• Non ci sono risultati.

The Steiner tree problem revisited through rectifiable G-currents

N/A
N/A
Protected

Academic year: 2021

Condividi "The Steiner tree problem revisited through rectifiable G-currents"

Copied!
21
0
0

Testo completo

(1)

Research Article

Andrea Marchese and Annalisa Massaccesi

The Steiner tree problem revisited through

rectifiable 𝐺-currents

Abstract: The Steiner tree problem can be stated in terms of finding a connected set of minimal length containing a given set of finitely many points. We show how to formulate it as a mass-minimization problem for 1-dimensional currents with coefficients in a suitable normed group. The representation used for these currents allows to state a calibration principle for this problem. We also exhibit calibrations in some examples.

Keywords: Steiner tree problem, rectifiable currents, calibration, flat 𝐺-chains MSC 2010: 49Q15, 49Q20

||

Andrea Marchese: Max-Planck-Institut für Mathematik in den Naturwissenschaften, Inselstraße 22, 04103 Leipzig, Germany,

e-mail: marchese@mis.mpg.de

Annalisa Massaccesi: Institut fΓΌr Mathematik der UniversitΓ€t ZΓΌrich, Winterthurerstrasse 190, CH-8057 ZΓΌrich, Switzerland,

e-mail: annalisa.massaccesi@math.uzh.ch

Communicated by: Giuseppe Mingione

Introduction

The classical Steiner tree problem consists in finding the shortest connected set containing 𝑛 given distinct points 𝑝1, . . . , 𝑝𝑛in ℝ𝑑. Some very well-known examples are shown in Figure 1.

DOI 10.1515/acv-2014-0022 | Adv. Calc. Var. 2014; ??? (???):1–22

Research Article

Andrea Marchese and Annalisa Massaccesi

The Steiner tree problem revisited through

rectifiable 𝐺-currents

Abstract: The Steiner tree problem can be stated in terms of finding a connected set of minimal length

containing a given set of finitely many points. We show how to formulate it as a mass-minimization problem for 1-dimensional currents with coefficients in a suitable normed group. The representation used for these currents allows to state a calibration principle for this problem. We also exhibit calibrations in some examples.

Keywords: Steiner tree problem, rectifiable currents, calibration, flat 𝐺-chains MSC 2010: 49Q15, 49Q20

||

Andrea Marchese: Max-Planck-Institut für Mathematik in den Naturwissenschaften, Inselstraße 22, 04103 Leipzig, Germany,

e-mail: marchese@mis.mpg.de

Annalisa Massaccesi: Institut fΓΌr Mathematik der UniversitΓ€t ZΓΌrich, Winterthurerstrasse 190, CH-8057 ZΓΌrich, Switzerland,

e-mail: annalisa.massaccesi@math.uzh.ch

Communicated by: Guiseppe Mingione

Introduction

The classical Steiner tree problem consists in finding the shortest connected set containing 𝑛 given distinct points 𝑝1, . . . , 𝑝𝑛in ℝ𝑑. Some very well-known examples are shown in Figure 1.

𝑝

4

𝑝

3

𝑝

1

𝑝

2

𝑂

𝑝

1

𝑝

2

𝑝

3

𝑂

Figure 1. Solutions for the vertices of an equilateral triangle and a square.

The problem is completely solved in ℝ2and there exists a wide literature on the subject, mainly devoted

to improving the efficiency of algorithms for the construction of solutions: see, for instance, [13] and [14] for a survey of the problem. The recent papers [22] and [23] witness the current studies on the problem and its generalizations.

Our aim is to rephrase the Steiner tree problem as an equivalent mass-minimization problem by replacing connected sets with 1-currents with coefficients in a more suitable group than β„€, in such a way that solutions

Figure 1. Solutions for the vertices of an equilateral triangle and a square.

The problem is completely solved in ℝ2and there exists a wide literature on the subject, mainly devoted to improving the efficiency of algorithms for the construction of solutions: see, for instance, [13] and [14] for a survey of the problem. The recent papers [22] and [23] witness the current studies on the problem and its generalizations.

Our aim is to rephrase the Steiner tree problem as an equivalent mass-minimization problem by replacing connected sets with 1-currents with coefficients in a more suitable group than β„€, in such a way that solutions of one problem correspond to solutions of the other, and vice-versa. The use of currents allows to exploit techniques and tools from the Calculus of Variations and the Geometric Measure Theory.

(2)

20 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

2 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

of one problem correspond to solutions of the other, and vice-versa. The use of currents allows to exploit techniques and tools from the Calculus of Variations and the Geometric Measure Theory.

Let us briefly point out a few facts suggesting that classical polyhedral chains with integer coefficients might not be the correct environment for our problem. First of all, one should make the given points 𝑝1, . . . , 𝑝𝑛

in the Steiner problem correspond to some integral polyhedral 0-chain supported on 𝑝1, . . . , 𝑝𝑛, with suitable

multiplicities π‘š1, . . . , π‘šπ‘›. One has to impose that π‘š1+ β‹… β‹… β‹… + π‘šπ‘›= 0 in order that this 0-chain is the boundary

of a compactly supported 1-chain. In the example of the equilateral triangle, see Figure 1, the condition π‘š3= βˆ’(π‘š1+ π‘š2) forces to break symmetry, leading to the minimizer in Figure 2. The desired solution is

instead depicted in Figure 1. In the second example from Figure 1, we get the β€œwrong” non-connected minimizer even though all boundary multiplicities have modulus 1; see Figure 2.

𝑂

𝑂

1

1

1

βˆ’1

βˆ’1

1

1

βˆ’2

1

Figure 2. Solutions for the mass-minimization problems among polyhedral chains with integer coefficients.

These examples show that β„€ is not the right group of coefficients.

Our framework will be that of currents with coefficients in a normed abelian group 𝐺 (briefly: 𝐺-currents), which we will introduce in Section 1.

Currents with coefficients in a group were introduced by W. Fleming. There is a vast literature on the subject: let us mention only the seminal paper [12], the work of B. White [26, 27], and the more recent papers by T. De Pauw and R. Hardt [8] and by L. Ambrosio and M. G. Katz [3]. A closure theorem holds for these flat 𝐺-chains, see [12] and [26].

In Section 2 we recast the Steiner problem in terms of a mass-minimization problem over currents with coefficients in a discrete group 𝐺, chosen only on the basis of the number of boundary points. As we already said, this construction provides a way to pass from a mass-minimizer to a Steiner solution and vice-versa.

This new formulation permits to initiate a study of calibrations as a sufficient condition for minimality; this is the subject of Section 3. Classically a calibration πœ” associated with a given oriented π‘˜-submanifold 𝑆 βŠ‚ ℝ𝑑is a unit closed π‘˜-form taking value 1 on the tangent space of 𝑆. The existence of a calibration

guaran-tees the minimality of 𝑆 among oriented submanifolds with the same boundary πœ•π‘†. Indeed, Stokes’ theorem and the assumptions on πœ” imply that

vol(𝑆) = ∫

𝑆

πœ” = ∫

𝑆󸀠

πœ” ≀ vol(𝑆󸀠),

for any submanifold 𝑆󸀠having the same boundary of 𝑆.

In order to define calibrations in the framework of 𝐺-currents, it is convenient to view currents as linear functionals on forms, which is not always possible in the usual setting of currents with coefficients in groups. This motivates the preliminary work in Section 1, where we embed the group 𝐺 in a normed linear space 𝐸 and we construct the currents with coefficients in 𝐸 in the classical way. In Definition 3.5, the notion of calibration is slightly weakened in order to include piecewise smooth forms, which appear in

Figure 2. Solutions for the mass-minimization problems among polyhedral chains with integer coefficients.

Let us briefly point out a few facts suggesting that classical polyhedral chains with integer coefficients might not be the correct environment for our problem. First of all, one should make the given points 𝑝1, . . . , 𝑝𝑛 in the Steiner problem correspond to some integral polyhedral 0-chain supported on 𝑝1, . . . , 𝑝𝑛, with suitable multiplicities π‘š1, . . . , π‘šπ‘›. One has to impose that π‘š1+ β‹… β‹… β‹… + π‘šπ‘›= 0 in order that this 0-chain is the boundary of a compactly supported 1-chain. In the example of the equilateral triangle, see Figure 1, the condition π‘š3= βˆ’(π‘š1+ π‘š2) forces to break symmetry, leading to the minimizer in Figure 2. The desired solution is instead depicted in Figure 1. In the second example from Figure 1, we get the β€œwrong” non-connected minimizer even though all boundary multiplicities have modulus 1; see Figure 2.

These examples show that β„€ is not the right group of coefficients.

Our framework will be that of currents with coefficients in a normed abelian group 𝐺 (briefly: 𝐺-currents), which we will introduce in Section 1.

Currents with coefficients in a group were introduced by W. Fleming. There is a vast literature on the subject: let us mention only the seminal paper [12], the work of B. White [26, 27], and the more recent papers by T. De Pauw and R. Hardt [8] and by L. Ambrosio and M. G. Katz [3]. A closure theorem holds for these flat 𝐺-chains, see [12] and [26].

In Section 2 we recast the Steiner problem in terms of a mass-minimization problem over currents with coefficients in a discrete group 𝐺, chosen only on the basis of the number of boundary points. As we already said, this construction provides a way to pass from a mass-minimizer to a Steiner solution and vice-versa.

This new formulation permits to initiate a study of calibrations as a sufficient condition for minimality; this is the subject of Section 3. Classically a calibration πœ” associated with a given oriented π‘˜-submanifold 𝑆 βŠ‚ ℝ𝑑is a unit closed π‘˜-form taking value 1 on the tangent space of 𝑆. The existence of a calibration guaran-tees the minimality of 𝑆 among oriented submanifolds with the same boundary πœ•π‘†. Indeed, Stokes’ theorem and the assumptions on πœ” imply that

vol(𝑆) = ∫ 𝑆

πœ” = ∫ 𝑆󸀠

πœ” ≀ vol(𝑆󸀠), for any submanifold 𝑆󸀠having the same boundary of 𝑆.

In order to define calibrations in the framework of 𝐺-currents, it is convenient to view currents as linear functionals on forms, which is not always possible in the usual setting of currents with coefficients in groups. This motivates the preliminary work in Section 1, where we embed the group 𝐺 in a normed linear space 𝐸 and we construct the currents with coefficients in 𝐸 in the classical way. In Definition 3.5, the notion of calibration is slightly weakened in order to include piecewise smooth forms, which appear in Examples 3.10 and 3.11, where we exhibit calibrations for the problem on the right of Figure 1 and for the Steiner tree problem on the vertices of a regular hexagon plus the center. It is worthwhile to note that our theory works for the Steiner tree problem in ℝ𝑑and for currents supported in ℝ𝑑; we made explicit computa-tions only on 2-dimensional configuracomputa-tions for simplicity reasons. We conclude Section 3 with some remarks concerning the use of calibrations in similar contexts, see for instance [19].

