Research article Special Issues

A fully discrete HDG ensemble Monte Carlo algorithm for a heat equation under uncertainty

  • This paper has introduced a novel fully discrete hybridizable discontinuous Galerkin (HDG) ensemble Monte Carlo method (FEMC-HDG) tailored for solving the heat equation with random diffusion and Robin coefficients. The FEMC-HDG method solves a single linear system with multiple right-hand side vectors per time step. We established stability analysis and error estimates that are optimal in the spatial and first-order accuracy in time for the L(0,T,L2(D))-norm error estimate. Numerical experiments were included to confirm the theoretical convergence and showcase the method's efficiency.

    Citation: JinJun Yong, Changlun Ye, Xianbing Luo. A fully discrete HDG ensemble Monte Carlo algorithm for a heat equation under uncertainty[J]. Networks and Heterogeneous Media, 2025, 20(1): 65-88. doi: 10.3934/nhm.2025005

    Related Papers:

    [1] Tingfu Yao, Changlun Ye, Xianbing Luo, Shuwen Xiang . A variational MAX ensemble numerical algorism for a transient heat model with random inputs. Networks and Heterogeneous Media, 2024, 19(3): 1013-1037. doi: 10.3934/nhm.2024045
    [2] Patrick Henning, Mario Ohlberger . The heterogeneous multiscale finite element method for advection-diffusion problems with rapidly oscillating coefficients and large expected drift. Networks and Heterogeneous Media, 2010, 5(4): 711-744. doi: 10.3934/nhm.2010.5.711
    [3] Salim Meddahi, Ricardo Ruiz-Baier . A new DG method for a pure–stress formulation of the Brinkman problem with strong symmetry. Networks and Heterogeneous Media, 2022, 17(6): 893-916. doi: 10.3934/nhm.2022031
    [4] Shi Jin, Min Tang, Houde Han . A uniformly second order numerical method for the one-dimensional discrete-ordinate transport equation and its diffusion limit with interface. Networks and Heterogeneous Media, 2009, 4(1): 35-65. doi: 10.3934/nhm.2009.4.35
    [5] Michael Herty, Lorenzo Pareschi, Mohammed Seaïd . Enskog-like discrete velocity models for vehicular traffic flow. Networks and Heterogeneous Media, 2007, 2(3): 481-496. doi: 10.3934/nhm.2007.2.481
    [6] Mengjun Yu, Kun Li . A data-driven reduced-order modeling approach for parameterized time-domain Maxwell's equations. Networks and Heterogeneous Media, 2024, 19(3): 1309-1335. doi: 10.3934/nhm.2024056
    [7] Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo . On boundary optimal control problem for an arterial system: First-order optimality conditions. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027
    [8] Xiongfa Mai, Ciwen Zhu, Libin Liu . An adaptive grid method for a singularly perturbed convection-diffusion equation with a discontinuous convection coefficient. Networks and Heterogeneous Media, 2023, 18(4): 1528-1538. doi: 10.3934/nhm.2023067
    [9] Olli-Pekka Tossavainen, Daniel B. Work . Markov Chain Monte Carlo based inverse modeling of traffic flows using GPS data. Networks and Heterogeneous Media, 2013, 8(3): 803-824. doi: 10.3934/nhm.2013.8.803
    [10] Andreas Hiltebrand, Siddhartha Mishra . Entropy stability and well-balancedness of space-time DG for the shallow water equations with bottom topography. Networks and Heterogeneous Media, 2016, 11(1): 145-162. doi: 10.3934/nhm.2016.11.145
  • This paper has introduced a novel fully discrete hybridizable discontinuous Galerkin (HDG) ensemble Monte Carlo method (FEMC-HDG) tailored for solving the heat equation with random diffusion and Robin coefficients. The FEMC-HDG method solves a single linear system with multiple right-hand side vectors per time step. We established stability analysis and error estimates that are optimal in the spatial and first-order accuracy in time for the L(0,T,L2(D))-norm error estimate. Numerical experiments were included to confirm the theoretical convergence and showcase the method's efficiency.



    This paper concentrates on the numerical simulation of a heat problem involving random diffusion and Robin coefficients. The problem is to find u(t,x,ω) and p(t,x,ω) such that

    {κ(t,x,ω)p(t,x,ω)u(t,x,ω)=0, in [0,T]×D×Ω,u(t,x,ω)tp(t,x,ω)=f(t,x,ω), in [0,T]×D×Ω,p(t,x,ω)n=0, on [0,T]×D0×Ω,p(t,x,ω)n=ρ(t,x,ω)(g(t,x,ω)u(t,x,ω)), on [0,T]×D1×Ω,u(0,x,ω)=u0(x,ω), in D×Ω. (1.1)

    Here, DRd (d=2,3) is a Lipschitz domain with boundary D=D0D1, where D0 and D1 are two non-overlapping parts. The vector n is the unit outward normal to D, and Ω denotes the sample space.

    Many problems in engineering and physics involve uncertainty. For example, in heat transfer dynamics, material properties, Robin boundary conditions (convective heat transfer coefficient), thermal conductivity (diffusion coefficient), etc., are affected by various forms of uncertainty. These uncertainties make the accurate prediction and analysis more challenging. For these problems, many numerical approaches have been devised [1,2]. In addition to stochastic finite element methods [3], stochastic collocation methods [4,5,6,7], and polynomial chaos methods [8], the Monte Carlo (MC) method is widely regarded as another very important approach (see [9,10,11,12]). The advantages of the MC method are that its convergence rate is independent of the dimensionality of the random parameters, and it is easy to implement. For the MC method, we first perform M independent and identically distributed (i.i.d.) samples of random variable ω. The i-th (i=1,2,,M) realization ωi satisfies

    {κi(t,x)pi(t,x)ui(t,x)=0, in [0,T]×D,ui(t,x)tpi(t,x)=fi(t,x), in [0,T]×D,pi(t,x)n=0, on [0,T]×D0,pi(t,x)n=ρi(t,x)(gi(t,x)ui(t,x)), on [0,T]×D1,ui(0,x)=u0i(x), in D, (1.2)

    where we assume that ρi(t,x)=ρ(t,x,ωi) and κi(t,x)=κ(t,x,ωi) for i=1,2,,M, with similar expressions to the other variables. The average

    1MMi=1ui(t,x)

    of the solution for problem (1.2) is used as an approximation of the solution for problem (1.1).

    Although the MC method is simple and easy to implement, its convergence speed is very slow. To improve this method, Jiang and Layton [13] proposed an ensemble approach for the random evolution equation. Since its proposal, this method has been widely promoted and applied [12,14,15,16,17,18]. In [16], a parabolic problem with a random diffusion coefficient is solved by the ensemble MC and finite element method. The error estimate is not optimal in space. To improve this, Yong et al. in [19] presented an optimal error estimate using the finite element method with ensemble MC. For the same problem as [16], Li and Luo used the ensemble MC and HDG method to approximate it and obtain the optimal error estimate about space. Similarly, for a parabolic problem with random diffusion and Robin coefficients, Yao et al. [18] used the ensemble MC and finite element method to approximate it and find the sub-optimal error estimate in space.

    The discontinuous Galerkin (DG) method is an excellent approximation method for problems concerned with partial differential equations (PDEs). The DG method is particularly well-suited for problems with discontinuous coefficients. However, the main disadvantage of DG methods is the large number of degrees of freedom. The HDG method (see, e.g., [20]) approximates the solution's flux and trace by introducing numerical fluxes and numerical traces in a mixed formulation. Compared to DG methods, HDG methods significantly reduce the number of globally coupled degrees of freedom. To date, the HDG method has been utilized for a variety of problems, including convection-diffusion problems, flow problems, and hyperbolic equations (see, e.g., [15,21,22,23]).

