Skip to main content
Log in

A Comparative Study from Spectral Analyses of High-Order Methods with Non-Constant Advection Velocities

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

Abstract

The response of numerical methods to flow properties dynamic changes is a key ingredient of numerical modeling calibration. In this context, a range of spectral analyses and canonical fluid mechanics problems are explored to compare the responses of the high-order spectral difference scheme and of the flux reconstruction methods. Spatial eigen-analysis (Hu et al. in J Comput Phys 151(2):921–946, 1999; Hu and Atkins in J Comput Phys 182(2):516–545, 2002), based on spatially evolving oscillations, is used to get useful insights for problems with inflow/outflow boundary conditions (i.e., non-periodic), while non-modal analysis (Fernandez et al. in Comput Methods Appl Mech Eng 346:43–62, 2019. https://doi.org/10.1016/j.cma.2018.11.027) focusses on the short-term dynamics of the discretised system. These two approaches are also extended to the general scalar conservation law with non-constant advection velocity. All these tools are used to obtain more insight about the expected behavior of the mentioned numerical methods for typical engineering applications. On this regard, despite the lack of a mathematical proof of stability in the non-linear case, the selected numerical schemes, for which an extensive literature is available for applications to Euler and Navier–Stokes equations, will be assumed to be stable. Findings are then verified considering one-dimensional linear advection, and two- and three-dimensional flows modeled with the compressible Euler equations. Implications in the accurate control of numerical dissipation for turbulent flows is also addressed. Within the context of high-order methods, this work complement the former temporal spectral analyses of the discontinuous Galerkin approach discretising the linear advection equation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29
Fig. 30
Fig. 31
Fig. 32
Fig. 33

Similar content being viewed by others

References

  1. Slotnick, J., Khodadoust, A., Alonso, J., Darmofal, D., Gropp, W., Lurie, E., Mavriplis, D.: CFD vision 2030 study: a path to revolutionary computational aerosciences. Technical Report, NASA Langley Research Center, Hampton, Virginia 23681-2199, NASA Contractor Report 218178 (Mar. 2014)

  2. Vincent, P.E., Jameson, A.: Facilitating the adoption of unstructured high-order methods amongst a wider community of fluid dynamicists. Math. Model. Nat. Phenomena 6(3), 97–140 (2011)

    Article  MathSciNet  Google Scholar 

  3. Fernandez, P., Nguyen, N.-C., Peraire, J.: On the ability of discontinuous Galerkin methods to simulate under-resolved turbulent flows. Preprint arXiv:1810.09435

  4. Fernandez, P., Nguyen, N., Peraire, J.: The hybridized discontinuous Galerkin method for implicit large-eddy simulation of transitional turbulent flows. J. Comput. Phys. 336, 308–329 (2017). https://doi.org/10.1016/j.jcp.2017.02.015

    Article  MathSciNet  MATH  Google Scholar 

  5. Fernandez, P., Nguyen, N.-C., Peraire, J.: Entropy-stable hybridized discontinuous Galerkin methods for the compressible Euler and Navier–Stokes equations. Preprint arXiv:1808.05066

  6. Beck, A.D., Bolemann, T., Flad, D., Frank, H., Gassner, G.J., Hindenlang, F., Munz, C.-D.: High-order discontinuous Galerkin spectral element methods for transitional and turbulent flow simulations. Int. J. Numer. Methods Fluids 76(8), 522–548 (2014). https://doi.org/10.1002/fld.3943

    Article  MathSciNet  Google Scholar 

  7. Flad, D., Gassner, G.: On the use of kinetic energy preserving dg-schemes for large eddy simulation. J. Comput. Phys. 350, 782–795 (2017). https://doi.org/10.1016/j.jcp.2017.09.004

    Article  MathSciNet  MATH  Google Scholar 

  8. Karniadakis, G., Sherwin, S.: Spectral/HP Element Methods for Computational Fluid Dynamics. Oxford University Press, Oxford (2013)

    MATH  Google Scholar 

  9. Lele, S.K.: Compact finite difference schemes with spectral-like resolution. J. Comput. Phys. 103(1), 16–42 (1992)

    Article  MathSciNet  Google Scholar 

  10. Bogey, C., Bailly, C.: A family of low dispersive and low dissipative explicit schemes for flow and noise computations. J. Comput. Phys. 194(1), 194–214 (2004)

    Article  Google Scholar 

  11. Van den Abeele, K., Lacor, C., Wang, Z.J.: On the stability and accuracy of the spectral difference method. J. Sci. Comput. 37(2), 162–188 (2008)

    Article  MathSciNet  Google Scholar 

  12. Vincent, P.E., Castonguay, P., Jameson, A.: Insights from von Neumann analysis of high-order flux reconstruction schemes. J. Comput. Phys. 230(22), 8134–8154 (2011)

    Article  MathSciNet  Google Scholar 

  13. Moura, R.C., Sherwin, S.J., Peiró, J.: Linear dispersion–diffusion analysis and its application to under-resolved turbulence simulations using discontinuous Galerkin spectral/HP methods. J. Comput. Phys. 298, 695–710 (2015)

    Article  MathSciNet  Google Scholar 

  14. Vanharen, J., Puigt, G., Vasseur, X., Boussuge, J.-F., Sagaut, P.: Revisiting the spectral analysis for high-order spectral discontinuous methods. J. Comput. Phys. 337, 379–402 (2017)

    Article  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

  16. Cockburn, B., Shu, C.: The local discontinuous Galerkin finite element method for convection–diffusion systems. SIAM J. Numer. Anal. 35, 2440–2463 (1998)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  18. Huynh, H.T.: A flux reconstruction approach to high-order schemes including discontinuous Galerkin methods. In: 18th AIAA Computational Fluid Dynamics Conference, p. 4079 (2007)

  19. Vincent, P.E., Castonguay, P., Jameson, A.: A new class of high-order energy stable flux reconstruction schemes. J. Sci. Comput. 47(1), 50–72 (2011)

    Article  MathSciNet  Google Scholar 

  20. Kopriva, D.A., Kolias, J.H.: A conservative staggered-grid Chebyshev multidomain method for compressible flows. J. Comput. Phys. 125(1), 244–261 (1996)

    Article  MathSciNet  Google Scholar 

  21. Liu, Y., Vinokur, M., Wang, Z.: Spectral difference method for unstructured grids I: basic formulation. J. Comput. Phys. 216(2), 780–801 (2006)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  23. Hu, F.Q., Atkins, H.L.: Eigensolution analysis of the discontinuous Galerkin method with nonuniform grids: I. One space dimension. J. Comput. Phys. 182(2), 516–545 (2002)

    Article  MathSciNet  Google Scholar 

  24. Mengaldo, G., Moura, R., Giralda, B., Peiró, J., Sherwin, S.: Spatial eigensolution analysis of discontinuous Galerkin schemes with practical insights for under-resolved computations and implicit LES. Comput. Fluids 169, 349–364 (2018)

    Article  MathSciNet  Google Scholar 

  25. Mengaldo, G., De Grazia, D., Moura, R.C., Sherwin, S.J.: Spatial eigensolution analysis of energy-stable flux reconstruction schemes and influence of the numerical flux on accuracy and robustness. J. Comput. Phys. 358, 1–20 (2018)

    Article  MathSciNet  Google Scholar 

  26. Moura, R., Aman, M., Peiró, J., Sherwin, S.: Spatial eigenanalysis of spectral/HP continuous Galerkin schemes and their stabilisation via DG-mimicking spectral vanishing viscosity: application to high Reynolds number flows. J. Comput. Phys. 406, 109112 (2019)

    Article  Google Scholar 

  27. Schmid, P.J.: Nonmodal stability theory. Annu. Rev. Fluid Mech. 39, 129–162 (2007)

    Article  MathSciNet  Google Scholar 

  28. Trefethen, L.N., Trefethen, A.E., Reddy, S.C., Driscoll, T.A.: Hydrodynamic stability without eigenvalues. Science 261(5121), 578–584 (1993)

    Article  MathSciNet  Google Scholar 

  29. Trefethen, L.N.: Pseudospectra of linear operators. SIAM Rev. 39(3), 383–406 (1997)

    Article  MathSciNet  Google Scholar 

  30. Manzanero, J., Rubio, G., Ferrer, E., Valero, E.: Dispersion-dissipation analysis for advection problems with nonconstant coefficients: applications to discontinuous Galerkin formulations. SIAM J. Sci. Comput. 40(2), A747–A768 (2018)

    Article  MathSciNet  Google Scholar 

  31. Manzanero, J., Rubio, G., Ferrer, E., Valero, E., Kopriva, D.A.: Insights on aliasing driven instabilities for advection equations with application to Gauss–Lobatto discontinuous Galerkin methods. J. Sci. Comput. 75(3), 1262–1281 (2017)

    Article  MathSciNet  Google Scholar 

  32. Jameson, A.: A proof of the stability of the spectral difference method for all orders of accuracy. J. Sci. Comput. 45(1), 348–358 (2010)

