Home The Yang–Mills–Higgs functional on complex line bundles: Asymptotics for critical points
Article Open Access

The Yang–Mills–Higgs functional on complex line bundles: Asymptotics for critical points

  • Giacomo Canevari , Federico Luigi Dipasquale ORCID logo EMAIL logo and Giandomenico Orlandi
Published/Copyright: September 3, 2024

Abstract

We consider a gauge-invariant Ginzburg–Landau functional (also known as Abelian Yang–Mills–Higgs model), on Hermitian line bundles over closed Riemannian manifolds of dimension n 3 . Assuming a logarithmic energy bound in the coupling parameter, we study the asymptotic behaviour of critical points in the London limit. After a convenient choice of the gauge, we show compactness of finite-energy critical points in Sobolev norms. Moreover, thanks to a suitable monotonicity formula, we prove that the energy densities of critical points, rescaled by the logarithm of the coupling parameter, converge to the weight measure of a stationary, rectifiable varifold of codimension 2.

Introduction

Let ( M , g ) be a smooth, compact, connected, oriented Riemannian manifold without boundary, of dimension n 3 . Let E M be a Hermitian line bundle over M, equipped with a (smooth) reference metric connection D 0 . For any ε > 0 , we consider the Ginzburg–Landau-type functional

(1) G ε ( u , A ) := M ( 1 2 | D A u | 2 + 1 2 | F A | 2 + 1 4 ε 2 ( 1 - | u | 2 ) 2 ) vol g .

Here u : M E is a section of the bundle, A is a real-valued 1-form on M, D A := D 0 - i A and F A is the curvature 2-form of D A . We denote the integrand of G ε as

(2) e ε ( u ε , A ε ) := 1 2 | D A u | 2 + 1 2 | F A | 2 + 1 4 ε 2 ( 1 - | u | 2 ) 2 .

The functional G ε is also known as the Abelian Yang–Mills–Higgs energy. In the present paper, we consider finite-energy critical points of G ε . More precisely, we notice that the functional G ε is well-defined and finite for ( u , A ) , where

:= { ( u , A ) L 4 ( M , E ) × L 2 ( M , T * M ) : D A u L 2 ( M , T * M E ) , F A L 2 ( M , Λ 2 T * M ) } .

One prominent feature of G ε is gauge-invariance: for any Φ W 1 , 2 ( M , 𝕊 1 ) and any pair ( u , A ) , each term of the energy density e ε is invariant under the gauge transformation

(3) ( u , A ) ( Φ u , Φ A ) := ( Φ u , A - i Φ - 1 d Φ ) ,

where Φ u is defined by the fibre-wise action of 𝕊 1 U ( 1 ) on E.

For any u and A, we define the forms (see, e.g., [33])

(4) j ( u , A ) := D A u , i u , J ( u , A ) := 1 2 d j ( u , A ) + 1 2 F A ,

called, respectively, the gauge-invariant pre-Jacobian and the gauge-invariant Jacobian of the pair ( u , A ) . If ( u , A ) is a pair with finite energy, both j ( u , A ) and J ( u , A ) are well-defined, both in the sense of distributions and pointwise, and invariant under gauge transformation. In particular, there holds the pointwise (a.e.) equality

(5) J ( u , A ) ( X , Y ) = i D A , X u , D A , Y u + 1 2 ( 1 - | u | 2 ) F A ( X , Y )

for any two smooth vector field X and Y on M.

Critical points of G ε are pairs ( u ε , A ε ) that satisfy the Euler–Lagrange equations

(6) D A ε * D A ε u ε + 1 ε 2 ( | u ε | 2 - 1 ) u ε = 0 ,
(7) d * F ε = j ( u ε , A ε )

in the sense of distributions. Here D A * is the L 2 -adjoint of D A , F ε := F A ε and d * denotes the codifferential. By taking the differential in both sides of (7), we obtain the London equation (at level ε)

(8) - Δ F ε + F ε = 2 J ( u ε , A ε ) .

Throughout the paper, we assume that { ( u ε , A ε ) } is a sequence of critical points of G ε satisfying the logarithmic energy bound

(9) G ε ( u ε , A ε ) Λ | log ε |

for some constant Λ > 0 that does not depend on ε. This assumption is natural, as the energy of minimisers of G ε is precisely of order | log ε | , whenever the bundle E M is non-trivial [13, Theorem A and Remark 2]. Under this assumption, we have this (preliminary) convergence result:

Theorem 0 ([13]).

Let { ( u ε , A ε ) } E be a sequence of critical points of G ε satisfying the logarithmic energy bound (9). Then there exist a (non-relabelled) subsequence, 2-forms J * , F * and a 1-form j * such that the following properties hold:

  1. J ( u ε , A ε ) J * strongly in W - 1 , p ( M , Λ 2 T * M ) for any p [ 1 , n n - 1 ) ,

  2. F A ε F * strongly in W 1 , p ( M , Λ 2 T * M ) for any p [ 1 , n n - 1 ) ,

  3. j ( u ε , A ε ) j * strongly in L p ( M , Λ 2 T * M ) for any p [ 1 , n n - 1 ) ,

  4. the Hodge dual J * is a boundary-less, integer-multiplicity rectifiable ( n - 2 ) - current in M , whose integral homology class 𝒞 H n - 2 ( M ; ) is Poincaré dual to the first Chern class of E M ,

  5. the limit curvature F * satisfies

    (10) d * F * = j * , - Δ F * + F * = 2 π J * ,

    in the sense of distributions.

The proof of Theorem 0 is contained in [13] (see, in particular, Theorem A and Remark 3.12 therin). Property (iii) is not mentioned explicitly in [13], but it is an immediate consequence of (7) and (ii). Note that Property (i) is completely independent of the Euler–Lagrange equations (6)–(7); it holds true for any sequence { ( u ε , A ε ) } that satisfies (9).

Remark 1.

If ( u ε , A ε ) is a critical point of G ε , then, in particular, u ε belongs to the space ( W 1 , 2 L ) ( M , E ) and satisfies | u ε | 1 in M (see Proposition 1.2 and Remark 1.5 below). Moreover, one immediately checks that d A ε L 2 ( M , Λ 2 T * M ) . Although in [13] we worked under the assumption that

( u ε , A ε ) W 1 , 2 ( M , E ) × W 1 , 2 ( M , T * M ) ,

a quick inspection of the proofs and a comparison with the notion of “optimised pair” ([13, Definition 3.1]) shows that, as a matter of fact, all that we really use in the proof of the convergence results in Theorem 0 are (i) the logarithmic energy bound (9); (ii) criticality; (iii) the bound | u ε | 1 in M. More generally, it is not difficult to check that, with Proposition 1.2 at hand, in [13] the assumption ( u ε , A ε ) W 1 , 2 ( M , E ) × W 1 , 2 ( M , T * M ) can be dropped everywhere and replaced by ( u ε , A ε ) .

Main results.

In this paper, we will address two more questions about the asymptotic behaviour, as ε 0 , of critical points satisfying (9). Our first main result describes the concentration of the rescaled energy densities. We denote by spt μ the support of a Radon measure μ and by n - 2 the ( n - 2 ) -dimensional Hausdorff measure.

Theorem 1.

Let { ( u ε , A ε ) } E be any sequence of critical points of G ε satisfying the logarithmic energy bound (9). Then, up to extraction of a subsequence, we have

e ε ( u ε , A ε ) | log ε | vol g * μ * as  ε 0

in the sense of Radon measures in M. The limit measure μ * satisfies H n - 2 ( spt μ * ) < + and, moreover, it is the weight measure of a stationary, rectifiable ( n - 2 ) -varifold in M.

Recall that varifolds are measure-theoretic generalisations of smooth submanifolds (see, e.g., [35] and Section 5 below for details). In particular, stationary rectifiable varifolds are generalisations of minimal surfaces. Constructing non-trivial stationary, rectifiable varifolds in Riemannian manifolds is a fundamental problem in Geometric Measure Theory [3, 31].

Remark 2.

By Γ-convergence [13, Theorem A and Remark 3.13], the limit measure μ * satisfies μ * π | J * | , where J * is the limit Jacobian, given by Theorem 0. In particular, μ * it is always nonzero, so long as the bundle E M is non-trivial. Moreover, if { ( u ε , A ε ) } is a sequence of minimisers, then μ * = π | J * | and the Hodge dual J * is area-minimising in its integral homology class 𝒞 (see [13, Corollary B]). However, we do not know whether the inequality μ * π | J * | may be strict, in general. A related question, which we do not address here, is whether or not μ * is an integral varifold (see the paragraph on open questions below).

Remark 3.

A more detailed description of μ * can be found in Theorem 5.3 in Section 5, of which Theorem 1 is an abridged version. Contrary to what is observed in related problems (e.g., [9, 38]), the measure μ * contains no diffuse part; see Remark 5.4.

The proof of Theorem 1 relies on a suitable monotonicity formula (see Theorem 3 below).

In our second main result, we prove compactness for the sequence { ( u ε , A ε ) } in W 1 , p ( M , E ) × W 2 , p ( M , T * M ) , for any p with 1 p < n n - 1 , so long as A ε is in Coulomb gauge, that is,

(11) A ε = d * ψ ε + ζ ε ,

where ψ ε is an exact 2-form, ζ ε is a harmonic 1-form and ζ ε L ( M ) C M for some constant C M that depends on M only (see, e.g., [13, Lemma 2.10]).

Theorem 2.

Let { ( u ε , A ε ) } E be a sequence of critical points of G ε . Assume that { ( u ε , A ε ) } satisfies the logarithmic energy bound (9) and that each A ε satisfies (11). Then, up to extraction of a (non-relabelled) subsequence, there holds

u ε u * strongly in  W 1 , p ( M , E ) , A ε A * strongly in  W 2 , p ( M , T * M )

as ε 0 , for any p with 1 p < n n - 1 . Moreover, u * and A * are smooth in M spt μ * , they satisfy | u * | = 1 in M spt μ * and

(12) j ( u * , A * ) = d * F A *

in the sense of distributions in M.

Remark 4.

By continuity of the Jacobian and the differential operator, the limits u * and A * are compatible with J * , F * , j * given by Theorem 0 – that is, F * = F A * , J * = J ( u * , A * ) , and j * = j ( u * , A * ) . Moreover, equation (12) implies that u * is an A * -harmonic unit section away from the support of μ * , where both u * and A * are smooth. In other words, there holds

(13) D A * * D A * u * = | D A * u * | 2 u * pointwise in  M spt μ *

(see Remark 2.4 for details). Finally, by passing to the limit in (11), it follows that A * = d * ψ * + ζ * , where ψ * is an exact 2-form and ζ * is a harmonic 1-form.

Remark 5.

Theorem 2 applies, in particular, to any sequence of minimisers of G ε in the class . Indeed, by [13, Remark 2], minimisers automatically satisfy assumption (9).

Before illustrating the ideas of the proof, let us motivate the interest towards our results.

Background and motivation.

Ginzburg–Landau functionals of the form (1) were originally proposed as a model of superconductivity. This theory gained much popularity, as it accounts for most commonly observed effects in superconductors, and it gradually became relevant to other areas of physics, such as particle physics and gauge theories. The asymptotic regime as ε 0 , which is known as the London limit in superconductivity theory or the strongly repulsive limit in particle physics, is characterised by the emergence of topological singularities: due to topological obstructions set by the structure of the bundle E M , critical points develop singularities, in the asymptotic limit, which are supported on a set of dimension n - 2 . The topologically-driven energy concentration phenomenon can be detected by the distributional Jacobian, which both enjoys remarkable compactness properties and identifies the topological singularities that emerge in the limit as ε 0 (see, e.g., [23, 1, 2]).

The asymptotic analysis of (1) in the limit as ε 0 relies on analogous results for the “non-magnetic” version of the functional, in which the variable A is set to zero and the energy is considered as a functional of u only. The analysis of critical points of Ginzburg–Landau-type functionals without magnetic field was initiated by Bethuel, Brezis, and Hélein [7] in the 2-dimensional, Euclidean setting. The analysis was then extended to critical points on 2-dimensional Riemann surfaces [5], higher-dimensional Euclidean domains [25, 8] and, more recently, higher-dimensional manifolds [14, 38, 15, 16]. In particular, our Theorems 2 and 1 generalise results that were first obtained in [8] in the Euclidean, non-magnetic case. In all these works, a main difficulty lies in the lack of uniform energy bounds, as the natural energy scaling considered is the logarithmic one (9).

As for the gauge-invariant, “magnetic” version of the functional, the first results addressed the asymptotic behaviour of minimisers in two-dimensional Euclidean domains [10] and on Riemann surfaces [28, 32]. A detailed analysis of the asymptotics of critical points for the gauge-invariant functional in two dimensions, in the Euclidean setting and in the logarithmic energy regime, has been carried out by Sandier and Serfaty – see, e.g., [33, Chapter 13] and the references therein. Apparently, less effort has been put in studying the asymptotic behaviour of critical points in higher dimensions. Recently, this problem has been addressed in [29], in the context of Hermitian line bundles, for the self-dual version G ε self of G ε , in which the curvature term M | F ε | 2 vol g is replaced by ε 2 M | F ε | 2 vol g . The main outcome of [29] is that, for a sequence { ( u ε , A ε ) } of critical points with bounded G ε self -energy, the energy densities e ε self ( u ε , A ε ) converge, up to subsequences, to the weight measure of a limiting stationary, rectifiable, integral varifold of codimension 2 in M. Conversely, in [16], it is shown that any non-degenerate minimal submanifold of codimension 2 in M can be obtained as energy concentration set of a sequences of bounded energy critical points of G ε self .

Proofs of the main results: A sketch.

In the paper, we first address the proof of Theorem 2, which ultimately relies on the L p -compactness for the prejacobians j ( u ε , A ε ) given by Theorem 0. Instead, the proof of Theorem 1 combines ideas from [8] in order to deal with the logarithmic energy regime and from [28, 29] to deal with the Hermitian line bundle setting. The most difficult part is to identify μ * as the weight measure of a stationary, rectifiable ( n - 2 ) -varifold. The key point is showing that, if μ * ( R ( x 0 ) ) is smaller than η 0 R n - 2 , where η 0 is suitable uniform “small” constant, then μ * ( R / 2 ( x 0 ) ) = 0 . This is shown in Lemma 5.6. The proof goes along the lines of that of [8, Proposition VIII.1] and it is based on a clearing-out result (see the analysis in Section 4, in particular Proposition 4.1 and Proposition 4.6).

A crucial rôle is played by the following (almost) ( n - 2 ) -monotonicity inequality. Let inj ( M ) be the injectivity radius of M. For x 0 M and r ( 0 , inj ( M ) ) , we denote by r ( x 0 ) the geodesic ball of radius r centred at x 0 . Given ( u ε , A ε ) , we define

(14) E ε ( x 0 , ρ ) := ρ ( x 0 ) e ε ( u ε , A ε ) vol g ,

and

(15) X ε ( x 0 , ρ ) := ρ ( x 0 ) ( | D A ε , ν u ε | 2 + | i ν F ε | 2 ) vol g ^ + 1 2 ε 2 ρ ρ ( x 0 ) ( 1 - | u ε | 2 ) 2 vol g ,

where ν is the exterior unit normal to B ρ ( x 0 ) , i ν denotes the contraction against ν (see (18) below), F ε := F A ε and g ^ is the metric induced by g on ρ ( x 0 ) .

Theorem 3.

There exist α ( 0 , 2 ) , R 0 ( 0 , inj ( M ) ) , and C > 0 , all depending only on M and g, such that the following holds. If ( u ε , A ε ) E is any critical point of G ε and x 0 is any point of M, then the inequality

(16) e C r 2 r n - 2 E ε ( x 0 , r ) + r R e C ρ 2 ρ n - 2 X ε ( x 0 , ρ ) d ρ e C R 2 R n - 2 E ε ( x 0 , R ) + 2 C α e C R 2 ( R α - r α ) R n - 2 + α E ε ( x 0 , R )

holds for any 0 < r < R < R 0 .

The proof of Theorem 3 is delicate. A classical argument (see, e.g., Proposition 3.3 below), based on the Pohozaev identity, proves that

(17) d d ρ ( e C ρ 2 E ε ( x 0 , ρ ) ρ n - 4 ) 0

for any x 0 M , ρ ( 0 , inj ( M ) ) and some uniform constant C > 0 . The exponent of n - 4 at the denominator appears for scaling reasons (heuristically, the curvature F ε scales as a second derivative, so the integral of | F ε | 2 scales as a length to the power of n - 4 ). Inequality (17) is known to be sharp in the context of non-Abelian Yang–Mills–Higgs theories [36]. However, in our Abelian setting, Theorem 3 gives an (almost) monotonicity formula for the energy rescaled by ρ n - 2 . A similar result is obtained in [29, Section 4], for the Ginzburg–Landau functional in the self-dual scaling. However, the arguments of [29] fall apart in our setting, because we work with a different scaling of the energy. Instead, we exploit the properties of the Euler–Lagrange equations. In particular, our arguments rely on equation (7), which can be reformulated as a uniformly elliptic equation for the variable A ε . By elliptic regularity estimates, we can control the curvature terms and improve on the monotonicity, obtaining in the end Theorem 3. As a byproduct of our arguments, we obtain a bound for the Morrey–Campanato L q , 2 -norm of F ε , in terms of the rescaled energy (14), for any q < n (see Proposition 3.8 below).

Back on the proof of Theorem 1, the clearing-out property combined with the energy bound (9) implies a two-sided estimate for the ( n - 2 ) -density of μ * , see Lemma 5.7. Following ideas in [8, pp. 498–499] (extended first in the gauge-invariant setting in [33] and to the context of Hermitian line bundles in [29, Section 6]), the conclusion is achieved once we characterise μ * in terms of the stress-energy tensor fields T ε , defined in (5.3) below. Indeed, by criticality of ( u ε , A ε ) , each tensor field T ε is stationary (i.e., divergence-free). Since stationarity is a linear condition, their weak limit T * is still stationary. On the other hand, T * is naturally identified with an ( n - 2 ) -generalised varifold in M whose weight measure | T * | is absolutely continuous with respect to μ * . Then, by Lemma 5.7, we can apply a suitable version of the Ambrosio-Soner rectifiability criterion [4, Theorem 3.1] to deduce that T * is also rectifiable and that μ * = | T * | , whence the conclusion of Theorem 1 follows.

Further comments and open questions.

We did not address specifically the problem of the existence, in full generality, of non-minimising critical points of G ε . We believe that they could be constructed by adapting the general framework developed in [24].

Another interesting question that we have not addressed in this paper is whether the varifold 1 π μ * is integral or not. For the magnetic Ginzburg–Landau functional in the self-dual scaling, it is known that the measure 1 π μ * is integral [29]. By contrast, for critical points of the non-magnetic Ginzburg–Landau functional (5.4) in Euclidean domains, the situation is more involved [30] – that is, the density of the limiting varifold 1 π μ * at a generic point is either 1 or possibly a non-integer number larger or equal than 2. Unfortunately, the arguments in [29] are really tailored to the bounded energy regime and do not carry over to our setting.

Finally, it would be interesting to study, along the lines of [16], whether any non-degenerate minimal codimension 2 submanifold of M can be obtained as energy concentration set of critical points of G ε , for an appropriate choice of the total space E of the bundle E M . We leave all these questions to future investigation.

Organisation of the paper.

The paper is organised as follows. In Section 1 we derive several a priori estimates for gauge-invariant quantities, such as | u ε | , | D A ε u ε | , and | F ε | , that are used throughout the whole paper. In Section 2 we prove Theorem 2. In Section 3 we prove the almost ( n - 2 ) -monotonicity inequality, i.e., Theorem 3. In Section 4, the longest and by far the most technical of the paper, we prove the clearing-out property with its consequences (Propositions 4.1 and 4.6). In the last section, Section 5, we complete the proof of Theorem 1.

The paper is completed by two appendices containing technical results concerning Poincaré and trace-type inequalities for differential forms (Appendix A) and elliptic estimates for the Hodge Laplacian (Appendix B). These results are certainly known to experts but we provide full proofs because they are crucial to our arguments and we could not find the exact statements we needed in the available literature.

Notation.

In this paper, we use the symbol { X ε } to denote a family of objects indexed by the parameter ε > 0 . Although in some instances { X ε } may also be considered as a continuous family, we always look at it as a sequence. In other words, { X ε } is a shorthand for the sequence { X ε k } k , where ε k 0 as k + . Usually, for the sake of a lighter notation, we do not relabel subsequences.

In inequalities like A B , the symbol means that there exists a constant C, independent of A and B, such that A C B . In particular, dealing with sequences indexed by ε, we use to denote inequality up to a constant independent of ε. Whenever it is relevant, we keep track of the dependences of the implicit constants. In most cases, they will depend on the manifold ( M , g ) and on the constant Λ in the energy bound (9). Since we only consider Riemannian manifolds ( M , g ) as base spaces of E M and the metric g on M is fixed from the very beginning, we will sometimes write M as a shorthand for ( M , g ) .

We denote Λ k T * M the bundle of k-forms on M. Sobolev spaces of sections of a bundle E M are denoted W k , p ( M , E ) . In particular, W 1 , 2 ( M , T * M ) is the Sobolev space of one-forms on M of class W 1 , 2 . More details on such spaces can be found, for instance, in [13, Appendix A]. If ω Λ k T * M is a k-form on M and X is a vector field, we denote by i X ω the contraction of ω with X, defined by setting

(18) i X ω ( X 1 , , X k - 1 ) := ω ( X , X 1 , , X k - 1 )

for any vector fields X 1 , , X k - 1 on M.

Finally, as mentioned above, we will always write inj ( M ) for the injectivity radius of M and r ( x 0 ) for the geodesic ball in M of centre x 0 M and r > 0 .

1 Preliminary estimates for gauge-invariant quantities

We recall the definition of the “energy space”,

:= { ( u , A ) L 4 ( M , E ) × L 2 ( M , T * M ) : D A u L 2 ( M , T * M E ) , F A L 2 ( M , Λ 2 T * M ) } .

We also recall the following decomposition of D A u which we will use several times.

Lemma 1.1 ([13, Lemma 2.2]).

For any ( u , A ) E , there holds

(1.1) D A u = d ( | u | ) | u | u + j ( u , A ) | u | 2 i u

a.e. on the set { u 0 } .

In [13], this result is stated for ( u , A ) W 1 , 2 ( M , E ) × W 1 , 2 ( M , T * M ) , but the proof carries over to the case ( u , A ) verbatim. Below, we collect some useful consequences of equations (6), (7), and (8) satisfied by critical points of the functionals G ε . All the results in this section are gauge-invariant and will be used repeatedly in the rest of the paper.

L -bounds.

