On the behaviour of the solution to a viscoplastic contact problem

Abstract

The present paper represents a continuation of [2]. There, a quasistatic contact problem for viscoplastic materials was considered, in which the contact was assumed to be frictionless and was described with normal compliance and unilateral constraint; the unique weak solvability of the problem was proved, a fully discrete scheme for the numerical approximation of the problem was described and numerical simulations were presented. In the present paper we analyse the dependence of the solution of the viscoplastic contact problem in [2] with respect to the data. We state and prove a convergence result, Theorem 3.1, then we illustrate its validity in the study of a two-dimensional numerical example.

Authors

Mikael Barboteu
(Laboratoire de MathΓ©matiques et Physique, UniversitΓ© de Perpignan)

Flavius Patrulescu
(Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy)

Ahmad Ramadan
(Laboratoire de MathΓ©matiques et Physique, UniversitΓ© de Perpignan)

Mircea Sofonea
(Laboratoire de MathΓ©matiques et Physique, UniversitΓ© de Perpignan)

Keywords

viscoplastic material; frictionless contact; normal compliance; unilateral constraint; weak-solution; convergence results; numerical simulations

Cite this paper as:

M. Barboteu, F. Pătrulescu, A. Ramadan, M. Sofonea, On the behaviour of the solution to a viscoplastic contact problem, in Advances in Mathematics, eds. L. Beznea, V. Brinzănescu, M. Iosifescu, G. Marinoschi, R. Purice, D. Timotin, pp. 75-88, The Publishing House of the Romanian Academy, 2013.

PDF

About this paper

Title

Advances in Mathematics

Publisher Name

Editura Academiei Romane
(The Publishing House of the Romanian Academy)

Editors

L. Beznea, V. Brinzănescu, M. Iosifescu, G. Marinoschi, R. Purice, D. Timotin

Print ISSN
Online ISSN

MR

3203417

ZBL

?

[1] Alart, A. Curnier, A mixed formulation for frictional contact problems prone to Newton like solution methods, Computer Methods in Applied Mechanics and Engineering, 92 (1991), 353–375.
[2] Barboteu, A. Matei, M. Sofonea, Analysis of quasistatic viscoplastic contact problems with normal compliance, Quarterly Jnl. of Mechanics and App. Maths. . ?????
[3] Cristescu, I. Suliciu, Viscoplasticity, Martinus Nijhoff Publishers, Editura Tehnica, Bucharest, (1982).
[4] Han, M. Sofonea, Quasistatic Contact Problems in Viscoelasticity and Viscoplasticity, Studies in Advanced Mathematics, 30, American Mathematical Society–International Press, Sommerville, MA (2002).
[5] R. Ionescu, M. Β Sofonea, Functional and Numerical Methods in Viscoplasticity, Oxford University Press, Oxford (1993).
[6] Jarusek, M. Sofonea, On the solvability of Β dynamic elastic-visco-plastic contact problems, Zeitschrift fur Angewandte Matematik und Mechanik Β (ZAMM), 88 (2008), 3-22.
[7] B. Khenous, P. Β Laborde, Y. Renard, On the discretization of contact problems in elastodynamics, Lecture Notes in Applied Computational Β Mechanics, 27 (2006), 31-38.
[8] B. Khenous, J.-C. Pommier, Y., Renard, Hybrid discretization of the Signorini problem with Coulomb friction. Theoretical aspects and comparison of some numerical solvers, Applied Numerical Mathematics, 56 (2006), 163-192.
[9] Laursen, Computational Contact and Impact Mechanics, Springer, Berlin (2002).
[10] Shillor, M. Sofonea, J.J. Telega, Models and Analysis of Quasistatic Contact, Lecture Notes in Physics, 655, Springer, Berlin (2004).
[11] Sofonea, A. Matei. History-dependent quasivariational inequalities arising in Contact Mechanics, Eur. J. Appl. Math., 22 (2011), 471-491.
[12] Sofonea, A. Matei, Mathematical Models in Contact Mechanics, London Mathematical Society Lecture Note Series, 398, Cambridge University Press, Cambridge (2012).
[13] Wriggers, Computational Contact Mechanics, Wiley, Chichester (2002).

Paper (preprint) in HTML form

ON THE BEHAVIOR OF THE SOLUTION TO A VISCOPLASTIC CONTACT PROBLEM

M. Barboteu, βˆ— F. PΔƒtrulescul [, A. Ramadan ‑ and M. Sofonea Β§
Abstract

The present paper represents a continuation of [2]. There, a quasistatic contact problem for viscoplastic materials was considered, in which the contact was assumed to be frictionless and was described with normal compliance and unilateral constraint; the unique weak solvability of the problem was proved, a fully discrete scheme for the numerical approximation of the problem was described and numerical simulations were presented. In the present paper we analyse the dependence of the solution of the viscoplastic contact problem in [2] with respect to the data. We state and prove a convergence result, Theorem 3.1, then we illustrate its validity in the study of a two-dimensional numerical example.

2010 Mathematics Subject Classification : 74M15, 74G30, 74G25, 74S05, 35Q74.
Keywords: viscoplastic material, frictionless contact, normal compliance, unilateral constraint, weak-solution, convergence results, numerical simulations.

1 Indroduction

The aim of this paper is to study the continuous dependence of the solution to a frictionless contact problems for rate-type viscoplastic materials. We model the behavior of the material with a constitutive law of the form

πˆΛ™=β„°Ξ΅(𝒖˙)+𝒒(𝝈,Ξ΅(𝒖)),\dot{\boldsymbol{\sigma}}=\mathcal{E}\varepsilon(\dot{\boldsymbol{u}})+\mathcal{G}(\boldsymbol{\sigma},\varepsilon(\boldsymbol{u})), (1.1)
00footnotetext: *Laboratoire de MathΓ©matiques et PhySique, UniversitΓ© de Perpignan Via Domitia, 52 Avenue de Paul Alduy, 66860 Perpignan, e-mail: barboteu@univ-perp.fr
† Tiberiu Popoviciu Institute of Numerical Analysis, P.O. Box 68-1, 400110 Cluj-Napoca, Romania, e-mail: fpatrulescu@ictp.acad.ro
‑ Laboratoire de MathΓ©matiques et PhySique, UniversitΓ© de Perpignan Via Domitia, 52 Avenue de Paul Alduy, 66860 Perpignan, e-mail: ahmad.ramadan@univ-perp.fr
Β§ Laboratoire de MathΓ©matiques et PhySique, UniversitΓ© de Perpignan Via Domitia, 52 Avenue de Paul Alduy, 66860 Perpignan, e-mail: sofonea@univ-perp.fr

where 𝒖\boldsymbol{u} denotes the displacement field, 𝝈\boldsymbol{\sigma} represents the stress and 𝜺(𝒖)\boldsymbol{\varepsilon}(\boldsymbol{u}) is the linearized strain tensor. Here β„°\mathcal{E} is a fourth order tensor which describes the elastic properties of the material and 𝒒\mathcal{G} is a nonlinear constitutive function which describes its visco-plastic behavior. In (1.1) and everywhere in this paper the dot above a variable represents the derivative with respect to the time variable tt.

Various results, examples and mechanical interpretations in the study of viscoplastic materials of the form (1.1) can be found in [3, 5] and the references therein. Displacementtraction boundary value problems with such materials were considered in [5], both in the dynamic and quasistatic case. Quasistatic frictionless and frictional contact problems for materials of the form (1.1) were studied in various papers, see 10] for a survey. There, various models of contact were stated and their variational analysis, including existence and uniqueness results, was provided. The numerical analysis of the corresponding models can be found in [4] and the references therein.

A quasistatic frictionless contact problem for viscoplastic materials of the form (1.1) was recently considered in [2]. There, the process was assumed to be quasistatic and the contact was modelled by using the normal compliance condition with unilateral constraint; the unique solvability of the solution was obtained by using new arguments on history-dependent variational inequalities obtained in [11]; a convergence result was provided, which shows that the weak solution of the problem may be approached as closely as one wishes by the solution of the viscoplastic contact problem with normal compliance and infinite penetration, with a sufficiently small deformability coefficient; finally, a fully discrete scheme for the numerical approximation of the problem was implemented and numerical simulations were presented. In the present paper we analyse the dependence of the solution of the viscoplastic contact problem in [2] with respect to the data. We state and prove a convergence result, Theorem 3.1, then we illustrate its validity in the study of a two-dimensional numerical example.

The rest of the paper is structured as follows. In Section 2 we introduce the contact problem and resume the results on its unique weak solvability obtained in 2. In Section 3 we state and prove our converge result, Theorem 3.1, which represents the main result of this paper. And, finally, in Section 4 we present a numerical validation of this convergence result.

Everywhere in this paper we use the notation β„•βˆ—\mathbb{N}^{*} for the set of positive integers and ℝ+\mathbb{R}_{+} will represent the set of non negative real numbers, i.e. ℝ+=[0,+∞)\mathbb{R}_{+}=[0,+\infty). We denote by π•Šd\mathbb{S}^{d} the space of second order symmetric tensors on ℝd\mathbb{R}^{d} or, equivalently, the space of symmetric matrices of order dd. The inner product and norm on ℝd\mathbb{R}^{d} and π•Šd\mathbb{S}^{d} are defined by

𝒖⋅𝒗=uivi,‖𝒗‖=(𝒗⋅𝒗)12βˆ€π’–=(ui),𝒗=(vi)βˆˆβ„dπˆβ‹…π‰=ΟƒijΟ„ij,‖𝝉‖=(𝝉⋅𝝉)12βˆ€πˆ=(Οƒij),𝝉=(Ο„ij)βˆˆπ•Šd\begin{array}[]{llr}\boldsymbol{u}\cdot\boldsymbol{v}=u_{i}v_{i},&\|\boldsymbol{v}\|=(\boldsymbol{v}\cdot\boldsymbol{v})^{\frac{1}{2}}&\forall\boldsymbol{u}=\left(u_{i}\right),\boldsymbol{v}=\left(v_{i}\right)\in\mathbb{R}^{d}\\ \boldsymbol{\sigma}\cdot\boldsymbol{\tau}=\sigma_{ij}\tau_{ij},&\|\boldsymbol{\tau}\|=(\boldsymbol{\tau}\cdot\boldsymbol{\tau})^{\frac{1}{2}}&\forall\boldsymbol{\sigma}=\left(\sigma_{ij}\right),\boldsymbol{\tau}=\left(\tau_{ij}\right)\in\mathbb{S}^{d}\end{array}

For each Banach space XX we use the notation C(ℝ+;X)C\left(\mathbb{R}_{+};X\right) for the space of continuously functions defined on ℝ+\mathbb{R}_{+}with values in XX and, for a subset KβŠ‚XK\subset X, we still use the symbol C(ℝ+;K)C\left(\mathbb{R}_{+};K\right) for the set of continuous functions defined on ℝ+\mathbb{R}_{+}with values in KK. It is well known that C(ℝ+;X)C\left(\mathbb{R}_{+};X\right) can be organized in a canonical way as a FrΓ©chet space, i.e. as a complete metric space in which the corresponding topology is induced by a countable family of seminorms; moreover, the convergence of a sequence (xm)m\left(x_{m}\right)_{m} to the element xx, in
the space C(ℝ+;X)C\left(\mathbb{R}_{+};X\right), can be described as follows:

{xmβ†’x in C(ℝ+;X) as mβ†’βˆž if and only if maxr∈[0,n]⁑‖xm(r)βˆ’x(r)β€–Xβ†’0 as mβ†’βˆž, for each nβˆˆβ„•βˆ—\left\{\begin{array}[]{l}x_{m}\rightarrow x\text{ in }C\left(\mathbb{R}_{+};X\right)\text{ as }m\rightarrow\infty\text{ if and only if }\\ \max_{r\in[0,n]}\left\|x_{m}(r)-x(r)\right\|_{X}\rightarrow 0\text{ as }m\rightarrow\infty,\text{ for each }n\in\mathbb{N}^{*}\end{array}\right.

Equivalence (1.2) will be used several times in Section 3 of the paper.

2 The model and preliminaries

The physical setting is as follows. A viscoplastic body occupies the domain Ξ©βŠ‚β„d\Omega\subset\mathbb{R}^{d} ( d=1,2,3d=1,2,3 ) with a Lipschitz continuous boundary Ξ“\Gamma, divided into three measurable parts Ξ“1\Gamma_{1}, Ξ“2\Gamma_{2} and Ξ“3\Gamma_{3}, such that meas (Ξ“1)>0\left(\Gamma_{1}\right)>0. We use the notation 𝒙=(xi)\boldsymbol{x}=\left(x_{i}\right) for a typical point in Ξ©βˆͺΞ“\Omega\cup\Gamma and we denote by 𝝂=(Ξ½i)\boldsymbol{\nu}=\left(\nu_{i}\right) the outward unit normal at Ξ“\Gamma. Here and below the indices i,j,k,li,j,k,l run between 1 and dd and, unless stated otherwise, the summation convention over repeated indices is used. An index that follows a comma represents the partial derivative with respect to the corresponding component of the spatial variable, e.g. vi,j=βˆ‚vi/βˆ‚xjv_{i,j}=\partial v_{i}/\partial x_{j}. The body is subject to the action of body forces of density 𝒇0\boldsymbol{f}_{0}. We also assume that it is fixed on Ξ“1\Gamma_{1} and surface tractions of density 𝒇2\boldsymbol{f}_{2} act on Ξ“2\Gamma_{2}. On Ξ“3\Gamma_{3}, the body is in frictionless contact with a deformable obstacle, the so-called foundation. We assume that the problem is quasistatic, and we study the contact process in the interval of time ℝ+=[0,∞)\mathbb{R}_{+}=[0,\infty). The contact is modelled with normal compliance and unilateral constraint. Therefore, the classical formulation of the problem is the following.

Problem 𝒫\mathcal{P}. Find a displacement field 𝒖:Ω×ℝ+→ℝd\boldsymbol{u}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{d} and a stress field 𝝈:Ω×ℝ+β†’π•Šd\boldsymbol{\sigma}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{S}^{d} such that

πˆΛ™=β„°Ξ΅(𝒖˙)+𝒒(𝝈,𝜺(𝒖)) in Ξ©Γ—(0,∞),Div⁑𝝈+𝒇0=𝟎 in Ξ©Γ—(0,∞),𝒖=𝟎 on Ξ“1Γ—(0,∞),πˆπ‚=𝒇2 on Ξ“2Γ—(0,∞),uν≀g,σν+p(uΞ½)≀0,(uΞ½βˆ’g)(σν+p(uΞ½))=0} on Ξ“3Γ—(0,∞),πˆΟ„=𝟎 on Ξ“3Γ—(0,∞),𝒖(0)=𝒖0,𝝈(0)=𝝈0 in Ξ©.\begin{array}[]{rll}\dot{\boldsymbol{\sigma}}=\mathcal{E}\varepsilon(\dot{\boldsymbol{u}})+\mathcal{G}(\boldsymbol{\sigma},\boldsymbol{\varepsilon}(\boldsymbol{u}))&\text{ in }&\Omega\times(0,\infty),\\ \operatorname{Div}\boldsymbol{\sigma}+\boldsymbol{f}_{0}=\mathbf{0}&\text{ in }&\Omega\times(0,\infty),\\ \boldsymbol{u}=\mathbf{0}&\text{ on }&\Gamma_{1}\times(0,\infty),\\ \boldsymbol{\sigma}\boldsymbol{\nu}=\boldsymbol{f}_{2}&\text{ on }&\Gamma_{2}\times(0,\infty),\\ \left.\begin{array}[]{l}u_{\nu}\leq g,\quad\sigma_{\nu}+p\left(u_{\nu}\right)\leq 0,\\ \left(u_{\nu}-g\right)\left(\sigma_{\nu}+p\left(u_{\nu}\right)\right)=0\end{array}\right\}&\text{ on }&\Gamma_{3}\times(0,\infty),\\ \boldsymbol{\sigma}_{\tau}=\mathbf{0}&\text{ on }&\Gamma_{3}\times(0,\infty),\\ \boldsymbol{u}(0)=\boldsymbol{u}_{0},\boldsymbol{\sigma}(0)=\boldsymbol{\sigma}_{0}&\text{ in }&\Omega.\end{array}

Here and below, in order to simplify the notation, we do not indicate explicitly the dependence of various functions on the variables 𝒙\boldsymbol{x} or tt. Equation (2.1) represents the viscoplastic constitutive law of the material introduced in Section 1. Equation (2.2) is the equilibrium equation in which Div denotes the divergence operator for tensor valued functions. Conditions (2.3) and (2.4) are the displacement and traction boundary conditions, respectively, and condition (2.5) represents the contact condition with normal compliance
and unilateral constraint, in which σν\sigma_{\nu} denotes the normal stress, uΞ½u_{\nu} is the normal displacement, gβ‰₯0g\geq 0 and pp is a given function. This condition was first introduced in [6] and then it was used in various papers, see [11, 12] and the references therein. Condition (2.6) shows that the tangential stress on the contact surface, denoted πˆΟ„\boldsymbol{\sigma}_{\tau}, vanishes. We use it here since we assume that the contact process is frictionless. Finally, (2.7) represents the initial conditions in which 𝒖0\boldsymbol{u}_{0} and 𝝈0\boldsymbol{\sigma}_{0} denote the initial displacement and the initial stress field, respectively.

