Research article Special Issues

Reduced-order modeling using the frequency-domain method for parabolic partial differential equations

  • This paper suggests reduced-order modeling using the Galerkin proper orthogonal decomposition (POD) to find approximate solutions for parabolic partial differential equations. We first transform a parabolic partial differential equation to the frequency-dependent elliptic equations using the Fourier integral transform in time. Such a frequency-domain method enables efficiently implementing a parallel computation to approximate the solutions because the frequency-variable elliptic equations have independent frequencies. Then, we introduce reduced-order modeling to determine approximate solutions of the frequency-variable elliptic equations quickly. A set of snapshots consists of the finite element solutions of the frequency-variable elliptic equations with some selected frequencies. The solutions are approximated using the general basis of the high-dimensional finite element space in a Hilbert space. reduced-order modeling employs the Galerkin POD for the snapshot subspace spanned by a set of snapshots. An orthonormal basis for the snapshot space can be easily computed using the spectral decomposition of the correlation matrix of the snapshots. Additionally, using an appropriate low-order basis of the snapshot space allows approximating the solutions of the frequency-variable elliptic equations quickly, where the approximate solutions are used for the inverse Fourier transforms to determine the approximated solutions in the time variable. Several numerical tests based on the finite element method are presented to asses the efficient performances of the suggested approaches.

    Citation: Jeong-Kweon Seo, Byeong-Chun Shin. Reduced-order modeling using the frequency-domain method for parabolic partial differential equations[J]. AIMS Mathematics, 2023, 8(7): 15255-15268. doi: 10.3934/math.2023779

    Related Papers:

    [1] Liang He, Yihui Sun, Zhenglong Chen, Fei Teng, Chao Shen, Zhendong Luo . The POD-based reduced-dimension study on the two-grid finite element method for the nonlinear time-fractional wave equation. AIMS Mathematics, 2025, 10(2): 3408-3427. doi: 10.3934/math.2025158
    [2] Justin W. L. Wan . Multigrid method for pricing European options under the CGMY process. AIMS Mathematics, 2019, 4(6): 1745-1767. doi: 10.3934/math.2019.6.1745
    [3] Aslı Alkan, Halil Anaç . The novel numerical solutions for time-fractional Fornberg-Whitham equation by using fractional natural transform decomposition method. AIMS Mathematics, 2024, 9(9): 25333-25359. doi: 10.3934/math.20241237
    [4] Bin He . Developing a leap-frog meshless methods with radial basis functions for modeling of electromagnetic concentrator. AIMS Mathematics, 2022, 7(9): 17133-17149. doi: 10.3934/math.2022943
    [5] Andrey Muravnik . Nonclassical dynamical behavior of solutions of partial differential-difference equations. AIMS Mathematics, 2025, 10(1): 1842-1858. doi: 10.3934/math.2025085
    [6] Cheng Chen . Hyperbolic function solutions of time-fractional Kadomtsev-Petviashvili equation with variable-coefficients. AIMS Mathematics, 2022, 7(6): 10378-10386. doi: 10.3934/math.2022578
    [7] Yashar Mehraliyev, Seriye Allahverdiyeva, Aysel Ramazanova . On one coefficient inverse boundary value problem for a linear pseudoparabolic equation of the fourth order. AIMS Mathematics, 2023, 8(2): 2622-2633. doi: 10.3934/math.2023136
    [8] Afraz Hussain Majeed, Sadia Irshad, Bagh Ali, Ahmed Kadhim Hussein, Nehad Ali Shah, Thongchai Botmart . Numerical investigations of nonlinear Maxwell fluid flow in the presence of non-Fourier heat flux theory: Keller box-based simulations. AIMS Mathematics, 2023, 8(5): 12559-12575. doi: 10.3934/math.2023631
    [9] Jafar Biazar, Fereshteh Goldoust . Multi-dimensional Legendre wavelets approach on the Black-Scholes and Heston Cox Ingersoll Ross equations. AIMS Mathematics, 2019, 4(4): 1046-1064. doi: 10.3934/math.2019.4.1046
    [10] Laiq Zada, Rashid Nawaz, Sumbal Ahsan, Kottakkaran Sooppy Nisar, Dumitru Baleanu . New iterative approach for the solutions of fractional order inhomogeneous partial differential equations. AIMS Mathematics, 2021, 6(2): 1348-1365. doi: 10.3934/math.2021084
  • This paper suggests reduced-order modeling using the Galerkin proper orthogonal decomposition (POD) to find approximate solutions for parabolic partial differential equations. We first transform a parabolic partial differential equation to the frequency-dependent elliptic equations using the Fourier integral transform in time. Such a frequency-domain method enables efficiently implementing a parallel computation to approximate the solutions because the frequency-variable elliptic equations have independent frequencies. Then, we introduce reduced-order modeling to determine approximate solutions of the frequency-variable elliptic equations quickly. A set of snapshots consists of the finite element solutions of the frequency-variable elliptic equations with some selected frequencies. The solutions are approximated using the general basis of the high-dimensional finite element space in a Hilbert space. reduced-order modeling employs the Galerkin POD for the snapshot subspace spanned by a set of snapshots. An orthonormal basis for the snapshot space can be easily computed using the spectral decomposition of the correlation matrix of the snapshots. Additionally, using an appropriate low-order basis of the snapshot space allows approximating the solutions of the frequency-variable elliptic equations quickly, where the approximate solutions are used for the inverse Fourier transforms to determine the approximated solutions in the time variable. Several numerical tests based on the finite element method are presented to asses the efficient performances of the suggested approaches.



    We consider the following parabolic problem to be the target of reduced-order modeling applied in the frequency-domain method. For a given u0H10(Ω),

    {1κut(σu)=finΩ×(0,),u=0onΩ×(0,),u(x,0)=u0inΩ (1.1)

    where Ω represents an open convex polygon in R2; κL2(Ω) and σW1,(Ω) denote positive functions of xΩ satisfying κκκ, σσσ, and |σ|σ with positive constants κ, κ, σ, and σ; f(,t)L2(Ω); and Ω denotes the boundary of Ω.

    Recently, many articles have addressed reduced-order modeling to determine low-order models of dynamical systems, such as very complex turbulence flows and problems of optimization or feedback control problems (see, e.g., [2,3]). As a tool for deriving low-order models for the given problems, many researchers have used the proper orthogonal decomposition method (POD) (see, e.g., [7,8,9,16]). The combination of the isogeometric analysis and POD was investigated for parabolic problems in [20] and unsteady convection-dominated convection-diffusion-reaction problems in [15]. The POD method provides a reduced-order basis for the modal decomposition of an ensemble of functions, such as data obtained during the course of experiments or numerical simulations. For example, suppose a finite series exists with a time step ΔT of finite element numerical solutions, so-called snapshots, of a time-dependent partial differential equation, in which the solutions are approximated using the general nodal basis of a high-dimensional finite element space in a Hilbert space. The space spanned by the snapshots is called the snapshot space. Then, using the Galerkin POD of the snapshot space, an appropriate low-order orthonormal basis, so-called a POD basis, can be employed for a low-dimensional subspace of the snapshot space. Such an orthonormal basis can be easily computed using the spectral decomposition of the correlation matrix of the snapshots in the Hilbert space. Note that the number of POD basis functions is much less than the dimension of the snapshot space in general. Once a low-order POD basis is determined, we can quickly compute approximate solutions of the time-dependent partial differential equation with time step Δt that are much less than ΔT. Thus, we employ the reduced-order modeling of Galerkin POD.

    In this paper, by applying the frequency-domain method to the time-dependent parabolic equation (1.1), we provide reduced-order modeling of the Galerkin POD to determine approximate solutions of frequency-dependent elliptic equations quickly. We first transform a parabolic equation to the frequency-variable elliptic equations using the Fourier integral transform in time. Such a frequency-domain method enables easily implementing a parallel computation algorithm to approximate the frequency-variable solutions because the frequency-variable elliptic equations have independent frequencies (see, e.g., [4,5,6,10,11,17]; see [12,13,18,19] for the case of using the Laplace transformation).

    Applying the Fourier transformation for the space-time problem (1.1), we have the following set of complex-valued elliptic equations depending on the frequency ω: for all ωR,

    {iω1κˆu(σˆu)=ˆfinΩ,ˆu=0onΩ (1.2)

    where f and u are extended by zero for t<0 and t>T for the Fourier transformation. The Fourier transform ˆu(,ω) of a function u(,t) in time and the Fourier inversion are given by

    ˆu(,ω)=u(,t)exp(iωt)dtandu(,t)=12πˆu(,ω)exp(iωt)dω.

    This paper investigates the combination of the frequency-domain method and the reduced-order modeling. We apply Galerkin POD method to determine approximate solutions of the frequency-variable elliptic equations (1.2) instead of the time-dependent parabolic equation (1.1). A set of snapshots consists of the finite element solutions of the frequency-variable elliptic equations with some sampled frequencies, in which the solutions are approximated using the general high-dimensional nodal basis of the finite element space. Then, from the spectral decomposition of the correlation matrix of the snapshots, we determine a low-order Galerkin POD basis for a subspace of the snapshot space spanned by the snapshots. Using a low-order Galerkin POD basis, we compute approximate solutions (POD-solutions) of the frequency-variable elliptic equations for sufficiently many frequencies to determine accurate inverse Fourier transforms for the solutions in the time variable. We use the Gaussian quadrature rule based on Legendre–Gauss–Lobatto (LGL) points for the accurate numerical integration of the inverse Fourier transformation. Thus, the selected frequencies for POD-solutions are sufficiently many LGL points on an appropriate interval. The number of sample snapshots must be much less than the number of the POD-solutions to reduce the total computational cost because the snapshots are approximated using full-dimensional basis functions, but the POD solutions are computed using low-order POD basis functions. Regarding numerical computation, a fast-solving parallel computation can be run to determine the snapshots to reduce the total computational time, which is merit of the Galerkin POD method applied to the frequency-variable equations (see, e.g., [7,8,14]).