Let ( u ε , A ε ) be a critical point of G ε . Elliptic regularity theory implies that, for each ε > 0 , the pair ( u ε , A ε ) is smooth up to a suitable gauge transformation (see, e.g., [22] or [29, Appendix]). In particular, gauge-invariant quantities such as | u ε | or | D A ε u ε | are continuous (and their squares are smooth). Below, we prove several a priori estimates on u ε , D A ε u ε , and F A ε . We will apply these estimates to obtain compactness results. For the reader’s convenience, we provide self-contained proofs of the estimates we need. First, we prove:

Proposition 1.2.

Let ε > 0 and let ( u ε , A ε ) E be any critical point of G ε . Then | u ε | satisfies

(1.2) - 1 2 Δ ( | u ε | 2 ) + 1 ε 2 ( | u ε | 2 - 1 ) | u ε | 2 + | D A ε u ε | 2 = 0

in the sense of distributions on M and

(1.3) | u ε | 1 a.e. in  M .

Remark 1.3.

The assumption ( u ε , A ε ) and (1.3) imply that | u ε | 2 W 1 , 2 ( M ) and | D A ε u ε | 2 L 1 ( M ) . In particular, it makes sense to test (1.2) against functions in ( L W 1 , 2 ) ( M ) .

In the proof Proposition 1.2, we will make use of an auxiliary result, Lemma 1.4 below. Before stating it, we recall that for any ( u , A ) , the covariant derivative D A u belongs to L 2 ( M , T * M E ) . This allows us to define a linear functional D A * D A u : C ( M , E ) by setting, for any test section v C ( M , E ) ,

(1.4) ( D A * D A u ) ( v ) := M D A u , D A v vol g .

Since such linear functional is sequentially continuous with respect to convergence of test sections on E, it defines a distribution on E M . Furthermore, the right-hand side of (1.4) still makes sense if we plug in it any measurable section v such that D A v L 2 ( M , T * M E ) .

Similarly, for ( u , A ) , we can define a linear functional D A * D A u , u : C ( M , ) by setting, for any test function φ C ( M , ) ,

(1.5) ( D A * D A u , u ) ( φ ) := M D A u , D A ( u φ ) vol g .

As it is immediately checked, the functional D A * D A u , u is sequentially continuous with respect to convergence of test functions, hence it is a well-defined distribution on M. In particular, both D A * D A u and D A * D A u , u are well-defined distributions for any ( u , A ) . We are now ready to state Lemma 1.4.

Lemma 1.4.

For any ( u , A ) E and any φ W 1 , 2 ( M ) such that either u L ( M , E ) or φ W 1 , ( M ) , there holds

(1.6) D A u , d φ u = 1 2 d ( | u | 2 ) , d φ

pointwise a.e. in M. Moreover, for any ( u , A ) E we have

(1.7) - 1 2 Δ ( | u | 2 ) = D A * D A u , u - | D A u | 2

in the sense of distributions on M.

Proof.

Let us focus on (1.6) first. Let ( u , A ) , φ W 1 , 2 ( M ) be as above. Let x M , and let { τ 1 , , τ n } be an orthonormal basis of T x M . The scalar product on T * M E is the one induced by the Hermitian form on E and the Riemannian metric on M; therefore, at the given point x, we have

(1.8) D A u , d φ u = j = 1 n D A τ j u , u φ τ j .

However, the right-hand side of (1.8) is also the scalar product between the (real-valued) 1-forms D A u , u and d φ , evaluated at x. As D A u , u L 1 ( M , T * M ) and the connection D A is compatible with the Hermitian form on E, we have D A u , u = d ( | u | 2 2 ) in the sense of distributions and pointwise a.e. As a consequence, we have

D A u , d φ u = D A u , u , d φ = 1 2 d ( | u | 2 ) , d φ ,

which proves (1.8). Now, for any ( u , A ) and any smooth φ, we have

( D A * D A u , u ) ( φ ) - M | D A u | 2 φ vol g = M D A u , d φ u vol g = 1 2 M d ( | u | 2 ) , d φ vol g

because of (1.5) and (1.6). The conclusion follows. ∎

Proof of Proposition 1.2.

Let ( u ε , A ε ) be a solution to system (6)–(7) in the sense of distributions. We fix a number k > 0 and define

u ε , k := { u ε where  | u ε | < k , k u ε | u ε | where  | u ε | k .

At almost every point of the set { | u ε | k } , we have

D A ε u ε , k = - k | u ε | 2 d ( | u ε | ) u ε + k | u ε | D A ε u ε = k | u ε | 3 j ( u ε , A ε ) i u ε ,

where the second equality follows from Lemma 1.1. As a consequence, we have

(1.9) | D A ε u ε , k | min ( 1 , k | u ε | ) | D A ε u ε | a.e. in  M .

Moreover, we have | D 0 u ε , k | | D A ε u ε , k | + k | A ε | a.e. in M, therefore

(1.10) u ε , k ( W 1 , 2 L ) ( M , E ) ,

as A ε L 2 ( M , T * M ) by the definition of . Now, we define

ξ ε := { 0 where  | u ε | < 1 , | u ε | 2 - 1 where  1 | u ε | < 2 , 3 where  | u ε | 2 .

We have d ξ ε = D A ε u ε , u ε η { 1 | u ε | 2 } , where η A denotes the indicator function of a set A, and hence

(1.11) ξ ε ( W 1 , 2 L ) ( M ) .

By formally testing equation (6) against ξ ε u ε , k , we obtain

(1.12) M ( D A ε u ε , D A ε u ε , k ξ ε + D A ε u ε , d ξ ε u ε , k ) vol g + 1 ε 2 M ( | u ε | 2 - 1 ) u ε , u ε , k ξ ε vol g = 0 .

More rigorously, equation (1.12) is deduced from (6) via an approximation argument, using (1.10) and (1.11). Now, we pass to the limit as k + in each term of (1.12). For the first term, we apply (1.9) and Lebesgue’s dominated convergence theorem:

(1.13) M D A ε u ε , D A ε u ε , k ξ ε , k M | D A ε u ε | 2 ξ ε 0 as  k + .

For the second term, when k 2 we have

(1.14) M D A ε u ε , d ξ ε u ε , k vol g = 1 2 { 1 | u ε | 2 } D A ε u ε , k , d ( | u ε , k | 2 ) u ε , k vol g = 1 2 { 1 | u ε | 2 } | D A ε u ε , k | 2 vol g 0

because of (1.6). Finally, we can pass to the limit in the third term of (1.12) by monotone convergence. Keeping (1.13) and (1.14) into account, we deduce

(1.15) M ( | u ε | 2 - 1 ) | u ε | 2 ξ ε vol g 0 .

As a consequence, we obtain that | u ε | 1 a.e. in M. By Lemma 1.4, ( u ε , A ε ) satisfies (1.7) in the sense of distributions on M. Combining (1.7) with (6), we immediately obtain (1.2). ∎

Remark 1.5.

Assume ( u ε , A ε ) is a critical point of G ε . From the bound | u ε | 1 in Proposition 1.2 and the assumption A L 2 ( M , T * M ) , it follows that D 0 u ε L 2 ( M , E ) and, therefore,

u ε ( W 1 , 2 L ) ( M , E ) with  | u ε | 1  in  M ,

whenever ( u ε , A ε ) is a critical pair of G ε .

Next, we state an L -bound for D A ε u ε . Precisely:

Lemma 1.6.

Let ε > 0 and let ( u ε , A ε ) E be a critical point of G ε satisfying (9). Then

(1.16) D A ε u ε L ( M ) ε - 1 .

The proof of Lemma 1.6 is postponed to Appendix C.

Now, we provide a bound on the L ( M ) -norm of { F ε } .

Lemma 1.7.

Let ε > 0 and let ( u ε , A ε ) E be a critical point of G ε . Then

F ε L ( M ) ε - 1 .

Proof.

The Euler–Lagrange equation (7), together with Proposition 1.2 and Lemma 1.6, implies

(1.17) d * F ε L ( M ) = j ( u ε , A ε ) L ( M ) u ε L ( M ) D A ε u ε L ( M ) ε - 1 .

Let F 0 be the curvature of the reference connection D 0 . The difference F ε - F 0 = d A ε is an exact form and, in particular, it is orthogonal to all harmonic 2-forms in M. Then, for any p ( n , + ) , the Gaffney inequality can be written in the form

F ε - F 0 W 1 , p ( M ) C p d * F ε - d * F 0 L p ( M )

for some constant C p that depend only on p and M (see, e.g., [21, Theorem 4.11]). Now, the form F 0 is smooth and does not depend on ε, so

F ε W 1 , p ( M ) C p ( d * F ε L p ( M ) + 1 ) C p ε - 1 .

The lemma follows by Sobolev embedding. ∎

Compactness for gauge-invariant quantities.

In this subsection, we apply the bounds obtained above, as well as results from [13], to prove compactness results for gauge-invariant quantities, such as the norm | u ε | and the pre-Jacobian j ( u ε , A ε ) of a sequence of critical points ( u ε , A ε ) that satisfies (9).

Proposition 1.8.

Let { ( u ε , A ε ) } E be a sequence of critical point of G ε satisfying the logarithmic energy bound (9). Then, for any p with 1 p < 2 , there exists a constant C p , depending only on Λ, p, M, and g, such that

(1.18) d ( | u ε | ) L p ( M ) C p .

Moreover, there holds

(1.19) M | d ( | u ε | ) | p vol g 0 as  ε 0 .

Proof.

We follow the argument of [7, Lemma X.13]. Set ρ ε := | u ε | and rewrite (1.2) in the form

(1.20) 1 2 Δ ρ ε 2 = 1 ε 2 ρ ε 2 ( ρ ε 2 - 1 ) + | D A ε u ε | 2 .

Next, for any 0 < ε < 1 , we define:

(1.21) ρ ~ ε := max { ρ ε , 1 - 1 | log ε | 2 } , M ε := { x M : ρ ε ~ ( x ) = ρ ε ( x ) } , N ε := M M ε .

Notice that on M ε there holds ρ ε = | u ε | 1 - 1 | log ε | 2 for any ε ( 0 , 1 ) . Moreover, as ρ ε belongs to ( L W 1 , 2 ) ( M ) , it follows that ρ ~ ε belongs to ( L W 1 , 2 ) ( M ) as well, for any ε ( 0 , 1 ) . Since M is compact, also 1 - ρ ~ ε belongs to ( L W 1 , 2 ) ( M ) and, by Remark 1.3, we can use 1 - ρ ~ ε as a test function in (1.20). Since M is compact, ρ ~ ε is bounded in M and moreover constant in N ε , we obtain

(1.22) M ε ρ ε | d ρ ε | 2 vol g M ( 1 - ρ ~ ε ) | D A ε u ε | 2 vol g .

Since 1 - ρ ~ ε 1 | log ε | 2 , by (1.22) and (9) we infer

(1.23) M ε ρ ε | d ρ ε | 2 vol g 1 | log ε | 0 as  ε 0 .

Being ρ ε = | u ε | 1 - 1 | log ε | 2 on M ε , (1.23) gives

(1.24) d ( | u ε | ) L p ( M ε ) 1

(where the implicit constant depends only on Λ, p, M, and g) and

(1.25) M ε | d ρ ε | 2 vol g 0 as  ε 0 .

Now, we show that the measure of the “bad” set N ε decreases at a quantitative rate as ε 0 . Indeed, the logarithmic energy bound (9) and (1.3) imply

| log ε | 1 ε 2 N ε ( 1 - ρ ε 2 ) 2 vol g 1 ε 2 N ε ( 1 - ρ ε ) 2 vol g | N ε | | log ε | 4 ε 2 ,

whence

(1.26) | N ε | | log ε | 5 ε 2 .

By (1.1), it is straightforward to see that, for any ε > 0 ,

(1.27) | d ρ ε | | D A ε u ε | a.e. on  M .

Combining (1.26) with (1.16) (which holds a.e. on M), we get

(1.28) D A ε u ε L p ( N ε ) ε 2 p - 1 | log ε | 5 p 0 as  ε 0 .

Thus, by (1.28) and (1.27),

(1.29) N ε | d ρ ε | p vol g 0 as  ε 0

for any p with 1 p < 2 . Moreover, along with (1.23) and the definition of M ε , (1.27) and (1.28) imply (1.18). Finally, combining (1.29) with (1.25) yields (1.19), completing the proof of the proposition. ∎

By combining the compactness result in [13, Theorem A] with the properties of the London equation, and reasoning along the lines of [13, Corollary B], we can extract a subsequence and find limits J * , F * such that

(1.30) J ( u ε , A ε ) π J * strongly in  W - 1 , p ( M ) , F ε F * strongly in  W 1 , p ( M )

for any p with 1 p < n n - 1 . In fact, F * and J * satisfy

(1.31) - Δ F * + F * = 2 π J * ,

that is, the London equation in (10). We claim that the curvature energy grows less than logarithmically in ε. This fact will be useful to our analysis of the properties of the energy concentration measure, in Section 5.

Lemma 1.9.

Let { ( u ε , A ε ) } ε > 0 E be a sequence of finite-energy solutions of (6)–(7) that satisfies (9). Then

lim ε 0 1 | log ε | M | F ε | 2 vol g = 0 .

The proof of Lemma 1.9 is very similar to arguments we used already in [13, Section 4].

Proof of Lemma 1.9.

The curvature form F ε is, by definition, closed. Therefore, by Hodge decomposition, we can write F ε = d φ ε + ξ ε , where ξ ε is a harmonic 2-form and φ ε is a co-exact 1-form such that

(1.32) φ ε W 1 , 2 ( M ) + ξ ε L 2 ( M ) F ε L 2 ( M ) ( 9 ) | log ε | 1 2

(see all the discussion from [13, Proposition A.14]). As the space of harmonic forms is finite-dimensional, the W 1 , 2 ( M ) -norm of ξ ε , too, is of order | log ε | 1 2 at most. From (7), we obtain

- Δ φ ε = d * d φ ε = d * F ε = j ( u ε , A ε ) .

Since φ ε is co-exact, hence orthogonal to all harmonic 1-forms, it follows from elliptic regularity theory that the W 2 , 2 ( M ) -norm of φ ε is bounded from above by the L 2 ( M ) -norm of - Δ φ ε . We deduce

(1.33) φ ε W 2 , 2 ( M ) j ( u ε , A ε ) L 2 ( M ) D A ε u ε L 2 ( M ) u ε L ( M ) | log ε | 1 2

(the last inequality follows from (9) and (1.3)). Therefore, we have

F ε W 1 , 2 ( M ) φ ε W 2 , 2 ( M ) + ξ ε W 1 , 2 ( M ) ( 1.32 ) , ( 1.33 ) | log ε | 1 2

and hence, by Sobolev embedding,

(1.34) F ε L q ( M ) | log ε | 1 2 with  q := 2 * = 2 n n - 2 .

From (1.30) and (1.34) we obtain, by interpolation,

(1.35) F ε L 2 ( M ) F ε L q ( M ) α F ε L 1 ( M ) 1 - α | log ε | α 2

with α := n n + 2 ( 0 , 1 ) . The lemma follows. ∎

Next, we notice that criticality and L p ( M ) -compactness of { d * F A ε } for any p with 1 p < n n - 1 (cf. [13, Proposition 3.11 (i) and Remark 3.12]) imply compactness of the prejacobians j ( u ε , A ε ) in L p ( M , T * M ) for any 1 p < n n - 1 .

Lemma 1.10.

Let { ( u ε , A ε ) } E be a sequence of critical points of G ε satisfying the logarithmic energy bound (9). Then there exists j * L p ( M , T * M ) for any p with 1 p < n n - 1 , such that, up to a (not relabelled) subsequence, there holds

(1.36) j ( u ε , A ε ) j * in  L p ( M , T * M ) as  ε 0

for any p with 1 p < n n - 1 . In particular, there holds

(1.37) j ( u ε , A ε ) L p ( M ) 1

for any p with 1 p < n n - 1 , where the implicit constant in front of the right-hand-side depends only on Λ, p, M and g.

Proof.

Set, for short, j ε := j ( u ε , A ε ) and F ε := F A ε . To prove (1.36), it is enough to show that { j ε } contains a Cauchy subsequence in (the Banach space) L p ( M , T * M ) , for any p with 1 p < n n - 1 . Due to [13, Proposition 3.11 (i) and Remark 3.12], we can extract a subsequence ε k 0 such that { F ε k } is a Cauchy sequence in W 1 , p ( M , Λ 2 T * M ) , for any p with 1 p < n n - 1 . But then, by (7), for any k, m we have

j ε k - j ε m L p ( M ) = d * ( F ε k - F ε m ) L p ( M ) .

This proves (1.36). If (1.37) were false, there would be a (countable) sequence of critical points { ( u ε , A ε ) } that satisfy (9) (for some ε-independent constant Λ) and are such that j ( u ε , A ε ) L p ( M ) + as ε 0 . This contradicts (1.36). Therefore, (1.37) is proved. ∎

We apply Proposition 1.8 and Lemma 1.10 to prove a uniform L p ( M ) -bound on { D A ε u ε } .

Lemma 1.11.

Let { ( u ε , A ε ) } E be a sequence of critical points of G ε satisfying the logarithmic energy bound (9). Then, for any p with 1 p < n n - 1 there exists a constant C p > 0 (depending on Λ, p, M and g) such that

(1.38) D A ε u ε L p ( M ) C p

for any ε > 0 . Moreover, there exists a (non-relabelled) subsequence such that

(1.39) D A ε u ε L p ( M ) j * L p ( M ) for any  1 p < n n - 1 , as  0 ,

where j * is given by Lemma 1.10.

Proof.

Define the sets M ε and N ε as in (1.21).

Proof of (1.38). Let p be fixed such that 1 p < n n - 1 . If (1.38) were false, we could extract a (non-relabelled) subsequence) in such a way that

(1.40) D A ε u ε L p ( M ) + as  ε 0

Let ε 2 = e - 2 so that 1 | log ε 2 | 2 = 1 2 , and notice that | u ε | > 1 2 on M ε for any ε with 0 < ε < ε 2 . Then, by (1.1),

| D A ε u ε | 2 ( | d ( | u ε | ) | + | j ( u ε , A ε ) | ) ) on  M ε ,

whence

(1.41) | D A ε u ε | p 2 p ( | d ( | u ε | ) | p + | j ( u ε , A ε ) | p ) on  M ε ,

for any ε with 0 < ε < ε 2 . Combining (1.41) with (1.18) and (1.37), we obtain that for any p with 1 p < n n - 1 there exists a constant C p > 0 such that

(1.42) D A ε u ε L p ( M ε ) C p

for any ε > 0 . By (1.42) and (1.28), we deduce that D A ε u ε is bounded in L p ( M ) . This contradicts (1.40) and proves (1.38).

Proof of (1.39). By Lemma 1.10, (1.28), and the fact that | u ε | 1 - 1 | log ε | 2 in N ε , it follows that

(1.43) M | j * | p vol g lim inf ε 0 M ε | j ( u ε , A ε ) | p | u ε | p vol g lim sup ε 0 M ε | j ( u ε , A ε ) | p | u ε | p vol g lim sup ε 0 ( 1 - 1 | log ε | 2 ) - p M ε | j ( u ε , A ε ) | p vol g lim ε 0 M | j ( u ε , A ε ) | p vol g = M | j * | p vol g .

Thus, by (1.1), (1.19), (1.36), and (1.43),

(1.44) j * L p ( M ) = lim ε 0 ( M ε | j ( u ε , A ε ) | p | u ε | p vol g ) 1 p ( 1.1 ) lim ε 0 ( M ε | D A ε u ε | p vol g ) 1 p lim ε 0 { ( M ε | d ρ ε | p vol g ) 1 p + ( 1 - 1 | log ε | 2 ) - 1 ( M ε | j ( u ε , A ε ) | p vol g ) 1 p } = ( 1.19 ) lim ε 0 ( M ε | j ( u ε , A ε ) | p vol g ) 1 p = ( 1.36 ) j * L p ( M )

for any p with 1 p < n n - 1 . But, by (1.28),

lim ε 0 ( M ε | D A ε u ε | p vol g ) 1 p = lim ε 0 ( M | D A ε u ε | p vol g ) 1 p

for any p with 1 p < n n - 1 , hence (1.39) follows by (1.43) and (1.44). ∎

Finally, the following lemma is classical in Ginzburg–Landau theory.

Lemma 1.12.

Let { ( u ε , A ε ) } E be a sequence satisfying the logarithmic bound (9). Then

(1.45) lim ε 0 M ( 1 - | u ε | 2 ) 2 vol g = 0 .

In particular, | u ε | 1 in L p ( M ) as ε 0 , for any finite p 1 .

Proof.

Equation (1.45) is an immediate consequence of (9). The L p ( M ) -convergence | u ε | 1 follows by (9) and (1.3), by interpolation. ∎

2 Compactness up to gauge equivalence

In this section, we prove Theorem 2. We assume that { ( u ε , A ε ) } is a sequence of critical points of G ε satisfying (9) and that each A ε takes the form (11), i.e., that each A ε writes as

A ε = d * ψ ε + ζ ε

for an exact 2-form ψ ε and a harmonic form ζ ε satisfying ζ ε L ( M ) C M , where C M depends only on M. As already mentioned in the Introduction, this is always possible. Indeed, a family of gauge transformations { Φ ε } W 1 , 2 ( M , 𝕊 1 ) with this property can be obtained by applying, for each ε > 0 , [13, Lemma 2.10]. We now recall that, by [13, Remark 3.12], we may apply [13, Proposition 3.11] to any sequence of critical pairs ( u ε , A ε ) satisfying the logarithmic energy bound (9). Then, as shown by (the proof of) Proposition 3.11 in [13], if A ε takes the form (11), we can find A * W 2 , p ( M , T * M ) for any p with 1 p < n n - 1 such that, possibly up to extraction of a (not relabelled) subsequence,

(2.1) Φ ε A ε A * strongly in  W 2 , p ( M ) as  ε 0 for any  p  with  1 p < n n - 1 .

Thus, to prove Theorem 2, it is enough to prove that there exists u * belonging to W 1 , p ( M , E ) for any p with 1 p < n n - 1 such that, up to a subsequence,

(2.2) Φ ε u ε u * strongly in  W 1 , p ( M ) as  ε 0 for any  p  with  1 p < n n - 1 .

In (2.1) and (2.2), Φ ε u ε , Φ ε A ε are defined as in (3); i.e., Φ ε u ε is given by fibre-wise action of 𝕊 1 on E and Φ ε A ε = A ε - i Φ ε - 1 d Φ ε . Since in the rest of this section we will always work in the gauge (11), we will omit the gauge transformations Φ ε from the notation; i.e., while writing A ε and u ε , we will actually mean Φ ε A ε and Φ ε u ε , respectively.

Remark 2.1.

In [13], we worked with gauge transformations of class W 2 , 2 ( M , 𝕊 1 ) . Routine arguments show that, in fact, one can enlarge the class of gauge transformations to the whole of W 1 , 2 ( M , 𝕊 1 ) . (In other words, if ( u ε , A ε ) , then its energy stays invariant under the action of Φ specified in (3), for any Φ W 1 , 2 ( M , 𝕊 1 ) .) A comparison with the proof of [13, Proposition 3.11] mentioned above shows that the W 2 , 2 ( M , 𝕊 1 ) -regularity of the gauge transformations Φ ε involved there is not really used. Indeed, the conclusion follows by (i) the L 2 -Hodge decomposition of A ε ; (ii) the fact that Φ ε puts A ε in Coulomb form (i.e., Φ ε A ε = d * ψ ε + ζ ε ); (iii) the fact that

d ( Φ ε 1 A ε 2 - Φ ε 2 A ε 2 ) = - Δ ( ψ ε 1 - ψ ε 2 ) = F ε 1 - F ε 2

for any ε 1 , ε 2 > 0 ; (iv) elliptic estimates for F ε in W 1 , p ( M , Λ 2 T * M ) and, in turn, W 3 , p -estimates for ψ ε , entailed by the fact that F ε satisfies the London equation (8) and implying that { F ε } , { ψ ε } are both Cauchy sequences. None of these ingredients require W 2 , 2 -regularity, but only the W 1 , 2 -regularity of Φ ε .

The following corollary is a straightforward consequence of Lemma 1.11.

Corollary 2.2.

Let { ( u ε , A ε ) } E be a sequence of critical points of G ε satisfying the logarithmic energy bound (9). Assume, in addition, that (11) holds. Then there exists a section u * W 1 , p ( M , E ) for any p with 1 p < n n - 1 satisfying | u * | = 1 almost everywhere on M such that, up to extraction of a (not relabelled) subsequence, for any p with 1 p < n n - 1 there holds

(2.3) u ε u * weakly in  W 1 , p ( M ) ,
(2.4) u ε u * strongly in  L p ( M ) ,
(2.5) | u ε | 1 strongly in  W 1 , p ( M ) ,

as ε 0 .

Proof.

By Lemma 1.11, (2.1), and the uniform bound u ε L ( M ) 1 (given by Proposition 1.2), it follows that the sequence { D 0 u ε } is uniformly bounded in L p ( M , T * M E ) , for any p with 1 p < n n - 1 . Consequently, the sequence { u ε } is uniformly bounded in W 1 , p ( M , E ) , for any p with 1 p < n n - 1 . Hence, there exists a section u * W 1 , p ( M , E ) for any p with 1 p < n n - 1 such that up to a (not relabelled) subsequence

u ε u * weakly in  W 1 , p ( M ) for any  p  with  1 p < n n - 1 as  ε 0 ,

i.e., (2.3). Moreover, since M is compact, by the Rellich–Kondrachov theorem it follows that, up to another (not relabelled) subsequence,

u ε u * strongly in  L p ( M ) for any  p  with  1 p < n n - 1 as  ε 0 ,

i.e., (2.4), and so, up to still another (not relabelled) subsequence, | u ε | | u * | pointwise almost everywhere on M as ε 0 . Then Lemma 1.12 implies

(2.6) | u * | = 1 a.e. on  M .

By (2.6), d ( | u * | ) = 0 almost everywhere on M. Therefore,

d ( | u * | ) - d ( | u ε | ) L p ( M ) = d ( | u ε | ) L p ( M ) ( 1.19 ) 0 as  ε 0 ,

for any p with 1 p < n n - 1 . Along with (2.4), this proves (2.5), and the corollary follows. ∎

Remark 2.3.

By [13, Proposition 3.11 and Remark 3.12], the uniform bound (1.3), Corollary 2.2, and the dominated convergence theorem, it follows that

(2.7) A ε u ε A * u * strongly in  L p ( M ) as  ε 0 for any  p  with  1 p < n n - 1 .

Indeed, for any p with 1 p < n n - 1 , we have

A ε u ε - A * u * L p ( M ) = ( A ε - A * ) u ε - A * ( u * - u ε ) L p ( M ) ( 1.3 ) A ε - A * L p ( M ) + | A * | | u * - u ε | L p ( M ) .

The first term at the right-hand side vanishes as ε 0 , because A ε A * in W 2 , p ( M ) , for any p with 1 p < n n - 1 . On the other hand, since | A * | | u * - u ε | 2 | A * | (again by (1.3)) and u ε u * pointwise a.e. in M (by (2.4)), the dominated convergence theorem yields that | A * | | u * - u ε | L p ( M ) 0 as ε 0 , proving (2.7).

We are now in a position to prove Theorem 2.

Proof of Theorem 2.

Since A ε satisfies (11), equations (2.3) and (2.7) imply

(2.8) D A ε u ε D A * u * weakly in  L p ( M ) for any  p  with  1 p < n n - 1 as  ε 0 .

We claim that

(2.9) M | D A ε u ε | p vol g M | D A * u * | p vol g for any  p  with  1 p < n n - 1 , as  ε 0 ,

Condition (2.9) implies the strong convergence D A ε u ε D A * u * in L p ( M ) for any 1 p < n n - 1 , and, in view of (2.7), that D 0 u ε D 0 u * strongly in L p ( M ) , for any 1 p < n n - 1 .

Towards the proof of (2.9), we observe that, by Lemma 1.1 and since | u * | = 1 a.e. on M (by Corollary 2.2), there holds

D A * u * = i j ( u * , A * ) u * a.e. on  M ,

where j ( u * , A * ) = D A * u * , i u * . On the other hand, equations (2.8), (2.4), and the uniform bound D A ε u ε L p ( M ) C p given, for any p with 1 p < n n - 1 , by Lemma 1.11, yield

(2.10) j ( u ε , A ε ) = D A ε u ε , i u ε D A * u * , i u * = j ( u * , A * )

weakly in L p ( M ) for any p with 1 p < n n - 1 as ε 0 . By the uniqueness of the weak limit, we must have j ( u * , A * ) = j * , where j * is given by Lemma 1.10. Recalling (1.39), we deduce (2.9).

By passing to the limit in the Euler–Lagrange equation (7), with the help of (1.30) and (2.10), we see that

(2.11) j ( u * , A * ) = d * F *

in the sense of distributions in M – that is, equation (12) holds. In order to complete the proof, it only remains to show that u * and A * are smooth in M spt μ * , where μ * is the limit of the rescaled energy densities, given by Theorem 1. By passing to the limit as ε 0 in (11) (with the help of [34, Proposition 5.6] or, say, [13, Proposition A.13]), we see that A * can be written in the form

(2.12) A * = d * ψ * + ζ * ,

where ψ * is an exact 2-form and ζ * a harmonic 1-form on M. By taking the differential of both sides, we obtain

- Δ ψ * = d d * ψ * = d A * = F * - F 0 ,

where F 0 is the curvature of the reference connection D 0 . Now, F * is a solution of the London equation (1.31), whose right-hand side 2 π J * is a measure concentrated on spt μ * , due to [13, Theorem A]. By elliptic regularity, it follows that F * and ψ * are smooth in M spt μ * and hence, A * is. Now, let B M spt μ * be a ball. We claim that there holds

(2.13) B | D A * u * | 2 vol g < +

Equation (2.13) is an almost immediate consequence of Proposition 4.1, which we will prove later on in Section 4, combined with a covering argument. As A * is smooth in a neighbourhood of B ¯ , from (2.13) we deduce that u * W 1 , 2 ( B , E ) . Therefore, up to identifying u * | B with a complex-valued map, we can write u * = e i θ * in B for some scalar function θ * W 1 , 2 ( B , ) . (This claim follows by lifting results that were originally proven in [11]; see also [12, Theorem 1.1].) The reference connection can locally be written as D 0 = d - i γ 0 on B, for some (smooth) 1-form γ 0 . Then a direct computation shows that

j ( u * , A * ) = d θ * - γ 0 - A * .

However, j ( u * , A * ) is smooth in B, due to (2.11). Therefore, d θ * is smooth in B, and hence u * = e i θ * is. This completes the proof of the theorem. ∎

Remark 2.4.

Let φ C c ( M spt μ * ) be a test function. From (2.11), we know that d * j ( u * , A * ) = 0 and hence, by integrating by parts,

0 = M j ( u * , A * ) , d φ vol g = M D A * u * , d φ i u * vol g = M D A * u * , D A * ( i φ u * ) vol g = M D A * * D A * u * , i u * φ vol g

where D A * * is the L 2 -adjoint of D A * . As φ is arbitrary, it follows that D A * * D A * u * = λ u * in M spt μ * , for some (real-valued) function λ : M spt μ * . Since | u * | = 1 , we obtain that λ = D A * * D A * u * , u * and hence, by Lemma 1.4,

D A * * D A * u * = | D A * u * | 2 u * in  M spt μ * .

In other words, u * is an A * -harmonic unit section away from the singular set spt μ * .

3 Monotonicity formula

This section is devoted to the proof of the monotonicity formula, Theorem 3. For the reader’s convenience, we first recall some notation. Let ( u ε , A ε ) be a critical point of G ε , fixed once and for all throughout this section. For x 0 M and r ( 0 , inj ( M ) ) , we define

(3.1) E ε ( x 0 , ρ ) := ρ ( x 0 ) e ε ( u ε , A ε ) vol g .

Moreover, if ν is the exterior unit normal field to ρ ( x 0 ) , we set

(3.2) X ε ( x 0 , ρ ) := ρ ( x 0 ) ( | D A ε , ν u ε | 2 + | i ν F ε | 2 ) vol g ^ + 1 2 ε 2 ρ ρ ( x 0 ) ( 1 - | u ε | 2 ) 2 vol g ,

where F ε := F A ε , i ν denotes the contraction against ν (as defined in (18)) and g ^ is the metric induced by g on ρ ( x 0 ) .

We will achieve the proof of Theorem 3 through a series of lemmas. The first one is a well-known consequence of the inner variation equation for critical points (see, e.g., [29, Section 4]).

Lemma 3.1.

Let ( u ε , A ε ) E be any critical point of G ε . For any x 0 M and any ρ ( 0 , inj ( M ) ) , there holds

(3.3) ρ ρ ( x 0 ) e ε ( u ε , A ε ) vol g ^ ρ X ε ( x 0 , ρ ) + ( n - 2 - C ( M ) ρ 2 ) E ε ( x 0 , ρ ) - ρ ( x 0 ) | F ε | 2 vol g ,

where C ( M ) > 0 is a constant depending only on M and g.

Lemma 3.1 follows by the very same computations contained in, e.g., [29] (cf. the computations starting at the beginning of Section 4 and ending right above equation (4.6) there). Therefore, we skip its proof.

The second lemma provides an estimate on the term ρ ( x 0 ) | F ε | 2 vol g in (3.3).

Lemma 3.2.

There exist ρ ¯ ( 0 , inj ( M ) ) and a constant λ > 0 , both depending only on M and g, such that the following holds. If ( u ε , A ε ) E is any critical point of G ε , then for any x 0 M and any ρ ( 0 , ρ ¯ ) there holds

(3.4) F ε L 2 ( ρ ( x 0 ) ) K ρ D A ε u ε L 2 ( ρ ( x 0 ) ) + λ ρ i ν F ε L 2 ( ρ ( x 0 ) ) ,

where K > 0 is a constant depending only on ( M , g ) .

Proof.

The idea is to exploit both the Euler–Lagrange equations and gauge transformations to rewrite the integral ρ ( x 0 ) | F ε | 2 vol g in terms of quantities that can be estimated more easily. To this purpose, we proceed step by step.

Step 1.

Let ( u ε , A ε ) be any critical point of G ε . Then the curvature F ε := F A ε satisfies the Euler–Lagrange equation

(3.5) d * F ε = j ( u ε , A ε )

in the sense of distributions on M, and hence in particular on ρ ( x 0 ) , for any fixed x 0 M . Recall that, by definition, D A ε = D 0 - i A ε , and that locally we may write D 0 = d - i γ 0 , where γ 0 is a 1-form. Thus, on the geodesic ball ρ ( x 0 ) we can write D A ε = d - i A ~ ε , where A ~ ε := A ε + γ 0 , and it holds d A ~ ε = F ε . From now on, we shall work locally and, for convenience, we will drop the tilde, writing A ε to mean, actually, A ~ ε .

We fix Coulomb gauge in ρ ( x 0 ) (without changing notation), so that A ε satisfies

(3.6) { d * A ε = 0 in  ρ ( x 0 ) , i ν A ε = 0 on  ρ ( x 0 ) , d A ε = F ε .

(Such a gauge exists: see, e.g., [28, Proposition 2.1].) Moreover, A ε is smooth up to the boundary of ρ ( x 0 ) , and hence so is F ε (recall that F ε is gauge-invariant, so it is actually globally smooth).

Step 2.

Using (3.6), we write

| F ε | 2 vol g = F ε , F ε = F ε , d A ε .

By the well-known integration-by-parts formula for differential forms (see, for instance, [21, (2.22) and (2.34)]), we obtain

(3.7) ρ ( x 0 ) | F ε | 2 vol g = ρ ( x 0 ) j ( u ε , A ε ) , A ε + ρ ( x 0 ) A ε ( F ε ) T ,

where ( F ε ) T denotes the part of F ε tangent to ρ ( x 0 ) . We are going to estimate the terms at the right-hand side of (3.7). Before doing this, we notice that the Poincaré and Gaffney inequalities yield the estimate

(3.8) A ε L 2 ( ρ ( x 0 ) ) ρ F ε L 2 ( ρ ( x 0 ) ) ,

up to constant depending only on M and g (see Proposition A.1 in the appendix for more details).

Step 3.

As it is easy to check, the following identity holds pointwise

(3.9) ( F ε ) T = ( i ν F ε ν )

on the boundary of any geodesic ball and, by (3.9), we obtain

(3.10) ρ ( x 0 ) A ε ( F ε ) T = ρ ( x 0 ) A ε , ( i ν F ε ) ν A ε L 2 ( ρ ( x 0 ) ) i ν F ε L 2 ( ρ ( x 0 ) ) .

On the other hand, the trace inequality and Gaffney’s inequality imply that there exists a constant λ > 0 , depending only on ( M , g ) , such that

(3.11) A ε L 2 ( ρ ( x 0 ) ) λ ρ F ε L 2 ( ρ ( x 0 ) )

(see (A.3) in the appendix for details). Recalling the easy estimate

(3.12) j ( u ε , A ε ) L 2 ( ρ ( x 0 ) ) D A ε u ε L 2 ( ρ ( x 0 ) ) ,

and combining (3.7), (3.10), (3.11), and (3.12) with (3.7), we obtain

(3.13) F ε L 2 ( ρ ( x 0 ) ) 2 ( K ρ D A ε u ε L 2 ( ρ ( x 0 ) ) + λ ρ i ν F ε L 2 ( B ρ ( x 0 ) ) ) F ε L 2 ( ρ ( x 0 ) ) .

As (3.4) is obvious if F ε L 2 ( ρ ( x 0 ) ) = 0 , we may assume F ε L 2 ( ρ ( x 0 ) ) > 0 . Then dividing both sides of (3.13) by F ε L 2 ( ρ ( x 0 ) ) yields the conclusion. ∎

Proposition 3.3.

There exist a number α ( 0 , 2 ] , a number ρ ¯ ( 0 , inj ( M ) ) and a constant C > 0 , depending only on M and g, such that the following holds. If ( u ε , A ε ) E is any critical point of G ε , then for any x 0 in M and any ρ ( 0 , ρ ¯ ) , there holds

(3.14) d d ρ ( e C ρ 2 E ε ( x 0 , ρ ) ρ n - 4 + α ) > 0 ,

where E ε ( x 0 , ρ ) is defined by (3.1).

Proof.

Assume, for the moment being, α is any (fixed) nonneg ative number. Using (3.3), we compute

d d ρ ( e C ρ 2 E ε ( x 0 , ρ ) ρ n - 4 + α ) = e C ρ 2 ρ n - 4 + α [ ( 2 C ρ - n - 4 + α ρ ) E ε ( x 0 , ρ ) + ρ ( x 0 ) e ε ( u ε , A ε ) ] ( 3.3 ) e C ρ 2 ρ n - 4 + α { ( 2 C - C ( M ) ) ρ E ε ( x 0 , ρ ) + X ε ( x 0 , ρ ) + 2 - α ρ ρ ( x 0 ) ( 1 2 | D A ε u ε | 2 + ( 1 - | u ε | 2 ) 2 4 ε 2 ) vol g - α 2 ρ ρ ( x 0 ) | F ε | 2 vol g } ,

for any ρ ( 0 , inj ( M ) ) . Then, by (3.4) and Young’s inequality we obtain

d d ρ ( e C ρ 2 ρ n - 4 + α B ρ ( x 0 ) e ε ( u ε , A ε ) vol g ) e C ρ 2 ρ n - 4 + α { ( 2 C - C ( M ) ) ρ E ε ( x 0 , ρ ) + X ε ( ρ , x 0 ) + 2 - α ρ ρ ( x 0 ) 1 2 | D A ε u ε | 2 + ( 1 - | u ε | 2 ) 2 4 ε 2 vol g - α K 2 ρ ρ ( x 0 ) | D A ε u ε | 2 vol g - α λ 2 ρ ( x 0 ) | i ν F ε | 2 vol g ^ }

for any ρ ( 0 , ρ ¯ ) , with ρ ¯ as in Lemma 3.2, and where K and λ are the same constants as in Lemma 3.2 (both of which depend only on M and g). Thus, recalling the definition (3.2) of X ε ( x 0 , ρ ) and taking α = min { 2 , 1 2 λ 2 , C ( M ) 2 K 2 } and (for instance) C = 2 C ( M ) , the proposition follows. ∎

With Proposition 3.3 at hand, we can prove a decay estimate for the L 2 -norm of the curvatures F ε in sufficiently small geodesic balls. In turn, such decay property is the key point to obtain the ( n - 2 ) -monotonicity formula (16).

Lemma 3.4.

There exist a number α ( 0 , 2 ) and a number R 0 ( 0 , inj ( M ) ) , both depending only on M and g, such that the following holds. If ( u ε , A ε ) E is any critical point of G ε , then for any x 0 M and any 0 < r < R < R 0 , there holds

(3.15) r ( x 0 ) | F ε | 2 vol g r n - 2 + α R n - 2 + α R ( x 0 ) e ε ( u ε , A ε ) vol g ,

where the implicit constant on the right-hand-side depends only on ( M , g ) and α.

Proof.

We start by fixing x 0 M and considering two geodesic balls r := r ( x 0 ) and s := s ( x 0 ) , with 0 < r < s < R and R ( 0 , inj ( M ) ) small enough (to be further specified later). Up to subtracting a (smooth, ε-independent) 1-form to A ε , we can write

d A ε = F ε in  R .

Moreover, we can fix Coulomb gauge (without changing notation), i.e., we can assume that

d * A ε = 0 in  R , i ν A ε = 0 on  R ,

where ν is the unit normal field to R . Then the Euler–Lagrange equation (7) becomes

- Δ A ε = j ( u ε , A ε ) in  R .

Standard decay estimates for elliptic equations with a right-hand side (see, e.g., Proposition B.4 for more details) imply that there exists R * > 0 (depending only on M and g) so that, for 0 < r < s < R < R * , there holds

r ( | d A ε | 2 + | d * A ε | 2 ) vol g r n s n s ( | d A ε | 2 + | d * A ε | 2 ) vol g + s 2 s | j ( u ε , D A ε ) | 2 vol g ,

up to a constant depending only on M and g. Keeping in mind that d * A ε = 0 and d A ε = F ε in R , and recalling the pointwise estimate | j ( u ε , A ε ) | | D A ε u ε | , it follows

(3.16) r | F ε | 2 vol g r n s n s | F ε | 2 vol g + s 2 s | D A ε u ε | 2 vol g ,

up to a constant depending only on M and g.

Taking now R 0 = min { R * , ρ ¯ } , where ρ ¯ is the distinguished radius (depending only on M and g) given by Proposition 3.3 and shrinking R so that 0 < r < s < R < R 0 , by (3.16) and (3.14) we obtain

(3.17) r | F ε | 2 vol g r n s n s | F ε | 2 vol g + s 2 s e ε ( u ε , A ε ) vol g r n s n s | F ε | 2 vol g + s n - 2 + α R n - 4 + α R e ε ( u ε , A ε ) vol g

for α ( 0 , 2 ] as given by Proposition 3.3.

Assume α ( 0 , 2 ) . Then, by a classical iteration lemma (e.g., [20, Lemma III.2.1] or [6, Lemma B.3]), it follows

(3.18) r | F ε | 2 vol g r n - 2 + α s n - 2 + α s | F ε | 2 vol g + r n - 2 + α R n - 4 + α R e ε ( u ε , A ε ) vol g r n - 2 + α s n - 2 + α s | F ε | 2 vol g + r n - 2 + α R n - 2 + α R e ε ( u ε , A ε ) vol g

for any 0 < r < s < R < R 0 . (The second inequality holds because R is certainly bounded from above.) In particular, by taking the limit s R , we get the desired decay estimate (3.15). ∎

Remark 3.5.

The restriction α < 2 is necessary in the argument from (3.17) to (3.18), as otherwise we could not use [20, Lemma III.2.1] or [6, Lemma B.3]. This inconvenience will produce the “error term”, i.e., the term depending on α, in (16). Nevertheless, the monotonicity formula (16) is enough for our purposes.

We are now ready to prove Theorem 3.

Proof of Theorem 3.

Denoting C the maximum between the constant C ( M ) in (3.3) and the constant given by Lemma 3.4, thanks to Lemma 3.4, inequality (3.3) yields

(3.19) ρ E ε ( x 0 , ρ ) ρ X ε ( x 0 , ρ ) + ( n - 2 ) E ε ( x 0 , ρ ) - C ρ 2 E ε ( x 0 , ρ ) - C ρ n - 2 + α R n - 2 + α E ε ( x 0 , R )

for any 0 < ρ < R < R 0 , where R 0 is the distinguished radius in ( 0 , inj ( M ) ) given by Lemma 3.4, any x 0 M . Note that the constant C depends only on M, g.

Dividing both sides of (3.19) by ρ n - 1 , we obtain

(3.20) ρ 2 - n E ε ( x 0 , ρ ) - ( n - 2 ) ρ 1 - n E ε ( x 0 , ρ ) + C ρ 3 - n E ε ( x 0 , ρ ) ρ 2 - n X ε ( x 0 , ρ ) - C ρ α - 1 R n - 2 + α E ε ( x 0 , R ) .

Multiplying (3.20) by exp ( C ρ 2 / 2 ) , where C is exactly the same constant as in (3.20), we get

(3.21) d d ρ { e C ρ 2 2 ρ 2 - n E ε ( x 0 , ρ ) } e C ρ 2 2 ρ 2 - n X ε ( x 0 , ρ ) - C e C ρ 2 2 ρ α - 1 R n - 2 + α E ε ( x 0 , R ) ,

and the conclusion follows by integrating both sides of (3.21) over ρ ( r , R ) , setting C := C 2 , and recalling that α depends only on ( M , g ) . ∎

Remark 3.6.

Multiplying (3.21) by ρ and recalling the definition (3.2) of X ε , we obtain the following variant of (3.21): for any 0 < ρ < R < R 0 ,

(3.22) ρ d d ρ ( e C ρ 2 E ( x 0 , ρ ) ρ n - 2 ) 1 2 ε 2 ρ n - 2 ρ ( 1 - | u ε | 2 ) 2 vol g - C ρ α E ( x 0 , R ) R n - 2 + α ,

where α ( 0 , 2 ) , C, C are absolute constants that depend only on M and g. We will use (3.22) in the proof of the clearing-out, in Section 4.

Remark 3.7.

All the arguments used in the proof of Theorem 3 are completely independent of Lemma 1.6. In fact, we will use Theorem 3 to prove Lemma 1.6.

As a byproduct of our arguments, we obtain the following bound on the curvature in terms of the rescaled energy:

Proposition 3.8.

For any q ( 0 , n ) there exists a constant C q > 0 , depending only on ( M , g ) and q, such that the following statement holds. If ( u ε , A ε ) E is any critical point of G ε , then

(3.23) r ( x 0 ) | F ε | 2 vol g C q r q R n - 2 E ε ( x 0 , R )

holds for any x 0 M and 0 < r < R .

Proof.

By a covering argument, there is no loss of generality in assuming that R < R 0 , where R 0 is a uniform constant. For instance, we can take R 0 exactly equal to the constant given by Lemma 3.4. Let α ( 0 , 2 ] be given by Proposition 3.3, and let 0 < r < s < R < R * . The estimate (3.17), obtained in the proof of Lemma 3.4, and the (almost) monotonicity formula (16) imply

r | F ε | 2 vol g r n s n s | F ε | 2 vol g + s 2 s e ε ( u ε , A ε ) vol g r n s n s | F ε | 2 vol g + s n R n - 2 R e ε ( u ε , A ε ) vol g ,

The proposition follows by an iteration argument, based on [20, Lemma III.2.1]. ∎

Finally, we conclude this section by stating and proving a corollary of the monotonicity formula, which will be useful in Section 4.

Corollary 3.9.

Let ε > 0 and let ( u ε , A ε ) E be any critical point of G ε . Moreover, let α ( 0 , 2 ) , R 0 ( 0 , inj ( M ) ) and C > 0 be the numbers (depending only on ( M , g ) ) given by Theorem 3. Then, for any x 0 M and any R ( 0 , R 0 ) , there holds

(3.24) 1 ε 2 B R ( x 0 ) ( 1 - | u ε ( y ) | 2 ) 2 dist n - 2 ( y , x 0 ) vol g ( y ) R 2 - n E ε ( x 0 , R ) ,

where the implicit constant in front of the right-hand side depends only on ( M , g ) .

Proof.

By Proposition 1.2, Lemma 1.6 and Lemma 1.7, the energy density is a bounded function (and its L ( M ) -norm is of order O ( ε - 2 ) ). Consequently, there holds

(3.25) lim r 0 r 2 - n E ε ( x 0 , r ) = 0 .

Let us set

W ε ( x 0 , ρ ) := ρ ( x 0 ) ( 1 - | u ε ( y ) | 2 ) 2 2 ε 2 vol g ( y ) .

Since the function ( 0 , R ) ρ W ε ( x 0 , ρ ) is absolutely continuous, for a.e. ρ ( 0 , R ) we have

W ε ( x 0 , ρ ) = ρ ( x 0 ) ( 1 - | u ε ( y ) | 2 ) 2 2 ε 2 vol g ^ ( y )

as well as

ρ 1 - n W ε ( x 0 , ρ ) = 1 n - 2 { - d d ρ ( ρ 2 - n W ε ( x 0 , ρ ) ) + ρ 2 - n W ε ( x 0 , ρ ) } .

By (3.21) and (3.2), we clearly obtain

d d ρ { e C ρ 2 2 ρ 2 - n E ε ( x 0 , ρ ) } ρ 1 - n ρ ( x 0 ) ( 1 - | u ε ( y ) | 2 ) 2 2 ε 2 vol g ( y ) - C e C ρ 2 2 ρ α - 1 R n - 2 + α E ε ( x 0 , R ) .

Hence, by combining the above relations, we find that

(3.26) d d ρ { e C ρ 2 2 ρ 2 - n E ε ( x 0 , ρ ) + 1 n - 2 ρ 2 - n W ε ( x 0 , ρ ) } ρ 2 - n n - 2 W ε ( x 0 , ρ ) - C e C ρ 2 2 ρ α - 1 R n - 2 + α E ε ( x 0 , R )

for a.e. ρ ( 0 , R ) . Integrating both sides of (3.26) over ( 0 , R ) , and taking (3.25) into account, yields

1 n - 2 R ( x 0 ) 1 2 ε 2 ( 1 - | u ε ( y ) | 2 ) 2 dist ( x 0 , y ) n - 2 vol g ( y ) ( 1 + C α ) e C R 2 2 R 2 - n E ε ( x 0 , R ) + R 2 - n n - 2 R ( x 0 ) ( 1 - | u ε ( y ) | 2 ) 2 2 ε 2 vol g ( y ) ( n n - 2 + C α ) R 2 - n E ε ( x 0 , R ) .

Recalling that α depends only on ( M , g ) , the conclusion follows. ∎

4 Clearing-out and its consequences

The aim of this section is to prove Proposition 4.1 below, which will be essential in characterising the support of the energy-concentration measure μ * , defined in (5.1) (see Section 5 and Lemma 5.6 in particular). As in the previous section, given a finite-energy pair ( u ε , A ε ) , a point x 0 M and a radius ρ > 0 , we write

E ε ( x 0 , ρ ) := G ε ( u ε , A ε ; ρ ( x 0 ) ) .

Proposition 4.1.

There exists constants η 0 > 0 , R * ( 0 , inj ( M ) ) and ε * > 0 , depending on ( M , g ) only, such that the following statement holds. Let x 0 M , R ( 0 , R * ) and let { ( u ε , A ε ) } E be a sequence of critical points of G ε that satisfy (9) and

(4.1) E ε ( x 0 , R ) η 0 R n - 2 log R ε .

Then, for any ε < ε * , there holds

(4.2) | u ε ( x ) | 1 2 for any  x 3 R / 4 ( x 0 ) .

Moreover, there holds

(4.3) sup ε > 0 E ε ( x 0 , R 2 ) < + .

Proposition 4.1 is the counterpart of [8, Proposition VII.1]. For its proof, we shall follow the same strategy as in [8], involving essentially three ingredients: energy decay estimates, a clearing-out property, and elliptic estimates.

The pivotal point of the argument is the “clearing-out” property, Proposition 4.6. Indeed, this property implies that equation (7) is uniformly elliptic in any geodesic ball in which (4.1) holds. In turn, uniform ellipticity implies several strong elliptic estimates that allow to achieve the conclusion of Proposition 4.1.

According to [8], the main concern towards the proof of the clearing-out property lies in obtaining a suitable, quantitative decay of the rescaled energy in small balls with respect to the radius. This is done in Section 4.1.

Throughout Section 4, we will consider a sequence of critical points { ( u ε , A ε ) } , a point x 0 M and a radius R > 0 (smaller than the injectivity radius of M). We will write indifferently r or r ( x 0 ) for a geodesic ball of center x 0 and radius r > 0 .

4.1 An energy decay estimate

Towards the proof of Proposition 4.1, the first step is an energy decay estimate, analogous to [8, Theorem 3].

Proposition 4.2.

There exist numbers R * > 0 and α ( 0 , 2 ) , depending on ( M , g ) only, such that the following statement holds. Let x 0 M , 0 < ε < R < R * and 0 < δ < 1 10 . Let ( u ε , A ε ) E be a critical point of G ε . Define

(4.4) p ε := 1 ε 2 R n - 2 R ( x 0 ) ( 1 - | u ε | 2 ) 2 vol g

Then there holds

E ε ( x 0 , δ R ) R n - 2 ( p ε 1 3 + δ n - 2 + α ) E ε ( x 0 , R ) R n - 2 + p ε 2 3 ,

where the implicit constant on the right-hand side depends only on ( M , g ) .

The rest of the section is devoted to the proof of Proposition 4.2. Let x 0 M , 0 < R < inj ( M ) , ε > 0 and 0 < δ < 1 10 be fixed, and let ( u ε , A ε ) be a critical point of G ε . First of all, we choose a radius

(4.5) r ( R 10 , R 5 )

such that

(4.6) r | D A ε u ε | 2 d n - 1 40 R R | D A ε u ε | 2 vol g ,
(4.7) r ( | u ε | 2 - 1 ) 2 d n - 1 40 R R ( | u ε | 2 - 1 ) 2 vol g .

Such a radius exist, because the set of radii r ( R 10 , R 5 ) that do not satisfy either of inequalities (4.6) has length R 30 at most. For such a choice of r, we have r > δ R . Let η r be the indicator function of the ball r (i.e., η r ( x ) := 1 if x r and η r ( x ) := 0 otherwise). By Hodge decomposition, there exist φ ε W 1 , 2 ( M ) , ψ ε W 1 , 2 ( M , Λ 2 T * M ) and ξ ε Harm 1 ( M ) such that

(4.8) η r j ( u ε , A ε ) = d φ ε + d * ψ ε + ξ ε ,

where the three terms on the right-hand side are orthogonal in L 2 ( M ) . Moreover, there is no loss of generality in assuming that ψ ε is exact (if not, we replace ψ ε by its projection onto exact forms). We prove decay estimates on φ ε , ψ ε and ξ ε separately.

Lemma 4.3.

We have

δ R ( | d φ ε | 2 + | ξ ε | 2 ) vol g δ n E ε ( x 0 , R ) .

Proof.

We consider φ ε first. In the interior of r , we have

- Δ φ ε = d * d φ ε = d * j ( u ε , A ε ) = d * d * F ε = 0 .

By differentiating both sides of - Δ φ ε = 0 , we deduce that - Δ ( d φ ε ) = 0 in the interior of R . Then Theorem B.3 implies

δ R | d φ ε | 2 vol g δ n R | d φ ε | 2 vol g δ n M | d φ ε | 2 vol g .

As the decomposition (4.8) is orthogonal in L 2 ( M ) , we deduce

δ R | d φ ε | 2 vol g δ n R | j ( u ε , A ε ) | 2 vol g δ n E ε ( x 0 , R ) ,

where we have used that | j ( u ε , A ε ) | | D A ε u ε | for the last inequality. An analogous argument applies to the harmonic form ξ ε . ∎

The proof of the decay estimate for ψ ε is more technical. It is convenient to further decompose ψ ε as a sum of several contributions, as in [8, Proof of Theorem 3, Step 2]. Let β ( 0 , 1 4 ) be a parameter, to be chosen later. Let f : [ 0 , + ) [ 0 , + ) be a smooth function such that

(4.9) f ( t ) = 1 if  0 t 1 - 2 β ,    f ( t ) = 1 t if  t 1 - β ,
(4.10) 1 f ( t ) 1 t if  1 - 2 β < t < 1 - β ,    | f ( t ) | 4 for any  t 0 .

We define

(4.11) ρ ε := f ( | u ε | ) , v ε := ρ ε u ε

By construction, the function ρ ε satisfies

(4.12) 0 ρ ε 2 - 1 4 β in  M .

Moreover, writing the reference connection as D 0 = d - i γ 0 for some (locally defined) 1-form γ 0 , we have

(4.13) j ( v ε , A ε ) = d ρ ε u ε + ρ ε d u ε - i ρ ε ( γ 0 + A ε ) u ε , i ρ ε u ε = ρ ε 2 j ( u ε , A ε ) .

Now, we consider the Hodge decomposition (4.8). As we have seen above, we can assume with no loss of generality that ψ ε is exact. Then, by differentiating both sides of (4.8), we obtain

- Δ ψ ε = d d * ψ ε = d ( η r j ( u ε , A ε ) ) = ( 4.13 ) d ( η r j ( v ε , A ε ) ) + d ( η r ( 1 - ρ ε 2 ) j ( u ε , A ε ) ) = η r d j ( v ε , A ε ) - σ r ν j ( v ε , A ε ) + d ( η r ( 1 - ρ ε 2 ) j ( u ε , A ε ) ) ,

where σ r denotes the ( n - 1 ) -dimensional Hausdorff measure on r and ν is the outward unit vector to r . Recalling the definition of Jacobian, equation (4), we deduce

- Δ ψ ε = 2 η r J ( v ε , A ε ) - η r F ε - σ r ν j ( v ε , A ε ) + d ( η r ( 1 - ρ ε 2 ) j ( u ε , A ε ) ) .

Therefore, we can write

(4.14) - Δ ψ ε = ω 1 + ω 2 + ω 3 + ω 4 ,

where

(4.15) ω 1 := 2 η r J ( v ε , A ε ) , ω 2 := - η r F ε
(4.16) ω 3 := - σ r ν j ( v ε , A ε ) , ω 4 := d ( η r ( 1 - ρ ε 2 ) j ( u ε , A ε ) ) .

We have ω k W - 1 , 2 ( M , Λ 2 T * M ) for each k – in fact, ω 1 , ω 2 belong to L 2 ( M , Λ 2 T * M ) , ω 4 is the differential of an L 2 -form, and ω 3 W - 1 , 2 ( M , Λ 2 T * M ) by continuity of the trace operator W 1 , 2 ( M ) L 2 ( r ) . Let

H : W - 1 , 2 ( M , Λ 2 T * M ) Harm 2 ( M )

be the orthogonal projection onto the space of harmonic 2-forms, defined by

(4.17) H ( ω ) := j = 1 ω , ζ j 𝒟 ( M ) , 𝒟 ( M ) ζ j ,

where ( ζ 1 , , ζ ) is an orthonormal basis of the finite-dimensional space Harm 2 ( M ) . By the Lax–Milgram lemma, for each k there exists a unique ψ k W 1 , 2 ( M , Λ 2 T * M ) such that

(4.18) { - Δ τ k = ω k - H ( ω k ) in the sense of distributions in  M , M τ k , ζ vol g = 0 for all harmonic  2 - forms  ζ .

in the sense of distributions on M. We have ψ ε = ψ 1 + ψ 2 + ψ 3 + ψ 4 because, by construction, ψ ε is exact and hence, orthogonal to all harmonic 2-forms. We shall prove a decay estimate for each d * ψ k separately.

Estimate for τ 4 .

We claim that

(4.19) M | d * τ 4 | 2 vol g β 2 E ε ( x 0 , R ) .

Indeed, as ω 4 is exact, it is orthogonal to all harmonic forms, i.e., H ( ω 4 ) = 0 . Then, by testing the equation for τ 4 against τ 4 , we obtain

M ( | d τ 4 | 2 + | d * τ 4 | 2 ) vol g = r ( 1 - ρ ε 2 ) j ( u ε , A ε ) , d * τ 4 vol g .

By applying the Young inequality at the right-hand side, we deduce

M ( | d τ 4 | 2 + | d * τ 4 | 2 ) vol g r ( 1 - ρ ε 2 ) 2 | j ( u ε , A ε ) | 2 vol g ( 4.12 ) β 2 r | j ( u ε , A ε ) | 2 vol g

and (4.19) follows.

Estimate for τ 3 .

We shall prove that

(4.20) δ R | d * τ 3 | 2 vol g δ n E ε ( x 0 , R ) .

To this end, we will need the following lemma.

Lemma 4.4.

For any τ W 1 , 2 ( M , Λ k T * M ) , any x 0 M and any r > 0 small enough, there holds

1 r r ( x 0 ) | τ | 2 d n - 1 M ( | d τ | 2 + | d * τ | 2 + | τ | 2 ) vol g ,

where the implicit constant at the right-hand side does not depend on r, x 0 .

Proof.

For simplicity of notation, we will write r instead of r ( x 0 ) . We claim that, for any scalar function f W 1 , 2 ( r ) , there holds

(4.21) 1 r r f 2 d n - 1 r | d f | 2 vol g + 1 r 2 r f 2 vol g .

Inequality (4.21) holds true if r is a ball in n , equipped with the Euclidean metric. In case r is a geodesic ball in M (with r < inj ( M ) 2 ), estimate (4.21) follows by composition with (normal geodesic) coordinate charts. The implicit constant in front of the right-hand side of (4.21) depends only on the metric and is bounded uniformly with respect to x 0 , r, because M is compact and smooth.

Now, take a form τ W 1 , 2 ( M , Λ k T * M ) . Let ( x 1 , , x n ) be normal geodesic charts in r ( x 0 ) . Let us write τ = α τ α d x α , where the sum is taken over all multi-indices α of order k. By applying (4.21) to each component τ α , we deduce

1 r r | τ | 2 d n - 1 α r | d τ α | 2 vol g + 1 r 2 r | τ | 2 vol g τ W 1 , 2 ( M ) 2 + 1 r 2 τ L 2 ( r ) 2 .

The last term on the right-hand side can be estimated by applying the Hölder inequality and Sobolev embeddings. Indeed, if p := 2 * = 2 n n - 2 then 1 2 = 1 p + 1 n and hence,

1 r τ L 2 ( r ) 2 1 r 2 τ L p ( M ) 2 | vol ( r ) | 2 n τ W 1 , 2 ( M ) 2

(where the implicit constants are uniform with respect to r, x 0 ). The lemma now follows by applying Gaffney’s inequality (see, e.g., [21, Theorem 4.8]). ∎

Remark 4.5.

If τ W 1 , 2 ( M , Λ k T * M ) is orthogonal to all harmonic k-forms, then the L 2 ( M ) -norm of τ is bounded by the L 2 ( M ) -norms of d τ and d * τ (see, e.g., [21, Theorem 4.11]), and we obtain

1 r 2 r ( x 0 ) | τ | 2 d n - 1 M ( | d τ | 2 + | d * τ | 2 ) vol g .

Now, we proceed to the proof of (4.20). In the interior of r , the form τ 3 satisfies - Δ τ 3 = 0 . By taking the codifferential of both sides of this equation, we deduce - Δ ( d * τ 3 ) = 0 in the interior of r . Then standard decay estimates for elliptic equations (see, e.g., Theorem B.3 in the appendix) give

(4.22) δ R | d * τ 3 | 2 vol g δ n R | d * τ 3 | 2 vol g .

It remains to estimate the right-hand side of (4.22). By testing the equation for τ 3 against τ 3 , we obtain

M ( | d τ 3 | 2 + | d * τ 3 | 2 ) vol g = - r ν j ( v ε , A ε ) , τ 3 d n - 1 + M H ( ω 3 ) , τ 3 vol g r 2 λ r ρ ε 4 | j ( u ε , A ε ) | 2 d n - 1 + λ 2 r r | τ 3 | 2 d n - 1 + 1 2 λ M | H ( ω 3 ) | 2 vol g + λ 2 M | τ 3 | 2 vol g

for any λ > 0 . If λ is small enough, the last term on the right-hand side can be absorbed into the left-hand side, because τ 3 is orthogonal to all harmonic 2-forms (see, e.g., [21, Theorem 4.11]). Moreover, due to Lemma 4.4 and Remark 4.5, we can choose a constant λ small enough (uniformly with respect to r), so as to obtain

(4.23) M ( | d τ 3 | 2 + | d * τ 3 | 2 ) vol g R r | j ( u ε , A ε ) | 2 d n - 1 + M | H ( ω 3 ) | 2 vol g .

Finally, we estimate the L 2 ( M ) -norm of H ( ω 3 ) . We observe that ω 3 is a bounded measure, whose total variation is given by

| ω 3 | ( M ) = r | ν j ( u ε , A ε ) | d n - 1 ( n - 1 ( r ) ) 1 2 j ( u ε , A ε ) L 2 ( r ) .

From the definition of the projection operator H, equation (4.17), we immediately obtain

(4.24) H ( ω 3 ) L 1 ( M ) | ω 3 | ( M ) r n 2 - 1 2 j ( u ε , A ε ) L 2 ( r ) .

Since all norms on the finite-dimensional space Harm 2 ( M ) are equivalent, equations (4.23) and (4.24) together imply

(4.25) M ( | d τ 3 | 2 + | d * τ 3 | 2 ) vol g ( R + R n - 1 ) r | j ( u ε , A ε ) | 2 d n - 1 ( 4.6 ) E ε ( x 0 , R ) .

By combining (4.22) with (4.25), the desired estimate (4.20) follows.

Estimate for τ 2 .

We claim that, for any μ ( 0 , 2 ) , there exists a constant C μ > 0 (independent on ε, R, δ and the point x 0 M ) such that

(4.26) δ R | d * τ 2 | 2 vol g C μ δ n - μ E ε ( x 0 , R ) .

We test the equation for τ 2 against τ 2 . By applying Young’s inequality at the right-hand side, we obtain

M ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g 1 2 λ r ( | F ε | 2 + | H ( ω 2 ) | 2 ) vol g + λ M | τ 2 | 2 vol g .

Since H is the L 2 ( M ) -orthogonal projection onto the space of harmonic forms, the L 2 ( M ) -norm of H ( ω 2 ) is not greater than the L 2 ( M ) -norm of ω 2 = - η r F ε . Moreover, as τ 2 is orthogonal to all harmonic 2-forms, the L 2 ( M ) -norm of τ 2 can be estimated by the L 2 ( M ) -norms of d τ 2 and d * τ 2 (see, e.g., [21, Theorem 4.11]). Therefore, choosing λ small enough, we obtain

(4.27) M ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g r | F ε | 2 vol g E ε ( x 0 , R ) .

Now, let s, t be numbers such that 0 < s < t < r . The decay estimate in Proposition B.4 implies

(4.28) s ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g s n t n t ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g + t 2 t ( | F ε | 2 + | H ( ω 2 ) | 2 ) vol g

(so long as we choose R < R * , where R * is a small number that depends on M only, as given by Lemma A.2). By definition, H ( ω 2 ) is harmonic and its L 2 ( M ) -norm does not exceed the L 2 -norm of F ε in r . By applying elliptic decay estimates (see, e.g., Theorem B.3) again, we obtain

t 2 t ( | F ε | 2 + | H ( ω 2 ) | 2 ) vol g t 2 t | F ε | 2 vol g + t n + 2 r n r | H ( ω 2 ) | 2 vol g t 2 t | F ε | 2 vol g + t n + 2 r n r | F ε | 2 vol g .

The right-hand side can be further bounded by applying (4.5) and the decay estimate (3.15): for each α ( 0 , 2 ) , there exists a constant C α such that

(4.29) t 2 t ( | F ε | 2 + | H ( ω 2 ) | 2 ) vol g C α t n + α E ε ( x 0 , R ) R n - 2 + α + t n + 2 E ε ( x 0 , R ) R n .

We choose, say, α = 1 . Combining (4.28) with (4.29) and noticing that t R 1 , we obtain

(4.30) s ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g C 1 s n t n t ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g + C 2 t n E ε ( x 0 , R ) R n - 2

for some positive constants C 1 , C 2 that depend only on M (in particular, not on ε, R, or x 0 ). Let μ ( 0 , 2 ) be given. We choose a number θ ( 0 , 1 ) small enough that C 1 θ n θ n - μ 2 . Then inequality (4.30) implies

θ t ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g θ n - μ 2 t ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g + C 3 t n - μ E ε ( x 0 , R ) R n - 2

for any t ( 0 , r ) and some constant C 3 that depends only on C 2 , μ and R * . By an iteration argument (see, e.g., [20, Lemma III.2.1] or [6, Lemma B.3]), we deduce

t ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g C μ t n - μ r n - μ r ( | d τ 2 | 2 + | d * τ 2 | 2 ) vol g + C μ t n - μ E ε ( x 0 , R ) R n - 2 ( 4.27 ) , ( 4.5 ) C μ t n - μ E ε ( x 0 , R ) R n - μ

for all 0 < t < r , where C μ is a constant that depends only on μ, n, θ and C 3 . The desired estimate (4.26) now follows by taking t = δ R .

Estimate for τ 1 .

We will prove that

(4.31) δ R | d * τ 1 | 2 β - 4 E ε ( x 0 , R ) ( 1 ε 2 R n - 2 R ( 1 - | u ε | 2 ) 2 vol g ) .

To this purpose, we need an estimate for the “Green function” of the Hodge Laplacian, which is provided by Proposition B.1 below. Moreover, we will need the basic bound on curvatures in Lemma 1.7.

We proceed exactly as in [8, pp. 453–454]. First of all, we observe that

(4.32) | ω 1 | 1 β 2 ε 2 ( 1 - | u ε | 2 ) 2 pointwise in  r .

Indeed, in the open set { x r : | u ε ( x ) | > 1 - β } , we have | v ε | = 1 (by construction) and hence, ω 1 = 2 J ( v ε , A ε ) = 0 . In the complement, { x r : | u ε ( x ) | 1 - β } , we have

| D A ε v ε | ε - 1

because of (4.9)–(4.10) and the L -estimate (1.16) for D A ε u ε . Taking Lemma 1.7 into account, we deduce

| ω 1 | | J ( u ε , A ε ) | ( 5 ) | D A ε u ε | 2 + | F ε | 1 ε 2 = 1 β 2 ε 2 β 2 1 β 2 ε 2 ( 1 - | u ε | 2 ) 2

at each point of { x r : | u ε ( x ) | 1 - β } . Therefore, (4.32) is proved.

Next, we show that

(4.33) τ 1 L ( M ) E ε ( x 0 , R ) β 2 R n - 2 .

Indeed, for any x M , Proposition B.1 implies

(4.34) | τ 1 ( x ) | r | ω 1 ( y ) | dist ( x , y ) n - 2 vol g ( y )

(we have used the fact that ω 1 = 0 out of r ). If x M 2 r , then

| τ 1 ( x ) | 1 r n - 2 r | ω 1 ( y ) | vol g ( y ) ( 4.32 ) 1 β 2 ε 2 R n - 2 r ( 1 - | u ε | 2 ) 2 vol g E ε ( x 0 , R ) β 2 R n - 2 .

On the other hand, if x 2 r , then we have r 3 r ( x ) and hence,

| τ 1 ( x ) | 3 r | ω 1 ( y ) | dist ( x , y ) n - 2 vol g ( y ) ( 4.32 ) 1 β 2 ε 2 3 r ( x ) ( 1 - | u ε ( y ) | 2 ) 2 dist ( x , y ) n - 2 vol g ( y ) .

By applying Corollary 3.9, we obtain

| τ 1 ( x ) | E ε ( u ε , A ε ; 3 r ( x ) ) β 2 r n - 2 E ε ( x 0 , R ) β 2 R n - 2

(for the last inequality, we have used that 3 r ( x ) R , because x 2 r and r < R 5 ). Therefore, (4.33) is proved.

Finally, testing the equation for τ 1 against τ 1 , we obtain

M ( | d τ 1 | 2 + | d * τ 1 | 2 ) vol g τ 1 L ( M ) ω 1 - H ( ω 1 ) L 1 ( M ) .

From the definition of the harmonic projector H, equation (4.17), we immediately deduce that

H ( ω 1 ) L 1 ( M ) ω 1 L 1 ( M ) .

Therefore, taking (4.32) and (4.33) into account, we obtain

M ( | d τ 1 | 2 + | d * τ 1 | 2 ) vol g β - 4 E ε ( x 0 , R ) ( 1 ε 2 R n - 2 R ( 1 - | u ε | 2 ) 2 vol g )

and (4.31) follows.

Combining Lemma 4.3 with (4.31), (4.26), (4.20) and (4.19), we obtain a decay estimate for the prejacobian: for any δ ( 0 , 1 10 ) , μ ( 0 , 2 ) and β ( 0 , 1 4 ) , there holds

(4.35) δ R | j ( u ε , A ε ) | 2 vol g C μ ( δ n - μ + β 2 + β - 4 ( 1 ε 2 R n - 2 R ( 1 - | u ε | 2 ) 2 vol g ) ) E ε ( x 0 , R ) ,

where C μ is a constant that depends only on μ and the ambient manifold M (not on ε, δ, β, R).

Completing the proof of Proposition 4.2.

With (4.35) at our disposal, we can complete the proof of Proposition 4.2. The arguments are largely similar to those in [8], so we omit some details. Let

p ε := 1 ε 2 R n - 2 R ( 1 - | u ε | 2 ) 2 vol g

By reasoning exactly as in [8, Proof of Theorem 3, Step 3, p. 455], we can show that

(4.36) r | d ( | u ε | 2 ) | 2 vol g β 2 R | D A ε u | 2 vol g + 1 β 2 ε 2 R ( 1 - | u ε | 2 ) 2 vol p β 2 E ε ( x 0 , R ) + β - 2 p ε R n - 2 .

Moreover, we have

(4.37) r ( 1 - | u ε | 2 ) | D A ε u | 2 vol g β 2 E ε ( x 0 , R ) + β - 2 p ε R n - 2

(as in [8, Equation (III.32)]). From Lemma 1.1, we have

| u ε | 2 | D A ε u ε | 2 | u ε | 2 | d ( | u ε | ) | 2 + | j ( u ε , A ε ) | 2 | d ( | u ε | 2 ) | 2 + | j ( u ε , A ε ) | 2

and hence,

(4.38) | D A ε u ε | 2 | d ( | u ε | 2 ) | 2 + | j ( u ε , A ε ) | 2 + ( 1 - | u ε | 2 ) | D A ε u ε | 2 .

Combining (4.38) with (4.35), (4.36) and (4.37), we deduce

(4.39) δ R | D A ε u ε | 2 vol g C μ ( δ n - μ + β 2 + β - 4 p ε ) E ε ( x 0 , R ) + β - 2 p ε R n - 2

for all μ ( 0 , 2 ) . On the other hand, by (3.15), there holds

(4.40) δ R | F ε | 2 vol g δ n - 2 + α E ε ( x 0 , R )

for some α > 0 that depends only on M (but is independent of R, δ, ε). Choosing μ = 2 - α in (4.39), and taking (4.40) into account, we immediately deduce

E ( δ R ) ( δ n - 2 + α + β 2 + β - 4 p ε ) E ε ( x 0 , R ) + β - 2 p ε R n - 2 .

Finally, choosing β = min { ( p ε ) 1 6 , 1 8 } , the proposition follows. ∎

4.2 Proof of Proposition 4.1

The energy decay estimate given by Proposition 4.2, and the monotonicity formula (Theorem 3), together, imply a “clearing-out” or “η-ellipticity” result (cf. [8, Theorem 2]).

Proposition 4.6.

There exist positive numbers R * , ε * , C and γ, depending on M only, such that the following statement holds. For any x 0 M , 0 < R < R * , 0 < ε < ε * R , and any critical point ( u ε , A ε ) E of G ε , there holds

(4.41) | u ε ( x 0 ) | 1 - C ( E ε ( x 0 , R ) log ( R ε ) R n - 2 ) γ .

Proposition 4.6 is only significant when the energy of ( u ε , A ε ) on the ball R ( x 0 ) is small compared to the logarithm of ε. Indeed, the estimate (4.41) can be rephrased as follows: if

E ε ( x 0 R ) η R n - 2 log R ε

for some η > 0 , then | u ( x 0 ) | 1 - C η γ . In particular, if η is small enough, then | u ε ( x 0 ) | 1 2 .

We proceed to the proof of Proposition 4.6. Let x 0 , R, ε be fixed, and let ( u ε , A ε ) ( u ε , A ε ) be a critical point of G ε . Let

(4.42) η := E ε ( x 0 , R ) log ( R ε ) R n - 2 .

As above, we omit x 0 from the notation when convenient.

Lemma 4.7.

There exists a small number ε * , depending on M only, such that the following statement holds. If 0 < ε < ε * R , then for any 0 < δ < 1 10 there exists a radius R * ( ε 3 4 R 1 4 , ε 1 2 R 1 2 ) such that

(4.43) E ε ( x 0 , R * ) R * n - 2 E ε ( x 0 , δ R * ) ( δ R * ) n - 2 + η | log δ | ,
(4.44) 1 ε 2 R * n - 2 R * ( 1 - | u ε | 2 ) 2 vol g η | log δ | .

Proof.

The proof follows along the lines of [8, Lemma III.1]. For any integer j 0 , let

R j := ε 1 2 R 1 2 ( δ 4 ) j .

Let k be the unique integer such that

(4.45) R k > ε 3 4 R 1 4 , R k + 1 ε 3 4 R 1 4 .

Let C > 0 be a constant. By the monotonicity formula (Theorem 3) and (4.42), for a suitable choice of C (depending on M only) we have

j = 0 k - 1 ( e C R j 2 E ε ( x 0 , R j ) R j n - 2 - e C R j + 1 2 E ε ( x 0 , R j + 1 ) R j + 1 n - 2 ) e C R 0 2 E ε ( x 0 , R 0 ) R 0 n - 2 E ε ( x 0 , R ) R n - 2 = η log ( R ε ) .

On the other hand, the second inequality in (4.45) implies, via some algebraic manipulation,

k + 1 log ( R ε ) 4 log ( 4 δ ) .

As a consequence, there must be an integer J { 0 , , k - 1 } such that

(4.46) e C R J 2 E ε ( x 0 , R J ) R J n - 2 - e C R J + 1 2 E ε ( x 0 , R J + 1 ) R J + 1 n - 2 η k log ( R ε ) η log ( 4 δ ) η | log δ | .

Due to the mean value theorem, there exists R * ( R J 2 , R J ) such that

(4.47) d d ρ | ρ = R * ( e C ρ 2 E ε ( x 0 , ρ ) ρ n - 2 ) η | log δ | R * .

Then the monotonicity formula (see equation (3.22)) and (4.42) imply

1 ε 2 R * n - 2 R * ( 1 - | u ε | 2 ) 2 vol g R * d d ρ | ρ = R * ( e C ρ 2 E ε ( x 0 , ρ ) ρ n - 2 ) + ( R * R ) α E ε ( x 0 , R ) R n - 2 η | log δ | + η ( ε R ) α 2 log ( R ε ) .

If the ratio ε R is small enough, then ( ε R ) α 2 log ( R ε ) 1 | log δ | for any δ ( 0 , 1 10 ) and (4.44) follows. As for the proof of (4.43), we observe that R J / 2 R * R J and hence, δ R * R J + 1 . Then the same monotonicity formula (3.22) implies

e C R * 2 E ε ( x 0 , R * ) R * n - 2 - e C ( δ R * ) 2 E ε ( x 0 , δ R * ) ( δ R * ) n - 2 + C E ε ( x 0 , R ) R n - 2 + α δ R * R * ρ α - 1 d ρ
e C R J 2 E ε ( x 0 , R J ) R J n - 2 - e C R J + 1 2 E ε ( x 0 , R J + 1 ) R J + 1 n - 2 + C E ε ( x 0 , R ) R n - 2 + α R j + 1 R j ρ α - 1 d ρ

for some constant C that depends only on M. Therefore, taking (4.42) and (4.46) into account, we deduce

e C R * 2 E ε ( x 0 , R * ) R * n - 2 - e C ( δ R * ) 2 E ε ( x 0 , δ R * ) ( δ R * ) n - 2 η | log δ | + ( R J R ) α E ε ( x 0 , R ) R n - 2 η | log δ | + η ( ε R ) α 2 log ( R ε ) .

If ε R is small enough, the desired estimate (4.43) follows. ∎

Proof of Proposition 4.6.

We proceed as in [8, Theorem 2, Part C, p. 456]. Let η 0 > 0 be a small parameter, to be chosen later on (depending on M only). Let η be defined as in (4.42). If η > η 0 , then Proposition 4.6 holds true for any value of γ > 0 , so long as we choose a constant C large enough (depending on γ, η 0 only). Therefore, it suffices to consider the case η η 0 . Let δ ( 0 , 1 10 ) be fixed, and let R * ( ε 3 4 R 1 4 , ε 1 2 R 1 2 ) be a radius that satisfies (4.43) and (4.44), as given by Lemma 4.7. For such a choice of R * , Proposition 4.2 implies

E ε ( x 0 , R * ) R * n - 2 E ε ( x 0 , δ R * ) ( δ R * ) n - 2 + η | log δ | ( ( η | log δ | ) 1 3 δ n - 2 + δ α ) E ε ( x 0 , R * ) R * n - 2 + ( η | log δ | ) 2 3 δ n - 2 + η | log δ | .

Assume δ is such that η | log δ | 1 . Then we obtain

( 1 - C ( η | log δ | ) 1 3 δ n - 2 - C δ α ) E ε ( x 0 , R * ) R * n - 2 ( η | log δ | ) 2 3 δ n - 2

for some uniform constant C. We choose δ = η 1 3 n and we take η 0 so small that for any η ( 0 , η 0 ] there holds η | log η | 1 and C η 2 3 n | log η | + C η α 3 n 1 2 . Then we deduce

E ε ( x 0 , R * ) R * n - 2 η n + 2 3 n | log η | 2 3

for any η η 0 . Moreover, assuming ε < R , we have R * > ε 3 4 R 1 4 > ε and hence, the monotonicity formula (Theorem 3) implies

1 ε n ε ( 1 - | u ε | 2 ) 2 vol g E ε ( x 0 , ε ) ε n - 2 E ε ( x 0 , R * ) R * n - 2 η n + 2 3 n | log η | 2 3 .

Now the proposition follows by repeating the arguments in [8, Lemma III.3]. ∎

Finally, we are in a position to prove the main result of this section, Proposition 4.1.

Proof of Proposition 4.1.

In view of Proposition 4.6, the property (4.2) follows by the very same argument as in [8, Proposition VII.1]. The proof of (4.3) is also inspired by [8, Proposition VII.1]. Suppose, towards a contradiction, that (4.3) fails. Then there exists a (non-relabelled) subsequence such that

(4.48) E ε ( x 0 , R 2 ) + as  ε 0 .

We split the rest of the proof into steps.

Step 1.

Locally, on the ball R := R ( x 0 ) , we may write the reference connection as D 0 = d - i γ 0 for some (smooth, ε-independent) 1-form γ 0 . Upon replacing A ε with A ε + γ 0 , from now on we assume that D A ε = d - i A ε on R . Moreover, we assume that A ε is in Coulomb gauge on R – that is, A ε satisfies (3.6). Then the Gaffney inequality (see, e.g., Proposition A.1 in the appendix for details) implies

A ε L 2 ( R ) R F ε L 2 ( R ) ( 4.1 ) R n 2 ( log R ε ) 1 2 .

Moreover, Sobolev embeddings and the Gaffney inequality again imply

(4.49) A ε L 2 * ( R ) C R A ε W 1 , 2 ( R ) C R ( A ε L 2 ( R ) + F ε L 2 ( R ) ) C R | log ε | 1 2

for 2 * := 2 n n - 2 and some constant C R depending on R (and, possibly, on x 0 ). On the other hand, the energy bound (4.1) and the L -estimate given by Proposition 1.2 imply, by interpolation,

(4.50) 1 - | u ε | 2 L n ( R ) 1 - | u ε | 2 L 2 ( R ) 2 n 1 - | u ε | 2 L ( R ) 1 - 2 n C R ε 2 n | log ε | 1 n .

From (4.49) and (4.50), we conclude that

(4.51) ( 1 - | u ε | 2 ) A ε L 2 ( R ) 0 as  ε 0 .

Step 2.

On the smaller ball 3 R / 4 := 3 R / 4 ( x 0 ) , the function ρ ε := | u ε | is bounded away from zero, thanks to (4.2). Therefore, by identifying u ε with a map 3 R / 4 , we find a (Lipschitz-continuous) function θ ε : 3 R / 4 such that u ε = ρ ε exp ( i θ ε ) . By direct computation, we see that

(4.52) j ( u ε , A ε ) = ρ ε 2 ( d θ ε - A ε ) .

Assume that ( u ε , A ε ) satisfies (9). Then, up to extraction of a subsequence, we know that j ( u ε , A ε ) converges in L p ( M ) and F ε converges in W 1 , p ( M ) for any p < n n - 1 (by Lemma 1.10 and (1.30), respectively). Since A ε is in Coulomb gauge (equation (3.6)), from the Gaffney inequality (Proposition A.1) we deduce that A ε is bounded in L p ( R ) for p < n n - 1 . As a consequence, from (4.2) and (4.52) we obtain that d θ ε is bounded in L p ( 3 R / 4 ) . Up to subtracting a constant multiple of 2 π , we can moreover assume that the average of θ ε on B 3 R / 4 belongs to the interval [ 0 , 2 π ) ; then it follows

(4.53) θ ε W 1 , p ( 3 R / 4 ) C R , p

for any p < n n - 1 and some constant C R , p depending on R and p (and possibly, on x 0 ), but not on ε.

Step 3.

The Euler–Lagrange equation (7) implies that the form j ( u ε , A ε ) is co-exact – in particular, co-closed. The form A ε is co-closed in 3 R / 4 , too, because of our choice of gauge (3.6). Therefore, we have

0 = d * j ( u ε , A ε ) + d * A ε = ( 4.52 ) d * ( ρ ε 2 d θ ε - ρ ε 2 A ε + A ε )

and hence,

(4.54) d * ( ρ ε 2 d θ ε ) = d * ( ( ρ ε 2 - 1 ) A ε ) .

By a suitable Caccioppoli inequality (see, e.g., Lemma B.5 in the appendix), we deduce

d θ ε L 2 ( 2 R / 3 ) R - n 2 - 1 θ ε L 1 ( 3 R / 4 ) + ( 1 - ρ ε 2 ) A ε L 2 ( 3 R / 4 )

and hence, thanks to (4.51) and (4.53),

(4.55) d θ ε L 2 ( 2 R / 3 ) C R

for some constant C R that depends on R (and possibly, on x 0 ), but not on ε.

Step 4.

Due to the choice of gauge (3.5), the Euler–Lagrange equation (7) rewrites as

(4.56) - Δ A ε = j ( u ε , A ε ) on  R .

Combining (4.56) with (4.52), we deduce

(4.57) - Δ A ε + ρ ε 2 A ε = ρ ε 2 d θ ε on  R .

The right-hand side of (4.57) is uniformly bounded in L 2 ( 3 R / 4 ) , because of (4.55). A Caccioppoli-type inequality (see Lemma B.7 in the appendix for details) now implies

(4.58) A ε L 2 ( 7 R / 12 ) + d A ε L 2 ( 7 R / 12 ) C R

for some constant C R that depends on R, but not on ε.

Step 5.

It remains to estimate the W 1 , 2 -norm of ρ ε := | u ε | . By explicit computation, recalling that u ε = ρ ε e i θ ε and D A ε = d - i A ε , we have

- 1 2 Δ ( ρ ε 2 ) + | D A ε u ε | 2 = - ρ ε Δ ρ ε + ρ ε 2 | d θ ε - A ε | 2 .

By injecting this identity into equation (1.2), and recalling that ρ ε 1 2 in 3 R / 4 , we obtain

(4.59) - Δ ρ ε + 1 ε 2 ρ ε ( ρ ε 2 - 1 ) + ρ ε | d θ ε - A ε | 2 = 0 in  3 R / 4 .

We test equation (4.59) against ( 1 - ρ ε ) ζ 2 , where ζ C c ( 7 R / 12 ) is a cut-off function, such that ζ = 1 in R / 2 and | ζ | R - 1 . We obtain

7 R / 12 ( | ρ ε | 2 + 1 ε 2 ρ ε ( 1 + ρ ε ) ( 1 - ρ ε ) 2 ) ζ 2 vol g = 7 R / 12 2 ζ ( 1 - ρ ε ) ρ ε , ζ vol g + 7 R / 12 ρ ε ( 1 - ρ ε ) ζ 2 | d θ ε - A ε | 2 vol g .

The second term on the right-hand side is bounded uniformly with respect to ε, due to (4.55), (4.58) and the bound | ρ ε | 1 . Therefore, upon applying the Young inequality at the right-hand side, we deduce that

(4.60) R / 2 ( | ρ ε | 2 + 1 ε 2 ( 1 - ρ ε 2 ) 2 ) vol g C R

for some ε-independent constant C R . From (4.55), (4.58) and (4.60), we obtain the energy bound E ε ( x 0 , R 2 ) C R , which contradicts (4.48). This completes the proof.∎

5 Convergence to the limiting varifold

In this section we consider the rescaled energy measures, defined as

(5.1) μ ε := e ε ( u ε , A ε ) π | log ε | vol g ,

where we identify vol g with a Radon measure on M, in a canonical way. We prove that the measures μ ε converge weakly * in the sense of Radon measures in M to a limiting Radon measure μ * and that μ * is the weight measure of a stationary, rectifiable ( n - 2 ) -varifold in M. For a precise formulation of this result, cf. Theorem 5.3 below.

A k-varifold V in a Riemannian n-manifold M is defined as a nonnegative Radon measure on G k ( T M ) , where k n is an integer and G k ( T M ) is the (total space of the) Grassmannian bundle on M, a bundle whose typical fibre is the Grassmannian manifold of k-planes in the tangent space to M. (Details on the construction of G k ( T M ) can be found, for instance, in [26, Section 6.5.7].) We will denote the set of k-varifolds in M by 𝒱 k ( M ) .

Remark 5.1.

Varifolds lack a boundary operator and there is no natural notion of orientation on them. Thus, varifolds are less regular objects than currents (indeed, any current induces a varifold but the converse does not hold), although they are both generalisations of the concept of smooth submanifold.

As in [29], one wishes to prove the rectifiability of the limiting varifold by applying (a suitable variant of) the Ambrosio-Soner rectifiability criterion for generalised varifolds, i.e., of [4, Theorem 3.8 ]. Since in [4] the authors work in the Euclidean setting, an extension of the Ambrosio-Soner theorem to the present setting is needed but, as explained in [29, p. 1058], easy to obtain. For the reader’s convenience, we recall the main points of the argument below.

Following [29, Section 6], we use the metric g to canonically identify tensors of rank ( 2 , 0 ) , ( 1 , 1 ) and ( 0 , 2 ) . We denote by End sym ( T M ) the space of symmetric endomorphisms of T M and we define a subbundle 𝒜 k , n ( M ) of End sym ( T M ) as follows:

𝒜 k , n ( M ) := { S End sym ( T M ) : - n g S g  and  Tr ( S ) k } .

The typical fibre of 𝒜 k , n ( M ) at x M is the set 𝒜 k , n ( T x M ) of symmetric endomorphisms S x : T x M T x M with trace k and such that - n g x S x g x as bilinear forms on T x M .

We define a generalised k-varifold in M as a nonnegative Radon measure on 𝒜 k , n ( M ) , thus extending in the most natural way the definition in [4, Section 3] (cf. [29, Section 6]). We denote 𝒱 k * ( M ) the set of generalised k-varifolds on 𝒜 k , n ( M )

Let V 𝒱 k * ( M ) and let π be the canonical projection 𝒜 k , n ( M ) M . We call V := π * V the weight measure of V, where π denotes the canonical projection 𝒜 k , n ( M ) M and π * V denotes the pushforward measure of V through π. We say that V has first variation if and only if there exists a vector Radon measure δ V C ( M , T M ) such that, for any X C 1 ( M , T M ) ,

(5.2) 𝒜 k , n ( M ) S x , X ( x ) d V ( x , S x ) = - M X , ν d δ V ,

where denotes the Levi Civita connection of M, S x 𝒜 k , n ( T x M ) , S x , X ( x ) denotes the scalar product of the matrices representing S x and X ( x ) , regarded as endomorphisms of T x M , and ν is a δ V -measurable vector field on M with | ν | = 1 δ V -a.e. in M. When (5.2) holds, we say that δ V is the first variation of V. We call V stationary if δ V ( X ) = 0 for all X C 1 ( M , T M ) . We say that a sequence { V h } 𝒱 k * ( M ) converges weakly * to V 𝒱 k * ( M ) if and only if { V h } converges weakly * in the sense of measures on 𝒜 k , n ( M ) to V. Next, we define the upper k-dimensional density of V at x M as

Θ k * ( V , x ) := lim sup r 0 V ( r ( x ) ) μ ( r ( x ) ) .

Keeping in mind the construction of G k ( T M ) , a standard localisation argument shows that the following version of [4, Theorem 3.8 (c)] holds (cf. [37, Proposition 5.1]):

Theorem 5.2 (Rectifiability criterion).

Let V V k * ( M ) be a generalised k-varifold with first variation and assume that

Θ k * ( V , x ) > 0 for  V -a.e.  x M .

Then there exists a (classical) k-rectifiable varifold V ~ such that

V = V ~ 𝑎𝑛𝑑 δ V = δ V ~ .

We follow the strategy of [29, Section 6], applying the modifications made necessary by the logarithmic energy regime.

For any given ( u , A ) , we define the ( 0 , 2 ) tensors D A u * D A u and F A * F A by

D A u * D A u ( E i , E j ) := i , j = 1 n D A , E i u , D A , E j u ,

and, respectively,

F A * F A ( E i , E j ) := i , j = 1 n F A ( E i , E j ) , F A ( E i , E j ) ,

where { E 1 , , E n } is any local orthonormal frame for E M . It is easily checked that D A u * D A u and F A * F A are gauge-invariant and smooth for any critical pair ( u , A ) .

Next, for any ε > 0 , we consider the rescaled stress-energy tensor field

(5.3) T ε T ε ( u ε , A ε ) := 1 | log ε | ( e ε ( u ε , A ε ) g - D A ε u * D A ε u - F A ε * F A ε ) ,

where ( u ε , A ε ) is any critical point of G ε . By the same computations as in [29, Section 4], we see that T ε is divergence-free, for any ε > 0 . By means of the metric, we can identify T ε with the induced End sym ( T M ) -valued Radon measure in M. If we can show that the sequence { T ε } converges to a limiting End sym ( T M ) -valued measure T * and that | T * | is absolutely continuous with respect to μ * , then we can represent T * by a generalised ( n - 2 ) -varifold P. The necessary estimates to this purpose are provided by Lemma 5.8. These estimates ensure, in addition, that we can apply Theorem 5.2 to obtain the rectifiability of P.

The main result of this section is Theorem 5.3 below (cf. [29, Proposition 6.4]).

Theorem 5.3.

Let { μ ε } , where μ ε is defined by (5.1), be the rescaled energy measures of a sequence { ( u ε , A ε ) } E of critical points of G ε satisfying the logarithmic energy bound (9). Then there exist a bounded Radon measure μ * on M and a (not relabelled) subsequence such that μ ε * μ * weakly * in the sense of measures. Moreover, the ( n - 2 ) -density Θ ( μ * , x 0 ) := lim r 0 r 2 - n μ * ( B r ( x 0 ) ) of μ * at x 0 is defined at every x 0 M and μ * is the weight measure of an associated stationary, rectifiable ( n - 2 ) -varifold V = v ( Σ , θ ) , which in turn is such that

lim ε 0 M T ε ( u ε , A ε ) , S vol g = Σ θ ( x ) T x Σ , S ( x ) d n - 2 ( x )

for all S C 0 ( M , End sym ( T M ) ) , where Σ := spt μ * , θ ( x ) := Θ n - 2 ( μ * , x ) , and μ * = θ H n - 2 Σ .

Remark 5.4.

The limiting measure μ * given by Theorem 5.3 is purely concentrated on a codimension-two varifold and it does not contain any “diffuse part” (i.e., a contribution absolutely continuous with respect to vol g ). The latter is instead typical of the “non-magnetic” Ginzburg–Landau functional

(5.4) I ε ( v ) = M 1 2 | d v | 2 + 1 4 ε 2 ( 1 - | v | 2 ) 2 vol g for  v W 1 , 2 ( M , )

without further constraints. For instance, it is proven in [38] that the rescaled energy measures of critical points v ε of I ε (satisfying I ε ( v ε ) O ( | log ε | ) as ε 0 ) concentrate towards a measure of the form μ ~ := V + | ψ | 2 vol g , where ψ is a possibly non-trivial harmonic 1-form on M. A similar phenomenon occurs also in the evolutionary case, even in the Euclidean setting (i.e., in n × + ). Indeed, it is proven in [9, Theorem A] that for solutions to the Ginzburg–Landau heat flow, in the logarithmic energy regime, the rescaled parabolic energy measures converge to measures of the form μ ^ ( x , t ) = V ^ + | Φ ^ | 2 d x , where Φ ^ is a solution of the heat equation in n × + . In rough terms, the diffuse part in the limiting measure arises because of wild oscillations in the phase made possible by the high energy at disposal. In the situation considered in [8], this phenomenon is ruled out by topological arguments. The same would occur if the manifold M is simply connected. In Theorem 5.3, there is no diffuse part in μ * because of the structure of the gauge-invariant functional, which allows us to remove the non-topological contributions to the energy, to leading order, by a suitable choice of the connection D A . This is independent of the topology of M; it is a feature of the energy functional G ε .

The first step in the proof of Theorem 5.3 is the characterisation of the limiting measure μ * .

Lemma 5.5.

Let { μ ε } , where μ ε is defined by (5.1), be the rescaled energy measures of a sequence { ( u ε , A ε ) } E of critical points of G ε satisfying the logarithmic energy bound (9). Then there exist a bounded Radon measure μ * on M and a (not relabelled) subsequence such that μ ε μ * weakly * in the sense of measures. Moreover, there exist constants C > 0 and R 0 ( 0 , inj ( M ) ) , depending on ( M , g ) only, such that the measure μ * satisfies

(5.5) e C r 2 r n - 2 μ * ( r ( x 0 ) ) + 2 C α e C R 2 R n - 2 + α r α μ * ( R ( x 0 ) ) e C R 2 R n - 2 μ * ( R ( x 0 ) ) ( 1 + 2 C α )

for every x 0 M , every 0 < r < R R 0 and some α > 0 depending on ( M , g ) only. As a consequence, the ( n - 2 ) -density

(5.6) Θ n - 2 ( μ * , x 0 ) := lim r 0 r 2 - n μ * ( r ( x 0 ) )

is defined at every x 0 M .

Proof.

Up to small modifications, the proof is as in [29, Section 6]. The existence of μ * is standard, because the uniform bound G ε ( u ε , A ε ) | log ε | Λ immediately yields a uniform bound on μ ε ( C 0 ( M ) ) * . The existence of the limit on the right-hand side of (5.6) is a straightforward consequence of (5.5). Hence, it suffices to prove (5.5). To the last purpose, fix arbitrarily x 0 M . Let C > 0 and let R 0 ( 0 , inj ( M ) ) be the same constants (depending on ( M , g ) only) as in Theorem 3. Then, by the definition on μ ε and Theorem 3,

e C R 2 R 2 - n μ * ( R ( x 0 ) ¯ ) ( 1 + 2 C α ) lim sup ε 0 e C R 2 R 2 - n μ ε ( R ( x 0 ) ¯ ) ( 1 + 2 C α ) = lim sup ε 0 e C R 2 R 2 - n μ ε ( R ( x 0 ) ) ( 1 + 2 C α ) lim inf ε 0 e C R 2 R 2 - n μ ε ( R ( x 0 ) ) ( 1 + 2 C α ) lim inf ε 0 { e C r 2 r 2 - n μ ε ( r ( x 0 ) ) + 2 C α e C R 2 R n - 2 + α r α μ ε ( R ( x 0 ) ) } e C r 2 r 2 - n μ * ( r ( x 0 ) ) + 2 C α e C R 2 R n - 2 + α r α μ * ( R ( x 0 ) )

for any 0 < r < R < R 0 . Then (5.5) follows by approximating R from below. Approximating R 0 with an increasing sequence { R k } , we see that (5.5) holds also for R = R 0 . This concludes the proof of the lemma. ∎

Lemma 5.6 below is the counterpart of [8, Proposition VIII.1] and it follows from Proposition 4.1 exactly as [8, Proposition VIII.1] follows from [8, Proposition VII.1].

Lemma 5.6.

There exists a constant η 0 > 0 , depending on Λ and ( M , g ) only, such that, if

μ * ( R ( x 0 ) ) < η 0 R n - 2 ,

then

μ * ( R / 2 ( x 0 ) ) = 0 ,

i.e., B R / 2 ( x 0 ) M spt μ * .

Lemma 5.7.

There exist a constant C * := C * ( M , Λ ) > 0 and a number R * ( 0 , inj ( M ) ) , depending on ( M , g ) only, such that

(5.7) C * - 1 r 2 - n μ * ( r ( x 0 ) ) C * for any  x 0 spt μ * and  r ( 0 , R * ] ,

whence

(5.8) C * - 1 Θ n - 2 ( μ , x 0 ) C * for any  x 0 spt μ * .

Proof.

The upper bound r 2 - n μ * ( B r ( x 0 ) ) C ~ , for some C ~ > 0 depending only on ( M , g ) and Λ, follows from (5.5) and the obvious inequality μ * ( r ( x 0 ) ) μ * ( M ) .

The lower bound is a straightforward consequence of Lemma 5.6. Indeed, were false, for every choice of c * > 0 , we could find x 0 spt μ * and r 0 ( 0 , R * ] such that μ * ( r 0 ( x 0 ) ) < c * r 0 n - 2 . By choosing c * < η 0 , where η 0 is the constant (depending on Λ and ( M , g ) only) of Lemma 5.6, we would deduce x 0 M spt μ * , a contradiction. Then (5.7) follows by letting C * := max { C ~ , η 0 - 1 } . In turn, (5.8) follows immediately from (5.7) and (5.6). ∎

We are now in a position to prove the following lemma, which is the counterpart of [29, Lemma 6.3] and, as in [29], is the key missing tool in the proof of Theorem 5.3.

Lemma 5.8.

Let { T ε } be the rescaled stress-energy tensor fields, defined by (5.3), of a sequence { ( u ε , A ε ) } E of critical points of G ε satisfying the logarithmic energy bound (9). Then there exists a bounded End sym ( M ) -valued Radon measure T * such that, up to extraction of a subsequence, T ε * T * , weakly * as End sym ( M ) -valued measures. Moreover, the following statements hold.

  1. T * is divergence-free, i.e., T * , X = 0 for all vector fields X C 1 ( M , T M ) (here, denotes the Levi Civita connection of M ).

  2. For every nonnegative φ C 0 ( M ) , there holds T * , φ g ( n - 2 ) μ * , φ .

  3. For all X C 0 ( M , T M ) , there holds

    - M | X | 2 d μ * T * , X X M | X | 2 d μ * .

Proof.

The existence of T * is standard, as the uniform bound G ε ( u ε , A ε ) | log ε | Λ immediately yields a uniform bound on T ε ( C 0 ( M ) ) * . Next, since each T ε is divergence-free (because ( u ε , A ε ) is a critical point of G ε ) and this property clearly passes to weak * -limits in the sense of measures, T * is divergence-free as well. This proves (i). The proof of (iii) proceeds exactly as in [29, Lemma 6.3]. Concerning (ii), take any φ C 0 ( M ) such that φ 0 . Then

M T ε , φ g vol g = 1 | log ε | M φ ( n e ε ( u ε , A ε ) - | D A ε | 2 - 2 | F A ε | 2 ) vol g = 1 | log ε | M ( n - 2 ) φ e ε ( u ε , A ε ) vol g + 1 | log ε | M φ { 1 2 ε 2 ( 1 - | u ε | 2 ) 2 - | F A ε | 2 } vol g ( n - 2 ) μ ε , φ - 1 | log ε | M | F A ε | 2 vol g .

By Lemma 1.9, 1 | log ε | M | F A ε | 2 vol g 0 as ε 0 , and (ii) follows. ∎

Proof of Theorem 5.3.

The existence of μ * and of its ( n - 2 ) -density Θ ( μ * , x 0 ) at every x 0 M follow by Lemma 5.5. Statement (iii) in Lemma 5.8 implies that the measure | T * | is absolutely continuous with respect to μ * . Hence, by the Radon–Nikodym theorem, we can represent T * by means of an L 1 ( μ * ) -section P : M End sym ( T M ) ; i.e., for all S C 0 ( M , End sym ( T M ) ) we can write

T * , S = M P ( x ) , S ( x ) d μ * ( x ) .

Still by Lemma 5.8, we infer that - g x P ( x ) g x (as bilinear forms) and Tr ( P ( x ) ) n - 2 for μ * -a.e. x M . This means that T * defines a generalised ( n - 2 ) -varifold in M, with weight measure μ * . Stationarity of such a varifold follows from (i) of Lemma 5.8 and rectifiability follows from Theorem 5.2, which we can apply thanks to (5.8). In particular, spt μ * is ( n - 2 ) -rectifiable, μ * coincides with θ n - 2 spt μ * , where θ ( x ) := Θ n - 2 ( μ * , x ) , and P ( x ) is given, at μ * -a.e. x M , by the orthogonal projection onto the ( n - 2 ) -subspace T x spt μ * T x M . This proves the theorem. ∎

Finally, we record here a further property of the limit measure μ * . Lemma 1.9 implies that μ * can be characterised in terms of u ε only, so long as we choose a convenient gauge.

Proposition 5.9.

Let { ( u ε , A ε ) } ε > 0 be a sequence of solutions of (6)–(7) that satisfies (9). Assume that A ε is in the form (11). Then

(5.9) μ * = lim ε 0 | D 0 u ε | 2 2 | log ε | = lim ε 0 | j ( u ε , 0 ) | 2 2 | log ε |

(the limit being taken weakly * in the sense of measures).

Proof.

By taking the differential in both sides of (11), we obtain

- Δ ψ ε = d d * ψ ε = d A ε = F ε - F 0 ,

where F 0 is the curvature of the reference connection. Moreover, the form ψ ε is exact, hence orthogonal to all harmonic 2-forms. By elliptic regularity theory, it follows that the L 2 ( M ) -norm of - Δ ψ ε bounds the W 2 , 2 ( M ) -norm of ψ ε from above (up to a constant factor). We deduce that

A ε W 1 , 2 ( M ) ψ ε W 2 , 2 ( M ) + ζ ε W 1 , 2 ( M ) F ε L 2 ( M ) + F 0 L 2 ( M ) + C M

(as ζ ε is a harmonic form, the W 1 , 2 ( M ) -norm of ζ ε is bounded by the L ( M ) -norm, up to a constant factor that depends on M only). Then Lemma 1.9 implies that the W 1 , 2 ( M ) -norm of A ε is of order lower than | log ε | 1 2 , as ε 0 . This implies

(5.10) μ * = lim ε 0 1 | log ε | ( 1 2 | D 0 u ε | 2 + 1 4 ε 2 ( 1 - | u ε | 2 ) 2 ) ,

where the limit is taken weakly * in the sense of measures.

Let β ( 1 2 , 1 ) be given. We define ρ ε := | u ε | and

M ε , β := { x M : ρ ε ( x ) 1 2 } , N ε , β := M M ε , β .

By reasoning exactly as in [8, Proposition 1], we can prove that

(5.11) 1 ε 2 N ε , β ( 1 - ρ ε 2 ) 2 vol g C β

for some constant C β that depends on β, but not on ε. Estimate (5.11) ultimately depends on the monotonicity formula given by Theorem 3 and the clearing out property, i.e., Proposition 4.6. Note that the arguments of [8] rely on Besicovitch’s covering theorem, but the latter holds true on closed Riemannian manifolds (see, for instance, [19, Theorem 2.8.14] for a statement that applies to a much larger class of metric spaces). Note that (5.11) implies vol g ( N ε , β ) C β ε 2 (for a different constant C β ) and hence, on account of Lemma 1.6,

(5.12) N ε , β | D A ε u ε | 2 vol g C β .

Next, we claim that

(5.13) lim sup ε 0 1 | log ε | M ( 1 2 | d ρ ε | 2 + 1 4 ε 2 ( 1 - ρ ε ) 2 ) vol g = 0 .

To prove (5.13), we first observe that | d ρ ε | | D A ε u ε | (by Lemma 1.1), so we have

(5.14) N ε , β ( 1 2 | d ρ ε | 2 + 1 ε 2 ( 1 - ρ ε ) 2 ) vol g C β

because of (5.12). Then, to estimate the integral on M ε , β , we take ρ ¯ ε := max ( ρ ε , β ) , we test equation (1.2) against 1 - ρ ¯ ε and we integrate over M:

(5.15) M ε , β ( ρ ε | d ρ ε | 2 + 1 ε 2 ρ ε ( 1 + ρ ε ) ( 1 - ρ ε ) 2 ) vol g M ( 1 - ρ ¯ ε ) | D A ε u ε | 2 vol g .

We combine (5.14) with (5.15), and we use (5.12) and assumption (9) to further estimate the right-hand side. We obtain

M ( 1 2 | d ρ ε | 2 + 1 ε 2 ( 1 - ρ ε ) 2 ) vol g C β + ( 1 - β ) M ε , β | D A ε u ε | 2 vol g C β + ( 1 - β ) | log ε | .

We divide both sides of this inequality by | log ε | and pass to the limit, first as ε 0 , then as β 1 . The claim (5.13) follows. In a similar way, we can prove that

(5.16) lim sup ε 0 1 | log ε | M ( 1 - ρ ε 2 ) | D A ε u ε | 2 vol g = 0 .

Finally, we have (see (1.1))

| D A ε u ε | 2 = ( 1 - ρ ε 2 ) | D A ε u ε | 2 + ρ ε 2 | D A ε u ε | 2 = ( 1 - ρ ε 2 ) | D A ε u ε | 2 + ρ ε 2 | d ρ ε | 2 + | j ( u ε , 0 ) | 2 .

Combining this identity with (5.10), (5.13), and (5.16), we obtain (5.9) and the proposition follows. ∎


Communicated by Guofang Wang


Award Identifier / Grant number: EXC-2047/1 - 390685813

Funding statement: Federico L. Dipasquale has been supported by the project Star Plus 2020 – Linea 1 (21-UNINA-EPIG-172) “New perspectives in the Variational modelling of Continuum Mechanics”. Giacomo Canevari and Federico L. Dipasquale thank the Hausdorff research Institute for Mathematics (HIM) for the warm hospitality during the Trimester Program “Mathematics of complex materials”, funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – EXC-2047/1 – 390685813, when part of this work was carried out. The authors have been supported by GNAMPA-INdAM.

A Poincaré- and trace-type inequalities for differential forms

In the appendix, we collect a few technical results on differential forms that are certainly part of the folklore. However, since these results are crucial to our analysis, we provide detailed proofs, for the convenience of the reader. The following proposition provides a Poincaré-type inequality and a trace-type inequality for co-closed k-forms normal to the boundary of geodesic balls. A crucial point is that in both inequalities we make explicit the dependence on the radius of the ball.

Proposition A.1.

Let M be a closed Riemannian manifold, x 0 M , B ρ ( x 0 ) the geodesic ball of radius ρ centred at x 0 , and let ω W 1 , p ( M , T * M ) , with p ( 1 , + ) , be a 1-form such that

(A.1) { d * ω = 0 in  ρ ( x 0 ) , i ν ω = 0 on  ρ ( x 0 ) .

There exists ρ ¯ > 0 , depending only on p and ( M , g ) , such that for all ρ ( 0 , ρ ¯ ) there holds

(A.2) ω L p ( ρ ( x 0 ) ) ρ d ω L p ( ρ ( x 0 ) ) ,

up to a constant depending only on M, p, and g. Moreover,

(A.3) ω L p ( ρ ( x 0 ) ) ρ 1 - 1 p d ω L p ( ρ ( x 0 ) ) ,

up to a constant depending only on M, p, and g.

Proof.

To prove (A.2), we start with the case of a Euclidean ball B ρ ( x 0 ) n with the Euclidean metric. First, we observe that in such case, by [21, Remark 4.10] and the Poincaré lemma, the only closed and co-closed 1-form h in B ρ ( x 0 ) satisfying i ν h = 0 on B ρ ( x 0 ) is the identically zero 1-form. Thus, by reasoning by contradiction similarly to as in the proof of the Gaffney-type inequality [21, Theorem 4.11], for any 1-form of class W 1 , p ( B ρ ( x 0 ) , ( n ) * ) that satisfies i ν ω = 0 on B ρ ( x 0 ) we get

(A.4) ω L p ( B ρ ( x 0 ) ) C p ( B ρ ( x 0 ) ) ( d ω L p ( B ρ ( x 0 ) ) + d 0 * ω L p ( B ρ ( x 0 ) ) ) ,

where ω := ( i ω j ) i , j = 1 n , d 0 * denotes the codifferential operator with respect to the Euclidean metric and C p ( B ρ ( x 0 ) ) is a constant depending only on p and B ρ ( x 0 ) . The dependence on p and B ρ ( x 0 ) can be made precise through a classical scaling argument: by (A.4) with ρ = 1 and x 0 = 0 and scaling, we obtain

(A.5) ω L p ( B ρ ( x 0 ) ) C p ( n ) ( d ω L p ( B ρ ( x 0 ) ) + d 0 * ω L p ( B ρ ( x 0 ) ) ) ,

where C p ( n ) is a positive constant depending only on n and p. In addition to (A.5), by the Gaffney-type inequality [21, Theorem 4.11], we obtain

(A.6) ω L p ( B ρ ( x 0 ) ) C p ( B ρ ( x 0 ) ) ( d ω L p ( B ρ ( x 0 ) ) + d 0 * ω L p ( B ρ ( x 0 ) ) ) ,

where C p ( B ρ ( x 0 ) ) is a constant depending only on p and B ρ ( x 0 ) . Like in the above, by (A.6) with ρ = 1 and x 0 = 0 and, again, a classical scaling argument, we easily obtain

(A.7) ω L p ( B ρ ( x 0 ) ) ρ ( d ω L p ( B ρ ( x 0 ) ) + d 0 * ω L p ( B ρ ( x 0 ) ) ) ,

up to constant depending only on n and p, which in particular establishes (A.2) in the Euclidean case. We remark explicitly that the absolute values and the integrals involved in the L p -norms in (A.7) are computed with respect to the Euclidean metric and the Euclidean volume form.

For the general case, we can work in normal coordinates centred at x 0 . Recall that the image of the geodesic ball ρ ( x 0 ) through the normal coordinates chart is the Euclidean ball B ρ (centred at the origin). Thus, it is enough to consider the case in which ρ ( x 0 ) = B ρ is a Euclidean ball centred at the origin endowed with the metric g of M.

Assume ω W 1 , p ( M , T * M ) is a 1-form on M satisfying (A.1). Since Gauss’ Lemma implies that the unit normal field is the same both when evaluated with respect to g and with respect to the Euclidean metric, (A.6) holds for ω. By hypothesis, we have d * ω = 0 . Writing d * ω in coordinates (employing Einstein notation), we see that

d * ω = 1 det g i ( det g g i j ω j ) = g i j i ω j + i ( det g g i j ) det g ω j = d 0 * ω + ( g i j - δ i j ) i ω j + i ( det g g i j ) det g ω j .

By the standard properties of normal coordinates, we have

(A.8) ( g i j - δ i j ) = O ( ρ 2 ) and i ( det g g i j ) det g = O ( ρ )

as ρ 0 . Thus, the assumption d * ω = 0 along with (A.8) leads to

(A.9) d 0 * ω L p ( B ρ ) ρ ( ω L p ( B ρ ) + ρ ω L p ( B ρ ) )

for any ρ sufficiently small and up to constant depending only on n, p and g. In (A.9), the norms are again computed with respect to the Euclidean metric and volume form. Choosing ρ sufficiently small (depending on p and ( M , g ) only), combining (A.5) and (A.9) we obtain

(A.10) ω L p ( B ρ ) C ( M ) ( ρ ω L p ( B ρ ) + d ω L p ( B ρ ) ) ,

where C p ( M ) is a constant depending only on p and ( M , g ) . Refining the choice of ρ if necessary (but still dependingly on p and ( M , g ) only), by (A.7), (A.9), and (A.10), we obtain

(A.11) ω L p ( B ρ ) ρ d ω L p ( B ρ )

for any ρ sufficiently small and up to constant depending only on n, p and ( M , g ) . Once more, the norms in (A.11) are computed with respect to the Euclidean metric and volume form. However, since M is compact, g is controlled from below and from above by the Euclidean metric. Similarly, vol g is controlled from below and from above by the Euclidean volume form, and hence (A.2) follows from (A.11) (up to replacing the implicit constant in (A.2) with a larger one, which however depends still only on p, M and g.

The proof of (A.3) follows a similar pattern. We begin again with a Euclidean ball endowed with the Euclidean metric. Then the inequality

(A.12) ω L p ( B ρ ) C p ( B ρ ) ( ω L p ( B ρ ) + d ω L p ( B ρ ) + d 0 * ω L p ( B ρ ) )

is proven by following the usual argument for the trace inequality for Sobolev functions and using the Gaffney-type inequality [21, Theorem 4.8]. Then, by scaling, we derive

(A.13) ω L p ( B ρ ) 1 ρ 1 p ω L p ( B ρ ) + ρ 1 - 1 p ( d ω L p ( B ρ ) + d 0 * ω L p ( B ρ ) ) ,

up to a constant depending only on n and p. In view of (A.1) and (A.7), this establishes (A.3) in the Euclidean case. For the general case, we obtain (A.3) from (A.13) arguing as in the proof of (A.2). Since the argument is very similar, we leave the details to the reader. ∎

Lemma A.2.

There exists a number R * > 0 such that, for any R ( 0 , R * ) , any x 0 M and any k-form ω W 1 , 2 ( B R ( x 0 ) , Λ k T * M ) with ω = 0 on B R ( x 0 ) , there holds

(A.14) R ( x 0 ) | ω | 2 vol g R 2 R ( x 0 ) ( | d ω | 2 + | d * ω | 2 ) vol g ,

where the implicit constant in front of the right-hand side is independent of R, x 0 .

Proof.

Let ω ¯ be the k-form on M defined by ω ¯ ( x ) := ω ( x ) if x R ( x 0 ) and ω ¯ ( x ) := 0 otherwise. As ω = 0 on R ( x 0 ) , we have ω ¯ W 1 , 2 ( M , Λ k T * M ) . Let p := 2 * = 2 n n - 2 . By applying Sobolev embeddings and the Gaffney inequality (in M, [34, Proposition 4.10]), we have

ω ¯ L p ( M ) ω ¯ W 1 , 2 ( M ) d ω ¯ L 2 ( M ) + d * ω ¯ L 2 ( M ) + ω ¯ L 2 ( M ) ,

where the implicit constants depend only on M (not on R or x 0 ). On the other hand, the Hölder inequality gives

ω L 2 ( R ( x 0 ) ) ( vol ( R ( x 0 ) ) ) 1 2 - 1 p ω ¯ L p ( M ) R d ω ¯ L 2 ( M ) + R d * ω ¯ L 2 ( M ) + R ω ¯ L 2 ( M ) .

If R < R * and R * is chosen small enough, then the last term on the right-hand side can be absorbed in the left-hand side, and the lemma follows. ∎

B Elliptic estimates for the Hodge Laplacian

In this section, we gather a few regularity estimates for elliptic problems involving differential forms. These results, as the ones contained in Appendix A, are classical, but we include detailed proofs for the reader’s convenience. We denote by Δ the Hodge–Laplace operator on differential forms, defined by - Δ := d d * + d * d , where d is the exterior differential and d * the codifferential.

Proposition B.1.

Let p > n 2 , let f L p ( M , Λ k T * M ) and let τ W 1 , 2 ( M , Λ k T * M ) be such that

(B.1) { - Δ τ = f - H ( f ) in the sense of distributions in  M , M τ , ξ vol g = 0 for any harmonic  k - form  ξ .

Then τ is continuous and, for any x M , there holds

(B.2) | τ ( x ) | M | f ( y ) | dist ( x , y ) n - 2 vol g ( y ) ,

where dist denotes the geodesic distance on M.

The proof of Proposition B.1 is based on the existence of a Green form for the Hodge–Laplace operator - Δ on k-forms. The Green form is a differential form 𝔤 on M × M of degree ( k , k ) – that is, at each point of M × M , 𝔤 can locally be written as

𝔤 ( x , y ) = I , J 𝔤 I J ( x , y ) d x I d y J ,

where ( x 1 , , x n ) , ( y 1 , , y n ) are local coordinate systems on M, the 𝔤 I J ( x , y ) are scalar coefficients, and the sum is taken over all multi-indices I, J of order k. The Green form 𝔤 is uniquely characterised by the following property: for any smooth k-form f on M, the unique solution τ of (B.1) can be written as

(B.3) τ ( x ) = M 𝔤 ( x , y ) f ( y ) for any  x M .

The existence of a Green form g was proved by de Rham (see [17, Chapter III, Section 21]), based on previous work by Bidal and de Rham who constructed a parametrix (i.e., an “approximate inverse”) for the Hodge–Laplace operator. The Green form is smooth in { ( x , y ) M × M : x y } and satisfies

(B.4) | 𝔤 ( x , y ) | dist ( x , y ) n - 2

for any x M , y M with x y (see [17]).

Proof of Proposition B.1.

If f is a smooth k-form, then estimate (B.2) is an immediate consequence of (B.3) and (B.4). Estimate (B.2) implies, via the Hölder inequality, an estimate for the L ( M ) -norm of τ in terms of the L p ( M ) -norm of f, for any p > n 2 . Indeed, if p > n 2 and q := p = p p - 1 , then q < n n - 2 and the function dist ( x , ) n - 2 belongs to L q ( M ) (in fact, the L q ( M ) -norm of dist ( x , ) n - 2 is uniformly bounded with respect to x). Then the proposition follows by a density argument. ∎

Lemma B.2.

Let x 0 M , 0 < R < inj ( M ) and let ω W 1 , 2 ( B R ( x 0 ) , Λ k T * M ) be such that - Δ ω = 0 in B R ( x 0 ) . Then

(B.5) ω W 1 , 2 ( R / 2 ) R - 1 ω L 2 ( R )

(where the implicit constant in front of the right-hand side is independent of x 0 , R).

Proof.

Let ζ C c ( R ) be a cut-off function such that ζ = 1 in R / 2 and | d ζ | R - 1 in R . There holds

(B.6) d ω , d ( ζ 2 ω ) = d ω , d ζ ζ ω + ζ d ( ζ ω ) = ζ d ω , d ζ ω + d ( ζ ω ) = d ( ζ ω ) - d ζ ω , d ζ ω + d ( ζ ω ) = | d ( ζ ω ) | 2 - | d ζ ω | 2

and similarly,

(B.7) d * ω , d * ( ζ 2 ω ) = d ( ω ) , d ( ζ 2 ω ) = | d ( ζ ω ) | 2 - | d ζ ω | 2 = | d * ( ζ ω ) | 2 - | d ζ ω | 2 .

Therefore, by testing the equation - Δ ω = 0 against ζ 2 ω , we obtain

R ( x 0 ) ( | d ( ζ ω ) | 2 + | d * ( ζ ω ) | 2 ) vol g = R ( x 0 ) ( | d ζ ω | 2 + | d ζ ω | 2 ) vol g R - 2 R ( x 0 ) | ω | 2 vol g .

By applying the Gaffney inequality (in M, [34, Proposition 4.10]), we obtain

ζ ω W 1 , 2 ( M ) R - 1 ω L 2 ( R ) + ζ ω L 2 ( M ) R - 1 ω L 2 ( R )

and the lemma follows. ∎

Thus, harmonic forms in geodesic balls satisfy a Caccioppoli-type inequality, exactly as in the Euclidean case. Moreover, exactly as in the case of harmonic functions on balls of n (see, e.g., [6, Lemma 4.11]), Caccioppoli inequality allows to prove a decay property of L 2 -norm of harmonic forms in geodesic balls.

Theorem B.3.

Let x 0 M , 0 < R < inj ( M ) and let ω W 1 , 2 ( B R ( x 0 ) , Λ k T * M ) be such that - Δ ω = 0 in B R ( x 0 ) . Then

(B.8) r ( x 0 ) | ω | 2 vol g r n R n R ( x 0 ) | ω | 2 vol g ,

where the implicit constant in front of the right-hand side is independent of x 0 , R.

Proof.

Let ζ C c ( R / 2 ) be a cut-off function such that ζ = 1 in R / 4 and | ζ | R - 1 , | 2 ζ | R - 2 . Let τ be the unique solution of

- Δ τ = - Δ ( ζ ω ) in  M , H ( τ ) = 0 .

(Note that τ exists, because Δ ( ζ ω ) is orthogonal to all harmonic forms.) Then ζ ω = τ + H ( ζ ω ) . Let x R / 8 . By Proposition B.1, we have

(B.9) | τ ( x ) | M | Δ ( ζ ω ) ( y ) | dist ( x , y ) n - 2 vol g ( y ) R 2 - n Δ ( ζ ω ) L 1 ( R / 2 R / 4 )

because Δ ( ζ ω ) = Δ ω = 0 in R / 4 and ζ ω = 0 out of R / 2 (and dist ( x , y ) R 8 for any x R / 8 , y R / 2 R / 4 ). By the Hölder inequality,

(B.10) | τ ( x ) | R 2 - n 2 Δ ( ζ ω ) L 2 ( R / 2 R / 4 ) .

Now, we have

- Δ ( ζ ω ) = d * ( d ζ ω + ζ d ω ) + d ( ± ( d ζ ω ) + ζ d * ω ) ) = = - ζ Δ ω + l . o . t . ,

where l.o.t. is a sum of terms that do not contain second derivatives of ω (but do contain derivatives of ζ). In particular, as Δ ω = 0 ,

(B.11) Δ ( ζ ω ) L 2 ( R / 2 ) R - 1 ω W 1 , 2 ( R / 2 ) + R - 2 ω L 2 ( R / 2 ) ( B .5 ) R - 2 ω L 2 ( R )

(the factors of R - 1 , R - 2 come from the derivatives of ζ). Combining (B.10) with (B.11), we deduce

(B.12) τ L ( R / 8 ) R - n 2 ω L 2 ( R ) .

The estimate of the L -norm of H ( ζ ω ) is immediate because H ( ζ ω ) is harmonic – and all norms are equivalent on Harm k ( M ) . In particular,

H ( ζ ω ) L ( M ) H ( ζ ω ) L 2 ( M ) ω L 2 ( M ) R - n 2 ω L 2 ( M )

(because R is bounded from above). Overall, we have proved

(B.13) ω L ( R / 8 ) R - n 2 ω L 2 ( R )

which immediately implies decay (cf., for instance, [6, Lemma 4.11].) ∎

In Proposition B.4 below, we extend the decay estimate in Theorem B.3 to forms solving Poisson-type equations with L 2 -integrable sources.

Proposition B.4.

There exists a number R * > 0 , depending only on M, such that the following statement holds. If x 0 M , 0 < R < R * and ω, f are k-form on B R ( x 0 ) such that - Δ ω = f in B R ( x 0 ) , then

(B.14) r ( x 0 ) ( | d ω | 2 + | d * ω | 2 ) vol g r n R n R ( x 0 ) ( | d ω | 2 + | d * ω | 2 ) vol g + R 2 R ( x 0 ) | f | 2 vol g

for every r ( 0 , R ) .

Proof.

The proof follows a very well-known pattern, typical for analogous estimates for elliptic systems in the Euclidean setting (see, e.g., [6, Lemma 4.13]). We provide details for completeness, as for differential forms on manifolds they are apparently lacking in the literature (to the best to our knowledge, at least).

Step 1.

Fix arbitrarily x 0 M . For the time being, let R * = inj ( M ) , and pick any R ( 0 , R * ) . During the whole proof, we omit to indicate the center of the balls, as it will be always x 0 . As a preliminary step, we write ω = ω 1 + ω 2 , where ω 1 satisfies

(B.15) { Δ ω 1 = 0 in  R , ω 1 = ω on  R ,

while ω 2 is the solution to

(B.16) { - Δ ω 2 = f in  R , ω 2 = 0 on  R .

In the next steps, we first obtain suitable estimates for ω 1 and ω 2 , and then we combine them to yield (B.14).

Step 2 (Estimate for ω 1 ).

Since Δ ω 1 = 0 in R , we have Δ ( d ω 1 ) = 0 and Δ ( d * ω 1 ) = 0 in R as well. Thus, the decay estimate (B.8) holds for both d ω 1 and d * ω 1 and yields

(B.17) r ( | d ω 1 | 2 + | d * ω 1 | 2 ) vol g r n R n R ( | d ω 1 | 2 + | d * ω 1 | 2 ) vol g

for any r ( 0 , R ) , where the implicit constant at right-hand-side depends only on M and g. On the other hand, ω 1 minimises the functional

(B.18) σ R ( | d σ | 2 + | d * σ | 2 ) vol g

among all k-forms σ such that σ = ω 1 on R . This fact, along with (B.17), yields

(B.19) r ( | d ω 1 | 2 + | d * ω 1 | 2 ) vol g r n R n R ( | d ω | 2 + | d * ω | 2 ) vol g .

Step 3 (Estimate for ω 2 ).

The estimate for ω 2 follows by testing (B.16) against ω 2 . We obtain

(B.20) R ( | d ω 2 | 2 + | d * ω 2 | 2 ) vol g f L 2 ( R ) ω 2 L 2 ( R ) .

Notice that there are no boundary terms in (B.20) because ω 2 = 0 on R . By Lemma A.2, up to shrink R * if necessary (but still according only to M and g), we get

(B.21) ω 2 L 2 ( R ) R ( d ω 2 L 2 ( R ) + d * ω 2 L 2 ( R ) ) ,

and therefore

(B.22) d ω 2 L 2 ( R ) + d * ω 2 L 2 ( R ) R f L 2 ( R ) ,

up to a constant depending only on M and g.

Step 4 (Conclusion).

We can now estimate

r ( | d ω | 2 + | d * ω | 2 ) vol g r ( | d ω 1 | 2 + | d * ω 1 | 2 + | d ω 2 | 2 + | d * ω 2 | 2 ) vol g
( B .19 ) , ( B .20 )     r n R n R ( | d ω 1 | 2 + | d * ω 1 | 2 ) vol g + R 2 R | f | 2 vol g
r n R n R ( | d ω | 2 + | d * ω | 2 ) vol g + R 2 R | f | 2 vol g ,

where the last line (which provides the desired inequality) follows because ω 1 minimises the functional (B.18) among k-forms with the same value on the boundary of R . Then the conclusion follows by the arbitrariness of x 0 and the compactness of M, which allows to choose a uniform critical radius R * .∎

We conclude this section by giving the proof of a few Caccioppoli-type inequalities, which appeared in the proof of Proposition 4.1.

Lemma B.5.

There exists a number R * > 0 , depending only on M, such that the following statement holds. Let x 0 M , 0 < r < R < R * , a L ( B R ( x 0 ) ) and a 0 > 0 be such that a 0 - 1 a ( x ) a 0 for a.e. x B R ( x 0 ) . Moreover, let f L 2 ( B R ( x 0 ) , T * M ) and let θ W 1 , 2 ( B R ( x 0 ) ) be a weak solution of the equation

(B.23) d * ( a d θ ) = d * f in  R ( x 0 ) .

Then

(B.24) d θ L 2 ( r ( x 0 ) ) ( R - r ) - n + 2 2 θ L 1 ( R ( x 0 ) ) + f L 2 ( R ( x 0 ) ) .

Estimate (B.24) is a Caccioppoli inequality, except that the right-hand side contains the L 1 -norm of θ instead of the usual L 2 norm. Therefore, we include a proof of Lemma B.5, for the reader’s convenience.

Proof of Lemma B.5.

In the proof, we will apply the following inequality: for any x 0 M , 0 < σ < inj ( M ) and θ W 1 , 2 ( σ ( x 0 ) ) , there holds

(B.25) θ L 2 ( σ ( x 0 ) ) d θ L 2 ( σ ( x 0 ) ) n n + 2 θ L 1 ( σ ( x 0 ) ) 2 n + 2 + σ - n 2 θ L 1 ( σ ( x 0 ) ) ,

where the implicit constant in front of the right-hand side does not depend on σ, x 0 . In case σ ( x 0 ) is a ball in n , equipped with the Euclidean metric, the estimate (B.25) is a special case of Gagliardo and Nirenberg’s interpolation inequality, also known as Nash’s inequality (see [27]). A scaling argument shows that the implicit constant in front of the right-hand side is independent of σ. When σ ( x 0 ) is a geodesic ball on M, of radius σ smaller than the injectivity radius of M, (B.25) follows from its Euclidean counterpart, up to composition with local (e.g., geodesic) coordinates.

Let ρ, σ be positive numbers such that r < ρ < σ < R . We consider the Caccioppoli inequality

(B.26) B ρ | d θ | 2 vol g ( σ - ρ ) - 2 σ θ 2 vol g + σ | f | 2 vol g .

For simplicity of notation, we have dropped the dependence on x 0 in all balls. Estimate (B.26) is obtained by considering a suitable cut-off function ζ C c ( σ ) and testing equation (B.23) against ζ 2 θ . Estimate (B.26), combined with (B.25) and the Young inequality, implies

(B.27) d θ L 2 ( ρ ) ( σ - ρ ) - 1 d θ L 2 ( σ ) n n + 2 θ L 1 ( σ ) 2 n + 2 + ( σ - ρ ) - n + 2 2 θ L 1 ( σ ) + f L 2 ( R ) δ d θ L 2 ( σ ) + C δ ( σ - ρ ) - n + 2 2 θ L 1 ( R ) + f L 2 ( R )

for an arbitrary δ ( 0 , 1 ) and a constant C δ that depends only on δ, a 0 and ( M , g ) . As inequality (B.27) is valid for arbitrary values of ρ, σ with r < ρ < σ < R , it can be applied iteratively. The lemma follows by an iteration argument (see [20, Lemma V.3.1] or [6, Lemma B.1]). ∎

There is another Caccioppoli-type inequality we need, which follows along the same lines of Lemma B.5 but applies to k-forms instead of scalar functions. As a preliminary result, we first provide a version of the Nash inequality (B.25) for differential forms.

Lemma B.6.

For any x 0 M , any 0 < r < R < inj ( M ) 2 and any ω W 1 , 2 ( B R ( x 0 ) , Λ k T * M ) , there holds

ω L 2 ( r ( x 0 ) ) d ω L 2 ( R ( x 0 ) ) n n + 2 ω L 1 ( R ( x 0 ) ) 2 n + 2 + d * ω L 2 ( R ( x 0 ) ) n n + 2 ω L 1 ( R ( x 0 ) ) 2 n + 2 + ( R - r ) - n 2 ω L 1 ( R ( x 0 ) ) ,

where the implicit constant in front of the right-hand side does not depend on r, R, x 0 .

Proof.

All the balls we consider are centred at x 0 , so we write r , R instead of r ( x 0 ) , R ( x 0 ) and so on. Without loss of generality, we can assume that r R 2 . Let ρ and σ be positive numbers such that r < ρ < σ < R . Let ω W 1 , 2 ( R , Λ k T * M ) . Using geodesic normal coordinates ( x 1 , , x n ) centred at x 0 , we can write ω component-wise as ω = α ω α d x α , where the sum is taken over all multi-indices α of order k. By applying inequality (B.25) (on the ball ρ ) to each component of ω, we obtain

(B.28) ω L 2 ( ρ ) ω W 1 , 2 ( ρ ) n n + 2 ω L 1 ( ρ ) 2 n + 2 + ρ - n 2 ω L 1 ( ρ ) .

The implicit constant in front of the right-hand side depends on the metric of g, but it can be estimated uniformly with respect to x 0 and ρ. Now, let ζ C c ( σ ) be a cut-off function such that ζ = 1 in ρ and | ζ | ( σ - ρ ) - 1 . By applying the Gaffney inequality [34, Proposition 4.10] to ζ ω W 1 , 2 ( M , Λ k T * M ) , we obtain

(B.29) ω W 1 , 2 ( ρ ) d ( ζ ω ) L 2 ( M ) + d * ( ζ ω ) L 2 ( M ) + ζ ω L 2 ( M ) d ω L 2 ( σ ) + d * ω L 2 ( σ ) + ( σ - ρ ) - 1 ω L 2 ( σ ) .

By combining (B.28) with (B.29) (and observing that ρ σ - ρ because we have assumed that r R 2 ), we obtain

ω L 2 ( ρ ) d ω L 2 ( R ) n n + 2 ω L 1 ( R ) 2 n + 2 + d * ω L 2 ( R ) n n + 2 ω L 1 ( R ) 2 n + 2
+ ( σ - ρ ) - n n + 2 ω L 2 ( σ ) n n + 2 ω L 1 ( R ) 2 n + 2 + ( σ - ρ ) - n 2 ω L 1 ( R ) .

By applying the Young inequality at the right-hand side, for each δ > 0 we find a constant C δ such that

(B.30) ω L 2 ( ρ ) d ω L 2 ( R ) n n + 2 ω L 1 ( R ) 2 n + 2 + d * ω L 2 ( R ) n n + 2 ω L 1 ( R ) 2 n + 2 + C δ ( σ - ρ ) - n 2 ω L 1 ( R ) + δ ω L 2 ( σ ) .

If we choose δ small enough (uniformly with respect to r, R, x 0 ), the term δ ω L 2 ( σ ) at the right-hand side can be “absorbed into the left-hand side”, in a manner of speaking, by means of an iteration argument (see [20, Lemma V.3.1] or [6, Lemma B.1]). The lemma follows. ∎

Lemma B.7.

There exists a number R * > 0 , depending only on M, such that the following statement holds. Let x 0 M , 0 < r < R < R * and c L ( B R ( x 0 ) ) be such that c 0 . Let f L 2 ( B R ( x 0 ) , Λ k T * M ) and let ω W 1 , 2 ( B R ( x 0 ) ) be a weak solution of the equation

(B.31) - Δ ω + c ω = f in  R ( x 0 ) .

Then

ω W 1 , 2 ( r ( x 0 ) ) ( R - r ) - n + 2 2 ω L 1 ( R ( x 0 ) ) + R f L 2 ( R ( x 0 ) ) .

Proof.

Once again, we write ρ for a generic (geodesic) ball of radius ρ and center x 0 . Let r := r + R 2 , let ρ, σ be positive numbers such that r < ρ < σ < R , and let τ := 1 2 ( ρ + σ ) . Let ζ C c ( τ ) be a cut-off function such that ζ = 1 in ρ and | ζ | ( τ - ρ ) - 1 . By testing (B.31) against ζ 2 ω , and keeping (B.6), (B.7) into account, we deduce

ρ ( | d ω | 2 + | d * ω | 2 ) vol g ( τ - ρ ) - 2 τ | ω | 2 vol g + τ 2 τ | f | 2 vol g .

By applying Lemma B.6 and the Young inequality, we obtain

d ω L 2 ( ρ ) + d * ω L 2 ( ρ ) ( σ - ρ ) - 1 d ω L 2 ( σ ) n n + 2 ω L 1 ( σ ) 2 n + 2 + ( σ - ρ ) - 1 d * ω L 2 ( σ ) n n + 2 ω L 1 ( σ ) 2 n + 2 + ( σ - ρ ) - n + 2 2 ω L 1 ( σ ) + R f L 2 ( R ) δ d ω L 2 ( σ ) + δ d * ω L 2 ( σ ) + C δ ( σ - ρ ) - n + 2 2 ω L 1 ( σ ) + R f L 2 ( R )

for an arbitrary δ > 0 and some constant C δ depending on δ, but not on ρ, σ. An iteration argument (see [20, Lemma V.3.1] or [6, Lemma B.1]) now gives

(B.32) d ω L 2 ( r ) + d * ω L 2 ( r ) ( R - r ) - n + 2 2 ω L 1 ( R ) + R f L 2 ( R ) ,

where, we recall, r = r + R 2 . On the other hand, setting r ′′ := r + r 2 , Lemma B.6 and the (weighted) Young inequality imply

(B.33) ω L 2 ( r ′′ ) ( r - r ′′ ) d ω L 2 ( r ) + ( r - r ′′ ) d * ω L 2 ( r ) + ( r - r ′′ ) - n 2 ω L 1 ( r ) .

Finally, the Gaffney-type inequality (B.29) gives

(B.34) ω W 1 , 2 ( r ) d ω L 2 ( r ′′ ) + d * ω L 2 ( r ′′ ) + ( r ′′ - r ) - 1 ω L 2 ( r ′′ ) .

Combining (B.32) with (B.33) and (B.34), the lemma follows. ∎

C Proof of Lemma 1.6

In this section, we prove Lemma 1.6. For the reader’s convenience, we recall the statement below.

Lemma C.1.

Let ε > 0 and let ( u ε , A ε ) E be a critical point of G ε satisfying (9). Then

(C.1) D A ε u ε L ( M ) ε - 1 .

Proof.

The proof is based on elliptic regularity estimates, along the lines of [10, Proposition II.6] and [29, Appendix]. We provide the details for the reader’s convenience.

Step 1 (Reduction of the problem).

Let x 0 M be fixed. Locally, in a neighbourhood of x 0 , the reference connection D 0 can be written as D 0 = d - i γ 0 , where γ 0 is a 1-form. Let A ~ ε := A ε + γ 0 , so that D A ε = d - i A ~ ε and d A ~ ε = F ε . By changing gauge (see, e.g., [28, Proposition 2.1]), we can always assume without loss of generality that A ~ ε satisfies

(C.2) { d * A ~ ε = 0 in  ε ( x 0 ) , i ν A ~ ε = 0 on  ε ( x 0 )

for any ε > 0 small enough. Taking (C.2) into account, the Euler–Lagrange equations (6)–(7) on the ball ε ( x 0 ) rewrite as

(C.3) - Δ u ε + 2 i d u ε , A ~ ε + | A ~ ε | 2 u ε + 1 ε 2 ( | u ε | 2 - 1 ) u ε = 0
(C.4) - Δ A ~ ε = ( d - i A ~ ε ) u ε , i u ε .

Moreover, the monotonicity formula – Theorem 3 – and the assumption (9) imply

(C.5) G ε ( u ε , A ε ; ε ( x 0 ) ) ε n - 2 G ε ( u ε , A ε ; R ( x 0 ) ) R n - 2 | log ε | ,

where R is any fixed radius such that ε < R < inj ( M ) . Note that we are in a position to apply Theorem 3, which is independent of Lemma C.1 – see Remark 3.7.

Step 2 (Scaling).

Next, we perform a scaling. To this end, we identify the tangent plane to M at the point x 0 with n , and consider the exponential map at x 0 , exp x 0 : U n M , defined in a neighbourhood U of the origin. Let D n n be the (open) unit ball. For ε small enough, we define v ε : D n , B ε : D n ( n ) * as

v ε ( x ) := u ε ( exp x 0 ( ε x ) ) ,
B ε ( x ) := ε A ~ ε ( exp x 0 ( ε x ) )

for x D n (up to composition with local trivialisations). The maximum principle for u ε (Proposition 1.2) gives

(C.6) | v ε | 1 in  D n .

In order to prove Lemma C.1, it suffices to show that

(C.7) | ( d - i B ε ) v ε ( 0 ) | C

for some uniform constant C; the bound (C.1) will follow from (C.7) by scaling back. We prove (C.7) by applying elliptic regularity theory. The equations for v ε , B ε are

(C.8) - Δ ε v ε + 2 i d v ε , B ε + | B ε | 2 v ε + ( | v ε | 2 - 1 ) v ε = 0 ,
(C.9) - 1 ε 2 Δ ε B ε = ( d - i B ε ) v ε , i v ε .

Here - Δ ε is the Hodge Laplacian on D n , for (complex-valued) 0-forms and 1-forms respectively, with respect to the metric g ε given by

(C.10) g ε ( x ) := g ( exp x 0 ( ε x ) )

for x D n . The operator - Δ ε is a small perturbation of the Euclidean Laplacian (see (A.8)). The form B ε is in Coulomb gauge, i.e., it satisfies (C.2).

Step 3 (The energy bound and its consequences).

The energy estimate (C.5) implies, by scaling,

(C.11) I ε ( v ε , B ε ) := D n ( 1 2 | ( d - i B ε ) v ε | 2 + 1 4 ( 1 - | v ε | 2 ) 2 + 1 2 ε 2 | d B ε | 2 ) vol g ε | log ε | .

This estimate, combined with (C.2) and Gaffney’s inequality (see, e.g., (A.7))

(C.12) B ε W 1 , 2 ( D n ) d B ε L 2 ( D n ) ε | log ε | 1 2 .

In particular, by Sobolev embedding, there exists an exponent p ( 1 , 2 ) – depending only on n – such that

(C.13) | B ε | p L 2 ( D n ) ε | log ε | 1 2 .

For future reference, we also note that

(C.14) d v ε L 2 ( D n ) ( d - i B ε ) v ε L 2 ( D n ) + B ε L 2 ( D n ) | log ε | 1 2 ,

because of (C.11) and (C.12).

Step 4 (A Bochner-type estimate).

Let

h ε := ε 2 + ε 2 | v ε | 2 + | B ε | 2 , f ε := h ε p 2 ,

where p ( 1 , 2 ) is precisely the same number as in (C.13). We claim that f ε satisfies

(C.15) - Δ ε f ε f ε

in the sense of distributions in D n . First, Weitzenböck-Bochner identity (see, e.g., [18, Equation (2.20)]) states

- Δ ε ( 1 2 | B ε | 2 ) = Δ ε B ε , B ε - | D ε B ε | 2 - Ric ε ( B ε # , B ε # ) ,

where D ε is the Levi-Civita connection induced by the metric (C.10), Ric ε the Ricci tensor induced by (C.10), and B ε # is the vector field associated with B ε . By injecting (C.9) and (C.6) into the right-hand side, we obtain

(C.16) - 1 2 Δ ε ( | B ε | 2 ) ε 2 | d v ε - i B ε | | B ε | - | D ε B ε | 2 + C | B ε | 2

for some constant C that depends only on M (in particular, C is independent of ε). Moreover, as in Proposition 1.2, we have

(C.17) - 1 2 Δ ε ( | v ε | 2 ) = ( 1 - | v ε | 2 ) | v ε | 2 - | ( d - i B ε ) v ε | 2 ( C .6 ) 1 - | ( d - i B ε ) v ε | 2 .

Combining (C.16) with (C.17) applying Young’s inequality, and recalling that 1 < p < 2 , we find a constant C (depending on p and M only) such that

(C.18) - 1 2 Δ ε h ε ( p - 2 ) ( ε 2 | d v ε | 2 + | D ε B ε | 2 ) + C ( | B ε | 2 + ε 2 )

for any ε small enough. On the other hand, the Cauchy–Schwarz inequality gives

(C.19) | d h ε | 2 4 ( ε 2 | v ε | | d v ε | + | B ε | | D ε B ε | ) 2 4 h ε ( ε 2 | d v ε | 2 + | D ε B ε | 2 ) .

From (C.18) and (C.19), we deduce

(C.20) - Δ ε h ε ( p 2 - 1 ) | d h ε | 2 h ε + 2 C h ε .

Now, a formal computation shows that

- Δ ε f ε = - p 2 ( p 2 - 1 ) h ε p 2 - 2 | d h ε | 2 - p 2 h ε p 2 - 1 Δ h ε ( C .20 ) C p h ε p 2 = C p f ε ,

which formally proves (C.15). This computation can be made rigourous by observing that f ε ( W 1 , 2 W loc 2 , 1 ) ( D n ) (see [29, Lemma A.2] for details).

Step 5 (Conclusion).

Let D r n be the open ball in n centred at the origin, with radius r > 0 . Inequality (C.15) implies, via Caccioppoli inequalities and Moser iteration, that

(C.21) f ε L ( D 1 / 2 n ) f ε L 2 ( D 1 / 2 n ) .

The implicit constant in front of the right-hand side of (C.21) does not depend on ε, even though the left-hand side of (C.15) does, because the coefficients of the metric g ε are bounded uniformly with respect to ε and so is the norm of the Sobolev embedding with respect to g ε . Now, using (C.6), (C.13), and (C.21), we obtain

(C.22) B ε L ( D 1 / 2 n ) p ε p + | B ε | p L 2 ( D n ) ε | log ε | 1 2 .

This estimate, combined with (C.6), (C.9), and (C.14), implies that

- Δ ε v ε L 2 ( D 1 / 2 n ) C

for any ε small enough. By elliptic regularity, it follows that the W 2 , 2 -norm of v ε in D 1 / 4 n is bounded uniformly with respect to ε. We are now in a position to apply a bootstrapping argument to system (C.8)–(C.9), so as to obtain uniform W 2 , q bounds for the solution on a small ball containing the origin, for arbitrary q < + . By Sobolev embedding, we deduce (C.7). ∎

References

[1] G. Alberti, S. Baldo and G. Orlandi, Functions with prescribed singularities, J. Eur. Math. Soc. (JEMS) 5 (2003), no. 3, 275–311. 10.1007/s10097-003-0053-5Search in Google Scholar

[2] G. Alberti, S. Baldo and G. Orlandi, Variational convergence for functionals of Ginzburg–Landau type, Indiana Univ. Math. J. 54 (2005), no. 5, 1411–1472. 10.1512/iumj.2005.54.2601Search in Google Scholar

[3] F. J. Almgren, The theory of varifolds, Mimeographed notes, Princeton University, Princeton, 1965. Search in Google Scholar

[4] L. Ambrosio and H. M. Soner, A measure-theoretic approach to higher codimension mean curvature flows, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 25 (1997), no. 1–2, 27–49. Search in Google Scholar

[5] S. Baraket, Critical points of the Ginzburg–Landau system on a Riemannian surface, Asymptot. Anal. 13 (1996), no. 3, 277–317. 10.3233/ASY-1996-13303Search in Google Scholar

[6] L. Beck, Elliptic Regularity Theory, Lect. Notes Unione Mat. Ital. 19, Springer, Cham, 2016. 10.1007/978-3-319-27485-0Search in Google Scholar

[7] F. Bethuel, H. Brezis and F. Hélein, Ginzburg–Landau Vortices, Progr. Nonlinear Differential Equations Appl. 13, Birkhäuser, Boston, 1994. 10.1007/978-1-4612-0287-5Search in Google Scholar

[8] F. Bethuel, H. Brezis and G. Orlandi, Asymptotics for the Ginzburg–Landau equation in arbitrary dimensions, J. Funct. Anal. 186 (2001), no. 2, 432–520. 10.1006/jfan.2001.3791Search in Google Scholar

[9] F. Bethuel, G. Orlandi and D. Smets, Convergence of the parabolic Ginzburg–Landau equation to motion by mean curvature, Ann. of Math. (2) 163 (2006), no. 1, 37–163. 10.4007/annals.2006.163.37Search in Google Scholar

[10] F. Bethuel and T. Rivière, Vortices for a variational problem related to superconductivity, Ann. Inst. H. Poincaré C Anal. Non Linéaire 12 (1995), no. 3, 243–303. 10.1016/s0294-1449(16)30157-3Search in Google Scholar

[11] F. Bethuel and X. M. Zheng, Density of smooth functions between two manifolds in Sobolev spaces, J. Funct. Anal. 80 (1988), no. 1, 60–75. 10.1016/0022-1236(88)90065-1Search in Google Scholar

[12] H. Brezis and P. Mironescu, Sobolev Maps to the Circle—from the Perspective of Analysis, Geometry, and Topology, Progr. Nonlinear Differential Equations Appl. 96, Birkhäuser, New York, 2021. 10.1007/978-1-0716-1512-6Search in Google Scholar

[13] G. Canevari, F. L. Dipasquale and G. Orlandi, The Yang–Mills–Higgs functional on complex line bundles: Γ-convergence and the London equation, Arch. Ration. Mech. Anal. 247 (2023), no. 6, Paper No. 104. 10.1007/s00205-023-01933-1Search in Google Scholar

[14] D. R. Cheng, Instability of solutions to the Ginzburg–Landau equation on S n and n , J. Funct. Anal. 279 (2020), no. 8, Article ID 108669. 10.1016/j.jfa.2020.108669Search in Google Scholar

[15] A. Colinet, R. Jerrard and P. Sternberg, Solutions of the Ginzburg–Landau equations with vorticity concentrating near a nondegenerate geodesic, J. Eur. Math. Soc. (JEMS) (2023), 10.4171/JEMS/1397. 10.4171/JEMS/1397Search in Google Scholar

[16] G. De Philippis and A. Pigati, Non-degenerate minimal submanifolds as energy concentration sets: A variational approach, Comm. Pure Appl. Math. 77 (2024), no. 8, 3581–3627. 10.1002/cpa.22193Search in Google Scholar

[17] G. de Rham and K. Kodaira, Harmonic Integrals, Institute for Advanced Study, Princeton, 1954. Search in Google Scholar

[18] J. Eells and L. Lemaire, A report on harmonic maps, Bull. Lond. Math. Soc. 10 (1978), no. 1, 1–68. 10.1112/blms/10.1.1Search in Google Scholar

[19] H. Federer, Geometric Measure Theory, Grundlehren Math. Wiss. 153, Springer, New York, 1969. Search in Google Scholar

[20] M. Giaquinta, Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems, Ann. of Math. Stud. 105, Princeton University, Princeton, 1983. 10.1515/9781400881628Search in Google Scholar

[21] T. Iwaniec, C. Scott and B. Stroffolini, Nonlinear Hodge theory on manifolds with boundary, Ann. Mat. Pura Appl. (4) 177 (1999), 37–115. 10.1007/BF02505905Search in Google Scholar

[22] A. Jaffe and C. Taubes, Vortices and Monopoles, Progr. Phys. 2, Birkhäuser, Boston, 1980. Search in Google Scholar

[23] R. L. Jerrard and H. M. Soner, The Jacobian and the Ginzburg–Landau energy, Calc. Var. Partial Differential Equations 14 (2002), no. 2, 151–191. 10.1007/s005260100093Search in Google Scholar

[24] R. L. Jerrard and P. Sternberg, Critical points via Γ-convergence: General theory and applications, J. Eur. Math. Soc. (JEMS) 11 (2009), no. 4, 705–753. 10.4171/jems/164Search in Google Scholar

[25] F.-H. Lin and T. Rivière, A quantization property for static Ginzburg–Landau vortices, Comm. Pure Appl. Math. 54 (2001), no. 2, 206–228. 10.1002/1097-0312(200102)54:2<206::AID-CPA3>3.0.CO;2-WSearch in Google Scholar

[26] D. Monclair, Groups and geometry. Search in Google Scholar

[27] J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer. J. Math. 80 (1958), 931–954. 10.2307/2372841Search in Google Scholar

[28] G. Orlandi, Asymptotic behavior of the Ginzburg–Landau functional on complex line bundles over compact Riemann surfaces, Rev. Math. Phys. 8 (1996), no. 3, 457–486. 10.1142/S0129055X96000159Search in Google Scholar

[29] A. Pigati and D. Stern, Minimal submanifolds from the abelian Higgs model, Invent. Math. 223 (2021), no. 3, 1027–1095. 10.1007/s00222-020-01000-6Search in Google Scholar

[30] A. Pigati and D. Stern, Quantization and non-quantization of energy for higher-dimensional Ginzburg–Landau vortices, Ars Inven. Anal. 2023 (2023), Paper No. 3. Search in Google Scholar

[31] J. T. Pitts, Existence and Regularity of Minimal Surfaces on Riemannian Manifolds, Math. Notes 27, Princeton University, Princeton, 1981. 10.1515/9781400856459Search in Google Scholar

[32] J. Qing, Renormalized energy for Ginzburg–Landau vortices on closed surfaces, Math. Z. 225 (1997), no. 1, 1–34. 10.1007/PL00004303Search in Google Scholar

[33] E. Sandier and S. Serfaty, Vortices in the Magnetic Ginzburg–Landau Model, Progr. Nonlinear Differential Equations Appl. 70, Birkhäuser, Boston, 2007. 10.1007/978-0-8176-4550-2Search in Google Scholar

[34] C. Scott, L p theory of differential forms on manifolds, Trans. Amer. Math. Soc. 347 (1995), no. 6, 2075–2096. 10.1090/S0002-9947-1995-1297538-7Search in Google Scholar

[35] L. Simon, Lectures on Geometric Measure Theory, Proc. Centre Math. Appl. Austral. Nat. Univ. 3, Australian National University, Canberra, 1984. Search in Google Scholar

[36] P. Smith and K. Uhlenbeck, Removeability of a codimension four singular set for solutions of a Yang Mills Higgs equation with small energy, Surveys in Differential Geometry 2019. Differential Geometry, Calabi–Yau Theory, and General Relativity. Part 2, Surv. Differ. Geom. 24, International Press, Boston, (2022), 257–291. 10.4310/SDG.2019.v24.n1.a7Search in Google Scholar

[37] D. Stern, p-harmonic maps to S 1 and stationary varifolds of codimension two, Calc. Var. Partial Differential Equations 59 (2020), no. 6, Paper No. 187. 10.1007/s00526-020-01859-6Search in Google Scholar

[38] D. Stern, Existence and limiting behavior of min-max solutions of the Ginzburg–Landau equations on compact manifolds, J. Differential Geom. 118 (2021), no. 2, 335–371. 10.4310/jdg/1622743143Search in Google Scholar

Received: 2023-05-25
Accepted: 2024-08-17
Published Online: 2024-09-03
Published in Print: 2025-01-01

© 2024 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 16.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2023-0064/html
Scroll to top button