Home A unilateral L2-gradient flow and its quasi-static limit in phase-field fracture by an alternate minimizing movement
Article Publicly Available

A unilateral L2-gradient flow and its quasi-static limit in phase-field fracture by an alternate minimizing movement

  • Matteo Negri ORCID logo EMAIL logo
Published/Copyright: June 4, 2017

Abstract

We consider an evolution in phase-field fracture which combines, in a system of PDEs, an irreversible gradient-flow for the phase-field variable with the equilibrium equation for the displacement field. We introduce a discretization in time and define a discrete solution by means of a 1-step alternate minimization scheme, with a quadratic L2-penalty in the phase-field variable (i.e. an alternate minimizing movement). First, we prove that discrete solutions converge to a solution of our system of PDEs. Then we show that the vanishing viscosity limit is a quasi-static (parametrized) BV-evolution. All these solutions are described both in terms of energy balance and, equivalently, by PDEs within the natural framework of W1,2(0,T;L2).

MSC 2010: 74R10; 35M86; 49M25

1 Introduction

Phase-field approaches are widely used to simulate crack propagation in academical and industrial applications: even within the linear-elastic setting there are plenty of phase-field models, based on different choices of potentials and evolution laws, see e.g. [2, 6, 10, 11, 18, 23, 25, 26, 34, 35] or the recent review [3]. In the present work, we will study in detail a couple of evolutions generated by a phase-field energy of the form

(t,u,v)=Ω(v2+η)W(Du~(t))𝑑x+12GcΩ(v-1)2+|v|2dx,

where Ω2 is a bounded Lipschitz domain, u~(t)=u+g(t) is the displacement fields with uH01(Ω) (so that u~(t)=g(t) on Ω), W(Du~) is a linear elastic energy density, vH1(Ω;[0,1]) is the phase-field variable, Gc>0 is toughness while η>0 is a regularization parameter.

Functionals like provide an elliptic regularization of free discontinuity functional: for instance, neglecting boundary conditions, for ε0+ and 0<ηε=o(ε) the Γ-limit of the functionals

ε(u,v)=Ω(v2+ηε)W(Du)𝑑x+GcΩ(4ε)-1(v-1)2+ε|v|2dx

is of the form

0(u)=ΩW(u)𝑑x+Gc1(J(u)),

where J(u) is the set of discontinuity points of the displacement field u and 1 denotes Hausdorff measure. Roughly speaking, the set J(u) represents the crack. Convergence has rigorously been proved in the framework of SBD2 and GSBD2 spaces respectively in [12] and [20] while in the scalar framework of the space GSBV2 it was proved in the well-known paper [5]. In our work, we will study an evolution in time, rather than a limit for ε0+, hence we will work with the functional , omitting, for simplicity of notation, the dependence on the “internal length” ε.

Our starting point is a time-discrete evolution generated by an alternate unilateral minimizing movement. For τ>0 let tn=nτ[0,T] for n be the time discretization. Then the incremental problem is the following: given (un-1,vn-1) (at time tn-1) the configuration (un,vn) at time tn is obtained by solving

(1.1){vnargmin{(tn,un,v)+12τv-vn-12:vvn-1,vH1},unargmin{(tn,u,vn-1):uH01}.

Similar discrete schemes have been used in applications, e.g. in [23, 35]. Note that the updated configuration is determined by a single iteration in each variable, that each minimization problem is well posed (thanks to η>0) and that the constraint vvn-1 models irreversibility. This scheme takes, of course, full advantage of the separate quadratic structure of (t,,). Our first result proves that (as τ0) the time-discrete evolutions converge to a time-continuous evolution t(u(t),v(t)), where v() is monotone non-increasing and satisfies for every t[0,T] the energy balance

(1.2)(t,u(t),v(t))=(0,u0,v0)-120tv˙(r)L22+|v-(r,u(r),v(r))|L22dr+0tt(r,u(r),v(r))𝑑r,

where |v-(t,u,v)|L2=sup{-v(t,u,v)[ξ]:ξ0,ξL21} is the unilateral slope. Note that, by irreversibility, only negative variations are allowed; for this reason a minus sign is added to the notation |v|L2 of the (unconstrained) slope. Equation (1.2) can be considered as De Giorgi’s integral characterization of gradient flows; indeed, for a.e. t[0,T] the time continuous limit solves also the following system of PDEs:

(1.3){v˙(t)=-[v(t)W(Du~(t))+Gc(v(t)-1)-GcΔv(t)]+,div(𝝈v(t)(u~(t)))=0,

where 𝝈v(u~)=(v2+η)𝝈(u~) denotes the phase-field stress. Technically, we will see that vW1,2(0,T;L2(Ω)) and that v(t)W(Du~(t))+Gc(v(t)-1)-GcΔv(t) is a finite Radon measure with positive part []+ in L2(Ω). To better understand the variational structure behind this evolution problem, note that formally the partial derivatives of read

v(t,u,v)[ξ]=Ω(vW(Du~(t))+Gc(v-1)-GcΔv)ξ𝑑x,
u(t,u,v)[ϕ]=-Ωdiv(𝝈v(u~(t)))ϕ𝑑x.

The positive part []+ appearing in (1.3) comes from the irreversibility constraint; more precisely, the term -[v(t)W(Du~(t))+Gc(v(t)-1)-GcΔv(t)]+ is (in a suitable sense) the “projection” of -v(t,u(t),v(t)) on the set of negative variations ξ, therefore the parabolic equation can be interpreted as a unilateral gradient flow, constrained by irreversibility.

In a larger perspective, an alternate minimizing scheme has been employed in different ways also in a dynamic visco-elastic setting [24], in another gradient flow setting [7] and in a quasi-static setting [21]. In general, alternate scheme are very useful in numerical simulation since they require only the minimization of convex (in our case, quadratic) functionals. On the other hand, different approaches can provide existence of solutions for (1.3) or similar problems. For instance: [22] obtains existence introducing a unilateral minimizing movement for a “reduced” (non-convex) energy, which in our setting would be of the form (t,ut,v,u) for ut,vargmin{(t,u,v):uH01}, while [9] proves existence, for a similar problem, by a fixed point argument. In the applications, “Ginzburg–Landau” models are used in phase-field fracture for instance [1, 18, 23].

In the second part of the paper we consider the vanishing viscosity limit of (1.3). More precisely, we start with the system

{εv˙ε(t)=-[vε(t)W(Du~ε(t))+Gc(vε(t)-1)-GcΔvε(t)]+,div(𝝈vε(t)(u~ε(t)))=0,

where ε>0 is a “mobility parameter” or “viscosity”. Our goal is the characterization of the quasi-static limit, obtained as ε0. Since in the limit we expect discontinuous evolutions and since the limit is rate independent, we first parametrize the evolutions by an arc-length parameter in L2. In this way, we get a Lipschitz map s(tε(s),wε(s),zε(s)), where wε(s)=uεtε(s) and zε(s)=vεtε(s). Passing to the limit, as ε0, we get a map s(t(s),w(s),z(s)) which satisfies w(s)argmin{(t(s),w,z(s)):w=gt(s) on Ω} together with the energy balance

(1.4)(t(s),w(s),z(s))=(0,w0,z0)-0s|z-(t(r),w(r),z(r))|L2𝑑r+0st(t(r),w(r),z(r))t(r)𝑑r.

It is noteworthy that this balance together with the Lipschitz continuity of parametrized solutions imply the main properties of the quasi-static parametrized solution s(t(s),w(s),z(s)). Indeed, if t(s)>0 (i.e. in continuity points), we have equilibrium, i.e.

{[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+=0,div(𝝈z(s)(w~(s)))=0,

if t(s) is constant in (s,s) (i.e. in discontinuity points) we have the following re-parametrization of (1.4)

(1.5){λ(s)z(s)=-[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+,div(𝝈z(s)(w~(s)))=0,

where λ(s)=[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+L2. Hence the vanishing viscosity limit is labelled “parametrized BV-evolution” [28, 32].

It is interesting to compare, at least qualitatively, the quasi-static limit obtained here with the one obtained in [21]. The latter is based on the alternate minimization scheme of [11], which reads, in the one iteration version,

{vnargmin{(tn,un,v):vvn-1,vH1},unargmin{(tn,u,vn-1):uH01}.

Note that in this scheme there is no additional viscosity. As a matter of fact, using the separate quadratic structure of the energy (t,,), the above minimization problems, in u and v, are recast in [21] as the minimization of linear energies with suitable (state depending) dissipation-norms. Such dissipation-norms are called “intrinsic”, as opposed to “artificial” dissipations appearing in the vanishing viscosity approach, like the L2-norm in (1.1). Roughly speaking, the quasi-static limit of [21] is a parametrized BV-evolution for the energy with respect to these intrinsic dissipation-norms; for the detailed characterization together with several fine properties we refer to [21].

Last, but not least, let us provide some technical considerations about our results. First of all, the proofs of (1.2) and (1.3) rely essentially on the separate convexity of the energy (t,,), the lower semi-continuity of the unilateral slope and a sort of upper gradient inequality, based on a measure theoretic argument employed in [15]. All these ingredients, put together, allow us to work with evolutions of class W1,2(0,T;L2), which seems to be the natural weak setting for (1.2) and, in perspective, for more complex systems, e.g. [1], and higher-dimensional problems. We remark that our proof of existence for the unilateral gradient flow does not rely on the chain rule in W1,2(0,T;H1) (see Lemma 7). However, in order to study the quasi-static limit, and prove (1.4), it is necessary to have a uniform bound on the length of the curves s(tε(s),wε(s),zε(s)). This is a delicate technical point, which is obtained by means of a discrete Gronwall argument, cf. [22, 33], and which gives, as a by-product, a uniform bound in W1,2(0,T;H1). This bound is used, together with the chain rule, in the proof (1.5). Finally, we remark that our analysis works for domains Ω contained in 2 since it relies on Sobolev embeddings, see for instance Lemma 6. For similar technical reasons, in the N-dimensional setting gradient flows of this type have been studied, e.g. [7] and [22], by modifying the energy with some dimension depending variants; for instance Ω|v-1|2+|v|2dx=v-1H12 is replaced by v-1W1,pp (for p>N) or by v-1L22+|v|H122 (for N=3).

2 Setting, energy and its derivatives

First of all we collect the set of assumptions used through the work.

Assumptions.

We assume that Ω is an open, bounded, connected domain in 2 with Lipschitz boundary Ω. Deformations are assumed to be of the form u~=u+g(t) for u𝒰=H01(Ω,2) and gC1([0,T];W1,p¯(Ω,2)) for p¯>2. The phase-field “space” 𝒱 is H1(Ω,[0,1]).

The potential energy :[0,T]×𝒰×𝒱[0,+) is given by the following (see [5, 11]) phase-field energy for brittle fracture

(t,u,v)=12Ω(v2+η)W(Du~(t))𝑑x+12GcΩ(v-1)2+|v|2dx,

where u~(t)=u+g(t) and W(Du~)=Du~:𝑪Du~=𝜺(u~):𝝈(u~) is the linear elastic energy density, Gc>0 is the toughness while η>0 is a (small) regularization parameter. For convenience of notation, let

(t,u,v)=12Ω(v2+η)W(Du~(t))𝑑x,𝒟(v)=12GcΩ(v-1)2+|v|2dx

denote respectively the elastic and the fracture (dissipated) phase-field energy.

For the sake of simplicity we will assume that the initial configuration u0,v0 (at time t=0) is in equilibrium; since the energy (t,,) is separately quadratic, equilibrium is equivalent to separate minimality, i.e.

u0argmin{(0,v0,):u𝒰},v0argmin{(0,,u0):vv0,v𝒱}.

Next, we provide the properties of energy and derivatives which will be used in the sequel.

Lemma 1.

If tnt, unu in U and vnv in V, then

(t,u,v)lim infn+(tn,un,vn).

Proof.

Since vnv in 𝒱, it is clear that 𝒟(v)lim infn+𝒟(vn). Thus, it is enough to show that

(t,u,v)lim infn+(tn,un,vn).

First, extract a subsequence (not relabeled) such that lim infn(tn,un,vn)=limn(tn,un,vn). Since vn is bounded, we can extract a further subsequence (again not relabeled) such that vnv a.e. in Ω. By Egorov’s theorem, for every ε1 there exists ΩεΩ with |ΩΩε|<ε such that vnv uniformly in Ωε. Hence for δ1 and n1 in Ωε it holds 0(v2+η)-δ(vn2+η). Then

12Ω(vn2+η)W(Du~n(tn))𝑑x12Ω(v2+η-δ)W(Du~n(tn))χΩε𝑑x.

Defining the density 0Wε(x,ξ)=(v2(x)+η-δ)W(ξ)χΩε(x), the weak lower semi-continuity of the right-hand side (see e.g. [14, Theorem 3.4]) yields

lim infn+(tn,un,vn)12Ω(v2+η-δ)W(Du~(t))χΩε𝑑x.

To conclude, it is sufficient to take first the supremum for δ0 and then the supremum for ε0. ∎

If the displacement field u is sufficiently regular (and this is the case for our evolutions), variations of energy take a simple form; more precisely, if uW1,p(Ω,2) for some p>2, then by Lemma 6, the energy (t,u,) is Gateaux differentiable with

v(t,u,v)[ξ]=ΩvξW(Du~(t))𝑑x+GcΩ(v-1)ξ+vξdxfor all ξH1(Ω).

In the evolution, irreversibility is modeled by monotonicity of v. For this reason, both the gradient flow and the BV-evolution will be defined in terms of the following unilateral L2-slope of (t,u,): if uW1,p(Ω,2) for some p>2, let

|v-(t,u,v)|L2=|inf{v(t,u,v)[ξ]:ξH1(Ω),ξ0,ξL21}|.

For future convenience denote Ξ={ξH1(Ω),ξ0,ξL21}, so that the unilateral slope is equivalently given by

(2.1)|v-(t,u,v)|L2=sup{-v(t,u,v)[ξ]:ξΞ}.

Lemma 2.

If tnt, unu in W1,p(Ω,R2) for p>2 and vnv in V, then

|v-(t,u,v)|L2lim infn+|v-(tn,un,vn)|L2.

Proof.

First, we show that for every ξΞ we have

(2.2)limn+v(tn,un,vn)[ξ]=v(t,u,v)[ξ].

By weak convergence in H1(Ω),

limn+Ω(vn-1)ξ+vnξdx=Ω(v-1)ξ+vξdx,

while

limn+ΩvnξW(Du~n(tn))𝑑x=ΩvξW(Du~(t))𝑑x

because vnv in Lq(Ω) for every q< (by compact embedding) while

Du~(tn)=Dun+Dg(tn)Du~(t)=Du+Dg(t)in Lr(Ω,2×2) for r=pp¯

and thus W(Du~n(tn))W(Du~(t)) in Lr2 for r2>1.

By (2.1) for every ξΞ,

|v-(tn,un,vn)|L2-v(tn,un,vn)[ξ]

and hence by (2.2) for every ξΞ we get

lim infn+|v-(tn,un,vn)|L2-limn+v(tn,un,vn)[ξ]=-v(t,u,v)[ξ].

Taking the supremum with respect to ξΞ concludes the proof. ∎

By the regularity in time of g the partial time derivative takes the form

(2.3)t(t,u,v)=Ω(v2+η)𝜺(u~(t)):𝝈(g˙(t))dx.

In particular, for v𝒱 by continuity and coercivity of the elastic energy we have

(2.4)|t(t,u,v)|C𝜺(u~(t))L2C(𝜺(u~(t))L22+1)C′′((t,u,v)+1).

Lemma 3.

If tnt, unu in U and vnv in V, then

limn+t(tn,un,vn)=t(t,u,v).

Proof.

It is sufficient to pass to the limit in equation (2.3) using the fact that Dg˙(tn)(vn2+η)Dg˙(t)(v2+η) in L2(Ω,2×2). ∎

Finally, the energy (t,,v) is Fréchet differentiable in H01(Ω,2), with respect to the natural norm of H01(Ω,2), and for every ζH01(Ω;2) we have

u(t,u,v)[ζ]=Ω(v2+η)𝝈(u):𝜺(ζ)dx=-div((v2+η)𝝈(u)),ζ,

where , denotes the duality between H01(Ω;2) and H-1(Ω;2).

3 Gradient flows

3.1 Incremental problems

Our gradient flow will be defined as an “alternate minimizing movement”, i.e. as the limit of an alternate implicit Euler discretization.

Fix τm=Tm>0 (for some m with m>0) and for k=0,,m consider the discrete times

tm,k=kτm[0,T].

Given um,k-1=u(tm,k-1) and vm,k-1=v(tm,k-1), the irreversible alternate minimizing movement is defined by

(3.1){vm,kargmin{(tm,k,um,k-1,v)+12τmv-vm,k-1L22:vvm,k-1,v𝒱},um,kargmin{(tm,k,u,vm,k):u𝒰}=argmin{(tm,k,u,vm,k):u𝒰}.

By Lemma 5 we get the regularity and the continuous dependence of um,k stated in the next lemma.

Lemma 1.

There exist 2<p~<p¯ and p(2,p¯), independent of τm and k, such that um,kW1,p(Ω,R2). Moreover, there exists C>0, independent of τm and k, such that

um,k-um,k-1W1,pC(|tm,k-tm,k-1|+vm,k-vm,k-1Lq)

for 1q=1p-1p~. Finally, um,k is bounded in W1,p(Ω,R2) uniformly with respect to τm and k.

The next two lemmas provide the main ingredients in the proof of the convergence Theorem 6.

Lemma 2.

For every k1 let v˙m,k=(vm,k-vm,k-1)/(tm,k-tm,k-1). Then

(3.2)v˙m,k,ξL2+v(tm,k,um,k-1,vm,k)[ξ]0for every ξΞ,
(3.3)v˙m,kL22=|v-(tm,k,um,k-1,vm,k)|L2v˙m,kL2=-v(tm,k,um,k-1,vm,k)[v˙m,k].

Proof.

First of all, by a standard truncation argument we know that we can replace 𝒱 with the whole H1(Ω) in (3.1), hence

vm,kargmin{(tm,k,um,k-1,v)+12τmv-vm,k-1L22:vvm,k-1,vH1(Ω)}.

By minimality, vm,k solves the variational inequality

(3.4)v(tm,k,um,k-1,vm,k)[w-vm,k]+v˙m,k,w-vm,kL20

for every wH1(Ω) with wvm,k-1. Inequality (3.2) follows. Choosing w-vm,k=±τmv˙m,k (corresponding to w=vm,k-1 and w=2vm,k-vm,k-1) provides

(3.5)v(tm,k,um,k-1,vm,k)[v˙m,k]+v˙m,kL22=0.

Next, by (2.1) and (3.4) with w-vm,k=ξΞ we get

|v-(tm,k,um,k-1,vm,k)|L2=sup{-v(tm,k,um,k-1,vm,k)[ξ]:ξΞ}
(3.6)sup{v˙m,k,ξL2:ξΞ}=v˙m,kL2.

If v˙m,k=0, there is nothing else to prove. Otherwise, ξ=v˙m,k/v˙m,kL2 is an admissible variation and by (3.5) the inequality in (3.6) becomes an equality, thus |v(tm,k,um,k-1,vm,k)|L2=v˙m,kL2. ∎

Lemma 3.

For every k1 the following energy estimate holds:

(tm,k,um,k,vm,k)(tm,k-1,um,k-1,vm,k-1)+tm,k-1tm,kt(t,um,k-1,vm,k-1)𝑑t
(3.7)-12tm,k-1tm,kv˙m,kL22+|v-(tm,k,um,k-1,vm,k)|L22dt.

Proof.

By the minimality of um,k and the convexity of (tm,k,um,k-1,) we get

(tm,k,um,k,vm,k)(tm,k,um,k-1,vm,k)
(3.8)(tm,k,um,k-1,vm,k-1)-v(tm,k,um,k-1,vm,k)[vm,k-1-vm,k].

By Lemma 2,

v(tm,k,um,k-1,vm,k)[vm,k-1-vm,k]=τmv˙m,kL22=τm12(v˙m,kL22+|v-(tm,k,um,k-1,vm,k)|L22)

and hence by (3.8),

(tm,k,um,k,vm,k)(tm,k,um,k-1,vm,k-1)-12tm,k-1tm,kv˙m,kL22+|v-(tm,k,um,k-1,vm,k)|L22dt.

Finally,

(tm,k,um,k-1,vm,k-1)=(tm,k-1,um,k-1,vm,k-1)+tm,k-1tm,kt(t,um,k-1,vm,k-1)𝑑t

and the proof is concluded. ∎

Lemma 4.

There exists C>0, independent of τm and k, such that

(tm,k,um,k,vm,k)C((t0,u0,v0)+1).

Proof.

By the minimality of um,k and vm,k ,

(tm,k,um,k,vm,k)(tm,k,um,k-1,vm,k)
(tm,k,um,k-1,vm,k)+12τmvm,k-vm,k-1L22(tm,k,um,k-1,vm,k-1).

Further,

(tm,k,um,k-1,vm,k-1)=(tm,k-1,um,k-1,vm,k-1)+tm,k-1tm,kt(t,um,k-1,vm,k-1)𝑑t

and

t(t,um,k-1,vm,k-1)=Ω(vm,k-12+η)𝜺(um,k-1+g(t)):𝝈(g˙(t))dx.

Clearly vm,k-12+1L(1+η) and 𝝈(g˙)L2C. Moreover, by (2.4) and by the Lipschitz continuity of g() we have

𝜺(um,k-1+g(t))L2𝜺(um,k-1+gm,k-1)L2+𝜺(g(t)-gm,k-1)L2
C((tm,k-1,um,k-1,vm,k-1)+1).

In summary, we can write

(tm,k,um,k,vm,k)(tm,k-1,um,k-1,vm,k-1)+Cτm((tm,k-1,um,k-1,vm,k-1)+1)

and then

((tm,k,um,k,vm,k)+1)(1+Cτm)((tm,k-1,um,k-1,vm,k-1)+1).

It follows that (tm,k,um,k,vm,k)(1+Cτm)k((t0,u0,v0)+1) and then, since τmTk,

(tm,k,um,k,vm,k)(1+CTk)k((t0,u0,v0)+1),

since (1+CTk)keCT the required estimate follows. ∎

3.2 Compactness and convergence

Let us denote by um:[0,T]𝒰 and vm:[0,T]𝒱 the evolutions obtained by piecewise affine interpolation of um,k=um(tm,k) and vm,k=vm(tm,k).

Lemma 5.

The sequence vm is bounded in L(0,T;H1(Ω)) and in H1(0,T;L2(Ω)) and thus, upon extracting a (non-relabeled) subsequence, vm*v in L(0,T;H1(Ω)) and vmv in H1(0,T;L2(Ω)).

As a consequence, if tmt, then vm(tm)v(t) in H1(Ω) and um(tm)u(t) in W1,p(Ω,R2), for some p>2, where u(t)argmin{E(t,u,v(t)):uU}.

Proof.

From Lemma 4 we know that (tm,k,um,k,vm.k) is uniformly bounded and thus vm is bounded in L(0,T;H1(Ω)) while um is bounded in L(0,T;H01(Ω,2)) by Korn’s inequality. Then from the energy balance (3.7), together with identity (3.3), and from the uniform bound on the time derivative (2.4) we get

(tm,k,um,k,vm,k)(tm,k-1,um,k-1,vm,k-1)-tm,k-1tm,kv˙m,kL22𝑑t+Cτm.

By induction we get

0Tv˙m,kL22𝑑t(0,u0,v0)+CT

and thus vm is bounded in H1(0,T;L2(Ω)). As a consequence, (up to subsequences) vmv in H1(0,T;L2(Ω)) and vm(tm)v(t) in L2(Ω) for tmt; since vm(tm) is bounded in H1(Ω) it turns out that vm(tm)v(t) in H1(Ω).

Let km such that tkmt<tkm+1. Being um(tm,km)=um,kmargmin{(tm,km,u,vm,km):u𝒰} we have

Ω(vm,km2+1)Du~m,kn:𝐂Dϕdx=0for all ϕ𝒰.

Since um,km is bounded in H01(Ω,2), there exists a subsequence (non-relabeled) weakly converging to some uH01(Ω,2). Clearly

Du~m,kn=Dum,km+Dg(tm,km)Du+Dg(t)=Du~(t)and(vm,km2+1)Dϕ(v2(t)+1)Dϕ

in L2(Ω,2×2), thus

Ω(v2(t)+1)Du~(t):𝐂Dϕdx=0for all ϕ𝒰

and u=u(t)argmin{(t,u,v(t)):u𝒰}. Since the limit is uniquely determined, the whole sequence converges. We can argue exactly in the same way for um(tm,km+1).

By compact embedding vm,kmv(t) in Lq(Ω) for every q<+. Then invoking Lemma 5, we have

u(t)-um,kmW1,pCg(t)-g(tm,km)Lq+Cv(t)-vm,kmLq

and similarly for um,km+1. As um(tm) is the affine interpolation of um,km and um,km+1, the strong convergence of displacements in W1,p(Ω,2) follows. ∎

Theorem 6.

Let v be a limit of vm (as in Lemma 5) and let u be the corresponding limit of um. Then vH1(0,T;L2(Ω))L(0,T;V) and for every t[0,T] it holds u(t)argmin={E(t,u,v(t)):uU} and

(t,u(t),v(t))=(0,u0,v0)-120tv˙(r)L22+|v-(r,u(r),v(r))|L22dr+0tt(r,u(r),v(r))𝑑r.

Moreover, for almost every t[0,T] we have

(3.9)v˙(t)L2=|v-(t,u(t),v(t))|L2.

For the sake of clarity the proof of the previous theorem will be split into a couple of propositions.

Proposition 7.

Under the hypotheses of Theorem 6 for every t[0,T] it holds

(3.10)(t,u(t),v(t))(0,u0,v0)-120tv˙(r)L22+|v-(r,u(r),v(r))|L22dr+0tt(r,u(r),v(r))𝑑r.

Proof.

Given t[0,T], let 1kmm such that tm,kmt. Then by induction (3.7) provides

(tm,km,u(tm,km),v(tm,km))+120tm,kmv˙m(r)L22𝑑r+12k=0km-1tm,ktm,k+1|v-(tm,k,um,k-1,vm,k)|L22𝑑r
(3.11)(0,u0,v0)+k=0km-1tm,ktm,k+1t(r,um,k-1,vm,k-1)𝑑r.

By assumption tm,kmt, then by Lemma 5 we have vm(tm,km)v(t) in H1(Ω) and um(tm,km)u(t) in W1,p(Ω,2). Then by Lemma 1 we get

(t,u(t),v(t))lim infm(tm,km,u(tm,km),v(tm,km)).

Since vmv in H1(0,T;L2(Ω)), we have

0tv˙(r)L22𝑑rlim infm0tm,kmv˙m(r)L22𝑑r.

Next, given r(0,t) let r[tm,km,tm,km+1) for kmkm. Clearly, both tm,kmr and tm,km-1r. By Lemma 5 we know that um(tm,km-1)u(r) strongly in W1,p(Ω,2) (for some p>2) while vm(tm,km)v(r) in H1(Ω) and then by Lemma 2 we get the pointwise estimate

|v-(r,u(r),v(r))|L2lim infm+|v-(tm,km,um(tm,km-1),vm(tm,km-1))|L2.

We remark that |v-(,u(),v())|L2 is measurable in the interval (0,T). Indeed, given ξΞ, the functional v(t,u,v)[ξ] is continuous in [0,T]×W1,p(Ω,2)×H1(Ω) and thus tv(t,u(t),v(t))[ξ] is measurable. To conclude, it is enough to employ (2.1) and write |v-(,u(),v())|L2 as the (pointwise) supremum of -v(,u(),v())[ξj] for ξj in a dense, countable subset of Ξ. So, by Fatou’s lemma we conclude that

0t|v-(r,u(r),v(r))|L22𝑑rlim infm+k=0km-1tm,ktm,k+1|v-(tm,k,um,k-1,vm,k)|L22𝑑r.

By Lemma 3 and (2.4) we get, by dominated convergence,

lim supm+k=0km-1tm,k-1tm,kt(r,um,k-1,vm,k-1)𝑑r0tt(r,u(r),v(r))𝑑r.

Taking respectively the liminf on the left-hand side and the limsup on the right-hand side of (3.11), we get the energy inequality

(t,u(t),v(t))+120tv˙(r)L22+|v-(r,u(r),v(r))|L22dr(0,u0,v0)+0tt(r,u(r),v(r))𝑑r,

which conclude the proof. ∎

There are different ways to prove the “upper gradient inequality”

(0,u0,v0)-(t,u(t),v(t))0t|v-(r,u(r),v(r))|L2v˙(r)L2𝑑r-0tt(r,u(r),v(r))𝑑r
120t|v-(r,u(r),v(r))|L22+v˙(r)L22dr-0tt(r,u(r),v(r))𝑑r.

For instance, the estimate will follow from the chain rule Lemma 7 once we will know (from Theorem 1) that the limit evolution v belongs to H1(0,T;H1(Ω)). Actually, Proposition 8 below provides the required inequality for v in H1(0,T;L2(Ω)); its proof, based only on measure theory and separate convexity, is inspired by [15, Theorem 4.12]. In some sense (3.12) corresponds to [4, Corollary 2.4.10].

Proposition 8.

Let vH1(0,T;L2(Ω))L(0,T;V) and let u(t)argmin{F(t,v(t),u):uU} such that t|v-F(t,u(t),v(t))| belongs to L2(0,T). Then for every t(0,T),

(3.12)(0,u0,v0)-(t,u(t),v(t))60t|v-(r,u(r),v(r))|L2v˙(r)L2𝑑r-0tt(r,u(r),v(r))𝑑r.

Proof.

We divide the proof into two steps.

Step I. Since the slope belongs to L2(0,t), there exists a sequence of finite subdivisions tj,i (for j and i=0,,Ij) of the time interval [0,t] with 0=tj,0<<tj,i<tj,i+1<<tj,Ij=t, with

limj+maxi{tj,i+1-tj,i}=0

and such that the piecewise constant functions

Fj(r)=i=0Ij-1χ(tj,i,tj,i+1)(r)|v-(tj,i,u(tj,i),v(tj,i))|L2

converge to |v-(,u(),v())|L2 strongly in L2(0,t) (cf. [15, Theorem 4.12] or [17]).

Denote for simplicity uj,i=u(tj,i) and χj,i=χ(tj,i,tj,i+1) etc. For each j and i=0,,Ij write

(tj,i,uj,i,vj,i)-(tj,i+1,uj,i+1,vj,i+1)=(tj,i,uj,i,vj,i)-(tj,i,uj,i,vj,i+1)
+(tj,i,uj,i,vj,i+1)-(tj,i+1,uj,i,vj,i+1)
+(tj,i+1,uj,i,vj,i+1)-(tj,i+1,uj,i+1,vj,i+1).

We will consider the three lines above separately, starting with the first. By the convexity of (tj,i,uj,i,) we get

(tj,i,uj,i,vj,i)-(tj,i,uj,i,vj,i+1)-v(tj,i,uj,i,vj,i)[vj,i+1-vj,i]
|v-(tj,i,uj,i,vj,i)|L2vj,i+1-vj,iL2
tj,itj,i+1|v-(tj,i,uj,i,vj,i)|L2v˙j,i+1L2𝑑r,

where v˙j,i=(vj,i+1-vj,i)/(tj,i+1-tj,i) denotes the “discrete” velocity. For the second term we will just write

(tj,i,uj,i,vj,i+1)-(tj,i+1,uj,i,vj,i+1)=-tj,itj,i+1t(r,uj,i,vj,i+1)𝑑r.

For the third term, remember that by minimality

Ω(vj,i+12+η)𝝈(uj,i+1+gj,i+1):𝜺(uj,i-uj,i+1)dx=0

and that (tj,i+1,,vj,i+1) is quadratic; then

(tj,i+1,uj,i,vj,i+1)-(tj,i+1,uj,i+1,vj,i+1)=12Ω(vj,i+12+η)(W(Duj,i+Dgj,i+1)-W(Duj,i+1+Dgj,i+1))𝑑x
=12Ω(vj,i+12+η)𝝈(uj,i+uj,i+1+2gj,i+1):𝜺(uj,i-uj,i+1)dx
=12Ω(vj,i+12+η)𝝈(uj,i-uj,i+1):𝜺(uj,i-uj,i+1)dx
Cuj,i+1-uj,iH12=Ctj,itj,i+1uj,i+1-uj,iH12𝑑r.

In conclusion,

(tj,i,uj,i,vj,i)-(tj,i+1,uj,i+1,vj,i+1)tj,itj,i+1|v-(tj,i,uj,i,vj,i)|L2v˙j,i+1L2𝑑r-tj,itj,i+1t(r,uj,i,vj,i+1)𝑑r
+Ctj,itj,i+1uj,i+1-uj,iH12𝑑r.

Taking the sum for i=0,,Ij yields

(0,u0,v0)-(t,u(t),v(t))i=0Ij-1tj,itj,i+1|v-(tj,i,uj,i,vj,i)|L2v˙j,i+1L2𝑑r-i=0Ij-1tj,itj,i+1t(r,uj,i,vj,i+1)𝑑r
(3.13)+i=0Ij-1tj,itj,i+1uj,i+1-uj,iH12𝑑r.

Step II. Let us re-write (3.13) as

(0,u0,v0)-(t,u(t),v(t))0tFj(r)Vj(r)-Pj(r)+Ej(r)dr

in terms of the piecewise constant functions Fj (defined above) and

Vj(r)=i=0Ij-1χj,i(r)v˙j,i+1L2,
Pj(r)=i=0Ij-1χj,i(r)t(r,uj,i,vj,i+1),
Ej(r)=i=0Ij-1χj,i(r)|tj,i+1-tj,i|-1uj,i+1-uj,iH12.

Since the above estimate holds for every subdivision tj,i, it must hold also

(0,u0,v0)-(t,u(t),v(t))limj+0tFj(r)Vj(r)-Pj(r)+Ej(r)dr.

We will show that

(3.14)limj0tFj(r)Vj(r)𝑑r=0t|v-(r,u(r),v(r))|L2v˙(r)L2𝑑r,
(3.15)limj0tPj(r)𝑑r=0tt(r,u(r),v(r))𝑑r,
(3.16)limj0tEj(r)𝑑r=0,

which will prove (3.12).

Since Fj converge strongly in L2(0,t) (by construction) to prove (3.14), it is enough to see that

Vj=i=0Ij-1χj,iv˙j,i+1L2v˙L2weakly in L2(0,t).

Note that Vjv˙ a.e. in [0,t] since vW1,2(0,t;L2). Thus, it is enough to check that Vj is bounded in L2(0,t). Write

v˙j,i+1L22=vj,i+1-vj,itj,i+1-tj,iL22=tj,itj,i+1v˙(r)dr2L2tj,itj,i+1v˙(r)2L2dr,

so that

0tVj2(r)𝑑ri=0Ij-1(tj,i+1-tj,i)tj,itj,i+1v˙(r)L22𝑑r=0tv˙(r)L22𝑑r.

Let us prove (3.15). Fix r(0,t) and let tj,ir<tj,i+1 (with i depending on j). For a.e. r(0,t) we have

t(r,uj,i,vj,i+1)=Ω(vj,i+12+η)𝝈(uj,i+g(r)):𝜺(g˙(r))dr.

Remember that vH1(0,T;L2(Ω))L(0,T;H1(Ω)). Using the arguments of Lemma 5, we get

vj,i+1=v(tj,i+1)v(r)andvj,i=v(tj,i)v(r)

in Lq(Ω) for every q<. Thus uj,i=u(tj,i)u(r) in W1,p(Ω,2) for p>2 (by Lemma 5). As a consequence

Ω(vj,i+12+η)𝝈(uj,i+g(r)):𝜺(g˙(r))drΩ(v2(r)+η)𝝈(u(r)+g(r)):𝜺(g˙(r))dr.

Therefore we have t(r,uj,i,vj,i+1)t(r,u(r),v(r)) a.e. in (0,t). Since vL(0,t;H1) by (2.4), we get that |t(r,uj,i,vj,i+1)| is uniformly bounded and thus (3.15) follows by dominated convergence.

Finally, let us prove (3.16). Since uj,iargmin{(tj,i,,vj,i)} by Lemma 1 we know that

uj,i+1-uj,iH12C|tj,i+1-tj,i|2+Cvj,i+1-vj,iLq2

for some q sufficiently large. Since vj,iLp(Ω) for every p< (by Sobolev embedding), we can apply the interpolation inequality

vj,i+1-vj,iLqvj,i+1-vj,iL2αvj,i+1-vj,iLq¯1-α

with 1q=α2+1-αq¯ (for a suitable q¯ depending on α). Hence, for α=12 we get

|tj,i+1-tj,i|-1vj,i+1-vj,iLq2v˙j,i+1L2vj,i+1-vj,iLq¯.

Then

0tEj(r)𝑑ri=0Ij-1tj,itj,i+1|tj,i+1-tj,i|-1uj,i+1-uj,iH12
i=0Ij-1tj,itj,i+1C|tj,i+1-tj,i|+Cv˙j,i+1L2vj,i+1-vj,iLq¯dr
(3.17)C0t|tj,i+1-tj,i|+Vj(r)Dj(r)dr,

where Vj has been defined before while Dj(r)=iχj,ivj,i+1-vj,iLq¯. We have already seen that Vjv˙L2 weakly in L2(0,t) and that vj,i+1-vj,iLq¯0 a.e. in (0,t). Since vj,i is uniformly bounded in H1(Ω), and thus in Lq¯(Ω), by dominated convergence vj,i+1-vj,iLq~0 in L2(0,t). As Ej0, from (3.17) follows (3.16). ∎

Theorem 9.

Let v,u be the limits obtained by Lemma 5, then vH1(0,T;L2(Ω))L(0,T;V) and for a.e. t[0,T] it holds

(3.18){v˙(t)=-[v(t)W(Du~(t))+Gc(v(t)-1)-GcΔv(t)]+,div(𝝈v(t)(u~(t)))=0.

Note that v(t)W(Du~(t))+Gc(v(t)-1)-GcΔv(t) is a finite Radon measure with positive part in L2(Ω) while 𝛔v(u~)=(v2+η)CDu~ is the phase-field stress (and u~(t) denotes the displacement u(t)+g(t)). In particular, the first equation holds in L2(Ω) and v is monotone non-increasing while the second holds in H-1(Ω,R2).

Proof.

By Lemma 2 we know that v˙m,k,ξ+v(tm,k,um,k-1,vm,k)[ξ]0 for every ξΞ and for every choice of the indices m and k. Besides the piecewise affine function vm we introduce the (auxiliary) piecewise constant functions tm, um and vm given by

tm(t)=tm,k,um(t)=um,k-1,vm(t)=vm,kfor t[tm,k-1,tm,k).

Then, for 0ta<tbT, we can write

tatbv˙m,ξL2𝑑ttatb-v(tm,um,vm)[ξ]dt.

Since v˙mv˙ in L2(0,T;L2(Ω)), we can easily pass to the limit in the first term. By Lemma 5 we know that vmv in H1(Ω) and umu strongly in W1,p(Ω,2) (for some p>2) pointwise in (0,T). Thus by (2.2),

tatbv˙(t),ξL2𝑑ttatb-v(t,u(t),v(t))[ξ]dt,

which holds for every 0ta<tbT. As a consequence v˙(t),ξ-v(t,u(t),v(t))[ξ] holds a.e. in time for every ξΞ.

From Theorem 6 we know that v˙(t)L2=|v-(t,u(t),v(t))|L2 holds a.e. in time; then by Lemma 3 it follows that v(t,u(t),v(t)) is a finite Radon measure μ with positive part μ+L2(Ω) and that

v˙(t)L2=|v-(t,u(t),v(t))|L2=μ+L2.

Moreover, for every ξΞC0(Ω),

-v˙(t),-ξL2v(t,u(t),v(t))[-ξ]=μ+-μ-,-ξ=Ωξ𝑑μ--Ωξμ+𝑑x,

where the duality is in the sense of distributions. Let Ω± be the disjoint supports of μ±. Let φL2(Ω) be a non-negative function supported in Ω+. Arguing as in the proof of Lemma 3, we can find a sequence φnC0(Ω) such that μ-,φn0 and μ+,φnμ+,φ. Hence, for every non-negative φL2(Ω) supported in Ω+ we get

-v˙(t),φL2μ+,φL2.

It follows that -v˙(t)μ+0 in Ω+. The same inequality hold (obviously) in Ω-, where μ+=0. Since -v˙(t)L2=μ+L2, we get -v˙(t)=μ+. To conclude, it is enough to represent the functional v(t,u(t),v(t))[ϕ]. For ϕC0(Ω) we have

v(t,u(t),v(t))[ϕ]=ΩvϕW(Du~(t))𝑑x+GcΩ(v(t)-1)ϕ+vϕdx
(3.19)=vW(Du~(t))+Gc(v(t)-1)-GcΔv(t),ϕ,

where the last duality is in the sense of distributions. Hence

μ+=[vW(Du~(t))+Gc(v(t)-1)-GcΔv(t)]+

and the proof is concluded. ∎

4 Time rescaling

In order to find the quasi-static limit we first rescale the time variable, in order to get a “slow” evolution in the rescaled physical time interval [0,Tε]. For ε>0 let us consider the boundary condition gε(t)=g(εt) defined in [0,Tε], for Tε=Tε. Clearly gε is Lipschitz continuous in W1,p¯(Ω,2) with

gε(t2)-gε(t1)W1,p¯εC|t2-t1|.

Next, we define ε:[0,Tε]×𝒰×𝒱[0,+) by

ε(t,u,v)=12Ω(v2+η)W(Du+Dgε(t))𝑑x+12GcΩ(v-1)2+|v|2dx.

As in Section 3 fix τm=Tεm>0 and let tm,k=kτm for k=0,,m. Given uε,m,k-1 and vε,m,k-1, define by induction

{vε,m,kargmin{ε(tm,k,uε,m,k-1,v)+12τmv-vε,m,k-1L22:vvε,m,k-1,v𝒱},uε,m,kargmin{ε(tm,k,u,vε,m,k):u𝒰}.

Denote by uε,m and vε,m the corresponding piecewise affine interpolate. By Lemma 5, Theorem 6 and Theorem 9 we easily get the following result.

Theorem 1.

There exists a subsequence (not relabelled) of vε,m such that vε,mvε in H1(0,Tε;L2(Ω)). Let uε be the corresponding limit of uε,m. Then, for every t[0,Tε], we have uε(t)argmin{E(t,vε(t),u):uU} and

(4.1)ε(t,uε(t),vε(t))=ε(0,u0,v0)-120tv˙ε(r)L22+|v-ε(r,uε(r),vε(r))|L22dr+t0ttε(r,uε(r),vε(r))𝑑r.

Moreover, for almost every t[0,Tε] we have

(4.2)v˙ε(t)L2=|v-ε(t,uε(t),vε(t))|L2

and

{v˙ε(t)=-[vε(t)W(Du~ε(t))+Gc(vε(t)-1)-GcΔvε(t)]+,div(𝝈vε(t)(u~ε(t)))=0.

Corollary 2.

For every λ[0,1] it holds

ε(t,uε(t),vε(t))=ε(0,u0,v0)+0ttε(r,uε(r),vε(r))𝑑r
(4.3)-0tλv˙ε(r)L22+(1-λ)|vε(r,uε(r),vε(r))|L22dr.

Proof.

It is sufficient to re-write (4.1) taking into account (4.2). ∎

Remark 3.

We remark that in general (4.3) does not provide a characterization of the gradient flow, unless it holds for λ=12. Moreover, introducing the rescaled variable t=(tε)[0,T], the functions vε(t)=vε(tε) and uε(t)=uε(tε) solve the system

{εv˙ε(t)=-[vε(t)W(Du~ε(t))+Gc(vε(t)-1)-GcΔvε(t)]+,div(𝝈vε(t)(u~ε(t)))=0.

5 Quasi-static parametrized limit

In this section we will apply the change of variable

tsε(t)=εt+0tv˙ε(r)L2𝑑r

in order to obtain a parametrization of vε (originally defined for t[0,Tε]) in terms of an arc-length parameter s[0,Sε]. This is a convenient way of writing the quasi-static (rate-independent) evolution and in particular to characterize its behavior in the discontinuity points. Note that here the parametrization is in L2 and not in H1 as in [22].

First of all, let us see that sε maps the physical time interval [0,Tε] onto a reference parametrization interval [0,Sε] with Sε=sε(Tε) uniformly bounded with respect to ε>0.

5.1 Finite length and boundedness

Theorem 1.

The length of the discrete curves vε,m is uniformly bounded in L2, i.e., there exists C>0 (independent of ε and m) such that for τ sufficiently small it holds

0Tεv˙ε,m(t)L2𝑑tC(T+|Ω|).

Moreover,

0Tεv˙ε,m(t)H12𝑑tC(εT+|Ω|).

Proof.

The proof is quite technical and is based on a discrete Gronwall argument, provided in Lemma 1, see also [33, 22, 31, 29]. We will prove this property employing again the time discretization scheme. However, for notational convenience, we will write vk instead of vε,m,k, etc.

Step I. By Lemma 2 for k0 we have

(5.1)vε(tk+1,uk,vk+1)[v˙k+1]+v˙k+1L22=0,

while for k1,

vε(tk,uk-1,vk)[w-vk]+v˙k,w-vkL20for every wH1(Ω) with wvk-1.

Choosing w=vk+v˙k+1 provides

(5.2)vε(tk,uk-1,vk)[v˙k+1]+v˙k,v˙k+1L20.

For k=0 we have, by equilibrium,

(5.3)vε(0,u0,v0)[v˙1]0.

Hence, from (5.1) and (5.2) we obtain, for k1,

(5.4)vε(tk,uk-1,vk)[v˙k+1]-vε(tk+1,uk,vk+1)[v˙k+1]v˙k+1L22-v˙k,v˙k+1L212v˙k+1L22-12v˙kL22.

For k=0 from (5.1) and (5.3) we get

vε(0,u0,v0)[v˙1]-vε(t1,u0,v1)[v˙1]v˙1L22.

Setting v˙0=0, we can also write

vε(0,u0,v0)[v˙1]-vε(t1,u0,v1)[v˙1]12v˙1L22-12v˙0L22

and thus (5.4) actually holds for every k0.

For the left-hand side of (5.4) we proceed as follows:

vε(tk,uk-1,vk)[v˙k+1]-vε(tk+1,uk,vk+1)[v˙k+1]=v(tk,uk-1,vk)[v˙k+1]-v(tk+1,uk,vk+1)[v˙k+1]
+v𝒟(vk)[v˙k+1]-v𝒟(vk+1)[v˙k+1].

For v𝒟(vk)[v˙k+1]-v𝒟(vk+1)[v˙k+1] we write

Ω(vk-1)v˙k+1+vkv˙k+1dx-Ω(vk+1-1)v˙k+1+vk+1v˙k+1dx
=Ω(vk-vk+1)v˙k+1+(vk-vk+1)v˙k+1dx=-τv˙k+1H12.

We can estimate v(tk,uk-1,vk)[v˙k+1]-v(tk+1,uk,vk+1)[v˙k+1] by

Ωvkv˙k+1W(Duk-1+Dgε,k)𝑑x-Ωvk+1v˙k+1W(Duk+Dgε,k+1)𝑑x
Ωvkv˙k+1W(Duk-1+Dgε,k)𝑑x-Ωvk+1v˙k+1W(Duk-1+Dgε,k)𝑑x
   +Ωvk+1v˙k+1W(Duk-1+Dgε,k)𝑑x-Ωvk+1v˙k+1W(Duk+Dgε,k+1)𝑑x
Ω(vk-vk+1)v˙k+1W(Duk-1+Dgε,k)𝑑x+Ωvk+1v˙k+1(W(Duk-1+Dgε,k)-W(Duk+Dgε,k+1))𝑑x
Ωvk+1v˙k+1(W(Duk-1+Dgε,k)-W(Duk+Dgε,k+1))𝑑x,

where last inequality follows from (vk-vk+1)v˙k+1W(Duk+Dgε,k)0. For 1s+2p=1 we get by the Hölder inequality

Ωv˙k+1vk+1(W(Duk-1+Dgε,k)-W(Duk+Dgε,k+1))𝑑xv˙k+1LsW(Duk-1+Dgε,k)-W(Duk+Dgε,k+1)Lp2.

Since the function uk is uniformly bounded in W1,p(Ω,2) for all p with 2<p<p¯ (by Lemma 5) and since gW1,(0,T;W1,p¯(Ω,2)), we get by Lemma 5,

supk(Duk-1+Dgε,k)+(Duk+Dgε,k+1)Lp<+

and

(Duk-1+Dgε,k)-(Duk+Dgε,k+1)LpC(ε|tk-tk-1|+vk-vk-1Lr+ε|tk+1-tk|)Cτ(ε+v˙kLr),

where 1r=1p-1p¯. Thus for q=sr we have

Ωv˙k+1vk+1(W(Duk-1+Dgε,k)-W(Duk+Dgε,k+1))𝑑xCτv˙k+1Lq(ε+v˙kLq).

In conclusion, for k0 we have

(5.5)12v˙k+1L22-12v˙kL22-τGcv˙k+1H12+Cτv˙k+1Lq(ε+v˙kLq).

Step II. By Young’s inequality, for 0<μ1 and Cμ>1,

v˙k+1Lq(ε+v˙kLq)εv˙k+1Lq+Cμv˙k+1Lq2+μv˙kLq2
Cμ(ε2+v˙k+1Lq2)+μv˙kLq2
(5.6)Cμ(ε2+v˙k+1Lq2)+Cμv˙kH12.

Write 1q=α+1-αq¯ for α(0,1) and q<q¯<+. Then, by interpolation and Young’s inequality, with p=1α, for 0<λ1 we have

v˙k+1Lq2v˙k+1L12αv˙k+1Lq¯2(1-α)
Cλv˙k+1L12+λv˙k+1Lq¯2
(5.7)Cλv˙k+1L1v˙k+1L2+Cλv˙k+1H12.

Hence, upon choosing μ and λ sufficiently small, respectively in (5.6) and in (5.7), we infer that for every 0<δ1 there exists Cδ such that

(5.8)Cv˙k+1Lq(ε+v˙kLq)Cδ(ε2+v˙k+1L1v˙k+1L2)+δv˙k+1H12+δv˙kH12.

Joining (5.5) and (5.8) yields the estimate

(5.9)12v˙k+1L22-12v˙kL22-τGcv˙k+1H12+τδv˙k+1H12+τδv˙kH12+τCδv˙k+1L1v˙k+1L2+τCδε2.

Step III. In order to apply the discrete Gronwall Lemma 1, we need to re-write (5.9). First,

12v˙k+1L22-12v˙kL22-τ(Gc-2δ)v˙k+1H12-δτv˙k+1H12+δτv˙kH12+τCδv˙k+1L1v˙k+1L2+τCδε2.

Hence,

(12v˙k+1L22+δτv˙k+1H12)-(12v˙kL22+δτv˙kH12)-τ(Gc-2δ)v˙k+1H12+τCδv˙k+1L1v˙k+1L2+τCδε2.

For τ,δ1 and γ>0 we can write

γ(12v˙k+1L22+δτv˙k+1H12)(Gc-2δ)v˙k+1H12.

Therefore, we get

(12v˙k+1L22+δτv˙k+1H12)-(12v˙kL22+δτv˙kH12)
(5.10)-γτ(12v˙k+1L22+δτv˙k+1H12)+Cδτv˙k+1L1(12v˙k+1L22+δτv˙k+1H12)12+Cδτε2.

Define

ak=(12v˙kL22+δτv˙kH12)12,bk=Cδv˙kL1,ck2=Cδε2.

Hence (5.10) reads: for every 0km-1,

ak+12-ak2-τγak+12+τak+1bk+1+τck+12.

Then, for 0<β<γ2 by Lemma 1 we get

ak(i=0kτe-2β(tk-ti)ci2)12+i=0kτe-β(tk-ti)bi.

Remembering the definition of ak, bk and ck, the previous estimate gives

12v˙kL2(12v˙kL22+δτv˙kH12)12
C(i=0ke-2β(tk-ti)τε2)12+Ci=0kτe-β(tk-ti)v˙iL1
Cε(0tke-2β(tk-r)𝑑r)12+C0tke-β(tk-r)v˙ε,m(r)L1𝑑r
Cε2β+C0tkv˙ε,m(r)L1𝑑rC(ε+|Ω|),

where the last inequality follows from monotonicity (in time) and boundedness of vε,m. Hence vε,m is bounded in W1,(0,Tε;L2). Moreover,

0Tεv˙ε,m(t)𝑑tk=0mτv˙kL2CεTε+Ck=0mτ0tke-β(tk-r)v˙ε,m(r)L1𝑑r.

Then, for t[tk,tk+1] we can write

0tke-β(tk-r)v˙ε,m(r)L1𝑑r0te-β(t-τ-r)v˙ε,m(r)L1𝑑r

and thus

k=0Kτ0tke-β(tk-r)v˙ε,m(r)L1𝑑r0Tε0te-β(t-τ-r)v˙ε,m(r)L1𝑑r𝑑t
eβτ0Tεv˙ε,m(r)L1rTεe-β(t-r)𝑑t𝑑r
C0Tεv˙ε,m(r)L1𝑑r=C|Ω|.

Step IV. Let us go back to (5.9), i.e.

12v˙k+1L22-12v˙kL22-τ(Gc-δ)v˙k+1H12+τδv˙kH12+τCδv˙k+1L1v˙k+1L2+τCδε2.

Let 0<C=Gc-δ for 0<δ1; being v˙kL2 uniformly bounded (by the previous step) the above estimate can be written as

τCv˙k+1H12(12v˙kL22-12v˙k+1L22)+τδv˙kH12+τCδ(ε2t+v˙k+1L1).

Taking the sum for k=0,,m-1 and remembering that vε,m is piecewise affine, we get

C0Tεv˙ε,mH12𝑑t=Ck=0m-1τv˙k+1H12
k=0m-1(12v˙kL22-12v˙k+1L22)+δk=0m-1τv˙kH12+Cδk=0m-1τ(ε2+v˙k+1L1)
12v˙0L22+δ0Tεv˙ε,mH12𝑑t+Cδ0Tεε2+v˙ε,mL1dt
δ0Tεv˙ε,mH12𝑑t+Cδ(εTε+|Ω|).

Choosing 0<δ<C, it follows that vε,m is bounded in H1(0,Tε;H1). ∎

Passing to the limit for τm0 in the previous theorem, we get the following result.

Corollary 2.

The limit evolution vε (provided by Theorem 1) satisfies

0Tεv˙ε(t)L2+v˙ε(t)H12dtC(T+|Ω|).

In particular, Sε=sε(Tε) is uniformly bounded.

5.2 Rescaled parametrized gradient flows

Let us go back to our parametrization

(5.11)tsε(t)=εt+0tv˙ε(r)L2𝑑r

from [0,Tε] onto [0,Sε]. The map tsε(t) is absolutely continuous and strictly monotone. We denote by tε(s) be its inverse; moreover, we denote

(5.12)tε(s)=εtε(s),zε(s)=vεtε(s),wε(s)=uεtε(s).

Accordingly, let w0=u0 and z0=v0.

Lemma 3.

The functions stε(s) and szε(s) are Lipschitz continuous in [0,Sε]; more precisely, for a.e. s[0,Sε] it holds

tε(s)+zε(s)L2=1,tε(s)=εε+|v-(tε(s),wε(s),zε(s))|L2.

Note that tε is onto [0,T].

Proof.

As tsε(t) is absolutely continuous with s˙ε(t)ε a.e. in [0,Tε], the inverse function stε(s) turns out to be Lipschitz continuous with (tε)(s)=1/s˙ε(tε(s)) a.e. in [0,Sε]. Hence, by (5.11) and (5.12),

1=s˙ε(tε(s))(tε)(s)=(ε+v˙ε(tε(s))L2)(tε)(s)=tε(s)+zε(s)L2.

Moreover, by (3.9) for a.e. t[0,Tε] we have

v˙ε(t)L2=|v-ε(t,uε(t),vε(t))|L2=|v-(εt,uε(t),vε(t))|L2.

Thus

v˙ε(tε(s))L2=|v-(εtε(s),uεtε(s),vεtε(s))|L2=|z-(tε(s),wε(s),zε(s))|L2.

Since sets of measure zero are mapped to sets of measure zero, both by stε(s) and by tsε(t), for a.e. s[0,Sε] we have

tε(s)=εs˙ε(tε(s))=εε+v˙ε(tε(s))L2=εε+|v(tε(s),wε(s),zε(s))|L2.

Since tε is the rescaled inverse of sε, it is surjective, taking values in [0,T]. ∎

Lemma 4.

For given ε>0, the rescaled parametrized evolutions (tε,zε) are (uniformly) bounded in both W1,(0,Sε;[0,T]×L2) and L(0,Sε;[0,T]×V) with tε0, zε0 and tε+zεL21. Further, for every s[0,Sε] and every λ[0,1] the following energy balance holds:

(tε(s),wε(s),zε(s))=(0,w0,z0)+0st(tε(r),wε(r),zε(r))tε(r)𝑑r
(5.13)-0sλΨε(zε(r)L2)+(1-λ)Φε(|z-(tε(r),wε(r),zε(r))|L2)dr,

where

Ψε(ξ)={εξ21-ξ,0ξ<1,+,ξ1,Φε(ξ)=ξ2ε+ξ.

We consider both Ψε and Φε to be defined in [0,+). Clearly, wε(s)argmin{E(tε(s),w,zε(s)):wU}

Proof.

By Corollary 2 we know that for every t¯[0,Tε] it holds

ε(t¯,uε(t¯),vε(t¯))=ε(0,u0,v0)+0t¯tε(t,uε(t),vε(t))𝑑t-0t¯λv˙ε(t)L22+(1-λ)|v-ε(t,uε(t),vε(t))|L22dt.

Remember that ε(t,u,v)=(εt,u,v) and thus

tε(t,u,v)=εt(εt,u,v),vε(t,u,v)=v(εt,u,v).

Hence, by the change of variable t=tε(s)=tε(s)ε, the energy balance in parametrized form reads: for a.e. s¯[0,Sε] it holds

(tε(s¯),wε(s¯),zε(s¯))=(0,w0,z0)+0s¯t(tε(s),wε(s),zε(s))tε(s)𝑑s
-0s[λv˙ε(tε(s))L22+(1-λ)|v-(tε(s),wε(s),zε(s))|L22](tε)(s)𝑑s.

Since v˙ε(tε(r))L2(tε)(r)=zε(r)L2 and (tε)(r)=tε(r)ε, it follows by Lemma 3 that

v˙ε(tε(r))L22(tε)(r)=εzε(r)L22tε(r)=εzε(r)L221-zε(r)L2=Ψε(zε(r)L2).

Again by Lemma 3, (tε)(r)=1/(ε+|z-(tε(r),wε(r),zε(r))|L2). ∎

Since Sε is uniformly bounded, by Corollary 2, we have S=lim infεSε<+. For compactness, it will be convenient to consider parametrized evolutions tε and zε to be defined in [0,S] with a constant extension in (Sε,S] (clearly only in the case Sε<S). In this way all the (possibly extended) evolutions enjoy the compactness properties of the previous lemma in the parametrization interval [0,S]. Note however that, with this simple extension, the energy balance is not true, in general, for s(Sε,S]. Note that, independently of ε>0, we have tε(0)=0 and tε(S)=T. Using Lemma 3 and Lemma 5, it is now immediate to prove the following compactness property.

Corollary 5.

For εn0 there exists a subsequence (not relabeled) such that

(tεn,zεn)*(t,z)in W1,(0,S;[0,T]×L2).

For a.e. s[0,S] we have

zεn(s)z(s)in H1

and thus wεn(s)w(s) in W1,p (for p>2), where w(s)argmin{E(t(s),w,z(s)):wU}. Finally, t(0)=0 and t(S)=T and thus t maps [0,S] onto [0,T].

5.3 Quasi-static parametrized BV-limit

Theorem 6.

Every limit evolution obtained by Corollary 5 satisfies z0, t0 and t+zL21, t(0)=0 and t(S)=T. Moreover, for every s[0,S) we have w(s)argmin{E(t(s),w,z(s)):wU} and the following energy balance:

(t(s),w(s),z(s))=(0,w0,z0)+0st(t(r),w(r),z(r))t(r)𝑑r
(5.14)-0s|z-(t(r),w(r),z(r))|L2𝑑r.

Any such limit will be called a parametrized BV-evolution.

Proof.

We divide the proof into two parts.

Part I. The proof follows closely that of [32, Theorem 4.4]. If s<S, then s[0,Sεn) for εn1; thus (5.13), with λ=0, provides

(tεn(s),wεn(s),zεn(s))=(0,w0,z0)+0st(tεn(r),wεn(r),zεn(t))tεn(r)𝑑r
(5.15)-0sΦεn(|z-(tεn(r),wεn(r),zεn(r))|L2)𝑑r.

By Corollary 5 we know that tεn(s)t(s), zεn(s)z(s) in H1 and wεn(s)w(s) in W1,p (for p>2). As a consequence, by Lemma 2

(5.16)(t(s),w(s),z(s))lim infεn 0(tεn(s),wεn(s),zεn(s)).

Next, taking lim supεn0 in (5.15), we get

lim supεn 0(tεn(s),wεn(s),zεn(s))(0,w0,z0)+lim supεn 00st(tεn(r),wεn(r),zεn(r))tεn(r)𝑑r
(5.17)-lim infεn 00sΦεn(|z(tεn(r),wεn(r),zεn(r))|L2)𝑑r.

First, let us see that

(5.18)limεn 00st(tεn(r),wεn(r),zεn(r))tεn(r)𝑑r=0st(t(r),w(r),z(r))t(r)𝑑r.

By Lemma 3 we know that t(tεn(r),wεn(r),zεn(t))t(t(r),w(r),z(r)) for a.e. r[0,s]. Moreover, t(tεn(r),wεn(r),zεn(r)) is uniformly bounded since

|t(tεn(r),wεn(r),zεn(r))|C((tεn(r),wεn(r),zεn(r))+1)<C¯.

Hence t(tεn(),wεn(),zεn()) converge to t(t(),w(),z()) strongly in L1(0,s) (by dominated convergence). Since tεn*t in L(0,s), we get (5.18).

Finally, let us show that

(5.19)0s|z-(t(r),w(r),z(r))|L2𝑑rlim infεn00sΦεn(|z-(tεn(r),wεn(r),zεn(r))|L2)𝑑r.

It is not difficult to check that Φε(ξ)ξ-ε for every ξ[0,+). Thus we can write

0sΦεn(|z-(tεn(r),wεn(r),zεn(r))|L2)𝑑r0s|z-(tεn(r),wεn(r),zεn(r))|L2𝑑r-εns.

By Lemma 2,

|z-(t(r),w(r),z(r))|L2lim infεn0|v-(tεn(r),wεn(r),zεn(r))|L2

and thus (5.19) follows from Fatou’s lemma. Joining (5.16)–(5.19) yields

(t(s),u(s),v(s))(0,w0,z0)-0s|z-(t(r),w(r),z(r))|L2𝑑r+0st(t(r),w(r),z(r))t(r)𝑑r.

Part II. To prove the opposite inequality, we employ the “upper gradient inequality” as in Proposition 8. In this setting, tW1,(0,s), zW1,(0,s;L2(Ω))L(0,s;𝒱), w(r)argmin{(t(r),w,z(r)):w𝒰} and r|z-(t(r),w(r),z(r))|L2 belongs to L1(0,s). Then following step by step the proof of Proposition 8 it is not difficult (but lengthy) to check that

(0,w0,z0)-(t(s),w(s),z(s))0s|z-(t(r),w(r),z(r))|L2z(r)L2𝑑r
(5.20)-0st(t(r),w(r),z(r))t(r)𝑑r.

Since z(r)L21, we get

(5.21)(0,w0,z0)-(t(s),w(s),z(s))0s|z-(t(r),w(r),z(r))|L2𝑑r-0st(t(r),w(r),z(r))t(r)𝑑r,

which concludes the proof. ∎

Corollary 7.

If sεns, then F(tεn(sεn),wεn(sεn),zεn(sεn))F(t(s),w(s),z(s)) and zεn(sεn)z(s) in H1(Ω). Moreover, sz(s) is continuous from (0,S) to H1(Ω).

Proof.

Following the proof of Theorem 6 it is easy to check, using (5.17)–(5.21), that

(t(s),w(s),z(s))lim infn+(tεn(sεn),wεn(sεn),zεn(sεn))
lim supn+(tεn(sεn),wεn(sεn),zεn(sεn))
(t(s),w(s),z(s)).

Thus limn+(tεn(sεn),wεn(sεn),zεn(sεn))=(t(s),w(s),z(s)).

If sεs, then by compactness (cf. Corollary 5) zεn(sεn) converge to z(s) weakly in H1 and thus, by compact embedding, strongly in Lq for every q<+. Since tεn*t, we get tεn(sεn)t(s). Then by Lemma 5 we have wε(sε)w(s) in W1,p for some p>2. Hence

Ω(zεn2(sεn)+η)W(Du~εn(sεn))𝑑xΩ(z2(s)+η)W(Du~(s))𝑑x,
Ω(zεn(sεn)-1)2𝑑xΩ(z(s)-1)2𝑑x.

By convergence of the energy it follows that

Ω|zεn(sεn)|2𝑑xΩ|z(s)|2𝑑x,

from which follows the strong convergence in H1. Since zεn(sεn)z(s) in H1 for every sequence sεns, we get that zεnz (strongly in H1) locally uniformly in (0,S).

Remember that, by Corollary 2, the evolution vε is bounded in W1,(0,Tε;L2)H1(0,Tε;H1) and thus it is continuous in H1. As a consequence szε(s)=vεtε(s) is continuous from [0,Sε] to H1. Since zε converge to z locally uniformly, its limit z is continuous as well. ∎

Remark 8.

Using the Legendre transform, it is possible to write (5.14) “em in gradient flow fashion”. Let

Ψ~(z)={0,z1,+,z>1,Φ~(z)={+,z<0,z,z0.

Note that Φ~(z)=Ψ~*(z). With this notation (5.14) reads

(t(s),w(s),z(s))=(0,w0,z0)+0st(t(r),w(r),z(r))t(r)𝑑r
-0sΨ~(z(r)L2)+Ψ~*(|v-(t(r),w(r),z(r))|L2)dr.

5.4 From energy balance to PDEs

In this last subsection we provide some properties, in terms of PDEs, of the parametrized evolution characterized by Theorem 6. Intuitively such an evolution is an “arc-length” parametrization of a BV-evolution [27, 30].

Remember that quasi-static evolutions for non-convex energies may have discontinuity in time and that characterization of these points makes the difference between different notion of quasi-static evolution, e.g. energetic, BV or local [27, 30]. Remember also that discontinuity points td (in time) correspond in the parametric picture to intervals (s,s) with t(s)=td, z(s)=z-(td) and z(s)=z+(td). “Vice versa” if t(sc)>0, then tc=t(sc) is a continuity point in time.

Most of the information are provided by the relationship between the derivative t(s) and the slope |v-(t(s),w(s),z(s))|L2, which is the subject of Proposition 9 ; its PDEs form is provided in Corollary 10. First, in order to employ the chain rule, we prove the following lemma.

Proposition 9.

Let (t,w,z) be a parametrized evolution (provided by Theorem 6). Then for a.e. s[0,S] the following holds:

  1. w(t(s),w(s),z(s))=w(t(s),w(s))=0,

  2. if t(s)>0, then |z-(t(s),w(s),z(s))|L2=0,

  3. if |z-(t(s),w(s),z(s))|L20, then t(s)=0 and

    z(s)argmin{z(t(s),w(s),z(s))[ξ]:ξΞ,ξL21};

    in particular, z(s)L2=1.

Proof.

Equilibrium for the displacement field follows from the minimality of w(s).

Since z(s)L21 for a.e. s[0,S] by (5.14) and (5.20), we can write

(t(s),w(s),z(s))=(0,w0,z0)-0s|z-(t(r),w(r),z(r))|L2𝑑r+0st(t(r),w(r),z(r))t(r)𝑑r
(0,w0,z0)-0s|v-(t(r),w(r),z(r))|L2z(r)L2𝑑r
+0st(t(r),u(r),v(r))t(r)𝑑r(t(s),u(s),v(s)).

Hence all inequalities becomes equalities and hold in every subinterval of (0,S). In particular, for a.e. s(0,S) we have

|z-(t(s),w(s),z(s))|L2(1-z(s)L2)=0.

Hence, if t(s)>0, then z(s)L2<1 (simply because t(s)+z(s)L21) and thus

|z-(t(s),w(s),z(s))|L2=0.

On the contrary, if |z-(t(s),w(s),z(s))|L20, then z(s)L2=1 and t(s)=0.

Let s¯[0,S] such that |v-(t(s¯),w(s¯),z(s¯))|L20, let us show that

z(s¯)argmin{v(t(s¯),w(s¯),z(s¯))[ξ]:ξΞ,ξL21}.

In order to apply the chain rule (A.4), we will show first that zW1,2(s1,s2;H1) for s¯(s1,s2). By the lower semi-continuity of the slope (cf. Lemma 2) for δ1 it holds |v-(t,w,z)|L2C>0 for

(t,z)Iδ×Bδ={|t-t(s¯)|δ}×{z-z(s¯)H1δ}

and wargmin{(t,,z)}. Since s(t(s),z(s)) is continuous in [0,T]×H1 and since tε and zε converge locally uniformly (cf. Corollaries 5 and 7), there exists s1<s2 such that both (t(s),z(s))Iδ×Bδ and (tε(s),zε(s))Iδ×Bδ for s[s1,s2]. Thus,

|v-(tε(s),wε(s),zε(s))|L2C>0for s[s1,s2].

In other terms, let t1ε=tε(s1) and t2ε=tε(s2). Then we have sε(t)[s1,s2] for t[t1ε,t2ε] and

|v-ε(t,uε(t),vε(t))|L2C>0for t[t1ε,t2ε].

Remember that zε(s)=vεtε(s) and that (tε)(s)=1/s˙ε(tε(s)) (being tε the inverse of sε); then, by the change of variable s=sε(t) we get

s1s2zε(s)H12𝑑s=s1s2v˙ε(tε(s))H12|(tε)(s)|2𝑑s
=t1εt2εv˙ε(t)H12s˙ε(t)𝑑t
=t1εt2εv˙ε(t)H12ε+v˙ε(t)L2𝑑t
=t1εt2εv˙ε(t)H12ε+|v-ε(t,uε(t),vε(t)|L2𝑑r
1ε+Ct1εt2εv˙ε(t)H12+,

where the last bound follows from Corollary 2. Thus, zε and its limit z belong to W1,2(s1,s2;H1).

Hence, by the chain rule

(t(s),w(s),z(s))=z(t(s),w(s),z(s))[z(s)]+t(t(s),w(s),z(s))t(s)

for a.e. in s(s1,s2).

On the other hand, by Theorem 6 for a.e. s(s1,s2) it holds

(t(s),w(s),z(s))=-|v(t(s),w(s),z(s)|L2+t(t(s),w(s),z(s))t(s).

Hence,

z(t(s),w(s),z(s))[z(s)]=-|v(t(s),w(s),z(s))|L2.

Therefore z(s)argmin{v(t(s),w(s),z(s))[ξ]:ξΞ,ξL21}. ∎

Corollary 10.

Let (t,w,z) be a parametrized evolution (provided by Theorem 6). Then for a.e. s[0,S] we have the following:

  1. if t(s)>0, then

    (5.22){[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+=0,div(𝝈z(s)(w~(s)))=0,

  2. if t(s)=td in (s,s), then

    (5.23){λ(s)z(s)=-[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+,div(𝝈z(s)(w~(s)))=0,

    where λ(s)=[z(s)W(Dw~(s))+Gc(z(s)-1)-GcΔz(s)]+L2.

Remember that the first case corresponds to a continuity point in time, the second describes instead the “instantaneous evolution” in the discontinuity point td. As in (3.18) the first equation in both the previous systems holds in L2(Ω) while the second holds in H-1(Ω,R2).

Proof.

If t(s)>0, then by Proposition 9 we have |z-(t(s),w(s),z(s))|L2=0, i.e.

z(t(s),w(s),z(s))[ξ]0for every ξH1 with ξ0.

In other terms, z(t(s),w(s),z(s)) is a negative Radon measure or, equivalently, a Radon measure μ with μ+=0. As in (3.19), writing v(t(s),w(s),z(s)) in the sense of distributions yields (5.22).

By Proposition 9 we know that z(s)argmin{z(t(s),w(s),z(s))[ξ]:ξΞ,ξL21} and thus by Lemma 4 we get (5.23). ∎


Communicated by Frank Duzaar


Award Identifier / Grant number: 290888 QuaDynEvoPro

Funding statement: Financial support was provided by INdAM-GNAMPA project “Flussi gradiente ed evoluzioni rate-independent: sviluppi dell’approccio variazionale ed applicazioni” and by the ERC Advanced Grant no. 290888 QuaDynEvoPro.

A Some lemmas

A.1 Discrete Gronwall

First of all let us provide the Gronwall estimate to be used in the proof of Theorem 1. Its proof originates from [33] and [22].

Lemma 1.

Let γ>0, ak,bk,ck0 and a0=0 such that

(A.1)ak+12-ak2-τγak+12+τak+1bk+1+τck+12for k.

Denote tk=kτ for kN. Then for 0<β<γ2 and τ1 it holds

ak(i=0kτe-2β(tk-ti)ci2)12+i=0kτe-β(tk-ti)bifor kN.

Proof.

For λ=(1+τγ)12 and for τ<1 let us re-write (A.1) as λ2ak+12-ak2-ak+1bk+1ck+12. Denote

Ak=λ-k(Ck+Bk),Ck=(i=0kλ2ici2)12,Bk=i=0kλibi.

Let us show that Ak satisfies

(A.2)λ2Ak+12-Ak2-Ak+1bk+1ck+12.

In terms of Ck and Bk, the left-hand side reads

λ-2(k+1)+2(Ck+12+Bk+12+2Ck+1Bk+1)-λ-2k(Ck2+Bk2+2CkBk)-λ-(k+1)(Ck+1+Bk+1)bk+1.

Let us see that (A.2) holds. First, since λ>1,

λ-2kCk+12-λ-2kCK2=λ-2k(Ck+12-Ck2)λ-2k+2(k+1)ck+12ck+12.

Next,

λ-2kBk+12-λ-2kBk2-λ-(k+1)Bk+1bk+1=λ-2k(Bk+λk+1bk+1)2-λ-2kBk2-λ-(k+1)(Bk+λk+1bk+1)bk+1
=(λ-2k+2(k+1)-1)bk+12+(2λ-2k+(k+1)-λ-(k+1))Bkbk+10,

where the last inequality follows again from λ>1. Finally,

2λ-2kCk+1(Bk+λk+1bk+1)-2λ-2kCkBk-λ-(k+1)Ck+1bk+1
=2λ-2k(Ck+1-Ck)Bk+(2λ-2k+(k+1)-λ-(k+1))Ck+1bk+10,

again because λ>1.

Since λ2ak+12-ak2-ak+1bk+1ck+12 and ak+10, we get

ak+112λ2(bk+1+bk+12+4λ2(ak2+ck+12)).

In the same way

Ak+112λ2(bk+1+bk+12+4λ2(Ak2+ck+12)).

Hence by induction, akAk for every k, i.e.

ak(i=0kτλ2(i-k)ci2)12+i=0kτλi-kbi.

Finally, it is not hard to check that for 0<β<γ2 and 0<τ1 it holds

λ-1=(1+τγ)-121-βτ.

Hence, for tk=kτ we have

λ(i-k)=λ-(k-i)(1-βτ)(k-i)=e(k-i)ln(1-βτ)e-β(k-i)τ=e-β(tk-ti).

Then

ak(i=0kτe-2β(tk-ti)ci2)12+i=0kτe-β(tk-ti)bi,

which concludes the proof. ∎

A.2 Representation of linear functionals

We provide here a couple of representations, to be used in Theorem 9 and in Corollary 10. The first, related to unilateral gradient flows, is already stated, without proof, in [16]. The second follows from [13, Lemma 4.4]. In the next lemmas we assume that Ω is an open, bounded set in n.

Lemma 2.

Let ζH-1(Ω). If

(A.3)sup{ζ,ξ:ξH01(Ω),ξ0,ξL21}<+,

then ζ is a (locally finite) Radon measure whose positive part belongs to L2(Ω).

Proof.

We introduce the indicator functions IB,I+:H01[0,+] given by

IB(ξ)={0,ξL21,+,otherwise,I+(ξ)={0,ξ0,+,otherwise.

By (A.3),

supξH01ζ,ξ-(I+(ξ)+IB(ξ))<+.

In other terms, ζ belongs to the proper domain of the Legendre transform (I++IB)* in H-1. In order to characterize the proper domain, let us write by inf-convolution, e.g. [8, Section 15.1],

(I++IB)*(ζ)=minφH-1I+*(φ)+IB*(ζ-φ).

Clearly, if (I++IB)*(ζ)<+, there exists μH-1 such that I+*(μ)+IB*(ζ-μ)<+ and hence I+*(μ)<+ and IB*(ζ-μ)<+. Choosing ξ=λξ^, for λ0 and ξ^0, yields

λμ,ξ^sup{μ,ξ:ξH01,ξ0}=I+*(μ)<+for every λ0,

thus μ,ξ^0 for every ξ^0 in H01. By the Riesz–Markov Theorem it follows that μ is a negative Radon measure. Further, since

IB*(ζ-μ)=sup{ζ-μ,ξ:ξH01,ξL21}<+,

the functional ζ-μ can be extended from H01 to the whole L2 (by the Hahn–Banach Theorem) and thus it can be represented as an element fL2 (by Riesz’s Representation Theorem). In summary, we write ζ=μ+f, where μ is a negative Radon measure, f is an L2-function and is the Lebesgue measure. Write μ=μac+μs where μac and μs are, respectively, absolutely continuous and singular with respect to . Then μac=-m (by the Radon–Nikodym Theorem), where mL1 and m0. Hence

ζ+=(f-m)++μs+=(f-m)+=(f-m)|A,

where A={f-m0}. In A we have fm0 and thus mL2(A). It follows that ζ+L2. ∎

Lemma 3.

Let ζ(H1(Ω))*. If

sup{ζ,ξ:ξH1(Ω),ξ0,ξL21}<+,

then ζ is a finite Radon measure whose positive part belongs to L2(Ω). Moreover,

sup{ζ,ξ:ξH1(Ω),ξ0,ξL21}=ζ+L2.

Corollary 4.

Let ζ(H1(Ω))* and let

ξMargmax{ζ,ξ:ξH1(Ω),ξ0,ξL21}.

Then ζ is a finite Radon measure and ζ+=ξMζ+L2. In particular, the positive part ζ+ belongs to H1(Ω).

Proof.

If ζ=0, there is nothing to prove because the identity ζ+=ξMζ+L2 becomes trivial. Otherwise, since ξ0, we have by the previous lemma and by density of smooth functions that

ζ+L2=ζ,ξM=sup{ζ,ξ:ξH1(Ω),ξ0,ξL21}
=sup{ζ,ξ:ξC,ξ0,ξL21}
sup{ζ+,ξ:ξL2,ξ0,ξL21}=ζ+L2.

It is now enough to note that ξ=ζ+/ζ+L2 is the unique maximizer in L2(Ω). ∎

A.3 Continuous dependence and differentiability

Finally, we collect, for the readers convenience, a few results from [22] adapted to our notation and framework; the first follows from [22, Lemma 2.2] (which in turn is based on a general regularity result proved in [19, Theorem 1.1]), the second from [22, Lemma 2.4] while the last from [22, Lemma 2.3].

Lemma 5.

Let gC1([0,T];W1,p¯(Ω,R2)) for p¯>2. For t[0,T] and vV denote

u(t,v)=argmin{(t,,v):u𝒰}.

There exist C>0 and 2<p~<p¯ such that for every 2p<p~, every t1,t2[0,T] and every v1,v2V it holds

u(t2,v2)-u(t1,v1)W1,pC(g(t2)-g(t1)W1,p+gL(0,T;W1,p)v2-v1Lq),

where 1q=1p-1p~. We remark that C>0 depends only on the linear elastic density W, on η>0 and on Ω; in particular, it is independent of the boundary condition.

Lemma 6.

If uW1,p(Ω,R2) for some p>2, then F(t,u,) is Gateaux differentiable (with respect to H1(Ω)) and

v(t,u,v)[ξ]=2ΩvξW(Du+Dg(t))𝑑x+GcΩ(v-1)ξ+vξdxfor all ξH1(Ω).

Note that the above integrals make sense thanks to the fact that, for ΩR2, ξLq for any 1q<+ while, by assumption, W(Du+Dg(t))Lp for some p>1.

Lemma 7.

If vW1,2(0,T;H1) and u(t)argmin{F(t,u,v(t)):uU}, then the energy tF(t,u(t),v(t)) is a.e. differentiable in (0,T) and the following chain rule holds:

(A.4)˙(t,u(t),v(t))=t(t,u(t),v(t))+v(t,u(t),v(t))[v˙(t)].

References

[1] A. Abdollahi and I. Arias, Phase-field modeling of crack propagation in piezoelectric and ferroelectric materials with different electromechanical crack conditions, J. Mech. Phys. Solids 60 (2012), no. 12, 2100–2126. 10.1016/j.jmps.2012.06.014Search in Google Scholar

[2] R. Alessi, J.-J. Marigo and S. Vidoli, Gradient damage models coupled with plasticity and nucleation of cohesive cracks, Arch. Ration. Mech. Anal. 214 (2014), no. 2, 575–615. 10.1007/s00205-014-0763-8Search in Google Scholar

[3] M. Ambati, T. Gerasimov and L. De Lorenzis, A review on phase-field models of brittle fracture and a new fast hybrid formulation, Comput. Mech. 55 (2015), no. 2, 383–405. 10.1007/s00466-014-1109-ySearch in Google Scholar

[4] L. Ambrosio, N. Gigli and G. Savaré, Gradient Flows in Metric Spaces and in the Space of Probability Measures, Lectures in Math. ETH Zürich, Birkhäuser, Basel, 2005. Search in Google Scholar

[5] L. Ambrosio and V. Tortorelli, On the approximation of free discontinuity problems, Boll. Unione Mat. Ital. B (7) 6 (1992), no. 1, 105–123. Search in Google Scholar

[6] M. Artina, M. Fornasier, S. Micheletti and S. Perotto, Anisotropic mesh adaptation for crack detection in brittle materials, SIAM J. Sci. Comput. 37 (2015), no. 4, 633–659. 10.1137/140970495Search in Google Scholar

[7] J.-F. Babadjian and V. Millot, Unilateral gradient flow of the Ambrosio–Tortorelli functional by minimizing movements, Ann. Inst. H. Poincaré Anal. Non Linéaire 31 (2014), no. 4, 779–822. 10.1016/j.anihpc.2013.07.005Search in Google Scholar

[8] H. Bauschke and P. Combettes, Convex Analysis and Monotone Operator Theory in Hilbert Spaces, CMS Books Math./Ouvrages Math. SMC, Springer, New York, 2011. 10.1007/978-1-4419-9467-7Search in Google Scholar

[9] E. Bonetti, F. Freddi and A. Segatti, An existence result for a model of complete damage in elastic materials with reversible evolution, Contin. Mech. Thermodyn. 29 (2017), no. 1, 31–50. 10.1007/s00161-016-0520-3Search in Google Scholar

[10] M. Borden, C. Verhoosel, M. Scott, T. Hughes and C. Landis, A phase-field description of dynamic brittle fracture, Comput. Methods Appl. Mech. Eng. 217 (2011), 77–95. 10.1016/j.cma.2012.01.008Search in Google Scholar

[11] B. Bourdin, G. Francfort and J.-J. Marigo, Numerical experiments in revisited brittle fracture, J. Mech. Phys. Solids 48 (2000), no. 4, 797–826. 10.1016/S0022-5096(99)00028-9Search in Google Scholar

[12] A. Chambolle, An approximation result for special functions with bounded deformation, J. Math. Pures Appl. (9) 83 (2004), no. 7, 929–954. 10.1016/j.matpur.2004.02.004Search in Google Scholar

[13] V. Crismale and G. Lazzaroni, Viscous approximation of quasistatic evolutions for a coupled elastoplastic-damage model, Calc. Var. Partial Differential Equations 55 (2016), no. 1, Paper no. 17. 10.1007/s00526-015-0947-6Search in Google Scholar

[14] B. Dacorogna, Direct Methods in the Calculus of Variations, Appl. Math. Sci. 78, Springer, Berlin, 1989. 10.1007/978-3-642-51440-1Search in Google Scholar

[15] G. Dal Maso, G. Francfort and R. Toader, Quasistatic crack growth in nonlinear elasticity, Arch. Ration. Mech. Anal. 176 (2005), no. 2, 165–225. 10.1007/s00205-004-0351-4Search in Google Scholar

[16] U. Gianazza and G. Savaré, Some results on minimizing movements, Rend. Accad. Naz. Sci. XL Mem. Mat. (5) 18 (1994), 57–80. Search in Google Scholar

[17] H. Hahn, Über Annäherung an Lebesguesche Integrale durch Riemannsche Summen, Wien. Ber. 123 (1914), 713–743. Search in Google Scholar

[18] V. Hakim and A. Karma, Laws of crack motion and phase-field models of fracture, J. Mech. Phys. Solids 57 (2009), no. 2, 342–368. 10.1016/j.jmps.2008.10.012Search in Google Scholar

[19] R. Herzog, C. Meyer and G. Wachsmuth, Integrability of displacement and stresses in linear and nonlinear elasticity with mixed boundary conditions, J. Math. Anal. Appl. 382 (2011), no. 2, 802–813. 10.1016/j.jmaa.2011.04.074Search in Google Scholar

[20] F. Iurlano, A density result for GSBD and its application to the approximation of brittle fracture energies, Calc. Var. Partial Differential Equation 51 (2014), 315–342. 10.1007/s00526-013-0676-7Search in Google Scholar

[21] D. Knees and M. Negri, Convergence of alternate minimization schemes for phase-field fracture and damage, Math. Models Methods Appl. Sci., to appear. 10.1142/S0218202517500312Search in Google Scholar

[22] D. Knees, R. Rossi and C. Zanini, A vanishing viscosity approach to a rate-independent damage model, Math. Models Methods Appl. Sci. 23 (2013), no. 4, 565–616. 10.1142/S021820251250056XSearch in Google Scholar

[23] C. Kuhn and R. Müller, A continuum phase field model for fracture, Engineering Fracture Mechanics 77 (2010), no. 18, 3625–3634. 10.1016/j.engfracmech.2010.08.009Search in Google Scholar

[24] C. Larsen, C. Ortner and E. Süli, Existence of solutions to a regularized model of dynamic fracture, Math. Models Methods Appl. Sci. 20 (2010), no. 7, 1021–1048. 10.1142/S0218202510004520Search in Google Scholar

[25] B. Li, C. Peco, D. Millán, I. Arias and M. Arroyo, Phase-field modeling and simulation of fracture in brittle materials with strongly anisotropic surface energy, Internat. J. Numer. Methods Engrg. 102 (2015), no. 3–4, 711–727. 10.1002/nme.4726Search in Google Scholar

[26] C. Miehe, F. Welschinger and M. Hofacker, Thermodynamically consistent phase-field models of fracture: Variational principles and multi-field FE implementations, Internat. J. Numer. Methods Engrg. 83 (2010), no. 10, 1273–1311. 10.1002/nme.2861Search in Google Scholar

[27] A. Mielke, Differential, energetic, and metric formulations for rate-independent processes, Nonlinear PDEs and Applications, Lecture Notes in Math. 2028, Springer, Berlin (2011), 87–170. 10.1007/978-3-642-21861-3_3Search in Google Scholar

[28] A. Mielke, R. Rossi and G. Savaré, BV solutions and viscosity approximations of rate-independent systems, ESAIM Control Optim. Calc. Var. 18 (2012), no. 1, 36–80. 10.1051/cocv/2010054Search in Google Scholar

[29] A. Mielke, R. Rossi and G. Savaré, Balanced viscosity (BV) solutions to infinite-dimensional rate-independent systems, J. Eur. Math. Soc. (JEMS) 18 (2016), no. 9, 2107–2165. 10.4171/JEMS/639Search in Google Scholar

[30] A. Mielke and T. Roubíček, Rate-Independent Systems: Theory and Application, Appl. Math. Sci. 193, Springer, New York, 2015. 10.1007/978-1-4939-2706-7Search in Google Scholar

[31] A. Mielke and S. Zelik, On the vanishing-viscosity limit in parabolic systems with rate-independent dissipation terms, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 13 (2014), no. 1, 67–135. 10.2422/2036-2145.201004_003Search in Google Scholar

[32] M. Negri, Quasi-static rate-independent evolutions: Characterization, existence, approximation and application to fracture mechanics, ESAIM Control Optim. Calc. Var. 20 (2014), no. 4, 983–1008. 10.1051/cocv/2014004Search in Google Scholar

[33] R. Nochetto, G. Savaré and C. Verdi, A posteriori error estimates for variable time-step discretizations of nonlinear evolution equations, Comm. Pure Appl. Math. 53 (2000), no. 5, 525–589. 10.1002/(SICI)1097-0312(200005)53:5<525::AID-CPA1>3.0.CO;2-MSearch in Google Scholar

[34] K. Pham, H. Amor, J. Marigo and C. Maurini, Gradient damage models and their use to approximate brittle fracture, Int. J. Damage Mech. 20 (2011), no. 4, 618–652. 10.1177/1056789510386852Search in Google Scholar

[35] T. Takaishi and M. Kimura, Phase field model for mode III crack growth in two-dimensional elasticity, Kybernetika 45 (2009), no. 4, 605–614. Search in Google Scholar

Received: 2016-06-15
Revised: 2017-05-04
Accepted: 2017-05-18
Published Online: 2017-06-04
Published in Print: 2019-01-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 11.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2016-0028/html
Scroll to top button