Skip to main content
Log in

Third-order sparse grid generalized spectral elements on hexagonal cells for uniform-speed advection in a plane

  • Original Paper
  • Published:
Meteorology and Atmospheric Physics Aims and scope Submit manuscript

Abstract

This paper investigates sparse grids on a hexagonal cell structure using a Local-Galerkin method (LGM) or generalized spectral element method (SEM). Such methods allow sparse grids to be used, known as serendipity grids in square cells. This means that not all points of the full grid are used. Using a high-order polynomial, some points of each cell are eliminated in the discretization, and thus saving Central Processing Unit (CPU) time. Here a sparse SEM scheme is proposed for hexagonal cells. It uses a representation of fields by second-order polynomials and achieves third-order accuracy. As SEM, LGM is strictly local for explicit time integration. This makes LGM more suitable for multiprocessing computers compared with classical Galerkin methods. The computer time depends on the possible timestep and program implementation. Assuming that these do not change when going to a sparse grid, the potential saving of computer time due to sparseness is 1:2. The projected CPU saving in 3-D from sparseness is by a factor of 3:8. A new spectral procedure is used in this paper, called the implied spectral equation (ISE). This procedure allows for some collocation points to use any finite difference scheme of high order and the time derivatives of other spectral coefficients are implied.

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

Similar content being viewed by others

References

  • Ahlberg JH, Nilson EN, Walsh JL (1967) The theory of splines and their application. Academic, New York

    Google Scholar 

  • Baumgardner JR, Frederickson PO (1985) Icosahedral discretization of the two-sphere. SIAM J Numer Anal 22:1107–1115

    Article  Google Scholar 

  • Cockburn B, Shu CW (2001) RungeKutta discontinuous Galerkin methods for convection-dominated problems. J Sci Comput 16(3):173–261

    Article  Google Scholar 

  • Cote J, Beland M, Staniforth A (1983) A spectral element shallow water model on spherical geodesic grids. Mon Wea Rev 111:1189–1207

    Article  Google Scholar 

  • Durran D (2010) Numerical methods of fluid dynamics: with applications to geophysics. Springer, New York

    Book  Google Scholar 

  • Giraldo FX (2001) A spectral element shallow water model on spherical geodesic grids. Int J Num Method Fluids 35:869–901

    Article  Google Scholar 

  • Kalnay E, Bayliss A, Storch J (1977) The 4th order GISS Model of the global atmosphere. Beitrage Phys Atmos 50:299–311

    Google Scholar 

  • Li J, Li J, Zheng J, Zhu J, Fang F, Pain CC, Steppeler J, Navon IM, Xiao H (2018) Performance of adaptive unstructured mesh modelling in idealized advection cases over steep Terrains. Atmosphere 9:444

    Article  Google Scholar 

  • Marras S, Kelly JF, Moragues M, Müller A, Kopera MA, Vázquez M, Giraldo FX, Houzeaux G, Jorba O (2016) A review of element-based Galerkin methods for numerical weather prediction: finite elements, spectral elements, and discontinuous Galerkin. Arch Comput Method E 23(4):673–722

    Article  Google Scholar 

  • Navon IM, Alperson Z (1978) Application of fourth-order finite differences to a baroclinic model of the atmosphere. Arch Meterol Geophyd Biol Serie A 27:1–19

    Google Scholar 

  • Peixoto PS, Barros SR (2013) Analysis of grid imprinting on geodesic spherical icosahedral grids. J Comput Phys 237:61–78

    Article  Google Scholar 

  • Rancic M, Purser RJ, Mesinger F (1996) A global shallow water model using an expanded spherical cube: gnomonic versus conformal coordinate. Q J R Meteorol Soc 122:959–982

    Article  Google Scholar 

  • Ringler TD, Heikes RH, Randall DA (2000) Modeling the atmospheric general circulation using a spherical geodesic grid: a new class of dynamical cores. Mon Wea Rev 128:2471–2485

    Article  Google Scholar 

  • Sadourny R (1972) Conservative finite-difference approximations of the primitive equations on quasi-uniform spherical grids. Mon Wea Rev 100:136–144

    Article  Google Scholar 

  • Satoh M, Masuno T, Tomita H, Miura H, Nasuno T, Iga S (2008) Nonhydrostatic icosahedaral atmosphericmodel (NICAM) for global cloud resolving simulations. J Comput Phys 227:3486–3514

    Article  Google Scholar 

  • Skamarock WC, Klemp JB (2008) A time-split nonhydrostatic atmospheric model for weather research and forecasting applications. J Comput Phys 227(7):3465–3485

    Article  Google Scholar 

  • Skamarock WC, Klemp JB, Duda MG, Fowler LD, Park SH (2012) A multiscale nonhydrostatic atmospheric model using centroidal voronoi tesselations and C-Grid staggering. Mon Wea Rev 140:3090–3105

    Article  Google Scholar 

  • Staniforth A, Thuburn J (2012) Horizontal grids for global weather and climate prediction models: a review. Q J R Meteorol Soc 138(662):1–26

    Article  Google Scholar 

  • Steppeler J (1976) The application of the second and third degree methods. J Comp Phys 22:295–318

    Article  Google Scholar 

  • Steppeler J (1979) Difference schemes with uniform second and third order accuracy and reduced smoothing. J Comput Phys 31:428–449

    Google Scholar 

  • Steppeler J (1987) Galerkin and finite element methods in numerical weather prediction. Duemmler, Bonn

    Google Scholar 

  • Steppeler J, Klemp JB (2017) Advection on cut-cell grids for an idealized mountain of constant slope. Mon Wea Rev 145:1765–1777

    Article  Google Scholar 

  • Steppeler J, Prohl P (1996) Application of finite volume methods to atmospheric models. Beitr Phys Atmos 69:297–306

    Google Scholar 

  • Steppeler J, Ripodas P, Thomas S (2008) Third order finite difference schemes on isocahedral-type grids on the sphere. Mon Wea Rev 136:2683–2698

    Article  Google Scholar 

  • Steppeler J, Li J, Fang F, Zhu J, Ullrich PA (2019) o3o3: a variant of spectral elements with a regular collocation grid. Mon Wea Rev 147(6):2067–2082. https://doi.org/10.1175/MWR-D-18-0288.1

    Article  Google Scholar 

  • Taylor M, Tribbia J, Iskandarani M (1997) The spectral element method for the shallow water equations on the sphere. J Comput Phys 130:92–108

    Article  Google Scholar 

  • Tomita H, Tsugawa M, Satoh M, Goto K (2001) Shallow water model on a modified isosahedral geodesic grid by using spring dynamics. J Comp Phys 174:579–613

    Article  Google Scholar 

  • Weller H, Thuburn J, Cotter CJ (2012) Computational modes and grid imprinting on five quasi-uniform spherical C grids. Mon Wea Rev 140(8):2734–2755

    Article  Google Scholar 

  • Williamson DL (1968) Integrations of the barotropic vorticity equation on a spherical geodesic grid. Tellus 20:643–653

    Article  Google Scholar 

  • Williamson DL (2007) The evolution of dynamical cores for global atmospheric models. J Meteorol Soc Jpn Ser II 85:241–269

    Article  Google Scholar 

  • Zängl G, Reinert D, Rípodas P, Baldauf M (2015) The ICON (ICOsahedral Non-hydrostatic) modelling framework of DWD and MPI-M: description of the non-hydrostatic dynamical core. Q J R Meteorol Soc 141(687):563–579

    Article  Google Scholar 

Download references

Acknowledgements

No authors reported any potential conflicts of interest. The paper was made possible by a number of research visits of the first author at NCAR, financed by NCAR. The city of Bad Orb supported this cooperation by providing office space. Dr. J. Li acknowledges the support of the National Natural Science Foundation of China (Grant No. 41905093), the China Postdoctoral Science Foundation Funded Project (Grant No. 2016M601101) and China Scholarship Council (No. 201904910136). Dr. F. Fang acknowledges the support of the Innovation of the Chinese Academy of Sciences International Partnership Project (Grant No. Y56601M601) and the EPSRC grant: Managing Air for Green Inner Cities (MAGIC) (EP/N010221/1).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Li.

Additional information

Responsible Editor: C. Simmer.

Publisher's Note

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

Appendix 1

Appendix 1

1.1 A system of indices for second-order hexagonal cells creating a compact grid representation and the 1-D grids for plane waves

This appendix introduces an alternative system of indices to that used in Sects. 2 and 3. The grid is shown in Fig. 12 and it differs from Fig. 1 by assigning different numbers to the points. Each point has two indices i and k, where i indicates the hexagon and k indicated the position of a point within the hexagon. k = 0 indicates the center point of the hexagonal cell. The index i can be a structured index i = (i′, j′) (i′, j′ = 0, 1, 2, 3, …) or unstructured. The structured case is shown in Fig. 12, where the grids and their center points are indicated by the double index (i’, j’). The advantage of a structured grid notation is that neighboring grids are always known. For example, in Fig. 12 the cell (i′, j′+ 1) is always above the cell (i′, j′). Our definition can also be applied to the unstructured grid case. In the unstructured case a list is needed indicating the six neighbouring hexagons to the hexagon i. The grid based compact grid pointed out in the following can be used with structures and unstructured index i, where Fig. 12 refers to the structured case i = (i′, j′).

Fig. 12
figure 12

As Fig. 1 except that with new indices creating a compact grid. Black points and points with thick circles are dynamic points, those with thin circles are diagnostic points. The points shown in black are the points for compact representation. They do not belong to any other hexagon and each dynamic point appears as a compact point of some hexagons

Let the side length of the hexagon be s. For the corner nodes \( {\mathbf{r}}_{i,k} \) (k = 1, 3, 5, 7, 9, 11) we have

$$ {\mathbf{r}}_{i,k} = {\mathbf{r}}_{i,0} + \left( {s \cdot \cos \left( {\pi + \left( {k - 1} \right)\tfrac{\pi }{3}} \right),s \cdot { \sin }\left( {\pi + \left( {k - 1} \right)\tfrac{\pi }{3}} \right)} \right) $$
(27)

The edge nodes k = 2, 4, 6, 8, 12 are obtained by

$$ {\mathbf{r}}_{i,k} = {\mathbf{r}}_{i,k - 1} + \tfrac{\sqrt 3 }{2}\left( {s \cdot \cos \tfrac{\pi }{3},s \cdot { \sin }\tfrac{\pi }{3}} \right) $$
(28)

For the diagnostic inner points k = 13 to 18, we have

$$ \begin{aligned} {\mathbf{r}}_{i,k} & = {\mathbf{r}}_{i,0} + \tfrac{1}{2}\left( {s \cdot \cos \left( {\pi + \left( {2k - 25 - 1} \right)\tfrac{\pi }{6}} \right),s \cdot { \sin }\left( {\pi + \left( {2k - 25 - 1} \right)\tfrac{\pi }{6}} \right)} \right) \\ & = {\mathbf{r}}_{i,0} - \tfrac{1}{2}\left( {s \cdot \cos \left( {\tfrac{k - 1}{3}\pi } \right),s \cdot { \sin }\left( {\tfrac{k - 1}{3}\pi } \right)} \right) \\ \end{aligned} $$
(29)

The full grid consists for each i of all points k = 0 to 18. The points numbering 13, 14, 15, 16, 17, 18 are unused in the sparse grid. The sparse grid collocation therefore consists of the points i, k (k = 0 to 12) which are the mid-point k = 0 and points on the boundary. The sparse grid is redundant, as all points with k > 0 belong to more than one hexagon. For example in Fig. 12, the point 9i,j, is also point 1i′+1,j′+1 and 5i,j′+1. The point 10i,j belongs only to one other hexagon: i′, j′+ 1. These redundant points of the collocation grid can be used to represent discontinuous functions.

The index is defined with respect to the orthogonal basis \( {\mathbf{i}}_{s} \) and \( {\mathbf{j}}_{s} \) shown in Fig. 12. This method of indexing is suitable for model areas being approximately square. A rhomboidal area can be achieved by changing the definition of the basis \( {\mathbf{i}}_{s} \) and \( {\mathbf{j}}_{s} \) shown in Fig. 12.

It was shown by Steppeler (1979) that the anisotropy of the square grid can lead to the deformation of structures, such as the destruction of the symmetry of a circular wave. It may be useful to show the connection to 1-D discretizations. In Fig. 13, two vectors \( {\mathbf{n}}_{1} \) and \( {\mathbf{n}}_{2} \) are given and for each of these directions grid lines are shown in Fig. 13b, c which meet hexagonal grid points and thus determine a 1-D grid. Figure 13 shows the lines and the 1-D grids created by them. \( {\mathbf{n}}_{1} \) and \( {\mathbf{n}}_{2} \) form a 30° angle. The situation is similar as with square grids, where two 1-D grids are generated by lines forming a 45° angle. For square grids, the implied grids have grid lengths of dx in x direction and for the 45° line the grid length \( dx' = \tfrac{dx}{\sqrt 2 } \). For the square the grids seen by a plane wave in the direction of either x or the diagonal are all the same, as each line determines the same grid.

Fig. 13
figure 13

Two examples of lines resulting in 1-D grids for two directions. The combined grid which is seen by a plane wave in the directions of the arrows \( {\mathbf{n}}_{1} = \left( {\cos 60^\circ ,\sin 60^\circ } \right) = \left( {\tfrac{1}{2},\tfrac{\sqrt 3 }{2}} \right) \) and \( {\mathbf{n}}_{2} = \left( {\cos 30^\circ ,\sin 30^\circ } \right) = \left( {\tfrac{\sqrt 3 }{2},\tfrac{1}{2}} \right) \) is shown in (d) and (e)

For 1-D grids in the direction of \( {\mathbf{n}}_{1} \) generated by hexagonal grids, three types of irregular grid appear, shown in Fig. 13b. A plane wave in the directions \( {\mathbf{n}}_{1} \) or \( {\mathbf{n}}_{2} \) sees one of the two grids (Fig. 13d, e). They have the average resolution dx′ (\( \tfrac{3}{8}s \) for Fig. 13d and \( \tfrac{\sqrt 3 }{4}s \) for Fig. 13e). This is of higher resolution than each of the original grids shown in Fig. 13b, c. It has also a higher resolution than the underlying regular square grid of resolution \( \tfrac{1}{2}s \). For the gradients of the fluxes at corner points in \( {\mathbf{n}}_{1} \) or \( {\mathbf{n}}_{2} \) direction, the o2o3 scheme proposed in this paper uses one-sided derivatives and averages them over the three directions shown in Fig. 5. For a structure being smooth in the direction vertical to \( {\mathbf{n}}_{1} \) or \( {\mathbf{n}}_{2} \), the derivative at one point uses points forming a different grid for each of the lines indicated in Fig. 13. Thus, the resolution of a plane wave is that of the 1-D grids shown in Fig. 13d, e. The irregularity of the plane wave grids is of a small scale which means that while some points are near to each other, they have large differences in grid length. The pattern of small and larger grid lengths repeats itself. These grids have a higher resolution than the plane wave grids for the square.

It was shown by Steppeler and Klemp (2017) for the example of centered differences that such small-scale irregularity gives a quality of simulation corresponding to the average grid length. It can be seen that the average grid for plane waves is smaller than the grid used on the hexagonal sides. The values of the average grids given above should be compared to those corresponding to the square grid. For the square grid, we obtain \( \tfrac{s}{2} \) and \( \tfrac{s}{2\sqrt 2 } \) for the averaged resolutions in two directions. The relation of the two grids is 1.17 for the hexagon and 1.47 for the square. These numbers are a measure of the isotropy of the grid, which is higher for hexagons. The maximum plane wave grid length is also smaller for the hexagon (0.433 s) than for the square (0.5 s). The maximum of the two grids for different directions of plane waves is the resolution of the scheme. A small resolution in just one direction is not useful except for the case that the nature of the solution requires a deformed grid.

The indices of the inherent 1-D grids and their computation from the hexagonal indices can be seen from Fig. 13b, d or c, e. The hexagonal version of the grid for the method o3o3 (Steppeler et al. 2019) is obtained from the o2o3 hexagonal grid by having four instead three points on each edge and the center point no longer carries an amplitude.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Steppeler, J., Li, J., Fang, F. et al. Third-order sparse grid generalized spectral elements on hexagonal cells for uniform-speed advection in a plane. Meteorol Atmos Phys 132, 703–719 (2020). https://doi.org/10.1007/s00703-019-00718-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00703-019-00718-0

Navigation