A mixed variational formulation of a contact problem with wear

Abstract

We consider a mathematical model which describes the sliding frictional contact between a viscoplastic body and an obstacle, the so-called foundation. The process is quasistatic, the material’s behavior is described with a viscoplastic constitutive law with internal state variable and the contact is modelled with normal compliance and unilateral constraint. The wear of the contact surfaces is taken into account, and is modelled with a version of Archard’s law.

We derive a mixed variational formulation of the problem which involve implicit history-dependent operators. Then, we prove the unique weak solvability of the contact model. The proof is based on a fixed point argument proved in Sofonea et al. (Commun. Pure Appl. Anal. 7:645–658, 2008), combined with a recent abstract existence and uniqueness result for mixed variational problems, obtained in Sofonea and Matei (J. Glob. Optim. 61:591–614, 2014).

Authors

Mircea Sofonea
Laboratoire de Mathématiques et Physique, Université de Perpignan

Flavius Patrulescu
Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy

Ahmad Ramadan
Laboratoire de Mathématiques et Physique, Université de Perpignan

Keywords

viscoplastic material, frictional contact, normal compliance, unilateral constraint, wear, mixed variational formulation, history-dependent operator, weak solution

Cite this paper as

M. Sofonea, F. Pătrulescu, A. Ramadan, A mixed variational formulation of a contact problem with wear, Acta Appl. Math., vol. 153 (2018), pp. 125-146.

PDF

About this paper

Publisher Name

Springer Netherlands, Dordrecht

Print ISSN

0167-8019

Online ISSN

1572-9036

MR

3745733

ZBL

1380.74089

Google Scholar

?

Paper (preprint) in HTML form

A Mixed Variational Formulation of a Contact Problem with Wear

Mircea Sofonea 1, Flavius Patrulescu 2, Ahmad Ramadan 1
Abstract

We consider a mathematical model which describes the sliding frictional contact between a viscoplastic body and an obstacle, the so-called foundation. The process is quasistatic, the material’s behavior is described with a viscoplastic constitutive law with internal state variable and the contact is modelled with normal compliance and unilateral constraint. The wear of the contact surfaces is taken into account, and is modelled with a version of Archard’s law. We derive a mixed variational formulation of the problem which involve implicit history-dependent operators. Then, we prove the unique weak solvability of the contact model. The proof is based on a fixed point argument proved in Sofonea et al. (Commun. Pure Appl. Anal. 7:645-658, 2008), combined with a recent abstract existence and uniqueness result for mixed variational problems, obtained in Sofonea and Matei (J. Glob. Optim. 61:591-614, 2014).

Keywords viscoplastic material • Frictional contact • Normal compliance • Unilateral constraint \cdot Wear \cdot Mixed variational formulation \cdot History-dependent operator \cdot Weak solution

Mathematics Subject Classification 74M15•74G25•74G30•49J40

1 Introduction

The mathematical modelling of contact phenomena is rather complex and, usually, leads to strongly nonlinear boundary value problems. The reason arise in the fact that, as shown in [ 6,7,13,23,24,29,316,7,13,23,24,29,31 ], accurate mathematical models need to take into consideration the additional phenomena involved in contact processes. These phenomena are the friction, the heat generation, the wear and the adhesion of contacting surfaces, among others. Wear

00footnotetext: M. Sofonea
sofonea@univ-perp.fr
1 Laboratoire de Mathématiques et Physique, Université de Perpignan, 52 Avenue de Paul Alduy, 66860 Perpignan, France
2 Tiberiu Popoviciu Institute of Numerical Analysis, P.O. Box 68-1, 400110 Cluj-Napoca, Romania

is defined as the material loss or change in surface texture occurring when two surfaces of mechanical components contact each other. As the contact process evolves, the contacting surfaces evolve too, via their wear. Wear in sliding systems is often very slow but it is persisting, continuous and cumulative. Its characterization represents one of the basic tasks in the study of machine elements. Indeed, in the process of design of machine elements and tools operating in contact conditions, engineers need to know areas of contact, contact stresses, and they need to predict wear of rubbing elements.

Wear of contact surfaces represents a complex phenomenon. Following [10,25], it is customary to distinguish among the following wear types: adhesive, abrasive, contact fatigue, freeting, oxidation, corrosion and erosion. In terms of the severity of wear on the wearing surfaces, two broad types of wear phenomena have been mentionned in [1]: severe wear and mild wear. Severe wear is characterized by high wear rates, extensive plastic deformation, transfer of material to the harder counter face, and flake-like metallic wear debris. Mild wear, by contrast, is characterized by low wear rates, minimal plastic deformation, formation of a surface film protecting against metal-to-metal contact, and oxide wear debris.

Due to its crucial role in various technological and biomechanical processes, the wear phenomenon subjects of numerous experimental and theoretical studies. For instance, the evolution of wear gaps in fretting problems was studied numerically in [35], by using the finite element method. Numerical simulations of wear shapes due to pitting phenomena for various operating conditions have been investigated in [9], by using arguments of fracture mechanics. A thermoelastic wheel-rail contact problem with wear has been studied in [4]. Numerical methods for wear problems with application to implanted knee joints has been developed in [28]. An original analytical approach to wear was performed in [10]. General models for frictional contact with wear could be find in [36, 37] as well as in the survey [38]. The mathematical analysis of various models of frictional contact with wear, including existence and uniqueness results of the weak solution, was carried out in [11, 12, 26-29].

A new mathematical model which describes the equilibrium of an elastic body in frictional contact with a moving foundation was recently considered in [34]. There, the contact was modeled with a normal compliance condition with unilateral constraint associated to a sliding version of Coulomb’s law of dry friction, and the wear of the foundation was described with a version of Archard’s law. A variational formulation of the problem was derived, in a form of the system which couples a time-dependent equation for the stress field, a time-dependent variational inequality for the displacement field and an integral equation for the wear function. The unique weak solvability of the model was proved, by using arguments on time-dependent variational inequalities and fixed point. This result was completed with a convergence result which shows that the solution of a penalized frictional contact problem with wear converges to the solution of the contact model, as the penalization parameter converges to zero.

The current paper represents a continuation of [34] and contains two main novelties. The first one concerns the mathematical model since, in contrast with [34], we consider here that the deformable body is viscoplastic and we model its behavior with a viscoplastic constitutive law with internal state variable. The analysis of this model could be carried out by using arguments similar to those used in [34], with a different choice of spaces and operators. Nevertheless, we choose to present here a different approach, which consists the second novelty of this paper. Thus, in contrast with [34], we derive a mixed variational formulation of the problem in which the unknowns are the stress field, the displacement field, the internal state variable, the wear function and the Lagrange multiplier, then we prove its unique solvability by using a recent abstract existence result in the study of mixed variational problems, proved in [32]. Mixed variational problems involving Lagrange multipliers have been
used both in analysis and mechanics, in the study of minimization problems. They provide a useful framework in which a large number of problems involving unilateral constraints can be cast and can be solved numerically. Their study is based on arguments on duality, saddle points theory and fixed point. The literature in the field is extensive, see for instance [3,8,14,18][3,8,14,18] and the references therein. The analysis of various mixed variational problems associated to contact models can be found in [15, 16, 19-21], for instance.

The rest of the manuscript is structured as follows. In Sect. 2 we present the notation and some preliminary material, including a new abstract result, Thereom 2.2. In Sect. 3 we introduce the model of sliding frictional contact with wear and list the assumption on the data. Then, in Sect. 4 we derive its mixed variational formulation. In Sect. 5 we state and prove our main existence and uniqueness result, Theorem 5.1, which provides the unique solvability of the viscoplastic contact problem with wear. Finally, in Sect. 6 we present relevant particular cases of our contact model and we comment on the corresponding existence and uniqueness results. We also provide a comparison between the fixed point method used in [34] and the Lagrange multiplier method used in the current paper. At the best of our knowledge, these two methods represent the main functional methods used in the study of contact problems with unilateral constraints.

2 Notations and Preliminaries

Everywhere in this paper we use the notation +\mathbb{R}_{+}for the set of positive real numbers and \mathbb{N} for the set of positive integers. Given two sets XX and YY we use the notation X×YX\times Y for their cartesian product and, ( x,yx,y ) will represent a typical point of the set X×YX\times Y. All the vector spaces considered below are real vector spaces and, for a vector space XX, we use the notation 0X0_{X} for the zero element of XX. In addition, if ( X,XX,\|\cdot\|_{X} ) and ( Y,YY,\|\cdot\|_{Y} ) are normed spaces, then X×Y\|\cdot\|_{X\times Y} represents the norm of the space X×YX\times Y given by

zX×Y=xX+yYz=(x,y)X×Y.\|z\|_{X\times Y}=\|x\|_{X}+\|y\|_{Y}\quad\forall z=(x,y)\in X\times Y.

We use similar notation for the product of more than two sets or spaces. For a normed space XX we use the notation C(+;X)C\left(\mathbb{R}_{+};X\right) for the space of continuous functions defined on +\mathbb{R}_{+} with values in XX and, for a subset KXK\subset X, we still use the symbol C(+;K)C\left(\mathbb{R}_{+};K\right) for the set of continuous functions defined on +\mathbb{R}_{+}with values in KK.

Now, assume that (X,X)\left(X,\|\cdot\|_{X}\right) and (Y,Y)\left(Y,\|\cdot\|_{Y}\right) are normed spaces and 𝒮:C(+;X)C(+;Y)\mathcal{S}:C\left(\mathbb{R}_{+};X\right)\rightarrow C\left(\mathbb{R}_{+};Y\right). Then, we recall that the operator 𝒮\mathcal{S} is called a history-dependent operator if the following property holds:

{ For each n there exists sn0 such that 𝒮u1(t)𝒮u2(t)Ysn0tu1(s)u2(s)X𝑑su1,u2C(+;X),t[0,n].\left\{\begin{array}[]{l}\text{ For each }n\in\mathbb{N}\text{ there exists }s_{n}\geq 0\text{ such that }\\ \left\|\mathcal{S}u_{1}(t)-\mathcal{S}u_{2}(t)\right\|_{Y}\leq s_{n}\int_{0}^{t}\left\|u_{1}(s)-u_{2}(s)\right\|_{X}ds\\ \quad\forall u_{1},u_{2}\in C\left(\mathbb{R}_{+};X\right),\forall t\in[0,n].\end{array}\right.

Note that in (2.1) and everywhere below the notation 𝒮η(t)\mathcal{S}\eta(t) represents the value of the function 𝒮η\mathcal{S}\eta at the point tt, i.e. 𝒮η(t)=(𝒮η)(t)\mathcal{S}\eta(t)=(\mathcal{S}\eta)(t). The notion of history-dependent operator was introduced in [30] and used in a number of papers, see for instance [31] and the references therein. Such kind of operators arise both in Functional Analysis, Theory of Partial Differential Equations and Solid Mechanics, as well. One of their main properties is given by the following fixed point result.

Theorem 2.1 Let ( X,XX,\|\cdot\|_{X} ) be a Banach space and let 𝒮:C(+;X)C(+;X)\mathcal{S}:C\left(\mathbb{R}_{+};X\right)\rightarrow C\left(\mathbb{R}_{+};X\right) be a history-dependent operator. Then the operator 𝒮\mathcal{S} has a unique fixed point ηC(+;X)\eta^{*}\in C\left(\mathbb{R}_{+};X\right).

Note that Theorem 2.1 represents a particular case of a more general result proved in [33]. Its proof is based on the fact that, if XX is a Banach space, then C(+;X)C\left(\mathbb{R}_{+};X\right) can be organized in a canonical way as a Fréchet space, i.e. as a complete metric space in which the corresponding topology is induced by a countable family of seminorms.

We turn now to an abstract result which represents a consequence of Theorem 2.1 and which will be used twice in Sect. 5 of this manuscript. Thus, we assume in what follows that (X,X)\left(X,\|\cdot\|_{X}\right) is a normed space, ( Y,YY,\|\cdot\|_{Y} ) is a Banach space and A:XYA:X\rightarrow Y and G:+×X×YYG:\mathbb{R}_{+}\times X\times Y\rightarrow Y are given operators, which satisfy the following conditions:

{ There exists LA>0 such that Ax1Ax2YLAx1x2Xx1,x2X.\displaystyle\left\{\begin{array}[]{l}\text{ There exists }L_{A}>0\text{ such that }\\ \left\|Ax_{1}-Ax_{2}\right\|_{Y}\leq L_{A}\left\|x_{1}-x_{2}\right\|_{X}\quad\forall x_{1},x_{2}\in X.\end{array}\right. (2.2)
{ (a) There exists LG>0 such that G(t,x1,y1)G(t,x2,y2)YLG(x1x2X+y1y2Y)x1,x2X,y1,y2Y,t+(b) The mapping tG(t,x,y) is measurable on +,for any xX,yY. (c) The mapping tG(t,0X,0Y) belongs to L(+).\displaystyle\left\{\begin{array}[]{l}\text{ (a) There exists }L_{G}>0\text{ such that }\\ \left\|G\left(t,x_{1},y_{1}\right)-G\left(t,x_{2},y_{2}\right)\right\|_{Y}\leq L_{G}\left(\left\|x_{1}-x_{2}\right\|_{X}+\left\|y_{1}-y_{2}\right\|_{Y}\right)\\ \forall x_{1},x_{2}\in X,y_{1},y_{2}\in Y,t\in\mathbb{R}_{+}\\ \text{(b) The mapping }t\mapsto G(t,x,y)\text{ is measurable on }\mathbb{R}_{+},\\ \text{for any }x\in X,y\in Y.\\ \text{ (c) The mapping }t\mapsto G\left(t,0_{X},0_{Y}\right)\text{ belongs to }L^{\infty}\left(\mathbb{R}_{+}\right).\end{array}\right.

Theorem 2.2 Let ( X,XX,\|\cdot\|_{X} ) be a normed space, ( Y,YY,\|\cdot\|_{Y} ) a Banach space and assume that (2.2)-(2.3) hold. Then, there exists an operator 𝒮:C(+;X)C(+;Y)\mathcal{S}:C\left(\mathbb{R}_{+};X\right)\rightarrow C\left(\mathbb{R}_{+};Y\right) such that for all functions xC(+;X)x\in C\left(\mathbb{R}_{+};X\right) and yC(+;Y)y\in C\left(\mathbb{R}_{+};Y\right), equality

y(t)=Ax(t)+0tG(s,x(s),y(s))𝑑st+y(t)=Ax(t)+\int_{0}^{t}G(s,x(s),y(s))ds\quad\forall t\in\mathbb{R}_{+} (2.4)

holds if and only if

y(t)=Ax(t)+𝒮x(t)t+.y(t)=Ax(t)+\mathcal{S}x(t)\quad\forall t\in\mathbb{R}_{+}. (2.5)

Moreover, the operator 𝒮:C(+;X)C(+;Y)\mathcal{S}:C\left(\mathbb{R}_{+};X\right)\rightarrow C\left(\mathbb{R}_{+};Y\right) is a history-dependent operator.
Proof Let xC(+;X)x\in C\left(\mathbb{R}_{+};X\right) and consider the operator Λ:C(+;Y)C(+;Y)\Lambda:C\left(\mathbb{R}_{+};Y\right)\rightarrow C\left(\mathbb{R}_{+};Y\right) defined by

Λτ(t)=0tG(s,x(s),Ax(s)+τ(s))𝑑s\Lambda\tau(t)=\int_{0}^{t}G(s,x(s),Ax(s)+\tau(s))ds (2.6)

for all τC(+;Y)\tau\in C\left(\mathbb{R}_{+};Y\right) and t+t\in\mathbb{R}_{+}. Note that, using the assumptions (2.2)-(2.3), it follows that the operator Λ\Lambda is well defined. Moreover, it depends on xx but, for simplicity, we do not indicate explicitly this dependence.

Let τ1,τ2C(+;Y)\tau_{1},\tau_{2}\in C\left(\mathbb{R}_{+};Y\right) and let t+t\in\mathbb{R}_{+}. Then, using definition (2.6) and assumption (2.3), we deduce that

Λτ1(t)Λτ2(t)Y\displaystyle\left\|\Lambda\tau_{1}(t)-\Lambda\tau_{2}(t)\right\|_{Y} 0tG(s,x(s),τ1(s)+Ax(s))G(s,x(s),τ2(s)+Ax(s))Y𝑑s\displaystyle\leq\int_{0}^{t}\left\|G\left(s,x(s),\tau_{1}(s)+Ax(s)\right)-G\left(s,x(s),\tau_{2}(s)+Ax(s)\right)\right\|_{Y}ds
LG0tτ1(s)τ2(s)Y𝑑s\displaystyle\leq L_{G}\int_{0}^{t}\left\|\tau_{1}(s)-\tau_{2}(s)\right\|_{Y}ds (2.7)

This inequality combined with Theorem 2.1 shows that the operator Λ\Lambda has a unique fixed point in C(+;Y)C\left(\mathbb{R}_{+};Y\right), denoted 𝒮x\mathcal{S}x. Moreover, combining (2.6) with equality Λ(𝒮x)=𝒮x\Lambda(\mathcal{S}x)=\mathcal{S}x we deduce that 𝒮x\mathcal{S}x is the unique element of the space C(+;Y)C\left(\mathbb{R}_{+};Y\right), which satisfies

𝒮x(t)=0tG(s,x(s),Ax(s)+𝒮x(s))𝑑st+\mathcal{S}x(t)=\int_{0}^{t}G(s,x(s),Ax(s)+\mathcal{S}x(s))ds\quad\forall t\in\mathbb{R}_{+} (2.8)

This implies the equivalence between equalities (2.4) and (2.5), for all functions xC(+;X)x\in C\left(\mathbb{R}_{+};X\right) and yC(+;Y)y\in C\left(\mathbb{R}_{+};Y\right).

To proceed, let x1,x2C(+;X),nx_{1},x_{2}\in C\left(\mathbb{R}_{+};X\right),n\in\mathbb{N} and t[0,n]t\in[0,n]. Then, using (2.8) and taking into account (2.2) and (2.3), we obtain that

𝒮x1(t)𝒮x2(t)Y\displaystyle\left\|\mathcal{S}x_{1}(t)-\mathcal{S}x_{2}(t)\right\|_{Y}
0tG(s,x1(s),Ax1(s)+𝒮x1(s))G(s,x2(s),Ax2(s)+𝒮x2(s))Y𝑑s\displaystyle\quad\leq\int_{0}^{t}\left\|G\left(s,x_{1}(s),Ax_{1}(s)+\mathcal{S}x_{1}(s)\right)-G\left(s,x_{2}(s),Ax_{2}(s)+\mathcal{S}x_{2}(s)\right)\right\|_{Y}ds
LG(LA+1)0tx1(s)x2(s)X𝑑s+LG0t𝒮x1(s)𝒮x2(s)Y𝑑s\displaystyle\quad\leq L_{G}\left(L_{A}+1\right)\int_{0}^{t}\left\|x_{1}(s)-x_{2}(s)\right\|_{X}ds+L_{G}\int_{0}^{t}\left\|\mathcal{S}x_{1}(s)-\mathcal{S}x_{2}(s)\right\|_{Y}ds

Using now the Gronwall argument we deduce that

𝒮x1(t)𝒮x2(t)YLG(LA+1)enLG0tx1(s)x2(s)X𝑑s\left\|\mathcal{S}x_{1}(t)-\mathcal{S}x_{2}(t)\right\|_{Y}\leq L_{G}\left(L_{A}+1\right)e^{nL_{G}}\int_{0}^{t}\left\|x_{1}(s)-x_{2}(s)\right\|_{X}ds (2.9)

This inequality shows that (2.1) holds with sn=LG(LA+1)enLGs_{n}=L_{G}\left(L_{A}+1\right)e^{nL_{G}}, which concludes the proof.

Note that Theorem 2.2 is important since it underlies the history-dependence feature of the solution of the implicit integral equation (2.4). It will be usefull in the study of viscoplastic constitutive laws, as explained in Sect. 5.

Next, we recall an existence and uniqueness result for mixed variational problems. To this end, let ( X,(,)X,XX,(\cdot,\cdot)_{X},\|\cdot\|_{X} ) and ( Y,(,)Y,YY,(\cdot,\cdot)_{Y},\|\cdot\|_{Y} ) be two real Hilbert spaces and we consider two operators A:XX,𝒮~:C(+;X)C(+;X)A:X\rightarrow X,\tilde{\mathcal{S}}:C\left(\mathbb{R}_{+};X\right)\rightarrow C\left(\mathbb{R}_{+};X\right), a bilinear form b:X×Yb:X\times Y\rightarrow\mathbb{R}, two functions f,h:+Xf,h:\mathbb{R}_{+}\rightarrow X and a set ΛY\Lambda\subset Y. We assume that the following conditions hold:

(a) There exists mA>0 such that\displaystyle\text{ (a) There exists }m_{A}>0\text{ such that } (2.10)
(Au1Au2,u1u2)XmAu1u2X2u1,u2X.\displaystyle\quad\left(Au_{1}-Au_{2},u_{1}-u_{2}\right)_{X}\geq m_{A}\left\|u_{1}-u_{2}\right\|_{X}^{2}\quad\forall u_{1},u_{2}\in X. (2.11)
(b) There exists LA>0 such that\displaystyle\text{ (b) There exists }L_{A}>0\text{ such that }
Au1Au2XLAu1u2Xu1,u2X.\displaystyle\quad\left\|Au_{1}-Au_{2}\right\|_{X}\leq L_{A}\left\|u_{1}-u_{2}\right\|_{X}\quad\forall u_{1},u_{2}\in X.
{ For each n there exists d~n0 and s~n0 such that 𝒮~u1(t)𝒮~u2(t)Xd~nu1(t)u2(t)X+s~n0tu1(s)u2(s)X𝑑su1,u2C(+;X),t[0,n].\displaystyle\left\{\begin{array}[]{l}\text{ For each }n\in\mathbb{N}\text{ there exists }\tilde{d}_{n}\geq 0\text{ and }\tilde{s}_{n}\geq 0\text{ such that }\\ \left\|\tilde{\mathcal{S}}u_{1}(t)-\tilde{\mathcal{S}}u_{2}(t)\right\|_{X}\leq\tilde{d}_{n}\left\|u_{1}(t)-u_{2}(t)\right\|_{X}+\tilde{s}_{n}\int_{0}^{t}\left\|u_{1}(s)-u_{2}(s)\right\|_{X}ds\\ \forall u_{1},u_{2}\in C\left(\mathbb{R}_{+};X\right),\forall t\in[0,n].\end{array}\right.
{b:X×Y is a bilinear form such that  (a) There exists Mb>0 such that |b(v;μ)|MbvXμYvX,μY. (b) There exists α>0 such that infμY,μ0YsupvX,v0Xb(v,μ)vXμYα.\displaystyle\left\{\begin{array}[]{l}b:X\times Y\rightarrow\mathbb{R}\text{ is a bilinear form such that }\\ \text{ (a) There exists }M_{b}>0\text{ such that }\\ \quad|b(v;\mu)|\leq M_{b}\|v\|_{X}\|\mu\|_{Y}\quad\forall v\in X,\mu\in Y.\\ \text{ (b) There exists }\alpha>0\text{ such that }\\ \quad\inf_{\mu\in Y,\mu\neq 0_{Y}}\sup_{v\in X,v\neq 0_{X}}\frac{b(v,\mu)}{\|v\|_{X}\|\mu\|_{Y}}\geq\alpha.\end{array}\right.
fC(+;X),hC(+;X).\displaystyle f\in C\left(\mathbb{R}_{+};X\right),\quad h\in C\left(\mathbb{R}_{+};X\right).

Λ\Lambda is a closed convex unbounded subset of YY that contains 0Y0_{Y}.

With these data we introduce the following evolutionary problem.
Problem 2.3 Find the functions u:+Xu:\mathbb{R}_{+}\rightarrow X and λ:+Λ\lambda:\mathbb{R}_{+}\rightarrow\Lambda such that

(Au(t),v)X+(𝒮~u(t),v)X+b(v,λ(t))=(f(t),v)XvX,\displaystyle(Au(t),v)_{X}+(\tilde{\mathcal{S}}u(t),v)_{X}+b(v,\lambda(t))=(f(t),v)_{X}\quad\forall v\in X, (2.15)
b(u(t),μλ(t))b(h(t),μλ(t))μΛ,\displaystyle b(u(t),\mu-\lambda(t))\leq b(h(t),\mu-\lambda(t))\quad\forall\mu\in\Lambda, (2.16)

for all t+t\in\mathbb{R}_{+}
The unique solvability of Problem 2.3 is provided in the next theorem.
Theorem 2.4 Assume (2.10)-(2.14). There exists d0>0d_{0}>0 which depends only on AA and bb such that, if d~n<d0\tilde{d}_{n}<d_{0} for all positive integers nn, then Problem 2.3 has a unique solution ( u,λu,\lambda ). Moreover, the solution satisfies uC(+;X)u\in C\left(\mathbb{R}_{+};X\right) and λC(+;Λ)\lambda\in C\left(\mathbb{R}_{+};\Lambda\right).

Theorem 2.4 was obtained in [32]. Its proof is based on results on generalized saddle point problems and various estimates, combined with a fixed point argument. The smalness assumption d~n<d0\tilde{d}_{n}<d_{0} in the statement of Theorem 2.4 is needed when using the fixed point argument which follows from the Banach contractions principle.

We end this section with further notation and preliminaries related to the contact model we are interested in. We denote by 𝕊d(d=1,2,3)\mathbb{S}^{d}(d=1,2,3) the space of second order symmetric tensors on d\mathbb{R}^{d} or, equivalently, the space of symmetric matrices of order dd. The inner product and norm on d\mathbb{R}^{d} and 𝕊d\mathbb{S}^{d} are defined by

𝒖𝒗=uivi,𝒗=(𝒗𝒗)12𝒖,𝒗d𝝈𝝉=σijτij,𝝉=(𝝉𝝉)12𝝈,𝝉𝕊d\begin{array}[]{ll}\boldsymbol{u}\cdot\boldsymbol{v}=u_{i}v_{i},&\|\boldsymbol{v}\|=(\boldsymbol{v}\cdot\boldsymbol{v})^{\frac{1}{2}}\quad\forall\boldsymbol{u},\boldsymbol{v}\in\mathbb{R}^{d}\\ \boldsymbol{\sigma}\cdot\boldsymbol{\tau}=\sigma_{ij}\tau_{ij},&\|\boldsymbol{\tau}\|=(\boldsymbol{\tau}\cdot\boldsymbol{\tau})^{\frac{1}{2}}\quad\forall\boldsymbol{\sigma},\boldsymbol{\tau}\in\mathbb{S}^{d}\end{array}

Also, we use the notation κ\|\kappa\| for the Euclidean norm of the element κm\kappa\in\mathbb{R}^{m}, where mm\in\mathbb{N}, and 𝟎\mathbf{0} for the zero element of the spaces d,𝕊d\mathbb{R}^{d},\mathbb{S}^{d} and m\mathbb{R}^{m}.

Let Ωd(d=1,2,3)\Omega\subset\mathbb{R}^{d}(d=1,2,3) be a bounded domain with Lipschitz continuous boundary Γ\Gamma and let Γ1,Γ2\Gamma_{1},\Gamma_{2} and Γ3\Gamma_{3} be three measurable parts of Γ\Gamma such that meas (Γ1)>0\left(\Gamma_{1}\right)>0. We use the notation 𝒙=(xi)\boldsymbol{x}=\left(x_{i}\right) for a typical point in ΩΓ\Omega\cup\Gamma and we denote by 𝒗=(vi)\boldsymbol{v}=\left(v_{i}\right) the outward unit normal at Γ\Gamma. Also, we use standard notation for the Lebesgue and Sobolev spaces associated to Ω\Omega and Γ\Gamma and, moreover, we consider the spaces

V={𝒗=(vi)H1(Ω)d:𝒗=𝟎 on Γ1}\displaystyle V=\left\{\boldsymbol{v}=\left(v_{i}\right)\in H^{1}(\Omega)^{d}:\boldsymbol{v}=\mathbf{0}\text{ on }\Gamma_{1}\right\}
Q={𝝉=(τij)L2(Ω)d×d:τij=τji}\displaystyle Q=\left\{\boldsymbol{\tau}=\left(\tau_{ij}\right)\in L^{2}(\Omega)^{d\times d}:\tau_{ij}=\tau_{ji}\right\}

These are real Hilbert spaces endowed with the inner products

(𝒖,𝒗)V=Ω𝜺(𝒖)𝜺(𝒗)𝑑x,(𝝈,𝝉)Q=Ω𝝈𝝉𝑑x(\boldsymbol{u},\boldsymbol{v})_{V}=\int_{\Omega}\boldsymbol{\varepsilon}(\boldsymbol{u})\cdot\boldsymbol{\varepsilon}(\boldsymbol{v})dx,\quad(\boldsymbol{\sigma},\boldsymbol{\tau})_{Q}=\int_{\Omega}\boldsymbol{\sigma}\cdot\boldsymbol{\tau}dx

and the associated norms V\|\cdot\|_{V} and Q\|\cdot\|_{Q}, respectively. Here 𝜺\boldsymbol{\varepsilon} represents the deformation operator given by

𝜺(𝒗)=(εij(𝒗)),εij(𝒗)=12(vi,j+vj,i)𝒗H1(Ω)d.\boldsymbol{\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)\forall\boldsymbol{v}\in H^{1}(\Omega)^{d}.

Completeness of the space ( V,VV,\|\cdot\|_{V} ) follows from the assumption meas (Γ1)>0\left(\Gamma_{1}\right)>0, which allows the use of Korn’s inequality.

