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

Likelihood-Free Dynamical Survival Analysis applied to the COVID-19 epidemic in Ohio

  • The Dynamical Survival Analysis (DSA) is a framework for modeling epidemics based on mean field dynamics applied to individual (agent) level history of infection and recovery. Recently, the Dynamical Survival Analysis (DSA) method has been shown to be an effective tool in analyzing complex non-Markovian epidemic processes that are otherwise difficult to handle using standard methods. One of the advantages of Dynamical Survival Analysis (DSA) is its representation of typical epidemic data in a simple although not explicit form that involves solutions of certain differential equations. In this work we describe how a complex non-Markovian Dynamical Survival Analysis (DSA) model may be applied to a specific data set with the help of appropriate numerical and statistical schemes. The ideas are illustrated with a data example of the COVID-19 epidemic in Ohio.

    Citation: Colin Klaus, Matthew Wascher, Wasiur R. KhudaBukhsh, Grzegorz A. Rempała. Likelihood-Free Dynamical Survival Analysis applied to the COVID-19 epidemic in Ohio[J]. Mathematical Biosciences and Engineering, 2023, 20(2): 4103-4127. doi: 10.3934/mbe.2023192

    Related Papers:

    [1] Fang Wang, Weiguo Li, Wendi Bao, Li Liu . Greedy randomized and maximal weighted residual Kaczmarz methods with oblique projection. Electronic Research Archive, 2022, 30(4): 1158-1186. doi: 10.3934/era.2022062
    [2] Ranran Li, Hao Liu . On global randomized block Kaczmarz method for image reconstruction. Electronic Research Archive, 2022, 30(4): 1442-1453. doi: 10.3934/era.2022075
    [3] Koung Hee Leem, Jun Liu, George Pelekanos . A regularized eigenmatrix method for unstructured sparse recovery. Electronic Research Archive, 2024, 32(7): 4365-4377. doi: 10.3934/era.2024196
    [4] Qin Guo, Binlei Cai . Learning capability of the rescaled pure greedy algorithm with non-iid sampling. Electronic Research Archive, 2023, 31(3): 1387-1404. doi: 10.3934/era.2023071
    [5] John Daugherty, Nate Kaduk, Elena Morgan, Dinh-Liem Nguyen, Peyton Snidanko, Trung Truong . On fast reconstruction of periodic structures with partial scattering data. Electronic Research Archive, 2024, 32(11): 6481-6502. doi: 10.3934/era.2024303
    [6] Zhengyu Liu, Yufei Bao, Changhai Wang, Xiaoxiao Chen, Qing Liu . A fast matrix completion method based on truncated L2,1 norm minimization. Electronic Research Archive, 2024, 32(3): 2099-2119. doi: 10.3934/era.2024095
    [7] Jingqian Xu, Ma Zhu, Baojun Qi, Jiangshan Li, Chunfang Yang . AENet: attention efficient network for cross-view image geo-localization. Electronic Research Archive, 2023, 31(7): 4119-4138. doi: 10.3934/era.2023210
    [8] Daochang Zhang, Dijana Mosić, Liangyun Chen . On the Drazin inverse of anti-triangular block matrices. Electronic Research Archive, 2022, 30(7): 2428-2445. doi: 10.3934/era.2022124
    [9] Huimin Qu, Haiyan Xie, Qianying Wang . Multi-convolutional neural network brain image denoising study based on feature distillation learning and dense residual attention. Electronic Research Archive, 2025, 33(3): 1231-1266. doi: 10.3934/era.2025055
    [10] Jiange Liu, Yu Chen, Xin Dai, Li Cao, Qingwu Li . MFCEN: A lightweight multi-scale feature cooperative enhancement network for single-image super-resolution. Electronic Research Archive, 2024, 32(10): 5783-5803. doi: 10.3934/era.2024267
  • The Dynamical Survival Analysis (DSA) is a framework for modeling epidemics based on mean field dynamics applied to individual (agent) level history of infection and recovery. Recently, the Dynamical Survival Analysis (DSA) method has been shown to be an effective tool in analyzing complex non-Markovian epidemic processes that are otherwise difficult to handle using standard methods. One of the advantages of Dynamical Survival Analysis (DSA) is its representation of typical epidemic data in a simple although not explicit form that involves solutions of certain differential equations. In this work we describe how a complex non-Markovian Dynamical Survival Analysis (DSA) model may be applied to a specific data set with the help of appropriate numerical and statistical schemes. The ideas are illustrated with a data example of the COVID-19 epidemic in Ohio.



    In this paper, we aim to develop a novel weak Galerkin (WG) finite element method for the biharmonic equation that is applicable to non-convex polytopal meshes and eliminates the need for traditional stabilizers. To this aim, we consider the biharmonic equation with Dirichlet and Neumann boundary conditions. The goal is to find an unknown function u satisfying

    Δ2u=f,inΩ,u=ξ,onΩ,un=ν,onΩ, (1.1)

    where ΩRd is an open bounded domain with a Lipschitz continuous boundary Ω. The domain Ω considered in this paper can be of any dimension d2. For the sake of clarity in presentation, we will focus on the case where d=2 throughout this paper. However, the analysis presented here can be readily extended to higher dimensions (d3) without significant modifications.

    The variational formulation of the model problem (1.1) is as follows: Find an unknown function uH2(Ω) satisfying u|Ω=ξ and un|Ω=ν, and the following equation

    2i,j=1(2iju,2ijv)=(f,v),vH20(Ω), (1.2)

    where 2ij denotes the second order partial derivative with respect to xi and xj, and H20(Ω)={vH2(Ω):v|Ω=0,v|Ω=0}.

    The WG finite element method offers an innovative framework for the numerical solution of partial differential equations (PDEs). This approach approximates differential operators within a structure inspired by the theory of distributions, particularly for piecewise polynomial functions. Unlike traditional methods, WG reduces the regularity requirements on approximating functions through the use of carefully designed stabilizers. Extensive studies have highlighted the versatility and effectiveness of WG methods across a wide range of model PDEs, as demonstrated by numerous references [1,2,3,4,5,6] for an incomplete list, establishing WG as a powerful tool in scientific computing. The defining feature of WG methods lies in their innovative use of weak derivatives and weak continuities to construct numerical schemes based on the weak forms of the underlying PDEs. This unique structure provides WG methods with exceptional flexibility, enabling them to address a wide variety of PDEs while ensuring both stability and accuracy in their numerical solutions.

    This paper presents a simplified formulation of the WG finite element method, capable of handling both convex and non-convex elements in finite element partitions. A key innovation of our method is the elimination of stabilizers through the use of higher-degree polynomials for computing weak second-order partial derivatives. This design preserves the size and global sparsity of the stiffness matrix while substantially reducing the programming complexity associated with traditional stabilizer-dependent methods. The method leverages bubble functions as a critical analytical tool, representing a significant improvement over existing stabilizer-free WG methods [7], which are limited to convex polytopal meshes. Our approach is versatile, accommodating arbitrary dimensions and polynomial degrees in the discretization process. In contrast, prior stabilizer-free WG methods [7] often require specific polynomial degree combinations and are restricted to 2D or 3D settings. Theoretical analysis establishes optimal error estimates for the WG approximations in both the discrete H2 norm and an L2 norm.

    This paper is organized as follows. Section 2 provides a brief review of the definition of the weak-second order partial derivative and its discrete counterpart. In Section 3, we introduce an efficient WG scheme that eliminates the need for stabilization terms. Section 4 establishes the existence and uniqueness of the solution. The error equation for the proposed WG scheme is derived in Section 5. Section 6 focuses on obtaining the error estimate for the numerical approximation in the discrete H2 norm, while Section 7 extends the analysis to derive the error estimate in the L2 norm.

    Throughout this paper, we adopt standard notations. Let D be any open, bounded domain with a Lipschitz continuous boundary in Rd. The inner product, semi-norm, and norm in the Sobolev space Hs(D) for any integer s0 are denoted by (,)s,D, ||s,D and s,D respectively. For simplicity, when the domain D is Ω, the subscript D is omitted from these notations. In the case s=0, the notations (,)0,D, ||0,D and 0,D are further simplified as (,)D, ||D and D, respectively.

    This section provides a brief review of the definition of weak weak-second partial derivatives and their discrete counterparts, as introduced in [5].

