Loading [MathJax]/extensions/TeX/boldsymbol.js
Research article

Breast conserving surgery (BCS) with adjuvant radiation therapy showed improved prognosis compared with mastectomy for early staged triple negative breast cancer patients

Running title: BCS had better prognosis than mastectomy for early TNBC patients
  • BackgroundTriple-negative breast cancer (TNBC) is a subtype of breast cancer with stronger invasive capacity. For the operation strategies of early staged (stage Ⅰ and stage Ⅱ) TNBC patients, BCS plus radiotherapy (BCS+RT), mastectomy only (MRM only) or MRM plus radiotherapy (MRM+RT) is feasible, but no clear conclusion has been made on the choice of these treatments.
    MethodsThe early staged TNBC patients (stage Ⅰ and stage Ⅱ) from the Surveillance, Epidemiology and End Results (SEER) program database between 1973 and 2014 were included in the study. Survival curves, univariate and multivariate cox proportional hazards models and propensity score weighting were applied to evaluate the prognostic impact among BCS+RT, MRM only and MRM+RT for patients.
    ResultsBoth overall and cancer-specific survival analysis showed that BCS+RT had better prognostic effect than MRM and MRM+RT in the cohort of early-staged triple-negative breast cancer patients (overall survival, P < 0.001; cancer-specific survival, P < 0.001). By taking all the risk factors into a multivariate cox proportional model, MRM and MRM+RT remained to have detrimental effect on the prognosis compared with BCS+RT as shown by either overall (HR = 1.742, CI = 1.387-2.188, P < 0.001; HR = 1.449, CI = 1.038-2.204, P = 0.029) or cancer-specific survival (HR = 1.876, CI = 1.415-2.489, P < 0.001; HR = 1.701, CI = 1.168-2.478, P = 0.006). After we performed propensity score weighting and integrated the weights for each covariate in the multivariate cox proportional model. BCS+RT remained to be prognostic beneficial compared to the other treatment options (P < 0.001).
    ConclusionBCS+RT demonstrated better prognosis than MRM only and MRM+RT treatments for early-staged TNBC patients.

    Citation: Shuoer Wang, Yidi Sun, Songjiao Zhao, Feng Wei, Gong Yang. Breast conserving surgery (BCS) with adjuvant radiation therapy showed improved prognosis compared with mastectomy for early staged triple negative breast cancer patients[J]. Mathematical Biosciences and Engineering, 2020, 17(1): 92-104. doi: 10.3934/mbe.2020005

    Related Papers:

    [1] JinJun Yong, Changlun Ye, Xianbing Luo . A fully discrete HDG ensemble Monte Carlo algorithm for a heat equation under uncertainty. Networks and Heterogeneous Media, 2025, 20(1): 65-88. doi: 10.3934/nhm.2025005
    [2] Serge Nicaise, Cristina Pignotti . Asymptotic analysis of a simple model of fluid-structure interaction. Networks and Heterogeneous Media, 2008, 3(4): 787-813. doi: 10.3934/nhm.2008.3.787
    [3] Seung-Yeal Ha, Shi Jin, Jinwook Jung . A local sensitivity analysis for the kinetic Kuramoto equation with random inputs. Networks and Heterogeneous Media, 2019, 14(2): 317-340. doi: 10.3934/nhm.2019013
    [4] Xiangdong Du, Martin Ostoja-Starzewski . On the scaling from statistical to representative volume element in thermoelasticity of random materials. Networks and Heterogeneous Media, 2006, 1(2): 259-274. doi: 10.3934/nhm.2006.1.259
    [5] Thomas Abballe, Grégoire Allaire, Éli Laucoin, Philippe Montarnal . Application of a coupled FV/FE multiscale method to cement media. Networks and Heterogeneous Media, 2010, 5(3): 603-615. doi: 10.3934/nhm.2010.5.603
    [6] Tasnim Fatima, Ekeoma Ijioma, Toshiyuki Ogawa, Adrian Muntean . Homogenization and dimension reduction of filtration combustion in heterogeneous thin layers. Networks and Heterogeneous Media, 2014, 9(4): 709-737. doi: 10.3934/nhm.2014.9.709
    [7] Martin Heida, Benedikt Jahnel, Anh Duc Vu . Regularized homogenization on irregularly perforated domains. Networks and Heterogeneous Media, 2025, 20(1): 165-212. doi: 10.3934/nhm.2025010
    [8] Tong Yan . The numerical solutions for the nonhomogeneous Burgers' equation with the generalized Hopf-Cole transformation. Networks and Heterogeneous Media, 2023, 18(1): 359-379. doi: 10.3934/nhm.2023014
    [9] Nimisha Pandey, Pramod Kumar Mishra . Detection of DDoS attack in IoT traffic using ensemble machine learning techniques. Networks and Heterogeneous Media, 2023, 18(4): 1393-1409. doi: 10.3934/nhm.2023061
    [10] Elisabeth Logak, Isabelle Passat . An epidemic model with nonlocal diffusion on networks. Networks and Heterogeneous Media, 2016, 11(4): 693-719. doi: 10.3934/nhm.2016014
  • BackgroundTriple-negative breast cancer (TNBC) is a subtype of breast cancer with stronger invasive capacity. For the operation strategies of early staged (stage Ⅰ and stage Ⅱ) TNBC patients, BCS plus radiotherapy (BCS+RT), mastectomy only (MRM only) or MRM plus radiotherapy (MRM+RT) is feasible, but no clear conclusion has been made on the choice of these treatments.
    MethodsThe early staged TNBC patients (stage Ⅰ and stage Ⅱ) from the Surveillance, Epidemiology and End Results (SEER) program database between 1973 and 2014 were included in the study. Survival curves, univariate and multivariate cox proportional hazards models and propensity score weighting were applied to evaluate the prognostic impact among BCS+RT, MRM only and MRM+RT for patients.
    ResultsBoth overall and cancer-specific survival analysis showed that BCS+RT had better prognostic effect than MRM and MRM+RT in the cohort of early-staged triple-negative breast cancer patients (overall survival, P < 0.001; cancer-specific survival, P < 0.001). By taking all the risk factors into a multivariate cox proportional model, MRM and MRM+RT remained to have detrimental effect on the prognosis compared with BCS+RT as shown by either overall (HR = 1.742, CI = 1.387-2.188, P < 0.001; HR = 1.449, CI = 1.038-2.204, P = 0.029) or cancer-specific survival (HR = 1.876, CI = 1.415-2.489, P < 0.001; HR = 1.701, CI = 1.168-2.478, P = 0.006). After we performed propensity score weighting and integrated the weights for each covariate in the multivariate cox proportional model. BCS+RT remained to be prognostic beneficial compared to the other treatment options (P < 0.001).
    ConclusionBCS+RT demonstrated better prognosis than MRM only and MRM+RT treatments for early-staged TNBC patients.


    Let (Ω,F,P) be a complete probability space, where Ω is a sample space, F2Ω is the σ-algebra of events, and P:F[0,1] is a probability measure. [0,T] is the temporal interval. DR2 is an open, bounded, Lipschitz physical space. D=D0D1 is the union of disjoint portions for the boundary. The unit outward normal vector to D is denoted by n. We carry out the numerical simulations of a transient heat model with random Robin coefficients and diffusion coefficients. That's to find a random function y:[0,T]×D×ΩR conforming to equations almost surely ωΩ,

    {yt[a(t,x,ω)y]=f(t,x,ω), in [0,T]×D×Ω,y(0,x,ω)=y0(x,ω), in D×Ω,ay(t,x,ω)n=0, on [0,T]×D0×Ω,ay(t,x,ω)n=α(t,x,ω)(u(t,x,ω)y(t,x,ω)), on [0,T]×D1×Ω, (1.1)

    where Robin coefficients α, diffusion coefficients a, boundary condition u, source term f, and initial condition y0 are random fields with continuous and bounded covariance functions.

    The model (1.1) describes the process of heat as well as mass transfer with random Robin factors and diffusion factors, which can be found in [3,4] and the references therein. These uncertain factors are involved in thermal environments. The temperature is quantified stochastically. In the review [3], this model can describe random high-cycle temperature fluctuations observed at the upper core structure of fast-breeder reactors or random variations in heat transfer coefficients around the stator vanes of gas turbines. The model (1.1) also appears in robust optimal boundary control problems as state equations, such as [25,26]. The effective implementation of these applications needs to solve (1.1) efficiently.

    To solve PDEs with random coefficients, there are many numerical algorithms that include polynomial chaos method, stochastic collocation method, stochastic finite element method, MC method, and so on (see, e.g., [1,9,11,21,22,27,29,31]). Where the MC method (see, e.g., [8,10,12,24]) is easy to implement and the convergence property of the MC method is not dependent on the dimension of the parameters in a random model. When the MC method is adopted, we first need independent sampling and then deal with independent numerical simulation per sample. These simulations are affected by independent diffusion factors, Robin factors, forces, initial and boundary conditions. To improve calculating efficiency, the ensemble algorithms are widely applied (see, e.g., [13,17,20,23,24]). In [23,24], the authors studied the parabolic problem with random coefficients by using the ensemble method, and obtained an error estimate. But the error estimate therein is not optimal in physical space. The authors of [18] combined the ensemble MC with the HDG method to obtain an optimal error estimate in space. For the heat conduction model with random Robin factors, we have not found other relevant results about ensemble algorithms besides our recent work [30].

    Look back at the numerical algorithm in the article [30]. ˉα(t,x) and ˉa(t,x) stand for the ensemble mean for the Robin factors as well as the diffusion factors, respectively, and we denote the fluctuations, αj:=ˉααj, aj:=ˉaaj. We adopt a uniform partition on the interval [0,T] as well as tn=nΔt,n=0,1,2,...,N. Let ynj,fnj,unj,anj, αnj, ˉan, ˉαn, anj and αnj be the values of functions yj,fj,uj,aj, αj, ˉa, ˉα, aj and αj at t=tn, respectively. yj=y(t,x;ωj) and the others are similar. In [30], based on the implicit Euler method, we developed a variational average ensemble method for the random transient heat problem (1.1). Conveniently, we called the variational average ensemble Monte Carlo (VAEMC) method: find yn+1j(j=1,2,,J;n=0,1,,N1) such that for all vH1(ˉD),

    (yn+1jynjΔt,v)+(ˉan+1yn+1j,v)+((an+1jˉan+1)ynj,v)+(ˉαn+1yn+1j,v)D1+((αn+1jˉαn+1)ynj,v)D1=(fn+1j,v)+(αn+1jun+1j,v)D1,

    After arrangement, that is

    (1Δtyn+1j,v)+(ˉan+1yn+1j,v)+(ˉαn+1yn+1j,v)D1=(fn+1j,v)+(αn+1jun+1j,v)D1+(1Δtynj,v)+((an+1j)ynj,v)+((αn+1j)ynj,v)D1. (1.2)

    Denote a=max1jJajˉa and α=max1jJαjˉα. From [30], we base on the following numerical stability condition a<1 and α<1 for the first-order ensemble time-stepping methods (1.2). In the ensemble algorithm in the literatures [17,18,19,20,23,24,30], both stability and convergence were conditionally dependent on the ratio between the fluctuating and average values of the relevant factors.

    Now, instead of employing ensemble means of Robin coefficients as well as diffusion coefficients in scheme (1.2), we propose a maximum ensemble of these coefficients at each time step. The process gets a shared coefficient matrix for the ensemble group to better improve computational efficiency. This is a new variational ensemble Monte Carlo numerical method, which we called the variational MAX ensemble Monte Carlo (VMEMC) method to conveniently distinguish it from other methods. In contrast with related methodologies, the novelty of this algorithm is that the numerical stability and convergence are unconditional.