    The paper is organized as follows. Section 2 provides an overview of the Galerkin POD, and Section 3 presents reduced-order modeling for the frequency-domain method to approximate the parabolic equation. Finally, Sections 4 presents some numerical experiments.

    The POD of order is to determine a set of ordered orthonormal basis functions, such that the snapshots can be expressed optimally using the selected first basis functions, where is a positive integer [7,8,16]. We briefly review the Galerkin POD in the context of the finite element method. Let Xh be a finite-dimensional subspace of a given Hilbert space X endowed with the inner product (,)X and norm X, and let {ϕp}np=1 be a basis of the space Xh. For example, we consider a nodal basis {ϕp}np=1 for a Galerkin finite element subspace Xh consisting of piecewise linear functions of the Sobolev space X=H1(Ω), where Ω is a given domain. For a set of snapshots S={y1,,ym}Xh, we define a snapshot subspace

    XS=span{y1,,ym}Xh,

    and consider the orthonormal basis {ψp}dp=1 of XS, where d=dimXS is the dimension of XS. Then, the method of Galerkin POD of order d consists of choosing the orthonormal basis such that the mean square X-norm error between the snapshots yq and the corresponding -th partial sums is minimized:

    min{ψp}p=1mq=1yqp=1(ψp,yq)Xψp2Xsubject to (ψp,ψq)X=δpqfor1p,q. (2.1)

    Using the orthogonality of ψp's yields that

    yqp=1(ψp,yq)Xψp2X=yq2Xp=1(ψp,yq)2X.

    Hence, the minimization problem (2.1) is equivalent to the following maximization problem:

    max{ψp}p=1mq=1p=1(ψp,yq)2Xsubject to (ψp,ψq)X=δpqfor1p,q. (2.2)

    Let ˆX=((ϕp,ϕq)X)Rn×n be the positive definite finite element matrix. The matrices containing the coefficients of yq and ψq are denoted by YRn×m and ΨRn×d, respectively, in the expansion with respect to the basis functions ϕp, that is,

    yq(x)=np=1Y(p,q)ϕp(x),(q=1,,m)andψq(x)=np=1Ψ(p,q)ϕp(x),(q=1,,d). (2.3)

    In addition, each snapshot can be expressed as a linear combination of the orthonormal basis functions such that

    yq(x)=dp=1B(p,q)ψp(x)whereB(p,q)=(ψp,yq)X. (2.4)

    Using (2.3) and (2.4) yields

    yq=dp=1(ψp,yq)Xψp=dp=1B(p,q)ψp=ns=1(dp=1Ψ(s,p)B(p,q))ϕs(x)

    and

    B(p,q)=nk=1Ψ(k,p)ns=1Y(s,q)(ϕk,ϕs)X=nk,s=1ΨT(p,k)ˆX(k,s)Y(s,q)

    so that we obtain the following identities

    Y=ΨB, (2.5)

    and

    B=ΨTˆXYRd×m. (2.6)

    Next, Q denotes the first columns and Q the first rows of a given matrix Q. Then, the problem (2.2) is equivalent to the following problem of in matrix form:

    maxΨmq=1p=1B(p,q)2=maxΨ(Ψ)TˆXY2Fsubject to (Ψ)TˆXΨ=I (2.7)

    where F denotes the Frobenius norm of the matrix and I denotes the identity × matrix.

    Let ˆY=ˆX1/2Y, let U and V be the left and right singular vectors, respectively, in the singular value decomposition (SVD) of ˆY:

    ˆY=UΣVT=UdD(Vd)TwithΣ=[D000],D=diag(σ1,,σd).

    For (2.7), using the Fritz John necessary conditions or Karush-Kuhn-Tucker conditions for the optimality of Ψ (see [1] for details) yields

    ˆYˆYTU=U(D2)

    and the optimal solution for (2.7) is given by

    Ψ=(ˆX1/2)1U.

    Hence, the optimal solution for the problem (2.1) is given by

    ψq(x)=np=1Ψ(p,q)ϕp(x),q=1,,. (2.8)

    In contrast, from the singular value decomposition of ˆY that

    Y=ˆX1/2ˆY=ˆX1/2UdD(Vd)T=ΨD(Vd)T.

