ON SUPERCONVERGENT DISCONTINUOUS GALERKIN METHODS FOR SCHR ¨ODINGER EQUATIONS AND SPARSE GRID CENTRAL DISCONTINUOUS GALERKIN METHOD By Anqi Chen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Applied Mathematics – Doctor of Philosophy 2019 ABSTRACT ON SUPERCONVERGENT DISCONTINUOUS GALERKIN METHODS FOR SCHR ¨ODINGER EQUATIONS AND SPARSE GRID CENTRAL DISCONTINUOUS GALERKIN METHOD By Anqi Chen In this thesis, we design and analyze a discontinuous Galerkin (DG) method for one- dimensional Schr¨odinger equations under a general class of numerical fluxes, and another efficient DG method for high-dimensional hyperbolic equations. In the first DG method, we develop an ultra-weak discontinuous Galerkin (UWDG) method to solve the one-dimensional nonlinear Schr¨odinger equation. Stability conditions and error estimates are derived for the scheme with a general class of numerical fluxes. The error estimates are based on detailed analysis of the projection operator associated with each individual flux choice. Depending on the parameters, we find out that in some cases, the projection can be defined element-wise, facilitating analysis. In most cases, the projection is global, and its analysis depends on the resulting 2 × 2 block-circulant matrix structures. For a large class of parameter choices, optimal a priori L2 error estimates can be obtained. Numerical examples are provided verifying theoretical results. In addition to the stability and error analysis, we analyze the superconvergence properties of the UWDG method for one-dimensional linear Schr¨odinger equation with various choices of flux parameters. Depending on the flux choices and if the polynomial degree k is even or odd, we prove 2k or (2k − 1)-th order superconvergence rate for cell averages and numerical flux of the function, as well as (2k−1) or (2k−2)-th order for numerical flux of the derivative. In addition, we prove superconvergence of (k + 2) or (k + 3)-th order of the UWDG solution towards a special projection. At a class of special points, the function values and the first and second order derivatives of the UWDG solution are superconvergent with order k +2, k +1, k, respectively. The proof relies on the correction function techniques initiated in [12], and applied to [10] for direct DG (DDG) methods for diffusion problems. By negative norm estimates, we apply the post-processing technique and show that the accuracy of our scheme can be enhanced to order 2k. Theoretical results are verified by numerical experiments. In the second DG method, we develop sparse grid central discontinuous Galerkin (CDG) scheme for linear hyperbolic systems with variable coefficients in high dimensions. The scheme combines the CDG framework with the sparse grid approach, with the aim of break- ing the curse of dimensionality. A new hierarchical representation of piecewise polynomials on the dual mesh is introduced and analyzed, resulting in a sparse finite element space that can be used for non-periodic problems. Theoretical results, such as L2 stability and error estimates are obtained for scalar problems. CFL conditions are studied numerically com- paring discontinuous Galerkin (DG), CDG, sparse grid DG and sparse grid CDG methods. Numerical results including scalar linear equations, acoustic and elastic waves are provided. Copyright by ANQI CHEN 2019 ACKNOWLEDGMENTS I would like to express my great appreciation to my advisor, Professor Yingda Cheng, for her invaluable guidance, inspiration, encouragement and patience during my five-year PhD study. Professor Cheng taught me how to think and express my ideas rigorously like a mathematician. She was always there whenever I needed advices, and she always provided me with kind and thoughtful help. I could never finish this thesis without her help. I would also like to thank all my committee members, Professor Andrew Christlieb, Professor Chichia Chiu and Professor Jianliang Qian for their guidance and advices. I would also like to thank my friends, Zhanjing Tao, Wei Guo, Yan Jiang, Menglun Wang, Xin Yang, Kedi Wu, Zixuan Cang, Xiao Feng, Qinfeng Gao, Chao Song and Ruochuan Zhang for their friendships and many precious moments. Last but not least, I am grateful to my family. My father Weibing Chen and mother Manzhen Deng always encouraged me to pursue what I wanted. I can’t say thanks enough for their support in my life. And I am grateful to my girlfriend Bixi Zhang for her love and support during the final stage of my PhD. v TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 UWDG method for one-dimensional Schr¨odinger equations . . . . . . . . . . 1.3 Superconvergence analysis of DG methods . . . . . . . . . . . . . . . . . . . 1.4 Sparse-grid DG methods for high-dimensional PDEs . . . . . . . . . . . . . . 1 1 2 4 5 Chapter 2 An UWDG method for Schr¨odinger equation in one dimension 7 8 15 16 20 27 32 32 39 2.1 Numerical scheme and stability . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Projection P (cid:63) h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Local projection condition . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2 P (cid:63) h properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Error estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 Numerical results of the projection operator P (cid:63) h . . . . . . . . . . . . 2.4.2 Numerical results of the DG scheme . . . . . . . . . . . . . . . . . . . Chapter 3 Superconvergence analysis of UWDG method on linear Schr¨odinger 46 equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.1 Numerical scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.2 Projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Superconvergence properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 55 . . . . . . . . . . . . . . 56 . . . . . . . . . . . . . . . 3.4 Superconvergence of postprocessed solution . . . . . . . . . . . . . . . . . . . 68 73 3.5 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 3.3.2 Correction functions and the main results Some intermediate superconvergence results Chapter 4 4.1 Numerical Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.1 Periodic problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.2 Discussions on implementations . . . . . . . . . . . . . . . . . . . . . 4.1.3 Discussions on CFL conditions . . . . . . . . . . . . . . . . . . . . . . 4.1.4 Non-periodic problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sparse grid central DG methods for linear hyperbolic systems 81 82 83 88 90 92 4.2 Stability and convergence 99 4.3 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 Scalar case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 System case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 4.3.1 4.3.2 vi APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 vii LIST OF TABLES Table 2.1: Notations for some frequently used quantities. Subscript j will be dropped . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . for uniform mesh. 21 Table 2.2: Interpretation of error estimate (2.29). . . . . . . . . . . . . . . . . . . . . 24 Table 2.3: Example 2.4.1. Error of local projection P (cid:63) h u − u on a nonuniform mesh. Flux parameters: α1 = 0.3, β1 = 0.4, β2 = 0.4. . . . . . . . . . . . . . . . . h u − u on a nonuniform mesh. Flux parameters: α1 = 0.3, β1 = 0.4/h, β2 = 0.4h. . . . . . . . . . . . . . . Table 2.4: Example 2.4.1. Error of local projection P (cid:63) Table 2.5: Example 2.4.1. Difference of local projection P (cid:63) h : P (cid:63) nonuniform mesh. Flux parameters: α1 = 0.5, β1 = 1, β2 = 0. h with P 1 h u− P 1 h u on a . . . . . . . 33 34 34 Table 2.6: Example 2.4.2. Error of global projection P (cid:63) α1 = 0.25, ˜β1 = 1, ˜β2 = 1, p1 = −0.5, p2 = 2. (A1.1) h u − u. Flux parameters: . . . . . . . . . . . . 35 Table 2.7: Example 2.4.2. Error of global projection P (cid:63) 8 2 + k(k+1) 0, β1 = 1 h u− u. Flux parameters: α1 = 2h, β2 = h. (A1.5) . . . . . . . . . . . . . . . . . . . . . . . . . . . h u− u. Flux parameters: α1 = 0.25, ˜β1 = k(k−1) , ˜β2 = 1.0, p1 = −1, p2 = 2, 3. Note here limh→0 λ1, λ2 = (−1)k. (A1.6.1) . . . . . . . . . . . . . . . . . . . . . . . h u − u. Flux parameters: 2k(k−1), p1 = −2,−3, p2 = 1. Note here α1 = 0.25, ˜β1 = 1, ˜β2 = limh→0 λ1, λ2 = (−1)k+1. (A1.7.2) . . . . . . . . . . . . . . . . . . . . . . h u − u. Flux parameters: 2k(k+1), p1 = −2,−3, p2 = 1. Note that α1 = 0.25, ˜β1 = −1, ˜β2 = limh→0 λ1, λ2 = 1. (A1.7.2) . . . . . . . . . . . . . . . . . . . . . . . . . . h u − u. (Central flux) Flux parameters: α1 = 0, β1 = 0, β2 = 0. . . . . . . . . . . . . . . . . . . . . . . h u − u. Flux parameters: 1 1 α1 = 0, β1 = 0, β2 = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 36 37 37 38 38 Table 2.9: Example 2.4.2. Error of global projection P (cid:63) Table 2.8: Example 2.4.2. Error of global projection P (cid:63) Table 2.10: Example 2.4.2. Error of global projection P (cid:63) Table 2.11: Example 2.4.3. Error of global projection P (cid:63) Table 2.12: Example 2.4.3. Error of global projection P (cid:63) viii 1 Table 2.13: Example 2.4.4. Error of global projection P (cid:63) and similar to A1.7.2 in Table 2.9): α1 = 0.25, ˜β1 = −1, ˜β2 = −2,−3, p2 = 1. Note here limh→0 λ1, λ2 = (−1)k+1. h u − u. Flux parameters (A3, 2k(k−1), p1 = . . . . . . . . . . . . Table 2.14: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) on a nonuniform mesh using flux parameters (corresponding to Table 2.3) α1 = 0.3, β1 = β2 = 0.4, ending time Te = 0.3. . . . . . . . . . . . . . Table 2.15: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) on a nonuniform mesh using flux parameters (corresponding to Table 2.4) α1 = 0.3, β1 = 0.4h, β2 = 0.4/hj, ending time Te = 1. . . . . . . . . . Table 2.16: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using central flux (corresponding to A2 in Table 2.11) α1 = β1 = β2 = 0, ending time Te = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . Table 2.17: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A2 in Table 2.12): α1 = β1 = 0, β2 = 1, ending time Te = 1. . . . . . . . . . . . . . . . . . . . . . . Table 2.18: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A1.6.1 in Table 2.8): α1 = 0.25, ˜β1 = k(k−1) , ˜β2 = 1.0, p1 = −1, p2 = 2, 3, ending time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Te = 1. Table 2.19: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A1.7.2 in Table 2.10): α1 = 2k(k+1), p1 = −2,−3, p2 = 1, ending time Te = 1. . . . 0.25, ˜β1 = −1, ˜β2 = + k(k+1) 2 8 1 Table 3.1: Example 3.5.1. Error table when using alternating flux on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. . . . . . . . . . . . . . . . . . . . . Table 3.2: Example 3.5.1. Error table when using flux parameters: α1 = 0.3, β1 = . . . h , β2 = 0.4h on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. 0.4 Table 3.3: Example 3.5.1. Error table when using central flux on uniform mesh. End- ing time Te = 1, x ∈ [0, 2π]. . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.4: Example 3.5.1. Error table when using flux parameters: α1 = 0.25, β1 = h, β2 = 0 on uniform mesh. Ending time Te = 1, x ∈ [0, 2π]. . . . . . 2 h , 5 h, 9 Table 3.5: Example 3.5.1. Error table for intermediate quantities when using alter- nating flux on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. . . . . . ix 39 42 42 43 43 44 44 75 76 77 78 79 Table 3.6: Example 3.5.1. Error table for intermediate quantities when using flux h , β2 = 0.4h on nonuniform mesh. Ending . . . . . . . . . . . . . . . . . . . . . . . . . . . . parameters: α1 = 0.3, β1 = 0.4 time Te = 1, x ∈ [0, 2π]. Table 3.7: Example 3.5.1. Error table for intermediate quantities when using central flux on uniform mesh. Ending time Te = 1, x ∈ [0, 2π]. . . . . . . . . . . . Table 3.8: Example 3.5.1. Postprocessing error table for the four sets of parameters. Ending time Te = 1, uniform mesh on x ∈ [0, 2π]. The first row below labels the parameters by (˜α1, ˜β1, ˜β2). . . . . . . . . . . . . . . . . . . . . . Table 4.1: CFL numbers of the DG method, CDG method, sparse grid DG method and sparse grid CDG method with piecewise degree k polynomials, Runge- Kutta method of order ν for Example 4.3.1 with d=2. The CFL numbers of the sparse grid DG/CDG methods are measured with regard to the most refined mesh hN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 4.2: L2 errors and orders of accuracy for L2 projection operator ˜PD of (4.9) onto ˆ˜Vk N,D when d = 2 and d = 3. N is the number of mesh levels, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 80 80 91 98 Table 4.3: L2 errors and orders of accuracy for Example 4.3.1 at T = 1 when d = 2, T = 2/3 when d = 3, and T = 0.5 when d = 4. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . . . . . . . 109 Table 4.4: L2 errors and orders of accuracy for Example 4.3.1 with Dirichlet boundary condition on the inflow edges at T = 1 when d = 2 and T = 2/3 when d = 3. N denotes mesh level, hN is the size of the smallest mesh on the primal mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . . . . . . . . . . . . . . . . . . . . 109 Table 4.5: L2 errors and orders of accuracy for Example 4.3.2 at T = 2π. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 Table 4.6: L2 errors and orders of accuracy for Example 4.3.3 at T = 1.5. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 x Table 4.7: L2 errors and orders of accuracy for Example 4.3.4 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 Table 4.8: L2 errors and orders of accuracy for Example 4.3.5 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, dimension d = 2. L2 order is calculated with respect to hN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 Table 4.9: L2 errors and orders of accuracy for Example 4.3.6 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 xi LIST OF FIGURES Figure 2.1: A sketch to illustrate the different cases parametrized by the values of p1, p2. 24 Figure 2.2: Example 2.4.5. Absolute difference of (cid:107)uh(t,·)(cid:107) with (cid:107)uh(0,·)(cid:107) with two sets of parameters (α1, α2, β1, β2) = (0.25,−0.25, 1 − i, 1 + i) (denoted by “imag”) and (α1, α2, β1, β2) = (0.25,−0.25, 1, 1) (denoted by “real”) . . . . . . . . . . . . . . . . when k = 2, N = 40, ending time Te = 100. 41 Figure 2.3: Example 2.4.7. Double soliton collision graphs at t = 0, 2.5, 5 and a x − t plot of the numerical solution. N = 250, P 2 elements with periodic boundary conditions on [-25,25]. Central flux (α1 = β1 = β2 = 0) is used. Figure 4.1: Illustration of one-dimensional bases on different levels for k = 0: non- periodic problems. Different colors represent different bases. . . . . . . . 45 99 Figure 4.2: Illustration of one-dimensional bases on different levels for k = 1: non- periodic problems. Different colors represent different bases. . . . . . . . 100 Figure 4.3: Example 4.3.1. The time evolution of the error of L2 norm of numerical solutions uh and vh of the sparse grid CDG method with d = 2. (a) k=1, (b) k=2, (c) k=3. N = 4, 5, 6. . . . . . . . . . . . . . . . . . . . . . . . 108 Figure 4.4: L2 errors and associated CPU times of DG, CDG, sparse grid DG and sparse grid CDG methods for Example 4.1 with initial condition (4.24) at T = 1 for d=2. (a) k=1, (b) k=2, (c) k=3. . . . . . . . . . . . . . . . . 111 Figure 4.5: Example 4.3.3. Deformational flow test. The contour plots of the numer- ical solutions on primal mesh at t = T /2 (a, c, e) and t = T (b, d, f). k = 1 (a, b), k = 2 (c, d), and k = 3 (e, f). N = 7. . . . . . . . . . . . . 114 xii Chapter 1 Introduction 1.1 Overview The discontinuous Galerkin (DG) method is a class of finite element methods using com- pletely discontinuous piecewise function space for test functions and numerical solution. The first DG method was introduced by Reed and Hill in [59]. A major development of DG method is the Runge-Kutta DG (RKDG) framework introduced for solving hyper- bolic conservation laws containing only first order spatial derivatives in a series of papers [25, 24, 22, 21, 26]. Because of the completely discontinuous basis, DG method has several attractive properties. It can be used on many types of meshes, even those with hanging nodes. The method has h-p adaptivity and very high parallel efficiency. A particular type of DG methods that is related to this thesis is central DG (CDG) scheme. The CDG schemes [48, 50, 51] are a class of DG schemes on overlapping cells that combine the idea of the central schemes [56, 45, 49] with the DG weak formulation. Such methods are intrinsically Riemann solver free, therefore no costly flux evaluations are needed in the computation. It is well known that the CDG schemes allow larger CFL numbers than the standard DG methods except for piecewise constant approximations [50, 60]. This compensates the increased cost caused by duplicate representation of the solution on the dual mesh. 1 In this thesis, we will focus on the design of a new DG method for one-dimensional Schr¨odinger equations and its error analysis, and superconvergence analysis, as well as a new sparse-grid central DG method for high-dimensional hyperbolic equations to make the simulation more efficient and accurate. 1.2 UWDG method for one-dimensional Schr¨odinger equations In this section, we introduce the one-dimensional time-dependent nonlinear Schr¨odinger (NLS) equation and review the numerical methods designed to solve this equation. The NLS equation is written as follows: iut + uxx + f (|u|2)u = 0, (1.1) where f (u) is a nonlinear real function and u is a complex function. The Schr¨odinger equation is the fundamental equation in quantum mechanics, reaching out to many applications in fluid dynamics, nonlinear optics and plasma physics. It is also called Schr¨odinger wave equation as it can describe how the wave functions of a physical system evolve over time. Many numerical methods have been applied to solve NLS equations [14, 28, 42, 43, 58, 71, 74]. In [14, 74], several important finite difference schemes are implemented, analyzed and compared. In [58], the author introduced a pseudo-spectral method for general NLS equations. Many finite element methods have been tested, such as quadratic B-spline for NLS in [28, 71] and space-time DG method for nonlinear (cubic) Schr¨odinger equation in [42, 43]. Various types of DG schemes has been applied to solve Schr¨odinger equations and they 2 have different discretization for the second order spatial derivative term. One group of such methods is the so-called local DG (LDG) method invented in [25] for convection-diffusion equations. The algorithm is based on introducing auxiliary variables and reformulating the equation into its first order form. In [80], an LDG method using alternating fluxes is developed with L2 stability and proved (k + 1 2)-th order of accuracy. Later in [81], Xu and Shu proved optimal accuracy for both the solution and the auxiliary variables in the LDG method for high order wave equations based on refined energy estimates. In [47], the authors presented an LDG method with exponential time differencing Runge-Kutta scheme and investigated the energy conservation performance of the scheme. Another group of method involves treating the second order spatial derivative directly in the weak formulations, such as IPDG method [77, 30] and NIPG method [62, 63]. Those schemes enforce a penalty jump term in the weak formulation, and they have been extensively applied to acoustic and elastic wave propagation [37, 3, 61]. As for Schr¨odinger equations, the direct DG (DDG) method was applied to Schr¨odinger equation in [52] and achieved energy conservation and optimal accuracy. In Chapter 2, we choose to discretize the second order spatial derivative term directly using UWDG method, which can be traced backed to [13], and refer to those DG methods [72] that rely on repeatedly applying integration by parts so all the spatial derivatives are shifted from the solution to test function in the weak formulation. In [18], Cheng and Shu developed UWDG methods for general time dependent problems with higher order spatial derivatives. In [6], Bona et. al. proposed an UWDG scheme for generalized KdV equation and performed error estimates. We investigate a most general form of the numerical flux functions that ensures stability along with our ultra-weak formulation. To estimate the convergence rate of our scheme, 3 we introduce special projections associated different flux parameters, and proved detailed estimates for the projections. With the results for the convergence rate of the projections, a priori L2 error estimates are obtained. Numerical tests are provided to verify the theoretical results for projection operators and solution convergence rates under various flux parameters. 1.3 Superconvergence analysis of DG methods The study of superconvergence is of importance because a posteriori error estimates can be derived guiding adaptive calculations. For superconvergence of DG methods, many results exist in the literature. We refer the readers to [2, 1] for ordinary differential equation re- sults. In [19], Cheng and Shu proved that the DG and LDG solutions are (k + 3/2)-th order superconvergent towards projections of exact solutions of hyperbolic conservation laws and convection-diffusion equations using specially designed test functions when piecewise poly- nomials of degree k are used. For linear hyperbolic problems, in [82], Yang and Shu proved that, under suitable initial discretization, the DG solutions of linear hyperbolic systems are convergent with optimal (k + 2)-th order at Radau points. More recently, in [12], Cao et al proved the (2k + 1)-th superconvergence rate for cell average and DG numerical fluxes by introducing a locally defined correction function. The correction function also helps simplify the proof for point wise (k + 2)-th superconvergence rate at Radau points and prove the derivative of DG solution has (k + 1)-th superconvergence rate at so-called “left Radau” points. Then this technique has been extended to prove the superconvergence of DG so- lutions for linear and nonlinear hyperbolic PDEs in [11, 9], DDG method for convection diffusion equations [10] and LDG method for linear Schr¨odinger equations [84]. Overall, for equations with higher order spatial derivatives, the same type of correction functions can 4 be used for the LDG method which is based on a reformulation into a first order system of equations. For DDG method, new correction functions are needed treating the second order derivative directly [10]. Another type of superconvergence of DG methods is achieved by postprocessing the solution by convolution with a kernel function, which is a linear combination of B-spline functions. For linear hyperbolic systems, [23] provided a framework for constructing such postprocessor and proving the superconvergence of the postprocessed DG solutions. Through the analysis of negative norm estimates and divided difference estimates, they showed that the postprocessed solution is superconvergent at a rate of 2k + 1. More recently, in [41, 54] the analysis are extended to scalar nonlinear hyperbolic equations. In Chapter 3, we study the superconvergence properties of the UWDG methods for linear Schr¨odinger equation with scale invariant flux parameters. Such choice include all commonly used fluxes, e.g. alternating, central, DDG and interior penalty DG (IPDG) fluxes. Using the correction function idea and negative norm estiamtes, we are able to prove superconvergence rate for solution and its derivatives at certain points, for cell averages and numerical fluxes, for solution towards a special projection, and for the postprocessed solution. 1.4 Sparse-grid DG methods for high-dimensional PDEs It has been a challenging problem for numerical simulations in high dimensions due to the so-called curse of dimensionality, which means the cost of computing and storing an approx- imation with a prescribed accuracy increases exponentially on dimension d. Many discretiza- tion techniques and computation techniques have been developed to alleviate the problem 5 to some extent. Among them, the sparse grid method has been a successful tool. It was originally developed to solve PDEs [83, 33] based on tensor product hierarchical basis rep- resentation. The method can reduce the full grid discretization complexity from O(h−d) to O(h−1| log2 h|d−1), where h is the uniform mesh size in each dimension, and only slightly deteriorate the accuracy. In recent years, sparse grid techniques have been incorporated in collocation methods for high-dimensional stochastic differential equations [79, 78, 57, 53], finite element methods [83, 8, 66], finite difference methods [34, 36], finite volume methods [40], and spectral methods [35, 32, 68, 69] for high-dimensional PDEs. Recently, our research group initiated a line of research on the development of sparse grid DG methods [76, 38, 39]. The sparse grid DG methods use the sparse finite element space, which has multidimensional multiwavelet bases constructed by tensor products from one-dimensional wavelet basis, in the DG framework to treat high-dimensional problems. The methods has been proven to reduce the degrees of freedom of O(h−d) in the standard full grid approximation space to O(h−1| log2 h|d−1) and remain a L2 convergence rate of O(hk+1/2| log2 h|d) for transport equations and a convergence rate of O(hk| log2 h|d) in the energy norm for elliptic equations. In Chapter 4, we incorporate sparse grid DG method with central DG scheme to de- velop a class of conservative numerical schemes with high computational efficiency for high- dimensional hyperbolic equations. Our work consists of construction of the sparse finite element space , L2 stability and error estimates, and numerical validation of the scheme. 6 Chapter 2 An UWDG method for Schr¨odinger equation in one dimension In this chapter, we develop and analyze a new ultra-weak discontinuous Galerkin (UWDG) method for solving one-dimensional nonlinear Schr¨odinger (NLS) equations (1.1). The method solves the equation without introducing any auxiliary variables or rewriting the equation into a larger system. The focus of this chapter is on the investigation of a most general form of the numerical flux functions that ensures stability along with our ultra-weak formulation. The fluxes under consideration include the alternating fluxes, and also the fluxes considered in [52], and therefore allows for flexibility for the design of the schemes. The analysis in this chapter relies on a detailed analysis of a special projection associated with different flux parameters, whose dependence on mesh size can be freely enforced. Under certain flux parameters, the projection can be defined locally. For other flux parameters, the projection is global and the projection analysis is based on a block-circulant matrix with 2×2 blocks. Our analysis reveals that under a large class of parameter choices, the UWDG method is optimally convergent in L2 norm, which is verified by extensive numerical tests for both the projection operators and the numerical schemes for (1.1). The remainder of this chapter is organized as follows. In Section 2.1, we introduce the 7 UWDG method with general flux definitions for one-dimensional nonlinear Schr¨odinger equa- tions and study its stability properties. We introduce a new projection operator and analyze its properties in Section 2.2, which is later used in Section 2.3 to obtain the convergence results of the schemes. The main body of this chapter, the error estimates, is contained in Section 2.3. Numerical validations are provided in Section 2.4. Some technical details, including proof of most lemmas are collected in Appendix. The major contents of this chapter has been published in [16]. 2.1 Numerical scheme and stability In this subsection, we formulate and discuss stability results of a DG scheme for one- dimensional NLS equation (1.1) on interval I = [a, b] with initial condition u(x, 0) = u0(x) and periodic boundary conditions. Here f (u) is a given real function. Our method can be defined for general boundary conditions, but the error analysis will require slightly different tools, and therefore we only consider periodic boundary conditions in this chapter. To facilitate the discussion, first we introduce some notations and definitions. For a 1-D interval I = [a, b], the usual DG meshes are defined as: and a = x 1 2 < x 3 2 < ··· < x N + 1 2 = b, Ij = (x j− 1 2 , x j+ 1 2 ), xj = 1 2 (x j− 1 2 + x ), j+ 1 2 hj = x j+ 1 2 − x , j− 1 2 h = max j hj, 8 with mesh regularity requirement < σ, σ is fixed during mesh refinement. h min hj Denote ZN = 1, 2,··· , N . The approximation space is defined as: h = {vh : vh|Ij V k ∈ P k(Ij), ∀j ∈ ZN}, h , we use (vh)− j− 1 2 meaning vh is a piecewise polynomial of x with degree up to k on each cell Ij. For a function vh ∈ V k Ij−1 and the right cell Ij respectively. The jump and average are defined as [vh] = v+ and {vh} = 1 to refer to the value of vh at x from the left cell h − v− and (vh)+ h ) at cell interfaces. h + v− 2(v+ j− 1 2 j− 1 2 h Throughout this chapter, we use the standard Sobolev norm notations (cid:107) · (cid:107)W s,p(I) and broken Sobolev space on mesh IN . We denote (cid:107)v(cid:107)2 Hs(Ij ) and (cid:107)v(cid:107)W s,∞(IN ) = maxj (cid:107)v(cid:107)W s,∞(Ij ). In Section 3.4, we consider negative norms and the def- =(cid:80)N inition is (cid:107)v(cid:107) . Additionally, we denote by (cid:107)v(cid:107)L2(∂IN ) the Hl(I) broken L2 norm on cell interfaces, i.e., (cid:107)v(cid:107)2 L2(∂Ij ) (v− )2. We also denote (cid:107)·(cid:107) = (cid:107)·(cid:107)L2(I) = (cid:107)·(cid:107)L2(IN ) to shorten the notation. I v(x)Φ(x)dx (cid:107)Φ(cid:107) , where(cid:107)v(cid:107)2 = supΦ∈C∞ j=1 (cid:107)v(cid:107)2 j=1 (cid:107)v(cid:107)2 L2(∂IN ) )2 + (v+ x H−l(I) L2(∂Ij ) 0 (I) (cid:82) = x j+ 1 2 Lastly, we recall inverse inequalities, ∀vh ∈ V k h , j− 1 2 Hs(IN ) = (cid:80)N (cid:107)(vh)x(cid:107)L2(Ij ) ≤ Ch−1 j (cid:107)vh(cid:107)L2(Ij ), − 1 2(cid:107)vh(cid:107)L2(Ij ), (cid:107)vh(cid:107)L2(∂Ij ) ≤ Ch − 1 (cid:107)vh(cid:107)L∞(Ij ) ≤ Ch 2(cid:107)vh(cid:107)L2(Ij ), (2.1) and trace inequalities (cid:107)v(cid:107)2 L2(∂Ij ) ≤ Ch−1 j (cid:107)v(cid:107)2 L2(Ij ) , (2.2) here and below C is a constant independent of the function u and the mesh size h. In this chapter, we consider a DG scheme motivated by [18] and based on integration by 9 parts twice, or the so-called ultra-weak formulation. In particular, we look for the unique function uh = uh(t) ∈ V k h , t ∈ (0, T ], such that (cid:90) Ij (cid:90) i (uh)tvhdx + uh(vh)xxdx − ˆuh(vh)− x | (cid:90) j+ 1 2 Ij +(cid:94)(uh)xv− h | j+ 1 2 − (cid:94)(uh)xv+ h | j− 1 2 + Ij x | + ˆuh(vh)+ j− 1 2 f (|uh|2)uhvhdx = 0 (2.3) holds for all vh ∈ V k h and all j = 1, ··· , N . Here, we require k ≥ 1, because k = 0 yields an inconsistent scheme. Notice that (2.3) can be written equivalently in a weak formulation by performing another integration by parts back as: (cid:90) i Ij (cid:90) Ij (uh)tvhdx − (uh)x(vh)xdx + (u− h − ˆuh)(vh)− x | j+ 1 2 − (cid:94)(uh)xv+ h | j− 1 2 + + (ˆuh − u+ (cid:90) x | h )(vh)+ j− 1 2 f (|uh|2)uhvhdx = 0 (2.4) Ij +(cid:94)(uh)xv− h | j+ 1 2 The “hat” and“tilde” terms are the numerical fluxes we pick for u and ux at cell bound- aries, which are single valued functions defined as: (cid:94)(uh)x = {(uh)x} + α1[(uh)x] + β1[uh], ˆuh = {uh} + α2[uh] + β2[(uh)x], (2.5) where α1, α2, β1, β2 are prescribed complex parameters. They may depend on the mesh parameter h. Commonly used fluxes such as the central flux (by setting α1 = α2 = β1 = β2 = 0) and alternating fluxes (by setting α1 = −α2 = ± 1 2 , β1 = β2 = 0) belong to this flux family. The direct DG scheme considered in [52] is a special case of our method when α1 = −α2, β1 = c framework as α1 = α2 = β2 = 0, β1 = c h, β2 = 0, c > 0, α1 ∈ R. The IPDG method can also be casted in this h, c > 0. 10 We write the scheme (2.3) in following short-hand notation: (cid:90) Ij aj(uh, vh) − i holds for all vh ∈ V k h , where f (|uh|2)uhvhdx = 0, ∀j ∈ ZN (2.6) (cid:90) Ij (uh)tvhdx − iAj(uh, vh), aj(uh, vh) = with Aj(uh, vh) = (cid:90) Ij uh(vh)xxdx − ˆuh(vh)− x | j+ 1 2 x | + ˆuh(vh)+ j− 1 2 + (cid:94)(uh)xv− h | j+ 1 2 − (cid:94)(uh)xv+ h | j− 1 2 as the UWDG spatial discretization for the second order derivative term. Using periodic boundary condition, we sum up on j for (2.6) and get (cid:90) I a(uh, vh) − i f (|uh|2)uhvhdx = 0, (2.7) where a(uh, vh) = A(uh, vh) = (cid:90) (uh)tvhdx − iA(uh, vh), N(cid:88) (cid:90) I Aj(uh, vh) = uh(vh)xxdx + I j=1 N(cid:88) j=1 (cid:16) ˆuh[(vh)x] − (cid:94)(uh)x[vh] (cid:17)(cid:12)(cid:12)j+ 1 2 . The following theorem contains the results on semi-discrete L2 stability. Theorem 2.1.1. (Stability) For u, v ∈ H2(IN ) satisfying periodic boundary condition, we have A(u, v) = A(v, u). 11 The solution of semi-discrete UWDG scheme (2.3) using numerical fluxes (2.5) satisfies L2 stability condition if d dt (cid:90) I |uh|2dx ≤ 0, Imβ2 ≥ 0, Imβ1 ≤ 0, |α1 + α2|2 ≤ −4Imβ1Imβ2. (2.8) In particular, when all parameters α1, α2, β1, β2 are restricted to be real, this condition amounts to without any requirement on β1, β2. α1 + α2 = 0 (2.9) Proof. From integration by parts, we have A(u, v) = − uxvxdx + Similarly, A(v, u) = −(cid:82) I I uxvxdx +(cid:80)N (cid:90) N(cid:88) (ˆu[vx] − [uvx] −(cid:102)ux[v])(cid:12)(cid:12)j+ 1 j=1 (ˆv[ux] − [vux] − (cid:101)vx[u])(cid:12)(cid:12)j+ 1 j=1 . 2 2 ,∀j ∈ ZN j+ 1 2 . Plugging in the defi- nition of the numerical fluxes in (2.5), we have at x ˆu[vx] − [uvx] −(cid:102)ux[v] =(cid:0){u} − α1[u] + β2[ux](cid:1)[vx] − ({u}[vx] + [u]{vx}) −(cid:0){ux} + α1[ux] + β1[u](cid:1)[v] = [ux](cid:0){v} − α1[v] + β2[vx](cid:1) − [u](cid:0){vx} + α1[vx] + β1[v](cid:1) −(cid:0)[ux]{v} + {ux}[v](cid:1) = [ux]ˆv − [u](cid:101)vx − [uxv], 12 then the proof for A(u, v) = A(v, u) is complete. From integration by parts, we have, for ∀vh ∈ V k N(cid:88) (cid:90) (cid:90) h (uh)tvhdx + i (uh)x(vh)xdx + i a(uh, vh) = I I j=1 ([uh(vh)x] − ˆuh[(vh)x] + (cid:94)(uh)x[vh])| . j+ 1 2 (cid:90) I A(uh, vh) = − N(cid:88) (uh)x(vh)xdx + j=1 (ˆuh[(vh)x] − [uh(vh)x] − (cid:94)(uh)x[vh])| j+ 1 2 Taking vh = ¯uh in (2.7) and compute its conjugate as well, we get (cid:90) (cid:90) I f (|uh|2)|uh|2dx 0 = i I f (|uh|2)|uh|2dx + i (cid:90) = a(uh, ¯uh) + a(uh, ¯uh) = d dt I |uh|2dx − iA(uh, uh) + iA(uh, uh). . (2.10) −iA(uh, uh) + iA(uh, uh) = −2Im N(cid:88) ([uh(¯uh)x] − ˆuh[(¯uh)x] + (cid:94)(uh)x[¯uh])| j=1 j+ 1 2 (cid:16){uh}[(¯uh)x] + [uh]{(¯uh)x} −(cid:0){uh} + α2[uh] + β2[(uh)x](cid:1)[(¯uh)x] (cid:0) − β2|[(uh)x]|2 + β1|[uh]|2 + α1[(uh)x][¯uh] − α2[uh][(¯uh)x](cid:1)| (cid:17)| j+ 1 2 j+ 1 2 + ({(uh)x} + α1[(uh)x] + β1[uh])[¯uh] = −2Im = −2Im N(cid:88) j=1 N(cid:88) N(cid:88) j=1 = 2Im (β2|[(uh)x]|2 − β1|[uh]|2 − (α1 + α2)[¯uh][(uh)x])| j+ 1 2 j=1 13 Plug it back into (2.10): (cid:90) I d dt N(cid:88) j=1 |uh|2dx + 2Imβ2|[(uh)x]|2 − 2Imβ1|[uh]|2 − 2Im{(α1 + α2)[¯uh][(uh)x]}| = 0. j+ 1 2 (2.11) If the stability condition (2.8) is satisfied, we have If all parameters are real and (2.9) is satisfied, then (2.11) further yields: d dt |uh|2dx ≤ 0. (cid:90) (cid:90) I I d dt which implies energy conservation. |uh|2dx = 0, (2.12) For simplicity of the discussion, in the contents below, we will only consider real param- eters, i.e. when α1, α2, β1, β2 are real and α1 + α2 = 0. This property of our scheme is consistent with the energy conservation property of Schr¨odinger equations. It is essential to have a symmetric A(uh, vh) for designing a finite element scheme which is energy-preserving for Schr¨odinger equations. Now the numerical fluxes are defined by three parameters as, (cid:94)(uh)x = {(uh)x} + α1[(uh)x] + β1[uh], ˆuh = {uh} − α1[uh] + β2[(uh)x], α1, β1, β2 ∈ R. (2.13) 14 Note we can rewrite the flux definition in a matrix form  ˆuh (cid:94)(uh)x  = G  u− h  + H  u+ h  , G =  1 (uh)− x (uh)+ x  , H =  1 2 − α1 β1  , β2 1 2 + α1 2 + α1 −β2 2 − α1 −β1 1 (2.14) where I2 denotes the 2 × 2 identity matrix. Note that G + H = I2, det G = det H = −(α2 4) and GH = −(det G)I2. 1 + β1β2 − 1 2.2 Projection P (cid:63) h In this section, we perform detailed studies of a projection operator that is key to the analysis of the UWDG scheme. Definition 2.2.1. For the UWDG scheme with flux choice (2.13), we define the associated h for any periodic function u ∈ W 1,∞(I) to be the unique polynomial projection operator P (cid:63) h u ∈ V k P (cid:63) h (when k ≥ 1) satisfying (cid:90) (cid:90) Ij P (cid:63) h u vhdx = Ij (cid:100)P (cid:63) h u = {P (cid:63) (cid:94)(P (cid:63) h u)x = {(P (cid:63) h u} − α1[P (cid:63) h u)x} + α1[(P (cid:63) h u] + β2[(P (cid:63) h u)x] = u h u)x] + β1[P (cid:63) h u] = ux u vhdx ∀vh ∈ P k−2(Ij), (2.15a) at x at x , , j+ 1 2 j+ 1 2 (2.15b) (2.15c) for all j. When k = 1, only conditions (2.15b)-(2.15c) are needed. This definition is to ensure u − (cid:100)P (cid:63) h u = 0 and ux − (cid:94)(P (cid:63) h u)x = 0 at cell boundaries, which will be used in error estimates for the scheme. In the following, we analyze the projection when the parameter choice reduces it to a local projection in Section 2.2.1, and then we 15 consider the more general global projection in Section 2.2.2. We can write (2.15b)-(2.15c) in vector form as  (cid:100)P (cid:63) h u (cid:94)(P (cid:63) h u)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  = G j+ 1 2  P (cid:63) h u (P (cid:63) h u)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   P (cid:63) h u (P (cid:63) h u)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  + x j+ 1 2  u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  = j+ 1 2 . (2.16) − x j+ 1 2 + H 2.2.1 Local projection condition In general, the projection P (cid:63) h is globally defined, and its existence, uniqueness and ap- proximation properties are quite complicated mathematically. However, with some special parameter choices, P (cid:63) h can be reduced to a local projection, meaning that it can be solved element-wise, and hence the analysis can be greatly simplified. For example, with the alternating fluxes α1 = ± 1 h for parameter choice α1 = 1 P 1 h and P 2 as: for each cell Ij, we find the unique polynomial of degree k, P 1 h defined below. P (cid:63) h = P 1 2 , β1 = β2 = 0, P (cid:63) h can be reduced to 2 , β1 = β2 = 0 is formulated h u, satisfying (cid:90) Ij (cid:90) P 1 h u vhdx = Ij h u)− = u (P 1 (P 1 h u)+ x = ux u vhdx ∀vh ∈ P k−2(Ij), at x j+ 1 2 at x j− 1 2 , . (2.17a) (2.17b) (2.17c) When k = 1, only conditions (2.17b)-(2.17c) are needed. Similarly, we can define P (cid:63) h = P 2 h for parameter choice α1 = − 1 2 , β1 = β2 = 0 as: for 16 each cell Ij, we find the unique polynomial of degree k, P 2 h u, satisfying (cid:90) Ij (cid:90) Ij P 2 h u vhdx = u vhdx ∀vh ∈ P k−2(Ij), h u)+ = u (P 2 h u)− (P 2 x = ux at x j− 1 2 at x j+ 1 2 , . (2.18a) (2.18b) (2.18c) When k = 1, only conditions (2.18b)-(2.18c) are needed. Similar local projections have been introduced and considered in [18]. It is obvious that h u can be solved element-wise, and their existence, uniqueness are straightforward. P 1 h u, P 2 From a standard scaling argument by Bramble-Hilbert lemma in [20], P 1 following error estimates: let u ∈ W k+1,p(Ij)(p = 2,∞), then h and P 2 h have the (cid:107)u − P ν (cid:107)ux − P ν h u(cid:107)Lp(Ij ) ≤ Chk+1 h ux(cid:107)Lp(Ij ) ≤ Chk j j|u| |u| W k+1,p(Ij ) W k+1,p(Ij ) p = 2,∞, ν = 1, 2, p = 2,∞, ν = 1, 2, , , (2.19) where here and below, C is a generic constant that is independent of the mesh size hj, the parameters α1, β1, β2 and the function u, but may take different value in each occurrence. Naturally, the immediate question is that if there are other parameter choices such that P (cid:63) h can be reduced to a local projection. The following lemma addresses this issue. Lemma 2.2.1 (The condition for reduction to a local projection). If α2 1 + β1β2 = 1 4, P (cid:63) h is 17 a local projection. Moreover, (2.15b) and (2.15c) is equivalent to  P (cid:63) h u (P (cid:63) h u)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  G − x j+ 1 2 + H  P (cid:63) h u (P (cid:63) h u)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  = G + x j− 1 2  u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  + H j+ 1 2  u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  j− 1 2 . (2.20) Proof. The definition (2.15a) provides k − 1 linearly independent equations for solving P (cid:63) h u on each cell. If (2.15b) and (2.15c) can be locally decoupled, P (cid:63) assumption α2 1 + β1β2 = 1 4, if β1 = β2 = 0, then α1 = ± 1 2 and P (cid:63) h is a local projection. By h u = P 1 h , and (2.20) h or P 2 holds. The rest of the cases are • if β1 (cid:54)= 0, left multiply (2.16) by a matrix, we have β1 1 2 + α1 2 − α1) β1 −( 1   u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x   0 β1 0 β1 −( 1 2 − α1) 0  1 2 + α1 0 h u  P (cid:63)  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   P (cid:63) h u)x (P (cid:63) (P (cid:63) h u h u)x + (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  − x j+ 1 2 , x j+ 1 2 j+ 1 2 = + which implies the following decoupled relations h u)+ + (P (cid:63) h u)− − 1 (P (cid:63) 1 2 + α1 β1 2 − α1 β1 h u)+ (P (cid:63) h u)− (P (cid:63) x = u + x = u − 1 1 2 + α1 β1 2 − α1 β1 ux at x ux at x , . j− 1 2 j+ 1 2 (2.21) 18 • if β2 (cid:54)= 0, by similar linear transformation, we have  1 2 − α1 −( 1 2 + α1) β2 β2   u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x    1 j+ 1 2 = + 0 h u  P (cid:63)  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   P (cid:63) h u)x (P (cid:63) (P (cid:63) h u h u)x +  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  − x j+ 1 2 , x j+ 1 2 0 −( 1 2 + α1) β2 2 − α1 β2 0 0 which implies (P (cid:63) h u)+ x + x − 1 h u)− (P (cid:63) 1 2 − α1 β2 2 + α1 β2 (P (cid:63) h u)+ = ux + h u)− = ux − 1 (P (cid:63) 1 2 − α1 β2 2 + α1 β2 u at x u at x , . j− 1 2 j+ 1 2 (2.22) (2.21), (2.22) are the desired decoupled conditions on each cell Ij, and it’s easy to verify that (2.21), (2.22) are equivalent to (2.20). Therefore the proof is complete. This lemma implies that for any parameter satisfying α2 1+β1β2 = 1 4, P (cid:63) h is locally defined. We remark that this condition turns out to be the same as the optimally convergent numerical flux families in [17] for two-way wave equations, although they arise in different contexts. Unfortunately, for the general definition of P (cid:63) h , unlike P 1 h and P 2 h , we cannot directly use the Bramble-Hilbert lemma and the standard scaling argument to obtain optimal approximation property, since the second and third relations in (2.21) and (2.22) may break the scaling. 19 2.2.2 P (cid:63) h properties Before moving on to detailed discussion on P (cid:63) h , we introduce some notations to facilitate the discussion. We define the Legendre expansion of a function u ∈ L2(I) on cell Ij as follows, ∞(cid:88) m=0 u|Ij = uj,mLj,m(x), (2.23) where Lj,m(x) := Lm(ξ), ξ = degree m on [−1, 1]. x−xj hj /2 , and Lm(·) is the standard Legendre polynomial of In what follows, we write Lj,m(x) as Lj,m, and Lm(ξ) as Lm for notational convenience. We can compute uj,m using orthogonality of Legendre polynomials (cid:90) 1 2m + 1 and Rodrigues’ formula, uj,m = = = 2m + 1 hj 2m + 1 2 2m + 1 2 (cid:90) Ij 1 2mm! (−1)l 2mm! (cid:90) 1 (cid:90) 1 −1 −1 u(x)Lj,mdx = ˆuj(ξ)Lmdξ 2 −1 dξm (ξ2 − 1)mdξ d ˆuj(ξ) d dξl ˆuj(ξ) d dξm−l (ξ2 − 1)mdξ, (2.24) where ˆuj(ξ) = u(x(ξ)) is defined as the function u|Ij [−1, 1]. By Holder’s inequality, if u ∈ W l,p(I), transformed to the reference domain (cid:12)(cid:12)uj,m (cid:12)(cid:12) ≤ Ch l− 1 p j |u| W l,p(Ij ) 0 ≤ l ≤ m. , (2.25) The L2 projection P 0 h is closely related to uj,m. By orthogonality of Legendre polynomi- als, we have P 0 h u = k(cid:88) uj,mLj,m. We collect some frequently used notations in Table 2.1 for quick reference. m=0 20 Table 2.1: Notations for some frequently used quantities. Subscript j will be dropped for uniform mesh. β1 + β2 h2 j 1 + β1β2 + 1 4) Notation G Γj L− j,m Aj Q Mj,m (cid:21)  ) 1 (cid:20) 1 Definition 2 + α1 −β2 −β1 2 − α1  Lj,m(x k2(k2 − 1) − 2k2 (α2 hj j+ 1 2 2 d dx Lj,m(x j+ 1 hj j,k−1, L− G[L− 2 j,k] −A−1B (Aj + Bj)−1(GL− ) j,m + HL+ j,m) − 2k hj Definition β2 (cid:20) 1 (cid:21) 2 − α1 1 2 + α1 β1   Lj,m(x 1 + β1β2 − 1 (α2 4) j− 1 2 2 d dx Lj,m(x j− 1 hj 2 H[L+ j,k−1, L+ j,k] Ql(I2 − QN )−1 ) ) Notation H Λj L+ j,m Bj rl Next, we will write the explicit formula of P (cid:63) h in order to get a clear view of the existence and uniqueness condition, as well as the error estimates. Suppose k(cid:88) m=0 P (cid:63) h u = ´uj,mLj,m. By the definition (2.15a), ´uj,m = uj,m, m ≤ k − 2, i.e., k(cid:88) k−2(cid:88) P (cid:63) h u = m=0 uj,mLj,m + ´uj,mLj,m. m=k−1 In what follows, we analyze the existence and uniqueness of P (cid:63) h , i.e., the existence and uniqueness of ´uj,k−1, ´uj,k based on the following assumptions on parameters: 4 and Γj (cid:54)= 0. • A0. (Local projection) α2 1 + β1β2 = 1 • A1. (Global projection) uniform mesh (hj = h,∀j), α2 (cid:12)(cid:12)(cid:12) > 1. (cid:12)(cid:12)(cid:12) Γ (cid:12)(cid:12)(cid:12) = 1. (cid:1)N (cid:54)= 1. If N is odd, and if k is odd, we require Γ = −Λ; if k is even, (Global projection) uniform mesh (hj = h,∀j), α2 1 + β1β2 (cid:54)= 1 4, (cid:0)(−1)k+1 Γ 1 + β1β2 (cid:54)= 1 (cid:12)(cid:12)(cid:12) Γ • A2. 4 and Λ Λ Λ 21 we require Γ = Λ. (Global projection) uniform mesh (hj = h,∀j), α2 1 + β1β2 (cid:54)= 1 4, • A3. (cid:32) (−1)k+1 Γ Λ + (cid:114)(cid:16) Γ Λ (cid:33)N (cid:17)2 − 1 (cid:54)= 1. (cid:12)(cid:12)(cid:12) Γ Λ (cid:12)(cid:12)(cid:12) < 1, Lemma 2.2.2 (P (cid:63) h existence, uniqueness and formula). If any of the assumptions above is satisfied, P (cid:63) h exists and is uniquely defined. Furthermore, if assumption A0 is satisfied, then ´uj,k−1  = uj,k−1  + ´uj,k uj,k ∞(cid:88) m=k+1 uj,mMj,m. (2.26) If any of the assumptions A1/A2/A3 is satisfied, then ´uj,k−1  = uj,k−1  + ´uj,k uj,k ∞(cid:88) (cid:0)uj,mV1,m + m=k+1 N−1(cid:88) l=0 (cid:1), uj+l,mrlV2,m (2.27) k−1, L− where V1,m = [L+ when j + l ≥ N , and rl is defined in Table 2.1. m, V2,m = [L− k ]−1L+ k−1, L+ k ]−1L− m−[L+ k−1, L+ k ]−1L+ m, uj+l = uj+l−N Proof. The proof of this lemma can be found in Appendix. Lemma 2.2.3. Suppose any of the assumptions A0/A1/A2/A3 holds and u satisfies the condition in Definition 2.2.1. For p = 2,∞, if assumption A0 is satisfied, (cid:18) (cid:19) |β1|, | 1 2 +α1| h | 1 2−α1| , minj |Γj| h |β2| h2 , . (2.28) (cid:107)P (cid:63) h u − u(cid:107)Lp(I) ≤ Chk+1|u| W k+1,∞(I) max 22 If assumption A1 is satisfied, (cid:107)P (cid:63) h u − u(cid:107)Lp(I) ≤ Chk+1|u| W k+1,∞(I) (1 + (cid:107)Q1(cid:107)∞ |1 − |λ1|| + (cid:107)I2 − Q1(cid:107)∞ |1 − |λ2|| ), (2.29) where λ1, λ2 are defined in (49), Q1 is defined in (56) and (57). If assumption A2 is satisfied, (cid:107)P (cid:63) h u − u(cid:107)Lp(I) ≤ Chk+1(cid:107)u(cid:107) W k+4,∞(I) (1 + (cid:107)Q2(cid:107)∞ |Γ| ), (2.30) where Q2 is defined in (59). If assumption A3 is satisfied, and assuming (cid:12)(cid:12)(cid:12)1 − λN 1 (cid:12)(cid:12)(cid:12) = O(hδ(cid:48) ), |1 − λ1| = O(hδ/2) with 0 ≤ δ ≤ 2, (cid:107)P (cid:63) h u − u(cid:107)Lp(I) ≤ Chk+1(cid:107)u(cid:107) W k+3,∞(I) (1 + h−δ(cid:48)−δ/2((cid:107)Q1(cid:107)∞ + (cid:107)I2 − Q1(cid:107)∞)). (2.31) Proof. Proof is given in Appendix. Above estimates provides error bound that can be computed once the parameters are given, yet its dependence on the mesh size h is not fully revealed, particularly when the parameters α1, β1, β2 also have h-dependence. To clarify such relations, next we will inter- pret (2.29) when considering the following common choice of parameters, where α1 has no dependence on h, β1 = ˜β1hp1, β2 = ˜β2hp2, ˜β1, ˜β2 are nonzero constants that do not depend on h. If indeed β1 or β2 is zero, it is equivalent to let p1, p2 → +∞ in the discussions below. We will discuss whether the parameter choice yields optimal (k + 1)-th order accuracy. To distinguish different cases, we illustrate the choice of parameters p1, p2 in Figure 2.1. 23 Figure 2.1: A sketch to illustrate the different cases parametrized by the values of p1, p2. For example, A1.1 means p1 > −1, p2 > 1, A1.5 means p1 = −1, p2 = 1 and A1.7.1 means p1 > −1, p2 = 1. The main results are summarized in Table 2.2. Table 2.2: Interpretation of error estimate (2.29). 1 2 If k = 1 and p2 < 1, then P (cid:63) h is suboptimal and is (k + p2)-th order accurate, 3 else 4 5 6 7 8 9 end if limh→0 |λ1, λ2| = 1 with |λ1, λ2| = 1 + O(hδ/2), then h is suboptimal and is (k + 1 − δ)-th order accurate, P (cid:63) else end P (cid:63) h has optimal (k + 1)-th order error estimates. The main reason of order reduction for k = 1, p2 < 1 in line 2 of Table 2.2 is that the term such as 1|λ1|−1(cid:107)Q1(cid:107)∞ is of O(hp2−1) instead of O(1), and this will cause (1 − p2)-th order reduction. The situation happens for A1.3, A1.4 and A1.6.2 when k = 1. 24 −3−2−10p1 −1012 p2A1.1A1.2A1.3A1.4A1.5A1.6.1A1.6.2A1.7.1A1.7.2 1 1 1 1−|λ1|, 1−|λ1|, (cid:12)(cid:12)(cid:12) Γ in (2.29). The fractions By definition of λ1, λ2 in (49), we know that The main reason of order reduction in line 5 is because of the terms such as |1 1−|λ2| 1−|λ2| cannot be bounded by a constant if limh→0 |λ2| = 1. (cid:12)(cid:12)(cid:12) = 1 + O(hδ), δ > 0, then |λ1, λ2| = 1 + O(hδ/2), then (cid:12)(cid:12)(cid:12) → 1 ⇔ |λ1, λ2| → 1. More precisely, if 1−|λ2| = O(h−δ/2). The relation Γ2 − Λ2 = (b1 − b2)(b1 + b2) + c2 2 also indicates that there is some cancellation of leading terms in b1 − b2 or b1 + b2, making (cid:107)Q1(cid:107)∞ ∼ O(h−δ/2), multiplying these factors together will result in δ-th order reduction in the error estimation of P (cid:63) (cid:12)(cid:12)(cid:12) Γ 1−|λ1|, Λ Λ 1 1 h . Note that b1, b2, c2 and Q1 are defined in (47), (48), (45) and (56). Then we look at what parameter choices make (cid:12)(cid:12)(cid:12) Γ Λ (cid:12)(cid:12)(cid:12) → 1. Since k + 1 + β1+ k2(k2−1) h2 Λ β2− k2 h β1− 1 h Λ k > 1, k = 1, Γ Λ = we have Λ Λ (cid:12)(cid:12)(cid:12) Γ (cid:12)(cid:12)(cid:12)(cid:12) 1 2 +2α2 1 2−2α2 1 1 (cid:12)(cid:12)(cid:12) → 1k(k ∓ 1), (cid:12)(cid:12)(cid:12)(cid:12) = 1. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)k + (cid:12)(cid:12)(cid:12) Γ (cid:12)(cid:12)(cid:12) → (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)k + (cid:12)(cid:12)(cid:12) Γ (cid:12)(cid:12)(cid:12) → (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)k + (cid:12)(cid:12)(cid:12) → (cid:12)(cid:12)(cid:12) Γ (cid:12)(cid:12)(cid:12) → 1. (cid:12)(cid:12)(cid:12) →(cid:12)(cid:12)(cid:12)k + β1 Λ 2α2 1 k(k±1), Λ Λ 2 , (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) → 1. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) → 1. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) → 1. − k2 h β1− k2 h Λ k2(k2−1)β2 h2 Λ k2(k2−1)β2 h2 Λ 1. A1.1 (p1 > −1, p2 > 1) with k = 1, α1 = 0, 2. A1.6.1 (p1 = −1, p2 > 1) ˜β1 = k(k±1) 2 + 2α2 3. A1.6.2 (p1 = −1, p2 < 1) with k > 1, ˜β1 = k(k±1) 4. A1.7.1 (p1 > −1, p2 = 1) ˜β2 = 1 2k(k∓1) + 5. A1.7.2 (p1 < −1, p2 = 1) ˜β2 = 1 2k(k±1) , (cid:12)(cid:12)(cid:12) Γ Λ 25 Remark 2.2.1. We only considered T given by (51) (when Q1 is given by (56)) in the discussion above. By Appendix, we can conclude that under the parameter conditions in assumption A1, (b1 + b2)(b1 − b2) = 0 only can happen if p1 = −1, p2 = 1 with (29) or (30). This is A1.5, for which we always have optimal error estimate. Remark 2.2.2. Through numerical tests, we found that (2.29) is mostly sharp with two ex- ceptions. When limh→0 |λ1, λ2| = 1, the estimates show that there will be order reduction for error of P (cid:63) h , while in numerical experiments (see e.g. Tables 2.8, 2.9), such order reduction is observed only when limh→0 λ1, λ2 = 1 but not −1. We believe when limh→0 λ1, λ2 = −1, a refined estimate can be obtained similar to (2.30) under assumption A1. We have not carried out this estimate in this thesis. Another example we find for which (2.29) is not sharp is k = 2, p1 = −2 or −3, p2 = 1, (α1, ˜β1, ˜β2) = (0.25,−1, 1 2k(k+1) and λ1, λ2 → 1+O(h−(1+p1)/2). The theoretical results predict accuracy order of (k +2+p1) but numerical 12), where parameters belong to A1.7.2, ˜β2 = 1 experiments in Table 2.10 show the order to be (k + 3 + p1). Our estimations can’t resolve this one order difference. This special parameter may trigger a cancellation we didn’t capture in analysis. We will improve this estimate in our future work. Remark 2.2.3. In most cases, (2.30) yields optimal accuracy order, except when k = 1, α1 = 0, β1 = 0, β2 = O(hp2), p2 < 1, where the P (cid:63) (cid:107)Q2(cid:107)∞ h2 β2+ 1 h is only (k + p2)-th order accurate because ∼ O(hp2−1) in (2.30). This is verified numerically in |−−4 | 2h |Λ| = |b1+b2| |Λ| = 1 2h Table 2.12. Remark 2.2.4. If δ/2 > 1, we can show δ/2 = δ(cid:48) + 1. This is because |1− λ1| = |1− eiθ| = 1 | = |1 − eiN θ| = 2| sin(N θ/2)|. When δ/2 > 1, one can assert that 2| sin(θ/2)|, and |1 − λN 1 | ∼ N θ, i.e. δ/2 = δ(cid:48) + 1. With this condition, we notice that (2.31) |1 − λ1| ∼ θ,|1 − λN 26 yields an reduction of δ-th order in convergence rate by checking the order of each term. This order reduction is consistent with numerical experiments in Example 2.4.4. Now we can summarize the estimation of P (cid:63) h for some frequently used flux parameters. For IPDG scheme with α1 = β2 = 0, β1 = c/h, and DDG scheme discussed in [52] with α1 = constant, β1 = c/h, β2 = 0, and the more general scale invariant parameter choice α1 = constant, β1 = c/h, β2 = ch, P (cid:63) h always have optimal error estimates. For those parameters, we can show that the eigenvalues λ1, λ2 are always constants independent of h, therefore, by Lemma 2.2.3, we will have optimal convergence rate. Corresponding numerical results are shown in Tables 2.4 and 2.7. For a natural parameter choice where α1, β1, β2 are all real constants, if β2 (cid:54)= 0, then P (cid:63) h has first order convergence rate when k = 1 and optimal convergence rate when k > 1 by Lemma 2.2.3. Corresponding numerical results are shown in Tables 2.3 and 2.12. Lastly, for central flux α1 = α2 = β1 = β2 = 0, this parameter choice satisfies assumption A3 when k = 1 and assumption A2 when k > 1, thus we can verify that P (cid:63) h has optimal convergence rate. Corresponding numerical results are shown in Table 2.11. 2.3 Error estimates In this section, we will derive error estimates of the DG scheme (2.3) for the model NLS equation (1.1). We will focus on the impact of the choice of the parameters α1, β1, β2 on the accuracy of the scheme. The error estimates rely on the projection error estimates to obtain convergence result. Theorem 2.3.1. Assume that the exact solution u and the nonlinear term f (|u|2) of (1.1) are sufficiently smooth with bounded derivatives for any time t ∈ (0, Te] and that the nu- 27 merical flux parameters in (2.13) satisfy the existence conditions of P (cid:63) h in Lemma 2.2.2. h u has at least first order convergence rate in L2 and L∞ Furthermore, assume h = u − P (cid:63) norm from the results in Lemma 2.2.3. With periodic boundary conditions solution space V k h (k ≥ 1), the following error estimation holds for uh, which is the numerical solution of (2.3) with flux (2.13): (cid:107)u − uh(cid:107)L2(I) ≤ C(cid:63) ((cid:107)(u − uh)|t=0(cid:107) + (cid:107)(h)t(cid:107) + (cid:107)h(cid:107)) , (2.32) where C(cid:63) depends on k,(cid:107)f(cid:107)W 2,∞, u as well as final time Te, but not on h. Proof. When P (cid:63) h exists, we can decompose the error into two parts. e = u − uh = u − P (cid:63) h u + P (cid:63) h u − uh := h + ζh. By Galerkin orthogonality, ∀vh ∈ V k h , (cid:90) f (|u|2)uvhdx + i 0 = a(e, vh) − i = a(h, vh) + a(ζh, vh) − i I (cid:90) I (cid:90) I f (|uh|2)uhvhdx (cid:90) I f (|u|2)uvhdx + i f (|uh|2)uhvhdx. By letting vh = ζh and taking conjugate of above equation, we have a(ζh, ζh) + a(ζh, ζh) = − a(h, ζh) − a(h, ζh) − 2 (cid:90) I f (|u|2)Im(uζh)dx + 2 (cid:90) I (2.33) f (|uh|2)Im(uhζh)dx. 28 By Taylor expansion f (|uh|2) = f (|u|2) + f(cid:48)(|u|2)E + ˆf(cid:48)(cid:48)E2, 1 2 where ˆf(cid:48)(cid:48) = f(cid:48)(cid:48)(c), c is a value between |uh|2 and |u|2. E = |uh|2 − |u|2 = −2Re(eu) + |e|2. Therefore, the nonlinear part becomes (cid:90) (cid:90) (cid:90) = f (|u|2)Im(uζh)dx − f (|uh|2)Im(cid:0)eζh f (|uh|2)Im(uhζh)dx (cid:1) +(cid:0)f (|u|2) − f (|uh|2)(cid:1)Im(uζh)dx I I I =N1 + N2 + N3, where N1 = N2 = N3 = (cid:90) (cid:90) (cid:90) I I I f (|u|2)Im(cid:0)eζh (cid:1) − f(cid:48)(|u|2)EIm(uζh)dx, (cid:1) − 1 f(cid:48)(|u|2)EIm(cid:0)eζh ˆf(cid:48)(cid:48)E2Im(cid:0)eζh (cid:1)dx, ˆf(cid:48)(cid:48)E2Im(uζh)dx, 2 1 2 will be estimated separately as follows. • N1 and N2 terms. Since eζh = hζh + |ζh|2,(cid:12)(cid:12)EIm(uζh)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(−2Re(eu) + |e|2)Im(uζh) (cid:12)(cid:12)(cid:12) ≤ C((cid:107)u(cid:107)2 L∞(I) + 29 (cid:107)u(cid:107)L∞(I)(cid:107)e(cid:107)L∞(I))(|h|2 + |ζh|2), we have (cid:16) |N1| ≤ C(cid:107)f(cid:107)W 1,∞ 1 + (cid:107)u(cid:107)2 |N2| ≤ C(cid:107)f(cid:107)W 2,∞(cid:107)E(cid:107)L∞(I) (cid:16) L∞(I) + (cid:107)u(cid:107)L∞(I)(cid:107)e(cid:107)L∞(I) 1 + (cid:107)u(cid:107)2 L∞(I) + (cid:107)u(cid:107)L∞(I)(cid:107)e(cid:107)L∞(I) (cid:17) ((cid:107)h(cid:107)2 + (cid:107)ζh(cid:107)2), (cid:17) ((cid:107)h(cid:107)2 + (cid:107)ζh(cid:107)2). • N3 term. |N3| ≤ C(cid:107)f(cid:48)(cid:48)(cid:107)L∞(cid:107)E(cid:107)2 L∞(I)((cid:107)h(cid:107)2 + (cid:107)ζh(cid:107)2). To conduct a proper estimate for the nonlinear part, we would like to make an a priori assumption that, for h small enough, (cid:107)e(cid:107) = (cid:107)u − uh(cid:107) ≤ h0.5. (2.34) h , (cid:107)h(cid:107)Lp(I) ≤ C1h, p = 2,∞, thus (cid:107)ζh(cid:107) ≤ C1h0.5 and (cid:107)ζh(cid:107)L∞(I) ≤ By our assumption on P (cid:63) C1 by inverse inequality, then (cid:107)e(cid:107)L∞(I) ≤ C1, (cid:107)E(cid:107)L∞(I) ≤ C1. Here and below, C1 is a generic constant that has no dependence on h, but may depend on u according to the lemma used to estimate h. Therefore, we get the estimate: |N1| + |N2| + |N3| ≤ C1((cid:107)h(cid:107)2 + (cid:107)ζh(cid:107)2), (2.35) where C1 depends on (cid:107)f(cid:107)W 2,∞ and u. 30 For linear part of the right hand side in (2.33), we have a(h, ζh) + a(h, ζh) = (cid:90) ((cid:98)h[(ζh)x] −(cid:103)(h)x[ζh])| (h)(ζh)xxdx I j+ 1 2 I (cid:90) (cid:90) (h)(ζh)xxdx − i N(cid:88) (cid:90) (h)tζh + (h)tζhdx − i N(cid:88) ((cid:98)h[(ζh)x] −(cid:103)(h)x[ζh])| Re(cid:0)(h)tζh (cid:1)dx. j=1 j=1 I + i + i = 2 I , j+ 1 2 The last equality holds because of the definition of P (cid:63) h u. For the left hand side of (2.33), by similar computation in stability analysis we have a(ζh, ζh) + a(ζh, ζh) = Combine these two equations with (2.35): (cid:90) I d dt |ζh|2dx. (2.36) (cid:107)ζh(cid:107)2 ≤ (cid:107)(h)t(cid:107)2 + (cid:107)ζh(cid:107)2 + C1((cid:107)h(cid:107)2 + (cid:107)ζh(cid:107)2). d dt Assuming ut, u have sufficient smoothness, then by Gronwall’s inequality, we can get: (cid:16)(cid:107)ζh|t=0(cid:107)2 L2(I) + (cid:107)(h)t(cid:107)2 + (cid:107)(h)(cid:107)2(cid:17) , (cid:107)ζh(cid:107)2 ≤ C1 and we obtain (2.32). To complete the proof, we shall justify the a priori assumption. To be more precise, we consider h0, s.t., ∀h < h0, C(cid:63)h ≤ 1 2 h0.5, where C(cid:63) is defined in (2.32), dependent on Te, but not on h. Suppose ∃ t∗ = sup{t : (cid:107)u(t∗)− uh(t∗)(cid:107) ≤ h0.5}, we would have (cid:107)u(t∗)− uh(t∗)(cid:107) = 31 h0.5 by continuity if t∗ is finite. By (2.32), we obtain (cid:107)e(cid:107) ≤ C(cid:63)h ≤ 1 2 h0.5 if t∗ ≤ Te, which contradicts the definition of t∗. Therefore, t∗ > Te and the a priori assumption is justified. Remark 2.3.1. If f is a constant function, we can prove the same error estimates without using the a prixori assumption. Therefore, the assumption that h = u − P (cid:63) first order convergence rate in L2 and L∞ norm is no longer needed. h u has at least Moreover, the estimates for (cid:107)h(cid:107) has been established in Lemma 2.2.3. In other words, the error of the DG scheme (2.3) has the same accuracy as P (cid:63) h u, as long as P (cid:63) h u is well-defined and the numerical initial condition is chosen sufficiently accurate. 2.4 Numerical results In this subsection, we present numerical experiments to validate our theoretical results. Particularly, in Section 2.4.1, we provide numerical validations of convergence rate for the projection P (cid:63) h as discussed in Lemma 2.2.3 with focus on the dependence of the errors on parameters α1, β1, β2 . Section 2.4.2 illustrates the energy conservation property and validates theoretical convergence rate of DG scheme for NLS equation (1.1). 2.4.1 Numerical results of the projection operator P (cid:63) h Example 2.4.1. In this example, we focus on local projection where α2 4, and verify the conclusions in Lemma 2.2.1 by considering a smooth test function u = cos(x), x ∈ 1 + β1β2 = 1 [0, 2π] on a nonuniform mesh and k = 1, 2, 3 for various sets of parameters (α1, β1, β2). The nonuniform mesh is generated by perturbing the nodes of a uniform mesh of N cells by at most 10%. 32 We first consider two sets of parameters (α1, β1, β2) = (0.3, 0.4, 0.4) and (α1, β1, β2) = (0.3, 0.4/h, 0.4h). The results with (α1, β1, β2) = (0.3, 0.4, 0.4) are listed in Table 2.3. By plugging in the parameters into (2.28), we have that when k = 1, the projection has sub- optimal first order convergence rate, while for k > 1, optimal (k + 1)-th order convergence rate should be achieved. For k = 1, Γj = β1 − 1 h , which does not depend on β2 any more. This technical difference cause the discrepancy of the convergence order between k = 1 and k > 1 in Table 2.3. Results in Table 2.3 agree well with the theoretical prediction. On the other hand, when we choose parameters (α1, β1, β2) = (0.3, 0.4/h, 0.4h), by Lemma 2.2.3, we should observe optimal convergence rate for all k ≥ 1, and this is verified by the numerical results in Table 2.4. In [16], we also proved that P (cid:63) 2 and β2/Γj ∼ o(1). We choose the parameters as (α1, β1, β2) = (0.5, 1, 0) to verify this claim, i.e., the h when β2 = 0, α1 = ± 1 h is superclose to P 1 difference between P (cid:63) h and P 1 h can have convergence rates higher than k + 1. The results are listed in Table 2.5. The difference of the two projections is indeed of (k + 2)-th order for any k ≥ 1 in all norms. Table 2.3: Example 2.4.1. Error of local projection P (cid:63) parameters: α1 = 0.3, β1 = 0.4, β2 = 0.4. h u − u on a nonuniform mesh. Flux N 160 320 640 1280 160 320 640 1280 160 320 640 1280 L1 error 1.98E-02 9.98E-03 5.01E-03 2.51E-03 2.18E-06 2.71E-07 3.37E-08 4.19E-09 2.82E-09 1.76E-10 1.10E-11 6.86E-13 P 1 P 2 P 3 - - 0.99 0.99 1.00 order L2 error 1.56E-02 7.87E-03 3.95E-03 1.98E-03 1.91E-06 2.39E-07 2.97E-08 3.69E-09 2.45E-09 1.53E-10 9.50E-12 5.93E-13 3.01 3.01 3.01 4.00 4.00 4.00 - - - 0.99 0.99 1.00 order L∞ error 1.81E-02 9.20E-03 4.55E-03 2.27E-03 3.73E-06 5.14E-07 6.71E-08 7.99E-09 5.67E-09 3.76E-10 2.25E-11 1.46E-12 3.00 3.01 3.01 4.00 4.01 4.00 - order - 0.97 1.02 1.00 - 2.86 2.94 3.07 - 3.92 4.06 3.95 33 Table 2.4: Example 2.4.1. Error of local projection P (cid:63) parameters: α1 = 0.3, β1 = 0.4/h, β2 = 0.4h. h u − u on a nonuniform mesh. Flux N 160 320 640 1280 160 320 640 1280 160 320 640 1280 L1 error 3.42E-04 8.55E-05 2.14E-05 5.34E-06 6.36E-06 8.17E-07 1.02E-07 1.27E-08 3.32E-09 2.08E-10 1.30E-11 8.09E-13 P 1 P 2 P 3 - - 2.00 2.00 2.00 order L2 error 3.50E-04 8.75E-05 2.19E-05 5.47E-06 6.06E-06 7.99E-07 1.00E-07 1.24E-08 2.93E-09 1.83E-10 1.14E-11 7.12E-13 4.00 4.00 4.00 2.96 3.00 3.01 - - - 2.00 2.00 2.00 order L∞ error 8.62E-04 2.21E-04 5.45E-05 1.36E-05 2.06E-05 3.09E-06 4.51E-07 5.12E-08 7.58E-09 5.08E-10 3.04E-11 1.99E-12 4.00 4.01 4.00 2.92 2.99 3.02 - order - 1.96 2.02 2.00 - 2.73 2.78 3.14 - 3.90 4.06 3.93 Table 2.5: Example 2.4.1. Difference of local projection P (cid:63) nonuniform mesh. Flux parameters: α1 = 0.5, β1 = 1, β2 = 0. h with P 1 h : P (cid:63) h u − P 1 h u on a N 160 320 640 1280 160 320 640 1280 160 320 640 1280 L1 error 2.09E-05 2.56E-06 3.17E-07 3.94E-08 5.00E-09 3.14E-10 1.96E-11 1.22E-12 2.91E-12 9.11E-14 2.84E-15 8.84E-17 P 1 P 2 P 3 - - 3.03 3.01 3.01 order L2 error 1.96E-05 2.40E-06 2.96E-07 3.67E-08 5.05E-09 3.21E-10 2.00E-11 1.24E-12 3.38E-12 1.06E-13 3.27E-15 1.02E-16 3.99 4.00 4.01 5.00 5.00 5.00 - - - 3.03 3.02 3.01 order L∞ error 4.66E-05 5.99E-06 7.17E-07 9.11E-08 1.82E-08 1.28E-09 8.56E-11 5.02E-12 1.40E-11 4.72E-13 1.40E-14 4.63E-16 3.98 4.00 4.01 5.00 5.01 5.00 - order - 2.96 3.06 2.98 - 3.83 3.90 4.09 - 4.89 5.08 4.92 Example 2.4.2. In this example, we consider global projection when the parameter choices satisfy assumption A1. We consider a smooth test function u = ecos(x) on [0, 2π] with a uniform mesh of size h = 2π/N and k = 1, 2, 3 for various sets of parameters (α1, β1, β2). We first test the situation when limh→0 |λ1, λ2| (cid:54)= 1 by setting the parameters (α1, ˜β1, ˜β2) = (0.25, 1, 1), p1 = −0.5, p2 = 2. Another example is (α1, β1, β2) = (0, 1 2h , h), for which the eigenvalues λ1, λ2 are constant dependent on k but not h. These two parameter choices belong to A1.1 and A1.5, respectively. The numerical results shown in Tables 2.6 and 2.7 34 verify the optimal (k + 1)-th order convergence rate predicted by (2.29). ters (α1, ˜β1, ˜β2) = (0.25, k(k−1) Then we test the situation when limh→0 |λ1, λ2| = 1 by using two sets of parame- , 1), p1 = −1, p2 = 2, 3, and (α1, ˜β1, ˜β2) = k(k−1), 1), p1 = −2,−3, p2 = 1. The first set of parameters belongs to A1.6.1 and (0.25, we can verify that limh→0 λ1, λ2 = (−1)k. (2.29) and Algorithm 2.2 imply (k + 2 − p2)-th + k(k+1) 2 2 8 convergence order. The numerical results listed in Table 2.8 show that the expected order reduction only happens when limh→0 λ1, λ2 = 1, but not for limh→0 λ1, λ2 = −1. The second set of parameters belongs to A1.7.2 and we can verify that limh→0 λ1, λ2 = (−1)k+1. (2.29) and Algorithm 2.2 imply (k + 2 + p1)-th convergence order. The numerical results listed in Table 2.9 also show that order reduction is only observed when limh→0 λ1, λ2 = 1. Lastly, we test (α1, ˜β1, ˜β2) = (0.25,−1, 1 12) with k = 2, p1 = −2,−3, p2 = 1, where our theoretical results predict accuracy order of (k + 2 + p1), but numerical experiments show the order to be (k + 3 + p1) in Table 2.10. This is one of the exceptions that (2.29) is not sharp and has been commented in Remark 2.2.2. Table 2.6: Example 2.4.2. Error of global projection P (cid:63) 0.25, ˜β1 = 1, ˜β2 = 1, p1 = −0.5, p2 = 2. (A1.1) h u − u. Flux parameters: α1 = N 160 320 640 1280 160 320 640 1280 320 640 1280 2560 L1 error 0.10E-03 0.26E-04 0.67E-05 0.17E-05 0.63E-06 0.88E-07 0.11E-07 0.14E-08 0.64E-10 0.45E-11 0.29E-12 0.19E-13 P 1 P 2 P 3 - - 1.93 1.98 1.99 order L2 error 0.69E-03 0.18E-03 0.46E-04 0.12E-04 0.52E-05 0.71E-06 0.91E-07 0.11E-07 0.49E-09 0.35E-10 0.23E-11 0.15E-12 2.85 2.95 2.99 3.82 3.93 3.97 - - - 1.93 1.97 1.99 order L∞ error 0.89E-03 0.23E-03 0.58E-04 0.15E-04 0.87E-05 0.11E-05 0.14E-06 0.17E-07 0.72E-09 0.52E-10 0.34E-11 0.22E-12 2.88 2.97 2.99 3.80 3.91 3.96 - order - 1.94 1.98 2.00 - 2.95 3.00 3.01 - 3.79 3.92 3.96 Example 2.4.3. In this example, we consider global projection when the parameter choices 35 Table 2.7: Example 2.4.2. Error of global projection P (cid:63) 1 2h, β2 = h. (A1.5) h u−u. Flux parameters: α1 = 0, β1 = N 320 640 1280 2560 320 640 1280 2560 320 640 1280 2560 L1 error 0.11E-03 0.28E-04 0.70E-05 0.18E-05 0.11E-06 0.14E-07 0.18E-08 0.22E-09 0.38E-10 0.24E-11 0.15E-12 0.92E-14 P 1 P 2 P 3 - - 2.00 2.00 2.00 order L2 error 0.63E-03 0.16E-03 0.39E-04 0.98E-05 0.71E-06 0.89E-07 0.11E-07 0.14E-08 0.25E-09 0.16E-10 0.99E-12 0.62E-13 3.00 3.00 3.00 4.00 4.00 4.00 - - - 2.00 2.00 2.00 order L∞ error 0.38E-03 0.95E-04 0.24E-04 0.60E-05 0.62E-06 0.77E-07 0.96E-08 0.12E-08 0.22E-09 0.14E-10 0.86E-12 0.54E-13 4.00 4.00 4.00 3.00 3.00 3.00 - order - 2.00 2.00 2.00 - 3.00 3.00 3.00 - 4.00 4.00 3.99 Table 2.8: Example 2.4.2. Error of global projection P (cid:63) 0.25, ˜β1 = k(k−1) h u − u. Flux parameters: α1 = , ˜β2 = 1.0, p1 = −1, p2 = 2, 3. Note here limh→0 λ1, λ2 = (−1)k. 2 + k(k+1) 8 (A1.6.1) P 1 p2 = 2 ˜β1 = 1 4 P 2 p2 = 2 ˜β1 = 7 4 P 2 p2 = 3 ˜β1 = 7 4 P 3 p2 = 2 ˜β1 = 9 2 N 640 1280 2560 5120 640 1280 2560 5120 640 1280 2560 5120 320 640 1280 2560 L1 error 0.75E-05 0.19E-05 0.48E-06 0.12E-06 0.15E-06 0.39E-07 0.98E-08 0.25E-08 0.14E-04 0.71E-05 0.35E-05 0.18E-05 0.12E-09 0.78E-11 0.49E-12 0.31E-13 - - 1.97 1.99 1.99 1.94 1.97 1.98 order L2 error 0.52E-04 0.13E-04 0.34E-05 0.84E-06 0.12E-05 0.32E-06 0.82E-07 0.21E-07 0.12E-03 0.58E-04 0.29E-04 0.15E-04 0.95E-09 0.60E-10 0.38E-11 0.24E-12 1.00 1.00 1.00 3.99 3.99 4.00 - - - 1.97 1.98 1.99 1.93 1.97 1.98 order L∞ error 0.66E-04 0.17E-04 0.42E-05 0.11E-05 0.23E-05 0.61E-06 0.16E-06 0.39E-07 0.21E-03 0.11E-03 0.54E-04 0.27E-04 0.20E-08 0.13E-09 0.80E-11 0.51E-12 1.00 1.00 1.00 3.99 3.99 3.99 - - order - 1.97 1.99 1.99 - 1.94 1.97 1.99 - 1.00 1.00 1.00 - 3.99 3.99 3.97 are similar to central fluxes, and satisfy assumptions A1 and A2, for smooth function u = ecos(x) on [0, 2π] with a uniform mesh of size h = 2π/N and k = 1, 2, 3. For central flux (α1, β1, β2) = (0, 0, 0), Γ = − k2 |Γ| |Λ| = k > 1, flux parameters satisfy to assumption A1, and if k = 1, Γ = −Λ and flux parameters satisfy 2h . If k > 1, 2h , Λ = k 36 Table 2.9: Example 2.4.2. Error of global projection P (cid:63) ˜β1 = 1, ˜β2 = 2k(k−1), p1 = −2,−3, p2 = 1. Note here limh→0 λ1, λ2 = (−1)k+1. (A1.7.2) h u − u. Flux parameters: α1 = 0.25, 1 P 2 p1 = −3 ˜β2 = 1 4 P 3 p1 = −2 ˜β2 = 1 12 P 3 p1 = −3 ˜β2 = 1 12 N 320 640 1280 2560 320 640 1280 2560 320 640 1280 2560 L1 error 0.28E-07 0.35E-08 0.44E-09 0.55E-10 0.70E-08 0.94E-09 0.12E-09 0.15E-10 0.16E-06 0.40E-07 0.10E-07 0.25E-08 - - 3.00 3.00 3.00 order L2 error 0.21E-06 0.27E-07 0.33E-08 0.41E-09 0.57E-07 0.77E-08 0.99E-09 0.13E-09 0.13E-05 0.32E-06 0.79E-07 0.20E-07 2.90 2.95 2.98 2.00 2.00 2.00 - - - 3.00 3.00 3.00 order L∞ error 0.24E-06 0.31E-07 0.38E-08 0.48E-09 0.12E-06 0.16E-07 0.20E-08 0.26E-09 0.24E-05 0.61E-06 0.15E-06 0.38E-07 2.00 2.00 2.00 2.90 2.95 2.98 - order - 3.00 3.00 3.00 - 2.91 2.95 2.98 - 2.00 2.00 2.00 Table 2.10: Example 2.4.2. Error of global projection P (cid:63) 0.25, ˜β1 = −1, ˜β2 = 2k(k+1), p1 = −2,−3, p2 = 1. Note that limh→0 λ1, λ2 = 1. (A1.7.2) h u − u. Flux parameters: α1 = 1 N 320 640 1280 2560 320 640 1280 2560 L1 error 0.72E-07 0.90E-08 0.11E-08 0.14E-09 0.80E-06 0.20E-06 0.50E-07 0.13E-07 order L2 error 0.56E-06 2.99 2.99 0.71E-07 0.89E-08 3.00 0.11E-08 3.00 2.01 0.63E-05 0.16E-05 2.00 0.39E-06 2.00 2.00 0.98E-07 order L∞ error 0.94E-06 2.98 2.99 0.12E-06 0.15E-07 3.00 0.19E-08 3.00 2.01 0.12E-04 0.30E-05 2.00 0.75E-06 2.00 2.00 0.19E-06 order 2.97 2.99 2.99 3.00 2.01 2.00 2.00 2.00 P 2 p1 = −2 ˜β2 = 1 12 P 2 p1 = −3 ˜β2 = 1 12 to assumption A2. We conclude that P (cid:63) h exists and is unique for k = 1 when N is odd and k > 1 for arbitrary N. P (cid:63) h has optimal error estimates as proved in Lemma 2.2.3. Our numerical test in Table 2.11 demonstrates optimal convergence rate for all k. A similar flux is (α1, β1, β2) = (0, 0, 1). When k = 1, this flux parameter set satisfies assumption A2 and (2.30) yields first order convergence rate as discussed in Remark 2.2.3. When k = 2, 3, similar to central flux, this parameter choice satisfies assumption A1, showing optimal convergence rate. The numerical test in Table 2.12 verifies the theoretical results. Example 2.4.4. In this example, we consider global projection when the parameter choices 37 Table 2.11: Example 2.4.3. Error of global projection P (cid:63) eters: α1 = 0, β1 = 0, β2 = 0. h u − u. (Central flux) Flux param- N 93 279 837 2511 160 320 640 1280 160 320 640 1280 L1 error 0.12E-03 0.13E-04 0.15E-05 0.17E-06 0.11E-05 0.14E-06 0.17E-07 0.22E-08 0.11E-08 0.68E-10 0.42E-11 0.27E-12 P 1 P 2 P 3 - - 2.00 2.00 2.00 order L2 error 0.74E-03 0.82E-04 0.91E-05 0.10E-05 0.85E-05 0.11E-05 0.13E-06 0.17E-07 0.83E-08 0.52E-09 0.32E-10 0.20E-11 4.00 4.00 4.00 3.00 3.00 3.00 - - - 2.00 2.00 2.00 order L∞ error 0.55E-03 0.61E-04 0.68E-05 0.76E-06 0.10E-04 0.13E-05 0.16E-06 0.20E-07 0.11E-07 0.68E-09 0.42E-10 0.26E-11 4.00 4.00 4.00 3.00 3.00 3.00 - order - 2.00 2.00 2.00 - 2.99 3.00 3.00 - 4.00 4.00 4.00 Table 2.12: Example 2.4.3. Error of global projection P (cid:63) 0, β1 = 0, β2 = 1. h u − u. Flux parameters: α1 = N 93 279 837 2511 160 320 640 1280 2560 160 320 640 1280 L1 error 0.21E-01 0.72E-02 0.24E-02 0.80E-03 0.11E-05 0.14E-06 0.17E-07 0.22E-08 0.27E-09 0.27E-08 0.17E-09 0.11E-10 0.66E-12 order - 1.00 1.00 1.00 - 3.00 3.00 3.00 3.00 - 4.00 4.00 4.00 L2 error 0.12E+00 0.40E-01 0.13E-01 0.44E-02 0.86E-05 0.11E-05 0.13E-06 0.17E-07 0.21E-08 0.23E-07 0.14E-08 0.89E-10 0.55E-11 P 1 P 2 P 3 - - 1.00 1.00 1.00 order L∞ error 0.68E-01 0.23E-01 0.75E-02 0.25E-02 0.10E-04 0.13E-05 0.16E-06 0.20E-07 0.25E-08 0.36E-07 0.22E-08 0.14E-09 0.87E-11 3.00 3.00 3.00 3.00 4.00 4.00 4.00 - order - 1.00 1.00 1.00 - 3.00 3.00 3.00 3.00 - 4.00 4.00 4.00 satisfy assumption A3 for the smooth function u = ecos(x) on [0, 2π] with uniform mesh size h = 2π/N and k = 1, 2, 3. 1 (0.25,−1, An example of A3 is shown in Table 2.13, where the parameters are (α1, ˜β1, ˜β2) = 2k(k−1)), p1 = −2,−3, p2 = 1, similar to the parameters in Table 2.9. The asymptotic behavior of λ1, λ2 when h approaches 0 is indeed similar to Table 2.9, that is, |λ1, λ2| = 1 + O(h−(p1+1)/2) and limh→0 λ1, λ2 = (−1)k+1. Same as previous examples, 38 order reductions are only observed when limh→0 λ1, λ2 = 1, that is for k = 3. We performed more numerical results under assumption A3, and all are similar to those of A1 as long as the eigenvalues λ1, λ2 are approaching 1 at the same rate. Hence, we will not show more examples under assumption A3. - P 2 p1 = −3 ˜β2 = 1 4 Table 2.13: Example 2.4.4. Error of global projection P (cid:63) similar to A1.7.2 in Table 2.9): α1 = 0.25, ˜β1 = −1, ˜β2 = Note here limh→0 λ1, λ2 = (−1)k+1. L1 error 0.28E-07 0.35E-08 0.44E-09 0.55E-10 0.70E-08 0.94E-09 0.12E-09 0.15E-10 0.16E-06 0.40E-07 0.10E-07 0.25E-08 order L2 error 0.21E-06 0.27E-07 0.33E-08 0.41E-09 0.57E-07 0.77E-08 0.99E-09 0.13E-09 0.13E-05 0.32E-06 0.79E-07 0.20E-07 N 320 640 1280 2560 320 640 1280 2560 320 640 1280 2560 P 3 p1 = −2 ˜β2 = 1 12 P 3 p1 = −3 ˜β2 = 1 12 3.00 3.00 3.00 2.90 2.95 2.98 2.00 2.00 2.00 - - h u − u. Flux parameters (A3, and 2k(k−1) , p1 = −2,−3, p2 = 1. 1 - - 3.00 3.00 3.00 order L∞ error 0.24E-06 0.31E-07 0.38E-08 0.48E-09 0.12E-06 0.16E-07 0.20E-08 0.26E-09 0.24E-05 0.61E-06 0.15E-06 0.38E-07 2.00 2.00 2.00 2.90 2.95 2.98 - order - 3.00 3.00 3.00 - 2.91 2.95 2.98 - 2.00 2.00 2.00 2.4.2 Numerical results of the DG scheme In this subsection, we show the numerical results of the DG scheme applied to the NLS equation. For the time discretization, we use third order IMEX Runge-Kutta method [5] and fix ∆t = 1/10000, which is small enough to guarantee that the spatial errors dominate. To be more precise, we treat the DG discretization of linear term uxx implicitly and nonlinear term f (|u|2)u explicitly. Example 2.4.5. In this example, we verify the energy conservation property of our scheme 39 by considering the following linear equation iut + uxx = 0, with the progressive plane wave solution: u(x, t) = Aexp(i(x − t)), with A = 1. We use L2 projection as the numerical initial condition. In the discussion of stabil- ity condition, we derive that when Imβ2 ≥ 0, Imβ1 ≤ 0,|α1 + α2|2 ≤ −4Imβ1Imβ2, our scheme for Schr¨odinger equation is stable. Furthermore, when α1 + α2 = 0, β1, β2 are real numbers, the scheme is energy conservative. In this example, we compare two different parameter choices to verify the energy conservation property. The parameter choices are (α1, α2, β1, β2) = (0.25,−0.25, 1 − i, 1 + i), and (α1, α2, β1, β2) = (0.25,−0.25, 1, 1) when k = 2, N = 40, ending time T = 100. Both are numerically stable flux parameters. For the first set of parameters, we expect energy decay due to the contributions from the imaginary part of β1, β2 as in (2.11). For the second set of parameter, energy should be conserved. In Fig. 2.2, we verify that as t increases from 0 to 100, the flux with only real parameters preserve (cid:107)uh(cid:107), while the flux with complex numbers have much larger errors. More precisely, for real parameters, (cid:107)uh(0,·)(cid:107)−(cid:107)uh(100,·)(cid:107) = 7.9E-09, for complex parameters, (cid:107)uh(0,·)(cid:107)− (cid:107)uh(100,·)(cid:107) = 5.7E-04. Example 2.4.6. Accuracy test for NLS equation iut + uxx + |u|2u + |u|4u = 0, (2.37) which admits a progressive plane wave solution: u(x, t) = Aexp(i(cx − ωt)), where ω = c2 − |A|2 − |A|4 with c = 1, A = 1. 40 Figure 2.2: Example 2.4.5. Absolute difference of (cid:107)uh(t,·)(cid:107) with (cid:107)uh(0,·)(cid:107) with two sets of parameters (α1, α2, β1, β2) = (0.25,−0.25, 1 − i, 1 + i) (denoted by “imag”) and (α1, α2, β1, β2) = (0.25,−0.25, 1, 1) (denoted by “real”) when k = 2, N = 40, ending time Te = 100. For numerical initial condition, P (cid:63) h is used when applicable, otherwise standard L2 pro- jection is applied. On uniform mesh, we use four sets of parameters. The numerical errors and orders are shown in Tables 2.14 - 2.19, where corresponding projection results are listed in Tables 2.3, 2.4, 2.11, 2.12, 2.8 and 2.10 respectively. Our numerical experiments show that the order of convergence for the scheme is the same as the order of error estimates for the projection P (cid:63) h . We would like to make some additional comments on Tables 2.16 and 2.17, whose pa- rameter choices satisfy assumption A2 when k = 1. The existence of P (cid:63) h requires N to be odd for this parameter assumption. However, this assumption is not needed for the optimal convergence rate of the numerical scheme for (2.37) as shown in Tables 2.16 and 2.17. Similar comments have been made in [6]. 41 Table 2.14: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) on a nonuniform mesh using flux parameters (corresponding to Table 2.3) α1 = 0.3, β1 = β2 = 0.4, ending time Te = 0.3. N 40 80 160 320 640 40 80 160 320 40 80 160 320 L1 error 2.86E-02 1.26E-02 6.34E-03 3.18E-03 1.58E-03 2.22E-04 1.99E-05 3.17E-06 3.49E-07 1.54E-06 4.96E-08 2.81E-09 1.61E-10 P 1 P 2 P 3 - - 1.18 1.00 1.00 1.01 order L2 error 2.48E-02 1.02E-02 4.99E-03 2.56E-03 1.27E-03 2.13E-04 2.13E-05 3.03E-06 3.34E-07 1.35E-06 4.36E-08 2.60E-09 1.57E-10 4.96 4.14 4.13 3.48 2.65 3.18 - - - 1.28 1.03 0.96 1.01 order L∞ error 3.92E-02 1.56E-02 6.77E-03 3.47E-03 1.85E-03 6.06E-04 7.28E-05 9.01E-06 1.23E-06 3.29E-06 1.29E-07 8.37E-09 7.68E-10 4.95 4.07 4.05 3.33 2.81 3.18 - order - 1.33 1.20 0.96 0.91 - 3.06 3.02 2.87 - 4.67 3.95 3.45 Table 2.15: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) on a nonuniform mesh using flux parameters (corresponding to Table 2.4) α1 = 0.3, β1 = 0.4h, β2 = 0.4/hj, ending time Te = 1. N 40 80 160 320 640 40 80 160 320 640 40 80 160 320 L1 error 7.47E-03 2.10E-03 4.82E-04 1.21E-04 3.12E-05 5.14E-04 6.81E-05 8.04E-06 9.53E-07 1.68E-07 1.30E-06 5.74E-08 4.44E-09 2.25E-10 P 1 P 2 P 3 - - 1.83 2.12 1.99 1.96 order L2 error 6.50E-03 1.76E-03 4.18E-04 1.05E-04 2.71E-05 5.37E-04 7.00E-05 8.06E-06 9.75E-07 1.61E-07 1.25E-06 6.00E-08 4.12E-09 2.13E-10 2.92 3.08 3.08 2.50 4.51 3.69 4.30 - - - 1.89 2.07 1.99 1.95 order L∞ error 1.29E-02 4.22E-03 1.16E-03 2.87E-04 7.40E-05 1.74E-03 2.99E-04 3.58E-05 3.92E-06 4.90E-07 4.09E-06 2.60E-07 1.49E-08 9.65E-10 2.94 3.12 3.05 2.60 4.38 3.86 4.28 - order - 1.62 1.86 2.01 1.96 - 2.54 3.06 3.19 3.00 - 3.98 4.13 3.94 Example 2.4.7. A simulation for the NLS equation iut + uxx + 2|u|2u = 0 (2.38) 42 Table 2.16: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using central flux (corresponding to A2 in Table 2.11) α1 = β1 = β2 = 0, ending time Te = 1. N 40 80 160 320 640 40 80 160 320 640 40 80 160 320 640 L1 error 0.28E-02 0.71E-03 0.18E-03 0.45E-04 0.11E-04 0.13E-03 0.16E-04 0.21E-05 0.26E-06 0.32E-07 0.22E-06 0.16E-07 0.10E-08 0.62E-10 0.39E-11 P 1 P 2 P 3 - - 2.00 2.00 2.00 2.00 order L2 error 0.22E-02 0.56E-03 0.14E-03 0.35E-04 0.88E-05 0.11E-03 0.14E-04 0.18E-05 0.22E-06 0.27E-07 0.18E-06 0.13E-07 0.79E-09 0.49E-10 0.31E-11 2.99 3.00 3.00 3.00 3.76 4.00 4.00 3.99 - - - 2.00 2.00 2.00 2.00 order L∞ error 0.27E-02 0.67E-03 0.17E-03 0.41E-04 0.10E-04 0.16E-03 0.20E-04 0.25E-05 0.31E-06 0.39E-07 0.24E-06 0.13E-07 0.84E-09 0.52E-10 0.33E-11 2.99 3.00 3.00 3.00 3.80 4.00 4.00 3.99 - order - 2.02 2.01 2.00 2.00 - 3.00 3.01 3.00 3.00 - 4.16 4.00 4.00 3.96 Table 2.17: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A2 in Table 2.12): α1 = β1 = 0, β2 = 1, ending time Te = 1. N 40 80 160 320 640 40 80 160 320 640 40 80 160 320 640 L1 error 0.17E+00 0.92E-01 0.48E-01 0.24E-01 0.12E-01 0.13E-03 0.16E-04 0.21E-05 0.26E-06 0.32E-07 0.68E-06 0.42E-07 0.26E-08 0.16E-09 0.10E-10 order - 0.90 0.94 0.97 0.98 - 3.00 3.00 3.00 3.00 - 4.00 4.00 4.00 4.00 L2 error 0.13E+00 0.72E-01 0.38E-01 0.19E-01 0.97E-02 0.11E-03 0.14E-04 0.18E-05 0.22E-06 0.27E-07 0.56E-06 0.35E-07 0.22E-08 0.14E-09 0.85E-11 P 1 P 2 P 3 - - 0.89 0.94 0.97 0.98 order L∞ error 0.14E+00 0.75E-01 0.38E-01 0.19E-01 0.98E-02 0.17E-03 0.20E-04 0.25E-05 0.31E-06 0.39E-07 0.83E-06 0.51E-07 0.32E-08 0.20E-09 0.13E-10 3.00 3.00 3.00 3.00 4.01 4.00 4.00 4.00 - order - 0.87 0.97 0.98 0.99 - 3.02 3.01 3.01 3.00 - 4.01 4.00 4.00 4.00 with double-soliton collision u(x, t) = sech(x + 10− 4t) exp(i(2(x + 10)− 3t)) + sech(x− 10 + 4t) exp(i(−2(x− 10)− 3t)). 43 (2.39) Table 2.18: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A1.6.1 in Table 2.8): α1 = 0.25, ˜β1 = k(k−1) 2 + k(k+1) 8 - - 1.77 1.93 1.86 1.93 P 2 p2 = 2 ˜β1 = 7 4 P 1 p2 = 2 ˜β1 = 1 4 , ˜β2 = 1.0, p1 = −1, p2 = 2, 3, ending time Te = 1. order L2 error 0.37E-02 0.10E-02 0.25E-03 0.69E-04 0.18E-04 0.49E-04 0.73E-05 0.29E-05 0.92E-06 0.25E-06 0.34E-03 0.20E-03 0.11E-03 0.53E-04 0.27E-04 0.19E-05 0.38E-07 0.15E-08 0.90E-10 0.57E-11 L1 error 0.41E-02 0.12E-02 0.31E-03 0.87E-04 0.23E-04 0.49E-04 0.83E-05 0.31E-05 0.95E-06 0.26E-06 0.36E-03 0.21E-03 0.11E-03 0.56E-04 0.28E-04 0.19E-05 0.43E-07 0.15E-08 0.91E-10 0.58E-11 N 40 80 160 320 640 40 80 160 320 640 40 80 160 320 640 40 80 160 320 640 P 2 p2 = 3 ˜β1 = 7 4 P 3 p2 = 2 ˜β1 = 9 2 2.55 1.44 1.69 1.85 0.78 0.92 1.00 1.00 5.50 4.88 4.00 3.96 - - - - 1.82 2.05 1.87 1.94 2.74 1.32 1.69 1.86 order L∞ error 0.72E-02 0.21E-02 0.39E-03 0.10E-03 0.26E-04 0.13E-03 0.14E-04 0.65E-05 0.20E-05 0.55E-06 0.74E-03 0.43E-03 0.23E-03 0.11E-03 0.58E-04 0.43E-05 0.84E-07 0.26E-08 0.17E-09 0.11E-10 0.76 0.92 1.00 1.00 5.65 4.68 4.02 3.99 - - order - 1.80 2.39 1.94 1.97 - 3.23 1.12 1.70 1.87 - 0.77 0.92 0.99 1.00 - 5.66 5.00 3.94 3.98 Table 2.19: Example 2.4.6. Error in L1, L2 and L∞ norm for solving NLS equation (2.37) using flux parameters (corresponding to A1.7.2 in Table 2.10): α1 = 0.25, ˜β1 = −1, ˜β2 = 2k(k+1), p1 = −2,−3, p2 = 1, ending time Te = 1. 1 P 2 p1 = −2 ˜β1 = 1 12 P 2 p1 = −3 ˜β1 = 1 12 N 40 80 160 320 640 40 80 160 320 640 L1 error 0.60E-04 0.76E-05 0.96E-06 0.12E-06 0.15E-07 0.95E-04 0.21E-04 0.49E-05 0.12E-05 0.29E-06 - 2.99 3.00 3.00 3.00 order L2 error 0.54E-04 0.68E-05 0.85E-06 0.11E-06 0.13E-07 0.85E-04 0.18E-04 0.44E-05 0.11E-05 0.27E-06 2.22 2.08 2.02 2.02 - - 2.98 3.00 3.00 3.00 order L∞ error 0.95E-04 0.12E-04 0.15E-05 0.19E-06 0.24E-07 0.15E-03 0.33E-04 0.79E-05 0.20E-05 0.48E-06 2.20 2.07 2.02 2.02 - order - 2.96 2.99 2.99 3.00 - 2.18 2.06 2.02 2.02 We use periodic boundary condition and L2 projection initialization to run the simulation for double-soliton collision solution. The two waves propagate in opposite directions and 44 collide at t = 2.5, after that, the two waves separate. Such behaviors are accurately captured by our numerical simulations, see Figure 2.3 for details. Figure 2.3: Example 2.4.7. Double soliton collision graphs at t = 0, 2.5, 5 and a x− t plot of the numerical solution. N = 250, P 2 elements with periodic boundary conditions on [-25,25]. Central flux (α1 = β1 = β2 = 0) is used. 45 -20-1001020X00.20.40.60.8|uh|T = 0-20-1001020X00.511.5|uh|T = 2.5-20-1001020X00.20.40.60.8|uh|T = 5 Chapter 3 Superconvergence analysis of UWDG method on linear Schr¨odinger equation In this chapter, we study the superconvergence properties of the UWDG method on solving the following linear Schr¨odinger equation: iut + uxx = 0, (x, t) ∈ I × (0, Te], u(x, 0) = u0(x), (3.1) where I = [a, b] and periodic boundary condition. We consider solving the equation with scale invariant flux parameters. Such choice include all commonly used fluxes, e.g. alternating, central, DDG and interior penalty DG (IPDG) fluxes. We study the superconvergence property in two types. One type is the superconvergence of cell averages, numerical fluxes, solution at special points and superconvergence towards the projection P (cid:63) h in Chapter 2. Depending on the flux choices and the evenness of oddness of the polynomial degree k, we obtain 2k or (2k−1)-th order superconvergence rate for cell averages and numerical flux of the function, as well as (2k−1) or (2k−2)-th order for numerical flux of derivative. The proof relies on the correction function techniques for second order derivatives 46 applied to [10] for DDG methods for diffusion problems. We also prove the UWDG solution is superconvergent with a rate of k + 3 to the projection P (cid:63) h we introduced in Chapter 2 if k ≥ 3. At interior points whose locations are determined by roots of certain polynomials associated with the flux parameters, we show that the function values and the first and second order derivatives of the DG solution are superconvergent with order k + 2, k + 1, k, respectively. Compared with [10] for solving diffusion problems, Schr¨odinger equation poses unique challenges for superconvergence proof because of the lack of the dissipation mechanism from the equation. One major highlight of our proof is that we introduce specially chosen test functions in the error equation and show the superconvergence of the second derivative and jump across the cell interfaces of the difference between numerical solution and projected exact solution. This technique was originally proposed in [19] and is essential to elevate the convergence order for our analysis. Another type of superconvergence is by postprocessing the UWDG solution such that the postprocessed solution is convergent faster than original solution. We introduce a dual problem and prove (2k)-th order negative norm estimate. The order is one order less than that in hyperbolic equations, due to the ultra-weak formulation which has boundary term of the product of derivatives and function values. With the negative norm estimates and divided difference estimates, we prove the (2k)-th order superconvergence rate for the postprocessed solution. The rest of this chapter is organized as follows. In Section 3.1, we recall the UWDG scheme for linear Schr¨odinger equations and define some new notations. In Section 3.2, we restate the projection results in Section 2.2 under scale invariant parameters and introduce another related projection. Section 3.3 contains the superconvergence results of the UWDG solution in various quantities. In Section 3.5, we provide numerical tests verifying theoretical 47 results. Some technical proof is provided in the Appendix. The major contents of this chapter has been published in [15]. 3.1 Numerical scheme In this chapter, the semi-discrete UWDG scheme for solving linear Schr¨odinger equation is defined as follows: solve for the unique function uh = uh(t) ∈ V k h , k ≥ 1, t ∈ (0, Te], such that aj(uh, vh) = 0, ∀j ∈ ZN (3.2) holds for all vh ∈ V k h , where aj is defined in (2.6) and the numerical fluxes are defined in (2.14). Some commonly used fluxes take the following choices of parameters. • central flux, α1 = β1 = β2 = 0; • alternating flux, α1 = ± 1 2 , β1 = β2 = 0; • IPDG like flux, α1 = β2 = 0, β1 = ˜β1h−1; • DDG like flux, α1 = ˜α1, β2 = 0, β1 = ˜β1h−1; • more generally, any scale invariant flux, α1 = ˜α1, β1 = ˜β1h−1, β2 = ˜β2h; where ˜α1, ˜β1, ˜β2 are prescribed constants independent of mesh size. In this chapter, we will only consider scale invariant flux choices. Compared with discretization for diffusion equations, we don’t have any extra diffusion term in (2.10) to help with the estimates. Therefore, superconvergence error estimates are more challenging compared with [10]. 48 To facilitate the discussion, we introduce notations that will be used in this chapter. Similar to [12], we define operator D−1 for any integrable function v on Ij by (cid:90) x x j− 1 2 (cid:90) ξ −1 ˆv(ξ)dξ, x ∈ Ij. (3.3) (cid:19) (Lj,k − Lj,k−2) , (3.4a) k ≥ 2, (3.4b) D−1v(x) = 2 hj v(x)dx = Using the property of Legendre polynomials, we have (cid:0)Lj,k+1 − Lj,k−1 (cid:18) 1 (cid:1) , k ≥ 1. 2k + 1 2k + 3 (Lj,k+2 − Lj,k) − 1 2k − 1 D−1Lj,k = D−2Lj,k = 1 2k + 1 1 where D−2 = D−1 ◦ D−1. 3.2 Projections Under scale invariant flux parameter assumption, we have more concise results for P (cid:63) h . To facilitate the superconvergence proof at special points, we introduce another projection op- erator P † h in this section. To shorten the notation, from here on we use two notations Cm and Cm,n to denote W k+3+m,∞(I) mesh independent constants. Cm may depend on |u| and (cid:107)u(cid:107) depend on |u| sumption A3 and on (cid:107)u(cid:107) for assumption A3, (cid:107)u(cid:107) W k+1+m+2n,∞(I) W k+4+m+4n,∞(I) for assumption A2. W k+1+m,∞(I) for assumptions A0/A1, W k+4+m,∞(I) for assumption A2. Cm,n may for assumptions A0/A1, on (cid:107)u(cid:107) W k+3+m+3n,∞(I) for as- The definition of P (cid:63) h is given in (2.15). When scale-invariant flux parameters are used, we have the following Lemma. 49 Lemma 3.2.1 (P (cid:63) h under scale invariant flux parameters). Suppose any of the assumptions A0/A1/A2/A3 holds, u satisfies the condition in Definition 2.2.1, and scale-invariant flux parameters are used. We have the following estimates (cid:12)(cid:12)´uj,m − uj,m (cid:12)(cid:12) ≤ C0hk+1, m = k − 1, k, (cid:107)u − P (cid:63) h u(cid:107)Lν (IN ) ≤ C0hk+1, ν = 2,∞. (3.5) In addition, if hj = hj+1, (cid:12)(cid:12)´uj,m − uj,m − (´uj+1,m − uj+1,m)(cid:12)(cid:12) ≤ C1hk+2, m = k − 1, k. (3.6) Proof. When scale invariant parameters are used, when assumption A0 is satisfied, (3.5) is a direct result of (2.29) in Lemma 2.2.3. On uniform mesh, under assumption A1/A2/A3, the matrices Q1, Q2 and eigenvalues λ1, λ2 are constants independent of h. Thus, (3.5) is a direct result of (2.30) and (2.31) in Lemma 2.2.3. To prove the estimates for ´uj,m − uj,m − (´uj+1,m − uj+1,m), m = k − 1, k, we denote ´uj,k−1 − uj,k−1 ´uj,k − uj,k  − ´uj+1,k−1 − uj+1,k−1 ´uj+1,k − uj+1,k  . Uj = Plug (2.26), (2.27) in above formula. With the use of (33), Uj can be estimated in the same way as Uj in the proof of Lemma 2.2.3, and then (3.6) is obtained. We omit the proof for brevity. With the optimal estimates of P (cid:63) h u, we proved the optimal L2 error estimate of the DG scheme in Theorem 2.3.1, which is restated in a more concise version below. Theorem 3.2.2 (Theorem 2.3.1 under scale-invariant flux parameters). Suppose any of the 50 assumptions A0/A1/A2/A3 holds, let the exact solution u of (3.1) be sufficiently smooth, satisfying periodic boundary condition and uh be the UWDG solution in (3.2), then (cid:107)P (cid:63) h u − uh(cid:107) ≤ C2hk+1, (cid:107)u − uh(cid:107) ≤ C2hk+1. (3.7) Next, we introduce a local projection P † h as a variant of P (cid:63) h and study its approximation properties, especially the superconvergence property at a special set of points. Such super- convergence estimates will help us reveal the superconvergence of UWDG solution at special points. Similar ideas have been employed in [9] for proving the superconvergence at the so-called generalized Radau points when using upwind-biased flux for hyperbolic equations. Definition 3.2.1. For DG scheme with flux choice (2.14), we define a local projection op- † h for any periodic function u ∈ W 1,∞(I) to be the unique polynomial P erator P (when k ≥ 1) satisfying  P (P † hu † hu)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  − x j+ 1 2 + H G (cid:90)  P (P (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  Ij † hu † hu)x † hu vhdx = P = G + x j− 1 2 (cid:90) Ij u vhdx,  u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  ∀vh ∈ P k−2 (Ij), c  u ux (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  j− 1 2 + H j+ 1 2 † hu ∈ V k h (3.8a) (3.8b) for all j ∈ ZN . When k = 1, only condition (3.8b) is needed. Projection P † h is always a local projection. Denote P † hu|Ij =(cid:80)k m=0 `uj,mLj,m, by (3.8a), `uj,m = uj,m, m ≤ k − 2. The similarities in definition imply that P (cid:63) h and P to each other, as shown in the following lemma. 51 † h are very close Lemma 3.2.3. For periodic function u ∈ W 1,∞(I), if assumption A0 is satisfied, P (cid:63) † hu. If any of the assumptions A1/A2/A3 is satisfied, P P (−1)k+1 Γj Λj † h exists and is uniquely defined if (cid:54)= 1 for all j ∈ ZN . Then, h u = (cid:107)u − P † hu(cid:107)Lν (Ij ) ≤ Chk+1|u| W k+1,ν (Ij ) ν = 2,∞. , If any of the assumptions A1/A2/A3 is satisfied, we have (cid:107)P (cid:63) h u − P † hu(cid:107)Lν (IN ) ≤ C1hk+2, ν = 2,∞. (3.9) (3.10) Proof. When assumption A0 is satisfied, due to (2.20), P (cid:63) h = P † h. The rest of the proof is given in Appendix. To analyze the superconvergence property at special points, we need to investigate the expansion of the projection error of P † h on every cell Ij, if A0/A1/A2/A3, (u − P † hu)|Ij = [Lj,k−1, Lj,k] uj,k−1 − `uj,k−1 uj,k − `uj,k  + ∞(cid:88) ∞(cid:88) uj,mRj,m, uj,mLj,m = m=k+1 m=k+1 where uj,m is defined in (2.24) and Rj,m = Lj,m − [Lj,k−1, Lj,k]Mm. (3.11) (3.12) 52 We write out the explicit expression of the leading term in expansions Rj,k+1 = Lj,k+1 + bLj,k + cLj,k−1, (3.13) where 2k+1 hj b = − 2α1 Γj + (−1)kΛj c = − β1 − 2(k+1)2 hj , (α2 1 + β1β2 + 1 4) − (−1)k+1 2(k+1) hj Γj + (−1)kΛj 1 + β1β2 − 1 (α2 4) − β2 2 k(k + 2)(k + 1)2 hj Γj + (−1)kΛj to determine the location of superconvergent points. dxs Rj,k+1, Ds =(cid:83)N For s = 0, 1, 2, denote Ds j as the roots of ds j=1 Ds j , then it follows from (3.11) and (2.25) that, for x ∈ Ds j , ∞(cid:88) m=k+2 ∂s(u − P † hu)(x) = uj,m ds dxs Rj,m ≤ Chk+ 3 2−s|u| W k+2,s(Ij ) , (3.14) indicating superconvergence at those points. We state such superconvergence results in Theorem 3.3.6. Since the expression of b, c depends on hj, on nonuniform mesh, Ds j , s = 0, 1, 2 have nodes with the different relative locations on each cell. For simplicity, below we discuss the locations of D0 j , D1 j , D2 j for special flux choices on uniform mesh. • Alternating fluxes: b = ± 2k+1 k , c = − (k+1)2 k . • Central flux: if k is even, then b = 0, c = − (k+1)(k+2) k(k−1) ; if k is odd, then b = 0, c = −1. 53 • IPDG fluxes: if k is even, then b = 0, c = − (k+1)(k+2)−2 ˜β1 k(k−1)−2 ˜β1 −1. ; if k is odd, then b = 0, c = For central and IPDG fluxes, if k is odd, Rj,k+1 = Lj,k+1−Lj,k−1, d dx Rj,k+1 = 4k+2 hj Lj,k, d2 dx2 Rj,k+1 = 8k+4 h2 j points of order k and Lobatto points of order k + 1 excluding end points, respectively, on j are Lobatto points of order k + 1, Gauss L(cid:48) j,k, implying that D0 j , D1 j , D2 interval Ij. Therefore, card(D0 j ) = k + 1, card(D1 j ) = k, card(D2 j ) = k − 1. Cao et al. proved there exists k + 1 superconvergence points (Radau points) when using upwind flux for linear hyperbolic problem in [12], k + 1 superconvergence points (Lobatto points) using special flux parameter in DDG method in [10] and k + 1 or k superconvergence points, depending on parameters, for using upwind-biased flux for linear hyperbolic problem in [9]. Analyzing the number and location of superconvergent points for our scheme is more challenging. We shall only provide lower bound estimates for the number of superconvergence points. For general parameters choices, when k ≥ 2, Rj,k+1 ⊥ P k−2 (Ij), by Theorem 3.3 and Corollary 3.4 in [67], we can easily show Rj,k+1 has at least k − 1 simple zeros, i.e., j ) ≥ k − 2, and card(D0 when k ≥ 4, card(D2 j ) ≥ k − 1. By the same approach, we can show when k ≥ 3, card(D1 j ) ≥ k − 3. For small k values, D1 j can possibly be empty sets. j , D2 c 3.3 Superconvergence properties In this section, we study superconvergence of the numerical solution. We investigate the superconvergence of UWDG fluxes, cell averages, towards a particular projection and at some special points. This analysis is done by decomposing the error into e = u − uh = h + ζh, h = u − uI , ζh = uI − uh (3.15) 54 for some uI ∈ V k h . For error analysis of DG schemes, uI is usually taken as a projection of u. While for our purpose of superconvergence analysis, uI needs to be carefully designed as illustrated in Section 3.3.2. Before that, we prove some intermediate superconvergence results in Section 3.3.1 without specifying uI . Then, the choice of uI is made in Section 3.3.2 and the main results are obtained. (cid:80)N j=1 |[ζh]|2 3.3.1 Some intermediate superconvergence results This subsection will collect superconvergence results of (cid:107)(ζh)xx(cid:107)L2(I), ( 1 (cid:80)N j=1 |[(ζh)x]|2 ( 1 N 1 2 without specifying uI . The main idea is to choose special test func- j+ 1 2 1 2 , ) N ) j+ 1 2 tions in error equation, similar to the techniques used in [19] for hyperbolic problems. This is an essential step to elevate the superconvergence order in Theorem 3.3.4 when k is even. Lemma 3.3.1. For k ≥ 2, let u be the exact solution to (3.1) and uh be the DG solution (cid:82) h, ζh are defined in (3.15). We choose sh to be a function in V k h , such that h . Then, when any of the assumptions A0/A1 is satisfied, I shvhdx = a(h, vh), ∀vh ∈ V k in (3.2). (3.16) (3.17) (3.18) Proof. The proof is given in Appendix. (cid:107)(ζh)xx(cid:107) ≤ C(cid:107)sh + (ζh)t(cid:107), |[ζh]|2 j+ 1 2 1 2 ≤ Ch2(cid:107)sh + (ζh)t(cid:107), ) ( N(cid:88) N(cid:88) 1 N j=1 j=1 ( 1 N |[(ζh)x]|2 j+ 1 2 1 2 ≤ Ch(cid:107)sh + (ζh)t(cid:107). ) 55 3.3.2 Correction functions and the main results In this section, we shall present the main superconvergence results. The proof depends on Lemma 3.3.1 and the correction function technique introduced by Cao et al. in [12, 10], which is essential for superconvergence. We let uI = P (cid:63) when k ≥ 3, where w ∈ V k h u when k = 2, and uI = P (cid:63) h u − w, h is a specially designed correction function defined below. Similar to [10], we start the construction by defining wq, 1 ≤ q ≤ (cid:98) k−1 2 (cid:99). For k ≥ 3, we denote w0 = u − P (cid:63) h u and define a series of functions wq ∈ V k h , as follows (cid:90) Ij (cid:90) Ij wq(vh)xxdx = −i (cid:99)wq = 0, (cid:94)(wq)x = 0, (wq−1)tvhdx, ∀vh ∈ P k c (Ij) \ P 1 c (Ij), at x at x , , j+ 1 2 j+ 1 2 (3.19a) (3.19b) (3.19c) for all j ∈ ZN . (3.19b) and (3.19c) is equivalent to  wq (wq)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  G  wq (wq)x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  + x j+ 1 2 = 0. (3.20) − x j+ 1 2 + H wq exists and is unique when any of the assumptions A0/A1/A2/A3 is satisfied for the same reason as the existence and uniqueness of P (cid:63) h . With the construction of wq, we define wq(x, t), (3.21) w(x, t) = (cid:98) k−1 2 (cid:99)(cid:88) q=1 56 then aj(h, vh) = aj(u − P (cid:63) h u, vh) + aj(wq, vh) (cid:98) k−1 q=1 (cid:98) k−1 2 (cid:99)(cid:88) (cid:32)(cid:90) 2 (cid:99)(cid:88) (cid:90) 2 (cid:99)(cid:88) q=1 (cid:98) k−1 (cid:90) (cid:90) (cid:90) Ij Ij Ij = = = (w0)tvhdx + (w0)tvhdx + (wq)tvhdx − i Ij (wq − wq−1)tvhdx q=1 (w(cid:98) k−1 2 (cid:99))tvhdx, Ij ∀vh ∈ V k h (I). (cid:90) Ij (cid:33) wq(vh)xxdx (3.22) The approximation property of wq and aj(h, vh) are presented in the following Lemma. Lemma 3.3.2. For k ≥ 3, suppose u satisfies the condition in Theorem 3.2.2. For wq, 1 ≤ q ≤ (cid:98) k−1 2 (cid:99) + 1, we have 2 (cid:99), q + r ≤ (cid:98) k−1 k(cid:88) t wq|Ij ∂r = m=k−1−2q and then For any vh ∈ V k h , t cq ∂r j,mLj,m, t cq ∂r j,k−1−2q = Ch2q j ∂q+r t (cid:12)(cid:12)(cid:12)∂r t cq j,m (uj,k−1 − ´uj,k−1), (cid:12)(cid:12)(cid:12) ≤ C2r,qhk+1+2q, (cid:107)∂r t wq(cid:107) ≤ C2r,qhk+1+2q. (3.23) (3.24) (3.25) |a(h, vh)| ≤ C Proof. The proof is given in Appendix. 2 (cid:99)hk+1+2(cid:98) k−1 2,(cid:98) k−1 2 (cid:99)(cid:107)vh(cid:107). Lemma 3.3.3. For k ≥ 2, suppose u satisfies the condition in Theorem 3.2.2. If the 57 parameters satisfy any of the assumptions A0/A1 and uh|t=0 = uI|t=0, we have N(cid:88) N(cid:88) j=1 j=1 ( 1 N ( 1 N (cid:107)(ζh)xx(cid:107) ≤ C |[(ζh)]|2 j+ 1 2 1 2 ≤ C ) 4+2(cid:98) k−1 4+2(cid:98) k−1 |[(ζh)x]|2 1 2 ≤ C ) j+ 1 2 4+2(cid:98) k−1 2 (cid:99)hk+1+2(cid:98) k−1 2 (cid:99) 2 (cid:99)hk+3+2(cid:98) k−1 2 (cid:99) 2 (cid:99)hk+2+2(cid:98) k−1 2 (cid:99) . . . (3.26) (3.27) (3.28) Proof. This Lemma is a direct result of Lemma 3.3.1, once the estiamtes of (cid:107)sh + (ζh)t(cid:107) is acquired. When k = 2, w = 0. a(h, vh) =(cid:82) h . That is, sh = (h)t in the condition of Lemma 3.3.1. To bound (cid:107)(ζh)t(cid:107), we take the time derivative I (h)tvhdx from the definition of P (cid:63) of the error equation and obtain a(et, vh) = a((h)t, vh) + a((ζh)t, vh) = 0. Let vh = (ζh)t, since a((ζh)t, (ζh)t) + a((ζh)t, (ζh)t) = d dt(cid:107)(ζh)t(cid:107)2, by the property of P (cid:63) h u, we have (cid:107)(ζh)t(cid:107)2 = −a((h)t, (ζh)t) − a((h)t, (ζh)t) ≤ 2(cid:107)(h)tt(cid:107)(cid:107)(ζh)t(cid:107), d dt dt(cid:107)(ζh)t(cid:107) ≤ (cid:107)(h)tt(cid:107). To estimate (cid:107)(ζh)t|t=0(cid:107), we let t = 0 in the error which implies d equation. Since ζh|t=0 = (uh − uI )|t=0 = 0, we have (cid:90) I a(h, vh) + (ζh)t|t=0vhdx = 0. 58 Let vh = (ζh)t|t=0, then (cid:107)(ζh)t|t=0(cid:107)2 ≤ (cid:107)(h)t(cid:107)(cid:107)(ζh)t|t=0(cid:107). Therefore, (cid:107)(ζh)t(cid:107) ≤ (cid:107)(h)t(cid:107) + t(cid:107)(h)tt(cid:107). By Lemma 3.3.1, estimates in (3.5) and the inequality above, we can get (3.26)-(3.28). For k ≥ 3, by (3.22), we have a(h, vh) =(cid:82) 2 (cid:99))t in the condition of Lemma 3.3.1. Then, following the same lines of proof as above, by replacing 2 (cid:99))tvhdx, that is, sh = (w(cid:98) k−1 I (w(cid:98) k−1 h with w(cid:98) k−1 2 (cid:99) and using Lemma 3.3.2, we are done. Now we are ready to state the following estimates of (cid:107)ζh(cid:107). Theorem 3.3.4. For k ≥ 2, suppose u satisfies the condition in Theorem 3.2.2. Assume uh|t=0 = uI|t=0, then ∀t ∈ (0, Te], C 2 (cid:107)ζh(cid:107) ≤  h2k 2, k−1 (Ck+2h4k +(cid:80) h2k−1, C 2, k−2 2 Ij⊂IN U Ckh4k−1) if k is odd and A0/A1/A2/A3, 1 2 if k is even and A0/A1, (3.29) if k is even and A2/A3, where IN U is the collection of cells in which the length of Ij is different with at least one of its neighbors. 59 Proof. From error equation, a(e, ζh) = a(h, ζh) + a(ζh, ζh) = 0, which gives us (cid:107)ζh(cid:107)2 = −a(h, ζh) − a(h, ζh) ≤ d dt 2(cid:107)(h)t(cid:107)(cid:107)ζh(cid:107), 2(cid:107)(w(cid:98) k−1 k = 2, (3.30) 2 (cid:99))t(cid:107)(cid:107)ζh(cid:107), k ≥ 3. By (3.5), (3.24) and Gronwall’s inequality, we have (cid:107)ζh(cid:107) ≤ C 2 (cid:99)thk+1+2(cid:98) k−1 2 (cid:99) 2,(cid:98) k−1 , ∀t ∈ (0, Te]. Therefore, when k is odd, or k is even and parameters satisfy any of the assumptions A2/A3, the proof is complete. When k is even and parameters satisfy any of the assumptions A0/A1, we make use of Lemma 3.3.3 to show the improved estimates. We let l = (cid:98) k−1 2 (cid:99) = k−2 2 , then (wl)tζhdx = Lj,mζhdx (cid:90) N(cid:88) I j=1 (cid:90) Ij ∂tcl j,1 N(cid:88) k(cid:88) j=1 m=1 Lj,1ζhdx + (cid:90) k(cid:88) Ij ∂tcl j,m N(cid:88) j=1 m=2 (cid:90) Ij a(h, ζh) = = ∂tcl j,m Lj,mζhdx = A1 + A2, . where we denote the first term in the summation by A1, and the other term in summation as A2. Note that D−1Lj,m ⊥ P 0, m ≥ 1 in the inner product sense, thus D−2Lj,m(±1) = 60 0, m ≥ 2. By integration by parts, we get (cid:90) Ij A2 = ∂tcl j,m D−2Lj,m(ζh)xxdx (cid:90) N(cid:88) k(cid:88) ( m=2 Ij j=1 |h2 j 4 ∂tcl j,m|2 + h N(cid:88) j=1 k(cid:88) N(cid:88) m=2 h2 j 4 ≤ Ch−1 j=1 ≤ Ck+2h4k D−2Lj,m(ζh)xxdx)2 ≤ Ckh4k + Ch2(cid:107)(ζh)xx(cid:107)2 where we have used (3.23) in the first inequality, and (3.26) in the third inequality. 2 ( ξ2 hj To estimate A1, we take the first and second antiderivative of Lj,1 = ξ as 2 )2 ξ3−ξ and apply integration by parts twice, hj ( 6 2 − 1 6), (cid:33) j=1 N(cid:88) N(cid:88) N(cid:88) N(cid:88) j=1 j=1 A1 = = = + hj 2 hj 2 ∂tcl j,1 ∂tcl j,1 j,1 ∂tcl hj 2 (cid:18)(cid:16)hj (cid:17) ∂tcl 2 j=1 (cid:90) (cid:12)(cid:12)x x ( )¯ζh ξ2 2 − 1 6 (cid:32) (cid:32) (cid:32) j,1 −(cid:16)hj+1 (¯ζh|− j+ 1 2 −1 3 j+ 1 2 [¯ζh] 1 3 2 − j+ 1 2 j− 1 2 ξ2 2 ( Ij − ¯ζh|+ j− 1 2 ) + hj 2 (cid:90) Ij hj 2 + (cid:17) ξ3 − ξ 6 (cid:19) 1 ∂tcl j+1,1 ¯ζh|+ j+ 1 2 , 3 (cid:33) )(¯ζh)xdx − 1 6 ξ3 − ξ (cid:90) (¯ζh)xxdx (cid:33) 6 Ij (¯ζh)xxdx where we have used the periodicity in the last equality. Therefore, |A1| ≤ 1 2 h N(cid:88) j=1 2 1 9 + h−1 (cid:18)(cid:16)hj (cid:19)2 (cid:17)|∂tcl N(cid:88) j,1| (cid:12)(cid:12)(cid:12)(cid:12)(cid:18)(cid:16) hj (cid:17) j,1 −(cid:16) hj+1 (cid:17) N(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:18)(cid:16)hj N(cid:88) ∂tcl j=1 j=1 2 2 2 ∂tcl j=1 ≤ Ck+2h4k + C(cid:107)ζh(cid:107)2 + Ch−1 + h−1 1 18 j=1 + Ch2 N(cid:88) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2 N(cid:88) j,1 −(cid:16)hj+1 + h j=1 2 |[¯ζh]|2 j+ 1 2 j+1,1 (cid:17) ∂tcl (cid:107)(ζh)xx(cid:107)2 L2(Ij ) (cid:107)ζh(cid:107)2 (cid:17) ∂tcl j+1,1 L2(∂Ij ) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2 61 where we used (3.23), inverse inequality, (3.26) and (3.27) in the last inequality. We estimate the last term in A1 by the estimation of the difference of uj,m in neighboring cells, similar to that in Proposition 3.1 of [6]. If hj (cid:54)= hj+1, then hj 2 ∂tcl j,1 − hj+1 2 ∂tcl j+1,1 ≤ Ckh2k. If hj = hj+1, by (3.23) and (3.6), (cid:16)hj (cid:17)2l+1 ≤ Ck+1h2k+1. 2 hj 2 ∂tcl j,1 − hj+1 2 ∂tcl j+1,1 = C ∂l+1 t (uj,k−1 − ´uj,k−1 − (uj+1,k−1 − ´uj+1,k−1)) Therefore, we have |A1| ≤ Ck+2h4k + C(cid:107)ζh(cid:107)2 + (cid:88) Ij⊂IN U Ckh4k−1. Combine with the estimates for A2, we have (cid:107)ζh(cid:107)2 ≤ Ck+2h4k + C(cid:107)ζh(cid:107)2 + d dt (cid:88) Ij⊂IN U Ckh4k−1. By Gronwall’s inequality and the numerical initial condition, we obtain (cid:107)ζh(cid:107) ≤ (Ck+2h4k + Ckh4k−1) 1 2 . (cid:88) Ij⊂IN U The proof is now complete. Above theorem states that for k ≥ 2, when k is odd or k is even and any of the as- 62 sumptions A0/A1 is satisfied, (cid:107)ζh(cid:107) has the desired 2k-th order convergence rate. When k is even and any of the assumptions A2/A3 is satisfied, we are unable to improve the order because that a middle step, Lemma 3.3.3, is proved only under assumption A0/A1. However, numerical result shows the superconvergence results of (cid:107)ζh(cid:107) and the quantities in following two theorems still hold when k is even and any of the assumptions A2/A3 is satisfied. There is room for improving the proof under such assumption. Our main superconvergence results are listed in the following two theorems. Theorem 3.3.5 (Superconvergence of numerical fluxes and cell averages). Let (cid:17) 1 (cid:16) 1 (cid:16) 1 N N(cid:88) N(cid:88) j=1 N j=1 (u −(cid:99)uh)|2 (cid:90) (cid:12)(cid:12)(cid:12) 1 hj Ij Ef = Ec = 2 , Efx = j+ 1 2 (cid:12)(cid:12)(cid:12)2(cid:17) 1 2 u − uhdx (cid:16) 1 N(cid:88) N j=1 (ux − (cid:94)(uh)x)|2 j+ 1 2 (cid:17) 1 2 , (3.31) be the errors in the two numerical fluxes and the cell averages, respectively. For k ≥ 2, suppose u satisfies the condition in Theorem 3.2.2. Assume uh|t=0 = uI|t=0, then ∀t ∈ (0, Te] • if k is odd, parameters satisfy any of the assumptions A0/A1/A2/A3, we have Ef ≤ C 2, k−1 2 h2k, Efx ≤ C 2, k−1 2 h2k−1, Ec ≤ C h2k, 2, k−1 2 (3.32) 63 • if k is even, parameters satisfy any of the assumptions A0/A1, we have Ef ≤ (Ck+2h4k + Ec ≤ (Ck+2h4k + (cid:88) Ij⊂IN U (cid:88) Ij⊂IN U (cid:88) Ij⊂IN U Ckh4k−1) 1 2 , Efx ≤ (Ck+2h4k + Ckh4k−1) 1 2 , Ckh4k−1) 1 2 h−1, (3.33) (3.34) where IN U is the collection of cells in which the length of Ij is different with at least one of its neighbors. • if k is even and parameters satisfy assumption A2/A3, we have Ef ≤ C 2, k−2 2 h2k−1, Efx ≤ C 2, k−2 2 h2k−2, Ec ≤ C h2k−1. 2, k−2 2 Proof. We first prove the estimates for Ef . By (3.19b) and the definition of P (cid:63) (cid:92)u − uI (x ) = 0, then j+ 1 2 (u −(cid:99)uh)| j+ 1 2 = ((cid:98)ζh)| j+ 1 2 =(cid:0){ζh} − α1[ζh] + β2[(ζh)x](cid:1)| . j+ 1 2 (3.35) ) = j+ 1 2 h ,(cid:98)h(x Therefore, by inverse inequality and the fact β2 = ˜β2h, (cid:16) 1 N Ef ≤ C (cid:107)ζh(cid:107)2 L2(∂IN ) (cid:17) 1 2 ≤ C(cid:107)ζh(cid:107), and the desired estimates for Ef is obtained by (3.29). The estimates for Efx can be obtained following same lines. 64 Next, we prove the estimates for Ec. If k is odd, then (cid:82) Ij (3.23) and orthogonality of Legendre polynomials. Thus, wqdx = 0, 1 ≤ q ≤ k−3 2 , by (cid:90) Ij Thus, u − uhdx = (cid:90) (cid:12)(cid:12)(cid:12) 1 If k is even, then (cid:82) hj (cid:90) Ij (cid:98) k−1 2 (cid:99)(cid:88) q=1 u − P (cid:63) h u + wq + ζhdx = (cid:90) Ij (cid:90) Ij ζhdx. w(cid:98) k−1 2 (cid:99)dx + (cid:12)(cid:12)(cid:12)2 ≤ 2 hj u − uhdx ((cid:107)ζh(cid:107)2 L2(Ij ) + (cid:107)w k−1 2 (cid:107)2 L2(Ij ) ). polynomials. Thus, by similar step, we have Ij Ij wqdx = 0, 1 ≤ q ≤ k−2 (cid:90) (cid:90) ζhdx, Ij Ij u − uhdx = (cid:90) Ij Therefore, 2 , by (3.23) and orthogonality of Legendre (cid:12)(cid:12)(cid:12) 1 hj (cid:12)(cid:12)(cid:12)2 ≤ 2 hj u − uhdx (cid:107)ζh(cid:107)2 L2(Ij ) . Ec ≤ C((cid:107)ζh(cid:107)2 + (cid:107)w k−1 2 (cid:107)2)1/2 if k is odd, Ec ≤ C(cid:107)ζh(cid:107) if k is even, and the desired estimate for Ec is obtained by (3.29) and (3.24). Theorem 3.3.6 (Superconvergence towards projections and at special points). Suppose u satisfies the condition in Theorem 3.2.2. Assume uh|t=0 = P (cid:63) h u0, then ∀t ∈ (0, Te], (C4h4k +(cid:80) C2,1(1 + t)hk+3 (cid:107)uh − P (cid:63) h u(cid:107) ≤ Ij⊂IN U C2h4k−1) 1 2 k = 2, if A0 or A1 k ≥ 3, (3.36) where IN U is the collection of cells in which the length of Ij is different with at least one of 65 its neighbors. Assume Ds, s = 0, 1, 2 defined in (3.14) are not empty sets. Let Eu = (cid:16) 1 (cid:16) 1 |D0| |(u − uh)(x)|2(cid:17) 1 (cid:88) |(u − uh)xx(x)|2(cid:17) 1 (cid:88) 2 x∈D0 2 , Eux = Euxx = |D2| x∈D2 (cid:16) 1 |D1| (cid:88) |(u − uh)x(x)|2(cid:17) 1 2 , x∈D1 (3.37) be the average point value error for the numerical solution, the derivative of solution and the second order derivative of solution at corresponding sets of points. Then • if k = 2 and any of the assumptions A0/A1 is satisfied, we have (cid:88) Eu ≤ (C4h4k + Ij⊂IN U Euxx ≤ h−2(C4h4k + C2h4k−1) (cid:88) C2h4k−1) 1 2 . 1 2 , Eux ≤ h−1(C4h4k + Ij⊂IN U (cid:88) Ij⊂IN U C2h4k−1) 1 2 , (3.38) • if k ≥ 3 and any of the assumptions A0/A1/A2/A3 is satisfied, we have Eu ≤ C2,1hk+2, Eux ≤ C2,1hk+1, Euxx ≤ C2,1hk. (3.39) Proof. When k = 2, we have uh − P (cid:63) h u = −ζh. If any of the assumptions A0/A1 is satisfied, by (3.29), we have (cid:107)uh − P (cid:63) h u(cid:107) ≤ (C4h4k + C2h4k−1) 1 2 . (cid:88) Ij⊂IN U When k ≥ 3, to relax the regularity requirement, we follow the same steps in Lemma 66 3.3.2, and change the definition of uI to uI = P (cid:63) h u − w1. Then h = u − uI , ζh = uI − uh and we obtain |a(h, vh)| ≤ C2,1hk+3(cid:107)vh(cid:107), ∀vh ∈ V k h . By the estimates above, (3.24) and the error equation, we obtain (cid:107)ζh(cid:107)2 ≤ 2(cid:107)(w1)t(cid:107)(cid:107)ζh(cid:107) ≤ C2,1hk+3. d dt By Gronwall’s inequality, (cid:107)ζh(cid:107) ≤ C2,1thk+3 + (cid:107)(ζh)|t=0(cid:107) = C2,1thk+3 + (cid:107)w1|t=0(cid:107) ≤ C2,1(1 + t)hk+3, ∀t ∈ (0, Te], where uh|t=0 = P (cid:63) that ∀t ∈ (0, Te], h u0 is used in the first equality. Since uh − P (cid:63) h u = −ζh − w1, it follows (cid:107)uh − P (cid:63) h u(cid:107) ≤ 2((cid:107)ζh(cid:107) + (cid:107)w1(cid:107)) ≤ C2,1(1 + t)hk+3. Then the proof for (3.36) is complete. If any of the assumptions A0/A1/A2/A3 is satisfied, then |(u − P † hu)(x)|2 + |(P (cid:63) h u − uh)(x)|2 + |(P (cid:63) h u − P hu)(x)|2(cid:17) 1 † 2 Eu ≤(cid:16) 1 |D0| (cid:88) x∈D0 ≤ Chk+2|u| W k+2,2(I) + C(cid:107)P (cid:63) h u − uh(cid:107) + C(cid:107)P (cid:63) h u − P † hu(cid:107), where (3.14), inverse inequality, and (3.10) are used in the last inequality. Then the estimates for Eu is proven by Lemma 3.2.3 and (3.36). The estimates for Eux and Euxx can be proven following the same lines. 67 Remark 3.3.1. If the initial discretization is taken as uh|t=0 = uI|t=0, the theorem above still holds. However, the regularity requirement will be higher. 3.4 Superconvergence of postprocessed solution In this section, we analyze the superconvergence property of the postprocessed DG so- lutions for linear Schr¨odinger equation (3.1) on uniform mesh by using negative Sobolev norm estimates. The postprocessor was originally introduced in [7, 55] for finite difference and finite element methods, and later applied to DG methods in [23]. The postprocessed solution is computed by the convolution of numerical solution uh and a kernel function h), where d is the number of spatial dimensions, and l is the index of H−l h (x) = 1 Kν,l hd Kν,l( x norm we’re trying to estimate later. The convolution kernel has three main properties. First, it has compact support, making post processing computationally advantageous. Second, it preserves polynomials of degree up to ν − 1 by convolution, thus the convergence rate is not deteriorated. Third, the kernel Kν,l is a linear combination of B-splines, which allows us to express the derivatives of kernel by difference quotients (see section 4.1 in [23]). We give the formula for the convolution kernel when the DG scheme uses approximation space V k h : K2(k+1),k+1(x) = k(cid:88) γ=−k 2(k+1),k+1 k γ ψ(k+1)(x − γ), 2(k+1),k+1 where ψ(k+1) are the B-spline bases and the computation of coefficients k γ can be 68 found in [65]. Then we can define the postprocessed DG solution as (cid:90) ∞ −∞ K u∗ = 2(k+1),k+1 h (y − x)uh(y)dy. (3.40) u∗ is an “averaged” version of uh such that it is closer as an approximation to the exact solution u. Lastly, we define divided difference as dhv(x) = 1 h (v(x + 1 2 h) − v(x − 1 2 h)). Now we are ready to state an approximation result showing the smoothness of u and negative Sobolev norm of divided difference lead to a bound on u − u∗. Theorem 3.4.1 (Bramble and Schatz [7]). Suppose u∗ is defined in (3.40) and K 2(k+1),k+1 h = 1 hK2(k+1),k+1( x h), where K2(k+1),k+1 is a kernel function as defined above. Let u be the exact solution of linear Schr¨odinger equation (3.1) satisfying periodic boundary condition, u ∈ H2k+2(I). Then for arbitrary time t ∈ (0, Te], h sufficiently small, we have (cid:88) α≤k+1 (cid:107)dα h(u − uh)(cid:107) H−(k+1)(IN ) , (3.41) (cid:107)u − u∗(cid:107) ≤ Ch2k+2|u| H2k+2(I) + where C is independent of u and h. The right hand side of (3.41) indicates that if (cid:107)dα h(u − uh)(cid:107) H−(k+1)(IN ) converges at a rate higher than k + 1, then we have superconvergence property for the postprocessed solution. In what follows, we estimate the negative-norm term following the steps in [23]. First, we introduce a dual problem: find a function v such that v(·, t) is periodic function 69 with period equal to the length of I, i.e., b − a for all t ∈ (0, Te] and ivt − vxx = 0, in R × (0, Te), where Φ is an arbitrary function in C∞ v(x, Te) = Φ(x), x ∈ R, 0 (I). We use the notation (φ, ψ) :=(cid:82) (3.42) I φ ψdx in this section. At final time Te, (u(Te) − uh(Te), Φ) = (u, v)(Te) − (uh, v)(Te) = (u, v)(0) + {(u, vt) + (ut, v)}dt − (uh, v)(Te) (cid:90) Te = (u, v)(0) − (uh, v)(0) − = (u − uh, v)(0) − {((uh)t, v) + (uh, vt)}dt, {((uh)t, v) + (uh, vt)}dt (cid:90) Te 0 (cid:90) Te 0 0 where the property uvt + utv = 0 is used to obtain the third equality. The DG solution uh satisfies (3.2). Therefore, we have ∀vh ∈ V k h ((uh)t, v) = ((uh)t, v − vh) + ((uh)t, vh) = ((uh)t, v − vh) + iA(uh, vh) = ((uh)t, v − vh) − iA(uh, v − vh) + iA(uh, v). Then we obtain (u(Te) − uh(Te), Φ) = ΘM + ΘN + ΘC , 70 where (cid:90) Te (cid:90) Te ΘM = (u − uh, v)(0), ΘN = − ΘC = − 0 0 {((uh)t, v − vh) − iA(uh, v − vh)}dt, {(uh, vt) + iA(uh, v)}dt. ∀vh ∈ V k h , By choosing the initial numerical discretization uh(0) = P 0 h u0 and vh = P 0 h v, we have ΘM = (u − uh, v)(0) = (u − uh, v − vh)(0) and |ΘM| ≤ (cid:107)(u − uh)(0)(cid:107) · (cid:107)(v − vh)(0)(cid:107) ≤ Ch2k+2(cid:107)u(cid:107) Hk+1(I) (cid:107)v(cid:107) Hk+1(I) . Since v is a smooth function, we have (cid:90) Te 0 ΘC = − {(uh, vt) + i(uh, vxx)}dt = 0. Choose vh = P 0 |ΘN| = 0 0 j=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:90) Te (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = A(v − vh, uh)dt h v and from the symmetry of the operator A(·,·), we get (cid:90) Te (cid:90) Te (cid:90) Te (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:0)(cid:92)v − vh[(uh)x] − (cid:94)(v − vh)x[uh](cid:1)(cid:12)(cid:12)j+ 1 (cid:0)(cid:92)v − vh[ux − (uh)x] − (cid:94)(v − vh)x[u − uh](cid:1)(cid:12)(cid:12)j+ 1 (cid:0)(cid:107)u − uh(cid:107)L2(∂IN )(cid:107) (cid:94)(v − vh)x(cid:107)L2(∂IN ) A(uh, v − vh)dt N(cid:88) N(cid:88) ≤ CTe max t∈(0,Te] + (cid:107)(u − uh)x(cid:107)L2(∂IN )(cid:107)(cid:92)v − vh(cid:107)L2(∂IN ) = = 0 j=1 (cid:1). dt 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt 2 71 By (3.7), (cid:107)u − uh(cid:107)L2(∂IN ) = (cid:107)u − P (cid:63) ≤ C0hk+ 1 2 + Ch − 1 2(cid:107)P (cid:63) h u(cid:107)L2(∂IN ) + (cid:107)P (cid:63) h u − uh(cid:107)L2(∂IN ) h u − uh(cid:107) ≤ C2hk+ 1 2 , where we have used Lemma 2.2.3 and Theorem 3.2.2. Similarly, we have (cid:107)ux−(uh)x(cid:107)L2(∂IN ) ≤ C2hk− 1 2(cid:107)vx − (vh)x(cid:107)L2(∂IN ) + h(cid:107)vx − (vh)x(cid:107) + 2(cid:107)v − vh(cid:107)L2(∂IN ) + (cid:107)v − vh(cid:107) ≤ Chk+1(cid:107)v(cid:107) 2 . By the property of L2 projection h . Then it is straightforward that for Hk+1(I) h 1 3 scale invariant fluxes, (cid:107)(cid:92)v − vh(cid:107)L2(∂IN ) ≤ Chk+ 1 2(cid:107)v(cid:107) Hk+1(I) , (cid:107) (cid:94)(v − vh)x(cid:107)L2(∂IN ) ≤ Chk− 1 2(cid:107)v(cid:107) Hk+1(I) . Therefore, we have |ΘN| ≤ C2h2k(cid:107)v(cid:107) Hk+1(I) . (3.43) Combine the above three estimate and the fact (cid:107)v(cid:107) = (cid:107)Φ(cid:107) Hk+1(I) Hk+1(I) , we have (cid:107)u(Te) − uh(Te)(cid:107) H−(k+1)(I) ≤ C2h2k. the divided difference dα Since we consider uh with optimal error estimates on uniform mesh with mesh size h, then hu0, α ≤ k+1 huh also satisfies the DG scheme (3.2) but with shifted mesh hu0. Then by the same proof for u − uh hu satisfies the linear Schr¨odinger but with initial data dα and initial numerical discretization dα on shifted mesh. Similarly, dα huh = P 0 h dα above, (cid:107)dα h(u − uh)(Te)(cid:107) H−(k+1)(I) ≤ C2+αh2k, (3.44) 72 where we used Taylor expansion to estimate dα hu to obtain the last inequality. The following theorem is a result of (3.44) and Theorem 3.4.1. Theorem 3.4.2. Let uh be the UWDG solution of (3.2), suppose the conditions in Theorem 3.4.1 and any of the assumptions A0/A1/A2/A3 is satisfied, then on a uniform mesh (cid:107)u(Te) − u∗(Te)(cid:107) ≤ Ck+3h2k. (3.45) 3.5 Numerical results In this section, we provide numerical tests demonstrating superconvergence properties. In the proof, we see that the initial value of uh matters in estimating (cid:107)uh − uI(cid:107), thus will impact the superconvergence estimation for Ef and Efx. Therefore, in our numerical tests, we apply two types of initial discretization for uh. For computing the postprocessed solution u∗, we use the standard L2 projection P 0 h u as numerical initialization to demonstrate the convergence enhancement ability of postprocessor. For verifying other superconvergence quantities, we apply the initial condition uh|t=0 = uI|t=0. In order not to deteriorate the high order convergence rates, for temporal discretization, we use explicit Runge-Kutta fourth order method with dt = c · h2.5, c = 0.05 when k = 2 and c = 0.01 when k = 3, 4. Example 3.5.1. We compute (3.1) on [0, 2π] with exact solution u(x, t) = exp(i3(x − 3t)) using UWDG scheme (3.2). We verify the results with several flux parameters. In the following tables, we show the convergence rate for quantities Ef , Efx, Ec, Eu, Eux, Euxx as defined in (3.31) and (3.37) as well as E∗ = (cid:107)u − u∗(cid:107), EP = (cid:107)uh − P (cid:63) h u(cid:107), (3.46) 73 which represent the error after postprocessing, and the error between numerical solution and the projected exact solution P (cid:63) h u. In addition, we test the superconvergence of the intermediate quantities ζh = P (cid:63) h u − w − uh as in Lemma 3.3.3, and introduce the following notations: N(cid:88) j=1 E[ζh] = ( 1 N |[ζh]|2 j+ 1 2 ) 1 2 , E[(ζh)x] = ( 1 N N(cid:88) j=1 |[(ζh)x]|2 j+ 1 2 1 2 . ) The numerical fluxes we tested include 1. Tables 3.1, 3.5: A0 parameters, alternating flux, α1 = 0.5, β1 = β2 = 0, with nonuni- form mesh; 2. Tables 3.2, 3.6: A0 parameters, a scale invariant flux, α1 = 0.3, β1 = 0.4 h , β2 = 0.4h, with nonuniform mesh; 3. Tables 3.3, 3.7: A1 parameters, central flux, α1 = β1 = β2 = 0, with uniform mesh; 4. Tables 3.4: A3 parameters, α1 = 0.25, β2 = 0, β1 = 2 h for k = 2, 3, 4, respectively, h , 5 h, 9 with uniform mesh; 5. Table 3.8: all parameters mentioned above, with uniform mesh, where the nonuniform mesh is generated by perturbing the location of the nodes of a uniform mesh by 10% of mesh size. We first verify the results in Theorems 3.3.5, 3.3.6 by examining Tables 3.1, 3.2, 3.3, 3.4, where the parameters satisfy assumption A0, A0, A1, A3, respectively. We observe that the scheme can achieve at least the theoretical order of convergence for the quantities in these two theorems. To be more specific, EP shows (k + min(3, k))-th order of convergence. 74 Table 3.1: Example 3.5.1. Error table when using alternating flux on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. N 10 20 40 80 160 10 20 40 80 160 10 20 40 80 160 N 10 20 40 80 160 10 20 40 80 160 10 20 40 80 160 L2 error 2.68E-01 2.68E-02 2.00E-03 1.91E-04 2.19E-05 1.02E-02 5.65E-04 2.94E-05 1.87E-06 1.18E-07 6.45E-04 2.06E-05 6.83E-07 2.04E-08 6.10E-10 Eu 2.83E-01 2.30E-02 1.47E-03 9.12E-05 5.78E-06 6.81E-03 1.40E-04 2.16E-06 3.73E-08 8.04E-10 1.09E-04 1.04E-06 1.28E-08 1.94E-10 2.85E-12 P 2 P 3 P 4 P 2 P 3 P 4 order EP order Euxx order Eux order - 3.32 3.75 3.39 3.12 - 4.18 4.26 3.98 3.98 - 4.97 4.91 5.07 5.06 order - 3.62 3.97 4.01 3.98 - 5.61 6.01 5.86 5.53 - 6.72 6.34 6.04 6.09 2.53E-01 2.47E-02 1.42E-03 9.22E-05 5.83E-06 7.51E-03 1.24E-04 2.13E-06 3.02E-08 4.57E-10 8.76E-05 6.12E-07 2.52E-09 1.55E-11 1.02E-13 Ef 2.79E-01 2.31E-02 1.46E-03 9.11E-05 5.77E-06 6.78E-03 1.36E-04 2.02E-06 3.05E-08 4.59E-10 9.69E-05 5.02E-07 1.76E-09 6.99E-12 2.79E-14 - 3.36 4.12 3.95 3.98 - 5.93 5.86 6.14 6.05 - 7.16 7.92 7.35 7.24 order - 3.60 3.98 4.01 3.98 - 5.64 6.07 6.05 6.05 - 7.59 8.16 7.97 7.97 2.36E+00 2.78E-01 6.02E-02 1.50E-02 3.78E-03 1.60E-01 2.00E-02 2.43E-03 2.99E-04 3.76E-05 1.79E-02 1.25E-03 7.45E-05 4.70E-06 2.85E-07 Efx 9.10E-01 6.94E-02 4.45E-03 2.75E-04 1.74E-05 1.98E-02 4.08E-04 6.06E-06 9.13E-08 1.38E-09 2.83E-04 1.57E-06 5.26E-09 2.09E-11 5.55E-13 - 3.09 2.21 2.00 1.99 - 3.00 3.04 3.02 2.99 - 3.84 4.06 3.99 4.04 order - 3.71 3.96 4.02 3.98 - 5.60 6.07 6.05 6.05 - 7.49 8.22 7.97 5.23 8.92E-01 7.43E-02 5.29E-03 3.83E-04 3.24E-05 2.50E-02 8.23E-04 3.83E-05 2.29E-06 1.45E-07 8.67E-04 3.06E-05 9.00E-07 2.78E-08 8.34E-10 Ec 5.00E-01 4.48E-02 2.91E-03 1.82E-04 1.15E-05 1.16E-02 2.64E-04 4.03E-06 6.11E-08 9.25E-10 1.62E-04 9.69E-07 3.50E-09 1.40E-11 5.07E-14 - 3.59 3.81 3.79 3.56 - 4.93 4.42 4.06 3.98 - 4.82 5.09 5.02 5.06 order - 3.48 3.94 4.00 3.98 - 5.45 6.03 6.04 6.05 - 7.39 8.11 7.97 8.10 Eu, Eux, Euxx are shown to have (k + 2)-th, (k + 1)-th and k-th order of convergence , respectively. Note that when k = 2, in Tables 3.2 and 3.4, there are situations when no superconvergence point exists. This finding shows an evidence to the assertion that Ds defined in (3.14) could be empty sets. The order of convergence for Ef , (Ef )x, Ec in all tables are 2k. In addition, Table 3.4 shows that when k is even and assumption A3 is satisfied, the convergence order for all quantities are the same as when any of assumption 75 A0/A1 is satisfied, which is one order higher than the estimates in Theorems 3.3.5, 3.3.6. Table 3.2: Example 3.5.1. Error table when using flux parameters: α1 = 0.3, β1 = 0.4 0.4h on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. h , β2 = N 40 80 160 320 640 10 20 40 80 160 10 20 40 80 160 N 40 80 160 320 640 10 20 40 80 160 10 20 40 80 160 L2 error 1.66E-02 1.50E-03 1.70E-04 2.16E-05 2.64E-06 2.08E-02 1.14E-03 5.91E-05 3.75E-06 2.39E-07 9.72E-04 3.17E-05 1.05E-06 3.14E-08 9.39E-10 Eu 1.26E-02 7.43E-04 4.82E-05 3.09E-06 1.94E-07 1.44E-02 3.08E-04 4.70E-06 7.56E-08 1.03E-09 1.77E-04 1.56E-06 1.83E-08 2.82E-10 4.12E-12 P 2 P 3 P 4 P 2 P 3 P 4 order EP order - 3.47 3.14 2.98 3.03 - 4.19 4.27 3.98 3.97 - 4.94 4.91 5.07 5.06 order - 4.08 3.95 3.96 3.99 - 5.54 6.04 5.96 6.20 - 6.83 6.41 6.02 6.10 1.28E-02 7.46E-04 4.91E-05 3.12E-06 1.96E-07 1.57E-02 2.75E-04 4.83E-06 6.81E-08 1.09E-09 1.48E-04 1.04E-06 4.15E-09 2.51E-11 1.63E-13 Ef 1.26E-02 7.42E-04 4.82E-05 3.09E-06 1.94E-07 1.41E-02 3.05E-04 4.56E-06 6.88E-08 1.10E-09 1.65E-04 8.66E-07 3.04E-09 1.21E-11 4.41E-14 - 4.10 3.92 3.98 3.99 - 5.84 5.83 6.15 5.96 - 7.16 7.97 7.37 7.27 order - 4.08 3.95 3.96 3.99 - 5.53 6.06 6.05 5.97 - 7.57 8.15 7.98 8.10 Euxx 2.38E-01 2.52E-02 4.51E-03 8.89E-04 2.00E-04 2.00E-01 2.23E-02 2.72E-03 3.36E-04 4.23E-05 1.93E-02 1.37E-03 8.18E-05 5.16E-06 3.13E-07 Efx 3.83E-02 2.28E-03 1.47E-04 9.39E-06 5.86E-07 4.38E-02 9.17E-04 1.37E-05 2.06E-07 3.29E-09 4.69E-04 2.52E-06 8.85E-09 3.76E-11 1.30E-13 order - 3.24 2.48 2.34 2.15 - 3.17 3.03 3.02 2.99 - 3.82 4.06 3.99 4.04 order - 4.07 3.95 3.97 4.00 - 5.58 6.06 6.05 5.97 - 7.54 8.15 7.88 8.18 Eux DNE DNE DNE DNE DNE 4.47E-02 1.42E-03 6.57E-05 3.91E-06 2.47E-07 1.32E-03 4.63E-05 1.35E-06 4.27E-08 1.27E-09 Ec 2.49E-02 1.48E-03 9.63E-05 6.18E-06 3.88E-07 2.49E-02 5.92E-04 9.12E-06 1.38E-07 2.21E-09 2.73E-04 1.66E-06 6.02E-09 2.41E-11 8.73E-14 order - - - - - - 4.98 4.43 4.07 3.98 - 4.83 5.10 4.99 5.07 order - 4.07 3.94 3.96 3.99 - 5.39 6.02 6.04 5.97 - 7.36 8.11 7.97 8.11 In Tables 3.1 and 3.2, we used nonuniform mesh in numerical test. The quantities tested have similar order of convergence compared to the order of convergence on uniform mesh. Another interesting observation is the order of convergence of Efx. Our numerical tests show that Efx converges at an order of 2k for all four sets of parameters, which is at least one order higher than the estimates in Theorem 3.3.5. 76 order order Eux order order EP 3.29 3.13 3.04 3.01 Table 3.3: Example 3.5.1. Error table when using central flux on uniform mesh. Ending time Te = 1, x ∈ [0, 2π]. L2 error 4.20E-03 4.31E-04 4.92E-05 5.99E-06 7.44E-07 3.18E-04 1.71E-05 1.03E-06 6.41E-08 4.00E-09 5.04E-04 2.10E-05 7.32E-07 2.36E-08 7.42E-10 3.21E-03 2.23E-04 1.43E-05 9.01E-07 5.60E-08 7.32E-05 1.02E-06 1.55E-08 2.41E-10 3.76E-12 7.75E-05 4.91E-07 2.65E-09 1.60E-11 1.13E-13 3.39E-02 4.49E-03 5.69E-04 7.14E-05 8.94E-06 3.34E-03 2.04E-04 1.27E-05 7.91E-07 4.94E-08 1.32E-02 6.17E-04 2.17E-05 6.76E-07 2.19E-08 2.97 2.99 3.00 3.00 1.87 1.97 1.99 2.00 2.92 2.98 2.99 3.00 3.85 3.96 3.99 4.01 6.16 6.04 6.01 6.00 4.03 4.01 4.00 4.00 4.21 4.05 4.01 4.00 - - - - - - - - - - - - Euxx 4.86E-01 1.33E-01 3.41E-02 8.57E-03 2.15E-03 4.31E-02 5.49E-03 6.89E-04 8.62E-05 1.08E-05 1.14E-01 1.11E-02 8.01E-04 5.21E-05 3.29E-06 Efx 9.58E-03 6.86E-04 3.90E-05 3.00E-06 1.51E-07 2.16E-04 3.07E-06 4.63E-08 7.18E-10 1.12E-11 2.15E-04 1.26E-06 6.17E-09 2.57E-11 9.96E-14 3.36 3.79 3.94 3.99 order - 3.80 4.14 3.70 4.31 - 6.14 6.05 6.01 6.00 - 7.42 7.67 7.91 8.01 4.42 4.83 5.01 4.95 order - 3.84 3.96 3.99 4.01 - 6.12 6.03 6.01 6.00 - 7.20 7.74 7.93 7.98 Ec 6.36E-03 4.44E-04 2.86E-05 1.80E-06 1.12E-07 1.41E-04 2.03E-06 3.10E-08 4.81E-10 7.51E-12 1.27E-04 8.66E-07 4.04E-09 1.66E-11 6.55E-14 N 40 80 160 320 640 20 40 80 160 320 10 20 40 80 160 N 40 80 160 320 640 20 40 80 160 320 10 20 40 80 160 P 2 P 3 P 4 P 2 P 3 P 4 4.58 4.84 4.96 4.99 order - 3.85 3.96 3.99 4.01 - 5.19 4.99 5.01 5.01 - 5.28 5.76 5.93 5.98 7.30 7.53 7.37 7.15 order - 3.85 3.96 3.99 4.01 - 6.16 6.04 6.01 6.00 - 7.40 7.78 7.94 7.99 Ef 3.21E-03 2.23E-04 1.43E-05 9.01E-07 5.60E-08 7.28E-05 1.02E-06 1.54E-08 2.39E-10 3.73E-12 7.63E-05 4.52E-07 2.05E-09 8.31E-12 3.27E-14 Eu 3.24E-03 2.25E-04 1.45E-05 9.10E-07 5.66E-08 1.88E-04 5.17E-06 1.63E-07 5.04E-09 1.57E-10 2.21E-04 5.70E-06 1.05E-07 1.72E-09 2.72E-11 Next, we test the order of convergence for quantities in Lemma 3.3.3. In Tables 3.5 and 3.6, we observe clean convergence order of 2k − 1, 2k + 1, 2k for (cid:107)(ζh)xx(cid:107), E[ζh], E[(ζh)x] when k is even and 2k, 2k + 2, 2k + 1 for these three quantities when k is odd. In Table 3.7, the order of convergence has some fluctuation, but the quantities are shown to have the same order of convergence as those in Tables 3.5 and 3.6. These convergence rates are consistent with the results in Lemma 3.3.3. 77 Table 3.4: Example 3.5.1. Error table when using flux parameters: α1 = 0.25, β1 = 2 h, 5 h , β2 = 0 on uniform mesh. Ending time Te = 1, x ∈ [0, 2π]. h, 9 N 80 160 320 640 1280 20 40 80 160 320 20 40 80 160 N 40 80 160 320 640 20 40 80 160 320 10 20 40 80 160 L2 error 1.41E-03 1.65E-04 2.03E-05 2.53E-06 3.16E-07 8.27E-04 3.92E-05 2.29E-06 1.40E-07 8.74E-09 5.10E-04 8.28E-06 1.87E-07 5.44E-09 Eu 3.24E-03 2.25E-04 1.45E-05 9.10E-07 5.66E-08 1.88E-04 5.17E-06 1.63E-07 5.04E-09 1.57E-10 2.21E-04 5.70E-06 1.05E-07 1.72E-09 2.72E-11 P 2 P 3 P 4 P 2 P 3 P 4 order EP order - 3.09 3.02 3.01 3.00 - 4.40 4.10 4.03 4.01 - 5.95 5.47 5.10 order - 3.85 3.96 3.99 4.01 - 5.19 4.99 5.01 5.01 - 5.28 5.76 5.93 5.98 8.17E-05 4.74E-06 2.92E-07 1.80E-08 1.22E-09 4.58E-05 5.20E-07 7.54E-09 1.16E-10 1.80E-12 2.08E-04 2.38E-07 1.04E-09 7.11E-12 Ef 3.21E-03 2.23E-04 1.43E-05 9.01E-07 5.60E-08 7.28E-05 1.02E-06 1.54E-08 2.39E-10 3.73E-12 7.63E-05 4.52E-07 2.05E-09 8.31E-12 3.27E-14 - 4.11 4.02 4.03 3.88 - 6.46 6.11 6.03 6.01 - 9.77 7.84 7.19 order - 3.85 3.96 3.99 4.01 - 6.16 6.04 6.01 6.00 - 7.40 7.78 7.94 7.99 Euxx DNE DNE DNE DNE DNE 2.98E-01 3.11E-02 3.72E-03 4.60E-04 5.74E-05 3.76E-01 1.24E-02 5.64E-04 3.29E-05 Efx 9.58E-03 6.86E-04 3.90E-05 3.00E-06 1.51E-07 2.16E-04 3.07E-06 4.63E-08 7.18E-10 1.12E-11 2.15E-04 1.26E-06 6.17E-09 2.57E-11 9.96E-14 order Eux order - - - - - - 3.26 3.06 3.02 3.00 - 4.92 4.47 4.10 order - 3.80 4.14 3.70 4.31 - 6.14 6.05 6.01 6.00 - 7.42 7.67 7.91 8.01 1.15E-02 1.34E-03 1.65E-04 2.05E-05 2.55E-06 6.63E-03 3.60E-04 2.18E-05 1.35E-06 8.43E-08 3.96E-03 6.76E-05 1.55E-06 4.53E-08 Ec 6.36E-03 4.44E-04 2.86E-05 1.80E-06 1.12E-07 1.41E-04 2.03E-06 3.10E-08 4.81E-10 7.51E-12 1.27E-04 8.66E-07 4.04E-09 1.66E-11 6.55E-14 - 3.11 3.02 3.01 3.01 - 4.20 4.05 4.01 4.00 - 5.87 5.45 5.09 order - 3.84 3.96 3.99 4.01 - 6.12 6.03 6.01 6.00 - 7.20 7.74 7.93 7.98 Lastly, we test the order of convergence for E∗ on uniform mesh for the four sets of parameters. Table 3.8 shows that E∗ has a convergence rate of at least 2k, and can go up to 2k + 2. Similar higher order of convergence behaviors exists in the literature [23, 65]. 78 Table 3.5: Example 3.5.1. Error table for intermediate quantities when using alternating flux on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. N 10 20 40 80 160 10 20 40 80 160 10 20 40 80 160 (cid:107)ζh(cid:107) error 3.96E-01 3.28E-02 2.08E-03 1.29E-04 8.17E-06 9.54E-03 1.93E-04 2.86E-06 4.31E-08 6.87E-10 1.35E-04 7.10E-07 2.50E-09 9.90E-12 3.58E-14 order - 3.60 3.98 4.01 3.98 - 5.63 6.08 6.05 5.97 - 7.57 8.15 7.98 8.11 (cid:107)(ζh)xx(cid:107) 3.02E+00 2.23E-01 1.42E-02 7.95E-04 9.54E-05 8.83E-02 1.76E-03 2.59E-05 3.91E-07 6.19E-09 1.41E-03 8.60E-06 4.24E-08 2.55E-10 2.22E-12 P 2 P 3 P 4 order - 3.76 3.98 4.15 3.06 - 5.65 6.09 6.05 5.98 - 7.36 7.66 7.38 6.85 E[ζh] 4.51E-02 1.57E-03 4.21E-05 1.17E-06 3.83E-08 7.52E-05 1.75E-07 3.42E-10 1.45E-12 6.76E-15 4.97E-07 1.20E-09 2.56E-12 2.98E-15 8.90E-18 order - 4.84 5.22 5.16 4.94 - 8.75 9.00 7.88 7.74 - 8.69 8.88 9.74 8.39 E[(ζh)x] 3.37E-01 1.82E-02 9.42E-04 5.50E-05 3.55E-06 1.31E-03 6.16E-06 2.45E-08 2.71E-10 2.59E-12 2.81E-05 9.88E-08 2.90E-10 7.48E-13 5.46E-15 order - 4.21 4.28 4.10 3.96 - 7.74 7.97 6.50 6.71 - 8.15 8.41 8.60 7.10 Table 3.6: Example 3.5.1. Error table for intermediate quantities when using flux parameters: α1 = 0.3, β1 = 0.4 order order E[ζh] order order 3.97 3.97 3.99 3.98 3.38 3.17 3.11 2.96 h , β2 = 0.4h on nonuniform mesh. Ending time Te = 1, x ∈ [0, 2π]. (cid:107)ζh(cid:107) error 1.46E-02 9.35E-04 5.96E-05 3.76E-06 2.38E-07 2.02E-02 4.31E-04 6.46E-06 9.72E-08 1.55E-09 2.27E-04 1.22E-06 4.30E-09 1.71E-11 6.17E-14 (cid:107)(ζh)xx(cid:107) 2.67E-01 2.57E-02 2.86E-03 3.30E-04 4.25E-05 1.78E-01 3.90E-03 5.83E-05 8.76E-07 1.40E-08 2.23E-03 1.30E-05 5.93E-08 4.66E-10 3.19E-12 E[(ζh)x] 2.13E-02 1.25E-03 7.56E-05 4.76E-06 3.19E-07 1.15E-03 5.48E-06 2.99E-08 1.84E-10 1.56E-12 2.06E-06 1.66E-08 6.95E-11 2.33E-13 8.34E-16 2.55E-03 7.74E-05 2.52E-06 7.74E-08 2.57E-09 5.73E-04 1.09E-06 3.27E-09 9.55E-12 4.28E-14 1.03E-06 4.13E-09 1.01E-11 1.82E-14 3.61E-17 5.04 4.94 5.02 4.91 9.04 8.38 8.42 7.80 5.52 6.06 6.06 5.97 5.55 6.06 6.05 5.97 7.96 8.67 9.12 8.98 7.42 7.78 6.99 7.19 7.54 8.15 7.98 8.11 4.09 4.05 3.99 3.90 7.71 7.52 7.34 6.89 - - - 6.96 7.90 8.22 8.13 - - - - - - - - - N 40 80 160 320 640 10 20 40 80 160 10 20 40 80 160 P 2 P 3 P 4 79 Table 3.7: Example 3.5.1. Error table for intermediate quantities when using central flux on uniform mesh. Ending time Te = 1, x ∈ [0, 2π]. N 40 80 160 320 640 20 40 80 160 320 10 20 40 80 160 (cid:107)ζh(cid:107) error 4.53E-03 3.15E-04 2.02E-05 1.27E-06 7.92E-08 1.03E-04 1.44E-06 2.18E-08 3.38E-10 5.28E-12 1.06E-04 6.37E-07 2.88E-09 1.17E-11 4.63E-14 order - 3.85 3.96 3.99 4.01 - 6.16 6.04 6.01 6.00 - 7.37 7.79 7.94 7.98 (cid:107)(ζh)xx(cid:107) 3.84E-02 3.03E-03 1.29E-04 1.59E-05 5.73E-07 9.27E-04 1.29E-05 1.99E-07 3.05E-09 4.76E-11 1.05E-03 7.12E-06 3.20E-08 1.84E-10 2.45E-12 P 2 P 3 P 4 order - 3.66 4.55 3.03 4.79 - 6.17 6.02 6.03 6.00 - 7.21 7.80 7.44 6.23 E[ζh] 3.04E-05 1.16E-06 1.79E-07 7.14E-09 2.90E-10 4.27E-08 1.75E-10 4.23E-13 2.91E-16 1.28E-18 7.67E-07 3.03E-09 5.78E-12 9.33E-15 3.04E-17 order - 4.71 2.70 4.65 4.62 - 7.93 8.69 10.50 7.83 - 7.99 9.03 9.28 8.26 E[(ζh)x] 2.79E-04 8.31E-06 1.31E-06 5.08E-08 2.11E-09 5.25E-06 4.32E-08 1.64E-10 3.24E-14 1.41E-15 2.26E-05 7.70E-08 6.36E-11 5.24E-13 8.82E-16 order - 5.07 2.66 4.69 4.59 - 6.93 8.04 12.31 4.52 - 8.20 10.24 6.92 9.21 Table 3.8: Example 3.5.1. Postprocessing error table for the four sets of parameters. Ending time Te = 1, uniform mesh on x ∈ [0, 2π]. The first row below labels the parameters by ( ˜α1, ˜β1, ˜β2). P 2 Fluxes N 10 20 40 80 160 10 20 40 80 160 10 20 40 80 160 P 3 P 4 (0.5,0,0) E∗ order (0, 0, 0) E∗ order (0.3, 0.4, 0.4) E∗ order (0.25, {2, 5, 9}, 0) order E∗ 1.00E+00 2.84E-01 2.11E-02 1.37E-03 8.69E-05 1.00E+00 6.04E-02 5.39E-04 3.28E-06 3.14E-08 1.00E+00 4.54E-02 1.32E-04 1.70E-07 1.79E-10 - 1.81 3.75 3.94 3.98 - 4.05 6.81 7.36 6.70 - 4.46 8.42 9.60 9.89 2.81E-01 3.71E-02 3.23E-03 2.24E-04 1.44E-05 1.00E+00 6.29E-02 6.05E-04 5.04E-06 6.49E-08 1.00E+00 4.54E-02 1.32E-04 1.70E-07 1.80E-10 - 2.92 3.52 3.85 3.96 - 3.99 6.70 6.91 6.28 - 4.46 8.42 9.60 9.89 80 1.00E+00 1.20E-01 9.63E-03 7.55E-04 5.13E-05 1.00E+00 6.05E-02 5.26E-04 2.82E-06 2.04E-08 1.00E+00 4.54E-02 1.32E-04 1.70E-07 1.79E-10 - 3.06 3.64 3.67 3.88 - 4.05 6.85 7.54 7.11 - 4.46 8.42 9.60 9.89 1.53E-01 8.05E-02 2.68E-03 1.20E-04 6.99E-06 1.00E+00 7.02E-02 5.46E-04 2.91E-06 1.79E-08 1.00E+00 4.54E-02 1.36E-04 1.66E-07 1.75E-10 - 0.93 4.91 4.49 4.10 - 3.83 7.01 7.55 7.34 - 4.46 8.39 9.67 9.89 Chapter 4 Sparse grid central DG methods for linear hyperbolic systems In this chapter, we develop sparse grid central discontinuous Galerkin (CDG) method for the following time-dependent linear hyperbolic system with variable coefficients ∂(Ai(t, x)u) = 0, x ∈ Ω, (4.1) d(cid:88) ∂u ∂t + i=1 ∂xi subject to appropriate initial and boundary conditions. In the expression above, d ≥ 2 is the spatial dimension of the problem, u(t, x) = (u1(t, x),··· , um(t, x))T is the unknown function, Ai(t, x) ∈ Rm×m, i = 1, . . . , d are the given smooth variable coefficients. We assume Ω = [0, 1]d in this chapter, but the discussion can be easily generalized to arbitrary box-shaped domains. The model (4.1) arises in many contexts [46], such as simulations of acoustic, elastic waves, and Maxwell’s equations in free space. The scheme we develop in this chapter can also apply to the case when Ai(t, x) is defined through another set of equations that can be nonlinearly coupled with u, such as the models in kinetic plasma waves and incompressible flows. Similar to [38], in this chapter, we restrict our attention to smooth solutions of (4.1). It is known that for non-smooth solutions, adaptivity should be invoked to capture discontinuity 81 like structures. This can be achieved using the idea in [39] and is left for our future work. Based on the sparse grid DG scheme constructed in [38], the goal of the present chapter is to design and analyze the sparse grid CDG method. Motivated by the Riemann-solver-free property and large CFL number allowance of CDG methods, we develop sparse grid CDG method to compute hyperbolic systems efficiently. We investigate stability, convergence rate and CFL condition of the resulting scheme. A novelty of this work is the design of the scheme for non-periodic problems, where a new hierarchical representation of the solution is presented, which results in a sparse finite element space that can be defined on the dual mesh. L2 projection results are studied for this space, which helps the convergence proof of the schemes for initial-boundary value problems. The rest of this chapter is organized as follows: in Section 4.1, we construct the sparse grid CDG formulations for periodic and non-periodic problems, and perform numerical study of the CFL conditions. In Section 4.2, we prove L2 stability and error estimates for scalar equations. The numerical performance is validated in Section 4.3 by several benchmark tests, including scalar transport equations, acoustic and elastic waves. The contents of this chapter has been published in [75]. 4.1 Numerical Scheme In this section, we define and discuss the properties of the proposed sparse grid CDG meth- ods. For convenience of notations, we rewrite (4.1) in a component-wise form as ∂ul ∂t + ∇ · (Al(t, x)u) = 0, l = 1,··· , m, x ∈ Ω, (4.2) 82 where Al(t, x) = (Al d(t, x))T ∈ Rd×m denotes a collection of the l-th row of each matrix Ai. The problem is solved with given initial value u(0, x) = u0(x), and periodic 1(t, x),··· , Al or Dirichlet type boundary conditions. We proceed as follows. First, we introduce the scheme for periodic problems. In this setting, the finite element space on the primal and dual mesh can be defined in similar ways. Then, we discuss the implementation details and perform numerical study of the CFL conditions. Finally, we consider the more complicated non-periodic problems, for which a new sparse finite element space will be introduced on the dual mesh. 4.1.1 Periodic problems To define the sparse finite element space, we first review the hierarchical decomposition of piecewise polynomial space in one dimension [76]. Consider a general interval [a, b], we define the n-th level mesh Ωn([a, b]) to be a uniform partition of 2n cells with length hn = 2−n(b−a) and Ij n = [a + jhn, a + (j + 1)hn], j = 0, . . . , 2n − 1, for any n ≥ 0. Let n ([a, b]) := {v : v ∈ P k(Ij V k n), ∀ j = 0, . . . , 2n − 1} be the usual piecewise polynomials of degree at most k on Ωn. Then, we have the nested structure 0 ([a, b]) ⊂ V k V k 1 ([a, b]) ⊂ V k 2 ([a, b]) ⊂ V k 3 ([a, b]) ⊂ ··· Similar to [76], we can now define the multiwavelet subspace W k n ([a, b]), n = 1, 2, . . . as the orthogonal complement of V k n−1([a, b]) in V k n ([a, b]) with respect to the L2 inner product 83 on [a, b], i.e., n−1([a, b]) ⊕ W k V k n ([a, b]) = V k n ([a, b]), W k n ([a, b]) ⊥ V k n−1([a, b]). For notational convenience, we let W k 0 ([a, b]) := V k 0 ([a, b]), which is the standard piecewise polynomial space of degree k on [a, b]. This gives the hierarchical decomposition V k n ([a, b]) n ([a, b]) =(cid:76) on Ωn as V k 0≤l≤n W k l ([a, b]). For a d dimensional domain [a, b]d, we recall some basic notations about multi-indices. 0, where N0 denotes the set of nonnegative integers, For a multi-index α = (α1,··· , αd) ∈ Nd the l1 and l∞ norms are defined as (cid:88)d |α|1 := αi, |α|∞ := max 1≤i≤d αi. i=1 The component-wise arithmetic operations and relational operations are defined as α · β := (α1β1, . . . , αdβd), c · α := (cα1, . . . , cαd), 2α := (2α1, . . . , 2αd), α ≤ β ⇔ αi ≤ βi, ∀i, α < β ⇔ α ≤ β and α (cid:54)= β. By making use of the multi-index notation, we denote by l = (l1,··· , ld) ∈ Nd 0 the mesh level in a multivariate sense. We define the tensor-product mesh grid Ωl([a, b]d) = ([a, b]) ⊗ ··· ⊗ Ωld Ωl1 the grid Ωl, we denote by Ij ([a, b]) and the corresponding mesh size hl = (hl1 ,··· , hld ). Based on l = {x : xi ∈ I ji li , i = 1,··· , d} as an elementary cell, and l ([a, b]d) := {v : v(x) ∈ Qk(Ij Vk l ), 0 ≤ j ≤ 2l − 1} = V k l1,x1 ([a, b]) × ··· × V k ld,xd ([a, b]) 84 as the standard tensor-product piecewise polynomial space on this mesh, where Qk(Ij l ) denotes the collection of polynomials of degree up to k in each dimension on cell Ij l . l = (N,··· , N ), the grid and space will be further denoted by ΩN ([a, b]d) and Vk N ([a, b]d), If respectively. Based on a tensor-product construction, the multi-dimensional increment space can be defined as Therefore, we have Vk mation space we consider, is defined by Wk l1,x1 l ([a, b]d) = W k N ([a, b]d) =(cid:76)|l|∞≤N l∈Nd 0 ˆVk N ([a, b]d) := ([a, b]) × ··· × W k ([a, b]). ld,xd Wk l ([a, b]d). The sparse finite element approxi- (cid:77) |l|1≤N l∈Nd 0 Wk l ([a, b]d). This is a subset of Vk 1)d2N N d−1) [76], which is significantly less than that of Vk N ([a, b]d), and its number of degrees of freedom scales as O((k + N ([a, b]d) with exponential de- pendence on N d. This is the key to computational savings in high dimensions. The standard CDG schemes [48, 50] is characterized by numerical approximations on two sets of overlapping grids: primal and dual meshes. Now, we are ready to incorporate the sparse finite element space defined above into the CDG framework. For the domain under consideration Ω = [0, 1]d, we let ΩN,P := ΩN ([0, 1]d) be the primal mesh and ΩN,D, which is the periodic extension of ΩN ([−hN /2, 1 − hN /2]d) restricted to [0, 1]d, be the dual mesh. Similarly, we let ˆVk N,D to be the periodic extension of N ([−hN /2, 1 − hN /2]d) restricted to [0, 1]d. Here and below, the subscripts P and D ˆVk N ([0, 1]d) and ˆVk N,P := ˆVk represent the quantities defined on the primal and dual mesh, respectively. 85 The approximation properties for the sparse finite element space have been established in previous work [76, 38]. By using a lemma in [38], we can have estimates for L2 projection operator onto the spaces ˆVk N,P , ˆVk N,D. To facilitate the discussion, below we introduce some notations about norms and semi- Hs(I j N,G) (cid:107)v(cid:107)2 0≤j≤2N−1 is the standard Sobolev norm on Ij Hs(ΩN,G) = (cid:80) norms. Let G = P, D, on primal or dual mesh ΩN,G, we use (cid:107) · (cid:107)Hs(ΩN,G) to denote the standard broken Sobolev norm, i.e. (cid:107)v(cid:107)2 (cid:107)v(cid:107) N,G, (and s = 0 is used to denote the L2 norm). Similarly, we use | · |Hs(ΩN,G) to denote the broken Sobolev semi-norm, and (cid:107) · (cid:107)Hs(Ωl,G),|·|Hs(Ωl,G) to denote the broken Sobolev norm and semi-norm that are supported on a general grid Ωl,G. For any set L = {i1, . . . ir} ⊂ {1, . . . d}, we define Lc to be the complement set of L in {1, . . . d}. For a non-negative integer α and set L, we define the semi-norm on any domain denoted by Ω(cid:48) j N,G) , where Hs(I |v| Hα,L(Ω(cid:48)) := and (cid:33) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L2(Ω(cid:48)) v , ··· ∂α ∂xα ir ∂α ∂xα i1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:32)  max |v|Hq+1(Ω(cid:48)) := max 1≤r≤d |v| Hq+1,L(Ω(cid:48)) L⊂{1,2,··· ,d} |L|=r  , which is the norm for the mixed derivative of v of at most degree q + 1 in each direction. In this chapter, we use the notation A (cid:46) B to represent A ≤ constant × B, where the constant is independent of N and the mesh level considered. The following results are obtained from Lemma 3.2 in [38]. Lemma 4.1.1 (L2 projection estimate). Let PP , PD be L2 projections onto the spaces N,D, respectively, then for k ≥ 1, 1 ≤ q ≤ min{p, k}, and v ∈ Hp+1(Ω), which is N,P , ˆVk ˆVk 86 periodic on Ω, N ≥ 1, d ≥ 2, we have for G = P, D, |PGv − v|Hs(ΩN,G) (cid:46) N d2−N (q+1)|v|Hq+1(Ω) 2−N q|v|Hq+1(Ω) s = 0, s = 1. (4.3) N,G, ΓN,G :=(cid:83) This lemma shows that the L2 norm and H1 semi-norm of the projection error scale like O(N d2−N (k+1)) and O(2−N k) with respect to N when the function v has bounded mixed derivatives up to enough degrees. This lemma will be used in Theorem 4.2.2 to establish convergence of the scheme. Now, we are ready to formulate the sparse grid CDG scheme. Below we review some standard notations about jumps and averages of piecewise functions. With G = P or D, let Th,G be the collection of all elementary cell Ij T∈ΩN,G ∂T be the union of the interfaces for all the elements in ΩN,G (here we have taken into account the periodic boundary condition when defining ΓN,G) and S(ΓG) := ΠT∈ΩN,G L2(∂T ) be the set of L2 functions defined on ΓN,G. For any q ∈ S(ΓN,G) and q ∈ [S(ΓN,G)]d, we define their averages {q},{q} and jumps [q], [q] on the interior edges as follows. Suppose e is an interior edge shared by elements T+ and T−, either on primal or dual mesh, we define the unit normal vectors n+ and n− on e pointing exterior of T+ and T−, respectively, then [q] = q−n− + q+n+, {q} = [q] = q− · n− + q+ · n+, {q} = (q− + q+), 1 2 (q− + q+). 1 2 The semi-discrete sparse grid CDG scheme for (4.2), based on the weak formulation introduced in [48, 50], is defined as follows: we find ul h ∈ ˆVk N,P and vl h ∈ ˆVk N,D, such that 87 ∀ l = 1,··· , m (cid:90) Ω (cid:90) (ul h)t ϕh dx = (vl h)t ψh dx = Ω τmax 1 Ω (cid:90) − (cid:88) (cid:90) − (cid:88) e∈ΓN,P 1 Ω τmax e∈ΓN,D e (cid:90) h − ul (vl h) ϕh dx + e Al(t, x)vh · [ϕh] ds, (cid:90) h) ψh dx + Ω Al(t, x)uh · [ψh] ds, (cid:90) h − vl (ul (cid:90) Ω Al(t, x)vh · ∇ϕh dx (4.4) Al(t, x)uh · ∇ψh dx (4.5) for any ϕh ∈ ˆVk N,P and ψh ∈ ˆVk N,D, where uh = (u1 h,··· , um h ), vh = (v1 h,··· , vm h ) and τmax is an upper bound for the time step due to the CFL restriction (see Section 4.1.3 for detailed discussions). 4.1.2 Discussions on implementations Here, we briefly discuss some details about the implementation of the scheme. We perform the computation by using orthonormal multiwavelet bases constructed by Alpert [4]. In 1D, the bases of W k l ([0, 1]) are denoted by vj p,l(x), p = 1,··· , k + 1, j = 0,··· , 2l−1 − 1 and they satisfy(cid:82) b p,l(x)vj(cid:48) a vj of the basis functions for k = 0, 1 and l = 0, 1, 2. The bases in W k l p(cid:48),l(cid:48)(x)dx = δpp(cid:48)δjj(cid:48)δll(cid:48). Figures 4.1a and 4.2a provide illustrations in multi-dimensions are defined by tensor products d(cid:89) i=1 vs = vj p,l := ji v pi,li (xi), pi = 1,··· , k + 1, ji = 0,··· , max(0, 2li−1 − 1), 88 where we have used the notation s = (l, j, p) and si = (li, ji, pi) to denote the multi-index for the bases. As for temporal schemes, we can use the total variation diminishing Runge-Kutta (TVD- RK) methods [73] to solve the ordinary differential equations for the coefficients resulting from the discretization. To calculate the right-hand-side of (4.4)-(4.5), the fast matrix-vector product by LU split or LU decomposition algorithms [69, 70, 64] can be applied, by which one can decompose all calculations into one dimensional operations. Below, we briefly describe the LU decomposition algorithm for the calculation of the following matrix-vector product which appears at the right-hand-side of (4.4)-(4.5) (cid:88) bj = s:|l|1≤N fst1 s1,j1 ··· td sd,jd , where fs can be the coefficient of the basis in sparse grid space and ti si,ji , i = 1,··· , d, are the corresponding one-dimensional transform of coefficients from basis vsi to basis vji i-th dimension in our scheme. Note that we have n = 2N (k + 1) one-dimensional bases in in the each dimension, and we use vsi to denote the si-th basis. The bases are ordered according to grid increment. Using Algorithm 1 in [70], we should calculate all the one-dimensional transform along each direction associated with a block lower triangular matrix, and then calculate all the one-dimensional transforms having a block upper triangular structure. The fast matrix-vector product fs → bj on sparse grid with LU decomposition can be proceeded as follows. 1. Calculate (block) LU decomposition ti m,j, s, j = 1,··· , n, for i = 1,··· , d, where P i, Qi are the permutation matrices, li, ui are lower and upper m=1(P l)i s,m(uQ)i s,j =(cid:80)n triangular matrices. 89 2. Compute the transform with a (block) lower triangular matrix for i = 1,··· , d, b (cid:48) s1,··· ,si−1,s i,si+1,··· ,sd si:l1+···+ld≤N fs(P l)i . (cid:48) si,s i 3. Compute the transform with a (block) upper triangular matrix for i = 1,··· , d, ←(cid:80) bs ←(cid:80) (cid:48) i:l1+···+li−1+l s (cid:48) i+li+1+···+ld≤N b (cid:48) s1,··· ,si−1,s i,si+1,··· ,sd (uQ)i (cid:48) s i,si . Note that in step 1, the LU decomposition pivots only from rows or columns in the same mesh level to maintain the hierarchical structure. This pivoting can be successfully done in the sparse grid CDG scheme, but not in the sparse grid DG scheme, for which additional splitting of the flux terms are deemed necessary for variable coefficient case. For the integrals involving variable-coefficient, we use Gaussian quadrature to compute these terms. Since these integrals are multi-dimensional integrations, we use the so-called unidirectional principle to separate the integration into multiplication of one-dimensional integrals. For example, if φ(x) = φ1(x1)··· φd(xd) is separable, (cid:90) (cid:90) (cid:90) φ(x) = Ω [a,b] φ1(x1)··· φd(xd). [a,b] When the variable coefficient Ai(t, x) is separable, we can use unidirectional principle directly. If it is not separable, we can find Ah i (t, x) as the L2 projection of Ai(t, x) onto the sparse grid finite element space, and then use Ah i (t, x) to compute the integrals. 4.1.3 Discussions on CFL conditions It is well known that the CDG schemes allow larger CFL numbers than the standard DG methods except for piecewise constant approximations [50, 60]. Here, we perform a numerical study of the CFL conditions of DG [27], CDG [51], sparse grid DG [38], and the sparse 90 Table 4.1: CFL numbers of the DG method, CDG method, sparse grid DG method and sparse grid CDG method with piecewise degree k polynomials, Runge-Kutta method of order ν for Example 4.3.1 with d=2. The CFL numbers of the sparse grid DG/CDG methods are measured with regard to the most refined mesh hN . k ν = 2 ν = 3 ν = 4 1 0.33 0.40 0.46 DG 2 – 3 – 0.20 0.23 0.13 0.14 CDG 2 – 3 – 0.36 0.52 0.24 0.35 1 0.48 0.66 0.90 2 – 3 – 0.41 0.46 0.25 0.28 0.66 0.81 0.92 sparse grid DG sparse grid CDG 1 1 2 – 3 – 0.65 0.94 0.44 0.62 0.87 1.17 1.58 grid CDG schemes. We only consider the two-dimensional case solving constant coefficient equation ut + ux1 + ux2 = 0 for now. The results are listed in Table 4.1. The CFL number of DG method is obtained from Table 2.2 in [27]. The rest of the table is computed by eigenvalue analysis of the discretization matrix, and by requiring the amplification of the eigenvalues to be bounded by 1 in magnitude. We observe that the sparse grid DG method has CFL number that is about two times the CFL number of the standard DG method. The sparse grid CDG method offers the largest CFL conditions among all four methods. Here, as a side note, we find that the CFL number for two-dimensional CDG method is larger than the CFL number for one-dimensional CDG method in [51]. This table shows that one advantage of the sparse grid CDG method is the ability to take large time steps for time evolution problems. In general, further numerical results suggest that for equation ut + c1ux1 + c2ux2 = 0, the CFL number for sparse grid DG and sparse grid CDG method will change with the value of the coefficients c1, c2. Results in higher dimensions are yet to be studied. A preliminary calculation shows that for equation ut + ux1 + ux2 + ux3 = 0 the CFL conditions for CDG, sparse grid DG and sparse grid CDG methods in 3D are all higher than those for the 2D case in Table 4.1. The sparse grid CDG method still possesses the largest CFL number among all four methods. Those interesting issues will be investigated in our future work. 91 4.1.4 Non-periodic problems Here, we consider non-periodic problems, where equation (4.1) or (4.2) is supplemented by Dirichlet boundary condition on the inflow edges. In this case, we can no longer use periodicity to define the finite element space on the dual mesh, and a new grid hierarchy needs to be introduced. Recall that for standard CDG methods with non-periodic boundary condition on the domain [0, 1], the finite element space on dual mesh with cell size hn = 1/2n is represented by n,D = {v : v ∈ P k(Ij V k n,D), ∀ j = 0, . . . , 2n}, (4.6) where the mesh is partitioned as I0 n,D = [0, hn 2 ], n,D = [(j − 1 Ij 2 )hn, (j + 1 2 )hn], j = 1, . . . , 2n − 1, n,D = [1 − hn I2n 2 , 1], which consists of 2n−1 cells of size hn, and two cells at the left and right ends of size hn/2. It is easy to see that this space does not have nested structures, i.e. V k n−1,D (cid:54)⊂ V k n,D. Therefore, we need a new hierarchy to define the increment polynomial spaces. For a fixed refined mesh level N, we define the following grid Ωl,N,D on level l, l = 0 . . . N, by a collection of cells as l,N,D = [0, hl − hN I0 2 ], l,N,D = [1 − hN I2l , 1], , (j + 1)hl − hN ], j = 1, . . . , 2l − 1, 2 2 l,N,D = [jhl − hN Ij 2 which consists of 2l − 1 cells of size hl, and a cell at the left end of size hl − hN 2 , and a cell at the right end of size hN 2 . This grid structure is naturally nested, and therefore V k l,N,D 92 which consists of piecewise polynomials of degree k defined on Ωl,N,D are also nested, and V k N,N,D = V k N,D as defined in (4.6). Then the definitions of sparse finite element space in Section 4.1.1 can be naturally extended here. We let W k l,N,D, l = 1, . . . N be a complement set of V k l−1,N,D in V k l,N,D, i.e. l−1,N,D ⊕ W k V k l,N,D = V k l,N,D. However, we no longer require W k l,N,D to be L2 orthogonal to V k l−1,N,D, because such defi- nition will be difficult to implement in practice. Instead, we define W k l,N,D to be a span of basis functions that are shifted basis functions of W k l space defined in Section 4.1.1, namely, W k l,N,D = W k l ([−hN 2 , 1 − hN 2 l ≥ 1. ])(cid:12)(cid:12)[0,1], N,D =(cid:76) By denoting W k 0,N,D = V k 0,N,D, we have decomposed V k 0≤l≤N W k l,N,D. Illustration of basis functions by such definitions for k = 0, 1 and l = 0, 1, 2 can be found in Figures 4.1b and 4.2b. The dimension of W k are 2l−1(k + 1). 0,N,D is 2(k + 1), while the dimensions of W k l,N,D, l = 1, . . . N Finally, the sparse finite element space on the dual mesh of domain [0, 1]d is defined as ˆ˜Vk N,D := Wk l,N,D, (cid:77) |l|1≤N l∈Nd 0 93 ×···×W k l,N,D = W k Wk N,D = l,N,D, and its number of degrees of freedom scales as O(2d−1(k + 1)d2N N d−1) . This is a subset of the full grid space Vk ld,N,D,xd l1,N,D,x1 where Wk (cid:76)|l|∞≤N l∈Nd 0 (the proof is similar to Lemma 2.3 in [76]), which is larger than that of ˆVk N,P , but still significantly less than that of Vk N,D with exponential dependence on N d. We will now investigate the approximation property of the space ˆ˜Vk N,D. We can obtain the following result, which essentially states that the L2 projection onto this newly constructed space has the same order of accuracy as PP , PD in Lemma 4.1.1. Lemma 4.1.2 (L2 projection estimate onto ˆ˜Vk N,D ). Let ˜PD be the L2 projection onto the N,D, then for k ≥ 1, 1 ≤ q ≤ min{p, k}, and v ∈ Hp+1(Ω), N ≥ 1, d ≥ 2, we have space ˆ˜Vk | ˜PDv − v|Hs(ΩN,D) (cid:46) N d2−N (q+1)|v|Hq+1(Ω) 2−N q|v|Hq+1(Ω) s = 0, s = 1. (4.7) Proof. The proof follows same procedure as Appendix A in [38]. We will mainly highlight the difference in the proof (see Steps 1 and 2 below). The main difference lies in the fact that all the hierarchical spaces (and associated projections) have dependence not only on l, but also on the finest mesh level N. Step 1: Decomposition of ˜PD into tensor products of one-dimensional increment pro- jections. We denote P k l,N,D as the standard L2 projection operator from L2([0, 1]) to V k l,N,D, and the induced increment projection P k P k 0,N,D, l,N,D − P k l−1,N,D, Qk l,N,D := if l = 1, . . . N, if l = 0, 94 and further denote ˜Pk N,D := (cid:88) |l|1≤N l∈Nd 0 Qk l1,N,D,x1 ⊗ ··· ⊗ Qk ld,N,D,xd , where the last subindex of Qk li,N,D,xi indicates that the increment operator is defined in xi- direction. We can verify that ˜PD = ˜Pk N,D. In fact, for any v, it’s clear that ˜Pk N,Dv ∈ ˆ˜Vk N,D. Therefore, we only need (cid:90) Ω ( ˜Pk N,Dv − v)w dx = 0, ∀ w ∈ ˆ˜Vk N,D. (4.8) It suffices to show (4.8) for v ∈ C∞(Ω) which is a dense subset of L2(Ω). In fact, we have v = Pk N,Dv + v − Pk N,Dv, ( ˜Pk N,Dv − v)wdx = N,Dv)wdx + is the L2 projection onto the full grid space (cid:90) Ω (v − Pk N,Dv)wdx Qk l1,N,D,x1 ⊗ ··· ⊗ Qk ld,N,D,xd v) w dx. where Pk N,D = P k N,N,D,x1 Vk N,D. Therefore, (cid:90) Ω ⊗ ··· ⊗ PN,N,D,xd (cid:90) (cid:90) Ω ( ˜Pk N,Dv − Pk (cid:88) = − ( |l|∞≤N,|l|1>N Ω l∈Nd 0 The last term in the first row of the equality above vanishes because w ∈ ˆ˜Vk N,D ⊂ Vk N,D. 95 In addition, for any l ≥ 1, φ ∈ L2([0, 1]), ϕ ∈ V k l−1,N,D (cid:90) Qk l,N,Dφ ϕdx = [0,1] [0,1] (I − P k l−1,N,D)φ ϕdx − (cid:90) [0,1] (I − P k l,N,D)φ ϕdx = 0, Therefore, by properties of the tensor product projections ( ˜Pk N,Dv − v)wdx = 0, ∀w ∈ ˆ˜Vk N,D, and the proof for ˜PD = ˜Pk N,D is complete. Step 2: Estimation of the increment projections. For a function v ∈ Hp+1([0, 1]), we have the convergence property of the L2 projection P k 1 ≤ q ≤ min{p, k}, s = 0, 1, l,N,D as follows: for any integer q with |P k l,N,Dv − v| Hs(I j l,N,D ) ≤ ck,s,q(hj l,N )(q+1−s)|v| Hq+1(I j l,N,D j = 1,··· , 2l − 1, , ) where the mesh size hj l,N = hl − hN /2, j = 0 hl, j = 1,··· , 2l − 1, hN /2, j = 2l. The estimation above directly applies for Qk 0,N,D = P k 0,N,D. For l ≥ 1, by simple algebra, (cid:90) (cid:90) Ω  96 we have |Qk l,N,Dv| Hs(I |Qk l,N,Dv| Hs(I j l,N,D ) j l,N,D ) ≤ ˜ck,s,q2−l(q+1−s)|v| Hq+1(I (cid:98)j/2(cid:99) l−1,N,D ≤ ck,s,q(hl)(q+1−s)|v| Hq+1(I j l,N,D ) j = 2,··· , 2l − 1, , ) + ck,s,q(hl−1 − hN /2)(q+1−s)|v| < ˜ck,s,q2−l(q+1−s)|v| Hq+1(I0 l−1,N,D , ) Hq+1(I0 l−1,N,D , ) j = 0, 1, |Qk l,N,Dv| Hs(I2l l,N,D = 0, ) with ˜ck,s,q = ck,s,q(1 + 2q+1−s). The rest of the proof is then very similar to Appendix A in [38], and is omitted. We now provide a numerical validation of Lemma 4.1.2 by considering the error of pro- jection ˜PD for a smooth function u(x) = exp  , x ∈ [0, 1]d. xi  d(cid:89) i=1 (4.9) In Table 4.2, we report the L2 errors and the associated orders of accuracy for k = 1, 2, 3, d = 2, 3. It is clear that the predicted order of accuracy is achieved. With the aid of this space, the semi-discrete scheme can now be defined similarly as in (4.4)-(4.5) by using the space on the dual mesh as ˆ˜Vk N,D, and replacing the numerical values on the boundary of the domain by corresponding functions in the Dirichlet boundary conditions. We now comment on the implementation of this algorithm. As can be seen from Figures 97 Table 4.2: L2 errors and orders of accuracy for L2 projection operator ˜PD of (4.9) onto ˆ˜Vk N,D when d = 2 and d = 3. N is the number of mesh levels, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . L2 error order L2 error order L2 error order N hN k = 1 3 4 5 6 7 3 4 5 6 7 1/8 1/16 1/32 1/64 1/128 1/8 1/16 1/32 1/64 1/128 8.93E-04 2.61E-04 7.34E-05 2.00E-05 5.35E-06 6.19E-04 1.90E-04 5.71E-05 1.67E-05 4.80E-06 – 1.77 1.83 1.88 1.90 – 1.70 1.73 1.77 1.80 k = 2 d = 2 9.14E-06 1.29E-06 1.77E-07 2.37E-08 3.11E-09 d = 3 4.93E-06 7.45E-07 1.10E-07 1.58E-08 2.24E-09 – 2.82 2.87 2.90 2.93 – 2.73 2.76 2.80 2.82 k = 3 6.40E-08 4.45E-09 3.01E-10 1.98E-11 1.29E-12 3.18E-08 2.36E-09 1.69E-10 1.18E-11 9.35E-13 – 3.85 3.89 3.93 3.94 – 3.75 3.80 3.84 3.66 4.1b and 4.2b, there are two types of basis functions in 1D for the dual space. • Type 1 bases (for l ≥ 0), which are the shifted and truncated multiwavelet bases. • Type 2 bases (for l = 0), which are the Legendre polynomials of degree up to k on [1 − hN 2 , 1]. Clearly, Type 1 bases are orthogonal to Type 2 bases, because their support do not overlap. Type 2 bases are orthogonal to each other due to the definition of Legendre polynomials. However, Type 1 bases are no longer orthogonal to each other, due to the domain shift and truncation. However, only the left-most element on each level are changed. For other bases in that level, they will still retain orthogonality. The bases on left-most element in all level are orthogonal to other bases, but not to each other, i.e., the bases defined on left-most element in different levels are not orthogonal. This implies that although the mass matrix is not identity here, it will have block structures and be sparse. 98 (a) Primal mesh. Number of bases for l = 0, 1, 2 are 1, 1, 2. (b) Dual mesh. Number of bases for l = 0, 1, 2 are 2, 1, 2. Figure 4.1: Illustration of one-dimensional bases on different levels for k = 0: non-periodic problems. Different colors represent different bases. 4.2 Stability and convergence In this section, we prove L2 stability and error estimates for the sparse grid CDG scheme for the scalar equation. We consider both periodic and non-periodic boundary conditions. For periodic problems, (4.2) reduces to + ∇ · (Au) = 0, x ∈ Ω, ∂u ∂t (4.10) where A = (A1(t, x),··· , Ad(t, x)), and (cid:107)A(cid:107)L∞(Ω) < ∞,(cid:107)∇ · A(cid:107)L∞(Ω) < ∞. We assume Ai (cid:54)= 0 to avoid the discussion of different boundary conditions for degenerating coefficients. However, there is no difficulty to extend the proof below to degenerating case. For non- 99 (a) Primal mesh. Number of bases for l = 0, 1, 2 are 2, 2, 4. (b) Dual mesh. Number of bases for l = 0, 1, 2 are 4, 2, 4. Figure 4.2: Illustration of one-dimensional bases on different levels for k = 1: non-periodic problems. Different colors represent different bases. periodic problems, the following inflow boundary conditions are prescribed, u(t, x)|∂Ω = gi(t,··· , xi−1, xi+1,··· , xd) where ∂Ω xin i := {x ∈ Ω|xi = 0}, {x ∈ Ω|xi = 1}, if Ai(t, x) > 0, if Ai(t, x) < 0. Correspondingly, we denote the outflow edges by xin i   ∂Ω xout i := {x ∈ Ω|xi = 1}, {x ∈ Ω|xi = 0}, if Ai(t, x) > 0, if Ai(t, x) < 0. 100 The scheme for periodic case reduces to: to find uh ∈ ˆVk (cid:90) N,P and vh ∈ ˆVk (cid:90) (cid:90) (cid:90) (uh)t ϕh dx = (vh − uh) ϕh dx + vhA · ∇ϕh dx − (cid:88) uhA · ∇ψh dx − (cid:88) e∈ΓN,P e∈ΓN,D e (cid:90) e Ω (cid:90) Ω N,D, such that vhA · [ϕh] ds, (4.11) uhA · [ψh] ds, (4.12) Ω (cid:90) Ω 1 τmax Ω (cid:90) 1 τmax Ω (vh)t ψh dx = (uh − vh) ψh dx + for any ϕh ∈ ˆVk and enforce uh|∂Ω N,P and ψh ∈ ˆVk = vh|∂Ω xin i xin i N,D. For non-periodic problems, we require vh, ψh ∈ ˆ˜Vk = gi on the boundary interface. N,D, We can prove that the schemes retain similar stability properties as the standard CDG schemes. Theorem 4.2.1 (L2 Stability). With periodic boundary condition, the numerical solutions uh and vh of the sparse grid CDG scheme (4.11)-(4.12) for the equation (4.10) satisfy the following L2 stability condition (cid:107)uh(cid:107)2 L2(ΩN,P ) + (cid:107)vh(cid:107)2 L2(ΩN,D) (cid:46) (cid:107)uh(0, x)(cid:107)2 L2(ΩN,P ) + (cid:107)vh(0, x)(cid:107)2 L2(ΩN,D) . (4.13) For non-periodic boundary condition, the corresponding numerical solutions satisfy (cid:107)uh(cid:107)2 L2(ΩN,P ) + (cid:107)vh(cid:107)2 L2(ΩN,D) + (cid:107)vh(0, x)(cid:107)2 (cid:46) (cid:107)uh(0, x)(cid:107)2 (cid:90) (cid:90) T d(cid:88) + L2(ΩN,P ) |Ai|g2 0 i=1 ∂Ω xin i i ds dt L2(ΩN,D) if τmax (cid:46) hN(cid:107)A(cid:107)1 (4.14) . Proof. For periodic boundary condition, let ϕh = uh in (4.11) and ψh = vh in (4.12), 101 summing the two equalities up, we have 1 2 d dt 1 = τmax (cid:90) + Ω (cid:90) (cid:90) uhA · ∇vh dx − (cid:88) (cid:90) (cid:90) Ω e∈ΓN,D Ω = − 1 τmax − (cid:88) Ω (uh − vh)2dx + (cid:90) uhA · [vh] ds. Ω e∈ΓN,D e ((uh)2 + (vh)2)dx vh uh − uh uh + uhvh − vhvh dx + (cid:90) vhA · ∇uh dx − (cid:88) (cid:90) e vhA · [uh] ds Ω uhA · [vh] ds (cid:90) A · ∇(uhvh)dx − (cid:88) e e∈ΓN,P e∈ΓN,P (cid:90) e vhA · [uh] ds (cid:90) Ω Apply divergence theorem, and by periodicity, we have A · ∇(uhvh)dx − (cid:88) (cid:90) Avh · [uh] ds − (cid:88) (cid:90) e∈ΓN,P e e∈ΓN,D e Auh · [vh] ds = − (cid:90) Ω ∇ · Auhvhdx. By the simple inequality ab ≤ 1 2(a2 + b2), (cid:90) 1 2 d dt Ω (cid:0)(uh)2 + (vh)2(cid:1)dx ≤ − 1 (cid:90) (uh − vh)2dx + (cid:107)∇ · A(cid:107)L∞(Ω) 1 2 τmax Ω (cid:90) Ω ((uh)2 + (vh)2)dx. and the proof for the periodic case is complete by using Gronwall’s inequality. For non-periodic boundary condition, we follow the same lines and plug in the corre- sponding boundary condition, (cid:90) d dt 1 2 = − 1 τmax Ω (cid:90) ((uh)2 + (vh)2)dx (uh − vh)2dx − Ω (cid:90) ∂Ω A · nuhvhds (cid:90) Ω ∇ · Auhvhdx + 102  A · nuhvhds (cid:90) ∂Ω xout i (cid:90) ∂Ω xout i (cid:90) Ω (cid:90) 1 2   (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) − d(cid:88) i=1 d(cid:88) i=1 d(cid:88) i=1 d(cid:88) i=1 d(cid:88) i=1 + + + + ≤ − 1 τmax Ω = − 1 τmax Ω A · ngi(uh + vh)ds + 2 ∂Ω xin i (uh − vh)2dx − (cid:90) Ω ∇ · Auhvhdx = − 1 τmax Ω |A · n|(−uhvh + gi(uh + vh))ds − |A · n|uhvhds ∂Ω xin i (uh − vh)2dx + ≤ − 1 τmax Ω (cid:107)∇ · A(cid:107)L∞(Ω) 1 2 ((uh)2 + (vh)2)dx |A · n|(uh − vh)2ds |A · n|(g2 i + ∂Ω xin i (uh − vh)2dx + |A · n|g2 i ds + ∂Ω xin i (uh − vh)2dx + 1 2 (uh − vh)2)ds + (cid:90) 1 2 (cid:107)∇ · A(cid:107)L∞(Ω) (cid:90) Ω ∪∂Ω ∂Ω xin i xout i (cid:107)∇ · A(cid:107)L∞(Ω) (cid:90) ∂Ω xout i ((uh)2 + (vh)2)dx |A · n|1 2 (uh − vh)2ds ((uh)2 + (vh)2)dx Ω  1 2 (cid:90) |Ai|g2 i ds + xin i ∂Ω ∂Ω xin i ∪∂Ω xout i |Ai|1 2 (uh − vh)2ds  . by noticing A · n|∂Ω < 0 and A · n|∂Ω xin i > 0. xout i Let T i N,D := {T ∈ ΩN,D|T ∩ ∂Ωxi (cid:54)= ∅} denote the cells on dual mesh adjacent to (cid:46) the boundary in the i-th direction. By inverse inequality, we have (cid:107)uh − vh(cid:107)2 h−1 N (cid:107)uh − vh(cid:107)2 N (cid:107)uh − vh(cid:107)2 L2(∂Ωxi ) ≤ h−1 L2(Ω) L2(T i N,D) (cid:90) Ω 1 2 d dt ((uh)2 + (vh)2)dx ≤ 1 2 (cid:107)∇ · A(cid:107)L∞(Ω) Ω . Therefore, if τmax (cid:46) hN(cid:107)A(cid:107)1 (cid:90) d(cid:88) ((uh)2 + (vh)2)dx + , (cid:90) |Ai|g2 i ds, i=1 ∂Ω xin i and the proof for the non-periodic case is complete by using Gronwall’s inequality. 103 Now we are ready to prove L2 error estimate of the sparse grid CDG scheme. Theorem 4.2.2 (L2 error estimate). Let u be the exact solution to (4.10) and uh, vh be the numerical solution to the semidiscrete scheme (4.11) and (4.12) with initial discretiza- tion uh(0, x) = PP u0, vh(0, x) = PDu0 for periodic boundary condition or uh(0, x) = PP u0, vh(0, x) = ˜PDu0 for non-periodic boundary condition. If τmax (cid:46) hN , then for k ≥ 1, u0 ∈ Hp+1(Ω), 1 ≤ q ≤ min{p, k}, N ≥ 1, d ≥ 2, we have for all t ≥ 0 (cid:107)u − uh(cid:107)L2(ΩN,P ) + (cid:107)u − vh(cid:107)L2(ΩN,D) (cid:46) N d2−N q |u|Hq+1(Ω) . (4.15) Proof. For periodic problems, we first introduce the standard notation of bilinear form B(uh, vh; ϕh, ψh) = + − Ω (cid:90) (cid:90) (cid:90) (cid:90) (uh)t ϕh dx − 1 (cid:88) τmax vhA · [ϕh] ds + (cid:90) (cid:88) uhA · ∇ψh dx + e∈ΓP e Ω Ω e∈ΓD e (cid:90) (vh − uh) ϕh dx − (vh)t ψh dx − 1 (cid:90) τmax uhA · [ψh] ds. Ω Ω (cid:90) vhA · ∇ϕh dx (uh − vh) ψh dx Ω By Galerkin orthogonality, we have the error equation B(u − uh, u − vh; ϕh, ψh) = 0, ∀ϕh ∈ ˆVk N,P , ψh ∈ ˆVk N,D. (4.16) We take ϕh = PP u − uh, ψh = PDu − uh, ψe = PDu − u, ϕe = PP u − u, 104 then the error equation (4.16) becomes B(ϕh, ψh; ϕh, ψh) = B(ϕe, ψe; ϕh, ψh). (4.17) From Theorem 4.2.1, we get (cid:90) 1 2 d dt Ω (cid:0)ϕ2 h + ψ2 h (cid:1)dx ≤ B(ϕe, ψe; ϕh, ψh) + (cid:90) Ω (cid:107)∇ · A(cid:107)L∞(Ω) 1 2 (ϕ2 h + ψ2 h)dx. (4.18) We write the bilinear form on the right-hand side as a sum of three terms B(ϕe, ψe; ϕh, ψh) = B1 + B2 + B3, (4.19) where B1 = (cid:90) B2 = − B3 = Ω (cid:90) (ϕe)t ϕh dx − 1 τmax (cid:90) ψeA · ∇ϕh dx − (cid:88) Ω ψeA · [ϕh] ds + (ψe − ϕe) ϕh dx + ϕeA · ∇ψh dx, (cid:88) (cid:90) ϕeA · [ψh] ds. e∈ΓN,P e e∈ΓN,D e (cid:90) Ω (ψe)t ψh dx − 1 τmax (cid:90) Ω (ϕe − ψe) ψh dx, (cid:90) (cid:90) Ω Ω (cid:90) Ω By Cauchy-Schwartz inequality, Lemma 4.1.1 and τmax (cid:46) hN , we have B1 (cid:46) (ϕ2 h + ψ2 h)dx + N 2d2−2N q |u|2Hq+1(Ω) . (4.20) To estimate B2, B3, we use the following inverse inequalities ∀wh ∈ ˆVk N,G, for G = P, D, |wh|H1(ΩN,G) (cid:46) h−1 N (cid:107)wh(cid:107)L2(ΩN,G), (cid:107)wh(cid:107)ΓN,G (cid:46) h − 1 N (cid:107)wh(cid:107)L2(ΩN,G) 2 105 and trace inequality, (cid:107)φ(cid:107)2 L2(∂T ) (cid:46) hN Then we have −1(cid:107)φ(cid:107)2 (cid:90) + hN|φ|H1(T ), ∀φ ∈ H1(T ), T ∈ ΩN,G. L2(T ) B2 (cid:46) Ω (ϕ2 h + ψ2 h)dx + N 2d2−2N q |u|2Hq+1(Ω) (4.21) and (cid:90) Ω B3 (cid:46) (ϕ2 h + ψ2 h)dx + N 2d2−2N q |u|2Hq+1(Ω) . (4.22) Combining (4.20), (4.21), (4.22) with (4.18), we obtain (cid:90) d dt Ω (cid:0)ϕ2 h + ψ2 h (cid:1)dx (cid:46) (cid:90) Ω (ϕ2 h + ψ2 h)dx + N 2d2−2N q |u|2Hq+1(Ω) . Together with the estimates for initial discretization and by Gronwall’s inequality, the proof is complete. For non-periodic problems, the argument is very similar as long as the stability result holds. The proof is omitted for brevity. This theorem proves L2 error of the scheme is O(N d2−N k) or O(|log hN|d hk N ) when the exact solution has enough smoothness in the mixed derivative norms. 4.3 Numerical results In this section, we present several numerical tests to validate the performance of the proposed sparse grid CDG schemes. Unless otherwise stated, we use the third-order TVD-RK temporal discretization [73] and choose the time step ∆t = , with c = 0.1 for k = 1, 2, where d(cid:88) c ci hN i=1 106 ci is the maximum wave propagation speed in xi-direction. To guarantee that the spatial error dominates for k = 3, we take ∆t = O(h 4/3 N ). τmax is taken as 1 2k+1 hN which is always smaller than the maximum time step allowed based on the CFL number in Table 4.1. For periodic problems, we only provide L2 errors on the primal mesh, because the results on the dual mesh are similar. For non-periodic problems, the L2 errors are the L2 average of the errors on the primal and dual meshes. 4.3.1 Scalar case In this subsection, we consider the scalar case, i.e. m = 1. Example 4.3.1 (Linear advection with constant coefficients). We consider  d(cid:88) ut + i=1 u(0, x) = sin uxi = 0, x ∈ [0, 1]d, 2π d(cid:88) i=1  , xi (4.23) with periodic or Dirichlet boundary conditions on the inflow edges corresponding to the given exact solution. The exact solution is a smooth function, 2π  . xi − d t  d(cid:88) i=1 u(t, x) = sin In the simulation, we compute the numerical solutions up to two periods in time, meaning that we let final time T = 1 for d = 2, T = 2/3 for d = 3, and T = 0.5 for d = 4. We first test the scheme with periodic boundary condition. In Table 4.3, we report the 107 L2 errors and orders of accuracy for k = 1, 2, 3 and up to dimension four. As for accuracy, we observe about half order reduction from the optimal (k + 1)-th order for high-dimensional computations (d = 4). The order is slightly better for lower dimensions. The convergence order is similar to the performance of the sparse grid DG scheme in [38]. In Figure 4.3, we plot the time evolution of the error of L2 norm of numerical solutions uh and vh, which is given by (cid:90) Ω (cid:0)(uh(t, x))2 + (vh(t, x))2(cid:1)dx − (cid:90) Ω (cid:0)(uh(0, x))2 + (vh(0, x))2(cid:1)dx for two-dimensional case for t = 0 to t = 100. From Theorem 4.2.1, such errors are propor- tional to the difference between uh and vh. We can clearly see that the higher order accurate scheme performs way better in conservation of L2 norm due to its higher order accuracy. (a) k=1 (b) k=2 (c) k=3 Figure 4.3: Example 4.3.1. The time evolution of the error of L2 norm of numerical solutions uh and vh of the sparse grid CDG method with d = 2. (a) k=1, (b) k=2, (c) k=3. N = 4, 5, 6. Then, we test the scheme with Dirichlet boundary condition prescribed at the inflow edge according to the exact solution. The results are listed in Table 4.4. The accuracy order is similar to the periodic case. Finally, we use this example to compare the performance of the DG, CDG, sparse grid 108 tK=1020406080100(cid:173)1.2(cid:173)1(cid:173)0.8(cid:173)0.6(cid:173)0.4(cid:173)0.20N=4N=5N=6tK=2020406080100(cid:173)0.06(cid:173)0.05(cid:173)0.04(cid:173)0.03(cid:173)0.02(cid:173)0.010N=4N=5N=6tK=3020406080100(cid:173)0.007(cid:173)0.006(cid:173)0.005(cid:173)0.004(cid:173)0.003(cid:173)0.002(cid:173)0.0010N=4N=5N=6 Table 4.3: L2 errors and orders of accuracy for Example 4.3.1 at T = 1 when d = 2, T = 2/3 when d = 3, and T = 0.5 when d = 4. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . L2 error order L2 error order L2 error order N hN k = 1 3 4 5 6 7 3 4 5 6 7 3 4 5 6 7 1/8 1/16 1/32 1/64 1/128 1/8 1/16 1/32 1/64 1/128 1/8 1/16 1/32 1/64 1/128 3.14E-01 6.99E-02 1.34E-02 3.43E-03 9.21E-04 6.77E-01 3.56E-01 1.05E-01 2.54E-02 7.45E-03 7.13E-01 6.48E-01 3.80E-01 1.37E-01 3.81E-02 – 2.17 2.38 1.97 1.90 – 0.93 1.76 2.05 1.77 – 0.14 0.77 1.47 1.85 k = 2 d = 2 1.20E-02 2.23E-03 4.87E-04 5.97E-05 9.33E-06 d = 3 5.27E-02 1.10E-02 1.82E-03 5.22E-04 6.89E-05 d = 4 1.26E-01 3.39E-02 6.91E-03 1.39E-03 3.56E-04 – 2.43 2.20 3.03 2.68 – 2.26 2.60 1.80 2.92 – 1.89 2.29 2.31 1.97 k = 3 5.84E-04 8.50E-05 3.84E-06 3.89E-07 1.80E-08 2.13E-03 2.62E-04 2.85E-05 2.01E-06 2.01E-07 4.41E-03 7.56E-04 9.82E-05 9.44E-06 8.16E-07 – 2.78 4.47 3.30 4.43 – 3.02 3.20 3.83 3.32 – 2.54 2.94 3.38 3.53 Table 4.4: L2 errors and orders of accuracy for Example 4.3.1 with Dirichlet boundary condition on the inflow edges at T = 1 when d = 2 and T = 2/3 when d = 3. N denotes mesh level, hN is the size of the smallest mesh on the primal mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . L2 error order L2 error order L2 error order N hN k = 1 3 4 5 6 7 3 4 5 6 7 1/8 1/16 1/32 1/64 1/128 1/8 1/16 1/32 1/64 1/128 2.66E-01 7.47E-02 1.94E-02 5.44E-03 1.49E-03 6.15E-01 2.86E-01 1.14E-01 3.23E-02 1.03E-02 – 1.83 1.95 1.83 1.87 – 1.10 1.33 1.82 1.65 k = 2 d = 2 1.66E-02 3.33E-03 5.97E-04 8.60E-05 1.35E-05 d = 3 5.34E-02 1.40E-02 2.57E-03 5.82E-04 9.81E-05 – 2.32 2.48 2.80 2.67 – 1.93 2.45 2.14 2.57 k = 3 8.21E-04 8.80E-05 4.79E-06 4.50E-07 2.20E-08 2.67E-03 2.87E-04 3.21E-05 2.60E-06 2.86E-07 – 3.22 4.20 3.41 4.35 – 3.22 3.16 3.63 3.18 109 DG and sparse grid CDG methods. We use the following non-separable initial condition sin 2π d(cid:88)  , x ∈ [0, 1]d, xi u(0, x) = exp (4.24) i=1 where d = 2. When k = 1, 2, 3, Runge-Kutta methods of order ν = 2, 3, 4, respectively, are used for time discretization. We take the time step according to the CFL numbers listed in Table 4.1. We plot the comparison of the methods measuring L2 errors vs. CPU times in Figure 4.4. The computations in this example are implemented by an OpenMP code using computational resources from the Institute for Cyber-Enabled Research in Michigan State University. We can see that the sparse grid CDG method outperforms the CDG method, and the sparse grid DG method outperforms the DG method particularly when the mesh level N is more refined. When the mesh level increases from N to N + 1, the CPU cost for sparse grid method grows with the rate of about 4 to 5, while the factor is about 8 to 10 for full grid calculations, respectively, for this 2D case. This shows the advantage of the sparse grid approach. When comparing the sparse grid CDG method with the sparse grid DG method, it seems that for this example, the sparse grid DG method is more efficient. It will be interesting to compare the results for fully nonlinear problems in higher dimensions, for which the CDG method is more advantageous, and this is currently under investigation. Example 4.3.2 (Solid body rotation). We consider solid-body-rotation problems, which are in the form of (4.1) with periodic boundary conditions and • d = 2, A1(t, x) = −x2 + 1 (cid:16) • d = 3, A1(t, x) = − √ 2 2 2 , A2(t, x) = x1 − 1 2 , √ 2 2 , A2(t, x) = x2 − 1 (cid:17) 2 (cid:16) (cid:17) √ 2 2 + (cid:16) x3 − 1 2 (cid:17) x1 − 1 2 , A3(t, x) = (cid:16) x2 − 1 2 (cid:17) √ 2 2 − . 110 (a) k=1 (b) k=2 (c) k=3 Figure 4.4: L2 errors and associated CPU times of DG, CDG, sparse grid DG and sparse grid CDG methods for Example 4.1 with initial condition (4.24) at T = 1 for d=2. (a) k=1, (b) k=2, (c) k=3. Such benchmark tests are commonly used in the literature to assess performance of transport schemes. Here, the initial profile traverses along circular trajectories centered at (1/2, 1/2) for d = 2 and about the axis {x1 = x3} ∩ {x2 = 1/2} for d = 3 without deformation, and it goes back to the initial state after 2π in time. The initial conditions are set to be the following smooth cosine bells (with C5 smoothness),  bd−1 cos6(cid:0) πr 0, u(0, x) = (cid:1) , 2b r ≤ b, if otherwise, (4.25) where b = 0.23 when d = 2 and b = 0.45 when d = 3, and r = |x − xc| denotes the distance between x and the center of the cosine bell with xc = (0.75, 0.5) for d = 2 and xc = (0.5, 0.55, 0.5) for d = 3. In Table 4.5, we summarize the convergence study of the numerical solutions computed by the sparse CDG method, including the L2 errors and orders of accuracy. For this variable coefficients equation, we observe at least k-th order convergence for all cases. The order is slightly lower than the corresponding ones in Example 4.3.1. 111 +++CPU timeL2 error10010110(cid:173)310(cid:173)210(cid:173)1Full DG Full CDG Sparse DG Sparse CDG ++++CPU timeL2 error10010110(cid:173)510(cid:173)410(cid:173)310(cid:173)2Full DGFull CDGSparse DGSparse CDG++++CPU timeL2 error10110210(cid:173)710(cid:173)610(cid:173)510(cid:173)410(cid:173)3Full DGFull CDGSparse DGSparse CDG+ Table 4.5: L2 errors and orders of accuracy for Example 4.3.2 at T = 2π. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . L2 error order L2 error order L2 error order N hN k = 1 5 6 7 8 5 6 7 8 1/32 1/64 1/128 1/256 1/32 1/64 1/128 1/256 1.53E-02 1.02E-02 4.66E-03 1.42E-03 4.83E-03 1.87E-03 7.46E-04 2.55E-04 – 0.58 1.13 1.71 – 1.37 1.33 1.55 k = 2 d = 2 5.81E-03 1.50E-03 1.46E-04 2.34E-05 d = 3 6.25E-04 1.20E-04 3.39E-05 8.11E-06 – 1.95 3.36 2.64 – 2.38 1.82 2.06 k = 3 1.34E-03 9.64E-05 1.16E-05 1.10E-06 7.35E-05 9.18E-06 1.36E-06 1.94E-07 – 3.80 3.05 3.40 – 3.00 2.75 2.81 Example 4.3.3 (Deformational flow). We consider the two-dimensional deformational flow with velocity field A1(t, x) = sin2(πx1) sin(2πx2)g(t), A2(t, x) = − sin2(πx2) sin(2πx1)g(t), where g(t) = cos(πt/T ) with T = 1.5, with periodic boundary condition. We still adopt the cosine bell (4.25) as the initial condition for this test, but with xc = (0.65, 0.5) and b = 0.35. Note that the deformational test is more challenging than the solid body rotation due to the space and time dependent flow field. In particular, along the direction of the flow, the cosine bell deforms into a crescent shape at t = T /2 , then goes back to its initial state at t = T as the flow reverses. In the simulations, we compute the solution up to t = T . The convergence study is summarized in Table 4.6. Similar orders are observed compared with Example 4.3.2. In Figure 4.5, we plot the contour plots of the numerical solutions on the primal mesh at t = T /2 when the shape of the bell is greatly deformed, and t = T when the solution is recovered into its initial state. It is observed that 112 the sparse CDG scheme with higher degree k can better resolve the highly deformed solution structure. Table 4.6: L2 errors and orders of accuracy for Example 4.3.3 at T = 1.5. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 2. N hN L2 error order L2 error order L2 error order k = 1 k = 2 k = 3 5 6 7 8 1/32 1/64 1/128 1/256 1.73E-02 8.06E-03 3.29E-03 1.08E-03 – 1.10 1.29 1.61 4.37E-03 1.17E-03 2.04E-04 2.78E-05 – 1.90 2.52 2.88 1.14E-03 2.44E-04 2.05E-05 2.75E-06 – 2.22 3.57 2.90 4.3.2 System case In this subsection, we consider system case, which means m > 1 in equation (4.1) or (4.2). Example 4.3.4 (Acoustic wave equation with constant wave speed). We consider ut = ∇ · v, x ∈ [0, 1]2, vt = ∇u, u(0, x) = u0(x), v(0, x) = v0(x). (4.26) with periodic boundary conditions. The initial conditions u0(x) and v0(x) are chosen ac- cording to the following two types of exact solutions: the standing wave u(t, x) 2πt) sin(2πx1) sin(2πx2)  v1(t, x) v2(t, x)  −√ √ 2 sin(2 √ cos(2 √ cos(2 2πt) cos(2πx1) sin(2πx2) 2πt) sin(2πx1) cos(2πx2)  ,   = 113 (a) (b) (c) (d) (e) (f) Figure 4.5: Example 4.3.3. Deformational flow test. The contour plots of the numerical solutions on primal mesh at t = T /2 (a, c, e) and t = T (b, d, f). k = 1 (a, b), k = 2 (c, d), and k = 3 (e, f). N = 7. 114 and the traveling wave  =  √ √ 2πt + 2πx1) cos(2πx2) 2 sin(2 √ sin(2 2πt + 2πx1) cos(2πx2) √ cos(2 2πt + 2πx1) sin(2πx2)  . u(t, x) v1(t, x) v2(t, x) We compute the solution until T = 1. Similar to the scalar case, we present the L2 errors (cid:20) (cid:21)T and orders of accuracy for u(t, x) = u(t, x), v1(t, x), v2(t, x) in Table 4.7. From the table, we still observe at least (k + 1/2)-th order for the solution. Table 4.7: L2 errors and orders of accuracy for Example 4.3.4 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 2. L2 error order L2 error order L2 error order N hN k = 1 k = 2 k = 3 3 4 5 6 7 3 4 5 6 7 1/8 1/16 1/32 1/64 1/128 1/8 1/16 1/32 1/64 1/128 3.56E-01 7.93E-02 1.50E-02 3.72E-03 1.01E-03 3.97E-01 8.58E-02 1.97E-02 5.36E-03 1.50E-03 – 2.17 2.40 2.01 1.88 – 2.21 2.12 1.88 1.84 standing wave 1.05E-02 – 2.51 1.84E-03 2.53 3.18E-04 2.68 4.95E-05 7.60E-06 2.70 traveling wave 1.85E-02 – 3.36E-03 6.07E-04 9.66E-05 1.45E-05 2.46 2.47 2.65 2.74 5.37E-04 4.31E-05 3.39E-06 2.77E-07 2.03E-08 7.75E-04 6.76E-05 5.68E-06 4.44E-07 3.39E-08 – 3.64 3.67 3.61 3.77 – 3.52 3.57 3.68 3.71 Example 4.3.5 (Two-dimensional homogeneous isotropic elastic wave [44]). The 2D elastic wave equation in homogeneous and isotropic medium in velocity-stress formulation without external source, is a linear hyperbolic system of the form ut + A1ux1 + A2ux2 = 0, (4.27) 115   0 0 0 λ + 2µ 0 0 0 0 0 0 0 1 ρ 0 0 0 0 1 ρ λ 0 0 0 0 µ 0 0 A1 = − , A2 = − 0 0 0 µ  0 0 0 0 λ 0 0 0 0 λ + 2µ 0 0 0 1 ρ 1 ρ 0 0 0  , 0 0 0 (cid:113) λ+2µ ρ (cid:20) (cid:21)T where u = σxx, σyy, σxy, v, w , σxx, σyy represents the normal stress and σxy rep- resents the shear stress and v, w are the velocity in x and y directions. where λ and µ are the Lam´e constants and ρ is the mass density of material. Eigenvalues of A1 and A2 are −cp,−cs, 0, cs, cp, which give us the wave speed cp = and cs = (cid:113) µ ρ for P-wave and S-wave respectively. We consider the homogeneous material parameters λ = 2, µ = 1, ρ = 1, then cp = 2, cs = 1. On domain Ω = [0, 1]2, we take the solutions √ 2 2 , √ 2 2 ) and a plane consisting of a plane P-wave traveling along diagonal direction n = ( S-wave traveling in the opposite direction, i.e., u(t, x) = Rsesin(k·x+kcst) + Rpesin(k·x−kcpt), where Rs = [−µ, µ, 0,− √ kn, k = 2 √ 2 2 cs, √ 2 cs]T , Rp = [λ + µ, λ + µ, µ,− 2 √ 2 cp,− 2 √ 2 2 cp]T and k = 2π. Periodic boundary condition is applied and the initial condition is chosen as u(0, x). We compute the solution until T = 1. The L2 errors and orders of accuracy for u(t, x) are shown in Table 4.8. We observe that the convergence order is close to k + 1. Example 4.3.6 (Three-dimensional isotropic elastic wave [31]). We extend the previous 116 Table 4.8: L2 errors and orders of accuracy for Example 4.3.5 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, dimension d = 2. L2 order is calculated with respect to hN . N 4 5 6 7 hN 1/16 1/32 1/64 1/128 L2 error order L2 error order L2 error order k = 1 k = 2 k = 3 1.09E+00 7.47E-01 2.41E-01 7.14E-02 – 0.55 1.63 1.76 2.72E-01 6.48E-02 9.65E-03 1.12E-03 – 2.07 2.75 3.11 5.71E-02 6.19E-03 4.77E-04 2.55E-05 – 3.21 3.70 4.23 example to 3D and obtain the following linear hyperbolic system (cid:20) ut + A1ux1 + A2ux2 + A3ux3 = 0, (cid:21)T (4.28) where u = σxx, σyy, σzz, σxy, σyz, σxz, u, v, w , σ is the stress tensor and u, v, w are the velocities in each spatial direction. 0 0 0 0 0 0 0 λ 0 0 0 0 0 0 0 0 λ + 2µ 0  0 0 0 0 0 0 0 0 0 0 0 0 0 µ , A2 = − 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 ρ 0 0 0 0 1 ρ 0 0 0 0 0 0 0 0 0 1 ρ 0 0  , λ 0 0 0 0 0 0 0 0 µ 0 0 0 0  0 0 0 0 0 0 λ + 2µ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 ρ 0 0 0 0 0 0 0 0 1 ρ 0 0 0 0 0 0 0 1 ρ λ λ 0 0 0 0 0 0 0 0 0 0 µ 0 0 0 0 µ 0 0 0 0 0 0 A1 = −  117  0 0 0 0 0 0 0 0 A3 = − 0 0 0 0 0 0 0 µ 0 0 0 0 0 0 µ 0 0 0 0 0 0 0 0 0 0 1 ρ 1 ρ 0 0 0 0 0 0 0 1 ρ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 λ + 2µ  , λ λ 0 0 0 0 0 0 where λ, µ and ρ take the same values as the previous example. Hence, we have the same values for cp and cs. Eigenvalues of A1, A2 and A3 are −cp,−cs,−cs, 0, 0, 0, cs, cs, cp, which describe the wave speed for P-wave and S-wave (with different polarizations). On domain Ω = [0, 1]3, we take the solutions consisting of a plane S-wave traveling along diagonal direction n = (− 1√ ) and a plane P-wave traveling in the opposite direction, i.e., ,− 1√ ,− 1√ 3 3 3 u(t, x) = Rs sin(k · x − kcst) + Rp sin(k · x + kcpt), where Rs = [−2 3 µ, 2 3 µ, 0, 0, 1 3 µ,−1 3 Rp = [λ + 2 3 µ, λ + 2 3 µ, λ + 2 3 µ,− 1√ 3 2 3 2 3 µ, µ, cs, µ, cs, 0]T , 1√ 3 µ,− 1√ 3 2 3 cp,− 1√ 3 cp,− 1√ 3 cp]T √ and k = kn, k = −2 3π. Similarly, we consider periodic boundary condition and u0(x) = u(0, x) as initial condition. 118 We present the numerical results at T = 1. In Table 4.9, we get at least (k + 1/2)-th order of accuracy for the solution u(t, x). Table 4.9: L2 errors and orders of accuracy for Example 4.3.6 at T = 1. N denotes mesh level, hN is the size of the smallest mesh in each direction, k is the polynomial order, d is the dimension. L2 order is calculated with respect to hN . d = 3. N 4 5 6 7 hN 1/16 1/32 1/64 1/128 L2 error order L2 error order L2 error order k = 1 k = 2 k = 3 2.49E+00 7.70E-01 1.76E-01 4.27E-02 – 1.69 2.13 2.04 4.93E-02 8.17E-03 1.59E-03 2.79E-04 – 2.59 2.36 2.51 8.91E-04 8.66E-05 7.12E-06 5.42E-07 – 3.36 3.60 3.72 119 APPENDIX 120 Detailed discussions on the choice of the T matrix as in (51) or (52) We discuss what parameters result in |b1 ± b2| = 0, under the assumption that α1 has no dependence on h, β1 = ˜β1hp1, β2 = ˜β2hp2, ˜β1, ˜β2 are nonzero constants that do not depend on h. b1 − b2 = (−β1 + k(k − 1) )(1 − β2 2h k(k − 1) 2k(k − 1) k(k − 1) ) + 2α2 1 h = (− ˜β1hp1 + h−1)(1 − 2k(k − 1) ˜β2hp2−1) + k(k − 1)2α2 1h−1, h h 2 2 b1 + b2 = (−β1 + k(k + 1) )(1 − β2 2k(k + 1) k(k + 1) ) + 2α2 1 h = (− ˜β1hp1 + h−1)(1 − 2k(k + 1) ˜β2hp2−1) + k(k + 1)2α2 1h−1. 2h k(k + 1) If b1 − b2 = 0,∀h < h0, then • α1 (cid:54)= 0, then p1 = −1, p2 = 1 and ˜β1, ˜β2 satisfies (− ˜β1 + k(k − 1) 2 )(1 − 2k(k − 1) ˜β2) + k(k − 1)2α2 1 = 0. (29) Similarly, for b1 + b2 = 0,∀h < h0, then • α1 (cid:54)= 0, p1 = −1, p2 = 1 and ˜β1, ˜β2 satisfies (− ˜β1 + k(k + 1) 2 )(1 − 2k(k + 1) ˜β2) + k(k + 1)2α2 1 = 0. (30) 121 Detailed discussions on assumption A2 Parameter choices for |Γ| = |Λ| imply Γ ± Λ = β1 + k2(k2 − 1) = (β1 − k(k ∓ 1) h2 2h β2 + )(1 − 2β2 k(k ± 1) h k(k ± 1) (−2α2 1 − 2β1β2) + ) − k(k ± 1) 2α2 h h −k2 ± k 2h 1 = 0, which indicates • if α1 (cid:54)= 0, then b1 ± b2 can be greatly simplified as follows. – If Γ + Λ = 0, then k is odd, and b1 + b2 = (cid:18) (cid:18) 1 − β2 (cid:18) b1 − b2 = − 2 k + 1 k h 2k(k + 1) (cid:19) β1 − k(k − 1) , k2(k2 − 1) (cid:19) 2h h , Λ = − 1 k + 1 β1 − k2 h + (cid:19) . β2 (cid:18) 122 – If Γ − Λ = 0, then k is even, and h2 (cid:19) (cid:19) , β1 − k(k + 1) 2k(k − 1) 2h 2 b1 + b2 = (cid:18) k − 1 (cid:18) b1 − b2 = − k 1 − β2 h β1 − k2 Λ = − 1 k − 1 h (cid:19) , h k2(k2 − 1) + h2 β2 , k > 1. • If α1 = 0, then k(k ± 1) 2h β1 = , or β2 = h 2k(k ± 1) . (31) More estimates related to Legendre coefficients We provide estimates of the Legendre coefficients, especially their difference in neighboring cells of equal size. If u ∈ W k+2+n,∞(I), then expand ˆuj(ξ) at ξ = −1 in (2.24) by Taylor series, we have for m ≥ k + 1, ∃z ∈ [−1, 1], s.t. (cid:90) 1 uj,m = C d −1 dξk+1 (cid:16) n(cid:88) s=0 (ξ + 1)s s! d dξs ˆuj(−1) (cid:17) d dξn+1 ˆuj(z) (ξ + 1)n+1 (n + 1)! dξm−k−1 (ξ2 − 1)mdξ, (32) θshk+1+s j u(k+1+s)(x ) + O(hk+2+n j j− 1 2 |u| W k+2+n,∞(Ij ) ), where θs are constants independent of u and hj. When hj = hj+1, we use Taylor expansion again, and compute the difference of two uj,m from neighboring cells uj,m − uj+1,m = µshk+1+s j u(k+1+s)(x ) + O(hk+2+n j j− 1 2 |u| W k+2+n,∞(Ij∪Ij+1) ). (33) + = d n(cid:88) s=0 n(cid:88) s=1 n(cid:88) s=1 Then we obtain the estimates |uj,m − uj+1,m + µshk+1+s j u(k+1+s)(x j− 1 2 )| ≤ Chk+2+n|u| W k+2+n,∞(Ij∪Ij+1) , (34) 123 where µs are constants independent of u and hj. Two convolution-like operators To facilitate the analysis in Chapter 2 and 3, we define two operators on a periodic functions u in L2(I): (cid:1)u(x) = (cid:2)λu(x) = N−1(cid:88) l=0 (−1)l−N + 2l N−1(cid:88) 2 1 1 − λN l=0 u(x + L l N ), N is odd, (35a) λlu(x + L l N ), |λ| = 1, (35b) n=−∞ ˆf (n)e2πinx/L, we have where L = b − a is the size of I and N is odd in (35a). Expand u by Fourier series, i.e., u(x) =(cid:80)∞ N−1(cid:88) ∞(cid:88) (−1)l−N + 2l ∞(cid:88) N−1(cid:88) ∞(cid:88) (cid:1)u(x) = n=−∞ n=−∞ 2π L inx ˆf (n)e l=0 l=0 = 2 −2e2πi n (1 + e2πi n N = n=−∞ ˆf (n)ein( 2π L x+2π l N ) −N + 2l 2 (−ei2π n N )l ˆf (n)e 2π L inx, (cid:2)λu(x) = 1 1 − λN N )2 ∞(cid:88) n=−∞ λl N−1(cid:88) ∞(cid:88) l=0 ˆf (n)ein( 2π L x+2π l N ) N−1(cid:88) ˆf (n)e 2π L inx (λei2π n N )l 1 − λN 1 ∞(cid:88) = = n=−∞ ˆf (n) n=−∞ 1 − λe2πi n N l=0 2π L inx. e 124 In addition, we can apply the operator on the same function recursively, we have 1 ··· 1 (1 − λne2πi n N )νn ˆf (n)ei 2π L inx, N )ν1 2π L inx, ˆf (n) (1 − λ1e2πi n (cid:16) −2e2πi n (1 − λe2πi n N N )ν (1 + e2πi n N )2 e (cid:17)ν ˆf (n)e 2π L inx. (cid:2)ν1 λ1 ··· (cid:2)νn λn u(x) = ((cid:2)λ)νu(x) = ((cid:1))νu(x) = ∞(cid:88) ∞(cid:88) ∞(cid:88) n=−∞ n=−∞ n=−∞ Next, we estimate the two operators. Assuming u ∈ W 3,1(I), then the Fourier coefficient ˆf (n) satisfies: (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) ≤ C |u| W k+4,1(I) 1 + |n|3 . (36) Since N is odd, then ωn = e2πi n N (cid:54)= −1,∀n. Hence, (cid:1)u(x) are well defined. We estimate (cid:1)u(x) by splitting it into blocks of size N as ∞(cid:88) l=−∞ (cid:1)u(x) = Sl, where Sl = lN + N−1 2(cid:88) n=lN− N−1 2 −2e2πi n (1 + e2πi n N N )2 ˆf (n)e 2π L inx. Let’s estimate S0 k+1 first. Denote W1(n) = −2e 2 |1+ωn|2 ≤ 2 |1+ei3π/4|2 2−√ = 2 2 . For other n, |W1(n)| ≤ |W1( N−1 (1+e 2πi n N 2πi n N )2 e 2π L inx. For |n| ≤ [ 3N 2 )| = 8 ], |W1(n)| = ≤ CN 2 |1+ω(N−1)/2|2 2 125 from Taylor expansions. N−1 (cid:12)(cid:12)(cid:12) ˆf (n)W1(n) (cid:12)(cid:12)(cid:12) + 2(cid:88) (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) + CN 2 8 ]−1(cid:88) n=[ 3N (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) ˆf (n)W1(n) (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) 2(cid:88) N−1 8 ]+1 n=[ 3N 8 ]+1 ( N 4 + 2)|u|W 3,1(I) |S0| ≤ ≤ ≤ n=−[ 3N 8 ] 2 2 [ 3N − N−1 n=−[ 3N 8 ](cid:88) [ 3N n=−[ 3N (cid:12)(cid:12)(cid:12) ˆf (n)W1(n) (cid:12)(cid:12)(cid:12) + 8 ]−1(cid:88) (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) + CN 2 8 ](cid:88) (cid:12)(cid:12)(cid:12) + CN 2 (cid:12)(cid:12)(cid:12) ˆf (n) 8 ](cid:88) |u|W 3,1(I).  N−1 2(cid:88) 1 + |n|3 + n=−[ 3N 8 ] n=−[ 3N 8 ] − N−1 1 + ( 3N 8 )3 [ 3N 1 2 1 2 1 ( 3 8)3 n=− N−1 2 2 2 − √ 2 2 − √ ≤ C Then, in a similar way, (37) . N (cid:12)(cid:12)(cid:12)1 − λN(cid:12)(cid:12)(cid:12) = |Sl| ≤ C n=lN− N−1  lN + N−1 2(cid:88)  ∞(cid:88) 1 2 1 + |n|3 + n=−∞ 8)3 |u|W 3,1(I). |u|W 3,1(I) ≤ C|u|W 3,1(I). 1 1 + |n|3 + 1 (|l| + 3 ∞(cid:88) l=−∞ 1 (|l| + 3 8)3 Therefore, |(cid:1)u(x)| ≤ C Similar to the estimation for (cid:1)u(x) above, we split (cid:2) into blocks of size N , ∞(cid:88) l=−∞ (cid:2)λu(x) = Sl, where Sl = (l+1)N−1(cid:88) n=lN ˆf (n)W2(n), W2(n) = e 2π L inx 1 − λe2πi n W2(n) is singular when λ is close to any n-th root of unity. Assuming 126 O(hδ(cid:48) ) and |λ − 1| = O(hδ/2) with 0 ≤ δ/2 ≤ 1. We can write λ = e±iθ and assume θ ∈ (0, π) without loss of generality. First, we establish a relation between δ and δ(cid:48). Since |λ| = |λN| = 1, we have δ, δ(cid:48) ≥ 0. Because 1 − λN = (e2πin/N )N − (eiθ)N = (e2πin/N − (cid:12)(cid:12)(cid:12)1 − λN(cid:12)(cid:12)(cid:12) ≤ N (cid:12)(cid:12)(cid:12)ωn − eiθ(cid:12)(cid:12)(cid:12) ,∀n. With the assumption (cid:12)(cid:12)(cid:12)ωn − eiθ(cid:12)(cid:12)(cid:12) ≥ Chδ(cid:48)+1. Particularly, when n = 0, we have |1 − λ| ≥ eiθ)((cid:80)N−1 (cid:12)(cid:12)(cid:12)1 − λN(cid:12)(cid:12)(cid:12) ∼ Chδ(cid:48) l=0 (e2πin/N )N−1−l(eiθ)l), thus , we get Chδ(cid:48)+1, hence δ/2 ≤ δ(cid:48) + 1. In addition, |W2(n)| = |λ − ωn|−1 ≤ Ch−(δ(cid:48)+1). With the assumption that 0 ≤ δ/2 ≤ 1, N . Let n1 = (cid:98)n0/2(cid:99), n2 = 2n0 − n1, n0+1 |u|W 2,1(I). For other n, |w2(n)| ≤ |w2(n1)| ≤ N ≤ θ < 2π n0 1 1+n2 1 |S0| ≤ Ch−δ/2 (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) ≤ C 1 there ∃ n0 ∼ O(hδ/2−1) s.t. 2π then for n1 ≤ n ≤ n2, 2|sin(πn1/N−θ/2)| ≤ Ch−δ/2. Thus, n1−1(cid:88) (cid:12)(cid:12)(cid:12) ˆf (n) (cid:12)(cid:12)(cid:12) N−1(cid:88) h−δ/2 N−1(cid:88) h−δ/2 N−1(cid:88) h−δ/2 N−1(cid:88)  + Ch−(δ(cid:48)+1) 1 + |n|2 + h−(δ(cid:48)+1)(n2 − n1 + 1) 1 + |n|2 + h−(δ(cid:48)+1)hδ/2−1h2−δ |u|W 2,1(I). 1 + |n|2 + h−δ(cid:48)−δ/2 n=n2+1 ≤ C ≤ C ≤ C n=0 1 1 1 + n=0 n=0 n=0 n=n1 n2(cid:88) (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) ˆf (n) |u|W 2,1(I) |u|W 2,1(I) 1 + n2 1 1 Using similar approaches, for l (cid:54)= 0, h−δ/2 (l+1)N−1(cid:88) n=lN |Sl| ≤ C 1 1 + |n|2 + h−δ(cid:48)+δ/2 1 |n1/N + l|2 |u|W 2,1(I). 127 Summing up, we reach the estimation h−δ/2 ∞(cid:88) n=−∞ 1 + |n|2 + h−δ(cid:48)−δ/2 + h−δ(cid:48)+δ/2 (cid:88) 1 l∈N,l(cid:54)=0 |u|W 2,1(I) 1 |l|2 (38) |(cid:2)λu(x)| ≤ C ≤ Ch−δ(cid:48)−δ/2|u|W 2,1(I). Corollary .0.7. When λi, i ≤ n is a complex number with |λi| = 1, independent of h, above estimates yields following results: (cid:12)(cid:12)(cid:12)(cid:2)ν1 λ1 (cid:12)(cid:12)(cid:12) ≤ C|u| ··· (cid:2)νn λn u(x) 1+(cid:80)n i=1 νi,1 , (I) |((cid:1)u(x))ν| ≤ C|u|W 1+2ν,1(I). (39) W Estimates for Mj,m Let’s recall the definition Mj,m = (Aj + Bj)−1(GL− j,m + HL+ j,m), ∀m ∈ Z+,∀j ∈ ZN , where G, H, Aj, Bj, L− j,m, L+ j,m, Γj, Λj are defined in Table 2.1. Aj + Bj = G[L− j,k−1, L− j,k] + H[L+ j,k−1, L+ j,k] = 1 0 1 2  M+ +  α1 −β2 −β1 −α1  1 0 0 1 hj  M−, 0 1 hj 128  [L− j,k−1 ± L+ j,k−1, L− j,k ± L+ j,k] 0 where M± = = 1  0 hj 1 ± (−1)k−1 1 ± (−1)k k(k − 1)(1 ± (−1)k) k(k + 1)(1 ± (−1)k+1)  . Therefore, (Aj + Bj)−1 = M−1− 1 D1  −α1 β1hj − k(k−(−1)k) 2  , β2 − hj 2k(k+(−1)k) α1hj where D1 = (−1)khj 2k(k+(−1)k) ((−1)kΓj + Λj). In what follows, we estimate Mj,m when scale-invariant flux parameters are used in (40), and when α2 1 + β1β2 = 1 4 in (41). • Scale-invariant flux parameters. 129 D1 is bounded by definitions of Γj, Λj and mesh regularity condition. Then ˜β2hh−1 j − 1 2k(k+(−1)k) α1   1 2 + α1 − ˜β1h−1hj  , − ˜β2hh−1 2 − α1 1 j (Aj + Bj)−1G M−1− 1 D1 ˜β1h−1hj − k(k−(−1)k) 2 (Aj + Bj)−1H 1 0 1 0  = 0 1 hj −α1  = 0 1 hj −α1   ˜β1h−1hj − k(k−(−1)k) 2 ˜β2hh−1 j − 1 2k(k+(−1)k) α1  1 2 − α1 ˜β1h−1hj  ˜β2hh−1 j 1 2 + α1 M−1− 1 D1 and   Mj,m =(Aj + Bj)−1G 1 0  0 1 hj + (−1)m(Aj + Bj)−1H  1 0 m(m + 1) 1 0 1 hj    . 1 −m(m + 1) If the mesh is uniform, the three formulas above are independent of mesh size h. For nonuniform mesh, by mesh regularity condition, ∃σ1, σ2, s.t., σ1 ≤ h−1hj ≤ σ2, therefore, (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(Aj + Bj)−1G 1 0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞  0 1 hj ≤ C, (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(Aj + Bj)−1H 1 0  (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞ 0 1 hj ≤ C, (cid:13)(cid:13)Mj,m (cid:13)(cid:13)∞ ≤ C. (40) 130 • α2 1 + β1β2 = 1 4. Λj = 0. D1 = hj Γj 2k(k+(−1)k) and above formulas simplifies to 0 1 hj −α1 − 1 2 + β1hj − k(k−(−1)k) β1hj k(k+(−1)k) − α1k(k − (−1)k) α1 − 1 1 β2h−1 j − 2 + β2h−1 2k(k+(−1)k) j k(k − (−1)k)  , 1 1 0  = 0 2  = 0 1 hj α1 − 1 (Aj + Bj)−1G M−1− 1 2D1 (Aj + Bj)−1H   M−1− 1 2D1 2 − β1hj − k(k−(−1)k) 2 β1hj k(k+(−1)k) + α1k(k − (−1)k) α1 + 1 1 β2h−1 j − 2 − β2h−1 2k(k+(−1)k) j k(k − (−1)k) Thus, we have the following estimation 1 + (cid:18) |β1|, 2 +α1| | 1 h |β2| h2 , | 1 2−α1| , |Γj| h (cid:19)  . max (cid:107)Mj,m(cid:107)∞ ≤  . (41) Proof of Lemma 2.2.2 =(cid:80)k−2 By Definition 2.2.1, P (cid:63) h u|Ij m=0 uj,mLj,m + ´uj,k−1Lj,k−1 + ´uj,kLj,k. We solve the two coefficients ´uj,k−1, ´uj,k on every cell Ij according to definition (2.16). If assumption A0 is satisfied, it has been shown in Lemma 2.2.1 that (2.16) is equivalent 131 to (2.20). Substitute u and ux by (2.23), we obtain the following equation ´uj,k−1  = (Aj + Bj) uj,k−1  + ´uj,k uj,k (Aj + Bj) ∞(cid:88) m=k+1 uj,m(GL− j,m + HL+ j,m), (42) the existence and uniqueness of the system above is ensured by assumption A0, that is, det(Aj + Bj) = 2(−1)kΓj (cid:54)= 0. And (2.26) is proven by multiplying (Aj + Bj)−1 on both sides of above equality. If any of the assumption A1/A2/A3 is satisfied, (2.16) can be written as k(cid:88) G k(cid:88) ´uj,mL− m + H ´uj+1,mL+ m = G uj,mL− m + H uj+1,mL+ m, m=0 m=0 m=0 m=0 ∞(cid:88) ∞(cid:88) where we used (2.23) and G + H = I2 in above equality. Since ´uj,m = uj,m when m ≤ k − 2, ´uj,k−1  + B ´uj+1,k−1  = A ´uj,k ´uj+1,k ∞(cid:88) m=k−1 uj,mGL− m + uj+1,mHL+ m. In order to solve for ´uj,k−1, ´uj,k, we group above coupled equations for all j in a 2N × 2N ´uN−1,k−1 = , (43) linear system as follows, M  η1 θ1 ··· ηN−1 θN−1 ηN θN   ´u1,k−1 ´u1,k ··· ´uN−1,k ´uN,k−1 ´uN,k  132 where  = G ηj θj ∞(cid:88) m=k−1 uj,mL− m + H ∞(cid:88) m=k−1 uj+1,mL+ m, and M = circ(A, B, 02,··· , 02), denoting a 2N × 2N block-circulant matrix with first two rows as (A, B, 02,··· , 02), with 02 as a 2 × 2 zero matrix. We can calculate that det A = det B = −2k h (α2 1 + β1β2 − 1 4 ) (cid:54)= 0. (44) It is clear that the existence and uniqueness of P (cid:63) h is equivalent to det M (cid:54)= 0. By a direct computation, det M = det AN det(I2− QN ), where I2 denotes the 2× 2 identity matrix, and Q = −A−1B = (−1)k+1 Λ c1 + c2 b1 − b2  , b1 + b2 c1 − c2 with c1 = β1 + k2(k2 − 1) h2 β2 − 2k2 h (α2 1 + β1β2 + 1 4 ) := Γ, k h (2α1), c2 = b1 = −β1 − k2(k2 + 1) b2 = −2k3 (α2 h2 β2 + h2 2k h β2 + 2k2 h (α2 1 + β1β2 + 1 4 ), 1 + β1β2 + 1 4 ). The eigenvalues of Q are (−1)(k+1) Λ (Γ + λ1 = (cid:112) Γ2 − Λ2), (Γ −(cid:112) Γ2 − Λ2). (−1)(k+1) Λ λ2 = 133 (45) (46) (47) (48) (49) Since det Q = det B/ det A = 1, we have the relations λ1λ2 = 1 and 1 − b2 b2 2 = Γ2 − Λ2 − c2 2. (50) Below we discuss the existence and uniqueness of P (cid:63) h based on the types of eigenvalues of Q. A1 If |Γ| > |Λ|, then λ1,2 are real and different. Therefore, we can perform eigenvalue decomposition of Q, Q = T DT−1, where and  T = − b1+b2 √ c2+ Γ2−Λ2 1 b1−b2 √ c2+ 1 Γ2−Λ2 √ Γ2−Λ2 √ Γ2−Λ2 where det T = 2 c2+ where T = λ1  , 0 0 λ2 D =  , T−1 = 1 det T  1 √ Γ2−Λ2 − b1−b2 c2+  , √ b1+b2 c2+ Γ2−Λ2 1 (51) , except for the case when (b1 − b2)(b1 + b2) = 0 and c2 < 0,  1 b1−b2 2c2  , T−1 =  − b1+b2 2c2 1 1 − b1−b2 2c2 b1+b2 2c2 1  . (52) 134 In both situations, we have det M = det AN det(I2 − λN 1 0 0 λN 2 ) = det AN det( 1 − λN 1 0 ). 0 1 − λN 2 det M (cid:54)= 0 if and only if (λ1)N (cid:54)= 1 and (λ2)N (cid:54)= 1, which is true since |λ1| (cid:54)= 1 and |λ2| (cid:54)= 1. A2 If |Γ| = |Λ|, then λ1 = λ2 = (−1)k+1 Γ Λ and we have two repeated eigenvalues. Perform Jordan decomposition:c1 + c2 b1 − b2  = T c1 0 T −1, 1 c1 b1 + b2 c1 − c2 and T =  c2  , 2b1 0  , 1 b1 − b2 0 T = if b1 (cid:54)= b2, if b1 = b2. (53) 0 1 We define c1 0  = λ1 0 1 c1 (−1)k+1 Λ λ1  , Q = T J T −1, J = (−1)k+1 Λ 135 then Qj = T J jT −1, J j =  , 1 κj 0 λj 1 λj T −1, 2 −κN 0 2 I2 − QN = T jΓj−1. where κj = (−1)(k+1)j Λj In both situations, det M (cid:54)= 0 if and only if (λ1)N (cid:54)= 1, meaning that we require N to be odd and further, if k is odd, we require Γ = −Λ; if k is even, we require Γ = Λ. In both cases, λ1 = λ2 = −1. A3 If |Γ| < |Λ|, then λ1,2 are complex, |λ1,2| = 1, λ1 = λ2, still Q is diagonalizable, and similar to A1, det M (cid:54)= 0 turns to (λ1)N (cid:54)= 1 and (λ2)N = (λ1)N (cid:54)= 1, i.e. we require  Γ Λ (cid:115)(cid:18) Γ Λ N (cid:19)2 − 1 (cid:54)= 1. (−1)(k+1)N + Summarize above results, we proved the existence and uniqueness for P (cid:63) h when any of the assumptions A0/A1/A2/A3 is satisfied. In order to obtain the exact formula of ´uj,k−1 and ´uj,k, we analyze the inverse of the matrix M. It is known that the inverse of a nonsingular circulant matrix is also circulant, so is a block-circulant matrix. In particular, M−1 = circ(r0, r1,··· , rN−1) ⊗ A−1 136 where ⊗ means Kronecker product for block matrices and rl is a 2 × 2 matrix defined as, rj = Qj(I2 − QN )−1, j = 0,··· , N − 1. (54) Therefore, if any of the assumptions A1/A2/A3 is satisfied, uj+l,mGL− m + uj+l+1,mHL+ m (cid:17) , ∞(cid:88) m=k+1 uj+l,m[L− k−1, L− k ]−1L− m A = l=0 ´uj,k uj+l,k  + ∞(cid:88) uj+l+1,k ´uj,k−1  uj+l,k−1  + B uj+l+1,k−1 rlA−1(cid:16) N−1(cid:88)  +  − Q (cid:16)uj+l,k−1 uj+l+1,k−1 N−1(cid:88) (cid:17)  + uj,k−1 k ]−1L+ − uj+l+1,mQ[L+ (cid:16) N−1(cid:88) ∞(cid:88) uj,k−1  + − uj+N,mrN [L− k−1, L− ∞(cid:88) k−1, L+ uj+l,mV2,m + uj,mr0[L− (cid:17) k ]−1L− (cid:0)uj,mV1,m + N−1(cid:88) uj+l,mrlV2,m uj+l+1,k uj+l,k m=k+1 m=k+1 l=1 = rl l=0 uj,k = = m m uj,k m=k+1 l=0 k−1, L− k ]−1L− m (cid:1), where rN = QN (I2 − QN )−1 = r0 − I2 is used in the third equality. And (2.27) is proven. 137 Proof of Lemma 2.2.3 Denote ´uj,k−1 − uj,k−1 ´uj,k − uj,k  , Uj = When assumption A0 is satisfied (2.28) is a direct result of (2.26) and (41). If any of the assumptions A1/A2/A3 is satisfied, we have ´uj,k−1 − uj,k−1 ´uj,k − uj,k  = Uj = ∞(cid:88) (uj,mV1,m + m=k+1 N−1(cid:88) l=0 uj+l,mrlV2,m). In order to estimate Uj, we first compute rl to get its detailed dependence on l. If A1/A3, Q is diagonalizable, then rl = T Dl(I2 − DN )−1T−1 =  T−1 + 1 0 0 0 T  T−1 0 0 0 1 T λl 2 1 − λN 2 λl 1 1 − λN 1 λl 1 1 − λN 1 Q1 + λl 2 1 − λN 2 (I2 − Q1), = where Q1 = √ 2 1 Γ2 − Λ2 when T is given by (51), and c2 + √ Γ2 − Λ2 b1 − b2  , b1 + b2 √ Γ2 − Λ2 −c2 + 138 (55) (56)  2c2 b1 − b2  , b1 + b2 0 Q1 = 1 2c2 when T is given by (52). If assumption A2 is satisfied, where rl = Ql(I2 − QN )−1 = (−1)l 2 I2 + (−1)l−N + 2l 4Γ Q2, T −1 = 0 1 0 0  c2 b1 + b2 b1 − b2 −c2  . Q2 = T (57) (58) (59) When assumption A1 is satisfied, eigenvalues λ1, λ2 are real. N−1(cid:88) l=0 1,2 | λl 1 − λN 1,2 | = 1 1 − |λ1,2| 1 − |λ1,2|N 1,2| . |1 − λN Without loss of generality, we assume |λ1| < 1 < |λ2|, then N−1(cid:88) N−1(cid:88) l=0 1 | λl 1 − λN | λl 1 − λN 2 1 2 | ≤ 1 1 − |λ1| = |λ2| |λ2| − 1 , | ≤ 1 |λ2| − 1 . And thus l=0 N−1(cid:88) l=0 (cid:107)rl(cid:107)∞ ≤ (cid:107)Q1(cid:107)∞ 1 − |λ1| + (cid:107)I2 − Q1(cid:107)∞ 1 − |λ2| . (60) 139 Then we have (cid:107)Uj(cid:107)∞ ≤ C(1 + N−1(cid:88) l=0 (cid:107)rl(cid:107)∞)|u| ≤ Chk+1|u| W k+1,∞(I) W k+1,∞(I) (cid:107)Q1(cid:107)∞ 1 − |λ1| + (1 + (cid:107)I2 − Q1(cid:107)∞ 1 − |λ2| ), where (2.25) and the fact that V1,m, V2,m,∀m ≥ 0 are constant matrices independent of h are used in above inequalities. When assumption A3 is satisfied, eigenvalues λ1, λ2 are complex, with |λ1,2| = 1 and above estimation does not apply. We perform more detailed computation and use Fourier analysis to bound Uj by utilizing the smoothness and periodicity of u. If u ∈ W k+2+n,∞(I), N−1(cid:88) N−1(cid:88) l=0 rl rl ∞(cid:88) ∞(cid:88) (cid:0)ul+j,m − n(cid:88) n(cid:88) s=0 m=k+1 l=0 m=k+1 s=0 Uj = + = O(hk+2+n|u| ∞(cid:88) n(cid:88) ) + W k+2+n,∞(Ij ) N−1(cid:88) θshk+1+s λl 1 ( 1 − λN 1 m=k+1 s=0 l=0 θshk+1+su(k+1+s)(x θshk+1+su(k+1+s)(x ∞(cid:88) j+l− 1 2 )V2,m + uj,mV1,m (cid:1) j+l− 1 2 ))V2,m ∞(cid:88) m=k+1 uj,mV1,m λl 2 1 − λN 2 (I2 − Q1))u(k+1+s)(x j+l− 1 2 )V2,m m=k+1 Q1 + ∞(cid:88) m=k+1 + + = O(hk+2+n|u| ∞(cid:88) n(cid:88) m=k+1 s=0 ≤ Chk+1(cid:107)u(cid:107) W k+2+n,∞(Ij ) ) + uj,mV1,m θshk+1+s(Q1 (cid:2)λ1 +(I2 − Q1)(cid:2)λ2 j− 1 2 (1 + h−δ(cid:48)−δ/2((cid:107)Q1(cid:107)∞ + (cid:107)I2 − Q1(cid:107)∞)), )u(k+1+s)(x W k+3,∞(I) )V2,m where (32) is used in the first equality and (38) is used in the last inequality. 140 When assumption A2 is satisfied, by similar computation, if u ∈ W k+2+n,∞(I), uj,mV1,m (cid:1)hk+1+su(k+1+s)(x Q2 j+l− 1 2 )V2,m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:1) u(k+1+s)(x )V2,m j− 1 2 ) + u(k+1+s)(x j+2l(cid:48)− 1 2 ) ∞(cid:88) Uj = O(hk+2+n|u| n(cid:88) ∞(cid:88) + 1 2 θs m=k+1 s=0 ≤ Chk+1|u| (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞(cid:88) + m=k+1 s=0 − u(k+1+s)(x + 2Γ l=0 ) + m=k+1 W k+2+n,∞(Ij ) N−1(cid:88) (cid:0)(−1)l + (−1)l−N + 2l (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞(cid:88) n(cid:88) θshk+1+s(cid:0)u(k+1+s)(x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) )(cid:1)V2,m j+N− 3 2 m=k+1 s=0 W k+1,∞(Ij ) n(cid:88) 1 2 j+2l(cid:48)+ 1 2 θshk+1+s Q2 2Γ N−1 2(cid:88) l(cid:48)=0 ≤ Chk+1(cid:107)u(cid:107) W k+4,∞(I) (1 + (cid:107)Q2(cid:107)∞ |Γ| ), where (34) and (39) are used in the last inequality. The estimates of Uj for assumptions A1, A2 and A3 are finished, and (2.29), (2.30) and (2.31) are direct results of the estimation of (cid:107)Uj(cid:107)∞. Proof of Lemma 3.2.3 Proof. Since P (cid:63) h u = P is, † hu when A0, the formula for `uj,k−1, `uj,k is the same as (2.26). That `uj,k−1  = uj,k−1  + `uj,k uj,k ∞(cid:88) m=k+1 uj,mMm. (61) Under assumption A1/A2/A3, above formula is well-defined if and only if Mm is not singular. By the analysis of Mj,m in Appendix, the existence and uniqueness condition is det(A+B) = 2((−1)kΓ + Λ) (cid:54)= 0. Thus, by (40) and (2.25), (3.9) is proven. 141 If any of the assumptions A1/A2/A3 is satisfied, then the difference of two projections can be written as W u|Ij = P (cid:63) h u|Ij − P † hu|Ij = (´uj,k−1 − `uj,k−1)Lj,k−1 + (´uj,k − `uj,k)Lj,k. The properties of P (cid:63) † hu yield the following coupled system A h u and P ´uj,k − `uj,k ´uj,k−1 − `uj,k−1  + B (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x  u   = τj  (u − P j+ 1 2 (u − P = G ux ιj ´uj+1,k − `uj+1,k ´uj+1,k−1 − `uj+1,k−1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)   P † hu † hu)x (P − x − G − H j+ 1 2 † hu)|− j+ 3 2 † hu)x|− x x − (u − P − (u − P x † hu)|− † hu)x|− x j+ 1 2 j+ 1 2  , † hu † hu)x ιj  = τj  P  , (P ∀j ∈ ZN , (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  + x j+ 1 2 j+ 3 2 † hu (3.8b). where the second equality was obtained by the definition of P Gather the relations above for all j results in a large 2N × 2N linear system with block circulant matrix M, defined in (43), as coefficient matrix, then the solution is ´uj,k−1 − `uj,k−1 ´uj,k − `uj,k  = N−1(cid:88) l=0 rlA−1 τl+j  , ιl+j j ∈ ZN , where by periodicity, when l + j > N , τl+j = τl+j−N , ιl+j = ιl+j−N . On uniform mesh, by the definition of Rj,m in (3.12), Rj,m(1) and (Rj,m)x(1) are independent of j, we denote the corresponding values as Rm(1) and (Rm)x(1) and let 142 R− m = [Rm(1), (Rm)x(1)]T . By (3.11), we have x † hu)|− † hu)x|− x j+ 1 2 j+ 1 2 − (u − P − (u − P x † hu)|− j+ 3 2 † hu)x|− x j+ 3 2  = ∞(cid:88) (uj,m − uj+1,m)R− m m=k+1 (u − P  (u − P and´uj,k−1 − `uj,k−1  = ´uj,k − `uj,k N−1(cid:88) (cid:0) ∞(cid:88) rl (ul+j,m − ul+j+1,m)A−1GR− m (cid:1) . = Sj, j ∈ ZN . (62) l=0 m=k+1 Using (33), we can estimate Sj by the same lines as the estimation of Uj in the proof of Lemma 2.2.3 in Appendix, and (3.10) is proven. Proof of Lemma 3.3.1 Proof. By error equation, the symmetry of A(·,·) and the definition of sh, we have 0 = a(e, vh) = a(h, vh) + a(ζh, vh) = shvhdx + (cid:90) (cid:90) I (ζh)tvhdx − iA(vh, ζh), ∀vh ∈ V k h . (63) I Now, we are going to choose three special test functions to extract superconvergence prop- erties (3.16)-(3.18) about ζh. We first prove (3.16). In order to have A(vh, ζh) = (cid:107)(ζh)xx(cid:107)2, (cid:82) we choose a function v1 ∈ V k = αj,k−1Lj,k−1 + αj,kLj,k + (ζh)xx, v1(ζh)xxdx = (cid:107)(ζh)xx(cid:107)2 h , such that ∀j ∈ ZN , v1|Ij = 0 and (cid:93)(v1)x| , ˆv1| = 0. L2(Ij ) j+ 1 2 j+ 1 2 Ij When the assumption A0 holds, the definition of v1 yields the following local system for 143 each pair of αj,k−1 and αj,k, (Aj + Bj) 2 xx xxx αj,k (ζh)− αj,k−1 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1  = −G  (ζh)− 1 αj,k−1  = − (Aj + Bj)−1G 1 − (Aj + Bj)−1H αj,k 0 0 , 2 xx xxx (ζh)+  (ζh)+  (ζh)−  (ζh)+ hj(ζh)− (cid:12)(cid:12)(cid:12)(cid:12)j− 1 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 (cid:12)(cid:12)(cid:12)(cid:12)j− 1 xxx xx xx hj(ζh)+ xxx 2 2 − H   0 1 hj 0 1 hj ∀j ∈ ZN . (64) , thus v1 is nontrivial and uniquely defined under assumption A0. By orthogonality of Leg- endre polynomials, it follows that (cid:90) Ij (cid:90) (cid:107)v1(cid:107)2 L2(Ij ) = |αj,k−1|2 j,k−1dx + |αj,k|2 L2 j,kdx + (cid:107)(ζh)xx(cid:107)2 L2 L2(Ij ) Ij j(cid:107)(ζh)xxx(cid:107)2 + h3 + (cid:107)(ζh)xx(cid:107)2 L2(Ij ) ) L2(∂Ij ) ≤ C(hj(cid:107)(ζh)xx(cid:107)2 ≤ C(cid:107)(ζh)xx(cid:107)2 , L2(Ij ) L2(∂Ij ) where (40), trace inequalities and inverse inequalities are used in above inequality. Let vh = v1, then (63) becomes (cid:90) I (cid:90) I (ζh)tv1dx − i(cid:107)(ζh)xx(cid:107)2. 0 = shv1dx + Hence (cid:107)(ζh)xx(cid:107)2 ≤ (cid:107)sh + (ζh)t(cid:107)·(cid:107)v1(cid:107) ≤ C(cid:107)sh + (ζh)t(cid:107)·(cid:107)(ζh)xx(cid:107). Therefore, (3.16) is proven when assumption A0 is satisfied. Similarly, in order to have A(vh, ζh) = −(cid:80)N j=1 |[ζh]|2 , we define v2 ∈ V k h , such that j+ 1 2 144 = αj,k−1Lj,k−1 + αj,kLj,k,(cid:82) ∀j ∈ ZN , v2|Ij [ζh]| for each pair of αj,k−1 and αj,k, j+ 1 2 . When assumption A0 is satisfied, this definition yields the following local system v2(ζh)xxdx = 0, ˆv2| j+ 1 2 Ij = 0 and (cid:93)(v2)x| j+ 1 2 = (Aj + Bj) αj,k αj,k−1 1  = G  0 0 1 hj  0 [ζh]  0 hj[ζh] 2 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 2 By same algebra as above, we have αj,k−1  = (Aj + Bj)−1G αj,k  0 [ζh] (cid:12)(cid:12)(cid:12)(cid:12)j− 1 2 ∀j ∈ ZN . , + H + (Aj + Bj)−1H 1 0  0 1 hj  0 hj[ζh] (cid:12)(cid:12)(cid:12)(cid:12)j− 1 2 . By (40), it follows directly that (cid:107)v2(cid:107)2 L2(Ij ) ≤ Ch3 j (|[ζh]|2 j+ 1 2 + |[ζh]|2 j− 1 2 ). Plug v2 in (63), we obtain N(cid:88) j=1 |[ζh]|2 j+ 1 2 = i (cid:90) I (cid:90) I shv2dx + i (ζh)tv2dx ≤ (cid:107)sh + (ζh)t(cid:107)(cid:107)v2(cid:107). Therefore, (3.17) is proven when assumption A0 is satisfied. Finally, in order to have A(vh, ζh) =(cid:80)N ∀j ∈ ZN , v3|Ij and (cid:93)(v3)x| = αj,k−1Lj,k−1+αj,kLj,k such that(cid:82) j+ 1 2 j=1 |[(ζh)x]|2 , we choose v3 ∈ V k j+ 1 2 v3(ζh)xxdx = 0, ˆv3| h , such that = [(ζh)x]| j+ 1 2 = 0. Follow the same lines as the estimates for v2, we end up with the j+ 1 2 Ij estimates (cid:107)v3(cid:107)2 L2(Ij ) ≤ Chj(|[(ζh)x]|2 j+ 1 2 + |[(ζh)x]|2 j− 1 2 ). 145 Plug v3 in (63), we obtain (3.18) when assumption A0 is satisfied. The definition of v1 yields the following coupled system Under assumption A1, we need to compute(cid:80)N (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 αj,k−1  (ζh)− αj+1,k−1  = −G  + B xx A (ζh)− xxx αj,k αj+1,k j=1(|αj,k−1|2 + |αj,k|2) to estimate (cid:107)v1(cid:107)2.  (ζh)+ xx (ζh)+ xxx (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 2 j ∈ ZN . (65) , − H 2 Write it in matrix form M α = b, α = [α1,··· , αN ]T , where M is defined in (43) and αj = [αj,k−1, αj,k], b = [b1,··· , bN ]T , bj = −G  (ζh)− (ζh)− xx xxx (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 2 − H  (ζh)+ xx (ζh)+ xxx (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 2 . Left multiply A−1 in (65), we get an equivalent system M(cid:48)α = b(cid:48), M(cid:48) = circ(I2, A−1B, 02,··· , 02), b(cid:48) = [b(cid:48) 1,··· , b(cid:48) N ]T , b(cid:48) j = A−1bj, and (M(cid:48))−1 = circ(r0,··· , rN−1). By Theorem 5.6.4 in [29] and similar to the proof in Lemma 3.1 in [10], M(cid:48) = (F∗ N ⊗ I2)Ω(FN ⊗ I2), where FN is the discrete Fourier transform matrix defined by (FN )ij = 1√ ω = ei 2π N . FN is symmetric and unitary and N ω(i−1)(j−1), Ω = diag(I2 + A−1B, I2 + ωA−1B,··· , I2 + ωN−1A−1B). 146 |Γ| |Λ| > 1 in A1 ensures that the eigenvalues of Q = −A−1B are not 1, thus The assumption I2 + ωjA−1B,∀j, is nonsingular and Ω is invertible. Then |ρ((M(cid:48))−1)| = (cid:107)(M(cid:48))−1(cid:107)2 ≤ (cid:107)F∗ N ⊗ I2(cid:107)2(cid:107)Ω(cid:107)2(cid:107)FN ⊗ I2(cid:107)2 ≤ C. (66) Therefore, N(cid:88) j=1 1 0 0 1 h Since A−1G (cid:107)b(cid:48) j(cid:107)2 2 ≤ C (|αj,k−1|2 + |αj,k|2) = αT α = (b(cid:48))T (M(cid:48))−T (M(cid:48))−1(b(cid:48))T j=1 (cid:107)b(cid:48) j(cid:107)2 2. 2(cid:107)b(cid:48)(cid:107)2 2 ≤ C ≤ (cid:107)(M(cid:48))−1(cid:107)2 N(cid:88)  are constant matrices, we have (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)2 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 (cid:12)(cid:12)(cid:12)(cid:12)j+ 1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)  (ζh)+ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)2 h(ζh)+ xxx xx + 2 2 xx xxx  L2(∂Ij+1) + (cid:107)(ζh)xx(cid:107)2 + h2(cid:107)(ζh)xxx(cid:107)2 + (cid:107)(ζh)xx(cid:107)2 L2(∂Ij ) L2(∂Ij ) L2(∂Ij ) L2(∂Ij+1) ) L2(∂Ij+1) ), 0 1 h 1 0  , A−1H  (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)  (ζh)− h(ζh)− ≤ C((cid:107)(ζh)xx(cid:107)2 + h2(cid:107)(ζh)xxx(cid:107)2 ≤ C((cid:107)(ζh)xx(cid:107)2 where inverse inequality is used to obtain the last inequality. Finally, we obtain the estimate N(cid:88) j=1 (cid:107)v1(cid:107)2 = N(cid:88) j=1 |αj,k−1|2(cid:107)Lj,k−1(cid:107)2 L2(Ij ) + 147 |αj,k|2(cid:107)Lj,k(cid:107)2 L2(Ij ) + (cid:107)(ζh)xx(cid:107)2 ≤ (cid:107)(ζh)xx(cid:107)2 + Ch ≤ (cid:107)(ζh)xx(cid:107)2 + Ch N(cid:88) N(cid:88) j=1 j=1 (|αj,k−1|2 + |αj,k|2) ((cid:107)(ζh)xx(cid:107)2 + (cid:107)(ζh)xx(cid:107)2 L2(∂Ij+1) ) L2(∂Ij ) ≤ (cid:107)(ζh)xx(cid:107)2 + Ch(cid:107)(ζh)xx(cid:107)2 L2(∂IN ) ≤ C(cid:107)(ζh)xx(cid:107)2, where inverse inequality is used to obtain the last inequality. Then the estimates for (3.16) hold true. (3.17) and (3.18) can be proven by the same procedure when assumption A1 is satisfied, and the steps are omitted for brevity. Remark .0.1. When assumption A2 or A3 is satisfied, the eigenvalues of Q are −1 or two complex number with magnitude 1, then a constant bound for ρ((M(cid:48))−1) as in (66) is not possible. Therefore, we cannot obtain similar results for assumption A2 and A3. Proof for Lemma 3.3.2 Proof. Since wq ∈ V k h , we have k(cid:88) m=0 wq|Ij = cq j,mLj,m. Let vh = D−2Lj,m, m ≤ k − 2 in (3.19a), we obtain j,m = −i cq 2m + 1 hj h2 j 4 ∂twq−1D−2Lj,mdx. (cid:90) Ij (67) (68) Since D−2Lj,m ∈ P m+2 c (Ij), by the property u − P (cid:63) h u ⊥ V k−2 h in the L2 inner product 148 sense, we have c1 j,m =  (cid:82) (cid:90) Ij h2 j 4 −i 2m+1 hj h2 j 4 −i 2m + 1 hj ∂t(u − P (cid:63) h u)D−2Lj,mdx = 0, m ≤ k − 4, ∂t((uj,k−1 − ´uj,k−1)Lj,k−1 Ij +(uj,k − ´uj,k)Lj,k)D−2Lj,mdx, m = k − 3, k − 2. (69) By induction using (67), (68), (69), for 0 ≤ m ≤ k − 2 − 2q, cq j,m = 0. Furthermore, the first nonzero coefficient can be written in a simpler form related to uj,k−1 by induction. When q = 1, we compute c1 j,k−3 by (69) and the definition of w0. That is (cid:16)hj (cid:17)2 j,k−3 = −i c1 2(k − 3) + 1 hj 2 = Ch2 j ∂t(uj,k−1 − ´uj,k−1). ∂t(uj,k−1 − ´uj,k−1) D−2Lj,k−3Lj,k−1dx (cid:90) Ij Suppose cq−1 k+1−2q = Ch2q−2 j ∂q−1 t j,k−1−2q = −i cq 2(k − 1 − 2q) + 1 hj 2 (uj,k−1 − ´uj,k−1), then (cid:90) (cid:16) hj (cid:17)2 ∂tcq−1 j,k+1−2q Ij D−2Lj,k−1−2qLj,k+1−2qdx = Chj 2q∂q t (uj,k−1 − ´uj,k−1). The induction is completed and the second formula in (3.23) is proven when r = 0. Next, we begin estimating the coefficient cq j,m, k − 1 − 2q ≤ m ≤ k − 2, the estimates for cq j,m. By Holder’s inequality and (68), we have (cid:12)(cid:12)(cid:12)cq j,m (cid:12)(cid:12)(cid:12) ≤ Ch2− 1 s(cid:107)∂twq−1(cid:107)Ls(Ij ). 149 To estimate the coefficients cq j,k−1, cq j,k, we need to discuss them under different assump- tions. If assumption A0 is satisfied, meaning (3.19b) and (3.19c) can be decoupled and therefore wq is locally defined by (3.19). By (3.20) and following the same algebra of solving the k-th and (k + 1)-th coefficients in (42), cq j,k−1 cq j,k  = − k−2(cid:88) m=0 Mj,mcq j,m. By (40), for all j ∈ ZN , (cid:12)(cid:12)(cid:12)cq (cid:12)(cid:12)(cid:12)2 (cid:12)(cid:12)(cid:12)cq (cid:12)(cid:12)(cid:12)cq + (cid:12)(cid:12)(cid:12) , j,k j,k (cid:12)(cid:12)(cid:12)2 ≤ C (cid:12)(cid:12)(cid:12)) ≤ C k−2(cid:88) m=k−2q−3 (cid:12)(cid:12)(cid:12)cq j,m max k−2q−3≤m≤k−2 (cid:12)(cid:12)(cid:12)2 ≤ Ch3(cid:107)∂twq−1(cid:107)2 (cid:12)(cid:12)(cid:12) ≤ Ch2(cid:107)∂twq−1(cid:107)L∞(Ij ). (cid:12)(cid:12)(cid:12)cq L2(Ij ) j,m . j,k−1 (cid:12)(cid:12)(cid:12)cq max( j,k−1 If any of the assumptions A1/A2/A3 is satisfied, (3.20) defines a coupled system. From the same lines for obtaining (2.27) in Appendix, the solution for cq j,k−1, cq j,k is C. We have the estimate for cq l=0 Under assumption A1 and scale invariant flux assumption, (60) implies(cid:80)N−1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) cq j,m, m = k − 1, k, that is N−1(cid:88) j+l,m| ≤ Ch2(cid:107)∂twq−1(cid:107)L∞(IN ). (cid:107)rl(cid:107)∞) max|cq ≤ C(1 + j,k−1 cq j,k  l=0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞ l=0 (cid:107)rl(cid:107)∞ ≤ 150  = − cq j,k−1 cq j,k = − N−1(cid:88) (cid:16) l=0 k−2(cid:88) k−2(cid:88) m=k−1−2q m=k−1−2q rlA−1(GL− mcq N−1(cid:88) j+l,m + HL+ mcq j+l+1,m) (cid:17) (70) . cq j,mV1,m + cq j+l,mrlV2,m Under assumption A2 or A3,(cid:80)N−1 l=0 (cid:107)rl(cid:107)∞ is unbounded. Thus we use Fourier analysis to bound the coefficients utilizing the smoothness and periodicity by similar steps of estimating Uj in the proof of Lemma 2.2.3. In the rest of the proof, we make use of two operators (cid:2) and (cid:1), which are defined in (35b) and (35a). When assumption A3 is satisfied, Q = −A−1B has two imaginary eigenvalues λ1, λ2 with (I2− Q1), where Q1 is a constant matrix independent |λ1| = |λ2| = 1. rl = Q1 + λl 1 1−λN 1 λl 2 1−λN 2 of h, and defined in (56) and (57). We perform more detailed computation of the coefficients. In (69), plug in (2.27), for m = k − 3, k − 2, when ut ∈ W k+2+n,∞(I), [Lj,k−1, Lj,k]∂t uj,pV1,p + uj+l,prlV2,p (cid:16) ∞(cid:88) N−1(cid:88) p=k+1 l=0 N−1(cid:88) (cid:17) l=0 uj,pF 1 p,m + uj+l,prlF 2 p,m (cid:17) D−2Lj,mdx c1 j,m = i 2m + 1 h = i 2m + 1 2 h2 4 h2 4 (cid:90) Ij (cid:16) ∞(cid:88) (cid:16) n(cid:88) ∞(cid:88) p=k+1 ∂t p=k+1 s=0 = i 2m + 1 N−1(cid:88) 8 (cid:0) λl 1 + 1 − λN 1 ∞(cid:88) Q1 + n(cid:88) l=0 2m + 1 8 p=k+1 s=0 = i µshk+3+su j− 1 2 (x (k+1+s) t (I2 − Q1)(cid:1) n(cid:88) λl 2 2 1 − λN µshk+3+s(cid:0)u (k+1+s) t (x j− 1 2 )F 1 )F 1 p,m + O(hk+n+3|ut| W k+2+n,∞(I) ) (cid:17) µshk+3+su (k+1+s) t (x )F 2 p,m j+l− 1 2 s=0 p,m (cid:1) + O(hk+3+n|ut| W k+2+n,∞(I) ), )u (k+1+s) t (x )F 2 p,m j− 1 2 [Lj,k−1, Lj,k]Vν,pD−2Lj,mdx, ν = 1, 2, are constants independent of h + (Q1 (cid:2)λ1 where F ν p,m = 2 h +(I2 − Q1)(cid:2)λ2 (cid:82) Ij and (32) is used in the third equality. Plug the formula above into (70), by similar computation, we have  = −i c1 j,k−1 c1 j,k k−2(cid:88) ∞(cid:88) n(cid:88) m=k−3 p=k+1 s=0 µshk+1+s(cid:0)u (k+1+s) t 2m + 1 8 h2 )F 1 p,mV1,m (x j− 1 2 151 +(I2 − Q1)(cid:2)λ2 +(I2 − Q1)(cid:2)λ2 + (Q1 (cid:2)λ1 + (Q1 (cid:2)λ1 + O(hk+2+n|ut| W k+2+n,∞(I) ). )u (k+1+s) t (x j− 1 2 )(F 2 p,mV1,m + F 1 p,mV2,m) (k+1+s) )2u t (x )F 2 p,mV2,m j− 1 2 (cid:1) By (39), we have (Q1 (cid:2)λ1 +(I2 − Q1)(cid:2)λ2 (k+1+s) )νu t (x j− 1 2 ) ≤ C|ut| W k+2+s+ν,1(I) ≤ C|u| W k+4+s+ν,1(I) . Therefore, |c1 j,m| ≤ C2hk+3, m = k − 3, k − 2, and |c1 j,m| ≤ C3hk+3, m = k − 1, k. By induction and similar computation, we can obtain the formula for cq j,m. For brevity, we omit the computation and directly show the estimates j,m| ≤ C3qhk+1+2q, |cq k − 1 − 2q ≤ m ≤ k. (−1)l 2 When assumption A2 is satisfied, Q = −A−1B has two repeated eigenvalues. rl = I2 + (−1)l −N +2l 4Γ Q2, where Q2/Γ is a constant matrix. For m = k − 3, k − 2, when j,m by the same procedure as previous case and obtain )F 2 p,m (x j− 1 2 152 ut ∈ W k+2+n,∞(I), we compute c1 n(cid:88) ∞(cid:88) 2m + 1 c1 j,m = i h2 8 µshk+1+s(cid:0)u + 1 2 ((cid:2)−1 + Q2 Γ s=0 p=k+1 (cid:1))u (k+1+s) t (k+1+s) t )F 1 (x j− 1 2 p,m (cid:1) + O(hk+3+n|ut| W k+2+n,∞(I) ). Plug formula above into (70), we have  = −i c1 j,k−1 c1 j,k k−2(cid:88) m=k−3 ∞(cid:88) n(cid:88) µshk+1+s(cid:0)u p=k+1 s=0 2m + 1 8 h2 (k+1+s) t (x j− 1 2 )F 1 p,mV1,m + + 1 2 1 4 ((cid:2)−1 + ((cid:2)−1 + Q2 Γ Q2 Γ By (39), we have (cid:1))u (k+1+s) t (x j− 1 2 )(F 2 p,mV1,m + F 1 (cid:1))2u (k+1+s) t (x )F 2 p,mV2,m j− 1 2 p,mV2,m) (cid:1) + O(hk+2+n|ut| W k+2+n,∞(I) ). ((cid:2)−1 + Q2 Γ (cid:1))νu (k+1+s) t (x j− 1 2 ) ≤ C|ut| W k+2+s+2ν,1(I) ≤ C|u| W k+4+s+2ν,1(I) and |c1 j,m| ≤ C2hk+3, m = k − 3, k − 2, and |c1 j,m| ≤ C4hk+3, m = k − 1, k. By induction and similar computation, we can obtain the formula for cq j,m. For brevity, we omit the computation and directly show the estimates |cq j,m| ≤ C4qhk+1+2q, k − 1 − 2q ≤ m ≤ k. All the analysis above works when we change definition of wq to ∂r t wq (and change t wq−1 accordingly) in (3.19). Summarize the estimates for cq j,m under all (wq−1)t to ∂r+1 three assumptions, for 1 ≤ q ≤ (cid:98) k−1 |∂r j,m| ≤ C2r,qhk+1+2q, t cq (cid:107)∂r 2 (cid:99), we have N(cid:88) t wq(cid:107) ≤ C( j=1 153 k(cid:88) m=k−2q−1 |∂r j,m|2hj) t cq 1 2 ≤ C2r,qhk+1+2q. Then (3.23), (3.24) is proven. And (3.25) is a direct result of above estimate and (3.22). 154 BIBLIOGRAPHY 155 BIBLIOGRAPHY [1] S. Adjerid, K. D. Devine, J. E. Flaherty, and L. Krivodonova. A posteriori error esti- mation for discontinuous Galerkin solutions of hyperbolic problems. Computer Methods in Applied Mechanics and Engineering, 191(11-12):1097–1112, 2002. [2] S. Adjerid and T. C. Massey. Superconvergence of discontinuous Galerkin solutions for a nonlinear scalar hyperbolic problem. Computer Methods in Applied Mechanics and Engineering, 195(25-28):3331–3346, 2006. [3] M. Ainsworth, P. Monk, and W. Muniz. Dispersive and dissipative properties of discon- tinuous Galerkin finite element methods for the second-order wave equation. Journal of Scientific Computing, 27(1):5–40, 2006. [4] B. Alpert. A class of bases in Lˆ2 for the sparse representation of integral operators. SIAM Journal on Mathematical Analysis, 24(1):246–262, 1993. [5] U. M. Ascher, S. J. Ruuth, and R. J. Spiteri. Implicit-explicit Runge-Kutta methods for time-dependent partial differential equations. Applied Numerical Mathematics, 25(2- 3):151–167, 1997. [6] J. Bona, H. Chen, O. Karakashian, and Y. Xing. Conservative, discontinuous Galerkin– methods for the generalized Korteweg–de Vries equation. Mathematics of Computation, 82(283):1401–1432, 2013. [7] J. H. Bramble and A. H. Schatz. Higher order local accuracy by averaging in the finite element method. Mathematics of Computation, 31(137):94–111, 1977. [8] H.-J. Bungartz and M. Griebel. Sparse grids. Acta Numerica, 13:147–269, 2004. [9] W. Cao, D. Li, Y. Yang, and Z. Zhang. Superconvergence of discontinuous Galerkin methods based on upwind-biased fluxes for 1d linear hyperbolic equations. ESAIM: Mathematical Modelling and Numerical Analysis, 51(2):467–486, 2017. [10] W. Cao, H. Liu, and Z. Zhang. Superconvergence of the direct discontinuous Galerkin method for convection-diffusion equations. Numerical Methods for Partial Differential Equations, 33(1):290–317, 2017. [11] W. Cao, C.-W. Shu, Y. Yang, and Z. Zhang. Superconvergence of discontinuous Galerkin method for scalar nonlinear hyperbolic equations. SIAM Journal on Numerical Analysis, 56(2):732–765, 2018. 156 [12] W. Cao, Z. Zhang, and Q. Zou. Superconvergence of discontinuous Galerkin methods for linear hyperbolic equations. SIAM Journal on Numerical Analysis, 52(5):2555–2573, 2014. [13] O. Cessenat and B. Despres. Application of an ultra weak variational formulation of elliptic PDEs to the two-dimensional helmholtz problem. SIAM Journal on Numerical Analysis, 35(1):255–299, 1998. [14] Q. Chang, E. Jia, and W. Sun. Difference schemes for solving the generalized nonlinear Schr¨odinger equation. Journal of Computational Physics, 148(2):397–415, 1999. [15] A. Chen, Y. Cheng, Y. Liu, and M. Zhang. Superconvergence of ultra-weak discon- tinuous Galerkin methods for the linear Schr¨odinger equation in one dimension. arXiv e-prints, page arXiv:1905.08161, May 2019. [16] A. Chen, F. Li, and Y. Cheng. An ultra-weak discontinuous Galerkin method for Schr¨odinger equation in one dimension. Journal of Scientific Computing, 78(2):772– 815, Feb 2019. [17] Y. Cheng, C.-S. Chou, F. Li, and Y. Xing. L2 stable discontinuous Galerkin methods for one-dimensional two-way wave equations. Mathematics of Computation, 86(303):121– 155, 2017. [18] Y. Cheng and C.-W. Shu. A discontinuous Galerkin finite element method for time dependent partial differential equations with higher order derivatives. Mathematics of Computation, 77(262):699–730, 2008. [19] Y. Cheng and C.-W. Shu. Superconvergence of discontinuous Galerkin and local dis- continuous Galerkin schemes for linear hyperbolic and convection-diffusion equations in one space dimension. SIAM Journal on Numerical Analysis, 47(6):4044–4072, 2010. [20] P. G. Ciarlet. The finite element method for elliptic problems. North-Holland Publishing Co., Amsterdam-New York-Oxford, 1978. Studies in Mathematics and its Applications, Vol. 4. [21] B. Cockburn, S. Hou, and C.-W. Shu. The Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws. iv. the multidimensional case. Mathematics of Computation, 54(190):545–581, 1990. [22] B. Cockburn, S.-Y. Lin, and C.-W. Shu. TVB Runge-Kutta local projection discontin- uous Galerkin finite element method for conservation laws iii: one-dimensional systems. Journal of Computational Physics, 84(1):90–113, 1989. 157 [23] B. Cockburn, M. Luskin, C.-W. Shu, and E. S¨uli. Enhanced accuracy by post-processing for finite element methods for hyperbolic equations. Mathematics of Computation, 72(242):577–606, 2003. [24] B. Cockburn and C.-W. Shu. TVB Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws. ii. general framework. Mathematics of Computation, 52(186):411–435, 1989. [25] B. Cockburn and C.-W. Shu. The local discontinuous Galerkin method for time- SIAM Journal on Numerical Analysis, dependent convection-diffusion systems. 35(6):2440–2463, 1998. [26] B. Cockburn and C.-W. Shu. The Runge-Kutta discontinuous Galerkin method for conservation laws v: multidimensional systems. Journal of Computational Physics, 141(2):199–224, 1998. [27] B. Cockburn and C.-W. Shu. Runge-Kutta discontinuous Galerkin methods for convection-dominated problems. Journal of Scientific Computing, 16:173–261, 2001. [28] I. Dag. A quadratic b-spline finite element method for solving nonlinear Schr¨odinger equation. Computer Methods in Applied Mechanics and Engineering, 174(1-2):247–258, 1999. [29] P. J. Davis. Circulant matrices. Wiley, New York, 1979. [30] J. Douglas and T. Dupont. Interior penalty procedures for elliptic and parabolic Galerkin methods. Computing Methods in Applied Sciences, pages 207–216, 1976. [31] M. Dumbser and M. K¨aser. An arbitrary high-order discontinuous Galerkin method for elastic waves on unstructured meshes—II. The three-dimensional isotropic case. Geophysical Journal International, 167(1):319–336, 2006. [32] V. Gradinaru. Fourier transform on sparse grids: code design and the time dependent Schr¨odinger equation. Computing, 80(1):1–22, 2007. [33] M. Griebel. A parallelizable and vectorizable multi-level algorithm on sparse grids. In W. Hackbusch, editor, Parallel Algorithms for Partial Differential Equations, Notes on Numerical Fluid Mechanics, volume 31, pages 241–251, Braunschweig, 1991. Vieweg- Verlag. [34] M. Griebel. Adaptive sparse grid multilevel methods for elliptic PDEs based on finite differences. Computing, 61(2):151–179, 1998. [35] M. Griebel and J. Hamaekers. Sparse grids for the Schr¨odinger equation. ESAIM: Mathematical Modelling and Numerical Analysis, 41(02):215–247, 2007. 158 [36] M. Griebel and G. Zumbusch. Adaptive sparse grids for hyperbolic conservation laws. In Hyperbolic Problems: Theory, Numerics, Applications, pages 411–422. Springer, 1999. [37] M. J. Grote, A. Schneebeli, and D. Sch¨otzau. Discontinuous Galerkin finite element method for the wave equation. SIAM Journal on Numerical Analysis, 44(6):2408–2431, 2006. [38] W. Guo and Y. Cheng. A sparse grid discontinuous galerkin method for high- dimensional transport equations and its application to kinetic simulations. SIAM Jour- nal on Scientific Computing, 38(6):A3381–A3409, 2016. [39] W. Guo and Y. Cheng. An adaptive multiresolution discontinuous Galerkin method for time-dependent transport equations in multidimensions. SIAM Journal on Scientific Computing, 39(6):A2962–A2992, 2017. [40] P. Hemker. Sparse-grid finite-volume multigrid for 3D-problems. Advances in Compu- tational Mathematics, 4(1):83–110, 1995. [41] L. Ji, Y. Xu, and J. K. Ryan. Negative-order norm estimates for nonlinear hyperbolic conservation laws. Journal of Scientific Computing, 54(2):531–548, Feb 2013. [42] O. Karakashian and C. Makridakis. A space-time finite element method for the nonlinear Schr¨odinger equation: the discontinuous Galerkin method. Mathematics of Computa- tion, 67(222):479–499, 1998. [43] O. Karakashian and C. Makridakis. A space-time finite element method for the nonlinear Schr¨odinger equation: the continuous Galerkin method. SIAM Journal on Numerical Analysis, 36(6):1779–1807, 1999. [44] M. K¨aser and M. Dumbser. An arbitrary high-order discontinuous Galerkin method for elastic waves on unstructured meshes—I. The two-dimensional isotropic case with external source terms. Geophysical Journal International, 166(2):855–877, 2006. [45] A. Kurganov and E. Tadmor. New high-resolution central schemes for nonlinear con- servation laws and convection–diffusion equations. Journal of Computational Physics, 160(1):241–282, 2000. [46] R. J. LeVeque. Finite volume methods for hyperbolic problems, volume 31. Cambridge university press, 2002. [47] X. Liang, A. Q. M. Khaliq, and Y. Xing. Fourth order exponential time differ- encing method with local discontinuous Galerkin approximation for coupled nonlin- ear Schr¨odinger equations. Communications in Computational Physics, 17(2):510–541, 2015. 159 [48] Y. Liu. Central schemes and central discontinuous Galerkin methods on overlapping cells. PhD thesis, Georgia Institute of Technology, 2005. [49] Y. Liu. Central schemes on overlapping cells. Journal of Computational Physics, 209(1):82–104, 2005. [50] Y. Liu, C.-W. Shu, E. Tadmor, and M. Zhang. Central discontinuous Galerkin methods on overlapping cells with a nonoscillatory hierarchical reconstruction. SIAM Journal on Numerical Analysis, 45(6):2442–2467, 2007. [51] Y. Liu, C.-W. Shu, E. Tadmor, and M. Zhang. L2 stability analysis of the central dis- continuous Galerkin method and a comparison between the central and regular discon- tinuous Galerkin methods. ESAIM: Mathematical Modelling and Numerical Analysis, 42(4):593–607, 2008. [52] W. Lu, Y. Huang, and H. Liu. Mass preserving discontinuous Galerkin methods for Schr¨odinger equations. Journal of Computational Physics, 282:210–226, 2015. [53] X. Ma and N. Zabaras. An adaptive hierarchical sparse grid collocation algorithm for the solution of stochastic differential equations. Journal of Computational Physics, 228(8):3084–3113, 2009. [54] X. Meng and J. K. Ryan. Discontinuous Galerkin methods for nonlinear scalar hy- perbolic conservation laws: divided difference estimates and accuracy enhancement. Numerische Mathematik, 136(1):27–73, May 2017. [55] M. S. Mock and P. D. Lax. The computation of discontinuous solutions of linear hy- perbolic equations. Communications on Pure and Applied Mathematics, 31(4):423–430, 1978. [56] H. Nessyahu and E. Tadmor. Non-oscillatory central differencing for hyperbolic conser- vation laws. Journal of Computational Physics, 87(2):408–463, 1990. [57] F. Nobile, R. Tempone, and C. Webster. A sparse grid stochastic collocation method for partial differential equations with random input data. SIAM Journal on Numerical Analysis, 46(5):2309–2345, 2008. [58] D. Pathria and J. L. Morris. Pseudo-spectral solution of nonlinear Schr¨odinger equa- tions. Journal of Computational Physics, 87(1):108–125, 1990. [59] W. H. Reed and T. Hill. Triangular mesh methods for the neutron transport equation. Technical report, Los Alamos Scientific Lab., N. Mex.(USA), 1973. [60] M. A. Reyna and F. Li. Operator bounds and time step conditions for the DG and central DG methods. Journal of Scientific Computing, 62(2):532–554, 2015. 160 [61] B. Riviere and M. F. Wheeler. Discontinuous finite element methods for acoustic and elastic wave problems. Contemporary Mathematics, 329:271–282, 2003. [62] B. Rivi`ere, M. F. Wheeler, and V. Girault. Improved energy estimates for interior penalty, constrained and discontinuous Galerkin methods for elliptic problems. part i. Computational Geosciences, 3(3):337–360, 1999. [63] B. Rivi`ere, M. F. Wheeler, and V. Girault. A priori error estimates for finite element methods based on discontinuous approximation spaces for elliptic problems. SIAM Journal on Numerical Analysis, 39(3):902–931, 2001. [64] Z. Rong, J. Shen, and H. Yu. A nodal sparse grid spectral element method for multi- dimensional elliptic partial differential equations. International Journal of Numerical Analysis and Modeling, 14(4-5):762–783, 2017. [65] J. Ryan, C.-W. Shu, and H. Atkins. Extension of a post processing technique for the discontinuous Galerkin method for hyperbolic equations with application to an aeroacoustic problem. SIAM Journal on Scientific Computing, 26(3):821–843, 2005. [66] C. Schwab, E. S¨uli, and R. Todor. Sparse finite element approximation of high- dimensional transport-dominated diffusion problems. ESAIM: Mathematical Modelling and Numerical Analysis, 42(05):777–819, 2008. [67] J. Shen, T. Tang, and L.-L. Wang. Spectral methods: algorithms, analysis and applica- tions, volume 41. Springer Science & Business Media, 2011. [68] J. Shen and L.-L. Wang. Sparse spectral approximations of high-dimensional problems based on hyperbolic cross. SIAM Journal on Numerical Analysis, 48(3):1087–1109, 2010. [69] J. Shen and H. Yu. Efficient spectral sparse grid methods and applications to high- dimensional elliptic problems. SIAM Journal of Scientific Computing, 32(6):3228–3250, 2010. [70] J. Shen and H. Yu. Efficient spectral sparse grid methods and applications to high- dimensional elliptic equations II. unbounded domains. SIAM Journal of Scientific Com- puting, 34(2):A1141–A1164, 2012. [71] Q. Sheng, A. Khaliq, and E. Al-Said. Solving the generalized nonlinear Schr¨odinger Journal of Computational Physics, equation via quartic spline approximation. 166(2):400–417, 2001. [72] C.-W. Shu. Discontinuous Galerkin methods for time-dependent convection dominated problems: Basics, recent developments and comparison with other methods. In A. C. G.R. Barrenechea, F. Brezzi and E. Georgoulis, editors, Building Bridges: Connections 161 and Challenges in Modern Approaches to Numerical Partial Differential Equations, vol- ume 114, pages 369–397. Springer, Switzerland, 2016. [73] C.-W. Shu and S. Osher. Efficient implementation of essentially non-oscillatory shock- capturing schemes. Journal of Computational Physics, 77:439–471, 1988. [74] T. R. Taha and M. I. Ablowitz. Analytical and numerical aspects of certain nonlinear evolution equations. ii. Numerical, nonlinear Schr¨odinger equation. Journal of Compu- tational Physics, 55(2):203–230, 1984. [75] Z. Tao, A. Chen, M. Zhang, and Y. Cheng. Sparse grid central discontinuous galerkin method for linear hyperbolic systems in high dimensions. SIAM Journal on Scientific Computing, 41(3):A1626–A1651, 2019. [76] Z. Wang, Q. Tang, W. Guo, and Y. Cheng. Sparse grid discontinuous Galerkin methods for high-dimensional elliptic equations. Journal of Computational Physics, 314:244— 263, 2016. [77] M. F. Wheeler. An elliptic collocation-finite element method with interior penalties. SIAM Journal on Numerical Analysis, 15(1):152–161, 1978. [78] D. Xiu. Efficient collocational approach for parametric uncertainty analysis. Commu- nications in Computational Physics, 2(2):293–309, 2007. [79] D. Xiu and J. Hesthaven. High-order collocation methods for differential equations with random inputs. SIAM Journal of Scientific Computing, 27(3):1118–1139, 2005. [80] Y. Xu and C.-W. Shu. Local discontinuous Galerkin methods for nonlinear Schr¨odinger equations. Journal of Computational Physics, 205(1):72–97, 2005. [81] Y. Xu and C.-W. Shu. Optimal error estimates of the semidiscrete local discontinuous Galerkin methods for high order wave equations. SIAM Journal on Numerical Analysis, 50(1):79–104, 2012. [82] Y. Yang and C.-W. Shu. Analysis of optimal superconvergence of discontinuous Galerkin method for linear hyperbolic equations. SIAM Journal on Numerical Analysis, 50(6):3110–3133, 2012. [83] C. Zenger. Sparse grids. In W. Hackbusch, editor, Parallel Algorithms for Partial Differential Equations, Notes on Numerical Fluid Mechanics, volume 31, pages 241– 251, Braunschweig, 1991. Vieweg-Verlag. [84] L. Zhou, Y. Xu, Z. Zhang, and W. Cao. Superconvergence of local discontinuous Galerkin method for one-dimensional linear schr¨odinger equations. Journal of Scientific Computing, 73(2-3):1290–1315, 2017. 162