Citation: Stefano Almi, Giuliano Lazzaroni, Ilaria Lucardesi. Crack growth by vanishing viscosity in planar elasticity[J]. Mathematics in Engineering, 2020, 2(1): 141-173. doi: 10.3934/mine.2020008
[1] | Manuel Friedrich . Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005 |
[2] | Hitoshi Ishii . The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 1: linear coupling. Mathematics in Engineering, 2021, 3(4): 1-21. doi: 10.3934/mine.2021032 |
[3] | Caterina Ida Zeppieri . Homogenisation of high-contrast brittle materials. Mathematics in Engineering, 2020, 2(1): 174-202. doi: 10.3934/mine.2020009 |
[4] | Paola F. Antonietti, Chiara Facciolà, Marco Verani . Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017 |
[5] | Miyuki Koiso . Stable anisotropic capillary hypersurfaces in a wedge. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029 |
[6] | Hitoshi Ishii . The vanishing discount problem for monotone systems of Hamilton-Jacobi equations: a counterexample to the full convergence. Mathematics in Engineering, 2023, 5(4): 1-10. doi: 10.3934/mine.2023072 |
[7] | Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043 |
[8] | Matteo Novaga, Marco Pozzetta . Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024 |
[9] | Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi . A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014 |
[10] | Ko-Shin Chen, Cyrill Muratov, Xiaodong Yan . Layered solutions for a nonlocal Ginzburg-Landau model with periodic modulation. Mathematics in Engineering, 2023, 5(5): 1-52. doi: 10.3934/mine.2023090 |
In many applications of engineering, it is crucial to predict the propagation of fracture in structures and to understand whether cracks are stable. When the external loading is very slow if compared with the time scale of internal oscillations (such as in a building in standard conditions), it is possible to ignore inertia and to assume that the system is always at equilibrium: The resulting model is called quasistatic. Quasistatic (or rate-independent) processes have been extensively analyzed in the mathematical literature both in the context of fracture and of other models (see [44] and references therein).
The first difficulties in modeling fracture are related to identifying equilibrium configurations. In fact, in order to state that a configuration is stable, one would have to use a derivative of the mechanical energy with respect to the crack set, which is not well defined. Thus one may prefer a derivative-free formulation where equilibria are restricted to global minimizers (of the sum of the mechanical energy and of the dissipated energy due to crack growth), in the context of energetic solutions to rate-independent systems, see, e.g., [7,15,16,19,24,25,26].
A second approach allows one to take into account of more equilibria by restricting the set of the admissible cracks. In fact, the problem is to select a class of regular curves and to prove the existence of a derivative of the mechanical energy with respect to the elongation of a crack in that class. The opposite of this derivative is called energy release rate and represents the gain in stored elastic energy due to an infinitesimal crack growth. Griffith's criterion [27] allows crack growth only when the energy release rate reaches the toughness of the material (i.e., the energy spent to produce an infinitesimal crack).
In this context, some existence results for crack evolution were first given in the case of a prescribed crack, i.e., before the evolution starts one already knows the set which is going to crack, see [33,48] in linear elasticity and [36] for a nonlinear model. An algorithm for predicting a stable crack path (chosen from a class of regular curves) was proposed in [38,40] in the case of antiplane linear elasticity, where the deformation is represented by a scalar function (that is the vertical displacement, depending on the two horizontal components, while the horizontal displacement is zero). This was extended in [12] to a class of curves with branches and kinks.
In this paper we prove an existence result for crack evolution based on Griffith's criterion, in the context of planar linear elasticity, in dimension two, as in [33]. In this case the displacement is a vector (with two components). Differently from [33], in our model the path followed by the crack is not a priori known. In fact, the crack is assumed to be the union of a fixed number of C1,1 curves and is selected among a class of (unions of) curves with bounded curvature, with no self-intersections, and with at most one point meeting the boundary of the domain (in the reference configuration). Some geometric constraints guarantee that this class is compact with respect to the Hausdorff convergence of sets. The same class of admissible cracks was employed in [38].
In order to write the flow rule for crack propagation, we need the expression of the energy release rate. The first step is by now standard and requires to prove that, when the crack is a prescribed curve, then the mechanical energy (i.e., the sum of the stored elastic energy and of the work of external volume and surface forces) is differentiable with respect to the arc length of the curve, and its derivative can be written as a surface integral depending on the deformation gradient. This is done in Proposition 3.1, by adapting to our framework (C1,1 crack, nonconstant elasticity tensor, volume forces) the classical techniques of [21,28], as done e.g., in [30,32,33,39]. Since we want a model predicting the crack path (not prescribed a priori), we need to prove that the energy release rate is independent of the extension of the crack (in the class of C1,1 curves). This is a crucial result in our analysis, shown in Theorem 3.6. Moreover, the energy release rate is continuous with respect to the Hausdorff convergence of cracks (see Remark 3.11). When there are more curves, there is an energy release rate for each crack tip.
Proving such properties of the energy release rate(s) is fundamental to study quasistatic crack evolution and is the major technical difficulty of this work. In fact, the strategy of the proof differs from the method used in the corresponding results in the antiplane case, cf. [38,39]. In planar elasticity, assuming that the crack is C∞, that there are no external forces, and that the elasticity tensor is constant, it was proven in [3] (see also [11,20]) that each energy release rate can be expressed in terms of two stress intensity factors, which characterize the singularities of the elastic equilibrium; since the stress intensity factors only depend on the current crack, it turns out that the energy release rate is independent of the crack's extension. In this paper we need a corresponding property for C1,1 cracks (in the class where we have compactness with respect to Hausdorff convergence) and for energies with external forces and nonconstant elasticity tensor. The same strategy of the antiplane case does not apply to the nonsmooth case, in particular we do not prove the existence of the stress intensity factors; nonetheless, we prove that the energy release rate is stable under Hausdorff convergence in the class of C1,1 cracks, so we can employ the results of [3] via some approximation arguments (see Section 3). For this reason we need a version of Korn's inequality in cracked domains, where a sequence of converging crack paths is given and the constant of Korn's inequality is uniform with respect to the cracked domains (Section 2).
We remark that a recent paper [29], extending [6], shows the independence of the energy release rates from the crack's extension with different methods and under stronger regularity assumptions on the cracks (which are required to be H3 curves). For an account on the wide literature regarding energy release rates and stress intensity factors we mention e.g., [4,18,30,31,45,46] and the references therein. Moreover, we point out that an energy release rate associated with a crack tip does exist also under much weaker regularity conditions on the crack set. For instance, the results of [5] apply to cracks that are merely closed and connected. However, in this setting energy release rates can be characterized just up to subsequences through a blow-up limit, thus uniqueness is not guaranteed and, ultimately, the independence on extensions may not hold. On the other hand, the results of [8] do not have this limitation, but the initial crack needs to be straight, which makes it impossible to use such characterization in the context of an evolution problem. (We also refer to [10] for related results in antiplane elasticity.) For these reasons in this paper we resort to the class of (unions of) C1,1 cracks where, as mentioned, better properties can be proven.
This allows us to employ the well known vanishing viscosity method for finding balanced viscosity solutions to rate-independent systems, see [12,33,38] for brittle fracture in linear elasticity and [44] for further references. We fix a time discretization and solve some incremental problems where we minimize the sum of the mechanical energy and of the dissipated energy in the class of non prescribed C1,1 cracks. Notice that in the present work the dissipated energy density is nonconstant and depends on the position of the crack tip in the reference configuration. In the minimum problems, the total energy is perturbed with a term penalizing brutal propagations between energy wells, multiplied by a parameter ε. Passing to the continuous time, we obtain a viscous version of Griffith's criterion, with a regularizing term multiplied by ε; a second passage to the limit as ε→0 leads to rate-independent solutions. It is also possible to characterize the time discontinuities of the resulting evolution using the reparametrization technique first proposed in [22] and then refined in [41,42,43,47]: Thanks to these methods we can also treat multiple non interacting cracks.
In our paper we extend the results of [38] to planar elasticity and the results of [33] to C1,1, non prescribed cracks. Our main outcome is the existence of a quasistatic evolution (more precisely, a balanced viscosity evolution) fulfilling Griffith's criterion: The length of each component of the crack is a nondecreasing function of time; at all continuity points of these functions, the energy release rate at each tip is less than or equal to the material's toughness at that tip (which is a stability condition); the length is increasing only if the energy release rate reaches the toughness. Moreover, time discontinuities (corresponding to brutal propagation) can be interpolated by a transition, characterized by a viscous flow rule, where the energy release rates are larger than or equal to the toughness (see e.g., [13,14,34,35] for corresponding results in damage and plasticity).
Notation
Given two vectors a,b∈Rd, their scalar product is denoted by a⋅b. We set Md the space of d×d square matrices, and we denote by Mdsym and Mdskw the subsets of symmetric and skew-symmetric ones, respectively. We set I the identity matrix in Md. Given A and B in Md, we write A:B to denote their Euclidean scalar product, namely A:B=AijBij. Here and in the rest of the paper we adopt the convention of summation over repeated indices. For every p≥1 we define the p-norm in Rd as |x|p:=(∑di=1|xi|p)1/p. The 2-norm will be simply denoted by |⋅|. The latter induces the distance dist(C,D):=inf{|x−y|:x∈C,y∈D} between two sets C and D. The maximal distance between two points of a set E, namely its diameter, is denoted by diam(C).
The symbol Bρ(x) denotes the open ball of radius ρ in R2, centred at x. The support of a function f, namely the closure of {f≠0}, is denoted by spt(f). For a tensor field V∈C1(Rd;Md), by divV we mean its divergence with respect to lines, namely (divV)i:=∂jVij. The symmetric gradient of a vector field u∈C1(Rd;Rd) is denoted by Eu, namely (Eu)ij:=(∂iuj+∂jui)/2.
We adopt standard notations for Lebesgue and Sobolev spaces on a bounded open set of Rd. The boundary values of a Sobolev function are always intended in the sense of traces. Boundary integrals on Lipschitz curves are done with respect to the 1-dimensional Hausdorff measure H1. Given an interval I⊂R and a Banach space X, Lp(I;X) is the space of Lp functions from I to X. Similarly, the sets of continuous and absolutely continuous functions from I to X are denoted by C0(I;X) and AC(I;X), respectively. Derivatives of functions depending on one variable are denoted by a prime or, when the variable is time, by a dot.
The identity map in a vector space is denoted by id. Given a normed vector space X the norm in X is denoted by ‖⋅‖X. We adopt the same notation also for vector valued functions in X. For brevity, the norm in Lp over an open set Ω of Rd is denoted by ‖⋅‖p,Ω or, when no ambiguity may arise, simply by ‖⋅‖p.
We describe a crack model in planar elasticity for a brittle body. The body is represented in its reference configuration by an infinite cylinder Ω×R, where Ω⊂R2 is a bounded connected open set, with Lipschitz boundary. By assumption, the displacement u produced by the external loading is horizontal and depends only on the two horizontal components: The deformation is then given by
Ω×R∋(x1,x2,x3)↦(x1+u1(x1,x2),x2+u2(x1,x2),x3),where u=(u1,u2):Ω→R2. |
The set of possible discontinuity points of u (the crack) is assumed to lie in a class of admissible regular cracks. We now define such class following [38,39]. It depends on a parameter η>0 that is thought as small, but is fixed throughout the paper.
Definition 1.1. Fixed η>0, the set R0η contains all closed subsets Γ⊆¯Ω such that
(a) Γ is a union of a finite number of arcs of C1,1 curves, each of them intersecting ∂Ω in at most one endpoint,
(b) H1(Γ∩Ω)>0 and Ω∖Γ is a connected open set, union of a finite number of Lipschitz domains,
(c) for every x∈Γ there exist two open balls B1η,B2η⊆R2 of radius η such that
¯B1η∩¯B2η={x} and (B1η∪B2η)∩Γ=∅. | (1.1) |
Furthermore, we denote with R0,1η the class of curves Γ∈R0η such that Γ is one arc of curve of class C1,1 intersecting ∂Ω in exactly one endpoint.
Notice that R0η⊆R0η′ if η>η′. The role of (1.1) is twofold: On the one hand it gives a uniform bound (depending on 1/η) on the curvature of each connected component of any set Γ∈R0η, on the other hand it ensures that each of these components is an arc of a simple curve, i.e., a curve with no self-intersections. Because of (a), each of the arcs has one or two endpoints contained in Ω; we say that these points are the crack tips.
Since quasistatic models are in general unable to predict crack initiation [9], i.e., nucleation of a new crack from sound material, we assume that there is an initial crack Γ0∈R0η. Each connected component of an admissible crack Γ will be the extension of a connected component of Γ0, starting from its crack tips. Let M be the number of crack tips of Γ0; notice that M may be larger than the number of connected components of Γ0. We parametrize Γ0 by introducing M injective functions γm of class C1,1, for m=1,…,M, in the following way:
● If a connected component Γm0 of Γ0 intersects ∂Ω in a single endpoint x0, we consider its arc-length parametrization γm:[0,H1(Γm0)]→Γm0 such that γm(0)∈∂Ω and γm(H1(Γm0)) is the crack tip of Γm0. In particular, a crack in R0,1η has exactly one tip.
● If a connected component of Γ0 is contained in Ω, we see it as the union of two curves Γm0, Γm+10, intersecting at a single point ˉx (that is not a tip of Γ0); then we consider two arc-length parametrizations γm:[0,H1(Γm0)]→Γm0, γm+1:[0,H1(Γm+10)]→Γm+10, such that γm(0)=γm+1(0)=ˉx and γm(H1(Γm0)) and γm+1(H1(Γm+10)) are the two crack tips.
We then have Γ0=⋃Mm=1Γm0. Analogous parametrizations will be used for the extensions of Γ0. In the next definition, M is the number fixed above.
Definition 1.2. The set Rη contains all subsets Γ∈R0η such that
(d) Γ is the union of M connected subsets Γ1,…,ΓM, such that any two of them intersect in up to a point,
(e) Γm⊇Γm0 for every m=1,…,M,
(f) for every m=1,…,M and for every x∈Γm∖Γm0
B2η(x)∩(∂Ω∪⋃l≠mΓl)=∅. |
Given a set Γ=⋃Mm=1Γm∈Rη, we extend the functions γm, m=1,…,M, defined above, to arc-length parametrizations γm:[0,H1(Γm)]→Γm; it turns out that they are injective and of class C1,1. Properties (a)–(f) ensure that the class Rη is sequentially compact with respect to Hausdorff convergence (see the next section for details), induced by the following distance.
Definition 1.3. Given two compact subsets Γ,Γ′⊂¯Ω, their Hausdorff distance is given by
dH(Γ′;Γ):=max{supx∈Γ′dist(x,Γ), supx∈Γdist(x,Γ′)}, |
with the conventions dH(x;∅)=diamΩ and sup∅=0. A sequence (Γn)n∈N of compact subsets of ¯Ω converges to Γ in the Hausdorff metric if dH(Γn;Γ)→0 as n→∞.
Remark 1.4. There are choices of Γ0 such that Rη contains no elements different from Γ0: We mention a few examples with Ω=[−1,1]2. Let Γ0=[−1,0]×{0,14}: Then Γ0∈R0η only if η≤1/8, thus Rη=∅ if and only if η>1/8. If instead Γ0=[−1,0]×{0}, we have Rη={Γ0} if and only if η≥1/2. However, given Γ0 such that Rη is trivial, one can find η′<η such that Rη′ contains nontrivial extensions of Γ0. Starting from an initial crack with nontrivial extensions, the model described in this paper is reliable as long as our algorithm finds a current configuration Γ(t) such that there are nontrivial extensions. If, during the evolution, some tip becomes (2η)-close to ∂Ω or to other connected components of the crack, the results should not be regarded as meaningful.
Since the body is brittle, the uncracked part Ω∖Γ is elastic; we assume that the displacements are small (so we adopt the setting of linear elasticity) and the crack is traction-free. We look for evolutions in the time interval [0,T], produced by the time-dependent external loading:
(H1) a boundary condition w∈AC([0,T];H1(Ω∖Γ0;R2)), to be satisfied on a relatively open subset ∂DΩ of ∂Ω, with a finite number of connected components,
(H2) a volume force f∈AC([0,T];L2(Ω∖Γ;R2)) and a surface force g∈AC([0,T];L2(∂SΩ;R2)), where ∂SΩ is a relatively open subset of ∂Ω such that ∂SΩ⋐∂Ω∖¯∂DΩ.
Without loss of generality, we assume that spt(w)⊆{x∈¯Ω : dist(x,∂Ω)≤η}, so w≡0 around any crack tip.
At each point x∈Ω, the stress tensor is C(x):M2sym→M2sym, where
(H3) C(x)A=λ(x)tr(A)I+2μ(x)A for every A∈M2sym, with λ,μ∈C0,1(¯Ω) such that μ(x)>0 and λ(x)+μ(x)>0 for every x∈¯Ω.
Notice that the standard conditions μ(x)>0 and λ(x)+μ(x)>0 ensure the positive definiteness of C(x), uniformly in x.
Given t∈[0,T] and Γ∈Rη, the minimum problem
min{12∫Ω1ΓCEv:Evdx−∫Ω∖Γf(t)⋅vdx−∫∂sΩg(t)⋅vdH1:v∈H1(Ω∖Γ;R2),v=w(t) on ∂DΩ} | (1.2) |
has a unique solution, denoted by u(t;Γ):Ω∖Γ→R2, with elastic energy
E(t;Γ):=12∫Ω∖ΓCEu(t;Γ):Eu(t;Γ)dx−∫Ω∖Γf(t)⋅u(t;Γ)dx−∫∂SΩg(t)⋅u(t;Γ)dH1. |
According to the assumption of brittle behavior, in order to produce a crack the system employs an energy depending (only) on the geometry of the crack itself, in the context of Griffith's theory [27]. The total energy of the configuration corresponding to a crack Γ at time t is
F(t;Γ):=E(t;Γ)+K(Γ),K(Γ):=∫ΓκdH1, |
where the surface energy density satisfies
(H4) κ∈C0(¯Ω;[κ1,κ2]),
where 0<κ1<κ2.
Starting from the initial condition Γ0 fixed above, we define a discrete-time evolution of stable states by solving some incremental minimum problems. For every k∈N we consider a subdivision of the time interval [0,T] in nodes {tk,i}0≤i≤k such that
0=tk,0<tk,1<⋯<tk,k=Tandlimk→∞max1≤i≤k(tk,i−tk,i−1)=0. | (1.3) |
Fixed ε>0, we define by recursion the sets Γε,k,i, i=0,…,k, as follows. We set Γε,k,0:=Γ0; for i≥1, Γε,k,i is a solution to the minimum problem
min{E(tk,i;Γ)+H1(Γ)+ε2M∑m=1H1(Γm‖Γmε,k,i−1)2tk,i−tk,i−1:Γ∈Rη,Γ⊇Γε,k,i−1}, | (1.4) |
where the role of the term multiplied by ε is to penalize transitions between energy wells. The existence of solutions to (1.4) is proven in Corollary 2.5 exploiting the compactness properties of Rη with respect to the Hausdorff convergence, see Section 2 for details.
We define a piecewise constant interpolation on [0,T] by
Γε,k(0):=Γ0,Γε,k(t):=Γε,k,ifor t∈(tk,i−1,tk,i]. | (1.5) |
The unilateral constraint Γ⊇Γε,k,i−1 in (1.4) enforces irreversibility of the crack growth, indeed the set function t↦Γε,k(t) is nondecreasing with respect to the inclusion.
Passing to the limit as k→∞ and exploiting again the compactness of Rη, we obtain a time-continuous evolution t↦Γε(t). In order to understand its properties, we need to define the energy release rate associated to a crack.
For simplicity, let us first consider the case of a prescribed curve with only one tip. Given an increasing family of cracks Γσ∈R0,1η parametrized by their arc length σ∈[0,S], we will prove that the map σ↦E(t;Γσ) is differentiable for every fixed t. Moreover, we will show that the derivative only depends on the current configuration Γs, and not on its possible extensions, i.e., if Γσ=ˆΓσ for σ≤s, then
dE(t;Γσ)dσ|σ=s=dE(t;ˆΓσ)dσ|σ=s. |
In particular, we are allowed to write −dE(t;Γσ)dσ|σ=s=:G(t;Γs) with no ambiguity. The quantity G(t;Γs) is the energy release rate corresponding to the crack Γs and represents the (partial) derivative of the energy E with respect to variations of crack in the set of all admissible curves R0,1η larger than Γs. For the details of these results, we refer to Section 3 below.
In the case of a curve with several connected components Γ∈Rη, for every tip indexed by m we define the m-th energy release rate Gm(t;Γ) as above, with respect to variations of the sole component Γm of Γ. The energy release rate will be in this case a vector G(t;Γ):=(G1(t;Γ),…,GM(t;Γ)).
The properties of the evolution t↦Γε(t) are summarized in the next proposition, whose proof is postponed to Section 4.
Proposition 1.5. Fix η>0, Γ0∈R0η, and ε>0. Assume (H1)–(H4). Let Γε,k be as in (1.5). Then there are a subsequence (not relabeled) of Γε,k and a set function t↦Γε(t)∈Rη such that Γε,k(t) converges to Γε(t) in the Hausdorff metric for every t∈[0,T].
Set Γε(t)=⋃Mm=1Γmε(t), with the conventions of Definition 1.2, and lmε(t):=H1(Γmε(t)). Then for every m=1,…,M and for a.e.\ t∈[0,T]
(G1) ε ˙lmε(t)≥0;
(G2) ε κ(Pmε(t))−Gmε(t)+ε˙lmε(t)≥0;
(G3) ε ˙lmε(t)[κ(Pmε(t))−Gmε(t)+ε˙lmε(t)]=0,
where Gmε(t) is the energy release rate corresponding to Γmε(t).
Moreover, along a suitable ε-subsequence, ε‖˙lmε‖22 is bounded uniformly w.r.t. ε.
Properties (G1)ε–(G3)ε show that the term multiplied by ε in (1.4) has a regularizing effect. Indeed, the flow rule for the evolution of lε:=(l1ε,…,lMε) features a time derivative of the unknown. For this reason the corresponding solutions are called viscous.
In the passage to the limit as ε→0, such viscous regularizing term vanishes, so the system follows an evolution of stable states. We thus obtain a balanced viscosity evolution. The next result is proven in Section 4.
Theorem 1.6. Fix η>0 and Γ0∈R0η. Assume (H1)–(H4). For every ε>0, let Γε be the evolution found in Proposition 1.5. Then there are a subsequence (not relabeled) of Γε and a set function t↦Γ(t)∈Rη such that Γε(t) converges to Γ(t) in the Hausdorff metric for every t∈[0,T].
Set Γ(t)=⋃Mm=1Γm(t), with the conventions of Definition 1.2, and lm(t):=H1(Γm(t)). Then for every m=1,…,M
(G1) for a.e. t∈[0,T], ˙lm(t)≥0;
(G2) for every t∈[0,T] of continuity for lm, κ(Pm(t))−Gm(t)≥0;
(G3) for a.e. t∈[0,T], ˙lm(t)[κ(Pm(t))−Gm(t)]=0,
where Gm(t) is the energy release rate corresponding to Γm(t).
Properties (G1)–(G3) are a formulation of Griffith's criterion for crack growth and show the stability of the evolution t↦l(t):=(l1(t),…,lM(t)) in its continuity points. However, the function t↦l(t) may have discontinuities and Theorem 1.5 does not provide a characterization of jumps in time. The existence result is refined in the following theorem, where we show that there are a reparametrization of the time interval and a parametrized evolution, continuous in time, that interpolates l and follows a viscous flow rule in the intervals corresponding to the discontinuities of l. The next theorem is proven in Section 5.
Theorem 1.7 (Griffith's criterion). Fix η>0 and Γ0∈R0η. Assume \textnormal{(H1)–(H4)}. There are absolutely continuous functions ˜t:[0,S]→[0,T] and ˜Γm:[0,S]→Rη, m∈{1,…,M}, such that for a.e. σ∈[0,S], setting ˜Γ(σ)=⋃Mm=1˜Γm(σ), with the conventions of Definition 1.2, and ˜lm(σ):=H1(˜Γm(σ)),
(pG1) ˜t′(σ)≥0 and (˜lm)′(σ)≥0 for every m=1,…,M;
(pG2) if ˜t′(σ)>0, then ˜Gm(σ)≤κ(˜Pm(σ)) for every m=1,…,M;
(pG3) if ˜t′(σ)>0 and (˜lm)′(σ)>0 for some m∈{1,…,M}, then ˜Gm(σ)=κ(˜Pm(σ));
(pG4) if ˜t′(σ)=0, then there is m∈{1,…,M} such that (˜lm)′(σ)>0; moreover, for every m with this property, we have ˜Gm(σ)≥κ(˜Pm(σ)),
where ˜Gm(σ) is the energy release rate corresponding to ˜Γm(σ). Moreover, denoting with ˜u(σ) the solution of (1.2) at time ˜t(σ) with a crack ˜Γ(σ), for every s∈[0,S] it holds
F(˜t(s);˜Γ(s))= F(0;Γ0)+∫s0∫ΩCE˜u(σ):E˙w(˜t(σ))˜t′(σ)dxdσ−M∑m=1∫s0(˜Gm(σ)−κ(˜Pm(σ)))(˜lm)′(σ)dσ−∫s0∫Ω˙f(˜t(σ))⋅˜u(σ)˜t′(σ)dxdσ−∫s0∫Ωf(˜t(σ))⋅˙w(˜t(σ))˜t′(σ)dxdσ−∫s0∫∂SΩ˙g(˜t(σ))⋅˜u(σ)˜t′(σ)dH1dσ−∫s0∫∂SΩg(˜t(σ))⋅˙w(˜t(σ))˜t′(σ)dH1dσ. | (1.6) |
Finally,
if ˜t′(σ)>0,then ˜Γ(σ)=Γ(˜t(σ)), |
where Γ is the balanced viscosity evolution found in Theorem 1.5.
In this section we collect some properties of the class of admissible cracks Rη and of the associated displacements. We recall that, given a crack, the associated displacement is the unique solution to the corresponding minimum problem (1.2).
As already mentioned in the previous section, the elements of Rη have no self-intersections, and, during the evolution, their crack tips stay uniformly far from the boundary and from the other connected components of the crack set. Moreover, it is easy to show that the curvature and the H1 measure of the elements of Rη are controlled from above by η−1 and by some constant C(Ω,Γ0,η), respectively. Finally, as proven in [39, Proposition 2.9 and Remark 2.10], the class of admissible cracks Rη is sequentially compact with respect to the Hausdorff convergence introduced in Definition 1.3.
Theorem 2.1. Every sequence (Γn)n∈N⊂Rη admits (up to a subsequence) a limit Γ∞∈Rη in the Hausdorff metric. Moreover, along the subsequence (not relabeled), we have H1(Γn)→H1(Γ) as n→∞.