(3)

The existence of a calibration is a sufficient condition for a manifold to be a minimizer; one could wonder whether this condition is necessary as well. In general, a smooth (or piecewise smooth, according to Definition 3.7) calibration might not exist; nevertheless, one can still search for some weak calibration, for instance a differential form with bounded measurable coefficients. In Section 4 we discuss a strategy in order to get the existence of such a weak calibration. A duality argument due to H. Federer [11] ensures that a weak calibration exists for mass-minimizing normal currents; the same argument works for mass-minimizing normal currents with coefficients in the normed vector space 𝐸. Therefore an equivalence principle between minima among normal and rectifiable 1-currents with coefficients in 𝐸 and 𝐺, respectively, is sufficient to conclude that a calibration exists. Proposition 4.3 guarantees that the equivalence between minima holds in the case of classical 1-currents with real coefficients; hence a weak calibration always exists. The proof of this result is subject to the validity of a homogeneity property for the candidate minimizer stated in Remark 4.4. Example 4.5 shows that for 1-dimensional 𝐺-currents an interesting new phenomenon occurs, since (at least in a non-Euclidean setting) this homogeneity property might not hold; the validity of the homogeneity prop-erty may be related to the ambient space. The problem of the existence of a calibration in the Euclidean space is still open.

1 Rectifiable currents over a coefficient group

In this section we provide definitions for currents over a coefficient group, with some basic examples. Fix an open set π‘ˆ βŠ‚ ℝ𝑑and a normed vector space (𝐸, β€– β‹… β€–

𝐸) with finite dimension π‘š β‰₯ 1. We will denote by (πΈβˆ—, β€– β‹… β€–

πΈβˆ—) its dual space endowed with the dual norm

β€–π‘“β€–πΈβˆ—:= sup

‖𝑣‖𝐸≀1

βŸ¨π‘“; π‘£βŸ©. Definition 1.1. We say that a map

πœ” : Ξ›π‘˜(ℝ𝑑) Γ— 𝐸 β†’ ℝ is an πΈβˆ—-valued π‘˜-covector in ℝ𝑑if

(i) for all 𝜏 ∈ Ξ›π‘˜(ℝ𝑑), πœ”(𝜏, β‹… ) ∈ πΈβˆ—, that is, πœ”(𝜏, β‹… ) : 𝐸 β†’ ℝ is a linear function. (ii) for all 𝑣 ∈ 𝐸, πœ”( β‹… , 𝑣) : Ξ›π‘˜(ℝ𝑑) β†’ ℝ is a (classical) π‘˜-covector.

Sometimes we will use βŸ¨πœ”; 𝜏, π‘£βŸ© instead of πœ”(𝜏, 𝑣), in order to simplify the notation. The space of πΈβˆ—-valued π‘˜-covectors in ℝ𝑑is denoted by Ξ›π‘˜

𝐸(ℝ𝑑) and it is endowed with the comass norm

β€–πœ”β€– := sup{β€–πœ”(𝜏, β‹… )β€–πΈβˆ—: |𝜏| ≀ 1, 𝜏 simple}. (1.1)

Remark 1.2. Fix an orthonormal system of coordinates in ℝ𝑑, (e

1, . . . , e𝑑); the corresponding dual base in (ℝ𝑑)βˆ—is (𝑑π‘₯

1, . . . , 𝑑π‘₯𝑑). Consider a complete biorthonormal system for 𝐸, i.e. a pair (𝑣1, . . . , π‘£π‘š) ∈ πΈπ‘š, (𝑀1, . . . , π‘€π‘š) βŠ‚ (πΈβˆ—)π‘š

such that ‖𝑣𝑖‖𝐸= 1, β€–π‘€π‘–β€–πΈβˆ— = 1 and βŸ¨π‘€π‘–; π‘£π‘—βŸ© = 𝛿𝑖𝑗. Given an πΈβˆ—-valued π‘˜-covector πœ”, we denote

πœ”π‘—:= πœ”( β‹… , 𝑣 𝑗).

For each 𝑗 ∈ {1, . . . , π‘š}, πœ”π‘— is a π‘˜-covector in the usual sense. Hence the biorthonormal system (𝑣

1, . . . , π‘£π‘š), (𝑀1, . . . , π‘€π‘š) allows to write πœ” in β€œcomponents”

πœ” = (πœ”1, . . . , πœ”π‘š), in fact we have πœ”(𝜏, 𝑣) =βˆ‘π‘š 𝑗=1βŸ¨πœ” 𝑗; πœβŸ©βŸ¨π‘€ 𝑗; π‘£βŸ©. In particular πœ”π‘—admits the usual representation

πœ”π‘— = βˆ‘

1≀𝑖1<β‹…β‹…β‹…<π‘–π‘˜β‰€π‘‘

(4)

22 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

Definition 1.3. An πΈβˆ—-valued differential π‘˜-form in π‘ˆ βŠ‚ ℝ𝑑, or just a π‘˜-form when it is clear which vector space we are referring to, is a map

πœ” : π‘ˆ β†’ Ξ›π‘˜πΈ(ℝ𝑑);

we say that πœ” isC∞-regular if every component πœ”π‘—is so (see Remark 1.2). We denote byCπ‘βˆž(π‘ˆ, Ξ›π‘˜πΈ(ℝ𝑑)) the

vector space ofC∞-regular πΈβˆ—-valued π‘˜-forms with compact support in π‘ˆ.

We are mainly interested in πΈβˆ—-valued 1-forms, nevertheless we analyze π‘˜-forms in wider generality, in order to ease other definitions, such as the differential of an πΈβˆ—-valued form and the boundary of an 𝐸-current.

Definition 1.4. We define the differential dπœ” of aC∞-regular πΈβˆ—-valued π‘˜-form πœ” by components:

dπœ”π‘—:= d(πœ”π‘—) : π‘ˆ β†’ Ξ›π‘˜+1(ℝ𝑑), 𝑗 = 1, . . . , π‘š,

Moreover,Cπ‘βˆž(π‘ˆ, Ξ›1𝐸(ℝ𝑑)) has a norm, denoted by β€– β‹… β€–, given by the supremum of the comass norm of the

form defined in (1.1). Hence we mean

β€–πœ”β€– := sup

π‘₯βˆˆπ‘ˆβ€–πœ”(π‘₯)β€–. (1.2)

Definition 1.5. A π‘˜-dimensional current 𝑇 in π‘ˆ βŠ‚ ℝ𝑑, with coefficients in 𝐸, or just an 𝐸-current when there is no doubt on the dimension, is a linear and continuous function

𝑇 :Cπ‘βˆž(π‘ˆ, Ξ›π‘˜πΈ(ℝ𝑑)) β†’ ℝ,

where the continuity is meant with respect to the locally convex topology on the space Cπ‘βˆž(π‘ˆ, Ξ›π‘˜πΈ(ℝ𝑑)),

built in analogy with the topology onCπ‘βˆž(ℝ𝑛), with respect to which distributions are dual. This defines the

weakβˆ—topology on the space of π‘˜-dimensional 𝐸-currents. Convergence in this topology is equivalent to the convergence of all the β€œcomponents” in the space of classicalΒΉ π‘˜-currents, by which we mean the following. We define for every π‘˜-dimensional 𝐸-current 𝑇 its components 𝑇𝑗, for 𝑗 = 1, . . . , π‘š, and we write

𝑇 = (𝑇1, . . . , π‘‡π‘š), denoting

βŸ¨π‘‡π‘—; πœ‘βŸ© := βŸ¨π‘‡; Μƒπœ‘π‘—βŸ©,

for every (classical) compactly supported differential π‘˜-form πœ‘ on ℝ𝑑. Here Μƒπœ‘

𝑗denotes the πΈβˆ—-valued differen-tial π‘˜-form on ℝ𝑑such that

Μƒπœ‘π‘—( β‹… , 𝑣𝑗) = πœ‘, (1.3)

Μƒπœ‘π‘—( β‹… , 𝑣𝑖) = 0 for 𝑖 ΜΈ= 𝑗. (1.4)

It turns out that a sequence of π‘˜-dimensional 𝐸-currents π‘‡β„Žweaklyβˆ—converges to an 𝐸-current 𝑇 (in this case we write π‘‡β„Žβ‡€ 𝑇) if and only if the sequence of the components π‘‡βˆ— β„Žπ‘— converge to 𝑇𝑗 in the space of classical π‘˜-currents, for 𝑗 = 1, . . . , π‘š.

Definition 1.6. For a π‘˜-current 𝑇 over 𝐸 we define the boundary operator

βŸ¨πœ•π‘‡; πœ‘βŸ© := βŸ¨π‘‡; dπœ‘βŸ© for all πœ‘ = (πœ‘1, . . . , πœ‘π‘š) ∈Cπ‘βˆž(π‘ˆ, Ξ›π‘˜βˆ’1𝐸 (ℝ𝑑))

and the mass

𝕄(𝑇) := sup β€–πœ”β€–β‰€1βŸ¨π‘‡; πœ”βŸ©.

As one can expect, the boundary πœ•(𝑇𝑗) of every component 𝑇𝑗 is the relative component (πœ•π‘‡)𝑗of the boundary πœ•π‘‡.

(5)

Definition 1.7. A π‘˜-dimensional normal 𝐸-current in π‘ˆ βŠ‚ ℝ𝑑is an 𝐸-current 𝑇 with the properties 𝕄(𝑇) < +∞ and 𝕄(πœ•π‘‡) < +∞. Thanks to the Riesz Theorem, 𝑇 admits the following representation:

βŸ¨π‘‡; πœ”βŸ© = ∫ π‘ˆ

βŸ¨πœ”(π‘₯); 𝜏(π‘₯), 𝑣(π‘₯)⟩ dπœ‡π‘‡(π‘₯) for all πœ” ∈Cπ‘βˆž(π‘ˆ, Ξ›π‘˜πΈ(ℝ𝑑)),

where πœ‡π‘‡is a Radon measure on π‘ˆ, 𝑣 : π‘ˆ β†’ 𝐸 is summable with respect to πœ‡π‘‡and |𝜏| = 1, πœ‡π‘‡-a.e. A similar representation holds for the boundary πœ•π‘‡.

Definition 1.8. A rectifiable π‘˜-current 𝑇 in π‘ˆ βŠ‚ ℝ𝑑, over 𝐸, or a rectifiable 𝐸-current is an 𝐸-current admitting the following representation:

βŸ¨π‘‡; πœ”βŸ© := ∫ Ξ£

βŸ¨πœ”(π‘₯); 𝜏(π‘₯), πœƒ(π‘₯)⟩ dHπ‘˜(π‘₯) for all πœ” ∈Cπ‘βˆž(ℝ𝑑, Ξ›π‘˜πΈ(π‘ˆ))

where Ξ£ is a countably π‘˜-rectifiable set (see [16, Definition 5.4.1]) contained in π‘ˆ, 𝜏(π‘₯) ∈ 𝑇π‘₯Ξ£ with |𝜏(π‘₯)| = 1 forHπ‘˜-a.e. π‘₯ ∈ Ξ£ and πœƒ ∈ 𝐿1(Hπ‘˜ Ξ£; 𝐸). We will refer to such a current as 𝑇 = 𝑇(Ξ£, 𝜏, πœƒ). If 𝐡 is a Borel set and

𝑇(Ξ£, 𝜏, πœƒ) is a rectifiable 𝐸-current, we denote by 𝑇 𝐡 the current 𝑇(Ξ£ ∩ 𝐡, 𝜏, πœƒ).

Consider now a discrete subgroup 𝐺 < 𝐸, endowed with the restriction of the norm β€– β‹… ‖𝐸. If the multiplicity πœƒ takes only values in 𝐺, and if the same holds in the representation of πœ•π‘‡, we call 𝑇 a rectifiable 𝐺-current. Pay attention to the fact that, in the framework of currents over the coefficient group 𝐸, rectifiable 𝐸-currents play the role of (classical) rectifiable current, while rectifiable 𝐺-currents correspond to (classical) integral currents. Actually this correspondence is an equality, when 𝐸 is the group ℝ (with the Euclidean norm) and 𝐺 is β„€.

