Processing math: 100%
Research article

Linearization and computation for large-strain visco-elasticity

  • Received: 29 October 2021 Revised: 23 March 2022 Accepted: 28 March 2022 Published: 06 May 2022
  • Time-discrete numerical minimization schemes for simple visco-elastic materials in the Kelvin-Voigt rheology at high strains are not well posed because of the non-quasi-convexity of the dissipation functional. A possible solution is to resort to non-simple material models with higher-order gradients of deformations. However, this makes numerical computations much more involved. Here, we propose another approach that relies on local minimizers of the simple material model. Computational tests are provided that show a very good agreement between our model and the original.

    Citation: Patrick Dondl, Martin Jesenko, Martin Kružík, Jan Valdman. Linearization and computation for large-strain visco-elasticity[J]. Mathematics in Engineering, 2023, 5(2): 1-15. doi: 10.3934/mine.2023030

    Related Papers:

    [1] Jayme Vaz Jr., Edmundo Capelas de Oliveira . On the fractional Kelvin-Voigt oscillator. Mathematics in Engineering, 2022, 4(1): 1-23. doi: 10.3934/mine.2022006
    [2] Francesco Maddalena, Danilo Percivale, Franco Tomarelli . Signorini problem as a variational limit of obstacle problems in nonlinear elasticity. Mathematics in Engineering, 2024, 6(2): 261-304. doi: 10.3934/mine.2024012
    [3] Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614
    [4] Manuel Friedrich . Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005
    [5] Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082
    [6] Stefano Almi, Giuliano Lazzaroni, Ilaria Lucardesi . Crack growth by vanishing viscosity in planar elasticity. Mathematics in Engineering, 2020, 2(1): 141-173. doi: 10.3934/mine.2020008
    [7] L. De Luca, M. Ponsiglione . Variational models in elasticity. Mathematics in Engineering, 2021, 3(2): 1-4. doi: 10.3934/mine.2021015
    [8] Antonello Gerbi, Luca Dedè, Alfio Quarteroni . A monolithic algorithm for the simulation of cardiac electromechanics in the human left ventricle. Mathematics in Engineering, 2019, 1(1): 1-37. doi: 10.3934/Mine.2018.1.1
    [9] Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048
    [10] Zeljko Kereta, Valeriya Naumova . On an unsupervised method for parameter selection for the elastic net. Mathematics in Engineering, 2022, 4(6): 1-36. doi: 10.3934/mine.2022053
  • Time-discrete numerical minimization schemes for simple visco-elastic materials in the Kelvin-Voigt rheology at high strains are not well posed because of the non-quasi-convexity of the dissipation functional. A possible solution is to resort to non-simple material models with higher-order gradients of deformations. However, this makes numerical computations much more involved. Here, we propose another approach that relies on local minimizers of the simple material model. Computational tests are provided that show a very good agreement between our model and the original.



    In-time-semidiscretized problems of nonlinear visco-elasticity in Kelvin-Voigt rheology lead to a sequence of minimization problems where a current solution depends on the previous one. More precisely, starting from an initial condition y0 we must find for k=1,T/τ

    ykargminyΩ(W(y)+12τD2(y,yk1))dx , (1.1)

    where τ>0 is a time step, T is a final time, W is an elastic energy density and D is a dissipation due to viscosity. However, physically acceptable dissipation functions are not (Morrey) quasiconvex [1]. As we are interested in the limit for τ0, the problem (1.1) does not necessarily have a minimizer since, for small τ, the non-quasiconvexity of D cannot be compensated for by any convexity of W. This suggests microstructure formation due to rapidly oscillating deformation gradients, which, however, is not observed in reality. The aim of this work is to explore this discrepancy.

    Due to dissipative effects, we expect yk to be found in a small neighborhood of yk1. Thus, the minimization in (1.1) should be local. For that reason, we linearize the energy functional in (1.1) in the spirit of [7]. Thus, we obtain a quadratic energy contribution that stems from the dissipation and stress of the elastic energy. The resulting functional is not coercive, but, in spite of this, minimizers exist, as we show.

    We perform numerical experiments by comparing the original nonlinear scheme in (1.1) with our linearized scheme for small time steps that show very good quantitative agreement. Moreover, both schemes satisfy an energy balance.

    If we neglect inertial effects, the deformation y:[0,T]×ΩRd of a nonlinear viscoelastic material in Kelvin-Voigt rheology obeys the following system of equations

    div(FW(y)+˙FR(y,˙y))=f in [0,T]×Ω, (2.1a)
    (FW(y)+˙FR(y,˙y))n=g on [0,T]×ΓN, (2.1b)
    y=yD on [0,T]×ΓD, (2.1c)
    y(0,)=y0 in Ω, (2.1d)

    Here, [0,T] is a process time interval, ΩRd (d=2 or d=3) is a smooth bounded domain representing the reference configuration, ΓDΓN is a disjoint partition of Ω such that ΓD has a positive (d1)-dimensional Hausdorff measure. In addition, f:[0,T]×ΩRd is a density of body forces and g:[0,T]×ΓNRd is a density of surface forces. We also impose boundary data yD:[0,T]×ΓDRd and an initial condition y0:ΩRd.

    The properties of the material are described by the stored energy density W:Rd×d[0,], where its gradient represents the first Piola-Kirchhoff stress tensor, and by a (pseudo)potential of dissipative forces R:Rd×d×Rd×d[0,).

    The stress tensor S(F,˙F):=˙FR(F,˙F) has its origin in the viscous dissipative mechanisms of the material. Naturally, we require R(F,˙F)R(F,0)=0. The viscous stress tensor must comply with the time-continuous frame-indifference principle, meaning that for all F,˙FRd×d

    S(F,˙F)=F˜S(C,˙C)

    with some ˜S:Rd×dsym×Rd×dsymRd×dsym where C=FF and ˙C=˙FF+F˙F. This condition constrains R so that (see [1,2,12])

    R(F,˙F)=˜R(C,˙C) (2.2)

    for some non-negative function ˜R such that ˜R(C,0)=0. In other words, R must depend on the right Cauchy-Green strain tensor C and its time derivative ˙C. On the other hand, this property makes the analysis of (2.1) very difficult. To our knowledge, the existence of a solution has not been established so far. If the stored energy is enriched by the second gradient of deformation, the existence results are available in various settings; see [8,11] in three dimensions and [3,12] in one dimension.

    In any case, a standard approach to obtain solutions for such evolution problems is to introduce an implicit time discretization. For simplicity, we will assume that yD is constant in time in the sequel. Let τ>0 be a time step such that T/τN. Given an initial condition y0τ, for k=1,,T/τ, we now want to find ykτ such that

    div(FW(ykτ)+˙FR(ykτ,ykτyk1ττ))=fkτ:=τ1kτ(k1)τf(s,)ds in Ω,(FW(ykτ)+˙FR(ykτ,ykτyk1ττ))n=gkτ:=τ1kτ(k1)τg(s,)ds on ΓN,ykτ=yD on ΓD, (2.3)

    that is, ykτ should solve a discretized version of (2.1). In the following, we summarize our setting and the basic assumptions used to treat this incremental time minimization problem.

    Stored elastic energy and body forces: We assume that the stored elastic energy W:Rd×d[0,] has the following properties:

    (i)  W is smooth on matrices with positive determinants,(ii)  W(QF)=W(F) for all FRd×d and all proper rotations QSO(d),(iii)  W(F)c(1+|F|p) for all FRd×d, and for some c>0,p>1,(iv)  W(F)=+ if detF0, and limdetF0W(F)=+. (2.4)

    We introduce the nonlinear elastic energy ϕkτ:W1,p(Ω;Rd)[0,] by

    ϕkτ(y)=ΩW(y(x))dxΩfkτ(x)y(x)dxΓNgkτ(x)y(x)dS (2.5)

    for y:W1,p(Ω;Rd)Rd. In what follows, we assume that ϕkτ:W1,p(Ω;Rd)R{+} is sequentially weakly lower semicontinuous.

    Dissipation potential and viscous stress: Consider a time-dependent deformation y:[0,T]×ΩRd. Viscosity is not only related to strain y(t,x), but also to strain rate ˙y(t,x) and can be expressed in terms of dissipation potential R(y,˙y). An admissible potential has to satisfy the frame indifference in the sense (see [1,12])

    R(F,˙F)=R(QF,Q(˙F+AF))   QSO(d),ASkew(d) (2.6)

    for all FGL+(d) and ˙FRd×d, where

    GL+(d)={FRd×d:detF>0},Skew(d)={ARd×d:A=A}.

    Following the discussion in [12,Section 2.2], from the point of view of modeling, it is much more convenient to postulate the existence of a (smooth) global distance D:GL+(d)×GL+(d)[0,) that satisfies D(F,F)=0 for all FGL+(d). From this distance, an associated dissipation potential R can be calculated by

    R(F,˙F):=limε012ε2D2(F+ε˙F,F)=142F21D2(F,F)[˙F,˙F] (2.7)

    for FGL+(d), ˙FRd×d. Above, the fourth-order tensor 2F21D2(F1,F2) denotes the Hessian of D2 in the direction of F1 at (F1,F2).

    For the sake of simplicity, in the present work we use

    D(F1,F2)=|C1C2|,

    where Ci=FiFi for i=1,2 and |A|2=tr(AA). In this case, a direct computation yields

    R(F,˙F)=2|sym(F˙F)|2. (2.8)

    For justification of this choice for D and for further examples of admissible dissipation distances, we refer the reader to [12,Section 2.3].

    Now, the solutions to (2.3) can formally be written as solutions of the following sequence of problems:

    For a given yk1τW1,p(Ω;Rd) with yk1τ=yD on ΓD minimize ϕkτ(y)+12τΩD2(y,yk1τ)dx subject to yW1,p(Ω;Rd), y=yD on ΓD. (2.9)

    The non-quasiconvexity of FD(F,G) for a given G, see [6], would suggest that an energy-minimizing microstructure is formed in the viscous relaxation of materials. However, such a microstructure has not been observed in experiments. Thus, we propose that, in such a viscous evolution, the energy landscape is only explored in a small neighborhood of the current state, instead of the global minimization suggested by (2.9). Therefore, our approach is to minimize in each time step a problem that is linearized around yk1τ in a suitable sense.

    To motivate our linearization approach for this implicit time-discretization problem, we first consider the case of (already) linearized visco-elasticity. Here, we simply recover the original evolution equations, which (neglecting any forces) are given by the linear Kelvin-Voigt visco-elasticity problem

    div(Cu+D˙u)=0in [0,T]×Ω,u=uDon ΓD,u(0,)=u0in Ω. (3.1)

    Here C and D are positive definite fourth-order tensors of elasticity and viscosity coefficients, respectively. Note that in this linearized setting, we work with a displacement field u instead of the deformation y.

    The variational formulation for the linear counterpart of (2.3) reads: For a given uk1τW1,2(Ω;Rd) with uk1τ=uD on ΓD, find the minimizer ukτ of

    u12ΩCu:udx+12τΩD(uuk1τ):(uuk1τ)dx. (3.2)

    under the constraints uW1,2(Ω;Rd) and u=uD on ΓD. Since the functional is quadratic and convex in u, the minimizer exists in fact and is unique. It is clear that this iterative minimization scheme, in the limit of small time steps τ, leads to a solution of (3.1).

    Our idea is to reformulate the minimization problem to (3.2) in terms of minimizing an increase in displacement. More precisely, we write u=uk1τ+τz and minimize

    z12Ω(Cuk1τ:uk1τ+2τCuk1τ:z+τ2Cz:z+τDz:z)dx, (3.3)

    or, equivalently, after subtracting the constant first term and rescaling by 1τ, minimize

    z12Ω(2Cuk1τ:z+τCz:z+Dz:z)dx (3.4)

    under the constraints zW1,2(Ω;Rd) and z=0 on ΓD. We observe that the second term is dominated by the third for small τ. Therefore, we minimize

    z12Ω(2Cuk1τ:z+Dz:z)dx (3.5)

    instead. This problem admits a unique minimizer z that solves the associated Euler-Lagrange equation

    div(Cuk1τ+Dz)=0. (3.6)

    Then ukτ=uk1τ+τz fulfills

    div(Cuk1τ+1τD(ukτuk1τ))=0. (3.7)

    We note that this is an explicit Euler time step for the original evolution equation (3.1), where the standard minimization of movement results in the implicit scheme

    div(Cukτ+1τD(ukτuk1τ))=0. (3.8)

    Note that we obtained this solution from the linearized functional (3.5) and not from (3.4). This suggests that we can write a minimization problem in the velocity z for the dissipated power and still obtain a reasonable time-stepping scheme for linearized Kelvin-Voigt visco-elasticity.