In what follows we show the continuity of the elastic energy E w.r.t. Hausdorff convergence of the crack set Γ∈Rη. This will in particular imply the existence of solutions for the incremental minimum problems (1.4).
We start with recalling in Proposition 2.2 a Korn inequality for Ω∖Γ. In Proposition 2.3, instead, we show that, along sequences of cracks Γn∈Rη converging in the Hausdorff metric, such an inequality is independent of n. The study is carried out disregarding the time variable, which for brevity is omitted. Accordingly, the elastic energy associated to a fracture Γ writes E(Γ).
Proposition 2.2. Let Γ∈Rη. Then, there exists a positive constant C=C(Ω,Γ) such that for every u∈H1(Ω∖Γ;R2)
‖∇u‖2≤C(‖u‖2+‖Eu‖2). |
Proof. Being Ω∖Γ connected by arcs (see Definition 1.2), it is possible to fix ˆΓ⊃Γ such that Ω∖ˆΓ is the union of N disjoint open sets Ωi with Lipschitz boundaries ∂Ωi such that H1(∂DΩ∩∂Ωi)>0 for i∈{1,…,N}, and apply the usual Korn inequality to u restricted to Ωi.
Proposition 2.3. Let Γn,Γ∞∈Rη be such that Γn converges to Γ∞ in the Hausdorff metric as n→∞. Then, there exists a positive constant C=C(Ω) (independent of n) such that for n sufficiently large
‖∇u‖2≤C(‖u‖2+‖Eu‖2)for every u∈H1(Ω∖Γn;R2). | (2.1) |
Moreover, for u∈H1(Ω∖Γn;R2) with u=0 H1-a.e. on ∂DΩ we have
‖∇u‖2≤C‖Eu‖2and‖u‖2≤C‖Eu‖2. | (2.2) |
Proof. At least for n sufficiently large, we may assume that there exists an extension ˆΓn of Γn such that Ω∖ˆΓn=⋃Ni=1Ωni, where Ωni (i=1,…,N) are open bounded disjoint sets with Lipschitz boundaries and Lipschitz constant L independent of n. Moreover, we can assume that H1(∂DΩ∩∂Ωni)>0 for i∈{1,…,N} and every n. The same construction can be repeated for n=∞ in such a way that Ωni converges to Ω∞i in the Hausdorff metric as n→∞.
Let us now fix Ω′⋐Ω∞1. For n large enough (including the case n=∞), we have that Ω′⋐Ωn1. Hence, applying Proposition 2.2 in Ω′ we deduce that there exists a positive constant C′ independent of n such that
‖∇u‖2,Ω′≤C′(‖u‖2,Ω′+‖Eu‖2,Ω′)for every u∈H1(Ω∖Γn;R2). | (2.3) |
Since Ωn1 and Ω∞1 share the same Lipschitz constant L, applying locally, close to the boundary of Ωn1 (resp. Ω∞1), the results of [23, Theorem 4.2], we also obtain that there exists a positive constant ˜C such that
‖∇u‖2,Ωn1∖¯Ω′≤˜C(‖u‖2,Ωn1∖¯Ω′+‖Eu‖2,Ωn1∖¯Ω′)for every u∈H1(Ω∖Γn;R2). | (2.4) |
The same inequality can be proven for Ωni, i≥2. Therefore, combining (2.3) and (2.4) we get (2.1) for some positive constant C independent of n∈N∪{∞}, n large enough.
To prove (2.2) it is enough to show that
‖u‖2≤C‖Eu‖2for every u∈H1(Ω∖Γn;R2) with u=0 H1-a.e. on ∂DΩ, n large enough. | (2.5) |
We proceed with the usual contradiction argument. Assume that (2.5) is false. Then, for every k∈N there exist nk>nk−1 and uk∈H1(Ω∖Γnk;R2) such that ‖uk‖2>k‖Euk‖2. Without loss of generality, we may assume that ‖uk‖2=1. By (2.1) we deduce that ‖∇uk‖2 is bounded. Hence, up to a subsequence, ∇uk⇀φ weakly in L2(Ω;M2) and uk→u in L2loc(Ω∖Γ∞;R2), which implies that u∈H1loc(Ω∖Γ∞;R2) with ∇u=φ. Since φ∈L2(Ω;M2), applying [17, Proposition 7.1] we deduce that u∈H1(Ω∖Γ∞;R2). Since Euk converges to 0 in L2(Ω;M2), we get that Eu=0 in Ω. Thus, u is a rigid movement in Ω, i.e., there exist A∈M2skw and b∈R2 such that u=Ax+b for x∈Ω. Moreover, setting Ωη:={x∈Ω:dist(x,∂Ω)<η}, by Definition 1.2 we have (Γnk∩Ωη)∖Γ0=∅ and uk⇀u in H1(Ωη∖Γ0;R2). Therefore, u=0 H1-a.e. on ∂DΩ, which implies that u=0. We claim that ‖uk‖2→‖u‖2. Indeed, ‖uk‖2,Ω′→‖u‖2,Ω′ for every Ω′⋐Ω∖Γ∞. By a simple reflection argument applied on both sides of the crack set Γnk, we instead obtain that ‖uk‖2,Ω∖¯Ω′→‖u‖2,Ω∖¯Ω′. Thus, 1=‖uk‖2→‖u‖2=0, which is a contradiction. This concludes the proof of (2.2).
We are now ready to prove the continuity of the energy E with respect to the crack set. The following lemma is actually stated in a more general setting. Indeed, we show the continuity of the displacement u solution of (1.2) not only w.r.t. the Hausdorff convergence of sets in Rη, but also w.r.t. the data of the problem, i.e., the applied forces, the boundary datum, and the elasticity tensor. Such a continuity result will be useful in the next section, where we prove the differentiability of E w.r.t. crack elongations by using some approximations. From now on, when explicitly needed, we highlight the dependence on the data by writing E(f,g,w,C;Γ) for E(Γ).
Lemma 2.4. Let fn,f∞∈L2(Ω;R2), gn,g∞∈L2(∂SΩ;R2), wn,w∞∈H1(∂SΩ;R2), Cn,C∞∈C0,1(¯Ω), Γn,Γ∞∈Rη, and n∈N be such that fn→f∞ strongly in L2(Ω;R2), wn→w∞ in H1(Ω∖Γ0;R2), gn⇀g∞ weakly in L2(∂SΩ;R2), Cn→C∞ uniformly in ¯Ω, and Γn→Γ∞ in the Hausdorff metric, as n→∞.
Then, the energies E(fn,gn,wn,Cn;Γn) converge to E(f∞,g∞,w∞,C∞;Γ∞) in the limit as n→∞. Moreover, the corresponding displacements un and u∞, solutions to the associated minimum problems (1.2), satisfy ∇un→∇u∞ strongly in L2(Ω;M2).
Proof. The proof is carried out following the steps of [49, Lemma 3.7] and of [2, Lemma 5.5]. The letter C will denote a positive constant, which can possibly change from line to line.
For the sake of clarity, we consider cracks in R0,1η. The proof can be easily generalized to the whole class Rη. For brevity, we set En:=E(fn,gn,wn,Cn;Γn) and E∞:=E(f∞,g∞,w∞,C∞;Γ∞); furthermore, along the proof we denote with En and E∞ the functionals appearing in the minimization (1.2) with data {fn,gn,wn,Cn,Γn} and {f∞,g∞,w∞,C∞,Γ∞}, respectively. Clearly, we have
En=En(un)=12∫ΩCEun:Eundx−∫Ωfn⋅undx−∫∂SΩgn⋅undH1for n∈N∪{∞}, |
where Eun are interpreted as functions defined a.e. in Ω.
Let γn∈C1,1([0,ℓn];R2) and γ∞∈C1,1([0,ℓ∞];R2) be the arc-length parametrizations of Γn and Γ∞, respectively, where ℓn and ℓ∞ denote the H1 measures of the crack sets. By a simple rescaling of γn, we construct a C1,1 parametrization ˆγn of Γn, defined in [0,ℓ∞]. The new parametrization, by definition of R0,1η, belongs to W2,∞([0,ℓ∞];R2) and its norm is bounded by a constant independent of n. From the Hausdorff convergence of Γn to Γ∞, we deduce that ˆγn converges to γ∞ weakly* in W2,∞([0,ℓ∞];R2) and strongly in W1,∞([0,ℓ∞];R2).
Let us fix ρ>0 sufficiently small, so that the projection ΠΓ∞ over Γ∞ is well defined in Iρ(Γ∞):={x∈Ω:d(x,Γ∞)<ρ)}. For n large enough we have Γn⊆Iρ(Γ∞). We want to construct a Lipschitz function Λn such that Λn(Γ∞)=Γn and Λn(x)=x for x∈R2∖Iρ(Γ∞). For every x∈Iρ(Γ∞) we define s(x)∈[0,ℓ∞] in such a way that γ∞(s(x))=ΠΓ∞(x). We notice that the map x↦s(x) is locally Lipschitz, while ΠΓ∞ is Lipschitz on Iρ(Γ∞). Moreover, we set dn:=‖ˆγn−γ∞‖1/2W1,∞ and λn(t):=(1−|t|dn)+, where (⋅)+ stands for the positive part. With this notation at hand, we define
Λn(x):=x+λn(|x−ΠΓ∞(x)|)(ˆγn(s(x))−γ∞(s(x)))for x∈R2. |
In particular, Λn is Lipschitz, ‖Λn−id‖W1,∞≤Cdn→0 as n→∞, and, for n large enough, Λn(Γ∞)=Γn and Λn=id out of Idn(Γ∞). Applying the Hadamard Theorem [37, Theorem 6.2.3], we deduce that Λn is globally invertible with ‖Λ−1n−id‖W1,∞→0 as n→∞.
Given v∈H1(Ω∖Γ∞;R2) with v=w∞ on ∂DΩ∖Γ0, we have that the function vn:=v∘Λ−1n+wn−w∞ belongs to H1(Ω∖Γn;R2) and satisfies vn=wn on ∂DΩ∖Γ0. Moreover, ∇vn→∇v in L2(Ω;M2), vn→v in L2(Ω;R2), and, by the continuity of the trace operator, vn→v strongly in L2(∂SΩ;R2). This asymptotic analysis implies that the sequence (En(vn))n∈N is bounded and converges to E∞(v) as n→∞.
By the minimality of un for En, we have
En(un)≤En(vn)<C. | (2.6) |
It is easy to see that the functionals En are equi-coercive in H1(Ω∖Γn;R2), so that inequality (2.6), together with Proposition 2.3, provides a uniform bound on the L2 norm of un, of its gradient, and of its trace. Therefore, up to a subsequence (not relabeled), we have un⇀φ weakly in L2(Ω;R2) for some φ∈L2(Ω;R2). Moreover, in a suitably small neighborhood U of the boundary, this convergence is stronger, since (Ω∖Γn)∩U=(Ω∖Γ0)∩U for every n. More precisely, we have un⇀φ weakly in H1((Ω∖Γ0)∩U;R2) and, therefore, un→φ strongly in L2(∂Ω;R2) and φ=w∞ on ∂DΩ. The above convergences imply that
limn→∞∫Ωfn⋅undx+∫∂SΩgn⋅undH1=∫Ωf∞⋅φdx+∫∂SΩg∞⋅φdH1. | (2.7) |
Hence, passing to the liminf in (2.6) we get
E∞(φ)≤E∞(v)for every v∈H1(Ω∖Γ∞;R2) with v=w∞ H1-a.e. on ∂DΩ. |
Thus, φ is a minimizer of E∞ in H1(Ω∖Γ∞;R2) with boundary condition w∞ and, by uniqueness of the minimizer, φ=u∞. The strong convergence of the gradients follows by considering (2.6) for v=u∞. Indeed, we have
E∞(u∞)≤lim infnEn(un)≤lim supnEn(un)≤limnEn(u∞∘Λ−1n+wn−w∞)=E∞(u∞), |
which implies, together with (2.7), that En→E∞ and Eun→Eu∞ in L2(Ω;M2sym). Applying Proposition 2.3 and recalling that wn→w∞ in H1(Ω∖Γ0;R2), we also obtain the strong convergence of ∇un to ∇u∞ in L2(Ω;M2). This concludes the proof of the lemma.
As a corollary of Lemma 2.4 we deduce the existence of solutions of the incremental minimum problems (1.4).
Corollary 2.5. Fix ε>0, k∈N, and i∈{1,…,k}. Then the minimum problem (1.4) admits a solution.
Proof. It is sufficient to apply the direct method. Let (Γn)n∈N⊆Rη be a minimizing sequence for (1.4). By Theorem 2.1, Γn converges in the Hausdorff metric, up to a subsequence (not relabeled), to some Γ∞∈Rη such that the constraint Γ∞⊇Γε,k,i−1 is preserved; moreover we have H1(Γn)→H1(Γ∞). Applying Lemma 2.4 with Cn=C∞=C, fn=f∞=f(tk,i), gn=g∞=g(tk,i), and wn=w∞=w(tk,i), we obtain the convergence of the corresponding energies E(tk,i;Γn)→E(tk,i;Γ∞). Hence, Γ∞ is a solution to the minimum problem.
This section is devoted to the definition of the energy release rate, i.e., the opposite of the derivative of the energy E(t;⋅) with respect to the crack elongation. The problem is clearly time-independent, therefore we omit the variable t, which is kept fixed. As in the previous section, the energy in (1.2) simply writes E(Γ).
We aim at generalizing the results of [3], where the energy release rate has been computed only in presence of smooth cracks Γ, in the absence of forces, and with a spatially constant elasticity tensor. Here we extend its definition to the case Γ∈Rη, non-zero volume and boundary forces f∈L2(Ω;R2) and g∈L2(∂SΩ;R2), boundary datum w∈H1(Ω∖Γ0;R2), and nonconstant tensor C∈C0,1(¯Ω).
As in [3], the fundamental steps are the following:
(i) Given an increasing family of cracks Γσ∈R0,1η parametrized by their arc length σ∈[0,S], we prove that the map σ↦E(Γσ) is differentiable, thus
dE(Γσ)dσ|σ=s:=limσ→sE(Γσ)−E(Γs)σ−s. |
(ii) We show that the above derivative only depends on the current configuration Γs, and not on its possible extensions, i.e., if Γσ=ˆΓσ for σ≤s, then
dE(Γσ)dσ|σ=s=dE(ˆΓσ)dσ|σ=s. |
In particular, we are allowed to write −dE(Γσ)dσ|σ=s=:G(Γs) with no ambiguity.
We point out a difference of our strategy with respect to the proof of [39] for the antiplane case. In that case, the energy release rate is first characterized via the stress intensity factor assuming that the volume force is null in a neighborhood of the crack tip; then, one treats general forces by approximation, using the property that the stress intensity factor is continuous with respect to the force. In this paper, in the planar case we do not prove the existence of stress intensity factors for nonsmooth curves. Hence, when expressing the energy release rate via integral forms, we have to deal carefully with the terms containing the external force. Once the existence of the energy release rate is guaranteed, we will reduce to the case of forces that are null close to the tip via some approximation arguments, see Lemma 3.8 below.
In order to rigorously proceed with (i), we follow the classical techniques of [21,28], refined e.g., in [30,32,33,39]. We first restrict our attention to cracks Γs∈R0,1η. We write Γs as
Γs:={γ(σ):0≤σ≤s}, | (3.1) |
where γ∈C1,1 is the arc-length parametrization of Γs. We will discuss in Remark 3.10 how to tackle the general case Γ∈Rη. For brevity, we denote with us∈H1(Ω∖Γs;R2) the minimizer of (1.2). As in the previous section, when explicitly needed, we will highlight the dependence on the data by writing E(f,g,w,C;Γs) for E(Γs).
In order to make explicit computations, for every s∈(0,S) and δ∈R with |δ| small we need to employ a C1,1 diffeomorphism Fs,δ, borrowed from [30,33], which transforms Γs+δ in Γs and maps Ω into itself. Precisely, for r>0 small enough we may assume that the curve Γσ∩Br(γ(s)), for |s−σ| small, is the graph of a C1,1 function ζ, with ζ′(γ1(s))=0, where we have denoted with γ1 the first component of the arc-length parametrization γ. We define the function Fs,δ:Br/2(γ(s))→R2 by
Fs,δ(x):=x+((γ1(s+δ)−γ1(s))φ(x)ζ(x1+(γ1(s+δ)−γ1(s))φ(x))−ζ(x1)), |
where φ∈C∞c(Br/2(γ(s))) is a suitable cut-off function equal to 1 close to γ(s). Notice that, for r small enough, spt(φ)∩spt(w)=∅. We extend Fs,δ to the identity in R2∖Br/2(γ(s)).
By the regularity of ζ, Fs,δ is a C1,1 diffeomorphism of R2 such that Fs,δ(Γs)=Γs+δ and Fs,0=id. Moreover, the following equalities hold:
ρs(x):=∂δ(Fs,δ(x))|δ=0=γ′1(s)φ(x)(1ζ′(x1)),∂δ(det∇Fs,δ)|δ=0=divρs,∂δ(∇Fs,δ)|δ=0=−∂δ(∇Fs,δ)−1|δ=0=∇ρs. | (3.2) |
Notice that ρs(x) only depends on s through the term γ′1(s), where γ1∈C1,1; when we come to derive in x, it turns out that s↦∇ρs(x) is continuous for a.e.\ x. On the other hand, for the continuity of the energy release rate we will need a bound on ‖∇ρs‖L∞, uniform in s: for this reason we need the assumption that the crack is C1,1.
With this notation at hand, we can write the derivative of s↦E(Γs) as a surface integral depending on the deformation gradient. Such formula follows arguing as in [33], where it is stated for C2 cracks and with elastic energy independent of the position in the reference configuration. For the reader's convenience, we include a proof adapted to our framework (C1,1 cracks, nonconstant elasticity tensor, volume forces).
Proposition 3.1. Let {Γσ}σ≥0⊆R0,1η be parametrized as in (3.1). Let f∈L2(Ω;R2), g∈L2(∂SΩ;R2), w∈H1(Ω∖Γ0;R2), and C∈C0,1(¯Ω). Then, the map σ↦E(Γσ) is differentiable and
dE(Γσ)dσ|σ=s= 12∫Ω∖Γs(DCρs)∇us:∇usdx−∫Ω∖ΓsC∇us∇ρs:∇usdx+12∫Ω∖ΓsC∇us:∇usdivρsdx+∫Ωf⋅∇usρsdx, | (3.3) |
where DC denotes the fourth order tensor
(DCρs)ijkl:=2∑m=1∂Cijkl∂xmρs,m,ρs=(ρs,1,ρs,2). |
Proof. The proof follows the lines of [21,28,30,33]. To prove (3.3), we compute explicitly the limits
limσ↘sE(Γσ)−E(Γs)σ−s=limδ↘0E(Γs+δ)−E(Γs)δ, | (3.4) |
limσ↗sE(Γσ)−E(Γs)σ−s=limδ↗0E(Γs+δ)−E(Γs)δ, | (3.5) |
and show that the two limits coincide.
Let us start with (3.4). For every δ>0, the function us∘F−1s,δ belongs to H1(Ω∖Γs+δ;R2) and us∘F−1s,δ=us on ∂Ω. Hence,
E(Γs+δ)−E(Γs)δ≤ 12δ(∫Ω∖Γs+δC∇(us∘F−1s,δ):∇(us∘F−1s,δ)dx−∫Ω∖ΓsCEus:Eusdx)−1δ∫Ωf⋅(us∘F−1s,δ−us)dx. | (3.6) |
By a change of coordinate in the first integral in (3.6) we deduce that
E(Γs+δ)−E(Γs)δ≤ 12δ(∫Ω∖ΓsC(Fs,δ)∇us(∇Fs,δ)−1:∇us(∇Fs,δ)−1det∇Fs,δdx−∫Ω∖ΓsCEus:Eusdx)−1δ∫Ωf⋅(us∘F−1s,δ−us)dx. | (3.7) |
By a simple computation, we can rewrite (3.7) as
E(Γs+δ)−E(Γs)δ≤12∫Ω∖Γs(C(Fs,δ)−C)δ∇us(∇Fs,δ)−1:∇us(∇Fs,δ)−1det∇Fs,δdx+12∫Ω∖ΓsC∇us((∇Fs,δ)−1−I)δ:∇us(∇Fs,δ)−1det∇Fs,δdx+12∫Ω∖ΓsC∇us:∇us((∇Fs,δ)−1−I)δdet∇Fs,δdx+12∫Ω∖ΓsC∇us:∇usdet∇Fs,δ−1δdx−1δ∫Ωf⋅(us∘F−1s,δ−us)dx=:Iδ,1+Iδ,2+Iδ,3+Iδ,4+Iδ,5. | (3.8) |
Since C(Fs,δ)−Cδ converges to DCρs weakly* in L∞(Ω) and
limδ→0(∇Fs,δ)−1−Iδ=∂δ(∇Fs,δ)−1|δ=0=−∇ρs,limδ→0det∇Fs,δ−1δ=∂δ(det∇Fs,δ)|δ=0=divρs, |
where the limits are uniform in δ, we obtain
limδ↘0Iδ,1=12∫Ω∖Γs(DCρs)∇us:∇usdx, | (3.9) |
limδ↘0Iδ,2=limδ↘0Iδ,3=−12∫Ω∖ΓsC∇us∇ρs:∇usdx, | (3.10) |
limδ↘0Iδ,4=12∫Ω∖ΓsC∇us:∇usdivρsdx. | (3.11) |
Applying e.g., [1, Lemma 3.8] (see also [32, Lemma 4.1]), we see that δ−1(us∘F−1s,δ−us)→−∇usρs in L2(Ω) as δ→0. Thus,
limδ↘0Iδ,5=∫Ωf⋅∇usρsdx. | (3.12) |
Combining (3.8)–(3.12) we get
lim supδ↘0E(Γs+δ)−E(Γs)δ≤ 12∫Ω∖Γs(DCρs)∇us:∇usdx−∫Ω∖ΓsC∇us∇ρs:∇usdx+12∫Ω∖ΓsC∇us:∇usdivρsdx+∫Ωf⋅∇usρsdx. | (3.13) |
In order to obtain the opposite inequality, we consider the function us+δ∘Fs,δ∈H1(Ω∖Γs;R2). By the minimality of us we have
E(Γs+δ)−E(Γs)δ≥ 12δ(∫Ω∖Γs+δCEus+δ:Eus+δdx−∫Ω∖ΓsC∇(us+δ∘Fs,δ):∇(us+δ∘Fs,δ)dx)−1δ∫Ωf⋅(us+δ−us+δ∘Fs,δ)dx. | (3.14) |
For simplicity of notation, we denote with Us,δ:=us+δ∘Fs,δ. By a change of variable in the first integral in (3.14) we deduce that
E(Γs+δ)−E(Γs)δ≥12δ(∫Ω∖ΓsC(Fs,δ)∇Us,δ(∇Fs,δ)−1:∇Us,δ(∇Fs,δ)−1det∇Fs,δdx−∫Ω∖ΓsC∇Us,δ:∇Us,δdx)−1δ∫Ωf⋅(us+δ−Us,δ)dx. |
Repeating the computations of (3.8)–(3.12) and taking into account that δ−1(us+δ−Us,δ)⇀−∇uρs weakly in L2(Ω;R2) (see, e.g., [1, Lemma 3.8]), we infer that
lim infδ↘0E(Γs+δ)−E(Γs)δ≥ 12∫Ω∖Γs(DCρs)∇us:∇usdx−∫Ω∖ΓsC∇us∇ρs:∇usdx+12∫Ω∖ΓsC∇us:∇usdivρsdx+∫Ωf⋅∇usρsdx, |
which, together with (3.13) implies that
limδ↘0E(Γs+δ)−E(Γs)δ= 12∫Ω∖Γs(DCρs)∇us:∇usdx−∫Ω∖ΓsC∇us∇ρs:∇usdx+12∫Ω∖ΓsC∇us:∇usdivρsdx+∫Ωf⋅∇usρsdx. |
Adapting the above argument to the case δ<0, cf. (3.5), it is also possible to prove that
limδ↗0E(Γs+δ)−E(Γs)δ= 12∫Ω∖Γs(DCρs)∇us:∇usdx−∫Ω∖ΓsC∇us∇ρs:∇usdx+12∫Ω∖ΓsC∇us:∇usdivρsdx+∫Ωf⋅∇usρsdx. |
This concludes the proof of (3.3).
The following corollary states the continuity of the derivative (3.3) w.r.t. the data f, g, w, C, and Γs.
Corollary 3.2. Let fn,f∈L2(Ω;R2), gn,g∈L2(∂SΩ;R2), wn,w∈H1(Ω∖Γ0;R2), and Cn,C∈C0,1(¯Ω) be such that fn→f strongly in L2(Ω;R2), gn⇀g weakly in L2(∂SΩ;R2), wn⇀w weakly in H1(Ω∖Γ0;R2), and Cn⇀C weakly* in W1,∞(Ω). Moreover, let S>0, let {Γs}s∈[0,S]⊆R0,1η be as in (3.1), and assume that there exists a sequence {Γns}s∈[0,S]⊆R0,1η such that Γns converges to Γs in the Hausdorff metric of sets for every s∈[0,S]. Then, for every s∈(0,S) we have
limndE(fn,gn,wn,Cn;Γnσ)dσ|σ=s=dE(f,g,w,C;Γσ)dσ|σ=s. | (3.15) |
Proof. Let us denote with un and u the displacements associated to E(fn,gn,wn,Cn;Γns) and to E(f,g,w,C;Γs), respectively. By Lemma 2.4 and by the hypotheses, it follows that ∇un converges to ∇u strongly in L2(Ω;M2). Let us denote by ρns the quantity defined in (3.2) and corresponding to Γns. Since Γns converges to Γs in the Hausdorff metric of sets for every s∈[0,S], we have that ρns→ρs uniformly in Ω and weakly* in W1,∞(Ω;R2) for every s∈(0,S). Thus (3.15) follows by (3.3).
We notice that the dependence of dE(Γσ)dσ|σ=s on {Γσ}σ∈[0,S] is encoded by the quantity ρs introduced in (3.2). The rest of this section is devoted to step (ii), namely at proving that the above derivative only depends on the current fracture Γs, and not on its possible extensions, i.e., on the choice of the family {Γσ}σ∈[0,S]. We start by recalling a result of [3] (cf. also [11,20]) stating that this very same property holds for C∞ cracks in absence of external forces and with constant elasticity tensor.
Theorem 3.3. ([3, Theorem 4.1]). Let f=g=0 and let C be constant in Ω. Let {Γσ}σ∈[0,S]⊆R0,1η be as in (3.1) and assume that there exists ˉs∈(0,S) such that Γσ is of class C∞ for every σ∈(0,ˉs]. Then, for every s∈(0,ˉs] there exist two constants Q1(Γs) and Q2(Γs) (independent of Γσ for σ>s) such that
dE(Γσ)dσ|σ=s=C(λ,μ)(Q21(Γs)+Q22(Γs)), | (3.16) |
where C(λ,μ) is a constant which depends only on the Lamé coefficients λ and μ.
Remark 3.4. The constants Q1(Γs) and Q2(Γs) are the so called stress intensity factors. Indeed, it has been proven in [3, Theorem 2.5] that, in the condition of Theorem 3.3, the displacement us can be written as
us=uR+Q1(Γs)Φ1+Q2(Γs)Φ2, | (3.17) |
for suitable functions uR∈H2(Ω∖Γs;R2) and Φ1,Φ2∈H1(Ω∖Γs;R2)∖H2(Ω∖Γs;R2). Moreover, the proof of formula (3.16) follows from the above decomposition.
The next proposition is a simple localization of Theorem 3.3.
Proposition 3.5. Let {Γσ}σ∈[0,S]⊆R0,1η be as in (3.1). Let s∈(0,S), f∈L2(Ω;R2), g∈L2(∂SΩ;R2), w∈H1(Ω∖Γ0;R2), and C∈C0,1(¯Ω) be such that Γs is C∞, f=0, and C is constant in a neighborhood of the tip γ(s) of Γs. Then, there exist two constants Q1(Γs) and Q2(Γs) (independent of Γσ for σ>s) such that
dE(Γσ)dσ|σ=s=C(λs,μs)(Q21(Γs)+Q22(Γs)), | (3.18) |
where C(λs,μs) coincides with the constant appearing in (3.16) and λs,μs denote the Lamé coefficients of C in γ(s).
Proof. As mentioned in Remark 3.4, the proof of formula (3.18) follows directly from a splitting of the form (3.17) for the displacement us solution of
min{12∫Ω∖ΓsCEu:Eudx−∫Ω∖Γsf⋅udx−∫∂SΩg⋅udH1:u∈H1(Ω∖Γs;R2),u=w on ∂DΩ}. |
close to the tip γ(s) of Γs. Indeed, given (3.17), we can simply repeat step by step the proof of [3, Theorem 4.1] and get (3.18). In order to obtain such a decomposition in a neighborhood of γ(s), we note that us is also solution of
min{12∫Bℓ(γ(s))∖ΓsCEu:Eudx:u∈H1(Bℓ(γ(s))∖Γs;R2),u=us on ∂Bℓ(γ(s)) } |
with ℓ chosen in such a way that Γs is smooth, f=0, and C is constant in Bℓ(γ(s)). This enables us to apply [3, Theorem 2.5] in the domain Bℓ(γ(s)) and to deduce the decomposition (3.18) in Bℓ(γ(s)).
We are now in a position to state and prove the main result of this section.
Theorem 3.6. Let {Γσ}σ∈[0,S],{ˆΓσ}σ∈[0,S]⊆R0,1η be as in (3.1). Let f∈L2(Ω;R2), g∈L2(∂SΩ;R2), w∈H1(Ω∖Γ0;R2), and C∈C0,1(¯Ω). Let s∈(0,S) and assume that Γσ=ˆΓσ for σ≤s. Then,
dE(Γσ)dσ|σ=s=dE(ˆΓσ)dσ|σ=s. | (3.19) |
Remark 3.7. The previous theorem states that the derivative dE(Γσ)dσ|σ=s computed in Proposition 3.1 does not depend on the possible extension of Γs in the class R0,1η. Hence, it represents the slope of the energy E with respect to variations of crack in the set of admissible curves R0,1η.
The proof of Theorem 3.6 is a corollary of the following lemma, where we use an approximation argument to reduce ourselves to the case of smooth cracks, constant elasticity tensor, and forces that are null close to the crack tip. In the latter case, the relation between the energy release rate and the stress intensity factors shows (3.19), cf. [3].
Lemma 3.8. Let {Γσ}σ∈[0,S], f, and C be as in the statement of Theorem 3.6, and let s∈(0,S). Then, there exist δ>0, fn∈L2(Ω;R2), Cn∈C0,1(¯Ω), and {Γnσ}σ∈[0,s+δ]⊆R0,1η such that fn→f strongly in L2(Ω;R2), Cn⇀C weakly* in W1,∞(Ω), Γnσ→Γσ in the Hausdorff metric of sets for every σ∈[0,s+δ], and, close to the tip of Γns, Γns is smooth, fn=0, and Cn is constant.
Moreover, if {ˆΓσ}σ∈[0,S] is another family of curves in R0,1η with ˆΓσ=Γσ for σ≤s, then the sequences {Γnσ}σ∈[0,s+δ], {ˆΓnσ}σ∈[0,s+δ], Cn, ˆCn, fn, and ˆfn can be chosen in such a way that ˆΓnσ=Γnσ for σ≤s, Cn=ˆCn, and fn=ˆfn.
Proof. We start with the construction of an approximating sequence for Γσ. Let δ>0 be such that s+δ<S. Let us fix a sequence sn↗s. By definition of the class R0,1η, for every n there exist two open balls B1η,n and B2η,n of radius η such that ¯B1η,n∩¯B2η,n={γ(sn)} and (B1η,n∪B2η,n)∩Γs=∅. Up to a redefinition of δ, for n large enough we may assume that the portion of curve {γ(σ):sn−δ≤σ≤s+δ}⊆Γs+δ can be represented, in a suitable coordinate system (x1,x2) possibly dependent on n, as graph of a function ψn of class C1,1 with ψ′n(xsn1)=0, where the point (xsn1,ψn(xsn1)) coincides with γ(sn). A similar notation is used for γ(s)=(xs1,ψn(xs1)). Without loss of generality, we assume that ψ′n(xs1)≥0.
The idea of our construction is to extend the curve Γsn with the arc of circumference of equation
x2=ψn(xsn1)+η−√η2−(x1−xsn1)2for x1∈[xsn1,ˉx1), | (3.20) |
where ˉx1 is the smallest x1≥xsn1 such that ψ′n(ˉx1)=ψ′n(xs1). We notice that (3.20) is the equation of the boundary of one of the two open balls Biη,n and that ˉx1=xsn1 whenever ψ′n(xs1)=0. We denote with Λn the extension of Γsn with the arc (3.20) and its tip with Pn. We also use the symbol Λnσ, σ∈[0,H1(Λn)], to indicate the piece of curve contained in Λn of length σ.
A direct computation gives H1(Λn)=sn+ηarctan(ψ′n(xs1)), which can also be written as follows:
H1(Λn)=sn+η∫ψ′n(xs1)011+x2dx. | (3.21) |
On the other hand, exploiting the upper bound η−1 on the curvature of the crack set, which reads |ψ″n|[1+(ψ′n)2]−3/2≤η−1 in terms of the graph parametrization, we get
H1(Γs)=sn+∫xs1xsn1√1+(ψ′n)2(x)dx≥sn+η∫xs1xsn1|ψ″n|1+(ψ′n)2(x)dx≥sn+η∫ψ′n(xs1)011+x2dx. | (3.22) |
Comparing (3.21) and (3.22), we conclude that H1(Λn)≤H1(Γs)=s.
If H1(Λn)<s, we denote with αnℓ the segment of length ℓ, initial point Pn and parallel to γ′(s), and we define {Γnσ}σ∈[0,s+δ] as follows:
Γnσ:={Λnσif σ∈[0,H1(Λn)],Λn∪αnσ−H1(Λn)if σ∈[H1(Λn),s],Λn∪αns−H1(Λn)∪((Pn+(s−H1(Λn))γ′(s)−γ(s))+Γσ∖Γs)if σ∈(H1(Λn),s+δ], |
where we have used the notation v+E:={v+e:e∈E} for v∈R2 and E⊆R2.
If H1(Λn)=s, we simply set
Γnσ:={Λnσif σ∈[0,H1(Λn)],Λn∪((Pn−γ(s))+Γσ∖Γs)if σ∈(H1(Λn),s+δ]. |
In both cases, we have that {Γnσ}σ∈[0,s+δ]⊆R0,1η, Γnσ→Γσ in the Hausdorff metric of sets for every σ∈[0,s+δ], and Γns is of class C∞ close to its tip.
The construction of fn is trivial, since the set of functions in L2(Ω;R2) that vanish close to γ(s) is dense in L2(Ω;R2) w.r.t. the L2-norm. We only have to ensure that fn is also null close to the tip of Γns, which is still possible because of the Hausdorff convergence.
As for the elasticity tensor C, for every r>0 we consider a cut off function φr in Br(γ(s)) with φr(x)=φr(|x−γ(s)|), φr=1 in ¯Br/2(γ(s)), and |∇φr|≤C/r for some positive constant C independent of r. Let us set Cs:=C(γ(s)) and Cr:=φrC+(1−φr)Cs. It is easy to see that Cr∈C0,1(¯Ω) with Lipschitz constant bounded by C(Lip(C)+1). Hence, Cr⇀C weakly* in W1,∞(Ω) as r↘0. To conclude, it is enough to choose a suitable sequence rn↘0 in such a way that Cn:=Crn is constant close to the tip of Γns. This is possible thanks to the Hausdorff convergence of Γns to Γs.
The last part of the lemma is a trivial consequence of the above construction.
Proof of Theorem 3.6. To prove (3.19), we apply Lemma 3.8 to both Γσ and ˆΓσ. Fixed s∈(0,S) and δ>0 small, let {Γnσ}σ∈[0,s+δ], {ˆΓσ}{σ∈[0,s+δ], Cn, and fn be as in Lemma 3.8. By Corollary 3.2 we have that
limndE(fn,g,w,Cn;Γnσ)dσ|σ=s=dE(f,g,w,C;Γσ)dσ|σ=s,limndE(fn,g,w,Cn;ˆΓnσ)dσ|σ=s=dE(f,g,w,C;ˆΓσ)dσ|σ=s. |
Taking into account Proposition 3.5, we have that
dE(fn,g,w,Cn;Γnσ)dσ|σ=s=dE(fn,g,w,Cn;ˆΓnσ)dσ|σ=sfor every n∈N |
and we deduce (3.19).
We are now in a position to give the precise definition of energy release rate for a crack of the form (3.1). We stress that this is now possible thanks to Theorem 3.6.
Definition 3.9. Let Γ∈R0,1η, s:=H1(Γ), f∈L2(Ω;R2), g∈L2(∂SΩ;R2), w∈H1(Ω∖Γ0;R2), and C∈C0,1(¯Ω). Let S>s and let {Γσ}σ∈[0,S]⊆R0,1η be such Γs=Γ. We define the energy release G(Γ) as
G(Γ):=−dE(Γσ)dσ|σ=s. |
Remark 3.10. Definition 3.9, stated for a curve Γ∈R0,1η, can be further generalized in order to consider general cracks in the class Rη. Indeed, given Γ∈Rη, it is enough to represent it as union of arcs of C1,1 curves Γm, m=1,…,M. In particular, each component belongs to R0,1η and can be written as in (3.1). Hence, for every m we define the m-th energy release rate Gm(Γ) as in Definition 3.9 w.r.t. variations of the sole component Γm of Γ. The energy release rate will be in this case the vector
G(Γ):=(G1(Γ),…,GM(Γ)). |
Remark 3.11. We collect here the main properties of the energy release rate G.
(a) G is continuous w.r.t. the Hausdorff convergence of cracks Γ∈Rη, strong convergence of volume forces f∈L2(Ω;R2), weak convergence of surface forces g∈L2(∂SΩ;R2), and convergence of Dirichlet boundary data w∈H1(Ω∖Γ0;R2).
(b) There exists a positive constant C=C(C,η) such that for every Γ∈Rη, every f∈L2(Ω;R2), every g∈L2(∂SΩ;R2), and every w∈H1(Ω∖Γ0;R2),
0≤G(Γ)≤C‖u‖2H1+‖f‖2‖u‖H1, |
where u∈H1(Ω∖Γ;R2) is the solution of (1.2) with data Γ, f, g, w, and C.
We will make use of these two properties in the proofs of Proposition 1.5 and Theorems 1.6 and 1.7.
In this section we focus on the proofs of existence of a viscous evolution Γε (see Proposition 1.4) and of a balanced viscosity evolution Γ (Theorem 1.5), the latter obtained as limit of Γε as the viscosity parameter ε tends to 0. To this end, we follow the method employed in a wide literature on rate-independent processes [44]. However, we point out that the abstract results of [41,42,43] do not directly apply to our setting.
Since the problem we analyze depends explicitly on time t through the applied loads f, g, and w, from now on we denote with G(t;Γ) the energy release rate defined in Definition 3.9 and Remark 3.10 for a crack Γ∈Rη at time t∈[0,T].
As anticipated in Section 1, the proofs of Proposition 1.5 and of Theorem 1.6 are based on a time-discretization procedure. Let the initial crack Γ0∈R0η and the viscosity parameter ε>0 be fixed, and let us set lm0:=H1(Γm0), where Γ0=⋃Mm=1Γm0, according to Definition 1.1. For every k∈N we fix a partition {tk,i}ki=0 of the time interval [0,T] as in (1.2). For i=0 we set Γε,k,0:=Γ0. For i∈{1,…,k} we denote with Γε,k,i a minimizer of the incremental minimum problem (1.4), whose existence is provided by Corollary 2.5. Recalling the conventions of Definition 1.2, we write Γε,k,i=⋃Mm=1Γmε,k,i, we set lmε,k,i:=H1(Γmε,k,i), and we denote with Pmε,k,i the tip of Γmε,k,i. Furthermore, we define the interpolation functions
lmε,k(t):=lmε,k,i−1+(lmε,k,i−lmε,k,i−1)t−tk,i−1tk,i−tk,i−1,Gmε,k(t):=Gm(tk,i;Γε,k,i),Gε,k(t):=G(tk,i;Γε,k,i),Γε,k(t):=Γε,k,i,Pmε,k(t):=Pmε,k,i,uε,k(t):=uεk,i,tk(t):=tk,i,fk(t):=f(tk,i),gk(t):=g(tk,i)for t∈(tk,i−1,tk,i],Γ_ε,k(t):=Γε,k,i−1,u_ε,k(t):=uεk,ifor t∈[tk,i−1,tk,i), |
where we denoted with uεk,i the function u(tk,i;Γε,k,i)∈H1(Ω∖Γε,k,i;R2).
In the following proposition we state a time discrete version of the Griffith's criterion (G1)ε–(G3)ε.
Proposition 4.1. For every ε>0, every k∈N, every m∈{1,…,M}, and a.e. t∈(0,T] it holds:
(G1)k ˙lmε,k(t)≥0;
(G2)k κ(Pmε,k(t))−Gmε,k(t)+ε˙lmε,k(t)≥0;
(G3)k ˙lmε,k(t)(κ(Pmε,k(t))−Gmε,k(t)+ε˙lε,k(t))=0.
Proof. By construction, lmε,k is a nondecreasing function, so that (G1)k is clearly satisfied. In order to show (G2)k–(G3)k we take into account the minimality of Γε,k,i. Let us fix i∈{1,…k}. For every ¯m∈{1,…,M}, let Γ¯m∈Rη be such that Γ¯m=⋃m≠¯mΓmε,k,i∪Λ¯m with Λ¯m⊇Γ¯mε,k,i, and let us set λ:=H1(Λ¯m)≥l¯mε,k,i. Then,
E(tk,i;Γε,k,i)+∫Γε,k,iκdH1+ε2M∑m=1H1(Γmε,k,i∖Γmε,k,i−1)2tk,i−tk,i−1≤E(tk,i;Γ¯m)+∫Γ¯mκdH1+ε2∑m≠¯mH1(Γmε,k,i∖Γmε,k,i−1)2tk,i−tk,i−1+ε2H1(Λ¯m∖Γ¯mε,k,i−1)2tk,i−tk,i−1, |
which implies
E(tk,i;Γε,k,i)+∫Γ¯mε,k,iκdH1+ε2(l¯mε,k,i−l¯mε,k,i−1)2tk,i−tk,i−1≤E(tk,i;Γ¯m)+∫Λ¯mκdH1+ε2(λ−l¯mε,k,i−1)2tk,i−tk,i−1. | (4.1) |
We divide (4.1) by λ−l¯mε,k,i and pass to the limit as λ→l¯mε,k,i, obtaining (G2)k as a consequence of (H4) and of Definition 3.9. If, moreover, Γ¯mε,k,i−1⊊Γ¯mε,k,i, we can consider as a competitor a set Γ¯m∈Rη as above, with Γ¯mε,k,i−1⊆Λ¯m⊆Γ¯mε,k,i, so that l¯mε,k,i−1≤λ≤l¯mε,k,i. Repeating the above computation we obtain
κ(P¯mε,k,i)−G¯m(tk,i;Γ¯mε,k,i)+εl¯mε,k,i−l¯mε,k,i−1tk,i−tk,i−1=0. |
This concludes the proof of (G3)k.
We now show an a priori bound on lε,k and on uε,k.
Proposition 4.2. The following facts hold:
(a) there exist two positive constants c and C independent of ε, k, and i such that for every ε>0, every k∈N, and every t∈[0,T],
ε2M∑m=1∫tk(t)0|˙lmε,k(τ)|2dτ+c(‖Euε,k(t)‖22−‖uε,k(t)‖H1)≤ F(0;Γ0)+∫tk(t)0∫ΩCEu_ε,k(τ):E˙w(τ)dxdτ+Ck∑i=1‖wk,i−wk,i−1‖2H1−∫tk(t)0∫Ω˙f(τ)⋅u_ε,k(τ)dxdτ−∫tk(t)0∫Ωfk(τ)⋅˙w(τ)dxdτ−∫tk(t)0∫∂SΩ˙g(τ)⋅u_ε,k(τ)dH1dτ−∫tk(t)0∫∂SΩgk(τ)⋅˙w(τ)dH1dτ; | (4.2) |
(b) for every ε>0, along a suitable (not relabeled) subsequence, ‖uε,k(t)‖2 and ‖∇uε,k(t)‖2 are bounded uniformly w.r.t. t∈[0,T] and k∈N;
(c)for every ε>0, along a suitable (not relabeled) subsequence, ε‖˙lmε,k‖22 is bounded uniformly w.r.t. k∈N and m∈{1,…,M}.
Proof. For the sake of simplicity, let us denote with wk,i, fk,i, and gk,i the functions w(tk,i), f(tk,i), and g(tk,i), respectively.
By definition of uεk,i, by hypothesis (H3), and by the regularity of the data of the problem f, g, and w, we have that
C1(‖Euεk,i‖22−‖uεk,i‖H1)≤E(tk,i;Γε,k,i)≤12∫ΩCEwk,i:Ewk,idx−∫Ωfk,i⋅wk,idx−∫∂SΩgk,i⋅wk,idH1≤C2, | (4.3) |
for some positive constants C1 and C2 depending only on f, g, w, and C.
Since, for every ε and k, the set function Γε,k:[0,T]→Rη is nondecreasing, we have that uεk,i∈H1(Ω∖Γε,k(T);R2). By definition of the class Rη, the curves Γε,k(T) have bounded length uniformly w.r.t. ε and k. Hence, we may assume that, up to a not relabeled subsequence, Γε,k(T)→ˆΓε∈Rη in the Hausdorff metric of sets as k→∞. We are therefore in a position to apply Proposition 2.3 to Γε,k(T), ˆΓε, and uεk,i, which, together with (4.3), implies (b).
By definition of Γε,k,i and of the energy E(tk,i;Γ) we have that
E(tk,i;Γε,k,i)+∫Γε,k,iκdH1+ε2∑Mm=1H1(Γmε,k,i∖Γmε,k,i−1)2tk,i−tk,i−1≤E(tk,i;Γε,k,i−1)+∫Γε,k,i−1κdH1≤12∫ΩCE(uεk,i−1+wk,i−wk,i−1):E(uεk,i−1+wk,i−wk,i−1)dx−∫Ωfk,i⋅(uεk,i−1+wk,i−wk,i−1)dx−∫∂SΩgk,i⋅(uεk,i−1+wk,i−wk,i−1)dH1+∫Γε,k,i−1κdH1=E(tk,i−1;Γε,k,i−1)+∫ΩCEuεk,i−1:E(wk,i−wk,i−1)dx+12∫ΩCE(wk,i−wk,i−1):E(wk,i−wk,i−1)dx−∫Ωfk,i⋅(wk,i−wk,i−1)dx−∫Ω(fk,i−fk,i−1)⋅uεk,i−1dx−∫∂SΩgk,i⋅(wk,i−wk,i−1)dH1−∫∂SΩ(gk,i−gk,i−1)⋅uεk,i−1dH1+∫Γε,k,i−1κdH1. |
Iterating the above chain of inequalities for i∈{1,…,k} and using (H2) we deduce (4.2), which, together with (b), implies (c).
In the following proposition we discuss the properties of the limit of the sequence Γε,k as k→∞.
Proposition 4.3. For every ε>0 there exists a subsequence (not relabeled) of Γε,k and a set function t↦Γε(t)∈Rη such that Γmε,k(t) converges to Γmε(t) in the Hausdorff metric for every t∈[0,T] and every m∈{1,…,M}, and
(a) Γε is nondecreasing in time;
(b) lmε,k⇀lmε weakly in H1(0,T) and lmε,k(t)→lmε(t) for every t∈[0,T] and every m, where lmε(t):=H1(Γmε(t));
(c) ∇uε,k(t)→∇uε(t) strongly in L2(Ω;M2) for every t∈[0,T], where uε(t):=u(t;Γε(t));
(d) Gε,k(t)→Gε(t) for every t∈[0,T], where Gε(t):=G(t;Γε(t));
(e) Gε,k→Gε in L2(0,T).
Moreover, along a suitable (not relabeled) subsequence, we have
(f) ε‖˙lmε‖22 is uniformly bounded in ε for every m∈{1,…,M};
(g) ‖∇uε(t)‖2 is uniformly bounded in ε and t.
Proof. For brevity, in the following we will not relabel subsequences. For ε>0 let us consider the subsequence Γε,k detected in (b) and (c) of Proposition 4.2. Since Γε,k is a sequence of increasing set functions with uniformly bounded length, there exists a nondecreasing set function Γε:[0,T]→Rη such that, up to a further subsequence, Γε,k(t) converges to Γε(t) in the Hausdorff metric of sets for every t∈[0,T]. Hence, for every m∈{1,…,M} and every t∈[0,T] it holds Γmε,k(t)→Γmε(t).
For ε>0 fixed, by Proposition 4.2 we have that lmε,k∈H1(0,T) is bounded w.r.t. k and m. Therefore, for every m∈{1,…,M} the sequence lmε,k converges weakly in H1(0,T) to a nondecreasing function lmε. Up to a further subsequence, we may assume that lmε,k(t)→lmε(t) for every t∈[0,T] and every m. In particular, lmε(t)=H1(Γmε(t)), so that (b) is proven. We also notice that, because of the continuity of lmε, we have that Γ_ε,k(t)→Γε(t) in the Hausdorff metric as k→∞.
The L2-convergence of ∇uε,k(t) to ∇uε(t) is a consequence of the convergence of Γε,k(t) to Γε(t) and of Lemma 2.4. In a similar way, since Γ_ε,k(t) converges to Γε(t), ∇u_ε,k(t)→∇uε(t) in L2(Ω;M2). Moreover, by Remark 3.11 we have that Gε,k(t)→G(t;Γε(t))=:Gε(t) for every t∈[0,T], so that (d) holds. Being ‖∇uε,k(t)‖2 and ‖uε,k(t)‖2 bounded uniformly w.r.t. t and k, again by Remark 3.11 we infer that Gε,k(t) is bounded, so that Gε,k→Gε in L2(0,T) and (e) is concluded.
In order to prove (f) and (g), we employ Proposition 4.2, obtaining
‖Euε(t)‖22−‖uε(t)‖H1≤Cfor every ε>0 and every t∈[0,T], | (4.4) |
where C>0 is independent of t and ε. Arguing as in the proof of Proposition 4.2, we have that uε(t)∈H1(Ω∖Γε(T);R2) for every t∈[0,T]. Since Γε(T)∈Rη has a uniformly bounded length, we may assume that, up to a subsequence, Γε(T)→ˆΓ∈Rη in the Hausdorff metric of sets. Thus, we can apply Proposition 2.3 to Γε(T), ˆΓ, and uε(t), to deduce from (4.4) that ‖∇uε(t)‖2 is bounded uniformly w.r.t. ε and t, so that (g) holds.
Finally, we pass to the liminf in (4.2) for t=T, obtaining
ε2M∑m=1∫T0|˙lmε(t)|2dt+C1(‖Euε(T)‖22−‖uε(T)‖2)≤E(0;Γ0)+∫Γ0κdH1+∫T0∫ΩCEuε(t):E˙w(t)dxdt−∫T0∫Ω˙f(t)⋅uε(t)dxdt−∫T0∫Ωf(t)⋅˙w(t)dxdt−∫T0∫∂SΩ˙g(t)⋅uε(t)dH1dt−∫T0∫∂SΩg(t)⋅˙w(t)dH1dt. |
By the boundedness of ‖∇uε(t)‖2 and of ‖uε(t)‖2 we immediately get (f), and the proof is thus concluded. We are now in a position to prove Proposition 1.5.
Proof of Proposition 1.5. Let Γε, lmε, and Gmε be the functions determined in Proposition 4.3. Since lmε is nondecreasing, (G1)ε is satisfied. In order to prove (G2)ε let us consider ψ∈L2(0,T) with ψ≥0. By (G2)k we have
∫T0(κ(Pmε,k(t))−Gmε,k(t)+ε˙lmε,k(t))ψ(t)dt≥0. | (4.5) |
From the Hausdorff convergence of Γmε,k(t) to Γmε(t) it follows that Pmε,k(t)→Pmε(t) for every t∈[0,T] and every m, where Pmε(t) stands for the tip of Γmε(t). By hypothesis (H4) we have that κ(Pmε,k)→κ(Pmε) in L2(0,T). Hence, passing to the limit in (4.5) as k→∞ and taking into account (e) of Proposition 4.3 we get
∫T0(κ(Pmε(t))−Gmε(t)+ε˙lmε(t))ψ(t)dt≥0. |
By the arbitrariness of ψ∈L2(0,T), ψ≥0, we infer (G2)ε.
As for (G3)ε, we integrate (G3)k over [0,T] and pass to the liminf as k→∞. By (b) and (e) of Proposition 4.3 and by the convergence of κ(Pmε,k) to κ(Pmε) we obtain
∫T0˙lmε(t)(κ(Pmε(t))−Gmε(t)+ε˙lmε(t))dt≤0. |
Combining the previous inequality with (G1)ε and (G2)ε we deduce (G3)ε. Finally, the uniform boundedness of ε‖˙lmε‖22 has been stated in (f) of Proposition 4.3.
Remark 4.4. Let Γε be as in Proposition 1.5. Then, for every t∈[0,T] it holds
F(t;Γε(t))= F(0;Γ0)−M∑m=1∫t0(Gmε(τ)−κ(Pmε))˙lmε(τ)dτ+∫t0∫ΩCEuε(τ):E˙w(τ)dxdτ−∫t0∫Ω˙f(τ)⋅uε(τ)dxdτ−∫t0∫Ωf(τ)⋅˙w(τ)dxdτ−∫t0∫∂SΩ˙g(τ)⋅u(τ)dH1dτ−∫t0∫∂SΩg(τ)⋅˙w(τ)dH1dτ. | (4.6) |
Indeed, being lε∈H1([0,T];RM), the function t↦F(t,Γε(t))=E(t;Γε(t))+K(Γε(t)) belongs to H1(0,T) with
ddtF(t;Γε(t))=∂tE(t;Γε(t))−M∑m=1[Gmε(t)−κ(Pmε(t))]˙lmε(t)for a.e. t∈[0,T]. |
We conclude with the proof of Theorem 1.6.
Proof of Theorem 1.6. For ε>0 and m∈{1,…,M} let Γε, Γmε, and lmε be the viscous evolutions determined in Proposition 1.5. Let us consider, without relabeling, the ε-subsequence satisfying (f) and (g) of Proposition 4.3. Since Γε is a sequence of nondecreasing set functions and H1(Γε(t)) is uniformly bounded w.r.t. t∈[0,T] and ε>0, there exists a nondecreasing set function Γ:[0,T]→Rη such that Γε(t)→Γ(t) in the Hausdorff metric of sets for every t∈[0,T]. In particular, Γmε(t)→Γm(t) for every t and every m∈{1,…,M}, where Γ(t)=⋃Mm=1Γm(t). Moreover, being lmε a sequence of bounded nondecreasing functions, we may assume that, up to a further subsequence, lmε(t)→lm(t) for every t∈[0,T] and lmε→lm in L2(0,T). In particular, lm(t)=H1(Γm(t)) and (G1) is proven.
In order to show (G2), let us consider ψ∈L2(0,T) with ψ≥0. In view of (G2)ε we have
∫T0(κ(Pmε(t))−Gmε(t)+ε˙lmε(t))ψ(t)dt≥0. | (4.7) |
Since Γmε(t)→Γm(t), we have that Pmε(t)→Pm(t) for every t and every m, where Pm(t) is the tip of Γm(t). Thus, by hypothesis (H4) we get that κ(Pmε)→κ(Pm) in L2(0,T) for every m. From (e) and (f) of Proposition 4.3 we deduce that ε˙lmε→0 and Gmε→Gm in L2(0,T). Hence, passing to the limit in (4.7) we get
∫T0(κ(Pm(t))−Gm(t))ψ(t)dt≥0for every ψ∈L2(0,T), ψ≥0. |
As a consequence, κ(Pm(t))−Gm(t)≥0 for a.e. t∈[0,T]. By continuity, this inequality holds in all the continuity points of Γm(t). Hence, (G2) is proven.
As for (G3), we integrate (G3)ε over the interval [0,T] and notice that the term ε(˙lmε)2 is positive, so that
∫T0˙lmε(t)(κ(Pmε(t))−Gmε(t))dt≤0. |
Passing to the limit in the previous inequality we get
∫T0˙lm(t)(κ(Pm(t))−Gm(t))dt≤0. | (4.8) |
Combining (4.8) with (G1) and (G2) we deduce (G3).
This section is devoted to the proof of Theorem 1.7. The strategy is to perform a change of variables which transforms the lengths lmε obtained in Proposition 1.5 in absolutely continuous functions. Loosely speaking, this is done by a reparametrization of time which continuously interpolates the time discontinuities of the solution lm.
Let us fix the sequence ε→0 determined in Proposition 4.3 and Theorem 1.6. For t∈[0,T] we set
σε(t):=t+M∑m=1(lmε(t)−lm0). | (5.1) |
Thanks to the properties of lmε (see Proposition 1.5), σε is strictly increasing, continuous, and ˙σε(t)≥1 for every ε>0 and a.e. t∈[0,T]. Therefore, σε is invertible and we denote its inverse with ˜tε:[0,σε(T)]→[0,T]. We deduce that ˜tε is strictly increasing, continuous, and 0<˜t′ε(σ)≤1 for every ε>0 and a.e. σ∈[0,σε(T)], where the symbol ′ stands for the derivative with respect to σ.
For m=1,…,M and σ∈[0,σε(T)], we set
˜lmε(σ):=lmε(˜tε(σ)),˜lε(σ):=(˜l1ε(σ),…,˜lMε(σ)),˜l′ε(σ):=((˜l1ε)′(σ),…,(˜lMε)′(σ)),˜Γε(σ):=Γε(˜tε(σ)),˜Γmε(σ):=Γmε(˜tε(σ)),˜Pmε(σ):=Pmε(˜tε(σ)). |
By (5.1) we have σ=˜tε(σ)+|˜lε(σ)|1−|l0|1. Differentiating this relation we get
˜t′ε(σ)+|˜l′ε(σ)|1=1 | (5.2) |
for every ε>0 and a.e. σ∈[0,σε(T)]. By (5.2) and the monotonicity of ˜lmε we have 0≤(˜lmε)′(σ)≤1 for every ε>0, every m=1,…,M, and a.e. σ∈[0,σε(T)]. Moreover, ˜tε and ˜lε are Lipschitz continuous.
We define ˜Gmε(σ):=Gm(˜tε(σ);˜Γε(σ)) for σ∈[0,σε(T)] and ˉS:=supε>0σε(T), which is bounded by a constant depending on T and on the class Rη. In order to deal with functions defined on the same interval, we extend ˜tε, ˜lε, ˜Γε, ˜Γmε, ˜t′ε, and ˜l′ε on (σε(T),ˉS] by ˜tε(σ):=˜tε(σε(T)), ˜lε(σ):=˜lε(σε(T)), ˜Γε(σ):=˜Γε(σε(T)), ˜Γmε(σ):=˜Γmε(σε(T)), ˜t′ε(σ):=0, and ˜s′ε(σ):=0.
Recalling that ˜t′ε(σ)>0 on [0,σε(T)], the Griffith's criterion stated in Proposition 1.4 reads in the new variables as
{(˜lmε)′(σ)≥0,κ(˜Pmε(σ))˜t′ε(σ)−˜Gmε(σ)˜t′ε(σ)+ε(˜lmε)′(σ)≥0,(˜lmε)′(σ)(κ(˜Pmε(σ))˜t′ε(σ)−˜Gmε(σ)˜t′ε(σ)+ε(˜lmε)′(σ))=0, | (5.3) |
for every m, every ε, and a.e. σ∈[0,ˉS].
Finally, we observe that by (f) of Proposition 4.3
ε∫σε(T)0|(˜lmε)′(σ)|22dσ=ε∫σε(T)0|˙lmε(˜tε(σ))|22(˜t′ε)2(σ)dσ≤ε∫σε(T)0|˙lmε(˜tε(σ))|22˜t′ε(σ)dσ=ε∫T0|˙lmε(t)|22dt≤C |
uniformly in ε and m∈{1,…,M}. Therefore, ε(˜lmε)′→0 in L2(0,ˉS).
Passing to the limit as ε→0, we are now able to prove Theorem 1.7, showing that the parametrized solution ˜Γ satisfies a generalized Griffith's criterion.
Proof of Theorem 1.7. Since ˜Γε:[0,ˉS]→Rη is a nondecreasing set function with uniformly bounded length, there exists ˜Γ:[0,ˉS]→Rη such that, up to a not relabeled subsequence, ˜Γε(σ)→˜Γ(σ) and ˜Γmε(σ)→˜Γm(σ) in the Hausdorff metric of sets for every σ∈[0,ˉS] and every m∈{1,…,M}. We denote with ˜Pm(σ) the tip of ˜Γm(σ) and we notice that ˜Pmε(σ)→˜Pm(σ) for every σ∈[0,ˉS] and every m∈{1,…,M}.
Being ˜tε,˜lmε bounded in W1,∞(0,ˉS), up to a further subsequence we have that ˜tε and ˜lmε converge weakly* in W1,∞(0,ˉS) to some functions ˜t and ˜lm, respectively, As a consequence, we have that ˜lm(σ)=H1(˜Γm(σ)), so that ˜Γm:[0,S]→Rη is continuous in the Hausdorff metric of sets. We can also assume that σε(T)→S and ˜t,˜lm∈W1,∞(0,S). Moreover, writing (5.2) in an integral form and passing to the limit, we deduce that for a.e. σ∈[0,S]
˜t′(σ)+|˜l′(σ)|1=˜t′(σ)+M∑m=1(˜lm)′(σ)=1, | (5.4) |
where we have set ˜l(σ):=(l1(σ),…,˜lM(σ)). For m=1…,M and σ∈[0,S] we define,
˜Gm(σ):=Gm(˜t(σ);˜Γ(σ)),˜G(σ):=(˜G1(σ),…,˜GM(σ)). |
We notice that, by Remark 3.11, ˜Gε(σ) converges to ˜G(σ) for every σ∈[0,S] and ˜Gε→˜G in L2(0,S), as ε→0.
By the monotonicity of ˜t and ˜lm, we have ˜t′(σ)≥0 and (˜lm)′(σ)≥0 for every m and a.e. σ∈[0,S]. Moreover, by (5.4) they can not be simultaneously zero.
Let us fix m∈{1,…,M} and ψ∈L2(0,S) with ψ≥0. Thanks to (5.3), for every ε we have
∫S0(κ(˜Pmε(σ))˜t′ε(σ)−˜Gmε(σ)˜t′ε(σ)+ε(˜lmε)′(σ))ψ(σ)dσ≥0. | (5.5) |
Since ˜t′ε converges to ˜t′ weakly* in L∞(0,S), ε(˜lmε)′→0 in L2(0,S), ˜Pmε(σ)→˜Pm(σ) for σ∈[0,S], and ˜Gmε→˜Gm in L2(0,S), passing to the limit in (5.5) as ε→0 we get
∫S0(κ(˜Pm(σ))˜t′(σ)−˜Gm(σ)˜t′(σ))ψ(σ)dσ≥0, |
which implies (pG2).
We notice that if (pG1), (pG2) and (5.4) hold, then (pG3) and (pG4) are equivalent to the following property:
if ˜Gm(ˉσ)<κ(˜Pm(ˉσ)) for some m and some ˉσ∈(0,S), then ˜lm is locally constant around ˉσ. |
Let us therefore assume that ˜Gm(ˉσ)<κ(˜Pm(ˉσ)). We first claim that there exist ˉε>0 and δ>0 such that ˜Gmε(σ)<κ(˜Pmε(σ)) for every σ∈(ˉσ−δ,ˉσ+δ) and every ε≤ˉε. By contradiction, suppose that this is not the case. Then, there exist σk→ˉσ and εk→0 such that ˜Gmεk(σk)≥κ(˜Pmεk(σk)). By continuity and monotonicity of ˜Γm, we have that ˜Γmεk(σk)→˜Γm(σ) in the Hausdorff metric of sets and ˜Pmεk(σk)→˜Pm(σ) as k→∞. Hence, the continuity of the energy release rate and the hypothesis (H4) lead us to the contradiction ˜Gm(ˉσ)≥κ(˜Pm(ˉσ)).
Let δ and ˉε be as above. We deduce from the Griffith's criterion (5.3) that ˜lmε is constant in (ˉσ−δ,ˉσ+δ) for every ε≤ˉε. Since ˜lmε converges to ˜lm weakly* in W1,∞(0,S), we get that ˜lm is locally constant around ˉσ, and this concludes the proof of (pG3) and (pG4).
In order to show that Γ(˜t(σ))=˜Γ(σ) for every σ∈[0,S] such that ˜t′(σ)>0, we define
s(t):=min{s∈[0,S]:˜t(s)=t}for every t∈[0,T]. |
If ˜t′(σ)>0, then we have s(˜t(σ))=σ and s(˜t(ˉσ))≠s(˜t(σ)) for ˉσ≠σ. Let us prove that ˜t(σ) is a continuity point for Γ, where the map t↦Γ(t) has been determined in Theorem 1.5. By contradiction, assume that ˜t(σ) is a discontinuity point of Γ. Then, there exist t1ε<t2ε such that t1ε,t2ε→˜t(σ) and Γε(t1ε)→Γ−(˜t(σ)) and Γε(t2ε)→Γ+(˜t(σ)) in the Hausdorff metric of sets, where we have denoted with Γ±(˜t(σ)) the left and right limits of Γ(t) in ˜t(σ). As a consequence, s(˜t(ˉσ))=s(˜t(σ)) for ˉσ∈(σ−H1(Γ+(˜t(σ))∖Γ−(˜t(σ))),σ+H1(Γ+(˜t(σ))∖Γ−(˜t(σ)))), which is a contradiction. Hence, ˜t(σ) is a continuity point of t↦Γ(t). Therefore, ˜Γε(σ)=Γε(˜tε(σ)) converges to Γ(˜t(σ)) in the Hausdorff metric of sets. This implies that Γ(˜t(σ))=˜Γ(σ).
We conclude with the energy balance (1.6). Starting from Eq. (4.6) and using the change of variable t=tε(σ), for ε>0 and for s∈[0,S] we get
F(˜tε(s);˜Γε(s))= F(0;Γ0)+∫s0∫ΩCE˜uε(σ):E˙w(˜tε(σ))˜t′ε(σ)dxdσ−M∑m=1∫s0(˜Gmε(σ)−κ(˜Pmε(σ)))(˜lmε)′(σ)dσ−∫s0∫Ω˙f(˜tε(σ))⋅˜uε(σ)˜t′ε(σ)dxdσ−∫s0∫Ωf(˜tε(σ))⋅˙w(τ)˜t′ε(σ)dxdσ−∫s0∫∂SΩ˙g(˜tε(σ))⋅˜u(σ)˜t′ε(σ)dH1dσ−∫s0∫∂SΩg(˜tε(σ))⋅˙w(˜tε(σ))˜t′ε(σ)dH1dσ, | (5.6) |
where we have set ˜uε(σ):=uε(˜tε(σ)). Since ˜tε and ˜lmε converge weakly* in W1,∞(0,S) to ˜t and ˜lm, respectively, ˜Gmε converges to ˜Gm in L2(0,S), and E˜uε(σ) converges to E˜u(σ)=Eu(t(σ),˜Γ(σ)) in L2(Ω;M2) (cf. Lemma 2.4), passing to the limit as ε→0 in (5.6) we get (1.6). This concludes the proof of the theorem.
The authors would like to acknowledge the kind hospitality of the Erwin Schrödinger International Institute for Mathematics and Physics (ESI), where part of this research was developed during the workshop New trends in the variational modeling of failure phenomena. All authors would like to acknowledge the kind hospitality of the University of Naples Federico II, to which GL was affiliated when this research was initiated. SA wishes to thank the Technical University of Munich, where he worked during the preparation of this paper, with partial support from the SFB project TRR109 Shearlet approximation of brittle fracture evolutions. GL and IL are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). IL acknowledges the partial support of the PEPS-JCJC project 2019 Évolution quasi-statique de la rupture cohésive à travers une approche de champ de phase. GL received support from the INdAM-GNAMPA 2019 Project Modellazione e studio di proprietà asintotiche per problemi variazionali in fenomeni anelastici.
The authors declare that they have no conflict of interest and guarantee the compliance with the Ethics Guidelines of the journal.
[1] |
Almi S (2017) Energy release rate and quasi-static evolution via vanishing viscosity in a fracture model depending on the crack opening. ESAIM Control Optim Calc Var 23: 791-826. doi: 10.1051/cocv/2016014
![]() |
[2] |
Almi S (2018) Quasi-static hydraulic crack growth driven by Darcy's law. Adv Calc Var 11: 161-191. doi: 10.1515/acv-2016-0029
![]() |
[3] |
Almi S, Lucardesi I (2018) Energy release rate and stress intensity factors in planar elasticity in presence of smooth cracks. Nonlinear Differ Equ Appl 25: 43. doi: 10.1007/s00030-018-0536-4
![]() |
[4] | Argatov II, Nazarov SA (2002) Energy release in the kinking of a crack in a plane anisotropic body. Prikl Mat Mekh 66: 502-514. |
[5] |
Babadjian JF, Chambolle A, Lemenant A (2015) Energy release rate for non-smooth cracks in planar elasticity. J Éc polytech Math 2: 117-152. doi: 10.5802/jep.19
![]() |
[6] |
Brokate M, Khludnev A (2004) On crack propagation shapes in elastic bodies. Z Angew Math Phys 55: 318-329. doi: 10.1007/s00033-003-3026-3
![]() |
[7] |
Chambolle A (2003) A density result in two-dimensional linearized elasticity, and applications. Arch Ration Mech Anal 167: 211-233. doi: 10.1007/s00205-002-0240-7
![]() |
[8] |
Chambolle A, Francfort GA, Marigo JJ (2010) Revisiting energy release rates in brittle fracture. J Nonlinear Sci 20: 395-424. doi: 10.1007/s00332-010-9061-2
![]() |
[9] |
Chambolle A, Giacomini A, Ponsiglione M (2008) Crack initiation in brittle materials. Arch Ration Mech Anal 188: 309-349. doi: 10.1007/s00205-007-0080-6
![]() |
[10] |
Chambolle A, Lemenant A (2013) The stress intensity factor for non-smooth fractures in antiplane elasticity. Calc Var Partial Dif 47: 589-610. doi: 10.1007/s00526-012-0529-9
![]() |
[11] |
Costabel M, Dauge M (2002) Crack singularities for general elliptic systems. Math Nachr 235: 29-49. doi: 10.1002/1522-2616(200202)235:1<29::AID-MANA29>3.0.CO;2-6
![]() |
[12] |
Crismale V, Lazzaroni G (2017) Quasistatic crack growth based on viscous approximation: A model with branching and kinking. Nonlinear Differ Equ Appl 24: 7. doi: 10.1007/s00030-016-0426-6
![]() |
[13] |
Dal Maso G, DeSimone A, Solombrino F (2011) Quasistatic evolution for Cam-Clay plasticity: A weak formulation via viscoplastic regularization and time rescaling. Calc Var Partial Dif 40: 125-181. doi: 10.1007/s00526-010-0336-0
![]() |
[14] |
Dal Maso G, DeSimone A, Solombrino F (2012) Quasistatic evolution for Cam-Clay plasticity: Properties of the viscosity solution. Calc Var Partial Dif 44: 495-541. doi: 10.1007/s00526-011-0443-6
![]() |
[15] |
Dal Maso G, Francfort GA, Toader R (2005) Quasistatic crack growth in nonlinear elasticity. Arch Ration Mech Anal 176: 165-225. doi: 10.1007/s00205-004-0351-4
![]() |
[16] |
Dal Maso G, Lazzaroni G (2010) Quasistatic crack growth in finite elasticity with noninterpenetration. Ann Inst H Poincaré Anal Non Linéaire 27: 257-290. doi: 10.1016/j.anihpc.2009.09.006
![]() |
[17] |
Dal Maso G, Morandotti M (2017) A model for the quasistatic growth of cracks with fractional dimension. Nonlinear Anal 154: 43-58. doi: 10.1016/j.na.2016.03.007
![]() |
[18] |
Dal Maso G, Orlando G, Toader R (2015) Laplace equation in a domain with a rectilinear crack: Higher order derivatives of the energy with respect to the crack length. Nonlinear Differ Equ Appl 22: 449-476. doi: 10.1007/s00030-014-0291-0
![]() |
[19] |
Dal Maso G, Toader R (2002) A model for the quasi-static growth of brittle fractures: Existence and approximation results. Arch Ration Mech Anal 162: 101-135. doi: 10.1007/s002050100187
![]() |
[20] | Dauge M (1988) Smoothness and asymptotics of solutions, In: Elliptic Boundary Value Problems on Corner Domains, Berlin: Springer-Verlag. |
[21] |
Destuynder P, Djaoua M (1981) Sur une interprétation mathématique de l'intégrale de Rice en théorie de la rupture fragile. Math Method Appl Sci 3: 70-87. doi: 10.1002/mma.1670030106
![]() |
[22] | Efendiev MA, Mielke A (2006) On the rate-independent limit of systems with dry friction and small viscosity. J Convex Anal 13: 151-167. |
[23] |
Fonseca I, Fusco N, Leoni G, et al. (2007) Equilibrium configurations of epitaxially strained crystalline films: Existence and regularity results. Arch Ration Mech Anal 186: 477-537. doi: 10.1007/s00205-007-0082-4
![]() |
[24] |
Francfort GA, Larsen CJ (2003) Existence and convergence for quasi-static evolution in brittle fracture. Commun Pur Appl Math 56: 1465-1500. doi: 10.1002/cpa.3039
![]() |
[25] |
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
![]() |
[26] |
Friedrich M, Solombrino F (2018) Quasistatic crack growth in 2d-linearized elasticity. Ann Inst H Poincaré Anal Non Linéaire 35: 27-64. doi: 10.1016/j.anihpc.2017.03.002
![]() |
[27] |
Griffith AA (1921) The phenomena of rupture and flow in solids. Philos Trans R Soc London A 221: 163-198. doi: 10.1098/rsta.1921.0006
![]() |
[28] | Grisvard P (1985) Monographs and Studies in Mathematics, In: Elliptic Problems in Consmooth Domains, Boston: Pitman. |
[29] |
Khludnev AM, Shcherbakov VV (2018) A note on crack propagation paths inside elastic bodies. Appl Math Lett 79: 80-84. doi: 10.1016/j.aml.2017.11.023
![]() |
[30] |
Khludnev AM, Sokolowski J (2000) Griffith formulae for elasticity systems with unilateral conditions in domains with cracks. Eur J Mech A Solids 19: 105-119. doi: 10.1016/S0997-7538(00)00138-8
![]() |
[31] | Knees D (2011) A short survey on energy release rates, In: Dal Maso G, Larsen CJ, Ortner C, Mini-Workshop: Mathematical Models, Analysis, and Numerical Methods for Dynamic Fracture, Oberwolfach Reports, 8: 1216-1219. |
[32] |
Knees D, Mielke A (2008) Energy release rate for cracks in finite-strain elasticity. Math Method Appl Sci 31: 501-528. doi: 10.1002/mma.922
![]() |
[33] |
Knees D, Mielke A, Zanini C (2008) On the inviscid limit of a model for crack propagation. Math Mod Meth Appl Sci 18: 1529-1569. doi: 10.1142/S0218202508003121
![]() |
[34] |
Knees D, Rossi R, Zanini C (2013) A vanishing viscosity approach to a rate-independent damage model. Math Mod Meth Appl Sci 23: 565-616. doi: 10.1142/S021820251250056X
![]() |
[35] |
Knees D, Rossi R, Zanini C (2015) A quasilinear differential inclusion for viscous and rateindependent damage systems in non-smooth domains. Nonlinear Anal Real 24: 126-162. doi: 10.1016/j.nonrwa.2015.02.001
![]() |
[36] |
Knees D, Zanini C, Mielke A (2010) Crack growth in polyconvex materials. Physica D 239: 1470-1484. doi: 10.1016/j.physd.2009.02.008
![]() |
[37] | Krantz SG, Parks HR (2002) The Implicit Function Theorem: History, theory, and applications, Boston: Birkhäuser Boston Inc. |
[38] |
Lazzaroni G, Toader R (2011) A model for crack propagation based on viscous approximation. Math Mod Meth Appl Sci 21: 2019-2047. doi: 10.1142/S0218202511005647
![]() |
[39] |
Lazzaroni G, Toader R (2011) Energy release rate and stress intensity factor in antiplane elasticity. J Math Pure Appl 95: 565-584. doi: 10.1016/j.matpur.2011.01.001
![]() |
[40] | Lazzaroni G, Toader R (2013) Some remarks on the viscous approximation of crack growth. Discrete Contin Dyn Syst Ser S 6: 131-146. |
[41] |
Mielke A, Rossi R, Savaré G (2009) Modeling solutions with jumps for rate-independent systems on metric spaces. Discrete Contin Dyn Syst 25: 585-615. doi: 10.3934/dcds.2009.25.585
![]() |
[42] |
Mielke A, Rossi R, Savaré G (2012) BV solutions and viscosity approximations of rate-independent systems. ESAIM Control Optim Calc Var 18: 36-80. doi: 10.1051/cocv/2010054
![]() |
[43] |
Mielke A, Rossi R, Savaré G (2016) Balanced viscosity (BV) solutions to infinite-dimensional rate-independent systems. J Eur Math Soc 18: 2107-2165. doi: 10.4171/JEMS/639
![]() |
[44] | Mielke A, Roubíček T (2015) Rate-Independent Systems, New York: Springer. |
[45] |
Nazarov SA, Specovius-Neugebauer M, Steigemann M (2014) Crack propagation in anisotropic composite structures. Asymptot Anal 86: 123-153. doi: 10.3233/ASY-131187
![]() |
[46] |
Negri M (2011) Energy release rate along a kinked path. Math Method Appl Sci 34: 384-396. doi: 10.1002/mma.1362
![]() |
[47] |
Negri M (2014) Quasi-static rate-independent evolutions: Characterization, existence, approximation and application to fracture mechanics. ESAIM Control Optim Calc Var 20: 983-1008. doi: 10.1051/cocv/2014004
![]() |
[48] |
Negri M, Ortner C (2008) Quasi-static crack propagation by Griffith's criterion. Math Mod Meth Appl Sci 18: 1895-1925. doi: 10.1142/S0218202508003236
![]() |
[49] |
Negri M, Toader R (2015) Scaling in fracture mechanics by Bažant law: From finite to linearized elasticity. Math Mod Meth Appl Sci 25: 1389-1420. doi: 10.1142/S0218202515500360
![]() |
1. | L. De Luca, M. Ponsiglione, Variational models in elasticity, 2021, 3, 2640-3501, 1, 10.3934/mine.2021015 | |
2. | Stefano Almi, Irreversibility and alternate minimization in phase field fracture: a viscosity approach, 2020, 71, 0044-2275, 10.1007/s00033-020-01357-x | |
3. | Marco Bonacini, Sergio Conti, Flaviana Iurlano, Cohesive Fracture in 1D: Quasi-static Evolution and Derivation from Static Phase-Field Models, 2021, 239, 0003-9527, 1501, 10.1007/s00205-020-01597-1 | |
4. | Sukhminder Singh, Lukas Pflug, Michael Stingl, Material optimization to enhance delamination resistance of composite structures using viscous regularization, 2021, 382, 00457825, 113881, 10.1016/j.cma.2021.113881 | |
5. | Vito Crismale, Riccarda Rossi, Balanced Viscosity Solutions to a Rate-Independent Coupled Elasto-plastic Damage System, 2021, 53, 0036-1410, 3420, 10.1137/19M1303563 | |
6. | Filippo Riva, Giovanni Scilla, Francesco Solombrino, The notions of inertial balanced viscosity and inertial virtual viscosity solution for rate-independent systems, 2022, 0, 1864-8266, 10.1515/acv-2021-0073 | |
7. | Maicol Caponi, Vito Crismale, A rate-independent model for geomaterials under compression coupling strain gradient plasticity and damage, 2025, 31, 1292-8119, 14, 10.1051/cocv/2025004 |