Citation: Claude Stolz. Moving surfaces and interfaces : application to damage, fracture and wearcontact[J]. AIMS Materials Science, 2016, 3(3): 881-907. doi: 10.3934/matersci.2016.3.881
[1] | Francesco Parrinello, Giuseppe Marannano . Cohesive delamination and frictional contact on joining surface via XFEM. AIMS Materials Science, 2018, 5(1): 127-144. doi: 10.3934/matersci.2018.1.127 |
[2] | Nikolay A. Voronin . Specialties of deformation and damage of the topocomposite on a ductile substrate during instrumental indentation. AIMS Materials Science, 2020, 7(4): 453-467. doi: 10.3934/matersci.2020.4.453 |
[3] | Clara Argerich, Ruben Ibáñez, Angel León, Anaïs Barasinski, Emmanuelle Abisset-Chavanne, Francisco Chinesta . Tape surface characterization and classification in automated tape placement processability: Modeling and numerical analysis. AIMS Materials Science, 2018, 5(5): 870-888. doi: 10.3934/matersci.2018.5.870 |
[4] | R. Citarella, M. Lepore, M. Perrella, C. Calì . Multiple crack propagation by DBEM in a riveted lap-joint. AIMS Materials Science, 2016, 3(4): 1577-1586. doi: 10.3934/matersci.2016.4.1577 |
[5] | Kai Friebertshäuser, Christian Wieners, Kerstin Weinberg . Dynamic fracture with continuum-kinematics-based peridynamics. AIMS Materials Science, 2022, 9(6): 791-807. doi: 10.3934/matersci.2022049 |
[6] | Serguei Chiriaev, Nis Dam Madsen, Horst-Günter Rubahn, Shuang Ma Andersen . Helium Ion Microscopy of proton exchange membrane fuel cell electrode structures. AIMS Materials Science, 2017, 4(6): 1289-1304. doi: 10.3934/matersci.2017.6.1289 |
[7] | Hanh C. Nguyen, Shigeru Nagasawa, Kensei Kaneko . Strength estimation of silicon nitride ceramics using a round-notched specimen subjected to shearing-tool indentation. AIMS Materials Science, 2020, 7(5): 518-533. doi: 10.3934/matersci.2020.5.518 |
[8] | Christoph Nick, Helmut F. Schlaak, Christiane Thielemann . Simulation and Measurement of Neuroelectrodes Characteristics with Integrated High Aspect Ratio Nano Structures. AIMS Materials Science, 2015, 2(3): 189-202. doi: 10.3934/matersci.2015.3.189 |
[9] | Bhushan Sopori, Srinivas Devayajanam, Prakash Basnyat . Surface characteristics and damage distributions of diamond wire sawn wafers for silicon solar cells. AIMS Materials Science, 2016, 3(2): 669-685. doi: 10.3934/matersci.2016.2.669 |
[10] | Mario Ceddia, Bartolomeo Trentadue . A review of carbon fiber-reinforced polymer composite used to solve stress shielding in total hip replacement. AIMS Materials Science, 2024, 11(3): 449-462. doi: 10.3934/matersci.2024023 |
Two main families of methods exist to model failure of quasi-brittle structures. One is based on crack models, another one uses continuous damage approach that leads to a local loss of stiffness. Initiation of crack can be investigated using Griffith law an a criterion of global stability as proposed in [1], but fracture mechanics is generally not sufficient to model the full scenario of the degradation of Omegaids. Another point of view is adopted here based on damage modelling.
The full scenario of the degradation of Omegaids under mechanical loading is described by modelling the gradual loss of rigidity. This common approach is purely local. Generally local damage induces localization. Several models were proposed to avoid spurious localization: non local approaches and higher order, damage based, gradient models. These approaches are compared in [2]. Other new approaches are also proposed, the phase field approach [3] and the so-called variational approach [4].
Another way to describe the damage evolution is to consider the propagation of the surface separating the sound material and damaged material, as proposed for brittle material in [5,6]. This point of view is adopted here.
The propagation of surface inside a body is analysed. The moving surface here is associated with a change of mechanical properties. This framework is used to describe damage or phase transformation. Variational formulations were performed to describe the evolution of the surface between undamaged and damaged material. Connection with the notion of configurational forces can be made [7]. When this surface is moving, a flux of matter is induced, that is useful for describing the loss of material during wear mechanisms or brittle fracture [8].
For elastic brittle material the evolution of the interface separating the undamaged material (d=0) from the total damaged (d=1) have been studied using an energetic description of the propagation of damage [9,10]. In this description the damage parameter jumps from 0 to 1.
The interface has no thickness and mechanical quantities present strong discontinuities. The evolution in terms of rate of displacement and velocity of propagation of the surface is governed by a variational inequality. Criteria of stability and uniqueness have been established. For example, nucleation of defects in this modelling can be considered as a bifurcation of equilibrium solution [11]. Description of moving interfaces and of moving layers is also a manner for studying loss of material. Such a description of thin or thick layers permits to describe complex processes of wear contact between two bodies in relative motion [8,12]. In a more recent paper [13], the transition between undamaged material to damaged material is continuous through a layer of finite thickness. The evolution of damage is then associated to a moving layer.
Two cases have been considered: the propagation of a interface and the propagation of a moving layer of finite thickness.
For sharp interface, the transition zone is very thin, the damage parameter has a discontinuity, it jumps from 0 to 1. The driving force associated to the propagation of the interface is a local release rate of energy G(s) as in [14]. Moreover, stability and non bifurcation conditions of the evolution are given when a normality rule based on G(s) governs the propagation of the interface. When an additional surface energy is considered along the interface, the propagation is more stable [15].
For a moving layer, the transition is more regular. The thickness of the transition provides a length scale in the model. For specific definition of the damage, a local condition of steady state is induced and a generalisation of the local energy release rate is obtained. This is the driving force associated to the motion of surface separating the sound material from the damage one.
The main ideas of the model is to consider that the strain energy changes from a sound material to a damaged one depending on a damage parameter, which is continuous or not, and to describe the motion of the interface by a complementary law. For particular description, the motion of the interface induces a dissipation depending on an energy release rate. A normality rule is then proposed to govern the propagation of the interface based on this mechanical quantities. This choice generalizes classical Griffith's law.
The purpose of this article is to describe the motion of a surface or interface, to propose an evolution law for this motion in terms of the local energy release rate. The rate boundary value problem is then analysed and conditions of stability and bifurcation are given. Some applications and comments of this modelling are proposed in the last section.
Consider a body Ω. The external boundary ∂Ω is decomposed in two complementary parts: ∂Ωu where the displacement u_ is prescribed u_d(t) and ∂ΩT where the loading T_d(t) is applied: ∂Ω = ∂Ωu∪∂ΩT,∅ = ∂Ωu∩∂ΩT. The displacement u_ is continuous across the surface Γo, separating the undamaged material and the damaged material. The internal state of stress σ is such that the stress vector is also continuous along the surface Γo. Then we have the relations of continuity:
[u_]=0,[σ].v_=0, | (1) |
where v_ is the normal vector to the interface. When the surface is moving, these conditions must be preserved.
Consider the surface Γo separating the sound material from the damaged material. The surface is parametrized by two parameters sα. For Mo on the surface, the tangent space is determined by the local basis T_α. The normal vector to the tangent space is v_. Then we have
T_α=∂Mo∂sα=∇αMo | (2) |
where K is the curvature tensor at point Mo.
When this surface is moving the local basis changes accordingly. At time t the point Mot belongs to the surface, it satisfies the equation ϕ(Mot,t)=0 defining Γo.
At time t+=t+dt, the geometrical point Mot comes in Mot+ such that ϕ(Mot+,t+)=0 then the new position is Mot+=Mto−dta(sα)v_t where the normal speed a satisfies the relation
∂ϕ∂t−a∇ϕ.v_t=0. | (3) |
Any quantities f defined at Mo is varying during the motion. Introducing the derivative of f Omegalowing the motion of Γo by
Daf=limΔt→0f(Mot+,t+)−f(Mot,t)dt. | (4) |
We obtain for the evolution of the geometry
DaT_α=−∇αav_,Dav_=∇aDaK=−∇∇a−K.Ka. | (5) |
When f is the value at Mo of a field defined on Ω, then the derivative Daf is the classical convective derivative
Daf=∂f∂t−a∇f.v_. | (6) |
This expression is useful for written conservation laws and continuities along the surface Γo.
Hadamard's relations. On account of perfect bounding between phases, displacement and stress vectors are continuous along Γo. Their rates have discontinuities according to the general compatibility conditions of Hadamard. These conditions are rewritten in terms of convected derivative:
0=[u_]Γ,Da([u_]Γ)=[v_]Γ−a[∇u_]Γ.v_=0, | (7) |
0=[σ]Γ.v_,Da([σ]Γ.v_)=[˙σ] Γ.v_+divΓo[aσ]Γ=0. | (8) |
The last equation is obtained by taking account of the conservation of the momentum. Indeed we have
Da([σ].v_)=Da[σ].v_+σ.Da(v_)=0, | (9) |
using the conservation of the momentum in the local frame relatively to Γo
T_α.∇.σ.T_α+v_.∇σ.v_=0, | (10) |
and the expression of the surface divergence
divΓof=divf−v_.∇f.v_, | (11) |
the required result is obtained.
In two dimensions. In two dimensions the expressions are:
DaT−=−dadS v−,Dav−=dadS T−,DaK=−d2adS2−K2 a, | (12) |
0=[v_]−a∇u_.v_,0=[˙σ]Γ.v_+ddS([σ.T_]Γa). | (13) |
The domain is composed of two distinct Omegaumes Ωo and Ω1, which are occupied by two materials with different mechanical characteristics. The perfect interface between them is assumed to be a regular surface and is denoted by Γo. Material 0 changes into material 1, along Γo by an irreversible process. Hence, Γo moves with a normal velocity a, the irreversibility is given by the positivity of a for a propagation in the direction external to 2.
The subscript i is used to denote material i. The actual state is characterized by the displacement field u_. The strain field ε is given by ε(u_)=12(∇u_+∇Tu_).
To simplify the presentation, we consider here that the materials are linear elastic. The total potential energy of the structure has the Omegalowing form:
ε(u_,Γo)=∑i∫Ωiψi(ε(u_))dΩ−∫∂ΩTT_d.u_dS, | (14) |
where ψi(ε) denotes the density of strain energy in the domain i: ψi(ε)=12ε:Ci:ε.
It is important to point out that the potential energy represents the global free energy in a thermodynamic description; then the position of the interface is an internal parameter for the global structure.
The displacement u_ is kinematically admissible, it satisfies particular boundary condition:
u_(x,t)∈C={u_/u_(x,t)=u_d(t),x∈∂Ωu,[u_]=0,x∈Γo(t)}. | (15) |
At any time, when the position of the interface is known, the behaviour of the body is those of a composite bi-materials, with perfect contact between two material phases. Then the solution of the problem of equilibrium is determined by the minimization of the total potential energy with respect to the displacement u_ among the set of admissible fields C.
Characteristic of an equilibrium state. For a state of equilibrium (a=0) the variations with respect to u_ are
∂ε∂u_.δu_=∑i∫Ωiσ:ε(δu_)dΩ−∫∂ΩTT_d.δu_dΩ. | (16) |
A state of equilibrium is defined by displacement u_ such that
σ=Ci:ε(u_), in Ωi,divσ=0,σ.n_=T_d along ∂ΩT. | (17) |
Variations with respect to the position of the interface. The total variation of ε near a position of equilibrium, must take into account of the continuity condition. The variations of the displacement and the variations of the position of the surface are coupled by Hadamard's relation.
[δu_]+δa[∇u_].v_=0. | (18) |
The total variation of the potential energy is:
∂ε∂u_.δu_+∂ε∂Γo.δΓo=∫Ωσ.ε(δu_)dΩ−∫Γo[ψ]δadΩ−∫∂ΩTT_d.δu_dS. | (19) |
Taking account of the equilibrium, the total variation of ε near an equilibrium point is reduced to
−∂ε∂Γo.δΓo=∫Γoδa([ψ]−v_.σ.[∇u_].v_)dS=∫ΓoG(Mo)δa(Mo)dS=Dm. | (20) |
This corresponds to the dissipation Dm of the system as pointed out in [10,12]. We recover the results of [7,14,16]. The definition of the release rate of energy can be written in a simpler form, taking account of continuity of the stress vector and of the fact the discontinuities of strain and stress are orthogonal [17]:
[σ].[∇u_]=0, | (21) |
then the release rate of energy G takes the final expression:
G=[ψ]−σ:[∇u_]. | (22) |
The release rate of energy depends upon the position of the point Mto and on the loading.
From a position of equilibrium, we applied an amount of loading. The body is deformed an simultaneously, the surface Γo is moving accordingly some constitutive law. A kinetic law defined by a direct relation between a and G can be chosen as proposed in [16]. To describe the irreversibility of the motion, we specify here a different relation.
The propagation law. Based on the form of the dissipation, a criterion on G is chosen as a generalized form of the well-known theory of Griffith. We assume a normality rule to govern the speed of propagation a
a≥0,G≤Gc,(G−Gc)a=0. | (23) |
The subset of Γo where the critical value Gc is reached is denoted Γo+. Then the propagation is only possible on point of Γo+ where G(Mto,t)=Gc, this is equation of a surface.
When the interface is moving, the consistency condition associated to the motion is given by the convected derivative of G: during the motion DaG=0. This leads to the consistency condition written for all points belonging to Γo+:
(a−a∗)DaG≥0,∀a∗≥0 on Γo+. | (24) |
Evaluation of DaG. To calculate the convected derivative, we derive term by term, using the Hadamard's relations on velocities
v_o+a∇u_o.v_=v_1+a∇u_1.v_. | (25) |
The first term of G is the jump of energy
Da[ψ]=−σ1:(∇v_1+a∇∇u_1.v_)+σo:(∇v_o+a∇∇u_o.v_). | (26) |
Then we get
DaG=[ψ]−Daσ1:[∇u_]−σ1:[Da∇u_]=[σ]:∇v_o−˙σ1:[∇u_]−aGn, | (27) |
where Gn=−[σ]:∇∇u_1.v_+∇σ1.v_:[∇u_].
The rate boundary value problem. A solution (v_,a) of the problem of evolution must satisfy
• the constitutive law :
˙σ=Ci:ε(v_) in Ωi, | (28) |
• the compatibility of the velocity :
ε(v−)=12(∇v−+∇tv−), | (29) |
and the boundary condition v_=v_d on dΩu,
• the conservation of the momentum:
div˙σ=0, in Ω, | (30) |
and ˙σ.n_=˙T_d on dΩT,
• the compatibility conditions on the moving interface:
[Dau_]Γ=0,[Da(σ.u_)]Γ=0, | (31) |
• the consistency condition:
∀a∗≥0 on Γo+,(a−a∗)DaG≥0. | (32) |
Theorem 1. The evolution is determined by the variational inequality
∂F∂v_.(v_∗−v_)+∂F∂a(a∗−a)≥0, | (33) |
among the set K.A of admissible fields (v_∗,a∗):
K.A={(v_,a)|v_=v_d on ∂Ωu,[v_]+a[∇u_]=0,a∈K},K={a|a≥0 on Γo+,a=0 otherwise }, |
and the functional F(v_,a,˙T_d) is defined as
F=∑i∫Ωi12ε(v_):Ci:ε(v_)dΩ−∫∂ΩT˙T_.v_dS−∫Γoa[σ]:∇v_1dS+∫Γo12a2GndS. | (34) |
The equations (28, 29, 30) are those of a classical problem of elasticity with non classical boundary conditions (31) on Γo+. Assume that a is a given function on Γo+ then the problem can be Omegaved with respect to v_. This solution v_ is a function of the boundary conditions and of the distribution of the speed a along the interface. We can define ψ(a) the value of F for this solution then:
ψ(a)=F(v_(a,˙T_d,u_d),a,˙T_d), | (35) |
and finally the problem of evolution is defined as a variational inequality based on ψ. It is obvious that ψ is quadratic in a and we have the properties
Stability If a.∂2ψ∂a∂a.a≥0 for all a≥0 over Γo+, the position a=0 is stable.
No- bifurcation If a.∂2ψ∂a∂a.a≥0 for all a over Γo+, there is no bifurcation.
When an additional surface energy is present on the interface Γo+ the release rate of energy is changed for a motion of the surface
δ∫ΓoβdS=∫ΓoβTrKδadS,Gβ=Go−βTrK. | (36) |
Now, we consider that the interface has a finite thickness. The damage parameter varies from 0 to 1 continuously. The material of the body has an elastic linear behaviour with moduli eOmegaving with damage. Then the local free energy ψ(ε,d) is a function of the strain and of the damage parameter d. The state equations are defined classically as:
σ=∂ψ∂ε=C(d):ε,Y=−∂ψ∂d. | (37) |
The elastic moduli are known functions of d. Driving force Y is associated to the damage parameter and the dissipation of the wOmegae system is reduced to
Dm=∫ΩY˙ddΩ. | (38) |
When damage is established, the wOmegae body is decomposed in three parts: the undamaged body Ωo where d=0, the transition zone Ωc and the damaged material Ω1 where d=1.
On the boundary ∂Ωc∩∂Ωo=Γo the moduli of elasticity are continuous, the displacement and the stress vector being continuous, the strain is also continuous then the free energy is continuous. Therefore, when the surface Γo is moving, there is no dissipation along the boundary of this layer.
The level-set ϕ=0 gives the position of Γo. We assume that the damage d is a continuous explicit function d(ϕ) of the distance ϕ to the surface Γo. In the domain Ω the damage parameter satisfies
0≤d(ϕ)≤1,{d=0,ϕ≤0,d′(ϕ)≥0,0≤ϕ≤lc,d(ϕ)=1,lc≤ϕ. | (39) |
The surface iso-damage d(M,t)=do is a level-set, it corresponds to the level-set ϕ(M,t)=z. The evolution of this level set is those of the motion of a surface then
˙ϕ+a∇ϕ.v_=0, | (40) |
as ϕ is a function distance v_=∇ϕ, ||v_||=1 and ˙ϕ+a=0 and we have
˙d+aΔd.v_=˙d+ad′(ϕ)=0,d′(ϕ)=Δd.v_. | (41) |
A point of the layer is described by the parameters sα along Γo and the distance to Γo.
M=Mo+zv_. | (42) |
At point M the local basis is (T_α,v_) with
T_α=T_α−zKβαT_β. | (43) |
Reciprocal basis is T_α such that T_α.T_β=δαβ. To simplify the analysis we consider now that the tangent vectors T_α are eigenvectors of the curvature K, then
T_1=(1−zK1)T_1,T_2=(1−zK2)T_2. | (44) |
The metric tensor in this basis is then
g11=(1−zK1)2,g22=(1−zK2)2,g12=0. | (45) |
And the curvature at point M is determined by
∇v_=∂v_∂sα⊗T_α=−KβαT_β⊗T_α=−∑αKα1−zKαT_α⊗T_α=−καT_α⊗T_α. | (46) |
We recover the classical result: Tr∇v_=ΔSϕ=−TrK. We define the covariant derivative of any quantity f by
∇Sf=∂f∂sαT_α=gαβ∂f∂sαT_β=∇αfT_α, | (47) |
then we obtain:
∇S∇Sf=∇γ(gαβ∂f∂sαT_β)⊗T_γ, | (48) |
ΔSf=1√g∂∂sα(gαβ√g∂f∂sβ), | (49) |
g=det(g),Γααβ=1√g∂√g∂sα, | (50) |
dΩ=√gdS1dS2dz. | (51) |
Where Γαβγ are the Christoffel symOmega associated to the tangent frame :
∂T_α∂sβ=ΓγαβT_γ+καβv_. | (52) |
When the surface Γo is moving with normal velocity a(sα) the local basis is changing accordingly and in particular
Dav_=−∇a,DaK=K.Ka+∇S∇Sa,Da(dΩ)=Da(√g)dΩ. | (53) |
In the same spirit ϕ(Mo+zv_,t)=z. As Da(Mo+zv_)=−av_+zDav_ the velocity of the level-set ϕ(M,t)=z for the point M=Mo+zv_ is the same than at the point Mo. All point on the normal direction to Γo has the same normal velocity a. Then the damage parameter d satisfies ˙d+aΔd.v_=0. The profile of damage d is conserved along the normal direction d′=Δd.v_.
The total potential energy is defined as previously
ε(u_,d,T_d)=∫Ωψ(ε(u_),d)dΩ−∫∂ΩTT_d.u_dS. | (54) |
The equilibrium position minimises the potential energy among the set of admissible displacement u_
u_∈C={u_|u_=u_d, over dΩu}, | (55) |
then
σ=C(d):ε(u_),0=divσ,σ.n_=T_ over dΩT. | (56) |
Combining the equilibrium position and the total variations of ε, we obtain
−∂ε∂d.δd=∫ΩYδddΩ=Dm, | (57) |
which correspond to the mechanical dissipation of the system.
The dissipation associated to the motion of the layer is associated to the motion of the surface Γo
Dm=∫ΩY˙ddΩ=∫ΩYd′˙ϕdΩ=∫Γo(∫loYd′j(z)dz)a(Mo)dS. | (58) |
This relation defines a generalized driving force associated to the motion of the layer ; the local thickness l depends on the point Mo and is limited by lc:
G(Mo)=∫loYd′j(z)dz. | (59) |
The Omegaume is described by the geometry of the surface Γo and the distance z to this surface. The tensor of curvature at Mo is K and the curvature at M=Mo+zv_ depends on K and z. The unit of area defined on the surface ϕ=z is related to the unit of area defined on ϕ=0 as
dSM=det(I−zK)dSo. | (60) |
Choice of a propagation law.
The velocity a is determined with respect to an evolution law based on the driving force G. The integration of the local normality rule:
˙d≥0,Y≤Yc,(Y−Yc)˙d=0, | (61) |
suggests that the velocity a satisfies the generalized normality rule:
a(Mo)≥0,G(Mo)≤Gc(Mo)=∫loYcΔd.v_j(z)dz;(G(Mo)−Gc(Mo))a(Mo)=0. | (62) |
Now the problem of evolution is investigated. During the motion of the layer, the critical value on G is conserved. We are interested by the evolution of integral of a continuous function f like
F=∫ΩcfdΩ→dFdt=∫Ωc(Daf+fTr∇DaM)dΩ. | (63) |
As M=Mo+zv_, we obtain DaM=av_−z∇a and Tr∇DaM=aTr∇v_−zΔa.
Consider the particular function f(ε,d)=(Y−Yc)d′ and a virtual velocity fields a∗, then the normality law implies
ddt∫Ωcfa∗dΩ=∫Ωc(Daf+fTr∇DaM)a∗dΩ. | (64) |
Due to the hypothesis d(ϕ) then Dad=0 and
Daf=∂f∂ε:Daε=∂f∂ε:(˙ε+∇ε.DaM)=∂f∂ε:(˙ε+a∇ε.v_)−z∂f∂ε:(∇ε.∇a). | (65) |
The last term is combined with fTr∇DaM and taking account of ∇a.v_=0
(z∂f∂ε:(∇ε.∇a)+f(aTrK+zΔa))a∗=∇(zf∇aa∗)−fz∇a.∇a∗+zfTrKaa∗. | (66) |
Combining the derivation of equilibrium equations and the consistency condition for the propagation, we obtain:
Theorem 2. The solution of the problem of evolution is given by the variational inequality
∂F∂v_.(v_∗−v_)+∂F∂a(a∗−a)≥0, | (67) |
where
F(v_,a)=∫Ω12ε(v_).C(d):ε(v_)dΩ−∫Ωca∂f∂ε:ε(v_)dΩ+∫Ωc12f(z||∇a||2−TrKa2)−12∂f∂ε:∂ε∂za2dΩ. | (68) |
The variations with respect to v_ implies the derivation of 56 with respect to time
{˙σ=C(0):ε(v−), in Ωo,˙σ=C(d):ε(v−)−a∂C(d)∂d:ε(u−)Δd.v−, in Ωc,˙σ=C(1):ε(v−), in Ω1, | (69) |
and the conservation of the momentum:
div˙σ=0, over Ω,˙σ.n_=˙T_d, along dΩT. | (70) |
The variations relatively to a(sα) gives
∫Ωc−∂f∂ε:˙εδa+f(z∇a.∇δa−TrKaδa)−∂f∂ε:∂ε∂zaδadΩ. | (71) |
By integration by part we obtain
∫Ωc−∂f∂ε:˙εδa+∇(fz∇aδa)−fzΔaδa−TrKaδa−∂f∂ε:∂ε∂zaδadΩ. | (72) |
And using the boundary condition δa=0 along ∂Γo+, we recover the consistency condition
∫Γo∫lco(∂f∂ε:Daε+fTr(∇DaM))j(z)dzδadS. | (73) |
During the phase of initiation of the layer, some additional terms due to ˙l are present.
Comparing to the case of sharp interface, the velocity must be more regular, due to the presence of its tangential gradient. The curvature plays a fundamental Omegae on the stability and bifurcation of the solution.
To illustrate the preceding results we propose to consider the case of a composite cylinder or a composite sphere, with a kernel of material 1 and a core of material 0 in the case of sharp interface, the radius of the boundary Γo is Ro, the external radius is Re. And we consider radial loading. The displacement solution of the problem of equilibrium is assumed to be radial
u_=u(r)e_r, | (74) |
and the strain are
• for the cylinder n=2, εr=dru,εt=ur,εz=0,
• for the sphere n=3, εr=dru,εt=εphi=ur.
In linear elasticity, with Lamés moduli λ,μ the conservation of the momentum, when Λ(r)=λ(r)+2μ with a shear modulus μ uniform, implies the differential equation on u(r):
dr(Λ(r)(dru+(n−1)ur))=0. | (75) |
The displacement solution is then obtained in the composite structure, assuming that the continuity of u at r=0 is satisfied, we have
rn−1u(r)=A∫rorn−1Λdr. | (76) |
The radial displacement for the global response is given by
u(Re)=ERe,RneE=AIn(Re). | (77) |
The total energy is
ε=2n−1π∫Reo12σ:εrn−1dr=2n−1πσrr(Re)u(Re)Rn−1e. | (78) |
As the radial stress is
σrr=λ(dru+(n−1)ur+2μdru=A−2μ(n−1)ur, | (79) |
the total strain energy is obtained:
ε=2n−1πRn−1eE2(RneIn(Re)−2μ). | (80) |
We consider now a two phase composite, the boundary Γo between the sound material and the partially damaged materiel is defined by the surface with radius Ro. When the radius Ro eOmegaves according to the normality rule, potential energy is eOmegaving and the dissipation is determined. For the composite system, the integral In depends upon Ro
In=∫Roorn−1Λ1dr+∫ReRorn−1Λodr. | (81) |
For given E, A and In satisfies ERn−1=AIn. Combining the derivation of In and E relatively to Ro we find the dissipation
Dm=−∂ε∂Roa=2n−1πRon−1Ga, | (82) |
where the energy release rate by unit of areas satisfies
G=G(E,Ro)=12A2InRo. | (83) |
During the motion of Γo the critical value Gc is conserved G=Gc, then A=Ac with
G=12A2cΛo−Λ1ΛoΛ1=Gc. | (84) |
Initially the radius is Ro=Ri and then In(Ro)=RnenΛo(1+c(Ro)Λo−Λ1ΛoΛ1) where c is the Omegaume fraction of material 1 c(Ro)=(RoRe)n−1. For increasing value E of the radial loading, the global response of the composite is decomposed in three state
• State I, A≤Ac then the answer is purely elastic, E≤Ec(c(Ri))
Σ=σrr(Re)=A−2n−1μE,E=AIn(Ri)Rne, | (85) |
• State II. The interface is moving, now A=Ac, the Omegaume fraction eOmegaves with Ro and
ERne=AcIn(Ro),Σ=ΛoAc−2n−1μE | (86) |
From this state, the relation between E and Σ is linear and Σ is a decreasing function of E.
When Ri tends to 0, the critical value Ec(c(Ri)) tends to Eoc which corresponds to the critical value for the propagation of an infinitesimal defect as proposed in [11].
To describe a continuous transition, we consider the damage law d(ϕ)=ϕlc and
1Λ=dΛ1+1−dΛo | (87) |
The values Λo,Λ1 are those used for the sharp interface. Initially the damaged material appears at r=0, and Ro increases from 0 to lc. After that the layer i moving keeping its thickness constant.
• State 0: Initially the response of the composite structure is purely elastic. When a critical value is reached, the transition zone is established.
• State I : Initiation of the transition zone. During the initiation of the layer, ϕ=(Ro−r)/lc,r≤lc, and the integral In is given by
In=∫Reorn−1Λ(r)dr=RnenΛo+Λo−Λ1ΛoΛ1Ron+1n(n+1)lc. | (88) |
During this state, the dissipation satisfies the relation
Dm=2n−1πΛo−Λ1ΛoΛ1A2lcRonn=2n−1πYc∫Roorn−1lcdr, | (89) |
and then
Λo−Λ1ΛoΛ1A2c=Yc. | (90) |
As for the case of sharp layer the parameter A=Ac is constant. The value Ac determines the critical strain Ec for the beginning of a kernel of damaged material: Ec=AcIn(0)/Rne.
During the phase of initiation of the layer A=Ac, Ro≤lc and
Σ=Ac−2n−1μE,E=AcIn(Ro)/Rne. | (91) |
After a purely elastic response during the increase of E from 0 to Ec, the global stress decreases linearly with E.
• State II : The finite thickness is reached, Ro=lc. From now, the layer is moving inside the structure keeping its thickness constant. d=(Ro−r)/lc,Ro−lc≤rRo. The integral In becomes
In(Ro)=RnenΛo+Λo−Λ1ΛoΛ11l _ c n(n+1)(Ron+1−(Ro−lc)n+1). | (92) |
The dissipation during the propagation gives the constrain
Dm=2n−1πΛo−Λ1ΛoΛ1A2nlc(Ron−(Ro−lc)n)=2n−1πYc∫Roorn−1lcdr. | (93) |
After simplification we recover the value A=Ac. Then after identification Yc=Gc, the response with a moving layer cannot be distinct from the sharp interface. Such a result is observed for the uni-axial case, like a bar in tension, for any constitutive law [13]. This is due for the particular choice of the function Λ(d).
For the same function d(ϕ) we can choose Λ(d)=dΛ1+(1−d)Λo. During the propagation of the layer when thickness is lc, we have
Λ(r)=Λ1Ro−rlc+(1−Ro−rlc)Λo=ΛM+ΔΛrlc. | (94) |
For this behaviour, we have for the cylinder
I2=(Ro−lc)22Λ1+R2e−Ro22Λo+l2cΔΛ−ΛMl2c(ΔΛ)2logΛoΛ1, | (95) |
and the constrain due to the normality law becomes
A2Λo−Λ1ΛoΛ1(Ro+lcΛoΔΛ)=2YcRo. | (96) |
In this case, the influence of the layer curvature is emphasized. It can be noticed that when lc/Ro tends to zero, the result for the sharp interface is recovered.
We consider now a model of blister test. Similar study is made in [18]. The damage parameter d=0 for the bonded part and d=1 for the domain of a membrane peeled of a rigid substrate.
The membrane is submitted to an internal pressure p and the deformation of the membrane is described by the vertical displacement u relatively to the substrate as shown in the figure.
The domain Ω is now a surface S where the membrane is peeled and the potential energy takes the form
ε(u,S,p)=∫S12K∇u2−pudS. | (97) |
On ∂S the membrane is bonded then u(x,t)=0,x∈∂S Under the internal pressure, the membrane is deformed and the equilibrium state is given by the minimum of the potential energy
∂ε∂u.δu=∫SK∇u∇δu−pδudS=0, | (98) |
for all δu defined on S and δu(x)=0,∈∂S. We obtain
KΔu+p=0, over S,u=0, along ∂S. | (99) |
For a variation of the loading ˙p, the boundary can move and the boundary condition u=0 is conserved Omegalowing the motion of ∂S with normal velocity a. The rate of the displacement for a point on the boundary satisfies the Hadamard relation
v+a∇u.n_=0, | (100) |
and the variation of potential energy satisfies simultaneously
dtε=∫SK∇u.∇v−pv−˙pudS+∫∂S(12K∇u2−pu)ϕds. | (101) |
The problem of evolution of a state of equilibrium is given by
KΔv+˙p=0, over S,v+a∇u.n_=0, along ∂S. | (102) |
The evolution of ε contains two terms, one due to v and another due to a The variation relatively to a defined the energy release rate
−∂ε∂SδS=∫∂SG(s)a(s)dS,G(s)=12K(∇u.n_)2. | (103) |
The propagation law is defined by the normality rule
G≤Gc,a≥0,a(G−Gc)=0. | (104) |
The convected derivative of G is easy to calculate
DaG=K∇u.n_(∇v.n_+ϕ∇∇un_.n_). | (105) |
As Dau=v+a∇u.n_=0, then any variation of u satisfies δu+δa∇u.n_=0.
The functional F(v,a) is reduced to
F=∫S12K∇v.∇v−˙pvdS−∫Γ12a2K∇u.n_∇∇u.n_.n_ds. | (106) |
For an initial circular geometry S with radius R, the solution of the problem is u=−p/4K(r2−R2) and G(R)≤Gc gives a critical pressure
pR<pcR=2σo. | (107) |
For p pressure greater than the critical value domain S varies according to the normality rule. The solution satisfies
∂F∂v(v∗−v)+∂F∂ϕ(a∗−a)≥0,∀a∗≥0. | (108) |
For given distribution of local speed a(s), the solution of the problem is v=v(˙p,a), we can introduce the functional Q
Q(a)=F(v_(˙p,a),a). | (109) |
To determine the set of possible solutions, solution v(˙p,a) is obtained by considering that the propagation speed a is developed according to the expansion a=αo+∑iαicos(iθ)+βisin(iθ).
• For prescribed pressure p we have
v=pr2K(αo+∑i(αicos(iθ)+βisin(iθ))(rR)i), | (110) |
the reduced functional
Q(a)=2πGc(−2α2o+∑i(i−1)(α2i+β2i)), | (111) |
is not positive definite, then the circular geometry is unstable.
• But when we contOmega the Omegaume the pressure varies according to the motion of Γo:
v=pr2K(αo+∑i(αicos(iθ)+βisin(iθ))(rR)i)+δpR2−r24K,∫SvdS=0. | (112) |
The variations of pressure is determined δp=−4pRαo and the functional is now
Q(a)=2πGc(6α2o+∑i(i−1)(α2i+β2i)). | (113) |
The positivity of Q for a≥0 is obtained, solutions a=˙RR+α1cosθ+β1sinθ≥0 are possible for sufficiently small α1,β1, the stability of the circular geometry is ensured. The functional Q is not strictly positive for all a and possible bifurcation, loss of circular geometry is obtained.