    Emboldened by this result for the linearized problem, for which existence of solutions is of course already well known, we attempt to treat the fully non-linear case.

    We now consider again the fully nonlinear implicit time-stepping problem for visco-elasticity.

    Neglecting forces, we start with the (ill-posed) time-step problem of finding ykτ minimizing the problem

    yΩ(W(y)+12τD2(y,yk1τ))dx. (4.1)

    This minimization problem can be restricted to include appropriate boundary conditions. Let us, as in the linear case, see (3.3), rewrite this problem in terms of minimizing increments. By substituting ykτ=yk1τ+τzkτ, we arrive at the problem of finding zkτ as a minimizer of

    zΩ(W(yk1τ+τz)+12τD2(yk1τ+τz,yk1τ))dx. (4.2)

    We notice that for small τ the terms that are linear in τ play the crucial role in the minimization. Therefore, let us for a suitable y:ΩRd and any τ>0 define a singularly perturbed functional

    Fy,τ(z):=1τΩ(W(y+τz)W(y)+12τD2(y+τz,y))dx. (4.3)

    Notice that the functionals in (4.2) and Fyk1τ,τ of (4.3) are equivalent with respect to minimization. Since τ is small, let us explore the limit limτ0Fy,τ for a fixed y. We begin with some pointwise assessments of the density of the functional (4.3). Therefore, for simplicity, we write Y:=y and Z:=z and assume detYc>0 and |Y|,|Z|M<.

    Since W is smooth, it holds

    W(Y+τZ)=W(Y)+FW(Y):τZ+122F2W(Y+ξZ)τZ:τZ

    for some ξ[0,τ]. Thus, the first two terms in (4.3) can be rewritten as

    W(Y+τZ)W(Y)τ=FW(Y):Z+τ22F2W(Y+ξZ)Z:Z. (4.4)

    For the third term, we have (considering D(F,G)=|FFGG|)

    12τ2D(Y+τZ,Y)2=12τ2|(Y+τZ)(Y+τZ)YY|2=12|ZY+YZ+τZZ|2=12|ZY+YZ|2+τ(ZY+YZ):(ZZ)+τ22|ZZ|2=2|sym(YZ)|2+2τ Y:ZZZ+τ22|ZZ|2

    We note that terms containing τ become negligible since we have a uniform upper bound on admissible |Y| and |Z|.

    Then

    W(Y+τZ)W(Y)τ+12τ2D(Y+τZ,Y)2=FW(Y):Z+2|sym(YZ)|2+R(τ,Y,Z) (4.5)

    with the remainder satisfying

    |R(τ,Y,Z)|Cτ|Z|2+2τ|Y||Z|3+τ22|Z|4 (4.6)

    for some C=Cc,M>0. Therefore, if y,zW1,(Ω;Rd) and detyc>0, we obtain the pointwise convergence

    limτ0Fy,τ(z)=Ω(FW(y):z+2|sym((y)z)|2)dx.

    This immediately leads to the following approximation result.

    Lemma 4.1. Let yW1,(Ω;Rd) with detyc>0. On the set {zW1,(Ω;Rd):zLM}, the functional Fy,τ defined in (4.3) can be uniformly approximated by

    Fy(z)=Ω(FW(y):z+2|sym((y)z)|2)dx (4.7)

    so that

    |Fy,τ(z)Fy(z)|τC

    where C depends only on Ω, yL, c, M, and W.

    Hence, the solution to the visco-elastic evolution problem should be well approximated by

    yk=yk1+ˆτz (4.8)

    for small ˆτ, with z calculated as the minimizer of (4.7) (with y=yk1) in each time step. The heuristic interpretation of this method is that we derived a linearized functional (depending on given state) whose minimizer is the velocity in the evolution problem corresponding to the state. Equation (4.8) then is nothing but a forward Euler time step with this velocity.

    Let us explore the properties of the functional in (4.7). In the following, restricting ourselves to the physically relevant case d=3, we fix yC2(¯Ω;R3) so that infxΩdety>0.

    We immediately notice that the quadratic form in the energy (4.7) is not fully coercive, as 2|sym((y)z)|2=0 for any skew-symmetric (y)z. This is in a sense similar to the problem of linearized elasticity. It appears that the presence of the first term FW(y):z means that Fy is not bounded from below for directions z where linearized dissipation vanishes. However, this is not the case since the second Piola-Kirchhoff stress tensor (y)1FW(y) is always symmetric.

    For the sake of clarity, let us denote the energy density by

    fY(Z)=FW(Y):Z+2|sym(YZ)|2. (4.9)

    If Y1FW(Y) is symmetric, for any ZR3×3 it holds

    FW(Y):Z=YY1FW(Y) :Z=Y1FW(Y) :YZ=Y1FW(Y) :sym(YZ). (4.10)

    Therefore,

    fY(Z)=FW(Y) :Z+2|sym(YZ)|2=Y1FW(Y) :sym(YZ)+2|sym(YZ)|2=2|14Y1FW(Y)+sym(YZ)|218|Y1FW(Y)|2.

    Therefore, the minimum of fY is 18|Y1FW(Y)|2 and is achieved at Z for which it holds that sym(YZ)=14Y1FW(Y), i.e., the set of minimizers is 14(YY)1FW(Y)+YSkew(3). Moreover, we have

    fY(Z)2|sym(YZ)|2|Y1FW(Y)||sym(YZ)|

    Therefore, the density fY and, consequently, the functional Fy are bounded from below.

    Now we consider the lack of coercivity since the functional depends only on sym((y)z). Here, we rely mainly on the results in [13]. Let us start by determining the subspace

    Ny={zW1,2(Ω;R3):sym((y)z)=0}.

    First, we notice

    sym((y)z)=0sym(z(y)1)=0.

    From Lemma 4.1 in [13] follows a higher regularity, namely zC2(¯Ω;R3) and z(y)1C1,1/2(¯Ω;R3×3). In Corollary 3.10 in [13] it was shown that if we have matrix fields A,BC1(¯Ω;R3×3) such that A is skew-symmetric and B=ψ, then

    curl(AB)=0  A is constant.

    For zNy the matrix field A=z(y)1 is skew-symmetric. If we take B=y, all assumptions are fulfilled, since

    curl(AB)=curl(z(y)1y)=curlz=0.

    Therefore, A is constant and therefore z=Ay. Consequently, z=a+Ay for some vector aR3. Equivalently, z=a+ω×y for some vectors a,ωR3. Thus, we have shown

    Ny={zW1,2(Ω;R3):sym((y)z)=0}=R3+Skew(3)y.

    It is a 6-dimensional subspace.

    Theorem 4.2. For a given yC2(¯Ω;R3) with dety>0, let us consider the functional

    Fy(z)=Ω(FW(y):z+2|sym((y)z)|2)dx.

    (a) Fy admits a unique minimizer on any subspace HW1,2(Ω;R3) with HNy={0}. This includes the case H={zW1,2(Ω;R3):z|ΓD=0} for some smooth ΓDΩ with H2(ΓD)>0.

    (b) On the whole W1,2(Ω;R3) minimizers exist and are unique up to an addition of elements from Ny.

    The proof will rely on generalized Korn's second inequality. For the proof, see Corollaries 4.6, 4.7, and Section 5 in [13].

    Lemma 4.3 (Generalized Korn's second inequality).Let ΩR3 be a Lipschitz domain, and yC1(¯Ω;R3) with dety>0. Then

    zsym((y)z)L2(Ω)+zL2(Ω)

    is a norm on W1,2(Ω;R3) that is equivalent to W1,2(Ω).

    Generalizing the standard Korn's inequality, one proves that for any subspace HW1,2(Ω;R3) with HSkew(3)y={0} there exists a constant C such that for every zH

    sym((y)z)L2(Ω)CzL2(Ω).

    Proof of Theorem 4.2. First, consider a subspace H with HNy={0}. The functional Fy is strictly convex on H. This can be easily seen since, for a given YR3×3, the function is

    fY(Z)=Y1FW(Y) :sym(YZ)+2|sym(YZ)|2

    is a composition of a linear function Zsym(YZ) and a strictly convex function XY1FW(Y):X+2|X|2. Moreover, as

    Fy(z)Csym((y)z)L2(Ω)CCzL2(Ω)C,

    there is a unique minimizer zH on H. We refer to [13,Theorem 4.3] for the proof that (a) applies to {zW1,2(Ω):z|ΓD=0}, that is, {zW1,2(Ω):z|ΓD=0}Ny={0}.

    For the whole space case, let H=Ny be the orthogonal complement of Ny in W1,2(Ω;R3). By (a), there is a unique minimizer zNy on Ny. Therefore, Fy admits a minimizer in W1,2(Ω;R3), and the set of all minimizers reads

    zNy+Ny.

    Remark 4.4. Since by (2.8)

    Fy(z)=Ω(FW(y):z+R(y,z))dx,

    the Euler-Lagrange equation for the functional Fy is

    0=div(FW(y)+˙FR(y,z)) (4.11)

    or

    0=div(FW(y)+2y(z)y+2y(y)z). (4.12)

    Let us return to (4.8) and suppose that z is a minimizer of Fyk1. By plugging z=(ykyk1)/τ in (4.12), we arrive at

    div(FW(yk1)+˙FR(yk1,ykyk1τ))=0. (4.13)

    Since this is a simple time discretization of the original evolution equation (2.1a), we conjecture that our linearization method yields a solution of the original nonlinear visco-elastic problem. The convergence study below already suggests that the energy-dissipation equality is satisfied by our method.

    We now consider a numerical implementation of the time-step method given in (4.8), where the velocity z is computed in each time step minimizing (4.7). For simplicity, we denote the time-step size here again by τ. Our implementation is based on finite elements, so we approximate

    yk1yk1h=j(yk1)jϕj

    with a vector yk1 of degrees of freedom and finite element basis functions ϕj. In each time step we thus solve a linear system of the form

    Kz=b,

    where the symmetric system matrix is given by

    Kij=Ωc((ϕi)yk1h+(yk1h)ϕi):((ϕj)yk1h+(yk1h)ϕj)dx (5.1)

    corresponding to the linearization of a nonlinear dissipation potential D(F,G)=c|FFGG|. Compared to the previous discussion, we introduce a scaling parameter c>0. The vector on the right-hand side is

    bi=Ω(FW(yk1h)):ϕidx.

    The linear system is solved using the sparse direct Cholesky solver CHOLMOD [5] after a small shift of K (which, without boundary conditions, is only positive semidefinite) by δM (with finite element mass matrix M) to ensure positive definiteness. In all numerical experiments below, the value of δ has been fixed at 1108.

    Finally we compute

    yk=yk1+τz (5.2)

    and iterate this procedure. The finite element system can be endowed with appropriate boundary conditions, body forces, and surface tractions in the usual way. In the following we use the Neo-Hookean material model

    W(F)=μ2(|F|232logdetF)+λ2(detF1)2 (5.3)

    with Lamé parameters μ and λ. All 3d simulations are performed using P2 finite elements on a tetrahedral mesh.

    Experiment 1: a visco-elastic plate. In [8] a rigorous thin-plate limit for nonlinear visco-elastic materials was derived in the Föppl–von Kármán-scaling regime. Numerical experiments for thin-plate equations were performed in [9] combining Bogner-Fox-Schmit (BFS) finite elements [4,14] for out-of-plane deformation v:ωR and Q1 elements for in-plane deformation u:ωR2 in the plate domain ω=(1,1)2R2. We use this setting as a first test case for our method, computing the visco-elastic relaxation of an initially deformed plate with clamped boundary conditions. As an initial condition, we choose

    v(x1,x2)=(x11)2(x21)2,u(x1,x2)=0.

    The thin-plate model minimizes the energy ϕ(u,v)+12τD2((un1,vn1),(u,v)), where the von Kármán energy functional ϕ is defined as

    ϕ(u,v):=S12Q2W(e(u)+12vv)+124Q2W(2v)dx (5.4)

    and the global dissipation distance D as

    ((u0,v0),(u1,v1)):=S(Q2D(e(u1)e(u0)+12v1v112v0v0)+112Q2D(2v12v0))1/2dx (5.5)

    respectively. As Q2W and Q2D we take the limiting Föppl–von Kármán plate material, i.e.,

    Q2W(G):=2μλ2μ+λtr2(G)+2μ|G|2,Q2D(G):=4c|G|2 , c>0 

    for every symmetric GR2×2sym. The constants λ,μ are the Lamé constants from (5.3) and c>0 is the same viscosity parameter as in (5.1).

    We then compare the solution of the thin-plate model with a full three-dimensional simulation using the algorithm from the beginning of Section 5 developed in this article, applied to a 3d plate with thickness h in a reference domain ω×(12,12) and x3 derivative rescaled by 1h. The initial configuration is given as the respective element of the recovery sequence for v, u above, as detailed in [10,Section 6.2], yielding approximately the same starting energy as the thin-plate limit. The simulation was performed using 201,514 tetrahedra resulting in 818,538 degrees of freedom. The wall time to full relaxation at t30 was approximately 6 days on 8 cores of an Intel Xeon Gold 6230.

    The parameters of this experiment are μ=λ=1.0103, c=3.0103 and h=0.1. The time-step size for the plate simulation was τplate=1, for the 3d simulation it was τ3d=0.01. A comparison of midplane deformation in the 3d simulation with the limiting plate model is shown in Figure 1. Figure 2 shows the 3d deformation at a t=2, as well as a comparison of the energy and cumulative dissipation of the plate model and the 3d simulation (see also the convergence study below). One can clearly see that our model and the plate limit yield very similar results, as one would expect. We also point out the near perfect adherence to energy-dissipation equality by the 3d simulations using the method developed here.

    Figure 1.  Comparison of the simulation results using the 3d visco-elasticity method introduced here (top row, shown are out-of-plane (v) and in-plane (u) deformation of the mid-plane of the 3d plate) and the results using the limiting viscoelastic Föppl-von Kármán plate using the method of [9] (bottom row). The in-plane deformation is exaggerated by a factor of 6.
    Figure 2.  The 3d-deformation state and the time-dependent elastic and dissipated energy.

    Experiment 2: Jello. We simulate the viscous evolution of a rectangular block of a material attached at the top under the influence of a gravitational body force. The jello has a reference configuration that occupies a domain Ω=(1,1)×(1,1)×(12,12) and an initial condition y(x)=x. A Dirichlet boundary condition fixes y|{x3=12}=x for all time. The energy is a sum of the Neo-Hookean energy in (5.3) and Ωfy3dx. The parameters in this simulation are μ=1.0103, λ=1.5103, f=2.0103, c=3.0103, and τ=0.01. Again, we see nearly perfect adherence to the energy-dissipation balance in this simulation. The number of degrees of freedom and the running time are the same as for Experiment 1 above. The results are illustrated in Figure 3.

    Figure 3.  Simulation results for the visco-elastic (jello) material. The plate at the top indicates where the specimen is fixed; a gravitational body force acts downwards.

    Convergence study. We performed a simple study testing the relative error in the energy-dissipation equality according to the time-step size τ. In the reference domain Ω=(1,1)3, we choose an initial condition

    y0=(1.200010001)x

    and perform the linearized time-stepping scheme in (5.2) until final time T=1, i.e., up to step Kτ=1τ for time steps τ{0.1,0.01,0.001}. Finally, we compute the relative error in the energy-dissipation equality with the nonlinear dissipation distance, i.e., we compute

    eτ=|W(y0)W(yKτ)Kτj=1D(yj,yj1)|Kτj=1D(yj,yj1).

    The simulation was performed using the parameters μ=1.0103, λ=1.5103, c=3.0103, no body force and stress-free boundary conditions on a domain discretized using 147 342 P2 finite elements. From Figure 4 one can clearly obtain that eτ=O(τ). The same scaling is also obtained using coarser discretizations with 71,369 and 33,216 P2 finite elements. We thus conclude that our linearized time-stepping scheme recovers the correct energy-dissipation equality for nonlinear Kelvin-Voigt visco-elasticity.

    Figure 4.  Simulation using 147,342 P2 finite elements. The slope of the linear fit is 1.00.

    PD and MJ are grateful for the hospitality afforded by The Institute of Information Theory and Automation of the Czech Academy of Sciences. MJ and MK acknowledge the support from the University of Freiburg during their stays there. PD acknowledges support by the state of Baden-Württemberg through bwHPC and from the DFG via project 441523275 in SPP 2256. MK and JV were supported by the GAČR project 21-06569K and by the MŠMT ČR project 8J21AT001 Model Reduction and Optimal Control in Thermomechanics. MK also thanks the ESI Vienna for its hospitality during his stay in January-February 2022.

    The authors declare no conflict of interest.