In the study of problem 𝒫\mathcal{P} we use the standard notation for Sobolev and Lebesgue spaces associated to Ξ©\Omega and Ξ“\Gamma. Moreover, we consider the spaces

V={π’—βˆˆH1(Ξ©)d:𝒗=𝟎 on Ξ“1},Q={𝝉=(Ο„ij)∈L2(Ξ©)d:Ο„ij=Ο„ji}V=\left\{\boldsymbol{v}\in H^{1}(\Omega)^{d}:\boldsymbol{v}=\mathbf{0}\text{ on }\Gamma_{1}\right\},\quad Q=\left\{\boldsymbol{\tau}=\left(\tau_{ij}\right)\in L^{2}(\Omega)^{d}:\tau_{ij}=\tau_{ji}\right\}

These are real Hilbert spaces endowed with the inner products

(𝒖,𝒗)V=∫ΩΡ(𝒖)β‹…Ξ΅(𝒗)𝑑x,(𝝈,𝝉)Q=βˆ«Ξ©πˆβ‹…π‰π‘‘x(\boldsymbol{u},\boldsymbol{v})_{V}=\int_{\Omega}\varepsilon(\boldsymbol{u})\cdot\varepsilon(\boldsymbol{v})dx,\quad(\boldsymbol{\sigma},\boldsymbol{\tau})_{Q}=\int_{\Omega}\boldsymbol{\sigma}\cdot\boldsymbol{\tau}dx

and the associated norms βˆ₯β‹…βˆ₯V\|\cdot\|_{V} and βˆ₯β‹…βˆ₯Q\|\cdot\|_{Q}, respectively. Here Ξ΅\varepsilon represents the deformation operator given by

Ξ΅(𝒗)=(Ξ΅ij(𝒗)),Ξ΅ij(𝒗)=12(vi,j+vj,i)βˆ€π’—βˆˆH1(Ξ©)d\varepsilon(\boldsymbol{v})=\left(\varepsilon_{ij}(\boldsymbol{v})\right),\quad\varepsilon_{ij}(\boldsymbol{v})=\frac{1}{2}\left(v_{i,j}+v_{j,i}\right)\quad\forall\boldsymbol{v}\in H^{1}(\Omega)^{d}

Completeness of the space ( V,βˆ₯β‹…βˆ₯VV,\|\cdot\|_{V} ) follows from the assumption meas (Ξ“1)>0\left(\Gamma_{1}\right)>0, which allows the use of Korn’s inequality.

For an element π’—βˆˆV\boldsymbol{v}\in V we still write 𝒗\boldsymbol{v} for the trace of VV and we denote by vΞ½v_{\nu} and 𝒗τ\boldsymbol{v}_{\tau} the normal and tangential components of 𝒗\boldsymbol{v} on Ξ“\Gamma given by vΞ½=𝒗⋅𝝂,𝒗τ=π’—βˆ’vν𝝂v_{\nu}=\boldsymbol{v}\cdot\boldsymbol{\nu},\boldsymbol{v}_{\tau}=\boldsymbol{v}-v_{\nu}\boldsymbol{\nu}. Let Ξ“3\Gamma_{3} be a measurable part of Ξ“\Gamma. Then, by the Sobolev trace theorem, there exists a positive constant c0c_{0} which depends on Ξ©,Ξ“1\Omega,\Gamma_{1} and Ξ“3\Gamma_{3} such that

‖𝒗‖L2(Ξ“3)d≀c0‖𝒗‖Vβˆ€π’—βˆˆV\|\boldsymbol{v}\|_{L^{2}\left(\Gamma_{3}\right)^{d}}\leq c_{0}\|\boldsymbol{v}\|_{V}\quad\forall\boldsymbol{v}\in V (2.8)

We turn now to the assumptions on the data. First, we assume that the elasticity tensor β„°\mathcal{E} and the nonlinear constitutive function 𝒒\mathcal{G} satisfy the following conditions.

{ (a) β„°=(β„°ijkl):Ξ©Γ—π•Šdβ†’π•Šd. (b) β„°ijkl=β„°klij=β„°jikl∈L∞(Ξ©),1≀i,j,k,l≀d. (c) There exists mβ„°>0 such that β„°π‰β‹…𝝉β‰₯mℰ‖𝝉‖2βˆ€π‰βˆˆπ•Šd, a.e. in Ξ©.\displaystyle\left\{\begin{array}[]{l}\text{ (a) }\mathcal{E}=\left(\mathcal{E}_{ijkl}\right):\Omega\times\mathbb{S}^{d}\rightarrow\mathbb{S}^{d}.\\ \text{ (b) }\mathcal{E}_{ijkl}=\mathcal{E}_{klij}=\mathcal{E}_{jikl}\in L^{\infty}(\Omega),1\leq i,j,k,l\leq d.\\ \text{ (c) There exists }m_{\mathcal{E}}>0\text{ such that }\\ \mathcal{E}\boldsymbol{\tau}\cdot\boldsymbol{\tau}\geq m_{\mathcal{E}}\|\boldsymbol{\tau}\|^{2}\forall\boldsymbol{\tau}\in\mathbb{S}^{d},\text{ a.e. in }\Omega.\end{array}\right. (2.9)
{ (a) π’’:Ξ©Γ—π•ŠdΓ—π•Šdβ†’π•Šd. (b) There exists L𝒒>0 such that β€–𝒒(𝒙,𝝈1,𝜺1)βˆ’π’’(𝒙,𝝈2,𝜺2)‖≀L𝒒(β€–πˆ1βˆ’πˆ2β€–+β€–πœΊ1βˆ’πœΊ2β€–)βˆ€πˆ1,𝝈2,𝜺1,𝜺2βˆˆπ•Šd, a.e. π’™βˆˆΞ©. (c) The mapping π’™β†¦π’’(𝒙,𝝈,𝜺) is measurable on Ξ©, for any πˆ,πœΊβˆˆπ•Šd. (d) The mapping π’™β†¦π’’(𝒙,𝟎,𝟎) belongs to Q.\displaystyle\left\{\begin{array}[]{l}\text{ (a) }\mathcal{G}:\Omega\times\mathbb{S}^{d}\times\mathbb{S}^{d}\rightarrow\mathbb{S}^{d}.\\ \text{ (b) There exists }L_{\mathcal{G}}>0\text{ such that }\\ \left\|\mathcal{G}\left(\boldsymbol{x},\boldsymbol{\sigma}_{1},\boldsymbol{\varepsilon}_{1}\right)-\mathcal{G}\left(\boldsymbol{x},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{2}\right)\right\|\\ \leq L_{\mathcal{G}}\left(\left\|\boldsymbol{\sigma}_{1}-\boldsymbol{\sigma}_{2}\right\|+\left\|\boldsymbol{\varepsilon}_{1}-\boldsymbol{\varepsilon}_{2}\right\|\right)\\ \forall\boldsymbol{\sigma}_{1},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{1},\boldsymbol{\varepsilon}_{2}\in\mathbb{S}^{d},\text{ a.e. }\boldsymbol{x}\in\Omega.\\ \text{ (c) The mapping }\boldsymbol{x}\mapsto\mathcal{G}(\boldsymbol{x},\boldsymbol{\sigma},\boldsymbol{\varepsilon})\text{ is measurable on }\Omega,\\ \text{ for any }\boldsymbol{\sigma},\boldsymbol{\varepsilon}\in\mathbb{S}^{d}.\\ \text{ (d) The mapping }\boldsymbol{x}\mapsto\mathcal{G}(\boldsymbol{x},\mathbf{0},\mathbf{0})\text{ belongs to }Q.\end{array}\right. (2.10)

Next, we assume that the normal compliance function pp is such that

{ (a) p:Ξ“3×ℝ→ℝ+β‹… (b) There exists Lp>0 such that |p(𝒙,r1)βˆ’p(𝒙,r2)|≀Lp|r1βˆ’r2|βˆ€r1,r2βˆˆβ„, a.e. π’™βˆˆΞ“3. (c) (p(𝒙,r1)βˆ’p(𝒙,r2))(r1βˆ’r2)β‰₯0βˆ€r1,r2βˆˆβ„, a.e. π’™βˆˆΞ“3. (d) The mapping π’™β†¦p(𝒙,r) is measurable on Ξ“3, for any rβˆˆβ„. (e) p(𝒙,r)=0 for all r≀0, a.e. π’™βˆˆΞ“3.\left\{\begin{array}[]{l}\text{ (a) }p:\Gamma_{3}\times\mathbb{R}\rightarrow\mathbb{R}_{+}\cdot\\ \text{ (b) There exists }L_{p}>0\text{ such that }\\ \quad\left|p\left(\boldsymbol{x},r_{1}\right)-p\left(\boldsymbol{x},r_{2}\right)\right|\leq L_{p}\left|r_{1}-r_{2}\right|\\ \quad\forall r_{1},r_{2}\in\mathbb{R},\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}.\\ \text{ (c) }\left(p\left(\boldsymbol{x},r_{1}\right)-p\left(\boldsymbol{x},r_{2}\right)\right)\left(r_{1}-r_{2}\right)\geq 0\\ \quad\forall r_{1},r_{2}\in\mathbb{R},\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}.\\ \text{ (d) The mapping }\boldsymbol{x}\mapsto p(\boldsymbol{x},r)\text{ is measurable on }\Gamma_{3},\\ \quad\text{ for any }r\in\mathbb{R}.\\ \text{ (e) }p(\boldsymbol{x},r)=0\text{ for all }r\leq 0,\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}.\end{array}\right.