The next proposition gives a formula to compute the mass of a 1-dimensional rectifiable 𝐸-current. Proposition 1.9. Let 𝑇 = 𝑇(Ξ£, 𝜏, πœƒ) be a 1-dimensional rectifiable 𝐸-current. Then

𝕄(𝑇) = ∫ Ξ£

β€–πœƒ(π‘₯)‖𝐸dH1(π‘₯).

Since the mass is lower semicontinuous, we can apply the direct method of the Calculus of Variations for the existence of minimizers with given boundary, once we provide the following compactness result. Here we assume for simplicity that 𝐺 is the subgroup of 𝐸 generated by 𝑣1, . . . , π‘£π‘š(see Remark 1.2). A similar argument works for every discrete subgroup 𝐺.

Theorem 1.10. Let (π‘‡β„Ž)β„Žβ‰₯1be a sequence of rectifiable 𝐺-currents such that there exists a positive finite

con-stant 𝐢 satisfying

𝕄(π‘‡β„Ž) + 𝕄(πœ•π‘‡β„Ž) ≀ 𝐢 for every β„Ž β‰₯ 1.

Then there exists a subsequence (π‘‡β„Žπ‘–)𝑖β‰₯1and a rectifiable 𝐺-current 𝑇 such that

π‘‡β„Žπ‘–

βˆ— ⇀ 𝑇.

Proof. The statement of the theorem can be proved component by component. In fact, let 𝑇1

β„Ž, . . . , π‘‡β„Žπ‘šbe the components of π‘‡β„Ž. Since (𝑣1, . . . , π‘£π‘š), (𝑀1, . . . , π‘€π‘š) is a biorthonormal system, we have

𝕄(π‘‡β„Žπ‘—) + 𝕄(πœ•π‘‡β„Žπ‘—) ≀ π‘š(𝕄(π‘‡β„Ž) + 𝕄(πœ•π‘‡β„Ž)) ≀ π‘šπΆ, hence, after a diagonal procedure, we can find a subsequence (π‘‡β„Žπ‘–)𝑖β‰₯1such that (𝑇

𝑗

β„Žπ‘–)𝑖β‰₯1weakly

βˆ—converges to some integral current 𝑇𝑗, for every 𝑗 = 1, . . . , π‘š. Denoting by 𝑇 the rectifiable 𝐺-current, whose components are 𝑇1, . . . , π‘‡π‘š, we have

π‘‡β„Žπ‘–

βˆ— ⇀ 𝑇.

We conclude this section with some notations and basic facts about certain classes of rectifiable 𝐸-currents. Given a Lipschitz path 𝛾 : [0, 1] β†’ ℝ2(parameterized with constant speed), and a coefficient 𝑔 ∈ 𝐺, we define the associated 1-dimensional rectifiable 𝐺-current 𝑇 = 𝑇(Ξ“, 𝜏, 𝑔), where Ξ“ is the curve 𝛾([0, 1]) and, denoting

(6)

24 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

by β„“(Ξ“) the length of the curve Ξ“, the orientation 𝜏 is defined by 𝜏(𝛾(𝑑)) := 𝛾󸀠(𝑑)/β„“(Ξ“) for a.e. 𝑑 ∈ [0, 1]. It turns out that the boundary of such a current is πœ•π‘‡ = 𝑔𝛿𝛾(1)βˆ’ 𝑔𝛿𝛾(0), where the notation means that for every smooth πΈβˆ—-valued map πœ”, there holds

βŸ¨πœ•π‘‡; πœ”βŸ© = βŸ¨πœ”(𝛾(1)); π‘”βŸ© βˆ’ βŸ¨πœ”(𝛾(0)); π‘”βŸ©.

Using this notation, we observe that, given some points 𝑃1, . . . , π‘ƒπ‘˜ and some multiplicities 𝑔1, . . . , π‘”π‘˜in 𝐺, the 0-dimensional rectifiable 𝐺-current 𝑆 = 𝑔1𝛿𝑃1+ β‹… β‹… β‹… + π‘”π‘˜π›Ώπ‘ƒπ‘˜is the boundary of some 1-dimensional

rectifi-able 𝐺-current with compact support 𝑇 if and only if 𝑔1+ β‹… β‹… β‹… + π‘”π‘˜= 0.

2 Steiner tree problem revisited

In this section we establish the equivalence between the Steiner tree problem and a mass-minimization problem in a family of 𝐺-currents. We firstly need to choose the right group of coefficients 𝐺. Once we fix the number 𝑛 of points in the Steiner problem, we construct a normed vector space (𝐸, β€– β‹… ‖𝐸) and a subgroup 𝐺 of 𝐸, satisfying the following properties:

(P1) There exist 𝑔1, . . . , π‘”π‘›βˆ’1 ∈ 𝐺 and β„Ž1, . . . , β„Žπ‘›βˆ’1∈ πΈβˆ—such that (𝑔1, . . . , π‘”π‘›βˆ’1) with (β„Ž1, . . . , β„Žπ‘›βˆ’1) is a complete biorthonormal system for 𝐸 and 𝐺 is generated by 𝑔1, . . . , π‘”π‘›βˆ’1.

(P2) ‖𝑔𝑖1+ β‹… β‹… β‹… + π‘”π‘–π‘˜β€–πΈ= 1 whenever 1 ≀ 𝑖1< β‹… β‹… β‹… < π‘–π‘˜β‰€ 𝑛 βˆ’ 1 and π‘˜ ≀ 𝑛 βˆ’ 1.

(P3) ‖𝑔‖𝐸β‰₯ 1 for every 𝑔 ∈ 𝐺 \ {0}. (P4) Let πœƒ = βˆ‘π‘›βˆ’1

𝑗=1πœƒπ‘—π‘”π‘—and Μƒπœƒ = βˆ‘π‘›βˆ’1𝑗=1 Μƒπœƒπ‘—π‘”π‘—satisfy the following condition: { { { 0 ≀ Μƒπœƒπ‘—β‰€ πœƒπ‘— when πœƒπ‘— β‰₯ 0, 0 β‰₯ Μƒπœƒπ‘—β‰₯ πœƒπ‘— otherwise. Then β€–Μƒπœƒβ€–πΈβ‰€ β€–πœƒβ€–πΈ.

For the moment we will assume the existence of 𝐺 and 𝐸. The proof of their existence and an explicit repre-sentation, useful for the computations, are given in Lemma 2.6.

The next lemma has a fundamental role: through it, we can give a nice structure of 1-dimensional recti-fiable 𝐺-current to every suitable competitor for the Steiner tree problem. From now on we will denote

𝑔𝑛:= βˆ’(𝑔1+ β‹… β‹… β‹… + π‘”π‘›βˆ’1).

Lemma 2.1. Let 𝐡 be a compact and connected set with finite length in ℝ𝑑, containing the points 𝑝

1, . . . , 𝑝𝑛.

Then there exists a connected set π΅σΈ€ βŠ‚ 𝐡 containing 𝑝

1, . . . , 𝑝𝑛 and a 1-dimensional rectifiable 𝐺-current 𝑇𝐡󸀠 = 𝑇(𝐡󸀠, 𝜏, πœƒ) such that

(i) β€–πœƒ(π‘₯)‖𝐸= 1 for a.e. π‘₯ ∈ 𝐡󸀠,

(ii) πœ•π‘‡π΅σΈ€ is the 0-dimensional 𝐺-current 𝑔1𝛿𝑝

1+ β‹… β‹… β‹… + 𝑔𝑛𝛿𝑝𝑛.

Proof. Since 𝐡 is a connected, compact set of finite length, it follows that 𝐡 is connected by paths of finite length (see [9, Lemma 3.12]). Consider a curve 𝐡1which is the image of an injective path contained in 𝐡 going from 𝑝1to 𝑝𝑛and associate to it the rectifiable 𝐺-current 𝑇1with multiplicity βˆ’π‘”1, as explained in Section 1. Repeat this procedure keeping the end-point 𝑝𝑛and replacing at each step 𝑝1with 𝑝2, . . . , π‘π‘›βˆ’1. To be precise, in this procedure, as soon as a curve 𝐡𝑖intersects another curve 𝐡𝑗with 𝑗 < 𝑖, we force 𝐡𝑖to coincide with 𝐡𝑗 from that intersection point to the end-point 𝑝𝑛. The set 𝐡󸀠= 𝐡1βˆͺ β‹… β‹… β‹… βˆͺ π΅π‘›βˆ’1 βŠ‚ 𝐡 is a connected set containing the points 𝑝1, . . . , 𝑝𝑛and the 1-dimensional rectifiable 𝐺-current 𝑇 = 𝑇1+ β‹… β‹… β‹… + π‘‡π‘›βˆ’1satisfies the requirements of the lemma, in particular condition (i) is a consequence of (P2).

Via the next lemma (Lemma 2.3), we can say that solutions to the mass-minimization problem defined in Theorem 2.4 have connected supports. For the proof we need the following theorem on the structure of classical integral 1-currents. This theorem has firstly been stated as a corollary of [10, Theorem 4.2.25]. It allows us to consider an integral 1-current as a countable sum of oriented simple Lipschitz curves with integer multiplicities.

(7)

Theorem 2.2. Let 𝑇 be an integral 1-current in ℝ𝑑. Then 𝑇 =βˆ‘πΎ π‘˜=1π‘‡π‘˜+ ∞ βˆ‘ β„“=1𝐢ℓ (2.1) with

(i) π‘‡π‘˜are integral 1-currents associated to injective Lipschitz paths for every π‘˜ = 1, . . . , 𝐾, and 𝐢ℓare integral 1-currents associated to Lipschitz paths which have the same value at 0 and 1 and are injective on (0, 1) for every β„“ β‰₯ 1,

(ii) πœ•πΆβ„“= 0 for every β„“ β‰₯ 1.

Moreover 𝕄(𝑇) =βˆ‘πΎ π‘˜=1 𝕄(π‘‡π‘˜) +βˆ‘βˆž β„“=1 𝕄(𝐢ℓ) (2.2) and 𝕄(πœ•π‘‡) =βˆ‘πΎ π‘˜=1 𝕄(πœ•π‘‡π‘˜). (2.3)

Lemma 2.3. Let 𝑇 = 𝑇(Ξ£, 𝜏, πœƒ) be a 1-dimensional rectifiable 𝐺-current such that the boundary πœ•π‘‡ is the 0-current 𝑔1𝛿𝑝1+ β‹… β‹… β‹… + 𝑔𝑛𝛿𝑝𝑛. Then there exists a rectifiable 𝐺-current ̃𝑇 = 𝑇(ΜƒΞ£, Μƒπœ, Μƒπœƒ) such that

(i) πœ•Μƒπ‘‡ = πœ•π‘‡ = 𝑔1𝛿𝑝1+ β‹… β‹… β‹… + 𝑔𝑛𝛿𝑝𝑛,

(ii) supp(̃𝑇) is a connected 1-rectifiable set containing {𝑝1, . . . , 𝑝𝑛} and it is contained in supp(𝑇), (iii)H1(supp(̃𝑇) \ ΜƒΞ£) = 0,

(iv) 𝕄(̃𝑇) ≀ 𝕄(𝑇) and, if equality holds, then supp(𝑇) = supp(̃𝑇).

Proof. Let 𝑇𝑗 = 𝑇(Σ𝑗, πœπ‘—, πœƒπ‘—) be the components of 𝑇, for 𝑗 = 1, . . . , 𝑛 βˆ’ 1 (with respect to the biorthonormal system (𝑔1, . . . , π‘”π‘›βˆ’1), (β„Ž1, . . . , β„Žπ‘›βˆ’1)).

For every 𝑗, we can use Theorem 2.2 and write 𝑇𝑗 =βˆ‘πΎπ‘— π‘˜=1 π‘‡π‘˜π‘—+βˆ‘βˆž β„“=1 𝐢𝑗ℓ. Moreover, since πœ•π‘‡π‘—= 𝛿

π‘π‘—βˆ’ 𝛿𝑝𝑛, by equation (2.3), we have 𝐾𝑗= 1 for every 𝑗. We choose ̃𝑇 the rectifiable

𝐺-current whose components are ̃𝑇𝑗:= 𝑇𝑗

1. Because of (2.2), we have supp(̃𝑇𝑗) βŠ‚ supp(𝑇𝑗) (the cyclic part of 𝑇𝑗 never cancels the acyclic one).

Property (i) is easy to check. Property (iii) is also easy to check, because the corresponding property holds for every component ̃𝑇𝑗. To prove property (ii), it is sufficient to observe that ̃𝑇 is a finite sum of currents associated to oriented curves with multiplicities, having the point 𝑝𝑛in the support and that, by property (P1), 𝑔1, . . . , π‘”π‘›βˆ’1are linearly independent, hence the support of ̃𝑇 is the union of the supports of ̃𝑇𝑗. The inequality in property (iv) follows from (2.2) and from property (P4): indeed (2.2) implies that for every index β„“ such that the support of 𝐢𝑗

β„“intersects the support of 𝑇1𝑗in a set of positive length, thenH1-a.e. on this set the orientation of 𝐢𝑗

β„“coincide with the orientation of 𝑇1𝑗. Moreover, if 𝕄(̃𝑇) = 𝕄(𝑇), then (2.2) implies that every cycle 𝐢𝑗

β„“is supported in supp(̃𝑇), hence the second part of (iv) follows.

Before stating the main theorem, let us point out that the existence of a solution to the mass-minimization problem is a consequence of Theorem 1.10.

Theorem 2.4. Assume that 𝑇0= 𝑇(Ξ£0, 𝜏0, πœƒ0) is a mass-minimizer among all 1-dimensional rectifiable 𝐺-currents

with boundary

𝐡 = 𝑔1𝛿𝑝1+ β‹… β‹… β‹… + 𝑔𝑛𝛿𝑝𝑛.

Then 𝑆0:= supp(𝑇0) is a solution of the Steiner tree problem. Conversely, given a set 𝐢 which is a solution of

the Steiner problem for the points 𝑝1, . . . , 𝑝𝑛, there exists a canonical 1-dimensional 𝐺-current, supported on 𝐢,

minimizing the mass among the currents with boundary 𝐡.

Proof. Since 𝑇0 is a mass-minimizer, the mass of 𝑇0 must coincide with that of the current ̃𝑇0 given by Lemma 2.3. In particular, properties (ii) and (iv) of Lemma 2.3 guarantee that 𝑆0is a connected set.

(8)

26 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

Let 𝑆 be a competitor for the Steiner tree problem and let 𝑆󸀠and 𝑇

𝑆󸀠be the connected set and the rectifiable

1-current given by Lemma 2.1, respectively. Hence we have

H1(𝑆) β‰₯H1(𝑆󸀠)(1)= 𝕄(𝑇𝑆󸀠) (2) β‰₯ 𝕄(𝑇0) (3) β‰₯ H1(Ξ£0)(4)= H1(𝑆0), indeed

(1) thanks to the second property of Lemma 2.1 and Proposition 1.9, we obtain 𝕄(𝑇𝑆󸀠) = ∫

𝑆󸀠

β€–πœƒπ‘†σΈ€ (π‘₯)‖𝐸dH1(π‘₯) =H1(𝑆󸀠),

(2) we assumed that 𝑇0is a mass-minimizer, (3) from property (P3), we get

𝕄(𝑇0) = ∫ Ξ£0

β€–πœƒ0(π‘₯)‖𝐸dH1(π‘₯) β‰₯ ∫ Ξ£0

1 dH1(π‘₯) =H1(Ξ£0),

(4) is property (iii) in Lemma 2.3.

To prove the second part of the theorem, apply Lemma 2.1 to the set 𝐢. Notice that with the procedure described in the lemma, the rectifiable 𝐺-current 𝑇𝐢󸀠 is uniquely determined, because for every point 𝑝𝑖,

the set 𝐢 contains exactly one path from 𝑝𝑖to 𝑝𝑛, in fact it is well known that solutions of the Steiner tree problem cannot contain cycles; this explains the adjective β€œcanonical”. Assume by contradiction there exists a 1-dimensional 𝐺-current 𝑇 with πœ•π‘‡ = 𝐡 and 𝕄(𝑇) < 𝕄(𝑇𝐢󸀠). The 1-dimensional 𝐺-current ̃𝑇 obtained

applying Lemma 2.3 to 𝑇 has a connected 1-rectifiable support containing {𝑝1, . . . , 𝑝𝑛} and satisfies

H1(supp(̃𝑇)) ≀ 𝕄(̃𝑇) ≀ 𝕄(𝑇) < 𝕄(𝑇𝐢󸀠) =H1(supp(𝑇𝐢󸀠) ≀H1(𝐢),

which is a contradiction.

Remark 2.5. The proof given in the previous theorem shows in particular that the solutions of the mass-minimization problem do not depend on the choice of 𝐸 and 𝐺, but are universal for every 𝐺 and 𝐸 satisfying properties (P1)–(P4).

Eventually, we give an explicit representation for 𝐺 and 𝐸.

Lemma 2.6. For every 𝑛 ∈ β„• there exist a normed vector space (𝐸, β€– β‹… ‖𝐸) and a subgroup 𝐺 of 𝐸 satisfying

prop-erties (P1)–(P4).

Proof. Let {e1, . . . , e𝑛} be the standard basis of ℝ𝑛and {dπ‘₯1, . . . , dπ‘₯𝑛} be the dual basis. Consider 𝐸 := {𝑣 ∈ ℝ𝑛: 𝑣 β‹… e𝑛= 0}

and the homomorphism πœ™ : ℝ𝑛→ 𝐸 such that

πœ™(𝑒1, . . . , 𝑒𝑛) := (𝑒1βˆ’ 𝑒𝑛, . . . , π‘’π‘›βˆ’1βˆ’ 𝑒𝑛, 0). (2.4) Consider on ℝ𝑛the seminorm

‖𝑒‖⋆:= max𝑖=1,...,𝑛𝑒 β‹… eπ‘–βˆ’ min𝑖=1,...,𝑛𝑒 β‹… e𝑖.

Observe that β€– β‹… ‖⋆induces via πœ™ a norm on 𝐸 that we denote β€– β‹… ‖𝐸. For every 𝑖 = 1, . . . , 𝑛 βˆ’ 1, define 𝑔𝑖:= πœ™(e𝑖) and define 𝑔𝑛:= βˆ’(𝑔1+β‹… β‹… β‹…+π‘”π‘›βˆ’1). Let 𝐺 be the subgroup of 𝐸 generated by 𝑔1, . . . , π‘”π‘›βˆ’1. For every 𝑖 = 1, . . . , π‘›βˆ’1 denote by β„Žπ‘–the element dπ‘₯𝑖of πΈβˆ—. The pair (𝑔1, . . . , π‘”π‘›βˆ’1), (β„Ž1, . . . , β„Žπ‘›βˆ’1) is a biorthonormal system and prop-erties (P1)–(P4) are easy to check.

Remark 2.7. The norm β€– β‹… β€–πΈβˆ—of an element 𝑀 = 𝑀1β„Ž1+β‹… β‹… β‹…+π‘€π‘›βˆ’1β„Žπ‘›βˆ’1∈ πΈβˆ—can be characterized in the following

way: let us abbreviate

𝑀𝑃 :=π‘›βˆ’1βˆ‘

𝑖=1(π‘€π‘–βˆ¨ 0) and 𝑀

𝑁:= βˆ’π‘›βˆ’1βˆ‘ 𝑖=1(π‘€π‘–βˆ§ 0)

and, for every 𝑣 = (𝑣1, . . . , π‘£π‘›βˆ’1, 0) ∈ 𝐸 with ‖𝑣‖𝐸= 1, πœ†(𝑣) := max𝑖=1,...,π‘›βˆ’1(π‘£π‘–βˆ¨ 0) ∈ [0, 1]. Then β€–π‘€β€–πΈβˆ—= sup ‖𝑣‖𝐸=1 π‘›βˆ’1 βˆ‘ 𝑖=1𝑀𝑖𝑣𝑖= sup‖𝑣‖𝐸=1 [πœ†(𝑣)𝑀𝑃+ (1 βˆ’ πœ†(𝑣))𝑀𝑁] = sup πœ†βˆˆ[0,1][(πœ†π‘€ 𝑃+ (1 βˆ’ πœ†)𝑀𝑁] = π‘€π‘ƒβˆ¨ 𝑀𝑁. (2.5)

(9)

Moreover we can also notice that, according to this representation of 𝐸 and 𝐺, the only extreme points of the unit ball in 𝐸 are all the points of 𝐺 of unit norm, i.e. all the points 𝑔 of the type 𝑔 = Β±(𝑔𝑖1+ β‹… β‹… β‹… + π‘”π‘–π‘˜) such

that 1 ≀ 𝑖1< β‹… β‹… β‹… < π‘–π‘˜β‰€ 𝑛 βˆ’ 1 and π‘˜ ≀ 𝑛 βˆ’ 1.

3 Calibrations

As we recalled in the Introduction, our interest in calibrations is the reason why we have chosen to provide an integral representation for 𝐸-currents, indeed the existence of a calibration guarantees the minimality of the associated current, as we will see in Proposition 3.2.

Definition 3.1. A smooth calibration associated with a π‘˜-dimensional rectifiable 𝐺-current 𝑇(Ξ£, 𝜏, πœƒ) in ℝ𝑑is a smooth compactly supported πΈβˆ—-valued differential π‘˜-form πœ” with the following properties:

(i) βŸ¨πœ”(π‘₯); 𝜏(π‘₯), πœƒ(π‘₯)⟩ = β€–πœƒ(π‘₯)‖𝐸forHπ‘˜-a.e. π‘₯ ∈ Ξ£, (ii) dπœ” = 0,

(iii) β€–πœ”β€– ≀ 1, where β€–πœ”β€– is the comass of πœ”, defined in (1.2).

Proposition 3.2. A rectifiable 𝐺-current 𝑇 which admits a smooth calibration πœ” is a minimizer for the mass among the normal 𝐸-currents with boundary πœ•π‘‡.