    Let T be a polygonal element with boundary T. A weak function on T is represented as v={v0,vb,vg}, where v0L2(T), vbL2(T) and vg[L2(T)]2. The first component, v0, denotes the value of v within the interior of T, while the second component, vb, represents the value of v on the boundary of T. The third component vgR2 with components vgi (i=1,2) approximates the gradient v on the boundary T. In general, vb and vg are treated as independent of the traces of v0 and v0, respectively.

    The space of all weak functions on T, denote by W(T), is defined as

    W(T)={v={v0,vb,vg}:v0L2(T),vbL2(T),vg[L2(T)]2}.

    The weak second order partial derivative, 2ij,w, is a linear operator mapping W(T) to the dual space of H2(T). For any vW(T), 2ij,wv is defined as a bounded linear functional on H2(T), given by:

    (2ij,wv,φ)T=(v0,2jiφ)Tvbni,jφT+vgi,φnjT,φH2(T),

    where n, with components ni(i=1,2), represents the unit outward normal vector to T.

    For any non-negative integer r0, let Pr(T) denote the space of polynomials on T with total degree at most r. A discrete weak second order partial derivative, 2ij,w,r,T, is a linear operator mapping W(T) to Pr(T). For any vW(T), 2ij,w,r,Tv is the unique polynomial in Pr(T) satisfying

    (2ij,w,r,Tv,φ)T=(v0,2jiφ)Tvbni,jφT+vgi,φnjT,φPr(T). (2.1)

    For a smooth v0H2(T), applying standard integration by parts to the first term on the right-hand side of (2.1) yields:

    (2ij,w,r,Tv,φ)T=(2ijv0,φ)T(vbv0)ni,jφT+vgiiv0,φnjT, (2.2)

    for any φPr(T).

    Let Th be a finite element partition of the domain ΩR2 into polygons. Assume that Th satisfies the shape regularity condition [8]. Let Eh represent the set of all edges in Th, and denote the set of interior edges by E0h=EhΩ. For any element TTh, let hT be its diameter, and define the mesh size as h=maxTThhT.

    Let k, p and q be integers such that kpq1. For any element TTh, the local weak finite element space is defined as:

    V(k,p,q,T)={{v0,vb,vg}:v0Pk(T),vbPp(e),vg[Pq(e)]2,eT}.

    By combining the local spaces V(k,p,q,T) across all elements TTh and ensuring continuity of vb and vg along the interior edges E0h, we obtain the global weak finite element space:

    Vh={{v0,vb,vg}: {v0,vb,vg}|TV(k,p,q,T),TTh}.

    The subspace of Vh consisting of functions with vanishing boundary values on Ω is defined as:

    V0h={vVh:vb|Ω=0,vg|Ω=0}.

    For simplicity, the discrete weak second order partial derivative 2ij,wv is used to denote the operator 2ij,w,r,Tv defined in (2.1) on each element TTh, as:

    (2ij,wv)|T=2ij,w,r,T(v|T),TTh.

    On each element TTh, let Q0 denote the L2 projection onto Pk(T). On each edge eT, let Qb and Qn denote the L2 projection operators onto Pp(e) and Pq(e), respectively. For any wH2(Ω), the L2 projection into the weak finite element space Vh is denoted by Qhw, defined as:

    (Qhw)|T:={Q0(w|T),Qb(w|T),Qn(w|T)},TTh.

    The simplified WG numerical scheme, free from stabilization terms, for solving the biharmonic equation (1.1) is formulated as follows:

    Weak Galerkin Algorithm 3.1. Find uh={u0,ub,ug}Vh such that ub=Qbξ, ugn=Qnν and {{\mathbf{u}}}_g\cdot{\boldsymbol{\tau}} = Q_n(\nabla\xi\cdot{\boldsymbol{\tau}}) on \partial\Omega , and satisfy:

    \begin{equation} (\partial^2_{w} u_h, \partial^2_{w} v) = (f, v_0), \qquad\forall v = \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h^0, \end{equation} (3.1)

    where {\boldsymbol{\tau}}\in \mathbb R^2 is the tangential direction along \partial\Omega , and the terms are defined as:

    (\partial^2_{w} u_h, \partial^2_{w} v) = \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 (\partial^2_{ij, w} u_h, \partial^2_{ij, w} v)_T,
    (f, v_0) = \sum\limits_{T\in {\mathcal{T}}_h}(f, v_0)_T.

    Recall that {\mathcal{T}}_h is a shape-regular finite element partition of the domain \Omega . Consequently, for any T\in {\mathcal{T}}_h and \phi\in H^1(T) , the following trace inequality holds [8]:

    \begin{equation} \|\phi\|^2_{\partial T} \leq C(h_T^{-1}\|\phi\|_T^2+h_T \|\nabla \phi\|_T^2). \end{equation} (4.1)

    If \phi is a polynomial on the element T\in {\mathcal{T}}_h , a simpler form of the trace inequality applies [8]:

    \begin{equation} \|\phi\|^2_{\partial T} \leq Ch_T^{-1}\|\phi\|_T^2. \end{equation} (4.2)

    For any v = \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h , define the norm:

    \begin{equation} ||| v||| = (\partial^2_{w} v, \partial^2_{ w} v) ^{\frac{1}{2}}, \end{equation} (4.3)

    and introduce the discrete H^2 - semi-norm:

    \begin{equation} \|v\|_{2, h} = \Big( \sum\limits_{T\in {\mathcal{T}}_h} \|\sum\limits_{i,j = 1}^2\partial^2_{ij} v_0\|_T^2+h_T^{-3}\|v_0-v_b\|_{\partial T}^2+h_T^{-1}\|\nabla v_0-{{\mathbf{v}}}_g\|_{\partial T}^2\Big)^{\frac{1}{2}}. \end{equation} (4.4)

    Lemma 4.1. For v = \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h , there exists a constant C such that for i, j = 1, 2 ,

    \|\partial^2_{ij} v_0\|_T\leq C\|\partial^2_{ij, w} v\|_T.

    Proof. Let T\in {\mathcal{T}}_h be a polytopal element with N edges denoted as e_1, \cdots , e_N . Importantly, T can be non-convex. On each edge e_i , construct a linear function l_i(x) satisfying l_i(x) = 0 on e_i as:

    l_i(x) = \frac{1}{h_T}\overrightarrow{AX}\cdot {{\mathbf{n}}}_i,

    where A is a fixed point on e_i , X is any point on e_i , {{\mathbf{n}}}_i is the normal vector to e_i , and h_T is the diameter of T .

    Define the bubble function for T as:

    \Phi_B = l^2_1(x)l^2_2(x)\cdots l^2_N(x) \in P_{2N}(T).

    It is straightforward to verify that \Phi_B = 0 on \partial T . The function \Phi_B can be scaled such that \Phi_B(M) = 1 where M is the barycenter of T . Additionally, there exists a subdomain \hat{T}\subset T such that \Phi_B\geq \rho_0 for some constant \rho_0 > 0 .

    For v = \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h , let r = 2N+k-2 and choose \varphi = \Phi_B \partial^2_{ij} v_0\in P_r(T) in (2.2). This yields:

    \begin{equation} \begin{split} &\; (\partial^2_{ij, w}v, \Phi_B \partial^2_{ij} v_0)_T\\ = &\; (\partial^2_{ij}v_0, \Phi_B \partial^2_{ij} v_0)_T- \langle (v_b-v_0) n_i, \partial_j (\Phi_B \partial^2_{ij} v_0 )\rangle_{\partial T} \\&+\langle v_{gi}-\partial_i v_0, \Phi_B \partial^2_{ij} v_0 n_j\rangle_{\partial T}\\ = &\; (\partial^2_{ij}v_0, \Phi_B \partial^2_{ij} v_0)_T, \end{split} \end{equation} (4.5)

    where we applied \Phi_B = 0 on \partial T .

    Using the domain inverse inequality [8], there exists a constant C such that

    \begin{equation} (\partial^2_{ij} v_0, \Phi_B \partial^2_{ij} v_0)_T \geq C (\partial^2_{ij} v_0, \partial^2_{ij} v_0)_T. \end{equation} (4.6)

    By applying the Cauchy-Schwarz inequality to (4.5) and (4.6), we obtain

    \begin{equation*} \begin{split} & (\partial^2_{ij} v_0, \partial^2_{ij} v_0)_T\leq C (\partial^2_{ij, w} v, \Phi_B \partial^2_{ij} v_0)_T \\ &\leq C \|\partial^2_{ij, w} v\|_T \|\Phi_B \partial^2_{ij} v_0\|_T \leq C \|\partial^2_{ij, w} v\|_T \|\partial^2_{ij} v_0\|_T, \end{split} \end{equation*}

    which implies:

    \|\partial^2_{ij}v_0\|_T\leq C\|\partial^2_{ij, w} v\|_T.