The discussion of delamination of laminates can be studied in this framework as proposed in [19,20,21]. The model of description of the laminate structure has a strong influence on the condition of stability and bifurcation of solution [22,23].
Neuber [24] has investigated the mechanism of stress-concentration near a notch and crack propagation. A solution is obtained for any non-linear stress-strain laws monotonically increasing and all loading intensities. The boundary of the notch consists of two parallel straight lines and a cycloid along which the amount shear is uniform. Bui and Ehrlacher [6] Omegaved the same problem for a linear elastic material; the thickness of the damaged zone is determined as a function of the stress intensity factor of the equivalent crack. Similar solution are obtained also in elastoplasticity by Bui [5] and for crack in [25]. More recently, the results have been generalized to different families of non-linear elastic brittle material [26] using the hodograph transformation proposed by [27,28]. To determine the mechanical quantities over the domain Ω, the hodograph transformation is used [27,29]. Another technique is also available [24].
In the transformation of hodograph, the components of the gradient ∇w become the new independent variables.
(x1,x2)→(ξ1,ξ2),ξα=w,α(x1,x2). | (114) |
The displacement appears as a potential (dw=ξαdxα). The mapping is invertible provided the Jacobian H=w,11w,22−(w,12)2 does not vanish. Denote by U the Legendre transformation of w with respect to ξα
U(ξ1,ξ2)=xαw,α(x1,x2)−w(x1,x2). | (115) |
By differentiating U with respect to ξ, the conjugate equations are obtained
xα=∂U∂ξα,w=ξ.∇U−U. | (116) |
In the hodograph plane, the Omegaar coordinates are used (ξ1=RcosΘ,ξ2=RsinΘ) then this inverse of the mapping is given by:
x1=cosΘ∂U∂R−sinΘR∂U∂Θ,x2=sinΘ∂U∂R+cosΘR∂U∂Θ. | (117) |
The stress field satisfies:
τ1=σ13=μ(R)RcosΘ,τ2=σ23=μ(R)RsinΘ. | (118) |
The equation of motion is rewritten in the hodograph plane (R,Θ) using the mapping 117. The notation ()′ indicates a derivation with respect to R. The differential equation obtained in statics takes the expression
∂∂R(μ(R)R∂U∂R)+(μ(R)R)′R∂2U∂Θ2=0. | (119) |
This equation is homogeneous of degree one in U. Using 115 the displacement w is
w=R2∂∂R(UR). | (120) |
For the considered example, the geometry of the crack is the straight line −∞≤x1≤0. The direction of the line is e_1. The applied loading is that of a crack obtained in classical linearised elasticity or embedded in neo-Hookean material. For an inner point of view, the local behaviour is non-linear and the amount of shear is limited by a critical value Rm, as described by the constitutive law:
{RRo,τ=R,RoRRm,τ=Ro(RRo)α=ˆμRα,RmR,τ=0. | (121) |
Then for R≤Ro the differential equation is elliptic, for Rm≥R≥Ro ellipticity is ensured if α≥0 otherwise the differential equation becomes hyperOmegaic. We only consider here the ellipticity case.
For the considered constitutive law, solution in the hodograph plane is build by the combination of solution for the linear part of the constitutive law and solution for the non linear part with the help of peculiar solutions. The potential U is searched as [26,30]:
{0≤R≤Ro,U=Uo(R,Θ),Uo=AoRcosΘ∫RoRdtμot3−BoR(logRcosΘ−ΘsinΘ)+CoRcosΘ,Ro≤R≤Rm,U1=ˆU(R,Θ),U1=A1RcosΘ∫RoRdt^μotα+2−B1R(2αα+1logRcosΘ−ΘsinΘ)+C1RcosΘ. | (122) |
The constants (Ao,A1,Bo,B1,Co,C1) are determined by:
• (a) continuity conditions of the potential in R=Ro,
• (b) continuity of the displacement w everywhere,
• (c) asymptotic behaviour when R→0,
• (d) traction free- boundary conditions along the boundary of the damaged zone.
• (a) : Continuity of the potential implies that Bo=B1=B, and the relation
B2αα+1RologRo+C1Ro=BRologRo+CoRo. | (123) |
Then the displacement is given by
{0≤R≤Ro,w=wo(R,Θ)=−(AoμoR+BR)cosΘ,Ro≤R≤Rm,w=ˆw(R,Θ)=−(A1^μoRα+2Bαα+1R)cosΘ. | (124) |
• (b) Displacement is continuous en R=Ro
A1+μoRo2Bα−1α+1=Ao. | (125) |
• (c) Displacement satisfies the matching condition at ∞, so we obtain
Ao=K2πμo. | (126) |
• (d) The boundary along the damaged zone is traction free.
To simplify the expression, we adopt an adimensional formulation. Let us define new parameters (a,a1,b)
AoμoRo2=K2πτ2o=a,A1μoR2o=aa1,B=ba. | (127) |
We adopt Co=2−B+Blog(Ro) in order to determine the position of the quasi-crack along x2=0, and we consider the notation Xi=xia,ρ=RRo. Then ρo=1,ρm=RmRo.
The condition (126) is then rewritten as
a1+bα−1α+1=1. | (128) |
With these notations, the traction-free boundary condition (d) is now explained.
For that, the image of the hodograph plane is determined.
• 0≤ρ≤1,
X1=−12−12ρ2cos(2Θ)−blogρ,X2=−12ρ2sin(2Θ)+b(Θ−π2),W=−(1ρ+bρ)cosΘ. |
• 1≤ρ≤ρm,
X1=−(a1ρα+1+bα−1α+1)cos2Θ2−(a1ρα+1+b)α−12(α+1)−2bαα+1logρ−a1α+1,X2=−(a1ρα+1+bα−1α+1)sin2Θ2+b(Θ−π2),W=−(a1ρα+2αα+1bρ)cosΘ. |
We have moved the frame along e_2 with a shift of −bπ/2, to emphasize the symmetry of the geometry with respect to e_1. For ρ=ρa, the curve X1(ρa,Θ),X2(ρa,Θ) is a cycloid. With this geometry, the traction- free boundary condition is given by the equation
τ1dX2−τ2dX1=0. | (129) |
The boundary is decomposed into three parts: the two horizontal lines Θ=π/2±π/2 and the cycloid where ρ=ρm. The traction-free boundary condition is satisfied for Θ=π/2±π/2 for all ρ, and implies a relation along the cycloid where ρ=ρm:
(2bα+1−a1ρα+1m)cosΘ=0. | (130) |
This relation can be combined with (126), then
1=2bα+1(ρα+1m+α−12) | (131) |
It can be noticed that the thickness H of the damaged zone is H=baπ.
For ρm→∞, the quantity b vanishes, the quasi-crack tends to the classical crack, the thickness H vanishes.
The strain energy for the critical shear Rm is given by
ψ(Rm)=∫Rmoτ(t)dt=1μoτ2oα+1(ρα+1m+α−12). | (132) |
As a is a function of the equivalent loading K the relation 131 determines the thickness of the damaged zone
12aπ=hα+1(ρα+1m+α−12)=12K2τ2o. | (133) |
The flux of the energy along the cycloid is given by
Dd=hψ(Rm)=πaτ2o2μo=K22μo. | (134) |
This flux is exactly the energy release rate for the equivalent crack given by the condition of loading at infinity
G=J=hψ(Rm). | (135) |
For different constitutive relations, a numerical scheme is proposed in [31] to Omegave the problem of propagation of a quasi-crack for some specific classes of non linear elastic brittle materials.
The system consist of two sliding and contacting bodies Ω and Ω2 separated by a contact interface Ω3. The mechanical behaviour of the two bodies are known. The behaviour of the interface Ω3 is more complex. Particles are detached from sound Omegaids when some local criteria are satisfied at the boundary. Wear leads to geometrical changes and modification of contact conditions. Wear debris induce a specific layer. The interface is a complex medium made of particles; eventually of a lubricant fluid and of damaged zones inside the two bodies near the surface of contact. The interface here is considered at a macroscopic level as an homogeneous body obtained by an averaging process through the thickness H=2h of Ω3. This thickness is so small compared to the size of the contact zone and of the triOmegaogical system that the condition of such homogeneity can be acceptable.
During wear process, the boundaries Γ1=∂Ω∩∂Ω3 and Γ2=∂Ω2∩∂Ω3 eOmegave. During this process a flux of matter is lost from the materials and the framework developed previously can be applied. Along the two interfaces Γi energy is dissipated due to the motion of the boundaries
Di=∫Γiai[ψ]−n_i.σ.[v_]dS=∫ΓiGiaidS, | (136) |
and inside the Omegaume of the layer Ω3 a Omegaume dissipation dm due to damage, plasticity or viscosity, is also present. By integration over the thickness of the layer Ω3, dissipation by unit of contact area is defined as
Dm=∑iGijiai+∫Hdm(z)j(z)dz. | (137) |
A point of the layer Ω3 is defined relatively to middle surface S(X,t)=0 with normal n_. Then a point of the layer has coordinates (X,z) such that x=X+zn_(X). For x1∈Γ1,z1=h and for x2∈Γ2,z2=−h. The unit area of surface at distance z from S(X,t)=0 are dS(z)=j(z)dS and j1=j(h),j2=j(−h).
The two first terms of the dissipation are associated to the lost of material. The last term can be interpreted as local friction [8]
friction D3=∫Hdmdz=∫H(σ:˙ε−˙ψ3)dz. | (138) |
The main difficulty of the model is to choose the behaviour of the interface which takes into account of wear debris.
All mechanical quantities are defined with respect to the middle surface : S(X,t)=0 with normal n_. And Γ1,Γ2 are defined by
x1=X+h(X,t)v_,x2=X−h(X,t)v_. | (139) |
We denote a the velocity of S.
Da(x1)=a1v_1,Da(x2)=a2v_2. | (140) |
The displacement along the boundary Γi is continuous
[u_(X±h(X,t)n_,t)]Γi=0, | (141) |
and also the stress vector [σi−σ3].v_=0.
The internal state of the layer is defined by the total strain energy by unit of contact area
ψS(u_1,u_2,α)=1ρS∫h−hρ(x,z)ψ3(ε(x+zv_),α)j(x,z)dz, | (142) |
the surface density of mass
ρS=∫h−hρ(x,z)j(x,z)dz | (143) |
the total potential of dissipation
DS(v_1,v_2,˙α)=1ρS∫h−hρ(x,z)dm(˙ε(x+zv_),˙α)j(x,z)dz | (144) |
To build model of interface, the displacement inside the layer is decomposed as an expansion relatively to z.
u_=u_o+zu_1+z2u_2+... | (145) |
Up to order 1 we have
u_1=u_(X−hv_),u_2=u_(X+hv_) | (146) |
and
u_1−u_2=∇u_(X,0).v_h=u_1h,u_o=12(u_1+u_2) | (147) |
It can be shown that the equilibrium along the interface satisfies
σ.n_i=ρS(∂ψS∂u_i+∂DS∂v_i). | (148) |
The rigid punch has a vertical displacement upy=δ+x2/2R and we assume that wear occurs only in the half space. Ahead the punch there is no debris, then the Omegaume fraction of debris f=0. The thickness of the interface is Ho, which corresponds to the sum of roughness of the Omegaids and contains the thickness of the incompressible fluid. Due to wear, the thickness eOmegaves. The mass conservation and the fluid incompressibility give the relations between the wear rate ϕ(x), the fraction of debris f(x) and the thickness of the thin layer X(x) [32].All the equations of conservation are written in the moving frame with the punch at the velocity Ve_x.
The free energy of the mixture is given by
ψs(w,f)=k(f)12(wn)2+kt(f)(wt−αt)2. | (149) |
A potential of dissipation is given to determine the irreversible contribution, essentially due to viscosity
Ψs(˙w,˙αt)=12ηn(f)˙w2n+12ηt(f)˙w2t+12ηa(f)˙α2t. | (150) |
Then the local state equations are
n_.σ=k(f)wnn_+kt(f)(wt−αt)τ+ηn˙wnn_+ηt˙wtτ,A=kt(f)(wt−αt)=ηa(f)˙αt. | (151) |
whereA is the driving force associated to viscosity. This constitutive law generalizes the law use in ( [32]) in which ηa=kt=0. For (ηa=kt=0) we have an interface behaviour given by
σyy=k(f)wn,σxy=ηt(f)˙wt, | (152) |
k(f) and η(f) are chosen from typical homogenized value of the phases.
1k(f)=fKs+(1−f)Kf,ηt=ηo(1+2.5f), | (153) |
that the homogenized Reuss's model for the stiffness and the Einstein's law for the viscosity.
Introducing these equations in the equilibrium equation determines for a given profile ϕ(x) the answer of the elastic half space. The wear rate ϕ must satisfy a complementary law as proposed before. For the sake of simplicity we take
ϕ=kpσyy. | (154) |
The half plane has an elastic linear behaviour. In plane strain, the displacement (ux,uy) of the interface Γ2 is given by Omegaving the Galin's equations:
co1ux,x(x)=co2σyy+Vp1π∫a−aσxy(s)s−xdS, | (155) |
co1uy,x(x)=−co2σxy+Vp1π∫a−aσyy(s)s−xdS, | (156) |
where σyy,σxy are the component of the stress vector applied on the half plane, and coi are coefficients given in terms of Young's modulus E and Poisson's ratio ν. Vp is principal value of integral in sense of Cauchy.
co1=E2(1−ν2),co2=1−2ν2(1−ν) | (157) |
The solution is obtained analytically by an asymptotic expansion in series of the Omegaume fraction f of particles.
• At zero order, the Hertz's solution is recovered.
• At first order, a dependence with f is obtained. Wear occurs, and the profile of the pressure σyy(x) eOmegaves. The presence of viscous fluid induces a displacement of the maximum of pressure like under the dry contact with friction ( [33]).
This analytical solution is studied in paper [32].
In this article, we propose a framework using the motion of interface or surface to describe particular irreversible processes : damage, fracture, wear, delamination.
The main idea is the change in the strain energy when the degradation process occurs.
The model is based on the definition of the potential energy and to a complementary law using a normality rule based on the driving force associated to the motion of the interface. The evolution of the system is given by the revolution of a variational inequality. This formulation of the rate boundary value problem leads to discussion of stability of the equilibrium configuration and to possible bifurcation from this state.
Some applications of the concept and examples have illustrated the potentiality of the approach.
The author declares that there is no conflict of interest regarding the publication of this manuscript.
[1] |
Francfort GA, Marigo JJ (1998) Revisiting brittle fracture as an energy minimization problem. J Mech Phys Solids 46: 1319-1342. doi: 10.1016/S0022-5096(98)00034-9
![]() |
[2] |
Peerlings R, Geers M, de Borst R, et al. (2001) A critical comparison of non-local and gradientenhanced softening continua. Int J Solids Struct 38: 7723-7746. doi: 10.1016/S0020-7683(01)00087-7
![]() |
[3] |
Hakim V, Karma A (2009) A law of crack motion and phase-field models of fracture. J Mech Phys Solids 57: 342-368. doi: 10.1016/j.jmps.2008.10.012
![]() |
[4] |
Bourdin B, Francfort GA, Marigo JJ (2008) The variational approach to fracture. J Elasticity 91: 5-148. doi: 10.1007/s10659-007-9107-3
![]() |
[5] | Bui HD (1980) Solution explicite d'un problème de frontière libre en élastoplasticité avec endommagement. C R Acad des Sciences de Paris, série B 290: 345-348. |
[6] | Bui HD, Ehrlacher A (1980) Propagation dynamique d'une zone endommagée dans un solide élastique fragile en mode III et en régime permanent. C R Acad des Sciences de Paris, Série B 290 : 273-276. |
[7] |
Gurtin ME (1995) The nature of configurational forces. Arch Rat Mech 131: 67-100. doi: 10.1007/BF00386071
![]() |
[8] | Dragon-Louiset M, Stolz C (1999) A thermodynamical approach to contact wear. C R Acad des Sciences de Paris, S´erie IIb 327: 1275-1280. |
[9] | Bui HD, Dang Van Ky, Stolz C (1981) Variational-principles applicable to rate boundary-valueproblems of elastic-brittle solids with damaged zone. C R Acad des Sciences de Paris, Série II 292: 251-254. |
[10] |
Pradeilles-Duval RM, Stolz C (1995) Mechanical transformations and discontinuities along a moving surface. J Mech Phys Solids 43: 91-121. doi: 10.1016/0022-5096(94)00061-9
![]() |
[11] |
Stolz C (2007) Bifurcation of equilibrium solutions and defects nucleation. Int J Fracture 147: 103-107. doi: 10.1007/s10704-007-9147-5
![]() |
[12] |
Stolz C (2010) Thermodynamical description of running discontinuties: application to friction and wear. Entropy 12: 1418-1439. doi: 10.3390/e12061418
![]() |
[13] |
Stolz C, Moës N (2012) A new model of damage: a moving thick layer approach. Int J Fracture 174: 49-60. doi: 10.1007/s10704-012-9693-3
![]() |
[14] |
Abeyaratne R, Knowles J (1990) On the driving traction acting on a surface of strain discontinuity in a continuum. J Mech Phys Solids 38: 345-360. doi: 10.1016/0022-5096(90)90003-M
![]() |
[15] | Son QN, Pradeilles RM, Stolz C (1989) On a regularized propagation law in fracture and brittle damage. C R Acad des Sciences de Paris, Série II 309: 1515-1520. |
[16] |
Gurtin ME, Podio-Giugli P (1996) Configurational forces and the basic laws for crack propagation. J Mech Phys Solids 44: 905-927. doi: 10.1016/0022-5096(96)00014-2
![]() |
[17] |
Hill R (1986) Energy momentum tensors in elastostatics: some reflections on the general theory. J Mech Phys Solids 34: 305-317. doi: 10.1016/0022-5096(86)90022-0
![]() |
[18] |
Hutchinson J, Thouless MD, Liniger EG (1992) Growth and configurational stability of circular buckling driven film delamination. Acta Metall Mat 40: 295-308. doi: 10.1016/0956-7151(92)90304-W
![]() |
[19] | Pradeilles RM, Stolz C (1991) Sur le problème d'évolution des solides à changement de phase irréversibles. C R Acad des Sciences de Paris, Série II 313: 297-302. |
[20] |
Storakers B, Andersson B (1988) Non linear theory applied to delamination in composites. J Mech Phys Solids 36: 689-718. doi: 10.1016/0022-5096(88)90004-X
![]() |
[21] |
Stolz C (2011) Analysis of stability and bifurcation in non-linear mechanics with dissipation. Entropy 13: 332-366. doi: 10.3390/e13020332
![]() |
[22] |
Pradeilles-Duval RM (2004) Quasi-static evolution of delaminated structures: analysis of stability and bifurcation. Int J Solids Struct 41: 103-130. doi: 10.1016/j.ijsolstr.2003.07.006
![]() |
[23] | Stolz C. (2004), Energy methods in Non linear Mechanics, Lectures Notes 11,Warsawa, IPPT Pan |
[24] |
Neuber H (1968) A physically non linear notch and crack model. J Mech Phys Solids 16: 289-294. doi: 10.1016/0022-5096(68)90037-9
![]() |
[25] | Rice J (1966) Contained plastic deformation near cracks and notches under longitudinal shear. Int J Fracture Mech 2: 426-447. |
[26] |
Stolz C (2015) Asymptotic fields ahead a crack for a class of non-linear materils under mode III. Mechanics of Materials 90: 102-110. doi: 10.1016/j.mechmat.2015.04.005
![]() |
[27] |
Knowles JK, Sternberg E (1980) Discontinuous deformation gradients near the tip of a crack in finite anti-plane shear: an example. J Elasticity 10: 81-110. doi: 10.1007/BF00043136
![]() |
[28] |
Rice J (1967) Stresses due to sharp notch in a work hardening elastic plastic material loaded by a longitudinal shear. J Appl Mech 34: 287-298. doi: 10.1115/1.3607681
![]() |
[29] |
Knowles JK, Sternberg E (1981) Anti-plane shear fields with discontinuous gradients near the tip of a crack in finite elastostatics. J Elasticity 11: 129-164. doi: 10.1007/BF00043857
![]() |
[30] | Stolz C (2010) Closed form solution for the finite anti-plane shear field for a class of incompressible brittle solids. Comptes Rendus Mécanique 338-12: 663-669. |
[31] |
Stolz C, Parrilla-Gomez A (2015) Antiplane shear field for a class of hyperelastic incompressible brittle material: analytical and numerical approaches. J Mech Mater Struct 10: 395-410. doi: 10.2140/jomms.2015.10.395
![]() |
[32] |
Dragon-Louiset M (2001) On a predictive macroscopic contact-sliding wear model based on micro-mechanical consideration.. Int J Solids Struct 38: 1625-1639. doi: 10.1016/S0020-7683(00)00065-2
![]() |
[33] | Bui HD, Dang Van K (1976) Contribution à l'étude théorique du contact élastique d'un frotteur cylindrique glissant sur un massif élastique. Industrie Minérale, No Sp´ecial Rh´eologie IV: 1-8. |