    [1] S. S. Antman, Physically unacceptable viscous stresses, Z. Angew. Math. Phys., 49 (1998), 980–988. https://doi.org/10.1007/s000330050134 doi: 10.1007/s000330050134
    [2] S. S. Antman, Nonlinear problems of elasticity, 2 Eds., New York: Springer, 2005. https://doi.org/10.1007/0-387-27649-1
    [3] J. M. Ball, Y. Şengül, Quasistatic nonlinear viscoelasticity and gradient flows, J. Dyn. Diff. Equat., 27 (2015), 405–442. https://doi.org/10.1007/s10884-014-9410-1 doi: 10.1007/s10884-014-9410-1
    [4] F. K. Bogner, R. L. Fox, L. A. Schmit, The generation of interelement compatible stiffness and mass matrices by the use of interpolation formulas, In: Proc. Conf. Matrix Methods in Struct. Mech, AirForce Inst. of Tech, Wright Patterson AF Base, Ohio, 1965,397–444.
    [5] Y. Chen, T. A. Davis, W. W. Hager, S. Rajamanickam, Algorithm 887: CHOLMOD, Supernodal sparse Cholesky factorization and update/downdate, ACM Trans. Math. Software, 35 (2008), 22. https://doi.org/10.1145/1391989.1391995 doi: 10.1145/1391989.1391995
    [6] B. Dacorogna, Direct methods in the calculus of variations, 2 Eds., New York: Springer, 2008. https://doi.org/10.1007/978-0-387-55249-1
    [7] G. Dal Maso, M. Negri, D. Percivale, Linearized elasticity as Γ-limit of finite elasticity, Set-Valued Analysis, 10 (2002), 165–183. https://doi.org/10.1023/A:1016577431636 doi: 10.1023/A:1016577431636
    [8] M. Friedrich, M. Kružík, On the passage from nonlinear to linearized viscoelasticity, SIAM J. Math. Anal., 50 (2018), 4426–4456. https://doi.org/10.1137/17M1131428 doi: 10.1137/17M1131428
    [9] M. Friedrich, M. Kružík, J. Valdman, Numerical approximation of von Kármán viscoelastic plates, Discrete Cont. Dyn. Syst. S, 14 (2021), 299–319. https://doi.org/10.3934/dcdss.2020322 doi: 10.3934/dcdss.2020322
    [10] G. Friesecke, R. D. James, S. Müller, A hierarchy of plate models derived from nonlinear elasticity by Gamma-convergence, Arch. Rational Mech. Anal., 180 (2006), 183–236. https://doi.org/10.1007/s00205-005-0400-7 doi: 10.1007/s00205-005-0400-7
    [11] S. Krömer, T. Roubíček, Quasistatic viscoelasticity with self-contact at large strains, J. Elast., 142 (2020), 433–445. https://doi.org/10.1007/s10659-020-09801-9 doi: 10.1007/s10659-020-09801-9
    [12] A. Mielke, C. Ortner, Y. Şengül, An approach to nonlinear viscoelasticity via metric gradient flows, SIAM J. Math. Anal., 46 (2014), 1317–1347. https://doi.org/10.1137/130927632 doi: 10.1137/130927632
    [13] P. Neff, On Korn's first inequality with non-constant coefficients, Proc. Roy. Soc. Edinb. A, 132 (2002), 221–243. https://doi.org/10.1017/S0308210500001591 doi: 10.1017/S0308210500001591
    [14] J. Valdman, MATLAB implementation of {C1 finite elements: Bogner-Fox-Schmit rectangle, In: Parallel processing and applied mathematics. PPAM 2019, Cham: Springer, 2020,256–266. https://doi.org/10.1007/978-3-030-43222-5_22
  • This article has been cited by:

    1. Julia Orlik, Viacheslav Khilkov, Stefan Rief, Heiko Andrä, Homogenization based heating control for moist paperboard with evaporation on the pore surface, 2024, 104, 0044-2267, 10.1002/zamm.202300673
    2. J. Orlik, D. Neusius, K. Steiner, M. Krier, On the ultimate strength of heterogeneous slender structures based on multi-scale stress decomposition, 2024, 195, 00207225, 104010, 10.1016/j.ijengsci.2023.104010
    3. Julia Orlik, Viacheslav Khilkov, Stefan Rief, Holger Schubert, Marek Hauptmann, Heiko Andrä, Deep Drawing of Paperboard Under Heat–Moisture Control, 2025, 13, 2227-9717, 780, 10.3390/pr13030780
  • 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(2133) PDF downloads(205) Cited by(3)

Figures and Tables

Figures(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog