ArticlePDF Available

Shape Optimization for a Class of Semilinear Variational Inequalities with Applications to Damage Models

Authors:

Abstract

The present contribution investigates shape optimisation problems for a class of semilinear elliptic variational inequalities with Neumann boundary conditions. Sensitivity estimates and material derivatives are firstly derived in an abstract operator setting where the operators are defined on polyhedral subsets of reflexive Banach spaces. The results are then refined for variational inequalities arising from minimisation problems for certain convex energy functionals considered over upper obstacle sets in H1H^1. One particularity is that we allow for dynamic obstacle functions which may arise from another optimisation problems. We prove a strong convergence property for the material derivative and establish state-shape derivatives under regularity assumptions. Finally, as a concrete application from continuum mechanics, we show how the dynamic obstacle case can be used to treat shape optimisation problems for time-discretised brittle damage models for elastic solids. We derive a necessary optimality system for optimal shapes whose state variables approximate desired damage patterns and/or displacement fields.
Shape optimisation for a class of semilinear variational
inequalities with applications to damage models
Christian Heinemann and Kevin Sturm
January 15, 2016
Abstract
The present contribution investigates shape optimisation problems for a class of
semilinear elliptic variational inequalities with Neumann boundary conditions. Sen-
sitivity estimates and material derivatives are firstly derived in an abstract operator
setting where the operators are defined on polyhedral subsets of reflexive Banach
spaces. The results are then refined for variational inequalities arising from minimi-
sation problems for certain convex energy functionals considered over upper obstacle
sets in H1. One particularity is that we allow for dynamic obstacle functions which
may arise from another optimisation problems. We prove a strong convergence
property for the material derivative and establish state-shape derivatives under reg-
ularity assumptions. Finally, as a concrete application from continuum mechanics,
we show how the dynamic obstacle case can be used to treat shape optimisation
problems for time-discretised brittle damage models for elastic solids. We derive a
necessary optimality system for optimal shapes whose state variables approximate
desired damage patterns and/or displacement fields.
Keywords: shape optimisation, semilinear elliptic variational inequalities, optimisation
problems in Banach spaces, obstacle problems, damage phase field models, elasticity;
AMS subject classification: 49J27, 49J40, 49Q10, 35J61, 49K20, 49K40, 74R05, 74B99.
Contents
1 Introduction 2
2 Preliminaries 4
2.1 Notation and basic relations . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Polyhedricity of upper obstacle sets in H1() ................ 6
2.3 Eulerian semi and shape derivative . . . . . . . . . . . . . . . . . . . . . . 11
3 Abstract sensitivity analysis 13
3.1 Sensitivity result for minimisers of energy functionals . . . . . . . . . . . . 13
3.2 Sensitivity result for uniformly monotone operators . . . . . . . . . . . . . 16
3.3 Variational inequality for the material derivative . . . . . . . . . . . . . . 17
1
4 A semilinear dynamic obstacle problem 19
4.1 Setting and state system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2 Perturbed problem and sensitivity estimates . . . . . . . . . . . . . . . . . 21
4.3 Limiting system for the transformed material derivative . . . . . . . . . . 25
4.4 Limiting system for the material derivative . . . . . . . . . . . . . . . . . 29
4.5 Limiting system for the state-shape derivative . . . . . . . . . . . . . . . . 31
4.6 Eulerian semi-derivative of certain shape functions . . . . . . . . . . . . . 34
5 Applications to damage phase field models 34
5.1 Physicalmodel ................................. 34
5.2 Setting up time-discretisation scheme and shape optimisation problem . . 36
5.3 Material derivative and necessary optimality system . . . . . . . . . . . . 39
1 Introduction
Finding optimal shapes such that a physical system exhibits an intended behaviour is
of great interest for plenty of engineering applications. For example design questions
arise in the construction of air- and spacecrafts, wind and combustion turbines, wave
guides and inductor coils. More examples can be found in [5] and references therein.
The physical system is usually modelled by a pde or a coupled pde system supplemented
with suitable boundary conditions. In certain cases the state is given as a minimiser of
an energy, e.g., an equilibrium state of an elastic membrane, which has to be in some
set of admissible states. The solution is then characterised by a variational inequality
holding for test-functions on the sets of admissible states.
The treatment of optimal shape and control problems for variational inequalities is
substantially more difficult as without constraints, where the sets of admissible states
is a linear space. For optimal control problems there exist a rapidly growing litera-
ture exploring different types of stationarity conditions and their approximations (see,
for instance, [14, 18]). However shape optimisation problems for systems described by
variational inequalities are less explored and reveal additional difficulties due to the in-
tricated structure of the set of admissible domains. Some results following the paradigm
first optimise—then discretise can be found in [22, 21, 15, 19] and for the first discretise–
then optimise approach we refer to [3, 1, 10].
The main aim of this paper is to establish sensitivity estimates and material deriva-
tives for certain nonlinear elliptic variational inequalities with respect to the domain.
Our approach is based on the paradigm first optimise then discretise, thus the sen-
sitivity is derived in the infinite dimensional setting. In order to highlight the main
arguments needed in the proof of these main results and to increase their applicability,
we investigate the optimisation problems firstly on an abstract operator level formulated
over a polyhedric subset Kof some reflexive Banach space V. The domain-to-state map
is there replaced by a parametrised family of operators (At) and sensitivity estimates
are shown in Theorem 3.2 and Theorem 3.3 under general assumption (see Assump-
tion (E) and Assumption (O1)). By strengthen the assumptions (see Assumption (O2))
2
differentiability with respect to the parameter thas been shown in Theorem 3.5. One
crucial requirement is the polyhedricity of the closed convex set Kon which the oper-
ators are defined. The results are applicable for optimal shape as well as for optimal
control problems.
Equipped with the proven abstract results we resort to shape optimisation problems
where the state system is a variational inequality of semilinear elliptic type given by
uKψand ϕKψ:dE(Ω, u;ϕu)0
with the energy
E(Ω, u) = Z
1
2|∇u|2+λ
2|u|2+W(x, u) dx(λ > 0)
and the upper obstacle set
Kψ=vH1(Ω) : vψa.e. in Ω.
In the classical theory of VI-constrained shape optimisation problems established in [23],
linear variational inequalities with constant obstacle and W(x, u) = f(x) for some given
fixed function f:DRdefined on a “larger set” DΩ have been investigated by
means of conical derivatives of projection operators in Hilbert spaces dating back to [18].
For results on topological sensitivity analysis for variational inequalities and numerical
implementations we refer to [13] as well as [2].
In our paper we allow for semilinear terms in the variational inequality by including
convex contributions to Wwith respect to uand also consider a dependence of W
and ψon Ω in a quit general sense. As presented in the last section of this work ψ
may itself be a solution of a variational inequality. Such general Ω-dependence of the
obstacle will be referred to as “dynamic obstacle” in constrast to the case of a “static
obstacle” where ψ(x) = g(x) for some fixed function g:DR.
To apply the abstract sensitivity results we perform the transformation u7→ y:=
uψsuch that the transformed problem is formulated over the cone H1
(Ω), i.e.,
the non-positive half space of H1(Ω). Existence of the material derivative ˙ywhich
turns out to be the unique solution of a variational inequality considered over the cone
Ty(H1
(Ω)) kern(dE(u;·)) and strong convergence of the corresponding difference quo-
tients are established in Theorem 4.8 and Corollary 4.9. The variational inequality
characterising the material derivative ˙uis then established in Corollary 4.11. Moreover
in the case of a static obstacle and H2(Ω)-regularity for uwe derive relations for the
state-shape derivative u0in Theorem 4.15 and Corollary 4.16.
The theorems for the abstract semilinear case are then applied to a specific model
problem from continuum damage mechanics. There one considers an elastic solid which
undergoes deformation and damage processes in a small strain setting. The state of
damage is modelled by a phase field variable χwhich influences the material stiffness and
which is described by an evolution inclusion forcing the variable χto be monotonically
decreasing in time. We consider a time-discretised version of the evolution system (but
3
we stay continuous in the spatial components) where the damage variable fulfills for all
time steps the constraints
χNχN1. . . χ01 a.e. in Ω.
Such constraints lead to N-coupled variational inequalities with dynamic obstacle sets
of the type
Kk1(Ω) = vH1(Ω) : vχk1a.e. in Ω, k = 1, . . . , N.
Our objective is to find an optimal shape Ω such that the associated displacement fields
(uk)N
k=1 and damage phase fields (χk)N
k=1 minimise a given tracking type cost functional.
We derive relations for the material derivative and establish necessary optimality condi-
tions for optimal shapes which are summarised in Proposition 5.3.
Structure of the paper
In Section 2 we recall some basics notions from convex analysis and shape optimisation
theory. For reader’s convenience and for the sake of clarity we derive tangential and
normal cones of Kψand prove polyhedricity of Kψby using arguments from [18, 4, 12].
In Section 3 we establish sensitivity and material derivative results in an abstract
operator setting (see Theorem 3.2, Theorem 3.3 and Theorem 3.5). Some results are
even applicable to quasi-linear problems such as to p-Laplace equations. The advantage
of this approach is that the theorems can be applied to a large class of optimisation
problems including shape optimisation and optimal control problems.
This flexibility is demonstrated in Section 4 where semilinear VI-constrained shape
optimisation problems with an energy and obstacle of type E(Ω, u) and Kψfrom above
are treated. By applying the abstract results from Section 3 we derive sensitivity esti-
mates for the shape-perturbed problem in Proposition 4.5, material derivatives in The-
orem 4.8 and state-shape derivatives in Theorem 4.15.
Finally, in Section 5, we apply the still abstract results from Section 4 to a particular
problem in continuum damage mechanics where dynamic obstacles occur.
2 Preliminaries
2.1 Notation and basic relations
For the treatment of variational inequalities we recall certain well-known cones from
convex analysis (the definitions can, for instance, be found in [4, Chapter 2.2.4] and [23,
Chapter 4.1]). Let KVbe a subset of a real Banach space Vand denote by Vits
topological dual space.
The radial cone at yKof the set Kis defined by
Cy(K) := {wV:t > 0, y +tw K},(1)
the tangent cone at yas
Ty(K) := Cy(K)V(2)
4
and the normal cone at yas
Ny(K) := {wV:vK, hw, v yiV0}.(3)
Furthermore we introduce the polar cone of a set Kas
[K]:= {wV:vK, hw, viV0},(4)
and the orthogonal complements of elements yVand yV
[y]:= {wV:hw, yiV= 0},
kern(y) := [y]:= {wV:hy, wiV= 0}.
The normal cone may also be written as
Ny(K)=[Ty(K)]= [Cy(K)].(5)
In combination with the bipolar theorem (see [4, Prop. 2.40]) we obtain
Ty(K) = [[Ty(K)]]= [Ny(K)].(6)
We recall that a closed convex set KVis polyhedric if (cf. [14])
yK, wNy(K), Cy(K)[w]V=Ty(K)[w].(7)
Note that the inclusion “” is always satisfied above. Due to Mazur’s lemma and the
convexity of the involved sets, the closure in Vcan also be taken in the weak topology.
The following lemma shows a useful implication of (7) involving variational inequal-
ities arising from (possibly non-)linear operators.
Lemma 2.1. Let KVbe a polyhedric subset.
(i) Let A:KVbe an operator and let ybe a solution of the following variational
inequality
yKand ϕK:hA(y), ϕ yiV0.(8)
Then it holds
Cy(K)kern(A(y)) = Ty(K)kern(A(y)).(9)
(ii) For all vVit holds
Cy(K)[vy]=Ty(K)[vy],
where ydenotes the projection of von K.
Proof. To (i): We infer from (8) that −A(y)Ny(K). Thus definition (7) implies
Cy(K)kern(−A(y)) = Ty(K)kern(−A(y)).
The identity kern(−A(y)) = kern(A(y)) completes the proof.
To (ii): This follows from vyNy(K).
5
2.2 Polyhedricity of upper obstacle sets in H1(Ω)
Let us consider an important class of polyhedral subsets which will be utilized in Section
4 where semilinear obstacle problems are treated. Let Ω Rdbe a Lipschitz domain
and V=H1(Ω). Moreover let ψVbe a given function. We define the upper obstacle
set as
Kψ:= {wH1(Ω) : wψa.e. in Ω}.(10)
In the remaining part of this subsection we will sketch the proofs for the characterisation
of the tangential and normal cones as well as of the polyhedricity of Kψfor reader’s
convenience since such obstacles sets are usually considered in the space
H1(Ω) in the
literature. The adaption to H1(Ω) requires some careful modifications in the proofs.
Furthermore we denote with M+(Ω) the Radon measures on Ω. The Riesz repre-
sentation theorem for local compact Hausdorff spaces (see [6, Theorem VIII.2.5]) states
that for each non-negative functional I:C(Ω) Rthere exists a unique Radon measure
µM+(Ω) such that for all fC(Ω)
I(f) = Z
fdµ. (11)
In the sequel we will use the following notation for the half space
H1
+(Ω) := vH1(Ω) : v0 a.e. in Ω.
With the help of the Riesz representation theorem we are now in the position to give a
characterisation of (cf. [4, Chapter 6.4.3] for
H1(Ω) instead of H1(Ω))
H1(Ω)
+:= IH1(Ω):hI, viH1(Ω) 0 for all vH1
+(Ω).
Lemma 2.2. We have
H1(Ω)
+=nIH1(Ω):!µIM+(Ω),vH1(Ω) C(Ω),hI, viH1(Ω) =Z
vdµIo.
(12)
Proof. Let I:H1(Ω) Rbe a non-negative functional. Then the restriction I|H1(Ω)C(Ω)
is a non-negative functional on the space H1(Ω) C(Ω) =: Y.
Now let yYbe arbitrary. Then y+:= max{0, y}and y:= min{0, y}(defined in
a pointwise sense) are also in Yand we find by non-negativity of L:
|Ly|=|L(y++y)|=|L(y+)
|{z }
0
+L(y)
|{z }
0
| | L(y+)
|{z }
0
L(y)
| {z }
0
|
≤ |L(y+y)|=L(|y|)
=L(|y| − 1kyk)
| {z }
0
+kykL(1)
6
≤ kykL(1).
Thus I|Yis continuous in the C(Ω)-topology. Since Yis also dense in C(Ω) the functional
I|Yhas a unique continuous and non-negative extension ˜
I:C(Ω) Rover C(Ω). By the
Riesz representation theorem (see (11)) we find a µM+(Ω) such that I(v) = Rvdµ
for all vC(Ω).
Conversely, let Ibe in the set on the right-hand side of (12). Then we know
hI, viH1(Ω) =RvdµI0 for all vY+:= {vY:v0 pointwise in Ω}. So
by density of Y+in H1
+(Ω) we obtain IH1(Ω)
+.
Remark 2.3. Note that, by an abuse of notation, the right-hand side of (12) is some-
times written as H1(Ω)M+(Ω) (see, e.g., [4, Chapter 6]).
For the notion of capacity of a set,quasi-everywhere (q.e.) and quasi-continuous
representant where refer to [12, Chapter 3.3]. The following result is an extension of
(11) valid for elements from H1(Ω)
+.
Lemma 2.4. For all IH1(Ω)
+and all fH1(Ω) we have ˜
fL1(Ω, µI)and
hI, f iH1(Ω) =Z
˜
fdµI,(13)
where ˜
f(defined on ) denotes a quasi-continuous representative of fand µIthe measure
from (12) of Lemma 2.2.
Proof. The proof of this lemma requires some modifications of [4, Lemma 6.56] and
references therein which were designed to the situation V=
H1(Ω). In our case we will
need the following auxiliary results:
(a) For an arbitrary DRdthe capacity of Dcalculates as
cap(D) = inf kvk2
H1(Rd):vH1(Rd) and v1 a.e. in a neighborhood of D.
See [12, Proposition 3.3.5] for a proof.
(b) Any function fH1(Ω) can be approximated by a sequence {fn} ⊆ C
c(Rd) in
the sense that fnfin H1(Rd) as n→ ∞ by extending fto Rdwith compact
support and then uses an approximation argument via Friedrichs mollifiers.
The proof carried out in the following steps on the basis of [4, Lemma 6.56] and the
references therein (see also [12, Th´eor`eme 3.3.29] for the case V=H1(Rd)):
Claim 1: There exists a sequence {fn} ⊆ C
c(Rd)s.t. fn|˜
fin H1(Ω) and q.e. in
Let {fn}be given by (b). By resorting to a subsequence (we omit the subscript) we may
find kfnfkH1(Rd)2nn1and therefore
X
n=1
4n+1kfn+1 fnk2
H1(Rd)
X
n=1
4n+1(kfn+1 fkH1(Rd)+kfnfkH1(Rd)2<+.
(14)
7
We define
Bn:= xRd:|fn+1(x)fn(x)| ≥ 2n.
Since |fn+1 fn|is a continuous with compact support in Rd, the set Bnis compact and
2n+1|fn+1 fn| ≥ 1 holds in a neighborhood of Bn.
Thus by (a)
cap(Bn)4n+1kfn+1 fnk2
H1(Rd).
Using this estimate, the sub-additivity of the capacity (see [12, Remarque 3.3.10]) and
(14), we obtain:
cap
[
k=n
Bk
X
k=n
cap(Bk)
X
k=n
4n+1kfn+1 fnk2
H1(Rd)0 as n→ ∞.(15)
Now let nNand x\S
k=nBkbe arbitrary. Then {fk(x)}knis a Cauchy sequence
since for all mn:
|fm(x)fn(x)| ≤
m1
X
k=n
|fk+1(x)fk(x)| ≤
m1
X
k=n
2k.
We denote the limit with ˜
f(x) and gain for all N, K n:
|˜
f(x)fN(x)| ≤ | ˜
f(x)fK+1(x)|
| {z }
0 as K→∞
+
K
X
k=N
|fk+1(x)fk(x)|
| {z }
2ksince x\S
k=nBk
A limit passage K→ ∞ then shows
|˜
f(x)fN(x)| ≤
X
k=N
2k.
This estimate implies that {fN}Nnconverges uniformly to ˜
fon the set \S
k=nBk.
Due to (15) we obtain Claim 1.
Claim 2: If cap(A)=0for a Borel set Athan µI(A)=0.
Let ε > 0 be arbitrary. By (a) we find a function uH1(Rd) such that kukH1(Ω) < ε
and u1 a.e. on Aεwhere Aεis a neighborhood of A. Thus there exists a Lipschitz
function fε:Rd[0,1] such that
fε(x) =
0 if xRd\Aε,
(0,1) if xAε\A,
1 if xA.
8
Then fεu0 a.e. in Ω and by Lemma 2.2
µI(A) = ZA
1dµIZ
fεdµI=hI, fεiH1(Ω) =hI, uiH1(Ω) +hI, fεuiH1(Ω)
| {z }
0 since feua.e. in Ω
≤ hI, uiH1(Ω)
εkIkH1(Ω).
The limit passage ε&0 yields to claim.
Claim 3: fn˜
fin L1(Ω, µI)
Lemma 2.2 implies for every n, m N
Z
|fnfm|dµI=hI, |fnfm|iH1(Ω) ≤ kIkH1(Ω)kfnfmkH1(Ω),(16)
where fnis the approximation sequence from Claim 1. Since fnfin H1(Ω) we obtain
from (16) that {fn}is a Cauchy sequence in L1(Ω, µI). Thus there exists a limit element
˜gL1(Ω, µI) and a subsequence (we omit the subscript) such that fn˜gin L1(Ω, µI)
and pointwise µI-a.e. on Ω. However, by Claim 1, we already know that fnconverges
q.e. to ˜
fon Ω and, by Claim 2, we find that this covergence is also µI-a.e. Thus ˜
f= ˜g
µI-a.e.
Conclusion:
Finally, Lemma 2.2 shows for every nN
hI, fniH1(Ω) =Z
fndµI.
With the properties proven above we can pass to the limit n→ ∞ and obtain (13).
We are now in a position to characterise the tangential and normal cones in Kψ.
The proofs of the following results are based on arguments from [18, Lemme 3.1-3.2,
Th´eor`eme 3.2].
Lemma 2.5. Let yKψand Kψbe as in (10). Then it holds
Ty(Kψ) = uH1(Ω) : ˜u0q.e. on {˜y=˜
ψ},(17a)
Ny(Kψ) = IH1(Ω):IH1(Ω)
+and µI({˜y < ˜
ψ})=0,(17b)
where ˜ydenotes a quasi-continuous representant of y(the same for ˜uand ˜
ψ) and µI
M+(Ω) the measure associated to Iby Lemma 2.2.
Please notice that the sets
{˜y=˜
ψ}:= {xΩ : ˜y(x) = ˜
ψ(x)},
{˜y < ˜
ψ}:= {xΩ : ˜y(x)<˜
ψ(x)}
are calculated for arguments in (not only in ).
9
Proof. From the definitions (1)-(3) we see that
Ty(Kψ) = Tyψ(K), Ny(Kψ) = Nyψ(K)
with K:= {wH1(Ω) : w0 a.e. in Ω}. Thus it suffices to prove the assertion for
Kψ=K.
We firstly prove (17b).
”: Let INy(K). Then by using definition (3) and choosing v=y+wfor an
arbitrary wH1(Ω) with w0 a.e. we obtain hI, wiH1(Ω) 0. Thus IH1(Ω)
+
and by Lemma 2.2 we find the associated measure µIfrom (12). On the other hand by
choosing v=ψand v= 2yin (3) yields hI, yiH1(Ω) = 0. From Lemma 2.4 we obtain
Z
˜ydµI= 0 with a quasi-continuous representant ˜yof y. (18)
Since y0 a.e. in Ω we find ˜y0 q.e. in Ω (see [12, Remarque 3.3.6]). This implies
in combination with (18) that R|˜y|dµI= 0. Thus R{˜y<0}|˜y|dµI= 0 and therefore
µI({˜y < 0}) = 0.
”: Let IH1(Ω)
+with µI({˜y < 0}) = 0. Now let vKbe arbitrary. The splitting
v= max{v, y}+ min{0, v y}implies
hI, v yiH1(Ω) =hI, max{v, y} − yiH1(Ω) +hI , min{0, v y}iH1(Ω)
| {z }
0
Z{˜y=0}
max{˜v, ˜y} − ˜ydµI+Z{˜y<0}
max{˜v, ˜y} − ˜ydµI
| {z }
=0 since µI({˜y<0})=0
Z{˜y=0}
max{˜v, 0}
| {z }
=0 since vK
dµI= 0.
Hence INy(K).
Now we prove (17a). By applying the bipolar theorem as in (6) as well as Lemma
2.4, we find
Ty(K) = nuH1(Ω) : Z
˜udµI0 for all IH1(Ω)
+with µI({˜y < 0})=0o
=nuH1(Ω) : Z{˜y=0}
˜udµI0 for all IH1(Ω)
+with µI({˜y < 0}) = 0o.
From this representation we see that the “”-inclusion in (17a) is fulfilled. Conversely,
let uTy(K). By definition of Ty(K) given in (2) we find a sequence vnKand tn>0
such that tn(vny)uin H1(Ω) as n→ ∞. This implies for a subsequence (we omit
the subindex) tnvn˜y)˜uq.e. in Ω. Since vnKwe see that
tnvn˜y) = tn˜vn0 q.e. on {˜y= 0}.
Thus ˜u0 q.e. on {˜y= 0}.
10
Theorem 2.6 (cf. [18, Th´eor`eme 3.2]).The set Kψis polyhedric.
Proof. Let yand was in (7) and let vTy(Kψ)[w]. Then there exists a sequence
vnvstrongly in H1(Ω) such that vnCy(Kψ). Define
v0
n:= max{vn, v}.
By resorting to quasi-continuous representants we find by Lemma 2.5
v0 q.e. in {y= 0}and vn0 q.e. in {y= 0}
and thus
v0
n0 q.e. in {y= 0}.
Moreover by definition of v0
n
vv0
n0 q.e. in Ω.
Invoking Lemma 2.5 again yield v0
nTy(Kψ) and vv0
nTy(Kψ). Since wNy(Kψ)
we see by (5) that
hw, v0
ni ≤ 0 and hw, v v0
ni ≤ 0.
Taking also hw, vi= 0 into account we obtain from above that hw, v0
ni= 0. Thus
v0
nCy(Kψ)[w].
2.3 Eulerian semi and shape derivative
We recall some preliminaries from shape optimisation theory. For more details we refer
to [5].
Let X:RdRdbe a vector field satisfying a global Lipschitz condition: there is a
constant L > 0 such that
|X(x)X(y)| ≤ L|xy|for all x, y Rd.
Then we associate with Xthe flow Φtby solving for all xRd
d
dtΦt(x) = Xt(x)) on [τ, τ ],Φ0(x) = x. (19)
The global existence of the flow Φ : R×RdRdis ensured by the theorem of Picard-
Lindel¨of.
Subsequently, we restrict ourselves to a special class of vector fields, namely Ck-
vector fields with compact support in some fixed set. To be more precise for a fixed
open set DRd, we consider vector fields belonging to Ck
c(D, Rd). We equip the space
Ck
c(D, Rd) respectively C
c(D, Rd) with the topology induced by the following family
of semi-norms: for each compact KDand muli-index αNdwith |α| ≤ kwe
define kfkK,α := supxK|αf(x)|.With this familiy of semi-norms the space Ck
c(D, Rd)
becomes a locally convex vector space.
Next, we recall the definition of the Eulerian semi-derivative.
11
Definition 2.7. Let DRdbe an open set. Let J: Ξ Rbe a shape function defined
on a set Ξof subsets of Dand fix k1. Let Ξand XCk
c(D, Rd)be such that
Φt(Ω) Ξfor all t > 0sufficiently small. Then the Eulerian semi-derivative of Jat
in direction Xis defined by
dJ(Ω)(X) := lim
t&0
Jt(Ω)) J(Ω)
t.(20)
(i) The function Jis said to be shape differentiable at if dJ (Ω)(X)exists for all
XC
c(D, Rd)and X7→ dJ(Ω)(X)is linear and continuous on C
c(D, Rd).
(ii) The smallest integer k0for which X7→ dJ(Ω)(X)is continuous with respect to
the Ck
c(D, Rd)-topology is called the order of dJ (Ω).
The set Din the previous definition is usually called hold-all domain or hold-all set
or universe.
In the case that the state system is given as a solution of a variational inequality
we cannot expect dJ(Ω)(X) to be linear in X. However we have the following general
result:
Lemma 2.8. Suppose that the Eulerian semi-derivative dJ(Ω)(X)exists for all X
Ck
c(D, Rd). Then dJ (Ω)(·)is positively 1-homogeneous.
Proof. Let λ > 0 be arbitrary. We write ΦλX
tfor the flow induced by λX. By definition
(19), we see that ΦλX
tand ΦX
λt solve
d
dtΦλX
t(x) = λXλX
t(x)),d
dtΦX
λt(x) = λXX
λt(x))
as well as ΦλX
0(x) = xand ΦX
0(x) = x. Uniqueness of the flow implies ΦλX
t= ΦX
λt.
Finally,
dJ(Ω)(λX) = lim
t&0
JλX
t(Ω)) J(Ω)
t= lim
t&0
JX
λt(Ω)) J(Ω)
t=λ dJ(Ω)(X).
The following result can be found for instance in [5]:
Lemma 2.9. Let DRdbe open and bounded and suppose XC1
c(D, Rd).
(i) We have
ΦtI
t∂X strongly in C(D, Rd,d)
Φ1
tI
t→ − ∂X strongly in C(D, Rd,d)
det(Φt)1
tdiv(X)strongly in C(D).
(ii) For all open sets Dand all ϕW1
µ(Ω),µ1, we have
ϕΦtϕ
t→∇ϕ·Xstrongly in Lµ(Ω).(21)
12
3 Abstract sensitivity analysis
In this section we will derive sensitivity estimates and relations for material derivatives
under general conditions. We start in Section 3.1 with minimisers of certain p-coercive
energy functionals and deduce a H¨older-type estimate with exponent 1/p. We present an
example which includes the quasi-linear p-Laplacian p(·) = div(|∇ ·|p2∇·). Then we
proceed in Section 3.2 with solutions of monotone operators where we are able to improve
the estimates from Subsection 3.1. For the case p= 2 we even establish a Lipschitz type
sensitivity estimate. Finally in Subsection 3.3 we strengthen the assumptions in order to
establish the weak material derivative. A crucial requirement will be the polyhedricity
of the underlying set.
In this whole section Vwill denote a Banach space, KVa closed convex subset
and τ > 0 a fixed constant.
3.1 Sensitivity result for minimisers of energy functionals
Our starting point is a family of energy functionals
E: [0, τ ]×VR,
where we denote the set of attained infima at t[0, τ] by
X(t) := utV: inf
ϕKE(t, ϕ) = E(t, ut).(22)
Our aim is to establish a general result showing the convergence of minimisers of E(t, ·)
to minimisers of E(0,·) as t&0. Before we state our abstract sensitivity result, we
recall [20, Theorem 1] which will be used in a subsequent proof:
Theorem 3.1 ([20, Theorem 1]).Let [·]be a seminorm on V. Let E:VRbe an
energy functional such that for all v, w Kthe mapping s7→ γ(s) := E(sw + (1 s)v))
is C1on [0,1]. Let us denote by A:KVthe Gateaux-differential of Ewhich is
supposed to be p-coercive on K:
α > 0,u, v K, hA(u)− A(v), u viVα[vw]p.
Then every minimum uof Eon Ksatisfies:
vK, α
p[uv]pE(u)E(v).
In what follows let Esatisfy the following assumption:
Assumption (E) Suppose that the energy functionals E(t, ·)satisfies for a given p1:
(i) c1>0,c2>0,ϕK, E(·, ϕ)is differentiable and
t[0, τ ], ∂tE(t, ϕ)c1kϕkp
V+c2;
13
(ii) c > 0,Λ>0,ϕK,t[0, τ ],
E(t, ϕ)ckϕkp
VΛ;
(iii) t[0, τ ],E(t, ·)is Gateaux-differentiable and
α > 0,u, v K, hAt(u)− At(v), u viVα[vw]p,
where hAt(v), wiV:= dE(t, v;w)and [·]is a semi-norm on V;
(iv) v, w K,t[0, τ ],
s7→ γ(s) := E(t, sv + (1 s)w)is C1([0,1])
Now we are in the position to state and prove our sensitivity result:
Theorem 3.2. Let E: [0, τ]×VRbe a family of energy functionals satisfying
Assumption (E) and let X(t)be non-empty for every t[0, τ]. Then X(t) = {ut}is a
singleton and there exists a constant c > 0such that for all t[0, τ ]:
[utu0]ct1/p.
Proof. Let t[0, τ ] and utX(t). Let us first show that utis bounded in Vuniformly
in t. According to Assumption (E) (i)-(ii), the definition of utand the mean value
theorem, we obtain ηt(0, t) such that
ckutkp
VΛE(t, ut)
E(t, u0)
=E(0, u0) + t∂tE(ηt, u0)
E(0, u0) + tc1ku0kp
V+c2.
(23)
This shows that kutkVCfor all t[0, τ] for some constant C > 0. Furthermore
applying Theorem 3.1 by using Assumption (E) (iii)-(iv) shows
c[utu0]pE(t, ut)E(t, u0),(24)
c[utu0]pE(0, u0)E(0, ut).(25)
Adding both inequalities, applying the mean value theorem twice with some ηt, ζt(0, t)
and using Assumption (E) (i) and the estimate (23) yields
2c[utu0]pE(t, ut)E(t, u0) + E(0, u0)E(0, ut)
ttE(ηt, ut)tE(ζt, u0)
tC(kutkp
V+ku0kp
V)
(23)
tC(1 + ku0kp
V).
(26)
This finishes the proof.
14
Example (p-Laplace equation)
As an application of Theorem 3.2 let us consider the p-Laplace equation
div(|∇u|p2u) = fin K=V=
W1
p(Ω)
on a bounded Lipschitz domain Ω and the associated energy given by
E(0, ϕ) = 1
pZ
|∇ϕ|pdxZ
f ϕ dx, ϕ
W1
p(Ω).
By applying integration by substitution, the energy of the perturbed equation trans-
ported to Ω via integration by substitution is of the form
E(t, ϕ) = 1
pZ
ξ(t)|B(t)ϕ|pdxZ
f(t)ϕdx,
More generally we assume that ξ: [0, τ]Rand B: [0, τ ]Rd×dare C1-functions
which satisfy ξ(0) = 1 and B(0) = I. Moreover let f(0) = fand f(·, x) be differentiable
and f0(t)Lp0(Ω) be uniformly bounded where p0=p/(p1) denotes the conjugate of
p. We check that the assumptions in (E) are satisfied:
Indeed, we have
tE(t, ϕ) = Z
ξ0(t)1
p|B(t)ϕ|p+ξ(t)|B(t)ϕ|p2B(t)ϕ·B0(t)ϕdxZ
f0(t)ϕdx.
Thus applying H¨older and Young’s inequalities we verify Assumption (E) (i):
tE(t, ϕ)Z
ξ0(t)1
p|B(t)ϕ|p+ξ(t)|B(t)ϕ|p2B(t)ϕ·B0(t)ϕdxZ
f0(t)ϕdx
cZ
|∇ϕ|p+|f0(t)||ϕ|dx
ck∇ϕkp
Lp+ 1/p0kf0(t)kp0
Lp0+1
pkϕkp
Lp.
On the other hand using Young’s and Poincar´e’s inequality with small ε > 0
E(t, ϕ)ck∇ϕkp
Lp1/p0()1
p1kf(t)kp0
Lp/(p1) εkϕkp
Lp
c1kϕkp
W1
pc2εkϕkp
Lp.
Thus we have verified Assumption (E) (ii). Assumption (E) (iii) follows from uniform
p-monotonicity of p(·) and Assumption (E) (iv) by direct calculations.
Finally we may use Theorem 3.2 and obtain kutukW1
p(Ω) ct1/p for some constant
c > 0 and all sufficiently small t > 0. In the case of the usual Laplace equation, that is
for p= 2, we get kutukH1(Ω) ct1/2.
15
3.2 Sensitivity result for uniformly monotone operators
In this section we develop sensitivity results for variational inequalites involving uni-
formly monotone operators. Let Vbe a normed space and KVbe a closed convex
subset.
Assumption (O1) Suppose that (At) : KV,t[0, τ]is a family of operators such
that for a given p1:
(i) α > 0,t[0, τ ],u, v K:
αkuvkp
V≤ hAt(u)− At(v), u viV;
(ii) uK,c > 0,t[0, τ ],vK,
|hAt(u)− A0(u), u viV| ≤ ctkuvkV.
Theorem 3.3. Suppose that (At) : KVis a family of operators satisfying As-
sumption (O1). For every t > 0we denote by utKa solution of the variational
inequality
utKand vK, hAt(ut), v utiV0.(27)
Then there exists a c > 0such that
t[0, τ ] : kutu0kVct 1
p1.
Proof. Taking into account Assumption (O1) and (27):
αkutu0kp
V≤ hAt(ut)− At(u0), utu0iV
≤ −hAt(u0), utu0iV
=hA0(u0), utu0iV+hA0(u0)− At(u0), utu0iV
≤ |hA0(u0)− At(u0), utu0iV|
ctkutu0kV.
Remark 3.4. In the important case p= 2 Theorem 3.3 yields a Lipschitz type estimates.
Example (p-Laplace equation)
It can be checked that the p-Laplace example from Subsection 3.1 where Atis given by
hAt(u), ϕi
W1
p
=Z
ξ(t)|B(t)u|p2B(t)u·B(t)ϕf(t)ϕdx
also fulfills Assumption (O1). Thus in this case Theorem 3.3 gives a sharper estimate
than Theorem 3.2.
16
3.3 Variational inequality for the material derivative
In the previous section we have shown that under certain conditions on (At) satisfied for
p= 2 the quotient (utu0)/t stays bounded. In this subsection we additionally assume
that Vis reflexive and that KVis a polyhedric subset. Then there will be a weakly
converging subsequence of (utu0)/t converging to some zV. If this zis unique the
whole sequence converges and additionally satisfies some limiting equation which is the
subject of this subsection.
Let (At) be as in Subsection 3.2 and define in accordance with (22) for all t[0, τ ]
the solution set of the associated variational inequality as
X(t) := utK:ϕK, hAt(ut), ϕ uti ≥ 0.(28)
We will write u:= u0and A:= A0. The variational inequality for the material derivative
will be deduced from the following assumptions:
Assumption (O2) Suppose that the family (At)satisfies
(i) for all v, w Vand all uK,
hA(u)w, viV:= lim
t&0A(u+tw)− A(u)
t, vV
and
hA0(u), viV:= lim
t&0At(u)− A(u)
t, vV
exist;
(ii) for all null-sequences (tn), for all sequences (vn)in Vconverging weakly to some
vV, for all utnX(tn)converging strongly to some uK, we have
hA0(u), viV= lim
n0Atn(utn)− A(utn)
tn
, vnV
;
(iii) for all null-sequences (tn), there exists a subsequence (still indexed the same) such
that utnX(tn)converges storngly to uKand (unu)/tnconverges weakly to
some zVand
hA(u)z, ziVlim inf
n0A(utn)− A(u)
tn
,utnu
tnV
and for all (vn)in Vconverging strongly to vV:
hA(u)z, viV= lim
n0A(utn)− A(u)
tn
, vnV
.
17
Theorem 3.5. Let Vbe a reflexive Banach space and KVa polyhedric subset.
Suppose that At:KV,t[0, τ]is a family of operators satisfying Assumption (O1)
for p= 2 and (O2). Suppose that utX(t), i.e., utsolves
utK, hAt(ut), ϕ utiV0ϕK. (29)
Then the material derivative ˙u:= weak limt&0(utu)/t exists and solves
˙uTu(K)kern(A(u)) and (30a)
ϕTu(K)kern(A(u)) : hA(u) ˙u, ϕ ˙uiV≥ −hA0(u), ϕ ˙uiV.(30b)
Proof. Let us firstly show (30a). We get by (29)
ϕK(Ω) : hAt(ut), ϕ uti ≥ 0,(31)
ϕK(Ω) : hA(u), ϕ ui ≥ 0.(32)
Thus testing (31) with uand (32) with utand dividing by t > 0, we obtain by setting
zt:= (utu)/t
hAt(ut), zti ≤ 0,hA(u), zti ≥ 0.(33)
By invoking Theorem 3.3 with p= 2 we know that utustrongly in Vand that ztis
bounded in Vwhich allows us to choose a weakly convergence subsequence with limit
˙uV. We find (by omitting the subscript)
hAt(ut), zti − hA(u),˙ui
=hAt(ut)− A(ut), zti
|{z }
0 by Assumption (O2) (ii)
+DA(ut)− A(u)
t, utuE
| {z }
0 by Assumption (O2) (iii)
+hA(u), zt˙ui
| {z }
0
Therefore passing to the limit in (33) gives 0 ≤ hA(u),˙ui ≤ 0 and thus ˙ukern(A(u)).
Furthermore we know by the definition of the radial cone that ztCu(K). Taking the
weak convergence zt˙uin Vand Mazur’s Lemma into account we find ˙uTu(K).
Thus (30a) is proven.
Now we will show (30b) by using (29) and obtain for every ϕV:
hA(ut)− A(u), ϕ uti=hA(ut)− At(ut), ϕ uti+hAt(ut)− A(u), ϕ uti
≥ hA(ut)− At(ut), ϕ uti − hA(u), ϕ uti.(34)
By definition of the radial cone Cu(K) (see (1)) we find for every ϕCu(K) a t>0
such that for all t[0, t]: u+K. Plugging this test-function into (34) we obtain
for all ϕCu(K)
hA(ut)− A(u), tϕ (utu)i ≥ hA(ut)− At(ut), tϕ (utu)i − hA(u), tϕ (utu)i.
(35)
18
Dividing the previous equation by t2and setting zt:= (utu)/t, we obtain
A(ut)− A(u)
t, ϕ zt At(ut)− A(ut)
t, ϕ zt1
thA(u), ϕ zti.(36)
Now let ϕCu(K)kern(A(u)). Then because of hA(u), ϕi= 0 and the definition of
uX(0) (testing the relation in (28) with ut), we find
−hA(u), ϕ zti ≥ 0.
Thus (36) reads
A(ut)− A(u)
t, ϕ zt At(ut)− A(ut)
t, ϕ zt.(37)
Using Assumption (O2) we may take the lim sup on both sides to obtain (note that
lim sup(...) = lim inf (...))
hA(u)z, ϕ zi ≥ −hA0(u), ϕ zi ∀ϕCu(K)kern(A(u)).
Via density arguments we obtain the inequality for all ϕCu(K)kern(A(u)). Finally
using polyhedricity of Kand Lemma 2.1 (i) finish the proof.
4 A semilinear dynamic obstacle problem
In this section we are going to apply the theorems from Section 3 to generalised obstacle
problems with convex energies. present a generalised obstacle problem. It also covers
previous results from [23] where the zero obstacle case has been treated as a special
case. A non-trivial example from continuum damage mechanics is presented afterward
in Section 5.
4.1 Setting and state system
Let DRdbe an open and bounded subset. We consider a convex energy of the
following type
E(Ω, ϕ) := Z
1
2|∇ϕ|2+λ
2|ϕ|2+W(x, ϕ) dx, ϕ H1(Ω),(38)
where Ω Dis a bounded Lipschitz domain and λ > 0. The energy is minimised over
the convex set
Kψ(Ω) := ϕH1(Ω) : ϕψa.e. in Ω.
A particularity of this setting is that, besides the density function W, also the obstacle
function ψis allowed to depend on the shape variable Ω (the precise assumptions are
stated below in Assumption (A1)):
dynamic density function: 7→ W
dynamic obstacle: 7→ ψH1(Ω)
In the special case ψ0 we write K(Ω) := K0(Ω).
19
Remark 4.1. (i) An important class which is covered by our setting are static obstacle
problems where ψ:= Ψ|with a given function ΨH2(D).
(ii) The energy E(Ω,·)is motivated by time-discretised parabolic problems, where an
additional λ-convex non-linearity may be included in E. By choosing a small time
step size, the incremental minimisation problem may take the form (38).
In context with time-discretised damage models in Section 5 we are faced with
iterative obstacle problems. In this case the obstacle ψitself is a solution of a
variational inequality describing the damage profile from the previous time step.
As we will see it suffices to have H1(Ω)-regularity of the damage profile provided
that the material derivative of the obstacle exists in H1(Ω) and the initial value is
in H2(Ω). We will present this application in the last section.
For later use we recall that the Sobolev exponent 2depending on the spatial dimen-
sion dto the space H1(Ω) is defined as
2:=
2d
d2if d > 2,
arbitrary in [1,+) if d= 2,
+if d= 1.
(39)
Its conjugate (2)0is given by 2
21with the convention that (2)0:= 1 for 2= +.
For well-posedness of the state system we require the following assumptions (note that
we restrict ourselves to the convex case which will be exploited in the next sections):
Assumption (A1) For all Lipschitz domains Dit holds:
(i) W(x, ·)is convex and in C1(R)for all x;
(ii) the following map H1(Ω) Ris assumed to be continuous (in particular the
integral exists)
y7→ Z
W(x, y(x)) dx
and bounded from below by
Z
W(x, y(x)) dx≥ −c(kykH1+ 1);
(iii) for all y, ϕ H1(Ω):
Z
W(x, y +)W(x, y)
tdxZ
yW(x, y)ϕdxas t&0
(in particular the integral on the right-hand side exists);
(iv) ψH1(Ω).
20
Remark 4.2. Assumption (A1) (iii) and the continuity property from (A1) (ii) are
satisfied if, e.g., the following growth condition holds: There exist constants , C > 0and
functions sL1(Ω) and rL(2)0(Ω) such that for all xand yR:
|W(x, y)| ≤ C|y|2+s(x),
|yW(x, y)| ≤ C|y|21+r(x).
The assumptions in (A1) in combination with the direct method in the calculus of
variations imply unique solvability of the variational inequality fulfilled by the minimisers
of E(Ω,·).
Lemma 4.3. Under Assumption (A1) the energy (38) admits for each Lipschitz domain
Da unique minimum u(depending on ) on Kψ(Ω) which is given as the unique
solution of
uKψ(Ω) and ϕKψ(Ω) :
Z
u· ∇(ϕu) + λu(ϕu) + w(x, u)(ϕu) dx0,(40)
where
w(x, y) := yW(x, y).
In the sequel we will treat the variational inequality (40) by making use of the
transformation for the state variable and its test-function:
y:= uψand ˜ϕ:= ϕψ.
The variation inequality becomes a problem involving the standard obstacle set
K(Ω) := ϕH1(Ω) : ϕ0 a.e. on Ω.
Substituting above tranformation into (40) we obtain the following variational inequality:
yK(Ω) and ϕK(Ω) :
Z
y· ∇(ϕy) + λy(ϕy) + w(x, y +ψ)(ϕy) dx
Z
ψ· ∇(ϕy) + λψ(ϕy) dx
(41)
Hence it will suffice to investigate the solution yto deduce properties of the function u.
4.2 Perturbed problem and sensitivity estimates
In this subsection we prove a shape sensitivity result for the variational inequality (41).
In what follows let us denote by Φtthe flow generated by a vector field XC1
c(D, Rd).
For Ω Ddenote by Ωt:= Φt(Ω), t0, the perturbed domains (see Subsection 2.3 for
more details).
21
Perturbed problem
The solution ytH1(Ωt) to the perturbed variational inequality to (41) satisfies
ytK(Ωt) and ϕK(Ωt) :
Zt
yt· ∇(ϕyt) + λyt(ϕyt) + wt(x, yt+ψt)(ϕyt) dx
Zt
ψt· ∇(ϕyt) + λψt(ϕyt) dx.
(42)
We will sometimes write yt(X) = ytto emphasise the dependence on X. Please note
that in general y0(X) = yt(X) for all t0 and for all vector fields XC1
c(D, R2) with
the property X·n= 0 on Ω. This implication will be used in the forthcoming Lemma
4.14. Throughout this work we will adopt the following abbreviations:
wt
X(x, ϕ) := wtt(x), ϕ), W t
X(x, ϕ) := Wtt(x), ϕ), ψt
X:= ψtΦt,
A(t) := ξ(t)(Φt)1(Φt)T, ξ(t) := det Φt, yt:= ytΦt
(43)
and (for t= 0)
ψ(x) := ψ(x), w(x, ϕ) := w0
X(x, ϕ).
From Lemma 2.9 we can directly infer the following convergences and estimates
Lemma 4.4. Let XC1
c(D, Rd)be given. Then it holds:
(i) the convergences as t&0:
A(t)I
tA0(0) = div(X)I∂X (X )Tstrongly in C(D, Rd,d),(44a)
ξ(t)1
tξ0(0) = div(X)strongly in C(D); (44b)
(ii) there is a constant t>0such that
t[0, t],xD, ζRd, A(t, x)ζ·ζ1/2|ζ|2,
t[0, t],xD, ξ(t, x)1/2.
Performing a change of variables and using (y)Φt= (Φt)T(yΦt) it is easy
to check that the transported function yt(which is defined on Ω) satisfies the relation
ytK(Ω) and ϕK(Ω) :
Z
A(t)yt· ∇(ϕyt) + ξ(t)λyt(ϕyt) + ξ(t)wt
X(x, yt+ψt
X)(ϕyt) dx
Z
A(t)ψt
X· ∇(ϕyt)ξ(t)λψt
X(ϕyt) dx.
(45)
22
For later usage let us introduce the bilinear form
at(v, w) := Z
A(t)v· ∇w+ξ(t)λvw dx,
the operator At:Kψ(Ω) H1(Ω)by
hAt(v), wiH1(Ω) := at(v, w) + Z
ξ(t)wt
X(x, v)wdx(46)
and the “shifted” operator ˜
At:K(Ω) H1(Ω)by
˜
At(v) := At(v+ψt
X).(47)
By making use of this notation the variational inequality (45) can be recasted as
ytK(Ω) and h˜
At(yt), ϕ ytiH10 for all ϕK(Ω).(48)
In the following we also write A:= A0and ˜
A:= ˜
A0.
Sensitivity estimate
Our goal is to apply Theorem 3.3 designed for abstract operators. For this reason we
make the following assumption in addition to (A1):
Assumption (A2)
(i) XC1
c(D, Rd),c > 0,t[0, τ ],χH1(Ω),
kwt
X(·, χ)w(·, χ)kL(2)0(Ω) ct;
(ii) XC1
c(D, Rd),c > 0,t[0, τ ],χ1, χ2H1(Ω),
kwt
X(·, χ1)wt
X(·, χ2)kL(2)0(Ω) ckχ1χ2kH1(Ω);
(iii) XC1
c(D, Rd),c > 0,t[0, τ ],
kψt
XψkH1(Ω) ct.
We are now in the position to prove the following sensitivity result:
Proposition 4.5. Let the Assumptions (A1)-(A2) be satisfied. Then the family of op-
erators (˜
At)defined by (47) fulfills
(i) α > 0,t>0,t[0, t],v, w K(Ω),
αkvwk2
H1(Ω) ≤ h ˜
At(v)˜
At(w), v wi; (49)
23
(ii) vK(Ω),c > 0,t>0,t[0, t],wK(Ω),
|h ˜
At(v)˜
A(v), v wi| ≤ ctkvwkH1(Ω).(50)
Proof. To (i): We first show the monotonicity estimate (49). With the help of Lemma 4.4
(ii) and monotonicity of wt
Xin the second variable (see Assumption (A1) (i)) we obtain
for all v, w H1(Ω) and all small t0
1
2Z
|∇(vw)|2+λ|vw|2dx
at(vw, v w)
+Z
ξ(t)wt
X(x, v +ψt
X)wt
X(x, w +ψt
X)(v+ψt
X)(w+ψt
X)dx
(51)
Thus (49) is shown.
To (ii): Let us fix vH1(Ω). Then by applying H¨older inequality, Sobolev embed-
dings and the assumptions in (A2) we find for all wH1(Ω)
h˜
At(v)˜
A(v), v wi
Z
(A(t)I)v· ∇(vw) dx
| {z }
≤kA(t)IkLk∇vkL2k∇(vw)kL2
+Z
λ(ξ(t)1)v(vw) dx
| {z }
λkξ(t)1kLkvkL2kvwkL2
+Z
(A(t)ψt
X− ∇ψ)· ∇(vw) + λ(ξ(t)ψt
Xψ)(vw) dx
| {z }
kA(t)IkLk∇ψt
XkL2+k∇ψt
X−∇ψkL2+λkξ(t)1kLkψt
XkL2+λkψt
XψkL2kvwkH1
+Z
(ξ(t)1)wt
X(x, v +ψt
X)(vw)dx
| {z }
≤kξ(t)1kLkwt
X(x,v+ψt
X)kL(2)0kvwkH1
+Z
(wt
X(x, v +ψt
X)wt
X(x, v +ψ))(vw) dx.
| {z }
≤kwt
X(x,v+ψt
X)wt(x,v+ψ)kL(2)0kvwkH1≤ kψt
XψkH1kvwkH1
+Z
(wt
X(x, v +ψ)w(x, v +ψ))(vw) dx.
| {z }
≤kwt
X(x,v+ψ)w(x,v+ψ)kL(2)0kvwkH1ctkvwkH1
Taking Lemma 4.4 into account and using Young’s inequality, we obtain the assertion.
The desired Lipschitz estimate immediately follows from Theorem 3.3 since Propo-
sition 4.5 proves that Assumption (O1) are satisfied for p= 2.
Corollary 4.6. Under the assumption of Proposition 4.5 there exist t>0and c > 0
such that
kytykH1(Ω) ct for all t[0, t].
24
4.3 Limiting system for the transformed material derivative
In Corollary 4.6 we have established a Lipschitz estimate for the mapping t7→ yt. In
this section we are going to prove that there is a unique element ˙yin H1(Ω) – called
the material derivative – such that (yty)/t converges strongly to ˙yin H1(Ω) which is
uniquely determined by a variational inequality.
In order to derive the differentiability of ytwe impose the additional assumptions to
(A1) and (A2):
Assumption (A3)
(i) w(x, ·)is of class C1(R)for all x;
(ii) for all XC1
c(D, Rd), there exists a function ˙wX: ×RRsuch that for all
ϕnϕstrongly in H1(Ω) we have ˙wX(·, ϕ)L(2)0(Ω) and for all tn&0
wtn
X(·, ϕn)w(·, ϕn)
tn
˙wX(·, ϕ)strongly in L(2)0(Ω) as n→ ∞;
(iii) for any given sequences ϕnϕin H1(Ω) and tn&0with (ϕnϕ)/tn z weakly
in H1(Ω):
w(·, ϕn)w(·, ϕ)
tn
yw(·, ϕ)zstrongly in L(2)0(Ω) as n→ ∞;
(iv) for all XC1
c(D, Rd)there exists a function ˙
ψXH1(Ω) such that
ψt
Xψ
t˙
ψXstrongly in H1(Ω) as t&0.
Remark 4.7. (i) Property (iii) from Assumption (A3) is satisfied if, e.g., there exist
a constant C > 0and a function sL21
22
(Ω) such that for all xand yR:
|yw(x, y)| ≤ C|y|α+s(x)
with the exponent α:= 2(21)
22. The constant αis chosen such that the function
x7→ yw(x, ϕ(x))z(x)and x7→ f0(ϕ(x))z(x)are in L(2)0(Ω) for given ϕ, z
H1(Ω).
(ii) A useful consequence of properties (ii) and (iii) is the following continuity
wtn
X(·, ϕn)w(·, ϕ)strongly in L(2)0(Ω) as n→ ∞.
for all ϕnϕstrongly in H1(Ω) and tn&0.
25
(iii) Let XC1
c(D, Rd)be given. Then we have by using property (iv) from Assumption
(A3)
A(t)ψt
X+ψ
t A0(0)ψ− ∇ ˙
ψXstrongly in L2(Ω,Rd),
ξ(t)ψt
X+ψ
t ξ0(0)ψ˙
ψXstrongly in L2(Ω,Rd).
We are now well-prepared for the derivation of the material derivative.
Theorem 4.8. Let (A1)-(A3) be satisfied. The weak material derivative ˙yof t7→ yt
exists in all directions XC1
c(D, Rd)and is characterised as the unique solution of the
following variational inequality
(˙y˜
Sy(K)and ϕ˜
Sy(K) :
h˜
A(y) ˙y , ϕ ˙yiH1≥ −h ˜
A0(y), ϕ ˙yiH1,(52)
where ˜
Sy(K)denotes the closed and convex cone
˜
Sy(K) = Ty(K)kern(˜
A(y)).(53)
The functional derivatives ˜
Aand ˜
A0are given by
h˜
A(y) ˙y , ϕi=a( ˙y+˙
ψX, ϕ) + Z
yw(x, y +ψ) ˙dx, (54)
h˜
A0(y), ϕi=Z
A0(0)y· ∇ϕ+ξ0(0)λy +w(x, y +ψ)ϕdx
+Z
˙wX(x, y +ψ)ϕ+yw(x, y +ψ)˙
ψXϕdx
+Z
A0(0)ψ· ∇ϕ+ξ0(0)λψϕ dx.
(55)
Proof.
Existence of ˙y:We want to apply Theorem 3.5. For this we need to check Assumption
(O2). To this end we notice that by Corollary 4.6 ytnustrongly and (ytny)/tn z
weakly in H1(Ω) for a suitable subsequence tn&0.
We check (O2) (ii): Let vn v be a given weakly convergence sequence in H1(Ω).
Then
*˜
Atn(ytn)˜
A(ytn)
tn
, vn+
=Z
A(tn)I
tn
ytn· ∇vndx
| {z }
RA0(0)y·∇vdx
+Z
ξ(tn)1
tnλytn+wtn
X(x, ytn+ψtn
X)vndx
| {z }
Rξ0(0)(λy+w(x,y+ψ))vdxby Remark 4.7 (ii)
26
+Z
wtn
X(x, ytn+ψtn
X)w(x, ytn+ψtn
X)
tn
vndx
| {z }
R˙wX(x,y+ψ)vdxby Assumption (A3) (ii) and (iv)
+Z
w(x, ytn+ψtn
X)w(x, y +ψ)
tn
vndxZ
w(x, ytn+ψ)w(x, y +ψ)
tn
vndx
| {z }
Ryw(x,y+ψ)(z+˙
ψX)vdxRyw(x,y+ψ)zv dx=Ryw(x,y+ψ)˙
ψXvdxby (A3) (iii)-(iv)
+Z
A(tn)I
tn
ψtn
X· ∇vn+ξ(tn)1
tn
ψtn
Xvndx
| {z }
→− RA0(0)ψ·∇v+ξ0(0)ψv dx
.
We check (O2) (iii):
*˜
A(ytn)˜
A(y)
tn
,ytny
tn+
=Zytny
tn2+λytny
tn2dx
| {z }
lim inf R|∇z|2+λ|z|2dx
+Z
w(x, ytn+ψ)w(x, y +ψ)
tn
ytny
tn
dx
| {z }
Ryw(x,y+ψ)|z|2by Assumption (A3) (iii)
and for all ϕnϕstrongly in H1(Ω):
*˜
A(ytn)˜
A(y)
tn
, ϕn+
=Z
ytny
tn
· ∇ϕn+λytny
tn
ϕndx
| {z }
Rz·∇ϕ+λzϕ dx
+Z
w(x, ytn+ψ)w(x, y +ψ)
tn
ϕndx
| {z }
Ryw(x,y+ψ)by Assumption (A3) (iii)
.
Property (O2) (i) follows from the above calculations.
Uniqueness of ˙y:Assume two solutions ˙yand ˙zfor (52). Testing their variational
inequalities with ˙zand ˙y, respectively, and adding the result yields
h˜
A(y) ˙y˜
A(y) ˙z, ˙y˙zi ≤ 0.
The left-hand side calculates as
h˜
A(y) ˙y˜
A(y) ˙z, ˙y˙zi
=a( ˙y˙z, ˙y˙z) + Z
yw(x, y +ψ)|˙y˙z|2dx.
Due to the convexity assumption in (A1) (i) we find yw0 and see that
a( ˙y˙z, ˙y˙z)0.
We obtain ˙y˙z= 0.
27
By exploiting the specific structure of ˜
Atand Assumption (A3) we can even show
that the strong material derivative exists.
Corollary 4.9. We have for all XC1
c(D, Rd)
yt
Xy
t˙yXstrongly in H1(Ω).(56)
Proof. We test the variational inequality (48) with ϕ=ytand for t= 0 with ϕ=y.
Adding both inequalities yields
h˜
At(yt)˜
A(y), ytyi ≤ 0.
Dividing by t2and rearranging the terms we obtain by setting zt:= (yty)/t
a(zt, zt)
Z
A(t)I
tyt· ∇ztdxZ
λξ(t)1
tytztdx
Zξ(t)1
twt
X(x, yt+ψt
X) + wt
X(x, yt+ψt
X)w(x, yt+ψt
X)
tztdx
Z
w(x, yt+ψt
X)w(x, y +ψ)
tztdx
Z
A(t)ψt
X− ∇ψ
t· ∇ztdxZ
λξ(t)ψt
Xψ
tztdx
=: B(t).
(57)
The known convergence properties shows as t&0 for a subsequence
B(t) −h ˜
A0(y),˙yi − Z
yw(x, y +ψ)|˙y|2dxZ
˙
ψX· ∇ ˙ydxZ
λ˙
ψX˙ydx
| {z }
=:B(0)
.
However testing (52) with ϕ= 2 ˙y˜
Sy(K) we also obtain h˜
A(y) ˙y, ˙yiH1≥ −h ˜
A0(y),˙yiH1
which is precisely
a( ˙y, ˙y)B(0).
All in all we get
lim sup
t&0
a(zt, zt)lim sup
t&0
B(t) = B(0) a( ˙y, ˙y).(58)
The weak convergence zt˙yin H1(Ω) implies lim inft&0a(zt, zt)a( ˙y, ˙y). Together
with (58) this gives a(zt, zt)a( ˙y, ˙y) as t&0. This finishes the proof.
Remark 4.10. If we assume that
˙wX(x, y) := T0(x, y)·X(x) + T1(x, y) : X(x) (59)
28
for functions T0(·,·):Ω×RRdand T1(·,·):Ω×RRd×dwe may rewrite the
variational inequality in (52) by using Lemma 4.4 as
a( ˙y , ϕ ˙y) + Z
yw(x, y +ψ) ˙y(ϕ˙y) dx
Z
L1(x, y +ψ;ϕ˙y) : X +L0(x, y +ψ;ϕ˙y)·Xdx
a(˙
ψ, ϕ ˙y) + Z
yw(x, y +ψ)˙
ψ(ϕ˙y) dx,
where we use the abbreviations
L1(x, y +ψ;ϕ) := (y+ψ)· ∇ϕ+λ(y+ψ) + w(x, y +ψ)ϕI
+ϕ⊗ ∇(y+ψ) + (ψ+y)⊗ ∇ϕT1(x, y +ψ)ϕ,
L0(x, y +ψ;ϕ) := T0(x, y +ψ)ϕ.
4.4 Limiting system for the material derivative
So far we have derived an equation for ˙y. Since we are interested in the original problem
(40), we may now use Theorem 4.8 and the transformation y=uψto obtain the
material derivative equation for (40). It is clear that ˙y= ˙u˙
ψXand we conclude with
the following result:
Corollary 4.11. Under the assumptions (A1)-(A3) the material deriative ˙u= ˙u(X)of
solutions of the perturbed problem to (40) in direction XC1
c(D, Rd)exists and is given
as the solution of the following variational inequality:
˙uSX
u(Kψ)and ϕSX
u(Kψ) :
a( ˙u, ϕ ˙u) + Z
yw(x, u) ˙u(ϕ˙u) dx
Z
A0(0)u· ∇(ϕ˙u) + ξ0(0)λu +w(x, u)(ϕ˙u) dx
Z
˙wX(x, u)(ϕ˙u) dx,
(60)
where
SX
u(Kψ) := Tu(Kψ)kern(A(u)) + ˙
ψX.
In particular under the additional assumption in Remark 4.10
a( ˙u, ϕ ˙u) + Z
yw(x, u) ˙u(ϕ˙u) dx
Z
L1(x, u;ϕ˙u) : ∂X +L0(x, u;ϕ˙u)·Xdx.
29
Proof. We obtain from Theorem 4.8 that ˙u˜
Sy(K) + ˙
ψXand for all ϕ˜
Sy(K) + ˙
ψX:
h˜
A(uψ)( ˙u˙
ψX), ϕ ˙uiH1≥ −h ˜
A0(uψ), ϕ ˙uiH1,
which is precisely the inequality in (60).
It remains to show SX
u(Kψ) = ˜
Sy(K)+ ˙
ψXwhich is equivalent to Tu(Kψ)kern(A(u)) =
Ty(K)kern(˜
A(y)). Indeed, by definition (47) we find
kern(A(u)) = kern(˜
A(y))
as well as by (1)-(3)
Tu(Kψ) = Tuψ(K) = Ty(K)
Note that we get the following characterisation of SX
uby using Lemma 2.5 and the
definition in (53):
ϕSX
u(Kψ)ϕ˙
ψXTu(Kψ)kern(A(u))
(ϕH1(Ω) with ϕ˙
ψXq.e. on {u=ψ},
hA(u), ϕ ˙
ψXi= 0.
Moreover under an additional assumptions we obtain the subsequent translation prop-
erty:
Lemma 4.12. Suppose that u, ψ H2(Ω) and let ζH1(Ω) be with
˜
ζ= 0 q.e. on the coincidence set {xΩ : ˜u(x) = ˜
ψ(x)},
where ˜
ζ,˜uand ˜
ψdenote quasi-continuous representatives for ζ,uand ψ. Then we have
±ζTu(Kψ)kern(A(u)).
In particular
ζ+SX
u(Kψ) = SX
u(Kψ).(61)
Proof. It is clear from the assumption that ±˜
ζ= 0 q.e. on the coincidence set {u=ψ}.
Thus ±ζTu(Kψ). Furthermore y=uψsatisfies the variational inequality (see (48)
with t= 0)
h˜
A(y), ϕ yi ≥ 0 for all ϕH1(Ω) and ϕ0 a.e. in Ω.
From the H2(Ω)-regularity of uand ψwe deduce that (in a pointwise formulation)
˜
A(y) = 0 a.e. in {xΩ : u(x)< ψ(x)}. In particular we see that
h˜
A(y), ϕi= 0 for all ϕH1(Ω) with {xΩ : ϕ(x)=0}⊇{xΩ : u(x) = ψ(x)}a.e.
Testing with ϕ=±ζyields ±ζkern(˜
A(y)) = kern(A(u)).
Finally, ζTu(Kψ)kern(A(u)) implies ζ+SX
u(Kψ)SX
u(Kψ), and ζTu(Kψ)
kern(A(u)) implies ζ+SX
u(Kψ)SX
u(Kψ).
30
In the following ψis referred to as a static obstacle if there exists a fixed function
ψH2(D) such that ψ˜
=ψ|˜
for all Lipschitz domains ˜
D.
Remark 4.13. Let XC1
c(D, Rd). Suppose that ψis a static obstacle, uH2(Ω)
and {X=0} ⊇ {˜u=˜
ψ}q.e. in . Then ˙
ψX=ψ·Xand the assumptions from
Lemma 4.12 are satisfied for ζ=˙
ψXand we obtain
±˙
ψXTu(Kψ)kern(A(u)).
In particular
SX
u(Kψ) = Tu(Kψ)kern(A(u))
and
ϕSX
u(Kψ)(ϕH1(Ω) with ϕ0q.e. on {u=ψ},
hA(u), ϕi= 0.
4.5 Limiting system for the state-shape derivative
The state shape derivative of uat Ω in direction XC1
c(D, Rd) is defined by
u0=u0(X) := ˙uXuon Ω (62)
where usolves (40), ˙usolves (60) and Xu:= u·X. It is clear that u0L2(Ω).
Thus in general the state shape derivative is less regular than the material derivative.
Another important observation is that the boundary conditions imposed on ˙uon Ω are
not carried over to u0.
Lemma 4.14. Let XC1
c(D, Rd)be a vector field satisfying X·n= 0 on . Then
the state shape derivative vanishes identically, that is, u0(X)=0a.e. on .
Proof. The X-flow Φtleaves the domain Ω unchanged, i.e., Φt(Ω) = Ω for all t[0, τ ].
Consequently, ut=u(Ωt) = u(Ω) = uand thus ut=utΦt=uΦtfor all t[0, τ ].
Hence by Lemma 2.9 (ii) we may calculate the material derivative ˙uas
utu
t=uΦtu
tXustrongly in L2(Ω).
Thus ˙u=Xuand consequently u0= 0.
Now we are prepared to prove the main result of this section which gives a simplified
variational inequality for the state-shape derivative u0under certain conditions. To
derive this result we will assume the enhanced regularity uH2(Ω). Preliminarily we
observe from Corollary 4.11 and by using the relation (62) that
u0ˆ
SX
u(Kψ) and ϕˆ
SX
u(Kψ) :
a(u0, ϕ u0) + Z
yw(x, u)u0(ϕu0) dx
Z
A0(0)u· ∇(ϕu0) + ξ0(0)λu +w(x, u)(ϕu0) dx
a(Xu, ϕ u0)Z
yw(x, u)Xu(ϕu0) dx,
(63)
31
where ˆ
SX
u(Kψ) := Tu(Kψ)kern(A(u)) + ˙
ψXXu.
We notice that in general the cone b
SX(K) depend on the vector field X. In the case of
a static obstacle problem (see Remark 4.1 (i)) we derive the following result:
Theorem 4.15. Suppose that (A1)-(A3), (59) and uH2(Ω) hold. Furthermore let
ψbe a static obstacle function.
Then ˆ
SX
uis independent of XC1
c(D, Rd)with
ˆ
SX
u(Kψ) = Tu(Kψ)kern(A(u)) =: Su(Kψ) (64)
and the state shape derivative is the unique solution of
u0Su(Kψ)and ϕSu(Kψ) :
a(u0, ϕ u0) + Z
yw(x, u)u0(ϕu0) dx
ZΓ
S1(x, u;ϕu0)n·n(X·n) ds,
(65)
where
S1(x, u;ϕ) := L1(x, u;ϕ)− ∇u⊗ ∇ϕ.
with L1from Remark 4.10.
Proof. By using the assumption ˙
ψX=ψ·Xwe find on the coincidence set {u=ψ}
(here we resort to quasi-continuous representants):
˙
ψXXu=˙
ψXXψ= 0.
Lemma 4.12 applied to ζ=˙
ψXXuyields ±(˙
ψXXu)Su(Kψ) and therefore (64).
Furthermore by using the notation in Remark 4.10 and the identity (note that u
H2(Ω) by assumption)
(Xu)=(∂X )T(u)+(2u)X,
the variational inequality in (63) rewrites to u0Su(Kψ) and for all ϕSu(Kψ):
a(u0, ϕ u0)+ Z
yw(x, u)u0(ϕu0) dx
Z
S1(x, u;ϕu0) : ∂X +S0(x, u;ϕu0)·Xdx,
(66)
where
S0(x, u, ϕ) := L0(x, u, ϕ)yw(x, u)ϕu(2u)ϕ,
S1(x, u, ϕ) := L1(x, u, ϕ)− ∇u⊗ ∇ϕ.
32
Picking any vector field XC1
c(D, Rd) with X·n= 0 on Γ we know from Lemma 4.14
that u0(±X) = 0 and it follows from (66)
Z
S1(x, u; ˜ϕ) : ∂X +S0(x, u; ˜ϕ)·Xdx= 0 (67)
for all ˜ϕSu(Kψ). Then integrating by parts in (67) shows the pointwise identity
div(S1(x, u(x); ˜ϕ(x))) + S0(x, u(x); ˜ϕ(x)) = 0 a.e. on Ω.(68)
Now for an arbitrary XC1
c(D, Rd) and ˜ϕSu(Kψ) we consider the additive splitting
X=Xn+XTfor Xn, XTC1
c(D, Rd) such that Xn=n(X·n) and XT=Xn(X·n)
on Γ. Then XT·n= 0 on Γ and we get
Z
S1(x, u; ˜ϕ) : X +S0(x, u; ˜ϕ)·Xdx
=Z
S1(x, u; ˜ϕ) : XT+S0(x, u; ˜ϕ)·XTdx
| {z }
=0 by (67)
+Z
S1(x, u; ˜ϕ) : Xn+S0(x, u; ˜ϕ)·Xndx
| {z }
partial integration and (68)
=ZΓ
S1(x, u; ˜ϕ)n·Xnds.
(69)
We may test (69) with ˜ϕ=u0since u0Su(Kψ). Then multiplying the resulting identity
with 1 and exploiting linearity of S0and S1with respect to ϕyields
Z
S1(x, u;u0) : ∂X +S0(x, u;u0)·Xdx=ZΓ
S1(x, u;u0)n·Xnds. (70)
Now we find by letting ˜ϕ=ϕSu(Kψ) be arbitrary, adding (69) and (70), and again
exploiting linearity
Z
S1(x, u;ϕu0) : ∂X +S0(x, u;ϕu0)·Xdx=ZΓ
S1(x, u;ϕu0)n·Xnds.
In combination with (66) we obtain (65). Uniqueness of u0is implied by uniqueness of
˙y(see Theorem 4.8).
It is readily checked that
S1(x, u;ϕ)n·n=−∇Γu· ∇Γϕλu +w(x, u)ϕfor all ϕSu(Kψ).
Thus we conclude this section with an explicit formula for the shape derivative in the
case of a static obstacle.
33
Corollary 4.16. Under the assumption of Theorem 4.15 the shape derivative u0is the
unique solution of the following variational inequality:
u0Su(Kψ),a(u0, ϕ u0) + Z
yw(x, u)u0(ϕu0) dx
ZΓΓu· ∇Γ(ϕu0) + λu +w(x, u)(X·n)(ϕu0) ds
for all ϕSu(Kψ).
4.6 Eulerian semi-derivative of certain shape functions
We adopt the notation from Subsection 2.3 and denote by J: Ξ Ra shape function.
Application of Corollary 4.11, Lemma 2.8 and the chain rule yield the following result:
Corollary 4.17. Let (A1)-(A3) be satisfied and let Ξbe a Lipschitz domain, X
C1
c(D, Rd)and Φt: Ω tbe the associated flow. Suppose that for all small t > 0
J(Ωt) = Jt, ut),
where
J=J, u) : C0,1(Ω; Rd)×H1(Ω) R
is assumed to be a Fr´echet differentiable functional and utH1(Ω) the transported state
ut=utΦtwith the unique solution utof (40) on t.
Then the Eulerian semi-derivative exists and is given as
dJ(Ω)(X) = hdΦJ(I d, u0), XiC0,1(Ω;Rd)+hduJ(Id, u0),˙uXiH1(Ω),
where ˙uXdenotes the unique solution of (60).
In particular dJ(Ω)(·)is positively 1-homogeneous.
5 Applications to damage phase field models
In this section we investigate shape optimization problems for a coupled inclusion/pde
system describing damage processes in linear elastic materials. Our aim is to apply
the abstract results from Section 4 designed for semilinear variational inequalities with
dynamic obstacles to such concrete application scenarios. In this way we demonstrate
how necessary optimality conditions for shape problems can be derived for relevant
engineering tasks.
5.1 Physical model
The physical model under consideration was derived in [9] and is described in the time-
continuous setting by the following relations:
utt divC(χ)ε(u)=`,(71a)
34
0∂I(−∞,0] (χt) + χtχ+1
2C0(χ)ε(u) : ε(u) + f0(χ),(71b)
with the damage-dependent stiffness tensor Cand the damage potential function f.
The variable udenotes the displacement field, ε(u) := 1
2(u+ (u)T) the linearised
strain tensor and χis an internal variable (a so-called phase field variable) indicating
the degree of damage. In terms of damage mechanics χis interpreted as the density of
micro-defects and is therefore valued in the unit interval (cf. [17]). In this spirit we may
use the following interpretation:
χ(x) =
1no damage in x,
(0,1) partial damage in x,
0maximal damage in x.
The system is supplemented with initial-time values for χ,uand ut, Dirichlet boundary
condition for uand homogeneous Neumann boundary condition for χ. The governing
state system (71) can be derived by balance equations and suitable constitutive relations
such that the laws of thermodynamics from continuum physics are fulfilled. We refer to
[9] for more details on the derivation of the model.
A main feature of the evolution system (71) is the uni-directionality constraint χt0
enforced by the subdifferential ∂I(−∞,0](χt). This leads to non-smooth/switching be-
haviour of the evolution law by noticing that (71b) rewrites as
χt=(d, if d0,
0,if d > 0with the driving force d= ∆χ1
2C0(χ)ε(u) : ε(u)g0(χ).
A weak formulation of (71) and existence of weak solution can be found in [11] with
minor adaption. Existence and uniqueness results for strong solutions for the above
system with higher-order viscous terms are established in [7]. For the analysis of quasi-
linear variants of (71) and for rate-independent as well as rate-dependent cases, we refer
to [16] and the references therein.
The following remark justifies that the phase field variable χtakes only admissible
values provided H1(0, T ;H1(Ω))-regularity and mild growth assumptions on Cand g. In
that case it is not