1. Introduction
Global weather forecasting and climate models still have grids too coarse to resolve thunderstorms. A common remedy is to embed a high-resolution window—a limited-area model (LAM)—within a global model. A common LAM task is to solve the Poisson equation to calculate the streamfunction from the vorticity and the velocity potential from the divergence.




















In spite of this rather modest predicted rate of convergence, Fourier LAM models and data analysis have been successful, spawning the large number of applications already cited. As noted in the abstract, part of the reason for the success is that accuracy in the meteorologically interesting region is an order of magnitude higher than predicted for the maximum error over the entire Fourier domain (i.e., the square).
The reasons that meteorologists discard a strip of width D around the boundary of the unit square as shown schematically in Fig. 1 are fully described in Boyd (2005) and the references he cites. In this article, we shall concentrate on the mathematics. However, when a storm straddles the boundary between the high-resolution LAM domain (here nondimensionalized to the unit square), and the coarser grid of the global model that fills the rest of the sphere, the dynamics of the storm will be unreliable. Only data that is a finite distance away from the discontinuity in resolution can be trusted. Furthermore, there are several issues about inflow and outflow boundary conditions and blending the limited-area and global approximations smoothly through “Davies relaxation” and so on. It is then silly to worry about numerical accuracy in such regions where the meteorology is already corrupted.

Schematic of the scaled, nondimensionalized Fourier domain (unit square). The buffer region (dotted) is not meteorologically trustworthy. The domain of meteorological interest is thus smaller, non-dotted square (x, y) ∈ [D, 1 − D] ⊗ [D, 1 − D].
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

Schematic of the scaled, nondimensionalized Fourier domain (unit square). The buffer region (dotted) is not meteorologically trustworthy. The domain of meteorological interest is thus smaller, non-dotted square (x, y) ∈ [D, 1 − D] ⊗ [D, 1 − D].
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Schematic of the scaled, nondimensionalized Fourier domain (unit square). The buffer region (dotted) is not meteorologically trustworthy. The domain of meteorological interest is thus smaller, non-dotted square (x, y) ∈ [D, 1 − D] ⊗ [D, 1 − D].
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
The reason that the exclusion of this boundary strip is important is illustrated in Fig. 2, which is a cross section of the error from the center of the domain to the boundary for a typical truncated, unaccelerated Poisson Fourier solution. The error is large near the boundary but diminishes very rapidly away from the boundary at x = 1. This extreme nonuniformity in error is the point of this article.

Cross section along the line segment y = 1/2, x ∈ [1/2, 1] of the absolute value of the error in approximating the solution of the Poisson equation when forced by a typical inhomogeneous function, f = fcheckerboard (x, y) where the “checkerboard” function is defined later in the paper, by an unaccelerated bivariate Fourier series truncated at N = 160. [The analytical solution is (6) to (9) plus (14).]
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

Cross section along the line segment y = 1/2, x ∈ [1/2, 1] of the absolute value of the error in approximating the solution of the Poisson equation when forced by a typical inhomogeneous function, f = fcheckerboard (x, y) where the “checkerboard” function is defined later in the paper, by an unaccelerated bivariate Fourier series truncated at N = 160. [The analytical solution is (6) to (9) plus (14).]
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Cross section along the line segment y = 1/2, x ∈ [1/2, 1] of the absolute value of the error in approximating the solution of the Poisson equation when forced by a typical inhomogeneous function, f = fcheckerboard (x, y) where the “checkerboard” function is defined later in the paper, by an unaccelerated bivariate Fourier series truncated at N = 160. [The analytical solution is (6) to (9) plus (14).]
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
In the rest of this article, we shall prove the higher rate of convergence in the interior in one dimension, demonstrate this in two dimensions, and describe how the order of convergence can be raised by simple modifications to the standard Fourier procedure. We concentrate primarily upon solving the Poisson equation with homogeneous boundary conditions. We shall show in the next-to-last section that when a boundary strip is deleted, the analogous Fourier series that solves the Laplace equation converges exponentially fast.
2. Rates of convergence
Why is the convergence so slow?















Maximum error over the unit square for the solution of uxx + uyy = fcheckerboard (top thin curve with black squares). As in top curve, but that errors at points within D = 1/5 of the boundary are excluded (bottom thick curve with black circles). The dashed lines are guidelines with slopes of 1/N2 and 1/N3, respectively.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

Maximum error over the unit square for the solution of uxx + uyy = fcheckerboard (top thin curve with black squares). As in top curve, but that errors at points within D = 1/5 of the boundary are excluded (bottom thick curve with black circles). The dashed lines are guidelines with slopes of 1/N2 and 1/N3, respectively.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Maximum error over the unit square for the solution of uxx + uyy = fcheckerboard (top thin curve with black squares). As in top curve, but that errors at points within D = 1/5 of the boundary are excluded (bottom thick curve with black circles). The dashed lines are guidelines with slopes of 1/N2 and 1/N3, respectively.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Figure 4 shows the errors for N = 40. The dashed square shows the boundary of the region used to compute Edeleted,strip. Third-order convergence is obviously better than second order, but can one do better?

Absolute values of errors in the N = 40 Fourier approximation to the solution of the Poisson equation when forced by fcheckerboard(x, y). The domain of meteorological interest is shown schematically as the interior of the dashed square.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

Absolute values of errors in the N = 40 Fourier approximation to the solution of the Poisson equation when forced by fcheckerboard(x, y). The domain of meteorological interest is shown schematically as the interior of the dashed square.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Absolute values of errors in the N = 40 Fourier approximation to the solution of the Poisson equation when forced by fcheckerboard(x, y). The domain of meteorological interest is shown schematically as the interior of the dashed square.
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
3. Raising the order of convergence by boundary subtractions

















































4. Fourier solution of the Laplace equation
























All this Fourier–Laplace mechanics are well described in appendix B of Chen and Kuo (1992a). The only novelty in this section is the proof of exponential convergence in the inner square of meteorological interest, given in the appendix.
5. Summary
We have shown that for applications to limited-area modeling (LAM), the domain of needed accuracy is not the whole LAM window, here normalized to the unit square, but is rather a smaller square embedded within the larger. In the smaller domain, we show that the error of the bivariate Poisson Fourier solution is O(1/N3). The previously published estimate O(1/N2), a factor of N more pessimistic, is correct only for the entire square including the meteorologically irrelevant boundary buffer.
We also explicitly constructed an improved Fourier algorithm that solves the Poisson equation with fifth-order accuracy. Because we seek only particular solutions, our order-raising procedure is significantly simpler than previously proposed methods.
The Fourier solution to the Laplace equation, also needed for the harmonic-sine LAM analysis of Chen and Kuo (1992a,b), is shown to converge exponentially fast on the square of meteorological interest.
An exponential rate of convergence can be recovered for the Poisson solution, too, in the domain with the deleted strip by applying a series acceleration method (Boyd 2011). The rate of convergence diminishes to nothing as the discontinuities of the boundary of the domain are approached, but again, because a boundary strip is of no interest, the exponential convergence rate of the accelerated series is always nonzero in the region of meteorological interest. [Another oceanic–atmospheric example where series acceleration on only part of the domain is needed—the tropics for Hermite series on the latitudinally infinite equatorial beta plane—is described in Boyd and Moore (1986).] Other periodize-with-windowing procedures appear promising (Boyd 2005; Termonia et al. 2012; Degrauwe et al. 2012). “Boundary subtraction” can be continued to arbitrary order (Averbuch et al. 1997, 1998; Huybrechs 2010; Huybrechs et al. 2011; Adcock and Hansen 2012).
Thus, we do not claim to have spoken the last word or offered the most accurate algorithm. We only claim that our third- and fifth-order Poisson solvers are simple, and sufficiently accurate for most practical LAM applications.
Acknowledgments
The first author was supported by the NSF through Grant OCE 1059703 and by the Distinguished International Visiting Scholar Program of the Chinese Academy of Sciences. The remaining authors are supported by the project from Research Plan 973 (2012CB417201) and the project from the Natural Science Foundation of China (40930950, 41075043).
APPENDIX
A Proof of the Interior Rate of Convergence for the Fourier Series of a Discontinuous Function
Gibbs Phenomenon and the ensuing O(1) error in the Fourier approximation of a general, nonperiodic function are discussed in standard textbooks including the first author's book (Boyd 2001) and especially in the collection of articles entirely on this topic (Jerri 2011). However, it is difficult to find a careful treatment that proves that the error improves to O(1/N) everywhere outside the “Gibbs layers,” which shrink in width inversely proportional to N, so we provide one. Note that in this appendix, the domain is [0, π] instead of [0, 1].
Our error analysis requires a strengthened form of a classic lemma.
a. Theorem 2 [Riemann–Lebesgue (strong form)]
Suppose f(x) is sufficiently nice as to allow one integration by parts in the integrals below. It is sufficient for f(x) to belong to





















As noted in the main body of this article, a function can always be written on x ∈ [0, 1] as the sum of a linear polynomial B(x) that interpolates f(0) and f(1) plus the function













The lemma shows that the error in the Nth partial sum of










(left) The periodic sine function σ(x) approximated by its sine series up to and including sin(19x). (right) As in (left), but for the piecewise linear sawtooth function with partial sums up to sin(20x).
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

(left) The periodic sine function σ(x) approximated by its sine series up to and including sin(19x). (right) As in (left), but for the piecewise linear sawtooth function with partial sums up to sin(20x).
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
(left) The periodic sine function σ(x) approximated by its sine series up to and including sin(19x). (right) As in (left), but for the piecewise linear sawtooth function with partial sums up to sin(20x).
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1










The limits of integration do not include the points y = 0, π where the denominator sin(y/2) = 0. It follows that for any fixed x on the interior of the interval [0, π], we can invoke the strong form of the Riemann–Lebesgue theorem to prove the following.
b. Theorem 3 (interior rate of convergence)
Suppose f(x) is sufficiently “nice” as to allow one integration by parts in the usual Fourier coefficient integrals below. It is sufficient for f(x) to belong to


There is no contradiction with the classical analysis of Gibbs Phenomenon. At fixed x near a discontinuity, the error is indeed large and O(1) for small N, but as the Gibbs layers shrink with N, the error at fixed x asymptotes to its predicted O(1/N) decay as illustrated in Fig. A2.

Absolute value of
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1

Absolute value of
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
Absolute value of
Citation: Monthly Weather Review 141, 11; 10.1175/MWR-D-13-00074.1
REFERENCES
Adcock, B., and A. C. Hansen, 2012: Stable reconstructions in Hilbert spaces and the resolution of the Gibbs phenomenon. Appl. Comput. Harmon. Anal., 32 (3), 357–388.
Averbuch, A., L. Vozovoi, and M. Israeli, 1997: On a fast direct elliptic solver by a modified Fourier method. Numer. Algorithms, 15 (3–4), 287–313.
Averbuch, A., L. Ioffe, M. Israeli, and L. Vozovoi, 1998: Two-dimensional parallel solver for the solution of Navier–Stokes equations with constant and variable coefficients using ADI on cells. Parallel Comput., 24 (5–6), 673–699.
Boyd, J. P., 2001: Chebyshev and Fourier Spectral Methods. 2nd ed. Dover, 665 pp.
Boyd, J. P., 2005: Limited-area Fourier spectral models and data analysis schemes: Windows, Fourier extension, Davies relaxation, and all that. Mon. Wea. Rev., 133, 2030–2042.
Boyd, J. P., 2009: Large-degree asymptotics and exponential asymptotics for Fourier coefficients and transforms, Chebyshev and other spectral coefficients. J. Eng. Math., 63 (2–4), 355–399.
Boyd, J. P., 2011: Defeating Gibbs Phenomenon in Fourier and Chebyshev spectral methods for solving differential equations. Gibbs Phenomenon, A. Jerri, Ed., Sampling Publishing, 127–172.
Boyd, J. P., and D. W. Moore, 1986: Summability methods for Hermite functions. Dyn. Atmos. Oceans, 10, 51–62.
Bubnová, R., G. Hello, P. Bénard, and J. Geleyn, 1995: Integration of the fully elastic equations cast in the hydrostatic pressure terrain-following coordinate in the framework of the ARPEGE/Aladin NWP system. Mon. Wea. Rev., 123, 515–535.
Chen, Q.-S., and Y.-H. Kuo, 1992a: A consistency condition for wind-field reconstruction in a limited area and a harmonic-cosine series expansion. Mon. Wea. Rev., 120, 2653–2670.
Chen, Q.-S., and Y.-H. Kuo, 1992b: A harmonic-sine series expansion and its application to partitioning and reconstruction problems in a limited area. Mon. Wea. Rev., 120, 91–112.
Chen, Q.-S., L.-S. Bai, and D. H. Bromwich, 1997: A harmonic-Fourier spectral limited-area model with an external wind lateral boundary condition. Mon. Wea. Rev., 125, 143–167.
Cocke, S., 1998: Case study of Erin using the FSU nested regional spectral model. Mon. Wea. Rev., 126, 1337–1346.
Cocke, S., and T. E. LaRow, 2000: Seasonal predictions using a regional spectral model embedded within a coupled ocean–atmosphere model. Mon. Wea. Rev., 128, 689–708.
Degrauwe, D., S. Caluwaerts, F. Voitus, R. Hamdi, and P. Termonia, 2012: Application of Boyd's periodization and relaxation method in a spectral atmospheric limited-area model. Part II: Accuracy analysis and detailed study of operational impact. Mon. Wea. Rev., 140, 3149–3162.
Denis, B., J. Côté, and R. Laprise, 2002: Spectral decomposition of two-dimensional atmospheric fields on limited-area domains using the Discrete Cosine Transform (DCT). Mon. Wea. Rev., 130, 1812–1829.
Errico, R. M., 1985: Spectra computed from a limited area grid. Mon. Wea. Rev., 113, 1554–1562.
Gordon, W. J., and C. A. Hall, 1973: Transfinite element methods: Blending-function interpolation over arbitrary domains. Numer. Math., 21 (2), 109–129.
Haugen, J. E., and B. Machenhauer, 1993: A spectral limited-area model formulation with time-dependent boundary conditions applied to the shallow-water equations. Mon. Wea. Rev., 121, 2618–2630.
Hong, S., and A. Leetmaa, 1999: An evaluation of the NCEP RSM for regional climate modeling. J. Climate, 12, 592–609.
Hong, S., H. H. Juang, and D. Lee, 1999: Evaluation of a regional spectral model for the East Asian monsoon case studies for July 1987 and 1988. J. Meteor. Soc. Japan, 77, 1–20.
Huybrechs, D., 2010: On the Fourier extension of non-periodic functions. SIAM J. Numer. Anal., 47, (6), 4326–4355.
Huybrechs, D., A. Iserles, and S. P. Norsett, 2011: From high oscillation to rapid approximation IV: Accelerating convergence. IMA J. Numer. Anal., 31 (2), 442–468.
Jackson, J. D., 1962: Classical Electrodynamics. John Wiley, 635 pp.
Jerri, A. J., Ed., 2011: Gibbs Phenomenon. Sample Publishing, 500 pp.
Juang, H.-M. H., 1992: A spectral fully compressible nonhydrostatic mesoscale model in hydrostatic sigma coordinates: Formulation and preliminary results. Meteor. Atmos. Phys., 50, 75–88.
Juang, H.-M. H., and M. Kanamitsu, 1994: The NMC nested regional spectral model. Mon. Wea. Rev., 122, 3–26.
Juang, H.-M. H., and S. Y. Hong, 2001: Sensitivity of the NCEP regional spectral model to domain size and nesting strategy. Mon. Wea. Rev., 129, 2904–2922.
Juang, H.-M. H., S. Hong, and M. Kanamitsu, 1997: The NCEP regional model: An update. Bull. Amer. Meteor. Soc., 78, 2125–2143.
Juang, H.-M. H., C. Shiao, and M. Cheng, 2003: The Taiwan Central Weather Bureau regional spectral model for seasonal prediction: Multiparallel implementation and preliminary results. Mon. Wea. Rev., 131, 1832–1847.
Kuo, H.-C., and R. T. Williams, 1992: Boundary effects in regional spectral models. Mon. Wea. Rev., 120, 2986–2992.
Kuo, H.-C., and R. T. Williams, 1998: Scale-dependent accuracy in regional spectral methods. Mon. Wea. Rev., 126, 2640–2647.
Park, J.-R., H.-B. Cheong, and H.-G. Kang, 2011: High-order spectral filter for the spherical-surface limited area. Mon. Wea. Rev., 139, 1256–1278.
Radnoti, G., 1995: Comments on “A spectral limited-area model formulation with time-dependent boundary conditions applied to the shallow-water equations.” Mon. Wea. Rev., 123, 3122–3123.
Sasaki, H., H. Kida, T. Koide, and M. Chiba, 1995: The performance of long-term integrations of a limited area model with the spectral boundary coupling method. J. Meteor. Soc. Japan, 73, 165–181.
Segami, A., K. Khara, H. Nakamura, M. Ueno, I. Takano, and Y. Tatsumi, 1989: Operational mesoscale weather prediction with Japan spectral model. J. Meteor. Soc. Japan, 67, 907–923.
Sköllermo, G., 1975: A Fourier method for the numerical solution of Poisson's equation. Math. Comput., 29 (131), 697–711.
Tatsumi, Y., 1986: A spectral limited-area model with time-dependent lateral boundary conditions and its application to a multi-level primitive equation model. J. Meteor. Soc. Japan, 64, 637–663.
Termonia, P., F. Voitus, D. Degrauwe, S. Caluwaerts, and R. Hamdi, 2012: Application of Boyd's periodization and relaxation method in a spectral atmospheric limited-area model. Part I: Implementation and reproducibility tests. Mon. Wea. Rev., 140, 3137–3148.
Titchmarsh, E. C., 1939: The Theory of Functions. 2nd ed. Oxford University Press, 452 pp.