Processing math: 54%
Research article

A class of thermal sub-differential contact problems

  • Received: 16 April 2017 Accepted: 28 November 2017 Published: 04 December 2017
  • We study a class of dynamic sub-differential contact problems with friction, and thermal e ects, for time depending long memory visco-elastic materials, with or without the clamped condition. We describe the mechanical problem, derive its variational formulation, and after specifying the assumptions on the data and operators, we prove an existence and uniqueness of weak solution on displacement and temperature fields. Then we present a fully discrete scheme for numerical approximations of the different solutions, and provide analysis of error order estimates. Finally various numerical computations in dimension two will be given.

    Citation: Oanh Chau. A class of thermal sub-differential contact problems[J]. AIMS Mathematics, 2017, 2(4): 658-681. doi: 10.3934/Math.2017.4.658

    Related Papers:

    [1] Fujun Cao, Dongfang Yuan . A class of HOC finite difference method for elliptic interface problems with imperfect contact. AIMS Mathematics, 2023, 8(3): 5789-5815. doi: 10.3934/math.2023292
    [2] Ibrahim-Elkhalil Ahmed, Ahmed E. Abouelregal, Doaa Atta, Meshari Alesemi . A fractional dual-phase-lag thermoelastic model for a solid half-space with changing thermophysical properties involving two-temperature and non-singular kernels. AIMS Mathematics, 2024, 9(3): 6964-6992. doi: 10.3934/math.2024340
    [3] Marion Darbas, Jérémy Heleine, Renier Mendoza, Arrianne Crystal Velasco . Sensitivity analysis of the complete electrode model for electrical impedance tomography. AIMS Mathematics, 2021, 6(7): 7333-7366. doi: 10.3934/math.2021431
    [4] Yakufu Kasimu, Gulijiamali Maimaitiaili . Non-fragile ${H_\infty }$ filter design for uncertain neutral Markovian jump systems with time-varying delays. AIMS Mathematics, 2024, 9(6): 15559-15583. doi: 10.3934/math.2024752
    [5] Austine Efut Ofem, Jacob Ashiwere Abuchu, Godwin Chidi Ugwunnadi, Hossam A. Nabwey, Abubakar Adamu, Ojen Kumar Narain . Double inertial steps extragadient-type methods for solving optimal control and image restoration problems. AIMS Mathematics, 2024, 9(5): 12870-12905. doi: 10.3934/math.2024629
    [6] Lin Fan, Shunchu Li, Dongfeng Shao, Xueqian Fu, Pan Liu, Qinmin Gui . Elastic transformation method for solving the initial value problem of variable coefficient nonlinear ordinary differential equations. AIMS Mathematics, 2022, 7(7): 11972-11991. doi: 10.3934/math.2022667
    [7] Qian Cao, Xin Long . New convergence on inertial neural networks with time-varying delays and continuously distributed delays. AIMS Mathematics, 2020, 5(6): 5955-5968. doi: 10.3934/math.2020381
    [8] Nadeem Abbas, Wasfi Shatanawi, Taqi A. M. Shatnawi . Innovation of prescribe conditions for radiative Casson micropolar hybrid nanofluid flow with inclined MHD over a stretching sheet/cylinder. AIMS Mathematics, 2025, 10(2): 3561-3580. doi: 10.3934/math.2025164
    [9] Zhi Guang Li . Global regularity and blowup for a class of non-Newtonian polytropic variation-inequality problem from investment-consumption problems. AIMS Mathematics, 2023, 8(8): 18174-18184. doi: 10.3934/math.2023923
    [10] Limei Xue, Jianmin Song, Shenghua Wang . A modified projection and contraction method for solving a variational inequality problem in Hilbert spaces. AIMS Mathematics, 2025, 10(3): 6128-6143. doi: 10.3934/math.2025279
  • We study a class of dynamic sub-differential contact problems with friction, and thermal e ects, for time depending long memory visco-elastic materials, with or without the clamped condition. We describe the mechanical problem, derive its variational formulation, and after specifying the assumptions on the data and operators, we prove an existence and uniqueness of weak solution on displacement and temperature fields. Then we present a fully discrete scheme for numerical approximations of the different solutions, and provide analysis of error order estimates. Finally various numerical computations in dimension two will be given.


    1. Introduction

    Phenomena of contact between deformable bodies abound since the dawn of time. In order to understand their inherent complexity, considerable efforts have been achieved in modeling, mathematical analysis and numerical simulations, within the weak distributional formulation framework, expressed in terms of evolutional variational inequalities and hemivariational inequalities. The literature dedicated to this field is increasing day by day. The state of the art can be found in the masterpieces [4], [6], [7], [8].

    This work is a continuation of the paper in [1], where the authors studied a dynamic frictional sub-differential contact for a short memory visco-elastic body, which was supposed to be fixed in some part on its boundary. There were no thermal effects and no numerical studies.

    Here we investigate the extension of this work to thermal contact with friction, for time depending long memory visco-elastic materials, with or without the clamped condition. Moreover, we propose a fully discrete scheme for numerical approximations of the different solutions, and elaborate a general numerical analysis of error estimates. Finally various corresponding numerical computations in dimension two will be given.

    The paper is organized as follows. In Section 2 we describe the mechanical problem, its corresponding variational formulation, and then we claim the main existence and uniqueness result under specific assumptions, that we prove in Section 3. In Section 4, we introduce a fully discrete approximation scheme, derive and prove an optimal order error estimate under certain solution regularity assumptions. Finally in Section 5, we present several numerical simulations, showing then the evolution of the displacement field and temperature, as well as of the Von Mise's stress norm.


    2. The contact problem

    In this section we study a class of thermal contact problems with sub-differential conditions, for long memory visco-elastic materials. We describe the mechanical problem, list the assumptions on the data and derive the corresponding variational formulation. Then we state an existence and uniqueness result on displacement field and temperature, which we will prove in the next section.

    The physical setting is as follows. A visco-elastic body occupies a domain Ω in Rd (d=1, d=2 or d=3) with a Lipschitz boundary Γ that is partionned into three disjoint measurable parts, Γ1, Γ2 and Γ3. We denote by ν the unit outward normal on Γ. Let [0,T] be the time interval of interest, where T>0. The body is clamped on Γ1×(0,T) and therefore the displacement field vanishes there. Here we suppose that meas(Γ1)=0 or meas(Γ1)>0, which means that Γ1 may be an empty set or reduced to a finite set of points. We assume that a volume force of density f0 acts in Ω×(0,T) and that surface tractions of density f2 act on Γ2×(0,T). The body may come in contact with an obstacle, the foundation, over the potential contact surface Γ3. The model of the contact is specified by a general sub-differential boundary condition, where thermal effects may occur in the frictional contact with the basis. We are interested in the dynamic evolution of the body.

    Let us recall now some classical notations, see e.g. [4] for further details. We denote by Sd the space of second order symmetric tensors on Rd, while `` " and || will represent the inner product and the Euclidean norm on Sd and Rd. Everywhere in the sequel the indices i and j run from 1 to d, summation over repeated indices is implied and the index that follows a comma represents the partial derivative with respect to the corresponding component of the independent variable. We also use the following notation:

    H=(L2(Ω))d,H={σ=(σij)|σij=σjiL2(Ω), 1i,jd},
    H1={uH|ε(u)H},H1={σH|DivσH}.

    Here ε:H1H and Div:H1H are the deformation and the divergence operators, respectively, defined by :

    ε(u)=(εij(u)),εij(u)=12(ui,j+uj,i),Divσ=(σij,j).

    The spaces H, H, H1 and H1 are real Hilbert spaces endowed with the canonical inner products given by :

    (u,v)H=Ωuividx,(σ,τ)H=Ωσijτijdx,
    (u,v)H1=(u,v)H+(ε(u),ε(v))H,(σ,τ)H1=(σ,τ)H+(Divσ,Divτ)H.

    Recall that D(Ω) denotes the set of infinitely differentiable real functions with compact support in Ω; and Wm,p(Ω), Hm(Ω):=Wm,2(Ω), mN, 1p+ for the classical real Sobolev spaces; Lp(U;X) the classical Lp spaces defined on U with values in X.

    To continue, the mechanical problem is then formulated as follows.

    Problem Q: Find a displacement field u:(0,T)×ΩRd and a stress field σ:(0,T)×ΩSd and a temperature field ξ:(0,T)×ΩR+ such that for a.e. t(0,T):

    σ(t)=A(t)ε(˙u(t))+G(t)ε(u(t))+t0B(ts)ε(u(s))ds+Ce(t,ξ(t))inΩ (2.1)
    ¨u(t)=Divσ(t)+f0(t)inΩ (2.2)
    u(t)=0onΓ1 (2.3)
    σ(t)ν=f2(t)onΓ2 (2.4)
    u(t)U,φu(t,w)φu(t,˙u(t))σ(t)ν(w˙u(t))wU on  Γ3 (2.5)
    ˙ξ(t)div(Kc(t,ξ(t)))=Ce(t,˙u(t))+q(t)inΩ, (2.6)
    Kc(t,x,ξ(t,x))ν:=Ξ(t,x,ξ(t,x))φ(t,x,ξ(t,x))a.e.xΓ3, (2.7)
    ξ(t)=θa(t)onΓ1Γ2 (2.8)
    ξ(0)=ξ0inΩ (2.9)
    u(0)=u0,˙u(0)=v0inΩ (2.10)

    Here, (2.1) is the Kelving Voigt's time-dependent long memory thermo-visco-elastic constitutive law of the body, where σ represents the stress tensor; A denotes the viscosity operator depending on the velocity of infinitesimal deformations ε(˙u), with the notation : for τSd, A(t)τ=A(t,,τ) some function defined on Ω; here a dot above a quantity represents the derivative of the quantity with respect to the time variable; G is the elastic operator depending on the linearized strain tensor ε(u) of infinitesimal deformations, with G(t)τ=G(t,,τ) which is defined on Ω. The term B(t)τ=B(t,,τ) represents the so called relaxation tensor which is time-depending on the linearized strain tensor and is defined on Ω. Recall that the visco-elastic short memory corresponds to the case B0. The last tensor Ce(t,θ):=Ce(t,,θ) denotes the thermal expansion tensor depending on time and on the temperature, defined on Ω. For example,

    Ce(t,θ):=θCexp(t)in Ω,

    where

    Cexp(t):=(cij(t,))

    is some time-depending expansion tensor, defined on Ω.