Finally, we assume that the body forces and the tractions have the regularity

𝒇0∈C(ℝ+;L2(Ξ©)d),𝒇2∈C(ℝ+;L2(Ξ“2)d),\boldsymbol{f}_{0}\in C\left(\mathbb{R}_{+};L^{2}(\Omega)^{d}\right),\quad\boldsymbol{f}_{2}\in C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{2}\right)^{d}\right), (2.12)

and the initial data satisfy

𝒖0∈V,𝝈0∈Q.\boldsymbol{u}_{0}\in V,\quad\boldsymbol{\sigma}_{0}\in Q. (2.13)

Consider now the subset UβŠ‚VU\subset V, the operator P:Vβ†’VP:V\rightarrow V and the function 𝒇:ℝ+β†’V\boldsymbol{f}:\mathbb{R}_{+}\rightarrow V defined by equalities

U={π’—βˆˆV:vν≀g on Ξ“3}\displaystyle U=\left\{\boldsymbol{v}\in V:v_{\nu}\leq g\text{ on }\Gamma_{3}\right\} (2.14)
(P𝒖,𝒗)V=βˆ«Ξ“3p(uΞ½)vν𝑑aβˆ€π’–,π’—βˆˆV\displaystyle(P\boldsymbol{u},\boldsymbol{v})_{V}=\int_{\Gamma_{3}}p\left(u_{\nu}\right)v_{\nu}da\quad\forall\boldsymbol{u},\boldsymbol{v}\in V (2.15)
(𝒇(t),𝒗)V=βˆ«Ξ©π’‡0(t)⋅𝒗𝑑x+βˆ«Ξ“2𝒇2(t)⋅𝒗𝑑aβˆ€π’—βˆˆV,tβˆˆβ„+\displaystyle(\boldsymbol{f}(t),\boldsymbol{v})_{V}=\int_{\Omega}\boldsymbol{f}_{0}(t)\cdot\boldsymbol{v}dx+\int_{\Gamma_{2}}\boldsymbol{f}_{2}(t)\cdot\boldsymbol{v}da\quad\forall\boldsymbol{v}\in V,t\in\mathbb{R}_{+} (2.16)

Then, the variational formulation of Problem 𝒫\mathcal{P}, derived in [2], is the following.
Problem 𝒫V\mathcal{P}^{V}. Find a displacement field 𝒖:ℝ+β†’U\boldsymbol{u}:\mathbb{R}_{+}\rightarrow U and a stress field 𝝈:ℝ+β†’Q\boldsymbol{\sigma}:\mathbb{R}_{+}\rightarrow Q such that, for all tβˆˆβ„+t\in\mathbb{R}_{+},

𝝈(t)=β„°πœΊ(𝒖(t))+∫0t𝒒(𝝈(s),𝜺(𝒖(s)))𝑑s+𝝈0βˆ’β„°πœΊ(𝒖0)\displaystyle\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t))+\int_{0}^{t}\mathcal{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)))ds+\boldsymbol{\sigma}_{0}-\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right) (2.17)
(𝝈(t),𝜺(𝒗)βˆ’πœΊ(𝒖(t)))Q+(P𝒖(t),π’—βˆ’π’–(t))V\displaystyle(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v})-\boldsymbol{\varepsilon}(\boldsymbol{u}(t)))_{Q}+(P\boldsymbol{u}(t),\boldsymbol{v}-\boldsymbol{u}(t))_{V} (2.18)
β‰₯(𝒇(t),π’—βˆ’π’–(t))Vβˆ€π’—βˆˆU\displaystyle\geq(\boldsymbol{f}(t),\boldsymbol{v}-\boldsymbol{u}(t))_{V}\quad\forall\boldsymbol{v}\in U

The unique solvability of Problem 𝒫V\mathcal{P}^{V} is given by the following result.

Theorem 2.1 Assume that (2.9)-(2.13) hold. Then Problem 𝒫V\mathcal{P}^{V} has a unique solution, which satisfies π’–βˆˆC(ℝ+;U),𝝈∈C(ℝ+;Q)\boldsymbol{u}\in C\left(\mathbb{R}_{+};U\right),\boldsymbol{\sigma}\in C\left(\mathbb{R}_{+};Q\right).

The proof of Theorem 2.1 can be found in [2] as well as in [12, Ch. 6]. It is based on arguments of history-dependent variational inequalities obtained in [1].

3 A convergence result

We now study the dependence of the solution of Problem 𝒫V\mathcal{P}^{V} with respect to perturbations of the data. To this end, we assume in what follows that (2.9)-(2.13) hold and we denote by ( 𝒖,𝝈\boldsymbol{u},\boldsymbol{\sigma} ) the solution of Problem 𝒫V\mathcal{P}^{V} obtained in Theorem 2.1. For each ρ>0\rho>0 let 𝒒ρ,pρ,𝒇0ρ\mathcal{G}_{\rho},p_{\rho},\boldsymbol{f}_{0\rho}, 𝒇2ρ,𝒖0ρ\boldsymbol{f}_{2\rho},\boldsymbol{u}_{0\rho} and 𝝈0ρ\boldsymbol{\sigma}_{0\rho} be perturbations of 𝒒,p,𝒇0,𝒇2,𝒖0\mathcal{G},p,\boldsymbol{f}_{0},\boldsymbol{f}_{2},\boldsymbol{u}_{0} and 𝝈0\boldsymbol{\sigma}_{0}, respectively, which satisfy conditions (2.10)-(2.13). We define the operator Pρ:Vβ†’VP_{\rho}:V\rightarrow V and the function 𝒇ρ:ℝ+β†’V\boldsymbol{f}_{\rho}:\mathbb{R}_{+}\rightarrow V by equalities

(Pρ𝒖,𝒗)V=βˆ«Ξ“3pρ(uΞ½)vν𝑑aβˆ€π’–,π’—βˆˆV\displaystyle\left(P_{\rho}\boldsymbol{u},\boldsymbol{v}\right)_{V}=\int_{\Gamma_{3}}p_{\rho}\left(u_{\nu}\right)v_{\nu}da\quad\forall\boldsymbol{u},\boldsymbol{v}\in V (3.1)
(𝒇ρ(t),𝒗)V=βˆ«Ξ©π’‡0ρ(t)⋅𝒗𝑑x+βˆ«Ξ“2𝒇2ρ(t)⋅𝒗𝑑aβˆ€π’—βˆˆV,tβˆˆβ„+\displaystyle\left(\boldsymbol{f}_{\rho}(t),\boldsymbol{v}\right)_{V}=\int_{\Omega}\boldsymbol{f}_{0\rho}(t)\cdot\boldsymbol{v}dx+\int_{\Gamma_{2}}\boldsymbol{f}_{2\rho}(t)\cdot\boldsymbol{v}da\quad\forall\boldsymbol{v}\in V,t\in\mathbb{R}_{+} (3.2)

and we consider the following perturbation of the variational Problem 𝒫V\mathcal{P}^{V}.
Problem 𝒫ρV\mathcal{P}_{\rho}^{V}. Find a displacement field 𝒖ρ:ℝ+β†’U\boldsymbol{u}_{\rho}:\mathbb{R}_{+}\rightarrow U and a stress field 𝝈ρ:ℝ+β†’Q\boldsymbol{\sigma}_{\rho}:\mathbb{R}_{+}\rightarrow Q such that, for all tβˆˆβ„+t\in\mathbb{R}_{+},

𝝈ρ(t)=β„°Ξ΅(𝒖ρ(t))+∫0t𝒒ρ(𝝈ρ(s),Ξ΅(𝒖ρ(s)))𝑑s+𝝈0Οβˆ’β„°Ξ΅(𝒖0ρ)\displaystyle\boldsymbol{\sigma}_{\rho}(t)=\mathcal{E}\varepsilon\left(\boldsymbol{u}_{\rho}(t)\right)+\int_{0}^{t}\mathcal{G}_{\rho}\left(\boldsymbol{\sigma}_{\rho}(s),\varepsilon\left(\boldsymbol{u}_{\rho}(s)\right)\right)ds+\boldsymbol{\sigma}_{0\rho}-\mathcal{E}\varepsilon\left(\boldsymbol{u}_{0\rho}\right) (3.3)
(𝝈ρ(t),Ξ΅(𝒗)βˆ’Ξ΅(𝒖ρ(t)))Q+(Pρ𝒖ρ(t),π’—βˆ’π’–Ο(t))V\displaystyle\left(\boldsymbol{\sigma}_{\rho}(t),\varepsilon(\boldsymbol{v})-\varepsilon\left(\boldsymbol{u}_{\rho}(t)\right)\right)_{Q}+\left(P_{\rho}\boldsymbol{u}_{\rho}(t),\boldsymbol{v}-\boldsymbol{u}_{\rho}(t)\right)_{V} (3.4)
β‰₯(𝒇ρ(t),π’—βˆ’π’–Ο(t))Vβˆ€π’—βˆˆU\displaystyle\quad\geq\left(\boldsymbol{f}_{\rho}(t),\boldsymbol{v}-\boldsymbol{u}_{\rho}(t)\right)_{V}\quad\forall\boldsymbol{v}\in U

It follows from Theorem 2.1 that, for each ρ>0\rho>0 Problem 𝒫ρV\mathcal{P}_{\rho}^{V} has a unique solution ( 𝒖ρ,𝝈ρ\boldsymbol{u}_{\rho},\boldsymbol{\sigma}_{\rho} ) with regularity π’–ΟβˆˆC(ℝ+;U),𝝈ρ∈C(ℝ+;Q)\boldsymbol{u}_{\rho}\in C\left(\mathbb{R}_{+};U\right),\boldsymbol{\sigma}_{\rho}\in C\left(\mathbb{R}_{+};Q\right). Consider now the following assumptions.

{ There exists G:ℝ+→ℝ+such that  (a) |𝒒ρ(𝒙,𝝈,𝜺)βˆ’π’’(𝒙,𝝈,𝜺)|≀G(ρ)βˆ€πˆ,πœΊβˆˆπ•Šd, a.e. π’™βˆˆΞ©, for each Ο>0. (b) G(ρ)β†’0 as Οβ†’0.\displaystyle\left\{\begin{array}[]{l}\text{ There exists }G:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}\text{such that }\\ \text{ (a) }\left|\mathcal{G}_{\rho}(\boldsymbol{x},\boldsymbol{\sigma},\boldsymbol{\varepsilon})-\mathcal{G}(\boldsymbol{x},\boldsymbol{\sigma},\boldsymbol{\varepsilon})\right|\leq G(\rho)\\ \forall\boldsymbol{\sigma},\boldsymbol{\varepsilon}\in\mathbb{S}^{d},\text{ a.e. }\boldsymbol{x}\in\Omega,\text{ for each }\rho>0.\\ \text{ (b) }G(\rho)\rightarrow 0\quad\text{ as }\quad\rho\rightarrow 0.\end{array}\right. (3.5)
{ There exists F:ℝ+→ℝ+such that  (a) |pρ(𝒙,r)βˆ’p(𝒙,r)|≀F(ρ)βˆ€r≀g, a.e. π’™βˆˆΞ“3, for each Ο>0. (b) F(ρ)β†’0 as Οβ†’0.\displaystyle\left\{\begin{array}[]{l}\text{ There exists }F:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}\text{such that }\\ \text{ (a) }\left|p_{\rho}(\boldsymbol{x},r)-p(\boldsymbol{x},r)\right|\leq F(\rho)\\ \forall r\leq g,\text{ a.e. }\boldsymbol{x}\in\Gamma_{3},\text{ for each }\rho>0.\\ \text{ (b) }F(\rho)\rightarrow 0\quad\text{ as }\quad\rho\rightarrow 0.\end{array}\right. (3.6)
𝒇0ρ→𝒇0 in C(ℝ+;L2(Ξ©)d) as Οβ†’0.\displaystyle\boldsymbol{f}_{0\rho}\rightarrow\boldsymbol{f}_{0}\quad\text{ in }C\left(\mathbb{R}_{+};L^{2}(\Omega)^{d}\right)\quad\text{ as }\quad\rho\rightarrow 0. (3.7)
𝒇2ρ→𝒇2 in C(ℝ+;L2(Ξ“2)d) as Οβ†’0.\displaystyle\boldsymbol{f}_{2\rho}\rightarrow\boldsymbol{f}_{2}\quad\text{ in }C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{2}\right)^{d}\right)\quad\text{ as }\quad\rho\rightarrow 0. (3.8)
𝒖0ρ→𝒖0 in V as Οβ†’0.\displaystyle\boldsymbol{u}_{0\rho}\rightarrow\boldsymbol{u}_{0}\quad\text{ in }V\quad\text{ as }\quad\rho\rightarrow 0. (3.9)
𝝈0Οβ†’πˆ0 in Q as Οβ†’0.\displaystyle\boldsymbol{\sigma}_{0\rho}\rightarrow\boldsymbol{\sigma}_{0}\quad\text{ in }Q\quad\text{ as }\quad\rho\rightarrow 0. (3.10)

We have the following convergence result.

Theorem 3.1 Under assumptions (3.5)-(3.10), the solution ( 𝒖ρ,𝝈ρ\boldsymbol{u}_{\rho},\boldsymbol{\sigma}_{\rho} ) of Problem 𝒫ρV\mathcal{P}_{\rho}^{V} converges to the solution ( 𝒖,𝝈\boldsymbol{u},\boldsymbol{\sigma} ) of Problem 𝒫V\mathcal{P}^{V}, i.e.,

𝒖ρ→𝒖 in C(ℝ+;V),πˆΟβ†’πˆ in C(ℝ+;Q)\boldsymbol{u}_{\rho}\rightarrow\boldsymbol{u}\quad\text{ in }\quad C\left(\mathbb{R}_{+};V\right),\quad\boldsymbol{\sigma}_{\rho}\rightarrow\boldsymbol{\sigma}\quad\text{ in }\quad C\left(\mathbb{R}_{+};Q\right) (3.11)

as ρ→0\rho\rightarrow 0.

Proof. Let ρ>0,nβˆˆβ„•βˆ—\rho>0,n\in\mathbb{N}^{*} and let t∈[0,n]t\in[0,n]. We take 𝒗=𝒖(t)\boldsymbol{v}=\boldsymbol{u}(t) in (3.4) and 𝒗=𝒖ρ(t)\boldsymbol{v}=\boldsymbol{u}_{\rho}(t) in (2.18) and add the resulting inequalities to obtain

(𝝈ρ(t)βˆ’πˆ(t),𝜺(𝒖ρ(t))βˆ’πœΊ(𝒖(t)))Q\displaystyle\left(\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)-\boldsymbol{\varepsilon}(\boldsymbol{u}(t))\right)_{Q} (3.12)
≀(Pρ𝒖ρ(t)βˆ’P𝒖(t),𝒖(t)βˆ’π’–Ο(t))V+(𝒇ρ(t)βˆ’π’‡(t),𝒖ρ(t)βˆ’π’–(t))V\displaystyle\quad\leq\left(P_{\rho}\boldsymbol{u}_{\rho}(t)-P\boldsymbol{u}(t),\boldsymbol{u}(t)-\boldsymbol{u}_{\rho}(t)\right)_{V}+\left(\boldsymbol{f}_{\rho}(t)-\boldsymbol{f}(t),\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right)_{V}

On the other hand, using (3.3) and (2.17) we find that

𝝈ρ(t)\displaystyle\boldsymbol{\sigma}_{\rho}(t) βˆ’πˆ(t)=β„°πœΊ(𝒖ρ(t))βˆ’β„°πœΊ(𝒖(t))\displaystyle-\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)-\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)) (3.13)
+∫0t(𝒒ρ(𝝈ρ(s),𝜺(𝒖ρ(s)))βˆ’π’’(𝝈(s),𝜺(𝒖(s))))𝑑s\displaystyle+\int_{0}^{t}\left(\mathcal{G}_{\rho}\left(\boldsymbol{\sigma}_{\rho}(s),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(s)\right)\right)-\mathcal{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)))\right)ds
+𝝈0Οβˆ’πˆ0βˆ’β„°πœΊ(𝒖0ρ)+β„°πœΊ(𝒖0)\displaystyle+\boldsymbol{\sigma}_{0\rho}-\boldsymbol{\sigma}_{0}-\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0\rho}\right)+\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right)

We substitute equality (3.13) in (3.12) and use assumption (2.9) to deduce that

mβ„°\displaystyle m_{\mathcal{E}} ‖𝒖ρ(t)βˆ’π’–(t)β€–V2\displaystyle\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}^{2} (3.14)
≀(Pρ𝒖ρ(t)βˆ’P𝒖(t),𝒖(t)βˆ’π’–Ο(t))V+(𝒇ρ(t)βˆ’π’‡(t),𝒖ρ(t)βˆ’π’–(t))V\displaystyle\leq\left(P_{\rho}\boldsymbol{u}_{\rho}(t)-P\boldsymbol{u}(t),\boldsymbol{u}(t)-\boldsymbol{u}_{\rho}(t)\right)_{V}+\left(\boldsymbol{f}_{\rho}(t)-\boldsymbol{f}(t),\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right)_{V}
+(∫0t(𝒒ρ(𝝈ρ(s),𝜺(𝒖ρ(s)))βˆ’π’’(𝝈(s),𝜺(𝒖(s))))𝑑s,𝜺(𝒖(t))βˆ’πœΊ(𝒖ρ(t)))Q\displaystyle+\left(\int_{0}^{t}\left(\mathcal{G}_{\rho}\left(\boldsymbol{\sigma}_{\rho}(s),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(s)\right)\right)-\mathcal{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)))\right)ds,\boldsymbol{\varepsilon}(\boldsymbol{u}(t))-\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)\right)_{Q}
+(𝝈0Οβˆ’πˆ0,𝜺(𝒖(t))βˆ’πœΊ(𝒖ρ(t)))Q+(β„°πœΊ(𝒖0ρ)βˆ’β„°πœΊ(𝒖0),𝜺(𝒖ρ(t))βˆ’πœΊ(𝒖(t)))Q\displaystyle+\left(\boldsymbol{\sigma}_{0\rho}-\boldsymbol{\sigma}_{0},\boldsymbol{\varepsilon}(\boldsymbol{u}(t))-\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)\right)_{Q}+\left(\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0\rho}\right)-\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)-\boldsymbol{\varepsilon}(\boldsymbol{u}(t))\right)_{Q}

Next, we use the definitions (3.1) and (2.15), the monotonicity of the function pρp_{\rho} and assumption (3.6) to see that

(Pρ𝒖ρ(t)βˆ’P𝒖(t),𝒖(t)βˆ’π’–Ο(t))V=βˆ«Ξ“3(pρ(uρν(t))βˆ’p(uΞ½(t)))(uΞ½(t)βˆ’uρν(t))𝑑a\displaystyle\left(P_{\rho}\boldsymbol{u}_{\rho}(t)-P\boldsymbol{u}(t),\boldsymbol{u}(t)-\boldsymbol{u}_{\rho}(t)\right)_{V}=\int_{\Gamma_{3}}\left(p_{\rho}\left(u_{\rho\nu}(t)\right)-p\left(u_{\nu}(t)\right)\right)\left(u_{\nu}(t)-u_{\rho\nu}(t)\right)da
β‰€βˆ«Ξ“3(pρ(uΞ½(t))βˆ’p(uΞ½(t)))(uΞ½(t)βˆ’uρν(t))𝑑a\displaystyle\quad\leq\int_{\Gamma_{3}}\left(p_{\rho}\left(u_{\nu}(t)\right)-p\left(u_{\nu}(t)\right)\right)\left(u_{\nu}(t)-u_{\rho\nu}(t)\right)da
β‰€βˆ«Ξ“3|pρ(uΞ½(t))βˆ’p(uΞ½(t))||uΞ½(t)βˆ’uρν(t)|𝑑a\displaystyle\quad\leq\int_{\Gamma_{3}}\left|p_{\rho}\left(u_{\nu}(t)\right)-p\left(u_{\nu}(t)\right)\right|\left|u_{\nu}(t)-u_{\rho\nu}(t)\right|da
β‰€βˆ«Ξ“3𝑭(ρ)|uΞ½(t)βˆ’uρν(t)|𝑑a≀F(ρ)( meas Ξ“3)12‖𝒖ρ(t)βˆ’π’–β€–L2(Ξ“3)d\displaystyle\quad\leq\int_{\Gamma_{3}}\boldsymbol{F}(\rho)\left|u_{\nu}(t)-u_{\rho\nu}(t)\right|da\leq F(\rho)\left(\text{ meas }\Gamma_{3}\right)^{\frac{1}{2}}\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}\right\|_{L^{2}\left(\Gamma_{3}\right)^{d}}

Therefore, using the trace inequality (2.8), we find that

(Pρ𝒖ρ(t)βˆ’P𝒖(t),𝒖(t)βˆ’π’–Ο(t))V≀cF(ρ)‖𝒖ρ(t)βˆ’π’–(t)β€–V\left(P_{\rho}\boldsymbol{u}_{\rho}(t)-P\boldsymbol{u}(t),\boldsymbol{u}(t)-\boldsymbol{u}_{\rho}(t)\right)_{V}\leq cF(\rho)\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V} (3.15)

where, here and below in this section, cc represents a positive constant which may depend on the data but it is independent on ρ,t\rho,t and nn, and whose value may change from line to line.

Finally, let δρn>0,Hρ:ℝ+→ℝ+\delta_{\rho n}>0,H_{\rho}:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}and Ο‰0ρ>0\omega_{0\rho}>0 be defined by

δρn=maxr∈[0,n]⁑‖𝒇ρ(r)βˆ’π’‡(r)β€–V\displaystyle\delta_{\rho n}=\max_{r\in[0,n]}\left\|\boldsymbol{f}_{\rho}(r)-\boldsymbol{f}(r)\right\|_{V} (3.16)
Hρ(s)=‖𝒒ρ(𝝈ρ(s),𝜺(𝒖ρ(s)))βˆ’π’’(𝝈(s),𝜺(𝒖(s)))β€–Qβˆ€sβˆˆβ„+\displaystyle H_{\rho}(s)=\left\|\mathcal{G}_{\rho}\left(\boldsymbol{\sigma}_{\rho}(s),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(s)\right)\right)-\mathcal{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)))\right\|_{Q}\quad\forall s\in\mathbb{R}_{+} (3.17)
Ο‰0ρ=β€–πˆ0Οβˆ’πˆ0β€–Q+‖𝒖0Οβˆ’π’–0β€–V\displaystyle\omega_{0\rho}=\left\|\boldsymbol{\sigma}_{0\rho}-\boldsymbol{\sigma}_{0}\right\|_{Q}+\left\|\boldsymbol{u}_{0\rho}-\boldsymbol{u}_{0}\right\|_{V} (3.18)

Then, it is easy to see that the following inequalities hold:

(𝒇ρ(t)βˆ’π’‡(t),𝒖ρ(t)βˆ’π’–(t))V≀δρn‖𝒖ρ(t)βˆ’π’–(t)β€–V\displaystyle\left(\boldsymbol{f}_{\rho}(t)-\boldsymbol{f}(t),\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right)_{V}\leq\delta_{\rho n}\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V} (3.19)
(∫0t(𝒒ρ(𝝈ρ(s),𝜺(𝒖ρ(s)))βˆ’π’’(𝝈(s),𝜺(𝒖(s))))𝑑s,𝜺(𝒖ρ(t))βˆ’πœΊ(𝒖(t)))Q\displaystyle\left(\int_{0}^{t}\left(\mathcal{G}_{\rho}\left(\boldsymbol{\sigma}_{\rho}(s),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(s)\right)\right)-\mathcal{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)))\right)ds,\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)-\boldsymbol{\varepsilon}(\boldsymbol{u}(t))\right)_{Q} (3.20)
≀(∫0tHρ(s)𝑑s)‖𝒖(t)βˆ’π’–Ο(t)β€–V\displaystyle\quad\leq\left(\int_{0}^{t}H_{\rho}(s)ds\right)\left\|\boldsymbol{u}(t)-\boldsymbol{u}_{\rho}(t)\right\|_{V}
(𝝈0Οβˆ’πˆ0,𝜺(𝒖(t))βˆ’πœΊ(𝒖ρ(t)))Q\displaystyle\left(\boldsymbol{\sigma}_{0\rho}-\boldsymbol{\sigma}_{0},\boldsymbol{\varepsilon}(\boldsymbol{u}(t))-\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)\right)_{Q} (3.21)
+(β„°πœΊ(𝒖0ρ)βˆ’β„°πœΊ(𝒖0),𝜺(𝒖ρ(t))βˆ’πœΊ(𝒖(t)))Q\displaystyle\quad+\left(\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0\rho}\right)-\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right),\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{\rho}(t)\right)-\boldsymbol{\varepsilon}(\boldsymbol{u}(t))\right)_{Q}
≀cΟ‰0ρ‖𝒖ρ(t)βˆ’π’–(t)β€–V\displaystyle\quad\leq c\omega_{0\rho}\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}

We now combine inequalities (3.14), (3.15), (3.19)-(3.21) to deduce that

‖𝒖ρ(t)βˆ’π’–(t)β€–V≀c(F(ρ)+δρn+Ο‰0ρ)+∫0tHρ(s)𝑑s\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}\leq c\left(F(\rho)+\delta_{\rho n}+\omega_{0\rho}\right)+\int_{0}^{t}H_{\rho}(s)ds (3.22)

On the other hand, using equality (3.13), assumptions (2.9), (2.10) and notation (3.17), (3.18) we find that

β€–πˆΟ(t)βˆ’πˆ(t)β€–Q≀c‖𝒖ρ(t)βˆ’π’–(t)β€–V+∫0tHρ(s)𝑑s+Ο‰0ρ\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\leq c\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}+\int_{0}^{t}H_{\rho}(s)ds+\omega_{0\rho}

and, using (3.22) yields

β€–πˆΟ(t)βˆ’πˆ(t)β€–Q≀c(F(ρ)+δρn+Ο‰0ρ)+c∫0tHρ(s)𝑑s\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\leq c\left(F(\rho)+\delta_{\rho n}+\omega_{0\rho}\right)+c\int_{0}^{t}H_{\rho}(s)ds (3.23)

We now add inequalities (3.22) and (3.23) to obtain

‖𝒖ρ(t)βˆ’π’–(t)β€–V+β€–πˆΟ(t)βˆ’πˆ(t)β€–Q≀c(F(ρ)+δρn+Ο‰0ρ)+c∫0tHρ(s)𝑑s\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\leq c\left(F(\rho)+\delta_{\rho n}+\omega_{0\rho}\right)+c\int_{0}^{t}H_{\rho}(s)ds (3.24)

and, using (3.17) and (2.10) we deduce that

Hρ(s)≀G(ρ)+L𝒒(‖𝒖ρ(s)βˆ’π’–(s)β€–V+β€–πˆΟ(s)βˆ’πˆ(s)β€–Q)βˆ€s∈[0,t]H_{\rho}(s)\leq G(\rho)+L_{\mathcal{G}}\left(\left\|\boldsymbol{u}_{\rho}(s)-\boldsymbol{u}(s)\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}(s)-\boldsymbol{\sigma}(s)\right\|_{Q}\right)\quad\forall s\in[0,t]\text{. } (3.25)

Next, we combine (3.24) and (3.25) then we use the Gronwall inequality to see that

‖𝒖ρ(t)βˆ’π’–(t)β€–V+β€–πˆΟ(t)βˆ’πˆ(t)β€–Q≀c(G(ρ)+F(ρ)+δρn+Ο‰0ρ)ect\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\leq c\left(G(\rho)+F(\rho)+\delta_{\rho n}+\omega_{0\rho}\right)e^{ct} (3.26)

We pas to the upper bound as t∈[0,n]t\in[0,n] in (3.26) to obtain

maxt∈[0,n]⁑(‖𝒖ρ(t)βˆ’π’–(t)β€–V+β€–πˆΟ(t)βˆ’πˆ(t)β€–Q)\displaystyle\max_{t\in[0,n]}\left(\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\right) (3.27)
≀c(G(ρ)+F(ρ)+δρn+Ο‰0ρ)ecn for all Ο>0\displaystyle\quad\leq c\left(G(\rho)+F(\rho)+\delta_{\rho n}+\omega_{0\rho}\right)e^{cn}\quad\text{ for all }\quad\rho>0

We now use equalities (3.2) and (2.16) to see that

‖𝒇ρ(s)βˆ’π’‡(s)β€–V≀c‖𝒇0ρ(s)βˆ’π’‡0(s)β€–L2(Ξ©)d+c‖𝒇2ρ(s)βˆ’π’‡2(s)β€–L2(Ξ“2)d\left\|\boldsymbol{f}_{\rho}(s)-\boldsymbol{f}(s)\right\|_{V}\leq c\left\|\boldsymbol{f}_{0\rho}(s)-\boldsymbol{f}_{0}(s)\right\|_{L^{2}(\Omega)^{d}}+c\left\|\boldsymbol{f}_{2\rho}(s)-\boldsymbol{f}_{2}(s)\right\|_{L^{2}\left(\Gamma_{2}\right)^{d}}

for all s∈[0,n]s\in[0,n] and, therefore, (3.7), (3.8) and (1.2) imply that

δρnβ†’0 as Οβ†’0\delta_{\rho n}\rightarrow 0\quad\text{ as }\rho\rightarrow 0 (3.28)

On the other hand, (3.9) and (3.10) yield

Ο‰0ρ→0 as Οβ†’0\omega_{0\rho}\rightarrow 0\quad\text{ as }\rho\rightarrow 0 (3.29)

We now combine the convergences (3.5) (b), (3.6) (b), (3.28) and (3.29) with inequality (3.27) to obtain that

maxt∈[0,n]⁑(‖𝒖ρ(t)βˆ’π’–(t)β€–V+β€–πˆΟ(t)βˆ’πˆ(t)β€–Q)β†’0 as Οβ†’0\max_{t\in[0,n]}\left(\left\|\boldsymbol{u}_{\rho}(t)-\boldsymbol{u}(t)\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}(t)-\boldsymbol{\sigma}(t)\right\|_{Q}\right)\rightarrow 0\quad\text{ as }\rho\rightarrow 0 (3.30)

Since the convergence (3.30) holds for each nβˆˆβ„•βˆ—n\in\mathbb{N}^{*}, we deduce from (1.2) that (3.11) holds, which concludes the proof.

In addition to the mathematical interest in the convergence result (3.11), it is of importance from mechanical point of view, since it states that the weak solution of the problem (2.1)- (2.7) depends continuously on the viscoplastic constitutive function, the normal compliance function, the densities of body forces and surface tractions and the initial data, as well.

4 Numerical validation

We proceed with the numerical validation of the convergence result in Theorem 3.1. Details on the numerical approximation of Problem 𝒫V\mathcal{P}^{V} can be found in [2. Here we restrict ourselves to recall that for the numerical treatment of the contact condition we use the penalized method for the compliance contact combined with the augmented Lagrangean approach for the unilateral constraint. To this end, we consider additional fictitious nodes for the Lagrange multiplier in the initial mesh. The construction of these nodes depends on the contact element used for the geometrical discretization of the interface Ξ“3\Gamma_{3}. In the case of the numerical example presented below, the discretization is based on "node-torigid" contact element, which is composed by one node of Ξ“3\Gamma_{3} and one Lagrange multiplier node. More details on this discretization step and the corresponding numerical method

can be found in [1, 7, 8, 9, 13. For the numerical simulations we consider the physical setting depicted in Figure 1 and described below. There, Ξ©=(0,2)Γ—(0,4)βŠ‚β„2\Omega=(0,2)\times(0,4)\subset\mathbb{R}^{2} with Ξ“1=[0,2]Γ—{4},Ξ“2=({0}Γ—[0,4])βˆͺ({2}Γ—[0,4]),Ξ“3=[0,2]Γ—{0}\Gamma_{1}=[0,2]\times\{4\},\Gamma_{2}=(\{0\}\times[0,4])\cup(\{2\}\times[0,4]),\Gamma_{3}=[0,2]\times\{0\}. The domain Ξ©\Omega represents the cross section of a three-dimensional deformable body subjected to the action of tractions in such a way that the plane stress hypothesis is assumed. On the part Ξ“1=[0,2]Γ—{4}\Gamma_{1}=[0,2]\times\{4\} the body is clamped and, therefore, the displacement field vanishes there. Horizontal tractions act on the part ({0}Γ—[1.5,4])βˆͺ({2}Γ—[1.5,4])(\{0\}\times[1.5,4])\cup(\{2\}\times[1.5,4]) of the boundary Ξ“2\Gamma_{2} and the part ({0}Γ—[0,1.5])βˆͺ({2}Γ—[0,1.5])(\{0\}\times[0,1.5])\cup(\{2\}\times[0,1.5]) is traction free. No body forces are assumed to act on the body during the process. The body is in frictionless contact with an obstacle on the part Ξ“3=[0,2]Γ—{0}\Gamma_{3}=[0,2]\times\{0\} of the boundary. For the discretization we use 13223 elastic finite elements and 129 contact elements. The total number of degrees of freedom is equal to 13712 and we take a time step kk equal to 0.01s0.01s.

We model the material’s behavior with a constitutive law of the form (1.1) in which elasticity tensor β„°\mathcal{E} satisfies

(ℰ𝝉)Ξ±Ξ²=EΞΊ1βˆ’ΞΊ2(Ο„11+Ο„22)δαβ+E1+κταβ,1≀α,β≀2,(\mathcal{E}\boldsymbol{\tau})_{\alpha\beta}=\frac{E\kappa}{1-\kappa^{2}}\left(\tau_{11}+\tau_{22}\right)\delta_{\alpha\beta}+\frac{E}{1+\kappa}\tau_{\alpha\beta},\quad 1\leq\alpha,\beta\leq 2, (4.1)

where EE is the Young modulus, ΞΊ\kappa the Poisson ratio of the material and δαβ\delta_{\alpha\beta} denotes the Kronecker symbol. Moreover, in order to facilitate the numerical implementation, we assume that 𝒒(𝝈,𝜺(𝒖))=𝒒ρ(𝝈,𝜺(𝒖))=π’žπœΊ(𝒖)\mathcal{G}(\boldsymbol{\sigma},\boldsymbol{\varepsilon}(\boldsymbol{u}))=\mathcal{G}_{\rho}(\boldsymbol{\sigma},\boldsymbol{\varepsilon}(\boldsymbol{u}))=\mathcal{C}\boldsymbol{\varepsilon}(\boldsymbol{u}), where the tensor π’ž\mathcal{C} satisfies

(π’žπ‰)Ξ±Ξ²=Ξ³1(Ο„11+Ο„22)δαβ+Ξ³2ταβ,1≀α,β≀2.(\mathcal{C}\boldsymbol{\tau})_{\alpha\beta}=\gamma_{1}\left(\tau_{11}+\tau_{22}\right)\delta_{\alpha\beta}+\gamma_{2}\tau_{\alpha\beta},\quad 1\leq\alpha,\beta\leq 2. (4.2)

For the computation below we use the following data:

t∈[0,T],T=1s,\displaystyle t\in[0,T],\quad T=1s,
E=20000N/m2,ΞΊ=0.4,Ξ³1=20N/m2,Ξ³2=30N/m2,\displaystyle E=0000N/m^{2},\quad\kappa=4,\quad\gamma_{1}=0N/m^{2},\quad\gamma_{2}=0N/m^{2},
𝒇0=𝒇0ρ=(0,0)N/m2,\displaystyle\boldsymbol{f}_{0}=\boldsymbol{f}_{0\rho}=(0,0)N/m^{2},
𝒇2={(0,0)N/m on ({0}Γ—[0,1.5])βˆͺ({2}Γ—[0,1.5]),(8000,0)N/m on {0}Γ—[1.5,4],(βˆ’8000,0)N/m on {2}Γ—[1.5,4],\displaystyle\boldsymbol{f}_{2}=\left\{\begin{array}[]{rr}(0,0)N/m&\text{ on }(\{0\}\times[0,1.5])\cup(\{2\}\times[0,1.5]),\\ (8000,0)N/m&\text{ on }\{0\}\times[1.5,4],\\ (-8000,0)N/m&\text{ on }\{2\}\times[1.5,4],\end{array}\right.
𝒇2ρ={(0,0)N/m on ({0}Γ—[0,1.5])βˆͺ({2}Γ—[0,1.5]),(βˆ’8000+ρ,0)N/m on {0}Γ—[1.5,4],\displaystyle\boldsymbol{f}_{2\rho}=\left\{\begin{array}[]{rr}(0,0)N/m&\text{ on }(\{0\}\times[0,1.5])\cup(\{2\}\times[0,1.5]),\\ (-8000+\rho,0)N/m&\text{ on }\{0\}\times[1.5,4],\end{array}\right.
p(r)=cνr+,pρ(r)=(cν+ρ)r+,cν=200N/m2,g=0.1m,\displaystyle p(r)=c_{\nu}r_{+},\quad p_{\rho}(r)=\left(c_{\nu}+\rho\right)r_{+},\quad c_{\nu}=00N/m^{2},\quad g=1m,
𝒖0=𝒖0ρ=𝟎m,𝝈0=𝝈0ρ=𝟎N/m2.\displaystyle\boldsymbol{u}_{0}=\boldsymbol{u}_{0\rho}=\mathbf{0}m,\quad\boldsymbol{\sigma}_{0}=\boldsymbol{\sigma}_{0\rho}=\mathbf{0}N/m^{2}.

Our results are presented in Figures 2, 3 and are described in what follows.
First, in Figure 2, the deformed configuration as well as the contact interface forces at t=1st=1s are plotted in the case ρ=0\rho=0, which corresponds to problem 𝒫V\mathcal{P}^{V}. We recall that the contact follows a normal compliance condition as far as the penetration is less than the limit g=0.1mg=0.1\mathrm{~m} and, when this limit is reached, it follows a unilateral condition. As it results from the zoom depicted in Figure 2, for part of the contact nodes the complete flattening of the asperities of size g=0.1mg=0.1\mathrm{~m} was reached; therefore, these nodes (situated in the central part of the boundary Ξ“3\Gamma_{3} ) are into unilateral contact. In contrast, the nodes on the extremities of the boundary Ξ“3\Gamma_{3} remain in the status of a contact with normal compliance.

Next, we denote by (𝒖ρhk,𝝈ρhk)\left(\boldsymbol{u}_{\rho}^{hk},\boldsymbol{\sigma}_{\rho}^{hk}\right) and (𝒖hk,𝝈hk)\left(\boldsymbol{u}^{hk},\boldsymbol{\sigma}^{hk}\right) the discrete solution of the contact problems 𝒫ρV\mathcal{P}_{\rho}^{V} and 𝒫V\mathcal{P}^{V}, respectively, for a given ρ>0\rho>0. The numerical estimations of the difference

‖𝒖ρhkβˆ’π’–hkβ€–V+β€–πˆΟhkβˆ’πˆhkβ€–Q\left\|\boldsymbol{u}_{\rho}^{hk}-\boldsymbol{u}^{hk}\right\|_{V}+\left\|\boldsymbol{\sigma}_{\rho}^{hk}-\boldsymbol{\sigma}^{hk}\right\|_{Q}

at the time t=1st=1s, for various values of the parameter ρ\rho, is presented in Figure 3. It results from here that this difference converges to zero as ρ\rho tends to zero. To highlight
this study, we plot 4 deformed meshes and the associated contact forces at t=1st=1s, for ρ=104,102,1,10βˆ’2\rho=10^{4},10^{2},1,10^{-2}, respectively. One can see that for ρ=104\rho=10^{4} both the deformed mesh and the contact interface forces are very different from those in Figure 2 which, recall, corresponds to the case ρ=0\rho=0. These differences progressively disappear as ρ→0\rho\rightarrow 0 in such a way that, for ρ=10βˆ’2\rho=10^{-2}, both the deformed mesh and the contact interface forces are very close to that obtained for ρ=0\rho=0. We conclude that our results in Figures 3, 4 represent a numerical validation of the theoretical convergence result obtained in Theorem 3.1,

References

[1] Alart, P., Curnier, A., A mixed formulation for frictional contact problems prone to Newton like solution methods, Computer Methods in Applied Mechanics and Engineering, 92 (1991), 353-375.
[2] Barboteu, M., Matei, A., Sofonea, M., Analysis of quasistatic viscoplastic contact problems with normal compliance, Quarterly Journal of Mechanics and Applied Mathematics, submitted.
[3] Cristescu, N., Suliciu, I., Viscoplasticity, Martinus Nijhoff Publishers, Editura Tehnica, Bucharest, 1982.
[4] Han, W., Sofonea, M., Quasistatic Contact Problems in Viscoelasticity and Viscoplasticity, Studies in Advanced Mathematics 30, Americal Mathematical SocietyInternational Press, 2002.
[5] Ionescu, I. R., Sofonea, M., Functional and Numerical Methods in Viscoplasticity, Oxford University Press, Oxford, 1993.
[6] JaruΕ‘ek, J., Sofonea, M., On the solvability of dynamic elastic-visco-plastic contact problems, Zeitschrift fΓΌr Angewandte Matematik und Mechanik (ZAMM), 88 (2008), 3-22.
[7] Khenous, H. B., Laborde, P., Renard, Y., On the discretization of contact problems in elastodynamics, Lecture Notes in Applied Computational Mechanics, 27 (2006), 31-38.
[8] Khenous, H. B., Pommier, J.-C., Renard, Y., Hybrid discretization of the Signorini problem with Coulomb friction. Theoretical aspects and comparison of some numerical solvers, Applied Numerical Mathematics, 56 (2006), 163-192.
[9] Laursen, T., Computational Contact and Impact Mechanics, Springer, Berlin, 2002.
[10] Shillor, M., Sofonea, M. and Telega, J. J., Models and Analysis of Quasistatic Contact, Lecture Notes in Physics, 655, Springer, Berlin, 2004.
[11] Sofonea, M., Matei, A., History-dependent quasivariational inequalities arising in Contact Mechanics, European Journal of Applied Mathematics, 22 (2011), 471-491.
[12] Sofonea, M., Matei, A., Mathematical Models in Contact Mechanics, London Mathematical Society Lecture Notes 398, Cambridge University Press, Cambridge, 2012.
[13] Wriggers, P., Computational Contact Mechanics, Wiley, Chichester, 2002.

2013

Related Posts