    Article  MathSciNet  Google Scholar 

  33. Jameson, A., Vincent, P., Castonguay, P.: On the non-linear stability of flux reconstruction schemes. J. Sci. Comput. 50(2), 434–445 (2012)

    Article  MathSciNet  Google Scholar 

  34. Fernandez, P., Moura, R.C., Mengaldo, G., Peraire, J.: Non-modal analysis of spectral element methods: towards accurate and robust large-eddy simulations. Comput. Methods Appl. Mech. Eng. 346, 43–62 (2019). https://doi.org/10.1016/j.cma.2018.11.027

    Article  MathSciNet  MATH  Google Scholar 

  35. Asthana, K., Jameson, A.: High-order flux reconstruction schemes with minimal dispersion and dissipation. J. Sci. Comput. 62(3), 913–944 (2015)

    Article  MathSciNet  Google Scholar 

  36. Liang, C., Premasuthan, S., Jameson, A.: High-order accurate simulation of low-Mach laminar flow past two side-by-side cylinders using spectral difference method. Comput. Struct. 87(11–12), 812–827 (2009)

    Article  Google Scholar 

  37. Liang, C., Jameson, A., Wang, Z.J.: Spectral difference method for compressible flow on unstructured grids with mixed elements. J. Comput. Phys. 228(8), 2847–2858 (2009)

    Article  Google Scholar 

  38. Cantwell, C., Moxey, D., Comerford, A., Bolis, A., Rocco, G., Mengaldo, G., Grazia, D.D., Yakovlev, S., Lombard, J.-E., Ekelschot, D., Jordi, B., Xu, H., Mohamied, Y., Eskilsson, C., Nelson, B., Vos, P., Biotto, C., Kirby, R., Sherwin, S.: Nektar++: an open-source spectral/HP element framework. Comput. Phys. Commun. 192, 205–219 (2015). https://doi.org/10.1016/j.cpc.2015.02.008

    Article  MATH  Google Scholar 

  39. Taylor, G.I., Green, A.E.: Mechanism of the production of small eddies from large ones. Proc. R. Soc. Lond. A 158(895), 499–521 (1937)

    Article  Google Scholar 

  40. Chapelier, J.-B., Lodato, G., Jameson, A.: A study on the numerical dissipation of the spectral difference method for freely decaying and wall-bounded turbulence. Comput. Fluids 139, 261–280 (2016). https://doi.org/10.1016/j.compfluid.2016.03.006

    Article  MathSciNet  MATH  Google Scholar 

  41. Manzanero, J., Ferrer, E., Rubio, G., Valero, E.: On the role of numerical dissipation in stabilising under-resolved turbulent simulations using discontinuous Galerkin methods (2018)

  42. Moura, R.C., Mengaldo, G., Peiró, J., Sherwin, S.J.: On the eddy-resolving capability of high-order discontinuous Galerkin approaches to implicit LES/under-resolved DNS of Euler turbulence. J. Comput. Phys. 330, 615–623 (2017)

    Article  MathSciNet  Google Scholar 

  43. Chapelier, J.-B., Lodato, G.: Optimal high-order Spectral Difference schemes for the computation of aeroacoustics and turbulence. AIAA Paper 2017-1228 (2017) 1–17, 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, Jan. 9–13 (2017). https://doi.org/10.2514/6.2017-1228

  44. Hu, F.Q., Hussaini, M., Rasetarinera, P.: An analysis of the discontinuous Galerkin method for wave propagation problems. J. Comput. Phys. 151(2), 921–946 (1999)

    Article  Google Scholar 

Download references

Acknowledgements

The use of the SD solver originally developed by Antony Jameson’s group at Stanford University is gratefully acknowledged. This work was granted access to the high-performance computing resources of CRIANN. The PhD scholarship of the first author is founded by the University of Rouen Normandie. Financial support to the second and third authors was provided by ANR under Grant No. ANR-18-CE05-0030.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Niccolò Tonicello.

Additional information

Publisher's Note

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

Appendices

Appendix

A: Flux Reconstruction Schemes

Collocation based nodal DG and SD have been widely used in the last few years due to their simplicity. The FR method [18, 19] provides a simple and general framework among which popular schemes like DG and SD can be recovered for linear advection problems. In this particular case, linear stability is ensured for a wide range of FR schemes [19], among which, DG-FR and SD-FR too.

The FR scheme, apart from being a robust and promising numerical scheme for aerodynamics simulations, is also a very useful tool to study a wide range of different high-order numerical schemes. Like nodal DG schemes, FR schemes exploit a high-order (nodal) polynomial basis to approximate the solution within each element of the computational domain, and like nodal DG schemes, inter-element continuity is not strongly enforced. However, unlike nodal DG schemes, FR methods are based solely on the governing system in a differential form. A description of the FR approach in one dimension is presented below.

For simplicity, let us consider a reference element \(\varOmega _{n}=\{\hat{x} | -1 \le \hat{x} \le 1\}\) in which the general one-dimensional conservation law is defined

$$\begin{aligned} \frac{\partial {u}}{\partial {t}}+\frac{\partial {(f(u))}}{\partial {\hat{x}}}=0. \end{aligned}$$
(33)

The FR approach to solve Eq. (33) within the standard element can be described in five stages. The first stage involves the representation of the approximated solution \(\hat{u}\) in terms of a nodal basis function:

$$\begin{aligned} \hat{u}(\hat{x})=\sum _{i=0}^{k}u_{i}l_{i}(\hat{x}), \end{aligned}$$
(34)

where \(l_{i}\) are the Lagrange polynomials defined on a certain set of points \(\hat{x}_{i}\) (with i=0 to k) called solution points. Notice that this implies a polynomial representation of \(\hat{u}\) of degree k. The second stage consists in the construction of a degree k polynomial flux \(\hat{f}^{D}=\hat{f}^{D}(\hat{x},t)\) which is based on the same collocation basis as \(\hat{u}\):

$$\begin{aligned} \hat{f}^{D}(\hat{x})=\sum _{i=0}^{k}f_{i}^{D}l_{i}(\hat{x}), \end{aligned}$$
(35)

where flux nodal values can be easily evaluated directly from the approximated solution. The third stage involves the evaluation of the approximate solution at the end points of the standard element (namely, \(\hat{u}(\pm 1)\)). Subsequently, this information is used to compute the numerical flux at the left \(\hat{f}^{I}_{L}\) and right \(\hat{f}^{I}_{R}\) interfaces. The fourth stage involves the construction of the degree \(k+1\) polynomial \(\hat{f}\), by adding a correction flux \(\hat{f}^{C}=\hat{f}^{C}(\hat{x},t)\) of degree \(k+1\) to \(\hat{f}^{D}\), such that their sum recovers the numerical interface flux at \(\hat{x}=\pm 1\), yet in some sense follows \(\hat{f}^{D}\) within the interior of \(\varOmega _{s}\). This is the key step of FR schemes since an higher order polynomial (degree \(k+1\)) is summed to the interpolated one, leading to a difference in approximation order that plays a fundamental role in the whole procedure.

In order to define \(\hat{f}^{C}\) such that it satisfies the above requirements, consider first two degree \(k+1\) correction functions \(g_{L}=g_{L}(\hat{x})\) and \(g_{R}=g_{R}(\hat{x})\) approximating zero (in some sense) within \(\varOmega _{s}\) and satisfying the following conditions:

$$\begin{aligned}&g_{L}(-1)=1, \quad g_{L}(1)=0, \end{aligned}$$
(36)
$$\begin{aligned}&g_{R}(-1)=0, \quad g_{R}(1)=1, \end{aligned}$$
(37)

and

$$\begin{aligned} g_{L}(\hat{x})=g_{R}(-\hat{x}). \end{aligned}$$
(38)

Then a suitable formulation for \(\hat{f}^{C}\) can now be written as

$$\begin{aligned} \hat{f}^{C}=(\hat{f}^{I}_{L}-\hat{f}^{D}_{L})g_{L}+(\hat{f}^{I}_{R}-\hat{f}^{D}_{R})g_{R}, \end{aligned}$$
(39)

where \(\hat{f}^{D}_{L}=\hat{f}^{D}(-1,t)\) and \(\hat{f}^{D}_{R}=\hat{f}^{D}(1,t)\). Using this expression, the degree \(k+1\) approximate total flux \(\hat{f}\) within \(\varOmega _{s}\) can be constructed from the discontinuous and correction fluxes as follows

$$\begin{aligned} \hat{f}=\hat{f}^{D}+\hat{f}^{C}=\hat{f}^{D}+(\hat{f}^{I}_{L}-\hat{f}^{D}_{L})g_{L}+(\hat{f}^{I}_{R}-\hat{f}^{D}_{R})g_{R}. \end{aligned}$$
(40)

The fifth and final stage of the FR approach involves the evaluation of the flux divergence at each solution point \(\hat{x}_{i}\) using the expression

$$\begin{aligned} \frac{\mathrm {d}{\hat{f}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i})=\sum _{j=0}^{k}\hat{f}^{D}_{j}\frac{\mathrm {d}{l_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i})+(\hat{f}^{I}_{L}-\hat{f}^{D}_{L})\frac{\mathrm {d}{g_{L}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i})+(\hat{f}^{I}_{R}-\hat{f}^{D}_{R})\frac{\mathrm {d}{g_{R}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i}). \end{aligned}$$
(41)

These values can then be used to advance \(\hat{u}\) in time via a suitable temporal discretisation of the following semi-discrete expression

$$\begin{aligned} \frac{\mathrm {d}{\hat{u}_{i}}}{\mathrm {d}{\text {t}}}=-\frac{\mathrm {d}{\hat{f}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i}). \end{aligned}$$
(42)

The nature of a particular FR scheme depends solely on three factors:

  1. 1.

    the location of the solution points \(\hat{x}_{i}\);

  2. 2.

    the methodology for calculating the interface fluxes \(\hat{f}^{I}_{L}\) and \(\hat{f}^{I}_{R}\);

  3. 3.

    the form of the flux correction functions \(g_{L}\) and \(g_{R}\).

It has been shown by Huynh [18] that, for particular choices of the correction functions, it is possible to recover some well known collocation methods (among which, nodal DG and SD) for the linear advection equation. In particular, Vincent–Catonguay–Jameson–Huynh Flux Reconstruction schemes can be obtained with the following choice of correction functions:

$$\begin{aligned} \begin{aligned} g_{L}(x)&=\frac{(-1)^{k}}{2} \bigg [ L_{k}(x) - \bigg ( \frac{\eta _{k}L_{k-1}(x)+L_{k+1}(x)}{1+\eta _{k}} \bigg )\bigg ], \\ g_{R}(x)&=\frac{1}{2} \bigg [ L_{k}(x) + \bigg ( \frac{\eta _{k}L_{k-1}(x)+L_{k+1}(x)}{1+\eta _{k}} \bigg )\bigg ], \\ \end{aligned} \end{aligned}$$
(43)

where

$$\begin{aligned} \eta _{k}=\frac{c(2k+1)(a_{k}k!)^{2}}{2} \quad \text {and}\quad a_{k}=\frac{(2k)!}{2^{k}(k!)^{2}}. \end{aligned}$$
(44)
Fig. 34
figure 34

FR correction functions (\(g_{L}\)) of degree \(k=5\) recovering DG and SD schemes: solid line, DG-FR; dotted line, SD-FR

In the above relations, \(L_{k}\) is a Legendre polynomial of degree k, and c is a free scalar parameter that must lie within the range

$$\begin{aligned} c_{-}<c<c_{\infty },\quad \text {with}\quad c_{-}=\frac{-2}{(2k+1)(a_{k}k!)^{2}} \quad \text {and} \quad c_{\infty }=\infty . \end{aligned}$$

For the particular choice of the parameter c nodal DG and SD schemes are recovered. In particular:

$$\begin{aligned} c_{DG}=0 \quad \text {and} \quad c_{SD}=\frac{2(k+1)}{(2k+1)k(a_{k}k!)^{2}}. \end{aligned}$$

In this way, the respective correction functions for DG and SD simplify in the following expressions:

$$\begin{aligned} \begin{aligned} g_{L}^{DG}(x)&=\frac{(-1)^{k}}{2}(L_{k}(x) - L_{k+1}(x)),&g_{R}^{DG}(x)&=\frac{(-1)^{k}}{2}(L_{k}(x) + L_{k+1}(x)), \\ g_{L}^{SD}(x)&=\frac{(-1)^{k}}{2}(1-x)L_{k}(x),&g_{R}^{SD}(x)&=\frac{1}{2}(1+x)L_{k}(x). \end{aligned} \end{aligned}$$
(45)

An example of correction functions recovering DG and SD schemes, namely, the DG-FR and SD-FR schemes, respectively, is shown in Fig. 34.

In the particular case of the linear advection equation with unitary advection velocity, the FR scheme can be expressed in a matrix form as

$$\begin{aligned} \frac{\mathrm {d}{\hat{\mathbf{u }}}}{\mathrm {d}{\text {t}}} = -2 \mathbf{D} \hat{\mathbf{u }}-(2\hat{f}^{I}_{L}-2 \mathbf{l} ^{T} \hat{\mathbf{u }})\mathbf{g} ^{L}-(2\hat{f}^{I}_{R}-2 \mathbf{r} ^{T} \hat{\mathbf{u }})\mathbf{g} ^{R}, \end{aligned}$$
(46)

where \(\hat{\mathbf{u }}_{i}=\hat{u}(\hat{x}_{i})\), \(\mathbf{D} _{ij}=\frac{\mathrm {d}{l_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i})\), \(\mathbf{g} ^{L/R}_{i}=\frac{\mathrm {d}{g_{L/R}}}{\mathrm {d}{\hat{x}}}(\hat{x}_{i})\) and \(\mathbf{r} _{i}=l_{i}(1)\), \(\mathbf{l} _{i}=l_{i}(-1)\).

B: The Spectral Difference Scheme

The SD method [20,21,22], like FR schemes, solves the strong form of the differential equation using piecewise continuous functions as approximation space. Consequently, the solution is assumed to be discontinuous at elements interface. In order to have a consistent discretisation, the solution is interpolated using a polynomial of degree k while the flux, which is connected to the conservative variables via a divergence operator, is approximated with a polynomial of degree \(k+1\). The most important ingredient of the SD discretisation is the definition of two different set of points: solution and flux points. The numerical solution is defined on the nodes \(x_i^s\) with i=0 to k. Fluxes, instead, are defined on a different set of nodes \(x_{i}^f\), with i=0 to \(k+1\), among which boundary points are included. It shall be noted that, in the present study, the solution points are set as the Gauss–Legendre points of order \(k+1\), a sensible choice to minimise aliasing errors in the nonlinear case while defining a well conditioned basis set for the solution interpolation [33], whereas the flux points are set as the Gauss–Legendre points of order k plus the two end points -1 and 1 to ensure linear stability [32]. An example of solution and flux points for a polynomial approximation of degree \(k=3\) is shown in Fig. 35.

Fig. 35
figure 35

Solution and flux points of SD discretisation in the reference element (\(k=3\))

As in the FR scheme, the solution is approximated with a polynomial of degree k:

$$\begin{aligned} \hat{u}(\hat{x})=\sum _{i=0}^{k}u_{i}l^{s}_{i}(\hat{x}). \end{aligned}$$
(47)

At this point the FR scheme would use an interpolated flux (of degree k) on the same set of points and subsequently add a correction flux (of degree \(k+1\)) defined on element extrema as well. In the SD scheme, instead, the values of the solution are extrapolated at the flux points

$$\begin{aligned} \hat{u}(\hat{x}^{f}_{j})=\sum _{i=0}^{k}u_{i}l^{s}_{i}(\hat{x}^{f}_{j}), \qquad j=0,\ldots ,k+1, \end{aligned}$$
(48)

and then used to compute fluxes on the same collocation basis:

$$\begin{aligned} f_{j}= \hat{f}(\hat{x}^f_{j})=\hat{f}(\hat{u}(\hat{x}^{f}_{j})). \end{aligned}$$
(49)

Then, a continuous flux polynomial of degree \(k+1\) is constructed, by Lagrange interpolation, using the fluxes evaluated from the interpolated solution at the interior flux points and the numerical fluxes at the element interfaces:

$$\begin{aligned} \hat{f}(\hat{x})= \hat{f}^{I}_{L}l^f_{0}(\hat{x}) +\sum _{j=1}^{k}f_{j}l^f_{j}(\hat{x}) +\hat{f}^{I}_{R}l^f_{k+1}(\hat{x}). \end{aligned}$$
(50)

In other words, the interpolated values of the flux at elements extrema are substituted by the interface numerical fluxes \(\hat{f}^{I}_{L}\) and \(\hat{f}^{I}_{R}\). Finally, the flux divergence is evaluated at the solution points,

$$\begin{aligned} \frac{\mathrm {d}{\hat{f}}}{\mathrm {d}{\hat{x}}}(\hat{x}^s_{i})= \hat{f}^{I}_{L}\frac{\mathrm {d}{l^f_{0}}}{\mathrm {d}{\hat{x}}}(\hat{x}^{s}_{i}) +\sum _{j=1}^{k}f_{j}\frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x}^{s}_{i}) +\hat{f}^{I}_{R}\frac{\mathrm {d}{l^f_{k+1}}}{\mathrm {d}{\hat{x}}}(\hat{x}^{s}_{i}), \end{aligned}$$
(51)

and, as in the FR schemes, the numerical solution can be advanced in time using a suitable time integration scheme discretising the following equation:

$$\begin{aligned} \frac{\mathrm {d}{\hat{u}}}{\mathrm {d}{\text {t}}}=-\frac{\mathrm {d}{\hat{f}}}{\mathrm {d}{\hat{x}}}(\hat{x}^s_{i}). \end{aligned}$$
(52)

In the case of linear advection equation, Eq. (52) can be written in matrix form as

$$\begin{aligned} \frac{\mathrm {d}{\hat{\mathbf{u }}_{n}}}{\mathrm {d}{\text {t}}} =-\,2 \mathbf{D} (\hat{f}_{L}\mathbb {1}^{0}+\mathbf{M} \hat{\mathbf{u }}_{n}+\hat{f}_{R}\mathbb {1}^{k+1}), \end{aligned}$$
(53)

where

$$\begin{aligned} \mathbf{D} _{ij}=\frac{\mathrm {d}{l^{f}_{j}(\hat{x}^{s}_{i})}}{\mathrm {d}{\hat{x}}}, \qquad \mathbf{M} _{ij}= {\left\{ \begin{array}{ll} 0 &{} \quad \mathrm {for} \quad i=0, \\ 0 &{} \quad \mathrm {for} \quad i=k+1, \\ l^{s}_{j}(\hat{x}^{f}_{i}) &{} \quad \text {otherwise,} \end{array}\right. } \end{aligned}$$
(54)

\(\mathbb {1}^{0}_{j}=\delta _{j0}\) and \(\mathbb {1}^{k+1}_{j}=\delta _{j(k+1)}\) for \(j=0\) to \(k+1\). In the above formalism, when the superscript s is used, the relevant index ranges from 0 to k, whereas, when the superscript f is adopted, the index goes from 0 to \(k+1\). Hence, \(\mathbf{D} \in \mathbb {R}^{(k+1) \times (k+2)}\) and \(\mathbf{M} \in \mathbb {R}^{(k+2) \times (k+1)}\) resulting in a local linear system of dimensions \((k+1) \times (k+1)\). The subscript \((\cdot )_{n}\), instead, denotes the element numbering. Notice how a clear distinction between the interior differential operator and the evaluation of the interface flux is evident in both Eqs. (46) and (53).

The core of the method is hidden inside the definition of the matrix \(\mathbf{M} \), for which both nodal sets are needed, while the interface information is contained in the \(\hat{f}_{L/R}\) terms. In analogy with FR schemes a SD method is completely defined by:

  1. 1.

    the location of both solution and flux points;

  2. 2.

    the expression of interface fluxes \(\hat{f}_{L/R}\).

Notice that the particular choice of FR correction functions is now substituted by the selection of the flux points set. The analogy between these choices is of key importance to prove the equivalency of SD-FR and SD schemes for the linear advection equation. The interpolation operator between the two set of nodes, in fact, implicitly defines the correction function of the corresponding FR recovering scheme. Nodes coordinates are somehow a useful degree of freedom that can be modified in order to increase accuracy and/or stability. In fact, some recent work has been done in optimised nodes location for SD methods [43].

C: SD Scheme Versus SD-FR Scheme

In order to highlight the connection between the SD method and the corresponding FR recovering scheme, it is interesting to rewrite Eq. (51) at the general location \(\hat{x}\) as

$$\begin{aligned} \frac{\mathrm {d}{\hat{f}}}{\mathrm {d}{\hat{x}}}(\hat{x})= \sum _{j=0}^{k+1}f_{j}\frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x}) +(\hat{f}^{I}_{L}-f_{0})\frac{\mathrm {d}{l^f_{0}}}{\mathrm {d}{\hat{x}}}(\hat{x})+(\hat{f}^{I}_{R}-f_{k+1})\frac{\mathrm {d}{l^f_{k+1}}}{\mathrm {d}{\hat{x}}}(\hat{x}), \end{aligned}$$
(55)

where the contributions of the interpolated fluxes at \(\hat{x}=\pm 1\) have been added and subtracted. Observing this formulation, it is clearly evident the analogy with Eq. (41). In particular, it is easy to show that the first summation of the two equations is exactly the same for a linear flux function. In fact, considering the linear advection equation with unitary advection velocity, namely \(f(\hat{x})\equiv u(\hat{x})\):

$$\begin{aligned} \begin{aligned} \sum _{j=0}^{k+1}f_{j}\frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x})&=\sum _{j=0}^{k+1} \bigg (\sum _{i=0}^{k} u_{i}l_{i}^{s}(\hat{x}_{j}^{f})\bigg ) \frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x})=\sum _{i=0}^{k}u_{i}\sum _{j=0}^{k+1}l_{i}^{s}(\hat{x}_{j}^{f})\frac{\mathrm {d}{l_{j}^{f}}}{\mathrm {d}{\hat{x}}}(\hat{x})=\\&=\sum _{i=0}^{k}u_{i}\frac{\mathrm {d}{}}{\mathrm {d}{\hat{x}}}\bigg (\sum _{j=0}^{k+1}l_{i}^{s}(\hat{x}_{j}^{f})l_{j}^{f}(\hat{x})\bigg )=\sum _{i=0}^{k}u_{i}\frac{\mathrm {d}{l_{i}^{s}}}{\mathrm {d}{\hat{x}}}(\hat{x}). \\ \end{aligned} \end{aligned}$$
(56)

The last equality has been obtained by noting that the Lagrange interpolation over \(k+2\) points of the Lagrange polynomial of order \(k+1\), namely \(l_{i}^{s}(\hat{x})\), coincides with this last.

Notice that, if a non-constant advection velocity is considered and the flux is defined as \(\hat{f}=a(\hat{x})u(\hat{x})\), following the SD case:

$$\begin{aligned} \begin{aligned} \sum _{j=0}^{k+1}f_{j}\frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x})&=\sum _{j=0}^{k+1} a(\hat{x}_{j}^{f})\bigg (\sum _{i=0}^{k} u_{i}l_{i}^{s}(\hat{x}_{j}^{f})\bigg ) \frac{\mathrm {d}{l^f_{j}}}{\mathrm {d}{\hat{x}}}(\hat{x})=\sum _{i=0}^{k}u_{i}\sum _{j=0}^{k+1} a(\hat{x}_{j}^{f}) l_{i}^{s}(\hat{x}_{j}^{f})\frac{\mathrm {d}{l_{j}^{f}}}{\mathrm {d}{\hat{x}}}(\hat{x})=\\&=\sum _{i=0}^{k}u_{i}\frac{\mathrm {d}{}}{\mathrm {d}{\hat{x}}}\bigg (\underbrace{\sum _{j=0}^{k+1} a(\hat{x}_{j}^{f})l_{i}^{s}(\hat{x}_{j}^{f})l_{j}^{f}(\hat{x})}_{\mathcal {I}^f[a(\hat{x})l_{i}^{s}(\hat{x})]}\bigg )=\sum _{i=0}^{k}u_{i}\frac{\mathrm {d}{(\mathcal {I}_l[a(\hat{x})l_{i}^{s}(\hat{x})])}}{\mathrm {d}{\hat{x}}}(\hat{x}), \\ \end{aligned} \end{aligned}$$
(57)

where \(\mathcal {I}^f[\cdot ]\) denotes the Lagrange interpolation operation over the \(k+2\) flux points. Equation (57) evidently differs from what would be obtained from the FR procedure:

$$\begin{aligned} \sum _{i=0}^{k}f_{i}\frac{\mathrm {d}{l_{i}^{s}}}{\mathrm {d}{\hat{x}}}(\hat{x})=\sum _{i=0}^{k}a(\hat{x}_{i}^{s})u_{i}\frac{\mathrm {d}{l_{i}^{s}}}{\mathrm {d}{\hat{x}}}(\hat{x}). \end{aligned}$$
(58)

Clearly the two right-hand sides are equal only when the advection velocity is constant.

Considering now the second part of Eq. (55), it is evident that, in order for the FR method to recover the SD scheme, \(g_{L}(\hat{x})=l^{f}_{0}(\hat{x})\) and \(g_{R}(\hat{x})=l^{f}_{k+1}(\hat{x})\). So, just changing polynomial basis, the classical expressions for \(g_{L}\) and \(g_{R}\) are recovered:

$$\begin{aligned} g_{L}^{SD}(\hat{x})=\frac{(-1)^{k}}{2}(1-\hat{x})L_{k}(\hat{x}) \quad \text {and} \quad g_{R}^{SD}(\hat{x})=\frac{1}{2}(1+\hat{x})L_{k}(\hat{x}), \end{aligned}$$
(59)

where \(L_{k}\) is the Legendre polynomial of degree k.

It is worthwhile stressing that the definition of flux points is a key ingredient in the SD method. In fact, considering the SD discretisation of the linear advection equation, according to the particular choice of flux nodes, a correction function in the FR framework is implicitly defined. This link will be of fundamental importance in the proof of equivalence between the two schemes under such conditions.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tonicello, N., Lodato, G. & Vervisch, L. A Comparative Study from Spectral Analyses of High-Order Methods with Non-Constant Advection Velocities. J Sci Comput 87, 65 (2021). https://doi.org/10.1007/s10915-021-01484-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10915-021-01484-1

Keywords

Navigation