
We discuss the design and implementation details of two conforming virtual element methods for the numerical approximation of two partial differential equations that emerge in phase-field modeling of fracture propagation in elastic material. The two partial differential equations are: (i) a linear hyperbolic equation describing the momentum balance and (ii) a fourth-order elliptic equation modeling the damage of the material. Inspired by [
Citation: Yu Leng, Lampros Svolos, Dibyendu Adak, Ismael Boureima, Gianmarco Manzini, Hashem Mourad, Jeeyeon Plohr. A guide to the design of the virtual element methods for second- and fourth-order partial differential equations[J]. Mathematics in Engineering, 2023, 5(6): 1-22. doi: 10.3934/mine.2023100
[1] | Claudio Canuto, Davide Fassino . Higher-order adaptive virtual element methods with contraction properties. Mathematics in Engineering, 2023, 5(6): 1-33. doi: 10.3934/mine.2023101 |
[2] | Andrea Borio, Martina Busetto, Francesca Marcon . SUPG-stabilized stabilization-free VEM: a numerical investigation. Mathematics in Engineering, 2024, 6(1): 173-191. doi: 10.3934/mine.2024008 |
[3] | Zaffar Mehdi Dar, M. Arrutselvi, Chandru Muthusamy, Sundararajan Natarajan, Gianmarco Manzini . Virtual element approximations of the time-fractional nonlinear convection-diffusion equation on polygonal meshes. Mathematics in Engineering, 2025, 7(2): 96-129. doi: 10.3934/mine.2025005 |
[4] | Massimo Frittelli, Ivonne Sgura, Benedetto Bozzini . Turing patterns in a 3D morpho-chemical bulk-surface reaction-diffusion system for battery modeling. Mathematics in Engineering, 2024, 6(2): 363-393. doi: 10.3934/mine.2024015 |
[5] | E. Artioli, G. Elefante, A. Sommariva, M. Vianello . Homogenization of composite materials reinforced with unidirectional fibres with complex curvilinear cross section: a virtual element approach. Mathematics in Engineering, 2024, 6(4): 510-535. doi: 10.3934/mine.2024021 |
[6] | Ludovica Cicci, Stefania Fresca, Stefano Pagani, Andrea Manzoni, Alfio Quarteroni . Projection-based reduced order models for parameterized nonlinear time-dependent problems arising in cardiac mechanics. Mathematics in Engineering, 2023, 5(2): 1-38. doi: 10.3934/mine.2023026 |
[7] | Pablo Blanc, Fernando Charro, Juan J. Manfredi, Julio D. Rossi . Games associated with products of eigenvalues of the Hessian. Mathematics in Engineering, 2023, 5(3): 1-26. doi: 10.3934/mine.2023066 |
[8] | Stefano Berrone, Stefano Scialò, Gioana Teora . The mixed virtual element discretization for highly-anisotropic problems: the role of the boundary degrees of freedom. Mathematics in Engineering, 2023, 5(6): 1-32. doi: 10.3934/mine.2023099 |
[9] | Federica Botta, Matteo Calafà, Pasquale C. Africa, Christian Vergara, Paola F. Antonietti . High-order discontinuous Galerkin methods for the monodomain and bidomain models. Mathematics in Engineering, 2024, 6(6): 726-741. doi: 10.3934/mine.2024028 |
[10] | Francesca Bonizzoni, Davide Pradovera, Michele Ruggeri . Rational-approximation-based model order reduction of Helmholtz frequency response problems with adaptive finite element snapshots. Mathematics in Engineering, 2023, 5(4): 1-38. doi: 10.3934/mine.2023074 |
We discuss the design and implementation details of two conforming virtual element methods for the numerical approximation of two partial differential equations that emerge in phase-field modeling of fracture propagation in elastic material. The two partial differential equations are: (i) a linear hyperbolic equation describing the momentum balance and (ii) a fourth-order elliptic equation modeling the damage of the material. Inspired by [
The virtual element method (VEM) is a generalization of the finite element method (FEM) that enables computational simulations using n-sided polygonal/polyhedral elements (with n≥3 in 2D and n≥4 in 3D), significantly reducing the difficulty of meshing geometrically complex domains. In addition, the VEM provides a means of designing approximations of an arbitrarily high order of accuracy and regularity [4], thus making it ideal for problems admitting C1-continuous solutions. The VEM is also known to be accurate, stable, and robust on highly deformed meshes, as shown in numerous theoretical [5,6,7] and numerical [8,9,10] studies on the treatment of large-deformation problems. The method preserves these desirable performance characteristics when the mesh includes elements with obtuse or re-entrant angles (i.e., nonconvexity) and hanging nodes, facilitating the development of adaptive mesh refinement (AMR) schemes to avoid excessive mesh refinement [11].
Despite these excellent properties and an established and solid mathematical foundation, which is well documented in the literature [12,13,14,15,16,17,18], the VEM lags behind the FEM's adoption in practical applications mainly because it might be perceived as being conceptually more complex and, therefore, more challenging to implement and apply. Several works in the literature help in addressing this issue (e.g., see [2,19]) sometimes by presenting simple virtual element formulations and detailing their implementations (e.g., see [20,21,22]). In this work, we address the implementation design of two distinct virtual element methods for approximating a time-dependent, second-order momentum balance equation and a stationary, fourth-order, general elliptic equation, which includes second- and zero-order terms. The VEM is developed as a numerical method to approximate different model problems on polygonal elements and could be seen as a generalization of the FEM. Unlike the FEM, the discrete functions are formally the solutions of a local PDE that is defined on every mesh element Instead of directly evaluating the discrete functions, which are never explicitly constructed, we employ suitable projection operators, which are computable from the degrees of freedom, and project such functions onto a polynomial space. Consequently, the matrices associated with the discrete-bilinear forms involve projection operators. The VEM allows us to compute the local matrices on different polygonal elements in a unified way and the scheme's formulation and implementation do not depend on the shape of the elements. These features make the VEM different from the existing techniques. However, some issues may exist, such as the stability related to the stiffness and mass matrices. If the polygonal element contains very small edges compared to the diameter of the element, the dofi-dofi stabilization will not work, and we need boundary integration of the trace of the basis functions, which is more technical to compute [5,23]. In this work, we have assumed the meshes do not contain small edges and all polygonal elements are star-shaped.
The momentum balance equation governs the linear elastodynamic initial boundary-value problem (IBVP); the second equation mathematically models crack propagation in materials. When coupled, these two PDEs constitute fourth-order phase-field models of quasi-brittle fractures, similar to those presented in [24,25]. Part of the authors has previously presented VEMs for the solution of the first equation in [26], while the virtual element discretization of the second equation is currently under study, see, for instance, the work in progress of [27]. These previous works focused on both methods' theoretical aspects and convergence analysis, providing convergence proofs, error estimates, and the numerical investigation of their convergence behavior in representative benchmark test case applications. Instead, herein we focus on the implementation design for each equation separately, emphasizing similar computational aspects. At the same time, a forthcoming publication will address the VEM-based coupled approach to solving the fracture model.
The remainder of this paper is as follows. We present the mathematical formulation of these two model problems in the strong and weak form in Section 2 and their numerical discretizations using the VEM in Section 3. We discuss the implementation design in Section 4 and offer our concluding remarks in Section 5.
In this section, we introduce the model problems that we consider in this paper, and their variational formulation, which we will discretize using the virtual element method. In Subsection 2.1, we introduce the notation that is used throughout the paper; in Subsection 2.2, we present the model problems in strong form; in Subsection 2.3, we present the model problems in variational form.
Throughout this paper, we adopt the notation of Sobolev spaces as in [28]. Accordingly, we denote the space of square-integrable functions defined on any open, bounded, connected domain Ω⊂R2 with boundary ∂Ω by L2(Ω), and the Hilbert space of functions in L2(Ω) with all partial derivatives up to a positive integer m also in L2(Ω) by Hm(Ω), see [28]. We endow Hm(Ω) with the usual norm and seminorm that we denote as ||⋅||m,Ω and |⋅|m,Ω, respectively.
The virtual element methods that we consider in the next sections are formulated on the mesh family {Ωh}h, where each mesh Ωh is a partition of the computational domain Ω into nonoverlapping star-shaped polygonal elements E with boundary ∂E, area |E|, center of gravity xE, and diameter hE=supx,y∈E|x−y|. The mesh elements of Ωh form a finite cover of Ω such that ¯Ω=∪E∈ΩhE and the mesh size labeling each mesh Ωh is defined by h=maxE∈ΩhhE. A mesh edge e has center xe and length he; a mesh vertex v has position vector xv. We denote the edges of the polygonal boundary ∂E by EE.
For any integer number ℓ≥0, we let Pℓ(E) and Pℓ(e) denote the space of polynomials defined on the element E and the edge e, respectively; Pℓ(Ωh) denotes the space of piecewise polynomials of degree ℓ on the mesh Ωh. For the convenience of exposition, we also use the notation. P−2(E)=P−1(E)={0}. Accordingly, it holds that q|E∈Pℓ(E) if E∈Ωh for all q∈Pℓ(Ωh). Throughout the paper, we use the multi-index notation, so that ν=(ν1,ν2) is a two-dimensional index defined by the two integer numbers ν1,ν2≥0. Moreover, Dνw=∂|ν|w/∂xν11∂xν22 denotes the partial derivative of order |ν|=ν1+ν2>0 of a sufficiently regular function w(x1,x2), and we use the conventional notation that D(0,0)w=w for ν=(0,0). We also denote the partial derivatives of w versus x and y by the shortcuts ∂xw and ∂yw, and the normal and tangential derivatives with respect to a given edge by ∂nw and ∂tw. Finally, we define the trace operator tr(A)=∑iaii for any matrix (two-index tensor) A=(aij). Given two tensors A=(aij) and B=(bij), we also define A:B:=∑i∑jaijbij.
The linear momentum balance equation for the displacement u:Ω×(0,T]→R2 on the 2-D computational domain Ω reads as
ρ¨u=∇⋅σ+f,ugNu=0v0=0onΩ×(0,T], | (2.1) |
where ρ:Ω→R is a scalar density field, ¨u denotes the second-order time derivative of u, σ:Ω×(0,T]→R2×2sym is the (time-dependent) stress tensor (R2×2sym being the set of symmetric 2×2-sized real-valued tensors), and f:Ω×(0,T]→R2 is a suitable vector-valued forcing term. In this paper, we assume the density field is constant. The corresponding stress tensor σ is given by
σ(u)=gu(x)σe,withσe=(λ+μ)εvol+2μεdev, | (2.2) |
where gu(x):R2→R is a given function modeling different material behaviors such as hardening or softening effects, for instance, due to damage or fracture, and εvol and εdev are the volumetric and deviatoric strain tensors.
Tensors εvol and εdev are respectively given by
εvol=12tr(ε(u))I,andεdev=ε(u)−εvol,withε(u)=12(∇u+∇uT), | (2.3) |
where I the 2×2 identity tensor, and the volumetric strain tensor can also be written as εvol=(∇⋅u)I.
To complete the mathematical formulation of this model problem, we supplement Eq (2.1) with the initial and boundary conditions. The initial conditions on the displacement and its derivative at t=0 are given by
u=u0,u=0inΩ×{0}, | (2.4) |
˙u=v0,u=0inΩ×{0}. | (2.5) |
Using (2.4)-(2.5) allows us to define the stress tensor σ at t=0 through definitions (2.2)-(2.3). To introduce the boundary conditions, we first assume that the domain boundary Γ is decomposed into a Dirichlet boundary, e.g., ΓD,u that is a closed subset of Γ in the Euclidean topology, and Neumann boundary, e.g., ΓN,u, such that Γ=ΓD,u∪ΓN,u and |ΓD,u∩ΓN,u|=0. Then the boundary conditions read as:
u=¯u,gNonΓD,u×(0,T],σn=gN,uonΓN,u×(0,T]. |
Herein, u0,v0,u, and gN are given data; n=(nx,ny)T is the unit vector orthogonal to Γ and pointing out of Ω.
The fourth-order equation for the phase field d on the 2-D computational domain Ω reads as
α2Δ2d−α1Δd+α0d+gd(d)Ht=0, | (2.6) |
where α2,α1, and α0 are model parameters, gd(d):R→R is an affine function, and the history variable Ht:Ω→R is an additional multiplicative term. We assume that α2 and α1 are strictly positive real numbers, and α0 is a non-negative real number. We remark that gd(d):=α3d+α4, where α3 and α4 are constants. For simplicity, we let α4=0 in this work. We refer to [12] for a detailed discussion about the physical meaning of gd(d) and Ht when modeling crack initiation and propagation phenomena. To set the boundary conditions that establish the mathematical model, we first split the domain boundary as Γ=ΓD,d∪ΓN,d, where ΓD,d and ΓN,d represents the Dirichlet and Neumann parts of the domain boundary, respectively. Subsets ΓD,d and ΓN,d are disjoint in the sense that their intersection has zero 2D measure, i.e., |ΓD,d∩ΓN,d|=0. Additionally, we split the Dirichlet boundary as ΓD,d=ΓD1,d∪ΓD2,d and the Neumann boundary as ΓN,d as ΓN,d=ΓN1,d∪ΓN2,d. We complete the mathematical formulation of the model problem (2.6) with the following boundary conditions
d=¯d on ΓD1,d,and∇d=0 on ΓD2,d,Δd=0 on ΓN1,d,and∇Δd=0 on ΓN2,d. |
Consider the affine functional space of the two-dimensional vector-valued fields whose components belong to the Sobolev space H1(Ω), and have an assigned value on the boundary Dirichlet boundary ΓD,u:
Vu={u∈[H1(Ω)]2:u=¯u on ΓD,u}. | (2.7) |
Denoting Vu,0 if ¯u=0 and testing Eq (2.1) with elements in Vu,0, we arrive at the variational formulation of the elastodynamics equation as
Findu∈Vus.t.Mu(¨u,w)+Au(u,w)=Fu(w),∀w∈Vu,0 | (2.8) |
with the definitions:
Mu(u,v):=∫Ωρu⋅vdV,andAu(u,v):=∫Ωσ(u):ε(v)dV,∀u∈Vu, and ∀v∈Vu,0, |
and
Fu(v)=∫Ωf⋅vdV+∫ΓN,ugN⋅vdS,∀v∈Vu,0, |
where we denote the scalar product of two (matrix) tensors by ":". In particular, for any tensor fields A=(ail)i,j=1,2, and B=(bil)i,j=1,2, we consider the standard scalar product of 2 × 2- matrices: A:B=∑i,j=1,2aijbij. As shown, for example, by Raviart and Thomas (see [13, Theorem 8-3.1]) the variational problem (2.8) is well-posed and its unique solution satisfies
u∈C0((0,T);Vu)∩C1((0,T);[L2(Ω)]2), |
assuming that the external load f∈L2((0,T);[L2(Ω)]2), the boundary function gN∈C1((0,T);[H120,ΓN]2), and the initial functions u0∈[H10,ΓD(Ω)]2, v0∈[L2(Ω)]2.
Similar to the previous section, we consider the affine functional space
Vd={d∈H2(Ω):d=¯d on ΓD1,d and ∇d=0 on ΓD2,d}, | (2.9) |
and a special case of Vd, namely, Vd,0, given by setting ¯d=0 in the previous definition. On testing (2.6) with elements in Vd,0, we obtain the variational formulation of the fourth-order phase-field equation.
Find d∈Vd, such that
Ad(d,c):=α2Ad,2(d,c)+α1Ad,1(d,c)+α0Ad,0(d,c)+Agd(gd(d),c;Ht)=0,∀c∈Vd,0, | (2.10) |
where
Ad,2(d,c):=∫ΩΔdΔcdV,Ad,1(d,c):=∫Ω∇d⋅∇cdV, | (2.11) |
Ad,0(d,c):=∫ΩdcdV,andAgd(gd(d),c;Ht):=∫Ωgd(d)HtcdV. | (2.12) |
The wellposedness of the variational formulation, i.e., existence and uniqueness of d∈H2(Ω) solving (2.10) with definitions (2.11)-(2.12), follows from applications of the Lax-Milgram theorem [14], since the bilinear form Ad(⋅,⋅) is coercive and continuous.
In this section, we briefly review the formulation of the virtual element method that approximates the second-order momentum balance equation and the high-order phase field equation introduced in the previous section.
Let Vhu,k be the virtual element approximation of the affine functional space Vu defined in (2.7). The semi-discrete virtual element approximation of (2.8) reads as:
For all t∈(0,T], find uh(t)∈Vhu,k, such that for t=0 it holds that
uh(0)=(u0)I,˙uh(0)=(u1)I,andMhu(¨uh,vh)+Ahu(uh,vh)=Fhu(vh)∀vh∈Vhu,k. | (3.1) |
Here, uh(t) is the virtual element approximation of u and vh the generic test function in Vhk, while (u0)I and (u1)I are the virtual element interpolants of the initial solution functions u(0) and ˙u(0), respectively. The bilinear forms Mhu(⋅,⋅), Ahu(⋅,⋅), and the linear functional Fhu(⋅) are the virtual element approximations of Mu(⋅,⋅), Au(⋅,⋅), and Fu(⋅), in (2.8).
The total time interval [0,T] is divided into NT subintervals with time step Δt=TNT and time instants t(n)=nΔt. The Newmark method [15] is widely used to solve elastodynamic problems because it allows us to solve the second-order equation directly without splitting the equation into two first-order equations. Therefore, we employ the Newmark method to advance the system in time. In this implicit scheme, accelerations and velocities at time t(n+1) are approximated as follows:
¨u(n+1)=1β(Δt)2(u(n+1)−u(n))−1βΔt˙u(n)−1−2β2β¨u(n), | (3.2) |
˙u(n+1)=γβΔt(u(n+1)−u(n))+(1−γβ)˙u(n)+(1−γ2β)¨u(n), | (3.3) |
where β in (0, 0.5] and γ∈(0,1] are Newmark parameters.
The vector-valued virtual element space is such that Vhu,k(E)=Vhu,k(E)×Vhu,k(E), where Vhu,k(E) denotes the local scalar conforming virtual element space (which will be introduced below). Accordingly, the vector-valued displacement field uh is provided by the two components uh=(ux,h,uy,h), where each component belongs to Vhu,k(E). Both components ui,h, for i∈{x,y}, are uniquely identified by the following degrees of freedom:
(U1) the values of ui,h at the vertices of E;
(U2) the edge polynomial moments of ui,h of order up to k−2 on each one-dimensional edge e∈EE:
1|e|∫eui,hmdS,∀m∈Mk−2(e),∀e∈EE; | (3.4) |
(U3) the cell polynomial moments of ui,h of order up to k−2 on E:
1|E|∫Eui,hmdV,∀m∈Mk−2(E). | (3.5) |
In this definition, we use the symbols Mk−2(e) and Mk−2(E) to denote a basis for the polynomial spaces Pk(e) and Pk−2(E), respectively. Suitably scaled monomials or orthogonal polynomials can provide this basis.
Figure 1 shows the degrees-of-freedom of the three scalar conforming virtual element spaces (k=1,2,3) defined on a pentagonal cell. Using these degrees of freedom, we compute the elliptic projection operator Π∇,Eu,k:Vhu,k(E)→Pk(E), which is defined as
∫E∇Π∇,Eu,kch⋅∇qdV,=∫E∇ch⋅∇qdV,∀q∈Pk(E), | (3.6a) |
∫∂EΠ∇,Eu,kchdS=∫∂EchdS, | (3.6b) |
where the second condition is needed to fix the kernel of the gradient operator. The local scalar virtual element space of order k≥1 is defined by
Vhu,k(E)={ch∈H1(E):Δch∈Pk(E) and ch|∂E∈B(0)k(∂E) with (ch−Π∇,Eu,kch,μh)E=0,∀μh∈Pk(E)∖Pk−2(E)} | (3.7) |
where Pk(E)∖Pk−2(E) denotes as the space spanned by the monomials of degree k and k−1 on E, and
B(0)k(∂E)={ch∈C0(∂E):ch|e∈Pk(e),∀e∈EE}. |
According to this definition, the orthogonal projection Π0,Eu,kch of a virtual element function ch onto the polynomial space of degree k, defined as,
∫E(Π0,Eu,kch−ch)q=0,∀q∈Pk(E), |
is also computable. The global conforming space of vector-valued virtual element field of order k≥1, i.e., the finite-dimensional space Vhu,k, is obtained by combining all the elemental spaces [Vhu,k(E)]2. Building upon the local spaces Vhu,k(E), k≥1 for all E∈Ωh, we define it as
Vhu,k:={u∈Vu:u|E∈Vhu,k(E)∀E∈Ωh}. | (3.8) |
In the virtual element setting, we define the bilinear forms Mhu(⋅,⋅) and Ahu(⋅,⋅) as the sum of elemental contributions, which are denoted by MEu(⋅,⋅) and AEu(⋅,⋅), respectively:
Mhu(⋅,⋅):Vhu,k×Vhu,k→R,withMhu(⋅,⋅)=∑E∈ΩhMEu(⋅,⋅),Ahu(⋅,⋅):Vhu,k×Vhu,k→R,withAhu(⋅,⋅)=∑E∈ΩhAEu(⋅,⋅). |
The local bilinear form Mh,Eu(⋅,⋅) is given by
MEu(uh,vh)=∫EρΠ0,Eu,kuh⋅Π0,Eu,kvhdV+SEm(uh,vh), | (3.9) |
where SEm(⋅,⋅) is the local stabilization term for Mh,Eu. The bilinear form MEu depends on the orthogonal projections Π0u,kuh and Π0u,kvh, which are computable from the degrees of freedom of uh and vh, respectively, see the previous section. The local bilinear form Ah,Eu(⋅,⋅) is given by
AEu(uh,vh)=∫EΠ0,Eu,k−1σ(uh):Π0,Eu,k−1ε(vh)dV+SEa(uh,vh), | (3.10) |
where SEa(⋅,⋅) is the local stabilization term for AEu. The bilinear form AEu depends on the orthogonal projections Π0,Eu,k−1σ(uh) and Π0,Eu,k−1ε(vh), which are computable from the degrees of freedom of uh and vh, respectively, see the previous section.
The local stabilization terms SEm(⋅,⋅),SEa(⋅,⋅):Vhu,k×Vhu,k→R can be any symmetric and coercive bilinear forms that are computable from the degrees of freedom, and suitably provides the k-consistency and stability property. The role of the stabilization in the VEM is discussed in detail in [23]. The stabilization used in our implementation is discussed in Subsection 4.2.
Finally, we approximate the right-hand side (3.1) of the semi-discrete formulation as follows:
Fhu(vh)=∑E∈ΩhFEu(vh)+∑e∈ΓNFeu(vh),∀vh∈Vhk, | (3.11) |
where
FEu(vh)=∫Ef⋅Π0,Eu,k(vh)dV,Feu(vh)=∫egN⋅vhdS. | (3.12) |
The linear functional Fhu(⋅) is clearly computable since the edge trace vh|e is a polynomial and Π0,Eu,k(vh) is computable from the degrees of freedom of vh.
The virtual element method approximates the variational formulation (2.10) as follows
Finddh∈Vhd,rsuchthatAhd(dh,ch)=0,∀ch∈Vhd,r. | (3.13) |
In this formulation, Vhd,r is the H2-conforming approximation of the space H2(Ω) provided by the VEM, dh and ch are the trial and test functions from this space.
On every mesh element E∈Ωh, the following set of values comprises the degrees of freedom of a virtual element function ch as shown in Figure 2:
(D1) for r≥2, ch(xv), ∂xch(xv), ∂ych(xv) for any vertex v of ∂E;
(D2) for r≥4, 1|e|∫eqchdS for any q∈Pr−4(e), and any edge e∈EE;
(D3) for r≥3, ∫eq∂nchdS for any q∈Pr−3(e), and any edge e∈EE;
(D4) for r≥2, 1|E|∫EqvdV for any q∈Pr−2(E).
Consider the integer r≥2 and the bilinear form B(u,v)=α2Ad,2(u,v)+α1Ad,1(u,v) with α2,α1>0. We define the elliptic projection operator ΠL,Ed,r:H2(E)→Pr(E) such that for every v∈H2(E), the r-degree polynomial ΠL,Ed,rv is the solution to the variational problem:
B(ΠL,Ed,rv−v,q)=0,∀q∈Pr(E), | (3.14) |
∫∂E(ΠL,Ed,rv−v)dS=0. | (3.15) |
The elliptic projection operator ΠL,Ed,r:H2(E)∩C1(ˉE)→Pr(E) for any integer number r≥k is such that ΠL,Ed,rd for all d∈H2(E)∩C1(ˉE) is the the solution of the finite-dimensional variational problem
B(ΠL,Ed,rd,q)=B(d,q),q∈Pr(E), | (3.16) |
with the following additional conditions
(ΠL,Ed,rd,1)L2(E)=(d,1)L2(E), | (3.17) |
and for i={x,y}:
(∂iΠL,Ed,rd,1)L2(E))=(∂id,1)L2(E)). | (3.18) |
For r≥3, the local virtual element space on element E is defined by
Vhd,r(E)={dh∈H2(E):Ldh∈Pr−4(E) and dh|∂E∈B(1)r(∂E) with (dh−ΠL,Ed,rdh,μh)E=0,∀μh∈Pr(E)∖Pr−2(E)} |
where
B(1)r(∂E)={dh∈C1(∂E):dh|e∈Pr(e),∂ndh|e∈Pr−1(e),∀e∈EE}. | (3.19) |
For r=2, we have the special case of the low-order virtual element space:
Vhd,r(E)={dh∈H2(E):Ldh=0 and dh∈P3(e),∂ndh∈P1(e),∀e∈∂E with (dh−ΠL,Ed,rdh,μh)E=0,∀μh∈P2(E)}. | (3.20) |
Note that Pr(E)⊂Vhd,r(E). Building upon the local spaces Vhd,r(E), r≥2, for all E∈Ωh the global conforming virtual element space Vhd,r is defined as
Vhd,r:={dh∈H2(Ω):dh|E∈Vhd,r(E)∀E∈Ωh}. | (3.21) |
This implies that Vhd,r⊂C1(Ω) because dh∈H2(E) in each element and C1-regularity across the internal mesh faces.
Let E∈Ωh be a mesh element, and consider the bilinear forms AEd,2,AEd,1,AEd,0:Vhd,r(E)×Vhd,r(E)→R given by integrating on E instead of Ω in the corresponding bilinear forms in (2.11)-(2.12). Let AEd(⋅,⋅)=α2AEd,2(⋅,⋅)+α1AEd,1(⋅,⋅)+α0AEd,0(⋅,⋅). We use the elliptic projection ΠL,Ed,r and the L2-orthogonal projection Π0,Ed,r to define the virtual element bilinear form AEd:Vhd,r(E)×Vhd,r(E)→R:
AEd(dh,ch):=α2AEd,2(ΠL,Ed,rdh,ΠL,Ed,rch)+α1AEd,1(ΠL,Ed,rdh,ΠL,Ed,rch)+α0AEd,0(Π0,Ed,rdh,Π0,Ed,rch)+SEh(dh−ΠL,Ed,rdh,ch−ΠL,Ed,rch). |
Herein, the stabilization term is also built by using the projection ΠL,Ed,r(ch), and the usual formula, so that the bilinear form SEh:Vhd,r(E)×Vhd,r(E)→R can be any symmetric, positive definite, bilinear form that suitably provides the k-consistency and stability properties. The stabilization used in our implementation is discussed in Subsection 4.3.
In this section, we follow the general guidelines of [19] and briefly describe how we implemented the VEM for the second-order elastodynamics equation and the fourth-order phase-field equation, respectively.
In this section, we introduce VEM vector and matrix notations that will be used in the following sections. We have introduced two virtual element spaces, H1-conforming Vhu,k and H2-conforming Vhd,r. For conciseness, we use the virtual element space Vhu,k as an example to introduce vector and matrix notations. Such notations can be easily extended to Vhd,r.
We consider the following compact notation. For all element E∈Ωh, we locally number the degrees of freedom (U1), (U2), and (U3) from 1 to Ndofs. Then, we introduce the bounded, linear functionals dofi:Vhu,k(E)→R, i=1,…,Ndofs, such that
dofi(vh):=i−th degree of freedom of vh |
for vh∈Vhu,k(E). Let ΛE={dofi(⋅)}i denote the set of such functionals and collect the degrees of freedom of vh in the vector vh=(dof1(vh),…,dofNdofs(vh))T. Since the degrees of freedom (U1), (U2), and (U3) are unisolvent in Vhu,k(E), the triplet (E,Vhu,k(E),ΛE) is a finite element in the sense of Ciarlet [16]. This property implies the existence of a Lagrangian basis set {φui}i, with φui∈Vhu,k(E), i=1,…,Ndofs, which satisfies
dofi(φuj)=δij,i,j=1,2,…,Ndofs. |
We refer to the basis function set {φui}i as the "canonical" basis of Vhu,k(E). We introduce the compact notation
φφu(x)=(φu1(x),…,φuNdofs(x))T, |
and write the expansion of a virtual element function vh on such a basis set as
vh(x)=φφu(x)Tvh=Ndofs∑i=1dofi(vh),φui(x)∀x∈E. |
We also introduce a compact notation for the basis of the polynomial subspace Pk(E)⊂Vhu,k(E), which reads as
mmu(x)=(mu1(x),…,munk(x))T, |
where nk is the cardinality of Pk(E). Since the polynomials muα(x) are also virtual element functions, we can expand them on the canonical basis φφ. We express such expansions as
mmu(x)T=φφu(x)TDDu, |
where matrix DDu has size Ndofs×nk and collects all the expansion coefficients
Duiℓ=dofi(muℓ), |
so that
muℓ(x)=Ndofs∑i=1φui(x)Duiℓℓ=1,…,nk. |
Following this notation, we also express the action of a differential operator D, e.g., D=Δ or D=∇, in a entry-wise way, so that
Dφφu(x)=(Dφu1(x),…,DφuNdofs(x))T, |
and
Dmmu(x)=(Dmu1(x),…,Dmunk(x))T. |
Similarly, we express the action of the projectors Π∇,Eu,k, and Π0,Eu,k on the canonical basis functions φφu and their expansion on the polynomial basis mmu as follows:
Π∇,Eu,kφφTu=[Π∇,Eu,kφu1,Π∇,Eu,kφu2,…Π∇,Eu,kφuNdofs]=mmTuΠΠ∇,Eu,k,Π0,Eu,kφφTu=[Π0,Eu,kφu1,Π0,Eu,kφu2,…Π0,Ed,kφuNdofs]=mmTuΠΠ0,Eu,k, |
where ΠΠ∇,Eu,k and ΠΠ0,Eu,k are the matrix representation of Π∇,Eu,k and Π0,Eu,k, respectively. The expansion coefficients for the three projection operators applied to the basis function φj are collected along the j-th column of the projection matrices ΠΠ∇,Eu,k, and ΠΠ0,Eu,k.
Remark 1. The vector and matrix notation for the fourth-order phase-field equation can be written in a similar fashion, such as, the virtual element space Vhd,r(E), degree of freedoms (D1)–(D4), canonical basis φφd, the polynomial basis mmd, and more.
In this section, we start with constructing elliptic and L2 projectors, followed by assembling local mass and stiffness matrices.
The elliptic projector Π∇,Eu,k is defined as, for vh∈Vhu,k(E),
∫E∇Π∇,Eu,kvh⋅∇qdV=∫E∇vh⋅∇qdV,∀q∈Pk(E). | (4.1) |
Using vector and matrix notation as introduced in Section 4.1, we have
∫E∇mmu⋅∇(Π∇,Eu,kφφTu)dV=∫E∇mmu⋅∇φφTudV. | (4.2) |
Substituting Π∇,Eu,kφφTu with mmTuΠΠ∇,Eu,k, we obtain
∫E∇mmu⋅∇φφTudV=∫E∇mmu⋅∇(mmTuΠΠ∇,Eu,k)dV=∫E∇mmu⋅∇mmTudVΠΠ∇,Eu,k. | (4.3) |
As a consequence, we have the following matrix equation
~GGuΠΠ∇,Eu,k=~BBu | (4.4) |
where
~GGu=∫E∇mmu⋅∇mmTudV,~BBu=∫E∇mmu⋅∇φφTudV=∫E−ΔmmuφφudV+∫∂E∂n(mmu)φφTudS. |
However, ~GGu is singular. We additionally define
GG0uΠΠ∇,Eu,k=BB0u,forGG0u=(∫∂EmmTudS0T⋮0T)andBB0u=(∫∂EφφTudS0T⋮0T). | (4.5) |
Collecting Eqs (4.4) and (4.5), we obtain
GGuΠΠ∇,Eu,k=BBuwithGGu=~GGu+GG0u,and BBu=~BBu+BB0u. | (4.6) |
It is worth noting that GGu is nonsingular by construction, and we can formally state that ΠΠ∇,Eu,k=GG−1uBBu.
Recall the definition of the L2-orthogonal projector Π0,Eu,k acting on a scalar vh∈Vhu,k(E), is
∫E(Π0,Eu,kvh)qdV=∫EvhqdV,∀q∈Pk(E). | (4.7) |
Rewrite Eq (4.7) using vector and matrix notation as
∫EmmuΠ0,Eu,kφφTudV=∫EmmuφφTudV. | (4.8) |
We replace Π0,Eu,kφφTu with mmTuΠΠ0,Eu,k, and obtain
∫EmmuφφTudV=∫EmmummTuΠΠ0,Eu,kdV=∫EmmummTudVΠΠ0,Eu,k. | (4.9) |
The above equation can be rewritten as
HuΠΠ0,Eu,k=Cu, | (4.10) |
where
Hu=∫EmmummTudV, |
and
Cu=∫EmmuφφTudV. |
We remark that Hu is nonsingular, therefore ΠΠ0,Eu,k=H−1uCu.
In addition, the L2-orthogonal projector acting on ∇vh, for vh∈Vhu,k(E) is defined as
∫E(Π0,Eu,k−1∇vh)⋅qdV=∫E∇vh⋅qdV=−∫Evh∇⋅qdV+∫∂Evhq⋅ndS,∀q∈[Pk−1(E)]2. | (4.11) |
We can choose q=[qup,0]T or q=[0,qdown]T where qup,qdown∈Pk−1(E). The above equation can be split into the following two equations,
∫E(Π0,Eu,k−1∂xvh)qupdV=∫E∂xvhqupdV,∀qup∈Pk−1(E), | (4.12) |
∫E(Π0,Eu,k−1∂yvh)qdowndV=∫E∂yvhqdowndV,∀qdown∈Pk−1(E). | (4.13) |
We start with the first equation,
∫Emmu,k−1(∂xφφu)TdV=∫Emmu,k−1(Π0,Eu,k−1(∂xφφu)T)dV=∫Emmu,k−1(∂xmmu)TdVΠΠ0,x,Eu,k−1. |
As a result, the above equation can be cast into
Hu,k−1ΠΠ0,x,Eu,k−1=Cxu,k−1, | (4.14) |
where
Hu,k−1=∫Emmu,k−1(∂xmmu)TdV=∫E(mmu,k−1)(mmu,k−1)TdV, | (4.15) |
and
Cxu,k−1=∫Emmu,k−1(∂xφφu)TdV=−∫E∂xmmu,k−1φφTudV+∫∂E(mmu,k−1)nxφφTudS. | (4.16) |
Following a similar procedure, we have the matrix equation for ΠΠ0,y,Eu,k−1
Hu,k−1ΠΠ0,y,Eu,k−1=Cyu,k−1, | (4.17) |
where
Cyu,k−1=∫Emmu,k−1(∂yφφu)TdV=−∫E∂ymmu,k−1φφTudV+∫∂E(mmu,k−1)nyφφTudS. | (4.18) |
It is easy to see that Hu,k−1 is invertible, therefore,
ΠΠ0,x,Eu,k−1=H−1u,k−1Cxu,k−1andΠΠ0,y,Eu,k−1=H−1u,k−1Cyu,k−1. |
After constructing projection matrices, we can assemble the local mass and stiffness matrices. We follow two major ideas in this section. The first idea is that we split the vector-valued function into two scalar-valued components, i.e.,
u∈Vhu,k(E)=[Vhu,k(E)]2,andu=[uup,0]T+[0,udown]T,where uup,udown∈Vhu,k(E). | (4.19) |
Second, it is well-known that we can split the local matrix into consistency and stability terms in VEM. The consistency term is calculated using polynomials, while the stability term is approximated using the local degrees of freedom.
Using the splitting idea Eq (4.19), we can rewrite the mass term as in the variational form Eq (2.8) as
MEu(u,v)=∫eρuupvup+ρudownvdowndV,∀u,v∈Vhu,k(E). | (4.20) |
The local mass matrix is defined as
MMuE:=MMu,consE+MMu,stabE. | (4.21) |
We neglect the stability term of the local mass matrix, namely, MMu,stabE=0. Therefore, the local mass matrix is given by
MMuE=MMu,consE:=[MMu,cons,up,upE00MMu,cons,down,downE]. |
Using the L2-projector Π0,Eu,k and adopting the vector and matrix notations, we have
MMu,cons,up,upE=MMu,cons,down,downE=∫Eρ(Π0,Eu,kφφu)⋅(Π0,Eu,kφφu)dV=ρ(ΠΠ0,Eu,k)THuΠΠ0,Eu,k, | (4.22) |
where we have assumed the density of the material, ρ, is a constant.
Next, substituting Eqs (2.2) and (2.3) into the stiffness term in the variational formulation Eq (2.8), we arrive at
AEu(u,v)=AE,devu(u,v)+AE,volu(u,v), |
where
AE,devu(u,v):=2μ∫Egu(x)σdev(u):ε(v)dV, and AE,volu(u,v):=(λ+μ)∫Egu(x)σvol(u):ε(v)dV. |
Moreover, using Eq (4.19) and expanding ε(u), we further obtain
AE,devu(u,v)=μ∫Egu(∂xvup)(∂xuup)+gu(∂yvup)(∂yuup)dV+μ∫Egu(∂yvup)(∂xudown)−gu(∂xvup)(∂yudown)dV+μ∫Egu(∂xvdown)(∂yuup)−gu(∂yvdown)(∂xuup)dV+μ∫Egu(∂xvdown)(∂xudown)+gu(∂yvdown)(∂yudown)dV, | (4.23) |
and
AE,volu(u,v)=(λ+μ)∫Egu(∂xvup)(∂xuup)dV+(λ+μ)∫Egu(∂xvup)(∂yudown)dV+(λ+μ)∫Egu(∂yvdown)(∂xuup)dV+(λ+μ)∫Egu(∂yvdown)(∂yudown)dV. | (4.24) |
Now, we are ready to assemble the stiffness matrix, which consists of consistency and stability terms as
KKuE:=KKu,consE+KKu,stabE. | (4.25) |
From (4.23), we obtain the consistency term
KKu,consE=[KKu,cons,up,upE,devKKu,cons,up,downE,dev[.2cm]KKu,cons,down,upE,devKKu,cons,down,downE,dev]+[KKu,cons,up,upE,volKKu,cons,up,downE,vol[.2cm]KKu,cons,down,upE,volKKu,cons,down,downE,vol]. | (4.26) |
We first address the deviatoric terms, and the volumetric components follow similarly. Using Π0,x,Eu,k−1, Π0,y,Eu,k−1, and Eq (4.23), we obtain
KKu,cons,up,upE,dev=μ(ΠΠ0,x,Eu,k−1)THxxgΠΠ0,x,Eu,k−1+μ(ΠΠ0,y,Eu,k−1)THyygΠΠ0,y,Eu,k−1,KKu,cons,up,downE,dev=μ(ΠΠ0,y,Eu,k−1)THyxgΠΠ0,x,Eu,k−1−μ(ΠΠ0,x,Eu,k−1)THxygΠΠ0,y,Eu,k−1,KKu,cons,down,upE,dev=μ(ΠΠ0,x,Eu,k−1)THxygΠΠ0,y,Eu,k−1−μ(ΠΠ0,y,Eu,k−1)THyxgΠΠ0,x,Eu,k−1,KKu,cons, down,downE,dev=μ(ΠΠ0,x,Ek−1)THxxgΠΠ0,x,Eu,k−1+μ(ΠΠ0,y,Eu,k−1)THyygΠΠ0,y,Eu,k−1, |
where
Hxxg:=∫Egu(x)(∂xmu)(∂xmu)TdV,Hxyg:=∫Egu(x)(∂xmu)(∂ymu)TdV,Hyxg:=∫Egu(x)(∂ymu)(∂xmu)TdV,Hyyg:=∫Egu(x)(∂ymu)(∂ymu)TdV. |
It is evident that
Hxxg=Hxyg=Hyxg=Hyyg=∫Egu(x)(mmu,k−1)(mmu,k−1)TdV. | (4.27) |
Then, the components of the volumetric term of the local stiffness matrix can be derived from Eq (4.24) as,
KKu,cons,up,upE,vol=(λ+μ)(ΠΠ0,x,Eu,k−1)THxxgΠΠ0,x,Eu,k−1,KKu,cons,up,downE,vol=(λ+μ)(ΠΠ0,x,Eu,k−1)THxygΠΠ0,y,Eu,k−1,KKu,cons,down,upE,vol=(λ+μ)(ΠΠ0,y,Eu,k−1)THyxgΠΠ0,x,Eu,k−1,KKu,cons,down,downE,vol=(λ+μ)(ΠΠ0,y,Eu,k−1)THyygΠΠ0,y,Eu,k−1. |
Finally, we follow [26] and adopt the stability term of the stiffness matrix as
KKu,stabE=max(2μ,λ)[(II−DDuΠΠ∇,Eu,k)T(II−DDuΠΠ∇,Eu,k)00(II−DDuΠΠ∇,Eu,k)T(II−DDuΠΠ∇,Eu,k)]. |
The stability matrix KKu,stabE should scale with the consistency matrix KKu,consE, which depends on the Lame parameters. Therefore, we let KKu,stabE scale with max(2μ,λ). We remark that more options to design the stability term can be found in [17,18], and [29,30] for nearly incompressible material, where the stabilization term depends on the Lame parameters.
Using the compact notation again, we rewrite the virtual element approximation of the first term of Fhu, see Eq (3.11), as
FEu(vh)=∫Ef⋅Π0,Eu,kvhdV=∫Ef⋅[Π0,Eu,kφφuΠ0,Eu,kφφu]dV=∫Ef⋅[mTuΠΠ0,Eu,kmTuΠΠ0,Eu,k]dV. | (4.28) |
In this section, we follow a similar procedure as in Section 4.2 and discuss how to construct elliptic and L2-orthogonal projectors, and the fourth-order equation's local mass and stiffness matrices.
In this section, we first construct Π∇,Ed,r, followed by ΠL,Ed,r. The elliptic projector Π∇,Ed,r is defined as, for vh∈Vhd,k(E)
AEd,1(Π∇,Ed,rvh,Π∇,Ed,rvh)=∫E(∇Π∇,Ed,rvh)⋅∇qdV=∫E∇vh⋅∇qdV,∀q∈Pr(E). |
The above equation can be recast into the following matrix equation:
~GGd,1ΠΠ∇,Ed,r=~BBd,1, | (4.29) |
where
~GGd,1:=∫E∇mmd⋅∇mmTddV,~BBd,1:=∫E∇mmd⋅∇φφTddV. |
To fix the kernel of the differential operator ∇, we additionally require
GG0dΠΠ∇,Ed,r=BB0d,withGG0d:=(∫∂EmmTddS0T⋮0T),andBB0d:=(∫∂EφφTddS0T⋮0T). | (4.30) |
Gathering Eqs (4.29) and (4.30), we arrive at
GGd,1ΠΠ∇,Eu,r=BBd,1withGGu=~GGd,1+GG0d,and BBd=~BBd,1+BB0d. | (4.31) |
Next, from the definition of ΠL,Ed,r,
α2AEd,2(ΠL,Ed,rvh,ΠL,Ed,rvh)+α1AEd,1(ΠL,Ed,rvh,ΠL,Ed,rvh)=α2∫EΔvhΔqdV+α1∫E∇vh⋅∇qdV,∀q∈Pr(E), |
we obtain
~GGdΠΠL,Ed,r=~BBd, | (4.32) |
where
~GGd:=α2~GGd,2+α1~GGd,1,~GGd,2:=∫EΔmmdΔmmTddV,~BBd:=α2~BBd,2+α1~BBd,1,~BBd,2:=∫EΔmmdΔφφTddV. |
A (double) integration by parts of the right-hand-side of the definition of ~BBd yields
α2~BBd,2+α1~BBd,1=α2∫EΔmmdΔφφTddV+α1∫E∇mmd⋅∇φφTddV=∫E[α2Δ2mmd−α1Δmmd]φφddV+∫∂E[(∂n(−α2mmd+α1mmd))φφTd+α2Δmmd(∂nφφTd)]dS. |
Collecting Eqs (4.30) and (4.32), we have
GGdΠΠL,Ed,r=BBd,withGGd=~GGd+GG0d,BBd=~BBd+BB0d. | (4.33) |
It is worth noting that Eq (4.30) also fixes the kernel of L, namely, GG0dΠΠL,Ed,r=BB0d. Therefore, GGd,1 and GGd are nonsingular by construction, and we can formally state that
ΠΠ∇,Ed,r=GG−1d,1BBd,1,andΠΠL,Ed,r=GG−1dBBd. | (4.34) |
Recall that Π0,Ed,r is defined as, for vh∈Vhd,k(E)
AEd,0(Π0,Ed,rvh,Π0,Ed,rvh)=∫E(Π0,Ed,rvh)qdV=∫EvhqdV,∀q∈Pr(E). |
The construction of the L2 projection matrix ΠΠ0,Ed,r is similar to the derivation of Eq (4.10), therefore
ΠΠ0,Ed,r=H−1dCd, | (4.35) |
where
Hd:=∫EmmdmmTddV, and Cd:=∫EmmdφφTddV. |
The construction of the local mass matrices uses the L2-orthogonal projection, Π0,Ed,r. Following a similar procedure to obtain Eq (4.22), we have
MMdE:=MMd,consE=α0Ad,0(Π0,Ed,kφφd,Π0,Ed,kφφTd)=α0(ΠΠ0,Ed,k)THdΠΠ0,Ed,k, |
and
MMd,HtE:=MMd,Ht,consE=Agd(Π0,Ed,kφφd,Π0,Ed,kφφTd)=(ΠΠ0,Ed,k)THd,HtΠΠ0,Ed,k, |
where we have assumed the linear function gd(d)=d and
Hd,Ht:=∫EHt(x)mmdmmTddV. |
Note that we do not need to specify stabilization terms in the local mass matrices.
The stiffness matrix is given by the sum of two terms: a rank-deficient consistency term, which guarantees the consistency of the approximation, and a stability term, which fixes the correct rank:
KKdE:=KKd,consE+KKd,stabE, |
where
KKd,consE=α2Ad,2(ΠL,Ed,rφφd,ΠL,Ed,rφφd)+α1Ad,1(ΠL,Ed,rφφd,ΠL,Ed,rφφd)=(ΠΠL,Ed,r)T~GGdΠΠL,Ed,r |
and
KKstabE=(II−DDdΠΠL,Ed,r)T(II−DDdΠΠL,Ed,r). |
It is worth noting that we can also use other stabilizations as in [31].
In this work, we present, in detail, using vector and matrix notation, how to implement two distinct virtual element methods for approximating a time-dependent, second-order momentum balance equation and a fourth-order elliptic equation. Specifically for the momentum equation, we discuss the implementation details by adopting a deviatoric and volumetric strain split. Such split is essential to model various material behavior such as hardening and softening.
The momentum balance equation governs the linear elastodynamic IBVP; the second equation mathematically models crack propagation in materials. Future works include coupling the second-order elastodynamic equation with the fourth-order phase-field equation. When connected, these two PDEs constitute fourth-order phase-field models of dynamic-brittle fractures, similar to those presented in [24,25]. We plan to solve the coupled equations using virtual element methods discussed in the work and study dynamic brittle fracture problems.
The authors declare they have not used any Artificial Intelligence (AI) tools in the creation of this article.
The authors gratefully acknowledge the support of the Laboratory Directed Research and Development (LDRD) program of Los Alamos National Laboratory under project number 20220129ER. Los Alamos National Laboratory is operated by Triad National Security, LLC, for the National Nuclear Security Administration of U.S. Department of Energy (Contract No. 89233218CNA000001).
All authors declare no conflicts of interest in this paper.
[1] |
F. Brezzi, L. D. Marini, Virtual element methods for plate bending problems, Comput. Methods Appl. Mech. Eng., 253 (2013), 455–462. https://doi.org/10.1016/j.cma.2012.09.012 doi: 10.1016/j.cma.2012.09.012
![]() |
[2] |
K. Berbatov, B. S. Lazarov, A. P. Jivkov, A guide to the finite and virtual element methods for elasticity, Appl. Numer. Math., 169 (2021), 351–395. https://doi.org/10.1016/j.apnum.2021.07.010 doi: 10.1016/j.apnum.2021.07.010
![]() |
[3] |
L. B. Da Veiga, F. Brezzi, L. D. Marini, Virtual elements for linear elasticity problems, SIAM J. Numer. Anal., 51 (2013), 794–812. https://doi.org/10.1137/120874746 doi: 10.1137/120874746
![]() |
[4] |
P. F. Antonietti, G. Manzini, S. Scacchi, M. Verani, A review on arbitrarily regular conforming virtual element methods for second-and higher-order elliptic partial differential equations, Math. Mod. Meth. Appl. Sci., 31 (2021), 2825–2853. https://doi.org/10.1142/S0218202521500627 doi: 10.1142/S0218202521500627
![]() |
[5] |
L. Beirão da Veiga, C. Lovadina, A. Russo, Stability analysis for the virtual element method, Math. Mod. Meth. Appl. Sci., 27 (2017), 2557–2594. https://doi.org/10.1142/S021820251750052X doi: 10.1142/S021820251750052X
![]() |
[6] |
S. C. Brenner, Q. Guan, L. Y. Sung, Some estimates for virtual element methods, Comput. Methods Appl. Math., 17 (2017), 553–574. https://doi.org/10.1515/cmam-2017-0008 doi: 10.1515/cmam-2017-0008
![]() |
[7] |
S. C. Brenner, L. Y. Sung, Virtual element methods on meshes with small edges or faces, Math. Mod. Meth. Appl. Sci., 28 (2018), 1291–1336. https://doi.org/10.1142/S0218202518500355 doi: 10.1142/S0218202518500355
![]() |
[8] |
H. Chi, L. B. Da Veiga, G. H. Paulino, Some basic formulations of the virtual element method (VEM) for finite deformations, Comput. Methods Appl. Mech. Eng., 318 (2017), 148–192. https://doi.org/10.1016/j.cma.2016.12.020 doi: 10.1016/j.cma.2016.12.020
![]() |
[9] |
L. B. Da Veiga, C. Lovadina, D. Mora, A virtual element method for elastic and inelastic problems on polytope meshes, Comput. Methods Appl. Mech. Eng., 295 (2015), 327–346. https://doi.org/10.1016/j.cma.2015.07.013 doi: 10.1016/j.cma.2015.07.013
![]() |
[10] |
E. Artioli, L. B. Da Veiga, C. Lovadina, E. Sacco, Arbitrary order 2D virtual elements for polygonal meshes: part I, elastic problem, Comput. Mech., 60 (2017), 355–377. https://doi.org/10.1007/s00466-017-1404-5 doi: 10.1007/s00466-017-1404-5
![]() |
[11] |
A. Cangiani, E. H. Georgoulis, T. Pryer, O. J. Sutton, A posteriori error estimates for the virtual element method, Numer. Math., 137 (2017), 857–893. https://doi.org/10.1007/s00211-017-0891-9 doi: 10.1007/s00211-017-0891-9
![]() |
[12] | Y. Leng, L. Svolos, I. D. Boureima, J. N. Plohr, G. Manzini, H. M. Mourad, Virtual element methods for the solution of the fourth-order phase-field model of quasi-brittle fracture, unpublished work, 2023. |
[13] | P. A. Raviart, J. M. Thomas, Introduction à l'analyse numérique des équations aux dérivées partielles, Collection Mathématiques Appliquées pour la Maîtrise, Paris: Masson, 1983. |
[14] | S. C. Brenner, L. R. Scott, The mathematical theory of finite element methods, New York: Springer Science & Business Media, 2008. https://doi.org/10.1007/978-0-387-75934-0 |
[15] |
N. M. Newmark, A method of computation for structural dynamics, J. Eng. Mech. Div., 85 (1959), 67–94. https://doi.org/10.1061/JMCEA3.0000098 doi: 10.1061/JMCEA3.0000098
![]() |
[16] | P. G. Ciarlet, The finite element method for elliptic problems, SIAM, 2002. https://doi.org/10.1137/1.9780898719208 |
[17] |
F. Dassi, L. Mascotto, Exploring high-order three dimensional virtual elements: bases and stabilizations, Comput. Math. Appl., 75 (2018), 3379–3401. https://doi.org/10.1016/j.camwa.2018.02.005 doi: 10.1016/j.camwa.2018.02.005
![]() |
[18] |
T. R. Liu, F. Aldakheel, M. H. Aliabadi, Virtual element method for phase field modeling of dynamic fracture, Comput. Methods Appl. Mech. Eng., 411 (2023), 116050. https://doi.org/10.1016/j.cma.2023.116050 doi: 10.1016/j.cma.2023.116050
![]() |
[19] |
L. B. Da Veiga, F. Brezzi, L. D. Marini, A. Russo, The hitchhiker's guide to the virtual element method, Math. Mod. Meth. Appl. Sci., 24 (2014), 1541–1573. https://doi.org/10.1142/S021820251440003X doi: 10.1142/S021820251440003X
![]() |
[20] |
O. J. Sutton, The virtual element method in 50 lines of MATLAB, Numer. Algor., 75 (2017), 1141–1159. https://doi.org/10.1007/s11075-016-0235-3 doi: 10.1007/s11075-016-0235-3
![]() |
[21] |
M. Mengolini, M. F. Benedetto, A. M. Aragón, An engineering perspective to the virtual element method and its interplay with the standard finite element method, Comput. Methods Appl. Mech. Eng., 350 (2019), 995–1023. https://doi.org/10.1016/j.cma.2019.02.043 doi: 10.1016/j.cma.2019.02.043
![]() |
[22] |
M. Frittelli, I. Sgura, Virtual element method for the Laplace-Beltrami equation on surfaces, ESAIM: Math. Modell. Numer. Anal., 52 (2018), 965–993. https://doi.org/10.1051/m2an/2017040 doi: 10.1051/m2an/2017040
![]() |
[23] |
L. Mascotto, The role of stabilization in the virtual element method: a survey, Comput. Math. Appl., 151 (2023), 244–251. https://doi.org/10.1016/j.camwa.2023.09.045 doi: 10.1016/j.camwa.2023.09.045
![]() |
[24] |
M. J. Borden, T. J. R. Hughes, C. M. Landis, C. V. Verhoosel, A higher-order phase-field model for brittle fracture: formulation and analysis within the isogeometric analysis framework, Comput. Methods Appl. Mech. Eng., 273 (2014), 100–118. https://doi.org/10.1016/j.cma.2014.01.016 doi: 10.1016/j.cma.2014.01.016
![]() |
[25] |
L. Svolos, H. M. Mourad, G. Manzini, K. Garikipati, A fourth-order phase-field fracture model: formulation and numerical solution using a continuous/discontinuous Galerkin method, J. Mech. Phys. Solids, 165 (2022), 104910. https://doi.org/10.1016/j.jmps.2022.104910 doi: 10.1016/j.jmps.2022.104910
![]() |
[26] |
P. F. Antonietti, G. Manzini, I. Mazzieri, H. M. Mourad, M. Verani, The arbitrary-order virtual element method for linear elastodynamics models: convergence, stability and dispersion-dissipation analysis, Int. J. Numer. Meth. Eng., 122 (2021), 934–971. https://doi.org/10.1002/nme.6569 doi: 10.1002/nme.6569
![]() |
[27] | D. Adak, G. Manzini, H. M. Mourad, J. N. Plohr, L. Svolos, A C1-conforming arbitrary-order two-dimensional virtual element method for the fourth-order phase-field equation, arXiv, 2023. https://doi.org/10.48550/arXiv.2307.16068 |
[28] | R. A. Adams, J. J. F. Fournier, Sobolev spaces: pure and applied mathematics, 2 Eds., Academic Press, 2003. |
[29] |
P. Wriggers, B. D. Reddy, W. Rust, B. Hudobivnik, Efficient virtual element formulations for compressible and incompressible finite deformations, Comput. Mech., 60 (2017), 253–268. https://doi.org/10.1007/s00466-017-1405-4 doi: 10.1007/s00466-017-1405-4
![]() |
[30] |
P. Krysl, Mean-strain 8-node hexahedron with optimized energy-sampling stabilization, Finite Elem. Anal. Des., 108 (2016), 41–53. https://doi.org/10.1016/j.finel.2015.09.008 doi: 10.1016/j.finel.2015.09.008
![]() |
[31] |
C. Chen, X. Huang, H. Wei, Hm-conforming virtual elements in arbitrary dimension, SIAM J. Numer. Anal., 60 (2022), 3099–3123. https://doi.org/10.1137/21M1440323 doi: 10.1137/21M1440323
![]() |
1. | Y. Leng, L. Svolos, I. Boureima, G. Manzini, J. N. Plohr, H. M. Mourad, Arbitrary Order Virtual Element Methods for High‐Order Phase‐Field Modeling of Dynamic Fracture, 2024, 0029-5981, 10.1002/nme.7605 |