In this paper we study the homogenization of a class of energies concentrated on lines. In dimension (i.e., in codimension ) the problem reduces to the homogenization of partition energies studied by [1]. There, the key tool is the representation of partitions in terms of functions with values in a discrete set. In our general case the key ingredient is the representation of closed loops with discrete multiplicity either as divergence-free matrix-valued measures supported on curves or with -currents with multiplicity in a lattice. In the dimensional case the main motivation for the analysis of this class of energies is the study of line defects in crystals, the so called dislocations.
Dedicated to Umberto Mosco in the occasion of his 80th birthday
1. Introduction
In the present work we consider energies concentrated on lines of the form
(1.1)
where is a -rectifiable set in given by the union of closed loops, the function is a vector-valued multiplicity, constant on each closed loop of , and is the tangent vector defined -a.e. on .
The main result of the paper concerns the homogenization of energies of this class. This can be expressed as the characterization of the limit of scaled energies
(1.2)
as when is periodic in the first variable.
The main motivation for the study of the energies above comes from the analysis of dislocations in crystals. Dislocations are defects in the crystalline structure of metals that are crucial for the understanding of plastic behaviours.
At a continuum level they can be interpreted as line singularities carrying an energy of the form above, where the multiplicity is the so-called Burgers vector which belongs to a lattice (which is a material property), that we can assume to be . The function represents the line tension energy density that can be computed à la Volterra (see e.g. [4]) using continuum elasticity.
The asymptotic behaviour of energies of the form (1.2) can be expressed in terms the computation of their -limit with respect to a suitable convergence. In a two-dimensional setting this has been carried over in [1] in a BV setting. Indeed, in that case the system of lines can be interpreted as the set of the interfaces of a Caccioppoli partition, or, equivalently, the set of discontinuity points of a BV-function taking values in a discrete set. In that functional setting energies of the form above can be analysed as -limits with respect to the BV-convergence.
In higher dimension, instead of identifying systems of loops with partitions, they can be interpreted as divergence-free measures or -rectifiable currents without boundary. Throughout the paper we will make use of those two standpoints interchangeably, taking advantage of the possibility of choosing the most suited of the two in technical points. The corresponding equivalent notions of convergence makes it possible to study energies (1.2) in terms of -convergence.
The main result of the paper is that, under suitable growth assumptions of the -limit of energies (1.2) as exists and can be written as
(1.3)
for a suitable function . Furthermore, this function can be characterized by an asymptotic formula.
In order to prove that the function given by the asymptotic homogenization formula gives a lower bound for the -limit, following [6], we make use of Fonseca and Müller blow-up technique [10]. This method, originally introduced to deal with relaxation problems, works nicely with homogenization problems as well. Here, as in [5], it will be useful to rephrase the problem in terms of closed 1-rectifiable currents, as this allows an easy treatment of the possibility of fixing boundary conditions, which is a technical point necessary to carry over the blow-up method. In order to prove the upper bound we proceed by density using the homogenization formula explicit construction of recovery sequences. We will need some results on divergence-free measures for which we refer to [5] instead.
2. Formulation of the problem
Let be an open set with Lipschitz boundary. Let be the energy density of the energy (1.3). We assume that ia a Borel function and satisfies
(2.1)
A convenient framework to represent the set of admissible configurations is the one of divergence-free matrix-valued measures or alternatively of 1-rectifiable currents without boundary.
Representation with measures:
Following [5], we will denote by the set of divergence-free (in the sense of distribution) measures of the form
where is a -integrable function, a 1-rectifiable set and its tangent vector defined -a.e. on .
The divergence-free conditions reads as
(2.2)
for all .
So that, for any the energy in (1.3) is denoted by
(2.3)
In particular by the growth condition (2.1)
we deduce that the energy
and it is coercive with respect to the weak∗ convergence of measures. Indeed a sequence with bounded energy is in particular bounded in the total variation, and therefore it is compact in the weak∗ converge.
The fact that is closed with respect to the weak∗ convergence can be seen, as already mentioned, in a very efficient way by using the approach of geometric measure theory, i.e., the setting of rectifiable currents à la Federer and Fleming extended to the case of currents with vector multiplicity.
Representation with integral -rectifiable currents:
We denote by the set of -valued 1-rectifiable currents. Let , , be as before, then is a functional on the space of smooth compactly supported 1-forms that admits the following representation
(2.4)
We recall that the boundary of a 1-rectifiable current is the 0-current for all ; finally a current is closed or without boundary if . We denote by the set of currents in which are closed.
Now there is a -to- correspondence between measures in and currents in with no boundary. Indeed it is immediate to see that the divergence free condition (2.2) translates in the condition of having zero boundary for the corresponding currents.
Therefore for any we denote by the corresponding current in and for any we denote by the corresponding measure in .
In particular given the mass of the associated current is given by
Moreover the weak∗ convergence of a sequence of measure translates exactly in the weak∗ convergence of the corresponding currents , i.e.,
The advantage of using the language of rectifiable currents relies on a rich theory which guarantees a structure result which allows to characterize all currents in as a countable family of Lipschitz closed loops with constant multiplicity, a compactness result for sequence with bounded mass in (see [9]), and a good approximation of currents (and therefore of the corresponding measures) with polyhedral currents (i.e. currents in supported on polyhedral curves). These results in the formulation that is needed here are recalled in the Appendix.
Our results will be mainly stated using the more familiar notation of measures, but throughout the paper, especially in the proofs, we will use the two notations interchangeably, making sure to highlight the advantages of one over the other.
3. The homogenization theorem
We now state the main result of this paper which concerns the homogenization of the energies concentrated on lines.
In the following is a bounded, open set with Lipschitz boundary and is a Borel function satisfying (2.1).
Additionally we assume that is -periodic in the first variable and we define
(3.1)
The main result of this paper is the characterization of the -limit of the functional as .
We stress that a -convergence result must be complemented by a compactness result: together they guarantee the convergence of minima. Here compactness in the class , as we noted, is a consequence of the divergence free constraint. This is the first point in which the already mentioned equivalence between the representation with measures and the one with 1-currents gives the its contribution.
Theorem 3.1(Compactness).
Let be a sequence of measures satisfying
for some , then there is a measure and a subsequence such that
Proof.
The result is an immediate consequence of the lower bound for the energy density given by (2.1). Indeed implies that
Then, up to a subsequence the sequence of measure weakly∗ converge to some matrix valued Radon measure . In order to conclude it is enough to show that the limit measure as the right structure, i.e., it is concentrated on a -rectifiable set and its density is of the form , in other words it belongs to .
This is an immediate consequence of the result of compactness of -rectifiable currents with integer multiplicity. Indeed the family of currents satisfy
and hence up to a subsequence it converges to a current in . Then and it is the weak∗ limit of the corresponding subsequence of .
∎
Remark 3.2.
This compactness result clarifies the importance of the divergence free condition for the measures . It is indeed easy to construct a sequence of measures supported on -rectifiable sets with integer multiplicities (equivalently a sequence of integer -currents) which converges to a measure which is not supported on curves, for instance which converges to the -dimensional Lebesgue measure. This can be done by taking a collection of many uniformly distributed short segments. In this example the corresponding current has large boundary and therefore the compactness result does not apply.
Moreover the above theorem sets the right topology in which we have to study the -limit of the .
Before stating the theorem, it is useful to introduce the following notation. For all choose a rotation with . Then, for every we define the rectangle
of height and a side , centred at the origin, and one side parallel to the direction . If the rectangle is centred in a point we denote it by .
Similarly we denote by the cube of side , and one side parallel to the direction , centred at , i.e., . If we drop the and write .
Theorem 3.3(Homogenization).
Assume that , satisfying (2.1), is -periodic in the first variable, then the functionals in 3.1 -converge as , with respect to the weak∗ convergence of measures, to the functional defined by
where
for all and the effective energy density is given by
(3.2)
Proof.
The proof of the lower bound is given in Subsection 3.2 and uses the characterization of the effective energy density by means of the asymptotic formula (studied in Section 3.1). The proof of the upper bound is instead given in Section 3.3.
∎
3.1. The cell problem formula
A key ingredient is the analysis of the cell problem formula in (3.2) which characterizes the effective energy.
Here and in the rest of the paper it is convenient to introduce the localised functionals for every Borel subset and a measure ,
(3.3)
In the next proposition we prove that the energy density is well defined through an asymptotic formula. Moreover we show that it is rather flexible and thanks to the periodicity of is not sensitive to the translations. As a consequence we will get the continuity of .
Proposition 3.4(Homogenization formula).
Let be a family of points in , with and as in Theorem 3.3. Then for all and the limit
exists and it is independent of . Therefore it coincides with .
Proof.
Let be the straight line parallel to passing through , i.e., and denote .
Note that, in general, the line may not intersect the set of points in .
Let . Fix and consider the family of equispaced points on with spacing . For each point consider the point such that
(3.4)
Let be a test measure for the minimum problem
(3.5)
such that
(3.6)
Figure 1.
Let be the set of indices such that and
Hence define the function and the measure
Note that the measure is nothing else than the measure obtained from translating the support from the cube to the cube , i.e.,
therefore by the choice of we have
From , we can obtain a divergence free measure in , by connecting through a segment each endpoint of on to . In doing so we have obtained a test measure for (see Figure 1). The conclusion follows if we prove
In fact, using we get
where we have used the choice of as in (3.6) and (3.4) to control the contribution of the segments in the measure . The conclusion follows taking the first and then the .
∎
Remark 3.5.
Once the existence of the limit is proved it is also easy to see that this is not only independent of the choice of but also uniform in . Indeed, by periodicity, we can assume that which is compact, and then we deduce the uniformity.
An important consequence of the above characterization of is its continuity.
Proposition 3.6(Continuity of ).
For all and let be as in Theorem 3.3. Then the following continuity property holds
with a constant that depends only on .
Proof.
We prove this property by means of the asymptotic formula. Suppose that . Fix and let be a test measure for the minimum problem in (3.5) (Proposition 1.12, with ), such that
(3.7)
From , we can obtain a test measure for the problem
proceeding as in Figure 2 (with support on the bold line), adding two segments to obtain the divergence free condition in for the measure and to get the right boundary condition.
Figure 2. The construction of the measure , for simplicity .
Then it is easy to see that
and using the estimate from above for and by construction of we have
We conclude taking the limit and by the arbitrariness of . To obtain the inequality with the opposite sign is enough to swap the roles of and .
∎
Finally we show that the asymptotic formula still holds if we replace the boundary condition with an approximate boundary condition. This point is essential in the proof of the lower bound.
To this aim we need the following technical lemma which allows one to modify the boundary value of a converging sequence of measures with a small change in the corresponding line energy.
Lemma 3.7.
Let be a sequence of divergence free measures in such that
(3.8)
and . Then, , there exist a sequence , and a sequence such that , where is such that and for every large enough (by we have denoted the interior of the cube ).
Proof.
Fix and two independent parameters and consider , the distance function from . We now slice the measure , with , through the function , Lipschitz continuous with .
Let and be the support and multiplicity of , respectively. Now the idea is to exploit the weak∗ convergence of to zero in order to find a for which is the boundary of weighted segments with small mass; this will allow us to construct the measure in the statement.
Even though this can done by hands using the measures in it is convenient also in this case to use the tools of currents which give a more general argument not confined to the case of dimensional objects.
We denote the current associated to and for every we consider the current obtained by slicing along the level sets of the function which we denote by
Since in , we have
, i.e.
for all .
For slicing of currents it holds
(3.9)
and
(3.10)
where denote the flat norm of the current (see the Appendix).
Since, by (3.8), the sequence is equi-bounded in mass, from (3.9) we get that, for a.e. , is finite. Hence, since is a 0-rectifiable current for almost every , we have
(3.11)
where the sum runs through a finite number of points , with multiplicity and positive orientation if exits from at (these oriented points together with their multiplicity give the boundary of ).
Moreover by (3.8), we can exploit the equivalence between weak∗ convergence and convergence in the flat norm [11, Theorem 31.2], and from the fact that to obtain the convergence to zero of the flat norm. Therefore we can find such that for all . Thus, by (3.10), for each , the sets
(3.12)
have positive -dimensional Lebesgue measure. In fact
hence
Now, for each choose a such that and the sum (3.11) runs through a finite number of points.
To get to the conclusion, we first show that the following minimum problems, well defined by our choice of , have solution. By definition, for every ,
Since is a -valued -rectifiable current, then and from
we can conclude that . It follows that, for every ,
It’s easily checked that, for all , the set over which we take the above infimum is not empty.
Hence , by means of the direct method, there exists a -rectifiable current that satisfies the following properties
(i)
;
(ii)
;
(iii)
.
Now
we conclude defining for and equal to otherwise.
∎
Remark 3.8.
By a scaling argument the result in Lemma 3.7 still hold in the domain .
Moreover by a diagonal argument under the assumptions of Lemma 3.7 one could show that, for every ,
there exists a sequence such that , with . Moreover, exploiting the construction of and its decomposition in mutually singular measures we also deduce that
(3.13)
A further diagonal argument produces a sequence such that , with and
(3.14)
Using the above lemma we finally prove the following proposition which is essentially the -convergence result when the limiting configuration is given by the measure , i.e., straight with constant multiplicity.
Proposition 3.9.
Let be a family of points in , with
and as in Theorem 3.3. For all and we define
Then
(3.15)
Proof.
In order to show one inequality we construct a sequence admissible for the definition of .
As above by the periodicity of the problem we can reduce to the case in which .
For any we consider a family of equispaced points with spacing on . Given , thanks to Remark 3.5, we can find large enough such that we find a test measure for the minimum problem in (3.5) with such that
(3.16)
In particular we have that . With this condition we can rescale and glue the measures and construct a sequence which is admissible for . Precisely, given we define the function and denote , , then the measure
is divergence free and satisfies as . Therefore using the notation we have
Thus
which gives one inequality.
In order to obtain the opposite inequality it is enough to observe that given and fixing a sequence admissible for the minimum problem , by means of Remark 3.8 and (3.14), we find a sequence satisfying such that
From the proof of the above proposition we also deduce that given and for all there exists a sequence such that and
3.2. Proof of the liminf inequality by blow-up
aa
We are ready to prove the Theorem 3.3. We start by proving the liminf inequality, i.e., we need to show that for all , and one has
Let , be as above; it’s not restrictive to suppose that is finite and, by compactness (Theorem 3.1), .
We head to the conclusion in three steps.
Step 1. (localization and decomposition)
For all we denote the positive measures
(3.18)
which is the energy density of the
localised functional
(3.19)
By the assumptions on , the sequence is equi-bounded hence, by compactness, there exists a positive Radon measure on such that, up to a subsequence, . Now consider the Radon-Nikodym decomposition of the measure with respect to the -dimensional Hausdorff measure restricted to , i.e.,
(3.20)
where is the density of the part of which is absolutely continuous with respect to and the singular part is denoted by (and it is a positive measure as well).
Step 2. (definition of the blow-up) Let be a Lebesgue point for with respect to . We can write
(3.21)
where and the last equality holds for -a.e by the Besicovitch-Marstrand-Mattila Theorem.
The Besicovitch derivation theorem ensures that -a.e is a Lebesgue point for with respect to , i.e., a Lebesgue point for ; moreover we can also assume that is a Lebesgue point for and . By definition of -rectifiability and approximate tangent space and then we can assume that -a.e satisfies the following property
(3.22)
with , in every open, bounded subset of ; in fact it follows from Theorem 3.13 that is the union of countably many closed Lipschitz curves on which is constant, hence is -measurable and integrable. Since is finite, we have up to at most countably many . For all such it holds
(3.23)
Moreover for every we have
(3.24)
as .
Then by a diagonalization argument on (3.21), (3.23), and (3.24) we can extract a subsequence such that
(3.25)
(3.26)
and as .
Step 3. (lower bound for the blow-up) Recalling the expression of and (3.21), (3.25) is equivalent to
(3.27)
By a change of variable we get
(3.28)
Then denoting by
which, in view of (3.26), satisfies
,
we have
Step 3. (conclusion) The liminf inequality is achieved by integrating over . In fact from and we get
Since , it follows that and so
as desired.
3.3. The limsup inequality
We complete the proof by exhibiting the construction of a recovery sequence, i.e., given a target measure we have to find a sequence such that and
for all . Using a standard diagonal argument it suffices to show the construction for a dense family. Here we consider the set of measures in which are supported on a polyhedral curve . This density result is a consequence of the corresponding result for currents (see Appendix).
Step 1: (polyhedral measures) Now let be a polyhedral measure, in the sense that the are disjoint segments (up to the endpoints), , for . Let . We cover , up to a -null set, with families of countably many disjoint cubes and which are contained in and have the property that with though the centre of the cube so that
for .
In particular for every .
By Proposition 3.9 and Remark 3.10, for all and , there exists a sequence such that , and
(3.30)
Finally define . By the properties of we have that , and
which concludes the proof for polyhedral.
Step 2: (general measures) Finally, we deduce the inequality for all measure . We first extend to a measure , with
where (see the Appendix). By Theorem 3.14 there exist a sequence of polyhedral measures and a sequence of and bi-Lipschitz maps such that
This implies . By Lemma 3.16, the continuity of in and the invariance under deformations of the multiplicity map one obtains
It then follows from the lower semicontinuity with respect to the weak∗ convergence of , the definition of , Step 1 and (3.32) that
as desired.
Appendix: Some results for rectifiable currents
For convenience of the reader here we give the basic definitions and properties for currents in the form that is used in the paper.
Mass of a current: The total variation of the rectifiable current in (2.4) is the measure , its mass is
and it gives the weighted length of the current with respect to the Euclidean norm on . Moreover for any open subset we denote and we can define the support of the current as
Flat norm: The flat norm of the -current is defined as follows: for all let
For example the flat norm of a -current given by to point and with multiplicity and respectively is the length of the segment connecting and .
Slicing of -currents:
For a current such that for all and a Lipschitz function , one can define the slice of through in as
Up to at most a countable set of point for which , it holds
These are the main properties of the slicing that we need in the paper (see Section 4.2.1 in [8]):
;
it holds
for all ;
is a 0-current for a.e. ;
if , a compact set, then
Push forward of -currents: For a bi-lipschitz map , the push-forward of is the current
(3.33)
where it the tangent to with the same orientation of , and denotes the tangential derivative of along , which exists -a.e. on since is Lipschitz on ; if is differentiable in then .
Finally we state some additional results on which the proof of the liminf inequality also relies. Although both are known results in the theory of scalar currents [8, Subsection 4.2.24, Theorems 4.2.16], we refer to [5, Theorem 2.4, Theorem 2.5] for their -valued version and proof. The first one is a compactness result in the class of 1-rectifiable currents without boundary, that, in the liminf, ensures that the limit measure belongs to the space .
Theorem 3.12(Compactness).
Let be a sequence of rectifiable 1-currents without boundary in . If
then there are a current and a subsequence such that
The second theorem gives a characterization of the support of a closed 1-rectifiable current.
Theorem 3.13(Structure).
Let with and . Then there are countably many oriented Lipschitz curves with tangent vector fields and multiplicities such that
Further,
Finally we recall the approximation result for currents with polyhedral currents (this is a classical result for scalar currents. For currents with vector value multiplicity the proof can be found in [5], while the corresponding result for -currents con be found in [3]).
Theorem 3.14(Density).
Fix and consider a -valued closed 1-current . Then there exist a bijective map , with inverse also , and a closed polyhedral 1-current such that
and
Moreover, whenever .
It is important to notice that the deformation result given above guarantees a current without boundary can be approximated by polyhedral currents without boundary, or, equivalently, divergence-free measures. In particular one should note that the multiplicity map is invariant under deformation through a bi-Lipschitz function.
The theorem is given on but a local version can be deduced using the extension lemma recalled below [5, Lemma 2.3].
Lemma 3.15(Extension).
Let be a bounded Lipschitz open set. For every closed rectifiable 1-current defined in , , there exists a closed rectifiable 1-current with and . The constant depends only on . Moreover, , where .
Finally we include the following lemma [5, Lemma 3.3], again useful in proving the limsup inequality.
Lemma 3.16.
Assume that is Borel measurable, obeys , and
Let . Then for any open set we have
Further, if is bi-Lipschitz then for any open set
(3.34)
References
[1] L. Ambrosio, A. Braides: Functionals defined on partitions in sets of finite perimeter. II. Semicontinuity, relaxation and homogenization. J. Math. Pures Appl. (9) 69 (1990), no. 3, 307-333.
[2] L. Ambrosio, A. Braides: Functionals defined on partitions in sets of finite perimeter. I. Integral representation and -convergence. J. Math. Pures Appl. (9) 69 (1990), no. 3, 285-305. 49J45 (49K10)
[3] A. Braides, S. Conti, A. Garroni: Density of polyhedral partitions. Calc. Var. Partial Differential Equations 56 (2017), no. 2, Paper No. 28, 10 pp.
[4]
S. Conti, A. Garroni, M. Ortiz: The line-tension approximation as the dilute limit of linear-elastic dislocations. Arch. Ration. Mech. Anal. 218 (2015), no. 2, 699-755.
[5]
S. Conti, A. Garroni, A. Massaccesi: Modeling of dislocations and relaxation of functionals on -currents with discrete multiplicity, Calc. Var. Partial Differential Equations, 54, 2, 1847-1874, 2015.
[6]
A. Braides, M. Maslennikov, L. Sigalotti: Homogenization by blow up, Applicable Anal., 87, 1341-1356, 2008.
[7]
A. Braides: -convergence for Beginners, Oxford University Press, Oxford, 2002.
[8]
H. Federer: Geometric Measure Theory, Springer-Verlag Berlin, Heidelberg, New York, 1969.
[9]
H. Federer, W. Fleming: Normal and integral currents, Annals of math., 72, 458-520, 1960.
[10]
I. Fonseca, S. Müller: Quasiconvex integrands and lower semicontinuity in , SIAM J. Math. Anal. 23, 1081-1098, 1992.
[11]
L. Simon: Lectures on Geometric Measure Theory, Proceedings of the Center for Mathematical Analysis, Australian National University, Volume 3, 1983.