    In (2.2) is the dynamic equation of motion where the mass density ϱ1. The equation in (2.3) is the clamped condition and in (2.4) is the traction condition. On the contact surface, the general relation (2.5) is a sub-differential boundary condition such that

    D(Ω)dU,

    where U represents the set of contact admissible test functions and D(Ω)d is the distribution space. Here σν denotes the Cauchy stress vector on the contact boundary and φu:(0,T)×Γ3×RdR is a given function. Various situations may be modelled by such a condition, see e.g. the monograph [6] or the Habilitation thesis [2] p. 117. The differential equation (2.6) describes the evolution of the temperature field, where Kc(t,ξ):=Kc(t,,ξ) is some nonlinear thermal conductivity function defined on Ω, depending on time and on the temperature gradient ξ. For axample, denote by

    Kc(t,):=(kij(t,))

    the thermal conductivity tensor defined on Ω, we could consider

    Kc(t,,ξ)=Kc(t,)ξ.

    In the second member, Ce(t,˙u(t)):=Ce(t,,˙u(t)) is some nonlinear function defied on Ω, depending on the displacement velocity, and q(t) represents the density of volume heat sources. For example,

    Ce(t,˙u(t))=Cexp(t):˙u(t)=cij(t,)˙uixj(t).

    The associated temperature boundary condition is given by (2.7) and (2.8), where Ξ and φ are some functions defined on (0,T)×Γ3×R and θa(t) represents the ambient temperature. Here

    φ(t,x,r):=φ(t,x,)(r), (t,x,r)(0,T)×Γ3×R

    denotes the sub-differential on the third variable of φ in the convex or maybe locally Lipschitz framework.

    We recall that for a locally Lipschitz function G:RR, at any point aR and for any vector dR, we can define the following directional derivative with respect to d :

    ¯limτ0+G(a+τd)G(a)τ:=G0(a;d). (2.11)

    We have for all a,dR, for all ξG(a):

    G0(a;d)ξd

    and

    |G0(a;d)||G0(a)|×|d|,|ξ||G0(a)|

    where

    ¯limh0,h0G(a+h)G(a)h:=G0(a).

    In the case where G is convex on R, we have

    G0(a;d)={Gr(a)difd>0Gl(a)difd<00 ifd=0, 

    and

    G0(a)=max{Gr(a),Gl(a)},

    where Gr and Gl denotes the right side and left side derivatives respectively.

    In the sequel, for a.e. (t,x)(0,T)×Γc, for all (r,s)R2, we use the notation

    φ0(t,x,r;s):=[φ(t,x,)]0(r;s),

    and

    φ0(t,x,r):=[φ(t,x,)]0(r).

    Taking the previous example for Kc, we have

    Kc(t,x,ξ)ν=kij(t,x)ξxjνi.

    Let consider the following standard example

    φ(t,x,r):=12ke(t,x)(rθR(t,x))2, (t,x,r)(0,T)×Γ3×R, (2.12)

    where θR is the temperature of the foundation, and ke is the heat exchange coefficient between the body and the obstacle. We obtain

    Ξ(t,x,r)=φ(t,x,r)=ke(t,x)(rθR(t,x)), (t,x,r)(0,T)×Γ3×R.

    Finally in (2.9) and (2.10), ξ0,u0,v0 represent the initial temperature, displacement and velocity respectively.

    One may remark that since φu is assumed real-valued, then unilateral contact, defined by indicator functions taking infinite values, is excluded. So the body is in fixed contact with the foundation of the body according to a friction law. This is consistent with the linear heat conduction modeled in (2.6). We insist that the new feature here is that we may have the absence of the usual claimed condition in the case where meas(Γ1)=0. However, there is coerciveness with regard to the temperature by (2.7). To derive the variational formulation of the mechanical problems (2.1)–(2.10) we need additional notations. Thus, let consider V the closed subspace of H1 defined by

    V={vH1 | v=0on Γ1}U.

    We remark that the subspace V may be different or not to the whole space H1, depending on the set U of admissible contact conditions.

    On V we consider the inner product given by

    (u,v)V=(ε(u),ε(v))H+(u,v)Hu,vV,

    and let V be the associated norm, i.e.

    v2V=ε(v)2H+v2HvV.

    It follows that H1 and V are equivalent norms on V and therefore (V,V) is a real Hilbert space. Moreover, by the Sobolev's trace theorem, we have a constant C0>0 depending only on Ω, and Γ3 such that

    vL2(Γ3)  C0vVvV.

    For functional reason, it is convenient to shift the ambient temperature to zero on Γ1Γ2. We introduce for this propose θ=ξθa, by assuming θaH1(0,T;H1(Ω)). Thus we have t[0,T]:

    ξ(t)=θa(t)θ(t)=0onΓ1Γ2.

    In what follows, we use the following change of variables:

    ξ=θ+θa,ξ0=θ0+θa(0).

    Consider then the following spaces for the temperature field:

    E={ηH1(Ω), η=0on Γ1Γ2};F=L2(Ω).

    The spaces E and F, endowed with their respective canonical inner product, are Hilbert spaces.

    Identifying then H and F with their own duals, we obtain two Gelfand evolution triples (see e.g. [9] Ⅱ/A p. 416):

    VHHV,EFFE

    where the inclusions are continuous and dense.

    In the study of the mechanical problem (2.1)-(2.10), we assume that the viscosity operator A:(0,T)×Ω×SdSd, (t,x,τ)A(t,x,τ) satisfies

    {(i)A(,,τ) is measurable on (0,T)×Ω, τSd;(ii)A(t,x,) is continuous on Sd for a.e. (t,x)(0,T)×Ω;(iii)there exists mA>0  such that (A(t,x,τ1)A(t,x,τ2))(τ1τ2)mA|τ1τ2|2,τ1,τ2Sd, for a.e. (t,x)(0,T)×Ω;(iv)there exists cA0L2((0,T)×Ω;R+), cA1>0 such that |A(t,x,τ)|cA0(t,x)+cA1|τ|,τSd, for a.e. (t,x)(0,T)×Ω. (2.13)

    In this paper for every t(0,T), τSd we denote by A(t)=A(t,,) a functional which is defined on Ω×Sd and A(t)τ=A(t,,τ) some function defined on Ω.

    The elasticity operator G:(0,T)×Ω×SdSd satisfies :

    {(i)there exists LG>0 such that|G(t,x,ε1)G(t,x,ε2)|LG|ε1ε2|ε1,ε2Sd,a.e.(t,x)(0,T)×Ω;(ii)G(,,ε) is Lebesgue measurable on (0,T)×Ω,εSd;(iii)the mapping G(,,0)H. (2.14)

    We put again G(t)τ=G(t,,τ) some function defined on Ω for every t(0,T), τSd.

    The relaxation tensor B:(0,T)×Ω×SdSd, (t,x,τ)(Bijkh(t,x)τkh) satisfies

    {(i) BijkhL((0,T)×Ω);(ii) B(t)στ=σB(t)τσ,τSd, a.e. t(0,T), a.e. in Ω (2.15)

    where we denote by B(t)τ=B(t,,τ) which is defined on Ω for every t(0,T), τSd.

    We suppose the body forces and surface tractions satisfy

    f0L2(0,T;H),f2L2(0,T;L2(Γ2)d) (2.16)

    On the contact surface, the following frictional contact function

    ψu(t,w):=Γ3φu(t,w)da

    verifies

    { (i) ψu:(0,T)×VRis well defined; (ii) t(0,T)ψu(t,w) is Lebesgue measurable wV;(iii)|ψu(t,w)|c(t)+dwV, wV, a.e. t(0,T);(iv)ψu(t,) is convex on V a.e. t(0,T), (2.17)

    where d>0 is some constante and cL2(0,T;R+).

    The thermal expansion tensor Ce:(0,T)×Ω×RSd verifies

    {(i)Ce(,,ϑ) is measurable on (0,T)×Ω, ϑR;(ii)there exists Le>0 such that|Ce(t,x,ϑ1)Ce(t,x,ϑ2)|Le|ϑ1ϑ2|ϑ1,ϑ2R,a.e.(t,x)(0,T)×Ω;(iii)there exists cCe0L((0,T)×Ω;R+), cCe10 such that |Ce(t,x,ϑ)|cCe0(t,x)+cCe1|ϑ|,ϑR, for a.e. (t,x)(0,T)×Ω. (2.18)

    Here we recall the notation Ce(t,ϑ)=Ce(t,,ϑ) some function defined on Ω, for all t(0,T) and ϑR.

    The nonlinear function Kc:(0,T)×Ω×RdR satisfies :

    {(i)Kc(,,ξ) is measurable on (0,T)×Ω, ξRd;(ii)Kc(t,x,) is continuous on Rd, a.e. (t,x)(0,T)×Ω;(iii)there exists cKc0L2((0,T)×Ω;R+),cKc10,  such that |Kc(t,x,ξ)|cKc0(t,x)+cKc1|ξ|,ξRd, a.e. (t,x)(0,T)×Ω;(iv) there exists mKc>0 such that(Kc(t,x,ξ1)Kc(t,x,ξ2))(ξ1ξ2)mKc|ξ1ξ2|2,ξ1,ξ2Rd,a.e.(t,x)(0,T)×Ω;(v) there exists nKc>0 such thatKc(t,x,ξ)ξnKc|ξ|2,ξRd,a.e.(t,x)(0,T)×Ω. (2.19)

    We suppose that the nonlinear function Ce:(0,T)×Ω×RdR satisfies :

    {(i)Ce(,,v) is measurable on (0,T)×Ω, vRd;(ii)there exists LCe>0 such that|Ce(t,x,v1)Ce(t,x,v2)|LCe|v1v2|,v1,v2Rd,a.e.(t,x)(0,T)×Ω. (2.20)

    We assume for the heat sources density, that

    qL2(0,T;L2(Ω)) (2.21)

    The nonlinear function Ξ,φ:(0,T)×Γc×RR verifies :

    {(i)Ξ(,,r),φ(,,r) are measurable on (0,T)×Γc, rR;(ii)φ(t,x,) is locally Lipschitz on on R for a.e. (t,x)(0,T)×Γc;(iii)there exists cφ0L2((0,T)×Γc;R+),cφ10,  such that |φ0(t,x,r)|cφ0(t,x)+cφ1|r|,rR, a.e. (t,x)(0,T)×Γc;(iv)(Ξ(t,x,r1)Ξ(t,x,r2))(r1r2)0,r1,r2R, a.e. (t,x)(0,T)×Γc. (2.22)

    We notice that these assumptions are verified for the example (2.12).

    Finally we assume that the initial data satisfy the conditions

    u0  H,v0  V,θ0E. (2.23)

    To continue, using Green's formula, we obtain the variational formulation of the mechanical problem Q in abstract form as follows.

    Problem QV : Find u:[0,T]V, θ:[0,T]E satisfying a.e. t(0,T):

    {¨u(t)+A(t)˙u(t)+B(t)u(t)+C(t)θ(t),w˙u(t)V×V+(t0B(ts)ε(u(s))ds,ε(w)ε(˙u(t)))H+ψu(t,w)ψu(t,˙u(t))f(t),w˙u(t)V×VwV. (2.24)
    {<˙θ(t),η>E×E+<K(t)θ(t),η>E×E+ψ(t,θ(t);η)<R(t)˙u(t),η>E×E+<Q(t),η>E×E, ηE. (2.25)
    u(0)=u0,˙u(0)=v0,θ(0)=θ0. (2.26)

    Here, the operators and functions A(t),B(t):VV, C(t):EV, f:[0,T]V, K(t):EE, ψ(t,;):E×ER, R(t):VE and Q:[0,T]E are defined by vV, wV, τE, ηE, a.e. t(0,T):

    A(t)v,wV×V=(A(t)εv,εw)H; (2.27)
    B(t)v,wV×V=(G(t)εv,εw)H; (2.28)
    C(t)ζ,wV×V=(Ce(t,ζ()+θa(t)),εw)H; (2.29)
    f(t),wV×V=(f0(t),w)H+(fF(t),w)(L2(Γ2))d; (2.30)
    K(t)ζ,ηE×E=ΩKc(t,ζ+θa(t))ηdx; (2.31)
    ψ(t,ζ;η)=Γ3φ0(t,x,ζ(x)+θa(t);η(x))da(x); (2.32)
    R(t)v,ηE×E=ΩCe(t,v)ηdx; (2.33)
    Q(t),ηE×E=Ω(q(t)˙θa(t))ηdx. (2.34)

    We verify that from (2.22) then the term ψ(t,ζ;η) is well defined for all ζE, ηE, for a.e. t(0,T).

    The inequality (2.25) is a consequence of the equation below,

    {<˙θ(t),η>E×E+<K(t)θ(t),η>E×E+Γ3Ξ(t,θ(t)+θa(t))ηda=<R(t)˙u(t),η>E×E+<Q(t),η>E×E, ηE, (2.35)

    where recall that Ξ(t,r):=Ξ(t,,r) defined on Γ3, for (t,r)(0,T)×R.

    Our main existence and uniqueness result is stated as follows, that we prove in the next Section.

    Theorem 2.1. Assume that t (2.13)–(2.23) hold, then there exists an unique solution {u,θ} to problem QV with the regularity :

    {uW1,2(0,T;V)W2,2(0,T;V)C1(0,T;H)θL2(0,T;E)W1,2(0,T;E)C(0,T;F). (2.36)

    3. roof of Theorem 2.1

    The idea is to bring the second order inequality to a first order inequality, using monotone operator, convexity and fixed point arguments, and will be carried out in several steps.

    Lvelocity variable

    v=˙u.

    The system in Problem QV is then written for a.e. t(0,T):

    {u(t)=u0+t0v(s)ds;˙v(t)+A(t)v(t)+B(t)u(t)+C(t)θ(t),wv(t)V×V+(t0B(ts)ε(u(s))ds,ε(w)ε(v(t)))H+ψu(t,w)ψu(t,v(t))f(t),wv(t)V×VwV;<˙θ(t),η>E×E+<K(t)θ(t),η>E×E+ψ(t,θ(t);η)<R(t)˙u(t),η>E×E+<Q(t),η>E×E, ηE;v(0)=v0,θ(0)=θ0,

    with the regularity

    {vvL2(0,T;V)W1,2(0,T;V)C(0,T;H)θL2(0,T;E)W1,2(0,T;E)C(0,T;F).

    To continue, we assume in the sequel that the conditions (2.13)–(2.17) of the Theorem 1 are satisfied. Let define

    W:=L2(0,T;H).

    We begin by

    Lemma 1. For all ηW, there exists an unique

    vηL2(0,T;V)W1,2(0,T;V)C(0,T;H)

    satisfying

    {˙vη(t)+A(t)vη(t),wvη(t)V×V+(η(t),ε(w)ε(vη(t)))H+ψu(t,w)ψu(t,vη(t))f(t),wvη(t)V×V,wV, a.e. t(0,T);vη(0)=v0. (3.1)

    Moreover, c>0 such that η1,η2W:

    vη2(t)vη1(t)2H+t0vη1vη22Vct0η1η22H,t[0,T]. (3.2)

    Proof. Let ηW. Using [9] Ⅱ/B p. 893, we deduce the existence and uniqueness of vη.

    Now let η1,η2W. In (3.1) we take (η=η1,w=vη2(t)), then (η=η2,w=vη1(t)). Adding the two inequalities, we deduce that for a.e. t(0;T):

    ˙vη2(t)˙vη1(t),vη2(t)vη1(t)V×V+A(t)vη2(t)A(t)vη1(t),vη2(t)vη1(t)V×V(η2(t)η1(t),ε(vη2(t))ε(vη1(t)))H.

    Then integrating over (0,t), from (2.13)(ⅲ) and from the initial condition on the velocity, we obtain:

    t[0,T],vη2(t)vη1(t)2H+mAt0vη2(s)vη1(s)2Vdst0(η2(s)η1(s),ε(vη2(s))ε(vη1(s)))Hds+mAt0vη2(s)vη1(s)2Hds.

    We conclude that c>0 such that η1,η2W, t[0,T]:

    vη2(t)vη1(t)2H+t0vη1(s)vη2(s)2Vdsct0η1(s)η2(s)2Hds+ct0vη2(s)vη1(s)2Hds. (3.3)

    Now let fix τ[0,T]. We have t[0,τ]:

    vη2(t)vη1(t)2Hcτ0η1(s)η2(s)2H+ct0vη2(s)vη1(s)2Hds.

    Using then Gronwall's inequality, we obtain τ[0,T]:

    vη2(τ)vη1(τ)2H(cτ0η1(s)η2(s)2H)ecT.

    Finally, integrating the last inequality and reporting the result in (3.3), we get (3.2).

    Here and below, we denote by c>0 a generic constant, which value may change from lines to lines.

    Lemma 2. For all ηW, thereexists an unique

    θηL2(0,T;E)W1,2(0,T;E)C(0,T;F)

    satisfying

    {<˙θη(t),η>E×E+<K(t)θη(t),η>E×E+ΓcΞ(t,θη(t))ηda=<R(t)vη(t),η>E×E+<Q(t),η>E×E, ηE,  a.e. t(0,T);θη(0)=θ0. (3.4)

    Moreover, c>0 such that η1,η2W:

    θη1(t)θη2(t)2Fct0vη1vη22V,t[0,T]. (3.5)

    Proof. The existence and uniqueness result verifying (3.3) follows from standard result on first order evolution equation (see e.g. [5]). Indeed we verify that from the expression of the operator R, we have

    \mathit{\boldsymbol{v}}_\eta \in L^{2}(0, T;V) \Longrightarrow R\, \mathit{\boldsymbol{v}}_\eta \in L^{2}(0, T;E'),

    as Q \in L^{2}(0, T; E') then R\, \mathit{\boldsymbol{v}}_\eta + Q \in L^{2}(0, T; E').

    Using the assumptions (2.19) and (2.22), the operator \Psi(t)\, :\, E \longrightarrow E' for a.e. t\in (0, T) defined by

    <\Psi(t)\, \xi, \eta>_{E'\times E} := <K(t)\, \xi, \eta>_{E'\times E} + \int_{\Gamma_3}\, \Xi(t, \xi) \eta \, da, \quad \forall\, \xi, \, \eta\, \in E

    is strongly monotone.

    Now for \eta_1, \, \eta_2 \in \mathcal{W}, we have for a.e. t\in (0;T):

    \begin{array}l \left\langle { \dot{\theta}_{\eta_1}(t)-\dot{\theta}_{\eta_2}(t), {\theta}_{\eta_1}(t) - {\theta}_{\eta_2}(t)} \right\rangle_{E'\times E} + \left\langle { K(t)\, {\theta}_{\eta_1}(t)- K(t)\, {\theta}_{\eta_2}(t), \, {\theta}_{\eta_1}(t)-{\theta}_{\eta_2}(t) } \right\rangle _{E'\times E} \\ = \left\langle {R(t)\, {\mathit{\boldsymbol{v}}}_{\eta_1}(t) - R(t)\, {\mathit{\boldsymbol{v}}}_{\eta_2}(t), \, {\theta}_{\eta_1}(t)-{\theta}_{\eta_2}(t) } \right\rangle _{E'\times E}. \end{array}

    Then integrating the last property over (0, t), using the strong monotonicity of K(t) and the Lipschitz continuity of R(t)\, :\, V \longrightarrow E', we deduce (3.5).

    Proof of Theorem 1.

    We have now all the ingredients to prove the Theorem 1.

    Consider the operator \Lambda\, :\, \mathcal{W} \to \mathcal{W} defined by for all \eta \in \mathcal{W}:

    \Lambda\, \eta\, (t) = {\mathcal G}(\mathit{\boldsymbol{\varepsilon }} (\mathit{\boldsymbol{u}}_{\eta}(t)) ) + \int_0^t\, B(t-s)\, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}_{\eta}(s))\, ds + C_e(t, {\theta}_{\eta}(t)), \quad \forall t \in [0, T],

    where

    \mathit{\boldsymbol{u}}_{\eta}(t) = \mathit{\boldsymbol{u}}_0+ \int_0^t\, {\mathit{\boldsymbol{v}}}_{\eta}(s)\, ds, \ \forall t \in [0, T]; \quad \mathit{\boldsymbol{u}}_{\eta} \in W^{1, 2}(0, T;V) \cap W^{2, 2}(0, T;V') \cap C^1(0, T;H).

    Then from (2.14), (2.15), and Lemma 2, we deduce that for all \eta_1, \, \eta_2 \in \mathcal{W}, for all t\in [0, T]:

    \label{ea1} \begin{array}l \| \Lambda\, \eta_1\, (t)-\Lambda\, \eta_2\, (t) \|_{\mathcal{H}}^2 \leq c\, \| {\theta}_{\eta_1}(t) - {\theta}_{\eta_2}(t) \|_F^2 + c\, \int_0^t\, \|{\mathit{\boldsymbol{v}}}_{\eta_1}(s)-{\mathit{\boldsymbol{v}}}_{\eta_2}(s)\|_V^2 \, ds \\ \leq c\, \int_0^t\, \|{\mathit{\boldsymbol{v}}}_{\eta_1}(s)-{\mathit{\boldsymbol{v}}}_{\eta_2}(s)\|_V^2 \, ds. \end{array} (3.6)

    Now using (3.6), after some algebraic manipulations, we have for any \beta >0:

    \int_0^T\, e^{-\beta \tau}\, \| \Lambda\, \eta_1\, (\tau)-\Lambda\, \eta_2\, (\tau) \|_{\mathcal{H}}^2 \leq \frac{c}{\beta}\, \int_0^T\, e^{-\beta \tau}\, \| \eta_1(\tau)- \eta_2(\tau) \|_{\mathcal{H}}^2 \, d\tau.

    We conclude from the last inequality by contracting principle that the operator \Lambda has a unique fixed point \eta^* \in \mathcal{W}. We verify then that the functions

    {\mathit{\boldsymbol{u}}}(t) := \mathit{\boldsymbol{u}}_0 + \int_0^t\, {\mathit{\boldsymbol{v}}}_{\eta^*}, \ \forall t\in [0, T], \quad \theta := \theta_{\eta^*}

    are solutions to problem QV with the regularity (2.36), the uniqueness follows from the uniqueness in Lemma 1 and Lemma 2.


    4. Analysis of a numerical scheme

    In this section, we study a fully-discrete numerical approximation scheme of the variational problem QV. For this purpose, let \{ \mathit{\boldsymbol{u}}, \theta \} be the unique solution of the problem QV, and introduce the velocity variable

    \mathit{\boldsymbol{v}}(t)=\dot{\mathit{\boldsymbol{u}}}(t), \quad \forall t\in [0, T].

    Then

    \mathit{\boldsymbol{u}}(t)= \mathit{\boldsymbol{u}}_0 + \int_0^t \mathit{\boldsymbol{v}}(s)\, ds, \quad \forall t\in [0, T].

    Here we make the following additional assumptions on the different data, operators and solution fields:

    A \in C([0, T]; \mathcal{L}(V, V')), \\ B \in C([0, T]; C(V, V')), \\ C \in C([0, T]; \mathcal{L}(V, V')), \\ \mathcal{B} \in C([0, T]; \mathcal{L}( \mathcal{H}, \mathcal{H})), \\ \psi_u \in C([0, T]; C(V, \mathbb{R})), \\ \psi \in C([0, T]; C(E\times E, \mathbb{R})), \\ \mathit{\boldsymbol{f}} \in C([0, T];V'), \\ K \in C([0, T]; \mathcal{L}(E, E')), \\ R \in C([0, T]; \mathcal{L}(E, E')), \\ Q \in C([0, T];E'), \\ \mathit{\boldsymbol{v}} \in C([0, T];V)\cap C^1([0, T];H), \\ \theta \in C([0, T];E)\cap C^1([0, T];F). (4.1)

    Moreover we assume that for all r, r_1, r_2 \in \mathbb{R}, a.e. (t, \mathit{\boldsymbol{x}}) \in(0, T)\times \Gamma_c :

    \begin{equation} \left\{ \begin{array}{l} {\rm (i)} \; \varphi'(t, \mathit{\boldsymbol{x}}, r; r_1+r_2) \leq \varphi'(t, \mathit{\boldsymbol{x}}, r; r_1) + \varphi'(t, \mathit{\boldsymbol{x}}, r; r_2);\\ {\rm (ii)} \; \varphi'(t, \mathit{\boldsymbol{x}}, r_2; r_1-r_2) + \varphi'(t, \mathit{\boldsymbol{x}}, r_1; r_2-r_1) \leq 0;\\ {\rm (iii)} \; \mbox{there exists } c^{\varphi} \geq 0 \ \mbox{ such that } \\[1mm] \qquad \varphi'(t, \mathit{\boldsymbol{x}}, r_1; r) + \varphi'(t, \mathit{\boldsymbol{x}}, r_2; -r) \leq c^{\varphi}\, |(r_1 - r_2)\, r|. \end{array} \right. \end{equation} (4.2)

    We remark that the example of \varphi given in (2.12) satisfies the hypotheses (4.2).

    From Theorem 1, \{ \mathit{\boldsymbol{v}}, \theta \} verify for all t\in [0, T]:

    \begin{equation} \label{4.1} \left\{ \begin{array}l \left\langle {\dot{\mathit{\boldsymbol{v}}}(t)+ A(t)\, {\mathit{\boldsymbol{v}}}(t)+ B(t)\, \mathit{\boldsymbol{u}}(t) + C(t)\, \theta(t), \mathit{\boldsymbol{w}} - \mathit{\boldsymbol{v}}(t) } \right\rangle_{V'\times V} \\ + (\int_0^t\, \mathcal{B}(t-s)\, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}(s))\, ds, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{w}})- \mathit{\boldsymbol{\varepsilon }}(\dot{\mathit{\boldsymbol{u}}}(t)))_{\mathcal H} + \psi_u(t, \mathit{\boldsymbol{w}}) - \psi_u(t, \mathit{\boldsymbol{v}}(t)) \\ \qquad \geq \left\langle {\mathit{\boldsymbol{f}}(t), \, \mathit{\boldsymbol{w}} - \mathit{\boldsymbol{v}}(t)} \right\rangle_{V'\times V}, \quad \forall\, \mathit{\boldsymbol{w}}\in V. \end{array} \right. \end{equation} (4.3)
    \begin{equation} \label{4.2} \begin{array}l <\dot{\theta}(t), \eta>_{E'\times E} + <K(t)\, \theta(t), \eta>_{E'\times E} +\psi(t, \theta(t);\eta) \\[1mm] \qquad \geq <R(t) \dot{\mathit{\boldsymbol{u}}}(t), \eta>_{E'\times E} + <Q(t), \eta>_{E'\times E}, \ \forall\, \eta\, \in E. \end{array} \end{equation} (4.4)
    \mathit{\boldsymbol{u}}(0)=\mathit{\boldsymbol{u}}_0, \quad \mathit{\boldsymbol{v}}(0) = \mathit{\boldsymbol{v}}_0, \quad \theta(0)=\theta_0, (4.5)

    Now let V^h\subset V and E^h\subset E be a family of finite dimensional subspaces, with h>0 a discretization parameter. We divide the time interval [0, T] into N equal parts: t_n=n\, k, n=0, 1, \dots, N, with the time step k=T/N.

    For a continuous operator or function U \in C([0, T]; X) with values in a space X, we use the notation U_n = U(t_n)\in X.

    Then from (4.3) and (4.4) we introduce the following fully discrete scheme.

    Problem P^{hk}. Find \mathit{\boldsymbol{v}}^{hk}=\{\mathit{\boldsymbol{v}}^{hk}_n\}_{n=0}^N \subset V^h, \theta^{hk}=\{\theta^{hk}_n\}_{n=0}^N \subset E^h such that

    \mathit{\boldsymbol{v}}^{hk}_0=\mathit{\boldsymbol{v}}_0^h, \quad \theta^{hk}_0=\theta_0^h (4.6)

    and for n=1, \cdots, N,

    \begin{equation} \label{4.4} \left\{ \begin{array}l \Big( \frac{\mathit{\boldsymbol{v}}^{hk}_n-\mathit{\boldsymbol{v}}^{hk}_{n-1}}{k}, \, \mathit{\boldsymbol{w}}^h - \mathit{\boldsymbol{v}}^{hk}_n \Big)_H + \left\langle {A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \mathit{\boldsymbol{w}}^h - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} +\left\langle {B_n\, \mathit{\boldsymbol{u}}_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} \\ + \left\langle {C_n\, \theta_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} + \psi_u(t_n, \mathit{\boldsymbol{w}}^h) - \psi_u(t_n, \mathit{\boldsymbol{v}}^{hk}_n) \\ + ( k\, \sum\limits_{m=0}^{n-1} \mathcal{B}(t_n-t_m) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}^{hk}_m), \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{w}}^h) - \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{v}}^{hk}_n) )_{\mathcal{H}} \\ \qquad \geq \left\langle { \mathit{\boldsymbol{f}}_n, \, \mathit{\boldsymbol{w}}^h - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{V^\prime\times V}, \quad \forall\, \mathit{\boldsymbol{w}}^h\in V^h. \end{array} \right. \end{equation} (4.7)
    \begin{equation} \label{4.5} \left\{ \begin{array}{l} \Big(\frac{\theta^{hk}_n-\theta^{hk}_{n-1}}{k}, \, \eta^h \Big)_{L^2(\Omega)} + \left\langle {K_n\, \theta^{hk}_n, \, \eta^h} \right\rangle_{ E^\prime\times E} +\psi(t_n, \theta^{hk}_n;\eta^h) \\ \qquad \geq \left\langle {R_n \, \mathit{\boldsymbol{v}}^{hk}_n, \, \eta^h} \right\rangle_{ E^\prime\times E} + \left\langle {Q_n, \eta^h} \right\rangle_{ E^\prime\times E}, \quad \forall\, \eta^h\in E^h. \end{array} \right. \end{equation} (4.8)

    where

    \mathit{\boldsymbol{u}}_{n}^{hk} = \mathit{\boldsymbol{u}}_{n-1}^{hk} + k\, \mathit{\boldsymbol{v}}^{hk}_n, \quad \mathit{\boldsymbol{u}}^{hk}_0 = \mathit{\boldsymbol{u}}_0^h. (4.9)

    Here \mathit{\boldsymbol{u}}^h_0 \in V^h , \mathit{\boldsymbol{v}}^h_0 \in V^h , \theta^{h}_0 \in E^h are suitable approximations of the initial values \mathit{\boldsymbol{u}}_0, \mathit{\boldsymbol{v}}_0, \theta_0.

    For n=1, \dots, N, suppose that \mathit{\boldsymbol{u}}^{hk}_{n-1}, \mathit{\boldsymbol{v}}^{hk}_{n-1}, \theta^{hk}_{n-1} are known, then by using standard result on elliptic variational inequalities of the second kind (see e.g. [9] p. 892), we calculate \mathit{\boldsymbol{v}}^{hk}_{n} by (4.7), \theta^{hk}_{n} by (4.8) and \mathit{\boldsymbol{u}}^{hk}_{n} by (4.9). Hence the discrete solution \mathit{\boldsymbol{v}}^{hk}\subset V^h, \theta^{hk}\subset E^h exists and is unique.

    We now turn to an error analysis of the numerical solution. The main result of this section is the following.

    Theorem 2. We keep the assumptions of Theorem 1. Under the additional assumptions (4.1), then for the unique solution \mathit{\boldsymbol{v}}^{hk}\subset V^h, \theta^{hk} \subset E^h of the discrete problem P^{hk}, we have the following error estimate

    \begin{equation} \label{est} \begin{array}l \max\limits_{1\le n\le N}{\|}\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}{\|}_H^2 +\Big(k\sum\limits_{n=1}^N \|\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}\|_V^2\Big) \\ \quad + \max\limits_{1\le n\le N}{\|}\theta_n-\theta_n^{hk}{\|}_{F}^2 +\Big(k\sum\limits_{n=1}^N \|\theta_n-\theta_n^{hk}\|_E^2\Big) \\ \le \ c\, \|\mathit{\boldsymbol{u}}_0-\mathit{\boldsymbol{u}}_0^{h}\|_V^2 + c\, {\|}\mathit{\boldsymbol{v}}_0-\mathit{\boldsymbol{v}}_0^{hk}{\|}_H^2 + c\, \|\theta_0-\theta_0^{h}\|_{F}^2 \\ \quad + c\, \max\limits_{1\le n\le N} \| \mathit{\boldsymbol{v}}_n - \mathit{\boldsymbol{w}}_n^{h} \|_{H} + c\, \max\limits_{1\le n\le N}{\|}\theta_n-\eta_n^{h}{\|}_{F}^2 \\ \quad +c\, k\sum\limits_{j=1}^N \|\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j \|_V^2 + c\, k\, \sum\limits_{j=1}^N{\|}\theta_j-\eta_j^{h}{\|}_{E}^2 \\ \quad + c\, \Big( \sum\limits_{j=1}^{N-1}{\|}(\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j) -(\mathit{\boldsymbol{v}}_{j+1}-\mathit{\boldsymbol{w}}^h_{j+1}){\|}_H \Big)^2 \\ \quad + c\, \Big( \sum\limits_{j=1}^{N-1} \| \theta_j-\eta_j^{h} - ( \theta_{j+1}-\eta_{j+1}^{h}) \|_{F} \Big)^2 \\ \quad + c\, k^2 + c\, k\, \sum\limits_{j=1}^N \, \| \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j \|_V, \end{array} \end{equation} (4.10)

    where for j=1, \dots, N, \mathit{\boldsymbol{w}}^h_j\in V^h, \eta^h_j\in E^h are arbitrary.

    Proof. The method takes again and generalize classical techniques used in [2], where we refer for details. Here we mention only the main steps of the proof.

    Denote

    \mathit{\boldsymbol{e }}_n=\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}^{hk}_n, \quad \varepsilon_n=\theta_n-\theta^{hk}_n, \quad 0\le n\le N,

    the numerical solution errors.

    We begin by the estimate of (\varepsilon_n).

    Let fix n = 1, \cdots, N. From (4.4) where we put t=t_n and \eta = -\eta^h , \eta^h\in E^h, then add to (4.8), we obtain

    \begin{array}l \Big( \dot{\theta}_n - \frac{\theta^{hk}_n-\theta^{hk}_{n-1}}{k}, \, \eta^h \Big)_{L^2(\Omega)} + \left\langle {K_n\, \theta_n - K_n\, \theta^{hk}_n, \, \eta^h} \right\rangle_{ E^\prime\times E}\\ \quad \leq \psi(t_n, \theta_n; -\eta^h ) + \psi(t_n, \theta_n^{hk}; \eta^h ) + \left\langle {R_n\, \mathit{\boldsymbol{v}}_n - R_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \eta^h} \right\rangle_{E^\prime\times E}. \end{array}

    Writing then

    \dot{\theta}_n - \frac{\theta^{hk}_n-\theta^{hk}_{n-1}}{k} = \dot{\theta}_n - \frac{\theta_n - \theta_{n-1}}{k} + \frac{ \varepsilon_n - \varepsilon_{n-1}}{k},

    and replacing \eta^h by \eta^h_n -\theta_n + \varepsilon_n we obtain

    \begin{array}l \Big( \frac{ \varepsilon_n - \varepsilon_{n-1}}{k}, \, \varepsilon_n \Big)_{L^2(\Omega)} + \left\langle {K_n\, \theta_n - K_n\, \theta^{hk}_n, \, \varepsilon_n} \right\rangle_{ E^\prime\times E} \\ \qquad \leq \left\langle {K_n\, \theta_n - K_n\, \theta^{hk}_n, \, \theta_n - \eta^h_n} \right\rangle_{ E^\prime\times E} \\ \qquad \quad + \left\langle {R_n\, \mathit{\boldsymbol{v}}_n - R_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \varepsilon_n } \right\rangle_{ E^\prime\times E} + \left\langle {R_n\, \mathit{\boldsymbol{v}}_n -\mathit{\boldsymbol{v}}^{hk}_n, \, \theta_n -\eta^h_n } \right\rangle_{ E^\prime\times E} \\ \qquad \quad + \Big( \dot{\theta}_n - \frac{\theta_n-\theta_{n-1}}{k} +\frac{\varepsilon_n- \varepsilon_{n-1}}{k}, \, \theta_n - \eta^h_n \Big)_{L^2(\Omega)} - \Big( \dot{\theta}_n - \frac{\theta_n-\theta_{n-1}}{k}, \, \varepsilon_n \Big)_{L^2(\Omega)} \\ \qquad \quad + \psi(t_n, \theta_n; -\eta^h ) + \psi(t_n, \theta_n^{hk}; \eta^h ). \end{array}

    From (2.19) we have

    | \left\langle {K_n\, \theta_n - K_n\, \theta^{hk}_n, \, \theta_n - \eta^h_n} \right\rangle_{ E^\prime\times E} | \leq c\, \| \theta_n - \theta^{hk}_n\|_E \times \| \theta_n - \eta^h_n\|_E.

    From (2.20) we have

    | \left\langle {R_n\, \mathit{\boldsymbol{v}}_n - R_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \eta^h} \right\rangle_{E^\prime\times E} | \leq c\, \| \mathit{\boldsymbol{v}}_n - \mathit{\boldsymbol{v}}^{hk}_n \|_{L^2(\Omega)} \times \| \eta^h \|_{L^2(\Omega)}.

    Then using (4.2) we obtain

    \psi(t_n, \theta_n; -\eta^h ) + \psi(t_n, \theta_n^{hk}; \eta^h ) \leq c\, \| \theta_n - \theta^{hk}_n \|_E\, \times \| \theta_n - \eta^h_n \|_E.

    Consider the quantity for n = 1, \cdots, N ,

    \Xi_n := \Big( \frac{ \varepsilon_n - \varepsilon_{n-1}}{k}, \, \varepsilon_n \Big)_{F} + \left\langle {K_n\, \theta_n - K_n\, \theta^{hk}_n, \, \varepsilon_n} \right\rangle_{ E^\prime\times E}.

    We have for some c_K > 0,

    \Xi_n \geq \frac{1}{2k}\, \Big( \| \varepsilon_n \|_F^2 - \| \varepsilon_{n-1} \|_F^2 \Big) + c_K\, \| \varepsilon_n \|_E^2.

    Summing then \Xi_j from j=1 to j=n , and after some manipulation, we obtain

    \begin{array}l \frac{1}{2k}\, \Big( \| \varepsilon_n \|_F^2 - \| \varepsilon_0 \|_F^2 \Big) + \sum\limits_{j=1}^n\, \| \varepsilon_j \|_E^2\\ \qquad \leq c\, \sum\limits_{j=1}^n\, \Big\| \dot{\theta}_j - \frac{\theta_j-\theta_{j-1}}{k} \Big\|_F^2 + c\, \sum\limits_{j=1}^n\, \| \theta_j -\eta^h_j \|_E^2 \\ \qquad \quad + c\, \sum\limits_{j=1}^n\, \| \mathit{\boldsymbol{e }}_j \|_V^2 + c\, \sum\limits_{j=1}^n\, \Big( \frac{\varepsilon_j-\varepsilon_{j-1}}{k}, \, \theta_j -\eta^h_j \Big)_{F}. \end{array}

    Denote now by

    \begin{array}l M_\varepsilon := \max\limits_{1\le n\le N} {\|}\varepsilon_n{\|}_{F}, \\ AT_0 := \max\limits_{1\le n\le N}{\|}\theta_n-\eta_n^{h}{\|}_{F}, \\ AT_1 := \sum\limits_{j=1}^N {\Big\|} \dot{\theta}_j - \frac{\theta_j-\theta_{j-1}}{k} {\Big\|}_{F}^2, \\ AT_2 := \sum\limits_{j=1}^N{\|}\theta_j-\eta_j^{h}{\|}_{E}^2, \\ AT_3 := \sum\limits_{j=1}^N{\|}\theta_j-\eta_j^{h} - ( \theta_{j+1}-\eta_{j+1}^{h}) {\|}_{F}. \end{array}

    We deduce from the last inequality that for some constant c>0 and for n=1, \cdots, N,

    \begin{equation} \label{estep} \begin{array}l \| \varepsilon_n {\|}_{F}^2+ k\sum\limits_{j=1}^n \|\varepsilon_j \|_E^2 \leq \ c\, {\|}\varepsilon_0{\|}_{F}^2 + c\, AT_0^2 + c\, k\, (AT_1 + AT_2) \\ + c\, AT_3\, M_\varepsilon + c\, k\, \sum\limits_{j=1}^n \, \| \mathit{\boldsymbol{e }}_{j} \|_{V}^2. \end{array} \end{equation} (4.11)

    We now turn to the estimate of (\mathit{\boldsymbol{e }}_n).

    Let fix n = 1, \cdots, N. Using (4.3) where we put t=t_n and \mathit{\boldsymbol{w}} = \mathit{\boldsymbol{v}}_n^{hk}, and adding to (4.7) where \mathit{\boldsymbol{w}}^h = \mathit{\boldsymbol{w}}_n^h, we have

    \begin{array}l ( \dot{\mathit{\boldsymbol{v}}}_n, \, \mathit{\boldsymbol{v}}^{hk}_n - \mathit{\boldsymbol{v}}_n )_H + \Big( \frac{\mathit{\boldsymbol{v}}^{hk}_n-\mathit{\boldsymbol{v}}^{hk}_{n-1}}{k}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}^{hk}_n \Big)_H \\ + \left\langle {A_n\, \mathit{\boldsymbol{v}}_n, \, \mathit{\boldsymbol{v}}^{hk}_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} + \left\langle {A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} \\ + \left\langle {B_n\, \mathit{\boldsymbol{u}}_n^{hk}, \, \mathit{\boldsymbol{v}}^{hk}_n -\mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} + \left\langle {B_n\, \mathit{\boldsymbol{u}}_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} \\ + \left\langle {C_n\, \theta_n, \, \mathit{\boldsymbol{v}}^{hk}_n -\mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} + \left\langle {C_n\, \theta_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}^{hk}_n} \right\rangle_{ V^\prime\times V} \\ + \mathcal{R}_n^{hk} + \psi_u(t_n, \mathit{\boldsymbol{w}}^h_n) - \psi_u(t_n, \mathit{\boldsymbol{v}}_n) \geq \left\langle {\mathit{\boldsymbol{f}}_n, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{V^\prime\times V}, \quad \forall\, \mathit{\boldsymbol{w}}^h_n\in V^h, \end{array}

    where for n = 1, \cdots, N:

    \begin{array}l \mathcal{R}_n^{hk} = \Big( \int_0^{t_n}\, \mathcal{B}(t_n-s) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}(s))\, ds - k\, \sum\limits_{m=0}^{n-1} \mathcal{B}(t_n-t_m) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}^{hk}_m), \, - \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{e }}_n) \Big)_{\mathcal{H}} \\ \qquad \quad + \Big( k\, \sum\limits_{m=0}^{n-1} \mathcal{B}(t_n-t_m) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}^{hk}_m), \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{w}}^h_n) - \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{v}}_n) \Big)_{\mathcal{H}}. \end{array}

    Writing then

    \left\{ \begin{array}l \dot{\mathit{\boldsymbol{v}}}_n = \dot{\mathit{\boldsymbol{v}}}_n - \frac{\mathit{\boldsymbol{v}}^{hk}_n-\mathit{\boldsymbol{v}}^{hk}_{n-1}}{k} + \frac{\mathit{\boldsymbol{v}}^{hk}_n-\mathit{\boldsymbol{v}}^{hk}_{n-1}}{k};\\ A_n\, \mathit{\boldsymbol{v}}_n = A_n\, \mathit{\boldsymbol{v}}_n - A_n\, \mathit{\boldsymbol{v}}^{hk}_n + A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \end{array} \right.

    we obtain

    \begin{array}l \Big( \frac{\mathit{\boldsymbol{e }}_n-\mathit{\boldsymbol{e }}_{n-1}}{k}, \, \mathit{\boldsymbol{e }}_n \Big)_H + \left\langle {A_n\, \mathit{\boldsymbol{v}}_n - A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \mathit{\boldsymbol{e }}_n} \right\rangle_{ V^\prime\times V} \\ \leq \left\langle {A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} \\ \quad + \Big( \dot{\mathit{\boldsymbol{v}}}_n - \frac{\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_{n-1}}{k}, \, - \mathit{\boldsymbol{e }}_n \Big)_H + \Big( \frac{\mathit{\boldsymbol{v}}^{hk}_n-\mathit{\boldsymbol{v}}^{hk}_{n-1}}{k}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n \Big)_H \\ \quad + \left\langle {B_n\, \mathit{\boldsymbol{u}}_n - B_n\, \mathit{\boldsymbol{u}}_{n-1}^{hk}, \, -\mathit{\boldsymbol{e }}_n} \right\rangle_{ V^\prime\times V} + \left\langle {B_n\, \mathit{\boldsymbol{u}}_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} \\ \quad + \left\langle {C_n\, \theta_n - C_n\, \theta_{n-1}^{hk}, \, -\mathit{\boldsymbol{e }}_n} \right\rangle_{ V^\prime\times V} + \left\langle {C_n\, \theta_{n-1}^{hk}, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{ V^\prime\times V} \\ \quad + \mathcal{R}_n^{hk} + \psi(t_n, \mathit{\boldsymbol{w}}^h_n) - \psi_u(t_n, \mathit{\boldsymbol{v}}_n) - \left\langle {\mathit{\boldsymbol{f}}_n, \, \mathit{\boldsymbol{w}}^h_n - \mathit{\boldsymbol{v}}_n} \right\rangle_{V^\prime\times V} \end{array}

    Consider the quantity for n = 1, \cdots, N ,

    E_n=\Big(\frac{\mathit{\boldsymbol{e }}_n-\mathit{\boldsymbol{e }}_{n-1}}{k}, \mathit{\boldsymbol{e }}_n \Big)_H + \left\langle {A_n\, \mathit{\boldsymbol{v}}_n -A_n\, \mathit{\boldsymbol{v}}^{hk}_n, \mathit{\boldsymbol{e }}_n} \right\rangle_{ V^\prime\times V}.

    We have

    E_n \geq \frac{1}{2\, k}\, \big({\|}\mathit{\boldsymbol{e }}_n{\|}_H^2-{\|}\mathit{\boldsymbol{e }}_{n-1}{\|}_H^2\big) +m_{\mathcal{A}}\, (\|\mathit{\boldsymbol{e }}_n\|_V^2 - \|\mathit{\boldsymbol{e }}_n\|_H^2 ).

    For N large enough (recall that k= \frac{T}{N} ), we have \frac{1}{2\, k} -m_{\mathcal{A}} \geq 1, then

    E_n\ge \frac{1}{2\, k}\, \big({\|}\mathit{\boldsymbol{e }}_n{\|}_H^2-{\|}\mathit{\boldsymbol{e }}_{n-1}{\|}_H^2\big) +m_{\mathcal{A}}\, \|\mathit{\boldsymbol{e }}_n\|_V^2.

    To continue, we denote in the sequel by

    \begin{array}l M_e := \max\limits_{1\le n\le N} {\|}\mathit{\boldsymbol{e }}_n{\|}_H, \\ BV_0 := \max\limits_{1\le n\le N}{\|}\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{w}}_n^{h}{\|}_{H}, \\ BV_1 := \sum\limits_{j=1}^N {\Big\|}\frac{\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{v}}_{j-1}}{k} -\dot{\mathit{\boldsymbol{v}}}_j{\Big\|}_H^2, \\ BV_2 := \sum\limits_{j=1}^N \|\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j\|_V^2; \quad \widehat{BV}_2 := \sum\limits_{j=1}^N \|\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j\|_V, \\ BV_3 := \sum\limits_{j=1}^{N-1}{\|}(\mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j) -(\mathit{\boldsymbol{v}}_{j+1}-\mathit{\boldsymbol{w}}^h_{j+1}){\|}_H, \\ I := \sum\limits_{j=1}^{N-1}\, \Big\| \int_0^{t_j}\, \mathit{\boldsymbol{v}} - k\, \sum\limits_{i=1}^{j} \mathit{\boldsymbol{v}}_i \Big\|_V^2, \\ BV_4 := \sum\limits_{j=1}^N\, {\|}\mathit{\boldsymbol{u}}_j - \mathit{\boldsymbol{u}}_{j-1} {\|}_{V}^2, \quad AT_4 := \sum\limits_{j=1}^N\, {\|}\theta_j - \theta_{j-1} {\|}_F^2, \\ BV_5 := \sum\limits_{j=1}^N | \left\langle {B_j\, \mathit{\boldsymbol{u}}_{j-1}, \, \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j} \right\rangle_{V'\times V} |, \quad AT_5 := \sum\limits_{j=1}^N | \left\langle {C_j\, \theta_{j-1}, \, \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j} \right\rangle_{V'\times V} |, \\ F := \sum\limits_{j=1}^N | \left\langle {A_j\, \mathit{\boldsymbol{v}}_j, \, \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j } \right\rangle _{V'\times V} + \psi_u(t_j, \mathit{\boldsymbol{v}}_j) - \psi_u(t_j, \mathit{\boldsymbol{w}}^h_j) - \left\langle {\mathit{\boldsymbol{f}}_j, \, \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j} \right\rangle_{V'\times V} |, \end{array}

    Then we sum E_j from j=1 to j=n . After some algebraic manipulations, we obtain for N large enough, for any small \varepsilon >0, for some constant c>0, for n=1, \cdots, N,

    \begin{array}l {\|}\mathit{\boldsymbol{e }}_n{\|}_H^2+ k\sum\limits_{j=1}^n \, \|\mathit{\boldsymbol{e }}_j\|_V^2 \leq \ c\, {\|}\mathit{\boldsymbol{e }}_0{\|}_H^2 + c\, \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^h \|_V^2 \\ + c\, BV_0 + c\, k\, ( BV_1 + BV_2 + BV_4 + BV_5 + I ) + c\, BV_3\, M_e \\ + c\, k\, ( AT_4 + AT_5 ) + c\, k\, F \\ + c\, k\sum\limits_{j=1}^{n} \, (\alpha^{hk}_j)^2 + c\, k\sum\limits_{j=1}^{n} \, \beta^{hk}_j\, \| \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j \|_V \\ + \varepsilon\, k\, \sum\limits_{j=0}^{n-1} \, \| \varepsilon_j \|_F^2 + c\, k\sum\limits_{j=1}^{n-1} \, \Big( k\sum\limits_{i=1}^{j} \, \|\mathit{\boldsymbol{e }}_i\|_V^2 \Big), \end{array}

    where for n = 1, \cdots, N:

    \begin{array}l \alpha^{hk}_n := \Big\| \int_0^{t_n}\, \mathcal{B}(t_n-s) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}(s))\, ds - k\, \sum\limits_{m=0}^{n-1} \mathcal{B}(t_n-t_m) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}^{hk}_m) \Big\|_{\mathcal{H}} \\ \beta^{hk}_n = \Big\| k\, \sum\limits_{m=0}^{n-1} \mathcal{B}(t_n-t_m) \, \mathit{\boldsymbol{\varepsilon }}(\mathit{\boldsymbol{u}}^{hk}_m) \Big\|_{\mathcal{H}}. \end{array}

    From (4.1) and after some upper bound computations, we obtain for n=1, \cdots, N:

    \begin{array}l k\sum\limits_{j=1}^{n} \, (\alpha^{hk}_j)^2 \leq c\, k^2 + c\, k\, \sum\limits_{m=0}^{n-1}\, \| \mathit{\boldsymbol{u}}^{hk}_m- \mathit{\boldsymbol{u}}_m\|_V^2 \\ \leq c\, k^2 + c\, \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^h \|_V^2 + c\, k\, I + c\, k\sum\limits_{j=1}^{n-1} \, \Big( k\sum\limits_{i=1}^{j} \, \|\mathit{\boldsymbol{e }}_i\|_V^2 \Big); \end{array}

    and

    \begin{array}l \, k\sum\limits_{j=1}^{n} \, \beta^{hk}_j\, \| \mathit{\boldsymbol{v}}_j-\mathit{\boldsymbol{w}}^h_j \|_V \leq c\, k\, \sum\limits_{m=0}^{n-1}\, \| \mathit{\boldsymbol{u}}^{hk}_m- \mathit{\boldsymbol{u}}_m\|_V^2 + c\, k\, (BV_2 + \widehat{BV}_2) \\ \leq c\, \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^h \|_V^2 + c\, k\, I + c\, k\sum\limits_{j=1}^{n-1} \, \Big( k\sum\limits_{i=1}^{j} \, \|\mathit{\boldsymbol{e }}_i\|_V^2 \Big) + c\, k\, (BV_2 + \widehat{BV}_2); \end{array}

    and

    \begin{array}l k\, I \leq c\, k^2; \\ k\, AT_1 \leq c\, k^2; \\ k\, BV_1 \leq c\, k^2; \\ k\, AT_4 \leq c\, k^2; \\ k\, BV_4 \leq c\, k^2; \\ AT_5 + BV_5 + F \leq c\, \widehat{BV}_2. \end{array}

    Using then (4.11), we deduce for some constant c>0, for n=1, \cdots, N,

    \begin{array}l \| \mathit{\boldsymbol{e }}_n{\|}_H^2 + k\sum\limits_{j=1}^n \|\mathit{\boldsymbol{e }}_j\|_V^2 \\ \ \leq c\, \| \mathit{\boldsymbol{e }}_0 \|_H^2 + c\, \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^h \|_V^2 + c\, \| \varepsilon_0 \|_{F}^2 + c\, BV_0 + c\, AT_0^2 \\ \quad + c\, k\, ( BV_2 + \widehat{BV}_2 ) + c\, k\, AT_2 + c\, k^2 \\ \quad + c\, BV_3\, M_e + c\, AT_3\, M_\varepsilon \\ \quad + c\, k\sum\limits_{j=1}^{n-1} \, \Big( k\, \sum\limits_{i=1}^{j} \, \|\mathit{\boldsymbol{e }}_i\|_V^2 \Big). \end{array}

    Using then Gronwall's inequality, and again the estimation (4.11), we conclude that for some constant c>0, and for n=1, \cdots, N,

    \begin{array}l \max\Big( \| \mathit{\boldsymbol{e }}_n{\|}_H^2 + k\sum\limits_{j=1}^n \|\mathit{\boldsymbol{e }}_j\|_V^2;\, \| \varepsilon_n {\|}_{F}^2+ k\sum\limits_{j=1}^n \|\varepsilon_j\|_E^2 \Big) \\ \ \leq c\, \| \mathit{\boldsymbol{e }}_0 \|_H^2 + c\, \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^h \|_V^2 + c\, \| \varepsilon_0 \|_{F}^2 + c\, BV_0 + c\, AT_0^2 \\ \quad + c\, k\, ( BV_2 + \widehat{BV}_2 ) + c\, k\, AT_2 \\ \quad + c\, k^2 + c\, BV_3^2 + c\, AT_3^2. \end{array}

    This gives the estimation (4.10) stated in Theorem 2.

    The inequality (4.10) is a basis for error estimates for particular choice of the finite-dimensional subspace V^h and under additional data and solution regularities.

    As a typical example, let us consider \Omega \subset \mathbb{R}^d, d\in \mathbb{N}^*, a polygonal domain. Let {\mathcal T}^h be a regular finite element partition of \Omega. Let V^h\subset V and E^h\subset E be the finite element space consisting of piecewise polynomials of degree \leq m-1, with m \geq 2, according to the partition {\mathcal T}^h. Denote by \Pi^h_V\, :\, H^m(\Omega)^d \to V^h and \Pi^h_E\, :\, H^m(\Omega)\to E^h the corresponding finite element interpolation operator. Recall (see e.g. [3]) that:

    \left\{ \begin{array}l \| \mathit{\boldsymbol{w}} - \Pi_V^h \mathit{\boldsymbol{w}} \|_{H^l(\Omega)^d} \leq c\, h^{m-l}\, |\mathit{\boldsymbol{w}}|_{H^m(\Omega)^d}, \quad \forall\, \mathit{\boldsymbol{w}} \in H^m(\Omega)^d; \\ \| \eta -\Pi_E^h \eta \|_{H^l(\Omega)} \leq c\, h^{m-l}\, |\eta|_{H^m(\Omega)}, \quad \forall\, \eta \in H^m(\Omega). \end{array} \right.

    where l=0 (for which H^0=L^2) or l=1.

    We assume more generally the following additional data and solution regularities for some \alpha \geq 1:

    \begin{equation} \label{addss} \left\{ \begin{array}l \mathit{\boldsymbol{u}}_0 \in H^{\alpha + 1}(\Omega)^d ; \\ \mathit{\boldsymbol{v}}\in C([0, T];H^{2\alpha +1}(\Omega)^d), \quad \dot{\mathit{\boldsymbol{v}}} \in L^1(0, T; H^{\alpha}(\Omega)^d); \\ \theta \in C([0, T];H^{\alpha +1}(\Omega)), \quad \dot{\theta} \in L^1(0, T;H^{\alpha}(\Omega)). \end{array} \right. \end{equation} (4.12)

    Then we choose in (4.10) the elements

    \mathit{\boldsymbol{u}}_0^{h}=\Pi^h_V\, \mathit{\boldsymbol{u}}_0, \quad \mathit{\boldsymbol{v}}_0^{h}=\Pi^h_V\, \mathit{\boldsymbol{v}}_0, \quad \theta_0^{h}=\Pi^h_E\, \theta_0,

    and

    \mathit{\boldsymbol{w}}_j^{h}=\Pi^h_V\, \mathit{\boldsymbol{v}}_j, \quad \eta_j^{h}=\Pi^h_E\, \theta_j, \quad j=1\cdots N.

    From the assumptions (4.12), we have:

    \| \mathit{\boldsymbol{u}}_0 - \mathit{\boldsymbol{u}}_0^{h} \|_V \leq c\, h^\alpha, \quad \| \mathit{\boldsymbol{e }}_0 \|_H \leq c\, h^\alpha, \quad \| \varepsilon_0 \|_F \leq c\, h^\alpha; \\ AT_0 \leq c\, h^\alpha, \quad BV_0 \leq c\, h^{2\alpha};\\ AT_3 \leq c\, h^\alpha, \quad BV_3 \leq c\, h^\alpha;\\ k\, AT_2 \leq c\, h^{2\alpha}, \quad k\, BV_2 \leq c\, h^{2\alpha}, \quad k\, \widehat{BV}_2 \leq c\, h^{2\alpha}.

    Using these estimates in (4.10), we conclude to the following error estimate result.

    Theorem 3. We keep the assumptions of Theorem 2. Under the additional assumptions (4.12), we obtain the error estimate for the corresponding discrete solution \mathit{\boldsymbol{v}}_n^{hk}, \theta_n^{hk}, n=1, \dots, N.

    \begin{array}l \max\limits_{0\le n\le N}{\|}\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}{\|}_H +\Big(k\sum\limits_{n=0}^N \|\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}\|_V^2\Big)^{1/2} \\ + \max\limits_{0\le n\le N}{\|}\theta_n-\theta_n^{hk}{\|}_{F} +\Big(k\sum\limits_{n=0}^N \|\theta_n-\theta_n^{hk}\|_E^2\Big)^{1/2} \le c\, (h^{\alpha} + k). \end{array}

    In particular, for \alpha = 1, we have

    \begin{array}l \max\limits_{0\le n\le N}{\|}\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}{\|}_H +\Big(k\sum\limits_{n=0}^N \|\mathit{\boldsymbol{v}}_n-\mathit{\boldsymbol{v}}_n^{hk}\|_V^2\Big)^{1/2} \\ + \max\limits_{0\le n\le N}{\|}\theta_n-\theta_n^{hk}{\|}_{F} +\Big(k\sum\limits_{n=0}^N \|\theta_n-\theta_n^{hk}\|_E^2\Big)^{1/2} \le c\, (h + k). \end{array}

    5. Numerical computations

    Here we provide numerical simulations derived from the previous discrete schemes, by using Matlab computation codes, in the case of bilateral contact with Tresca's friction law. Here the following classical examples are taken :

    \begin{array}l {C}_e(t, \theta) := - \theta\, (c_{ij}(t) ) \quad \mbox{in} \quad \Omega; \\ {\cal K}_c(t, \nabla \theta) = (k_{ij}(t))\, \nabla \theta \quad \mbox{in} \quad \Omega; \\ {\cal C}_e(t, \mathit{\boldsymbol{v}} ) = -c_{ij}(t)\, \frac{\partial\, v_i}{\partial\, x_j} \quad \mbox{in} \quad \Omega;\\ \varphi(t, r) = \frac{1}{2} k_e(t)\, (r -\theta_R(t) )^2 \quad \mbox{on} \quad \Gamma_3. \end{array}

    On the contact surface \Gamma_3, we consider a bilateral condition and satisfies (see e.g. [4], [6]):

    \left\{ \begin{array}{l} u_\nu=0, \quad |{\mathit{\boldsymbol{\sigma }}}_{\tau}|\leq g(t), \\ |{\mathit{\boldsymbol{\sigma }}}_{\tau}|<g(t)\, \Longrightarrow \, \dot {\mathit{\boldsymbol{u}}}_{\tau}={\bf{0}}, \\ |{\mathit{\boldsymbol{\sigma }}}_{\tau}|=g(t)\, \Longrightarrow\, \dot {\mathit{\boldsymbol{u}}}_{\tau}=-\lambda{\mathit{\boldsymbol{\sigma }}}_{\tau}, \mbox{ for some } \lambda\geq 0, \end{array} \right.\quad \mbox{ on } (0, T) \times \Gamma_3.

    Here g(t) represents the friction bound, i.e., the magnitude of the limiting frictional traction at which slip begins, with g\in L^{\infty}((0, T) \times \Gamma_3), g(t) \geq 0 a.e. on \Gamma_3. We deduce the admissible displacement space:

    V := \{ \mathit{\boldsymbol{w}} \in H_1; \ w_\nu = 0 \ \mbox{on} \ \Gamma_3 \},

    and the sub-differential contact function independent on time:

    \varphi_u(t, \mathit{\boldsymbol{x}}, \mathit{\boldsymbol{y }}) = g(t, \mathit{\boldsymbol{x}})|\mathit{\boldsymbol{y }}_{\tau(x)}|\quad\forall \mathit{\boldsymbol{x}}\in \Gamma_3, \ \mathit{\boldsymbol{y }}\in \mathbb{R}^d,

    where \mathit{\boldsymbol{y }}_{\tau(x)} := \mathit{\boldsymbol{y }} -y_{\nu(x)}\mathit{\boldsymbol{\nu}}(\mathit{\boldsymbol{x}}), y_{\nu(x)} := \mathit{\boldsymbol{y }} \cdot \mathit{\boldsymbol{\nu}}(\mathit{\boldsymbol{x}}), with \mathit{\boldsymbol{\nu}}(\mathit{\boldsymbol{x}}) the unit normal at \mathit{\boldsymbol{x}}\in\Gamma_3. We have then

    \psi_u(t, \mathit{\boldsymbol{v}}) := \int_{\Gamma_3} g(t)\, | \mathit{\boldsymbol{v}}_\tau |\, da, \quad \forall \mathit{\boldsymbol{v}} \in V

    is well defined on V and independent on time with the property: for some c>0,

    | \psi_u(t, \mathit{\boldsymbol{w}}) -\psi_u(t, \mathit{\boldsymbol{v}})| \leq c\, \| \mathit{\boldsymbol{v}}-\mathit{\boldsymbol{w}} \|_{ L^2(\Gamma_3)^d}, \quad \forall \mathit{\boldsymbol{v}}, \, \mathit{\boldsymbol{w}} \in V.

    Thus from the definition it is clear that \psi_u(t, \cdot)\, :\, V \longrightarrow \mathbb{R} is convex. By using the continuous embedding from V into L^2(\Gamma_3)^d and the last inequality, we find that

    \psi_u(t, \cdot)\ \mbox{is Lipschitz continuous on} \ V.

    Then the assumptions in (2.17) are verified.

    We consider for simulations a rectangular open set, linear elastic and long memory viscoelastic operators, with non clamped condition.

    \begin{array}l \Omega = (0, L_1) \times (0, L_2);\quad \Gamma_1 = \emptyset; \\ \Gamma_2 = (\{0\}\times [0, L_2]) \cup ([0, L_1] \times \{L_2\} ) \cup (\{L_1\}\times [0, L_2]); \quad \Gamma_3 = [0, L_1] \times \{0\};\\ ({\mathcal G}(t)\, \mathit{\boldsymbol{\tau }})_{ij}= \frac{E(t)\, \kappa(t)}{1-\kappa(t)^2}(\tau_{11}+\tau_{22})\, \delta_{ij}+\frac{E(t)}{1+\kappa(t)}\tau_{ij}, \quad 1\leq i, \, j\leq 2, \ \mathit{\boldsymbol{\tau }} \in S_2;\\ ({\mathcal A}(t)\, \mathit{\boldsymbol{\tau }})(t)_{ij}= \mu(t)\, (\tau_{11}+\tau_{22})\, \delta_{ij}+ \eta(t)\, \tau_{ij}, \quad 1\leq i, \, j\leq 2, \ \mathit{\boldsymbol{\tau }} \in S_2;\\ ({\mathcal B}(t)\, \mathit{\boldsymbol{\tau }})_{ij}= B_1(t)\, (\tau_{11}+\tau_{22})\, \delta_{ij}+ B_2(t)\, \tau_{ij}, \quad 1\leq i, \, j\leq 2, \ \mathit{\boldsymbol{\tau }} \in S_2, \ t\in [0, T] \end{array}

    Here E(t) is the Young's modulus, \kappa(t) the Poisson's ratio of the material, \delta_{i j} denotes the Kronecker symbol, \mu(t) and \eta(t) are viscosity constants, at each time t \in [0, T].

    We refer to the previous numerical scheme, and use spaces of continuous piecewise affine functions V^h\subset V and E^h\subset E as families of approximating subspaces. %, with the time and space steps k=h=1/8. For computations we considered the following data (IS unity), \forall t\in [0, T]:

    \begin{array}{l} L_1 = L_2 = 1, \quad T = 1\\ \displaystyle \mu(t) = 8\, e^{t}, \quad \eta(t) = \frac{20}{1+ t^2}, \quad E(t) = \frac{4}{1+ t}, \quad \kappa(t) = \frac{0, 2}{1+ t^2}, \quad \mathit{\boldsymbol{f}}_0(\mathit{\boldsymbol{x}}, t) = (0, \, -5 t) \\ \mathit{\boldsymbol{f}}_2(t, \mathit{\boldsymbol{x}})=(0, \, 0), \quad \forall \mathit{\boldsymbol{x}} \in \{0\}\times [0, L_2] \\ \mathit{\boldsymbol{f}}_2(t, \mathit{\boldsymbol{x}})=(1+t, \, 0), \quad \forall \mathit{\boldsymbol{x}} \in ([0, L_1] \times \{L_2\}) \cup (\{L_1\}\times [0, L_2]) \\ c_{11}(t) = c_{12}(t) = c_{21}(t) = t, \quad c_{22}(t)=t^2 \\ k_{11}(t) = k_{22}(t) = 1+t, \quad k_{12}(t) = k_{21}(t) = t \\ k_e(t) = 1+t, \quad q(t)=e^t \\ B_1(t)=B_2(t)= 10^{-2}\, e^{-t} \\ \mathit{\boldsymbol{u}}_0=(0, 0), \quad \mathit{\boldsymbol{v}}_0=(0, 0), \quad \theta_0 = 0 \end{array}

    In Figure 1 (see figures below) is representing the initial configuration.

    Figure 1. Initial configuration.

    In Figure 2 we have the deformed configuration at final time, where \theta_R(t) = 0, \theta_a(t) = 5\, e^{-t}, 0\leq t\leq 1, for two different types of Tresca's friction bounds. For smaller friction bound with respect to the surface fraction on \Gamma_2, where g(t, x)= \frac{t+x}{4}, 0\leq t \leq 1, 0\leq x \leq 1, we observe on the contact surface a slip phenomena in the direction of the fraction: which means that the friction bound has been obtained. Whereas for large friction bound, e.g. for g(t, x)= 9\, (t+x), 0\leq t \leq 1, 0\leq x \leq 1, then slip in the direction of the traction could be less easy, friction in this case appears more important as well as for the deformation.

    Figure 2. Deformed configurations at final time, \theta_R(t) = 0, \theta_a(t) = 5\, e^{-t}, 0\leq t\leq 1.

    In Figures 3, we compute the Von Mises' norm which gives a global measure of the stress in the body. We observe that stress is more important for larger friction bounds, at the contact surface or at the traction surface.

    Figure 3. Von Mises' norm in long memory deformed configurations, \theta_R(t) = 0, \theta_a(t) = 5\, e^{-t}.

    In Figures 4, we show the influence of the different temperatures of the foundation (\theta_R(t) = 0 or \theta_R(t) = 30) on the temperature field \xi of the body.

    Figure 4. Temperature \xi at final time in long memory, g(t, x)= 9\, (t+x).

    Acknowledgments

    I am immensely grateful for the reviewer to have read this work and for the AIMS Mathematics to publish the paper.


    Conflict of Interest

    The author declares no conflicts of interest in this paper.


    [1] O. Chau, V. V. Motreanu, Dynamic contact problems with velocity conditions, Int. J. Ap. Mat. Com-Pol, 12 (2002): 17-26.
    [2] O. Chau, Quelques problèmes d'évolution en mécanique de contact et en biochimie, Habilitation thesis, Saint Denis La Réunion: University Press, 2010.
    [3] P. G. Ciarlet, The Finite Element Method for Elliptic Problems, Contributions to Nonlinear Functional Analysis, Amsterdam: North Holland, 1978.
    [4] G. Duvaut, J. L. Lions, Les Inéquations en Mécanique et en Physique, Paris: Dunod, 1972.
    [5] J. L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Contributions to Nonlinear Functional Analysis, Dunod et Gauthier-Villars, 1969.
    [6] P. D. Panagiotopoulos, Inequality Problems in Mechanics and Applications, Contributions to Nonlinear Functional Analysis, Basel: Birkhäuser, 1985.
    [7] P. D. Panagiotopoulos, Hemivariational Inequalities, applications in Mechanics and Engineering, Contributions to Nonlinear Functional Analysis, Springer-Verlag, 1993.
    [8] N. Costea, V. Radulescu, Hartman-Stampacchia results for stably pseudomonotone operators and non-linear hemivariational inequalities, Appl. Anal., 89 (2010): 175-88.
    [9] E. Zeidler, Nonlinear Functional Analysis and its Applications, Contributions to Nonlinear Functional Analysis, Springer Verlag, 1997.
  • This article has been cited by:

    1. Lamia Chouchane, Lynda Selmani, A HISTORY-DEPENDENT FRICTIONAL CONTACT PROBLEM WITH WEAR FOR THERMOVISCOELASTIC MATERIALS, 2019, 24, 1392-6292, 351, 10.3846/mma.2019.022
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(4493) PDF downloads(1019) Cited by(1)

Article outline

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog