}\{\!\{v_3\}\!\} + \frac{1}{2}\rho ^{\ln }\{\!\{v_3\}\!\}\overline{||\mathbf {v} ||^2}\end{bmatrix}}. \end{aligned}$$
(A.7)

The matrix dissipation term remains similar where the matrix of right eigenvectors are given by

$$\begin{aligned} {{\hat{\varvec{\mathcal {R}}_2}}}= & {} {\begin{bmatrix} 1&\quad 0&\quad 1&\quad 0&\quad 1 \\ \{\!\{v_1\}\!\}&\quad 1&\quad \{\!\{v_1\}\!\}&\quad 0&\quad \{\!\{v_1\}\!\} \\ \{\!\{v_2\}\!\} - \bar{a}&\quad 0&\quad \{\!\{v_2\}\!\}&\quad 0&\quad \{\!\{v_2\}\!\} + \bar{a} \\ \{\!\{v_3\}\!\}&\quad 0&\quad \{\!\{v_3\}\!\}&\quad 1&\quad \{\!\{v_3\}\!\} \\ \bar{h} - \{\!\{v_2\}\!\}\bar{a}&\quad \{\!\{v_1\}\!\}&\quad \frac{1}{2}\overline{\Vert \mathbf v \Vert ^2}&\quad \{\!\{v_3\}\!\}&\quad \bar{h} + \{\!\{v_2\}\!\}\bar{a} \\ \end{bmatrix}} ,\nonumber \\ {{\hat{\varvec{\mathcal {R}}_3}}}= & {} {\begin{bmatrix} 1&\quad 0&\quad 0&\quad 1&\quad 1 \\ \{\!\{v_1\}\!\}&\quad 1&\quad 0&\quad \{\!\{v_1\}\!\}&\quad \{\!\{v_1\}\!\} \\ \{\!\{v_2\}\!\}&\quad 0&\quad 1&\quad \{\!\{v_2\}\!\}&\quad \{\!\{v_2\}\!\} \\ \{\!\{v_3\}\!\} - \bar{a}&\quad 0&\quad 0&\quad \{\!\{v_3\}\!\}&\quad \{\!\{v_3\}\!\} + \bar{a} \\ \bar{h} - \{\!\{v_3\}\!\}\bar{a}&\quad \{\!\{v_1\}\!\}&\quad \{\!\{v_2\}\!\}&\quad \frac{1}{2}\overline{\Vert \mathbf v \Vert ^2}&\quad \bar{h} + \{\!\{v_3\}\!\}\bar{a} \\ \end{bmatrix}} \end{aligned}$$
(A.8)

respectively. The diagonal matrix of eigenvalues uses the appropriate value of the velocity depending on the spatial direction.

To create the contravariant entropy conservative fluxes we incorperate the average of the metric terms, e.g. at the \((\xi =1)\) element face

$$\begin{aligned} \tilde{\mathbf {F}}^{*,\mathrm{ec}} = \{\!\{Ja_1^1\}\!\}\mathbf {F}_1^{*,\mathrm{ec}} + \{\!\{Ja_2^1\}\!\}\mathbf {F}_2^{*,\mathrm{ec}} + \{\!\{Ja_3^1\}\!\}\mathbf {F}_3^{*,\mathrm{ec}}. \end{aligned}$$
(A.9)

The dissipation terms remain unchanged for the contravariant entropy stable approximations.

Proofs of Entropy Conservation for Non-Linear Advection Terms

1.1 Proof of Entropy Conservation for the One-Dimensional Volume Integral

We first show the property (4.21), rewritten here as

$$\begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N = \left. F^{\,\epsilon }\right| _{-1}^1. \end{aligned}$$
(B.1)

For convenience we introduce the entropy potential

$$\begin{aligned} \varPsi ^f_i \equiv W_i\,F_i - F^{\,\epsilon }_i \end{aligned}$$
(B.2)

at each GL node \(i=0,...,N\) to rewrite the entropy-conservation condition (4.18) on the two-point flux

$$\begin{aligned} F^{\mathrm {ec}}\,\llbracket W\rrbracket - \llbracket F\,W\rrbracket +\llbracket F^{\,\epsilon }\rrbracket = 0 \end{aligned}$$
(B.3)

as

$$\begin{aligned} F^{\mathrm {ec}}\,\llbracket W\rrbracket = \llbracket \varPsi ^f\rrbracket . \end{aligned}$$
(B.4)

We then explicitly write the volume integral in (B.1) as a sum

$$\begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N = \sum \limits _{i=0}^N \omega _i W_i\,2\,\sum _{m=0}^N D_{im} F^{\mathrm {ec}}(U_i,U_m). \end{aligned}$$
(B.5)

Let us now introduce the summation-by-parts matrix \(Q = MD\) with entries \(Q_{im}\equiv \omega _{i}D_{im}\), which has the property

$$\begin{aligned} Q+Q^T=B\,, \end{aligned}$$
(B.6)

where \(B=\mathrm {diag}([-1,0,...,0,1])\) is the boundary evaluation matrix. Because the derivative of a constant is exact, the Q matrix satisfies

$$\begin{aligned} \sum \limits _{m=0}^N Q_{im} V_i = 0 \end{aligned}$$
(B.7)

for all \(V_{i}\).

In terms of Q,

$$\begin{aligned} \begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N = \sum \limits _{i=0}^N W_i\,2\,\sum _{m=0}^N Q_{im} F^{\mathrm {ec}}(U_i,U_m)\,. \end{aligned} \end{aligned}$$
(B.8)

But from (B.6), \(2\,Q_{im} = Q_{im} - Q_{mi} + B_{im}\) so

$$\begin{aligned} \begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N&= \sum \limits _{i=0}^N \sum _{m=0}^N W_i\,(Q_{im} - Q_{mi} + B_{im}) F^{\mathrm {ec}}(U_i,U_m)\\ {}&= \sum \limits _{i,m} {{W_i}{Q_{im}}{F^{\mathrm {ec}}}\left( {{U_i},{U_m}} \right) } + \sum \limits _{i,m} {{W_i}{Q_{mi}}{F^{\mathrm {ec}}}\left( {{U_i},{U_m}} \right) } \\&\quad + \sum \limits _{i,m} {{B_{im}}{W_i}{F^{\mathrm {ec}}}\left( {{U_i},{U_m}} \right) }\,. \end{aligned} \end{aligned}$$
(B.9)

We now re-index the second sum, \(i\leftrightarrow m\), use the fact that \(F^{\mathrm {ec}}\) is symmetric in its arguments, and recombine the sums to get

$$\begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N = \sum \limits _{i=0}^N \sum _{m=0}^N \left\{ (W_i - W_m)\,F^{\mathrm {ec}}(U_i,U_m)\,Q_{im} + B_{im}\,W_i\,F^{\mathrm {ec}}(U_i,U_m)\right\} \,.\nonumber \\ \end{aligned}$$
(B.10)

The definition of the entropy potential (B.4) says that we can write the argument

$$\begin{aligned} (W_i - W_m)\,F^{\mathrm {ec}}(U_i,U_m) = \varPsi ^f_i - \varPsi ^f_m\,. \end{aligned}$$
(B.11)

We further note that \(B_{im}\) only has entries for \(i=m=0\) and \(i=m=N\) so with B.2 and the consistency condition on the entropy conserving flux \(W_i\,F^{\mathrm {ec}}(U_i,U_i) = W_i\,F(U_i) = \varPsi ^f_i+F^{\,\epsilon }_i\),

$$\begin{aligned} B_{im}\,W_i\,F^{\mathrm {ec}}(U_i,U_m) = B_{im}\,\left( \varPsi ^f_i+F^{\,\epsilon }_i\right) \,. \end{aligned}$$
(B.12)

Inserting (B.11) and (B.12) into (B.10),

$$\begin{aligned} \begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N&=\sum \limits _{i=0}^N \sum _{m=0}^N \left( \varPsi ^f_i - \varPsi ^f_m\right) \,Q_{im} + B_{im}\,\left( \varPsi ^f_i+F^{\,\epsilon }_i\right) \\ {}&= \sum \limits _i {\varPsi _i^f\sum \limits _m {{Q_{im}}} } - \sum \limits _{i,m}^{} {\varPsi _m^f{Q_{im}}} + \sum \limits _{i,m} {{B_{im}}\varPsi _i^f} + \sum \limits _{i,m} {{B_{im}}F_i^{\,\epsilon }} \,. \end{aligned} \end{aligned}$$
(B.13)

By (B.7), the first term in the second line is zero. Next, with \(B_{im}=Q_{im}+Q_{mi}\), re-indexing \(i\leftrightarrow m\), and using (B.7),

$$\begin{aligned} \sum \limits _{i,m} {{B_{im}}\varPsi _i^f} = \sum \limits _{i,m} {{Q_{im}}\varPsi _i^f} + \sum \limits _{i,m} {{Q_{mi}}\varPsi _i^f} = \sum \limits _{i,m} {{Q_{im}}\varPsi _m^f}\,. \end{aligned}$$
(B.14)

Therefore the second and third sums in the second line of (B.13) cancel. What is left is what we set out to show, namely

$$\begin{aligned} \begin{aligned} \left\langle \mathbb {D}(F)^{\mathrm {ec}},W\right\rangle _N&= \sum \limits _{i=0}^N \sum _{m=0}^N B_{im}\,F^{\,\epsilon }_i = F^{\,\epsilon }_N - F^{\,\epsilon }_0 = \left. F^{\,\epsilon }\right| _{-1}^1\,. \end{aligned} \end{aligned}$$
(B.15)

Remark 4

The result (B.1) is a general one that depends only on the summation-by-parts property. It is therefore not specific to the discontinuous Galerkin approximation, per se.

1.2 Proof of Entropy Conservation at Interelement Interfaces for Curvilinear Elements

We now show the property (4.80),

$$\begin{aligned} \sum \limits _{\begin{array}{c} {\mathrm {interior}} \\ {\mathrm {faces}} \end{array}} \int \limits _{N} \left( \left( \tilde{\mathbf {F}}_n^{\mathrm {ec},*}\right) ^T\llbracket \mathbf {W}\rrbracket - \llbracket \left( \tilde{\mathbf {F}}_n\right) ^T\,\mathbf {W}\rrbracket + \llbracket \tilde{ F}_n^{\,\epsilon }\rrbracket \right) \,\,{\text {dS}} = 0\,. \end{aligned}$$
(B.16)

For the approximate 2D surface integral, we evaluate the integrand at \((N+1)^2\) GL quadrature points, see (2.40). Therefore we only need to prove that the integrand vanishes discretely at each interior face quadrature point. We assume that the following derivations are restricted to one interior face quadrature point, and skip quadrature point indices.

The general three-dimensional conditions on the Cartesian components of the two-point numerical flux are

$$\begin{aligned} \begin{aligned}&\left( \mathbf {F}^{\mathrm {ec},*} \right) ^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {F}^T\,\mathbf {W}\rrbracket +\llbracket F^{\,\epsilon }\rrbracket = 0,\\&\left( \mathbf {G}^{\mathrm {ec},*} \right) ^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {G}^T\,\mathbf {W}\rrbracket +\llbracket G^{\,\epsilon }\rrbracket = 0,\\&\left( \mathbf {H}^{\mathrm {ec},*} \right) ^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {H}^T\,\mathbf {W}\rrbracket +\llbracket H^{\,\epsilon }\rrbracket = 0, \end{aligned} \end{aligned}$$
(B.17)

where we use the slave–master jump definition from (3.76),

$$\begin{aligned} \llbracket \mathbf {W}\rrbracket = \mathbf {W}_{\mathrm {slave}} - \mathbf {W}_ {\mathrm {master}} , \end{aligned}$$
(B.18)

and each interior face has the master element side orientation, so that quadrature points of the slave and the master map on each other.

We make the assumption that the mesh is watertight, i.e. that the normal vector and the surface element are continuous across element interfaces. For a conforming hexahedral mesh, the condition holds discretely at the surface quadrature points if we ensure the discrete metric identities (2.32) and if the unit outward facing normal vector and surface element on the element side are constructed from the element metrics by

$$\begin{aligned} \hat{s}= \left| \sum \limits _{l=1}^3\left( J\vec {a}^{\,l}\right) \hat{n}^l\right| \,,\quad \vec {n} = \frac{1}{\hat{s}} \sum \limits _{l=1}^3\left( J\vec {a}^{\,d}\right) \hat{n}^l. \end{aligned}$$
(B.19)

The continuity of the surface metric allows us to use only the metric of the master element side, so that we are able to move the metric into the jump.

We can write any normal contravariant surface flux using the Cartesian fluxes and the metric

$$\begin{aligned} \left( \tilde{\mathbf {F}}_n\right)= & {} \left( \hat{s}\vec {n}\right) \cdot {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }} =\hat{s}\left( \mathbf {F} n_1 +\mathbf {G} n_2 + \mathbf {H} n_3\right) = \sum \limits _{l=1}^3\hat{n}^l \left( \left( Ja_1^{\,l}\right) \mathbf {F} +\left( Ja_2^{\,l}\right) \mathbf {G} + \left( Ja_3^{\,l}\right) \mathbf {H} \right) \nonumber \\= & {} \left( \mathcal {M}^T {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }} \right) \cdot \hat{n}\,. \end{aligned}$$
(B.20)

We combine the three Cartesian equations (B.17) with the surface metric \(\hat{s}\vec {n}\), defined in (B.19), leading to

$$\begin{aligned} \begin{aligned}&\hat{s}\Bigg (n_1 \left[ (\mathbf {F}^{\mathrm {ec},*} )^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {F}^T\,\mathbf {W}\rrbracket +\llbracket F^{\,\epsilon }\rrbracket \right] \\&\quad +n_2 \left[ (\mathbf {G}^{\mathrm {ec},*} )^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {G}^T\,\mathbf {W}\rrbracket +\llbracket G^{\,\epsilon }\rrbracket \right] \\&\quad +n_3 \left[ (\mathbf {H}^{\mathrm {ec},*} )^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \mathbf {H}^T\,\mathbf {W}\rrbracket +\llbracket H^{\,\epsilon }\rrbracket \right] \Bigg )= 0\,. \end{aligned} \end{aligned}$$
(B.21)

If we define the contravariant numerical flux as

$$\begin{aligned} \tilde{\mathbf {F}}^{\mathrm {ec},*}_{n} \equiv \hat{s}\left( n_1\mathbf {F}^{\mathrm {ec},*} + n_2\mathbf {G}^{\mathrm {ec},*} + n_3\mathbf {H}^{\mathrm {ec},*} \right) \,, \end{aligned}$$
(B.22)

then

$$\begin{aligned} (\tilde{\mathbf {F}}^{\mathrm {ec},*}_{n})^T\,\llbracket \mathbf {W}\rrbracket +\left( \hat{s}\vec {n}\,\right) \cdot \left( - \llbracket {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^T\,\mathbf {W}\rrbracket + \llbracket \vec {F}^{\,\epsilon }\rrbracket \right) = 0\,. \end{aligned}$$
(B.23)

Finally, using the continuity of the surface metric, we move it inside the jump terms, yielding

$$\begin{aligned} \begin{aligned} \left( \mathbf {F}^{\mathrm {ec},*}_{n}\right) ^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \left( \left( \mathcal {M}^T{\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}\right) \cdot \hat{n}\right) ^T\,\mathbf {W}\rrbracket +\llbracket \left( \mathcal {M}^T\vec {F}^{\,\epsilon }\right) \cdot \hat{n}\rrbracket&\\ =\left( \tilde{\mathbf {F}}^{\mathrm {ec},*}_{n}\right) ^T\,\llbracket \mathbf {W}\rrbracket - \llbracket \left( \tilde{\mathbf {F}}_n\right) ^T\,\mathbf {W}\rrbracket +\llbracket \tilde{ F}_n^{\,\epsilon }\rrbracket&= 0 \,,\\ \end{aligned} \end{aligned}$$
(B.24)

proving that in (4.80), the integrand vanishes at each quadrature point of an interior face individually.

1.3 Proof of Entropy Conservation in 3D Curvilinear Coordinates

We show in this section that the property (4.75) holds, reproduced here for convenience as

$$\begin{aligned} \left\langle \vec {\mathbb {D}}({\mathop {\tilde{\mathbf {F}}}\limits ^{\leftrightarrow }})^{\mathrm {ec}},\mathbf {W}\right\rangle _N = \int \limits _{\partial E,N} \left( \vec {\tilde{F}}^{\,\epsilon }\cdot \hat{n}\right) \,{\text {dS}} \,, \end{aligned}$$
(B.25)

provided that the discrete metric identities (2.32) are satisfied.

Similar to what was done in “Appendix B.1”, we introduce three entropy potentials, one for each Cartesian space direction

$$\begin{aligned} \begin{aligned} \varPsi ^l_{ijk}&\equiv \left( \mathbf {F}^{l}_{ijk}\right) ^{T}\mathbf {W}_{ijk} - F^{l,{\,\epsilon }}_{ijk},\; l = 1,2,3, \end{aligned} \end{aligned}$$
(B.26)

for each GL node \(i,j,k=0,...,N\). We use them to rewrite the entropy-conservation condition on the two-point fluxes \(F^{l,\mathrm {ec},*}\), c.f. (B.17), into

$$\begin{aligned} \left( \mathbf {F}^{l,\mathrm {ec}}_{(i,m)jk}\right) ^T\,\llbracket \mathbf {W}\rrbracket _{(i,m)jk} = \llbracket \varPsi ^l\rrbracket _{(i,m)jk} ,\; l = 1,2,3, \end{aligned}$$
(B.27)

for \(i,j,k,m=0,...,N\).

To match terms, we expand all the terms of the volume integral approximation

$$\begin{aligned} \begin{aligned} \left\langle \vec {\mathbb {D}}({\mathop {\tilde{\mathbf {F}}}\limits ^{\leftrightarrow }})^{\mathrm {ec}},\mathbf {W}\right\rangle _N&\equiv \,\sum \limits _{i,j,k=0}^N\omega _{ijk}\,\mathbf {W}^T_{ijk}\Bigg [ 2\sum _{m=0}^N D_{im}\left( {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\right) \\&\quad +2\sum _{m=0}^N D_{jm}\left( {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{imk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,2}\right\} \!\!\right\} _{i(j,m)k}\right) \\&\quad +2\sum _{m=0}^N D_{km}\left( {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{ijm})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,3}\right\} \!\!\right\} _{ij(k,m)}\right) \Bigg ]\,, \end{aligned} \end{aligned}$$
(B.28)

(c.f. (4.69)) and of the surface integral approximation

$$\begin{aligned} \begin{aligned} \int \limits _{\partial E,N} \left( \vec {\tilde{F}}^{\,\epsilon }\cdot \hat{n}\right) \,\,{\text {dS}} =&\quad \sum \limits _{j,k=0}^N\omega _{jk}\left[ \left( \left( J\vec {a}^{\,1}\right) _{Njk}\cdot \vec {F}^{\,\epsilon }_{Njk}\right) -\left( \left( J\vec {a}^{\,1}\right) _{0jk}\cdot \vec {F}^{\,\epsilon }_{0jk}\right) \right] \\&+\sum \limits _{i,k=0}^N\omega _{ik}\left[ \left( \left( J\vec {a}^{\,2}\right) _{iNk}\cdot \vec {F}^{\,\epsilon }_{iNk}\right) -\left( \left( J\vec {a}^{\,2}\right) _{i0k}\cdot \vec {F}^{\,\epsilon }_{i0k}\right) \right] \\&+\sum \limits _{i,j=0}^N\omega _{ij}\left[ \left( \left( J\vec {a}^{\,3}\right) _{ijN}\cdot \vec {F}^{\,\epsilon }_{ijN}\right) - \left( \left( J\vec {a}^{\,3}\right) _{ij0}\cdot \vec {F}^{\,\epsilon }_{ij0}\right) \right] \,. \end{aligned} \end{aligned}$$
(B.29)

We then focus on the first (\(\xi \) direction) term of the volume integral approximation, which allows us to follow the one dimensional proof as closely as possible. The sum can be written in terms of \(Q_{im}=w_{i}D_{im}\),

$$\begin{aligned} \begin{aligned}&\sum \limits _{i,j,k=0}^N\omega _{ijk}\,\mathbf {W}^T_{ijk} 2\sum _{m=0}^N D_{im}\, {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} \\&\qquad \qquad =\sum \limits _{j,k=0}^N\omega _{jk}\sum \limits _{i=0}^N \mathbf {W}^T_{ijk}2 \sum _{m=0}^N \omega _iD_{im}\,{\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} \\&\qquad \qquad =\sum \limits _{j,k=0}^N\omega _{jk}\sum \limits _{i=0}^N \mathbf {W}^T_{ijk}\, \sum _{m=0}^N 2Q_{im}\,{\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\,. \end{aligned} \end{aligned}$$
(B.30)

Therefore, we can use the same steps as in one dimension: We use the summation-by-parts property \(2\,Q_{im}=Q_{im} - Q_{mi} + B_{im}\), a re-indexing of i and m to subsume the \(Q_{mi}\) term, and the facts that \(F^{\mathrm {ec}}(U_{ijk}, U_{mjk})\) and the jump operator of the metric term \(\left\{ \!\!\left\{ Ja^1_1\right\} \!\!\right\} _{(i,m)jk}\) are symmetric with respect to the index i and m to rewrite the \(\xi \) direction contribution to the volume integral approximation as

$$\begin{aligned} \begin{aligned}&\sum \limits _{i=0}^N\mathbf {W}^T_{ijk}\,2\sum _{m=0}^N Q_{im}\, {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} \\&\quad = \sum \limits _{i=0}^N \sum _{m=0}^N \mathbf {W}^T_{ijk}\,(Q_{im} - Q_{mi} + B_{im}) {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk},U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\\&\quad = \sum \limits _{i=0}^N \sum _{m=0}^N Q_{im}\left( \mathbf {W}_{ijk} - \mathbf {W}_{mjk}\right) ^T\, {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk},U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\, \\&\quad + B_{im}\,\mathbf {W}^T_{ijk}\,{\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk},U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}. \end{aligned} \end{aligned}$$
(B.31)

Next, we use the relations (B.27) of the Cartesian two-point entropy-conserving flux \(\mathbf {F}^{l,\mathrm {ec}}\) to note that

$$\begin{aligned} \left( \mathbf {W}_{ijk} - \mathbf {W}_{mjk}\right) ^T\,\mathbf {F}^{l,\mathrm {ec}}(U_{ijk},U_{mjk}) = \varPsi ^l_{ijk} - \varPsi ^l_{mjk}\,, \; l = 1,2,3. \end{aligned}$$
(B.32)

We further note that \(B_{im}\) only has entries for \(i=m=0\) and \(i=m=N\) (c.f. (B.12)) so

$$\begin{aligned} B_{im}\,\mathbf {W}^T_{ijk}\,\mathbf {F}^{l,\mathrm {ec}}(U_{ijk},U_{mjk}) = B_{im}\,\left( \varPsi ^l_{ijk}+ F^{l,{\,\epsilon }}_{ijk}\right) \,, \; l = 1,2,3. \end{aligned}$$
(B.33)

Finally, we can exploit the consistency of the two-point flux and (B.26) so that for \(i=m=0\) and \(i=m=N\),

$$\begin{aligned} \mathbf {W}^T_{ijk}\,\mathbf {F}^{l,\mathrm {ec}}(U_{ijk},U_{ijk}) = \mathbf {W}^T_{ijk}\,\mathbf {F}(U_{ijk}) = \varPsi ^l_{ijk}+ F^{l,{\,\epsilon }}_{ijk}\,, \; l = 1,2,3. \end{aligned}$$
(B.34)

Inserting relations (B.32)-(B.34) into the last line of (B.31) says that

$$\begin{aligned} \begin{aligned}&\sum \limits _{i=0}^N \mathbf {W}^T_{ijk}\,2\sum _{m=0}^N Q_{im}\,{\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\\&\quad = \sum \limits _{i=0}^N \sum _{m=0}^N Q_{im} \left( \vec {\varPsi }_{ijk} - \vec {\varPsi }_{mjk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} +B_{im}\,\left( \vec {\varPsi }_{ijk}+\vec {F}^{\,\epsilon }_{ijk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}. \end{aligned} \end{aligned}$$
(B.35)

Now, since the derivative of a constant is zero,

$$\begin{aligned} \begin{aligned} \sum \limits _{i,m = 0}^N {{Q_{im}}{\vec {\varPsi }_{ijk}}\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}}&= \frac{1}{2}\sum \limits _{i = 0}^N {{\vec {\varPsi }_{ijk}}\cdot \left( J\vec {a}^{\,1}\right) _{ijk}\sum \limits _{m = 0}^N {{Q_{im}}} } \\&\quad + \frac{1}{2}\sum \limits _{i = 0}^N {\sum \limits _{m = 0}^N {{{\vec {\varPsi }_{ijk}}\cdot \left( J\vec {a}^{\,1}\right) _{mjk}Q_{im}}} } \\&= \frac{1}{2}\sum \limits _{i = 0}^N {\sum \limits _{m = 0}^N {{{\vec {\varPsi }_{ijk}}\cdot \left( J\vec {a}^{1}\right) _{mjk}Q_{im}}} }. \end{aligned}\end{aligned}$$

Next, since \(B_{im}=0\) unless \(i=m=0\) or \(i=m=N\),

$$\begin{aligned} \sum \limits _{i=0}^N \sum _{m=0}^N B_{im}\vec {F}^{\,\epsilon }_{ijk}\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}=\sum \limits _{i=0}^N \sum _{m=0}^N B_{im}\vec {F}^{\,\epsilon }_{ijk}\cdot \left( J\vec {a}^{\,1}\right) _{ijk}. \end{aligned}$$

Finally, \(B_{im}\) being diagonal and symmetric, we can swap the \(i\leftrightarrow m\) in \(B_{im}\vec {\varPsi }_{ijk}\) to rewrite the second line of (B.35)

$$\begin{aligned} \begin{aligned}&\sum \limits _{i=0}^N \sum \limits _{m=0}^N Q_{im}\left( \vec {\varPsi }_{ijk} - \vec {\varPsi }_{mjk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} + B_{im}\,\left( \vec {\varPsi }_{ijk}+\vec {F}^{\,\epsilon }_{ijk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} \\&\quad =\sum \limits _{i=0}^N \sum \limits _{m=0}^N \frac{1}{2}Q_{im}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,1}\right) _{mjk} + B_{im}\vec {F}^{\,\epsilon }_{ijk}\cdot \left( J\vec {a}^{\,1}\right) _{ijk}\\&\qquad + (B_{im}-Q_{im})\vec {\varPsi }_{mjk}\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}. \\ \end{aligned} \end{aligned}$$
(B.36)

The last step is to use the summation-by-parts property \(B_{im}-Q_{im}=Q_{mi}\) and re-index to see that

$$\begin{aligned} \begin{aligned} \sum \limits _{i=0}^N \sum \limits _{m=0}^N(B_{im}-Q_{im})\vec {\varPsi }_{mjk}\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}&=\sum \limits _{i=0}^N \sum \limits _{m=0}^N \frac{1}{2}(Q_{mi})\vec {\varPsi }_{mjk}\cdot \left( J\vec {a}^{\,1}\right) _{ijk} \\&= \sum \limits _{i=0}^N \sum \limits _{m=0}^N \frac{1}{2}(Q_{im})\vec {\varPsi }_{ijk}\cdot \left( J\vec {a}^{\,1}\right) _{mjk}\,. \end{aligned} \end{aligned}$$
(B.37)

Therefore,

$$\begin{aligned} \begin{aligned}&\sum \limits _{i=0}^N \sum \limits _{m=0}^N Q_{im}\left( \vec {\varPsi }_{ijk} - \vec {\varPsi }_{mjk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} + B_{im}\,\left( \vec {\varPsi }_{ijk}+\vec {F}^{\,\epsilon }_{ijk}\right) \cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk}\\&\quad =\sum \limits _{i=0}^N \sum \limits _{m=0}^N Q_{im}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,1}\right) _{mjk} + B_{im}\vec {F}^{\,\epsilon }_{ijk}\cdot \left( J\vec {a}^{\,1}\right) _{ijk} . \end{aligned} \end{aligned}$$
(B.38)

To summarize, the first (\(\xi \) direction) part of the volume integral approximation (B.30) is

$$\begin{aligned} \begin{aligned} \sum \limits _{i,j,k=0}^N&\omega _{ijk}\,\mathbf {W}^T_{ijk} 2\sum _{m=0}^N D_{im}\, {\mathop {{\mathbf {F}}}\limits ^{\leftrightarrow }}^{\mathrm {ec}}(U_{ijk}, U_{mjk})\cdot \left\{ \!\!\left\{ J\vec {a}^{\,1}\right\} \!\!\right\} _{(i,m)jk} \\&= \sum \limits _{j,k=0}^N \omega _{jk}\left( \vec {F}^{\,\epsilon }_{Njk}\cdot \left( J\vec {a}^{\,1}\right) _{Njk}-\vec {F}^{\,\epsilon }_{0jk}\cdot \left( J\vec {a}^{\,1}\right) _{0jk}+ \sum \limits _{i,m=0}^N\omega _i D_{im}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,1}\right) _{mjk} \right) \,, \end{aligned} \end{aligned}$$
(B.39)

where we have returned \(Q_{im}=\omega _i D_{im}\) and represented the boundary terms explicitly.

Similar results hold for the second and third parts of the volume integral approximation, leading to

$$\begin{aligned} \begin{aligned}&\left\langle \vec {\mathbb {D}}({\mathop {\tilde{\mathbf {F}}}\limits ^{\leftrightarrow }})^{\mathrm {ec}},\mathbf {W}\right\rangle _N\\&\quad = \sum \limits _{j,k=0}^N \omega _{jk}\left( \vec {F}^{\,\epsilon }_{Njk} \cdot \left( J\vec {a}^{\,1}\right) _{Njk} -\vec {F}^{\,\epsilon }_{0jk} \cdot \left( J\vec {a}^{\,1}\right) _{0jk} + \sum \limits _{i,m=0}^N \omega _i D_{im}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,1}\right) _{mjk} \right) \\&\quad + \sum \limits _{i,k=0}^N \omega _{ik}\left( \vec {F}^{\,\epsilon }_{iNk} \cdot \left( J\vec {a}^{\,2}\right) _{iNk} -\vec {F}^{\,\epsilon }_{i0k} \cdot \left( J\vec {a}^{\,2}\right) _{i0k} + \sum \limits _{j,m=0}^N \omega _j D_{jm}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,2}\right) _{imk} \right) \\&\quad + \sum \limits _{i,j=0}^N \omega _{ij}\left( \vec {F}^{\,\epsilon }_{ijN} \cdot \left( J\vec {a}^{\,3}\right) _{ijN} -\vec {F}^{\,\epsilon }_{ij0} \cdot \left( J\vec {a}^{\,3}\right) _{ij0} + \sum \limits _{k,m=0}^N \omega _k D_{km}\vec {\varPsi }_{ijk} \cdot \left( J\vec {a}^{\,3}\right) _{ijm} \right) \,. \end{aligned} \end{aligned}$$
(B.40)

Regrouping the boundary terms and the volume terms, we have shown that

$$\begin{aligned} \left\langle \vec {\mathbb {D}}({\mathop {\tilde{\mathbf {F}}}\limits ^{\leftrightarrow }})^{\mathrm {ec}},\mathbf {W}\right\rangle _N= & {} \int \limits _{\partial E,N} \left( \vec {\tilde{F}}^{\,\epsilon }\cdot \hat{n}\right) \,\,{\text {dS}} \nonumber \\&+ \sum \limits _{i,j,k=0}^N\omega _{ijk}\varPsi ^1_{ijk}\, \left\{ \sum _{m=0}^N D_{im}\,\left( Ja^1_1\right) _{mjk}+D_{jm}\,\left( Ja^2_1\right) _{imk}+D_{km}\,\left( Ja^3_1\right) _{ijm}\right\} \nonumber \\&+ \sum \limits _{i,j,k=0}^N\omega _{ijk}\varPsi ^2_{ijk}\, \left\{ \sum _{m=0}^N D_{im}\,\left( Ja^1_2\right) _{mjk}+D_{jm}\,\left( Ja^2_2\right) _{imk}+D_{km}\,\left( Ja^3_2\right) _{ijm}\right\} \nonumber \\&+ \sum \limits _{i,j,k=0}^N\omega _{ijk}\varPsi ^3_{ijk}\, \left\{ \sum _{m=0}^N D_{im}\,\left( Ja^1_3\right) _{mjk}+D_{jm}\,\left( Ja^2_3\right) _{imk}+D_{km}\,\left( Ja^3_3\right) _{ijm}\right\} ,\nonumber \\ \end{aligned}$$
(B.41)

which gives the desired entropy condition (B.25), provided that the metric identities (2.32) hold discretely, i.e. that

$$\begin{aligned} \sum _{m=0}^N D_{im}\,\left( J a^1_n\right) _{mjk}+D_{jm}\,\left( Ja^2_n\right) _{imk}+D_{km}\,\left( Ja^3_n\right) _{ijm}=\left( \sum _{l=1}^{3}\frac{\partial }{\partial \xi ^{l}}\mathbb {I}^{N}\left( Ja^l_n\right) \right) _{ijk} = 0 \,,\nonumber \\ \end{aligned}$$
(B.42)

for \(n=1,2,3\) and all GL points \(i,j,k,=0,...,N\) within an element. For the strategy on how to properly approximate the metric terms so that the metric identities are satisfied, see [25].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gassner, G.J., Winters, A.R., Hindenlang, F.J. et al. The BR1 Scheme is Stable for the Compressible Navier–Stokes Equations. J Sci Comput 77, 154–200 (2018). https://doi.org/10.1007/s10915-018-0702-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-018-0702-1

Keywords

Navigation