Proof. Fix a competitor 𝑇󸀠which is a normal 𝐸-current associated with the vectorfield πœσΈ€ , the multiplicity πœƒσΈ€  and the measure πœ‡π‘‡σΈ€ (according to Definition 1.7), with πœ•π‘‡σΈ€ = πœ•π‘‡. Since πœ•(𝑇 βˆ’ 𝑇󸀠) = 0, it follows that 𝑇 βˆ’ 𝑇󸀠is

a boundary of some 𝐸-current 𝑆 in ℝ𝑑, and then 𝕄(𝑇) = ∫ Ξ£ β€–πœƒβ€–πΈdHπ‘˜ (3.1) (i) = ∫ Ξ£ βŸ¨πœ”(π‘₯); 𝜏(π‘₯), πœƒ(π‘₯)⟩ dHπ‘˜= βŸ¨π‘‡; πœ”βŸ© (3.2) (ii)= βŸ¨π‘‡σΈ€ ; πœ”βŸ© = ∫ ℝ𝑑 βŸ¨πœ”(π‘₯); πœσΈ€ (π‘₯), πœƒσΈ€ (π‘₯)⟩ dπœ‡ 𝑇󸀠 (3.3) (iii) ≀ ∫ ℝ𝑑 β€–πœƒσΈ€ β€–πΈdπœ‡π‘‡σΈ€  = 𝕄(𝑇󸀠), (3.4)

where each equality (respectively inequality) holds because of the corresponding property of πœ”, as established in Definition 3.1. In particular, equality in (ii) follows from

βŸ¨π‘‡ βˆ’ 𝑇󸀠; πœ”βŸ© = βŸ¨πœ•π‘†; πœ”βŸ© = βŸ¨π‘†; dπœ”βŸ© = 0.

Remark 3.3. If 𝑇 is a rectifiable 𝐺-current calibrated by πœ”, then every mass-minimizer with boundary πœ•π‘‡ is calibrated by the same form πœ”. In fact, choose a mass-minimizer 𝑇󸀠= 𝑇(Ξ£σΈ€ , πœσΈ€ , πœƒσΈ€ ) with boundary πœ•π‘‡σΈ€ = πœ•π‘‡: obviously we have 𝕄(𝑇) = 𝕄(𝑇󸀠), then equality holds in (3.4), which means

βŸ¨πœ”(π‘₯); πœσΈ€ (π‘₯), πœƒσΈ€ (π‘₯)⟩ = β€–πœƒσΈ€ (π‘₯)‖𝐸 forHπ‘˜-a.e. π‘₯ ∈ Ξ£σΈ€ .

At this point we need a short digression on the representation of a πΈβˆ—-valued 1-form πœ”; we will consider the case 𝑑 = 2, all our examples being for the Steiner tree problem in ℝ2. Remember that in Section 2 we fixed a basis (β„Ž1, . . . , β„Žπ‘›βˆ’1) for πΈβˆ—, dual to the basis (𝑔1, . . . , π‘”π‘›βˆ’1) for 𝐸. We represent

πœ” = (

πœ”1,1dπ‘₯1+ πœ”1,2dπ‘₯2 ...

πœ”π‘›βˆ’1,1dπ‘₯1+ πœ”π‘›βˆ’1,2dπ‘₯2 ) , so that, if 𝜏 = 𝜏1e1+ 𝜏2e2∈ Ξ›1(ℝ2) and 𝑣 = 𝑣1𝑔1+ β‹… β‹… β‹… + π‘£π‘›βˆ’1π‘”π‘›βˆ’1∈ 𝐸, then

βŸ¨πœ”; 𝜏, π‘£βŸ© =π‘›βˆ’1βˆ‘

(10)

28 | A. Marchese and A. Massaccesi, The Steiner tree problem revisitedA. Marchese and A. Massaccesi, The Steiner tree problem revisited | 11

𝑝

1

𝑝

2

𝑝

3

𝑝

0

𝑔

3

𝑔

1

𝑔

2

Figure 3. Solution for the problem with boundary on the vertices of an equilateral triangle.

In Definition 3.1 we intentionally kept vague the regularity of the form πœ”. Indeed πœ” has to be a compactly supportedΒ² smooth form, a priori, in order to fit Definition 1.5. Nevertheless, in some situations it will be useful to consider calibrations with lower regularity, for instance piecewise constant forms. As long as (3.1)–(3.4) remain valid, it is meaningful to do so; for this reason we introduce the following very general definition.

Definition 3.5. A generalized calibration associated with a π‘˜-dimensional normal 𝐸-current 𝑇 is a linear and

bounded functional πœ™ on the space of normal 𝐸-currents satisfying the following conditions: (i) πœ™(𝑇) = 𝕄(𝑇),

(ii) πœ™(πœ•π‘…) = 0 for any (π‘˜ + 1)-dimensional normal 𝐸-current 𝑅, (iii) β€–πœ™β€– ≀ 1.

Remark 3.6. Proposition 3.2 still holds, since for every competitor 𝑇󸀠with πœ•π‘‡ = πœ•π‘‡σΈ€ , there holds

𝕄(𝑇) = πœ™(𝑇) = πœ™(𝑇󸀠) + πœ™(πœ•π‘…) ≀ 𝕄(𝑇󸀠),

where 𝑅 is chosen such that 𝑇 βˆ’ 𝑇󸀠= πœ•π‘…. Suchan𝑅 exists because 𝑇 and 𝑇󸀠are in the same homology class.

As examples, we present the calibrations for two well-known Steiner tree problems in ℝ2. Both β€œcalibrations”

in Example 3.10 and in Example 3.11 are piecewise constant 1-forms (with values in normed vector spaces of dimension 3 and 6, respectively). So firstly we need to show that certain piecewise constant forms provide generalized calibrations in the sense of Definition 3.5.

Definition 3.7. Fix a 1-dimensional rectifiable 𝐺-current 𝑇 in ℝ2, 𝑇 = 𝑇(Ξ£, 𝜏, πœƒ). Assume we have a

collec-tion {πΆπ‘Ÿ}π‘Ÿβ‰₯1which is a locally finite, Lipschitz partition of ℝ2, where the sets πΆπ‘Ÿhave non-empty connected

interior, the boundary of every set πΆπ‘Ÿis a Lipschitz curve (of finite length, unless πΆπ‘Ÿis unbounded) and

πΆπ‘Ÿβˆ© 𝐢𝑠= 0 whenever π‘Ÿ ΜΈ= 𝑠. Assume moreover that 𝐢1is a closed set and for every π‘Ÿ > 1,

πΆπ‘ŸβŠƒ (πΆπ‘Ÿ\ ⋃ 𝑖<π‘ŸπΆπ‘–).

Let us consider a compactly supported piecewise constant πΈβˆ—-valued 1-form πœ” with

πœ” ≑ πœ”π‘Ÿ on πΆπ‘Ÿ,

where πœ”π‘Ÿβˆˆ Ξ›1𝐸(ℝ2) for every π‘Ÿ. In particular πœ” ΜΈ= 0 only on finitely many elements of the partition. Then we

say that πœ” represents a compatible calibration for 𝑇 if the following conditions hold:

2 Since we deal with currents that are compactly supported, we can easily drop the assumption that πœ” has compact support.

Figure 3. Solution for the problem with boundary on the vertices of an

equilateral triangle.

Example 3.4. Consider the vector space 𝐸 and the group 𝐺 defined in Lemma 2.6 with 𝑛 = 3; let 𝑝0= (0, 0), 𝑝1= (1/2, √3/2), 𝑝2= (1/2, βˆ’βˆš3/2), 𝑝3= (βˆ’1, 0)

(see Figure 3). Consider the rectifiable 𝐺-current 𝑇 supported in the cone over (𝑝1, 𝑝2, 𝑝3), with respect to 𝑝0, with piecewise constant weights 𝑔1, 𝑔2, 𝑔3:= βˆ’(𝑔1+ 𝑔2) on 𝑝0𝑝1, 𝑝0𝑝2, 𝑝0𝑝3respectively (see Figure 3 for the orientation). This current 𝑇 is a minimizer for the mass. In fact, a constant 𝐺-calibration πœ” associated with 𝑇 is

πœ” := (12dπ‘₯1+ √32 dπ‘₯2 1

2dπ‘₯1βˆ’ √32 dπ‘₯2 ) .

Condition (i) is easy to check and condition (ii) is trivially verified because πœ” is constant. To check condi-tion (iii) we note that, for the vector 𝜏 = cos 𝛼 e1+ sin 𝛼 e2, we have

βŸ¨πœ”; 𝜏, β‹… ⟩ = (12cos 𝛼 +√32 sin 𝛼 1

2cos 𝛼 βˆ’βˆš32 sin 𝛼 ) .

In order to compute the comass norm of πœ”, we could use the characterization of the norm β€– β‹… β€–πΈβˆ—given in

Remark 2.7, but for 𝑛 = 3 computations are simpler. Since the unit ball of 𝐸 is convex, and its extreme points are the unit points of 𝐺, it is sufficient to evaluate βŸ¨πœ”; 𝜏, β‹… ⟩ on ±𝑔1, ±𝑔2, Β±(𝑔1+ 𝑔2). We have

|βŸ¨πœ”; 𝜏, 𝑔1⟩| = |βŸ¨πœ”; 𝜏, βˆ’π‘”1⟩| =󡄨󡄨󡄨󡄨󡄨󡄨󡄨sin(𝛼 +πœ‹ 6)󡄨󡄨󡄨󡄨󡄨󡄨󡄨 ≀ 1, |βŸ¨πœ”; 𝜏, 𝑔2⟩| = |βŸ¨πœ”; 𝜏, βˆ’π‘”2⟩| =󡄨󡄨󡄨󡄨󡄨󡄨󡄨sin(𝛼 +5

6πœ‹)󡄨󡄨󡄨󡄨󡄨󡄨󡄨 ≀ 1, |βŸ¨πœ”; 𝜏, 𝑔1+ 𝑔2⟩| = |βŸ¨πœ”; 𝜏, βˆ’(𝑔1+ 𝑔2)⟩| = |cos 𝛼| ≀ 1.

In Definition 3.1 we intentionally kept vague the regularity of the form πœ”. Indeed πœ” has to be a compactly supportedΒ² smooth form, a priori, in order to fit Definition 1.5. Nevertheless, in some situations it will be useful to consider calibrations with lower regularity, for instance piecewise constant forms. As long as (3.1)–(3.4) remain valid, it is meaningful to do so; for this reason we introduce the following very general definition. Definition 3.5. A generalized calibration associated with a π‘˜-dimensional normal 𝐸-current 𝑇 is a linear and bounded functional πœ™ on the space of normal 𝐸-currents satisfying the following conditions:

(i) πœ™(𝑇) = 𝕄(𝑇),

(ii) πœ™(πœ•π‘…) = 0 for any (π‘˜ + 1)-dimensional normal 𝐸-current 𝑅, (iii) β€–πœ™β€– ≀ 1.

(11)

Remark 3.6. Proposition 3.2 still holds, since for every competitor 𝑇󸀠with πœ•π‘‡ = πœ•π‘‡σΈ€ , there holds 𝕄(𝑇) = πœ™(𝑇) = πœ™(𝑇󸀠) + πœ™(πœ•π‘…) ≀ 𝕄(𝑇󸀠),

where 𝑅 is chosen such that 𝑇 βˆ’ 𝑇󸀠= πœ•π‘…. Such an 𝑅 exists because 𝑇 and 𝑇󸀠are in the same homology class. As examples, we present the calibrations for two well-known Steiner tree problems in ℝ2. Both β€œcalibrations” in Example 3.10 and in Example 3.11 are piecewise constant 1-forms (with values in normed vector spaces of dimension 3 and 6, respectively). So firstly we need to show that certain piecewise constant forms provide generalized calibrations in the sense of Definition 3.5.

Definition 3.7. Fix a 1-dimensional rectifiable 𝐺-current 𝑇 in ℝ2, 𝑇 = 𝑇(Ξ£, 𝜏, πœƒ). Assume we have a collec-tion {πΆπ‘Ÿ}π‘Ÿβ‰₯1which is a locally finite, Lipschitz partition of ℝ2, where the sets πΆπ‘Ÿhave non-empty connected interior, the boundary of every set πΆπ‘Ÿ is a Lipschitz curve (of finite length, unless πΆπ‘Ÿ is unbounded) and πΆπ‘Ÿβˆ© 𝐢𝑠= 0 whenever π‘Ÿ ΜΈ= 𝑠. Assume moreover that 𝐢1is a closed set and for every π‘Ÿ > 1,

πΆπ‘ŸβŠƒ (πΆπ‘Ÿ\ ⋃ 𝑖<π‘ŸπΆπ‘–).

Let us consider a compactly supported piecewise constant πΈβˆ—-valued 1-form πœ” with πœ” ≑ πœ”

π‘Ÿ on πΆπ‘Ÿ, where πœ”π‘Ÿ ∈ Ξ›1𝐸(ℝ2) for every π‘Ÿ. In particular πœ” ΜΈ= 0 only on finitely many elements of the partition. Then we say that πœ” represents a compatible calibration for 𝑇 if the following conditions hold:

(i) forH1-almost every point π‘₯ ∈ Ξ£, βŸ¨πœ”(π‘₯); 𝜏(π‘₯), πœƒ(π‘₯)⟩ = β€–πœƒ(π‘₯)‖𝐸,

(ii) forH1-almost every point π‘₯ ∈ πœ•πΆπ‘Ÿβˆ© πœ•πΆπ‘ we have

βŸ¨πœ”π‘Ÿβˆ’ πœ”π‘ ; 𝜏(π‘₯), β‹… ⟩ = 0, where 𝜏 is tangent to πœ•πΆπ‘Ÿ,

(iii) β€–πœ”π‘Ÿβ€– ≀ 1 for every π‘Ÿ.

We will refer to condition (ii) with the expression of compatibility condition for a piecewise constant form. Proposition 3.8. Let πœ” be a compatible calibration for the rectifiable 𝐺-current 𝑇. Then 𝑇 minimizes the mass among the normal 𝐸-currents with boundary πœ•π‘‡.

To prove this proposition we need the following result of decomposition of classical normal 1-currents, see [24] for the classical result and [21] for its generalization to metric spaces. Given a compact measure space (𝑋, πœ‡) and a family of π‘˜-currents {𝑇π‘₯}π‘₯βˆˆπ‘‹in ℝ𝑑such that

∫ 𝑋 𝕄(𝑇π‘₯) dπœ‡(π‘₯) < +∞, we denote by 𝑇 := ∫ 𝑋 𝑇π‘₯dπœ‡(π‘₯) the π‘˜-current 𝑇 satisfying

βŸ¨π‘‡, πœ”βŸ© = ∫ 𝑋

βŸ¨π‘‡π‘₯, πœ”βŸ© dπœ‡(π‘₯) for every smooth compactly supported π‘˜-form πœ”.

Proposition 3.9. Every normal 1-current 𝑇 in ℝ𝑑can be written as 𝑇 =

𝑀 ∫ 0

𝑇𝑑d𝑑,

where 𝑇𝑑is an integral current with 𝕄(𝑇𝑑) ≀ 2 and 𝕄(πœ•π‘‡π‘‘) ≀ 2 for every 𝑑, and 𝑀 is a positive number depending

only on 𝕄(𝑇) and 𝕄(πœ•π‘‡). Moreover

𝕄(𝑇) = 𝑀 ∫ 0

(12)

30 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

Proof of Proposition 3.8. Firstly we see that a suitable counterpart of Stokes’ theorem holds. Namely, given a component πœ”π‘—of πœ” and a classical integral 1-current 𝑇 = 𝑇(Ξ£, 𝜏, 1) in ℝ2, without boundary, then we claim that

βŸ¨πœ”π‘—; π‘‡βŸ© := ∫ Ξ£

βŸ¨πœ”π‘—(π‘₯); 𝜏(π‘₯)⟩ dH1(π‘₯) = 0. (3.5)

To prove the claim, note that it is possible to find at most countably many unit multiplicity integral 1-currents 𝑇𝑖= 𝑇(Σ𝑖, πœπ‘–, 1) in ℝ2, without boundary, each one supported in a single set 𝐢

π‘Ÿ, such that βˆ‘π‘–π‘‡π‘–= 𝑇. Since πœ”π‘—β‰‘ πœ”π‘—

π‘Ÿon πΆπ‘Ÿand since (ii) holds, we obtain ∫ Σ𝑖 βŸ¨πœ”π‘—(π‘₯); 𝜏 𝑖(π‘₯)⟩ dH1(π‘₯) = ∫ Σ𝑖 βŸ¨πœ”π‘— π‘Ÿ(π‘₯); πœπ‘–(π‘₯)⟩ dH1(π‘₯) = 0 for every 𝑖; then the claim follows.

As a consequence of (3.5) we can find a family of β€œpotentials”, i.e. Lipschitz functions πœ™π‘—: ℝ2β†’ ℝ such that for every (classical) integral 1-current 𝑆 associated to a Lipschitz path 𝛾 with 𝛾(1) = π‘₯𝑆and 𝛾(0) = 𝑦𝑆, there holds

βŸ¨πœ”π‘—; π‘†βŸ© = πœ™π‘—(π‘₯𝑆) βˆ’ πœ™π‘—(𝑦𝑆) for every 𝑗.

Indeed, by (3.5) the above integral does not depend on the path 𝛾 but only on the points π‘₯𝑆and 𝑦𝑆. Therefore, in order to construct such potentials, it is sufficient to choose πœ™π‘—(0) = 0 and

πœ™π‘—(π‘₯) = |π‘₯| 1 ∫ 0

βŸ¨πœ”π‘—(𝑑π‘₯); π‘₯|π‘₯|⟩ d𝑑.

Moreover it is easy to see that every πœ™π‘—is constant outside of the support of πœ”π‘—, so we can assume, possibly subtracting a constant, that πœ™π‘—is compactly supported.

Now, consider any 2-dimensional normal 𝐸-current 𝑇. Let {𝑇𝑗}

𝑗be the components of 𝑇. For every 𝑗, use Proposition 3.9 to write 𝑆𝑗:= πœ•π‘‡π‘—= 𝑀𝑗 ∫ 0 𝑆𝑗𝑑d𝑑. Then we have βŸ¨πœ”; πœ•π‘‡βŸ© = βˆ‘ 𝑗 𝑀𝑗 ∫ 0 βŸ¨πœ”π‘—; π‘†π‘—π‘‘βŸ© d𝑑 = βˆ‘ 𝑗 𝑀𝑗 ∫ 0 πœ™π‘—(π‘₯𝑆𝑗𝑑) βˆ’ πœ™π‘—(𝑦𝑆𝑗𝑑) d𝑑.

Since for every 𝑗 we have

0 = πœ•(πœ•π‘‡π‘—) = 𝑀𝑗 ∫ 0 𝛿π‘₯ π‘†π‘—π‘‘βˆ’ 𝛿𝑦𝑆𝑗𝑑d𝑑, it follows that 𝑀𝑗 ∫ 0 𝑔(π‘₯𝑆𝑗 𝑑) βˆ’ 𝑔(𝑦𝑆𝑗𝑑) d𝑑 = 0

for every 𝑗 and for every compactly supported Lipschitz function 𝑔, in particular for 𝑔 = πœ™π‘—. Hence we have βŸ¨πœ”; πœ•π‘‡βŸ© = 0.

Example 3.10. Consider the points 𝑝1= (1, 1), 𝑝2= (1, βˆ’1), 𝑝3= (βˆ’1, βˆ’1), 𝑝4= (βˆ’1, 1) ∈ ℝ2. The corresponding solution of the Steiner tree problemΒ³ are those represented in Figure 1. We associate with each point 𝑝𝑗 with 𝑗 = 1, . . . , 4 the coefficients π‘”π‘—βˆˆ 𝐺, where 𝐺 is the group defined in Lemma 2.6 with 𝑛 = 4: let us call

𝐡 := 𝑔1𝛿𝑝1+ 𝑔2𝛿𝑝2+ 𝑔3𝛿𝑝3+ 𝑔4𝛿𝑝4.

3 In dimension 𝑑 > 2, an interesting question related to this problem is the following: is the cone over the (𝑑 βˆ’ 2)-skeleton of

the hypercube in ℝ𝑑area minimizing, among hypersurfaces separating the faces? The question has a positive answer if and only

(13)

This 0-dimensional current is our boundary. Intuitively our mass-minimizing candidates among 1-dimen-sional rectifiable 𝐺-currents are those represented in Figure 4: these currents 𝑇hor, 𝑇verare supported in the sets drawn, respectively, with continuous and dashed lines in Figure 4 and have piecewise constant coeffi-cients intended to satisfy the boundary condition πœ•π‘‡14 | A. Marchese and A. Massaccesi, The Steiner tree problem revisitedhor= 𝐡 = πœ•π‘‡ver.

𝑔

1

𝑔

2

𝑔

3

𝑔

4

𝑔

1

𝑔

4

𝑔

2

𝑔

1

+ 𝑔

2

𝑔

1

+ 𝑔

4

𝑔

1

𝑔

4

𝑔

3

𝑔

3

𝑔

2

𝑇

ver

πœ”

1

πœ”

3

πœ”

2

πœ”

4

𝑇

hor

part.

Figure 4. Solution for the mass-minimization problem.

It is easy to check that πœ” satisfies both condition (i) and the compatibility condition of Definition 3.7. To check that condition (iii) is satisfied, we can use formula (2.5).

Example 3.11. Consider the vertices of a regular hexagon plus the center, namely

𝑝1= (1/2, √3/2), 𝑝2= (1, 0), 𝑝3= (1/2, βˆ’βˆš3/2),

𝑝4= (βˆ’1/2, βˆ’βˆš3/2), 𝑝5= (βˆ’1, 0), 𝑝6= (βˆ’1/2, √3/2), 𝑝7= (0, 0),

and associate with each point 𝑝𝑗the corresponding multiplicity π‘”π‘—βˆˆ 𝐺, where 𝐺 is the group defined in

Lemma 2.6 with 𝑛 = 4. A mass-minimizer for the problem with boundary 𝐡 =βˆ‘7

𝑗=1𝑔𝑗𝛿𝑝𝑗

is illustrated in Figure 5, the other one can be obtained with a πœ‹/3-rotation of the picture.

Figure 5. Solution for the mass-minimization problem.

Let us divide ℝ2intosix cones of angle πœ‹/3, as in Figure 5; we will label each cone with a number

from 1 to 6, starting from that containing (0, 1) and moving clockwise. A compatible calibration for the two

minimizers is the following: Note 4:

Display: Shall I replace = by :=?

Figure 4. Solution for the mass-minimization problem.

In this case, a compatible calibration for both 𝑇horand 𝑇veris defined piecewise as follows (the notation is the same as in Example 3.4 and the partition is delimited by the dotted lines):

πœ”1≑ ( √3 2 dπ‘₯1+ 12dπ‘₯2 (1 βˆ’βˆš32 )dπ‘₯1βˆ’12dπ‘₯2 (βˆ’1 +√32 )dπ‘₯1βˆ’1 2dπ‘₯2 ) , πœ”2≑ ( 1 2dπ‘₯1+ √32dπ‘₯2 1 2dπ‘₯1βˆ’ √32dπ‘₯2 βˆ’1 2dπ‘₯1βˆ’ (1 βˆ’βˆš32 )dπ‘₯2 ) , πœ”3≑ ( (1 βˆ’βˆš32)dπ‘₯1+ 12dπ‘₯2 √3 2dπ‘₯1βˆ’12dπ‘₯2 βˆ’βˆš32 dπ‘₯1βˆ’ 12dπ‘₯2 ) , πœ”4≑ ( 1 2dπ‘₯1+ (1 βˆ’βˆš32)dπ‘₯2 1 2dπ‘₯1βˆ’ (1 βˆ’βˆš32)dπ‘₯2 βˆ’1 2dπ‘₯1βˆ’ √32dπ‘₯2 ) .

It is easy to check that πœ” satisfies both condition (i) and the compatibility condition of Definition 3.7. To check that condition (iii) is satisfied, we can use formula (2.5).

Example 3.11. Consider the vertices of a regular hexagon plus the center, namely 𝑝1= (1/2, √3/2), 𝑝2= (1, 0),

𝑝3= (1/2, βˆ’βˆš3/2), 𝑝4= (βˆ’1/2, βˆ’βˆš3/2), 𝑝5= (βˆ’1, 0), 𝑝6= (βˆ’1/2, √3/2), 𝑝7= (0, 0),

and associate with each point 𝑝𝑗 the corresponding multiplicity π‘”π‘—βˆˆ 𝐺, where 𝐺 is the group defined in Lemma 2.6 with 𝑛 = 4. A mass-minimizer for the problem with boundary

𝐡 =βˆ‘7 𝑗=1𝑔𝑗𝛿𝑝𝑗

(14)

32 | A. Marchese and A. Massaccesi, The Steiner tree problem revisitedA. Marchese and A. Massaccesi, The Steiner tree problem revisited | 15

𝑔

1

𝑔

6

𝑔

2

𝑔

3

𝑔

4

𝑔

5

𝑔

7

Figure 5. Solution for the mass-minimization problem.

Again, it is not difficult to check that πœ” satisfies both condition (i) and the compatibility condition of Definition 3.7. To check that condition (iii) is satisfied, we use formula (2.5).

Remark 3.12. We may wonder whether or not the calibration given in Example 3.11 can be adjusted so to work

for the set of the vertices of the hexagon (without the seventh point in the center): the answer is negative, in fact the support of the current in Figure 5 is not a solution for the Steiner tree problem on the six points, the perimeter of the hexagon minus one side being the shortest graph, as proved in [15].

Remark 3.13. In both Examples 3.10 and 3.11, once we fixed the partition and we decided to look for a

piece-wise constant calibration for our candidates, the construction of πœ” was forced by both conditions (i) of Definition 3.1 and the compatibility condition of Definition 3.7. Notice that the calibration for Example 3.11 has evident analogies with the one exhibited in Example 3.4. Actually we obtained the first one simply pasting suitably β€œrotated” copies of the second one.

In the following remarks we intend to underline the analogies and the connections with calibrations in similar contexts. See [20, Chapter 6] for an overview on the subject of calibrations.

Remark 3.14 (Functionals defined on partitions and null lagrangians). There is an interesting and deep

anal-ogy between calibrations and null lagrangians, analanal-ogy that still holds in the group-valued coefficients framework. Consider some points {πœ‚1, . . . , πœ‚π‘›} βŠ‚ β„π‘šwith

|πœ‚π‘–βˆ’ πœ‚π‘—| = 1 for all 𝑖 ΜΈ= 𝑗, (3.7)

and fix an open set with Lipschitz boundary Ξ© βŠ‚ ℝ𝑑. It is natural to study the variational problem

inf{∫

Ξ©

|𝐷𝑒| : 𝑒 ∈ BV(Ξ©; {πœ‚1, . . . , πœ‚π‘›}), 𝑒|πœ•Ξ©β‰‘ 𝑒0}. (3.8)

It turns out that ∫Ω|𝐷𝑒| is the same energy we want to minimize in the Steiner tree problem, ∫Ω|𝐷𝑒| being the length of the jump set of 𝑒.

This problem concerns the theory of partitions of an open set Ξ© in a finite number of sets of finite perimeter. This theory was developed by Ambrosio and Braides in [1, 2], which we refer to for a complete exposition.

Figure 5. Solution for the mass-minimization problem. Let us divide ℝ2 into six cones of angle πœ‹/3, as in Figure 5; we will label each cone with a number from 1 to 6, starting from that containing (0, 1) and moving clockwise. A compatible calibration for the two minimizers is the following:

πœ”1≑ ( ( ( ( ( ( ( βˆ’βˆš32 dπ‘₯1+ 1 2dπ‘₯2 √3 2dπ‘₯1+12dπ‘₯2 0 0 0 0 ) ) ) ) ) ) ) , πœ”2≑ ( ( ( ( ( ( ( 0 dπ‘₯2 √3 2 dπ‘₯1βˆ’ 12dπ‘₯2 0 0 0 ) ) ) ) ) ) ) , πœ”3≑ ( ( ( ( ( ( ( 0 0 √3 2dπ‘₯1+12dπ‘₯2 βˆ’dπ‘₯2 0 0 ) ) ) ) ) ) ) , πœ”4≑ ( ( ( ( ( ( ( 0 0 0 √3 2dπ‘₯1βˆ’12dπ‘₯2 βˆ’βˆš32 dπ‘₯1βˆ’ 12dπ‘₯2 0 ) ) ) ) ) ) ) , πœ”5≑ ( ( ( ( ( ( ( 0 0 0 0 βˆ’dπ‘₯2 βˆ’βˆš32dπ‘₯1+1 2dπ‘₯2 ) ) ) ) ) ) ) , πœ”6≑ ( ( ( ( ( ( ( dπ‘₯2 0 0 0 0 βˆ’βˆš32 dπ‘₯1βˆ’ 1 2dπ‘₯2 ) ) ) ) ) ) ) . (3.6)

Again, it is not difficult to check that πœ” satisfies both condition (i) and the compatibility condition of Definition 3.7. To check that condition (iii) is satisfied, we use formula (2.5).

Remark 3.12. We may wonder whether or not the calibration given in Example 3.11 can be adjusted so to work for the set of the vertices of the hexagon (without the seventh point in the center): the answer is negative, in fact the support of the current in Figure 5 is not a solution for the Steiner tree problem on the six points, the perimeter of the hexagon minus one side being the shortest graph, as proved in [15].

Remark 3.13. In both Examples 3.10 and 3.11, once we fixed the partition and we decided to look for a piece-wise constant calibration for our candidates, the construction of πœ” was forced by both conditions (i) of Definition 3.1 and the compatibility condition of Definition 3.7. Notice that the calibration for Example 3.11 has evident analogies with the one exhibited in Example 3.4. Actually we obtained the first one simply pasting suitably β€œrotated” copies of the second one.

In the following remarks we intend to underline the analogies and the connections with calibrations in similar contexts. See [20, Chapter 6] for an overview on the subject of calibrations.

Remark 3.14 (Functionals defined on partitions and null lagrangians). There is an interesting and deep anal-ogy between calibrations and null lagrangians, analanal-ogy that still holds in the group-valued coefficients

(15)

framework. Consider some points {πœ‚1, . . . , πœ‚π‘›} βŠ‚ β„π‘šwith

|πœ‚π‘–βˆ’ πœ‚π‘—| = 1 for all 𝑖 ΜΈ= 𝑗, (3.7)

and fix an open set with Lipschitz boundary Ξ© βŠ‚ ℝ𝑑. It is natural to study the variational problem inf{∫

Ξ©

|𝐷𝑒| : 𝑒 ∈ BV(Ξ©; {πœ‚1, . . . , πœ‚π‘›}), 𝑒|πœ•Ξ© ≑ 𝑒0}. (3.8) It turns out that ∫Ω|𝐷𝑒| is the same energy we want to minimize in the Steiner tree problem, ∫Ω|𝐷𝑒| being the length of the jump set of 𝑒.16 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

πœ‚

4

πœ‚

2

πœ‚

1

πœ‚

3

Ξ©

Figure 6. Boundary data.

Remark 3.15 (Clusters with multiplicities). In [19], F. Morgan applies flat chains with coefficients in a group 𝐺

to soap bubble clusters and immiscible fluids, following the idea of B. White in [25]. For a detailed comparison of [19] with our technique, see [18,Section3.2.3]. Here we just notice that the definition of calibration in [19] works well in the case of free abelian groups and this is the main difference with our approach.

Remark 3.16 (Paired calibrations). It is worth mentioning another analogy between the technique of

cal-ibrations (for currents with coefficients in a group) illustrated in this paper and the technique of paired calibrations in [17]. In particular, in the specific example of the truncated cone over the 1-skeleton of the tetrahedron in ℝ3(the surface with least area among those separating the faces of the tetrahedron), one can

detect a correspondence even at the level of the main computations. See [18,Section3.2.3] for the details. Following an idea of Federer (see [11]), in [17, 19] (and in [5, 6], as well) one can observe the exploitation of the duality between minimal surfaces and maximal flows through the same boundary. We will examine this duality in Section 4, but we conclude the present section with a remark closely related to this idea.

Remark 3.17 (Covering spaces and calibrations for soap films). In [6] Brakke develops new tools in

Geomet-ric Measure Theory for the analysis of soap films: as the underlying physical problem suggests, one can represent a soap film as the superposition of two oppositely oriented currents. In order to avoid cancellations of multiplicities, the currents are defined in a covering space and, as stated in [6], the calibration technique still holds.

Let us remark that cancellations between multiplicities were a significant obstacle for the Steiner tree problem, too. The representation of currents in a covering space goes in the same direction of currents with coefficients in a group, though, as in Remark 3.16, a sort of PoincarΓ© duality occurs in the formulation of the Steiner tree problem (1-dimensional currents in ℝ𝑑) with respect to the soap film problem (currents of

codimension 1 in ℝ𝑑).

4 Existence of the calibration and open problems

Once we established that the existence of a calibration is a sufficient condition for a rectifiable 𝐺-current to be a mass-minimizer, we may wonder if the converse is also true: does a calibration (of some sort) exist for

Figure 6. Boundary data.

This problem concerns the theory of partitions of an open set Ξ© in a finite number of sets of finite perimeter. This theory was developed by Ambrosio and Braides in [1, 2], which we refer to for a complete exposition.

The analog of a calibration in this context is a null lagrangian⁴ with some special properties: again, the existence of such an object, associated with a function 𝑒, is a sufficient condition for 𝑒 to be a minimizer for the variational problem (3.8) with a given boundary condition.

We refer to [18, Section 3.2.4] for a detailed survey of the analogy.

Remark 3.15 (Clusters with multiplicities). In [19], F. Morgan applies flat chains with coefficients in a group 𝐺 to soap bubble clusters and immiscible fluids, following the idea of B. White in [25]. For a detailed comparison of [19] with our technique, see [18, Section 3.2.3]. Here we just notice that the definition of calibration in [19] works well in the case of free abelian groups and this is the main difference with our approach.

Remark 3.16 (Paired calibrations). It is worth mentioning another analogy between the technique of cal-ibrations (for currents with coefficients in a group) illustrated in this paper and the technique of paired calibrations in [17]. In particular, in the specific example of the truncated cone over the 1-skeleton of the tetrahedron in ℝ3(the surface with least area among those separating the faces of the tetrahedron), one can detect a correspondence even at the level of the main computations. See [18, Section 3.2.3] for the details. Following an idea of Federer (see [11]), in [17, 19] (and in [5, 6], as well) one can observe the exploitation of the duality between minimal surfaces and maximal flows through the same boundary. We will examine this duality in Section 4, but we conclude the present section with a remark closely related to this idea.

Remark 3.17 (Covering spaces and calibrations for soap films). In [6] Brakke develops new tools in Geomet-ric Measure Theory for the analysis of soap films: as the underlying physical problem suggests, one can 4 See [7] for an overview on null lagrangians.

(16)

34 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

represent a soap film as the superposition of two oppositely oriented currents. In order to avoid cancellations of multiplicities, the currents are defined in a covering space and, as stated in [6], the calibration technique still holds.

Let us remark that cancellations between multiplicities were a significant obstacle for the Steiner tree problem, too. The representation of currents in a covering space goes in the same direction of currents with coefficients in a group, though, as in Remark 3.16, a sort of PoincarΓ© duality occurs in the formulation of the Steiner tree problem (1-dimensional currents in ℝ𝑑) with respect to the soap film problem (currents of codimension 1 in ℝ𝑑).

4 Existence of the calibration and open problems

Once we established that the existence of a calibration is a sufficient condition for a rectifiable 𝐺-current to be a mass-minimizer, we may wonder if the converse is also true: does a calibration (of some sort) exist for every mass-minimizing rectifiable 𝐺-current?

Let us step backward: does it occur for classical integral currents? The answer is quite articulate, but we can briefly summarize the state of the art we will rely upon.

We consider a boundary 𝐡0, that is, a (π‘˜ βˆ’ 1)-dimensional rectifiable 𝐺-current without boundary, and we compare the following minima:

M𝐸(𝐡0) := min{𝕄(𝑇) : 𝑇 is a normal π‘˜-dimensional 𝐸-current, πœ•π‘‡ = 𝐡0}

and

M𝐺(𝐡0) := min{𝕄(𝑇) : 𝑇 is a rectifiable π‘˜-dimensional 𝐺-current, πœ•π‘‡ = 𝐡0}.

ObviouslyM𝐸(𝐡0) ≀M𝐺(𝐡0), the main issue is to establish whether they coincide or not. In fact, a normal

𝐸-current 𝑇 with boundary 𝐡0admits a generalized calibration if and only if 𝕄(𝑇) =M𝐸(𝐡0), as we recall in Proposition 4.2. In the classical case (𝐸 = ℝ and 𝐺 = β„€) it is known that

(i) Mℝ(𝐡0) may be strictly less thanMβ„€(𝐡0) (and, if this happens, a solution forMβ„€(𝐡0) cannot be

cali-brated),

(ii) Mβ„€(𝐡0) =Mℝ(𝐡0) if π‘˜ = 1, as we prove in Proposition 4.3.

At the end of this section, we show that this outlook changes significantly when we replace the ambient space ℝ𝑑with a suitable metric space.

Remark 4.1. For every mass-minimizing classical normal π‘˜-current 𝑇, there exists a generalized calibration πœ™ in the sense of Definition 3.5. Moreover, by means of the Riesz Representation Theorem, πœ™ can be represented by a measurable map from π‘ˆ to Ξ›π‘˜(ℝ𝑑). This result is contained in [11].

In particular, Remark 4.1 provides a positive answer to the question of the existence of a generalized cali-bration for mass-minimizing integral currents of dimension π‘˜ = 1, because minima among both normal and integral currents coincide, as we prove in Proposition 4.3. It is possible to apply the same technique in the class of normal 𝐸-currents, therefore we have the following proposition.

Proposition 4.2. For every mass-minimizing normal 𝐸-current 𝑇, there exists a generalized calibration. The following fact is probably in the folklore, unfortunately we were not able to find any literature on it. We give a proof here in order to enlighten the problems arising in the case of currents with coefficients in a group.

Proposition 4.3. Consider the boundary of an integral 1-current in ℝ𝑑, represented as 𝐡0= βˆ’ π‘βˆ’ βˆ‘ 𝑖=1π‘Žπ‘–π›Ώπ‘₯𝑖+ 𝑁+ βˆ‘ 𝑗=1𝑏𝑗𝛿𝑦𝑗, π‘Žπ‘–, 𝑏𝑗 ∈ β„•. (4.1) ThenMℝ(𝐡0) =Mβ„€(𝐡0).

(17)

Proof. Let us assume that the minimum among normal currents is attained at some current 𝑇0, that is, 𝕄(𝑇0) =Mℝ(𝐡0).

Let {π‘‡β„Ž}β„Žβˆˆβ„•be an approximation of 𝑇0made by polyhedral 1-currents such that βˆ™ 𝕄(π‘‡β„Ž) β†’ 𝕄(𝑇0) as β„Ž β†’ ∞,

βˆ™ πœ•π‘‡β„Ž= 𝐡0for all β„Ž ∈ β„•,

βˆ™ the multiplicities allowed in π‘‡β„Žare only integer multiples of 1β„Ž.

The existence of such a sequence is a consequence of the Polyhedral Approximation Theorem (see [10, Theo-rem 4.2.24] or [16] for the detailed statement and the proof). Thanks to TheoTheo-rem 2.2, it is possible to decompose such a π‘‡β„Žas a sum of two addenda:

π‘‡β„Ž= π‘ƒβ„Ž+ πΆβ„Ž, (4.2)

so that

𝕄(π‘‡β„Ž) = 𝕄(π‘ƒβ„Ž) + 𝕄(πΆβ„Ž) for all β„Ž β‰₯ 1 and

βˆ™ πœ•πΆβ„Ž= 0, so πΆβ„Žcollects the cyclical part of π‘‡β„Ž,

βˆ™ π‘ƒβ„Ždoes not admit any decomposition π‘ƒβ„Ž= 𝐴 + 𝐡 satisfying πœ•π΄ = 0 and 𝕄(π‘ƒβ„Ž) = 𝕄(𝐴) + 𝕄(𝐡).

It is clear that π‘ƒβ„Žis the sum of a certain number of polyhedral currents π‘ƒβ„Žπ‘–,𝑗each one having boundary a non-negative multiple of βˆ’1 β„Žπ›Ώπ‘₯𝑖+ 1 β„Žπ›Ώπ‘¦π‘—and satisfying 𝕄(π‘ƒβ„Ž) = βˆ‘ 𝑖,𝑗𝕄(𝑃 𝑖,𝑗 β„Ž ). We replace each 𝑃𝑖,𝑗

β„Ž with the oriented segment 𝑄𝑖,𝑗, from π‘₯𝑖to 𝑦𝑗having the same boundary as π‘ƒβ„Žπ‘–,𝑗(therefore having multiplicity a non-negative multiple of1

β„Ž). This replacement is represented in Figure 7. 18 | A. Marchese and A. Massaccesi, The Steiner tree problem revisited

It is clear that π‘ƒβ„Žis the sum of a certain number of polyhedral currents π‘ƒβ„Žπ‘–,𝑗each one having boundary a

non-negative multiple of βˆ’1 β„Žπ›Ώπ‘₯𝑖+ 1 β„Žπ›Ώπ‘¦π‘—and satisfying 𝕄(π‘ƒβ„Ž) = βˆ‘ 𝑖,𝑗𝕄(𝑃 𝑖,𝑗 β„Ž ). We replace each 𝑃𝑖,𝑗

β„Ž with the oriented segment 𝑄𝑖,𝑗, from π‘₯𝑖to 𝑦𝑗having the same boundary as π‘ƒβ„Žπ‘–,𝑗(therefore

having multiplicity a non-negative multiple of1

β„Ž). This replacement is represented in Figure 7.

𝑦

𝑗

𝑃

β„Žπ‘–,𝑗

𝑄

𝑖,π‘—β„Ž

𝐢

β„Ž

π‘₯

𝑖

Figure 7. Replacement with a segment.

Since this replacement obviously does not increase the mass, there holds

𝕄(π‘ƒβ„Ž) β‰₯ 𝕄(π‘„β„Ž),

where π‘„β„Ž= βˆ‘π‘–,𝑗𝑄𝑖,π‘—β„Ž. In other words we can write

π‘„β„Ž= ∫ 𝐼

𝑇 dπœ†β„Ž,

as an integral of currents, with respect to a discrete measure πœ†β„Žsupported on the finite set 𝐼 of unit

multi-plicity oriented segments with the first extreme among the points π‘₯1, . . . , π‘₯π‘βˆ’and second extreme among the

points 𝑦1, . . . , 𝑦𝑁+. It is also easy to see that the total variation of πœ†β„Žhas eventually the following bound from

above:

β€–πœ†β„Žβ€– ≀ min𝕄(π‘‡β„Ž) 𝑖 ΜΈ=𝑗𝑑(π‘₯𝑖, 𝑦𝑗)≀

𝕄(𝑇0) + 1

min𝑖 ΜΈ=𝑗𝑑(π‘₯𝑖, 𝑦𝑗).

Hence, up to subsequences, πœ†β„Žconverges to some positive measure πœ† on 𝐼 and so the normal 1-current

𝑄 = ∫ π‘‡βˆˆπΌ 𝑇 dπœ† satisfies πœ•π‘„ = 𝐡0 (4.3) and 𝕄(𝑄) ≀ 𝕄(𝑇0) =M𝑁(𝐡0).

In order to conclude the proof of the theorem, we need to show that 𝑄 can be replaced by an integral current 𝑅 with same boundary and mass 𝕄(𝑅) = 𝕄(𝑄) ≀M𝑁(𝐡0). Since 𝐼 is the set of unit multiplicity

Figure 7. Replacement with a segment.

Since this replacement obviously does not increase the mass, there holds 𝕄(π‘ƒβ„Ž) β‰₯ 𝕄(π‘„β„Ž),

where π‘„β„Ž= βˆ‘π‘–,𝑗𝑄𝑖,π‘—β„Ž . In other words we can write π‘„β„Ž= ∫

𝐼 𝑇 dπœ†β„Ž,

as an integral of currents, with respect to a discrete measure πœ†β„Žsupported on the finite set 𝐼 of unit multi-plicity oriented segments with the first extreme among the points π‘₯1, . . . , π‘₯π‘βˆ’and second extreme among the

- 10.1515/acv-2014-0022 Downloaded from PubFactory at 08/24/2016 10:17:03AM via UZH Hauptbibliothek / Zentralbibliothek ZΓΌrich

Riferimenti

Documenti correlati

However, the strengthening effects ob- served in the pull-out response of a single treated fibre may only par- tially extend to the macroscopic flexural behaviour of macro

In this paper we presented different state of the art feature weight- ing methods, compared their performance and proposed a more general framework to effectively apply machine

From the previous discussion it is clear that the best possible constant portfolio rule, the rule that gains all wealth when trading with (different) constant rules, is the rule

US Uber politics: elite-driven politics of concentrated actors and surrogate representation of dispersed actorsβ€”workers and consumers... While the platform economy is disruptive

transformations that, starting from the logical characterization of the problem, allow one to first replace it with a characterization based on automata on infinite words

transformations that, starting from the logical characterization of the problem, allow one to first replace it with a characterization based on automata on infinite words

We will show that the O (n) term in the preprocessing comes from the time needed to build the Cartesian Tree of A and the O (1) term in the query comes from the time needed to

Szulkin, Multiple solutions to a nonlinear Schr¨ odinger equation with Aharonov- Bohm magnetic potential, Nonlinear Differ.. Lenzmann, Mean-field limit of quantum Bose gases