    This completes the proof.

    Remark 4.1. If the polytopal element T is convex, the bubble function in Lemma 4.1 can be simplified to:

    \Phi_B = l_1(x)l_2(x)\cdots l_N(x).

    This simplified bubble function satisfies 1) \Phi_B = 0 on \partial T , 2) there exists a subdomain \hat{T}\subset T such that \Phi_B\geq \rho_0 for some constant \rho_0 > 0 . The proof of Lemma 4.1 follows the same approach, using this simplified bubble function. In this case, we set r = N+k-2 .

    By constructing an edge-based bubble function,

    \varphi_{e_k} = \Pi_{i = 1, \cdots, N, i\neq k}l_i^2(x),

    it can be easily verified that 1) \varphi_{e_k} = 0 on each edge e_i for i \neq k , and 2) there exists a subdomain \widehat{e_k}\subset e_k such that \varphi_{e_k} \geq \rho_1 for some constant \rho_1 > 0 . Let \varphi = (v_b-v_0)l_k \varphi_{e_k} . It is straightforward to verify the following properties: 1) \varphi = 0 on each edge e_i for i = 1, \cdots, N , 2) \nabla \varphi = 0 on each edge e_i for i \neq k , and 3) \nabla \varphi = (v_0-v_b)(\nabla l_k) \varphi_{e_k} = \mathcal{O}(\frac{ (v_0-v_b)\varphi_{e_k}}{h_T}\textbf{C}) on e_k , where \textbf{C} is a constant vector.

    Lemma 4.2. [9] For \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h , let \varphi = (v_b-v_0)l_k \varphi_{e_k} . The following inequality holds:

    \begin{equation} \|\varphi\|_T ^2 \leq Ch_T \int_{e_k}(v_b-v_0)^2ds. \end{equation} (4.7)

    Lemma 4.3. For \{v_0, v_b, {{\mathbf{v}}}_g\}\in V_h , let \varphi = (v_{gi} -\partial_i v_0) \varphi_{e_k} . The following inequality holds:

    \begin{equation} \|\varphi\|_T ^2 \leq Ch_T \int_{e_k}(v_{gi} -\partial_i v_0 )^2ds. \end{equation} (4.8)

    Proof. Define the extension of {{\mathbf{v}}}_g , originally defined on the edge e_k , to the entire polytopal element T as:

    {{\mathbf{v}}}_g(X) = {{\mathbf{v}}}_g(Proj_{e_k} (X)),

    where X = (x_1, x_2) is any point in T , Proj_{e_k} (X) denotes the orthogonal projection of X onto the plane H\subset\mathbb R^2 containing e_k . If Proj_{e_k} (X) is not on e_k , {{\mathbf{v}}}_g(Proj_{e_k} (X)) is defined as the extension of {{\mathbf{v}}}_g from e_k to H . The extension preserves the polynomial nature of {{\mathbf{v}}}_g as demonstrated in [9].

    Let v_{trace} denote the trace of v_0 on e_k . Define its extension to T as:

    v_{trace} (X) = v_{trace}(Proj_{e_k} (X)).

    This extension is also polynomial, as demonstrated in [9].

    Let \varphi = (v_{gi} -\partial_i v_0) \varphi_{e_k} . Then,

    \begin{equation*} \begin{split} \|\varphi\|^2_T = \int_T \varphi^2dT = &\; \int_T ((v_{gi} -\partial_i v_0 )(X) \varphi_{e_k})^2dT\\ \leq &\; Ch_T \int_{e_k} ((v_{gi} -\partial_i v_0 )(Proj_{e_k} (X)) \varphi_{e_k})^2dT\\ \\\leq &\; Ch_T \int_{e_k} (v_{gi} -\partial_i v_0 ) ^2ds, \end{split} \end{equation*}

    where we used the facts that 1) \varphi_{e_k} = 0 on each edge e_i for i \neq k , 2) there exists a subdomain \widehat{e_k}\subset e_k such that \varphi_{e_k} \geq \rho_1 for some constant \rho_1 > 0 , and applied the properties of the projection.

    This completes the proof of the lemma.

    Lemma 4.4. There exist two positive constants, C_1 and C_2 , such that for any v = \{v_0, v_b, {{\mathbf{v}}}_g\} \in V_h , the following equivalence holds:

    \begin{equation} C_1\|v\|_{2, h}\leq ||| v||| \leq C_2\|v\|_{2, h}. \end{equation} (4.9)

    Proof. Consider the edge-based bubble function defined as

    \varphi_{e_k} = \Pi_{i = 1, \cdots, N, i\neq k}l_i^2(x).

    First, extend v_b from the edge e_k to the element T . Similarly, let v_{trace} denote the trace of v_0 on the edge e_k and extend v_{trace} to the element T . For simplicity, we continue to use v_b and v_0 to represent their respective extensions. Details of these extensions can be found in Lemma 4.3. Substituting \varphi = (v_b-v_0)l_k\varphi_{e_k} into (2.2), we obtain

    \begin{equation} \begin{split} (\partial^2_{ij, w}v, \varphi)_T = &\; (\partial^2_{ij}v_0, \varphi)_T- \langle (v_b-v_0) n_i, \partial_j \varphi \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, \varphi n_j\rangle_{\partial T}\\ = &\; (\partial^2_{ij}v_0, \varphi)_T + Ch_T^{-1}\int_{e_k} |v_b- v_0|^2 \varphi_{e_k} ds, \end{split} \end{equation} (4.10)

    where we used 1) \varphi = 0 on each edge e_i for i = 1 , \cdots , N , 2) \nabla \varphi = 0 on each edge e_i for i \neq k , and 3) \nabla \varphi = (v_0-v_b)(\nabla l_k) \varphi_{e_k} = \mathcal{O}(\frac{ (v_0-v_b)\varphi_{e_k}}{h_T}\textbf{C}) on e_k , where \textbf{C} is a constant vector.

    Recall that 1) \varphi_{e_k} = 0 on each edge e_i for i \neq k , and 2) there exists a subdomain \widehat{e_k}\subset e_k such that \varphi_{e_k} \geq \rho_1 for some constant \rho_1 > 0 . Using Cauchy-Schwarz inequality, the domain inverse inequality [8], (4.10) and Lemma 4.2, we deduce:

    \begin{equation*} \begin{split} \int_{e_k}|v_b- v_0|^2 ds\leq &\; C\int_{e_k} |v_b- v_0|^2 \varphi_{e_k} ds \\ \leq&\; C h_T(\|\partial^2_{ij, w} v\|_T+\|\partial^2_{ij} v_0\|_T){ \| \varphi\|_T}\\ \leq &\; C { h_T^{\frac{3}{2}}} (\|\partial^2_{ij, w} v\|_T+\|\partial^2_{ij} v_0\|_T){ (\int_{e_k}|v_b- v_0|^2ds)^{\frac{1}{2}}}, \end{split} \end{equation*}

    which, from Lemma 4.1, leads to:

    \begin{equation} h_T^{-3}\int_{e_k}|v_b- v_0|^2 ds \leq C (\|\partial^2_{ij, w} v\|^2_T+\|\partial^2_{ij} v_0\|^2_T)\leq C\|\partial^2_{ij, w} v\|^2_T. \end{equation} (4.11)

    Next, extend {{\mathbf{v}}}_g from the edge e_k to the element T , denoting the extension by the same symbol for simplicity. Details of this extension are in Lemma 4.3. Substituting \varphi = (v_{gi}-\partial_i v_0)\varphi_{e_k} into (2.2), we obtain:

    \begin{equation} \begin{split} &\; (\partial^2_{ij, w}v, \varphi)_T\\ = &\; (\partial^2_{ij}v_0, \varphi)_T- \langle (v_b-v_0) n_i, \partial_j \varphi \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, \varphi n_j\rangle_{\partial T}\\ = & \; (\partial^2_{ij}v_0, \varphi)_T - \langle (v_b-v_0) n_i, \partial_j \varphi \rangle_{\partial T}+\int_{e_k} |v_{gi}-\partial_i v_0|^2 \varphi_{e_k}ds, \end{split} \end{equation} (4.12)

    where we used \varphi_{e_k} = 0 on edge e_i for i \neq k , and the fact that there exists a sub-domain \widehat{e_k}\subset e_k such that \varphi_{e_k} \geq \rho_1 for some constant \rho_1 > 0 . This, together with Cauchy-Schwarz inequality, the domain inverse inequality [8], the inverse inequality, the trace inequality (4.2), (4.11) and Lemma 4.3, gives

    \begin{equation*} \begin{split} & \; \int_{e_k}|v_{gi}-\partial_i v_0|^2 ds\\\leq &\; C \int_{e_k}|v_{gi}-\partial_i v_0|^2 \varphi_{e_k}ds\\ \leq &\; C (\|\partial^2_{ij, w} v\|_T+\|\partial^2_{ij} v_0\|_T)\| \varphi\|_T+ C\|v_0-v_b\|_{\partial T}\|\partial_j \phi\|_{\partial T}\\ \leq & \; C h_T^{\frac{1}{2}} (\|\partial^2_{ij, w} v\|_T+\|\partial^2_{ij} v_0\|_T)(\int_{e_k}|v_{gi}-\partial_i v_0|^2ds)^{\frac{1}{2}} + C h_T^{\frac{3}{2}} \|\partial^2_{ij, w} v\|_T h_T^{-1}(\int_{e_k}|v_{gi}-\partial_i v_0|^2ds)^{\frac{1}{2}}. \end{split} \end{equation*}

    Applying Lemma 4.1, gives

    \begin{equation} h_T^{-1}\int_{e_k}|v_{gi}-\partial_i v_0|^2 ds \leq C (\|\partial^2_{ij, w} v\|^2_T+\|\partial^2_{ij} v_0\|^2_T)\leq C\|\partial^2_{ij, w} v\|^2_T. \end{equation} (4.13)

    Using Lemma 4.1, Eqs (4.11), (4.13), (4.3) and (4.4), we deduce:

    C_1\|v\|_{2, h}\leq ||| v|||.

    Finally, using the Cauchy-Schwarz inequality, inverse inequalities, and the trace inequality (4.2) in (2.2), we derive:

    \begin{equation*} \begin{split} \Big| (\partial^2_{ij, w}v, \varphi)_T\Big| \leq &\; \|\partial^2_{ij}v_0\|_T \| \varphi\|_T+ \|(v_b-v_0) n_i\|_{\partial T} \| \partial_j\varphi\|_{\partial T}+\|v_{gi}-\partial_i v_0\|_{\partial T} \|\varphi n_j\|_{\partial T} \\ \leq &\; \|\partial^2_{ij}v_0\|_T \| \varphi\|_T+ h_T^{-\frac{3}{2}}\|v_b-v_0\|_{\partial T} \| \varphi\|_{ T}+h_T^{-\frac{1}{2}}\|v_{gi}-\partial_i v_0\|_{\partial T} \|\varphi \|_{T}, \end{split} \end{equation*}

    which gives:

    \| \partial^2_{ij, w}v\|_T^2\leq C( \|\partial^2_{ij}v_0\|^2_T + h_T^{-3}\|v_b-v_0\|^2_{\partial T}+h_T^{-1}\|v_{gi}-\partial_i v_0\|^2_{\partial T}),

    and further gives

    ||| v||| \leq C_2\|v\|_{2, h}.

    This completes the proof.

    Theorem 4.5. The WG scheme 3.1 admits a unique solution.

    Proof. Assume that u_h^{(1)}\in V_h and u_h^{(2)}\in V_h are two distinct solutions of the WG scheme 3.1. Define \eta_h = u_h^{(1)}-u_h^{(2)}\in V_h^0 . Then, \eta_h satisfies

    ( \partial^2_{ij, w} \eta_h, \partial^2_{ij, w} v) = 0, \qquad \forall v\in V_h^0.

    Choosing v = \eta_h yields ||| \eta_h||| = 0 . From the equivalence of norms in (4.9), it follows that \|\eta_h\|_{2, h} = 0 , which yields \partial^2_{ij} \eta_0 = 0 for i, j = 1, 2 on each T , \eta_0 = \eta_b and \nabla \eta_0 = {\boldsymbol{\eta}}_g on each \partial T . Consequently, \eta_0 is a linear function on each element T and \nabla \eta_0 = C on each T .

    Since \nabla \eta_0 = {\boldsymbol{\eta}}_g on each \partial T , it follows that \nabla \eta_0 is continuous across the entire domain \Omega . Thus, \nabla \eta_0 = C throughout \Omega . Furthermore, the condition {\boldsymbol{\eta}}_g = 0 on \partial\Omega implies \nabla \eta_0 = 0 in \Omega and {\boldsymbol{\eta}}_g = 0 on each \partial T . Therefore, \eta_0 is a constant on each element T .

    Since \eta_0 = \eta_b on \partial T , the continuity of \eta_0 over \Omega implies \eta_0 is globally constant. From \eta_b = 0 on \partial\Omega , we conclude \eta_0 = 0 throughout \Omega . Consequently, \eta_b = \eta_0 = 0 on each \partial T , which implies \eta_h\equiv 0 in \Omega . Thus, u_h^{(1)}\equiv u_h^{(2)} , proving the uniqueness of the solution.

    Let Q_r denote the L^2 projection operator onto the finite element space of piecewise polynomials of degree at most r .

    Lemma 5.1. The following property holds:

    \begin{equation} \partial^2_{ij, w}u = Q_r(\partial^2_{ij} u), \qquad \forall u\in H^2(T). \end{equation} (5.1)

    Proof. For any u\in H^2(T) , using (2.2), we have

    \begin{equation*} \begin{split} &\; (\partial^2_{ij, w}u, \varphi)_T\\ = &\; (\partial^2_{ij}u, \varphi)_T- \langle (u|_{\partial T}-u|_T) n_i, \partial_j \varphi \rangle_{\partial T}+\langle (\nabla u|_{\partial T})_{i} -\partial_i (u|_{T}), \varphi n_j\rangle_{\partial T}\\ = &\; (\partial^2_{ij}u, \varphi)_T = (Q_r(\partial^2_{ij}u), \varphi)_T, \end{split} \end{equation*}

    for all \varphi\in P_r(T) . This completes the proof.

    Let u be the exact solution of the biharmonic equation (1.1), and u_h \in V_{h} its numerical approximation obtained from the WG scheme 3.1. The error function, denoted by e_h , is defined as

    \begin{equation} e_h = u-u_h. \end{equation} (5.2)

    Lemma 5.2. The error function e_h defined in (5.2) satisfies the following error equation:

    \begin{equation} (\partial_{w}^2 e_h, \partial_{w}^2 v) = \ell (u, v), \qquad \forall v\in V_h^0, \end{equation} (5.3)

    where

    \ell (u, v) = \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (v_b-v_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, (Q_r-I) \partial_{ij}^2 u n_j\rangle_{\partial T}.

    Proof. Using (5.1), standard integration by parts, and substituting \varphi = Q_r \partial_{ij}^2 u into (2.2), we obtain

    \begin{equation} \begin{split} &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 u, \partial_{ij, w}^2 v)_T\\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(Q_r \partial_{ij}^2 u, \partial_{ij, w}^2 v)_T\\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial^2_{ij}v_0, Q_r \partial_{ij}^2 u)_T- \langle (v_b-v_0) n_i, \partial_j (Q_r \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, Q_r \partial_{ij}^2 u n_j\rangle_{\partial T}\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial^2_{ij}v_0, \partial_{ij}^2 u)_T- \langle (v_b-v_0) n_i, \partial_j (Q_r \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, Q_r \partial_{ij}^2 u n_j\rangle_{\partial T}\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 ((\partial^2_{ij})^2u, v_0)_T+\langle \partial_{ij}^2 u, \partial_i v_0\cdot n_j\rangle_{\partial T}-\langle \partial_j(\partial_{ij}^2u)\cdot n_i, v_0\rangle_{\partial T}\\ &- \langle (v_b-v_0) n_i, \partial_j (Q_r \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, Q_r \partial_{ij}^2 u n_j\rangle_{\partial T}\\ = &\; (f, v_0)+\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (v_b-v_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, (Q_r-I) \partial_{ij}^2 u n_j\rangle_{\partial T}, \end{split} \end{equation} (5.4)

    where we used (1.1), \partial_{ij}^2 v_0\in P_{k-2}(T) and r = 2N+k-2\geq k-2 , \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial_{ij}^2 u, v_{gi}\cdot n_j\rangle_{\partial T} = \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial_{ij}^2 u, v_{gi}\cdot n_j\rangle_{\partial \Omega} = 0 since v_{gi} = 0 on \partial \Omega , and \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial_j(\partial_{ij}^2u)\cdot n_i, v_b\rangle_{\partial T} = \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial_j(\partial_{ij}^2u)\cdot n_i, v_b\rangle_{\partial \Omega} = 0 since v_{b} = 0 on \partial \Omega .

    Subtracting (3.1) from (5.4) yields

    \begin{equation*} \begin{split} &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 e_h, \partial_{ij, w}^2 v)_T\\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (v_b-v_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 u) \rangle_{\partial T}+\langle v_{gi}-\partial_i v_0, (Q_r-I) \partial_{ij}^2 u n_j\rangle_{\partial T}. \end{split} \end{equation*}

    This concludes the proof.

    Lemma 6.1. [5] Let {\mathcal{T}}_h be a finite element partition of the domain \Omega satisfying the shape regularity assumption specified in [8]. For any 0\leq s \leq 2 and 1\leq m \leq k , the following estimates hold:

    \begin{eqnarray} \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 h_T^{2s}\|\partial_{ij}^2 u- Q_r \partial_{ij}^2 u\|^2_{s,T}&\leq& C h^{2(m-1)}\|u\|^2_{m+1}, \end{eqnarray} (6.1)
    \begin{eqnarray} \sum\limits_{T\in {\mathcal{T}}_h}h_T^{2s}\|u- Q _0u\|^2_{s,T}&\leq& C h^{2(m+1)}\|u\|^2_{m+1}. \end{eqnarray} (6.2)

    Lemma 6.2. If u\in H^{k+1}(\Omega) , then there exists a constant C such that

    \begin{equation} ||| u-Q_hu ||| \leq Ch^{k-1}\|u\|_{k+1}. \end{equation} (6.3)

    Proof. Utilizing (2.2), the trace inequalities (4.1) and (4.2), the inverse inequality, and the estimate (6.2) for m = k and s = 0, 1, 2 , we analyze the following summation for any \varphi\in P_r(T) :

    \begin{equation*} \begin{split} &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(\partial^2_{ij, w}(u-Q_hu), \varphi)_T\\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2 (\partial^2_{ij}(u-Q_0u), \varphi)_T- \langle (Q_0u-Q_bu) n_i, \partial_j \varphi \rangle_{\partial T}\\&+\langle (\partial_i u- Q_n (\partial_i u))-\partial_i (u-Q_0u), \varphi n_j\rangle_{\partial T}\\ \leq &\; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2\|\partial^2_{ij}(u-Q_0u)\|^2_T\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h} \|\varphi\|_T^2\Big)^{\frac{1}{2}}\\& + \Big(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i = 1}^2\|(Q_0u-Q_bu) n_i\|_{\partial T} ^2\Big)^{\frac{1}{2}}\Big(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{ j = 1}^2\|\partial_j \varphi\|_{\partial T}^2\Big)^{\frac{1}{2}}\\ &+ \Big(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i = 1}^2\| \partial_i (Q_0u)- Q_n (\partial_i u) \|_{\partial T} ^2\Big)^{\frac{1}{2}}\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{ j = 1}^2 \| \varphi n_j\|_{\partial T}^2\Big)^{\frac{1}{2}}\\ \leq &\; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2\|\partial^2_{ij}(u-Q_0u)\|^2_T\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h} \|\varphi\|_T^2\Big)^{\frac{1}{2}}\\& + \Big(\sum\limits_{T\in {\mathcal{T}}_h} h_T^{-1}\| Q_0u- u \|_{ T}+h_T \| Q_0u- u \|_{1, T} ^2\Big)^{\frac{1}{2}}\Big(\sum\limits_{T\in {\mathcal{T}}_h} h_T^{-3}\| \varphi\|_{ T}^2\Big)^{\frac{1}{2}}\\ &+ \Big(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i = 1}^2h_T^{-1}\| \partial_i (Q_0u)- \partial_i u \|_{ T} ^2+h_T \| \partial_i (Q_0u)- \partial_i u \|_{1, T} ^2\Big)^{\frac{1}{2}}\Big(\sum\limits_{T\in {\mathcal{T}}_h} h_T^{-1} \| \varphi \|_{T}^2\Big)^{\frac{1}{2}}\\ \leq&\; Ch^{k-1}\|u\|_{k+1}\Big(\sum\limits_{T\in {\mathcal{T}}_h} \|\varphi\|_T^2\Big)^{\frac{1}{2}}. \end{split} \end{equation*}

    Letting \varphi = \partial^2_{ij, w}(u-Q_hu) gives

    \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial^2_{ij, w}(u-Q_hu), \partial^2_{ij, w}(u-Q_hu))_T\leq Ch^{k-1}\|u\|_{k+1}||| u-Q_hu |||.

    This completes the proof.

    Theorem 6.3. Suppose the exact solution u of the biharmonic equation (1.1) satisfies u\in H^{k+1}(\Omega) . Then, the error estimate satisfies:

    \begin{equation} ||| u-u_h||| \leq Ch^{k-1}\|u\|_{k+1}. \end{equation} (6.4)

    Proof. Note that r\geq 1 . For the first term on the right-hand side of the error equation (5.3), using Cauchy-Schwarz inequality, the trace inequality (4.1), the estimate (6.1) with m = k and s = 1, 2 , and (4.9), we have

    \begin{equation} \begin{split} &\; \Big|\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (v_b-v_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 u) \rangle_{\partial T}\Big|\\ \leq &\; C(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2h_T^{-3}\|(v_b-v_0) n_i\|^2_{\partial T} )^{\frac{1}{2}} \cdot(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i,j = 1}^2h_T^3\|\partial_j ((Q_r-I) \partial_{ij}^2 u) \|^2_{\partial T})^{\frac{1}{2}}\\\leq &\; C \| v\|_{2,h} (\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i,j = 1}^2h_T^2\|\partial_j ((Q_r-I) \partial_{ij}^2 u) \|^2_{T}+h_T^4\|\partial_j ((Q_r-I) \partial_{ij}^2 u) \|^2_{1, T})^{\frac{1}{2}}\\ \leq &\; Ch^{k-1} \|u\|_{k+1} ||| v|||. \end{split} \end{equation} (6.5)

    For the second term on the right-hand side of the error equation (5.3), using the Cauchy-Schwarz inequality, the trace inequality (4.1), the estimate (6.1) with m = k and s = 0, 1 , and (4.9), we have

    \begin{equation} \begin{split} &\; \Big|\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 \langle v_{gi}-\partial_i v_0, (Q_r-I) \partial_{ij}^2 u n_j\rangle_{\partial T}\Big|\\ \leq &\; C(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2h_T^{-1}\| v_{gi}-\partial_i v_0\|^2_{\partial T} )^{\frac{1}{2}} (\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i,j = 1}^2h_T\|(Q_r-I) \partial_{ij}^2 u n_j\|^2_{\partial T})^{\frac{1}{2}}\\ \leq &\; C \| v\|_{2,h} (\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i,j = 1}^2\|(Q_r-I) \partial_{ij}^2 u n_j\|^2_{ T}+h_T^2\|(Q_r-I) \partial_{ij}^2 u n_j\|^2_{1, T})^{\frac{1}{2}}\\ \leq &\; C \| v\|_{2,h} h^{k-1}\|u\|_{k+1}\\ \leq &\; Ch^{k-1}\|u\|_{k+1} ||| v|||. \end{split} \end{equation} (6.6)

    Substituting (6.5) and (6.6) into (5.3) gives

    \begin{equation} (\partial^2_{ij, w} e_h, \partial^2_{ij, w} v)\leq Ch^{k-1} \|u\|_{k+1} ||| v|||. \end{equation} (6.7)

    Using Cauchy-Schwarz inequality, letting v = Q_hu-u_h in (6.7), the error estimate (6.3) gives

    \begin{equation*} \begin{split} & \; ||| u-u_h|||^2\\ = &\; \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial^2_{ij, w} (u-u_h), \partial^2_{ij, w} (u-Q_hu))_T+(\partial^2_{ij, w} (u-u_h), \partial^2_{ij, w} (Q_hu-u_h))_T\\ \leq &\; ||| u-u_h ||| ||| u-Q_hu |||+ Ch^{k-1} \|u\|_{k+1} ||| Q_hu-u_h||| \\ \leq &\; ||| u-u_h ||| ||| u-Q_hu ||| + Ch^{k-1} \|u\|_{k+1} (||| Q_hu-u|||+||| u-u_h|||) \\ \leq &\; ||| u-u_h ||| ||| u-Q_hu ||| + Ch^{k-1}\|u\|_{k+1} h^{k-1} \|u\|_{k+1} +Ch^{k-1} \|u\|_{k+1} ||| u-u_h|||, \end{split} \end{equation*}

    which further gives

    \begin{equation*} \begin{split} ||| u-u_h||| \leq ||| u-Q_hu |||+Ch^{k-1} \|u\|_{k+1} \leq Ch^{k-1} \|u\|_{k+1}. \end{split} \end{equation*}

    This completes the proof.

    To derive the error estimate in the L^2 norm, we use the standard duality argument. The error is expressed as e_h = u-u_h = \{e_0, e_b, {{\mathbf{e}}}_g\} , and we define \zeta_h = Q_hu - u_h = \{\zeta_0, \zeta_b, {{\boldsymbol{\zeta}}}_g\}\in V_h^0 . Consider the dual problem associated with the biharmonic equation (1.1), which seeks a function w \in H_0^2(\Omega) satisfying:

    \begin{equation} \begin{split} \Delta^2 w& = \zeta_0, \qquad \text{in}\ \Omega,\\ w& = 0, \qquad \text{on}\ \partial\Omega,\\ \frac{\partial w}{\partial {{\mathbf{n}}}}& = 0, \qquad \text{on}\ \partial\Omega. \end{split} \end{equation} (7.1)

    We assume the following regularity condition for the dual problem:

    \begin{equation} \|w\|_4\leq C\|\zeta_0\|. \end{equation} (7.2)

    Theorem 7.1. Let u\in H^{k+1}(\Omega) be the exact solution of the biharmonic equation (1.1), and let u_h\in V_h denote the numerical solution obtained using the weak Galerkin scheme 3.1. Assume that the H^4 -regularity condition (7.2) holds. Then, there exists a constant C such that

    \begin{equation*} \|e_0\|\leq Ch^{k+1}\|u\|_{k+1}. \end{equation*}

    Proof. Testing the dual problem (7.1) with \zeta_0 and applying integration by parts, we derive:

    \begin{equation} \begin{split} \|\zeta_0\|^2 = &\; (\Delta^2 w, \zeta_0)\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(\partial^2_{ij} w, \partial^2_{ij}\zeta_0)_T-\langle \partial^2_{ij} w, \partial_i\zeta_0 \cdot n_j \rangle_{\partial T}+\langle \partial_j(\partial^2_{ij} w)\cdot n_i, \zeta_0 \rangle_{\partial T}\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(\partial^2_{ij} w, \partial^2_{ij}\zeta_0)_T-\langle \partial^2_{ij} w, (\partial_i\zeta_0-\zeta_{gi}) \cdot n_j \rangle_{\partial T}+\langle \partial_j(\partial^2_{ij} w)\cdot n_i, \zeta_0-\zeta_b \rangle_{\partial T}, \end{split} \end{equation} (7.3)

    where we used \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial^2_{ij} w, \zeta_{gi} \cdot n_j \rangle_{\partial T} = \sum_{i, j = 1}^2\langle \partial^2_{ij} w, \zeta_{gi} \cdot n_j \rangle_{\partial \Omega} = 0 due to {{\boldsymbol{\zeta}}}_g = 0 on \partial\Omega , and \sum_{T\in {\mathcal{T}}_h} \sum_{i, j = 1}^2 \langle \partial_j(\partial^2_{ij} w)\cdot n_i, \zeta_b \rangle_{\partial T} = \sum_{i, j = 1}^2\langle \partial_j(\partial^2_{ij} w)\cdot n_i, \zeta_b \rangle_{\partial \Omega} = 0 due to \zeta_b = 0 on \partial\Omega .

    Letting u = w and v = \zeta_h in (5.4) gives

    \begin{equation*} \begin{split} & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 w, \partial_{ij, w}^2 \zeta_h)_T \\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij}^2 w, \partial_{ij}^2 \zeta_0)_T - \langle (\zeta_b-\zeta_0) n_i, \partial_j (Q_r \partial_{ij}^2 w) \rangle_{\partial T}+\langle \zeta_{gi}-\partial_i \zeta_0, Q_r \partial_{ij}^2 w n_j\rangle_{\partial T}, \end{split} \end{equation*}

    which is equivalent to

    \begin{equation*} \begin{split} & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij}^2 w, \partial_{ij}^2 \zeta_0)_T \\ = &\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 w, \partial_{ij, w}^2 \zeta_h)_T +\langle (\zeta_b-\zeta_0) n_i, \partial_j (Q_r \partial_{ij}^2 w) \rangle_{\partial T}-\langle \zeta_{gi}-\partial_i \zeta_0, Q_r \partial_{ij}^2 w n_j\rangle_{\partial T}. \end{split} \end{equation*}

    Substituting the above equation into (7.3) and using (5.3) gives

    \begin{equation} \begin{split} \|\zeta_0\|^2 = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 w, \partial_{ij, w}^2 \zeta_h)_T +\langle (\zeta_b-\zeta_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 w) \rangle_{\partial T}\\&-\langle \zeta_{gi} -\partial_i \zeta_0, (Q_r-I) \partial_{ij}^2 w n_j\rangle_{\partial T}\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 w, \partial_{ij, w}^2 e_h)_T+(\partial_{ij, w}^2 w, \partial_{ij, w}^2 (Q_hu-u))_T-\ell(w, \zeta_h)\\ = & \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 Q_hw, \partial_{ij, w}^2 e_h)_T+(\partial_{ij, w}^2 (w-Q_hw), \partial_{ij, w}^2 e_h)_T\\&+(\partial_{ij, w}^2 w, \partial_{ij, w}^2 (Q_hu-u))_T-\ell(w, \zeta_h)\\ = &\; \ell(u, Q_hw) + \sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2(\partial_{ij, w}^2 (w-Q_hw), \partial_{ij, w}^2 e_h)_T\\&+(\partial_{ij, w}^2w, \partial_{ij, w}^2(Q_hu-u))_T-\ell(w, \zeta_h)\\ = &\; J_1+J_2+J_3+J_4. \end{split} \end{equation} (7.4)

    We will estimate the four terms J_i \; (i = 1 , \cdots , 4) on the last line of (7.4) individually.

    For J_1 , using the Cauchy-Schwarz inequality, the trace inequality (4.1), the inverse inequality, the estimate (6.1) with m = k and s = 0, 1, 2 , the estimate (6.2) with m = 3 and s = 0, 1, 2 , gives

    \begin{equation} \begin{split} &J_1 = \ell(u, Q_hw)\\ \leq &\; \Big|\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (Q_bw-Q_0w) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 u) \rangle_{\partial T}\\&+\langle Q_n(\partial_i w) -\partial_i Q_0w, (Q_r-I) \partial_{ij}^2 u n_j\rangle_{\partial T}\Big|\\ \leq&\; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2\|(Q_bw-Q_0w) n_i\|_{\partial T}^2\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2\|\partial_j ((Q_r-I) \partial_{ij}^2 u)\|_{\partial T}^2\Big)^{\frac{1}{2}} \\ &+\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2\|Q_n(\partial_i w) -\partial_i Q_0w\|_{\partial T}^2\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2\|(Q_r-I) \partial_{ij}^2 u n_j\|_{\partial T}^2\Big)^{\frac{1}{2}} \\ \leq&\; \Big(\sum\limits_{T\in {\mathcal{T}}_h} h_T^{-1}\| w-Q_0w \|_{ T}^2+h_T \|w-Q_0w \|_{1, T}^2\Big)^{\frac{1}{2}} \\&\cdot\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2h_T^{-1}\|\partial_j ((Q_r-I) \partial_{ij}^2 u)\|_{T}^2+h_T\|\partial_j ((Q_r-I) \partial_{ij}^2 u)\|_{1, T}^2\Big)^{\frac{1}{2}} \\ &+\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2h_T^{-1}\| \partial_i w -\partial_i Q_0w\|_{T}^2+h_T \| \partial_i w -\partial_i Q_0w\|_{1, T}^2\Big)^{\frac{1}{2}} \\&\cdot \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2h_T^{-1}\|(Q_r-I) \partial_{ij}^2 u n_j\|_{T}^2+h_T\|(Q_r-I) \partial_{ij}^2 u n_j\|_{1, T}^2\Big)^{\frac{1}{2}} \\ \leq &\; Ch^{k+1}\|u\|_{k+1}\|w\|_4. \end{split} \end{equation} (7.5)

    For J_2 , using Cauchy-Schwarz inequality, (6.3) with k = 3 and (6.4) gives

    \begin{equation} \begin{split} J_2\leq ||| w-Q_hw||| ||| e_h|||\leq Ch^{k-1}\|u\|_{k+1}h^2\|w\|_4\leq Ch^{k+1}\|u\|_{k+1}\|w\|_4. \end{split} \end{equation} (7.6)

    For J_3 , denote by Q^1 a L^2 projection onto P_1(T) . Using (2.1) gives

    \begin{equation} \begin{split} &\; (\partial^2_{ij, w}(Q_hu-u), Q^1\partial^2_{ij, w} w)_T\\ = &\; (Q_0u-u, \partial^2_{ji} ( Q^1\partial^2_{ij, w} w))_T-\langle Q_bu-u, \partial_j (Q^1\partial^2_{ij, w} w)\rangle_{\partial T}+ \langle Q_n(\partial_i u)-\partial_i u, Q^1\partial^2_{ij, w} w n_j\rangle_{\partial T}\\ = &\; 0, \end{split} \end{equation} (7.7)

    where we used \partial^2_{ji} (Q^1\partial^2_{ij, w} w) = 0 , \partial_j (Q^1\partial^2_{ij, w} w) = C and the property of the projection operators Q_b and Q_n and p\geq q\geq 1 .

    Using (7.7), Cauchy-Schwarz inequality, (5.1) and (6.3), gives

    \begin{equation} \begin{split} J_3\leq &\; |\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(\partial^2_{ij, w} w, \partial^2_{ij, w} (Q_hu-u))_T| \\ = &\; |\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(\partial^2_{ij, w} w-Q^1\partial^2_{ij, w} w, \partial^2_{ij, w} (Q_hu-u))_T|\\ = &\; |\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2(Q_r\partial^2_{ij} w-Q^1 Q_r\partial^2_{ij} w, \partial^2_{ij, w} (Q_hu-u))_T|\\ \leq & \; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i, j = 1}^2\|Q_r\partial^2_{ij} w-Q^1 Q^r\partial^2_{ij} w\|_T^2\Big)^{\frac{1}{2}} ||| Q_hu-u |||\\ \leq & \; Ch^{k+1}\|u\|_{k+1} \|w\|_4. \end{split} \end{equation} (7.8)

    For J_4 , using Cauchy-Schwarz inequality, the trace inequality (4.1), Lemma 4.4, the estimate (6.1) with m = 3 and s = 0, 1 , (6.3), (6.4) gives

    \begin{equation} \begin{split} J_4 = &\; \ell(w, \zeta_h)\\ \leq &\; \Big|\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2 - \langle (\zeta_b-\zeta_0) n_i, \partial_j ((Q_r-I) \partial_{ij}^2 w) \rangle_{\partial T}\\&+\langle \zeta_{gi}-\partial_i \zeta_0, (Q_r-I) \partial_{ij}^2 w n_j\rangle_{\partial T}\Big| \\ \leq&\; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2\|(\zeta_b-\zeta_0) n_i\|_{\partial T}^2\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2\|\partial_j ((Q_r-I) \partial_{ij}^2 w) \|_{\partial T}^2\Big)^{\frac{1}{2}} \\ &+\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2\|\zeta_{gi}-\partial_i \zeta_0\|_{\partial T}^2\Big)^{\frac{1}{2}} \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2\|(Q_r-I) \partial_{ij}^2 w n_j\|_{\partial T}^2\Big)^{\frac{1}{2}} \\ \leq&\; \Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i,j = 1}^2h_T^2 \|\partial_j ((Q_r-I) \partial_{ij}^2 w)\|_{T}^2+h_T^4 \| \partial_j ((Q_r-I) \partial_{ij}^2 w)\|_{1, T}^2\Big)^{\frac{1}{2}} \\&\cdot\Big(\sum\limits_{T\in {\mathcal{T}}_h}h_T^{-3}\|\zeta_0-\zeta_b\|_{\partial T}^2\Big)^{\frac{1}{2}} \\ &+ \Big(\sum\limits_{T\in {\mathcal{T}}_h} \sum\limits_{i,j = 1}^2 \|(Q_r-I) \partial_{ij}^2 w n_j\|_{T}^2+h_T^2 \|(Q_r-I) \partial_{ij}^2 w n_j\|_{1, T}^2\Big)^{\frac{1}{2}} \\ &\cdot\Big(\sum\limits_{T\in {\mathcal{T}}_h}\sum\limits_{i = 1}^2 h_T^{-1}\|\zeta_{gi}-\partial_i \zeta_0\|_{\partial T}^2\Big)^{\frac{1}{2}} \\\leq &\; Ch^2\|w\|_{4}|||\zeta_h||| \\ \leq &\; Ch^2\|w\|_{4}(||| u-u_h|||+||| u-Q_hu|||) \\ \leq &\; Ch^{k+1}\|w\|_4\|u\|_{k+1}. \end{split} \end{equation} (7.9)

    Substituting (7.5), (7.6), (7.8) and (7.9) into (7.4), and using (7.2), gives

    \|\zeta_0\|^2\leq Ch^{k+1}\|w\|_4\|u\|_{k+1}\leq Ch^{k+1} \|u\|_{k+1} \|\zeta_0\|.

    This gives

    \|\zeta_0\|\leq Ch^{k+1} \|u\|_{k+1},

    which, using the triangle inequality and (6.2) with m = k and s = 0 , gives

    \|e_0\|\leq \|\zeta_0\|+\|u-Q_0u\|\leq Ch^{k+1}\|u\|_{k+1}.

    This completes the proof of the theorem.

    The author declares she has not used Artificial Intelligence (AI) tools in the creation of this article.

    The research of Chunmei Wang was partially supported by National Science Foundation Grant DMS-2136380.

    The author declares there is no conflict of interest.



    [1] L. An, V. R. Grimm, A. Sullivan, B. L. Turner II, N. Malleson, A. Heppenstall, et al., Challenges, tasks, and opportunities in modeling agent-based complex systems, Ecolog. Model., 457 (2021), 109685. https://doi.org/10.1016/j.ecolmodel.2021.109685 doi: 10.1016/j.ecolmodel.2021.109685
    [2] W. R. KhudaBukhsh, B. S. Choi, E. Kenah, G. A. Rempała, Survival dynamical systems: Individual-level survival analysis from population-level epidemic models, Interface Focus, 10 (2020). https://doi.org/10.1098/rsfs.2019.0048
    [3] F. Di Lauro, W. R. KhudaBukhsh, I. Z. Kiss, E. Kenah, M. Jensen, G. A. Rempała, Dynamic survival analysis for non-markovian epidemic models. J. Royal Soc. Interf., 19 (2022), 20220124. https://doi.org/10.1098/rsif.2022.0124
    [4] B. S. Choi, S. Busch, D. Kazadi, B. Ilunga, E. Okitolonda, Y. Dai, et al., Modeling outbreak data: Analysis of a 2012 Ebola virus disease epidemic in DRC, Biomath, 8 (2019). https://doi.org/10.11145/j.biomath.2019.10.037
    [5] H. Vossler, P. Akilimali, Y. H. Pan, W. R. KhudaBukhsh, E. Kenah, G. A. Rempała, Analysis of individual-level epidemic data: Study of 2018-2020 Ebola outbreak in Democratic Republic of the Congo, Sci. Rep., 12 (2022). https://doi.org/10.1038/s41598-022-09564-4
    [6] W. R. KhudaBukhsh, S. K. Khalsa, E. Kenah, G. A. Rempala, J. H. Tien, COVID-19 dynamics in an Ohio prison,, medRxiv, 2021. Available from: https://www.medrxiv.org/content/early/2021/01/15/2021.01.14.21249782
    [7] M. Wascher, P. M. Schnell, W. R. KhudaBukhsh, M. Quam, J. H. Tien, G. A. Rempała, Monitoring sars-cov-2 transmission and prevalence in populations under repeated testing, 2021. Available from: https://www.medrxiv.org/content/10.1101/2021.06.22.21259342v1
    [8] I. Somekh, W. R. KhudaBukhsh, E. D. Root, G. A. Rempała, E. Sim{ o}es, E. Somekh, Quantifying the population-level effect of the COVID-19 mass vaccination campaign in Israel: A modeling study, Open Forum Infect. Diseases, 9 (2022). https://doi.org/10.1093/ofid/ofac087
    [9] E. Kenah, Contact intervals, survival analysis of epidemic data, and estimation of R_0, Biostatistics, 12 (2011), 548–566. https://doi.org/10.1093/biostatistics/kxq068 doi: 10.1093/biostatistics/kxq068
    [10] N. G. van Kampen, Remarks on Non-Markov Processes, Brazilian J. Phys., 28 (1998).
    [11] R. Ferrière, V. C. Tran, Stochastic and deterministic models for age-structured populations with genetically variable traits, In CANUM 2008, volume 27 of ESAIM Proc., pages 289–310. EDP Sci., Les Ulis, 2009. https://doi.org/10.1051/proc/2009033
    [12] E. Franco, M. Gyllenberg, O. Diekmann, One dimensional reduction of a renewal equation for a measure-valued function of time describing population dynamics, Acta Appl. Math., 175 (2021), 12. https://doi.org/10.1007/s10440-021-00440-3 doi: 10.1007/s10440-021-00440-3
    [13] J. M. Hyman, J. Li, Infection-age structured epidemic models with behavior change or treatment, J. Biol. Dynam., 1 (2007), 109–131. https://doi.org/10.1080/17513750601040383 doi: 10.1080/17513750601040383
    [14] N. Sherborne, J. C. Miller, K. B. Blyuss, I. Z. Kiss, Mean-field models for non-markovian epidemics on networks, J. Math. Biol., 76 (2018), 755–778. https://doi.org/10.1007/s00285-017-1155-0 doi: 10.1007/s00285-017-1155-0
    [15] V. C. Tran, Large population limit and time behaviour of a stochastic particle model describing an age-structured population, ESAIM. Probab. Stat., 12 (2008), 345–386. https://doi.org/10.1051/ps:2007052 doi: 10.1051/ps:2007052
    [16] N. Fournier, S. Méléard, A microscopic probabilistic description of a locally regulated population and macroscopic approximations, Ann. Appl. Probab., 14 (2004), 1880–1919.
    [17] W. R. KhudaBukhsh, H.-W. Kang, E. Kenah, G. Rempała, Incorporating age and delay into models for biophysical systems, Phys. Biol., 18 (2021), 10. https://doi.org/10.1088/1478-3975/abc2ab doi: 10.1088/1478-3975/abc2ab
    [18] W. R. KhudaBukhsh, C. D Bastian, M. Wascher, C. Klaus, S. Y. Sahai, M. H. Weir, et al., Projecting covid-19 cases and subsequent hospital burden in ohio, medRxiv, 2022. Available from: https://www.medrxiv.org/content/10.1101/2022.07.27.22278117v1.full.pdf+html
    [19] C. D. Bastian, G. A. Rempala, Throwing stones and collecting bones: Looking for poisson-like random measures, Math. Methods Appl. Sci., 43 (2020), 4658–4668. https://doi.org/10.1002/mma.6224 doi: 10.1002/mma.6224
    [20] G. F. Webb, Theory of nonlinear age-dependent population dynamics, volume 89 of Monographs and Textbooks in Pure and Applied Mathematics. Marcel Dekker, Inc., New York, 1985. https://doi.org/10.1007/BF00250793
    [21] M. Iannelli, M. Martcheva, F. A. Milner, Gender-structured population modeling, volume 31 of Frontiers in Applied Mathematics, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2005.
    [22] L. C. Evans, Partial differential equations, volume 19 of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, second edition, 2010.
    [23] E. DiBenedetto, Partial differential equations, Cornerstones. Birkhäuser Boston, Ltd., Boston, MA, second edition, 2010.
    [24] C Klaus, PDE-DSA github repository, 2022. Available from: https://github.com/klauscj68/PDE-Vax
    [25] J. Bezanson, A. Edelman, S. Karpinski, V. B. Shah, Julia: A fresh approach to numerical computing, SIAM Rev., 59 (2017), 65–98.
    [26] Ohio Department of Health, Ohio Department of Health COVID Dashboard, Available from: https://coronavirus.ohio.gov/wps/portal/gov/covid-19/dashboards/overview
    [27] Centers for Disease Control and Prevention (CDC), US Centers for Disease Control and Prevention: COVID-19 vaccinations in the United States, County, Available from: https://data.cdc.gov/Vaccinations/COVID-19-Vaccinations-in-the-United-States-County/8xkx-amqh
    [28] U. S. Census, County Population Totals 2010-2019, Available from: https://www.census.gov/data/datasets/time-series/demo/popest/2010s-counties-total.html
    [29] L. L. Schumaker, Spline functions: Computational Methods, Society for Industrial and Applied Mathematics, Philadelphia, PA, 2015. https://doi.org/10.1137/1.9781611973907
    [30] T. Kypraios, P. Neal, D. Prangle, A tutorial introduction to bayesian inference for stochastic epidemic models using approximate bayesian computation, {Math. Biosci.}, 287 (2017), 42–53. https://doi.org/10.1016/j.mbs.2016.07.001 doi: 10.1016/j.mbs.2016.07.001
    [31] S..A. Sisson, Y. N. Fan, M. A. Beaumont, Handbook of Approximate Bayesian Computation, CRC Press, Boca Raton, FL, 2020.
    [32] Centers for Disease Control and Prevention (CDC), US Centers for Disease Control and Prevention: SARS-CoV-2 Variant Classifications and Definitions, Available from: https://www.cdc.gov/coronavirus/2019-ncov/variants/variant-classifications.html
    [33] K. Koelle, M. A. Martin, R. Antia, B. Lopman, N. E. Dean, The changing epidemiology of sars-cov-2, Science, 375 (2022), 1116–1121. https://doi.org/10.1126/science.abm4915 doi: 10.1126/science.abm4915
    [34] M. O'Driscoll, G. R. Dos Santos, L. Wang, D. A. T. Cummings, A. S. Azman, J. Paireau, et al., Age-specific mortality and immunity patterns of SARS-CoV-2, Nature, 590 (2021), 140–145. https://doi.org/10.1038/s41586-020-2918-0 doi: 10.1038/s41586-020-2918-0
    [35] I. Holmdahl, C. Buckee, Wrong but useful—what covid-19 epidemiologic models can and cannot tell us, New England J. Med., 383 (2020), 303–305.
    [36] N. P. Jewell, J. A. Lewnard, B. L. Jewell, Predictive mathematical models of the COVID-19 Pandemic: Underlying principles and value of projections, JAMA, 323 (2020), 1893–1894. https://doi.org/10.1001/jama.2020.6585 doi: 10.1001/jama.2020.6585
    [37] N. Barda, D. Riesel, A. Akriv, J. Levy, U. Finkel, G. Yona, et al. Developing a COVID-19 mortality risk prediction model when individual-level data are not available, Nat. Commun., 11 (2020), 4439. https://doi.org/10.1038/s41467-020-18297-9 doi: 10.1038/s41467-020-18297-9
    [38] C. Klaus, M. Wascher, W. R. KhudaBukhsh, J. H. Tien, G. A. Rempała, E. Kenah, Assortative mixing among vaccination groups and biased estimation of reproduction numbers, Lancet Infect. Diseases, 22 (2022), 579–581. https://doi.org/10.1016/S1473-3099(22)00155-4 doi: 10.1016/S1473-3099(22)00155-4
    [39] O. A. Ladyženskaja, V. A. Solonnikov, N. N. Ural'ceva, Linear and quasilinear equations of parabolic type, in Translations of Mathematical Monographs, Vol. 23. American Mathematical Society, Providence, R.I., 1968. Translated from the Russian by S. Smith.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(2881) PDF downloads(72) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog