Processing math: 52%
Research article Special Issues

A high-order convergence analysis for semi-Lagrangian scheme of the Burgers' equation

  • In this article, we provide a comprehensive convergence and stability analysis of a semi-Lagrangian scheme for solving nonlinear Burgers' equations with a high-order spatial discretization. The analysis is for the iteration-free semi-Lagrangian scheme comprising the second-order backward finite difference formula (BDF2) for total derivative and the fourth-order central finite difference for diffusion term along the trajectory. The main highlight of the study is to thoroughly analyze the order of convergence of the discrete 2-norm error O(h2+x4+xp+1/h) by managing the relationship between the local truncation errors from each discretization procedure and the interpolation properties with a symmetric high-order discretization of the diffusion term. Furthermore, stability is established by the uniform boundedness of the numerical solution using the discrete Grönwall's Lemma. We provide numerical examples to support the validity of the theoretical convergence and stability analysis for the propounded backward semi-Lagrangian scheme.

    Citation: Philsu Kim, Seongook Heo, Dojin Kim. A high-order convergence analysis for semi-Lagrangian scheme of the Burgers' equation[J]. AIMS Mathematics, 2023, 8(5): 11270-11296. doi: 10.3934/math.2023571

    Related Papers:

    [1] Ning Yang, Yang Liu . Mixed finite element method for a time-fractional generalized Rosenau-RLW-Burgers equation. AIMS Mathematics, 2025, 10(1): 1757-1778. doi: 10.3934/math.2025080
    [2] Shahbaz Ahmad, Faisal Fairag, Adel M. Al-Mahdi, Jamshaid ul Rahman . Preconditioned augmented Lagrangian method for mean curvature image deblurring. AIMS Mathematics, 2022, 7(10): 17989-18009. doi: 10.3934/math.2022991
    [3] Xumei Zhang, Junying Cao . A high order numerical method for solving Caputo nonlinear fractional ordinary differential equations. AIMS Mathematics, 2021, 6(12): 13187-13209. doi: 10.3934/math.2021762
    [4] Weiwen Wan, Rong An . Convergence analysis of Euler and BDF2 grad-div stabilization methods for the time-dependent penetrative convection model. AIMS Mathematics, 2024, 9(1): 453-480. doi: 10.3934/math.2024025
    [5] Saulo Orizaga, Maurice Fabien, Michael Millard . Efficient numerical approaches with accelerated graphics processing unit (GPU) computations for Poisson problems and Cahn-Hilliard equations. AIMS Mathematics, 2024, 9(10): 27471-27496. doi: 10.3934/math.20241334
    [6] Zhengang Zhao, Yunying Zheng, Xianglin Zeng . Finite element approximation of fractional hyperbolic integro-differential equation. AIMS Mathematics, 2022, 7(8): 15348-15369. doi: 10.3934/math.2022841
    [7] Ziqiang Wang, Qin Liu, Junying Cao . A higher-order numerical scheme for system of two-dimensional nonlinear fractional Volterra integral equations with uniform accuracy. AIMS Mathematics, 2023, 8(6): 13096-13122. doi: 10.3934/math.2023661
    [8] Junying Cao, Zhongqing Wang, Ziqiang Wang . Stability and convergence analysis for a uniform temporal high accuracy of the time-fractional diffusion equation with 1D and 2D spatial compact finite difference method. AIMS Mathematics, 2024, 9(6): 14697-14730. doi: 10.3934/math.2024715
    [9] Zongning Zhang, Chunguang Li, Jianqiang Dong . A class of lattice Boltzmann models for the Burgers equation with variable coefficient in space and time. AIMS Mathematics, 2022, 7(3): 4502-4516. doi: 10.3934/math.2022251
    [10] Lanyin Sun, Kunkun Pang . Numerical solution of unsteady elastic equations with C-Bézier basis functions. AIMS Mathematics, 2024, 9(1): 702-722. doi: 10.3934/math.2024036
  • In this article, we provide a comprehensive convergence and stability analysis of a semi-Lagrangian scheme for solving nonlinear Burgers' equations with a high-order spatial discretization. The analysis is for the iteration-free semi-Lagrangian scheme comprising the second-order backward finite difference formula (BDF2) for total derivative and the fourth-order central finite difference for diffusion term along the trajectory. The main highlight of the study is to thoroughly analyze the order of convergence of the discrete 2-norm error O(h2+x4+xp+1/h) by managing the relationship between the local truncation errors from each discretization procedure and the interpolation properties with a symmetric high-order discretization of the diffusion term. Furthermore, stability is established by the uniform boundedness of the numerical solution using the discrete Grönwall's Lemma. We provide numerical examples to support the validity of the theoretical convergence and stability analysis for the propounded backward semi-Lagrangian scheme.



    One of the most popular techniques for solving advection-type equations is the backward semi-Lagrangian method (BSLM), which has the advantage over conventional Eulerian approaches of using a larger temporal step size while avoiding the mesh deformation of pure Lagrangian methods. Due to these attributes, BSLMs have been much studied in a vast range of models since their introduction in the early 1980s [29] such as weather forecasting [31,33], oceanography [29,38], engineering control [16], and various important fluid dynamics applications ([3,4,5,6,15,23,30,32,36,37] and the references therein). In addition, numerous research efforts have been recently made to develop more efficient BSLM algorithms in computational aspects [10,21,24] as well as in mass conservation and multi-dimension [11,12,13].

    Their convergence analyses have been conducted in various problems including a second-order BSLM in linear advection–diffusion equations [19] and nonlinear advection-type equations [27] using the conservation property; second-order finite element BSLMs in the Navier–Stokes equations given a priori estimates [8,9], and second-order finite difference BSLMs in advection–diffusion equations with the Dirichlet boundary condition [22]. Despite these convergence analysis studies, there remained unaddressed nonlinear advection-type problems of finite difference BSLMs with a high-order spatial discretization due to the complexity and difficulties of understanding the structure for the Cauchy problem represented by a characteristic curve.

    The BSLM offers a way of dealing with the nonlinear advection term by splitting the model equation into a Cauchy problem and a total differential equation along characteristic curves. Even though this nonlinearity appears to vanish in the Lagrangian procedure, it is expressed as the velocity of the characteristic curve, which creates the uncertainty of its departure points. This uncertainty makes the convergence analysis more challenging because the truncation error of the departure points of Cauchy problem contains the errors of the solution from the previous temporal steps in the time discretization (see (2.16) and (2.17) for example). Analyzing the error arising from the diffusion term along the characteristic curves is another difficulty in high-order spatial discretizations. The second-order spatial discretization gives a positive-definite matrix for the diffusion term satisfying the Dirichlet boundary condition [22], which allows the Cholesky decomposition to split the errors into consecutive temporal steps. The other issue of the convergence analysis is in the temporal evolution of the truncation errors and bounds of the error from each time step. For the Dirichlet problem with the second-order spatial discretization, the discrete Poincaré inequality allows to evaluate the error at the final step using the spatial finite difference of the error. However, the mentioned techniques are no longer possible to analyze the BSLM with high-order spatial discretizations in the Dirichlet problem, which is the concern of this paper.

    This work is a part of the sequential research project on the convergence analysis of various proposed finite difference BSLMs due to [25,26,27,28] as well as a subsequent result of [22]. The periodic problem is permitted to discretize a diffusion term with a symmetric high-order spatial approximation. From this symmetry, as the second-order analysis, the Cholesky decomposition allows that the errors related the diffusion term can be split into the ones at two consecutive temporal steps, (see Lemma 4). For the analysis of the finite difference BSLM with a high-order spatial discretization, we apply the Cholesky decomposition of the positive definite matrix for a discretization of the Laplacian in the periodic problem. Instead of using the discrete Poincaré inequality which is no longer allowed in the periodic problem, we apply the telescoping summation technique in order to obtain the bound of the error at the final step using the finite difference of error in time. This not only allows a high-order approximation analysis for spatial discretization but also provides the general convergence analysis without boundary restriction for nonlinear advection–diffusion-type equations. More precisely, we consider Burgers' equation with the periodic boundary condition, which is a prototype of nonlinear advection-diffusion equations, as follows:

    {ut+uux=νuxx,a<x<b,0<tT,u(0,x)=u0(x),xI:=[a,b], (1.1)

    where u(t,x) represents the particle velocity, ν stands for the kinematic viscosity, and u0(x) is a given smooth function. To illustrate the BSLM for the model problem (1.1), we assume that the temporal steps are uniformly distributed throughout the domain [0,T]; that is, the grid points are formed by calculating

    tn:=nh,0nN

    with the temporal step size h=T/N. For each temporal step tn, let χ(x,tn+1;t) be the characteristic curve arriving the spatial point x at time tn+1 that satisfies the nonlinear Cauchy problem described as

    {ddtχ(x,tn+1;t)=u(t,χ(x,tn+1;t)),t<tn+1,χ(x,tn+1;tn+1)=x. (1.2)

    Combining (1.1) and (1.2), the total derivative of u(t,χ(x,tn+1;t)) along the characteristic curve χ(x,tn+1;t) provides the following diffusion equation given by

    ddtu(t,χ(x,tn+1;t))=νuxx(t,χ(x,tn+1;t)),0<t<tn+1,xI. (1.3)

    Notice that the characteristic curve χ(x,tn+1;t) satisfying the Cauchy problem (1.2) is used to evolve the solution u of (1.3) along the characteristic curve, while the solution u is used as a velocity for the characteristic curves with a given initial value at the fixed grid points. This self-consistency requirement of the system generates the main aforementioned hindrance to analyzing the constructed method. The nonlinear Cauchy problem can be solved by several approaches, including iteration methods [1,34,36], iteration-free methods [26,27], and so on. The solution of (1.3) is usually obtained by combining three procedures: a backward finite difference method for the total derivative, an interpolation technique for the solution at off-grid departure points, and a solver for the diffusion term, for instance, a finite difference method [14], a finite element method [3,4,14], a spectral finite element method [36], or a discontinuous Galerkin finite element method [35].

    In this paper, we present a concrete convergence analysis of the finite difference BSLM that adopts an iteration-free error correction method (ECM) [26,27] to solve the Cauchy problem (1.2), the second-order backward difference formula (BDF2) for the total derivative, and the fourth-order central finite difference scheme for the diffusion term in the discretization of the Eq (1.3). The substantial contribution of this study is to provide not only the comprehensive analysis of convergence and stability of a high-order spatial discretization BSLM but also the experiments to support the theoretical results. For the prerequisite process, we review the BSLM, which is modified from [26,27], to determine the numerical solution of the model problem while providing analysis of the local truncation errors of departure points obtained by the ECM, the temporal and spatial discretizations from the finite difference methods, BDF2 and the fourth-order central difference formula. Using the relations between these truncation errors and the boundedness properties of the interpolation, we provide the error equation of the proposed scheme at each time step with the help of the symmetric discretization for the diffusion term. Finally, after taking the telescoping summation of the error equations through all the temporal steps, we obtain the error of the numerical solution at the final time (see Theorem 3.6). The resulting order of convergence O(h2+x4+xp+1/h) with respect to the discrete 2-norm supports convergence analyses for linear advection-type equations [19] and finite element BSLMs [8,9], where p represents the degree of the interpolation. The result also validates the claim shown by the numerical results, including the non-monotonic property [17] of O(xp+1/h). We further investigate the stability of the proposed scheme by presenting that the discrete norm of the numerical solution is uniformly bounded without any restriction of the temporal step size h, which means that the proposed BSLM is unconditionally stable (see Theorem 3.7).

    The paper is organized as follows. The BSLM, comprising the ECM and finite difference schemes for the discretization of Eqs (1.2) and (1.3), is reviewed and local truncation errors are analyzed in Section 2. Based on the error equation, the main result of the paper, the convergence and stability analysis of the BSLM for Burgers' equation, are presented in Section 3. In Section 4, numerical experiments are presented to support the theoretical analysis. Finally, Section 5 presents our conclusions and further discussion. Additionally, in Appendix, we review the Lagrange interpolation formula and its properties.

    For the derivation of the BSLM, we review the ECM for solving the Cauchy problem in Subsection 2.1 and the fourth-order spatial discretization for the Helmholtz equation in Subsection 2.3. Especially, Subsection 2.2 is devoted to the estimation of a concrete bound for the truncation errors of its departure points.

    We start this subsection with a concise review of the ECM for solving the nonlinear self-consistent Cauchy problem given by

    {dχj(t)dt=u(t,χj(t)),t[tn1,tn+1),χj(tn+1)=xj, (2.1)

    where χj(t):=χj(xj,tn+1;t) is the characteristic curve that arrives at the grid point xj(1jJ) at a fixed temporal time step tn+1 and u is the solution of the problem (1.1). Here, we assume that spatial grids xj(1jJ) are uniformly spaced in the domain I:=[a,b] as follows:

    xj:=a+jx,x:=baJ. (2.2)

    The ECM then focuses on finding the departure points χj(tnk)=χ(xj,tn+1;tnk), k=0,1, which will be used for the BSLM. For notational simplicity, we denote um(x):=u(tm,x), umj:=um(xj), and um:=[um1,,umJ]T. Furthermore, we assume that approximations Um:=[Um1,,UmJ]T for the solutions um (mn) are already calculated at all spatial grids xj(1jJ) in the domain I.

    Let ϕ(t) be a perturbed characteristic curve given by an Euler polygon y(t) such that

    ϕ(t):=χj(t)y(t),y(t):=xj+(ttn+1)unj,t[tn1,tn+1]. (2.3)

    Then, by the Taylor expansion of χj(t) about t=tn+1 one can see that

    ϕ(t)=(un+1junj)(ttn+1)+12χj(ξ1)(ttn+1)2

    for some ξ1(t,tn+1). Again, the Taylor expansion of un+1j about t=tn implies that

    |ϕ(t)|h|ut(ξ2,xj)||ttn+1|+12|χj(ξ1)|(ttn+1)2Ch2, (2.4)

    where ξ2 belongs to the interval between tn and tn+1 and a constant C depends only on the bounds of u, ut and ux. Also, by (2.1)–(2.4), the derivative ϕ(t) satisfies

    ϕ(t)=unx(y(tn))ϕ(t)+u(t,y(t))unj+T1,ttn+1, (2.5)

    where T1 is bounded by

    |T1|Ch3 (2.6)

    for a constant C depending only on u,ut,ux,uxx, and uxt. We now use the bound (2.6) and the mid-point quadrature rule for integrating (2.5) over [tn1,tn+1] to obtain

    ϕ(tn+1)ϕ(tn1)=2h(unx(y(tn))ϕ(tn)+un(y(tn))unj)+O(h3).

    The property that ϕ(tn)=12(ϕ(tn+1)+ϕ(tn1))+h2ϕ(ξ3) for some ξ3(tn1,tn+1) with ϕ(tn+1)=0 yields

    (1+hunx(y(tn)))ϕ(tn1)=2h(unjun(y(tn)))+O(h3). (2.7)

    Since y(tn) is not a grid point in the usual sense, after truncating the asymptotic term of (2.7) the ECM uses the Lagrange interpolation to evaluate unx(y(tn)) and un(y(tn)). We then define the approximation ϕn1 of ϕ(tn1) by

    ϕn1:=2h(1+h˜DxLUn(yn))1(UnjLUn(yn)),yn:=xjhUnj, (2.8)

    where L and ˜DxL represent the Lagrangian interpolation polynomial of degree p and its piecewise average derivative at the grid points, respectively (see (A.1) for details). Here, the approximation yn of the Euler polygon y(t) at t=tn satisfies

    y(tn)yn=henj, (2.9)

    where enj is the error of un at xj such as

    enj:=unjUnj,en:=[en1,,enj,,enJ]T. (2.10)

    Instead of solving (2.1) directly, the iteration-free ECM uses the approximation ϕn1 of (2.8) and the relation (2.3) to obtain an approximation χn1j of χj(tn1) defined by

    χn1j:=yn1+ϕn1=xj2hUnj+ϕn1. (2.11)

    In addition, to find an approximation χnj for χj(tn) we use the Taylor expansion of χj(t) at tn1, given by

    χj(tn)=14(xj+3χj(tn1)+2hun1(χj(tn1)))+O(h3). (2.12)

    By using χn1j and LUn1(χn1j) and (2.12), we finally define χnj by

    χnj:=14(xj+3χn1j+2hLUn1(χn1j)). (2.13)

    In this subsection, we focus on estimating the truncation errors τnkj of the approximations χnkj for two departure points χj(tnk) defined by

    τnkj:=χj(tnk)χnkj,k=0,1. (2.14)

    From (2.7) and (2.8), the approximation error of the perturbed characteristic curve ϕ(t) given by (2.4) satisfies

    (1+h˜DxLUn(yn))(ϕ(tn1)ϕn1)=h(˜DxLUn(yn)unx(y(tn)))ϕ(tn1)+2h(enj+LUn(yn)un(y(tn)))+O(h3). (2.15)

    Therefore, Eqs (2.3), (2.9), (2.11), and (2.15) make up the truncation error for the departure points

    τn1j:=(1+h˜DxLUn(yn))1ϵτ2henj, (2.16)

    where ϵτ is given by

    ϵτ:=h(˜DxLUn(yn)unx(y(tn)))ϕ(tn1)+2h(enj+LUn(yn)un(y(tn)))+O(h3). (2.17)

    Moreover, Eqs (2.12) and (2.13) imply that the truncation error at t=tn satisfies

    τnj:=34τn1j+h2(un1(χj(tn1))LUn1(χn1j))+O(h3). (2.18)

    To estimate the truncation errors of τnkj, we introduce δx that is a first-order backward spatial finite difference operator given by

    δxvj:=1x(vjvj1),j=1,,J. (2.19)

    Lemma 2.1. Assume that the Euler polygon and its approximation yn and y(tn) belong to the same subinterval Ij at t=tn for each j. Then we have

    |LUn(yn)un(y(tn))|C(jΛj|enj|+h|enj|+xp+1),|˜DxLUn(yn)unx(y(tn))|C(jΛj|δxenj|+h|enj|+xp),

    where C and C depend only on the degree p of the interpolation and the bounds of the first and second partial derivatives of u.

    Proof. Using en in (2.10), we split the quantity |LUn(yn)un(y(tn))| into three terms such as

    I1:=|Len(yn)|,I2:=|Lun(yn)un(yn)|, and I3:=|un(yn)un(y(tn))|.

    The definition (A.1) of interpolation and its associated property (A.3), respectively, provide that

    I1CjΛj|enj| and I2Cxp+1,

    where C is independent of the indices j and j. Also, the mean value theorem and the property (2.9) give I3Ch|enj| for some constant C that depends only on the bounds of the partial derivatives of u, which implies that the first inequality holds. To estimate ˜DxLUn(yn)unx(y(tn)), we assume that yn is in the interior of Ij and split it into three terms. By virtue of Lemma A.1, (2.9), the mean value theorem, and the interpolation error (A.5), we have

    |DxLUn(yn)unx(y(tn))||DxLen(yn)|+|DxLun(yn)unx(yn)|+|unx(yn)unx(y(tn))|C(jΛj|δxenj|+h|enj|+xp),

    where C depends only on the bounds of the partial derivatives of u. This establishes the second inequality. The estimates of the previous lemma give the invertibility of 1+h˜Dx(LUn)(yn) as follows.

    Corollary 2.2. Assume that for n<N, en and δxen are sufficiently small. Then, there exists a temporal step size h<1 such that 1+h˜DxLUn(yn) is invertible and

    |(1+h˜DxLUn(yn))1|<1+ϵ,ϵ1.

    Proof. To show the invertibility of |˜DxLUn(yn)| for a small h<1, it is enough to show that it is uniformly bounded. This follows easily from the second result in Lemma 2.1 by considering

    |˜DxLUn(yn)||unx(y(tn))|+|˜DxLUn(yn)unx(y(tn))|.

    Combining the outcomes of Lemma 2.1 and Corollary 2.2 yields the estimates for the truncation errors of the departure points at the previous two steps:

    Corollary 2.3. Assume that both y(tnk) and ynk, the Euler polygon and its approximation as defined by (2.3), (2.8), and (2.11), belong to the same jth subinterval Ij for k=0,1. Then the truncation error τnkj (k=0,1) of the departure points in (2.14) can be estimated as follows:

    |τnkj|C[h3(jΛj|δxenj|+h|enj|+xp)+h(jΛj(|enj|+|en1j|)+xp+1)],n<N

    for some constant C depending on p and on the bounds of u,ut and ux.

    Proof. Note that the quantity ϵτ defined in (2.17) can be estimated from the bounds of (2.4) and Lemma 2.1. Thus, using the uniform bound of Corollary 2.2 yields the desired order of the estimate for τn1j. By applying the result for τn1j in (2.18) and Lemma 2.1, we can also have the bound of τnj from the estimation

    |un1(χj(tn1))LUn1(χn1j)|C(jΛj|en1j|+xp+1+|τn1j|),

    which can be obtained by splitting into three terms as performed in the proof of Lemma 2.1.

    In this subsection, we review the BSLM [25,26,27] for solving (1.3). To do this, the fourth-order central finite difference method is chosen to discretize the diffusion term in space and the second-order BDF (BDF2) is used to discretize the total derivative using the departure points obtained in Subsection 2.1 with the interpolation L which is reviewed in Appendix.

    To this end, we first evaluate the diffusion equation (1.3) at time t=tn+1 with the setting s=tn+1 and then apply BDF2 to the total derivative, resulting the asymptotic one-dimensional Helmholtz equation

    32hun+1(x)νun+1xx(x)=12h(4un(χ(x,tn+1;tn))un1(χ(x,tn+1;tn1)))+O(h2). (2.20)

    For the discretization of the diffusion term unxx, we employ the finite difference operator δ4x with the fourth-order accuracy defined by

    δ4xunj:=unj2+16unj130unj+16unj+1unj+212x2. (2.21)

    Using the difference operator δ4x and the approximations χnkj(k=0,1) for the departure points χj(tnk):=χ(xj,tn+1;tnk) developed in the previous subsection, the Eq (2.20) can be discretized at each grid point xj, as follows:

    32hun+1jνδ4xun+1j=12h(4un(χj(tn))un1(χj(tn1)))+O(h2+x4)=12h(4un(χnj)un1(χn1j))+Tn+1j,Tn+1j:=12h(4unx(ξ0j)τnjun1x(ξ1j)τn1j)+O(h2+x4) (2.22)

    for some ξkj(k=0,1) between χj(tnk) and χnkj. Using the interpolation polynomials LUnk(χnkj) for unk(χnkj) (k=0,1) and truncating the asymptotic term Tn+1j in (2.22), which is bounded by the result of Corollary 2.3, we introduce a discrete system for the approximation Un+1j of the solution un+1j given by

    32hUn+1jνδ4xUn+1j=12h(4LUn(χnj)LUn1(χn1j)),1jJ. (2.23)

    For convenience, let us define a vector LUnk(χnk) by

    LUnk(χnk):=[LUnk(χnk1),,LUnk(χnkJ)]TRJ,k=0,1

    and a symmetric circulant matrix A for the periodic model problem by

    A:=(ai,j)J×J,ai,j=112x2{30ifi=j,16if|ij|=1 or |ij|=J1,1if|ij|=2 or |ij|=J2,0otherwise. (2.24)

    Using these vectors and the matrix A, the system of Eq (2.23) can be simplified as

    (32hIνA)Un+1=12h(4LUn(χn)LUn1(χn1)),n=1,,N1, (2.25)

    where I denotes the identity matrix of size J×J.

    Remark 2.4. The eigenvalue computations in [20,Theorem 3.1] for A involving the periodic boundary condition can be used as an efficient solver for the system (2.25). Since the matrix A defined by (2.24) is symmetric circulant its eigenvalue decomposition is given by

    A=QΛQT,Λ:=diag(λ1,,λj,,λJ),Q:=(q1,,qj,,qJ),λj=16x2(cos(4(j1)πJ)16cos(2(j1)πJ)+15),qj=1J[1,cos(2πjJ),,cos(2πj(J1)J)]T, (2.26)

    where qj is the eigenvector corresponding to the eigenvalue λj. With the decomposition in (2.26), the system (2.25) can be solved by the following procedure:

    ˜vn+1=12hQT(4LUn(χn)LUn1(χn1)),(32hIνΛ)vn+1=˜vn+1,Un+1=Qvn+1. (2.27)

    Remark 2.4. Note that since cos(2θ)16cos(θ)+150 for 0θ2π, all eigenvalues λj in (2.26) are non-positive real values. Thus, the entries of the diagonal matrix 32hIνΛ in (2.27) are strictly positive. Therefore, the matrix 32hIνΛ is invertible and hence the discretized system (2.27) is uniquely solvable.

    This section mainly aims to present the convergence analysis of the BSLM by manipulating the bounds of the truncation errors along the discrete time evolution. To do this, we first introduce several definitions and hypotheses to be used in the subsequent analysis. Let δt be the first-order backward temporal finite difference operator for the vector vn+1:=[vn+11,,vn+1J]TRJ defined by

    δtvn+1:=1h(vn+1vn). (3.1)

    Furthermore, let f,g be the discrete 2-inner product defined by

    f,g:=x1jJfjgj,f:=[f1,,fJ]T,g:=[g1,,gJ]T

    and let g2 be the corresponding discrete 2-norm for the vector g defined by

    g22=g,g=x1jJg2j. (3.2)

    The convergence analysis for the BSLM presented here is based on an induction technique under mesh restriction (see [8,9,22]); we assume the induction hypothesis for errors at each temporal step and mesh size restriction as follows:

    (A1) For N2, there exists a constant CA1>0 independent of h and x satisfying

    max0nN1en2CA1h2 and max0nN1δxen2CA1h2

    for efficiently small x1.

    (A2) We assume that the ratio between the spatial mesh size and temporal mesh size is satisfied by

    h2x=CA2

    for some constant CA2>0.

    Remark 3.1. Applying the previous hypotheses to the result of Corollary 2.3, we have

    |τnkj|C1h(h3+jΛj(|enj|+|en1j|)+xp+1),k=0,1 (3.3)

    for C1 depending only on p, CA1 and CA2, since

    jΛj|δxenj|p(jΛj|δxenj|2)1/21xpδxen2Ch,C=pCA1CA2.

    We first derive an error equation for the proposed method using (2.22) and (2.23) as follows. From (2.22) and the fact that um(x)Lum(x)=O(xp+1) for a sufficiently smooth solution u, subtracting (2.23) from (2.22) and multiplying the resulting equation by 23 lead to an equation for the error en+1:

    (1hI23νA)en+1=13h(4Len(χn)Len1(χn1))+rn+1, (3.4)

    where

    Lenk(χnk):=[Lenk(χnk1),,Lenk(χnkJ)]T,k=0,1,n1

    and

    rn+1:=[rn+11,,rn+1j,,rn+1J]T,rn+1j:=13h(4unx(ξ0j)τnjun1x(ξ1j)τn1j)+OΓ,
    OΓ:=O(h2+x4+xp+1h). (3.5)

    Note that from Remark 3.1, each component rn+1j can be estimated by

    |rn+1j|C2jΛj(|enj|+|en1j|)+OΓ, (3.6)

    where C2:=max{43C1unx,13C1un1x}. Taking the discrete 2-inner product with δten+1 after some manipulation of (3.4), one can obtain

    δten+1222ν3Aen+1,δten+1=13δten,δten+1+43hLen(χn+1,n)en,δten+113hLen1(χn+1,n1)en1,δten+1+rn+1,δten+1. (3.7)

    The basic idea for estimating en+12 is to use an induction hypothesis with a bound of δxen+12 obtained by estimating each term of the Eq (3.7). We begin with an estimation of the last term rn+1,δten+1 of (3.7) in the following lemma.

    Lemma 3.2. Suppose that all assumptions in Lemma 2.1 are satisfied. Then rn+1,δten+1 can be estimated by

    |rn+1,δten+1|C3k=0,1enk22+16δten+122+O2Γ

    for a constant C3:=6C22(p+1)2, where OΓ is defined by (3.5).

    Proof. Applying the triangle inequality to (3.6) shows that

    |rn+1j|24C22(p+1)jΛj(|enj|2+|en1j|2)+2O2Γ.

    By the periodicity of enkj and the definition of discrete 2-norm, we have

    rn+1224C22(p+1)2(en22+en122)+2O2Γ.

    Using Young's inequality for the inner product rn+1,δten+1 completes the proof.

    Next, for the estimation of Lenk(χnk)enk,δten+1 (k=0,1) in (3.7), let us define ˜χj(t) to be an arbitrary approximation of the characteristic curve χj(t) over [tn1,tn+1] satisfying

    ˜χj(tnk)=χnkj,k=1,0,1,suptnkttn+1|ddt˜χj(t)|M<j, (3.8)

    where χn+1j=xj and χnkj (k=0,1) are the approximations of the departure points χj(tnk) and M is a fixed constant. Then the property (3.8) and the fundamental theorem for a given vector enk imply that

    |Lenk(χnkj)enkj|=|Lenk(˜χj(tnk))Lenk(˜χj(tn+1))|tn+1tnk|ddxLenk(˜χj(t))||d˜χjdt(t)|dtMtn+1tnk|ddxLenk(˜χj(t))|dt. (3.9)

    Thus, if we assume ˜χj(t)Ij for all t[tn1,tn+1], then by applying Hölder's inequality and using the results of Lemma A.1 on the right-hand side of (3.9), we obtain

    tn+1tnk|ddxLenk(˜χj(t))|dtd1(k+1)h(jΛj|δxenkj|2)12, (3.10)

    where d1 is a constant defined in Lemma A.1. Using this bound, we estimate the middle two terms of the right-hand side in (3.7) in the following lemma.

    Lemma 3.3. Suppose that the approximation ˜χj(t) is included in the subinterval Ij for 1jJ and the derivatives d˜χjdt(t) satisfy the conditions of (3.8). Then there exist constants Ck4>0,k=1,2 depending on p and M for vectors δxenk:=[δxenk1,δxenk2,,δxenkJ]T(k=0,1) such that

    |bk3hLenk(χnk)enk,δten+1|32Ck4δxenk22+16δten+122,b0=4,b1=1. (3.11)

    Proof. By the definition of the discrete 2-norm and the periodicity of δxenkj, the inequality (3.10) shows

    Lenk(χnk)enk22pd21M2(k+1)2h2δxenk22,k=0,1.

    Thus, applying Young's inequality and choosing Ck4:=19pd21M2b2k(k+1)2 complete the proof.

    We now estimate the term Aen+1,δten+1 in (3.7). For this, let δ2x be the second-order central difference operator given by

    δ2xenj:=enj12enj+enj+1x2.

    Then the finite difference operator δ4xen+1j defined by (2.21) can be written in terms of the operators δx and δ2x as follows

    δ4xen+1j=112δxEn+1j+1,En+1j:=x(δ2xen+1jδ2xen+1j1)+12δxen+1j. (3.12)

    Using the expression in (3.12), we have the following result.

    Lemma 3.4. For n0, the symmetric circulant matrix A defined in (2.24) satisfies

    Aen+1,δten+112h(δxen22δxen+122)+x224h(δ2xen22δ2xen+122).

    Proof. From discrete summation by parts with the periodicity of en+1, the term Aen+1,δten+1 can be expressed as

    Aen+1,δten+1=x1jJ(δ4xen+1j)(δten+1j)=x121jJ(δxEn+1j+1)(δten+1j)=x121jJEn+1j+1(δxδten+1j+1)+112(En+1J+1δten+1J+1En+11δten+11)=x121jJEn+1j+1(δxδten+1j+1). (3.13)

    Also, from the definition of En+1j and the property δxδten+1=δtδxen+1, the term Aen+1,δten+1 in (3.13) can be split into two terms as follows:

    Aen+1,δten+1:=A1+A2, (3.14)
    A1:=x2121jJ(δ2xen+1j+1δ2xen+1j)(δtδxen+1j+1),A2:=x1jJ(δxen+1j+1)(δtδxen+1j+1).

    Using the periodicity of en+1 and discrete summation by parts again, the term A1 can be estimated as follows:

    A1=x212[1jJ(δ2xen+1j+1)(δtδxen+1j+2δtδxen+1j+1)(δ2xen+1J+1)(δtδxen+1J+1)+(δ2xen+11)(δtδxen+11)]=x312h1jJδ2xen+1j+1(δ2xen+1j+1δ2xenj+1)=x212hδ2xen+122+x212hδ2xen+1,δ2xen. (3.15)

    Similarly, from the definition of the operator δt, the term A2 can be estimated as

    A2=xh1jJδxen+1j+1(δxen+1j+1δxenj+1)=1hδxen+122+1hδxen+1,δxen. (3.16)

    Finally, combining (3.15) and (3.16), the term Aen+1,δten+1 can be bounded as follows:

    hAen+1,δten+1=(δxen+122δxen+1,δxen)x212(δ2xen+122δ2xen+1,δ2xen)12(δxen22δxen+122)+x224(δ2xen22δ2xen+122) (3.17)

    with the help of Young's inequality. This completes the proof.

    Remark 3.5. The negative of A is positive-definite, which allows the Choleksy decomposition to split the errors as shown in the first equation of (3.17). This is the discrete version of a common integration by parts for the Laplacian with the periodic boundary condition.

    Combining the results of Lemmas 3.2–3.4, we can establish our main theorem, namely the following convergence theorem for the proposed scheme.

    Theorem 3.6. Suppose that δte1, δxe1, and δ2xe1 are bounded and the hypotheses in Lemmas 3.2 and 3.3 are satisfied. Then there exists constant C5 such that

    eN2C5(hδte12+νδxe12+νxδ2xe12+h2+x4+xp+1h),

    where C5 depends only on C3,Ck4,CA1,CA2,p,ν,M, and T.

    Proof. We apply Young's inequality on the first term of (3.7), i.e., |δten,δten+1|12δten22+12δten+122 and combine the results of Lemmas 3.2–3.4 into (3.7) to obtain

    13δten+122+ν3h(δxen+122+x212δ2xen+122)16δten22+ν3h(δxen22+x212δ2xen22)+C5k=0,1δxenk22+O2Γ, (3.18)

    where C5:=max{C3,32Ck4,k=0,1}. By adding 16δte122 on both sides after summing from n=0 to N1 in (3.18) under the induction hypothesis (A1) on the term δxem2 for m<N, we have

    16Nn=1δten22+ν3h(δxeN22+x212δ2xeN22)13δte122+ν3h(δxe122+x212δ2xe122)+2C5N1n=1δxen22+NO2Γ, (3.19)

    in which the asymptotic order h2 from the induction hypothesis is incorporated in O2Γ. Since 1hTeN22=1h2NeNe022=1NNn=1δten22Nn=1δten22 from the telescoping summation, we can obtain that multiplying by 6hT

    eN22+2νT(δxeN22+x212δ2xeN22)2T(hδte122+νδxe122+νx212δ2xe122+3TO2Γ), (3.20)

    which yields the desired estimate of eN2 with the generic constant C5 from T and OΓ.

    Remark 3.7. (1) In (3.19), the estimations of δxem2 for m<N from the induction hypothesis are applied to obtain the bound of eN2. Equations (3.19) and (3.20) show that the bound of δxeN2 increases when either T or ν0, and so does eN2, cognate to a result shown in [1,5,22].

    (2) Note that for the Dirichlet problem [22,Lemma 2], it satisfies eN(xj)=0 for all xjI={a,b}, so that eN2CδxeN2 for constant C=12(ba)2, which is the discrete Poincaré inequality. But, the analysis of the model problem with the periodic boundary condition is no longer allowed to use this inequality as mentioned in the introduction. To overcome this structural drawback, we have instead investigated the error δten using the temporal finite difference operator in each time step. By applying the property of the telescoping summation of δten for nN, the error eN at the final step can be estimated without the use of the bound δxeN2. This approach gives more general convergence analysis of the BSLMs for solving advection-type equations without restrictions of boundary conditions compared to the strategy in [22].

    (3) The error expansion presented here is similar to the result for cases of linear advection problems [17] and also nonlinear problems [22] with Dirichlet boundary condition. In particular, the non-monotonic dependence of the temporal step size O(xp+1/h) is a well-known result, and indicates that the superior accuracy of semi-Lagrangian schemes with smaller temporal step size may not always be achieved. As expected, the other two terms in O(h2+x4) are derived from global truncation errors in the ECM strategy equipped with BDF2 for the Cauchy problem and fourth-order central finite difference for the diffusion term. Therefore, the precise convergence error of BSLM depends on the scheme of backward integration for solving the Cauchy problem as well as the interpolation formula for estimating the solution at departure points.

    (4) The above convergence analysis can be naturally extended in higher-dimensional problems: For example, the truncation errors of the characteristic curves on departure points in the two-dimensional problem satisfy consistent results with Lemma 2.1, Corollaries 2.2 and 2.3, (see [22] for details). With the help of induction hypotheses, one can show that the error equation corresponding to (3.4) satisfies the same asymptotic behavior in (3.5). Furthermore, the positive definite matrix, the negative of A for diffusion term (2.24) in a higher-dimensional problem can be expressed as a sum of the tensor product of the identity matrix and a positive definite matrix, so that the result of Lemma 3.4 can be easily extended. The rest of the process to obtain the error at the final step is achieved by the temporal evolution as Theorem 3.6 in a similar way. Due to space limitations, we only provide numerical results, (see Example 4.2 in Section 4).

    In this subsection, we analyze the numerical stability of the proposed semi-Lagrangian scheme by showing the boundedness of the numerical solution for the model problem (1.1) with respect to the discrete 2-norm without any restriction of temporal step size h and spatial step size x. As mentioned in the Introduction, the stability analysis of the Dirichlet Problem can be obtained by using the discrete Poincaré inequality from the boundary condition, which is not permissible for the periodic problem. For the stability of the BSLM with periodic boundary, we instead begin by applying the discrete Grönwall's lemma to achieve the bound of δxUm for each m<N and then obtain the bound of UN by summation of δxUm2 from m=0 to N1.

    Theorem 3.8. Assume that the trajectories of the characteristic curves satisfy that χnkjIj, (k=0,1) and the condition (3.8) for all j=1,,J. If U0 is bounded, so that δxU0 and δ2xU0 are also bounded, then the discrete solution UN of the proposed BSLM is unconditionally stable in the following sense:

    UN2Cs(U02+δxU02+xδ2xU02),N2,

    where Cs depends only on ν, p, M and T and is independent of h and x.

    Proof. We multiply the numerical system (2.25) by 2/3 and add Un/h to obtain that

    δtUn+12ν3AUn+1=13δtUn43h(UnLUn(χn))+13h(Un1LUn1(χn1)). (3.21)

    Applying the inner product with δtUn+1 in (3.21) yields

    δtUn+1222ν3AUn+1,δtUn+1=13δtUn,δtUn+143hUnLUn(χn),δtUn+1+13hUn1LUn1(χn1),δtUn+1. (3.22)

    We apply Un instead of using en as done in Lemma 3.3 and (3.17) respectively, using similar techniques to obtain

    |bk3hLUnk(χn+1,nk)Unk,δtUn+1|32Ck4δxUnk22+16δtUn+122,b0=4,b1=1 (3.23)

    and

    hAUn+1,δtUn+1=(δxUn+122δxUn+1,δxUn)x212(δ2xUn+122δ2xUn+1,δ2xUn)12δxUn2212δxUn+122, (3.24)

    where δxUm22:=δxUm22+x212δ2xUm22 and Ck4(k=0,1) are the constants in Lemma 3.3. Young's inequality applying on the first term of the right-hand side of (3.22) provides

    δtUn,δtUn+112δtUn22+12δtUn+122, (3.25)

    one can rearrange Eq (3.22) using the bounds of (3.23) and (3.24), so that

    12δtUn+122+ν3hδxUn+12216δtUn22+ν3hδxUn22+C4k=0,1δxUnk22, (3.26)

    where C4:=32max{C04,C14}. Summing on both sides of (3.26) from n=1 to N1 yields

    13Nn=1δtUn22+ν3hδxUN22ν3hδxU022+2C4N1n=0δxUn22. (3.27)

    Note that UN2U02+hNn=1δtUn2, so that UN222U022+2Nh2Nn=1δtUn22. Multiplying by 6hT in (3.27) and applying the previous lower bound for Nn=1δtUn22, we obtain

    UN22+2νTδxUN222U022+2νTδxU022+12C4hTN1n=0δxUn22. (3.28)

    Dividing (3.28) by 2νT induces that

    δxUN221νTU022+δxU022+6C4hνN1n=0δxUn22.

    Hence, from the discrete Grönwall's lemma we conclude that

    δxUN22e6C4hN/ν(1νTU022+δxU022),

    which implies

    N1n=0δxUn22C4(1νTU022+δxU022), (3.29)

    where C4 satisfies

    C4:=N1n=0exp(6C4hnν)Nexp(6C4Tν).

    Again, applying (3.29) to (3.28), we have

    UN222U022+2νTδxU022+12C4T2e6C4Tν(1νTU022+δxU022).

    We finally complete the proof by choosing C2s=max{2+12νTC4T2e6C4Tν,2νT+12C4T2e6C4Tν}.

    Remark 3.9. (1) The stability analysis is performed to obtain the generic constant Cs, which is independent of x and h, using neither the induction hypothesis nor mesh restrictions. Therefore, the result says that the numerical algorithm is unconditionally stable, which is often numerically asserted. We will present numerical experiments to support this stability analysis.

    (2) This proof procedure offers a more general approach to achieving the stability of the BSLM than the previous stability presentation in [22], so that the obtained results can be applied to the BSLM in Dirichlet problems.

    This section provides several experiments with two simple problems to support the theoretical convergence result and the numerical stability in Theorems 3.6 and 3.8, respectively. To verify the order of convergence of the proposed scheme, we exhibit rate-varying spatial mesh sizes and temporal step sizes for varying viscosity ν: the temporal convergence order, the spatial convergence order, and the reciprocal term. For the second-order temporal convergence, we perform numerical tests with different sizes of h for a fixed sufficiently small size of x, so that the error is not much affected by the terms x2 and xp+1/h, and calculate the rate of temporal convergence (RTC) via the formula

    log2(eN(h)2/eN(h/2)2).

    Here, eN(h)2 denotes the discrete 2-norm error corresponding to the temporal grid size h at the end time tN. To investigate the order of spatial convergence from the two terms x4 and xp+1/h, we restrict the relation between two step sizes such as h=x2 using the quintic Lagrange interpolation, p=5, in order that

    O(h2+x4+x6h)=O(x4). (4.1)

    Reducing spatial and temporal step sizes by half under the restricted relation, we compute the rate of spatial convergence (RSC) via the formula

    log2(eN(x)2/eN(x/2)2),

    where eN(x) represents the error corresponding to the spatial grid size x at the end time tN. Finally, to investigate the non-monotonic property of the term xp+1/h, we apply three different orders p=1,2,3 of Lagrangian interpolation. For each interpolation order, we demonstrate the error and its ratio by decreasing the temporal step sizes by half for a fixed spatial step size to determine the behavior of the fractional term involving temporal step size h based on the analytical results already obtained.

    Example 4.1. As the first experiment, we consider the Burgers' equation of form

    ut+uux=νuxx,x(1,1),t(0,1]

    with the initial condition given by

    u0(x)=sin(πx).

    With the help of a Cole–Hopf transform, the periodic analytic solution [39] is given by

    u(x,t)=sin(π(x4νts))exp(cos(π(x4νts))2νπ)exp(s2)dsexp(cos(π(x4νts))2νπ)exp(s2)ds.

    We have numerically integrated to estimate the analytic solution using Gauss–Hermite quadrature with 200 nodes for the integral form. For the first test, we experiment with the errors and the ratio of the changes while varying temporal step sizes h = 1/2^k, k = 6, 7, \cdots, 10 with a fixed small spatial step size \triangle x = 1/2^{11} \leq h in the different viscosity coefficients \nu = 0.01, 0.1, 1.0 . It is easily noted that h^2 is the dominant term of the error, since h^2 > \triangle x^4/h > \triangle x^4 for \triangle x < h \ll 1 , so the error \| \mathbf{e}^N\|_2 is dominated by \mathcal{O}(h^2) . Thus, we can check the temporal order of convergence for the method by halving temporal step sizes under the condition \triangle x < h \ll 1 . The numerical results for the errors and their ratios are displayed for the different viscosities in Table 1. The figures for RTC in the table guarantee that the method has the required second-order temporal convergence rate.

    Table 1.  Errors and RTC with cubic Lagrangian interpolation and \triangle x = {1}/{2^{11}} .
    h \nu=0.01 \nu=0.1 \nu=1.0
    \| \mathbf{e}^N\|_2 RTC \| \mathbf{e}^N\|_2 RTC \| \mathbf{e}^N\|_2 RTC
    1/2^6 2.52\times 10^{-3} - 5.55\times 10^{-5} - 4.32\times 10^{-6} -
    1/2^7 7.58\times 10^{-4} 1.73 1.36 \times 10^{-5} 2.03 1.04\times 10^{-6} 2.05
    1/2^8 2.06\times 10^{-4} 1.88 3.35\times 10^{-6} 2.02 2.56 \times 10^{-7} 2.05
    1/2^9 5.38 \times 10^{-5} 1.94 8.32\times 10^{-7} 2.01 6.37\times 10^{-8} 2.01
    1/2^{10} 1.37\times 10^{-5} 1.97 2.07\times 10^{-7} 2.00 1.71\times 10^{-8} 1.90

     | Show Table
    DownLoad: CSV

    To check the order of the spatial convergence we perform a test problem with three different viscosity coefficients \nu = 0.01, 0.5, 1.0 , by varying step sizes under the relation h = \triangle x^2 such as (h, \triangle x) = (2^{-2k}, 2^{-k}) , k = 1, 2, \cdots, 9 . The numerical results are displayed in Table 2. The figures show that the method has an almost fourth-order spatial convergence since each term in the error expression has fourth-order convergence under the relation h = \triangle x^2 as seen in (4.1). This supports the theoretical order of convergence analyzed in Theorem 3.6.

    Table 2.  Errors with quintic Lagrangian interpolation for (h, \triangle x) = (4^{-k}, 2^{-k}) under h = \triangle x^2 .
    \nu=0.01 \nu=0.5 \nu=1.0
    (4^{-k}, 2^{-k}) \| \mathbf{e}^N\|_{2} Rate (4^{-k}, 2^{-k}) \| \mathbf{e}^N\|_{2} Rate (4^{-k}, 2^{-k}) \| \mathbf{e}^N\|_{2} Rate
    (4^{-5}, 2^{-5}) 8.05\times 10^{-3} - (4^{-1}, 2^{-1}) 2.28\times 10^{-2} - (4^{-2}, 2^{-2}) 6.29\times 10^{-5} -
    (4^{-6}, 2^{-6}) 7.05\times 10^{-4} 3.51 (4^{-2}, 2^{-2}) 1.15 \times 10^{-3} 4.31 (4^{-3}, 2^{-3}) 4.24\times 10^{-6} 3.89
    (4^{-7}, 2^{-7}) 3.72 \times 10^{-5} 4.24 (4^{-3}, 2^{-3}) 7.29\times 10^{-5} 3.98 (4^{-4}, 2^{-4}) 2.54\times 10^{-7} 4.06
    (4^{-8}, 2^{-8}) 2.04 \times 10^{-6} 4.19 (4^{-4}, 2^{-4}) 5.30 \times 10^{-6} 3.78 (4^{-5}, 2^{-5}) 1.60\times 10^{-8} 3.99
    (4^{-9}, 2^{-9}) 1.26\times 10^{-7} 4.01 (4^{-5}, 2^{-5}) 4.32\times 10^{-7} 3.62 (4^{-6}, 2^{-6}) 1.05\times 10^{-9} 3.94

     | Show Table
    DownLoad: CSV

    Finally, in order to check the reciprocal term of the temporal step size \mathcal{O}({\triangle x^{p+1}}/{h}) , we apply Lagrange interpolations with different orders p = 1, 2, 3 . It can be seen that the term {\triangle x^{p+1}}/{h} dominates the error as soon as h < \triangle x^{(p+1)/3} \ll 1 , since

    \frac{\triangle x^{p+1}}{h} = \frac{\triangle x^{\frac{p+1}{3}}}{h} \triangle x^{\frac{2(p+1)}{3}} \geq \triangle x^{\frac{2(p+1)}{3}} > \begin{cases} h^2 \\ \triangle x^4 \end{cases} \text{ for } p = 1,2,3.

    We display the behavior of the error for each different interpolation order p = 1, 2, 3 in Figure 1 while varying h = 1/2^k, k = 3, 4, \cdots, 11 for a fixed spatial grid size \triangle x = 1/2^6, 1/2^{7} with viscosity \nu = 0.1, 0.01 . The horizontal and vertical axes are represented by -\ln(h) and \ln(\|\mathbf{e}^N(h)\|_2) , respectively. The behavior of the error confirms to a slop of approximately 2 when h \geq \triangle x^{(p+1)/3} , since \mathcal{O}(h^2) dominates the error, as seen in Figure 1. However, the error stops decreasing after the point at which the {\triangle x^{p+1}}/{h} term begins to dominate it. Moreover, we can see that the error is slowly increasing to certain levels as h is decreasing with h \leq \triangle x^{(p+1)/3} . This behavior comes from the term \mathcal{O}(1/h) and thus it is not useful for the accuracy of the scheme to choose a temporal step size such as h < \triangle x^{(p+1)/3} .

    Figure 1.  The error with respect to h for different interpolations.

    Example 4.2. This example is the unsteady two-dimensional Burgers' equation with periodic boundary condition described as

    \begin{equation*} \label{num:Ex5} u_{t}(t,x,y )+\vec{u}(t,x,y)\cdot \nabla u(t,x,y) = \nu\Delta u(t,x,y),\quad (x,y)\in (-1,1)\times(-1,1), \; t\in(0,1], \end{equation*}

    where \vec{u}(t, x, y) = [{u}(t, x, y), {u}(t, x, y)]^{\mathtt T} and the initial value is given as u(0, x, y) = -\sin\left(\pi y \right).

    In this example, we uniformly discretize the spatial domain given by \triangle x = \triangle y and present the resulting convergence orders in Theorem 3.6 as extended in the two-dimensional problem. With the lack of an exact solution for the problem, we compute reference solutions by setting (h, \triangle x) = ({1}/{2^{8}}, {1}/{2^{8}}) with cubic Lagrangian interpolation and (h, \triangle x) = ({1}/{2^{17}}, {1}/{2^{8}}) with quintic Lagrangian interpolation to estimate the temporal convergence and spatial convergence rates, respectively. To estimate the temporal order of convergence, we approach the test problem similarly to Example 1 by varying h = {1}/{ 2^{k}} , k = 1, 2, \cdots, 7 for a fixed small spatial grid size \triangle x = {1}/{2^{11}} , which guarantees that the two dependent spatial terms \triangle x^4 and \triangle x^{p+1}/h become sufficiently smaller than h^2 . The numerical results for three different viscosities \nu = 0.01, 0.1, 1.0 are displayed in Figure 2(a). The figures indicate that the method has an almost second-order convergence rate in time, which allows us to use a much larger temporal than spatial step size, guaranteeing the theoretical temporal order of convergence shown in Theorem 3.6 with good stability.

    Figure 2.  (a) Rate Temporal Convergence (RTC) with cubic Lagrangian interpolation and \triangle x = {1}/{2^{11}} . (b) Errors with quintic Lagrange interpolation and h = \triangle x ^2 .

    To check the spatial order of convergence, we reduce spatial step sizes by half based on the relation h = \triangle x^2 , for instance (h, \triangle x) = (1/2^{2k}, 1/2^{k}) , k = 1, 2, \cdots, 7 while applying the quintic Lagrange interpolation for different viscosity coefficients \nu = 0.01, 0.1, 1.0 , as displayed in Figure 2(b). The figure shows that the proposed scheme has a fourth-order spatial convergence rate, as seen in the previous example.

    In the next numerical experiment, we support the numerical stability result of Theorem 3.8 for the proposed BSLM by estimating the generic constant C_{s} in the theorem as well as showing the behavior of the numerical solution over a sufficiently long time T with no restriction of step sizes \triangle x and h . The constant C_{s} is computed as the ratio of the solution at between the initial time t_0 = 0 with its spatial differences and the final time T = 1 such as

    C_{s} = \frac{ \| \mathbf{U}^N\|_2}{ \left\|{{ \mathbf{U}}}^{0}\right\|_2+ \left\|{\delta}_x {{ \mathbf{U}}}^{0}\right\|_2+ \triangle x \left\|{\delta}^2_x {{ \mathbf{U}}}^{0}\right\|_2 }.

    We present the estimates of C_{s} while varying \triangle x = 1/2^k, \; k = 4, 5, \cdots, 10 for fixed h = 1/2^4 and while varying h = 1/2^k, \; k = 4, 5, \cdots, 10 for fixed \triangle x = 1/2^4 when \nu = 0.01, 0.05, 0.1, 0.5 , and 1 , which are plotted in Figure 3(a) and 3(b), respectively. The presented figures show that the constant C_{s} is independent of the sizes of \triangle x and h as claimed in Theorem 3.8. We next exhibit the behavior of the discrete \ell_2 -norm of the solution \mathbf{U}^n versus time t^n < T = 10 with different spatial step sizes \triangle x = 1/2^k, k = 1, 5, 9, 13 for fixed h = 1/2 when \nu = 0.01 and \nu = 1.0 , which are displayed in Figure 4, respectively. The experiments show that the solutions consistently decrease for a long time without blowing up under various spatial step sizes in both cases of \nu , even when h is 2^{12}( = 4096) times larger than \triangle x in the case k = 13 ; this provides strong evidence for the numerical stability of the scheme. Therefore, the experiment results guarantee that the numerical scheme of the proposed BSLM is stable without any restriction of the relation between mesh and temporal discretizations, as we claimed.

    Figure 3.  (a) Estimates of C_{s} for fixed h = 1/2^4 . (b) Estimates of C_{s} for fixed \triangle x = 1/2^4 .
    Figure 4.  (a) Estimates of \| \mathbf{U}^n\|_2 for 0\leq t^n \leq 10 when \nu = 0.01 . (b) Estimates of \| \mathbf{U}^n\|_2 for 0\leq t^n \leq 10 when \nu = 1.0 .

    We end the section with a comment about the performance of the solution with respect to the viscosity. It is well known that one of the difficulties for solving Burgers' equation occurs when a small viscosity \nu is applied, which generates solution curves steepen as a shock like discontinuity as time increases. This creates spurious oscillations called Gibbs phenomena near the shock front [18]. The design of the schemes to overcome this problem with sharp front solutions is one of the challenges [7].

    In Figure 5, we present the behaviors of the numerical solution obtained by the proposed BSLM when viscosity coefficients \nu = 1, 0.1, 0.01, 0.001 . The figures show the proposed scheme does not suffer from the sharp front phenomena up to the viscosity \nu = 0.001 . For more related results or comparisons of the proposed method, we refer the interested readers to [22,24,26,27,28] and references therein.

    Figure 5.  Behavior of the numerical solution u(t, x) when \nu = 1, 0.1, 0.01, 0.001 for h = 1/2^9 , \triangle x = 1/2^9 , p = 3.

    We have proposed a BSLM for solving Burgers' equation with the periodic boundary condition and have proved that the scheme has second-order and fourth-order convergence in time and space, respectively. In particular, the asymptotic expression possesses a reciprocal term of temporal step size that is inversely proportional to interpolation order. Furthermore, we have shown that the proposed BSLM is unconditionally stable, allowing the use of a large size of temporal step compared to spatial step size, with examples of supporting numerical experiments. The analysis of the model problem suggests an extension of linear advection–diffusion problems and high-dimensional model problems with no restriction due to boundary conditions using the same order of discretization, implying the cognate results for order of convergence. The paper contains a theoretical result, namely, the asymptotic order of convergence and numerical stability of the scheme with respect to the discrete \ell_2 -norm with varying viscosity coefficients. The overall theoretical order of convergence and numerical stability of the scheme are also numerically verified with examples.

    The first author (P.Kim) was supported by the R & D program through Korea Institute of Fusion Energy (KFE) funded by the Government funds, Republic of Korea (No. EN2241-3) and the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2020R1A4A1018190).

    The authors declare no conflicts of interest.

    Lagrange interpolation and its properties

    In this appendix, we review the Lagrange interpolation polynomial used to analyze the convergence and stability of the proposed BSLM and also introduce its properties. We consider the domain I to be uniformly discretized by a spatial grid size \triangle x and a fixed number J as shown in (2.2).

    For a given periodic function w({x}) defined in I and each local cell I_j: = [x_{j-1}, x_j] , let {\mathfrak L}_{j}w({x}) be a j^{th} local Lagrangian interpolation polynomial of degree p\geq1 defined by

    \begin{equation} \begin{aligned} & {\mathfrak L}_{j}w({x}): = \sum\limits_{k\in {\Lambda}_j} w_k { L}_{j,k}(x),\quad x\in I_j,\\ & L_{j, k}( x) : = \prod\limits_{\substack{j_\ell\in {\Lambda}_j\backslash \{j_k\}}} \frac{x-{x}_{j_\ell} }{ x_{j_k} -x_{j_\ell} },\quad {\Lambda}_j: = \left\{ {j_\ell}: = j+\ell-\left[\frac{p+1}{2}\right],\quad 0\leq \ell \leq p \right\}, \end{aligned} \end{equation} (A.1)

    where w_k: = w(x_k) and [m] denotes the integer satisfying [m]\leq m < [m]+1 . For the interpolation polynomial near boundary points, we consider w_{J+ k} = w_{ k} and w_{-k} = w_{J- k} for the periodicity when J+k , -k \in {\Lambda}_j . Using the local interpolation {\mathfrak L}_{j}w(x) of (A.1), a sufficiently smooth function w(x) can be approximated as

    \begin{equation} w({x}) \approx {{\mathfrak L}} \mathbf{w}({x}): = \sum\limits_{1\leq j\leq J}\mathbf{1}_{{I}_{j}}({x}){{\mathfrak L}_{j}}w({x}), \quad \mathbf{w}: = [w_{1},\cdots,w_{J}]^{\mathtt T}, \end{equation} (A.2)

    and for x\in I_j , its truncation error (see [2] for detail) satisfies

    \begin{equation} \Bigl|w({x})- {{\mathfrak L}} \mathbf{w}({x})\Bigr| = \frac{ \big| \prod\limits_{j_\ell\in {\Lambda}_{j}} (x-x_{j_\ell})\big|}{(p+1)!} \big|w^{(p+1)}(\xi_j)\big| \leq \frac{\triangle x^{p+1}}{4(p+1)}\big\|w^{(p+1)}\big\|_{\infty} \end{equation} (A.3)

    for some \xi_j \in I_j , since |\prod_{{j_\ell\in {\Lambda}_{j} }} (x-x_{j_\ell})| \leq\frac{1}{4} {\triangle x ^{p+1}\left[\frac{p+1}{2}\right]! \left[\frac{p+2}{2}\right]!} for uniformly spaced. Here, \mathbf{1}_{{I}_{j}} denotes the indicator function and \|\cdot\|_{\infty} is the infinity norm. Note that the interpolation {\mathfrak L} \mathbf{w}({x}) is piece-wise continuous in I but in general not differentiable at the grid points. To overcome this non-differentiability of {\mathfrak L} \mathbf{w}({x}) for {x}\in{I}_{j} on the interface, we define a derivative \widetilde{D}_x {\mathfrak L} \mathbf{w}({x}) such as

    \begin{equation} {\widetilde{D}_x}({\mathfrak L} \mathbf{w}(x)): = \begin{cases} \frac{1}{2}\Big( {D_x}({\mathfrak L}_{j}w(x))+{D_x} ({\mathfrak L}_{j+1} w(x))\Big) & \text{if } x = x_j, \\ {D_x}({\mathfrak L}_{j}w(x)) & \text{otherwise.} \end{cases} \end{equation} (A.4)

    We note [27,Lemma 2.3] that for a sufficiently smooth function w({x}) defined on I_j , the derivative \widetilde D_x defined in (A.4) satisfies

    \begin{equation} {D_x }w(x)-{ \widetilde D_x}({\mathfrak L} \mathbf{w}(x)) = \mathcal{O}(\triangle x^{p}). \end{equation} (A.5)

    Lastly, we introduce some boundedness properties for the interpolation polynomial.

    Lemma A.1. The derivative of {\mathfrak L}_{j}w({x}) defined in (A.1) can be bounded by

    \bigl|{D_x}({\mathfrak L}_{j}w(x))\bigr|\leq d_1\sum\limits_{k\in {\Lambda'_j} } \bigl|\delta_x w_k\bigr|,\quad\delta_x w_k: = \frac{w_{k}-w_{k-1}}{\triangle x},

    where {\Lambda'_j}: = \{j_\ell\in \Lambda_j | 1\leq \ell\leq p \}, \; d_1: = d_0 p(p+1), \; d_0: = \max \limits_{x\in I_j} \Big(\prod\limits_{ \substack{k, \ell \in {\Lambda}_j\\ k\neq \ell} } \left| \frac{ x-x_k}{x_\ell-x_k}\right|\Big) . Furthermore, for x\in I_j , we have

    \Bigl|{D_x}({\mathfrak L}_{j}w({x}))\Bigr|\leq \frac{d_2}{\triangle x} \| \mathbf{w}\|^2, \quad \mathbf{w} = [w_1,\cdots,w_J]^{\mathtt T},

    where d_2: = 2d_0p^2(p+1) , and \|\cdot\| denotes the standard Euclidean L_2 -norm.

    Proof. First note that the Lagrange interpolation basis L_{j, k}(x) of degree p given by (A.1) satisfies

    \begin{equation} \sum\limits_{k \in {\Lambda}_j} L_{j,k}(x) = 1,\quad {D_x}\Big(\sum\limits_{k\in {\Lambda}_j}L_{j,k}(x)\Big) = 0. \end{equation} (A.6)

    Hence, using the property (A.6), one can easily check that

    \begin{equation} \begin{aligned} {D_x}({\mathfrak L}_{j} w({x}))& = \sum\limits_{k \in {\Lambda}_j} w_{k} {D_x}(L_{j,k}(x)) = \sum\limits_{k \in \Lambda'_j} \sum\limits_{\ell = j_1}^{k} {D_x} (L_{j,k}(x))\Bigl( w_{\ell} - w_{\ell-1} \Bigr). \end{aligned} \end{equation} (A.7)

    From the definition of (A.1), we have that for any {k \in {\Lambda}_j}

    \begin{equation} \begin{aligned} \Bigl|{D_x}(L_{j,k}(x))\Bigr|\leq & \sum\limits_{i \in {\Lambda}_j\backslash \{k\}}\left| \frac{ \prod\limits_{\ell\in {\Lambda}_j\backslash \{ i,k\}}(x-x_{\ell}) }{ \prod\limits_{\ell\in {\Lambda}_j\backslash \{ k\}} (x_k-x_\ell)}\right|\\ = & \sum\limits_{i \in {\Lambda}_j\backslash \{ k\}}\frac{1}{\left|x_k-x_i\right|}\prod\limits_{\ell \in {\Lambda}_j \backslash \{ i,k\}} \left| \frac{ x-x_\ell}{x_k-x_\ell}\right| \\ \leq & \frac{p}{\triangle x}d_0, \end{aligned} \end{equation} (A.8)

    and Eqs (A.7) and (A.8) imply

    \begin{aligned} \Bigl|{D_x}({\mathfrak L}_{j}w({ x}))\Bigr|& \leq d_0 p (p +1)\sum\limits_{k \in {\Lambda'_j}}\Bigl|\frac{ w_{k}-w_{k-1}}{\triangle x}\Bigr| \leq d_1 \sum\limits_{k \in {\Lambda'_j}} \Bigl|\frac{ w_{k}-w_{k-1}}{\triangle x}\Bigr|, \end{aligned}

    which proves the first assertion. For the second assertion, we apply the triangle inequality and the Cauchy-Schwartz inequality to obtain

    \begin{aligned} \Bigl|{D_x}{\mathfrak L}_{j} w({ x})\Bigr|& \leq \frac{d_0p(p+1)}{\triangle x} \sum\limits_{k \in \Lambda'_j}\Bigl(|w_{k}| +| w_{k-1}| \Bigr)\\ &\leq \frac{d_2}{\triangle x} \| \mathbf{w}\|^2,\\ \end{aligned}

    which completes the proof.



    [1] A. Allievi, R. Bermejo, Finite element modified method of characteristics for the Navier–Stokes equations, Int. J. Numer. Methods Fluids, 32 (2000), 439–463.
    [2] K. E. Atkinson, An introduction to numerical analysis, 2 Eds., Canada: Wiley, 1989.
    [3] A. Bermúdez, M. R. Nogueiras, C. Vázquez, Numerical analysis of convection-diffusion-reaction problems with higher order characteristics/finite elements. Part I: time discretization, SIAM J. Numer. Anal., 44 (2006), 1829–1853. https://doi.org/10.1137/04061201 doi: 10.1137/04061201
    [4] A. Bermúdez, M. R. Nogueiras, C. Vázquez, Numerical analysis of convection-diffusion-reaction problems with higher order characteristics/finite elements. Part II: fully discretized scheme and quadrature formulas, SIAM J. Numer. Anal., 44 (2006), 1854–1876. https://doi.org/10.1137/040615109 doi: 10.1137/040615109
    [5] K. Boukir, Y. Maday, B. Métivet, A high order characteristics method for the incompressible Navier-Stokes equations, Comput. Methods Appl. Mech. Eng., 116 (1994), 211–218. https://doi.org/10.1016/S0045-7825(94)80025-1 doi: 10.1016/S0045-7825(94)80025-1
    [6] K. Boukir, Y. Maday, B. Métivet, E. Razafindrakoto, A high-order characteristics/finite element method for the incompressible Navier–Stokes equations, Int. J. Numer. Methods Fluids, 25 (1998), 1421–1454.
    [7] H. P. Bhatt, A. Q. M. Khaliq, Fourth-order compact schemes for the numerical simulation of coupled Burgers' equation, Comput. Phys. Commun., 200 (2016), 117–138. https://doi.org/10.1016/j.cpc.2015.11.007 doi: 10.1016/j.cpc.2015.11.007
    [8] R. Bermejo, P. del Sastre, L. Saavedra, A second order in time modified Lagrange–Galerkin finite element method for the incompressible Navier–Stokes equations, SIAM J. Numer. Anal., 50 (2012), 3084–3109. https://doi.org/10.1137/11085548X doi: 10.1137/11085548X
    [9] R. Bermejo, L. Saavedra, Modified Lagrange–Galerkin methods to integrate time dependent incompressible Navier–Stokes equations, SIAM J. Sci. Comput., 37 (2015), B779–B803. https://doi.org/10.1137/140973967 doi: 10.1137/140973967
    [10] S. Bak, P. Kim, S. Park, Development of a parallel CUDA algorithm for solving 3D guiding center problems, Comput. Phys. Commun., 276 (2022), 108331. https://doi.org/10.1016/j.cpc.2022.108331 doi: 10.1016/j.cpc.2022.108331
    [11] W. Boscheri, M. Tavelli, L. Pareschi, On the construction of conservative semi-Lagrangian IMEX advection schemes for multiscale time dependent PDEs, J. Sci. Comput., 90 (2022), 97. https://doi.org/10.1007/s10915-022-01768-0 doi: 10.1007/s10915-022-01768-0
    [12] S. Y. Cho, S. Boscarino, G. Russo, S. B. Yun, Conservative semi-Lagrangian schemes for kinetic equations Part I: Reconstruction, J. Comput. Phys., 432 (2021), 110159. https://doi.org/10.1016/j.jcp.2021.110159 doi: 10.1016/j.jcp.2021.110159
    [13] S. Y. Cho, S. Boscarino, G. Russo, S. B. Yun, Conservative semi-Lagrangian schemes for kinetic equations Part II: applications, J. Comput. Phys., 436 (2021), 110281. https://doi.org/10.1016/j.jcp.2021.110281 doi: 10.1016/j.jcp.2021.110281
    [14] J. Douglas, T. F. Russell, Numerical methods for convection-dominated diffusion problems based on combining the method of characteristics with finite element or finite difference procedures, SIAM J. Numer. Anal., 19 (1982), 871–885. https://doi.org/10.1137/0719063 doi: 10.1137/0719063
    [15] R. E. Ewing, T. F. Russell, Multistep Galerkin method along characteristics for convection-diffusion problems, Adv. Comput. Methods Partial Differ. Equ. IV, 4 (1981), 28–36.
    [16] G. Fourestey, S. Piperno, A second-order time-accurate ALE Lagrange–Galerkin method applied to wind engineering and control of bridge profiles, Comput. Methods Appl. Mech. Eng., 193 (2004), 4117–4137. https://doi.org/10.1016/j.cma.2003.12.060 doi: 10.1016/j.cma.2003.12.060
    [17] M. Falcone, R. Ferretti, Convergence analysis for a class of high-order semi-Lagrangian advection schemes, SIAM J. Numer. Anal., 35 (1998), 909–940. https://doi.org/10.1137/S0036142994273513 doi: 10.1137/S0036142994273513
    [18] J. Grétarsson, R. Fedkiw, Fully conservative leak-proof treatment of thin solid structures immersed in compressible fluids, J. Sci. Comput., 245 (2013), 160–204. https://doi.org/10.1016/j.jcp.2013.02.017 doi: 10.1016/j.jcp.2013.02.017
    [19] P. Galán del Sastre, R. Bermejo, Error analysis for hp-FEM semi-Lagrangian second order BDF method for convection-dominated diffusion problems, J. Sci. Comput., 49 (2011), 211–237. https://doi.org/10.1007/s10915-010-9454-2 doi: 10.1007/s10915-010-9454-2
    [20] R. M. Gray, Toeplitz and circulant matrices: a review, Found. Trends Commun. Inf. Theory, 2 (2006), 155–239. http://dx.doi.org/10.1561/0100000006 doi: 10.1561/0100000006
    [21] P. Kim, S. Bak, Algorithm for a cost-reducing time-integration scheme for solving incompressible Navier–Stokes equations, Comput. Methods Appl. Mech. Eng., 373 (2021), 113546. https://doi.org/10.1016/j.cma.2020.113546 doi: 10.1016/j.cma.2020.113546
    [22] P. Kim, D. Kim, Convergence and stability of a BSLM for advection–diffusion models with Dirichlet boundary conditions, Appl. Math. Comput., 366 (2020), 124744. https://doi.org/10.1016/j.amc.2019.124744 doi: 10.1016/j.amc.2019.124744
    [23] H. Notsu, M. Tabata, A single-step characteristic-curve finite element scheme of second order in time for the incompressible Navier–Stokes equations, J. Sci. Comput., 38 (2009), 1–14. https://doi.org/10.1007/s10915-008-9217-5 doi: 10.1007/s10915-008-9217-5
    [24] S. Park, P. Kim, Y. Jeon, S. Bak, An economical robust algorithm for solving 1D coupled Burgers' equations in a semi-Lagrangian framework, Appl. Math. Comput., 428 (2022), 127185. https://doi.org/10.1016/j.amc.2022.127185 doi: 10.1016/j.amc.2022.127185
    [25] X. Piao, S. Kim, P. Kim, D. Kim, A new time stepping method for solving one dimensional Burgers' equations, Kyungpook Math. J., 52 (2012), 327–346. http://dx.doi.org/10.5666/KMJ.2012.52.3.327 doi: 10.5666/KMJ.2012.52.3.327
    [26] X. Piao, S. Bu, S. Bak, P. Kim, An iteration free backward semi-Lagrangian scheme for solving incompressible Navier–Stokes equations, J. Comput. Phys., 283 (2015), 189–204. https://doi.org/10.1016/j.jcp.2014.11.040 doi: 10.1016/j.jcp.2014.11.040
    [27] X. Piao, S. Kim, P. Kim, J. Kwon, D. Yi, An iteration free backward semi-Lagrangian scheme for guiding center problems, SIAM J. Numer. Anal., 53 (2015), 619–643. https://doi.org/10.1137/130942218 doi: 10.1137/130942218
    [28] X. Piao, P. Kim, D. Kim, One-step L (\alpha)-stable temporal integration for the backward semi-Lagrangian scheme and its application in guiding center problems, J. Comput. Phys., 366 (2018), 327–340. https://doi.org/10.1016/j.jcp.2018.04.019 doi: 10.1016/j.jcp.2018.04.019
    [29] A. Robert, A stable numerical integration scheme for the primitive meteorological equations, Atmos. Ocean, 19 (1981), 35–46. https://doi.org/10.1080/07055900.1981.9649098 doi: 10.1080/07055900.1981.9649098
    [30] H. Rui, M. Tabata, A second order characteristic finite element scheme for convection-diffusion problems, Numer. Math., 92 (2002), 161–177. https://doi.org/10.1007/s002110100364 doi: 10.1007/s002110100364
    [31] A. Staniforth, J. Côté, Semi-Lagrangian integration schemes for atmospheric models—a review, Mon. Weather Rev., 119 (1991), 2206–2223.
    [32] P. K. Smolarkiewicz, J. A. Pudykiewicz, A class of semi-Lagrangian approximations for fluids, J. Atmos. Sci., 49 (1992), 2082–2096.
    [33] C. Temperton, M. Hortal, A. Simmons, A two-time-level semi-Lagrangian global spectral model, Quart. J. Roy. Meteor. Soc., 127 (1991), 111–127. https://doi.org/10.1002/qj.49712757107 doi: 10.1002/qj.49712757107
    [34] C. Temperton, A. Staniforth, An efficient two-time-level semi-Lagrangian semi-implicit integration scheme, Quart. J. Roy. Meteor. Soc., 113 (1987), 1025–1039. https://doi.org/10.1002/qj.49711347714 doi: 10.1002/qj.49711347714
    [35] D. Vít, On the discontinuous Galerkin method for the numerical solution of the Navier–Stokes equations, Int. J. Numer. Methods Fluids, 45 (2004), 1083–1106. https://doi.org/10.1002/fld.730 doi: 10.1002/fld.730
    [36] D. Xiu, G. E. Karniadakis, A semi-Lagrangian high-order method for Navier–Stokes equations, J. Comput. Phys., 172 (2001), 658–684. https://doi.org/10.1006/jcph.2001.6847 doi: 10.1006/jcph.2001.6847
    [37] D. Xiu, S. J. Sherwin, S. Dong, G. E. Karniadakis, Strong and auxiliary forms of the semi-Lagrangian method for incompressible flows, J. Sci. Comput., 25 (2005), 323–346. https://doi.org/10.1007/s10915-004-4647-1 doi: 10.1007/s10915-004-4647-1
    [38] J. R. Yearsley, A semi-Lagrangian water temperature model for advection-dominated river systems, Water Resour. Res., 45 (2009), W12405. https://doi.org/10.1029/2008WR007629 doi: 10.1029/2008WR007629
    [39] G. B. Whitham, Linear and nonlinear waves, Wiley and Sons, 2011.
  • This article has been cited by:

    1. João Gabriel Gomes de Oliveira, Pedro Galán del Sastre, Pablo Alcaide Fernández, Jaime Carpio Huertas, Stabilized quasi-monotone semi-Lagrangian high-order finite element method for a nonhydrostatic ocean model, 2025, 481, 01672789, 134766, 10.1016/j.physd.2025.134766
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2203) PDF downloads(107) Cited by(1)

Figures and Tables

Figures(5)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog