Skip to main content
Log in

Trefftz discontinuous Galerkin basis functions for a class of Friedrichs systems coming from linear transport

  • Published:
Advances in Computational Mathematics Aims and scope Submit manuscript

Abstract

The Trefftz discontinuous Galerkin (TDG) method provides natural well-balanced (WB)and asymptotic preserving (AP) discretization, since exact solutions are used locally in the basis functions. However, one difficult point may be the construction of such solutions, which is a necessary first step in order to apply the TDG scheme. This work deals with the construction of solutions to Friedrichs systems with relaxation with application to the spherical harmonics approximation of the transport equation (the so-called PN models). Various exponential and polynomial solutions are constructed. Two numerical tests on the P3 model illustrate the good accuracy of the TDG method. They show that the exponential solutions lead to accurate schemes to capture boundary layers on a coarse mesh and that a combination of exponential and polynomial solutions is efficient in a regime with vanishing absorption coefficient.

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

References

  1. Abdelaziz, Y., Hamouine, A.: A survey of the extended finite element. Computers & Structures 86(11), 1141–1151 (2008)

    Google Scholar 

  2. Anderson, J.: A secular equation for the eigenvalues of a diagonal matrix perturbation. Linear Algebra Appl. 246(0), 49–70 (1996)

    MathSciNet  MATH  Google Scholar 

  3. Berthon, C., Turpault, R.: Asymptotic preserving HLL schemes. Numer. Methods Partial Differ. Equations 27(6), 1396–1422 (2011)

    MathSciNet  MATH  Google Scholar 

  4. Birkhoff, G., Abu-Shumays, I.: Harmonic solutions of transport equations. J. Math. Anal Appl. 28, 211–221 (1969)

    MathSciNet  MATH  Google Scholar 

  5. Birkhoff, G., Abu-Shumays, I.K.: Exact analytic solutions of transport equations. J. Math. Anal Appl. 32, 468–481 (1970)

    MathSciNet  MATH  Google Scholar 

  6. Blanco, A.M., Florez, M., Bermejo, M.: Evaluation of the rotation matrices in the basis of real spherical harmonics. Journal of Molecular Structure-theochem 419, 19–27, 12 (1997)

    Google Scholar 

  7. Brunner, T.A., Holloway, J.P.: One-dimensional Riemann solvers and the maximum entropy closure. J. Quant. Spectros. Radiat. Transfer 69(5), 386–399 (2005)

    MATH  Google Scholar 

  8. Buet, C., Després, B., Franck, E.: Asymptotic preserving schemes on distorted meshes for Friedrichs systems with stiff relaxation: application to angular models in linear transport. J. Sci. Comput. 62(2), 371–398 (2015)

    MathSciNet  MATH  Google Scholar 

  9. Buffa, A., Monk, P.: Error estimates for the ultra weak variational formulation of the Helmholtz equation. ESAIM: Mathematical Modelling and Numerical Analysis 8(6), 925–940 (2008)

    MathSciNet  MATH  Google Scholar 

  10. Case, K.M.: Elementary solutions of the transport equation and their applications. Ann. Phys. 9, 1–23 (1960)

    MathSciNet  MATH  Google Scholar 

  11. Cessenat, O., Després, B.: Application of an ultra weak variational formulation of Elliptic PDEs to the two-dimensional Helmholtz problem. SIAM J. Numer Anal. 35(1), 255–299 (1998)

    MathSciNet  MATH  Google Scholar 

  12. Dai, F., Xu, Y.: Approximation Theory and Harmonic Analysis on Spheres and Balls. Springer, New York (2013)

    MATH  Google Scholar 

  13. Després, B., Buet, C.: The structure of well-balanced schemes for Friedrichs systems with linear relaxation. Appl. Math. Comput. 272, 440–459 (2016)

    MathSciNet  MATH  Google Scholar 

  14. Ern, A., Guermond, J.-L.: Discontinuous Galerkin methods for Friedrichs’ systems. I.general theory. SIAM J. Numerical Analysis 44(2), 753–778 (2006)

    MathSciNet  MATH  Google Scholar 

  15. Fries, T.-P., Belytschko, T.: The extended/generalized finite element method: an overview of the method and its applications. Int. J. Numer. Methods Eng. 84(3), 253–304 (2010)

    MathSciNet  MATH  Google Scholar 

  16. Gabard, G.: Exact integration of polynomial-exponential products with application to wave-based numerical methods. Comm. Numer. Methods Engrg. 25(3), 237–246 (2009)

    MathSciNet  MATH  Google Scholar 

  17. Garrett, C.K., Hauck, C.D.: On the eigenstructure of spherical harmonic equations for radiative transport. Comput. Math. Appl. 72(2), 264–270 (2016)

    MathSciNet  MATH  Google Scholar 

  18. Gittelson, C.J., Hiptmair, R., Perugia, I.: Plane wave discontinuous Galerkin methods: analysis of the h-version. ESAIM Math. Model. Numer. Anal. 43 (2), 297–331 (2009)

    MathSciNet  MATH  Google Scholar 

  19. Gopalakrishnan, J., Monk, P., Sepúlveda, P.: A tent pitching scheme motivated by Friedrichs theory. Comput. Math. Appl. 70(5), 1114–1135 (2015)

    MathSciNet  Google Scholar 

  20. Gosse, L.: Computing Qualitatively Correct Approximations of Balance Laws. Exponential-Fit, Well-Balanced and Asymptotic-Preserving. Springer, Milano (2013)

    MATH  Google Scholar 

  21. Gosse, L., Toscani, G.: An asymptotic-preserving well-balanced scheme for the hyperbolic heat equations. C. R., Math., Acad. Sci Paris 334(4), 337–342 (2002)

    MathSciNet  MATH  Google Scholar 

  22. Greenberg, J.M., Leroux, A.Y.: A well-balanced scheme for the numerical processing of source terms in hyperbolic equations. SIAM J. Numer. Anal. 33(1), 1–16 (1996)

    MathSciNet  Google Scholar 

  23. Hermeline, F.: A discretization of the multigroup pN radiative transfer equation on general meshes. J. Comput. Phys. 313, 549–582 (2016)

    MathSciNet  MATH  Google Scholar 

  24. Hiptmair, R., Moiola, A., Perugia, I.: Plane wave discontinuous Galerkin methods for the 2D Helmholtz equation: analysis of the p-version. SIAM J. Numer. Anal. 49(1), 264–284 (2011)

    MathSciNet  MATH  Google Scholar 

  25. Hiptmair, R., Moiola, A., Perugia, I.: A survey of Trefftz methods for the Helmholtz equation. In: Building Bridges: Connections and Challenges in Modern Approaches to Numerical Partial Differential Equations, volume 114 of Lect. Notes Comput. Sci. Eng., pp 237–278. Springer, Cham (2016)

  26. Ivanic, J., Ruedenberg, K.: Rotation matrices for real spherical harmonics. direct determination by recursion. J. Phys. Chem. 100(15) (1996)

  27. Jin, S.: A steady-state capturing method for hyperbolic systems with geometrical source terms. In: Naoufel Ben et al Abdallah (ed.) Transport in Transition Regimes, pp 177–183. Springer, New York (2004)

  28. Jin, S., Levermore, C.D.: Numerical schemes for hyperbolic conservation laws with stiff relaxation terms. J Comput. Phys. 126(2), 449–467 (1996)

    MathSciNet  MATH  Google Scholar 

  29. Jin, S., Levermore, D.: The discrete-ordinate method in diffusive regimes. Transport Theory and Statistical Physics 20, 413–439, 10 (1991)

    MathSciNet  MATH  Google Scholar 

  30. Kretzschmar, F., Moiola, A., Perugia, I., Schnepp, S.M.: A priori error analysis of space–time Trefftz discontinuous Galerkin methods for wave problems. IMA J. Numer. Anal. 36(4), 1599 (2016)

    MathSciNet  MATH  Google Scholar 

  31. Monk, P., Richter, G.R.: A discontinuous Galerkin method for linear symmetric hyperbolic systems in inhomogeneous media. J. Sci Comput. 22-23, 443–477 (2005)

    MathSciNet  MATH  Google Scholar 

  32. Morel, G.: Asymptotic-preserving and well-balanced schemes for transport models using Trefftz discontinuous Galerkin method. Theses, Sorbonne Université (2018)

  33. Morel, G., Buet, C., Després, B.: Trefftz discontinuous Galerkin method for Friedrichs systems with linear relaxation: application to the p1 model. Computational Methods in Applied Mathematics, 521–557 (2018)

  34. Pinchon, D., Hoggan, P.E.: Rotation matrices for real spherical harmonics: general rotations of atomic orbitals in space-fixed axes. J. Phys. A, Math. Theor. 40 (7), 1597–1610 (2007)

    MathSciNet  MATH  Google Scholar 

  35. Tezaur, I.K., Farhat, C.: A discontinuous enrichment method for the finite element solution of high Péclet advection-diffusion problems. Finite Elements in Analysis and Design 45(4), 238–250 (2009)

    MathSciNet  Google Scholar 

  36. Zienkiewicz, O.C.: Origins, milestones and directions of the finite element method? A personal view. Archives of Computational Methods in Engineering 2(1), 1–48 (1995)

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christophe Buet.

Additional information

Communicated by: Ilaria Perugia

Publisher’s note

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

Appendices

Appendix A: Time dependent solutions of equation (1)

We give some possible ways to get time dependent solutions to the PN model. These solutions can also be used as basis functions for the TDG method when considering space-time meshes. Many other time-dependent solutions can be constructed starting from [5, 10].

  • A first possibility is to consider one-dimensional solution under the form as follows:

    $$ \mathbf{v}(t,x) = \mathbf{q}(t,x) e^{\lambda x}, $$

    where \(\mathbf {q}(t,x) \in \mathbb {R}^{m}\) is polynomial vector in x and t. A concrete example is given in [33] [Proposition 4.2], for the case of the P1 model. Then, one can use the rotational invariance of the PN model and gets the following solutions:

    $$ \mathbf{v}(t,\mathbf{x}) = U_{\theta}\mathbf{q}(t,x \cos \theta + y \sin \theta) e^{\lambda (x \cos \theta + y \sin \theta)}. $$
    (29)

    Another possibility is to search directly for two-dimensional solutions under the form as follows:

    $$ \mathbf{v}(t,\mathbf{x}) = \mathbf{p}(t,\mathbf{x}) e^{\lambda (x \cos \theta + y \sin \theta)}, $$
    (30)

    where \(\mathbf {p}(t,\mathbf {x}) \in \mathbb {R}^{m}\) is polynomial vector in x, y ant t. Note that the functions obtained with (30) may differ from the functions (29). A complete example is given in [32], [Section 5.2.2] for the P1 model.

  • Another possibility is to consider time dependent solutions under the form as follows:

    $$ \mathbf{v}(t,\mathbf{x}) = \mathbf{g}(\mathbf{x})e^{\alpha t}, $$
    (31)

    with \(\alpha \in \mathbb {R}\). One can inject this solution in the PN model (1). One gets after removing the exponentials as follows:

    $$ \left( A_{1} \partial_{x} + A_{2} \partial_{y} + (R+ \varepsilon\alpha I_{m})\right) \mathbf{g}(\mathbf{x}) = \mathbf{0}, $$

    where Im is the identity matrix of \(\mathbb {R}^{m \times m}\). The function g(x) is very similar to the stationary solutions already calculated in Section 3.1. The matrix R is just replaced by the matrix \(\widetilde {R}:=R+\varepsilon \alpha I_{m}\).

    For example, if one takes α such that σa + εα > 0, then g(x) is one of the exponential solutions (10). In particular, if α > 0, the functions (31) naturally degenerate toward nontrivial time-dependent solutions when \(\sigma _{a} \rightarrow 0\). This is one asset of the functions (31).

Appendix B: A variational Trefftz discontinuous Galerkin method

The objective of this Appendix is to provide the basic material for the construction of the TDG discussed in this work. The notations are taken from [33]. We write v = (v1,...,vm)T where T denotes the transpose. The domain Ω is partitioned: let \(\mathcal {T}_{S,h}\) be a partition of ΩS into polyhedral subdomains ΩS,k. For a stationary problem, we set \(\mathcal {T}_{h} = \mathcal {T}_{S,h}\). For time-dependent problems, we consider \(N \in \mathbb {N}^{*}\) time steps [ti,ti+ 1], 0 = t0 < t1 < ... < tN = T. Then, \(\mathcal {T}_{h}\) is the partition of Ω made of the polyhedral subdomains \({\Omega }_{k^{n}}={\Omega }_{S,k} \times [t_{n},t_{n+1}]\); here, the index k is related to \(\mathcal {T}_{S,h}\) and n to the time step. In both cases, we use the general notation Ωk when designing a cell of \(\mathcal {T}_{h}\). With this notations, the time variable is treated as the space variables in the domain Ω. One must therefore be careful that \(\mathcal {T}_{h}\) denotes a spatial mesh for stationary model and a space-time mesh for time-dependent model.

The matrix R is constant in each cell. The broken Sobolev space is as follows:

$$ H^{1}(\mathcal{T}_{h}) := \{\mathbf{v} \in L^{2}({\Omega}), \ \mathbf{v}_{|{\Omega}_{k}} \in H^{1}({\Omega}_{k}) \ \forall {\Omega}_{k} \in \mathcal{T}_{h} \}. $$

In the following, we assume \(\mathbf {u} \in H^{1}(\mathcal {T}_{h})\). For convenience, we rewrite the model under the form \( L = {\sum }_{i} A_{i}\partial _{i} + R\) and Lu = 0. We consider the adjoint operator \( L^{*} = - {\sum }_{i} A_{i}\partial _{i} + R\). Multiplying the equation by \(\mathbf {v} \in H^{1}(\mathcal {T}_{h})\) and integrating gives on Ω as follows:

$$ \sum\limits_{k} {\int}_{{\Omega}_{k}} \mathbf{v}_{k}^{T} L \mathbf{u}_{k} = 0, $$
(32)

where \(\mathbf {v}_{k} = \mathbf {v}_{|{\Omega }_{k}}\), \(\mathbf {u}_{k} = \mathbf {u}_{|{\Omega }_{k}}\). Integrating by parts one gets the following:

$$ \sum\limits_{k} {\int}_{{\Omega}_{k}} \left( L^{*} \mathbf{v}_{k}\right)^{T} \mathbf{u}_{k} + \sum\limits_{k} {\int}_{\partial{\Omega}_{k}}\mathbf{v}_{k}^{T}M_{k} \mathbf{u}_{k} = 0, $$

where Ωk is the contour of the element Ωk with an outward unit normal \(\mathbf {n}_{k} = (n_{t},n_{x_{1}},...,n_{x_{d}})^{T}\), \(M = A_{0} n_{t} + {\sum }_{i} A_{i} n_{i}\) and Mk = M(nk(x)). Denoting Σkj = ΩkΩj (nk + nj = 0 on Σkj), one can writes the following:

$$ \begin{array}{@{}rcl@{}} \sum\limits_{k} {\int}_{{\Omega}_{k}} \left( L^{*} \mathbf{v}_{k}\right)^{T} \mathbf{u}_{k} &+& \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} \left( \mathbf{v}^{T} M \mathbf{u}\right)_{k} + \left( \mathbf{v}^{T} M \mathbf{u}\right)_{j} \\ &+& \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{+} \mathbf{u}_{k} = - \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{-} \mathbf{g} . \end{array} $$

For u satisfying the equation, the normal flux is continuous in a weak sense as follows:

$$ M_{k} \mathbf{u}_{k} = - M_{j} \mathbf{u}_{j} = f_{\text{kj}}(\mathbf{u}_{k},\mathbf{u}_{j}), \quad \text{on } {\Sigma}_{\text{kj}} $$
(33)

where fkj is a numerical flux on Σkj defined below. One has the following:

$$ \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (\mathbf{v}^{T} M \mathbf{u})_{k} + (\mathbf{v}^{T} M \mathbf{u})_{j} = \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (\mathbf{v}_{k} - \mathbf{v}_{j})^{T} f_{\text{kj}}(\mathbf{u}_{k},\mathbf{u}_{j}) . $$

Because M is symmetric one can decompose M under the form M = M+ + M where M+ is a non-negative matrix and M is a non-positive matrix. In the following, we will consider the upwind flux \(f_{\text {kj}}(\mathbf {u}_{k},\mathbf {u}_{j}) = M_{\text {kj}}^{+}\mathbf {u}_{k} + M_{\text {kj}}^{-}\mathbf {u}_{j},\) where \(M_{\text {kj}} = M_{k|{\Sigma }_{\text {kj}}}\). Finally, one has the following:

$$ \begin{array}{@{}rcl@{}} \sum\limits_{k} {\int}_{{\Omega}_{k}} \left( L^{*} \mathbf{v}_{k}\right)^{T} \mathbf{u}_{k} &+& \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (\mathbf{v}_{k} - \mathbf{v}_{j})^{T} (M_{\text{kj}}^{+} \mathbf{u}_{k} + M_{\text{kj}}^{-} \mathbf{u}_{j}) \end{array} $$
(34)
$$ \begin{array}{@{}rcl@{}} &+& \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{+} \mathbf{u}_{k} = -\sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{-} \mathbf{g} . \end{array} $$
(35)

We define the bilinear form \(a_{\text {DG}}:H^{1}(\mathcal {T}_{h}) \times H^{1}(\mathcal {T}_{h}) \rightarrow \mathbb {R}\) and the linear form \(l:H^{1}(\mathcal {T}_{h}) \rightarrow \mathbb {R}\) as follows:

$$ \begin{array}{@{}rcl@{}} a_{\text{DG}}(\mathbf{u},\mathbf{v}) &=& \sum\limits_{k} {\int}_{{\Omega}_{k}} (L^{*} \mathbf{v}_{k})^{T} \mathbf{u}_{k} + \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (\mathbf{v}_{k} - \mathbf{v}_{j})^{T} (M_{\text{kj}}^{+} \mathbf{u}_{k} + M_{\text{kj}}^{-} \mathbf{u}_{j})\\ &&+ \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{+} \mathbf{u}_{k} , \quad \mathbf{u},\mathbf{v} \in H^{1}(\mathcal{T}_{h}),\\ l(\mathbf{v}) &=& -\sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{-} \mathbf{g} , \quad \mathbf{v} \in H^{1}(\mathcal{T}_{h}). \end{array} $$
(36)

One can rewrite (34) as \(a_{\text {DG}}(\mathbf {u},\mathbf {v}) = l(\mathbf {v}), \ \forall \mathbf {v} \in H^{1}(\mathcal {T}_{h})\). The standard discontinuous Galerkin method is based on the use of polynomial basis functions [14, 31]. Define \(\mathbb {P}_{q}^{d}\), the space of polynomials of d variables, of total degree at most q and the broken polynomial space as follows:

$$ \mathbb{P}_{q}^{d}(\mathcal{T}_{h}) := \{\mathbf{v} \in L^{2}({\Omega}), \mathbf{v}_{|{\Omega}_{k}} \in \mathbb{P}_{q}^{d} \ \forall {\Omega}_{k} \in \mathcal{T}_{h} \} \subset H^{1}(\mathcal{T}_{h}). $$

Definition B.1

Assume \(P_{m}(\mathcal {T}_{h})\) is a finite subspace of \(H^{1}(\mathcal {T}_{h})\), for example, \(P_{m}(\mathcal {T}_{h})=\mathbb {P}_{q}^{d}(\mathcal {T}_{h})\). The standard upwind discontinuous Galerkin method for Friedrichs systems is formulated as follows:

$$ \left\{\begin{array}{ll} \text{find } \mathbf{u}_{h} \in P_{m}(\mathcal{T}_{h}) \text{ such that }\\ a_{\text{DG}}(\mathbf{u}_{h},\mathbf{v}_{h}) =l(\mathbf{v}_{h}), \quad \forall \mathbf{v}_{h} \in P_{m}(\mathcal{T}_{h}). \end{array}\right. $$
(37)

For Trefftz methods, one takes basis functions which are exact solutions to the equations in each cell as follows:

$$ V(\mathcal{T}_{h}) = \{\mathbf{v} \in H^{1}(\mathcal{T}_{h}), L \mathbf{v}_{k} = \mathbf{0} \quad \forall {\Omega}_{k} \in \mathcal{T}_{h}\} \subset H^{1}(\mathcal{T}_{h}). $$

The space \(V(\mathcal {T}_{h})\) is a genuine subspace of \(H^{1}(\mathcal {T}_{h})\) except in the case L = 0 which is of no interest. Starting from the bilinear form aDG from (36), one notices that the volume term can be written for all functions in \(V(\mathcal {T}_{h})\) as follows:

$$ {\int}_{{\Omega}_{k}} \left( L^{*} \mathbf{v}_{k}\right)^{T} \mathbf{u}_{k} = 2 {\int}_{{\Omega}_{k}} \mathbf{v}_{k}^{T} R\mathbf{u}_{k} , \quad \forall \mathbf{u},\mathbf{v} \in V(\mathcal{T}_{h}). $$
(38)

One can therefore define a bilinear form \(a_{T}:V(\mathcal {T}_{h}) \times V(\mathcal {T}_{h}) \rightarrow \mathbb {R}\) as follows:

$$ \begin{array}{@{}rcl@{}} a_{T}(\mathbf{u},\mathbf{v}) &=& \sum\limits_{k} 2 {\int}_{{\Omega}_{k}} \mathbf{v}_{k}^{T} R\mathbf{u}_{k} + \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (\mathbf{v}_{k} - \mathbf{v}_{j})^{T} (M_{\text{kj}}^{+} \mathbf{u}_{k} + M_{\text{kj}}^{-} \mathbf{u}_{j})\\ &&+ \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{+} \mathbf{u}_{k} , \quad \mathbf{u},\mathbf{v} \in V(\mathcal{T}_{h}). \end{array} $$
(39)

Thanks to an integration by part, one gets an equivalent formulation of the bilinear form aT(⋅,⋅) as follows:

$$ a_{T}(\mathbf{u},\mathbf{v}) = - \sum\limits_{k} \sum\limits_{j<k} {\int}_{{\Sigma}_{\text{kj}}} (M_{\text{kj}}^{-}\mathbf{v}_{k} + M_{kj}^{+} \mathbf{v}_{j})^{T} (\mathbf{u}_{k} - \mathbf{u}_{j}) -\!\! \sum\limits_{k} {\int}_{{\Sigma}_{\text{kk}}} \mathbf{v}_{k}^{T} M_{k}^{-} \mathbf{u}_{k} ,\!\!\!\!\!\! \quad \mathbf{u},\mathbf{v} \in\! V(\mathcal{T}_{h}). $$
(40)

The main advantage to the formulation (40) compare to (39) is that there is no volume term; thus, it can save some computational time and may be easier to code. Note also that in the formulation (40), the relaxation term R completely disappears. This is not a problem for the TDG method because the information about R will be contained in the basis functions. For the bilinear form \(l:V(\mathcal {T}_{h}) \rightarrow \mathbb {R}\), it is the same as in (36), that is, \(l(\mathbf {v}) = -{\sum }_{k} {\int \limits }_{{\Sigma }_{kk}} \mathbf {v}_{k}^{T} M_{k}^{-} \mathbf {g} \) for all \( \mathbf {v} \in V(\mathcal {T}_{h})\).

Definition B.2

Assume \(V_{m}(\mathcal {T}_{h})\) is a finite subspace of \(V(\mathcal {T}_{h})\). The TDG method is as follows:

$$ \left\{\begin{array}{ll} \text{find } \mathbf{u}_{h} \in V_{m}(\mathcal{T}_{h}) \text{ such that }\\ a_{T}(\mathbf{u}_{h},\mathbf{v}_{h}) = l(\mathbf{v}_{h}), \quad \forall \mathbf{v}_{h} \in V_{m}(\mathcal{T}_{h}). \end{array}\right. $$
(41)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Buet, C., Despres, B. & Morel, G. Trefftz discontinuous Galerkin basis functions for a class of Friedrichs systems coming from linear transport. Adv Comput Math 46, 41 (2020). https://doi.org/10.1007/s10444-020-09755-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10444-020-09755-5

Keywords

Mathematics Subject Classification (2010)

Navigation