    The following is the outline of this work. Some mathematical preliminaries and notations are listed in Section 2. The full discretization VMEMC scheme for the random transient heat model (1.1) is given in Section 3. In Section 4, the unconditional numerical stability and error estimate are analyzed. Several numerical tests are shown in Section 5. Related discussions of this algorithm and corresponding summaries are presented in Section 6.

    In this section, we will give some notations and mathematical preliminaries. For simplicity, dx,ds, and dt in some expressions will be omitted when there is no confusion. The boundaries D0 and D1 concern the experimentally accessible and inaccessible parts, respectively. Throughout this work C is a positive constant; it has different values in different places and does not rely on time step Δt, sample size J, or mesh size h.

    Let and (,) be the L2(D) norm as well as inner product, respectively. Simultaneously, D and (,)D stand for the L2(D) norm and the inner product. The Sobolev space Ws,q(D) with the norm vWs,q(D), here sN (natural number set) and q1. We denote Hs(D)=Ws,2(D). Particularly, H1(D) is equipped with the norm 1=1,D, which is defined by

    y1,D=(y2+y2)12.

    Let Hs(D) be the dual space of the bounded linear functions on Hs(D), with norm fs=sup0vHs(D)(f,v)/vs. Denote

    L(D)={v:v is a measurable function and |v|<+},

    where |v|=esssupxD|v|.

    (Ω,F,P) is a complete probability space. ZL1P(Ω) is a random variable. Denote

    E[Z]=ΩZ(ω)dP(ω).

    Let δ=(δ1,,δd) is a d-tuple with the length |δ|=di=1δi,δiN+. The stochastic Sobolev spaces ˜Ws,q(D)= LqP(Ω,Ws,q(D)) contains stochastic function v:Ω×DR, which is measurable with respect to the product σ-algebra FB(D). The norm of ˜Ws,q(D) is defined by

    v˜Ws,q(D)=(E[vqWs,q(D)])1/q=(E[|δs|D|δv|q])1/q,1q<+.

    Let ˜Hs(D)=˜Ws,2(D) L2P(Ω)Hs(D).

    We will use the tensor product Hilbert space

    X=˜L2(0,T;H1(D))L2P(0,T;H1(D);Ω)

    with its inner product

    (v,u)XE[T0D(vu+vu)].

    The induced norm is given by

    vX=(E[T0D(|v|2+v2)])1/2.

    Lemma 2.1. The norm 1,D1 is defined by

    y21,D1=D|y|2+D1|y|2.

    This norm and the standard norm 1 in H1(D) are equivalent. Particularly, there exist constants CP,CP>0 such that vH1(D)

    CPv1v1,D1,CPv1,D1v1.

    Proof. cf. [14,15].

    The weak solution of the problem (1.1) is defined as follows: a function yX is a weak solution of (1.1) if it satisfies the initial condition y(0,x,ω)=y0(x,ω)L2P(H1(D);Ω) and for T>0,

    T0(yt,v)dt+T0ΩDayvdxdP(ω)dt+T0ΩD1αyvdsdP(ω)dt=T0ΩDfvdxdP(ω)dt+T0ΩD1αuvdsdP(ω)dt (2.1)

    for all vX.

    We emphasize that the assumed regularity only requires the random fields to be square integrable. Assumptions on the finite-dimensional noise, that is, all the involved random input data depend on a finite number of real-valued random variables. Assumptions of input data for coefficients are considered:

    (A1) a=a(t,x,ω)L2P(0,T;L2(D);Ω) is uniformly bounded and coercive, i.e., amin, amax>0 such that Pro{ωΩ:a(t,x,ω)[amin,amax],(t,x)[0,T]×D}=1.

    (A2) α=α(t,x,ω)L2P(0,T;L2(D1);Ω) is uniformly bounded and coercive too, i.e., αmin, αmax>0 such that Pro{ωΩ:α(t,x,ω)[αmin,αmax],(t,x)[0,T]×D1}=1.

    Denote the ensemble group of solution variables for the equation (1.1) by yj(t,x)=y(t,x;ωj), j=1,2,...,J, with corresponding boundary conditions. [0,T]= N1n=0[tn,tn+1] is an isometric partition of the interval. These norms are utilized in the numerical error discussion: 1s<,

    \begin{align*} \boldsymbol{\||} v \boldsymbol{\||}_{\infty, s}: = \max _{0 \leq n \leq N}\left\|v^{n}\right\|_{s}, \; \; \; \boldsymbol{\||} v \boldsymbol{\||}_{p, s}: = \left(\Delta t \sum\limits_{n = 0}^{N}\left\|v^{n}\right\|_{s}^{p}\right)^{1/p} . \end{align*}

    We use the MC method for random sampling because it is non-intrusive, easy to implement, and its convergence is independent of the dimension of the uncertain model parameters. For j\in\{1, 2, ..., J\} , we let a_{\max}^{n} = \max \limits_{1 \leq j\leq J} \sup\limits_{x\in\Omega} a_{j}\left(t_{n}, \mathbf{x}\right) and \alpha _{\max }^{n} = \max \limits_{1\leq j\leq J}\sup \limits_{x \in \Omega}\alpha_{j}\left(t_{n}, \mathbf{x}\right) , where a_{j}\left(t, \mathbf{x}\right) = a\left(t, \mathbf{x}, \omega_{j}\right) , \alpha_{j}\left(t, \mathbf{x}\right) = \alpha\left(t, \mathbf{x}, \omega_{j}\right) . Without causing confusion, we abuse the fluctuating component symbols a_{j}^{\prime \; n}: = a_{\max }^{n}-a_{j}\left(t_{n}, \mathbf{x}\right) and \alpha_{j}^{\prime \; n}: = \alpha_{\max }^{n}-a_{j}\left(t_{n}, \mathbf{x}\right) . Suppressing the spatial discretization, utilizing an implicit-explicit for time-direct to the model (1.1), we propose the VMEMC scheme: find y_{j}^{n+1} \in H^{1}(\bar{D}) such that

    \begin{align} \;\;\;\;\; \left(\frac{y_{j}^{n+1}-y_{j}^{n}}{\Delta t},v\right) +\left(a_{\max }^{n+1} \nabla y_{j}^{n+1},\nabla v\right) -\left(a_{j}^{\prime \; n+1} \nabla y_{j}^{n},\nabla v\right) +\left(\alpha _{\max }^{n+1}y_{j}^{n+1},v\right)_{\partial D_{1}}\\ -\left(\alpha_{j}^{\prime \; n+1}y_{j}^{n},v \right)_{\partial D_{1}} = \left(f_{j}^{n+1},v\right)+\left(\alpha_{j}^{n+1}u_{j}^{n+1},v\right)_{\partial D_{1}}, j = 1, \ldots, J, \quad \forall v \in H^{1}(\bar{D}). \end{align} (2.2)

    Using a standard finite element (FE) method to discretize the space, we obtain the following block linear system for each ensemble member j :

    \begin{align} &\left(\frac{1}{\Delta t} M+a_{\max }^{n+1} D+\alpha_{\max }^{n+1} M_{\partial D_{1}}\right)y_{j}^{n+1}\\ = &\left(f_{j}^{n+1}+\alpha_{j}^{n+1}u_{j}^{n+1}+\frac{1}{\Delta t} M+a_{j }^{\prime n+1}D+\alpha_{j }^{\prime n+1}M_{\partial D_{1}}\right) y_{j}^{n}, \end{align} (2.3)

    where M is the mass matrix, M_{\partial D_{1}} is the boundary mass matrix, and D is the diffusion matrix.

    The above linear system is equivalent to the following:

    \begin{align} [A]\left[y_{1}\left|y_{2}\right| \ldots \mid y_{J}\right] = \left[b_{1}\left|b_{2}\right| \ldots \mid b_{J}\right], \end{align} (2.4)

    where A is the resulting coefficient matrix (independent of j ). The matrix A is symmetric positive definite (SPD) since both \frac{1}{\Delta t} M , a_{\max }^{n+1} D , and \alpha_{\max }^{n+1} M_{\partial D_{1}} are SPD. The system (2.4) can be solved with efficient block solvers [7,16,28]. Further, since only one coefficient matrix is required for computation each time step, the storage requirement is thereby reduced.

    Suppose \mathcal{T}_{h} is a quasi-uniform triangulation of the domain \bar{D} , such that \bar{D} = \bigcup_{K \in \mathcal{T}_{h}} \bar{K} . Let h_{K} be the diameter of the element K . Define h = \max _{K \in \mathcal{T}_{h}} h_{K} . Let \mathbf{P}_{k} be the set of polynomials of degree k . Denote the finite elements (FEs) space V_{h} of the domain \bar{D} by

    \begin{equation*} V_{h} = span\left\{\varphi_{k}\right\}_{1}^{n_{\bar{D}}} \subset\left\{v_{h} \in H^{1}(D) \cap H^{k+1}(\bar{D}) : \left.v_{h}\right|_{\bar{K}} \in \mathbf{P}_{k}, \forall K \in \mathcal{T}_{h}\right\}, \end{equation*}

    the FEs space S_{h} of the trace space corresponding to V_{h} by S_{h} = span\left\{\phi_{i}\right\}_{1}^{n_{\partial D_{1}}} \subset V_{h, \partial D_{1}}, here the space V_{h, \partial D_{1}} is defined as the restriction of V_{h} on \partial D_{1} , n_{\bar{D}} is the number of nodes on \bar{D} , n_{\partial D_{1}} is the number of nodes on \partial D_{1} . The spaces above satisfy the following approximation properties [2,5] : \forall 1 \leq l \leq k ,

    \begin{align} \inf _{v_{h} \in H^{1}(\bar{D})}\left\{\left\|y-v_{h}\right\|+h\left\|\nabla\left(y-v_{h}\right)\right\|\right\} \leq C h^{l+1}|y|_{l+1}, \quad y \in H^{1}(\bar{D}) \cap H^{l+1}(\bar{D}). \end{align} (2.5)

    I_{h}: C(\bar{D}) \rightarrow V_{h} or I_{h}: C(\partial D_{1}) \rightarrow S_{h} denotes the interpolation operator related to the spaces V_{h} or S_{h} .

    We first present a full discretization VMEMC method with finite elements for the VMEMC scheme (2.2) of the random transient heat model (1.1). And then the VMEMC algorithm is proposed. The full-discrete VMEMC scheme: find y_{j, h}^{n+1} \in V_{h} such that

    \begin{align} &\left(\frac{y_{j,h}^{n+1}-y_{j,h}^{n}}{\Delta t},v_{h}\right) +\left(a_{\max }^{n+1} \nabla y_{j,h}^{n+1},\nabla v_{h}\right) -\left(a_{j}^{\prime \; n+1} \nabla y_{j,h}^{n},\nabla v_{h}\right) +\left(\alpha _{\max }^{n+1}y_{j,h}^{n+1},v_{h}\right)_{\partial D_{1}}\\ -&\left(\alpha_{j}^{\prime \; n+1}y_{j,h}^{n},v_{h}\right)_{\partial D_{1}} = \left(f_{j}^{n+1},v_{h}\right)+\left(\alpha_{j}^{n+1}u_{j}^{n+1},v_{h}\right)_{\partial D_{1}}, \quad \forall v_{h} \in V_{h}, \end{align} (3.1)

    here the initial value y_{j, h}^{0} \in V_{h} which \left(y_{j, h}^{0}, v_{h}\right) = \left(y_{j}^{0}, v_{h}\right), \forall v_{h} \in V_{h} .

    As a comparison, here let's list the full-discrete VAEMC scheme [30] : find y_{j, h}^{n+1} \in V_{h} such that,

    \begin{align} &\left(\frac{y_{j, h}^{n+1}-y_{j, h}^{n}}{\Delta t},v_{h}\right) +\left(\bar{a}^{n+1} \nabla y_{j, h}^{n+1},\nabla v_{h}\right) -\left(a_{j}^{\prime \; n+1} \nabla y_{j, h}^{n},\nabla v_{h}\right) +\left(\bar{\alpha}^{n+1}y_{j, h}^{n+1},v_{h}\right)_{\partial D_{1}}\\ -&\left(\alpha_{j}^{\prime \; n+1}y_{j, h}^{n},v_{h} \right)_{\partial D_{1}} = \left(f_{j}^{n+1},v_{h}\right)+\left(\alpha_{j}^{n+1}u_{j}^{n+1},v_{h}\right)_{\partial D_{1}}, \forall v_{h} \in V_{h}, \end{align} (3.2)

    here the initial value y_{j, h}^{0}\in V_{h} which \left(y_{j, h}^{0}, v_{h}\right) = \left(y_{j}^{0}, v_{h}\right), \forall v_{h}\in V_{h} .

    By the way, the full-discrete variational Monte Carlo (VMC) scheme: find y_{j, h}^{n+1}\in V_{h} such that,

    \begin{align} &\left(\frac{y_{j, h}^{n+1}-y_{j, h}^{n}}{\Delta t},v_{h}\right) +\left(a_{j}^{n+1} \nabla y_{j, h}^{n+1},\nabla v_{h}\right) +\left(\alpha_{j}^{n+1}y_{j, h}^{n+1},v_{h}\right)_{\partial D_{1}}\\ = &\left(f_{j}^{n+1},v_{h}\right)+\left(\alpha_{j}^{n+1}u_{j}^{n+1},v_{h}\right)_{\partial D_{1}}, \forall v_{h} \in V_{h}, \end{align} (3.3)

    here the initial value y_{j, h}^{0} \in V_{h} which \left(y_{j, h}^{0}, v_{h}\right) = \left(y_{j}^{0}, v_{h}\right), \forall v_{h} \in V_{h} .

    Remark 3.1. (1) The symbols \alpha_{j}^{\prime n+1} and a_{j}^{\prime n+1} both appear in VMEMC and VAEMC schemes. Although they all represent fluctuations in corresponding random coefficients, they all mean different things.

    (2) Although the VMEMC and VAEMC schemes are formally similar, the novelty of the VMEMC scheme is its unconditional numerical stability and convergence. And the VMEMC scheme is more efficient in the numerical simulation.

    (3) If the random thermal conductivity and convective coefficient are independent of time, such as, \alpha: = \alpha(x, \omega) , a: = a(x, \omega) , the accordingly analyzed analysis is analogous to that presented below. It (VMEMC) is unconditionally stable as well as the first-order accurate approach.

    (4) Assume that the random thermal conductivity and convective coefficient are provided with uncertain temperature-dependent values as [6], e.g., a: = a(y) , \alpha: = \alpha(y) . Let y_{h}^{n} be the fully discrete approximate solution at time t_{n} , a_{h}^{n} = a\left(y_{h}^{n}\right) , a_{h}^{\prime \; n} = a_{\max }-a_{h}^{n} , \alpha_{h}^{n} = \alpha\left(y_{h}^{n}\right) , \alpha_{h}^{\prime \; n} = \alpha_{\max }-\alpha_{h}^{n} . Then we can replace them accordingly in the VMEMC scheme. Similarly, unconditionally stable and first-order accurate methods under assumptions that two coefficients are continuously differentiable can also be obtained.

    Applying the variational MAX ensemble scheme to the stochastic model (1.1), we first take the MC approach for random sampling. After sampling with independent identically distributed (i.i.d.), we solve the corresponding deterministic PDE for each sample. Furthermore, the solution for model (1.1) is made from the expectation of the solutions of the deterministic PDEs. A variational MAX ensemble-based Monte-Carlo algorithm (i.e., VMEMC algorithm) is proposed to quantifies uncertainty and improve its computational efficiency. The VMEMC algorithm includes the following three steps (S):

    VMEMC algorithm:

    \mathbf{S1.} Choose a group of random samples for the stochastic Robin coefficient, diffusion coefficient, source term, boundary, and initial conditions, \alpha_{j} \equiv \alpha(\cdot, \cdot, \omega_{j}) , a_{j} \equiv a(\cdot, \cdot, \omega_{j}) , f_{j} \equiv f(\cdot, \cdot, \omega_{j}) , u_{j} \equiv u(\cdot, \cdot, \omega_{j}) , and y_{j}^{0} \equiv y^{0}(\cdot, \omega_{j}) for per j\in \{1, 2, ..., J\} . Therefore, the solutions y(\cdot, \cdot, \omega_{j}) are i.i.d..

    \mathbf{S2.} Denote y_{j}^{n} = y(t_{n}, \mathbf{x}, \omega_{j}) , a _{\max }^{n} = \max \limits_{1 \leq j \leq J} \sup \limits_{x \in \Omega} a_{j}(t_{n}, \mathbf{x}) and \alpha _{\max }^{n} = \max \limits_{1 \leq j \leq J} \sup \limits_{x \in \Omega} \alpha_{j}(t_{n}, \mathbf{x}) . For n = 0, \ldots, N-1 and the j-th member, we find y_{j}^{n+1} to satisfy the VMEMC scheme. One finds the FE solution y_{h}(\cdot, \cdot, \omega_{j}) after choosing an approximate FE space.

    \mathbf{S3.} Take the VMEMC sample average \frac{1}{J} \sum_{j = 1}^{J} y_{h}(\cdot, \cdot, \omega_{j}) as an approximation to the expectation \mathbb{E}[y] . Given an object of interest G(y) , we are going to analyze these outputs for the variational ensemble systems, G(y_{h}(\cdot, \cdot, \omega_{1})) , \ldots , G\left(y_{h}\left(\cdot, \cdot, \omega_{J}\right)\right) , to obtain the stochastic information.

    The unconditional stability and error estimates are presented in this section.

    The unconditional stability for the VMEMC method is below.

    Theorem 4.1. Assume f_{j} \in L^{2}\left(0, T; H^{-1}(D) \right) , u_{j}\in L^{2}\left(0, T; L^{2}(\partial D_{1}) \right) , and hypothesises (\mathbf{A1}) and (\mathbf{A2}) are satisfied. Then the VMEMC method is unconditionally stable. Furthermore, the numerical solution of (3.1) holds

    \begin{align} & \mathbb{E} \left[\left\|y_{j, h}^{N}\right\|^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{N}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{N}\right\|_{\partial D_{1}}^{2}\right]\\ &+\sum\limits_{n = 0}^{N-1}\left( \mathbb{E} \left[\left\|y_{j, h}^{n+1}-y_{j, h}^{n}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|^{2}\right]\right.\\ &+\left.\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|_{\partial D_{1}}^{2}\right] \right) +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2} \right]\\ \leq &\mathbb{E} \left[\left\|y_{j, h}^{0}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{0}\right\|^{2}\right] +\Delta t \mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{0}\right\|_{\partial D_{1}}^{2}\right]\\ &+\frac{8 \Delta t}{C_{P}^{2} a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|f^{n+1}\right\|_{-1}^{2}\right] +\frac{8 \alpha^2_{\max }\Delta t}{C_{P}^{2} a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\|u_{j}^{n+1}\|_{\partial D_{1}}^{2}\right]. \end{align} (4.1)

    Proof. Taking v_{h} = 2\Delta t y_{j, h}^{n+1} in (3.1), we obtain

    \begin{align} 2\left(y_{j, h}^{n+1}-y_{j, h}^{n},y_{j, h}^{n+1}\right) &+2\Delta t \left(a_{\max }^{n+1} \nabla y_{j, h}^{n+1},\nabla y_{j, h}^{n+1}\right) -2\Delta t \left(a_{j}^{\prime \; n+1} \nabla y_{j, h}^{n}, \nabla y_{j, h}^{n+1}\right)\\ &+2\Delta t \left(\alpha _{\max }^{n+1}y_{j, h}^{n+1},y_{j, h}^{n+1}\right)_{\partial D_{1}} -2\Delta t\left(\alpha_{j}^{\prime \; n+1}y_{j, h}^{n},y_{j, h}^{n+1} \right)_{\partial D_{1}}\\ = &\left(f_{j}^{n+1}, 2\Delta t y_{j, h}^{n+1}\right)+\left(\alpha_{j}^{n+1}u_{j}^{n+1}, 2\Delta ty_{j, h}^{n+1}\right)_{\partial D_{1}}. \end{align} (4.2)

    Noting that

    \begin{align} &2\Delta t \left(a_{\max }^{n+1} \nabla y_{j, h}^{n+1},\nabla y_{j, h}^{n+1}\right) -2\Delta t \left(a_{j}^{\prime \; n+1} \nabla y_{j, h}^{n}, \nabla y_{j, h}^{n+1}\right) = 2 \Delta t\left(a_{\max }^{n+1} \nabla y_{j, h}^{n+1}, \nabla\left(y_{j, h}^{n+1}-y_{j, h}^{n}\right)\right)\\ &+2 \Delta t\left(a_{j}^{n+1} \nabla y_{j, h}^{n}, \nabla\left(y_{j, h}^{n+1}-y_{j, h}^{n}\right)\right) +2 \Delta t\left(a_{j}^{n+1} \nabla y_{j, h}^{n}, \nabla y_{j, h}^{n}\right) \end{align} (4.3)

    and

    \begin{align} &2\Delta t \left(\alpha _{\max }^{n+1}y_{j, h}^{n+1},y_{j, h}^{n+1}\right)_{\partial D_{1}} -2\Delta t\left(\alpha_{j}^{\prime \; n+1}y_{j, h}^{n},y_{j, h}^{n+1} \right)_{\partial D_{1}} = 2 \Delta t\left(\alpha_{\max }^{n+1} y_{j, h}^{n+1}, y_{j, h}^{n+1}-y_{j, h}^{n}\right)_{\partial D_{1}}\\ &+2 \Delta t\left(\alpha_{j}^{n+1} y_{j, h}^{n}, y_{j, h}^{n+1}-y_{j, h}^{n}\right)_{\partial D_{1}} +2 \Delta t\left(\alpha_{j}^{n+1} y_{j, h}^{n}, y_{j, h}^{n}\right)_{\partial D_{1}}, \end{align} (4.4)

    applying the polarization identity, integrating over the probability space, and rearranging terms in equation (4.2), we obtain

    \begin{align} & \mathbb{E} \left[\left\|y_{j, h}^{n+1}\right\|^{2}\right] -\mathbb{E} \left[\left\|y_{j, h}^{n}\right\|^{2}\right] +\mathbb{E} \left[\left\|y_{j, h}^{n+1}-y_{j, h}^{n}\right\|^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{n+1}\right\|^{2}\right]\\ & - \Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{n}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}}\nabla y_{j, h}^{n+1}\right\|^{2}\right]\\ &+ \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}}\nabla y_{j, h}^{n}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] -\Delta t \mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right]\\ &+\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|_{\partial D_{1}}^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} y_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right]\\ = &2\Delta t\mathbb{E} \left[\left(f_{j}^{n+1}, y_{j, h}^{n+1}\right)\right] +2\Delta t\mathbb{E} \left[\left(\alpha_{j}^{n+1}u_{j}^{n+1}, y_{j, h}^{n+1}\right)_{\partial D_{1}}\right]. \end{align} (4.5)