    Thus, the optimal solution for the POD-basis problem (2.1) can be more easily given by using the right singular vectors of ˆY:

    Ψ=YV(D)1,=1,,d. (2.9)

    The error between the snapshots and their POD solutions is given by (see [8,16])

    mq=1yqp=1(ψp,yq)Xψp2X=dp=+1σ2p. (2.10)

    Let K be the correlation matrix corresponding to the snapshots {yq}mq=1 in the Hilbert space X:

    K=((yq,yp)X)Rm×m.

    Then, the right singular vectors V holds the following spectral decomposition of K:

    K=YTˆXY=ˆYTˆY=VΣVT=VdD2(Vd)T. (2.11)

    In the context of the POD-basis approach for the finite element methods, the eigenvalue problem for m×m matrix ˆYTˆY is more practical to solve than the eigenvalue problem for n×n matrix ˆYˆYT in cases where the size of the input collection m is significantly smaller than the number of coefficients n needed to represent each function for the general basis functions of finite element space.

    Finally, we report on reduced-order modeling for the finite element method. If we have the following linear system of a Galerkin finite element discretization in the space XhX:

    Ay=f,

    then using the linear transformation y=ΨxRn for xR, we obtain the following reduced-order modeling of order to approximate the solution for the above linear system:

    Ax=(Ψ)TfwhereA:=(Ψ)TAΨR×. (2.12)

    This section provides reduced-order modeling using the Galerkin POD for the following complex-valued elliptic equations depending on ω: for all ωR

    {iω1κˆu(σˆu)=ˆfinΩ,ˆu=0onΩ. (3.1)

    This paper applies the standard notation and definitions for the real-valued Sobolev spaces Hs(Ω), associated with the scalar product (,)s and norm s, s0. Nevertheless, H0(Ω) coincides with L2(Ω), in which the associated inner product and norm are denoted by (,) and , respectively. The real and imaginary parts of a complex-valued vector or scalar function ˆv are denoted by ˆvr and ˆvi respectively. Then, the L2(Ω) inner product and norm for complex-valued functions ˆu=ˆur+iˆui and ˆv=ˆvr+iˆvi are given by

    (ˆu,ˆv)c:=Ωˆuˉˆvdx=(ˆu,ˉˆv)andˆvc:=(ˆv,ˉˆv)12=(ˆvr2+ˆvi2)12.

    From now on, we denote by Hsc(Ω):=Hs(Ω)×Hs(Ω), L2c(Ω):=L2(Ω)×L2(Ω), and H10,c(Ω):=H10(Ω)×H10(Ω) where H10(Ω) is the subspace of H10(Ω) vanishing on the boundary of Ω. We identify H10,c(Ω) with V and define the sesquilinear form aω(,):V×VC for ωR as

    aω(ˆu,ˆv)=iω(1κˆu,ˆv)c+(σˆu,ˆv)c.

    Then, the variational formulation of the equation (3.1) can determine ˆu(,ω)V such that

    aω(ˆu,ˆv)=(ˆf,ˆv)c,ˆvV. (3.2)

    For the finite element approximation of (3.1), Let Th be a quasi-regular partition of Ω into triangles with a diameter bounded by h<1. We take a standard finite element subspace VhV such that

    infˆϕVh{ˆvˆϕc+h|ˆvˆϕ|1,c}Ch2|ˆv|2,c,ˆvH2c(Ω),

    where the positive constant C is independent of h and v, and ||1,c and ||2,c denote the seminorms of H1c(Ω) and H2c(Ω), respectively. In this paper, we take the standard piecewise linear finite element space for the space Vh. We assume that Vh=PhiPh where Ph is the standard piecewise linear real value finite element subspace of H10(Ω) over Th with dim(Ph)=n and {ϕj}nj=1 is the nodal basis of Ph. Then, the Galerkin finite element approximation is to find ˆuh(ω)=ˆuh(,ω)Vh such that

    aω(ˆuh(ω),ˆvh)=(ˆf,v)c,ˆvhVh. (3.3)

    In [10], the authors provided the existence of the solution in Theorem 2.1, the stability in Theorem 2.2 and the error estimation in Theorem 3.1. The error estimation is given as follows. A generic positive constant denoted by C may differ from place to place.

    Theorem 3.1. Let ˆu(ω) be the approximate solutions of the Eq (3.2) and let ˆuh(ω) be the approximate solutions of the Eq (3.3). Then the following estimations hold:

    ˆu(ω)ˆuh(ω)1,cCh(ω+1ω)ˆfc (3.4)

    and

    ˆu(ω)ˆuh(ω)cCh2(ω2+1ω)ˆfc. (3.5)

    For the approximate Fourier inversion of frequency-variable solutions to the time variable solutions, we apply the Gaussian quadrature rule based on the LGL-points on an appropriate interval [0,ω] with a sufficiently large ω>0 such that ˆu(ω)=ˆu(,ω) is negligible for |ω|>ω. Let Gω={ωj}Nωj=1 be the set of LGL-points on the interval [0,ω]. Then, the time variable approximate solution uh(x,t) for the real-valued solution u(x,t) of the problem (1.1) is approximated by

    uh(x,t)=1πRe(Nωj=1ˆuh(x,ωj)exp(iωjt)wj), (3.6)

    where wj denote the Gaussian quadrature weights corresponding the LGL-points ωj. The error estimation between the approximate solution uh,Δω and u(x,t) is given in [10] where the approximate Fourier inversion uh,Δω is given by using the composite mid-point rule. This paper applies the Gaussian quadrature rule for the approximate Fourier inversion to reduce the computational cost. We do not provide an error estimation but we focus on the performance of reduced-order modeling using Galerkin POD. The error estimation can be proved following the similar arguments given in [10].

    To determine a Galerkin POD basis, we must construct a set of snapshots consisting of Galerkin finite element solutions ˆuh(ξp)Vh for the problem (3.3) with some selected frequencies, for example, {ξp}Nsp=1, where Ns denotes the number of snapshots. For the efficiency of reduced-order modeling of the frequency-domain problem, the number of samples Ns must be less than the number of approximate solutions Nω to be used for the Fourier inversion. The sets of the real and imaginary parts of the snapshots are denoted as

    XR:={Re(ˆuh(ξp))}Nsp=1andXI:={Im(ˆuh(ξp))}Nsp=1.

    We define two correlation matrices for two Galerkin POD-bases:

    KR=((Re(ˆuh(ξp)),Re(ˆuh(ξq)))L2(Ω))RNs×Ns

    and

    KI=((Im(ˆuh(ξp)),Im(ˆuh(ξq)))L2(Ω))RNs×Ns.

    This paper employs the L2(Ω)-inner product for the correlation matrices. The H1(Ω)-inner product could be used, but we obtained similar results in the experiments. In addition the 2-Euclidean inner product for the coefficient vectors of the snapshots could also be applied (see [8,9,17]).

    We let MhRn×n be the mass matrix over the finite element space Ph such that Mh(p,q)=(ϕq,ϕp)L2(Ω), and YRRn×Ns and YIRn×Ns represent the matrices containing the coefficients of Re(ˆuh(ξp)) and Im(ˆuh(ξp)) with respect to the basis functions ϕp of Ph, respectively. Following the argument in (2.11), suppose that we have the following spectral decompositions for the correlation matrices KR and KI:

    KR=YRMhYTR=VdRRD2R(VdRR)TandKI=YIMhYTI=VdIID2I(VdII)T, (3.7)

    where dR=dim(XR) and dI=dim(XI), and DR=diag(σR1,,σRdR) and DI=diag(σI1,,σIdR). Then, the Galerkin POD bases of order for XR and XI are given by, for q=1,,,

    ψRq(x)=np=1ΨR(p,q)ϕp(x)andψIq(x)=np=1ΨI(p,q)ϕp(x), (3.8)

    where min{dR,dI}, and

    ΨR=YRVdRR(DR)1Rn×dRandΨI=YIVdII(DI)1Rn×dI. (3.9)

    Now we have the reduced-order subspace Vh of Vh of order such that

    Vh=XRiXI

    where

    XR=span{ψRq}q=1andXI=span{ψIq}q=1. (3.10)

    For any ωR, a function ˆuh(ω)Vh=PhiPh can be represented by

    ˆuh(ω)=np=1uRp(ω)ϕp+inp=1uIp(ω)ϕp.

    Then, from the problem (3.3) we have the following linear system of order 2n

    [AR(ω)AI(ω)][uR(ω)uI(ω)]=[fRfI] (3.11)

    where AR(ω),AI(ω)R2n×n are given by

    AR(ω)(p,q)={aω(ϕp,ϕq),p=1,,n,q=1,,n,aω(ϕp,iϕq),p=n+1,,2n,q=1,,n,
    AI(ω)(p,q)={aω(iϕp,ϕq),p=1,,n,q=1,,n,aω(iϕp,iϕq),p=n+1,,2n,q=1,,n,

    and fR,fIRn×1 are given by

    fR(p)=(ˆf,ϕp)c,fI(p)=(ˆf,iϕp)c,p=1,,n.

    Following the arguments of reduced-order modeling of order given in (2.12), we obtain the following reduced system of order 2:

    A(ω)x(ω)=f (3.12)

    where

    A(ω)=[(ΨR)T00(ΨI)T][AR(ω)AI(ω)][ΨR00ΨI],

    and

    [uR(ω)uI(ω)]=[ΨR00ΨI]x(ω)andf=[(ΨR)TfR(ΨI)TfI].

    Using the reduced-order system (3.12), we can quickly compute POD approximate solutions ˆuh(x,ωj) for the LGL-points {ωj}Nωj=1, in which the approximate solutions are used for the Fourier inversion uh(x,t), the POD solution of order , like the approximate formulation (3.6):

    uh(x,t)=1πRe(Nωj=1ˆuh(x,ωj)exp(iωjt)wj). (3.13)

    This section presents computational experiments with three examples focused on the error discretization and elapsed time for the efficient performances of using the POD-basis. For the measure of approximate errors, we define several error types:

    eh0:=u(x,t)uh(x,t)L2(0,T;L2(Ω)),eh1:=u(x,t)uh(x,t)L2(0,T;H1(Ω)),
    eh0:=u(x,t)uh(x,t)L2(0,T;L2(Ω)),eh1:=u(x,t)uh(x,t)L2(0,T;H1(Ω)),

    where uh(x,t) and uh(x,t) denote the approximate solution using the general full-basis of Vh and the approximate POD-solution using the Galerkin POD-basis of order , respectively, and the errors are approximately computed where T=1 and Δt=1/100 are used. The order of the POD-basis is selected from the error estimation between the snapshots and their POD-approximations given in (2.10). The marks 'Time(Full)' and 'Time(POD)' indicate the total CPU elapsed times to find approximate solutions using the full and POD bases, respectively.

    For every example, we simply take Ω=(0,1)2, and we set κ=1 and σ=1 in the problem (1.1):

    {utu=finΩ×(0,T),u=0onΩ×(0,T),u(x,0)=u0inΩ.

    All simulations are computed by using the Matlab program on a 3 GHz Intel Xeon dual-core 64-bit CPU processor with 4.00 GB RAM, without parallel computation, but parallel computation can be used for large-scale real-world problems.

    Example 1. For the first test problem, we take an elementary separable example:

    u(x,y,t)=h(t)g(x,y), where h(t)=1πtt2+1,g(x,y)=sin(3π(x+1)/4)sin(3π(y+1)/4),ˆu(x,y,ω)=ˆh(ω)g(x,y) where ˆh(ω)=iexp(ω).

    In this example, taking ω=20 and Nω=12 is sufficient to illustrate good performance in the approximate results. For snapshots, we set Ns=2 (i.e., two snapshots) with two frequencies (ξj=(j12)Δω, j=1,2 with Δω=ω/2). From the snapshot space, we employed only one POD-basis (i.e., =1). Although the order is too small, the solution is very well approximated because of the simplicity of the problem (see Table 1). The error discretization is similar for the two cases, but the elapsed CPU times are very different, and that of the POD basis is much less than that of full basis. The outcome is because Nω=12 LGL frequency approximate solutions using the full-basis are needed for the full-basis case. But two frequency snapshots using the full-basis and Nω=12 LGL frequency approximate solutions of POD basis solution are needed for the POD basis case, where the cost of the spectral decomposition of the correlation matrix must be added, but it is relatively cheap.

    Table 1.  Discretization errors and CPU time for Example 1.
    h eh0 eh1 Time(Full) eh0 eh1 Time(POD)
    1/16 1.76e04 8.37e04 0.03 1.12e04 5.51e04 0.03
    1/32 4.18e05 1.98e04 0.20 2.65e05 1.30e04 0.06
    1/64 1.02e05 4.80e05 1.34 6.44e06 3.15e05 0.16
    1/128 2.52e06 1.19e05 8.03 1.60e06 7.81e06 1.17
    1/256 6.41e07 2.99e06 46.03 4.18e07 1.99e06 5.27

     | Show Table
    DownLoad: CSV

    Table 2 compares the numerical results for the trapezoidal midpoint quadrature rule and Gaussian quadrature rule for the Fourier inversion. Although we employed 240 trapezoidal midpoints, the performance of this case is much worse than using the Gaussian quadrature rule on the error discretization and elapsed CPU time. Thus, we apply the Gaussian quadrature rule for the Fourier inversion.

    Table 2.  Discretization errors and CPU time using the midpoint quadrature rule for Example 1.
    h eh0 eh1 Time(Full)
    1/16 1.71e04 8.24e04 0.92
    1/32 4.00e05 2.00e04 5.16
    1/64 2.01e05 9.27e05 30.53
    1/128 2.04e05 8.81e05 180.78
    1/256 2.09e05 8.91e05 1010.55

     | Show Table
    DownLoad: CSV

    Example 2. In this example, we take the exact solution that has two separable components.

    u(x,y,t)=h1(t)g1(x,y)+h2(t)g2(x,y), where h1(t)=1πtt2+1,g1(x,y)=sin(3π(x+1)/4)sin(3π(y+1)/4),h2(t)=exp(t2),g2(x,y)=sin(πx)cos(πy),ˆu(x,y,ω)=ˆh1(ω)g1(x,y)+ˆh2(ω)g2(x,y) where ˆh1(ω)=iexp(ω),ˆh2(ω)=πexp(ω24).

    In this example, we set ω=20 and Nω=20. For the snapshots, we set Ns=4 with four frequencies ξj=(j12)Δω, j=1,,4 where Δω=ω/4. We employ POD basis of order =2. Table 3 lists the numerical results, which are very similar to those in Example 1.

    Table 3.  Discretization errors and CPU time for Example 2.
    h eh0 eh1 Time(Full) eh0 eh1 Time(POD)
    1/16 1.71e03 1.25e02 0.08 1.72e03 1.26e02 0.03
    1/32 4.11e04 2.98e03 0.38 4.13e04 3.00e03 0.08
    1/64 1.00e04 7.24e04 2.27 1.00e04 7.28e04 0.28
    1/128 2.46e05 1.78e04 13.88 2.47e05 1.79e04 2.06
    1/256 6.07e06 4.41e05 78.38 6.09e06 4.43e05 9.00

     | Show Table
    DownLoad: CSV

    Example 3. In this example, we take an exact solution that is not separable.

    u(x,y,t)=1πt(txy)2+1sin(πx)sin(πy),ˆu(x,y,ω)=(xyi)exp(ω(1+ixy))sin(πx)sin(πy).

    In this example, we set ω=25 and Nω=24. For the snapshots, we set Ns=20 with four frequencies: ξj=(j12)Δω, j=1,,20 where Δω=ω/20. We take POD basis of order =15. Table 4 presents the numerical results, which are very similar to those in Examples 1 and 2. However, we require more POD-basis functions than in the cases of Examples 1 and 2 to achieve better results.

    Table 4.  Discretization errors and CPU time for Example 2.
    h eh0 eh1 Time(Full) eh0 eh1 Time(POD)
    1/16 8.10e04 3.92e03 0.13 8.10e04 3.92e03 0.09
    1/32 1.92e04 9.27e04 0.53 1.93e04 9.28e04 0.17
    1/64 4.67e05 2.25e04 3.02 4.70e05 2.26e04 0.77
    1/128 1.15e05 5.54e05 16.94 1.17e05 5.66e05 4.17
    1/256 2.85e06 1.38e05 95.00 3.15e06 1.80e05 18.14

     | Show Table
    DownLoad: CSV

    We investigate reduced-order modeling using Galerkin POD to determine approximate solutions of time-dependent parabolic problems. We apply the frequency-domain method to the time-domain problem using the Fourier transformation for a more efficient and fast approximation. This approach of the frequency-domain method enables easily and efficiently implementing the parallel computation. Hence, we transformed the time-dependent parabolic problem into a frequency-dependent elliptic problems. However, these elliptic problems are independent. Next, we provided reduced-order modeling using Galerkin POD to determine a very small-dimensional subspace with the orthonormal POD basis. Using such a POD basis, we quickly approximated the elliptic problems for selected frequencies, in which the frequencies are LGL-points on a given interval.

    The approximate solution of the time-dependent parabolic problem was approximated by the inverse Fourier transformation using the fast Gaussian quadrature rule. In the experiments, we demonstrated that the proposed reduced-order modeling performs very well in error discretization and reducing computational costs.

    This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2023R1A2C1006588).

    The authors declare that there are no conflicts of interest.



    [1] J. A. Atwell, B. B. King, Reduced-order controllers for spatially distributed systems via proper orthogonal decomposition, SIAM J. Sci. Comput., 26 (2004), 128–151. https://doi.org/10.1137/S1064827599360091 doi: 10.1137/S1064827599360091
    [2] J. Burkardt, M. Gunzburger, H. C. Lee, Centroidal Voronoi tessellation-based reduced-order modeling of complex systems, SIAM J. Sci. Comput., 28 (2006), 459–484. https://doi.org/10.1137/5106482750342221x doi: 10.1137/5106482750342221x
    [3] J. Burkardt, M. Gunzburger, H. C. Lee, POD and CVT-based reduced-order modeling of Navier-Stokes flows, Comput. Meth. Appl. Mech. Eng., 196 (2006), 337–355. https://doi.org/10.1016/j.cma.2006.04.004 doi: 10.1016/j.cma.2006.04.004
    [4] J. Douglas, Jr., J. E. Santos, D. Sheen, Approximation of scalar waves in the space-frequency domain, Math. Models Meth. Appl. Sci., 4 (1994), 509–531. https://doi.org/10.1142/S0218202594000297 doi: 10.1142/S0218202594000297
    [5] J. Douglas, Jr., J. E. Santos, D. Sheen, L. S. Bennethum, Frequency domain treatment of one-dimensional scalar waves, Math. Model Meth. Appl. Sci., 3 (1993), 171–194. https://doi.org/10.1142/S0218202593000102 doi: 10.1142/S0218202593000102
    [6] X. Feng, D. Sheen, An elliptic regularity estimate for a problem arising from the frequency domain treatment of waves, Trans. Ams. Math. Soc., 346 (1994), 475–487.
    [7] K. Kunisch, S. Volkwein, Control of the Burgers equation by a reduced-order approach using proper orthogonal decomposition, J. Optim. Theory Appl., 102 (1999), 345–371. https://doi.org/10.1023/A:1021732508059 doi: 10.1023/A:1021732508059
    [8] K. Kunisch, S. Volkwein, Galerkin proper orthogonal decomposition methods for parabolic equations, Numer. Math., 90 (2001), 117–148. https://doi.org/10.1007/s002110100282 doi: 10.1007/s002110100282
    [9] K. Kunisch, S. Volkwein, Galerkin proper orthogonal decomposition methods for a general equation in fluid dynamics, SIAM J. Numer. Anal., 40 (2002), 492–515. https://doi.org/10.1137/S0036142900382612 doi: 10.1137/S0036142900382612
    [10] C. O. Lee, J. Lee, D. Sheen, Y. Yeom, A frequency-domain parallel method for the numerical approximation of parabolic problems, Comput. Meth. Appl. Mech. Eng., 169 (1999), 19–29. https://doi.org/10.1016/S0045-7825(98)00168-6 doi: 10.1016/S0045-7825(98)00168-6
    [11] C. O. Lee, J. Lee, D. Sheen, Frequency domain formulation of linearized Navier-Stokes equations, Comput. Meth. Appl. Mech. Eng., 187 (2000), 351–362. https://doi.org/10.1016/S0045-7825(99)00132-2 doi: 10.1016/S0045-7825(99)00132-2
    [12] J. Lee, D. Sheen, An accurate numerical inversion of Laplace transforms based on the location of their poles, Comput. Math. Appl., 48 (2004), 1415-1423. https://doi.org/10.1016/j.camwa.2004.08.003 doi: 10.1016/j.camwa.2004.08.003
    [13] J. Lee, D. Sheen, A parallel method for backward parabolic problems based on the Laplace transformation, SIAM J. Numer. Anal., 44 (2006), 1466–1486. https://doi.org/10.1137/050624649 doi: 10.1137/050624649
    [14] H. C. Lee, S. W. Lee, G. R. Piao, Reduced-order modeling of Burgers equations based on centroidal Voronoi tessellation, Int. J. Numer. Anal. Model., 4 (2007), 559–583.
    [15] R. Li, Q. Wu, S. Zhu, Proper orthogonal decomposition with SUPG-stabilized isogeometric analysis for reduced order modeling of unsteady convection-dominated convection-diffusion-reaction problems, J. Comput. Phys., 387 (2019), 280–302. https://doi.org/10.1016/j.jcp.2019.02.051 doi: 10.1016/j.jcp.2019.02.051
    [16] Y. C. Liang, H. P. Lee, S. P. Lim, W. Z. Lin, K. H. Lee, C. G. Wu, Proper orthogonal decomposition and its application–part 1: theory, J. Sound Vib., 252 (2002), 527–544. https://doi.org/10.1006/jsvi.2001.4041 doi: 10.1006/jsvi.2001.4041
    [17] J. K. Seo, B. C. Shin, Numerical solutions of Burgers equation by reduced-order modeling based on pseudo-spectral collocation method, J. Korean Soc. Ind. Appl. Math., 19 (2015), 123–135. https://doi.org/10.12941/jksiam.2015.19.123 doi: 10.12941/jksiam.2015.19.123
    [18] D. Sheen, I. H. Sloan, V. Thomée, A parallel method for time-discretization of parabolic problems based on contour integral representation and quadrature, Math. Comp., 69 (2000), 177-195.
    [19] D. Sheen, I. H. Sloan, V. Thomée, A parallel method for time-discretization of parabolic equations based on Laplace transformation and quadrature, IMA J. Numer. Anal., 23 (2003), 269–299. https://doi.org/10.1093/imanum/23.2.269 doi: 10.1093/imanum/23.2.269
    [20] S. Zhu, L. Dede, A. Quarteroni, Isogeometric analysis and proper orthogonal decomposition for parabolic problems, Numer. Math., 135 (2017), 333–370. https://doi.org/10.1007/s00211-016-0802-5 doi: 10.1007/s00211-016-0802-5
  • 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(1566) PDF downloads(63) Cited by(0)

Figures and Tables

Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog