This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Homogenization of energies defined on 11-rectifiable currents

A. Garroni Dipartimento di Matematica “Guido Castelnuovo”, Sapienza Università di Roma, P.le Aldo Moro 5, I-00185 Roma, Italy garroni@mat.uniroma1.it  and  P. Vermicelli pietro@pietrovermicelli.com
Abstract.

In this paper we study the homogenization of a class of energies concentrated on lines. In dimension 22 (i.e., in codimension 11) the problem reduces to the homogenization of partition energies studied by [1]. There, the key tool is the representation of partitions in terms of BVBV 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 11-currents with multiplicity in a lattice. In the 33 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

Ωγψ(y,ϑ(y),τ(y))𝑑1(y),\int_{\Omega\cap\gamma}\psi(y,\vartheta(y),\tau(y))\,d\mathcal{H}^{1}(y)\,, (1.1)

where γ\gamma is a 11-rectifiable set in Ωn\Omega\subset\mathbb{R}^{n} given by the union of closed loops, the function ϑ:γm\vartheta:\gamma\rightarrow\mathbb{Z}^{m} is a vector-valued multiplicity, constant on each closed loop of γ\gamma, and τ:γ𝒮n1\tau:\gamma\rightarrow\mathcal{S}^{n-1} is the tangent vector defined 1\mathcal{H}^{1}-a.e. on γ\gamma. 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

Ωγψ(xε,ϑ(x),τ(x))𝑑1(x),\int_{\Omega\cap\gamma}\psi\Bigl{(}{x\over\varepsilon},\vartheta(x),\tau(x)\Bigr{)}\,d\mathcal{H}^{1}(x)\,, (1.2)

as ε0\varepsilon\to 0 when ψ\psi 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 ϑ\vartheta is the so-called Burgers vector which belongs to a lattice (which is a material property), that we can assume to be m\mathbb{Z}^{m}. The function ψ\psi 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 Γ\Gamma-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 Γ\Gamma-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 11-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 Γ\Gamma-convergence.

The main result of the paper is that, under suitable growth assumptions of ψ\psi the Γ\Gamma-limit of energies (1.2) as ε0\varepsilon\to 0 exists and can be written as

Ωγψhom(ϑ(y),τ(y))𝑑1(y),\int_{\Omega\cap\gamma}\psi_{\mathrm{hom}}(\vartheta(y),\tau(y))\,d\mathcal{H}^{1}(y)\,, (1.3)

for a suitable function ψhom\psi_{\mathrm{hom}}. 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 Γ\Gamma-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 Ωn\Omega\subset\mathbb{R}^{n} be an open set with Lipschitz boundary. Let ψ:n×m×𝒮n1[0,)\psi:\mathbb{R}^{n}\times\mathbb{Z}^{m}\times\mathcal{S}^{n-1}\rightarrow[0,\infty) be the energy density of the energy (1.3). We assume that ψ\psi ia a Borel function and satisfies

c0|ϑ|ψ(y,ϑ,τ)c1|ϑ|c0,c1>0,yn,τ𝒮n1,ϑn.c_{0}\lvert\vartheta\lvert\leq\psi(y,\vartheta,\tau)\leq c_{1}\lvert\vartheta\lvert\qquad c_{0},c_{1}>0,\,\,\forall y\in\mathbb{R}^{n},\,\,\forall\tau\in\mathcal{S}^{n-1},\ \ \forall\vartheta\in\mathbb{Z}^{n}. (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 df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) the set of divergence-free (in the sense of distribution) measures μ(Ω;m×n)\mu\in\mathcal{M}(\Omega;\mathbb{R}^{m\times n}) of the form

μ=ϑτ1    γ,\mu=\vartheta\otimes\tau\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma,

where ϑ:γm\vartheta:\gamma\rightarrow\mathbb{Z}^{m} is a 1    γ\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma-integrable function, γ\gamma a 1-rectifiable set and τ:γ𝒮n1\tau:\gamma\rightarrow\mathcal{S}^{n-1} its tangent vector defined 1\mathcal{H}^{1}-a.e. on γ\gamma. The divergence-free conditions reads as

γϑ(Dφτ)𝑑1=0,\int_{\gamma}\vartheta\cdot(D\varphi\tau)d\mathcal{H}^{1}=0, (2.2)

for all φC0(Ω;m)\varphi\in C_{0}^{\infty}(\Omega;\mathbb{R}^{m}).

So that, for any μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) the energy in (1.3) is denoted by

F(μ)=Ωγψ(y,ϑ(y),τ(y))𝑑1(y).F(\mu)=\int_{\Omega\cap\gamma}\psi(y,\vartheta(y),\tau(y))\,d\mathcal{H}^{1}(y)\,. (2.3)

In particular by the growth condition (2.1) we deduce that the energy

F(μ)c0|μ|(Ω),F(\mu)\geq c_{0}|\mu|(\Omega),

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 df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) 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 11-rectifiable currents:

We denote by 1(Ω,m)\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) the set of m\mathbb{Z}^{m}-valued 1-rectifiable currents. Let ϑ\vartheta, γ\gamma, τ\tau be as before, then T1(Ω,m)T\in\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) is a functional on the space of smooth compactly supported 1-forms that admits the following representation

T,φ=γϑφ;τ𝑑1m,φCc(Ω,n).\langle T,\varphi\rangle=\int_{\gamma}\vartheta\langle\varphi;\tau\rangle d\mathcal{H}^{1}\in\mathbb{R}^{m},\quad\forall\varphi\in C^{\infty}_{c}(\Omega,\mathbb{R}^{n}). (2.4)

We recall that the boundary of a 1-rectifiable current TT is the 0-current T,φ=T,dφ\langle\partial T,\varphi\rangle=\langle T,d\varphi\rangle for all φCc(Ω)\varphi\in C^{\infty}_{c}(\Omega); finally a current is closed or without boundary if T=0\partial T=0. We denote by 1cl(Ω,m)\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) the set of currents in 1(Ω,m)\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) which are closed.

Now there is a 11-to-11 correspondence between measures in df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) and currents in 1(Ω,m)\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) 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 μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) we denote by T(μ)T(\mu) the corresponding current in 1cl(Ω,m)\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) and for any T1cl(Ω,m)T\in\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) we denote by μ(T)\mu(T) the corresponding measure in df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega).

In particular given μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) the mass of the associated current T(μ)T(\mu) is given by

𝐌(T(μ))=|μ|(Ω).\mathbf{M}(T(\mu))=|\mu|(\Omega).

Moreover the weak convergence of a sequence of measure μj\mu_{j} translates exactly in the weak convergence of the corresponding currents TjT_{j}, i.e.,

limjTj,φ=T,φφCc(Ω,n).\lim_{j}\,\langle T_{j},\varphi\rangle=\langle T,\varphi\rangle\quad\forall\varphi\in C_{c}(\Omega,\mathbb{R}^{n}).

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 1cl(Ω,m)\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) as a countable family of Lipschitz closed loops with constant multiplicity, a compactness result for sequence with bounded mass in 1cl(Ω,m)\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) (see [9]), and a good approximation of currents (and therefore of the corresponding measures) with polyhedral currents (i.e. currents in 1cl(Ω,m)\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}) 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 Ωn\Omega\subset\mathbb{R}^{n} is a bounded, open set with Lipschitz boundary and ψ:n×m×𝒮n1[0,)\psi:\mathbb{R}^{n}\times\mathbb{Z}^{m}\times\mathcal{S}^{n-1}\rightarrow[0,\infty) is a Borel function satisfying (2.1).

Additionally we assume that ψ\psi is 11-periodic in the first variable and we define

Fε(μ)=Ωγψ(yε,ϑ(y),τ(y))𝑑1(y)μdf(m)(Ω).F_{\varepsilon}(\mu)=\int_{\Omega\cap\gamma}\psi\bigg{(}\frac{y}{\varepsilon},\vartheta(y),\tau(y)\bigg{)}d\mathcal{H}^{1}(y)\qquad\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega). (3.1)

The main result of this paper is the characterization of the Γ\Gamma-limit of the functional FεF_{\varepsilon} as ε0\varepsilon\to 0.

We stress that a Γ\Gamma-convergence result must be complemented by a compactness result: together they guarantee the convergence of minima. Here compactness in the class df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega), 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 μεdf(m)(Ω)\mu_{\varepsilon}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) be a sequence of measures satisfying

Fε(με)CF_{\varepsilon}(\mu_{\varepsilon})\leq C

for some C>0C>0, then there is a measure μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) and a subsequence (μεj)(\mu_{\varepsilon_{j}}) such that

μεjμ.\mu_{\varepsilon_{j}}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu.
Proof.

The result is an immediate consequence of the lower bound for the energy density ψ\psi given by (2.1). Indeed Fε(με)CF_{\varepsilon}(\mu_{\varepsilon})\leq C implies that

c0|με|(Ω)=c0Ωγε|ϑε|(y)𝑑1(y)C.c_{0}|\mu_{\varepsilon}|(\Omega)=c_{0}\int_{\Omega\cap\gamma_{\varepsilon}}|\vartheta_{\varepsilon}|(y)d\mathcal{H}^{1}(y)\leq C.

Then, up to a subsequence the sequence of measure με\mu_{\varepsilon} weakly converge to some matrix valued Radon measure μ\mu. In order to conclude it is enough to show that the limit measure as the right structure, i.e., it is concentrated on a 11-rectifiable set γ\gamma and its density is of the form ϑτ\vartheta\otimes\tau, in other words it belongs to df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega).

This is an immediate consequence of the result of compactness of 11-rectifiable currents with integer multiplicity. Indeed the family of currents T(με)T(\mu_{\varepsilon}) satisfy

𝐌(T(με))C𝐌(T(με))=0\mathbf{M}(T(\mu_{\varepsilon}))\leq C\qquad\mathbf{M}(\partial T(\mu_{\varepsilon}))=0

and hence up to a subsequence it converges to a current in T1cl(Ω,m)T\in\mathcal{R}_{1}^{\mathrm{cl}}(\Omega,\mathbb{Z}^{m}). Then μ(T)df(m)(Ω)\mu(T)\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) and it is the weak limit of the corresponding subsequence of με\mu_{\varepsilon}. ∎

Remark 3.2.

This compactness result clarifies the importance of the divergence free condition for the measures μ\mu. It is indeed easy to construct a sequence of measures supported on 11-rectifiable sets with integer multiplicities (equivalently a sequence of integer 11-currents) which converges to a measure which is not supported on curves, for instance which converges to the nn-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 Γ\Gamma-limit of the FεF_{\varepsilon}.

Before stating the theorem, it is useful to introduce the following notation. For all t𝒮n1t\in\mathcal{S}^{n-1} choose a rotation OtSO(n)O_{t}\in\mathrm{SO}(n) with Ote1=tO_{t}e_{1}=t. Then, for every h,l>0h,l>0 we define the rectangle

Rl,ht=Ot[[l2,l2]×[h2,h2]n1]R_{l,h}^{t}=O_{t}\bigg{[}\bigg{[}-\frac{l}{2},\frac{l}{2}\bigg{]}\times\bigg{[}-\frac{h}{2},\frac{h}{2}\bigg{]}^{n-1}\bigg{]}

of height hh and a side ll, centred at the origin, and one side parallel to the direction tt. If the rectangle is centred in a point xnx\in\mathbb{R}^{n} we denote it by Rl,ht(x)R_{l,h}^{t}(x). Similarly we denote by Qht(x)Q_{h}^{t}(x) the cube of side hh, and one side parallel to the direction tt, centred at xΩx\in\Omega, i.e., Qht(x)=Rh,ht(x)Q_{h}^{t}(x)=R_{h,h}^{t}(x). If x=0x=0 we drop the xx and write QhtQ_{h}^{t}.

Theorem 3.3 (Homogenization).

Assume that ψ\psi, satisfying (2.1), is 11-periodic in the first variable, then the functionals FεF_{\varepsilon} in ((3.1)) Γ\Gamma-converge as ε0\varepsilon\to 0, with respect to the weak convergence of measures, to the functional defined by

Fhom(μ)={γΩψhom(ϑ,τ)𝑑1μdf(m)(Ω)+otherwise,F_{\mathrm{hom}}(\mu)=\begin{cases}\displaystyle{\int_{\gamma\cap\Omega}\psi_{\mathrm{hom}}(\vartheta,\tau)d\mathcal{H}^{1}}&\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega)\\ +\infty&\hbox{otherwise},\end{cases}

where for all bm\textrm{b}\in\mathbb{Z}^{m} and t𝒮n1\textrm{t}\in\mathcal{S}^{n-1} the effective energy density is given by

ψhom(b,t)=limT1Tinf{QTtγ\displaystyle\psi_{\mathrm{hom}}(b,t)=\lim_{T\to\infty}\frac{1}{T}\inf\Bigg{\{}\int_{Q_{T}^{t}\cap\gamma} ψ(y,ϑ(y),τ(y))d1(y):μdf(m)(QTt),\displaystyle\psi(y,\vartheta(y),\tau(y))d\mathcal{H}^{1}(y):\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{T}^{t}), (3.2)
supp(μbt1 (tQTt))QTt}.\displaystyle\mathrm{supp}(\mu-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}\cap Q_{T}^{t}))\subset Q_{T}^{t}\Bigg{\}}.
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 AnA\subseteq\mathbb{R}^{n} and a measure μ=ϑτ1    γdf(m)(A)\mu=\vartheta\otimes\tau\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma\in\mathcal{M}_{\mathrm{df}}^{(m)}(A),

Fε(μ,A)=Aγψ(yε,ϑ,τ)𝑑1.F_{\varepsilon}(\mu,A)=\int_{A\cap\gamma}\psi\left(\frac{y}{\varepsilon},\vartheta,\tau\right)d\mathcal{H}^{1}. (3.3)

In the next proposition we prove that the energy density ψhom\psi_{\mathrm{hom}} is well defined through an asymptotic formula. Moreover we show that it is rather flexible and thanks to the periodicity of ψ\psi is not sensitive to the translations. As a consequence we will get the continuity of ψhom\psi_{\mathrm{\hom}}.

Proposition 3.4 (Homogenization formula).

Let (xT)(x_{T}) be a family of points in n\mathbb{R}^{n}, with TT\in\mathbb{R} and ψ\psi as in Theorem 3.3. Then for all bmb\in\mathbb{Z}^{m} and t𝒮n1t\in\mathcal{S}^{n-1} the limit

limT1Tinf{QTt(xT)γ\displaystyle\lim_{T\to\infty}\frac{1}{T}\inf\Bigg{\{}\int_{Q_{T}^{t}(x_{T})\cap\gamma} ψ(y,ϑ,τ)d1:μ=ϑτ1 γdf(m)(QTt(xT))\displaystyle\psi(y,\vartheta,\tau)d\mathcal{H}^{1}:\mu=\vartheta\otimes\tau\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{T}^{t}(x_{T}))
supp(μbt1 (xT+t))QTt(xT)}\displaystyle\mathrm{supp}(\mu-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(x_{T}+t\mathbb{R}))\subset Q_{T}^{t}(x_{T})\Bigg{\}}

exists and it is independent of (xT)(x_{T}). Therefore it coincides with ψhom(b,t)\psi_{\hom}(b,t).

Proof.

Let γt¯(xT)\bar{\gamma_{t}}(x_{T}) be the straight line parallel to tt passing through xTx_{T}, i.e., γt¯(xT)=xT+t\bar{\gamma_{t}}(x_{T})=x_{T}+t\mathbb{R} and denote γt¯=γt¯(0)\bar{\gamma_{t}}=\bar{\gamma_{t}}(0). Note that, in general, the line γt¯(xT)\bar{\gamma_{t}}(x_{T}) may not intersect the set of points in n\mathbb{Z}^{n}.

Let STS\gg T. Fix L>nL>n and consider the family of equispaced points (zj)(z_{j}) on γt¯(xS)\bar{\gamma_{t}}(x_{S}) with spacing T+LT+L. For each point zjz_{j} consider the point yjxT+ny_{j}\in x_{T}+\mathbb{Z}^{n} such that

|yjzj|n.\lvert y_{j}-z_{j}\lvert\leq\sqrt{n}. (3.4)

Let μ(T)=ϑ(T)τ(T)1    γ(T)\mu^{(T)}=\vartheta^{(T)}\otimes\tau^{(T)}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma^{(T)} be a test measure for the minimum problem

m(T)=inf{F1(μ,QTt(xT)):μdf(m)(n),supp(μbt1    γt¯(xT))QTt(xT)}m^{(T)}=\inf\Bigg{\{}F_{1}(\mu,Q_{T}^{t}(x_{T})):\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n})\ ,\,\ \mathrm{supp}(\mu-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\bar{\gamma_{t}}(x_{T}))\subset Q_{T}^{t}(x_{T})\Bigg{\}} (3.5)

such that

F1(μ(T),QTt(xT))=QTt(xT)γ(T)ψ(y,ϑ(T),τ(T))𝑑1<m(T)+1T.F_{1}(\mu^{(T)},Q_{T}^{t}(x_{T}))=\int_{Q_{T}^{t}(x_{T})\cap\gamma^{(T)}}\psi(y,\vartheta^{(T)},\tau^{(T)})d\mathcal{H}^{1}<m^{(T)}+\frac{1}{T}. (3.6)
Refer to caption
Figure 1.

Let JJ be the set of indices such that QTt(yj)QSt(xS)Q_{T}^{t}(y_{j})\subset Q_{S}^{t}(x_{S}) and

RS,T=QSt(xS)jJ(QTt(yj)).R_{S,T}=Q_{S}^{t}(x_{S})\setminus\bigcup_{j\in J}(Q_{T}^{t}(y_{j})).

Hence define the function hj(x)=xxT+yjh^{j}(x)=x-x_{T}+y_{j} and the measure

μ~(S)=jJhj(μ(T))    QTt(yj)+bt1    (γt¯(xS)RS,T).\widetilde{\mu}^{(S)}=\sum_{j\in J}h^{j}_{\sharp}(\mu^{(T)}){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t}(y_{j})+b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(\bar{\gamma_{t}}(x_{S})\cap R_{S,T}).

Note that the measure hjμ(T)h_{\sharp}^{j}\mu^{(T)} is nothing else than the measure obtained from μ(T)\mu^{(T)} translating the support from the cube QTt(xT)Q^{t}_{T}(x_{T}) to the cube QTt(yj)Q^{t}_{T}(y_{j}), i.e.,

hjμ(T)=ϑ(T)(xT+yj)τ(T)(xT+yj)1    (γ(T)xT+yj),h_{\sharp}^{j}\mu^{(T)}=\vartheta^{(T)}(\cdot-x_{T}+y_{j})\otimes\tau^{(T)}(\cdot-x_{T}+y_{j})\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(\gamma^{(T)}-x_{T}+y_{j}),

therefore by the choice of yjy_{j} we have

F1(hjμ(T),QTt(yj))<m(T)+1T.F_{1}(h_{\sharp}^{j}\mu^{(T)},Q_{T}^{t}(y_{j}))<m^{(T)}+\frac{1}{T}.

From μ~(S)\widetilde{\mu}^{(S)}, we can obtain a divergence free measure μ^(S)=ϑ^(S)τ^(S)1    γ^(S)\widehat{\mu}^{(S)}=\widehat{\vartheta}^{(S)}\otimes\widehat{\tau}^{(S)}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\widehat{\gamma}^{(S)} in QSt(xS)Q_{S}^{t}(x_{S}), by connecting through a segment each endpoint of γ(T)+yixT\gamma^{(T)}+y_{i}-x_{T} on yi+QTty_{i}+Q_{T}^{t} to γ¯t(xS)\bar{\gamma}_{t}(x_{S}). In doing so we have obtained a test measure for m(S)m^{(S)} (see Figure 1). The conclusion follows if we prove

lim supS1Sm(S)lim infT1Tm(T).\limsup_{S\to\infty}\frac{1}{S}m^{(S)}\leq\liminf_{T\to\infty}\frac{1}{T}m^{(T)}.

In fact, using μ^(S)\widehat{\mu}^{(S)} we get

1S\displaystyle\frac{1}{S} m(S)1SQSt(xS)γ^(S)ψ(y,ϑ^(S),τ^(S))𝑑1\displaystyle m^{(S)}\leq\frac{1}{S}\int_{Q_{S}^{t}(x_{S})\cap\widehat{\gamma}^{(S)}}\psi(y,\widehat{\vartheta}^{(S)},\widehat{\tau}^{(S)})d\mathcal{H}^{1}
=1SjJF1(hjμ(T),QTt(yj))+1SRS,Tγt¯(xS)ψ(y,b,t)𝑑1\displaystyle=\frac{1}{S}\sum_{j\in J}F_{1}(h_{\sharp}^{j}\mu^{(T)},Q_{T}^{t}(y_{j}))+\frac{1}{S}\int_{R_{S,T}\cap\bar{\gamma_{t}}(x_{S})}\psi(y,b,t)d\mathcal{H}^{1}
1S[ST+L](m(T)+1T)+1S1(RS,Tγt¯(xS))c+2Cn1S[ST+L]\displaystyle\leq\frac{1}{S}\bigg{[}\frac{S}{T+L}\bigg{]}\bigg{(}m^{(T)}+\frac{1}{T}\bigg{)}+\frac{1}{S}\mathcal{H}^{1}(R_{S,T}\cap\bar{\gamma_{t}}(x_{S}))c+2C\sqrt{n}\frac{1}{S}\bigg{[}\frac{S}{T+L}\bigg{]}
1S[ST+L](m(T)+1T)+1S[ST(|[ST+L]2|)]c\displaystyle\leq\frac{1}{S}\bigg{[}\frac{S}{T+L}\bigg{]}\bigg{(}m^{(T)}+\frac{1}{T}\bigg{)}+\frac{1}{S}\bigg{[}S-T\bigg{(}\bigg{\lvert}\bigg{[}\frac{S}{T+L}\bigg{]}-2\bigg{\lvert}\bigg{)}\bigg{]}c
+2Cn1S[ST+L]\displaystyle+2C\sqrt{n}\frac{1}{S}\bigg{[}\frac{S}{T+L}\bigg{]}

where we have used the choice of μ(T)\mu^{(T)} as in (3.6) and (3.4) to control the contribution of the segments in the measure μ^(S)\widehat{\mu}^{(S)}. The conclusion follows taking the lim supS\limsup_{S\to\infty} first and then the lim infT\liminf_{T\to\infty}. ∎

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 xTx_{T} but also uniform in xTx_{T}. Indeed, by periodicity, we can assume that xTQ1(0)x_{T}\in Q_{1}(0) which is compact, and then we deduce the uniformity.

An important consequence of the above characterization of ψhom\psi_{\mathrm{hom}} is its continuity.

Proposition 3.6 (Continuity of ψhom\psi_{\mathrm{hom}}).

For all bmb\in\mathbb{Z}^{m} and t,t𝒮n1t,t^{\prime}\in\mathcal{S}^{n-1} let ψhom\psi_{\hom} be as in Theorem 3.3. Then the following continuity property holds

|ψhom(b,t)ψhom(b,t)|c|b||tt|,\lvert\psi_{\mathrm{hom}}(b,t)-\psi_{\mathrm{hom}}(b,t^{\prime})\lvert\leq c\lvert b\lvert\lvert t-t^{\prime}\lvert,

with c>0c>0 a constant that depends only on ψhom\psi_{\mathrm{hom}}.

Proof.

We prove this property by means of the asymptotic formula. Suppose that ψhom(b,t)ψhom(b,t)\psi_{\mathrm{hom}}(b,t^{\prime})\geq\psi_{\mathrm{hom}}(b,t). Fix ε>0\varepsilon>0 and let μt(T)\mu_{t}^{(T)} be a test measure for the minimum problem mt(T)m_{t}^{(T)} in (3.5) (Proposition 1.12, with xT=0x_{T}=0), such that

QTtγt(T)ψ(y,ϑt(T),τt(T))𝑑1<mt(T)+ε\int_{Q_{T}^{t}\cap\gamma_{t}^{(T)}}\psi(y,\vartheta_{t}^{(T)},\tau_{t}^{(T)})d\mathcal{H}^{1}<m_{t}^{(T)}+\varepsilon (3.7)

From μt(T)\mu_{t}^{(T)}, we can obtain a test measure μ^t(T)\widehat{\mu}_{t^{\prime}}^{(T)} for the problem

inf{QT+T|tt|tγψ(y,ϑ,τ)𝑑1:μdf(m)(n),supp(μbt1    (t))QT+T|tt|t}\inf\bigg{\{}\int_{Q_{T+T\lvert t-t^{\prime}\lvert}^{t^{\prime}}\cap\gamma}\psi(y,\vartheta,\tau)d\mathcal{H}^{1}:\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n}),\ \mathrm{supp}(\mu-b\otimes t^{\prime}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t^{\prime}\mathbb{R}))\subset Q_{T+T\lvert t-t^{\prime}\lvert}^{t^{\prime}}\bigg{\}}

proceeding as in Figure 2 (with support on the bold line), adding two segments to obtain the divergence free condition in QT+T|tt|tQTtQ_{T+T\lvert t-t^{\prime}\lvert}^{t^{\prime}}\supset Q_{T}^{t} for the measure μ^t(T)\widehat{\mu}_{t^{\prime}}^{(T)} and to get the right boundary condition.

Refer to caption
Figure 2. The construction of the measure μ^t(T)\widehat{\mu}_{t^{\prime}}^{(T)}, for simplicity t=e1t^{\prime}=e_{1}.

Then it is easy to see that

ψhom(b,t)1Tmt(T)1T(\displaystyle\psi_{\mathrm{hom}}(b,t^{\prime})-\frac{1}{T}m_{t}^{(T)}\leq\frac{1}{T}\bigg{(} QT+T|tt|tγ^t(T)ψ(y,ϑ^t(T),τ^t(T))𝑑1\displaystyle\int_{Q_{T+T\lvert t-t^{\prime}\lvert}^{t^{\prime}}\cap\widehat{\gamma}_{t^{\prime}}^{(T)}}\psi(y,\widehat{\vartheta}_{t^{\prime}}^{(T)},\widehat{\tau}_{t^{\prime}}^{(T)})d\mathcal{H}^{1}
QTtγt(T)ψ(y,ϑt(T),τt(T))d1)+εT\displaystyle-\int_{Q_{T}^{t}\cap\gamma_{t}^{(T)}}\psi(y,\vartheta_{t}^{(T)},\tau_{t}^{(T)})d\mathcal{H}^{1}\bigg{)}+\frac{\varepsilon}{T}

and using the estimate from above for ψ\psi and by construction of μ^tT\widehat{\mu}_{t}^{T} we have

ψhom(b,t)1Tmt(T)1Tc|b|T|tt|+εT.\psi_{\mathrm{hom}}(b,t^{\prime})-\frac{1}{T}m_{t}^{(T)}\leq\frac{1}{T}c\lvert b\lvert T\lvert t-t^{\prime}\lvert+\frac{\varepsilon}{T}.

We conclude taking the limit TT\to\infty and by the arbitrariness of ε\varepsilon. To obtain the inequality with the opposite sign is enough to swap the roles of tt and tt^{\prime}. ∎

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 μjdf(m)(Q1t)\mu_{j}\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{1}^{t}) be a sequence of divergence free measures in Q1tQ_{1}^{t} such that

supj|μj|(Q1t)M<\sup_{j}\lvert\mu_{j}\lvert(Q_{1}^{t})\leq M<\infty (3.8)

and μjμ=bt1    (tQ1t)df(m)(Q1t)\mu_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu=b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}\cap Q_{1}^{t})\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{1}^{t}). Then, ε>0\forall\varepsilon>0, δ0(0,1/2)\forall\delta_{0}\in(0,1/2) there exist a sequence δj\delta_{j}, δ0/2<δj<δ0\delta_{0}/2<\delta_{j}<\delta_{0} and a sequence μ~jdf(m)(Q1t)\widetilde{\mu}_{j}\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{1}^{t}) such that μ~j=μj    (Q1δjt)+bt1    (Q1tQ¯1δjt)+νj\widetilde{\mu}_{j}=\mu_{j}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(Q_{1-\delta_{j}}^{t})^{\circ}+b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(Q_{1}^{t}\setminus\bar{Q}_{1-\delta_{j}}^{t})+\nu_{j}, where νj[(Q1t)]m×n\nu_{j}\in[\mathcal{M}(Q_{1}^{t})]^{m\times n} is such that supp(νj)Q1δj\mathrm{supp}(\nu_{j})\subset\partial Q_{1-\delta_{j}} and |νj|(Q1t)<ε\lvert\nu_{j}\lvert(Q_{1}^{t})<\varepsilon for every jj large enough (by (Q1δjt)(Q_{1-\delta_{j}}^{t})^{\circ} we have denoted the interior of the cube (Q1δjt)(Q_{1-\delta_{j}}^{t})).

Proof.

Fix ε(0,1)\varepsilon\in(0,1) and δ0(0,1/2)\delta_{0}\in(0,1/2) two independent parameters and consider dd, the distance function from Q1t\partial Q_{1}^{t}. We now slice the measure μ^j=μjμb,t\widehat{\mu}_{j}=\mu_{j}-\mu_{b,t}, with μb,t=bt1    (t)\mu_{b,t}=b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}), through the function dd, Lipschitz continuous with Lip(d)=1\mathrm{Lip}(d)=1.

Let γ^j\widehat{\gamma}_{j} and ϑ^j\widehat{\vartheta}_{j} be the support and multiplicity of μ^j\widehat{\mu}_{j}, respectively. Now the idea is to exploit the weak convergence of μ^j\widehat{\mu}_{j} to zero in order to find a δj\delta_{j} for which suppμ^jQ1δj\mathrm{supp}\,\widehat{\mu}_{j}\cap\partial Q_{1-\delta_{j}} is the boundary of weighted segments with small mass; this will allow us to construct the measure νj\nu_{j} in the statement. Even though this can done by hands using the measures in df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) 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 11 dimensional objects.

We denote T^j=T(μ^j)\widehat{T}_{j}=T(\widehat{\mu}_{j}) the current associated to μ^j\widehat{\mu}_{j} and for every δ[0,δ0]\delta\in[0,\delta_{0}] we consider the current obtained by slicing T^j\widehat{T}_{j} along the level sets of the function dd which we denote by

T^j[d,δ+]=(T^j)    {x:d(x)>δ}(T^j    {x:d(x)>δ}).\widehat{T}_{j}[d,\delta+]=(\partial\widehat{T}_{j}){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:d(x)>\delta\}-\partial(\widehat{T}_{j}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:d(x)>\delta\}).

Since T^j=0\partial\widehat{T}_{j}=0 in Q1tQ_{1}^{t}, we have T^j[d,δ+]=(T^j    {x:d(x)>δ})\widehat{T}_{j}[d,\delta+]=-\partial(\widehat{T}_{j}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:d(x)>\delta\}), i.e.

T^j[d,δ+],φ=γj{d>δ}ϑjφ,τj𝑑1\langle\widehat{T}_{j}[d,\delta+],\varphi\rangle=\int_{\gamma_{j}\cap\{d>\delta\}}\vartheta_{j}\langle\nabla\varphi,\tau_{j}\rangle d\mathcal{H}^{1}

for all φC(Q1t)\varphi\in C^{\infty}(Q_{1}^{t}). For slicing of currents it holds

δ02δ0𝐌(T^j[d,s+])𝑑sLip(d)𝐌Q1δ0/2tQ1δ0t(T^j)<𝐌Q1δ0/2t(T^j)\int_{\frac{\delta_{0}}{2}}^{\delta_{0}}\mathbf{M}(\widehat{T}_{j}[d,s+])\,ds\leq\mathrm{Lip}(d)\mathbf{M}_{Q_{1-\delta_{0}/2}^{t}\setminus Q_{1-\delta_{0}}^{t}}(\widehat{T}_{j})<\mathbf{M}_{Q_{1-\delta_{0}/2}^{t}}(\widehat{T}_{j}) (3.9)

and

δ02δ0𝐅Q¯1δ0/2t{d(x)=s}(T^j[d,s+])𝑑sLip(d)𝐅Q¯1δ0/2t(T^j),\int_{\frac{\delta_{0}}{2}}^{\delta_{0}}\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}\cap\{d(x)=s\}}(\widehat{T}_{j}[d,s+])\,ds\leq\mathrm{Lip}(d)\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}}(\widehat{T}_{j}), (3.10)

where 𝐅Q(T)\mathbf{F}_{Q}(T) denote the flat norm of the current TT (see the Appendix). Since, by (3.8), the sequence T^j\widehat{T}_{j} is equi-bounded in mass, from (3.9) we get that, for a.e. δ(δ0/2,δ0)\delta\in(\delta_{0}/2,\delta_{0}), 𝐌(T^j[d,δ+])\mathbf{M}(\widehat{T}_{j}[d,\delta+]) is finite. Hence, since T^j[d,δ+]\widehat{T}_{j}[d,\delta+] is a 0-rectifiable current for almost every δ\delta, we have

𝐌(T^j[d,s+])=xQ1δγ^j|ϑ^j(x)|,|ϑ^j(x)|1\mathbf{M}(\widehat{T}_{j}[d,s+])=\sum_{x\in\partial Q_{1-\delta}\cap\widehat{\gamma}_{j}}\lvert\widehat{\vartheta}_{j}(x)\lvert,\quad\lvert\widehat{\vartheta}_{j}(x)\lvert\geq 1 (3.11)

where the sum runs through a finite number of points x1,,xMx_{1},\ldots,x_{M}, with multiplicity ϑ^j(xi)\widehat{\vartheta}_{j}(x_{i}) and positive orientation if γ^j\widehat{\gamma}_{j} exits from Q1δQ_{1-\delta} at xix_{i} (these oriented points together with their multiplicity give the boundary of T^j    Q1δ\widehat{T}_{j}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{1-\delta}).

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 T^j0\widehat{T}_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}0 to obtain the convergence to zero of the flat norm. Therefore σ<δ04ε\forall\sigma<\frac{\delta_{0}}{4}\varepsilon we can find JσJ_{\sigma}\in\mathbb{N} such that 𝐅Q¯1δ0/2t(T^j)<σ\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}}(\widehat{T}_{j})<\sigma for all jJσj\geq J_{\sigma}. Thus, by (3.10), for each jJσj\geq J_{\sigma}, the sets

Aj={δ(δ0/2,δ0):𝐅Q¯1δ0/2t{d(x)=δ}(T^j[d,δ+])<4δ0σ}A_{j}=\bigg{\{}\delta\in(\delta_{0}/2,\delta_{0}):\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}\cap\{d(x)=\delta\}}(\widehat{T}_{j}[d,\delta+])<\frac{4}{\delta_{0}}\sigma\bigg{\}} (3.12)

have positive 11-dimensional Lebesgue measure. In fact

|Ajc|\displaystyle|A_{j}^{c}| δ04σδ02δ0𝐅Q¯1δ0/2t{d(x)=s}(T^j[d,s+])𝑑s\displaystyle\leq\frac{\delta_{0}}{4\sigma}\int_{\frac{\delta_{0}}{2}}^{\delta_{0}}\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}\cap\{d(x)=s\}}(\widehat{T}_{j}[d,s+])\,ds
δ04σ𝐅Q1δ0/2t(T^j)<δ04,\displaystyle\leq\frac{\delta_{0}}{4\sigma}\mathbf{F}_{Q_{1-\delta_{0}/2}^{t}}(\widehat{T}_{j})<\frac{\delta_{0}}{4},

hence

|Aj|=δ02|Ajc|>δ04>0.|A_{j}|=\frac{\delta_{0}}{2}-|A_{j}^{c}|>\frac{\delta_{0}}{4}>0.

Now, for each jJσj\geq J_{\sigma} choose a δj\delta_{j} such that δjAj\delta_{j}\in A_{j} 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 δj\delta_{j}, have solution. By definition, for every jJσj\geq J_{\sigma},

𝐅Q¯1δ0/2t{d(x)=δj}(T^j[d,δj+])\displaystyle\mathbf{F}_{\bar{Q}_{1-\delta{0}/2}^{t}\cap\{d(x)=\delta_{j}\}}(\widehat{T}_{j}[d,\delta_{j}+])
=inf{𝐌(Rj)+𝐌(Bj):supp(Rj),supp(Bj)Q1δj,Rj+Bj=T^j[d,δj+]}.\displaystyle\quad=\inf\{\mathbf{M}(R_{j})+\mathbf{M}(B_{j}):\mathrm{supp}(R_{j}),\,\mathrm{supp}(B_{j})\subset\partial Q_{1-\delta_{j}},R_{j}+\partial B_{j}=\widehat{T}_{j}[d,\delta_{j}+]\}.

Since RjR_{j} is a m\mathbb{Z}^{m}-valued 0-rectifiable current, then 𝐌(Rj)>1\mathbf{M}(R_{j})>1 and from

𝐅Q¯1δ0/2t{d(x)=δj}(T^j[d,δj+])<ε<1\mathbf{F}_{\bar{Q}_{1-\delta{0}/2}^{t}\cap\{d(x)=\delta_{j}\}}(\widehat{T}_{j}[d,\delta_{j}+])<\varepsilon<1

we can conclude that Rj=0R_{j}=0. It follows that, for every jJσj\geq J_{\sigma},

𝐅Q¯1δ0/2t{d(x)=δj}(T^j[d,δj+])\displaystyle\mathbf{F}_{\bar{Q}_{1-\delta{0}/2}^{t}\cap\{d(x)=\delta_{j}\}}(\widehat{T}_{j}[d,\delta_{j}+])
=inf{𝐌(Bj):Bj1(Q1t;m),supp(Bj)Q1δj,Bj=T^j[d,δj+]}\displaystyle\quad=\inf\{\mathbf{M}(B_{j}):\ B_{j}\in\mathcal{R}_{1}(Q_{1}^{t};\mathbb{Z}^{m}),\mathrm{supp}(B_{j})\subset\partial Q_{1-\delta_{j}},\,\partial B_{j}=\widehat{T}_{j}[d,\delta_{j}+]\}

It’s easily checked that, for all jJσj\geq J_{\sigma}, the set over which we take the above infimum is not empty. Hence jJσ\forall j\geq J_{\sigma}, by means of the direct method, there exists a 11-rectifiable current Sj1(Q1t,m)S_{j}\in\mathcal{R}_{1}(Q_{1}^{t},\mathbb{Z}^{m}) that satisfies the following properties

  1. (i)

    supp(Sj)Q1δj\mathrm{supp}(S_{j})\subset\partial Q_{1-\delta_{j}};

  2. (ii)

    𝐌(Sj)=𝐅Q¯1δ0/2t{d(x)=δj}(T^j[d,δj+])<4δ0σ<ε\mathbf{M}(S_{j})=\mathbf{F}_{\bar{Q}_{1-\delta_{0}/2}^{t}\cap\{d(x)=\delta_{j}\}}(\widehat{T}_{j}[d,\delta_{j}+])<\frac{4}{\delta_{0}}\sigma<\varepsilon;

  3. (iii)

    Sj=T^j[d,δj+]\partial S_{j}=\widehat{T}_{j}[d,\delta_{j}+].

Now we conclude defining νj=μ(Sj)\nu_{j}=\mu(S_{j}) for jJσj\geq J_{\sigma} and equal to νj=bt1    (Q1δjt)μj    (Q1δjt)\nu_{j}=b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(Q_{1-\delta_{j}}^{t})^{\circ}-\mu_{j}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(Q_{1-\delta_{j}}^{t})^{\circ} otherwise.

Remark 3.8.

By a scaling argument the result in Lemma 3.7 still hold in the domain QTtQ_{T}^{t}.

Moreover by a diagonal argument under the assumptions of Lemma 3.7 one could show that, for every δ\delta, there exists a sequence μ~jdf(m)(QTt)\widetilde{\mu}_{j}\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q_{T}^{t}) such that μ~jbt1    QTt\widetilde{\mu}_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t}, with supp(μ~jbt1    QTt+νj)QT(1δ)t\mathrm{supp}(\widetilde{\mu}_{j}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t}+\nu_{j})\subset Q^{t}_{T(1-\delta)}. Moreover, exploiting the construction of μ~j\widetilde{\mu}_{j} and its decomposition in mutually singular measures we also deduce that

limjF1(μ~j,QTt)=limjF1(μj,QT(1δ)t)+F1(bt1    QTt,QTtQT(1δ)t).\lim_{j}F_{1}(\widetilde{\mu}_{j},Q_{T}^{t})=\lim_{j}F_{1}(\mu_{j},Q^{t}_{T(1-\delta)})+F_{1}(b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t},Q_{T}^{t}\setminus Q^{t}_{T(1-\delta)}). (3.13)

A further diagonal argument produces a sequence μ^j\hat{\mu}_{j} such that μ^jbt1    QTt\hat{\mu}_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t}, with supp(μ^jbt1    QTt+νj)QTt\mathrm{supp}(\hat{\mu}_{j}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T}^{t}+\nu_{j})\subset\subset Q^{t}_{T} and

limjF1(μ^j,QTt)limjF1(μj,QTt).\lim_{j}F_{1}(\hat{\mu}_{j},Q_{T}^{t})\leq\lim_{j}F_{1}(\mu_{j},Q^{t}_{T}). (3.14)

Using the above lemma we finally prove the following proposition which is essentially the Γ\Gamma-convergence result when the limiting configuration is given by the measure μb,t:=bt1    (t)\mu_{b,t}:=b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}), i.e., straight with constant multiplicity.

Proposition 3.9.

Let (yT)(y_{T}) be a family of points in n\mathbb{R}^{n}, with TT\in\mathbb{R} and ψ\psi as in Theorem 3.3. For all bmb\in\mathbb{Z}^{m} and t𝒮n1t\in\mathcal{S}^{n-1} we define

ψ(b,t):=inf{lim infL+F1L(μL,Q1t(yL)):μLdf(m)(n),μLbt1    (yL+t)0}.\psi^{*}(b,t):=\inf\left\{\liminf_{L\to+\infty}F_{\frac{1}{L}}(\mu_{L},Q_{1}^{t}(y_{L})):\mu_{L}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n}),\,\ \mu_{L}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(y_{L}+t\mathbb{R})\stackrel{{\scriptstyle*}}{{\rightharpoonup}}0\right\}\,.

Then

ψhom(b,t)=ψ(b,t).\psi_{\mathrm{\hom}}(b,t)=\psi^{*}(b,t). (3.15)
Proof.

In order to show one inequality we construct a sequence μL\mu_{L} admissible for the definition of ψ(b,t)\psi^{*}(b,t).

As above by the periodicity of the problem we can reduce to the case in which yL=0y_{L}=0. For any T>0T>0 we consider a family of equispaced points (xi)(x_{i}) with spacing TT on tt\mathbb{R}. Given ε>0\varepsilon>0, thanks to Remark 3.5, we can find TT large enough such that we find μi(T)=ϑi(T)τi(T)d1    γi(T)\mu_{i}^{(T)}=\vartheta_{i}^{(T)}\otimes\tau_{i}^{(T)}d\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma_{i}^{(T)} a test measure for the minimum problem m(T)m^{(T)} in (3.5) with xT=xix_{T}=x_{i} such that

1TQTt(xi)γi(T)ψ(y,ϑi(T),τi(T))𝑑1ψhom(b,t)+ε.\frac{1}{T}\int_{Q_{T}^{t}(x_{i})\cap\gamma_{i}^{(T)}}\psi(y,\vartheta_{i}^{(T)},\tau_{i}^{(T)})\,d\mathcal{H}^{1}\leq\psi_{\mathrm{hom}}(b,t)+\varepsilon. (3.16)

In particular we have that supp(μi(T)bt1    (t))QTt(xi)\mathrm{supp}(\mu_{i}^{(T)}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}))\subset Q_{T}^{t}(x_{i}). With this condition we can rescale and glue the measures μi(T)\mu_{i}^{(T)} and construct a sequence which is admissible for ψ(b,t)\psi^{*}(b,t). Precisely, given L>0L>0 we define the function f1/L(x)=xLf^{1/L}(x)=\frac{x}{L} and denote IL={i:QT/Lt(xi/L)Q1t}I_{L}=\{i:\ Q_{T/L}^{t}(x_{i}/L)\subset Q_{1}^{t}\}, RL=tiITQT/Lt(xi/L)R^{L}=t\mathbb{R}\setminus\cup_{i\in I_{T}}Q_{T/L}^{t}(x_{i}/L), then the measure

μL=iILf1/L(μi(T))    QT/Lt(xi/L)+bt1    (tRL)\mu_{L}=\sum_{i\in I_{L}}f_{\sharp}^{1/L}(\mu_{i}^{(T)}){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q_{T/L}^{t}(x_{i}/L)+b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}\cap R^{L})

is divergence free and satisfies μLbt1    (t)\mu_{L}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}) as LL\to\infty. Therefore using the notation μL=ϑLτL1    γL\mu_{L}=\vartheta_{L}\otimes\tau_{L}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma_{L} we have

F1L(μL,Q1t)=\displaystyle F_{\frac{1}{L}}(\mu_{L},Q_{1}^{t})= Q1tγLψ(Ly,ϑL(y),τL(y))𝑑1(y)=\displaystyle\int_{Q_{1}^{t}\cap\gamma_{L}}\psi\big{(}Ly,\vartheta_{L}(y),\tau_{L}(y)\big{)}d\mathcal{H}^{1}(y)=
=iILF1L(f1/L(μi(T)),QT/Lt(xi/L))+RLQ1tψ(y,b,t)𝑑1(y)\displaystyle=\sum_{i\in I_{L}}F_{\frac{1}{L}}(f_{\sharp}^{1/L}(\mu_{i}^{(T)}),Q_{T/L}^{t}(x_{i}/L))+\int_{R^{L}\cap Q_{1}^{t}}\psi(y,b,t)d\mathcal{H}^{1}(y)
=iIL1LF1(μi(T),QTt(xi))+RLQ1tψ(y,b,t)𝑑1(y)\displaystyle=\sum_{i\in I_{L}}\frac{1}{L}F_{1}(\mu_{i}^{(T)},Q_{T}^{t}(x_{i}))+\int_{R^{L}\cap Q_{1}^{t}}\psi(y,b,t)d\mathcal{H}^{1}(y)
(IL)TL(ψhom(b,t)+ε)+c1|b|1(RLQ1t)\displaystyle\leq{\sharp(I_{L})}\frac{T}{L}(\psi_{\mathrm{\hom}}(b,t)+\varepsilon)+c_{1}|b|\mathcal{H}^{1}(R^{L}\cap Q_{1}^{t})
[LT]TL(ψhom(b,t)+ε)+c1|b|TL.\displaystyle\leq\bigg{[}\frac{L}{T}\bigg{]}\frac{T}{L}(\psi_{\mathrm{\hom}}(b,t)+\varepsilon)+c_{1}|b|\frac{T}{L}.

Thus

ψ(b,t)ψhom(b,t)+ε\psi^{*}(b,t)\leq\psi_{\mathrm{\hom}}(b,t)+\varepsilon

which gives one inequality.

In order to obtain the opposite inequality it is enough to observe that given LL and fixing a sequence μL\mu_{L} admissible for the minimum problem ψ(b,t)\psi^{*}(b,t), by means of Remark 3.8 and (3.14), we find a sequence μ^L\hat{\mu}_{L} satisfying supp(μ^Lbt1    t)Q1t\mathrm{supp}(\hat{\mu}_{L}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}t\mathbb{R})\subset\subset Q^{t}_{1} such that

lim infLF1L(μ^L,Q1t)lim infLF1L(μL,Q1t).\liminf_{L}F_{\frac{1}{L}}(\hat{\mu}_{L},Q^{t}_{1})\leq\liminf_{L}F_{\frac{1}{L}}({\mu}_{L},Q_{1}^{t}). (3.17)

Then rescaling by LL the measure μ~L:=fL(μ^L)\tilde{\mu}_{L}:=f_{\sharp}^{L}(\hat{\mu}_{L}) satisfies supp(μ~Lbt1    t)QLt\mathrm{supp}(\tilde{\mu}_{L}-b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}t\mathbb{R})\subset\subset Q^{t}_{L}, and

m(L)1LF1(μ~L,QLt)=F1L(μ^L,Q1t)m^{(L)}\leq\frac{1}{L}F_{1}(\tilde{\mu}_{L},Q^{t}_{L})=F_{\frac{1}{L}}(\hat{\mu}_{L},Q_{1}^{t})

which together with (3.17) concludes the proof. ∎

Remark 3.10.

From the proof of the above proposition we also deduce that given bmb\in\mathbb{Z}^{m} and t𝒮n1t\in\mathcal{S}^{n-1} for all R>0R>0 there exists a sequence μjbtd1    t\mu_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes td\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}t\mathbb{R} such that supp(μjbtd1    t)QRt(xj){\mathrm{supp}}(\mu_{j}-b\otimes td\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}t\mathbb{R})\subset Q^{t}_{R}(x_{j}) and

ψhom(b,t)R=limjFεj(μj,QR(xj)).\psi_{\hom}(b,t)R=\lim_{j\to\infty}F_{\varepsilon_{j}}(\mu_{j},Q_{R}(x_{j})).

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 μjdf(m)(Ω)\mu_{j}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega), μjμ\mu_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu and εj0+\varepsilon_{j}\to 0^{+} one has

Fhom(μ)lim infjFεj(μj).F_{\mathrm{hom}}(\mu)\leq\liminf_{j\to\infty}F_{\varepsilon_{j}}(\mu_{j}).

Let εj\varepsilon_{j}, μj\mu_{j} be as above; it’s not restrictive to suppose that lim infjFεj(μj)\liminf_{j}F_{\varepsilon_{j}}(\mu_{j}) is finite and, by compactness (Theorem 3.1), μ=ϑτ1    γdf(m)(Ω)\mu=\vartheta\otimes\tau\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega).

We head to the conclusion in three steps.

Step 1. (localization and decomposition) For all jj\in\mathbb{N} we denote νj\nu_{j} the positive measures

νj=ψ(yεj,ϑj,τj)1    γj.\nu_{j}=\psi\bigg{(}\frac{y}{\varepsilon_{j}},\vartheta_{j},\tau_{j}\bigg{)}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma_{j}. (3.18)

which is the energy density of the localised functional

Fεj(μj,A)=Aγjψ(yεj,ϑj,τj)𝑑1.F_{\varepsilon_{j}}(\mu_{j},A)=\int_{A\cap\gamma_{j}}\psi\bigg{(}\frac{y}{\varepsilon_{j}},\vartheta_{j},\tau_{j}\bigg{)}d\mathcal{H}^{1}. (3.19)

By the assumptions on ψ\psi, the sequence νj\nu_{j} is equi-bounded hence, by compactness, there exists a positive Radon measure ν\nu on Ω\Omega such that, up to a subsequence, νjν\nu_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\nu. Now consider the Radon-Nikodym decomposition of the measure ν\nu with respect to the 11-dimensional Hausdorff measure restricted to γ\gamma, i.e.,

ν=g1    γ+νs\nu=g\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma+\nu^{s} (3.20)

where g=dνd1    γg=\frac{d\nu}{d\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma} is the density of the part of ν\nu which is absolutely continuous with respect to 1    γ\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma and the singular part is denoted by νs\nu^{s} (and it is a positive measure as well).

Step 2. (definition of the blow-up) Let x0Ωγx_{0}\in\Omega\cap\gamma be a Lebesgue point for gg with respect to 1    γ\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma. We can write

g(x0)=limϱ0+ν(Qϱt(x0))1(Qϱt(x0)γ)=limϱ0+ν(Qϱt(x0))ϱg(x_{0})=\lim_{\varrho\to 0^{+}}\frac{\nu(Q_{\varrho}^{t}(x_{0}))}{\mathcal{H}^{1}(Q_{\varrho}^{t}(x_{0})\cap\gamma)}=\lim_{\varrho\to 0^{+}}\frac{\nu(Q_{\varrho}^{t}(x_{0}))}{\varrho} (3.21)

where t=τ(x0)t=\tau(x_{0}) and the last equality holds for 1\mathcal{H}^{1}-a.e x0γx_{0}\in\gamma by the Besicovitch-Marstrand-Mattila Theorem. The Besicovitch derivation theorem ensures that 1\mathcal{H}^{1}-a.e x0Ωγx_{0}\in\Omega\cap\gamma is a Lebesgue point for ν\nu with respect to 1    γ\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma, i.e., a Lebesgue point for gg; moreover we can also assume that x0x_{0} is a Lebesgue point for ϑ\vartheta and τ\tau. By definition of 1\mathcal{H}^{1}-rectifiability and approximate tangent space and then we can assume that 1\mathcal{H}^{1}-a.e x0γΩx_{0}\in\gamma\cap\Omega satisfies the following property

ϑ(x0+ϱy)τ(x0+ϱy)1    γx0ϱbt1    (t)\vartheta(x_{0}+\varrho y)\otimes\tau(x_{0}+\varrho y)\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\frac{\gamma-x_{0}}{\varrho}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}) (3.22)

with b=ϑ(x0)b=\vartheta(x_{0}), in every open, bounded subset of n\mathbb{R}^{n}; in fact it follows from Theorem 3.13 that γ\gamma is the union of countably many closed Lipschitz curves on which ϑ\vartheta is constant, hence ϑ\vartheta is 1    γ\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma-measurable and integrable. Since ν\nu is finite, we have ν(Qϱt(x0))=0\nu(\partial Q_{\varrho}^{t}(x_{0}))=0 up to at most countably many ϱ>0\varrho>0. For all such ϱ\varrho it holds

ν(Qϱt(x0))=limjνj(Qϱt(x0)).\nu(Q_{\varrho}^{t}(x_{0}))=\lim_{j\to\infty}\nu_{j}(Q_{\varrho}^{t}(x_{0})). (3.23)

Moreover for every ϱ>0\varrho>0 we have

ϑj(x0+ϱy)τj(x0+ϱy)1    γjx0ϱϑ(x0+ϱy)τ(x0+ϱy)1    γx0ϱ\vartheta_{j}(x_{0}+\varrho y)\otimes\tau_{j}(x_{0}+\varrho y)\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\frac{\gamma_{j}-x_{0}}{\varrho}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\vartheta(x_{0}+\varrho y)\otimes\tau(x_{0}+\varrho y)\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\frac{\gamma-x_{0}}{\varrho} (3.24)

as j+j\to+\infty.

Then by a diagonalization argument on (3.21), (3.23), and (3.24) we can extract a subsequence ϱj\varrho_{j} such that

g(x0)=dνd1    γ(x0)=limjνj(Qϱjt(x0))1(Qϱjt(x0)γ).g(x_{0})=\frac{d\nu}{d\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma}(x_{0})=\lim_{j\to\infty}\frac{\nu_{j}(Q_{\varrho_{j}}^{t}(x_{0}))}{\mathcal{H}^{1}(Q_{\varrho_{j}}^{t}(x_{0})\cap\gamma)}. (3.25)
ϑj(x0+ϱjy)τj(x0+ϱjy)1    γjx0ϱjbt1    (t)\vartheta_{j}(x_{0}+\varrho_{j}y)\otimes\tau_{j}(x_{0}+\varrho_{j}y)\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\frac{\gamma_{j}-x_{0}}{\varrho_{j}}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}) (3.26)

and ϱjεj\frac{\varrho_{j}}{\varepsilon_{j}}\rightarrow\infty as jj\to\infty.

Step 3. (lower bound for the blow-up) Recalling the expression of νj\nu_{j} and (3.21), (3.25) is equivalent to

g(x0)=limj1ϱjQϱjt(x0)γjψ(yεj,ϑj,τj)𝑑1.g(x_{0})=\lim_{j\to\infty}\frac{1}{\varrho_{j}}\int_{Q_{\varrho_{j}}^{t}(x_{0})\cap\gamma_{j}}\psi\bigg{(}\frac{y}{\varepsilon_{j}},\vartheta_{j},\tau_{j}\bigg{)}d\mathcal{H}^{1}. (3.27)

By a change of variable we get

1ϱjQϱjt(x0)γjψ(yεj,ϑj,τj)𝑑1=Q1t(x0ϱj)γjψ(ϱjεjx,ϑj(ϱjx),τj(ϱjx))𝑑1.\frac{1}{\varrho_{j}}\int_{Q_{\varrho_{j}}^{t}(x_{0})\cap\gamma_{j}}\psi\bigg{(}\frac{y}{\varepsilon_{j}},\vartheta_{j},\tau_{j}\bigg{)}d\mathcal{H}^{1}=\int_{Q_{1}^{t}(\frac{x_{0}}{\varrho_{j}})\cap\gamma_{j}}\psi\bigg{(}\frac{\varrho_{j}}{\varepsilon_{j}}x,\vartheta_{j}(\varrho_{j}x),\tau_{j}(\varrho_{j}x)\bigg{)}d\mathcal{H}^{1}. (3.28)

Then denoting by μ~j=ϑj(ϱjx)τj(ϱjx)1    γϱj\tilde{\mu}_{j}=\vartheta_{j}(\varrho_{j}x)\otimes\tau_{j}(\varrho_{j}x)\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\frac{\gamma}{\varrho_{j}} which, in view of (3.26), satisfies μ~jbt1    (t)\widetilde{\mu}_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}b\otimes t\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(t\mathbb{R}), we have

g(x0)=limjFϱjεj(μ~j,Q1t(x0ϱj)).g(x_{0})=\lim_{j\to\infty}F_{\frac{\varrho_{j}}{\varepsilon_{j}}}\left(\tilde{\mu}_{j},Q_{1}^{t}\left(\displaystyle{\frac{x_{0}}{\varrho_{j}}}\right)\right).

Now applying Proposition 3.9 we conclude that

g(x0)ψhom(b,t).g(x_{0})\geq\psi_{\mathrm{hom}}(b,t).

Therefore for 1\mathcal{H}^{1}-a.e. x0Ωγx_{0}\in\Omega\cap\gamma one has

dνd1    γ(x0)ψhom(ϑ(x0),τ(x0)).\frac{d\nu}{d\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma}(x_{0})\geq\psi_{\mathrm{hom}}(\vartheta(x_{0}),\tau(x_{0})). (3.29)

Step 3. (conclusion) The liminf inequality is achieved by integrating over Ωγ\Omega\cap\gamma. In fact from (3.18)(\ref{eqo}) and (3.29)(\ref{eq7}) we get

ν(Ω)Ωγdνd1    γ𝑑1Ωγψhom(ϑ,τ)𝑑1.\nu(\Omega)\geq\int_{\Omega\cap\gamma}\frac{d\nu}{d\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma}d\mathcal{H}^{1}\geq\int_{\Omega\cap\gamma}\psi_{\mathrm{hom}}(\vartheta,\tau)d\mathcal{H}^{1}.

Since νjν\nu_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\nu, it follows that lim infjνj(Ω)ν(Ω)\liminf_{j\to\infty}\nu_{j}(\Omega)\geq\nu(\Omega) and so

lim infjFεj(μj)=lim infjνj(Ω)\displaystyle\liminf_{j\to\infty}F_{\varepsilon_{j}}(\mu_{j})=\liminf_{j\to\infty}\nu_{j}(\Omega) ν(Ω)\displaystyle\geq\nu(\Omega)
Ωγψhom(ϑ,τ)𝑑1=Fhom(μ)\displaystyle\geq\int_{\Omega\cap\gamma}\psi_{\mathrm{hom}}(\vartheta,\tau)d\mathcal{H}^{1}=F_{\mathrm{hom}}(\mu)

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 μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) we have to find a sequence μεjdf(m)(Ω)\mu_{\varepsilon_{j}}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) such that μεjμ\mu_{\varepsilon_{j}}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu and

lim supjFεj(μεj)Fhom(μ)\limsup_{j\to\infty}F_{\varepsilon_{j}}(\mu_{\varepsilon_{j}})\leq F_{\mathrm{hom}}(\mu)

for all εj0+\varepsilon_{j}\to 0^{+}. Using a standard diagonal argument it suffices to show the construction for a dense family. Here we consider the set of measures in df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega) which are supported on a polyhedral curve γ\gamma. This density result is a consequence of the corresponding result for currents (see Appendix).

Step 1: (polyhedral measures) Now let μ=i=1Nbiti1    γidf(m)(n)\mu=\sum_{i=1}^{N}b_{i}\otimes t_{i}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma_{i}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n}) be a polyhedral measure, in the sense that the γi\gamma_{i} are disjoint segments (up to the endpoints), bimb_{i}\in\mathbb{Z}^{m}, ti𝒮n1t_{i}\in\mathcal{S}^{n-1} for i=1,,Ni=1,\ldots,N. Let γ=i=1Nγi\gamma=\cup_{i=1}^{N}\gamma_{i}. We cover γΩ\gamma\cap\Omega, up to a 1\mathcal{H}^{1}-null set, with NN families of countably many disjoint cubes {Qik=Qrk,iti(xk,i)}k\{Q^{k}_{i}=Q^{t_{i}}_{r_{k,i}}(x_{k,i})\}_{k\in\mathbb{N}} and i=1,,Ni=1,\ldots,N which are contained in Ω\Omega and have the property that γQik=γiQik\gamma\cap Q^{k}_{i}=\gamma_{i}\cap Q^{k}_{i} with γi\gamma_{i} though the centre of the cube so that

μ    Qik=biti1    (xk,i+ti)\mu{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q^{k}_{i}=b_{i}\otimes t_{i}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(x_{k,i}+t_{i}\mathbb{R})

for i{1,,N}i\in\{1,\ldots,N\}. In particular krk,i=1(γi)\sum_{k}r_{k,i}=\mathcal{H}^{1}(\gamma_{i}) for every i=1,,Ni=1,\ldots,N.

By Proposition 3.9 and Remark 3.10, for all kk\in\mathbb{N} and i{1,,N}i\in\{1,\ldots,N\}, there exists a sequence μεjk,idf(m)(Qik)\mu_{\varepsilon_{j}}^{k,i}\in\mathcal{M}_{\mathrm{df}}^{(m)}(Q^{k}_{i}) such that μεjk,iμ    Qik\mu_{\varepsilon_{j}}^{k,i}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}Q^{k}_{i}, supp(μεjk,ibiti1    (xk,i+ti))Qik\mathrm{supp}(\mu_{\varepsilon_{j}}^{k,i}-b_{i}\otimes t_{i}\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(x_{k,i}+\mathbb{R}t_{i}))\subset Q^{k}_{i} and

limjFεj(μεjk,i,Qik)rk,iψhom(bi,ti).\lim_{j}F_{\varepsilon_{j}}(\mu_{\varepsilon_{j}}^{k,i},Q^{k}_{i})\leq r_{k,i}\psi_{\mathrm{hom}}(b_{i},t_{i}). (3.30)

Finally define μεj=i=1Nkμεjk,i\mu_{\varepsilon_{j}}=\sum_{i=1}^{N}\sum_{k\in\mathbb{N}}\mu_{\varepsilon_{j}}^{k,i}. By the properties of μjk,i\mu_{j}^{k,i} we have that μεjdf(m)(Ω)\mu_{\varepsilon_{j}}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega), μεjμ\mu_{\varepsilon_{j}}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mu and

limjFεj(μεj)Fhom(μ)\lim_{j}F_{\varepsilon_{j}}(\mu_{\varepsilon_{j}})\leq F_{\mathrm{hom}}(\mu)

which concludes the proof for μ\mu polyhedral.

Step 2: (general measures) Finally, we deduce the inequality for all measure μdf(m)(Ω)\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega). We first extend μ\mu to a measure μdf(m)(n)\mathcal{E}\mu\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n}), with

limδ0|μ|(ΩδΩ)=0\lim_{\delta\to 0}\lvert\mathcal{E}\mu\lvert(\Omega_{\delta}\setminus\Omega)=0

where Ωδ={x:dist(x,Ω)<δ}\Omega_{\delta}=\{x:\mathrm{dist}(x,\Omega)<\delta\} (see the Appendix). By Theorem 3.14 there exist a sequence of polyhedral measures μkdf(m)(n)\mu_{k}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\mathbb{R}^{n}) and a sequence of C1C^{1} and bi-Lipschitz maps fkf_{k} such that

|μk(fk)μ|(n)0,fkxL0,DfkIdL0.\lvert\mu_{k}-(f_{k})_{\sharp}\mathcal{E}\mu\lvert(\mathbb{R}^{n})\to 0,\,\,\|f_{k}-x\|_{L^{\infty}}\to 0,\,\,\|Df_{k}-\mathrm{Id}\|_{L^{\infty}}\to 0.

This implies μkμ\mu_{k}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}\mathcal{E}\mu. By Lemma 3.16, the continuity of ψhom\psi_{\mathrm{hom}} in t𝒮n1t\in\mathcal{S}^{n-1} and the invariance under deformations of the multiplicity map one obtains

Fhom(μk,Ω)\displaystyle F_{\mathrm{hom}}(\mu_{k},\Omega) Fhom((fk)μ,Ω)+cμk(fk)μ\displaystyle\leq F_{\mathrm{hom}}((f_{k})_{\sharp}\mathcal{E}\mu,\Omega)+c\|\mu_{k}-(f_{k})_{\sharp}\mathcal{E}\mu\|
Fhom(μk,Ωδk)(1+cDfkIdL)+cμk(fk)μ\displaystyle\leq F_{\mathrm{hom}}(\mathcal{E}\mu_{k},\Omega_{\delta_{k}})(1+c\|Df_{k}-\mathrm{Id}\|_{L^{\infty}})+c\|\mu_{k}-(f_{k})_{\sharp}\mathcal{E}\mu\| (3.31)

where δk=fkxL0\delta_{k}=\|f_{k}-x\|_{L^{\infty}}\to 0. Taking the limit in (3.3) we get

lim supkFhom(μk,Ω)Fhom(μ,Ω)=Fhom(μ).\limsup_{k\to\infty}F_{\mathrm{hom}}(\mu_{k},\Omega)\leq F_{\mathrm{hom}}(\mu,\Omega)=F_{\mathrm{hom}}(\mu). (3.32)

It then follows from the lower semicontinuity with respect to the weak convergence of Γ-lim supjFεj\Gamma\textrm{-}\limsup_{j}F_{\varepsilon_{j}}, the definition of Γ-lim sup\Gamma\textrm{-}\limsup, Step 1 and (3.32) that

Γ-lim supjFεj(μ)\displaystyle\Gamma{\operatorname{\mathit{-}}}\limsup_{j}F_{\varepsilon_{j}}(\mu) lim infk(Γ-lim supjFεj(μk))\displaystyle\leq\liminf_{k}(\Gamma{\operatorname{\mathit{-}}}\limsup_{j}F_{\varepsilon_{j}}(\mu_{k}))
lim infkFhom(μk)lim supkFhom(μk)Fhom(μ)\displaystyle\leq\liminf_{k}F_{\mathrm{hom}}(\mu_{k})\leq\limsup_{k}F_{\mathrm{hom}}(\mu_{k})\leq F_{\mathrm{hom}}(\mu)

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 T=|ϑ|1    γ\|T\|=\lvert\vartheta\lvert\mathcal{H}^{1}{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\gamma, its mass is

𝐌(T)=T(Ω)=γ|ϑ|d1,\mathbf{M}(T)=\|T\|(\Omega)=\int_{\gamma}\lvert\vartheta\lvert d\mathcal{H}^{1},

and it gives the weighted length of the current TT with respect to the Euclidean norm ||\lvert\cdot\lvert on m\mathbb{Z}^{m}. Moreover for any open subset WΩW\subset\Omega we denote 𝐌W(T)=T(W)\mathbf{M}_{W}(T)=\|T\|(W) and we can define the support of the current TT as suppT\mathrm{supp}\|T\|

Flat norm: The flat norm of the kk-current TT is defined as follows: for all WΩW\subset\subset\Omega let

𝐅W(T):=inf{𝐌W(A)+\displaystyle\mathbf{F}_{W}(T):=\inf\{\mathbf{M}_{W}(A)+ 𝐌W(B):T=A+B\displaystyle\mathbf{M}_{W}(B):T=A+\partial B
Ak(Ω,m),Bk+1(Ω,m)}.\displaystyle A\in\mathcal{R}_{k}(\Omega,\mathbb{Z}^{m}),\,B\in\mathcal{R}_{k+1}(\Omega,\mathbb{Z}^{m})\}.

For example the flat norm of a 0-current given by to point x1x_{1} and x2x_{2} with multiplicity +1+1 and 1-1 respectively is the length of the segment connecting x1x_{1} and x2x_{2}.

Theorem 3.11 (Theorem 31.2, [11]).

Let Tj,T1(Ω,m)T_{j},T\in\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) be such that

supj1(𝐌W(Tj)+𝐌W(Tj))<\sup_{j\geq 1}(\mathbf{M}_{W}(T_{j})+\mathbf{M}_{W}(\partial T_{j}))<\infty

for all WΩW\subset\subset\Omega. Then TjTT_{j}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}T if and only if 𝐅W(TjT)0\mathbf{F}_{W}(T_{j}-T)\to 0 for every WΩW\subset\subset\Omega.

Slicing of 11-currents: For a current T1(Ω,m)T\in\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}) such that 𝐌W(T)+𝐌W(T)<\mathbf{M}_{W}(T)+\mathbf{M}_{W}(\partial T)<\infty for all WΩW\subset\subset\Omega and a Lipschitz function f:nf:\mathbb{R}^{n}\rightarrow\mathbb{R}, one can define the slice of TT through ff in tt\in\mathbb{R} as

T[f,t]=(T    {x:f(x)<t})(T)    {x:f(x)<t}T[f,t-]=\partial(T{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)<t\})-(\partial T){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)<t\}
T[f,t+]=(T)    {x:f(x)>t}(T    {x:f(x)>t}).T[f,t+]=(\partial T){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)>t\}-\partial(T{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)>t\}).

Up to at most a countable set of point tt\in\mathbb{R} for which 𝐌(T    {x:f(x)=t})+𝐌((T)    {x:f(x)=t})>0\mathbf{M}(T{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)=t\})+\mathbf{M}((\partial T){\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\{x:f(x)=t\})>0, it holds

T[f,t]=T[f,t+]=:T[f,t].T[f,t-]=T[f,t+]=:T[f,t].

These are the main properties of the slicing that we need in the paper (see Section 4.2.1 in [8]):

  • (i)\mathrm{(i)}

    suppT[f,t+]f1{t}suppT\mathrm{supp}T[f,t+]\subset f^{-1}\{t\}\cap\mathrm{supp}T;

  • (ii)\mathrm{(ii)}

    it holds

    ab𝐌(T[f,t+])𝑑tLip(f)T{x:a<f(x)<b}.\int_{a}^{b}\mathbf{M}(T[f,t+])\,dt\leq\mathrm{Lip}(f)\|T\|\{x:a<f(x)<b\}.

    for all a<b-\infty\leq a<b\leq\infty;

  • (iii)\mathrm{(iii)}

    T[f,t+]T[f,t+] is a 0-current for a.e. tt\in\mathbb{R};

  • (iv)\mathrm{(iv)}

    if suppTK\mathrm{supp}T\subset K, KnK\subset\mathbb{R}^{n} a compact set, then

    𝐅K{x:f(x)=t}(T[f,t+])𝑑tLip(f)𝐅K(T).\int\mathbf{F}_{K\cap\{x:f(x)=t\}}(T[f,t+])\,dt\leq\mathrm{Lip}(f)\mathbf{F}_{K}(T).

Push forward of 11-currents: For a bi-lipschitz map f:nnf:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n}, the push-forward fTf_{\sharp}T of TT is the current

fT,φ=f(γ)ϑ(f1(y))φ(y);τ(y)𝑑1(y),\langle f_{\sharp}T,\varphi\rangle=\int_{f(\gamma)}\vartheta(f^{-1}(y))\langle\varphi(y);\tau^{\prime}(y)\rangle d\mathcal{H}^{1}(y), (3.33)

where τ\tau^{\prime} it the tangent to f(γ)f(\gamma) with the same orientation of τ\tau, τ(f(x))=Dτf(x)/|Dτf(x)|\tau^{\prime}(f(x))=D_{\tau}f(x)/\lvert D_{\tau}f(x)\lvert and Dτf(x)D_{\tau}f(x) denotes the tangential derivative of ff along γ\gamma, which exists 1\mathcal{H}^{1}-a.e. on γ\gamma since ff is Lipschitz on γ\gamma; if ff is differentiable in xx then Dτf(x)=Df(x)τ(x)D_{\tau}f(x)=Df(x)\tau(x).


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 m\mathbb{Z}^{m}-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 df(m)(Ω)\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega).

Theorem 3.12 (Compactness).

Let (Tj)j(T_{j})_{j\in\mathbb{N}} be a sequence of rectifiable 1-currents without boundary in 1(n,m)\mathcal{R}_{1}(\mathbb{R}^{n},\mathbb{Z}^{m}). If

supj𝐌(Tj)<\sup_{j\in\mathbb{N}}\mathbf{M}(T_{j})<\infty

then there are a current T1(n,m)T\in\mathcal{R}_{1}(\mathbb{R}^{n},\mathbb{Z}^{m}) and a subsequence (Tjk)k(T_{j_{k}})_{k\in\mathbb{N}} such that

TjkT.T_{j_{k}}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}T.

The second theorem gives a characterization of the support of a closed 1-rectifiable current.

Theorem 3.13 (Structure).

Let T1(n;m)T\in\mathcal{R}_{1}(\mathbb{R}^{n};\mathbb{Z}^{m}) with T=0\partial T=0 and 𝐌(T)<\mathbf{M}(T)<\infty. Then there are countably many oriented Lipschitz curves γi\gamma_{i} with tangent vector fields τi:γi𝒮n1\tau_{i}:\gamma_{i}\rightarrow\mathcal{S}^{n-1} and multiplicities ϑim\vartheta_{i}\in\mathbb{Z}^{m} such that

T,φ=iϑiγiφ(x);τi(x)𝑑1(x)φCc(n,n).\langle T,\varphi\rangle=\sum_{i\in\mathbb{N}}\vartheta_{i}\int_{\gamma_{i}}\langle\varphi(x);\tau_{i}(x)\rangle d\mathcal{H}^{1}(x)\quad\forall\varphi\in C_{c}^{\infty}(\mathbb{R}^{n},\mathbb{R}^{n}).

Further,

i|ϑi|1(γi)m𝐌(T).\sum_{i}\lvert\vartheta_{i}\lvert\mathcal{H}^{1}(\gamma_{i})\leq\sqrt{m}\mathbf{M}(T).

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 kk-currents con be found in [3]).

Theorem 3.14 (Density).

Fix ε>0\varepsilon>0 and consider a m\mathbb{Z}^{m}-valued closed 1-current T1(Ω,m)T\in\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}). Then there exist a bijective map fC1(n;n)f\in C^{1}(\mathbb{R}^{n};\mathbb{R}^{n}), with inverse also C1C^{1}, and a closed polyhedral 1-current P𝒫1(n,m)P\in\mathcal{P}_{1}(\mathbb{R}^{n},\mathbb{Z}^{m}) such that

𝐌(fTP)ε\mathbf{M}(f_{\sharp}T-P)\leq\varepsilon

and

|Df(x)Id|+|f(x)x|εxn.\lvert Df(x)-\mathrm{Id}\lvert+\lvert f(x)-x\lvert\leq\varepsilon\quad\forall x\in\mathbb{R}^{n}.

Moreover, f(x)=xf(x)=x whenever dist(x,suppT)ε\mathrm{dist}(x,\mathrm{supp}T)\geq\varepsilon.

It is important to notice that the deformation result given above guarantees a current TT 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 n\mathbb{R}^{n} but a local version can be deduced using the extension lemma recalled below [5, Lemma 2.3].

Lemma 3.15 (Extension).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz open set. For every closed rectifiable 1-current defined in Ω\Omega, T1(Ω,m)T\in\mathcal{R}_{1}(\Omega,\mathbb{Z}^{m}), there exists a closed rectifiable 1-current T1(n;m)\mathcal{E}T\in\mathcal{R}_{1}(\mathbb{R}^{n};\mathbb{Z}^{m}) with T    Ω=T\mathcal{E}T{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}\Omega=T and 𝐌(T)c𝐌(T)\mathbf{M}(\mathcal{E}T)\leq c\mathbf{M}(T). The constant cc depends only on Ω\Omega. Moreover, limδ0𝐌(T    (ΩδΩ))=0\lim_{\delta\to 0}\mathbf{M}(\mathcal{E}T{\mathchoice{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to7.0pt{\vfill\hrule width=7.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to5.0pt{\vfill\hrule width=5.0pt,height=0.2pt}}\,}{\,\hbox{\vrule width=0.2pt\vbox to3.35pt{\vfill\hrule width=3.35pt,height=0.2pt}}\,}}(\Omega_{\delta}\setminus\Omega))=0, where Ωδ={x:dist(x,Ω)<δ}\Omega_{\delta}=\{x:\mathrm{dist}(x,\Omega)<\delta\}.

Finally we include the following lemma [5, Lemma 3.3], again useful in proving the limsup inequality.

Lemma 3.16.

Assume that ψ:m×𝒮n1[0,)\psi:\mathbb{Z}^{m}\times\mathcal{S}^{n-1}\rightarrow[0,\infty) is Borel measurable, obeys ψ(0,t)=0\psi(0,t)=0, ψ(b,t)c|b|\psi(b,t)\geq c\lvert b\lvert and

|ψ¯(b,t)ψ¯(b,t)|c|bb|+c|b||tt|.\lvert\bar{\psi}(b,t)-\bar{\psi}(b^{\prime},t^{\prime})\lvert\leq c\lvert b-b^{\prime}\lvert+c\lvert b\lvert\lvert t-t^{\prime}\lvert.

Let μ,μdf(m)(Ω)\mu,\mu^{\prime}\in\mathcal{M}_{\mathrm{df}}^{(m)}(\Omega). Then for any open set ωΩ\omega\subset\Omega we have

|E(μ,ω)E(μ,ω)|c|μμ|(ω).\lvert E(\mu,\omega)-E(\mu^{\prime},\omega)\lvert\leq c\lvert\mu-\mu^{\prime}\lvert(\omega).

Further, if f:nnf:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n} is bi-Lipschitz then for any open set ωn\omega\subset\mathbb{R}^{n}

|E(μ,ω)E(fμ,f(ω))|cE(μ,ω)DfIdL.\lvert E(\mu,\omega)-E(f_{\sharp}\mu,f(\omega))\lvert\leq cE(\mu,\omega)\|Df-\mathrm{Id}\|_{L^{\infty}}. (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 Γ\Gamma-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 11-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: Γ\Gamma-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 L1L^{1}, 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.