1. Introduction
In many physical problems the spatial scale of the solution varies significantly over the domain. For example, a tropical cyclone is a relatively small-scale vortex embedded in a larger-scale flow, and accurate prediction of the vortex track may require resolving the flow both within and around the storm. Solving such problems numerically often requires variable resolution to accurately resolve the small-scale features without compromising the overall efficiency. Conventional nested-grid techniques (Kurihara et al. 1979; Berger and Oliger 1984) achieve local mesh refinement by superimposing nested uniform grids of varying mesh sizes. Such methods have been widely used for many years (e.g., Kurihara and Bender 1980; Juang and Hoke 1992;Skamarock and Klemp 1993). The success of conventional nested-grid tropical cyclone models such as the GFDL (Kurihara et al. 1998) and VICBAR (DeMaria et al. 1992) models demonstrates the value of this approach.
In contrast to nested-grid models, multigrid methods (Brandt 1977) use multiple grids of different mesh size covering the same region. The goal here is to solve the discretized equations with optimum efficiency: using simple relaxation on each grid to smooth the error on the scale of that grid produces the solution (on the finest grid) to the level of truncation error in computational work proportional to the number of unknowns (usually in the work of just a few relaxation sweeps). In addition to speeding up the solution process, multigrid processing provides accurate truncation error estimates, which can be used to determine where to refine (or coarsen) the grids in a local refinement scheme or to achieve higher-order accuracy via extrapolation. For an introduction to multigrid methods, see Briggs et al. (2000) or Fulton et al. (1986).
While the multigrid literature has many examples of combining multigrid processing with nested grids (e.g., Ciesielski et al. 1986; Bai and Brandt 1987; Saleh 1994;Ruge et al. 1995), in nested-grid meteorological models the interaction between the solution on different grid levels—where they overlap—has in general not been fully exploited. Indeed, some of the complexity of the VICBAR model is due to the fact that the computational grids are specifically not allowed to overlap.
The purpose of this paper is to demonstrate the potential of adaptive multigrid methods for time-dependent problems requiring local mesh refinement by describing an adaptive multigrid tropical cyclone track model. In the tradition of the barotropic models SANBAR (Sanders et al. 1975) and VICBAR, we refer to this multigrid barotropic model as MUDBAR. Our primary focus here is on the numerical method: describing the adaptive multigrid approach in detail and documenting its performance. Indeed, the MUDBAR model currently has the simplest possible dynamics (nondivergent barotropic equation) and no physics. Nevertheless, the model is fast and accurate, and could be useful for operational forecasting if augmented with boundary data from a global model and an appropriate initialization scheme. In particular, it is fast enough for real-time ensemble forecasting.
The remainder of the paper is organized as follows. Section 2 describes the model equations and solution method. The grid adaptation algorithm is presented in section 3. Numerical results appear in section 4, and our conclusions are summarized in section 5. An appendix gives the details of the grid size adjustment algorithm used in the mesh refinement scheme.
2. Model description
In this section we briefly describe the governing equations, discretization, grid structure, and solution method of the MUDBAR model. A preliminary version of the model was described in Fulton (1997), to which the reader is referred for some of the more basic details.
a. Governing equations
b. Discretization
The time discretization uses the classical fourth-order Runge–Kutta (RK4) scheme; this is highly accurate, allows relatively large time steps for stability, and—since it is a one-step scheme—is easy to implement and has no spurious computational modes. This scheme involves four stages per time step, each of which predicts a new value for ζh via (4) and solves (5) for the corresponding ψh.
c. Grid structure and local time stepping
The above discretizations are embedded in an adaptive method that includes local refinements in both space and time, superimposing nested uniform grids with different mesh sizes to adapt the resolution near the storm. The base grid G1 with mesh size h1 covers the entire computational domain, while successively finer patches Gl with mesh sizes hl (l = 2, . . . , n) cover smaller nested areas as shown in the example in Fig. 1. We use the mesh ratio hl−1/hl = 2 (to facilitate multigrid processing), and require the patches to be strictly nested (fine grid contained in the interior of the next coarser grid) and aligned (patch boundaries coincide with coarse-grid lines). For simplicity, we allow only one region of refinement (i.e., one grid patch per mesh size), but this is not a fundamental limitation of the method.
The time stepping algorithm uses local time stepping, that is, smaller time steps on finer grids; this algorithm is similar to that used in most nested-grid models (e.g., Berger and Oliger 1984; DeMaria et al. 1992). Specifically, with two grids (coarse and fine) one full time step is executed as follows:
one step (length Δt) on the coarse grid;
two steps (length Δt/2) on the fine grid, using boundary values interpolated from the coarse grid in space (ζh linearly and ψh by cubic interpolation) and time (linear interpolation); and
transfer the fine-grid solution to the coarse grid where they overlap (ψh by injection and ζh by full weighting).
d. Multigrid solution for streamfunction
To solve (5) efficiently for the streamfunction ψh on any computational grid Gl we use a multigrid method. This uses point Gauss–Seidel relaxation with red–black ordering, full weighting of residuals, bilinear interpolation of corrections, and the full multigrid (FMG) algorithm with one V(1, 1)-cycle per level and bicubic initial interpolation. As these elements are all standard, the reader is referred to Fulton et al. (1986) and the references therein for details. The resulting algorithm solves accurately for the streamfunction, producing residuals smaller than the truncation error; the computational work is comparable to about eight relaxation sweeps on the finest grid and, overall, accounts for about two-thirds of the total execution time of the model.1 Numerical results show that using more sweeps per cycle (or more cycles) produces smaller residuals but no significant improvement in the track error and, thus, is not worth the extra computational expense.
When used in the context of the adaptive grid structure described above, the standard multigrid approach is modified in several ways that deserve mention. First, with local time stepping, the next coarser grid Gl−1 may be at a different time t (e.g., at the end of time steps 2, 3, and 6 in Fig. 2). In this situation the main computational grids cannot be used for multigrid processing. Instead, for each computational grid Gl we use an additional set of coarse grids with mesh sizes 2hl, 4hl, 8hl, etc., covering the same domain as Gl (Fulton 1997). These “local coarse grids” may be used in two ways. The Berger–Oliger (BO) method uses them at all times, resulting in only one-way interaction between the grids:the fine grid uses boundary values from the coarse grid (step 2 in the algorithm above), but its effect is felt on the coarse grid only through the solution transfer in the overlap region at the end of the time step (step 3 above). In contrast, the multigrid (MG) method uses the full computational grid Gl−1 instead of the first local coarse grid during multigrid cycling in the final RK4 stage of the second fine-grid time step (e.g., at the end of time steps 4, 5, and 7 in Fig. 2). This results in automatic two-way interaction between the computational grids, since fine-grid information can affect the entire coarse grid immediately through the relaxation process. Preliminary results (Fulton 1997) showed that the differences between these two methods are generally small; further experimentation (see section 4d) shows some improvement with the MG approach, so that is used in all other results presented here.
3. The grid adaptation algorithm
The preliminary version of the model described in Fulton (1997) used movable grids, that is, patches with sizes fixed in advance were simply moved when necessary to keep them approximately centered on the vortex. The results indicated that some choices of patch sizes substantially increased the accuracy or efficiency (relative to using a single uniform grid), while others led to significantly less improvement. Since the optimal combination of patch sizes will depend on both the vortex and the surrounding flow and may vary as the solution evolves, choosing fixed patch sizes “manually” is less than ideal: a fully automatic algorithm would be advantageous. This section details such an algorithm; we refer to the grids it chooses as adaptive grids.
a. Basic regrid strategy
The MUDBAR model regrids, that is, chooses patch sizes and locations adaptively, by estimating the truncation error and placing the grid patches where the truncation error is large. The regridding algorithm is embedded in the local time stepping algorithm of section 2c as follows (see Fig. 3). Before starting a time step on a given grid with mesh size h—here considered to be the “coarse grid”—the truncation error is estimated (see section 3b). Points where the truncation error is large (see section 3c) are flagged for refinement, as indicated by × in Fig. 3a. Second, after the time step is complete the truncation error is again estimated and large values are flagged for refinement, as shown in Fig. 3b. Third, the smallest acceptable patches at the beginning and end of the step are determined (dotted and dashed outlines, respectively, in Fig. 3), and a suitable fine-grid patch with mesh size h/2 is chosen to cover both of these (solid outlines in Fig. 3b). Finally, that patch is “filled” with values of the vorticity ζh/2 and streamfunction ψh/2. The patch so constructed is then used for the two fine-grid time steps that coincide with the single time step originally taken on the coarse grid.
This process is repeated recursively; that is, during each of the two fine-grid steps a size and location are chosen for the next finer patch (if needed), proceeding to increasingly finer grids until either further refinement is not needed or a specified maximum number of grids is reached. This recursion is illustrated in Fig. 2, where the solid circles denote times at which the truncation error is estimated. For example, the truncation error on level l = 2 is computed before and after time step 5 (at t = Δt/2 and t = Δt, respectively) and used to determine the location and size of the next finer patch l = 3 for time steps 6 and 7 (from t = Δt/2 to t = Δt). Since grid patches are chosen “from scratch” at each time step, refinements are produced locally in both space and time:if at any time a new level of refinement is needed (or a current level is no longer needed), this is automatically detected and the grid patch is created (or deleted) as needed.
To fill a given patch, the vorticity ζh/2 is set by copying values from the previous version of that grid (if any) where they coincide and using bilinear interpolation from the coarse grid where the fine grid is new; at t = 0 the specified initial vorticity is used instead. The streamfunction ψh/2 is set likewise except that bicubic interpolation is used; if the patch is new (e.g., at t = 0), then ψh/2 is obtained instead by solving (5) with boundary values interpolated (cubically) from the coarse grid. During model initialization at t = 0, adaptive grids are constructed by starting with a specified base grid and constructing patches as dictated by the truncation error as outlined above. If the first patch covers most of the base grid (i.e., the base grid resolution is inadequate), this patch is extended to the full domain to replace the base grid and the process is repeated. Once the initial grids are set—and thus the fine-grid information has influenced the solution on all coarser levels—a second pass through the grid adaptation algorithm is made to take advantage of that fine-grid information.
Even though new grids are chosen for each time step on each level, the regridding process adds surprisingly little overhead to the execution time. All calculations required for regridding (estimating the truncation error, choosing the grid, and filling it by copying and/or interpolation) typically amount to less than 2% of the total execution time of the model. This slight overhead is more than made up for by the fact that the grid chosen at each step on each level is precisely the size needed for the entire step, and no larger. Indeed, a simpler scheme based on regridding on all levels only at the end of each base-grid time step regrids much less often but must choose larger patches (since they must be adequate for several fine-grid time steps) and thus takes more time (typically 20%–30% more).
The model also includes an option for using movable grids, that is, fixed patch sizes specified in advance. This differs from the preliminary version described in Fulton (1997) in that the patch locations are now chosen for an interval in time, rather than an instant. This is done by finding the vortex center at the beginning and end of the coarse-grid time step and centering the fine-grid patch for the corresponding two fine-grid time steps halfway between these two points. Except for this choice of patch location (and the fixed size); the regrid strategy with movable grids is identical to that described above for adaptive grids.
b. Truncation error estimates
c. Patch location and size
Once the points where the truncation error on the coarse grid (mesh spacing h) is large—in the sense of (14)—are located, the boundaries of the fine-grid patch (mesh spacing h/2) must be determined. For simplicity, the model currently uses only rectangular grids (and only one grid per value of h, so only one region is refined), so we can simply determine the smallest rectangles enclosing the large truncation errors before and after the step, as shown in Fig. 3. The tentative boundary of the patch for the full time step is simply the smallest rectangle enclosing both of these (i.e., their convex hull). This boundary may then be adjusted by adding a “buffer” of several grid intervals (usually two) to give some flexibility when the truncation error is highly localized. Finally, the patch size may be increased (or possibly slightly decreased) if needed to achieve good coarsenability, that is, so that the numbers of grid intervals in both directions contain many factors of 2 so the coarsest local coarse grid will be small, thus increasing the efficiency of solving for the streamfunction on the patch. The algorithm employed here uses simple work estimates as detailed in the appendix. In any case, strict nesting of the patch is enforced.
4. Results
This section summarizes the results of many model runs designed to explore the performance of the model for various situations.
a. Initial conditions
b. Sample solution
Figure 4 illustrates the solution of the model for the weak hurricane case. The left-hand panels of Fig. 4 show the streamfunction ψ and the right-hand panels show the corresponding vorticity ζ. For this run, the base grid has size 64 × 64 (h1 = 64 km), the exchange rate is λ = 1000, and the grid refinement is limited to three patches, so the finest mesh used is h4 = 8 km. The boundaries of the grid patches chosen by the adaptive algorithm are shown in the figures. It can be seen that the patch sizes tend to grow slightly during the model run; the apparent slight bias toward refining the grids behind the vortex is likely due to the larger truncation errors in the wake that are characteristic of centered finite differences. This effect could probably be lessened by including a small amount of dissipation; however, the model run shows that no dissipation is needed to achieve reasonable performance. In particular, no problems are observed at the grid interfaces.4
c. Accuracy and efficiency
To measure the accuracy of the model in the absence of an analytical solution, we compare the tropical cyclone track to that obtained in a high-resolution reference run with uniform resolution h = 4 km. The track error is defined as the distance between the location of the vortex center (vorticity maximum) and that of the reference run, with the center located by quadratic interpolation using four points surrounding the vorticity maximum on the finest grid (Smith et al. 1990); the mean forecast error is computed over a 72-h model run (using trapezoidal quadrature with a step size of 1 h). Previous experiments (Fulton 1997) showed that with uniform grids the forecast error is O(h2) as expected; with local refinement, the error (as a function of the finest mesh size) is somewhat larger, but still decreases at approximately the proper rate. Here we focus instead on the trade-off between accuracy and efficiency, measuring computational work by the CPU time required for the model run (on a Sun Ultra60 workstation with one 360-MHz processor).
Figure 5 shows the performance of the model for a large number of model runs for the weak hurricane case. Each point shows the mean forecast error obtained for a single run, plotted as a function of the computer time used for that run. The points plotted with plus signs are for uniform-resolution model runs (i.e., a base grid with no finer grid patches) with the indicated resolution. The line joining these points indicates a baseline (uniform grid) performance. The isolated points to the left of this line are for model runs with various combinations of movable grids (i.e., fixed size patches), using essentially every reasonable combination of patches with mesh sizes from h = 128 km to h = 4 km. The points plotted with × are for the adaptive model. These are connected by lines according to the finest mesh size allowed in the run; different points on each line correspond to different values of the exchange rate λ from 104 to 101. As the execution time increases monotonically with λ, points on the left side (small time) correspond to large λ, and points on the right side (large time) correspond to small λ.
Figure 5 shows that using local mesh refinement can save a factor of 2 to 50 or more in execution time, depending on the choice of grids used. Some combinations of movable grids work very well; others lead to relatively little improvement. The adaptive grid results show that the errors generally decrease as λ decreases (and CPU time increases), as suggested by the refinement criterion (14). However, there are exceptions. For large λ the error is sometimes unexpectedly low (e.g., the first few points on the left for the h ≥ 16 km curve); these cases probably represent fortuitous combinations of grids obtained in the region before the asymptotic convergence rate O(h2) on which the criterion (14) is based sets in. Similarly, for small λ the convergence “stalls” (e.g., the last few points on the right for the h ≥ 8 km curve); in such cases, the error is as small as possible with the finest mesh size allowed, and obtaining increased accuracy requires allowing finer grids, rather than using smaller λ. Overall, the performance is not particularly sensitive to λ, with values in the range λ ≈ 200–1000 giving good results. The adaptive algorithm usually chooses reasonable grids, although not always as good as the best choices of movable grids shown.
Figure 6 shows analogous results for the strong hurricane case. Since the vortex is both smaller and stronger, errors for a given amount of work are generally larger than in the previous case. Again the adaptive multigrid method chooses reasonable grids; in some cases the savings of execution time (compared to using uniform resolution) approach a factor of 100. Since the vortex is small, adaptive runs limited to h ≥ 16 km and h ≥ 8 km produce little improvement with decreasing exchange rate λ; in order to obtain smaller errors one must allow finer grids. The apparently erratic behavior for the h ≥ 4 km curve is due to two points with particularly good combinations of grids that gave smaller errors than would be expected by the general trend.
These results show that the accuracy and efficiency of the adaptive model depends on two factors: the exchange rate λ and the finest mesh size h allowed. Indeed, one could eliminate the mesh size as a factor by simply allowing very fine meshes (e.g., h = 1 km) and leaving it for the algorithm to decide whether or not to actually use grids this fine. For example, in Fig. 5 the adaptive grid curves for h ≥ 4 km and h ≥ 8 km start out identical (for large λ), since the grid with h = 4 km is not needed until higher accuracy is requested (smaller λ). It would be ideal if there were a direct relation between λ and the resulting error, but that appears to be problem dependent. However, this leaves the adaptive model with only one parameter to adjust experimentally to get the desired combination of accuracy and efficiency; in contrast, with movable grids, there are many combinations of patch sizes from which to choose, and the model performance depends heavily on that choice.
d. Effect of two-way interaction
Figure 8 demonstrates the effect of including two-way interaction between the computational grids. Each point represents the difference in mean forecast error between the BO method and the MG method as described in section 2d. The points are plotted as a function of the error produced by the MG method for many runs with different choices of movable grids. The differences between the accuracy of the two methods are small—and can go either way—but when one method is noticeably better (e.g., with error at least one kilometer less), it is usually the multigrid method, which includes two-way interaction.
e. A barotropically unstable vortex
The adaptive multigrid method might be especially advantageous for modeling a vortex that not only moves but has small-scale structure that evolves with time. While the nondivergent barotropic model employed here is incapable of representing the baroclinic effects and convection related to tropical cyclone intensification, it can represent the basic physical process of potential vorticity (PV) mixing. Thus, we consider the case of a barotropically unstable vortex that develops small-scale asymmetries due to chaotic nonlinear PV mixing and evolves toward a smooth symmetric vortex as it moves.
As an initial condition we use the “PV ring” vortex of Schubert et al. (1999) embedded in the zonal current (16). All parameters have the values given in that paper, except for the diffusion parameter ν for which we use the larger value 300 m2 s−1, corresponding to an e-folding time of 338 s for a wavelength of 2 km. For this case we run the model with spherical geometry using the Mercator projection (1). We present results of two model runs: one using movable grids and the other using adaptive grid patches (with exchange rate λ = 20). In both cases we start with the base grid size 128 × 128 with mesh spacing h = 32 km and corresponding time step Δt = 20 min. The movable grid run uses six square patches with side lengths ranging from 1/4 to 1/32 of the domain size. The adaptive run is limited to six adaptive grid patches; however, the model rejects the base grid as inadequate and consequently uses the base grid mesh spacing h = 16 km and up to five adaptive grids. Thus, for both runs the model domain is 4096 km square and the finest mesh spacing is h = 0.5 km.
Figure 9 shows the details of the solution computed using movable grids (left panels) and adaptive grids (right panels). Only a region 256 km square centered on the vortex is shown. Since the vorticity is not always maximum at the center of the ring, we define the vortex center as the vorticity centroid (taken over the region where ζ > f, to eliminate bias from the environmental vorticity). These results are quite similar to those of Schubert et al. (1999) up to at least t = 6 h, verifying that the characteristic wavenumber-4 pattern observed in that study is due to the dynamics, and is not an artifact of the periodicity required for the spectral model used in that study. At later times, the solutions are qualitatively similar but differ in detail, as expected due to the chaotic nature of the PV mixing. Comparing the two solutions, we find that the details of the small-scale structure are quite similar; also, the vortex tracks (not shown) differ by only about 7 km over the course of the 72-h model run. However, the run using adaptive grids was six times faster (about 16 min on a Sun Ultra60), since as the solution evolves toward a smooth symmetric vortex the finest grid patches are no longer needed and are automatically discarded.
5. Conclusions and remarks
We have used the problem of tropical cyclone track prediction to investigate the potential of the adaptive multigrid method. Combining nesting of uniform grids, multigrid processing, and adaptive mesh refinement based on truncation error estimates, the method provides a seamless discretization with enhanced resolution precisely where it is needed. Numerical results from the nondivergent barotropic model MUDBAR show the method performs well, generally reducing the execution time by an order of magnitude or more (relative to using uniform grids) and choosing reasonable grids with a minimum of user intervention. Compared to conventional nested-grid models, the treatment of grid interfaces is particularly simple (although that simplicity may be more a result of using simple dynamics rather than the numerical method). These results are encouraging and suggest that adaptive multigrid methods may be useful in other problems requiring time-dependent local mesh refinement.
While the model as it stands is fast and accurate, there is much room for improvement. We are currently working on two modifications that should substantially improve the inherent accuracy, namely, converting the model to the shallow-water equations and implementing fourth-order space differencing. To be useful for actual prediction, the model should be supplied with time-dependent boundary data from a global model and an appropriate initialization scheme. Given its speed and accuracy, the resulting model could be especially appropriate for ensemble forecasting techniques. For example, hundreds of model runs with slightly different initial conditions could be run in less than a half an hour (even on a PC) to investigate the sensitivity of the track to uncertainties in the initial data.
The numerical method presented here may in fact be overkill: for tropical cyclone track prediction, discretization errors may be less important than observational uncertainties or incomplete representations of the physics. Indeed, since the discretization is the only source of error considered here, our numerical results show far more accuracy than could be hoped for in operational prediction. Nevertheless, the adaptive multigrid approach allows one to effectively remove discretization error from the problem, and does so in a natural, robust, and efficient manner.
Acknowledgments
The helpful comments and advice of Mark DeMaria are gratefully acknowledged. Three anonymous reviewers made suggestions that significantly improved this paper. Mark Loftis wrote the original (nonadaptive) version of the model, and Brittany Mitchell, Nicole Burgess, and Hu Miao helped with the testing. This work was supported by the Office of Naval Research under Grants N00014-98-1-0103 and N00014-98-1-0368.
REFERENCES
Arakawa, A., 1966: Computational design for long-term numerical integration of the equations of fluid motion: Two-dimensional incompressible flow. Part I. J. Comput. Phys.,1, 119–143.
Bai, D., and A. Brandt, 1987: Local mesh refinement multilevel techniques. SIAM J. Sci. Stat. Comput.,8, 109–134.
Berger, M. J., and J. Oliger, 1984: Adaptive mesh refinement for hyperbolic partial differential equations. J. Comput. Phys.,53, 484–512.
Bernert, C., 1997: τ-extrapolation—Theoretical foundation, numerical experiment, and application to Navier–Stokes equations. SIAM J. Sci. Comput.,18, 460–478.
Brandt, A., 1977: Multi-level adaptive solutions to boundary-value problems. Math. Comput.,31, 333–390.
Briggs, W. L., V. E. Henson, and S. F. McCormich, 2000: A Multigrid Tutorial. 2d ed. Society for Industrial and Applied Mathematics, 193 pp.
Ciesielski, P. E., S. R. Fulton, and W. H. Schubert, 1986: Multigrid solution of an elliptic boundary value problem from tropical cyclone theory. Mon. Wea. Rev.,114, 797–807.
DeMaria, M., 1985: Tropical cyclone motion in a nondivergent barotropic model. Mon. Wea. Rev.,113, 1199–1210.
——, S. D. Aberson, K. V. Ooyama, and S. J. Lord, 1992: A nested spectral model for hurricane track forecasting. Mon. Wea. Rev.,120, 1628–1643.
Fulton, S. R., 1989: Multigrid solution of the semigeostrophic invertibility relation. Mon. Wea. Rev.,117, 2059–2066.
——, 1997: A comparison of multilevel adaptive methods for hurricane track prediction. Electron. Trans. Numer. Anal.,6, 120–132.
——, P. E. Ciesielski, and W. H. Schubert, 1986: Multigrid methods for elliptic problems: A review. Mon. Wea. Rev.,114, 943–959.
Guinn, T. A., and W. H. Schubert, 1993: Hurricane spiral bands. J. Atmos. Sci.,50, 3380–3403.
Juang, H.-M. H., and J. E. Hoke, 1992: Application of fourth-order finite differencing to the NMC nested grid model. Mon. Wea. Rev.,120, 1767–1782.
Kurihara, Y. M., and M. A. Bender, 1980: Use of a movable nested-mesh model for tracking a small vortex. Mon. Wea. Rev.,108, 1792–1809.
——, G. J. Tripoli, and M. A. Bender, 1979: Design of a movable nested-mesh primitive equation model. Mon. Wea. Rev.,107, 239–249.
——, R. E. Tuleya, and M. A. Bender, 1998: The GFDL hurricane prediction system and its performance in the 1995 hurricane season. Mon. Wea. Rev.,126, 1306–1322.
Oliger, J., and A. Sundström, 1978: Theoretical and practical aspects of some initial boundary value problems in fluid dynamics. SIAM J. Appl. Math.,35, 419–446.
Ruge, J. W., S. F. McCormick, and S. Y. K. Yee, 1995: Multilevel adaptive methods for semi-implicit solution of shallow-water equations on a sphere. Mon. Wea. Rev.,123, 2197–2205.
Saleh, A. M., 1994: Self-adaptive multilevel methods for fluid flow problems. Ph.D. dissertation, Clarkson University, 109 pp. [Available from Dept. of Mathematics and Computer Science, Clarkson University, Potsdam, NY 13699-5815.].
Sanders, F., A. C. Pike, and J. P. Gaertner, 1975: A barotropic model for operational prediction of tracks of tropical storms. J. Appl. Meteor.,14, 265–280.
Schubert, W. H., M. T. Montgomery, R. K. Taft, T. A. Guinn, S. R. Fulton, J. P. Kossin, and J. P. Edwards, 1999: Polygonal eyewalls, asymmetric eye contraction, and potential vorticity mixing in hurricanes. J. Atmos. Sci.,56, 1197–1223.
Skamarock, W. C., 1989: Truncation error estimates for refinement criteria in nested and adaptive models. Mon. Wea. Rev.,117, 872–976.
——, and J. B. Klemp, 1993: Adaptive grid refinement for two-dimensional and three-dimensional nonhydrostatic atmospheric flow. Mon. Wea. Rev.,121, 788–804.
Smith, R. K., W. Ulrich, and G. Dietachmayer, 1990: A numerical study of tropical cyclone motion using a barotropic model. I: The role of vortex asymmetries. Quart. J. Roy. Meteor. Soc.,116, 337–362.
APPENDIX
Grid Size Adjustment
Once the desired grid size is determined based on the truncation error as described in section 3c, it may be adjusted to ensure good coarsenability. For example, if the desired grid is 38 × 28 (grid intervals in x and y), this can be coarsened only once (to 19 × 14) for multigrid processing; since the coarsest grid is large, the multigrid solver is inefficient. In such a situation it will pay to increase the grid size slightly, for example, to 40 × 28 (which can be coarsened twice down to 10 × 7) or even to 40 × 32 (which can be coarsened three times down to 5 × 4). We do this as follows. Starting with the desired size Nx × Ny (grid intervals in x and y),
1) search for a grid slightly larger than desired by increasing Nx and/or Ny so they have at least one more factor of 2 in common, thus allowing at least one more coarse grid for multigrid processing, and accept the larger grid only if it can be strictly nested and gives a predicted gain in efficiency; repeat until no gain in efficiency is produced; and
2) test grid sizes slightly smaller than desired (while keeping at least as many coarse grid levels as obtained in step 1), and accept such a grid only if it results in a substantial predicted gain in efficiency.
Sample computational grids (h = 128, 64, 32, 16 km).
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Example of time stepping algorithm for one full time step Δt on three computational grids. The steps are executed in the numbered order, with left arrows denoting the transfer of the solution from the fine grid to the coarse grid where they overlap. Solid circles mark times and grids where the truncation error is estimated for regridding.
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Choice of a fine-grid patch. The contours show truncation errors (large values flagged by ×) at (a) the start of the time step, and (b) the end of the time step. The boundaries of the minimum acceptable patch at the start (dotted lines) and end (dashed lines) are shown, along with the patch chosen for the step (solid lines).
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Sample solution: the weak hurricane case.
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
(Continued)
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Error vs CPU time for the weak hurricane case.
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Error vs CPU time the for strong hurricane case.
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Same as Fig. 6 except for environmental flow given by (17).
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Comparison of accuracy achieved by the BO (one-way interaction) and MG (two-way interaction) methods for the weak hurricane case.
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Vorticity for the PV ring case computed using movable grids (left panels) and adaptive grids (right panels). Solid rectangles show patch boundaries. The finest patch shown has mesh size h = 0.5 km in each case except the adaptive grids at t = 24 h (h = 1 km) and t = 72 h (h = 2 km).
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
(Continued)
Citation: Monthly Weather Review 129, 1; 10.1175/1520-0493(2001)129<0138:AAMBTC>2.0.CO;2
Achieving the same accuracy with single-grid relaxation (SOR with optimal relaxation parameter) requires about 50–100 sweeps and roughly 10 times as much total computer time.
In Bai and Brandt (1987) the sign of
The exchange rate λ originates as a Lagrange multiplier in the problem of minimizing the error subject to the constraint of constant work (or minimizing the work subject to constant error).
The small oscillations in the zero contour for vorticity are associated with the finite-difference wake, rather than the grid interfaces; with a uniform resolution of h = 8 km the oscillations are actually larger.