For an element 𝒗V\boldsymbol{v}\in V we still write 𝒗\boldsymbol{v} for the trace of 𝒗\boldsymbol{v} on the boundary and we denote by vνv_{\nu} and 𝒗τ\boldsymbol{v}_{\tau} the normal and tangential components of 𝒗\boldsymbol{v} on Γ\Gamma, given by vν=𝒗𝒗,𝒗τ=𝒗vν𝒗v_{\nu}=\boldsymbol{v}\cdot\boldsymbol{v},\boldsymbol{v}_{\tau}=\boldsymbol{v}-v_{\nu}\boldsymbol{v}. Let Γ3\Gamma_{3} be a measurable part of Γ\Gamma. Then, by the Sobolev trace theorem, there exists a positive constant c0c_{0} which depends on Ω,Γ1\Omega,\Gamma_{1} and Γ3\Gamma_{3} such that

𝒗L2(Γ3)dc0𝒗V𝒗V.\|\boldsymbol{v}\|_{L^{2}\left(\Gamma_{3}\right)^{d}}\leq c_{0}\|\boldsymbol{v}\|_{V}\quad\forall\boldsymbol{v}\in V. (2.17)

As in [32] we consider the space

W={z=𝒗|Γ3:𝒗V},W=\left\{z=\left.\boldsymbol{v}\right|_{\Gamma_{3}}:\boldsymbol{v}\in V\right\},

where 𝒗|Γ3\left.\boldsymbol{v}\right|_{\Gamma_{3}} denotes the restriction of the trace of the element 𝒗V\boldsymbol{v}\in V to Γ3\Gamma_{3}. We recall that WH1/2(Γ3;d)W\subset H^{1/2}\left(\Gamma_{3};\mathbb{R}^{d}\right), where H1/2(Γ3;d)H^{1/2}\left(\Gamma_{3};\mathbb{R}^{d}\right) is the space of the restriction on Γ3\Gamma_{3} of traces on Γ\Gamma of functions of H1(Ω)dH^{1}(\Omega)^{d}. We denote by DD the dual of the space WW, and by ,Γ3\langle\cdot,\cdot\rangle_{\Gamma_{3}} the duality pairing between DD and WW. Nevertheless, for simplicity, we write 𝝁,𝒗Γ3\langle\boldsymbol{\mu},\boldsymbol{v}\rangle_{\Gamma_{3}} instead of 𝝁,𝒗|Γ3Γ3\left\langle\boldsymbol{\mu},\left.\boldsymbol{v}\right|_{\Gamma_{3}}\right\rangle_{\Gamma_{3}}, when 𝝁D\boldsymbol{\mu}\in D and 𝒗V\boldsymbol{v}\in V.

For a regular function 𝝈Q\boldsymbol{\sigma}\in Q we use the notation σν\sigma_{\nu} and 𝝈τ\boldsymbol{\sigma}_{\tau} for the normal and the tangential traces, i.e. σν=(𝝈𝒗)𝒗\sigma_{\nu}=(\boldsymbol{\sigma}\boldsymbol{v})\cdot\boldsymbol{v} and 𝝈τ=𝝈𝒗σν𝒗\boldsymbol{\sigma}_{\tau}=\boldsymbol{\sigma}\boldsymbol{v}-\sigma_{\nu}\boldsymbol{v}. Moreover, we recall that the divergence operator is defined by the equality Div𝝈=(σij,j)\operatorname{Div}\boldsymbol{\sigma}=\left(\sigma_{ij,j}\right) and, finally, the following Green’s formula holds:

Ω𝝈𝜺(𝒗)𝑑x+ΩDiv𝝈𝒗dx=Γ𝝈𝒗𝒗𝑑a𝒗V\int_{\Omega}\boldsymbol{\sigma}\cdot\boldsymbol{\varepsilon}(\boldsymbol{v})dx+\int_{\Omega}\operatorname{Div}\boldsymbol{\sigma}\cdot\boldsymbol{v}dx=\int_{\Gamma}\boldsymbol{\sigma}\boldsymbol{v}\cdot\boldsymbol{v}da\quad\forall\boldsymbol{v}\in V (2.18)

Finally, we denote by 𝐐\mathbf{Q}_{\infty} the space of fourth order tensor fields given by

𝐐={=(ijkl):ijkl=jikl=klijL(Ω),1i,j,k,ld}\mathbf{Q}_{\infty}=\left\{\mathcal{E}=\left(\mathcal{E}_{ijkl}\right):\mathcal{E}_{ijkl}=\mathcal{E}_{jikl}=\mathcal{E}_{klij}\in L^{\infty}(\Omega),1\leq i,j,k,l\leq d\right\}

and we recall that 𝐐\mathbf{Q}_{\infty} is a real Banach space with the norm

𝐐=max1i,j,k,ldijklL(Ω).\|\mathcal{E}\|_{\mathbf{Q}_{\infty}}=\max_{1\leq i,j,k,l\leq d}\left\|\mathcal{E}_{ijkl}\right\|_{L^{\infty}(\Omega)}.

Moreover, a simple calculation shows that

𝝉Qd𝐐𝝉Q𝐐,𝝉Q.\|\mathcal{E}\boldsymbol{\tau}\|_{Q}\leq d\|\mathcal{E}\|_{\mathbf{Q}_{\infty}}\|\boldsymbol{\tau}\|_{Q}\quad\forall\mathcal{E}\in\mathbf{Q}_{\infty},\boldsymbol{\tau}\in Q. (2.19)

This inequality will be used in several places, in Sect. 5.

3 Problem Statement

The physical setting is similar to that considered in [34] and can be resumed as follows. A viscoplastic body occupies a bounded domain Ωd\Omega\subset\mathbb{R}^{d} with a Lipschitz continuous boundary Γ\Gamma, divided into three measurable parts Γ1,Γ2\Gamma_{1},\Gamma_{2} and Γ3\Gamma_{3} such that meas (Γ1)>0\left(\Gamma_{1}\right)>0 and, in addition, Γ3\Gamma_{3} is plane. The body is subject to the action of body forces of density 𝒇0\boldsymbol{f}_{0}. It is fixed on Γ1\Gamma_{1} and time-dependent surfaces tractions of density 𝒇2\boldsymbol{f}_{2} act on Γ2\Gamma_{2}. On Γ3\Gamma_{3}, the body is in sliding frictional contact with a moving obstacle, the so-called foundation, which is made of a hard material covered by a layer of soft material of thickness gg. The friction implies the wear of the foundation that we model it with a surface variable, the wear function. Then, the classical formulation of the contact problem is the following.

Problem 𝒫\mathcal{P} Find a stress field 𝝈:Ω×+𝕊d\boldsymbol{\sigma}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{S}^{d}, a displacement field 𝒖:Ω×+d\boldsymbol{u}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{d}, an internal state variable κ:Ω×+m\kappa:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{m} and a wear function w:Γ3×+w:\Gamma_{3}\times\mathbb{R}_{+}\rightarrow\mathbb{R} such that

𝝈˙(t)=𝜺(𝒖˙(t))+𝒢(𝝈(t),𝜺(𝒖(t)),𝜿(t)) in Ω,\displaystyle\dot{\boldsymbol{\sigma}}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\dot{\boldsymbol{u}}(t))+\mathcal{G}(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{u}(t)),\boldsymbol{\kappa}(t))\text{ in }\Omega, (3.1)
𝜿˙(t)=𝑮(𝝈(t),𝜺(𝒖(t)),𝜿(t)) in Ω,\displaystyle\dot{\boldsymbol{\kappa}}(t)=\boldsymbol{G}(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{u}(t)),\boldsymbol{\kappa}(t))\text{ in }\Omega, (3.2)
Div𝝈(t)+𝒇0(t)=𝟎 in Ω,\displaystyle\operatorname{Div}\boldsymbol{\sigma}(t)+\boldsymbol{f}_{0}(t)=\mathbf{0}\text{ in }\Omega, (3.3)
𝒖(t)=𝟎 on Γ1,\displaystyle\boldsymbol{u}(t)=\mathbf{0}\text{ on }\Gamma_{1}, (3.4)
𝝈(t)𝒗=𝒇2(t) on Γ2,\displaystyle\boldsymbol{\sigma}(t)\boldsymbol{v}=\boldsymbol{f}_{2}(t)\text{ on }\Gamma_{2}, (3.5)
uv(t)g,σv(t)+p(uv(t)w(t))0,(uv(t)g)(σv(t)+p(uv(t)w(t))=0 on Γ3,𝝈τ(t)=ηp(uv(t)w(t))𝒏(t) on Γ3,w˙(t)=α(t)p(uv(t)w(t)) on Γ3,w(0)=0 in Γ3,𝝈(0)=𝝈0,𝒖(0)=𝒖0,𝜿(0)=𝜿0 in Ω.\displaystyle\begin{array}[]{ll}u_{v}(t)\leq g,&\sigma_{v}(t)+p\left(u_{v}(t)-w(t)\right)\leq 0,\\ \left(u_{v}(t)-g\right)\left(\sigma_{v}(t)+p\left(u_{v}(t)-w(t)\right)=0\right.&\text{ on }\Gamma_{3},\\ -\boldsymbol{\sigma}_{\tau}(t)=\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t)&\text{ on }\Gamma_{3},\\ \dot{w}(t)=\alpha(t)p\left(u_{v}(t)-w(t)\right)&\text{ on }\Gamma_{3},\\ w(0)=0&\text{ in }\Gamma_{3},\\ \boldsymbol{\sigma}(0)=\boldsymbol{\sigma}_{0},&\boldsymbol{u}(0)=\boldsymbol{u}_{0},\\ \boldsymbol{\kappa}(0)=\boldsymbol{\kappa}_{0}&\text{ in }\Omega.\end{array} (3.6)

We now provide a brief description of the equations and conditions in Problem 𝒫\mathcal{P}. Here and below, in order to simplify the notation, we do not indicate explicitly the dependence of various functions on the spatial variable 𝒙\boldsymbol{x}.

First, (3.1) and (3.2) represent the rate-type viscoplastic constitutive law with internal state variable in which we assume that elasticity tensor \mathcal{E} and the constitutive functions 𝒢\mathcal{G} and 𝑮\boldsymbol{G} satisfy the following conditions:

{ (a) =(ijkl):Ω×𝕊d𝕊d. (b) ijkl=klij=jiklL(Ω),1i,j,k,ld. (c) There exists m>0 such that 𝝉𝝉m𝝉2𝝉𝕊d, a.e. in Ω.\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.
{ (a) 𝒢:Ω×𝕊d×𝕊d×m𝕊d. (b) There exists L𝒢>0 such that 𝒢(𝒙,𝝈1,𝜺1,𝜿1)𝒢(𝒙,𝝈2,𝜺2,𝜿2)L𝒢(𝝈1𝝈2+𝜺1𝜺2+𝜿1𝜿2)𝝈1,𝝈2,𝜺1,𝜺2𝕊d,𝜿1,𝜿2m, a.e. 𝒙Ω. (c) The mapping 𝒙𝒢(𝒙,𝝈,𝜺,𝜿) is measurable on Ω, for any 𝝈,𝜺𝕊d and 𝜿m. (d) The mapping 𝒙𝒢(𝒙,𝟎,𝟎,𝟎) belongs to Q.\displaystyle\left\{\begin{array}[]{l}\text{ (a) }\mathcal{G}:\Omega\times\mathbb{S}^{d}\times\mathbb{S}^{d}\times\mathbb{R}^{m}\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},\boldsymbol{\kappa}_{1}\right)-\mathcal{G}\left(\boldsymbol{x},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{2},\boldsymbol{\kappa}_{2}\right)\right\|\\ \quad\leq L_{\mathcal{G}}\left(\left\|\boldsymbol{\sigma}_{1}-\boldsymbol{\sigma}_{2}\right\|+\left\|\boldsymbol{\varepsilon}_{1}-\boldsymbol{\varepsilon}_{2}\right\|+\left\|\boldsymbol{\kappa}_{1}-\boldsymbol{\kappa}_{2}\right\|\right)\\ \quad\forall\boldsymbol{\sigma}_{1},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{1},\boldsymbol{\varepsilon}_{2}\in\mathbb{S}^{d},\boldsymbol{\kappa}_{1},\boldsymbol{\kappa}_{2}\in\mathbb{R}^{m},\text{ a.e. }\boldsymbol{x}\in\Omega.\\ \text{ (c) The mapping }\boldsymbol{x}\mapsto\mathcal{G}(\boldsymbol{x},\boldsymbol{\sigma},\boldsymbol{\varepsilon},\boldsymbol{\kappa})\text{ is measurable on }\Omega,\\ \quad\text{ for any }\boldsymbol{\sigma},\boldsymbol{\varepsilon}\in\mathbb{S}^{d}\text{ and }\boldsymbol{\kappa}\in\mathbb{R}^{m}.\\ \text{ (d) The mapping }\boldsymbol{x}\mapsto\mathcal{G}(\boldsymbol{x},\mathbf{0},\mathbf{0},\mathbf{0})\text{ belongs to }Q.\end{array}\right.
{ (a) 𝑮:Ω×𝕊d×𝕊d×mm. (b) There exists LG>0 such that 𝑮(𝒙,𝝈1,𝜺1,𝜿1)𝑮(𝒙,𝝈2,𝜺2,𝜿2)LG(𝝈1𝝈2+𝜺1𝜺2+𝜿1𝜿2)𝝈1,𝝈2,𝜺1,𝜺2𝕊d,𝜿1,𝜿2m, a.e. 𝒙Ω. (c) The mapping 𝒙𝑮(𝒙,𝝈,𝜺,𝜿) is measurable on Ω, for any 𝝈,𝜺𝕊d and 𝜿m. (d) The mapping 𝒙𝑮(𝒙,𝟎,𝟎,𝟎) belongs to L2(Ω)m.\displaystyle\left\{\begin{array}[]{l}\text{ (a) }\boldsymbol{G}:\Omega\times\mathbb{S}^{d}\times\mathbb{S}^{d}\times\mathbb{R}^{m}\rightarrow\mathbb{R}^{m}.\\ \text{ (b) There exists }L_{G}>0\text{ such that }\\ \left\|\boldsymbol{G}\left(\boldsymbol{x},\boldsymbol{\sigma}_{1},\boldsymbol{\varepsilon}_{1},\boldsymbol{\kappa}_{1}\right)-\boldsymbol{G}\left(\boldsymbol{x},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{2},\boldsymbol{\kappa}_{2}\right)\right\|\\ \quad\leq L_{G}\left(\left\|\boldsymbol{\sigma}_{1}-\boldsymbol{\sigma}_{2}\right\|+\left\|\boldsymbol{\varepsilon}_{1}-\boldsymbol{\varepsilon}_{2}\right\|+\left\|\boldsymbol{\kappa}_{1}-\boldsymbol{\kappa}_{2}\right\|\right)\\ \quad\forall\boldsymbol{\sigma}_{1},\boldsymbol{\sigma}_{2},\boldsymbol{\varepsilon}_{1},\boldsymbol{\varepsilon}_{2}\in\mathbb{S}^{d},\boldsymbol{\kappa}_{1},\boldsymbol{\kappa}_{2}\in\mathbb{R}^{m},\text{ a.e. }\boldsymbol{x}\in\Omega.\\ \text{ (c) The mapping }\boldsymbol{x}\mapsto\boldsymbol{G}(\boldsymbol{x},\boldsymbol{\sigma},\boldsymbol{\varepsilon},\boldsymbol{\kappa})\text{ is measurable on }\Omega,\\ \text{ for any }\boldsymbol{\sigma},\boldsymbol{\varepsilon}\in\mathbb{S}^{d}\text{ and }\boldsymbol{\kappa}\in\mathbb{R}^{m}.\\ \text{ (d) The mapping }\boldsymbol{x}\mapsto\boldsymbol{G}(\boldsymbol{x},\mathbf{0},\mathbf{0},\mathbf{0})\text{ belongs to }L^{2}(\Omega)^{m}.\end{array}\right.

Constitutive equations of the form (3.1)-(3.2) could describe both elasticity, plasticity, creep, relaxation, hardening and softening phenomena. For this reason they have been considered in the literature in order to model the behavior of real materials like rubbers, metals, pastes, rocks and so on. Various results and mechanical interpretation concerning constitutive laws of this form may be found in [5] and [17], for instance. Here we restrict ourselves to provide three clasical examples of such equations, with our without internal state variables.

The first example is one-dimensional and does not involve internal state variable. It is of the form

σ˙=Eε˙+𝒢(σ,ε)\dot{\sigma}=E\dot{\varepsilon}+\mathcal{G}(\sigma,\varepsilon) (3.14)

with

𝒢(σ,ε)={k1F1(σf(ε)) if σ>f(ε)0 if g(ε)σf(ε)k2F2(g(ε)σ) if σ<g(ε)\mathcal{G}(\sigma,\varepsilon)=\begin{cases}-k_{1}F_{1}(\sigma-f(\varepsilon))&\text{ if }\sigma>f(\varepsilon)\\ 0&\text{ if }g(\varepsilon)\leq\sigma\leq f(\varepsilon)\\ k_{2}F_{2}(g(\varepsilon)-\sigma)&\text{ if }\sigma<g(\varepsilon)\end{cases}

Here where E>0E>0 is the Young modulus, k1,k2>0k_{1},k_{2}>0 are viscosity constants, ff and gg are Lipschitz continuous functions with g(ε)<f(ε)g(\varepsilon)<f(\varepsilon), and F1,F2:+F_{1},F_{2}:\mathbb{R}_{+}\rightarrow\mathbb{R} are increasing functions with F1(0)=F2(0)=0F_{1}(0)=F_{2}(0)=0. Note that the domain of elastic behavior of the material is characterized by the inequalities g(ε)σf(ε)g(\varepsilon)\leq\sigma\leq f(\varepsilon). Plastic deformations occur only for σ>f(ε)\sigma>f(\varepsilon) in extension or for σ<g(ε)\sigma<g(\varepsilon) in compression. Therefore, since the yield limit (in extension and in compression) depends on the deformation, we conclude that the model (3.14), (3.15) represents a model with hardening.

A second example of an elastic-viscoplastic constitutive law without internal state variable is Perzyna’s law given by

𝜺˙=1𝝈˙+1δ(σ𝒫𝒦𝝈)\dot{\boldsymbol{\varepsilon}}=\mathcal{E}^{-1}\dot{\boldsymbol{\sigma}}+\frac{1}{\delta}\left(\sigma-\mathcal{P}_{\mathcal{K}}\boldsymbol{\sigma}\right) (3.16)

Here \mathcal{E} is a fourth order tensor satisfying (3.11), 1\mathcal{E}^{-1} denotes its inverse, δ>0\delta>0 is a viscosity constant, 𝒦\mathcal{K} is a nonempty, closed, convex set in the space 𝕊d\mathbb{S}^{d} of symmetric tensors and 𝒫𝒦\mathcal{P}_{\mathcal{K}} represents the projection operator. Notice that in this case the function 𝒢\mathcal{G} does not depend on 𝜺\boldsymbol{\varepsilon} and is given by

𝒢(𝝈,𝜺)=1δ(𝝈𝒫𝒦𝝈)\mathcal{G}(\boldsymbol{\sigma},\boldsymbol{\varepsilon})=-\frac{1}{\delta}\mathcal{E}\left(\boldsymbol{\sigma}-\mathcal{P}_{\mathcal{K}}\boldsymbol{\sigma}\right)

Since σ=𝒫𝒦σ\sigma=\mathcal{P}_{\mathcal{K}}\sigma iff σ𝒦\sigma\in\mathcal{K}, from (3.16) we see that viscoplastic deformations occur only for the stress tensors σ\sigma outside the set 𝒦\mathcal{K}. Thus, the set 𝒦\mathcal{K} represents the domain of elastic behavior of the material. It is usually defined by

𝒦={σ𝕊d:(σ)0}\mathcal{K}=\left\{\sigma\in\mathbb{S}^{d}:\mathcal{F}(\sigma)\leq 0\right\} (3.17)

where :𝕊d\mathcal{F}:\mathbb{S}^{d}\rightarrow\mathbb{R} is a convex function such that (𝟎)<0\mathcal{F}(\mathbf{0})<0. The function \mathcal{F} is called the yield function and the equation (𝝈)=0\mathcal{F}(\boldsymbol{\sigma})=0 represents the yield condition.

A concrete example of an elastic-viscoplastic constitutive law of the form (3.1), (3.2) is given by the Perzyna’s law with internal state variable,

𝜺˙=1𝝈˙+1δ(𝝈𝒫𝒦(κ)𝝈)\displaystyle\dot{\boldsymbol{\varepsilon}}=\mathcal{E}^{-1}\dot{\boldsymbol{\sigma}}+\frac{1}{\delta}\left(\boldsymbol{\sigma}-\mathcal{P}_{\mathcal{K}(\kappa)}\boldsymbol{\sigma}\right) (3.18)
𝜿˙(t)=23δ𝝈P𝒦(κ)𝝈\displaystyle\dot{\boldsymbol{\kappa}}(t)=\frac{2}{3\delta}\left\|\boldsymbol{\sigma}-P_{\mathcal{K}(\kappa)}\boldsymbol{\sigma}\right\| (3.19)

Here 𝒫𝒦(𝜿)\mathcal{P}_{\mathcal{K}(\boldsymbol{\kappa})} represents the projection mapping on the von Mises convex set 𝒦(𝜿)\mathcal{K}(\boldsymbol{\kappa}) defined by equality

𝒦(𝜿)={𝝈𝕊d:𝝈Dω(𝜿)2}\mathcal{K}(\boldsymbol{\kappa})=\left\{\boldsymbol{\sigma}\in\mathbb{S}^{d}:\left\|\boldsymbol{\sigma}^{D}\right\|\leq\omega(\boldsymbol{\kappa})\sqrt{2}\right\}

𝝈D\boldsymbol{\sigma}^{D} being the deviator of 𝝈\boldsymbol{\sigma}, and ω:\omega:\mathbb{R}\rightarrow\mathbb{R} is a given positive function. Note that, as explained in [13], the variable κ\kappa given by (3.19) represents the irreversible equivalent strain.

Equation (3.3) is the equilibrium equation and we use it here since we assume that the process is quasistatic. Conditions (3.4) and (3.5) are the displacement boundary condition and traction boundary condition, respectively. We assume that the densities of body forces and surface tractions are such that

f0C(+;L2(Ω)d),f2C(+;L2(Γ2)d)f_{0}\in C\left(\mathbb{R}_{+};L^{2}(\Omega)^{d}\right),\quad f_{2}\in C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{2}\right)^{d}\right) (3.20)

Conditions (3.6)-(3.8) were introduced and justified in [34] and, for this reason, we do not present here in detail. We restrict ourselves to mention that (3.6) represents the contact condition in which the normal compliance function pp satisfies
(a) p:Γ3×+p:\Gamma_{3}\times\mathbb{R}\rightarrow\mathbb{R}_{+}.
(b) There exists Lp>0L_{p}>0 such that

|p(𝒙,r1)p(𝒙,r2)|Lp|r1r2|\displaystyle\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|
r1,r2, a.e. 𝒙Γ3\displaystyle\forall r_{1},r_{2}\in\mathbb{R},\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}
(p(𝒙,r1)p(𝒙,r2))(r1r2)0\displaystyle\left(p\left(\boldsymbol{x},r_{1}\right)-p\left(\boldsymbol{x},r_{2}\right)\right)\left(r_{1}-r_{2}\right)\geq 0 (3.21)
r1,r2, a.e. 𝒙Γ3\displaystyle\forall r_{1},r_{2}\in\mathbb{R},\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}

(d) The mapping 𝒙p(𝒙,r)\boldsymbol{x}\mapsto p(\boldsymbol{x},r) is measurable on Γ3\Gamma_{3}, for any rr\in\mathbb{R}.
(e) p(𝒙,r)=0p(\boldsymbol{x},r)=0 for all r0r\leq 0, a.e. 𝒙Γ3\boldsymbol{x}\in\Gamma_{3}.

This condition was derived by assuming an additive decomposition of the normal stress into two components which satisfy the Signorini condition in the form with a gap function and the normal compliance contact condition with wear, respectively.

Condition (3.7) represents a sliding version of the classical Coulomb law of dry friction. Here η\eta represents the friction coefficient, 𝒏\boldsymbol{n}^{*} is the unitary vector defined by

𝒏(t)=𝒗(t)𝒗(t)\boldsymbol{n}^{*}(t)=-\frac{\boldsymbol{v}^{*}(t)}{\left\|\boldsymbol{v}^{*}(t)\right\|}

where 𝒗\boldsymbol{v}^{*} is the velocity of the foundation, supposed to be a non vanishing time-dependent function in the plane of Γ3\Gamma_{3}. This condition was derived under the assumption that the velocity of the foundation 𝒗(t)\boldsymbol{v}^{*}(t) is large in comparison with the tangential velocity 𝒖˙τ(t)\dot{\boldsymbol{u}}_{\tau}(t). Here, we assume that the coefficient of friction and velocity of the foundation verify two following conditions:

ηL(Γ3),η(𝒙)0 a.e. 𝒙Γ3,\displaystyle\quad\eta\in L^{\infty}\left(\Gamma_{3}\right),\quad\eta(\boldsymbol{x})\geq 0\quad\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}, (3.22)
{𝒗C(+;3) and there exist v1,v2>0 such that v1𝒗(t)v2t+.\displaystyle\left\{\begin{array}[]{l}\boldsymbol{v}^{*}\in C\left(\mathbb{R}_{+};\mathbb{R}^{3}\right)\text{ and there exist }v_{1},v_{2}>0\text{ such that }\\ v_{1}\leq\left\|\boldsymbol{v}^{*}(t)\right\|\leq v_{2}\quad\forall t\in\mathbb{R}_{+}.\end{array}\right. (3.23)

The differential equation (3.8) represents a version of Archard’s law which governs the evolution of the wear function and, again, it was derived under the assumption that the velocity of the foundation 𝒗(t)\boldsymbol{v}^{*}(t) is large in comparison with the tangential velocity 𝒖˙τ(t)\dot{\boldsymbol{u}}_{\tau}(t). Here

α(t)=k𝒗(t),\alpha(t)=k\left\|\boldsymbol{v}^{*}(t)\right\|,

kk being the wear coefficient, assumed to be such that

kL(Γ3),k(𝒙)0 a.e. 𝒙Γ3.k\in L^{\infty}\left(\Gamma_{3}\right),\quad k(\boldsymbol{x})\geq 0\quad\text{ a.e. }\boldsymbol{x}\in\Gamma_{3}. (3.24)

Condition (3.9) represents the initial condition for the wear function and shows that at the initial moment the materials involved in the process are new. Next, (3.10) represent the initial conditions for the rest of the unknowns in which 𝒖0,𝝈0,𝜿0\boldsymbol{u}_{0},\boldsymbol{\sigma}_{0},\boldsymbol{\kappa}_{0} denote the initial displacement, the initial stress field and the initial state variable, respectively. We assume in what follows that these initial data have the regularity

𝒖0V,𝝈0Q,𝜿0L2(Ω)m.\boldsymbol{u}_{0}\in V,\quad\boldsymbol{\sigma}_{0}\in Q,\quad\boldsymbol{\kappa}_{0}\in L^{2}(\Omega)^{m}. (3.25)

Finally, we assume that

 there exists 𝜽~V such that θ~v=1 a.e. on Γ3\text{ there exists }\tilde{\boldsymbol{\theta}}\in V\text{ such that }\tilde{\theta}_{v}=1\text{ a.e. on }\Gamma_{3} (3.26)

where, we recall, θ~v=𝜽~𝒗\tilde{\theta}_{v}=\tilde{\boldsymbol{\theta}}\cdot\boldsymbol{v}. This assumption concerns only the geometry of the problem and was already used in [2], for instance. It is needed in order to derive a mixed variational formulation to Problem 𝒫\mathcal{P}.

4 A Mixed Variational Formulation

We now derive a mixed variational formulation of Problem 𝒫\mathcal{P}. To this end, we define the sets KVK\subset V and ΛD\Lambda\subset D, the bilinear form b:V×Db:V\times D\rightarrow\mathbb{R} and the function f:+Vf:\mathbb{R}_{+}\rightarrow V by
equalities

K={𝒗V:vv0 a.e. on Γ3}\displaystyle K=\left\{\boldsymbol{v}\in V:v_{v}\leq 0\text{ a.e. on }\Gamma_{3}\right\} (4.1)
Λ={𝝁D:𝝁,𝒗Γ30𝒗K}\displaystyle\Lambda=\left\{\boldsymbol{\mu}\in D:\langle\boldsymbol{\mu},\boldsymbol{v}\rangle_{\Gamma_{3}}\leq 0\forall\boldsymbol{v}\in K\right\} (4.2)
b(𝒗,𝝁)=𝝁,𝒗Γ3,𝒗V,𝝁D\displaystyle b(\boldsymbol{v},\boldsymbol{\mu})=\langle\boldsymbol{\mu},\boldsymbol{v}\rangle_{\Gamma_{3}},\quad\forall\boldsymbol{v}\in V,\boldsymbol{\mu}\in D (4.3)
(𝒇(t),𝒗)V=Ω𝒇0(t)𝒗𝑑x+Γ2𝒇2(t)𝒗𝑑a𝒗V,t+\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}_{+} (4.4)

Next, we assume that 𝝈,𝒖,𝜿\boldsymbol{\sigma},\boldsymbol{u},\boldsymbol{\kappa} and ww are regular functions which verify (3.1)-(3.10). Let t+,𝒗Vt\in\mathbb{R}_{+},\boldsymbol{v}\in V and 𝝁Λ\boldsymbol{\mu}\in\Lambda. We integrate (3.1), (3.2) with initial conditions (3.10) to find that

𝝈(t)=𝜺(𝒖(t))+0t𝒢(𝝈(s),𝜺(𝒖(s)),𝜿(s))𝑑s+𝝈0𝑬𝜺(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)),\boldsymbol{\kappa}(s))ds+\boldsymbol{\sigma}_{0}-\boldsymbol{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right) (4.5)
𝜿(t)=0t𝑮(𝝈(s),𝜺(𝒖(s)),𝜿(s))𝑑s+𝜿0\displaystyle\boldsymbol{\kappa}(t)=\int_{0}^{t}\boldsymbol{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)),\boldsymbol{\kappa}(s))ds+\boldsymbol{\kappa}_{0} (4.6)

Moreover, we integrate (3.8) with the initial condition (3.9) to obtain

w(t)=0tα(s)p(uv(s)w(s))𝑑sw(t)=\int_{0}^{t}\alpha(s)p\left(u_{v}(s)-w(s)\right)ds (4.7)

Next, we use Green formula (2.18) and the equilibrium equation (3.3) to see that

(𝝈(t),𝜺(𝒗))Q=(𝒇0(t),𝒗)L2(Ω)d+Γ𝝈(t)𝒗𝒗𝑑a𝒗V(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}=\left(\boldsymbol{f}_{0}(t),\boldsymbol{v}\right)_{L^{2}(\Omega)^{d}}+\int_{\Gamma}\boldsymbol{\sigma}(t)\boldsymbol{v}\cdot\boldsymbol{v}da\quad\forall\boldsymbol{v}\in V (4.8)

We split the surface integral over Γ1,Γ2\Gamma_{1},\Gamma_{2} and Γ3\Gamma_{3}. Then we use the equalities 𝒗=𝟎\boldsymbol{v}=\mathbf{0} on Γ1\Gamma_{1}, 𝝈(t)𝒗=𝒇2(t)\boldsymbol{\sigma}(t)\boldsymbol{v}=\boldsymbol{f}_{2}(t) on Γ2,𝝈(t)𝒗𝒗=σv(t)vv+𝝈τ(t)𝒗τ\Gamma_{2},\boldsymbol{\sigma}(t)\boldsymbol{v}\cdot\boldsymbol{v}=\sigma_{v}(t)v_{v}+\boldsymbol{\sigma}_{\tau}(t)\cdot\boldsymbol{v}_{\tau} on Γ3\Gamma_{3}, and definition (4.4) to obtain that

(𝝈(t),𝜺(𝒗))Q=(𝒇(t),𝒗)V+Γ3(σv(t)vv+𝝈τ(t)𝒗τ)𝑑a𝒗V(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}=(\boldsymbol{f}(t),\boldsymbol{v})_{V}+\int_{\Gamma_{3}}\left(\sigma_{v}(t)v_{v}+\boldsymbol{\sigma}_{\tau}(t)\cdot\boldsymbol{v}_{\tau}\right)da\quad\forall\boldsymbol{v}\in V (4.9)

Let λ(t)D\lambda(t)\in D be the Lagrange multiplier defined by

λ(t),zΓ3=Γ3(σν(t)+p(uν(t)w(t)))zν𝑑azW\langle\lambda(t),z\rangle_{\Gamma_{3}}=-\int_{\Gamma_{3}}\left(\sigma_{\nu}(t)+p\left(u_{\nu}(t)-w(t)\right)\right)z_{\nu}da\quad\forall z\in W (4.10)

Then, taking into account (4.3) we can write

Γ3σν(t)vν𝑑a=b(𝒗,𝝀(t))Γ3p(uν(t)w(t))vν𝑑a𝒗V\int_{\Gamma_{3}}\sigma_{\nu}(t)v_{\nu}da=-b(\boldsymbol{v},\boldsymbol{\lambda}(t))-\int_{\Gamma_{3}}p\left(u_{\nu}(t)-w(t)\right)v_{\nu}da\quad\forall\boldsymbol{v}\in V (4.11)

and, combining this equality with (4.9) and (3.7) we obtain that

(𝝈(t),𝜺(𝒗))Q+b(𝒗,𝝀(t))+Γ3p(uv(t)w(t))vv𝑑a\displaystyle(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}+b(\boldsymbol{v},\boldsymbol{\lambda}(t))+\int_{\Gamma_{3}}p\left(u_{v}(t)-w(t)\right)v_{v}da
+Γ3ηp(uv(t)w(t))𝒏(t)𝒗τ𝑑a=(𝒇(t),𝒗)V𝒗V\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da=(\boldsymbol{f}(t),\boldsymbol{v})_{V}\quad\forall\boldsymbol{v}\in V (4.12)

On the other hand, (4.10), (3.6), (4.1) and (4.2) imply that λ(t)Λ\lambda(t)\in\Lambda. Moreover, using (3.26) and definition (4.3) we deduce that

b(𝒖(t),𝝁λ(t))\displaystyle b(\boldsymbol{u}(t),\boldsymbol{\mu}-\lambda(t)) =b(𝒖(t)g𝜽~,𝝁λ(t))+b(g𝜽~,𝝁λ(t))\displaystyle=b(\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\lambda(t))+b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\lambda(t))
=𝝁,𝒖(t)g𝜽~Γ3λ(t),𝒖(t)g𝜽~Γ3+b(g𝜽~,𝝁λ(t))𝝁Λ\displaystyle=\langle\boldsymbol{\mu},\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}}\rangle_{\Gamma_{3}}-\langle\lambda(t),\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}}\rangle_{\Gamma_{3}}+b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\lambda(t))\quad\forall\boldsymbol{\mu}\in\Lambda (4.13)

In addition, the contact condition (3.6), assumption (3.26) and definitions (4.1), (4.2), (4.10) imply that

𝒖(t)g𝜽~K,𝝁,𝒖(t)g𝜽~Γ30,λ(t),𝒖(t)g𝜽~Γ3=0𝝁Λ.\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}}\in K,\quad\langle\boldsymbol{\mu},\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}}\rangle_{\Gamma_{3}}\leq 0,\quad\langle\lambda(t),\boldsymbol{u}(t)-g\tilde{\boldsymbol{\theta}}\rangle_{\Gamma_{3}}=0\quad\forall\boldsymbol{\mu}\in\Lambda. (4.14)

We combine now (4.13) and (4.14) to deduce that

b(𝒖(t),𝝁λ(t))b(g𝜽~,𝝁λ(t))𝝁Λ.b(\boldsymbol{u}(t),\boldsymbol{\mu}-\lambda(t))\leq b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\lambda(t))\quad\forall\boldsymbol{\mu}\in\Lambda. (4.15)

Finally, we gather equalities (4.5)-(4.7), (4.12), and inequality (4.15) to obtain the following mixed variational formulation of Problem 𝒫\mathcal{P}.

Problem 𝒫V\mathcal{P}^{V} Find a stress field 𝝈:+Q\boldsymbol{\sigma}:\mathbb{R}_{+}\rightarrow Q, a displacement field 𝒖:+V\boldsymbol{u}:\mathbb{R}_{+}\rightarrow V, an internal state variable 𝜿:+L2(Ω)m\boldsymbol{\kappa}:\mathbb{R}_{+}\rightarrow L^{2}(\Omega)^{m}, a wear function w:+L2(Γ3)w:\mathbb{R}_{+}\rightarrow L^{2}\left(\Gamma_{3}\right) and a Lagrange multiplier λ:+Λ\lambda:\mathbb{R}_{+}\rightarrow\Lambda such that

𝝈(t)=𝜺(𝒖(t))+0t𝒢(𝝈(s),𝜺(𝒖(s)),𝜿(s))𝑑s+𝝈0𝑬𝜺(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)),\boldsymbol{\kappa}(s))ds+\boldsymbol{\sigma}_{0}-\boldsymbol{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right) (4.16)
𝜿(t)=0t𝑮(𝝈(s),𝜺(𝒖(s)),𝜿(s))𝑑s+𝜿0\displaystyle\boldsymbol{\kappa}(t)=\int_{0}^{t}\boldsymbol{G}(\boldsymbol{\sigma}(s),\boldsymbol{\varepsilon}(\boldsymbol{u}(s)),\boldsymbol{\kappa}(s))ds+\boldsymbol{\kappa}_{0} (4.17)
w(t)=0tα(s)p(uv(s)w(s))𝑑s\displaystyle w(t)=\int_{0}^{t}\alpha(s)p\left(u_{v}(s)-w(s)\right)ds (4.18)
(𝝈(t),𝜺(𝒗))Q+Γ3p(uv(t)w(t))vv𝑑a+b(𝒗,𝝀(t))\displaystyle(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}+\int_{\Gamma_{3}}p\left(u_{v}(t)-w(t)\right)v_{v}da+b(\boldsymbol{v},\boldsymbol{\lambda}(t))
+Γ3ηp(uv(t)w(t))𝒏(t)𝒗τ𝑑a=(𝒇(t),𝒗)V𝒗V\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da=(\boldsymbol{f}(t),\boldsymbol{v})_{V}\quad\forall\boldsymbol{v}\in V (4.19)
b(𝒖(t),𝝁𝝀(t))b(g𝜽~,𝝁𝝀(t))𝝁Λ\displaystyle b(\boldsymbol{u}(t),\boldsymbol{\mu}-\boldsymbol{\lambda}(t))\leq b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\boldsymbol{\lambda}(t))\quad\forall\boldsymbol{\mu}\in\Lambda (4.20)

for all t+t\in\mathbb{R}_{+}.
Note that Problem 𝒫V\mathcal{P}^{V} represents a system which couples three nonlinear implicit integral equations for the stress field, the internal state variable and the wear function, respectively, with a history-dependent variational equation for displacement field, and a first-order timedependent variational inequality for the Lagrange multiplier.

5 Existence and Uniqueness

In this section we state and prove the following existence and uniqueness result concerning problem 𝒫V\mathcal{P}^{V}.

Theorem 5.1 Assume (3.11)-(3.13), (3.20)-(3.26). Then, there exists e0>0e_{0}>0 which depends only on ,Ω,Γ1\mathcal{E},\Omega,\Gamma_{1} and Γ3\Gamma_{3} such that Problem 𝒫V\mathcal{P}^{V} has a unique solution ( 𝝈,𝒖,𝜿,w,𝝀\boldsymbol{\sigma},\boldsymbol{u},\boldsymbol{\kappa},w,\boldsymbol{\lambda} ), if Lp(1+ηL(Γ3))<e0L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0}. Moreover, the solution satisfies

(𝝈,𝒖,𝜿,w,𝝀)C(+;Q×V×L2(Ω)m×L2(Γ3)×Λ)(\boldsymbol{\sigma},\boldsymbol{u},\boldsymbol{\kappa},w,\boldsymbol{\lambda})\in C\left(\mathbb{R}_{+};Q\times V\times L^{2}(\Omega)^{m}\times L^{2}\left(\Gamma_{3}\right)\times\Lambda\right) (5.1)

The proof of Theorem 5.1 will be carried out in several steps, based on the abstract results presented in Sect. 2. To present it, we assume in what follows that (3.11)-(3.13), (3.20)-(3.26) hold. The first step of the proof is the following.

Lemma 5.2 There exists an operator 𝒮=(𝒮1,𝒮2):C(+;V)C(+;Q×L2(Ω)m)\mathcal{S}=\left(\mathcal{S}_{1},\mathcal{S}_{2}\right):C\left(\mathbb{R}_{+};V\right)\rightarrow C\left(\mathbb{R}_{+};Q\times L^{2}(\Omega)^{m}\right) such that for all functions 𝒖C(+;V)\boldsymbol{u}\in C\left(\mathbb{R}_{+};V\right) and (𝝈,𝜿)C(+;Q×L2(Ω)m)(\boldsymbol{\sigma},\boldsymbol{\kappa})\in C\left(\mathbb{R}_{+};Q\times L^{2}(\Omega)^{m}\right), equalities (4.16), (4.17) hold for all t+t\in\mathbb{R}_{+}if and only if

𝝈(t)\displaystyle\boldsymbol{\sigma}(t) =𝜺(𝒖(t))+𝒮1𝒖(t)\displaystyle=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t))+\mathcal{S}_{1}\boldsymbol{u}(t) (5.2)
𝜿(t)\displaystyle\boldsymbol{\kappa}(t) =𝒮2𝒖(t)\displaystyle=\mathcal{S}_{2}\boldsymbol{u}(t) (5.3)

for all t+t\in\mathbb{R}_{+}. Moreover, the operator 𝒮:C(+;V)C(+;Q×L2(Ω)m)\mathcal{S}:C\left(\mathbb{R}_{+};V\right)\rightarrow C\left(\mathbb{R}_{+};Q\times L^{2}(\Omega)^{m}\right) is a historydependent operator.

Proof Lemma 5.2 is a direct consequence of Theorem 2.2 applied with X=V,Y=Q×L2(Ω)mX=V,Y=Q\times L^{2}(\Omega)^{m},

A𝒖=(𝜺(𝒖)+𝝈0𝜺(𝒖0),𝜿0)\displaystyle A\boldsymbol{u}=\left(\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u})+\boldsymbol{\sigma}_{0}-\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right),\boldsymbol{\kappa}_{0}\right)
G(t,𝒖,(𝝈,𝜿))=(𝒢(𝝈,𝜺(𝒖),𝜿),𝑮(𝝈,𝜺(𝒖),𝜿))\displaystyle G(t,\boldsymbol{u},(\boldsymbol{\sigma},\boldsymbol{\kappa}))=(\mathcal{G}(\boldsymbol{\sigma},\boldsymbol{\varepsilon}(\boldsymbol{u}),\boldsymbol{\kappa}),\boldsymbol{G}(\boldsymbol{\sigma},\boldsymbol{\varepsilon}(\boldsymbol{u}),\boldsymbol{\kappa}))

for all 𝒖V,(𝝈,𝜿)Q×L2(Ω)m\boldsymbol{u}\in V,(\boldsymbol{\sigma},\boldsymbol{\kappa})\in Q\times L^{2}(\Omega)^{m} and t+t\in\mathbb{R}_{+}. Indeed, it is easy to see that assumptions (3.11)-(3.13) and (3.25) imply that the operators above are well defined and, moreover, they satisfy conditions (2.2) and (2.3), respectively.

The next step consists in the following result concerning the wear function.

Lemma 5.3 There exists an operator :C(+;V)C(+;L2(Γ3))\mathcal{R}:C\left(\mathbb{R}_{+};V\right)\rightarrow C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right) such that for all functions 𝒖C(+;V)\boldsymbol{u}\in C\left(\mathbb{R}_{+};V\right) and wC(+;L2(Γ3))w\in C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right), equality (4.18) holds for all t+t\in\mathbb{R}_{+}if and only if

w(t)=𝒖(t)w(t)=\mathcal{R}\boldsymbol{u}(t) (5.4)

for all t+t\in\mathbb{R}_{+}. Moreover, the operator :C(+;V)C(+;L2(Γ3))\mathcal{R}:C\left(\mathbb{R}_{+};V\right)\rightarrow C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right) is a historydependent operator.

Proof Lemma 5.3 is a direct consequence of Theorem 2.2 applied with X=V,Y=L2(Γ3)X=V,Y=L^{2}\left(\Gamma_{3}\right),

A𝒖=0L2(Γ3),G(t,𝒖,w)=α(t)p(uvw)A\boldsymbol{u}=0_{L^{2}\left(\Gamma_{3}\right)},\quad G(t,\boldsymbol{u},w)=\alpha(t)p\left(u_{v}-w\right)

for all 𝒖V,wL2(Γ3)\boldsymbol{u}\in V,w\in L^{2}\left(\Gamma_{3}\right) and t+t\in\mathbb{R}_{+}. Indeed, it is easy to see that assumptions (3.21)(3.24) imply that the operators above are well defined and, moreover, they satisfy conditions (2.2) and (2.3), respectively.

We now complete the statement of Lemmas 5.2 and 5.3 with some estimates concerning the constants involved on the inequalities which provide the history-dependence of the operators 𝒮\mathcal{S} and \mathcal{R}. Let 𝒦\mathcal{K} and MnM_{n} be given by

𝒦=(L𝒢+LG)(d𝑬𝐐+1)\displaystyle\mathcal{K}=\left(L_{\mathcal{G}}+L_{G}\right)\left(d\|\boldsymbol{E}\|_{\mathbf{Q}_{\infty}}+1\right) (5.5)
Mn=Lpmax{1,c0}maxs[0,n]α(s)L(Γ3)n\displaystyle M_{n}=L_{p}\max\left\{1,c_{0}\right\}\max_{s\in[0,n]}\|\alpha(s)\|_{L^{\infty}\left(\Gamma_{3}\right)}\quad\forall n\in\mathbb{N} (5.6)

Then, a simple computation shows that for each nn\in\mathbb{N} the inequalities below holds:

𝒮𝒖(t)𝒮𝒗(t)Q×L2(Ω)m𝒦en𝒦0t𝒖(s)𝒗(s)V𝑑s\displaystyle\|\mathcal{S}\boldsymbol{u}(t)-\mathcal{S}\boldsymbol{v}(t)\|_{Q\times L^{2}(\Omega)^{m}}\leq\mathcal{K}e^{n\mathcal{K}}\int_{0}^{t}\|\boldsymbol{u}(s)-\boldsymbol{v}(s)\|_{V}ds
𝒖,𝒗C(+;V)t[0,n]\displaystyle\quad\forall\boldsymbol{u},\boldsymbol{v}\in C\left(\mathbb{R}_{+};V\right)\forall t\in[0,n] (5.7)
𝒖(t)𝒗(t)L2(Γ3)MnenMn0t𝒖(s)𝒗(s)V𝑑s\displaystyle\|\mathcal{R}\boldsymbol{u}(t)-\mathcal{R}\boldsymbol{v}(t)\|_{L^{2}\left(\Gamma_{3}\right)}\leq M_{n}e^{nM_{n}}\int_{0}^{t}\|\boldsymbol{u}(s)-\boldsymbol{v}(s)\|_{V}ds
𝒖,𝒗C(+;V)t[0,n]\displaystyle\quad\forall\boldsymbol{u},\boldsymbol{v}\in C\left(\mathbb{R}_{+};V\right)\forall t\in[0,n] (5.8)

We now state the following equivalence result whose proof represents a direct consequence of Lemmas 5.2 and 5.3.

Lemma 5.4 Let ( 𝝈,𝒖,𝜿,w,𝝀\boldsymbol{\sigma},\boldsymbol{u},\boldsymbol{\kappa},w,\boldsymbol{\lambda} ) be functions with regularity (5.1). Then ( 𝝈,𝒖,𝜿,w,𝝀\boldsymbol{\sigma},\boldsymbol{u},\boldsymbol{\kappa},w,\boldsymbol{\lambda} ) is a solution of Problem 𝒫V\mathcal{P}^{V} if and only if

𝝈(t)=𝜺(𝒖(t))+𝒮1𝒖(t)\displaystyle\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t))+\mathcal{S}_{1}\boldsymbol{u}(t) (5.9)
𝜿(t)=𝒮2𝒖(t)\displaystyle\boldsymbol{\kappa}(t)=\mathcal{S}_{2}\boldsymbol{u}(t) (5.10)
w(t)=𝒖(t)\displaystyle w(t)=\mathcal{R}\boldsymbol{u}(t) (5.11)
(𝜺(𝒖(t)),𝜺(𝒗))Q+(𝒮1𝒖(t),𝜺(𝒗))Q+Γ3p(uv(t)𝒖(t))vv𝑑a\displaystyle(\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}+\left(\mathcal{S}_{1}\boldsymbol{u}(t),\boldsymbol{\varepsilon}(\boldsymbol{v})\right)_{Q}+\int_{\Gamma_{3}}p\left(u_{v}(t)-\mathcal{R}\boldsymbol{u}(t)\right)v_{v}da (5.12)
+Γ3ηp(uv(t)𝒖(t))𝒏(t)𝒗τ𝑑a+b(𝒗,𝝀(t))=(𝒇(t),𝒗)V𝒗V\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-\mathcal{R}\boldsymbol{u}(t)\right)\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da+b(\boldsymbol{v},\boldsymbol{\lambda}(t))=(\boldsymbol{f}(t),\boldsymbol{v})_{V}\quad\forall\boldsymbol{v}\in V
b(𝒖(t),𝝁λ(t))b(g𝜽~,𝝁λ(t))𝝁Λ\displaystyle b(\boldsymbol{u}(t),\boldsymbol{\mu}-\lambda(t))\leq b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\lambda(t))\quad\forall\boldsymbol{\mu}\in\Lambda (5.13)

for all t+t\in\mathbb{R}_{+}.
Note that the interest of Lemma 5.4 arises in the fact that it decouples the unknowns of the Problem 𝒫V\mathcal{P}^{V}. Indeed, a careful examination of the system (5.9)-(5.13) shows that the unknowns 𝝈,𝜿\boldsymbol{\sigma},\boldsymbol{\kappa} and ww do not appear in the system (5.12)-(5.13), which contains only the unknowns 𝒖\boldsymbol{u} and 𝝀\boldsymbol{\lambda}. For this reason, the next step in the proof of Theorem 5.1 consists to obtain the unique solvability of the system (5.12)-(5.13).

Lemma 5.5 There exists e0>0e_{0}>0 which depends only on ,Ω,Γ1\mathcal{E},\Omega,\Gamma_{1} and Γ3\Gamma_{3} such that if Lp(1+ηL(Γ3))<e0L_{p}(1+\left.\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0} then there exists a unique couple of functions ( 𝒖,𝝀\boldsymbol{u},\boldsymbol{\lambda} ) which satisfies (5.12)(5.13) for all t+t\in\mathbb{R}_{+}. Moreover,

(𝒖,λ)C(+;V×Λ)(\boldsymbol{u},\lambda)\in C\left(\mathbb{R}_{+};V\times\Lambda\right) (5.14)

Proof We use the Riesz representation theorem to define the operators A:VVA:V\rightarrow V and 𝒮~:C(+;V)C(+;V)\tilde{\mathcal{S}}:C\left(\mathbb{R}_{+};V\right)\rightarrow C\left(\mathbb{R}_{+};V\right) by equalities

(A𝒖,𝒗)V=(𝜺(𝒖),𝜺(𝒗))Q𝒖,𝒗V\displaystyle(A\boldsymbol{u},\boldsymbol{v})_{V}=(\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}\quad\forall\boldsymbol{u},\boldsymbol{v}\in V (5.15)
(𝒮~𝒖(t),𝒗)V=(𝒮1𝒖(t),𝜺(𝒗))Q+Γ3p(uv(t)𝒖(t))vv𝑑a\displaystyle(\tilde{\mathcal{S}}\boldsymbol{u}(t),\boldsymbol{v})_{V}=\left(\mathcal{S}_{1}\boldsymbol{u}(t),\boldsymbol{\varepsilon}(\boldsymbol{v})\right)_{Q}+\int_{\Gamma_{3}}p\left(u_{v}(t)-\mathcal{R}\boldsymbol{u}(t)\right)v_{v}da
+Γ3ηp(uv(t)𝒖(t))𝒏(t)𝒗τ𝑑a𝒖C(+;V)𝒗V,t+\displaystyle+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-\mathcal{R}\boldsymbol{u}(t)\right)\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da\quad\forall\boldsymbol{u}\in C\left(\mathbb{R}_{+};V\right)\forall\boldsymbol{v}\in V,t\in\mathbb{R}_{+} (5.16)

Note that, by assumptions (3.21)-(3.23), the surface integrals in (5.16) are well defined. With these notation it is easy to see that the variational equation (5.12) is equivalent with

(A𝒖(t),𝒗)V+(𝒮~𝒖(t),𝒗)V+b(𝒗,λ(t))=(𝒇(t),𝒗)V𝒗V(A\boldsymbol{u}(t),\boldsymbol{v})_{V}+(\tilde{\mathcal{S}}\boldsymbol{u}(t),\boldsymbol{v})_{V}+b(\boldsymbol{v},\lambda(t))=(\boldsymbol{f}(t),\boldsymbol{v})_{V}\quad\forall\boldsymbol{v}\in V (5.17)

Therefore, to conclude the proof it is sufficient to show that there exists a unique couple of functions ( 𝒖,𝝀\boldsymbol{u},\boldsymbol{\lambda} ) with regularity (5.14), which satisfies (5.17) and (5.13) for all t+t\in\mathbb{R}_{+}. The main ingredient in the solution of this system is Theorem 2.4 and, to this end, we check in what follows the assumptions of this theorem.

First, using (3.11) we deduce that the operator AA, defined by (5.15), verifies (2.10). Let 𝒖1,𝒖2C(+;V),𝒗V,n\boldsymbol{u}_{1},\boldsymbol{u}_{2}\in C\left(\mathbb{R}_{+};V\right),\boldsymbol{v}\in V,n\in\mathbb{N} and t[0,n]t\in[0,n]. According to definition (5.16) of operator 𝒮~\tilde{\mathcal{S}} we have

(𝒮~𝒖1\displaystyle\left(\tilde{\mathcal{S}}\boldsymbol{u}_{1}\right. (t)𝒮~𝒖2(t),𝒗)V\displaystyle\left.(t)-\tilde{\mathcal{S}}\boldsymbol{u}_{2}(t),\boldsymbol{v}\right)_{V}
=\displaystyle= (𝒮1𝒖1(t)𝒮1𝒖2(t),𝜺(𝒗))Q\displaystyle\left(\mathcal{S}_{1}\boldsymbol{u}_{1}(t)-\mathcal{S}_{1}\boldsymbol{u}_{2}(t),\boldsymbol{\varepsilon}(\boldsymbol{v})\right)_{Q}
+Γ3[p(u1v(t)𝒖1(t))p(u2v(t)𝒖2(t))]vv𝑑a\displaystyle+\int_{\Gamma_{3}}\left[p\left(u_{1v}(t)-\mathcal{R}\boldsymbol{u}_{1}(t)\right)-p\left(u_{2v}(t)-\mathcal{R}\boldsymbol{u}_{2}(t)\right)\right]v_{v}da
+Γ3η[p(u1v(t)𝒖1(t))p(u2v(t)𝒖2(t))]𝒏(t)𝒗τ𝑑a\displaystyle+\int_{\Gamma_{3}}\eta\left[p\left(u_{1v}(t)-\mathcal{R}\boldsymbol{u}_{1}(t)\right)-p\left(u_{2v}(t)-\mathcal{R}\boldsymbol{u}_{2}(t)\right)\right]\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da (5.18)

Using assumptions (3.21)-(3.23), inequality (2.17) and estimates (5.7) and (5.8) we obtain

|(𝒮~𝒖1(t)𝒮~𝒖2(t),𝒗)V|\displaystyle\left|\left(\tilde{\mathcal{S}}\boldsymbol{u}_{1}(t)-\tilde{\mathcal{S}}\boldsymbol{u}_{2}(t),\boldsymbol{v}\right)_{V}\right|
[c02Lp(1+ηL(Γ3))]𝒖1(t)𝒖2(t)V𝒗V\displaystyle\quad\leq\left[c_{0}^{2}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)\right]\left\|\boldsymbol{u}_{1}(t)-\boldsymbol{u}_{2}(t)\right\|_{V}\|\boldsymbol{v}\|_{V}
+[𝒦en𝒦+c0Lp(1+ηL(Γ3))MnenMn]0t𝒖1(s)𝒖2(s)V𝑑s𝒗V\displaystyle\quad+\left[\mathcal{K}e^{n\mathcal{K}}+c_{0}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)M_{n}e^{nM_{n}}\right]\int_{0}^{t}\left\|\boldsymbol{u}_{1}(s)-\boldsymbol{u}_{2}(s)\right\|_{V}ds\|\boldsymbol{v}\|_{V} (5.19)

Thus,

𝒮~𝒖1(t)𝒮~𝒖2(t)V\displaystyle\left\|\tilde{\mathcal{S}}\boldsymbol{u}_{1}(t)-\tilde{\mathcal{S}}\boldsymbol{u}_{2}(t)\right\|_{V}
[c02Lp(1+ηL(Γ3))]𝒖1(t)𝒖2(t)V\displaystyle\leq\left[c_{0}^{2}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)\right]\left\|\boldsymbol{u}_{1}(t)-\boldsymbol{u}_{2}(t)\right\|_{V}
+[𝒦en𝒦+c0Lp(1+ηL(Γ3))MnenMn]0t𝒖1(s)𝒖2(s)V𝑑s\displaystyle\quad+\left[\mathcal{K}e^{n\mathcal{K}}+c_{0}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)M_{n}e^{nM_{n}}\right]\int_{0}^{t}\left\|\boldsymbol{u}_{1}(s)-\boldsymbol{u}_{2}(s)\right\|_{V}ds (5.20)

The previous inequality implies that the operator 𝒮~\tilde{\mathcal{S}} satisfies condition (2.11) with

d~n=c02Lp(1+ηL(Γ3))\tilde{d}_{n}=c_{0}^{2}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right) (5.21)

and

s~n=𝒦en𝒦+c0Lp(1+ηL(Γ3))MnenMn.\tilde{s}_{n}=\mathcal{K}e^{n\mathcal{K}}+c_{0}L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)M_{n}e^{nM_{n}}.

Next, as it was shown in [21,22], definition (4.3), implies that the bilinear form b(,)b(\cdot,\cdot) satisfies condition (2.12), i.e. there exist constants Mb>0M_{b}>0 and α>0\alpha>0 such that

|b(𝒗,𝝁)|Mb𝒗V𝝁D𝒗V,𝝁D|b(\boldsymbol{v},\boldsymbol{\mu})|\leq M_{b}\|\boldsymbol{v}\|_{V}\|\boldsymbol{\mu}\|_{D}\quad\forall\boldsymbol{v}\in V,\boldsymbol{\mu}\in D (5.22)

and

inf𝝁D,𝝁𝟎Dsup𝒗V,𝒗𝟎Vb(𝒗,𝝁)𝒗V𝝁Dα.\inf_{\boldsymbol{\mu}\in D,\boldsymbol{\mu}\neq\mathbf{0}_{D}}\sup_{\boldsymbol{v}\in V,\boldsymbol{v}\neq\mathbf{0}_{V}}\frac{b(\boldsymbol{v},\boldsymbol{\mu})}{\|\boldsymbol{v}\|_{V}\|\boldsymbol{\mu}\|_{D}}\geq\alpha. (5.23)

Finally, definition (4.4) and assumptions (3.20), (3.26) yield

𝒇C(+;V) and g𝜽~V.\boldsymbol{f}\in C\left(\mathbb{R}_{+};V\right)\quad\text{ and }\quad g\tilde{\boldsymbol{\theta}}\in V. (5.24)

The previous results allow us to apply Theorem 2.4 with X=V,Y=DX=V,Y=D and 𝒉=g𝜽~\boldsymbol{h}=g\tilde{\boldsymbol{\theta}}. According to this theorem there exists d0>0d_{0}>0 which depends only on 𝑬,Ω,Γ1\boldsymbol{E},\Omega,\Gamma_{1} and Γ3\Gamma_{3} such that, if d~n<d0\tilde{d}_{n}<d_{0}, for all positive integers nn, then there exists a unique couple of functions ( 𝒖,𝝀\boldsymbol{u},\boldsymbol{\lambda} ) with regularity (5.14), which satisfies (5.17) and (5.13) for all t+t\in\mathbb{R}_{+}. Denote

e0=d0c02e_{0}=d_{0}c_{0}^{-2} (5.25)

which, clearly, depends only on 𝑬,Ω,Γ1\boldsymbol{E},\Omega,\Gamma_{1} and Γ3\Gamma_{3}. Then, it follows from (5.21) and (5.25) that d~n<d0\tilde{d}_{n}<d_{0} iff Lp(1+ηL(Γ3))<e0L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0} which concludes the proof.

We now have all ingredients to prove our main existence and uniqueness result.

Proof of Theorem 5.1 Existence. Assume that Lp(1+ηL(Γ3))<e0L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0}, where e0e_{0} is defined by (5.25). Then, using Lemma 5.5 we deduce that there exists a unique couple of functions ( 𝒖,𝝀\boldsymbol{u},\boldsymbol{\lambda} ) such that (5.12)-(5.13) hold, for all t+t\in\mathbb{R}_{+}. Moreover, the solution has the regularity (5.14). Next, we introduce the functions 𝝈,𝜿\boldsymbol{\sigma},\boldsymbol{\kappa} and ww defined by (5.9)-(5.11). Taking into account assumption (3.11) and the regularity of operators 𝒮\mathcal{S} and \mathcal{R} we conclude that the triple ( 𝝈,𝜿,w\boldsymbol{\sigma},\boldsymbol{\kappa},w ) has the regularity ( 𝝈,𝜿,w)C(+;Q×L2(Ω)m×L2(Γ3)\boldsymbol{\sigma},\boldsymbol{\kappa},w)\in C\left(\mathbb{R}_{+};Q\times L^{2}(\Omega)^{m}\times L^{2}\left(\Gamma_{3}\right)\right. ). It follows from here that (5.1) holds. Lemma 5.4 implies now the existence part of the theorem.

Uniqueness. The uniqueness of the solution is now a consequence of the unique solvability of system (5.12)-(5.13), guaranteed by Lemma 5.5, combined with Lemma 5.4.

We conclude from above that, under the assumptions of Theorem 5.1, the contact problem 𝒫\mathcal{P} has a unique weak solution. Note that inequality Lp(1+ηL(Γ3))<e0L_{p}\left(1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0}, which guarantees the unique weak solvability of Problem 𝒫V\mathcal{P}^{V}, is verified if either the Lipschitz constant LpL_{p} or 1+ηL(Γ3)1+\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)} is small enough. Therefore, this condition represents a smallness condition on the normal compliance function and/or the coefficient of friction.

6 Particular Cases

The aim of this section is twofold. The first one is to provide examples of contact problems which represent particular cases of Problem 𝒫\mathcal{P} and whose unique weak solvability could be obtained by using Theorem 5.1. The second one is to compare the mixed variational formulation used in this paper with a different aproach, used in [34].

Elastic contact with wear We start by considering Problem 𝒫\mathcal{P} in the particular case when the material is elastic, i.e. when 𝒢𝟎,G𝟎\mathcal{G}\equiv\mathbf{0},G\equiv\mathbf{0} and 𝝈0=𝜺(𝒖0)\boldsymbol{\sigma}_{0}=\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{0}\right). The classical formulation of this problem is the following.

Problem 𝒫𝒆\mathcal{P}_{\boldsymbol{e}} Find a stress field 𝝈:Ω×+𝕊d\boldsymbol{\sigma}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{S}^{d}, a displacement field 𝒖:Ω×+d\boldsymbol{u}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{d} and a wear function w:Γ3×+w:\Gamma_{3}\times\mathbb{R}_{+}\rightarrow\mathbb{R} such that

𝝈(t)=𝜺(𝒖(t))\displaystyle\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)) in Ω,\displaystyle\text{ in }\Omega, (6.1)
Div𝝈(t)+𝒇0(t)=𝟎\displaystyle\operatorname{Div}\boldsymbol{\sigma}(t)+\boldsymbol{f}_{0}(t)=\mathbf{0} in Ω,\displaystyle\text{ in }\Omega, (6.2)
𝒖(t)=𝟎\displaystyle\boldsymbol{u}(t)=\mathbf{0} on Γ1,\displaystyle\text{ on }\Gamma_{1}, (6.3)
𝝈(t)𝒗=𝒇2(t)\displaystyle\boldsymbol{\sigma}(t)\boldsymbol{v}=\boldsymbol{f}_{2}(t) on Γ2,\displaystyle\text{ on }\Gamma_{2}, (6.4)
uv(t)g,(uv(t)g)(σv(t)+p(uv(t)w(t)))=0}\displaystyle\left.\begin{array}[]{l}u_{v}(t)\leq g,\\ \left(u_{v}(t)-g\right)\left(\sigma_{v}(t)+p\left(u_{v}(t)-w(t)\right)\right)=0\end{array}\right\} on Γ3,\displaystyle\text{ on }\Gamma_{3}, (6.5)
𝝈τ(t)=ηp(uv(t)w(t))𝒏(t)\displaystyle-\boldsymbol{\sigma}_{\tau}(t)=\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t) on Γ3,\displaystyle\text{ on }\Gamma_{3}, (6.6)
w˙(t)=α(t)p(uv(t)w(t))\displaystyle\dot{w}(t)=\alpha(t)p\left(u_{v}(t)-w(t)\right) on Γ3,\displaystyle\text{ on }\Gamma_{3}, (6.7)
w(0)=0\displaystyle w(0)=0 in Γ3,\displaystyle\text{ in }\Gamma_{3}, (6.8)

The mixed variational formulation of Problem 𝒫e\mathcal{P}_{e} follows from Sect. 4 and can be formulated as follows.

Problem 𝒫𝒆𝑽\mathcal{P}_{\boldsymbol{e}}^{\boldsymbol{V}}. Find a stress field 𝝈:+Q\boldsymbol{\sigma}:\mathbb{R}_{+}\rightarrow Q, a displacement field 𝒖:+V\boldsymbol{u}:\mathbb{R}_{+}\rightarrow V, a wear function w:+L2(Γ3)w:\mathbb{R}_{+}\rightarrow L^{2}\left(\Gamma_{3}\right) and a Lagrange multiplier λ:+Λ\lambda:\mathbb{R}_{+}\rightarrow\Lambda such that

𝝈(t)=𝜺(𝒖(t))\displaystyle\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)) (6.9)
w(t)=0tα(s)p(uv(s)w(s))𝑑s\displaystyle w(t)=\int_{0}^{t}\alpha(s)p\left(u_{v}(s)-w(s)\right)ds (6.10)
(𝝈(t),𝜺(𝒗))Q+Γ3p(uv(t)w(t))vv𝑑a+b(𝒗,𝝀(t))\displaystyle(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v}))_{Q}+\int_{\Gamma_{3}}p\left(u_{v}(t)-w(t)\right)v_{v}da+b(\boldsymbol{v},\boldsymbol{\lambda}(t))
+Γ3ηp(uv(t)w(t))𝒏(t)𝒗τ𝑑a=(𝒇(t),𝒗)V𝒗V\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t)\cdot\boldsymbol{v}_{\tau}da=(\boldsymbol{f}(t),\boldsymbol{v})_{V}\quad\forall\boldsymbol{v}\in V (6.11)
b(𝒖(t),𝝁𝝀(t))b(g𝜽~,𝝁𝝀(t))𝝁Λ\displaystyle b(\boldsymbol{u}(t),\boldsymbol{\mu}-\boldsymbol{\lambda}(t))\leq b(g\tilde{\boldsymbol{\theta}},\boldsymbol{\mu}-\boldsymbol{\lambda}(t))\quad\forall\boldsymbol{\mu}\in\Lambda (6.12)

for all t+t\in\mathbb{R}_{+}.
The unique solvability of this problem follows from Theorem 5.1, under the assumptions (3.11), (3.20)-( 3.24), (3.26), combined with a smallness assumption of the form Lp(1+ηL(Γ3))<e0L_{p}(1+\left.\|\eta\|_{L^{\infty}\left(\Gamma_{3}\right)}\right)<e_{0} for the coefficient of friction.

Note that the elastic contact problem 𝒫e\mathcal{P}_{e} was studied in [34], under the same assumptions. There, using the set of admissible displacement fields given by

U={𝒗V:vvg a.e. on Γ3}U=\left\{\boldsymbol{v}\in V:v_{v}\leq g\text{ a.e. on }\Gamma_{3}\right\}

the following three-fields variational formulation of the problem was derived.
Problem 𝒫~𝒆𝑽\tilde{\mathcal{P}}_{\boldsymbol{e}}^{\boldsymbol{V}} Find a stress field 𝝈:+Q\boldsymbol{\sigma}:\mathbb{R}_{+}\rightarrow Q, a displacement field 𝒖:+U\boldsymbol{u}:\mathbb{R}_{+}\rightarrow U and a wear function w:+L2(Γ3)w:\mathbb{R}_{+}\rightarrow L^{2}\left(\Gamma_{3}\right) such that

𝝈(t)=𝜺(𝒖(t))\displaystyle\boldsymbol{\sigma}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)) (6.13)
(𝝈(t),𝜺(𝒗)𝜺(𝒖(t)))Q+Γ3p(uv(t)w(t))(vvuv(t))𝑑a\displaystyle(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{v})-\boldsymbol{\varepsilon}(\boldsymbol{u}(t)))_{Q}+\int_{\Gamma_{3}}p\left(u_{v}(t)-w(t)\right)\left(v_{v}-u_{v}(t)\right)da
+Γ3ηp(uv(t)w(t))𝒏(t)(𝒗τ𝒖τ(t))𝑑a(𝒇(t),𝒗𝒖(t))V𝒗U\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{v}(t)-w(t)\right)\boldsymbol{n}^{*}(t)\cdot\left(\boldsymbol{v}_{\tau}-\boldsymbol{u}_{\tau}(t)\right)da\geq(\boldsymbol{f}(t),\boldsymbol{v}-\boldsymbol{u}(t))_{V}\quad\forall\boldsymbol{v}\in U (6.14)
w(t)=0tα(s)p(uv(s)w(s))𝑑s\displaystyle w(t)=\int_{0}^{t}\alpha(s)p\left(u_{v}(s)-w(s)\right)ds (6.15)

for all t+t\in\mathbb{R}_{+}.
Then, the existence of a unique solution of the problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} was derived in several steps, which could be resumed as follows.
(i) In the first step it is proved that, for a given wear function wC(+;L2(Γ3))w\in C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right), there exists a unique displacement filed 𝒖wC(+;U)\boldsymbol{u}_{w}\in C\left(\mathbb{R}_{+};U\right) such that

(𝜺(𝒖w(t)),𝜺(𝒗)𝜺(𝒖w(t)))Q+Γ3p(uwν(t)w(t))(vνuwν(t))𝑑a\displaystyle\left(\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{w}(t)\right),\boldsymbol{\varepsilon}(\boldsymbol{v})-\boldsymbol{\varepsilon}\left(\boldsymbol{u}_{w}(t)\right)\right)_{Q}+\int_{\Gamma_{3}}p\left(u_{w\nu}(t)-w(t)\right)\left(v_{\nu}-u_{w\nu}(t)\right)da
+Γ3ηp(uwν(t)w(t))𝒏(t)(𝒗τ𝒖wτ(t))𝑑a(𝒇(t),𝒗𝒖w(t))V𝒗U\displaystyle\quad+\int_{\Gamma_{3}}\eta p\left(u_{w\nu}(t)-w(t)\right)\boldsymbol{n}^{*}(t)\cdot\left(\boldsymbol{v}_{\tau}-\boldsymbol{u}_{w\tau}(t)\right)da\geq\left(\boldsymbol{f}(t),\boldsymbol{v}-\boldsymbol{u}_{w}(t)\right)_{V}\quad\forall\boldsymbol{v}\in U (6.16)

for all t+t\in\mathbb{R}_{+}.
(ii) Then, it was shown that the operator Λ:C(+;L2(Γ3))C(+;L2(Γ3))\Lambda:C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right)\rightarrow C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right) defined by

Λw(t)=0tα(s)p(uwv(s)w(s))𝑑s\Lambda w(t)=\int_{0}^{t}\alpha(s)p\left(u_{wv}(s)-w(s)\right)ds (6.17)

has a unique fixed point wC(+;L2(Γ3))w^{*}\in C\left(\mathbb{R}_{+};L^{2}\left(\Gamma_{3}\right)\right).
(iii) Finally, defining 𝒖\boldsymbol{u}^{*} and 𝝈\boldsymbol{\sigma}^{*} by equalities 𝒖=𝒖w,𝝈=𝜺(𝒖)\boldsymbol{u}^{*}=\boldsymbol{u}_{w^{*}},\boldsymbol{\sigma}^{*}=\mathcal{E}\boldsymbol{\varepsilon}\left(\boldsymbol{u}^{*}\right), it was proved that the triple (𝝈,𝒖,w)\left(\boldsymbol{\sigma}^{*},\boldsymbol{u}^{*},w^{*}\right) is the uniqe weak solution to Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V}.
A brief comparasion between problems Problem 𝒫eV\mathcal{P}_{e}^{V} and Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} leads to the following comments.
(a) Problem 𝒫eV\mathcal{P}_{e}^{V} represents a four-fields variational formulations of the mechanical contact problem 𝒫e\mathcal{P}_{e} while, in contrast, Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} represent a three-fields variational formulations of the same problem.
(b) Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} involves a variational inequality with contraints, (6.14). In contrast, Problem 𝒫eV\mathcal{P}_{e}^{V} involves a variational equation without contraints, (6.11). Removing the constraints in (6.14) was possible by introducing a new variable, the Lagrange multiplier λ\lambda. Considering mixed formulations based on the Lagrange multiplier has important advantages in the numerical solution of the contact problems, as explained in [14, 16, 18].
(c) Under the same assumption on the data, both Problem 𝒫eV\mathcal{P}_{e}^{V} and Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} have a unique weak solution, with the same regularity. Moreover, their solvability is guaranteed by a smallness assumption on the coefficient of friction. Nevertheless, the question if this assumption represents an intrinsic feature of the contact Problem 𝒫e\mathcal{P}_{e} or it describes a limitation of the mathematical methods used in solving the problems 𝒫eV\mathcal{P}_{e}^{V} and 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} is left open.
(d) The solution of Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} is based on arguments of time-dependent variational inequalities of the first kind, combined with the fixed point argument provided by Theorem 2.1. In contrast, the solution of Problem 𝒫eV\mathcal{P}_{e}^{V} is based on the more elaborate result provided by Theorem 2.4, which already integrates a fixed point argument.
(e) The equivalence between Problems 𝒫eV\mathcal{P}_{e}^{V} and Problem 𝒫~eV\tilde{\mathcal{P}}_{e}^{V} represents an open question. As far as this equivalence is not proved, we conclude that a contact problem could have different variational formulations and, therefore, the concept of weak solution for such a problem is not an intrinsic one.

Viscoplastic contact with normal compliance and unilateral constraint We now consider Problem 𝒫\mathcal{P} in the particular case when the material is viscoplastic, without internal state variable, the contact is frictionless and the wear of the contact surfaces in neglecting. Therefore, we take G𝟎,η0G\equiv\mathbf{0},\eta\equiv 0 and α0\alpha\equiv 0 to obtain the following contact model.

Problem 𝒫𝒗𝒑\mathcal{P}_{\boldsymbol{v}\boldsymbol{p}} Find a stress field 𝝈:Ω×+𝕊d\boldsymbol{\sigma}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{S}^{d} and a displacement field 𝒖:Ω×+d\boldsymbol{u}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{d} such that

𝝈˙(t)=𝜺(𝒖˙(t))+𝒢(𝝈(t),𝜺(𝒖(t))) in Ω,Div𝝈(t)+𝒇0(t)=𝟎 in Ω,𝒖(t)=𝟎 on Γ1,𝝈(t)𝒗=𝒇2(t) on Γ2,uv(t)g,σv(t)+p(uv(t))0,(uv(t)g)(σv(t)+p(uv(t)))=0} on Γ3,,𝝈τ(t)=𝟎 on Γ3,,𝒖(0)=𝒖0 in Ω.,\left.\begin{array}[]{rl}\dot{\boldsymbol{\sigma}}(t)=\mathcal{E}\boldsymbol{\varepsilon}(\dot{\boldsymbol{u}}(t))+\mathcal{G}(\boldsymbol{\sigma}(t),\boldsymbol{\varepsilon}(\boldsymbol{u}(t)))&\text{ in }\Omega,\\ \operatorname{Div}\boldsymbol{\sigma}(t)+\boldsymbol{f}_{0}(t)=\mathbf{0}&\text{ in }\Omega,\\ \boldsymbol{u}(t)=\mathbf{0}&\text{ on }\Gamma_{1},\\ \boldsymbol{\sigma}(t)\boldsymbol{v}=\boldsymbol{f}_{2}(t)&\text{ on }\Gamma_{2},\\ u_{v}(t)\leq g,\quad\sigma_{v}(t)+p\left(u_{v}(t)\right)\leq 0,\\ \left(u_{v}(t)-g\right)\left(\sigma_{v}(t)+p\left(u_{v}(t)\right)\right)=0\end{array}\right\}\text{ on }\Gamma_{3},,\quad\boldsymbol{\sigma}_{\tau}(t)=\mathbf{0}\quad\text{ on }\Gamma_{3},,\quad\boldsymbol{u}(0)=\boldsymbol{u}_{0}\quad\text{ in }\Omega.,

Problem 𝒫vp\mathcal{P}_{vp} was considered in [2]. There, besides the unique solvability of the problem, the continuous dependence of the weak solution with respect to both the normal compliance function and the penetration bound was proved. Numerical simulations which provide a numerical evidence of this continuous dependence result were also perfomed.

Signorini frictionless problem with gap We finally consider Problem 𝒫vp\mathcal{P}_{vp} in the particular case when the material is elastic and the normal compliance vanishes. Therefore, taking 𝒢𝟎,p0\mathcal{G}\equiv\mathbf{0},p\equiv 0 and 𝝈=(𝜺𝒖0)\boldsymbol{\sigma}=\mathcal{E}\left(\boldsymbol{\varepsilon}\boldsymbol{u}_{0}\right) in (6.18)-(6.24) we obtain the following contact model.

Problem 𝒫S\mathcal{P}_{S} Find a stress field 𝝈:Ω×+𝕊d\boldsymbol{\sigma}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{S}^{d} and a displacement field 𝒖:Ω×+d\boldsymbol{u}:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{d} such that

𝝈=𝜺(𝒖(t))\displaystyle\boldsymbol{\sigma}=\mathcal{E}\boldsymbol{\varepsilon}(\boldsymbol{u}(t)) in Ω,\displaystyle\text{ in }\Omega, (6.25)
Div𝝈(t)+𝒇0(t)=𝟎\displaystyle\operatorname{Div}\boldsymbol{\sigma}(t)+\boldsymbol{f}_{0}(t)=\mathbf{0} in Ω,\displaystyle\text{ in }\Omega, (6.26)
𝒖(t)=𝟎\displaystyle\boldsymbol{u}(t)=\mathbf{0} on Γ1,\displaystyle\text{ on }\Gamma_{1}, (6.27)
𝝈(t)𝒗=𝒇2(t)\displaystyle\boldsymbol{\sigma}(t)\boldsymbol{v}=\boldsymbol{f}_{2}(t) on Γ2,\displaystyle\text{ on }\Gamma_{2}, (6.28)
uv(t)g,σv(t)0,\displaystyle\begin{array}[]{l}u_{v}(t)\leq g,\\ \sigma_{v}(t)\leq 0,\end{array} on Γ3,\displaystyle\text{ on }\Gamma_{3}, (6.29)
uv(t)g)σv(t)=0\displaystyle\left.u_{v}(t)-g\right)\sigma_{v}(t)=0 (6.30)
𝝈τ(t)=𝟎\displaystyle\boldsymbol{\sigma}_{\tau}(t)=\mathbf{0} on Γ3,\displaystyle\text{ on }\Gamma_{3},

Note that Problem 𝒫S\mathcal{P}_{S} represents the time-dependent version of the famous Signorini frictionless contact problem, see for instance [29] and the references therein. The mixed variational method presented in this paper could be applied to in the study of Problems 𝒫vp\mathcal{P}_{vp} and 𝒫S\mathcal{P}_{S} in order to provide the unique solvability of these problems. It also provides the background for their numerical simulations.

Acknowledgements The work of the authors has been partially supported by project LEA Math Mode 2014/2015. The work of the second author has been partially supported by project POSDRU/159/1.5/S/ 132400: Young successful researchers-professional development in an international and interdisciplinary environment at Babeş-Bolyai University.

References

  1. 1.

    Archard, J.F., Hirst, W.: Wear of metals under unlubricated conditions. Proc. R. Soc. Lond. Ser. A 236, 3-55 (1956)

  2. 2.

    Barboteu, M., Matei, A., Sofonea, M.: On the behavior of the solution of a viscoplastic contact problem. Q. Appl. Math. 72, 625-647 (2014)

  3. 3.

    Céa, J.: Optimization. Théorie et Algorithmes. Dunod, Paris (1971)

  4. 4.

    Chudzikiewicz, A., Myśliński, A.: Thermoelastic wheel-rail contact problem with elastic gradedmaterials. Wear 271, 417-425 (2011)

  5. 5.

    Cristescu, N., Suliciu, I.: Viscoplasticity. Nijhoff/Editura Tehnică, Bucharest (1982)

  6. 6.

    Duvaut, G., Lions, J.-L.: Inequalities in Mechanics and Physics. Springer, Berlin (1976)

  7. 7.

    Eck, C., Jarušek, J., Krbeč, M.: Unilateral Contact Problems: Variational Methods and Existence Theorems. Pure and Applied Mathematics, vol. 270. Chapman/CRC Press, New York (2005)

  8. 8.

    Ekeland, I., Temam, R.: Convex Analysis and Variational Problems. Classics in Applied Mathematics, vol. 28. SIAM, Philadelphia (1999)

  9. 9.

    Glodež, S., Ren, Z., Flašker, J.: Simulation of surface pitting due to contact loading. Int. J. Numer. Methods Eng. 43, 33-50 (1998)

  10. 10.

    Goryacheva, I.G.: Contact Mechanics in Tribology. Kluwer, Dordrecht (1988)

  11. 11.

    Gu, R.J., Shillor, M.: Thermal and wear analysis of an elastic beam in sliding contact. Int. J. Solids Struct. 38, 2323-2333 (2001)

  12. 12.

    Gu, R.J., Kuttler, K.L., Shillor, M.: Frictional wear of a thermoelastic beam. J. Math. Anal. Appl. 242, 212-236 (2000)

  13. 13.

    Han, W., Sofonea, M.: Quasistatic Contact Problems in Viscoelasticity and Viscoplasticity. Studies in Advanced Mathematics, vol. 30. Am. Math. Soc./International Press, Providence/Somerville (2002)

  14. 14.

    Haslinger, J., Hlaváček, I., Nečas, J.: Numerical methods for unilateral problems in solid mechanics. In: Ciarlet, P.G., Lions, J.-L. (eds.) Handbook of Numerical Analysis, vol. IV, pp. 313-485. North-Holland, Amsterdam (1996)

  15. 15.

    Hild, P., Renard, Y.: A stabilized Lagrange multiplier method for the finite element approximation of contact problems in elastostatics. Numer. Math. 115, 101-129 (2010)

  16. 16.

    Hüeber, S., Matei, A., Wohlmuth, B.: Efficient algorithms for problems with friction. SIAM J. Sci. Comput. 29, 70-92 (2007)

  17. 17.

    Ionescu, I.R., Sofonea, M.: Functional and Numerical Methods in Viscoplasticity. Oxford Univ. Press, Oxford (1993)

  18. 18.

    Lions, J.-L., Glowinski, R., Trémolières, R.: Numerical Analysis of Variational Inequalities. NorthHolland, Amsterdam (1981)

  19. 19.

    Matei, A.: On the solvability of mixed variational problems with solution-dependent sets of Lagrange multipliers. Proc. R. Soc. Edinb., Sect. A, Math. 143, 1047-1059 (2013)

  20. 20.

    Matei, A.: An evolutionary mixed variational problem arising from frictional contact mechanics. Math. Mech. Solids 19, 223-239 (2014)

  21. 21.

    Matei, A., Ciurcea, R.: Contact problems for nonlinearity elastic materials: weak solvability involving dual Lagrange multipliers. ANZIAM J. 52, 160-178 (2010)

  22. 22.

    Matei, A., Ciurcea, R.: Weak solutions for contact problems involving viscoelastic materials with long memory. Math. Mech. Solids 16, 393-405 (2011)

  23. 23.

    Panagiotopoulos, P.D.: Inequality Problems in Mechanics and Applications. Birkhauser, Boston (1985)

  24. 24.

    Panagiotopoulos, P.D.: Hemivariational Inequalities, Applications in Mechanics and Engineering. Springer, Berlin (1993)

  25. 25.

    Rabinowiz, E.: Friction and Wear of Materials, 2nd edn. Wiley, New York (1995)

  26. 26.

    Rochdi, M., Shillor, M., Sofonea, M.: Quasistatic viscoelastic contact with normal compliance and friction. J. Elast. 51, 105-126 (1998)

  27. 27.

    Rodriguez-Arós, A., Viaño, J.M., Sofonea, M.: Asymptotic derivation of quasistatic frictional contact models with wear for elastic rods. J. Math. Anal. Appl. 401, 641-653 (2013)

  28. 28.

    Rojek, J., Telega, J.J., Stupkiewicz, S.: Contact problems with friction, adhesion and wear in orthopaedic biomechanics, II: numerical implementation and application to implanted knee joints. J. Theor. Appl. Mech. 39, 679-706 (2001)

  29. 29.

    Shillor, M., Sofonea, M., Telega, J.J.: Models and Analysis of Quasistatic Contact. Lecture Notes in Physics, vol. 655. Springer, Berlin (2004)

  30. 30.

    Sofonea, M., Matei, A.: History-dependent quasivariational inequalities arising in contact mechanics. Eur. J. Appl. Math. 22, 471-491 (2011)

  31. 31.

    Sofonea, M., Matei, A.: Mathematical Models in Contact Mechanics. London Mathematical Society Lecture Note Series, vol. 398. Cambridge Univ. Press, Cambridge (2012)

  32. 32.

    Sofonea, M., Matei, A.: History-dependent mixed variational problems in contact mechanics. J. Glob. Optim. 61, 591-614 (2015)

  33. 33.

    Sofonea, M., Avramescu, C., Matei, A.: A fixed point result with applications in the study of viscoplastic frictionless contact problems. Commun. Pure Appl. Anal. 7, 645-658 (2008)

  34. 34.

    Sofonea, M., Pătrulescu, F., Souleiman, Y.: Analysis of a contact problem with wear and unilateral constraint. Appl. Anal. 95, 2602-2619 (2016)

  35. 35.

    Strömberg, N.: A Newton method for three-dimensional fretting problems. Int. J. Solids Struct. 36, 20752090 (1999)

  36. 36.

    Strömberg, N., Johansson, L., Klarbring, A.: Derivation and analysis of a generalized standard model for contact, friction and wear. Int. J. Solids Struct. 33, 1817-1836 (1996)

  37. 37.

    Zmitrowicz, A.: Variational descriptions of wearing out solids and wear particles in contact mechanics. J. Theor. Appl. Mech. 39, 791-808 (2001)

  38. 38.

    Zmitrowicz, A.: Wear patterns and laws of wear-a review. J. Theor. Appl. Mech. 44, 219-253 (2006)

2018

Related Posts