    Combing Cauchy-Schwarz, Young's inequalities with the trace theorem and the Sobolev embedding theorem to the right-hand side (RHS) of (4.5), we have

    \begin{equation} 2\Delta t\mathbb{E} \left[\left(f_{j}^{n+1}, y_{j, h}^{n+1}\right)\right] \leq \frac{\Delta t}{\epsilon_{1} a_{\min }}\mathbb{E} \left[\left\|f^{n+1}\right\|_{-1}^{2}\right] +\epsilon_{1} \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2}\right], \end{equation} (4.6)
    \begin{equation} 2\Delta t\mathbb{E} \left[\left(\alpha_{j}^{n+1}u_{j}^{n+1}, y_{j, h}^{n+1}\right)_{\partial D_{1}}\right] \leq \frac{\Delta t\alpha^2_{\max }}{\epsilon_{2} a_{\min }}\mathbb{E} \left[\|u_{j}^{n+1}\|_{\partial D_{1}}^{2}\right] +\epsilon_{2} \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2}\right] . \end{equation} (4.7)

    Under Lemma 2.1, we obtain \left\|\sqrt{a_{j}^{n+1}} \nabla y_{j, h}^{n+1}\right\|^{2}+ \left\|\sqrt{\alpha_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2} \geq \frac{C_{P}^{2}}{2}\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2} . Multiplying by \Delta t and integrating over the probability space, we can yield

    \begin{equation} \frac{C_{P}^{2}\Delta t}{2}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2} \right] \leq \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \nabla y_{j, h}^{n+1}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] . \end{equation} (4.8)

    For Eq (4.5), dropping two positive terms \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}}\nabla y_{j, h}^{n}\right\|^{2}\right] and \Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} y_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right] , useing estimate Eqs (4.6) and (4.7) with \epsilon_{1} = \epsilon_{2} = \frac{C_{P}^{2}}{8} , we can obtain

    \begin{align} & \mathbb{E} \left[\left\|y_{j, h}^{n+1}\right\|^{2}\right] -\mathbb{E} \left[\left\|y_{j, h}^{n}\right\|^{2}\right] +\mathbb{E} \left[\left\|y_{j, h}^{n+1}-y_{j, h}^{n}\right\|^{2}\right] + \Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{n+1}\right\|^{2}\right]\\ -& \Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{n}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right]\\ -&\Delta t \mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|_{\partial D_{1}}^{2}\right] +\frac{C_{P}^{2}\Delta t}{4}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} y_{j, h}^{n+1}\right\|_{1}^{2} \right]\\ \leq &\frac{8 \Delta t}{C_{P}^{2} a_{\min }}\mathbb{E} \left[\left\|f^{n+1}\right\|_{-1}^{2}\right] +\frac{8 \alpha^2_{\max }\Delta t}{C_{P}^{2} a_{\min }}\mathbb{E} \left[\|u_{j}^{n+1}\|_{\partial D_{1}}^{2}\right] . \end{align} (4.9)

    Summing from n = 0 to n = N-1 , one obtains the stability result Eq (4.1).

    Remark 4.1. (1) The stability result above shows that we have controlled over the temperature approximation in both L^{\infty}\left(0, T; L^{2}(\Omega)\right) and L^{2}\left(0, T; H^{1}(\Omega)\right) , unconditionally. Specifically, this numerical stability is independent of both the time step size and the grid size constraints, as well as the fluctuations in the random coefficients. This feature is different from the VAEMC approach ((see, e.g., [23,30]).

    (2)Compared to the standard Backward Euler method, we see that the numerical dissipation is enhanced with the additional term

    \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (y_{j, h}^{n+1}-y_{j, h}^{n})\right\|_{\partial D_{1}}^{2}\right].

    (3) If we consider the assumption of minimum regularity, that is,

    f\in \widetilde{L}^{2}\left(0, T; H^{-1}(D)\right), u\in \widetilde{L}^{2}\left(0,T;H^{-1}(\partial D_{1})\right),

    hypotheses (\mathbf{A1}) and (\mathbf{A2}) are satisfied, then the VMEMC method also possesses unconditional stability.

    In this subsection, we first decompose the overall error of the VMEMC algorithm into the statistical error and the discretization error. Next, we sequentially estimate the statistical error (Theorem 4.2) and the discretization error (Theorem 4.3). Finally, we combine these two error estimates to obtain the overall error estimate of the VMEMC algorithm (Theorem 4.4).

    The full-discrete VMEMC approximate solution of (1.1) is defined as \Psi_{h}^{n} \equiv \frac{1}{J} \sum_{j = 1}^{J} y_{j, h}^{n} . Noting that \mathbb{E}\left[y^{n}\right] = \mathbb{E}\left[y_{j}^{n}\right] , we can divide the overall error \mathbb{E}\left[y^{n}\right]-\Psi_{h}^{n} as

    \begin{aligned} \mathbb{E}\left[y^{n}\right]-\Psi_{h}^{n} = \left(\mathbb{E}\left[y_{j, h}^{n}\right]-\Psi_{h}^{n}\right)+\left(\mathbb{E}\left[y_{j}^{n}\right]-\mathbb{E}\left[y_{j, h}^{n}\right]\right) : = \mathcal{E}_{S}^{n}+\mathcal{E}_{h}^{n}. \end{aligned}

    Here the statistical error, \mathcal{E}_{S}^{n} = \mathbb{E}\left[y_{j, h}^{n}\right]-\Psi_{h}^{n} , is estimated using the sample number J . The second term, \mathcal{E}_{h}^{n} = \mathbb{E}\left[y_{j}^{n}-y_{j, h}^{n}\right] , that is corresponding to the total error of the time discretization and the FE discretization, is controlled by the time step as well as mesh size. Next, we examine the bounds of \mathcal{E}_{S}^{n} and \mathcal{E}_{h}^{n} .

    Theorem 4.2. Assume f_{j} \in L^{2}\left(0, T; H^{-1}(D) \right) , u_{j}\in L^{2}\left(0, T; L^{2}(\partial D_{1}) \right) , and hypotheses (\mathbf{A1}) and (\mathbf{A2}) are held, for the numerical solution to the VMEMC (3.1), then

    \begin{align*} & \mathbb{E} \left[\left\|\mathcal{E}_{S}^{N}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla \mathcal{E}_{S}^{N}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} \mathcal{E}_{S}^{N}\right\|_{\partial D_{1}}^{2}\right]\nonumber\\ &+\sum\limits_{n = 0}^{N-1}\left( \mathbb{E} \left[\left\|\mathcal{E}_{S}^{n+1}-\mathcal{E}_{S}^{n}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (\mathcal{E}_{S}^{n+1}-\mathcal{E}_{S}^{n})\right\|^{2}\right]\right.\nonumber\\ &+\left.\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (\mathcal{E}_{S}^{n+1}-\mathcal{E}_{S}^{n})\right\|_{\partial D_{1}}^{2}\right] \right) +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \mathcal{E}_{S}^{n+1}\right\|_{1}^{2} \right]\nonumber\\ \leq & \frac{C}{J} \left(\mathbb{E} \left[\left\|y_{j, h}^{0}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{0}\right\|^{2}\right] +\Delta t \mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{0}\right\|_{\partial D_{1}}^{2}\right]\right.\nonumber\\ &\left.+\frac{8 \Delta t}{C_{P}^{2} a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|f^{n+1}\right\|_{-1}^{2}\right] +\frac{8 \alpha^2_{\max }\Delta t}{C_{P}^{2} a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\|u_{j}^{n+1}\|_{\partial D_{1}}^{2}\right]\right) . \end{align*}

    Proof. The process of this proof is similar to the Theorem 5 in [30].

    The \mathcal{E}_{h}^{n} is examined in the following. Recall, the true solution satisfies

    \begin{align} & \;\;\;\;\left(\frac{y_{j}^{n+1}-y_{j}^{n}}{\Delta t}, v\right)+\left(a_{j}^{n+1} \nabla y_{j}^{n+1}, \nabla v\right)+\left(\alpha_{j}^{n+1} y_{j}^{n+1}, v\right)_{\partial D_{1}}\\ &= \left(f_{j}^{n+1}, v\right)+ \left(\frac{y_{j}^{n+1}-y_{j}^{n}}{\Delta t}-y_{j, t}^{n+1}, v\right) +(\alpha_{j}^{n+1} u_{j}^{n+1}, v)_{\partial D_{1}} \; \; \; ,\forall v \in H^{1}(D). \end{align} (4.10)

    The error at the time t = t_{n} is denoted by e_{j}^{n} = \left(y_{j}^{n}-I_{h} y_{j}^{n}\right)-\left(y_{j, h}^{n}-I_{h} y_{j}^{n}\right) = \phi_{j}^{n}-\psi_{j, h}^{n} , where I_{h} y_{j}^{n} is a Lagrange interpolation [2,5] of y_{j}^{n} . Taking v = v_{h} \in V_{h} and subtracting Eq (3.1) from (4.10), respectively, we have the error equation:

    \begin{align} \left(\frac{e_{j}^{n+1}-e_{j}^{n}}{\Delta t}, v_{h}\right) +&\left(a_{\max }^{n+1} \nabla e_{j}^{n+1}, \nabla v_{h}\right) -\left(a_{j}^{\prime \; n+1} \nabla e_{j}^{n}, \nabla v_{h}\right)\\ +&\left(\alpha_{\max }^{n+1} e_{j}^{n+1} , v_{h}\right)_{\partial D_{1}} -\left(\alpha_{j}^{\prime\; n+1} e_{j}^{n}, v_{h}\right)_{\partial D_{1}} = \xi\left(y_{j}^{n+1}, v_{h}\right), \end{align} (4.11)

    here \xi\left(y_{j}^{n+1}, v_{h}\right) is defined as

    \begin{align} \xi\left(y_{j}^{n+1}, v_{h}\right): = &\left(\frac{y_{j}^{n+1}-y_{j}^{n}}{\Delta t}-y_{j, t}^{n+1}, v_{h}\right) +\left(\left(a_{\max }^{n+1}-a_{j}^{n+1}\right) \nabla y_{j}^{n+1}, \nabla v_{h}\right) -\left(a_{j}^{\prime \; n+1} \nabla y_{j}^{n}, \nabla v_{h}\right) \\ +&\left(\left(\alpha_{\max }^{n+1}-\alpha_{j}^{n+1}\right) y_{j}^{n+1}, v_{h}\right)_{\partial D_{1}} -\left(\alpha_{j}^{\prime \; n+1} y_{j}^{n}, v_{h}\right)_{\partial D_{1}}. \end{align} (4.12)

    The following regularity assumptions are needed:

    \begin{align} &y_{j} \in L^{\infty}\left(0, T ; H^{1}(D) \cap H^{k+1}(D)\right), \nabla y_{j} \in L^{\infty}\left(0, T ; L^{\infty}(D)\right), \\ &y_{j, t} \in L^{2}\left(0, T ; H^{1}(D)\right), y_{j, tt} \in L^{2}\left(0, T ; H^{-1}(D)\right) . \end{align} (4.13)

    As a preparatory step for estimating the discretization error, we first estimate each term of the \xi\left(y_{j}^{n+1}, v_{h}\right) in the following lemma.

    Lemma 4.1. Suppose every y_{j} fulfills the equation (1.1) and the regularity assumptions (4.13), then,

    \begin{align} \mathbb{E} \left[\xi\left(y_{j}^{n+1}, v_{h}\right) \right] \leq & \left(\sum\limits_{i = 3}^{7} \frac{\epsilon_{i}}{2}\right)\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} v_{h}\right\|_{1}^{2}\right] +\frac{\Delta t}{2 a_{\min } \epsilon_{3}}\mathbb{E} \left[\left\|y_{j, t t}\right\|_{L^{2}\left(t_{n},t_{n+1} ; H^{-1}(D)\right)}^{2}\right]\\ +&\left(\epsilon_{4}^{-1}+\epsilon_{5}^{-1}\right)\Delta t \frac{a_{\max}^{2} }{2 a_{\min } }\mathbb{E} \left[\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(D)\right)}\right]\\ +&\left(\epsilon_{6}^{-1}+\epsilon_{7}^{-1}\right)\Delta t \frac{\alpha_{\max}^{2} }{2 \alpha_{\min } }\mathbb{E} \left[\left\| y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(\partial D_{1})\right)}\right] . \end{align} (4.14)

    Proof. Applying Lemma 2.1, the Cauchy-Schwarz and Young's inequalities, the Taylor's Theorem with integral remainder, we obtain the following results for \xi\left(y_{j}^{n+1}, v_{h}\right) is defined by Eq (4.12). The first term of \xi\left(y_{j}^{n+1}, v_{h}\right) is estimated as

    \begin{equation} \left(\frac{y_{j}^{n+1}-y_{j}^{n}}{\Delta t}-y_{j, t}^{n+1}, v_{h}\right) \leq \frac{\Delta t}{2 a_{\min } \epsilon_{3}}\left\|y_{j, t t}\right\|_{L^{2}\left(t_{n},t_{n+1} ; H^{-1}(D)\right)}^{2} +\frac{\epsilon_{3}}{2}\left\|\sqrt{a_{j}^{n+1}} v_{h}\right\|_{1}^{2} . \end{equation} (4.15)

    The second and third terms of \xi\left(y_{j}^{n+1}, v_{h}\right) are estimated as

    \begin{align} &\left(\left(a_{\max }^{n+1}-a_{j}^{n+1}\right) \nabla y_{j}^{n+1}, \nabla v_{h}\right) -\left(a_{j}^{\prime \; n+1} \nabla y_{j}^{n}, \nabla v_{h}\right)\\ = &\left(a_{\max }^{n+1} \nabla\left(y_{j}^{n+1}-y_{j}^{n} \right), \nabla v_{h}\right) -\left(a_{j}^{n+1} \nabla\left(y_{j}^{n+1}-y_{j}^{n}\right), \nabla v_{h}\right), \end{align} (4.16)
    \begin{equation} \left(a_{\max }^{n+1} \nabla\left(y_{j}^{n+1}-y_{j}^{n} \right), \nabla v_{h}\right) \leq \frac{a_{\max}^{2} \Delta t}{2 a_{\min } \epsilon_{4}}\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(D)\right)} +\frac{\epsilon_{4}}{2}\left\|\sqrt{a_{j}^{n+1}} \nabla v_{h}\right\|^{2}, \end{equation} (4.17)
    \begin{equation} -\left(a_{j}^{n+1} \nabla\left(y_{j}^{n+1}-y_{j}^{n}\right), \nabla v_{h}\right) \leq \frac{a_{\max}^{2} \Delta t}{2 a_{\min } \epsilon_{5}}\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(D)\right)} +\frac{\epsilon_{5}}{2}\left\|\sqrt{a_{j}^{n+1}} \nabla v_{h}\right\|^{2}. \end{equation} (4.18)

    The last two terms of \xi\left(y_{j}^{n+1}, v_{h}\right) are estimated as

    \begin{align} &\left(\left(\alpha_{\max }^{n+1}-\alpha_{j}^{n+1}\right) y_{j}^{n+1}, v_{h}\right)_{\partial D_{1}} -\left(\alpha_{j}^{\prime \; n+1} y_{j}^{n}, v_{h}\right)_{\partial D_{1}}\\ = &\left(\alpha_{\max }^{n+1} \left(y_{j}^{n+1}-y_{j}^{n} \right), v_{h}\right)_{\partial D_{1}} -\left(\alpha_{j}^{n+1} \left(y_{j}^{n+1}-y_{j}^{n}\right), v_{h}\right)_{\partial D_{1}}, \end{align} (4.19)
    \begin{equation} \left(\alpha_{\max }^{n+1} \left(y_{j}^{n+1}-y_{j}^{n} \right), v_{h}\right)_{\partial D_{1}} \leq \frac{\alpha_{\max}^{2} \Delta t}{2 \alpha_{\min } \epsilon_{6}}\left\| y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(\partial D_{1})\right)} +\frac{\epsilon_{6}}{2}\left\|\sqrt{\alpha_{j}^{n+1}} v_{h}\right\|_{\partial D_{1}}^{2}, \end{equation} (4.20)
    \begin{equation} -\left(\alpha_{j}^{n+1} \left(y_{j}^{n+1}-y_{j}^{n}\right), v_{h}\right)_{\partial D_{1}} \leq \frac{\alpha_{\max}^{2} \Delta t}{2 \alpha_{\min } \epsilon_{7}}\left\| y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(\partial D_{1})\right)} +\frac{\epsilon_{7}}{2}\left\|\sqrt{\alpha_{j}^{n+1}} v_{h}\right\|_{\partial D_{1}}^{2}. \end{equation} (4.21)

    Combining the above estimates Eqs (4.15)–(4.21), using \left\|\nabla v_{h}\right\| \leq\left\|v_{h}\right\|_{1} and \left\| v_{h}\right\|_{\partial D_{1}} \leq\left\|v_{h}\right\|_{1} , and integrating over the probability space, we can yield Eq (4.13).

    With the results of Lemma 4.1, the major convergence result of this work can be proved next.

    Theorem 4.3. Suppose y_{j} satisfies the assumptions of Lemma 4.1. Moreover, suppose y_{j, h}^{0} \in V_{h} is an approximation of y_{j}^{0} with the accuracy of the interpolation. Then, the VMEMC solution of (3.1) satisfies

    \begin{align} & \mathbb{E} \left[\left\|e_{j}^{N}\right\|^{2}\right] +\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|e_{j}^{n+1}-e_{j}^{n}\right\|^{2}\right] +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} e_{j}^{n+1}\right\|_{1}^{2}\right]\\ & +\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{N}} \nabla e_{j}^{N}\right\|^{2} \right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{a_{j}^{\prime\; n+1}} \nabla\left(e_{j}^{n+1}-e_{j}^{n}\right)\right\|^{2}\right]\\ +&\Delta t\mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{N}} e_{j, h}^{N}\right\|_{\partial D_{1}}^{2}\right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}}\left(e_{j, h}^{n+1}-e_{j, h}^{n}\right)\right\|_{\partial D_{1}}^{2}\right] \\ \leq &C\left(\frac{1}{a_{\min}}+\frac{a_{\max }^2}{a_{\min }}+\frac{\alpha_{\max }^2}{\alpha_{\min }} \right)\Delta t^{2} +C\left(\frac{a_{\max }^2}{a_{\min }}+\frac{a_{\max }^2}{a_{\min }} \Delta t^2 +a_{\max }\Delta t+ a_{\max }\Delta t^{2} \right) h^{2k}\\ &+C\left(\frac{\alpha_{\max }^2}{a_{\min }}+\frac{1}{a_{\min }}+1 +\Delta t+ a_{\max }\Delta t+\alpha_{\max }\Delta t+\alpha_{\max }\Delta t^{2}\right)h^{2k+2}. \end{align} (4.22)

    Proof. Noting e_{j}^{n} = \phi_{j}^{n}-\psi_{j, h}^{n} , rearranging the equation (4.11) with v_{h} = 2 \Delta t \psi_{j, h}^{n+1} , we have

    \begin{align} \left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}, 2\psi_{j, h}^{n+1}\right) &+2\Delta t\left(a_{\max}^{n+1} \nabla \psi_{j, h}^{n+1},\nabla \psi_{j, h}^{n+1}\right) -2\Delta t\left(a_{j}^{\prime \; n+1} \nabla \psi_{j, h}^{n},\nabla \psi_{j, h}^{n+1}\right) \\ &+2\Delta t\left(\alpha_{\max }^{n+1} \psi_{j , h}^{n+1},\psi_{j,h}^{n+1}\right)_{\partial D_{1}} -2\Delta t\left(\alpha_{j}^{\prime \; n+1} \psi_{j, h}^{n},\psi_{j,h}^{n+1}\right)_{\partial D_{1}} \\ = &\left(\phi_{j}^{n+1}-\phi_{j}^{n}, 2\psi_{j, h}^{n+1}\right) -\xi\left(y_{j}^{n+1},2\Delta t \psi_{j, h}^{n+1}\right)\\ &+2\Delta t\left(a_{\max}^{n+1}\nabla \phi_{j}^{n+1},\nabla \psi_{j,h}^{n+1}\right) -2\Delta t\left(a_{j}^{\prime \; n+1}\nabla \phi_{j}^{n},\nabla \psi_{j,h}^{n+1}\right) \\ &+2\Delta t\left(\alpha_{\max}^{n+1}\phi_{j}^{n+1}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}} -2\Delta t\left(\alpha_{j}^{\prime \; n+1} \phi_{j}^{n}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}}. \end{align} (4.23)

    Integrating over the probability space and applying the polarization identity, we arrive at

    \begin{align} &\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}\right\|^{2}\right] -\mathbb{E} \left[\left\|\psi_{j, h}^{n}\right\|^{2}\right] +\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}-\psi_{j , h}^{n}\right\|^{2}\right]\\ &+\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2} \right] -\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{n+1}} \nabla \psi_{j, h}^{n}\right\|^{2} \right]\\ &+\Delta t \mathbb{E}\left[\left\|\sqrt{a_{j}^{\prime\; n+1}} \nabla\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|^{2}\right] +\Delta t \mathbb{E}\left[\left\|\sqrt{a_{j}^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2}\right] \\ &+\Delta t \mathbb{E}\left[\left\|\sqrt{a_{j}^{n+1}} \nabla \psi_{j, h}^{n}\right\|^{2}\right] +\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{n+1}} \psi_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right]\\ &-\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{n+1}} \psi_{j , h}^{n}\right\|_{\partial D_{1}}^{2}\right] +\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}}\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|_{\partial D_{1}}^{2}\right]\\ &+\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] +\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{n+1}} \psi_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right]\\ = &-2\Delta t\mathbb{E} \left[ \xi\left(y_{j}^{n+1}, \psi_{j, h}^{n+1}\right)\right] +\mathbb{E} \left[\left(\phi_{j}^{n+1}-\phi_{j}^{n}, 2\psi_{j, h}^{n+1}\right)\right]\\ &+2\Delta t\mathbb{E} \left[ \left(a_{\max }^{n+1} \nabla \phi_{j}^{n+1}, \nabla \psi_{j, h}^{n+1}\right)\right] -2\Delta t\mathbb{E} \left[ \left(a_{j}^{\prime \; n+1} \nabla \phi_{j}^{n}, \nabla \psi_{j, h}^{n+1}\right)\right]\\ &+2\Delta t\mathbb{E} \left[ \left(\alpha_{\max }^{n+1} \phi_{j}^{n+1}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}}\right] -2\Delta t\mathbb{E} \left[ \left(\alpha_{j}^{\prime \; n+1} \phi_{j}^{n}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}}\right]. \end{align} (4.24)

    From Lemma 4.1 with v_{h} = \psi_{j, h}^{n+1} , the first term on the RHS of Eq (4.24) can be expressed as

    \begin{align} &-2\Delta t\mathbb{E} \left[\xi\left(y_{j}^{n+1}, \psi_{j, h}^{n+1}\right) \right] \leq\left(\sum\limits_{i = 3}^{7} \epsilon_{i}\right)\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{1}^{2}\right]\\ &+\frac{\Delta t^{2}}{ a_{\min } \epsilon_{3}}\mathbb{E} \left[\left\|y_{j, t t}\right\|_{L^{2}\left(t_{n},t_{n+1} ; H^{-1}(D)\right)}^{2}\right] +\left(\epsilon_{4}^{-1}+\epsilon_{5}^{-1}\right)\Delta t^{2} \frac{a_{\max}^{2} }{ a_{\min } }\mathbb{E} \left[\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(D)\right)}\right]\\ &+\left(\epsilon_{6}^{-1}+\epsilon_{7}^{-1}\right)\Delta t^{2} \frac{\alpha_{\max}^{2} }{ \alpha_{\min } }\mathbb{E} \left[\left\| y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(\partial D_{1})\right)}\right] . \end{align} (4.25)

    Applying Lemma 2.1, the Taylor's theorem with integral remainder, the Young's inequality, and the Cauchy-Schwarz inequality on the second term of RHS of (4.24), we have

    \begin{equation} \mathbb{E} \left[\left(\phi_{j}^{n+1}-\phi_{j}^{n}, 2\psi_{j, h}^{n+1}\right)\right] \leq\frac{1}{\epsilon_{8}a_{\min }}\mathbb{E} \left[\left\|\phi_{t}\right\|_{L^{2}\left(t_{n}, t_{n+1}; L^{2}(D)\right)}^{2}\right] +\epsilon_{8}\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|^{2}\right]. \end{equation} (4.26)

    Using the Young's inequality and the Cauchy-Schwarz inequality to the third and fourth terms, we obtain

    \begin{equation} 2 \Delta t\mathbb{E} \left[ \left(a_{\max }^{n+1} \nabla \phi_{j}^{n+1}, \nabla \psi_{j, h}^{n+1}\right)\right] \leq\frac{a_{\max}^{2} \Delta t}{ \epsilon_{9} a_{\min }}\mathbb{E} \left[\left\|\nabla \phi_{j}^{n+1}\right\|^{2}\right] +\epsilon_{9}\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2}\right], \end{equation} (4.27)
    \begin{equation} -2\Delta t\mathbb{E} \left[ \left(a_{j}^{\prime \; n+1} \nabla \phi_{j}^{n}, \nabla \psi_{j, h}^{n+1}\right)\right] \leq\frac{a_{\max}^{2} \Delta t}{ \epsilon_{10} a_{\min }}\mathbb{E} \left[\left\|\nabla \phi_{j}^{n}\right\|^{2}\right] +\epsilon_{10}\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2}\right]. \end{equation} (4.28)

    Using the Young's inequality and the Cauchy-Schwarz inequality as well as the trace theorem to the two boundary terms, we have

    \begin{equation} 2 \Delta t\mathbb{E} \left[ \left(\alpha_{\max }^{n+1} \phi_{j}^{n+1}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}}\right] \leq\frac{\alpha_{\max}^{2} \Delta t}{ \epsilon_{11} a_{\min }}\mathbb{E} \left[\left\| \phi_{j}^{n+1}\right\|_{\partial D_{1}}^{2}\right] +\epsilon_{11}\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|^{2}\right], \end{equation} (4.29)
    \begin{equation} -2 \Delta t\mathbb{E} \left[ \left(\alpha_{j}^{\prime \; n+1} \phi_{j}^{n}, \psi_{j, h}^{n+1}\right)_{\partial D_{1}}\right] \leq\frac{\alpha_{\max}^{2} \Delta t}{ \epsilon_{12} a_{\min }}\mathbb{E} \left[\left\| \phi_{j}^{n}\right\|_{\partial D_{1}}^{2}\right] +\epsilon_{12}\Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|^{2}\right]. \end{equation} (4.30)

    Under Lemma 2.1, proceeding in similar fashion with equation (4.8), we obtain

    \begin{equation} \frac{C_{P}^{2}\Delta t}{2}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{1}^{2} \right] \leq \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] . \end{equation} (4.31)

    Dropping the positive terms \Delta t\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}}\nabla \psi_{j, h}^{n}\right\|^{2}\right] and \Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{n+1}} \psi_{j, h}^{n}\right\|_{\partial D_{1}}^{2}\right] , using the estimate (4.25)-(4.31), noting that \|\nabla v\| \leq\|v\|_{1} , \| v\| \leq\|v\|_{1} , selecting \epsilon_{i} = C_{P}^{2} / 40 for i = 3, 4, ..., 12 , in equation (4.24), then we can obtain,

    \begin{align} &\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}\right\|^{2}\right]-\mathbb{E} \left[\left\|\psi_{j, h}^{n}\right\|^{2}\right] +\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}-\psi_{j , h}^{n}\right\|^{2}\right] +\frac{C_{P}^{2}\Delta t}{4}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{1}^{2}\right]\\ &+\Delta t\left(\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{n+1}} \nabla \psi_{j, h}^{n+1}\right\|^{2} \right] -\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{n+1}} \nabla \psi_{j, h}^{n}\right\|^{2} \right]\right)\\ &+\Delta t\mathbb{E}\left[\left\|\sqrt{a_{j}^{\prime\; n+1}} \nabla\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|^{2}\right] + \Delta t\mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}}\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|_{\partial D_{1}}^{2}\right]\\ &+\Delta t \left(\mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{n+1}} \psi_{j, h}^{n+1}\right\|_{\partial D_{1}}^{2}\right] - \mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{n+1}} \psi_{j , h}^{n}\right\|_{\partial D_{1}}^{2}\right]\right)\\ \leq&\frac{40\Delta t^{2}}{ a_{\min } C_{P}^{2}}\mathbb{E} \left[\left\|y_{j, t t}\right\|_{L^{2}\left(t_{n},t_{n+1} ; H^{-1}(D)\right)}^{2}\right] +\frac{40}{C_{P}^{2}a_{\min }}\mathbb{E} \left[\left\|\phi_{t}\right\|_{L^{2}\left(t_{n}, t_{n+1}; L^{2}(D)\right)}^{2}\right]\\ &+\frac{80\Delta t^{2}}{ C_{P}^{2}} \frac{a_{\max}^{2} }{ a_{\min }} \mathbb{E} \left[\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(D)\right)}\right] +\frac{80\Delta t^{2}}{ C_{P}^{2}} \frac{\alpha_{\max}^{2} }{ \alpha_{\min }}\mathbb{E} \left[\left\| y_{j, t}\right\|^{2}_{L^{2}\left(t_{n}, t_{n+1} ; L^{2}(\partial D_{1})\right)}\right] \\ &+\frac{40a_{\max}^{2} \Delta t}{ a_{\min }C_{P}^{2}}\left(\mathbb{E} \left[\left\|\nabla \phi_{j}^{n+1}\right\|^{2}\right] +\mathbb{E} \left[\left\|\nabla \phi_{j}^{n}\right\|^{2}\right]\right)\\ &+\frac{40\alpha_{\max}^{2} \Delta t}{ a_{\min }C_{P}^{2}}\left(\mathbb{E} \left[\left\| \phi_{j}^{n+1}\right\|_{\partial D_{1}}^{2}\right] +\mathbb{E} \left[\left\|\phi_{j}^{n}\right\|_{\partial D_{1}}^{2}\right]\right). \end{align} (4.32)

    Summing n = 0 to n = N-1 , collecting the constants, and reorganizing the items, we obtain

    \begin{align} & \mathbb{E} \left[\left\|\psi_{j, h}^{N}\right\|^{2}\right] +\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}-\psi_{j , h}^{n}\right\|^{2}\right] +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{1}^{2}\right]\\ & +\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{N}} \nabla \psi_{j, h}^{N}\right\|^{2} \right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{a_{j}^{\prime\; n+1}} \nabla\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|^{2}\right]\\ & +\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{N}} \psi_{j, h}^{N}\right\|_{\partial D_{1}}^{2}\right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}}\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|_{\partial D_{1}}^{2}\right]\\ & \leq\mathbb{E} \left[\left\|\psi_{j, h}^{0}\right\|^{2}\right] +\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{0}} \nabla \psi_{j, h}^{0}\right\|^{2} \right] +\Delta t\mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{0}} \psi_{j , h}^{0}\right\|_{\partial D_{1}}^{2}\right]\\ & +\frac{C\Delta t^{2}}{ a_{\min }}\mathbb{E} \left[\left\|y_{j, t t}\right\|_{L^{2}\left(0,T ; H^{-1}(D)\right)}^{2}\right] +\frac{C}{a_{\min }}\mathbb{E} \left[\left\|\phi_{t}\right\|_{L^{2}\left(0,T ; L^{2}(D)\right)}^{2}\right]\\ & +\frac{C\Delta t^{2}a_{\max}^{2} }{ a_{\min }} \mathbb{E} \left[\left\|\nabla y_{j, t}\right\|^{2}_{L^{2}\left(0,T ; L^{2}(D)\right)}\right] +\frac{C\Delta t^{2}\alpha_{\max}^{2} }{ \alpha_{\min }}\mathbb{E} \left[\left\| y_{j, t}\right\|^{2}_{L^{2}\left(0,T ; L^{2}(\partial D_{1})\right)}\right]\\ & +\frac{C a_{\max}^{2} \Delta t}{ a_{\min }}\sum\limits_{n = 0}^{N-1}\left(\mathbb{E} \left[\left\|\nabla \phi_{j}^{n+1}\right\|^{2}\right] +\mathbb{E} \left[\left\|\nabla \phi_{j}^{n}\right\|^{2}\right]\right)\\& +\frac{C\alpha_{\max}^{2} \Delta t}{ a_{\min }}\sum\limits_{n = 0}^{N-1}\left(\mathbb{E} \left[\left\| \phi_{j}^{n+1}\right\|^{2}_{\partial D_{1}}\right] +\mathbb{E} \left[\left\| \phi_{j}^{n}\right\|^{2}_{\partial D_{1}}\right]\right). \end{align} (4.33)

    Then,

    \begin{align} &\mathbb{E} \left[\left\|\psi_{j, h}^{N}\right\|^{2}\right] +\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\psi_{j, h}^{n+1}-\psi_{j , h}^{n}\right\|^{2}\right] +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \psi_{j, h}^{n+1}\right\|_{1}^{2}\right]\\ &+\Delta t\mathbb{E} \left[ \left\|\sqrt{a_{\max }^{N}} \nabla \psi_{j, h}^{N}\right\|^{2} \right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{a_{j}^{\prime\; n+1}} \nabla\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|^{2}\right]\\ &+\Delta t \mathbb{E}\left[\left\|\sqrt{\alpha_{\max }^{N}} \psi_{j, h}^{N}\right\|_{\partial D_{1}}^{2}\right] + \Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E}\left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}}\left(\psi_{j, h}^{n+1}-\psi_{j, h}^{n}\right)\right\|_{\partial D_{1}}^{2}\right]\\ \leq & C\left( \frac{1}{a_{\min }}+\frac{a_{\max }^2}{a_{\min }}+\frac{\alpha_{\max }^2}{\alpha_{\min }} \right)\Delta t^{2} +C\frac{a_{\max }^2}{a_{\min }} h^{2k} +C \frac{\alpha_{\max }^2}{a_{\min }} h^{2k+2}+\frac{C}{a_{\min }}h^{2k+2}. \end{align} (4.34)

    Inequalities \left\|e_{j }^{n}\right\|^{2} \leq 2\left\|\phi_{j }^{n}\right\|^{2}+2\left\|\psi_{j, h}^{n}\right\|^{2} and \left\|\phi_{j}^{n+1}-\phi_{j}^{n}\right\|^{2}\leq C \Delta t \int_{t_n}^{t_{n+1}} \left\|\phi_{j, t}(s)\right\|^{2} d s imply the result (4.22).

    Combining the MC sampling error and the discretization error, one can acquire the total error of the VMEMC FE approximation.

    Theorem 4.4. Assume that f_{j} \in L^{2}\left(0, T; H^{-1}(D) \right) , u_{j}\in L^{2}\left(0, T; L^{2}(\partial D_{1})\right) , y_{j} satisfies the assumptions of Lemma 4.1. Moreover, suppose y_{j, h}^{0} \in V_{h} is an approximation of y_{j}^{0} with the accuracy of the interpolation. Then,

    \begin{align} & \mathbb{E}\left[\left\|\mathbb{E}\left[y^{N}\right]-\Psi_{h}^{N}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla \left(\mathbb{E}\left[y^{N}\right]-\Psi_{h}^{N}\right)\right\|^{2}\right] \\ & +\Delta t\mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} \left(\mathbb{E}\left[y^{N}\right]-\Psi_{h}^{N}\right)\right\|_{\partial D_{1}}^{2}\right] +\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\left(\mathbb{E}\left[y^{n+1}\right]-\Psi_{h}^{n+1}\right)-\left(\mathbb{E}\left[y^{n}\right] -\Psi_{h}^{n}\right)\right\|^{2}\right] \\ & +\Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{\prime \; n+1}}\nabla (\left(\mathbb{E}\left[y^{n+1}\right]-\Psi_{h}^{n+1}\right)-\left(\mathbb{E}\left[y^{n}\right] -\Psi_{h}^{n}\right))\right\|^{2}\right] \\ & +\Delta t\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{\alpha_{j}^{\prime \; n+1}} (\left(\mathbb{E}\left[y^{n+1}\right]-\Psi_{h}^{n+1}\right)-\left(\mathbb{E}\left[y^{n}\right] -\Psi_{h}^{n}\right))\right\|_{\partial D_{1}}^{2}\right] \\ & +\frac{C_{P}^{2}\Delta t}{4}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|\sqrt{a_{j}^{n+1}} \left(\mathbb{E}\left[y^{n+1}\right]-\Psi_{h}^{n+1}\right)\right\|_{1}^{2} \right] \\ & \leq \frac{C}{J} \left(\mathbb{E} \left[\left\|y_{j, h}^{0}\right\|^{2}\right] +\Delta t\mathbb{E} \left[\left\|\sqrt{a_{\max }^{n+1}}\nabla y_{j, h}^{0}\right\|^{2}\right] +\Delta t \mathbb{E} \left[\left\|\sqrt{\alpha_{\max }^{n+1}} y_{j, h}^{0}\right\|_{\partial D_{1}}^{2}\right]\right. \\ & +\left.\frac{ \Delta t}{ a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\left\|f_{j}^{n+1}\right\|_{-1}^{2}\right] +\frac{ \alpha^2_{\max }\Delta t}{ a_{\min }}\sum\limits_{n = 0}^{N-1}\mathbb{E} \left[\|u_{j}^{n+1}\|_{\partial D_{1}}^{2}\right]\right) \\ &+C \left( \frac{1}{a_{\min }}+\frac{a_{\max }^2}{a_{\min }}+\frac{\alpha_{\max }^2}{\alpha_{\min }} \right)\Delta t^{2} +C \left(\frac{a_{\max }^2}{a_{\min }}+\frac{a_{\max }^2}{a_{\min }} \Delta t^{2}+a_{\max }\Delta t+ a_{\max }\Delta t^{2} \right) h^{2k}\\ &+C\left(\frac{\alpha_{\max }^2}{a_{\min }}+\frac{1}{a_{\min }}+1 +\Delta t+ a_{\max }\Delta t+\alpha_{\max }\Delta t+\alpha_{\max }\Delta t^{2}\right)h^{2k+2}. \end{align} (4.35)

    Proof. In view of the first term on the left-hand side (LHS) of (4.35), the triangle inequality and Young inequality suggest

    \mathbb{E}\left[\left\|\mathbb{E}\left[y^{N}\right]-\Psi_{h}^{N}\right\|^{2}\right] \leq 2\left(\mathbb{E}\left[\left\|\mathbb{E}\left[y^{N}\right] -\mathbb{E}\left[y_{h}^{N}\right]\right\|^{2}\right] +\mathbb{E}\left[\left\|\mathbb{E}\left[y_{h}^{N}\right]-\Psi_{h}^{N}\right\|^{2}\right]\right).

    Using Jensen inequality to the first term on the RHS of the above inequality, one gets

    \mathbb{E}\left[\left\|\mathbb{E}\left[y^{N}\right]-\mathbb{E}\left[y_{h}^{N}\right]\right\|^{2}\right] \leq \mathbb{E}\left[\mathbb{E}\left[\left\|y^{N}-y_{h}^{N}\right\|^{2}\right]\right] = \mathbb{E}\left[\left\|y^{N}-y_{h}^{N}\right\|^{2}\right].

    Thus, the result can be obtained by Theorem 4.2 and Theorem 4.3. The remaining terms of the LHS of (4.35) can be handled similarly.

    Remark 4.2. Collecting these constants, the RHS of (4.35) can be expressed as

    \frac{C}{J} +C\Delta t^{2}+ Ch^{2k}.

    For comparison purposes, here we use the numerical examples in our recent work [30] to test the various performances of the VMEMC algorithm. Example One is a deterministic heat transfer model with a known exact solution, which is intended to show the convergence rate of time and space, respectively. A random transient heat equation is in Example Two, which is used to test Theorem 4.4 and reveal the effectiveness of the VMEMC algorithm.

    We take y_j = (1+\omega_j)\exp(t+1)\cos(2\pi x_1)\cos(2\pi x_2) as the manufactured solution; here j \in \{1, 2, 3\} , \omega_j\in [0, 1] is a random variable, (x_1, x_2)\in[0, 1]^2 , and t \in [0, 1] . The upper and lower boundary of the domain are zero Neumann boundary; the left and right boundary are Robin boundaries. The Robin coefficient and thermal conductivity diffusion factor are specified by \alpha_j = a_j = 2+(1+\omega_j)\sin(x_1x_2)\sin(t) . The source term, initial condition, and Robin boundary function should be adjusted appropriately.

    In this experiment, the ensemble schemes (3.1) and (3.2) are applied to simulate the ensemble set that includes three samples where \omega_1 = 0.2 , \omega_2 = 0.7 , \omega_3 = 0.8 . We consider \mathcal{E}^{E, j}_{L^2} = \boldsymbol{\||} y_j^n-y^n_{j, h} \boldsymbol{\||}_{2, 0}, j = 1, 2, 3 . We use isometric time partition and linear FEs (k = 1) on uniform space partition. Moreover, we take \Delta t from 1/2 to 1/2^{4} with h = 1/2^{9} to verify the convergence order in the temporal direction. The numerical results on the VMEMC and the VAEMC schemes are shown in Table 1. Meanwhile, we select h from 1/2^{4} to 1/2^{7} with \Delta t = 1/2^{9} to consider the convergence order in space. The resulted numerical errors on the VMEMC and the VAEMC schemes are listed in Table 2.

    Table 1.  Errors and convergence orders on temporal step.
    \Delta t 1/2 1/2^2 1/2^3 1/2^4
    \omega_1=0.2 , h=1/2^9
    \mathcal{E}^{E, 1}_{L^2}\; of\; VAEMC 0.04729 0.02254 0.01091 0.00537
    Rate 1.068 1.046 1.022
    \mathcal{E}^{E, 1}_{L^2}\; of\; VMEMC 0.07742 0.03724 0.01812 0.00893
    Rate 1.055 1.039 1.020
    \omega_1=0.7 , h=1/2^9
    \mathcal{E}^{E, 2}_{L^2}\; of\; VAEMC 0.02952 0.01384 0.00659 0.00317
    Rate 1.093 1.069 1.055
    \mathcal{E}^{E, 2}_{L^2}\; of\; VMEMC 0.01504 0.00711 0.00344 0.00171
    Rate 1.079 1.048 1.009
    \omega_1=0.8 , h=1/2^9
    \mathcal{E}^{E, 3}_{L^2}\; of\; VAEMC 0.05090 0.02377 0.01132 0.00546
    Rate 1.098 1.070 1.049
    \mathcal{E}^{E, 3}_{L^2}\; of\; VMEMC 0.00600 0.00286 0.00134 0.00060
    Rate 1.068 1.093 1.150

     | Show Table
    DownLoad: CSV
    Table 2.  Errors and convergence orders on spatial step.
    h 1/2^4 1/2^5 1/2^6 1/2^7
    \omega_1=0.2 , \Delta t=1/2^9
    \mathcal{E}^{E, 1}_{L^2}\; of\; VAEMC 0.14645 0.03808 0.00965 0.00246
    Rate 1.943 1.979 1.969
    \mathcal{E}^{E, 1}_{L^2}\; of\; VMEMC 0.14648 0.03812 0.00970 0.00251
    Rate 1.942 1.974 1.946
    \omega_1=0.7 , \Delta t=1/2^9
    \mathcal{E}^{E, 2}_{L^2}\; of\; VAEMC 0.20886 0.05419 0.01364 0.00338
    Rate 1.946 1.989 2.011
    \mathcal{E}^{E, 2}_{L^2}\; of\; VMEMC 0.20891 0.05424 0.01370 0.00344
    Rate 1.945 1.985 1.993
    \omega_1=0.8 , \Delta t=1/2^9
    \mathcal{E}^{E, 3}_{L^2}\; of\; VAEMC 0.22145 0.05743 0.01444 0.00356
    Rate 1.946 1.991 2.018
    \mathcal{E}^{E, 3}_{L^2} \; of\; VMEMC 0.22150 0.05749 0.01450 0.00362
    Rate 1.945 1.987 2.001

     | Show Table
    DownLoad: CSV

    Both the VMEMC and the VAEMC algorithms, the convergence orders on temporal direction are about one from Table 1, and coincide with theoretical findings. From Table 2, the rates of convergence on the VMEMC scheme in spatial direction are close to 2, which is higher than the theoretical result (Theorem 4.4). This may be the cause of the higher regularity of source term f and boundary function u in this numerical test.

    The transient heat model (1.1) with random factors is considered on [0, 1]^2 . The upper and lower boundary of the domain are zero Neumann boundary; the left and right boundary are Robin boundaries. Inspired by numerical examples in literature [25,30], we choose the exact solution, the diffusion factor and Robin factor are as follow:

    \begin{equation*} y(t,x,\mathbf{\omega}) = (1+\omega_1+\sum\limits_{i = 2}^{5} \omega_i)\exp(t+1)\cos(2\pi x_1)\cos(2\pi x_2), \end{equation*}
    \begin{equation*} a(x,\omega_1) = 2+(1+\omega_1)\sin(x_1x_2), \end{equation*}
    \begin{equation*} \begin{aligned} \alpha(x,\omega_2,\cdots,\omega_5) = &10+\exp((\omega_2\cos(\pi x_2)+\omega_3\sin(\pi x_2))\exp(-\frac{3}{4})\\ &+(\omega_4\cos(\pi x_1)+\omega_5\sin(\pi x_1))\exp(-\frac{3}{4})). \end{aligned} \end{equation*}

    The random variables \omega_1, \cdots, \omega_5 are independent. \omega_1 is uniformly distributed on [0, 1] . \omega_2, \cdots, \omega_5 are uniformly distributed in the interval [-0.5, 0.5] . The Robin boundary condition, the source term, and initial conditions, are chosen to match the exact solution. Let \mathcal{E}_{L^2} = \sqrt{\frac{1}{J}\sum_{j = 1}^{J}\boldsymbol{\||} y_j^n-y^n_{j, h} \boldsymbol{\||}^{2}_{2, 0}}, j = 1, 2, 3.

    (1) Convergence orders for time and physic space. Analogy to Example One, we use linear FEs (k = 1) on uniform partition physical space and isometric partition in time direction. The corresponding average numerical errors on VAEMC and VMEMC schemes are shown in Table 3 ( h = \frac{1}{2^8}, J = 50 ) and Table 4 ( \Delta t = \frac{1}{2^9}, J = 50 ).

    Table 3.  Average errors and convergence orders on temporal step.
    \Delta t 1/2 1/2^2 1/2^3 1/2^4
    \mathcal{E}_{L^2}\; of\; VAEMC 0.01472 0.00727 0.00358 0.00180
    Rate 1.017 1.020 0.990
    \mathcal{E}_{L^2}\; of\; VMEMC 0.02233 0.01134 0.00574 0.00294
    Rate 0.977 0.982 0.962

     | Show Table
    DownLoad: CSV
    Table 4.  Average errors and convergence orders on spatial step.
    h 1/2^4 1/2^5 1/2^6 1/2^7
    \mathcal{E}_{L^2}\; of\; VAEMC 0.16484 0.04296 0.01084 0.00270
    Rate 1.939 1.985 2.001
    \mathcal{E}_{L^2}\; of\; VMEMC 0.18079 0.04713 0.01191 0.00299
    Rate 1.939 1.984 1.993

     | Show Table
    DownLoad: CSV

    From Table 3, the convergence order on temporal direction is according to Theorem 4.4. Looking at Table 4, the orders of convergence on the VMEMC scheme on spatial direction are close to 2 , which is higher than the theoretical result (Theorem 4.4). This may be the cause of the higher regularity of source term f and boundary function u .

    (2) Convergence order for MC sampling. We take the VMEMC solution choosing J_{0} = 5000 samples as our benchmark, vary the values of J in the VMEMC simulations, and then evaluate the approximation errors based on the benchmark. We repeat such error analysis for R = 12 independent replicas and compute the average of the output errors. Denote the VMEMC solution at time t_n in the r- th independent replica by \Psi_{J, h}^{n, r} \equiv \frac{1}{J} \sum_{j = 1}^{J} y_{j, h}^{n, r}, where y_{j, h}^{n, r} is result of the VMEMC (3.1) in the r- th experiment. Denote

    \mathcal{E}^{*}_{L^2} = \sqrt{\frac{1}{R}\sum\limits_{r = 1}^{R}\boldsymbol{\||} \Psi_{J_0, h}^{n, r}-\Psi_{J, h}^{n, r} \boldsymbol{\||}^{2}_{2, 0}}.

    We run R = 12 times and take the logarithm for the numerical results of \mathcal{E}^{*}_{L^2} at J = 10, 20, 40, 80,160 are depicted in the left subgraph of Figure 1. One group of values of \mathcal{E}^{*}_{L^2} can be seen in Table 5 and the corresponding linear regression model is shown in the right subgraph of Figure 1.

    Figure 1.  Convergence order of the VMEMC algorithm for sampling number J . Left: the 12 numerical results of \mathcal{E}^{*}_{L^2} . Right: the linear regression for one group of numerical results of \mathcal{E}^{*}_{L^2} .
    Table 5.  One group of values of \mathcal{E}^{*}_{L^2} ( h = \Delta t = \frac{1}{2^6} ).
    J 10 20 40 80 160
    \mathcal{E}^{*}_{L^2} 0.03431 0.01763 0.02839 0.01077 0.00661

     | Show Table
    DownLoad: CSV

    We apply linear regression analysis on the numerical errors, which show

    \mathcal{E}^{*}_{L^2} \approx 0.1237 J^{-0.546}.

    These results shows that the rate of convergence with respect to J is close to -0.5 , which coincides with our theoretical results in Theorem 4.4.

    (3) Comparison. Using the mesh size h = \Delta t = 1/2^{6} and the sample size J = 100 , we obtain the mean and the variance of the solutions at t = 0.2 . The results are plotted in Figure 2 and Figure 3 for different algorithms.

    Figure 2.  Three means. Left: mean of the VMEMC solution. Middle: mean of the VAEMC solution. Right: mean of the IFE-MC solution.
    Figure 3.  Variances. Left: variance of the VMEMC solution. Middle: variance of the VAEMC solution. Right: variance of the IFE-MC solution.

    From Figure 2 and Figure 3, it is seen that both VMEMC and VAEMC approximations are excellent.

    To check out the performance of the VMEMC algorithm, we compare the result with that of individual finite element MC (IFE-MC) simulations applying the same groups of sample values. The difference from the mean to the IFE-MC solution is plotted in Figure 4.

    Figure 4.  Difference of the VMEMC mean and the IFE-MC solution.

    The accurate approximation to the IFE-MC implemented for the VMEMC scheme is very nice. From Figure 4, the difference on the order of 10^{-3} signifies that the VMEMC approximation can also achieve basically accurate approximation as the IFE-MC implements.

    (4) Efficiency. To test the efficiency, setting the same time as well as space size, we record the CPU running times of the VMEMC, VAEMC, and MC algorithms in Table 6. Here we apply LU factorization because the discrete system's size is small.

    Table 6.  CPU time (s) ( \Delta t = \frac{1}{2^4}, h = \frac{1}{2^8} ).
    J 10 50 100 200
    CPU time for VMEMC 37.65 150.21 295.59 569.66
    CPU time for VAEMC 38.37 152.47 285.91 607.01
    CPU time for MC 65.27 288.10 559.29 1100.71

     | Show Table
    DownLoad: CSV

    From Table 6, when J > 1 , we can show that both the VMEMC and the VAEMC algorithms take fewer CPU times than the MC algorithm. In contrast with the non-ensemble scheme, these ensemble methods improve the calculational efficiency by about 70%–80%. As the size of the ensemble increases, the efficiency of these ensemble algorithms improves, particularly for the VMEMC algorithm.

    The core idea of the ensemble algorithm is to develop a semi-implicit and semi-explicit temporal discretization scheme for PDEs. It will select a representative \hat{a} of the random coefficient as the implicit term, and take the fluctuation \| a_{j}-\hat{a}\| between random coefficients and this representation as the explicit term. This results in a shared coefficient matrix for each ensemble member at every time step. Consequently, storage requirements are reduced, and solution speed is increased. In general, as long as the relative fluctuation \frac{ \| a_{j}-\hat{a}\|}{ \hat{a}} is not very large, such as \frac{ \| a_{j}-\hat{a}\|}{ \hat{a}} < 1 , the ensemble numerical scheme can be stable. A large amount of literature [17,18,19,20,23,24,30] discussed the numerical stability conditions and error analysis of the ensemble algorithm based on taking the mean as representation \hat{a} .

    Based on taking the maximum value of the random coefficient as representation, an unconditionally stable ensemble scheme, VMEMC, is established in this work. First, we apply the VMEMC algorithm to deal with the transient heat equation with the stochastic diffusion factors and the Robin factors. Without any stability conditions, we prove that the algorithm is numerically stable. Additionally, we also provide the overall error estimate for the algorithm. Moreover, numerical experiments are conducted to demonstrate the application of the ensemble schemes and to validate the established properties of the algorithm.

    Important next steps include accommodating phase changes in the solid material (e.g., transitions to a liquid phase) and incorporating additional physical processes into the boundary conditions (e.g., surface-to-ambient radiation). The stochastic Robin boundary control problem is worth investigating. The VMEMC method can also be developed for nonlinear stochastic parabolic equations.

    Tingfu Yao: Methodology, writing original draft, software; Changlun Ye: Software and supervision; Xianbing Luo: Review, editing, and supervision; Shuwen Xiang: Supervision and validation.

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

    This work is supported by the National Natural Science Foundation of China (Granted No. 11961008, 12361033) and The Natural Science Research Foundation of Education Department of Guizhou Province (Granted NO. QJJ[2024]190).

    The authors declare there is no conflict of interest.



    [1] R. L. Siegel, K. D. Miller and A. Jemal, Cancer statistics, 2017, Ca-Cancer J. Clin., 67 (2017), 7-30.
    [2] R. Ismail-Khan and M. M. Bui, A review of triple-negative breast cancer, Cancer Control: J. Moffitt Cancer Center, 17 (2010), 173-176.
    [3] P. Boyle, Triple-negative breast cancer: Epidemiological considerations and recommendations, Ann. Oncol.: Off. J. Eur. Soc. Med. Oncol., 23 (2012), 7-12.
    [4] N. U. Lin, A. Vanderplas, M. E. Hughes, et al, Clinicopathologic features, patterns of recurrence, and survival among women with triple-negative breast cancer in the National Comprehensive Cancer Network, Cancer, 118 (2012), 5463-5472.
    [5] X. Chen, F. Xia, J. Luo, et al, Postmastectomy radiotherapy reduces locoregional and disease recurrence in patients with stage Ⅱ-Ⅲ triple-negative breast cancer treated with neoadjuvant chemotherapy and mastectomy, OncoTargets Ther., 11 (2018), 1973-1980.
    [6] R. W. Carlson, D. C. Allred, B. O. Anderson, et al, Breast cancer. Clinical practice guidelines in oncology, J. Nat. Compr. Cancer Network: JNCCN, 7 (2009), 122-192.
    [7] W. Gradishar, K. E. Salerno, NCCN guidelines update: Breast cancer, J. Natl. Compr. Cancer Network, 14 (2016), 641-644.
    [8] U. Veronesi, N. Cascinelli, L. Mariani, et al., Twenty-year follow-up of a randomized study comparing breast-conserving surgery with radical mastectomy for early breast cancer, New Eng. J. Med., 347 (2002), 1227-1232.
    [9] B. S. Abdulkarim, J. Cuartero, J. Hanson, et al., Increased risk of locoregional recurrence for women with T1-2N0 triple-negative breast cancer treated with modified radical mastectomy without adjuvant radiation therapy compared with breast-conserving therapy, J. Clin. Oncol.: Off. J. Am. Soc. Clin. Oncol., 29 (2011), 2852-2858.
    [10] F. C. Adkins, A. M. Gonzalez-Angulo, X. Lei, et al, Triple-negative breast cancer is not a contraindication for breast conservation, Ann. Surg. Oncol., 18 (2011), 3164-3173.
    [11] M. C. van Maaren, L. de Munck, G. H. de Bock, et al., 10 year survival after breast-conserving surgery plus radiotherapy compared with mastectomy in early breast cancer in the Netherlands: A population-based study, Lancet Oncol., 17 (2016), 1158-1170.
    [12] X. Chen, X. Yu, J. Chen, et al., Analysis in early stage triple-negative breast cancer treated with mastectomy without adjuvant radiotherapy: Patterns of failure and prognostic factors, Cancer, 119 (2013), 2366-2374.
    [13] Q. X. Chen, X. X. Wang, P. Y. Lin, et al., The different outcomes between breast-conserving surgery and mastectomy in triple-negative breast cancer: A population-based study from the SEER 18 database, Oncotarget, 8 (2017), 4773-4780. doi: 10.18632/oncotarget.13976
    [14] J. K. Horton, R. Jagsi, W. A. Woodward, et al., Breast cancer biology: Clinical implications for breast radiation therapy, Int. J. Radiat. Oncol., Biol., Phys., 100 (2018), 23-37.
    [15] R. C. Brianese, K. D. M. Nakamura, F. Almeida, et al, BRCA1 deficiency is a recurrent event in early-onset triple-negative breast cancer: A comprehensive analysis of germline mutations and somatic promoter methylation, Breast Cancer Res. Treatment., 167 (2018), 803-814.
    [16] S. Y. Phuah, L. M. Looi, N. Hassan, et al., Triple-negative breast cancer and PTEN (phosphatase and tensin homologue) loss are predictors of BRCA1 germline mutations in women with early-onset and familial breast cancer, but not in women with isolated late-onset breast cancer, Breast Cancer Res., 14 (2012), R142.
    [17] M. Kalimutho, K. Parsons, D. Mittal, et al., Targeted therapies for triple-negative breast cancer: Combating a stubborn disease, Trends Pharmacol. Sci., 36 (2015), 822-846. doi: 10.1016/j.tips.2015.08.009
    [18] H. Jia, C. I. Truica, B. Wang, et al., Immunotherapy for triple-negative breast cancer: Existing challenges and exciting prospects, Drug Resist. Updates: Rev. Comment. Antimicrob. Anticancer Chemother., 32 (2017), 1-15.
  • Reader Comments
  • © 2020 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(6040) PDF downloads(865) Cited by(19)

Figures and Tables

Figures(1)  /  Tables(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog