Skip to main content
Log in

A Unifying Algebraic Framework for Discontinuous Galerkin and Flux Reconstruction Methods Based on the Summation-by-Parts Property

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

Abstract

We propose a unifying framework for the matrix-based formulation and analysis of discontinuous Galerkin (DG) and flux reconstruction (FR) methods for conservation laws on general unstructured grids. Within such an algebraic framework, the multidimensional summation-by-parts (SBP) property is used to establish the discrete equivalence of strong and weak formulations, as well as the conservation and energy stability properties of a broad class of DG and FR schemes. Specifically, the analysis enables the extension of the equivalence between the strong and weak forms of the discontinuous Galerkin collocation spectral-element method demonstrated by Kopriva and Gassner (J Sci Comput 44:136–155, 2010) to more general nodal and modal DG formulations, as well as to the Vincent-Castonguay-Jameson-Huynh (VCJH) family of FR methods. Moreover, new algebraic proofs of conservation and energy stability for DG and VCJH schemes with respect to suitable quadrature rules and discrete norms are presented in which the SBP property serves as a unifying mechanism for establishing such results. Numerical experiments are provided for the two-dimensional linear advection and Euler equations, highlighting the design choices afforded for methods within the proposed framework and corroborating the theoretical analysis.

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

Similar content being viewed by others

Data Availability

The data generated from the numerical experiments described in Sect. 5 as well as the scripts used to process and tabulate the results are available within the above repository.

Code Availability

All computations described in this work were performed using the Generalized High-Order Solver Toolbox (GHOST), which was developed by the first author and is available under the GNU General Public License at https://github.com/tristanmontoya/GHOST.

Notes

  1. For notational convenience, it is assumed that all elements are of the same type and therefore have an equal number of facets, although this is not a limitation of the analysis.

  2. For the purposes of the present analysis, the regularity of such functions is is not of concern; we simply require data to be available at all nodes in \(\mathcal {{S}}\) in order to form an interpolant.

  3. This is also true for the Legendre-Gauss-Radau (LGR) nodes, which contain one endpoint of the domain and support a quadrature rule of degree 2p (see, for example, [20,  Sect. 7.2]).

  4. In such contexts, the weight matrices are generally assumed to be diagonal; Boland and Duris [36] as well as Hunkins [37] provide more general analyses of such approximations.

  5. Recalling the formulation in (3.20), as well as those appearing elsewhere in the literature (e.g. [11,  Eq. (4.35)] and [12,  Eq. (29)]), it is the correction fields \({h}^{(\zeta ,j)}\), and not the correction functions \(\varvec{{g}}^{(\zeta ,j)}\), that appear explicitly in the implementation of a multidimensional FR scheme.

  6. As noted in [26,  Sect. 4] in the context of the DGSEM-LGL, such a modification to the facet terms has no effect on the discretization when \(N= N^*\) and \(\mathcal {{S}}^{(\zeta )} \subset \mathcal {{S}}\) for all \(\zeta \in \{1,\ldots , N_f\}\).

  7. Considering, for example, the quadrature-based discrete inner products in Appendix B.1 and the total-degree polynomial space \({\mathbb {P}}_p({\hat{\varOmega }})\), volume quadrature rules of total degree \(p-1\) or greater and facet quadrature rules of total degree p or greater are generally insufficient for Assumption 3.1 to hold, but nonetheless satisfy (4.16).

  8. Available at https://github.com/tristanmontoya/GHOST.

  9. The latter choice is merely used as a reference value in order to demonstrate that the results in Sect. 4 hold for \(\underline{\underline{K}}{} \ne \underline{\underline{0}}\); the optimality of such a choice in terms of time step size is highly dependent on the time-marching method, the mesh, and the particular problem being solved.

  10. Specifically, we take \(\varDelta t = T/N_t\), where the total number of time steps is given by \(N_t = \lfloor T/ \varDelta t^* \rfloor \) in terms of the target time step \(\varDelta t^* {:}{=}C_t h/a\), with \(\lfloor \cdot \rfloor \) denoting the floor operator.

  11. Such an ordering may be chosen arbitrarily, provided that \(\sigma \) is a bijection; for example, Hesthaven and Warburton [38,  Sect. 6.1] consider \(\sigma (\alpha ) {:}{=}\alpha _2 + (p+1)\alpha _1 + 1 - \alpha _1(\alpha _1-1)/2\).

  12. Taking \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }}) = {\mathbb {P}}_q({\hat{\varOmega }})\), this would require the nodes in \(\mathcal {{S}}\) to be associated with a volume quadrature rule of at least degree 2q, as is the case for the collocated LG quadrature rule in Remark 2.6, for which the mass matrix is the same whether computed as in (2.21) or (2.22).

References

  1. Kreiss, H.-O., Oliger, J.: Comparison of accurate methods for the integration of hyperbolic equations. Tellus 24, 199–215 (1972)

    Article  MathSciNet  Google Scholar 

  2. Wang, Z.J., Fidkowski, K., Abgrall, R., Bassi, F., Caraeni, D., Cary, A., Deconinck, H., Hartmann, R., Hillewaert, K., Huynh, H.T., Kroll, N., May, G., Persson, P.-O., van Leer, B., Visbal, M.: High-order CFD methods: Current status and perspective. International Journal for Numerical Methods in Fluids 72, 811–845 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  3. Reed, W.H., Hill, T.R.: Triangular mesh methods for the neutron transport equation. Tech. Rep. LA-UR-73-479, Los Alamos Scientific Laboratory, USA (1973)

    Google Scholar 

  4. Cockburn, B., Shu, C.-W.: TVB Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws II: General framework. Mathematics of Computation 52, 411–435 (1989)

    MathSciNet  MATH  Google Scholar 

  5. 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. Journal of Computational Physics 84, 90–113 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cockburn, B., Hou, S., Shu, C.-W.: The Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws. IV: The multidimensional case. Mathematics of Computation 54, 545–581 (1990)

    MathSciNet  MATH  Google Scholar 

  7. Cockburn, B., Shu, C.-W.: The Runge-Kutta discontinuous Galerkin method for conservation laws V: Multidimensional systems. Journal of Computational Physics 141, 199–224 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  8. Huynh, H.T.: A flux reconstruction approach to high-order schemes including discontinuous Galerkin methods. In: 18th AIAA Computational Fluid Dynamics Conference, American Institute of Aeronautics and Astronautics, Miami, FL (2007)

  9. Wang, Z.J., Gao, H.: A unifying lifting collocation penalty formulation including the discontinuous Galerkin, spectral volume/difference methods for conservation laws on mixed grids. Journal of Computational Physics 228, 8161–8186 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  10. Vincent, P.E., Castonguay, P., Jameson, A.: A new class of high-order energy stable flux reconstruction schemes. Journal of Scientific Computing 47, 50–72 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  11. Castonguay, P., Vincent, P.E., Jameson, A.: A new class of high-order energy stable flux reconstruction schemes for triangular elements. Journal of Scientific Computing 51, 224–256 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  12. Williams, D.M., Jameson, A.: Energy stable flux reconstruction schemes for advection–diffusion problems on tetrahedra. Journal of Scientific Computing 59, 721–759 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Allaneau, Y., Jameson, A.: Connections between the filtered discontinuous Galerkin method and the flux reconstruction approach to high order discretizations. Computer Methods in Applied Mechanics and Engineering 75, 3628–3636 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  14. De Grazia, D., Mengaldo, G., Moxey, D., Vincent, P.E., Sherwin, S.J.: Connections between the discontinuous Galerkin method and high-order flux reconstruction schemes. International Journal for Numerical Methods in Fluids 75, 860–877 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Mengaldo, G., De Grazia, D., Vincent, P.E., Sherwin, S.J.: On the connections between discontinuous Galerkin and flux reconstruction schemes: Extension to curvilinear meshes. Journal of Scientific Computing 67, 1272–1292 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  16. Zwanenburg, P., Nadarajah, S.: Equivalence between the energy stable flux reconstruction and filtered discontinuous Galerkin schemes. Journal of Computational Physics 306, 343–369 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  17. Del Rey Fernández, D.C., Hicken, J.E., Zingg, D.W.: Review of summation-by-parts operators with simultaneous approximation terms for the numerical solution of partial differential equations. Computers & Fluids 95, 171–196 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  18. Svärd, M., Nordström, J.: Review of summation-by-parts schemes for initial-boundary-value problems. Journal of Computational Physics 268, 17–38 (2014)

    Article  MathSciNet  MATH  Google Scholar 

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

  20. Del Rey Fernández, D.C., Boom, P.D., Zingg, D.W.: A generalized framework for nodal first derivative summation-by-parts operators. Journal of Computational Physics 266, 214–239 (2014)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  23. Ranocha, H., Öffner, P., Sonar, T.: Summation-by-parts operators for correction procedure via reconstruction. Journal of Computational Physics 311, 299–328 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  24. Gassner, G.J., Winters, A.R., Kopriva, D.A.: Split form nodal discontinuous Galerkin schemes with summation-by-parts property for the compressible Euler equations. Journal of Computational Physics 327, 39–66 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  25. Chan, J.: On discretely entropy conservative and entropy stable discontinuous Galerkin methods. Journal of Computational Physics 362, 346–374 (2018)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  27. Zhang, X., Shu, C.-W.: Maximum-principle-satisfying and positivity-preserving high-order schemes for conservation laws: Survey and new developments. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 467, 2752–2776 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  28. Vinokur, M.: Conservation equations of gasdynamics in curvilinear coordinate systems. Journal of Computational Physics 14, 105–125 (1974)

    Article  MathSciNet  MATH  Google Scholar 

  29. Gurtin, M.E., Fried, E., Anand, L.: The Mechanics and Thermodynamics of Continua. Cambridge University Press, UK (2010)

    Book  Google Scholar 

  30. Cohen, A., Migliorati, G.: Multivariate approximation in downward closed polynomial spaces. In: Contemporary Computational Mathematics – A Celebration of the 80th Birthday of Ian Sloan (J. Dick, F. Y. Kuo, and H. Woźniakowski, eds.), pp. 233–282, Springer, Cham (2018)

  31. Yu, M., Wang, Z.J., Liu, Y.: On the accuracy and efficiency of discontinuous Galerkin, spectral difference and correction procedure via reconstruction methods. Journal of Computational Physics 259, 70–95 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  32. Toro, E.F.: Riemann Solvers and Numerical Methods for Fluid Dynamics: A Practical Introduction, 3rd edn. Springer, Berlin Heidelberg (2009)

    Book  MATH  Google Scholar 

  33. Carpenter, M.H., Gottlieb, D.: Spectral methods on arbitrary grids. Journal of Computational Physics 129, 74–86 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  34. Vincent, P.E., Farrington, A.M., Witherden, F.D., Jameson, A.: An extended range of stable-symmetric-conservative flux reconstruction correction functions. Computer Methods in Applied Mechanics and Engineering 296, 248–272 (2015)

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  36. Boland, W.R., Duris, C.S.: Product type quadrature formulas. BIT 11, 139–158 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  37. Hunkins, D.R.: Product type multiple integration formulas. BIT 13, 408–414 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  38. Hesthaven, J.S., Warburton, T.: Nodal Discontinuous Galerkin Methods: Algorithms, Analysis, and Applications. Springer, New York (2008)

    Book  MATH  Google Scholar 

  39. Karniadakis, G.E., Sherwin, S.J.: Spectral/hp Element Methods for Computational Fluid Dynamics, 2nd edn. Oxford University Press, UK (2005)

    Book  MATH  Google Scholar 

  40. Chen, T., Shu, C.-W.: Review of entropy stable discontinuous Galerkin methods for systems of conservation laws on unstructured simplex meshes. CSIAM Transactions on Applied Mathematics 1, 1–52 (2020)

    Google Scholar 

  41. Carpenter, M.H., Gottlieb, D., Abarbanel, S.: Time-stable boundary conditions for finite-difference schemes solving hyperbolic systems: Methodology and application to high-order compact schemes. Journal of Computational Physics 111, 220–236 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  42. Funaro, D., Gottlieb, D.: A new method of imposing boundary conditions in pseudospectral approximations of hyperbolic equations. Mathematics of Computation 51, 599–599 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  43. Del Rey Fernández, D.C., Hicken, J.E., Zingg, D.W.: Simultaneous approximation terms for multi-dimensional summation-by-parts operators. Journal of Scientific Computing 75, 83–110 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  44. Horn, R.A., Johnson, C.R.: Matrix Analysis, 2nd edn. Cambridge University Press, UK (2013)

    MATH  Google Scholar 

  45. Jameson, A.: A proof of the stability of the spectral difference method for all orders of accuracy. Journal of Scientific Computing 45, 348–358 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  46. Gustafsson, B., Kreiss, H.-O., Oliger, J.: Time-Dependent Problems and Difference Methods, 2nd edn. John Wiley & Sons Inc, Hoboken, NJ (2013)

  47. Fisher, T.C., Carpenter, M.H.: High-order entropy stable finite difference schemes for nonlinear conservation laws: Finite domains. Journal of Computational Physics 252, 518–557 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  48. Carpenter, M.H., Fisher, T.C., Nielsen, E.J., Frankel, S.H.: Entropy stable spectral collocation schemes for the Navier-Stokes equations: Discontinuous interfaces. SIAM Journal on Scientific Computing 36(5), B835–B867 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  49. Crean, J., Hicken, J.E., Del Rey Fernández, D.C., Zingg, D.W., Carpenter, M.H.: Entropy-stable summation-by-parts discretization of the Euler equations on general curved elements. Journal of Computational Physics 356, 410–438 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  50. Chen, T., Shu, C.-W.: Entropy stable high order discontinuous Galerkin methods with suitable quadrature rules for hyperbolic conservation laws. Journal of Computational Physics 345, 427–461 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  51. Xiao, H., Gimbutas, Z.: A numerical algorithm for the construction of efficient quadrature rules in two and higher dimensions. Computers & Mathematics with Applications 59, 663–676 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  52. Warburton, T.: An explicit construction of interpolation nodes on the simplex. Journal of Engineering Mathematics 56, 247–262 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  53. Roe, P.L.: Approximate Riemann solvers, parameter vectors, and difference schemes. Journal of Computational Physics 43, 357–372 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  54. Del Rey Fernández, D.C., Boom, P.D., Shademan, M., Zingg, D.W.: Numerical investigation of tensor-product summation-by-parts discretization strategies and operators. In: 55th AIAA Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, US (2017)

    Google Scholar 

  55. Cockburn, B., Shu, C.-W.: Runge-Kutta discontinuous Galerkin methods for convection-dominated problems. Journal of Scientific Computing 16, 173–261 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  56. Shu, C.-W.: Essentially non-oscillatory and weighted essentially non-oscillatory schemes for hyperbolic conservation laws. In: Advanced Numerical Approximation of Nonlinear Hyperbolic Equations: Lectures given at the 2nd Session of the Centro Internazionale Matematico Estivo (C.I.M.E.) held in Cetraro, Italy, June 23–28, 1997 (A. Quarteroni, ed.), pp. 325–432, Springer, Berlin Heidelberg (1998)

  57. 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, American Institute of Aeronautics and Astronautics, Dallas, TX (2015)

  58. Ponce, M., van Zon, R., Northrup, S., Gruner, D., Chen, J., Ertinaz, F., Fedoseev, A., Groer, L., Mao, F., Mundim, B. C., Nolta, M., Pinto, J., Saldarriaga, M., Slavnic, V., Spence, E., Yu, C.-H., Peltier, W. R.: Deploying a top-100 supercomputer for large parallel workloads: The Niagara supercomputer. In: Proceedings of the Practice and Experience in Advanced Research Computing on Rise of the Machines (Learning), Association for Computing Machinery, New York (2019)

  59. Proriol, J.: Sur une famille de polynomes à deux variables orthogonaux dans un triangle. Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences 245, 2459–2461 (1957)

    MathSciNet  MATH  Google Scholar 

  60. Koornwinder, T.: Two-variable analogues of the classical orthogonal polynomials. In: Theory and Application of Special Functions (R. Askey, ed.), pp. 435–495, Academic Press, New York (1975)

  61. Dubiner, M.: Spectral methods on triangles and other domains. Journal of Scientific Computing 6, 345–390 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  62. Marchildon, A. L., Zingg, D. W.: Unisolvency for polynomial interpolation in simplices with symmetrical nodal distributions. Journal of Scientific Computing 92, 50 (2022)

  63. Cicchino, A., Nadarajah, S.: A new norm and stability condition for tensor product flux reconstruction schemes. Journal of Computational Physics 429, 110025 (2021)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Funding

The authors acknowledge the financial support provided by the University of Toronto, the Natural Sciences and Engineering Research Council of Canada, and the Government of Ontario. Computations were performed on the Niagara supercomputer at the SciNet HPC Consortium [58], which is funded by the Canada Foundation for Innovation, the Government of Ontario, the Ontario Research Fund – Research Excellence, and the University of Toronto.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tristan Montoya.

Ethics declarations

Competing interests

The authors declare that there are no conflicts of interest or competing interests which could have influenced this work.

Additional information

Publisher's Note

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

Appendices

Appendix

1.1 Appendix A Polynomial Bases

Although the theoretical analysis in this work is independent of the choice of basis \(\mathcal {{B}}\) employed for a given discretization, the concrete implementation of a DG or FR method nevertheless requires a basis to be chosen. Two families of such bases, both of which are used for the computations in Sect. 5, are thus described in this appendix.

1.2 A.1 Modal (\(L^2\)-Orthogonal) Basis

A basis \(\mathcal {{B}}_0 {:}{=}\{\phi _0^{(1)}, \ldots ,\phi _0^{(N^*)} \}\) is said to be orthogonal with respect to the \(L^2\) inner product on the reference element if any two basis functions \(\phi _0^{(i)}, \phi _0^{(j)}\in \mathcal {{B}}\) satisfy

$$\begin{aligned} \int _{{\hat{\varOmega }}}\phi _0^{(i)}(\varvec{\xi })\phi _0^{(j)}(\varvec{\xi })\, \text {d} \varvec{\xi } = 0, \quad \forall \, i \ne j. \end{aligned}$$
(A.1)

Analytical expressions for such bases are well known for certain spaces and element types; for example, an orthogonal basis for the space \({\mathbb {P}}_p(\hat{\mathcal {{T}}}^2)\) was introduced by Proriol [59] (see also Koornwinder [60,  Sect. 3.3] and Dubiner [61,  Sect. 5]) as a triangular analogue of the Legendre polynomial system. Defining \({P}_r^{(a,b)} \in {\mathbb {P}}_r([-1,1])\) as the degree \(r \in {\mathbb {N}}_0\) Jacobi polynomial satisfying

$$\begin{aligned} \int _{-1}^1 {P}_q^{(a,b)}(\xi ){P}_r^{(a,b)}(\xi )(1-\xi )^a(1+\xi )^b \, \text {d} \xi = 0, \quad \forall \, q \ne r \end{aligned}$$
(A.2)

the normalized Proriol-Koornwinder-Dubiner (PKD) basis functions are given by

$$\begin{aligned} \phi _0^{(\sigma (\alpha ))}(\varvec{\chi }(\varvec{\eta })) {:}{=}\sqrt{\big (\alpha _1 + \textstyle {\frac{1}{2}}\big )\big (\alpha _1 + \alpha _2 + 1\big )} {P}_{\alpha _1}^{(0,0)}(\eta _1)\Big (\frac{1-\eta _2}{2}\Big )^{\alpha _1}{P}_{\alpha _2}^{(2\alpha _1+1,0)}(\eta _2), \end{aligned}$$
(A.3)

where \(\sigma : \mathcal {{N}} \rightarrow \{1,\ldots , (p+1)(p+2)/2\}\) defines an orderingFootnote 11 of the multi-index set \(\mathcal {{N}} {:}{=}\{\varvec{\alpha } \in {\mathbb {N}}_{0}^2 : \alpha _1 + \alpha _2 \le p \}\), and the mapping \(\varvec{\chi } : [-1,1]^2 \rightarrow \hat{\mathcal {{T}}}^2\) is given by

$$\begin{aligned} \varvec{\chi }(\varvec{\eta }) {:}{=}\left[ \begin{array}{c} \frac{1}{2}(1+\eta _1)(1-\eta _2) - 1 \\ \eta _2\end{array}\right] . \end{aligned}$$
(A.4)

Similar bases are described for quadrilateral, hexahedral, tetrahedral, prismatic, and pyramidal elements in [39,  Ch. 3].

1.3 A.2 Nodal (Lagrange) Basis

While it is possible to directly employ a modal basis to represent the numerical solution such that \(\mathcal {{B}} = \mathcal {{B}}_0\), it can often be more computationally efficient to use a Lagrange basis \(\mathcal {{B}} {:}{=}\{\phi ^{(1)}, \ldots , \phi ^{(N^*)}\}\) associated with a nodal set \(\tilde{\mathcal {{S}}} {:}{=}\{\tilde{\varvec{\xi }}{}^{(1)}, \ldots ,\tilde{\varvec{\xi }}{}^{(N^*)} \} \subset {\hat{\varOmega }}\) such that \(\phi ^{(i)}({\tilde{\varvec{\xi }}}{}^{(j)}) = \delta _{ij}\). Although the corresponding basis functions are given explicitly by (2.17) for \(d=1\), analytical expressions do not necessarily exist for nodal bases on arbitrary non-tensorial nodal sets in the case of \(d \ge 2\). We therefore take a more general approach by constructing a matrix \(\tilde{\underline{\underline{V}}{}} \in {\mathbb {R}}^{N^*\times N^*}\) with entries

$$\begin{aligned} {\tilde{V}}_{ij} {:}{=}\phi _0^{(j)}(\tilde{\varvec{\xi }}{}^{(i)}) \end{aligned}$$
(A.5)

for a modal basis \(\mathcal {{B}}_0\) as described in Appendix A.1, which is invertible when \(\tilde{\mathcal {{S}}}\) is unisolvent for the space \({\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }})\). The Lagrange polynomials constituting the basis \(\mathcal {{B}}\) may then be obtained in terms of the modal basis as

$$\begin{aligned} \phi ^{(i)}(\varvec{\xi }) {:}{=}\sum _{j=1}^{N^*}\Big [\tilde{\underline{\underline{V}}{}}^{-\mathrm {T}}\Big ]_{ij}\phi _0^{(j)}(\varvec{\xi }). \end{aligned}$$
(A.6)

The entries of the corresponding derivative matrix satisfying (3.10) are therefore

$$\begin{aligned} D_{ij}^{(m)} = \frac{\partial \phi ^{(j)}}{\partial \xi _m}(\tilde{\varvec{\xi }}{}^{(i)}) = \sum _{k=1}^{N^*}\Big [\tilde{\underline{\underline{V}}{}}^{-\mathrm {T}}\Big ]_{jk}\frac{\partial \phi _0^{(k)}}{\partial \xi _m}(\tilde{\varvec{\xi }}{}^{(i)}), \quad \forall \, i,j \in \{1,\ldots , N^*\}, \end{aligned}$$
(A.7)

defining a nodal SBP operator on \(\tilde{\mathcal {{S}}}\) in the sense of [21,  Definition 2.1] in the case of discrete inner products satisfying Assumptions 3.1 and 3.2.

Remark A.1

In the case of \(d \ge 2\), unisolvency is not guaranteed for arbitrary nodal sets \(\tilde{\mathcal {{S}}}\) containing \(N^*\) distinct nodes. Special care is therefore needed to ensure that \(\tilde{\underline{\underline{V}}{}}\) is invertible, where we refer to Marchildon and Zingg [62] for the derivation of necessary conditions for obtaining unisolvent symmetrical nodal sets for total-degree polynomial spaces on triangles and tetrahedra. Further details regarding nodal bases for high-order methods may be found, for example, in [38,  Sect. 3.1, Sect. 6.1, Sect. 10.1].

Appendix B Discrete Inner Products

Implementations of the DG and FR methods presented in Sect.  2.5 can often be characterized by the techniques employed for numerical integration or projection, which may be formalized in terms of discrete inner products defined as in (3.1) and (3.2). In this appendix, we situate the standard quadrature-based integration approach commonly employed for DG methods (e.g. those in [4,5,6,7]) as well as the collocation-based approach employed for the nodal DG formulations described in [38] and the FR schemes in [8,9,10,11,12] within the general context of the present framework.

1.1 B.1 Quadrature-Based Approximation

Considering quadrature rules on \(\mathcal {{S}}\) and \(\mathcal {{S}}^{(\zeta )}\) with positive weights given by \(\{\omega ^{(i)}\}_{i=1}^{N}\) and \(\{\omega ^{(\zeta ,i)}\}_{i=1}^{N_\zeta }\), respectively, we may define the diagonal weight matrices

$$\begin{aligned} \underline{\underline{W}}{} {:}{=}{\text {diag}}\big (\omega ^{(1)}, \ldots , \omega ^{(N)}\big ) \quad \text {and} \quad \underline{\underline{B}}{}^{(\zeta )} {:}{=}{\text {diag}}\big (\omega ^{(\zeta ,1)}, \ldots , \omega ^{(\zeta ,N_\zeta )}\big ), \end{aligned}$$
(B.1)

such that the double sums in (3.1) and (3.2) reduce to

$$\begin{aligned} \big \langle {u},{v}\big \rangle _{W} {:}{=}\sum _{i=1}^{N}{u}(\varvec{\xi }^{(i)}) {v}(\varvec{\xi }^{(i)})\,\omega ^{(i)} \end{aligned}$$
(B.2)

and

$$\begin{aligned} \big \langle {u},{v}\big \rangle _{B^{(\zeta )}} {:}{=}\sum _{i=1}^{N_\zeta }{u}(\varvec{\xi }^{(\zeta ,i)}) {v}(\varvec{\xi }^{(\zeta ,i)})\,\omega ^{(\zeta ,i)}, \end{aligned}$$
(B.3)

respectively. The following lemma provides sufficient conditions on such quadrature rules for Assumption 3.1 to be satisfied.

Lemma B.1

Assumption 3.1 is satisfied with \({\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }})={\mathbb {P}}_p({\hat{\varOmega }})\) for any \(p \in {\mathbb {N}}_0\), on any polytopal reference element, if the discrete inner products are computed as in (B.2) and (B.3) using quadrature rules of at least total degree \(2p-1\) and 2p, respectively.

Proof

For any \({u},{v} \in {\mathbb {P}}_p({\hat{\varOmega }})\), the volume integrals in (3.3) contain functions in \({\mathbb {P}}_{2p-1}({\hat{\varOmega }})\), while the facet integrals contain functions in \({\mathbb {P}}_{2p}({\hat{\varGamma }}^{(\zeta )})\). For volume and facet quadrature rules of total degree \(2p-1\) or greater and 2p or greater, respectively, all terms in (3.3) are computed exactly, and hence (3.4) holds for \(m \in \{1,\ldots , d\}\). \(\square \)

Remark B.1

In addition to the positive-definiteness of \(\underline{\underline{W}}{}\) and \(\underline{\underline{B}}{}^{(\zeta )}\), which is clearly the case if and only if the volume and facet quadrature weights are strictly positive, Assumption 3.2 requires \(\underline{\underline{V}}{}\) to be of rank \(N^*\), which is difficult to ensure theoretically, particularly in the case of \(N > N^*\). We have nevertheless numerically verified that such a property is satisfied for all schemes employed for the computations in Sect. 5.

1.2 B.2 Collocation-Based Approximation

If \(\mathcal {{S}}\) is unisolvent for a space \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }}) \supseteq {\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }})\) of dimension \(N\ge N^*\), we may define a nodal basis \(\{ \ell ^{(1)}, \ldots , \ell ^{(N)}\}\) for \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }})\) satisfying \(\ell ^{(i)}(\varvec{\xi }^{(j)}) = \delta _{ij}\) as in Appendix A.2. The collocation projection \({I}_{\mathcal {{K}}}{u} \in {\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }})\) of a function \({u} : {\hat{\varOmega }} \rightarrow {\mathbb {R}}\) is then given by

$$\begin{aligned} ({I}_{\mathcal {{K}}}{u})(\varvec{\xi }) {:}{=}\sum _{i=1}^{N}{u}(\varvec{\xi }^{(i)})\ell ^{(i)}(\varvec{\xi }), \end{aligned}$$
(B.4)

recovering the projection in (2.13) when \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }}) = {\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }})\). Similarly, if \(\mathcal {{S}}^{(\zeta )}\) is unisolvent for a polynomial space on \({\hat{\varGamma }}^{(\zeta )} \subset \partial {\hat{\varOmega }}\) which is of dimension \(N_\zeta \ge N_\zeta ^*\) and contains \({\mathbb {P}}_{\mathcal {{N}}}({\hat{\varGamma }}^{(\zeta )})\), we may define a projection operator analogously to (B.4) as

$$\begin{aligned} ({I}^{(\zeta )}{u})(\varvec{\xi }) {:}{=}\sum _{i=1}^{N_\zeta }{u}(\varvec{\xi }^{(\zeta ,i)})\ell ^{(\zeta ,i)}(\varvec{\xi }), \end{aligned}$$
(B.5)

where the nodal basis functions \(\{ \ell ^{(\zeta ,1)}, \ldots , \ell ^{(\zeta ,N_\zeta )}\}\) satisfy \(\ell ^{(\zeta ,i)}(\varvec{\xi }^{(\zeta ,j)}) = \delta _{ij}\). Integrating the products of such projections, we obtain discrete inner products given by

$$\begin{aligned} \big \langle {u},{v}\big \rangle _{W} {:}{=}\int _{{\hat{\varOmega }}} ({I}_{\mathcal {{K}}}{u})(\varvec{\xi })({I}_{\mathcal {{K}}}{v})(\varvec{\xi }) \, \text {d} \varvec{\xi } \end{aligned}$$
(B.6)

and

$$\begin{aligned} \big \langle {u},{v}\big \rangle _{B^{(\zeta )}} {:}{=}\int _{{\hat{\varGamma }}^{(\zeta )}} ({I}^{(\zeta )}{u})(\varvec{\xi })({I}^{(\zeta )}{v})(\varvec{\xi }) \, \text {d} {\hat{s}}, \end{aligned}$$
(B.7)

which take the forms in (3.1) and (3.2), respectively, with

$$\begin{aligned} W_{ij} {:}{=}\int _{{\hat{\varOmega }}}\ell ^{(i)}(\varvec{\xi })\ell ^{(j)}(\varvec{\xi })\, \text {d} \varvec{\xi } \quad \text {and} \quad B_{ij}^{(\zeta )} {:}{=}\int _{{\hat{\varGamma }}^{(\zeta )}}\ell ^{(\zeta ,i)}(\varvec{\xi })\ell ^{(\zeta ,j)}(\varvec{\xi })\, \text {d} {\hat{s}}, \end{aligned}$$
(B.8)

where the above integrals may be evaluated as part of a preprocessing stage.

Remark B.2

Noting that the matrices \(\underline{\underline{W}}{}\) and \(\underline{\underline{B}}{}^{(\zeta )}\) defined in (B.8) are generally dense, and are SPD due to the linear independence of the nodal bases used to define the projection operators (see, for example, [44,  Theorem 7.2.10]), Assumptions 3.1 and 3.2 (as well as Assumption 3.3 for meshes satisfying Assumption 2.1, at least in the case of \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }}) = {\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }})\) considered for the computations in Sect. 5) are therefore satisfied by construction. Such a quadrature-free formulation enables the construction of conservative and energy-stable schemes on arbitrary unisolvent nodal sets for which the quadrature accuracy requirements of Lemma B.1 may not be met.

Remark B.3

Recalling (4.15), the weights of the interpolatory quadrature rule on the abscissae \(\mathcal {{S}}\) under which conservation may be proven for collocation-based approximations may be expressed in terms of the Lagrange basis functions as

$$\begin{aligned} \omega ^{(i)} = \sum _{j=1}^{N} W_{ij} = \int _{{\hat{\varOmega }}}\ell ^{(i)}(\varvec{\xi })\, \text {d} \varvec{\xi }, \quad \forall \, i \in \{1,\ldots , N\}, \end{aligned}$$
(B.9)

where such a quadrature is exact (at least) for all functions in \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }})\). If, however, the nodes in \(\mathcal {{S}}\) are chosen such that the quadrature rule with weights given as in (B.9) is exact for all products of two functions in \({\mathbb {P}}_{\mathcal {{K}}}({\hat{\varOmega }})\),Footnote 12 it can be shown (see, for example, [37]) that \(\underline{\underline{W}}{}\) reduces to a diagonal matrix of quadrature weights, recovering an approximation identical to that in Appendix B.1. An analogous equivalence holds for the matrices \(\underline{\underline{B}}{}^{(\zeta )}\) and the associated facet quadrature rules.

Appendix C Parametrization of the \(\underline{\underline{K}}{}\) Matrix

The following lemma establishes suitable choices of multi-index sets \(\mathcal {{M}} \subset \mathcal {{N}}\) satisfying the first part of Assumption 3.4 for schemes employing total-degree polynomial spaces (the case of tensor-product elements is discussed by Cicchino and Nadarajah in [63]).

Lemma C.1

Taking \({\mathbb {P}}_{\mathcal {{N}}}({\hat{\varOmega }}) = {\mathbb {P}}_p({\hat{\varOmega }})\) for any \(p \in {\mathbb {N}}_0\), the choice of \(\mathcal {{M}} {:}{=}\{\varvec{\alpha } \in {\mathbb {N}}_0^d : |\varvec{\alpha } |= p\}\) in (3.26) ensures that \(\underline{\underline{K}}{}\underline{\underline{D}}{}^{(m)} = \underline{\underline{0}}\) is satisfied for all \(m \in \{1,\ldots , d\}\).

Proof

Noting from (3.26) that \(\underline{\underline{K}}{}\underline{\underline{D}}{}^{(m)} = \underline{\underline{0}}\) is satisfied for a given \(m \in \{1,\ldots , d\}\) if \(\underline{\underline{D}}{}^{\varvec{\alpha }}\underline{\underline{D}}{}^{(m)} = \underline{\underline{0}}\) for all \(\varvec{\alpha } \in \mathcal {{M}}\), it is sufficient to require the image of any \({v} \in {\mathbb {P}}_p({\hat{\varOmega }})\) under the corresponding differential operator \(\partial ^{{|{\varvec{\alpha }}|}+1}/\partial \xi _1^{\alpha _1} \cdots \partial \xi _m^{\alpha _m + 1} \cdots \partial \xi _d^{\alpha _d}\) to be the zero function. Since differentiating \({v} \in {\mathbb {P}}_p({\hat{\varOmega }})\) a total of \(p+1\) times always yields the zero function, and the action of the above operator consists of differentiating a total of \(|\varvec{\alpha } |+ 1\) times, constraining \(\mathcal {{M}}\) to satisfy \(|\varvec{\alpha }|= p\) leads to the desired result. \(\square \)

Based on symmetry considerations discussed in [11,  Sect. 5.3] and [12,  Sect. 4.1], the set of multi-indices \(\mathcal {{M}}\) defined in Lemma C.1 may be parametrized in terms of \(d-1\) scalar indices. The matrix \(\underline{\underline{K}}{}\) in (3.26) is then given in two and three space dimensions by

$$\begin{aligned} \underline{\underline{K}}{} {:}{=}\frac{c}{|{\hat{\varOmega }}|}\sum _{q=0}^p\left( \begin{array}{c}p \\ q\end{array} \right) \big (\underline{\underline{D}}{}^{(p-q,q)}\big )^\mathrm {T}\underline{\underline{M}}{}\underline{\underline{D}}{}^{(p-q,q)} \end{aligned}$$
(C.1)

and

$$\begin{aligned} \underline{\underline{K}}{} {:}{=}\frac{c}{|{\hat{\varOmega }}|}\sum _{q_1=0}^p\sum _{q_2=0}^{q_1}\left( \begin{array}{c}p \\ q_1\end{array} \right) \left( \begin{array}{c}q_1 \\ q_2\end{array} \right) \big (\underline{\underline{D}}{}^{(p-q_1,q_1-q_2,q_2)}\big )^\mathrm {T}\underline{\underline{M}}{}\underline{\underline{D}}{}^{(p-q_1,q_1-q_2,q_2)}, \end{aligned}$$
(C.2)

respectively, recovering (2.27) in the one-dimensional case, where \(c \in {\mathbb {R}}\) determines the properties of the resulting schemes. With respect to the formalism of Theorem 3.1, the parametrizations defining \(\underline{\underline{K}}{}\) in (C.1) and (C.2) correspond to \(c_{\varvec{\alpha }} = c\,\left( {\begin{array}{c}p\\ \alpha _2\end{array}}\right) \) and \(c_{\varvec{\alpha }} = c\, \left( {\begin{array}{c}p\\ p-\alpha _1\end{array}}\right) \left( {\begin{array}{c}p-\alpha _1\\ \alpha _3\end{array}}\right) \), respectively, and lead to \(\underline{\underline{K}}{}\) being SPSD for all \(c \ge 0\), which is sufficient under Assumption 3.2 for \(\underline{\underline{M}}{}+\underline{\underline{K}}{}\) to be SPD, implying that the second part of Assumption 3.4 is satisfied under such conditions.

Remark C.1

While \(c \ge 0\) is typically assumed for simplicial elements with \(d \ge 2\), it has been shown for \(d=1\) (see, for example, [10,  Sect. 3.3], [13,  Sect. 3.2], and [23,  Sect. 3.6]) that there exists \(c_- < 0\) depending only on the polynomial degree p (and on the volume quadrature rule, if (3.24) is not satisfied) such that \(\underline{\underline{M}}{}+\underline{\underline{K}}{}\) is SPD for \( c_-< c < \infty \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Montoya, T., Zingg, D.W. A Unifying Algebraic Framework for Discontinuous Galerkin and Flux Reconstruction Methods Based on the Summation-by-Parts Property. J Sci Comput 92, 87 (2022). https://doi.org/10.1007/s10915-022-01935-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10915-022-01935-3

Keywords

Mathematics Subject Classification

Navigation