    In this work, we study the numerical approximate of a parabolic problem with random diffusion and Robin coefficients by the ensemble MC and HDG methods. By introducing the flux p, the parabolic problem with random Robin coefficients and diffusion coefficients can be transformed into problem (1.1). After using MC sampling, the problem involves solving a large number of problems (1.2). Introduce two averages for the Robin coefficient and diffusion coefficient, respectively, and construct an ensemble format. Then in the calculation of each time step, a coefficient matrix can be shared, which reduces the computational complexity (the details can be seen in Eq (3.12)). The FEMC-HDG method has first-order temporal accuracy and optimal L2 convergence in spatial space.

    The paper is structured as follows: In Section 2, the necessary notations and preliminaries are provided. In Section 3, we introduce the fully discrete HDG ensemble scheme for problem (1.2), along with a comprehensive analysis of its stability and convergence. In Section 4, two numerical examples are provided to illustrate the effectiveness of the proposed method. Finally, concluding remarks are presented.

    In this section, we introduce some notations that will be used throughout this work.

    Let (,) and be the inner product and the L2(D) norm, respectively. We adopt the standard Sobolev space notation Ws,q(D), with the corresponding norm Ws,q(D) and seminorm ||Ws,q(D), where s0 and 1q. For convenience, we use Hs(D):=Ws,2(D). Specifically, the norm Hs(D) and semi-norm ||Hs(D) correspond to Hs(D).

    We assume that (Ω,F,P) is a complete probability space, where F2Ω represents the σ-algebra of events, and P:F[0,1] denotes the probability measure. HL1P(Ω) is a random variable. The expectation of H can be expressed as

    E[H]:=ΩH(ω)dP(ω).

    Consider a d-tuple δ=(δ1,,δd) with length |δ|=di=1δi, where each δiN+. Let the space ˜Ws,p(D)=LpP(Ω,Ws,p(D)) consist of random functions u:Ω×DR, which are measurable with respect to the σ-algebra FB(D), where B(D) denotes the Borel σ-algebra on D. The norm in ˜Ws,p(D) is defined as

    u˜Ws,p(D):=(E[upWs,p(D)])1/p=(E[|δ|sD|δu|pdx])1/p,1p<+.

    If u˜Ws,p(D), then for almost every ω, u(ω,)Ws,p(D). Additionally, for |δ|s, the derivatives δu(,x) are in LpP(Ω) for almost every point on the domain D. Lastly, we define ˜Hs(D)=L2P(Ω,Hs(D)).

    We introduce the tensor product Hilbert space

    X:=˜L2(0,T;H1(D))L2P(0,T;H1(D);Ω),

    where the inner product is defined as

    (u,w)X:=E[T0D(uw+uw)dxdt].

    The norm is expressed as

    uX:=(E[T0D(|u|2+u2)dxdt])1/2.

    Let Th be a quasi-uniform triangulation of D. ¯D0¯D0 are the triangulation nodes. Thus, D can be written as the union D=KThK, where each element K has a diameter hK. We further define h as h=maxKThhK. The set Th is defined as {K:KTh}, and Fh represents the collection of faces F from elements KTh. The space of polynomials of degree on K is denoted by P(K). We then introduce the following discontinuous finite element spaces:

    Vh:={v[L2(D)]d:v|K[P(K)]d,KTh},Wh:={wL2(D):w|KP(K),KTh},Mh:={μL2(Fh):μ|FP(F),FFh}.

    The inner products are defined as follows:

    (v,w)Th:=KTh(v,w)K,   ζ,ρTh:=KThζ,ρK,   ζ,ρD1:=KThζ,ρKD1,

    where (v,w)D:=Dvwdx, and v,wD:=Dvwds. So we define w2Th=(w,w)Th.

    We assume that the constant C is positive and changes throughout the paper for different occurrences. Crucially, it does not depend on the discrete parameters M, Δt, or h.

    In this section, we begin with outlining several assumptions, then develop an FEMC-HDG scheme for problem (1.2), and proceed to design the FEMC-HDG algorithm. Following this, we present results on stability and error estimates for both the FEMC-HDG scheme and the FEMC-HDG algorithms.

    The mean values of the Robin boundary conditions and diffusion coefficients across the ensemble can be defined as follows:

    ˉρ(t,x):=1MMi=1ρi(t,x), (3.1)

    and

    ˉκ(t,x):=1MMi=1κi(t,x), (3.2)

    respectively. In order to derive the stability and error estimates for the FEMC-HDG algorithm, we draw on the work of [16] and make the following two assumptions:

    (H1): Let κmax, κmin, ρmin, and ρmax be four positive constants such that for any t[0,T], the probability is expressed as

    P{ωΩ;minxˉDκ(t,x,ω)>κmin}=1, (3.3)
    P{ωΩ;maxxˉDκ(t,x,ω)<κmax}=1, (3.4)

    and

    P{ωΩ;minx¯D1ρ(t,x,ω)>ρmin}=1, (3.5)
    P{ωΩ;maxx¯D1ρ(t,x,ω)<ρmax}=1. (3.6)

    (H2): Let κ+, κ, ρ+, and ρ be four positive constants, such that for every t[0,T], the probability is stated as follows:

    P{ωiΩ;κ|κ(t,x,ωi)ˉκ|,Dκ+}=1, (3.7)

    and

    P{ωiΩ;ρ|ρ(t,x,ωi)ˉρ|,D1ρ+}=1. (3.8)

    Hypothesis (H1) ensures almost sure uniform coercivity, while hypothesis (H2) imposes a uniform bound on |κ(t,x,ωi)ˉκ(t,x)| with a probability close to certainty. Similar properties are also assumed to hold for ρ(t,x,ωi).

    For a discretized physical space, problem (1.2) seeks (pih(t,),uih(t,),ˆuih(t,))Vh×Wh×Mh such that, for all (rh,vh,μh)Vh×Wh×Mh,

    {(κipih,rh)Th+(uih,rh)Thˆuih,rhnTh=0,(tuih,wh)Th+(pih,wh)Thˆpihn,whTh=(fi,wh)Th,ˆpihn,μhTh+ρiˆuih,μhD1=ρigi,μhD1,i=1,2,,M. (3.9)

    The numerical flux ˆpih is defined by:

    ˆpih:=pihτ(uihˆuih)n, on Th, (3.10)

    where τ represents the stabilization parameter. In this work, we set τ=1 since we do not address multiple physical scales. Substituting Eq (3.10) into Eq (3.9), we obtain the semidiscrete HDG scheme of problem (1.2), which seeks (pih(t,),uih(t,),ˆuih(t,))Vh×Wh×Mh such that, for all (rh,vh,μh)Vh×Wh×Mh, i=1,2,,M,

    {(κipih,rh)Th+(uih,rh)Thˆuih,rhnTh=0,(tuih,wh)Th+(pih,wh)Thpihn,whTh+τ(uihˆuih),whTh=(fi,wh)Th,pihn,μhThτ(uihˆuih),μhTh+ρiˆuih,μhD1=ρigi,μhD1. (3.11)

    We divide the interval [0,T] into N equal parts, with a length of Δt for each part. Denote tn=nΔt for n=1,2,,N. At each time t=tn, the functions ui, fi, gi, ρi, κi, ˉρ, and ˉκ are denoted by uni, fni, gni, ρni, κni, ˉρn, and ˉκn, respectively. Using the backward Euler method to approximate the time derivative of Eq (3.11), we can get the fully discrete non-ensemble MC HDG (FNEMC-HDG) scheme. Here, we omit the details. By applying Eqs (3.1) and (3.2) to (3.11) and performing simple algebraic calculations, we can obtain the FEMC-HDG scheme: Seek (pnih,unih,ˆunih)Vh×Wh×Mh such that, for all (rh,vh,μh)Vh×Wh×Mh,

    (unih,rh)Th+(ˉκnpnih,rh)Thˆunih,rhnTh=((κniˉκn)pn1ih,rh)Th,n=1,2,N, (3.12a)
    (unihun1ihΔt,wh)Th(pnih,wh)Th+pnihn,μhTh+τ(unihˆunih),whμhTh+ˉρnˆunih,μhD1=(fni,wh)Th+ρnigni,μhD1(ρniˉρn)ˆun1ih,μhD1,n=1,2,N. (3.12b)

    The initial conditions are given by u0ih=Π+1u0i, p0ih=1κiu0ih, and ˆu0ih=ˆΠu0i, where Π+1 and ˆΠ represent the L2 projection operators Π+1:L2(K)P+1(K) and ˆΠ:L2(F)P(F), respectively, which satisfy:

    (Π+1w,vh)K=(w,vh)K,vhP+1(K), (3.13a)
    ˆΠ+1w,μhF=w,μhF,μhP+1(F). (3.13b)

    To solve the stochastic partial differential equation (SPDE) (1.1) using the FEMC-HDG scheme, we first employ the MC method to generate i.i.d. samples. Subsequently, we apply the FEMC-HDG scheme to solve the resulting Eq (1.2). The solution of Eq (1.2) is used to compute the expected value of the solution to the SPDE (1.1). We present the FEMC-HDG algorithm which consists of three main steps:

    Step 1. Generate an i.i.d. sample set for the initial conditions, boundary conditions, source term, diffusion coefficients, and Robin coefficients. For the i-th realization, these samples are denoted as u0i=u0(,ωi), gi=g(,,ωi), fi=f(,,ωi), κi=κ(,,ωi), and ρi=ρ(,,ωi), respectively.

    Step 2. Calculate ˉρn=1MMi=1ρ(tn,x,ωi) and ˉκn=1MMi=1κ(tn,x,ωi). For the i-th sample, solve Eq (3.12a) and (3.12b) to determine uni,h and pni,h for n=1,,N.

    Step 3. For n=1,,N, compute 1MMi=1unih, 1MMi=1pnih to approximate the expectation E[un], E[pn], respectively.

    The FEMC-HDG algorithm effectively combines the scheme (3.12a)–(3.12b) with the random sampling method. It preserves the benefits of the ensemble approach used for deterministic PDEs, such as employing the same coefficient matrix for all simulations at each time step. As a result, this approach involves solving a linear system with several right-hand side vectors, significantly reducing computational expenses (refer to [16]).

    In this section, we carry out the stability analysis and error estimates.

    The FEMC-HDG scheme (3.12a)–(3.12b) has the following stability results.

    Theorem 4.1. Assume that fi˜L2(0,T;L2(D)), gi˜L2(0,T;L2(D1)), and u0i˜L2(H1(D)), and that hypotheses (H1) and (H2) are met. The FEMC-HDG scheme (3.12a)–(3.12b) is stable if

    κminκ+>0,    and    ρminρ+>0. (4.1)

    Moreover, the numerical solution to (3.12a)–(3.12b) satisfies the following inequality:

    max1nNE[unih2Th]+Nn=1E[unihun1ih2Th]+ΔtNn=1E[τ(unihˆunih)2Th]+(κminκ+)ΔtNn=1E[pnih2Th]+κminΔtmax1nNE[pnih2Th]+ρminρ+2ΔtNn=1E[ˆunih2D1]+μminΔtmax1nNE[ˆunih2D1]C(ΔtNn=1E[gni2D1]+ΔtNn=1E[fni2Th]+E[u0ih2Th]+ΔtE[ˆu0ih2D1]+ΔtE[p0ih2Th]). (4.2)

    Proof. By selecting (rh,wh,μh)=(pnih,unih,ˆunih) in scheme (3.12a)–(3.12b), we obtain

    (ˉκnpnih,pnih)Th+(unih,pnih)Thˆunih,pnihnTh=((κniˉκn)pn1ih,pnih)Th, (4.3)
    (unihun1ihΔt,unih)Th(pnih,unih)Thpnihn,ˆunihTh+τ(unihˆunih),unihˆunihTh+ ˉρnˆunih,ˆunihD1=(fni,unih)Th+ρnigni,ˆunihD1(ρniˉρn)ˆun1ih,ˆunihD1. (4.4)

    Applying the polarization identity (ab)a=12[a2b2+(ab)2] to Eqs (4.3) and (4.4), and then integrating over the probability space, we derive that

    12Δt(E[unih2Th]E[un1ih2Th])+12ΔtE[unihun1jh2Th]+E[τ(unihˆunih)2Th]+E[(ˉκnpnih,pnih)Th]+E[ˉρnˆunih,ˆunihD1]=E[((κniˉκn)pn1ih,pnih)Th]E[(ρniˉρn)ˆun1ih,ˆunihD1]+E[(fni,unih)Th]+E[ρnigni,ˆunihD1]. (4.5)

    By utilizing the condition (3.6), and applying Young's inequality and the Cauchy-Schwarz inequality on the right-hand side of Eq (4.5), we obtain for any ϵi>0 (i=1,2,3) that

    E[((κniˉκn)pn1ih,pnih)Th]E[(ρniˉρn)ˆun1ih,ˆunihD1]E[|κniˉκn|,Dpn1ihThpnihTh]+E[|ρniˉρn|,D1ˆun1ihD1ˆunihD1]E[|κniˉκn|,D(12ϵ1pn1ih2Th+ϵ12pnih2Th)]+E[|ρniˉρn|,D1(12ϵ2ˆun1ih2D1+ϵ22ˆunih2D1)], (4.6)
    E[(fni,unih)Th]+E[ρnigni,ˆunihD1]]E[fniThunihTh]+ρmaxE[gniD1ˆunihD1]12E[fni2Th]+12E[unih2Th]+ρmax2ϵ3E[gni2D1]+ϵ32E[ˆunih2D1]. (4.7)

    By combining the estimates given in Eqs (4.6) and (4.7), applying the conditions (3.3), (3.5), (3.7), and (3.8), and multiplying Eq (4.5) by Δt, we find that for all 0<σ11,

    12(E[unih2Th]E[un1ih2Th])+12E[unihun1ih2Th]+ΔtE[τ(unihˆunih)2Th]+Δt(κminσ1κ+2ϵ1)E[pnih2Th]+Δt(ρminσ1ρ+2ϵ2ϵ32)E[ˆunih2D1]+Δtκmin(1σ1)(E[pnih2Th]E[pn1ih2Th])+Δt(κmin(1σ1)κ+2ϵ1)E[pn1ih2Th]+Δtρmin(1σ1)(E[ˆunih2D1]E[ˆun1ih2D1])+Δt(ρmin(1σ1)ρ+2ϵ2)E[ˆun1ih2D1]12ΔtE[unih2]+ρmax2ϵ3ΔtE[gni2D1]+Δt2E[fni2Th]. (4.8)

    According to the condition (4.1), choosing

    σ1=12,ϵ1=ϵ2=1,ϵ3=ρminρ+2,

    from Eq (4.8), we obtain

    12(E[unih2Th]E[un1ih2Th])+12E[unihun1ih2Th]+ΔtE[τ(unihˆunih)2Th]+Δtκminκ+2E[pnih2Th]+Δtρminρ+4E[ˆunih2D1]+Δtκmin2(E[pnih2Th]E[pn1ih2Th])+Δtκminκ+2E[pn1ih2Th]+Δtρmin2(E[ˆunih2D1]E[ˆun1ih2D1])+Δtρminρ+2E[ˆun1ih2D1]12ΔtE[unih2]+ρmaxρminρ+ΔtE[gni2D1]+Δt2E[fni2Th]. (4.9)

    Summing Eq (4.9) up from n=1 to n=N, we get

    max1nNE[unih2Th]+Nn=1E[unihun1ih2Th]+2ΔtNn=1E[τ(unihˆunih)2Th]+Δt(κminκ+)Nn=1E[pnih2Th]+Δtκminmax1nNE[pnih2Th]+Δtρminρ+2Nn=1E[ˆunih2D1]+Δtρminmax1nNE[ˆunih2D1]ΔtNn=1E[unih2Th]+ρmaxρminρ+ΔtNn=1E[gni2D1]+ΔtNn=1E[fni2Th]+E[u0ih2Th]+ΔtρminE[ˆu0ih2D1]+ΔtκminE[p0ih2Th]. (4.10)

    Finally, applying Gronwall's inequality to the inequality in (4.10) yields the desired result.

    Remark 4.1. According to condition (4.1), for the sequence {κi}Mi=1, the gap between κi and ˉκ should not exceed the fixed value of κmin. A similar condition applies to the sequence {ρi}Mi=1. If this criterion is not satisfied, it becomes imperative to partition the ensemble into multiple smaller groups and employ the FEMC-HDG algorithm on each subset. Throughout this procedure, upholding the stability requirement within each subgroup is crucial.

    Under the assumption of sufficiently smooth solutions to the heat conduction equation, we provide error estimates for the FEMC-HDG algorithm. Let ˉUnMh=1MMi=1unih and ˉQnMh=1MMi=1pnih denote the results obtained from the FEMC-HDG algorithm, which approximate the expected exact solutions. We now proceed to estimate the error E[ui(tn)]ˉUnMh and E[pi(tn)]ˉQnMh in various norms. For simplicity, we drop the subscript i in E[ui(tn)]ˉUnMh and E[pi(tn)]ˉQnMh and consider these errors in two separate terms:

    E[ui(tn)]ˉUnMh=(E[u(tn)]E[unh])+(E[unh]ˉUnMh)=Eunh+EunM, (4.11)
    E[pi(tn)]ˉQnMh=(E[p(tn)]E[pnh])+(E[pnh]ˉQnMh)=Epnh+EpnM. (4.12)

    The term Eunh=E[u(tn)unh] represents the combined error due to HDG and time discretization. The second term, EunM=E[unh]ˉUnMh, denotes the statistical error. Epnh and EpnM are analogous. Before deriving an error estimate for the FEMC-HDG algorithm, we undertake some preparatory work. First, we review standard error estimates for the L2 projections Π and ˆΠ.

    Lemma 4.1. [24] Assume 0. Then there exists a constant C, which is not dependent on KTh, such that

    uiΠuiKCh+1|ui|H+1(K),uiH+1(K), (4.13a)
    uiˆΠuiKCh+1|ui|H+32(K),ui∈∈H+32(K). (4.13b)

    Next, for any t[0,T], let (ΠVpi,ΠWui) represent the HDG projection of (pi,ui). Here, ΠVpi and ΠWui are the components of the HDG projection of pi and ui into Vh and Wh, respectively. For each KTh, the pair (ΠVpi,ΠWui) satisfies the following equations:

    (ΠVpi,r)K=(pi,r)K,r[P1(K)]d, (4.14a)
    (ΠWui,w)K=(ui,w)K,wP1(K), (4.14b)
    ΠVpinτΠWui,μF=pinτui,μF,μP(F), (4.14c)

    for every face F of the simplex K. When =0, the Eq (4.14a) and (4.14b) become trivial, and the projections are solely governed by Eq (4.14c). For these projections, they have the following property (Lemma 4.2), which is established in [20].

    Lemma 4.2. Suppose the polynomial degree meets the requirements 0 and τ>0. Then the system (4.14) has a unique solution (ΠVpi,ΠWui). Moreover, there exists a constant C>0, which is independent of both K and τ, such that

    ΠVpipiKC(hpi+1|pi|Hpi+1(K)+hui+1|u|Hui+1(K)),ΠWuiuiKC(hui+1|u|Hui+1(K)+hpi+1|pi|Hpi(K)),

    for ui, pi in [0,].

    Lemma 4.3. For every n=1,2,,N, the subsequent equations:

    (ˉκnΠVpni,rh)Th+(ΠWuni,rh)ThˆΠuni,rhnTh=((κniˉκn)ΠVpni,rh)Th+(κni(ΠVpnipni),rh)Th (4.15)

    and

    (ΠWuniΠWun1iΔt,wh)Th(ΠVpni,wh)Th+ΠVpnin,μhTh+τ(ΠWuniˆΠuni),whμhTh+ˉρnˆΠuni,μhD1=(fni,wh)Th+ρnigni,μhD1+(ΠWuniΠWun1iΔt,wh)Th(tuni,wh)Th+ρni(ˆΠuniuni),μhD1(ρniˉρn)ˆΠuni,μhD1 (4.16)

    hold for all (rh,vh,μh)Vh×Wh×Mh and i=1,2,,M.

    Proof. By the definition of ΠV in Eq (4.14a), ˆΠ in Eq (3.13b), and ΠW in Eq (4.14b), we get

    (ˉκnΠVpni,rh)Th+(ΠWuni,rh)ThˆΠuni,rhnTh=(ˉκnΠVpni,rh)Th+(uni,rh)Thuni,rhnTh=(ˉκnΠVpni,rh)Th(uni,rh)Th=(ˉκnΠVpni,rh)Th(κnipni,rh)Th=(ˉκnΠVpni,rh)Th(κniΠVpni,rh)Th+(κniΠVpni,rh)Th(κnipni,rh)Th.

    The initial identity has been established.

    Going forward, we shall demonstrate the second identity. First, by the definition of ΠW and ΠV in Eq (4.14c), we have

    ΠVpnin,μhThτ(ΠWuiˆΠuni),μTh+ρniˆΠuni,μhD1=ρni(ˆΠuniuni),μhD1+ρnigni,μhD1, (4.17)

    and

    (ΠVpni,wh)Th+τ(ΠWuniˆΠuni),whTh=(fni,wh)Th(tuni,wh)Th. (4.18)

    Adding Eqs (4.17) and (4.18) yields

    (ΠVpni,wh)Th+ΠVpnin,μhTh+τ(ΠWuniˆΠuni),whμTh+ρniˆΠuni,μhD1=(fni,wh)Th(tuni,wh)Th+ρnigni,μhD1+ρni(ˆΠuniuni),μhD1. (4.19)

    Therefore, combining with Eq (4.19), we have

    (ΠWuniΠWun1iΔt,wh)Th(ΠVpni,wh)Th+ΠVpnin,μhTh+τ(ΠWuniˆΠuni),whμhTh+ˉρnˆΠuni,μhD1=(fni,wh)Th+ρnigni,μhD1+(ΠWuniΠWun1iΔt,wh)Th(tuni,wh)Th+ρni(ˆΠuniuni),μhD1+ˉρnˆΠuni,μhD1ρniˆΠuni,μhD1.

    This proves the second identity.

    Subtracting the outcome of Lemma (4.3) from the FEMC-HDG system (3.12a)–(3.12b) yields the subsequent error equations.

    Lemma 4.4. Let ξunih=unihΠWuni, ξpnih=pnihΠVpni, and ξˆunih=ˆunihˆΠuni. Then we have that the following error equations:

    (ˉκnξpnih,rh)Th+(ξunih,rh)Thξˆunih,rhnTh=((κniˉκn)(ΠVpnipn1ih),rh)Th(κni(ΠVpnipni),rh)Th (4.20)

    and

    (ξunihξun1ihΔt,wh)Th(ξpnih,wh)Th+ξpnihn,μhTh+τ(ξunihξˆunih),whμhTh+ˉρnξˆunih,μhD1=(tuni,wh)Th(ΠWuniΠWun1iΔt,wh)Thρni(ˆΠuniuni),μhD1+(ρniˉρn)(ˆΠuniun1ih),μhD1 (4.21)

    hold for all (rh,vh,μh)Vh×Wh×Mh and i=1,2,,M.

    Here, we determine the estimates of Eunh, EunM, Epnh, and EpnM to establish an error assessment for the FEMC-HDG algorithm.

    Theorem 4.2. Let (uni,pni) and (unih,pnih) denote the solutions to systems (1.2) and (3.12) at time tn, respectively. Assume that fi˜L2(0,T;L2(D)), gi˜L2(0,T;L2(D1)), u0˜L2(H+2(D)), and that the condition (H1) and (H2) are met. Consequently, there exists a positive constant C such that

    max1nNEunh2+(κminκ+)ΔtNn=1Epnh2ThC(h2+2+Δt2), (4.22)

    provided the stability conditions κminκ+>0 and ρminρ+>0 are satisfied.

    Proof. First, by selecting (rh,wh,μh)=(ξpnih,ξunih,ξˆunih) in Eqs (4.20) and (4.21), we obtain

    (ˉκnξpnih,ξpnih)Th+(ξunih,ξpnih)Thξˆunih,ξpnihnTh=((κniˉκn)(ΠVpnipn1ih),ξpnih)Th(κni(ΠVpnipni),ξpnih)Th, (4.23)

    and

    (ξunihξun1ihΔt,ξunih)Th(ξpnih,ξunih)Th+ξpnihn,ξˆunihTh+τ(ξunihξˆunih),ξunihξˆunihTh+ˉρnξˆunih,ξˆunihD1=(tuni,ξunih)Th(ΠiuniΠiun1iΔt,ξunih)Thρni(ˆΠuniuni),ξˆunihD1+(ρniˉρn)(ˆΠuniun1ih),ξˆunihD1.

    By summing Eqs (4.23) and (4.24), applying the polarization identity, and integrating over the probability space, we derive

    12Δt(E[ξunih2Th]E[ξun1ih2Th])+12ΔtE[ξunihξun1ih2Th]+E[τ(ξunihξˆunih)2Th]+E[(ˉκnξpnih,ξpnih)Th]+E[ˉρnξˆunih,ξˆunihD1]=E[((κniˉκn)(ΠVpnipn1ih),ξpnih)Th]E[(κni(ΠVpnipni),ξpnih)Th]+E[(tuni,ξunih)Th(ΠWuniΠWun1iΔt,ξunih)Th]E[ρni(ˆΠuniuni),ξˆunihD1]+E[(ρniˉρn)(ˆΠuniun1ih),ξˆunihD1]. (4.24)

    Utilizing the Cauchy-Schwarz inequality, the conditions (3.6) and (3.4), and Young's inequality on the right-hand side of Eq (4.24), we obtain the following five inequalities for any ϵi>0 (i=1,2,,6):

    E[((κniˉκn)(ΠVpnipn1ih),ξpnih)Th]=E[((κniˉκn)(ΠV(pnipn1i)ξpn1ih),ξpnih)Th]=E[((κniˉκn)ΠV(pnipn1i),ξpnih)Th]E[((κniˉκn)ξpn1ih,ξpnih)Th]E[|κniˉκn|,DΠV(pnipn1i)ThξpnihTh]+E[|κniˉκn|,Dξpn1ihThξqnihTh]14ϵ1E[|κniˉκn|2,DΠV(pnipn1i)2Th]+ϵ1E[ξpnih2Th]+E[(|κniˉκn|,D)(12ϵ2ξpn1ih2Th+ϵ22ξpnih2Th)], (4.25)
    E[(κni(ΠVpnipni),ξpnih)Th]κmax4ϵ3E[ΠVpnipni2Th]+ϵ3E[ξpnih2Th], (4.26)
    E[(tuni,ξunih)Th(ΠWuniΠWun1iΔt,ξunih)Th]=E[(tuniuniun1iΔt,ξunih)]+1Δt(uniun1i(ΠWuniΠWun1i),ξunih)E[tuniuniun1iΔt2Th]+1Δt2E[uniun1i(ΠWuniΠWun1i)2Th]+12E[ξunih2Th], (4.27)
    E[ρni(ˆΠuniuni),ξˆunihD1]ρmaxE[ˆΠuniuniD1ξˆunihD1]ρ2max4ϵ4E[ˆΠuniuni2D1]+ϵ4E[ξˆunih2D1], (4.28)

    and

    E[(ρniˉρn)(ˆΠuniun1ih),ξˆunihD1]=E[(ρniˉρn)ˆΠ(uniun1i),ξˆunihD1]E[(ρniˉρn)ξˆun1ih,ξˆunihD1]14ϵ5E[|ρniˉρn|2,D1ˆΠ(uniun1i)2D1]+ϵ5E[ξˆunih2D1]+12ϵ6E[|ρniˉρn|2,D1ξˆun1ih2D1]+ϵ62E[|ρniˉρn|2,D1ξˆunih2D1]. (4.29)

    By substituting Eqs (4.25) through (4.29) into Eq (4.24), and additionally applying the conditions (3.7), (3.8), (3.3), and (3.5), we obtain the following for any 0<σ1<1:

    12(E[ξunih2Th]E[ξun1ih2Th])+12E[ξunihξun1ih2Th]+ΔtE[τ(ξunihξˆunih)2Th]+Δt(κminσ1ϵ22κ+ϵ1ϵ3)E[ξpnih2Th]+Δtκmin(1σ1)(E[ξpnih2Th]E[ξpn1ih2Th])+Δt(κmin(1σ1)κ+2ϵ2)E[ξpn1ih2Th]+Δt(ρminσ1ϵ62ρ+ϵ4ϵ5)E[ξˆunih2D1]+Δtρmin(1σ1)(E[ξˆunih2D1]E[ξˆun1ih2D1])+Δt(ρmin(1σ1)ρ+2ϵ6)E[ξˆun1ih2D1]Δtρ2+4ϵ1E[ΠV(pnipn1i)2Th]+Δtκmax4ϵ3E[ΠVpnipni2Th]+ΔtE[tuniuniun1iΔt2Th]+1ΔtE[uniun1i(ΠWuniΠWun1i)2Th]+Δt2E[ξunih2Th]+Δtρ2+4ϵ5E[ˆΠ(uniun1i)2D1]+Δtρ2max4ϵ4E[ˆΠuniuni2D1]. (4.30)

    According to the condition (4.1), we choose

    σ1=12,ϵ2=ϵ6=1,ϵ1=κminκ+4,ϵ3=κminκ+8,ϵ4=ρminρ+4,ϵ5=ρminρ+8.

    Then, from Eq (4.30), we obtain

    12(E[ξunih2Th]E[ξun1ih2Th])+12E[ξunihξun1ih2Th]+ΔtE[τ(ξunihξˆunih)2Th]+Δtκminκ+8E[ξpnih2Th]+Δtκmin2(E[ξpnih2Th]E[ξpn1ih2Th])+Δtκminκ+2E[ξqn1ih2Th]+ρminρ+8ΔtE[ξˆunih2D1]+Δtρmin2(E[ξˆunih2D1]E[ξˆun1ih2D1])+Δtρminρ+2E[ξˆun1ih2D1]Δtρ2+4(κminκ+)E[ΠV(pnipn1i)2Th]+Δt2κmaxκminκ+E[ΠVpnipni2Th]+ΔtE[tuniuniun1iΔt2Th]+1ΔtE[uniun1i(ΠWuniΠWun1i)2Th]+Δt2E[ξunih2Th]+Δt2ρ2+ρminρ+E[ˆΠ(uniun1i)2D1]+Δtρ2maxρminρ+E[ˆΠuniuni2D1]. (4.31)

    Additionally, the following estimates pertain to temporal errors:

    ΔtE[ˆΠ(uniun1i)2D1]=ΔtE[D1[tntn1tˆΠuidt]2ds]CΔt2E[tˆΠui2L2(tn1,tn;L2(D1))], (4.32)
    ΔtE[ΠV(pnipn1i)2Th]=ΔtE[D[tntn1tΠVpnidt]2dx]CΔt2E[tΠVpni2L2(tn1,tn;L2(D))], (4.33)

    and

    1ΔtE[uniun1i(ΠWuniΠWun1i)2Th]=Δt1E[D[tntn1t(uiΠWui)dt]2dx]CE[tntn1t(uiΠWui)2Thdt], (4.34)
    ΔtE[tuniuniun1iΔt2Th]=Δt1E[D[tntn1tntttui(s)dsdt]2dx]CΔt2E[ttui2L2(tn1,tn;L2(D))]. (4.35)

    Summing Eq (4.31) up from n=1 to N and using Eqs (4.32)–(4.35), we arrive at

    max1nNE[ξunih2Th]+Nn=1E[ξunihξun1ih2Th]+2ΔtNn=1E[τ(ξunihξˆunih)2Th]+κminκ+4ΔtNn=1E[ξqnih2Th]+ρminρ+4ΔtNn=1E[ξˆunih2D1]ΔtNn=1E[ξunih2Th]+ρ2+C2(κminκ+)Δt2tΠVpni2L2(0,T;L2(D))+4κmaxκminκ+ΔtNn=1E[ΠVpnipnipni2Th]+Δt2E[ttui2L2(0,T;L2(D))]+CE[tntn1t(uiΠWui)2Thdt]+4ρ2+ρminρ+Δt2E[tˆΠui2L2(0,T;L2(D1))]+ρ2maxρminρ+ΔtNn=1E[ˆΠuniuni2D1]. (4.36)

    Applying Gronwall's inequality to Eq (4.36), and utilizing Lemma 4.1 along with Lemma 4.2, we obtain

    max1nNE[ξunih2Th]+(κminκ+)ΔtNn=1E[ξpnih2Th]C(h2+2+Δt2). (4.37)

    Decomposing Eunh as

    E[uniunih]=E[uniΠWuni]E[ξunih],

    and Epnh as

    E[pnipnih]=E[pniΠVpni]ξpnih,

    and applying the triangle inequality, Jensen's inequality, and Lemma 4.2, we derive the expected result (4.22).

    Remark 4.2. As far as we know, the previous other research has only provided a suboptimal L2 convergence rate for ensemble solutions uih for the heat equation with a Robin coefficient. In contrast, our result in Eq (4.22) achieves the optimal L(0,T;L2(D)) convergence rate on a general polygonal domain D.

    The statistical errors EunM and EpnM can be obtained using the standard estimation method (refer to [1]).

    Theorem 4.3. Given that the condition (H1), (H2), and the stability condition (4.1) hold, along with fi˜L2(0,T;L2(D)), gi˜L2(0,T;L2(D1)), and u0i˜L2(H+2(D)), then there exists a positive constant C such that

    max1nNE[EunM2Th]+(κminκ+)ΔtNn=1E[EpnM2Th]CM(ΔtNn=1E[gni2D1]+ΔtNn=1E[fni2Th]+E[u0ih2Th]+ΔtE[ˆu0ih2D1]+ΔtE[p0ih2Th]). (4.38)

    Proof. We first analyze E[EunM2Th], for all 1nN. It is easy to see that

    E[EunM2Th]=E[(1MMi=1(E[unh]unih),1MMj=1(E[unh]unjh))Th]=1M2Mi=1Mj=1E[(E[unh]unih,E[unh]unjh)Th]. (4.39)

    Given that unh(,ω1),,unh(,ωM) are i.i.d., the expectation of (E[unh]unih,E[unh]unjh) is zero when ij. Thus, we obtain

    E[EunM2Th]=1M2Mi=1E[(E[unh]unih,E[unh]unih)Th].

    Let M=unh and ˉM=E[M], from which we can infer

    E[(MˉM,MˉM)Th]=E[M2Th2(M,ˉM)Th+ˉM2Th]=E[M2Th]ˉM2ThE[M2Th].

    As a result, we arrive at

    E[EunM2Th]1ME[unih2Th].

    Regarding E[EqnM2Th], the situation is similar. From Theorem 4.1, we derive the result (4.38).

    By combining the space error, time error, and the MC sampling error, we can derive the total error of the FEMC-HDG algorithm.

    Theorem 4.4. Let f˜L2(0,T;L2(D)), g˜L2(0,T;L2(D1)), and u0˜L2(H+2(D)). Assume that the condition (H1), (H2), and the stability condition (4.1) are satisfied. Then there exists a constant C>0 such that

    max1nNE[E[ui(tn)]ˉUnMh2Th]+(κminκ+)ΔtNn=1E[E[pi(tn)]ˉQnMh2Th]CM(ΔtNn=1E[gni2D1]+ΔtNn=1E[fni2Th]+E[u0ih2Th]+ΔtE[ˆu0ih2D1]+ΔtE[p0ih2Th])+C(h2+2+Δt2). (4.40)

    Proof. Employing Young's inequality along with the triangle inequality on the first term of the left-hand side of Eq (4.40), we obtain

    E[E[un]ˉUnMh2]2(E[E[un]E[unh]2]+E[E[unh]ˉUnMh2]).

    Applying Jensen's inequality to the first term on the right-hand side of the preceding inequality yields

    E[E[un]E[unh]2]E[E[ununh2]]=E[ununh2].

    Thus, by employing Theorem 4.1, Theorem 4.2, and Theorem 4.3, the desired outcome is obtained. Likewise, the term

    ΔtNn=1E[E[pi(tn)]ˉQnMh2Th]

    on the left-hand side of Eq (4.40) can be deduced in a similar way.

    Remark 4.3. The method outlined above attains first-order accuracy in time. Higher-order temporal accuracy can be achieved using numerical algorithms like the BDF(k) (backward differentiation formula of order k) scheme, with k2 [25,26,27], which can be adapted for use with ensemble algorithms for the uncertain heat equation. However, these may demand more stringent stability conditions than those outlined in Eq (4.1). A stable numerical scheme can be constructed using similar methods, and its convergence theory is also analogous.

    As defined in [28], the element-by-element postprocessing is as follows: Find unihP+1(K) such that for all

    (zh,wh)[P+1(K)]×P0(K),
    (unih,zh)K=(κnipnih,zh)K,(unih,wh)K=(unih,wh)K,

    where

    [P+1(K)]={zhP+1(K)|(zh,1)K=0},n=1,,N,i=1,,M.

    After such postprocessing, we can use 1MMi=1unih in place of 1MMi=1unih in the FEMC-HDG algorithm. From the numerical experiments presented later, it can be observed that, after such postprocessing, the discrete solution achieves a super-convergent rate under certain conditions on the domain. For instance, a convex domain is adequate.

    In this section, we verify Theorem 4.4 through numerical simulations and highlight the advantages of the FEMC-HDG algorithm over the FNEMC-HDG approach. In particular, we examine the heat equation (1.1) with random coefficients, defined on the unit square [0,1]2, with boundary conditions D0={x2=0}{x2=1} and D1={x1=0}{x1=1}. Our numerical simulations utilize the open-source software NGSolve [29], which can be accessed at https://ngsolve.org/.

    We conduct the experiment using the exact solution

    u(t,x,ω)=(1+ω)cos(2πx1)cos(2πx2)sin(t),

    where ω is uniformly distributed over [0,1] and t[0,1]. The coefficients are set as

    κ(x,ω)=ρ(x,ω)=8+(1+ω)cos(x1x2).

    The FEMC-HDG scheme is applied in this experiment to simulate the ensemble with M=30. Define

    EunE:=max1nN1MMi=1ui(tn)unih,E2Th,EpnE:=Nn=11MMi=1pi(tn)pnih,E2Th,
    EunN:=max1nN1MMi=1ui(tn)unih,N2Th,EpnN:=Nn=11MMi=1pi(tn)pnih,N2Th,
    EunE:=max1nN1MMi=1ui(tn)unih,E2Th,EunN:=max1nN1MMi=1ui(tn)unih,N2Th,

    where (unih,N,pnih,N) and (unih,E,pnih,E) denote the non-ensemble solution and ensemble solution, respectively. In the experiment, both time and space are partitioned uniformly. We use =0, where the time step is set to Δt=h. For =1, the time step is set to Δt=h3. The spatial step size h is varied from 121 to 125. The corresponding errors and convergence rates are shown in Table 1.

    Table 1.  Errors and convergence rates (Δt=h for =0; Δt=h3 for =1; M=30).
    (a) FEMC-HDG method
    Degree h2 EunE EpnE EunE
    Error Rate Error Rate Error Rate
    =0 21 1.72147E+00 6.63799E-01 8.17718E-01
    22 3.35280E-01 2.36 2.82738E-01 1.23 8.64218E-02 3.24
    23 1.73353E-01 0.95 1.45335E-01 0.96 4.98378E-02 0.79
    24 8.68186E-02 1.00 7.13858E-02 1.03 2.62287E-02 0.93
    25 4.51054E-02 0.94 3.66211E-02 0.96 1.43155E-02 0.87
    =1 21 1.51705E+00 4.56279E-01 5.81619E-01
    22 1.11939E-01 3.76 7.44930E-02 2.61 3.97105E-02 3.87
    23 2.47416E-02 2.18 1.89130E-02 1.98 4.88681E-03 3.02
    24 5.42893E-03 2.19 4.40070E-03 2.10 4.96631E-04 3.30
    25 1.36441E-03 1.99 1.11908E-03 1.98 6.44694E-05 2.95
    (b) FNEMC-HDG method
    Degree h2 EunN EpnN EunN
    Error Rate Error Rate Error Rate
    =0 21 1.72230E+00 6.63280E-01 8.18437E-01
    22 3.35085E-01 2.36 2.82734E-01 1.23 8.59583E-02 3.25
    23 1.73201E-01 0.95 1.45360E-01 0.96 4.94796E-02 0.80
    24 8.67049E-02 1.00 7.13961E-02 1.03 2.59940E-02 0.93
    25 4.50339E-02 0.95 3.66238E-02 0.96 1.41820E-02 0.87
    =1 21 1.51785E+00 4.56149E-01 5.81860E-01
    22 1.11923E-01 3.76 7.44951E-02 2.61 3.96944E-02 3.87
    23 2.47406E-02 2.18 1.89131E-02 1.98 4.88301E-03 3.02
    24 5.42888E-03 2.19 4.40070E-03 2.10 4.96169E-04 3.30
    25 1.36440E-03 1.99 1.11908E-03 1.98 6.44069E-05 2.95

     | Show Table
    DownLoad: CSV

    From Table 1 it can be seen that when Δt=h and =0, the convergence rate of both the FEMC-HDG and FNEMC-HDG schemes is O(h+Δt)=O(h). Additionally, when Δt=h3 and =1, the convergence rate for both schemes is O(h2+Δt)=O(h2). Further observations reveal that, under the same spatial discretization parameter h, the errors for both the non-ensemble and ensemble schemes are of the same magnitude. Notably, after postprocessing, the convergence rate of uih reaches O(h3+Δt)=O(h3) when Δt=h3, which demonstrates spatial super-convergence.

    We chose a sample size of M=100, with a mesh size of h=126, =1, and a time step Δt=126. The mean solution at t=0.5 is calculated. The results are displayed in Figure 1 (left). To assess the efficiency of the FEMC-HDG algorithm, we compare its results with those from simulations using the FNEMC-HDG algorithm, employing the same sample values. The difference in the mean solutions between FEMC-HDG and FNEMC-HDG is depicted in Figure 1 (right).

    Figure 1.  FEMC-HDG algorithm simulations. Left: Mean. Right: Difference between the mean of FEMC-HDG and FNEMC-HDG.

    The difference between the FEMC-HDG and FNEMC-HDG algorithms is around 106, which indicates that the FEMC-HDG algorithm delivers a similar level of accuracy to that of the FNEMC-HDG algorithm. This suggests that the FEMC-HDG approach performs comparably in terms of precision.

    For further evaluation of the FEMC-HDG algorithm, we adopt piecewise constant elements (=0). We conduct a comparative analysis with the FNEMC-HDG algorithm. For a fair comparison, results from both algorithms are obtained under the identical conditions that the same time step and spatial dimensions are adopted. As illustrated in Table 2, the evaluations are performed with a mesh size Δt=124, and a time size of h=124. Given the moderate size of the discrete system, LU decomposition was employed.

    Table 2.  CPU time (s) (Δt=h=124,=0).
    M 30 60 120 240 480
    FNEMC-HDG 332.62 523.69 974.76 1672.21 4625.02
    FEMC-HDG 91.29 171.33 346.98 602.20 1029.98

     | Show Table
    DownLoad: CSV

    From Table 2, it is evident that the FEMC-HDG algorithm surpasses the FNEMC-HDG algorithm in terms of CPU time. The FEMC-HDG algorithm improves computational efficiency by approximately 70% compared to the FNEMC-HDG algorithm.

    Next, we investigate the convergence rate of MC for the FEMC-HDG algorithm. Using the FEMC-HDG algorithm, we calculate the mean of the solution with M0=12,000 samples as the reference benchmark, where =0 and h=Δt=124.

    By adjusting the values of M in the FEMC-HDG simulations, we assess the approximation errors to the benchmark. Furthermore, this error analysis is conducted for J=10 independent runs, and the mean of the resulting errors is calculated.

    Let Un,mM,h represent the FEMC-HDG solution at time tn for the m-th independent trial, given by

    Un,mM,h:=1MMi=1un,mih,

    where un,mih is the result of the m-th experiment using the FEMC-HDG scheme. We also define the error measure

    EMC:=max1nN1JJm=1UnM0,hUn,mM,h2.

    We performed 10 independent runs and reported the experimental results of EMC for M=10,30,90, and 270 in Figure 2. The convergence rate observed in the experiments with respect to M is approximately 0.5, which is consistent with the results derived in Theorem 4.4.

    Figure 2.  Convergence rate of the MC sample.

    In this study, we introduced a fully discrete HDG ensemble Monte Carlo algorithm that effectively tackles Robin boundary conditions and stochastic diffusion in heat equations. In the preparatory phase, the coefficient matrix for the linear systems is computed and preserved once, albeit requiring substantial computational effort. This methodology facilitates the swift assembly of linear systems in the operational phase, independent of the ensemble size. This marks the inaugural application of such an algorithm to address heat problems characterized by random diffusion and Robin coefficients. The performance of this method has been thoroughly validated and assessed.

    The model described by Eq (1.1) is also relevant in robust optimal boundary control challenges as state equations (see [30,31]). Efficient resolution of the model in Eq (1.1) is paramount for numerically addressing the associated optimal Robin boundary control issues. Both theoretical analyses and numerical tests confirm the efficacy of the FEMC-HDG algorithm in resolving (1.1). Additionally, the FEMC-HDG algorithm extends to nonlinear random heat equations, demanding further exploration into stability conditions. Therefore, it is pertinent to investigate additional applications of the FEMC-HDG algorithm in random contexts, which we highlight as a valuable direction for subsequent studies.

    Jinjun Yong: Methodology, analysis and writing original draft; Changlun Ye: Software and analysis; Xianbing Luo: Review, editing, and supervision.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work is partially supported by the National Natural Science Foundation of China (12461076) and funding from the Scientific Research Fund Project of Guizhou Education University (2024ZD007).

    The authors declare there is no conflict of interest.



    [1] K. Liu, B. M. Riviere, Discontinuous Galerkin methods for elliptic partial differential equations with random coefficients, Int. J. Comput. Math., 90 (2013), 2477–2490. https://doi.org/10.1080/00207160.2013.784280 doi: 10.1080/00207160.2013.784280
    [2] G. J. Lord, C. E. Powell, T. Shardlow, An Introduction to Computational Stochastic PDEs, New York: Cambridge University Press, 2014. https://doi.org/10.1017/CBO9781139017329
    [3] M. Gunzburger, C. G. Webster, G. Zhang, Stochastic finite element methods for partial differential equations with random input data, Acta Numer., 23 (2014), 521–650. https://doi.org/10.1017/S0962492914000075 doi: 10.1017/S0962492914000075
    [4] I. Babuska, F. Nobile, R. Tempone, A stochastic collocation method for elliptic partial differential equations with random input data, SIAM Rev., 52 (2010), 317–355. https://doi.org/10.1137/100786356 doi: 10.1137/100786356
    [5] B. Ganapathysubramanian, N. Zabaras, Sparse grid collocation schemes for stochastic natural convection problems, J. Comput. Phys., 225 (2007), 652–685. https://doi.org/10.1016/j.jcp.2006.12.014 doi: 10.1016/j.jcp.2006.12.014
    [6] D. Xiu, J. S. Hesthaven, High-order collocation methods for differential equations with random inputs, SIAM J. Sci. Comput., 27 (2005), 1118–1139. https://doi.org/10.1137/040615201 doi: 10.1137/040615201
    [7] X. Zhu, E. M. Linebarger, D. Xiu, Multi-fidelity stochastic collocation method for computation of statistical moments, J. Comput. Phys., 341 (2017), 386–396. https://doi.org/10.1016/j.jcp.2017.04.022 doi: 10.1016/j.jcp.2017.04.022
    [8] L. Mathelin, M. Y. Hussaini, T. A. Zang, Stochastic approaches to uncertainty quantification in CFD simulations, Numer. Algorithms., 38 (2005), 209–236. https://doi.org/10.1007/BF02810624 doi: 10.1007/BF02810624
    [9] G. Fishman, Monte Carlo: Concepts, Algorithms, and Applications, New York: Springer, 1996. https://doi.org/10.1007/978-1-4757-2553-7
    [10] M. B. Giles, Multilevel monte carlo methods, Acta Numer., 24 (2015), 259–328. https://doi.org/10.1017/S096249291500001X doi: 10.1017/S096249291500001X
    [11] J. C. Helton, F. J. Davis, Latin hypercube sampling and the propagation of uncertainty in analyses of complex systems, Reliab. Eng. Syst. Safe., 81 (2003), 23–69. https://doi.org/10.1016/S0951-8320(03)00058-9 doi: 10.1016/S0951-8320(03)00058-9
    [12] Y. Luo, Z. Wang, A multilevel Monte Carlo ensemble scheme for solving random parabolic PDEs, SIAM J. Sci. Comput., 41 (2019), A622–A642. https://doi.org/10.1137/18M1174635 doi: 10.1137/18M1174635
    [13] N. Jiang, W. Layton, An algorithm for fast calculation of flow ensembles, Int. J. Uncertain. Quan., 4 (2014), 273–301. https://doi.org/10.1615/Int.J.UncertaintyQuantification.2014007691 doi: 10.1615/Int.J.UncertaintyQuantification.2014007691
    [14] N. Jiang, A higher order ensemble simulation algorithm for fluid flows, J. Sci. Comput., 64 (2015), 264–288. https://doi.org/10.1007/s10915-014-9932-z doi: 10.1007/s10915-014-9932-z
    [15] M. Li, X. Luo, An MLMCE-HDG method for the convection diffusion equation with random diffusivity, Comput. Math. with Appl., 127 (2022), 127–143. https://doi.org/10.1016/j.camwa.2022.10.002 doi: 10.1016/j.camwa.2022.10.002
    [16] Y. Luo, Z. Wang, An ensemble algorithm for numerical solutions to deterministic and random parabolic PDEs, SIAM J. Numer. Anal., 56 (2018), 859–876. https://doi.org/10.1137/17M1131489 doi: 10.1137/17M1131489
    [17] T. Yao, C. Ye, X. Luo, S. Xiang, An ensemble scheme for the numerical solution of a random transient heat equation with uncertain inputs, Numer. Algorithms, 94 (2023), 643–668. https://doi.org/10.1007/s11075-023-01514-z doi: 10.1007/s11075-023-01514-z
    [18] C. Ye, T. Yao, H. Bi, X. Luo, A variational Crank-Nicolson ensemble Monte Carlo algorithm for a heat equation under uncertainty, J. Comput. Appl. Math., 451 (2024), 116068. https://doi.org/10.1016/j.cam.2024.116068 doi: 10.1016/j.cam.2024.116068
    [19] J. Yong, C. Ye, X. Luo, S. Sun, Improved error estimates of ensemble Monte Carlo methods for random transient heat equations with uncertain inputs, Comp. Appl. Math, 44 (2025), 58. https://doi.org/10.1007/s40314-024-03022-9 doi: 10.1007/s40314-024-03022-9
    [20] S. Du, F. J. Sayas, An Invitation to the Theory of the Hybridizable Discontinuous Galerkin Method: Projections, Estimates, Tools, Cham: Springer, 2019. https://doi.org/10.1007/978-3-030-27230-2
    [21] B. Cockburn, F. J. Sayas, Divergence-conforming HDG methods for Stokes flows, Math. Comp., 83 (2014), 1571–1598. https://doi.org/10.1090/S0025-5718-2014-02802-0 doi: 10.1090/S0025-5718-2014-02802-0
    [22] S. Rhebergen, B. Cockburn, J. J. W. Van Der Vegt, A space-time discontinuous Galerkin method for the incompressible Navier-Stokes equations, J. Comput. Phys., 233 (2013), 339–358. https://doi.org/10.1016/j.jcp.2012.08.052 doi: 10.1016/j.jcp.2012.08.052
    [23] M. Stanglmeier, N. C. Nguyen, J. Peraire, B. Cockburn, An explicit hybridizable discontinuous Galerkin method for the acoustic wave equation, Comput. Methods Appl. Mech. Eng., 300 (2016), 748–769. https://doi.org/10.1016/j.cma.2015.12.003 doi: 10.1016/j.cma.2015.12.003
    [24] S. C. Brenner, L. R. Scott, The Mathematical Theory of Finite Element Methods, New York: Springer, 2008. https://doi.org/10.1007/978-0-387-75934-0
    [25] Y. H. Hao, Q. M. Huang, C. Wang, A third order BDF energy stable linear scheme for the no-slope-selection thin film model, Commun. Comput. Phys., 29 (2021), 905–929. https://doi.org/10.4208/cicp.OA-2020-0074 doi: 10.4208/cicp.OA-2020-0074
    [26] Z. Y. Li, H. L. Liao, Stability of variable-step BDF2 and BDF3 methods, SIAM J. Numer. Anal., 60 (2022), 2253–2272. https://doi.org/10.1137/21M1462398 doi: 10.1137/21M1462398
    [27] K. L. Zheng, C. Wang, S. M. Wise, Y. Wu, A third order accurate in time, BDF-type energy stable scheme for the Cahn-Hilliard equation, Numer. Math. Theor. Meth. Appl., 15 (2022), 279–303. https://doi.org/10.4208/nmtma.OA-2021-0165 doi: 10.4208/nmtma.OA-2021-0165
    [28] B. Cockburn, J. Gopalakrishnan, F. J. Sayas, A projection-based error analysis of HDG methods, Math. Comp., 79 (2010), 1351–1367. https://doi.org/10.1090/S0025-5718-10-02334-3 doi: 10.1090/S0025-5718-10-02334-3
    [29] J. Schöberl, C++ 11 Implementation of Finite Elements in NGSolve, Institute for Analysis and Scientific Computing, Vienna University of Technology, 2014.
    [30] J. Martinez-Frutos, M. Kessler, A. Mnch, F. Periago, Robust optimal Robin boundary control for the transient heat equation with random input data, Int. J. Numer. Methods Eng., 108 (2016), 116–135. https://doi.org/10.1002/nme.5210 doi: 10.1002/nme.5210
    [31] J. Martinez-Frutos, F. P. Esparza, Optimal Control of PDEs Under Uncertainty: An introduction with Application to Optimal Shape Design of Structures, Cham: Springer, 2018. https://doi.org/10.1007/978-3-319-98210-6
  • Reader Comments
  • © 2025 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(627) PDF downloads(33) Cited by(0)

Figures and Tables

Figures(2)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog